15.02.2013 Views

Nature Immunology

Nature Immunology

Nature Immunology

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

www.nature.com/natureimmunology<br />

EDITORIAL OFFICE immunology@natureny.com<br />

75 Varick Street, Fl 9, New York, NY 10013-1917<br />

Tel: (212) 726 9207, Fax: (212) 696 9752<br />

Editor: Jamie D. K. Wilson<br />

Senior Editors: Christine Borowski, Laurie A. Dempsey<br />

Copy Editor: Jennifer Fosmire<br />

Production Editors: Marina Corral, Matt Hansen<br />

Senior Illustrator: Katie Vicari<br />

Illustrator: Kimberly Caesar<br />

Senior Editorial Assistant: Adam Lipkin<br />

MANAGEMENT OFFICES<br />

NPG New York<br />

75 Varick Street, Fl 9, New York, NY 10013-1917<br />

Tel: (212) 726 9200, Fax: (212) 696 9006<br />

Executive Editor: Linda Miller<br />

Chief Technology Officer: Howard Ratner<br />

Head of <strong>Nature</strong> Research & Reviews Marketing: Sara Girard<br />

Marketing Manager: Leah Rodriguez<br />

Assistant Production Coordinator: Karen Wilson<br />

Head of Web Services: Anthony Barrera<br />

Web Production Editor: Colleen Diez<br />

NPG London<br />

The Macmillan Building, 4 Crinan Street, London N1 9XW<br />

Tel: 44 207 833 4000, Fax: 44 207 843 4996<br />

Managing Director: Steven Inchcoombe<br />

Publishing Director: Alison Mitchell<br />

Publisher: Stephanie Diment<br />

Editor-in-Chief, <strong>Nature</strong> Publications: Philip Campbell<br />

Marketing Director: Della Sar<br />

Director of Web Publishing: Timo Hannay<br />

NPG <strong>Nature</strong> Asia-Pacific<br />

Chiyoda Building, 2-37 Ichigayatamachi, Shinjuku-ku, Tokyo 162-0843<br />

Tel: 81 3 3267 8751, Fax: 81 3 3267 8746<br />

Publishing Director — Asia-Pacific: David Swinbanks<br />

Associate Director: Antoine E. Bocquet<br />

Manager: Koichi Nakamura<br />

Senior Marketing Manager: Peter Yoshihara<br />

Asia-Pacific Sales Director: Kate Yoneyama<br />

Asia-Pacific Sales Manager: Ken Mikami<br />

DISPLAY ADVERTISING display@natureny.com (US/Canada)<br />

display@nature.com (Europe)<br />

nature@natureasia.com (Asia)<br />

Global Head of Display Advertising Sales: John Michael, Tel: 44 207 843 4960, Fax: 44 207<br />

843 4996<br />

Asia-Pacific Sales Manager: Ken Mikami, Tel: 81 3 3267 8765, Fax: 81 3 3267 8746<br />

Display Account Managers:<br />

New England: Sheila Reardon, Tel: (617) 399 4098, Fax: (617) 426 3717<br />

New York/Mid-Atlantic/Southeast: Jim Breault, Tel: (212) 726 9334, Fax: (212) 696 9481<br />

Midwest: Mike Rossi, Tel: (212) 726 9255, Fax: (212) 696 9481<br />

West Coast South: George Lui, Tel: (415) 781 3804, Fax: (415) 781 3805<br />

West Coast North: Bruce Shaver, Tel: (415) 781 6422, Fax: (415) 781 3805<br />

Germany/Switzerland/Austria: Sabine Hugi-Fürst, Tel: 41 52761 3386, Fax: 41 52761 3419<br />

United Kingdom/Ireland/France/Belgium/Eastern Europe: Jeremy Betts, Tel: 44 207 843 4968,<br />

Fax: 44 207 843 4749<br />

Scandinavia/The Netherlands/Italy/Spain/Portugal/Israel/Iceland: Graham Combe,<br />

Tel: 44 207 843 4914, Fax: 44 207 843 4749<br />

Greater China/Singapore: Gloria To, Tel: 852 2811 7191, Fax: 852 2811 0743<br />

NATUREJOBS naturejobs@natureny.com (US/Canada)<br />

naturejobs@nature.com (Europe)<br />

nature@natureasia.com (Asia)<br />

US Sales Manager: Peter Bless, Tel: (212) 726 9248, Fax: (212) 696 9482<br />

European Sales Manager: Andrew Douglas, Tel: 44 207 843 4975, Fax: 44 207 843 4996<br />

Asia-Pacific Sales Manager: Ayako Watanabe, Tel: 81 3 3267 8765, Fax: 81 3 3267 8746<br />

SITE LICENSE BUSINESS UNIT<br />

Americas: Tel: (888) 331 6288 institutions@natureny.com<br />

Asia/Pacific: Tel: 81 3 3267 8751 institutions@natureasia.com<br />

Australia/New Zealand: Tel: 61 3 9825 1160 nature@macmillan.com.au<br />

India: Tel: 91 124 2881054/55 npgindia@nature.com<br />

ROW: Tel: 44 207 843 4759 institutions@nature.com<br />

CUSTOMER SERVICE www.nature.com/help<br />

Senior Global Customer Service Manager: Gerald Coppin<br />

For all print and online assistance, please visit www.nature.com/help<br />

Purchase subscriptions:<br />

Americas: <strong>Nature</strong> <strong>Immunology</strong>, Subscription Dept., 342 Broadway, PMB 301, New York, NY<br />

10013-3910. Tel: (866) 363 7860, Fax: (212) 689 9108<br />

Europe/ROW: <strong>Nature</strong> <strong>Immunology</strong>, Subscription Dept., Macmillan Magazines Ltd., Brunel<br />

Road, Houndmills, Basingstoke, RG21 6XS, United Kingdom. Tel: 44 1256 329 242, Fax: 44<br />

1256 812 358<br />

Japan: <strong>Nature</strong> <strong>Immunology</strong>, NPG <strong>Nature</strong> Asia-Pacific, Chiyoda Building, 2-37<br />

Ichigayatamachi, Shinjuku-ku, Tokyo 162-0843. Tel: 81 3 3267 8751, Fax: 81 3 3267 8746<br />

India: <strong>Nature</strong> <strong>Immunology</strong>, NPG India, 3A, 4th Floor, DLF Corporate Park, Gurgaon 122002,<br />

India. Tel: 91 124 2881054/55, Fax: 91 124 2881052<br />

REPRINTS reprint@natureny.com<br />

<strong>Nature</strong> <strong>Immunology</strong>, Reprint Department, <strong>Nature</strong> Publishing Group, 75 Varick Street, Fl 9,<br />

New York, NY 10013-1917, USA.<br />

For commercial reprint orders of 600 or more, please contact:<br />

UK Reprints: Tel: 44 1256 302 923, Fax: 44 1256 321 531<br />

US Reprints: Tel: (212) 726 9278, Fax: (212) 679 0843


experimental autoimmune<br />

encephalomyelitis requires the entry of<br />

disease-inducing T cells into the brain.<br />

reboldi and colleagues (p 514; see also<br />

news and views by Steinman,<br />

p 453) find that T H -17 cells initiate<br />

this disease by entering the brain<br />

through the choroid plexus. The original<br />

image shows human brain tissue in<br />

which choroid plexus epithelial cells<br />

are stained with antibody to CCl20<br />

(fuchsia) and astrocytes are ‘decorated’<br />

by antibody to glial fibrillary acidic<br />

protein (brown). original image by<br />

andrew elston (lifeSpan bioSciences).<br />

artwork by lewis long.<br />

brahms and inteferon (p 447)<br />

Editorial<br />

445 The final push?<br />

Essay<br />

447 Aimez-vous Brahms? A story capriccioso from the discovery of a cytokine family<br />

and its regulators<br />

tadatsugu taniguchi<br />

nEws and viEws<br />

451 Local advantage: skin DCs prime; skin memory T cells protect<br />

akiko iwasaki see also pp 488 & 524<br />

453 Gaining entry to an uninflamed brain<br />

robert C axtell & lawrence steinman see also p 514<br />

455 Crohn’s disease–associated Nod2 mutants reduce IL10 transcription<br />

dana J Philpott & stephen E Girardin see also p 471<br />

457 The Foxo and the hound: chasing the in vivo regulation of T cell populations<br />

during infection<br />

Elia d tait & Christopher a Hunter see also p 504<br />

459 rEsEarCH HiGHliGHts<br />

rEviEw<br />

461 Autophagy genes in immunity<br />

Herbert w virgin & Beth levine<br />

artiClEs<br />

volume 10 number 5 may 2009<br />

471 A Crohn’s disease–associated NOD2 mutation suppresses transcription of human<br />

IL10 by inhibiting activity of the nuclear ribonucleoprotein hnRNP-A1<br />

Eiichiro noguchi, yoichiro Homma, Xiaoyan Kang, Mihai G netea & Xiaojing Ma<br />

see also p 455<br />

480 Autophagy enhances the presentation of endogenous viral antigens on MHC class I<br />

molecules during HSV-1 infection<br />

luc English, Magali Chemali, Johanne duron, Christiane rondeau, annie laplante,<br />

diane Gingras, diane alexander, david leib, Christopher norbury, roger lippé &<br />

Michel desjardins<br />

<strong>Nature</strong> <strong>Immunology</strong> (issn 1529-2908) is published monthly by nature Publishing Group, a trading name of nature america inc. located at 75 varick<br />

street, Fl 9, new york, ny 10013-1917. Periodicals postage paid at new york, ny and additional mailing post offices. Editorial Office: 75 varick street,<br />

Fl 9, new york, ny 10013-1917. tel: (212) 726 9207, Fax: (212) 696 9752. Annual subscription rates: Usa/Canada: Us$225 (personal), Us$3,060<br />

(institution). Canada add 7% Gst #104911595rt001; Euro-zone: €287 (personal), €2,430 (institution); rest of world (excluding China, Japan, Korea):<br />

£185 (personal), £1,570 (institution); Japan: Contact nPG nature asia-Pacific, Chiyoda Building, 2-37 ichigayatamachi, shinjuku-ku, tokyo 162-0843.<br />

tel: 81 (03) 3267 8751, Fax: 81 (03) 3267 8746. POSTMASTER: send address changes to <strong>Nature</strong> <strong>Immunology</strong>, subscriptions department, 342 Broadway,<br />

PMB 301, new york, ny 10013-3910. Authorization to photocopy material for internal or personal use, or internal or personal use of specific clients,<br />

is granted by nature Publishing Group to libraries and others registered with the Copyright Clearance Center (CCC) transactional reporting service,<br />

provided the relevant copyright fee is paid direct to CCC, 222 rosewood drive, danvers, Ma 01923, Usa. identification code for <strong>Nature</strong> <strong>Immunology</strong>:<br />

1529-2908/04. Back issues: Us$45, Canada add 7% for Gst. CPC PUB aGrEEMEnt #40032744. Printed on acid-free paper by dartmouth Journal<br />

services, Hanover, nH, Usa. Copyright © 2009 nature Publishing Group. Printed in Usa.<br />

i


nature immunology<br />

Getting aID<br />

(p 540)<br />

The Foxo factor<br />

(pp 457 and 504)<br />

Skin dendritic cells<br />

(pp 451 and 488)<br />

488 Cross-presentation of viral and self antigens by skin-derived CD103 + dendritic<br />

cells<br />

sammy Bedoui, Paul G whitney, Jason waithman, liv Eidsmo, linda wakim,<br />

irina Caminschi, rhys s allan, Magdalena wojtasiak, Ken shortman,<br />

Francis r Carbone, andrew G Brooks & william r Heath see also pp 451 & 524<br />

496 Divergent functions for airway epithelial matrix metalloproteinase 7 and retinoic<br />

acid in experimental asthma<br />

sangeeta Goswami, Pornpimon angkasekwinai, Ming shan, Kendra J Greenlee,<br />

wade t Barranco, sumanth Polikepahad, alexander seryshev, li-zhen song,<br />

david redding, Bhupinder singh, sanjiv sur, Prescott woodruff, Chen dong,<br />

david B Corry & Farrah Kheradmand<br />

504 Transcription factor Foxo3 controls the magnitude of T cell immune responses by<br />

modulating the function of dendritic cells<br />

anne s dejean, daniel r Beisner, irene l Ch’en, yann M Kerdiles, anna Babour,<br />

Karen C arden, diego H Castrillon, ronald a dePinho & stephen M Hedrick<br />

see also p 457<br />

514 C-C chemokine receptor 6–regulated entry of T H-17 cells into the CNS through<br />

the choroid plexus is required for the initiation of EAE<br />

andrea reboldi, Caroline Coisne, dirk Baumjohann, Federica Benvenuto,<br />

denise Bottinelli, sergio lira, antonio Uccelli, antonio lanzavecchia,<br />

Britta Engelhardt & Federica sallusto see also p 453<br />

524 Memory T cells in nonlymphoid tissue that provide enhanced local immunity<br />

during infection with herpes simplex virus<br />

thomas Gebhardt, linda M wakim, liv Eidsmo, Patrick C reading, william r Heath &<br />

Francis r Carbone see also pp 451 & 488<br />

531 T cell antigen receptor signaling and immunological synapse stability require<br />

myosin IIA<br />

tal ilani, Gaia vasiliver-shamis, santosh vardhana, anthony Bretscher &<br />

Michael l dustin<br />

540 HoxC4 binds to the promoter of the cytidine deaminase AID gene to induce AID<br />

expression, class-switch DNA recombination and somatic hypermutation<br />

seok-rae Park, Hong Zan, Zsuzsanna Pal, Jinsong Zhang, ahmed al-Qahtani,<br />

Egest J Pone, Zhenming Xu, thach Mai & Paolo Casali<br />

551 CorriGEnda<br />

natUrE iMMUnoloGy ClassiFiEd<br />

See back pages.<br />

volume 10 number 5 may 2009<br />

iii


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

The final push?<br />

Over 20 years ago, the Global Polio Eradication Initiative was launched. Today, polio is still endemic in four countries.<br />

With great fanfare, the World Health Assembly launched the<br />

Global Polio Eradication Initiative (GPEI) just over 20 years<br />

ago. In what was described as a “magnificent gift from the 20th<br />

century to future generations of children,” public health officials and<br />

volunteers committed themselves to ridding the planet of polio by the<br />

year 2000. Unfortunately, there are still new cases of infection in 2009.<br />

Although worldwide cases of polio have fallen from 350,000 in 1988 to<br />

under 2,000 in 2008, some experts question whether eradication will<br />

ever be possible. Many elements are conspiring against the efforts to<br />

eliminate disease, including the present political and economic crisis.<br />

By 2001, it almost seemed that the battle against polio had been won—<br />

worldwide cases fell to an all-time low of 791. Poliomyelitis type II had<br />

been completely eradicated by 1999. However, since 2003 the initiative<br />

has suffered several setbacks. In Nigeria, rumors began circulating that<br />

the polio vaccine contained the AIDS virus and was part of a Western plot<br />

to sterilize Muslim girls. The vaccine was described as “tainted by evildoers<br />

from America” and was “America’s revenge for September 11th.”<br />

As a result, the Nigerian government halted the vaccination campaign<br />

until the vaccine could be proven safe. As a consequence, 18 formerly<br />

polio-free countries suffered outbreaks traceable to Nigeria. A few countries,<br />

such as Sudan, are still struggling to regain their polio-free status.<br />

At present, polio is endemic in four countries: Nigeria, Afghanistan,<br />

Pakistan and India. In 2007, an intensified eradication effort was<br />

launched by the GPEI to circumvent the remaining technical, financial<br />

and operational barriers that were preventing polio eradication in<br />

these countries (http://www.polioeradication.org/content/publications/<br />

PolioStrategicPlan09-13_Framework.pdf). By mid-2008, two independent<br />

World Health Organization (WHO) advisory bodies concluded<br />

that the intensified efforts could overcome these challenges, and consequently<br />

a new Strategic Plan 2009–2013 has been endorsed. This 5-year<br />

plan combines proven eradication strategies with recently developed<br />

tools and tactics, including improving the efficacy of the vaccine. But<br />

even with a new strategic plan in place, is it realistic to expect a poliofree<br />

world by 2013?<br />

The GPEI will need to overcome many obstacles, not least the political<br />

instability of many regions in the polio-endemic countries. Afghanistan<br />

and Pakistan will probably pose the greatest challenge. Because of<br />

successful vaccine drives, no cases of polio have been reported in the<br />

relatively peaceful northern provinces of Afghanistan, but the conflict–<br />

ridden south and southeastern provinces still suffer recurrent outbreaks.<br />

Access of aid workers to vulnerable communities in these areas has been<br />

increasingly limited. Many aid workers have been killed, abducted or<br />

threatened by criminal gangs and Taliban insurgents in recent years,<br />

according to the Afghanistan NGO Safety Office. Only a few months<br />

ago, a day of tranquility organized by the United Nations Assistance<br />

Mission in Afghanistan to enable immunizations in remote areas was<br />

ediTorial<br />

cancelled by the WHO after two Afghan doctors were killed by a suicide<br />

car bombing caused by the Taliban.<br />

Pakistan has also become increasingly unstable. The Taliban and<br />

Al-Qaeda resurgence in the North-West Frontier Province of Pakistan<br />

is allowing the virus to cross borders between the two countries unimpeded.<br />

Cases of polio have exploded in Pakistan, causing outbreaks<br />

in previously polio-free areas. In 2007, the cleric Mufti Khalid Shah<br />

declared a fatwa on employees of the United Nations, WHO and all<br />

other foreign organizations in 2007, and aid workers in Bannu were<br />

sent a 500-rupee note with a letter stating they could either stop or buy<br />

their own coffin.<br />

Cultural differences also present problems for the vaccine program<br />

in these regions. Finding sufficient numbers of women to join the vaccination<br />

team has been difficult. However, the participation of women<br />

is essential, as by tradition only women can enter a Muslim household<br />

when the husband is away, and women with children are usually better<br />

at persuading other mothers to vaccinate. The GPEI must also persuade<br />

people that the vaccine is safe, improve poor sanitation that can interfere<br />

with uptake of the oral vaccine, and offset fatigue that can set in among<br />

volunteers, donors and the general populace. It only takes one unvaccinated<br />

child to trigger a new epidemic.<br />

Fortunately, fatigue has not affected the organizers of the eradication<br />

movement. In January 2009, the Bill & Melinda Gates Foundation,<br />

Rotary International and the governments of the UK and Germany<br />

pledged US$630 million over 5 years for a massive final push to rid<br />

humanity of this scourge. But this may not be enough. Heidemarie<br />

Wieczorek-Zeul, the German Federal Minister for Economic Cooperation<br />

and Development, estimated that even with the new injection of funds,<br />

the global initiative is still some US$340 million short of its budget for<br />

2009–2010. Given that funding agencies and charities, including the<br />

Gates Foundation, are having to prune grant growth because of the recession,<br />

these are uncertain times for the eradication program.<br />

Nevertheless, there are many reasons to be optimistic. The government<br />

in Nigeria is now firmly behind the GPEI, and Muslim clerics who<br />

initially shunned the vaccine program are now actively campaigning for<br />

acceptance of the polio vaccine. In March, President Barack Obama<br />

announced the mobilization of more troops in Afghanistan to “disrupt,<br />

dismantle and defeat” the terrorist Al-Qaeda network in Afghanistan<br />

and neighboring Pakistan. Although the final push for eradication will<br />

be costly and difficult, it certainly will not compare with the costs the<br />

world would face if polio were allowed to reemerge. Since 1988, more<br />

than 2 million children have been immunized, and it is estimated that<br />

as a result, 5 million fewer people have been paralyzed by polio. It is<br />

a testament to the success of the GPEI that worldwide cases of polio<br />

have fallen by over 99%. Now is not the time to give up, even when the<br />

going seems to be getting tough.<br />

nature immunology volume 10 number 5 may 2009 445


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Aimez-vous Brahms? A story capriccioso<br />

from the discovery of a cytokine family and<br />

its regulators<br />

Tadatsugu Taniguchi<br />

Do you delight in Brahms? Do you delight in immunology? Tada Taniguchi recounts the story of Type 1 interferon and<br />

its downstream regulators.<br />

“The Master said ‘Those who know it are not<br />

comparable to those who love it; those who love<br />

it are not compatible to those who delight in<br />

it.’” —Analects of Confucius<br />

Music aficionados may recognize the title of<br />

this article as a phrase originating from<br />

Françoise Sagan’s romantic novel of the same<br />

name (published in English as Goodbye Again).<br />

A typical laboratory scene on a Sunday afternoon<br />

in the 1970s was one in which I would<br />

whistle the first movement of the Brahms second<br />

piano concerto and then the next portion<br />

of the melody would be whistled back from<br />

the nearby office of my graduate school mentor<br />

Charles Weissmann. In this environment, I<br />

began the scientific achievements that would<br />

shape my career. To the credit of Massimo<br />

Libonati, my mentor from two prior years<br />

in Napoli, and my very rudimentary Italian,<br />

I had made my debut at a scientific meeting<br />

held in Roma in 1972 (ref. 1) before meeting<br />

Charles, my subsequent friend and mentor. I<br />

had left Tokyo for Naples without even a master’s<br />

degree, partly because I was appassionato<br />

about classic music. Indeed, it was through my<br />

desire to do science in Italy that I was able to<br />

derive tremendous enjoyment from my personal<br />

meetings with many artists, including the<br />

tenor Mario del Monaco and a retired German<br />

prima donna who lived in a friend’s villa in<br />

Tadatsugu Taniguchi is in the Department of<br />

<strong>Immunology</strong>, Graduate School of Medicine and<br />

Faculty of Medicine, at Tokyo University, Tokyo,<br />

Japan.<br />

e-mail: tada@m.u-tokyo.ac.jp<br />

Capri and recounted to me many episodes<br />

of her—sometimes romantic—experiences<br />

with the conductors Bruno Walter, Wilhelm<br />

Furtwängler, Herbert von Karajan and others.<br />

In Charles’ lab at the Institut für<br />

Molekularbiologie I der Universität Zürich,<br />

the project that was to become my PhD thesis<br />

involved site-directed mutagenesis of the RNA<br />

phage Qβ, a technique invented by Charles and<br />

one of my brothers in science, Richard Flavell 2 ,<br />

thereby spawning the concept of ‘reverse<br />

genetics’ that was only later applied to DNA.<br />

Although it was not very easy to catch up with<br />

the other members of the lab—who came to<br />

Zurich from top institutes in the world—in<br />

terms of either language or science, their<br />

exceptional kindness helped me to adjust in a<br />

relatively short time. I was fortunate enough<br />

to publish several papers on Qβ, including a<br />

description of the use of recombinant DNA<br />

technology to generate a plasmid encoding<br />

the genome of this virus as a means to produce<br />

an infectious RNA virus 3 . On weekends,<br />

I was invited to Charles’s home to write papers:<br />

as I recall well, writing was accompanied by<br />

music, often the chamber music of Brahms<br />

and Beethoven but sometimes an opera such<br />

as Berlioz’s Les Troyens which, as many readers<br />

may have experienced, was quite beneficial to<br />

scientific writing.<br />

It was in 1978 that the hitherto unheard-of<br />

word “interferon” (IFN) came to my ears<br />

at a lecture given by Peter Lengyel of Yale<br />

University, a good friend of Charles and a<br />

pioneer of IFN research. It was a stormy experience,<br />

and it made me very curious about the<br />

phenomenon by which this molecule with<br />

Silhouette of Brahms.<br />

ESSAy<br />

antiviral activity is produced in mammalian<br />

cells infected by viruses; in my mind, at the<br />

time, I envisioned IFN as a ‘Sleeping Beauty’<br />

gene(s) awakened upon ‘un bacio’ (a kiss) of<br />

a virus. By this point, I had finished my PhD,<br />

and I was about to return to Tokyo for a new<br />

position when Charles asked me to begin<br />

work on the initial isolation of human leukocyte<br />

IFN (afterward renamed IFN-α) mRNA<br />

and the development of strategies for its eventual<br />

gene cloning, in collaboration with Kari<br />

Cantell in Helsinki.<br />

My curiosity about the nature of IFN was<br />

strongly sostenuto after my return to the Cancer<br />

Institute in Tokyo at the end of 1978, and so I<br />

decided to work on the fibroblast IFN (afterward<br />

renamed IFN-β, or IFNB) gene, which<br />

was presumed to be distinct from leukocyte<br />

IFN 4 , mainly in order to avoid unnecessary<br />

overlap and competition with the project in<br />

Charles’s lab. As it turned out, this decision led<br />

to the identification of IFN-α and IFN-β by<br />

Charles’s and my group, respectively, resulting<br />

in the first recognition of a cytokine gene fam-<br />

nature immunology volume 10 number 5 may 2009 447


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

eSSay<br />

Tada Taniguchi, his mentor Charles Weissmann<br />

and colleagues in Zurich (1978). Charles (second<br />

from right) is wearing a uniform because he had<br />

just returned from military service.<br />

ily. My initial inclination was to search around<br />

for someone working on fibroblast IFN; the<br />

gene was extremely ricercato by numerous<br />

pharmaceutical companies, and there were<br />

many aspirations for its clinical use, so I was<br />

not very optimistic about getting support<br />

from them because of their desire to protect<br />

trade secrets and potential patents. Although<br />

IFNs and other soluble mediators of immune<br />

responses, later collectively termed ‘cytokines’,<br />

had already garnered considerable attention<br />

in the biological sciences and medicine at that<br />

time, the structure and function of these molecules<br />

remained elusive, along with the underlying<br />

mechanisms of signal transmission and<br />

the regulation of their expression. Indeed,<br />

addressing these issues was hampered by the<br />

fact that these molecules have pleiotropic, or<br />

multiple, biological activities, are usually produced<br />

simultaneously and are expressed only<br />

weakly by a variety of different cell types, making<br />

it difficult to obtain a single cytokine in<br />

pure form.<br />

My initial efforts to obtain a consistent<br />

source of starting material were in vain until<br />

I read a book about IFN, written in Japanese<br />

with great scientific enthusiasm and dedication<br />

by Shigeyasu Kobayashi of Toray, Inc. I<br />

visited Dr. Kobayashi in Toray’s Basic Research<br />

Institute and explained my enthusiasm for<br />

elucidating the structure of fibroblast IFN and<br />

the prospect that such work could lead to the<br />

discovery of the gene-switching mechanism in<br />

mammalian cells, an idea that was beginning<br />

to be addressed worldwide. He listened to me<br />

attentively, and then promptly told me that he<br />

would be willing to provide large amounts of<br />

his fibroblast cell line for the gene cloning. I was<br />

honestly surprised not only by the speed with<br />

which he made his decision but also because he<br />

placed no conditions on my receiving his help. I<br />

then realized that this was a kind of gentlemen’s<br />

agreement, although nowadays such an agreement<br />

is difficult to imagine: perhaps it was our<br />

shared scientific enthusiasm for the discovery<br />

of IFN, which was almost totally unknown at<br />

that time, that helped forge our collaboration.<br />

After all, “Curiosity has its own reason for<br />

existence,” as Albert Einstein (a fellow Zurich<br />

alumnus) once wrote.<br />

There was another big obstacle to my work:<br />

human gene cloning had not yet been done in<br />

Japan and the regulation of recombinant DNA<br />

technology was so strict that all work had to<br />

be done in a P3 facility using only the safest<br />

bacteria, that is, those that are the most difficult<br />

to grow (for example, Escherichia coli strain<br />

χ1776). Fortunately for me, the Institute had<br />

such a facility almost ready to use, though interestingly<br />

I was required, before the start of my<br />

experiments, to spray Neurospora crassa spores<br />

into the cabinet and monitor their capture so<br />

as to demonstrate the facility’s suitability for<br />

P3-level experiments. Finally, with the strong<br />

support and leadership of the Institute’s magnanimous<br />

director Haruo Sugano, I was soon<br />

allowed to proceed with the gene cloning.<br />

I was aware at the time that the competition<br />

for this project worldwide was very strong.<br />

Moreover, because I had to work solo, I thought<br />

it necessary to devise a strategy that would minimize<br />

my time and effort as much as possible.<br />

The most notable difficulty at that time was<br />

to identify cDNA clones for mRNAs that were<br />

very weakly expressed, of which IFN cDNA<br />

was a typical example. As a result, I came upon<br />

a two-step approach by which I would first<br />

search for all suspected clones before proceeding<br />

to identify the desired clone among them.<br />

I first prepared ‘hot’ ( 32 P-labeled) cDNA from<br />

mRNA isolated from poly(I:C)-stimulated<br />

fibroblast cells, and then hybridized the cDNA<br />

with mRNA from unstimulated cells. After<br />

separation of unhybridized cDNA (probe A)<br />

from hybridized cDNA (probe B), I subjected<br />

the unhybridized cDNA to Grunstein-Hogness<br />

in situ colony hybridization with two identical<br />

sets of filters harboring ~3,600 χ1776 colonies.<br />

I then assayed the four clones that preferentially<br />

hybridized to probe A by a ‘hybridizationtranslation<br />

assay’ to search more rigorously for<br />

the cDNA clone that contained the IFN mRNA<br />

sequence. This strategy worked well for me,<br />

and by the end of the summer of 1979, I was<br />

done 5 . Although I can no longer recall this, my<br />

wife Yoko says that I seldom came back home<br />

before midnight, even on holidays.<br />

I sequenced the cDNA (again solo) by the<br />

Maxam-Gilbert method, though the information<br />

I could obtain from one piece of endlabeled<br />

DNA was very limited at that time.<br />

Finally, after a few months, I elucidated the<br />

coding sequence of fibroblast IFN; it was on 2<br />

February 1980, while I was listening to one of<br />

the Beethoven’s Rasumovsky quartets at home<br />

with my son, who celebrated his second birth-<br />

day that day. In the interim, Kathy Zoon and<br />

Ernest Knight, together with Mike Hunkapiller,<br />

Lee Hood and colleagues, had determined the<br />

N-terminal sequences of human IFN-α and<br />

IFN-β, respectively, and Shigekazu Nagata,<br />

another brother in science in Charles’s lab,<br />

and his colleagues had cloned and expressed<br />

a human IFN-α (IFNA) cDNA 6–8 .<br />

I then received two international phone calls,<br />

one from Charles and the other from Mark<br />

Ptashne, already one of the most renowned<br />

molecular biologists around, who was then<br />

working at Harvard; Mark is also a renowned<br />

violinist and so might perhaps be reclassified<br />

as a ‘molecular violinist’. Charles and I talked<br />

about the structures of IFN-α and IFN-β, and<br />

we eventually published papers back to back<br />

on their sequences 9,10 . Naturally, we were very<br />

interested in comparing their sequences. Mark,<br />

meanwhile, was very keen to exploit the new<br />

technique developed by Lenny Guarente and<br />

Tom Roberts of expressing cDNA in E. coli and<br />

so invited me to Harvard so that we could work<br />

together. At the end of February, Yoko and I<br />

left Tokyo for Zurich with a final destination of<br />

Boston. Charles and I, sitting together again in<br />

his room where we had written so many papers,<br />

began the comparison between our IFN-α and<br />

IFN-β sequences by hand. I remember that it<br />

was he who noticed the first portion of the<br />

sequence similarities, which got us very excited,<br />

though only later would we become aware that<br />

this was just the beginning of the discovery of<br />

many cytokine families.<br />

Charles and I then decided to write a paper<br />

together. He kindly said something to the effect<br />

of, “Well, Tada, you are still young and need further<br />

development.” So, after carefully discussing<br />

the issue further with his colleagues Marco<br />

Schwarzstein and Ned Mantei, who sequenced<br />

the IFN-α cDNA, Charles generously gave me<br />

the first authorship. Our paper was eventually<br />

published as an Article in <strong>Nature</strong> back to back<br />

with the paper by Rik Derynck, Walter Fiers<br />

and colleagues, who also cloned IFN-β 11,12 .<br />

This was for us a truly singular experience<br />

with <strong>Nature</strong> in that we originally submitted the<br />

manuscript as a Letter, which the editors then<br />

converted into the longer Article format. Yoko<br />

and I then continued on to Boston, where we<br />

received a warm welcome from Mark and his<br />

colleagues. Thanks to the kind help provided<br />

by Jan Vilcek (IFN assay), Alice Wong (virus),<br />

Vicki Sato (human cells) and many others, the<br />

work was done in a relatively short period of<br />

time 13 . During our sojourn, we were given<br />

many opportunities to interact socially with<br />

many scientists and artists, including Wally<br />

Gilbert, Lew Cantley, Hidde Ploegh, (the late)<br />

pianist Patricia Zander and cellist Yo-Yo Ma.<br />

Of the many concerts we enjoyed, perhaps the<br />

448 volume 10 number 5 may 2009 nature immunology


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

most unforgettable was Sergey Rachmaninov’s<br />

second sonata for piano played by his friend<br />

Vladimir Horowitz at Boston Symphony Hall.<br />

Overall, my friendships and experiences from<br />

this time have endured ever since then and are<br />

a great treasure of my life.<br />

After returning to Tokyo, I started working<br />

on another class of cytokines, now known as<br />

the interleukins (ILs), and specifically on T<br />

cell growth factor 14 (since renamed IL-2), with<br />

Junji Hamuro, Hiroshi Matsui and colleagues<br />

at Ajinomoto Inc. IL-2 was also the subject<br />

of considerable hopes for its use in basic and<br />

clinical immunology, and we succeeded in<br />

elucidating the structure of human IL-2 and<br />

producing recombinant IL-2 in 1982 (ref. 15).<br />

It is remarkable that, since then, more than<br />

30 ILs have been cloned and characterized. In<br />

the interim, my colleague Shigeo Ohno joined<br />

my lab and discovered that the 5′ region of the<br />

human IFN-β gene confers virus infection–<br />

dependent activation of a reporter gene—the<br />

inception of the virus-inducible promoter<br />

analysis 16 . In parallel, both the IFN-α and<br />

IFN-β promoters were subsequently analyzed<br />

by many investigators, most notably by Charles<br />

and by Tom Maniatis and colleagues (reviewed<br />

in ref. 17). Indeed, the study of the IFN-β gene<br />

has led to the concept of the ‘enhanceosome’<br />

developed by Tom and his colleagues, which<br />

has become a model for understanding the<br />

mechanisms whereby genes are turned on and<br />

off in mammalian cells 18 .<br />

After I moved my lab and family to Osaka,<br />

another talented colleague, Takashi Fujita,<br />

joined my lab and identified a minimal DNA<br />

consensus sequence that functions as virusinducible<br />

enhancer 19 . Then, working together<br />

with Takashi, a master’s degree graduate student,<br />

Masaaki Miyamoto used an expression-<br />

cloning strategy to screen a cDNA library for<br />

gene products that would bind to Takashi’s<br />

consensus sequence. For several reasons, we<br />

originally proposed to name the newly discovered<br />

factor ‘cytokine regulatory factor-1’, but<br />

Ben Lewin, then editor of Cell, rejected our<br />

proposal. So the gene was instead named ‘interferon<br />

regulatory factor-1’ or IRF1 (ref. 20). Soon<br />

after, another master’s degree student, Hisashi<br />

Harada, discovered a related gene, termed IRF2,<br />

which officially signified the recognition of an<br />

IRF family. The family has since been extended<br />

to nine members in mammals by a number of<br />

other investigators (reviewed in ref. 21). In retrospect,<br />

the term ‘CRF’ gene family might indeed<br />

have been more accurate in the light of what we<br />

know now of this family’s role, but c’est la vie.<br />

The roles of IRFs have since been extensively<br />

studied by a number of groups worldwide.<br />

Indeed, interest in IRFs has only grown, particularly<br />

with the recent discovery of patternrecognition<br />

receptors whose activation evokes<br />

type I IFN and other innate immune responses,<br />

such as Toll-like receptors (pioneered by Jules<br />

Hoffman’s group working on Drosophila Toll<br />

and Ruslan Medzhitov, Charles Janeway and<br />

Bruce Beutler on TLR4 in mammals) and cytosolic<br />

receptors for nucleic acids (by Mitsutoshi<br />

Yoneyama and Takashi Fujita on RIG-I and<br />

MDA5, by my laboratory on DAI/DLM-1/ZBP1<br />

and recently by several groups on AIM2). We<br />

now know that IRF1, IRF3 and IRF7 (and possibly<br />

also IRF5) play critical roles in the regulation<br />

of type I IFN (IFN-α and IFN-β) genes,<br />

wherein the contribution of each depends<br />

on the nature of the stimuli and the cell type<br />

(reviewed in ref. 22). The role of IRFs in other<br />

biological responses, for example, in the regulation<br />

of oncogenesis, has also received much<br />

attention: accordingly, the numbers of articles<br />

describing IRFs are rapidly increasing. I am particularly<br />

grateful to Tak Mak, Shigeru Noguchi<br />

and Nobuaki Yoshida, who so kindly collaborated<br />

with us on the generation and analyses of<br />

a number of IRF knockout mice.<br />

Upon reflection, I find it quite remarkable<br />

that, since its independent discovery by Isaacs<br />

and Lindenman and by Nagano and Kojima<br />

more than 50 years ago, the type I IFN family<br />

of cytokines has become the prototype for<br />

cytokine research over the past three decades.<br />

Indeed, IFN research has led directly to the<br />

discovery of Janus-family (JAK) kinases,<br />

signal transducers and activators of transcription<br />

(STATs), the enhanceosome, IRFs<br />

and IFN-inducible genes; and, as key regulators<br />

of pathological and protective immune<br />

responses, IFNs are likely to lead to new<br />

research discoveries in the years to come.<br />

ACKNOWLEDGMENTS<br />

I thank J. Vilcek and D. Savitsky for their kind<br />

reading of this essay. My particular appreciation<br />

goes to my mentor C. Weissmann, from whom I<br />

learned so much about science and music (and<br />

good jokes), and to the colleagues who spent time<br />

in my laboratory through the many facets of the<br />

studies described above. I also thank my wife Yoko<br />

and my friends, either mentioned herein or not, for<br />

their continuous support. Our studies were mostly<br />

supported by Kakenhi, Grants-in-Aid for Scientific<br />

Research from the Ministry of Education, Culture,<br />

Sports, Science and Technology, Japan. Finally, I<br />

also thank baseball’s Hanshin Tigers for their longstanding<br />

roller coaster of hope and disappointment,<br />

which in addition to music has significantly<br />

enriched my life.<br />

eSSay<br />

1. Taniguchi, T., Libonati, M. & Leone, e. azione della<br />

subtilisina sulla ribonucleasi BS-1. Boll. Soc. Ital. Biol.<br />

Sper. XLVIII, 1115–1119 (1972).<br />

2. Flavell, R.a., Sabo, D.L., Bandle, e.F. & Weissmann,<br />

C. Site-directed mutagenesis: generation of an extracistronic<br />

mutation in bacteriophage Qβ RNa. J. Mol. Biol.<br />

89, 255–272 (1974).<br />

3. Taniguchi, T., Palmieri, M. & Weissmann, C.Q. β DNa–<br />

containing hybrid plasmids giving rise to Qβ phage<br />

formation in the bacterial host. <strong>Nature</strong> 274, 223–228<br />

(1978).<br />

4. Cavalieri, R.L., Havell, e.a., Vilcek, J. & Pestka, S.<br />

Synthesis of human interferon by Xenopus laevis oocytes:<br />

two structural genes for interferons in human cells. Proc.<br />

Natl. Acad. Sci. USA 74, 3287–3291 (1977).<br />

5. Taniguchi, T. et al. Construction and identification of a<br />

bacterial plasmid containing the human fibroblast interferon<br />

gene sequence. Proc. Jpn. Acad. 55B, 464–469<br />

(1979).<br />

6. Knight, e., Jr, Hunkapiller, M.W., Korant, B.D., Hardy,<br />

R.W. & Hood, L.e. Human fibroblast interferon: amino<br />

acid analysis and amino terminal amino acid sequence.<br />

Science 207, 525–526 (1980).<br />

7. Nagata, S. et al. Synthesis in E. coli of a polypeptide<br />

with human leukocyte interferon activity. <strong>Nature</strong> 284,<br />

316–320 (1980).<br />

8. Zoon, K.C. et al. amino terminal sequence of the major<br />

component of human lymphoblastoid interferon. Science<br />

207, 527–528 (1980).<br />

9. Mantei, N. et al. The nucleotide sequence of a cloned<br />

human leukocyte interferon cDNa. Gene 10, 1–10<br />

(1980).<br />

10. Taniguchi, T., Ohno, S., Fujii-Kuriyama, y. & Muramatsu,<br />

M. The nucleotide sequence of human fibroblast interferon<br />

cDNa. Gene 10, 11–15 (1980).<br />

11. Derynck, R. et al. Isolation and structure of a human<br />

fibroblast interferon gene. <strong>Nature</strong> 285, 542–547<br />

(1980).<br />

12. Taniguchi, T. et al. Human leukocyte and fibroblast interferons<br />

are structurally related. <strong>Nature</strong> 285, 547–549<br />

(1980).<br />

13. Taniguchi, T. et al. expression of the human fibroblast<br />

interferon gene in Escherichia coli. Proc. Natl. Acad. Sci.<br />

USA 77, 5230–5233 (1980).<br />

14. Morgan, D.a., Ruscetti, F.W. & Gallo, R. Selective in<br />

vitro growth of T lymphocytes from normal human bone<br />

marrows. Science 193, 1007–1008 (1976).<br />

15. Taniguchi, T. et al. Structure and expression of a cloned<br />

cDNa for human interleukin-2. <strong>Nature</strong> 302, 305–310<br />

(1983).<br />

16. Ohno, S. & Taniguchi, T. Inducer-responsive expression<br />

of the cloned human interferon β1 gene introduced into<br />

cultured mouse cells. Nucleic Acids Res. 10, 967–977<br />

(1982).<br />

17. Honda, K., Takaoka, a. & Taniguchi, T. Type I interferon<br />

gene induction by the interferon regulatory factor family<br />

of transcription factors. Immunity 25, 349–360<br />

(2006).<br />

18. Kim, T.K. & Maniatis, T. The mechanism of transcriptional<br />

synergy of an in vitro assembled interferon-β<br />

enhanceosome. Mol. Cell 1, 119–129 (1997).<br />

19. Fujita, T., Shibuya, H., Hotta, H., yamanishi, K. &<br />

Taniguchi, T. Interferon-β gene regulation: tandemly<br />

repeated sequences of a synthetic 6 bp oligomer function<br />

as a virus-inducible enhancer. Cell 49, 357–367<br />

(1987).<br />

20. Miyamoto, M. et al. Regulated expression of a gene<br />

encoding a nuclear factor, IRF-1, that specifically binds<br />

to IFN-β gene regulatory elements. pCell 54, 903–913<br />

(1988).<br />

21. Tamura, T., yanai, H., Savitsky, D. & Taniguchi, T. The<br />

IRF family transcription factors in immunity and oncogenesis.<br />

Annu. Rev. Immunol. 26, 535–584 (2008).<br />

22. Honda, K. & Taniguchi, T. IRFs: master regulators of<br />

signalling by Toll-like receptors and cytosolic patternrecognition<br />

receptors. Nat. Rev. Immunol. 6, 644–658<br />

(2006).<br />

nature immunology volume 10 number 5 may 2009 449


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Local advantage: skin DCs prime; skin memory T cells<br />

protect<br />

Akiko Iwasaki<br />

How the immune system responds to local infection and establishes protective immunity in susceptible tissues<br />

remains unclear. Two new studies show that local tissue-resident dendritic cells prime cytotoxic T lymphocyte<br />

responses and that memory cytotoxic T lymphocytes remain in the tissue to provide antiviral immunity.<br />

Cytotoxic T lymphocytes (CTLs) are key<br />

effector cells that provide protection from<br />

viral infection. CD8 + T cells bearing T cell antigen<br />

receptors specific for a given viral antigen<br />

are primed in the secondary lymphoid organs<br />

by dendritic cells (DCs). Exactly which subset<br />

of DCs primes CTLs in response to viral infection<br />

has been an area of intense debate. Many<br />

pathogens enter the host via a specific niche,<br />

most commonly the mucosal surfaces, where<br />

they establish local acute and chronic infection.<br />

Herpes virus family members represent<br />

a good example of such a pathogen. In particular,<br />

herpes simplex virus type 1 (HSV-1) enters<br />

the human host through the oral mucosa and<br />

establishes latency in the trigeminal ganglion.<br />

Reactivation of latent HSV-1 in the ganglion<br />

leads to anterograde transport of virus back<br />

to the skin, causing ‘cold sore’ lesions around<br />

the mouth. CTLs are important both in controlling<br />

HSV-1 reactivation1 and in limiting<br />

viral replication in the peripheral site of replication2<br />

. Given the restricted nature of HSV-1<br />

replication in the epithelial cells and latency<br />

in the innervating ganglia, which cells prime<br />

CD8 + T cells in the draining lymph node and<br />

how protection is afforded by memory T cells<br />

are important unresolved issues. In this issue<br />

of <strong>Nature</strong> <strong>Immunology</strong>, two papers show that<br />

local tissue-resident cells do both: Heath and<br />

colleagues demonstrate that CTL priming is<br />

accomplished mainly by langerin-positive<br />

CD103 + dermal DCs3 , whereas Carbone<br />

and colleagues report that memory CTLs in<br />

the skin remain in the tissue and maximize<br />

Akiko Iwasaki is in the Department of<br />

Immunobiology, Yale University School of Medicine,<br />

New Haven, Connecticut, USA.<br />

e-mail: akiko.iwasaki@yale.edu<br />

NewS AND vIewS<br />

protective immunity to subsequent challenge<br />

with HSV-1 (ref. 4).<br />

There are three DC subsets in the skin:<br />

Langerhans cells in the epidermis, and two<br />

subsets of dermal DCs, consisting of langerinnegative<br />

dermal DCs and the newly described<br />

langerin-positive CD103 + dermal DCs 5–7<br />

(Fig. 1). Traditionally, it has been thought<br />

that local tissue-resident DCs ingest microbial<br />

antigens by phagocytosis and migrate to the<br />

draining lymph node to prime naive T cells.<br />

However, that paradigm has been challenged<br />

by many studies showing that lymph noderesident<br />

DCs are the only cells that present<br />

antigens to T cells. As for the DCs involved in<br />

CTL priming, the importance of the lymph<br />

node–resident CD8α + DCs has become well<br />

accepted 8 . CD8α + DCs are the main antigenpresenting<br />

cells for CTLs that are generated<br />

after infection by HSV-1, influenza virus,<br />

vaccinia virus, lymphocytic choriomeningitis<br />

virus and Listeria monocytogenes. CD8α +<br />

DCs can exclusively prime CTL responses<br />

whether HSV-1 is injected by needle into the<br />

footpad, by the intravenous route or by dermal<br />

abrasion 9 (Fig. 1a). In their study presented<br />

here, Bedoui et al. make the intriguing<br />

observation that reactivating HSV-1 causes a<br />

second phase of infection of the entire skin<br />

dermatome innervated by the infected ganglion<br />

3 (Fig. 1b), which results in a second<br />

wave of antigen presentation in the lymph<br />

nodes draining the new site of viral replication.<br />

Reactivating HSV-1 replicates in the<br />

epithelium, notably in the absence of any<br />

artificial manipulation of the skin, allowing<br />

the authors to study the course of natural<br />

infection. Taking advantage of this system,<br />

they assess DC subsets for their ability to<br />

stimulate CTLs. They find that whereas cross-<br />

presentation of viral antigen during primary<br />

infection after scarification is mediated by the<br />

lymph node–resident CD8α + DCs 3,9 , antigen<br />

presentation after natural infection with HSV-1<br />

in the skin during recrudescence is handled<br />

almost exclusively by the CD103 + dermal DCs<br />

(Fig. 1b). These results are consistent with the<br />

fact that dermal DCs are the main antigen-<br />

presenting cells after natural infection of<br />

vaginal mucosa with HSV-2 (ref. 10) and are also<br />

supported by a study showing differences in the<br />

participation of migrant versus lymph node–<br />

resident DCs in CTL priming after natural<br />

mucosal infection versus skin abrasion with<br />

HSV-1, respectively 11 .<br />

Several intriguing questions arise from this<br />

study 3 . First, why are different DCs involved<br />

in CTL priming during the primary and secondary<br />

infection? Is it possible that scarification<br />

allows HSV-1 to be carried by the lymph,<br />

circumventing the requirement for presentation<br />

by migrant DCs? This is unlikely, as<br />

migrant DCs are still needed for the CD8α +<br />

DCs to prime CTL immunity after scarification<br />

9 , which indicates that even if direct entry<br />

of the virus into the lymph node does occur,<br />

it is insufficient for priming by CD8α + DCs.<br />

Because scarification causes considerable tissue<br />

damage, it is conceivable that the CD103 +<br />

dermal DC functions may be suppressed by<br />

tissue-derived factors, rendering them unable<br />

to prime CD8 + T cells. Langerin-positive dermal<br />

DCs are shown to be responsible for crosspresenting<br />

epidermally expressed self antigen 3<br />

and are required for contact-hypersensitivity<br />

responses 7 , which suggests that these cells are<br />

able to present a diverse set of antigens. In<br />

contrast, langerin-negative dermal DCs are<br />

the main antigen-presenting cell for CD4 +<br />

T cells after scarification-induced HSV-1<br />

nature immunology volume 10 number 5 may 2009 451


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

nEwS AnD ViEwS<br />

a Primary infection<br />

b Secondary infection<br />

c<br />

Scarification<br />

Basement<br />

membrane<br />

Spinal column<br />

Dorsal root<br />

ganglion<br />

HSV-1<br />

8<br />

4<br />

Skin<br />

Epidermis<br />

Dermis<br />

Draining lymph<br />

node<br />

infection and after natural reactivation<br />

(Fig. 1a,b). Thus, it will also be important to<br />

understand the cellular and molecular mechanisms<br />

by which CD103 + dermal DCs versus<br />

CD103 – dermal DCs prime CD8 + T and CD4 +<br />

T cells. Second, which wave of T cell priming<br />

results in the establishment of protective<br />

immunity? Are the effector and memory CTLs<br />

induced by the first and second waves of HSV-1<br />

infection quantitatively and qualitatively similar?<br />

The answers to these questions will not<br />

only be important for development of vaccines<br />

against HSV-1 but also provide key clues to the<br />

findings reported by Carbone et al. 4 .<br />

Once primed by DCs, the virus-specific<br />

CD8 + T cells differentiate into at least two<br />

types of memory cells: central memory T cells<br />

home to lymphoid organs and have limited<br />

effector functions, whereas effector memory<br />

T cells home to peripheral tissues and rapidly<br />

secrete cytokines 12 . In their study presented<br />

here, Carbone and colleagues propose the<br />

existence of another type of memory T cells:<br />

tissue-resident memory T cells 4 . The authors<br />

carry out an elegant set of transplantation<br />

studies to demonstrate that these cells reside<br />

both near the latently infected ganglia and<br />

in the skin near the primary site of HSV-1<br />

infection (Fig. 1c). Unlike central and effec-<br />

4<br />

Langerhans cells CD103 4<br />

– dDC Langerin-positive CD103 + dDC CD8α + lymph node DC CD4 + T cell 8 CD8 + T cell<br />

tor memory T cells, tissue-resident memory<br />

T cells do not readily enter circulation once<br />

they establish residency in a given tissue and<br />

can proliferate locally after secondary viral<br />

challenge. Notably, tissue-resident memory T<br />

cells enter and remain in the tissue even in the<br />

absence of virus and, presumably, viral antigens.<br />

This is demonstrated by the finding that<br />

scarification alone induces the recruitment<br />

of these cells to damaged tissue and that skin<br />

containing these cells, when transplanted into<br />

a naive recipient, retains these cells for over 3<br />

weeks even though it is separated from the<br />

latently infected ganglia. Most importantly,<br />

when secondary HSV-1 challenge is applied to<br />

previously infected flank (containing tissueresident<br />

memory T cells) or to the opposite<br />

flank (able to recruit only effector memory<br />

T cells), the tissue containing tissue-resident<br />

memory T cells has 1% as much virus as is<br />

present in the site in which only the effector<br />

memory T cells are newly recruited. However,<br />

effector memory T cells still provide some<br />

protection relative to that afforded by unimmunized<br />

control by diminishing viral load to<br />

1% as much as naive control. These data show<br />

that complete protection from a viral challenge<br />

requires not only systemic CTL memory<br />

but also that tissue-resident memory T cells<br />

be mobilized to the site of potential viral<br />

encounter before infection.<br />

Tissue-resident memory T cells could<br />

represent a distinct lineage of memory cells<br />

that arise from the effector CTL pool, or they<br />

may differentiate from effector memory T<br />

cells once they arrive in the infected and/or<br />

damaged tissue. Tissue-resident memory T<br />

cells have high expression of integrin α 1 β 1<br />

(VLA-1) and CD69 but are CD62L lo and<br />

CD122 lo . However, this expression pattern<br />

is also shared by effector memory T cells.<br />

If tissue-resident memory T cells are a distinct<br />

lineage of memory cells, what factors<br />

influence their development? Or if effector<br />

memory T cells do differentiate into<br />

tissue-resident memory T cells, this would<br />

indicate that the chemokines that recruit<br />

effector memory T cells are responsible for<br />

the eventual existence of these cells in a given<br />

tissue. The cues that are responsible for the<br />

conversion of effector memory T cells into<br />

tissue-resident memory T cells and how long<br />

this takes will be important areas of research.<br />

Another question that needs an answer is<br />

how tissue-resident memory T cells are<br />

retained in a particular tissue, given that viral<br />

antigen is not required. This is particularly<br />

relevant for autoimmune diseases in which<br />

452 volume 10 number 5 may 2009 nature immunology<br />

8<br />

4<br />

4<br />

T RM<br />

T RM<br />

T RM<br />

T RM<br />

Memory<br />

T RM T RM<br />

T EM<br />

T RM<br />

T CM<br />

Blood vessel<br />

Figure 1 The induction and execution of immune responses to HSV-1 are coordinated by local DCs and memory T cells. (a) Primary infection by HSV-1<br />

is initiated by mechanical scarification of the superficial layer of the epidermis, which allows the virus to enter and replicate locally. HSV-1 infects the<br />

innervating ganglion by retrograde transport from the nerve endings in the skin and establishes latency. Virus introduced by this route is taken up by local<br />

skin-resident DCs, which, after migrating, present antigens to CD4 + T cells. However, cross-priming of CD8 + T cells is done uniquely by the lymph node–<br />

resident CD8α + DCs. (b) Reactivation of latent virus in the ganglion results in anterograde migration of infectious virions to the skin and infection of epithelial<br />

cells throughout the dermatome innervated by the ganglion. After this natural route of reinfection, the viral antigens are cross-presented to CD8 + T cells by<br />

langerin-positive CD103 + dermal DCs, not CD8α + DCs. (c) CTLs induced by DCs differentiate into three kinds of memory cells: central memory T cells (T CM ),<br />

effector memory T cells (T EM ) and tissue-resident memory T cells (T RM ). Only tissue-resident memory T cells take up residency in the skin (at the previous<br />

site of virus infection) and near the latently infected ganglion. After tertiary infection by HSV-1, tissue-resident memory T cells provide bulk of the protection,<br />

although effector memory T cells can also be recruited to the site from systemic circulation.<br />

Kim Caesar


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

tissue-resident memory T cells may cause<br />

chronic tissue destruction.<br />

Those questions aside, the findings<br />

reported by Gebhardt et al. 4 have important<br />

implications for vaccine development.<br />

Most pathogens gain a foothold in the host<br />

through specific portals of entry. For example,<br />

HIV-1 enters through the genital and<br />

rectal mucosa, whereas Mycobacterium tuberculosis<br />

gains access through the respiratory<br />

mucosa. Thus, provision of effective protection<br />

requires mobilization of tissue-resident<br />

memory T cells to the appropriate mucosal<br />

surfaces. Once again, the mechanism of the<br />

recruitment and retention of these cells<br />

in a given tissue needs to be investigated.<br />

One unconventional proposal would be to<br />

‘scratch and save (a life)’: creating minor<br />

scarification to recruit these cells to the<br />

potential site of pathogen encounter after<br />

conventional parenteral immunization. Of<br />

course, not all surfaces are conducive to this<br />

type of manipulation. A more universal and<br />

powerful approach would be to immunize at<br />

the potential site of pathogen encounter. The<br />

development of safe mucosal vaccines able to<br />

establish tissue-resident memory T cells at<br />

the site of pathogen entry might be the way<br />

to prevent the transmission of deadly virus<br />

infection in humans.<br />

In conclusion, these studies by the groups<br />

of Heath 3 and Carbone 4 provide important<br />

insight into the priming and execution of<br />

antiviral immunity to a local viral infection<br />

and open new avenues of investigation<br />

and, possibly, therapeutic approaches. With<br />

the development of a new in vivo approach<br />

for temporally and selectively depleting the<br />

skin of langerin-positive dermal DCs 7 , future<br />

studies should identify the function of these<br />

cells in immune responses to a variety of<br />

antigens. In addition, elucidating the biol-<br />

Gaining entry to an uninflamed brain<br />

Robert C Axtell & Lawrence Steinman<br />

ogy of tissue-resident memory T cells will<br />

provide clues about the generation of tissuespecific<br />

memory that is needed for vaccine<br />

development and for immune intervention<br />

in autoimmune diseases.<br />

1. Divito, S., Cherpes, T.L. & Hendricks, R.L. Immunol.<br />

Res. 36, 119–126 (2006).<br />

2. Zhu, J. et al. J. Exp. Med. 204, 595–603 (2007).<br />

3. Bedoui, S. et al. Nat. Immunol. 10, 488–495<br />

(2009).<br />

4. Gebhardt, T. et al. Nat. Immunol. 10, 524–530<br />

(2009).<br />

5. Poulin, L.F. et al. J. Exp. Med. 204, 3119–3131<br />

(2007).<br />

6. Ginhoux, F. et al. J. Exp. Med. 204, 3133–3146<br />

(2007).<br />

7. Bursch, L.S. et al. J. Exp. Med. 204, 3147–3156<br />

(2007).<br />

8. Heath, w.R. et al. Immunol. Rev. 199, 9–26<br />

(2004).<br />

9. Allan, R.S. et al. Immunity 25, 153–162 (2006).<br />

10. Zhao, X. et al. J. Exp. Med. 197, 153–162 (2003).<br />

11. Lee, H.K. et al. J. Exp. Med. 206, 359–370<br />

(2009).<br />

12. Lanzavecchia, A. & Sallusto, F. Curr. Opin. Immunol.<br />

17, 326–332 (2005).<br />

Little is known about how pathogenic T cells gain access to the uninflamed brain in multiple sclerosis and<br />

experimental autoimmune encephalomyelitis. A new study reports that interleukin 17–producing T helper cells enter<br />

the uninflamed central nervous system through the choroid plexus by a CCR6-CCL20–dependent mechanism.<br />

The blood-brain barrier is a protective<br />

barricade that limits the entry of large<br />

molecules and cells such as erythrocytes,<br />

platelets and lymphoid cells into the central<br />

nervous system (CNS) in normal conditions.<br />

Although there are well known ‘windows’<br />

in this barrier, particularly adjacent to the<br />

hypothalamus, which allow cytokines such as<br />

interleukin 1 (IL-1), IL-6 and tumor necrosis<br />

factor to elicit fever1 , in general, the brain is<br />

very selective about allowing large molecules<br />

and cells to gain entry. Yet, how autoreactive<br />

T cells gain access to the uninflamed brain<br />

to initiate disease has remained unknown.<br />

In this issue of <strong>Nature</strong> <strong>Immunology</strong>, Reboldi<br />

and colleagues2 demonstrate that autoreactive<br />

IL-17-producing T helper cells (TH-17 cells) enter the CNS through a ‘chink in<br />

the armor’ of the blood-brain barrier. In a<br />

special location called the ‘choroid plexus’,<br />

these TH-17 cells initiate the inflammatory<br />

cascade, causing demyelination through an<br />

Robert C. Axtell and Lawrence Steinman are in the<br />

Department of Neurological Sciences and Neurology,<br />

Stanford University, Stanford, California, USA.<br />

e-mail: steinman@stanford.edu<br />

interaction between the chemokine CCL20<br />

and its receptor, CCR6.<br />

For years, the brain has been considered a<br />

site of ‘immune privilege’, a place that tends to<br />

exclude members, both cells and molecules, of<br />

the immune community. ‘Privilege’ has its price,<br />

and of course, the immune system is any case<br />

skilled at outwitting protective ‘covenants’. In<br />

multiple sclerosis, the blood-brain barrier is<br />

breached by a variety of immune cells, including<br />

macrophages, dendritic cells, B cells and autoreactive<br />

T cells 3 . T cells are thought to be the<br />

earliest sentinels that penetrate the blood-brain<br />

barrier, but data to support this idea are scant. So<br />

far, much has been reported on the cellular and<br />

molecular processes involved in the migration of<br />

lymphocytes to the brain through inflamed vessels.<br />

Research on homing through the inflamed<br />

blood-brain barrier has shown that the integrin<br />

α 4 β 1 is critical for this 3 . Such studies of the<br />

blood-brain barrier in inflammation have led<br />

to the development of the most potent drug so<br />

far approved for treatment relapsing remitting<br />

multiple sclerosis, natalizumab, a humanized<br />

monoclonal antibody specific for α 4 β 1 .<br />

The migration of T H-17 cells to the CNS has<br />

been linked to the induction of inflammation<br />

nEwS AnD ViEwS<br />

in multiple sclerosis and in the animal model<br />

of experimental autoimmune encephalomyelitis<br />

(EAE) 4,5 . It has been shown in humans<br />

that T H -17 cells ‘preferentially’ express CCR6<br />

(ref. 6). Reboldi and colleagues now confirm<br />

that mice have a similar expression pattern<br />

in which CCR6 expression is restricted to<br />

T H -17 cells and is not found on the surface<br />

of T helper type 1 (T H 1) or T H 2 cells 2 . Given<br />

those data, the authors next explore the contribution<br />

of this chemokine receptor to EAE.<br />

The authors find that Ccr6 –/– mice are completely<br />

resistant to EAE 2 . This effect is not due<br />

to a block in the differentiation of T H -17 or<br />

T H 1 cells in the lymph nodes after induction<br />

of EAE but is associated with a lower capacity<br />

of T H 1 and T H -17 cells to migrate to the CNS.<br />

EAE disease signs are restored in the Ccr6 –/–<br />

mice by the transfer of myelin-specific T cells<br />

from Ccr6 +/+ 2D2 mice. Notably, in this transfer<br />

system, most of the T cells in the spinal<br />

cords and brain at the peak of disease are not<br />

of the Ccr6 +/+ donor origin but are almost all<br />

from the Ccr6 –/– recipient. This finding indicates<br />

that the initial trigger of inflammation is<br />

caused by CCR6-dependent infiltration of the<br />

uninflamed CNS by autoreactive T H-17 cells<br />

nature immunology volume 10 number 5 may 2009 453


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

nEwS AnD ViEwS<br />

Subarachnoid<br />

space<br />

Dendritic<br />

cell<br />

CCR6 + T H -17<br />

cell<br />

that then causes a ‘second wave’ of infiltration<br />

by a CCR6-independent mechanism.<br />

The choroid plexus is a highly vascularized<br />

region in the brain where the cerebrospinal fluid<br />

is formed. It is also a chief site for the entry of<br />

lymphocytes into the CNS, where they perform<br />

their normal immune surveillance of the fluid<br />

that surrounds the brain in the subarachnoid<br />

space 7 . Furthermore, there is speculation that<br />

the choroid plexus is an area in which autoreactive<br />

lymphocytes gain access to the CNS to cause<br />

multiple sclerosis 8 . Notably, the authors establish<br />

that CCL20, one of the ligands for CCR6, has<br />

higher expression in the choroid plexus than in<br />

other regions of the brain in both healthy mice<br />

and mice with EAE 2 . Additionally, they find that<br />

Ccr6 –/– cells of the immune system are unable to<br />

pass through the epithelial barrier of the choroid<br />

plexus to gain access to the CNS during EAE.<br />

These elegant mouse experiments are successfully<br />

‘translated’ to the human arena.<br />

The investigators find that this mechanism<br />

may actually occur during the early stages of<br />

multiple sclerosis. In patients experiencing<br />

their first neurological episode, memory T<br />

cells from the cerebrospinal fluid have much<br />

higher expression of CCR6 on their surface<br />

than do cells from peripheral blood 2 . In<br />

addition, CCL20 protein is present in greater<br />

abundance in the choroid plexus in brains of<br />

both healthy subjects and patients with multiple<br />

sclerosis than in the parenchyma of the<br />

brain. Such ‘translation’ of experiments in the<br />

EAE model to the human disease of multiple<br />

Choroid plexus<br />

Venule<br />

CCL20<br />

Choroid plexus<br />

epithelial cell<br />

Ventricular<br />

epithelial cell<br />

Figure 1 CCR6-dependent entry into the CnS. Epithelial cells of the uninflamed choroid plexus<br />

constitutively express CCL20, which attracts T H -17 cells and facilitates the crossing of T cells into<br />

the subarachnoid space of the CnS. in the subarachnoid space, the autoreactive T cells engage their<br />

cognate peptide–major histocompatibility complex expressed on dendritic cells to initiate inflammation.<br />

sclerosis is to be applauded and represents an<br />

important message for immunologists: findings<br />

in mice beg for confirmation in man.<br />

When results can be compared and found to<br />

be concordant from mouse to man, it is reassuring,<br />

and the importance of any such study<br />

is elevated above those far more numerous<br />

investigations restricted to the mouse alone.<br />

‘Translation’ to man is critical; the most<br />

potent therapy so far for the treatment of<br />

relapsing-remitting multiple sclerosis, natalizumab,<br />

came from initial experiments on the<br />

adhesion of human lymphocytes to inflamed<br />

brains of rodents with EAE 8 .<br />

These data are compelling and the evidence<br />

suggests that the earliest event of inflammation<br />

of the CNS in multiple sclerosis occurs<br />

through the interaction of CCR6 + T H -17 cells<br />

with CCL20 expressed by the epithelium of<br />

the choroid plexus 2 (Fig. 1). After this initial<br />

insult, the vasculature in the deep white<br />

matter becomes inflamed and recruits other<br />

immune cells to the CNS by the classic integrin<br />

VLA-4 (α 4 β 1 )–VCAM adhesion molecule–<br />

dependent mechanism 3,8,9 .<br />

There has been an active debate about<br />

which subset of helper T cells is most critical<br />

for the pathogenesis of multiple sclerosis and<br />

EAE. Originally, multiple sclerosis and EAE<br />

were regarded as T H 1 diseases. Interferon-γ<br />

(IFN-γ) is found in considerable abundance<br />

in the cerebrospinal fluid of patients with<br />

multiple sclerosis and mice with EAE 10 . Mice<br />

deficient in the ‘master’ transcription factor<br />

critical for T H 1 differentiation, T-bet, are<br />

resistant to the EAE 11 . Moreover, an early<br />

clinical trial reported that IFN-γ treatment<br />

exacerbated symptoms in multiple sclerosis<br />

patients 12 . However, subsequent studies<br />

have shown that the T H 1 cytokines IFN-γ<br />

and IL-12 have anti-inflammatory effects in<br />

EAE 13,14 . Mice deficient in these cytokines<br />

develop an aggressive atypical form of EAE<br />

that is indicative of greater infiltration of cells<br />

of the immune system into the brain. Those<br />

data catalyzed the discovery of T H -17 cells,<br />

and now, this population has gained notoriety<br />

as the pathogenic population in EAE and<br />

multiple sclerosis. IL-17 is also found in the<br />

inflamed CNS, and altering T H -17 differentiation<br />

can inhibit signs of EAE in mice 4,5 .<br />

But to complicate the issue, two provocative<br />

papers have provided data that question<br />

the pathogenic potential of T H -17 cells in<br />

neuroinflammation: first, mice with conditional<br />

deletion of IL-17 in T cells develop EAE<br />

normally 15 ; second, experimental uveitis is<br />

cured in rats by treatment with recombinant<br />

IL-17 (ref. 16). Assessing the data reported<br />

by Reboldi and colleagues 2 in the context<br />

of those other observations illuminates the<br />

problems of placing functional importance<br />

on a particular cytokine that might define a<br />

particular subset of autoreactive effector T<br />

cells. With this in mind, the implications of<br />

the data from Reboldi et al. 2 suggest that the<br />

important autoreactive effector T cells are not<br />

those that express IL-17 or IFN-γ but actually<br />

those T cells with an autoreactive brain<br />

tissue–specific T cell antigen receptor that also<br />

concomitantly express CCR6. In other words,<br />

the pathogenic function of these cells is due<br />

to their ability to infiltrate the target tissue<br />

and engage major histocompatibility complex<br />

to generate an immune response against the<br />

nervous system. This feature is independent<br />

of whether the pathogenic T cells are T H 1,<br />

T H -17, T H 2 or ‘T H -9’. These studies, illuminating<br />

how pathogenic T cells gain entry into<br />

the normal brain, are marvelous both for their<br />

implications for understanding immune surveillance<br />

of the nervous system and for the<br />

elegant ‘translation’ of experimental work all<br />

the way from mice to man.<br />

1. Conti, B., Tabarean, i., Andrei, C. & Bartfai, T. Front.<br />

Biosci. 9, 1433–1449 (2004).<br />

2. Reboldi, A. et al. Nat. Immunol. 10, 514–523<br />

(2009).<br />

3. Steinman, L. Nat. Rev. Drug Discov. 4, 510–519<br />

(2005).<br />

4. Tzartos, J.S. et al. Am. J. Pathol. 172, 146–155<br />

(2008).<br />

5. Langrish, C.L. et al. J. Exp. Med. 201, 233–240<br />

(2005).<br />

6. Acosta-Rodriguez, E.V. et al. Nat. Immunol. 8, 639–<br />

646 (2007).<br />

7. Ransohoff, R.M., Kivisakk, P. & Kidd, G. Nat. Rev.<br />

454 volume 10 number 5 may 2009 nature immunology<br />

Kim Caesar


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Immunol. 3, 569–581 (2003).<br />

8. Yednock, T. et al. <strong>Nature</strong> 356, 63–66 (1992).<br />

9. Engelhardt, B., wolburg-Buchholz, K. & wolburg, H.<br />

Microsc. Res. Tech. 52, 112–129 (2001).<br />

10. Sospedra, M. & Martin, R. Annu. Rev. Immunol. 23,<br />

683–747 (2005).<br />

11. Bettelli, E. et al. J. Exp. Med. 200, 79–87 (2004).<br />

12. Panitch, H.S., Hirsch, R.L., Haley, A.S. & Johnson, K.P.<br />

Lancet 1, 893–895 (1987).<br />

13. willenborg, D.O., Fordham, S., Bernard, C.C., Cowden,<br />

w.B. & Ramshaw, i.A. J. Immunol. 157, 3223–3227<br />

(1996).<br />

14. Gran, B. et al. J. Immunol. 169, 7104–7110 (2002).<br />

15. Haak, S. et al. J. Clin. Invest. 119, 61–69 (2009).<br />

16. Ke, Y. et al. J. Immunol. 182, 3183–3190 (2009).<br />

Crohn’s disease-associated Nod2 mutants reduce IL10<br />

transcription<br />

Dana J Philpott & Stephen E Girardin<br />

The 3020insC mutation in Nod2 is associated with Crohn’s disease, but how it influences disease pathogenesis is<br />

unknown. A new study shows that the 3020insC mutant protein fails to activate a key transcription factor that drives<br />

interleukin 10 expression, resulting in reduced production of this anti-inflammatory cytokine.<br />

Crohn’s disease, a chronic inflammatory<br />

disorder that can affect the entire gastrointestinal<br />

tract1 , is generally believed to<br />

result from an inappropriate immune reaction<br />

to the normal commensal flora of the<br />

intestine. The most persuasive argument<br />

for this conclusion is that in many mouse<br />

models of inflammatory bowel disease,<br />

regardless of the underlying cause, intestinal<br />

inflammation does not develop if the animals<br />

are housed under germ-free conditions.<br />

Moreover, there is some improvement in<br />

patients who take antibiotics or have intestinal<br />

shunts, which divert intestinal contents<br />

away from the inflamed areas. The molecular<br />

basis for this overexuberant response<br />

to commensal flora is not understood, but<br />

the gene encoding nucleotide oligomerization<br />

domain 2 (Nod2) is present within a<br />

previously described Crohn’s ‘hot-spot’ on<br />

chromosome 16 called IBD1, and three Nod2<br />

coding region polymorphisms are associated<br />

with increased susceptibility to disease2,3 . In<br />

this issue of <strong>Nature</strong> <strong>Immunology</strong>, Noguchi<br />

et al. find that some of these mutations—<br />

specifically, those in the leucine-rich repeat<br />

(LRR) domain of Nod2—result in suppression<br />

of the production of the immunosuppressive<br />

cytokine interleukin 10 (IL-10).<br />

One of these polymorphisms is a frameshift<br />

mutation resulting from an insertion<br />

of a cytosine at nucleotide position 3020 in<br />

Nod2 (3020insC). The nucleotide insertion<br />

generates a premature stop codon, resulting<br />

in a truncated protein lacking part of the last<br />

Dana J. Philpott is in the Department of<br />

<strong>Immunology</strong> and Stephen e. Girardin is in<br />

the Department of Laboratory Medicine and<br />

Pathobiology at the University of Toronto, Toronto,<br />

Ontario, Canada.<br />

e-mail: dana.philpott@utoronto.ca<br />

LRR near the C terminus of the protein. The<br />

relative risk for developing Crohn’s disease<br />

in individuals bearing a homozygous frameshift<br />

mutation in Nod2 has been calculated<br />

to be approximately 30 times that of normal<br />

individuals. Moreover, these individuals<br />

show a more severe disease phenotype, with<br />

increased risk of ileal stenosis and the need<br />

for surgical intervention 4 .<br />

Since the discovery of the link between<br />

Nod2 and Crohn’s disease, many studies<br />

have focused on trying to understand how<br />

mutations in Nod2 influence Crohn’s disease<br />

development. Nod2 is an Apaf-like<br />

molecule with a distinct domain organization<br />

comprising two N-terminal caspase-activating<br />

and recruitment domains<br />

(CARDs), a central nucleotide binding site<br />

(NBS) and a C-terminal series of LRRs 5 .<br />

The domain architecture of Nod2 is similar<br />

to plant R (for resistance) proteins, which<br />

mediate defense against invading pathogens.<br />

Indeed, Nod2, a member of a family<br />

of cytosolic innate immune molecules called<br />

NBS-LRR receptors (NLRs), is also involved<br />

in host defense. Comprising 22 members<br />

so far, the NLR family includes proteins<br />

that are involved in sensing both microbe-<br />

associated and danger-associated molecular<br />

patterns within cells and commencing innate<br />

immune responses 6 .<br />

Nod2 responds to a motif called muramyl<br />

dipeptide (MDP), which is found in the cell<br />

wall of Gram-positive and Gram-negative<br />

bacteria 7,8 and is a component of peptidoglycan.<br />

It is not yet known how Nod2 senses<br />

MDP; indeed, no study so far has been able<br />

to show whether MDP binds directly to<br />

Nod2. On the basis of studies in the Tolllike<br />

receptor field and plant R proteins, it has<br />

been postulated that the LRR region is likely<br />

to represent the site at which ligand sens-<br />

nEwS AnD ViEwS<br />

ing occurs, either directly or through a coreceptor.<br />

MDP sensing is thought to unfold<br />

Nod2, leading to oligomerization through<br />

the NBS domain. This process exposes the<br />

CARDs, which then provide a platform for<br />

the recruitment of Rip2, a CARD-containing<br />

signaling kinase that triggers the NF-κB<br />

pathway 5 . Rip2 also activates the p38 and Erk<br />

MAP kinases upon MDP stimulation 9 . These<br />

inflammatory pathways lead to the transcription<br />

and expression of several genes, many of<br />

which are proinflammatory or provide some<br />

defensive function.<br />

In relation to Crohn’s disease, the 3020insC<br />

mutation in Nod2 results in a protein product<br />

that, when expressed in vitro, no longer<br />

reacts to MDP to drive NF-κB signaling 7,8 .<br />

These findings are supported by studies<br />

showing that when cells are isolated from the<br />

blood of humans with Crohn’s disease who<br />

are homozygous for the 3020insC mutation<br />

and treated in vitro with MDP, they do not<br />

produce certain cytokines in response to this<br />

bacterial product 10 . A key point, however, is<br />

that no study so far has proven that a mutated<br />

Nod2 protein is actually present in tissues<br />

from subjects with Crohn’s disease. It is possible<br />

that this mutation results in an unstable<br />

protein product. If this is the case, then the<br />

Nod2-deficient mouse is an acceptable model<br />

for the disease. However, if the protein is<br />

present, it might have ligand-independent<br />

functions. Indeed, the paper by Noguchi et<br />

al. supports the latter possibility 11 .<br />

Over the years, much research has gone<br />

into trying to unravel the apparent paradox:<br />

the 3020insC mutation in Nod2 results in a<br />

protein product incapable of responding to<br />

a bacterial ligand, yet Crohn’s disease results<br />

in overt inflammation that probably is triggered<br />

by bacterial products. One explanation<br />

that might help resolve this paradox<br />

nature immunology volume 10 number 5 may 2009 455


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

nEwS AnD ViEwS<br />

MDP stimulated<br />

MDP<br />

IκB-α<br />

Nod2<br />

Rip2<br />

IKK<br />

Nod2<br />

NF-κB<br />

Proinflammatory<br />

gene transcription<br />

P<br />

Healthy human:<br />

normal Nod2<br />

Basal Basal MDP stimulated<br />

P<br />

p38<br />

Nod2<br />

P<br />

Nod2<br />

hnRNP-A1 hnRNP-A1<br />

P<br />

p38<br />

P P<br />

IL10<br />

transcription<br />

relies on the observation that cells or tissues<br />

isolated from Crohn’s disease patients show<br />

depressed IL-10 production 12 . IL-10 is produced<br />

by various hematopoietic cells, including<br />

macrophages, dendritic cells and T cells,<br />

and acts to contain and suppress inflammatory<br />

responses and thereby to downmodulate<br />

adaptive immune responses and minimize<br />

tissue damage in response to microbes. It<br />

seems that enhancing IL-10 production,<br />

which can occur after treatment with certain<br />

probiotic bacteria, helps to calm inflammation<br />

in Crohn’s disease 12 . Moreover, reduced<br />

intestinal colonization with an IL-10promoting<br />

member of the commensal flora,<br />

Fecalibacterium prausnitzii, is associated with<br />

enhanced risk of recurrence of Crohn’s disease<br />

after surgical resection 13 . Noguchi et al.<br />

shed light on why IL-10 expression is affected<br />

in Crohn’s disease.<br />

Noguchi et al. begin with the observation<br />

that bacterially induced IL-10 production<br />

is dampened in primary human monocytes<br />

from subjects with Crohn’s disease having<br />

the 3020insC Nod2 mutation. Although<br />

?<br />

?<br />

Mouse:<br />

normal Nod2<br />

P<br />

MDP<br />

3020insC<br />

Nod2<br />

No<br />

signaling<br />

P<br />

p38<br />

Crohn’s disease:<br />

mutant Nod2<br />

Basal Basal<br />

hnRNP-A1<br />

No basal Il10<br />

transcription<br />

IL10<br />

transcription<br />

actual suppression of IL-10 production is<br />

not clear from their studies with these cells,<br />

synergy between Toll-like receptors (TLRs)<br />

and Nod2 is clearly lost. Consistent with this<br />

finding, several previous studies have shown<br />

a synergistic production of cytokines upon<br />

coactivation of TLR2 and Nod2, which is lost<br />

in the 3020insC mutants. Transduction of<br />

human monocytes with a cDNA encoding the<br />

3020insC mutant Nod2 protein blocks basal<br />

transcription and expression of IL-10, but<br />

leaves other cytokines, such as IL-1β, unaffected.<br />

Expression of the 3020insC form of<br />

Nod2 blocks IL10 promoter activity as measured<br />

from a luciferase reporter. Notably, two<br />

other Crohn’s disease–associated Nod2 variants,<br />

R702W and G908R, similarly inhibit<br />

IL10 promoter activity. The 3020insC Nod2<br />

mutant has no affect on the activity of mouse<br />

Il10 reporter constructs and, similarly, the<br />

equivalent mutation made in the mouse<br />

Nod2 gene, referred to as 2939insC, does<br />

not affect human IL10 transcription. Taken<br />

together, these findings demonstrate that the<br />

human 3020insC mutant of Nod2 acts spe-<br />

cifically to block the human IL10 promoter<br />

(Fig. 1). Moreover, it may explain why the<br />

2939insC mouse mutation, when knocked<br />

into the mouse Nod2 locus, does not behave<br />

like the human 3020insC mutant with regard<br />

to proinflammatory cytokine production 14 .<br />

Noguchi et al. show that heterogeneous<br />

nuclear ribonucleoprotein A1 (hnRNP A1)<br />

binds constitutively to the IL10 promoter.<br />

When 3020insC is expressed in cells, the<br />

binding of hnRNP A1 to IL10 is reduced.<br />

Exogenously expressed hnRNP A1 stimulates<br />

IL10 promoter activity, and knocking down<br />

hnRNP A1 expression in primary human<br />

monocytes reduces lipopolysaccharide-<br />

stimulated production of IL-10. Convincingly,<br />

through chromatin immunoprecipitation<br />

analysis, the authors show reduced binding<br />

of hnRNP A1 to the IL10 promoter region in<br />

humans with the 3020insC Nod2 genotype,<br />

who are highly deficient in IL-10 production.<br />

But how does the presence of the 3020insC<br />

mutation in Nod2 result in reduced transcription<br />

factor binding at the IL10 promoter?<br />

Noguchi et al. found that wild-type<br />

Nod2 forms a trimeric complex in the cytoplasm<br />

with hnRNP A1 as well as active,<br />

phosphorylated p38, and it appears that p38<br />

is able to serine-phosphorylate hnRNP A1.<br />

This phosphorylation event is required for<br />

cleavage of hnRNP A1 into its active form,<br />

which can then translocate to the nucleus to<br />

act on the IL10 promoter (Fig. 1). However,<br />

the 3020insC Nod2 mutant protein interacts<br />

with p38 but not with p38hnRNP A1. The<br />

lack of this interaction reduces the presence<br />

of phosphorylated hnRNP A1 in the nucleus,<br />

thereby affecting basal transcription at the<br />

IL10 promoter. Interestingly, the mouse<br />

2939insC Nod2 mutant protein resembles<br />

wild-type mouse Nod2 in its influence on<br />

the interaction between phospho-p38 and<br />

hnRNP A1; this observation provides mechanistic<br />

insight into why the mouse mutant<br />

does not affect IL10 transcription.<br />

Taken together, the findings of Noguchi et<br />

al. stress an important point with regard to<br />

the 3020insC mutation in Nod2 associated<br />

with Crohn’s disease: this mutation might be<br />

a loss of function or gain of function, depending<br />

on the response examined. Indeed, it certainly<br />

seems that the mutant protein is unable<br />

to sense MDP. One might argue that the phenotype<br />

caused by the 3020insC mutation in<br />

Nod2 described by Noguchi et al. also represents<br />

a loss of function. Indeed, the 3020insC<br />

mutant Nod2 protein is unable to form a trimolecular<br />

complex that leads to the activation<br />

of hnRNP A1 and thereby affects the ability<br />

of hnRNP A1 to bind to the IL10 promoter.<br />

However, the authors’ point is well taken that<br />

456 volume 10 number 5 may 2009 nature immunology<br />

p38<br />

2939insC<br />

Nod2<br />

P<br />

?<br />

?<br />

Mouse:<br />

mutant Nod2<br />

P<br />

hnRNP-A1<br />

Figure 1 Reduced production of iL-10 in cells expressing Crohn’s disease–associated mutations in<br />

Nod2. in cells from humans and mice expressing wild-type Nod2, muramyl dipeptide (MDP) triggers<br />

nod2 unfolding, allowing its interaction with Rip2 and the subsequent activation of the nF-κB pathway,<br />

which drives proinflammatory cytokine gene transcription. noguchi et al. show that wild-type nod2<br />

forms a trimolecular complex with active, phosphorylated p38 and the transcription factor hnRnP A1.<br />

This association drives IL10 transcription during the steady-state and after bacterial stimulation (left<br />

panel). The Crohn’s disease–associated 3020insC mutant nod2 (right panel) fails to detect MDP and<br />

activate the nF-κB pathway, and does not efficiently interact with hnRnP A1; as a result, hnRnP A1<br />

activation and subsequent transcriptional control of the IL10 promoter is impaired. Reduced iL-10<br />

production is therefore a consequence of this mutation and may contribute to the hyperinflammatory<br />

response in the intestine that is characteristic of Crohn’s disease. The equivalent mutation translated<br />

to mouse nod2 results in a protein that retains its ability to activate Il10 transcription. These findings<br />

provide a possible explanation of why cells isolated from a knock-in mouse with this equivalent<br />

mutation do not behave like their human counterparts.<br />

P


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

if the mutant protein is indeed expressed, it<br />

might have acquired a new, proinflammatory<br />

“gain of function” that is distinct from<br />

the activity of the wild-type version of Nod2.<br />

Finally, these studies bring ideas for possible<br />

treatments for Crohn’s disease: targeting<br />

hnRNP A1 directly to influence its ability to<br />

drive IL10 transcription might bypass the<br />

block erected by mutant Nod2.<br />

1. Xavier, R.J. & Podolsky, D.K. <strong>Nature</strong> 448, 427–434<br />

(2007).<br />

2. Hugot, J.P. et al. <strong>Nature</strong> 411, 599–603 (2001).<br />

3. Ogura, Y. et al. <strong>Nature</strong> 411, 603–606 (2001).<br />

4. Seiderer, J. et al. Inflamm. Bowel Dis. 12, 1114–1121<br />

(2006).<br />

5. Ogura, Y. et al. J. Biol. Chem. 276, 4812–4818<br />

(2001).<br />

6. Franchi, L., warner, n., Viani, K. & nunez, G. Immunol.<br />

Rev. 227, 106–128 (2009).<br />

7. Girardin, S.E. et al. J. Biol. Chem. 278, 8869–8872<br />

(2003).<br />

8. inohara, n. et al. J. Biol. Chem. 278, 5509–5512<br />

(2003).<br />

9. Park, J.H. et al. J. Immunol. 178, 2380–2386 (2007).<br />

10. netea, M.G. et al. J. Biol. Chem. 280, 35859–35867<br />

(2005).<br />

11. noguchi, E. et al. Nat. Immunol. 10, 471–479<br />

(2009).<br />

12. Baumgart, D.C. & Sandborn, w.J. Lancet 369, 1641–<br />

1657 (2007).<br />

13. Sokol, H. et al. Proc. Natl. Acad. Sci. USA 105, 16731–<br />

16736 (2008).<br />

14. Maeda, S. et al. Science 307, 734–738 (2005).<br />

The Foxo and the hound: chasing the in vivo regulation<br />

of T cell populations during infection<br />

Elia D Tait & Christopher A Hunter<br />

T cell expansion and contraction during the immune response to pathogens are regulated by a wide variety of cellintrinsic<br />

and cell-extrinsic factors. A new study identifies a role for CTLA-4 signaling and activation of the Foxo3<br />

transcription factor in modulating T cell populations.<br />

In this issue of <strong>Nature</strong> <strong>Immunology</strong>, Dejean<br />

et al. provide new insight into the role of the<br />

forkhead box (Fox) transcription factor Foxo3<br />

in limiting the magnitude of T cell responses1 .<br />

The Fox family of transcription factors participate<br />

in a multitude of biological processes,<br />

including the cell cycle, metabolism and the<br />

stress response. The activity of proteins in the<br />

Foxo subfamily is predominantly regulated by<br />

phosphorylation status and subcellular localization,<br />

but these transcription factors are also<br />

subject to a number of post-translational modifications,<br />

including acetylation and ubiquitination.<br />

Growth factors mobilize numerous kinases,<br />

including Akt, that phosphorylate Foxo proteins,<br />

leading to their export from the nucleus and<br />

inactivation. In contrast, stress stimuli mobilize<br />

a different set of kinases, notably Jun N-terminal<br />

kinase (Jnk), that promote nuclear import of<br />

Foxo and the initiation of Foxo-dependent<br />

gene regulation programs2 . These transcription<br />

factors can modulate gene expression by binding<br />

promoters at Foxo-binding sites to recruit<br />

transcriptional machinery, by competing for<br />

promoter binding sites with other factors, and<br />

by dynamically regulating interactions between<br />

cofactors and various elements of the transcriptional<br />

machinery3 .<br />

In the past decade, immunologists have come<br />

to appreciate the importance of the Fox family<br />

in coordinating immune responses, particularly<br />

elia D. Tait and Christopher A. Hunter are in the<br />

Department of Pathobiology, School of veterinary<br />

Medicine, University of Pennsylvania, Philadelphia,<br />

Pennsylvania, USA.<br />

e-mail: chunter@vet.upenn.edu<br />

in the context of Foxp3 expression by T regulatory<br />

(T reg ) cells. The Foxo subfamily members<br />

(in mammals, Foxo1, Foxo3, Foxo4 and Foxo6)<br />

have also been shown to influence T cell function.<br />

T cell receptor (TCR) and co-receptor<br />

ligation modulate the activity of Foxo proteins<br />

that upregulate prosurvival programs in the<br />

presence of growth factors and initiate apoptotic<br />

pathways in the absence of mitogen or<br />

cytokine 2 . In 2004, Pandiyan et al. demonstrated<br />

that signaling through the inhibitory receptor<br />

cytotoxic T lymphocyte–associated antigen-4<br />

(CTLA-4) protected T helper type 2 (T H 2)<br />

cells from apoptosis via Foxo3 inactivation and<br />

subsequent upregulation of the survival protein<br />

Bcl-2 (ref. 4). Consistent with a proapoptotic,<br />

regulatory role for Foxo3, Lin et al. reported<br />

that Foxo3a antagonizes signaling by the transcription<br />

factor NF-κB and that mice carrying a<br />

mutated Foxo3a allele underwent spontaneous<br />

T cell hyperproliferation and inflammation 5 . In<br />

2007, it was shown that Foxo3a participates in<br />

the maintenance of memory CD4 + T cells by<br />

coordinating signals from cytokine receptors<br />

and the TCR 6 . Thus, Foxo proteins appear to<br />

have an intrinsic role in T cell homeostasis and<br />

in limiting T cell activity.<br />

Dejean et al. examined how Foxo3 influences<br />

T cell population expansion and contraction<br />

during infection by lymphocytic choriomeningitis<br />

virus (LCMV). In this model, the absence<br />

of Foxo3 led to an exaggerated antigen-specific<br />

T cell accumulation, with largely normal contraction<br />

1 . Whereas a previous study had suggested<br />

that Foxo3 was “largely dispensable” 5 in<br />

modulating T cell apoptosis and was involved<br />

exclusively in spontaneous T cell population<br />

nEwS AnD ViEwS<br />

expansion, Dejean et al. showed that antigenspecific<br />

T cells in Foxo3-deficient mice had<br />

enhanced expression of the prosurvival factor<br />

Bcl-2 and decreased binding to the apoptosis<br />

marker annexin-5. Notably, the expansion<br />

of T cell populations observed in the Foxo3deficient<br />

mice was not T cell intrinsic and was<br />

instead dependent on increased production<br />

of the cytokine interleukin-6 (IL-6) by Foxo3deficient<br />

dendritic cells (DC) 1 . IL-6 has long<br />

been regarded as proinflammatory, and it also<br />

has a strong prosurvival effect on activated T<br />

cells 7 . Blockade of IL-6R α-chain restored the<br />

magnitude of the LCMV-driven T cell response<br />

in the Foxo3-deficient environment to a level<br />

comparable to that observed in wild-type mice.<br />

Finally, Dejean et al. showed that CTLA-4 ligation<br />

induced nuclear localization and activation<br />

of Foxo3 and inhibited the production of<br />

IL-6 in DCs in vitro 1 . These findings provide a<br />

previously unsuspected mechanism by which<br />

CTLA-4–expressing T cells can limit their own<br />

survival; by initiating Foxo3 activation and<br />

nuclear import in DCs, T cells modulate DC<br />

cytokine production and thereby influence their<br />

own viability (Fig. 1).<br />

These findings illustrate the importance<br />

of the interactions that take place between<br />

DCs and T cells in regulating the expansion<br />

and contraction of T cell populations during<br />

infection. In addition, this work identifies a<br />

new role for Foxo3 in modulating this process.<br />

Understanding how T cell responses are<br />

regulated and maintained during infection<br />

has obvious implications for vaccine development.<br />

To this end, multiple DC–T cell interactions<br />

that result in the activation of these two<br />

nature immunology volume 10 number 5 may 2009 457


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

nEwS AnD ViEwS<br />

CTLA-4 signals:<br />

↓ IL-2 production<br />

↓ Activation<br />

↓ Cell cycle<br />

Lack of IL-6 signals:<br />

↓ Bcl-2<br />

↓ Viability<br />

IL-6R<br />

CTLA-4<br />

TCR<br />

CD4<br />

IL-6<br />

populations have been characterized. The<br />

CD40-CD40L interaction is one example of<br />

a mechanism through which T cells influence<br />

their own activation and accumulation<br />

by optimizing DC responses. In contrast,<br />

the interactions between T cells and DCs<br />

that govern the contraction and control of<br />

T cell numbers are less well defined 8 . DCs<br />

may imprint T cells during activation with<br />

apoptotic programs, and T cell population<br />

expansion is certainly limited by T cell–<br />

mediated DC killing 9 . Also, interactions<br />

of CTLA-4 and programmed death-1 with<br />

their respective ligands intriusically modulate<br />

T cell responses and are important in<br />

controlling autoimmunity 8 . Specifically,<br />

mice lacking CTLA-4 develop spontaneous<br />

lymphoproliferative disease (similar to<br />

that of the Foxo3-deficient mice described<br />

previously 5 ), indicating that this molecule<br />

is vital in maintaining T cell homeostasis 10 .<br />

Overall, however, the signals that orchestrate<br />

the fine balance between too much and too<br />

little T cell population expansion during an<br />

immune response are still unclear.<br />

The findings of Dejean et al. suggest a scenario<br />

in which activated T cells that express<br />

CTLA-4 (or T reg cells that constitutively<br />

express CTLA-4; ref. 11) can induce nuclear<br />

localization of Foxo3, and this active Foxo3<br />

then limits the production of IL-6 by DCs<br />

B7<br />

MHC II<br />

Jnk<br />

Foxo3<br />

P<br />

and thus controls T cell population expansion<br />

(Fig. 1) 1 . Certain aspects of this model<br />

need to be tested further, however. Some<br />

previous studies have shown that antibodymediated<br />

blockade of CTLA-4 can lead to<br />

enhanced T cell responses during infection 12<br />

and cancer 13 , and the model outlined in<br />

Figure 1 suggests that antagonizing CTLA-4<br />

in vivo during LCMV infection should<br />

lead to an increase in the expansion of T<br />

cell populations 1 . However, in agreement<br />

with other studies 14 , Dejean et al. found<br />

that this was not the case: LCMV-driven T<br />

cell responses were unaffected by CTLA-4<br />

blockade in vivo 1 . These negative data serve<br />

as a reminder that blocking CTLA-4 has<br />

different effects during different types of<br />

immune responses 1,4,8,12–14 . Thus, the effect<br />

of CTLA-4 signaling on the expansion and<br />

contraction of the T cell population overall<br />

might vary depending on context. The<br />

apparent contradiction between the model<br />

described in Figure 1 and the negative in vivo<br />

data of Dejean et al. 1 also indicates that there<br />

may be ligands other than CTLA-4 that lead<br />

to Foxo3 activation in vivo. The model proposed<br />

by Dejean et al. prompts other questions<br />

regarding the biology and function of<br />

Foxo3 in response to CTLA-4 signaling. The<br />

lack of Foxo3-binding sites in the Il6 promoter<br />

1 implies that Foxo3 may regulate Il6<br />

transcription through indirect means or by<br />

upregulating a broader anti-inflammatory<br />

gene expression program that includes<br />

downregulation of IL-6 production. Overall,<br />

the principles established by Dejean et al. are<br />

important in advancing our understanding<br />

of the role of DC–T cell interactions in controlling<br />

T cell populations, and they highlight<br />

some of the contradictions and future<br />

directions in this field.<br />

In total, the current literature suggests that<br />

CTLA-4 and Foxo3 have important roles in<br />

the control of spontaneous and antigeninduced<br />

T cell population expansion. In this<br />

issue, Dejean et al. make an important contribution<br />

by providing evidence linking these<br />

pathways 1 . In the context of treating human<br />

disease, particularly autoimmune disease,<br />

controlling the immune response using<br />

therapies that mimic natural cell-intrinsic<br />

immunoregulation is an appealing strategy.<br />

CTLA-4–Ig, a fusion between CTLA-4<br />

and IgG that mimics CTLA-4, was initially<br />

thought to act as an antagonist of B7-CD28–<br />

mediated signaling that would downregulate<br />

inappropriate T cell population expansion in<br />

the context of autoimmunity; we now appreciate<br />

that it has additional effects, including<br />

inducing the immunosuppressive indoleamine-2,3-dioxygenase<br />

pathway 8 . This fusion<br />

protein is now used for the management of<br />

rheumatoid arthritis, but the true biological<br />

mechanism behind its clinical success is<br />

not entirely clear. Whether Foxo3 signaling<br />

is also influenced during treatment with<br />

CTLA-4–Ig remains to be tested. Although<br />

CTLA-4 signaling and the in vivo biology<br />

of Foxo3 are still not fully understood, the<br />

work of Dejean et al. 1 prompts exciting new<br />

questions about the basic biology and clinical<br />

relevance of these factors.<br />

1. Dejean, A.S. et al. Nat. Immunol. 10, 504–513<br />

(2009).<br />

2. Peng, S.L. Oncogene 27, 2337–2344 (2008).<br />

3. Fu, Z. & Tindall, D.J. Oncogene 27, 2312–2319<br />

(2008).<br />

4. Pandiyan, P. et al. J. Exp. Med. 199, 831–842<br />

(2004).<br />

5. Lin, L., Hron, J.D. & Peng, S.L. Immunity 21, 203–<br />

213 (2004).<br />

6. Riou, C. et al. J. Exp. Med. 204, 79–91 (2007).<br />

7. Teague, T.K. et al. J. Exp. Med. 191, 915–926<br />

(2000).<br />

8. Fife, B.T. & Bluestone, J.A. Immunol. Rev. 224,<br />

166–182 (2008).<br />

9. Masson, F., Mount, A.M., wilson, n.S. & Belz, G.T.<br />

Immunol. Cell Biol. 86, 333–342 (2008).<br />

10. waterhouse, P. et al. Science 270, 985–988 (1995).<br />

11. wing, K. et al. Science 322, 271–275 (2008).<br />

12. Kaufmann, D.E. et al. Nat. Immunol. 8, 1246–1254<br />

(2007).<br />

13. Fong, L. & Small, E.J. J. Clin. Oncol. 26, 5275–5283<br />

(2008).<br />

14. Bachmann, M.F. et al. J. Immunol. 160, 95–100<br />

(1998).<br />

458 volume 10 number 5 may 2009 nature immunology<br />

Foxo3<br />

IL-6 IL-6<br />

Figure 1 T cells can limit their own accumulation by multiple mechanisms mediated by CTLA-4<br />

signaling and the Foxo3 transcription factor. During antigen presentation, B7 molecules on DCs engage<br />

CTLA-4 expressed on activated T cells or T reg cells. The resulting signals promote direct control of T cell<br />

activation and expansion through downregulation of iL-2 production, decreased T cell activation and<br />

decreased cell cycle progression. Also, the B7–CTLA-4 interaction initiates signaling in the DCs that<br />

mobilizes nuclear import and activation of Foxo3, which indirectly controls T cell population expansion<br />

by limiting DC production of the prosurvival cytokine iL-6. T cells with limited access to iL-6 are more<br />

susceptible to apoptosis and have lower expression of Bcl-2. MHC ii, major histocompatibility class ii.<br />

P<br />

Kim Caesar


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Dimers dampen T cell activity<br />

The transcription factor NFAT influences T cell function by<br />

acting together with other transcription factors, including<br />

AP-1, T-bet GATA-3 and Foxp3. In the Journal of Experimental<br />

Medicine, Macian and colleagues show that NFAT homodimers<br />

are needed for the transcription of some genes associated with<br />

T cell anergy. In T cells stimulated with ionomycin, mutant NFAT<br />

proteins that bind to AP-1 but not to other NFAT proteins induce<br />

activation of some (Tle4 and Dgka) but not other (Rnf128 and<br />

Casp3) anergy-associated genes. Tandem κB-like binding sites<br />

in the Rnf128 promoter—to which NFAT homodimers bind—are<br />

needed for ionomycin-induced transcription of this gene.<br />

T cells expressing the NFAT mutant that does not form dimers<br />

produce more interleukin 2 and have a lower anergy index score<br />

than do T cells transduced with a constitutively active form of<br />

NFAT. Whether NFAT proteins form dimers with other proteins<br />

to facilitate expression of anergy-related genes such as Tle4 and<br />

Dgka remains to be determined. CB<br />

J. Exp. Med. (23 March 2009) doi:10.1084/jem20082731<br />

Facilitating cross-presentation<br />

Although macrophages, neutrophils and dendritic cells (DCs) are known<br />

to have phagosomal pathways that differ in terms of oxidation, pH and<br />

degradation, it is not known whether phagocytic organization differs<br />

among DC subsets. In Immunity, Amigorena and colleagues compare the<br />

phagocytic pathways of splenic CD8 + and CD8 – DCs. The former subset<br />

is known to cross-present antigens, and cross-presentation requires a high<br />

pH in the phagosome. Consistent with that, CD8 + DCs restrict assembly<br />

of the NADPH oxidase complex to the phagosome, which results in the<br />

production of reactive oxygen species and local alkalization. CD8 – DCs,<br />

however, are unable to assemble this complex on the phagosomal membrane.<br />

The ability of CD8 + DCs to assemble the NADPH oxidase complex<br />

depends on the GTPas Rac2, and deficiency of Rac2 results in a lower<br />

phagosomal pH and cross-presentation efficacy. These data show that DC<br />

subpopulations have differences in their endophagocytic pathways that<br />

can explain their different abilities to cross-present antigen. JDKW<br />

Immunity (26 March 2009) doi:10.1016/j.immuni.2009.01.013<br />

Systemic suppression<br />

Retinoic acid (RA) induces regulatory T cells in the intestine,<br />

but whether it is involved in the systemic immune response<br />

is unclear. In <strong>Nature</strong> Medicine, Pulendran and colleagues<br />

now find that zymosan induces, through TLR2, expression<br />

of RA–metabolizing enzyme in splenic DCs. This Erk kinase–<br />

dependent pathway produces RA that functions in an autocrine<br />

way to upregulate SOCS3, a negative regulator of the kinase<br />

p38 and proinflammatory cytokines. Consistent with that,<br />

RA, in conjunction with IL-10, acts synergistically to induce<br />

regulatory T cells and suppress T helper type 1 (T H 1) and T H -17<br />

autoimmune responses in vivo. In the absence of TLR2, dectin-1,<br />

a C-type lectin receptor that also recognizes zymosan, induces a<br />

proinflammatory cytokine response in splenic DCs, which promotes<br />

T H1 and T H-17 responses. Identification of the physiological<br />

conditions in which zymosan triggers dectin-1 but not TLR2 in<br />

splenic DCs is needed. JDKW<br />

Nat. Med. (1 March 2009) doi:10.1038/nm.1925<br />

Written by Christine Borowski, Laurie A. Dempsey & Jamie D.K. Wilson<br />

Distinguishing danger<br />

How the body distinguishes sterile inflammation from pathogen-induced<br />

inflammatory responses has been unclear. In Science, Chen et al. show<br />

that the recognition of endogenous ‘danger-associated molecular pattern’<br />

(DAMP) proteins by CD24 and the signaling molecule Siglec-10 inhibits<br />

activation of the transcription factor NF-κB, thereby braking inflammation<br />

due to non–microbe-induced tissue damage. CD24, a glycosylphosphoinositol-anchored<br />

protein, in association with Siglec-10, binds DAMP<br />

proteins such as high-mobility group box 1 and heat-shock proteins 70<br />

and 90. Mice lacking either CD24 or Siglec-10 succumb to sublethal doses<br />

of acetaminophen, which is toxic to liver tissue and induces hepatocyte<br />

necrosis. Siglec-10 interacts with the SHP-1 tyrosine phosphatase to block<br />

activation and nuclear translocation of the NF-κB family member p65.<br />

Despite having more NF-κB activation in response to DAMP proteins,<br />

neither CD24- or Siglec-10-deficient dendritic cells show enhanced NF-κB<br />

activation in response to pathogen-associated molecular pattern motifs<br />

such as lipopolysaccharide. Thus, CD24–Siglec-10 regulation acts to limit<br />

tissue damage in response to non-pathogen threats. LAD<br />

Science 323, 1722–1725 (2009)<br />

Regulating stress<br />

ReseARch highlights<br />

Cellular stress triggered by an abundance of unfolded proteins<br />

in the endoplasmic reticulum activates the stress sensor<br />

IREα1. IREα1 contains an endoribonuclease whose splicing<br />

activity produces functional mRNA encoding the transcription<br />

factor XBP-1 (called ‘XBP-1s’), which, among other functions,<br />

is required for immunoglobulin secretion in plasma cells.<br />

In Molecular Cell, Glimcher and colleagues identify Bax<br />

inhibitor 1 (BI-1) as a negative regulator of this stress-induced<br />

activation of IREα1 and XBP-1s. Like IREα1, BI-1 localizes<br />

to the endoplasmic reticulum membrane and physiologically<br />

associates with IREα1 through its conserved carboxy-terminal<br />

cytoplasmic tail and inhibits the endoribonuclease activity of<br />

IREα1. After stress induction, cells that lack BI-1 have higher<br />

expression of XBP-1s and proteins encoded by its target genes.<br />

Lipopolysaccharide stimulation yields much higher titers of<br />

IgM in BI-1-deficient B cells than in wild-type B cells. How<br />

endoplasmic reticulum stress alters the BI-1–IREα1 interaction<br />

remains unknown. LAD<br />

Mol. Cell 33, 679–691 (2009)<br />

Degrading inflammatory mRNA<br />

If left unrestrained, TLR-driven immune responses induce unwanted tissue<br />

damage. In <strong>Nature</strong>, Akira and colleagues identify Zc3h12a, a CCCHtype<br />

zinc-finger protein that increases in abundance after TLR signaling, as<br />

an essential feedback inhibitor of TLR-induced inflammatory responses.<br />

Zc3h12a –/– mice show impaired survival and rampant inflammation<br />

characterized by excessive accumulation of plasma cells and activated T<br />

cells in several organs, granuloma formation in lymph nodes, and serum<br />

hyperimmunoglobulinemia. Macrophages from Zc3h12a –/– mice contain<br />

more Il6 and Il12p40 mRNA but similar amounts of Tnf and Cxcl1<br />

mRNA after TLR stimulation. Zc3h12a degrades—apparently through<br />

an endonuclease activity—Il6 and Il12p40 mRNA by a mechanism that<br />

depends on the 3′ untranslated regions of these molecules. Additional<br />

work is needed to determine if other ‘proinflammatory’ mRNA transcripts<br />

are degraded by Zc3h12a and contribute to the severely detrimental phenotype<br />

of Zc3h12a –/– mice. CB<br />

<strong>Nature</strong> (25 March 2009) doi: 10.1038/nature07924<br />

nature immunology volume 10 number 5 may 2009 459


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Autophagy genes in immunity<br />

Herbert W Virgin 1 & Beth Levine 2<br />

In its classical form, autophagy is a pathway by which cytoplasmic constituents, including intracellular pathogens, are<br />

sequestered in a double-membrane–bound autophagosome and delivered to the lysosome for degradation. This pathway has been<br />

linked to diverse aspects of innate and adaptive immunity, including pathogen resistance, production of type I interferon, antigen<br />

presentation, tolerance and lymphocyte development, as well as the negative regulation of cytokine signaling and inflammation.<br />

Most of these links have emerged from studies in which genes encoding molecules involved in autophagy are inactivated in<br />

immune effector cells. However, it is not yet known whether all of the critical functions of such genes in immunity represent<br />

‘classical autophagy’ or possible as-yet-undefined autophagolysosome-independent functions of these genes. This review<br />

summarizes phenotypes that result from the inactivation of autophagy genes in the immune system and discusses the pleiotropic<br />

functions of autophagy genes in immunity.<br />

Classical macroautophagy (called ‘autophagy’ here) involves the sequestration<br />

of cytoplasmic contents in a characteristic double-membraned<br />

vacuole, the autophagosome. Fusion of the outer autophagosomal<br />

membrane with the lysosome and the subsequent breakdown of the<br />

inner membrane results in the exposure of the sequestered cytoplasmic<br />

material to lysosomal hydrolases 1,2 (Fig. 1). A related term, ‘xenophagy’,<br />

refers to the use of the autophagy pathway to digest foreign rather than<br />

self constituents 3 . The autophagy pathway has two main physiological<br />

functions: it rids the cell of unwanted constituents, and it recycles cytoplasmic<br />

material so that cells can maintain macromolecular synthesis<br />

and energy homeostasis during stressful conditions. These functions<br />

probably underlie the well described roles of autophagy in cell survival<br />

and in the prevention of neurodegeneration, cancer and aging and, at<br />

least in part, the protective roles of autophagy in pathogen-infected<br />

cells 1,2 .<br />

The autophagy pathway involves the concerted action of evolutionarily<br />

conserved gene products involved in the initiation of autophagy,<br />

elongation and closure of the autophagosome, and lysosomal fusion<br />

(Fig. 1). Beclin 1, in complex with the class III phosphatidylinositol-<br />

3-OH kinase (PI(3)K) Vps34 and other proteins, is important for<br />

the nucleation of autophagosomes. This macromolecular complex is<br />

critical for integrating multiple signals that positively and negatively<br />

regulate autophagy. The autophagy (‘Atg’) protein Atg7 is required for<br />

two ubiquitin-like conjugation pathways that in concert result in the<br />

covalent conjugation of Atg5 to Atg12 and the conversion of LC3 (the<br />

mammalian ortholog of yeast Atg8) to its phosphatidylethanolamine-<br />

1 Department of Pathology and <strong>Immunology</strong>, Department of Molecular<br />

Microbiology and Department of Medicine, Washington University School<br />

of Medicine, St. Louis, Missouri, USA. 2Howard Hughes Medical Institute,<br />

Department of Internal Medicine and Department of Microbiology, University<br />

of Texas Southwestern Medical Center, Dallas, Texas, USA. Correspondence<br />

should be addressed to H.W.V. (virgin@wustl.edu) or B.L. (beth.levine@<br />

utsouthwestern.edu).<br />

Published online 20 April 2009; doi:10.1038/ni.1726<br />

reVIeW<br />

conjugated LC3-II form. The Atg5-Atg12 conjugate forms part of a large<br />

complex with Atg16L1 that is responsible for the membrane localization<br />

of the autophagic machinery and the generation of the autophagosome<br />

4 . At present, there are more than 20 recognized Atg proteins 1 , and<br />

it is likely that more will be identified. For example, FIP200, a protein<br />

reported to interact with the kinases FAK, Pyk2 and ASK1, the tumor<br />

suppressor TSC1, the tumor suppressor p53 and tumor necrosis factor<br />

receptor–associated factor 2, has been found to be critical for the initiation<br />

of autophagy through interaction with the kinases ULK1 and ULK2<br />

(mammalian orthologs of yeast Atg1) 5 . Furthermore, four independent<br />

groups have identified a candidate functional mammalian ortholog of<br />

yeast Atg14, known as either human Atg14 or Barkor, that interacts with<br />

Beclin 1 and the Vps34 class III PI(3)K complex to initiate autophagic<br />

vesicle nucleation 6–9 . Of the many proteins involved in autophagy, Beclin<br />

1, Atg5, Atg7, LC3, Atg12 and Atg16L1 have been studied in one or more<br />

aspects of immunity.<br />

The identification of this core molecular machinery has revolutionized<br />

the ability to detect and genetically manipulate the autophagy pathway.<br />

This has led to considerable advances in understanding the function<br />

of genes encoding molecules involved in autophagy (called ‘autophagy<br />

genes’ here) in immunity and other biological processes 1,2 . In terms of<br />

immunity, studies have shown that autophagy is regulated by pathways<br />

critical for the function and differentiation of cells of the immune system,<br />

including Toll-like receptors (TLRs), p47 GTPases, eIF2α kinases,<br />

and cytokines such as interferon-γ (IFN-γ) 10–13 . Furthermore, autophagy<br />

genes are important for thymic selection, lymphocyte development and<br />

survival, antigen presentation, killing of intracellular pathogens, and tissue<br />

homeostasis. There is also increasing evidence that autophagy genes<br />

may regulate innate immune signaling in a cell type–specific way.<br />

Although some of the functions described above, such as the targeting<br />

of microbes for lysosomal degradation (xenophagy), may unequivocally<br />

reflect the involvement of classical autophagy in immunity (Fig.<br />

2), it is not yet clear whether all functions of autophagy genes in immunity<br />

relate to the autophagy pathway itself (Fig. 3). This uncertainty<br />

arises from both the assays commonly used to detect autophagy and<br />

the lack of universal criteria for interpreting reverse-genetic studies.<br />

nature immunology volume 10 number 5 may 2009 461


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

REVIEw<br />

For example, the most commonly used<br />

autophagy assays, biochemical detection of<br />

the lipidated form of LC3 (LC3-II) and detection<br />

of the localization of LC3-II to punctate<br />

dots by light microscopy, are subject to many<br />

interpretations. Most investigators now recognize<br />

that detection of increased LC3-II can<br />

indicate either more autophagosome formation<br />

or a block in autophagosomal maturation,<br />

which necessitates the use of ancillary<br />

approaches to distinguish between these<br />

possibilities 14 . Perhaps less well appreciated<br />

is the idea that LC3 dots may represent the<br />

targeting of LC3 to structures other than<br />

autophagosomes, such as phagosomes, double-membraned<br />

scaffolds for the replication<br />

complexes of positive-strand RNA viruses, or<br />

even protein aggregates 15–17 . Similarly, LC3<br />

may become lipidated, forming LC3-II, in<br />

the absence of autophagosome formation 18,19 .<br />

Thus, whereas LC3-II formation and localization<br />

of LC3 punctae are hallmark features of<br />

autophagosome formation (and sensitive<br />

parameters for the detection of autophagy),<br />

they lack complete specificity as markers of<br />

classical autophagy. The direct demonstration<br />

of a function for autophagosomes in a<br />

process is the gold standard for proving that<br />

autophagy is involved. An extensive discussion<br />

of the various assays used in autophagy<br />

research, as well as the criteria for their use and<br />

interpretation, has been provided in a consensus<br />

paper published by many of the authorities<br />

in the field 14 .<br />

The genetic knockout or knockdown of<br />

core autophagy genes is an effective way to<br />

turn off the autophagy pathway, but it is often<br />

Autophagy induction<br />

Starvation<br />

Growth factor deprivation<br />

Immune signals:<br />

IFN-γ<br />

TNF<br />

TLRs<br />

PKR-elF2α kinase<br />

Jnk<br />

FADD<br />

Immunity-related GTPases<br />

Autophagy suppression<br />

Nutrient abundance<br />

Insulin-Akt-TOR signaling<br />

Immune signals:<br />

IL-4<br />

IL-13<br />

FADD<br />

Immunity-related GTPases<br />

unclear whether resulting phenotypes are due to a deficiency of classical<br />

autophagy or autophagy-independent functions of the autophagy genes.<br />

As reviewed elsewhere, autophagy genes are known to have alternative<br />

functions, including those in other membrane-trafficking events, axonal<br />

elongation and cell death 20,21 . Furthermore, investigators often assume<br />

that phenotypes noted in autophagy gene–deficient cells or organisms<br />

are due to lack of classical autophagy without providing direct experimental<br />

proof. Notably, for many studies of the involvement of autophagy<br />

genes in immunity (Fig. 3), it is not clear how classical autophagy, involving<br />

the autophagolysosomal degradation of sequestered contents, could<br />

contribute to the described functions of autophagy genes. Thus, it is<br />

possible that autophagy genes act in other cellular processes that affect<br />

immune effector cell development and function. Here we will discuss<br />

advances related to the function of autophagy genes in the immune<br />

system.<br />

Autophagy genes in antimicrobial defense<br />

More than a decade ago, the first published paper on Beclin 1, a mammalian<br />

autophagy protein, demonstrated an antiviral function for<br />

exogenous neuronal expression of Beclin 1 in mice with alphavirus<br />

encephalitis 22 . Subsequently, endogenous autophagy genes were shown<br />

to be critical for the successful innate immune response to fungal, bacterial<br />

and viral pathogens in plants 23 , and viral evasion of Beclin 1 function<br />

by a herpes simplex virus neurovirulence factor was found to be essential<br />

Vesicle<br />

nucleation<br />

Isolation<br />

membrane<br />

Vesicle<br />

elongation<br />

Docking &<br />

fusion<br />

Autophagosome<br />

AUTOPHAGY<br />

Atg16L1 Atg12 Atg12<br />

Atg10 E2 Atg12<br />

Atg5 Atg5<br />

Atg5<br />

Atg5 Atg16L1 Atg16L1<br />

Atg12<br />

Atg16L1 Atg16L1<br />

Atg7 E1<br />

Atg5<br />

Atg12<br />

Atg5<br />

Atg12<br />

PE<br />

LC3 -Gly LC3<br />

Atg3 E2<br />

Atg4 LC3<br />

Lysosome<br />

Class III PI(3)K complex<br />

Other<br />

regulators<br />

Beclin 1 Atg14<br />

Bcl-2–Bcl-XL Vps34 Vps15<br />

Ubiquitin-like conjugation systems<br />

Vesicle breakdown<br />

& degradation<br />

Autolysosome<br />

Figure 1 The autophagy pathway and its regulation. The autophagy pathway proceeds through a series<br />

of stages, including nucleation of the autophagic vesicle, elongation and closure of the autophagosome<br />

membrane to envelop cytoplasmic constituents, docking of the autophagosome with the lysosome, and<br />

degradation of the cytoplasmic material inside the autophagosome. Vesicle nucleation depends on a<br />

class III PI(3)K complex that contains various proteins (in light yellow box at right), as well as additional<br />

proteins that regulate the activity of this complex (such as rubicon, UVRAG, Ambra-1 and Bif-1). Vesicle<br />

elongation and completion involves the activity of two ubiquitin-like conjugation systems (light blue<br />

box at right). The autophagy pathway is positively regulated (green box at left) and negatively regulated<br />

(red box at left) by diverse environmental and immunological signals. TNF, tumor necrosis factor; PE,<br />

phosphatidylethanolamine; Gly, glycine; E1 and E2, ligases for ubiquitin-like conjugation systems.<br />

for lethal encephalitis in mice 24 . These studies collectively suggested<br />

a likely function for autophagy genes in pathogen defense in vivo. In<br />

parallel, many studies demonstrated a function for autophagy in vitro<br />

in defense against invading pathogens, including group A Streptococcus,<br />

Shigella flexneri, Mycobacterium tuberculosis, Salmonella typhimurium<br />

and Toxoplasma gondii 10,11,13 .<br />

Two recent studies have further confirmed an antimicrobial function<br />

for autophagy genes in host defense in vivo against intracellular<br />

pathogens and have identified previously unknown relationships among<br />

innate immune signaling, autophagy genes and potential autophagyindependent<br />

functions of autophagy genes. An innate microbial sensor,<br />

the peptidoglycan-recognition protein PRGP-LE, which recognizes bacterial<br />

diaminopimelic acid–type peptidoglycan, is crucial for autophagic<br />

control of Listeria monocytogenes infection in fly hemocytes and for host<br />

survival 25 . PRGP-LE-mediated resistance is associated with autophagy<br />

induction, requires the autophagy gene Atg5 and presumably involves<br />

the xenophagic degradation of bacteria (as bacteria are visualized in<br />

autophagosomes in wild-type animals). In this case, xenophagy targets<br />

cytoplasmic bacteria that have escaped from the endosome for<br />

envelopment in an autophagosome and destruction. In other cases<br />

in which a pathogen inside a vesicular structure is targeted (such as<br />

mycobacteria inside phagosomes, or salmonella inside vacuoles), the<br />

membrane dynamics involved are incompletely defined. It may be that<br />

an autophagosome can envelop the entire vesicular structure containing<br />

462 volume 10 number 5 may 2009 nature immunology<br />

P<br />

LC3


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Antigen presentation<br />

Antigen<br />

Autophagosome<br />

Isolation<br />

membrane<br />

MIIC<br />

MHC class II<br />

antigen presentation<br />

Activation of<br />

interferon production<br />

in plasmacytoid<br />

dendritic cells<br />

Isolation<br />

membrane<br />

Viral<br />

nucleic acid<br />

the pathogen or that the autophagic machinery somehow enhances<br />

phagolyosomal maturation. An alternative mechanism is discussed<br />

below for T. gondii.<br />

Although TLR signaling has been linked to autophagy induction (discussed<br />

below), the consequences of this autophagy induction have not<br />

been studied so far; therefore, the finding of for a function for PRGP-LE<br />

in listeria infection in flies is the first to demonstrate involvement of a<br />

cytoplasmic pattern-recognition receptor in the delivery of microbes to<br />

autophagosomes for degradation and in autophagy-mediated pathogen<br />

defense. One speculation is that other pattern-recognition receptors<br />

function similarly in innate immunity to target diverse microbes<br />

or microbial products to the autophagosome. A related issue is whether<br />

autophagosomal targeting motifs, such as monoubiquitination or<br />

polyubiquitination, that can target cellular proteins and organelles to<br />

the autophagosome through the adaptor protein SQSTM1 (p62) 26 also<br />

target microbial products. It will be useful to determine how the selective<br />

delivery of microbes, through cytoplasmic recognition by innate sensors<br />

and/or ubiquitination, may result in the selective removal or killing of<br />

pathogens by xenophagy, the activation of innate immune signaling<br />

and/or the generation of antigenic peptides from these pathogens for<br />

presentation to T cells.<br />

In mice, phagocytic cell–specific deletion of Atg5 results in greater<br />

susceptibility to infection with two different types of intracellular pathogens,<br />

the bacterium L. monocytogenes and the protozoan T. gondii 27 .<br />

Such work has provided important confirmation that autophagy genes<br />

are involved in antimicrobial host defense in mammals. However, in<br />

contrast to studies in the fly, in which autophagy is believed to degrade<br />

bacteria that invade professional phagocytes called hemocytes, Atg5 is<br />

Virus<br />

Isolation<br />

membrane<br />

Autophagosome<br />

Endosome<br />

TLR7<br />

Type 1 IFN<br />

AutophagosomeLysosome<br />

Suppression of<br />

interferon production<br />

in fibroblasts<br />

and macrophages<br />

Damaged<br />

mitochondria<br />

Reactive<br />

oxygen<br />

species<br />

RIG-I-like<br />

receptor signaling<br />

Autolysosome<br />

Reactive<br />

oxygen<br />

species<br />

Autophagosome<br />

Cell<br />

survival and death<br />

‘decisions’<br />

Pathogen degradation<br />

Isolation<br />

membrane<br />

Autophagosome<br />

proposed to control T. gondii by a mechanism<br />

independent of autophagosome formation.<br />

Published work suggests that IFN-γ-mediated<br />

control of T. gondii involves the localization<br />

of immunity-related GTPases to the parasitophorous<br />

vacuole, vacuolar membrane<br />

stripping, and either the induction of green<br />

fluorescent protein–LC3–positive vesicular<br />

structures that localize together with GTPases 28<br />

or the formation of autophagic membranes<br />

near the damaged parasitophorous vacuoles 29 .<br />

In Atg5 –/– macrophages, the GTPase IIGP1<br />

(Irga6) fails to localize to the parasitophorous<br />

vacuole, which results in a lack of disruption<br />

of parasitophorous vacuole membranes. Thus,<br />

the autophagy machinery, or at least Atg5, may<br />

be critical in the destruction of a parasitecontaining<br />

vesicular structure through the<br />

recruitment of immunity-related GTPases to<br />

the parasitophorous vacuole.<br />

As for T. gondii–infected mouse astrocytes<br />

28 , autophagosomal membranes are not<br />

found surrounding the damaged parasitophorous<br />

vacuole in wild-type macrophages,<br />

which suggests that the GTPase-recruitment<br />

function of Atg5, rather than autophagic<br />

entrapment of T. gondii, may be the main<br />

mechanism by which Atg5 functions ‘downstream’<br />

of IFN-γ to mediate intracellular<br />

resistance to this parasite. Studies using<br />

real-time microscopy have documented that<br />

killing of T. gondii is not consistently associated<br />

with localization of LC3 to the parasi-<br />

tophorous vacuole 30 . These data collectively provide support for a<br />

model in which an autophagy gene, Atg5, is involved in host defense<br />

independently of autophagosome formation. An important question<br />

is whether recruitment of GTPase to the parasitophorous vacuole<br />

requires autophagy proteins other than Atg5 and how Atg5, and perhaps<br />

the autophagic machinery, mediates this trafficking event.<br />

Immunity-related GTPases and autophagy<br />

Several different immune signals positively regulate autophagy, including<br />

PKR, TLRs, tumor necrosis factor, the CD40–CD40 ligand interaction,<br />

IFN-γ and the immunity-related GTPases 10,13,27,31,32 , whereas T helper<br />

type 2 cytokines (IL-4 and 1L-13) negatively regulate autophagy 33 . One<br />

of the most rapidly evolving areas of research centers on the relationship<br />

between autophagy and immunity-related GTPases (IRGs or p47<br />

GTPases), a family of proteins crucial for IFN-γ-mediated resistance<br />

to intracellular pathogens 34 . Although only the expression of mouse<br />

p47 GTPases seems to be regulated by IFN-γ, both mouse (LRG47) and<br />

human (Irgm1) p47 GTPases are thought to be required for IFN-γinduced<br />

autophagy and antimycobacterial activity in macrophages 31,32 .<br />

Furthermore, delivery of T. gondii to autophagosomes in macrophages<br />

may occur in an Irgm3 (IGTP)-dependent way 29 . Thus, a consensus<br />

is emerging that p47 GTPases are important in the IFN-γ-dependent<br />

autophagic elimination of intracellular pathogens. The precise<br />

mechanisms by which p47 GTPases exert this effect, however, remain<br />

unknown. One theory is that their localization to pathogen-containing<br />

phagosomes or vacuoles somehow facilitates damage to the pathogencontaining<br />

compartment with subsequent targeting of pathogen to the<br />

autophagosome.<br />

nature immunology volume 10 number 5 may 2009 463<br />

Lysosome<br />

Lymphocyte<br />

development<br />

Figure 2 Immunological processes involving autophagy. The colored boxes present five immunological<br />

processes that, on the basis of data available at present, involve classical autophagy. MIIC, MHC class<br />

II loading compartment.<br />

REVIEw


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

REVIEw<br />

Beyond the proposed involvement of p47 GTPases in the autophagic<br />

control of intracellular pathogens, evidence suggests potentially more<br />

complex crosstalk between this family of immune effectors and<br />

autophagy. As noted above, the autophagy protein Atg5 is needed<br />

for recruitment of the GTPase Irga6 to the T. gondii parasitophorous<br />

vacuole 27 . Thus, autophagy proteins can function ‘upstream’ of p47<br />

GTPases to regulate their trafficking to pathogen-containing compartments.<br />

As Irga6 that is associated with the parasitophorous vacuole<br />

is in an active GTP-bound state and Irga6 in uninfected cells is in an<br />

inactive GDP-bound state 35 , one obvious question is whether the<br />

effects of Atg5 on Irga6 localization occur through regulation of the<br />

GDP- or GTP-bound state of the GTPase. Another question is whether<br />

the Atg5-Atg12-Atg16L1 complex enhances the ability of Irga6, and<br />

perhaps other p47 GTPases, to form multimers on intracellular membranes.<br />

Furthermore, it is not yet known whether pathogen-containing<br />

phagosomes, in addition to parasitophorous vacuoles, may be sites of<br />

autophagy gene–dependent recruitment of p47 GTPase. Along these<br />

lines, it will be useful to determine whether Atg5 and other autophagy<br />

genes are required for the recruitment of p47 GTPases to mycobacteriacontaining<br />

phagosomes.<br />

A somewhat paradoxical function for Irgm1 in negatively regulating<br />

IFN-γ-induced autophagy has also been proposed. Mice deficient in<br />

Irgm1 have less proliferation of mature effector CD4 + T cell populations<br />

in the presence of IFN-γ 36 . This diminished expansion seems to be due<br />

to the induction of a caspase-independent cell death program that is<br />

blocked by PI(3)K inhibitors such as wortmanin or LY294002 or small<br />

interfering RNA specific for the autophagy gene encoding Beclin 1. Thus,<br />

Irgm1 may have dual interactions with the autophagy pathway that act<br />

in synergy to promote IFN-γ antimicrobial interactions; Irgm1 may signal<br />

autophagic sequestration of intracellular pathogens in macrophages<br />

while simultaneously limiting IFN-γ-induced autophagic death of effector<br />

T lymphocytes. That suggests the possibility of divergent functions<br />

for Irgm1 and potentially other p47 GTPases in autophagy regulation<br />

in different cell types. At present, it is unknown how Irgm1 might negatively<br />

regulate IFN-γ-induced autophagy to preserve activated CD4 +<br />

T cell populations. Furthermore, the possibility has not yet been formally<br />

excluded that the greater vacuolization in Irgm1 –/– lymphocytes<br />

is a function of aberrant membrane trafficking rather than enhanced<br />

autophagic flux. As reviewed elsewhere, whether autophagy is truly<br />

involved in mediating cell death also remains controversial 21 .<br />

TLRs and autophagy<br />

TLRs and the autophagy pathway intersect at many different levels 37 :<br />

TLRs can regulate autophagy induction 38,39 , the autophagy machinery<br />

can be used to deliver viral genetic material to endosomal TLRs for efficient<br />

induction of type I interferon 40 , and TLRs may act in the recruitment<br />

of autophagy proteins to phagosomal membranes 16 . Initially it<br />

was shown that bacterial lipopolysaccharide (LPS) induces autophagy<br />

through its cognate receptor TLR4 in macrophages by a mechanism<br />

that requires the TLR adaptor TRIF but not the adaptor MyD88 and<br />

requires the ‘downstream’ molecules receptor-interacting protein 1 and<br />

p38 mitogen-activated protein kinase 38 . Subsequently, another group<br />

confirmed induction of autophagy by LPS-TLR4 and also reported<br />

induction of autophagy by single-stranded RNA, poly(I:C) and imiquimod<br />

through TLR3 or TLR7 39 . TLR7-dependent induction of<br />

autophagy was found to be MyD88 dependent, and another study has<br />

proposed that in response to TLR activation, both TRIF and MyD88<br />

trigger autophagy through a direct interaction with Beclin 1 (ref. 41).<br />

However, it is not yet known how this interaction stimulates autophagy,<br />

and further studies are needed to more fully elucidate the molecular<br />

mechanism(s) by which TLR signaling activates autophagy.<br />

It is noteworthy that more robust TLR-dependent stimulation of<br />

autophagy in RAW264.7 mouse macrophages than in primary macrophages<br />

has been reported 39 and that autophagy induction has not<br />

been reported in primary fetal liver–derived macrophages treated with<br />

LPS or several other TLR stimuli 42 . These primary fetal liver–derived<br />

macrophages, however, are able to mount an autophagic response to<br />

starvation or invasive bacteria, which indicates that this lack of responsiveness<br />

to TLR stimuli does not reflect a deficiency in their general<br />

ability to undergo stress-induced autophagy. These data conflict with<br />

the data outlined above showing positive regulation of autophagy by<br />

TLR signaling. One possibility is that the function of TLR signaling in<br />

autophagy induction has been examined over different time frames in<br />

different studies. Alternatively, there may be TLR-independent mechanisms<br />

for the induction of autophagy in phagocytic cells, the regulation<br />

of autophagy by TLRs may be cell type specific or the function of TLRs<br />

may be tightly regulated by the differentiation state of the macrophage.<br />

The physiological importance of TLR regulation of autophagy requires<br />

further studies in primary cells, in cells other than macrophages and<br />

dendritic cells and, perhaps most importantly, in different in vivo models<br />

of microbial infection.<br />

Although the physiological function of TLR-mediated induction of<br />

autophagy is not yet known, it seems reasonable to speculate that it<br />

may function either in the autophagic control of intracellular pathogens<br />

and/or in other autophagy gene–dependent functions in immunity.<br />

As noted above, the autophagic control of L. monocytogenes in flies<br />

requires another type of pattern-recognition receptor, PGRP-LE 25 ; a<br />

critical question is whether such findings represent a general paradigm<br />

that applies to other families of pattern-recognition receptors in mammalian<br />

hosts, including TLRs and Nod-like receptors. At least in vitro,<br />

LPS stimulation of macrophages (which activates TLR4) or TLR7 activation<br />

results in greater localization of M. tuberculosis in autophagosomes<br />

and lower mycobacterial survival 38,39 , which raises the possibility that<br />

TLR induction of autophagy may similarly participate in autophagic<br />

control of intracellular pathogens in vivo. It is also possible that TLR<br />

induction of autophagy represents a type of feed-forward mechanism<br />

for priming the autophagy machinery to deliver viral nucleic acids to<br />

endosomal TLRs to activate innate immune signaling. Furthermore,<br />

another unexplored possibility is that induction of autophagy by TLRs<br />

or pattern-recognition receptors enhances autophagy-mediated presentation<br />

of endogenous antigens to major histocompatibility complex<br />

(MHC) class II loading compartments (Fig. 2).<br />

Studies also suggest that TLR signaling may not only induce classical<br />

autophagy but also recruit autophagy proteins to the phagosomal<br />

membrane and deliver phagocytosed material to the lysosome (Fig. 3).<br />

Engagement of TLRs during the phagocytosis of LPS-coated beads or<br />

zymosan particles induces the rapid recruitment of Beclin 1 and LC3<br />

to phagosomal membranes in an Atg5- and Atg7-dependent way 16 .<br />

Furthermore, proteomic analyses of latex bead–containing phagosomes<br />

from macrophages shows enhanced recruitment of LC3 to membranes<br />

in response to autophagy inducers 43 . The translocation of Beclin<br />

1 and LC3 to phagosomal membranes in macrophages treated with<br />

LPS-coated beads or zymosan particles is associated with enhanced<br />

phagosome fusion with lysosomes but not with detectable doublemembraned<br />

structures characteristic of autophagosomes 16 . Thus, it<br />

is postulated that TLR signaling, through recruitment of autophagy<br />

proteins to the phagosome, may constitute another mechanism for<br />

expediting lysosomal microbial elimination. Similar to the failure to<br />

detect autophagosomes surrounding T. gondii parasitophorous vacuoles<br />

in some studies 27,28,30 , it is not yet known whether the elimination<br />

of pathogen-containing phagosomes is truly a process that requires<br />

autophagy proteins but not the formation of autophagosomes or,<br />

464 volume 10 number 5 may 2009 nature immunology


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Figure 3 Immunological processes involving autophagy genes. The colored boxes present six<br />

immunologic processes that involve various autophagy genes but that are not yet proven to involve<br />

classical autophagy.<br />

alternatively, whether autophagosomes are sufficiently evanescent to<br />

have escaped detection in ultrastructural studies so far. Nevertheless,<br />

the findings obtained with T. gondii and with TLR signaling, phagocytosis<br />

and autophagy proteins do raise the possibility that components<br />

of the autophagy machinery may be involved in membrane-mediated<br />

events that do not involve the classical delivery of cytoplasmic constituents<br />

or pathogens to the lysosome (Fig. 3).<br />

Autophagy genes and cytokine secretion<br />

The first connection between autophagy and innate immune signaling<br />

emerged after the demonstration that Atg5 is essential for the production<br />

of type I interferon in plasmacytoid dendritic cells infected with vesicular<br />

stomatitis virus by a mechanism presumed to involve autophagymediated<br />

delivery of viral genetic material to endosomal TLRs 40 (Fig.<br />

2). Somewhat paradoxically, several studies have shown that absent or<br />

hypomorphic expression of autophagy genes in certain cell types can<br />

result in enhanced production of type I interferon or other cytokines,<br />

including proinflammatory molecules such as IL-1β and IL-18, or adipocytokines,<br />

such as leptin and adiponectin 42,44–46 (Fig. 3). As discussed<br />

below, enhanced secretion of such molecules may be involved in certain<br />

aspects of Crohn’s disease in mice deficient in the expression of the<br />

autophagy gene Atg16l1. Thus, the autophagic machinery may serve a<br />

dual function in innate immune signaling, acting not only to stimulate<br />

antiviral type I interferon responses in dendritic cells but also to ensure<br />

homeostatic balance by preventing excess innate immune activation in<br />

other cell types.<br />

Two distinct models have been proposed to explain the higher production<br />

of type I interferon in autophagy gene–deficient fibroblasts and<br />

primary macrophages (Fig. 3). Enhanced production of type I interferon<br />

in Atg5 –/– and Atg7 –/– mouse embryonic fibroblasts in response<br />

to infection with vesicular stomatitis virus or double-stranded RNA has<br />

REVIEw<br />

been reported 44 . In this model, the Atg5-Atg12<br />

conjugate can bind to cytoplasmic viral sensors,<br />

the retinoic acid–inducible helicases RIG-1 and<br />

Mda5 and their adaptor protein IPS-1, to suppress<br />

the activity of such helicases in stimulating<br />

the production of type I interferon. Thus, it<br />

is possible that the autophagic machinery may<br />

suppress innate immune signaling by direct<br />

inhibitory interactions with these helicases<br />

and their adaptor proteins. Of note, however, a<br />

completely different, but not mutually exclusive,<br />

mechanism for the higher production of type I<br />

interferon in Atg5-deficient mouse embryonic<br />

fibroblasts and primary macrophages has also<br />

been proposed 45 . Dysfunctional mitochondria<br />

have been shown to accumulate in the absence<br />

of Atg5, leading to higher concentrations of<br />

IPS-1 and a reactive oxygen species–dependent<br />

increase in retinoic acid–inducible helicase signaling<br />

and production of type I interferon in<br />

response to infection with vesicular stomatitis<br />

virus (Fig. 3). Therefore, it is also possible<br />

that the homeostatic functions of constitutive<br />

autophagy, especially mitochondrial turnover<br />

and quality control, indirectly regulate the production<br />

of type I interferon.<br />

It will be useful to determine whether this<br />

enhanced response to viral infection is present<br />

in cells deficient in autophagy genes that act<br />

in other steps of the autophagy pathway and<br />

whether the underlying mechanism reflects a deficiency of the cellular<br />

process of autophagy or alternative, as-yet-undefined functions of<br />

autophagy proteins in regulating interferon signaling. Because Atg5deficient<br />

plasmacytoid dendritic cells with impaired type I interferon<br />

responses are more sensitive to infection with vesicular stomatitis virus<br />

and Atg5-deficient mouse embryonic fibroblasts and primary macrophages<br />

with enhanced type I interferon responses are more resistant<br />

to such infection, another critical question is how potential cell type–<br />

specific differences in autophagy gene–dependent regulation of type I<br />

interferon signaling may interact in vivo to determine the fate of viral<br />

infection. It is not yet clear whether the autophagic machinery truly<br />

has a direct immunosuppressive function that increases susceptibility<br />

to viral infection by inhibiting cytokine release or if it merely functions<br />

as a negative feedback mechanism that provides a brake on unneeded<br />

innate antiviral signaling.<br />

Another study has shown an important function for the autophagy<br />

gene Atg16l1 in regulating the secretion of proinflammatory cytokines<br />

(Figs. 3,4). The Nod-like receptor protein cyropryrin (also called NALP<br />

or NLRP3) forms a complex known as the ‘inflammasome’, which contains<br />

the adaptor protein ASC (apoptosis-associated speck-like protein<br />

containing a caspase-activating and caspase-recruitment domain) and<br />

caspase 1 and is responsible for the processing of pro-IL-1β to its mature,<br />

secreted form 47,48 . Macrophages from mice lacking Atg16l1 produce<br />

more of the proinflammatory cytokines IL-1β and IL-18 after endotoxin<br />

stimulation of TLR4 (ref. 42). Similarly, mouse chimeras engrafted<br />

with Atg16l1 –/– fetal liver hematopoietic progenitors have higher serum<br />

concentrations of IL-1β and IL-18 after treatment with dextran sodium<br />

sulfate; this enhanced cytokine secretion probably contributes to pathology,<br />

as treatment with antibody to IL-1β and antibody to IL-18 results<br />

in lower susceptibility of Atg16l1 –/– chimeric mice to dextran sodium<br />

sulfate–induced colitis.<br />

nature immunology volume 10 number 5 may 2009 465


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

REVIEw<br />

The precise mechanisms responsible for the enhanced secretion of<br />

IL-1β and IL-18 by Atg16l1-deficient cells are not yet clear, but TRIFdependent<br />

activation of caspase 1 and consequent enhanced processing<br />

of IL-1β in Atg16l1 –/– macrophages has been noted. In addition, similar<br />

to results obtained with virus-infected Atg5-deficient cells 45 , higher production<br />

of reactive oxygen species in LPS-stimulated Atg16l1 –/– macrophages<br />

has been found 42 . Thus, a gain of function in cytokine release<br />

may be obtained with deficiency in both Atg5 and Atg16l1, and in both<br />

cases it may result directly from effects of the autophagy machinery on<br />

signaling pathways that regulate cytokine production and/or indirectly<br />

from effects of autophagy on the cellular redox state.<br />

Another study has emphasized the involvement of Atg16L1 in regulating<br />

cytokine expression in a specialized intestinal epithelial cell that is<br />

important in innate immunity, the Paneth cell (Figs. 3,4). Substantially<br />

enhanced gene expression of acute-phase reactants, PPAR signaling molecules<br />

and adipocytokines such as leptin and adiponectin was observed<br />

after transcriptional profiling by microarray of Paneth cell RNA from<br />

mice with a hypomorphic Atg16l1 allele 46 . Similar alterations have not<br />

been found in thymocytes from these mice, which suggests that Atg16l1<br />

deficiency may upregulate cytokine production in a cell type–specific<br />

way. Furthermore, this study has emphasized the potential importance<br />

of the autophagic machinery in Paneth cells, as specialized cells of the<br />

innate immune system, in regulating the expression of molecules with a<br />

broad range of systemic pro- and anti-inflammatory effects. The mechanism<br />

by which deficiency in an autophagy gene results in these transcriptional<br />

changes in Paneth cells remains unknown. It is also unknown<br />

whether the polymorphism in human ATG16L1 that is associated with<br />

Crohn’s disease (ATG16L1 T300A) results in similar changes and, if so,<br />

whether such changes contribute to the pathophysiology of Crohn’s<br />

diseases and/or other types of inflammatory diseases.<br />

Autophagy, antigen presentation and thymic selection<br />

There is growing evidence that autophagy may be involved in the MHC<br />

class II presentation of certain endogenously synthesized peptides 10,49<br />

(Fig. 2). The involvement of autophagy in MHC class I–restricted antigen<br />

presentation remains more speculative 50 ; one paper has suggested<br />

involvement of Atg5 in MHC class I antigen presentation 51 . Although<br />

it is clear that autophagosomes can deliver peptides to MHC class II<br />

loading compartments for presentation to CD4 + T cells 52,53 , it is not yet<br />

known whether this represents a major pathway for antigen presentation<br />

during the generation of adaptive immune responses in physiological<br />

conditions in primary cells or in vivo. It will therefore be necessary to<br />

assess the involvement of autophagy in the function of professional<br />

antigen-presenting cells during immune responses in vivo. Regardless<br />

of the results, the specific targeting of antigens to autophagosomes by<br />

fusion with the LC3 autophagy protein may represent an effective vaccine<br />

strategy for enhancing CD4 + T cell responses; for example, a 20-fold<br />

enhancement of MHC class II presentation to CD4 + T cell clones has<br />

been noted when influenza virus matrix protein is targeted to autophagosomes<br />

by fusion with LC3 (ref. 53).<br />

One notable physiological context in which autophagy pathway–<br />

dependent endogenous loading of MHC class II may be important<br />

is in shaping the T cell repertoire during thymic selection. Thymic<br />

epithelial cells that function in positive and negative selection show<br />

constitutive autophagy, as demonstrated by the presence of Atg5dependent<br />

green fluorescent protein–LC3 dots 54,55 . Normal thymic T<br />

cell selection requires Atg5 expression in the stromal cells in thymic<br />

allografts 55 . Experiments with thymic transplantation have shown that<br />

the selection of certain MHC class II–dependent TCRs but not MHC<br />

class I–dependent TCRs requires Atg5. Furthermore, mice with Atg5 –/–<br />

thymic implants develop autoreactive CD4 + T cells, as measured both<br />

by the proportion of CD62L lo peripheral cells and the development of<br />

inflammatory infiltrates in many organs, including the intestine. This<br />

last observation supports speculation that abnormal negative selection<br />

might explain the association between a polymorphism in the autophagy<br />

protein Atg16L1 and susceptibility to Crohn’s disease (discussed below;<br />

Fig. 4). The development of autoreactive CD4 + T cells in mice with<br />

Atg5 –/– thymic implants is thought to reflect a function of Atg5 and,<br />

presumably, the autophagy pathway in controlling the repertoire of self<br />

peptides presented by thymic epithelial cells that are responsible for<br />

normal thymic selection and the generation of T cell tolerance. However,<br />

autophagy genes other than Atg5 have not yet been analyzed in similar<br />

studies, and thus the function of the cellular process of autophagy in<br />

thymic selection remains to be defined. It will be useful to determine<br />

whether the mechanism responsible for the function of Atg5 in thymic<br />

selection reflects the generation of peptides or a function for nutritional,<br />

survival or transcriptional responses regulated by Atg5 or other<br />

autophagy genes.<br />

Another important question is whether autophagy, in either the<br />

antigen-presenting cell or the cell donating antigen, is involved in efficient<br />

antigen cross-presentation. Indeed, one study has shown that small<br />

interfering RNA–mediated knockdown of either Beclin 1 or Atg12 in<br />

antigen-donor tumor cells results in less cross-presentation 56 . The<br />

authors postulate a mechanism involving direct involvement of the<br />

autophagosome as a carrier of protein antigens from tumor cells. More<br />

studies are needed to confirm this mechanism, as well as explore its<br />

potential clinical implications for cancer vaccines.<br />

An additional potential link between autophagy and cross-presentation<br />

has emerged from the observations that autophagy is required for<br />

the removal of apoptotic corpses in mouse embryoid body and chick<br />

retinal development 57,58 . In these settings, autophagy is required for<br />

dying cells to have sufficient energy to generate the engulfment signals<br />

necessary for the clearance of such corpses by phagocytes. Therefore, one<br />

interesting, as-yet-unexplored theory is that autophagy in dying cells<br />

may also be required for the generation of signals that induce efficient<br />

uptake of cell corpses for subsequent cross-presentation. This hypothesis<br />

is potentially important for understanding and enhancing the immunogenicity<br />

of cell-based tumor vaccines. The function of autophagy in<br />

the clearance of dead cells during normal tissue turnover might also<br />

contribute to peripheral tolerance.<br />

Autophagy genes and lymphocytes<br />

Lymphocytes undergo extensive rearrangement of cytoplasm and<br />

organelles during the selection, expansion and contraction of antigenspecific<br />

clones and frequently confront life-or-death decisions during<br />

their development and activation. Therefore, it is not unexpected that<br />

autophagy might have critical cell-intrinsic functions in lymphocyte<br />

biology. Indeed, several studies have shown that deletion of autophagy<br />

genes in T lymphocytes and B lymphocytes alters lymphocyte development,<br />

survival and/or function (Fig. 2).<br />

The thymus is a site of constitutive autophagy 54,55 , and the expression<br />

of Beclin 1 is regulated during T cell and B cell development and T cell<br />

activation 59 , which provides indirect evidence of potential links between<br />

autophagy and lymphocyte development. Of note, loss of Atg5 or Atg7<br />

impairs the survival and proliferation of mature T lymphocytes in<br />

vivo 60–62 , and Atg5 is required in B lymphocytes for the survival of developing<br />

pre–B cells in the bone marrow and of mature B-1a cells in the<br />

periphery 63 . On the basis of studies of lethally irradiated mice reconstituted<br />

with Atg5 –/– fetal liver progenitors, it seems that macrophages, plasmacytoid<br />

dendritic cells, neutrophils, erythrocytes and immature T cells<br />

develop and survive normally in the absence of Atg5. Therefore, a critical<br />

question is how deletion of Atg5 and Atg7 confers lymphocyte-specific<br />

466 volume 10 number 5 may 2009 nature immunology


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Abnormal secretion into<br />

intestinal lumen<br />

Abnormal<br />

granules<br />

mRNAs for adipocytokines<br />

and acute phase reactants<br />

Abnormal<br />

small<br />

vesicles<br />

Autophagy gene mutation<br />

(Atg5, Atg7, Atg16l1)<br />

in Paneth cells<br />

Degenerating<br />

mitochondria<br />

Intestinal inflammation<br />

Release of autoimmune T cells<br />

Abnormal<br />

thymus<br />

gland<br />

Abnormal<br />

negative<br />

selection<br />

Autophagy gene mutation<br />

(Atg5) in<br />

thymic epithelial cells<br />

and lymphocyte lineage–specific (for example, T cell and B-1a cell but<br />

not B2 B cell) effects on development and survival.<br />

It is not yet completely clear whether the effects of the deletion<br />

of Atg5 and Atg7 on lymphocyte development are mediated through<br />

autophagy (as proposed in Fig. 2) or other effects of these genes; studies<br />

of mice with mutations in genes whose products act at other stages in<br />

the autophagy pathway may be helpful in elucidating this. One mechanism<br />

postulated for the involvement of Atg7 and Atg5 in the survival of<br />

mature T cells involves a classical autophagy function in the clearance<br />

of mitochondria and consequent prevention of the accumulation of<br />

reactive oxygen species and imbalance of the expression of pro- and<br />

antiapoptotic proteins 60,62 . Notably, the developmental pattern of<br />

Beclin 1 in B lymphocytes, similar to that of the antiapoptotic protein<br />

Bcl-2, shows downregulation at the transition from pro–B cell to<br />

pre–B cell 58 . Thus, it will be useful to determine whether suppression<br />

of autophagy, in addition to apoptosis, is necessary for the physiological<br />

cell death that occurs at this stage in B lymphocyte development<br />

and whether null mutations in autophagy genes result in excessive cell<br />

death that is incompatible with efficient development of pre–B cells<br />

in the bone marrow.<br />

A controversial area is whether autophagy is exclusively a prosurvival<br />

pathway in lymphocytes or whether it also can be used as a death pathway.<br />

As noted before, deletion of Atg5 or Atg7 results in the impaired<br />

survival of certain populations of B cells or T cells 60–63 , which suggests a<br />

major function in lymphocyte survival. However, as discussed earlier, it<br />

has been postulated that IFN-γ may induce Irgm1-regulated autophagic<br />

cell death in CD4 + T cells 36 . In addition, knockdown of the gene encoding<br />

Beclin 1 or Atg7 results in less death of CD4 + T cells induced by<br />

growth-factor withdrawal 64 , and autophagy genes are required for the<br />

death of bystander CD4 + T cells triggered by the human immunodeficiency<br />

virus envelope protein 65 . Thus, a consensus is growing that<br />

autophagy may mediate activation-induced death of CD4 + T cells.<br />

Failure to control<br />

intracellular<br />

bacterial replication<br />

Increased ROS production<br />

Increased LPS-induced<br />

IL-1β and IL-18 secretion<br />

Autophagy gene mutation<br />

(Atg7, Atg16l1)<br />

in macrophages<br />

However, several caveats must be considered when determining<br />

whether autophagy is truly a cause of cell death (death by autophagy)<br />

rather than simply being associated with cell death (death with<br />

autophagy) 21 . An example of some of the complexities in determining<br />

the contribution of autophagy to T cell death is provided by a<br />

study examining the inter-relationships among signaling by the death<br />

domain protein FADD, autophagy and the death of primary CD8 + T<br />

cells 66 . Transgenic expression of dominant negative FADD in CD8 +<br />

T cells leads to enhanced autophagy and cell death in response to<br />

mitogenic signals, which can be prevented by either targeting of the<br />

autophagy pathway (as with pharmacological inhibitors, dominant<br />

negative Vps34 or short hairpin RNA specific for Atg7) or by use of the<br />

necropoptosis inhibitor Nec-1, which blocks the kinase RIPK1. Thus,<br />

FADD signaling may somehow dampen the autophagy response to<br />

prevent RIPK1-dependent necroptotic cell death, but it is as yet unclear<br />

whether autophagy is a true cell death pathway for proliferating CD8 +<br />

T cells or whether the autophagic machinery functions in signaling<br />

necroptosis. If autophagy does prove to be a cell death pathway for T<br />

cells, one possible explanation for the apparently conflicting observations<br />

is that basal autophagy is required for T cell survival, whereas<br />

unrestrained autophagy, which may occur in response to IFN-γ stimulation<br />

or compromise of FADD signaling, may be deleterious.<br />

The stress-activated signaling molecules Jnk1 and Jnk2 may also be<br />

involved in the dual integration of autophagy signaling with T cell differentiation,<br />

activation and function. Jnk1 and Jnk2 both serve important<br />

but nonredundant functions in CD4 + and CD8 + T cells 67–69 . Notably, several<br />

studies have indicated that Jnk kinases are important in autophagic<br />

signaling. Although one study has reported more green fluorescent<br />

protein–LC3 dots in T cells lacking Jnk1 or Jnk2 (ref. 64), another study<br />

has found that Jnk1 positively regulates autophagy by a mechanism that<br />

involves phosphorylation of Bcl-2 and disruption of the Bcl-2–Beclin 1<br />

complex, but Jnk2 does not 70 . The last study showed that Jnk2-deficient<br />

nature immunology volume 10 number 5 may 2009 467<br />

ROS<br />

IL-1β<br />

IL-18<br />

ROS<br />

Altered gut<br />

microbiota<br />

Natural<br />

antibody<br />

B1a<br />

B cell<br />

Atg5 mutation<br />

B cell<br />

precursor<br />

Altered<br />

intestinal<br />

immunity<br />

Altered T cell<br />

homeostasis<br />

Atg5, Atg7,<br />

Irgm1 mutations<br />

T cell<br />

Autophagy gene mutation<br />

(Atg5, Atg7, Irgm1)<br />

in lymphocytes<br />

REVIEw<br />

Figure 4 Possible mechanisms by which alterations in autophagy may be involved in the pathogenesis of human Crohn’s disease. The colored boxes present<br />

four potential mechanisms by which the mutation of an autophagy gene could contribute to Crohn’s disease, proposed on the basis of studies of model<br />

organisms in which various autophagy genes have been mutated in various cell types (boxes at bottom). So far, only ATG16L1 and IRGM1 have been linked to<br />

susceptibility to Crohn’s disease. ROS, reactive oxygen species.


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

REVIEw<br />

mouse embryonic fibroblasts have enhanced Jnk1 signaling and hyperactive<br />

autophagy 70 . Therefore, it is plausible that the lack of autophagy<br />

in Jnk1-deficient mice (or perhaps even enhanced autophagy in Jnk2deficient<br />

mice) may contribute to alterations in T cell differentiation.<br />

Further assessment of the function of Jnk signaling in the regulation of<br />

autophagy in T cells and the effects of Jnk signaling–mediated autophagy<br />

regulation in T cell differentiation is clearly needed.<br />

Autophagy genes and Crohn’s disease<br />

One of the most exciting recent developments in the field of autophagy<br />

genes and immunity was catalyzed by genome-wide association studies<br />

identifying strong links between polymorphisms in two human<br />

autophagy genes, IRGM (called ‘IRGM1’ here) and ATG16L1, and susceptibility<br />

to inflammatory bowel disease 71,72 (Fig. 4). A polymorphism<br />

in IRGM1 is associated with both Crohn’s disease and ulcerative colitis;<br />

this association is due to the absence of specific single-nucleotide<br />

polymorphisms associated with a 2.7-kilobase deletion upstream of the<br />

IRGM1 transcriptional start site 71,73,74 . This raises the possibility that<br />

Irgm1-mediated regulation of autophagy may contribute to susceptibility<br />

to inflammatory bowel disease; however, so far there are no studies<br />

that mechanistically link Irgm1 to features of inflammatory bowel disease<br />

in animal models, nor is it known whether the IRGM1 risk allele<br />

affects autophagy regulation.<br />

The Crohn’s disease–associated ATG16L1 risk allele encodes a protein<br />

with a threonine-to-alanine substitution (T300A) in the carboxyterminal<br />

domain containing tryptophan–aspartic acid dipeptide (WD)<br />

repeats. Although this domain of Atg16L1 is not conserved in yeast, is<br />

not required for the formation of multimeric complexes with Atg5-<br />

Atg12 and is not required for starvation-induced autophagy, the Atg16L1<br />

T300A mutant protein may have lower stability and may be defective in<br />

localizing the autophagic machinery to intracellular bacteria 75,76 . Most<br />

notably, two studies of mice with genetically engineered mutations in<br />

Atg16l1 have identified critically important functions of this autophagy<br />

protein in innate immunity 42,46 .<br />

Mice with deletion of the coiled-coil domain of Atg16L1, which<br />

results in the lack of a functional protein, have been generated 42 . As<br />

reported for null mutations of Atg5 and Atg7, these mice die in the first<br />

day of life, presumably because of starvation or a bioenergetic crisis<br />

during the postnatal period. As noted above, studies with macrophages<br />

from these mice have shown enhanced cytokine responsiveness after<br />

stimulation with LPS (TLR4) or stimulation of other TLRs, and chimeric<br />

mice engrafted with Atg16l1 –/– hematopoietic progenitors from<br />

fetal liver have increased mortality, weight loss, colitis and serum concentrations<br />

of IL-1β and IL-18 in response to dextran sodium sulfate.<br />

This Atg16L1-dependent regulation of endotoxin-induced activation<br />

of inflammasomes demonstrates a previously unknown function for<br />

a component of the autophagic machinery in controlling inflammatory<br />

immune responses. A key question is whether the Atg16L1 T300A–<br />

encoding risk allele of ATG16L1 results in similar hyperexpression of<br />

proinflammatory genes and thus represents a potential mechanism by<br />

which this variant protein increases the risk for Crohn’s disease and<br />

potentially other inflammatory disorders (Fig. 4). It will also be useful<br />

to determine if this gain-of-function increase in cytokine release occurs<br />

with deficiency of additional autophagy proteins.<br />

Viable mice have been bred that are hypomorphic for Atg16L1 expression<br />

by means of intronic insertion of a gene-trap vector, which has<br />

allowed direct assessment of the effects of lower Atg16L1 expression<br />

on intestinal pathology in adult mice 46 (Fig. 4). As function is not<br />

completely abolished (but may be lower because of altered stability)<br />

by the T300A substitution of Atg16L1, this model of lowered mouse<br />

Atg16L1 protein expression may be relevant to humans carrying the<br />

allele encoding Atg16L1 T300A. Notably, in addition to the transcriptional<br />

changes in adipocytokines and other inflammatory regulators<br />

in Paneth cells described above, prominent histological abnormalities<br />

in Paneth cells of Atg16L1-hypomorphic mice have been reported that<br />

resemble changes in resected ileal tissue from patients with Crohn’s<br />

disease who are homozygous for the allele encoding Atg16L1 T300A.<br />

In both settings, Paneth cells show morphological defects in secretory<br />

granules and granule exocytosis, and in the Atg16L1-hypomorphic mice,<br />

deficient Paneth cell secretion of the antibacterial protein lysozyme has<br />

been noted in the small intestine.<br />

As Paneth cells function as a barrier to bacterial invasion (in part by<br />

secreting granule contents that contain antimicrobial peptides) and as<br />

regulators of intestinal inflammation 77,78 , the findings reported above<br />

suggest that defects in Atgl6L1 function in Paneth cells may contribute to<br />

the pathogenesis of Crohn’s disease. Studies of mice with intestinal epithelial<br />

cell–specific deletion of other autophagy genes, including Atg5 or<br />

Atg7, have demonstrated similar abnormalities in Paneth cells 46,79 . Thus,<br />

several components of the autophagic machinery that act at the membrane<br />

expansion-completion stage (such as Atg16L1, Atg5 and Atg7)<br />

are involved in the maintenance of Paneth cell function. At present, it is<br />

not yet known whether this function of Atg16L1, Atg5 and Atg7 reflects<br />

involvement of classical autophagy in Paneth cell function or perhaps an<br />

as-yet-uncharacterized function for the autophagy protein–conjugation<br />

machinery in granule exocytosis (Fig. 3).<br />

A model is thus emerging in which dysregulation of autophagy gene<br />

function may contribute to the pathogenesis of Crohn’s disease or other<br />

types of intestinal inflammatory disorders by several possible routes<br />

(Fig. 4). Lack of Atg5 in the thymus results in autoreactive CD4 + T cells<br />

and intestinal inflammatory infiltrates 55 ; lack of Atg16l1 in macrophages<br />

results in enhanced endotoxin-induced inflammatory signaling 42 , and<br />

lower expression of Atg16l1 in Paneth cells results in transcriptional<br />

alterations in molecules that regulate inflammation and in defects in<br />

granule exocytosis 46 . More speculatively, defects in other autophagy<br />

gene–dependent processes, such as the maintenance of normal numbers<br />

of B-1a cells that make natural antibody to bacterial carbohydrate<br />

antigens 62 or the survival of T cells 60–62 , might be involved in the pathogenesis<br />

of Crohn’s disease. Although much more work is needed to elucidate<br />

the precise effect of polymorphisms in IRGM1 and ATG16L1 on<br />

the pathogenesis of Crohn’s disease, these observations are beginning<br />

to delineate previously unknown functions for autophagy genes in the<br />

regulation of innate immunity.<br />

Therapeutic implications<br />

The many functions of autophagy and autophagy genes in immunity<br />

provide both opportunities and risks for manipulating autophagy therapeutically.<br />

For example, there is the opportunity to enhance CD4 + T<br />

cell–dependent vaccines by targeting antigens to the autophagic pathway.<br />

A notable idea that has arisen from the finding that autophagy is<br />

important in host resistance to viral, bacterial and parasitic infection is<br />

that pharmacological enhancers of autophagy may function as broadspectrum<br />

antimicrobial agents. Furthermore, enhancing autophagy<br />

might ameliorate Crohn’s disease if agents that overcome deficiencies<br />

in the function of Atg16L1 or Irgm1 can be identified. The potential<br />

beneficial effects of autophagy in immunity must be taken into consideration<br />

in the clinical development of agents for other diseases, such<br />

as cancer, that are targeted at autophagy inhibition. Although inhibition<br />

of autophagy-dependent cell survival may be beneficial in cancer<br />

therapy, the adverse potential immunological effects of such approaches<br />

include diminished negative selection of thymocytes, with consequent<br />

autoimmunity, enhanced secretion of proinflammatory cytokines by<br />

hyper-responsive innate immune cells, greater susceptibility to infection<br />

468 volume 10 number 5 may 2009 nature immunology


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

with intracellular pathogens and lower survival of B-1a cells and T cells.<br />

Perhaps additional understanding of the molecular basis of the diverse<br />

functions of autophagy genes in immunity will open new doors for<br />

rational intervention in a variety of diseases without untoward immunological<br />

consequences.<br />

Conclusion<br />

Studies of autophagy genes and immunity have identified important<br />

potential functions for autophagy and unexpected functions for<br />

autophagy proteins in the regulation of innate immunity and inflammation.<br />

Autophagy proteins are involved in immunity-related processes<br />

that may not involve the formation of autophagosomes, such as the<br />

recruitment of immunity-related GTPases to pathogen-containing compartments,<br />

TLR-mediated phagolysosomal maturation, the regulation<br />

of RNA helicases and the inflammasome, and the exocytosis of Paneth<br />

cell granules (Fig. 3). Converging evidence suggests that autophagy proteins,<br />

although critical for normal immune responses, may also function<br />

to prevent immunological hyper-responsiveness in certain cell types,<br />

such as macrophages, fibroblasts and Paneth cells; this may at least in<br />

part underlie the associations between polymorphisms in autophagy<br />

genes and Crohn’s disease. Another emerging theme is that deletion<br />

of autophagy genes can have different effects in different immune cell<br />

populations, which may reflect an effect of the milieu of differentiated<br />

immune cells on the function of autophagy genes and/or cell type–specific<br />

differences in the complex interaction between autophagy proteins<br />

and the specialized functions of various cells of the immune system.<br />

Distinguishing between the effects of autophagy proteins that are mediated<br />

by classical autophagy and those that are perhaps mediated by other<br />

non–autophagosome-dependent mechanisms may help elucidate how<br />

proteins of a primordial stress-response pathway may have functionally<br />

diversified to shape the immunological responses of complex higher<br />

eukaryotic organisms.<br />

ACKNOWLEDGMENTS<br />

We thank A. Diehl for medical illustration. Supported by the US National Institutes<br />

of Health (R01 CA074730, R01 CA096511, R01 AI054483, R01 AI065982 and U54<br />

AI057160 to H.W.V., and R01 AI151267 and R01 CA109618 to B.L.), the Howard<br />

Hughes Medical Institute (B.L.) and the Ellison Medical Foundation (B.L.).<br />

Published online at http://www.nature.com/natureimmunology/<br />

reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Levine, B. & Kroemer, G. Autophagy in the pathogenesis of disease. Cell 132, 27–42<br />

(2008).<br />

2. Mizushima, N., Levine, B., Cuervo, A.M. & Klionsky, D.J. Autophagy fights disease<br />

through cellular self-digestion. <strong>Nature</strong> 451, 1069–1075 (2008).<br />

3. Levine, B. Eating oneself and uninvited guests: autophagy-related pathways in cellular<br />

defense. Cell 120, 159–162 (2005).<br />

4. Kuma, A., Mizushima, N., Ishihara, N. & Ohsumi, Y. Formation of the approximately<br />

350-kDa Apg12-Apg5.Apg16 multimeric complex, mediated by Apg16 oligomerization,<br />

is essential for autophagy in yeast. J. Biol. Chem. 277, 18619–18625 (2002).<br />

5. Hara, T. et al. FIP200, a ULK-interacting protein, is required for autophagosome<br />

formation in mammalian cells. J. Cell Biol. 181, 497–510 (2008).<br />

6. Sun, Q. et al. Identification of Barkor as a mammalian autophagy-specific factor for<br />

Beclin 1 and class III phosphatidylinositol 3-kinase. Proc. Natl. Acad. Sci. USA 105,<br />

19211–19216 (2008).<br />

7. Itakura, E., Kishi, C., Inoue, K. & Mizushima, N. Beclin 1 forms two distinct phosphatidylinositol<br />

3-kinase complexes with mammalian Atg14 and UVRAG. Mol. Biol.<br />

Cell 19, 5360–5372 (2008).<br />

8. Zhong, Y. et al. Distinct regulation of autophagic activity by Atg14L and Rubicon<br />

associated with Beclin 1–phosphatidylinositol-3-kinase complex. Nat. Cell. Biol. 11,<br />

468–476 (2009).<br />

9. Matsunaga,K. et al. Two Beclin 1-binding proteins, Atg14L and Rubicon, reciprocally<br />

regulate autophagy at different stages. Nat. Cell Biol. 11, 385–396 (2009).<br />

10. Munz, C. Enhancing immunity through autophagy. Annu. Rev. Immunol. 27, 423–449<br />

(2009).<br />

11. Levine, B. & Deretic, V. Unveiling the roles of autophagy in innate and adaptive immunity.<br />

Nat. Rev. Immunol. 7, 767–777 (2007).<br />

12. Schmid, D. & Munz, C. Innate and adaptive immunity through autophagy. Immunity<br />

27, 11–21 (2007).<br />

REVIEw<br />

13. Orvedahl, A. & Levine, B. Eating the enemy within: autophagy in infectious diseases.<br />

Cell Death Differ. 16, 57–69 (2009).<br />

14. Klionsky, D.J. et al. Guidelines for the use and interpretation of assays for monitoring<br />

autophagy in higher eukaryotes. Autophagy 4, 151–175 (2008).<br />

15. Kuma, A., Matsui, M. & Mizushima, N. LC3, an autophagosome marker, can be incorporated<br />

into protein aggregates independent of autophagy: caution in the interpretation<br />

of LC3 localization. Autophagy 3, 323–328 (2007).<br />

16. Sanjuan, M.A. et al. Toll-like receptor signaling in macrophages links the autophagy<br />

pathway to phagocytosis. <strong>Nature</strong> 450, 1253–1257 (2007).<br />

17. Jackson, w.T. et al. Subversion of cellular autophagosomal machinery by RNA viruses.<br />

PLoS Biol. 3, 861–871 (2005).<br />

18. Suzuki, K., Noda, T. & Ohsumi, Y. Interrelationships among Atg proteins during<br />

autophagy in Saccharomyces cerevisiae. Yeast 21, 1057–1065 (2004).<br />

19. Matsui, Y. et al. Distinct roles of autophagy in the heart during ischemia and reperfusion:<br />

roles of AMP-activated protein kinase and Beclin 1 in mediating autophagy. Circ.<br />

Res. 100, 914–922 (2007).<br />

20. Codogno, P. & Meijer, A.J. Atg5: more than an autophagy factor. Nat. Cell Biol. 8,<br />

1045–1047 (2006).<br />

21. Kroemer, G. & Levine, B. Autophagic cell death: the story of a misnomer. Nat. Rev.<br />

Mol. Cell Biol. 9, 1004–1010 (2008).<br />

22. Liang, X.H. et al. Protection against fatal Sindbis virus encephalitis by beclin, a novel<br />

Bcl-2-interacting protein. J. Virol. 72, 8586–8596 (1998).<br />

23. Liu, Y. et al. Autophagy regulates programmed cell death during the plant innate<br />

immune response. Cell 121, 567–577 (2005).<br />

24. Orvedahl, A. et al. HSV-1 ICP34.5 Confers neurovirulence by targeting the beclin 1<br />

autophagy protein. Cell Host Microbe 1, 23–35 (2007).<br />

25. Yano, T. et al. Autophagic control of listeria through intracellular innate immune recognition<br />

in Drosophila. Nat. Immunol. 9, 908–916 (2008).<br />

26. Kim, P.K., Hailey, D.w., Mullen, R.T. & Lippincott-Schwartz, J. Ubiquitin signals<br />

autophagic degradation of cytosolic proteins and peroxisomes. Proc. Natl. Acad. Sci.<br />

USA 105, 20567–20574 (2008).<br />

27. Zhao, Z. et al. Autophagosome-independent essential function for the autophagy protein<br />

Atg5 in cellular immunity to intracellular pathogens. Cell Host Microbe 4, 458–469<br />

(2008).<br />

28. Martens, S. et al. Disruption of Toxoplasma gondii parasitophorous vacuoles by the<br />

mouse p47-resistance GTPases. PLoS Pathog. 1, e24 (2005).<br />

29. Ling, Y.M. et al. Vacuolar and plasma membrane stripping and autophagic elimination<br />

of Toxoplasma gondii in primed effector macrophages. J. Exp. Med. 203, 2063–2071<br />

(2006).<br />

30. Zhao, Y.O., Khaminets, A., Hunn, J.P. & Howard, J.C. Disruption of the Toxoplasma<br />

gondii parasitophorous vacuole by IFNγ−inducible immunity-related GTPases (IRG<br />

proteins) triggers necrotic cell death. PLoS Pathog. 5, e1000288 (2009).<br />

31. Gutierrez, M.G. et al. Autophagy is a defense mechanism inhibiting BCG and<br />

Mycobacterium tuberculosis survival in infected macrophages. Cell 119, 753–766<br />

(2004).<br />

32. Singh, S.B., Davis, A.S., Taylor, G.A. & Deretic, V. Human IRGM induces autophagy to<br />

eliminate intracellular mycobacteria. Science 313, 1438–1441 (2006).<br />

33. Harris, J. et al. T helper 2 cytokines inhibit autophagic control of intracellular<br />

Mycobacterium tuberculosis. Immunity 27, 505–517 (2007).<br />

34. Taylor, G.A., Feng, C.G. & Sher, A. p47 GTPases: regulators of immunity to intracellular<br />

pathogens. Nat. Rev. Immunol. 4, 100–109 (2004).<br />

35. Papic, N., Hunn, J.P., Pawlowski, N., Zerrahn, J. & Howard, J.C. Inactive and active<br />

states of the interferon-inducible resistance GTPase, Irga6, in vivo. J. Biol. Chem. 283,<br />

32143–32151 (2008).<br />

36. Feng, C.G. et al. The immunity-related GTPase Irgm1 promotes the expansion of<br />

activated CD4 + T cell populations by preventing interferon-γ-induced cell death. Nat.<br />

Immunol. 9, 1279–1287 (2008).<br />

37. Delgado, M. et al. Autophagy and pattern recognition receptors in innate immunity.<br />

Immunol. Rev. 227, 189–202 (2009).<br />

38. Xu, Y. et al. Toll-like receptor 4 is a sensor for autophagy associated with innate immunity.<br />

Immunity 27, 135–144 (2007).<br />

39. Delgado, M.A., Elmaoued, R.A., Davis, A.S., Kyei, G. & Deretic, V. Toll-like receptors<br />

control autophagy. EMBO J. 27, 1110–1121 (2008).<br />

40. Lee, H.K., Lund, J.M., Ramanathan, B., Mizushima, N. & Iwasaki, A. Autophagydependent<br />

viral recognition by plasmacytoid dendritic cells. Science 315, 1398–1401<br />

(2007).<br />

41. Shi, C.S. & Kehrl, J.H. MyD88 and Trif target Beclin 1 to trigger autophagy in macrophages.<br />

J. Biol. Chem. 283, 33175–33182 (2008).<br />

42. Saitoh, T. et al. Loss of the autophagy protein Atg16L1 enhances endotoxin-induced<br />

IL-1β production. <strong>Nature</strong> 456, 264–268 (2008).<br />

43. Shui, w. et al. Membrane proteomics of phagosomes suggests a connection to<br />

autophagy. Proc. Natl. Acad. Sci. USA 105, 16952–16957 (2008).<br />

44. Jounai, N. et al. The Atg5 Atg12 conjugate associates with innate antiviral immune<br />

responses. Proc. Natl. Acad. Sci. USA 104, 14050–14055 (2007).<br />

45. Tal, M.C. et al. Absence of autophagy results in reactive oxygen species-dependent<br />

amplification of RLR signaling. Proc. Natl. Acad. Sci. USA 106, 2770–2775<br />

(2009).<br />

46. Cadwell, K. et al. A key role for autophagy and the autophagy gene Atg16l1 in mouse<br />

and human intestinal Paneth cells. <strong>Nature</strong> 456, 259–263 (2008).<br />

47. Petrilli, V., Dostert, C., Muruve, D.A. & Tschopp, J. The inflammasome: a danger sensing<br />

complex triggering innate immunity. Curr. Opin. Immunol. 19, 615–622 (2007).<br />

48. Kanneganti, T.D., Lamkanfi, M. & Nunez, G. Intracellular NOD-like receptors in host<br />

defense and disease. Immunity 27, 549–559 (2007).<br />

nature immunology volume 10 number 5 may 2009 469


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

REVIEw<br />

49. Lunemann, J.D. & Munz, C. Autophagy in CD4 + T-cell immunity and tolerance. Cell<br />

Death Differ. 16, 79–86 (2009).<br />

50. Vyas, J.M. van,d., V, & Ploegh,H.L. The known unknowns of antigen processing and<br />

presentation. Nat. Rev. Immunol. 8, 607–618 (2008).<br />

51. English, D. et al. Autophagy enhances the presentation of endogenous viral antigens<br />

on MHC class I molecules during HSV-1 infection. Nat. Immunol. 10, 480–487<br />

(2009).<br />

52. Paludan, C. et al. Endogenous MHC class II processing of a viral nuclear antigen after<br />

autophagy. Science 307, 593–596 (2005).<br />

53. Schmid, D., Pypaert, M. & Munz, C. Antigen-loading compartments for major histocompatibility<br />

complex class II molecules continuously receive input from autophagosomes.<br />

Immunity 26, 79–92 (2007).<br />

54. Mizushima, N., Yamamoto, A., Matsui, M., Yoshimori, T. & Ohsumi, Y. In vivo analysis<br />

of autophagy in response to nutrient starvation using transgenic mice expressing a<br />

fluorescent autophagosome marker. Mol. Biol. Cell 15, 1101–1111 (2004).<br />

55. Nedjic, J., Aichinger, M., Emmerich, J., Mizushima, N. & Klein, L. Autophagy in thymic<br />

epithelium shapes the T-cell repertoire and is essential for tolerance. <strong>Nature</strong> 455,<br />

396–400 (2008).<br />

56. Li, Y. et al. Efficient cross-presentation depends on autophagy in tumor cells. Cancer<br />

Res. 68, 6889–6895 (2008).<br />

57. Qu, X. et al. Autophagy gene-dependent clearance of apoptotic cells during embryonic<br />

development. Cell 128, 931–946 (2007).<br />

58. Mellen, M.A., de la Rosa, E.J. & Boya, P. The autophagic machinery is necessary for<br />

removal of cell corpses from the developing retinal neuroepithelium. Cell Death Differ.<br />

15, 1279–1290 (2008).<br />

59. Arsov, I. et al. BAC-mediated transgenic expression of fluorescent autophagic protein<br />

Beclin 1 reveals a role for Beclin 1 in lymphocyte development. Cell Death Differ. 15,<br />

1385–1395 (2008).<br />

60. Stephenson, L.M. et al. Identification of Atg5-dependent transcriptional changes and<br />

increases in mitochondrial mass in Atg5-deficient T lymphocytes. Autophagy (in the<br />

press).<br />

61. Pua, H.H., Dzhagalov, I., Chuck, M., Mizushima, N. & He, Y.w. A critical role for the<br />

autophagy gene Atg5 in T cell survival and proliferation. J. Exp. Med. 204, 25–31<br />

(2007).<br />

62. Pua, H.H., Guo, J., Komatsu, M. & He, Y.w. Autophagy is essential for mitochondrial<br />

clearance in mature T lymphocytes. J. Immunol. 182, 4046–4055 (2009).<br />

63. Miller, B.C. et al. The autophagy gene ATG5 plays an essential role in B lymphocyte<br />

development. Autophagy 4, 309–314 (2007).<br />

64. Li, C. et al. Autophagy is induced in CD4 + T cells and important for the growth factor-<br />

withdrawal cell death. J. Immunol. 177, 5163–5168 (2006).<br />

65. Espert, L. et al. Autophagy is involved in T cell death after binding of HIV-1 envelope<br />

proteins to CXCR4. J. Clin. Invest. 116, 2161–2172 (2006).<br />

66. Bell, B.D. et al. FADD and caspase-8 control the outcome of autophagic signaling in<br />

proliferating T cells. Proc. Natl. Acad. Sci. USA 105, 16677–16682 (2008).<br />

67. Arbour, N. et al. c-Jun NH 2 -terminal kinase (JNK)1 and JNK2 signaling pathways have<br />

divergent roles in CD8 + T cell-mediated antiviral immunity. J. Exp. Med. 195, 801–810<br />

(2002).<br />

68. Conze, D. et al. c-Jun NH 2 -terminal kinase (JNK)1 and JNK2 have distinct roles in<br />

CD8 + T cell activation. J. Exp. Med. 195, 811–823 (2002).<br />

69. Dong, C. et al. Defective T cell differentiation in the absence of Jnk1. Science 282,<br />

2092–2095 (1998).<br />

70. wei, Y., Pattingre, S., Sinha, S., Bassik, M. & Levine, B. JNK1-mediated phosphorylation<br />

of Bcl-2 regulates starvation-induced autophagy. Mol. Cell 30, 678–688<br />

(2008).<br />

71. Barrett, J.C. et al. Genome-wide association defines more than 30 distinct susceptibility<br />

loci for Crohn’s disease. Nat. Genet. 40, 955–962 (2008).<br />

72. Massey, D.C. & Parkes, M. Genome-wide association scanning highlights two autophagy<br />

genes, ATG16L1 and IRGM, as being significantly associated with Crohn’s disease.<br />

Autophagy 3, 649–651 (2007).<br />

73. The wellcome Trust Case Control Consortium. Genome-wide association study of<br />

14,000 cases of seven common diseases and 3,000 shared controls. <strong>Nature</strong> 447,<br />

661–678 (2007).<br />

74. McCarroll, S.A. et al. Deletion polymorphism upstream of IRGM associated with altered<br />

IRGM expression and Crohn’s disease. Nat. Genet. 40, 1107–1112 (2008).<br />

75. Rioux, J.D. et al. Genome-wide association study identifies new susceptibility loci<br />

for Crohn disease and implicates autophagy in disease pathogenesis. Nat. Genet. 39,<br />

596–604 (2007).<br />

76. Kuballa, P., Huett, A., Rioux, J.D., Daly, M.J. & Xavier, R.J. Impaired autophagy of an<br />

intracellular pathogen induced by a Crohn’s disease associated ATG16L1 variant. PLoS<br />

ONE 3, e3391 (2008).<br />

77. Vaishnava, S., Behrendt, C.L., Ismail, A.S., Eckmann, L. & Hooper, L.V. Paneth cells<br />

directly sense gut commensals and maintain homeostasis at the intestinal host-microbial<br />

interface. Proc. Natl. Acad. Sci. USA 105, 20858–20863 (2008).<br />

78. Kaser, A. et al. XBP1 links ER stress to intestinal inflammation and confers genetic<br />

risk for human inflammatory bowel disease. Cell 134, 743–756 (2008).<br />

79. Cadwell, K., Patel, K.K., Komatsu, M., Virgin, H.w. & Stappenbeck, T.S. A common<br />

role for Atg16L1, Atg5, and Atg7 in small intestinal Paneth cells and Crohn’s disease.<br />

Autophagy 5, 250–252 (2008).<br />

470 volume 10 number 5 may 2009 nature immunology


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

A Crohn’s disease–associated NOD2 mutation<br />

suppresses transcription of human IL10 by inhibiting<br />

activity of the nuclear ribonucleoprotein hnRNP-A1<br />

Eiichiro Noguchi 1 , Yoichiro Homma 1 , Xiaoyan Kang 2 , Mihai G Netea 3 & Xiaojing Ma 2<br />

A common mutation in the gene encoding the cytoplasmic sensor Nod2, involving a frameshift insertion at nucleotide 3020<br />

(3020insC), is strongly associated with Crohn’s disease. How 3020insC contributes to this disease is a controversial issue.<br />

Clinical studies have identified defective production of interleukin 10 (IL-10) in patients with Crohn’s disease who bear the<br />

3020insC mutation, which suggests that 3020insC may be a loss-of-function mutation. However, here we found that 3020insC<br />

Nod2 mutant protein actively inhibited IL10 transcription. The 3020insC Nod2 mutant suppressed IL10 transcription by<br />

blocking phosphorylation of the nuclear ribonucleoprotein hnRNP-A1 via the mitogen-activated protein kinase p38. We<br />

confirmed impairment in phosphorylation of hnRNP-A1 and binding of hnRNP-A1 to the IL10 locus in peripheral blood<br />

mononuclear cells from patients with Crohn’s disease who bear the 3020insC mutation and have lower production of IL-10.<br />

Crohn’s disease is a chronic inflammatory bowel disease that affects<br />

mainly the terminal ileum, cecum, colon and perianal area. Inflammatory<br />

bowel disease is thought to result from inappropriate and<br />

continuing activation of the mucosal immune system driven by<br />

normal luminal flora. This aberrant response is probably facilitated<br />

by defects in both the barrier function of the intestinal epithelium and<br />

the mucosal immune system. These defects are in turn caused by<br />

genetic and environmental factors 1 .<br />

Interleukin 10 (IL-10; A001243) is a pleiotropic cytokine produced<br />

by T cells, B cells and macrophages. IL-10-deficient mice show<br />

enhanced development of several inflammatory and autoimmune<br />

diseases, which suggests that IL-10 serves a central function in vivo<br />

in restricting inflammatory responses 2 . Although they are healthy in<br />

germ-free conditions, some IL-10-deficient mouse strains develop<br />

colitis when given a normal, specific pathogen–free bowel flora,<br />

probably as a result of a poorly controlled mucosal immune response.<br />

Innate immune responses are initiated by the detection of microbial<br />

invaders by several distinct host systems collectively called ‘patternrecognition<br />

receptors’. The nucleotide-binding oligomerization<br />

domain (Nod)-like receptors, Toll-like receptors (TLRs), C-type lectin<br />

receptors and RIG-I-like receptors are examples of pattern-recognition<br />

receptors. Two members of the Nod-like receptor family of proteins,<br />

Nod1 and Nod2, are believed to be cytoplasmic sensors of microbial<br />

products 3 . Both Nod1 and Nod2 recognize peptidoglycan, but each<br />

protein senses distinct molecular motifs in peptidoglycan. Nod1<br />

recognizes a naturally occurring muropeptide of peptidoglycan that<br />

presents a unique amino acid at its terminus called ‘diaminopimelic<br />

Received 26 January; accepted 24 February; published online 6 April 2009; doi:10.1038/ni.1722<br />

ARTICLES<br />

acid’. This amino acid is found mainly in the peptidoglycan of<br />

Gram-negative bacteria 4 . In contrast, Nod2 can detect the minimal<br />

bioactive fragment of peptidoglycan called ‘muramyl dipeptide’. Thus,<br />

whereas Nod1 detects mainly Gram-negative bacteria, Nod2 is a more<br />

general sensor of bacterial peptidoglycan 5 .<br />

Nod2 expression is restricted to monocytes 6 andisrecognizedasan<br />

important initiator of inflammation. Models hypothesize that after<br />

binding to muramyl dipeptide, Nod2 forms oligomers and binds to the<br />

caspase-recruitment domain–containing serine-threonine kinase RICK<br />

(also called RIP2, CARDIAK, CCK and Ripk2). RICK then forms<br />

oligomers and facilitates the ubiquitination of a site on the modulator<br />

of transcription factor NF-kB NEMO (also called IKKg) 7 ,whichresults<br />

in activation of the inhibitor of NF-kB kinase complex and NF-kB 8 .<br />

The gene encoding Nod2 on human chromosome 16 (NOD2) was<br />

the first gene identified as being associated with susceptibility to<br />

Crohn’s disease 9 . Between 30% and 50% of patients with Crohn’s<br />

disease in the Western hemisphere carry NOD2 mutations on at least<br />

one allele. Mutations resulting in substitution of tryptophan for<br />

arginine at amino acid residue 702 (R702W) or arginine for glycine<br />

at amino acid residue 908 (G908R) and a frameshift substitution at<br />

amino acid position 1007 are estimated to represent 32%, 18% and<br />

31%, respectively, of all disease-causing mutations in NOD2 and are<br />

independently associated with susceptibility to Crohn’s disease 10 .All<br />

three mutations are located in the leucine-rich repeat (LRR) domain<br />

of NOD2. The frameshift substitution at amino acid 1007 associated<br />

with Crohn’s disease stems from an insertion mutation at nucleotide<br />

3020; this mutation (3020insC) results in partial deletion of the<br />

1 Department of Surgery II, Tokyo Women’s Medical University, Tokyo, Japan. 2 Department of Microbiology and <strong>Immunology</strong>, Weill Medical College of Cornell University,<br />

New York, New York, USA. 3 Department of Medicine (463), Radboud University Nijmegen Medical Center, Nijmegen, The Netherlands. Correspondence should be sent to<br />

X.M. (xim2002@med.cornell.edu).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 471


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a b c d e<br />

Mouse IL-10 (ng/ml)<br />

3<br />

2<br />

1<br />

0<br />

WT<br />

Nod2-KO<br />

Medium<br />

LPS<br />

PGN<br />

MDP<br />

Pam3Cys Mouse IL-12p40 (ng/ml)<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

WT<br />

Nod2-KO<br />

Medium<br />

LPS<br />

PGN<br />

MDP<br />

Pam3Cys terminal LRR domain of the protein. The genotype relative risk for<br />

developing Crohn’s disease in people heterozygous or homozygous for<br />

this mutation alone is calculated to be 1.5 2.6 or 17.6 42.1,<br />

respectively 11 . Overall, patients homozygous for 3020insC demonstrate<br />

a much more severe disease phenotype than other patients with<br />

Crohn’s disease and have a higher risk for ileal stenosis and surgical<br />

intervention 12 . It is postulated that the LRR domain of Crohn’s<br />

disease–associated Nod variants is impaired in its ability to recognize<br />

microbial components and/or in its ability to inhibit the formation of<br />

Nod2 dimers, which results in lack of activation of NF-kB in<br />

monocytes 9 . However, the higher NF-kB activity in patients with<br />

Crohn’s disease and the clinical effectiveness of NF-kB inhibitors in<br />

ameliorating symptoms of Crohn’s disease 1 suggest that diseasecausing<br />

variants of Nod2 may promote activation of NF-kB.<br />

IL-10-deficient mice develop a chronic enterocolitis that shares<br />

histopathological features with human Crohn’s disease 13 .Colitisin<br />

IL-10-deficient mice is driven by T helper type 1 (TH1) cells but not by<br />

B cells 14 and is dependent on the presence of resident enteric<br />

bacteria 15 . In accordance with those findings, Helicobacter hepaticus<br />

triggers colitis in specific pathogen–free IL-10-deficient mice by<br />

a mechanism that depends on T H1 proinflammatory cytokines 16 .<br />

The relationship between Nod2 and IL-10 is not completely understood.<br />

Peripheral blood mononuclear cells (PBMCs) from patients<br />

with Crohn’s disease who are homozygous for the 3020insC mutation<br />

show defective release of IL-10 after stimulation with enteric microorganisms<br />

and TLR ligands 17 . Those findings contrast with studies of<br />

mice in which an artificially created mouse version (2939insC) of<br />

human 3020insC replaces the endogenous wild-type mouse Nod2 gene<br />

(Nod2 2939insC ) 18 . In these mice, macrophages stimulated by the TLR4<br />

ligand lipopolysaccharide (LPS), peptidoglycan or muramyl dipeptide<br />

release quantities of IL-10 resembling those released by wild-type<br />

macrophages. We hypothesized that the human 3020insC and mouse<br />

Nod2 2939insC alleles may have different functional properties because of<br />

species-specific physiology or evolutionary adaptation. We undertook<br />

this study to explore that possibility and the potential target(s) of the<br />

mutant Nod2 protein produced by the 3020insC mutation (called<br />

‘3020insC Nod2’ here).<br />

RESULTS<br />

Suppression of IL10 production by 3020insC Nod2<br />

To investigate the influence of endogenous Nod2 on cytokine production,<br />

we measured the production of IL-10 and IL-12p40 by mouse<br />

bone marrow–derived macrophages (BMDMs) from wild-type and<br />

Mouse IL-10 (ng/ml)<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

WT<br />

RIP2-KO<br />

Medium<br />

LPS<br />

PGN<br />

Pam3Cys Human IL-10 (pg/ml)<br />

900<br />

Nod2 / mice 19 . We detected little difference in the amount of IL-10<br />

(Fig. 1a) orIL-12p40(Fig. 1b) released by wild-type and Nod2 /<br />

BMDMs after stimulation with LPS, peptidoglycan or Pam 3Cys (a<br />

synthetic TLR2 ligand). Muramyl dipeptide alone stimulated little<br />

production of IL-10 or IL-12p40 in this experiment, consistent with<br />

results obtained with human PBMCs, in which muramyl dipeptide<br />

augments only TLR2 ligand–mediated induction of IL-12 and IL-10<br />

(ref. 20). The patterns of production of IL-10 and IL-12p40 by<br />

peritoneal macrophages and splenocytes were very similar to those<br />

in BMDMs (data not shown). Furthermore, RICK (RIP2) was also<br />

dispensable for IL-10 production in primary macrophages stimulated<br />

by LPS, peptidoglycan or Pam 3Cys (Fig. 1c).<br />

Next we tested the response of primary monocytes to muramyl<br />

dipeptide, Pam3Cys or a combination of muramyl dipeptide and<br />

Pam 3Cys in cells from patients with Crohn’s disease who were<br />

homozygous for the 3020insC mutation, as well as cells from control<br />

volunteers bearing the wild-type NOD2 allele. In contrast to the data<br />

obtained with mice, IL-10 production was much lower in cells from<br />

patients homozygous for 3020insC (Fig. 1d). Production of tumor<br />

necrosis factor was also significantly lower in cells from patients<br />

homozygous for 3020insC after stimulation with muramyl dipeptide<br />

alone but was similar with that of wild-type cells after stimulation with<br />

Pam 3Cys or muramyl dipeptide plus Pam 3Cys (Fig. 1e).<br />

Thus, we hypothesized that the 3020insC Nod2 mutant may be able<br />

to regulate expression of the gene encoding IL-10 (IL10). To determine<br />

the effect of 3020insC Nod2 on endogenous IL10 expression, we<br />

transduced primary human monocytes with lentivirus expressing<br />

wild-type or 3020insC Nod2 (Fig. 2a). We found that 3020insC<br />

Nod2 but not wild-type Nod2 inhibited both basal as well as LPSinduced<br />

expression of IL-10 mRNA and protein (Fig. 2b,c). In<br />

contrast, we found no difference in IL-1b production by cells overexpressing<br />

wild-type or 3020insC Nod2 (Fig. 2d), and IL-12p40<br />

production was suppressed by high expression of wild-type Nod2<br />

but not 3020insC Nod2 (Fig. 2e). However, it is possible that the lack<br />

of influence of 3020insC Nod2 on IL-12p40 production was secondary<br />

to its suppression of IL-10 production. Thus, wild-type and 3020insC<br />

Nod2 seem to affect IL-10, IL-1b and IL-12p40 expression differently<br />

in human monocytes. We obtained similar results when we used<br />

peptidoglycan and Pam 3Cys instead of LPS (data not shown).<br />

To confirm our finding of an influence of Nod2 on IL10 transcription,<br />

we transfected primary human monocytes with a human IL10<br />

promoter–luciferase reporter construct (containing the sequence<br />

between positions 1044 and +30 relative to the transcription start<br />

Human TNF (pg/ml)<br />

400<br />

600<br />

300<br />

0 *<br />

*<br />

200<br />

0 *<br />

Medium<br />

WT Nod2<br />

3020insC Nod2<br />

MDP<br />

Pam3Cys MDP +<br />

Pam<br />

Medium<br />

WT Nod2<br />

3020insC Nod2<br />

Figure 1 Influence of Nod2 signaling on macrophage cytokine production. (a–c) ELISA of mouse IL-10 (a,c) and IL-12p40 (b) in supernatants of wild-type<br />

(WT), Nod2 –/– (Nod2-KO) and Ripk2 –/– (RIP2-deficient; RIP2-KO) mouse BMDMs stimulated for 24 h in vitro with medium, LPS (1 mg/ml), peptidoglycan<br />

(PGN; 10 mg/ml), muramyl dipeptide (MDP; 10 mg/ml) or Pam 3Cys (1 mg/ml); results are normalized to total cellular protein. Data represent three<br />

independent experiments (error bars, s.e.m.). (d,e) ELISA of human IL-10 (d) and tumor necrosis factor (e) in supernatants of monocytes obtained from<br />

healthy people expressing wild-type Nod2 (n ¼ 9) and patients with Crohn’s disease who were homozygous for 3020insC (n ¼ 5) and then stimulated for<br />

24 h with medium, muramyl dipeptide, Pam 3Cys or a combination of muramyl dipeptide and Pam 3Cys (MDP + Pam); results are normalized to total cellular<br />

protein. *, P o 0.05. Data are representative of one experiment with each donor sample measured in triplicate (error bars, s.d.).<br />

472 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

MDP<br />

Pam3Cys MDP +<br />

Pam


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b<br />

IL-10 (ng/ml)<br />

IL-12p40 (ng/ml)<br />

Medium<br />

WT CARD CARD Nod<br />

1,040<br />

LRR<br />

IL-10<br />

3020insC CARD CARD Nod<br />

CV WT Nod2<br />

c<br />

3020insC Nod2<br />

4<br />

3<br />

LRR<br />

1,007<br />

WT or<br />

3020insC Nod2<br />

β-actin<br />

CV WT Nod2<br />

3020insC Nod2<br />

d 4<br />

2<br />

1 * *<br />

3<br />

2<br />

1<br />

0<br />

Supernatant<br />

(µl)<br />

5 50<br />

Medium<br />

5 50<br />

LPS<br />

0<br />

Supernatant<br />

(µl)<br />

5 50<br />

Medium<br />

5 50<br />

LPS<br />

e 10<br />

CV WT Nod2<br />

3020insC Nod2 f 1.4<br />

8<br />

6<br />

4<br />

2<br />

*<br />

1<br />

0.6<br />

0.2<br />

*<br />

0<br />

Supernatant<br />

(µl)<br />

5 50<br />

Medium<br />

5 50<br />

LPS<br />

pCDNA3 WT 3020insC<br />

Nod2 Nod2<br />

site) 21 together with vectors expressing wild-type or 3020insC Nod2.<br />

Neither the empty vector (pCDNA3) nor wild-type Nod2 altered<br />

constitutive transcription of IL10; however, 3020insC Nod2 suppressed<br />

constitutive IL10 transcription by about 50% (Fig. 2f). These results<br />

collectively demonstrate that the 3020insC mutation conferred on<br />

Nod2 a new function, the ability to inhibit IL-10 production, and loss<br />

of the ability to repress IL-12p40 expression.<br />

Inhibition of IL10 promoter activity by 3020insC<br />

To further explore the molecular mechanism whereby 3020insC Nod2<br />

inhibited IL10 expression, we transfected the IL10 promoter–luciferase<br />

reporter and the wild-type or 3020insC Nod2 construct into the<br />

RAW264.7 mouse macrophage cell line. We verified expression of<br />

each construct by RT-PCR (Supplementary Fig. 1 online). Although<br />

the IL10 promoter was constitutively active, peptidoglycan and<br />

Pam 3Cys further boosted luciferase activity (Fig. 3a). Expression<br />

of 3020insC Nod2 strongly inhibited IL10 transcription in all conditions,<br />

even in the absence of any microbial stimuli, but wild-type<br />

Nod2 did not.<br />

To determine if wild-type and 3020insC Nod2 could functionally<br />

compete (for example, in the setting of a heterozygous person), we<br />

transfected wild-type and 3020insC Nod2 together, at various<br />

ratios, into RAW264.7 cells, along with the IL10 promoter–luciferase<br />

reporter. When wild-type Nod2 was more abundant than 3020insC<br />

IL-1β (ng/ml)<br />

Stimulation index<br />

LPS<br />

CV<br />

WT<br />

3020insC<br />

CV<br />

WT<br />

3020insC<br />

Figure 3 Inhibition of IL10 transcription by 3020insC Nod2. (a) Luciferase<br />

activity of lysates of RAW264.7 cells transfected for 1 d with the human<br />

IL10 promoter–lucifersase reporter together with empty vector (pCDNA3) or<br />

vector encoding wild-type or 3020insC Nod2 at an effector/reporter molar<br />

ratio of 0.3:1, then stimulated for 24 h with medium (Med), peptidoglycan<br />

or Pam 3Cys. RLU, raw luciferase units. Expression of the transfected Nod2<br />

and 3020insC mutant in RAW264.7 cells was verified by RT-PCR with<br />

primers selective for human but not mouse Nod2 (Supplementary Fig. 1).<br />

(b,c) Luciferase activity in lysates of cells transfected for 1 d with the<br />

human IL10 promoter–luciferase reporter, together with various molar ratios<br />

(horizontal axes) of vectors encoding wild-type and 3020insC Nod2 (b) or<br />

wild-type Nod2 and empty vector (pCDNA3; c), then stimulated for 7 h with<br />

medium or peptidoglycan. Reporter amount remained constant, set as 1.<br />

(d) Immunoassay of extracts of HEK293T cells transfected for 24 h (Input)<br />

with Flag- or Myc-tagged wild-type Nod2 (WT) or 3020insC Nod2 (frameshift<br />

(FS)), then immunoprecipitated (IP) with anti-Flag (a-Flag) or anti-c-Myc<br />

(a-Myc) and analyzed by immunoblot (IB) with anti-Flag. Arrows indicate that<br />

3020insC Nod2 is smaller than wild-type Nod2. Data represent the pooled<br />

results of at least three independent experiments (error bars (a–c), s.d.).<br />

Nod2, it countered the inhibitory effect of 3020insC Nod2 on IL10<br />

transcription; this counter-inhibitory effect was lost as 3020insC Nod2<br />

increased in abundance (Fig. 3b,c). Competition between wild-type<br />

and 3020insC Nod2 could have been due to direct interaction between<br />

the two molecules. To investigate that possibility, we generated wildtype<br />

and 3020insC Nod2 constructs tagged with Flag and c-Myc<br />

epitopes. After being transfected into human embryonic kidney<br />

(HEK293) cells, wild-type and 3020insC Nod2 constructs engaged in<br />

homotypic and heterotypic interactions (Fig. 3d). Thus, wild-type<br />

Nod2 may inhibit 3020insC Nod2–mediated suppression of IL10<br />

transcription through direct physical interference.<br />

Because only a fraction of patients with Crohn’s disease are<br />

homozygous carriers of the 3020insC mutation, we also analyzed the<br />

effect of the other two common Nod2 variants (R702W and G908R)<br />

on IL10 expression. Like the 3020insC Nod2 mutant, the R702W<br />

and G908R Nod2 mutants exerted an inhibitory effect on IL10<br />

promoter activity (data not shown). In summary, these data suggest<br />

that all three LRR mutants of Nod2 are able to inhibit basal and<br />

TLR-induced IL10 transcription.<br />

Gain of function is unique to human Nod2 and IL10<br />

AstudyofNod2 2939insC mice (in which an artificially created mouse<br />

version of human 3020insC replaces endogenous wild-type mouse<br />

Nod2) did not detect alteration in IL-10 expression in macrophages 18 .<br />

a pCDNA3<br />

b<br />

Luciferase activity<br />

(10 3 RLU/s)<br />

c<br />

Luciferase activity<br />

(10 2 RLU/s)<br />

18<br />

14<br />

10<br />

6<br />

2<br />

24<br />

16<br />

8<br />

0<br />

WT Nod2<br />

3020insC Nod2<br />

*<br />

Med<br />

Med<br />

PGN<br />

4:0 3:1 2:2 1:3 0:4<br />

WT:pCDNA3<br />

*<br />

*<br />

PGN Pam 3 Cys<br />

Luciferase activity<br />

(10 2 RLU/s)<br />

d<br />

16<br />

12<br />

8<br />

4<br />

Med<br />

PGN<br />

ARTICLES<br />

Figure 2 Different effects of wild-type and 3020insC Nod2 on endogenous<br />

IL-10 expression in primary human monocytes. (a) Wild-type and 3020insC<br />

Nod2 constructs. CARD, caspase-recruitment domain. (b) RT-PCR analysis<br />

of the expression of mRNA encoding IL-10, Nod2 and b-actin by primary<br />

human monocytes transduced for 4 d with empty control vector (CV) or<br />

lentivirus encoding wild-type or 3020insC Nod2 and then stimulated for 5 d<br />

with medium or LPS. On the basis of lentiviral expression of green fluorescent<br />

protein, an average infection rate of 38% was achieved. ‘Background’<br />

bands represent endogenous Nod2 mRNA in infected cells. (c–e) ELISAof<br />

the secretion of IL-10 (c), IL-1b (d) and IL-12p40 (e) by human monocytes<br />

infected with 5 or 50 ml lentivirus-containing supernatant (key) and treated<br />

with medium or LPS, analyzed on day 5 after infection and normalized to<br />

cytokine produced (pg/ml) per 1 10 5 live cells. *, P o 0.03. (f) Luciferase<br />

activity in lysates of primary human monocytes transfected for 6 h with a<br />

human IL10 promoter–luciferase reporter, together with empty vector<br />

(pCDNA3) or vector encoding wild-type or 3020insC Nod2 constructs<br />

(optimized response); results are presented as the stimulation index, derived<br />

from the ratio of the luciferase activity in each experimental condition to that<br />

of cells transfected with empty vector. Mean transfection efficiency, 35%.<br />

Data are representative of three experiments with one donor each (error<br />

bars (c–f), s.d.).<br />

0<br />

4:0 3:1 2:2 1:3 0:4<br />

WT:3020insC<br />

IP: α-Flag α-Myc<br />

Flag WT WT FS FS WT WT FS FS<br />

Input<br />

Myc WT FS WT FS WT FS WT FS<br />

IB:<br />

α-FLAG<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 473


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a b<br />

Luciferase activity (10 3 RLU/s)<br />

c d<br />

Mouse IL-10 (pg/ml)<br />

Human IL10 reporter Mouse Il10 reporter<br />

30<br />

25<br />

120<br />

100<br />

CV<br />

Human WT Nod2<br />

3020insC Nod2<br />

20<br />

80<br />

15<br />

60<br />

10<br />

40<br />

5<br />

20<br />

0<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

CV<br />

Mouse WT Nod2<br />

2939insC Nod2<br />

0<br />

Med MDP LPS PGN<br />

0<br />

Med MDP LPS PGN<br />

1,400<br />

CV<br />

1,200<br />

WT Nod2<br />

1,000<br />

3020insC<br />

800<br />

Nod2<br />

600<br />

400<br />

200<br />

0<br />

Med LPS PGN Pam3Cys Luciferase activity (10 3 RLU/s)<br />

Luciferase activity<br />

(10 3 RLU/s)<br />

That is inconsistent with our experimental data presented above and<br />

with clinical observations of human patients with Crohn’s disease 17,20 .<br />

We hypothesized that the human and mouse Nod2 mutants may have<br />

different functional properties, as the latter was created artificially and<br />

thus was not subject to evolutionary adaptation in mammalian<br />

physiology. We tested our hypothesis by comparing the effects of the<br />

human mutant (3020insC Nod2) and mouse mutant (2939insC Nod2)<br />

on the expression of a human IL10 promoter–luciferase reporter<br />

construct or mouse Il10 promoter–luciferase reporter construct.<br />

Human 3020insC Nod2 suppressed basal and TLR-induced expression<br />

of the human reporter but not of the mouse reporter (Fig. 4a,b). In<br />

contrast, mouse 2939insC Nod2 did not inhibit expression of either<br />

the human or mouse reporter (Fig. 4a,b). Furthermore, retrovirusmediated<br />

expression of 3020insC Nod2 into BMDMs derived from<br />

Nod2 / mice had little effect on either basal or TLR-stimulated IL-10<br />

production (Fig. 4c). These results indicate that the human and<br />

mouse Nod2 mutants have different functional properties and that<br />

only the human Nod2 mutant can inhibit expression of human IL-10.<br />

Blockade of hnRNP-A1–IL10 binding by 3020insC Nod2<br />

Next we searched for IL10 promoter elements (3020insC Nod2–<br />

response elements (NREs)) required for 3020insC Nod2–mediated<br />

inhibition of IL-10 production. Using an extensive and systematic<br />

3<br />

2<br />

1<br />

0<br />

Med<br />

LPS<br />

WT IL10<br />

WT Nod2<br />

3020insC Nod2<br />

PGN<br />

Med<br />

LPS<br />

PGN<br />

Mut IL10<br />

Figure 5 Binding of nuclear proteins to the NRE. (a) Immunoassay of<br />

lysates of RAW264.7 cells stably transfected with empty vector or vector<br />

encoding wild-type or 3020insC Nod2, then immunoprecipitated and<br />

analyzed by immunoblot with anti-Nod2. Results are representative of more<br />

than five separate experiments. (b) Sequences of probes used for EMSA. The<br />

mutant NRE probe bears the same mutations as those between positions<br />

30 and 25 in the IL10 promoter mutant construct in Figure 4d.<br />

(c) EMSA (with the probes in b) of nuclear extracts of the stable RAW264.7<br />

transfectants in a, stimulated with medium, LPS or peptidoglycan. FP, free<br />

probe. Results are representative of four separate experiments. (d) DNA<br />

precipitation of nuclear extracts of the stable RAW264.7 transfectants in a<br />

with biotinylated oligonucleotides encompassing the IL10 NRE, followed by<br />

elution at increasing NaCl concentrations (above gel), separation by 12%<br />

SDS-PAGE and silver staining. Arrows indicated bands that were excised and<br />

analyzed. M, molecular size markers (in kilodaltons (kDa)). Results are<br />

representative of three independent experiments with identical results.<br />

Figure 4 Human and mouse Nod2 mutant proteins have different<br />

transcriptional effects. (a,b) Luciferase activity in lysates of RAW264.7 cells<br />

transfected for 1 d with a human IL10 promoter–luciferase reporter (a) ora<br />

mouse Il10 promoter–luciferase reporter (b) together with empty control<br />

vector or vector expressing human wild-type or 3020insC Nod2 (top) or<br />

mouse wild-type or 2939insC Nod2 (bottom) at a reporter/effector molar<br />

ratio of 1:1, then stimulated for 24 h with medium, muramyl dipeptide, LPS<br />

or peptidoglycan. *, P o 0.03. Data represent three to four independent<br />

experiments (mean and s.d.). (c) ELISA of mouse IL-10 in supernatants<br />

of Nod2 –/– BMDMs infected twice with empty control vector (CV) or<br />

retrovirus encoding wild-type or 3020insC Nod2 and then, on day 5 after<br />

infection, stimulated for 24 h with medium, LPS, peptidoglycan or<br />

Pam 3Cys. Equivalent expression of transduced human wild-type and<br />

3020insC Nod2 mRNA in unstimulated Nod2 –/– BMDMs was ensured<br />

(Supplementary Fig. 3 online). Data are representative of three experiments<br />

with one mouse each (error bars, s.d. of triplicate samples). (d) Luciferase<br />

activity in lysates of RAW264.7 cells transiently transfected with constructs<br />

encoding wild-type or 3020insC Nod2 (key) plus luciferase reporter<br />

constructs driven by the IL10 promoter fragment (positions 782 to +12)<br />

with (Mut IL10) or without (WT IL10) point mutations in the region between<br />

positions 30 and 25 and then stimulated for 7 h with medium, LPS<br />

or peptidoglycan. Data are representative of four individual experiments<br />

(mean and s.d.).<br />

deletion- and mutagenesis-based search strategy, we focused on a<br />

narrow region of the IL10 promoter between positions 30 and 25<br />

upstream of the transcription-initiation site. A version containing point<br />

mutations introduced into the region between positions 30 and 25<br />

was not inhibited by 3020insC Nod2, in contrast to the inhibition of a<br />

construct containing a wild-type IL10 promoter fragment (Fig. 4d).<br />

Notably, the mutant also lost a substantial portion of its basal<br />

transcriptional activity. These results suggest that the TACACA<br />

sequence in the region between positions 30 and 25 may form<br />

the core of an NRE and that a nuclear DNA-binding protein may<br />

constitutively interact with this region. It is noteworthy that the NRE is<br />

predicted to be a positive functional element because its mutation<br />

resulted in less basal and microbe-stimulated IL10 transcription.<br />

We hypothesized that the NRE may bind to transcription factors<br />

involved in IL10 transcription and that this binding may be regulated<br />

by 3020insC Nod2. To test our hypothesis, we established stable clones<br />

of RAW264.7 cells constitutively expressing wild-type or 3020insC<br />

Nod2 (Fig. 5a). We stimulated these clones with LPS or peptidoglycan,<br />

then isolated nuclear extracts and assessed by electrophoretic<br />

mobility-shift assay (EMSA) their binding to a probe containing the<br />

wild-type or mutated human IL10 NRE (Fig. 5b,c). We detected a<br />

nuclear factor bound to the wild-type but not mutant probe in<br />

unstimulated cells; this binding did not increase after stimulation<br />

a CV WT 3020insC b<br />

IP: α-NOD2<br />

IB: α-NOD2<br />

c d<br />

Effector: FP<br />

TF-X<br />

WT NRE Mutant NRE<br />

Med LPS PGN Med<br />

FP<br />

CV<br />

WT3020insC<br />

CV<br />

WT<br />

3020insC<br />

CV<br />

WT<br />

3020insC<br />

CV<br />

WT<br />

3020insC<br />

NaCl (M)<br />

Human WT NRE (–36 to –19)<br />

5′– AAGGTCTACACATCAGGG–3′<br />

Human mutant NRE<br />

5′– AAGGTCCGTGTGTCAGGG–3′<br />

0.1<br />

(kDa)<br />

250 M CV<br />

WT<br />

150<br />

100<br />

75<br />

50<br />

37<br />

25<br />

0.2 0.4<br />

3020insC<br />

WT<br />

3020insC<br />

WT<br />

3020insC<br />

474 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b c d e<br />

Luciferase activity (10 3 RLU/s)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

*<br />

*<br />

*<br />

10<br />

8<br />

6<br />

4<br />

*<br />

*<br />

*<br />

*<br />

2.0<br />

1.5<br />

1.0<br />

*<br />

2<br />

1<br />

2<br />

0.5<br />

0<br />

1:0 1:0.25 1:0.5 1:0.75 1:1<br />

0<br />

1:0 1:0.25 1:0.5 1:0.75 1:1<br />

0<br />

WT NRE WT NRE Mut NRE<br />

Reporter:effector Reporter:effector<br />

pCDNA3 hnRNPA1<br />

Luciferase activity (10 3 RLU/s)<br />

with LPS or peptidoglycan (Fig. 5c). We designated this binding<br />

factor, which may represent a putative transcription factor involved in<br />

maintaining basal IL10 transcription, ‘TF-X’. Consistent with our<br />

prediction, 3020insC Nod2 resulted in much less binding of TF-X to<br />

the NRE, but wild-type Nod2 did not (Fig. 5c).<br />

We then did a series of biochemical experiments to identify TF-X.<br />

First we used biotinylated oligonucleotides encoding the NRE to<br />

precipitate TF-X from nuclear extracts derived from RAW264.7 clones<br />

expressing wild-type or 3020insC Nod2 (Fig. 5d). Next we excised the<br />

approximately 26-kilodalton TF-X band and analyzed it by nanoflow<br />

liquid chromatography–tandem mass spectrometry. The result with<br />

the highest score (30.98) for bands excised from cells expressing wildtype<br />

Nod2 was heterogeneous nuclear ribonucleoprotein A1 (hnRNP-<br />

A1; A001137); these hnRNP-A1 peptides (1,218.64 and 1,784.91<br />

daltons in size, with intensities of 1.89 10 6 and 5.57 10 5 ,<br />

respectively), which covered 8% of the amino acid residues of the<br />

full-length protein, were undetectable for the bands excised from<br />

extracts of cells expressing 3020insC Nod2.<br />

Activation of IL10 transcription by hnRNP-A1<br />

The hnRNPs are among the most abundant proteins in the eukaryotic<br />

cell nucleus and are directly involved in DNA repair and telomere<br />

elongation; chromatin remodeling and transcription; RNA splicing<br />

and stability; export of mature RNA; and translation. Approximately<br />

30 hnRNPs have been identified. Autoantibodies specific for A and B<br />

proteins of the hnRNP complex have been detected in rheumatoid<br />

arthritis, systemic lupus erythematosus and mixed connective tissue<br />

disease 22,23 . Moreover, hnRNP-A1 is also found as a potential autoantigen<br />

in 38% of psoriasis patients 24 .<br />

To analyze the influence of hnRNP-A1 on IL10 transcription, we<br />

transfected a mouse hnRNP-A1 expression vector (effector) together<br />

with the IL10 promoter–luciferase reporter into RAW264.7 cells. We<br />

found that hnRNP-A1 stimulated Il10 promoter activity in a dosedependent<br />

way at fairly low ratios of reporter to effector (Fig. 6a).<br />

Moreover, hnRNP-A1 also augmented LPS- and peptidoglycan-stimulated<br />

IL10 promoter–luciferase reporter activity (data not shown).<br />

Notably, mouse hnRNP-A1 also stimulated a mouse Il10 promoter–<br />

luciferase reporter construct (Fig. 6b). However, hnRNP-A1 had little<br />

effect on luciferase reporters driven by human IL-12p40 or ‘generic’<br />

NF-kB promoters, even at high ratios of reporter to effector (data not<br />

shown). Furthermore, the stimulatory effect of human hnRNP-A1<br />

Stimulation index<br />

3<br />

2<br />

IL-10<br />

IL-12p40<br />

*<br />

700<br />

500<br />

IL-10 *<br />

IL-12p40<br />

1<br />

300<br />

100<br />

0<br />

0<br />

hnRNP-A1<br />

hnRNP-A1<br />

β-actin<br />

β-actin<br />

siRNA Mock 4 1-3 Vector pCDNA3 hnRNP-A1<br />

Figure 6 Stimulation of IL10 transcription by hnRNP-A1. (a,b) Luciferase activity of RAW264.7 cells transfected for 48 h with a human (a) ormouse<br />

(b) IL-10 reporter together with pCDNA3 and/or mouse hnRNP-A1 at various reporter/effector ratios (horizontal axes). (c) Luciferase activity of RAW264.7<br />

cells transfected with reporter constructs driven by a wild-type (WT NRE) or mutant (Mut NRE) IL10 NRE, together with empty vector (pcDNA3) or vector<br />

encoding mouse hnRNP-A1, presented relative to that in pCDNA3-transfected cells. (d) ELISA of IL-10 and IL-12p40 in supernatants of primary human<br />

monocytes transfected with siRNA specific for hnRNP-A1 (4 and 1–3) or mock transfected (Mock; d) or transfected with with empty vector (pCDNA3) or<br />

expression vector encoding hnRNP-A1 (e), then, 6 h after transfection, stimulated for 12 h with LPS. Below, RT-PCR analysis of the expression of hnRNP-A1<br />

and b-actin mRNA demonstrates the efficiency of siRNA-mediated knockdown (d) or vector transfection (e). *, P o 0.05. Data represent from three to five<br />

independent experiments (mean and s.d.).<br />

Cytokine (ng/ml)<br />

was mediated through the NRE in the IL10 promoter, as hnRNP-A1<br />

had no effect on the NRE-mutant IL10 promoter–luciferase<br />

reporter (Fig. 6c).<br />

To demonstrate the physiological function of hnRNP-A1 in the<br />

regulation of endogenous IL-10 production in primary cells, we transfected<br />

human peripheral blood–derived monocytes with small interfering<br />

RNA (siRNA) specific for human hnRNP-A1. A pool of three<br />

siRNA molecules that diminished expression of hnRNP-A1 inhibited<br />

LPS-stimulated IL-10 production by about 45%; a fourth hnRNP-A1specific<br />

siRNA resembled the mock siRNA, as it failed to suppress<br />

hnRNP-A1 and did not influence IL-10 production (Fig. 6d). These<br />

siRNAs had no effect on LPS-induced production of IL-12p40. The<br />

degree of inhibition of IL-10 production by hnRNP-A1-specific siRNA<br />

was notable, given that fewer than 40% of the monocytes were transfected<br />

with the siRNA (data not shown) and that hnRNP-A1 expression<br />

was only partially ‘knocked down’. Conversely, overexpression of<br />

hnRNP-A1 in these cells strongly enhanced LPS-stimulated production<br />

of IL-10 but had no effect on IL-12p40 secretion (Fig. 6e). These results<br />

suggest that hnRNP-A1 promotes transcription of human IL10.<br />

Suppression of hnRNP-A1–NRE binding by 3020insC Nod2<br />

To confirm the identification of hnRNP-A1 as an NRE-binding<br />

nuclear protein by mass spectrometry, we used EMSA and chromatin<br />

immunoprecipitation (ChIP). As shown by EMSA, overexpression of<br />

hnRNP-A1 in RAW264.7 cells resulted in binding of protein to the<br />

human NRE probe; this binding activity was diminished by an<br />

antibody to hnRNP-A1 but not by an isotype-matched control antibody<br />

(Fig. 7a). ChIP analysis of primary human monocytes showed<br />

strong and specific binding of hnRNP-A1 in the region between<br />

positions 121 and +42 of IL10, which contains the NRE, but not<br />

in an irrelevant upstream region between positions 3158 and 2947<br />

(Fig. 7b). The binding seemed to be constitutive and was not<br />

regulated by stimulation of primary human monocytes with LPS<br />

(Fig. 7c). These data indicate that hnRNP-A1 interacts with the IL10<br />

NRE in human monocytes.<br />

Next we analyzed PBMCs from three patients with Crohn’s<br />

disease who were homozygous for 3020insC and had profoundly<br />

diminished IL-10 production 25 . The binding of hnRNP-A1 to IL10<br />

around the NRE was much lower in these patients than in age- and<br />

sex-matched healthy people and patients with Crohn’s disease<br />

who lacked the 3020insC mutation (Fig. 7d). The diminished<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 475<br />

Cytokine (pg/ml)<br />

ARTICLES


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a c<br />

b<br />

FP<br />

–121 to +42<br />

–3158 to –2947<br />

pCDNA3<br />

hnRNP-A1<br />

Control IgG<br />

α-hnRNP-A1<br />

hnRNP-A1<br />

hnRNP-A1<br />

e<br />

hnRNP-A1 binding (relative)<br />

Input<br />

Control lg<br />

α-hnRNP<br />

d<br />

Input Rat IgG α-hnRNP<br />

Med LPS Med LPS Med LPS<br />

Input Rat IgG α-hnRNP<br />

CD CD CD<br />

Ctrl WT FS Ctrl WT FS Ctrl WT FS<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

Ctrl WT FS Ctrl WT FS Ctrl WT FS<br />

CD<br />

CD<br />

CD<br />

Input rlgG α-hnRNP<br />

hnRNP-A1 binding in the patients with Crohn’s disease who<br />

were homozygous for 3020insC was even more prominent in<br />

experiments using real-time quantitative PCR (Fig. 7e).<br />

We further investigated the molecular basis for the impaired<br />

nuclear binding activity of hnRNP-A1 in patients with Crohn’s<br />

disease who were homozygous for 3020insC. For this, we overexpressed<br />

Flag-tagged wild-type or 3020insC Nod2 and Flag-tagged<br />

hnRNP-A1 in HEK293 cells. Wild-type and 3020insC Nod2 were<br />

expressed in the cytoplasm but not in the nucleus, whereas hnRNP-<br />

A1 was expressed in both compartments (Fig. 8a). Notably, in<br />

Figure 8 Nod2–hnRNP-A1 interaction and<br />

phosphorylation of hnRNP-A1 by p38.<br />

(a) Immunoassay of HEK293 cells transiently<br />

transfected for 1 d with empty vector (C) or<br />

with vector encoding Flag-tagged wild-type (W)<br />

or 3020insC (M) Nod2, together with vector<br />

encoding Flag-tagged hnRNP-A1, followed by<br />

immunoprecipitation of cytoplasmic and nuclear<br />

extracts with anti-Flag and immunoblot analysis<br />

with anti-Flag, anti-hnRNP-A1 (a-hnRNP) and<br />

antibody to phosphorylated serine (a-p-Ser). **,<br />

immunoglobulin heavy chain; *, immunoglobulin<br />

light chain. (b) Immunoassay of HEK293 cells<br />

transiently transfected for 1 d with the Flagtagged<br />

constructs described in a, followedby<br />

immunoprecipitation and immunoblot analysis of<br />

whole-cell lysates with anti-Flag or anti-hnRNP-<br />

A1. (c) Immunoassay of HEK293 cells<br />

transfected with human wild-type Nod2 (W) or<br />

3020insC Nod2 (M; left) or with mouse wild-type<br />

Nod2 (W) or 2939insC Nod2 (M; right), followed<br />

by immunoprecipitation of whole-cell lysates with<br />

antibody to serine-phosphorylated p38 (a-p-p38)<br />

and immunoblot analysis with anti-Flag.<br />

(d) Immunoassay of HEK293 cells treated for<br />

60 min with the p38 inhibitor SB203580,<br />

a<br />

d<br />

IP: α-Flag<br />

Cytoplasm Nucleus<br />

FLAG-NOD2 C W M C W M<br />

IB: α-Flag<br />

IB: α-hnRNP<br />

IB: α-p-Ser<br />

addition to the intact form of hnRNP-A1 (about 36 kilodaltons),<br />

we noted three shorter forms of hnRNP-A1, which have been<br />

described as products of hnRNP-A1 cleavage mediated by caspase<br />

3 activated during apoptosis of the human Burkitt lymphoma cell<br />

line BL60 induced by antibody to immunoglobulin M (anti-IgM) 26 .<br />

There was less cleaved hnRNP-A1 in cells expressing 3020insC<br />

Nod2 than in cells expressing empty vector or wild-type Nod2.<br />

In addition, intact hnRNP-A1 was serine-phosphorylated exclusively<br />

in the nucleus, whereas we detected serine-phosphorylated<br />

forms of cleavage products only in the cytoplasm. Expression<br />

b C W M e<br />

IB: α-Flag<br />

c<br />

*<br />

IP: α-p-p38<br />

Human<br />

Mouse<br />

Flag-Nod2 C W M C W M C W M<br />

Flag–hnRNP-A1<br />

Nod2<br />

– – – + + + + + +<br />

IgH<br />

hnRNP-A1<br />

TPA (µM) 0 1 1 1<br />

SB203580 (µM) 0 0 1 10<br />

IP: α-p-Ser<br />

IB: α-hnRNP<br />

Figure 7 Binding of hnRNP-A1 to the IL10 NRE. (a) EMSAofnuclear<br />

extracts of RAW264.7 cells stably transfected with empty vector (pCDNA3)<br />

or vector encoding hnRNP-A1, analyzed with a probe containing wild-type<br />

IL10 NRE. Right, extracts incubated with anti-hnRNP-A1 or control<br />

immunoglobulin (IgG) before incubation with probe. Results are representative<br />

of three independent experiments. (b) ChIP of primary human monocytes<br />

with anti-hnRNP-A1 or control immunoglobulin (Ig), followed by PCR<br />

amplification of IL10 promoter regions between positions 121 and +42<br />

(top) or positions 3158 and 2947 (bottom) in input DNA or DNA extracted<br />

from immunoprecipitates. Results are representative of three separate<br />

experiments. (c) ChIP of unstimulated (Med) and LPS-stimulated (LPS)<br />

primary human monocytes, followed by PCR amplification of the IL10<br />

promoter region between positions 121 and +42. Rat IgG serves as a<br />

control. Results are representative of three separate experiments. (d) ChIP<br />

analysis of unstimulated human PBMCs from healthy control subjects (Ctrl)<br />

or patients with Crohn’s disease (CD) with wild-type NOD2 (WT) or homozygous<br />

for the 3020insC mutation (FS). Results are from one representative<br />

of three experiments. (e) Real-time quantitative PCR analysis of the binding<br />

of hnRNP-A1 in each sample in d. Results for immunoprecipitated DNA are<br />

presented relative to those for genomic input DNA. Data are representative of<br />

experiments done in triplicate (mean and s.d. of three donors per group).<br />

IB: α-hnRNP<br />

hnRNP-A1<br />

Flag-<br />

IP: α-Flag<br />

IP: α-hnRNP<br />

IP: α-hnRNP<br />

IP: α-Flag<br />

followed by the addition of TPA (12-O-tetradecanoylphorbol-1,3-acetate) for 30 min (concentrations, above lanes) and immunoprecipitation and immunoblot<br />

analysis of nuclear extracts. Results in a–d represent at least three independent experiments. (e) Immunoblot analysis of whole-cell lysates of BMDMs<br />

obtained from wild-type mice (WT; n ¼ 3) and p38a-deficient mice (p38-KO; n ¼ 3) and then stimulated with LPS. Results are representative of two<br />

separate experiments. (f) Luciferase activity of lysates of RAW264.7 cells transfected with the human IL10 promoter–luciferase reporter together with empty<br />

vector (pCDNA3) or vector encoding wild-type, S310312A or S192A hnRNP-A1, then stimulated for 24 h with medium or LPS. *, P o 0.03. Data represent<br />

three independent experiments (mean and s.d.). (g) Immunoprecipitation and immunoblot analysis of cytoplasmic and nuclear extracts of PBMCs isolated<br />

from patients with Crohn’s disease homozygous for 3020insC (M; n ¼ 4) or expressing wild-type Nod2 (W1; n ¼ 4) and healthy people expressing wild-type<br />

Nod2 (W2; n ¼ 6). Results are representative of one experiment.<br />

**<br />

cleaved<br />

f<br />

Luciferase activity<br />

(10 3 RLU/s)<br />

g<br />

IB:<br />

8<br />

6<br />

4<br />

2<br />

*<br />

* *<br />

0 pCDNA3 WT S310–<br />

312A<br />

α-p38<br />

α-hnRNP-A1<br />

IP: α-p-Ser<br />

IB: α-hnRNP<br />

IP: α-p-p38<br />

IB: α-hnRNP<br />

p38<br />

hnRNP-A1<br />

p-hnRNP-A1<br />

*<br />

M W1 W2<br />

WT p38-KO<br />

Med<br />

LPS<br />

S192A<br />

Cytoplasm<br />

Nuclear<br />

Cytoplasm<br />

476 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

of 3020insC Nod2 selectively diminished the presence of<br />

phosphorylated hnRNP-A1 in the nucleus. By coimmunoprecipitation,<br />

we found evidence of direct physical interaction between<br />

endogenous hnRNP-A1 and exogenous wild-type Nod2; hnRNP-A1<br />

interacted less efficiently with 3020insC Nod2 (Fig. 8b). This<br />

interaction seemed to take place exclusively in the cytoplasm<br />

(data not shown), presumably because Nod2 is present only in<br />

the cytoplasm.<br />

Next we investigated whether the mitogen-activated protein kinase<br />

p38a (A001717) was responsible for phosphorylation of hnRNP-A1.<br />

Flag-tagged wild-type and 3020insC Nod2 interacted equally well with<br />

phosphorylated p38 (Fig. 8c). However, we detected more hnRNP-A1<br />

in complexes of serine-phosphorylated p38 and wild-type Nod2 than<br />

in complexes of serine-phosphorylated p38 and 3020insC Nod2.<br />

Notably, the 2939insC Nod2 mouse counterpart of 3020insC Nod2<br />

did not act like the human mutant in that it resembled wild-type<br />

Nod2 in its ability to facilitate interaction between serine-phosphorylated<br />

p38 and hnRNP-A1.<br />

In support of the idea that p38 was responsible for hnRNP-A1<br />

phosphorylation, we found that SB203580, a small chemical inhibitor<br />

of p38, blocked phorbol ester–induced phosphorylation of cleaved<br />

hnRNP-A1 in HEK293 cells in a dose-dependent way (Fig. 8d). As<br />

chemical inhibitors of p38 may have additional unidentified targets,<br />

we confirmed our findings in mice lacking p38a specifically in<br />

myeloid cells 27 . BMDMs from p38a-deficient and wild-type mice<br />

expressed similar amounts of total hnRNP-A1, but LPS-induced<br />

serine phosphorylation of hnRNP-A1 was much lower in p38adeficient<br />

macrophages (Fig. 8e). These data support the idea that<br />

p38, particularly the a-isoform, is critical for phosphorylation<br />

of hnRNP-A1.<br />

It has been shown that hnRNP-A1 is phosphorylated at the serine<br />

residue at position 192 (Ser192), Ser310, Ser311 and Ser312 after T cell<br />

stimulation 28 . To determine if these serine residues are critical for<br />

hnRNP-A1-mediated IL10 transcription, we generated mutants of<br />

hnRNP-A1 in which these serine residues were replaced with alanine<br />

(S192A and S310–312A). We transfected cells expressing the IL10<br />

promoter–luciferase reporter with empty vector or vector encoding<br />

wild-type, S192A or S310–312A hnRNP-A1 constructs. Wild-type<br />

hnRNP-A1 enhanced both basal and LPS-stimulated IL10 promoter<br />

activity relative to the promoter activity resulting from transfection of<br />

empty vector (Fig. 8f). S192A hnRNP-A1 also induced promoter<br />

activity, albeit less efficiently than wild-type hnRNP-A1 did; in<br />

contrast, S310–312A hnRNP-A1 failed to stimulate IL10 reporter<br />

activity (Fig. 8f). These data collectively suggest that phosphorylation<br />

of hnRNP-A1 at Ser310–Ser312, possibly by p38, is important for<br />

its ability to enhance transcription of human IL10; moreover,<br />

3020insC Nod2 may interfere with p38-mediated phosphorylation<br />

of hnRNP-A1.<br />

In agreement with the hypothesis proposed above, we noted<br />

considerable impairment in the phosphorylation of cleaved and fulllength<br />

hnRNP-A1 in the nucleus of cells from patients with Crohn’s<br />

disease who were homozygous for 3020insC, relative to that of cells<br />

from healthy subjects or patients with Crohn’s disease who lacked<br />

3020insC (Fig. 8g). Expression of total p38 and hnRNP-A1 was similar<br />

in all samples. The average extent of phosphorylation of hnRNP-A1 in<br />

control subjects (n ¼ 6) was 3.2-fold higher than that of patients<br />

with Crohn’s disease who were homozygous for 3020insC (n ¼ 6;<br />

P ¼ 0.02467). Consistent with the impaired phosphorylation of<br />

hnRNP-A1, there was less interaction between serine-phosphorylated<br />

p38 and hnRNP-A1 in these patients (Fig. 8g). These data collectively<br />

indicate that p38 phosphorylates hnRNP-A1 and that 3020insC Nod2<br />

ARTICLES<br />

inhibits this event by blocking the interaction between p38 and<br />

hnRNP-A1 (Supplementary Fig. 2 online).<br />

DISCUSSION<br />

In this study we have investigated the controversial issue of the<br />

influence of the NOD2 mutation 3020insC on host defense and<br />

inflammation. We have shown that 3020insC Nod2 acted as an<br />

inhibitor of IL10 expression in human monocytes. Because of the<br />

critical function of IL-10 in controlling mucosal inflammation caused<br />

by intestinal microflora, this finding links 3020insC to a probable<br />

route leading to the development and pathogenesis of Crohn’s disease.<br />

Our observations indicate that 3020insC Nod2 can interfere with the<br />

steady-state intracellular signaling responsible for constitutive IL10<br />

transcription. We have identified hnRNP-A1 as a chief target of<br />

3020insC Nod2 in this pathway and have shown that hnRNP-A1<br />

acted as a physiological inducer of IL10 transcription. Although<br />

3020insC Nod2 did not inhibit the expression or nuclear localization<br />

of hnRNP-A1, it did suppress the DNA-binding activity of hnRNP-A1<br />

by interfering with p38-mediated phosphorylation of hnRNP-A1.<br />

Notably, 3020insC Nod2 blocked transcription of human IL10 but<br />

not mouse Il10, and the engineered mouse counterpart of the<br />

3020insC Nod2 mutant, 2939insC Nod2, did not modulate expression<br />

of either mouse or human IL-10. Thus, the 3020insC-p38-hnRNP<br />

pathway is operative only in humans. This species specificity seems to<br />

be due to the ability of 3020insC Nod2 but not 2939insC Nod2 to<br />

block the interaction between p38 and hnRNP-A1.<br />

We did not find evidence that 3020insC Nod2 could enhance<br />

secretion of IL-1b from primary human monocytes. That finding is<br />

inconsistent with studies of the Nod2 2939insC knock-in mouse, which<br />

shows exacerbated IL-1b production due to hyperactivation of<br />

NF-kB 18 . It suggests that the discrepancy between those results and<br />

our finding of a lack of effect on IL-1b production in primary human<br />

monocytes transfected with 3020insC Nod2 is due to the possibility<br />

that 2939insC Nod2 has a gain of activity that the human mutant does<br />

not have. The data collectively demonstrate that the mouse 2939insC<br />

Nod2 mutant is not equivalent to human 3020insC Nod2 and that<br />

lacking Nod2 is not the same as expressing 3020insC. These findings<br />

call for caution in comparisons of mouse models and human disease.<br />

As 3020insC Nod2blockedsteady-stateaswellasmicrobe-induced<br />

IL10 expression, people with the 3020insC mutation may have chronically<br />

lower IL-10 production. This diminution in IL-10, when combined<br />

with other contributing factors over a lengthy period of time,<br />

could cumulatively cause persistent inflammation and homeostatic<br />

imbalance in the intestinal mucosa, leading to the development and<br />

pathogenesis of Crohn’s disease. In people heterozygous for 3020insC,<br />

wild-type Nod2 can presumably directly bind to 3020insC Nod2 and<br />

may thereby diminish its IL-10-inhibiting and disease-causing activities.<br />

This interaction could explain the much lower susceptibility to<br />

Crohn’s disease of heterozygous patients. Emphasizing the influence of<br />

other contributing factors, not all people with the 3020insC mutation<br />

on both chromosomes suffer from Crohn’s disease 29 . Moreover, none<br />

of the three common mutations in the LRR domain of NOD2 have<br />

been found in Japanese patients with Crohn’s disease 30 . That observation<br />

provides evidence for the presence of genetic heterogeneity among<br />

patients of different ethnic and racial backgrounds.<br />

It should be emphasized that published research documenting lower<br />

defensin expression 31 , dysfunction of dendritic cells 32 and diminished<br />

production of other cytokines 25,33 supports the idea that 3020insC<br />

represents a loss-of-function mutation. It is likely that the cumulative<br />

phenotypic effect of the 3020insC variant in patients with Crohn’s<br />

disease may be attributable to other alterations in addition to<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 477


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

diminished IL-10 production. Whether the 3020insC Nod2 mutant<br />

represents a loss of function or gain of function may depend on the<br />

response being measured. For example, because the 3020insC mutation<br />

is in the LRR domain of NOD2,the3020insC Nod2 mutant may not be<br />

able to bind muramyl dipeptide; thus, all muramyl dipeptide–induced<br />

responses could be lost. In that context, 3020insC Nod2 is a loss-offunction<br />

mutant. In agreement with that conclusion, monocytes<br />

derived from patients with Crohn’s disease who bear other mutations<br />

in the LRR domain, like those from patients with the 3020insC<br />

mutation, show impaired responses to muramyl dipeptide and several<br />

TLR ligands in terms of secretion of IL-8 and tumor necrosis factor,<br />

respectively 33 . However, our evidence suggests that 3020insC Nod2 and<br />

other Nod2 variants bearing mutations in the LRR domain have also<br />

acquired the ability to inhibit IL-10 transcription.<br />

In summary, our study challenges the present paradigms about the<br />

influence of the 3020insC mutation on Crohn’s disease. Notably,<br />

3020insC is also strongly associated with graft-versus-host disease 34 ,<br />

which is similar to Crohn’s disease in that both disorders are thought<br />

to be exacerbated by TH1 responses and suppressed by IL-10. Thus,<br />

our findings might be useful in efforts to identify therapeutic targets<br />

for the treatment of Crohn’s disease and other T H1-mediated autoimmune<br />

diseases associated with the 3020insC mutation.<br />

METHODS<br />

Mice. Nod2-knockout and control littermate mice were from P.J. Murray.<br />

RICK-knockout (RIP2-knockout) and control mice were from E. Pamer. Mice<br />

with conditional knockout of p38a were provided by J.M. Park 27 .<br />

Antibodies and reagents. Antibody to phosphorylated serine (AB1603) was<br />

from Chemicon International; anti-Flag (M2) was from Sigma; anti-p38 (9212)<br />

and antibody to phosphorylated p38 (c9211) were from Cell Signaling; and<br />

other antibodies for immunoblot analysis, EMSA and ChIP (anti-hnRNP-A1<br />

(10030), anti-Nod2 (30199)) were from Santa Cruz Biotechnology. Recombinant<br />

human and mouse interferon-g were from Genzyme. Muramyl dipeptide,<br />

peptidoglycan and LPS from Escherichia coli were from Sigma. The peptidoglycan,<br />

muramyl dipeptide and Pam3Cys (tripalmitoyl cysteinyl seryl tetralysine)<br />

used in all experiments were from the same source and were used at the<br />

concentrations described before 35 . They were free of endotoxin contamination,<br />

as shown by analysis with polymixin B (data not shown).<br />

Expression vectors. The mouse Nod2 expression vector and the mouse<br />

2939insC Nod2 mutant were provided by G. Nunez. The 3020insC and<br />

other human Nod2 mutants were generated by overlapping PCR. Human<br />

hnRNP-A1 cDNA was amplified by RT-PCR of total RNA from human<br />

monocytes. The sequences of all vectors were verified and all plasmids<br />

were isolated with an EndoFree Maxi kit (Qiagen).<br />

DNA affinity binding assay. The DNA affinity binding assay was done<br />

essentially as described 36 with 2 mg biotinylated DNA oligonucleotides encompassing<br />

the IL10 NRE conjugated to 100 ml streptavidin-bound Dynabeads.<br />

Bound proteins were eluted by increasing the strength of the elution buffer and<br />

were separated by 10% SDS-PAGE gel and visualized by silver staining.<br />

Retroviral and lentiviral packaging and macrophage transduction. GP2-293<br />

packaging cells (Clontech Laboratories) at a confluency of about 50 70% in<br />

100-mm culture plates were transfected, using Lipofectamine (Invitorogen),<br />

with 5 mg each of plasmids (pVSV-G; Clontech Laboratories) encoding human<br />

wild-type Nod2 or 3020insC Nod2. Then, 2 d after transfection, supernatants<br />

were cleared by centrifugation for 10 min at 1,000g and were pelleted for<br />

90 min at 50,000g. Pelleted viruses were resuspended at 4 1C inDMEM.Bone<br />

marrow cells were infected with retroviral particles on day 2 and again on day 4<br />

during their maturation in the presence of 20% (vol/vol) conditioned medium<br />

from mouse L cells. For lentivirus generation, HEK293TN cells (System<br />

Biosciences) at a confluency of 50% were transfected, using FuGENE 6 (Roche),<br />

with 3 mg of plasmid encoding viral envelope protein (G glycoprotein of<br />

vesicular stomatitis virus), 5 mg of the pMDLg/pRRE lentivirus packaging<br />

plasmid, 2.5 mg pRSV-REV (expressing human immunodeficiency virus 1<br />

reverse transcriptase under control of the Rous sarcoma virus U3 promoter),<br />

and 10 mg Nod2-MA1 or 3020insC Nod2–MA1 expression plasmid (the<br />

bidirectional vector MA1). Then, 2 d after transfection, supernatants were<br />

cleared by centrifugation for 10 min at 1,000g and were pelleted for 90 min at<br />

50,000g. Pelleted viruses were resuspended at 4 1C in DMEM. Human<br />

monocytes (0.5 10 6 cells per condition) were infected with 5 ml or50ml<br />

of lentivirus-containing supernatant on day 1 or day 3, respectively, before<br />

being stimulated with LPS on day 4.<br />

Patients and genotyping of NOD2 mutations. Blood was collected from<br />

patients with Crohn’s disease (n ¼ 74) and healthy volunteers (n ¼ 20). NOD2<br />

gene fragments containing the 3020insC site were amplified by PCR in 50-ml<br />

reaction volumes containing 100 200 ng genomic DNA and the appropriate<br />

primers in various concentrations (Supplementary Table 1 online) in 10 mM<br />

Tris-HCl pH 9.0, 50 mM KCl, 1.5 mM MgCl 2, 0.01% (wt/vol) gelatin, 0.1%<br />

(vol/vol) Triton X-100, 0.35 mM dNTPs and 2 U Taq DNA polymerase<br />

(Invitrogen). Samples were denatured at 92 1C for 5 min and then were<br />

amplified in a PTC-200 thermal cycler (MJ Research; Biozym) by 35 cycles of<br />

92 1C for1min,59.81C for 1 min and 72 1C for 1 min, followed by a final<br />

extension step of 72 1Cfor3min.The3020insC polymorphism was analyzed by<br />

Genescan on an ABI Prism 3100 Genetic Analyzer according to the manufacturer’s<br />

protocol (Applied Biosystems). Three patients with Crohn’s disease<br />

homozygous for the 3020insC mutation were selected for further studies. Three<br />

patients with Crohn’s disease bearing the wild-type NOD2 allele and three<br />

healthy people, matched for age and sex (23–38 years of age; male), served as<br />

controls. All patients with Crohn’s disease were in remission and were not<br />

treated with steroid drugs or antibiological therapy (such as anti–tumor<br />

necrosis factor) during the 2 months before the study. None of the people<br />

selected had the NOD2 mutations of C to T at position 2104 (R702W) or G to<br />

C at position 2722 (G908R), also known to be associated with Crohn’s disease<br />

(data not shown) 37 .<br />

Additional methods. Information on cell culture, cytokine measurement,<br />

reporter plasmids, preparation of nuclear extracts, immunoblot and immunoprecipitation,<br />

ChIP assay, siRNA, nanoflow liquid chromatography–tandem<br />

mass spectrometry analysis and the database search of tandem mass spectrometry<br />

data for the identification of peptide sequences is available in the<br />

Supplementary Methods online.<br />

Statistical analysis. Student’s t-test (one-tailed) was used for data analysis<br />

where appropriate.<br />

Accession codes. UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signalinggateway.org):<br />

A001243, A001137 and A001717.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank P.J. Murray (St. Jude Children’s Research Hospital) for Nod2knockout<br />

and control littermate mice; E. Pamer (Memorial Sloan-Kettering<br />

Cancer Center) for RICK-knockout (RIP2-knockout) and control mice;<br />

J.M. Park (Harvard University School of Medicine) for mice with conditional<br />

knockout of p38a; and G. Nunez (University of Michigan) for the mouse Nod2<br />

expression vector and mouse 2939insC Nod2. Supported by the Broad<br />

Medical Research Program (IBD-210R2 to X.M.).<br />

AUTHOR CONTRIBUTIONS<br />

E.N. contributed to the work in Figures 3,4,6,8; Y.H. contributed to the work in<br />

Figures 1–7; X.K. contributed to Figure 4; M.G.N. contributed to Figures 1,7,8;<br />

and X.M. contributed to the overall project.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Podolsky, D.K. Inflammatory bowel disease. N. Engl. J. Med. 347, 417–429 (2002).<br />

2. Fuss, I.J., Boirivant, M., Lacy, B. & Strober, W. The interrelated roles of TGF-b and<br />

IL-10 in the regulation of experimental colitis. J. Immunol. 168, 900–908 (2002).<br />

478 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

3. Inohara, N. & Nunez, G. NODs: intracellular proteins involved in inflammation and<br />

apoptosis. Nat. Rev. Immunol. 3, 371–382 (2003).<br />

4. Girardin, S.E. et al. Nod1 detects a unique muropeptide from gram-negative bacterial<br />

peptidoglycan. Science 300, 1584–1587 (2003).<br />

5. Girardin, S.E. et al. Nod2 is a general sensor of peptidoglycan through muramyl<br />

dipeptide (MDP) detection. J. Biol. Chem. 278, 8869–8872 (2003).<br />

6. Ogura, Y. et al. Nod2, a Nod1/Apaf-1 family member that is restricted to monocytes and<br />

activates NF-kB. J. Biol. Chem. 276, 4812–4818 (2001).<br />

7. Abbott, D.W., Wilkins, A., Asara, J.M. & Cantley, L.C. The Crohn’s disease protein,<br />

NOD2, requires RIP2 in order to induce ubiquitinylation of a novel site on NEMO. Curr.<br />

Biol. 14, 2217–2227 (2004).<br />

8. Inohara, N. et al. An induced proximity model for NF-kB activationintheNod1/RICK<br />

and RIP signaling pathways. J. Biol. Chem. 275, 27823–27831 (2000).<br />

9. Hugot, J.P. et al. Association of NOD2 leucine-rich repeat variants with susceptibility to<br />

Crohn’s disease. <strong>Nature</strong> 411, 599–603 (2001).<br />

10. Lesage, S. et al. CARD15/NOD2 mutational analysis and genotype-phenotype correlation<br />

in 612 patients with inflammatory bowel disease. Am. J. Hum. Genet. 70,<br />

845–857 (2002).<br />

11. Ogura, Y. et al. A frameshift mutation in NOD2 associated with susceptibility to Crohn’s<br />

disease. <strong>Nature</strong> 411, 603–606 (2001).<br />

12. Seiderer, J. et al. Predictive value of the CARD15 variant 1007fs for the diagnosis<br />

of intestinal stenoses and the need for surgery in Crohn’s disease in clinical<br />

practice: results of a prospective study. Inflamm. Bowel Dis. 12, 1114–1121<br />

(2006).<br />

13. Kuhn, R., Lohler, J., Rennick, D., Rajewsky, K. & Muller, W. Interleukin-10-deficient<br />

mice develop chronic enterocolitis. Cell 75, 263–274 (1993).<br />

14. Davidson, N.J. et al. T helper cell 1-type CD4 + T cells, but not B cells, mediate colitis in<br />

interleukin 10-deficient mice. J. Exp. Med. 184, 241–251 (1996).<br />

15. Sellon, R.K. et al. Resident enteric bacteria are necessary for development of<br />

spontaneous colitis and immune system activation in interleukin-10-deficient mice.<br />

Infect. Immun. 66, 5224–5231 (1998).<br />

16. Kullberg, M.C. et al. Helicobacter hepaticus triggers colitis in specific-pathogen-free<br />

interleukin-10 (IL-10)-deficient mice through an IL-12- and g interferon-dependent<br />

mechanism. Infect. Immun. 66, 5157–5166 (1998).<br />

17. Netea, M.G. et al. NOD2 mediates anti-inflammatory signals induced by TLR2<br />

ligands: implications for Crohn’s disease. Eur. J. Immunol. 34, 2052–2059<br />

(2004).<br />

18. Maeda, S. et al. Nod2 mutation in Crohn’s disease potentiates NF-kB activity and IL-1b<br />

processing. Science 307, 734–738 (2005).<br />

19. Pauleau, A.L. & Murray, P.J. Role of nod2 in the response of macrophages to toll-like<br />

receptor agonists. Mol. Cell. Biol. 23, 7531–7539 (2003).<br />

20. Netea, M.G. et al. Nucleotide-binding oligomerization domain-2 modulates specific<br />

TLR pathways for the induction of cytokine release. J. Immunol. 174, 6518–6523<br />

(2005).<br />

ARTICLES<br />

21. Cao, S., Liu, J., Song, L. & Ma, X. The protooncogene c-Maf is an essential transcription<br />

factor for IL-10 gene expression in macrophages. J. Immunol. 174, 3484–3492<br />

(2005).<br />

22. Steiner, G., Skriner, K., Hassfeld, W. & Smolen, J.S. Clinical and immunological<br />

aspects of autoantibodies to RA33/hnRNP-A/B proteins–a link between RA, SLE and<br />

MCTD. Mol. Biol. Rep. 23, 167–171 (1996).<br />

23. Caporali, R., Bugatti, S., Bruschi, E., Cavagna, L. & Montecucco, C. Autoantibodies to<br />

heterogeneous nuclear ribonucleoproteins. Autoimmunity 38, 25–32 (2005).<br />

24. Jones, D.A. et al. Identification of autoantigens in psoriatic plaques using expression<br />

cloning. J. Invest. Dermatol. 123, 93–100 (2004).<br />

25. Netea, M.G. et al. NOD2 3020insC mutation and the pathogenesis of Crohn’s disease:<br />

impaired IL-1b production points to a loss-of-function phenotype. Neth. J. Med. 63,<br />

305–308 (2005).<br />

26. Brockstedt, E. et al. Identification of apoptosis-associated proteins in a human Burkitt<br />

lymphoma cell line. Cleavage of heterogeneous nuclear ribonucleoprotein A1 by<br />

caspase 3. J. Biol. Chem. 273, 28057–28064 (1998).<br />

27. Kim, C. et al. The kinase p38a serves cell type-specific inflammatory functions in skin<br />

injury and coordinates pro- and anti-inflammatory gene expression. Nat. Immunol. 9,<br />

1019–1027 (2008).<br />

28. Buxade, M. et al. The Mnks are novel components in the control of TNF a biosynthesis<br />

and phosphorylate and regulate hnRNP A1. Immunity 23, 177–189 (2005).<br />

29. Linde, K. et al. Card15 and Crohn’s disease: healthy homozygous carriers of the<br />

3020insC frameshift mutation. Am. J. Gastroenterol. 98, 613–617 (2003).<br />

30. Yamazaki, K., Takazoe, M., Tanaka, T., Kazumori, T. & Nakamura, Y. Absence of<br />

mutation in the NOD2/CARD15 gene among 483 Japanese patients with Crohn’s<br />

disease. J. Hum. Genet. 47, 469–472 (2002).<br />

31. Wehkamp, J. et al. NOD2 (CARD15) mutations in Crohn’s disease are associated with<br />

diminished mucosal a-defensin expression. Gut 53, 1658–1664 (2004).<br />

32. Kramer, M., Netea, M.G., de Jong, D.J., Kullberg, B.J. & Adema, G.J. Impaired<br />

dendritic cell function in Crohn’s disease patients with NOD2 3020insC mutation.<br />

J. Leukoc. Biol. 79, 860–866 (2006).<br />

33. van Heel, D.A. et al. Muramyl dipeptide and toll-like receptor sensitivity in NOD2associated<br />

Crohn’s disease. Lancet 365, 1794–1796 (2005).<br />

34. Holler, E. et al. Both donor and recipient NOD2/CARD15 mutations associate with<br />

transplant-related mortality and GvHD following allogeneic stem cell transplantation.<br />

Blood 104, 889–894 (2004).<br />

35. Watanabe, T., Kitani, A., Murray, P.J. & Strober, W. NOD2 is a negative regulator of Tolllike<br />

receptor 2–mediated T helper type 1 responses. Nat. Immunol. 5, 800–808 (2004).<br />

36. Liu, J., Cao, S., Herman, L.M. & Ma, X. Differential regulation of interleukin (IL)-12<br />

p35 and p40 gene expression and interferon (IFN)-g-primed IL-12 production by IFN<br />

regulatory factor 1. J. Exp. Med. 198, 1265–1276 (2003).<br />

37. Netea, M.G. et al. NOD2 mediates induction of the antiinflammatory signals induced by<br />

TLR2-ligands: implications for Crohn’s disease. Eur. J. Immunol. 34, 2052–2059<br />

(2004).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 479


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Autophagy enhances the presentation of endogenous<br />

viralantigensonMHCclassImoleculesduring<br />

HSV-1 infection<br />

Luc English 1 , Magali Chemali 1 , Johanne Duron 1 , Christiane Rondeau 1 , Annie Laplante 1 , Diane Gingras 1 ,<br />

Diane Alexander 2 , David Leib 2 , Christopher Norbury 3 , Roger Lippé 1 & Michel Desjardins 1,4,5<br />

Viral proteins are usually processed by the ‘classical’ major histocompatibility complex (MHC) class I presentation pathway.<br />

Here we showed that although macrophages infected with herpes simplex virus type 1 (HSV-1) initially stimulated CD8 + T cells<br />

by this pathway, a second pathway involving a vacuolar compartment was triggered later during infection. Morphological and<br />

functional analyses indicated that distinct forms of autophagy facilitated the presentation of HSV-1 antigens on MHC class I<br />

molecules. One form of autophagy involved a previously unknown type of autophagosome that originated from the nuclear<br />

envelope. Whereas interferon-c stimulated classical MHC class I presentation, fever-like hyperthermia and the pyrogenic<br />

cytokine interleukin 1b activated autophagy and the vacuolar processing of viral peptides. Viral peptides in autophagosomes<br />

were further processed by the proteasome, which suggests a complex interaction between the vacuolar and MHC class I<br />

presentation pathways.<br />

The elaboration of an efficient immune response against pathogens<br />

involves complex intracellular antigen-processing events. Endogenous<br />

antigens such as viral proteins synthesized by infected host cells are<br />

degraded in the cytoplasm by the proteasome, and the resulting<br />

peptides are translocated into the endoplasmic reticulum, where<br />

they are loaded onto major histocompatibility complex (MHC) class I<br />

molecules. In contrast, exogenous antigens are processed by hydrolases<br />

in lytic endovacuolar compartments and are loaded on MHC class II<br />

molecules that reach the cell surface using recycling machineries<br />

associated with these organelles. Although initially thought<br />

to be strictly segregated, these pathways are actually functionally<br />

interconnected, as shown by the ability of cells to present exogenous<br />

antigens on MHC class I molecules; this process is referred<br />

to as ‘cross-presentation’ 1 .<br />

Autophagy, a process that allows the transfer of endogenous cellular<br />

components into lytic vacuolar compartments, has been shown to be<br />

essential to both innate and adaptive immunity 2–4 . This process can<br />

be used to eliminate intracellular bacteria and viruses 5–11 . Furthermore,<br />

autophagy can facilitate the presentation of endogenous<br />

antigens on MHC class II molecules, thereby leading to activation of<br />

CD4 + T cells 12,13 . The nature of the endovacuolar membranes<br />

involved in the formation of autophagosomes is still a matter of active<br />

investigation, but the endoplasmic reticulum is one potential source<br />

of autophagosome membrane 14 . As substantial amounts of viral<br />

Received 23 December 2008; accepted 12 February 2009; published online 22 March 2009; doi:10.1038/ni.1720<br />

membrane glycoproteins are synthesized in the endoplasmic reticulum<br />

of infected cells, it is likely that some of these antigens will reach lytic<br />

vacuolar organelles during autophagy. Despite the fact that several<br />

reports have demonstrated that antigens generated in the lumen of<br />

phagosomes can be loaded (or ‘cross-presented’) on MHC class I<br />

molecules and trigger a CD8 + T cell response 15–17 , it is not known<br />

whether a similar process could occur after autophagy. Here we<br />

provide evidence that infection of macrophages with herpes simplex<br />

virus type 1 (HSV-1) triggered a vacuolar response that increased<br />

the presentation of a peptide of HSV-1 glycoprotein B (gB) to<br />

CD8 + T cells on MHC class I molecules. This vacuolar response,<br />

linked to autophagy, could be modulated by various cytokines and<br />

stress conditions.<br />

RESULTS<br />

Two phases of MHC class I presentation<br />

IncubationofBMA3.1A7(called‘BMA’here)macrophageswithHSV-1<br />

expressing a green fluorescent protein (GFP)-tagged capsid protein<br />

(K26-GFP) resulted in the infection of about 35% of the cells, as<br />

determined by flow cytometry (data not shown). Fluorescence microscopy<br />

showed that K26-GFP capsids assembled in the nucleus of<br />

infected cells at about 6–8 h after infection, reaching a maximum at<br />

about 12 h after infection (data not shown). Starting at 6 h after<br />

infection, infected macrophages stimulated a CD8 + T cell hybridoma<br />

1 Département de Pathologie et Biologie Cellulaire, Université de Montréal, Succursale Centre-Ville, Montreal, Quebec, Canada. 2 Department of Ophthalmology and Visual<br />

Sciences, Washington University School of Medicine, St. Louis, Missouri, USA. 3 Department of Microbiology and <strong>Immunology</strong>, Pennsylvania State University, Milton S.<br />

Hershey College of Medicine, Hershey, Pennsylvania, USA. 4 Département de microbiologie et immunologie, Université de Montréal, Succursale Centre-Ville, Montreal,<br />

Quebec, Canada. 5 Caprion Pharmaceuticals, Montréal, Québec, Canada. Correspondence should be addressed to M.D. (michel.desjardins@umontreal.ca)<br />

480 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Figure 1 A vacuolar pathway participates in the processing of<br />

endogenous viral proteins for presentation on MHC class I molecules.<br />

(a) Activation of the gB-specific CD8 + T cell hybridoma (which<br />

expresses b-galactosidase as an indicator of T cell activation)<br />

by macrophages infected for various times (horizontal axis) with HSV-1,<br />

then incubated for 12 h at 37 1C with the hybridoma. A595, absorbance<br />

at 595 nm. (b) Activation of the hybridoma as described in a, withthe<br />

addition of dimethyl sulfoxide (DMSO; negative control), bafilomycin A<br />

(Baf), brefeldin A (BFA) or MG-132 at 2 h after macrophage infection.<br />

(c) Activation of the hybridoma as described in a, with the addition of<br />

bafilomycin A at 2 h after macrophage infection. (d) Immunofluorescence<br />

microscopy of gB (blue) and LAMP-1 (red) in HSV-1infected<br />

macrophages; pink indicates colocalization. Original<br />

magnification, 40. (e) CD8 + T cell–stimulatory capacity (as described<br />

in a) of uninfected macrophages (Mock), of BMA (BMA-HSV) or J774<br />

(J774-HSV) macrophages infected for 8 h with HSV-1, and of cocultures<br />

of J774 macrophages (H-2 d ) infected for 1 h with HSV-1, then mixed<br />

with uninfected BMA (H-2 b ) macrophages at a ratio of 1:1 and cultured<br />

together for 8 h (J774-HSV + BMA). Results in b,c,e are normalized to<br />

results obtained for CD8 + T cells stimulated with macrophages treated<br />

with DMSO (b,c) or infected BMA macrophages (e) and are presented in<br />

arbitrary units. Data are from one representative of three independent experiments (mean and s.d. of triplicate samples; a), are from three independent<br />

experiments (mean and s.e.m. of triplicate samples error bars; b,c,e) or are representative of three independent experiments (d).<br />

specific for a peptide of amino acids 498–505 of HSV-1 gB 18,19 ,as<br />

measured by the release of b-galactosidase after T cell activation<br />

(Fig. 1a). The capacity of macrophages to stimulate CD8 + T cells<br />

continued to increase up to 12 h after infection, the time at which<br />

cellular mortality induced by the viral infection began to occur (data<br />

not shown). This macrophage cell death probably explains the<br />

decrease in CD8 + T cell stimulation between 12 h and 14 h after<br />

infection (Fig. 1a).<br />

Stimulation of the CD8 + T cell hybridoma was much lower after<br />

treatment of macrophages with the proteasome inhibitor MG-132 or<br />

with brefeldin A, a drug that inhibits the transport of molecules<br />

through the biosynthetic pathway (Fig. 1b). These results indicate that<br />

processing and presentation of the viral antigen gB in infected<br />

a<br />

Time (h)<br />

2<br />

4<br />

6<br />

8<br />

e<br />

β-galactosidase<br />

activity (relative)<br />

2<br />

1<br />

0<br />

LC3 Phase-contrast b<br />

Basal<br />

Basal + Baf<br />

Rapa<br />

Rapa + Baf<br />

HS<br />

HS + Baf<br />

β-galactosidase<br />

activity (relative)<br />

c<br />

β-galactosidase<br />

activity (relative)<br />

f<br />

β-galactosidase activity<br />

(relative)<br />

1.5<br />

1.0<br />

0.5<br />

2<br />

1<br />

0<br />

0<br />

2<br />

1<br />

0<br />

Rapa:<br />

DMEM<br />

DMEM + 3-MA<br />

8 10 12<br />

Time after infection (h)<br />

Ctrl siRNA<br />

Atg5 siRNA<br />

8 12 8 12<br />

Time after<br />

infection (h)<br />

– + – +<br />

d<br />

Ctrl<br />

siRNA Atg5<br />

siRNA<br />

Ctrl siRNA<br />

Atg5 siRNA<br />

Atg5<br />

Tubulin<br />

a b c 2<br />

β-galactosidase activity<br />

(A595 )<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

4 6 8 101214<br />

Time after infection (h)<br />

β-galactosidase activity<br />

(relative)<br />

d e<br />

1.0<br />

0.5<br />

0<br />

DMSO<br />

Baf<br />

BFA<br />

MG-132<br />

β-galactosidase activity<br />

(relative)<br />

gB LAMP-1 Merge<br />

ARTICLES<br />

0<br />

8 10 12<br />

Time after infection (h)<br />

macrophages involves the ‘classical’ endogenous pathway of MHC<br />

class I presentation. However, bafilomycin A, a drug that inhibits the<br />

vacuolar proton pump and the acidification of endosomes and<br />

lysosomes, had only a minimal effect during the early period of<br />

infection (up to 8 h after infection) but strongly inhibited the capacity<br />

of infected macrophages to stimulate CD8 + T cells at 10 h and 12 h<br />

after infection (Fig. 1c). These results suggest that the initial processing<br />

of endogenous gB by the classical pathway is followed by the<br />

engagement of a vacuolar pathway that considerably improves the<br />

processing of gB and the activation of CD8 + T cells. The possibility of<br />

a contribution by a vacuolar pathway was supported by immunofluorescence<br />

analyses that indicated that endogenous gB localized<br />

together with the endo-lysosomal marker LAMP-1 during infection<br />

(Fig. 1d). The possibility of the presence of gB in degradative<br />

compartments was further supported by the finding of higher gB<br />

expression in infected macrophages treated with bafilomycin (data not<br />

shown). These results raised the issue of how gB reaches the lysosomal<br />

1<br />

β-galactosidase<br />

activity (relative)<br />

DMSO<br />

Baf<br />

1.0<br />

0.5<br />

0<br />

Mock<br />

BMA-HSV<br />

J774-HSV + BMA<br />

Figure 2 Autophagy induced during HSV-1 infection contributes to the<br />

processing and presentation of endogenous viral antigens on MHC class I<br />

molecules. (a) Immunofluorescence microscopy of LC3 expression in<br />

macrophages infected for 2–8 h (left margin) with HSV-1. Original<br />

magnification, 20. (b) Activation of the gB-specific CD8 + T cell hybridoma<br />

(described in Fig. 1a) by macrophages infected for various times (horizontal<br />

axis) with HSV-1, with (DMEM + 3-MA) or without (DMEM) the addition of<br />

3-methyladenine 2 h after infection, then incubated for 12 h at 37 1C with<br />

the hybridoma. (c) Activation of the hybridoma (as described in b) by<br />

macrophages transfected for 60 h with control (Ctrl) siRNA or Atg5-specific<br />

siRNA, then infected for 8 h or 12 h with HSV-1. (d) Immunoblot analysis<br />

of Atg-5 in siRNA-treated macrophages. (e) Activation of the hybridoma<br />

(as described in b) by macrophages infected with HSV-1 and incubated at<br />

37 1C (Basal), incubated for 12 h at 39 1C before being infected with<br />

HSV-1 (heat shock (HS)), or treated with rapamycin during HSV-1 infection<br />

(Rapa), with (+ Baf) or without the addition of bafilmycin A at 2 h after<br />

infection. (f) Activation of the hybridoma (as described in b) by macrophages<br />

transfected for 60 h with control siRNA or Atg5-specific siRNA, then infected<br />

for 8 h with HSV-1, with (+) or without (–) the addition of rapamycin at<br />

2 h after infection. Results in b,c,e,f are normalized to results obtained for<br />

CD8 + T cells stimulated with macrophages infected for 8 h at 37 1C without<br />

further treatment (b,e) or infected macrophages treated with control siRNA<br />

(c,f) and are presented in arbitrary units. Data are representative of three<br />

independent experiments (a,d) or are from three independent experiments<br />

(mean and s.e.m. of triplicate samples; b,c,e,f).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 481


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

degradative pathway. The possibility of transfer of gB to lysosomes<br />

through phagocytosis of infected cells (cross-presentation) was ruled<br />

out by results showing that incubation of H-2 b BMA macrophages<br />

together with HSV-1-infected H-2 d J774 macrophages did not result<br />

in activation of the H-2 b -specific CD8 + T cell hybridoma (Fig. 1e).<br />

Hence, vacuolar processing of gB in the later period of infection was<br />

more likely to involve membrane-trafficking events that occurred<br />

exclusively in the infected cell.<br />

Data have shown that endogenous viral proteins can be presented<br />

on MHC class II molecules by a process involving autophagy 9 ,which<br />

indicates that trafficking events that enable the transport of endogenous<br />

proteins to vacuolar degradative organelles can occur in virusinfected<br />

cells. To determine if autophagy was involved in the late<br />

processing of endogenous gB and its presentation on MHC class I<br />

molecules, we first monitored the presence of LC3, a marker of<br />

autophagy, in macrophages at various times after infection<br />

(Fig. 2a). Although we did not detect it in uninfected cells, we<br />

found LC3 beginning at 4–6 h after infection, which indicated that<br />

an autophagic response occurred during the late phase of HSV-1<br />

infection in macrophages. Whereas the inhibitory effect of bafilomycin<br />

indicated that a vacuolar response of some kind was triggered during<br />

the late phase of infection in macrophages (Fig. 1c), the possibility<br />

of a contribution of autophagy to this process was suggested by<br />

the substantial inhibition of the CD8 + T cell–stimulatory capacity<br />

of macrophages treated with 3-methyladenine, a commonly used<br />

a<br />

b<br />

c<br />

g<br />

Control<br />

HS<br />

Rapa<br />

Time (h)<br />

IB: α-HSV<br />

d Time (h)<br />

WT ∆34.5<br />

0 6 8 10 6 8 10<br />

4<br />

6<br />

8<br />

10<br />

(kDa)<br />

250<br />

100<br />

75<br />

50<br />

37<br />

25<br />

20<br />

VP26 gB LC3 Merge<br />

h<br />

Time (h)<br />

1,200<br />

Counts<br />

960<br />

620<br />

480<br />

240<br />

6 8<br />

1,200<br />

960<br />

620<br />

480<br />

240<br />

inhibitor of autophagy (Fig. 2b). Confirmation of the involvement of<br />

autophagy in the processing and presentation of gB peptides on MHC<br />

class I molecules was provided by experiments involving small interfering<br />

RNA (siRNA)-mediated silencing of Atg5, a protein involved in<br />

the formation of autophagosomes 20 . Macrophages treated with a<br />

control siRNA had a greater capacity to stimulate CD8 + T cells<br />

between 8 h and 12 h after infection, but macrophages treated with<br />

Atg5-specific siRNA did not (Fig. 2c),whichlinkedthislategainin<br />

stimulation to the induction of autophagy.<br />

Further support for the idea that autophagy contributes to the<br />

vacuolar processing and presentation of gB on MHC class I molecules<br />

was provided by results indicating that treatment of infected macrophages<br />

with rapamycin, an inhibitor of the kinase mTOR that<br />

stimulates autophagy 21 ,considerablyimprovedCD8 + T cell stimulation<br />

(Fig. 2d). We obtained similar results with macrophages exposed<br />

to a mild heat shock before infection (39 1C for 12 h), a condition also<br />

known to induce autophagy 22 (Fig. 2d). The enhanced CD8 + T cell<br />

stimulation induced by mTOR or heat shock was abolished by the<br />

addition of bafilomycin (Fig. 2e). These data further link the vacuolar<br />

processing of gB to autophagy. We obtained similar results with mouse<br />

embryonic fibroblasts isolated from Atg5 –/– mice 20 (Supplementary<br />

Fig. 1a online). Silencing of Atg5 in infected macrophages with siRNA<br />

abolished the effect of rapamycin (Fig. 2f), which confirmed the<br />

specificity of this drug for the autophagic pathway. Similarly, neither<br />

bafilomycin (Supplementary Fig. 1c,d) nor 3-methyladenine (data<br />

e<br />

6 h<br />

8 h<br />

∆34.5<br />

WT<br />

∆34.5<br />

WT<br />

1,200<br />

gB LC3 Overlay<br />

10<br />

Ctrl<br />

WT 17+<br />

∆34.5<br />

0<br />

0<br />

0<br />

100 101 102 103 104 100 101 102 103 104 100 101 102 103 104 gB<br />

960<br />

620<br />

480<br />

240<br />

f<br />

β-galactosidase<br />

activity (relative)<br />

2.5<br />

WT 17+<br />

∆34.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

6 8 10<br />

Time after<br />

infection (h)<br />

i WT 17+<br />

∆34.5<br />

1.00<br />

0.75<br />

0.50<br />

0.25<br />

0<br />

Baf: – + – + – + – + – + – +<br />

6 8 10<br />

Time after infection (h)<br />

Figure 3 Both gB and LC3 accumulate in perinuclear regions during HSV-1 infection. (a–c) Immunofluorescence microscopy of uninfected macrophages<br />

incubated at 37 1C (control; a), subjected to mild heat shock (b) or treated with rapamycin (c), then stained with anti-LC3. (d) Immunofluorescence<br />

microscopy of the expression of LC3, gB and GFP (VP26) by macrophages infected for various times (left margin) with HSV-1. White indicates colocalization.<br />

(e) Immunofluorescence microscopy of macrophages infected for 6 h or 8 h (left margin) with wild-type HSV-1 (WT) or HSV-1 lacking ICP34.5 (D34.5).<br />

Blue, staining of nuclei with DAPI (4,6-diamidino-2-phenylindole). Original magnification, 100 (a–c) or 63 (d,e). Results in a–e are representative of<br />

three independent experiments. (f) Activation of the gB-specific CD8 + T cell hybridoma (as described in Fig. 1a) by macrophages infected for various times<br />

(horizontal axis) with wild-type HSV-1 strain 17+ (WT 17+) or D34.5 HSV-1. Data are from three independent experiments (mean and s.e.m. of triplicate<br />

samples). (g,h) Immunoblot analysis (IB; g) and flow cytometry (h) of the expression of HSV-1 proteins (g) and gB (h) in macrophages infected for various<br />

times (above lanes (g) or plots (h)) with wild-type or D34.5 HSV-1. Ctrl, control (uninfected BMA macrophages; h). Data are representative of two (g) or<br />

three (h) independent experiments. (i) Activation of the gB-specific CD8 + T cell hybridoma (as described in Fig. 1a) by macrophages infected for various<br />

times (below graph) with wild-type or D34.5 HSV-1, with (+) or without (–) the addition of bafilomycin A at 2 h after infection. Results in f,i are normalized<br />

to results obtained for CD8 + T cells stimulated with macrophages infected for 6 h with wild-type virus (f) or with infected macrophages incubated without<br />

bafilomycin (i) and are presented in arbitrary units. Data are from three independent experiments (mean and s.e.m. of triplicate samples).<br />

482 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

β-galactosidase<br />

activity (relative)


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

b<br />

N<br />

VP<br />

c d<br />

e f<br />

N<br />

not shown) affected the capacity of infected macrophages to stimulate<br />

CD8 + T cells when autophagy was inhibited by silencing of Atg5. Our<br />

results so far indicated that a vacuolar pathway linked to autophagy,<br />

triggered within 8–10 h of infection, enhanced the ability of infected<br />

macrophages to stimulate gB-specific CD8 + T cells.<br />

Late-stage autophagy involving the nuclear envelope<br />

To study the autophagic response associated with HSV-1 infection in<br />

macrophages, we used immunofluorescence and electron microscopy.<br />

We first analyzed the induction of autophagy by assessing the<br />

autophagic marker LC3. We noted a weak signal for LC3 in uninfected<br />

macrophages (Fig. 3a), but uninfected cells submitted to a mild heat<br />

shock (Fig. 3b) or treated with rapamycin (Fig. 3c) had strong<br />

punctate LC3 signals in the cytoplasm. In contrast, there was a strong<br />

LC3 signal in close association with the nuclear envelope in cells 6–8 h<br />

after HSV-1 infection (Fig. 3d,e). In many cases, vesicles strongly<br />

labeled for LC3 seemed to be connected to the nuclear envelope. The<br />

colocalization of LC3 and gB suggested that autophagic structures<br />

containing viral proteins might originate in the vicinity of the nucleus<br />

at a late phase of infection. The difference between the typical labeling<br />

noted when autophagy was triggered by rapamycin and that in<br />

infected cells might be linked to the fact that HSV-1 infection has<br />

been shown to inhibit macroautophagy 23 .<br />

The accumulation of LC3 around the nucleus was possibly associated<br />

with a cellular process distinct from classical macroautophagy<br />

and induced as a late response to infection. To test that hypothesis, we<br />

compared the distribution of gB and LC3 in macrophages infected<br />

with wild type HSV-1 and a mutant HSV-1 lacking the ICP34.5<br />

protein (D34.5); unlike the wild-type virus, this mutant virus is unable<br />

N<br />

N<br />

VP<br />

VP<br />

N<br />

ARTICLES<br />

Figure 4 HSV-1 induces the formation of autophagosome-like structures<br />

from the nuclear envelopes of infected macrophages. Electron microscopy<br />

of macrophages 10 h after infection with HSV-1. (a) Arrows indicate<br />

membrane-coiled structures emerging from the nucleus of an infected cell.<br />

(b–d) Four-layered membrane structures formed by coiling of the nuclear<br />

membrane. (e,f) Glucose-6-phosphatase (black deposits) on autophagosomelike<br />

structures emerging from the nuclear envelope or free in the<br />

cytoplasm, and viral capsids in the cytoplasm engulfed in the lumen of an<br />

autophagosome-like compartment. N, nucleus; VP, viral particles. Scale<br />

bars, 1 mm (a), 0.25 mm (b–d,f) or0.4mm (e). Images are representative<br />

of three independent experiments with at least 100 cell profiles in each.<br />

to inhibit macroautophagy. As expected, the D34.5 virus failed to<br />

express any detectable ICP34.5 protein (data not shown). In addition,<br />

consistent with the ability of ICP34.5 to mediate dephosphorylation of<br />

the translation-initiation factor eIF2a 24 ,theD34.5 virus induced more<br />

phosphorylated eIF2a than did wild-type HSV-1 (data not shown).<br />

Macrophages infected with the D34.5 virus showed considerable<br />

accumulation of LC3 on vesicular structures, which also contained<br />

large amounts of gB and were present throughout the cytoplasm<br />

(Fig. 3e). We found no apparent labeling for LC3 in the vicinity of the<br />

nuclear envelope. In contrast, infection with the corresponding wildtype<br />

virus (strain 17+) led to the accumulation of LC3 near the<br />

nuclear envelope between 6 h and 8 h after infection (Fig. 3e); these<br />

results are in agreement with results obtained with the KOS wild-type<br />

strain (Fig. 3d). These findings collectively suggested that distinct<br />

types of autophagic structures were induced in response to the D34.5<br />

and wild type viruses. Macrophages infected with D34.5 or wild-type<br />

HSV-1 (at an identical multiplicity of infection) were able to stimulate<br />

gB-specific CD8 + T cells to a similar extent (Fig. 3f). However, the<br />

expression of HSV-1 proteins (Fig. 3g) and gB (Fig. 3h) was much<br />

lower in macrophages infected with the D34.5 virus, in agreement with<br />

published studies 25,26 . Thus, macrophages infected with the D34.5<br />

virus stimulated gB-specific CD8 + T cells much more efficiently.<br />

Bafilomycin strongly inhibited the stimulatory ability of macrophages<br />

infected with either D34.5 or wild-type HSV-1 (Fig. 3i), which<br />

emphasizes the involvement of a vacuolar pathway in both cases.<br />

To characterize the autophagosomal structures induced during the<br />

late phase of infection, we used electron microscopy for detailed<br />

morphological analysis (Fig. 4 and Supplementary Fig. 2 online). We<br />

noted two types of prominent structures in infected cells. The first was<br />

characterized by the presence of double-membraned structures<br />

reminiscent of the morphology of autophagosomes found in cells<br />

treated with rapamycin (Supplementary Fig. 2a). These structures<br />

often surrounded viral particles in the cytoplasm (Supplementary<br />

Fig. 2b). The degradation of microbial invaders by autophagy has<br />

been described before and is referred to as ‘xenophagy’ 27 . These<br />

structures showed substantial labeling for LC3 and, to a lesser extent,<br />

for gB (Supplementary Fig. 2c,d), which confirmed the presence of<br />

viral proteins in autophagosomes. The second type of organelle had<br />

multiple membranes either connected to the nuclear envelope or<br />

present in the cytoplasm (Fig. 4a,e). These structures seemed to<br />

emerge through a coiling process of the inner and outer nuclear<br />

membrane, forming four-layered structures that engulfed part of the<br />

nearby cytoplasm (Fig. 4b,c). In several examples, similar structures<br />

containing cytoplasm and unenveloped viral capsids seemed to be<br />

disconnected from the nucleus (Fig. 4d).Thesestructureswerenot<br />

present in uninfected or rapamycin-treated cells (Supplementary<br />

Fig. 2e). However, there were more two- and four-membraned structures<br />

in infected macrophages at the late phase of infection (Supplementary<br />

Fig. 2e,f). To determine whether the four-layered membrane<br />

structures present in the cytoplasm were similar to those that emerged<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 483


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a LC3 b LC3<br />

gB<br />

c d<br />

N<br />

BSA-gold<br />

Figure 5 The four-layered membrane structures that emerge from the<br />

nuclear envelope have autophagosome-like features. Immunoelectron<br />

microscopy of macrophages 10 h after infection with HSV-1.<br />

(a–c) Accumulation of LC3 (a,b) and gB (c). (d) Fusion of four-layered<br />

membrane structures and lysosomes preloaded with bovine serum albumin–<br />

gold (BSA-gold; black dots). Original magnification, 54,800 (a,b),<br />

69,000 (c) or 38,000 (d). Images are representative of three (a–c) or<br />

two (d) independent experiments.<br />

from the nuclear envelope, we used electron microscopy to locate the<br />

product of the enzyme glucose-6-phosphatase, a specific marker of the<br />

endoplasmic reticulum and nuclear envelope. The results indicated<br />

that the product of glucose-6-phosphatase was restricted to the<br />

endoplasmic reticulum and nuclear envelope (Fig. 4e). The fourlayered<br />

membrane structures connected to the nuclear envelope, as<br />

well as those present in the cytoplasm, were also positive for glucose-<br />

6-phosphatase (Fig. 4f), which confirmed their origin in the endoplasmic<br />

reticulum and/or nuclear envelope.<br />

To determine whether the four-layered membrane structures had<br />

features of autophagosomes, we used immunoelectron microscopy to<br />

detect the autophagosome marker LC3. We found accumulation of<br />

LC3 on membrane structures emerging from the nuclear envelope, as<br />

well as on four-layered membrane structures apparently disconnected<br />

from the nucleus and present in the cytoplasm (Fig. 5a,b). The<br />

association between LC3 and the membrane of autophagosomes<br />

Figure 6 Involvement of lytic vacuolar compartments in the processing and<br />

presentation of endogenous antigens on MHC class I molecules after<br />

treatment with proinflammatory cytokines. (a) Activation of the gB-specific<br />

CD8 + T cell hybridoma (as described in Fig. 1a) by macrophages exposed<br />

to DMSO (negative control), mild heat shock, IL-1b or IFN-g.<br />

(b) Dansylcadaverin staining of untreated macrophages (Basal) or<br />

macrophages exposed to mild heat shock, IFN-g or IL-1b and infected<br />

for 8 h with wild-type HSV-1, normalized to the signal obtained in basal<br />

conditions and presented in arbitrary units. (c–f) Activation of the gBspecific<br />

CD8 + T cell hybridoma (as described in Fig. 1a) by macrophages<br />

incubated at 37 1C (c) or exposed to mild heat shock (d), IL-1b (e) or<br />

IFN-g (f) and infected for 8 h with wild-type HSV-1 with the addition of<br />

DMSO (negative control), 3-methyladenine, bafilomycin, brefeldin A or<br />

MG-132 at 2 h after infection. Results in a,c–f are normalized to results<br />

obtained for CD8 + T cells stimulated with macrophages incubated with<br />

DMSO in each condition and are presented in arbitrary units. Data are from<br />

three independent experiments (mean and s.e.m. of triplicate samples).<br />

indicated that our antibody recognized the LC3-II cleaved form of<br />

the protein 28 . Quantitative analysis of the immunolabeling for LC3<br />

showed there was an average (± s.e.m.) of 3.03 ± 0.47 gold particles<br />

per mm membrane on autophagosomes, compared with 0.15 ± 0.02<br />

and 0.70 ± 0.15 gold particles per mm membrane for the plasma<br />

membrane and nuclear membrane, respectively. We also found gB on<br />

the membrane of the nuclear envelope (data not shown), as well as on<br />

the four-layered membrane structures (Fig. 5c). The ability of these<br />

structures to fuse with lytic organelles was confirmed by the presence<br />

of bovine serum albumin–gold particles, transferred from lysosomes,<br />

in these structures (Fig. 5d). These data collectively confirm the<br />

autophagosomal nature of the four-layered membrane structures<br />

originating from the nuclear envelope.<br />

Cytokines engage classical and vacuolar responses<br />

Treatment of macrophages with the proinflammatory cytokine interferon-g<br />

(IFN-g) stimulates the clearance of mycobacteria by a process<br />

involving autophagy 6 . Therefore, we tested whether this cytokine<br />

might promote the vacuolar processing of gB and improve the ability<br />

of HSV-1-infected macrophages to stimulate CD8 + T cells. As heat<br />

treatment of macrophages stimulated autophagy, we also tested the<br />

potential effect of the pyrogenic cytokine interleukin 1b (IL-1b). IFN-g<br />

considerably enhanced the ability of HSV-1-infected macrophages to<br />

stimulate CD8 + T cells (Fig. 6a). IL-1b and mild heat-shock treatment<br />

also augmented the stimulatory capacity of macrophages (Fig. 6a).<br />

Notably, staining with dansylcadaverin, a marker of autophagy, was<br />

greater after exposure to mild heat shock or IL-1b but not after<br />

treatment with IFN-g (Fig. 6b). Electron microscopy also showed that<br />

cells stimulated with heat shock or IL-1b had more autophagosomes<br />

(data not shown). These results suggest that the enhanced ability of<br />

macrophages to stimulate CD8 + T cells after treatment with mild heat<br />

shock or IL-1b was linked to the contribution of a vacuolar pathway<br />

related to autophagy. We confirmed that hypothesis by showing that<br />

treatment of IL-1b- or heat shock–treated macrophages with either<br />

3-methyladenine, the inhibitor of autophagy, or bafilomycin, the<br />

inhibitor of the vacuolar proton pump, resulted in a much lower<br />

capacity of macrophages to stimulate CD8 + T cells than that of control<br />

cells (Fig. 6c–e). In contrast, the very strong stimulation of CD8 +<br />

T cells induced by infected macrophages kept at 37 1C or treated with<br />

a<br />

β-galactosidase activity (relative)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

DMSO<br />

HS<br />

1.0<br />

0.5<br />

0<br />

HS<br />

IL-1β<br />

IFN-γ<br />

DMSO<br />

3-MA<br />

Baf<br />

BFA<br />

MG-132<br />

b<br />

Dansylcadaverin intensity<br />

(relative)<br />

1.3<br />

1.2<br />

1.1<br />

1.0<br />

0.9<br />

0.8<br />

0<br />

IL-1β<br />

1.0<br />

0.5<br />

0<br />

Basal<br />

HS<br />

IL-1β<br />

IFN-γ<br />

c<br />

DMSO<br />

3-MA<br />

Baf<br />

BFA<br />

MG-132<br />

β-galactosidase<br />

activity (relative)<br />

d e f<br />

β-galactosidase<br />

activity (relative)<br />

β-galactosidase<br />

activity (relative)<br />

37 °C<br />

1.0<br />

0.5<br />

0<br />

β-galactosidase<br />

activity (relative)<br />

IFN-γ<br />

1.0<br />

0.5<br />

0<br />

DMSO<br />

3-MA<br />

Baf<br />

BFA<br />

MG-132<br />

DMSO<br />

3-MA<br />

Baf<br />

BFA<br />

MG-132<br />

484 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

IFN-g was not strongly affected by 3-methyladenine or bafilomycin<br />

and therefore was not due to vacuolar processing (Fig. 6c,f). Confirmation<br />

that autophagy did not contribute to the stimulatory effect<br />

of IFN-g was provided by experiments showing that this cytokine<br />

enhanced the capacity of infected embryonic fibroblasts isolated from<br />

wild-type and Atg5 –/– mice to stimulate CD8 + T cells (Supplementary<br />

Fig. 2a). However, the stimulatory effect induced by IL-1b and mild<br />

heat shock was completely abolished by brefeldin A and MG-132<br />

(Fig. 6d,e). The effect of these two inhibitors of the ‘classical’ pathway<br />

of MHC class I presentation indicated a close interaction occurring<br />

between this pathway and the vacuolar pathway in the processing of<br />

gB (Supplementary Fig. 3 online).<br />

DISCUSSION<br />

One of the main components of the complex cell-entry ‘machinery’ of<br />

herpes viruses is gB 29 . This transmembrane protein of the viral<br />

envelope is synthesized in the endoplasmic reticulum of infected<br />

cells. In agreement with published work 18 , we found that gB was<br />

expressed within 2 h after infection in BMA macrophages and was<br />

present in the perinuclear region of infected cells. gB can also<br />

accumulate in the inner membrane as well as the outer membrane<br />

of the nuclear envelope 30,31 . Our results indicated that the processing<br />

of gB by infected macrophages has two distinct phases. During the<br />

first phase, which occurs 6–8 h after infection, gB is processed mainly<br />

by the classical pathway of MHC class I presentation, which involves<br />

proteasome-mediated degradation and transport steps through the<br />

biosynthetic apparatus. During the second phase of infection, which<br />

occurs 8–12 h after infection, a vacuolar pathway is triggered and<br />

contributes substantially to the capacity of infected macrophages to<br />

stimulate CD8 + T cells. The cellular trafficking events that enable the<br />

transport of gB in the lysosomal degradative pathway are poorly<br />

understood. We were able to rule out the possibility of a contribution<br />

by a cross-presentation pathway involving the phagocytosis of infected<br />

cells by neighboring macrophages, which emphasized the fact that gB<br />

reaches the vacuolar processing pathway by an endogenous route.<br />

Instead, our results indicated the involvement of autophagy in facilitating<br />

the processing and presentation of endogenous viral peptides on<br />

MHC class I molecules.<br />

Our findings may seem to contradict earlier reports indicating<br />

that HSV-1 inhibits macroautophagy 23 ; this inhibition is key to the<br />

neurovirulence of the virus 11 . However, a closer look at infected<br />

macrophages suggests that a form of autophagy distinct from<br />

macroautophagy is triggered. Macroautophagy can be induced<br />

in a variety of cells by mild heat shock or treatment with the<br />

mTOR inhibitor rapamycin 21,22 . In uninfected macrophages, these<br />

conditions induced throughout the cytoplasm the formation of<br />

autophagosomes with a double-membraned structure. In contrast,<br />

HSV-1-infected macrophages had two types of structures. In addition<br />

to the double-membraned structures, four-layered membrane<br />

structures that emerged from the nuclear envelope and accumulated<br />

in the cytoplasm at around 8 h after infection were present in<br />

most infected macrophages. We did not find such structures in<br />

uninfected cells treated with rapamycin, which suggested that<br />

they arose from a specific host response to HSV-1 infection distinct<br />

from macroautophagy. Although four-layered membrane structures<br />

have been reported before in the cytoplasm of HSV-1-infected<br />

mouse embryonic fibroblasts 25 , the autophagosomal nature of<br />

those structures was not documented. Here we have shown that<br />

these four-layered membrane structures were ‘decorated’ with LC3,<br />

a protein that is key to the formation of autophagosomes 28 ,and<br />

that these structures fused with lysosomes filled with bovine<br />

ARTICLES<br />

serum albumin–gold, thereby generating an environment suitable<br />

for the hydrolytic degradation of gB.<br />

We conclude that the ability of HSV-1 to inhibit macroautophagy<br />

early after infection, and thus to potentially limit the presentation of<br />

viral peptides, is counterbalanced by a host response involving the<br />

induction of a previously unknown autophagy pathway at a later time<br />

after infection. Our conclusion is supported by the results obtained<br />

with the D34.5 mutant virus. Macrophages infected with this mutant<br />

were able to trigger a strong immune response, like macrophages<br />

infected with wild-type viruses. Although the apparent magnitude of<br />

activation of CD8 + T cells induced by wild-type and D34.5 viruses was<br />

similar, the quantity of gB and viral proteins expressed in D34.5infected<br />

cells was much lower than that in macrophages infected<br />

with wild-type HSV-1. Macrophages infected with either D34.5 or<br />

wild-type HSV-1 engaged strong vacuolar responses. The distinct<br />

localization of LC3 to autophagosomes in the cytoplasm in macrophages<br />

infected with D34.5 virus, in contrast to its localization<br />

to the nuclear membrane in macrophages infected with wildtype<br />

virus, indicates that both types of autophagic responses can<br />

participate in the processing of viral proteins for presentation on<br />

MHC class I molecules.<br />

It has been reported that IFN-g stimulates autophagy and the<br />

clearance of mycobacteria in macrophages 3 . In our studies, IFN-g had<br />

no substantial effect on the induction of autophagy, a discrepancy that<br />

might be explained by the cell type and/or concentration of the<br />

cytokine used for stimulation 32 . Nevertheless, IFN-g-treated macrophages<br />

were much more efficient at stimulating CD8 + T cells after<br />

HSV-1 infection. As cotreatment with bafilomycin or 3-methyladenine<br />

had no effect on this IFN-g-induced stimulatory capacity, we conclude<br />

that vacuolar processing and autophagy were not essential to the<br />

response induced by IFN-g. The stimulatory effect of IFN-g on<br />

antigen presentation is well established. This cytokine upregulates<br />

assembly of the immunoproteasome 33 and stimulates expression of<br />

TAP1, the transporter associated with antigen presentation 34 , two<br />

key components of the classical MHC class I presentation pathway.<br />

Therefore, it was not unexpected to find strong inhibition of the<br />

stimulation of CD8 + T cells when we treated IFN-g-stimulated<br />

macrophages with MG-132 and BFA. In contrast, the IL-1b-induced<br />

improvement in the ability of infected macrophages to stimulate<br />

CD8 + T cells was inhibited considerably by 3-methyladenine<br />

and bafilomycin, which confirmed the contribution of autophagy to<br />

this process.<br />

The similarity in the results obtained with IL-1b and mild heatshock<br />

treatment suggests that IL-1b induces a cellular response similar<br />

to the one that occurs during fever-like conditions. Indeed, IL-1b is<br />

well known for its ability to induce fever 35 . Tumor necrosis factor, a<br />

second pyrogenic cytokine, also stimulated autophagy and the vacuolar<br />

processing of gB in our system (data not shown). These results<br />

suggest that the stress induced during fever conditions or after<br />

stimulation with pyrogenic cytokines triggers defense mechanisms,<br />

promoting more-efficient processing of viral antigens in vacuolar<br />

organelles. Notably, cytomegalovirus, a member of the herpesviridae<br />

family, can block signaling by IL-1b and tumor necrosis factor<br />

during the early phase of infection 36 , a process that might protect<br />

the virus by inhibiting the induction of antigen processing through a<br />

vacuolar response.<br />

Our results showing that lytic organelles associated with the<br />

processing of antigens for presentation on MHC class II molecules 12<br />

participated in the presentation of endogenous viral peptides on MHC<br />

class I molecules emphasize the dynamic cooperation between the<br />

‘classical’ and ‘vacuolar’ pathways of antigen presentation. This close<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 485


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

interaction is further emphasized by results showing that even in<br />

conditions in which vacuolar processing contributes to most of the<br />

viral antigen processing, such as after IL-1b or heat-shock stimulation,<br />

MG-132 and BFA still had a strong inhibitory effect. These results<br />

support a model in which the vacuolar processing of viral proteins in<br />

autophagosomes is followed by processing by the proteasome and<br />

peptide loading onto MHC class I molecules in the endoplasmic<br />

reticulum. The molecular mechanisms that enable the transfer of viral<br />

peptides from autophagosomes to proteasomes in the cytoplasm are<br />

unknown. A similar transport step has been shown to take place on<br />

phagosomes 37 . Although the molecular ‘machines’ that enable these<br />

translocation events have not been identified, it has been proposed<br />

that endoplasmic reticulum translocons such as Sec61 and/or derlin-1<br />

could be involved 16,17,38 . Although the nature of the endomembranes<br />

involved in the formation of autophagosomes during macroautophagy<br />

is still a matter of debate, our results have shown that the nuclear<br />

envelope, made of endoplasmic reticulum, is the membrane used to<br />

form the gB-enriched autophagosomes with four-layered membranes<br />

in HSV-1-infected macrophages. The endoplasmic reticulum has been<br />

shown to participate in a form of autophagy in yeasts referred to as<br />

‘ER-phagy’ 14 . Whether this process is homologous with the nuclear<br />

envelope–derived autophagic process documented here remains to be<br />

investigated. Published studies have shown that viruses take advantage<br />

of macroautophagy to find refuge in organelles, where they freely<br />

assemble 39 . Our results indicate that autophagy can also benefit the<br />

host by providing an additional pathway for the degradation of<br />

endogenous viral proteins for antigen presentation.<br />

METHODS<br />

Cells, viruses, antibodies and reagents. The BMA3.1A7 macrophage cell<br />

line was derived from C57BL/6 mice as described 37 and was cultured in<br />

complete DMEM (10% (vol/vol) FCS, penicillin (100 units/ml) and streptomycin<br />

(100 mg/ml)). The mouse macrophage cell line J774 was from American<br />

Type Culture Collection. The b-galactosidase–inducible HSV gB–specific CD8 +<br />

T cell hybridoma HSV-2.3.2E2 (provided by W. Heath, University of<br />

Melbourne) was maintained in RPMI-1640 medium supplemented with 5%<br />

(vol/vol) FCS, glutamine (2 mM), penicillin (100 units/ml), streptomycin<br />

(100 mg/ml), the aminoglycoside G418 (0.5 mg/ml) and hygromycin B<br />

(100 mg/ml). The HSV-1 K26-GFP mutant (strain KOS), carrying a GFPtagged<br />

capsid protein VP26, was provided by P. Desai 40 . The ICP34.5-null virus<br />

D34.5 was constructed by a strategy similar to that used to make the null<br />

mutant 17termA 41 . A 3,333–base pair DpnII fragment containing the gene<br />

encoding ICP34.5 with a nonsense mutation inserted at the sequence encoding<br />

the amino acid at position 30 was transfected into Vero cells together with<br />

infectious DNA from HSV-1 strain 17 (ref. 25). Individual plaques resulting<br />

from this transfection were screened by PCR amplification, followed by<br />

screening by SpeI digestion. After three rounds of plaque purification, viral<br />

stock was generated from which infectious DNA was prepared and ‘marker<br />

rescue’ was done. The ICP34.5-null D34.5 virus was characterized by immunoblot<br />

analysis for ICP34.5 and phosphorylated eIF2a as described 25 . Primary<br />

antibodies used were as follows: rabbit polyclonal antibody to LC3a (anti-LC3a;<br />

AP1801a; Abgent), rabbit polyclonal anti-LC3b (AP1802a; Abgent), rabbit<br />

polyclonal antibody to cleaved LC3b (AP1806a; Abgent), mouse monoclonal<br />

anti-gB (M612449; Fitzgerald), rat anti-LAMP-1 (1D4B; Developmental Studies<br />

Hybridoma Bank), rabbit polyclonal anti-Atg5 (NB-110-53818; Novus Biologicals),<br />

mouse monoclonal anti-tubulin (B-5-1-2; Sigma) and rabbit polyclonal<br />

anti-HSV (RB-1425-A; Neomarkers). The secondary antibodies Alexa Fluor<br />

568–conjugated goat anti-rabbit, Alexa Fluor 633–conjugated goat anti-mouse<br />

and Alexa Fluor 488–conjugated goat anti-mouse were from Invitrogen.<br />

Brefeldin A, bafilomycin A, MG-132, 3-methyladenine and rapamicyn were<br />

from Sigma.<br />

Heat shock, cytokine treatment and infection. For heat-shock treatment,<br />

macrophages (1 10 5 cells per well in 24-well plates) were incubated for 12 h<br />

at 39 1C, followed by a recovery period of 2 h at 37 1C before infection. The<br />

cytokines IL-1b (5 ng/ml; R&D Systems) and IFN-g (200 U/ml; PBL) were<br />

added 18–24 h before infection and were kept in the medium during viral<br />

infection. Macrophages were infected by incubation for 30 min with virus at a<br />

multiplicity of infection of 10. Cells were then washed and were incubated in<br />

fresh medium for a total of 8 h unless indicated otherwise. Drugs were added to<br />

the medium from 2 h after infection until the end of infection at the following<br />

concentrations: brefeldin A, 5 mg/ml; bafilomycin A, 0.5 mM; MG-132, 5 mM;<br />

3-methyladenine, 10 mM; and rapamicyn, 10 mg/ml.<br />

CD8 + T cell hybridoma assay. Mock- or HSV-1-infected macrophages<br />

(2 105 ) were washed in Dulbecco’s PBS and were fixed for 10 min at<br />

23 1C with 1% (wt/vol) paraformadehyde, followed by three washes in<br />

complete DMEM. Antigen-presenting cells were then cultured for 12 h at<br />

37 1C together with 4 105 HSV-2.3.2E2 cells (the b-galactosidase–inducible,<br />

gB-specific CD8 + T cell hybridoma) for analysis of the activation of T cells.<br />

Cells were then washed in Dulbecco’s PBS and lysed (0.125 M Tris base, 0.01 M<br />

cyclohexane diaminotetraacetic acid, 50% (vol/vol) glycerol, 0.025% (vol/vol)<br />

Triton X-100 and 0.003 M dithiothreitol, pH 7.8). A b-galactosidase substrate<br />

buffer (0.001 M MgSO4 7H2O, 0.01 M KCl, 0.39 M NaH2PO4 H2O, 0.6 M<br />

Na2HPO4 7H2O, 100 mM 2-mercaptoethanol and 0.15 mM chlorophenol<br />

red b-D-galactopyranoside, pH 7.8) was added for 2–4 h at 37 1C. Cleavage<br />

of the chromogenic substrate chlorophenol red-b-D-galactopyranoside was<br />

quantified in a spectrophotometer as absorbance at 595 nm.<br />

Immunufluorescence and dansylcadaverin labeling. For immunofluorescence<br />

analysis, control, treated and/or infected macrophages were fixed and made<br />

permeable with a Cytofix/Cytoperm kit according to the manufacturer’s<br />

recommendations (BD Biosciences). Cells were then incubated for 60 min<br />

at 25 1C with anti-LC3, anti-gB or anti-LAMP-1. For analysis of infection with<br />

HSV-1 K26-GFP, infected cells were visualized by detection of GFP fluorescence<br />

at 488 nm. Cells were analyzed with a confocal laser-scanning microscope<br />

(LSM 510Meta Axiovert; Carl Zeiss) or standard Axiophot fluorescent microscope<br />

(Zeiss) or by flow cytometry with a FACSCalibur (BD). For labeling<br />

with dansylcadaverin (Sigma), macrophages left untreated or exposed to IL-1b,<br />

IFN-g or mild heat shock were infected for 8 h and then stained for 15 min at<br />

37 1C with 50 mM dansylcadaverin. Cells were then washed in Dulbecco’s PBS<br />

and were lysed in 200 ml lysis buffer (described above). Total fluorescence was<br />

quantified with a SpectraMax Gemini electron microscopy spectrophotometer<br />

(excitation, 380 nm; emission, 525 nm).<br />

Electron microscopy. For morphological analysis, cells were fixed in 2.5%<br />

(vol/vol) glutaraldehyde and were embedded in Epon (Mecalab). Glucose-<br />

6-phosphatase was detected by electron microscopy cytochemistry as described<br />

42 . For immunocytochemistry after embedding, cells were fixed in 1%<br />

(vol/vol) glutaraldehyde and were embedded at –20 1C in Lowicryl (Canemco).<br />

Lowicryl ultrathin sections were incubated overnight with antibodies and were<br />

visualized by 60 min of incubation with protein A–gold complex (10 nm).<br />

Rabbit anti-LC3 and mouse anti-gB were used at dilution of 1:10.<br />

Analysis with siRNA. Control siRNA (nontargeting; siCONTROL) and siRNA<br />

specific for mouse Atg5 (L-064838-00-0005; ON-TARGETplus SMARTpool)<br />

were from Dharmacon. Cells were transfected with 100 nM siRNA using the<br />

DermaFECT 4 siRNA transfection reagent according to the manufacturer’s<br />

recommendations (Dharmacon). After 24 h, transfection medium was replaced<br />

by complete medium.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank J. Thibodeau and C. Perreault for critical reading of the manuscript;<br />

K. Rock (University of Massachusetts Medical School) for BMA cells;<br />

W. Heath (University of Melbourne) for the HSV-2.3.2E2 hybridoma; G. Arthur<br />

(University of Manitoba) for the wild-type and Atg5 –/– mouse embryonic<br />

fibroblasts produced by N. Mizushima (Medical and Dental University,<br />

Tokyo); P. Desai (Johns Hopkins University) for the HSV-1 K26-GFP<br />

mutant; and M. Bendayan for assistance with electron microscopy. Supported<br />

by the Canadian Institutes for Health Research (R.L. and M.D.), the Natural<br />

Science and Engineering Research Council of Canada (L.E.), Fonds de la<br />

486 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Recherche en Santé du Québec (L.E.), the US National Institutes of Health<br />

(EY09083) and Research to Prevent Blindness (D.L.)<br />

AUTHOR CONTRIBUTIONS<br />

L.E. planned and did most of the experiments and actively participated in<br />

writing the manuscript; M.C. did the experiments with mouse embryonic<br />

fibroblasts; J.D. maintained viral stocks; C.R. did the technical work for Epon<br />

electron microscopy; A.L. provided technical assistance for immunoblot analysis<br />

and immunofluorescence; D.G. did the immunogold labeling and morphological<br />

quantification; D.A. and D.L. produced the ICP34.5 HSV-1 mutant; C.N.<br />

participated in the planning and development of the antigen-presentation assay<br />

and provided help in writing the manuscript; R.L. provided HSV-1 virus stocks<br />

and expertise with the infection system and helped write the manuscript; and<br />

M.D. planned and directed the work and wrote the manuscript.<br />

COMPETING INTERESTS STATEMENT<br />

The authors declare competing financial interests: details accompany the full-text<br />

HTML version of the paper at http://www.nature.com/natureimmunology/.<br />

Published online at http://www.nature.com/naturegenetics/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Bevan, M.J. Cross-priming for a secondary cytotoxic response to minor H antigens with<br />

H-2 congenic cells which do not cross-react in the cytotoxic assay. J. Exp. Med. 143,<br />

1283–1288 (1976).<br />

2. Deretic, V. Autophagy in innate and adaptive immunity. Trends Immunol. 26, 523–528<br />

(2005).<br />

3. Levine, B. & Deretic, V. Unveiling the roles of autophagy in innate and adaptive<br />

immunity. Nat. Rev. Immunol. 7, 767–777 (2007).<br />

4. Schmid, D. & Munz, C. Innate and adaptive immunity through autophagy. Immunity 27,<br />

11–21 (2007).<br />

5. Andrade, R.M., Wessendarp, M., Gubbels, M.J., Striepen, B. & Subauste, C.S. CD40<br />

induces macrophage anti-Toxoplasma gondii activity by triggering autophagy-dependent<br />

fusion of pathogen-containing vacuoles and lysosomes. J. Clin. Invest. 116,<br />

2366–2377 (2006).<br />

6. Gutierrez, M.G. et al. Autophagy is a defense mechanism inhibiting BCG and Mycobacterium<br />

tuberculosis survival in infected macrophages. Cell 119, 753–766 (2004).<br />

7. Ling, Y.M. et al. Vacuolar and plasma membrane stripping and autophagic elimination<br />

of Toxoplasma gondii in primed effector macrophages. J. Exp. Med. 203, 2063–2071<br />

(2006).<br />

8. Nakagawa, I. et al. Autophagy defends cells against invading group A Streptococcus.<br />

Science 306, 1037–1040 (2004).<br />

9. Ogawa, M. et al. Escape of intracellular Shigella from autophagy. Science 307,<br />

727–731 (2005).<br />

10. Liu, Y. et al. Autophagy regulates programmed cell death during the plant innate<br />

immune response. Cell 121, 567–577 (2005).<br />

11. Orvedahl, A. et al. HSV-1 ICP34.5 confers neurovirulence by targeting the Beclin 1<br />

autophagy protein. Cell Host Microbe 1, 23–35 (2007).<br />

12. Paludan, C. et al. Endogenous MHC class II processing of a viral nuclear antigen after<br />

autophagy. Science 307, 593–596 (2005).<br />

13. Dengjel, J. et al. Autophagy promotes MHC class II presentation of peptides from<br />

intracellular source proteins. Proc. Natl. Acad. Sci. USA 102, 7922–7927<br />

(2005).<br />

14. Bernales, S., Schuck, S. & Walter, P. ER-phagy: selective autophagy of the endoplasmic<br />

reticulum. Autophagy 3, 285–287 (2007).<br />

15. Ackerman, A.L., Kyritsis, C., Tampe, R. & Cresswell, P. Early phagosomes in dendritic<br />

cells form a cellular compartment sufficient for cross presentation of exogenous<br />

antigens. Proc. Natl. Acad. Sci. USA 100, 12889–12894 (2003).<br />

16. Houde, M. et al. Phagosomes are competent organelles for antigen cross-presentation.<br />

<strong>Nature</strong> 425, 402–406 (2003).<br />

17. Guermonprez, P. et al. ER-phagosome fusion defines an MHC class I cross-presentation<br />

compartment in dendritic cells. <strong>Nature</strong> 425, 397–402 (2003).<br />

ARTICLES<br />

18. Mueller, S.N. et al. The early expression of glycoprotein B from herpes simplex virus<br />

canbedetectedbyantigen-specificCD8 + Tcells.J. Virol. 77, 2445–2451 (2003).<br />

19. Mueller, S.N., Jones, C.M., Smith, C.M., Heath, W.R. & Carbone, F.R. Rapid cytotoxic<br />

T lymphocyte activation occurs in the draining lymph nodes after cutaneous herpes<br />

simplex virus infection as a result of early antigen presentation and not the presence of<br />

virus. J. Exp. Med. 195, 651–656 (2002).<br />

20. Kuma, A. et al. The role of autophagy during the early neonatal starvation period.<br />

<strong>Nature</strong> 432, 1032–1036 (2004).<br />

21. Noda, T. & Ohsumi, Y. Tor, a phosphatidylinositol kinase homologue, controls autophagy<br />

in yeast. J. Biol. Chem. 273, 3963–3966 (1998).<br />

22. Komata, T. et al. Mild heat shock induces autophagic growth arrest, but not apoptosis in<br />

U251-MG and U87-MG human malignant glioma cells. J. Neurooncol. 68, 101–111<br />

(2004).<br />

23. Talloczy, Z. et al. Regulation of starvation- and virus-induced autophagy by the eIF2a<br />

kinase signaling pathway. Proc. Natl. Acad. Sci. USA 99, 190–195 (2002).<br />

24. He, B., Gross, M. & Roizman, B. The g134.5 protein of herpes simplex virus 1<br />

complexes with protein phosphatase 1a to dephosphorylate the alpha subunit of the<br />

eukaryotic translation initiation factor 2 and preclude the shutoff of protein synthesis<br />

by double-stranded RNA-activated protein kinase. Proc. Natl. Acad. Sci. USA 94,<br />

843–848 (1997).<br />

25. Alexander, D.E., Ward, S.L., Mizushima, N., Levine, B. & Leib, D.A. Analysis of the role<br />

of autophagy in replication of herpes simplex virus in cell culture. J. Virol. 81,<br />

12128–12134 (2007).<br />

26. Talloczy, Z., Virgin, H.W. IV. & Levine, B. PKR-dependent autophagic degradation of<br />

herpes simplex virus type 1. Autophagy 2, 24–29 (2006).<br />

27. Levine, B. Eating oneself and uninvited guests: autophagy-related pathways in cellular<br />

defense. Cell 120, 159–162 (2005).<br />

28. Kabeya, Y. et al. LC3, a mammalian homologue of yeast Apg8p, is localized in<br />

autophagosome membranes after processing. EMBO J. 19, 5720–5728 (2000).<br />

29. Heldwein, E.E. et al. Crystal structure of glycoprotein B from herpes simplex virus 1.<br />

Science 313, 217–220 (2006).<br />

30. Stannard, L.M., Himmelhoch, S. & Wynchank, S. Intra-nuclear localization of two<br />

envelope proteins, gB and gD, of herpes simplex virus. Arch. Virol. 141, 505–524<br />

(1996).<br />

31. Farnsworth, A. et al. Herpes simplex virus glycoproteins gB and gH function in fusion<br />

between the virion envelope and the outer nuclear membrane. Proc. Natl. Acad. Sci.<br />

USA 104, 10187–10192 (2007).<br />

32. Khalkhali-Ellis, Z. et al. IFN-g regulation of vacuolar pH, cathepsin D processing<br />

and autophagy in mammary epithelial cells. J. Cell. Biochem. 105, 208–218<br />

(2008).<br />

33. Griffin, T.A. et al. Immunoproteasome assembly: cooperative incorporation of interferon<br />

g (IFN-g)-inducible subunits. J. Exp. Med. 187, 97–104(1998).<br />

34. Epperson, D.E. et al. Cytokines increase transporter in antigen processing-1 expression<br />

more rapidly than HLA class I expression in endothelial cells. J. Immunol. 149,<br />

3297–3301 (1992).<br />

35. Conti, B., Tabarean, I., Andrei, C. & Bartfai, T. Cytokines and fever. Front. Biosci. 9,<br />

1433–1449 (2004).<br />

36. Jarvis, M.A. et al. Human cytomegalovirus attenuates interleukin-1b and tumor<br />

necrosis factor a proinflammatory signaling by inhibition of NF-kB activation.<br />

J. Virol. 80, 5588–5598 (2006).<br />

37. Kovacsovics-Bankowski, M. & Rock, K.L. A phagosome-to-cytosol pathway for exogenous<br />

antigens presented on MHC class I molecules. Science 267, 243–246<br />

(1995).<br />

38. Ackerman, A.L., Giodini, A. & Cresswell, P. A role for the endoplasmic reticulum protein<br />

retrotranslocation machinery during crosspresentation by dendritic cells. Immunity 25,<br />

607–617 (2006).<br />

39. Jackson, W.T. et al. Subversion of cellular autophagosomal machinery by RNA viruses.<br />

PLoS Biol. 3, e156 (2005).<br />

40. Desai, P. & Person, S. Incorporation of the green fluorescent protein into the herpes<br />

simplex virus type 1 capsid. J. Virol. 72, 7563–7568 (1998).<br />

41. Bolovan, C.A., Sawtell, N.M. & Thompson, R.L. ICP34.5 mutants of herpes simplex<br />

virus type 1 strain 17syn+ are attenuated for neurovirulence in mice and for replication<br />

in confluent primary mouse embryo cell cultures. J. Virol. 68, 48–55 (1994).<br />

42. Griffiths, G., Quinn, P. & Warren, G. Dissection of the Golgi complex. I. Monensin<br />

inhibits the transport of viral membrane proteins from medial to trans Golgi cisternae in<br />

baby hamster kidney cells infected with Semliki Forest virus. J. Cell Biol. 96, 835–850<br />

(1983).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 487


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Cross-presentation of viral and self antigens by<br />

skin-derived CD103 + dendritic cells<br />

Sammy Bedoui 1–3 , Paul G Whitney 1,3 , Jason Waithman 1–3 , Liv Eidsmo 1 , Linda Wakim 1 , Irina Caminschi 2 ,<br />

Rhys S Allan 1 , Magdalena Wojtasiak 1 , Ken Shortman 2 , Francis R Carbone 1 , Andrew G Brooks 1 &<br />

William R Heath 1,2<br />

Skin-derived dendritic cells (DCs) include Langerhans cells, classical dermal DCs and a langerin-positive CD103 + dermal subset.<br />

We examined their involvement in the presentation of skin-associated viral and self antigens. Only the CD103 + subset efficiently<br />

presented antigens of herpes simplex virus type 1 to naive CD8 + T cells, although all subsets presented these antigens to CD4 +<br />

T cells. This showed that CD103 + DCs were the migratory subset most efficient at processing viral antigens into the major<br />

histocompatibility complex class I pathway, potentially through cross-presentation. This was supported by data showing only<br />

CD103 + DCs efficiently cross-presented skin-derived self antigens. This indicates CD103 + DCs are the main migratory subtype<br />

able to cross-present viral and self antigens, which identifies another level of specialization for skin DCs.<br />

Dendritic cells (DCs) serve an essential antigen-presenting function in<br />

the initiation of T cell responses 1,2 . DCs can be categorized into many<br />

independent subsets that seem to represent end-stage populations 3–5 .<br />

One main subclassification can be made of plasmacytoid and conventional<br />

DCs. In the spleen, both plasmacytoid DCs and at least three<br />

subsets of conventional DCs can be found in the steady state, the latter<br />

being loosely categorized as CD8a + DCs and CD8a – DCs, and the<br />

CD8a – DCs being further classified into CD4 + and CD4 – populations.<br />

These DCs enter the spleen from the blood either as mature plasmacytoid<br />

DCs or as precursors of conventional DCs. The lymph nodes<br />

also contain these blood-derived DC subsets but in addition have a<br />

cohort of DCs that have migrated from peripheral tissues 6 . Depending<br />

on the site of drainage for a particular lymph node, it will contain a<br />

variety of migratory DCs. For skin-draining lymph nodes, the migratory<br />

population consists of both epidermis-derived Langerhans cells<br />

and dermis-derived DCs. This represents an abbreviated list of DC<br />

subsets, with many intricacies associated with particular tissues, but it<br />

can be further extended to include monocyte-derived DCs generated<br />

in inflammatory conditions such as infection 4,7 .<br />

Although a variety of DC subsets have been identified, understanding<br />

of the specific functions of individual groups, particularly in<br />

the skin, where at least three different subsets reside, is limited. Clearly,<br />

migratory DCs are critical for trafficking antigen from peripheral sites<br />

to the lymph nodes 8,9 , and plasmacytoid DCs are a chief source of<br />

interferon-a during viral infection 10,11 . One property that varies<br />

extensively among subsets is the ability to introduce exogenous<br />

antigens into the major histocompatibility complex (MHC) class I<br />

Received 10 December 2008; accepted 9 March 2009; published online 6 April 2009; doi:10.1038/ni.1724<br />

pathway, a process referred to as ‘cross-presentation’ 12–14 . In particular,<br />

CD8a + DCs are the dominant splenic population responsible for<br />

cross-presentation of cell-associated antigens 12,14 or those antigens<br />

targeted by monoclonal antibodies 13 . This same subset has been<br />

strongly suggested to be involved in the presentation of viral antigens<br />

for many different infections 15–19 , with some evidence that this relates<br />

to their ability to cross-present 20 .<br />

The importance of CD8a + DCs in viral immunity has been<br />

emphasized by studies examining the initiation of CD8 + T cell<br />

immunity to infection of the skin with herpes simplex virus type 1<br />

(HSV-1) 9,16 . These studies have shown that CD8a + DCs are the sole<br />

DC subset responsible for antigen presentation in the draining lymph<br />

nodes shortly after infection, which challenges the long-held paradigm<br />

that Langerhans cells control immunity to epidermal antigens. The<br />

limited involvement of Langerhans cells has been emphasized by<br />

examination of cytotoxic T lymphocyte (CTL) immunity in bone<br />

marrow chimeras in which radioresistant Langerhans cells are the only<br />

DCs expressing the correct MHC-restriction elements 16 .Inthiscase,<br />

CTL immunity is abrogated, which further challenges the importance<br />

of Langerhans cells in the initiation of skin immunity. This view has<br />

been reinforced for skin infection with other viruses, such as vaccinia<br />

18,21 and influenza 18 , although there is some evidence for the<br />

involvement of skin-derived DCs in the generation of CTL immunity<br />

to lentivirus infection 21 , the only noncytopathic virus tested.<br />

Several subsequent studies have provided some evidence that<br />

Langerhans cells contribute to immunity induced by contact<br />

sensitization 22 and to self tolerance 23 , although even these conclusions<br />

1 The Department of Microbiology and <strong>Immunology</strong>, The University of Melbourne, Parkville, Victoria, Australia. 2 The Walter and Eliza Hall Institute of Medical Research,<br />

Parkville, Victoria, Australia. 3 These authors contributed equally to this work. Correspondence should be addressed to W.R.H. (wrheath@unimelb.edu.au), F.R.C.<br />

(fcarbone@unimelb.edu.au) or A.G.B. (agbrooks@unimelb.edu.au).<br />

488 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

b<br />

c<br />

1° site, day 2<br />

HSV-1 DAPI<br />

2° site, day 2<br />

are not universal 24–26 . Some studies have used transgenic expression of<br />

fluorescent proteins and/or the diphtheria toxin receptor under<br />

control of the promoter of the gene encoding langerin for the<br />

identification or depletion of Langerhans cells 22,24–26 . However, this<br />

approach has led to the identification of a second skin-associated<br />

population of langerin-positive cells that, unlike Langerhans cells,<br />

are radiosensitive 26–28 and, in contrast to other skin-derived DC<br />

populations, express the integrin CD103 (ref. 26). The identification<br />

of this previously unknown population of langerin-positive cells<br />

suggests a need to reevaluate their involvement in immunity to<br />

HSV-1 (ref. 27).<br />

Here we examine the contribution of this migratory CD103 + DC<br />

population to the presentation of epidermis-expressed self antigen or<br />

to the presentation of HSV-1 antigen during infection of the skin. We<br />

provide evidence that despite the presence of three distinct DC<br />

subtypes in the skin, CD103 + DCs dominated the MHC class I–<br />

restricted cross-presentation of both pathogen and self antigens. This<br />

provides new insight into the specialization of DC subtypes in the skin<br />

and suggests that the CD103 + DC subset, present in various tissues,<br />

may be broadly responsible for generating CTL immunity and<br />

tolerance to tissue-associated pathogens and antigens, respectively.<br />

RESULTS<br />

Two phases of acute HSV-1 infection<br />

Acute cutaneous infection with HSV-1 can be achieved by scarification<br />

of mouse flank skin 29,30 . Infection initially results in viral replication<br />

in keratinocytes surrounding the site of scarification (Fig. 1a), before<br />

the virus enters neuronal cells that innervate this area. HSV-1 then<br />

travels by retrograde axonal transport to the cell body, where it<br />

replicates extensively, spreading throughout several local dorsal root<br />

ganglia. By days 3–4 of infection, virions begin to travel by anterograde<br />

transport along the many axons derived from the infected<br />

ganglia (Fig. 1b), reaching the epidermis, where they infect the entire<br />

dermatome innervated by the ganglia. In this site, HSV-1 again<br />

replicates extensively in epidermal cells (Fig. 1c), causing a zosteriform<br />

band beginning at day 6 after infection. Of critical relevance here,<br />

HSV-1 has two phases of acute viral replication in the skin: a primary<br />

infection that is limited to the site of scarification, and a secondary<br />

growth phase involving the entire innervated dermatome. The dose of<br />

virus used to infect mice had little influence on viral titer at either the<br />

ARTICLES<br />

Figure 1 Progress of infection after inoculation of HSV-1 on scarified flank<br />

skin. (a) Confocal microscopy (left) and direct photography (right; box<br />

outlines primary (11) site) of HSV-1 in flank skin of C57BL/6 mice on day 2<br />

after infection with HSV-1 by flank scarification. Left: sections 8 mm in<br />

thickness acquired with a 10 objective; red, viral proteins; blue, nuclei<br />

(stained with the DNA-intercalating dye DAPI). Scale bar, 100 mm.<br />

(b) Confocal microscopy (z-stack images) of nerves expressing viral proteins<br />

on day 4 of infection as described in a, assessing five to seven sections<br />

1 mm in thickness acquired with a 63 oil objective and stacked according<br />

to maximum intensity. Scale bar, 20 mm. (c) Confocal microscopy of flank<br />

skin on day 5 of infection as described in a, assessing sections 8 mm in<br />

thickness acquired with a 10 objective (left), or photography of flank skin<br />

on day 7 of infection as described in a (right; box outlines secondary (21)<br />

site). Scale bar, 100 mm. For these studies, uninfected skin was examined<br />

visually and control serum was used to confirm the validity of positive<br />

staining. No viral staining was detected in healthy skin. Areas on either<br />

side of the primary infection site in a represent uninfected regions that<br />

act as an internal negative control. Data are representative of at least<br />

three experiments.<br />

primary site (day 2) or the secondary site (day 5), but did affect the<br />

chance of being infected (Supplementary Fig. 1 online).<br />

T cell activation occurs in phases<br />

After infection of the flank with HSV-1, antigen presentation occurs<br />

rapidly in the draining brachial lymph nodes, with strong presentation<br />

evident at 48 h (ref. 16). Because viral recrudescence causes a second<br />

phase of infection of the entire skin dermatome beginning about 72 h<br />

after infection 30 , we investigated whether this led to a second wave of<br />

antigen presentation. To monitor antigen presentation after flank<br />

infection, we tracked the proliferation of CD8 + gBT-I T cells specific<br />

for an H-2K b -restricted epitope of glycoprotein B (gB). We infected<br />

C57BL/6 mice with HSV-1 by scarification of the upper flank and<br />

then, on day 2 or day 5 after infection, intravenously injected the mice<br />

with gBT-I cells labeled with the cytosolic dye CFSE to monitor<br />

proliferation 42 h later. We chose the 42-hour time frame to allow<br />

proliferation without sufficient time for T cells to exit the lymph node<br />

in which they responded. In this way, any proliferation detected would<br />

mirror antigen presentation in that lymph node. With this approach,<br />

T cell proliferation was evident in the brachial lymph nodes but not in<br />

the closely associated axillary lymph nodes on day 2 (Fig. 2a). On day<br />

5, however, viral recrudescence led to antigen presentation in both the<br />

brachial and axillary lymph nodes, as measured by proliferation of<br />

gBT-I cells in each site. We found no such proliferation in the<br />

mesenteric lymph nodes (Supplementary Fig. 2 online), which<br />

indicated that the response was not systemic. As the delayed presentation<br />

in the axillary lymph nodes coincided with a greater antigen load<br />

associated with viral recrudescence, antigen-bearing DCs may have<br />

drained from the secondary site directly to the axillary lymph nodes<br />

or, alternatively, may have passed through the brachial lymph node,<br />

overflowing into the axillary lymph node. To distinguish between<br />

those possibilities, we altered the initial site of scarification to the<br />

lower flank. In this case, the pattern of antigen presentation was<br />

reversed: T cell proliferation occurred initially in the axillary but not<br />

the brachial lymph nodes on day 2 and occurred in both lymph nodes<br />

on day 5 (Fig. 2b). These findings indicate that the first phase of<br />

infection initiated antigen presentation in a single lymph node,<br />

whereas the second phase contributed a later wave of presentation<br />

involving at least two lymph nodes.<br />

We confirmed the view that the second phase of T cell activation<br />

required viral recrudescence by using a thymidine kinase–deficient<br />

form of HSV-1 (HSV-TK – ), which does not replicate in the dorsal root<br />

ganglia or form secondary lesions. We found that this thymidine<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 489


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

Ax LN<br />

Br LN<br />

Day 2 Day 5<br />

6.5 ± 2.1 82.8 ± 4.8<br />

48.4 ± 11.0 75.1 ± 5.7<br />

CFSE<br />

Ax LN<br />

Br LN<br />

kinase–deficient form of the virus had a primary growth phase in the<br />

skin similar to that of wild-type virus, with titers diverging only after<br />

the wild-type virus returned from the ganglia during the zosteriform<br />

phase of disease (Supplementary Fig. 3 online). Infection of the upper<br />

flank with HSV-TK – did not result in T cell activation in the axillary<br />

lymph nodes on day 2 or day 5, although activation in the brachial<br />

lymph nodes was evident on days 2 and 5 as a consequence of the<br />

primary infection (Fig. 2c).<br />

Presentation by migratory DCs after viral recrudescence<br />

There are several subsets of DCs; studies have indicated involvement of<br />

a lymph node–resident population of CD8a + DCs in the initiation of<br />

T cell activation after HSV-1 infection 9,16 . Those studies and our<br />

studies here used ex vivo isolation of DCs to show extensive presentation<br />

by CD8a + DCs on day 2 (refs. 9,16; Fig. 3a,b) with no evidence of<br />

presentation by any DC subset on day 4 (Fig. 3a,c). As in the earlier<br />

studies 9,16 , we prepared DCs by first depleting lymph node suspensions<br />

of other cell types and then staining for CD11c, CD205 and<br />

CD8 (Fig. 3). Then, we sorted DCs by flow cytometry on the basis<br />

of expression of CD11c and various combinations of CD8a and<br />

CD205 (Fig. 3a); Langerhans cells were CD205 hi CD8a – ,dermalDCs<br />

were CD205 int CD8a – , lymph node–resident CD8a + DCs were<br />

CD205 int CD8a + , and we called the remaining mixed population<br />

‘double-negative DCs’. These studies initially led us to assume that<br />

presentation had mostly ended by day 4 (Fig. 3c), although it is<br />

b<br />

Day 2 Day 5<br />

53.9 ± 8.9 77.9 ± 7.1<br />

3.5 ± 1.5 76.1 ± 7.7<br />

CFSE<br />

c<br />

Ax LN<br />

Br LN<br />

Day 2 Day 5<br />

6.4 ± 3.2 10.5 ± 5.0<br />

38.4 ± 8.7 57.4 ± 10.2<br />

Figure 2 In vivo antigen presentation to HSV-specific CD8 + T cells after cutaneous HSV-1 infection. (a,b) Proliferation in the axillary lymph nodes<br />

(Ax LN; top) and brachial lymph nodes (Br LN; bottom) at 42 h after transfer of 1 10 6 CFSE-labeled gBT-I cells intravenously into mice infected with<br />

HSV-1 on the upper flank (a) or lower flank (b), 2 d earlier (left) or 5 d earlier (right). Numbers in plots indicate percent proliferated cells (± s.e.m.).<br />

Data are from three independent experiments. (c) Proliferation analysis as described in a, but for mice infected with HSV-TK – . Data are from two<br />

independent experiments.<br />

CD205<br />

a<br />

LC<br />

CD8α +<br />

DC<br />

dDC<br />

DN DC<br />

CD8<br />

b<br />

Divided gBT-I/well (×10 3 )<br />

Day 2, br LN<br />

30<br />

20<br />

10<br />

c<br />

Divided gBT-I/well (×10 3 )<br />

Day 4, br LN<br />

30<br />

20<br />

10<br />

CFSE<br />

worth noting that studies using a T cell hybridoma that produces<br />

b-galactosidase after stimulation suggest that presentation may extend<br />

beyond this time 9 .<br />

Given the evidence presented above for viral recrudescence–<br />

dependent T cell activation on day 5 in the axillary lymph nodes<br />

(Fig. 2a), we questioned whether ex vivo presentation by DC subsets<br />

might be evident on day 5 in the axillary lymph nodes. This was of<br />

particular interest, as such secondary-site-dependent presentation<br />

would not require mechanical scarring of the skin but instead<br />

would be a consequence of the natural movement of virus from the<br />

nerves to the epidermis. We infected mice by scarification of the upper<br />

flank and then, 5 d later, collected axillary lymph nodes and assessed<br />

the isolated DC subsets for their ability to stimulate CD8 + gBT-I cells<br />

in vitro. In this case, extensive presentation was evident in the axillary<br />

lymph nodes on day 5 (Fig. 3d), but it was not apparent when we used<br />

nonrecrudescent HSV-TK – as the infectious agent (Supplementary<br />

Fig. 4 online). In contrast to the presentation on day 2 in the brachial<br />

lymph node, dermal DCs seemed to dominate presentation in the<br />

axillary lymph node at this later time point. Presentation by CD8a +<br />

DCs still occurred at an amount similar to that in the brachial lymph<br />

nodes on day 2, and there was also some stimulation by Langerhans<br />

cells, but it was difficult to exclude the possibility of contamination of<br />

this population by the highly stimulatory dermal DCs that overlapped<br />

in CD205 expression. Consistent with the proposal that a new wave of<br />

DCs entered the lymph nodes from the secondary site of infection,<br />

d<br />

Divided gBT-I/well (×10 3 )<br />

Day 5, ax LN<br />

120<br />

e<br />

Divided gBT-I/well (×10 3 )<br />

LC<br />

dDC<br />

Day 5, br LN<br />

30<br />

CD8α + DC<br />

DN DC<br />

0<br />

0 1.9 3.8 7.5 15 30 0 1.9 3.8 7.5 15 30 0 1.9 3.8 7.5 15 30 0 1.9 3.8 7.5 15 30<br />

DC/well (×10 3 0<br />

0<br />

0<br />

)<br />

DC/well (×10 3 ) DC/well (×10 3 ) DC/well (×10 3 )<br />

Figure 3 Migratory DCs are involved in antigen presentation to CD8 + T cells after secondary viral infection of the skin. (a) Purification of DC subsets from<br />

lymph nodes of infected mice: after DC enrichment, expression of CD205 and CD8a by CD11c + DCs was assessed, and Langerhans cells (LC; CD205 hi CD8a – ),<br />

dermal DCs (dDC; CD205 int CD8a – ), double-negative DCs (DN DC; CD205 – CD8a – )andCD8a + DCs (CD205 int CD8a + ) were isolated with a cell sorter after<br />

gating on CD11c + cells. (b–e) Proliferation of 5 10 4 CFSE-labeled gBT-I cells cultured for 60 h together with serial dilutions of DC subsets (identified<br />

as in a) isolated from brachial lymph nodes (b,c,e) or axillary lymph nodes (d) of mice infected 2 d earlier (b), 4 d earlier (c) or 5 d earlier (d,e). Day-4<br />

assays were done on the same day as day-2 assays. Data are pooled from three to four individual experiments (mean ± s.e.m.).<br />

490 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

90<br />

60<br />

30<br />

20<br />

10


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

CD11c<br />

CD205<br />

a<br />

CD8<br />

CD103<br />

dDC<br />

CD326<br />

LC<br />

CD8α +<br />

DC<br />

CD103 +<br />

DC<br />

Divided gBT-I/well (×10 3 )<br />

Divided gBT-I/well (×10 3 )<br />

b<br />

c<br />

25<br />

20<br />

15<br />

10<br />

0<br />

DC/well (×10 3 0 1.9 3.8 7.5 15<br />

)<br />

50<br />

40<br />

20<br />

LC<br />

dDC<br />

CD8α + CD103<br />

DC<br />

+ DC<br />

analysis of presentation in the brachial lymph nodes on day 5 showed<br />

presentation by dermal DCs (Fig. 3e) that was not evident on day 4<br />

(Fig. 3c). In this case, presentation by CD8a + DCs was minimal and<br />

presentation by dermal DCs was somewhat less than in the axillary<br />

lymph nodes, possibly because effector CTLs were already actively<br />

killing antigen-presenting cells in this site.<br />

CD103 + DCs present after viral recrudescence<br />

Several groups have reported a third population of skin-derived DCs<br />

that, like Langerhans cells, express langerin 26–28 but, unlike other skinderived<br />

DCs, express CD103 (ref. 26). These cells are restricted mainly<br />

to the dermis and have therefore been called ‘langerin-positive dermal<br />

DCs’, but for simplicity, we will call them ‘CD103 + DCs’ here. These<br />

DCs are resident in the skin and migrate to the lymph nodes after the<br />

skin is painted with tetramethylrhodamine isothiocyanate plus irritant<br />

26 . Both dermal DCs and Langerhans cells migrate after skin is<br />

painted with fluorescein isothiocyanate (FITC) 9 ; here we used painting<br />

with FITC to show that CD103 + DCs migrated in the context of<br />

irritant or infection with HSV-1 (Supplementary Fig. 5 online). These<br />

studies also confirmed the migratory nature of Langerhans cells and<br />

the classical CD103 – dermal DCs and showed a lack of migration by<br />

lymph node–resident CD8a + DCs.<br />

Given the strong presentation by dermal DCs on day 5 reported<br />

above (Fig. 3d), it became imperative to examine the antigenpresenting<br />

activity of contaminating CD103 + DCs in these groups.<br />

To achieve this, we included CD103 staining in our sorting protocol.<br />

Because of limitations in the number of groups that could be sorted,<br />

in subsequent studies we did not include the double-negative DCs<br />

identified above, which represented mainly lymph node–resident<br />

CD8a – DCs that did not seem to contribute to HSV-1-specific<br />

5<br />

30<br />

10<br />

0<br />

Day 2, br LN<br />

Day 5, ax LN<br />

DC/well (×10 3 0 1.9 3.8 7.5 15<br />

)<br />

Figure 5 Many DC subsets present MHC class II–restricted viral antigen to<br />

gDT-II CD4 + T cells. (a,b) Proliferation of 5 10 4 CFSE-labeled HSV-1specific<br />

CD4 + T cells (gDT-II) after 60 h of culture together with serial<br />

dilutions of DC subsets isolated (as described in Fig. 4a) frombrachial<br />

lymph nodes (a) or axillary lymph nodes (b) of mice infected 2 d earlier.<br />

(c,d) Proliferation analysis as described in a,b for DC subsets isolated<br />

from brachial lymph nodes (c) or axillary lymph nodes (d) of mice infected<br />

5 d earlier. Data are pooled from two to four individual experiments<br />

(mean ± s.e.m.).<br />

Figure 4 CD103 + DCs present HSV-1 antigens to CD8 + T cells after<br />

secondary viral infection of the skin. (a) Gating strategy for isolation of<br />

CD103 + DCs: after DC enrichment, CD8a + DCs were purified on the<br />

basis of expression of CD11c and CD8a (top; right gate); expression of<br />

CD205 and CD103 (middle) was assessed in CD11c + CD8a – cells (top; left<br />

gate) for the isolation of CD103 + DCs (middle; right gate); and CD103 –<br />

CD11c + CD8a – DCs (middle; left gate) were categorized as Langerhans cells<br />

(CD326 hi ) and dermal DCs (CD326 – ) on the basis of CD326 expression<br />

(bottom; sort gates outlined). Cells of DC subsets purified from lymph<br />

nodes of naive and infected mice are counted in Supplementary Figure 8.<br />

(b,c) Proliferation of 5 10 4 CFSE-labeled gBT-I cells cultured for 60 h<br />

together with serial dilutions of DC subsets (identified as in a) sorted from<br />

brachial lymph nodes 2 d after infection (b) or from axillary lymph nodes<br />

5 d after infection (c). Data are pooled from two to four individual<br />

experiments (mean ± s.e.m.).<br />

CD8 + T cell activation. In this sorting scheme (Fig. 4a), we isolated<br />

CD8a + DCs directly from the CD11c + cells, then isolated CD103 + cells<br />

and separated Langerhans cells and dermal DCs on the basis of CD205<br />

expression and differences in expression of CD326 (Ep-CAM;<br />

expressed by Langerhans cells). We isolated CD8a + DCs first, as<br />

they had low expression of CD103 (Supplementary Fig. 6 online)<br />

that would otherwise potentially result in contamination of the<br />

CD103 + population by this subset.<br />

After sorting the cells as described above, we examined these DC<br />

subsets for their presentation of antigen on day 2 in the brachial<br />

lymph nodes (Fig. 4b) and on day 5 in the axillary lymph nodes<br />

(Fig. 4c), which represent the primary and secondary sites, respectively.<br />

This showed that CD103 + DCs were the dominant migratory<br />

DCs that presented viral antigen on day 5 in the axillary lymph nodes,<br />

whereas little activity by any migratory DC was evident on day 2 in the<br />

brachial lymph nodes. In contrast, CD8a + DCs presented viral antigen<br />

during both phases. Some presentation by dermal DCs and, to a lesser<br />

extent, Langerhans cells was also present on day 5 in the axillary<br />

lymph nodes, but it is difficult to exclude the possibility that this was<br />

not due to contaminating CD103 + DCs, which showed detectable<br />

presentation even with as few as 467 cells (data not shown). These<br />

findings suggest that the CD103 + DCs were dominant among migratory<br />

DCs for the ability to cross-present.<br />

To formally address whether CD103 + DCs that had migrated from<br />

the skin presented viral antigen, we examined presentation on day 5 in<br />

the axillary lymph nodes after labeling the flank skin with FITC at 2 d<br />

Divided gDT-II/well (×10 3 )<br />

Divided gDT-II/well (×10 3 )<br />

a<br />

c<br />

Day 2, br LN<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 1.9 3.8 7.5 15<br />

Day 5, br LN<br />

40<br />

30<br />

20<br />

10<br />

DC/well (×10 3 )<br />

LC<br />

dDC<br />

CD8α + CD103<br />

DC<br />

+ DC<br />

0 1.9 3.8 7.5 15<br />

DC/well (×10 3 )<br />

0<br />

0 1.9 3.8 7.5 15 0 1.9 3.8 7.5 15<br />

DC/well (×10 3 ) DC/well (×10 3 0<br />

)<br />

Divided gDT-II/well (×10 3 )<br />

Divided gDT-II/well (×10 3 )<br />

b<br />

d<br />

25<br />

20<br />

15<br />

10<br />

5<br />

40<br />

30<br />

20<br />

10<br />

Day 2, ax LN<br />

Day 5, ax LN<br />

ARTICLES<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 491


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Figure 6 CD103 + DCs present<br />

skin-derived self antigen to<br />

CD8 + T cells. Proliferation of<br />

5 10 4 CFSE-labeled OVAspecific<br />

CD8 + T cells (OT-I)<br />

after 60 h of culture together<br />

with serial dilutions of DC<br />

subsets isolated (as described<br />

in Fig. 4a) from the skindraining<br />

lymph nodes of K5mOVA<br />

mice. Data are pooled<br />

from two individual experiments<br />

(mean ± s.e.m.).<br />

Divided OT-I/well (×10 3 )<br />

LC<br />

dDC<br />

CD8α + CD103<br />

DC<br />

+ DC<br />

after infection (Supplementary Fig. 7 online). In this case,<br />

FITC + CD103 + DCs, which had migrated from the skin after viral<br />

recrudescence, presented viral antigens to CD8 + T cells. Notably,<br />

FITC – CD103 + DCs also efficiently presented viral antigen. Presentation<br />

by this latter population most likely reflected the fact that only a<br />

small proportion of the CD103 + DCs are amenable to labeling<br />

with FITC. That idea was reinforced by the finding that although<br />

the number of CD103 + DCs in the axillary lymph nodes<br />

increased fourfold between day 2 and day 5 (Supplementary Fig. 8<br />

online), only about 6% of the cells were labeled with FITC (Supplementary<br />

Fig. 5e).<br />

Access to viral antigen is not limited to CD103 + DCs<br />

‘Preferential’ MHC class I–restricted presentation of viral antigen by<br />

CD103 + DCs could relate to access to viral antigens rather than a<br />

dominant ability to cross-present. To investigate that possibility, we<br />

explored whether other migratory DCs had access to viral antigen for<br />

MHC class II–restricted presentation. To achieve this, we generated a<br />

new line of T cell antigen receptor–transgenic mice that produce CD4 +<br />

T cells specific for glycoprotein D (gD) of HSV-1 (gDT-II mice;<br />

P.G.W., S.B., G. Davey, W.R.H., F.R.C., A.G.B., unpublished observations).<br />

We used CD4 + gDT-II T cells as responders to examine<br />

presentation in the brachial and axillary lymph nodes on days 2 and<br />

5(Fig. 5). This showed that all DCs could present antigens to virusspecific<br />

CD4 + T cells on day 5 in the axillary lymph nodes (Fig. 5d)<br />

and brachial lymph nodes (Fig. 5c) and all but the Langerhans cells<br />

presented on day 2 in the brachial lymph nodes (Fig. 5a). As expected,<br />

there was no presentation on day 2 in the axillary lymph nodes<br />

(Fig. 5b). These studies support the view that the dominant ability of<br />

CD103 + DCs to present MHC class I–restricted antigens on day 5 is<br />

related to their ability to process viral antigen into this pathway, rather<br />

than their ability to access the antigen. This was especially evident by<br />

comparison of MHC class I– and class II–restricted presentation on<br />

day 5 by dermal DCs and CD103 + DCs. Analysis of cytokines<br />

produced by gDT-II cells on day 5 in response to viral antigens<br />

presented by the various DC subsets showed a bias toward production<br />

of T helper type 1 cytokines (interferon-g, interleukin 2 and tumor<br />

necrosis factor) in amounts that mirrored proliferation but did not<br />

differ among DC subtypes (Supplementary Fig. 9 online).<br />

CD103 + DCs cross-present epidermal antigens<br />

It is difficult to distinguish between cross-presentation and direct<br />

infection of DCs during viral responses. As an alternative approach to<br />

examine the cross-presentation ability of skin-derived migratory DCs,<br />

we used the DC-sorting protocol described above for isolation of<br />

CD103 + DCs to examine the presentation of a membrane-associated<br />

form of ovalbumin (OVA) by DCs from mice expressing OVA in<br />

epidermal keratinocytes under control of the promoter of the gene<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

DC/well (×10 3 0 1.6 3.1 6.3 12.5 25<br />

)<br />

encoding keratin 5 (K5-mOVA mice). In accordance with our studies<br />

of HSV-1 infection, this showed that CD103 + DCs were the dominant<br />

DC population presenting MHC class I–restricted epidermal skin<br />

antigen in the steady state (Fig. 6). As this antigen was expressed by<br />

keratinocytes but not by DCs 23 , it was apparent that in this case,<br />

CD103 + DCs used cross-presentation.<br />

DISCUSSION<br />

This study has shown that langerin-positive CD103 + dermal DCs are<br />

the dominant migratory DCs of the skin that present MHC class I–<br />

restricted antigens. This dominance most likely relates to their ability<br />

to cross-present, at least for the types of antigens tested here. An<br />

alternative explanation is that presentation of viral antigens by<br />

CD103 + DCs occurs as a consequence of ‘preferential’ infection.<br />

Definitive resolution of this possibility is problematic by most<br />

approaches available, as it is almost impossible to distinguish between<br />

an infected cell and a cell that has captured infected cellular material.<br />

Although some doubt may exist over whether CD103 + DCs use crosspresentation<br />

for HSV-1 antigens, it is likely that this function was<br />

responsible for the presentation of OVA in K5-mOVA mice. The<br />

failure of classical dermal DCs to cross-present OVA in this case might<br />

be explained by limited access to antigen, as OVA is expressed by<br />

epidermal cells. This explanation cannot apply to Langerhans cells,<br />

however, as they reside directly in the region of OVA expression in<br />

K5-mOVA mice 31 . Unfortunately, we have not been able to detect<br />

presentation of OVA on MHC class II by any DC subset in K5-mOVA<br />

mice, which precludes a definitive statement about access by dermal<br />

DCs. However, the ability of all migratory DCs to present viral<br />

antigens on MHC class II during HSV infection of the skin indicates<br />

that all had access to viral antigens, yet only the CD103 + DCs<br />

presented these antigens efficiently on MHC class I.<br />

It is worth considering whether migratory DCs might access viral<br />

antigens in the draining lymph nodes rather than the skin. As mature<br />

DCs are very poor at capturing antigen for presentation 20 , and<br />

migratory DCs have a mature phenotype in the lymph nodes, it is<br />

most likely that viral antigens presented by migratory DCs are<br />

captured in the skin. More importantly, if CD103 + DCs captured<br />

viral antigen in the lymph nodes, then they might be expected to<br />

cross-present it efficiently on both days 2 and 5, as reported for the<br />

CD8a + DCs, yet cross-presentation by CD103 + DCs was absent on day<br />

2. Furthermore, the ability of CD103 + DCs to cross-present skinderived<br />

OVA in the K5-mOVA mice indicated that this is related to<br />

capture in the skin, as capture in the lymph nodes should lead to<br />

similar access and cross-presentation by CD8a + DCs. Although the<br />

possibility that virus was captured in the draining lymph nodes cannot<br />

be completely excluded, the finding that all migratory DCs presented<br />

MHC class II–restricted viral antigens yet only CD103 + DCs presented<br />

these antigens in an MHC class I–restricted way supports the main<br />

point of this report, which is that CD103 + DCs are the dominant<br />

migratory subset that cross-presents viral antigen. Our data are mostly<br />

compatible with published studies of HSV-2 (ref. 32) and HSV-1<br />

(ref. 33) showing that a dermal-like population of DCs provides a<br />

principal antigen-presenting contribution to both CD4 + and CD8 +<br />

T cell antiviral responses. Notably, we specifically separated CD103 +<br />

DCs from classical dermal DCs and identified the CD103 + DCsasthe<br />

key dominant cross-presenting migratory population.<br />

A corollary of the conclusion that CD103 + DCs are the dominant<br />

cross-presenting population of migratory DCs from the skin is that<br />

both Langerhans cells and classical dermal DCs are inefficient at crosspresentation.<br />

Earlier studies have provided evidence that Langerhans<br />

cells do not prime CD8 + T cell responses to HSV 9,16 .Thatdoesnot<br />

492 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

exclude the possibility that they could transport virus or even<br />

present viral antigens to CD4 + T cells, as we have now shown here,<br />

which raises the possibility that these classical epidermal DCs contribute<br />

to priming. Earlier studies linked Langerhans cells to the crosspresentation<br />

of both cellular and soluble OVA expressed or introduced<br />

into the skin, respectively 23,34 . However, the authors of those studies<br />

were unaware of CD103 + DCs, which also express langerin. Thus,<br />

much of the earlier data indicating involvement of Langerhans cells<br />

can be explained by functions associated with contaminating CD103 +<br />

DCs. Notably, however, the use of bone marrow chimeras has shown<br />

that radioresistant cells are able to present antigen and induce deletion<br />

of naive CD8 + T cells in the K5-mOVA model, which suggests some<br />

role for Langerhans cells 23 . Similarly, here we have shown weak but<br />

detectable MHC class I–restricted presentation of either HSV-1 or<br />

OVA by Langerhans cells, although, as in the other studies, it was<br />

difficult to exclude the possibility of residual presentation by contaminating<br />

CD103 + DCs.<br />

The disparate cross-presenting abilities of classical dermal DCs and<br />

CD103 + DCs is reminiscent of differences between lymphoid tissue–<br />

resident CD8a + and CD8a – DC subtypes, in which CD8a + DCs<br />

dominate 13,35 . Notably, CD8a + DCs and CD103 + DCs have many<br />

similarities 26 , including expression of langerin and CD103, relatively<br />

low expression of CD11b, responsiveness to Toll-like receptor 3<br />

ligands 36 and ablation in mice deficient for the transcription factor<br />

Batf3 (ref. 37). Given such similarities, it is plausible that the<br />

precursors of these DCs are related, with differences between<br />

CD103 + DCs and CD8a + DCs being determined by the type of tissue<br />

their precursor seeds (lymphoid versus skin).<br />

Although langerin-positive CD103 + CD11b lo DCs have been identified<br />

in the skin only recently, several studies have identified a similar<br />

DC type in other organs or their draining lymph nodes 19,26,36 .<br />

Examination of the presentation of influenza viral antigens by<br />

those DC subsets after lung infection has shown that both CD8a +<br />

DCs and CD103 + DCs present viral antigens to CD8 + Tcells 18,38 ,<br />

although only the CD103 + subset migrates from the lung 18 .Wehave<br />

defined this migratory subset mainly on the basis of their low<br />

expression of CD11b relative to that of other migratory DCs. Similar<br />

DCs have been found in lymph nodes draining the kidney, liver,<br />

pancreas and lung 18 . One group, using langerin and CD103 to<br />

identify this subset in the lungs and mediastinal lymph nodes, has<br />

found that whereas they present viral antigens to CD8 + T cells, they<br />

are not the main subset containing viral material; this is present in<br />

both plasmacytoid DCs and a migratory CD11b + DC population 38 .<br />

Such findings are consistent with our conclusion that MHC class I–<br />

restricted presentation by CD103 + DCs is not a consequence of<br />

‘preferential’ access to viral material but instead is a consequence of<br />

their dominant cross-presenting ability. Our view is further substantiated<br />

by studies examining cross-presentation in the lung of introduced<br />

innocuous antigen in the form of soluble OVA 39 . Again, these<br />

studies have found CD103 + DCs are the dominant subset crosspresenting<br />

OVA, whereas other CD11b + DCs ‘preferentially’ present it<br />

to CD4 + T cells 39 . Another group extensively examining the presentation<br />

of particulate and soluble OVA introduced into the lung has<br />

concluded that CD103 + DCs are the most effective at cross-presenting<br />

antigens associated with latex beads 40 . However, in that case, ‘preferential’<br />

capture might have been responsible. Our study has<br />

demonstrated the dominant cross-presenting ability of CD103 +<br />

DCs among the migratory populations in the skin but also raises<br />

the possibility that this DC subtype might be broadly represented in<br />

many tissues, mainly for its function in the generation of CD8 + T cell<br />

immunity and tolerance.<br />

ARTICLES<br />

The ability of CD103 + DCs to present viral antigens raises the issue<br />

of whether they are fully able to prime effective viral immunity. It is<br />

difficult to address this issue for HSV-1 infection, as CD8a + DCs also<br />

contribute to presentation. It is evident that CD8a + DCs can prime<br />

CTL immunity on their own when HSV-1 is introduced intravenously<br />

18 or when infection is cut short by excision of the site of<br />

infection at 8 h (ref. 41). Thus, it is unlikely that removal of CD103 +<br />

DCs by approaches such as use of the langerin–diphtheria toxin<br />

receptor system 22,24–26 will substantially ablate priming. To address<br />

this issue directly, a system is needed that will allow removal of CD8a +<br />

DCs while leaving CD103 + DCs intact but able to be depleted.<br />

As a final point of discussion, it is worth revisiting the function of<br />

Langerhans cells in immunity to HSV-1. It has been shown that<br />

Langerhans cells are not able to generate CTL immunity to this<br />

virus 16 . That finding is consistent with their poor ability to present<br />

viral antigen to CD8 + T cells shown here on both day 2 and day 5.<br />

Why Langerhans cells fail to effectively prime HSV-1-specific CTL<br />

responses remains a mystery, but if they are poorly endowed with the<br />

ability to cross-present, then presentation of MHC class I–restricted<br />

viral antigens would depend entirely on infection and access to the<br />

classical MHC class I pathway. As extensive evidence indicates that<br />

DCs infected with HSV-1 respond poorly to chemokines 42 ,failto<br />

upregulate costimulatory molecules 42 , die rapidly and are poorly<br />

stimulatory 43,44 , it is perhaps not unexpected that Langerhans cells,<br />

even if infected, would be ineffective at priming HSV-specific CTL<br />

immunity. Whether infected and dying Langerhans cells represent a<br />

likely source of antigenic material for CD103 + DCs to cross-present as<br />

they attempt to migrate to the draining lymph nodes remains an open<br />

issue. Our data indicate that CD103 + DCs can sample antigens from<br />

the epidermis, despite their proposed residence in the dermis 26–28 ,as<br />

they cross-presented OVA from the epithelia of K5-mOVA mice. That<br />

view is further supported by evidence that CD103 + DCs may be able<br />

to reach their dendrites through to the epidermis 26 .<br />

Our studies have indicated that CD103 + DCs are the dominant<br />

migratory subset involved in MHC class I–restricted presentation of<br />

both self and pathogen antigens. This property probably relates to a<br />

specialized ability to cross-present cell-associated antigens. Such a<br />

‘division of labor’ raises the issue of what alternative specializations<br />

other DC subsets might be endowed with. It also suggests that<br />

vaccination strategies targeting the generation of CTL immunity<br />

might require careful consideration of how to best ‘bait’ vaccine<br />

antigens for the CD103 + subset.<br />

METHODS<br />

Mice. C57BL/6 gBT-I mice 45 on a C57BL/6.SJL-PtprcaPep3b/BoyJ background<br />

(Ly5.1 gBT-I), gDT-II mice, OT-I mice (ovalbumin-specific TCR-transgenic<br />

mice) 46 and K5.mOVA mice 31 were bred and maintained at The Walter and<br />

Eliza Hall Institute of Medical Research animal facility or the Department of<br />

Microbiology and <strong>Immunology</strong> at the University of Melbourne. The gDT-II.1<br />

mice express transgenes encoding T cell antigen receptors obtained from<br />

a Va3.2 + Vb2 + I-A b -restricted HSV-1-specific T cell hybridoma (P.W., S.B.,<br />

G. Davey, W.R.H., F.R.C., A.G.B., unpublished data). Animal experiments were<br />

approved by the University of Melbourne Animal Ethics Committee or the<br />

Walter and Eliza Hall Institute Animal Ethics Committee.<br />

Viruses and viral infection. The wild-type parental HSV-1 strain KOS (HSV-1)<br />

and thymidine kinase–deficient HSV KOS.Cre (HSV-TK – ) 47 were grown in<br />

minimal essential medium with 10% (vol/vol) heat-inactivated FCS and were<br />

titrated on Vero cells (CSL). Unless stated otherwise, mice were infected with<br />

1 10 6 plaque-forming units of HSV-1 or HSV-TK – as described 30 .Insome<br />

experiments, mice were infected on the upper flank proximal to the dorsal<br />

midline on one side or both sides of the mice or on the lower flank proximal to<br />

the ventral midline.<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 493


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

DC isolation and sorting strategy. Single-cell suspensions were prepared from<br />

brachial or axillary lymph nodes of mice infected 2, 4 or 5 d earlier with HSV-1<br />

(for HSV experiments) or from pooled inguinal, cervical and brachial lymph<br />

nodes of naive K5-mOVA mice (for K5-mOVA experiments). Lymph node<br />

suspensions were enriched for conventional DCs (excluding plasmacytoid DCs)<br />

with antibody to Gr-1 (anti-Gr-1; RB6-8C5), anti-CD3e (KT3), anti-CD19<br />

(ID6), anti-Thy-1 (T24), anti-B220 (RA36B2) and antibody to erythrocytes<br />

(anti-Ter119), followed by depletion with magnetic beads as described 9,48 .<br />

Enrichment of the cell suspension usually yielded between 80% and 85%<br />

CD11c + cells. Cells were stained with phycoerythrin-indotricarbocyanine–<br />

labeled anti-CD11c (N418; BD Biosciences), phycoerythrin-labeled anti-CD8<br />

(53-6.7; BD Biosciences), allophycocyanin-labeled anti-CD205 (NLDC; prepared<br />

‘in house’), biotinylated anti-CD326 (anti-Ep-CAM; G8.8 BioLegend)<br />

and FITC-labeled anti-CD103 (2E7; BD Biosciences). DC subtypes from<br />

propidium iodide–negative CD11c + events were sorted by flow cytometry<br />

(FACSAria; BD Biosciences) according to the following two strategies: in one,<br />

expression of CD205 and CD8 was analyzed in CD11c + cells for the identification<br />

of Langerhans cells, dermal DCs, double-negative DCs and CD8a + DCs<br />

(Fig. 3a); in the other, three sorting steps were used (Fig. 4a), the first to<br />

identify CD8a + DCs (CD11c versus CD8), the second to identify CD103 + DCs<br />

among all non-CD8a + DCs (CD205 versus CD103) and the third to categorize<br />

CD8a – CD103 – DCs as Langerhans cells and dermal DCs on the basis of CD326<br />

expression. After sorting, DC subtypes were washed and resuspended in RPMI<br />

medium supplemented with 10% (vol/vol) FCS, 2-mercaptoethanol (50 mM),<br />

L-glutamine (2 mM), penicillin (100 U/ml) and streptomycin (100 mg/ml) for<br />

culture together with T cells. The purity of sorted subsets was routinely<br />

over 94%.<br />

Preparation of T cells. OT-I, gBT-I and gDT-II T cells were purified as<br />

described 23 . T cells were isolated from lymph nodes and spleens, and samples<br />

were enriched by incubation for 30 min with the following antibodies: anti-<br />

Mac-1 (M1/70), anti-F4/80 (F4/80), antibody to erythrocytes (anti-Ter119),<br />

anti-Gr-1 (RB6-8C5), anti-I-A/E (M5114) and either anti-CD4 (GK1.5)<br />

for enrichment for CD8 + T cells or anti-CD8 (53.6-7) for enrichment for<br />

CD4 + T cells (gDT-II). Cells that bound antibodies were removed with magnetic<br />

beads coupled to goat anti–rat immunoglobulin G (Qiagen, Victoria,<br />

Australia). Cells were routinely 90–95% pure, as determined by flow cytometry.<br />

Proliferation of T cells. T cells were labeled with 2.5 mM CFSE (carboxyfluorescein<br />

diacetate succinimidyl ester; Sigma) as described 48 . Forin vivo<br />

proliferation, 1 10 6 CFSE-labeled CD8 + Ly5.1 gBT-I T cells were<br />

transferred intravenously into C57BL/6 mice on day 2 or day 5 after flank<br />

infection with HSV-1 or HSV-TK – . T cell proliferation in the brachial, axillary<br />

or mesenteric lymph nodes was assessed 42 h later by flow cytometry<br />

(FACSCalibur or LSRII) after staining of single-cell suspensions with phycoerythrin-labeled<br />

anti-Ly5.1 (A20; BD Biosciences) and allophycocyanin-labeled<br />

anti-CD8 (53-6.7; BD Biosciences). For ex vivo antigen-presentation assays, 5<br />

10 4 CFSE-labeled gBT-I, gDT-II or OT-I cells were cultured together for 60 h<br />

with various concentrations of purified DC subtypes from HSV-1-infected or<br />

K5-mOVA mice as described 16,23 . Proliferation was measured by flow cytometry<br />

(FACSCalibur or LSRII) as CFSE dilution by CD8 + (allophycocyaninlabeled<br />

anti-CD8; 53-6.7; BD Biosciences), Va2 + (phycoerythrin-labeled<br />

anti-Va 2; B20.1; BD Biosciences) T cells for gBT-I and OT-I; or as CD4 +<br />

(allophycocyanin-labeled anti-CD4; GK1.5; BD Biosciences), V 3.2 + (biotinylated<br />

RR3-16; BD Biosciences) cells for gDT-II.<br />

Immunofluorescence. Skin samples were collected from primary and secondary<br />

sites of infection and were immediately frozen in Tissue-Tek optimal<br />

cutting temperature compound (Sakura Finetek). Acetone-fixed sections 8 mm<br />

in thickness were stained with polyclonal rabbit anti–HSV immunoglobulin<br />

(N1562; Dako) or control rabbit immunoglobulin (X0903; Dako) and were<br />

visualized with Alexa Fluor 594–conjugated anti–rabbit immunoglobulin<br />

(a21207; Molecular Probes; Invitrogen). Slides were mounted with Vectashield<br />

containing DAPI (4,6-diamidino-2-phenylindole; Vector Laboratories). Confocal<br />

images were obtained with a Meta 5110 confocal microscope (Leica) and<br />

were analyzed with ImageJ software (US National Institutes of Health).<br />

Painting with FITC. For monitoring of the migration of DCs from the skin<br />

during HSV infection, at 2 d after viral inoculation, an area of flank<br />

skin corresponding to the secondary site was painted with 20 ml of a 1%<br />

(vol/vol) solution of FITC (Isomer 1; Sigma-Aldrich) prepared in acetone<br />

(Sigma-Aldrich). FITC was first dissolved as a 10% (wt/vol) solution in<br />

dimethylsulfoxide (Sigma-Aldrich) and then was diluted to 1% (wt/vol) in<br />

acetone. For experiments involving monitoring of DCs after skin irritation,<br />

anaesthetized mice were depilated and were painted with the 1% FITC solution<br />

described above, or they were painted with a 1% (vol/vol) FITC solution<br />

prepared in acetone and dibutyl phthalate (1:1) as described 9 . DCs were<br />

enriched from the axillary lymph nodes as described above and migration<br />

after either HSV infection or skin irritation was assessed by flow cytometry 3 d<br />

after FITC application after cells were stained with Alexa Fluor 700–conjugated<br />

anti-CD11c (N418; BD Biosciences), phycoerythrin-indotricarbocyanine–<br />

labeled anti-CD8 (53-6.7; BD Biosciences), allophycocyanin-labeled anti-<br />

CD205 (NLDC), biotinylated anti-CD326 (G8.8) and phycoerythrin-labeled<br />

anti-CD103 (2E7; BD Biosciences).<br />

Cytokine bead array. The secretion of cytokines into supernatants after culture<br />

of gDT-II T cells together with various DC subsets was analyzed by cytometric<br />

bead array (Mouse Th1/Th2 Cytokine kit (interleukins 2, 4 and 5); BD<br />

Biosciences) according to the manufacturer’s instructions.<br />

Measurement of viral titers. Mice were infected with 1 10 6 plaque-forming<br />

units of HSV-TK – or wild-type HSV-1 (strain KOS) by flank scarification. The<br />

primary inoculation site, defined as a 5-mm 5-mm full-thickness piece of<br />

skin encompassing the area of scarification, was excised and homogenized. Skin<br />

removed from the secondary site was 5 mm in width and extended from 5 mm<br />

below the inoculation site to the ventral midline. Infectious virus in the tissue<br />

was measured by standard assay of plaque-forming units on confluent Vero cell<br />

monolayers (CSL) as described 30 .<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank the flow cytometry facilities and animal facility staff of the Walter<br />

and Eliza Hall Institute of Medical Research and K. Field; we also thank<br />

B. Davies and J. Langley for technical assistance. Supported by Deutsche<br />

Forschungsgemeinschaft (BE 3285-1/2 to S.B.), the National Health and<br />

Medical Research Council of Australia (W.R.H., F.R.C., A.G.B. and P.G.W.)<br />

and the Howard Hughes Medical Institute (W.R.H.).<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Steinman, R.M. Lasker Basic Medical Research Award. Dendritic cells: versatile<br />

controllers of the immune system. Nat. Med. 13, 1155–1159 (2007).<br />

2. Villadangos, J.A. & Heath, W.R. Life cycle, migration and antigen presenting functions<br />

of spleen and lymph node dendritic cells: limitations of the Langerhans cells paradigm.<br />

Semin. Immunol. 17, 262–272 (2005).<br />

3. Shortman, K. & Liu, Y. Mouse and human dendritic cell subtypes. Nat. Rev. Immunol.<br />

2, 151–161 (2002).<br />

4. Shortman, K. & Naik, S.H. Steady-state and inflammatory dendritic-cell development.<br />

Nat. Rev. Immunol. 7, 19–30 (2007).<br />

5. Heath, W.R. et al. Cross-presentation, dendritic cell subsets, and the generation of<br />

immunity to cellular antigens. Immunol. Rev. 199, 9–26(2004).<br />

6. Jakubzick, C. et al. Lymph-migrating, tissue-derived dendritic cells are minor constituents<br />

within steady-state lymph nodes. J. Exp. Med. 205, 2839–2850 (2008).<br />

7. Wakim, L.M., Waithman, J., van Rooijen, N., Heath, W.R. & Carbone, F.R. Dendritic<br />

cell-induced memory T cell activation in nonlymphoid tissues. Science 319, 198–202<br />

(2008).<br />

8. Itano, A.A. et al. Distinct dendritic cell populations sequentially present antigen to<br />

CD4 + T cells and stimulate different aspects of cell-mediated immunity. Immunity 19,<br />

47–57 (2003).<br />

9. Allan, R.S. et al. Migratory dendritic cells transfer antigen to a lymph node-resident<br />

dendritic cell population for efficient CTL priming. Immunity 25, 153–162 (2006).<br />

10. Cella, M. et al. Plasmacytoid monocytes migrate to inflamed lymph nodes and produce<br />

large amounts of type I interferon. Nat. Med. 5, 919–923 (1999).<br />

11. Diebold, S.S. et al. Viral infection switches non-plasmacytoid dendritic cells into high<br />

interferon producers. <strong>Nature</strong> 424, 324–328 (2003).<br />

12. den Haan, J.M. & Bevan, M.J. Constitutive versus activation-dependent crosspresentation<br />

of immune complexes by CD8 + and CD8 – dendritic cells in vivo. J. Exp.<br />

Med. 196, 817–827 (2002).<br />

494 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

13. Dudziak, D. et al. Differential antigen processing by dendritic cell subsets in vivo.<br />

Science 315, 107–111 (2007).<br />

14. Schnorrer, P. et al. The dominant role of CD8 + dendritic cells in cross-presentation is<br />

not dictated by antigen capture. Proc. Natl. Acad. Sci. USA 103, 10729–10734<br />

(2006).<br />

15. Smith, C.M. et al. Cutting edge: conventional CD8a + dendritic cells are preferentially<br />

involved in CTL priming after footpad infection with herpes simplex virus-1.<br />

J. Immunol. 170, 4437–4440 (2003).<br />

16. Allan, R.S. et al. Epidermal viral immunity induced by CD8a + dendritic cells but not by<br />

Langerhans cells. Science 301, 1925–1928 (2003).<br />

17. Belz, G.T., Shortman, K., Bevan, M.J. & Heath, W.R. CD8a + dendritic cells selectively<br />

present MHC class I-restricted noncytolytic viral and intracellular bacterial antigens in<br />

vivo. J. Immunol. 175, 196–200 (2005).<br />

18. Belz, G.T. et al. Cutting edge: Conventional CD8a + dendritic cells are generally involved<br />

in priming CTL immunity to viruses. J. Immunol. 172, 1996–2000 (2004).<br />

19. Belz, G.T. et al. Distinct migrating and nonmigrating dendritic cell populations are<br />

involved in MHC class I-restricted antigen presentation after lung infection with virus.<br />

Proc. Natl. Acad. Sci. USA 101, 8670–8675 (2004).<br />

20. Wilson, N.S. et al. Systemic activation of dendritic cells by Toll-like receptor ligands or<br />

malaria infection impairs cross-presentation and antiviral immunity. Nat. Immunol. 7,<br />

165–172 (2006).<br />

21. He, Y., Zhang, J., Donahue, C. & Falo, L.D. Jr. Skin-derived dendritic cells induce<br />

potent CD8 + T cell immunity in recombinant lentivector-mediated genetic immunization.<br />

Immunity 24, 643–656 (2006).<br />

22. Bennett, C.L. et al. Inducible ablation of mouse Langerhans cells diminishes but fails<br />

to abrogate contact hypersensitivity. J. Cell Biol. 169, 569–576 (2005).<br />

23. Waithman, J. et al. Skin-derived dendritic cells can mediate deletional tolerance of<br />

class I-restricted self-reactive T cells. J. Immunol. 179, 4535–4541 (2007).<br />

24. Kaplan, D.H., Jenison, M.C., Saeland, S., Shlomchik, W.D. & Shlomchik, M.J.<br />

Epidermal langerhans cell-deficient mice develop enhanced contact hypersensitivity.<br />

Immunity 23, 611–620 (2005).<br />

25. Kissenpfennig, A. et al. Dynamics and function of Langerhans cells in vivo dermal<br />

dendritic cells colonize lymph node areas distinct from slower migrating Langerhans<br />

cells. Immunity 22, 643–654 (2005).<br />

26. Bursch, L.S. et al. Identification of a novel population of Langerin + dendritic cells.<br />

J. Exp. Med. 204, 3147–3156 (2007).<br />

27. Poulin, L.F. et al. The dermis contains langerin + dendritic cells that develop and<br />

function independently of epidermal Langerhans cells. J. Exp. Med. 204, 3119–3131<br />

(2007).<br />

28. Ginhoux, F. et al. Blood-derived dermal langerin+ dendritic cells survey the skin in the<br />

steady state. J. Exp. Med. 204, 3133–3146 (2007).<br />

29. Simmons, A. & Nash, A.A. Zosteriform spread of herpes simplex virus as a model of<br />

recrudescence and its use to investigate the role of immune cells in prevention of<br />

recurrent disease. J. Virol. 52, 816–821 (1984).<br />

30. van Lint, A. et al. Herpes simplex virus-specific CD8 + T cells can clear established lytic<br />

infections from skin and nerves and can partially limit the early spread of virus after<br />

cutaneous inoculation. J. Immunol. 172, 392–397 (2004).<br />

ARTICLES<br />

31. Azukizawa, H. et al. Induction of T-cell-mediated skin disease specific for antigen<br />

transgenically expressed in keratinocytes. Eur. J. Immunol. 33, 1879–1888<br />

(2003).<br />

32. Zhao, X. et al. Vaginal submucosal dendritic cells, but not Langerhans cells, induce<br />

protective Th1 responses to herpes simplex virus-2. J. Exp. Med. 197, 153–162<br />

(2003).<br />

33. Lee, H.K. et al. Differential roles of migratory and resident DCs in T cell priming after<br />

mucosal or skin HSV-1 infection. J. Exp. Med. 206, 359–370 (2009).<br />

34. Stoitzner, P. et al. Langerhans cells cross-present antigen derived from skin. Proc. Natl.<br />

Acad. Sci. USA 103, 7783–7788 (2006).<br />

35. den Haan, J.M., Lehar, S.M. & Bevan, M.J. CD8 + but not CD8 – dendritic cells crossprime<br />

cytotoxic T cells in vivo. J. Exp. Med. 192, 1685–1696 (2000).<br />

36. Sung, S.S. et al. A major lung CD103 (a E)-b 7 integrin-positive epithelial dendritic cell<br />

population expressing Langerin and tight junction proteins. J. Immunol. 176,<br />

2161–2172 (2006).<br />

37. Hildner, K. et al. Batf3 deficiency reveals a critical role for CD8a + dendritic cells in<br />

cytotoxic T cell immunity. Science 322, 1097–1100 (2008).<br />

38. GeurtsvanKessel, C.H. et al. Clearance of influenza virus from the lung depends on<br />

migratory langerin + CD11b – but not plasmacytoid dendritic cells. J. Exp. Med. 205,<br />

1621–1634 (2008).<br />

39. del Rio, M.L., Rodriguez-Barbosa, J.I., Kremmer, E. & Forster, R. CD103 – and CD103 +<br />

bronchial lymph node dendritic cells are specialized in presenting and cross-presenting<br />

innocuous antigen to CD4 + and CD8 + T cells. J. Immunol. 178, 6861–6866<br />

(2007).<br />

40. Jakubzick, C., Helft, J., Kaplan, T.J. & Randolph, G.J. Optimization of methods to study<br />

pulmonary dendritic cell migration reveals distinct capacities of DC subsets to acquire<br />

soluble versus particulate antigen. J. Immunol. Methods 337, 121–131 (2008).<br />

41. Stock, A.T., Mueller, S.N., van Lint, A.L., Heath, W.R. & Carbone, F.R. Cutting edge:<br />

prolonged antigen presentation after herpes simplex virus-1 skin infection. J. Immunol.<br />

173, 2241–2244 (2004).<br />

42. Salio, M., Cella, M., Suter, M. & Lanzavecchia, A. Inhibition of dendritic cell<br />

maturation by herpes simplex virus. Eur. J. Immunol. 29, 3245–3253 (1999).<br />

43. Kruse, M. et al. Mature dendritic cells infected with herpes simplex virus type 1 exhibit<br />

inhibited T-cell stimulatory capacity. J. Virol. 74, 7127–7136 (2000).<br />

44. Jones, C.A. et al. Herpes simplex virus type 2 induces rapid cell death and functional<br />

impairment of murine dendritic cells in vitro. J. Virol. 77, 11139–11149 (2003).<br />

45. Mueller, S.N., Heath, W., McLain, J.D., Carbone, F.R. & Jones, C.M. Characterization of<br />

two TCR transgenic mouse lines specific for herpes simplex virus. Immunol. Cell Biol.<br />

80, 156–163 (2002).<br />

46. Hogquist, K.A. et al. T cell receptor antagonist peptides induce positive selection.<br />

Cell 76, 17–27 (1994).<br />

47. Wakim, L.M., Jones, C.M., Gebhardt, T., Preston, C.M. & Carbone, F.R. CD8 + T-cell<br />

attenuation of cutaneous herpes simplex virus infection reduces the average viral copy<br />

number of the ensuing latent infection. Immunol. Cell Biol. 86, 666–675 (2008).<br />

48. Belz, G.T., Bedoui, S., Kupresanin, F., Carbone, F.R. & Heath, W.R. Minimal activation<br />

of memory CD8 + T cell by tissue-derived dendritic cells favors the stimulation of naive<br />

CD8 + Tcells.Nat. Immunol. 8, 1060–1066 (2007).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 495


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Divergent functions for airway epithelial<br />

matrix metalloproteinase 7 and retinoic acid<br />

in experimental asthma<br />

Sangeeta Goswami 1 , Pornpimon Angkasekwinai 2 , Ming Shan 1 , Kendra J Greenlee 3 , Wade T Barranco 4 ,<br />

Sumanth Polikepahad 4 , Alexander Seryshev 4 , Li-zhen Song 4 , David Redding 5 , Bhupinder Singh 5 , Sanjiv Sur 5 ,<br />

Prescott Woodruff 6 , Chen Dong 2 , David B Corry 1,4 & Farrah Kheradmand 1,4<br />

The innate immune response of airway epithelial cells to airborne allergens initiates the development of T cell responses that<br />

are central to allergic inflammation. Although proteinase allergens induce the expression of interleukin 25, we show here that<br />

epithelial matrix metalloproteinase 7 (MMP7) was expressed during asthma and was required for the maximum activity of<br />

interleukin 25 in promoting the differentiation of T helper type 2 cells. Allergen-challenged Mmp7 –/– mice had less airway<br />

hyper-reactivity and production of allergic inflammatory cytokines and higher expression of retinal dehydrogenase 1. Inhibition of<br />

retinal dehydrogenase 1 restored the asthma phenotype of Mmp7 –/– mice and inhibited the responses of lung regulatory T cells,<br />

whereas exogenous administration of retinoic acid attenuated the asthma phenotype. Thus, MMP7 coordinates allergic lung<br />

inflammation by activating interleukin 25 while simultaneously inhibiting retinoid-dependent development of regulatory T cells.<br />

Proximal and distal airway epithelial cells directly communicate with<br />

the outside environment, and thus their response to inhaled allergens<br />

could be integral to the initiation of allergic lung inflammation.<br />

Identification of the critical epithelium-derived mediators that drive<br />

allergic lung responses is at an early stage but is essential to complete<br />

understanding of the molecular mechanisms that underlie diseases<br />

such as asthma. After being inhaled, proteinase allergens (such as<br />

airborne allergens with active proteinase properties) rapidly induce a<br />

powerful inflammatory response that initiates the recruitment of<br />

allergic immune cells into the lung 1,2 . Although the function of<br />

cytokines expressed in the airway epithelium in asthma has been<br />

studied extensively 3,4 , the mechanism by which allergens initiate the<br />

differentiation of airway T helper type 2 (TH2) cells is poorly understood.<br />

Interleukin 25 (IL-25; also called IL-17E) is a member of the<br />

IL-17 family of cytokines that is produced mainly in the airway<br />

epithelium in response to proteinase allergens 5,6 . Transgenic expression<br />

of human and mouse IL-25 induces T H2 immune responses<br />

marked by considerable lung eosinophilia, mucus hyperproduction<br />

and higher expression of IL-4, IL-5, IL-13 and CCL11 (eotaxin 1) in<br />

the airway 7,8 . Neutralization of IL-25 by blocking antibodies has been<br />

shown to reverse airway hyper-reactivity, a pathognomonic feature of<br />

asthma, in a mouse model of asthma 9 . Moreover, IL-25-deficient mice<br />

are unable to develop a strong TH2-biased immune response and fail<br />

Received 12 January; accepted 12 February; published online 29 March 2009; doi:10.1038/ni.1719<br />

to clear gut parasitic infection with Nippostrongylus brasiliensis or<br />

Trichuris muris 10 . Such studies suggest a critical function for epithelial<br />

cells in providing the IL-25 signals that may be required for effective<br />

T H2 responses. Although the function of IL-25 in allergic lung disease<br />

is well known, a broader understanding of how IL-25 interacts with<br />

other airway epithelial factors to coordinately regulate allergic<br />

responses is needed.<br />

Matrix metalloproteinases (MMPs) are zinc-dependent enzymes<br />

that are induced in response to many stimuli. Although MMPs were<br />

originally shown to mediate the turnover of extracellular matrix<br />

molecules, it is now apparent that they control diverse biological<br />

processes unrelated to matrix degradation 11 . Although the exact<br />

function of MMPs in chronic models of lung inflammation is<br />

debatable, it is clear that several members of this family are acutely<br />

induced either locally or systemically at the onset of allergen challenge<br />

12–16 . MMP2 and MMP9 are not required for the development of<br />

allergic lung disease but serve key functions in the clearance of<br />

inflammatory cells from lung parenchyma by establishing the necessary<br />

trans-epithelial chemokine gradients 15,16 . However, whereas<br />

neither MMP2 nor MMP9 is expressed in the airway epithelium 17 ,<br />

MMP7 (also called matrilysin; A001477), an epithelial cell–specific<br />

MMP, is induced in the lung and gut during inflammation 18,19 .<br />

MMP7 has been shown to modify pro-a-defensin, the ligand for the<br />

1 Department of <strong>Immunology</strong>, Baylor College of Medicine, Houston, Texas, USA. 2 Department of <strong>Immunology</strong>, University of Texas MD Anderson Cancer Center, Houston,<br />

Texas, USA. 3 Department of Biological Sciences, North Dakota State University, Fargo, North Dakota, USA. 4 Department of Medicine, Baylor College of Medicine, Houston<br />

Texas, USA. 5 Department of Medicine, University of Texas Medical Branch Galveston, Galveston, Texas, USA. 6 Department of Medicine, University of California San<br />

Francisco, San Francisco, California, USA. Correspondence should be addressed to F.K. (farrahk@bcm.edu) or D.B.C. (dcorry@bcm.edu).<br />

496 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Figure 1 MMP7 is induced in allergic<br />

inflammation and modulates IL-25 function.<br />

(a) Immunohistochemistry of MMP7 in lungs of<br />

PBS-treated (WT PBS) and complete Aspergillus<br />

allergen (CAA)-immunized (WT CAA) C57BL/6<br />

mice. Arrows and arrowheads indicate MMP7<br />

expression; insets, 40 magnification of the<br />

airway. Scale bars, 100 mm (main images) and<br />

25 mm (insets). Data are representative of at<br />

least three independent experiments. (b) Silver<br />

staining of rIL-25 (5 mg), rIL-13 (5 mg) and<br />

rTSLP (5 mg) left untreated or incubated for 2 h<br />

at 37 1C with 4-aminophenylmercuric acetate–<br />

activated MMP7 (0.5 mg; + MMP7) and then<br />

separated by electrophoresis through a 16.5%<br />

tricine gel. Arrowheads indicate MMP7-cleaved<br />

fragments of rIL-25. Results are representative<br />

of at least three independent experiments.<br />

(c) ELISA of IL-4, IL-5, IL-13 and interferon-g<br />

(IFN-g) in supernatants of naive lymph node cells<br />

(left) and spleen cells (right) isolated from<br />

C57BL/6 mice and stimulated for 2 d with platebound<br />

anti-CD3 (2 mg/ml) in the presence of<br />

equal amounts of native rIL-25 (250 ng/ml)<br />

or rIL-25’C (250 ng/ml). Media alone and the<br />

4-aminophenylmercuric acetate–containing<br />

buffer used to activate MMP7 (Vehicle) serve as<br />

controls. *, P o 0.05, versus rIL-25 (one-way<br />

ANOVA). Data are representative of three<br />

different experiments (mean and s.d.;<br />

Lymph<br />

*<br />

n ¼ 3 samples). (d) Flow cytometry of the expression of IL-4 and IL-5 by lymphocytes and spleen cells cultured for 3 d in the conditions described in c,<br />

then restimulated for 5 h with ionomycin (500 ng/ml) and phorbol 12-myristate 13-acetate (50 ng/ml) in the presence of GolgiStop, followed by<br />

intracytoplasmic staining. Numbers adjacent to outlined areas indicate percent IL-5 + Il-4 – cells (top left) or IL-5 – Il-4 + cells (bottom right). Data are<br />

representative of three independent experiments.<br />

cell surface receptor Fas (FasL; CD95L) and tumor necrosis factor in<br />

the gut during defense against enteric pathogens, but its function in<br />

allergic inflammation has yet to be explored 19,20 .<br />

In this study, we examine how airway epithelial IL-25 coordinates<br />

proteinase-dependent allergic lung disease in the context of other<br />

potentially relevant airway epithelium-derived factors, including<br />

MMP7. We show that in response to proteinase allergen or recombinant<br />

IL-25 (rIL-25), mouse airway epithelial cells expressed MMP7.<br />

Activation of MMP7 was critical for IL-25 function because MMP7cleaved<br />

rIL-25 (called ‘rIL-25’C’ here) enhanced T H2 differentiation.<br />

Furthermore, we show that humans with chronic asthma also<br />

expressed MMP7 and IL-25 in their distal airspaces and, in response<br />

to ragweed extract, a potent allergenic stimuli, patients with a history<br />

of allergy had significantly more nasal secretion of MMP7. In the<br />

absence of MMP7, mice showed attenuated allergic responses to<br />

challenge with proteinase allergen and enhanced expression of retinal<br />

dehydrogenase 1 (RALDH-1), a rate-limiting enzyme for the production<br />

of retinoic acid. Moreover, Mmp7 –/– mice had more regulatory<br />

T cells in the lung parenchyma. Our results collectively support the<br />

idea that MMP7 serves as a proinflammatory mediator that specifically<br />

enhances the function of IL-25 that is necessary for robust TH2<br />

responses. Finally, we demonstrate a tolerogenic mechanism initiated<br />

by airway epithelial cells in response to challenge with proteinase<br />

allergen in which RALDH-1 was induced to promote immunosuppressive<br />

regulatory T cells.<br />

RESULTS<br />

MMP7 mediates enhanced function of IL-25<br />

MMP2 and MMP9 are not expressed in airway epithelial cells after<br />

allergen challenge 17 . In contrast, we found that a fungus-derived<br />

proteinase allergen strongly enhanced airway epithelial expression of<br />

a b<br />

IL-4 (ng/ml)<br />

IL-5 (ng/ml)<br />

IL-13 (ng/ml)<br />

IFN-γ (ng/ml)<br />

2<br />

1<br />

0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

150<br />

100<br />

50<br />

0<br />

150<br />

100<br />

50<br />

0<br />

Media<br />

WT PBS<br />

*<br />

*<br />

rIL-25<br />

rIL-25′C<br />

Vehicle<br />

4<br />

3<br />

2<br />

1<br />

0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

200<br />

100<br />

0<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Media<br />

WT CAA<br />

c d<br />

Spleen<br />

*<br />

*<br />

*<br />

rIL-25<br />

rIL-25′C<br />

Vehicle<br />

(kDa) (kDa)<br />

37<br />

20<br />

17.5<br />

15<br />

25<br />

20<br />

10<br />

12.5<br />

10<br />

Media<br />

rIL-25<br />

rIL-25′C<br />

IL-5<br />

MMP7<br />

rIL-25<br />

rIL-25 + MMP7<br />

ARTICLES<br />

rTSLP<br />

rTSLP + MMP7<br />

rIL-13<br />

rIL-13 + MMP7<br />

10<br />

Lymph<br />

Spleen<br />

0.14<br />

0.18<br />

0.27 0.88<br />

0.23 0.43<br />

0.43 2.34<br />

0.55 0.48<br />

0.95<br />

3.5<br />

IL-4<br />

4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

10 1<br />

10 1 10 0<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

MMP7 (Fig. 1a). Whereas induction of MMP2 and MMP9 is IL-13<br />

dependent 15 , MMP7 induction was independent of IL-13, as there<br />

was equivalent induction of MMP7 in wild-type mice and allergenchallenged<br />

mice deficient in STAT6, the principal mediator of the<br />

effects of IL-13 in the lung (data not shown). Because MMP7 is known<br />

to cleave and activate cytokines in the gut 13 , we next determined if<br />

MMP7 modifies cytokines critical for initiating allergic lung responses.<br />

In addition to IL-25, both IL-13 and thymic stromal lymphopoietin<br />

(TSLP) are critical participants in the molecular pathways that<br />

underlie allergic lung disease 5,21 . We found that although rIL-13 and<br />

rTSLP were not substrates for activated MMP7, in the same conditions,<br />

rIL-25 was cleaved at multiple sites (Fig. 1b and Supplementary<br />

Table 1 and Fig. 1 online). To determine the function of cleaved IL-25,<br />

we examined the responses of naive spleen and lymph node cells after<br />

activation by crosslinking of T cell antigen receptors in vitro.Wefound<br />

that rIL-25’C induced significantly more secretion of T H2 cytokines<br />

(IL-4, IL-5 and IL-13) than did native rIL-25 and it showed enhanced<br />

binding to a fusion protein of IL-17 receptor B and the Fc fragment<br />

in vitro (Fig. 1c,d and Supplementary Fig. 2a,b online). Indeed,<br />

native IL-25 had little effect on IL-4 secretion, but rIL-25’C enhanced<br />

IL-4 secretion tenfold, which suggested that the main active form of<br />

IL-25 is the MMP7-cleaved form. We found no effect of rIL-25 or<br />

rIL-25’C on the induction of interferon-g, a canonical T H1 cytokine<br />

(Fig. 1c). Collectively, these data indicate that IL-25 is a substrate for<br />

MMP7 and that cleavage of IL-25 by MMP7 is needed to drive robust<br />

T H2 responses.<br />

Attenuated asthma phenotype of Mmp7 –/– mice<br />

The greater potency of rIL-25’C in activating T H2 cells in vitro raised<br />

the possibility of an essential function for MMP7 in mediating allergic<br />

inflammation in vivo. To investigate this, we compared the responses<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 497


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

b<br />

WT<br />

60<br />

WT CAA<br />

Mmp7 *<br />

WT PBS<br />

*<br />

**<br />

PBS<br />

**<br />

0 0.03 0.1<br />

Ach (µg/g)<br />

0.3 1.0<br />

WT<br />

**<br />

CAA<br />

*<br />

*<br />

**<br />

*<br />

**<br />

–/– CAA<br />

Mmp7 –/– PBS<br />

Mmp7 –/–<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

40<br />

20<br />

0<br />

3<br />

2<br />

1<br />

0<br />

CAA – + – +<br />

60<br />

40<br />

20<br />

f 5<br />

4<br />

3<br />

2<br />

1<br />

*<br />

4<br />

3<br />

2<br />

1<br />

*<br />

0<br />

0<br />

0<br />

CAA<br />

25<br />

20<br />

15<br />

10<br />

– + – + rlL-25<br />

g 8<br />

6<br />

4<br />

– + + rlL-25<br />

200<br />

150<br />

100<br />

– + +<br />

5<br />

0<br />

2<br />

0<br />

ND *<br />

50<br />

0<br />

ND *<br />

CAA – + – +<br />

rlL-25 – + +<br />

rlL-25 – + +<br />

c<br />

R RS (cm H 2 O s/ml)<br />

PC200 (µg/g)<br />

Glycoprotein<br />

(µg/ml)<br />

d<br />

BAL eosinophils<br />

(×10 5 /ml)<br />

BAL total cells<br />

(×10 5 /ml)<br />

of Mmp7 –/– and wild-type mice to the proteinase allergen CAA<br />

(complete aspergillus allergen) extracted from fungi, a powerful<br />

allergen used before to elicit allergic lung disease in mice 1 . This airway<br />

inflammation model is characterized by a predominantly eosinophilic<br />

influx and airway obstruction due to enhanced glycoprotein secretion,<br />

as well as more airway hyper-reactivity to secondary challenge with<br />

acetylcholine 22,23 . Mmp7 –/– mice immunized with CAA had less airway<br />

hyper-reactivity than did wild-type mice, as shown by both their<br />

lower respiratory system resistance in response to many doses of<br />

acetylcholine and higher stimulating concentration of acetylcholine<br />

required to elicit a 200% change from the baseline respiratory system<br />

resistance. Moreover, Mmp7 –/– mice had less secretion of glycoproteins<br />

in bronchoalveolar lavage (BAL) fluid after CAA challenge (Fig. 2a–c).<br />

Airway eosinophilia was also significantly lower in CAA-immunized<br />

Mmp7 –/– mice (Fig. 2d). Because the egress of lung parenchymal<br />

eosinophils into the airways is impaired in Mmp2 –/– and Mmp9 –/–<br />

mice 15,16 , we assessed lung parenchymal eosinophils to rule out the<br />

possibility of a similar defect in Mmp7 –/– mice. Analysis of the total<br />

number of lung parenchymal inflammatory cells showed that the<br />

IL-4 (pg/ml)<br />

Figure 3 Attenuated T H2 cytokines and<br />

chemokines in Mmp7 –/– mice. (a–c) Luminex<br />

assay of IL-4 (a), IL-5 (b) and IL-13 (c) inBAL<br />

fluid from age- and sex-matched wild-type and<br />

Mmp7 –/– mice (n ¼ 5 mice per group)<br />

immunized intranasally with PBS (–) or CAA (+)<br />

every 4 d for a total of five doses and assessed<br />

24 h after the final immunization. (d) ELISA of<br />

CCL11 in BAL fluid of mice treated as described<br />

in a–c. (e) Real-time RT-PCR analysis of the<br />

expression of Il25 mRNA in the lungs of<br />

e<br />

IL-4 (pg/ml)<br />

BAL eosinophils<br />

(×10 5 /ml)<br />

IL-5 (pg/ml)<br />

lungs of CAA-immunized Mmp7 –/– mice had fewer eosinophils than<br />

did those of wild-type mice (Fig. 2e and Supplementary Fig. 3<br />

online), which indicated that the lower cell number in the airway<br />

was not due to less cell trafficking.<br />

We next assessed whether MMP7-mediated modification of IL-25<br />

was influential in determining allergic lung responses. We administered<br />

rIL-25 intranasally to wild-type and Mmp7 –/– mice and assessed<br />

eosinophil recruitment. In this assay, rIL-25 alone readily induced<br />

lung eosinophilia and new expression of MMP7 in airway epithelial<br />

cells of wild-type mice, but we found significantly fewer eosinophils<br />

and lower concentrations of IL-4 and IL-5 in the BAL fluid of<br />

Mmp7 –/– mice (Fig. 2f,g and Supplementary Fig. 4 online). Collectively,<br />

these findings support the idea that induction of MMP7 in<br />

response to allergens is critical for enhancing IL-25 function during<br />

initiation of the allergic immune response.<br />

Diminished TH2 responses of Mmp7 –/– mice<br />

We next determined if the diminished airway hyper-reactivity and<br />

BAL inflammatory cells in Mmp7 –/– mice were secondary to lower<br />

a b c d e<br />

WT<br />

Mmp7<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

*<br />

ND<br />

**<br />

ND<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

*<br />

**<br />

150<br />

100<br />

50<br />

0<br />

*<br />

**<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

*<br />

**<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

*<br />

**<br />

CAA – + – + CAA – + – + CAA – + – + CAA – + – + CAA – + +<br />

–/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

WT<br />

Mmp7 –/–<br />

unimmunized (–) and CAA-immunized (+) wild-type and Mmp7 –/– mice, presented relative to Actb expression. *, P o 0.05, versus PBS-challenged<br />

wild-type; **, P o 0.05, versus CAA-immunized wild-type (one-way ANOVA and t-test). Data are representative of three independent experiments<br />

(mean and s.d.).<br />

IL-5 (pg/ml)<br />

IL-13 (pg/ml)<br />

Figure 2 MmP7 –/– mice have an attenuated<br />

asthma phenotype. (a,b) Airway hyper-reactivity in<br />

age- and sex-matched wild-type mice (WT; n ¼ 4)<br />

and Mmp7 –/– mice (n ¼ 4) immunized<br />

intranasally with PBS (–) or CAA (+) every 4 d<br />

for a total of five doses and assessed 24 h after<br />

the final immunization as respiratory system<br />

resistance (RRS) in response to various doses<br />

of acetylcholine (Ach; a) or concentration of<br />

acetylcholine needed to elicit a 200% change<br />

from baseline respiratory system resistance<br />

(PC200; b). (c,d) ELISA of glycoproteins (c) and<br />

eosinophil cell counts (d) in BAL fluid from the<br />

mice in a,b. Ina–d, *,P o 0.05, versus PBSchallenged<br />

wild-type; **, P o 0.05, versus<br />

CAA-immunized wild-type (one-way ANOVA and<br />

t-test). Data are representative of at least three<br />

independent experiments (mean and s.d.).<br />

(e) Photomicrographs of bronchovascular bundles<br />

stained with hematoxylin and eosin (n ¼ 4 mice<br />

per group). Scale bar, 100 mm. Data are<br />

representative of at least three independent<br />

experiments. (f,g) Total cell and eosinophil counts<br />

(f) and ELISA of IL-4 and IL-5 (g) inBALfluid<br />

of wild-type and MmP7 –/– mice given PBS (–) or<br />

rIL-25 (+; 5 mg) intranasally twice daily for 3 d,<br />

assessed 24 h after the final dose. *, P o 0.05,<br />

versus rIL-25-treated wild-type mice (one-way<br />

ANOVA). Data are representative of three<br />

independent experiments (mean and s.d.;<br />

n ¼ 4 mice per group).<br />

498 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

CCL11 (ng/ml)<br />

IL-25 mRNA<br />

(relative expression)


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

c<br />

ATRA (AU)<br />

BAL total cells<br />

(×10 5 /ml)<br />

WT CAA Mmp7 Superimposed<br />

–/– CAA<br />

BAL eosinophils<br />

(×10 5 /ml)<br />

IL-25 mRNA<br />

(relative expression)<br />

T H2 responses in Mmp7 –/– mice. As expected, CAA-challenged<br />

Mmp7 –/– mice had less IL-4, IL-5, IL-13, IL-6 and tumor necrosis<br />

factor in BAL fluid than did wild-type mice (Fig. 3a–c and Supplementary<br />

Fig. 5 online), whereas we found no significant differences in<br />

the concentrations of CCL3, IL-9 and IL-1b (data not shown).<br />

Consistent with the diminished lung eosinophilia, we found significantly<br />

lower CCL11 expression in CAA-immunized Mmp7 –/– mice<br />

than in CAA-immunized wild-type mice, but we detected no significant<br />

differences the concentrations of CCL7 and CCL17 (Fig. 3d,<br />

and data not shown). Furthermore, whole-lung analysis of RNA<br />

showed that immunized Mmp7 –/– mice had lower expression of Il25<br />

than did wild-type mice (Fig. 3e), which suggested that in addition to<br />

proteolytic modification of IL-25, activation and function of MMP7 in<br />

the lung is required for the transcriptional expression of Il25 in this<br />

model of allergic airway disease.<br />

More retinoic acid in allergen-challenged Mmp7 –/– mice<br />

To identify additional events that might regulate the airway epithelial<br />

response to allergen, we did proteomic analyses of BAL fluid from<br />

ATRA (AU)<br />

CCL11 mRNA<br />

(relative expression)<br />

RALDH-1 abundance (log)<br />

ATRA (AU)<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

–0.1<br />

–0.2<br />

–0.3<br />

–0.4<br />

–0.5<br />

IL-4 (pg/ml)<br />

WT PBS<br />

WT CAA<br />

Mmp7 –/– CAA<br />

Mmp7 –/– WT PBS<br />

WT CAA<br />

CAA<br />

–0.00000<br />

–0.00002<br />

–0.00004<br />

–0.00006<br />

–0.00008<br />

–0.00010<br />

–0.00012<br />

–0.00000<br />

–0.00002<br />

–0.00004<br />

–0.00006<br />

–0.00008<br />

–0.00010<br />

–0.00012<br />

8.081<br />

–0.00000<br />

–0.00002<br />

–0.00004<br />

–0.00006<br />

–0.00008<br />

–0.00010<br />

–0.00012<br />

8.053<br />

–0.00014<br />

–0.00014<br />

–0.00014<br />

–0.00016<br />

–0.00016<br />

–0.00016<br />

–0.00018<br />

–0.00018<br />

–0.00018<br />

–0.00020<br />

–0.00020<br />

–0.00020<br />

0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14 0 2 4 6 8 10 12 14<br />

Time (min) Time (min) Time (min)<br />

d 60<br />

40<br />

20<br />

*<br />

**<br />

e 50<br />

40<br />

30<br />

20<br />

10<br />

*<br />

**<br />

f<br />

25<br />

20<br />

15<br />

10<br />

5<br />

*<br />

**<br />

g 4<br />

3<br />

2<br />

1<br />

*<br />

**<br />

h 30<br />

20<br />

10<br />

*<br />

**<br />

i 0.4<br />

0.3<br />

0.2<br />

0.1<br />

*<br />

**<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0.0<br />

CAA – – + + CAA – – + + CAA – – + + CAA – – + + CAA – – + + CAA – – + +<br />

ATRA – + – + ATRA – + – + ATRA – + – + ATRA – + – + ATRA – + – + ATRA – + – +<br />

Figure 4 Proteomics analysis of BAL fluid from CAA-immunized wild-type and Mmp7 –/– mice. (a) Two-dimensional gel electrophoresis assessing the<br />

processing of proteins in pooled BAL fluid from CAA-immunized wild-type mice (green; indocarbocyanine) and Mmp7 –/– mice (red; indodicarbocyanine;<br />

Supplementary Table 2). Superimposed images (right) show differences in proteins: yellow, no change; red spots, more abundant in Mmp7 –/– ; green spots,<br />

more abundant in wild-type; inset, enlargement of the area with RALDH-1. Arrows indicate RALDH-1 protein. Data are representative of two independent<br />

experiments (n ¼ 3 mice per group). (b) RALDH-1 abundance in BAL fluid from wild-type and Mmp7 –/– mice immunized with PBS or CAA, assessed as spot<br />

volumes standardized and log-transformed with DeCyder software, Biological Variation Analysis. Each point represents the spot volume of a pooled sample<br />

(n ¼ 3 mice). *, P o 0.05, versus saline-challenged wild-type; **, P o 0.05, versus CAA-immunized wild-type mice (t-test). Data are representative of at<br />

least three independent experiments (mean ± s.d.). (c) HPLC of ATRA in BAL fluid from wild-type and Mmp7 –/– mice (n ¼ 3) immunized with PBS or CAA,<br />

presented as absorbance units (AU). Numbers in graphs indicate time of ATRA peak; area under the curve for each peak is 344 (WT CAA) and 755<br />

(Mmp7 –/– CAA). (d,e) Total cells (d) and eosinophils (e) in BAL fluid from wild-type mice treated with liposomal ATRA (5 mg per g body weight) 1 h before<br />

intranasal immunization with CAA. Negative control, liposomal ATRA and PBS (n ¼ 5); positive control, CAA alone (n ¼ 5). (f,g) Real-time RT-PCR analysis<br />

of the expression of Il25 mRNA (f) andCcl11 mRNA (g) inthelungsofthemiceind,e, presented relative to the expression of Actb (f) or 18S RNA (g).<br />

(h,i) Luminex assay of IL-4 (h) and IL-13 (i) in BAL fluid. *, P o 0.05, versus PBS-challenged wild-type; **, P o 0.05, versus CAA-immunized wild-type<br />

(one-way ANOVA). Data are representative of three independent experiments (mean and s.d.).<br />

CAA-immunized wild-type and Mmp7 –/– mice 17 . Among the proteins<br />

that were significantly and differently modulated in CAA-immunized<br />

Mmp7 –/– mice (P o 0.05; Supplementary Table 2 online), we found<br />

that they had more RALDH-1 (Fig. 4a,b and Supplementary Fig. 6<br />

online). RALDH-1 is the rate-limiting enzyme for the conversion of<br />

retinaldehyde to all-trans retinoic acid (ATRA), which is critical for<br />

maintenance of the tolerogenic environment of mucosal surfaces 24 .In<br />

support of our proteomic-based finding of more RALDH-1, analysis<br />

of BAL fluid showed that CAA-immunized Mmp7 –/– mice had a<br />

higher ATRA concentration than did CAA-immunized wild-type<br />

mice (Fig. 4c).<br />

Next, we determined if more production of ATRA in the lungs of<br />

Mmp7 –/– mice might account in part for their attenuated response to<br />

allergen. We delivered liposomal ATRA by means of respiratory<br />

aerosol to wild-type mice 1 h before each immunization with CAA,<br />

then assessed lung allergic responses. Mice that received ATRA had an<br />

attenuated asthma phenotype, as demonstrated by fewer total cells in<br />

BAL fluid, especially eosinophils, and lower expression of Il25 mRNA<br />

than that of mice given aerosol vehicle (empty liposome; Fig. 4d–f).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 499<br />

b<br />

*<br />

IL-13 (ng/ml)<br />

ARTICLES<br />

**


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

WT<br />

10 4<br />

Mmp7 –/–<br />

WT Mmp7 –/–<br />

a b c<br />

PBS<br />

Consistent with lower expression of inflammatory markers, we also<br />

detected less CCL11, IL-4 and IL-13, which act ‘downstream’ of IL-25<br />

signaling (Fig. 4g–i and Supplementary Fig. 7 online). These data<br />

confirm that ATRA has potent anti-inflammatory activity when<br />

administered to the lung and that this activity can be enhanced<br />

through inhibition of Il25 expression.<br />

Epithelial RALDH-1 and ATRA induce regulatory T cells<br />

We found expression of RALDH-1 protein in airway epithelial cells<br />

and alveolar macrophages of CAA-immunized wild-type and Mmp7 –/–<br />

mice (Fig. 5a). Higher expression of RALDH-1 and concentrations of<br />

ATRA have been linked to enhanced immune tolerance and more<br />

production of inducible regulatory T cells 25 . Therefore, we determined<br />

if higher expression of RALDH-1 acted as a negative regulator in<br />

response to allergens and increased the abundance of regulatory T cells<br />

in the lungs of wild-type and Mmp7 –/– mice after immunization with<br />

CAA. CAA-immunized wild-type and Mmp7 –/– mice had more<br />

PBS<br />

Foxp3<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

0.14 0.31<br />

BAL total cells<br />

(×10 5 /ml)<br />

BAL eosinophils<br />

(×10 5 /ml)<br />

CD25 + Foxp3 + regulatory T cells in lungs than did naive mice (Fig. 5b<br />

and Supplementary Fig. 8 online). To further examine the function<br />

of RALDH-1 in inducing the production of regulatory T cells, we<br />

administered citral, an inhibitor of RALDH-1, to mice immunized<br />

with CAA 26 . Inhibition of the synthesis of retinoic acid by citral<br />

resulted in fewer CD25 + Foxp3 + regulatory T cells in the lung (Fig. 5b<br />

and Supplementary Fig. 9 online), which suggested possible involvement<br />

of RALDH-1 and ATRA in the development of this inhibitory<br />

subset of lung T cells. Moreover, wild-type and Mmp7 –/– mice<br />

immunized with CAA and challenged with citral had more eosinophils,<br />

IL-5 and CCL11 in BAL fluid than did mice immunized with<br />

CAA alone (Fig. 5c–e and data not shown).<br />

MMP7 expression in response to allergen in human asthma<br />

To address the relevance of MMP7 expression to human asthma, we<br />

studied bronchial biopsy specimens from human volunteers with and<br />

without asthma. Notably, whereas we detected IL-25 in airway<br />

d<br />

IL-5 (pg/ml)<br />

12.5<br />

10.0<br />

7.5<br />

5.0<br />

2.5<br />

0.0<br />

CAA –<br />

Citral +<br />

*<br />

WT<br />

Mmp7 –/–<br />

**<br />

+ + + +<br />

– + – +<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

10 1<br />

10 1<br />

1.26<br />

3.01<br />

0.45<br />

0.7<br />

10<br />

CD25<br />

0<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

CAA<br />

CAA<br />

10<br />

8<br />

6<br />

4<br />

WT<br />

Mmp7<br />

2<br />

0<br />

CAA – + + + +<br />

Citral + – + – +<br />

CAA<br />

+<br />

citral<br />

200<br />

150<br />

100<br />

50<br />

0<br />

CAA<br />

Citral<br />

–<br />

+<br />

+<br />

–<br />

+<br />

+<br />

+<br />

–<br />

+<br />

+<br />

–/–<br />

WT<br />

Mmp7 –/–<br />

*<br />

**<br />

Figure 5 Epithelial expression of RALDH-1 in response to<br />

allergens initiates a negative regulatory response.<br />

(a) Immunohistochemistry of RALDH-1 (arrowheads)<br />

in lung sections of wild-type and Mmp7<br />

e<br />

*<br />

**<br />

/ mice immunized<br />

intranasally with PBS or CAA. Scale bars, 100 mm. Data are<br />

representative of two independent experiments. (b) Flow<br />

cytometry of regulatory T cells obtained from the lungs of<br />

wild-type and Mmp7 / mice immunized intranasally with<br />

PBS or CAA (same dose for all), with (bottom row) or without<br />

(top and middle rows) the administration of citral 1 h before CAA immunization, stained with anti-CD4, anti-CD25 and anti-Foxp3. Numbers in outlined<br />

areas indicate percent Foxp3 + CD25 + cells. Data are representative of two independent experiments (n ¼ 4 mice per group). (c–e) Total cells (c), eosinophils<br />

(d) and IL-5 (e) in BAL fluid from mice treated as described in b. Intranasal administration of citral or CAA alone serves as a negative or positive control,<br />

respectively. *, P o 0.05, versus CAA-challenged wild-type; **, P o 0.05, versus CAA-challenged Mmp7 / independent experiments (mean and s.d.; n ¼ 4 mice per group).<br />

mice (t-test). Data are representative of three<br />

a Nonasthmatic, MMP7<br />

Nonasthmatic, IL-25 b<br />

Asthmatic, MMP7<br />

Asthmatic, IL-25<br />

MMP7 (pg/ml)<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

RW (30 min) RW (5 h)<br />

*<br />

Figure 6 Expression of MMP7 and IL-25 in<br />

humans with asthma. (a) Serial photomicrographs<br />

of bronchial biopsies from a nonasthmatic subject<br />

(top row) and an asthmatic patient (bottom row),<br />

stained with anti-MMP7 (left) or anti-IL-25 (right).<br />

Arrows indicate positive immunoreactivity; arrowhead<br />

indicates lack of MMP7 expression. Data are<br />

representative of at least three independent experiments<br />

(n ¼ 5 subjects per group). (b) ELISA<br />

of MMP7 responses in nasal washings from seven<br />

atopic-allergic volunteers (n ¼ 7) challenged<br />

intranasally with increasing doses of ragweed<br />

(RW) or saline (data not shown). *, P o 0.05,<br />

versus 30-minute ragweed challenge (two-tailed<br />

paired t-test). Data are representative of at least<br />

three independent experiments.<br />

500 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

epithelia of subjects with and without asthma and thus it seemed to be<br />

constitutively present, we found MMP7 expression exclusively in the<br />

airways of asthmatic subjects (Fig. 6a). Consistent with those findings,<br />

analysis of nasal washings collected from volunteers with allergic<br />

rhinitis 30 min and 5 h after intranasal challenge with ragweed (a<br />

potent proteinase allergen) showed significantly more MMP7 secretion<br />

in response to ragweed (Fig. 6b) but not in response to the<br />

normal saline control (data not shown). We did not detect secretion of<br />

IL-25 in nasal washes. Nonetheless, these data suggest that constitutive<br />

IL-25 was activated by inducible expression of MMP7 during allergen<br />

challenge of the human airway.<br />

DISCUSSION<br />

We have shown here how proteinase allergens simultaneously activated<br />

pro- and anti-inflammatory programs in airway epithelial cells.<br />

Allergens enhanced MMP7 expression in airway epithelial cells to<br />

maximize allergic lung inflammation; conversely, in the absence of<br />

MMP7, mice developed an attenuated asthmatic phenotype. The<br />

mechanisms responsible for the diminished allergic inflammation in<br />

Mmp7 –/– mice seemed to be in part related to lower expression of<br />

IL-25, which has been shown to be critical in initiating robust T H2<br />

responses 5,7,27,28 . We found that in addition to modulating Il25<br />

expression, MMP7 mediated IL-25 cleavage that was required for<br />

most of the biological activity of IL-25 in allergic inflammation.<br />

Enhanced expression of MMP family members is a frequent<br />

manifestation of active or chronic inflammatory process, but the<br />

precise function of MMPs in these conditions remains poorly understood<br />

13 . Studies have documented that proteinase allergens are able to<br />

induce an innate immune response that further modulates the adaptive<br />

immune response during allergic inflammation 1,2 .Wehaveshownhere<br />

that airway epithelial expression of MMP7 was critical for the development<br />

of asthma-like disease, a finding in contrast to the functions of<br />

MMP2 and MMP9, two gelatinases that are also upregulated in the<br />

same model, as absence of these enzymes leads to exaggerated lung<br />

allergic inflammation 15–17 . Furthermore, we have shown that MMP7<br />

was expressed in the distal lung parenchyma of untreated asthmatics<br />

and that in response to a proteinase allergen derived from ragweed,<br />

MMP7 expression was significantly induced. Although the secretion of<br />

IL-25 in the same conditions was below the detection limit of our<br />

enzyme-linked immunosorbent assay (ELISA), we found IL-25 expression<br />

in the distal lung space in normal and asthmatic volunteers.<br />

Collectively, these findings demonstrate a temporal relation between<br />

exposure to allergens and induction of MMP7 in human allergic<br />

disease and may indicate involvement of activation of IL-25 by<br />

MMP7 in conditions of allergen exposure. It is important to note<br />

that during allergic inflammation, many endogenous proteinases,<br />

including members of the serine and cysteine family, are abundantly<br />

present in the airway 1,29 . However, our finding that MMP7 was<br />

required for the amplification of T H2 responses emphasizes the<br />

nonredundant functions that MMPs may serve in the course of<br />

inflammation. The specificity of this response was exemplified by the<br />

specific proteolysis of IL-25, but not of other allergy-related cytokines<br />

such as TSLP and IL-13, by MMP7. Similarly, MMP7-mediated<br />

proteolytic modification of defensins, apoptotic ligands and cytokines<br />

has been shown to be important in diverse biological functions, such as<br />

innate immunity in the gut to cancer cell metatasis 19,30–32 . The findings<br />

that rIL-25 induced expression of lung MMP7 and intranasal administration<br />

of rIL-25 to Mmp7 –/– mice only weakly induced allergic<br />

inflammation suggest that the bioactivity of IL-25 depends mainly<br />

on proteolysis by MMP7. These data further suggest the existence of a<br />

feedback loop in which IL-25 upregulates lung MMP7 expression,<br />

ARTICLES<br />

followed by cleavage and activation of IL-25 that further enhances both<br />

MMP7 expression and allergic lung disease.<br />

Although regulation of gene expression has been studied for other<br />

members of the IL-17 family, little is known about IL-25 regulation 12 .<br />

We found that although there was little or no RALDH-1 in the airways<br />

of naive mice, this enzyme had prominent expression in the airway<br />

epithelial cells of both wild-type and Mmp7 –/– mice in response to a<br />

proteinase allergen. By high-throughput proteomic analysis, we found<br />

excess RALDH-1 enzyme in CAA-immunized Mmp7 –/– mice that was<br />

functionally relevant because it resulted in higher BAL fluid concentrations<br />

of ATRA and less allergic inflammation. The function of ATRA<br />

in the regulation of adaptive immunity, especially TH2 responses, is<br />

controversial. For example, IL-4-induced eotaxin production, eosinophil<br />

lineage commitment and IL-5 receptor expression are all inhibited<br />

by the addition of ATRA in vitro 33,34 .Furthermore,retinoicacidinthe<br />

presence of transforming growth factor-b1 inhibits the binding of<br />

STAT6 to the promoter of the gene encoding the transcription factor<br />

Foxp3 and effectively inhibits IL-4 signaling in T cells 35 . In other<br />

models of allergic lung disease, deficiency in vitamin A results in lower<br />

expression of the muscarinic M2 receptor, which results in more airway<br />

hyper-reactivity 36 . In contrast, intranasal administration of high concentrations<br />

of ATRA (over 1,500 mg) has been shown to fail to alter<br />

T H2 responses in another asthma model 37 , and vitamin A deficiency<br />

has been shown to attenuate the production of T H2 cytokines in<br />

mice 38 . In our study, allergen-challenged mice that received 100 mg<br />

ATRA showed much lower allergic inflammation and attenuation of<br />

the asthmatic phenotype. Such divergent findings are not unexpected,<br />

given the known idiosyncrasies of retinoid, which can act as a hormone<br />

and, depending on the physiological or pharmaceutical context, can<br />

mediate diverse effects on cells of the immune system 39 .<br />

Studies have indicated involvement of ATRA in the development<br />

and differentiation of inducible regulatory T cells, an antiinflammatory<br />

subset of T cells that are important in the homeostasis<br />

of lymphocytes in various organs 40 . In particular, dendritic cells in the<br />

lamina propria have been found to provide ATRA that is critical for<br />

the differentiation of regulatory T cells and homing of these cells to<br />

the intestinal mucosa 41,42 . Although we did not identify the cellular<br />

origin of ATRA, we found that RALDH-1, the rate-limiting enzyme<br />

that converts retinaldehyde to ATRA, was upregulated in response to<br />

allergic stimulation in the airway epithelium, which indicates a<br />

possible site for ATRA production in the lung. Consistent with the<br />

requirement for ATRA in the differentiation of regulatory T cells<br />

in vivo 25,41,43 , we found more CD25 + Foxp3 + T cells in the lungs<br />

of Mmp7 –/– mice immunized with CAA than in the lungs of<br />

CAA-immunized wild-type mice. Moreover, we found that inhibition<br />

of the synthesis of retinoic acid by citral inhibited the development of<br />

regulatory T cells. The higher percentage of regulatory T cells after<br />

allergen challenge in both wild-type and Mmp7 –/– mice suggested that<br />

RALDH-1 is a negative regulator of allergic inflammation, a plausible<br />

protective response regulated by airway epithelial cells that we found<br />

to be enhanced in the absence of MMP7. What remains unclear is the<br />

mechanism responsible for the higher expression and function of<br />

RALDH-1 in the lung in the absence of MMP7. Our finding of<br />

RALDH-1 expression in airway epithelial cells and macrophages in<br />

response to allergen challenge is in agreement with published reports<br />

indicating an anti-inflammatory function for these cells in allergic<br />

inflammation 44 . Furthermore, depletion of alveolar macrophages<br />

results in an exaggerated allergic response that can be inhibited by<br />

adoptive transfer of alveolar macrophages in brown Norway rats 45 .<br />

Our data suggest that alveolar macrophages that mediate antiinflammatory<br />

responses during allergic inflammation might be acting<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 501


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

through RALDH-1, an idea that is also consistent with the tolerogenic<br />

function of RALDH-1 in gut lamina propria 41,42 .<br />

In summary, we have described here a proinflammatory function<br />

of MMP7 that was activated in response to proteinase<br />

allergens and involved proteolytic modification of the expression<br />

and function of IL-25. In this model, we found activation of<br />

RALDH-1 was ‘downstream’ of allergen-induced inflammation<br />

and that MMP7 actively inhibited the function of this enzyme<br />

in vivo. Thisuniquemechanisminvolved,atleastinpart,inhibition<br />

of regulatory T cells that inhibit lung inflammation. Our<br />

findings thus substantially expand the understanding of the regulation<br />

of IL-25, a critical innate cytokine in allergic responses,<br />

and further identify pro- and anti-inflammatory targets that can<br />

be explored for therapeutic benefit.<br />

METHODS<br />

Mice. Mmp7 –/– mice backcrossed nine generations with C57BL/6 mice were<br />

provided by P.W. Park and were bred in the transgenic animal facility at Baylor<br />

College of Medicine (accredited by the Association for Assessment and<br />

Accreditation of Laboratory Animal Care). All experiments were in accordance<br />

with protocols approved by the Institutional Animal Care and Use Committee<br />

of Baylor College of Medicine. Information on the genotyping of Mmp7 –/– mice<br />

and description of control mice are in the Supplementary Methods online.<br />

Stat6 –/– mice (BALB/c background) and Il13 –/– mice (C57BL/6 background)<br />

from the Jackson Laboratories were bred for ten to twelve generations onto the<br />

BALB/c background.<br />

Experimental model of asthma. CAA, composed of Aspergillus oryzae proteinase<br />

(1 mg/ml; Sigma) and ovalbumin (0.5 mg/ml; Sigma) in a volume of 50 ml,<br />

was administered intranasally every 4 d for a total of five doses as described<br />

before 1,16 . In some experiments, rIL-25 (5 mg in50ml PBS)wasadministered<br />

intranasally twice daily for 3 d. In some experiments, liposomal ATRA or citral<br />

was given by aerosol administration 1 h before CAA challenge. Additional<br />

information on the liposome formulation and administration of ATRA is in the<br />

Supplementary Methods.<br />

Experimental ragweed allergen challenge in human volunteers. Nasal ‘provocation’<br />

with ragweed pollen extract was done as described 46 . The Institutional<br />

Review Board at the University of Texas Medical Branch reviewed and<br />

approved the protocol and consent forms. The challenge procedure and<br />

analysis of nasal lavage fluid and bronchial biopsies are described in the<br />

Supplementary Methods.<br />

Analysis of asthmatic phenotype. All data were collected 24 h after the final<br />

allergen challenge. Airway hyper-reactivity was assessed by estimation of the<br />

concentration of acetylcholine (in mg per g body weight) that caused a 200%<br />

increase in airway resistance over the baseline, calculated by linear interpolation<br />

of the appropriate dose-response curves as described 15 . Collection and analysis<br />

of BAL fluid and lung tissue is described in the Supplementary Methods.<br />

Immunohistochemistry. A standard published protocol was used for immunohistochemistry<br />

47 . Deparaffinized sections were washed with 3% (vol/vol)<br />

H2O2 and 2% (vol/vol) Triton X-100 in PBS, followed by antigen retrieval<br />

(Target Retrieval Solution; Dako). Slides were incubated overnight at 4 1C with<br />

rabbit polyclonal antibody to mouse ALDH1A1 (24343; Abcam), rat monoclonal<br />

antibody to mouse MMP7 (377314; R&D Systems), mouse monoclonal<br />

antibody to human MMP7 (377314 (R&D Systems) or ID2 (Calbiochem)) and<br />

mouse monoclonal antibody to human IL-17E. Additional information on<br />

staining and analysis is in the Supplementary Methods.<br />

Measurement of cytokines and chemokines in BAL fluid. CCL11, CCL7 and<br />

CCL17 in BAL fluid were measured by standard antibody-based ELISA and/or a<br />

multiplex bead-based cytokine detection kit (BioRad) as described 16 .Reagents<br />

and the LINCOplex assay are described in the Supplementary Methods.<br />

Proteomics analysis of BAL protein. Proteomics analysis of BAL fluid was<br />

done as described 17 . BAL samples from wild-type and Mmp7 –/– mice were<br />

pooled and were concentrated with AMICON filters, and protein concentrations<br />

were measured by BCA protein assay (bicinchoninic acid; Pierce). An<br />

equal amount of protein from each pooled sample (50 mg) was labeled with<br />

400 pmol fluorescent dye (CyDye; Amersham Biosciences) matched by<br />

charge and molecular size and was brought to a final volume of 250 ml with<br />

sample buffer that also contained 0.5% (wt/vol) ampholytes (pH 3–11, nonlinear),<br />

0.1% (wt/vol) bromophenol blue and DeStreak reagent (12 ml/ml;<br />

GE Healthcare). Additional information on proteomics analysis is in the<br />

Supplementary Methods.<br />

Quantitative RT-PCR. The RNeasy mini kit (Qiagen) was used for extraction of<br />

total cellular RNA from mouse lung tissues stabilized with the RNA Later<br />

reagent (Qiagen). One-step real-time quantitative RT-PCR (Real-Time PCR<br />

system 7300; Applied Biosystems) was used to assess the expression of Ccl11<br />

(encoding CCL11 (eotaxin 1)) and Aldh1a1 (encoding RALDH-1). The primer<br />

and probe mixtures of target genes (Ccl11, Mm00441238_m1; Raldh-1,<br />

Mm01194995_mH; Applied Biosystems identifiers) and 18S rRNA were from<br />

Applied Biosystems. Additional information on RT-PCR is in the Supplementary<br />

Methods.<br />

MMP activation and in vitro cleavage assay. Carrier-free mouse rIL-25 (5 mg;<br />

R&D), mouse rIL-13 (Peprotech) and mouse rTSLP (R&D Systems) were<br />

incubated for 2 h at 37 1C with0.5mg 4-aminophenylmercuric acetate–activated<br />

MMP7 in the presence or absence of the MMP inhibitor 1,10-phenanthroline<br />

(10 mM; Sigma). After incubation, equal volumes of each sample (16.5 ml) were<br />

reduced and were resolved by separation through a 16.5% tricine gel. Cleaved<br />

and uncleaved proteins were visualized with the Proteosilver Plus Silver Stain kit<br />

according to the manufacturer’s instructions (Sigma-Aldrich). Additional information<br />

on these assays is in the Supplementary Methods. Amino-terminal<br />

sequencing of IL-25 was done by standard protein-sequencing methods (Procise<br />

492cLC; Applied Biosystems).<br />

Cell culture. Lymph node cells and splenocytes isolated from C57BL/6 mice<br />

were stimulated with plate-bound antibody to CD3 (anti-CD3; 2 mg/ml) and<br />

human IL-2 (50 U/ml) in the presence of mouse rIL-25 (250 ng/ml; R&D<br />

Systems) or rIL-25’C (250 ng/ml). After 2 d, culture supernatants were collected<br />

and analyzed for IL-4, IL-5, IL-13 and interferon-g by ELISA as described 5 .On<br />

day 3, cells were restimulated for 5 h with ionomycin (500 ng/ml) and phorbol<br />

12-myristate 13-acetate (50 ng/ml) in the presence of GolgiStop (BD Biosciences).<br />

Cells were made permeable with a Cytofix/Cytoperm kit (BD<br />

Biosciences) and were analyzed for intracellular expression of IL-4 (with<br />

phycoerythrin-conjugated anti-IL-4; 11B11) and IL-5 (with allophycocyaninconjugated<br />

anti-IL-5; TRFK5; both from BD Biosciences).<br />

Flow cytometry. Total lung eosinophils were quantified by bead-enhanced flow<br />

cytometry as described 48 . Total mouse lung cells were stained with fluorescein<br />

isothiocyanate–conjugated conjugated antibody to major histocompatibility<br />

class II (2G9; 553623; BD Biosciences) and phycoerythrin-conjugated anti-<br />

Siglec-F (1 mg per10 6 cells; E50-2440; BD Biosciences), followed by incubation<br />

with fluorescent beads (Flow-Check Fluorospheres; Beckman Coulter). Samples<br />

were acquired with a flow cytometer (XL2; Beckman Coulter) and cell and bead<br />

counts were measured. Absolute numbers of eosinophils were calculated with<br />

formulas as described 48 . Additional information on flow cytometry is in the<br />

Supplementary Methods.<br />

High-performance liquid chromatography (HPLC). Pooled BAL samples<br />

from wild-type and Mmp7 –/– mice were evaporated under nitrogen and the<br />

dry residue was dissolved in 100 ml acetonitrile (Fisher). Aliquots of 25 ml were<br />

then analyzed with the Waters HPLC system, consisting of a WISP autosampler<br />

(model 717 Plus), dual-absorbance detector (model 2487), pump (model 515)<br />

and Nova-Pak C18 columns (3.9 mm 150 mm). ATRA was separated with a<br />

solvent system of 57.5% (vol/vol) acetonitrile, 25% (vol/vol) acetic acid (2%<br />

solution) at a flow rate of 1.3 ml/min. Data were analyzed with Waters<br />

Millennium software (version 3.2).<br />

Binding assay. Analysis of the binding of IL-25 to the fusion protein of IL-17<br />

receptor B and Fc fragment is described in the Supplementary Methods.<br />

502 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Statistics. All data were analyzed with Prism 4.0a statistical analysis software.<br />

Significant differences (P o 0.05) between treatment groups were identified<br />

with a t-test, one-way analysis of variance (ANOVA) or Bonferroni multiplecomparison<br />

test.<br />

Accession code. UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signalinggateway.org):<br />

A001477.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank P. Woo Park (Children’s Hospital, Harvard School of Medicine) for<br />

Mmp7 –/– mice backcrossed to C57BL/6 mice; D.A. Engler, R.K. Matsunami and<br />

K. Gonzalez for technical help with proteomics analysis; N. Barrows for editorial<br />

support; and all members of the Kheradmand and Corry laboratory for<br />

comments and criticisms. Supported by the US National Institutes of Health<br />

(AI070973 and HL082487 to F.K.; AI071130 and AR050772 to C.D.; and<br />

HL075243, AI057696 and AI070973 to D.B.C.), the American Lung Association<br />

(C.D.), the Leukemia and Lymphoma Society (C.D.) and MD Anderson Cancer<br />

Center (C.D.).<br />

AUTHOR CONTRIBUTIONS<br />

S.G. did animal experiments, in vitro cleavage assay, immunohistochemistry,<br />

flow cytometry, RT-PCR, ELISA and Luminex assays; P.A. did in vitro T cell<br />

experiments; W.T.B. and S.P. did airway physiology experiments; K.J.G. did<br />

proteomics analysis; A.S. made liposomal ATRA and helped with HPLC; M.S.,<br />

L.S., D.R., B.S. and S.S. did the ragweed challenges of humans and ELISA of<br />

samples from allergic volunteers; P.W. obtained bronchial biopsies of asthmatic<br />

and control volunteers; S.G., F.K., D.B.C. and C.D. designed experiments;<br />

S.G. and F.K. wrote the manuscript; and P.A., D.B.C. and C.D. critically<br />

reviewed the manuscript.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Kheradmand, F. et al. A protease-activated pathway underlying Th cell type 2 activation<br />

and allergic lung disease. J. Immunol. 169, 5904–5911 (2002).<br />

2. Kiss, A. et al. A new mechanism regulating the initiation of allergic airway inflammation.<br />

J. Allergy Clin. Immunol. 120, 334–342 (2007).<br />

3. Li, L. et al. Effects of Th2 cytokines on chemokine expression in the lung: IL-13<br />

potently induces eotaxin expression by airway epithelial cells. J. Immunol. 162,<br />

2477–2487 (1999).<br />

4. Zhu, J. et al. Exacerbations of bronchitis: bronchial eosinophilia and gene expression for<br />

interleukin-4, interleukin-5, and eosinophil chemoattractants. Am. J. Respir. Crit. Care<br />

Med. 164, 109–116 (2001).<br />

5. Angkasekwinai, P. et al. Interleukin 25 promotes the initiation of proallergic type 2<br />

responses. J. Exp. Med. 204, 1509–1517 (2007).<br />

6. Moseley, T.A., Haudenschild, D.R., Rose, L. & Reddi, A.H. Interleukin-17 family and<br />

IL-17 receptors. Cytokine Growth Factor Rev. 14, 155–174 (2003).<br />

7. Pan, G. et al. Forced expression of murine IL-17E induces growth retardation, jaundice,<br />

a Th2-biased response, and multiorgan inflammation in mice. J. Immunol. 167,<br />

6559–6567 (2001).<br />

8. Kim, M.R. et al. Transgenic overexpression of human IL-17E results in eosinophilia, Blymphocyte<br />

hyperplasia, and altered antibody production. Blood 100, 2330–2340<br />

(2002).<br />

9. Ballantyne, S.J. et al. Blocking IL-25 prevents airway hyperresponsiveness in allergic<br />

asthma. J. Allergy Clin. Immunol. 120, 1324–1331 (2007).<br />

10. Fallon, P.G. et al. Identification of an interleukin (IL)-25-dependent cell population that<br />

provides IL-4, IL-5, and IL-13 at the onset of helminth expulsion. J. Exp. Med. 203,<br />

1105–1116 (2006).<br />

11. Greenlee, K.J., Werb, Z. & Kheradmand, F. Matrix metalloproteinases in lung: multiple,<br />

multifarious, and multifaceted. Physiol. Rev. 87, 69–98 (2007).<br />

12. Dong, C. TH17 cells in development: an updated view of their molecular identity and<br />

genetic programming. Nat. Rev. Immunol. 8, 337–348 (2008).<br />

13. Parks, W.C., Wilson, C.L. & Lopez-Boado, Y.S. Matrix metalloproteinases as<br />

modulators of inflammation and innate immunity. Nat. Rev. Immunol. 4, 617–629<br />

(2004).<br />

14. McMillan, S.J. et al. Matrix metalloproteinase-9 deficiency results in enhanced<br />

allergen-induced airway inflammation. J. Immunol. 172, 2586–2594 (2004).<br />

15. Corry, D.B. et al. Overlapping and independent contributions of MMP2 and MMP9 to<br />

lung allergic inflammatory cell egression through decreased CC chemokines. FASEB J.<br />

18, 995–997 (2004).<br />

16. Corry, D.B. et al. Decreased allergic lung inflammatory cell egression and increased<br />

susceptibility to asphyxiation in MMP2-deficiency. Nat. Immunol. 3, 347–353<br />

(2002).<br />

ARTICLES<br />

17. Greenlee, K.J. et al. Proteomic identification of in vivo substrates for matrix metalloproteinases<br />

2 and 9 reveals a mechanism for resolution of inflammation. J. Immunol.<br />

177, 7312–7321 (2006).<br />

18. Li, Q., Park, P.W., Wilson, C.L. & Parks, W.C. Matrilysin shedding of syndecan-1<br />

regulates chemokine mobilization and transepithelial efflux of neutrophils in acute lung<br />

injury. Cell 111, 635–646 (2002).<br />

19. Wilson, C.L. et al. Regulation of intestinal a-defensin activation by the metalloproteinase<br />

matrilysin in innate host defense. Science 286, 113–117 (1999).<br />

20. Gearing, A.J. et al. Matrix metalloproteinases and processing of pro-TNF-a. J. Leukoc.<br />

Biol. 57, 774–777 (1995).<br />

21. Corry, D.B. IL-13 in allergy: home at last. Curr. Opin. Immunol. 11, 610–614 (1999).<br />

22. Corry, D.B. et al. Requirements for allergen- induced airway hyperreactivity in T and B<br />

cell-deficient mice. Mol. Med. 4, 344–355 (1998).<br />

23. Grunig, G. et al. Requirement for IL-13 independently of IL-4 in experimental asthma.<br />

Science 282, 2261–2263 (1998).<br />

24. Denning, T.L., Wang, Y.C., Patel, S.R., Williams, I.R. & Pulendran, B. Lamina propria<br />

macrophages and dendritic cells differentially induce regulatory and interleukin 17producing<br />

T cell responses. Nat. Immunol. 8, 1086–1094 (2007).<br />

25. Mucida, D. et al. Reciprocal Th-17 and regulatory T cell differentiation mediated by<br />

retinoic acid. Science 317, 256–260 (2007).<br />

26. Connor, M.J. & Smit, M.H. Terminal-group oxidation of retinol by mouse epidermis.<br />

Inhibition in vitro and in vivo. Biochem. J. 244, 489–492 (1987).<br />

27. Cheung, P.F., Wong, C.K., Ip, W.K. & Lam, C.W. IL-25 regulates the expression of<br />

adhesion molecules on eosinophils: mechanism of eosinophilia in allergic inflammation.<br />

Allergy 61, 878–885 (2006).<br />

28. Fort, M.M. et al. IL-25 induces IL-4, IL-5, and IL-13 and Th2-associated pathologies<br />

in vivo. Immunity 15, 985–995 (2001).<br />

29. Zheng, T. et al. Inducible targeting of IL-13 to the adult lung causes matrix<br />

metalloproteinase- and cathepsin-dependent emphysema. J. Clin. Invest. 106,<br />

1081–1093 (2000).<br />

30. Swee, M., Wilson, C.L., Wang, Y., McGuire, J.K. & Parks, W.C. Matrix metalloproteinase-7<br />

(matrilysin) controls neutrophil egress by generating chemokine gradients.<br />

J. Leukoc. Biol. 83, 1404–1412 (2008).<br />

31. Liu, D. et al. Overexpression of matrix metalloproteinase-7 (MMP-7) correlates with<br />

tumor proliferation, and a poor prognosis in non-small cell lung cancer. Lung Cancer<br />

58, 384–391 (2007).<br />

32. Ito, T.K., Ishii, G., Chiba, H. & Ochiai, A. The VEGF angiogenic switch of<br />

fibroblasts is regulated by MMP-7 from cancer cells. Oncogene 26, 7194–7203<br />

(2007).<br />

33. Upham, J.W. et al. Retinoic acid modulates IL-5 receptor expression and selectively<br />

inhibits eosinophil-basophil differentiation of hemopoietic progenitor cells. J. Allergy<br />

Clin. Immunol. 109, 307–313 (2002).<br />

34. Takamura, K. et al. Retinoic acid inhibits interleukin-4-induced eotaxin production in a<br />

human bronchial epithelial cell line. Am. J. Physiol. Lung Cell. Mol. Physiol. 286,<br />

L777–L785 (2004).<br />

35. Takaki, H. et al. STAT6 Inhibits TGF-b1-mediated Foxp3 induction through direct<br />

binding to the Foxp3 promoter, which is reverted by retinoic acid receptor. J. Biol.<br />

Chem. 283, 14955–14962 (2008).<br />

36. McGowan, S.E., Holmes, A.J. & Smith, J. Retinoic acid reverses the airway hyperresponsiveness<br />

but not the parenchymal defect that is associated with<br />

vitamin A deficiency. Am. J. Physiol. Lung Cell. Mol. Physiol. 286, L437–L444<br />

(2004).<br />

37. Maret, M. et al. Liposomal retinoic acids modulate asthma manifestations in mice.<br />

J. Nutr. 137, 2730–2736 (2007).<br />

38. Schuster, G.U., Kenyon, N.J. & Stephensen, C.B. Vitamin A deficiency decreases and<br />

high dietary vitamin A increases disease severity in the mouse model of asthma.<br />

J. Immunol. 180, 1834–1842 (2008).<br />

39. Mora, J.R., Iwata, M. & von Andrian, U.H. Vitamin effects on the immune system:<br />

vitamins A and D take center stage. Nat. Rev. Immunol. 8, 685–698 (2008).<br />

40. Coombes, J.L. & Powrie, F. Dendritic cells in intestinal immune regulation. Nat. Rev.<br />

Immunol. 8, 435–446 (2008).<br />

41. Sun, C.-M. Small intestine lamina propria dendritic cells promote de novo generation of<br />

Foxp3 T reg cells via retinoic acid. J. Exp. Med. 204, 1775–1785 (2007).<br />

42. Benson, M.J., Pino-Lagos, K., Rosemblatt, M. & Noelle, R.J. All-trans retinoic acid<br />

mediates enhanced T reg cell growth, differentiation, and gut homing in the face of high<br />

levels of co-stimulation. J. Exp. Med. 204, 1765–1774 (2007).<br />

43. Elias, K.M. et al. Retinoic acid inhibits Th17 polarization and enhances FoxP3<br />

expression through a Stat-3/Stat-5 independent signaling pathway. Blood 111,<br />

1013–1020 (2008).<br />

44. Peters-Golden, M. The alveolar macrophage: the forgotten cell in asthma. Am.<br />

J. Respir. Cell Mol. Biol. 31, 3–7 (2004).<br />

45. Careau, E. et al. Antigen sensitization modulates alveolar macrophage functions in<br />

an asthma model. Am. J. Physiol. Lung Cell. Mol. Physiol. 290, L871–L879<br />

(2006).<br />

46. Diaz-Sanchez, D., Tsien, A., Fleming, J. & Saxon, A. Combined diesel exhaust<br />

particulate and ragweed allergen challenge markedly enhances human in vivo nasal<br />

ragweed-specific IgE and skews cytokine production to a T helper cell 2-type pattern.<br />

J. Immunol. 158, 2406–2413 (1997).<br />

47. Grumelli, S. et al. An immune basis for lung parenchymal destruction in chronic<br />

obstructive pulmonary disease and emphysema. PLoS Med. 1, e8(2004).<br />

48. Montes, M., Jaensson, E.A., Orozco, A.F., Lewis, D.E. & Corry, D.B. A general method<br />

for bead-enhanced quantitation by flow cytometry. J. Immunol. Methods 317, 45–55<br />

(2006).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 503


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Transcription factor Foxo3 controls the magnitude of<br />

T cell immune responses by modulating the function<br />

of dendritic cells<br />

Anne S Dejean 1 , Daniel R Beisner 1,4 , Irene L Ch’en 1 , Yann M Kerdiles 1 , Anna Babour 1 , Karen C Arden 2 ,<br />

Diego H Castrillon 3,4 , Ronald A DePinho 3 & Stephen M Hedrick 1<br />

Foxo transcription factors regulate cell cycle progression, cell survival and DNA-repair pathways. Here we demonstrate that<br />

deficiency in Foxo3 resulted in greater expansion of T cell populations after viral infection. This exaggerated expansion was not<br />

T cell intrinsic. Instead, it was caused by the enhanced capacity of Foxo3-deficient dendritic cells to sustain T cell viability by<br />

producing more interleukin 6. Stimulation of dendritic cells mediated by the coinhibitory molecule CTLA-4 induced nuclear<br />

localization of Foxo3, which in turn inhibited the production of interleukin 6 and tumor necrosis factor. Thus, Foxo3 acts to<br />

constrain the production of key inflammatory cytokines by dendritic cells and to control T cell survival.<br />

Foxo transcription factors guide the cellular response to growth factors,<br />

nutrients and stress. This information is ‘encoded’ as post-translational<br />

modifications, referred to as the ‘Foxo code’, that govern Foxo<br />

intracellular localization, cofactor associations and transcriptional<br />

activity. The resulting program of gene expression regulates cell cycle<br />

arrest, repair, apoptosis and autophagy and many other aspects of<br />

cellular homeostasis 1 . In mammals, four Foxo members have been<br />

identified. Foxo1 (FKH1 and FKHR) 2 , Foxo3 (FKHRL1; A000945) 3 ,<br />

and Foxo4 (AFX) 4 are widely expressed and similarly regulated,<br />

whereas expression of Foxo6 (ref. 5) is confined to specific structures<br />

of the brain and is subject to distinct regulatory mechanisms.<br />

Insulin, insulin-like growth factor and other growth factors induce<br />

activation of phosphatidylinositol-3-OH kinase and the kinase Akt,<br />

which phosphorylate three Foxo amino acids 6 . These modifications<br />

result in the association of the Foxo protein with the adaptor protein<br />

14-3-3 and the nuclear exclusion and eventual degradation of Foxo3<br />

(refs. 7–9). This process can be actively opposed by stress-induced<br />

signals that activate the Jnk mitogen-activated protein kinase, which<br />

result in Foxo3 nuclear localization 10 .Foxofactortargetspecificitycan<br />

be further refined by the SIRT1 deacetylase 11,12 .<br />

Given their involvement in coordinating cellular growth, proliferation<br />

and survival, we predicted that Foxo factors would be central to<br />

the highly dynamic, infection-mediated expansion and contraction of<br />

antigen-specific T cell populations. Indeed, Foxo activity is regulated<br />

by signaling by the T cell antigen receptor (TCR) and CD28, as well as<br />

by cytokines such as interleukin 2 (IL-2) 13 , IL-3 (ref. 14) and IL-7<br />

Received 22 September 2008; accepted 19 March 2009; published online 12 April 2009; doi:10.1038/ni.1729<br />

(refs. 15,16). These stimuli result in the phosphorylation of Foxo by<br />

Akt, the kinase Sgk or inhibitor of transcription factor NF-kB (IkB)<br />

kinase, and in Foxo3 nuclear exclusion 17 . Withdrawal of growth factor<br />

causes dephosphorylation of nuclear Foxo3, binding of Foxo3 to the<br />

promoters of the genes encoding the proapoptotic molecules Bim and<br />

Puma, induction of transcription of those genes, and T cell apoptosis<br />

18 . In contrast, enforced expression of a constitutively nuclear<br />

form of Foxo3 in T cell lines causes cell cycle arrest 13 . Furthermore,<br />

Foxo3 is involved in the persistence of CD4 + central memory T cells in<br />

mice and humans 19 .<br />

Published studies have shown spontaneous T cell activation, lymphoproliferative<br />

disease and organ infiltration in ‘Foxo3 Trap ’mice,which<br />

have a mutated Foxo3 allele generated by gene-trap technology 20 .Such<br />

studies indicate an important function for Foxo3 in immune regulation,<br />

although the underlying molecular mechanisms are not understood. In<br />

addition, Foxo3 regulates superoxide dismutase in dendritic cells (DCs);<br />

this signaling pathway may affect the suppressive or stimulating<br />

characteristics of plasmacytoid DCs 21,22 .<br />

In this study, we sought to determine whether Foxo3 is involved in<br />

the quiescence of cells of the immune system and the dynamics of<br />

T cell population expansion and contraction. We found that in<br />

response to LCMV infection, Foxo3 deficiency caused superabundant<br />

expansion of antigen-specific T cell populations. However, contrary to<br />

our expectations, this phenotype was not intrinsic to T cells but<br />

instead arose from altered stimulatory properties of Foxo3-deficient<br />

DCs. Foxo3 deficiency resulted in a DC-specific higher production of<br />

1 Molecular Biology Section, Division of Biological Sciences, and Department of Cellular and Molecular Medicine, University of California, San Diego, San Diego, California,<br />

USA. 2 Ludwig Institute for Cancer Research, University of California, San Diego School of Medicine, La Jolla, California, USA. 3 Center for Applied Cancer Science, Belfer<br />

Institute for Innovative Cancer Science, Departments of Adult Oncology, Medicine and Genetics, Dana-Farber Cancer Institute, Harvard Medical School, Boston,<br />

Massachusetts, USA. 4 Present addresses: Genetics Institute of the Novartis Foundation, San Diego, California, USA (D.R.B.) and Department of Pathology, University of<br />

Texas, Southwestern Medical Center at Dallas, Dallas, Texas, USA (D.H.C.). Correspondence should be addressed to S.M.H. (shedrick@ucsd.edu).<br />

504 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Table 1 Leukocyte subsets<br />

Spleen B6 B6 FVB FVB<br />

Foxo3 +/+ Foxo3Kca Foxo3 +/+ Foxo3 –/–<br />

CD4 + 15.0 ± 2.0 17.3 ± 1.6 29.3 ± 5.1 31.0 ± 5.9<br />

CD8 + 13.4 ± 3.5 9.7 ± 1.3 13.0 ± 2.2 14.9 ± 1.9<br />

B220 + 46.9 ± 6.7 51.7 ± 9.2 58.2 ± 9.7 56.3 ± 12.1<br />

Gr-1hiCD11bhi Lymph node<br />

1.4 ± 0.2 6.0 ± 1.1* 2.0 ± 0.2 5.4 ± 0.3*<br />

CD4 + 10.3 ± 1.4 9.4 ± 1.2 11.0 ± 2.5 12.8 ± 2.4<br />

CD8 + 7.9 ± 1.0 7.3 ± 0.7 5.5 ± 1.2 7.0 ± 1.4<br />

B220 + 13.8 ± 1.4 12.4 ± 1.1 4.0 ± 1.2 3.6 ± 0.7<br />

Results are presented as the number of cells 10 6 (± s.e.m.). *, P o 0.001, wild-type<br />

versus Foxo3-deficient mice (two-sample t-test assuming equal variances). Data are<br />

representative of two experiments with five to six mice per group.<br />

IL-6. Nuclear localization of Foxo3 induced by signals initiated by<br />

cytotoxic T lymphocyte–associated antigen 4 (CTLA-4; A000706)<br />

suppressed Toll-like receptor–induced production of IL-6. Our results<br />

collectively emphasize the importance of Foxo3 in restraining the<br />

production of inflammatory cytokines by DCs and T cell viability.<br />

RESULTS<br />

No spontaneous activation of Foxo3-deficient T cells<br />

To examine the function of Foxo3 in the immune system, we used two<br />

independently derived Foxo3-null strains distinct from the Foxo3 Trap<br />

mutant 20 . One strain, called ‘Foxo3 Kca ’ here, was backcrossed to the<br />

C57BL/6 strain, whereas mice of the other strain, called ‘Foxo3 –/– ’here,<br />

were maintained as congenic FVB mice. Although we detected strainspecific<br />

differences, deletion of Foxo3 had no effect on the proportion<br />

of T cells or B cells in the spleens and lymph nodes of 6- to 10-weekold<br />

mice (Table 1). However, we noted a greater proportion of<br />

Gr-1 hi CD11b hi cells, which include granulocytes and macrophages,<br />

in the spleens of both Foxo3-deficient strains. We detected neither<br />

lymphocytic organ infiltration nor lymphadenopathy 23 (data not<br />

shown) in 6- to 10-week-old Foxo3-deficient mice. However 3- to<br />

6-month-old Foxo3-deficient mice had enlargement of the spleen<br />

a b<br />

CD44<br />

CD69<br />

B6<br />

Foxo3 +/+<br />

B6<br />

Foxo3 Kca<br />

FVB<br />

Foxo3 +/+<br />

FVB<br />

Foxo3 –/–<br />

7.2 ± 0.6 7 ± 1.1 4.1 ± 0.6 4.9 ± 0.5<br />

CD62L<br />

14.5 ± 0.9 14.3 ± 1.1 12 ± 0.9 13.4 ± 1.3<br />

CD4<br />

CD8<br />

CD69<br />

B6<br />

Foxo3 +/+<br />

B6<br />

Foxo3 Kca<br />

FVB<br />

Foxo3 +/+<br />

FVB<br />

Foxo3 –/–<br />

13.2 ± 0.7 14.4 ± 1.5 13.4 ± 1.1 15.8 ± 1.4<br />

CD44<br />

7.4 ± 0.3 7.4 ± 0.3 8 ± 0.8 8.1 ± 1.0<br />

CD8<br />

Figure 1 Foxo3-deficient mice show no spontaneous T cell<br />

activation. (a) Flow cytometry of the expression of CD44, CD62L<br />

and CD69 by T cells from C57BL/6 (B6) Foxo3 Kca and FVB<br />

Foxo –/– mice and their wild-type (Foxo +/+ ) congenic littermates<br />

(n ¼ 5–6 mice per genotype), gated on CD4 + T cells (left) or CD8 +<br />

T cells (right). Numbers adjacent to outlined areas indicate percent<br />

cells in gate (mean ± s.e.m.). (b) Proliferation of CFSE-labeled<br />

CD4 + T cells from Foxo3 Kca and Foxo –/– mice and their wild-type<br />

littermates, activated in the presence of anti-CD3 (a-CD3) alone or<br />

CD4 + c T cells<br />

0.6 ± 0.1<br />

IL-4<br />

10 ± 1<br />

0.7 ± 0.1<br />

IFN-γ<br />

and more erythrocyte progenitors (Ter119 + ; data not shown). That<br />

observation probably relates to the shorter erythrocyte lifespan<br />

and lower rate of erythrocyte maturation in Foxo3-deficient mice<br />

reported before 24 . Next we measured the expression of CD69, CD44<br />

and CD62L (L-selectin) on T cells from Foxo3-deficient mice and<br />

their wild-type littermates. Although we again detected differences<br />

between the C57BL/6 and FVB strains, inactivation of Foxo3 had no<br />

effect on the proportion of activated (CD69 hi ) or memory-effector<br />

(CD44 hi CD62L lo )Tcells(Fig. 1a).<br />

Published studies have shown constitutive activation of the NF-kB<br />

pathway in Foxo3 Trap mice, as reflected by the absence of IkB 20 .<br />

However, T cells purified from Foxo3 Kca and Foxo3 –/– mice showed<br />

no change in the expression of IkBa, IkBb or IkBe protein relative to<br />

that of wild-type T cells (Supplementary Fig. 1 online). To measure<br />

the effect of Foxo3 deficiency on T cell proliferation, we labeled<br />

purified lymph node T cells with the cytosolic dye CFSE and then<br />

cultured the cells with plate-bound antibody to CD3 (anti-CD3) with<br />

or without anti-CD28. Neither the number of cell divisions nor the<br />

accumulation of cells in each division was altered by Foxo3 deficiency<br />

(Fig. 1b). Also, freshly explanted Foxo3 Kca and wild-type CD4 + and<br />

CD8 + T cells produced similar amounts of cytokines (Fig. 1c). As FasL<br />

(CD95L), the ligand for the cell surface receptor Fas, is a known target<br />

of Foxo3, we also analyzed FasL expression on unstimulated and<br />

antibody-stimulated T cells from wild-type and Foxo3 Kca mice; however<br />

we found no differences related to Foxo3 status (data not shown).<br />

These results suggest that loss of Foxo3 alone is not sufficient to elicit<br />

manifestations of T cell activation and spontaneous autoimmunity.<br />

Foxo3 Kca mice show enhanced T cell accumulation<br />

As Foxo3 is involved in cell cycle progression and apoptosis 25 ,we<br />

sought to investigate its function in the dynamics of T cell population<br />

expansion and contraction in response to viral infection. Historically,<br />

the C57BL/6 strain has been used to study the progression of viral<br />

responses, and as the two mutant strains homozygous for Foxo3<br />

deficiency had identical profiles here and in published results 23,26 ,our<br />

further efforts at characterization focused on the C57BL/6 Foxo3 Kca<br />

mice. We infected Foxo3 Kca mice and their wild-type littermates with<br />

12 ± 2<br />

IL-17<br />

0.5 ± 0.1<br />

10 ± 1<br />

0.5 ± 0.1<br />

IFN-γ<br />

12 ± 2<br />

CD4<br />

IL-2<br />

B6<br />

FVB<br />

Cells<br />

Control α-CD3<br />

α-CD3 +<br />

α-CD28<br />

500 400 1,000<br />

1,000 400 2,000<br />

CFSE<br />

Foxo3 +/+<br />

20 ± 2<br />

Foxo3 Kca<br />

22 ± 1<br />

TNF<br />

CD8 + T cells<br />

32 ± 3 22 ± 2<br />

4 ± 1<br />

27 ± 6 22 ± 2<br />

CD8<br />

6 ± 2<br />

IFN-γ IL-2<br />

ARTICLES<br />

12 ± 2<br />

13 ± 3<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 –/–<br />

Foxo3 +/+<br />

with anti-CD28 (a-CD28) and assessed by CFSE dilution after 72 h of stimulation. Numbers in plots indicate number of accumulated CFSE + cells.<br />

(c) Cytokine secretion by splenocytes from Foxo3 Kca mice and their wild-type littermates (n ¼ 6 mice per group), activated for 3 h in the presence of phorbol<br />

12-myristate 13-acetate and ionomycin and analyzed by intracellular staining with gating on CD4 + T cells (left) or CD8 + T cells (right). Numbers in quadrants<br />

and adjacent to outlined areas indicate percent cells in each (mean ± s.e.m.). Similar results were obtained with lymph node T cells. Data are from five<br />

separate experiments (a) or are representative of three independent experiments (b,c).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 505<br />

Foxo3 +/+<br />

Foxo3 Kca


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a b LCMV No infection c<br />

T cells (× 10 6 )<br />

gp61-specific<br />

CD4<br />

4<br />

+ T cells<br />

**<br />

3<br />

2<br />

1<br />

**<br />

*<br />

T cells (× 10 6 )<br />

gp33-specific<br />

CD8 + T cells<br />

30<br />

**<br />

Foxo3 Kca<br />

0<br />

0<br />

0 10 20 50 60 0 10 20 50 60<br />

Time (d) Time (d)<br />

20<br />

10<br />

Foxo3 +/+<br />

Adoptive transfer<br />

P14 Foxo3 +/+<br />

P14 Foxo3 Kca<br />

Adoptive transfer<br />

lymphocytic choriomeningitis virus (LCMV) and assessed LCMVresponsive<br />

T cell populations at various times after infection. Whereas<br />

the Foxo3 genotype had no effect on the kinetics of T cell population<br />

expansion, Foxo3 Kca mice developed a threefold greater accumulation<br />

of LCMV-specific T cells than did their wild-type littermates (Fig. 2).<br />

Measurement of virus in the liver at day 8 after infection showed<br />

complete viral clearance in both strains (data not shown).<br />

The lack of an effect of Foxo3 deficiency on T cells stimulated in<br />

culture prompted us to test whether the enhanced T cell accumulation<br />

during an LCMV response was T cell intrinsic. We transferred T cells<br />

from LCMV-specific TCR-transgenic P14 Foxo3 Kca or P14 wild-type<br />

mice into wild-type C57BL/6 mice and measured the accumulation of<br />

P14 T cells after infection of the recipient mice with LCMV. We<br />

detected no significant difference in the accumulation of Foxo3 Kca and<br />

wild-type P14 T cells (Fig. 2b). To further examine this issue, we<br />

transferred wild-type P14 T cells into either wild-type or Foxo3 Kca<br />

mice and infected recipients with LCMV. P14 T cells transferred into<br />

Foxo3 Kca mice accumulated in greater numbers than did P14 T cells<br />

transferred into wild-type hosts (Fig. 2b). We reproduced that result<br />

with the transfer of ovalbumin (OVA)-specific OT-I T cells into wildtype<br />

and Foxo3 Kca hosts that we subsequently infected with OVAexpressing<br />

vesicular stomatitis virus (Supplementary Fig. 2 online).<br />

To further investigate the cellular cause of the enhanced T cell<br />

population expansion, we reconstituted irradiated wild-type mice with<br />

bone marrow from wild-type or Foxo3 Kca mice. After 8 weeks, we<br />

infected the recipient mice with LCMV and analyzed expansion of the<br />

LCMV-specific T cell population. Excessive T cell accumulation was<br />

apparent in recipients of Foxo3 Kca bone marrow (Fig. 2c), which<br />

indicated involvement of a bone marrow–derived cell type in this<br />

phenotype. Thus, a non–T cell bone marrow–derived cell type was<br />

responsible for the enhanced T cell proliferation and/or survival in<br />

Foxo3-deficient mice.<br />

Stimulatory capacity of DCs from Foxo3 Kca mice<br />

The efficiency of antigen presentation can affect the magnitude<br />

of a T cell response to viral infection 27 . Therefore, we assessed<br />

whether Foxo3 regulates the number, phenotype and/or function of<br />

P14<br />

P14<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

T cells (× 10 6 )<br />

T cells (× 10 6 )<br />

4<br />

3<br />

2<br />

1<br />

0<br />

60<br />

40<br />

20<br />

Foxo3 Kca<br />

LCMV No infection<br />

**<br />

antigen-presenting DCs. Naive Foxo3 Kca mice had significantly more<br />

DCs than did their wild-type littermates (Fig. 3a), and there was a<br />

slightly greater proportion of DCs with high expression of the<br />

costimulatory molecules B7-1 (CD80) and B7-2 (CD86). We found<br />

no difference in the expression of major histocompatibility complex<br />

(MHC) class I, MHC class II or CD40 on DCs from these mice<br />

(Fig. 3b), which suggested that a small number of Foxo3-deficient<br />

DCs had a more mature phenotype.<br />

Analysis of various DC subsets showed that Foxo3-deficient mice<br />

had more CD11c + CD11b + CD8 – , CD11c + CD11b – CD8 + and<br />

CD11c + B220 + DCs (Fig. 3c). All subsets showed a moderate shift<br />

toward high expression of B7-1 and B7-2 (Fig. 3d), although the<br />

biological importance of this shift is unclear. To further characterize<br />

the activation status of DCs from naive mice, we cultured splenic DCs<br />

overnight and then measured secreted cytokines. Foxo3 Kca DCs<br />

produced more IL-6, tumor necrosis factor (TNF) and chemokine<br />

CCL2 (MCP-1) than did wild-type DCs (Fig. 3e). Interferon-g<br />

(IFN-g), IL-10 and IL-12 were undetectable (data not shown).<br />

To determine whether DCs from Foxo3 Kca mice had enhanced<br />

antigen presentation, we infected mice with LCMV and analyzed<br />

DCs 3 d after infection, when virus is still present and T cell<br />

populations are expanding exponentially 28 . Compared with wildtype<br />

DCs, Foxo3 Kca DCs had slightly higher expression of B7-1,<br />

B7-2 and MHC class II (Fig. 4a) and more effectively stimulated<br />

the accumulation of P14 T cells, as measured by CFSE dilution and<br />

staining with annexin V and 7-amino-actinomycin D (7-AAD;<br />

Fig. 4b). This enhanced stimulatory capacity was not due to a<br />

difference in antigen processing, as differences between cultures with<br />

Foxo3 Kca and wild-type DCs remained even after DCs were pulsed<br />

with LCMV glycoprotein 33 (gp33) peptide (Fig. 4b).<br />

To rule out the possibility that Foxo3-deficient DCs from LCMVinfected<br />

mice can cause proliferation of bystander T cells independently<br />

of the presence of antigen, we also cultured DCs from<br />

LCMV-infected mice with OT-I T cells with or without OVA<br />

peptide (amino acids 257–264; OVA (257–264)). In the absence of<br />

OVA (257–264), there was no detectable division of OT-I T cells<br />

(Fig. 4b), which ruled out the possibility of bystander effects. Finally,<br />

T cells (× 10 6 )<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

gp61-specific<br />

CD4 + T cells<br />

2.5<br />

**<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Radiation chimera<br />

gp33-specific<br />

CD8 + T cells<br />

5 *<br />

4<br />

Figure 2 Foxo3 regulates the magnitude of an LCMV-induced immune<br />

0<br />

response. (a) TheCD4 + and CD8 + T cell responses to LCMV in Foxo3 Kca<br />

mice and their wild-type littermates infected with LCMV, Armstrong strain<br />

(2 105 plaque-forming units), analyzed after restimulation with gp61 or<br />

gp33 peptide, respectively, by intracellular staining for IFN-g. (b) P14<br />

T cells in recipient mice given CD45.2 congenic wild-type or Foxo3 Kca P14 T cells (transferred into (-) CD45.1 wild-type mice; top) and CD45.1 congenic<br />

wild-type P14 T cells (transferred into CD45.2 wild-type or CD45.2 Foxo3Kca mice; bottom) and then left uninfected (right) or infected with LCMV (left) and<br />

assessed with a congenic marker 8 d later. (c) Virus-specific T cells in lethally irradiated (gray ‘lightning bolt’ around) CD45.1 wild-type mice reconstituted<br />

with CD45.2 congenic wild-type or Foxo3Kca bone marrow and then infected with LCMV 8 weeks later, assessed by specific peptide restimulation and<br />

intracellular IFN-g staining. *, P o 0.01, and **, P o 0.005 (unpaired two-tailed Student’s t-test). Data are representative of four (a) ortwo(b,c)<br />

independent experiments with at least three mice per group (error bars, s.e.m.).<br />

506 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

3<br />

2<br />

1<br />

0<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 Kca


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b<br />

CD11c + cells/spleen (× 10 5 )<br />

40<br />

20<br />

0<br />

Foxo3 +/+<br />

***<br />

Foxo3 Kca<br />

B7-1 hi (%)<br />

B7-1 hi<br />

B7-1 B7-2 MHC II<br />

MHC I CD40<br />

20 ** 20 ** 60<br />

20<br />

20<br />

15<br />

15<br />

40<br />

15<br />

15<br />

10<br />

5<br />

10<br />

5<br />

20<br />

10<br />

5<br />

10<br />

5<br />

0<br />

0<br />

0<br />

0<br />

0<br />

B7-2 hi (%)<br />

B7-2 hi<br />

Figure 3 Foxo3 Kca mice have more DCs and greater<br />

activation of DCs. (a) Total DCs among splenocytes<br />

obtained from Foxo3 Kca mice and their wild-type<br />

littermates (n ¼ 10 mice per group) and stained for<br />

CD11c. (b) Flow cytometry of CD11c + DCs obtained<br />

from naive Foxo3 Kca mice and their wild-type littermates<br />

(n ¼ 5 mice per group) and stained with antibodies<br />

specific for various markers (below plots). Bottom row,<br />

accumulated data. MHCII, MHC class II; MHCI, MHC<br />

class I. (c) Absolute numbers of CD11c + CD11b + CD8 – ,<br />

CD11c + CD11b – CD8 + and CD11c + B220 + DCs in spleens<br />

we measured the capacity of DCs from uninfected mice to stimulate<br />

P14 T cells in the presence of gp33 peptide. We noted slightly greater<br />

accumulation and viability of T cells from cultures containing<br />

Foxo3 Kca DCs than from those containing wild-type DCs (Fig. 4c).<br />

To determine if the enhanced function of Foxo3-deficient DCs was<br />

cell autonomous, we generated DCs in vitro. Forthis,wecultured<br />

bone marrow cells for 8 d with granulocyte-macrophage colonystimulating<br />

factor. The number and proportion of the resulting<br />

CD11c + cells (called ‘BMDCs’ here) were unaffected by Foxo3 status,<br />

and none of the BMDCs had higher expression of B7-1, B7-2, CD40,<br />

MHC class I or MHC class II (Supplementary Fig. 3a online).<br />

Consistent with published reports of cells stimulated with granulocyte-macrophage<br />

colony-stimulating factor, wild-type BMDCs had a<br />

uniformly cytoplasmic Foxo3 localization 29 (Supplementary Fig. 3b).<br />

Next we used BMDCs to stimulate OT-II and P14 T cells at various<br />

ratios of T cells to DCs. Foxo3 Kca BMDCs induced accumulation of<br />

OT-II and P14 T cells more effectively than did wild-type BMDCs<br />

(Fig. 5a). Moreover, the Foxo3-deficient BMDCs also more effectively<br />

sustained T cell viability, regardless of cell division (Fig. 5b,c). To<br />

determine whether the differences in viability correlated with changes<br />

in Bcl-2 family members, we analyzed expression of the prosurvival<br />

factors Bcl-2 and Bcl-x L. Total and naive CD44 lo T cells had higher<br />

expression of both when stimulated with Foxo3-deficient BMDCs<br />

(Fig. 5d).<br />

To assess the capacity of BMDCs to enhance T cell survival in vivo,<br />

we transferred CFSE-labeled OT-II T cells into naive CD45.1 congenic<br />

hosts. Then, 1 d later, we transferred wild-type or Foxo3-deficient<br />

BMDCs, with or without preloaded OVA peptide (amino acids<br />

323–339; OVA(323–339), into the footpads of these mice. After<br />

2 additional days, Foxo3 Kca BMDCs induced a greater accumulation<br />

of OT-II T cells in the draining lymph nodes but a similar number of<br />

OT-II T cell divisions, compared with wild-type DCs (Fig. 5e).<br />

Notably, Foxo3 Kca DCs facilitated greater OT-II T cell survival, even<br />

MHC II hi (%)<br />

d<br />

MHC I hi (%)<br />

CD40 hi (%)<br />

CD11c + CD11b + CD11c + B220 + CD11c + CD8 +<br />

B7-1<br />

B7-2<br />

MHCII hi<br />

MHCI hi<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

CD40 hi<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

c<br />

CD11c + CD11b +<br />

cells/spleen (× 10 5 )<br />

e<br />

IL-6 (pg/ml)<br />

15<br />

10<br />

5<br />

0<br />

Foxo3 +/+<br />

60<br />

40<br />

20<br />

0<br />

*<br />

Foxo3 Kca<br />

CD11c + B220 +<br />

cells/spleen (× 10 5 )<br />

CD11c + CD8 +<br />

cells/spleen (× 10 5 )<br />

in the absence of OVA(323–339). These data collectively establish that<br />

Foxo3-deficient DCs produce greater antigen-induced T cell population<br />

expansion by enhancing survival.<br />

Enhanced IL-6 secretion by Foxo3-deficient DCs<br />

To determine if the enhanced T cell survival induced by Foxo3deficient<br />

DCs was due to altered cytokine secretion, we infected<br />

wild-type and Foxo3 Kca mice with LCMV and measured cytokine<br />

concentrations in blood plasma at various times after infection. The<br />

concentrations of IL-12, IL-17, CCL2 (MCP-1), IL-1b, IL-5, IL-10<br />

were similar in wild-type and Foxo3 Kca mice (Supplementary Fig. 4a<br />

online), but the concentrations of IL-6 and TNF were higher in Foxo3deficient<br />

mice (Fig. 6a and Supplementary Fig. 4a). To determine if<br />

DCs were the source of this abundant IL-6 and TNF, we collected<br />

splenic DCs from mice on day 3 after infection and cultured the cells<br />

for 24 h. Supernatants of Foxo3 Kca DC cultures had significantly higher<br />

concentrations of IL-6, TNF, CCL2 (MCP-1) and IFN-g (Fig. 6b and<br />

Supplementary Fig. 4b), whereas we detected no difference in the<br />

concentration of IL-10 or IL-12. We also found higher expression of<br />

IL-6 mRNA in Foxo3 Kca DCs (Fig. 6b).<br />

To determine whether cells other than DCs could be responsible for<br />

the greater IL-6 production, we collected T cells, B cells and macrophages<br />

from mice 3 d after LCMV infection and cultured the cells for<br />

24 h. Neither T cells nor B cells had detectable production of IL-6<br />

(data not shown). Macrophages did produce IL-6, although there was<br />

no difference related to Foxo3 status (data not shown). However, given<br />

the finding of more macrophages in Foxo3-deficient mice, these cells<br />

may have contributed to the excess IL-6 production in vivo.<br />

We next determined whether BMDCs also produced excess IL-6 in<br />

the absence of Foxo3. We detected significantly more IL-6 in cultures<br />

containing OT-II CD4 + T cells or OT-I CD8 + T cells and Foxo3 Kca<br />

BMDCs than in those containing T cells and wild-type DCs (Fig. 6c).<br />

Confirming the DC-specific nature of this abundant IL-6, RT-PCR<br />

TNF (pg/ml)<br />

25<br />

0<br />

6<br />

3<br />

0<br />

IL-6<br />

**<br />

*** 50 * *<br />

60<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

ARTICLES<br />

of Foxo3 Kca mice and their wild-type littermates (n ¼ 5 mice per group). (d) Flow cytometry of the expression of B7-1 and B7-2 on DC subsets (above plots)<br />

from Foxo3 Kca mice and their wild-type littermates. (e) Cytokine bead array analysis of the concentration of IL-6, TNF and CCL2 (MCP-1) in supernatants of<br />

splenic DCs (CD11c + ) purified from wild-type or Foxo3 Kca mice and cultured for 24 h (4 10 5 cells). In a,c, each symbol represents an individual mouse;<br />

small horizontal lines indicate the mean. *, P o 0.01; **, P o 0.005; and ***, P o 0.001 (unpaired two-tailed Student’s t-test). Data are representative of<br />

three independent experiments (a,b) or two independent experiments with at least three mice per group (c–e; error bars (b,e), s.e.m.).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 507<br />

MCP-1 (pg/ml)<br />

30<br />

0<br />

4<br />

2<br />

0<br />

Foxo3 +/+<br />

**<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 Kca


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

b<br />

Foxo3 +/+<br />

Foxo3<br />

Annexin V<br />

Kca<br />

Foxo3<br />

OTI, no peptide OTI, OVAp<br />

Kca<br />

Foxo3 +/+<br />

B7-1 B7-2 MHCII MHCI<br />

P14, no peptide P14, gp33p<br />

CFSE<br />

9<br />

23 5 14<br />

24<br />

7-AAD<br />

44<br />

9<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

showed that Foxo3 Kca BMDCs had twofold more IL-6 mRNA than did<br />

wild-type BMDCs; IL-6 mRNA was at least tenfold greater in BMDCs<br />

than in CD4 + T cells, and IL-6 mRNA was undetectable in CD8 +<br />

T cells (Fig. 6d).<br />

In combination with transforming growth factor-b, IL-6induces<br />

the differentiation of IL-17-producing T cells 30 . Given the results<br />

presented above, we analyzed the balance between IL-17-producing<br />

T cells and Foxp3 + regulatory T cells in Foxo3-deficient mice.<br />

OT-II<br />

P14<br />

Cells<br />

24<br />

c<br />

Foxo3 +/+<br />

Foxo3<br />

Uninfected mice<br />

P14, no peptide P14, gp33p<br />

Kca<br />

Foxo3 +/+<br />

Foxo3<br />

Annexin V<br />

Kca<br />

CFSE<br />

a b<br />

d<br />

Cells<br />

30:1 90:1 270:1<br />

200 150 130<br />

1,200 800 400<br />

CFSE<br />

Bcl-2<br />

OT-II<br />

P14<br />

OT-II, CD44 lo<br />

P14, CD44 lo<br />

Bcl-x L<br />

P14<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

2.5<br />

49<br />

6 60<br />

7-AAD<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

7-AAD<br />

OT-II OT-II, CD44 lo<br />

P14, CD44 lo<br />

CFSE<br />

The proportion of Foxp3 + and IL-17 + CD4 + T cells in naive<br />

LCMV-infected Foxo3 Kca mice was similar to that of their wild-type<br />

littermates (Supplementary Fig. 5 online). We did not detect<br />

IL-17-producing cells in the CD8 + lineage (data not shown).<br />

To evaluate whether the excess IL-6 produced by Foxo3-deficient<br />

DCs was required for their ability to induce enhanced T cell survival,<br />

we cultured OT-II and P14 T cells with BMDCs in the presence of an<br />

IL-6-specific blocking antibody or an immunoglobulin G1 (IgG1)<br />

isotype-matched control antibody. IL-6 blockade diminished the<br />

expansion of T cell populations cultured with Foxo3 Kca BMDCs to<br />

an amount resembling that of T cells cultured with wild-type DCs<br />

(Fig. 6e). Counting of dead cells stained with 7-AAD demonstrated<br />

OT-II<br />

P14<br />

OVAp No peptide gp33p No peptide<br />

24 48<br />

16 27<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Figure 4 Greater immunogenicity of LCMV-infected Foxo3-deficient DCs.<br />

(a) Flow cytometry of the expression of activation markers on wild-type and<br />

Foxo3 Kca DCs on day 3 after LCMV infection. (b) Proliferation and viability<br />

of CFSE-labeled wild-type P14 CD8 + T cells stimulated by DCs isolated from<br />

LCMV-infected Foxo3 Kca mice or their wild-type littermates (day 3) in the<br />

presence or absence of gp33 peptide (gp33p; left), or of CFSE-labeled<br />

wild-type OT-I CD8 + T cells stimulated in the presence or absence of<br />

OVA (323–339) (OVAp; bystander proliferation; right), assessed by CFSE<br />

dilution (proliferation; top row) or staining with annexin V and 7-AAD after<br />

3 d of culture (viability; middle and bottom rows). Numbers in dot plots<br />

indicate percent live CD8 + Tcells.(c) Proliferation and viability of T cells<br />

from P14 mice stimulated by DCs purified from the spleens of uninfected<br />

mice, cultured and analyzed as described in b. Data are representative of<br />

two independent experiments with at least three mice per group (a,c) or<br />

three independent experiments (b).<br />

65 87<br />

23 68<br />

e CD45.1 mice<br />

0 1 2 3<br />

CFSE-labeled<br />

CD45.2<br />

OT-II T cells<br />

CFSE<br />

OVAppulsed<br />

BMDC<br />

footpad<br />

BMDC BMDC + OVAp<br />

CD4 + CD45.2 + cells (× 10 5 )<br />

c<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Annexin V<br />

Analysis of<br />

T cells in<br />

draining LN<br />

***<br />

BMDC<br />

32<br />

78<br />

7-AAD<br />

OT-II<br />

BMDC +<br />

OVAp<br />

21<br />

46<br />

P14<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Figure 5 Enhanced T cell responses induced by Foxo3 Kca BMDCs. (a) CFSE-dilution analysis of the accumulation of wild-type OT-II CD4 + T cells or wild-type<br />

P14 CD8 + T cells stimulated for 3 d at various ratios (above plots) with wild-type or Foxo3 Kca BMDCs in the presence of OVA(323–339) (for OT-II T cells)<br />

or gp33 peptide (for P14 T cells). Numbers in plots indicate number of accumulated CSFE + cells. (b) Death of OT-II CD4 + T cells or P14 CD8 + Tcells<br />

(1 10 5 ) cultured with wild-type or Foxo3 Kca BMDCs (3.3 10 2 ) in the presence or absence of the appropriate peptide, assessed by 7-AAD staining.<br />

Numbers adjacent to outlined areas indicate percent 7-AAD + Tcells.(c) Viability of OT-II CD4 + TcellsorP14CD8 + T cells cultured for 3 d with BMDCs in<br />

the presence or absence of the appropriate peptide and stained with annexin V and 7-AAD. Numbers in outlined areas indicate percent viable CD4 + or CD8 +<br />

Tcells.(d) Flow cytometry of the expression of Bcl-2 and Bcl-x L in total T cells or gated CD44 lo T cells among OT-II or P14 T cells cultured as described in<br />

b. Dashed lines, isotype-matched control antibody. (e) Proliferation (left) and number (right) of CD4 + T cells purified from wild-type CD45.2 OT-II mice,<br />

labeled with CFSE and injected intravenously into CD45.1 recipient mice, followed by subcutaneous injection of unpulsed or OVA(323–339)-pulsed<br />

(+ OVAp) BMDCs from Foxo3 Kca mice or their wild-type littermates (n ¼ 4 mice per group) into the footpads of the same recipient mice and analysis 3 d<br />

later. Left, CFSE-dilution analysis; right, CD45.1 cells in draining lymph nodes (LN). Data are representative of at least six independent experiments (a,b) or<br />

two (c,d) orthree(e) independent experiments (error bars (e), s.e.m.).<br />

508 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b c Foxo3 d<br />

e<br />

+/+<br />

Foxo3 +/+<br />

Foxo3 +/+<br />

Serum IL-6 (pg/ml)<br />

250<br />

200<br />

150<br />

100<br />

50<br />

***<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

0 2 4 6 8 10<br />

IL-6 (pg/ml)<br />

100<br />

75<br />

50<br />

25<br />

0<br />

0 3<br />

that this reversion was due to lower T cell survival (Fig. 6f). In<br />

addition, we determined whether exogenous IL-6 was sufficient to<br />

enhance the survival of T cells cultured with wild-type BMDCs.<br />

Indeed, we noted a direct correlation between the amount of IL-6<br />

added and the proportion of live T cells in each culture 31 (Fig. 6g).<br />

To directly ascertain if IL-6 was responsible for the greater expansion<br />

of T cell populations in response to LCMV, we treated mice on<br />

days –1 and +4 (relative to LCMV infection at day 0) with 100 mg of<br />

blocking antibody to IL-6 receptor a-chain. Blockade of this receptor<br />

resulted in LCMV-specific T cell responses of substantially smaller<br />

magnitude in Foxo3-deficient mice but did not affect the number of<br />

LCMV-specific T cells in wild-type mice (Fig. 6h). These experiments<br />

collectively suggest that wild-type DCs are restrained from abundant<br />

IL-6 production by Foxo3 and thus that the factors that affect the<br />

localization and transcriptional activity of Foxo3 will influence the<br />

magnitude of a T cell–mediated immune response.<br />

Foxo3 acts ‘downstream’ of CTLA-4-induced signals<br />

Stimulation of B7 receptors by CTLA-4 promotes nuclear localization<br />

of Foxo3 and activation of DCs 32 . As cell surface CTLA-4<br />

expression is induced on activated T cells, we analyzed whether<br />

signaling through the B7 receptor inhibits IL-6 production in a<br />

Foxo3-dependent way. We stimulated splenic DCs with loxoribine,<br />

a Toll-like receptor 7 agonist, in the presence of varying amounts<br />

of a fusion of CTLA-4 and immunoglobulin (CTLA-4–Ig).<br />

The production of IL-6 and TNF stimulated by loxoribine was<br />

strongly inhibited by CTLA-4–Ig in wild-type DCs but not in<br />

***<br />

Il6 mRNA<br />

(relative expression)<br />

f<br />

7-AAD<br />

0 3<br />

Time (d) Time (d) Time (d)<br />

Figure 6 IL-6 synthesis by Foxo3-deficient<br />

DCs is involved in enhanced T cell survival.<br />

(a) Cytokine bead array analysis of IL-6 in<br />

plasma from wild-type and Foxo3 Kca mice<br />

(n ¼ 4 mice per group) at various times after<br />

LCMV infection. (b) Cytokine bead array<br />

analysis (left) of IL-6 secreted by DCs<br />

(CD11c + ) purified from wild-type and<br />

Foxo3 Kca mice without infection (0 d) or 3 d<br />

after LCMV infection (n ¼ 4 mice per group)<br />

and then cultured for 24 h (2 10 5 cells).<br />

Right, RT-PCR analysis of Il6 mRNA<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0.0<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

Foxo3 Kca<br />

CFSE<br />

**<br />

IL-6 (pg/ml)<br />

200<br />

150<br />

100<br />

50<br />

Isotype Anti-IL-6<br />

64 66<br />

42 61<br />

0<br />

Foxo3 Kca<br />

**<br />

**<br />

OT-II P14<br />

Il6 mRNA<br />

(relative expression)<br />

g<br />

Live cells (%)<br />

60<br />

40<br />

20<br />

0<br />

2<br />

1<br />

0<br />

**<br />

Foxo3 Kca<br />

DC OT-II P14<br />

0 0.4 4 40 α-IL-6<br />

rIL-6 (ng/ml)<br />

h<br />

T cells × 10 6<br />

OT-II<br />

P14<br />

CFSE<br />

Isotype<br />

gp33-specific<br />

CD8<br />

40<br />

+ T cells<br />

Anti-IL-6<br />

0<br />

0<br />

IgG1 α-IL-6R IgG1 α-IL-6R<br />

Foxo3-deficient DCs (Fig. 7a). Although treatment with CTLA-4–Ig<br />

resulted in a slightly greater proportion of dead cells, its toxic effects<br />

were equivalent in wild-type and Foxo3 Kca splenic DCs showed the<br />

same toxicity (Fig. 7a).<br />

To confirm the involvement of Foxo3 in the inhibition of cytokine<br />

production induced by CTLA-4–Ig, we analyzed the intracellular<br />

localization of Foxo3 after stimulation with loxoribine or CTLA-4–Ig.<br />

Foxo3 remained in the cytosol of unstimulated and loxoribine-treated<br />

BMDCs; however, CTLA-4–Ig stimulation strongly promoted the<br />

nuclear accumulation of Foxo3 (ref. 32; Fig. 7b). Notably, signaling<br />

through B7 was dominant over Toll-like receptor 7 signals, as DCs<br />

treated with both loxoribine and CTLA-4–Ig showed Foxo3 nuclear<br />

localization. These experiments collectively show that CTLA-4–Ig<br />

treatment profoundly suppressed Toll-like receptor–induced production<br />

of IL-6 and TNF in wild-type DCs by promoting the nuclear<br />

accumulation of Foxo3.<br />

If CTLA-4 signals DCs through B7-1 and B7-2, then blocking<br />

CTLA-4 should allow wild-type DCs, like Foxo3-deficient DCs,<br />

to enhance T cell survival. To investigate this issue, we cultured<br />

TCR-transgenic T cells in the presence of wild-type or Foxo3 Kca DCs<br />

and specific peptides with or without the addition of blocking antibody<br />

to CTLA-4. As expected, cultures with Foxo3 Kca BMDCs<br />

included more viable T cells than did cultures with wild-type<br />

BMDCs (Fig. 7c). However, the addition of anti-CTLA-4 essentially<br />

eliminated the difference in T cell accumulation and viability in these<br />

cultures. The results of these experiments are consistent with the idea<br />

that CTLA-4 is involved in promoting the nuclear localization of<br />

30<br />

20<br />

10<br />

ARTICLES<br />

gp61-specific<br />

CD4 + T cells<br />

2<br />

1<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

expression, presented relative to the expression in uninfected wild-type cells, set as 1. (c) Enzyme-linked immunosorbent assay of IL-6 in supernatants<br />

of wild-type OT-II or P14 T cells cultured for 3 d in the presence of wild-type or Foxo3 Kca BMDCs loaded with OVA(323–339) or gp33 peptide. (d) RT-PCR<br />

analysis of Il6 mRNA in BMDCs (DC), OT-II T cells and P14 T cells purified after 2 d of culture as described in c, presented relative to Gapdh mRNA<br />

(encoding glyceraldehyde phosphate dehydrogenase). (e,f) Accumulation (e) anddeath(f) of wild-type OT-II or P14 T cells activated by BMDCs generated<br />

from wild-type or Foxo3 Kca mice pulsed with OVA(323–339) (for OT-II CD4 + T cells) or gp33 peptide (for P14 CD8 + T cells), in the presence of an IgG1<br />

isotype-matched control antibody or IL6-specific blocking antibody (10 mg/ml), assessed after 3 d of culture by CFSE dilution (accumulation) or 7-AAD<br />

staining (death). Numbers adjacent to outlined areas (f) indicate percent dead 7-AAD + CFSE + cells. (g) Death of P14 T cells stimulated by wild-type BMDCs<br />

in the presence of increasing amounts of recombinant IL-6 (rIL-6) or IL-6-specific blocking antibody (a-IL-6), assessed after 3 d of culture by 7-AAD<br />

staining. (h) LCMV-specific T cell responses of Foxo3 Kca mice and their wild-type littermates (n ¼ 4 mice per group) treated on days –1 and +4 with 100 mg<br />

of antibody to IL-6 receptor a-chain (a-IL-6R) or isotype-matched control antibody (IgG1) and infected on day 0 with LCMV, Armstrong strain, and then<br />

analyzed on day +8 by counting of P14 CD8 + T cells and OT-II CD4 + T cells producing IFN-g after restimulation with OVA(323–339) (for OT-II CD4 + Tcells)<br />

or gp33 peptide (for P14 CD8 + T cells). Data are representative of three (a–f) ortwo(g,h) independent experiments (error bars (a–d,g,h), s.e.m.).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 509


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

IL-6 (pg/ml)<br />

0<br />

Loxo<br />

CTLA-4–Ig<br />

b<br />

Media<br />

Loxo<br />

600<br />

400<br />

200<br />

**<br />

– + + +<br />

0 0 20 40<br />

TNF (pg/ml)<br />

Control Ig<br />

600<br />

400<br />

200<br />

0<br />

Loxo – + + +<br />

0<br />

Loxo – + + +<br />

CTLA-4–Ig 0 0 20 40 CTLA-4–Ig 0 0 20 40<br />

CTLA-4–Ig<br />

Foxo3 and the resulting inhibition of inflammatory cytokines<br />

produced by DCs. Published experiments have shown that antibody<br />

specific for CTLA-4 can enhance an immune response in vivo 33,34 ,<br />

although, to our knowledge, this has not been demonstrated before in<br />

an immune response to LCMV. Nonetheless, we sought to determine<br />

whether we could detect Foxo3-dependent enhanced T cell population<br />

expansion. We inoculated mice with 100 mg of a CTLA-4-specific<br />

antibody and, 4 h later, infected the mice with LCMV. We further<br />

inoculated mice with a CTLA-4-specific antibody every other day and<br />

assessed the results on day 8. We did not find enhancement of T cell<br />

accumulation in either wild-type or Foxo3-deficient mice (Supplementary<br />

Fig. 6 online).<br />

DISCUSSION<br />

In this report, we have established an essential function for Foxo3 in<br />

modulating the magnitude of antigen-specific T cell immune responses.<br />

Foxo3-deficient mice showed enhanced T cell proliferation in response<br />

to viral infection, an enhancement that was not T cell intrinsic but<br />

instead depended on the augmented capacity of DCs to sustain T cell<br />

viability. Despite the many potential targets of Foxo3, including<br />

regulators of cell cycle, apoptosis and reactive oxygen detoxification,<br />

Foxo3 loss-of-function mutation alone had no effect on the steady-state<br />

homeostasis, survival or proliferation of T cells. That finding is<br />

consistent with studies showing that Foxo1 is essential for normal<br />

T cell homeostasis and self tolerance 16 (data not shown). Instead, we<br />

found a critical DC-intrinsic function for Foxo3 in the control of T cell<br />

responses. Foxo3 acted ‘downstream’ of CTLA-4-induced signals to<br />

constrainIL-6productionbyDCs.The natural conditions that would<br />

override the CTLA-4-mediated nuclear localization of Foxo3 are not<br />

known; however, a new immunostimulatory cancer treatment regimen<br />

that involves blocking CTLA-4 may act in part by this mechanism 35 .<br />

**<br />

7-AAD + cells (%)<br />

40<br />

30<br />

20<br />

10<br />

Foxo3 Kca<br />

Foxo3 +/+<br />

c<br />

P14 OT-II<br />

Foxo3<br />

Anti-CTLA-4<br />

Anti-CTLA-4<br />

900 900<br />

150<br />

150<br />

Kca<br />

Foxo3 +/+<br />

Foxo3 37<br />

+/+ 66<br />

Foxo3 Kca<br />

Annexin V<br />

CFSE<br />

60 67<br />

7-AAD<br />

Our results differ substantially from a published characterization of<br />

Foxo3 Trap mice 20 . Foxo3 Trap mutant mice were derived from an<br />

independent insertional mutant embryonic stem line obtained from<br />

BayGenomics and were backcrossed to the 129 strain for three<br />

generations. Foxo3 Trap mice show spontaneous T cell activation,<br />

lymphoproliferation and organ infiltration associated with constitutive<br />

activation of NF-kB. In contrast, we found none of those<br />

characteristics in Foxo3 –/– or Foxo3 Kca mice. T cells isolated from<br />

these two strains of mice seemed to be indistinguishable from wildtype<br />

T cells in terms of NF-kB activation, expression of activation<br />

markers and response to mitogenic stimulation. Our results are<br />

consistent with published analyses of Foxo3 Kca and Foxo3 –/– mice in<br />

that extensive histological analysis has not shown lymphocytic infiltration<br />

of organs 23,26 .<br />

Differences in autoimmunity are not uncommon in mixed<br />

C57BL/6 and 129 mouse strains. The 129 mice have autoimmunitysusceptibility<br />

loci, one of which (Sle16) induces humoral autoimmunity<br />

in congenic C57BL/6 mice. Several studies have noted modifier<br />

genes in 129 strains that confer enhanced autoimmune disease to 129<br />

C57BL/6 mice 36–38 . Foxo Trap mice are apparently inbred 129 mice<br />

without a contribution from the C57BL/6 strain, but perhaps a 129<br />

susceptibility locus that modifies the Foxo3 deficiency is present.<br />

IL-6 is a pleiotropic cytokine that regulates many aspects of the<br />

immune system, including antibody production, hematopoiesis,<br />

inflammation and, most relevant to our study, T cell survival 39 .IL-6<br />

‘rescues’ resting T cells from apoptosis by inhibiting the downregulation<br />

of Bcl-2 in a dose-dependent way 40 ,andIL-6increasesthesurvival<br />

of antigen-stimulated T cells 31 . Consistent with that observation, we<br />

noted higher expression of Bcl-2 and Bcl-x L in T cells from LCMVinfected<br />

Foxo3 Kca mice than in those from wild-type mice. In addition,<br />

the division rate of T cells stimulated with Foxo3 Kca DCs was not<br />

43<br />

61<br />

73 75<br />

Figure 7 Production of IL-6 and TNF by stimulated DCs is inhibited by CTLA-4–Ig stimulation in a Foxo3-dependent way. (a) Immunoassay of the<br />

concentration of IL-6 (left) and TNF (middle) in supernatants of DCs (2 10 5 ) purified from the spleens of Foxo3 Kca mice and their wild-type littermates<br />

and stimulated for 18 h with (+) or without (–) loxoribine (Loxo) plus increasing amounts of CTLA-4–Ig (below graphs, in mg/ml). Right, DC viability, assessed<br />

by 7-AAD staining. (b) Immunofluorescence analysis of Foxo3 localization (green) after 18 h of stimulation of BMDCs from wild-type and Foxo3 Kca mice with<br />

loxoribine (bottom) or without loxoribine (Media; top) in presence of control immunoglobulin or CTLA-4–Ig. Blue, DAPI staining of nuclei. Original<br />

magnification, 63. (c) Accumulation (top) and death (below) of OT-II CD4 + T cells or P14 CD8 + Tcells(1 10 5 ) cultured for 3 d with wild-type or<br />

Foxo3 Kca BMDCs (3.3 10 2 ) in the presence of the appropriate peptide plus anti-CTLA-4 (50 mg/ml; 9D9), assessed as CFSE dilution (accumulation) and<br />

staining with annexin V and 7-AAD (death). Numbers in plots indicate number of accumulated cells (top) or percent live OT-II CD4 + T cells or P14 CD8 +<br />

T cells (middle and bottom). Data are representative of at least two independent experiments (error bars (a), s.e.m.).<br />

510 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

different from that of T cells stimulated with wild-type DCs; instead,<br />

the effect was manifested as improved T cell survival. Moreover,<br />

Foxo3 Kca DCs sustained T cell viability both in vitro and in vivo even<br />

in the absence of specific peptide, which suggests that the enhanced<br />

T cell survival induced by IL-6 is not restricted to TCR-engaged T cells.<br />

Notably, IL-6-specific blocking antibody did not affect the viability of<br />

wild-type T cells in vitro, nor did it result in lower T cell accumulation<br />

in vivo. That observation indicates that in the conditions tested, the<br />

amount of IL-6 produced by wild-type DCs was not sufficient to affect<br />

T cell viability. It would be useful to determine whether immune<br />

responses to other agents are sensitive to IL-6 concentration and, in<br />

those infections, whether CTLA-4, regulatory T cells and Foxo3 affect<br />

the magnitude and dynamics of the T cell response.<br />

CTLA-4 acts as a counterbalance to costimulation of CD28 by<br />

antagonizing TCR signals in activated T cells and by promoting<br />

tolerance 41 . CTLA-4 is also expressed on regulatory T cells and is<br />

essential for their effectiveness 42 . CTLA-4 can also modify function of<br />

cells of the immune system by triggering ‘reverse signaling’ through B7<br />

receptors; these signals promote the DC production of stimulatory<br />

and/or suppressive mediators 21 . Reverse signals through B7 also<br />

activate the immunosuppressive pathway of tryptophan catabolism<br />

in plasmacytoid DCs involving indoleamine 2,3-dioxygenase (IDO) 43 .<br />

Foxo3 is involved in the IDO pathway in plasmacytoid DCs, as CTLA-<br />

4–Ig leads to nuclear accumulation of Foxo3 and expression of<br />

superoxide dismutase 2 (ref. 32). As IDO induces T cell death through<br />

the generation of kinurenin and tryptophan privation 43 ,wespeculated<br />

that a potential IDO defect in Foxo3 Kca DCs could explain the<br />

improved T cell survival; however, addition of the IDO inhibitor<br />

1-MT did not enhance T cell survival in cultures containing wild-type<br />

or Foxo3 Kca BMDCs (data not shown). Moreover, IDO is expressed<br />

mainly by plasmacytoid and CD8a + DC subsets 22 ,andlowornoIDO<br />

protein is detected in resting DCs derived from bone marrow 44 .<br />

Presumably CTLA-4 expressed on regulatory T cells constitutes the<br />

main B7-stimulatory ligand, but the possibility that activated T cells or<br />

other cells contribute to B7 ligation in vivo has not been investigated.<br />

In addition, the relative contribution of each CTLA-4-mediated<br />

suppressive mechanism to the regulation of immune responses to<br />

various infectious agents is not well understood.<br />

Our experiments collectively indicate that nucleus-localized Foxo3<br />

functions in DCs to inhibit the production of inflammatory cytokines<br />

that would otherwise be activated in response to infection. We deduce<br />

that the IL-6-induced T cell survival in Foxo3 Kca mice resulted from an<br />

inability of Foxo3-deficient DCs to shut down cytokine production in<br />

response to CTLA-4 signals. Conditions that alter Foxo3 activity, such<br />

as stress, growth factors or nutrients, would be expected to substantially<br />

influence the activity of DCs and perhaps macrophages, and such<br />

conditions might be expected to alter the outcome of responses to<br />

infectious agents.<br />

METHODS<br />

Mice. C57BL/6, C57BL/6-CD45.1, P14, OT-I and OT-II mice were maintained<br />

in specific pathogen–free conditions at the University of California, San Diego.<br />

C57BL/6 mice lacking Foxo3 were generated from embryonic stem cell clones<br />

from the OmniBank embryonic stem cell library of randomly targeted cell lines<br />

(Lexicon Genetics) 23 . These mice (Foxo3 Gt(VICTR20)1Kca ; called ‘Foxo3 Kca ’ here)<br />

were backcrossed to the C57BL/6J strain for 12 generations and were then<br />

intercrossed to generate congenic C57BL/6 Foxo3 Kca mice. ‘Foxo L/L ’mice,with<br />

conditional targeting of Foxo3, were produced in FVB embryonic stem cells, and<br />

the loxP-flanked allele was deleted by breeding with male EIIa-Cre–transgenic<br />

mice 26 . Offspring with complete excision were bred to exclude EIIa-Cre; these<br />

mice are called ‘Foxo3 –/– ’ here 26 . TCR-transgenic mice with a Foxo3-null<br />

genotype were produced by breeding of C57BL/6 Foxo3 Kca mice with either<br />

ARTICLES<br />

P14 or OT-II mice. All procedures were approved by the Institutional Animal<br />

Care and Use Committee of the University of California, San Diego.<br />

Virus infection and analysis. Foxo3 Kca mice and their wild-type littermates<br />

6–12 weeks of age were infected by intraperitoneal injection of<br />

2 10 5 plaque-forming units of LCMV, Armstrong strain, in 0.2 ml PBS.<br />

Alternatively, mice were infected with 1 10 5 plaque-forming units of vesicular<br />

stomatitis virus 45 expressing OVA (provided by L. Lefrançois). At various times,<br />

spleens were collected from LCMV-infected mice and uninfected control mice<br />

and splenocytes were stimulated with gp33 peptide, an MHC class I–restricted<br />

LCMV epitope (gp33-41), or gp61 peptide, an MHC class II–restricted LCMV<br />

epitope (gp61-80; Genemed Synthesis) and brefeldin A (1 mg/ml; Golgistop; BD<br />

PharMingen). After 5 h of stimulation, cells were stained for intracellular IFN-g<br />

(XMG1.2; eBioscience). For transfer experiments, CD45.2 congenic wild-type<br />

P14 or Foxo3 Kca P14 T cells (2 10 4 ) were injected intravenously into CD45.1<br />

wild-type hosts, which were infected with LCMV 24 h later. Conversely, CD45.1<br />

congenic wild-type P14 T cells were transferred into congenic CD45.2 wild-type<br />

littermate or Foxo3 Kca mice. P14 T cells in the spleen were counted at day 8 after<br />

LCMV infection. Infection with vesicular stomatitis virus was done in an<br />

identical way except that T cells from CD45.2 OT-I mice were transferred. For<br />

bone marrow–chimera experiments, congenic CD45.1 recipient mice were<br />

lethally irradiated (1,200 rads) and were injected intravenously with 2 10 6<br />

bone marrow cells from Foxo3 Kca or wild-type littermate donor mice (both<br />

CD45.2). After 2 months, the reconstitution of peripheral blood lymphocytes<br />

was over 90%. These radiation chimeras were then infected with LCMV and the<br />

magnitude of the LCMV response was assessed at day 8 after infection. For<br />

antibody blockade of IL-6 receptor a-chain, wild-type or Foxo3-deficient mice<br />

were inoculated intraperitoneally with 100 mg of antibody to IL-6 receptor<br />

a-chain (purified antibody to mouse and rat CD126; D7715A7; BioLegend) on<br />

day –1 and day +4 and were infected with LCMV, Armstrong strain, on day 0.<br />

Cell isolation and purification. Bone marrow cells were isolated by flushing of<br />

the femur and tibia with RPMI medium (containing 10% (vol/vol) FBS, 2 mM<br />

L-glutamine, 100 U/ml of penicillin, 100 mg/ml of streptomycin and 50 mM<br />

b-mercaptoethanol). Spleens were removed and were incubated for 20 min at<br />

37 1C with collagenase D (1 mg/ml; Roche), and splenocytes were collected by<br />

homogenization through a 100-mm tissue strainer. Cells were resuspended in<br />

Tris–ammonium chloride buffer for lysis of red blood cells. For purification<br />

of splenic DCs, cells were first incubated for 15 min with supernatants of<br />

2.4G2 hybridoma cultures (anti–mouse FcgRII-III), then were incubated for<br />

20 min with anti–mouse ‘pan-DC’ microbeads (Miltenyi Biotec), followed<br />

by positive selection. The positive fraction was typically over 95% CD11c + .<br />

For T cell purification, splenocytes from P14 or OT-II mice were incubated<br />

with a mixture of biotinylated anti-B220 (RA3-6B2), anti-CD19, anti-Gr-1<br />

(RB6-8C5), anti–MHC class II, anti-DX5 and anti-CD11b (M1/70; all from<br />

eBioscience) and were negatively selected to over 95% purity with strepavidincoupled<br />

microbeads (Miltenyi Biotec).<br />

BMDC cultures. Bone marrow cells were isolated and were cultured at a<br />

density of 2 10 6 cells/ml in RPMI medium containing recombinant mouse<br />

granulocyte-macrophage colony-stimulating factor (20 ng/ml; Peprotech). On<br />

days 2, 4 and 6, half the culture was removed and centrifuged and the cell pellet<br />

was resuspended in 10 ml fresh media containing recombinant mouse<br />

granulocyte-macrophage colony-stimulating factor (20 ng/ml). Cultures were<br />

collected on day 8 and CD11c + cells were selected with microbeads (Miltenyi<br />

Biotec). The final population was 98% CD11c + CD11b + B220 – DCs.<br />

Flow cytometry. Cells were incubated for 15 min with anti–mouse FcgRII-III<br />

and then were stained with primary antibodies for 20 min on ice. The<br />

following mouse monoclonal antibodies were used (all from eBioscience):<br />

allophycocyanin- or phycoerythrin-conjugated anti-CD11c (N418); fluorescein<br />

isothiocyanate– or phycoerythrin-conjugated anti-B220 (RA3-6B2);<br />

phycoerythrin-indotricarbocyanine– or peridin chlorophyll protein–conjugated<br />

anti-CD11b (M1/70); allophycocyanin-conjugated anti-CD8a (53-6.7); fluorescein<br />

isothiocyanate–conjugated anti–MHC class I and anti-Gr-1 (RB6-8C5);<br />

peridin chlorophyll protein–conjugated anti-CD3; and phycoerythrinconjugated<br />

anti-CD86, anti-CD80, anti–MHC class II, anti-Ly6C, anti-CD19,<br />

anti-CD3 and anti-NK1.1.<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 511


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

T cell proliferation and death assays. For in vitro experiments, DCs were<br />

purified from wild-type or Foxo3 Kca mice on day 3 after LCMV infection. DCs<br />

were also purified from uninfected mice or were derived from the bone marrow<br />

of wild-type or Foxo3 Kca mice. These DCs were used to stimulated P14 CD8 +<br />

T cells, OT-II CD4 + T cells or OT-I CD8 + T cells (1 10 5 cells per well) labeled<br />

with 1 mM CFSE (carboxyfluorescein diacetate succinimidyl ester; Molecular<br />

Probes) in the presence of gp33 peptide (0.1 mg/ml) or OVA(323–339);<br />

(1 mg/ml). Proliferation was analyzed by flow cytometry by measurement of<br />

CFSE dilution, and T cell death was measured by staining with annexin V and<br />

7-AAD. For in vivo transfer experiments, CFSE-labeled CD4 + T cells purified<br />

from OT-II mice were injected retro-orbitally into congenic CD45.1 C57BL/6<br />

mice. Wild-type or Foxo3 Kca BMDCs were pulsed with OVA(323–339) and were<br />

injected intradermally after 24 h. Then, 3 d later, draining lymph nodes<br />

were collected, were stained with allophycocyanin-conjugated anti-CD4<br />

(GK1.5) and phycoerythrin-conjugated anti-CD45.1 (A20; both from<br />

eBioscience) and were analyzed by flow cytometry.<br />

In vitro stimulation of DCs and quantification of cytokine production.<br />

Sorted DCs (2 10 5 ) were cultured in 96-well round-bottomed culture<br />

plates (Nunc) and were stimulated with medium alone or 0.03 mM<br />

loxoribine (InvivoGen) alone or in the presence of a recombinant mouse<br />

fusion protein consisting of the ectodomain of CTLA-4 linked to the<br />

Fc portion of IgG2a (CTLA-4–Ig) at a concentration of 20 or 40 mg/ml<br />

(R&D Systems). After culture for 18 h, supernatants were collected and<br />

cytokine concentrations were determined by immunoassay. Enzyme-linked<br />

immunosorbent assay kits were used for the detection of TNF or IL-6<br />

(Ready-Set-Go; eBioscience) unless otherwise specified. Cytokines were also<br />

measured by cytokine bead array (BD Biosciences) or Luminex technology<br />

(Invitrogen).<br />

Fluorescence microscopy. DCs derived from bone marrow of wild-type or<br />

Foxo3 Kca mice were cultivated for 12 h on sterile poly-L-lysine-coated coverslips<br />

(BioCoat; BD). Cells were fixed in 4% (vol/vol) formaldehyde, then were made<br />

permeable with 0.02% (vol/vol) Triton X-100. Coverslips were blocked for 2 h<br />

in PBS containing 5% (wt/vol) BSA and anti–mouse FcgRII-III. Anti-Foxo3<br />

(FKHRL1; Cell Signaling) was added for 1 h, followed by a secondary antibody<br />

conjugated to fluorescein isothiocyanate (goat anti–rabbit IgG; 65-6111;<br />

Zymex). For staining of nuclei, DAPI (4¢-6-diamidino-2-phenylindole) was<br />

added at a concentration of 0.04 mg/ml. All cells were visualized with a<br />

microscope (Axiovert 200M; Carl Zeiss MicroImaging) with a 63 objective.<br />

Images were captured with an Axiocam monochrome digital camera and were<br />

analyzed with Axiovision software (Carl Zeiss MicroImaging).<br />

Semiquantitative RT –PCR. Total RNA was extracted from DCs with Trizol<br />

reagent according to the manufacturer’s instructions (Invitrogen), then cDNA<br />

was synthesized with SuperScript First-Strand synthesis System for RT-PCR<br />

(Invitrogen). Il6 and the gene encoding cyclophilin A (Ppia) wereamplifiedby<br />

PCR with the following primers: Il6 forward, 5¢-ACCTGGAGTACATGAAGAA<br />

CAACTT-3¢ and reverse, 5¢-GGAAGCACTCACCTCTTGGT-3¢; Ppia forward,<br />

5¢-CACCGTGTTCTTCGACATC-3¢ and reverse, 5¢-ATTCTGTGAAAGGAG<br />

GAACC-3¢.<br />

Immunoblot analysis. Equal amounts of protein from whole-cell extracts were<br />

resolved by 4–12% SDS-PAGE (Invitrogen) and were transferred to a polyvinylidene<br />

difluoride membrane (Millipore) by semidry transfer (Bio-Rad).<br />

Blots were blocked and incubated with primary antibody overnight at 4 1C,<br />

followed by 2 h of incubation at 25 1C with the appropriate horseradish<br />

peroxidase–conjugated secondary antibody. Rabbit polyclonal anti-Foxo3a was<br />

provided by A. Brunet (Stanford University). Antibodies specific for IkB<br />

proteins were from Santa Cruz Biotechnology.<br />

Accession codes. UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signalinggateway.org):<br />

A000945 and A000706.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank J.P. Allison (Sloan-Kettering Memorial Hospital) and J.A. Bluestone<br />

(University of California at San Diego) for CTLA-4-specific blocking antibodies;<br />

A. Brunet (Stanford University) for Foxo3-specific antibody; L. Lefrançois<br />

(University of Connecticut Health Center) for OVA-expressing vesicular<br />

stomatitis virus; L. Mack and E. Zuniga for assistance with virus titers; and<br />

M. Niwa for microscope facility use. Supported by the American Cancer Society<br />

(R.A.D.), the Robert A. and Renee E. Belfer Institute for Innovative Cancer<br />

Research (R.A.D.), the US National Cancer Institute (R.A.D.), the Division of<br />

Biological Sciences of the University of California, San Diego (S.M.H.) and the<br />

Fondation pour la Recherche Médicale (A.S.D.).<br />

AUTHOR CONTRIBUTIONS<br />

D.R.B. initiated the project and did lymphocyte characterization and LCMV<br />

infection under the supervision of S.M.H.; A.S.D. designed and did the<br />

remaining experiments in collaboration with D.R.B., I.L.C., Y.M.K. and S.M.H.;<br />

A.B. provided expertise in fluorescence microscopy analysis; R.A.D. and D.H.C.<br />

produced Foxo3 –/– mice and provided intellectual input on the data; K.C.A.<br />

provided Foxo3 Kca mice; S.M.H. initiated the project with input from R.A.D.<br />

and K.C.A. and supervised the experiments; and A.S.D. and S.M.H. wrote the<br />

manuscript with editorial and intellectual contributions from the other authors.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Calnan, D.R. & Brunet, A. The FoxO code. Oncogene 27, 2276–2288 (2008).<br />

2. Galili, N. et al. Fusion of a fork head domain gene to PAX3 in the solid tumour alveolar<br />

rhabdomyosarcoma. Nat. Genet. 5, 230–235 (1993).<br />

3. Hillion, J., Le Coniat, M., Jonveaux, P., Berger, R. & Bernard, O.A. AF6q21, a novel<br />

partner of the MLL gene in t(6;11)(q21;q23), defines a forkhead transcriptional factor<br />

subfamily. Blood 90, 3714–3719 (1997).<br />

4. Corral, J. et al. Acute leukemias of different lineages have similar MLL gene fusions<br />

encoding related chimeric proteins resulting from chromosomal translocation. Proc.<br />

Natl. Acad. Sci. USA 90, 8538–8542 (1993).<br />

5. Jacobs, F.M. et al. FoxO6, a novel member of the FoxO class of transcription factors<br />

with distinct shuttling dynamics. J. Biol. Chem. 278, 35959–35967 (2003).<br />

6. Barthel, A., Schmoll, D. & Unterman, T.G. FoxO proteins in insulin action and<br />

metabolism. Trends Endocrinol. Metab. 16, 183–189 (2005).<br />

7. Biggs, W.H. III, Meisenhelder, J., Hunter, T., Cavenee, W.K. & Arden, K. C. Protein<br />

kinase B/Akt-mediated phosphorylation promotes nuclear exclusion of the winged helix<br />

transcription factor FKHR1. Proc. Natl. Acad. Sci. USA 96, 7421–7426 (1999).<br />

8. Brunet, A. et al. Akt promotes cell survival by phosphorylating and inhibiting a Forkhead<br />

transcription factor. Cell 96, 857–868 (1999).<br />

9. Kops, G.J. et al. Direct control of the Forkhead transcription factor AFX by protein<br />

kinase B. <strong>Nature</strong> 398, 630–634 (1999).<br />

10. van der Horst, A. & Burgering, B.M. Stressing the role of FoxO proteins in lifespan and<br />

disease. Nat. Rev. Mol. Cell Biol. 8, 440–450 (2007).<br />

11. Brunet, A. et al. Stress-dependent regulation of FOXO transcription factors by the<br />

SIRT1 deacetylase. Science 303, 2011–2015 (2004).<br />

12. Motta, M.C. et al. Mammalian SIRT1 represses forkhead transcription factors. Cell<br />

116, 551–563 (2004).<br />

13. Stahl, M. et al. The forkhead transcription factor FoxO regulates transcription of<br />

p27Kip1 and Bim in response to IL-2. J. Immunol. 168, 5024–5031 (2002).<br />

14. Dijkers, P.F. et al. Forkhead transcription factor FKHR-L1 modulates cytokine-dependent<br />

transcriptional regulation of p27 KIP1 . Mol. Cell. Biol. 20, 9138–9148 (2000).<br />

15. Barata, J.T. et al. Activation of PI3K is indispensable for interleukin 7-mediated<br />

viability, proliferation, glucose use, and growth of T cell acute lymphoblastic leukemia<br />

cells. J. Exp. Med. 200, 659–669 (2004).<br />

16. Kerdiles, Y.M. et al. Foxo1 links homing and survival of naive T cells by regulating<br />

L-selectin, CCR7 and interleukin 7 receptor. Nat. Immunol. 10, 176–184 (2009).<br />

17. Peng, S.L. Foxo in the immune system. Oncogene 27, 2337–2344 (2008).<br />

18. You, H. et al. FOXO3a-dependent regulation of Puma in response to cytokine/growth<br />

factor withdrawal. J. Exp. Med. 203, 1657–1663 (2006).<br />

19. Riou, C. et al. Convergence of TCR and cytokine signaling leads to FOXO3a phosphorylation<br />

and drives the survival of CD4 + central memory T cells. J. Exp. Med. 204,<br />

79–91 (2007).<br />

20. Lin, L., Hron, J.D. & Peng, S.L. Regulation of NF-kB, Th activation, and autoinflammation<br />

by the forkhead transcription factor Foxo3a. Immunity 21, 203–213 (2004).<br />

21. Orabona, C. et al. CD28 induces immunostimulatory signals in dendritic cells via CD80<br />

and CD86. Nat. Immunol. 5, 1134–1142 (2004).<br />

22. Mellor, A.L. & Munn, D.H. IDO expression by dendritic cells: tolerance and tryptophan<br />

catabolism. Nat. Rev. Immunol. 4, 762–774 (2004).<br />

23. Hosaka, T. et al. Disruption of forkhead transcription factor (FOXO) family members in<br />

mice reveals their functional diversification. Proc. Natl. Acad. Sci. USA 101,<br />

2975–2980 (2004).<br />

24. Marinkovic, D. et al. Foxo3 is required for the regulation of oxidative stress in<br />

erythropoiesis. J. Clin. Invest. 117, 2133–2144 (2007).<br />

25. Greer, E.L. & Brunet, A. FOXO transcription factors at the interface between longevity<br />

and tumor suppression. Oncogene 24, 7410–7425 (2005).<br />

26. Castrillon, D.H., Miao, L., Kollipara, R., Horner, J.W. & DePinho, R.A. Suppression of<br />

ovarian follicle activation in mice by the transcription factor Foxo3a. Science 301,<br />

215–218 (2003).<br />

512 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

27. Villadangos, J.A. & Schnorrer, P. Intrinsic and cooperative antigen-presenting functions<br />

of dendritic-cell subsets in vivo. Nat. Rev. Immunol. 7, 543–555 (2007).<br />

28. Oldstone, M.B. Biology and pathogenesis of lymphocytic choriomeningitis virus infection.<br />

Curr. Top. Microbiol. Immunol. 263, 83–117 (2002).<br />

29. Rosas, M. et al. IL-5-mediated eosinophil survival requires inhibition of GSK-3 and<br />

correlates with b-catenin relocalization. J. Leukoc. Biol. 80, 186–195 (2006).<br />

30. Korn, T., Bettelli, E., Oukka, M. & Kuchroo, V.K. IL-17 and Th17 cells. Annu. Rev.<br />

Immunol. 27, 485–517 (2009).<br />

31. Rochman, I., Paul, W.E. & Ben-Sasson, S.Z. IL-6 increases primed cell expansion and<br />

survival. J. Immunol. 174, 4761–4767 (2005).<br />

32. Fallarino, F. et al. CTLA-4-Ig activates forkhead transcription factors and protects<br />

dendritic cells from oxidative stress in nonobese diabetic mice. J. Exp. Med. 200,<br />

1051–1062 (2004).<br />

33. Leach, D.R., Krummel, M.F. & Allison, J.P. Enhancement of antitumor immunity by<br />

CTLA-4 blockade. Science 271, 1734–1736 (1996).<br />

34. Karandikar, N.J., Vanderlugt, C.L., Walunas, T.L., Miller, S.D. & Bluestone, J.A. CTLA-4:<br />

a negative regulator of autoimmune disease. J. Exp. Med. 184, 783–788 (1996).<br />

35. Zang, X. & Allison, J.P. The B7 family and cancer therapy: costimulation and<br />

coinhibition. Clin. Cancer Res. 13, 5271–5279 (2007).<br />

36. Botto, M. et al. Homozygous C1q deficiency causes glomerulonephritis associated with<br />

multiple apoptotic bodies. Nat. Genet. 19, 56–59 (1998).<br />

ARTICLES<br />

37. Bickerstaff, M.C. et al. Serum amyloid P component controls chromatin<br />

degradation and prevents antinuclear autoimmunity. Nat. Med. 5, 694–697<br />

(1999).<br />

38. Santiago-Raber, M.L. et al. Role of cyclin kinase inhibitor p21 in systemic autoimmunity.<br />

J. Immunol. 167, 4067–4074 (2001).<br />

39. Kishimoto, T. Interleukin-6: from basic science to medicine–40 years in immunology.<br />

Annu. Rev. Immunol. 23, 1–21(2005).<br />

40. Teague, T.K., Marrack, P., Kappler, J.W. & Vella, A.T. IL-6 rescues resting mouse T cells<br />

from apoptosis. J. Immunol. 158, 5791–5796 (1997).<br />

41. Keir, M.E. & Sharpe, A.H. The B7/CD28 costimulatory family in autoimmunity.<br />

Immunol. Rev. 204, 128–143 (2005).<br />

42. Wing, K. et al. CTLA-4 control over Foxp3 + regulatory T cell function. Science 322,<br />

271–275 (2008).<br />

43. Grohmann, U. et al. CTLA-4-Ig regulates tryptophan catabolism in vivo. Nat. Immunol.<br />

3, 1097–1101 (2002).<br />

44. Hara, T. et al. High-affinity uptake of kynurenine and nitric oxide-mediated inhibition of<br />

indoleamine 2,3-dioxygenase in bone marrow-derived myeloid dendritic cells. Immunol.<br />

Lett. 116, 95–102 (2008).<br />

45. van Stipdonk, M.J., Lemmens, E.E. & Schoenberger, S.P. Naive CTLs require a single<br />

brief period of antigenic stimulation for clonal expansion and differentiation. Nat.<br />

Immunol. 2, 423–429 (2001).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 513


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

C-C chemokine receptor 6–regulated entry of TH-17<br />

cells into the CNS through the choroid plexus is<br />

required for the initiation of EAE<br />

Andrea Reboldi 1 , Caroline Coisne 2 , Dirk Baumjohann 1 , Federica Benvenuto 3,4 , Denise Bottinelli 1 , Sergio Lira 5 ,<br />

Antonio Uccelli 3,4,6 , Antonio Lanzavecchia 1 , Britta Engelhardt 2 & Federica Sallusto 1<br />

Interleukin 17–producing T helper cells (TH-17 cells) are important in experimental autoimmune encephalomyelitis, but their<br />

route of entry into the central nervous system (CNS) and their contribution relative to that of other effector T cells remain to<br />

be determined. Here we found that mice lacking CCR6, a chemokine receptor characteristic of T H-17 cells, developed T H-17<br />

responses but were highly resistant to the induction of experimental autoimmune encephalomyelitis. Disease susceptibility<br />

was reconstituted by transfer of wild-type T cells that entered into the CNS before disease onset and triggered massive<br />

CCR6-independent recruitment of effector T cells across activated parenchymal vessels. The CCR6 ligand CCL20 was<br />

constitutively expressed in epithelial cells of choroid plexus in mice and humans. Our results identify distinct molecular<br />

requirements and ports of lymphocyte entry into uninflamed versus inflamed CNS and suggest that the CCR6-CCL20<br />

axis in the choroid plexus controls immune surveillance of the CNS.<br />

The identification of specialized subsets of effector CD4 + T cells has<br />

provided a paradigm for understanding immunity and immunopathology.<br />

Interleukin 17 (IL-17)-producing T cells (T H-17 cells)<br />

have been characterized in the mouse as a distinct lineage of<br />

CD4 + T cells that can differentiate from uncommitted naive T cell<br />

precursors under the aegis of the transcription factors RORgt and<br />

RORa and the polarizing cytokines transforming growth factor-b,<br />

IL-6 and IL-23 (ref. 1–3). IL-17 can mediate protection against<br />

extracellular pathogens by promoting neutrophil recruitment but<br />

has also been shown to cause immunopathology in various models<br />

of autoimmunity 4,5 .<br />

Experimental autoimmune encephalomyelitis (EAE) is a CD4 +<br />

T cell–mediated disease of the central nervous system (CNS) that is<br />

used as a model of multiple sclerosis, a devastating inflammatory<br />

demyelinating disease of the human CNS. Several lines of evidence<br />

indicate that T H-17 cells are involved in the onset and maintenance of<br />

EAE 6 . Thus, mice lacking RORgt, IL-17 or IL-23 as well as mice<br />

treated with IL-17-blocking antibodies are less susceptible to EAE than<br />

are wild-type or untreated mice 7–10 . In addition, IL-17 + T cells have<br />

been found in lesions in brain tissues from patients with multiple<br />

sclerosis 11 . However, it has also been shown that T helper type 1 (T H1)<br />

cells are present in lesions of EAE and in multiple sclerosis during the<br />

active phase of the disease and that mice lacking T-bet, the TH1<br />

‘master’ transcription factor, are resistant to the development of<br />

Received 4 November 2008; accepted 10 February 2009; published online 22 March 2009; doi:10.1038/ni.1716<br />

EAE 12 . EAE can be induced by transfer of either T H-17 or T H1<br />

cells 13,14 and the T H-17/T H1 ratio of infiltrating cells determines<br />

where inflammation occurs in the CNS 15,16 . Together these studies<br />

suggest the possibility that TH-17 and TH1 cells may be involved in<br />

pathogenesis at different times or at different sites.<br />

In the model of active EAE, autoreactive myelin-specific effector<br />

T cells are primed in peripheral lymph nodes and must migrate into<br />

uninflamed CNS to initiate tissue inflammation. The molecular<br />

determinants that control this initial step of cell migration, which is<br />

probably the same used for constitutive immune surveillance in the<br />

brain, remain to be determined. In contrast, the molecular requirements<br />

for lymphocyte rolling and adhesion to activated vessels of the<br />

inflamed blood-brain barrier have been intensively investigated in<br />

both active and passive models of EAE 17 . The integrin a4b1 (VLA-4)<br />

serves a key function in controlling the entry of lymphocytes into<br />

the CNS by interacting with the adhesion molecule VCAM expressed<br />

by inflamed endothelial cells 18 . P-selectin does not seem to be<br />

necessary for the recruitment of inflammatory cells during active<br />

EAE 19 but can be expressed in small amounts on resting brain<br />

endothelium or can be rapidly induced on endothelial cells by<br />

inflammatory stimuli 20,21 . Chemokine receptors are required for the<br />

entry of lymphocytes into the CNS 21,22 ,butthenatureofthereceptors<br />

has not been identified and may vary depending on the inflammatory<br />

conditionorlocation.<br />

1 Institute for Research in Biomedicine, Bellinzona, Switzerland. 2 Theodor Kocher Institute, University of Bern, Bern, Switzerland. 3 Neuroimmunology Unit and 4 Center of<br />

Excellence for Biomedical Research, University of Genoa, Genoa, Italy. 5 Mount Sinai School of Medicine, New York, New York, USA. 6 Advanced Biotechnology Center,<br />

Genoa, Italy. Correspondence should be addressed to F.S. (federica.sallusto@irb.unisi.ch).<br />

514 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Lymphocytes can also enter into the CNS<br />

through the choroid plexus 23,24 . These<br />

plexuses are villous structures extending into<br />

the cerebrospinal fluid–filled ventricular<br />

spaces that establish a blood–cerebrospinal<br />

fluid barrier at the level of apical tight junctions<br />

between epithelial cells of the choroid<br />

plexus. The specific enrichment for memory<br />

T cells in the cerebrospinal fluid of healthy<br />

people 24 and patients with multiple sclerosis<br />

25 , as well as the regulated functional<br />

expression of adhesion molecules on the<br />

choroid plexus epithelium 26 , suggest that<br />

T cells may use these structures to enter the<br />

cerebrospinal fluid and disseminate to the<br />

meningeal and perivascular spaces 27 . However,<br />

the homing determinants that regulate<br />

the entry of lymphocytes through the choroid<br />

plexus and the function of this port of entry<br />

relative to that of the blood-brain barrier<br />

remain to be defined.<br />

It is well established that both in humans<br />

and mice, chemokine receptors have differences<br />

in expression on subsets of effector and<br />

memory T cells and provide specificity to cell<br />

trafficking both in the steady state and<br />

inflammation 28 . For example, CCR7 endows<br />

naive and central memory T cells with the<br />

ability to migrate into peripheral lymph<br />

nodes, whereas CCR9 and CCR4 direct the<br />

migration of memory T cells into the gut and<br />

skin, respectively. In addition, some receptors,<br />

such as CXCR3 and CCR5, are ‘preferentially’<br />

expressed on T H1 cells, whereas<br />

others, such as CCR3, CCR8 and CRTH2,<br />

are ‘preferentially’ expressed on T H2 cells. The finding that in humans,<br />

CCR6 (A000629), the receptor for CCL20 (a chemokine expressed in<br />

the liver, lungs and Peyer’s patches) 29 , is expressed on IL-17-producing<br />

T cells (including some that also produce interferon-g (IFN-g)) 30<br />

prompted us to investigate the function of CCR6 in regulating T H-<br />

17-mediated immune pathology.<br />

Here we report that CCR6-deficient (CCR6-knockout) mice were<br />

highly resistant to EAE induction but became susceptible when given<br />

transfer of small numbers of CCR6-sufficient T cells. CCR6 was<br />

required on the first wave of TH-17 cells that entered the CNS through<br />

epithelial cells of the choroid plexus, which constitutively expressed<br />

CCL20 in both mice and humans. CCR6 + T cells triggered the entry of<br />

a second wave of T cells that migrated in large numbers into the CNS<br />

by crossing activated parenchymal vessels. Our results demonstrate<br />

distinct molecular requirements and anatomical sites for lymphocyte<br />

entry during the development of EAE and suggest that the CCR6-<br />

CCL20 axis controls an evolutionary conserved pathway of immune<br />

surveillance in the brain.<br />

RESULTS<br />

CCR6-knockout mice are highly resistant to EAE<br />

To study the function of CCR6 in T H-17-mediated immunopathology,<br />

we compared the susceptibility of wild-type and CCR6-knockout<br />

mice 31 to EAE induction. When immunized by subcutaneous injection<br />

of a peptide consisting of amino acids 35–55 of myelin oligodendrocyte<br />

glycoprotein (MOG(35–55)) in complete Freund’s<br />

a<br />

c<br />

Clinical score<br />

* * *<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

WT<br />

CCR6-KO<br />

1.0<br />

0.5<br />

0<br />

* * * * * * * *<br />

0 5 10<br />

Time (d)<br />

15 20<br />

WT<br />

CCR6-KO<br />

20 µm<br />

20 µm<br />

20 µm<br />

20 µm<br />

b<br />

WT<br />

CCR6-KO<br />

d<br />

IL-17 (ng/ml)<br />

20 µm 20 µm<br />

20 µm<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0<br />

WT KO WT KO WT KO WT KO<br />

Spleen Brain Spleen Brain<br />

adjuvant (CFA) and pertussis toxin, wild-type mice developed a<br />

monophasic disease characterized by ascending paralysis 9–16 d<br />

after immunization and prominent leukocyte infiltration and microglial<br />

activation in the CNS (Fig. 1a–c and data not shown). Notably,<br />

most CCR6-knockout mice were completely resistant to the development<br />

of EAE and did not show leukocyte infiltration in the CNS,<br />

whereas a few (8 of 26) had only minimal disease (clinical score, 0.17<br />

± 0.28 (mean ± s.d.)) that developed with similar kinetics but had a<br />

much lower score (Table 1). On day 20, we detected MOG-specific<br />

T cells able to produce IL-17 and IFN-g in the spleens of both wildtype<br />

and CCR6-knockout mice (Fig. 1d). In contrast, we detected<br />

MOG-specific T cells that produced IL-17 and IFN-g only in the<br />

brains of wild-type diseased mice. These results indicate that in CCR6knockout<br />

mice, MOG-reactive T H-17 and T H1 cells are primed in<br />

lymph nodes and enter the circulation but fail to migrate into the CNS<br />

and induce EAE.<br />

CCR6 is not required for TH-17 priming<br />

To rule out a possible contribution of CCR6 in the priming and<br />

differentiation of T cells, we assessed the ability of T cells from CCR6knockout<br />

mice to polarize into effector T cells in vitro. We stimulated<br />

wild-type and CCR6-knockout CD4 + naive T cells in vitro in T H1-,<br />

TH2- or TH-17-polarizing conditions (Supplementary Methods<br />

online). Both wild-type and CCR6-knockout T cells showed a similar<br />

capacity to upregulate mRNA encoding the transcription factors<br />

T-bet, GATA-3 and RORgt and to produce IFN-g, IL-4 and IL-17,<br />

IFN-γ (ng/ml)<br />

ARTICLES<br />

Figure 1 CCR6-deficient mice are resistant to EAE induction. (a) Clinical scores of wild-type mice<br />

(WT; n ¼ 5) and CCR6-knockout mice (CCR6-KO; n ¼ 5) at various times after immunization with<br />

MOG(33–55) in CFA. *, P o 0.01 (Student’s t-test). Data are representative of six experiments (mean<br />

and s.d.). (b,c) Immunofluorescence and histology of cryosections of brains and spinal cords from wildtype<br />

and CCR6-knockout mice with clinical scores of 2 and 0, respectively, collected on day 13 after<br />

perfusion. (b) Staining with anti-laminin (red) and anti-CD45 (green). (c) Immunoperoxidase staining<br />

for CD45 with a hemalaun counterstain. Scale bars, 20 mm. Results are representative of three<br />

experiments. (d) IL-17 and IFN-g in supernatants of total splenic cells (spleen; 2.5 10 6 ) and CD45 +<br />

cells enriched from brains (brain; 0.5 10 6 to 1 10 6 ) obtained from MOG(35–55) immunized<br />

wild-type mice (clinical score, 3) and CCR6-knockout mice (KO; clinical score, 0), restimulated in vitro<br />

with MOG(33–55) and assessed at 72 h of culture. IL-17 and IFN-g were not detected in supernatants<br />

of unstimulated cultures (data not shown). Data are representative of three independent experiments<br />

with three mice per group in each (mean and s.d.).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 515<br />

20 µm<br />

25<br />

20<br />

15<br />

10<br />

5


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Table 1 EAE in wild-type and CCR6-knockout mice<br />

Mouse genotype Incidence<br />

Day of onset<br />

(mean ± s.d.)<br />

Maximum<br />

clinical score<br />

(mean ± s.d.)<br />

Wild type 26 of 29 (89.65%) 16.67 (±0.81) 2.78 (±0.42)<br />

CCR6-knockout 8 of 26 (30.77%) 16.33 (±0.52) 0.17 (±0.28)<br />

Results are cumulative data from six different experiments.<br />

respectively (Supplementary Fig. 1a online). As shown before in<br />

humans 30 , CCR6 was selectively upregulated on wild-type TH-17 cells<br />

but not T H1 or T H2 cells (Supplementary Fig. 1b). To address<br />

whether the lack of CCR6 affected T cell priming in vivo, for example,<br />

by affecting antigen presentation by dendritic cells that express CCR6<br />

(ref. 32), we immunized wild-type and CCR6-knockout mice by<br />

subcutaneous injection of ovalbumin admixed with lipopolysaccharide–monophosphoryl<br />

lipid A or CFA (Supplementary Methods). The<br />

magnitude and quality of the ovalbumin-specific recall T cell<br />

responses were similar in wild-type and CCR6-knockout mice, with<br />

T H1 responses prevailing in mice primed with lipopolysaccharide–<br />

monophosphoryl lipid A and TH-17 responses prevailing in mice<br />

primed with CFA (Supplementary Fig. 1c). Furthermore, OT-II<br />

T cell antigen receptor (TCR)–transgenic T cells labeled with the<br />

cytosolic dye CFSE (carboxyfluorescein diacetate succinimidyl ester)<br />

and adoptively transferred into wild-type or CCR6-knockout mice<br />

proliferated and differentiated into T H-17 cells to a similar extent in<br />

both types of mice after immunization with ovalbumin in CFA<br />

(Supplementary Fig. 2 online). These results collectively indicate<br />

that CCR6 is selectively upregulated in developing mouse T H-17<br />

cells and that its expression is not required for T H-17 priming and<br />

differentiation in vitro and in vivo.<br />

CCR6-knockout mice given T cells are susceptible to EAE<br />

To determine whether CCR6 was required only on T cells, we<br />

adoptively transferred green fluorescent protein–positive (GFP + )<br />

MOG-specific naive 2D2-transgenic T cells into CCR6-knockout<br />

mice or wild-type control mice. After immunizing mice with<br />

MOG(35–55) in CFA, we found CCR6 + IL-17-producing 2D2<br />

T cells in similar proportions in the spleens of CCR6-knockout and<br />

wild-type mice (Fig. 2a,b), which indicated that in CCR6-knockout<br />

mice, MOG-reactive T cells differentiated into CCR6 + TH-17 effector<br />

Figure 2 Transfer of wild-type 2D2 T cells reconstitutes EAE susceptibility<br />

in CCR6-knockout mice. Analysis of wild-type and CCR6-knockout mice<br />

given sham treatment (PBS) or adoptive transfer of naive GFP + CD4 + 2D2<br />

T cells or naive polyclonal CD4 + T cells from wild-type mice or mice of<br />

various knockout strains, then immunized 16 h later with MOG(33–55)<br />

in CFA for EAE induction. (a) Expression of CCR6 (left) and production of<br />

IL-17 and IFN-g (right) by GFP-gated 2D2 T cells transferred into wild-type<br />

and CCR6-knockout mice and primed 7 d earlier (left). Numbers in<br />

quadrants indicate percent CCR6 + cells on GFP-gated 2D2 T cells (left)<br />

or percent cells in each (right). Data are representative of three different<br />

experiments. (b) Production of IL-17 and IFN-g by sorted CCR6 + and CCR6 –<br />

2D2 T cells primed in wild-type mice. Numbers in quadrants indicate<br />

percent cells in each. (c) Clinical scores of wild-type and CCR6-knockout<br />

mice given sham treatment or adoptive transfer of 2D2 T cells (+ 2D2) or<br />

with 2D2 CXCR6-knockout T cells (+ 2D2 CXCR6-KO). (d) Clinical scores<br />

of CCR6-knockout mice given sham treatment or adoptive transfer of CD4 + T<br />

cells (T) from wild-type, CXCR3-knockout, IFN-g-knockout or CCR7-knockout<br />

mice. Data are representative of at least three different experiments with<br />

groups of four or five mice per condition (b,c; mean and s.d.).<br />

cells. Notably, when given 2D2 T cells, both CCR6-knockout and wildtype<br />

mice developed EAE with same kinetics and with greater severity<br />

relative to that of mice that did not receive 2D2 T cells (Fig. 2c).<br />

Finally, CCR6-knockout mice given CCR6-deficient 2D2 T cells<br />

(obtained by crossing of 2D2 TCR–transgenic mice with CCR6-knockout<br />

mice) did not develop EAE (Fig. 2c). These results indicate that<br />

CCR6 expression on transferred T cells is both necessary and sufficient<br />

to reconstitute disease susceptibility in CCR6-knockout mice.<br />

We further investigated the molecular requirements for the reconstitution<br />

of susceptibility to EAE induction in CCR6-knockout mice.<br />

Adoptive transfer of wild-type polyclonal naive CD4 + T cells reconstituted<br />

disease susceptibility in CCR6-knockout mice, as did the<br />

transfer of naive CD4 + T cells from CXCR3-knockout and IFN-gknockout<br />

mice (Fig. 2d). Transfer of naive CD4 + T cells from CCR7knockout<br />

mice induced the development of EAE, which was delayed<br />

approximately 2 weeks (Fig. 2d), possibly due to inefficient priming of<br />

CCR7-deficient naive T cells in lymph nodes. The results presented<br />

above collectively indicate that T cells able to reconstitute disease<br />

susceptibility in CCR6-knockout mice require CCR6, which is characteristic<br />

of T H-17 cells, but not CXCR3 or IFN-g, which are<br />

characteristic of T H1 cells.<br />

CCR6-knockout T cells enter the CNS during active EAE<br />

We next analyzed the cells that infiltrated the CNS at the peak of the<br />

disease. We isolated CD45 + cells from the perfused brains and spinal<br />

cords of CCR6-knockout and wild-type mice given 2D2 T cells and<br />

counted endogenous and transferred 2D2 T cells on the basis of the<br />

expression of GFP and CD45 congenic markers. Unexpectedly, in<br />

both wild-type and CCR6-knockout mice, GFP + 2D2 T cells were<br />

almost completely undetectable in the brain, and we detected only a<br />

few in the spinal cord on day 20 (Fig. 3a). Notably, the abundant<br />

cellular infiltrate in both CCR6-knockout and wild-type mice was<br />

mainly endogenous CCR6-knockout CD4 + and CD8 + T cells, B cells,<br />

neutrophils and inflammatory monocytes (Fig. 3a,b and data not<br />

shown). We confirmed those findings by immunohistology that<br />

showed only rare GFP + cells in inflammatory clusters and several<br />

inflammatory foci with apparently no infiltrating GFP + cells (Fig. 3c).<br />

After in vitro stimulation with MOG(35–55), brain-infiltrating endogenous<br />

CD4 + T cells produced IL-17 and IFN-g (Fig. 3d), which<br />

demonstrated that they were antigen specific and may have been<br />

T H-17 cells, T H1 cells and T cells producing both IL-17 and<br />

IFN-g. The results reported above suggest that both CCR6-sufficient<br />

a<br />

CD4<br />

CD4<br />

b<br />

IL-17<br />

GFP<br />

12.4 0.7<br />

+ 2D2 in WT<br />

10.4<br />

CCR6<br />

GFP<br />

14.8 1.3<br />

+ IFN-γ<br />

2D2 in CCR6-KO<br />

11.6<br />

84.1<br />

81<br />

2.7<br />

2.9<br />

CCR6<br />

IFN-γ<br />

Sorted<br />

GFP<br />

7.3 0.1 85.3 2.5<br />

+ CCR6 – Sorted<br />

2D2 GFP + CCR6 + 2D2<br />

91.4<br />

IFN-γ<br />

1.2<br />

IL-17<br />

IL-17<br />

IL-17<br />

11.9<br />

IFN-γ<br />

0.2<br />

c<br />

Clinical score<br />

d<br />

Clinical score<br />

WT<br />

WT + 2D2<br />

CCR6-KO<br />

CCR6-KO + 2D2<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

CCR6-KO + 2D2 CCR6-KO<br />

0<br />

0 5 10<br />

Time (d)<br />

15 20<br />

4.0<br />

3.0<br />

2.0<br />

1.0<br />

WT T in CCR6-KO<br />

CXCR3-KO T in CCR6-KO<br />

IFN-γ-KO T in CCR6-KO<br />

CCR7-KO T in CCR6-KO<br />

CCR6-KO<br />

0<br />

0 5 10 15<br />

Time (d)<br />

20<br />

25 30<br />

516 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

CD4<br />

Spleen<br />

WT + 2D2 CCR6-KO + 2D2<br />

70.8 0.05 71.3 0.05<br />

GFP-2D2<br />

Brain<br />

WT + 2D2 CCR6-KO + 2D2<br />

84.1 0 84.6 0<br />

Spinal cord<br />

WT + 2D2 CCR6-KO + 2D2<br />

69.0 0.089 73.9 0.01<br />

0<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

WT CCR6-KO WT CCR6-KO<br />

c d<br />

20 µm 20 µm 20 µm 10 µm<br />

and CCR6-deficient T cells can enter the CNS in mice with active EAE<br />

and that at least some of these cells are MOG specific.<br />

CCR6-knockout T cells roll and adhere to inflamed CNS venules<br />

To visualize the interaction of effector T cells with CNS postcapillary<br />

venules, we injected in vitro–primed wild-type and CCR6-knockout<br />

TH-17 cells into wild-type mice in which EAE had been induced 12 d<br />

before and measured cell rolling and adhesion by intravital microscopy.<br />

Activated wild-type and CCR6-knockout CD4 + T cells rolled<br />

and adhered to spinal cord postcapillary venules to a similar extent<br />

(Fig. 4a). In addition, the fraction of permanently adhering T cells was<br />

similar and did not change over time (Fig. 4b), which suggested that<br />

the adhering cells began to enter the spinal cord during the time of<br />

observation. In contrast, when we injected T cells into CCR6-knockout<br />

mice that had been primed with MOG(35–55) and CFA 12 d<br />

before but did not develop disease, the cells failed to interact with<br />

endothelial cells (data not shown). Similarly, and consistent with a<br />

published report 21 , in healthy mice, activated T cells of either<br />

wild-type or CCR6-knockout origin failed to show adhesive interaction<br />

with brain endothelium (data not shown). These results<br />

indicate that CCR6 is not sufficient to mediate adhesion of TH-17<br />

cells to uninflamed endothelial cells (even in CCR6-knockout mice<br />

immunized with CFA and pertussis toxin) and that it is not required<br />

for the entry of T cells into the CNS once tissue inflammation<br />

is established.<br />

T cells migrate into the CNS in two waves<br />

On the basis of the results reported above, we hypothesized that a first<br />

wave of T H-17 cells migrating into the uninflamed CNS in a CCR6dependent<br />

way is needed to trigger activation of parenchymal<br />

b Brain<br />

Figure 3 Recruitment of endogenous effector T cells into the CNS of CCR6-knockout mice after<br />

transfer of 2D2 T cells. (a,b) Seven-color staining (with lineage-specific antibodies) of infiltrating<br />

cells isolated from various organs (above plots) of wild-type and CCR6-knockout mice immunized<br />

20 d before. (a) Expression of CD4 and GFP on gated CD3 + T cells from spleens, brains and spinal<br />

cords of mice after transfer of GFP + 2D2 T cells. Numbers in quadrants indicate percent CD4 + GFP –<br />

cells (top left) or CD4 + GFP + cells (top right). Data are representative of three independent<br />

experiments with three mice each. (b) Absolute number of endogenous (GFP – )CD3 + CD4 + Tcells,<br />

CD3 + CD8 + T cells and B220 + B cells in brains of wild-type and CCR6-knockout mice with (+ 2D2)<br />

or without (–) adoptive transfer of GFP + 2D2 T cells. NS, not significant. P values, Student’s t-test.<br />

Data are representative of three independent experiments (mean and s.d. of three mice per group).<br />

CD4 + T cells (×10 4 )<br />

4<br />

3<br />

2<br />

1<br />

NS<br />

P < 0.01<br />

CD8 + T cells (×10 3 )<br />

IL-17 (ng/ml)<br />

IFN-γ (ng/ml)<br />

Brain<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

Spleen<br />

4<br />

3<br />

2<br />

1<br />

0<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

Spleen<br />

15<br />

10<br />

5<br />

0<br />

NS<br />

P < 0.01<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

0<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

4<br />

Brain<br />

NS<br />

3<br />

P < 0.01<br />

0<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

Brain<br />

4<br />

NS<br />

P < 0.01<br />

3<br />

endothelial cells and the recruitment of a second wave of effector<br />

cells, composed of T H-17 and T H1 cells, which can migrate in a CCR6independent<br />

way across the activated blood-brain barrier. To test<br />

our hypothesis, we transferred into wild-type mice a mixture of<br />

naive 2D2 T cells from wild-type and CCR6-knockout mice that could<br />

be identified by their expression of GFP and CD45 and, after<br />

immunizing the recipient mice with MOG(35–55), we measured<br />

the relative proportions of transferred cells in the brain during<br />

disease development (Fig. 5a). On day 10, wild-type T cells<br />

predominated over CCR6-knockout T cells. In contrast, on day 16,<br />

more cells of both populations were present in the brain in<br />

similar proportions.<br />

We also adoptively transferred a mixture of in vitro–primed T H-17<br />

cells from wild-type mice (which had more than 60% CCR6 + cells)<br />

and CCR6-knockout mice and, 48 h later, measured their migration<br />

into the spleens and brains of mice that developed EAE (Fig. 5b). We<br />

recovered wild-type and CCR6-knockout T cells in similar proportions<br />

from spleens at all time points tested. In contrast, in the brain,<br />

wild-type T cells were present at higher frequency than were CCR6knockout<br />

T cells on day 7, but the proportion of CCR6-knockout cells<br />

steadily increased at later time points (days 10 and 16; Fig. 5b). These<br />

results are collectively consistent with the hypothesis that early<br />

migration of T cells in the brain is CCR6 dependent, whereas late<br />

migration can occur in a CCR6-independent way.<br />

Epithelial cells of the choroid plexus express CCL20<br />

To identify the initial port of entry of CCR6 + T cells, we analyzed in<br />

the CNS of healthy wild-type and CCR6-knockout mice and diseased<br />

wild-type mice expression of the CCR6 ligand CCL20, which is<br />

expressed in liver and Peyer’s patches 29 . We found CCL20 on scattered<br />

B cells (×10 3 )<br />

IL-17 (ng/ml)<br />

IFN-γ (ng/ml)<br />

ARTICLES<br />

Brain<br />

15<br />

10<br />

5<br />

2<br />

1<br />

2<br />

1<br />

0<br />

NS<br />

P < 0.01<br />

– + 2D2 – + 2D2<br />

WT CCR6-KO<br />

(c) Immunofluorescence staining for CD45 (red) in spinal cords from a wild-type mouse and a CCR6-knockout mouse (two images from each) after adoptive<br />

transfer of GFP + 2D2 T cells at day 20 of EAE, showing inflammatory cuffs with or without GFP + 2D2 T cells. Scale bars, 20 mm. Results are representative<br />

of three independent experiments. (d) IL-17 and IFN-g in supernatants of CD45 + cells enriched from spleens and brains of immunized mice with (+ 2D2) or<br />

without ( ) adoptive transfer of GFP + 2D2 T cells, and stimulated in vitro with MOG(33–55), assessed at 72 h of culture. IL-17 and IFN-g were not<br />

detected in unstimulated cultures (data not shown). P values, Student’s t-test. Data are representative of three separate experiments (mean and s.d.).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 517


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Figure 4 CCR6 is not required for T cell rolling and the adhesion of T cells<br />

to inflamed brain endothelia. Intravital microscopy (epi-illumination) of<br />

in vitro–primed T H-17 cells obtained from wild-type and CCR6-knockout<br />

mice, labeled with CellTracker green and injected into the catheterized right<br />

carotid arteries of wild-type mice (with a clinical score of 0.5–1) that had<br />

undergone laminectomy from C2–C7 and removal of the dura. (a) Initial<br />

contact, rolling and capture of T cells in postcapillary venules, among total<br />

T cells passing through a given venule during a 1-min observation period.<br />

Each dot represents one postcapillary venule; small horizontal lines indicate<br />

the median with the interquartile range. Statistical analysis, Mann-Whitney<br />

U-test. Data are representative of three experiments analyzing 20 venules in<br />

five mice that received wild-type T H-17 cells and 18 venules in four mice that received CCR6-knockout T H-17 cells. (b) Adherent and plugging T cells,<br />

presented as cell number per field of view (Cells/FOV), normalized to the number of fields of view. Data are representative of three experiments analyzing four<br />

to six fields of view per mouse (mean and s.d. of five mice that received wild-type T H-17 cells and four mice that received CCR6-knockout T H-17 cells).<br />

cells in several regions of the brain but not on normal or inflamed<br />

endothelial cells of the brain (data not shown). Notably, however, we<br />

found very high and uniform expression of CCL20 in epithelial cells of<br />

the choroid plexus of healthy wild-type and CCR6-knockout mice and<br />

of wild-type mice that developed EAE (Fig. 5c and data not shown).<br />

There was no staining of the choroid plexus parenchyma (data not<br />

shown), which suggested accumulation of CCL20 in and around<br />

choroid plexus epithelium. We detected CCL20 mRNA in liver and<br />

Peyer’s patches, as expected, as well as comparable expression in<br />

choroid plexuses of healthy and EAE mice, whereas we did not<br />

detect it in the brain parenchyma (Supplementary Methods and<br />

Supplementary Fig. 3 online).<br />

The findings reported above suggested that CCL20 might be<br />

required for the initial entry of CCR6 + cells into uninflamed CNS<br />

through the choroid plexus and cerebrospinal fluid and that the same<br />

pathway might be used for the constitutive migration of lymphocytes<br />

for immunosurveillance of the CNS. That hypothesis is consistent<br />

with two observations. First, the choroid plexus of MOG-immunized<br />

CCR6-knockout mice had an accumulation of CD45 + cells greater<br />

than that of wild-type mice (Fig. 5d), which supports the idea that the<br />

epithelial layer of the choroid plexus is the barrier that EAE-initiating<br />

T cells must pass through. In addition, CD45 + cells accumulated in the<br />

choroid plexus in CCR6-deficient mice in the parenchyma between<br />

Figure 5 CCR6 is required for the migration of T cells into the CNS through<br />

CCL20-expressing epithelial cells of the choroid plexus in the steady state<br />

and at early time points of EAE. (a,b) Recruitment of wild-type and CCR6knockout<br />

T cells in developing EAE. (a) Recovery of naive CD4 + Tcells<br />

(2.5 10 6 ) sorted from wild-type (GFP + ) mice and CCR6-knockout<br />

(CD45.1 + ) mice, mixed at a ratio of 1:1 and transferred into wild-type<br />

CD45.2 + mice, which were then immunized with MOG(33–55) in CFA,<br />

presented as the proportion of transferred wild-type and CCR6-knockout<br />

T cells recovered from the brain 10 d or 16 d after immunization.<br />

(b) Recovery of naive CD4 + T cells sorted from wild-type (GFP + ) and CCR6knockout<br />

(CD45.1 + ) mice, stimulated in vitro in T H-17 conditions, mixed at<br />

a ratio of 1:1 and transferred into CD45.2 + wild-type mice in which EAE<br />

was induced 7, 10 or 16 d earlier, presented as the proportion of transferred<br />

T cells recovered from the spleen and brain at 48 h after transfer. Data are<br />

representative of three separate experiments (mean and s.d.). (c) Immunofluorescence<br />

of CCL20 staining of cryosections of brains from wild-type and<br />

CCR6-knockout mice collected after perfusion. Scale bars, 50 mm. Results<br />

are representative of three experiments. (d) Immunoperoxidase staining and<br />

hemalaun counterstaining of CD45 + cells in the choroid plexus parenchyma<br />

of wild-type and CCR6-knockout mice in which EAE was induced 13 d<br />

earlier. Scale bars, 50 mm (top row and bottom left) or 20 mm (bottom,<br />

right). Results are representative of at least three separate experiments.<br />

(e) Absolute numbers of CD4 + Tcells,CD8 + T cells and CD19 + B cells in<br />

brains of wild-type and CCR6-knockout mice in the steady state. Each symbol<br />

represents an individual mouse; small horizontal lines indicate the median.<br />

P values, Student’s t-test. Data are representative of two experiments.<br />

a 20<br />

b<br />

Event (%)<br />

15<br />

10<br />

5<br />

NS NS NS<br />

Cells/FOV<br />

0<br />

0<br />

WT CCR6-KO WT CCR6-KO WT CCR6-KO WT CCR6-KO WT CCR6-KO WT CCR6-KO<br />

Initial contact Rolling Capture 10 min 30 min 60 min<br />

laminin-positive endothelial and epithelial basement membranes (data<br />

not shown). Finally, in the steady state, CCR6-knockout mice had<br />

significantly fewer CNS-associated CD4 + T cells, CD8 + T cells and<br />

B cells than did wild-type mice (Fig. 5e).<br />

CCR6 and CCL20 in human cerebrospinal fluid and brain<br />

To determine whether the findings reported above obtained with the<br />

mouse model could be extended to the human disease of multiple<br />

sclerosis, we analyzed CCR6 expression in T cells from patients with<br />

multiple sclerosis and analyzed CCL20 expression in the brains of<br />

controls and patients with multiple sclerosis. In eight patients with an<br />

initial demyelinating event (the first clinical episode of multiple<br />

sclerosis), CCR6 + CD25 – CD4 + inflammatory T cells were present at<br />

significantly higher frequencies in the cerebrospinal fluid than in<br />

peripheral blood (Fig. 6a). This finding demonstrates that many<br />

cells detectable in the cerebrospinal fluid at the earliest clinical<br />

demyelinating event express CCR6.<br />

Immunohistology of normal healthy tissues showed that CCL20<br />

was expressed in the liver, mainly in Kupffer cells, and in scattered cells<br />

a<br />

Frequency (%)<br />

WT<br />

CCR6-KO<br />

Brain<br />

90<br />

60<br />

30<br />

Cells (×10 2 )<br />

WT<br />

CCR6-KO<br />

b<br />

c d<br />

e<br />

Frequency (%)<br />

CD4<br />

15 P < 0.05<br />

15 50<br />

40<br />

10<br />

10<br />

30<br />

5<br />

5<br />

20<br />

10<br />

0<br />

WT CCR6-KO<br />

0<br />

WT CCR6-KO<br />

0<br />

WT CCR6-KO<br />

+ T cells CD8 + T cells CD19 + B cells<br />

P < 0.05<br />

P < 0.05<br />

Cells (×10 2 )<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Spleen Brain<br />

60 90<br />

40<br />

20<br />

CCR6-KO WT<br />

60<br />

30<br />

Cells (×10 2 )<br />

WT<br />

CCR6-KO<br />

0<br />

0<br />

0<br />

Day 10 Day 16 Day 7 Day 10 Day 16 Day 7 Day 10 Day 16<br />

Frequency (%)<br />

50 µm 50 µm 50 µm<br />

518 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

50 µm<br />

50 µm<br />

20 µm


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

CD4 + CCR6 + CD25 –<br />

T cells (%)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

P = 0.004<br />

0<br />

PBMC CSF<br />

b Healthy tissues c<br />

Liver<br />

Ileum<br />

Choroid plexus Choroid plexus<br />

in the crypts of the ileum (Fig. 6b). Notably, there was strong and<br />

uniform staining for CCL20 in epithelial cells of the choroid plexus in<br />

the brain (Fig. 6b), whereas normal brain parenchyma contained<br />

some cells that stained faintly for CCL20, including astrocytes and cells<br />

with neuronal and microglial morphology (data not shown). In tissues<br />

from patients with multiple sclerosis, we detected high CCL20<br />

expression in inflamed areas, in astrocytes positive for glial fibrillary<br />

acidic protein and in the choroid plexus (Fig. 6c and data not shown).<br />

Accordingly, we detected some CD3 + CCR6 + T cells in inflamed<br />

parenchyma near astrocytes that were CCL20 + (data not shown).<br />

These findings suggest that in humans, recruitment of CCR6 + cells<br />

into the CNS may occur initially through epithelial cells of the choroid<br />

plexus that constitutively express CCL20, whereas at a later stage,<br />

CCL20 production by activated astrocytes may contribute to the<br />

recruitment of CCR6 + T cells into the brain parenchyma. Thus,<br />

CCR6 and CCL20 may represent an evolutionary conserved axis<br />

that regulates the CNS entry and dissemination of T cells in the<br />

steady state and during inflammation.<br />

DISCUSSION<br />

We have shown here that the CCR6 serves an essential function in the<br />

initiation of EAE by controlling the migration of a first wave of<br />

autoreactive T H-17 cells in the uninflamed CNS. The entry of CCR6 +<br />

T cells into the CNS probably occurs through the blood–cerebrospinal<br />

fluid barrier, as epithelial cells of the choroid plexus constitutively<br />

expressed the CCR6 ligand CCL20. The first wave of migratory T cells<br />

was required for the recruitment of a second wave of T cells that<br />

entered the CNS parenchyma in a CCR6-independent way through<br />

activated parenchymal postcapillary venules.<br />

The conclusions above were based on three main findings. First,<br />

when immunized by a standard protocol MOG(35–55) in CFA plus<br />

pertussis toxin, CCR6-knockout mice developed TH-17 responses but<br />

failed to develop EAE. In particular, we did not recover MOG-reactive<br />

effector T cells from the CNS of these mice and their parenchymal<br />

venules did not support the extravasation of adoptively transferred<br />

T cell blasts. Second, the transfer of wild-type naive T cells into CCR6knockout<br />

mice was sufficient to reconstitute disease susceptibility.<br />

However, wild-type CCR6 + T cells predominated in the CNS only at<br />

an initial asymptomatic stage of the disease, whereas during active<br />

disease, the cellular infiltrate in the CNS parenchyma was composed of<br />

endogenous CCR6-knockout T cells, including MOG-reactive TH-17<br />

and T H1 cells, that efficiently adhered to<br />

activated CNS endothelial cells. Third, the<br />

CCR6 ligand CCL20 had high and constitutive<br />

expression by epithelial cells of the<br />

choroid plexus but not by parenchymal<br />

endothelial cells. In addition, in CCR6knockout<br />

mice, T cells seemed to be trapped<br />

between the endothelial and epithelial basement<br />

membranes of the choroid plexus.<br />

Our findings provide a molecular and<br />

anatomical basis for distinguishing between<br />

constitutive and inflammatory pathways of<br />

T cell entry into the CNS and support a twostep<br />

model of EAE pathogenesis in which a<br />

first wave of CCR6 + TH-17 cells leads to the<br />

CCR6-independent recruitment in the CNS<br />

of a second wave of T cells, including T H1<br />

cells, and inflammatory leukocytes. The twostep<br />

model of EAE pathogenesis is consistent<br />

with the finding that in the early phases of<br />

EAE, CD4 + T cells accumulate first in the subarachnoid space and<br />

subsequently appear in the CNS parenchyma 33,34 . In those studies, the<br />

nature of the T cells that initially infiltrate the subarachnoid space and<br />

their port of entry were not defined, although it was suggested that<br />

they might enter directly through the meningeal vessels or disseminate<br />

through the cerebrospinal fluid after crossing the epithelial cell layer of<br />

the choroid plexus. The model is also consistent with the published<br />

description of ‘pioneer’ lymphocytes that migrate into uninflamed<br />

CNS through the choroid plexus in a P-selectin-dependent way 20 and<br />

with studies emphasizing the importance of the activation state and<br />

antigenic specificity of T cells that migrate in the CNS in the initial<br />

phases of EAE 35,36 .Atwo-wavemigrationofTH-17 and TH1 cells<br />

into tissues may apply to other autoimmune diseases such as collageninduced<br />

arthritis 37 . In addition, in a model of Mycobacterium<br />

tuberculosis infection, it has been shown that vaccination induces<br />

IL-17-producing CD4 + T cells that populate the lung and, after<br />

challenge, trigger the production of chemokines that recruit CD4 + T<br />

cells that produce IFN-g, which ultimately restrict bacterial growth 38 .<br />

The findings that CCR6 is essential for the entry of T cells into the<br />

CNS in the early phase of EAE and the expression of CCL20 in<br />

epithelial cells of the choroid plexus provide a molecular determinant<br />

and an anatomical site for the first triggering step in EAE pathogenesis.<br />

It remains to be established which adhesion molecules are needed<br />

to T cells to cross the blood–cerebrospinal fluid barrier and whether<br />

chemokines other than CCL20 might be expressed on epithelial cells<br />

of the choroid plexus. Consistent with our finding that CXCR3 was<br />

not needed to mediate the entry of T cells in the initial phase of EAE,<br />

we found that its ligands, CXCL9 and CXC10, were not expressed in<br />

the choroid plexus of C57BL/6 mice (D.B., unpublished data), which<br />

also do not express CXCL11 because of a genetic defect.<br />

An implication of our findings is that one function of the first wave<br />

of CCR6 + TH-17 cell is to activate the postcapillary venules in the CNS<br />

parenchyma, which in normal circumstances are inefficient in sustaining<br />

rolling and adhesion of activated leukocytes and hence the entry of<br />

cells into the CNS parenchyma 21 . It is conceivable that once they have<br />

entered through the choroid plexus into the cerebrospinal fluid, T H-17<br />

cells may disseminate at the pial surface and in the enlarged perivascular<br />

Virchow-Robin spaces, where they may recognize self antigens<br />

displayed on resident antigen-presenting cells. Once activated, T H-17<br />

cells produce cytokines and chemokines that act locally to trigger<br />

activation of the blood-brain barrier and to initiate the influx of large<br />

Brain (MS)<br />

Figure 6 Expression of CCL20 and CCR6 in human normal tissue and multiple sclerosis tissues.<br />

(a) CCR6 + cells among gated CD4 + CD25 – T cells from matched samples of peripheral blood (PBMC)<br />

and cerebrospinal fluid (CSF) obtained from eight patients with multiple sclerosis. P value, Mann-<br />

Whitney U-test. Data are representative of eight experiments. (b,c) CCL20 expression on normal human<br />

tissues (b) and in various areas of the brains of patients with multiple sclerosis (MS; c), assessed by<br />

immunohistochemistry of formalin-fixed, paraffin-embedded sections with a CCL20-specific antibody.<br />

Original magnification, 40 (b, all images; c, top left and bottom row) or 60 (c, top right). Results<br />

are representative of two separate experiments on two different tissue samples.<br />

ARTICLES<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 519


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

numbers of inflammatory cells, including T H-17 and T H1 cells,<br />

neutrophils and inflammatory monocytes, that build up the lesions<br />

characteristic of EAE. Although we have shown that the T cell infiltrate<br />

was composed of MOG-reactive T cells, it is likely that antigennonspecific<br />

bystander T cells are also recruited through the bloodbrain<br />

barrier and may constitute a large fraction of the total cellular<br />

infiltrate, as has been shown in delayed-type hypersensitivity reactions<br />

as well as several autoimmune diseases 39 .<br />

It is worth noting that in the experimental conditions of active EAE<br />

used in our studies, effector T cells were induced in peripheral lymph<br />

nodes by subcutaneous immunization and had to enter uninflamed<br />

CNS, as the inflammatory stimuli were limited to the local effect of<br />

CFA and the systemic effect of pertussis toxin administered at the time<br />

of immunization. It can be anticipated that the requirement for the<br />

CCR6-dependent entry of autoreactive T cells through the choroid<br />

plexus, which seems to be rate-limiting in steady-state conditions, may<br />

be bypassed whenever the CNS vasculature is activated by local or<br />

systemic inflammatory stimuli 40 or in models of passive EAE induced<br />

by the injection of highly activated T cell blasts.<br />

An open issue relates to the nature of the cytokines produced by the<br />

first wave of T H-17 cells needed for the initiation of EAE. T H-17 cells<br />

produce IL-17, IL-22, tumor necrosis factor and, in some cases, IFN-g,<br />

which can act on microglial cells and endothelial cells 41 . We found that<br />

IFN-g expression was not required in the CCR6 + T cells that initiated<br />

EAE, and we have preliminary evidence that IL-17A may also be<br />

dispensable (A.R., unpublished data). In addition, studies have shown<br />

that IL-17A and IL-17F, as well as IL-22, may not be required for the<br />

development of EAE 42,43 ; such findings would be consistent with<br />

either a redundant function of these cytokines in disease development<br />

or with an essential involvement of another functional property<br />

(cytokine or chemokine production; chemokine receptor expression)<br />

of TH-17-lineage cells. Experiments with CCR6-knockout mice given<br />

adoptive transfer of CCR6-sufficient T cells carrying selective genetic<br />

defects will help to resolve this issue.<br />

We have shown here that once EAE was triggered, effector T cells of<br />

both wild-type and CCR6-knockout origin efficiently rolled and<br />

adhered to inflamed CNS postcapillary venules and migrated with<br />

similar efficiency into the CNS. These findings indicate that at a late<br />

stage, the CCR6-CCL20 axis is dispensable and the recruitment of<br />

inflammatory cells into the CNS parenchyma can be mediated by<br />

other chemokine receptors 44 . Several inflammatory chemokines are<br />

upregulated in the CNS during EAE 45–47 , and some chemokine<br />

receptors, such as CCR1, CCR2 and CXCR3, have been shown to be<br />

involved in EAE 48–52 . In addition, a CCR5 receptor antagonist and<br />

antibodies to CCL20 have been reported to diminish disease severity<br />

53,54 . Such findings are consistent with the presence of multiple<br />

redundant mechanisms that regulate the entry of leukocytes into the<br />

CNS once inflammation has been established.<br />

In the context of its function in migration into the CNS, CCR6<br />

expression on activated and memory T cells and other cell types<br />

deserves mention. We have shown that CCR6 was selectively upregulated<br />

in developing mouse TH-17 cells but not TH1 orTH2 cells both<br />

in vitro and in vivo. In addition, we have shown that in mice, CCR6 was<br />

selectively induced by CFA, the adjuvant typically used for EAE<br />

induction. It is notable that in humans, CCR6 is expressed not only<br />

on T H-17 cells, where it is expressed together with CCR4, but also on<br />

cells that produce both IL-17 and IFN-g, aswellasonasubsetofT H1<br />

cells characterized by CXCR3 expression 30 . It remains to be determined<br />

whether CCR6 can be induced on mouse T H1 cells in certain conditions,<br />

a possibility that may explain the finding that T H1 cells can enter<br />

uninflamed brain 40 , although in that case 40 , the port of entry into the<br />

CNS was not identified. CCR6 is also expressed on B cells, which have<br />

been linked to the pathogenesis of multiple sclerosis 55 , and on a subset<br />

of regulatory T cells 56 . Indeed, it has been shown that CCR6 is<br />

important in regulating the recruitment of T H-17 cells as well as<br />

regulatory T cells into inflammatory tissues 57 . Thus, it is possible that<br />

CCR6 is used by a variety of cells to enter the brain through the<br />

choroid plexus and participate in immunoregulatory circuits.<br />

Our findings help address a controversial aspect of the relative<br />

functions of TH1 andTH-17 cells in brain inflammation. Although<br />

studies have demonstrated a requirement for IL-17-producing T cells<br />

in EAE 10 , it was puzzling that T H-17 cells often accounted for a minor<br />

fraction of infiltrating T cells, at least at the peak of the disease. Indeed,<br />

subsequent studies have led to a reevaluation of the function of T H1<br />

cells 58 . The two-step model of EAE pathogenesis provides a way to<br />

reconcile those findings with the following considerations. First, the<br />

proportion of T H1andT H-17 cells may be highly variable depending<br />

on the priming conditions. Second, a few autoreactive CCR6 + T cells<br />

may be able to promote the vascular recruitment of large numbers of<br />

inflammatory cells. Third, CCR6 is expressed in subsets of TH1 cells in<br />

humans and possibly in mice.<br />

It is well appreciated that chemokines and their receptors, together<br />

with adhesion molecules, control the constitutive homing of lymphocytes<br />

to lymphoid and nonlymphoid tissues to accomplish their<br />

immunosurveillance function. For example, CCR9 defines a subset of<br />

gut-homing lymphocytes, whereas CCR4 and CCR10 direct skin-tropic<br />

T-cell trafficking 59 . Similar selective mechanisms that target T cells to<br />

theCNSarethoughttoexist 60 , but they have not been identified. On<br />

the basis of our findings, we propose that CCR6 is a brain-specific<br />

determinant for the constitutive trafficking of patrolling T cells and<br />

B cells in the CNS. The rapid distribution of T cells throughout the<br />

choroid plexus and the cerebrospinal fluid into the meningeal space<br />

seems to be instrumental for broad surveillance over the entire surface<br />

of the CNS, thus maximizing the chance that antigen-bearing antigenpresenting<br />

cells will be detected at superficial sites while at the same<br />

time the entry of inflammatory cells into the parenchyma, where this is<br />

needed, will be elicited. The cytokine-mediated activation of endothelial<br />

cells of the blood-brain barrier executed by patrolling T cells might<br />

provide a means for targeting large numbers of effector cells to a<br />

precise region of the CNS parenchyma.<br />

Some aspects of the mouse model also apply to humans. These<br />

include the fact that in humans, CCR6 is expressed by T H-17 cells 30 ,as<br />

well as the enrichment for CCR6 + cells in the cerebrospinal fluid of<br />

patients presenting with the first clinical symptom of multiple<br />

sclerosis, a condition that may parallel the earliest phase of EAE.<br />

Moreover, in noninflammatory conditions in humans, CCL20 is<br />

expressed almost exclusively by epithelial cells of the choroid plexus,<br />

whereas in multiple sclerosis tissues, CCL20 can also be expressed by<br />

astrocytes positive for glial fibrillary acidic protein (A.U., unpublished<br />

data), which suggests that in multiple sclerosis, selective recruitment of<br />

CCR6 + cells may occur through the choroid plexus in the early phase<br />

of disease, whereas at a later stage, astrocytes may contribute to the<br />

recruitment of CCR6 + T cells in brain parenchyma. These findings<br />

indicate involvement of the CCR6-CCL20 axis in initiating brain<br />

inflammation in humans and possibly an evolutionary conserved<br />

mechanism of immune surveillance in the CNS. The relapsingremitting<br />

form of human multiple sclerosis and the anatomical<br />

distribution of multiple sclerosis lesions may be consistent with either<br />

distinct waves of migration or asynchronous activation of already<br />

resident CCR6 + T cells. Distinguishing between these possibilities is<br />

relevant to understanding the potential therapeutic use of CCR6blocking<br />

drugs in human multiple sclerosis.<br />

520 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

METHODS<br />

Mice. C57BL/6 mice were from Harlan; 2D2 TCR–transgenic (006912), UBC-<br />

GFP (004353) and Ifng –/– (002287) mice were from the Jackson Laboratory.<br />

OT-II TCR–transgenic mice (provided by J. Kirberg) were bred onto backgrounds<br />

of various Cd45 alleles in the animal facility of the Institute for<br />

Research in Biomedicine. Ccr6 –/– mice have been described 31 , Cxcr3 –/– mice<br />

were provided by C. Gerard, and Ccr7 –/– mice were provided by M. Lipp. Mice<br />

were treated in accordance with guidelines of the Swiss Federal Veterinary<br />

Office and experiments were approved by the Dipartimento della Sanità<br />

eSocialità.<br />

EAE model. Groups of female C57BL/6 mice 8–10 weeks of age were<br />

immunized subcutaneously on day 0 with 100 mg MOG(35–55) (MEVG<br />

WYRSPFSRVVHLYRNGK; Servei de Proteòmica, Pompeu Fabra University,<br />

Barcelona) emulsified in CFA (Difco). Pertussis toxin in 100 ml salinewas<br />

injected intravenously twice, on days 0 and 2 or days 1 and 3. Disease severity<br />

was assigned scores on the following scale: 0, no disease; 1, tail weakness; 2,<br />

paraparesis; 3, paraplegia; 4, paraplegia with forelimb weakness or paralysis; 5,<br />

moribund or dead. In some EAE experiments, naive T cells (1 10 5 )from<br />

GFP + 2D2-transgenic mice or 2D2 CCR6-knockout mice or total CD4 +<br />

T cells (10 10 6 ) from wild-type, CXCR3-knockout, IFN-g-knockout or<br />

CCR7-knockout mice were transferred intravenously into mice 16 h before<br />

immunization. In some experiments, a mixture of GFP + 2D2-transgenic T cells<br />

(2.5 10 6 ) and 2D2 CCR6-knockout T cells (at a ratio of 1:1) were<br />

transferred intravenously into mice 16 h before immunization. Alternatively,<br />

effector TH-17 cells were generated in vitro from GFP + 2D2-transgenic T cells<br />

(2.5 10 6 ) and 2D2 CCR6-knockout T cells (at a ratio of 1:1) and were<br />

transferred intravenously in wild-type mice at various times after EAE induction.<br />

For the preparation of CNS lymphocytes, mice were perfused through the<br />

left cardiac ventricle with cold PBS. The forebrain and cerebellum were<br />

dissected, were cut into pieces and were digested for 45 min at 37 1C with<br />

collagenase D (1 mg/ml; Roche Diagnostics) and DNaseI (1 mg/ml; Sigma).<br />

CD45 + cells were isolated by passage of the tissue through a cell strainer<br />

(70 mm), followed by incubation with beads coated with antibody to CD45<br />

(anti-CD45; 130-052-301; Milteny). After passage through the column, CD45 +<br />

cells were washed and resuspended in culture medium for further analysis.<br />

Flow cytometry analysis and in vitro stimulation. For analysis of mouse<br />

phenotypes, the following monoclonal antibodies were used: anti-L-selectin<br />

(MEL14), anti-CD44 (IM7), anti-CD4 (RM4-5), anti-CD8a (53-6.7), anti-CD3<br />

(145-2C11), anti-CD28 (37.51), anti-B220 (RA3-6B2), anti-CD19 (1D3), anti-<br />

CD11b (M1/70), anti-CD45.1 (A20), anti-CD45.2 (104), anti-CD127 (A7R34)<br />

and anti-IFN-g (XMG1.2; all from eBiosciences); and anti-IL-17A (TC11-<br />

18H10) and anti-CCR6 (140706; both from BD Biosciences). For intracellular<br />

cytokine staining, cells were stimulated for 4 h with phorbol 12-myristate<br />

13-acetate (100 nM; Sigma) and ionomycin (1 mg/ml; Sigma), with the final 2 h<br />

of culture in the presence of brefeldin A (10 mg/ml; Sigma). Labeled antibodies<br />

were used after cells were fixed in 4% (wt/vol) paraformaldehyde and made<br />

permeable with 0.5% (wt/vol) saponin (Sigma-Aldrich). Six-color staining of<br />

the cell surface was done with the appropriate combinations of antibodies<br />

conjugated to fluorescein isothiocyanate, phycoerythrin, peridinine chlorophyll<br />

protein complex, phycoerythrin-indotricarbocyanine, allophycocyanin, allophycocyanin-indotricarbocyanine<br />

or biotin, and with streptavidin labeled with<br />

phycoerythrin-indotricarbocyanine or allophycocyanin-indotricarbocyanine<br />

(BD Biosciences). Samples were acquired on a FACSCanto (BD Biosciences)<br />

and were analyzed with FlowJo software (TreeStar).<br />

For studies of human cell phenotypes, peripheral blood and cerebrospinal<br />

fluid were obtained from eight patients (after informed consent was provided)<br />

who presented with a first demyelinating event suggestive of multiple sclerosis<br />

and underwent venipuncture and lumbar puncture for diagnostic purposes.<br />

The study was approved by the Ethical Committee and Board of the Department<br />

of Neurosciences, Ophthalmology and Genetics, University of Genoa.<br />

Peripheral blood mononuclear cells were isolated with Ficoll and cerebrospinal<br />

fluid leukocytes were collected after centrifugation, then these cells were<br />

incubated for 30 min with the following monoclonal antibodies: fluorescein<br />

isothiocyanate–conjugated anti-CD4 (RPA-T4), phycoerythrin-conjugated<br />

anti-CCR6 (11A9), phycoerythrin-indodicarbocyanine–conjugated anti-CD25<br />

ARTICLES<br />

(M-A251) and allophycocyanin-conjugated anti-CD45RO (UCHL1; all from<br />

BD Biosciences). At the end, cells were washed with PBS, were resuspended,<br />

were counterstained with propidium iodide (1 mg/ml; Sigma Aldrich) and were<br />

analyzed by flow cytometry of the propidium iodide–negative population. A<br />

FACSCanto (BD Biosciences) was used for all flow cytometry and data were<br />

analyzed with FACSDiva & FloJo software.<br />

For antigen-specific restimulation of mouse T cells, 5 10 6 splenocytes or<br />

2.5 10 6 lymph node cells were cultured for 3 d in presence of MOG(35–55).<br />

Cells purified from the brain were cultured for 3 d in presence of fixed<br />

splenocytes loaded with MOG(35–55). For fixation, spleens were removed from<br />

naive C57BL/6 mice and erythrocytes were lysed. Splenocytes were incubated<br />

for 90 min at 37 1C with MOG(35–55) (50 mg/ml), then were washed in 1%<br />

(vol/vol) FCS in PBS and were fixed for 30 s at 20 1C in 0.05% (vol/vol)<br />

glutaraldehyde (Merck). An equal volume of a solution of glycine (0.2 M;<br />

Sigma-Aldrich) was added for 30 s and then cells were washed and used for<br />

coculture. Cytokines in culture supernatants were measured by enzyme-linked<br />

immunosorbent assay according to the manufacturer’s protocols (BD Biosciences).<br />

Data were analyzed with the SoftMax program.<br />

Immunohistology and immunofluorescence. Mice were perfused through the<br />

left cardiac ventricle with 1% (vol/vol) formaldehyde (Grogg Chemie) in PBS.<br />

Brains and spinal cords were removed, were embedded in Tissue-Tek optimum<br />

cutting temperature compound (Haslab) and were ‘snap-frozen’ in a bath of<br />

dry ice and isopentane (Grogg Chemie). Cryostat sections 6 mm inthickness<br />

were air-dried overnight, were fixed in acetone and were stained for immunohistology<br />

with a three-step immunoperoxidase staining kit according to the<br />

manufacturer’s protocol (Vectastain; Reactolab). For immunofluorescence<br />

staining, sections were blocked for 20 min with skim milk and then were<br />

incubated for 1 h each with primary and secondary antibody diluted in skim<br />

milk, with washing steps of Tris-buffered saline between incubations. After a<br />

final wash in Tris-buffered saline, sections were mounted in Mowiol solution<br />

(Calbiochem). Antibodies used were as follows: fluorescein isothiocyanate–<br />

conjugated anti-CD45 (30F11; BD Biosciences), anti-CCL20 (AB9829; Abcam),<br />

anti-laminin (Z0097; Dako), anti-PECAM-1 (Mec13.3; BD Biosciences) and<br />

anti–‘pan cytokeratin’ (C2562; Sigma).<br />

LifeSpan Biosciences analyzed CCL20 expression in healthy and diseased<br />

human tissues. Formalin-fixed, paraffin-embedded tissues were stained with<br />

rabbit polyclonal anti-CCL20 (15 mg/ml; AB9829; Abcam). The detection<br />

system consisted of anti-rabbit secondary antibody (BA-1000; Vector) and an<br />

ABC-AP kit (avidin–biotinylated enzyme complex–alkaline phosphatase;<br />

AK-5000; Vector) with a Red substrate kit (SK-5100; Vector), which was used<br />

to produce a fuchsia-colored deposit. Only tissues that were positive for<br />

staining of CD31 and vimentin were used. The negative control consisted<br />

of immunohistochemistry of adjacent sections in the absence of primary<br />

antibody. Slides were imaged with a DVC 1310C digital camera coupled to a<br />

Nikon microscope.<br />

Intravital microscopy. Active EAE was induced by immunization with<br />

MOG(35–55) in CFA. Mice with a score of 0.5 (limp tail) to 1 (hind leg<br />

weakness) were used for intravital fluorescence videomicroscopy experiments.<br />

The spinal cord window was created as described in the Supplementary<br />

Methods. Normal body temperature was maintained throughout the entire<br />

experiment. Epi-illumination techniques were used for intravital fluorescence<br />

videomicroscopy with an IVM500 microscope (Mikron Instruments) coupled<br />

to a 50-Watt mercury lamp (HBO 50 microscope illuminator; Zeiss) and<br />

combined with blue filter blocks (exciter, 455DF70; dichroic, 515DRLP;<br />

emitter, 515ALP) and green filter blocks (exciter, 525DF45; dichroic, 560DRLP;<br />

emitter, 565ALP). Observations were made with 4 , 10 and 20 longdistance<br />

working objectives (Zeiss). All experiments were recorded by means of<br />

a low-light silicon-intensified target video camera with an optional image<br />

intensifier for weak fluorescence (Dage-MTI of Michigan City). Data were<br />

transferred to a digital video system for later offline analysis of the interaction<br />

of cells with CNS microvessels. The injection of 1% (vol/vol) tetramethylrhodamine<br />

isothiocyanate–conjugated dextran in 0.9% (wt/vol) NaCl allowed<br />

visualization of the spinal cord microvasculature. In parallel, mouse T cells were<br />

stained for 45 min at 37 1C with2.5mM CellTracker green (Molecular Probes)<br />

and were injected into the carotid artery (4 10 6 T cells in three<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 521


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

0.1-ml aliquots). Injection into this catheter resulted in direct transport into<br />

observed vessels, where interactions were recorded by digital video for subsequent<br />

offline analysis. Several adjacent areas were scanned in a ‘stepwise’ way<br />

for recording of permanently adhering T cells at 10 min, 30 min and 1 h after<br />

injection. Recorded videos were subsequently analyzed offline as described<br />

(Supplementary Methods).<br />

Statistics. Differences between data sets were analyzed by Student’s t-test or a<br />

Mann-Whitney U-test.<br />

Accession code. UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signaling-gate<br />

way.org): A000629<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank D. Jarrossay for cell sorting; T. Périnat and S. Minghelli for<br />

immunohistology analysis; E. Mira Catò and L. Perlini for technical assistance;<br />

B. Becher (University Hospital Zurich), C. Gerard (Harvard Medical School),<br />

J. Kirberg (Max-Plank Institute of Immunobiology) and M. Lipp (Max<br />

Delbruck Center) for mouse strains; L. Sallusto for discussions and support;<br />

and A. Almeida, A. Martín-Fontecha, G. Napolitani, U. Schenk and<br />

M. Uguccioni for discussions. Supported by the Swiss National Science<br />

Foundation (31-101962 to F.S.), the European Commission Sixth Framework<br />

Programme (LSB-CT-2005-518167 IINOCHEM and LSHG-CT-2005-005203<br />

MUGEN), the US National Multiple Sclerosis Society (B.E. and C.C.), the Swiss<br />

Multiple Sclerosis Society, the Italian Foundation for Multiple Sclerosis (Stem<br />

Cell Project 2007-2009), the European Union-funded International Graduate<br />

Program in Molecular Medicine (A.R.), Boehringer Ingelheim Fonds (D.B.) and<br />

the Helmut Horten Foundation (to The Institute for Research in Biomedicine).<br />

AUTHOR CONTRIBUTIONS<br />

A.R. did most of the experiments and contributed to experimental design;<br />

C.C., D. Baumjohann, F.B. and D. Bottinelli did experiments; S.L. generated the<br />

CCR6-knockout mice and provided intellectual input; B.E. and A.U. interpreted<br />

data, provided intellectual input and contributed to writing the manuscript;<br />

A.L. provided intellectual input and wrote the manuscript; and F.S. conceived<br />

the study, interpreted the data and wrote the manuscript.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. McGeachy, M.J. & Cua, D.J. Th17 cell differentiation: the long and winding road.<br />

Immunity 28, 445–453 (2008).<br />

2. Bettelli, E., Korn, T., Oukka, M. & Kuchroo, V.K. Induction and effector functions of<br />

T H17 cells. <strong>Nature</strong> 453, 1051–1057 (2008).<br />

3. Dong, C. TH17 cells in development: an updated view of their molecular identity and<br />

genetic programming. Nat. Rev. Immunol. 8, 337–348 (2008).<br />

4. Weaver, C.T., Hatton, R.D., Mangan, P.R. & Harrington, L.E. IL-17 family cytokines and<br />

the expanding diversity of effector T cell lineages. Annu. Rev. Immunol. 25, 821–852<br />

(2007).<br />

5. Gaffen, S.L. An overview of IL-17 function and signaling. Cytokine 43, 402–407 (2008).<br />

6. Steinman, L. A brief history of T H17, the first major revision in the T H1/T H2hypothesis<br />

of T cell-mediated tissue damage. Nat. Med. 13, 139–145 (2007).<br />

7. Ivanov, I.I. et al. The orphan nuclear receptor RORgt directs the differentiation program<br />

of proinflammatory IL-17 + T helper cells. Cell 126, 1121–1133 (2006).<br />

8. Komiyama, Y. et al. IL-17 plays an important role in the development of experimental<br />

autoimmune encephalomyelitis. J. Immunol. 177, 566–573 (2006).<br />

9. Hofstetter, H.H. et al. Therapeutic efficacy of IL-17 neutralization in murine experimental<br />

autoimmune encephalomyelitis. Cell. Immunol. 237, 123–130 (2005).<br />

10. Langrish, C.L. et al. IL-23 drives a pathogenic T cell population that induces<br />

autoimmune inflammation. J. Exp. Med. 201, 233–240 (2005).<br />

11. Tzartos, J.S. et al. Interleukin-17 production in central nervous system-infiltrating<br />

T cells and glial cells is associated with active disease in multiple sclerosis.<br />

Am. J. Pathol. 172, 146–155 (2008).<br />

12. Bettelli, E. et al. Loss of T-bet, but not STAT1, prevents the development of experimental<br />

autoimmune encephalomyelitis. J. Exp. Med. 200, 79–87 (2004).<br />

13. Luger, D. et al. Either a Th17 or a Th1 effector response can drive autoimmunity:<br />

conditions of disease induction affect dominant effector category. J. Exp. Med. 205,<br />

799–810 (2008).<br />

14. Kroenke, M.A., Carlson, T.J., Andjelkovic, A.V. & Segal, B.M. IL-12- and IL-23modulated<br />

T cells induce distinct types of EAE based on histology, CNS chemokine<br />

profile, and response to cytokine inhibition. J. Exp. Med. 205, 1535–1541 (2008).<br />

15. Stromnes, I.M., Cerretti, L.M., Liggitt, D., Harris, R.A. & Goverman, J.M. Differential<br />

regulation of central nervous system autoimmunity by T H1andT H17 cells. Nat. Med.<br />

14, 337–342 (2008).<br />

16. Lees, J.R., Golumbek, P.T., Sim, J., Dorsey, D. & Russell, J.H. Regional CNS responses<br />

to IFN-g determine lesion localization patterns during EAE pathogenesis. J. Exp. Med.<br />

205, 2633–2642 (2008).<br />

17. Engelhardt, B. & Ransohoff, R.M. The ins and outs of T-lymphocyte trafficking to the<br />

CNS: anatomical sites and molecular mechanisms. Trends Immunol. 26, 485–495<br />

(2005).<br />

18. Yednock, T.A. et al. Prevention of experimental autoimmune encephalomyelitis by<br />

antibodies against a 4b 1 integrin. <strong>Nature</strong> 356, 63–66 (1992).<br />

19. Engelhardt, B., Vestweber, D., Hallmann, R. & Schulz, M. E- and P-selectin are not<br />

involved in the recruitment of inflammatory cells across the blood-brain barrier in<br />

experimental autoimmune encephalomyelitis. Blood 90, 4459–4472 (1997).<br />

20. Carrithers, M.D., Visintin, I., Kang, S.J. & Janeway, C.A. Jr. Differential adhesion<br />

molecule requirements for immune surveillance and inflammatory recruitment. Brain<br />

123, 1092–1101 (2000).<br />

21. Piccio, L. et al. Molecular mechanisms involved in lymphocyte recruitment in inflamed<br />

brain microvessels: critical roles for P-selectin glycoprotein ligand-1 and heterotrimeric<br />

Gi-linked receptors. J. Immunol. 168, 1940–1949 (2002).<br />

22. Vajkoczy, P., Laschinger, M. & Engelhardt, B. a4-integrin-VCAM-1 binding mediates<br />

G protein-independent capture of encephalitogenic T cell blasts to CNS white matter<br />

microvessels. J. Clin. Invest. 108, 557–565 (2001).<br />

23. Engelhardt, B., Wolburg-Buchholz, K. & Wolburg, H. Involvement of the choroid<br />

plexus in central nervous system inflammation. Microsc. Res. Tech. 52, 112–129<br />

(2001).<br />

24. Ransohoff, R.M., Kivisakk, P. & Kidd, G. Three or more routes for leukocyte migration<br />

into the central nervous system. Nat. Rev. Immunol. 3, 569–581 (2003).<br />

25. Giunti, D. et al. Phenotypic and functional analysis of T cells homing into the CSF of<br />

subjects with inflammatory diseases of the CNS. J. Leukoc. Biol. 73, 584–590<br />

(2003).<br />

26. Steffen, B.J., Breier, G., Butcher, E.C., Schulz, M. & Engelhardt, B. ICAM-1, VCAM-1,<br />

and MAdCAM-1 are expressed on choroid plexus epithelium but not endothelium and<br />

mediate binding of lymphocytes in vitro. Am. J. Pathol. 148, 1819–1838 (1996).<br />

27. Pedemonte, E. et al. Mechanisms of the adaptive immune response inside the central<br />

nervous system during inflammatory and autoimmune diseases. Pharmacol. Ther. 111,<br />

555–566 (2006).<br />

28. Bromley, S.K., Mempel, T.R. & Luster, A.D. Orchestrating the orchestrators: chemokines<br />

in control of T cell traffic. Nat. Immunol. 9, 970–980 (2008).<br />

29. Rossi, D.L., Vicari, A.P., Franz-Bacon, K., McClanahan, T.K. & Zlotnik, A. Identification<br />

through bioinformatics of two new macrophage proinflammatory human chemokines:<br />

MIP-3a and MIP-3b. J. Immunol. 158, 1033–1036 (1997).<br />

30. Acosta-Rodriguez, E.V. et al. Surface phenotype and antigenic specificity of human<br />

interleukin 17–producing T helper memory cells. Nat. Immunol. 8, 639–646<br />

(2007).<br />

31. Cook, D.N. et al. CCR6 mediates dendritic cell localization, lymphocyte homeostasis,<br />

and immune responses in mucosal tissue. Immunity 12, 495–503 (2000).<br />

32. Greaves, D.R. et al. CCR6, a CC chemokine receptor that interacts with macrophage<br />

inflammatory protein 3a and is highly expressed in human dendritic cells. J. Exp. Med.<br />

186, 837–844 (1997).<br />

33. Brown, D.A. & Sawchenko, P.E. Time course and distribution of inflammatory and<br />

neurodegenerative events suggest structural bases for the pathogenesis of experimental<br />

autoimmune encephalomyelitis. J. Comp. Neurol. 502, 236–260 (2007).<br />

34. Kivisakk, P. et al. Localizing central nervous system immune surveillance: Meningeal<br />

antigen-presenting cells activate T cells during experimental autoimmune encephalomyelitis.<br />

Ann. Neurol. published online, doi:10.1002/ana.21379 (21 May 2008).<br />

35. Hickey, W.F., Hsu, B.L. & Kimura, H. T-lymphocyte entry into the central nervous<br />

system. J. Neurosci. Res. 28, 254–260 (1991).<br />

36. Flugel, A. et al. Migratory activity and functional changes of green fluorescent effector<br />

cells before and during experimental autoimmune encephalomyelitis. Immunity 14,<br />

547–560 (2001).<br />

37. Hirota, K. et al. Preferential recruitment of CCR6-expressing Th17 cells to inflamed<br />

joints via CCL20 in rheumatoid arthritis and its animal model. J. Exp. Med. 204,<br />

2803–2812 (2007).<br />

38. Khader, S.A. et al. IL-23 and IL-17 in the establishment of protective pulmonary CD4 +<br />

T cell responses after vaccination and during Mycobacterium tuberculosis challenge.<br />

Nat. Immunol. 8, 369–377 (2007).<br />

39. Steinman, L. A few autoreactive cells in an autoimmune infiltrate control a vast<br />

population of nonspecific cells: a tale of smart bombs and the infantry. Proc. Natl.<br />

Acad. Sci. USA 93, 2253–2256 (1996).<br />

40. O’Connor, R.A. et al. Cutting Edge: Th1 cells facilitate the entry of Th17 cells<br />

to the central nervous system during experimental autoimmune encephalomyelitis.<br />

J. Immunol. 181, 3750–3754 (2008).<br />

41. Kebir, H. et al. Human T H17 lymphocytes promote blood-brain barrier disruption and<br />

central nervous system inflammation. Nat. Med. 13, 1173–1175 (2007).<br />

42. Kreymborg, K. et al. IL-22 is expressed by Th17 cells in an IL-23-dependent fashion,<br />

but not required for the development of autoimmune encephalomyelitis. J. Immunol.<br />

179, 8098–8104 (2007).<br />

43. Haak, S. et al. IL-17A and IL-17F do not contribute vitally to autoimmune neuroinflammation<br />

in mice. J. Clin. Invest. 119, 61–69 (2009).<br />

44. Charo, I.F. & Ransohoff, R.M. The many roles of chemokines and chemokine receptors<br />

in inflammation. N. Engl. J. Med. 354, 610–621 (2006).<br />

45. Balashov, K.E., Rottman, J.B., Weiner, H.L. & Hancock, W.W. CCR5 + and CXCR3 +<br />

T cells are increased in multiple sclerosis and their ligands MIP-1a and IP-10 are<br />

expressed in demyelinating brain lesions. Proc. Natl. Acad. Sci. USA 96, 6873–6878<br />

(1999).<br />

522 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

46. Ambrosini, E., Columba-Cabezas, S., Serafini, B., Muscella, A. & Aloisi, F. Astrocytes<br />

are the major intracerebral source of macrophage inflammatory protein-3a/CCL20 in<br />

relapsing experimental autoimmune encephalomyelitis and in vitro. Glia 41, 290–300<br />

(2003).<br />

47. Columba-Cabezas, S., Serafini, B., Ambrosini, E. & Aloisi, F. Lymphoid chemokines<br />

CCL19 and CCL21 are expressed in the central nervous system during experimental<br />

autoimmune encephalomyelitis: implications for the maintenance of chronic neuroinflammation.<br />

Brain Pathol. 13, 38–51 (2003).<br />

48. Rottman, J.B. et al. Leukocyte recruitment during onset of experimental allergic<br />

encephalomyelitis is CCR1 dependent. Eur. J. Immunol. 30, 2372–2377<br />

(2000).<br />

49. Muller, M. et al. CXCR3 signaling reduces the severity of experimental autoimmune<br />

encephalomyelitis by controlling the parenchymal distribution of effector and regulatory<br />

T cells in the central nervous system. J. Immunol. 179, 2774–2786<br />

(2007).<br />

50. Liu, L. et al. Severe disease, unaltered leukocyte migration, and reduced IFN-g<br />

production in CXCR3 / mice with experimental autoimmune encephalomyelitis.<br />

J. Immunol. 176, 4399–4409 (2006).<br />

51. Gaupp, S., Pitt, D., Kuziel, W.A., Cannella, B. & Raine, C.S. Experimental autoimmune<br />

encephalomyelitis (EAE) in CCR2 / mice: susceptibility in multiple strains. Am. J.<br />

Pathol. 162, 139–150 (2003).<br />

ARTICLES<br />

52. Eltayeb, S. et al. Temporal expression and cellular origin of CC chemokine receptors<br />

CCR1, CCR2 and CCR5 in the central nervous system: insight into mechanisms of<br />

MOG-induced EAE. J. Neuroinflammation 4, 14(2007).<br />

53. Matsui, M. et al. Treatment of experimental autoimmune encephalomyelitis with the<br />

chemokine receptor antagonist Met-RANTES. J. Neuroimmunol. 128, 16–22 (2002).<br />

54. Kohler, R.E., Caon, A.C., Willenborg, D.O., Clark-Lewis, I. & McColl, S.R. A role for<br />

macrophage inflammatory protein-3a/CC chemokine ligand 20 in immune priming<br />

during T cell-mediated inflammation of the central nervous system. J. Immunol. 170,<br />

6298–6306 (2003).<br />

55. Franciotta, D., Salvetti, M., Lolli, F., Serafini, B. & Aloisi, F. B cells and multiple<br />

sclerosis. Lancet Neurol. 7, 852–858 (2008).<br />

56. Kleinewietfeld, M. et al. CCR6 expression defines regulatory effector/memory-like cells<br />

within the CD25 + CD4 + T-cell subset. Blood 105, 2877–2886 (2005).<br />

57. Yamazaki, T. et al. CCR6 regulates the migration of inflammatory and regulatory T cells.<br />

J. Immunol. 181, 8391–8401 (2008).<br />

58. Steinman, L. A rush to judgment on Th17. J. Exp. Med. 205, 1517–1522 (2008).<br />

59. Sigmundsdottir, H. & Butcher, E.C. Environmental cues, dendritic cells and the<br />

programming of tissue-selective lymphocyte trafficking. Nat. Immunol. 9, 981–987<br />

(2008).<br />

60. Hickey, W.F. Basic principles of immunological surveillance of the normal central<br />

nervous system. Glia 36, 118–124 (2001).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 523


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Memory T cells in nonlymphoid tissue that provide<br />

enhanced local immunity during infection with herpes<br />

simplex virus<br />

Thomas Gebhardt, Linda M Wakim, Liv Eidsmo, Patrick C Reading, William R Heath & Francis R Carbone<br />

Effective immunity is dependent on long-surviving memory T cells. Various memory subsets make distinct contributions to<br />

immune protection, especially in peripheral infection. It has been suggested that T cells in nonlymphoid tissues are important<br />

during local infection, although their relationship with populations in the circulation remains poorly defined. Here we describe a<br />

unique memory T cell subset present after acute infection with herpes simplex virus that remained resident in the skin and in<br />

latently infected sensory ganglia. These T cells were in disequilibrium with the circulating lymphocyte pool and controlled new<br />

infection with this virus. Thus, these cells represent an example of tissue-resident memory T cells that can provide protective<br />

immunity at points of pathogen entry.<br />

Immune memory results from the action of various cellular components<br />

that combine to control infectious pathogens 1 .Tcellsarekey<br />

mediators of the memory response, although uncertainty remains as<br />

to the exact function of individual memory T cell subsets 2–6 . Memory<br />

cells have been categorized as effector memory and central memory<br />

T cells, depending on their respective ability to recirculate between<br />

secondary lymphoid organs or to enter peripheral tissues 5,7,8 . Localized<br />

infection results in rapid population expansion of effector T cells<br />

in draining lymph nodes, and although most are lost, a substantial<br />

population of circulating memory T cells survives this contraction 9 .In<br />

addition, many T cells are also present in nonlymphoid tissues 7,8 ,<br />

where their numbers may actually exceed those in the circulation 10 .<br />

The relationship between peripheral and recirculating memory cells<br />

remains mostly undefined. Peripheral T cells can be replaced from the<br />

circulation in the steady state and can be further supplemented by<br />

effector memory T cells newly recruited during infection 11 . However,<br />

an active mechanism of T cell retention may exist in nonlymphoid<br />

tissues 12–14 , and this could explain the accumulation noted at sites that<br />

have cleared an infectious virus 14–16 . Thus, once in the periphery, at<br />

least a subset of memory cells may be separated from the circulating<br />

Tcellpool.<br />

Herpes simplex virus (HSV) can infect the skin of mice, causing a<br />

primary infection that rapidly moves to the innervating sensory<br />

ganglia 17,18 . There replication is controlled in about a week and<br />

then persists as a tightly controlled latent infection 19 . The acute<br />

infection induces a robust CD8 + T cell effector and memory response<br />

that is able to control virus in the ganglia 20 . The requirement for<br />

memory T cell responses in peripheral tissues has been examined with<br />

Received 5 December 2008; accepted 10 February 2009; published online 22 March 2009; doi:10.1038/ni.1718<br />

latently infected ganglia containing a long-lived population of<br />

virus-specific memory T cells 21,22 . It was found that these memory<br />

cells were not terminally differentiated but could instead mount a<br />

secondary proliferative response after virus reactivation 22 . Notably,<br />

this occurred in the ganglia itself, in the absence of any involvement of<br />

lymphoid tissue.<br />

One issue that arose from those studies was whether the T cells<br />

residing in the ganglia were circulating memory cells or, alternatively,<br />

represented an autonomous memory population separate from the<br />

circulating pool. Although in the case of HSV, such local memory cells<br />

act to control a preexisting latent infection 21,23 , this type of T cell<br />

sequestration could conceivably result in more efficient protection<br />

against reinfection by an environmentally derived pathogenic agent.<br />

Here we provide evidence of a population of peripheral memory T cells<br />

that seemed to be in disequilibrium with the circulation. Notably, we<br />

show that such tissue-resident memory cells contributed to peripheral<br />

HSV control, thus effectively limiting the extent of renewed infection.<br />

RESULTS<br />

Nonmigrating peripheral tissue–resident memory T cells<br />

Stimulation of T cells occurs after transplantation of sensory ganglia<br />

containing both a persisting latent form of HSV and a population of<br />

CD8 + T cells specific for this pathogen 22 . In this model, the local<br />

T cells undergo secondary restimulation exclusively in the grafts<br />

during virus reactivation, with no apparent involvement of lymph<br />

nodes. This suggests that either the ganglionic cells form part of a true<br />

nonmigrating memory population separate from the recirculating<br />

memory pool, or the excision and transplantation of the ganglia<br />

Department of Microbiology and <strong>Immunology</strong>, The University of Melbourne, Melbourne Victoria, Australia. Correspondence should be addressed to W.R.H. (wrheath@<br />

unimelb.edu.au) or F.R.C. (fcarbone@unimelb.edu.au).<br />

524 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Figure 1 Tissue-resident memory T cells fail to<br />

recirculate after localized or systemic viral<br />

infection. (a) Flow cytometry of short-term BrdU<br />

uptake 7 d after transplantation of latently<br />

infected dorsal root ganglia containing gBT-I.GFP<br />

memory CD8 + T cells (resident) into recipient<br />

mice containing circulating OT-I.CD45.1 (Control)<br />

and gBT-I.CD45.1 (recruited) CD8 + memory<br />

T cells. Numbers above lines indicate percent<br />

BrdU + cells (mean ± s.e.m.). Data are from two<br />

independent experiments with three to four mice.<br />

(b–d) Analysis of gBT-I populations in grafts and<br />

spleens after transplantation of latently infected<br />

ganglia containing gBT-I.GFP memory CD8 +<br />

T cells (resident (res)) into mice containing<br />

gBT-I.CD45.1 memory CD8 + T cells (recruited<br />

(rec)), followed 6 d later by retransplantation into naive recipients and, 20 d later, intravenous challenge of these recipient mice with HSV. (b) Numbers<br />

above outlined areas indicate percent recruited cells (left) and resident cells (right) among all CD8 + cells. (c) Proportion of gBT-I populations among all CD8 +<br />

cells in the spleen. (d) Recovery of gBT-I cells from grafts. Statistical analyses, paired Student’s t-test. Data are from two independent experiments with five<br />

to six mice per group (mean and s.e.m. in c,d). (e) Recruited and tissue-resident gBT-I memory CD8 + T cells 9 d after retransplantation of latent ganglia,<br />

manipulated as described in b–d (original), together with a fresh latently infected ganglia graft containing no gBT-I cells (fresh). Statistical analyses, paired<br />

Student’s t-test. Data are from two independent experiments with six to eight mice per group (mean and s.e.m.).<br />

under the kidney capsule of recipient mice affects the normal egress of<br />

T cells from this tissue. To show that recirculating memory T cells can<br />

leave the transplanted tissues, we first recruited such a migrating<br />

population into the grafts by transplanting latent ganglia into mice<br />

that had circulating HSV-specific CD8 + memory cells. We generated<br />

these primary recipients by infecting them with a recombinant<br />

influenza virus carrying the immunodominant major histocompatibility<br />

complex class I–restricted determinant from HSV glycoprotein B<br />

(gB; Supplementary Fig. 1 online). In these experiments, we ‘genetically<br />

marked’ the graft-resident T cells with green fluorescent protein<br />

(GFP) by using cells from gBT-I.GFP-transgenic mice, whose T cells<br />

are all specific for the immunodominant HSV gB determinant 24 .<br />

T cells recruited from the circulation were from gBT-I.CD45.1transgenic<br />

mice, which meant that we could use differences in the<br />

expression of CD45 and GFP to track the respective recruited and<br />

resident memory T cell populations. It should be noted that ganglia<br />

transplantation resulted in recruitment of the circulating gBT-I cells as<br />

CD45.1<br />

Acute<br />

2.0<br />

L 8.1<br />

R 1.2 R 0.1<br />

CD8<br />

0.5 0.3 P < 0.03<br />

Original Fresh<br />

8 d<br />

30 d > 100 d 8 d 30 d > 100 d<br />

P < 0.01 P < 0.005 P < 0.01<br />

P < 0.01 P > 0.1 P > 0.1<br />

100<br />

60 4,000<br />

800<br />

800<br />

50<br />

30<br />

Endogenous<br />

CD8 + /cm 2<br />

0<br />

0<br />

0<br />

0<br />

0<br />

0<br />

L R L R L R<br />

L R L R L R<br />

8 d<br />

30 d<br />

> 100 d<br />

75<br />

P < 0.001<br />

P < 0.05<br />

30<br />

P < 0.001<br />

P < 0.001<br />

30<br />

P < 0.001<br />

P < 0.001<br />

75<br />

50<br />

*<br />

**<br />

50<br />

20<br />

20<br />

25<br />

25<br />

10<br />

10<br />

0<br />

0<br />

0<br />

0<br />

8 30 50<br />

L R Spleen L R Spleen L R Spleen Time after HSV (d)<br />

Skin Skin Skin<br />

2,000<br />

d e<br />

gBT-l of TCRtg (%)<br />

400<br />

ARTICLES<br />

400<br />

P > 0.7<br />

GFP + CD45.1 +<br />

Figure 2 Long-term retention of virus-specific CD8 + T cells at the site of previous viral infection. (a–d) Analysis of naive gBT-I.CD45.1 CD8 + T cells<br />

transferred into C57BL/6 mice 1 d before HSV-1 infection of the left flank skin at various times after inoculation (above graphs). L, left flank (skin); R, right<br />

flank (skin); C, control skin from mock-infected mice. Statistical analyses, paired Student’s t-test (b,c) or analysis of variance followed by Tukey’s post-test<br />

comparison (d). Data are from at least three independent experiments with two to four mice per group (mean and s.e.m. in b,c). (a) Numbers adjacent to<br />

outlined areas indicate percent gBT-I.CD45.1 CD8 + T cells among all events. (b,c) Isolation of gBT-I.CD45.1 and endogenous CD8 + T cells. The number<br />

of gBT-I.CD45.1 T cells in skin from mock-infected mice was below the limit of detection. (d) Frequency of gBT-I.CD45.1 cells among all CD8 + Tcells.<br />

(e) Proportion of gBT-I.CD45.1 T cells among all T cell antigen receptor–transgenic T cells (TCRtg; gBT-I plus OT-I) after transfer of in vitro–activated gBT-<br />

I.CD45.1 and OT-I.CD45.1 CD8 + T cells together into C57BL/6 mice 2–4 d after HSV skin infection. *, P o 0.01; **, P o 0.001 (paired Student’s t-test).<br />

Data are from two to three independent experiments with three mice per group (mean and s.e.m.).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 525<br />

Rec<br />

Res<br />

Skin<br />

Spleen


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Figure 3 HSV-specific CD8 + Tcells<br />

‘preferentially’ localize to the epidermal skin<br />

layer covering the previously infected site and<br />

express the intraepithelial cell marker CD103.<br />

(a,b) Microscopy of skin after transfer of naive<br />

gBT-I.GFP CD8 + T cells into C57BL/6 mice<br />

1 d before skin infection with HSV-1.<br />

(a) Immunohistochemistry of the distribution<br />

of virus-specific gBT-I.GFP CD8 + T cells in skin<br />

obtained from previously infected areas 13 d after<br />

infection, then stained with DAPI and antibody to<br />

keratin 1 expressed in keratinocytes in the upper<br />

layer of the epidermis. Control, tissue from<br />

contralateral flanks. Arrowheads indicate gBT-<br />

I.GFP CD8 + T cells in the dermal-epidermal<br />

region. Scale bars, 100 mm. (b) Epidermal sheets<br />

from previously infected flank skin 25 d after<br />

infection, when the epithelium had reformed. The<br />

density of gBT-I.GFP CD8 + T cells is shown in the<br />

central area of lesion resolution (lesion center), on the edge of the lesion (lesion periphery) and in an unaffected area (control). Virus-specific T cells could not<br />

be detected by this method in unaffected areas in infected mice (control). Scale bars, 50 mm. (c) CD103 expression by memory gBT-I.CD45.1 CD8 + T cells<br />

isolated from spleen and epidermal sheets 13 weeks after infection of the skin with HSV. Numbers above lines (right) indicate percent CD103 + Infected<br />

Control<br />

Spleen<br />

1.5<br />

b Lesion center<br />

gBT-l (GFP) Keratin 1 DAPl<br />

Lesion periphery Control<br />

Vα2 CD103<br />

Epidermis<br />

98.1<br />

Vα2 CD103<br />

gBT-l (GFP) DAPl<br />

cells. Data are<br />

from one experiment representative of three.<br />

HSV-infected recipients of primary grafts confirmed that only the<br />

recruited T cells underwent considerable population expansion after<br />

systemic challenge (Supplementary Fig. 2 online).<br />

To demonstrate that the resident cells had a migration defect and<br />

did not simply lack the ability to mount a systemic response, we again<br />

recruited memory T cells into reactivating ganglia and, on day 6 after<br />

transplantation, collected the graft and retransplanted it into a naive<br />

recipient together with ‘fresh’ latent ganglia in a separate site under the<br />

same kidney capsule (Supplementary Fig. 3 online). On day 9 after<br />

retransplantation, we collected each graft and assessed if the infiltrating<br />

memory T cells (CD45.1 + ) and tissue-residing T cells (GFP + ) were<br />

able to enter the neighboring ganglia, which represented a fresh site of<br />

HSV infection. We detected a substantial population of recruited<br />

memory CD8 + T cells in the fresh graft, in contrast to those resident<br />

memory T cells carried by the original graft (Fig. 1e). Thus, the<br />

resident memory T cells did not migrate to a nearby yet anatomically<br />

distinct site, even though it was equivalent to the tissue of origin for<br />

these cells.<br />

Enrichment for ‘biased’ T cells after skin infection<br />

The data reported above suggested that HSV-specific T cells resident<br />

in the ganglia seemed functionally different from those found in<br />

Figure 4 Skin-resident, virus-specific CD8 + T cells are phenotypically and<br />

functionally different from their circulating counterparts. Analysis of naive<br />

gBT-I.CD45.1 CD8 + T cells transferred into C57BL/6 mice 1 d before<br />

infection of the skin with HSV-1, followed by a period of rest for the<br />

establishment of immunological memory. (a) Flow cytometry of memory gBT-<br />

I.CD45.1 CD8 + T cells isolated from brachial lymph nodes (bLN), spleen<br />

and skin 30–386 d after infection (median, 51 d). Data represent four<br />

independent experiments with pooled cell preparations from six to seventeen<br />

mice per group (mean and s.e.m.). (b) Expression of CD49a (VLA-1) on<br />

gBT-I.CD45.1 CD8 + T cells isolated 50 d after infection. Numbers above<br />

outlined areas indicate percent CD49a + cells. Isotype, isotype-matched<br />

control antibody. Data are from three independent experiments with two to<br />

four mice per group (mean ± s.e.m.). (c) Flow cytometry of the homeostatic<br />

proliferation of cells from ‘memory mice’ (43–428 d after infection;<br />

median, 72 d) treated for 7 d with BrdU. Data are from six individual<br />

experiments with three to eight mice per group (mean and s.e.m.).<br />

Statistical analyses, repeated measures analysis of variance followed<br />

by Tukey’s post-test comparison.<br />

a c<br />

CD45.1<br />

CD45.1<br />

the circulation. We reasoned that an analogous population of tissueresident<br />

memory T cells would also be found in other peripheral sites<br />

of infection, notably the skin. Evidence for this took the form of<br />

‘preferential’ retention of virus-specific T cells in flank skin involved in<br />

overt HSV disease relative to that in the uninvolved contralateral<br />

flanks. There was ‘preferential’ enrichment for gBT-I T cells in<br />

flank skin ipsilateral to the origin of infection at all times assessed<br />

(Fig. 2a,b). In contrast, there was no significant difference in the<br />

number of endogenous CD8 + T cells in opposing flanks after clearance<br />

of replicating virus, which occurs by day 8 after infection 18 (Fig. 2c). It<br />

should be noted that gBT-I cells represented a relatively small<br />

proportion of all CD8 + T cells in the skin, especially at these later<br />

times (day 30 or beyond; Fig. 2d). The accumulation of HSV-specific<br />

T cells was not restricted to transgenic cells, because repeat experiments<br />

with tetramer staining to detect HSV-specific T cells showed<br />

similar biases (Supplementary Fig. 4 online). Retention of T cells in<br />

the skin had a nonspecific component, as demonstrated by the use of<br />

attenuated recombinant viruses (HSV strains rgB and rgB-L8A) 25<br />

(Supplementary Fig. 5 online). The number of long-term gB-specific<br />

T cells was the same in flanks infected with gB-sufficient rgB virus and<br />

control gB-deficient rgB-L8A in dual-inoculated mice. Even scarification<br />

alone led to equivalent infiltration into rgB-infected and control<br />

a<br />

CD62L hi gBT-l (%)<br />

b<br />

Isotype CD49a<br />

100 P < 0.01<br />

75<br />

50<br />

25<br />

0<br />

bLN Spleen Skin<br />

bLN<br />

35.2 ± 16.2<br />

CD45.1<br />

P < 0.001 P < 0.05<br />

Spleen<br />

CD122 hi gBT-l (%)<br />

100<br />

75<br />

50<br />

25<br />

0<br />

Skin<br />

42.5 ± 9.8 85.1 ± 1.7<br />

Cells<br />

Cells<br />

P < 0.01<br />

P > 0.05 100<br />

75<br />

50<br />

25<br />

0<br />

P < 0.001<br />

bLN Spleen Skin bLN Spleen Skin<br />

CD69 + gBT-l (%)<br />

c<br />

BrdU + gBT-l (%)<br />

20<br />

10<br />

0<br />

P < 0.01<br />

P < 0.05<br />

bLN Spleen Skin<br />

526 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

CD45.1<br />

b<br />

gBT-l/cm 2<br />

Recipient<br />

G 0.02 C 0 L 0.6<br />

0.1<br />

750 6<br />

6<br />

500<br />

4<br />

4<br />

250<br />

0<br />

2<br />

2<br />

L R C C C L R<br />

C C L R L R<br />

1° HSV 1° HSV: – – +<br />

1° HSV: – – + +<br />

gBT-l: – + +<br />

gBT-l: – + + +<br />

α-CD8: – – – +<br />

described in d or also with depleting anti-CD8 before reinfection with HSV. Data are pooled from three independent experiments with two to five mice per<br />

group. Each symbol represents an individual mouse; small horizontal lines indicate the mean (a,b,d,e). Statistical analyses, one-way analysis of variance<br />

followed by Tukey’s post-test comparison.<br />

6<br />

4<br />

2<br />

50 d<br />

P < 0.001<br />

P < 0.001 P < 0.001<br />

P < 0.001<br />

6<br />

4<br />

2<br />

P < 0.001<br />

PFU/sample (log)<br />

>100 d<br />

P < 0.001<br />

c d e<br />

b<br />

PFU/sample (log)<br />

PFU/sample (log)<br />

6<br />

4<br />

2<br />

ARTICLES<br />

P < 0.001<br />

P < 0.001<br />

P > 0.05<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 527


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

later times (day 35; Supplementary Fig. 7b,c) showed that T cells were<br />

present only in the transplanted tissue. This was consistent with the<br />

possibility that tissue-resident memory cells are an autonomous<br />

population. Moreover, severing of the nerves that link the skin to<br />

the infected ganglia also indicated that the persistence of memory<br />

T cells was not due to continuous feeding of virus or antigen to the<br />

skin from latently infected sensory nerve bodies undergoing spontaneous<br />

bouts of reactivation, which in any case is not a property of<br />

mouse infection with HSV 19 .<br />

Immune protection by tissue-resident memory T cells<br />

As well having as ‘preferential’ accumulation of memory T cells,<br />

ipsilateral flanks had enhanced protection against subsequent infection.<br />

To demonstrate this without antibody interference, we infected B cell–<br />

deficient mMT mice in the flank with HSV and challenged them at<br />

various times afterward. The ipsilateral flank showed 100-fold better<br />

control of infection than that of the control contralateral flank up to<br />

50 d after the first infection (Fig. 6a). We noted a bias toward ipsilateral<br />

protection even at more than 100 d after infection, although this was<br />

less than that present at the earlier times. The greater protection of<br />

ipsilateral flanks was virus specific, as it did not extend to heterologous<br />

infection with vaccinia virus (Supplementary Fig. 8 online).<br />

Flank immunity was T cell dependent, because elimination of CD4 +<br />

and CD8 + cells by in vivo antibody depletion diminished antiviral<br />

protection (Fig. 6b). Antibody to CD4 (anti-CD4) had a much greater<br />

effect than did depletion with anti-CD8, which suggested that CD4 +<br />

T cells dominated control of skin infection 33 or that help was critical<br />

to the contribution of CD8 + T cells to virus elimination 34,35 . To show<br />

that local CD8 + T cells could provide antiviral protection, we injected<br />

mice deficient in recombination-activating gene 1 (Rag1 –/– mice) with<br />

in vitro–activated gBT-I CD8 + T cells 2–4 d after flank infection with<br />

HSV. Previously infected skin retained more T cells than did the<br />

contralateral flank (Fig. 6c), and this was associated with ‘preferential’<br />

protection against subsequent infection with HSV (Fig. 6d). Such<br />

biased protection was eliminated by treatment with anti-CD8<br />

(Fig. 6e), which showed that it was indeed mediated by the transferred<br />

T cells. In the Rag1 –/– mice, CD8 + T cells act in the complete absence<br />

of CD4 + T cell help, most likely because their numbers are in excess of<br />

those that normally persist long term after infection of wild-type<br />

hosts 36 . Nevertheless, these data collectively show enhanced local<br />

protection by T cells that accumulate in tissues after resolution of a<br />

previous infection.<br />

DISCUSSION<br />

There are obvious advantages provided by the existence of a sequestered<br />

T cell population for persisting, reactivating infections such as<br />

the one we used here. CD8 + T cells are known to maintain HSV in a<br />

latent state in the ganglia by constant expression of effector molecules<br />

such as granzyme B and interferon-g 37,38 . Whether such populations<br />

exist in acute infection systems, especially after the virus has been<br />

completely cleared, remains to be determined. There are reports of<br />

enrichment for memory T cells 7,8 at sites of initial pathogen infection<br />

well after clearance has been achieved 14–16 . That is similar to what we<br />

found after infection of the skin with HSV, with a notable imbalance<br />

of antigen-specific memory T cells in regions ipsilateral or contralateral<br />

to the original site of inoculation. Published studies using<br />

rapidly cleared viruses have excluded the possibility of prolonged<br />

infection as the mechanism of persistent accumulation 13,14,16,30 .An<br />

argument could conceivably be mounted for the possibility that<br />

ongoing gB-specific stimulation impedes the release of circulating<br />

memory T cells, given the nature of HSV latency. However, experiments<br />

with bone marrow–chimeric mice have shown that even in the sensory<br />

ganglia, T cells persist in the absence of specific recognition of infected<br />

neurons 39 . Retention in the skin by ongoing virus reactivation seems<br />

even less likely, given our demonstration that resident T cells survived<br />

transplantation and thus separation from the neuronal source of<br />

infection. Unlike humans, mice do not show spontaneous reactivation<br />

of HSV 19 , and skin-resident T cells show no signs of continuous<br />

antigen stimulation, such as the upregulation of granzymes, which is<br />

evident in T cells persisting in the ganglia 40 . Thus, although ongoing<br />

antigen presentation remains a formal possibility, it is unlikely to fully<br />

explain the retention of HSV-specific peripheral T cells, especially in<br />

the skin.<br />

Although localized infection can ‘imprint’ ‘preferential’ T cell<br />

migration 41,42 , and this could explain other cases of local accumulation<br />

of T cells, it is difficult to see how such tissue-tropic infiltration<br />

would contribute to the ‘side-biased’ T cell accumulation in the flank<br />

skin shown here. Moreover, the lack of T cell migration between the<br />

two ganglia transplanted under the same kidney capsule indicates that<br />

these tissue-resident T cells simply did not migrate between anatomically<br />

distinct but otherwise equivalent tissues. As a consequence, we<br />

propose these HSV-specific T cells represent a sessile or nonmigratory<br />

memory subset that we define as ‘tissue-resident memory T cells’.<br />

These cells are either separate from the recirculating pool or retained<br />

long enough in peripheral tissues to infrequently exchange with this<br />

population. Therefore, they are different from other memory T cells,<br />

including the central and effector memory T cells found in the<br />

wider circulation 4,5 .<br />

CD8 + T cells in the ganglia can mount a stand-alone new immune<br />

response that includes helper T cell–dependent proliferation and<br />

stimulation by dendritic cells recruited from the circulation 22 .<br />

Together with our data here, this suggests that tissue-resident memory<br />

T cells are not terminally differentiated but can self-renew either<br />

through homeostatic or antigen-stimulatory mechanisms. Furthermore,<br />

they suggest that studies confined to circulating effector<br />

memory and central memory T cell populations may underestimate<br />

the totality of immunological memory in a given person.<br />

Persisting T cells have been found in human skin that has cleared<br />

HSV infection 43 , and embedded T cells are associated with recurring<br />

allergic responses in humans, known as ‘fixed drug eruptions’, driven<br />

by CD8 + T cells in the epidermis that do not disseminate throughout<br />

the skin 44,45 . In addition, pathogenic T cells persist in pre-psoriatic<br />

human skin, and these cells can be transferred during mouse xenotransplantation<br />

46 . In those last two examples, the residing T cells<br />

express markers common to the tissue-resident memory T cells in our<br />

study, such as VLA-1 and CD103; the latter are important for<br />

epithelial localization 26 . Finally, we have shown that these memory<br />

T cells can efficiently inhibit HSV replication in peripheral nonlymphoid<br />

tissues, including those at the body surface. Thus, our<br />

study raises the possibility that resident memory cells can be exploited<br />

as a means of controlling infection at body surfaces that act as portals<br />

of entry for invading pathogens.<br />

METHODS<br />

Mice. C57BL/6, B6.SJL-PtprcaPep3b/BoyJ (B6.CD45.1), gBT-I, gBT-I<br />

B6.CD45.1 (gBT-I.CD45.1), gBT-I.GFP, OT-I B6.CD45.1 (OT-I.CD45.1),<br />

mMT and Rag1 –/– /Je (Rag1 –/– ) mice were bred in the Department of Microbiology<br />

and <strong>Immunology</strong> of The University of Melbourne in Parkville,<br />

Australia. The gBT-I and OT-I mice are CD8 + T cell receptor-transgenic mice<br />

that recognize the H-2K b -restricted HSV-1 gB epitope of amino acids 498–505<br />

(gB(498–505)) and the ovalbumin-derived epitope of amino acids 257–264<br />

(OVA(257–264)), respectively. B6.CD45.1 mice express the congenic marker<br />

528 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

CD45.1, in contrast to C57BL/6, mMT and Rag1 –/– mice, which express CD45.2.<br />

The gBT-I.CD45.1 and OT-I.CD45.1 mice were F1 offspring expressing both<br />

CD45.1 and CD45.2. Animal experiments were approved by The University of<br />

Melbourne Animal Ethics Committee.<br />

Virus infection. Viruses used were HSV-1 KOS, KOS.CreTK – , KOS.rgB 25 (rgB)<br />

and KOS.rgB-L8A (rgB-L8A) 25 , as well as WSN/NA/gB (flu-gB; from<br />

S. Tevethia), WSN/NA/OVA (flu-OVA; from P. Doherty) and vaccinia-NP<br />

(from J. Yewdell). The recombinant influenza viruses express gB(498–505)<br />

(flu-gB) or OVA(257–264) (flu-OVA) in their neuraminidase stalk. Routes<br />

of infections were epidermal (skin; 1 10 6 plaque-forming units (PFU)),<br />

subcutaneous (5 10 2 PFU influenza virus and 1 10 6 PFU HSV) and<br />

intravenous (1 10 5 PFU). Skin was infected with HSV-1 KOS and<br />

KOS.CreTK – by scarification as described 18 . The HSV-1 KOS strain caused<br />

typical zosteriform lesions spreading from the primary inoculation site along<br />

the affected dermatomes on the same flank. The other HSV strains used had<br />

attenuated skin replication and did not cause any visible zosteriform lesions.<br />

Skin was infected with vaccinia-NP (5 10 7 PFU) by inoculation of flank skin<br />

as for HSV infection. For subcutaneous infection, 25 ml of a virus suspension<br />

was injected into foot hocks.<br />

In vitro activation and adoptive transfer of transgenic CD8 + T cells. All<br />

adoptive transfers of gBT-I and OT-I cells were done intravenously with<br />

lymph node suspensions or in vitro–generated effector splenocytes activated<br />

by peptide-pulsed targets as described 18 . First, gBT-I or OT-I splenocytes<br />

were cultured for 4 d together with gB(498–505)- or OVA(257–264)-pulsed<br />

(1 mg/ml) C57BL/6 splenocytes, respectively. On days 2 and 3, cultures were split<br />

1: 2 and interleukin 2 (10 U/ml; PeproTech) was added. Then, 1 10 6 to 1.5<br />

10 6 effector cells derived from those cultures were transferred intravenously<br />

into mice whose skin was infected with HSV-1. Adoptive transfer of naive<br />

gBT-I.CD45.1 cells (5 10 4 ) resulted in a frequency of gBT-I cells roughly ten<br />

times higher than that of their endogenous HSV-1-specific counterparts 47 .<br />

Flow cytometry of tissue cells. Mice were killed by carbon dioxide administration<br />

and were perfused with 10 ml Hank’s buffered-salt solution (Media<br />

Preparation Unit, The University of Melbourne). Lymph nodes and spleens<br />

were collected and disrupted by passage through a wire mesh. Skin tissue<br />

(1–2 cm 2 ) and ganglia were removed, were chopped into small fragments and<br />

were incubated for 90 min at 37 1C in Eagle’s minimum essential medium<br />

containing 2% (vol/vol) FCS (Gibco BRL), collagenase (3 mg/ml; Worthington)<br />

and DNase (5 mg/ml; Sigma). For minimization of the digestion of certain<br />

surface receptors, this incubation period was decreased to 30 min in some<br />

phenotyping experiments. Thereafter, cell suspensions were filtered twice<br />

through nylon meshes (pore size, 70 mm and 20 mm) and were stained for<br />

30 min with antibodies for flow cytometry. The following antibodies were from<br />

BD Pharmingen: allophycocyanin- and phycoerythrin-indotricarbocyanine–<br />

conjugated anti-CD8a (53-6.7), fluorescein isothiocyanate– and phycoerythrin-conjugated<br />

anti-CD45.1 (A20), fluorescein isothiocyanate–conjugated<br />

anti-CD45.2 (104), phycoerythrin-conjugated anti-CD4 (GK1.5), anti-CD49a<br />

(Ha31/8), anti–keyhole limpet hemocyanin hamster IgG2 553962, fluorescein<br />

isothiocyanate–conjugated anti-CD44 (IM7), phycoerythrin-conjugated anti-<br />

Va2 (B20.1), phycoerythrin-conjugated anti-CD69 (H1.2F3), phycoerythrinconjugated<br />

anti-CD122 (TM-b1), fluorescein isothiocyanate–conjugated<br />

anti-CD62L (Mel-14), fluorescein isothiocyanate–conjugated antibody to<br />

Armenian and Syrian hamster IgG (G192-1), and anti-CD16/32 (2.4G2).<br />

Allophycocyanin-conjugated anti-CD45.1 (A20), phycoerythrin- and allophycocyanin–Alexa<br />

Fluor 750–conjugated anti-CD45.2 (104) and fluorescein<br />

isothiocyanate–conjugated anti-CD103 (2E7) were from eBiosciences. H-2K b<br />

gB(498–505)–phycoerythrin tetramer was generated at the Department of<br />

Microbiology and <strong>Immunology</strong> of The University of Melbourne. Dead cells<br />

were excluded by propidium iodide staining. Sphero calibration particles (BD<br />

Pharmingen) were added to samples to allow calculation of cell numbers.<br />

For analysis of the homeostatic turnover of memory cells, 1 mg sterile BrdU<br />

(5-bromodeoxyuridine) was injected intraperitoneally on 7 consecutive days.<br />

For short-pulse experiments, 1.25 mg BrdU was injected and mice were<br />

analyzed 1 h later. Uptake was detected with a BrdU Flow kit according<br />

to the manufacturer’s instructions (BD Pharmingen). A FACSCalibur or<br />

ARTICLES<br />

FACSCanto II, plus Cell QuestPro software (BD Biosciences) and FlowJo<br />

software (TreeStar), were used for analysis.<br />

Transplantation experiments. Latently infected dorsal root ganglia were<br />

transplanted under the kidney capsules of syngeneic recipients as described 22 .<br />

For transplantation of previously HSV-infected skin tissue, donor mice were<br />

killed by carbon dioxide administration 2–4 weeks after HSV infection of the<br />

left flank skin. After flank skin was clipped and then treated with Veet<br />

depilation cream (Reckitt Benckiser), mice were perfused through the left<br />

ventricle with 10 ml Hank’s balanced-salt solution. Donor skin tissue (previously<br />

infected skin with an area of 1–1.5 cm 1 cm) was removed in aseptic<br />

conditions for storage in ice-cold buffer. Naive recipient mice were anesthetized<br />

and the skin of their flanks and backs was clipped, was treated with Veet cream<br />

and was disinfected with 70% (vol/vol) ethanol. Graft beds were prepared by<br />

carefully snipping away of the skin layer with sterile curved scissors. Finally,<br />

graft tissue was placed onto the graft bed and was secured with Jelonet gauze<br />

dressing (Smith & Nephew) as well as Micropore and Transpore tape (both<br />

from 3M Health Care). For the next 4 d, recipient mice were given analgesics in<br />

their drinking water (paracetamol; 1.3 mg/ml). After 8–10 d, bandages were<br />

removed and engraftment of transplanted skin was monitored daily.<br />

Immunohistochemistry. Skin tissues were fixed for 2 h on ice in 4% (vol/vol)<br />

paraformaldehyde and 10% (wt/vol) sucrose in PBS and were frozen in Tissue<br />

Tek (Sakura Finetek). Sections were cut on a cryomicrotome (CM3050S; Leica),<br />

were air-dried and were fixed in ice-cold acetone. Epidermal sheets were<br />

prepared by incubation of full-thickness skin with Dispase II (2.5 mg/ml;<br />

Roche), followed by manual separation of the dermal and epidermal layers and<br />

fixation in 4% (vol/vol) paraformaldehyde and 10% (wt/vol) sucrose in PBS.<br />

Keratin expression was visualized by incubation with polyclonal rabbitanti–mouse<br />

keratin 1 (AF 109; Covance) followed by Alexa Fluor 594–<br />

conjugated donkey anti-rabbit (A21207; Molecular Probes). Slides were<br />

mounted with Vectashield containing DAPI (4,6-diamidino-2-phenylindole;<br />

Vector Laboratories). Images were acquired with a fluorescence microscope<br />

(DMI 4000B) and digital camera (DFC 490) with IM50 software (all from<br />

Leica) and the Adobe Photoshop program.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank S. Tevethia (Pennsylvania State University) for WSN/NA/gB;<br />

P. Doherty (The University of Melbourne) for WSN/NA/OVA; J. Yewdell (U.S.<br />

National Institutes of Health) for vaccinia-NP.; D. Masopust for discussions; and<br />

H. Kosaka for suggesting that fixed drug eruptions are caused by resident<br />

T cells. Supported by the Australian National Health and Medical Research<br />

Council, the Howard Hughes Medical Institute and the German Research<br />

Foundation (GE1666/1-1 to T.G.).<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Welsh, R.M., Selin, L.K. & Szomolanyi-Tsuda, E. Immunological memory to viral<br />

infections. Annu. Rev. Immunol. 22, 711–743 (2004).<br />

2. Bachmann, M.F., Kundig, T.M., Hengartner, H. & Zinkernagel, R.M. Protection against<br />

immunopathological consequences of a viral infection by activated but not resting<br />

cytotoxic T cells: T cell memory without ‘‘memory T cells’’? Proc. Natl. Acad. Sci. USA<br />

94, 640–645 (1997).<br />

3. Bachmann, M.F., Wolint, P., Schwarz, K. & Oxenius, A. Recall proliferation potential of<br />

memory CD8 + T cells and antiviral protection. J. Immunol. 175, 4677–4685<br />

(2005).<br />

4. Sallusto, F., Geginat, J. & Lanzavecchia, A. Central memory and effector memory T cell<br />

subsets: function, generation, and maintenance. Annu. Rev. Immunol. 22, 745–763<br />

(2004).<br />

5. Sallusto, F., Lenig, D., Forster, R., Lipp, M. & Lanzavecchia, A. Two subsets of memory<br />

T lymphocytes with distinct homing potentials and effector functions. <strong>Nature</strong> 401,<br />

708–712 (1999).<br />

6. Wherry, E.J. et al. Lineage relationship and protective immunity of memory CD8 T cell<br />

subsets. Nat. Immunol. 4, 225–234 (2003).<br />

7. Masopust, D., Vezys, V., Marzo, A.L. & Lefrancois, L. Preferential localization of effector<br />

memory cells in nonlymphoid tissue. Science 291, 2413–2417 (2001).<br />

8. Reinhardt, R.L., Khoruts, A., Merica, R., Zell, T. & Jenkins, M.K. Visualizing the<br />

generation of memory CD4 T cells in the whole body. <strong>Nature</strong> 410, 101–105 (2001).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 529


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

9. Kaech, S.M., Wherry, E.J. & Ahmed, R. Effector and memory T-cell differentiation:<br />

implications for vaccine development. Nat. Rev. Immunol. 2, 251–262<br />

(2002).<br />

10. Clark, R.A. et al. The vast majority of CLA + T cells are resident in normal skin.<br />

J. Immunol. 176, 4431–4439 (2006).<br />

11. Hikono, H. et al. T-cell memory and recall responses to respiratory virus infections.<br />

Immunol. Rev. 211, 119–132 (2006).<br />

12. Harris, N.L., Watt, V., Ronchese, F. & Le Gros, G. Differential T cell function and fate in<br />

lymph node and nonlymphoid tissues. J. Exp. Med. 195, 317–326 (2002).<br />

13. Klonowski, K.D. et al. Dynamics of blood-borne CD8 memory T cell migration in vivo.<br />

Immunity 20, 551–562 (2004).<br />

14. Masopust, D. et al. Activated primary and memory CD8 T cells migrate to nonlymphoid<br />

tissues regardless of site of activation or tissue of origin. J. Immunol. 172, 4875–4882<br />

(2004).<br />

15. Hawke, S., Stevenson, P.G., Freeman, S. & Bangham, C.R. Long-term persistence of<br />

activated cytotoxic T lymphocytes after viral infection of the central nervous system.<br />

J. Exp. Med. 187, 1575–1582 (1998).<br />

16. Hogan, R.J. et al. Activated antigen-specific CD8 + T cells persist in the lungs<br />

following recovery from respiratory virus infections. J. Immunol. 166, 1813–1822<br />

(2001).<br />

17. Simmons, A. & Nash, A.A. Zosteriform spread of herpes simplex virus as a model of<br />

recrudescence and its use to investigate the role of immune cells in prevention of<br />

recurrent disease. J. Virol. 52, 816–821 (1984).<br />

18. van Lint, A. et al. Herpes simplex virus-specific CD8 + T cells can clear established lytic<br />

infections from skin and nerves and can partially limit the early spread of virus after<br />

cutaneous inoculation. J. Immunol. 172, 392–397 (2004).<br />

19. Stanberry, L.R. Pathogenesis of herpes simplex virus infection and animal models for<br />

its study. Curr.Top.Microbiol.Immunol.179, 15–30 (1992).<br />

20. Simmons, A. & Tscharke, D.C. Anti-CD8 impairs clearance of herpes simplex virus from<br />

the nervous system: implications for the fate of virally infected neurons. J. Exp. Med.<br />

175, 1337–1344 (1992).<br />

21. Khanna, K.M., Bonneau, R.H., Kinchington, P.R. & Hendricks, R.L. Herpes simplex<br />

virus-specific memory CD8 + T cells are selectively activated and retained in latently<br />

infected sensory ganglia. Immunity 18, 593–603 (2003).<br />

22. Wakim, L.M., Waithman, J., van Rooijen, N., Heath, W.R. & Carbone, F.R. Dendritic<br />

cell-induced memory T cell activation in nonlymphoid tissues. Science 319, 198–202<br />

(2008).<br />

23. Liu, T., Khanna, K.M., Chen, X., Fink, D.J. & Hendricks, R.L. CD8 + Tcellscanblock<br />

herpes simplex virus type 1 (HSV-1) reactivation from latency in sensory neurons.<br />

J. Exp. Med. 191, 1459–1466 (2000).<br />

24. Mueller, S.N., Heath, W., McLain, J.D., Carbone, F.R. & Jones, C.M. Characterization of<br />

two TCR transgenic mouse lines specific for herpes simplex virus. Immunol. Cell Biol.<br />

80, 156–163 (2002).<br />

25. Stock, A.T., Jones, C.M., Heath, W.R. & Carbone, F.R. CTL response compensation for<br />

the loss of an immunodominant class I-restricted HSV-1 determinant. Immunol. Cell<br />

Biol. 84, 543–550 (2006).<br />

26. Cepek, K.L. et al. Adhesion between epithelial cells and T lymphocytes mediated by<br />

E-cadherin and the aEb7 integrin. <strong>Nature</strong> 372, 190–193 (1994).<br />

27. Pauls, K. et al. Role of integrin a E(CD103)b 7 for tissue-specific epidermal localization<br />

of CD8 + T lymphocytes. J. Invest. Dermatol. 117, 569–575 (2001).<br />

28. Hemler, M.E. VLA proteins in the integrin family: structures, functions, and their role on<br />

leukocytes. Annu. Rev. Immunol. 8, 365–400 (1990).<br />

29. Ray, S.J. et al. The collagen binding a1b1 integrin VLA-1 regulates CD8 T cellmediated<br />

immune protection against heterologous influenza infection. Immunity 20,<br />

167–179 (2004).<br />

30. Ely, K.H., Cookenham, T., Roberts, A.D. & Woodland, D.L. Memory T cell populations in<br />

the lung airways are maintained by continual recruitment. J. Immunol. 176, 537–543<br />

(2006).<br />

31. Hogan, R.J. et al. Long-term maintenance of virus-specific effector memory CD8 + Tcells<br />

in the lung airways depends on proliferation. J. Immunol. 169, 4976–4981 (2002).<br />

32. Zammit, D.J., Turner, D.L., Klonowski, K.D., Lefrancois, L. & Cauley, L.S. Residual<br />

antigen presentation after influenza virus infection affects CD8 T cell activation and<br />

migration. Immunity 24, 439–449 (2006).<br />

33. Nash, A.A. et al. Different roles for L3T4 + and Lyt2 + T cell subsets in the control of an<br />

acute herpes simplex virus infection of the skin and nervous system. J. Gen. Virol. 68,<br />

825–833 (1987).<br />

34. Jennings, S.R., Bonneau, R.H., Smith, P.M., Wolcott, R.M. & Chervenak, R. CD4positive<br />

T lymphocytes are required for the generation of the primary but not the<br />

secondary CD8-positive cytolytic T lymphocyte response to herpes simplex virus in<br />

C57BL/6 mice. Cell. Immunol. 133, 234–252 (1991).<br />

35. Smith, C.M. et al. Cognate CD4 + T cell licensing of dendritic cells in CD8 + Tcell<br />

immunity. Nat. Immunol. 5, 1143–1148 (2004).<br />

36. Mintern, J.D., Davey, G.M., Belz, G.T., Carbone, F.R. & Heath, W.R. Cutting edge:<br />

precursor frequency affects the helper dependence of cytotoxic T cells. J. Immunol.<br />

168, 977–980 (2002).<br />

37. Knickelbein, J.E. et al. Noncytotoxic lytic granule-mediated CD8 + T cell inhibition of<br />

HSV-1 reactivation from neuronal latency. Science 322, 268–271 (2008).<br />

38. Liu, T., Khanna, K.M., Carriere, B.N. & Hendricks, R.L. Gamma interferon can prevent<br />

herpes simplex virus type 1 reactivation from latency in sensory neurons. J. Virol. 75,<br />

11178–11184 (2001).<br />

39. van Lint, A.L. et al. Latent infection with herpes simplex virus is associated with<br />

ongoing CD8 + T-cell stimulation by parenchymal cells within sensory ganglia. J. Virol.<br />

79, 14843–14851 (2005).<br />

40. Mintern, J.D., Guillonneau, C., Carbone, F.R., Doherty, P.C. & Turner, S.J. Cutting edge:<br />

tissue-resident memory CTL down-regulate cytolytic molecule expression following<br />

virus clearance. J. Immunol. 179, 7220–7224 (2007).<br />

41. Campbell, D.J. & Butcher, E.C. Rapid acquisition of tissue-specific homing phenotypes<br />

by CD4 + T cells activated in cutaneous or mucosal lymphoid tissues. J. Exp. Med. 195,<br />

135–141 (2002).<br />

42. Mora, J.R. & von Andrian, U.H. T-cell homing specificity and plasticity: new concepts<br />

and future challenges. Trends Immunol. 27, 235–243 (2006).<br />

43. Zhu, J. et al. Virus-specific CD8 + T cells accumulate near sensory nerve endings in<br />

genital skin during subclinical HSV-2 reactivation. J. Exp. Med. 204, 595–603<br />

(2007).<br />

44. Mizukawa, Y. et al. Direct evidence for interferon-g production by effector-memory-type<br />

intraepidermal T cells residing at an effector site of immunopathology in fixed drug<br />

eruption. Am. J. Pathol. 161, 1337–1347 (2002).<br />

45. Shiohara, T. & Moriya, N. Epidermal T cells: their functional role and disease relevance<br />

for dermatologists. J. Invest. Dermatol. 109, 271–275 (1997).<br />

46. Boyman, O. et al. Spontaneous development of psoriasis in a new animal model shows<br />

an essential role for resident T cells and tumor necrosis factor-a. J. Exp. Med. 199,<br />

731–736 (2004).<br />

47. Stock, A.T. et al. Optimization of TCR transgenic T cells for in vivo tracking of immune<br />

responses. Immunol. Cell Biol. 85, 394–396 (2007).<br />

530 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

T cell antigen receptor signaling and immunological<br />

synapse stability require myosin IIA<br />

Tal Ilani 1 , Gaia Vasiliver-Shamis 2 , Santosh Vardhana 2 , Anthony Bretscher 1,3 & Michael L Dustin 2,3<br />

Immunological synapses are initiated by signaling in discrete T cell antigen receptor microclusters and are important for the<br />

differentiation and effector functions of T cells. Synapse formation involves the orchestrated movement of microclusters toward<br />

the center of the contact area with the antigen-presenting cell. Microcluster movement is associated with centripetal actin flow,<br />

but the function of motor proteins is unknown. Here we show that myosin IIA was necessary for complete assembly and<br />

movement of T cell antigen receptor microclusters. In the absence of myosin IIA or its ATPase activity, T cell signaling was<br />

interrupted ‘downstream’ of the kinase Lck and the synapse was destabilized. Thus, T cell antigen receptor signaling and the<br />

subsequent formation of immunological synapses are active processes dependent on myosin IIA.<br />

The specific and long-lasting interface between a T cell and an antigenpresenting<br />

cell (APC), called the ‘immunological synapse’, is critical<br />

for the afferent and efferent limbs of the adaptive immune response 1,2 .<br />

The supramolecular organization of the immunological synapse was<br />

described more than a decade ago 3–5 , yet the mechanisms leading to<br />

its formation and persistence are unknown. No function for motor<br />

proteins in signaling and synapse formation by cells of the immune<br />

system has been established 6,7 .<br />

The first step in synapse formation is engagement of the T cell<br />

antigen receptor (TCR) with the appropriate major histocompatibility<br />

complex–antigenic peptide complex, which leads to actin-dependent<br />

microcluster formation and recruitment of signaling components<br />

to form a signalosome within seconds 8–10 . The TCR signalosome<br />

includes tyrosine-phosphorylated forms of the kinase Lck (A001394),<br />

the kinase Zap70 (A002396) and the membrane adaptor Lat<br />

(A001392) and excludes the transmembrane phosphatase CD45<br />

(refs. 8,9,11–13). The contact area expands by integrin-mediated<br />

spreading as TCR microclusters continue to form at the outer<br />

edge 11,13 . Over a period of minutes, the microclusters move to the<br />

center of the contact area, where they fuse into larger clusters and<br />

become part of the nonmotile central supramolecular activation<br />

cluster (cSMAC) 13 . As there is less tyrosine phosphorylation in the<br />

cSMAC, it has been suggested to be the site of inactivation of old<br />

clusters, whereas new microclusters form at the periphery 9,13,14 .The<br />

formation and movement of new TCR microcluster–based signalosomes<br />

toward the cSMAC sustains signaling 13 .<br />

The driving force for protein rearrangement in the immunological<br />

synapse is unknown, although actomyosin-driven contraction has<br />

been proposed to drive TCR movement 15 . An alternative has been<br />

Received 11 December 2008; accepted 5 March 2009; published online 6 April 2009; doi:10.1038/ni.1723<br />

ARTICLES<br />

proposed on the basis of size-dependent segregation of proteins<br />

coupled with receptor-ligand interaction kinetics and membrane<br />

dynamics 16 . T cell synapses have been shown to have a centripetal<br />

version of retrograde actin flow 2,17 , a process that propels growth<br />

cones of neurons and other motile cells 18 . A close examination of the<br />

centripetal movement of TCR microclusters shows that it is F-actin<br />

dependent and that they move at about half of the speed of the<br />

underlying actin cytoskeleton (140 nm/s versus 320 nm/s, respectively)<br />

and can change course to move around barriers 2,17 . It has been<br />

proposed that intermittent coupling between the retrograde actin<br />

flow and the microclusters may drive centripetal movement, but the<br />

function of ‘motors’ in this process is not known.<br />

Members of the nonmuscle myosin II subfamily are critical to many<br />

cellular functions, including cell polarization, migration, adhesion and<br />

cytokinesis 19 . Members of the myosin II family are composed of a<br />

heavy-chain dimer; each heavy chain is associated with two myosin<br />

light chains (MLCs). Nonmuscle myosin II is activated by phosphorylation<br />

of the MLCs to induce assembly into bipolar filaments and<br />

contraction after interaction with actin filaments 19,20 . Three genes<br />

encode mammalian nonmuscle myosin II heavy chains, producing the<br />

following three isoforms: MyH9 (A004003), MyH10 and MyH14<br />

(refs. 21,22). Of those three isoforms, only MyH9 is dominant in<br />

T cells 6,23 . MyH9 pairs with regulatory MLCs to form a complex we<br />

refer to here by its common name, myosin IIA. T cell crawling and the<br />

movement of beads attached to the surface of T cells have been<br />

shown to require myosin IIA–mediated contractility 6,24 .Inbothof<br />

those studies, the immunological synapse seemed to form in<br />

the absence of myosin IIA activity or in cells depleted of myosin<br />

IIA by small interfering RNA (siRNA). Myosin IIA was recruited<br />

1 Department of Molecular Biology and Genetics, Weill Institute for Cell and Molecular Biology, Cornell University, Ithaca, New York, USA. 2 Molecular Pathogenesis<br />

Program, Helen L. and Martin S. Kimmel Center for Biology and Medicine of the Skirball Institute of Biomolecular Medicine, New York University School of Medicine,<br />

New York, New York, USA. 3 These authors contributed equally to this work. Correspondence should be addressed to A.B. (apb5@cornell.edu) or M.L.D.<br />

(michael.dustin@med.nyu.edu).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 531


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

Control<br />

Blebb<br />

0 s 6 s 12 s 18 s 24 s 72 s<br />

to the synapse 6 , but its activation and function in signaling and<br />

synapse formation were not firmly established.<br />

Here we show that the actin-based ‘molecular motor’ myosin IIA is<br />

an essential participant in the formation and persistence of immunological<br />

synapses and TCR signaling. Myosin IIA was rapidly activated<br />

after TCR engagement, and its activity was essential for the centripetal<br />

movement of TCR microclusters. Additionally, both immunological<br />

synapse stability and signaling ‘downstream’ of the TCR required<br />

intact myosin IIA.<br />

RESULTS<br />

TCR microcluster movement requires myosin IIA<br />

As translocation of TCR microclusters is an essential part of the<br />

formation of the immunological synapse, we first assessed whether<br />

myosin IIA was required for this motion. TCR microclusters can be<br />

tracked with a supported planar bilayer system and total internal<br />

reflection fluorescence (TIRF) microscopy 11,13 .WeusedTIRFmicroscopy<br />

to image the motion of TCR microclusters in Jurkat T cells<br />

on supported planar bilayers containing laterally mobile Alexa Fluor<br />

568–labeled antibody to TCR (anti-TCR; OKT3) and intercellular<br />

adhesion molecule 1 (ICAM-1) 17 . In agreement with published<br />

studies 17 , TCR microclusters in Jurkat T cells moved centripetally<br />

with an average velocity (± s.e.m.) of 0.15 ± 0.05 mm/s (P o 0.0001;<br />

Fig. 1a and Supplementary Movie 1 online) to generate the cSMAC.<br />

The average displacement of a microcluster from its point of formation<br />

to the cSMAC was 2.6 ± 0.8 mm (P o 0.0001), and the<br />

meandering index, calculated as displacement divided by track length,<br />

was 0.83 ± 0.09 (P o 0.0001); both are consistent with published<br />

values 17 . To assess the involvement of myosin IIA activity in microcluster<br />

translocation, we first treated the Jurkat T cells with blebbistatin,<br />

a well established specific inhibitor of myosin II ATPase<br />

activity 25 . Jurkat T cells pretreated for 10 min with blebbistatin<br />

(50 mM) formed microclusters but showed less directed microcluster<br />

movement, with an average speed of 0.06 ± 0.02 mm/s, a displacement<br />

of 0.25 ± 0.13 mm and a meandering index of 0.17 ± 0.09 (P o 0.0001<br />

for all measurements; Fig. 1a and Supplementary Fig. 1 and Supplementary<br />

Movie 2 online). We detected equivalent blockade of<br />

microcluster centripetal motion with ML-7 an inhibitor of myosin<br />

light-chain kinase (MLCK; Supplementary Movie 3 online). We<br />

noted similar effects on the inhibition of microcluster movement<br />

when we inhibited myosin IIA activity in primary human CD4 + T cells<br />

by treatment with blebbistatin (Supplementary Movies 4–6 online).<br />

b<br />

TCR Myosin<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Thus, the microclusters continued to move at 40% the speed of<br />

control Jurkat T cells but with over fourfold more meandering and<br />

only 10% of the displacement of control Jurkat T cells (treated with<br />

DMSO). In mature synapses with formed cSMACs, blebbistatin<br />

treatment did not disrupt the cSMAC, but the peripheral TCR<br />

microclusters ceased directed movement shortly after addition of the<br />

drug (Supplementary Movie 7 online). These results suggest that<br />

myosin IIA activity is required for the centripetal movement of TCR<br />

microclusters but not for microcluster formation.<br />

To further assess the involvement of myosin IIA in the translocation<br />

of TCR microclusters, we targeted MyH9 by siRNA. Jurkat<br />

T cells did not recover sufficiently from nucleofection of control<br />

siRNA to form mature synapses (data not shown). As myosin II is<br />

required for cytokinesis 19 , siRNA vectors that require growth and<br />

selection would also not be usable. Therefore, we ‘knocked down’<br />

MyH9 in primary activated human CD4 + T cells, which recover well<br />

from nucleofection. The best knockdown efficiency achieved in the<br />

primary T cells was 35%, as assessed by immunoblot analysis (data<br />

not shown). However, immunofluorescence analysis showed that<br />

this was due to nearly complete knockdown of MyH9 in one third of<br />

cells (data not shown). We analyzed microcluster tracking on planar<br />

bilayers of all cells in several microscopic fields while indexing the<br />

x-y coordinates of the fields, then fixed the cells and stained for<br />

intracellular MyH9, which allowed us to identify the cells in which<br />

MyH9 was ‘knocked down’ in the previously tracked and indexed<br />

cells. Primary T cells depleted of MyH9 failed to form the typical<br />

condensed cSMAC and instead had small, scattered TCR microclusters<br />

(Fig. 1b). TCR microclusters in cells treated with control<br />

siRNA had an average centripetal velocity of 0.12 ± 0.034 mm/s, with<br />

an average displacement of 2.2 ± 0.53 mm and a meandering index of<br />

0.85 ± 0.07 (P o 0.0001 for all measurements). TCR microclusters<br />

in MyH9-deficient cells had a speed of 0.062 ± 22 mm/s, a displacement<br />

of 0.26 ± 0.11 mm and a meandering index of 0.25 ± 0.11<br />

(P o 0.0001 for all measurements; Supplementary Fig. 1). Notably,<br />

there was significantly less TCR accumulation at the cSMAC<br />

(P o 0.0001) but only slightly, nonsignificantly less total TCR<br />

in the entire contact area in cells depleted of MyH9 (Fig. 1c).<br />

These results obtained by siRNA knockdown of MyH9 expression<br />

reproduced the results obtained by inhibition of myosin II activity<br />

with blebbistatin and ML7. Thus, myosin IIA activity is required for<br />

the translocation of TCR microclusters to form a cSMAC but not for<br />

the formation of TCR microclusters.<br />

c<br />

TCR (% of control)<br />

Control<br />

siRNA<br />

Total cSMAC<br />

Figure 1 Effect of the inhibition or depletion of myosin IIA on the centripetal motion of TCR microclusters. (a) TIRF microscopy of microclusters in control<br />

Jurkat T cells (top) or cells pretreated with blebbistatin (Blebb; bottom) added to a planar lipid bilayer containing Alexa Fluor 568–labeled anti-TCR and<br />

ICAM-1 and imaged during the initial minutes of synapse formation. Yellow plus symbols indicate initial microcluster localization; red circles indicate<br />

microcluster location; red lines indicate tracks of individual clusters. (b) Microscopy of primary human CD4 + cells unaffected by siRNA treatment (left;<br />

n ¼ 1 cell per image) or treated with MYH9-specific siRNA (arrows; n ¼ 1 cell per image), then added for 20 min to a planar lipid bilayer containing Alexa<br />

Fluor 568–labeled anti-TCR (red) and ICAM-1, and then fixed and stained for myosin IIA (green). Scale bars (a,b), 5 mm. (c) Total TCRs and TCRs at the<br />

center of the contact area (cSMAC) in control and myosin IIA–depleted (siRNA) cells (n ¼ 30 cells per condition). Data are representative of seven (a) or<br />

three (b) experiments with 26 samples per condition or three experiments (c error bars, s.e.m.).<br />

532 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

b<br />

Stim (min):<br />

p-MLC<br />

Coomassie<br />

MLC<br />

0<br />

0.5 1 3 5 10 30<br />

F-actin Myosin HC Merge<br />

Myosin IIA is activated during T cell stimulation<br />

Our initial results indicated that myosin IIA participates in the<br />

formation of the immunological synapse. Activation of myosin IIA<br />

through phosphorylation of MLCs during the formation of the<br />

immunological synapse has not been evaluated so far. We therefore<br />

examined the phosphorylation status of MLCs in Jurkat T cells<br />

stimulated either with soluble OKT3, which activates the TCR only,<br />

or with ‘superantigen’ presented by Raji B cells as APCs, which engages<br />

the TCR and multiple adhesion and costimulatory molecules. MLCs<br />

were not detectably phosphorylated in resting Jurkat T cells, but within<br />

30 s of stimulation by soluble OKT3, they became phosphorylated and<br />

this phosphorylation was sustained for at least 30 min (Fig. 2a). In<br />

resting Jurkat T cells, myosin IIA was uniformly distributed in the<br />

cytoplasm, whereas after stimulation with soluble OKT3, myosin IIA<br />

and its phosphorylated MLCs rapidly became enriched at the area of<br />

TCR clusters at the plasma membrane (Fig. 2b,c).<br />

In synapses formed between Jurkat T cells and ‘superantigen’loaded<br />

Raji B cells, we detected the typical accumulation of TCR,<br />

F-actin and ezrin at the contact site as described before 26,27 .Inatwocell<br />

system, either the T cell or B cell could contribute to this protein<br />

redistribution, yet the results obtained with immune synapses between<br />

Jurkat and Raji B cells were identical to results obtained with Jurkat<br />

T cells stimulated with soluble OKT3 and were indicative of a<br />

seemingly normal immune synapse (Figs. 2 and 3). The synapse<br />

was also highly enriched for myosin IIA, with a distribution very<br />

similar to that of the TCR (Fig. 3a). We obtained similar results with<br />

primary human CD4 + T cells (Supplementary Fig. 2 online). The<br />

recruitment of activated myosin IIA to the immunological synapse is<br />

consistent with the observed function of myosin IIA in the movement<br />

of TCR microclusters and cSMAC formation.<br />

Figure 3 Effect of the inhibition of myosin IIA<br />

activity on the formation of the immunological<br />

synapse. (a,b) Microscopy of Jurkat T cells<br />

pretreated for 10 min with DMSO (a) or50mM<br />

blebbistatin (b), incubated for 5 min with SEE<br />

superantigen–loaded B cells (prestained with<br />

CMTPX; red), then fixed and stained for TCR,<br />

ezrin, F-actin or myosin II heavy chain (green).<br />

Numbers above images indicate percent cells<br />

with similar protein distribution (n ¼ 30 cells per<br />

condition). Scale bars, 5 mm. (c) Conjugate<br />

formation by cells treated as described in a,b<br />

(n ¼ 50 cells per condition). Data are representative<br />

of three experiments (error bars (c), s.d.).<br />

c<br />

TCR Myosin HC Merge<br />

TCR p-MLC Merge<br />

Myosin HC p-MLC Merge<br />

DMSO<br />

Blebb<br />

Figure 2 Phosphorylation and redistribution<br />

of myosin IIA during activation of T cells.<br />

(a) Abundance of phosphorylated MLC (p-MLC)<br />

and total MLC (MLC) in total T cell lysates at<br />

various times during stimulation (Stim) by soluble<br />

anti-TCR (OKT3). (b) Microscopy of resting Jurkat<br />

T cells fixed and stained for F-actin (green) and<br />

myosin IIA heavy chain (HC; red). (c) Microscopy<br />

of Jurkat T cells stimulated for 1 min with OKT3,<br />

then fixed and stained for TCR, myosin heavy<br />

chain and phosphorylated MLC. Cells showing<br />

colocalization: 83%, TCR and myosin IIA heavy<br />

chain; 92%, TCR and phosphorylated MLC; 90%,<br />

myosin IIA heavy chain and phosphorylated MLC.<br />

Scale bars (b,c), 5 mm. Data are representative of<br />

three experiments (n ¼ 30 cells per condition).<br />

Immunological synapse stability requires myosin IIA<br />

To understand the consequences of myosin IIA activity for the<br />

immunological synapse, we determined the effect of pretreating<br />

Jurkat T cell with 50 mM blebbistatin on synapse formation with<br />

‘superantigen’-loaded Raji B cells. Unexpectedly, we found that<br />

inhibition of myosin II activity did not inhibit the concentration of<br />

TCR, ezrin, F-actin or myosin IIA itself at the contact site between the<br />

two cells (Fig. 3b). As we could not use siRNA in the Jurkat model, we<br />

used both ML7 and an additional inhibitor, Y27632, which inhibits<br />

Rho-associated kinase (ROCK). Both ROCK and MLCK phosphorylated<br />

and activated MLCs, and both ML7 and Y27632 inhibited<br />

phosphorylation of MLCs during T cell stimulation with soluble<br />

anti-TCR (Supplementary Fig. 3 online). Conjugates between ‘superantigen’-loaded<br />

Raji B cells and Jurkat T cells pretreated with either of<br />

those drugs had apparently normal accumulation of myosin IIA<br />

(Supplementary Fig. 4 online). We obtained similar results with<br />

primary human CD4 + cells pretreated with blebbistatin and incubated<br />

for 5 min with Raji B cells (Supplementary Fig. 2). These data<br />

confirm and extend earlier indications that the first attachment step of<br />

immunological synapse formation does not require myosin IIA<br />

activity 6 . These results were also consistent with the ability of<br />

blebbistatin-treated or myosin IIA–depleted cells to form immature<br />

immunological synapses on planar bilayers containing TCR microclusters<br />

and ICAM-1 (Fig. 1).<br />

Although the conjugates formed after inhibition of myosin IIA<br />

seemed normal, we found that fewer total conjugates formed with<br />

T cells pretreated with 50 mM blebbistatin than with control cells<br />

treated with DMSO (Fig. 3c). Conjugate formation was not further<br />

decreased by pretreatment with 100 mM blebbistatin (data not shown),<br />

which suggested that the residual conjugate formation was not simply<br />

a TCR (89%) Ezrin (86%) F-actin (93%) Myosin HC (83%) c<br />

b<br />

TCR (92%) Ezrin (79%) F-actin (90%) Myosin HC (81%)<br />

T cells in<br />

conjugates (%)<br />

ARTICLES<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 533<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

DMSO<br />

Blebb


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

DMSO<br />

(pretreat)<br />

Blebb<br />

(pretreat)<br />

Blebb<br />

(1 min)<br />

Blebb<br />

(12 min)<br />

Before blebb After blebb<br />

T<br />

T<br />

T<br />

T<br />

T<br />

T<br />

B<br />

B<br />

T<br />

B<br />

B B<br />

T<br />

T<br />

T<br />

B<br />

B<br />

an effect of partial inhibition of myosin IIA activity. To explore the<br />

basis for the lower conjugate number, we examined the effect of<br />

blebbistatin addition before and after conjugate formation. We first<br />

immobilized ‘superantigen’-loaded Raji B cells in dishes with coverslip<br />

inserts and then added Jurkat T cells. We monitored conjugate<br />

formation and stability by differential interference contrast (DIC)<br />

microscopy, adding blebbistatin at various times relative to conjugate<br />

formation (Fig. 4a). The addition of blebbistatin resulted in lower<br />

stability of formed conjugates so that only about 20% remained 2 min<br />

after drug addition (Fig. 4b). Jurkat T cells treated with blebbistatin<br />

formed unstable synapses that only lasted for 102 ± 14 s (Fig. 4c),<br />

whereas control T cells formed stable synapses that persisted for over<br />

20 min (data not shown). The addition of blebbistatin at various times<br />

after conjugate formation resulted in instability and detachment within<br />

1–2 min of drug addition, with an average time of 109 s (Fig. 4c).<br />

As blebbistatin resulted in the same instability regardless of the time of<br />

addition after synapse formation, we concluded that myosin IIA activity<br />

is needed to maintain the stability of both early and mature<br />

synapses. We obtained similar results by inhibiting myosin IIA activation<br />

with 10 mM ML7(Supplementary Movie 8 online) and with<br />

primary human CD4 + cells (Supplementary Movies 9,10 online).<br />

B<br />

b<br />

Synapses (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

170<br />

150<br />

130<br />

110<br />

90<br />

70<br />

50<br />

30<br />

Synapse duration (s) c<br />

DMSO<br />

Blebb<br />

0 1 2 3 4 5 10 12<br />

Blebb addition (min)<br />

Figure 5 Effect of the inhibition of myosin IIA activity on intracellular Ca 2+<br />

concentrations. (a) Microscopy of Jurkat T cells incubated with Fluo-LoJo,<br />

then mixed with SEE superantigen–loaded B cells and allowed to form<br />

immunological synapses, followed by the addition of DMSO or blebbistatin.<br />

Scale bars, 5 mm. (b) Change in intensity over time of Fluo-LoJo in the cells<br />

in a treated with DMSO (control) or blebbistatin at time 0 (n ¼ 3cells<br />

per condition), presented relative to the average sustained signal in<br />

‘superantigen’-activated cells, set as 100%. Numbers in parentheses (key)<br />

identify different samples. (c) Emission ratios of Jurkat T cells incubated<br />

with Fura-2AM, then added to a planar lipid bilayer containing anti-TCR and<br />

ICAM-1 for 15 min before imaging, assessed every 15 s by fluorescence<br />

microscopy and presented as the ratio of absorbance at 340 nm to<br />

absorbance at 340 nm (340/380). Gray shaded area indicates the addition<br />

of DMSO or blebbistatin 15 min after the addition of cells to bilayer (except<br />

for pretreated cells, which did not receive additional blebbistatin). The low<br />

and high calcium ratios corresponding to cells in EGTA with Mg 2+ and<br />

without Ca 2+ or ionomycin were also determined (data not shown). Data are<br />

representative of three experiments (a,b) or two independent experiments (c;<br />

n ¼ 17 cells per condition).<br />

Figure 4 Effect of the inhibition of myosin IIA activity on the stability of the<br />

immunological synapse. (a) DIC imaging of conjugate formation and stability<br />

of Jurkat T cells pretreated (pretreat) with DMSO or blebbistatin and added<br />

to SEE superantigen–loaded B cells immobilized in dishes with coverslip<br />

inserts, followed by the formation of immunological synapses (top two rows),<br />

or of Jurkat T cells without pretreatment added to the B cells described<br />

above, followed by the addition of blebbistatin 1 min or 12 min after synapse<br />

formation (bottom two rows), assessed before (left) and 1–2 min after (right)<br />

treatment with blebbistatin. T, T cell; B, B cell; white arrows, immunological<br />

synapses; black arrows, loss of synapse. (b) Synapses in the cells in a<br />

present 2 min after the addition of blebbistatin. (c) Duration of synapses in<br />

the cells in a after the addition of blebbistatin at various times after synapse<br />

formation (horizontal axis). Data are representative of five experiments (a),<br />

or five experiments with 35 cells per condition (b,c; error bars, s.d.).<br />

We next assessed whether synapse breakdown resulted from<br />

perturbation of the typical accumulation of the adhesion proteins<br />

LFA-1 and ICAM-1 at the peripheral SMAC 5 after inhibition of<br />

myosin IIA activity with blebbistatin. Pretreatment with blebbistatin<br />

led to a more peripheral distribution of these interactions, consistent<br />

with impaired transport toward the center. However, we did not detect<br />

a difference in the intensity of these interactions relative to that of<br />

control cells treated with DMSO (Supplementary Fig. 5 online).<br />

Thus, instability of the immunological synapse after the inhibition of<br />

myosin IIA activity was not due to initial failure of LFA-1 activation.<br />

Ca 2+ signaling requires myosin IIA activity<br />

One of the earliest and most easily monitored signaling events after<br />

T cell activation is a rapid increase in cytoplasmic Ca 2+ (ref. 28).<br />

A published study has shown that treatment with butanedione<br />

monoxide, a less specific inhibitor of myosin II than blebbistatin, in<br />

activated primary CD4 + T cells leads to a less sustained increase in<br />

Ca 2+ after stimulation and a partial blockade of the movement of<br />

membrane proteins to the synapse 24 . To explore if synapse instability<br />

correlated with loss of Ca 2+ signaling, we preloaded Jurkat T cells with<br />

the fluorescent, cytoplasmic, Ca 2+ -sensitive indicator dye Fluo-LoJo<br />

and assessed the effect of blebbistatin on cytoplasmic Ca 2+ in response<br />

to ‘superantigen’-loaded Raji B cells. Although control Jurkat T cells<br />

maintained higher cytoplasmic Ca 2+ concentrations (Fig. 5a,b), the<br />

addition of blebbistatin (50 mM) to an existing immunological synapse<br />

led to a rapid decrease in Ca 2+ concentrations within 1 min (Fig. 5a,b<br />

a b<br />

Time (s) DMSO Blebb<br />

0<br />

9<br />

18<br />

27<br />

33<br />

B<br />

B<br />

B<br />

B<br />

B<br />

B<br />

B<br />

B<br />

B<br />

B<br />

Intensity (%)<br />

c<br />

Fura-2 (340/380)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

0 15 30 45<br />

Time (s)<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1 2 3 4 5 6<br />

Time (min)<br />

DMSO (1)<br />

DMSO (2)<br />

DMSO (3)<br />

Blebb (1)<br />

Blebb (2)<br />

Blebb (3)<br />

DMSO<br />

Blebb<br />

Blebb<br />

(pretreat)<br />

534 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

0 min<br />

1 min<br />

5 min<br />

b 700<br />

1 min c<br />

600<br />

5 min<br />

TCR cluster (nm)<br />

500<br />

400<br />

Control<br />

300<br />

200<br />

100<br />

0<br />

Blebb: None Before<br />

stim<br />

Blebb pretreat Blebb after TCR<br />

After<br />

stim<br />

α-TCR:<br />

Blebb:<br />

p-Src<br />

p-Lat<br />

p-Zap70<br />

Actin<br />

and Supplementary Fig. 6b online). We obtained similar results with<br />

ML-7 (Supplementary Movie 11 and Supplementary Fig. 7 online).<br />

We detected a similar decrease in Ca 2+ concentrations in primary<br />

human CD4 + cells after inhibition of myosin IIA (Supplementary<br />

Movies 12–14 online). For a more quantitative measurement of<br />

cytoplasmic Ca 2+ changes, we loaded Jurkat T cells with the ratiometric<br />

Ca 2+ -indicator dye Fura-2AM and calculated emission ratios.<br />

The addition of 50 mM blebbistatin to cells with preformed synapses<br />

diminished cytoplasmic Ca 2+ concentrations to baseline within less<br />

than 2 min, whereas control cells maintained high Ca 2+ concentrations<br />

(Fig. 5c). Pretreatment with 50 mM blebbistatin blocked the<br />

TCR-inducedincreaseinCa 2+ altogether (Fig. 5c). To rule out the<br />

possibility that emission-intensity changes resulted from autofluorescence<br />

of blebbistatin, we preloaded T cells with Fluo-LoJo and added<br />

blebbistatin without any TCR stimulation. Blebbistatin fluorescence<br />

was negligible in our assays (Supplementary Fig. 6). Moreover, we<br />

found that the addition of 50 mM blebbistatin to the cells, followed by<br />

illumination, had no toxic effect (data not shown). Notably, in all<br />

these experiments, the decrease in cytoplasmic Ca 2+ concentration<br />

preceded the detachment of the immunological synapse, which<br />

indicated that myosin IIA activity is necessary for sustained Ca 2+<br />

signaling ‘downstream’ of the TCR in the immunological synapse<br />

independently of any effects on adhesion.<br />

The serial-triggering model holds that one major histocompatibility<br />

complex–bound antigenic peptide engages a large number of TCRs in<br />

successive rounds, contacting about 50–200 receptors per antigenic<br />

peptide 29 . This model is compatible with the demonstration that ten<br />

complexes of peptide and major histocompatibility complex in the<br />

T cell–APC interface can sustain signaling long enough to generate<br />

interleukin 2 (ref. 30). If myosin IIA is needed only to promote an<br />

active process of serial triggering, then increasing the number of TCRs<br />

triggered in parallel might overcome the requirement for myosin IIA<br />

activity. To test that possibility, we explored whether more activating<br />

anti-TCR could overcome the effect of blebbistatin on Ca 2+ signaling.<br />

We preloaded Jurkat T cells with Fluo-LoJo and then stimulated the<br />

cells with increasing concentrations of anti-TCR. Once the cytoplasmic<br />

Ca 2+ concentrations had risen, we added 50 mM blebbistatinand<br />

monitored Ca 2+ concentrations for an additional 1 min. The decrease<br />

in cytoplasmic Ca 2+ concentration was independent of the concentration<br />

of activating antibody, with a similar decrease in cells stimulated<br />

–<br />

–<br />

+<br />

–<br />

+<br />

+<br />

ARTICLES<br />

Figure 6 Effect of the inhibition of myosin IIA activity on TCR microclusters.<br />

(a) Microscopy of Jurkat T cells stimulated with Alexa Fluor 488–conjugated<br />

anti-CD3 without additional treatment (Control), after pretreatment for<br />

10 min with 50 mM blebbistatin (Blebb pretreat) or treatment with 50 mM<br />

blebbistatin after TCR stimulation (Blebb after TCR), then imaged<br />

immediately (0 min) or at 1 and 5 min after stimulation. Scale bars, 5 mm.<br />

Data are representative of five experiments with two cells per condition per<br />

time point. (b) Quantitative analysis of the results in a. Dataare<br />

representative of five experiments (error bars, s.e.m.; n ¼ 35 clusters per<br />

bar). (c) Immunoblot analysis of Src phosphorylated at tyrosine 416 (p-Src),<br />

Zap70 phosphorylated at tyrosine 319 (p-Zap70) and Lat phosphorylated at<br />

tyrosine 191 (p-Lat), in Jurkat T cells left untreated (a-TCR ) or treated for<br />

2 min with anti-CD3 (a-TCR +; OKT3) with (+) or without ( ) blebbistatin<br />

pretreatment. Bottom, actin protein abundance (loading control). Results<br />

are representative of three independent experiments.<br />

with between 10 mg/ml and 500 mg/ml of antibody (Supplementary<br />

Fig. 8 online). This result challenges the possibility of insufficient<br />

TCR engagement as a mechanism to account for the decrease in<br />

Ca 2+ signaling.<br />

TCR signaling requires myosin IIA activity<br />

Our results suggested that myosin IIA might be important for the<br />

function of the TCR signalosome. The simplest way to activate the<br />

formation of TCR signalosomes is based on the addition of soluble<br />

anti-CD3e to Jurkat T cells 31 , shown above to activate MLC phosphorylation.<br />

We incubated Jurkat T cells with fluorescence-tagged<br />

anti-CD3e and monitored TCR distribution and biochemical indicators<br />

of TCR signalosome assembly (phosphorylation of Lck, Zap70<br />

and Lat). Control Jurkat T cells initially showed a uniform surface<br />

fluorescence that aggregated into microclusters with a diameter of<br />

280 ± 70 nm by 1 min, followed by coalescence into larger clusters of<br />

456 ± 88 nm after 5 min of stimulation (Fig. 6a,b). When we<br />

pretreated Jurkat T cells for 5 min with 50 mM blebbistatin and<br />

then stimulated the cells for 1 min with labeled anti-TCR, the TCR<br />

clusters were slightly smaller, with a diameter of 217 ± 63 nm.<br />

However, progression in cluster size in the blebbistatin-treated cells<br />

was minimal, reaching a diameter of 247 ± 66 nm after 5 min of<br />

stimulation (Fig. 6a,b). We next explored the effect on microclusters<br />

when we added blebbistatin at 1 or 5 min after stimulation. In both<br />

cases, 5 min after the addition of blebbistatin, the cluster was smaller,<br />

with a diameter of 217 ± 64 nm or 258 ± 59 nm for 1 min or 5 min,<br />

respectively (Fig. 6a,b). These results collectively show that TCR<br />

microclusters about 217 nm in diameter can form in the absence of<br />

myosin IIA activity, yet their coalescence into larger clusters and their<br />

maintenance in larger clusters require myosin IIA activity. We<br />

obtained similar results with primary human CD4 + cells (Supplementary<br />

Fig. 2). When we treated Jurkat T cells for 2 min with anti-<br />

CD3e and then analyzed phosphorylated signalosome components by<br />

direct immunoblot of lysates, we found that phosphorylation of Src<br />

kinases, which probably included phosphorylated Lck, was similar<br />

with or without blebbistatin pretreatment (Fig. 6c). In contrast,<br />

phosphorylation of Zap70 and Lat was substantially lower after<br />

blebbistatin pretreatment (Fig. 6c). We obtained similar results with<br />

primary CD4 + T cells (Supplementary Fig. 2). We also assessed<br />

whether Jurkat T cells pretreated with blebbistatin increased their<br />

Ca 2+ in response to stimulation with soluble anti-CD3e. T cells<br />

preloaded with Fluo-LoJo and stimulated with soluble anti-TCR<br />

underwent a robust Ca 2+ response, whereas cells pretreated with<br />

blebbistatin failed to increase Ca 2+ concentrations in response<br />

to stimulation (Fig. 5c and Supplementary Fig. 8). These results<br />

indicate quantitative defects in TCR microcluster size and defective<br />

signalosome function in a synapse-free assay.<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 535


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a CD3 Src (pY416) Merge<br />

b<br />

DMSO<br />

Blebb<br />

DMSO<br />

Blebb<br />

DMSO<br />

Blebb<br />

CD3<br />

CD3<br />

Zap70 (pY319) Merge<br />

Lat (pY191) Merge<br />

Src (pY416)<br />

intensity (%)<br />

Zap70 (pY319)<br />

intensity (%)<br />

Lat (pY191)<br />

intensity (%)<br />

120<br />

100<br />

80<br />

TCR signalosome function can also be evaluated in a synapse-based<br />

system with supported planar bilayers presenting OKT3 (ref. 17).<br />

T cells interacting with a planar bilayer containing OKT3 and ICAM-1<br />

for 5 min had a central condensed TCR cluster surrounded by<br />

peripheral microclusters containing TCRs, as well as phosphorylated<br />

Zap70 and Lat (Fig. 7), similar to published studies 9 .Whenweadded<br />

Jurkat T cells pretreated with 50 mM blebbistatin to the bilayers,<br />

followed by staining with a specific antibody to each phosphorylated<br />

protein, we found that phosphorylated Src was localized together with<br />

TCR microclusters, but the abundance of phosphorylated Zap70 and<br />

Lat, as measured by fluorescence intensity, was significantly lower, by<br />

80% each (P o 0.0001; Fig. 7). We also extended this analysis to<br />

primary CD4 + T cells treated with control and MyH9-specific siRNA<br />

during activation; this resulted in nearly complete knockdown of<br />

myosin IIA in one third of the cells. We found that knockdown<br />

of myosin IIA diminished phosphorylation of Src by only 25%<br />

(P o 0.0001) but lowered phosphorylation of Zap70 at tyrosine 319<br />

by 80% (P o 0.0001) and phosphorylation of Lat by 70%<br />

(P o 0.0001). These data demonstrate, by both pharmacological and<br />

reverse-genetic approaches, that myosin IIA is required for amplification<br />

of TCR signaling between the Lck and Zap70 activation steps.<br />

DISCUSSION<br />

Here we have provided evidence that myosin IIA is central to synapse<br />

assembly and signaling, being necessary for TCR signaling, the<br />

centripetal motion and fusion of microclusters during the formation<br />

60<br />

40<br />

20 0<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20 0<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20 0<br />

DMSO<br />

Blebb<br />

DMSO<br />

Blebb<br />

DMSO<br />

Blebb<br />

Control<br />

siRNA<br />

Myosin IIA<br />

siRNA<br />

Control<br />

siRNA<br />

Myosin IIA<br />

siRNA<br />

Control<br />

siRNA<br />

Myosin IIA<br />

siRNA<br />

CD3<br />

Src (pY416) Merge<br />

CD3 Zap70 (pY319) Merge<br />

CD3 LatT (pY191) Merge<br />

Src (pY416)<br />

intensity (%)<br />

Zap70 (pY319)<br />

intensity (%)<br />

Lat (pY191)<br />

intensity (%)<br />

120<br />

100<br />

80<br />

of the immunological synapse, and synapse persistence. Published<br />

work has shown that the F-actin cytoskeleton is required for all of<br />

these processes 17,32 and that TCR engagement induces actin polymerization<br />

through recruitment of the adaptor protein Nck and Wiskott-<br />

Aldrich syndrome protein to the TCR microclusters 33 . Our study has<br />

shown that after T cell engagement, myosin IIA was activated by<br />

phosphorylation of MLCs and its activity was necessary for proper<br />

assembly of the signalosome. Inhibition of myosin IIA activity with<br />

the highly specific myosin II inhibitor blebbistatin or depletion of<br />

myosin IIA expression with specific siRNA resulted in a complete<br />

halting of directed motion of microclusters, prevented the formation<br />

of the cSMAC and prevented amplification of TCR signals after Lck<br />

activation. Whether myosin IIA activity was inhibited pharmacologically,<br />

in which case myosin IIA was still recruited to the synapse, or if<br />

its expression was diminished by siRNA, in which case it was<br />

profoundly depleted from the synapse, the formation of initial small<br />

TCR microclusters remained intact. However, these clusters did not<br />

increase in size, did not fully signal and did not undergo directed<br />

translocation. Thus, we have defined distinct F-actin-dependent and<br />

actomyosin-dependent phases of T cell activation and immunological<br />

synapse formation.<br />

The potential involvement of myosin II in the formation of<br />

immunological synapses has been reported before. One study showed<br />

that movement of ICAM-1-coated beads on T cells after activation by<br />

a B cell is inhibited by butanedione monoxime with concurrent<br />

decrease in Ca 2+ signaling, although the B cell–T cell conjugates<br />

60<br />

40<br />

20<br />

0<br />

Control Myosin IIA<br />

siRNA siRNA<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Control Myosin IIA<br />

siRNA siRNA<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Control Myosin IIA<br />

siRNA siRNA<br />

Figure 7 Effect of the inhibition or depletion of myosin IIA on signaling in T cells. (a) Localization of TCR microclusters and phosphorylated Src, Zap70<br />

and Lat (left) in Jurkat T cells (with or without blebbistatin pretreatment; left margin) added for 25 min to a planar lipid bilayer containing Alexa Fluor<br />

568–labeled anti-TCR and ICAM-1, then fixed and stained with antibody to Src phosphorylated at tyrosine 416 (Src(pY416)), Zap70 phosphorylated at<br />

tyrosine 319 (Zap70(pY319)) or Lat phosphorylated at tyrosine 191 (Lat(pY191)). Right, quantification of protein phosphorylation. Data are representative of<br />

three experiments (n ¼ 15 cells per bar; error bars, s.d.). (b) Localization of TCR microclusters and phosphorylated Src, Zap70 and Lat (left) in primary<br />

human CD4 + cells treated with control or MYH9-specific siRNA and added for 25 min to a planar lipid bilayer containing Alexa Fluor 568–labeled anti-TCR<br />

and ICAM-1, then fixed and stained as described in a. Myosin IIA–depleted cells were identified by the lack of central TCR clustering (as in Fig. 6b). Right,<br />

quantification of protein phosphorylation. Data are representative of two experiments (n ¼ 15 cells per bar; error bars, s.d.).<br />

536 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

remain stable 24 . The authors hypothesized that myosin II–mediated<br />

transport is delivering components to the immunological synapse that<br />

are needed for sustained signaling. Another study showed that myosin<br />

IIA is necessary for T cell motility and uropod maintenance and<br />

postulated that inhibition of myosin IIA filament formation is<br />

required for the T cell ‘stop signal’ after antigen encounter 6 .These<br />

authors also reported that immunological synapse formation seemed<br />

unaffected by pretreatment with blebbistatin. That result is in agreement<br />

with our finding that immunological synapses formed with<br />

blebbistatin-treated T cells were initially similar to synapses with<br />

control cells. The T cell blasts used in the earlier study 6 have high<br />

constitutive LFA-1 activity, so myosin IIA–dependent signaling is not<br />

required for conjugate formation. We focused on two systems here,<br />

Jurkat T cells and primary human T cells, in which basal LFA-1<br />

activity is low and ‘inside-out’ signaling through the TCR is required<br />

for conjugate formation 34 . In retrospect, evidence of spreading and<br />

contraction in the immunological synapse formation process was<br />

visible in earlier studies 5,9 and has been explicitly described for<br />

B cell synapse formation without indicating involvement of myosin<br />

II (ref. 35). Contractile oscillations have been reported at the outer<br />

edge of the immunological synapse formed by T cells 32 .Contractile<br />

oscillations require myosin IIA in fibroblasts. Our results suggest that<br />

this is also probably true in lymphocytes 36 .<br />

Myosin II–based cortical movement has been documented in<br />

several other situations. Myosin II is necessary for cortical tension<br />

and functions in the contractile ring during cytokinesis 37,38 .Several<br />

studies have suggested that an imbalance in cortical tension contributes<br />

to cytokinesis, with cortical loosening at the cell poles and<br />

enhanced tension at the cell equator, leading to equatorial movement,<br />

assembly and contraction of the contractile ring 39 . In a related<br />

mechanism, anterior-posterior polarity in the one-cell nematode<br />

embryo is established by myosin II–mediated cortical contraction,<br />

which moves granules and fate determinants toward the future<br />

anterior pole 40 . It is possible that a related myosin II–dependent<br />

cortical tension may move TCR microclusters toward the center of the<br />

immunological synapse. Such cortical tension seemed to be required<br />

for TCR signalosome function even in the absence of a synapse, on the<br />

basis of results obtained with soluble OKT3. The reported particle-size<br />

requirements for T cell stimulation may arise from the need for<br />

myosin IIA–mediated tension across an interface or crosslinked<br />

protein network 41,42 . Myosin IIA–mediated cortical tension may be<br />

required for the rearrangement of cytoskeletally associated ‘protein<br />

islands’ into functional signalosomes 43 .<br />

The activation of myosin II by phosphorylation of its MLCs can<br />

be mediated by several different kinases, including the calciumcalmodulin–dependent<br />

MLCK 44 , ROCK and protein kinase C 45 .<br />

Shortly after stimulation of T cells, Vav1, a Rho guanine-exchange<br />

factor, is recruited to TCR microclusters through interaction with the<br />

adaptor protein SLP-76, which is then followed by the recruitment of<br />

the GTP-binding protein Cdc42 and ROCK 46,47 . T cell stimulation also<br />

results in more cytoplasmic Ca 2+ , which is known to activate MLCK 44 .<br />

We have shown that treatment with either the ROCK inhibitor Y27632<br />

or the MLCK inhibitor ML-7 inhibited phosphorylation of MLCs after<br />

T cell stimulation. Thus, both kinases take part in activating myosin II<br />

even when the TCR is triggered by OKT3. As myosin IIA activity was<br />

necessary to maintain higher Ca 2+ concentrations, a plausible model is<br />

that Rho-GTP–activated ROCK initially phosphorylates MLCs, then<br />

Ca 2+ concentrations increase, which maintains light-chain phosphorylation<br />

through persistent activation of MLCK. Thus, one crucial function<br />

of myosin IIA activity is to maintain signaling that then feeds back<br />

to maintain higher Ca 2+ concentrations and active myosin IIA.<br />

ARTICLES<br />

This is the first report to our knowledge to link myosin II activity to<br />

signaling through an immunoreceptor. In examining the ‘downstream’<br />

signaling pathway, we found that phosphorylation of the Src family<br />

kinases was unimpaired by either inhibition or depletion of myosin<br />

IIA, whereas ‘downstream’ signaling, including phosphorylation of<br />

Zap70 and Lat and an increase in cytosolic Ca 2+ , were much more<br />

dependent on myosin IIA activity. The truncation in signaling ‘downstream’<br />

of Lck was not due to defects in adhesion, as inhibition of<br />

myosin IIA activity in Jurkat T cells stimulated with soluble OKT3 also<br />

resulted in less phosphorylation of Zap70 and Lat and a decrease in<br />

intracellular Ca 2+ concentrations to baseline. Our data support a twostep<br />

model in which initial conjugate formation involving the formation<br />

of TCR microclusters, recruitment of myosin IIA and activation<br />

of Lck are all independent of myosin IIA activity, whereas amplification<br />

of signaling and microcluster movement are dependent on<br />

myosin IIA activity. Our work here and published work indicates<br />

that there is careful ‘tuning’ of myosin IIA activity during T cell<br />

activation, with negative regulation through inhibition of the formation<br />

of thick filaments 6 and positive regulation through phosphorylation<br />

of MLCs, which leads to maintenance of the cortical tension<br />

needed for TCR signaling and synapse stabilization.<br />

METHODS<br />

Cells and antibodies. Jurkat T cells and Raji B cells were from American Type<br />

Culture Collection. Human peripheral blood lymphocytes were isolated from<br />

citrate-anticoagulated whole blood by dextran sedimentation (Blood Centers<br />

of America–hemerica), followed by density separation over Ficoll-Hypaque<br />

(Sigma). The resulting mononuclear cells were washed in PBS and were further<br />

purified by nylon wool and plastic adherence as described48 . Human peripheral<br />

CD4 + blasts were prepared as described49 . Anti-ezrin and anti–myosin II heavy<br />

chain have been described50 . Affinity-purified polyclonal antibodies to MLC<br />

phosphorylated at serine 19 (3671), Src phosphorylated at tyrosine 416 (used to<br />

measure phosphorylated Lck, the most abundant Src member in T cells; 2101),<br />

Zap 70 phosphorylated at tyrosine 319 (2701), and Lat phosphorylated at tyrosine<br />

191 (3584) were from Cell Signaling. OKT3 mouse anti–human CD3 was<br />

purified from an OKT3 hybridoma cell line (14-0037; eBioscience). Rhodaminephalloidin,<br />

Alexa Fluor 568–phalloidin, Alexa Fluor 488–conjugated donkey<br />

anti-mouse (21202) and Alexa Fluor 568–conjugated goat anti-rabbit (11011)<br />

and goat anti-mouse (11004) were from Invitrogen. Horseradish peroxidase–<br />

conjugated goat anti-rabbit (W4011) was from Promega. All procedures were<br />

approved by the Health and Safety Committee of Cornell University.<br />

Immunofluorescence. Cells were plated on poly-L-lysine-coated glass slides,<br />

were fixed for 30 min at 25 1C with 3.7% (wt/vol) formaldehyde, were made<br />

permeable by treatment for 2 min with 0.1% (vol/vol) Triton X-100 in PBS and<br />

then were rinsed three times in PBS. Cells were then incubated for 1 h with 5%<br />

(vol/vol) BSA in PBS, then were incubated for 1 h with primary antibody in 5%<br />

(vol/vol) BSA in PBS, washed in PBS and incubated for 1 h with the<br />

appropriate secondary antibody (or phalloidin) in 5% (vol/vol) BSA in PBS.<br />

After additional washes, 5 ml Vectashield (Vector Labs) was added to cells and<br />

slides were covered with coverslips. Cells were viewed with a Nikon Eclipse<br />

TE2000-U (100 objective; numerical aperture, 1.4) with the Perkin Elmer<br />

UltraVIEW LCI spinning-disk confocal imaging system and a Hamamatsu<br />

12-bit C4742-95 digital charge-coupled device camera.<br />

Immunoblot analysis. Jurkat and primary T cells were lysed and were resolved<br />

by SDS-PAGE, followed by transfer to polyvinylidene difluoride membranes<br />

(Immobilon-P; Millipore) with a semidry transfer system (Integrated Separation<br />

Systems). After 1 h of blocking in 5% (wt/vol) milk in Tris-buffered saline<br />

with Tween, membranes were incubated for 1 h with primary antibody, then<br />

were washed and were incubated for 1 h with the appropriate horseradish<br />

peroxidase–conjugated secondary antibody. Blots were developed by the<br />

enhanced chemiluminescence system (Amersham).<br />

Cell stimulation and conjugate formation. Jurkat and primary human T cells<br />

were activated for various times with the antibody OKT3 (10 mg/ml). For<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 537


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

stimulation with B cells, Raji B cells were labeled with the fluorescent dye<br />

CellTracker Red (CMTPX; Molecular Probes) and were loaded with the<br />

‘superantigen’ staphylococcus enterotoxin E (SEE; 2 mg/ml; Toxin Technology).<br />

Equal numbers of T cells and B cells were incubated together.<br />

Conjugate stability and DIC microscopy. Raji B cells were loaded with SEE<br />

superantigen and then were immobilized in dishes containing coverslip inserts<br />

(MatTek) and viewed on an Axiovert 100 TV microscope (Carl Zeiss) equipped<br />

with a charge-coupled device (C4742-95-12ERG; Hamamatsu) with a DIC<br />

prism and Openlab 4.0 software (Improvision). After initial B cell imaging,<br />

Jurkat T cells or primary human T cells were added to plates and cells were<br />

allowed to form conjugates. Blebbistatin (50 mM), ML7 (10 mM) or DMSO was<br />

added at various times and conjugates were continuously imaged. Movies were<br />

analyzed with ImageJ software.<br />

Ca 2+ assays. For nonratiometric assays, Jurkat T cells and primary human<br />

T cells were loaded with 1 mM Fluo-LoJo (TefLabs), then were added to<br />

SEE-loaded Raji B cells and allowed to form synapses or were stimulated with<br />

OKT3. Blebbistatin, ML7 or DMSO was added at various times and fluorescence<br />

intensity was measured with the spinning-disk confocal imaging system.<br />

For ratiometric analysis, Jurkat T cells were loaded with 2.5 mM Fura-2AM<br />

(acetoxymethyl ester; Molecular Probes) as described 13 .<br />

Bilayer assembly and TIRF microscopy. Glass-supported bilayers of dioleoyl<br />

phosphatidylcholine incorporating 0.01% (wt/vol) 1,2-dioleoyl-sn-glycero-3phosphatidylcholine,<br />

1,2-dioleoyl-sn-glycero-3-phosphoethanolamine-N-(cap<br />

biotinyl) were prepared in flow chambers (Bioptechs) as described 5 . Bilayers<br />

were loaded with monobiotinylated-564-OKT3. Cells were allowed to settle and<br />

form contacts with the bilayer before imaging. An Olympus inverted IX-70<br />

microscope equipped with Hamamatsu 12-bit C9100 1.1B charge-coupled<br />

device and a TIRF objective from Olympus were used for all bilayer imaging.<br />

Microclusters were analyzed with Volocity 4.2 (Improvision).<br />

Transfection with siRNA. CD4 + human T cell blasts (3 10 6 )atday4after<br />

isolation were transfected by electroporation with Amaxa nucleofector technology<br />

according to the manufacturer’s instructions. Two siRNA duplexes specific<br />

for human MYH9 and negative control siRNA were used (Dharmacon). Cells<br />

were cultured for 48 h and were analyzed by immunoblot or immunofluorescence.<br />

Suppression of the target protein was verified by immunoblot.<br />

Statistical analysis. Prism software was used for nonparametric t-tests.<br />

Accession codes UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signalinggateway.org):<br />

A001394, A002396, A001392 and A004003.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank D. Garbett for help with data analysis with Volocity and for<br />

comments, and D.W. Pruyne for help in setting up DIC microscopy. Supported<br />

by the European Molecular Biology Organization (T.I.) and the US National<br />

Institutes of Health (GM36652 to A.B.; and AI44931 and Nanomedicine<br />

Development Center EY16586 to M.L.D.).<br />

AUTHOR CONTRIBUTIONS<br />

The laboratories of A.B. and M.L.D. did independent work on the involvement of<br />

myosin IIA in the formation of immunological synapses and continued the work<br />

collaboratively focusing on studies in the human system initiated by T.I. and A.B.;<br />

T.I. conceived and did the experiments in Figures 2–4, 5a,b and 6; T.I., G.V.-S.<br />

and S.V. collaborated on Figures 1, 5c and 7; and T.I. and A.B. wrote the first<br />

draft of the manuscript, which M.L.D. extensively edited and revised.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Davis, M.M. The ab T cell repertoire comes into focus. Immunity 27, 179–180 (2007).<br />

2. DeMond, A.L., Mossman, K.D., Starr, T., Dustin, M.L. & Groves, J.T. T cell receptor<br />

microcluster transport through molecular mazes reveals mechanism of translocation.<br />

Biophys. J. 94, 3286–3292 (2008).<br />

3. Monks, C.R., Freiberg, B.A., Kupfer, H., Sciaky, N. & Kupfer, A. Three-dimensional<br />

segregation of supramolecular activation clusters in T cells. <strong>Nature</strong> 395, 82–86<br />

(1998).<br />

4. Dustin, M.L. et al. A novel adaptor protein orchestrates receptor patterning and<br />

cytoskeletal polarity in T-cell contacts. Cell 94, 667–677 (1998).<br />

5. Grakoui, A. et al. The immunological synapse: a molecular machine controlling T cell<br />

activation. Science 285, 221–227 (1999).<br />

6. Jacobelli, J., Chmura, S.A., Buxton, D.B., Davis, M.M. & Krummel, M.F. A single class<br />

II myosin modulates T cell motility and stopping, but not synapse formation. Nat.<br />

Immunol. 5, 531–538 (2004).<br />

7. Combs, J. et al. Recruitment of dynein to the Jurkat immunological synapse. Proc. Natl.<br />

Acad. Sci. USA 103, 14883–14888 (2006).<br />

8. Bunnell, S.C. et al. T cell receptor ligation induces the formation of dynamically<br />

regulated signaling assemblies. J. Cell Biol. 158, 1263–1275 (2002).<br />

9. Campi, G., Varma, R. & Dustin, M.L. Actin and agonist MHC-peptide complexdependent<br />

T cell receptor microclusters as scaffolds for signaling. J. Exp. Med. 202,<br />

1031–1036 (2005).<br />

10. Huse, M. et al. Spatial and temporal dynamics of T cell receptor signaling with a<br />

photoactivatable agonist. Immunity 27, 76–88 (2007).<br />

11. Yokosuka, T. et al. Newly generated T cell receptor microclusters initiate and sustain T<br />

cell activation by recruitment of Zap70 and SLP-76. Nat. Immunol. 6, 1253–1262<br />

(2005).<br />

12. Douglass, A.D. & Vale, R.D. Single-molecule microscopy reveals plasma membrane<br />

microdomains created by protein-protein networks that exclude or trap signaling<br />

molecules in T cells. Cell 121, 937–950 (2005).<br />

13. Varma, R., Campi, G., Yokosuka, T., Saito, T. & Dustin, M.L. T cell receptor-proximal<br />

signals are sustained in peripheral microclusters and terminated in the central<br />

supramolecular activation cluster. Immunity 25, 117–127 (2006).<br />

14. Lee, K.H. et al. The immunological synapse balances T cell receptor signaling and<br />

degradation. Science 302, 1218–1222 (2003).<br />

15. Dustin, M.L. & Cooper, J.A. The immunological synapse and the actin cytoskeleton:<br />

molecular hardware for T cell signaling. Nat. Immunol. 1, 23–29 (2000).<br />

16. Chakraborty, A.K. How and why does the immunological synapse form? Physical<br />

chemistry meets cell biology. Sci. STKE 2002, PE10 (2002).<br />

17. Kaizuka, Y., Douglass, A.D., Varma, R., Dustin, M.L. & Vale, R.D. Mechanisms for<br />

segregating T cell receptor and adhesion molecules during immunological synapse<br />

formation in Jurkat T cells. Proc. Natl. Acad. Sci. USA 104, 20296–20301 (2007).<br />

18. Lin, C.H., Espreafico, E.M., Mooseker, M.S. & Forscher, P. Myosin drives retrograde Factin<br />

flow in neuronal growth cones. Neuron 16, 769–782 (1996).<br />

19. Conti, M.A. & Adelstein, R.S. Nonmuscle myosin II moves in new directions. J. Cell Sci.<br />

121, 11–18 (2008).<br />

20. Tan, J.L., Ravid, S. & Spudich, J.A. Control of nonmuscle myosins by phosphorylation.<br />

Annu. Rev. Biochem. 61, 721–759 (1992).<br />

21. Simons, M. et al. Human nonmuscle myosin heavy chains are encoded by two genes<br />

located on different chromosomes. Circ. Res. 69, 530–539 (1991).<br />

22. Golomb, E. et al. Identification and characterization of nonmuscle myosin II–C,<br />

a new member of the myosin II family. J. Biol. Chem. 279, 2800–2808 (2004).<br />

23. Maupin, P., Phillips, C.L., Adelstein, R.S. & Pollard, T.D. Differential localization of<br />

myosin-II isozymes in human cultured cells and blood cells. J. Cell Sci. 107,<br />

3077–3090 (1994).<br />

24. Wulfing, C. & Davis, M.M. A receptor/cytoskeletal movement triggered by costimulation<br />

during T cell activation. Science 282, 2266–2269 (1998).<br />

25. Straight, A.F. et al. Dissecting temporal and spatial control of cytokinesis with a myosin<br />

II Inhibitor. Science 299, 1743–1747 (2003).<br />

26. Blanchard, N. et al. Strong and durable TCR clustering at the T/dendritic cell immune<br />

synapse is not required for NFAT activation and IFN-g production in human CD4 +<br />

Tcells.J. Immunol. 173, 3062–3072 (2004).<br />

27. Ilani, T., Khanna, C., Zhou, M., Veenstra, T.D. & Bretscher, A. Immune synapse<br />

formation requires ZAP-70 recruitment by ezrin and CD43 removal by moesin.<br />

J. Cell Biol. 179, 733–746 (2007).<br />

28. Weiss, A., Imboden, J., Shoback, D. & Stobo, J. Role of T3 surface molecules in human<br />

T-cell activation: T3-dependent activation results in an increase in cytoplasmic free<br />

calcium. Proc. Natl. Acad. Sci. USA 81, 4169–4173 (1984).<br />

29. Valitutti, S., Muller, S., Cella, M., Padovan, E. & Lanzavecchia, A. Serial triggering of<br />

many T-cell receptors by a few peptide-MHC complexes. <strong>Nature</strong> 375, 148–151<br />

(1995).<br />

30. Krogsgaard, M. et al. Agonist/endogenous peptide-MHC heterodimers drive T cell<br />

activation and sensitivity. <strong>Nature</strong> 434, 238–243 (2005).<br />

31. Janeway, C.A. Jr. & Bottomly, K. Responses of T cells to ligands for the T-cell receptor.<br />

Semin. Immunol. 8, 108–115 (1996).<br />

32. Sims, T.N. et al. Opposing effects of PKCy andWASponsymmetrybreakingand<br />

relocation of the immunological synapse. Cell 129, 773–785 (2007).<br />

33. Barda-Saad, M. et al. Dynamic molecular interactions linking the T cell antigen<br />

receptor to the actin cytoskeleton. Nat. Immunol. 6, 80–89 (2005).<br />

34. Dustin, M.L. & Springer, T.A. T-cell receptor cross-linking transiently stimulates<br />

adhesiveness through LFA-1. <strong>Nature</strong> 341, 619–624 (1989).<br />

35. Fleire, S.J. et al. B cell ligand discrimination through a spreading and contraction<br />

response. Science 312, 738–741 (2006).<br />

36. Giannone, G. et al. Lamellipodial actin mechanically links myosin activity with<br />

adhesion-site formation. Cell 128, 561–575 (2007).<br />

37. Pasternak, C., Spudich, J.A. & Elson, E.L. Capping of surface receptors and concomitant<br />

cortical tension are generated by conventional myosin. <strong>Nature</strong> 341, 549–551<br />

(1989).<br />

538 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

38. Mabuchi, I. & Okuno, M. The effect of myosin antibody on the division of starfish<br />

blastomeres. J. Cell Biol. 74, 251–263 (1977).<br />

39. Matsumura, F. Regulation of myosin II during cytokinesis in higher eukaryotes. Trends<br />

Cell Biol. 15, 371–377 (2005).<br />

40. Munro, E., Nance, J. & Priess, J.R. Cortical flows powered by asymmetrical contraction<br />

transport PAR proteins to establish and maintain anterior-posterior polarity in the early<br />

C. elegans embryo. Dev. Cell 7, 413–424 (2004).<br />

41. Mescher, M.F. Surface contact requirements for activation of cytotoxic T lymphocytes.<br />

J. Immunol. 149, 2402–2405 (1992).<br />

42. Galbraith, C.G., Yamada, K.M. & Sheetz, M.P. The relationship between force and focal<br />

complex development. J. Cell Biol. 159, 695–705 (2002).<br />

43. Lillemeier, B.F., Pfeiffer, J.R., Surviladze, Z., Wilson, B.S. & Davis, M.M. Plasma<br />

membrane-associated proteins are clustered into islands attached to the cytoskeleton.<br />

Proc. Natl. Acad. Sci. USA 103, 18992–18997 (2006).<br />

44. Gallagher, P.J., Herring, B.P., Griffin, S.A. & Stull, J.T. Molecular characterization of a<br />

mammalian smooth muscle myosin light chain kinase. J. Biol. Chem. 266,<br />

23936–23944 (1991).<br />

ARTICLES<br />

45. Ludowyke, R.I., Peleg, I., Beaven, M.A. & Adelstein, R.S. Antigen-induced secretion of<br />

histamine and the phosphorylation of myosin by protein kinase C in rat basophilic<br />

leukemia cells. J. Biol. Chem. 264, 12492–12501 (1989).<br />

46. Koretzky, G.A., Abtahian, F. & Silverman, M.A. SLP76 and SLP65: complex regulation<br />

of signalling in lymphocytes and beyond. Nat. Rev. Immunol. 6, 67–78 (2006).<br />

47. Zeng, R. et al. SLP-76 coordinates Nck-dependent Wiskott-Aldrich syndrome protein<br />

recruitment with Vav-1/Cdc42-dependent Wiskott-Aldrich syndrome protein activation<br />

at the T cell-APC contact site. J. Immunol. 171, 1360–1368 (2003).<br />

48. Dustin, M.L. & Springer, T.A. Lymphocyte function-associated antigen-1 (LFA-1)<br />

interaction with intercellular adhesion molecule-1 (ICAM-1) is one of at least three<br />

mechanisms for lymphocyte adhesion to cultured endothelial cells. J. Cell Biol. 107,<br />

321–331 (1988).<br />

49. Vasiliver-Shamis, G. et al. HIV-1 envelope gp120 induces a stop signal and virological<br />

synapse formation in non-infected CD4 + Tcells.J. Virol. 82, 9445–9457 (2008).<br />

50. Bretscher, A. Rapid phosphorylation and reorganization of ezrin and spectrin accompany<br />

morphological changes induced in A-431 cells by epidermal growth factor. J. Cell<br />

Biol. 108, 921–930 (1989).<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 539


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

HoxC4 binds to the promoter of the cytidine<br />

deaminase AID gene to induce AID expression, classswitch<br />

DNA recombination and somatic hypermutation<br />

Seok-Rae Park 1,2 , Hong Zan 1,2 , Zsuzsanna Pal 1 , Jinsong Zhang 1 , Ahmed Al-Qahtani 1 , Egest J Pone 1 ,<br />

Zhenming Xu 1 , Thach Mai 1 & Paolo Casali 1<br />

The cytidine deaminase AID (encoded by Aicda in mice and AICDA in humans) is critical for immunoglobulin class-switch<br />

recombination (CSR) and somatic hypermutation (SHM). Here we show that AID expression was induced by the HoxC4<br />

homeodomain transcription factor, which bound to a highly conserved HoxC4-Oct site in the Aicda or AICDA promoter. This site<br />

functioned in synergy with a conserved binding site for the transcription factors Sp1, Sp3 and NF-jB. HoxC4 was ‘preferentially’<br />

expressed in germinal center B cells and was upregulated by engagement of CD40 by CD154, as well as by lipopolysaccharide<br />

and interleukin 4. HoxC4 deficiency resulted in impaired CSR and SHM because of lower AID expression and not some other<br />

putative HoxC4-dependent activity. Enforced expression of AID in Hoxc4 –/– B cells fully restored CSR. Thus, HoxC4 directly<br />

activates the Aicda promoter, thereby inducing AID expression, CSR and SHM.<br />

Class-switch recombination (CSR) and somatic hypermutation<br />

(SHM) are critical for the maturation of antibody responses to foreign<br />

and self antigens. CSR recombines switch-region DNA located<br />

upstream of exons of the constant heavy-chain (C H) region, thereby<br />

altering immunoglobulin CH regions and endowing antibodies with<br />

new biological effector functions. SHM introduces mainly point<br />

mutations in immunoglobulin variable regions, thereby providing<br />

the structural substrate for the selection by antigen of antibody<br />

mutants with higher affinity. Despite advances in the identification<br />

of some factors involved in CSR and SHM, details of the mechanisms<br />

of these processes remain elusive. Both CSR and SHM require<br />

activation-induced cytidine deaminase (AID), which is expressed by<br />

activated B cells, mainly in germinal centers of peripheral lymphoid<br />

organs 1,2 . AID initiates CSR and SHM by deaminating deoxycytosine<br />

residues to yield deoxyuracil-deoxyguanine mispairings in<br />

DNA 3–8 . These mispairings trigger DNA-repair processes, which<br />

entails the introduction of mutations in variable-(diversity)-joining<br />

(V(D)J) regions or DNA breaks, including double-stranded DNA<br />

breaks, which lead to nonclassical nonhomologous end joining<br />

and CSR 3,5,9–14 .<br />

The mechanisms governing the transcriptional regulation of the<br />

gene encoding AID (AICDA in humans and Aicda in mice) remain to<br />

be elucidated. A conserved region in the first intron of Aicda containing<br />

two E-boxes, the consensus sequence for binding of the transcription<br />

factor E2A (A000804), has been suggested to contribute to the<br />

regulation of Aicda transcription through recruitment of the E2A<br />

Received 17 December 2008; accepted 9 March 2009; published online 12 April 2009; doi:10.1038/ni.1725<br />

helix-loop-helix transcription factor E47 and the inhibitor of DNA<br />

binding helix-loop-helix protein Id3, respectively 15 . The B cell lineage–<br />

specific transcription factor Pax5 is thought to act with E2A proteins<br />

in controlling Aicda transcription 16 . However, this was not confirmed<br />

by another study, which has instead suggested involvement of the Sp1<br />

family of ubiquitous zinc-finger transcription factors. These regulate<br />

various promoters by binding to deoxyguanine-deoxycytosine,<br />

deoxyguanine-deoxyadenine or deoxyguanine-deoxythymine boxes in<br />

activating the Aicda promoter 17 .<br />

Hox proteins are highly conserved helix-loop-helix homeodomaincontaining<br />

transcription factors that regulate cellular differentiation<br />

and organogenesis 18,19 . Genes encoding these proteins, which are<br />

clustered chromosomally, are expressed in a temporally and spatially<br />

regulated way 20,21 . Among the human genes, those encoding the<br />

HoxC proteins, particularly HOXC4, are expressed mainly in lymphoid<br />

cells 22 . HOXC4 expression increases through sequential stages<br />

of B cell development 22–25 , from uncommitted hematopoietic progenitors<br />

in the bone marrow to mature B cells in the periphery,<br />

particularly when they are activated and proliferating. Malignant<br />

B cells, including those of mantle cell lymphoma, Burkitt’s lymphoma<br />

and B cell chronic lymphocytic leukemia, have aberrant AID expression<br />

26,27 and abundant HoxC4 expression 22,28 . HoxC4 induces the<br />

human 3¢ E a enhancer elements, particularly DNAse I–hypersensitive<br />

sites 1 and 2 (called ‘hs1,2’ here), in a B cell development stage–<br />

specific way 25 . HoxC4 binds to a HoxC4-Oct motif with the sequence<br />

5¢-ATTTGCAT-3¢ in hs1,2 (refs. 24,25), which is conserved in humans,<br />

1 Institute for <strong>Immunology</strong>, School of Medicine and School of Biological Sciences, University of California, Irvine, California, USA. 2 These authors contributed equally to<br />

this work. Correspondence should be addressed to P.C. (pcasali@uci.edu).<br />

540 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

Aicda expression (fold) Hoxc4 expression (fold)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

660<br />

600<br />

480<br />

360<br />

240<br />

120<br />

0<br />

Bone<br />

marrow<br />

Thymus<br />

Spleen<br />

Peyer’s<br />

patch<br />

Liver<br />

Heart<br />

b<br />

HoxC4<br />

AID<br />

PCNA<br />

β-actin<br />

PNA hi B220 +<br />

PNA lo B220 +<br />

mice, rats and rabbits, and acts in synergy with homeodomain<br />

transcription factors Oct1 and Oct2 (A001704; collectively called<br />

‘Oct’ here) and the Oca-B coactivator (A001696) to induce hs1,2 in<br />

B cells 24,25 . HOXC4 expression is induced by stimuli that induce the<br />

differentiation of germinal center B cells and AICDA expression 24,25 ,<br />

such as CD154 (A000536) and interleukin 4 (IL-4; A001262), which<br />

suggests involvement of HoxC4 in CSR and SHM.<br />

In this study, we test the hypothesis that HoxC4 regulates AID<br />

expression in human and mouse B cells. We show that HoxC4 bound<br />

to a HoxC4-Oct motif in the AICDA and Aicda promoters that is<br />

conserved in humans, chimps, mice, rats, dogs and cows. Binding of<br />

HoxC4 to this cis element activated the AICDA and Aicda promoters<br />

and induced AID expression, thereby inducing CSR and SHM. In this<br />

function, HoxC4 acted in synergy with an equally conserved upstream<br />

binding site for the transcription factors Sp1 and Sp3 (collectively<br />

called ‘Sp’ here) and NF-kB intheAICDA and Aicda promoters.<br />

RESULTS<br />

HoxC4 and AID are induced in germinal center B cells<br />

HoxC4 is upregulated in human immunoglobulin D–negative (IgD – )<br />

CD38 + germinal center B cells 24,25 , which express AID and undergo<br />

CSR and SHM. Stimulation of human IgD + CD38 – naive B cells with<br />

an agonistic monoclonal antibody (mAb) to CD40 and human IL-4<br />

upregulates HoxC4 and induced AID expression 24,25 .Herewefurther<br />

analyzed the expression of Hoxc4 and Aicda in the bone marrow,<br />

thymuses, spleens, Peyer’s patches, livers and hearts of wild-type<br />

C57BL/6 mice. Real-time quantitative RT-PCR showed that like<br />

Aicda, Hoxc4 was expressed mainly in the spleen and Peyer’s patches,<br />

which contain a large proportion of hypermutating and switching<br />

B cells but not in nonlymphoid organs such as the liver or heart<br />

(Fig. 1a). To further address the correlation between the expression of<br />

HoxC4 and AID, we isolated PNA hi B220 + (germinal center) B cells<br />

and PNA lo B220 + (non–germinal center) B cells from the spleens of<br />

8- to 10-week-old C57BL/6 mice 14 d after immunizing the mice with<br />

NP 16-CGG (16 molecules of 4-hydroxy-3-nitrophenyl acetyl coupled<br />

to 1 molecule of chicken g-globulin) and measured HoxC4 and AID,<br />

as well as PCNA, a multifunctional protein critical for DNA replication<br />

and repair that has high expression in actively dividing cells.<br />

HoxC4 was specifically expressed in PNA hi B220 + germinal center<br />

B cells, which also had high expression of AID and PCNA (Fig. 1b).<br />

Stimulation of mouse spleen B cells with bacterial lipopolysaccharide<br />

(LPS) and IL-4 or CD154 and IL-4, which induce the differentiation of<br />

germinal center B cells and Aicda expression, upregulated Hoxc4<br />

expression by 10- to 15-fold (Fig. 1c) and induced CSR to IgG1<br />

(data not shown), which indicates that HoxC4 is involved in inducing<br />

AID expression.<br />

c<br />

Aicda expression (fold) Hoxc4 expression (fold)<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

+ nil<br />

+ LPS + IL-4<br />

+ CD154 + IL-4<br />

1 2<br />

Experiment<br />

3<br />

ARTICLES<br />

Figure 1 Hoxc4 expression correlates with Aicda expression. (a) Real-time<br />

quantitative RT-PCR analysis of Hoxc4 and Aicda transcripts in bone<br />

marrow, thymus, spleen, Peyer’s patches, liver and heart of C57BL/6 mice,<br />

normalized to expression of Gapdh (encoding glyceraldehyde phosphate<br />

dehydrogenase) and presented relative to mRNA in bone marrow, set as 1.<br />

Data are from three independent experiments (mean and s.e.m. of triplicate<br />

samples). (b) Immunoblot analysis of HoxC4, AID, PCNA and b-actin in<br />

PNA hi B220 + (germinal center) B cells and PNA lo B220 + (non–germinal<br />

center) B cells from spleens of mice immunized with NP 16-CGG. Data are<br />

representative of two independent experiments. (c) Real-time quantitative<br />

RT-PCR analysis of the expression of Hoxc4 and Aicda mRNA in C57BL/6<br />

splenic B cells stimulated for 3 d with LPS plus IL-4 or with CD154 plus<br />

IL-4, normalized to CD79b transcripts and presented relative to expression<br />

in unstimulated B cells (+ nil), set as 1. Data are from three independent<br />

experiments (mean and s.e.m.; n ¼ 3 mice).<br />

HoxC4 deficiency impairs the antibody response to NP-CGG<br />

We used Hoxc4 –/– mice to address the function of HoxC4 in CSR and<br />

SHM. Two laboratories independently generated two lines of HoxC4deficient<br />

mice; homozygous mutants of both lines had esophageal<br />

defects and abnormal cervical and thoracic vertebral development and<br />

suffered high postnatal mortality rates 29,30 . In these mice, the expression<br />

of Hoxc5 and Hoxc6, which are in the same gene cluster as Hoxc4,<br />

was lower, probably because of the effect of the neighboring neomycinresistance<br />

selection cassette inserted in the Hoxc4 locus 29,30 .Toobviate<br />

that effect, another laboratory generated a third Hoxc4 –/– strain in<br />

which the neomycin-resistance cassette was deleted by Cre recombinase<br />

through two flanking loxP sites in the germline (Supplementary Fig. 1<br />

online), thereby leaving the expression of Hoxc5 and Hoxc6 unaltered<br />

(A.M. Boulet and M.R. Capecchi, unpublished data). Those mice have<br />

since been lost, but frozen Hoxc4 +/– sperm on the C57BL/6 background<br />

was preserved. Using that Hoxc4 +/– sperm, we rederived Hoxc4 +/– mice<br />

by in vitro fertilization and bred them to obtain Hoxc4 –/– mice. These<br />

Hoxc4 –/– mice were born at the expected mendelian ratios, did not<br />

suffer the high postnatal mortality rate of the earlier HoxC4-deficient<br />

mouse lines 29,30 and developed to adulthood.<br />

In unimmunized Hoxc4 –/– mice, serum IgM titers were normal.<br />

However, the average serum IgG1 concentration was less than<br />

0.6 mg/ml, compared with 1.2 mg/ml in their Hoxc4 +/+ littermates<br />

(data not shown), which suggested impairment of CSR. We immunized<br />

four pairs of 8- to 10-week-old littermate Hoxc4 –/– and Hoxc4 +/+<br />

mice with NP 16-CGG and analyzed blood from these mice for titers of<br />

IgM and IgG1; IgM and IgG1 bound to NP30 (30 molecules of NP<br />

coupled to BSA); and IgM and IgG1 bound with high affinity to NP 3<br />

(3 molecules of NP coupled to BSA; Fig. 2a). Titers of total IgM and<br />

NP30-binding IgM were not significantly different in Hoxc4 –/– and<br />

Hoxc4 +/+ mice. Hoxc4 –/– mice, however, had slightly less NP 3-binding<br />

IgM and significantly lower titers of total IgG1 and NP 30-binding<br />

IgG1, as well as of high-affinity NP3-binding IgG1. The defective<br />

antibody response to NP-CGG did not reflect altered development of<br />

plasma cells or memory B cells, as the proportions of B220 lo CD138 +<br />

cells and CD38 hi B cells among NP-binding IgG1 B cells in NP16-<br />

CGG-immunized Hoxc4 –/– mice were similar to those of their<br />

Hoxc4 +/+ littermates (Fig. 2b,c). Instead, it reflected the lower overall<br />

IgG1 and, together with the slightly lower NP 3-binding IgM activity, a<br />

lower binding affinity for NP.<br />

HoxC4 deficiency does not alter germinal center formation<br />

The defective antibody response to NP-CGG in Hoxc4 –/– mice was not<br />

due to obvious defects in lymphoid differentiation. In these mice, the<br />

size of the spleen and the number and size of Peyer’s patches were<br />

similar to those of Hoxc4 +/+ mice (data not shown). Moreover, the<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 541


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

IgM (mgeq/ml) IgG1 (mgeq/ml)<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0.0<br />

P = 0.019 P = 0.022 P = 0.026<br />

NP30-binding IgG1<br />

EC50 (× 103 )<br />

1.6<br />

1.2<br />

0.8<br />

0.4<br />

0.0<br />

Hoxc4 +/+ Hoxc4 –/–<br />

Hoxc4 +/+ Hoxc4 –/–<br />

NS<br />

80<br />

NS<br />

60<br />

40<br />

20<br />

0<br />

NP30-binding IgM<br />

EC50 (× 103 )<br />

b c<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

B220-PE<br />

80<br />

60<br />

40<br />

20<br />

0<br />

NP3-binding IgG1<br />

EC50 (× 103 )<br />

NP3-binding IgM<br />

EC50 (× 103 )<br />

Hoxc4 +/+ Hoxc4 –/–<br />

80<br />

60<br />

40<br />

20<br />

0<br />

P = 0.048<br />

80<br />

60<br />

40<br />

20<br />

0<br />

10 4<br />

10 3<br />

10 2<br />

10 1.8<br />

1.9<br />

1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

33.3<br />

34.3<br />

10 0<br />

10<br />

CD138-FITC IgG1-APC CD38-PECY7<br />

1 10 2 10 3 10 4 10 0 10 1 10 2 10 3 10 4<br />

10 0 10 1 10 2 10 3 10 4<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

NP-PE<br />

number of B cells and T cells, the proportion of CD4 + T cells, and the<br />

viability of B cells and T cells in the spleen and Peyer’s patches, as<br />

analyzed by staining with 7-amino-actinomycin D, were also similar to<br />

those of Hoxc4 +/+ mice (Fig. 3a–d). After stimulation with LPS and<br />

IL-4, Hoxc4 –/– B lymphocytes were similar to Hoxc4 +/+ cells in terms<br />

of cell cycle, as analyzed by staining with propidium iodide, and cell<br />

division rate, as measured by incorporation of the vital dye CFSE<br />

(Fig. 3e,f). In Hoxc4 –/– mice, the number and architecture of germinal<br />

centers in the spleen, the proportion of proliferating B cells, as shown<br />

by in vivo incorporation of the thymidine analog BrdU, as well as the<br />

proportion of PNA hi germinal center B cells in both spleen and Peyer’s<br />

patches, were similar to those of Hoxc4 +/+ mice (Fig. 3g–i), which<br />

suggested that the defective antibody response to NP 16-CGG in<br />

Hoxc4 –/– mice reflected an intrinsic impairment of the CSR and<br />

SHM ‘machinery’.<br />

HoxC4 deficiency impairs CSR and SHM<br />

To assess the effect of HoxC4 deficiency on CSR, we stimulated<br />

spleen B cells with LPS (to induce switching to IgG2b and IgG3),<br />

with LPS or CD154 plus IL-4 (to induce switching to IgG1 and<br />

IgE), with LPS or CD154 plus IFN-g (A001238; to induce<br />

switching to IgG2a), or with LPS or CD154 in the presence of<br />

transforming growth factor-b1 (TGF-b1; A002271), IL-5, IL-4 and<br />

dextran-conjugated mAb to d-chain (to induce switching to IgA).<br />

After 4 d of stimulation with LPS alone or LPS plus cytokines, the<br />

proportion of Hoxc4 –/– B cells positive for surface IgG1, IgG2a,<br />

IgG2b, IgG3, IgA and IgE was 54%, 49%, 46%, 36%, 43% and 57%<br />

lower, respectively, than that of the corresponding Hoxc4 +/+<br />

B cells (Supplementary Fig. 2 online); in cultures stimulated<br />

with CD154 and cytokines, the proportion of Hoxc4 –/– B cells<br />

positive for surface IgG1, IgG2a and IgE was 49%, 69% and 79%<br />

lower, respectively. Consistent with those findings, after 7 d,<br />

secretion of IgG1, IgG2a, IgG2b, IgG3, IgA and IgE by Hoxc4 –/–<br />

B cells stimulated with LPS alone or LPS plus cytokines was as<br />

much as 48%, 55%, 37%, 35%, 40% and 66% lower, respectively<br />

Cells<br />

Figure 2 Impaired antibody response in Hoxc4 –/– mice. (a) Titers of<br />

circulating IgM and IgG1 (left), NP 30-binding IgM and IgG1 (middle) and<br />

high-affinity NP 3-binding IgM and IgG1 (right) in Hoxc4 +/+ and Hoxc4 –/–<br />

littermates (n ¼ 4 pairs of mice; each symbol represents an individual<br />

mouse) immunized with NP16-CGG and ‘boosted’ 21 d later, measured 7 d<br />

after the boost injection and presented as milligram equivalents (mgeq; left)<br />

or the number of dilutions needed to reach 50% of saturation binding<br />

(EC50; middle and right). NS, not significant (P values, paired t-test).<br />

(b,c) Development of plasma cells and memory B cells in Hoxc4 +/+ and<br />

Hoxc4 –/– littermates 14 d after immunization with NP 16-CGG. (b) Surface<br />

expression of B220 and CD138 on spleen cells. Numbers in outlined areas<br />

indicate percent B220 lo CD138 + (plasma) cells as percentage of total B220 +<br />

cells. (c) Flow cytometry of spleen cells stained with FITC-labeled PNA,<br />

phycoerythrin (PE)-labeled NP, allophycocyanin (APC)-labeled mAb to<br />

mouse IgG1 and phycoerythrin-indotricarbocyanine (PECy7)-labeled mAb to<br />

mouse CD38. Insets (left), NP-binding surface IgG1 + B cells; right, numbers<br />

above bracketed lines indicate percent CD38 + cells among those gated<br />

NP-binding IgG1 + B cells. Data are representative of four (a) or three<br />

(b,c) independent experiments.<br />

(Fig. 4a). Impaired CSR in Hoxc4 –/– B cells was not due to altered<br />

proliferation, as after 2, 3 or 4 d of culture with LPS plus IL-4 or<br />

with CD154 plus IL-4, Hoxc4 –/– B cells completed the same<br />

number of divisions as their Hoxc4 +/+ counterparts did (Figs. 3f<br />

and 4b).Itwasalsonotduetoalteredplasma cell differentiation,<br />

as after 4 d of culture with LPS alone, LPS plus IL-4, or CD154<br />

plus IL-4, the number of CD138 + B220 lo plasma cells that emerged<br />

from Hoxc4 –/– B cells was similar to that of their Hoxc4 +/+<br />

counterparts (Fig. 4c). Consistent with those findings, expression<br />

of the transcription factors Blimp-1 and IRF4, which are<br />

critical for plasma cell differentiation, was similar in Hoxc4 –/–<br />

and Hoxc4 +/+ B cells, as determined by real-time quantitative<br />

RT-PCR, at 5 d after stimulation with LPS and IL-4 (data not<br />

shown). Further, lower CSR in Hoxc4 –/– Bcellswasnotdueto<br />

impairment of germline transcription of the intervening heavychainregionandconstantheavy-chainregion(I<br />

H-C H), which is<br />

necessary for CSR. Real-time quantitative RT-PCR showed that the<br />

abundance of germline transcripts of I m-C m, I g3-C g3, I g1-C g1,<br />

I g2b-C g2b, I g2a-C g2a, I e-C e and I a-C a in Hoxc4 –/– B cells stimulated<br />

for 3 d with LPS alone or LPS plus cytokines was similar to<br />

that in their Hoxc4 +/+ B cell counterparts (Fig. 5 and Supplementary<br />

Fig. 3 online), whereas post-recombination I m-C H transcripts,<br />

which are generated by CSR, were significantly less<br />

abundant in Hoxc4 –/– B cells, by as much as 89.2%. Thus,<br />

HoxC4 deficiency impairs CSR to all isotypes without affecting<br />

germline I H-C H transcription.<br />

To asses the effect of HoxC4 deficiency on SHM, we analyzed the<br />

IgH J H4–intronic enhancer (iE m) sequence downstream of rearranged<br />

V J558DJ H4 DNA in PNA hi B220 + germinal center B cells from<br />

Peyer’s patches of 12-week-old unimmunized littermate Hoxc4 –/– and<br />

Hoxc4 +/+ mice. In these mice, the proportion of Peyer’s patch<br />

PNA hi B220 + germinal center B cells was similar (Fig. 3h). Analysis<br />

of 324 and 319 JH4-iEm intronic DNA sequences (720 base pairs (bp))<br />

from Hoxc4 –/– and Hoxc4 +/+ mice showed that Hoxc4 –/– mice had<br />

59% fewer mutations (P o 0.00001; Fig. 6a). The lower mutation<br />

frequency was associated with similarly fewer mutations of<br />

deoxyguanine-deoxycytosine and deoxyadenine-deoxythymine (Supplementary<br />

Fig. 4 online) and was not due to impaired transcription<br />

of the rearranged V J558DJ H genes, as shown by specific real-time<br />

quantitative RT-PCR of Peyer’s patch B cells from these Hoxc4 –/– mice<br />

and their Hoxc4 +/+ littermates (Fig. 6b). Thus, HoxC4 deficiency<br />

substantially impairs SHM without altering the spectrum of the<br />

residual mutations or VHDJH transcription.<br />

542 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a<br />

Hoxc4 +/+<br />

Hoxc4<br />

Cells<br />

–/–<br />

Spleen<br />

Peyer’s<br />

patch<br />

B220-PE<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

51.9 2.7 53.9 3.5<br />

5.8<br />

59.1<br />

39.6<br />

6.4<br />

5.8<br />

55.9<br />

37.6<br />

5.0<br />

2.4 32.1 3.4 35.6<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

CD3-FITC CD3-FITC CD3-FITC B220-PE<br />

HoxC4 deficiency impairs AID expression<br />

CSR and SHM require transcription of the Igh locus and AID<br />

expression. The much lower CSR and SHM, together with the normal<br />

abundance of mature VHDJH-Cm and germline IH-CH transcripts, in<br />

Hoxc4 –/– B cells prompted us to hypothesize modulation of AID<br />

expression by HoxC4. We stimulated spleen Hoxc4 +/+ and Hoxc4 –/–<br />

B cells for 0, 12, 24, 48 and 72 h with LPS plus IL-4 or with CD154<br />

plus IL-4. Expression of Aicda mRNA could be detected by real-time<br />

quantitative RT-PCR as early as 24 h and peaked within 48–72 h of<br />

stimulation in Hoxc4 +/+ B cells. In Hoxc4 –/– B cells, Aicda expression<br />

was more than 70% lower after 72 h of stimulation (Fig. 7a). The<br />

lower abundance of Aicda transcripts was associated with much less<br />

AID protein (Fig. 7b), which further suggests that HoxC4 regulates<br />

Aicda expression.<br />

Binding of transcription factors to the Aicda promoter<br />

To address the possibility that HoxC4 modulates Aicda expression<br />

by binding to a cis-regulatory element of this gene, we analyzed the<br />

sequence upstream of the putative transcription-initiation site of<br />

Aicda (Supplementary Fig. 5 online). We identified eight motifs<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

b c d<br />

CD4-PerCP<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

7-AAD<br />

B220-PE<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

e f<br />

g<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Cells<br />

Day 1<br />

Day 2 Day 2 Day 3<br />

250<br />

200<br />

150<br />

100<br />

50<br />

G0-G1 G0-G1 = 90.8%<br />

S S = 3.1%<br />

G2-M G2-M = 4.9%<br />

G0-G1 = 41.4%<br />

G0-G1<br />

S<br />

S = 25.2%<br />

G2-M<br />

G2-M = 16.1%<br />

250<br />

200<br />

150<br />

100<br />

50<br />

765 4321<br />

0 76543210<br />

0<br />

250<br />

200 G0-G1<br />

G0-G1 = 90.4%<br />

G0-G1 = 43.3%<br />

G0-G1<br />

150 S S = 3.1%<br />

S = 28.5%<br />

S<br />

100 G2-M G2-M = 4.3%<br />

G2-M = 14.3%<br />

50<br />

G2-M<br />

0<br />

0 200 400 600 800 1,000 0 200 400 600 800 1,000<br />

0<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

76543<br />

2 1 0<br />

76543210<br />

DNA CFSE<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

h i<br />

Spleen<br />

Peyer’s<br />

patch<br />

Figure 3 HoxC4 deficiency does not affect B cell, T cell or CD4 + T cell<br />

numbers, death of B cells and T cells in spleens and Peyer’s patches or<br />

B cell cycle or division or alter germinal center formation and in vivo B cell<br />

proliferation. (a–d) Flow cytometry of spleen and Peyer’s patch cells stained<br />

with phycoerythrin-labeled mAb to B220 and FITC-labeled mAb to CD3 (a),<br />

FITC-labeled mAb to CD4 and peridinine chlorophyll protein complex<br />

(PerCP)-labeled mAb to CD4 (b), 7-amino-actinomycin D (7-AAD) and<br />

FITC-labeled mAb to CD3 (c), or 7-amino-actinomycin and phycoerythrinlabeled<br />

mAb to B220 (d). Data are representative of three experiments.<br />

(e) Cell cycle analysis of Hoxc4 +/+ and Hoxc4 –/– splenic B cells stimulated<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

conserved in humans, chimps, mice, rats, dogs and cows (Supplementary<br />

Fig. 5). The first six motifs did not fulfill the minimal<br />

criteria for known transcription factor–binding sites by weightmatrix<br />

search with the Match program (threshold score, 0.75). The<br />

final two were a HoxC4- or Oct-binding site with a sequence of<br />

5¢-ATTTGAAT-3¢ (residues –29 to –22 in humans and mice; scores,<br />

1.0 for HoxC4 and 0.93 for Oct), which was nearly identical to the<br />

conserved HoxC4-Oct motif that is critical in inducing the human<br />

IGH 3¢ E a enhancer elements 24,25 , and an upstream Sp–NF-kB–<br />

binding site with a sequence of 5¢-GGGGAGGAGCC-3¢ (residues<br />

–57 to –47 in humans and mice 17 ; scores, 0.93 for Sp1 and Sp3, and<br />

0.96 for NF-kB). This sequence (5¢-GGGGAGGAGCC-3¢) hasbeen<br />

suggested to be a Pax5-binding site 16 , but it did not satisfy, in our<br />

analysis, the minimum requirement for such a binding site (a score<br />

of 0.61). Both the putative HoxC4-Oct–binding site and the<br />

Sp–NF-kB–binding site were identical in all six species analyzed.<br />

To determine the function of the region from position –349 to<br />

position –1 in the Aicda promoter (tentatively defined as the Aicda<br />

promoter on the basis of its high conservation) in the regulation of<br />

Aicda transcription, we constructed pGL3–luciferase reporter vectors<br />

7-AAD<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

B220-PE<br />

Hoxc4 +/+<br />

PNA-FITC BrdU-APC<br />

ARTICLES<br />

with LPS plus IL-4 and collected after 1 and 2 d for propidium iodide staining and flow cytometry for measurement of DNA content and cells in the G0-G1,<br />

S and G2-M phases. Data are representative of three experiments. (f) Cell division by Hoxc4 +/+ and Hoxc4 –/– splenic B cells labeled with CFSE, cultured with<br />

LPS plus IL-4, and collected 2 and 3 d later for flow cytometry. Progressive left shift of fluorescence indicates cell division; numbers in plots indicate cell<br />

generations. Data are from one representative of three experiments. (g) Staining of germinal centers in spleen sections prepared 10 d after immunization of<br />

Hoxc4 +/+ and Hoxc4 –/– mice with NP16-CGG. Scale bars, 200 mm. Results are representative of three experiments. (h) Flow cytometry of cells obtained from<br />

Peyer’s patches from NP16-CGG–immunized Hoxc4 +/+ and Hoxc4 –/– mice and stained with phycoerythrin-labeled mAb to B220 and FITC-labeled PNA. Data<br />

are representative of three experiments (n ¼ 3 pairs of mice). (i) In vivo proliferation of B cells from spleens and Peyer’s patches of 10-week-old Hoxc4 +/+<br />

and Hoxc4 –/– mice (n ¼ 3 pairs) immunized with NP16-CGG, then, 10 d later, injected with BrdU (1 mg) twice within 16 h and analyzed 4 h after the final<br />

injection; cells were stained with phycoerythrin-labeled mAb to B220 or that mAb plus FITC-labeled PNA; incorporated BrdU was detected by flow cytometry<br />

with allophycocyanin-labeled mAb to BrdU. Data are representative of three experiments. Numbers in quadrants (a–d,h,i) indicate percent cells in each.<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 543<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Hoxc4 +/+<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Hoxc4 –/–<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

Hoxc4 –/–<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a P = 0.034 P = 0.0049 P = 0.010 P = 0.020 P = 0.049 P = 0.0068 b<br />

IgG1 (ngeq/ml)<br />

1,400<br />

1,200<br />

1,000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

IgG2b (ngeq/ml)<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

IgG3 (ngeq/ml)<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

consisting of the 349-bp Aicda promoter and/or the 490-bp flanking 5¢<br />

region (Fig. 7c) upstream of the gene encoding the luciferase reporter,<br />

combined with nothing more or with conserved region1 (cr1) and/or<br />

cr2, which are in the first intron in the Aicda 15,16 .Weusedthese<br />

vectors to transfect human Ramos B cells, which spontaneously<br />

express AICDA and undergo SHM, and mouse CH12F3-2A B cells,<br />

which express Aicda and undergo CSR after stimulation with LPS,<br />

IL-4 and TGF-b1. We cultured these B cells and measured luciferase<br />

activity after 16 h (Ramos) or 24 h (CH12F3-2A). The construct that<br />

contained both the Aicda promoter and its flanking 5¢ region had<br />

11- to 15-fold more activity than did the empty vector in Ramos and<br />

CH12F3-2A B cells (Fig. 7d). Neither cr1 nor cr2 showed substantial<br />

enhancer activity in either Ramos or CH12F3-2A B cells. In addition,<br />

whereas the construct that contained only the Aicda promoter<br />

promoted transcription as efficiently as that containing both the<br />

Aicda promoter and its flanking 5¢ region, the construct that included<br />

only the flanking 5¢ region had only ‘background’ activity similar to<br />

that of empty vector in both Ramos B cells and stimulated CH12F3-<br />

2A B cells. These experiments show that the 349-bp Aicda promoter<br />

region had full promoter activity, whereas neither cr1 nor cr2<br />

enhanced Aicda promoter activity in human or mouse B cells.<br />

To address the specificity of the two evolutionarily conserved<br />

5¢-ATTTGAAT-3¢ and 5¢-GGGGAGGAGCC-3¢ motifs, we did<br />

electrophoretic mobility-shift assay (EMSA) with wild-type and<br />

mutated oligonucleotide probes encompassing residues –65 to –14<br />

of the human AICDA promoter sequence and containing both the<br />

HoxC4-Oct–binding site and the Sp–NF-kB–binding site (Sp-Hox),<br />

or an oligonucleotide probe encompassing residues –65 to –44 of the<br />

AICDA promoter sequence containing only the Sp–NF-kB–binding<br />

site (Sp; Supplementary Fig. 6a online). Incubation of nuclear<br />

extracts of human 4B6 B cells, which spontaneously express AICDA<br />

and undergo CSR, or Ramos B cells, with the Sp-Hox probe gave<br />

rise to four main protein–DNA complexes: A and A¢, and B and<br />

B¢ (Supplementary Fig. 6b,c), specific for the binding of HoxC4-Oct<br />

IgG2a (ngeq/ml)<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

c<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

B220-PE<br />

IgA (ngeq/ml)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

IgE (ngeq/ml)<br />

and Sp–NF-kB, respectively. The mutated oligonucleotides Sp-Hox mut<br />

(in which the HoxC4-Oct–binding site was disrupted), Sp mut -Hox (in<br />

which the Sp–NF-kB–binding site was disrupted) and Sp mut -Hox mut<br />

(in which both sites were disrupted) did not efficiently achieve<br />

competition for the formation of complexes B and B¢, AandA¢, or<br />

all four complexes, respectively. We confirmed those results by EMSA<br />

with the mutated oligonucleotides as radiolabeled probes; Sp-Hox mut<br />

gave rise only to complexes B and B¢, Sp mut -Hox yielded only A and<br />

A¢, and Sp mut -Hox mut gave rise to none of the four complexes.<br />

Incubation of nuclear extracts of 4B6 or Ramos B cells with the Sp<br />

mRNA (normalized)<br />

50<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

NS NS NS NS NS NS NS<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

10 1<br />

10 1<br />

10 0<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

10 1<br />

10 1<br />

10 0<br />

10 0<br />

654321 0<br />

20<br />

16<br />

12<br />

8<br />

4<br />

CFSE<br />

0<br />

0 1 2 3 4<br />

Cell division<br />

5 6<br />

654321 0<br />

LPS + IL-4<br />

12<br />

Hoxc4<br />

10<br />

8<br />

6<br />

4<br />

2<br />

CFSE<br />

0<br />

0 1 2 3 4<br />

Cell division<br />

5 6<br />

CD154 + IL-4<br />

+/+<br />

Hoxc4 –/–<br />

10<br />

CD138-FITC<br />

4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 4<br />

10 3<br />

10 3<br />

10 2<br />

10 2<br />

10 1<br />

10 1<br />

10 0<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

10 4<br />

10 3<br />

10 2<br />

10 1<br />

10 0<br />

LPS LPS + IL-4 CD154 + IL-4<br />

Figure 4 Impaired CSR in Hoxc4 –/– B cells.<br />

(a) Immunoglobulin titers in supernatants of<br />

cultures of Hoxc4 +/+ and Hoxc4 –/– splenic<br />

B cells stimulated for 7 d with LPS alone<br />

(IgG2b and IgG3), or with LPS plus IL-4<br />

(IgG1 and IgE), IFN-g (IgG2a) or TGF-b1,<br />

IL-4, IL-5 and dextran-conjugated mAb to<br />

d-chain (IgA). P values, paired t-test. Data are<br />

from three experiments with four pairs of<br />

Hoxc4 +/+ and Hoxc4 –/– mice. (b) Proliferation<br />

of Hoxc4 +/+ and Hoxc4 –/– B cells labeled with<br />

CFSE and stimulated with LPS plus IL-4 (top) or with CD154 plus IL-4 (bottom) to induce switching to IgG1. P ¼ 0.0051, for LPS plus IL-4, or<br />

P ¼ 0.0069, for CD154 plus IL-4 (overall; paired t-test). Numbers above plots (left) indicate cell divisions. Data are from two independent experiments.<br />

(c) Plasma cell–differentiation of Hoxc4 +/+ and Hoxc4 –/– B cells stimulated for 4 d with LPS alone, LPS plus IL-4, or CD154 plus IL-4, then stained with<br />

phycoerythrin-labeled mAb to mouse B220 and FITC-labeled mAb to mouse CD138 and analyzed by flow cytometry. Numbers in outlined areas indicate<br />

percent B220loCD138 + (plasma) cells among total cells. Data are from one representative of three independent experiments.<br />

40<br />

30<br />

20<br />

10<br />

0<br />

IgG1-APC<br />

IgG1-APC<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

I µ -C µ<br />

Iγ3-Cγ3 Iγ1-Cγ1 Iγ2b-Cγ2b P = 0.00026<br />

Iγ2a-Cγ2a P = 0.000046<br />

Iε-Cε P = 0.00034<br />

Iα-Cα P = 0.0022<br />

I µ -Cγ3 P = 0.0019<br />

I µ -Cγ1 P = 0.0028<br />

I µ -Cγ2b I µ -Cγ2a I µ -Cε I µ -Cα Germline I H -C H transcripts Post-recombination I µ -C H transcripts<br />

Figure 5 HoxC4 deficiency does not alter germline IH-CH transcripts but<br />

results in lower expression of post-recombination Im-CH transcripts. Realtime<br />

quantitative RT-PCR analysis of germline IH-CH transcripts (left) and<br />

post-recombination I m-C H transcripts (right) in Hoxc4 +/+ and Hoxc4 /<br />

splenic B cells cultured for 3 d with LPS alone (I g2b-C g2b, I g3-C g3, I m-C g2b<br />

and I m-C g3) or with LPS plus IL-4 (I m-C m,I g1-C g1, I E-C E,I mC g1 and I m-C E),<br />

IFN-g (I g2a-C g2a and I m-C g2a), or TGF-b1, IL-4, IL-5 and dextran-conjugated<br />

mAb to m-chain (I a-C a and I m-C a); expression is normalized to Cd79b<br />

expression and is presented relative to the expression in Hoxc4 +/+ B cells,<br />

set as 1. P values, paired t-test. Data are representative of three experiments<br />

(mean and s.e.m. of triplicate samples.) with one representative of three<br />

pairs of Hoxc4 +/+ and Hoxc4 –/– mice (other pairs, Supplementary Fig. 3).<br />

544 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

IgG1 (%)<br />

IgG1 (%)


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b<br />

Hoxc4 +/+<br />

Frequency<br />

of mutations<br />

Hoxc4 –/–<br />

Frequency<br />

of mutations<br />

Pair 1<br />

>10<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

9 >10<br />

7<br />

8<br />

6<br />

4<br />

3<br />

2<br />

1<br />

0<br />

>10<br />

8<br />

5<br />

4<br />

1<br />

0<br />

3.53 × 10<br />

578<br />

4 ≥10<br />

3<br />

2<br />

4 68<br />

>10<br />

3<br />

2<br />

6<br />

2<br />

9<br />

1<br />

1<br />

1<br />

–3<br />

2.70 × 10 –3<br />

2.56 × 10 –3<br />

Pair 2 Pair 3<br />

105<br />

107 107<br />

108 106 110<br />

0<br />

1.35 × 10 –3<br />

1.15 × 10 –3<br />

probe gave rise to three main protein-DNA complexes, C, C¢ and C¢¢,<br />

specific for the binding of Sp–NF-kB. We further confirmed the<br />

binding specificity of HoxC4, Oct1 and Oct2 to the 5¢-ATTTGAAT-<br />

3¢ site, and the binding specificity of Sp1, Sp3 and NF-kB tothe5¢-<br />

GGGGAGGAGCC-3¢ site, by supershift or inhibition of the formation<br />

of the respective protein-DNA complexes by using a specific mAb to<br />

HoxC4 and specific antibody to Oct1 (anti-Oct1), anti-Oct2, anti-Sp1,<br />

anti-Sp3 or antibody to the p52 subunit of NF-kB. No supershift or<br />

inhibition of protein-DNA complex involving the Sp-Hox or the Sp<br />

probe was achieved with a Pax5-specific antibody. These experiments<br />

show that HoxC4, Oct1 and Oct2 bind specifically to the conserved<br />

5¢-ATTTGAAT-3¢ site in the AICDA promoter, whereas Sp1, Sp3 and<br />

NF-kB, but not Pax5, bind specifically to the conserved 5¢-GGGGAG<br />

GAGCC-3¢ site in the AICDA promoter.<br />

Cis elements critical for Aicda and AICDA promoter activation<br />

To determine the contribution of the conserved cis elements, including<br />

HoxC4-Oct– and Sp–NF-kB–binding sites to the activity of the Aicda<br />

0<br />

1.12 × 10 –3<br />

0<br />

V J558 DJ H -C µ transcripts (fold)<br />

1.6<br />

Hoxc4<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

NS<br />

0<br />

Pair 1<br />

+/+<br />

Hoxc4 –/–<br />

NS NS<br />

Pair 2 Pair 3<br />

Figure 6 Somatic mutation in the immunoglobulin heavy-chain intronic J H4-iEm DNA of Peyer’s patch<br />

PNA hi B220 + (germinal center) B cells. (a) Proportion of sequences carrying various numbers (along<br />

margins) of mutations over the 720-bp J H4-iE m DNA from Peyer’s patch PNA hi B220 + (germinal center)<br />

B cells from 12-week-old Hoxc4 –/– and Hoxc4 +/+ littermates (n ¼ 3 pairs of mice); below, frequency of<br />

somatic mutation; center, number of sequences analyzed (spectrum of mutation, Supplementary<br />

Fig. 4). Data are from three experiments. (b) Real-time quantitative RT-PCR analysis of V J558DJ H-C m<br />

transcripts in Peyer’s patch B cells from the mice in a, normalized to CD79b expression and presented<br />

relative to the expression in Hoxc4 +/+ B cells. Data are from three experiments (mean and s.e.m. of<br />

triplicate samples) with three pairs of Hoxc4 +/+ and Hoxc4 –/– mice.<br />

Figure 7 HoxC4 deficiency impairs AID expression, which depends on the<br />

conserved HoxC4-Oct–binding site in the Aicda promoter. (a) Quantitative<br />

RT-PCR analysis of Aicda expression in RNA (2 mg) from Hoxc4 +/+ and<br />

Hoxc4 –/– splenic B cells cultured for 0, 12, 24, 48 and 72 h in the<br />

presence of LPS plus IL-4 (left) or CD154 plus IL-4 (right), normalized<br />

to Cd79b expression and presented relative to expression in Hoxc4 +/+<br />

B cells at time 0. P values, paired t-test. Data are representative of three<br />

independent experiments (mean and s.e.m.). (b) Immunoblot analysis of AID<br />

and b-actin in the cells in a. Data are representative of three independent<br />

experiments. (c) Promoter regions of human AICDA and mouse Aicda (AIDpro),<br />

the flanking 5¢ region (5¢R), cr1 and cr2 in the first intron, and exons<br />

1–5 (coding regions, light blue). Numbers above indicate the number of<br />

nucleotides in each region. (d) Aicda promoter activity in mouse CH12F3-2A<br />

B cells (CSR inducible) and human Ramos B cells (spontaneous SHM)<br />

transfected with pGL3-Basic luciferase (luc) reporter vectors containing<br />

various combinations of elements of the Aicda locus (left) and then treated<br />

with LPS, IL-4 and TGF-b1 to induce CSR (CH12F3-2A) or given no further<br />

treatment (nil; Ramos), assessed as luciferase activity measured after or<br />

24 h (CH12F3-2A) or 16 h (Ramos) culture and presented relative to the<br />

luciferase activity of cells transfected with empty vector (pGL3). Data are<br />

representative of three independent experiments (mean and s.e.m.).<br />

promoter, we constructed luciferase reporter<br />

vectors containing the mouse Aicda promoter<br />

sequence (residues –349 to –1) with mutation<br />

or deletion of the HoxC4-Oct–binding site<br />

and/or the Sp–NF-kB–binding site. In addition<br />

to the conserved HoxC4-Oct–binding<br />

site with the sequence 5¢-ATTTGAAT-3¢, we<br />

identified a putative HoxC4-binding site with<br />

a sequence of 5¢-ATTT-3¢ in the mouse and<br />

rat promoter (residues –155 to –158 of<br />

mouse Aicda) but not in the human, chimp<br />

or dog promoter (Supplementary Fig. 5).<br />

Deletion of this site did not alter the Aicda<br />

promoter activity (Fig. 8a). In contrast, deletion<br />

of the Hoxc4-Oct motif resulted in<br />

promoter activity that was 71%, 64% and<br />

88% lower in mouse CH12F3-2A B cells<br />

induced with LPS, IL-4 and TGF-b, in<br />

human 4B6 cells, and in Ramos B cells,<br />

respectively. To determine the relative contribution<br />

of the binding of HoxC4 and Oct to<br />

the promoter activity of the HoxC4-Oct<br />

motif as a whole, we mutated 5¢-ATTT-<br />

GAAT-3¢ to 5¢-CTTTGAAT-3¢, thereby disrupting the binding of<br />

HoxC4 but retaining the binding of Oct 25 ,orto5¢-ATTTGCCG-3¢,<br />

thereby abrogating the binding of Oct1-Oct2 but not of HoxC4<br />

(mutated residues underlined) 25 . Mutation of the HoxC4 motif in<br />

the HoxC4-Oct site resulted in transcription that was 47%, 36% and<br />

55% lower, whereas mutation of the Oct site resulted in transcription<br />

that was 28%, 21% and 55% lower, in CH12F3-2A, 4B6 and Ramos B<br />

cells, respectively. Thus, both the HoxC4 and Oct motifs of the<br />

HoxC4-Oct–binding site contribute to the Aicda promoter activity,<br />

as further confirmed by the up-to-88% loss of Aicda promoter activity<br />

a<br />

b<br />

c<br />

16<br />

Hoxc4 +/+<br />

Hoxc4 +/+<br />

Hoxc4 –/–<br />

Hoxc4 –/–<br />

P = 0.00067<br />

12<br />

10<br />

P = 0.0017<br />

P = 0.0046<br />

12<br />

P = 0.019<br />

8<br />

8<br />

4<br />

6<br />

4<br />

2<br />

P = 0.0047<br />

0<br />

0<br />

0 12 24 48 72<br />

Time (h) Time (h)<br />

Hoxc4<br />

AID<br />

β-actin<br />

Time (h) 0 12 24 48 72 0 12 24 48 72 0 12 24 48 72 0 12 24 48 72<br />

LPS + IL-4<br />

CD154 + IL-4<br />

+/+<br />

Hoxc4 –/–<br />

0 12 24 48 72<br />

Aicda expression<br />

Human AICDA<br />

ARTICLES<br />

5′R AID-pro 84<br />

5,748 1471,380 270 294 115 470 2,104<br />

1 cr1 cr2<br />

2 3 4 5<br />

5′R AID-pro 84<br />

Mouse Aicda<br />

1 cr1<br />

–839 –349 –1<br />

5,369<br />

cr2<br />

147 1,423 270<br />

2 3<br />

530 115 379<br />

4<br />

1,759<br />

5<br />

d<br />

CH12F3-2A<br />

(LPS + IL-4 + TGF-β1)<br />

Ramos<br />

(nil)<br />

pGL3–<br />

luc<br />

pGL3–5′R–AID-pro–cr1–cr2<br />

5′R AID-pro<br />

luc cr1 cr2<br />

pGL3–5′R–AID-pro–cr1<br />

luc cr1<br />

pGL3–5′R–AID-pro–cr2<br />

luc cr2<br />

pGL3–5′R–AID-pro<br />

luc<br />

pGL3–5′R<br />

luc<br />

pGL3–AID-pro<br />

luc<br />

0 2 4 6 8 10 12 0 5 10 15 20<br />

Luciferase activity Luciferase activity<br />

(fold)<br />

(fold)<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 545


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

a<br />

b<br />

pGL3–AID-pro–enh<br />

–349<br />

HoxC4 Sp–NF-κB Hoxc4 Oct<br />

–349 –158<br />

Empty vector<br />

–155 –57 –45 –29 –22 –1<br />

Human<br />

Mouse<br />

WT<br />

Mut0<br />

Mut1<br />

Mut2<br />

Mut3<br />

Mut4<br />

Mut5<br />

Mut6<br />

Mut7<br />

Aicda<br />

promoter<br />

when the entire Hoxc4-Oct–binding site was deleted. Further, mutation<br />

of the conserved Sp–NF-kB–binding site to 5¢-AAAAAGGAAA-3¢<br />

resulted in promoter activity that was 73%, 80% and 63% lower in<br />

CH12F3-2A, 4B6 and Ramos B cells, respectively. In agreement with<br />

that result, deletion of this site resulted in promoter activity that<br />

was 85%, 68% and 82% lower in CH12F3-2A, 4B6 and Ramos<br />

B cells, respectively. Finally, deletion of the HoxC4-Oct–binding site<br />

combined with mutation or deletion of the Sp–NF-kB–binding site<br />

abrogated Aicda promoter activity. These experiments show that the<br />

conserved HoxC4-Oct–binding site is important to Aicda promoter<br />

activity and, together with the conserved Sp–NF-kB–binding site, is<br />

indispensable for full transcriptional activation of Aicda.<br />

To confirm the relevance of our EMSA and luciferase reporter<br />

experiments, we precipitated chromatin from human 4B6 and<br />

–1<br />

luc<br />

4B6<br />

Ramos<br />

2E2 (nil)<br />

2E2 (anti-huCD40 + IL-4)<br />

CH12F3-2A (nil)<br />

CH12F3-2A (LPS + IL-4 + TGF-β1)<br />

Spleen (nil)<br />

Spleen (LPS + IL-4)<br />

Spleen (CD154 + IL-4)<br />

SV40 enhancer<br />

CH12F3-2A 4B6<br />

(LPS + IL-4 + TGF-β1)<br />

(nil)<br />

0<br />

100<br />

200<br />

300<br />

400<br />

500<br />

Luciferase activity<br />

(relative)<br />

0<br />

100<br />

200<br />

c<br />

400<br />

500<br />

Luciferase activity<br />

(relative)<br />

Ramos<br />

(nil)<br />

Luciferase activity<br />

(relative)<br />

Input<br />

Mo IgG<br />

Rab IgG<br />

Oct1<br />

Oct2<br />

HOXC4 or Hoxc4 AICDA or Aicda GAPDH or Gapdh AICDA or Aicda promoter<br />

Figure 8 The conserved HoxC4-Oct– and Sp–NF-kB–binding sites are essential for full Aicda promoter activity, and HoxC4, Oct1, Oct2, Oca-B, Pax5, Sp1,<br />

Sp3 and NF-kB (p52) are recruited to the Aicda promoter in B cells expressing AICDA or Aicda and undergoing CSR or SHM. (a) Luciferase activity of<br />

mouse CH12F3-2A B cells (CSR inducible), human 4B6 B cells (spontaneous CSR) and human Ramos B cells (spontaneous SHM) transfected with pGL3-<br />

Enhancer luciferase reporter constructs containing wild-type Aicda promoter (WT) or mutated Aicda promoter (Mut1–Mut7) in which the HoxC4-Oct– and/or<br />

Sp–NF-kB–binding sites were deleted or disrupted by site-directed mutagenesis (left; lower case indicates mutated nucleotides), and then treated with LPS,<br />

IL-4 and TGF-b1 to induce CSR (CH12F3-2A) or given no further treatment (nil; 4B6 and Ramos), assessed after 24 h (CH12F3-2A) or 16 h (Ramos and<br />

4B6) of culture and presented relative to the luciferase activity of cells transfected with empty vector. Data are representative of three independent<br />

experiments (mean and s.e.m.). (b) Semiquantitative RT-PCR of the expression of the genes encoding HoxC4 (left), AID (middle) and GAPDH (loading<br />

control; right) in human 4B6 and Ramos B cells; human 2E2 B cells left unstimulated (nil) or stimulated with mAb to human CD40 (anti-huDC40) plus<br />

IL-4; mouse CH12F2-2A B cells left unstimulated or stimulated with LPS, IL-4 and TGF-b1; and C57BL/6 wild-type splenic B cells left unstimulated or<br />

stimulated with LPS plus IL-4, or CD154 plus IL-4. Wedges (top) indicate serial twofold dilution of template cDNA. Results are representative of three<br />

independent experiments. (c) PCR of crosslinked chromatin precipitated from the cells in b with preimmune control mouse (Mo) or rabbit (Rab) IgG, mouse<br />

mAb to HoxC4, or rabbit anti-Oct1, anti-Oct2, anti-Oca-B, anti-Pax5, anti-Sp1, anti-Sp3 or anti–NF-kB p52, amplified with primers for the AICDA or Aicda<br />

promoter. Results are representative of three independent experiments.<br />

Ramos B cells with anti-HoxC4, anti-Oct1, anti-Oct2, anti-Oca-B,<br />

anti-Sp1, anti-Sp3 or anti–NF-kB p52. In DNA precipitated by<br />

each antibody, we readily identified the AICDA promoter sequence<br />

(Fig. 8b,c). The specificity of those findings was further proven by<br />

our ability to readily detect AICDA or Aicda promoter DNA in<br />

chromatin-immunoprecipitation assays involving human 2E2 B cells,<br />

which can be induced to express AID and undergo CSR after treatment<br />

with mAb to CD40 and cytokines (such as IL-4), mouse CH12F3-2A B<br />

cells, which can be induced to express AID and undergo CSR after<br />

treatment with LPS, IL-4 and TGF-b1, as well as spleen B cells from<br />

wild-type C57BL/6 mice activated by LPS plus IL-4 or by CD154 plus<br />

IL-4. Induction of CSR in 2E2 or CH12F3-2A B cells or primary mouse<br />

spleen B cells by these stimuli resulted in substantial Hoxc4 expression<br />

and recruitment of HoxC4, Oct1, Oct2, Oca-B, Sp1, Sp3 and NF-kB to<br />

546 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY<br />

300<br />

0<br />

40<br />

80<br />

Oca-B<br />

120<br />

IgG<br />

HoxC4<br />

Pax5<br />

Sp1<br />

Sp3<br />

NF-κB


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

a b<br />

TAC-Aicda TAC<br />

B220-PE<br />

Aicda expression (relative) Aicda expression (relative)<br />

10 0<br />

10 1<br />

10 2<br />

10 3<br />

10 4<br />

Hoxc4<br />

72.3<br />

+/+<br />

Hoxc4 –/–<br />

22.0 76.9 9.7<br />

5.6 0.05 12.3 0.07<br />

59.1 33.2 58.4 34.6<br />

10<br />

7.5 0.2 0.2<br />

0<br />

10 1<br />

10 2<br />

10 3<br />

10 4<br />

6.7<br />

10 0<br />

10 1<br />

10 2<br />

IgG1-FITC<br />

10 3<br />

10 4<br />

10 0<br />

10 1<br />

10 2<br />

10 3<br />

10 4<br />

Real-time<br />

P = 0.00041<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

TAC<br />

TAC-Aicda<br />

TAC<br />

TAC-Aicda<br />

Germline Iµ-Cµ expression<br />

(relative)<br />

Germline Iγ1-Cγ1 expression<br />

(relative)<br />

Real-time<br />

NS<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

TAC<br />

TAC-Aicda<br />

TAC<br />

TAC-Aicda<br />

Semiquantitative<br />

P = 0.000047<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Circle Iγ1-Cµ expression<br />

(relative)<br />

Post-rec Iµ-Cγ1 expression<br />

(relative)<br />

TAC<br />

TAC-Aicda<br />

Semiquantitative Real-time Real-time<br />

P = 0.0066<br />

1.2<br />

1.2<br />

NS<br />

P = 0.00021<br />

1.2<br />

1.0<br />

1.0<br />

1.0<br />

0.8<br />

0.8<br />

0.8<br />

0.6<br />

0.6<br />

0.6<br />

0.4<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

0.2<br />

0<br />

0<br />

0<br />

TAC<br />

TAC-Aicda<br />

the AICDA and Aicda promoters. Because it has been suggested that<br />

Pax5 is (indirectly) recruited to the Aicda promoter 16 ,weprecipitated<br />

chromatin in these human and mouse B cells with anti-Pax5; in these<br />

immunoprecipitated DNA complexes, we readily detected AICDA and<br />

Aicda promoter sequences, respectively. These findings show that<br />

HoxC4, Oct1, Oct2, Oca-B, Sp1, Sp3 and NF-kB are recruited to the<br />

AICDA and Aicda promoters in human and mouse B cells, respectively,<br />

that express AID and undergo CSR or SHM.<br />

Enforced AID expression restores CSR in Hoxc4 –/– B cells<br />

We then sought to demonstrate that the defective CSR in HoxC4deficient<br />

B cells was actually due to impairment of AID expression and<br />

not some other HoxC4-dependent activity. For this, we enforced<br />

expression of AID in Hoxc4 –/– B cells to restore CSR. We used a<br />

control retroviral vector containing human IL2RA, which encodes the<br />

human IL-2 receptor (CD25; TAC), and an AID-expression retroviral<br />

construct containing human IL2RA and Aicda (TAC-Aicda) 15 (Supplementary<br />

Fig. 7 online). We transduced LPS-activated spleen<br />

Hoxc4 +/+ and Hoxc4 –/– B cells with TAC and TAC-Aicda and stimulated<br />

them with LPS and IL-4 for 72 h and 96 h before analyzing CSR.<br />

Consistent with the results we obtained with untransduced Hoxc4 +/+<br />

and Hoxc4 –/– B cells, CSR was much lower in Hoxc4 –/– B cells<br />

transduced with the TAC control retrovirus than in their Hoxc4 +/+<br />

counterparts (Fig. 9a,b). In Hoxc4 +/+ B cells, enforced expression of<br />

AID increased CSR to IgG1 by about 50%. Transduction of Hoxc4 –/–<br />

B cells with TAC-Aicda retrovirus increased Aicda expression. It did<br />

not modulate the expression of germline I m-C m and I g1-C g1 transcripts<br />

but did increase CSR to IgG1 to an extent similar to that of Hoxc4 +/+<br />

B cells transduced with TAC-Aicda retrovirus, as shown by the greater<br />

proportion of B cells positive for surface IgG1 and more circle I g1-C m<br />

and post-recombination I m-C g1 transcripts. These experiments show<br />

that the defective AID expression and CSR of Hoxc4 –/– B cells are<br />

restored by enforced AID expression, which further indicates that<br />

HoxC4 modulates CSR by regulating AID expression.<br />

DISCUSSION<br />

In B cells, AID expression is tightly regulated 31–34 , possibly in an<br />

activation-dependent way in conditions in which CSR and SHM<br />

unfold. The specificity and amount of AID expression are probably<br />

controlled by a complex combination of various tissue-specific transcription<br />

factors, both activators and repressors. Here we have<br />

provided evidence that by binding to the highly conserved 5¢-ATTT<br />

GAAT-3¢ motif in the Aicda promoter, HoxC4 activates this promoter,<br />

thereby modulating AID expression, CSR and SHM. In this function,<br />

HoxC4 acts in synergy with Oct1-Oct2, which also bind to the<br />

ARTICLES<br />

Figure 9 Enforced expression of AID restores CSR in Hoxc4 –/– Bcells.<br />

(a) Flow cytometry of the surface expression of B220 and IgG1 in Hoxc4 +/+<br />

and Hoxc4 –/– B cells activated with LPS and transduced with TAC or<br />

TAC-Aicda retrovirus, then cultured for 3 or 4 d with LPS plus IL-4.<br />

Numbers in quadrants indicate percent cells in each. Data are from one<br />

representative of five independent experiments. (b) Real-time or<br />

semiquantitative RT-PCR analysis (above graphs) of Aicda expression,<br />

germline Im-Cm and Ig1-Cg1 transcripts, circle Ig1-Cm transcripts and postrecombination<br />

(Post-rec) I m-C g1 transcripts in the cells in a, normalized to<br />

CD79b transcripts and presented as the ratio of the expression in Hoxc4 –/–<br />

cells to the expression in Hoxc4 +/+ cells. P values, t-test. Data are<br />

representative of three independent experiments (mean and s.e.m.).<br />

5¢-ATTTGAAT-3¢ motif, by a mechanism similar to that reported<br />

for HoxC4-Oct–mediated activation of the human 3¢ E a hs1,2<br />

enhancer element 25 . In agreement with that, Hoxc4 –/– mice or<br />

B cells show defective CSR and SHM despite normal expression of<br />

mature V HDJ H and germline I H-C H transcripts 31 . The full restoration<br />

of CSR by enforced expression of AID in Hoxc4 –/– B cells indicates<br />

that induction of AID is the main pathway through which HoxC4<br />

regulates CSR and, probably, SHM. Mutation of the 5¢-ATTTGAAT-3¢<br />

site to 5¢-GCTTGAATT-3¢, which does not disrupt the binding<br />

of Oct and introduces a putative Sp-binding site, does not seem to<br />

alter Aicda or AICDA promoter activity in mouse B lymphoma<br />

M12 and CH33 cells or human embryonic kidney 293 cells, respectively<br />

16 . In our experiments, mutation of 5¢-ATTTGAAT-3¢ to disrupt<br />

HoxC4-binding activity while preserving the Oct-binding activity<br />

or to abrogate the binding of Oct1-Oct2 but not HoxC4 binding 25<br />

resulted in less Aicda transcription, thereby emphasizing the function<br />

of these homeodomain transcription factors in AID expression<br />

and adding another dimension to the function of Oct1-Oct2 in<br />

B cell differentiation.<br />

As we have shown in human B cells, HoxC4 serves an important<br />

and complex function in the regulation of the IgH locus 24,25,35 .The<br />

putative dampening effect of HoxC4 on the baseline activity of the<br />

I g and I e promoters would be effectively lifted and overridden by<br />

the strong activation of these promoters by CD40 signaling and<br />

CD154-induced HoxC4-mediated activation of the hs1,2 enhancer<br />

element 24,25,35 . The idea that HoxC4 is involved in the induction of<br />

AID expression is further strengthened by the demonstration that<br />

like AID, HoxC4 is expressed mainly in germinal center B cells of<br />

both humans 24 and mice (as reported here), and that engagement<br />

of CD40 by CD154 and treatment with cytokines, which induce the<br />

expression of AICDA and Aicda, also induces HOXC4 and Hoxc4 in<br />

human B cells 24 and mouse B cells (data presented here), respectively.<br />

Furthermore, NF-kB-binding sites in both the human and<br />

mouse HOXC4 and Hoxc4 promoters underpin the upregulation of<br />

HoxC4 by CD40 signaling (unpublished data).<br />

Consistent with the idea that the defect in CSR and SHM<br />

manifested by Hoxc4 –/– B cells was due to a failure to induce<br />

Aicda expression, both Aicda transcripts and AID protein, as<br />

induced by LPS and IL-4 or CD154 and IL-4, were much less<br />

abundant in Hoxc4 –/– B cells. The tissue and differentiation-stage<br />

specificity of HoxC4 expression would account to a great extent for<br />

the precise regulation of AID expression. In Hoxc4 –/– mice, the<br />

lower AID expression was reflected in vivo in the impairment of the<br />

maturation of the T cell–dependent antibody response. In these<br />

mice, although Aicda expression was much lower, germinal center<br />

formation seemed normal. That phenotype is reminiscent of that of<br />

Aicda +/– mice 36 , which also have normal germinal centers in the<br />

presence of much lower Aicda expression 37,38 , and contrasts with<br />

that of Aicda / mice, in which activated B cells accumulate and<br />

form giant germinal centers 36 .<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 547


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

Pax5 has been suggested to serve a function in Aicda expression by<br />

binding to the 5¢-GGGGAGGAGCC-3¢ site in the Aicda promoter 16 .<br />

This cis element, however, does not fulfill the requirements of a<br />

consensus Pax5-binding motif and did not bind Pax5 in our experiments<br />

and those by others 17 . Our finding of a lack of specificity of the<br />

conserved 5¢-GGGGAGGAGCC-3¢ cis element for Pax5 suggests<br />

that recruitment of Pax5 to the Aicda promoter, as showed by<br />

chromatin-immunoprecipitation assay, occurred indirectly through<br />

other DNA-binding transcription factors, perhaps Sp1 or Sp3, by a<br />

mechanism similar to the interaction between the estrogen receptor<br />

and Sp1 on certain estrogen-responsive promoters 39 .<br />

The B cell development–related transcription factor E47 has been<br />

suggested to contribute to the enhancement of Aicda promoter activity<br />

by binding to E-boxes in cr2 of the first intron of Aicda 21,22 .E47induced<br />

enhancement of Aicda promoter activity would be modulated<br />

by the E2A inhibitor Id3 (ref. 15). The presence of cr2 in human<br />

Ramos B cells and mouse CH12F3-2A B cells (as reported here) and in<br />

mouse BaF/3 pro–B cells and M12 B cells 16 did not enhance transcription<br />

of the luciferase reporter driven by the Aicda promoter. This<br />

might reflect the dispensability of E2A transcription factors in the<br />

expression of AICDA and Aicda, CSR and, possibly, SHM, as shown by<br />

analysis of E2A-deficient B cells 40,41 or the muted baseline activity of<br />

cr2 that resulted from ‘preferential’ binding of Id2 and/or Id3 to this<br />

region. These experiments, however, cannot rule out the possibility<br />

that like the IgH and immunoglobulin-k intronic enhancers, which<br />

contain many E-box sites and show a high enhancer activity in<br />

germinal center B cells 42 , cr2 and E proteins act together with<br />

HoxC4 to synergistically induce AID expression.<br />

The HoxC4-mediated activation of the Aicda promoter was<br />

further enhanced by the upstream conserved 5¢-GGGGAG<br />

GAGCC-3¢ site, which, as we have also shown, recruited Sp1, Sp3<br />

and NF-kB. Sp1 and Sp3 bind directly to DNA through their zincfinger<br />

motifs and enhance gene transcription. These proteins are<br />

ubiquitously expressed and are directly involved not only in the<br />

regulation of basal transcription and expression of ‘housekeeping’<br />

genes but also in the expression of developmentally controlled<br />

genes. In our experiments, the Sp–NF-kB–binding site was able to<br />

partially mediate Aicda promoter activity in the absence of the<br />

HoxC4-Oct–binding site and possibly accounted for the residual<br />

AID expression, CSR and SHM in Hoxc4 –/– B cells and mice.<br />

We have shown that Oca-B was also recruited to the AICDA and<br />

Aicda promoters in B cells undergoing CSR or SHM, probably<br />

through interaction with Oct1 and/or Oct2 (ref. 43). By clamping<br />

together the POU H and POU S subdomains of Oct1 and Oct2, Oca-B<br />

would increase the affinity of these homeodomain proteins for DNA,<br />

thereby potentiating HoxC4- and Oct1-Oct2-mediated activation of<br />

the Aicda promoter. Oca-B is important in HoxC4- and Oct1-Oct2mediated<br />

activation of the human 3¢ Ea enhancer hs1,2 element 25 .In<br />

agreement with that, mice that lack Oca-B have impaired CSR 43 .Our<br />

results here have identified another function for Oct1-Oct2 and Oca-B<br />

activity: regulating AID expression. Overall, our findings offer fundamental<br />

insights into the mechanisms of activation of the AICDA and<br />

Aicda promoters and induction of AID expression, CSR and SHM.<br />

The possibility that the induction of HoxC4 by stimuli other than<br />

CD40 signaling, LPS and cytokines, such as hormones (unpublished<br />

data), modulates AID expression and, therefore, antibody diversification<br />

in health and disease should be addressed.<br />

METHODS<br />

Hoxc4 –/– mice. In the Hoxc4 –/– mice used here (A.M. Boulet and<br />

M.R. Capecchi), Hoxc4 was disrupted through insertion of a loxP siteinexon<br />

2ofHoxc4 at the sequence encoding the amino-terminal end (between the<br />

third and the fourth codons) of the homeobox, which introduces a stop codon<br />

at the insertion site and yields a nonfunctional truncated protein (lacking 95%<br />

of the homeodomain; unpublished observations). We obtained Hoxc4 +/– frozen<br />

sperm and rederived Hoxc4 +/– mice by in vitro fertilization through the services<br />

of the transgenic mouse facility of the University of California at Irvine. These<br />

Hoxc4 +/– mice were on a C57BL/6 background after backcrossing of the<br />

129Sv/Ev founder strain with C57BL/6 mice. Hoxc4 –/– mice and their wildtype<br />

littermates were bred in pathogen-free conditions. The Institutional<br />

Animal Care and Use Committee of University of California at Irvine approved<br />

all animal experiments.<br />

B cells and T cells. The number of B cells (B220 + ) and T cells (CD3 + ), the<br />

proportion CD4 + T cells and dead B cells and T cells, and the proportion of<br />

PNA hi B cells, plasma cells (B220 lo CD138 + ) and NP-binding CD38 hi IgG1 +<br />

memory B cells was assessed by flow cytometry with a FACSCalibur (BD<br />

Biosciences). Single-cell suspensions were prepared from spleens or Peyer’s<br />

patches of Hoxc4 –/– and Hoxc4 +/+ mice and were stained with phycoerythrinlabeled<br />

mAb to mouse B220 (RA3-6B2; BD Biosciences), fluorescein isothiocyanate<br />

(FTIC)-labeled mAb to mouse CD3 (17A2; BioLegend), peridinine<br />

chlorophyll protein complex–labeled mAb to mouse CD4 (GK1.5; BioLegend),<br />

7-amino-actinomycin D (BD Biosciences), FITC-labeled peanut agglutinin<br />

(PNA), FITC-labeled mAb to mouse CD138 (281-2) or allophycocyaninlabeled<br />

anti–mouse IgG1 (X56; BD Biosciences), phycoerythrin-labeled NP<br />

(Biosearch Technologies) and phycoerythrin-indotricarbocyanine–labeled mAb<br />

to mouse CD38 (90; eBiosciences). Single-cell suspensions of B220 + cells were<br />

prepared from spleens or Peyer’s patches with the EasySep Mouse B Cell<br />

Enrichment kit (StemCell Technologies). For the preparation of PNA hi (germinal<br />

center) B cells, spleen or Peyer’s patch B cells were stained with<br />

phycoerythrin-labeled mAb to mouse CD45R and FITC-labeled PNA. Labeled<br />

lymphocytes were then sorted with a MoFlo cell sorter (Dako), which yielded<br />

PNA hi B220 + cells that were 95% pure.<br />

B cell lines. Ramos B cells were monitored for spontaneous SHM 44 .Monoclonal<br />

4B6 and 2E2 B cell lines were derived from the CSR- and SHM-inducible<br />

human monoclonal IgM + IgD + CL-01 B cell line 45–50 by sequential subcloning<br />

and selection for spontaneous and inducible CSR, respectively: 4B6 B cells are<br />

IgM + IgD + with an ‘early’ germinal center phenotype and undergo spontaneous<br />

CSR to IgG, IgE and IgA 24 ; 2E2 B cells are IgM + IgD + and undergo CSR to IgG,<br />

IgE and IgA after stimulation with an agonistic mAb to human CD40 and<br />

appropriate cytokines 51 . The mouse B lymphoma cell line CH12F3-2A was<br />

obtained from T. Honjo. CH12F3-2A cells have surface expression of IgM and<br />

switch to IgA after stimulation with CD154 or LPS in the presence of IL-4 and<br />

TGF-b1 (ref. 52). All these monoclonal B cells were cultured in FBS-RPMI<br />

(RPMI-1640 medium (Invitrogen) supplemented with 10% (vol/vol) heatinactivated<br />

FBS (Hyclone), 2 mM L-glutamine and 1 antibiotic-antimycotic<br />

mixture (penicillin (100 units/ml), streptomycin (100 mg/ml) and amphotericin<br />

B fungizone (0.25 mg/ml); Invitrogen)).<br />

B cell cycle and proliferation. Cell cycle was analyzed by staining with<br />

propidium iodide 50 . Proliferation was analyzed with the CellTrace CFSE Cell<br />

Proliferation kit (Molecular Probes). Cells were washed in serum-free Hank’s<br />

balanced-salt solution (Invitrogen) and were resuspended at a density of 1<br />

10 6 cells per ml. After the addition of an equal volume of 2.4 mM CFSE<br />

(carboxyfluorescein succinimidyl ester), cells were incubated for 12 min at<br />

37 1C and then were washed in FBS-RPMI. Cells were then diluted and were<br />

cultured in the presence or absence of LPS (20 mg/ml) from Escherichia coli<br />

(serotype 055:B5; Sigma-Aldrich) and recombinant mouse IL-4 (5 ng/ml; R&D<br />

Systems), then were collected at various times after activation and analyzed by<br />

flow cytometry. For in vivo B cell proliferation, mice were immunized<br />

with NP 16-CGG. After 10 d, they were injected intraperitoneally twice within<br />

16 h with 1 mg BrdU (5-bromodeoxyuridine) and were killed 4 h after the last<br />

injection. Cells from the spleen or Peyer’s patches were stained with<br />

phycoerythrin-labeled mAb to mouse B220 (BD Biosciences) or that mAb<br />

together with FITC-labeled PNA. Incorporated BrdU was stained with<br />

allophycocyanin-labeled mAb to BrdU with the APC BrdU Flow kit (BD<br />

Biosciences) and were analyzed by flow cytometry.<br />

548 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Analysis of in vitro CSR. Enriched spleen B cells were cultured at a density of<br />

1 10 6 cell/ml in FBS-RPMI with 0.05 mM b-mercaptoethanol. Cells were<br />

stimulated with LPS (20 mg/ml) or CD154-expressing membrane fragments of<br />

baculovirus-infected Sf21 insect cells (called ‘CD154’ here) 53 and the following<br />

reagents: nothing more for CSR to IgG3 and IgG2b; recombinant mouse IL-4<br />

(5 ng/ml) for CSR to IgG1 and IgE; IFN-g (50 ng/ml; PeproTech) for CSR to<br />

IgG2a; or TGF-b1 (1 ng/ml; R&D Systems), recombinant mouse IL-5 (5 ng/ml;<br />

R&D Systems) and dextran-conjugated mAb to d-chain (3 ng/ml; provided by<br />

C.M. Snapper) for CSR to IgA. Cells were collected on day 4 for analysis of<br />

surface immunoglobulins after being stained with FITC-labeled rat mAb to<br />

mouse IgG1 (A85-1), mouse IgG2a (R19-15), mouse IgG2b (R12-3), mouse<br />

IgG3 (R40-82) or mouse IgA (C10-3), or phycoerythrin-labeled rat mAb to<br />

mouse CD45R (B220; RA3-6B2; all from BD Biosciences). Cells were fixed with<br />

1% (vol/vol) paraformaldehyde in PBS and were analyzed by flow cytometry.<br />

Specific enzyme-linked immunosorbent assays involving 96-well plates coated<br />

with polyclonal goat antibody F(ab¢) 2 to the appropriate mouse isotype<br />

(Southern Biotechnology Associates) were used to measure IgG1, IgG2a, IgG2b,<br />

IgG3, IgA and IgE in culture supernatants of in vitro–stimulated spleen<br />

Hoxc4 +/+ and Hoxc4 –/– B cells. Supernatants were serially diluted twofold from<br />

1:5 to 1:640 and then were added to the plates (100 ml per well), which were<br />

incubated for 1 h at 25 1C. After plates were washed, biotin-labeled isotypespecific<br />

mAbs were added, followed by visualization with horseradish<br />

peroxidase–streptavidin (Supplementary Methods online). Concentrations of<br />

immunoglobulin isotypes were determined by interpolation with a calibrated<br />

standard curve for each isotype. Assays were done in triplicate.<br />

Quantitative real-time RT-PCR and semiquantitative RT-PCR. RNA was<br />

extracted with the RNeasy Mini Kit according to the manufacturer’s protocol<br />

(Qiagen). Residual DNA was removed by treatment with DNase I (Invitrogen).<br />

First-strand cDNA was synthesized from total RNA (2 mg for each experiment)<br />

with the SuperScript Preamplification system and oligo(dT) primer (Invitrogen).<br />

The expression of germline IH-CH, post-recombination Im-CH, mature<br />

VJ558DJH-Cm, Hoxc4 and Aicda transcripts was quantified by real-time quantitative<br />

RT-PCR 54 with the appropriate primers (Supplementary Table 1 online;<br />

Operon). In some cases, circle I g1-C m transcripts, HOXC4, Hoxc4, AICDA and<br />

Aicda transcripts were analyzed by specific semiquantitative RT-PCR with serial<br />

twofold dilutions so there was a nearly linear relationship between the amount<br />

of cDNA used and the intensity of the PCR product. A DNA Engine Opticon 2<br />

Real-Time PCR Detection system (Bio-Rad Laboratories) was used for realtime<br />

quantitative RT-PCR to measure the incorporation of SYBR Green<br />

(DyNAmo HS kit; New England Biolabs) according to the following protocol:<br />

50 1C for 2 min and 95 1C for 10 min; then 40 cycles of 95 1C for 10 s, 60 1Cfor<br />

20 s, 72 1C for 30 s and 80 1C for 1 s; data acquisition at 80 1C; and then 72 1C<br />

for 10 min. Melting-curve analysis was done from 72 1C to951C and samples<br />

were incubated for another 5 min at 72 1C. The change in cycling threshold<br />

(DDC T) method was used for data analysis.<br />

Analysis of somatic mutations in intronic JH4-iEl DNA. Peyer’s patch B cells<br />

were obtained from unimmunized 12-week-old littermate Hoxc4 –/– and<br />

Hoxc4 +/+ C57BL/6 mice and were used for analysis of somatic mutations in<br />

intronic DNA downstream of rearranged VJ558DJH4 genes. Platinum Pfx DNA<br />

polymerase (Invitrogen) was used for amplification of genomic DNA. The<br />

intronic IgH region downstream of rearranged V J558DJ H was amplified by<br />

nested PCR 55 involving two V H J558 framework region 3–specific forward<br />

primers and two reverse primers specific for sequences downstream of JH4<br />

(Supplementary Table 1), which yielded DNA of approximately 900 bp if<br />

aJ H4 rearrangement occurred. PCR conditions were 35 cycles of 94 1C for45s,<br />

58 1C for 45 s and 68 1C for 1 min. PCR products were cloned into the<br />

pCR-Blunt II-TOPO vector (Invitrogen) and sequenced. Only sequences from<br />

rearrangements involving JH4-iEm were analyzed. Sequences were analyzed with<br />

MacVector 7.2.3 software.<br />

Identification of putative transcription factor–binding sites. Putative transcription<br />

factor–binding sites in Aicda promoter sequence were identified<br />

by weight-matrix search with Match software (BIOBASE) which integrates<br />

the TRANSFAC 6.0 database and uses its positional weight matrices<br />

for analysis. Scores indicate the degree of fitness of the putative binding<br />

ARTICLES<br />

site with the consensus sequence according to the following parameters:<br />

a score 1.0 is 100%; the cut-off score is 0.75.<br />

Aicda promoter–luciferase reporter assays. Reporter constructs consisted of<br />

the pGL3-Basic (Fig. 7d) or pGL3-Enhancer (Fig. 8d) firefly luciferase reporter<br />

vector (Promega) and various sequences from the Aicda locus. Sequences<br />

consisting of 839 bp of flanking 5¢ region plus the Aicda promoter, 490 bp of<br />

flanking 5¢ region, or 349 bp of the Aicda promoter, as well as cr1 and cr2 of the<br />

first intron, were amplified by PCR from C57BL/6 mouse genomic DNA and<br />

were inserted upstream and/or downstream of the luciferase gene in pGL3<br />

vector. Various mutant reporters were constructed with the QuikChange Site-<br />

Directed Mutagenesis kit (Stratagene). Sequences of constructs were confirmed<br />

by at least two sequencing reactions. The reporter construct and the constitutively<br />

active Renilla reniformis luciferase–producing vector pRL-TK (Promega)<br />

were transfected together into human Ramos and 4B6 B cells and mouse<br />

CH12F3-2A B cells by electroporation24,25,35 . Firefly and renilla luciferase<br />

activity was measured with the Dual-Luciferase Reporter Assay system according<br />

to the manufacturer’s instructions (Promega).<br />

Chromatin immunoprecipitation. B cells (5 10 7 )weretreatedfor10minat<br />

25 1C with 1% (vol/vol) formaldehyde for crosslinkage of chromatin. After cells<br />

were washed with cold PBS containing protease inhibitors (Roche), chromatin<br />

was separated with nuclear lysis buffer (10 mM Tris-HCl, 1 mM EDTA, 0.5 M<br />

NaCl, 1% (vol/vol) Triton X-100, 0.5% (wt/vol) sodium deoxycholate and 0.5%<br />

(wt/vol) sarcosyl, pH 8.0) and was resuspended in IP-1 buffer (20 mM Tris-<br />

HCl, pH 8.0, 200 mM NaCl, 2 mM EDTA, 0.1% (wt/vol) sodium deoxycholate,<br />

0.1% (wt/vol) SDS and protease inhibitors). Chromatin was sonicated to yield<br />

DNA fragments approximately 200–1,000 bp in size, was precleared with<br />

agarose beads bearing protein G (Santa Cruz Biotechnology) and then was<br />

incubated overnight at 4 1C with mAb to HoxC4 (1E9; Novus Biologicals) or<br />

rabbit polyclonal anti-Oct1 (C-21), anti-Oct2 (C-20), Oca-B (C-20), anti-Pax5<br />

(N-19), anti-Sp1 (H-225), anti-Sp3 (D-20) or anti–NF-kB p52 (K-27; all from<br />

Santa Cruz Biotechnology). After that incubation, immune complexes were<br />

isolated with agarose beads bearing protein G, were eluted with elution buffer<br />

(50 mM Tris-HCl, 0.5% (vol/vol SDS, 200 mM NaCl and 100 mg/ml of<br />

proteinase K, pH 8.0) and then were incubated overnight at 65 1C forreversal<br />

of the formaldehyde crosslinks. DNA was extracted with phenol-chloroform<br />

and precipitated with ethanol and then was resuspended in 10 mM Tris-HCl,<br />

pH 8.0, and 1 mM EDTA. DNA sequences were confirmed by PCR with the<br />

appropriate primers (Supplementary Table 1).<br />

Retroviral transduction of B cells. The TAC and TAC-Aicda retroviral<br />

constructs 15 were obtained from C. Murre. For the generation of retrovirus,<br />

pCSretTAC-based constructs were transfected into the HEK293T packaging cell<br />

line with the ProFection Mammalian Transfection system (Promega). The<br />

retroviral constructs were used to transduce mouse spleen B cells as reported 15 .<br />

Additional methods. Information on immunoblot analysis, immunization<br />

with NP16-CGG and titration of NP-binding IgM and IgG1, histology, and<br />

EMSA and EMSA supershift-blocking assays is available in the Supplementary<br />

Methods online.<br />

Statistical analyses. Differences in the frequency and spectrum of mutations in<br />

Hoxc4 –/– and Hoxc4 +/+ mice were analyzed with the chi-squared test. Differences<br />

in immunoglobulin titers, CSR and mRNA expression were analyzed with<br />

paired t-tests.<br />

Accession codes. UCSD-<strong>Nature</strong> Signaling Gateway (http://www.signalinggateway.org):<br />

A000804, A001704, A001696, A000536, A001262, A001238 and<br />

A002271.<br />

Note: Supplementary information is available on the <strong>Nature</strong> <strong>Immunology</strong> website.<br />

ACKNOWLEDGMENTS<br />

We thank M.R. Capecchi and A.M. Boulet (University of Utah) for Hoxc4 +/–<br />

frozen mouse sperm; T. Fielder for technical efforts; L. Khidr, L. Phan,<br />

B. Gupta, J. Feng and Y. Zhong for handling Hoxc4 +/– mice; C. Murre<br />

(University of California, San Diego) for the Aicda retroviral construct,<br />

T. Honjo (Kyoto University) for CH12F3-2A cells; C.M. Snapper (Uniformed<br />

Services University of the Health Sciences) for dextran-conjugated mAb to<br />

NATURE IMMUNOLOGY VOLUME 10 NUMBER 5 MAY 2009 549


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

ARTICLES<br />

d-chain; Z. Yu for statistic analysis; A. Schaffer for discussions; and S. Sabet and<br />

M. Kang for technical assistance. Supported by the US National Institutes of<br />

Health (AI 045011, AI 079705 and AI 060573 to P.C.).<br />

AUTHOR CONTRIBUTIONS<br />

S.-R.P., H.Z., Z.P., J.Z., A.A.-Q., E.J.P., Z.X. and T.M. did experiments; H.Z.<br />

designed experiments, analyzed data and prepared the manuscript; and P.C.<br />

designed all experiments, analyzed the data, supervised the work and prepared<br />

the manuscript.<br />

Published online at http://www.nature.com/natureimmunology/<br />

Reprints and permissions information is available online at http://npg.nature.com/<br />

reprintsandpermissions/<br />

1. Honjo, T., Kinoshita, K. & Muramatsu, M. Molecular mechanism of class switch<br />

recombination: linkage with somatic hypermutation. Annu. Rev. Immunol. 20,<br />

165–196 (2002).<br />

2. Honjo, T., Muramatsu, M. & Fagarasan, S. Aid: How does it aid antibody diversity?<br />

Immunity 20, 659–668 (2004).<br />

3. Neuberger, M.S., Harris, R.S., Di Noia, J. & Petersen-Mahrt, S.K. Immunity through<br />

DNA deamination. Trends Biochem. Sci. 28, 305–312 (2003).<br />

4. Chaudhuri, J. & Alt, F.W. Class-switch recombination: interplay of transcription, DNA<br />

deamination and DNA repair. Nat. Rev. Immunol. 4, 541–552 (2004).<br />

5. Rada, C., Di Noia, J.M. & Neuberger, M.S. Mismatch recognition and uracil excision<br />

provide complementary paths to both Ig switching and the A/T-focused phase of<br />

somatic mutation. Mol. Cell 16, 163–171 (2004).<br />

6. Maizels, N. Immunoglobulin gene diversification. Annu. Rev. Genet. 39, 23–46<br />

(2005).<br />

7. Di Noia, J.M. & Neuberger, M.S. Molecular mechanisms of antibody somatic hypermutation.<br />

Annu. Rev. Biochem. 76, 1–22(2007).<br />

8. Peled, J.U. et al. The biochemistry of somatic hypermutation. Annu. Rev. Immunol. 26,<br />

481–511 (2008).<br />

9. Diaz, M. & Casali, P. Somatic immunoglobulin hypermutation. Curr. Opin. Immunol.<br />

14, 235–240 (2002).<br />

10. Papavasiliou, F.N. & Schatz, D.G. Somatic hypermutation of immunoglobulin genes;<br />

merging mechanisms for genetic diversity. Cell 109, s35–s44 (2002).<br />

11. Wu, X. et al. Immunoglobulin somatic hypermutation: double-strand DNA breaks, AID<br />

and error-prone DNA repair. J. Clin. Immunol. 23, 235–246 (2003).<br />

12. Xu, Z. et al. DNA lesions and repair in immunoglobulin class switch recombination and<br />

somatic hypermutation. Ann. NY Acad. Sci. 1050, 146–162 (2005).<br />

13. Casali, P., Pal, Z., Xu, Z. & Zan, H. DNA repair in antibody somatic hypermutation.<br />

Trends Immunol. 27, 313–321 (2006).<br />

14. Yan, C.T. et al. IgH class switching and translocations use a robust non-classical endjoining<br />

pathway. <strong>Nature</strong> 449, 478–482 (2007).<br />

15. Sayegh, C.E., Quong, M.W., Agata, Y. & Murre, C. E-proteins directly regulate expression<br />

of activation-induced deaminase in mature B cells. Nat. Immunol. 4, 586–593<br />

(2003).<br />

16. Gonda, H. et al. The balance between Pax5 and Id2 activities is the key to AID gene<br />

expression. J. Exp. Med. 198, 1427–1437 (2003).<br />

17. Yadav, A. et al. Identification of a ubiquitously active promoter of the murine<br />

activation-induced cytidine deaminase (AICDA) gene. Mol. Immunol. 43, 529–541<br />

(2006).<br />

18. Pearson, J.C., Lemons, D. & McGinnis, W. Modulating Hox gene functions during<br />

animal body patterning. Nat. Rev. Genet. 6, 839–904 (2005).<br />

19. Zakany, J. & Duboule, D. The role of Hox genes during vertebrate limb development.<br />

Curr. Opin. Genet. Dev. 17, 359–366 (2007).<br />

20. Deschamps, J. & van Nes, J. Developmental regulation of the Hox genes during axial<br />

morphogenesis in the mouse. Development 132, 2931–2942 (2005).<br />

21. Lemons, D. & McGinnis, W. Genomic evolution of Hox gene clusters. Science 313,<br />

1918–1922 (2006).<br />

22. Bijl, J. et al. Expression of HOXC4, HOXC5, and HOXC6 in human lymphoid cell lines,<br />

leukemias, and benign and malignant lymphoid tissue. Blood 87, 1737–1745<br />

(1996).<br />

23. Meazza, R. et al. Expression of HOXC4 homeoprotein in the nucleus of activated human<br />

lymphocytes. Blood 85, 2084–2090 (1995).<br />

24. Schaffer, A. et al. Selective inhibition of class switching to IgG and IgE by recruitment<br />

of the HoxC4 and Oct-1 homeodomain proteins and Ku70/Ku86 to newly identified<br />

ATTT cis-elements. J. Biol. Chem. 278, 23141–23150 (2003).<br />

25. Kim, E.C., Edmonston, C.R., Wu, X., Schaffer, A. & Casali, P. The HoxC4 homeodomain<br />

protein mediates activation of the immunoglobulin heavy chain 3¢ hs1,2 enhancer in<br />

human B cells. Relevance to class switch DNA recombination. J. Biol. Chem. 279,<br />

42258–42269 (2004).<br />

26. Pasqualucci, L. et al. Expression of the AID protein in normal and neoplastic B cells.<br />

Blood 104, 3318–3325 (2004).<br />

27. Guikema, J.E. et al. Quantitative RT-PCR analysis of activation-induced cytidine<br />

deaminase expression in tissue samples from mantle cell lymphoma and B-cell chronic<br />

lymphocytic leukemia patients. Blood 105, 2997–2998 (2005).<br />

28. Bijl, J.J. et al. HOXC4, HOXC5, and HOXC6 expression in non-Hodgkin’s lymphoma:<br />

preferential expression of the HOXC5 gene in primary cutaneous anaplastic T-cell and<br />

oro-gastrointestinal tract mucosa-associated B-cell lymphomas. Blood 90, 4116–4125<br />

(1997).<br />

29. Boulet, A.M. & Capecchi, M.R. Targeted disruption of hoxc-4 causes esophageal<br />

defects and vertebral transformations. Dev. Biol. 177, 232–249 (1996).<br />

30. Saegusa, H., Takahashi, N., Noguchi, S. & Suemori, H. Targeted disruption in the<br />

mouse Hoxc-4 locus results in axial skeleton homeosis and malformation of the xiphoid<br />

process. Dev. Biol. 174, 55–64 (1996).<br />

31. Xu, Z. et al. Regulation of aicda expression and AID activity: relevance to somatic<br />

hypermutation and class switch DNA recombination. Crit.Rev.Immunol.27, 367–397<br />

(2007).<br />

32. Crouch, E.E. et al. Regulation of AID expression in the immune response. J. Exp. Med.<br />

204, 1145–1156 (2007).<br />

33. Teng, G. et al. MicroRNA-155 is a negative regulator of activation-induced cytidine<br />

deaminase. Immunity 28, 621–629 (2008).<br />

34. Dedeoglu, F., Horwitz, B., Chaudhuri, J., Alt, F.W. & Geha, R.S. Induction of activationinduced<br />

cytidine deaminase gene expression by IL-4 and CD40 ligation is dependent<br />

on STAT6 and NFkB. Int. Immunol. 16, 395–404 (2004).<br />

35. Schaffer, A., Cerutti, A., Shah, S., Zan, H. & Casali, P. The evolutionary conserved<br />

sequence upstream of the human Ig Sg3 region is an inducible promoter: Synergistic<br />

activation by CD40 ligand and IL-4 via cooperative NF-kB and STAT-6 binding sites.<br />

J. Immunol. 162, 5327–5336 (1999).<br />

36. Muramatsu, M. et al. Class switch recombination and hypermutation require activationinduced<br />

cytidine deaminase (AID), a potential RNA editing enzyme. Cell 102,<br />

553–563 (2000).<br />

37. Jiang, C., Zhao, M.L. & Diaz, M. Activation-induced deaminase heterozygous MRL/lpr<br />

mice are delayed in the production of high-affinity pathogenic antibodies and in the<br />

development of lupus nephritis. <strong>Immunology</strong> 126, 102–113 (2009).<br />

38. Takizawa, M. et al. AID expression levels determine the extent of cMyc oncogenic<br />

translocations and the incidence of B cell tumor development. J. Exp. Med. 205,<br />

1949–1957 (2008).<br />

39. Saville, B. et al. Ligand-, cell-, and estrogen receptor subtype (a/b)-dependent<br />

activation<br />

at GC-rich (Sp1) promoter elements. J. Biol. Chem. 275, 5379–5387 (2000).<br />

40. Kwon, K. et al. Instructive role of the transcription factor E2A in early B lymphopoiesis<br />

and germinal center B cell development. Immunity 28, 751–762 (2008).<br />

41. Schoetz, U., Cervelli, M., Wang, Y.-D., Fiedler, P. & Buerstedde, J.-M. E2A expression<br />

stimulates Ig hypermutation. J. Immunol. 177, 395–400 (2006).<br />

42. Odegard, V.H., Kim, S.T., Anderson, S.M., Shlomchik, M.J. & Schatz, D.G. Histone<br />

modifications associated with somatic hypermutation. Immunity 23, 101–110 (2005).<br />

43. Teitell, M.A. OCA-B regulation of B-cell development and function. Trends Immunol.<br />

24, 546–553 (2003).<br />

44. Sale, J.E. & Neuberger, M.S. TdT-accessibe breaks are scattered over the immunoglobulin<br />

V domain in a constitutively hypermutating B cell line. Immunity 9, 859–869<br />

(1998).<br />

45. Cerutti, A. et al. CD40 ligand and appropriate cytokines induce switching to IgG, IgA,<br />

and IgE and coordinated germinal center and plasmacytoid phenotypic differentiation<br />

in a human monoclonal IgM + IgD + Bcellline.J. Immunol. 160, 2145–2157 (1998).<br />

46. Zan, H. et al. Induction of Ig somatic hypermutation and class switching in a human<br />

monoclonal IgM + IgD + cell line in vitro: Definition of the requirements and the<br />

modalities of hypermutation. J. Immunol. 162, 3437–3447 (1999).<br />

47. Zan, H., Cerutti, A., Schaffer, A., Dramitinos, P. & Casali, P. CD40 engagement triggers<br />

switching to IgA1 and IgA2 in human B cells through induction of endogenous TGF-b.<br />

Evidence for TGF-b but not IL-10-dependent direct Sm-Sa and sequential Sm-Sg,<br />

Sg-Sa DNA recombination. J. Immunol. 161, 5217–5225 (1998).<br />

48. Zan, H. et al. The translesion DNA polymerase z plays a major role in Ig and Bcl-6<br />

somatic hypermutation. Immunity 14, 643–653 (2001).<br />

49. Zan, H. et al. BCR engagement and T cell contact induce Bcl-6 hypermutation in<br />

human B cells: association with initiation of transcription and identity with immunoglobulin<br />

hypermutation. J. Immunol. 165, 830–839 (2000).<br />

50. Zan, H., Wu, X., Komori, A., Holloman, W.K. & Casali, P. AID-dependent generation of<br />

resected double-strand DNA breaks and recruitment of Rad52/Rad51 in somatic<br />

hypermutation. Immunity 18, 727–738 (2003).<br />

51. Zan, H. & Casali, P. AID- and Ung-dependent generation of staggered double-strand<br />

DNA breaks in immunoglobulin class switch DNA recombination: A post-cleavage role<br />

for AID. Mol. Immunol. 46, 45–61 (2008).<br />

52. Muramatsu, M. et al. Specific expression of activation-induced cytidine deaminase<br />

(AID), a novel member of the RNA-editing deaminase family in germinal center B cells.<br />

J. Biol. Chem. 274, 18470–18476 (1999).<br />

53. Rush, J.S., Liu, M., Odegard, V.H., Unniraman, S. & Schatz, D.G. Expression of<br />

activation-induced cytidine deaminase is regulated by cell division, providing a<br />

mechanistic basis for division-linked class switch recombination. Proc. Natl. Acad.<br />

Sci. USA 102, 13242–13247 (2005).<br />

54. Zan, H. et al. Lupus-prone MRL/faslpr/lpr mice display increased AID expression and<br />

extensive DNA lesions, comprising deletions and insertions, in the immunoglobulin<br />

locus: Concurrent upregulation of somatic hypermutation and class switch DNA<br />

recombination. Autoimmunity 42, 89–103 (2009).<br />

55. Zan, H. et al. The translesion DNA polymerase y plays a major role in immunoglobulin<br />

gene somatic hypermutation. EMBO J. 24, 3757–3769 (2005).<br />

550 VOLUME 10 NUMBER 5 MAY 2009 NATURE IMMUNOLOGY


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Corrigendum: ADAR1 is essential for the maintenance of hematopoiesis and<br />

suppression of interferon signaling<br />

Jochen C Hartner, Carl R Walkley, Jun Lu & Stuart H Orkin<br />

Nat. Immunol. 10, 109–115 (2009); published online 7 December 2008; corrected after print 9 April 2009<br />

CORRIGENDA AND ERRAtA<br />

In the version of this article initially published, the Cre-transgenic mouse is identified incorrectly as Tg(SV40-cre)1Jrg. The correct mouse strain<br />

should be Tg(SCL6E5-Cre)1Jrg, and the citation describing this mouse (ref. 29) should be as follows: Gothert, J.R. et al. In vivo fate-tracing studies<br />

using the Scl stem cell enhancer: embryonic hematopoietic stem cells significantly contribute to adult hematopoiesis. Blood 105, 2724–2732 (2005).<br />

The error has been corrected in the HTML and PDF versions of the article.<br />

Erratum: IL-4 inhibits TGF-β-induced Foxp3 + T cells and, together with<br />

TGF-β, generates IL-9 + Foxp3 – effector T cells<br />

Valérie Dardalhon, Amit Awasthi, Hyoung Kwon, George Galileos, Wenda Gao, Raymond A Sobel, Meike Mitsdoerffer, Terry B Strom,<br />

Wassim Elyaman, I-Cheng Ho, Samia Khoury, Mohamed Oukka & Vijay K Kuchroo<br />

Nat. Immunol. 9, 1347–1355 (2008); published online 9 November 2008; corrected after print 9 April 2009<br />

In the version of this article initially published, graph axes are mislabeled in Figures 1e, 2d and 2e, and a gene symbol is misidentified in the legend<br />

to Figure 2. The correct axis labels should be as follows: Figure 1e, middle right and far right vertical axes should end “(relative × 10 3 )”; Figure<br />

2d,e, left vertical axes should read “Il9 mRNA (relative × 10 2 )”; Figure 2e, right vertical axis should read “Il10 mRNA (relative)”; and Figure 2e,<br />

horizontal axes should read “WT” (in place of “Foxp3-GFP”) and “GATA3-KO” (in place of “STAT6-KO.Fox3-GFP”). The legend for Figure 2d<br />

should state “relative to Hprt1 mRNA” and the legend to Figure 2e should have that phrase removed. The errors have been corrected in the HTML<br />

and PDF versions of the article.<br />

nature immunology volume 10 number 5 may 2009 551


© 2009 <strong>Nature</strong> America, Inc. All rights reserved.<br />

Corrigendum: ADAR1 is essential for the maintenance of hematopoiesis and<br />

suppression of interferon signaling<br />

Jochen C Hartner, Carl R Walkley, Jun Lu & Stuart H Orkin<br />

Nat. Immunol. 10, 109–115 (2009); published online 7 December 2008; corrected after print 9 April 2009<br />

CORRIGENDA AND ERRAtA<br />

In the version of this article initially published, the Cre-transgenic mouse is identified incorrectly as Tg(SV40-cre)1Jrg. The correct mouse strain<br />

should be Tg(SCL6E5-Cre)1Jrg, and the citation describing this mouse (ref. 29) should be as follows: Gothert, J.R. et al. In vivo fate-tracing studies<br />

using the Scl stem cell enhancer: embryonic hematopoietic stem cells significantly contribute to adult hematopoiesis. Blood 105, 2724–2732 (2005).<br />

The error has been corrected in the HTML and PDF versions of the article.<br />

Erratum: IL-4 inhibits TGF-β-induced Foxp3 + T cells and, together with<br />

TGF-β, generates IL-9 + Foxp3 – effector T cells<br />

Valérie Dardalhon, Amit Awasthi, Hyoung Kwon, George Galileos, Wenda Gao, Raymond A Sobel, Meike Mitsdoerffer, Terry B Strom,<br />

Wassim Elyaman, I-Cheng Ho, Samia Khoury, Mohamed Oukka & Vijay K Kuchroo<br />

Nat. Immunol. 9, 1347–1355 (2008); published online 9 November 2008; corrected after print 9 April 2009<br />

In the version of this article initially published, graph axes are mislabeled in Figures 1e, 2d and 2e, and a gene symbol is misidentified in the legend<br />

to Figure 2. The correct axis labels should be as follows: Figure 1e, middle right and far right vertical axes should end “(relative × 10 3 )”; Figure<br />

2d,e, left vertical axes should read “Il9 mRNA (relative × 10 2 )”; Figure 2e, right vertical axis should read “Il10 mRNA (relative)”; and Figure 2e,<br />

horizontal axes should read “WT” (in place of “Foxp3-GFP”) and “GATA3-KO” (in place of “STAT6-KO.Fox3-GFP”). The legend for Figure 2d<br />

should state “relative to Hprt1 mRNA” and the legend to Figure 2e should have that phrase removed. The errors have been corrected in the HTML<br />

and PDF versions of the article.<br />

nature immunology volume 10 number 5 may 2009 551

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!