11.01.2013 Views

Chapter 1 - Stanford University

Chapter 1 - Stanford University

Chapter 1 - Stanford University

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.

MOLECULAR MECHANISMS OF THE INNATE IMMUNE<br />

RESPONSE TO<br />

FRANCISELLA TULARENSIS<br />

A DISSERTATION<br />

SUBMITTED TO<br />

THE DEPARTMENT OF MICROBIOLOGY AND IMMUNOLOGY<br />

AND THE COMMITTEE ON GRADUATE STUDIES<br />

OF STANFORD UNIVERSITY<br />

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS<br />

FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY<br />

Jonathan Wiley Jones<br />

August 2010


© 2010 by Jonathan Wiley Jones. All Rights Reserved.<br />

Re-distributed by <strong>Stanford</strong> <strong>University</strong> under license with the author.<br />

This work is licensed under a Creative Commons Attribution-<br />

Noncommercial 3.0 United States License.<br />

http://creativecommons.org/licenses/by-nc/3.0/us/<br />

This dissertation is online at: http://purl.stanford.edu/rq941zv2693<br />

ii


I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree of Doctor of Philosophy.<br />

Denise Monack, Primary Adviser<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree of Doctor of Philosophy.<br />

Manuel Amieva<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree of Doctor of Philosophy.<br />

Stanley Falkow<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree of Doctor of Philosophy.<br />

William Nelson<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree of Doctor of Philosophy.<br />

Approved for the <strong>Stanford</strong> <strong>University</strong> Committee on Graduate Studies.<br />

David Schneider<br />

Patricia J. Gumport, Vice Provost Graduate Education<br />

This signature page was generated electronically upon submission of this dissertation in<br />

electronic format. An original signed hard copy of the signature page is on file in<br />

<strong>University</strong> Archives.<br />

iii


Abstract<br />

Francisella tularensis is a facultative intracellular pathogen that causes the<br />

disease tularemia. The ability of F. tularensis to escape phagosomal degradation and<br />

replicate in the macrophage cytosol is central to its pathogenesis. The macrophage<br />

responds to the presence of cytosolic F. tularensis with the production of type-I<br />

interferons (IFN) and subsequent activation of the inflammasome, which leads to host<br />

cell death that eliminates the bacterium’s replicative niche. Very little is known about<br />

the molecular mechanisms that lead to cytosolic recognition of F. tularensis or<br />

bacterial factors that modulate this process and allow the bacterium to establish a<br />

niche in the macrophage. The bacterial ligand(s) that trigger the cytosolic response as<br />

well as the host pattern recognition receptors that lead to type-I IFN production and<br />

inflammasome activation are unknown. By investigating the molecular mechanisms of<br />

F. tularensis recognition in the cytosol we hypothesized that we would uncover novel<br />

bacterial PAMPs and novel host PRRs and greatly broaden our understanding of<br />

innate immunity.<br />

To this end we conducted a forward genetic screen of a F. novicida transposon<br />

library to identify mutants that resulted in an increased or decreased cytosolic response<br />

in macrophages. We identified 164 F. novicida mutants that lead to increased type-I<br />

IFN production and inflammasome activation in macrophages. These included a<br />

major outer membrane protein, fopA, as well as genes involved in LPS/capsule/cell<br />

wall biosynthesis, FTN_1212, lpcC, wbtA, kdsA, and lpxH. We also identified 74<br />

mutants that resulted in decreased type-I IFN and inflammasome responses in<br />

macrophages. These included 17 of the 19 genes in the Francisella Pathogenicity<br />

iv


Island, and several purine and pyrimidine biosynthesis genes. These mutants<br />

demonstrated a correlation between intracellular replication and induction of the<br />

cytosolic responses. The fact that we did not identify a single mutant that replicated<br />

intracellularly as efficiently as wild-type F. novicida but did not activate the cytosolic<br />

responses suggested that the presence of the bacterial ligand was associated with<br />

bacterial replication and possibly an essential gene.<br />

Finally, we identified AIM2 as the host receptor responsible for inflammasome<br />

activation in response to cytosolic F. novicida. We showed that lysing cytosolic F.<br />

novicida leads to release of bacterial DNA that triggers type-IFN through a pathway<br />

involving the adaptor STING. STING-dependent type-I IFN production increases the<br />

expression of AIM2, which complexes with the bacterial DNA and initiates<br />

inflammasome activation. We further demonstrate that AIM2 is critical for innate<br />

immunity to F. novicida infection in vivo. Thus we identified a novel bacterial ligand<br />

and novel cytosolic sensing components that play a role in the host defense to bacterial<br />

infections.<br />

v


Acknowledgements<br />

First and foremost I have to thank God for all of His blessings. I know that<br />

everything I have comes from Him. I want to thank my family for their continued<br />

love and support throughout my life. My mom, Carolyn Jones, instilled in me the<br />

importance of education, and encouraged me to pursue my dreams no matter how far<br />

from home they took me. My father, Wiley Jones, gave me strength, and continues to<br />

be my inspiration even after his passing. I love you both very much. My<br />

grandparents, sisters, aunts, uncles, nieces and nephews remind me that life is bigger<br />

than graduate school, and no matter where I go I will always have a place to call<br />

home. I want to thank my friends from Detroit, Troy Walls, Tarik & Crystal Green,<br />

Blair Parkman, Nathan Hood, Cordelia Ziraldo, and Konadu & Kim Addai for never<br />

letting me forget my roots.<br />

I have had an amazing time at <strong>Stanford</strong>. I have to thank Denise Monack for<br />

guiding me through this crazy process of grad school. You are a great mentor,<br />

colleague, and friend. I’d like to thank the Falkow, Amieva, and Monack labs for<br />

making lab fun, even when science wasn’t. Elizabeth Joyce knows everything and is<br />

an amazing teacher. To my classmates, you guys are wonderful. Jeff & Lacey<br />

Margolis helped me keep my head on straight, and Beth Ponder helped make Cookies<br />

& Cream a success every year. You all are the best. It’s been a wild ride. Peace and<br />

I’m out!!!<br />

vi


Table of Contents<br />

Section Page<br />

Abstract................................................................................................................... iv<br />

Table of Contents................................................................................................... vii<br />

Table of Figures ...................................................................................................... ix<br />

List of Tables.............................................................................................................x<br />

<strong>Chapter</strong> 1: General Introduction .............................................................................1<br />

Innate immunity.......................................................................................................................................... 1<br />

Pattern Recognition Receptors (PRRs) ..................................................................................................... 3<br />

Type-I IFNs................................................................................................................................................. 6<br />

Inflammasomes........................................................................................................................................... 8<br />

Tularemia .................................................................................................................................................. 14<br />

Intracellular lifestyle of F. tularensis...................................................................................................... 17<br />

Innate immunity to F. tularensis ............................................................................................................. 19<br />

<strong>Chapter</strong> 2: A genetic screen identifies novel F. novicida genes that modulate the<br />

cytosolic innate immune responses in macrophages..............................................22<br />

2.1 CHAPTER 2 SUMMARY ................................................................................................................ 23<br />

2.2 INTRODUCTION.............................................................................................................................. 25<br />

2.3 RESULTS ........................................................................................................................................... 28<br />

2.3.1 A transposon screen identifies fopA as a suppressor of the macrophage cytosolic innate<br />

immune response. ................................................................................................................................ 28<br />

2.3.2 ΔfopA stimulates the same cytosolic surveillance pathway as wild-type F. novicida........... 31<br />

2.3.3 A ΔfopA mutant has reduced fitness in mice............................................................................ 36<br />

2.3.4 Identification of F. tularensis mutants that differentially induce the cytosolic response in<br />

macrophages by a genome-wide forward genetic screen.................................................................. 37<br />

2.3.5 F. novicida LPS mutants hyper-induce the cytosolic responses. ............................................ 56<br />

2.3.6 F. novicida LPS mutants stimulate increased TLR2-depedent signaling............................... 60<br />

2.3.7 F. novicida LPS mutants hyper stimulate the inflammasome................................................. 64<br />

2.3.8 LPS mutants have reduced fitness in vivo................................................................................ 66<br />

2.3.9 Phagosomal escape is required for induction of the cytosolic response by LPS mutants.... 67<br />

2.3.10 LPS mutants induce increased proinflammatory cytokine signaling in the phagosome.... 70<br />

2.3.11 Surface-exposed PAMPs mediate recognition of LPS mutants............................................ 72<br />

2.3.12 Cytosolic localization is necessary but not sufficient to induce the cytosolic responses... 75<br />

2.3.13 Bacterial DNA and protein synthesis are required to induce the cytosolic responses........ 79<br />

2.3.14 The cytosolic response to F. novicida shares characteristics with the response to<br />

transfected dsDNA. ............................................................................................................................. 85<br />

2.4 DISCUSSION..................................................................................................................................... 88<br />

<strong>Chapter</strong> 3: AIM2 is required for innate immune recognition of Francisella<br />

tularensis..................................................................................................................90<br />

3.1 CHAPTER 3 SUMMARY ................................................................................................................ 91<br />

3.2 INTRODUCTION.............................................................................................................................. 92<br />

3.3 RESULTS ........................................................................................................................................... 94<br />

vii


3.3.1 AIM2 is essential for inflammasome activation in response to cytosolic dsDNA. ............... 94<br />

3.3.2 AIM2 is required for inflammasome activation in response to F. tularensis....................... 101<br />

3.3.3 AIM2 and ASC form a complex with F. tularensis DNA. ................................................... 104<br />

3.3.4 AIM2 is required for the formation of an ASC focus............................................................ 107<br />

3.3.5 Type I IFN increases AIM2 protein levels and inflammasome activity............................... 109<br />

3.4 Discussion......................................................................................................................................... 115<br />

<strong>Chapter</strong> 4: Discussion ...........................................................................................119<br />

<strong>Chapter</strong> 5: Materials and Methods ......................................................................129<br />

5.1 BEIR “two-allele” transposon library screen. ................................................................................ 129<br />

5.2 Bacterial strains and growth conditions.......................................................................................... 130<br />

5.3 Bacterial Mutagenesis...................................................................................................................... 130<br />

5.4 Bone marrow-derived macrophage culture and infections............................................................ 131<br />

5.5 Macrophage gene expression analysis. ........................................................................................... 132<br />

5.6 ISRE-L929 assays. ........................................................................................................................... 133<br />

5.7 NF-kB reporter cell assays............................................................................................................... 133<br />

5.8 Mice, bacteria, and reagents ............................................................................................................ 133<br />

5.9 Immunofluorescence Microscopy ................................................................................................... 134<br />

REFERENCES .....................................................................................................137<br />

viii


Table of Figures<br />

Figure Page<br />

Figure 1 - An F. novicida ΔfopA mutant hyper-induces the cytosolic innate immune response in<br />

macrophages. ....................................................................................................................................... 30<br />

Figure 2 - The cytosolic response to ΔfopA is IRF3- and IFNAR- dependent but TLR- and NOD-<br />

independent.......................................................................................................................................... 33<br />

Figure 3- The hyper-cytotoxicity ofΔfopA is ASC- and caspase-1 dependent but NLRP3-independent.<br />

.............................................................................................................................................................. 35<br />

Figure 4 - ΔfopA is less fit in vivo than wild-type F.novicida. .................................................................. 37<br />

Figure 5 - LPS mutants hyper-induce the type-I IFN response in macrophages....................................... 58<br />

Figure 6 - LPS mutants hyper-induce IL-1β release and host cell death................................................... 59<br />

Figure 7 - TLR signaling contributes to the early but not late responses to LPS mutants........................ 61<br />

Figure 8- TLR2 contributes to the early but not late cytosolic responses to LPS mutants....................... 63<br />

Figure 9 - ASC, caspase-1, and IFNAR are required for LPS mutant inflammasome activation............ 65<br />

Figure 10 – LPS mutants have reduced fitness in vivo. .............................................................................. 67<br />

Figure 11 - Phagosomal escape is required for LPS mutants to hyper-induce the cytosolic responses. . 69<br />

Figure 12 - LPS mutants hyper-induce NF-κB-dependent cytokines in the phagosome.......................... 71<br />

Figure 13 - Surface exposed PAMPs mediate recognition of LPS mutants. ............................................. 74<br />

Figure 14 - Cytosolic localization is necessary but insufficient to induce the cytosolic responses......... 78<br />

Figure 15 - Bacterial protein synthesis and DNA synthesis are required to induce the cytosolic<br />

responses. ............................................................................................................................................. 82<br />

Figure 16 - Replication is not required to induce the cytosolic response with high bacterial load.......... 84<br />

Figure 17 - The cytoslic response to F. novicida shares characteristics with that of dsDNA.................. 86<br />

Figure 18 - Generation of Aim2 -/- mice. ....................................................................................................... 96<br />

Figure 19 - AIM2 is essential for inflammasome activation in response to cytosolic dsDNA. ............... 97<br />

Figure 20 - AIM2 is essential for inflammasome activation in response to dsDNA in peritoneal<br />

macrophages. ....................................................................................................................................... 98<br />

Figure 21 - AIM2 is dispensable for IFN-β and TNF-α production in response to dsDNA.................. 101<br />

Figure 22 - AIM2 is required for inflammasome activation in response to F. tularensis. ..................... 103<br />

Figure 23 - AIM2 and ASC form a complex with F. tularensis DNA. ................................................... 105<br />

Figure 24 - F. tularensis DNA triggers IL-1β secretion. .......................................................................... 107<br />

Figure 25 - AIM2 is required for the formation of an ASC focus............................................................ 109<br />

Figure 26 - Type-I IFN increases AIM2 protein levels and inflammasome activity. ............................. 111<br />

Figure 27 - Type-I IFN facilitates formation of the AIM2-containing inflammasome. ......................... 113<br />

Figure 28 - AIM2 is required for host defense against F. tularensis ....................................................... 115<br />

ix


List of Tables<br />

Table Page<br />

- Table 1 – Mutants that hyper-induce the cytosolic responses ............................................................... 39<br />

- Table 2 – Mutants that hypo-induce the cytosolic responses ................................................................ 51<br />

- Table 3 – Mutants attenuated for intracellular replication are hypo-stimlate the cytosolic responses76<br />

- Table 4 – Primers for F. novicida cloning and mutagenesis ............................................................... 135<br />

x


<strong>Chapter</strong> 1: General Introduction<br />

Innate immunity<br />

As scientists studying microbial pathogenesis we are uncovering an<br />

evolutionary war that has been going on since the first time a single-celled organism<br />

ate a microbe. Since then there has been an enormous selective pressure for microbes<br />

to evolve strategies to resist detection and killing by host cells. From the microbe’s<br />

perspective, it either needs the host cell in order to multiply (such is the case with<br />

viruses, and obligate intracellular bacteria/parasites), or it simply is trying to take<br />

advantage of a niche free of competition for nutrients from other microbes, and avoid<br />

detection and killing by other cells. Few microbes have the ability to survive inside of<br />

a host, but those that can are termed pathogens or symbionts. Interacting with another<br />

organism provides a significant number of challenges: you have to be able to attach to<br />

the surface or enter a perspective host, you have to be able to maintain your location,<br />

and you must be able to obtain nutrients to either grow or persist in the face of an<br />

immune system bent on your destruction.<br />

Two arms of the immune system, the innate and adaptive responses, work<br />

together to detect and clear pathogens. Innate immunity is the first line of defense<br />

against invading microbes, and usually clears most infections without a problem.<br />

However if the innate responses prove insufficient, the orgasnism activates adaptive<br />

immunity to specifically target the pathogen. In contrast to adaptive immune<br />

mechanisms, which are only observed in vertebrates, innate immune defenses are<br />

found in all multi-cellular organisms. Therefore, it is evolutionarily ancient. When<br />

1


observing the characteristics of the innate and adaptive immune responses it is easy to<br />

see how the two work in concert. First, unlike adaptive immunity, which takes time to<br />

mount an antimicrobial response, innate immune defenses are constitutively in place<br />

and react rapidly upon infection. Second, the adaptive immune system responds to a<br />

specific microbial challenge, with the ability to generate T and B cell responses to<br />

specific pathogens, while the innate immune system is non-specific and instead<br />

recognizes “patterns” or “themes” in molecules unique to microbes. Finally, the<br />

adaptive immune system displays immunological memory, in that it “remembers” the<br />

microbe that it eliminated and responds even faster upon subsequent challenge with<br />

the same microbe. In contrast, the innate immune system has no memory, and reacts<br />

with the same speed and efficacy after repeated challenge with a microbe. The innate<br />

immune system works to fight off the initial infection and prime the adaptive immune<br />

system should the infection persist.<br />

The innate immune system includes anatomical barriers to infection, secretory<br />

components, and cellular mediators of pathogen detection and clearance. Anatomical<br />

barriers include the skin, which prevents penetration of pathogens, the movement of<br />

cilia that clears pathogens from air passages, the flushing action of tears and saliva,<br />

and mucous that lines the lungs and gastrointestinal tract, which traps and clears<br />

invading pathogens. Tears and saliva also contain antimicrobial factors such as<br />

lysozyme, which can dissolve microbial cell walls. If these anatomical barriers are<br />

breached, serum also contains secreted factors like lactoferrin that can sequester<br />

essential iron, eliminating an important co-factor for many microbial processes. The<br />

serum also contains complement, which can lyse infectious organisms, leading to their<br />

2


phagocytosis, and lead to inflammation and recruitment of other cellular responses.<br />

Cellular mediators of innate immunity are derived from the bone marrow and include<br />

macrophages, dendritic cells, neutrophils, natural killer cells, mast cells, basophils and<br />

eosinophils. These cells can kill microbes by phagocytosis and degradation, or<br />

through secretion of toxic molecules.<br />

Pattern Recognition Receptors (PRRs)<br />

The critical function of the immune system is the ability to discriminate<br />

between self and non-self. Being the first line of defense to infection, innate immune<br />

mechanisms must be able to protect against a diverse array of potentially threatening<br />

microbes, like viruses, bacteria, and fungi, all having different microbial factors to<br />

promote survival and virulence. With such a monumental task, the innate immune<br />

system takes a broad approach to identifying and eliminating all manner of microbes,<br />

using a relatively small set of germline-encoded sensors called pattern recognition<br />

receptors (PRRs). These PRRs are expressed on antigen presenting cells (APCs) like<br />

macrophages and dendritic cells, and take advantage of conserved molecules that are<br />

unique to microbes and often essential to microbial existence. Thus they are difficult<br />

for the microbes to alter significantly to avoid detection. The microbial molecules<br />

sensed by PRRs have been termed pathogen associated molecular patterns or PAMPs.<br />

Discovery of the Toll-like receptors (TLRs) was a breakthrough in the field of host-<br />

pathogen interaction.<br />

3


In 1996, studies in the fruit fly Drosophila, which only has an innate immune<br />

system, led to the identification of Toll as the central receptor in the host defense to<br />

fungal infection (92). One year later a homologous human Toll-like receptor was<br />

identified (109). At present there are 11 TLRs that have been identified and<br />

implicated in innate immunity (162). TLRs 1,2,4,5 and 6 recognize components of<br />

microbial cell wall while TLRs 3,7,8, and 9 recognize microbial nucleic acids. TLR4<br />

is probably the best characterized of the TLRs and is essential for recognizing<br />

bacterial lipopolysaccharide (LPS) (136). TLR1, TLR2, and TLR6 recognize<br />

lipoproteins from microbial cell wall (163, 164) and TLR5 recognizes bacterial<br />

flagellin (60). Of the nucleic acid sensors TLR9 recognizes unmethylated CpG motifs<br />

of bacterial DNA (63). TLR3 responds to viral double stranded-RNA (5), while TLR7<br />

and TLR8 respond to single stranded RNA from viruses (61). The TLRs can also be<br />

grouped according to their cellular location with TLR1, 2, 4, 5, and 6 located at the<br />

cell surface and TLR7, 8, and 9 localized to endosomes. The differential localization<br />

of TLRs allows cells of the innate immune system to survey multiple environments for<br />

potentially harmful pathogens.<br />

TLR signaling leads to the production of proinflammatory cytokines like<br />

Tumor necrosis factor-α (TNF-α), interferon-α/β (IFN)-α/β, IFN-γ, and interleukin-<br />

12 (IL-12). All TLRs except TLR3 use an adaptor protein myeloid differentiation<br />

factor 88 (MyD88), which leads to the activation of nuclear factor kappa-light-chain-<br />

enhancer of activated B cells (NF-κB) (2). TLR3 and TLR4 use the adaptor TIR-<br />

domain-containing adaptor protein-inducing IFN-β (TRIF) to activate interferon<br />

regulatory factor 3 (IRF3) and NF-κB in a MyD88-independent manner (3, 162, 182).<br />

4


Whether a pathogen’s niche is on the surface of a cell, inside a phagosome, or<br />

in the cell cytosol, it must be able to manipulate the host cell in order to establish an<br />

environment to promote survival and replication. To this end, pathogenic bacteria<br />

have acquired/evolved secretion systems and bacterial toxins that allow the pathogen<br />

to penetrate cell membranes and deliver effector molecules directly into the cell<br />

cytosol. As an evolutionary counterpart to bacterial secretion systems, host cells have<br />

evolved cytosolic surveillance mechanisms to detect danger signals in this fragile<br />

compartment of the cell. The nucleotide-binding oligomerization domain (NOD)<br />

proteins recognize components of the microbial cell wall in the cytosol (7, 24). NOD1<br />

senses γ-D-glutamyl-meso-diaminopimelic acid, and NOD2 senses muramyl dipeptide<br />

(MDP), both fragments of peptidoglycan in bacterial cell wall. The NODs activate<br />

NF-κB through the adaptor Rip-like interacting caspase-like apoptosis-regulatory<br />

protein kinase (RICK, also known as RIP2) to stimulate pro-inflammatory cytokine<br />

production (71).<br />

Additionally, two RNA helicases, retinoic acid inducible gene I (RIG-I) and<br />

melanoma differentiation-associated gene 5 (MDA5) detect viral dsRNA in the<br />

cytosol independent of TLR3 detection (184, 185). These two helicases uses the<br />

adaptor molecule mitochondrial antiviral signaling (MAVS) (also called IFN-β<br />

stimulator 1 (IPS-1), virus induced signaling adaptor (VISA), or Cardif) to stimulate<br />

production of type-I IFNs (IFN-α/β) through NF-κB and IRF-3 activation. There also<br />

appears to be an intracellular sensor(s) for dsDNA that acts independently from RIG-I<br />

and MDA5 to produce type-I IFNs (74, 127, 153). In this case, the production of type-<br />

I IFNs is still IRF3 dependent but is independent of MAVS signaling. One such<br />

5


dsDNA sensor, DNA-dependent activator of interferon regulatory factors (DAI) (also<br />

called DLM-1 or ZBP-1) was recently identified though lack of in vivo relevance<br />

suggests that redundant sensors must exist (75, 161). Other cytosolic dsDNA sensors<br />

that trigger type-I IFN production remain unknown.<br />

Type-I IFNs<br />

The type-I IFNs were the first cytokines discovered and represent an<br />

interesting class of cytokines in host defense. All vertebrates examined to date have a<br />

gene that encodes IFN-β and at least two that encode IFN-α, together with natural<br />

killer (NK) cells, T cells, and B cells (154). With the exception of LPS induced type-I<br />

IFN production through TLR4, type-I IFN is produced exclusively in response to<br />

nucleic acids. Furthermore, whereas the TLRs are only present on a subset of<br />

specialized immune cells, all nucleated cells can produce type-I IFNs and all nucleated<br />

cells express the type-I IFN receptor (IFNAR), which allows them to respond to type-I<br />

IFNs in the environment. Thus, we can surmise that all nucleated cells express one or<br />

more cytosolic PRRs that can signal the presence of intracellular infections through<br />

the detection of nucleic acids and lead to the production of type-I IFNs.<br />

Type-I IFNs were first described over 50 years ago as a soluble factor<br />

produced by cells treated with inactivated virus that blocked subsequent infection with<br />

live virus (72, 73, 120). It was later discovered that type-I IFNs signal through<br />

IFNAR in an autocrine and paracrine fashion to induce expression of hundreds of<br />

genes that establish an “antiviral state” in the cell (152, 172). We do not really<br />

6


understand the mechanisms behind this antiviral state, as most of the IFN-induced<br />

genes have not been ascribed a function. However, it is thought that many of these<br />

proteins cooperate with other signaling pathways to fully potentiate innate immunity.<br />

For example, treating cells with type-IFNs sensitizes them to apoptosis upon<br />

subsequent viral infection (152) or treatment with pore-forming bacterial toxins (188).<br />

This mechanism of cell suicide would effectively eliminate the replicative niche for<br />

intracellular infections. Perhaps for this reason several viruses have evolved<br />

mechanisms to block the apoptotic process (151).<br />

Although mainly studied as an antiviral mechanism, a number of recent reports<br />

demonstrate that type-I IFN is induced by a number of intracellular bacteria such as<br />

Mycobaterium tuberculosis (52, 176), Listeria monocytogenes (124), Legionella<br />

pneumophila (129), and Francisella tularensis (64). The bacterial PAMP that<br />

stimulates type-IFN production in these infections is unknown, though in the case of<br />

Listeria monocytogenes infection and Francisella tularensis infection it is suspected<br />

that a cytosolic DNA sensor could be involved (64, 153). Similar to its effects on viral<br />

infections, type-I IFN can acts to in concert with other cytosolic sensing pathways like<br />

the inflammasome to trigger a protective host cell death (32, 64). The mechanism<br />

linking the type-I IFN pathway and the inflammasome is currently unknown and is a<br />

central question of this thesis work.<br />

The existence of a host cell mediated death pathway represents an interesting<br />

paradox in host defense. The host cell effectively sacrifices itself for the survival of<br />

the whole organism. Theoretically this would only benefit an organism when tissue<br />

renewal can preserve the hosts physiological functions in the face of irreplaceable cell<br />

7


loss. Perhaps this is why only vertebrates have evolved type-I IFN signaling. Most<br />

cells in vertebrates are renewable, while most cells in adult invertebrates are<br />

postmitotic (154). Therefore loss of even a few key cells in invertebrates could be<br />

deleterious for the organism.<br />

Inflammasomes<br />

Protection from invading microbes and eliminating and replacing dying cells<br />

are two of the greatest challenges that multicellular organisms must have faced during<br />

evolution. Our innate immune system has learned to cope with these challenges by<br />

evolving alarm systems that can detect invading microbes and injured tissue, and<br />

initiate inflammation through the activity of cytokines and chemokines. Inflammation<br />

is one of the first responses of the innate immune system to infection and is<br />

characterized by redness, heat, swelling, and pain at the site of the infection. While<br />

inflammation is essential for fighting off infections and repairing damaged tissue, it<br />

can lead to further tissue damage if not tightly regulated. The same immune cells,<br />

molecules, and mechanisms involved in fighting pathogens can be harmful to<br />

uninfected tissue is not controlled. Tissue damage sustained during an infection can<br />

result from either direct action of a microbe or microbial on tissues (such as factors<br />

used for colonization/ toxins produced) or it can be the indirect result of an<br />

inflammatory response aimed at eliminating the pathogen, or both. Thus, intricate<br />

coordination of the immune response is required to strike a delicate balance between<br />

8


eliminating invading microbes and limiting inflammation to maintain tissue integrity.<br />

Too little of a response can lead to proliferation of the pathogen and the onset of<br />

disease, while too much of a response can lead to local and systemic tissue damage<br />

and even death in severe cases. An aberrant inflammatory response is the cause of<br />

several autoimmune diseases. Only an appropriate response eliminates the threat and<br />

goes undetected by the host.<br />

The key regulators of inflammation are enzymes known as inflammatory<br />

caspases. These inflammatory caspases are activated in complexes known as<br />

inflammasomes, which are cellular sensors of danger signals located in the cytosol of<br />

certain immune cells. Once the inflammasome complex activates caspases, they<br />

cleave pro-inflammatory cytokines into their mature, active forms, and direct their<br />

unconventional secretion from the cell. Additionally, the caspases control a form of<br />

cellular suicide. As this cell death is accompanied by the aforementioned released of<br />

pro-inflammatory cytokines, it has been termed “pyroptosis”, from the Greek roots<br />

“pyro” for fire denoting the inflammation, and “ptosis” meaning falling, denoting the<br />

death of the cell (89). For the purposes of this thesis, caspase-dependent cell death<br />

will just be referred to as cell death.<br />

Caspase-1 is a member of a family of intracellular aspartate-specific cysteine<br />

proteases. One of the subfamily of “inflammatory caspases”, its subfamily members<br />

include caspase-4, -5, -11 (which exists in rodents), and -12(149). Caspase-1 exists as<br />

an inactive 45kDa precursor, but once activated it undergoes autocatalytic cleavage<br />

into 20kDa and 10kDa subunits (p20 and p10) (9, 169, 180). In the cytosol, active<br />

caspase-1 processes pro-IL-1β(13, 87), pro-IL-18 (51), and pro IL-33 (147), into their<br />

9


mature, active forms that are secreted and regulate inflammation (in the case of IL-1β<br />

and IL-18), and generation of TH2 responses (in the case of IL-33). When regulated<br />

properly, IL-1β is critical for the host response to infection, but excessive levels of Il-<br />

1β are associated with several inflammatory diseases such as rheumatoid arthritis,<br />

inflammatory bowel disease, and septic shock to name a few (40). Clearly<br />

inflammation can be a double-edged sword but it is essential to combat infection and<br />

restore tissue homeostasis after infection.<br />

In 2002 it was discovered that a multiprotein complex termed the<br />

“inflammasome” was responsible for activating caspase-1(101). The inflammasome is<br />

composed of a NOD-like receptor (NLR), and the adaptor protein apoptosis associated<br />

spec-like protein containing a caspase recruitment domain (ASC). The NLRs are a<br />

family of cytosolic pattern recognition receptors (PRRs) that activate inflammatory<br />

and antimicrobial responses by sensing “danger signals” or danger-associated<br />

molecular patterns (DAMPs) (105) and conserved microbial products termed pathogen<br />

associated molecular patterns (PAMPs) (70, 171). There are 14 members of the NLRs<br />

subfamily, which are characterized by a C-terminal leucine rich repeat (LRR) domain,<br />

a central oligomerization domain (NACHT), and an N-terminus that is either a caspase<br />

recruitment domain (CARD) (as in NOD1, NOD2, and the NLRC family), three<br />

Baculovirus IAP Repeats (BIR) (as in the NAIPs), or a pyrin domain (PYD) (as in the<br />

NLRP family). The adaptor ASC (also known as PYCARD) has an N-terminal PYD<br />

that facilitates interactions with the PYD domain of NLRs, and a C-terminal CARD<br />

domain that recruits caspase-1 through CARD-CARD interactions (102). LRRs are<br />

also found in the Toll-like receptors (TLRs), which sense danger signals and microbial<br />

10


patterns on the cell surface and in endosomes (78, 162). Interestingly, the TLRs<br />

cannot distinguish between pathogenic and non-pathogenic microbes because they<br />

sense conserved microbial patterns that are present in both and they are located<br />

extracellularly and in endosomes, where both pathogenic and non-pathogenic<br />

microbes reside. The NLRs also sense conserved microbial patterns, but the only way<br />

these microbial patterns can reach the cytosol is if they are delivered there by<br />

disruption of cell membranes by toxins, specialized secretion systems of pathogenic<br />

microbes, and/or cytosolic pathogens. Therefore the location of the sensors and not<br />

the ligands they sense makes the NLRs specific for detecting pathogens (16).<br />

The vast number of NLR proteins allows the inflammasome to respond to<br />

numerous pathogens and danger signals. The cytosolic DAMP or PAMP sensed<br />

determines which NLR forms the complex. An inflammasome is named after the<br />

NLR that forms it (i.e. the NLRP3 inflammasome, the NLRC4 inflammasome, etc.),<br />

and so it can come in many different flavors. The inflammasome has been implicated<br />

in the host response to numerous pathogens and danger signals, as well as a number of<br />

autoimmune and auto-inflammatory diseases. For more a more thorough discussion of<br />

inflammasome regulation and its role in autoimmunity I suggest the reviews by<br />

Brodsky et al. (16), McIntire et al. (108), and Martinon, et al. (103). For my thesis<br />

work I have focused on the role of inflammasomes in pathogen recognition.<br />

Conserved microbial structures, such as bacterial cell wall components are<br />

potent activators of innate immunity. Muramyl dipeptide (MDP) is a breakdown<br />

product of bacterial peptidoglycan that is recognized by NOD2 and induces<br />

transcriptional activation of proinflammatory cytokines through the adaptor RIP2.<br />

11


MDP is also a potent activator of the inflammasome, which results in release of<br />

mature IL-1β. Inflammasome activation by MDP involves both NOD2 and NLRP3,<br />

suggesting that NLRP3 is an additional sensor of MDP (99, 100, 131). NOD2 has<br />

also been shown to cooperate with NLRP1b in inflammasome activation in response<br />

to MDP and anthrax lethal toxin (19, 69). These results suggest that inflammasome<br />

complexes may contain of multiple NLRs that act synergistically to activate caspase-1<br />

in response to PAMPs.<br />

Many bacteria employ pore-forming toxins in their pathogenic arsenal to<br />

establish infections. NLRP3 has been implicated in the inflammasome response to<br />

listeriolysin O form Listeria monocytogenes, α-toxin from Staphylococcus aureus and<br />

aerolysin from Aeromonas hydrophila (49, 98). The precise mechanism by which<br />

NLRP3 recognizes these toxins is unknown. NLRP3 also activates the inflammasome<br />

in response to high extracellular concentrations of ATP (98), various crystalline<br />

compounds (23, 41, 59, 104), as well as viral and bacterial nucleic acids (81, 119).<br />

With such a diverse array of stimuli, the hypothesis is that these ligands potentiate a<br />

common terminal signal that activates NLRP3. Potassium efflux, membrane damage,<br />

and stimulation of reactive oxygen species have all been implicated as the terminal<br />

signal for NLRP3, but each has caveats. Therefore, the precise mechanism of NLRP3<br />

activation remains a mystery.<br />

NLRC4 was the first NLR shown to activate caspase-1 in response to bacterial<br />

infection with Salmonella typhimurium in a type-III secretion system (T3SS)-<br />

dependent manner (96). It was later revealed that NLRC4 sensed bacterial flagellin<br />

that was secreted into the host cytosol through the T3SS, likely due to the evolutional<br />

12


similarity of the T3SS with the flagellar biosynthesis machinery (50, 113). Detection<br />

of cytosolic flagellin was later linked to the inflammasome response to Pseudomonas<br />

aeruginosa, Listeria monocytogenes, and Legionella pneumophila (114, 116, 141,<br />

174), though in the case of Legionella the type-IV secretion system (T4SS) was<br />

required and Naip5 was required in addition to NLRC4 (141, 186). NLRC4 was also<br />

implicated in recognition of Shigella flexneri (159), which encodes a T3SS but does<br />

not express flagellin (170), as well as non-flagellated strains of Pseudomonas (157),<br />

suggesting that NLRC4 must recognize more than just flagellin. The answer was<br />

recently provided when it was demonstrated that NLRC4 could also detect a basal<br />

body rod component of the T3SS, which shares a sequence motif with flagellin (115).<br />

This motif is essential to NLRC4-mediated recognition and explains how NLRC4 can<br />

respond to diverse pathogens that express either a T3SS or flagellin.<br />

Studies of NLRP1, NLRP3, and NLRC4 have expanded our understanding of<br />

the role of inflammasomes in the host response to pathogens. However, many<br />

pathogenic signals for inflammasome activation have yet to be elucidated.<br />

Transfection of bacterial, viral, or mammalian double-stranded DNA (dsDNA)<br />

triggers caspase-1 activation in a manner that requires ASC, but not any of the known<br />

NLRs (119, 139). Similarly, infection with Francisella tularensis activates an ASC-<br />

dependent inflammasome that does not require NLRP1, NLRP3, or NLRC4 (97, 98).<br />

Moreover, inflammasome activation by Francisella requires type-I IFN signaling,<br />

establishing a novel mechanism of inflammasome regulation (64). Thus, Francisella<br />

tularensis is a useful tool to expand our understanding of inflammasome activation<br />

13


and will allow us to uncover new NLRs and hence new stimuli involved in the host<br />

response to pathogens.<br />

Tularemia<br />

Tularemia is a zoonotic disease caused by the bacterium Francisella tularensis.<br />

The bacterium was first isolated from ground squirrels in 1911 in Tulare County<br />

California (107) after it was associated with an outbreak of a plague-like disease in<br />

rodents following the San Francisco earthquake. Further characterization of<br />

Francisella revealed that it is a pathogen with one of the widest host ranges of any<br />

known bacteria. It can infect rodents, hares, rabbits, muskrats, beavers, and voles<br />

among mammals, and can be transmitted by ticks, mosquitoes, and biting flies (123).<br />

Disease outbreaks in human populations will often parallel outbreaks in wild animals.<br />

Reports from Sweden have demonstrated a correlation between outbreaks of tularemia<br />

in vole and hare populations and outbreaks in humans (168).<br />

Although rabbits, hares, and rodents are important sources for human infection,<br />

tularemia is an acute disease in these animals and therefore they are unlikely to be a<br />

reservoir. Thus, the true environmental reservoir for Francisella remains a mystery.<br />

Outbreaks of tularemia in Europe have often been associated with fresh water (12, 37,<br />

62) where it may exist in biofilms (95) or associated with amoebae (1). Whatever the<br />

reservoir, it is clear that Francisella have evolved mechanisms to survive in a wide<br />

range of environments, with limited nutrient sources, and in the face of immune<br />

mechanisms.<br />

14


Francisella are pleomorphic, gram-negative coccobacilli ranging from 0.5-<br />

2um in size. Francisella tularensis is subdivided into four subspecies (ssp.):<br />

tularensis, holarctica, mediasiatica, and novicida. An additional Francisella species,<br />

Francisella philomiragia has also been identified. Tularemia seems to be<br />

geographically restricted to the Northern hemisphere. Subspecies tularensis, also<br />

known as type A, is found almost exclusively in North America, with a single reported<br />

case in Europe (56). F. tularensis is the most virulent subspecies, with an infectious<br />

dose in humans as low as 10 organisms (44, 112). Subspecies holarctica, or type B, is<br />

found mainly in North America, Europe, and Asia (117, 128). The Live Vaccine<br />

Strain (LVS) is derived from an isolate of subspecies holarctica and causes a<br />

tularemia like disease in mice although it is non-pathogenic to humans. The specific<br />

cause of the attenuation of LVS is unknown. Subspecies mediasiatica is found mainly<br />

in central Asia and subspecies novicida is found primarily in North America, although<br />

is has recently been isolated in Australia (178). Subspecies novicida is genetically<br />

tractable and causes a tularemia like disease in mice making it an ideal laboratory<br />

strain for studies of Francisella biology.<br />

F. tularensis garnered much interest in the 1950s when the United States and<br />

U.S.S.R began evaluating the organism’s potential as a biological weapon. The ability<br />

to aerosolize the organism, the low infectious dose, and the high incidence of mortality<br />

if left untreated make F. tularensis an ideal candidate for biological warfare. As a<br />

result, the CDC has categorized Francisella as a Category A bioterrorism agent. In<br />

the host, tularemia has an average incubation period of 3-5 days, but can range from 1-<br />

14 days. The clinical manifestations of tularemia may vary depending on the route of<br />

15


infection, the bacterial subspecies, and host genetics. The most severe form of the<br />

disease is pneumonic tularemia where the bacteria are inhaled directly into the lungs<br />

by aerosol. This usually occurs during farming activities where dust is generating<br />

moving hay that had been inhabited by infected rodents (66, 155, 160, 168). Cases of<br />

pneumonic tularemia have also recently been reported when a lawnmower ran over the<br />

carcass of a rabbit that had been infected with Francisella (46). This results in a very<br />

acute infection whose symptoms include high fever, chills, malaise, and cough (45).<br />

Pneumonic tularemia is extremely severe and has a mortality rate of 30% if untreated<br />

(38).<br />

Direct contact of Francisella with the conjunctiva can result in the<br />

oculoglandular form of the disease. This presentation is rare, accounting for 1-4% of<br />

cases (130). Other rare presentations of the disease are oropharyngeal or<br />

gastrointestinal tularemia, which is contracted by ingestion of contaminated food or<br />

water (62). Symptoms can include sore throat, diarrhea, and even laceration of the<br />

bowel.<br />

The most common presentation of tularemia is classified as ulceroglandular,<br />

where the bacterium is introduced through the skin, usually from a tick bite (126, 167)<br />

or from an abrasion during skinning of an infected animal. This accounts for 90% of<br />

tularemia cases (166). A characteristic ulcer forms at the site of infection. The<br />

bacteria then move to the draining lymph nodes, which can lead to lymphadenopathy,<br />

similar to the bubo observed in Plague patients. From the lymph nodes the bacteria<br />

can spread systemically to the reticuloendothelial organs, colonizing the lung, liver,<br />

spleen, bone marrow, and blood. Symptoms of tularemia may include fever, chills,<br />

16


headaches, diarrhea, muscle aches, joint pain, dry cough, and progressive weakness<br />

(29, 45). This form of the disease is rarely fatal, with a mortality rate of less than 3%<br />

of cases (45), although it may take time to resolve without treatment. However, an<br />

acute form of the disease, typhoidal tularemia, can occur with ssp. tularensis infection,<br />

which is characterized by septicemia without the appearance of an ulcer or bubo and<br />

carries a mortality rate between 30% and 60% (45, 53). Although tularemia can be a<br />

severe and debilitating disease, in most cases it can be effectively treated with<br />

antibiotics. Streptomycin and gentamicin are the first drugs of choice, though<br />

chloramphenicol, tetracycline, and ciprofloxacin have also been used to effectively<br />

treat tularemia. Antibiotic treatment has reduced the overall case fatality to<br />

approximately 2% (45, 142).<br />

Intracellular lifestyle of F. tularensis<br />

Tularemia is characterized by both significant extracellular and intracellular<br />

bacterial phases of F. tularensis. F. tularensis is a facultative intracellular pathogen,<br />

whose primary intracellular niche in the mammalian host is the macrophage, though it<br />

has been shown to replicate inside dendritic cells, hepatocytes, neutrophils, and type II<br />

alveolar epithelial cells in the host (20, 58, 93, 106, 179). Intracellular replication is<br />

crucial to F. tularensis pathogenesis as mutants that fail to replicate intracellularly are<br />

avirulent in mice. Because F. tularensis is found primarily in macrophages in vivo<br />

and because the macrophage represents an important mediator of host defense, the<br />

17


macrophage serves as the most widely used in vitro model to study F. tularensis<br />

infection.<br />

F. tularensis is engulfed by the macrophage via a spacious psuedopod loop<br />

mechanism (30). This mechanism is visually and mechanistically distinct from<br />

conventional, coiling, or ruffling phagocytosis that is observed with other pathogens<br />

and inert particles. Efficient bacterial uptake depends on complement factor C3 and<br />

the corresponding complement receptor (CR3) (30, 148). The mannose receptor (MR)<br />

and scavenger receptor class A (SRA) also play a minor role in bacterial uptake (10,<br />

135, 148). The PI3K pathway does not seem to play a significant role in uptake, but<br />

the Syk/Erk axis appears to play a role through an undefined mechanism (132, 133).<br />

Lipid rafts also play a role in bacterial entry into macrophages and may determine the<br />

biogenesis of the initial phagosome that contains F. tularensis (165).<br />

After uptake, F. tularensis initially resides in a membrane bound vacuole<br />

termed the Francisella-containing Phagosome (FCP). The FCP acquires the early<br />

endosomal antigen 1 (EEA1) within 5 minutes of uptake, but this marker rapidly<br />

dissociates from the phagosome followed by the acquisition of late endosomal markers<br />

Lamp1, Lamp2, and the Rab7 GTPase within 15-30 minutes (26, 31, 145). The FCP<br />

does not significantly fuse with lysosomes and though not absolutely required,<br />

acidification of the phagosomal acts as a cue for F. tularensis to escape the phagosome<br />

into the host cell cytosol (28, 143). Phagosomal escape is rapid, occurring within 60<br />

minutes of macrophage infection. Once in the cytosol F. tularensis replicates to high<br />

numbers. Both phagosomal escape and intracellular replication are mediated by a<br />

locus of F. tularensis genes known as the Francisella Pathogenicity Island (FPI) (122).<br />

18


FPI mutants remain in the initial phagosome, which progresses to lysosomes (14). FPI<br />

mutants are also avirulent in vivo (18, 177). At late stages of infection F. tularensis<br />

can be found again inside a membrane bound vacuole that exhibits characteristics of<br />

autophagosomes (26). The role of autophagy in F. tularensis pathogenesis is<br />

unknown, but F. tularensis downregulates several host proteins required for the<br />

formation of autophagosomes (22).<br />

Innate immunity to F. tularensis<br />

The intracellular lifestyle of F. tularensis brings it in contact with distinct<br />

environments of the macrophage, namely the surface during uptake, the initial<br />

phagosome, and the host cell cytosol. As discussed earlier, the macrophage is armed<br />

with innate immune defenses in each of these compartments that can respond to the<br />

presence of F. tularensis. Furthermore, there is crosstalk between signaling pathways<br />

that link sensing in one compartment to innate responses in other compartments.<br />

Although complement factors are able to kill a number of bacteria, F.<br />

tularensis species are resistant to direct killing by complement. It was initially<br />

thought that complement protection was derived from the presence of a bacterial<br />

capsule (67), although the existence of capsule has been difficult to demonstrate<br />

biochemically. Moreover, F. novicida and F. philomiragia, which lack an apparent<br />

capsule, are also resistant to serum killing (43).<br />

At the macrophage surface, host TLRs engage F. tularensis. Unlike LPS from<br />

enteric bacteria, F. tularensis LPS is only mildly inflammatory and stimulates a low<br />

19


level of proinflammatory cytokine production (57). This likely is due to the unique<br />

structure of F. tularensis lipid A. Unlike the hexa-acylated lipid A from E. coli and<br />

other gram-negative enterics, lipid A from F. tularensis is tetra-acylated (139). It is<br />

thought that this altered structure makes it unrecognizable to LPS-binding protein, and<br />

therefore subverts TLR4 recognition (11, 33, 57). Consistent with this observation,<br />

TLR4 deficient mice are not more susceptible to infection with F. tularensis than<br />

WILD-TYPE mice (27, 36). F. tularensis does significantly stimulate TLR2 signaling<br />

resulting in proinflammatory cytokine production, and intracellular bacteria colocalize<br />

with TLR2 and MyD88 (35, 83, 94). In vivo, TLR2 does not seem to play a role in<br />

host protection during intradermal challenge, but is important in an intranasal model of<br />

infection (36, 94). However, Myd88 deficient mice are completely susceptible to<br />

sublethal doses of LVS when given intradermally (36).<br />

After phagosomal escape F. tularensis is subject to cytosolic innate immune<br />

recognition. In the cytosol, F. tularensis activates a cytosolic surveillance pathway<br />

that is characterized by the production of type-I IFNs (64). The cytosolic sensor<br />

responsible for type-I IFN production is unknown. Autocrine and paracrine signaling<br />

through IFNAR leads to inflammasome activation, resulting in release of mature IL-<br />

1β and IL-18, and host cell death (64). Inflammasome activation is critical to host<br />

defense in vivo as mice lacking inflammasome components have increased bacterial<br />

burden and succumb to infection much faster than wild-type mice (97). The NLR<br />

involved in sensing cytosolic F. tularensis also remains a mystery. Interestingly, an in<br />

vivo genetic screen identified two proteins FTT0584 and FTT0748 that modulate<br />

inflammasome activation by an unknown mechanism (177). Mutants in both of these<br />

20


genes have reduced fitness in vivo. Further studies on the cytosolic stage of F.<br />

tularensis are required to elucidate mechanisms of bacterial virulence and host<br />

defense.<br />

Several features make tularemia an ideal model for studies of innate immunity<br />

and host pathogen interactions. First, tularemia is an acute disease, with the host<br />

either resolving the infection or succumbing to the infection before significant onset of<br />

the adaptive response. Second, F. tularensis is one of only five currently described<br />

bacterial pathogens that reside in the host cell cytosol, and the only one that does not<br />

use actin-based motility for cell-to-cell spread. Lastly, F. tularensis lacks homologs of<br />

a number of canonical toxins, secretion systems, virulence factors, and even PAMPs<br />

possessed by other mammalian pathogens. This provides the opportunity to discover<br />

new bacterial PAMPs, and new mechanisms of virulence and pathogen recognition.<br />

In my thesis work I set out to elucidate the molecular mechanisms of the cytosolic<br />

response to F. tularensis.<br />

21


<strong>Chapter</strong> 2: A genetic screen identifies novel F.<br />

novicida genes that modulate the cytosolic innate<br />

immune responses in macrophages<br />

Jonathan W. Jones 1 and Denise M. Monack 1<br />

1 Department of Microbiology and Immunology, <strong>Stanford</strong> <strong>University</strong> School of<br />

Medicine, <strong>Stanford</strong>, CA 94305, USA.<br />

22


2.1 CHAPTER 2 SUMMARY<br />

Francisella tularensis is a facultative intracellular pathogen that replicates in<br />

the macrophage cytosol. Macrophages respond to the presence of cytosolic F.<br />

tularensis with the production of type-I IFN, and activation of the inflammasome,<br />

which leads to release of proinflammatory cytokines and host cell death. Little is<br />

known about the bacterial ligands that induce the macrophage cytosolic reponse, or<br />

virulence mechanisms the bacteria use to subvert this response. We wanted to identify<br />

bacterial factors that modulate the cytosolic response in macrophages.<br />

We conducted a forward genetic screen to identify F. novicida genes that<br />

resulted in increased or decreased type-I IFN production and host cell death in<br />

macrophages. In total, we identified 236 genes with an altered cytosolic response.<br />

Deletions of the outer membrane protein fopA as well as several genes involved in<br />

LPS/capsule/cell wall resulted in hyper-stimulation of the cytosolic responses. This<br />

phenotype was partially due to an altered membrane surface that exposed TLR2-<br />

stimulating PAMPs. Deletions of genes in the Francisella Pathogenicity Island, as<br />

well as those involved in purine and pyrimidine biosynthesis were crucial for<br />

intracellular replication and induction of the cytosolic response in macrophages.<br />

We further demonstrate that cytosolic localization is necessary but not<br />

sufficient to induce the cytsolic response. Bacterial protein and DNA synthesis are<br />

required to induce the host response, but bacterial replication is dispensible. Finally,<br />

we show that the host response to cytosolic F. novicida shares characteristics with the<br />

response to transfected dsDNA and we hypothesize that recognition of cytosolic F.<br />

23


novicida DNA mediates type-I IFN production and inflammasome activation in<br />

response to infection. Bacterial mutant that hyper-stimulate this pathway may release<br />

more dsDNA through intracellular lysis.<br />

24


2.2 INTRODUCTION<br />

The innate immune system is the first line of defense against invading<br />

microbes. As part of their arsenal, immune effector cells have evolved a number of<br />

sensors, termed pattern recognition receptors (PRRs) that can detect conserved<br />

microbial patterns or pathogen associated molecular patterns (PAMPs) derived form<br />

microbes. These PAMPs are not unique to any given pathogen, but are usually<br />

structural components of microbial cell wall, and nucleic acids that are evolutionarily<br />

constrained and difficult for microbes to significantly alter to avoid detection. Instead,<br />

many host-adapted pathogens have evolved strategies to modulate host sensing, either<br />

by using bacterial proteins to mask PAMPs or by secreting effector proteins that target<br />

and disrupt host-signaling pathways. This constant interplay of host sensors and<br />

microbial PAMPs and effectors determines the outcome of the infection<br />

As pathogens have evolved to survive in various locations in the cell, so too<br />

have PRRs evolved that reside in and survey diverse cellular compartments. The Toll-<br />

like receptors (TLRs) are located are located on the macrophage surface and in<br />

endosomes, and respond to conserved molecules like bacterial LPS, lipoproteins,<br />

nucleic acids, and flagellin. The TLRs mediate the production of proinflammatory<br />

cytokines through the adaptors MyD88 and Trif. In the cytosol there are several<br />

groups of sensors that mediate pathogen recognition. The RNA helicases, RIG-I and<br />

MDA5, respond to viral RNA and trigger the “antiviral state” characterized by the<br />

production of type-I IFNs through the adaptor MAVS. NOD1 and NOD2 detect<br />

bacterial peptidoglycan fragments and initiate proinflammatory cytokine production<br />

25


through the adaptor RIP2. The cytosol is also equipped with several Nod-like<br />

receptors (NLRs) that initiate the formation of a multiprotein complex containing ASC<br />

and caspase-1, termed the inflammasome. Various stimuli such as pore-forming<br />

toxins, peptidoglycan fragments, and flagellin trigger inflammasome activation.<br />

Inflammasome activation leads to autocatalytic cleavage of capsase-1, which leads to<br />

processing and release of mature IL-1β and IL-18, and host cell death termed<br />

pyroptosis.<br />

Francisella tularensis is an intracellular pathogen that parasitizes host<br />

macrophages. Upon uptake into host macrophages the bacterium initially resides in a<br />

membrane bound vacuole. However, it rapidly escapes the phagosome and resides in<br />

the macrophage cytosol where it can replicate to high numbers. Lipid A from many<br />

enteric pathogens stimulates TLR4 signaling at the host surface. F. tularensis<br />

possesses a unique lipid A that is not recognized by TLR4. Instead TLR2 is involved<br />

in sensing F. tularensis at the surface and in the vacuole, and leads to the production<br />

of proinflammatory cytokines like proIL-1β and TNF-α. Macrophages respond to the<br />

presence of cytosolic F. tularensis with the production of type-I IFNs in an IRF3<br />

dependent but TLR-, RIG-I, MDA-5, and NOD-independent manner. This response is<br />

reminiscent of the host response to cytosolic nucleic acids. Autocrine and paracrine<br />

signaling of type-I IFN through the type-I IFN receptor (IFNAR) is crucial to<br />

subsequent activation of an ASC-containing inflammasome, which leads to release of<br />

mature IL-1b and IL-18 and host cell death. We refer to the combination of the type-I<br />

IFN response and the macrophage cell death response as the cytosolic responses.<br />

None of the known inflammasome Nod-like receptors (NLRs) play a role in<br />

26


inflammasome activation in response to F. tularensis. The bacterial ligand(s) and host<br />

receptor(s) responsible for recognizing cytosolic F. tularensis are unknown.<br />

Furthermore, there is evidence that F. tularensis has mechanisms to dampen the host<br />

response and delay the activation of the inflammasome (177).<br />

In order to elucidate the mechanisms by which F. tularensis modulates the<br />

cytosolic responses we performed a forward genetic screen of two F. novicida<br />

transposon libraries to identify mutants that induced increased to decreased type-I IFN<br />

and cell death responses from macrophages. We found that a gene encoding a major<br />

F. tularensis outer membrane protein, fopA, as well mutants in genes involved in<br />

LPS/capsule/cell wall synthesis were responsible for dampening the cytosolic response<br />

to F. tularensis. We also replicated previous findings that mutants that failed to<br />

escape the phagosome do not trigger the cytosolic responses. Further study<br />

demonstrated that cytosolic localization is necessary but not sufficient to induce the<br />

cytosolic responses. Finally, we show that induction of the cytosolic response by F.<br />

novicida does not require bacterial replication, but is dose dependent and shares many<br />

characteristics with the cytosolic response dsDNA.<br />

27


2.3 RESULTS<br />

2.3.1 A transposon screen identifies fopA as a suppressor of the macrophage<br />

cytosolic innate immune response.<br />

To identify bacterial genes involved in modulating the cytosolic response to F.<br />

tularensis, we screened a F. novicida transposon library (177) for mutants that<br />

exhibited enhanced or diminished type-I IFN and cell death responses upon infection<br />

in macrophages. Approximately 10,000 individual transposon insertion mutants were<br />

used to infect Pam3CSK pre-stimulated bone marrow-derived macrophages (BMDMs)<br />

in 96-well plates. The amount of type-IFN in macrophage supernatant was measured<br />

by a type-I IFN reporter cell line that produces luciferase in response to type-I IFN<br />

(ISRE-L929) (79). To confirm the initial results from the ISRE screen we infected<br />

Pam3CSK pre-stimulated BMDMs with individual transposon insertion mutants from<br />

the F. novicida transposon library (177) and measured the macrophage IFN-β<br />

response by quantitative RT-PCR. Transposon insertions in the pdpA and clpB genes<br />

were attenuated in inducing and IFN-β response in macrophages similar to the control<br />

ΔFPI mutant (Fig. 1A). These results are consistent with previous reports that<br />

demonstrate that the ΔpdpA mutant is attenuated for phagosomal escape and therefore<br />

would not induce the cytosolic response. A ΔclpB mutant is attenuated for intracellular<br />

replication in BMDMs and virulence in mice (55, 111). Furthermore, ClpB displays<br />

homology to ClpV proteins from type VI secretion systems, and we speculate that it<br />

may play a role in the function of the FPI, which would make a ΔclpB mutant<br />

attenuated for phagosomal escape, though this has not been formally shown. These<br />

28


esults were proof of principle that our screen would identify F. novicida genes that<br />

modulate the cytosolic responses.<br />

Surprisingly, library mutant Tn96G6, which was identified as a transposon<br />

insertion in the fopA gene, induced a 17-fold higher induction of IFN-β in<br />

macrophages than wild-type F. novicida (Fig. 1A), similar to a control ΔFTT0584<br />

mutant that hyper-induces the cytosolic response (177). This result was confirmed<br />

with an isogenic deletion of fopA (Fig. 1B). Consistent with the link between the<br />

type-I IFN response and the host cell death response (34, 64), a ΔfopA mutant killed<br />

77% of infected macrophages by 6hrs PI, while wild-type F. novicida induced<br />

negligible killing (Fig. 1C). We were able to complement these phenotypes by re-<br />

introducing a wild-type copy of the fopA gene in trans (Fig. 1B,C). This hyper-<br />

induction of the cytosolic responses was not due to increased replication of the<br />

Tn96G6 or ΔfopA mutants (Fig.1D). These results indicate that the fopA gene<br />

suppresses the cytosolic response to F. novicida.<br />

29


A B<br />

C D<br />

- Figure 1 - An F. novicida ΔfopA mutant hyper-induces the cytosolic innate immune response<br />

in macrophages.<br />

IFN-β mRNA was quantified by quantitative RT-PCR at 5hrs post-infection (PI) in wild-type BMDMs<br />

that were either Pam3CSK4 pre-stimulated (A) or unstimulated (B) and infected with the indicated<br />

strains of F. novicida at an MOI of 10:1. (C) Host cell death was quantified by lactate dehydrogenase<br />

(LDH) release at 6hrs PI in unstimulated BMDMs infected at an MOI of 10:1. (D) Intracellular<br />

replication was assayed by gentamicin protection assay over 8hrs in unstimulated BMDMs infected at<br />

an MOI of 1:10. Error bars for quantitative RT-PCR represent standard deviation of 3 technical<br />

30


eplicates. Error bars for cell death assay represent standard deviation of infections done in triplicate.<br />

Graphs are representative of three independent experiments.<br />

2.3.2 ΔfopA stimulates the same cytosolic surveillance pathway as wild-type F.<br />

novicida.<br />

FopA is an abundant outer membrane protein that may interact with LPS and<br />

peptidoglycan, which are considered PAMPs for TLR and NOD receptors<br />

respectively. Previous studies demonstrated that the IFN-β response to wild-type F.<br />

novicida was independent of TLR and NOD signaling, (64). However, we were<br />

curious to know if deletion of fopA would lead to an increased release of TLR and<br />

NOD ligands and a contribution of these PRRs to the cytosolic response. We<br />

investigated the roles of TLR and NOD signaling to the cytosolic response<br />

respectively by infecting myd88/trif -/- and rip2 -/- macrophages with wild-type F.<br />

novicida and ΔfopA, and measuring the IFN-β and host cell death responses. We<br />

observed only a slight decrease in IFN-β mRNA levels in rip2 -/- macrophages and no<br />

difference in myd88/trif -/- macrophages compared to wt (Fig. 2A, C). We also saw no<br />

effect on the cell death response in the absence of TLR or NOD signaling (Fig. 2D, E).<br />

This indicated that deletion of fopA did not lead to increased cytosolic recognition<br />

through release of LPS or peptidoglycan fragments.<br />

31


A B<br />

C D<br />

E<br />

32


- Figure 2 - The cytosolic response to ΔfopA is IRF3- and IFNAR- dependent but TLR- and<br />

NOD- independent.<br />

Unstimulated wt, irf3 -/- , myd88/trif -/- , ifnar -/- or rip2 -/- BMDM were infected with the indicated<br />

strains of F. novicida at an MOI of 10:1 and IFN-β mRNA was quantified by quantitative RT-PCR at<br />

5hrs PI (A, B, and C) and cell death was measured by LDH assay at 7hrs PI (D and E). Error bars for<br />

quantitative RT-PCR represent standard deviation of 3 technical replicates, while error bars for cell<br />

death represent standard deviation of triplicate infections. Graphs are representative of three<br />

independent experiments.<br />

The IFN-β response to wild-type F. novicida is dependent on IRF3 and<br />

autocrine and paracrine signaling through IFNAR acts to amplify IFN-β production<br />

(64). While the IFN-β response to ΔfopA was completely IRF3 dependent (Fig. 2A)<br />

we saw only a modest reduction in IFN-β mRNA in the absence of IFNAR signaling<br />

(Fig. 2B). Additionally, type-I IFN signaling was required for the full host cell death<br />

response to ΔfopA, although the mutant also induced higher levels of type-I IFN-<br />

independent cell death than wild-type F. novicida (Fig. 2D). Therefore we conclude<br />

that the IFNAR-dependent amplification of IFN-β observed in response to wild-type<br />

F. novicida is not required for the response to ΔfopA, possibly because fopA either<br />

suppresses the amplification loop, or because ΔfopA releases more of the IFN-β<br />

stimulating ligand, and thus amplification of the signal is not required. This leads to<br />

an increase in type-IFN dependent and independent host cell death responses by the<br />

macrophage.<br />

Cytosolic F. novicida induces an inflammasome-dependent host cell death that<br />

is triggered by recognition from an unknown NLR (97, 98). To determine if ΔfopA<br />

33


induced cell death through the same pathway as wild-type F. novicida we infected wt,<br />

asc-/-, and casp-1-/- BMDM with wild-type F. novicida or ΔfopA and compared the<br />

early and late cell death responses. The ΔfopA mutant induced host cell death in an<br />

ASC- and caspase-1-depedent manner, just like wild-type F. novicida (Fig. 3A).<br />

Furthermore, ΔfopA hyper-stimulated the ASC-dependent-caspase-1-independent late<br />

cell death observed in response to wild-type F. novicida (Fig. 3B). Finally, since<br />

NLRP3 has been implicated in the inflammasome response to peptidoglycan<br />

fragments (100) and FopA could interact with peptidoglycan in the outer membrane,<br />

we investigated if NLR3 contributed to the cell death response to ΔfopA comparing the<br />

cell death responses of wt and nlrp3 -/- BMDMs to infection with ΔfopA. Again, like<br />

wild-type F. novicida host cell death to ΔfopA was completely NLRP3-independent<br />

(Fig. 3C). These results further suggest cytosolic recognition of ΔfopA occurs through<br />

the same mechanism as wild-type F. novicida. Thus, fopA could suppress the<br />

cytosolic sensor leading to IFN-β or it could suppress release of the ligand from the<br />

bacteria.<br />

34


A B<br />

C<br />

- Figure 3- The hyper-cytotoxicity ofΔfopA is ASC- and caspase-1 dependent but NLRP3-<br />

independent.<br />

BMDM from wt, asc -/- , casp-1 -/- , or nlrp3 -/- mice were Pam3CSK4 pre-stimulated (A and B) or<br />

unstimulated (C) and infected with the indicated strains of F. novicida for 3.5hrs (A), 5.5hrs (B), or 8hrs<br />

(C) and cell death was measured by LDH assay. Means and standard deviations are shown for<br />

infections done in triplicate. Representative graphs of three independent experiments are shown.<br />

35


2.3.3 A ΔfopA mutant has reduced fitness in mice.<br />

To investigate the role of fopA in virulence we performed a competitive index<br />

experiment where we infected wt C57/BL6J mice intradermally (i.d.) with an equal<br />

ratio of wild-type F. novicida and ΔfopA mutant for two days and harvested skin and<br />

spleen to enumerate CFU. If the mutant and wild-type are equally fit in vivo we<br />

would expect a competitive index of 1. However, we observed a competitive index<br />

for ΔfopA that was significantly less than 1 (Fig. 4), indicating that this mutant has<br />

reduced fitness in mouse infections relative to wild-type F. novicida. The hyper-<br />

stimulation of the cytosolic immune responses in macrophages could contribute to the<br />

attenuation of the mutant, though we cannot rule out the possibility that a mutant in an<br />

outer membrane protein could be more susceptible to killing by complement or other<br />

antimicrobial defenses as well.<br />

36


- Figure 4 - ΔfopA is less fit in vivo than wild-type F.novicida.<br />

wt C57BL6/J mice were infected intradermally (i.d.) with 10 5 cfu of an equal ratio of wt F. novicida<br />

and ΔfopA. At 2 days post infection skin and spleen was harvested and bacterial loads were determined<br />

by plating serial dilutions of homogenized tissue on selective and non-selective MH agar.<br />

2.3.4 Identification of F. tularensis mutants that differentially induce the<br />

cytosolic response in macrophages by a genome-wide forward genetic screen.<br />

Although we had identified fopA in a screen of our transposon library, we<br />

experienced difficulty in identifying other transposon insertions by genomic<br />

sequencing. Therefore, we decided to repeat the forward genetic screen using a<br />

commercially available “two-allele” F. novicida transposon library from BEI<br />

Resources to search for mutants that induced an increased or decreased type-I IFN<br />

response and/or an increased or decreased host cell death response in bone marrow-<br />

derived macrophages. This simplified the identification of screen hits as the position<br />

37


of transposon insertion in each mutant was provided. Furthermore, we were confident<br />

that this screen gave us complete coverage of the F. novicida genome. The amount of<br />

type-I IFN in macrophage supernatants was measured by incubating culture<br />

supernatant from infected macrophages with a reporter cell line that produces<br />

luciferase in response to type-I IFNs (79). Macrophage cell death was measured by<br />

lactate dehydrogenase release assay (LDH). In total we identified 278 transposon<br />

insertions representing 236 F. novicida genes that differentially induced the cytosolic<br />

response in macrophages relative to the wild-type strain (Table 1 and 2). This<br />

represents approximately 17% of the non-essential F. novicida genome.<br />

38


- Table 1 – Mutants that hyper-induce the cytosolic responses<br />

FTN # Plate Well Gene Description Functional class<br />

FTN_0007 9 D10 hypothetical protein (predicted<br />

secretion signal, transmembrane<br />

domain)<br />

hypothetical - novel<br />

FTN_0018 26 C7 sdaC serine permease transport - amino-acid<br />

FTN_0073 6 B10 yidC conserved membrane protein of<br />

unknown function<br />

FTN_0074 2 B8 conserved hypothetical protein<br />

(signal peptide)<br />

FTN_0088 27 C4 ybgL protein of unknown function,<br />

LamB/YcsF family<br />

FTN_0137 3 H12 protein of unknown function<br />

(predicted secretion signal)<br />

39<br />

unknown function -<br />

conserved<br />

hypothetical -<br />

conserved<br />

unknown function -<br />

conserved<br />

unknown function -<br />

novel<br />

FTN_0139 22 F2 hypothetical protein hypothetical - novel<br />

FTN_0148 6 A7 hypothetical membrane protein<br />

(predicted signal sequence)<br />

FTN_0157 8 C2 membrane protein of unknown<br />

function<br />

FTN_0193 5 B5 cydA cytochrome bd-I terminal oxidase<br />

subunit I<br />

FTN_0203 19 D10 protein of unknown function<br />

(predicted secretion signal)<br />

hypothetical - novel<br />

unknown function -<br />

novel<br />

energy metabolism<br />

unknown function -<br />

novel<br />

FTN_0215 9 E8 hypothetical protein hypothetical - novel<br />

FTN_0234 6 B1 pgsA phosphatidylglycerophosphate<br />

synthetase<br />

FTN_0276 26 B2 mviN multidrug/oligosaccharidyllipid/polysaccharide<br />

(MOP)<br />

transporter<br />

fatty acids and lipids<br />

metabolism<br />

transport - drugs /<br />

antibacterial<br />

compounds<br />

FTN_0280 3 H10 hypothetical protein hypothetical - novel


FTN_0284 5 D9 prophage maintenance system<br />

killer protein (DOC)<br />

40<br />

mobile and<br />

extrachromosomal<br />

element functions -<br />

phage or plasmid<br />

related proteins<br />

FTN_0292 10 E5 protein of unknown function unknown function -<br />

novel<br />

FTN_0311 32 F2 hemK modification methylase, HemK<br />

family<br />

FTN_0330 9 C9 minD septum formation inhibitoractivating<br />

ATPase<br />

translation, ribosomal<br />

structure and<br />

biogenesis<br />

cell cycle<br />

FTN_0340 16 H7 protein of unknown function unknown function -<br />

novel<br />

FTN_0341 1 B7 protein of unknown function<br />

(predicted secretion signal)<br />

unknown function<br />

novel<br />

FTN_0345 8 B3 DNA uptake protein, SMF family transport<br />

FTN_0388 3 C7 protein of unknown function unknown function -<br />

novel<br />

FTN_0391 13 D11 LemA-like protein putative enzymes<br />

FTN_0409 17 C4 adhC Zn-dependent alcohol<br />

dehydrogenase<br />

FTN_0415 19 C3 pilA Type IV pili, pilus assembly<br />

protein<br />

energy metabolism<br />

motility, attachment<br />

and secretion structure<br />

FTN_0416 9 E11 lpxE lipid A 1-phosphatase fatty acids and lipids<br />

metabolism<br />

FTN_0444 27 C10 membrane protein of unknown<br />

function<br />

FTN_0449 11 C8 conserved protein of unknown<br />

function (predicted secretion<br />

signal)<br />

unknown function -<br />

novel<br />

unknown function -<br />

conserved


FTN_0463 9 F4 hypothetical protein hypothetical - novel<br />

FTN_0487 4 A1 30S ribosomal protein S21 translation, ribosomal<br />

structure and<br />

biogenesis<br />

FTN_0528 12 F4 lpxH UDP-2,3-diacylglucosamine<br />

hydrolase<br />

FTN_0530 8 C8 mpl UDP-N-acetylmuramate:L-alanylgamma-D-glutamyl-mesodiaminopimelate<br />

ligase<br />

FTN_0533 31 B6 drug:H+ antiporter-1 (DHA1)<br />

family protein<br />

FTN_0546 16 D9 flmK dolichyl-phosphate-mannoseprotein<br />

mannosyltransferase<br />

family protein<br />

FTN_0550 19 C9 sohB peptidase family S49 protein<br />

(predicted secretion signal,<br />

periplasmic) stalk mutant?<br />

FTN_0558 25 C3 ostA1 organic solvent tolerance protein,<br />

OstA<br />

41<br />

fatty acids and lipids<br />

metabolism<br />

cell wall / LPS /<br />

capsule<br />

transport - drugs /<br />

antibacterial<br />

compounds<br />

cell wall / LPS /<br />

capsule<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

cell wall / LPS /<br />

capsule<br />

FTN_0568 22 B3 birA birA-like protein post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

FTN_0576 21 H8 conserved protein of unknown<br />

function (predicted secretion<br />

signal)<br />

unknown function -<br />

conserved<br />

FTN_0582 9 G12 gph phosphoglycolate phosphatase putative enzymes<br />

FTN_0595 18 F1 outer membrane protein of<br />

unknown function<br />

unknown function -<br />

novel


FTN_0611 17 D6 kdsA 3-deoxy-D-manno-octulosonic<br />

acid 8-phosphate synthase<br />

42<br />

fatty acids and lipids<br />

metabolism<br />

FTN_0633 18 D4 katG peroxidase/catalase other metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_0638 24 C8 sulfate permease family protein transport<br />

FTN_0656 12 D6 Zn-dependent peptidase, M16<br />

family<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

FTN_0657 31 B5 metallopeptidase, M16 family post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

FTN_0660 29 F1 pepA cytosol aminopeptidase amino acid metabolism<br />

FTN_0672 12 G3 secA preprotein translocase, subunit A<br />

(ATPase, RNA helicase)<br />

FTN_0709 14 E11 hypothetical protein (predicted<br />

secretion signal, transmembrane<br />

domain)<br />

FTN_0714 12 G1 protein of unknown function<br />

(predicted secretion signal)<br />

motility, attachment<br />

and secretion structure<br />

hypothetical - novel<br />

unknown function -<br />

novel<br />

FTN_0742 27 D11 serB phosphoserine phosphatase amino acid metabolism<br />

- biosynthesis


FTN_0746 29 D12 alr alanine racemase amino acid metabolism<br />

- degradation,<br />

utilization,<br />

assimilation<br />

FTN_0747 13 A2 amino acid-polyamineorganocation<br />

(APC) superfamily<br />

protein<br />

43<br />

transport - amino-acid<br />

FTN_0756 4 G11 fopA OmpA family protein cell wall / LPS /<br />

capsule<br />

FTN_0757 22 D11 membrane protein of unknown<br />

function<br />

FTN_0772 22 C2 conserved protein of unknown<br />

function<br />

unknown function -<br />

novel<br />

unknown function -<br />

conserved<br />

FTN_0785 5 B6 isochorismatase family protein putative enzymes<br />

FTN_0812 17 C10 bioD dethiobiotin synthetase cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_0814 7 E11 bioF 8-amino-7-oxononanoate synthase cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_0828 13 B4 protein of unknown function<br />

(predicted secretion signal)<br />

pseudogene in Schu4<br />

FTN_0839 32 E10 conserved protein of unknown<br />

function<br />

unknown function -<br />

novel<br />

unknown function -<br />

conserved<br />

FTN_0841 8 D10 ThiJ/PfpI family protein putative enzymes<br />

FTN_0842 7 G10 aroG phospho-2-dehydro-3deoxyheptonate<br />

aldolase<br />

amino acid metabolism<br />

- biosynthesis<br />

FTN_0863 7 B2 hypothetical membrane protein hypothetical - novel<br />

FTN_0877 26 C10 cls; ybhO cardiolipin synthetase fatty acids and lipids<br />

metabolism


FTN_0901 10 C8 isomerase putative enzymes<br />

FTN_0905 4 B10 yrbI 3-deoxy-D-manno-octulosonate 8phosphate<br />

phosphatase<br />

FTN_0907 16 E6 dacD D-alanyl-D-alanine<br />

carboxypeptidase<br />

44<br />

fatty acids and lipids<br />

metabolism (lnvolved<br />

in LPS synthesis)<br />

cell wall / LPS /<br />

capsule<br />

FTN_0939 8 B10 hypothetical protein hypothetical - novel<br />

FTN_0943 8 C12 rimI ribosomal-protein-alanine<br />

acetyltransferase<br />

FTN_0967 29 G11 vanY D-alanyl-D-alanine<br />

carboxypeptidase<br />

translation, ribosomal<br />

structure and<br />

biogenesis<br />

cell wall / LPS /<br />

capsule<br />

FTN_1007 2 H8 rplY 50S ribosomal protein L25 translation, ribosomal<br />

structure and<br />

biogenesis<br />

FTN_1027 12 H7 ruvC holliday junction<br />

endodeoxyribonuclease<br />

FTN_1029 4 F11 elbB conserved protein of unknown<br />

function<br />

DNA replication,<br />

recombination,<br />

modification and<br />

repair -<br />

restriction/modificatio<br />

n<br />

unknown function -<br />

conserved<br />

FTN_1030 3 B5 lipA lipoic acid synthetase cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_1036 4 H11 protein of unknown function unknown function -<br />

conserved<br />

FTN_1037 16 E4 hypothetical membrane protein hypothetical - novel


FTN_1046 2 D9 wzb low molecular weight (LMW)<br />

phosphotyrosine protein<br />

phosphatase (exopolysaccharide<br />

production)<br />

FTN_1058 6 A6 tig trigger factor (TF) protein<br />

(peptidyl-prolyl cis/trans<br />

isomerase<br />

FTN_1064 1 A4 PhoH family protein, putative<br />

ATPase<br />

FTN_1066 11 B5 transporter-associated protein,<br />

HlyC/CorC family<br />

45<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

signal transduction and<br />

regulation<br />

transport<br />

FTN_1070 7 D7 ospD2 protein of unknown function unknown function -<br />

novel<br />

FTN_1073 30 F2 DNA/RNA endonuclease G nucleotides and<br />

nucleosides<br />

metabolism<br />

FTN_1074 4 H10 X-prolyl aminopeptidase 2 post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

FTN_1087 8 G1 cynT carbonic anhydrase other metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_1093 18 A5 protein of unknown function unknown function -<br />

novel<br />

FTN_1104 11 H2 hypothetical protein (predicted<br />

secretion signal)<br />

hypothetical - novel<br />

FTN_1109 16 C7 rhodanese-like family protein putative enzymes<br />

FTN_1123 13 B6 conserved hypothetical protein hypothetical -<br />

conserved<br />

FTN_1133 12 E4 protein of unknown function unknown function -<br />

novel


FTN_1135 10 B10 aroB 3-dehydroquinate synthetase amino acid metabolism<br />

- biosynthesis<br />

FTN_1146 2 B7 aspC2 aspartate aminotransferase other metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_1148 4 B11 glycoprotease family protein post-translational<br />

modification, protein<br />

turnover, chaperones<br />

FTN_1149 5 A9 nagA N-acetylglucosamine-6-phosphate<br />

deacetylase<br />

FTN_1154 13 B3 type I restriction-modification<br />

system, subunit S<br />

46<br />

carbohydrate<br />

metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

DNA replication,<br />

recombination,<br />

modification and<br />

repair - repair<br />

FTN_1187 27 E6 protein of unknown function unknown function -<br />

conserved<br />

FTN_1212 28 A1 glycosyl transferase, group 1 cell wall / LPS /<br />

capsule<br />

FTN_1213 3 D7 glycosyl transferase, family 2 cell wall / LPS /<br />

capsule<br />

FTN_1214 19 B7 glycosyl transferase, family 2 cell wall / LPS /<br />

capsule<br />

FTN_1215 27 C6 kpsC capsule polysaccharide export<br />

protein KpsC<br />

FTN_1217 19 G1 ATP-binding cassette (ABC)<br />

superfamily protein<br />

cell wall / LPS /<br />

capsule<br />

transport<br />

FTN_1218 7 C4 glycosyl transferase, group 1 cell wall / LPS /<br />

capsule<br />

FTN_1219 30 C10 galE UDP-glucose 4-epimerase carbohydrate<br />

metabolism -<br />

degradation,<br />

utilization,<br />

assimilation


FTN_1221 8 E6 rpe D-ribulose-phosphate 3-epimerase energy metabolism<br />

FTN_1223 10 C4 conserved hypothetical membrane<br />

protein DoxD like family protein<br />

FTN_1239 10 D7 5-formyltetrahydrofolate<br />

cycloligase<br />

47<br />

hypothetical -<br />

conserved<br />

putative enzymes<br />

FTN_1240 24 C11 BolA family protein cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_1241 14 F10 DedA family protein (predicted<br />

secretion signal)<br />

putative enzymes<br />

FTN_1242 18 B12 DedA family protein putative enzymes<br />

FTN_1253 25 C10 lpcC glycosyl transferase, group 1 cell wall / LPS /<br />

capsule<br />

FTN_1254 2 G7 protein of unknown function unknown function -<br />

novel<br />

FTN_1255 5 C11 glycosyl transferase, family 8 cell wall / LPS /<br />

capsule<br />

FTN_1256 7 B6 membrane protein of unknown<br />

function<br />

unknown function -<br />

novel<br />

FTN_1273 32 A7 fadD1 long chain fatty acid CoA ligase other metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_1296 2 F3 yhbH sigma54 modulation protein translation, ribosomal<br />

structure and<br />

biogenesis<br />

FTN_1330 10 C5 pyk pyruvate kinase carbohydrate<br />

metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_1368 14 E7 feoA Fe2+ transport system protein A transport<br />

FTN_1369 3 F7 protein of unknown function (cell<br />

division protein)<br />

unknown function -<br />

novel


FTN_1404 11 B6 yadH ATP-binding cassette (ABC)<br />

superfamily protein<br />

48<br />

transport<br />

FTN_1417 31 F11 manB phosphomannomutase carbohydrate<br />

metabolism -<br />

biosynthesis<br />

FTN_1418 5 C3 manC mannose-1-phosphate<br />

guanylyltransferase<br />

FTN_1421 6 C3 wbtH glutamine<br />

amidotransferase/asparagine<br />

synthase<br />

cell wall / LPS /<br />

capsule<br />

amino acid metabolism<br />

- biosynthesis<br />

FTN_1423 3 A12 wbtG glycosyl transferase, group 1 cell wall / LPS /<br />

capsule<br />

FTN_1426 20 H10 wbtE UDP-glucose/GDP-mannose<br />

dehydrogenase family protein<br />

cell wall / LPS /<br />

capsule<br />

FTN_1427 1 B5 wbtD glycosyl tranferase, group 1 cell wall/ LPS/ capsule<br />

FTN_1428 3 B11 wbtO transferase cell wall / LPS /<br />

capsule<br />

FTN_1429 21 B1 wbtP galactosyl transferase cell wall / LPS /<br />

capsule<br />

FTN_1431 6 G3 wbtA dTDP-glucose 4,6-dehydratase cell wall / LPS /<br />

capsule<br />

FTN_1452 12 E7 two-component response<br />

regulator<br />

FTN_1470 12 B3 ispA geranyl diphosphate<br />

synthase/farnesyl diphosphate<br />

synthase<br />

FTN_1472 17 D5 conserved protein of unknown<br />

function<br />

signal transduction and<br />

regulation<br />

cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

unknown function -<br />

conserved<br />

FTN_1476 26 A3 protein of unknown function unknown function -<br />

novel


FTN_1496 8 F10 coaE dephospho-CoA kinase cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_1500 30 C5 protein of unknown function unknown function -<br />

novel<br />

FTN_1504 20 D8 glucokinase regulatory protein signal transduction and<br />

regulation<br />

FTN_1515 3 F5 hypothetical membrane protein hypothetical - novel<br />

FTN_1533 21 A7 conserved protein of unknown<br />

function<br />

49<br />

unknown function -<br />

conserved<br />

FTN_1537 16 H9 hypothetical protein hypothetical - novel<br />

FTN_1548 9 A1 yfgL conserved protein of unknown<br />

function<br />

unknown function -<br />

conserved<br />

FTN_1602 4 E5 deoB phosphopentomutase other metabolism -<br />

degradation,<br />

utilization,<br />

assimilation<br />

FTN_1603 10 E6 regulatory factor, Bvg accessory<br />

factor family<br />

FTN_1613 12 D11 peptidase, U61 family (predicted<br />

secretion signal, peptidoglycan<br />

recycling)<br />

signal transduction and<br />

regulation<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein degradation<br />

FTN_1616 12 E11 protein of unknown function unknown function -<br />

novel<br />

FTN_1617 20 A1 qseC two-component regulator, sensor<br />

histidine kinase<br />

FTN_1630 21 C1 secG preprotein translocase, subunit G,<br />

membrane protein<br />

signal transduction and<br />

regulation<br />

motility, attachment<br />

and secretion structure<br />

FTN_1653 13 A11 hypothetical membrane protein hypothetical - novel<br />

FTN_1656 16 B11 conserved hypothetical protein hypothetical -<br />

conserved


FTN_1665 6 C4 magnesium chelatase cofactors, prosthetic<br />

groups, electron<br />

carriers metabolism<br />

FTN_1686 26 D10 hypothetical membrane protein hypothetical - novel<br />

FTN_1704 20 B4 pcm protein-L-isoaspartate Omethyltransferase<br />

50<br />

post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

protein modification<br />

FTN_1732 11 D10 Mg-dependent Dnase DNA replication,<br />

recombination,<br />

modification and<br />

repair - degradation<br />

FTN_1734 13 E11 protein of unknown function<br />

(predicted secretion signal)<br />

unknown function -<br />

novel<br />

FTN_1735 17 C9 protein of unknown function unknown function -<br />

novel<br />

FTN_1757 4 D5 D-isomer specific 2-hydroxyacid<br />

dehydrogenase<br />

FTN_1766 19 C6 drug/metabolite transporter<br />

(DMT) superfamily protein<br />

energy metabolism<br />

transport - drugs /<br />

antibacterial<br />

compounds<br />

intergenic 19 B4<br />

intergenic 21 A1<br />

intergenic 25 G11<br />

intergenic 26 H7<br />

intergenic 3 H11<br />

13 B5 intergenic<br />

3 B3 isftu3 isftu3 IS element<br />

16 D10 isftu2 isftu2 IS element<br />

30 B7 isftu6 isftu6 IS element


- Table 2 – Mutants that hypo-induce the cytosolic responses<br />

FTN # Plate Well Gene Description Functional class<br />

FTN_0019 29 A2 pyrB aspartate carbamoyltransferase post-translational<br />

modification, protein<br />

turnover, chaperones -<br />

chaperones<br />

FTN_0020 8 D3 carB carbamoyl-phosphate synthase<br />

large chain<br />

FTN_0021 30 D8 carA carbamoyl-phosphate synthase<br />

small chain<br />

51<br />

signal transduction and<br />

regulation<br />

hypothetical - novel<br />

FTN_0024 14 E2 pyrC dihydroorotase amino acid metabolism<br />

FTN_0035 29 E4 pyrF orotidine-5'-phosphate<br />

decarboxylase<br />

transport - drugs /<br />

antibacterial<br />

compounds<br />

FTN_0036 32 G12 pyrD dihydroorotate oxidase nucleotides and<br />

nucleosides<br />

metabolism<br />

FTN_0085 16 F12 uspA universal stress protein nucleotides and<br />

nucleosides<br />

FTN_0108 16 B12 trmU tRNA(5-methylaminomethyl-2thiouridylate)<br />

methyltransferase<br />

FTN_0124 11 G7 ssb single-strand DNA binding<br />

protein<br />

FTN_0140 23 C7 ABC-type anion transport system,<br />

duplicated permease component<br />

FTN_0156 1 G2 plsC 1-acylglycerol-3-phosphate<br />

acyltransferase<br />

FTN_0266 16 F2 htpG chaperone Hsp90, heat shock<br />

protein HtpG<br />

FTN_0289 2 D1 proQ activator of osmoprotectant<br />

transporter ProP<br />

FTN_0336 21 C9 hypothetical protein (predicted<br />

secretion signal, transmembrane<br />

domain)<br />

metabolism<br />

nucleotides and<br />

nucleosides<br />

metabolism<br />

amino acid metabolism<br />

- degradation,<br />

utilization,<br />

assimilation<br />

nucleotides and<br />

nucleosides<br />

metabolism<br />

transport - drugs /<br />

antibacterial<br />

compounds<br />

energy metabolism<br />

transport<br />

transport - amino-acid<br />

FTN_0343 32 E7 aminotransferase putative enzymes


FTN_0357 25 G7 pal peptidoglycan-associated<br />

lipoprotein, OmpA family<br />

FTN_0419 8 E12 purM phosphoribosylformylglycinamide<br />

cyclo-ligase<br />

FTN_0420 6 E2 SAICAR<br />

synthetase/phosphoribosylamineglycine<br />

FTN_0422 32 H9 purE N5-carboxyaminoimidazole<br />

ribonucleotide mutase<br />

52<br />

translation, ribosomal<br />

structure and<br />

biogenesis<br />

putative enzymes<br />

hypothetical -<br />

conserved<br />

transport<br />

FTN_0504 5 B11 cadA lysine decarboxylase transport<br />

FTN_0529 29 C4 pyrE orotate phosphoribosyltransferase transport<br />

FTN_0535 4 B4 drug:H+ antiporter-1 (DHA1)<br />

family protein<br />

FTN_0540 5 E6 pckA phosphoenolpyruvate<br />

carboxykinase<br />

fatty acids and lipids<br />

metabolism<br />

unknown function -<br />

conserved<br />

FTN_0570 7 A9 perM PerM family protein motility, attachment<br />

and secretion structure<br />

FTN_0571 12 C12 xasA amino acid-polyamineorganocation<br />

(APC) superfamily<br />

protein<br />

transport<br />

FTN_0604 25 G6 AMP-binding protein unknown function -<br />

novel<br />

FTN_0616 30 H2 rumA RNA methyltransferase, trmA<br />

family<br />

putative enzymes<br />

FTN_0618 21 F11 glk1 ROK family protein translation, ribosomal<br />

structure and<br />

FTN_0654 29 D6 conserved hypothetical membrane<br />

protein<br />

FTN_0728 14 H4 predicted Co/Zn/Cd cation<br />

transporter<br />

FTN_0737 22 A12 potI ATP-binding cassette putrescine<br />

uptake system, membrane protein,<br />

subunit I<br />

FTN_0738 27 G12 potH ATP-binding cassette putrescine<br />

uptake system, membrane protein,<br />

subunit H<br />

biogenesis<br />

cell wall / LPS /<br />

capsule<br />

motility, attachment<br />

and secretion structure<br />

unknown function -<br />

conserved<br />

cell wall / LPS /<br />

capsule<br />

FTN_0818 12 H9 lipase/esterase translation, ribosomal<br />

structure and<br />

FTN_0915 4 E1 yqeY conserved protein of unknown<br />

function (cytosolic)<br />

biogenesis<br />

transport - drugs /<br />

antibacterial<br />

compounds


FTN_0946 27 D6 pilF Type IV pili, pilus assembly<br />

protein<br />

FTN_1013 22 G3 monovalent cation:proton<br />

antiporter family protein<br />

FTN_1022 27 D4 protein of unknown function<br />

(possible transcriptional regulator<br />

arsR)<br />

53<br />

signal transduction and<br />

regulation<br />

putative enzymes<br />

unknown function -<br />

novel<br />

FTN_1034 7 B1 rnfB iron-sulfur cluster-binding protein unknown function -<br />

novel<br />

FTN_1063 9 H6 yleA tRNA-methylthiotransferase<br />

MiaB protein<br />

unknown function -<br />

novel<br />

FTN_1112 11 H4 cphA cyanophycin synthetase hypothetical -<br />

conserved<br />

FTN_1137 3 A2 pilQ type IV pili secretion component hypothetical - novel<br />

FTN_1199 20 G10 conserved protein of unknown<br />

function (predited secretion<br />

signal, transmembrane domain)<br />

hypothetical -<br />

conserved<br />

FTN_1201 10 C1 capB capsule biosynthesis protein CapB hypothetical - novel<br />

FTN_1264 31 C11 rluD ribosomal large subunit<br />

pseudouridine synthase D<br />

FTN_1275 32 D6 emrB drug:H+ antiporter-1 (DHA2)<br />

family protein<br />

FTN_1290 24 E4 mglA macrophage growth locus, protein<br />

A<br />

unknown function -<br />

novel<br />

unknown function -<br />

conserved<br />

unknown function -<br />

novel<br />

FTN_1298 17 A7 trmE GTPase of unknown function hypothetical - novel<br />

FTN_1309 5 F4 pdpA protein of unknown function hypothetical novel<br />

FTN_1310 18 A9 pdpB protein of unknown function unknown function -<br />

novel<br />

FTN_1311 9 F12 iglE protein of unknown function unknown function -<br />

novel<br />

FTN_1312 23 H4 vgrG conserved hypothetical protein unknown function -<br />

conserved<br />

FTN_1313 9 D6 iglF hypothetical protein unknown function -<br />

conserved<br />

FTN_1314 20 A6 iglG conserved hypothetical protein unknown function -<br />

novel<br />

FTN_1314 23 F2 iglF hypothetical protein cell wall / LPS /<br />

capsule


FTN_1315 14 G8 iglH protein of unknown function signal transduction and<br />

regulation<br />

FTN_1316 21 A4 dotU protein of unknown function other metabolism -<br />

biosynthesis<br />

FTN_1317 2 B6 iglI protein of unknown function transport - drugs /<br />

antibacterial<br />

compounds<br />

FTN_1318 20 F8 iglJ hypothetical protein translation, ribosomal<br />

structure and<br />

biogenesis<br />

FTN_1319 1 G9 pdpC hypothetical protein transport -<br />

carbohydrates (sugars,<br />

FTN_1321 27 E5 iglD intracellular growth locus protein<br />

D<br />

FTN_1322 14 G2 iglC intracellular growth locus protein<br />

C<br />

FTN_1323 10 G2 iglB intracellular growth locus protein<br />

B<br />

FTN_1324 5 C8 iglA intracellular growth locus protein<br />

A<br />

FTN_1325 2 F2 pdpD protein of unknown function<br />

FTN_1422 28 B6 wbtN glycosyl transferase, group 1<br />

FTN_1465 24 H5 pmrA two-component response<br />

regulator<br />

FTN_1518 21 F10 relA GDP pyrophosphokinase/GTP<br />

pyrophosphokinase<br />

FTN_1521 17 F8 10 TMS drug/metabolite exporter<br />

protein<br />

FTN_1559 4 A10 rplS 50S ribosomal protein L19<br />

FTN_1586 9 B2 sugar transporter, MFS<br />

superfamily<br />

FTN_1699 6 D2 purL phosphoribosylformylglycinamide<br />

synthase<br />

FTN_1743 11 F3 clpB chaperone clpB<br />

intergenic 28 E5<br />

8 H2<br />

54<br />

polysaccharides)<br />

nucleotides and<br />

nucleosides<br />

metabolism<br />

post-translational<br />

modification, protein<br />

turnover, chaperones<br />

intergenic


We identified 72 F. novicida mutants that resulted in a decreased macrophage<br />

cytosolic response (Table 2). The 72 gene insertions included several known F.<br />

tularensis virulence factors such as the transcriptional regulators mglA and pmrA and<br />

17 of the genes in the FPI. Mutants in the FPI and its transcriptional regulators fail to<br />

escape the phagosome and therefore fail to induce the cytosolic response in<br />

macrophages and are attenuated in a mouse model of tularemia (Brotcke et al., 2006,<br />

Henry et al., 2007, Weiss et al. 2007). Interestingly, two genes located in the FPI,<br />

pdpE and anmK, were not identified in our screen. Further analysis of these mutants<br />

showed that neither ΔpdpE nor ΔanmK were attenuated for intracellular replication in<br />

RAW267.4 macrophages, thus we surmise that these mutants escape from the<br />

phagosome as efficiently as wild-type. Furthermore, these mutants induced<br />

macrophage cell death as efficiently as wild type. In total, these results validated our<br />

screening method. We also identified 164 mutants that hyper-induced the macrophage<br />

cytosolic response, resulting in increased kinetics of type-I IFN secretion and<br />

macrophage cytotoxicity. These mutants fell into several functional categories<br />

including proteins of unknown function, transport, post-translational modification,<br />

fatty acid and lipid metabolism, and cell wall/LPS/capsule. This result suggests that<br />

either the induction of the cytosolic response is the result of the interaction of multiple<br />

bacterial factors with the host cytosolic surveillance pathway, or that an essential<br />

bacterial molecule is responsible for induction of the cytosolic response. Furthermore,<br />

we observed a positive correlation of the type-I IFN response and macrophage<br />

cytotoxicity with all mutants identified except one. A ΔkdsA mutant induced higher<br />

55


levels of type-I IFN relative to wild type but induced lower levels of macrophage<br />

cytotoxicity. Taken together, these results suggest that several F. tularensis gene<br />

products modulate the host cytosolic response and these genes fall into several<br />

functional categories. Moreover, the macrophage type-I IFN response and cell death<br />

is tightly linked and positively correlated upon infection with F. tularensis such that<br />

the kinetics of type-I IFN production determine the kinetics of host cell death.<br />

2.3.5 F. novicida LPS mutants hyper-induce the cytosolic responses.<br />

We chose to focus on genes characterized as LPS/capsule/cell wall because<br />

several clusters of these genes were identified in our screen. Furthermore, wild-type<br />

F. novicida LPS does not activate TLR4-dependent signaling, so we hypothesized that<br />

characterizing mutants with altered LPS/capsule/cell wall that hyper stimulate immune<br />

responses might lead us to the ligand(s) and the molecular mechanism of the cytosolic<br />

response. Therefore we generated clean deletion mutants in FTN_1212, lpcC, wbtA,<br />

kdsA, and lpxH and measured type-I IFN induction and host cell death during<br />

macrophage infections. ΔFTN_1212, ΔlpcC, ΔwbtA, ΔkdsA, and ΔlpxH induced IFN-<br />

β transcript and protein secretion in macrophages with increased magnitude and<br />

kinetics relative to wild-type F. novicida (Fig. 5A-D). Additionally, all of these<br />

mutants induced increased secretion of IL-1β in Pam3CSK4 pre-stimulated<br />

macrophages relative to wild-type F. novicida (Fig. 6A). We observed a similar<br />

increase in IL-1β secretion in unstimulated macrophages. Increased IL-1β secretion<br />

was accompanied by increased kinetics of host cell death with all mutants tested save<br />

56


ΔkdsA, which induced a higher level of cell death than wild-type F. novicida at 7hrs<br />

post-infection, but only killed 20% of the macrophages by 10hrs post-infection while<br />

50% of the macrophages were killed by wild-type over the same time (Fig. 6B, C).<br />

57


A B<br />

C D<br />

- Figure 5 - LPS mutants hyper-induce the type-I IFN response in macrophages.<br />

Unstimulated wild-type BMDMs were infected with the indicated strains of F. novicida at a multiplicity<br />

of infection of 10 bacteria per macrophage. At the indicated timepoints (A and B) IFN-β mRNA levels<br />

were determined by quantitative RT-PCR and normalized to the level of IFN-β mRNA in uninfected<br />

BMDM or (C and D) total type-I IFN in macrophage supernatant was determined ISRE-L929 reporter<br />

cell assay. Luciferase levels were normalized to the level from uninfected macrophage supernatant.<br />

Means and standard deviations are plotted for experiments done in triplicate. Graphs are representative<br />

of three independent experiments.<br />

58


A<br />

B<br />

C<br />

- Figure 6 - LPS mutants hyper-induce IL-1β release and host cell death.<br />

Pam3CSK4 pre-stimulated wt BMDM were infected with the indicated strains of F. novicida and IL-1β<br />

release was measured at 5.5hrs PI (A). The kinetics of host cell death was measured by LDH release<br />

assay (B and C) in unstimulated wt BMDM. Means and standard deviations of triplicate infections are<br />

shown. Graphs are representative of three independent experiments.<br />

59


2.3.6 F. novicida LPS mutants stimulate increased TLR2-depedent signaling.<br />

Type-I IFN can be induced by TLR signaling. F. novicida induces TLR2<br />

dependent signaling, resulting in the production of proinflammatory cytokines. We<br />

posited that mutants with altered LPS may lead to increased TLR recognition,<br />

accounting for the increased cytosolic responses. To this end we infected wt and<br />

myd88/trif -/- BMDM with wild-type F. novicida or the LPS mutants and measured<br />

type-I IFN secretion and host cell death. We found that the increased type-I IFN<br />

production observed with the LPS mutants at 2hrs post-infection was completely<br />

MyD88/Trif dependent (Fig. 7A). However, at 8hr post-infection, there was a<br />

significant production of type-I IFN from the LPS mutants in the absence of TLR<br />

signaling (Fig. 7B). Furthermore, we observed very little contribution of TLR<br />

signaling to the host cell death response (Fig. 7C). This suggests that TLR signaling<br />

contributes to the early hyper-induction of type-I IFN, but cytosolic signaling is<br />

responsible for later type-I IFN production and host cell death.<br />

60


A B<br />

C<br />

- Figure 7 - TLR signaling contributes to the early but not late responses to LPS mutants.<br />

wt or myd88/trif -/- BMDM were infected with the indicated strains of F. novicida at an MOI of 10:1.<br />

Type-I IFN in macrophage supernatant was measured by ISRE-L929 reporter cell assay at 2hrs post-<br />

infection (A) or 8hrs PI (B). Host cell death was measured by LDH assay at 8hr PI (C). Means and<br />

standard deviations of triplicate infections are shown. Graphs are representative of three independent<br />

experiments.<br />

61


TLR2-dependent signaling contributes to proinflammatory cytokine production<br />

in response to wild-type F. tularensis ssp. To see if the MyD88/Trif-depedent hyper-<br />

IFN phenotype of the LPS mutants was TLR2-dependent we infected wt or tlr2 -/-<br />

BMDM with wild-type F. novicida or the LPS mutants an measured type-I IFN<br />

secretion and host cell death. We found that the hyper-induction of type-I IFN<br />

observed with the LPS mutants at 2hrs post-infection was TLR2-dependent (Fig. 8A).<br />

However, in agreement with the results from the myd88/trif-/- BMDM, the hyper-<br />

simulation of type-I IFN induction and host cell death at 8hrs post-infection was<br />

TLR2-independent (Fig. 8B, C). Thus, we conclude that F. novicida LPS mutants<br />

hyper-stimulate TLR2 dependent type-IFN production, but this is only partially<br />

contributes to the induction of the cytosolic responses.<br />

62


A B<br />

C<br />

- Figure 8- TLR2 contributes to the early but not late cytosolic responses to LPS mutants.<br />

wt or tlr2 -/- BMDM were infected with the indicated strains of F. novicida at an MOI of 10:1. Type-I<br />

IFN in macrophage supernatant was measured by ISRE-L929 reporter cell assay at 2hrs (A) or 8hrs (B)<br />

post-infection. Host cell death was measured by LDH assay at 8hr PI (C). Means and standard<br />

deviations of triplicate infections are shown. Graphs are representative of three independent<br />

experiments.<br />

63


2.3.7 F. novicida LPS mutants hyper stimulate the inflammasome.<br />

F. novicida infection induces ASC and caspase-1 inflammasome-dependent<br />

IL-1β secretion. Furthermore, this inflammasome activation requires type-I IFN<br />

signaling. To determine if the hyper-inducing mutants were hyper-stimulating an<br />

inflammasome-dependent pathway we infected wt, ifnar -/- , asc -/- , and caspase-1 -/-<br />

macrophages with each of the mutants and measured IL-1β release and host cell death.<br />

Similar to wild-type F. novicida, all mutants induced IL-1β release and host cell death<br />

in an ifnar-, asc-, and caspase-1-dependent manner (Fig. 9A-C). Unlike previous<br />

findings with ΔoppB and ΔFTT_1209c hyper-cytotoxic mutants (18), ΔFTN_1212,<br />

ΔlpcC, ΔwbtA, ΔkdsA, and ΔlpxH did not exhibit increased intracellular replication in<br />

macrophages compared to wild-type F. novicida (Fig. 9D). The hyper-IFN inducing<br />

phenotype of these mutants remained intact in asc -/- and caspase-1 -/- macrophages<br />

(Fig. 9E). A decrease in type-I IFN was observed in ifnar -/- macrophages infected<br />

with wild-type F. novicida or ΔFTN_1212 relative to wild-type macrophages (Fig.<br />

9E), consistent with previously published reports (34, 64). Surprisingly, ΔlpcC and<br />

ΔwbtA induced a type-I IFN response in an ifnar-independent manner (Fig. 9E).<br />

Together these results suggest that ΔFTN_1212, ΔlpcC, ΔwbtA, ΔkdsA, and ΔlpxH<br />

hyper-stimulate inflammasome activation by a mechanism independent of bacterial<br />

replication.<br />

64


A B<br />

C D<br />

E<br />

- Figure 9 - ASC, caspase-1, and IFNAR are required for LPS mutant inflammasome<br />

activation.<br />

BMDM from wt, asc-/-, casp-1 -/- , or ifnar -/- were infected with the indicated strains of F. novicida at an<br />

MOI of 10:1 (A, B, C, and D) or 100:1 (E). IL-1β release was measured by ELISA in Pam3CSK4 pre-<br />

65


stimulated at 5.5 hrs (A) or in unstimulated BMDM at 7hrs post-infection (B) . To avoid complications<br />

with host cell death intracellular replication was measured in asc -/- BMDM by gentamicin protection<br />

assay (D). Replication was normalized to cfu counts at 30min post-infection. Type-I IFN secretion was<br />

measured by ISRE-L929 reporter cell assay at 5hrs post-infection (E). Means and standard deviations of<br />

triplicate infections are shown. Graphs are representative of three independent experiments.<br />

2.3.8 LPS mutants have reduced fitness in vivo.<br />

Lipid A is an essential molecule for bacterial viability and many F. novicida<br />

mutants with alterations in LPS have reduced fitness in mouse infection models (80,<br />

150, 173, 177). Furthermore, other F. novicida mutants that display hyper-<br />

cytotoxicity in macrophages are attenuated in vivo (Weiss et al, 2007). We determined<br />

the relative fitness of each of our hyper-inducing mutants in a mouse model of<br />

infection by measuring the competitive index after intradermal challenge. At 48-hours<br />

post-infection all hyper-inducing mutants displayed a competitive index significantly<br />

less than 1 in the skin and spleen of infected mice, indicating that these mutants are<br />

less fit than wild-type F. novicida in vivo (Fig. 10). The competitive index for each<br />

mutant was significantly lower in the spleen than in the skin, suggesting a defect in<br />

either dissemination from the site of infection or an inability to colonize systemic<br />

sites. These results confirm that FTN_1212, lpcC, wbtA, kdsA, and lpxH are F.<br />

novicida virulence factors and further demonstrate the correlation between the ability<br />

to limit the macrophage cytosolic responses and survival in vivo.<br />

66


- Figure 10 – LPS mutants have reduced fitness in vivo.<br />

wt C57BL6/J mice were infected intradermally (i.d.) with 10 5 cfu of an equal ratio of wild-type F.<br />

novicida and individual LPS mutants. At 2 days post-infection skin and spleen was harvested and<br />

bacterial loads were determined by plating serial dilutions of homogenized tissue on selective and non-<br />

selective MH agar.<br />

2.3.9 Phagosomal escape is required for induction of the cytosolic response by<br />

LPS mutants.<br />

Cytosolic localization is required for induction of the type-I IFN and cell death<br />

responses in F. novicida, as mutants in FPI genes are restricted to the phagosome and<br />

thus do not induce the cytosolic responses. To determine if the hyper-inducing<br />

mutants still required phagosomal escape to induce the cytosolic response we<br />

constructed FPI deletions in the ΔFTN_1212, ΔlpcC, ΔwbtA, and ΔkdsA backgrounds<br />

67


and measured the kinetics of the type-I IFN response and inflammasome activation in<br />

bone marrow-derived macrophages. Similar to a ΔFPI deletion mutant<br />

ΔFTN_1212ΔFPI, ΔlpcCΔFPI, and ΔwbtAΔFPI double mutants were attenuated in<br />

their ability to elicit type-I IFN production in macrophages (Fig. 11C). The double<br />

mutants were also failed to activate the inflammasome characterized by attenuated<br />

induction of IL-1β secretion and host cell death (Fig. 11A, B). From these results we<br />

conclude that FPI-dependent phagosomal escape is required for induction of the<br />

cytosolic response by the hyper-inducing mutants.<br />

68


A B<br />

C D<br />

- Figure 11 - Phagosomal escape is required for LPS mutants to hyper-induce the cytosolic<br />

responses.<br />

Unstimulated wt BMDM were infected with the indicated strains of F. novicida at an MOI of 10:1. At<br />

8hrs post-infection IL-1β release was measured by ELISA at (A) and cell death was measured by LDH<br />

assay (B). Kinetics of type-I IFN secretion was measured by ISRE-L929 reporter cell assay (C). (D) is<br />

a bar graph representing the luciferase levels at 2hrs post-infection in (C). Means and standard<br />

deviations of triplicate infections are shown. Graphs are representative of three independent<br />

experiments.<br />

69


2.3.10 LPS mutants induce increased proinflammatory cytokine signaling in the<br />

phagosome.<br />

Although the hyper-inducing FPI double mutants were unable to stimulate an<br />

increase in macrophage type-I IFN production over time, we observed an initial<br />

increase in the magnitude of type-I IFN produced at 2hrs post-infection relative to<br />

wild-type F. novicida and a single ΔFPI mutant (Fig. 11C, D). Interestingly, at 2hrs<br />

post-infection hyper-inducing single mutants and their ΔFPI-double mutant<br />

counterparts induced a similar level of type-I IFN in macrophages (Fig. 11D). This<br />

result suggests that the type-I IFN response has a vacuolar component that is FPI<br />

independent at 2hrs post-infection and that mutants that hyper-induce the cytosolic<br />

response also hyper-induce this vacuolar response.<br />

Previous reports demonstrated that retention of F. novicida in the vacuole is<br />

characterized by production of pro-inflammatory cytokines such as TNF-α and IL-1β<br />

(35). To extend our study of the vacuolar responses induced by the hyper-inducing<br />

mutants we investigated macrophage production of these pro-inflammatory cytokines<br />

in response to the hyper-inducing ΔFPI double mutants. Macrophages secreted<br />

increased amounts of TNF-α and produced higher levels of pro-IL-1β transcript in<br />

response to ΔFTN_1212ΔFPI, ΔlpcCΔFPI, ΔwbtAΔFPI, and ΔkdsAΔFPI double<br />

mutants relative to a ΔFPI single mutant (Fig . 12A, B). Thus we have demonstrated<br />

that in addition to an increased IFN-β response, these mutants hyper-induce the<br />

production of TNF-α and pro IL-1β.<br />

70


A B<br />

C<br />

- Figure 12 - LPS mutants hyper-induce NF-κB-dependent cytokines in the phagosome.<br />

wt BMDM were infected with the indicated strains of F. novicida at an MOI of 10:1 for 8hrs. Pro-IL-<br />

1β mRNA levels were measured by quantitative RT-PCR and normalized to the level in uninfected<br />

BMDM (A). TNF-α in supernatants was measured by ELISA (B). (C) RAW-κB luciferase reporter<br />

cells were infected with the indicated strains of F. novicida for 8hrs at an MOI of 10:1. Means and<br />

standard deviations of triplicate infections are shown. Graphs are representative of three independent<br />

experiments.<br />

71


Since IL-1β and TNF-α production both require nuclear translocation of NF-<br />

κB, we hypothesized that the hyper-inducing ΔFPI double mutants stimulated higher<br />

levels of NF-κB. To test our hypothesis we used our hyper-inducing single mutants<br />

and hyper-inducing ΔFPI double mutants to infect a RAW 264.7 macrophage cell line<br />

containing a luciferase-linked NF-κB reporter. At 8hrs post-infection we observed a<br />

significant increase in luciferase production from RAW macrophages infected with<br />

ΔlpcC, ΔwbtA, ΔkdsA, and their ΔFPI double mutant counterparts relative to wild-type<br />

F. novicida and a ΔFPI single mutant (Fig. 12C). The ΔFTN_1212 and<br />

ΔFTN_1212ΔFPI mutants did not induce more luciferase production than wild-type F.<br />

novicida. These results suggest that ΔlpcC, ΔwbtA, ΔkdsA, and their ΔFPI double<br />

mutant counterparts stimulate increased translocation of NF-κB relative to wild-type<br />

and ΔFPI single mutant F. novicida, resulting in increased magnitude and kinetics of<br />

production of several NF-κB-dependent pro-inflammatory cytokines.<br />

2.3.11 Surface-exposed PAMPs mediate recognition of LPS mutants.<br />

Deletion of genes involved in LPS/cell wall/capsule production could lead to<br />

an altered outer membrane surface and expose PAMPs not detectable in wild-type F.<br />

novicida. To determine if the increased cytokine production in response to the LPS<br />

mutants could be due to exposure of PAMPs on the bacterial surface we exposed wt<br />

BMDM to UV-killed wild-type F. novicida or LPS mutants and measured TNF-α and<br />

72


type-I IFN secretion in macrophage supernatants at 2hrs post-infection. We observed<br />

increased stimulation of TNF-α and type-I IFN with UV killed LPS mutants compared<br />

to wild-type. F. novicida (Fig 13A, C). Furthermore, we saw no difference in TNF-α<br />

or type-I IFN production from macrophages infected with wild-type F. novicida or<br />

LPS mutant strains that had been heat-killed (HK) (Fig. 13B, D). These results<br />

suggest that hyper-induction of cytokines by the LPS mutants is not dependent on<br />

bacterial viability but is dependent on an intact outer membrane. Therefore we<br />

conclude that an altered outer membrane in the LPS mutants exposes PAMPs that are<br />

not accessible on wild-type F. novicida.<br />

73


A B<br />

C D<br />

- Figure 13 - Surface exposed PAMPs mediate recognition of LPS mutants.<br />

wt BMDMs were infected for 2hrs at an MOI of 10:1 with the indicated strains of UV-killed (A and C)<br />

or heat-killed (B and D) F. novicida. TNF-α in macrophage supernatant was measured by ELISA (A<br />

and B). Type-I IFN in macrophage supernatant was measured by ISRE-L929 reporter cell assay (C and<br />

D). Means and standard deviations of triplicate infections are shown. Graphs are representative of three<br />

independent experiments.<br />

74


2.3.12 Cytosolic localization is necessary but not sufficient to induce the cytosolic<br />

responses.<br />

Mutants that hypo-induced the cytosolic response fell into several broad<br />

predicted functional categories such as chaperones, transporters, amino acid<br />

metabolism, and nucleotide metabolism. Importantly, we did not identify a F. novicida<br />

mutant that replicated as efficiently as wild-type but did not induce the cytosolic<br />

response. Additional screening of selected transposon insertions as well as review of<br />

the literature revealed a positive correlation between intracellular replication and<br />

induction of the cytosolic response (Table 3); mutants that failed to replicate<br />

intracellularly did not induce a measurable cytosolic response and mutants that were<br />

able to replicate but to a lower extent than wild-type also hypo-induced the cytosolic<br />

response. No mutant tested replicated to higher levels than wild-type, though it should<br />

be noted that ΔoppB and ΔFTT_1209c mutants display increased intracellular<br />

replication and increased macrophage cytotoxicity relative to a wild-type F. novicida<br />

strain (18). These genes were not identified in our screen.<br />

75


- Table 3 – Mutants attenuated for intracellular replication are hypo-stimlate the cytosolic<br />

responses<br />

Gene IFN Cell Death Intracellular replication in<br />

macrophages<br />

Reference<br />

Wild-type<br />

F. novicida<br />

+ + + Henry et al., 2007<br />

pdpA - - - Brotcke et al, 2006<br />

pdpB - - - Brotcke et al., 2006<br />

pdpD +/- +/- +/- Ludu et al., 2008<br />

caiC/migR +/- +/- +/- Buchan et al., 2009, Brotcke<br />

unpublished<br />

pmrA +/- +/- +/- Mohapatra et al., 2007<br />

htpG - - - Weiss et al., 2007<br />

carA - - - Schulert et al., 2009<br />

carB - - - Schulert et al., 2009<br />

proQ +/- +/- +/- This study<br />

pal ? +/- +/- This study<br />

purMCD - - - Pechous et al., 2006<br />

perM +/- +/- +/- This study<br />

emrB +/- +/- +/- This study<br />

clpB - - - Gray et al., 2002<br />

cphA +/- +/- +/- Brotcke, unpublished<br />

We further investigated the correlation between bacterial escape, cytosolic<br />

replication, and induction of the cytosolic response. Previous work demonstrated that<br />

bacterial localization in the cytosol was required to induce the cytosolic response (64,<br />

97). To determine is cytosolic localization is sufficient to induce the cytosolic<br />

response we took advantage of a purine auxotroph, ΔpurMCD, which has been<br />

reported to escape the phagosome as efficiently as wild-type but fails to replicate in<br />

the macrophage cytosol in the absence of exogenous purines (134). In total, 10 purine<br />

or pyrimidine biosynthesis mutants were identified in our screen that hypo-induced the<br />

cytosolic responses. We infected bone marrow derived macrophages with wild-type<br />

F. novicida, ΔpurMCD, or ΔiglC and measured the kinetics of the type-I IFN and cell<br />

76


death responses. Whereas wild-type F. novicida induced robust type-I IFN and cell<br />

death responses from the macrophage a both the ΔpurMCD and ΔiglC mutants failed<br />

to induce either of these responses (Fig. 14A, B). This data suggests that bacterial<br />

localization in the macrophage cytosol is necessary but not sufficient to induce the<br />

cytosolic responses.<br />

77


A<br />

B<br />

- Figure 14 - Cytosolic localization is necessary but insufficient to induce the cytosolic<br />

responses.<br />

Unstimulated BMDM were infected with the indicated strains of F. novicida at an MOI of 10:1.<br />

Kinetics of type-I IFN secretion in supernatant was measured by ISRE-L929 assay (A). Kinetics of<br />

host cell death was measured by LDH release assay (B). Means and standard deviations are shown for<br />

infections done in triplicate. Representative graphs of three independent experiments are shown.<br />

78


2.3.13 Bacterial DNA and protein synthesis are required to induce the cytosolic<br />

responses.<br />

During a macrophage infection with F. novicida greater than 60% of bacteria<br />

are present in the cytosol within 2hrs post-infection (28). However, we do not observe<br />

intracellular replication until 4hrs post-infection, consistent with published findings<br />

(175). During this lag phase in intracellular replication the bacterium exhibits<br />

transcriptional changes that may be required in order to establish a replicative niche in<br />

the cytosol (175). Since bacterial localization in the cytosol was insufficient to induce<br />

the cytosolic responses and the results of the F. novicida transposon screen<br />

demonstrated a correlation between intracellular replication and induction of the<br />

cytosolic responses we sought to determine if intracellular replication, protein<br />

synthesis, and/or DNA synthesis were necessary to induce the cytosolic responses.<br />

We infected macrophages with wild-type F. novicida at a multiplicity of infection<br />

(MOI) of 10 and at 1, 2, or 3hrs post-infection we treated the macrophages with<br />

bacteriostatic antibiotics to block intracellular replication and measured the kinetics of<br />

type-I IFN induction and macrophage cell death. Both chloramphenicol and nalidixic<br />

acid were bacteriostatic at the concentrations used and no intracellular replication was<br />

observed in the presence of antibiotics over 8hrs of infection (Fig. 15I). Antibiotic<br />

treatment at 0, 1, and 2hrs post-infection led to attenuated induction of both the type-I<br />

IFN and cell death responses (Fig. 15A-F). Interestingly, treatment at 3hrs post-<br />

infection resulted in induction of the type-I IFN response with similar magnitude and<br />

kinetics as untreated samples (Fig. 15G). However, treatment at 3hrs post-infection<br />

led to a significant attenuation of the host cell death response (Fig. 15H). These<br />

79


esults suggest that a minimum of 3 hours of protein and DNA synthesis are required<br />

by the bacterium to induce the cytosolic response. Furthermore, intracellular<br />

replication was not required for induction of type-I IFN, but host cell death was<br />

attenuated at this MOI in the absence of replication. This was the first time we were<br />

able to observe a type-I IFN response that did not result in a cell death response.<br />

80


A B<br />

C D<br />

E F<br />

81


G H<br />

I<br />

- Figure 15 - Bacterial protein synthesis and DNA synthesis are required to induce the cytosolic<br />

responses.<br />

wt BMDM were infected with wt F. novicida at an MOI of 10:1 and bacteriostatic antibiotics were<br />

added at 0hr (A and B), 1hr (C and D), 2hr (E and F), or 3hr post-infection (G and H). Type-I IFN in<br />

macrophage supernatant was measured by ISRE-L929 reporter assay (A,C,E, and G). Host cell death<br />

was measured by LDH assay (B,D,F and H). Intracellular replication was measured by gentamicin<br />

protection assay (I). Means and standard deviations are shown for infections done in triplicate.<br />

Representative graphs of three independent experiments are shown.<br />

Intracellular replication is dispensable for induction of type-I IFN, but required<br />

for the induction of cell death at an MOI of 10. This could suggest that the ligand<br />

responsible for the type-I IFN response is expressed during the lag phase, while the<br />

82


ligand responsible for host cell death is expressed during the replicative phase.<br />

Alternatively, the threshold for induction of type-I IFN may be lower than the<br />

threshold for induction of host cell death, and bacterial replication may serve as a<br />

means to increase the concentration of the ligand(s). To test this hypothesis we<br />

infected macrophages at an MOI of 100 to increase the initial concentration of all<br />

bacterial ligands and treated the macrophages with bacteriostatic antibiotics at 3hrs<br />

post-infection to block replication. Similar to the results at an MOI of 10, treatment<br />

with bacteriostatic antibiotics had no effect on the magnitude and kinetics of type-I<br />

IFN production (Fig. 16A). Additionally at this higher MOI, the bacteria induced host<br />

cell death in the absence of replication, though there was a slight decrease in kinetics<br />

(Fig. 16B). These results suggest that the ligand(s) for induction of both type-I IFN<br />

and host cell death are present in the cytosol by 3hrs post-infection, and bacterial<br />

replication increases the concentration of the ligand(s) such that it can be recognized<br />

by the inflammasome.<br />

83


A<br />

B<br />

- Figure 16 - Replication is not required to induce the cytosolic response with high bacterial<br />

load.<br />

wt BMDM were infected with wild-type F. novicida at an MOI of 100:1. Bacteriostatic antibiotics<br />

were added at 3hr PI. Type-I IFN in macrophage supernatant was measured by ISRE-L929 reporter<br />

asay (A). Host cell death was measured by LDH assay (B). Means and standard deviations are shown<br />

for infections done in triplicate. Representative graphs of three independent experiments are shown.<br />

84


2.3.14 The cytosolic response to F. novicida shares characteristics with the<br />

response to transfected dsDNA.<br />

Type-I IFN induction and host cell death are both induced by transfection of<br />

dsDNA into the macrophage cytosol (119, 153). We wanted to know if we observed<br />

the same dose dependent activation of the two responses with dsDNA transfection as<br />

we observed with F. novicida infection. To this end we infected wt BMDM with<br />

increasing MOIs of F. novicida or transfected increasing doses of poly(dA:dT) and<br />

measure type-I IFN secretion and host cell death. The characteristics of the responses<br />

were very similar between the different stimuli. An 8 hr infection with an initial dose<br />

of less that 30 bacteria per macrophage resulted in induction of the type-I IFN<br />

response without induction of the cell death response, similar to what we had observed<br />

at an MOI of 10 in the presence of bacteriostatic antibiotics (Fig. 17A). At an MOI of<br />

30 or higher, we observed induction of the type-I IFN response and host cell death<br />

(Fig. 17A). Similarly, transfection of less than 15ng/well of poly(dA:dT) induced<br />

type-I IFN secretion from macrophage with no host cell death (Fig. 17B). However<br />

transfection with poly(dA:dT) at concentrations of 15ng/well or higher induced a<br />

robust type-I IFN response and host cell death (Fig. 17B). Importantly, we did not<br />

observed induction of host cell death without accompanying type-I IFN induction.<br />

85


A B<br />

C<br />

- Figure 17 - The cytoslic response to F. novicida shares characteristics with that of dsDNA.<br />

wt BMDM were infected with an increasing MOI of F. novicida for 8hrs(A) or transfected with<br />

increasing concentrations of poly(dA:dT) for 5 hrs (B) and type-I IFN in macrophage supernatant was<br />

measured by luciferase reporter assay and cell death was measured by LDH assay. (C) wt or ifnar -/-<br />

BMDM were infected with F. novicida or transfected with poly(dA:dT) at the indicated concentrations<br />

and host cell death was measured by LDH assay. Means and standard deviations are shown for<br />

infections done in triplicate. Representative graphs of three independent experiments are shown.<br />

86


Since host cell death in response to F. novicida requires type-I IFN signaling,<br />

we wanted to know if transfection of dsDNA displayed a similar dependence on type-I<br />

IFN signaling. Therefore we infected wild-type and ifnar-/- macrophages with F.<br />

novicida at MOIs of 62:1 and 125:1 or transfected the same strains of macrophages<br />

with poly(dA:dT) at concentration of 62ng/well and 125ng/well. We observed a<br />

significant decrease in host cell death in ifnar -/- macrophages compared to wt<br />

macrophages that were infected with F. novicida or transfected with poly(dA:dT) (Fig.<br />

17C). These results, combined with the results from our screen, support the idea that<br />

F. novicida dsDNA triggers the cytosolic responses in macrophages.<br />

87


2.4 DISCUSSION<br />

In this report we identified several F. novicida genes that modulate<br />

macrophage innate immune responses. Mutants in a gene encoding a major outer<br />

membrane protein, fopA, and the LPS genes FTN_1212, lpcC, wbtA, kdsA, and lpxH<br />

induce increased type-I IFN and inflammasome activation in macrophages. Further<br />

study revealed that these mutants induce the cytosolic responses through an identical<br />

pathway as wild-type F. novicida, dependent on type-I IFN signaling and<br />

inflammasome components ASC and caspase-1. Alterations in the outer membrane<br />

and LPS led to increased stimulation of TLRs, and resulted increased proinflammatory<br />

cytokine production. This increased TLR signaling is likely due to detection of surface<br />

exposed PAMPs, and not due to the production of a new PAMP that is not present in<br />

wild-type F. novicida. The increased TLR2-dependent type-I IFN production could<br />

synergize with the type-I IFN produced from the unknown cytosolic receptor and lead<br />

to the increased inflammasome activation observed with these mutants. However, our<br />

results also indicate that the fopA and LPS mutants hyper-stimulate the cytosolic<br />

receptor as well.<br />

The mutants that alter the cytosolic response in macrophages represent 17% of<br />

the F. novicida genome and fall into several broad functional categories. This would<br />

represent and extremely large number of genes dedicated to target a specific host<br />

pathway and since F. novicida is not host adapted, we surmise that the mutants that<br />

induce an increase in the cytosolic response relative to wild-type F. novicida do so by<br />

converging on a specific pathway or release of the same PAMP. Therefore we<br />

88


conclude that fopA and F. novicida LPS act to dampen the cytosolic response by<br />

limiting the release or exposure of a common F. novicida PAMP.<br />

We observed a positive correlation between intracellular replication and<br />

induction of the cytosolic responses. We did not identify an F. novicida mutant that<br />

replicates intracellularly as efficient as wild-type F. novicida but does not induce the<br />

cytosolic response. However, it appears that the macrophage does not sense<br />

intracellular replication, as we were able to induce the cytosolic response in the<br />

presence of bacteriostatic antibiotics. On the other hand, replication provided a link<br />

between type-I IFN and host cell death at low infectious doses. We propose that the<br />

same PAMP induces the type-I IFN response and cell death responses in a dose<br />

dependent manner, and replication increases the cytosolic concentration of this PAMP.<br />

In support of this hypothesis we observed a similar dose dependent and type-I IFN<br />

receptor-dependent response to transfected dsDNA. The concentration of cytosolic<br />

dsDNA would increase during bacterial replication. However, release of bacterial<br />

DNA during intracellular replication has not been demonstrated previously. DNA<br />

release could result from bacterial lysis induced by a host response, or it could be part<br />

of the intracellular replications cycle that had not been previously described. Mutants<br />

in outer membrane proteins or LPS could have an unstable outer membrane and be<br />

more susceptible to intracellular lysis, resulting in increased release of dsDNA. We<br />

will not be able to definitely say that bacterial DNA is the PAMP that triggers the<br />

cytosolic response until we identify the host receptors for type-I IFN or inflammasome<br />

activation.<br />

89


<strong>Chapter</strong> 3: AIM2 is required for innate immune<br />

recognition of Francisella tularensis<br />

Jonathan W. Jones 1 *, Nobuhiko Kayagaki 2 *, Petr Broz 1 , Thomas Henry 1 , Kim<br />

Newton 2 , Karen O'Rourke 2 , Salina Chan 2 , Jennifer Dong 2 , Yan Qu 2 , Meron<br />

Roose-Girma 3 , Vishva M. Dixit 2 , Denise M. Monack 1<br />

1 Department of Microbiology and Immunology, <strong>Stanford</strong> School of Medicine, <strong>Stanford</strong><br />

<strong>University</strong>, California, USA<br />

2 Department of Physiological Chemistry, 3 Department of Molecular Biology,<br />

Genentech Inc., South San Francisco, California, USA.<br />

This chapter has been accepted for publication in the Proceedings of the National<br />

Academy of Sciences.<br />

90


3.1 CHAPTER 3 SUMMARY<br />

Macrophages respond to cytosolic nucleic acids by activating cysteine protease<br />

caspase-1 within a complex called the inflammasome. Subsequent cleavage and<br />

secretion of proinflammatory cytokines interleukin (IL)-1β and IL-18 is critical for<br />

innate immunity. Here we show that macrophages from mice lacking absent in<br />

melanoma 2 (AIM2) cannot sense cytosolic double-stranded DNA and fail to trigger<br />

inflammasome assembly. Caspase-1 activation in response to intracellular pathogen<br />

Francisella tularensis also required AIM2. Immunofluorescence microscopy of<br />

macrophages infected with F. tularensis revealed striking co-localization of bacterial<br />

DNA with endogenous AIM2, and inflammasome adaptor ASC. By contrast, type I<br />

interferon (Type-I IFN; IFN-α and -β) secretion in response to F. tularensis did not<br />

require AIM2. Type-I IFN did, however, boost AIM2-dependent caspase-1 activation<br />

by increasing AIM2 protein levels. Thus, inflammasome activation was reduced in<br />

infected macrophages lacking either the Type-I IFN receptor (IFNAR) or stimulator of<br />

interferon genes (STING). Finally, AIM2-deficient mice displayed increased<br />

susceptibility to F. tularensis infection compared to wild-type mice. Their increased<br />

bacterial burden in vivo confirmed that AIM2 is essential for an effective innate<br />

immune response.<br />

91


3.2 INTRODUCTION<br />

The innate immune system reacts to diverse molecules, which are collectively<br />

termed pathogen-associated molecular patterns (PAMPs) and damage-associated<br />

molecular patterns (DAMPs) (4, 86). These molecules include nucleic acids. RNA,<br />

for example, is recognized by several toll-like receptors, as well as the RNA helicases<br />

retinoic acid inducible gene-I (RIG-I, also called DDX58), melanoma differentiation-<br />

associated gene-5 (MDA5, also called IFIH1), and laboratory of genetics and<br />

physiology 2 (LGP2) (4). DNA recognition mechanisms have proved more elusive.<br />

Toll-like receptor (TLR)9 is located in phagosomes and recognizes DNA with CpG<br />

motifs, leading to NF-κB dependent inflammatory responses (63). DNA-dependent<br />

activator of IFN-regulatory factors (DAI, also known as DLM-1 and ZBP1), the first<br />

identified cytosolic DNA sensor, binds cytosolic dsDNA and leads to the production<br />

of type-I IFN, although the lack of demonstrated relevance in vivo has lead to the<br />

hypothesis that redundant cytosolic DNA sensors exist (75, 161). Additionally, the<br />

recently identified adapter stimulator of interferon genes (STING) or mediator of IRF3<br />

activation (MITA) [hereafter referred to as STING] mediates type-I IFN production in<br />

response to DNA transfection, as well bacterial and viral infection (76, 77, 137, 187).<br />

DNA is also a potent activator of a multiprotein complex known as the<br />

inflammasome, which contains a nucleotide-binding oligomerization domain (NOD)-<br />

like receptor (NLR), the adapter apoptosis-associated speck-like protein containing a<br />

CARD (ASC, also known as PYCARD) and the cysteine protease caspase-1 (119).<br />

Recent overexpression and knockdown studies in cell lines suggested that DNA<br />

engages AIM2, a novel NLR, which then interacts with ASC to promote<br />

92


inflammasome assembly and caspase-1 activation (21, 68). AIM2 is a type-I IFN-<br />

inducible cytosolic protein containing PYRIN and HIN200 domains (21, 90)<br />

(Supplementary Fig. 1C). The HIN domain facilitates binding of DNA, while the<br />

PYRIN domain allows for the association with ASC and formation of a caspase-1<br />

activating inflammasome, leading to the processing and release of mature IL-1β and<br />

IL-18, and host cell death (21, 47, 68). The role of AIM2 in innate immunity is<br />

unknown.<br />

The causative agent of tularemia, Francisella tularensis is a facultative<br />

intracellular gram-negative pathogen that escapes phagosomal degradation in<br />

macrophages and replicates in the host cell cytosol. Cytosolic replication is required<br />

for bacterial virulence, as F. tularensis mutants that fail to escape the vacuole cannot<br />

replicate in macrophages and are avirulent in mice (17, 54, 91, 122, 177). Moreover,<br />

cytosolic F. tularensis sequentially activates pro-inflammatory host responses,<br />

characterized by the initial production of type-I IFN, such as IFN-β, which is required<br />

for the subsequent activation of an ASC inflammasome (34, 64). Inflammasome<br />

activation is critical to host defense against F. tularensis, as mice lacking<br />

inflammasome components are more susceptible to infection (97). The PAMPs<br />

produced during F. tularensis infection and the host pattern recognition receptors<br />

(PRRs) required for pathogen recognition remain a mystery.<br />

93


3.3 RESULTS<br />

3.3.1 AIM2 is essential for inflammasome activation in response to cytosolic<br />

dsDNA.<br />

We investigated the role of AIM2 in vivo with gene-targeted aim2 -/- mice (Fig.<br />

18A, B). Western blotting with an antibody raised against amino acids 2-274 of<br />

mouse AIM2 confirmed that aim2 -/- bone marrow-derived macrophages (BMDMs)<br />

lacked AIM2 protein (Fig. 18C). First, we compared double-stranded DNA (dsDNA)-<br />

induced IL-1β secretion from aim2 -/- , asc -/- , nalp3 -/- ipaf -/- , and wild-type BMDMs.<br />

Cells were primed with lipopolysaccharide (LPS) to induce pro-IL-1β expression and<br />

then transfected with either poly(dA-dT) • poly(dA-dT) [hereafter referred to as<br />

poly(dA:dT)], poly(dG-dC) • poly(dG-dC) [hereafter referred to as poly(dG:dC)],<br />

pcDNA3 plasmid DNA, calf thymus DNA, or Listeria monocytogenes DNA. All of<br />

these dsDNAs induced AIM2- and ASC-dependent IL-1β secretion (Fig. 19A). In<br />

contrast, Nalp3 (98, 104) (also called NLRP3), which engages the caspase-1 adaptor<br />

protein ASC in response to a variety of PAMPs and DAMPs, and Ipaf (96) (also called<br />

NLRC4), which engages ASC in response to Salmonella typhimurium (S.<br />

typhimurium), were dispensable for dsDNA-induced IL-1β secretion. Loss of AIM2,<br />

unlike ASC deficiency, did not cause a general defect in IL-1β secretion since aim2 -/-<br />

and wild-type BMDMs secreted equivalent amounts of IL-1β after infection with S.<br />

typhimurium or treatment with LPS plus ATP (Fig. 19A). Similar results were<br />

obtained with peritoneal macrophages (Fig. 20). AIM2 deficiency also blocked IL-18<br />

94


secretion in response to dsDNA but not ATP (Fig. 19B). Therefore, AIM2 is essential<br />

for IL-18 and IL-1β secretion in response to dsDNA.<br />

95


- Figure 18 - Generation of Aim2 -/- mice.<br />

A, Strategy for deleting exon 5 of mouse aim2, which encodes the initiating methionine (ATG) and the<br />

entire PYRIN domain. Gene targeting was performed in C57BL/6 C2 embryonic stem cells. B, PCR<br />

genotyping of aim2 -/- mice. Primers 5’CCA GTG TTT CTC AAC TGT ACT GCT AT, 5’TAG GAG<br />

TGC CCT CCC TTA ATG, and 5’TTG GAG ACA GAC TCT GGT GAA G yield a 197 bp DNA<br />

fragment for the wild-type allele and a 397 bp DNA fragment for the knockout allele. C, BMDMs were<br />

incubated with 1000 U/mL IFN-b for 5 h. AIM2 was western blotted with 4G9.1.4 rat anti-mouse<br />

AIM2 monoclonal antibody.<br />

96


- Figure 19 - AIM2 is essential for inflammasome activation in response to cytosolic dsDNA.<br />

� = asc -/- , � = nalp3 -/- ipaf -/- , � = aim2 -/- , � = wt. A, IL-1b secretion by LPS-primed BMDMs treated<br />

with 5 mM ATP or transfected with 1 ug/mL of the indicated dsDNAs for 16 h. BMDMs were infected<br />

with S. typhimurium (multiplicity of infection = 100) without LPS priming. B, IL-18 secretion by<br />

BMDMs treated as in (A). LPS priming was used ATP stimulation only. C, Upper panels show mature<br />

IL-1b and cleaved caspase-1 secreted from LPS-primed BMDMs after stimulation with ATP or<br />

transfection with dsDNA for 5 h. Lower panels show pro-caspase-1 and pro-IL-1β in the cell lysate. D,<br />

IFN-β secretion by BMDMs treated as in (A), but without LPS priming. Graphs show the mean ±<br />

standard deviation of triplicate wells and are representative of 3 independent experiments.<br />

97


- Figure 20 - AIM2 is essential for inflammasome activation in response to dsDNA in peritoneal<br />

macrophages.<br />

� = asc -/- , � = nalp3 -/- ipaf -/- , � = aim2 -/- , � = wt. Peritoneal macrophages were harvested 5 days after<br />

intraperitoneal injection of 4% thioglycollate (DIFCO). These macrophages were primed with 500<br />

ng/mL LPS for 5 h, then transfected with 1 ug/mL of the dsDNAs indicated for 16 h. As controls, LPS-<br />

primed cells were cultured in medium alone (cont) or stimulated with 5 mM ATP. Additional<br />

macrophages were not primed (cont) or were infected with S. typhimurium (multiplicity of infection,<br />

100). IL-1β secreted into the culture supernatant was measured by ELISA. Graphs show the mean ±<br />

standard deviation of triplicate wells and are representative of 3 independent experiments.<br />

Processing of pro-IL-1β and pro-IL-18 by caspase-1 is necessary for secretion<br />

of biologically active IL-1β and IL-18 (39, 88) and so we compared caspase-1<br />

activation in wild-type and aim2 -/- BMDMs by western blotting for the p20 and p10<br />

caspase-1 subunits that are generated by autocatalytic cleavage and released from the<br />

98


cell by a poorly defined mechanism. Consistent with AIM2 promoting caspase-1<br />

activation in response to dsDNA, culture supernatants from LPS-primed wild-type<br />

BMDMs contained mature IL-1β plus the caspase-1 p10 and p20 subunits after<br />

transfection with poly(dA:dT) or poly(dG:dC), but supernatants from aim2 -/- BMDMs<br />

did not (Fig. 19C). aim2 -/- BMDMs expressed wild-type levels of pro-caspase-1 and<br />

pro-IL-1β, and they released IL-1β and processed caspase-1 normally in response to<br />

LPS plus ATP. These data indicate a specific requirement for AIM2 in caspase-1<br />

activation by dsDNA.<br />

Next, we determined whether aim2 -/- BMDMs produce inflammasome-<br />

independent proinflammatory cytokines such as type-I IFN and TNF-α in response to<br />

dsDNA. TANK-binding kinase 1 (TBK1) and the transcription factors IRF3 and IRF7<br />

signal type-I IFN synthesis in response to dsDNA (74, 153), but how dsDNA engages<br />

this pathway is unclear. Unexpectedly, aim2 -/- , asc -/- , and caspase-1 -/- BMDMs<br />

produced significantly more TNF-α (Fig. 21A) and IFN-β (Fig. 19D, Fig. 21C) than<br />

wild-type or nalp3 -/- ipaf -/- BMDMs after transfection with poly(dA:dT) or pcDNA3.<br />

There was little or no difference in IFN-β and TNF-α production, however, when the<br />

cells were transfected with poly(dG:dC) or a 45 base pair IFN-stimulatory DNA (ISD)<br />

(153). In addition, wild-type, aim2 -/- , asc -/- , and nalp3 -/- ipaf -/- BMDMs produced<br />

equivalent TNF-α in response to LPS (Fig. 21A). Therefore, AIM2 is dispensable for<br />

IFN-β and TNF-α secretion. Increased IFN-β and TNF-α production by aim2 -/- and<br />

caspase-1 -/- BMDMs in response to poly(dA:dT) or pcDNA3 correlated with enhanced<br />

cell survival. Between 35-45% of wild-type BMDMs had released lactate<br />

dehydrogenase (LDH) at 5 hours after transfection with poly(dA:dT) or pcDNA3,<br />

99


whereas most caspase-1 -/- and aim2 -/- BMDMs remained viable (Fig. 21B).<br />

Poly(dG:dC) was less cytotoxic, killing ~15% of wild-type BMDMs. Consistent with<br />

AIM2 engaging ASC and caspase-1 in response to dsDNA, but not all stimuli, aim2 -/-<br />

BMDMs were as sensitive as wild-type BMDMs to caspase-1-dependent death after<br />

infection with S. typhimurium.<br />

100


- Figure 21 - AIM2 is dispensable for IFN-β and TNF-α production in response to dsDNA.<br />

BMDMs were unstimulated (cont), treated with 1 mg/mL LPS, or transfected with 1 ug/mL of the<br />

dsDNAs indicated for 16 h. A, � = asc -/- , � = nalp3 -/- ipaf -/- , � = aim2 -/- , � = wt. TNF-α secreted into<br />

the culture supernatant was measured by Bio-Plex Cytokine Assay (Bio-Rad). B, � = casp-1 -/- , � =<br />

aim2 -/- , � = wt. Cytotoxicity was measured by LDH release. C, � = casp-1 -/- , � = wt. IFN-β secreted<br />

into the culture supernatant was measured by ELISA. Graphs show the mean ± standard deviation of<br />

triplicate wells and are representative of 3 independent experiments.<br />

3.3.2 AIM2 is required for inflammasome activation in response to F. tularensis.<br />

We then explored the contribution of AIM2 to innate immunity to bacterial<br />

infection. We focused on inflammasome activation in response to F. tularensis (84,<br />

110) because this intracellular pathogen escapes phagosomal degradation, replicates in<br />

the cytosol, and triggers ASC-dependent, but Nalp3- and Ipaf-independent, caspase-1<br />

activation (97). When wild-type, aim2 -/- , asc -/- , and caspase-1 -/- BMDMs were<br />

infected with F. tularensis subspecies novicida, only wild-type cells secreted IL-1β<br />

101


(Fig. 22A) and died (Fig. 22B), indicating that AIM2, like ASC and caspase-1, is<br />

essential for inflammasome activity. We propose that AIM2 recognizes cytosolic F.<br />

tularensis because the avirulent mutant ΔFPI (177), which cannot escape phagocytic<br />

vacuoles, failed to stimulate IL-1β secretion (Fig. 22A) or macrophage death (Fig.<br />

22B).<br />

102


- Figure 22 - AIM2 is required for inflammasome activation in response to F. tularensis.<br />

� = asc -/- , � = casp-1 -/- , � = aim2 -/- , � = wt. A, IL-1β secretion by BMDMs infected with F.<br />

tularensis ssp. novidica strain U112 or isogenic mutant ΔFPI for 5 h. (multiplicity of infection, moi)<br />

BMDMs treated with 5 mM ATP for 4 h were primed with 500 ng/mL Pam3CSK4 for 16 h.<br />

multiplicity of infection, moi. B, Cytotoxicity as measured by LDH release. Graphs show the mean ±<br />

standard deviation of triplicate wells and are representative of 3 independent experiments.<br />

103


3.3.3 AIM2 and ASC form a complex with F. tularensis DNA.<br />

The HIN200 domain of AIM2 recognizes dsDNA, and its PYRIN domain can<br />

engage ASC (47, 68). We visualized inflammasome assembly by<br />

immunofluorescence confocal microscopy of BMDMs infected with F. tularensis and<br />

stained with antibodies detecting endogenous AIM2, ASC, and F. tularensis. At 5.5<br />

hours post-infection, wild-type, asc -/- , and caspase-1 -/- BMDMs contained multiple<br />

AIM2 specks tightly associated with bright DAPI-staining material, likely reflecting<br />

leaked bacterial DNA due to its proximity to irregular-shaped bacterial remnants (Fig.<br />

23A). Regular shaped bacteria stained dimly with DAPI. Infections with F. tularensis<br />

pre-labeled with Hoechst 33342 nucleic acid stain confirmed that AIM2 was recruited<br />

to bacterial DNA (Fig. 23B). Merged images revealed that AIM2 overlapped almost<br />

completely with the bacterial DNA (Fig. 23A, 23B, Fig. 25A), indicating that DNA<br />

leaked from F. tularensis likely is the PAMP recognized by AIM2 during an infection.<br />

Consistent with this notion, IL-1b secretion from wild-type BMDMs primed with<br />

Pam3CSK4 and then transfected with F. tularensis extract was abolished when the<br />

extract was pre-incubated with DNase I (Fig. 24A). Purified F. tularensis DNA<br />

transfected into Pam3CSK4-primed BMDMs also stimulated AIM2-dependent IL-1β<br />

secretion (Fig. 24B).<br />

104


- Figure 23 - AIM2 and ASC form a complex with F. tularensis DNA.<br />

A, Immunofluorescence microscopy of F. novidica U112-infected BMDMs at 5.5 h post-infection.<br />

Scale bars, 10 mm. Differential interference contrast, DIC. Arrows indicate co-localization of DNA,<br />

degraded bacteria, AIM2, and ASC. Asterisks label diffuse AIM2 accumulation with DNA. Images are<br />

representative of at least 3 independent biological replicates. B, BMDMs were infected with F.<br />

105


novidica U112 pre-labeled with Hoechst 33342 nucleic acid stain. Upper panel scale bar, 10 mm; lower<br />

panel, 2 mm. Arrows and asterisks indicate co-localization of bacterial DNA and AIM2.<br />

106


- Figure 24 - F. tularensis DNA triggers IL-1β secretion.<br />

A, Extract from F. tularensis ssp. novicida strain U112 was prepared in 10 mM Tris-HCl pH 7.5 with a<br />

French press. Extract corresponding to 2x10 6 cfu was treated with nothing (cont), proteinase K<br />

(Qiagen) for 16 h at 55°C, or DNase I (Qiagen) for 2h at 25°C, and then transfected into BMDMs<br />

primed with 500 ng/mL Pam3CSK4 for 5 h. IL-1β secreted into the culture supernatant was assayed 3<br />

h later. B, � = aim2 -/- , � = wt. Pam3CSK4-primed BMDMs were transfected with 1 mg/mL F.<br />

tularensis DNA, which was isolated from bacteria with a Qiagen DNeasy Blood and Tissue Kit. IL-1β<br />

secreted into the culture supernatant was assayed 16 h later. Graphs show the mean ± standard<br />

deviation of triplicate wells and are representative of 3 independent experiments.<br />

3.3.4 AIM2 is required for the formation of an ASC focus.<br />

Despite multiple AIM2 specks forming in an infected cell adjacent to bacterial<br />

remnants, ASC was recruited to a single AIM2 speck in ~15-22% of wild-type or<br />

caspase-1 -/- BMDMs (Fig. 23A, Fig. 25B). The vacuole-restricted F. tularensis<br />

107


mutant ΔFPI did not stimulate ASC focus formation, consistent with its inability to<br />

stimulate IL-1β secretion (Fig. 22A). Importantly, ASC focus formation required<br />

AIM2 because no foci were detected in aim2 -/- BMDMs. We, and others, have shown<br />

that similar ASC foci are formed in macrophages infected with S. typhimurium upon<br />

activation of NOD-like receptors ((49); P.B., K.N., M. Lamkanfi, S. Mariathasan,<br />

V.M.D., and D.M.M., manuscript in preparation). Our data suggests that although<br />

AIM2 recognizes cytosolic DNA at multiple sites, only one of these sites will form the<br />

platform on which the ASC-containing inflammasome is built.<br />

108


- Figure 25 - AIM2 is required for the formation of an ASC focus.<br />

A, Three-dimensional reconstruction of a confocal image taken of a wild-type BMDM from (c). Scale<br />

bar, 0.5 mm. B, � = F. tularensis, � = ΔFPI. Graph showing the percentage of infected BMDMs<br />

containing an ASC focus in (Fig. 23A). Bars represent the mean ± standard deviation of 2 independent<br />

experiments. At least 300 cells of each genotype were examined per infection.<br />

3.3.5 Type I IFN increases AIM2 protein levels and inflammasome activity.<br />

Given that AIM2 protein expression is increased in BMDMs treated with IFN-<br />

β (Fig. 18C) or infected with F. tularensis (Fig. 26A) and that type-I IFN signaling is<br />

required for efficient inflammasome signaling in response to F. tularensis (64), we<br />

sought to delineate the signaling pathway(s) driving type-I IFN and AIM2 synthesis<br />

109


after F. tularensis infection. We investigated the contribution of STING because it<br />

complexes with TBK1 and mediates type-I IFN production in response to DNA (77).<br />

Unlike wild-type BMDMs, which synthesized IFN-β mRNA in response to F.<br />

tularensis, but not the avirulent mutant ΔFPI, sting -/- BMDMs did not upregulate IFN-<br />

β gene expression after infection (Fig. 26B). sting -/- BMDMs synthesized IFN-β<br />

mRNA normally in response to transfected poly I:C, which engages RIG-I and MDA5<br />

(82), excluding a general defect in IFN-β transcription. We speculate that bacterial<br />

DNA leaked from lysing F. tularensis leads to STING-dependent Type-I IFN<br />

production, although the sensor that recognizes the DNA remains unknown. Just as<br />

AIM2 was not required for IFN-β secretion from BMDMs transfected with dsDNA<br />

(Fig. 19D), AIM2 deficiency did not compromise IFN-β secretion from BMDMs<br />

infected with F. tularensis (Fig. 27A).<br />

110


- Figure 26 - Type-I IFN increases AIM2 protein levels and inflammasome activity.<br />

� = ifnar -/- , � = sting -/- , � = wt. BMDMs were infected with S. typhimurium (moi = 100), F. tularensis<br />

ssp. novidica strain U112, or isogenic mutant ΔFPI for 5 h. (multiplicity of infection, moi). A, Western<br />

blot of AIM2 protein expression. B, IFN-β mRNA expression quantified by RT-PCR. As a control,<br />

BMDMs were transfected with 1 mg/mL polyI:C. C, IL-1β secretion into the culture supernatant. D,<br />

Cytotoxicity as measured by LDH release. Where indicated, BMDMs were treated with 1000 U/mL<br />

recombinant IFN-β at 1 h post-infection. Graphs show the mean ± standard deviation of triplicate wells<br />

and are representative of 3 independent experiments.<br />

111


The increased AIM2 expression observed in F. tularensis-infected wild-type<br />

BMDMs was not observed in sting -/- or type-I IFN receptor-deficient ifnar -/- BMDMs<br />

(Fig. 26A). In addition, failure to upregulate AIM2 correlated with abrogated IL-1b<br />

secretion (Fig. 26C) and reduced cell death (Fig. 26D). Forced expression of AIM2 in<br />

ifnar -/- BMDMs restored IL-1β secretion in response to F. tularensis (Fig. 27B) and<br />

exogenous IFN-β restored cell death in infected sting -/- , but not ifnar -/- BMDM<br />

cultures (Fig. 26D). We conclude that STING-dependent type-I IFN production<br />

boosts inflammasome activity during F. tularensis infection by increasing AIM2<br />

expression.<br />

112


- Figure 27 - Type-I IFN facilitates formation of the AIM2-containing inflammasome.<br />

A, � = asc -/- , � = caspase-1 -/- , � = aim2 -/- , � = wt. IFN-β secretion by BMDMs that were untreated<br />

(cont) or infected with F. tularensis ssp. novicida strain U112 and isogenic mutant ΔFPI for 5 h.<br />

multiplicity of infection, moi. B, � = ifnar -/- + AIM2/GFP, � = ifnar -/- + GFP. ifnar -/- bone marrow<br />

was transduced with pMSCV2.2-IRES-GFP encoding mouse AIM2 or the empty parental vector, after<br />

retroviral particles were generated with the Phoenix.Eco packaging cell line. Three days later,<br />

macrophages were differentiated with M-CSF-containing medium, and GFP-positive cells were sorted<br />

in a FACS Aria (Becton Dickinson). These BMDMs were primed with 500 ng/mL Pam3CSK4 for 16 h<br />

and then infected with F. tularensis or treated with 5 mM ATP. IL-1β secreted into the supernatant was<br />

assayed after 5 h. Graphs show the mean ± standard deviation of triplicate wells and are representative<br />

of 3 independent experiments.<br />

113


3.3.6 AIM2 is required for host defense against F. tularensis.<br />

Finally, to extend our findings on the role of AIM2 in cultured macrophages to<br />

an in vivo setting, we challenged wild-type, aim2 -/- , and caspase-1 -/- mice with F.<br />

tularensis. caspase-1 -/- mice fail to control F. tularensis infections (97), and aim2 -/-<br />

mice were equally impaired at limiting F. tularensis replication (Fig. 28). Average<br />

bacterial loads in liver, lung, and spleen of aim2 -/- or caspase-1 -/- mice at 36 hours<br />

post-infection were 120- to 19,000-fold higher than in wild-type mice. These data<br />

demonstrate that AIM2 is essential for innate immunity to F. tularensis in vivo.<br />

114


- Figure 28 - AIM2 is required for host defense against F. tularensis<br />

Mice were infected intradermally with 1x10 5 colony forming units (cfu) of F. tularensis ssp. novidica<br />

strain U112F. Organs were harvested after 36 h, homogenized, and cfu were determined by plating<br />

serially diluted tissue extracts on modified MH agar. Bars indicate the geometric mean cfu per<br />

genotype.<br />

3.4 Discussion<br />

Collectively, our results provide genetic evidence that AIM2 is an essential<br />

DNA sensor of the innate immune system. Furthermore, AIM2 plays a critical role in<br />

defense against F. tularensis. To our knowledge we also provide the first visualization<br />

of an endogenous inflammasome NLR complexed with its ligand in the context of an<br />

infection. In our model, F. tularensis escapes phagosomal degradation into the<br />

macrophage cytosol where some bacteria lyse, releasing DNA into the cytosol. We<br />

support this model with visualization of Hoechst pre-labeled bacterial DNA observed<br />

115


outside of aberrantly shaped F. tularensis. An unknown sensor(s) recognizes cytosolic<br />

bacterial DNA and signals through the adapter STING to produce Type-I IFN.<br />

Autocrine and paracrine signaling through IFNAR leads to an increase in AIM2<br />

protein levels, which accelerates recognition of bacterial DNA by AIM2. We observe<br />

colocalization of AIM2 specks with bacterial DNA, one of which acts as a nucleus for<br />

formation of an ASC focus. We believe this complex represents a mature<br />

inflammasome that leads to secretion of mature IL-1β and host cell death. This model<br />

represents coordination of two independent DNA sensing pathways to produce a<br />

complete host response to a bacterial infection.<br />

Previous DNA transfection studies have demonstrated inflammasome<br />

activation in the absence of type-I IFN signaling, suggesting that endogenous levels of<br />

AIM2 are sufficient for recognition of transfected DNA (119). Although we observe a<br />

dependence on type-I IFN signaling for F. tularensis inflammasome activation, we<br />

demonstrate that we can restore inflammasome activation in the absence of type-I IFN<br />

signaling through exogenous expression of AIM2, suggesting that AIM2 is sufficient<br />

for recognition of F. tularensis. Considering these results we hypothesize that either<br />

the mechanism of DNA delivery or concentration of DNA delivered during F.<br />

tularensis infection is insufficient to be recognized by the endogenous levels of AIM2<br />

present in macrophages. The sequential activation of the type-I IFN and<br />

inflammasome responses also lead us to speculate that the threshold of DNA required<br />

to induce type-I IFN signaling is less than that required to induce AIM2<br />

inflammasome activation.<br />

116


Bacterial pathogens can exploit a wide range of niches within a host yet very<br />

few bacteria replicate inside the host cytosol, namely L. monocytogenes, Shigella<br />

flexneri, and F. tularensis. The inflammasome has been implicated in innate<br />

immunity to all of the aforementioned pathogens, although different NLRs, and<br />

different bacterial ligands mediate these events. Inflammasome activation by L.<br />

monocytogenes has been attributed to production of the pore-forming toxin LLO,<br />

which is recognized by the Nalp3 inflammasome, as well as bacterial flagellin, which<br />

is engaged by the IPAF inflammasome (158, 174). Recent studies have elucidated a<br />

shared motif between bacterial flagellin and components of the type III secretion<br />

system (T3SS) (115), which explains the ability of IPAF to recognize S. flexneri,<br />

which contains a T3SS and lacks flagellin. It is not clear whether the AIM2<br />

inflammasome senses either of these pathogens. On the other hand F. tularensis is<br />

non-flagellated, lacks a T3SS, and has not been shown to produce pore-forming<br />

toxins. Thus, DNA seems to be the only known inflammasome ligand possessed by F.<br />

tularensis.<br />

Cross-talk between the type-I IFN, NF-κB, and inflammasome pathways is<br />

poorly understood. We observed enhanced TNF-α and IFN-β secretions in asc-/-<br />

compared to wild-type BMDMs when transfected with poly(dA:dT) or pcDNA3 but<br />

not when transfected with poly(dG:dC) or ISD. One possible explanation for this<br />

difference is that there are at least two DNA sensors for type-I IFN (and TNF-α),<br />

which recognize different types of DNAs. One such receptor-mediated pathway may<br />

receive negative feedback from the inflammasome, while the other does not.<br />

Additionally, type-I IFN positively regulates the AIM2 inflammasome in response to<br />

117


F. tularensis infection. Although a role for the AIM2 inflammasome has not been<br />

demonstrated during Listeria monocytogenes infection it should be noted that type-I<br />

IFN signaling also accelerates inflammasome activation in response to this pathogen<br />

(156). AIM2 is also likely required for recognition of dsDNA viruses, such as<br />

Vaccinia virus (68). Interestingly, Vaccinia actively inhibits the antiviral effects of<br />

type-I IFN with the viral E3L protein, which may delay inflammasome activation and<br />

hence the innate immune response to the pathogen (25). In addition, AIM2 may<br />

contribute to aberrant IL-1β production in response to host DNA, leading to arthritis-<br />

like autoimmune disease pathology. The coordination between the type-I IFN<br />

response and the AIM2 inflammasome in the context of pathogen infection and auto-<br />

immunity warrants further investigation and is likely to have broad significance in our<br />

understanding of innate immunity.<br />

118


<strong>Chapter</strong> 4: Discussion<br />

Francisella is able to infect several cell types in vitro, and has been<br />

associated with epithelial cells and hepatocytes in vivo, as well as its well-<br />

characterized infection of immune cells like macrophages and dendritic cells. Only<br />

certain subsets of host cells express inflammasome components, namely macrophages,<br />

dendritic cells, natural killer cells, and epithelial cells. The observation that mice<br />

lacking inflammasome components are unable to control the bacterial burden<br />

emphasizes the importance of the inflammasome in promoting host defense against<br />

bacterial challenge (97). Inflammasome activation has been observed upon infection<br />

with F. novicida, and F. tularensis LVS, but not with the most virulent F. tularensis<br />

type A. Instead of activation and inflammasome dependent cell death, type A<br />

activates a caspase-3 dependent apoptosis. Furthermore, F. tularensis type A actively<br />

suppresses pulmonary dendritic cells and macrophages in the lung following aerosol<br />

challenge (15). This leads to a suppression of early proinflammatory cytokines such<br />

as TNF-α, IL-1β, and IL-12. Furthermore, this strain seems to promote an anti-<br />

inflammatory environment by inducing cytokines like TGF-β, which may aid the<br />

bacterium in subverting host defenses. It is interesting to speculate that type A strains<br />

are more virulent as a result of their ability to either suppress or avoid inflammasome<br />

activation. Given the high degree of genome sequence identity across all Francisella<br />

strains we set out to identify bacterial genes that might play a role in subverting host<br />

defenses, as well as identify the bacterial ligand(s) responsible for activating host<br />

responses by screening an F. novicida library. We decided to use a bacterial genetic<br />

119


screen because similar methods had proved useful in identifying flagellin as the<br />

inflammasome activating ligand of Salmonella species (113). Since flagellin is not an<br />

essential molecule of Salmonella, and does not affect its ability to replicate in<br />

macrophages we hypothesized that we would be able to identify a Francisella mutant<br />

that would replicate intracellularly as efficiently as wild-type, but fail to activate the<br />

immune responses. This turned out to be not the case. To our surprise, our screen in<br />

chapter 2 identified over 200 bacterial genes that modulated the host response, from<br />

several different functional classes. In reviewing the list it became apparent that there<br />

was a connection between the bacterium ability to reach the cytosol and induction of<br />

the host response. Also, we noticed that an abundance of mutants with potential outer<br />

membrane/LPS defects hyper-stimulated immune responses. The characterization of<br />

these mutants in chapter 2 was difficult to interpret due to the pleotropic effects<br />

resulting from deleting of outer membrane components. In retrospect, isolation of<br />

theses mutants was giving us a clue that an unstable outer membrane would lead to<br />

increased release of DNA, which we would later identify as the stimulator of the<br />

cytosolic responses. Therefore, the suppression of the cytosolic response that we<br />

attributed to these genes was likely due to their role in maintaining cell wall integrity<br />

and not likely due to direct inhibition of host pathways. These results suggest that the<br />

most virulent type-A strain must have other mechanisms to avoid inflammasome<br />

activation, either with a tougher cell wall, or with other genes that block DNA<br />

recognition.<br />

F. tularensis is a water pathogen and not adapted to the mammalian host<br />

therefore it lacks genes that specifically target mammalian innate immune pathways.<br />

120


In contrast, viruses like Vaccinia virus actively block type-I IFN production. Also,<br />

host adapted bacteria like Shigella flexneri, and Salmonella typhi secrete effector<br />

proteins that target and subvert host pathway to establish a replicative niche. Many F.<br />

tularensis genes are annotated as hypothetical, with no homology to proteins of known<br />

function. The ability of F. tularensis to infect mammalian macrophages likely arises<br />

from its ability to replicate inside of amoebae. A similar notion could be applied to<br />

studies of Legionella species, which are also natural parasite of amoeba and only<br />

infect humans accidentally. It may be that to truly understand the pathogenic<br />

functions of several F. tularensis genes we may need to look at it’s interactions with<br />

fresh water amoeba, which share a few characteristics of macrophages but will present<br />

distinct survival challenges for the bacterium. Though some work has been done in<br />

Acanthamoeba castellanii, Dictostylium may serve as a better tool because of it<br />

genetic tractability. Studies on the interaction of F. tularensis with amoeba could<br />

greatly enhance our understanding of its pathogenic strategies and mechanisms by<br />

which it persists in environmental reservoirs.<br />

The findings in chapter 3 suggest that bacterial lysis in the cytosol leads to<br />

release of F. tularensis DNA, induction of the type-I IFN pathway, and activation of<br />

the AIM2 inflammasome. The molecular mechanism that causes bacterial lysis is<br />

unclear but several hypotheses emerge from this work. First, the observation that<br />

induction of the cytosolic responses requires bacterial protein and DNA synthesis<br />

suggests that lysis in the macrophage cytosol is a bacterial mediated process. In broth<br />

culture in rich media, bacteria exhibit a life cycle characterized by a lag phase of no<br />

bacterial replication, a log phase with a net increase in bacterial multiplication, a<br />

121


stationary phase of limited nutrient availability where replication plateaus, and a death<br />

phase with a net decrease in bacterial numbers. During this death phase many bacteria<br />

lyse, releasing their contents into the culture media. Therefore, bacterial lysis is a<br />

natural part of the bacterial life cycle and this could be the mechanism of DNA release<br />

in the host cytosol. Additionally, the macrophage phagosome is a professional<br />

microbe-killing machine, and although F. tularensis is well equipped to escape with<br />

its life that does not mean that it is not wounded in the battle. Studies to date on the<br />

intracellular trafficking of the Francisella containing vacuole suggest that the bacteria<br />

escape before acquiring markers of lysosomes or degradative enzymes (26, 144, 145).<br />

However, slight perturbations in the bacterial envelope during the vacuolar stage may<br />

be enough to induce lysis once the bacteria reach the cytosol. This hypothesis is<br />

supported by the large number genes involved in outer membrane and LPS synthesis<br />

identified in <strong>Chapter</strong> 2 that led to an increased cytosolic response in the macrophage.<br />

If these mutants have an unstable outer membrane they may lyse at a higher frequency<br />

than wild-type F. novicida and lead to increased cytosolic sensing by the DNA<br />

pathway. In support o this idea, recent reports show that Listeria monocytogenes lyses<br />

at a low frequency in the macrophage cytosol and induces AIM2-dependent<br />

inflammasome activation (146). Furthermore, L. monocytogenes mutants that induced<br />

higher levels of inflammasome activation were shown to lyse with increased<br />

frequency than wild-type L. monocytogenes.<br />

Yet another hypothesis is the existence of antimicrobial defenses in the cytosol<br />

itself. Little is known about the cytosolic environment except that it is pH neutral.<br />

Also we known little about changes to this environment after macrophages are<br />

122


stimulated with proinflammatory cytokines such as interferons. The interferons were<br />

first described for their ability to induce and antiviral state in cells, but whether or not<br />

this includes defenses in the cytosol is unknown. IFN-γ induces bacterial killing in the<br />

cytosol (42), and INF-β induces a similar transcriptional response as IFN-γ in<br />

macrophages so we hypothesize that type-I IFNs may induce bacterial lysis itself.<br />

These hypotheses are not mutually exclusive and it’s likely that multiple mechanisms<br />

contribute to bacterial lysis.<br />

Although much work has been focused on the antimicrobial defenses of<br />

macrophages much of that work has focused on the killing mechanisms of the<br />

phagosome, which is highly efficient at eliminating the majority of invading<br />

organisms. However, for bacteria that escape phagosomal degradation little is known<br />

about potential antimicrobial mechanisms in the host cell cytosol. One mechanism<br />

that has gained interest in recent years is autophagy, which has the potential to<br />

reintroduce cytosolic bacteria into a degradative environment. Autophagy has been<br />

implicated in host defense against Group A Streptococcus (121), Shigella flexneri<br />

(125), and L. monocytogenes (138). The case of Shigella is interesting in that an<br />

autophagy protein, Atg5, is thought to target a specific Shigella protein VirG and<br />

target the bacteria to an autophagosome. Francisella does in fact enter into an<br />

autophagous vacuole during the late stages of macrophage infection (26), though the<br />

molecular mechanism and the impact of this finding on the outcome of the infection<br />

are unclear.<br />

DNA release is detrimental to the bacterium in this infection model but it may<br />

be beneficial to the bacterium in other environments. Francisella is a competent<br />

123


organism so release of bacterial DNA may serve as a way to exchange genetic<br />

information and drive evolution. Furthermore, recent work has demonstrated the<br />

ability of Francisella to establish biofilms (95), which are bacterial communities<br />

characterized by subpopulations in heterogeneous metabolic states. Biofilms often<br />

contain dead bacteria, which can provide a source of nutrients for living populations.<br />

Furthermore, Pseudomonas aeruginosa release bacterial DNA, which forms an<br />

adhesive network to promote biofilm structure (6). Additionally, type-IFN signaling<br />

resulting from the release of bacterial DNA is beneficial to the bacterium in vivo (65).<br />

Therefore, lysis of a small percentage of bacteria may benefit the overall bacterial<br />

community, much like pyroptosis of individual macrophages is important to survival<br />

of the host during infections.<br />

In this thesis work we have identified AIM2 as a critical component of the<br />

innate immune response to F. tularensis infection. The Alnemri group (48)<br />

independently came to the same conclusion, thus supporting our results. Recent repots<br />

have also shown that AIM2 is critical for host defense against viral infection, and<br />

plays a role in detecting intracellular L. monocytogenes (85, 140, 181). Clearly, our<br />

understanding of the role of AIM2 in innate immunity, as well as innate immune<br />

mechanisms for sensing intracellular pathogens is just beginning. It’s interesting that<br />

the AIM2 inflammasome does not recognize a molecule that is unique to microbes,<br />

but instead recognizes any dsDNA with a minimum length of 45 base pairs. It is the<br />

aberrant localization of dsDNA in the cytosol that acts as the danger signal rather than<br />

the molecule itself. It is interesting that IPAF, which recognizes the bacterial specific<br />

flagellin monomers, is able to recruit caspase-1 directly, while other NLRs (AIM2,<br />

124


NLRP3, and NLRP1), which responds to host and microbial molecules (DNA, ATP,<br />

gout crystals, and LLO share NLRP3) require the adapter protein ASC. This may<br />

provide another level of regulation to prevent aberrant inflammation from the<br />

recognition of self. In any case, the host innate immune response has used the<br />

compartmentalization of signals as additional method to distinguish safety from<br />

danger.<br />

Innate immune responses have mostly been studied with purified ligands such<br />

as LSP, MDP, or DNA, but studies of immune recognition in the context of an<br />

infection is lacking. In this thesis work we demonstrate innate immune recognition of<br />

bacterial DNA in the context of an intracellular infection. This brings out several<br />

differences from previous reports using DNA transfection. First, in chapter 3 and in<br />

previous reports (34, 64) we see a strong dependence on type-I IFN signaling for<br />

inflammasome activation in response to cytosolic F. tularensis infection. In contrast,<br />

type-I IFN signaling is dispensable for inflammasome activation in response to<br />

transfected dsDNA (119). The difference may be in the concentration of DNA<br />

delivered by the two methods. During a F. tularensis infection, we observe<br />

logarithmic intracellular replication by the bacteria, and thus we only observe a small<br />

percentage of bacterial lysis events leading to DNA release. Each rare lysis event will<br />

release approximately 4-5 femtograms of bacterial DNA, based on the genome size of<br />

F. tularensis. This is a relatively small intracellular concentration of DNA compared<br />

to studies with transfected dsDNA, where concentrations of 1µg/mL are typically<br />

used. In chapter 2, we demonstrate that lower concentration of transfected dsDNA<br />

lead to a type-I IFN dependent host cell death. But we also show that mutants that<br />

125


hyper-induce the cytosolic response can activate a type-I IFN-independent cell death,<br />

possible because they release more DNA than wild-type F. tularensis. Thus, the<br />

coordination between the type-I IFN response and the inflammasome may be a factor<br />

of the concentration of cytosolic DNA. We cannot however rule out the importance of<br />

other differences between the two stimuli, such as DNA size and charge. We do, in<br />

fact, observe different degrees of type-I IFN and inflammasome activation with<br />

poly(dG:dG) and poly(dA:dT). Thus, the nature of the DNA may dictate the host<br />

receptor that produces type-I IFN, and different receptors may assert different effects<br />

on the inflammasome.<br />

We observed a positive feedback loop between type-I IFN signaling and<br />

inflammasome activation in macrophages, but the link in vivo seems to be less clear.<br />

Shown in both previous reports (97) and in chapter 3, mice deficient in inflammasome<br />

components are more susceptible to infection with F. tularensis. However, mice<br />

deficient in type-I IFN signaling are more resistant to infection (65). Similar results<br />

were also obtained for L. monocytogenes infections. This apparent discrepancy<br />

between our in vitro results and the in vivo results is partially explained by the control<br />

of type-I IFN signaling on Il-17 production (65). Mice deficient in type-I IFN<br />

signaling produce more IL-17, which leads to a greater influx of neutrophils that can<br />

better control bacterial infections (65). Additionally, IFN-γ can restore inflammasome<br />

activation in vivo in an IFNAR-deficient mouse by signaling through the IFN-γ<br />

receptor, which would result in increased expression of AIM2 and subsequent<br />

inflammasome activation. Thus an interesting paradox exists, where type-I IFN is<br />

beneficial to the host in vitro and detrimental in vivo during bacterial infections.<br />

126


There is significant crosstalk between the type-I IFN and host cell dearth<br />

pathways. In both chapter 2 and chapter 3 we observed that lack of type-I IFN<br />

signaling results in decreased inflammasome activation, with STING serving as a<br />

critical adaptor for the production of type-I IFN downstream of DNA sensing.<br />

Additionally, we observe an increase in type-I IFN production in macrophages<br />

deficient for inflammasome components that are stimulated with dsDNA. These<br />

results suggest not only positive feedback but also negative feedback loops between<br />

the two pathways. To better understand the molecular mechanisms of the coordination<br />

between these two pathways we need to identify the host receptors that lead to type-I<br />

IFN production. One such receptor, DAI (161), is either not active in macrophages, or<br />

there are redundant receptors since DAI-deficient macrophages and mice respond<br />

normally to stimulation with dsDNA (75). However, a new cytosolic DNA sensor has<br />

been identified, LRRFIP1 (183), that mediates type-I IFN production in macrophages<br />

in response to L. monocytogenes and vesicular stomatitus virus. If LRRFIP1 were<br />

also involved in the macrophage type-I IFN response to F. tularensis we would have a<br />

system to study the coordination of the IFN pathway and inflammasome pathway in a<br />

biologically relevant model.<br />

Recent work has clearly demonstrated the importance of inflammasome in<br />

pathogen detection and innate immunity to infection. However, we can use these<br />

pathogens as tools to better understand aberrant inflammation associated with<br />

autoimmunity and auto-inflammatory diseases. Future studies on AIM2 could help us<br />

understand the mechanisms behind systemic lupus erythematosus (SLE). Thus F.<br />

127


tularensis may a useful tool for probing innate immune responses and better<br />

understanding host biology.<br />

128


<strong>Chapter</strong> 5: Materials and Methods<br />

5.1 BEIR “two-allele” transposon library screen.<br />

A sequenced two-allele transposon mutant library was used to test for F. novicida<br />

transposon mutants that elicited increased or decreased type-I IFN and macrophage<br />

cell death responses (the following reagent was obtained through the NIH Biodefense<br />

and Emerging Infections Research Resources Repository, NIAID, NIH: F. tularensis<br />

subsp. novicida, “Two-Allele” Transposon Mutant Library Plates 1-14, 16-32). The<br />

library represents two or more transposon insertions in all non-essential genes. At the<br />

time of screening, Plate 15 of the library was unavailable due to quality control issues,<br />

resulting in a library size of 2,954 mutants. The two-allele library was received frozen<br />

in 96-well format. The two-allele library was grown overnight in 96-well plates in<br />

TSB supplemented with 0.2% L-cysteine at 37°C with aeration. Bone marrow-derived<br />

macrophages from C57BL/6 mice were seeded in 96-well plates at a density of 10 5<br />

macrophages per well and cultured overnight at 37°C with 5% CO2. Individual<br />

transposon mutants were diluted into complete macrophage media and used to infect<br />

BMDM at an MOI of ~500:1. Infected macrophages were centrifuged for 15min at<br />

730 x g and incubated for 30min at 37°C. The infected media was then removed and<br />

replaced with fresh complete macrophage media containing 10ug/mL gentamicin and<br />

further incubated at 37°C with 5%CO2 for the duration of the experiment. At 4.5hrs<br />

post infection 75uL of macrophage supernatant was collected and frozen at -80°C.<br />

The amount of type-I interferon in the supernatant was later determined using the<br />

reporter cell line ISRE-L929. At 6.5 hrs post-infection 50uL of supernatant was<br />

129


collected and assayed for macrophage cell death by CytoTox96 non-radioactive<br />

cytotoxicity assay (Promega) according to the manufacturers instructions.<br />

5.2 Bacterial strains and growth conditions.<br />

Francisella novicida U112 and all mutant strains were grown to stationary phase in<br />

TSB supplemented with 0.2% cysteine at 37°C with aeration.<br />

5.3 Bacterial Mutagenesis.<br />

Targeted deletions were generated in the U112 strain by splicing by overlap extension<br />

(SOE) PCR (Liu, 2007), as previously described (Brotcke,, 2008) using the primers in<br />

Table A1. Briefly, SOE PCR was used to generate a construct containing a<br />

kanamycin resistance cassette expressed by the groEL promoter flanked by ~600bp<br />

regions of the chromosome 5' and 3' to the gene of interest. The resulting PCR<br />

product was transformed into U112 by chemical transformation and transformants<br />

were selected on MMH agar with 30µg/ml kanamycin. Gene deletions were<br />

confirmed by sequencing. Bacterial mutants were complemented in trans by deletion<br />

mutants were complemented in trans by introducing the wild-type gene, as well as the<br />

CAT cassette, into gro-gfp pFNLTP6 (Maier, 2004). Complementation constructs<br />

were generated by digestion of the pFNLTP6 at the NotI and BamHI sites, removing<br />

the gfp gene, and ligation of the vector with the gene of interest. The resulting<br />

plasmid expressed the complementing gene under the regulation of the constitutive<br />

groEL promoter. Complemented strains were selected for growth on 5µg/ml<br />

130


chloramphenicol and also confirmed by sequencing. All complementation primers are<br />

listed in Table A1.<br />

5.4 Bone marrow-derived macrophage culture and infections.<br />

Bone marrow-derived macrophages were prepared from mouse femurs and cultured in<br />

DMEM supplemented with 10% heat-inactivated fetal calf serum and 10%<br />

conditioned L929-MCSF supernatant at 37°C with 5% CO2. For cytokine<br />

measurements and cytotoxicity assays 10 5 macrophages per well were seeded into 96-<br />

well TC plates (BD) and bacteria were diluted into macrophage media to reach an<br />

MOI of 10 bacteria per macrophage. Macrophages were centrifuged at 730 x g for 15<br />

minutes and incubated at 37°C with 5% CO2 (Time zero). After 30 minutes infected<br />

media was removed and replaced with macrophage media containing10ug/mL<br />

gentamicin for the duration of the experiment. For intracellular replication assays 2.5<br />

x 105 macrophages were seeded in 24-well TC plates and infected as described above.<br />

At the indicated timepoints macrophages were washed 3 times with PBS, lysed in 1%<br />

saponin solution, and serial dilutions were plated on MMH agar for determination of<br />

cfu. Bone marrow cells were stimulated or infected with S. typhimurium or F.<br />

tularensis as described (97). dsDNAs and polyI:C were transfected with<br />

Lipofectamine 2000 (Invitrogen). For priming, BMDMs were cultured with 500<br />

ng/mL ultra-pure LPS or Pam3CSK4 (Invivogen) for 5 h. Priming was used to induce<br />

production of pro-IL-1β in macrophages. Priming with either LPS or Pam3CSK4<br />

induces similar levels of pro-IL-1β, however, Pam3CSK4 does not induce Trif-<br />

131


dependent IFN-I production as observed with LPS. IL-1β (Meso), IFN-β (PBL), and<br />

IL-18 (MBL) were measured by ELISA, TNF-a and IL-6 by Bioplex-23 cytokine<br />

assay (Bio-Rad). Cytotoxicity was measured by LDH release (Promega). Caspase-1<br />

was immunoblotted with 4B4 rat anti-mouse caspase p20 (Genentech) or rabbit anti-<br />

caspase p10 (Santa Cruz, sc-514), IL-1b with a rabbit polyclonal (GeneTex).<br />

5.5 Macrophage gene expression analysis.<br />

Bone marrow-derived macrophages were seeded at 10 6 macrophages per well in a 6-<br />

well dish and infected at an MOI of 10:1 with the appropriate strain. At the timepoints<br />

indicated infected macrophages were incubated with 1mL of TRIzol reagent<br />

(Invitrogen) for 5 minutes with shaking and frozen at -80°C. RNA was isolated using<br />

the RNEasy mini kit (Quiagen) as per the manufacturers instructions. IFN-β, pro IL-<br />

1β, and β-actin gene expression was determined by real time qRT-PCR analysis. Real-<br />

time Quantitative RT-PCR was performed on a 7300 Real Time PCR system (Applied<br />

Biosystems) using rTth enzyme (Applied Biosystems), SYBR green, and primers for<br />

IFN-β, pro IL-1β, and β-actin. Gene specific transcript levels were normalized to the<br />

level of β-actin mRNA. Primers used for IFN-β mRNA quantification are described<br />

(8). Experiments were performed with an iCycler (Bio-Rad) using SYBR green<br />

(Applied Biosystems).<br />

132


5.6 ISRE-L929 assays.<br />

ISRE-L929 cells were seeded at a density of 10 5 cells per well in 96-well viewplate 96<br />

TC plates (Perkin Elmer) and cultured in DMEM supplemented with 10% heat-<br />

inactivated fetal calf serum overnight at 37°C with 5% CO2. ISRE cells were<br />

incubated with 60uL of macrophage supernatant for 4 hours at 37°C with 5% CO2.<br />

After 4 hours the supernatant was removed and the ISRE cells were lysed with 40uL<br />

per well of Glo lysis buffer (Promega) for 7 minutes with shaking. Bright Glow<br />

luciferase reagent (Promega) was added at 40uL per well and the plates were read<br />

immediately on a luminometer using the Veritas software.<br />

5.7 NF-kB reporter cell assays.<br />

RAW 264.7 macrophages containing a NF-kb luciferase reporter were seeded at a<br />

density of 10 5 cells per well in 96-well luminometer plates (Nunc). RAW cells were<br />

infected at an MOI of 10:1 as described above and luciferase was measured at 2hrs<br />

post-infection using the Bright Glow luciferase assay (Promega).<br />

5.8 Mice, bacteria, and reagents<br />

asc -/- , caspase-1 -/- , ifnar -/- , ipaf -/- , nalp3 -/- , and sting -/- mice have been described (77,<br />

96, 98, 118). nalp3 -/- ipaf -/- mice were generated by nalp3 -/- and ipaf -/- intercrosses. All<br />

mice were backcrossed to C57BL/6 for at least 10 generations. Competitive index<br />

experiments were conducted by mixing a 1:1 ratio of wild-type F. novicida strain<br />

U112 and mutant bacteria in PBS, and inoculating the mice intradermally with a total<br />

133


of 10 5 cfu in 50uL of bacterial suspension. After 48 hours mice were sacrificed, skin<br />

and spleen samples were homogenized in sterile PBS, and serial dilutions were plated<br />

on MMH agar and MMH agar supplemented with 30ug/mL kanamycin. The<br />

competitive index was calculated as the ratio of mutant to U112 after 48 hours relative<br />

to the ratio of mutant to U112 in the input. Intradermal infections with F. tularensis<br />

ssp. novicida strain U112 and isogenic mutant ΔFPI (177) were performed as<br />

described (97). The Genentech and <strong>Stanford</strong> <strong>University</strong> animal care and use<br />

committees approved all mouse studies. S. typhimurium was from ATCC. Reagents<br />

included poly(dA:dT), poly(dG:dC), calf thymus DNA, polyI:C, and ATP (Sigma),<br />

pcDNA3.1(+) (Invitrogen), IFN-b (R&D Systems), and Pam3CSK4 (Invivogen).<br />

Sense (5’TAC AGA TCT ACT AGT GAT CTA TGA CTG ATC TGT ACA TGA<br />

TCT ACA) and anti-sense ISD (5’TGT AGA TCA TGT ACA GAT CAG TCA TAG<br />

ATC ACT AGT AGA TCT GTA) (153) were synthesized and annealed at Genentech.<br />

5.9 Immunofluorescence Microscopy<br />

BMDMs (1.25x10 5 ) were seeded onto glass cover slips in 24-well plates for infection.<br />

Where indicated, F. tularensis was incubated for 30 min at 37°C with 50 mg/mL<br />

Hoechst 33342, then washed 7 times prior to infection. Cells were washed twice with<br />

PBS, fixed for 15 min at 37°C with 4% paraformaldehyde in PBS, washed 3 times<br />

with PBS, and stained with rabbit anti-mouse AIM2 at 1/500 (Genentech), 8E4.1 rat<br />

anti-mouse ASC at 1/2000 (Genentech), and chicken anti-Francisella at 1/2000<br />

(Monack Laboratory) for 30 min in blocking buffer (3% BSA, 0.1% Saponin in PBS).<br />

134


Cells were washed 3 times with PBS and incubated 30 min with Alexa488 anti-rat,<br />

Alexa594 anti-rabbit, and Alexa647 anti-chicken antibodies (Invitrogen). Cells<br />

washed 4 times with PBS and stained with DAPI were imaged with a Zeiss LSM700<br />

confocal microscope. DAPI was omitted in samples with Hoechst-stained bacteria.<br />

- Table 4 – Primers for F. novicida cloning and mutagenesis<br />

Primer Sequence<br />

ftn_1212 checkF TAT TGA TAG TGA TGA TTG GG<br />

ftn_1212 F1 TAA GCA AAT AAA AGC TGC TG<br />

GCT TAT CGA TAC CGT CGA CCT CAA TTA ACT TCT AGT AAT TCT<br />

ftn_1212 inv1 TTT<br />

GAT ATC GAT CCT GCA GCT ATG CAA AAT TTT AAG GAA TGA AAT<br />

ftn_1212 inv2 GAA<br />

ftn_1212 R1 AAT GAC TCA ACA TCT GCT AC<br />

ftn_1212 checkR AGC ATT AGC AAT GAC TAT AC<br />

lpcC checkF TTA ATT GGA ACT GTG ATA GC<br />

lpcC F1 TAT CAA TTT CAT GTT CAA CG<br />

GCT TAT CGA TAC CGT CGA CCT CGA TTT ATT TAT ATT AAA ATA<br />

lpcC inv1<br />

TTC<br />

lpcC inv2 GAT ATC GAT CCT GCA GCT ATG CCT AGG TTA TAA GAT TAG CCG<br />

lpcC R1 AAA TGG TAA AGG GCT AGT TG<br />

lpcC checkR AGA GCA AGT CAA ACA AGC TC<br />

wbtA checkF TCG ATT AGA TAA GGC AAA AC<br />

wbtA F1 AAA GCT TGT TGC TAA ACA CC<br />

CGC TTA TCG ATA CCG TCG ACC TCT TGT TAA TTT TTA GAA AAT<br />

wbtA inv1 ATC<br />

GAT ATC GAT CCT GCA GCT ATG CAA TAT ATG AAA GAC AGA ATT<br />

wbtA inv2 TAT<br />

wbtA R1 TAA AAC CTT GCC TTA TCT GC<br />

wbtA checkR AGC ACA AAC ATT ACT CAT CC<br />

FPI checkF AAT CAG CTA TGG ATC GTA GC<br />

FPI f1 TAA TCC ACA GAT ATT ATG CG<br />

AAA TAC GAT GAG TGA CAA CCT GTC TAC TTA ATT AGA ACA TAA<br />

FPI cm inv1 C<br />

AGT GGC AGG GCG GGG CGT AAA CTT ACT ACT CTT ACA AGT AAA<br />

FPI cm inv2 C<br />

FPI r1 TAT GGA AGT TCT GTT TAA CC<br />

FPI checkR AGC AAA CAC TAC AAT TAT TCC<br />

kdsA checkF TGG TGA AGT TAA GGT TTT TG<br />

kdsA f1 ATG ATA GAA GAA ATC GTT GC<br />

GCT TAT CGA TAC CGT CGA CCT CTT GTG ATA ATT ATA CAG AAA<br />

kdsA inv1 AAG<br />

kdsA inv2 GAT ATC GAT CCT GCA GCT ATG CGG CAT AAA TAA TGG CTG GTA<br />

kdsA r1 TTT ATT TTG CGC ACC ATC TG<br />

135


kdsA checkR ACT CTA GCC ATT TTT TAC TC<br />

purMCD checkF TTA TCA TGG GTA GTC ATA CC<br />

purMCD f1 TAT ACT AAT GGG TCA AAT CG<br />

purMCD inv1 cm AAA TAC GAT GAG TGA CAA CCT TTA TTT CCT TTT AAT CAA T<br />

purMCD inv2 cm AGT GGC AGG GCG GGG CGT AAA TAA TGT CTA AGC TAA ATC T<br />

purMCD r1 ATT CTA CGC TCA AAT CGT AG<br />

purMCD checkR ACC CTA TGC TTA AAC TAT AG<br />

purMCD inv1<br />

kanfrt GCT TAT CGA TAC CGT CGA CCT CTT TAT TTC CTT TTA ATC AAT<br />

purMCD inv2<br />

kanfrt GAT ATC GAT CCT GCA GCT ATG CAT AAT GTC TAA GCT AAA TCT<br />

lpxH checkF ATA GAT ATC CTA ACT TAA CC<br />

lpxH f1 TGT ATG TTT ATA GAG TTT GC<br />

lpxH inv1 GCT TAT CGA TAC CGT CGA CCT CAT TTT TGA CGG TAC TGT TTA<br />

lpxH inv2 GAT ATC GAT CCT GCA GCT ATG CTT TAA CTT CAG AGC TGA ATT<br />

lpxH r1 ACT ATA TAG TTC CAT CTG GC<br />

lpxH checkR TTA TGC TTA TAC ATC GTG GC<br />

136


REFERENCES<br />

1. Abd, H., T. Johansson, I. Golovliov, G. Sandstrom, and M. Forsman.<br />

2003. Survival and growth of Francisella tularensis in Acanthamoeba<br />

castellanii. Appl Environ Microbiol 69:600-6.<br />

2. Adachi, O., T. Kawai, K. Takeda, M. Matsumoto, H. Tsutsui, M.<br />

Sakagami, K. Nakanishi, and S. Akira. 1998. Targeted disruption of the<br />

MyD88 gene results in loss of IL-1- and IL-18-mediated function. Immunity<br />

9:143-50.<br />

3. Akira, S., and K. Takeda. 2004. Toll-like receptor signalling. Nat Rev<br />

Immunol 4:499-511.<br />

4. Akira, S., S. Uematsu, and O. Takeuchi. 2006. Pathogen recognition and<br />

innate immunity. Cell 124:783-801.<br />

5. Alexopoulou, L., A. C. Holt, R. Medzhitov, and R. A. Flavell. 2001.<br />

Recognition of double-stranded RNA and activation of NF-kappaB by Tolllike<br />

receptor 3. Nature 413:732-8.<br />

6. Allesen-Holm, M., K. B. Barken, L. Yang, M. Klausen, J. S. Webb, S.<br />

Kjelleberg, S. Molin, M. Givskov, and T. Tolker-Nielsen. 2006. A<br />

characterization of DNA release in Pseudomonas aeruginosa cultures and<br />

biofilms. Mol Microbiol 59:1114-28.<br />

7. Athman, R., and D. Philpott. 2004. Innate immunity via Toll-like receptors<br />

and Nod proteins. Curr Opin Microbiol 7:25-32.<br />

8. Auerbuch, V., D. G. Brockstedt, N. Meyer-Morse, M. O'Riordan, and D.<br />

A. Portnoy. 2004. Mice lacking the type I interferon receptor are resistant to<br />

Listeria monocytogenes. J Exp Med 200:527-33.<br />

9. Ayala, J. M., T. T. Yamin, L. A. Egger, J. Chin, M. J. Kostura, and D. K.<br />

Miller. 1994. IL-1 beta-converting enzyme is present in monocytic cells as an<br />

inactive 45-kDa precursor. J Immunol 153:2592-9.<br />

10. Balagopal, A., A. S. MacFarlane, N. Mohapatra, S. Soni, J. S. Gunn, and<br />

L. S. Schlesinger. 2006. Characterization of the receptor-ligand pathways<br />

important for entry and survival of Francisella tularensis in human<br />

macrophages. Infect Immun 74:5114-25.<br />

11. Barker, J. H., J. Weiss, M. A. Apicella, and W. M. Nauseef. 2006. Basis for<br />

the failure of Francisella tularensis lipopolysaccharide to prime human<br />

polymorphonuclear leukocytes. Infect Immun 74:3277-84.<br />

12. Berdal, B. P., R. Mehl, H. Haaheim, M. Loksa, R. Grunow, J. Burans, C.<br />

Morgan, and H. Meyer. 2000. Field detection of Francisella tularensis. Scand<br />

J Infect Dis 32:287-91.<br />

13. Black, R. A., S. R. Kronheim, J. E. Merriam, C. J. March, and T. P. Hopp.<br />

1989. A pre-aspartate-specific protease from human leukocytes that cleaves<br />

pro-interleukin-1 beta. J Biol Chem 264:5323-6.<br />

14. Bonquist, L., H. Lindgren, I. Golovliov, T. Guina, and A. Sjostedt. 2008.<br />

MglA and Igl proteins contribute to the modulation of Francisella tularensis<br />

137


live vaccine strain-containing phagosomes in murine macrophages. Infect<br />

Immun 76:3502-10.<br />

15. Bosio, C. M., H. Bielefeldt-Ohmann, and J. T. Belisle. 2007. Active<br />

suppression of the pulmonary immune response by Francisella tularensis<br />

Schu4. J Immunol 178:4538-47.<br />

16. Brodsky, I. E., and D. Monack. 2009. NLR-mediated control of<br />

inflammasome assembly in the host response against bacterial pathogens.<br />

Semin Immunol 21:199-207.<br />

17. Brotcke, A., and D. M. Monack. 2008. Identification of fevR, a novel<br />

regulator of virulence gene expression in Francisella novicida. Infect Immun<br />

76:3473-80.<br />

18. Brotcke, A., D. S. Weiss, C. C. Kim, P. Chain, S. Malfatti, E. Garcia, and<br />

D. M. Monack. 2006. Identification of MglA-regulated genes reveals novel<br />

virulence factors in Francisella tularensis. Infect Immun 74:6642-55.<br />

19. Bruey, J. M., N. Bruey-Sedano, F. Luciano, D. Zhai, R. Balpai, C. Xu, C.<br />

L. Kress, B. Bailly-Maitre, X. Li, A. Osterman, S. Matsuzawa, A. V.<br />

Terskikh, B. Faustin, and J. C. Reed. 2007. Bcl-2 and Bcl-XL regulate<br />

proinflammatory caspase-1 activation by interaction with NALP1. Cell<br />

129:45-56.<br />

20. Buddingh, G. J., and F. C. Womack. 1941. Observations on the Infection of<br />

Chick Embryos with Bacterium Tularense, Brucella, and Pasteurella Pestis. J<br />

Exp Med 74:213-222.<br />

21. Burckstummer, T., C. Baumann, S. Bluml, E. Dixit, G. Durnberger, H.<br />

Jahn, M. Planyavsky, M. Bilban, J. Colinge, K. L. Bennett, and G.<br />

Superti-Furga. 2009. An orthogonal proteomic-genomic screen identifies<br />

AIM2 as a cytoplasmic DNA sensor for the inflammasome. Nat Immunol<br />

10:266-72.<br />

22. Butchar, J. P., T. J. Cremer, C. D. Clay, M. A. Gavrilin, M. D. Wewers, C.<br />

B. Marsh, L. S. Schlesinger, and S. Tridandapani. 2008. Microarray<br />

analysis of human monocytes infected with Francisella tularensis identifies<br />

new targets of host response subversion. PLoS One 3:e2924.<br />

23. Cassel, S. L., S. C. Eisenbarth, S. S. Iyer, J. J. Sadler, O. R. Colegio, L. A.<br />

Tephly, A. B. Carter, P. B. Rothman, R. A. Flavell, and F. S. Sutterwala.<br />

2008. The Nalp3 inflammasome is essential for the development of silicosis.<br />

Proc Natl Acad Sci U S A 105:9035-40.<br />

24. Chamaillard, M., S. E. Girardin, J. Viala, and D. J. Philpott. 2003. Nods,<br />

Nalps and Naip: intracellular regulators of bacterial-induced inflammation.<br />

Cell Microbiol 5:581-92.<br />

25. Chang, H. W., J. C. Watson, and B. L. Jacobs. 1992. The E3L gene of<br />

vaccinia virus encodes an inhibitor of the interferon-induced, double-stranded<br />

RNA-dependent protein kinase. Proc Natl Acad Sci U S A 89:4825-9.<br />

26. Checroun, C., T. D. Wehrly, E. R. Fischer, S. F. Hayes, and J. Celli. 2006.<br />

Autophagy-mediated reentry of Francisella tularensis into the endocytic<br />

compartment after cytoplasmic replication. Proc Natl Acad Sci U S A<br />

103:14578-83.<br />

138


27. Chen, W., R. KuoLee, H. Shen, M. Busa, and J. W. Conlan. 2004. Toll-like<br />

receptor 4 (TLR4) does not confer a resistance advantage on mice against lowdose<br />

aerosol infection with virulent type A Francisella tularensis. Microb<br />

Pathog 37:185-91.<br />

28. Chong, A., T. D. Wehrly, V. Nair, E. R. Fischer, J. R. Barker, K. E. Klose,<br />

and J. Celli. 2008. The early phagosomal stage of Francisella tularensis<br />

determines optimal phagosomal escape and Francisella pathogenicity island<br />

protein expression. Infect Immun 76:5488-99.<br />

29. Christenson, B. 1984. An outbreak of tularemia in the northern part of central<br />

Sweden. Scand J Infect Dis 16:285-90.<br />

30. Clemens, D. L., B. Y. Lee, and M. A. Horwitz. 2005. Francisella tularensis<br />

enters macrophages via a novel process involving pseudopod loops. Infect<br />

Immun 73:5892-902.<br />

31. Clemens, D. L., B. Y. Lee, and M. A. Horwitz. 2004. Virulent and avirulent<br />

strains of Francisella tularensis prevent acidification and maturation of their<br />

phagosomes and escape into the cytoplasm in human macrophages. Infect<br />

Immun 72:3204-17.<br />

32. Coers, J., R. E. Vance, M. F. Fontana, and W. F. Dietrich. 2007. Restriction<br />

of Legionella pneumophila growth in macrophages requires the concerted<br />

action of cytokine and Naip5/Ipaf signalling pathways. Cell Microbiol 9:2344-<br />

57.<br />

33. Cole, L. E., K. L. Elkins, S. M. Michalek, N. Qureshi, L. J. Eaton, P.<br />

Rallabhandi, N. Cuesta, and S. N. Vogel. 2006. Immunologic consequences<br />

of Francisella tularensis live vaccine strain infection: role of the innate immune<br />

response in infection and immunity. J Immunol 176:6888-99.<br />

34. Cole, L. E., A. Santiago, E. Barry, T. J. Kang, K. A. Shirey, Z. J. Roberts,<br />

K. L. Elkins, A. S. Cross, and S. N. Vogel. 2008. Macrophage<br />

proinflammatory response to Francisella tularensis live vaccine strain requires<br />

coordination of multiple signaling pathways. J Immunol 180:6885-91.<br />

35. Cole, L. E., K. A. Shirey, E. Barry, A. Santiago, P. Rallabhandi, K. L.<br />

Elkins, A. C. Puche, S. M. Michalek, and S. N. Vogel. 2007. Toll-like<br />

receptor 2-mediated signaling requirements for Francisella tularensis live<br />

vaccine strain infection of murine macrophages. Infect Immun 75:4127-37.<br />

36. Collazo, C. M., A. Sher, A. I. Meierovics, and K. L. Elkins. 2006. Myeloid<br />

differentiation factor-88 (MyD88) is essential for control of primary in vivo<br />

Francisella tularensis LVS infection, but not for control of intra-macrophage<br />

bacterial replication. Microbes Infect 8:779-90.<br />

37. Diaz de Tuesta, A. M., Q. Chow, M. P. Geijo Martinez, J. Dimas Nunez, F.<br />

J. Diaz de Tuesta, C. R. Herranz, and E. Val Perez. 2001. [Tularemia<br />

outbreak in the province of Cuenca associated with crab handling]. Rev Clin<br />

Esp 201:385-9.<br />

38. Dienst, F. T., Jr. 1963. Tularemia: a perusal of three hundred thirty-nine<br />

cases. J La State Med Soc 115:114-27.<br />

39. Dinarello, C. A. 2009. Immunological and inflammatory functions of the<br />

interleukin-1 family. Annu Rev Immunol 27:519-50.<br />

139


40. Dinarello, C. A., and S. M. Wolff. 1993. The role of interleukin-1 in disease.<br />

N Engl J Med 328:106-13.<br />

41. Dostert, C., V. Petrilli, R. Van Bruggen, C. Steele, B. T. Mossman, and J.<br />

Tschopp. 2008. Innate immune activation through Nalp3 inflammasome<br />

sensing of asbestos and silica. Science 320:674-7.<br />

42. Edwards, J. A., D. Rockx-Brouwer, V. Nair, and J. Celli. Restricted<br />

cytosolic growth of Francisella tularensis subsp. tularensis by IFN-gamma<br />

activation of macrophages. Microbiology 156:327-39.<br />

43. Elkins, K. L., S. C. Cowley, and C. M. Bosio. 2007. Innate and adaptive<br />

immunity to Francisella. Ann N Y Acad Sci 1105:284-324.<br />

44. Evans, M. E. 1985. Francisella tularensis. Infect Control 6:381-3.<br />

45. Evans, M. E., D. W. Gregory, W. Schaffner, and Z. A. McGee. 1985.<br />

Tularemia: a 30-year experience with 88 cases. Medicine (Baltimore) 64:251-<br />

69.<br />

46. Feldman, K. A., R. E. Enscore, S. L. Lathrop, B. T. Matyas, M. McGuill,<br />

M. E. Schriefer, D. Stiles-Enos, D. T. Dennis, L. R. Petersen, and E. B.<br />

Hayes. 2001. An outbreak of primary pneumonic tularemia on Martha's<br />

Vineyard. N Engl J Med 345:1601-6.<br />

47. Fernandes-Alnemri, T., J. W. Yu, P. Datta, J. Wu, and E. S. Alnemri.<br />

2009. AIM2 activates the inflammasome and cell death in response to<br />

cytoplasmic DNA. Nature 458:509-13.<br />

48. Fernandes-Alnemri, T., J. W. Yu, C. Juliana, L. Solorzano, S. Kang, J.<br />

Wu, P. Datta, M. McCormick, L. Huang, E. McDermott, L. Eisenlohr, C.<br />

P. Landel, and E. S. Alnemri. The AIM2 inflammasome is critical for innate<br />

immunity to Francisella tularensis. Nat Immunol 11:385-93.<br />

49. Fink, S. L., T. Bergsbaken, and B. T. Cookson. 2008. Anthrax lethal toxin<br />

and Salmonella elicit the common cell death pathway of caspase-1-dependent<br />

pyroptosis via distinct mechanisms. Proc Natl Acad Sci U S A 105:4312-7.<br />

50. Franchi, L., A. Amer, M. Body-Malapel, T. D. Kanneganti, N. Ozoren, R.<br />

Jagirdar, N. Inohara, P. Vandenabeele, J. Bertin, A. Coyle, E. P. Grant,<br />

and G. Nunez. 2006. Cytosolic flagellin requires Ipaf for activation of<br />

caspase-1 and interleukin 1beta in salmonella-infected macrophages. Nat<br />

Immunol 7:576-82.<br />

51. Ghayur, T., S. Banerjee, M. Hugunin, D. Butler, L. Herzog, A. Carter, L.<br />

Quintal, L. Sekut, R. Talanian, M. Paskind, W. Wong, R. Kamen, D.<br />

Tracey, and H. Allen. 1997. Caspase-1 processes IFN-gamma-inducing factor<br />

and regulates LPS-induced IFN-gamma production. Nature 386:619-23.<br />

52. Giacomini, E., E. Iona, L. Ferroni, M. Miettinen, L. Fattorini, G. Orefici,<br />

I. Julkunen, and E. M. Coccia. 2001. Infection of human macrophages and<br />

dendritic cells with Mycobacterium tuberculosis induces a differential cytokine<br />

gene expression that modulates T cell response. J Immunol 166:7033-41.<br />

53. Gill, V., and B. A. Cunha. 1997. Tularemia pneumonia. Semin Respir Infect<br />

12:61-7.<br />

54. Golovliov, I., V. Baranov, Z. Krocova, H. Kovarova, and A. Sjostedt. 2003.<br />

An attenuated strain of the facultative intracellular bacterium Francisella<br />

140


tularensis can escape the phagosome of monocytic cells. Infect Immun<br />

71:5940-50.<br />

55. Gray, C. G., S. C. Cowley, K. K. Cheung, and F. E. Nano. 2002. The<br />

identification of five genetic loci of Francisella novicida associated with<br />

intracellular growth. FEMS Microbiol Lett 215:53-6.<br />

56. Gurycova, D. 1998. First isolation of Francisella tularensis subsp. tularensis in<br />

Europe. Eur J Epidemiol 14:797-802.<br />

57. Hajjar, A. M., M. D. Harvey, S. A. Shaffer, D. R. Goodlett, A. Sjostedt, H.<br />

Edebro, M. Forsman, M. Bystrom, M. Pelletier, C. B. Wilson, S. I. Miller,<br />

S. J. Skerrett, and R. K. Ernst. 2006. Lack of in vitro and in vivo recognition<br />

of Francisella tularensis subspecies lipopolysaccharide by Toll-like receptors.<br />

Infect Immun 74:6730-8.<br />

58. Hall, J. D., R. R. Craven, J. R. Fuller, R. J. Pickles, and T. H. Kawula.<br />

2007. Francisella tularensis replicates within alveolar type II epithelial cells in<br />

vitro and in vivo following inhalation. Infect Immun 75:1034-9.<br />

59. Halle, A., V. Hornung, G. C. Petzold, C. R. Stewart, B. G. Monks, T.<br />

Reinheckel, K. A. Fitzgerald, E. Latz, K. J. Moore, and D. T. Golenbock.<br />

2008. The NALP3 inflammasome is involved in the innate immune response to<br />

amyloid-beta. Nat Immunol 9:857-65.<br />

60. Hayashi, F., K. D. Smith, A. Ozinsky, T. R. Hawn, E. C. Yi, D. R.<br />

Goodlett, J. K. Eng, S. Akira, D. M. Underhill, and A. Aderem. 2001. The<br />

innate immune response to bacterial flagellin is mediated by Toll-like receptor<br />

5. Nature 410:1099-103.<br />

61. Heil, F., H. Hemmi, H. Hochrein, F. Ampenberger, C. Kirschning, S.<br />

Akira, G. Lipford, H. Wagner, and S. Bauer. 2004. Species-specific<br />

recognition of single-stranded RNA via toll-like receptor 7 and 8. Science<br />

303:1526-9.<br />

62. Helvaci, S., S. Gedikoglu, H. Akalin, and H. B. Oral. 2000. Tularemia in<br />

Bursa, Turkey: 205 cases in ten years. Eur J Epidemiol 16:271-6.<br />

63. Hemmi, H., O. Takeuchi, T. Kawai, T. Kaisho, S. Sato, H. Sanjo, M.<br />

Matsumoto, K. Hoshino, H. Wagner, K. Takeda, and S. Akira. 2000. A<br />

Toll-like receptor recognizes bacterial DNA. Nature 408:740-5.<br />

64. Henry, T., A. Brotcke, D. S. Weiss, L. J. Thompson, and D. M. Monack.<br />

2007. Type I interferon signaling is required for activation of the<br />

inflammasome during Francisella infection. J Exp Med 204:987-94.<br />

65. Henry, T., G. S. Kirimanjeswara, T. Ruby, J. W. Jones, K. Peng, M.<br />

Perret, L. Ho, J. D. Sauer, Y. Iwakura, D. W. Metzger, and D. M.<br />

Monack. Type I IFN signaling constrains IL-17A/F secretion by gammadelta<br />

T cells during bacterial infections. J Immunol 184:3755-67.<br />

66. Hoel, T., O. Scheel, S. H. Nordahl, and T. Sandvik. 1991. Water- and<br />

airborne Francisella tularensis biovar palaearctica isolated from human blood.<br />

Infection 19:348-50.<br />

67. Hood, A. M. 1977. Virulence factors of Francisella tularensis. J Hyg (Lond)<br />

79:47-60.<br />

141


68. Hornung, V., A. Ablasser, M. Charrel-Dennis, F. Bauernfeind, G.<br />

Horvath, D. R. Caffrey, E. Latz, and K. A. Fitzgerald. 2009. AIM2<br />

recognizes cytosolic dsDNA and forms a caspase-1-activating inflammasome<br />

with ASC. Nature 458:514-8.<br />

69. Hsu, L. C., S. R. Ali, S. McGillivray, P. H. Tseng, S. Mariathasan, E. W.<br />

Humke, L. Eckmann, J. J. Powell, V. Nizet, V. M. Dixit, and M. Karin.<br />

2008. A NOD2-NALP1 complex mediates caspase-1-dependent IL-1beta<br />

secretion in response to Bacillus anthracis infection and muramyl dipeptide.<br />

Proc Natl Acad Sci U S A 105:7803-8.<br />

70. Inohara, Chamaillard, C. McDonald, and G. Nunez. 2005. NOD-LRR<br />

proteins: role in host-microbial interactions and inflammatory disease. Annu<br />

Rev Biochem 74:355-83.<br />

71. Inohara, N., and G. Nunez. 2003. NODs: intracellular proteins involved in<br />

inflammation and apoptosis. Nat Rev Immunol 3:371-82.<br />

72. Isaacs, A., and J. Lindenmann. 1957. Virus interference. I. The interferon.<br />

Proc R Soc Lond B Biol Sci 147:258-67.<br />

73. Isaacs, A., J. Lindenmann, and R. C. Valentine. 1957. Virus interference. II.<br />

Some properties of interferon. Proc R Soc Lond B Biol Sci 147:268-73.<br />

74. Ishii, K. J., C. Coban, H. Kato, K. Takahashi, Y. Torii, F. Takeshita, H.<br />

Ludwig, G. Sutter, K. Suzuki, H. Hemmi, S. Sato, M. Yamamoto, S.<br />

Uematsu, T. Kawai, O. Takeuchi, and S. Akira. 2006. A Toll-like receptorindependent<br />

antiviral response induced by double-stranded B-form DNA. Nat<br />

Immunol 7:40-8.<br />

75. Ishii, K. J., T. Kawagoe, S. Koyama, K. Matsui, H. Kumar, T. Kawai, S.<br />

Uematsu, O. Takeuchi, F. Takeshita, C. Coban, and S. Akira. 2008.<br />

TANK-binding kinase-1 delineates innate and adaptive immune responses to<br />

DNA vaccines. Nature 451:725-9.<br />

76. Ishikawa, H., and G. N. Barber. 2008. STING is an endoplasmic reticulum<br />

adaptor that facilitates innate immune signalling. Nature 455:674-8.<br />

77. Ishikawa, H., Z. Ma, and G. N. Barber. 2009. STING regulates intracellular<br />

DNA-mediated, type I interferon-dependent innate immunity. Nature.<br />

78. Iwasaki, A., and R. Medzhitov. 2004. Toll-like receptor control of the<br />

adaptive immune responses. Nat Immunol 5:987-95.<br />

79. Jiang, Z., P. Georgel, X. Du, L. Shamel, S. Sovath, S. Mudd, M. Huber, C.<br />

Kalis, S. Keck, C. Galanos, M. Freudenberg, and B. Beutler. 2005. CD14 is<br />

required for MyD88-independent LPS signaling. Nat Immunol 6:565-70.<br />

80. Kanistanon, D., A. M. Hajjar, M. R. Pelletier, L. A. Gallagher, T.<br />

Kalhorn, S. A. Shaffer, D. R. Goodlett, L. Rohmer, M. J. Brittnacher, S. J.<br />

Skerrett, and R. K. Ernst. 2008. A Francisella mutant in lipid A carbohydrate<br />

modification elicits protective immunity. PLoS Pathog 4:e24.<br />

81. Kanneganti, T. D., M. Body-Malapel, A. Amer, J. H. Park, J. Whitfield, L.<br />

Franchi, Z. F. Taraporewala, D. Miller, J. T. Patton, N. Inohara, and G.<br />

Nunez. 2006. Critical role for Cryopyrin/Nalp3 in activation of caspase-1 in<br />

response to viral infection and double-stranded RNA. J Biol Chem 281:36560-<br />

8.<br />

142


82. Kato, H., O. Takeuchi, E. Mikamo-Satoh, R. Hirai, T. Kawai, K.<br />

Matsushita, A. Hiiragi, T. S. Dermody, T. Fujita, and S. Akira. 2008.<br />

Length-dependent recognition of double-stranded ribonucleic acids by retinoic<br />

acid-inducible gene-I and melanoma differentiation-associated gene 5. J Exp<br />

Med 205:1601-10.<br />

83. Katz, J., P. Zhang, M. Martin, S. N. Vogel, and S. M. Michalek. 2006. Tolllike<br />

receptor 2 is required for inflammatory responses to Francisella tularensis<br />

LVS. Infect Immun 74:2809-16.<br />

84. Keim, P., A. Johansson, and D. M. Wagner. 2007. Molecular epidemiology,<br />

evolution, and ecology of Francisella. Ann N Y Acad Sci 1105:30-66.<br />

85. Kim, S., F. Bauernfeind, A. Ablasser, G. Hartmann, K. A. Fitzgerald, E.<br />

Latz, and V. Hornung. Listeria monocytogenes is sensed by the NLRP3 and<br />

AIM2 inflammasome. Eur J Immunol.<br />

86. Kono, H., and K. L. Rock. 2008. How dying cells alert the immune system to<br />

danger. Nat Rev Immunol 8:279-89.<br />

87. Kostura, M. J., M. J. Tocci, G. Limjuco, J. Chin, P. Cameron, A. G.<br />

Hillman, N. A. Chartrain, and J. A. Schmidt. 1989. Identification of a<br />

monocyte specific pre-interleukin 1 beta convertase activity. Proc Natl Acad<br />

Sci U S A 86:5227-31.<br />

88. Kuida, K., J. A. Lippke, G. Ku, M. W. Harding, D. J. Livingston, M. S. Su,<br />

and R. A. Flavell. 1995. Altered cytokine export and apoptosis in mice<br />

deficient in interleukin-1 beta converting enzyme. Science 267:2000-3.<br />

89. Labbe, K., and M. Saleh. 2008. Cell death in the host response to infection.<br />

Cell Death Differ 15:1339-49.<br />

90. Landolfo, S., M. Gariglio, G. Gribaudo, and D. Lembo. 1998. The Ifi 200<br />

genes: an emerging family of IFN-inducible genes. Biochimie 80:721-8.<br />

91. Lauriano, C. M., J. R. Barker, S. S. Yoon, F. E. Nano, B. P. Arulanandam,<br />

D. J. Hassett, and K. E. Klose. 2004. MglA regulates transcription of<br />

virulence factors necessary for Francisella tularensis intraamoebae and<br />

intramacrophage survival. Proc Natl Acad Sci U S A 101:4246-9.<br />

92. Lemaitre, B., E. Nicolas, L. Michaut, J. M. Reichhart, and J. A.<br />

Hoffmann. 1996. The dorsoventral regulatory gene cassette spatzle/Toll/cactus<br />

controls the potent antifungal response in Drosophila adults. Cell 86:973-83.<br />

93. Long, G. W., J. J. Oprandy, R. B. Narayanan, A. H. Fortier, K. R. Porter,<br />

and C. A. Nacy. 1993. Detection of Francisella tularensis in blood by<br />

polymerase chain reaction. J Clin Microbiol 31:152-4.<br />

94. Malik, M., C. S. Bakshi, B. Sahay, A. Shah, S. A. Lotz, and T. J. Sellati.<br />

2006. Toll-like receptor 2 is required for control of pulmonary infection with<br />

Francisella tularensis. Infect Immun 74:3657-62.<br />

95. Margolis, J. J., S. El-Etr, L. M. Joubert, E. Moore, R. Robison, A. Rasley,<br />

A. M. Spormann, and D. M. Monack. Contributions of Francisella tularensis<br />

subsp. novicida chitinases and Sec secretion system to biofilm formation on<br />

chitin. Appl Environ Microbiol 76:596-608.<br />

96. Mariathasan, S., K. Newton, D. M. Monack, D. Vucic, D. M. French, W. P.<br />

Lee, M. Roose-Girma, S. Erickson, and V. M. Dixit. 2004. Differential<br />

143


activation of the inflammasome by caspase-1 adaptors ASC and Ipaf. Nature<br />

430:213-8.<br />

97. Mariathasan, S., D. S. Weiss, V. M. Dixit, and D. M. Monack. 2005. Innate<br />

immunity against Francisella tularensis is dependent on the ASC/caspase-1<br />

axis. J Exp Med 202:1043-9.<br />

98. Mariathasan, S., D. S. Weiss, K. Newton, J. McBride, K. O'Rourke, M.<br />

Roose-Girma, W. P. Lee, Y. Weinrauch, D. M. Monack, and V. M. Dixit.<br />

2006. Cryopyrin activates the inflammasome in response to toxins and ATP.<br />

Nature 440:228-32.<br />

99. Marina-Garcia, N., L. Franchi, Y. G. Kim, D. Miller, C. McDonald, G. J.<br />

Boons, and G. Nunez. 2008. Pannexin-1-mediated intracellular delivery of<br />

muramyl dipeptide induces caspase-1 activation via cryopyrin/NLRP3<br />

independently of Nod2. J Immunol 180:4050-7.<br />

100. Martinon, F., L. Agostini, E. Meylan, and J. Tschopp. 2004. Identification<br />

of bacterial muramyl dipeptide as activator of the NALP3/cryopyrin<br />

inflammasome. Curr Biol 14:1929-34.<br />

101. Martinon, F., K. Burns, and J. Tschopp. 2002. The inflammasome: a<br />

molecular platform triggering activation of inflammatory caspases and<br />

processing of proIL-beta. Mol Cell 10:417-26.<br />

102. Martinon, F., K. Hofmann, and J. Tschopp. 2001. The pyrin domain: a<br />

possible member of the death domain-fold family implicated in apoptosis and<br />

inflammation. Curr Biol 11:R118-20.<br />

103. Martinon, F., A. Mayor, and J. Tschopp. 2009. The inflammasomes:<br />

guardians of the body. Annu Rev Immunol 27:229-65.<br />

104. Martinon, F., V. Petrilli, A. Mayor, A. Tardivel, and J. Tschopp. 2006.<br />

Gout-associated uric acid crystals activate the NALP3 inflammasome. Nature<br />

440:237-41.<br />

105. Matzinger, P. 2002. The danger model: a renewed sense of self. Science<br />

296:301-5.<br />

106. McCaffrey, R. L., and L. A. Allen. 2006. Francisella tularensis LVS evades<br />

killing by human neutrophils via inhibition of the respiratory burst and<br />

phagosome escape. J Leukoc Biol 80:1224-30.<br />

107. McCoy, G. W., and C. W. Chapin. 1911. Tuberculosis among Ground<br />

Squirrels (Citellus Beecheyi, Richardson). J Med Res 25:189-198.<br />

108. McIntire, C. R., G. Yeretssian, and M. Saleh. 2009. Inflammasomes in<br />

infection and inflammation. Apoptosis 14:522-35.<br />

109. Medzhitov, R., P. Preston-Hurlburt, and C. A. Janeway, Jr. 1997. A<br />

human homologue of the Drosophila Toll protein signals activation of adaptive<br />

immunity. Nature 388:394-7.<br />

110. Meibom, K. L., and A. Charbit. 2010. The unraveling panoply of Francisella<br />

tularensis virulence attributes. Curr Opin Microbiol 13:11-7.<br />

111. Meibom, K. L., I. Dubail, M. Dupuis, M. Barel, J. Lenco, J. Stulik, I.<br />

Golovliov, A. Sjostedt, and A. Charbit. 2008. The heat-shock protein ClpB<br />

of Francisella tularensis is involved in stress tolerance and is required for<br />

multiplication in target organs of infected mice. Mol Microbiol 67:1384-401.<br />

144


112. Metzger, D. W., C. S. Bakshi, and G. Kirimanjeswara. 2007. Mucosal<br />

immunopathogenesis of Francisella tularensis. Ann N Y Acad Sci 1105:266-<br />

83.<br />

113. Miao, E. A., C. M. Alpuche-Aranda, M. Dors, A. E. Clark, M. W. Bader,<br />

S. I. Miller, and A. Aderem. 2006. Cytoplasmic flagellin activates caspase-1<br />

and secretion of interleukin 1beta via Ipaf. Nat Immunol 7:569-75.<br />

114. Miao, E. A., R. K. Ernst, M. Dors, D. P. Mao, and A. Aderem. 2008.<br />

Pseudomonas aeruginosa activates caspase 1 through Ipaf. Proc Natl Acad Sci<br />

U S A 105:2562-7.<br />

115. Miao, E. A., D. P. Mao, N. Yudkovsky, R. Bonneau, C. G. Lorang, S. E.<br />

Warren, I. A. Leaf, and A. Aderem. 2010. Innate immune detection of the<br />

type III secretion apparatus through the NLRC4 inflammasome. Proc Natl<br />

Acad Sci U S A 107:3076-80.<br />

116. Molofsky, A. B., B. G. Byrne, N. N. Whitfield, C. A. Madigan, E. T. Fuse,<br />

K. Tateda, and M. S. Swanson. 2006. Cytosolic recognition of flagellin by<br />

mouse macrophages restricts Legionella pneumophila infection. J Exp Med<br />

203:1093-104.<br />

117. Morner, T., R. Mattsson, M. Forsman, K. E. Johansson, and G.<br />

Sandstrom. 1993. Identification and classification of different isolates of<br />

Francisella tularensis. Zentralbl Veterinarmed B 40:613-20.<br />

118. Muller, U., U. Steinhoff, L. F. Reis, S. Hemmi, J. Pavlovic, R. M.<br />

Zinkernagel, and M. Aguet. 1994. Functional role of type I and type II<br />

interferons in antiviral defense. Science 264:1918-21.<br />

119. Muruve, D. A., V. Petrilli, A. K. Zaiss, L. R. White, S. A. Clark, P. J. Ross,<br />

R. J. Parks, and J. Tschopp. 2008. The inflammasome recognizes cytosolic<br />

microbial and host DNA and triggers an innate immune response. Nature<br />

452:103-7.<br />

120. Nagano, Y., and Y. Kojima. 1958. [Inhibition of vaccinia infection by a<br />

liquid factor in tissues infected by homologous virus.]. C R Seances Soc Biol<br />

Fil 152:1627-9.<br />

121. Nakagawa, I., A. Amano, N. Mizushima, A. Yamamoto, H. Yamaguchi, T.<br />

Kamimoto, A. Nara, J. Funao, M. Nakata, K. Tsuda, S. Hamada, and T.<br />

Yoshimori. 2004. Autophagy defends cells against invading group A<br />

Streptococcus. Science 306:1037-40.<br />

122. Nano, F. E., N. Zhang, S. C. Cowley, K. E. Klose, K. K. Cheung, M. J.<br />

Roberts, J. S. Ludu, G. W. Letendre, A. I. Meierovics, G. Stephens, and K.<br />

L. Elkins. 2004. A Francisella tularensis pathogenicity island required for<br />

intramacrophage growth. J Bacteriol 186:6430-6.<br />

123. Nigrovic, L. E., and S. L. Wingerter. 2008. Tularemia. Infect Dis Clin North<br />

Am 22:489-504, ix.<br />

124. O'Riordan, M., C. H. Yi, R. Gonzales, K. D. Lee, and D. A. Portnoy. 2002.<br />

Innate recognition of bacteria by a macrophage cytosolic surveillance pathway.<br />

Proc Natl Acad Sci U S A 99:13861-6.<br />

145


125. Ogawa, M., T. Yoshimori, T. Suzuki, H. Sagara, N. Mizushima, and C.<br />

Sasakawa. 2005. Escape of intracellular Shigella from autophagy. Science<br />

307:727-31.<br />

126. Ohara, Y., T. Sato, and M. Homma. 1998. Arthropod-borne tularemia in<br />

Japan: clinical analysis of 1,374 cases observed between 1924 and 1996. J Med<br />

Entomol 35:471-3.<br />

127. Okabe, Y., K. Kawane, S. Akira, T. Taniguchi, and S. Nagata. 2005. Tolllike<br />

receptor-independent gene induction program activated by mammalian<br />

DNA escaped from apoptotic DNA degradation. J Exp Med 202:1333-9.<br />

128. Olsufjev, N. G., and I. S. Meshcheryakova. 1982. Infraspecific taxonomy of<br />

tularemia agent Francisella tularensis McCoy et Chapin. J Hyg Epidemiol<br />

Microbiol Immunol 26:291-9.<br />

129. Opitz, B., M. Vinzing, V. van Laak, B. Schmeck, G. Heine, S. Gunther, R.<br />

Preissner, H. Slevogt, P. D. N'Guessan, J. Eitel, T. Goldmann, A. Flieger,<br />

N. Suttorp, and S. Hippenstiel. 2006. Legionella pneumophila induces<br />

IFNbeta in lung epithelial cells via IPS-1 and IRF3, which also control<br />

bacterial replication. J Biol Chem 281:36173-9.<br />

130. Oyston, P. C., A. Sjostedt, and R. W. Titball. 2004. Tularaemia:<br />

bioterrorism defence renews interest in Francisella tularensis. Nat Rev<br />

Microbiol 2:967-78.<br />

131. Pan, Q., J. Mathison, C. Fearns, V. V. Kravchenko, J. Da Silva Correia,<br />

H. M. Hoffman, K. S. Kobayashi, J. Bertin, E. P. Grant, A. J. Coyle, F. S.<br />

Sutterwala, Y. Ogura, R. A. Flavell, and R. J. Ulevitch. 2007. MDPinduced<br />

interleukin-1beta processing requires Nod2 and CIAS1/NALP3. J<br />

Leukoc Biol 82:177-83.<br />

132. Parsa, K. V., J. P. Butchar, M. V. Rajaram, T. J. Cremer, and S.<br />

Tridandapani. 2008. The tyrosine kinase Syk promotes phagocytosis of<br />

Francisella through the activation of Erk. Mol Immunol 45:3012-21.<br />

133. Parsa, K. V., L. P. Ganesan, M. V. Rajaram, M. A. Gavrilin, A. Balagopal,<br />

N. P. Mohapatra, M. D. Wewers, L. S. Schlesinger, J. S. Gunn, and S.<br />

Tridandapani. 2006. Macrophage pro-inflammatory response to Francisella<br />

novicida infection is regulated by SHIP. PLoS Pathog 2:e71.<br />

134. Pechous, R., J. Celli, R. Penoske, S. F. Hayes, D. W. Frank, and T. C.<br />

Zahrt. 2006. Construction and characterization of an attenuated purine<br />

auxotroph in a Francisella tularensis live vaccine strain. Infect Immun<br />

74:4452-61.<br />

135. Pierini, L. M. 2006. Uptake of serum-opsonized Francisella tularensis by<br />

macrophages can be mediated by class A scavenger receptors. Cell Microbiol<br />

8:1361-70.<br />

136. Poltorak, A., X. He, I. Smirnova, M. Y. Liu, C. Van Huffel, X. Du, D.<br />

Birdwell, E. Alejos, M. Silva, C. Galanos, M. Freudenberg, P. Ricciardi-<br />

Castagnoli, B. Layton, and B. Beutler. 1998. Defective LPS signaling in<br />

C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282:2085-<br />

8.<br />

146


137. Prantner, D., T. Darville, and U. M. Nagarajan. 2010. Stimulator of IFN<br />

gene is critical for induction of IFN-beta during Chlamydia muridarum<br />

infection. J Immunol 184:2551-60.<br />

138. Py, B. F., M. M. Lipinski, and J. Yuan. 2007. Autophagy limits Listeria<br />

monocytogenes intracellular growth in the early phase of primary infection.<br />

Autophagy 3:117-25.<br />

139. Raetz, C. R., Z. Guan, B. O. Ingram, D. A. Six, F. Song, X. Wang, and J.<br />

Zhao. 2009. Discovery of new biosynthetic pathways: the lipid A story. J<br />

Lipid Res 50 Suppl:S103-8.<br />

140. Rathinam, V. A., Z. Jiang, S. N. Waggoner, S. Sharma, L. E. Cole, L.<br />

Waggoner, S. K. Vanaja, B. G. Monks, S. Ganesan, E. Latz, V. Hornung,<br />

S. N. Vogel, E. Szomolanyi-Tsuda, and K. A. Fitzgerald. The AIM2<br />

inflammasome is essential for host defense against cytosolic bacteria and DNA<br />

viruses. Nat Immunol 11:395-402.<br />

141. Ren, T., D. S. Zamboni, C. R. Roy, W. F. Dietrich, and R. E. Vance. 2006.<br />

Flagellin-deficient Legionella mutants evade caspase-1- and Naip5-mediated<br />

macrophage immunity. PLoS Pathog 2:e18.<br />

142. Sanders, C. V., and R. Hahn. 1968. Analysis of 106 cases of tularemia. J La<br />

State Med Soc 120:391-3.<br />

143. Santic, M., R. Asare, I. Skrobonja, S. Jones, and Y. Abu Kwaik. 2008.<br />

Acquisition of the vacuolar ATPase proton pump and phagosome acidification<br />

are essential for escape of Francisella tularensis into the macrophage cytosol.<br />

Infect Immun 76:2671-7.<br />

144. Santic, M., M. Molmeret, and Y. Abu Kwaik. 2005. Modulation of<br />

biogenesis of the Francisella tularensis subsp. novicida-containing phagosome<br />

in quiescent human macrophages and its maturation into a phagolysosome<br />

upon activation by IFN-gamma. Cell Microbiol 7:957-67.<br />

145. Santic, M., M. Molmeret, K. E. Klose, S. Jones, and Y. A. Kwaik. 2005.<br />

The Francisella tularensis pathogenicity island protein IglC and its regulator<br />

MglA are essential for modulating phagosome biogenesis and subsequent<br />

bacterial escape into the cytoplasm. Cell Microbiol 7:969-79.<br />

146. Sauer, J. D., C. E. Witte, J. Zemansky, B. Hanson, P. Lauer, and D. A.<br />

Portnoy. Listeria monocytogenes triggers AIM2-mediated pyroptosis upon<br />

infrequent bacteriolysis in the macrophage cytosol. Cell Host Microbe 7:412-9.<br />

147. Schmitz, J., A. Owyang, E. Oldham, Y. Song, E. Murphy, T. K.<br />

McClanahan, G. Zurawski, M. Moshrefi, J. Qin, X. Li, D. M. Gorman, J.<br />

F. Bazan, and R. A. Kastelein. 2005. IL-33, an interleukin-1-like cytokine<br />

that signals via the IL-1 receptor-related protein ST2 and induces T helper type<br />

2-associated cytokines. Immunity 23:479-90.<br />

148. Schulert, G. S., and L. A. Allen. 2006. Differential infection of mononuclear<br />

phagocytes by Francisella tularensis: role of the macrophage mannose<br />

receptor. J Leukoc Biol 80:563-71.<br />

149. Scott, A. M., and M. Saleh. 2007. The inflammatory caspases: guardians<br />

against infections and sepsis. Cell Death Differ 14:23-31.<br />

147


150. Sebastian, S., S. T. Dillon, J. G. Lynch, L. T. Blalock, E. Balon, K. T. Lee,<br />

L. E. Comstock, J. W. Conlan, E. J. Rubin, A. O. Tzianabos, and D. L.<br />

Kasper. 2007. A defined O-antigen polysaccharide mutant of Francisella<br />

tularensis live vaccine strain has attenuated virulence while retaining its<br />

protective capacity. Infect Immun 75:2591-602.<br />

151. Seet, B. T., J. B. Johnston, C. R. Brunetti, J. W. Barrett, H. Everett, C.<br />

Cameron, J. Sypula, S. H. Nazarian, A. Lucas, and G. McFadden. 2003.<br />

Poxviruses and immune evasion. Annu Rev Immunol 21:377-423.<br />

152. Stark, G. R., I. M. Kerr, B. R. Williams, R. H. Silverman, and R. D.<br />

Schreiber. 1998. How cells respond to interferons. Annu Rev Biochem<br />

67:227-64.<br />

153. Stetson, D. B., and R. Medzhitov. 2006. Recognition of cytosolic DNA<br />

activates an IRF3-dependent innate immune response. Immunity 24:93-103.<br />

154. Stetson, D. B., and R. Medzhitov. 2006. Type I interferons in host defense.<br />

Immunity 25:373-81.<br />

155. Stewart, S. J. 1996. Tularemia: association with hunting and farming. FEMS<br />

Immunol Med Microbiol 13:197-99.<br />

156. Stockinger, S., T. Materna, D. Stoiber, L. Bayr, R. Steinborn, T. Kolbe, H.<br />

Unger, T. Chakraborty, D. E. Levy, M. Muller, and T. Decker. 2002.<br />

Production of type I IFN sensitizes macrophages to cell death induced by<br />

Listeria monocytogenes. J Immunol 169:6522-9.<br />

157. Sutterwala, F. S., L. A. Mijares, L. Li, Y. Ogura, B. I. Kazmierczak, and<br />

R. A. Flavell. 2007. Immune recognition of Pseudomonas aeruginosa mediated<br />

by the IPAF/NLRC4 inflammasome. J Exp Med 204:3235-45.<br />

158. Sutterwala, F. S., Y. Ogura, and R. A. Flavell. 2007. The inflammasome in<br />

pathogen recognition and inflammation. J Leukoc Biol 82:259-64.<br />

159. Suzuki, T., L. Franchi, C. Toma, H. Ashida, M. Ogawa, Y. Yoshikawa, H.<br />

Mimuro, N. Inohara, C. Sasakawa, and G. Nunez. 2007. Differential<br />

regulation of caspase-1 activation, pyroptosis, and autophagy via Ipaf and ASC<br />

in Shigella-infected macrophages. PLoS Pathog 3:e111.<br />

160. Syrjala, H., P. Kujala, V. Myllyla, and A. Salminen. 1985. Airborne<br />

transmission of tularemia in farmers. Scand J Infect Dis 17:371-5.<br />

161. Takaoka, A., Z. Wang, M. K. Choi, H. Yanai, H. Negishi, T. Ban, Y. Lu,<br />

M. Miyagishi, T. Kodama, K. Honda, Y. Ohba, and T. Taniguchi. 2007.<br />

DAI (DLM-1/ZBP1) is a cytosolic DNA sensor and an activator of innate<br />

immune response. Nature 448:501-5.<br />

162. Takeda, K., and S. Akira. 2005. Toll-like receptors in innate immunity. Int<br />

Immunol 17:1-14.<br />

163. Takeuchi, O., T. Kawai, P. F. Muhlradt, M. Morr, J. D. Radolf, A.<br />

Zychlinsky, K. Takeda, and S. Akira. 2001. Discrimination of bacterial<br />

lipoproteins by Toll-like receptor 6. Int Immunol 13:933-40.<br />

164. Takeuchi, O., S. Sato, T. Horiuchi, K. Hoshino, K. Takeda, Z. Dong, R. L.<br />

Modlin, and S. Akira. 2002. Cutting edge: role of Toll-like receptor 1 in<br />

mediating immune response to microbial lipoproteins. J Immunol 169:10-4.<br />

148


165. Tamilselvam, B., and S. Daefler. 2008. Francisella targets cholesterol-rich<br />

host cell membrane domains for entry into macrophages. J Immunol 180:8262-<br />

71.<br />

166. Tarnvik, A., and L. Berglund. 2003. Tularaemia. Eur Respir J 21:361-73.<br />

167. Tarnvik, A., M. Ericsson, I. Golovliov, G. Sandstrom, and A. Sjostedt.<br />

1996. Orchestration of the protective immune response to intracellular<br />

bacteria: Francisella tularensis as a model organism. FEMS Immunol Med<br />

Microbiol 13:221-5.<br />

168. Tarnvik, A., G. Sandstrom, and A. Sjostedt. 1996. Epidemiological analysis<br />

of tularemia in Sweden 1931-1993. FEMS Immunol Med Microbiol 13:201-4.<br />

169. Thornberry, N. A., H. G. Bull, J. R. Calaycay, K. T. Chapman, A. D.<br />

Howard, M. J. Kostura, D. K. Miller, S. M. Molineaux, J. R. Weidner, J.<br />

Aunins, and et al. 1992. A novel heterodimeric cysteine protease is required<br />

for interleukin-1 beta processing in monocytes. Nature 356:768-74.<br />

170. Tominaga, A., M. A. Mahmoud, T. Mukaihara, and M. Enomoto. 1994.<br />

Molecular characterization of intact, but cryptic, flagellin genes in the genus<br />

Shigella. Mol Microbiol 12:277-85.<br />

171. Tschopp, J., F. Martinon, and K. Burns. 2003. NALPs: a novel protein<br />

family involved in inflammation. Nat Rev Mol Cell Biol 4:95-104.<br />

172. van Boxel-Dezaire, A. H., M. R. Rani, and G. R. Stark. 2006. Complex<br />

modulation of cell type-specific signaling in response to type I interferons.<br />

Immunity 25:361-72.<br />

173. Wang, X., A. A. Ribeiro, Z. Guan, S. N. Abraham, and C. R. Raetz. 2007.<br />

Attenuated virulence of a Francisella mutant lacking the lipid A 4'phosphatase.<br />

Proc Natl Acad Sci U S A 104:4136-41.<br />

174. Warren, S. E., D. P. Mao, A. E. Rodriguez, E. A. Miao, and A. Aderem.<br />

2008. Multiple Nod-like receptors activate caspase 1 during Listeria<br />

monocytogenes infection. J Immunol 180:7558-64.<br />

175. Wehrly, T. D., A. Chong, K. Virtaneva, D. E. Sturdevant, R. Child, J. A.<br />

Edwards, D. Brouwer, V. Nair, E. R. Fischer, L. Wicke, A. J. Curda, J. J.<br />

Kupko, 3rd, C. Martens, D. D. Crane, C. M. Bosio, S. F. Porcella, and J.<br />

Celli. 2009. Intracellular biology and virulence determinants of Francisella<br />

tularensis revealed by transcriptional profiling inside macrophages. Cell<br />

Microbiol 11:1128-50.<br />

176. Weiden, M., N. Tanaka, Y. Qiao, B. Y. Zhao, Y. Honda, K. Nakata, A.<br />

Canova, D. E. Levy, W. N. Rom, and R. Pine. 2000. Differentiation of<br />

monocytes to macrophages switches the Mycobacterium tuberculosis effect on<br />

HIV-1 replication from stimulation to inhibition: modulation of interferon<br />

response and CCAAT/enhancer binding protein beta expression. J Immunol<br />

165:2028-39.<br />

177. Weiss, D. S., A. Brotcke, T. Henry, J. J. Margolis, K. Chan, and D. M.<br />

Monack. 2007. In vivo negative selection screen identifies genes required for<br />

Francisella virulence. Proc Natl Acad Sci U S A 104:6037-42.<br />

178. Whipp, M. J., J. M. Davis, G. Lum, J. de Boer, Y. Zhou, S. W. Bearden, J.<br />

M. Petersen, M. C. Chu, and G. Hogg. 2003. Characterization of a novicida-<br />

149


like subspecies of Francisella tularensis isolated in Australia. J Med Microbiol<br />

52:839-42.<br />

179. White, J. D., J. R. Rooney, P. A. Prickett, E. B. Derrenbacher, C. W.<br />

Beard, and W. R. Griffith. 1964. Pathogenesis of Experimental Respiratory<br />

Tularemia in Monkeys. J Infect Dis 114:277-83.<br />

180. Wilson, K. P., J. A. Black, J. A. Thomson, E. E. Kim, J. P. Griffith, M. A.<br />

Navia, M. A. Murcko, S. P. Chambers, R. A. Aldape, S. A. Raybuck, and<br />

et al. 1994. Structure and mechanism of interleukin-1 beta converting enzyme.<br />

Nature 370:270-5.<br />

181. Wu, J., T. Fernandes-Alnemri, and E. S. Alnemri. Involvement of the<br />

AIM2, NLRC4, and NLRP3 Inflammasomes in Caspase-1 Activation by<br />

Listeria monocytogenes. J Clin Immunol.<br />

182. Yamamoto, M., S. Sato, H. Hemmi, K. Hoshino, T. Kaisho, H. Sanjo, O.<br />

Takeuchi, M. Sugiyama, M. Okabe, K. Takeda, and S. Akira. 2003. Role<br />

of adaptor TRIF in the MyD88-independent toll-like receptor signaling<br />

pathway. Science 301:640-3.<br />

183. Yang, P., H. An, X. Liu, M. Wen, Y. Zheng, Y. Rui, and X. Cao. The<br />

cytosolic nucleic acid sensor LRRFIP1 mediates the production of type I<br />

interferon via a beta-catenin-dependent pathway. Nat Immunol 11:487-94.<br />

184. Yoneyama, M., M. Kikuchi, K. Matsumoto, T. Imaizumi, M. Miyagishi, K.<br />

Taira, E. Foy, Y. M. Loo, M. Gale, Jr., S. Akira, S. Yonehara, A. Kato,<br />

and T. Fujita. 2005. Shared and unique functions of the DExD/H-box<br />

helicases RIG-I, MDA5, and LGP2 in antiviral innate immunity. J Immunol<br />

175:2851-8.<br />

185. Yoneyama, M., M. Kikuchi, T. Natsukawa, N. Shinobu, T. Imaizumi, M.<br />

Miyagishi, K. Taira, S. Akira, and T. Fujita. 2004. The RNA helicase RIG-I<br />

has an essential function in double-stranded RNA-induced innate antiviral<br />

responses. Nat Immunol 5:730-7.<br />

186. Zamboni, D. S., K. S. Kobayashi, T. Kohlsdorf, Y. Ogura, E. M. Long, R.<br />

E. Vance, K. Kuida, S. Mariathasan, V. M. Dixit, R. A. Flavell, W. F.<br />

Dietrich, and C. R. Roy. 2006. The Birc1e cytosolic pattern-recognition<br />

receptor contributes to the detection and control of Legionella pneumophila<br />

infection. Nat Immunol 7:318-25.<br />

187. Zhong, B., Y. Yang, S. Li, Y. Y. Wang, Y. Li, F. Diao, C. Lei, X. He, L.<br />

Zhang, P. Tien, and H. B. Shu. 2008. The adaptor protein MITA links virussensing<br />

receptors to IRF3 transcription factor activation. Immunity 29:538-50.<br />

188. Zwaferink, H., S. Stockinger, P. Hazemi, R. Lemmens-Gruber, and T.<br />

Decker. 2008. IFN-beta increases listeriolysin O-induced membrane<br />

permeabilization and death of macrophages. J Immunol 180:4116-23.<br />

150

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!