nucleotide-binding oligomerization domain containing 2

195
Nucleotide-binding Oligomerization Domain Containing 2 Characterization and Function during Mycobacterium tuberculosis Infection of Human Macrophages DISSERTATION Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University By Michelle Nichole Brooks Graduate Program in Microbiology The Ohio State University 2011 Dissertation Committee: Professor Larry Schlesinger, M.D., Advisor Professor John Gunn, Ph.D. Professor Chad Rappleye, Ph.D. Professor Mark Wewers, M.D.

Upload: khangminh22

Post on 22-Jan-2023

0 views

Category:

Documents


0 download

TRANSCRIPT

Nucleotide-binding Oligomerization Domain Containing 2 Characterization and Function during Mycobacterium tuberculosis Infection of Human Macrophages

DISSERTATION

Presented in Partial Fulfillment of the Requirements for the Degree Doctor of Philosophy in the Graduate School of The Ohio State University

By

Michelle Nichole Brooks

Graduate Program in Microbiology

The Ohio State University

2011

Dissertation Committee:

Professor Larry Schlesinger, M.D., Advisor

Professor John Gunn, Ph.D.

Professor Chad Rappleye, Ph.D.

Professor Mark Wewers, M.D.

                

Copyright by

Michelle Nichole Brooks

2011

  

ii

    

Abstract

Macrophages are important innate immune regulators in the recognition and clearance of

invading pathogens. Mycobacterium tuberculosis (M.tb), which causes tuberculosis, is a

host-adapted intracellular pathogen of macrophages. Intracellular pattern recognition

receptors in macrophages such as nucleotide-binding oligomerization domain (NOD)

proteins regulate pro-inflammatory cytokine production. NOD2-mediated signaling

pathways in response to M.tb have been studied primarily in mouse models and cell lines

but not in primary human macrophages. Work in this dissertation examined whether

NOD2 is expressed in human macrophages, controls the growth of M.tb, and functions in

inducible nitric oxide synthase (iNOS) and novel protein kinase c (PKC) signaling

pathways.

We examined NOD2 expression during monocyte differentiation and observed a marked

increase in NOD2 transcript and protein following 2-3 days in culture. Pre-treatment of

human monocyte-derived and alveolar macrophages with the NOD2 agonist muramyl

dipeptide enhanced production of TNF-α and IL-1β in response to M.tb and M. bovis

BCG (BCG) in a receptor interacting protein 2 (RIP2)-dependent fashion. The NOD2-

mediated cytokine response was significantly reduced following knockdown of NOD2

expression by using small interfering RNA (siRNA) in human macrophages. Finally,

NOD2 controlled the growth of both M.tb and BCG in human macrophages, whereas

iii

controlling only BCG growth in murine macrophages. Together, our results provide

evidence that NOD2 is an important intracellular receptor in regulating the host response

to M.tb and BCG infection in human macrophages.

Our observation of enhanced growth in NOD2 deficient cells at 24 hours (h) is of interest

because during this time M.tb is becoming acclimated to the macrophage. This led us to

examine the involvement of NOD2 in early innate immune responses. In order to study

new biological functions and molecular mechanisms of NOD2 during early M.tb

infection of human macrophages, we performed expression profiling of M.tb-infected

NOD2 deficient macrophages using a microarray. Notably, we identified that inducible

nitric oxide synthase (iNOS) expression is regulated by NOD2 during M.tb infection.

iNOS expression is enhanced at 24 and 48 h after M.tb and BCG incubation in human

macrophages. Our data show iNOS expression is decreased in NOD2 deficient cells

stimulated with M.tb and the M.tb-specific NOD2 agonist N-glycolylated muramyl

dipeptide (GMDP), indicating that NOD2 is involved in iNOS protein expression. These

data provide evidence for a novel pathway involving NOD2 that controls M.tb growth in

human macrophages through the regulation of iNOS expression.

We examined other potential signaling partners for NOD2 in human macrophages. Data

show that the NOD2 agonist, GMDP, stimulates phosphorylation and activation of PKCδ.

Corroborating these data, NOD2 deficient cells stimulated with PMA are significantly

reduced in PKCδ phosphorylation. PMA is a known stimulus for classical and novel

iv

PKC isoforms. Additionally, the kinase known to interact with NOD2, RIP2, is activated

by PMA. This activation is deficient upon knockdown of NOD2 in human macrophages.

Thus, our findings strongly support a role for NOD2 in human macrophages during M.tb

infection as well as in early innate immune responses that include iNOS and PKC

pathways.

v

Dedication

This document is dedicated to my family and friends who encouraged me throughout life

and molded me into the person I am today. I would like to specifically dedicate this to

my mother, Katherine L. Brooks, who struggled in life so that I may succeed. My father,

Michael S. Brooks, who taught me happiness and perseverance are the keys to life. My

fiancé, Samuel E. Landes, who I am eternally grateful to for providing a shoulder to lean

on during difficult times, listening to endless scientific presentations, and making my life

during graduate school filled with fun and laughter. Thank you!

vi

Acknowledgments

I am truly grateful to my advisor, Dr. Larry Schlesinger, M.D., for the

encouragement and expectation of scientific rigor, which led to much success during my

dissertation research. I am positive the skills obtained during my Ph.D. in his laboratory

will be invaluable to my career in science.

Foremost I would like to thank past and present laboratory members, who

everyday gave me scientific advice and encouragement. Specifically, I would like to give

thanks to Dr. Murugesan Rajaram, Ph.D., who undoubtedly guided the bulk of my work

in the laboratory and taught me how to execute experiments efficiently. Without him

instilling the excitement of obtaining the next result, none of my scientific success would

have been possible. I am also appreciative of Drs. Abul Azad, Ph.D. and Jordi Torrelles,

Ph.D., who helped me begin my training and taught me fundamental Schlesinger

laboratory techniques, while always taking time to answer my scientific questions.

I thank my graduate dissertation committee, John Gunn, Ph.D., Chad Rappleye,

Ph.D, Mark Wewers, M.D., for their contributions to my research. I am forever indebted

to them for shaping me into a true scientist who is confident and able to convey my

science. In addition, I would like to thank members of the CMIB for their support and

scientific advice for my dissertation research. Specifically, Drs. Amal Amer, M.D.,

vii

Ph.D, Joanne Turner, Ph.D., and Erin Rottinghaus, Ph.D., who devoted a great deal of

personal time to enhance my confidence and ability to communicate my science. I would

not be the scientist I am today without you.

Finally, I would like to acknowledge my funding sources including the

Microbiology graduate student fellowship, NSF Graduate Research Fellowship Program,

and the CMIB NIH/NIAID T32 training grant.

viii

Vita

August 14, 1984.............................................Born – Atlanta, Georgia

2006................................................................B.S.A. Biological Sciences,

University of Georgia

2006 to present ..............................................Graduate Research Fellow, Department of

Microbiology, The Ohio State University

2006 to 2007………………………………..Microbiology Graduate Student Fellowship

2007 to 2010………………………………..NSF Graduate Research Fellowship Program

2009………………………………………….1st place pre doctoral graduate poster

competition, DHLRI

2009………………………………………… Travel Award Winner 9th Annual OSUMC

Trainee Research Day poster competition

2010…………………………………………2nd place pre doctoral graduate poster

competition, DHLRI

2011…………………………………………Keystone Symposium Underrepresented

Minority Scholarship; TB Immunology and

Vaccine Design

2011 to present………………………………CMIB NIH/NIAID T32 training grant

ix

Publications

Brooks, M, M. Rajaram, A. Azad, A. Amer, M. Valdivia-Arenas, J. Park, G. Nuñez, and

L. Schlesinger. 2011. NOD2 controls the nature of the inflammatory response and

subsequent fate of Mycobacterium tuberculosis and M. bovis BCG in human

macrophages. Cellular Microbiology. 13(3):402-418

Carlson, T, M. Brooks, M. Rajaram, L. Henning, D. Myer, and L. Schlesinger. 2011.

Pulmonary innate immunity: soluble and cellular host defenses of the lung. In:

Regulation of Innate Immune Functions Book Chapter. Research Signpost, STM Books.

Rajaram, M, M. Brooks, J. Morris, J. Torrelles, A. Azad, L. Schlesinger. 2010.

Mycobacterium tuberculosis activates human macrophage peroxisome proliferator-

Activated receptor γ linking mannose receptor recognition to regulation of immune

responses. Journal of Immunology. 185:929-942

Fields of Study

Major Field: Microbiology

x

Table of Contents

Abstract ............................................................................................................................... ii

Dedication ........................................................................................................................... v

Acknowledgments.............................................................................................................. vi

Vita................................................................................................................................... viii

List of Figures ................................................................................................................. xvii

CHAPTER 1 : Introduction ............................................................................................... 1

1.1 The immune system .................................................................................................. 1

1.2 Innate immunity ........................................................................................................ 1

1.2.1 PKC isoforms ..................................................................................................... 3

1.2.2 iNOS and NO production ................................................................................... 5

1.3 Adaptive immunity.................................................................................................... 6

1.4 Pulmonary innate immunity ...................................................................................... 8

1.5 Alveolar macrophage ................................................................................................ 9

1.6 Macrophage activation .............................................................................................. 9

1.7 Macrophage receptors ............................................................................................. 11

xi

1.8 Pattern recognition receptors................................................................................... 12

1.9 NOD-like receptors ................................................................................................. 14

1.9.1 Discovery.......................................................................................................... 14

1.9.2 NLR structure ................................................................................................... 15

1.9.3 LRR domain ..................................................................................................... 15

1.9.4 NOD domain..................................................................................................... 16

1.9.5 CARD domain .................................................................................................. 16

1.9.6 NOD agonists ................................................................................................... 18

1.9.7 Generation of NOD agonists ............................................................................ 18

1.10 NLRs and innate immunity ................................................................................... 19

1.11 NOD2 signaling cascade ....................................................................................... 19

1.12 NOD2 and TLR crosstalk...................................................................................... 22

1.13 NOD2 and microbial immunity............................................................................. 23

1.14 NOD2 and inflammatory diseases......................................................................... 24

1.15 NOD2: human versus mouse................................................................................. 25

1.16 Tuberculosis .......................................................................................................... 26

1.16.1 Drug therapy and vaccines ............................................................................. 27

1.16.2 Infection with M.tb ......................................................................................... 28

1.16.3 M.tb and alveolus interaction.......................................................................... 29

xii

1.16.4 M.tb and AM receptor interaction .................................................................. 30

1.16.5 M.tb and phagosome environment.................................................................. 31

1.16.6 M.tb secretion systems.................................................................................... 32

1.16.7 M.tb and entry into the cytosol ....................................................................... 35

1.17 M.tb and NOD2..................................................................................................... 36

1.18 Mouse and human differences in the immune system .......................................... 38

1.19 Summary ............................................................................................................... 39

1.20 Specific aims ......................................................................................................... 40

CHAPTER 2 : NOD2 controls the nature of the inflammatory response and subsequent

fate of Mycobacterium tuberculosis and M. bovis BCG in human macrophages............. 42

2.1 Introduction ............................................................................................................. 42

2.2 Materials and Methods ............................................................................................ 45

2.2.1 Buffers and reagents ......................................................................................... 45

2.2.2 Antibodies......................................................................................................... 46

2.2.3 Bacterial strains and culture ............................................................................. 47

2.2.4 Isolation and culture of human monocytes and macrophages.......................... 47

2.2.5 Mice and mouse macrophages.......................................................................... 48

2.2.6 RNA isolation................................................................................................... 48

2.2.7 Gene expression studies by qRT-PCR ............................................................. 48

xiii

2.2.8 Confocal microscopy........................................................................................ 49

2.2.9 Transfection of MDMs ..................................................................................... 49

2.2.10 Macrophage stimulation / infection, cell lysis, immunoprecipitation, and

western blotting ......................................................................................................... 50

2.2.11 ELISAs ........................................................................................................... 52

2.2.12 M.tb and BCG intracellular growth assays in macrophages........................... 52

2.2.13 M.tb and BCG cell association assay within macrophages ............................ 53

2.2.14 Cell enumeration of NOD2 siRNA transfected MDMs in monolayer culture53

2.2.15 NOD2 antibody peptide competition assay .................................................... 54

2.2.16 Statistics.......................................................................................................... 54

2.3 Results ..................................................................................................................... 55

2.3.1 NOD2 mRNA and protein expression increase as monocytes differentiate into

macrophages .............................................................................................................. 55

2.3.2 Activation of NOD2 enhances the M.tb- and BCG-mediated pro-inflammatory

cytokine response through the NF-κB pathway ........................................................ 58

2.3.3 siRNA-mediated knockdown of NOD2 activity in human macrophages leads to

loss of its activity....................................................................................................... 61

2.3.4 NOD2 regulates the macrophage TNF- and IL-1β cytokine response to M.tb

and BCG .................................................................................................................... 63

xiv

2.3.5 Expression of NOD2 in human AMs and its role in regulating the production of

TNF- and IL-1β in response to M.tb ....................................................................... 66

2.3.6 NOD2 controls the intracellular growth of M.tb and BCG in human

macrophages .............................................................................................................. 68

2.3.7 Effects of NOD2 on M.tb and BCG growth in mouse macrophages and on the

cytokine response to infection ................................................................................... 70

2.4 Discussion ............................................................................................................... 72

CHAPTER 3 : Critical role for NOD2 in Mycobacterium tuberculosis induced iNOS

expression in human macrophages ................................................................................... 81

3.1 Introduction ............................................................................................................. 81

3.2 Materials and Methods ............................................................................................ 84

3.2.1 Buffers and reagents ......................................................................................... 84

3.2.2 Antibodies......................................................................................................... 85

3.2.3 Bacterial strains and culture ............................................................................. 85

3.2.4 Isolation and culture of human macrophages ................................................... 86

3.2.5 RNA isolation................................................................................................... 86

3.2.6 Gene expression studies by qRT-PCR ............................................................. 86

3.2.7 Transfection of MDMs ..................................................................................... 87

3.2.8 Microarray analysis .......................................................................................... 87

xv

3.2.9 Macrophage stimulation / infection, cell lysis, and Western blotting .............. 88

3.2.10 DAF-FM diacetate staining ............................................................................ 89

3.2.11 Statistics.......................................................................................................... 90

3.3 Results ..................................................................................................................... 90

3.3.1 NOD2 regulates many genes during M.tb infection of human macrophages... 90

3.3.2 NOD2 agonists regulate iNOS expression and NO production in human

macrophages .............................................................................................................. 93

3.3.3 M.tb-mediated iNOS expression is NOD2-dependent ..................................... 96

3.4 Discussion ............................................................................................................... 99

CHAPTER 4 : The role of PKCδ in the NOD2 signaling pathway in human macrophages

......................................................................................................................................... 106

4.1 Introduction ........................................................................................................... 106

4.2 Materials and Methods .......................................................................................... 109

4.2.1 Buffers and reagents ....................................................................................... 109

4.2.2 Antibodies....................................................................................................... 110

4.2.3 Isolation and culture of human macrophages ................................................. 110

4.2.4 Transfection of MDMs ................................................................................... 111

4.2.5 Macrophage stimulation and preparation for ELISAs and Western blotting . 111

4.2.6 Kinase assays.................................................................................................. 112

xvi

4.2.7 ELISA............................................................................................................. 113

4.2.8 Statistics.......................................................................................................... 113

4.3 Results ................................................................................................................... 114

4.3.1 Activation of NOD2 enhances the PMA-mediated pro-inflammatory cytokine

response through its signaling cascade intermediates ............................................. 114

4.3.2 Phosphorylated PKCδ expression and activity is NOD2-dependent.............. 117

4.3.3 RIP2 activation is PKCδ-dependent ............................................................... 121

4.4 Discussion ............................................................................................................. 123

CHAPTER 5 : Synthesis................................................................................................ 128

Bibliography ................................................................................................................... 140

xvii

List of Figures

Figure 1.1: NOD1 and NOD2 structure and domain functions ....................................... 17

Figure 1.2: NOD2 signaling cascades.............................................................................. 21

Figure 2.1: NOD2 expression increases during monocyte differentiation to macrophages.

........................................................................................................................................... 57

Figure 2.2: The NOD2 ligand MDP enhances M.tb- and BCG-mediated cytokine

production in human macrophages, a process involving activation of RIP2 and IkBα.... 60

Figure 2.3: Loss of NOD2 activity in human macrophages by siRNA knockdown........ 62

Figure 2.4: NOD2 controls M.tb- and BCG-mediated cytokine production and RIP2

activation in human macrophages..................................................................................... 65

Figure 2.5: M.tb-mediated cytokine production in human AMs treated with MDP........ 67

Figure 2.6: NOD2 controls the intracellular growth of M.tb and BCG in human

macrophages. .................................................................................................................... 69

Figure 2.7: Effect of NOD2 on M.tb and BCG growth and cytokine response in mouse

WT and NOD2-/- BMMs. .................................................................................................. 71

Figure 3.1: NOD2 regulates gene expression in human macrophages during M.tb

infection. ........................................................................................................................... 92

xviii

Figure 3.2: NOD2 agonists regulate the mRNA and protein expression of iNOS and NO

production in human macrophages. .................................................................................. 95

Figure 3.3: M.tb-mediated induction of iNOS expression and NO production is NOD2-

dependent in human macrophages. ................................................................................... 98

Figure 4.1: The NOD2 signaling pathway is involved in PMA-stimulated cytokine

production. ...................................................................................................................... 116

Figure 4.2: NOD2 and PKCδ agonists stimulate PKCδ phosphorylation and activity in a

NOD2-dependent manner. .............................................................................................. 120

Figure 4.3: RIP2 phosphorylation mediated by PMA and GMDP is PKCδ-dependent.122

Figure 4.4: PKCδ is necessary for optimal NOD2 activity............................................ 127

xix

Abbreviations

ABC = ATP-binding cassette

AM = alveolar macrophage

APC = antigen presenting cell

BAL = bronchoalveolar lavage

BCG = Mycobacterium bovis BCG

BMM = bone marrow derived macrophage

BSA = bovine serum albumin

CARD = caspase activation and recruitment domain

CFP10 = culture filtrate proteins of 10 kDa

CR = complement receptor

CRD = carbohydrate recognition domain

DAF-FM = 4-amino-5-methylamino-2’,7’-difluorofluorescein diacetate

DAG = diacylglycerol

DAN = diaminonapthalene

DAP = diaminopimelic acid

DC-SIGN = dendritic cell-specific intercellular adhesion molecule-2 grabbing non-integrin

DUOX2 = dual oxidase 2

xx

EBD = effector binding domain

ELISA = enzyme-linked immunosorbent assay

ESAT6 = early secreted antigen target of 6 kDa

ESX-1 = early secretory antigenic target 6 system 1

GlcNAc = N-acetylglucosamine

GMDP = N-glycolyl-MDP

HIV = human immunodeficiency virus

IFN = interferon

Ig = immunoglobulin

IL = interleukin

iNOS = inducible nitric oxide synthase

IRG = immunity-related guanosine triphosphatase

JAK = janus kinases

LM = lipomannan

LPS = lipopolysaccharide LRD = ligand recognition domain LRR = leucine rich repeats ManLAM = mannose capped lipoarabinomannan

MAP = mitogen-activated protein

MDP = muramyl dipeptide

MDM = monocyte derived macrophage

MDR-TB = multi-drug resistant tuberculosis

xxi

MHC = major histocompatibility complex

MOI = multiplicity of infection

MR = mannose receptor

M.tb = Mycobacterium tuberculosis

MurNAc = N-acetylmuramic acid

NADPH = nicotinamide adenine dinucleotide phosphate hydrogen

NamH = N-acetyl muramic acid hydroxylase

NF-κB = nuclear factor kappa-light-chain-enhancer of activated B cells

NK = natural killer

NLR = nod-like receptor

NO = nitric oxide

NOD = nucleotide-binding oligomerization domain

PAMPs = pathogen associated molecular patterns

PBMC = peripheral blood mononuclear cell

PKC = protein kinase c

PPARγ = peroxisome proliferator activated receptor

PRK = PKC related kinases

PRR = pattern recognition receptor

PMA = phorbol 12-myristate 13-acetate

qRT-PCR = quantitative real-time PCR

R = resistance

RCN = relative copy number

xxii

RD1 = region of difference 1

RIP2 = receptor interacting protein 2

RNI = reactive nitrogen intermediates

ROI = reactive oxygen intermediates

SDS-PAGE = sodium dodecyl sulfate polyacrylamide gel electrophoresis

siRNA = small interfering RNA

SP-A = surfactant protein A

SP-D = surfactant protein D

SR = scavenger receptor

STAT = signal transducers and activators of transcription

TAK1 = TGF activated kinase 1

Tat = twin-arginase transport

TB = tuberculosis Th1 = type 1 helper T cells Th2 = type 2 helper T cells

TIR = toll IL-1 receptor

TLR = toll like receptor TNF- = Tumor necrosis factor alpha XDR-TB = extensively drug resistant WHO = World Health Organization

1

CHAPTER 1 : Introduction

1.1 The immune system

The immune system is highly complex, involving soluble components, cells, tissues, and

organs that are responsible for keeping the host healthy from disease. There are many

different cell types that make up host immunity including T cells, neutrophils,

eosinophils, basophils, mast cells, and antigen presenting cells (1). These cell types are

each unique in their function and work together to keep balance and order in the host.

The immune system is made of two arms, adaptive and innate, that work together in order

to keep the host healthy (2).

1.2 Innate immunity

The innate immune response is the first line of defense against invading pathogens, and is

also known as native immunity. This response occurs rapidly upon encountering an

infectious agent. There are four main principle components of innate immunity: physical

and chemical barriers, phagocytic cells and natural killer (NK) cells, blood proteins, and

cytokines (1). The major components that work in concert in order to protect the host

include: soluble and membrane bound molecules, cells, and processes. The soluble

2

components include complement, collectins, and acute phase proteins that are known to

decorate the surface of infectious agents and elicit a response. Membrane bound

components are those found on innate immune cells such as pattern recognition receptors

(PRR) (3). The cells of the innate immune response play a critical role in defending

against pathogens. These cells include mast cells, eosinophils, basophils, macrophages,

neutrophils, dendritic cells, and NK cells. The processes that occur include phagocytosis,

reactive oxygen/nitrogen intermediate killing, phagosome acidification, and antimicrobial

protein and peptide production (4). Macrophages are an important cell type in innate

immunity. These cells use three mechanisms in order to uptake small particles or

infectious agents. Pinocytosis is the uptake of substances that are < 1 μm, and is also

known as fluid phase endocytosis. Receptor-mediated endocytosis is the uptake of

specific substances that bind to surface receptors and this is mediated by clathrin.

Phagocytosis is the uptake of microbes that are 1 μm or larger, and this mechanism is

mediated by actin. Endocytosis uses clathrin-coated pits while phagocytosis uses

cytoskeletal components such as paxillin, talin, and vinculin (5).

Once engulfed inside the macrophage, the pathogen is in a phagosome. This is known as

an early phagosome characterized by Rab5 and the presence of cathepsin H. The

phagosome matures and acquires Rab7 and cathepsin S. This is known as the late

phagosome. Finally, the phagosome fuses with a lysosome that contains antimicrobial

proteins and peptides accompanied by a lower pH (6).

3

Upon phagocytosis the macrophage uses a mechanism called the respiratory burst as its

first line of defense, which causes the production of superoxide and hydrogen peroxide.

Upon ligation of phagocytic receptor(s), phospholipase C releases diacylglycerol (DAG)

and IP3. This is followed by calcium release from the endoplasmic reticulum (ER), which

in conjunction with DAG activates protein kinase C (PKC) isoforms. PKC then

phosphorylates the cytosolic component p47phox. This then activates the formation of the

NADPH oxidase, allowing its components to translocate to the phagosome or plasma

membrane, where the complex binds flavocytochrome subunits gp91 phox and p22 phox

(7). One of the end products of the respiratory burst, superoxide anion, can dismutate to

hydrogen peroxide, and can generate hydroxyl radicals and singlet oxygen. In

neutrophils, hydrogen peroxide can also be converted by myeloperoxidase into

hypochlorous acid and chloramines (4).

1.2.1 PKC isoforms

PKCs are a family of enzymes and were one of the first kinases identified. The major

function of these kinases is to phosphorylate hydroxyl groups on the serine and threonine

residues of other proteins in order for them to function properly. PKCs are most notably

known for their function in the respiratory burst but are also important in cytokine

signaling (8,9). PKCs are activated by phosphotidylserine or DAG in a calcium-

dependent manner as well as by tumor-promoting phorbol esters such as phorbol 12-

myristate 13-acetate (PMA) (10). There are 12 identified PKC isotypes in the

mammalian PKC family, and this contributes to the complexity of understanding PKC

4

function. There are four subfamilies of isoforms that are classified based on the

secondary messenger that is required for activation. Classical (conventional) PKCs have

been the most thoroughly studied and require DAG, calcium, and phospholipid for

activation. They include the isoforms PKCα, PKCβI, PKCβII, and PKCγ (11). PKCs that

require DAG but not calcium for activation are known as novel PKCs and include the

isoforms PKCδ, PKCθ, PKCε, and PKCη (12). Atypical PKCs require neither calcium

nor DAG for activation and include PKCι, PKCζ, PK-N1, PK-N2 (13). Both classical

and novel PKC subfamilies can be activated by phorbol esters in the presence of

phosphatidylserine. PMA is able to activate the enzymes by abolishing the requirement

of DAG and decreasing the concentration of calcium needed for activation (14). A final

group, PKC-related kinases (PRK), are similar to atypical PKCs; however, they were

discovered to bind and activate RhoA GTPase (15-18).

The novel PKCs are of interest here, specifically PKCδ. PKCδ is involved in B-cell

signaling, and the regulation of growth, apoptosis, and differentiation of a variety of cell

types (19-21). Phosphorylation of various tyrosine residues is necessary for activation of

the enzyme; one specific residue is Tyrosine 311 (22). Interactions between this kinase

include, but are not limited to, phosphoinositide-dependent kinase-1, STAT1, and the

insulin receptor (23-26). PKCθ functions to activate T-cells, and is necessary in the

activation of NF-κB and AP-1 through the T-cell receptor signaling complex (27,28).

This kinase has been shown to interact with a few proteins including, RAC-alpha

serine/threonine-protein kinase (AKT1) and Proto-oncogene tyrosine-protein kinase

5

(FYN) (29,30). Interestingly, PKCθ has been shown to phosphorylate CARD11

(CARMA1) and activation of NF-κB is mediated by these two proteins and B-cell

lymphoma/leukemia 10 (BCL10) in T cells (31).

1.2.2 iNOS and NO production

Reactive nitrogen species (RNS) are also prominent in macrophages. Inducible nitric

oxide synthase (iNOS) is regulated at the transcriptional level and is necessary in order

for nitric oxide (NO) to be produced (4). NO is produced on the cytoplasmic side of the

phagosome and can diffuse across membranes to reach intraphagosomal targets (32).

Upon encountering ROS in the luminal environment NO can undergo spontaneous or

catalytic conversion to RNS products that are important mediators in bacterial killing

such as nitric dioxide, peryoxynitrite, dinitrogen trioxide, and dinitrosyl iron complexes.

The products can cause protein inactivation and lipid conversion by oxidative damage,

causing impaired bacterial metabolism and inhibition of replication (33). Signal

transduction is one fundamental role RNS plays in biology and has been shown

experimentally during infections (34,35). It is now widely accepted that NO signaling is

important in the cardiovascular system as well as in vascular homeostasis (36). Another

role for NO is in controlling inflammation. While the mechanism is still unclear as to

how exogenous NO puts forth an anti-inflammatory effect, it is thought that it inhibits

NF-κB (37).

6

iNOS and NO production are dependent on many factors. iNOS mRNA is regulated at

multiple steps, including transcription, stability, and translation (38). The regulation of

iNOS protein may occur by calmodulin binding, dimer formation, substrate depletion (L-

arginine), substrate recycling, tetrahydrobiopterin availability, end product inhibition,

phosphorylation, and subcellular localization. Tetrahydrobiopterin plays a major role in

iNOS enzymatic activity; however, production can be influenced by cytokines and LPS

(38,39). Upon NO binding to the iron in heme, it acts as a feedback inhibitor of iNOS

(40,41). iNOS mRNA can be induced by Interferon-gamma (IFNγ) through the binding

of Interferon receptor 1 (IFNR1) and IFNR2 complex, which activates janus kinases

(JAK) and signal transducers and activators of transcription (STAT) pathways. The

transcription factor interferon response factor-1 (IRF1) is activated and iNOS mRNA is

produced (42). Production of iNOS mRNA is also induced by toll-like receptor (TLR)

ligands and use NF-κB transcription factors (43). While initial studies showed lower

levels of iNOS expression and NO production in human cells, the availability of better

techniques to study NO production has led to more convincing data for its presence in

human cells (44).

1.3 Adaptive immunity

When the innate immune response cannot control an infection, it also serves to give

warning that an infection is present and in turn activates the adaptive immune response.

While the innate immune system is notorious for its immediate response to an infectious

agent, the adaptive immune response is valued for its specificity and ability to recognize

7

a re-occurring previous antigen. Activation of this system is dependent on co-stimulatory

molecules on antigen presenting cells (APCs) and T lymphocytes (T cells). There are

two major types of adaptive immune responses, humoral and cell-mediated immunity.

Humoral immunity is the production of antibodies by B lymphocytes (B cells) and cell-

mediated is the interaction of T cells with macrophages. Extracellular microbes are

neutralized and targeted for elimination by antibodies, using mechanisms such as

phagocytosis or releasing inflammatory mediators from immune cells to remove them.

T cells are responsible for infectious agents that are able to subvert the innate immune

response and reside inside the macrophage where antibodies cannot reach. A major

function of T cells is to mediate the killing of infected cells in order to control the

infection. Activation of the adaptive immune response can elicit the production of

several cytokines including, IL-2, IL-4, and IFNγ. The T cell response to an infection is

dependent on the cytokine produced, e.g. TH1 (IL-12) or TH2 (IL-4) (1).

Other characteristics of adaptive immunity include specificity and diversity, memory,

specialization, self-limitation, and non-reactivity to self. Specificity is the ability of

individual lymphocytes to recognize specific epitopes of antigens. The lymphocyte

repertoire of an individual is great and this is due to the diversity of the structure of the

antigen-binding site. Memory is another key characteristic of adaptive immunity. The

second time an individual is exposed to an antigen, the response is quicker and greater

compared to the primary interaction. Specialization refers to the variety of responses

elicited, either humoral or cell-mediated, which allows the system to be more effective at

8

clearing an infection. The ability of the adaptive immune system to return to basal

conditions and only functioning when an antigen is present is called self-limitation.

Finally, nonreactivity to self may be one of the most important features because it

controls the system from reacting to an individual’s own antigenic substances (45).

1.4 Pulmonary innate immunity

While the lung is known for its role in respiration, it is also a major player in innate

immunity. The lung is constantly bombarded with particulates and microbes and serves

as a major interface between the host and the external environment. An intricate

pulmonary immune system is in place in order to protect the host, and this system

encompasses innate and adaptive responses (46).

Upon inhalation, a particulate or microbe (< 5 μm) is able to avoid ciliary beat, the cough

reflex, and mucus clearance to travel down the trachea and through the bronchi where it

eventually settles in the alveolus. The alveolus is where the protection of the host occurs.

Pulmonary innate immunity is controlled by cellular and soluble components. Airway

and alveolar epithelial cells, as well as leukocytes are cellular components, while the

antimicrobial products (e.g. collectins, defensins, lactoferrin, and cathelicidins) secreted

into the epithelial lining fluid compose the soluble components of pulmonary innate

immunity (47).

9

1.5 Alveolar macrophage

An important mediator of pulmonary immunity and one of the first lines of cellular

defense in the alveolus is the alveolar macrophage (AM). There are only one to two AMs

per alveolus which range in size from 9-40 μm in diameter (48). There are two major

ways AMs are generated. The majority of AMs are thought to originate from monocytes

circulating in the blood that eventually find their place in the alveolus, where they

differentiate into macrophages. The second way, considered more controversial, is that

the mononuclear phagocytes present in the lung replicate in the alveolus (48). In the

alveolus, AMs are continuously bathed in surfactant, which is an important immune

modulator that is produced by type II epithelial cells (49).

Two important functions of the AM are phagocytosis and inhibiting excessive pro-

inflammatory signaling. It is necessary for these cells to have enhanced phagocytosis in

order to clear the particles and microbes that are encountered by the alveolus. It is

important to maintain homeostasis within the alveolus, and this is achieved by a balanced

production of pro- and anti-inflammatory cytokines. AMs are unique in this ability to

regulate cytokine production and are labeled alternatively activated for this purpose

(48,50,51).

1.6 Macrophage activation

There are currently three types of activation states for macrophages based on the stimulus

encountered: type I, type II, and alternative activation (52,53). Type I, also known as

10

classically activated macrophages or M1 macrophages, are those that encounter IFNγ and

a stimulus that induces TNF-α, such as a microbe. Upon stimulation these cells produce

pro-inflammatory cytokines such as TNF-α, IL-12, and IL-6, along with oxidants,

increased MHCII expression, and a decrease in mannose receptor (MR) expression.

Macrophages that encounter immune complexes are known as type II macrophages.

These are similar to type I macrophages in that they produce pro-inflammatory cytokines,

oxidants, and increased MHCII expression; however, they also produce the anti-

inflammatory cytokine IL-10, and have less IL-12 production when compared to type I

macrophages (53).

Alternatively activated, also known as M2 macrophages, are the third type of activation

state. In vitro TH2 cytokines, IL-4 and IL-13, induce this macrophage type, which is

consistent with the phenotype of AMs. These cells are characterized by having

enhanced PRR expression, such as the MR, and decreased CD14 expression. They also

produce fewer amounts of pro-inflammatory cytokines and oxidants, but instead secrete

anti-inflammatory cytokines IL-10 and TGFβ more readily. Other markers of alternative

activation found through mouse studies include receptors (CD23 and CD163), metabolic

factors (L-arginase and lipoxygenase), and soluble proteins (FIZZI and Ym1/2) (52). Our

laboratory has identified several features of alternatively activated human macrophages

such as increased MR and TLR2 expression (54), decreased NADPH oxidase (55) and

TLR activities (54), and increased activity of the nuclear receptor PPARγ which is linked

to the MR (56).

11

1.7 Macrophage receptors

Phagocytosis is mediated by several different receptors. Opsonic phagocytosis occurs

through the complement receptors (CR), CR1 and CR3, and Fcγ receptors. CR1 can bind

mannose binding lectin (MBL), C1q, C3b, C4b, and requires additional signals to

mediate phagocytosis. CR3 (Mac 1 or CD11b/CD18) binds C3bi-opspnized bacteria

(57). The Fcγ receptors recognize particles opsonized with IgG antibodies (58). Other

receptors are involved in non-opsonic phagocytosis. The MR is a C-type lectin,

membrane bound Pattern Recognition Receptor (PRRs, see below) found on the cell

membrane as well as on endosomes within the cell. Its function is normally to bind

terminal mannose and fucose residues of host glycoproteins and glycolipids, which are

also common components of bacterial cell walls. The tail of the receptor is necessary in

order to induce phagocytosis (59). The scavenger receptor (SR) is known to phagocytose

low-density lipoproteins (LDL) circulating in the body that can no longer interact with

the LDL receptor, but has been shown to bind microbes as well (60).

There are several other membrane receptors that are not directly involved in

phagocytosis, but instead signal in order to activate the phagocyte and in some cases

regulate phagocytic receptors. These include dectin-1, TLRs, and G protein-coupled

receptors. Dectin-1 is involved in signaling after recognition of non-opsonic beta glucan

particles found in fungi (61). TLRs are trans-membrane PRRs that recognize various

pathogen-associated molecular patterns (PAMPS) and induce a signaling cascade that

12

typically results in the production of inflammatory cytokines (62). G protein-coupled

receptors recognize short peptides containing N-formylmethionyl residues and function

to stimulate the recruitment of lymphocytes to the site of infection. The macrophage also

has a secondary line of foreign molecule recognition with receptors located in the cytosol.

These PRRs are known as the Nucleotide-binding Oligomerization Domain (NOD)-Like

Receptors (NLRs) and function to induce an inflammatory response upon recognition of

their PAMP intracellularly (63).

1.8 Pattern recognition receptors

As noted above, PRRs are unique in that they recognize specific patterns or PAMPS.

These can range from bacterial peptidoglycan fragments, carbohydrates, lipids or flagella

to intracellular DNA or single stranded RNA. They play a major role in innate immunity

through functions such as phagocytosis and signaling an inflammatory response. PRRs

can have a transmembrane domain and be located on the cell surface or endosomes, the

cytosol, or secreted.

Two major membrane-bound PRRs are the MR and TLRs. The MR is 180 kDa protein

that, along with its transmembrane domain, has four other domains: an extracellular

amino terminal cysteine rich domain, a fibronectin type II repeat domain, eight tandem

cytsteine rich lectin-like carbohydrate recognition domains (CRD), and a cytoplasmic

carboxy-terminal domain (64). TLRs are an important class of PRRs. There are 10

identified mammalian TLRs that are distinctly membrane-associated type I receptors (65-

13

67). The exterior amino terminus contains variable arrangements of leucine-rich repeats

(LRRs) which serve to recognize PAMPs. The cytosolic carboxy terminus of TLRs is

highly homologous to the IL-1 receptor and contains a Toll-IL-1R (TIR) domain that

forms the scaffold for the assembly of signaling intermediates such as Myd88, IRAK1,

IRAK4, IRAKM and Mal/Tiram. TLR4 and TLR2 are the most thoroughly studied

TLRs. TLR4 has been found to recognize the endotoxin (lipopolysaccharide, LPS) of

Gram-negative organisms, while TLR2 recognizes Gram-positive PAMPS, such as

lipoteichoic acid. In general, TLR signaling drives NF-κB activation via phosphorylation

of IκBα (67-69), although several negative regulators such as IRAKM have been

identified (70).

Cytoplasmic PRRs do not contain a transmembrane domain and are associated with

cytoplasmic networks. NLRs are involved in processes such as apoptosis and pro-

inflammatory signaling. RNA helicases recognize single stranded or double stranded

RNA. The two recognized in mammalian cells are RIG-1 and MDA5, which activate

antiviral signaling (71). Secreted PRRs include MBL, serum amyloid protein, and C -

reactive protein (1,2).

14

1.9 NOD-like receptors

1.9.1 Discovery

NLRs were initially discovered in plants. Plants are able to recognize pathogens by

disease resistance (R) genes, and in turn induce a response that allows the infected area to

be isolated through cell wall remodeling, production of reactive oxygen species, cell

death, and production of anti-pathogenic molecules at the site of infection (72). There are

two classes of R genes, one that produces transmembrane proteins similar to TLRs and

another that produces cytosolic NOD proteins. These cytosolic NOD proteins generally

contain amino-terminal α-helix-rich or TIR domains, a central NOD, and carboxyl-

terminal LRRs. Since there are known similarities between plants and animals, genome

wide searches for homology between these plant R proteins and the mammalian system

were embarked upon by investigators. The first searches yielded apoptosis regulator,

Apaf-1 and its nematode homolog, CED-4, which revealed two proteins NOD1 (CARD4)

and NOD2 (CARD15) (73). Further studies have described 23 mammalian NLRs in the

innate immune recognition of intracellular pathogens (74). The current literature is

heavily focused on NOD1 and NOD2. They are mainly expressed in APCs and epithelial

cells. NOD1 protein is abundantly found in intestinal epithelial cells, while NOD2 protein

is low or undetectable in these cells. NOD2 protein expression has been found in Paneth

cells found at the base of intestinal crypts (75). NOD2 mRNA is present in monocytes

but protein expression is low; however, NOD2 protein expression is abundant in human

macrophages (76,77).

15

1.9.2 NLR structure

The structure of NOD1 and NOD2 are very similar to Apaf-1 and CED-4. They both

contain amino terminal caspase activating and recruitment domains (CARDs) linked to a

NOD domain; however, the mammalian NOD proteins contain LRRs in their carboxy

termini (73). Most mammalian NOD family members contain 3 distinct functional

domains, an amino terminal effector binding domain (EBD), a centrally located NOD,

and a carboxy-terminal ligand recognition domain (LRD). Depending on the downstream

signaling partner, the EBD is involved in homophilic and heterophilic interactions. The

NLR family can be divided into four subfamilies depending on the composition of their

N-terminal EBD. The differing N-terminal domains are as follows with subfamily name:

acidic transactivation domains (NLRAs), CARD (NLRCs), pyrin domains (NLRPs), and

baculovirus IPA repeat domains (NLRBs). NOD1 and NOD2 are part of the NLRC

subfamily and contain a LRR domain, NOD domain, and CARD domains (78).

1.9.3 LRR domain

The LRR domain is made up of leucine rich repeats that are needed for the sensing of

agonists. There are approximately 11 LRRs in a horseshoe shape made up of beta strands

that constitute the concave face of the LRRs. Functional analysis of NOD2 revealed

that the most C-terminal LRRs, 6-11, are critical for bacterial recognition (79). These

residues are confined to the β-strand/β-turn (xxLxLxx) motif (L = leucine, x = amino

16

acid) (80). There are six non-conserved amino acids in the corresponding regions of the

LRRs for each protein. These include G879, W907, V935, E959, K989, and S991. These

amino acids face outward and are thought to be important in recognition of the specific

agonist for each NOD protein (79). Currently, no one has been able to show direct

binding of NOD proteins with their specific agonist. Some speculate that there may be an

intermediate that has not been discovered that causes activation.

1.9.4 NOD domain

This region contains an ATP-binding cassette (ABC) that includes catalytic residues with

canonical binding site motifs for phosphate residues (Walker’s A box) and magnesium

ions (Walker’s B box). Similarities have been found between the AAA+ family of

ATPases oligomerization module and ABC region of NOD-LRR proteins (81). In this

region, there are also motifs involved in hydrolysis of nucleosides, ATP, deoxyATP

and/or GTP. Nucleotide hydrolysis is necessary for protein activation of NOD1 and

NOD2. Mutational analysis identified 21 critical residues in the NOD domain important

for signaling (79).

1.9.5 CARD domain

The CARD domain is necessary for interactions with the CARD domain of the

downstream kinase receptor interacting protein 2 (RIP2), which is necessary for the

signaling cascade to occur and NF-κB activation. The CARD effector domain is

involved in homophilic interactions. There are 6 helices and a hydrophobic core. On

helices 2 and 3 there is an acidic patch formed by residues E53, E54, and E56. This

acidic patch has been found to be involved in interactions between NOD1 and RIP2 (82).

The crystal structure of NOD1 revealed a dimer of two monomers where the sixth helix is

swapped between two monomers (83). NOD1 contains one CARD domain, while

NOD2 contains two CARD domains (Figure 1.1) (84).

Figure 1.1: NOD1 and NOD2 structure and domain functions

(A) NOD1 contains one CARD domain that is necessary for signaling, a NOD domain involved in activation of the protein, and a LRR domain essential for agonist recognition. (B) NOD2 contains all of the same domains as NOD1, but contains an additional CARD domain.

17

18

1.9.6 NOD agonists

The NLRC proteins, NOD1 and NOD2, recognize specific muropeptides found in the

peptidoglycan layer of Gram-positive and Gram-negative bacteria. The repeating units of

N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc) are arranged in

parallel by stem peptides made up of amino acids. Cleavage of these chains leads to the

agonists for NOD1 and NOD2. NOD2 recognizes muramyl dipeptide (MDP), which is

found in Gram-positive and Gram-negative bacteria, while NOD1 recognizes meso-

diaminopimelic acid (meso-DAP) only found in Gram-negative bacteria (85). Recently,

it was found that NOD2 recognizes an N-glycolylated form of MDP (GMDP) that can be

found in Mycobacterium tuberculosis (M.tb). The chemically synthesized GMDP is a

more potent stimulator of the immune response (86). Likewise the synthetic truncated

form of DAP, γ-D-glutamyl-meso-DAP, is sufficient to activate NOD1 (87,88).

1.9.7 Generation of NOD agonists

Even though the agonists are positioned in the peptidoglycan layer there are several ways

these muropeptides can be generated by the bacteria and host. As bacteria replicate, cell

division occurs, peptidoglycan remodeling occurs, and small fragments are released (89).

During peptidoglycan biosynthesis bacteria are also able to produce small molecules such

as MDP and iE-DAP (89-91). Lysozyme is secreted from the host in tears, saliva,

mucous, and is also found inside phagocytic lysosomes in order to breakdown the cell

wall of bacteria. This glycoside hydrolase catalyzes the hydrolysis of MurNAc and

GlcNAc thereby generates muropeptides, which include MDP and iE-DAP (92). N-

19

acetylmuramoyl-L-alanine amidases are found in bacteria and host cells. This amidase

cleaves the bond between MurNAc and the stem peptide in order to generate MDP.

There are proposed models for the agonists to get into the cytosol from inside the

phagosome in order to be recognized by NLRs. One mechanism is bacteria can divide

inside the phagosome and cause it to burst, in turn releasing the bacteria as well as any

peptidoglycan fragments that may have been generated by bacterial replication or host

response factors. The other method is through bacterial secretion systems (93).

1.10 NLRs and innate immunity

NLRs play a major role in innate immunity. They are a second line of recognition and

are responsible for inducing a pro-inflammatory response to bacteria that get inside the

macrophage. While there are 23 NLRs with a wide range of known functions including

cell signaling and cell death/survival pathways, the NLRC family’s established function

is in inducing an inflammatory response through NF-κB upon recognition of specific

muropeptides from bacteria.

1.11 NOD2 signaling cascade

The determining region of whether a NOD protein is active or inactive has been revealed

by mutational analysis to be the most carboxy-terminal region of the NOD domain also

known as the “switching region” (79). Upon recognition of MDP, the NOD domain of

NOD2 undergoes oligomerization, which allows the CARD domain to be exposed

20

leading to an electrostatic interaction with the CARD domain of the adapter protein RIP2

(85,88). This in turn causes autophosphorylation of RIP2 at serine 176 in cell lines and

murine macrophages (94). The NOD2-RIP2 complex then binds TGF activated kinase 1

(TAK1), which can then either directly activate by phosphorylation of the inhibitor of

NF-κB, IκBα, and or activate the MAPKs and thus indirectly allow for NF-κB and AP-1

activation leading to the production of pro-inflammatory cytokines (95,96). More

recently researchers discovered that NOD2 can directly bind and activate caspase-1 and

interact with the NALP1/NALP3 inflammasome, which causes the activation of caspase-

1, a necessary enzyme needed to cleave pro-IL-1β into its active secreted form (97)

(Figure 1.2). Also, a truncated form of NOD2 exists that is able to negatively regulate

NOD2 activity (98).

Figure 1.2: NOD2 signaling cascades

Upon bacterial phagocytosis, MDP is released into the cytosol by various mechanisms and is sensed by the NOD2 LRR domain. (A) The NOD2 protein oligomerizes and recruits RIP2 and this NOD2/RIP2 complex interacts with TAK1. After TAK1 activation, either MAPKs are activated to induce transcription of pro-inflammatory cytokines or the IKK complex is activated, which in turn removes the inhibitor or NF-κB, IκBα, and allows for transcription of pro-inflammatory cytokines to occur. (B) The NOD2 protein interacts with the NALP1 or NALP3 inflammasome or directly binds pro-caspase1. Pro-caspase-1 is then cleaved into its active form which enables cleavage of pro-IL-1β into the released form, IL-1β.

21

22

1.12 NOD2 and TLR crosstalk

There are many reports on the synergy between TLR agonists and NOD2 agonists in pro-

and anti-inflammatory cytokine release (99-106). Proper signaling through NF-κB by

NOD2 and TLRs involves the same mechanism of K63-linked ubiquitination of NEMO

(107). A biochemical mechanism for the synergy between TLRs and NOD2 has been

reported. Both signaling cascades utilize TAK1/TAB/Ubc13 to activate NF-κB, which

allows TLR and NOD2 signaling to synergistically induce cytokine release. However the

primary source of this activation is different, NOD2 uses RIP2 while TLRs utilize

TRAF6 (108).

On the level of transcription it has been shown that TLR activation and cytokines induce

NOD2 mRNA expression (109,110). MDP enhances MyD88 expression, which is a

necessary adapter protein in TLR signaling and may explain the high sensitivity of TLR

responses after peptidoglycan priming (111). The location of the NOD proteins may also

be a factor in crosstalk between TLRs and NOD2. While NOD2 is found in the cytosol,

studies have shown that it can be found localized at the cell-membrane and this

localization may be important in signaling (112-114). The proximity of NOD2 at the

membrane where TLRs can be found also suggests a mechanism for crosstalk.

23

1.13 NOD2 and microbial immunity

The biological function of NOD2 in innate and adaptive immune responses is poorly

understood; however, studies over the past decades on purified or synthetic MDP have

given some insight. In animal models the administration of MDP induces activity against

bacteria, fungi, viruses, protozoa, and tumor growth (115). The microbicidal activity of

MDP can be explained by the induction of TNF-α and IL-1β. This was thoroughly tested

by antibody neutralization studies (116-121). Further studies proved that MDP activates

NF-κB and that all MDP-inducible genes require NOD2 activity (122). MDP has been

described to be involved in the secretion of multiple immune mediators that are important

in pathogen clearance: pro-inflammatory cytokines (TNF-α, IL-1α, IL-1β), chemokines

(IL-8, CCL3, CCL4), and hematopoietic growth/survival factors (IL-6, G-CSF, GM-CSF)

(116-119,121-123). Adhesion molecules (ICAM1, CD44, TNFAIP6) known to be

important in the recruitment of inflammatory cells are also induced upon MDP

stimulation. NOD2 has also been shown to play a role in linking the innate and adaptive

immune response due to its adjuvant activity. MDP generates antigen specific T- and B-

cell responses, delayed-type hypersensitivity, and antibody production (115). This

activity is due to the ability of MDP to induce the expression of costimulatory molecules

(CD40, CD80, CD86) in monocytes and dendritic cells therefore differentiating naive T

cells into effector T cells (123-125).

24

Several microbial models have substantiated the importance of NLRs in the recognition

of pathogens including Shigella flexneri, Chlamydophila pneumonia, Pseudomonas

aeruginosa, Helicobacter pylori, Streptocooccus pneumoniae, Mycobacterium

tuberculosis, Bacillus antracis, Listeria monocytogenes, Legionella pneumophila, and

Staphylococcus aureus (77,97,126-137).

1.14 NOD2 and inflammatory diseases

Mutations in the NOD2 gene, CARD15, have been linked to many inflammatory diseases

including: Crohn’s disease, Blau syndrome, and early-onset sarcoidosis (138-141).

Mutations are found in the NOD domain for Blau syndrome and early-onset sarcoidosis,

while mutations in the LRR domain are linked to Crohn’s disease. Crohn’s disease

presents as inflammation of the gut along with granulomas, and is mediated by over

production of IL-1β, IL-6, IL-12, and TNF-α. There are two effective therapies; one is

neutralizing antibodies against TNF-α, IL-6, and IL-12 or NF-κB inhibitors and the

second is antibiotics (142). While debated, a correlative study reported that patients with

Crohn’s disease are also likely to be infected with Mycobacterium avium subspecies

paratuberculosis (143,144).

Studies have described mutations in the LRR domain as resulting in a ‘loss-of-function’

and ‘gain-of-function’. A ‘loss-of-function’ mutation results in a NOD2 protein that

cannot activate the immune response. One hypothesis for the ‘loss of function’

phenotype is that this leads to excessive inflammation to the bacterial burden due to the

25

lack of NOD2-induced defense mechanisms. Another hypothesis is that NOD2 is

involved in inhibiting the TLR2-mediated macrophage response and the lack of inhibition

leads to excessive activation of NF-κB and IL-12 expression (104). A ‘gain-of-function’

mutation results in a new or abnormal function. The ‘gain-of-function’ mutation

hypothesis is derived from ‘knockout-knockin’ CARD15 mouse studies which have a

homologous mutation found in human Crohn’s disease. Stimulation of these mouse

macrophages with MDP resulted in increased activation of NF-κB and caspase-1

mediated IL-1β processing and release (145). The excessive IL-1β production could lead

to enhanced inflammation in the intestines. This study is debated as recent studies using

human peripheral blood mononuclear cells (PBMCs) from patients with the homozygous

NOD2 mutation displayed decreased IL-1β production in response to MDP (146). These

differences reported are likely due to the difference in cell type, i.e. from mouse and

human.

1.15 NOD2: human versus mouse

Researchers discovered that the promoter regions of NOD2 in humans and mice are

different. Sequence alignment analysis revealed the canonical NF-κB binding site (-16 to

-25 bp) is not found in the mouse or rat NOD2 core promoter (147). Furthermore, the

NOD2 promoter (-500 to +300 bp) sequence alignment showed that there is only 49.7%

similarity to mouse. Hu et al found that the NOD2 proximal promoter between human

and chimpanzees is 98% identical (147). This indicates that human and chimpanzee

NOD2 genes are regulated by a highly similar transcriptional mechanism, while the lack

26

of similarity between human and mouse is due to divergent mechanisms in NOD2

regulation. These differences are important to remember when studying the function of

NOD2 in the mouse model.

1.16 Tuberculosis

Mycobacterium tuberculosis (M.tb) is the infectious agent that causes tuberculosis (TB).

One third of the world’s population is currently infected with this bacillus, however only

5-10% of the people will develop clinical disease during their lifetime. In 2008, the

World Health Organization (WHO) revealed that sub-Saharan Africa and the South-East

Asia region are where TB cases are most heavily reported (148). In 2009, 1.7 million

people died from TB with most deaths in the African region. This high incidence of TB

deaths is due to the enhanced number of human immunodeficiency virus (HIV) positive

people in Africa. HIV causes patients to be severely immuno-compromised (particularly

in CD4 T cell number and function), and thus the probability of developing TB is greater.

The WHO reported that TB is the leading cause of death among people who are HIV

positive in Africa, and since 1990 HIV has been the single largest contributor to the

increase in TB cases (148). The prevalence of TB combined with the increasing lack of

effective drug therapy give reason to study the interaction of M.tb with the host in more

depth. Understanding the molecular interactions of M.tb with host immune cells will

potentially yield greater insight into the development of new targets for drug therapy as

well as vaccine design.

27

1.16.1 Drug therapy and vaccines

The first TB vaccine was attempted by Robert Koch, who identified M.tb as the causative

agent of TB. His vaccine used sterile filtrates from M.tb culture but it was not protective

against TB; however it did open the door to further studies. Two French scientists,

Albert Calmette and Camille Guerin, used a strain of mycobacteria that infects cattle, M.

bovis made attenuated, by making serial passages for 13 years. This strain, M. bovis

BCG (BCG), became avirulent in mice and vaccination of infants resulted in 90%

reduction of mortality. Vaccination by BCG is effective against M.tb in children;

however, its efficacy in adults varies from 80% to 0% (149,150). While the vaccine is

not as effective in adults, it protects children, who are a highly susceptible group to

disseminated and meningeal TB (151). While the exact reason is unknown, host and

environmental factors are attributed to the lack of efficacy (152). The search continues

for a cost effective vaccine that is protective for adults in order to one day eradicate TB.

Researchers are attempting to synthesize different types of vaccines including DNA, sub-

unit, and living-attenuated vaccines. While there is currently no new TB vaccine in phase

III clinical trials, a non-profit organization, the Aeras Global TB Vaccine Foundation, has

a portfolio of promising vaccines. With its location in high incidence areas (South

Africa and India), the company is capable of pursuing phase III clinical trials (153).

Although there are drugs to treat TB, incomplete or erratic usage has led to the

development of drug resistant strains. Currently there are two categories of drug

resistance, multi-drug resistant (MDR-TB) and extensively drug resistant tuberculosis

28

(XDR-TB) strains. MDR-TB strains are resistant to the two major first-line drugs,

Isoniazid and Rifampin, while XDR-TB strains are resistant to the same drugs as well as

fluoroquinolones, and any second-line injectible drugs (154). With 68 countries

reporting in 2010 at least one case of XDR-TB, and the cost and long treatment regimen

of current available drugs, it is clear that TB is a major public health concern and new

drugs need to be identified to control the spread of this disease (155). As of 2010, there

are ten new compounds for the treatment of active TB progressing through the clinical

development pipeline. These drugs focus on novel regimens with 2-4 months of

treatment (156).

1.16.2 Infection with M.tb

Individuals with active pulmonary TB can expel bacilli on minute aerosol droplets by

coughing. A droplet containing as few as 1-10 bacilli is enough to cause an infection in a

healthy individual (157). These droplets are small enough to subvert upper airway

immune mechanisms such as ciliary beat and mucus clearance in order to reach the

alveolus. In the alveolus the bacilli encounter antimicrobial products as well as the major

cellular component of pulmonary innate immunity, the AM. The AM engulfs the

bacteria, but M.tb is able to subvert the immune response through many different

mechanisms (158).

The unique cell wall of M.tb, consisting of glycoproteins, glycolipids, lipoglycans and

long-chain fatty acids, enables the bacteria to survive in the AM, while at the same time

29

these factors can be processed for presentation to T-cells. Over 2-12 weeks following

primary infection the bacteria spread through the lymphatics and become systemic while

the immune system recruits cells, notably T-cells, to the infected macrophages forming a

granuloma. This cellular mass allows the healthy individual to contain the bacteria and

remain asymptomatic; the individual is termed as having a latent infection. Active

disease occurs when an individual becomes immunocompromised such as by old age or

acquiring another infection, particularly HIV. At this time, the bacteria are able to

replicate within the granuloma and induce a clinically evident disease state that the host

cannot control. Reactivation occurs in the lung 80% of the time [most notably seen in the

apex of the upper lobes of the lung (159)]; the remainder of cases are called

extrapulmonary TB and occur in the lymph nodes (scrofula), spine (Potts disease), or

systemic (miliary TB) among others.

1.16.3 M.tb and alveolus interaction

Upon entry into the alveolus, M.tb encounters surfactant, and two surfactant-associated

components that regulate the early interaction of M.tb in the lung, SP-A and SP-D. SPA

has been shown in vitro to enhance PRR activity, increase phagocytosis, alter production

of pro-inflammatory cytokines, and decrease RNI and ROIs in response to stimuli

(55,160). SPA enhances phagocytosis of M.tb in mouse and human macrophages, and

also has been shown to inhibit M.tb-mediated NO production from IFNγ stimulated

mouse macrophages (161-163). In contrast, SPD decreases phagocytosis of M.tb by

binding to the mannose caps of ManLAM and inhibiting the normal interaction with the

30

MR (164). In vitro studies have shown that SPD-opsonized M.tb that do enter the

macrophage have reduced intracellular growth that is due in part to the increased

phagosome-lysosome fusion (164,165).

1.16.4 M.tb and AM receptor interaction

M.tb interacts with a number of receptors on the AM surface as well as in the cytosol. In

vitro studies have shown that non-opsonic and serum opsonized bacteria coated with

complement C3 are taken up by the CR3 on macrophages (166). While CR3 plays a

clear role in phagocytosis, its role in pathogenesis is unclear. No affect was seen on the

production of ROIs, NO or bacterial survival in CR3 deficient murine BMMs infected

with M.tb (167) or in vivo in the mouse (168). Whether CR3 functions differently in the

mouse compared to man with regards to TB pathogenesis is unknown. The importance of

the MR during M.tb infection is better understood. M.tb is recognized by the MR through

its mannosylated lipoarabinomannan (ManLAM) found on the cell wall. Upon

phagocytosis by the MR, M.tb resides in a unique phagosome that does not fuse with the

lysosome. Scavenger receptor A (SR-A) and CD-14 have been shown to mediate

phagocytosis of unopsonized M.tb in different cell types (169,170). Avirulent and

attenuated mycobacterial strains have been shown to induce dectin-1/TLR2 TNF-α

production in murine macrophages (171). FcγRs mediate the phagocytosis of M.tb

opsonized with immunoglobulin in the immune host (172,173). The 19 kDa lipoprotein,

lipomannan (LM), and lower order PIMs found on the surface of the mycobacterial cell

wall have all been shown to interact with TLR2 (174-176).

31

Recent studies have provided data to support M.tb activating intracellular receptors, i.e.

NLRs. Masters et al published that M.tb inhibits inflammasome activation and IL-1β

production in primary murine macrophages; however, using a human monocytic cell line

researchers were able to show M.tb does activate the NLRP3/ASC inflammasome

(177,178). Although the ligand has not been determined, several reports indicate that

NOD1 and NOD2 recognize M.tb (127,135,179-181). Our recent data presented in this

dissertation provide evidence that NOD2 is important in controlling the growth of M.tb in

human macrophages and this is may be due to the role of NOD2 in inflammatory

cytokine production (77).

1.16.5 M.tb and phagosome environment

Once inside the macrophage M.tb resides in a unique phagosome. The environment in

the phagosome has a pH of 6.4, due to limited fusion of pre-formed lysosomes (182,183).

Phagosomes containing M.tb resemble early endosomes. The M.tb cell wall component,

ManLAM, is able to inhibit calcium increase in the cytosol that occurs normally during

phagocytosis and without this calcium PI3K does not become activated (184). M.tb also

inhibits sphingosine kinase, which inhibits calcium efflux (185). Due to the ability of

M.tb to inhibit calcium- and Rab5-dependent recruitment of PI3K, PI3P is not formed on

the membrane and there is a lack of Rab5 activity. This lack of activity inhibits full

maturation of the phagosome by the lack of recruitment of Rab5 effector proteins EEA1

and syntaxin-6 (186). Entry through the MR leads to an anti-inflammatory program

32

along with a unique phagosome that has reduced fusion with the lysosome (187,188).

This phagosome does not acidify normally in part because it does not acquire all subunits

of the vacuolar proton ATPase (189,190).

1.16.6 M.tb secretion systems

The complex, thick and relatively impermeable cell wall of mycobacteria suggests the

need for a secretion system, and in fact there are many secretion systems identified in

mycobacteria including early secretory antigenic target 6 system 1 (ESX-1) through

ESX-5. Many reports provide evidence for the release of proteins by M.tb into the

supernatants of in vitro broth cultures (191). Specifically, culture filtrate proteins of 10

kDa (CFP10) and early secreted antigen target of 6 kDa (ESAT6), which are products of

the ESX-1 secretion system, are particularly important. These two proteins are encoded

by genes found in the region of difference 1 (RD1) and are attributed to M.tb virulence

(192,193). During serial passage, the attenuated BCG lost 38 open reading frames,

including the RD1 (192,193).

Putative analysis of the clustering of genes that encode membrane associated proteins

with the genes that encode CFP10 and ESAT6 led to the idea and eventual confirmation

of the ESX-1 secretion system (194-196). Addition of the RD1 region back into BCG

restored the production of ESAT6. Also, disruption of genes in the ESX-1 locus,

Rv3870, Rv3871, and Rv3877 inhibited secretion of CFP10 and ESAT6. Further studies

indicate that there are 14 possible genes in the RD1 necessary for optimal function of

33

ESX-1, and a number of these proteins have been associated with known functional

domains. These include Rv3868, a putative cytoplasmic chaperone with an AAA+

ATPase domain, Rv3883c (MycP1), a subtilisin-like serine protease, and

Rv3870/Rv3871, which together form an FtsK/SpoIIIE-like ATPase. Rv3860, Rv3877,

Rv3881c, and Rv3882 are involved in ESAT6 secretion but have no known functions due

to lack of homology with described proteins. These proteins are thought to be located in

the cytoplasmic membrane. Rv3877 is predicted to be a multi-trans-membrane-spanning

protein and hypothetically could be the translocation pore in the cytosolic membrane

(191). The discovery of a second gene cluster, the Rv3614c-Rv3616c locus, involved in

ESAT6 secretion has given further complexity to the ESX-1 system (197,198). The lack

of similarity to type I- type VI secretion systems and the identification of additional ESX-

1 systems within mycobacteria and other Gram-positive species has led to the

characterization of a type VII secretion system (191).

While there is no structure determined yet for the type VII secretion system there are

protein-protein interaction studies that indicate the mechanism of secretion. In order for

secretion to occur, ESAT6 and CFP10 form a tight dimer held together by hydrophobic

interactions. These two proteins are dependent on each other for secretion (199-201).

Yeast two-hybrid screening showed Rv3871 interacts with Rv3870 along with CFP10.

One hypothesis is that Rv3871 is involved in recognition of the CFP10-ESAT6 substrate

pair and delivers it in an ATP-dependent manner to Rv3870 and thereby to the cell

membrane secretion machinery. This is mediated by an interaction of Rv3871 with the

34

carboxy-terminus of CFP10 (199). This carboxy-terminus is similar to the secretion

signal found in the type IV secretion system (202-204). EspA or Rv3616c and Rv3615c

are also substrates that are secreted similarly to CFP10 and ESAT6 (198). All of the

substrates of the ESX-1 system are dependent on each other for secretion. While the

mechanism is not yet understood, in the absence of EspA or Rv3615c, the CFP10/ESAT6

dimer is not secreted (197,198).

Although their roles in virulence are not as clear as the type VII system, mycobacteria

contain two other types of secretion systems, general secretion pathways (Sec) and the

twin-arginase transport (Tat) pathway. Mycobacteria contain a general secretion

pathway, like all bacteria, in order to secrete unfolded proteins across the cytosolic

membrane. SecA2 was identified in mycobacteria as the non-essential homologue of

SecA and its role in pathogenesis is unclear. However, SecA2 functions in L.

monocytogenes as an accessory protein to secrete enzymes that remodel the

peptidoglycan layer and these broken down products are shown to modify the innate

immune response to favor bacterial colonization (205). The mycobacterial SecA2 mutant

induces a more robust inflammatory immune response in infected macrophages

indicating a role for it in suppressing the immune response (206). The Tat pathway is a

Sec-independent pathway that has the ability to transport folded protein substrates across

the plasma membrane (207,208). In the mouse model, four M.tb phospholipase C

enzymes are secreted by Tat and are required for full M.tb virulence (207,209).

35

Secretion systems may play a key role in the ability of intracellular NLRs to sense

mycobacteria, since the bacteria reside in a distinctive phagosome separate from the

cystosol. Given that mycobacteria secrete proteins, the potential exists for these proteins

to get across the phagosomal membrane and into the cytosol of infected macrophages to

interact with NLRs. Secretion systems are reported to be important in enabling NLRs to

sense H. pylori and Legionella pneumophila (93,137,210,211).

1.16.7 M.tb and entry into the cytosol

It is widely accepted that M.tb resides in a unique phagosome, which allows for its

survival in antigen presenting cells along with the recruitment of CD4 T cells though

MHCII-based antigen presentation. However, the mechanisms underlying the ability of

virulent M.tb to use MHCI-based antigen presentation in order to recruit CD8 T cells

remains less clear.

While still debated, there is evidence that M.tb can escape from the phagosome and reside

in the cytosol. The first report, using electron microscopy (EM), found that at 18-24 h 60

– 100% of endocytosed M.tb H37Rv by rabbit alveolar macrophages resulted in bacteria in

the cytosol. While the less virulent strain, M.tb H37Ra was located in intact phagosomes

99% of the time (212). Another report found, also using EM, that M.tb H37Rv was

located in the cytosol or vesicles 50% of the time four days post infection. The authors

found the M.tb H37Ra was intermediate, while BCG was never found in the cytosol (213).

These results are debated due to the cell types as well as the harsh fixation techniques

36

used to prepare samples for EM. The cells in this study were J744.16, a mouse

macrophage cell line, and also human monocytes. It is important to note that monocytes

do not express the MR, and M.tb entry through this receptor causes a decrease in

phagosome-lysosome fusion (188,214).

Recently, another group reported visualizing M.tb in the cytosol using cryogenic-EM, a

more credible method for the preparation of samples when using EM. Cryogenic-EM is

the use of low temperatures in order to keep the sample in its native state. They also

used immunogold labeling, which allowed visualization of phagolysosomes. Using

monocyte-derived dendritic cells, the authors report visualizing M.tb and M. leprae in the

cytosol 48 to 96 hours post infection. While they did visualize BCG in a phagolysosome

7 days post infection, no bacteria were found in the cytosol. The data also indicate the

genes encoding CFP-10 and ESAT6 are necessary for mycobacterial entry into the

cytosol (215). The presence of M.tb in the cystosol has implication for recognition by

intracellular NLRs.

1.17 M.tb and NOD2

Recent studies provide evidence that NOD2 is involved in the pathogenesis of

mycobacterial infections. For example, a genome analysis of patients with leprosy found

a single nucleotide polymorphism in the NOD2 gene (216) which is linked to

Mycobacterium leprae susceptibility. Furthermore, African American patients with TB

have variants in their NOD2 gene rendering them more susceptible to M.tb infection

37

(217). In addition, polymorphisms in CARD15, which encodes NOD2, have been linked

to inflammatory diseases, including Crohn’s disease (138,218), whose cause is speculated

to be of bacterial origin (219). The association of Crohn’s disease with Mycobacterium

avium subspecies paratuberculosis continues to be reported (144).

The majority of studies to examine the role of NOD2 in the host response to M.tb

infection have been done in mice, murine macrophages, or in transfected cell lines

(127,135,179-181). There are limited reports using either normal PBMCs, PBMCs from

patients with the NOD2 mutation associated with Crohn’s disease, mixed

bronchoalveolar lavage (BAL) leukocytes, and none prior to our work using human

monocyte derived macrophages (MDMs) (127,220). A recent report indicates that NOD2

and TLR pathways are non-redundant in the recognition of M.tb, but can synergize to

induce a robust pro-inflammatory response (127). In addition, one study shows that RIP2

polyubiquitination occurs during M.tb incubation in a NOD2-dependent manner in

BMMs (180). Other studies in mouse macrophages have shown that NOD2 does not

have a significant role in controlling M.tb growth during early infection (179) but may

have during late infection (135). A decrease in pro-inflammatory cytokine production

was observed in mouse NOD2 knockout BMMs and naive murine AMs in response to

M.tb without affecting intracellular bacterial growth (135,179). Studies using human

macrophages report the same role for NOD2 in cytokine production in response to M.tb;

however in contrast, data indicate an important role for NOD2 in controlling the growth

38

of M.tb (77). This finding is yet another example of the difference between human and

mouse.

1.18 Mouse and human differences in the immune system

While the human and mouse genomes only differ in approximately 300 genes, they differ

greatly on development, activation, and response to challenge in the innate and adaptive

immune system (221). For example, there are differences in the bronchus-associated

lymphoid tissue that is present in the mouse but not human (222). Though it is not clear

the consequence, mice have more lymphocyte rich blood compared to humans who have

more neutrophils present (223). In the area of microbial defense, defensins are produced

by neutrophils in humans and by Paneth cells in mice. Humans produce 2 defensins from

Paneth cells, compared to mice, which produce over 20 defensins from this cell type

(224). NO production is easily detected in mouse cells whereas it is more difficult to

measure in human macrophages. It is debated whether or not the assays used to detect

NO are sensitive enough to measure the NO produced in human macrophages, but it is

clear that mouse macrophages produce a great deal of this antimicrobial mediator

compared to human macrophages (225). There are five structurally distinct homologues

of the receptor Dendritic Cell-Specific Intercellular adhesion molecule-3-Grabbing (DC-

SIGN) in mice and only two found in humans (226). DC-SIGN is an adhesion, signaling,

and uptake receptor that functions to recognize mannose type carbohydrates found on

bacteria and fungi (227). Non-integrin FcR receptors are important links between the

innate and adaptive immune response and there are numerous differences between mice

39

and humans. Mice lack the FcαRI receptor, which is an important IgA receptor found in

humans, but they likely use other receptors for this function. FcγRIIA and FcγRIIC are

two IgG receptors expressed in humans and not mice (228,229). It has also been found

that depending on the mouse strain there are differences in the expression of IgG isotypes

as well as class switching. For example IL-4 in humans induces class switching of IgG4

and IgE, whereas in mice it induces switching to IgG1 and IgE (230). A few other

significant differences to be noted between mice and humans are in B cell development,

development and regulation of T cells, and IFN-α signaling (231-233). Finally, mice and

humans differ in immunity-related guanosine triphosphatases (IRGs), which are

important in innate immunity against pathogens and play a role in autophagy. There are

23 identified IRG genes in mice and only 3 in humans (234,235). The enhanced number

of IRG genes in mice could potentially lead to different functional outcomes in mice

when compared to humans.

1.19 Summary

The mammalian immune system, adaptive and innate, is uniquely designed to recognize

and respond to invading pathogens. The innate immune system plays an important role

in this function through the deployment of phagocytes that have extracellular and

intracellular PRRs.

One bacterial pathogen in particular that has found mechanisms to evade the innate

immune system is M.tb. The necessity to find better methods to control this pathogen is

40

at an all time high due to the emergence of MDR- and XDR-TB strains. In order to do

this, the mechanisms for how this bacterium is recognized inside of the macrophage is

important to understand because it will allow for the development of new ideas for the

generation of novel microbe- and host-directed therapeutic approaches. The intracellular

PRR, NOD2, has emerged as an important innate immune player in the recognition and

control of pathogens in the host and data suggest that NOD2 may be important in the

pathogenesis of M.tb.

1.20 Specific aims

It is important to study the interaction between M.tb and the immune system in order to

understand its pathogenesis and find new potential drug targets. Studying M.tb’s

interaction with the first major innate immune cell it comes in contact with in the

alveolus, the macrophage, is of particular relevance. The interaction of M.tb and

membrane receptors has been well studied; however, M.tb’s interaction with cytosolic

receptors has yet to be thoroughly studied in human macrophages. We hypothesize that

NOD2 is important in controlling the growth of M.tb in human macrophages by

regulating the nature of the inflammatory response in order to contain infection. We also

hypothesize that NOD2 plays a role in several novel early innate immune responses in

human macrophages and focus on its role in iNOS expression and PKCδ activation. The

specific aims of this dissertation are as follows:

(1) Characterize NOD2 expression in human macrophages and determine the role

of NOD2 during M.tb infection of human macrophages

41

(2) Determine if NOD2 is necessary for iNOS expression in human macrophages

in response to agonists as well as M.tb

(3) Determine NOD2 signaling partners in the innate immune response,

specifically the function of NOD2 in the novel PKC signaling cascade

42

CHAPTER 2 : NOD2 controls the nature of the inflammatory response and subsequent

fate of Mycobacterium tuberculosis and M. bovis BCG in human macrophages

2.1 Introduction

Tuberculosis (TB) is currently the leading cause of death in HIV-infected persons (236).

Multi-drug resistant Mycobacterium tuberculosis (M.tb) strains have limited our ability to

use the current antibiotic supply, and the development of new antibiotics requires a better

understanding of the biological mechanisms underlying M.tb growth and survival within

the host. Humans are the only known natural host for M.tb and alveolar macrophages

(AMs) in the lung microenvironment are the first immune cells that encounter inhaled

mycobacteria (237). Thus, knowledge about pathogenic events for M.tb in primary

human macrophages is particularly important.

Macrophages are equipped with an array of defense mechanisms to recognize and clear

invading pathogens. Several studies demonstrate that M.tb interacts with macrophage

membrane receptors such as the mannose receptor (MR), complement receptors, Toll-like

receptors (TLR) and scavenger receptors (237). Engagement of specific receptor

43

pathways during phagocytosis of mycobacteria regulates the early host response through

defined signaling pathways (158,175,238). Among these are the group of cytosolic

receptors called Nucleotide-binding Oligomerization Domain (NOD)-like receptors

(NLRs) (93) which function as intracellular pattern recognition receptors (PRRs) that

recognize pathogens or pathogen secreted molecules (239).

There are 23 NLRs which can be divided into four subfamilies depending on the

composition of their N-terminal effector domain. For example, those containing a

caspase recruitment domain (CARD) as their effector domain are known as the NLRC

subfamily and consist of proteins that initiate a pro-inflammatory response by signaling

through their CARD domain. NOD2, a member of this subfamily, is found in epithelial

cells and antigen presenting cells such as macrophages (76,240,241). NOD2 regulates the

production of inflammatory mediators in response to muramyl dipeptide (MDP) which is

found in Gram-negative and Gram-positive bacteria, and mycobacteria (86,88). Many

studies have found that NOD2 is able to recognize and generate a response to a variety of

microorganisms (97,126,127,136,242-248). NOD2 senses bacteria through its leucine

rich repeats (LRR), then oligomerizes through its NOD domain and is able to recruit

receptor interacting protein 2 (RIP2) through an electrostatic interaction of the CARD

domain in each protein. This leads to a signaling cascade in which activation and

translocation of nuclear factor-κB (NF-κB) into the nucleus occurs in order to transcribe

pro-inflammatory cytokines (108,147,241,249,250).

44

As reported earlier in the introduction section (See: 1.17 M.tb and NOD2), the literature

indicates that during mycobacterial infections NOD2 is an important player in the host

response. Two genomic analysis studies revealed an association of NOD2 with

susceptibility to mycobacterial infections. Patients with Mycobacterium leprae or

African American patients with TB had polymorphisms or variants in their NOD2 gene

(216,217). Patients with Crohn’s disease, an inflammatory granulomatous disease, are

reported to have polymorphisms in their NOD2 gene (CARD15) (138,218). While

debated, it is reported that the cause of Crohn’s disease is Mycobacterium avium

subspecies paratuberculosis (144).

The role of NOD2 in the host response to M.tb infection has mainly been studied in mice,

murine macrophages, or transfected cell lines, with limited reports using PBMCs from

healthy donors, PBMCs from patients with mutations in their NOD2 gene, and mixed

BAL leukocytes (127,127,135,179-181,220). NOD2 and TLR pathways are non-

redundant in the recognition of M.tb, but can synergize to induce a robust pro-

inflammatory response (127). In addition, RIP2 polyubiquitination occurs during M.tb

incubation with BMMs in a NOD2 dependent manner (180). While a decrease in pro-

inflammatory cytokine production in NOD2 knockout BMMs and naïve murine AMs in

response to M.tb was found, NOD2 did not control intracellular bacterial growth during

early in vivo infection in mice (179) but did during late infection (135).

45

Interest in the potential interactions between M.tb and cytosolic PRRs such as NOD2

continues to grow based on recent reports that M.tb secretes virulence factors and

immune activators or suppressors through a type VII secretion system (191,251,252).

Other reports provide evidence that M.tb may be able to escape the phagosome to the

cytosol under specific conditions (213,215).

Based on the known differences between human and mouse macrophages in receptor

expression, signaling pathways, and innate immune function (253-256), we decided to

study NOD2 in human macrophages and in response to M.tb by optimizing an antibody

to NOD2 and by accomplishing effective small interfering RNA (siRNA)-mediated

knockdown of NOD2 in these cells. We show that NOD2 expression increases during

monocyte differentiation. In contrast to previously published work using mouse

macrophages, we report that NOD2 is important in controlling both pro-inflammatory

cytokine production and the growth of M.tb in human macrophages; whereas NOD2

controls the growth of the related attenuated BCG strain in both human and mouse

macrophages.

2.2 Materials and Methods

2.2.1 Buffers and reagents

Dulbecco’s PBS with and without Ca2+ and Mg2+ ions, Iscoves, FBS, TRIzol, and RPMI

1640 medium with L-glutamine were purchased from Invitrogen (Invitrogen Life

46

Technologies). RHH medium [RPMI 1640 plus 10 mM HEPES (Invitrogen) plus 0.4%

human serum albumin (CSL Behring LLC)] was used for cell culture experiments. N-

Acetylmuramyl-L-alanyl-D-isoglutamine hydrate (MDP; Sigma-Aldrich) and H-Ala-D-γ-

Glu-diaminopimelic acid (iE-DAP; AnaSpec) were used for functional assays involving

ELISAs. 7H11 agar was prepared with Bacto Middlelebrook 7H11 agar, oleic acid-

albumin-dextrose-catalase enrichment medium, and glycerol (Difco Laboratories).

2.2.2 Antibodies

Unconjugated rabbit anti-human NOD2 was purchased from ProSci Inc. and another anti-

human NOD2 antibody was obtained from Dr. Nunez, University of Michigan, and

available by ebioscience. Syk (N-19), TLR2, actin, goat anti-rabbit IgG-HRP and

donkey anti-goat IgG-HRP antibodies were purchased from Santa Cruz Biotechnology.

The aforementioned antibodies were used for Western blot. The NOD2 antibody from

ProSci Inc., Alexa Fluor 488-conjugated goat anti-rabbit IgG (Molecular Probes and

Invitrogen) and normal rabbit IgG (Santa Cruz Biotechnology) were used for confocal

microscopy experiments. Monoclonal anti-human NOD2 (Cayman Chemical) and anti-

mouse IgG1 kappa (Abcam) were used for IP experiments. Antibodies specific for

phospho-RIP2 (Ser176) and phospho-IκB (Ser32) were purchased from Cell Signaling

Technology. The caspase-1 antibody was a kind gift from Dr. Mark Wewers, Ohio

State University.

2.2.3 Bacterial strains and culture

Lyophilized M.tb H37Rv (ATCC #25618) and BCG (ATCC #35734) were obtained from

American Tissue Culture Collection, reconstituted, and used as described (172). The

concentration of bacteria (1–2 x 108 bacteria /mL) and the degree of clumping ( 10%)

were determined by counting in a Petroff-Hausser chamber. Bacteria prepared in this

fashion are 90% viable by CFU assay.

2.2.4 Isolation and culture of human monocytes and macrophages

Blood obtained from healthy, Purified Protein Derivative negative human volunteers

using an approved protocol by The Ohio State University Institutional Review Board

(IRB) was processed as described (257). Briefly, PBMCs were isolated from heparinized

blood on a Ficoll-Paque cushion (Amersham Biosciences/GE Healthcare) and cultured in

Teflon wells (Savillex) for 1 (monocytes) through 5 (MDMs) days in the presence of

RPMI 1640 medium containing 20% autologous serum (2.0 x106 PBMCs/mL) at

37ºC/5% CO2 (258). Human AMs were isolated from BALs of healthy human donors

as previously described (161) using an approved protocol by The Ohio State University

IRB. Briefly, the BAL was centrifuged (200 x g, 4°C, 10 min), supernatant was removed,

and the pellet re-suspended in RPMI and washed two more times with RPMI. AMs were

plated in tissue culture plates with 10% human serum supplemented with penicillin

(1000U/mL), incubated in CO2 incubator at 37ºC for 2 h, washed with warm RPMI to

remove non-adherent cells and penicillin, and used for experiments.

47

48

2.2.5 Mice and mouse macrophages

Wild-type C57BL/6 and NOD2-/- mice were backcrossed to C57BL/6 background for 10

generations (242). BMMs were prepared from femurs of five to eight week-old mice as

previously described (137,259).

2.2.6 RNA isolation

One to 5 day-old PBMCs in Teflon wells were harvested and mononuclear phagocytes

adhered to a 6-well tissue culture plate with 10% autologous serum (8 x 106 PBMCs/mL).

After washing away lymphocytes, the monocytes or MDMs (8 x 105 cells) were lysed in

TRIzol and total RNA was isolated by using the Qiagen RNAeasy column method. The

Experion (Bio-Rad) was used to determine the quality and quantity of RNA samples.

2.2.7 Gene expression studies by qRT-PCR

RNA (100ng) was reverse transcribed to cDNA by reverse transcriptase enzyme,

(SuperScript II, Invitrogen) and qRT-PCR was performed using a human NOD2 or IL-1β

TaqMan gene expression kit (Applied Biosystems). NOD2 and IL-1β amplifications

were normalized to β actin as a housekeeping gene (∆Ct). Relative copy number (RCN)

and fold change were determined. RCN was calculated as follows: RCN = E-∆Ct x 100,

where E is the efficiency (2 = 100% efficiency) and Ct = Ct (target) – Ct (reference)

(260). Fold change was calculated from the RCN values compared to day 1 for NOD2

49

mRNA experiments, and compared to the siRNA resting group for IL-1β mRNA

experiments. Triplicate samples were analyzed in duplicate wells in each experiment.

2.2.8 Confocal microscopy

Untransfected MDMs or transfected (scramble siRNA or NOD2 siRNA) MDMs (2 x 105)

were adhered to a glass coverslip in a 24-well tissue culture plate for 2 h at 37°C and

washed to remove lymphocytes. The cells were washed, fixed with 2% paraformaldehyde

(10 min at room temperature), permeabilized by treatment with 100% methanol for 4

min, washed and then incubated overnight in 10% non-fat milk in PBS as blocking

reagent. The cells were then incubated with NOD2 (2.5µg/mL) or the appropriate isotype

(2.5µg/mL) control antibody for 1 h at room temperature, washed with blocking

reagent, and counterstained with an Alexa Fluor 488-conjugated secondary antibody

(1:500 dilution) for 1 h at room temperature. Nuclei were labeled with 0.1g/mL of the

DNA stain 4,6-diamidino-2-phenylindole (DAPI; Invitrogen/Molecular Probes) in PBS

for 5 min at room temperature. The cells were next washed with PBS and coverslips

were mounted on glass slides. Slides were viewed using an Olympus Flowview 1000

Laser Scanning Confocal microscope.

2.2.9 Transfection of MDMs

PBMCs were transfected with Accell scramble siRNA or NOD2 siRNA purchased from

Thermo Scientific Dharmacon RNAi technologies that targeted the sequence

50

CUUUAGGAUGUACAGUUA (368nM) using the Amaxa Nucleofector (Lonza Group

Ltd) Briefly, PBMCs (1× 107) were resuspended in 100 μl of nucleofector solution

followed by the addition of 5ug/mL of scramble siRNA or NOD2 siRNA, incubated at

room temperature for 5 min, and nucleofected according to the manufacturer’s

instructions. PBMCs were then seeded on 12-well plates containing 1.0 ml of RPMI

supplemented with 10% autologous serum and incubated for 2 h at 37°C and 5% CO2.

After 2 h, adhered transfected cells (MDMs) were washed and repleted with warm RPMI

containing 20% autologous serum for 72 h. Transfected cells were used for subsequent

experiments, such as CFU assays, cytokine assays, confocal microscopy, cell

enumeration assays, and Western blotting. Assessment of monolayer density in

experiments was performed as described below.

2.2.10 Macrophage stimulation / infection, cell lysis, immunoprecipitation, and western

blotting

Day 5 MDMs or MDMs transfected with scramble siRNA or NOD2 siRNA (2.0 x 105)

were incubated with M.tb or BCG (MOI 5:1; triplicate wells) in RHH for 30 min at 37°C

in 5% CO2 on a platform shaker for equal dispersion of bacteria followed by an additional

incubation for 90 min without shaking. The cells were then washed three times with

warm RPMI, repleted with RPMI containing 1% human autologous serum and incubated

for 24 h at which time supernatants were collected for ELISAs (see below). For

experiments using the NOD2 ligand MDP, 2.0 x 105 day 5 MDMs were adhered to a 24-

well plate for 2 h at 37°C in 5% CO2, washed to remove lymphocytes, and then incubated

51

with MDP (5ug/mL) for 90 min. Cells were then incubated with bacteria as described

above (2 h total), washed and repleted with MDP in 1% autologous serum for 24 h. Cell

culture supernatants were harvested and used for ELISAs. The monolayers were washed

once with PBS and lysed in TN-1 lysis buffer [50 mM Tris (pH 8.0), 10 mM EDTA, 10

mM Na4PO7, 10 mM NaF, 1% Triton X-100, 125 mM NaCl, 10 mM Na3VO4, 10

µg/ml aprotinin, and 10 µg/ml leupeptin)], incubated on ice for 10 min and then

centrifuged at 17,949 x g for 10 min at 4°C to remove cell debris (261). For infections,

MDM monolayers were incubated with M.tb or BCG in RHH for 15 min, 30 min, 1 h, 1.5

h, 2 h, and 3 h (MOI 5:1). At each time point, cells were washed with PBS and lysed as

above. For NOD2 knockdown assessment, lysates were collected as above at 72 h and 96

h post-knockdown. Protein concentrations of the cleared lysates were measured using the

BCA-protein assay kit (Pierce). Proteins were separated by SDS-PAGE and analyzed by

Western blot by probing with the primary and secondary antibody of interest and

development using ECL (Amersham Biosciences). For the IP experiment whole cell

lysates were collected from HEK293T (empty vector) and NOD2-expressing HEK293T

cells. Equal amounts of protein were incubated overnight with NOD2 antibody (Cayman

Chemicals) or control IgG1, kappa. Then cells were incubated for 2 h with Protein G

agarose beads (Invitrogen), washed, pellets boiled in Laemmli buffer and loaded on a

7.5% SDS-PAGE gel, and Western blot was performed using a NOD2 antibody (ProSci

Inc.). The protein band intensities were measured using the online ImageJ software

provided by the NIH. Background intensity was subtracted from each sample and then

52

normalized to β-4actin. Fold change was determined as follows: (treated sample band

intensity) / (untreated sample band intensity).

2.2.11 ELISAs

For MDM experiments, cell free culture supernatants were collected at 24 h and used to

measure TNF-α and IL-1β by ELISA (R&D Systems). For mouse macrophage studies,

WT and NOD2-/- macrophage monolayers were incubated with serum-opsonized

mycobacteria (M.tb MOI 1:1 and BCG MOI 2.5:1) in the same manner as described for

MDMs. Cell free culture supernatants were collected at 24 h, 72 h, and 168 h and

measured for IL-1β by ELISA (BD Biosciences).

2.2.12 M.tb and BCG intracellular growth assays in macrophages

A CFU assay using infected macrophage lysates was performed as described (262).

Briefly, scramble siRNA or NOD2 siRNA transfected MDM monolayers were incubated

with M.tb or BCG (MOI 5:1; triplicate wells) in RHH for 30 min at 37°C in 5% CO2 on a

platform shaker for equal dispersion of bacteria and for an additional 90 min stationary.

The cells were then washed three times with warm RPMI and either lysed or repleted

with RPMI containing 1% human autologous serum and incubated for 2 h or 24 h. The

supernatants and MDM monolayers were then processed for ELISA or enumeration of

CFUs, respectively (triplicate samples in each test group). The number of colonies was

counted after 4 weeks and data analyzed. For mouse macrophage experiments, WT and

53

NOD2-/- macrophages were incubated with serum-opsonized mycobacteria (M.tb MOI

1:1 and BCG MOI 2.5:1) in Iscoves plus 10% FBS using the same incubation protocol as

for MDMs and then lysed at 2 h, 24 h, 72 h, and 168 h.

2.2.13 M.tb and BCG cell association assay within macrophages

Scramble siRNA or NOD2 siRNA transfected MDM monolayers (2.0 x 105), or WT or

NOD2-/- mouse macrophages monolayers (2.5 x 105) were incubated with M.tb or BCG

(MOI 10:1; triplicate wells) in RHH for 30 min at 37°C in 5% CO2 on a platform shaker

for equal dispersion of bacteria and then for an additional 90 min without shaking. The

cells were then washed three times with warm RPMI and fixed with 10% formalin for 10

min at room temperature. Coverslips were then washed three times with PBS, dried and

cell-associated M.tb or BCG was stained with auramine-rhodamine (BD Biosciences).

Three hundred consecutive macrophages/coverslip/test group were counted using phase-

contrast and fluorescence microscopy (166).

2.2.14 Cell enumeration of NOD2 siRNA transfected MDMs in monolayer culture

Scramble siRNA and NOD2 siRNA transfected MDMs were plated in 24 well plates and

after 2 h, washed and repleted with 20% autologous serum for 72 h. Cells were then

washed one time with PBS and treated with 1% Cetavlon in 0.1M citric Acid with 0.05%

Napthol blue black (Sigma-Aldrich), pH 2.2, for 15 min at room temperature. Cell

54

lysates were then loaded on a hemacytometer and stained nuclei enumerated using phase

contrast microscopy (263).

2.2.15 NOD2 antibody peptide competition assay

The NOD2 antibody and the NOD2 peptide (ProSci Inc.) used to produce the antibody

were incubated in a ratio of 2:1 (peptide:antibody) at 37°C for 30 min in a 5% milk

solution in low retention tubes. Day 5 MDM lysates were run in parallel on a 7.5% SDS-

PAGE gel, transferred to a nitrocellulose membrane and the membrane cut so that each

lane could be probed with either the antibody-peptide mixture or antibody alone

overnight at 4°C. Western blotting was performed as described above. Membranes were

then developed using ECL.

2.2.16 Statistics

A paired one tailed student’s t test was used to analyze the differences between two

groups in figures showing one representative figure of three independent experiments.

An unpaired one tailed student’s t test was used to analyze differences between two

groups in figures showing cumulative data from three independent experiments.

Significance was P < 0.05.

55

2.3 Results

2.3.1 NOD2 mRNA and protein expression increase as monocytes differentiate into

macrophages

The expression of NOD2 in human macrophages has not been studied and there are few

reports on NOD2 mRNA expression in human monocytes (76,220). We began our

studies by determining the expression of NOD2 during monocyte differentiation to

MDMs in an established model by quantitative real-time PCR (qRT-PCR) and Western

blot (54). Results from qRT-PCR provide evidence that NOD2 transcript levels are

increased during the differentiation of monocytes to macrophages. While we saw

production of mRNA in day 1 monocytes, the results in Fig. 2.1A show a significant

increase in mRNA at day 3 of differentiation which remains elevated as the cells

differentiate further into mature macrophages. Next, we compared the mRNA levels with

the amount of protein produced during cell differentiation by Western blotting. Due to

the sensitivity and specificity issues in using antibodies specific for human NOD2, we

optimized an available commercial antibody (ProSci Inc.) with human macrophage

lysates. The specificity of this antibody was confirmed by a peptide competition assay

(Figure 2.1B), an immunoprecipitation (IP) experiment (Figure 2.1C), and in siRNA

experiments shown later. For IP experiments, proteins in cell lysates from HEK293T

cells expressing human NOD2 or HEK293T cells containing the empty vector were

immunoprecipitated with NOD2 antibody (Cayman Chemical) and then probed with

NOD2 antibody from ProSci Inc. As shown in Figure 2.1C, NOD2 antibody specifically

binds to the human NOD2 protein expressed in the HEK293 cell line. We were also able

56

to detect NOD2 in human lysates using the antibody available from eBioscience,

although with more background (data not shown). Thus, in all subsequent experiments

we used the antibody from ProSci Inc. In accordance with the mRNA studies, NOD2

protein levels in cell lysates increased markedly at day 3 of differentiation and remained

high on day 5 MDMs (Figure 2.1D upper panel). The densitometric analysis of Western

blots from three independent experiments shows that NOD2 protein expression levels

increased during cell differentiation (Figure 2.1D lower panel). Finally, our results were

confirmed using confocal microscopy where NOD2 protein was easily detected

throughout MDMs in a punctuate appearance (Figure 2.1E).

Figure 2.1: NOD2 expression increases during monocyte differentiation to macrophages.

Continued on next page

PBMCs were harvested from Teflon wells each day from day 1 through day 5 and monocytes/MDMs plated on tissue culture plates, and incubated at 37°C / 5% CO2 for 2 h. The resultant monocyte/MDM monolayers were either lysed with TRIzol for extraction of total RNA or TN-1 lysis buffer for cell lysate preparation. (A) NOD2 steady-state mRNA levels were measured during monocyte to macrophage differentiation by real time PCR. Data were normalized to the β actin gene and RCN was determined. Shown are cumulative data obtained from three independent experiments each performed in triplicate (mean ± SEM, *P < 0.05). (B) NOD2 antibody specificity was confirmed by peptide blocking using 2-fold excess NOD2 peptide (used to generate the NOD2 antibody) and Western blotting using anti-NOD2 antibody. Shown is a representative blot from two independent experiments. (C) HEK293T vector (lane 1) and HEK293T cells expressing NOD2 (lane 2) were lysed and 100 µg of total protein in each case was immunoprecipitated with NOD2 antibody (Cayman Chemicals) or isotype control (IgG1 kappa) and subsequent Western blots of the immuoprecipitants probed

57

58

with NOD2 antibody (ProSci Inc.). Shown is a representative Western blot from two independent experiments. (D) Monocyte/macrophage cell lysates were analyzed for NOD2 protein expression by Western blot using NOD2 antibody. The same membrane was re-probed with actin to verify equal loading of lysate proteins. Shown is a representative Western blot and the bar graph shows the cumulative densitometric analysis of Western blots from three different donors (mean ± SEM, *P < 0.05 ** P < 0.005). (E) Day 5 MDM monolayers on coverslips were fixed, permeabilized, and stained using either NOD2 antibody or isotype control antibody followed by Alexa flour 488 conjugated secondary antibody. Cells were then examined by confocal microscopy (Olympus FV1000). Shown is a representative section (63x mag) of a Z series image from one of three independent experiments performed on duplicate coverslips.

2.3.2 Activation of NOD2 enhances the M.tb- and BCG-mediated pro-inflammatory

cytokine response through the NF-κB pathway

We next investigated whether activation of NOD2 leads to an increase in the M.tb- and

BCG-mediated pro-inflammatory cytokine response in macrophages. Incubation of

macrophages with MDP, a NOD2 agonist, is known to induce the production of pro-

inflammatory cytokines such as IL-1β (122). A recent report indicates that in some TB

patients, NOD2 mRNA is enhanced in PBMCs; furthermore, patients on anti-tuberculosis

treatment have enhanced NOD2 levels in leukocytes (220). MDMs were pre-incubated

with the NOD2 ligand MDP for 90 min followed by incubation with M.tb or BCG as

described in the methods in the presence of MDP. Cell free culture supernatants were

analyzed for TNF-α and IL-1β secretion at 24 h. The data show an additive (TNF-α) or

synergistic (IL-1β) effect on cytokine production in cells pre-treated with MDP compared

to cells incubated with mycobacteria alone (Figure 2.2A and B). Thus, our results

indicate that pre-engagement of NOD2 by its ligand significantly enhances M.tb- or

59

BCG-mediated pro-inflammatory cytokine production. This effect is likely due to the

activation of NF-κB through the phosphorylation of IκBα and its subsequent degradation

(180,264).

To further explore the signaling pathway(s) involved, we examined whether M.tb- or

BCG-mediated signaling events include those known for the NOD2 pathway. NOD2

engagement leads to recruitment of a serine threonine kinase adapter protein called RIP2

by an electrostatic interaction through their CARD domains, which in turn causes

autophosphorylation of RIP2 at serine 176 (94) in cell lines and murine macrophages.

Activated RIP2 phosphorylates inhibitor of kBα (IkBα) resulting in the activation and

translocation of NF-κB. MDMs were incubated with M.tb or BCG for different time

points and the cell lysates were analyzed for activation of RIP2 and IkBα by Western

blotting of the phosphorylated proteins. The data show that both M.tb and BCG

phosphorylate RIP2 during incubation but with different kinetics and pattern. BCG- and

M.tb-mediated phosphorylation of RIP2 peaks at 30 min and 1 h, respectively, and the

phosphorylation decreased at 3 h during M.tb incubation (Figure 2.2C). M.tb and BCG

infection also leads to phosphorylation of serine 32 of IkBα (Figure 2.2D) which is

known to result in its degradation and subsequent activation of NF-B. We found that

phosphorylation occurs as early as 15 min for both mycobacteria and continues through 2

h for BCG but not M.tb. In addition, there is a decrease at 30 min which can be explained

by the degradation of IκBα and recruitment of newly re-synthesized pools of IκBα from

the cytoplasm, as previously described (264).

Figure 2.2: The NOD2 ligand MDP enhances M.tb- and BCG-mediated cytokine production in human macrophages, a process involving activation of RIP2 and IkBα.

MDM monolayers in a 24 well tissue culture plates were pre-incubated with or without MDP (5μg/mL) for 90 min and then incubated with M.tb or BCG (MOI 5:1) for 2 h, washed 3 times and repleted with MDP (5μg/mL) in RPMI with 1% autologous serum for 24 h. Supernatants were collected and the amounts of TNF-α (A) and IL-1β (B) were measured by ELISA. Shown is the mean ± SD of a representative experiment performed in triplicate (n = 5, * P < 0.05). In parallel, MDMs were incubated with M.tb or BCG (MOI 5:1) for different time points and cells were lysed with TN-1 lysis buffer for RIP2 (C) and IkBα (D) phosphorylation, and re-probed with actin as a loading control. Shown is a representative Western blot from three independent experiments.

60

61

2.3.3 siRNA-mediated knockdown of NOD2 activity in human macrophages leads to loss

of its activity

In order to more definitively examine the causal role of NOD2 in controlling the nature

of the inflammatory response and growth of mycobacteria in human macrophages, we

developed an effective and reliable protocol for knockdown of NOD2 in human

macrophages by using NOD2-specific siRNA and Amaxa nucleofector reagent. MDMs

were transfected with scramble siRNA or NOD2-specific siRNA under optimized

conditions, plated in tissue culture plates, and lysed after 72 h and 96 h. The cell lysates

were analyzed for NOD2 protein levels by Western blotting. Results shown in Figure

2.3A demonstrate the loss of NOD2 protein in MDMs transfected with NOD2-specific

siRNA at both time points. These results were confirmed by confocal microscopy

(Figure 2.3B). Cell enumeration of scramble siRNA- and NOD2 siRNA-treated MDMs

evaluated at 72 h post nucleofection by napthol blue black nuclear staining of cells

demonstrated no loss of cells on the monolayer in either experimental group (Figure

2.3C). There was also no difference in monolayer cell viability as seen by trypan blue

exclusion staining (95% viable in both groups, data not shown). An assay was performed

in order to verify that NOD2 protein knockdown led to loss of its function. Scramble

siRNA and NOD2 siRNA transfected macrophages were treated with MDP and the cell

free culture supernatants were analyzed for TNF-α production (Figure 2.3D). NOD2

knockdown led to a significant reduction in TNF-α production in response to MDP. We

also tested the specificity of NOD2 knockdown by stimulating the NOD2 siRNA

transfected MDMs with the NOD1 ligand H-Ala-D-γ-Glu-diaminopimelic acid (iE-DAP)

and measured TNF-α levels in the supernatant. There was no significant change in

NOD1 mediated TNF-α production between scramble siRNA and NOD2 siRNA

transfected cells providing evidence that NOD2 knockdown by siRNA did not affect the

activity of NOD1, another NLR family member (Figure 2.3D). We also found that

expression of the effector proteins TLR2, Syk and total RIP2 is unaffected in scramble

siRNA and NOD2 siRNA transfected MDMs during M.tb and BCG infection (Figure

2.3E).

Figure 2.3: Loss of NOD2 activity in human macrophages by siRNA knockdown.

Continued on next page

62

MDMs were transfected with either 368nM of scramble siRNA or NOD2 siRNA using the Amaxa nucleofector. (A) The transfected macrophages were plated in RPMI containing 20% autologous serum and incubated for the indicated time points (72 h and 96 h). At each time point macrophages were lysed and NOD2 or the loading control actin protein levels were examined by Western blot (representative blot from n = 3). (B) NOD2 or scramble siRNA trasfected MDM monolayers on coverslips were incubated for 72 h, washed, fixed, permeabilized, and stained using either NOD2 antibody or isotype control antibody followed by Alexa flour 488 conjugated secondary antibody. Cells were then examined by confocal microscopy. Shown is a representative micrograph from the Z series (63x mag for isotype; 126x mag for scramble/NOD2 siRNA, n = 3). (C) Cell numbers were assessed 72 h post transfection to ensure equal numbers of cells between scramble siRNA and NOD2 siRNA transfected cells. Nuclei in cell lysates were counted on a hemacytometer after nuclei were stained with Napthol blue black. Shown are cumulative data obtained from three independent experiments each performed in triplicate (mean ± SEM) (D) Scramble siRNA or NOD2 siRNA transfected MDM monolayers in culture for 72 h were treated with MDP (5μg/mL) or DAP (10μg/mL) for 24 h. Cell supernatants were analyzed for TNF-α by ELISA. Shown is a representative experiment performed in triplicate (mean ± SD, n =2, ** P < 0.005; N.S. P = 0.1519). (E) Scramble siRNA or NOD2 siRNA transfected MDM monolayers were incubated with M.tb or BCG (MOI 5:1) for the indicated time points (15 min and 2 h) and cells were lysed with TN-1 buffer. TLR2, Syk, and RIP2 were analyzed by Western blot. Shown is a representative experiment of two independent experiments.

2.3.4 NOD2 regulates the macrophage TNF- and IL-1β cytokine response to M.tb and

BCG

We next investigated whether NOD2 regulates the production of pro-inflammatory

cytokines during mycobacterial infection which can control the host defense response

through autocrine and paracrine functions (265). Results shown in Figure 2 had already

indicated a potential link between NOD2 and the cytokine response to mycobacteria.

Scramble and NOD2 siRNA transfected MDM monolayers were incubated for 72 h to

achieve the maximum NOD2 knockdown and subsequently incubated with M.tb or BCG

63

64

(MOI 5:1) for 24 h. The cell free culture supernatants were analyzed for cytokine levels

by ELISA. We found that NOD2 deficiency leads to a significant decrease in the

production of TNF-α and IL-1β in response to M.tb and BCG (Figure 2.4 A and B). Next,

we examined RIP2 activation in NOD2 knockdown MDMs in response to M.tb and BCG

infection. NOD2 deficient MDM monolayers were incubated with M.tb or BCG for short

time periods and cell lysates were analyzed for RIP2 phosphorylation by Western blot.

Our results show that NOD2 siRNA transfected MDMs have a partial but reproducible

decrease in RIP2 phosphorylation (Figure 2.4C: upper panel, Western blot; lower panel,

band densitometry), particularly at 2 h, compared to scramble siRNA transfected MDMs

in the case of both M.tb and BCG. The partial reduction of RIP2 phosphorylation may be

explained by the fact that the mycobacteria also produce NOD1 agonists that activate

RIP2 independently of NOD2. The smaller reduction of RIP2 phosphorylation in NOD2

knockdown cells incubated with M.tb compared to BCG may relate to the amount of the

NOD1 agonist DAP in the former (266).

Next we elected to study whether M.tb and BCG regulate IL-1β at the transcriptional

level or through caspase-1, the enzyme necessary for cleavage of pro IL-1β to the

secreted form. During BCG and M.tb infection, NOD2 knockdown macrophages

demonstrate either no effect or an increase in pro IL-1β transcription (Figure 2.4D). On

the other hand, since NOD2 knockdown leads to a significant reduction in active

(cleaved) caspase-1, NOD2 does appear to play a role in caspase-1 activation during M.tb

infection (Figure 2.4E). Similar results were found for caspase-1 using TCA

precipitation of infected supernatants (data not shown). Interestingly, there was a lack of

caspase-1 activation during BCG infection. However, this represents the results from

only one experiment and thus further conclusions cannot be reached.

Figure 2.4: NOD2 controls M.tb- and BCG-mediated cytokine production and RIP2 activation in human macrophages.

Continued on next page

65

Scramble or NOD2 siRNA transfected MDM monolayers were incubated with M.tb or BCG (MOI 5:1) for 2 h, washed 3 times with warm RPMI, and repleted in 1% autologous serum for 24 h. Cell culture supernatants were analyzed for TNF-α (A) and IL-1β (B) production by ELISA. The graphs shown are from a representative experiment of n = 3. *P < 0.05 or **P < 0.005 (C) Scramble or NOD2 transfected MDM monolayers were incubated with M.tb or BCG (MOI 5:1) for the indicated time points (15 min and 2 h) and cells were lysed with TN-1 buffer. NOD2-dependent RIP2 phosphorylation was analyzed by Western blot of cell lysates from scramble and NOD2 deficient MDMs (upper panel). Shown is a representative experiment of three independent experiments. Densitometry of the band intensities in this experiment is shown as a bar graph (lower panel). (D) IL-1β steady-state mRNA levels were measured during BCG and M.tb infection by real time PCR. Data were normalized to the β actin gene and RCN was determined. Shown is a representative experiment from two independent experiments each performed in triplicate. (E) Macrophage cell lysates were analyzed for caspase-1 protein expression by Western blot using caspase-1 antibody. The same membrane was re-probed with actin to verify equal loading of lysate proteins. Shown is a representative Western blot, and the bar graph below shows the densitometric analysis during M.tb infection (M.tb n = 2; BCG n = 1).

2.3.5 Expression of NOD2 in human AMs and its role in regulating the production of

TNF- and IL-1β in response to M.tb

A recent report showed the presence of NOD2 transcripts in human alveolar lavage cells,

but to date there is no evidence for the presence of NOD2 protein in AMs (220).

Therefore we examined whether NOD2 protein is expressed in human AMs. Cell lysates

were prepared from MDMs and AMs, the latter were isolated from healthy donors and

both cell types were analyzed for NOD2 expression by Western blot. Results shown in

Figure 2.5A demonstrate that AMs express a detectable amount of NOD2 protein but the

level is less than that in an equivalent number of MDMs. To further examine the

potential involvement of NOD2 in the production of inflammatory cytokines in response

66

to M.tb, we incubated AMs with or without MDP and then with M.tb (MOI 1:1) for 24 h.

The cell culture supernatants were analyzed for TNF-α and IL-1β production. We found

that MDP or M.tb treatment induces the production of TNF-α whereas only M.tb induces

the production of IL-1β, and that the levels of both cytokines were further enhanced by

the addition of both MDP and M.tb (Figure 2.5 B and C). The magnitude of the responses

was relatively low when compared with the response seen with MDMs (Figure 2.2).

Together, these results provide evidence that AMs express functional NOD2 protein in

the lung microenvironment, albeit with lower activity than in MDMs.

Figure 2.5: M.tb-mediated cytokine production in human AMs treated with MDP.

(A) The NOD2 expression level in healthy donor AMs was compared with day 5 MDMs using protein-matched cell lysates by Western blot with a NOD2-specific antibody and actin as a loading control. Shown is a representative blot of two independent experiments. AMs were pre-incubated with MDP (5μg/mL) or media alone for 90 min and then incubated with M.tb (MOI 1:1) for 2 h, washed 3 times and repleted with MDP in 1% autologous serum for 24 h. Culture supernatants were measured for TNF-α (B) and IL-1β (C) production by ELISA. Shown is a representative experiment of the mean ± SD of triplicate wells in each test group (n = 2 * P < 0.05).

67

68

2.3.6 NOD2 controls the intracellular growth of M.tb and BCG in human macrophages

We next focused our studies on determining the role of NOD2 in controlling the growth

of mycobacteria in human macrophages. MDMs were transfected with scramble or

NOD2 siRNA and after 72 h of transfection, macrophages were incubated with M.tb or

BCG (MOI 5:1) for 24 h. Experiments were performed for 24 h due to the fact that

NOD2 protein levels in macrophages began to reappear after 120 h of knockdown.

Intracellular growth of M.tb and BCG was determined by a colony forming unit (CFU)

assay (262). The results in Figure 2.6A show that silencing of NOD2 in macrophages

significantly enhances the growth of M.tb (84.6 ± 8.0%, n = 3). A similar result was

observed for BCG growth in Figure 2.6B (108.4 ± 30.4%, n = 5). Western blotting

confirmed NOD2 knockdown in each experiment (Figure 2.6C). We next determined

whether the increase in mycobacterial growth at 24 h was the result of increased cell

association (both bacterial attachment and phagocytosis). We compared the association of

bacteria with scramble and NOD2 siRNA transfected MDMs by phase contrast and

fluorescence microscopy after 2 h infection. Our results show that bacterial association is

not altered by NOD2 siRNA transfection (Figure 2.6D and E). Thus, these results

indicate that NOD2 plays a role in regulating the early host defense response during

infection of both M.tb and BCG.

Figure 2.6: NOD2 controls the intracellular growth of M.tb and BCG in human macrophages.

(A) Scramble and NOD2 siRNA transfected MDMs in monolayer culture for 72 h were incubated with M.tb or BCG (MOI 5:1) for 2 h, washed 3 times with warm RPMI, repleted in 1% autologous serum for 24 h and lysed to analyze CFUs for (A) M.tb growth and (B) BCG growth. Shown is a representative experiment (mean ± SD of triplicate wells in each test group) for each bacterium (for M.tb n = 3 and for BCG n =5; *P < 0.05). (C) Western blot in the experiment shown confirming NOD2 knockdown. Mycobacterial association with macrophages was assessed by phase contrast and fluorescent microscopy. Scramble and NOD2 siRNA transfected MDMs were plated on coverslips for 72 h and then incubated with M.tb (D) or BCG (E) for 2 h (MOI 10:1). Cells were then washed, fixed in 10% formalin, mycobacteria were stained with auramine-rhodamine and enumerated as bacteria per macrophage. Cumulative data from 3 independent experiments, each performed in triplicate, are shown as fold change relative to the scramble siRNA condition (mean ± SEM).

69

70

2.3.7 Effects of NOD2 on M.tb and BCG growth in mouse macrophages and on the

cytokine response to infection

Previous studies have reported that NOD2 is not critical for controlling the growth of

M.tb and BCG in mouse macrophages (135,179). Based on our results in human

macrophages, we elected to examine the host response to these mycobacteria in mouse

macrophage experiments. We reproduced the published finding that NOD2 does not

control the growth of M.tb in mouse macrophages using WT and NOD2-/- BMMs (Figure

2.7A). In contrast to the published work with BCG that analyzed growth in naive murine

AMs at 48 h (135), our data show that NOD2 is important in controlling the growth of

BCG at the later time points of 72 h and 168 h in WT and NOD2-/- BMMs (Figure 2.7B).

We also examined the cell association of bacteria with WT and NOD2-/- BMMs by

microscopy and found no difference with M.tb and only a small difference with BCG

(Figure 2.7C and D). Finally, consistent with the literature, we found a reduction in the

pro-inflammatory cytokine IL-1β in response to M.tb and BCG at 24 h, 72 h, and 168 h in

NOD2-/- BMMs (Figure 2.7E).

Figure 2.7: Effect of NOD2 on M.tb and BCG growth and cytokine response in mouse WT and NOD2-/- BMMs.

Continued on next page

71

BMMs from WT and NOD2-/- mice were plated in 24 well tissue culture plates and incubated with M.tb (1:1) or BCG (2.5:1) for 2 h, washed 3 times, and immediately lysed or incubated for different time points (2 h, 24 h, 72 h and 168 h) prior to lysis for assessment of CFUs for M.tb (A) and BCG (B). Shown is a representative experiment for each bacterium (mean ± variance of duplicate wells in each test group, n = 2). Cell association of M.tb (C) and BCG (D) with WT and NOD2-/- BMMs was analyzed at 2 h post-infection by microscopy. Cumulative data from 3 independent experiments, each performed in triplicate, are shown as fold change relative to WT condition (mean ± SEM, *P<0.05). The cytokine response IL-1β (E) was analyzed by ELISA in culture supernatants harvested from M.tb and BCG infected WT and NOD2-

/- BMMs. The data shown are from a representative experiment (mean ± SD, n = 2, with four test results in each group * P < 0.05; ** P < 0.005, ***P < 0.0005).

2.4 Discussion

Macrophages play a critical role in the recognition and removal of pathogens through

various PRRs, including intracellular receptors such as NLR family CARD-containing

domain 4 (NLRC4), Baculoviral IAP repeat-containing protein 1 (NAIP), NOD1, and

NOD2. Studies on host recognition of pathogens by human macrophages have

demonstrated the importance of transmembrane receptors; however the field is relatively

lacking in defining the role of intracellular receptors in these cells. NOD2 is an

intracellular PRR that senses bacteria through its LRR domain and induces a pro-

inflammatory response. While this inflammatory response can activate the macrophage,

it also allows for the recruitment and activation of other immune cells to the site of

infection. NOD2 has been shown to serve as a microbial sensor and activator of the pro-

inflammatory response to several pathogens, including M.tb (126,127,135,179,242-244).

In the case of H. pylori, it is thought that the bacterium is able to use a secretion system

to secrete peptidoglycan fragments that are then recognized by NOD2 in the cytosol (93).

72

73

Since M.tb is an intracellular pathogen, and known to secrete virulence factors, it is

necessary to understand the role of intracellular receptors and immune functions in

human macrophages with regards to this pathogen. Demonstrating a role for NOD2 can

lead to the development of new methods for targeting this host defense molecule/pathway

for therapy in the case of M.tb.

The presence of NOD2 mRNA has only been reported in monocytes (76). Thus, we

initiated studies of NOD2 mRNA transcript levels as monocytes differentiate into

macrophages using an established in vitro model. Our results indicate that NOD2 steady

state mRNA transcripts increase as monocytes differentiate into macrophages,

specifically on day three of differentiation. This result is similar to our previously

published data on a membrane PRR that recognizes M.tb, the MR, whose mRNA

transcript levels are also enhanced in macrophages when compared to monocytes. On the

other hand, TLR2 and TLR4 expression follows a different pattern on these cells (54).

Since mRNA levels do not always correlate with protein expression, we elected to

explore NOD2 protein expression in human macrophages compared to monocytes.

Studies in the field have been limited to assaying for mRNA levels in primary human

cells; whereas protein expression has been studied in cell lines by flag tagging NOD2

(114,267). Thus, a major contribution to the field by the current work is the optimization

of available antibodies for characterization of NOD2 protein expression in primary

human monocytes/macrophages by Western blot and confocal microscopy. Our findings

74

demonstrate that NOD2 protein is expressed in human macrophages and levels increase

during monocyte to macrophage differentiation, suggesting a particularly important role

for NOD2 in human macrophages. Our microscopy results show NOD2 protein in a

punctuate pattern in cells, suggesting the possibility that it resides within organelles or in

complexes present in the cytoplasm.

We also detected NOD2 protein in human AMs from healthy donors. An interesting

finding in our work is that AMs appear to express less of NOD2 protein than MDMs. It

is possible that reduced NOD2 levels in AMs represent another phenotypic marker of

alternatively activated macrophages of which AMs are a prototype and consequently

leads to less effective killing of M.tb on first contact during the innate immune response.

On the other hand, later during TB disease, NOD2 may play an important role in host

defense since NOD2 mRNA levels are increased in patients with TB disease (220).

NOD2 is engaged in response to pathogens which leads to the production of pro-

inflammatory cytokines. In cell lines expressing NOD2, incubation with M.tb enhances

NF-κB activation (86,127). We found that pre-stimulation of human macrophages (both

MDMs and AMs) with the NOD2 ligand MDP followed by M.tb or BCG incubation

caused a significant increase in TNF-α and IL-1β production. Although the addition of

MDP is not the paradigm for human infection, the result provided evidence for the

potential involvement of NOD2 in augmenting the cytokine response to M.tb and BCG.

This result raised the question as to what are the major downstream signaling kinases

75

activated in response to M.tb or BCG. RIP2 is a kinase that autophosphorylates at serine

176, and its kinase activity is necessary for NOD2-mediated pro-inflammatory cytokine

production (94). We found that M.tb and BCG both activate RIP2 by phosphorylation at

serine 176 in a NOD2-dependent manner; however the kinetics and pattern of activation

were different for these two bacteria. RIP2 phosphorylation peaked earlier for BCG (30

min) and continued throughout the time course; whereas RIP2 phosphorylation peaked at

1 h for M.tb and the level of phosphorylation decreased by 3 h. Consistent with this we

observed that BCG continues to activate downstream IκBα [thereby enabling the

transcriptional activity of NF-κB and the resultant up-regulation of expression of

macrophage pro-inflammatory response genes (264)] through 2 h whereas activation of

IκBα decreases for M.tb at this time point. Thus, it appears that M.tb but not BCG can

down-regulate RIP2 and IκBα activation perhaps as a mechanism to dampen the innate

immune response. One possibility for this down-regulation is that M.tb possesses

virulence factors that affect this pathway that are not present in BCG such as those

present in the Region of Difference-1 (RD1) which is responsible for the production of

several virulence-associated proteins in M.tb (191). For example, culture filtrate protein-

10 kDa (CFP-10) and early secreted antigenic target-6 kDa (ESAT-6) in M.tb but not in

BCG have been shown to inhibit lipopolysaccharide-induced NF-kB transactivation by

inhibiting reactive oxygen species production (251). Apart from RIP2, recent work from

our laboratory has identified a PPAR-mediated pathway for M.tb that is not operative for

BCG in human macrophages (56).

76

To more directly assess a causal role for NOD2 in regulating the inflammatory response

to mycobacteria in primary human macrophages, we developed a reliable knockdown

protocol for NOD2 that is specific and does not appreciably affect the viability of the

cells on the monolayer. To date, studies on the role of NOD2 in response to different

pathogens in primary human cells have relied on either expressing the NOD2 allele with

the Crohn’s disease mutation in cell lines (139) or by obtaining cells from individuals

homozygous for the NOD2 mutation, which was shown to render this protein reduced in

activity or non-functional (85). Obtaining cells from Crohn’s disease patients with this

mutation is difficult and the results may not be consistent among patients. Also, since

these patients have a mutation in this protein, their cells may be primed in other ways that

may alter the immunological response. The accomplished knockdown of NOD2 in

human macrophages is an important advance enabling the study of NOD2 more

comprehensively in primary monocytes/macrophages.

Our results are consistent with those found with mouse macrophages with regards to the

role of NOD2 in controlling the inflammatory response to mycobacteria. Gandotra et al

found that TNF-α and IL-12 production is decreased in NOD2 knockout BMMs in

response to M.tb (179) and Divingahi et al found that these cytokines are decreased in

NOD2 knockout naive alveolar mouse macrophages in response to BCG (135). We

observed that both TNF-α and IL-1β production are significantly reduced in NOD2-

deficient human macrophages and found a decrease in RIP2 phosphorylation in response

to M.tb and BCG infection. It is noteworthy that RIP2 activity is only partially reduced

77

rather than abolished by the knockdown of NOD2, which may be due to the activation of

RIP2 by other intracellular receptors. For example, NOD1 is involved in the activation

of RIP2 through the CARD-CARD interaction (94).

The role of NOD2 in IL-1β production during M.tb and BCG infection is of interest. A

recent publication reports that PBMCs from Crohn’s disease patients that are

homozygous for the 3020insC NOD2 mutation exhibit decreased production of IL-1β in

response to M.tb (268). The mechanism for the production of mature IL-1β in response

to M.tb involved the activation of ERK, p38 and RIP2 following recognition by

TLR2/TLR6 and NOD2 in PBMCs; however, the mechanism was not explored in human

macrophages. In PBMCs, NOD2 is reported to have a dual effect on pro IL-1β mRNA

transcription and secretion of mature IL-1β (269). Our preliminary observations propose

that during M.tb infection, NOD2 may play a small role in down-regulating pro IL-1β

transcription, suggesting that NOD2 may play a negative regulatory role (Figure 2.4D).

However, NOD2 is responsible for enhancing caspase-1 activation during M.tb infection,

since we observed a decrease in active caspase-1 expression in NOD2 deficient

macrophages (Figure 2.4E). In our experiments in which we added MDP to un-

transfected macrophages followed by M.tb infection, we observed an additive effect on

TNF-α secretion, but a synergistic effect on IL-1 secretion (Figure. 2.2 and 2.5). We

speculate that the differential response of TNF-α and IL-1β may be explained by NOD2

being relatively more important in the release of mature IL-1β due to a role in processing

78

pro IL-1β through caspase-1. The relationship between NOD2 and caspase-1 has been

reported (97,270).

Previous studies have reported that NOD2 knockout in mouse macrophages has no effect

on controlling M.tb growth (135,179). In contrast, we demonstrate here for the first time

that NOD2 plays a role in controlling intra-macrophage growth of M.tb and BCG in

human macrophages. The difference in the role of NOD2 in controlling mycobacterial

growth is validated by doing a direct side by side comparison of mycobacterial growth in

NOD2-deficient human macrophages and macrophages from NOD2 knockout mice. This

difference can be generally explained by the growing list of differences seen in the innate

immune response of human versus mouse macrophages (225,254,271-275). It is relevant

that the NOD2 knockout mouse does not show any signs of abnormalities in the intestinal

tract, which is inconsistent with the NOD2 mutations leading to Crohn’s disease and

inflammatory bowel disease in humans (276). In addition, differences in promoter

structure between humans and mice indicate divergent mechanisms in NOD2 regulation

(147).

There have been reports over time providing evidence that M.tb can translocate into the

cytosol of macrophages under defined conditions (212,213,215,277) and also that

mycobacterial phagosomes are porous (278). Nonetheless, the weight of the evidence to

date is that mycobacteria replicate in phagosomes in human and mouse macrophages. In

the context of the current work, we speculate that M.tb interacts with NOD2 by either

79

secreting virulence factors from the phagosome into the cytosol via a secretion system(s)

(181,191), or possibly translocates proteins across the phagosome through the action of a

yet to be defined flippase. Alternatively, a NOD2-RIP2 complex has been reported to

signal from the cell membrane along with the need for NOD2 to be present at the

membrane of intestinal epithelial cells for signaling to occur (114,279). These reports

raise the possibility that NOD2 may localize with the phagosome of M.tb leading to

engagement of this cytosolic PRR during residence of the bacterium within the

phagosome.

The mechanism(s) involved in the recognition of NOD2 by M.tb and BCG may differ

since BCG lacks the ESX-1 secretion system and is reported to be unable to stimulate

NOD2 dependent RIP2 polyubiquitination in murine macrophages (181). Nonetheless,

BCG does generate TNF-α in a NOD2-dependent manner [this report and (86)]. Of

relevance, mycobacteria contain the N-acetyl muramic acid hydroxylase, NamH, which

changes the acetyl group on MDP to a glycolyl group, causing it to be a more potent

stimulator of the immune response via NOD2 suggesting that NOD2 may be more tuned

at sensing mycobacterial infections in generating an immune response (86).

In summary, here we report that NOD2 is present and functional in human macrophages,

including AMs. We provide evidence that NOD2 plays a role in controlling the growth

of M.tb and BCG in human macrophages along with regulating the nature of the

inflammatory response. The role for NOD2 differs with respect to growth of M.tb and

80

BCG in human and mouse macrophages. Finally, the assays established and described in

this report now enable a more thorough evaluation of the role of NOD2 in regulating the

innate immune response of human macrophages to mycobacteria and other infectious

agents.

81

CHAPTER 3 : Critical role for NOD2 in Mycobacterium tuberculosis induced iNOS

expression in human macrophages 

3.1 Introduction

Mycobacterium tuberculosis (M.tb) is the causative agent of tuberculosis (TB). TB

infects approximately one-third of the world’s populations, and was responsible for 1.7

million deaths in 2009. The high incidence of death is attributed to the increase in HIV

infections, drug resistant strain development, and length/cost of drug treatment (155). It

is imperative to understand the mechanism that underlies the survival of M.tb in humans

in order to effectively treat patients for this disease as well as create useful vaccines.

An important immune cell that is involved in recognizing and aiding in the clearance of

pathogens is the macrophage. Macrophages use a myriad of defense mechanisms in

order to protect the host including the production of reactive oxygen/nitrogen

intermediates (ROI/RNI), pro-inflammatory cytokines, and proteases, as well as

acidification of the phagosome. The type of response induced during the interaction of

M.tb with the macrophage is dependent on the receptor activated (186). For example,

82

entry through the mannose receptor leads to limited phagosome-lysosome fusion and an

anti-inflammatory response (238).

A group of intracellular receptors, Nucleotide-binding oligomerization domain (NOD)-

like receptors (NLR), have been described to recognize and induce a pro-inflammatory

response to specific portions of the peptidoglycan layer of Gram-negative and Gram-

positive bacteria cell walls (85). NOD2 is a NLR that was originally found to recognize

N-acetyl muramyl dipeptide (MDP) found in Gram- negative and positive bacteria. M.tb

produces N-glycolyl-MDP (GMDP), and this change in functional groups (acetyl to

glycolyl) enhances the NOD2 stimulating activity (86). As discussed in chapter 2, our

laboratory has recently provided evidence that in primary human macrophages NOD2

plays an important role in controlling the inflammatory response as well as the growth of

M.tb (77).

The role of NOD2 during early innate immune responses is of interest to our laboratory,

specifically the role of NOD2 in RNI production. Inducible nitric oxide synthase (iNOS

or NOS2) is responsible for the production of the RNI, nitric oxide (NO). NO is known

to play a role in signaling during inflammation, which suggests a potential role for it in

the NOD2 biochemical pathway (37,280). NOD1 is a close family member of NOD2,

and has been found to be involved in iNOS induction in rat vascular smooth muscle cells

and a murine macrophage cell line (281). Furthermore, interferon gamma (IFNγ)

activated mouse macrophages stimulated with MDP induces NO production (282). It has

83

also been shown that in mouse macrophages MDP is able to induce iNOS expression

(283). Little is known about the relationship of NOD2 and iNOS during M.tb infection.

A study using IFNγ-activated NOD2 deficient mouse macrophages showed a decrease in

NO production in response to M.tb (179). In the mouse model for TB NO plays an

important roles in controlling M.tb growth (284). However, it has been difficult to

duplicate these data using human macrophages, which may be due to differences in the

amount of NO produced by mouse macrophages and human macrophages as well as a

lack in sensitivity of assays (285). There are also differences in upstream pathways of

NO production that exist between human and mouse macrophages. L-arginine is

required for the production of NO in mouse macrophages, but not in human macrophages

(286,287). This is possibly due to the lack of tetrahydrobiopterin in human

macrophages, which is a necessary cofactor in the conversion of L-arginine to NO;

however it has been shown that addition of this cofactor does not induce NO production

(288). Recent studies indicate that, while low in comparison to mouse macrophages,

human monocytes produce NO in response to M.tb (289). Furthermore, clinical studies

indicate that patients with TB have increased iNOS expression in the inflammatory zone

of granulomas (290). More specifically, alveolar macrophages of patients with TB have

enhanced iNOS protein expression (291).

Our recent finding of the ability of NOD2 to control M.tb growth at 24 h is of interest,

and led us to believe that NOD2 is involved in early innate immune mechanisms. We

performed a microarray at this time point using NOD2-deficient human macrophages and

84

found inducible nitric oxide synthase (iNOS) to be down-regulated. Based on reports of

M.tb inducing iNOS and NO using in vitro cell models as well as clinical specimens, the

recent evidence that NOD2 plays a role in NO production during M.tb infection of IFNγ

activated mouse macrophages, and our microarray data, we decided to study the role of

NOD2 in the iNOS pathway during M.tb infection of human macrophages. Here we

show that NOD2 agonists increase the expression of iNOS mRNA and protein expression

as well as NO production. In coordination with previous reports we show that M.tb

induces iNOS expression and NO production, however, we show for the first time in

human macrophages that this phenomena is NOD2-dependent. Together these findings

provide evidence for a novel role for NOD2 and its importance in the macrophage

induction of NO in response to pathogens.

3.2 Materials and Methods

3.2.1 Buffers and reagents

Dulbecco’s PBS with and without Ca2+ and Mg2+ ions, Iscoves, FBS, TRIzol, and RPMI

1640 medium with L-glutamine containing phenol red or phenol red deficient were

purchased from Invitrogen (Invitrogen Life Technologies). RHH medium [RPMI 1640

plus 10 mM HEPES (Invitrogen) plus 0.4% human serum albumin (CSL Behring LLC)]

was used for cell culture experiments. N-Acetylmuramyl-L-alanyl-D-isoglutamine

hydrate (MDP; Sigma-Aldrich), N-Glycolyl-muramyldipeptide (GMDP; Invivogen), and

Interferon gamma (IFNγ; BD Biosciences) were used for assays involving iNOS

expression and NO production. 4-amino-5-methylamino-2’,7’-difluorofluorescein

diacetate (DAF-FM diacetate) was purchased from invitrogen for NO production assays.

7H11 agar was prepared with Bacto Middlelebrook 7H11 agar, oleic acid-albumin-

dextrose-catalase enrichment medium, and glycerol (Difco Laboratories).

3.2.2 Antibodies

Unconjugated rabbit anti-human NOD2 was purchased from ProSci and optimization for

human macrophages can be found in the previous report (77). actin, goat anti-rabbit

IgG-HRP and donkey anti-goat IgG-HRP antibodies were purchased from Santa Cruz

Biotechnology. Antibodies specific for iNOS were purchased from Cell Signaling

Technology. The aforementioned antibodies were used for Western blot.

3.2.3 Bacterial strains and culture

Lyophilized M.tb H37Rv (ATCC #25618) and BCG (ATCC #35734) was obtained from

the American Tissue Culture Collection, reconstituted, and used as described (172). The

concentration of bacteria (1–2 x 108 bacteria /mL) and the degree of clumping ( 10%)

were determined by counting in a Petroff-Hausser chamber. Bacteria prepared in this

fashion are 90% viable by CFU assay.

85

86

3.2.4 Isolation and culture of human macrophages

Blood obtained from healthy, Purified Protein Derivative negative human volunteers

using an approved protocol by The Ohio State University Institutional Review Board

(IRB) was processed as described (292). Briefly, PBMCs were isolated from heparinized

blood on a Ficoll-Paque cushion (Amersham Biosciences/GE Healthcare) and cultured in

Teflon wells (Savillex) for 1 (monocytes) through 5 (MDMs) days in the presence of

RPMI 1640 medium containing 20% autologous serum (2.0 x106 PBMCs/mL) at

37ºC/5% CO2 (258).

3.2.5 RNA isolation

Day 5 PBMCs in Teflon wells were harvested and mononuclear phagocytes adhered to a

12-well tissue culture plate with 10% autologous serum (4 x 106 PBMCs/mL). After

washing away lymphocytes, the MDMs (4 x 105 cells) were lysed in TRIzol and total

RNA was isolated by using the Qiagen RNAeasy column method. For microarray studies

RNA was not processed through the Qiagen RNAeasy column. The Experion (Bio-

Rad) was used to determine the quality and quantity of all RNA samples.

3.2.6 Gene expression studies by qRT-PCR

RNA (100ng) was reverse transcribed to cDNA by reverse transcriptase enzyme,

(SuperScript II, Invitrogen) and qRT-PCR was performed using a human NOS2 TaqMan

gene expression kit (Applied Biosystems). NOS2 amplification was normalized to β

87

actin as a housekeeping gene (∆Ct). Relative copy number (RCN) and fold change were

determined. RCN was calculated as follows: RCN = E-∆Ct x 100, where E is the

efficiency (2 = 100% efficiency) and Ct = Ct (target) – Ct (reference) (260). Fold change

was calculated from RCN compared to resting group for NOS2 mRNA experiments.

Triplicate samples were analyzed in duplicate wells in each experiment.

3.2.7 Transfection of MDMs

PBMCs were transfected with Accell scramble siRNA or NOD2 siRNA purchased from

Thermo Scientific Dharmacon RNAi technologies that targeted the sequence

CUUUAGGAUGUACAGUUA (368nM) using the Amaxa Nucleofector (Lonza Group

Ltd) Transfected cells were used for subsequent experiments, such as the DAF-FM

diacetate assay and Western blotting. See section 2.2.9 for more details of the transfection

protocol. Assessment of the NOD2 siRNA specificity and effects on monolayer

density/viability were performed and can be found in chapter 2 (77).

3.2.8 Microarray analysis

Scramble and NOD2 siRNA-transfected cells were incubated with M.tb for 24 h in 1%

autologous serum. Cells were then lysed with TRIzol and RNA was isolated as described

above. Subsequent labeling and hybridization to Affymetrix hgu133plus2 chips was

performed at The Ohio State University (OSU) Comprehensive Cancer Center microarray

facility. Analysis was performed at the OSU Comprehnsive Cancer Center Biomedical

88

Informatics core laboratory. The Venn diagram was generated using Bioinformatics

Organization’s Gene List Venn Diagram program

(http://www.bioinformatics.org/gvenn/).

3.2.9 Macrophage stimulation / infection, cell lysis, and Western blotting

Day 12 MDMs or day 5 MDMs transfected with scramble or NOD2 siRNA (2.0 x 105)

for 72 h were incubated with M.tb or BCG (MOI 5:1; triplicate wells) in RHH for 30 min

at 37°C in 5% CO2 on a platform shaker for equal dispersion of bacteria followed by an

additional incubation for 90 min without shaking. The cells were then washed three times

with warm RPMI, repleted with RPMI containing 1-2% human autologous serum and

incubated for 24 h. For experiments using the NOD2 agonists MDP, GMDP, or the

cytokine IFNγ, 2.0 x 105 day 5 MDMs were adhered to a 24-well plate for 2 h at 37°C in

5% CO2, washed to remove lymphocytes, and then incubated with MDP (5ug/mL),

GMDP (5ug/mL), or IFNγ (10ng/mL), for 24 h in RPMI containing 1 % autologous

serum. The monolayers were washed once with PBS and lysed in TN-1 lysis buffer [50

mM Tris (pH 8.0), 10 mM EDTA, 10 mM Na4PO7, 10 mM NaF, 1% Triton X-100, 125

mM NaCl, 10 mM Na3VO4, 10 µg/ml aprotinin, and 10 µg/ml leupeptin)], incubated on

ice for 10 min and then centrifuged at 17,949 x g for 10 min at 4°C to remove cell debris

(261). For NOD2 knockdown assessment, lysates were collected as above at 72 h or 96 h

post-knockdown. Protein concentrations of the cleared lysates were measured using the

BCA-protein assay kit (Pierce). Proteins were separated by SDS-PAGE and analyzed by

Western blot by probing with the primary and secondary antibody of interest and

89

development using ECL (Amersham Biosciences). The protein band intensities were

measured using the online ImageJ software provided by the NIH. Background intensity

was subtracted from each sample and then normalized to β-actin. Fold change was

determined as follows: (treated sample band intensity) / (untreated sample band

intensity).

3.2.10 DAF-FM diacetate staining

Untransfected MDMs or transfected (scramble or NOD2 siRNA) MDMs (2 x 105) were

adhered to glass coverslips in a 24-well tissue culture plate for 2 h at 37°C and washed to

remove lymphocytes. The cells were treated with GMDP, IFNγ, or M.tb for 24 h in no

phenol red RPMI containing 1% autologous serum. At 24 h cells were washed one time

with warm no phenol red RPMI and repleted with 5 μM/mL DAF-FM diacetate for 30

min at 37°C. Next cells were washed with warm no phenol red RPMI 3 times and left in

the third wash for 15 min at 37°C to allow for deesterification of the dye. Finally, cells

were fixed with 10% formalin (10 min at room temperature), washed with PBS, and

mounted on glass slides. Slides were viewed using an Olympus Flowview 1000 Laser

Scanning Confocal microscope at an excitation/emissions maxima of 495/515 nm. One

hundred and fifty to three hundred consecutive macrophages/test group were counted and

Mean fluorescence intensity (MFI) was calculated using Olympus FluoView FV1000

software.

90

3.2.11 Statistics

A paired one tailed student’s t test was used to analyze the differences between two

groups in figures showing one representative figure of three independent experiments.

An unpaired one tailed student’s t test was used to analyze differences between two

groups in figures showing cumulative data from three independent experiments.

Significance was P < 0.05.

3.3 Results

3.3.1 NOD2 regulates many genes during M.tb infection of human macrophages

As discussed in chapter 2, our laboratory has published that NOD2 plays an important

role in controlling the growth of M.tb in human macrophages at 24 h post infection.

Between initial infection and 24 h there is limited bacterial replication; the bacteria are

becoming acclimated to the phagosome of the macrophages. Thus, enhanced bacterial

viability at 24 h in NOD2-deficient macrophages suggests an alteration in primary host

defense mechanisms of the cell. In order to study this enhanced viability at 24 h, we used

our established method to knock-down NOD2 in human macrophages and performed a

microarray of scramble and NOD2 siRNA-transfected MDMs 24 h post infection. In

total, 1,583 transcripts were found to be significantly different with a fold difference of 2

or more, which is depicted by scatterplot (Figure 3.1A).

In order to quantify the overlap in the number of genes NOD2 regulates under resting

conditions and during M.tb infection, a Venn diagram was created from the 1,583

91

regulated transcripts (Figure 3.1B). Of these, 22% were regulated by NOD2 under

resting conditions, (scramble versus NOD2 siRNA-transfected MDMs) indicating that

NOD2 plays a role in regulating genes involved in homeostasis of macrophages. 54% of

the genes were regulated by NOD2 during M.tb incubation for 24 h (scramble versus

NOD2 siRNA-transfected MDMs incubated with M.tb) and 24% of the genes were

regulated under both conditions (resting and M.tb incubation combined). While it is of

interest that NOD2 regulates genes involved in normal macrophage homeostasis, we

focused on the role of NOD2 during M.tb infection due to the fact that approximately half

of the genes obtained in the microarray were regulated by NOD2 under this condition.

We first analyzed the data by assessing genes regulated in major macrophage signaling

pathways. We found genes affected by NOD2 deficiency during M.tb infection in the

iNOS, JAK/STAT, TLR, autophagy, TGF-β, and apoptosis signaling pathways (Figure

3.1C). While the percentage of genes in each pathway affected in the microarray may

differ depending on the number of genes involved, it is of interest to note that a large

number of genes were identified to be affected in the apoptosis signaling cascade.

When looking further at the fold change of up- or down-regulated genes in each cascade,

we found that iNOS was down-regulated 3.07-fold. We were able to validate that NOD2

deficiency in human macrophages led to a decrease in iNOS mRNA expression during

M.tb infection using qRT-PCR (Figure 3.1D). This was of interest to us because of

previous reports indicating that patients with TB have enhanced iNOS expression in

macrophages when compared to non-TB patients (291).

Figure 3.1: NOD2 regulates gene expression in human macrophages during M.tb infection.

Continued on next page

92

Scramble or NOD2 siRNA-transfected MDM monolayers were incubated with M.tb (MOI 5:1) for 2 h, washed 3 times with warm RPMI, and repleted in 1% autologous serum for 24 h. Cells were lysed with TRIzol to obtain total RNA . An Affymetrix Human Genome U133 plus 2.0 array was performed and processed. (A) Shown is a scatter plot (log base 2 scale) with NOD2 siRNA treated cells incubated with M.tb (y-axis) plotted against scramble siRNA treated cells incubated with M.tb (x-axis). (B) A Venn Diagram showing all significantly different genes that were up- or down-regulated twofold or more, with 380 genes common to both conditions (Resting = Scramble vs NOD2 siRNA treated cells and M.tb = Scramble vs NOD2 siRNA treated cells incubated with M.tb). (C) A pie chart showing the major signaling pathways analyzed, and the proportion of genes in each pathway that are regulated by NOD2 during M.tb infection of human macrophages. (D) Scramble or NOD2 siRNA transfected MDM monolayers were incubated with M.tb (MOI 5:1) for 24 h, and then monolayers were lysed with TRIzol for extraction of total RNA. Quantitative real time RT-PCR was used to determine iNOS mRNA levels. Shown is the mean ± SD of a representative graph of two independent experiments.

3.3.2 NOD2 agonists regulate iNOS expression and NO production in human

macrophages

We next determined whether NOD2 agonists were able to induce iNOS mRNA in human

macrophages by using quantitative real-time PCR (qRT-PCR). MDM monolayers were

incubated with MDP or GMDP for 24 h and then lysed in order to isolate RNA for qRT-

PCR analysis or cell lysates were collected for Western blotting analysis of iNOS.

Results from qRT-PCR provide evidence that MDP, the NOD2 agonist found in Gram

negative and Gram positive bacteria, induces iNOS mRNA expression (Figure 3.2A).

However the agonist for NOD2 generated from the cell wall of mycobacteria, GMDP,

induces iNOS mRNA at a significantly higher level. This is not surprising because it has

93

94

been shown that GMDP is a more potent stimulator of the immune response (86). Next,

we compared the iNOS mRNA levels with the amount of protein expressed during MDM

stimulation with NOD2 agonists by Western blotting. While there were no differences in

the amount of iNOS protein expression between agonists, we did observe that expression

is induced by NOD2 agonists (Figure 3.2B). IFNγ was used as a positive control for the

ability of MDMs to induce iNOS protein expression (294). We next investigated whether

the NOD2 agonist GMDP and IFNγ activation of iNOS is NOD2-dependent. Scramble

and NOD2 siRNA-transfected MDM monolayers were incubated for 72 h in order to

achieve the maximum NOD2 knockdown and then incubated with GMDP or IFNγ for 24

h. At this time cell lysates were collected and analyzed for iNOS expression. We found

that NOD2 deficiency leads to the loss of iNOS expression by GMDP and IFNγ (Figure

3.2C). This result indicates that iNOS expression is NOD2-dependent. Finally, we

sought to confirm whether the iNOS expression we observed in response to these

agonists in MDMs resulted in NO production. It was previously observed that IFNγ

activated mouse macrophages stimulated with MDP led to the production of NO, and this

was NOD2-dependent (179). In our MDM model, we used a membrane permeable, NO-

sensitive fluorescent dye, DAF-FM diacetate, in order to measure the production of NO

(293). Figure 2D shows that exposure of MDMs to GMDP for 24 h led to a 45% increase

in NO production. IFNγ stimulation for 24 h resulted in a 57% increase in NO

production, similar to previously reported data using primary human alveolar

macrophages (294). Two representative images from each test group are shown in figure

3.2E.

Figure 3.2: NOD2 agonists regulate the mRNA and protein expression of iNOS and NO production in human macrophages.

(A) MDM monolayers were treated with MDP (5 μg/mL), GMDP (5 μg/mL) or IFNγ (10ng/mL) for 24 h, then monolayers were lysed with TRIzol for extraction of total RNA. Quantitative real time RT-PCR was used to determine iNOS mRNA levels. Data were normalized to the β actin gene and RCN was determined. Shown are cumulative data from five independent experiments (mean +/- SEM; *P < 0.05, **P < 0.005). (B) Cell lysates were analyzed for iNOS expression or loading control β actin by Western blot at 24 h. Shown is a representative experiment of three independent experiments. (C) Scramble or NOD2 siRNA-transfected MDM monolayers were incubated with GMDP (5 μg/mL) or IFNy (10 ng/mL) in 1% autologous serum for 24 h. Cells were lysed with TN-1 buffer and iNOS or the loading control β actin protein levels were examined by Western blot. Shown is a representative experiment of three independent experiments. (D) Day 5 MDM monolayers on coverslips were treated with GMDP or IFNγ for 24 h, and then cells were incubated with DAF-FM diacetate in RPMI containing no phenol red for 30 min. Cells were washed 3 times with warm RPMI containing no phenol red and then allowed to sit for 15 min in order for

Continued on next page 95

deesterification to occur. Next, cells were fixed and mounted on slides. Cells were then examined by confocal microscopy with excitation/emission maxima of 495/515 nm (Olympus FV1000). Shown are cumulative data of the mean fluorescence intensities for each test group obtained from three independent experiments (mean +/- SEM; **P < 0.005, ***P < 0.0005). (E) Two representative micrographs (40x) are shown of DAF-FM diacetate staining for each test group of three independent experiments performed on duplicate coverslips.

3.3.3 M.tb-mediated iNOS expression is NOD2-dependent

It is known that the attenuated M.tb strain, H37Ra, enhances iNOS mRNA and protein

expression as well as NO production in human alveolar macrophages (295). The virulent

strain, Erdman, was shown to induce NO production in human monocytes and monocytic

cell lines (289). Corroborating these data, it was discovered that patients with TB have

higher levels of iNOS protein expression in alveolar macrophages (295). We next

examined whether the virulent strain of M.tb, H37RV, and the attenuated vaccine strain,

M. bovis BCG (BCG), caused increased iNOS expression in MDMs. Day 12 MDM

monolayers were incubated with M.tb or BCG (MOI 5:1) for 24, 48, 72, and 96 h at

which time cell lysates were analyzed for iNOS expression by Western blotting (Figure

3.3A). We found iNOS protein expression is enhanced at 24 and 48 h for both bacteria

and decreased at 72 h. A notable difference between the virulent and attenuated strain is

seen at 24 h, M.tb increases iNOS expression to a > level than BCG. We next confirmed

that M.tb-mediated induction of iNOS resulted in NO production in MDMs. MDM

monolayers were incubated with M.tb for 24 h and then exposed to DAF-FM diacetate in

order to quantify NO production. Our results show a 71% increase of NO production in

96

97

MDMs infected with M.tb (Figure 3.3B). Two representative images of each test group

are shown in figure 3.3C. The previous observation of a decrease in iNOS expression

upon stimulation of NOD2 deficient MDMs with the component of peptidoglycan that

can be found in M.tb, GMDP, led us to study whether NOD2 is responsible for the M.tb-

induced iNOS expression. MDMs were transfected with scramble or NOD2 siRNA, and

after 72 h of transfection MDMs were incubated with M.tb. At 24 h cells were lysed and

analyzed for iNOS expression by Western blotting (Figure 3.3D). Silencing of NOD2 in

MDMs and incubation with M.tb resulted in decreased iNOS protein expression. Thus,

our results indicate that in human macrophages, NOD2 is an important mediator of iNOS

expression during M.tb infection.

Figure 3.3: M.tb-mediated induction of iNOS expression and NO production is NOD2-dependent in human macrophages.

(A) Day 12 MDMs were incubated with M.tb or BCG (MOI 5:1) for 2 h, washed 3 times and repleted with 2% autologous serum for 24, 48, 72, and 96 h. Cells were lysed with TN-1 buffer at each time point and iNOS or β actin protein levels were examined by Western blot. Shown is a representative experiment of three independent experiments. (B) Day 12 MDM monolayers on coverslips were incubated with M.tb as previously described for 24 h, and then cells were incubated with DAF-FM diacetate as described in the methods, fixed and examined by confocal microscopy with excitation/emission maxima of 495/515 nm (Olympus FV1000). Shown are cumulative data of the mean fluorescence intensities for each test group obtained from three independent experiments (mean +/- SEM; ***P < 0.0005). (C) Two representative micrographs (40x) are shown of DAF-FM diacetate staining for each test group of three independent experiments performed on duplicate coverslips. (D) Scramble and NOD2 siRNA-transfected monolayers were incubated with M.tb (MOI 5:1) for 2 h, washed 3 times and repleted with 1% autologous serum for 24 h. Cells were lysed with TN-1 buffer and iNOS or β actin protein levels were examined by Western blot. Shown is a representative experiment of three independent experiments.

98

99

3.4 Discussion

While NOD2 is one of the most well characterized members of the NLR family, there are

limited reports of the functions of NOD2 in other innate immune responses besides

cytokine production. These recent reports indicate a role for NOD2 in autophagy and

MHC class I and class II presentation (296-298). Our own findings of the growth of M.tb

being controlled at an early time point, led us to question whether other early innate

immune responses were being affected by NOD2 (77). Since NOD2 plays such a major

role in bacterial recognition and clearance, we sought to study other roles of NOD2

through a microarray (299).

Our microarray results show that NOD2 regulates 1,583 transcripts in human

macrophages. While we did not pursue this avenue, our microarray results indicate that

NOD2 is important for normal macrophage homeostasis (independent of M.tb). The role

of NOD2 in homeostatic macrophage functions has not been studied and may be an

important aspect of the receptor to study. Over half of the genes are up- or down-

regulated during M.tb infection of NOD2 deficient macrophages, which suggests that

NOD2 plays a role in many more macrophage responses during infection than previously

reported. Our array analysis revealed that during M.tb infection of human macrophages

NOD2 plays a role in many signaling cascades including: iNOS, JAK/STAT, TLR,

autophagy, TGFβ, and apoptosis. Thus a major contribution to the field by this current

work is the identification of important genes involved in inflammation/immunity affected

by NOD2 deficiency in human macrophages.

100

We chose to study the role of NOD2 in the iNOS signaling pathway due to its relevance

to M.tb infection. In the mouse model, NOS2 is identified as a protective locus against

TB. Mice deficient in the NOS2 gene succumb to infection within 33-45 days compared

to WT mice which survive an average of 160 days. Large, granulomatous lesions were

found in the lung, liver, and spleens at 30 days post infection, while few were found in

WT mice at this time point (284). Similar results of mortality and tissue damage were

found using iNOS inhibitors, aminoguanidine and NG-monomethyl-L-arginine, in the in

vivo mouse model (300). The role of NO during M.tb pathogenesis in humans has been a

controversial topic in the past; however, there is now more evidence in the literature that

supports its impact. Initial reports testing human cells for the production of NO were

negative, due to the low amounts of NO produced by these cells and the usage of

insensitive assays. The discovery of a more sensitive assay using the fluorescent probe

diaminonapthalene (DAN), which detects as little as one pmol of nitrite, allowed

researchers to detect nitrite in human monocytes incubated with M.tb (301). Jagannath et

al reported using the DAN assay that M.tb is able to induce NO production in human

monocytes and this NO is bacteriostatic (289). Whether or not NO is bacteriostatic or

bactericidal is dependent on the specificity of the RNI and the strain of mycobacteria

(302-304). Although the mechanism has not been identified, the ability of M.tb to resist

the effects of RNI is due to a novel gene, noxR1 (305). It is interesting to note that iNOS

deficient mice treated with the same drugs that cure M.tb infected immunocompetent

101

mice are ineffective. This suggests that the effectiveness in vivo of tuberculocidal drugs

may be dependent on iNOS-derived NO (306).

The precise mechanism(s) underlying how NO antagonizes M.tb is unknown; however, it

may include disruption of bacterial DNA, proteins, signaling, or induction of apoptosis.

Since we recently found that NOD2 is an important modulator of the immune response to

M.tb and is necessary to control its growth in primary human macrophages, we decided to

study whether NOD2 plays a role in the expression of the necessary mediator of NO

production, iNOS. Our microarray results indicated that iNOS mRNA expression is

down-regulated during M.tb infection of NOD2 deficient human macrophages. We

verified this result using quantitative RT-PCR. Recently, it was found in mouse

macrophages that MDP induces iNOS protein expression in a RIP2-dependent manner

(283). Thus, we initiated our studies by investigating whether or not NOD2 agonists

induced iNOS mRNA expression in human macrophages. Our results indicate that the

agonist found in Gram negative and positive bacteria, MDP, is a relatively poor inducer

of iNOS mRNA at 24 h. However, the agonist produced by M.tb, GMDP, induces robust

mRNA expression of iNOS. This is not surprising because comparative studies of MDP

and GMDP indicate that GMDP is a more robust stimulator of the immune response (86).

mRNA expression does not always translate to protein expression in eukaryotic cells, so

we explored the ability of NOD2 agonists to induce iNOS protein expression in human

macrophages. Our findings demonstrate that both NOD2 agonists are effective at

102

inducing iNOS protein expression and this is at the same level as a known inducer of

iNOS in human alveolar macrophages, IFNγ (294). These NOD2 agonists are known to

induce the production of pro-inflammatory cytokines, TNF-α and IL-1β. As discussed in

chapter 2, our laboratory has shown that the M.tb-induced cytokine production is NOD2-

dependent in human macrophages (77). Alveolar macrophages from TB patients produce

TNF-α and IL-1β at a higher level than healthy patients and the usage of an inhibitor of

NO led to a reduction in TNF-α and IL-1β (307). It is known that iNOS expression can

be induced by pro-inflammatory cytokines (308,309). The lack of these cytokines being

produced is a possible reason why iNOS is not expressed when NOD2 is deficient.

In order to directly evaluate the role of NOD2 in regulating iNOS expression, we used a

well established method in our laboratory to knockdown NOD2 in human macrophages

(77). To date the role of NOD2 in iNOS and NO production has relied solely on the use

of primary mouse macrophages or the iNOS knockout mouse model (179,282,283). Our

results show that NOD2 is necessary to induce the protein expression of iNOS in human

macrophages. Since iNOS is a critical enzyme that catalyzes the production of NO, this

suggests that NOD2 plays a role in iNOS-induced NO production. NO is an effective

host-defense molecule against microbial pathogens. NO can damage DNA as well as

interact with accessory protein targets and result in enzymatic inactivation (310). Using a

compound that is non-fluorescent until it binds NO to form a fluorescent benzotrizole, we

found that the GMDP-induced iNOS expression consequently results in NO production.

103

Since GMDP is a portion of the cell wall of M.tb, it is possible that it mediates M.tb-

induced iNOS expression and NO production.

Our results are consistent with previous findings that M.tb enhances iNOS expression.

However, here we compare virulent M.tb to the attenuated vaccine strain, BCG. We

observed that M.tb and BCG induce iNOS at 24 h, but M.tb enhances iNOS expression at

a > level when compared to BCG. However, the levels decrease at 72 h for both strains.

In this study, we were able to confirm that the M.tb-induced iNOS expression results in

NO production using a different technique than previously published (DAF-FM

diacetate). This NO could potentially be bacteriostatic or bactericidal as previously

mentioned. Human alveolar macrophages isolated from patients with fibrosis and

challenged with BCG for 24 h displayed enhanced iNOS expression. This led to

peroxynitrate production and the subsequent killing of BCG (311).

For the first time we show that the induction of iNOS by M.tb in human macrophages is

NOD2-dependent. This cytosolic PRR is an important protein in the iNOS signaling

pathway, which leads to the production of NO. The role of a PRR in the induction of an

inflammatory mediator such as iNOS is not reported in human macrophages. In mouse

macrophages, iNOS expression is mediated by TLR4 in MyD88-dependent and

independent manners (312); however its expression in human macrophages has not been

reported. For example in mouse macrophages bacterial lipoprotein activation of TLR2

leads to NO-dependent killing of M.tb, but in human alveolar macrophages and

104

monocytes, the TLR2 activated antimicrobial pathway was found to be NO-independent

(313). This finding is important because it identifies a cytosolic PRR potentially

involved in the antimicrobial killing effects of NO. NOD2 is thought to be a secondary

line of defense in macrophages to fight bacteria that are able to avoid the initial host

defense mechanisms at the surface, but now these data suggest an even more important

role for this receptor in inducing a pathway that results in bacterial killing. This is of

great importance during M.tb infection, where the bacteria are able to modulate the host

immune response and reside in a unique phagosome. Clinical samples have high levels

of iNOS expression in the inflammatory zone surrounding tuberculous granulomas and

NOD2 could be the mechanism for this expression (290).

Patients with Crohn’s disease have inflamed mucosa. They are homozygous for the

3020insC NOD2 mutation (loss-of-function) and are reported to have enhanced levels of

iNOS expression and NO production in the active mucosa (314). This is somewhat

contradictory to the results in this manuscript, since NOD2 knockdown led to decreased

iNOS expression. However, the enhanced iNOS expression and NO production in

Crohn’s disease patients is likely due to other infiltrating cells such as neutrophils or

epithelial cells found in the area. It has been shown that elicited but not circulating

neutrophils produce larger amounts of NO (315,316). Furthermore, we show here that

NOD2 is important in iNOS expression but we did not study endothelial NOS, which is

expressed in cultured human bronchiolar epithelium and is also responsible for NO

105

production (317). These reasons could explain why patients with Crohn’s disease have

increased iNOS and NO levels.

In summary, we report results of a microarray indicating that NOD2 plays a role in many

signaling pathways important in inflammation/immunity. We confirm published work in

that human macrophages are able to produce NO, dependent on NOD2 agonists as well

as M.tb. Finally, we provide evidence that the cytosolic PRR, NOD2, is a mediator in the

production of iNOS in response to M.tb. These findings uncover a new mechanism for

how NOD2 plays a role during M.tb pathogenesis and point towards GMDP as being

clinically useful for controlling M.tb infection in an NO-dependent manner.

106

CHAPTER 4 : The role of PKCδ in the NOD2 signaling pathway in human macrophages

4.1 Introduction

Inflammation is a process that involves the production of cytokines that are important to

control infections and involves various regulators. The most notable transcription factor

that is involved in pro-inflammatory cytokine production is NF-κB. There are several

signaling cascades responsible for NF-κB activation, including Protein kinase C (PKC)

signaling (318). The PKC family of enzymes consists of several isoforms that are

important kinases in ensuring that other proteins function properly through the

phosphorylation of hydroxyl groups on serine and threonine residues. The PKC

subfamilies of isoforms are classified based on their activation stimulus and are called

classical, atypical, and novel. Classical PKCs require DAG, calcium, and phospholipid

for activation and include the isoforms PKCα, PKCβI, PKCβII, and PKCγ (11). Atypical

PKCs require neither calcium nor DAG for activation. They include the isoforms PKCι,

PKCζ, PK-N1, PK-N2 (13). The novel PKCs are activated by diacylglycerol and do not

require calcium. They include the isoforms PKCδ, PKCθ, PKCε, and PKCη (12). PKCδ

is of interest because it is expressed in many cell types including macrophages and is

107

thought to be important in microbial infections (319,320). During Listeria

monocytogenes infection PKCδ is necessary for the confinement of bacteria to the

phagosome (320). PKCδ is also necessary for the control of Raf and MAPK signaling in

response to the bacterium H. pylori (321). The role of PKCδ in intracellular events

during infection points to the possibility that it may be involved in intracellular signaling

pathways that involve cytosolic PRRs.

A more recently discovered intracellular group of receptors are the Nucleotide-binding

oligomerization domain (NOD)-like receptors (NLR). These receptors recognize specific

muropeptides from the peptidoglycan of Gram-negative and Gram-positive bacteria and

induce a pro-inflammatory cytokine response through activation of NF-κB and MAPK

(85). Of the 23 NLRs, NOD2 is one of the best characterized. It recognizes the

muropeptide, muramyl dipeptide (MDP), which is found in both Gram-negative and

Gram-positive bacteria, as well as a glycolyl-MDP (GMDP) found in mycobacteria (86).

While it is a well characterized NLR, there are still several fundamental questions that

need to be answered regarding the signaling pathways that it is involved in. For example,

it is established that MDP is an agonist of NOD2 and that NOD2 “senses” the

muropeptide, but the underlying biochemistry for this recognition is still not understood.

It is hypothesized that another molecule may be involved in the recognition of MDP

which ultimately binds to NOD2; however it is yet to be discovered (74,322).

108

It is important to study the signaling cascades involving NOD2 in order to have a clearer

understanding of how this PRR functions during bacterial infection. The current model

of NOD2 signaling includes recruitment of a key kinase, receptor interacting protein 2

(RIP2). This kinase has a CARD domain that interacts with the CARD domain of NOD2

by a homotypic interaction and induced proximity signaling (76). The interaction only

occurs upon activation and oligomerization of the NOD2 protein (85,88). The next

interaction that occurs is the binding of TGF activated kinase 1 (TAK1) by the NOD2-

RIP2 complex, which leads to subsequent NF-κB and AP-1 mediated activation and pro-

inflammatory cytokine production (95,96). It was previously thought using cell lines that

RIP2 kinase activity was not necessary for proper NOD2 signaling (107). However a

recent study using RIP2 knockout mice, RIP2 kinase-dead knockin mice, as well as RIP2

deficient primary mouse cells proved that RIP2 kinase activity is necessary for protein

stability and downstream signaling of NOD1 and NOD2 (94). However, the exact role of

RIP2 kinase activity in the activation of signaling molecules is still unknown.

A short isoform of NOD2 has been shown to inhibit the NOD2-RIP2 assembly complex

by interacting with these two proteins and interfering with the oligomerization of NOD2

(98). The protein Erbin, a member of the leucine rich repeat and PDZ domain (LAP)

family, also binds NOD2 through the CARD domain and inhibits NOD2-mediated NF-

κB activation (112,113). Another protein, CARD8, has also been shown to inhibit NOD2

signaling through a physical interaction with NOD2 (323). While these studies give

insight into possible mechanisms of proteins that interfere with NOD2 signaling, they are

109

all done in transfected cell lines and thus provide only indirect evidence for the signaling

pathways that may be occurring in primary cells.

Studies indicate that NOD2 may play a role in PKC signaling. In epithelial cells, NOD2

interacts with dual oxidase 2 (DUOX2) and co-localizes at the plasma membrane.

DUOX2 is involved in NOD2-dependent ROS production (324). PKC activation with

low concentrations of PMA causes DUOX2 to be highly induced (325). A more recent

study using mouse macrophages indicates that MDP stimulates PKCβII phosphorylation

in a RIP2-dependent manner (283). These studies along with the lack of full

understanding of NOD2 signaling in primary cells led us to study the role of PKCδ in the

NOD2 signaling cascade. PKCδ has been shown to play an important role in signal

transduction pathways active during infection. We show that NOD2 agonists activate

PKCδ kinase activity and PKCδ-induced cytokine production through a NOD2-

dependent mechanism. We also provide evidence to support that PKCδ is needed for

RIP2 phosphorylation. Together these findings indicate that PKCδ is a newly identified

important mediator in the NOD2 signaling cascade.

4.2 Materials and Methods

4.2.1 Buffers and reagents

Dulbecco’s PBS with and without Ca2+ and Mg2+ ions, and RPMI 1640 medium with L-

glutamine containing phenol red were purchased from Invitrogen (Invitrogen Life

110

Technologies). RHH medium [RPMI 1640 plus 10 mM HEPES (Invitrogen) plus 0.4%

human serum albumin (CSL Behring LLC)] was used for cell culture experiments.

Phorbol myristate acetate (PMA; Sigma-Aldrich), N-Acetylmuramyl-L-alanyl-D-

isoglutamine hydrate (MDP; Sigma-Aldrich), and N-Glycolyl-muramyldipeptide

(GMDP; Invivogen) were used for assays involving PKCδ expression and kinase activity.

4.2.2 Antibodies

Unconjugated rabbit anti-human NOD2 was purchased from ProSci and optimization for

human macrophages can be found in chapter 2 and our recently published work (77).

Total PKCδ, actin, goat anti-rabbit IgG-HRP and donkey anti-goat IgG-HRP antibodies

were purchased from Santa Cruz Biotechnology. Phospho-RIP2, Phospho-PKCδ,

Phospho-PKCθ, and Phospho-PKC/βII antibodies were purchased from Cell Signaling

Technology. The aforementioned antibodies were used for Western blot.

4.2.3 Isolation and culture of human macrophages

Blood obtained from healthy, Purified Protein Derivative negative human volunteers

using an approved protocol by The Ohio State University Institutional Review Board

(IRB) was processed as described (292). Briefly, PBMCs were isolated from heparinized

blood on a Ficoll-Paque cushion (Amersham Biosciences/GE Healthcare) and cultured in

Teflon wells (Savillex) for 1 (monocytes) through 5 (MDMs) days in the presence of

RPMI 1640 medium containing 20% autologous serum (2.0 x106 PBMCs/mL) at

37ºC/5% CO2 (258).

111

4.2.4 Transfection of MDMs

PBMCs were transfected with Accell scramble siRNA, NOD2 siRNA, or PKCδ siRNA

purchased from Thermo Scientific Dharmacon RNAi technologies that targeted for

NOD2 the sequence CUUUAGGAUGUACAGUUA (368nM) and for PKCδ the

sequence GGCUGAGUUCUGGCUGGACTT (250nM) using the Amaxa Nucleofector

(Lonza Group Ltd) Transfected cells were used for subsequent experiments, such as

ELISAs and Western blotting. See section 2.2.9 for more details on the transfection

protocol. Assessment of the NOD2 siRNA specificity and effects on monolayer

density/viability were performed and can be found in chapter 2 and our recently

published work (77). Assessment of the PKCδ siRNA specificity can be found in

previous reports (326). PKCδ siRNA did not affect monolayer density/viability.

4.2.5 Macrophage stimulation and preparation for ELISAs and Western blotting

Day 5 MDMs or day 5 MDMs transfected with scramble, NOD2 siRNA, or PKCδ siRNA

(supernatants, 2.0 x 105; lysates 4.0 x 105 ) for 72 h were incubated with NOD2 agonists

MDP (5ug/mL), GMDP (5ug/mL), or the phorbol diester PMA (0.123ug/mL in RHH for

various time points between 0.5 and 4 h at 37°C in 5% CO2. For day 5 MDM ELISA

experiments using the NOD2 agonist MDP for pre-stimulation, 2.0 x 105 day 5 MDMs

were adhered to a 24-well plate for 2 h at 37°C in 5% CO2, washed to remove

lymphocytes, and then incubated with MDP (5ug/mL) for 90 min. Cells were then

112

washed and incubated with PMA for 1 h, 2 h, 4 h, 6 h, or 24 h. Cell culture supernatants

were harvested and used for ELISAs (see below). For Western blotting experiments, the

monolayers were washed once with PBS at the time points indicated and lysed in TN-1

lysis buffer [50 mM Tris (pH 8.0), 10 mM EDTA, 10 mM Na4PO7, 10 mM NaF,

1% Triton X-100, 125 mM NaCl, 10 mM Na3VO4, 10 µg/ml aprotinin, and 10 µg/ml

leupeptin)], incubated on ice for 10 min and then centrifuged at 17,949 x g for 10 min at

4°C to remove cell debris (261). For NOD2 knockdown assessment, lysates were

collected as above at 72 h post-knockdown. Protein concentrations of the cleared lysates

were measured using the BCA-protein assay kit (Pierce). Proteins were separated by

SDS-PAGE and analyzed by Western blot by probing with the primary and secondary

antibody of interest and development using ECL (Amersham Biosciences). The protein

band intensities were measured using the online ImageJ software provided by the NIH.

Background intensity was subtracted from each sample and then normalized to β-actin.

Fold change was determined as follows: (treated sample band intensity) /

(untreated sample band intensity).

4.2.6 Kinase assays

MDMs were lysed in TN-1 lysis buffer and 250 μg of cell lysates were

immunoprecipitated overnight at 4°C, with 200 ng of anti-PKCδ or PKC (SC-937) or

IgG isotype control (SC-2027) antibodies followed by 1 h incubation with 25 μl of

protein G-agarose beads (Amersham Biosciences). Immunoprecipitates were rinsed four

times with 700 ul of TN-1 buffer followed by one rinse with PKCδ-kinase buffer (25 mM

HEPES pH 7.4, 10 mM MnCl2, 1 mM MgCl2, 1 mM DTT, 0.1 mM PMSF) and

113

subjected to in vitro kinase assays by incubating the immunoprecipitates for 1 h at 37°C

in the presence of 20 μl of kinase assay buffer containing 2 μCi of [γ-32P]-ATP (Perkin

Elmer, Boston, MA), 0.5 mM ATP, 200 μg/ml phosphatidylserine, 20 μg/ml

diacylglycerol, and 1 μg histone 2B (H2B) (Roche Applied Science, Mannheim,

Germany), the latter as exogenous substrate.

4.2.7 ELISA

For MDM experiments, cell free culture supernatants were collected at 1 h, 2 h, 4h, 6h,

and 24 h and used to measure TNF-α by ELISA (R&D Systems).

4.2.8 Statistics

A paired one tailed student’s t test was used to analyze the differences between two

groups in figures showing one representative figure of three independent experiments.

An unpaired one tailed student’s t test was used to analyze differences between two

groups in figures showing cumulative data from three independent experiments.

Significance was P < 0.05.

114

4.3 Results

4.3.1 Activation of NOD2 enhances the PMA-mediated pro-inflammatory cytokine

response through its signaling cascade intermediates

We began our studies by investigating whether activation of NOD2 leads to an increase

in the phorbol-myristate acetate (PMA)-mediated pro-inflammatory cytokine response in

macrophages. PMA is a phorbol ester that activates various PKC isoforms by mimicking

the structure of diacylglycerol (DAG). Incubation of macrophages with MDP, a NOD2

agonist, is known to induce the production of pro-inflammatory cytokines such as TNF-α

(85). MDMs were pre-incubated with the NOD2 agonist MDP for 90 min followed by

incubation with PMA. Cell free culture supernatants were analyzed for TNF-α secretion

at 1 h, 2 h, and 4 h. The data show a synergistic effect on cytokine production in cells

pre-treated with MDP compared to cells incubated with PMA alone (Figure 4.1A). Since

our agonist studies indicated a potential link between NOD2 and the cytokine response to

PMA, we next wanted to explore if the PMA-induced TNF-α production was dependent

on NOD2. Scramble and NOD2 siRNA transfected MDM monolayers were incubated

for 72 h to achieve the maximum NOD2 knockdown and subsequently incubated with

PMA for 3 h. The cell free culture supernatants were analyzed for cytokine levels by

ELISA. We found that NOD2 deficiency leads to a significant decrease in the production

of TNF-α in response to PMA (Figure 4.1B). Similar results were found at 6 h and 24 h

(data not shown). Thus, our results indicate that pre-engagement and hence activation of

NOD2 by its agonist significantly enhances PMA-mediated pro-inflammatory cytokine

115

production. Furthermore, we show that PMA-mediated cytokine production is NOD2-

dependent.

The role of NOD2 in PMA-mediated TNF-α production led us to study the downstream

signaling molecules during PMA stimulation of macrophages. RIP2 has been identified

as a signaling partner for NOD2 that is autophosphorylated upon interaction with the

CARD domain of NOD2 which leads to NF-κB-mediated pro-inflammatory cytokine

production (94). MDMs were incubated with PMA or MDP for 0.5 h, 1.5 h, and 3 h.

Cell lysates were analyzed for phosphorylation of RIP2 by Western blotting of the

phosphorylated protein. The data show that both PMA and MDP phosphorylate RIP2

during incubation but with different intensities. The NOD2 agonist MDP has a distinct

kinetic pattern in the activation of RIP2 peaking at 0.5 h and decreasing over time, while

PMA more robustly stimulates the phosphorylation of RIP2 and does not cause a wane in

expression (Figure 4.1C). In order to understand whether the enhanced expression of

RIP2 phosphorylation was NOD2-dependent or mediated by a different mechanism, we

used siRNA to knockdown NOD2 in macrophages. Scramble and NOD2 siRNA

transfected MDM monolayers were incubated with PMA for 3 h and cell lysates were

analyzed for phosphorylation of RIP2 by Western blotting. We observed that PMA-

mediated phosphorylation of RIP2 in day 5 MDMs was greatly enhanced when compared

to scramble siRNA transfected cells incubated with PMA (Figure 4.1C and D). Even

though the level of PMA-mediated RIP2 phosphorylation of scramble siRNA treated

cells was lower than day 5 MDMs, we did find that NOD2 deficiency led to a decrease in

RIP2 phosphorylation (Figure 4.1D). Taken together, these results provide evidence for

an important role for NOD2 and RIP2 in PMA-mediated signaling pathways such as

those involving PKC.

Figure 4.1: The NOD2 signaling pathway is involved in PMA-stimulated cytokine production.

Continued on next page

116

(A) MDM monolayers in 24 well tissue culture plates were pre-incubated with or without MDP (5 μg/mL) for 90 min and then incubated with PMA (0.123 μg/mL) in RHH for the indicated time points. Supernatants were collected and the amount of TNF-α was measured by ELISA. Shown is the mean ± SD of a representative experiment performed in triplicate (n = 2, * P < 0.05; **P < 0.005). (B) Scramble or NOD2 siRNA-transfected MDM monolayers were incubated with PMA in RHH for 3 h. Supernatants were collected and the amount of TNF-α was measured by ELISA. Shown is the mean ± SD of a representative experiment performed in triplicate (n = 2, ***P < 0.0005). (C) Day 5 MDM monolayers were incubated with PMA or MDP for the indicated time points, and cells were lysed with TN-1 lysis buffer and probed for RIP2 phosphorylation, and then re-probed for actin as a loading control. Shown is a representative Western blot from two independent experiments. (D) Scramble or NOD2 siRNA-transfected MDM monolayers were incubated with PMA in RHH for 3 h and cells were lysed with TN-1 lysis buffer and probed for RIP2 phosphorylation, and then re-probed for actin as a loading control. Shown is a representative Western blot from three independent experiments.

4.3.2 Phosphorylated PKCδ expression and activity is NOD2-dependent

The ability of PMA to induce pro-inflammatory cytokine production and robustly

activate RIP2 in a NOD2-dependent manner led us to formally investigate whether PKC

isoforms were involved in this pathway. PMA is a known activator of several PKC

isoforms including PKCα, PKCβ, PKCδ, and PKCθ (14). We elected to examine

whether the NOD2 agonist, MDP, led to increased phosphorylation of PKC isoforms by

using a PKC phospho-array of isoforms, δ, θ, and, α / βII (Figure 4.2A, B, C). MDP

caused enhanced phosphorylation of PKC δ, PKCθ, and PKCα / βII. PKCθ is not well

characterized in macrophages (327,328). So, we focused our studies on PKCδ and PKCα

/ βII due to their presence in macrophages and roles in signaling during microbial entry of

host cells (320,329). We first compared the NOD2 agonists MDP and GMDP, the latter

117

118

is a NOD2 agonist from the cell wall of mycobacteria (see chapter 2), for their abilities to

induce PKCδ phosphorylation. MDMs were incubated with MDP or GMDP for 0.5 h,

1.5 h, and 3 h. Cell lysates were analyzed for phosphorylation of PKCδ by Western

blotting of the phosphorylated protein. The data show that GMDP leads to

phosphorylation of PKCδ at early time points (0.5 h and 1.5 h) and is lost at the later time

point (3 h), while MDP only slightly increased the phosphorylation of PKCδ at 0.5 h

(Figure 4.2D). Since GMDP was a more potent stimulator of PKCδ phosphorylation, we

proceeded in our studies using this agonist.

Detection of phosphorylation of PKC isoforms by Western blotting is not an accurate

indicator of the magnitude of its kinase activity because in order to be active the kinase

must be phosphorylated at multiple sites. In order to test for PKCδ kinase activity, we

performed a PKCδ kinase assay in immunoprecipitates using GMDP and PMA. The

results indicate that PKCδ activity is non-existent at 0.5 h and peaks at 1.5 h when

stimulated with GMDP (Figure 4.2E). This difference in kinetics of phosphorylation and

kinase activity of PKCδ is not surprising, as it may take time for all of the relevant sites

to be phosphorylated and the kinase to be fully functional. We also elected to study the

kinase activity of PKCα specifically during PMA and GMDP stimulation; however the

results were less consistent when compared to delta. Despite phosphorylation of

PKCα/βII at 1.5 h and 3 h (Figure 4.2C), the kinase assay shows PKCα is active only at

0.5 h (Figure 4.2F). Also, the kinetics of activation does not match RIP2. PKCα is

activated by GMDP at 0.5 h, which is at the same time that RIP2 is phosphorylated,

119

indicating that they would be working simultaneously. While this may occur it suggests

that other mechanisms are involved, making it more complicated to understand the

general role of PKCs in NOD2 signaling. Thus, we elected to move forward with

studying only PKCδ.

The ability of a NOD2 agonist to fully activate PKCδ led us to study whether NOD2 is

important in mediating PKCδ phosphorylation. Scramble and NOD2 siRNA transfected

MDM monolayers were incubated with PMA for 3 h and cell lysates were analyzed for

phosphorylation of PKCδ by Western blotting. Our results show decreased expression of

phosphorylated PKCδ in response to PMA in NOD2 deficient MDMs (Figure 4.2G). We

next wanted to determine whether the NOD2-dependent PKCδ phosphorylation was only

through the PMA-mediated pathway or also through the NOD2-mediated pathway.

Scramble and NOD2 siRNA transfected MDM monolayers were incubated with GMDP

for 3 h and cell lysates were analyzed for phosphorylation of PKCδ by Western blotting.

Our results show a marked decrease in phosphorylated PKCδ in NOD2 deficient MDMs

(Figure 4.2H). These results indicate that NOD2 is important in the classical PKCδ

signaling cascade (PMA) and the NOD2-signaling cascade (GMDP).

Figure 4.2: NOD2 and PKCδ agonists stimulate PKCδ phosphorylation and activity in a NOD2-dependent manner.

Day 5 MDM monolayers were incubated with PMA or MDP in RHH for the indicated time points, and cells were lysed with TN-1 lysis buffer and probed for either PKCδ (A), PKCθ (B), or PKCα/βII (C) phosphorylation, and then re-probed for the total proteins, and/or actin as a loading control. Shown is a representative Western blot from two independent experiments. (D) Day 5 MDM monolayers were incubated with MDP or GMDP in RHH for the indicated time points, and cells were lysed with TN-1 lysis buffer and probed for PKCδ phosphorylation, and then re-probed for total PKCδ. Shown is a representative experiment of two independent experiments. Cell lysates were collected and 250 μg of total protein were immunoprecipitated with

Continued on next page

120

PKCδ (E) or PKCα (F) antibody or isotype control. Immunoprecipitates were subjected to in vitro kinase assays. Shown are representative kinase assays of phosphorylation of recombinant H2B visualized by autoradiography from two independent experiments. Scramble or NOD2 siRNA-transfected MDM monolayers were incubated with PMA (G) or GMDP (H) in RHH for 3 h and cells were lysed with TN-1 lysis buffer and probed for PKCδ phosphorylation, and then re-probed for total PKCδ and actin as a loading control. Shown are representative Western blots from three independent experiments.

4.3.3 RIP2 activation is PKCδ-dependent

Our results indicate that NOD2 is necessary for the activation of PKCδ and this result

points to it being a potential player in the NOD2 signaling cascade. Therefore we next

sought to determine where PKCδ activation fits into the known NOD2 signaling cascade.

RIP2 is felt to be the next identified kinase to interact with NOD2, so we studied whether

PKCδ is necessary for RIP2 phosphorylation. Scramble and PKCδ siRNA transfected

MDM monolayers were incubated with PMA for 3 h and cell lysates were analyzed for

phosphorylation of RIP2 by Western blotting. Our results show decreased amounts of

phosphorylated RIP2 in PKCδ deficient MDMs (Figure 4.3A). We next wanted know if

this PKCδ dependent RIP2 phosphorylation was only through the PMA-induced pathway

or also through NOD2-induced pathway. Scramble and PKCδ siRNA transfected MDM

monolayers were incubated with GMDP for 1.5 h and cell lysates were analyzed for

phosphorylation of RIP2 by Western blotting. Our results show a small but consistent

decrease in amounts of phosphorylated RIP2 in PKCδ deficient MDMs (Figure 4.3B).

121

These results indicate that PKCδ is necessary for RIP2 phosphorylation and suggests that

it may be activated upstream of RIP2 during NOD2-mediated signaling.

Figure 4.3: RIP2 phosphorylation mediated by PMA and GMDP is PKCδ-dependent.

Scramble or PKCδ siRNA-transfected MDM monolayers were incubated with PMA for 3 h (A) or GMDP for 1.5 h (B) in RHH and cells were lysed with TN-1 lysis buffer and probed for RIP2 phosphorylation, and then re-probed for actin as a loading control. Shown are representative Western blots from three independent experiments.

122

123

4.4 Discussion

NOD2 is an important intracellular receptor in controlling bacterial infections (77,299).

There are many reports on the function of this receptor during bacterial infection, but

much less is known about the biochemical mechanisms (i.e. signaling pathways)

underlying its activation and ability to mediate its functions. Increased knowledge in this

area will not only expand our understanding of the role of NOD2 in host defense but may

also provide new targets for host-directed therapies that alter the activity of this receptor.

Currently there are reports of several proteins that potentially interact with NOD2 in in

vitro cell systems, and these reports point towards the ability of NOD2 to interact with

proteins other than its main regulator RIP2 (98,112,113,323).

The PKC enzyme family plays a role in pro-inflammatory cytokine production through

the transcription factor NF-κB (318). They also play a major role in signal transduction

cascades through their ability to phosphorylate hydroxyl groups of serine and threonine

residues. A member of the novel PKC family, PKCδ, is of interest in this chapter.

Diacylglycerol is necessary in order to activate this novel PKC, however, PMA is used

instead of diacylglycerol in vitro to mimic its actions (12). Our interest lies in PKCδ

because it is found in macrophages and recent findings provide evidence that PKCδ is

involved in microbial infections, including those due to Listeria monocytogenes and

Helicobacter pylori (319-321). PKCδ’s established role in signal transduction pathways

(including those that regulate ROS production) along with the discovery of its role during

microbial infection makes it plausible that it may intersect with pathways involving

124

intracellular receptors involved in microbial recognition such as NOD2. Thus, we sought

to study the role of PKCδ in the NOD2 signaling cascade in human macrophages

(283,325).

We initially found that using the NOD2 agonist MDP as an activator of NOD2 followed

by PMA stimulation caused a significant increase in TNF-α production at early time

points. We also found that PMA-induced cytokine production was NOD2-dependent.

PMA is a phorbol ester that mimics DAG, a lipid involved in second messenger

signaling. A major function of DAG is activation of PKCs. Upon activation, PKCs

translocate to the plasma membrane and perform their function of phosphorylating other

proteins that mediate the induction of cytokines. The substrates that PKCs activate are

dependent on the cell type. The recent finding that NOD2 co-localizes with an activator

of PKCs, DUOX2, at the plasma membrane indicates a role for PKCs in NOD2 signaling

(324).

In our model of human macrophages, we found that PMA led to the phosphorylation of

PKCs as expected, but were surprised to find that PMA also led to the phosphorylation of

the downstream signaling partner of NOD2, RIP2. The level of RIP2 phosphorylation

was in fact much higher than the known inducer of RIP2 phosphorylation, MDP. We

also found that this phosphorylation is partially NOD2-dependent. RIP2 is a serine

threonine kinase that is felt to act as a scaffolding protein more than a kinase during NF-

κB activation; however in NOD2 signaling the kinase activity is important for protein

125

stability (330). RIP2 has been shown to associate with TNFR1, TRAF1, TRAF5,

TRAF6, CIAP1, Cflip, TAK1, as well as caspase-1 (330-333). It is interesting to note

that PKCδ interacting protein kinase (DIK), also known as RIP4, shares 45% similarity to

the RIP2 kinase domain (334). RIP4 is a membrane-associated kinase with ankyrin

repeats that is known to bind PKCβ and contributes to activation of NF-κB (334,335).

This activation is dependent on the catalytic activity of RIP4. The ability of RIP2 to

interact with other proteins as well as the capability of a RIP family member to interact

with PKCs suggested to us that RIP2 may interact with PKCδ. Our findings along with

the literature provide evidence for the convergence of the NOD2 and PKC signaling

pathways. This discovery raises new possibilities for the role of NOD2 in a number of

macrophage host defense functions.

The ability of a NOD2 agonist to induce novel PKC isoforms is of interest. The

enhancement of PKCδ activity by NOD2 agonists indicates that NOD2 and PKCδ

signaling may be dependent on one another. Our studies show that PKCδ activity is

NOD2-dependent when stimulated with its classical agonist, PMA, or the NOD2 agonist

GMDP. The discovery that NOD2 is necessary for optimal PKCδ kinase activity is a

novel finding and may be an important mechanism for PKC-dependent cytokine

production. In order to evaluate the location of PKCδ in the NOD2 pathway, we used a

well-established method of knocking down PKCδ (326). We also show that the

downstream kinase of NOD2 is dependent on PKCδ, which is of interest. Our results

indicate that NOD2 is necessary for PKCδ activity and PKCδ is needed for RIP2

126

phosphorylation. We speculate in Figure 4.4 that NOD2, PKCδ, and RIP2 may be in

complex at the membrane that in turn binds TAK1 for proper NF-κB induced cytokine

production to occur. Upon PKC activation, RACK proteins translocate the kinase to the

membrane. It has been previously shown that NOD2 and RIP2 are at the plasma

membrane and signaling can occur here as well (112,114,279). In addition, we found that

PKCα is activated at an earlier time point which implies that this isoform and RIP2 are

activated simultaneously. While we did not pursue this area of interest, the data suggest

other possible signaling pathways may be involved. For example, NOD2 and PKCα may

be involved in earlier events, like p47phox activation for reactive oxygen species

production through the NADPH oxidase.

In summary, here we report a new addition to the NOD2 signaling pathway and provide

new insight into PKC-induced cytokine production. We show that NOD2 is important for

PKCδ activity as well as PMA-induced cytokine production. In addition, we show that

RIP2 phosphorylation is PKCδ-dependent. These findings divulge a new mechanism for

the NOD2 signaling cascade and for PKC-mediated cytokine production (Figure 4.4). It

will be interesting to further define the importance of PKCδ to the NOD2 signaling

pathway. These data provide useful knowledge for targeting NOD2 in host-directed

therapies to control bacterial infections.

Figure 4.4: PKCδ is necessary for optimal NOD2 activity.

GMDP is able to stimulate PKCδ activity. PMA and GMDP induce PKCδ activation and the release of cytokines, both of which rely on NOD2 for maximal effectiveness. PKCδ is necessary for RIP2 phosphorylation during PMA stimulation and to a lesser extent during GMDP stimulation. It is plausible that upon stimulation, NOD2, RIP2, and PKCδ form a complex that is dependent on all members. This complex may then bind TAK1 and allow for the production of pro-inflammatory cytokines.

127

128

CHAPTER 5 : Synthesis

The human immune system is a well defined entity that humans depend on to protect

them against pathogens. Humans are equipped with a complex immune system that

encompasses an adaptive and innate immune response. While both components are

necessary in the clearance and eventual immunity against pathogens, the innate immune

system is the first line of defense against invading organisms. Epithelial cells,

neutrophils, monocytes, and macrophages are cell types that play key roles in protecting

the host. Important components of these cells involved in the recognition and clearance

of microorganisms are PRRs (1). The innate immune system is not always enough to

eliminate pathogens, wherein lies the role of the adaptive immune response. The

adaptive immune response protects the host by remembering previous invaders and

producing an enhanced immune response in order to control the infection (1,336).

Our laboratory is interested in the innate immune response to the pathogen M.tb. More

specifically we study the interaction of M.tb with human macrophages. M.tb infects one-

third of the world’s population and is a great public health concern. It is easily

transmissible by aerosol droplets with a low infectious dose of ten bacilli causing

129

infection (157). One of the first immune cells that M.tb comes in contact with in the

alveolus is the alveolar macrophage. While treatments are available, their expense and

long regimens make it difficult for under developed countries to control the spread of

disease. The recent increase in multi- and extensive-drug resistant strains makes it

difficult for even developed countries to control. It is important to gain more

understanding of the pathogenesis of M.tb in order to be successful in the development of

more effective drug therapies and possible vaccines. This study focuses on one of the

recently described PRRs found in macrophages, NOD2, and its role during M.tb

pathogenesis.

In order to study the interaction of NOD2 with M.tb, our laboratory uses the most

clinically relevant model for M.tb, human MDMs and alveolar macrophages. These two

cell types are critical during M.tb infection and are relevant since this bacterium only

naturally infects humans. While most of our studies use MDMs, this work and previous

work from our laboratory show that they have similarities to alveolar macrophages. For

example, they both express common PRRs such as the MR, TLRs, CRs, and NOD2

(54,77,258). Therefore MDMs are an excellent model to study the first interaction of

M.tb with the host and understand the early innate immune response.

Prior to our NOD2 studies, data were lacking regarding basal mRNA and protein

expression on mononuclear phagocytes during monocyte differentiation into

macrophages. Previous NOD2 studies were limited to cell lines, murine macrophages,

130

and some human monocyte and mixed BAL studies (127,127,135,179-181,220).

Importantly, there are significant differences between the mouse and human promoter of

NOD2 (147). Furthermore, the field was only able to identify NOD2 mRNA expression

in human PBMCs and not protein. Therefore, we began our research by examining

NOD2 expression during monocyte differentiation into macrophages using our MDM

model [Chapter 2]. Interestingly, we observed NOD2 mRNA and protein expression is

low during day one and two of differentiation and significantly increases at day three of

differentiation after which it remains relatively unchanged (Figure 2.1). These data

suggest that NOD2 plays a relatively more important role during an effective innate

immune response in macrophages, while monocytes more readily rely on other PRRs to

control infection. We also observed that NOD2 protein has a punctate rather than diffuse

appearance in the cytosol (Figure 2.1, 2.3). This appearance suggests that NOD2 may be

located either in organelles or as complexes in the cytoplasm. There also appeared to be

less NOD2 protein in human alveolar macrophages than in MDMs (Figure 2.5). Since

NOD2 is an important receptor in controlling early infection, this finding provides one

explanation for why alveolar macrophages are less effective at controlling initial M.tb

infection. However, we were able to show that pre-treatment of alveolar macrophages

with the agonist for NOD2, MDP, followed by infection leads to enhanced pro-

inflammatory cytokine production (Figure 2.5). Even though NOD2 expression is less in

HAMs when compared to MDMs (Figure 2.5), this result implies that this receptor can be

up-regulated, at least to some degree, in alveolar macrophages to potentially control

infection.

131

After determining NOD2 expression during the differentiation into macrophages, we

examined whether MDP pre-treatment followed by M.tb or BCG incubation leads to

enhanced cytokine production. We found that stimulating NOD2 with its agonist

followed by M.tb or BCG leads to increased pro-inflammatory cytokine production.

Additionally, these two bacteria are able to activate downstream mediators (RIP2 and

IκBα) in the NOD2 signaling cascade, albeit with different expression patterns (Figure

2.2). The differences in expression indicate that M.tb, which has a shorter and less robust

activation compared to BCG, is able to dampen the immune response. One possibility is

that M.tb utilizes secreted virulence proteins found in the RD1 to dampen NOD2-

mediated signaling. These proteins are absent in the attenuated BCG strain (192,193).

The enhanced pro-inflammatory cytokine production after NOD2 pre-engagement,

suggests that NOD2 is involved during mycobacterial infection of macrophages.

However, use of an agonist is not the most direct method to study a receptor’s activity.

Thus, we next used siRNA knockdown of NOD2 in human macrophages and observed

decreased RIP2 phosphorylation and pro-inflammatory cytokine production in response

to the two mycobacteria (Figure 2.4). In addition, we found enhanced M.tb and BCG

growth in NOD2 deficient macrophages (Fig. 2.6). This increase in growth is different

from previously published data (135,179), as well as our own data, using NOD2

knockout BMMs incubated with M.tb (Figure 2.7). In contrast, BCG showed enhanced

growth in NOD2 knockout BMMs suggesting that M.tb has acquired a mechanism for

132

evading the immune system of the mouse, while BCG has not. One possibility is that

under normal conditions M.tb utilizes virulence proteins, that BCG is lacking (such as

those in the RD1 in the genome), in order to inhibit proper NOD2 functioning. These

virulence proteins may be more effective in inhibiting NOD2 function in mice because

they target the promoter of NOD2, which has been found to be structurally different in

mice (147).

It is of interest that NOD2 plays a greater role in IL-1β production during M.tb infection,

with slight enhancement of pro-IL-1β transcription and a decrease in caspase-1

expression in NOD2 deficient MDMs (Figure 2.4). As one explanation, NOD2 may play

a relatively more important role in enabling the release of mature IL-1β through caspase-

1 activation in response to M.tb than in increasing TNF-α production, since NOD2 has

been shown to interact with inflammasomes (97,270).

Our studies of NOD2 involvement during M.tb infection lead to unanswered questions,

such as how does NOD2 recognize M.tb and what molecule(s) does NOD2 regulate in

order to affect viability of mycobacteria at 24 h? Several potential mechanisms exist for

NOD2 recognition of M.tb, including the ability of M.tb to escape the phagosome, the

bacteria’s ability to secrete virulence proteins, and the possibility that NOD2 signals from

the phagosomal membrane. Since our data show that NOD2 regulates cytokine

production and M.tb growth at 24 h, we reason that M.tb is not escaping the phagosome,

which has been shown to begin to occur in human dendritic cells at 48 h post infection

133

(215). In addition, our results with the attenuated vaccine strain, BCG, argue that

secreted virulence proteins are not the sole mechanism for NOD2-mediated recognition

of M.tb; however, we cannot rule out the possibility that M.tb down-regulates

downstream signaling molecules important to NOD2 signaling when compared to BCG.

Mycobacteria express a hydroxylase, NamH, which glycolylates MDP (GMDP) causing

it to induce a strong immune response (86). Our NOD2 confocal data support the

possibility that NOD2 localizes at the M.tb phagosome membrane. NOD2 signaling in

epithelial cells occurs at the plasma membrane and NOD2 must be localized there for this

signaling to occur (114,324). Therefore, we hypothesize that NOD2 localizes at the

mycobacterium-containing phagosome and is engaged by bacterial cell wall products

such as GMDP released during replication of the bacteria inside the phagosome. Such a

hypothesis would require that bacterial cell wall products be locally transported to the

cytosolic face of the phagosome, possibly through a secretion or transport system. In

order to confirm or refute this hypothesis, we would need to either isolate

mycobacterium-containing phagosomes to look for the presence of NOD2 by Western

blotting or perform immunofluorescence microscopy studies to see if NOD2 is located on

the M.tb-containing phagosome membrane.

While we observed that NOD2 regulates the viability of M.tb and BCG in human

macrophages, the mechanism(s) behind this role remains unknown. One possible

mechanism that our data support is through the production of pro-inflammatory cytokines

which can activate the macrophages through autocrine and paracrine means. We

134

performed experiments using MDP to “prime” the macrophage to induce early innate

immune responses such as ROIs / RNIs in a NOD2-dependent manner. We found a

consistent decrease in the growth of BCG and M.tb under these conditions, but the results

were variable (10-26% decrease for BCG and 4-29% for M.tb). Thus, the results are

generally consistent with NOD2 being involved in controlling the viability of these

mycobacteria, however it is possible that the experimental conditions were not optimized,

e.g. other parameters such as time and dose should be further tested. In order to

determine the role of NOD2 on other early innate immune responses during M.tb

infection, we would need to examine which processes are involved. Such studies would

broaden the roles of NOD2 in innate immunity beyond those that involve enhancement of

pro-inflammatory cytokine production.

Related to the discussion above, we became interested in what processes NOD2 regulates

during infection [Chapter 3]. We began these studies by performing a microarray of

NOD2 deficient macrophages infected with M.tb for 24 h. Interestingly, we observed

many of the genes in signaling pathways involved in host defense were regulated by

NOD2 in human macrophages including iNOS, autophagy, and apoptosis (Figure 3.1).

Although somewhat controversial, studies in the literature support involvement of iNOS,

and thus its product NO, during M.tb infection of human cells (290,291). Thus, we

decided to study this enzyme in the context of NOD2. NOD2 agonists have previously

been shown to induce iNOS protein expression in mouse macrophages (283), but no one

has studied the role of NOD2 in iNOS activity in human macrophages. We observed that

135

iNOS mRNA and protein are expressed in response to NOD2 agonists (Figure 3.2). It is

of interest that higher mRNA expression was observed in response to the mycobacterial

NOD2 agonist, GMDP, when compared to the Gram negative / Gram positive NOD2

agonist, MDP, but no differences were observed in iNOS protein expression. One

potential explanation for this result is that GMDP is able to induce iNOS protein

expression for a longer time period as a result of increased stability of transcript levels.

This idea was not studied in the context of the current work but if this was found to be the

case, the result would add to our understanding of iNOS transcriptional regulation.

After determining that NOD2 may be involved in iNOS expression through the usage of

NOD2 agonists, we more directly assessed the role of NOD2 using siRNA knockdown.

NOD2 deficient macrophages treated with GMDP or IFNγ resulted in decreased iNOS

expression (Figure 3.2). Thus, these data indicate that NOD2 is involved in iNOS

expression in human macrophages. In addition, we observed that NO was enhanced in

response to the NOD2 agonist GMDP (Figure 3.2). These results raise the possibility that

NOD2 is necessary for NO production in response to invading organisms. Ongoing

studies involve assessing NO production in NOD2 deficient cells stimulated with these

agonists.

We next explored whether M.tb induces iNOS expression and NO production in our

MDM model, since the literature is not in agreement on this subject matter (337). We

observed enhanced iNOS expression and NO production in response to M.tb (Figure 3.3).

136

In addition we found that M.tb-induced iNOS expression is NOD2-dependent (Figure

3.3). It is of interest that NOD2 regulates M.tb-induced iNOS expression and potentially

NO production. This links a cytosolic PRR with the induction of iNOS and the

antimicrobial killing effects of NO. It would be of interest to study which molecules in

the iNOS signaling pathway are regulated by NOD2 during M.tb infection. This

experiment would add to our understanding of the relationship between NOD2 and an

antimicrobial mediator, iNOS. It would also be interesting to identify if GMDP is the

main component of M.tb responsible for iNOS expression. To study this we are currently

using M.tb deficient in NamH, the gene necessary for the production of GMDP (86), to

analyze iNOS expression and NO production.

Since NOD2 is a relatively new PRR and its ability to recognize its agonist is still

unknown, we decided to study the signaling cascade to get a better understanding of how

NOD2 signals in response to pathogens [Chapter 4]. We began these studies by

stimulating NOD2 deficient cells with PMA and observed a decrease in cytokine

production. In addition, we found that PMA induces the NOD2 downstream signaling

kinase RIP2 in a NOD2-dependent manner (Figure 4.1). This is of interest because PMA

induces many PKC isoforms by mimicking DAG, and suggests that NOD2 and PKC

signal in concert to produce pro-inflammatory cytokines.

A PKC phospho-array identified PKCδ as a kinase phosphorylated by the NOD2

agonists, MDP and GMDP, and kinase assays confirmed that PKCδ is active (Figure 4.2).

137

In addition, we observed that PMA and GMDP-induced PKCδ phosphorylation is NOD2-

dependent (Figure 4.2). These data indicate that NOD2 is a necessary molecule in

optimally activating PKCδ.

To delineate PKCδ’s involvement in PMA-induced RIP2 phosphorylation and where

PKCδ belongs in the NOD2 signaling cascade, we incubated PKCδ deficient

macrophages with PMA or GMDP. We observed a decrease in RIP2 phosphorylation,

indicating that PKCδ must be activated prior to RIP2 (Figure 4.3). These data provide

new evidence for PKCδ being a necessary kinase in the NOD2 signaling pathway. We

hypothesize that PKCδ is in the NOD2-RIP2 complex, and therefore it would be of

interest to perform immunoprecipitations with NOD2 and PKCδ or RIP2 and PKCδ in

order to confirm. Additionally, we would like to know whether this complex binds to the

downstream molecule TAK1, which the NOD2-RIP2 complex is known to bind (95,96).

Finally we would like to understand the importance to PKCδ in the NOD2 signaling

cascade. Since PKCδ is known to phosphorylate hydroxyl groups on serine and

threonine residues, we hypothesize that it does this to RIP2 in order for RIP2 to be active.

Currently the literature suggests that RIP2 is autophosphorylated when recruited by its

CARD domain to NOD2 (338). More experiments in this area would add to our

understanding of the NOD2 signaling cascade that contributes to the innate immune

response to invading organisms.

138

New approaches to the treatment and prevention of inflammatory or infectious diseases

could involve the identification of novel activators or inhibitors of NOD2. Since NOD2

mutations (gain and loss of function) are associated with Blau syndrome and Crohn’s

disease, future treatment for these diseases could include NOD2 replacement or

modifying therapy (depending on the disease) (339). NOD2 polymorphisms in certain

populations are associated with susceptibility to TB and leprosy, so NOD2 replacement

or modifying therapy could also be given to high risk individuals as a preventative. Our

findings that NOD2 is important in controlling early M.tb infection of macrophages

suggest that activators such as MDP, could be given to individuals with primary infection

or potentially to those with reactivation disease. MDP has been used for decades as a key

component in complete Freund’s adjuvant to elicit adjuvant activity to injected antigens,

and the use of MDP in immunotherapy of infections and cancer is currently being studied

(340,341). However, since NOD2 gain of function mutations result in Blau syndrome,

MDP therapy should be used in moderation and studies should be done to test for the

correct dosage and length of administration.

In summary, work in this dissertation adds to our understanding of the complex role of

the cytosolic receptor, NOD2, in innate immunity. We focused on the role of NOD2

during M.tb infection. This PRR has been shown to regulate NF-κB and subsequent pro-

inflammatory cytokine secretion in response to specific bacterial products. In this

dissertation, we observed in human macrophages the role of NOD2 in cytokine secretion

and viability of M.tb, its importance in M.tb-induced iNOS expression, and its

139

relationship to a new molecule in the NOD2 signaling cascade. In addition, we

optimized antibodies for the detection of NOD2 in human cells, which led to the first

report of NOD2 protein expression in human macrophages. Furthermore, the

contribution of a new model, siRNA mediated knockdown of NOD2, to study NOD2 in

human cells is of great importance to the field. We would argue that previous studies

using mouse models have given misleading results on the role of NOD2 in TB. As a

result of work presented in this dissertation, we have elucidated the role of NOD2 using a

more clinically relevant model, which has the potential of identifying new drug targets to

treat or prevent the disease. Our studies on the relationship between NOD2 and iNOS

bring us a step closer to understanding the role of NOD2 in controlling the viability of

M.tb in human macrophages. Moreover, our reports on the involvement of PKCδ in the

NOD2 signaling cascade will be beneficial in future drug development studies. In all, our

data provide strong evidence for the role of NOD2 during M.tb infection of human

macrophages. Further studies will identify how NOD2 recognizes M.tb, the molecules

involved in controlling M.tb viability, whether NOD2 is directly involved in NO

production, and the importance of PKCδ in the NOD2 signaling cascade, thus further

increasing our knowledge of the role of NOD2 in innate immunity.

140

Bibliography

1. Janeway, C. A., Jr., and R. Medzhitov. 2002. Innate immune recognition. Annu. Rev. Immunol. 20: 197-216

2. Medzhitov, R. 2007. Recognition of microorganisms and activation of the immune response. Nature 449: 819-826

3. Takeda, K., T. Kaisho, and S. Akira. 2003. Toll-like receptors. Annu. Rev. Immunol. 21: 335-376

4. Flannagan, R. S., G. Cosio, and S. Grinstein. 2009. Antimicrobial mechanisms of phagocytes and bacterial evasion strategies. Nat. Rev. Microbiol. 7: 355-366

5. Underhill, D. M., and A. Ozinsky. 2002. Phagocytosis of microbes: complexity in action. Annu. Rev. Immunol. 20: 825-852

6. Underhill, D. M. 2005. Phagosome maturation: steady as she goes. Immunity. 23: 343-344

7. Forman, H. J., and M. Torres. 2002. Reactive oxygen species and cell signaling: respiratory burst in macrophage signaling. Am. J Respir. Crit Care Med 166: S4-S8

8. Kontny, E., M. Kurowska, K. Szczepanska, and W. Maslinski. 2000. Rottlerin, a PKC isozyme-selective inhibitor, affects signaling events and cytokine production in human monocytes. J Leukoc. Biol. 67: 249-258

9. Kontny, E., M. Ziolkowska, A. Ryzewska, and W. Maslinski. 1999. Protein kinase c-dependent pathway is critical for the production of pro-inflammatory cytokines (TNF-alpha, IL-1beta, IL-6). Cytokine 11: 839-848

10. Nishizuka, Y. 1984. The role of protein kinase C in cell surface signal transduction and tumour promotion. Nature 308: 693-698

141

11. Takai, Y., A. Kishimoto, Y. Iwasa, Y. Kawahara, T. Mori, and Y. Nishizuka. 1979. Calcium-dependent activation of a multifunctional protein kinase by membrane phospholipids. J Biol. Chem 254: 3692-3695

12. Ono, Y., T. Fujii, K. Ogita, U. Kikkawa, K. Igarashi, and Y. Nishizuka. 1988. The structure, expression, and properties of additional members of the protein kinase C family. J Biol. Chem 263: 6927-6932

13. Ono, Y., T. Fujii, K. Igarashi, T. Kuno, C. Tanaka, U. Kikkawa, and Y. Nishizuka. 1989. Phorbol ester binding to protein kinase C requires a cysteine-rich zinc-finger-like sequence. Proc. Natl. Acad. Sci. U. S. A 86: 4868-4871

14. Castagna, M., Y. Takai, K. Kaibuchi, K. Sano, U. Kikkawa, and Y. Nishizuka. 1982. Direct activation of calcium-activated, phospholipid-dependent protein kinase by tumor-promoting phorbol esters. J Biol. Chem. 257: 7847-7851

15. Palmer, R. H., J. Ridden, and P. J. Parker. 1995. Cloning and expression patterns of two members of a novel protein-kinase-C-related kinase family. Eur. J Biochem. 227: 344-351

16. Palmer, R. H., L. V. Dekker, R. Woscholski, J. A. Le Good, R. Gigg, and P. J. Parker. 1995. Activation of PRK1 by phosphatidylinositol 4,5-bisphosphate and phosphatidylinositol 3,4,5-trisphosphate. A comparison with protein kinase C isotypes. J Biol. Chem 270: 22412-22416

17. Amano, M., H. Mukai, Y. Ono, K. Chihara, T. Matsui, Y. Hamajima, K. Okawa, A. Iwamatsu, and K. Kaibuchi. 1996. Identification of a putative target for Rho as the serine-threonine kinase protein kinase N. Science 271: 648-650

18. Watanabe, G., Y. Saito, P. Madaule, T. Ishizaki, K. Fujisawa, N. Morii, H. Mukai, Y. Ono, A. Kakizuka, and S. Narumiya. 1996. Protein kinase N (PKN) and PKN-related protein rhophilin as targets of small GTPase Rho. Science 271: 645-648

19. Mecklenbrauker, I., K. Saijo, N. Y. Zheng, M. Leitges, and A. Tarakhovsky. 2002. Protein kinase Cdelta controls self-antigen-induced B-cell tolerance. Nature 416: 860-865

20. Brodie, C., and P. M. Blumberg. 2003. Regulation of cell apoptosis by protein kinase c delta. Apoptosis. 8: 19-27

21. Ohba, M., K. Ishino, M. Kashiwagi, S. Kawabe, K. Chida, N. H. Huh, and T. Kuroki. 1998. Induction of differentiation in normal human keratinocytes by adenovirus-mediated introduction of the eta and delta isoforms of protein kinase C. Mol. Cell Biol. 18: 5199-5207

142

22. Blake, R. A., P. Garcia-Paramio, P. J. Parker, and S. A. Courtneidge. 1999. Src promotes PKCdelta degradation. Cell Growth Differ. 10: 231-241

23. Hodgkinson, C. P., E. M. Sale, and G. J. Sale. 2002. Characterization of PDK2 activity against protein kinase B gamma. Biochemistry 41: 10351-10359

24. Kristof, A. S., J. Marks-Konczalik, E. Billings, and J. Moss. 2003. Stimulation of signal transducer and activator of transcription-1 (STAT1)-dependent gene transcription by lipopolysaccharide and interferon-gamma is regulated by mammalian target of rapamycin. J Biol. Chem. 278: 33637-33644

25. Braiman, L., A. Alt, T. Kuroki, M. Ohba, A. Bak, T. Tennenbaum, and S. R. Sampson. 2001. Insulin induces specific interaction between insulin receptor and protein kinase C delta in primary cultured skeletal muscle. Mol. Endocrinol. 15: 565-574

26. Rosenzweig, T., L. Braiman, A. Bak, A. Alt, T. Kuroki, and S. R. Sampson. 2002. Differential effects of tumor necrosis factor-alpha on protein kinase C isoforms alpha and delta mediate inhibition of insulin receptor signaling. Diabetes 51: 1921-1930

27. Marsland, B. J., and M. Kopf. 2008. T-cell fate and function: PKC-theta and beyond. Trends Immunol. 29: 179-185

28. Sun, Z., C. W. Arendt, W. Ellmeier, E. M. Schaeffer, M. J. Sunshine, L. Gandhi, J. Annes, D. Petrzilka, A. Kupfer, P. L. Schwartzberg, and D. R. Littman. 2000. PKC-theta is required for TCR-induced NF-kappaB activation in mature but not immature T lymphocytes. Nature 404: 402-407

29. Ron, D., E. W. Napolitano, A. Voronova, N. J. Vasquez, D. N. Roberts, B. L. Calio, R. H. Caothien, S. M. Pettiford, S. Wellik, J. B. Mandac, and L. M. Kauvar. 1999. Direct interaction in T-cells between thetaPKC and the tyrosine kinase p59fyn. J Biol. Chem. 274: 19003-19010

30. Bauer, B., N. Krumbock, F. Fresser, F. Hochholdinger, M. Spitaler, A. Simm, F. Uberall, B. Schraven, and G. Baier. 2001. Complex formation and cooperation of protein kinase C theta and Akt1/protein kinase B alpha in the NF-kappa B transactivation cascade in Jurkat T cells. J Biol. Chem. 276: 31627-31634

31. Takeda, K., Y. Harada, R. Watanabe, Y. Inutake, S. Ogawa, K. Onuki, S. Kagaya, K. Tanabe, H. Kishimoto, and R. Abe. 2008. CD28 stimulation triggers NF-kappaB activation through the CARMA1-PKCtheta-Grb2/Gads axis. Int. Immunol. 20: 1507-1515

143

32. Webb, J. L., M. W. Harvey, D. W. Holden, and T. J. Evans. 2001. Macrophage nitric oxide synthase associates with cortical actin but is not recruited to phagosomes. Infect. Immun. 69: 6391-6400

33. Fang, F. C. 2004. Antimicrobial reactive oxygen and nitrogen species: concepts and controversies. Nat. Rev. Microbiol. 2: 820-832

34. Cooper, A. M., L. B. Adams, D. K. Dalton, R. Appelberg, and S. Ehlers. 2002. IFN-gamma and NO in mycobacterial disease: new jobs for old hands. Trends Microbiol. 10: 221-226

35. Alam, M. S., T. Akaike, S. Okamoto, T. Kubota, J. Yoshitake, T. Sawa, Y. Miyamoto, F. Tamura, and H. Maeda. 2002. Role of nitric oxide in host defense in murine salmonellosis as a function of its antibacterial and antiapoptotic activities. Infect. Immun. 70: 3130-3142

36. Li, H., and U. Forstermann. 2000. Nitric oxide in the pathogenesis of vascular disease. J Pathol. 190: 244-254

37. Laroux, F. S., D. J. Lefer, S. Kawachi, R. Scalia, A. S. Cockrell, L. Gray, H. H. Van der, J. M. Hoffman, and M. B. Grisham. 2000. Role of nitric oxide in the regulation of acute and chronic inflammation. Antioxid. Redox. Signal. 2: 391-396

38. Nathan, C., and Q.-W. Xie. 1994. Regulation of biosynthesis of nitric oxide. J. Biol. Chem. 269: 13725-13728

39. Rosenkranz-Weiss, P., W. C. Sessa, S. Milstien, S. Kaufman, C. A. Watson, and J. S. Pober. 1994. Regulation of nitric oxide synthesis by proinflammatory cytokines in human umbilical vein endothelial cells. Elevations in tetrahydrobiopterin levels enhance endothelial nitric oxide synthase specific activity. J Clin. Invest 93: 2236-2243

40. Assreuy, J., F. Q. Cunha, F. Y. Liew, and S. Moncada. 1993. Feedback inhibition of nitric oxide synthase activity by nitric oxide. Br. J Pharmacol. 108: 833-837

41. Rogers, N. E., and L. J. Ignarro. 1992. Constitutive nitric oxide synthase from cerebellum is reversibly inhibited by nitric oxide formed from L-arginine. Biochem. Biophys. Res. Commun. 189: 242-249

42. Martin, E., C. Nathan, and Q. Xie. 1994. Role of interferon regulatory factor 1 in induction of nitric oxide synthase. J. Exp. Med. 180: 977-984

43. Davis, R. L., A. C. Sanchez, D. J. Lindley, S. C. Williams, and P. J. Syapin. 2005. Effects of mechanistically distinct NF-kappaB inhibitors on glial inducible nitric-oxide synthase expression. Nitric. Oxide. 12: 200-209

144

44. Weinberg, J. B. 1998. Nitric oxide production and nitric oxide synthase type 2 expression by human mononuclear phagocytes: a review. Mol. Med. 4: 557-591

45. Abbas, A. K., and A. H. Lichtman. 2005. Cellular and Molecular Immunology.

46. Zhang, P., W. R. Summer, G. J. Bagby, and S. Nelson. 2000. Innate immunity and pulmonary host defense. Immunol. Rev. 173: 39-51

47. Zaas, A. K., and D. A. Schwartz. 2005. Innate immunity and the lung: defense at the interface between host and environment. Trends Cardiovasc. Med. 15: 195-202

48. Fels, A., and Z. A. Cohn. 1986. The alveolar macrophage. J. Appl. Physiol. 60: 353-369

49. Williams, M. C. 2003. Alveolar type I cells: molecular phenotype and development. Annu. Rev. Physiol 65: 669-695

50. Hoidal, J. R., D. Schmeling, and P. K. Peterson. 1981. Phagocytosis, bacterial killing, and metabolism by purified human lung phagocytes. J. Infect. Dis. 144: 61-71

51. Lohmann-Matthes, M. L., C. Steinmuller, and G. Franke-Ullmann. 1994. Pulmonary macrophages. Eur. Respir. J 7: 1678-1689

52. Gordon, S. 2003. Alternative activation of macrophages. Nat. Rev. Immunol. 3: 23-35

53. Mosser, D. M. 2003. The many faces of macrophage activation. J Leukoc. Biol 73: 209-212

54. Henning, L. N., A. K. Azad, K. V. Parsa, J. E. Crowther, S. Tridandapani, and L. S. Schlesinger. 2008. Pulmonary surfactant protein A regulates TLR expression and activity in human macrophages. J. Immunol. 180: 7847-7858

55. Crowther, J. E., V. K. Kutala, P. Kuppusamy, J. S. Ferguson, A. A. Beharka, J. L. Zweier, F. X. McCormack, and L. S. Schlesinger. 2004. Pulmonary surfactant protein a inhibits macrophage reactive oxygen intermediate production in response to stimuli by reducing NADPH oxidase activity. J Immunol. 172: 6866-6874

56. Rajaram, M. V., M. N. Brooks, J. D. Morris, J. B. Torrelles, A. K. Azad, and L. S. Schlesinger. 2010. Mycobacterium tuberculosis activates human macrophage peroxisome proliferator-activated receptor gamma linking mannose receptor recognition to regulation of immune responses. J Immunol. 185: 929-942

145

57. Brown E.J. 1991. Complement Receptors and Phagocytosis. Curr. Opin. Immunol. 3: 76-82

58. Ravetch, J. V., and S. Bolland. 2001. IgG Fc receptors. Annu. Rev. Immunol. 19: 275-290

59. Stephenson, J. D., and V. L. Shepherd. 1987. Purification of the human alveolar macrophage mannose receptor. Biochem. Biophys. Res. Commun. 148: 883-889

60. Palecanda, A., and L. Kobzik. 2001. Receptors for unopsonized particles: the role of alveolar macrophage scavenger receptors. Curr. Mol. Med 1: 589-595

61. Taylor, P. R., G. D. Brown, D. M. Reid, J. A. Willment, L. Martinez-Pomares, S. Gordon, and S. Y. Wong. 2002. The beta-glucan receptor, dectin-1, is predominantly expressed on the surface of cells of the monocyte/macrophage and neutrophil lineages. J Immunol. 169: 3876-3882

62. Underhill, D. M. 2003. Toll-like receptors: networking for success. Eur. J Immunol. 33: 1767-1775

63. Chen, S. K. a. N. n. 2009. NOD-Like Receptors: Role in Innate Immunity and Inflammatory Disease. The Annual Review of Pathology: Mechanisms of Disease

64. Stahl, P. D., and R. A. Ezekowitz. 1998. The mannose receptor is a pattern recognition receptor involved in host defense. Curr. Opin. Immunol. 10: 50-55

65. Gibson, G. a. W. 2010. Expression and Functions of Innate Pattern Recognition Receptors in T and B Cells. Immun. , Endoc. & Metab. Agents in Med. Chem 10: 11-20

66. Basu, S., and M. J. Fenton. 2004. Toll-like receptors: function and roles in lung disease. Am. J. Physiol Lung Cell Mol. Physiol 286: L887-L892

67. Akira, S., S. Uematsu, and O. Takeuchi. 2006. Pathogen recognition and innate immunity. Cell 124: 783-801

68. Poltorak, A., X. He, I. Smirnova, M. Y. Liu, H. C. Van, X. Du, D. Birdwell, E. Alejos, M. Silva, C. Galanos, M. Freudenberg, P. Ricciardi-Castagnoli, B. Layton, and B. Beutler. 1998. Defective LPS signaling in C3H/HeJ and C57BL/10ScCr mice: mutations in Tlr4 gene. Science 282: 2085-2088

69. Underhill, D. M., A. Ozinsky, A. M. Hajjar, A. Stevens, C. B. Wilson, M. Bassetti, and A. Aderem. 1999. The toll-like receptor 2 is recruited to macrophage phagosomes and discriminates between pathogens. Nature 401: 811-815

146

70. Kobayashi, K., L. D. Hernandez, J. E. Galan, C. A. Janeway, Jr., R. Medzhitov, and R. A. Flavell. 2002. IRAK-M is a negative regulator of Toll-like receptor signaling. Cell 110: 191-202

71. Creagh, E. M., and L. A. O'Neill. 2006. TLRs, NLRs and RLRs: a trinity of pathogen sensors that co-operate in innate immunity. Trends Immunol. 27: 352-357

72. Hulbert, S. H., C. A. Webb, S. M. Smith, and Q. Sun. 2001. Resistance gene complexes: evolution and utilization. Annu. Rev. Phytopathol. 39: 285-312

73. Harton, J. A., M. W. Linhoff, J. Zhang, and J. P. Ting. 2002. Cutting edge: CATERPILLER: a large family of mammalian genes containing CARD, pyrin, nucleotide-binding, and leucine-rich repeat domains. J Immunol. 169: 4088-4093

74. Inohara, Chamaillard, C. McDonald, and G. Nunez. 2005. NOD-LRR proteins: role in host-microbial interactions and inflammatory disease. Annu. Rev. Biochem. 74: 355-383

75. Ogura, Y., S. Lala, W. Xin, E. Smith, T. A. Dowds, F. F. Chen, E. Zimmermann, M. Tretiakova, J. H. Cho, J. Hart, J. K. Greenson, S. Keshav, and G. Nunez. 2003. Expression of NOD2 in Paneth cells: a possible link to Crohn's ileitis. Gut 52: 1591-1597

76. Ogura, Y., N. Inohara, A. Benito, F. F. Chen, S. Yamaoka, and G. Nunez. 2001. Nod2, a Nod1/Apaf-1 family member that is restricted to monocytes and activates NF-kappaB. J Biol Chem 276: 4812-4818

77. Brooks, M. N., M. V. Rajaram, A. K. Azad, A. O. Amer, M. A. Valdivia-Arenas, J. H. Park, G. Nunez, and L. S. Schlesinger. 2011. NOD2 controls the nature of the inflammatory response and subsequent fate of Mycobacterium tuberculosis and M. bovis BCG in human macrophages. Cell Microbiol. 13: 402-418

78. Ting, J. P., R. C. Lovering, E. S. Alnemri, J. Bertin, J. M. Boss, B. K. Davis, R. A. Flavell, S. E. Girardin, A. Godzik, J. A. Harton, H. M. Hoffman, J. P. Hugot, N. Inohara, A. MacKenzie, L. J. Maltais, G. Nunez, Y. Ogura, L. A. Otten, D. Philpott, J. C. Reed, W. Reith, S. Schreiber, V. Steimle, and P. A. Ward. 2008. The NLR gene family: a standard nomenclature. Immunity. 28: 285-287

79. Tanabe, T., M. Chamaillard, Y. Ogura, L. Zhu, S. Qiu, J. Masumoto, P. Ghosh, A. Moran, M. M. Predergast, G. Tromp, C. J. Williams, N. Inohara, and G. Nunez. 2004. Regulatory regions and critical residues of NOD2 involved in muramyl dipeptide recognition. EMBO J 23: 1587-1597

147

80. Kajava, A. V., and B. Kobe. 2002. Assessment of the ability to model proteins with leucine-rich repeats in light of the latest structural information. Protein Sci. 11: 1082-1090

81. Martinon, F., and J. Tschopp. 2004. Inflammatory caspases: linking an intracellular innate immune system to autoinflammatory diseases. Cell 117: 561-574

82. Manon, F., A. Favier, G. Nunez, J. P. Simorre, and S. Cusack. 2007. Solution structure of NOD1 CARD and mutational analysis of its interaction with the CARD of downstream kinase RICK. J Mol. Biol. 365: 160-174

83. Coussens, N. P., J. C. Mowers, C. McDonald, G. Nunez, and S. Ramaswamy. 2007. Crystal structure of the Nod1 caspase activation and recruitment domain. Biochem. Biophys. Res. Commun. 353: 1-5

84. Rosenstiel, P., A. Till, and S. Schreiber. 2007. NOD-like receptors and human diseases. Microbes. Infect. 9: 648-657

85. Inohara, N., Y. Ogura, A. Fontalba, O. Gutierrez, F. Pons, J. Crespo, K. Fukase, S. Inamura, S. Kusumoto, M. Hashimoto, S. J. Foster, A. P. Moran, J. L. Fernandez-Luna, and G. Nunez. 2003. Host recognition of bacterial muramyl dipeptide mediated through NOD2. Implications for Crohn's disease. J Biol Chem 278: 5509-5512

86. Coulombe, F., M. Divangahi, F. Veyrier, L. L. de, J. L. Gleason, Y. Yang, M. A. Kelliher, A. K. Pandey, C. M. Sassetti, M. B. Reed, and M. A. Behr. 2009. Increased NOD2-mediated recognition of N-glycolyl muramyl dipeptide. J. Exp. Med. 206: 1709-1716

87. Chamaillard, M., M. Hashimoto, Y. Horie, J. Masumoto, S. Qiu, L. Saab, Y. Ogura, A. Kawasaki, K. Fukase, S. Kusumoto, M. A. Valvano, S. J. Foster, T. W. Mak, G. Nunez, and N. Inohara. 2003. An essential role for NOD1 in host recognition of bacterial peptidoglycan containing diaminopimelic acid. Nat. Immunol. 4: 702-707

88. Girardin, S. E., I. G. Boneca, J. Viala, M. Chamaillard, A. Labigne, G. Thomas, D. J. Philpott, and P. J. Sansonetti. 2003. Nod2 is a general sensor of peptidoglycan through muramyl dipeptide (MDP) detection. J Biol Chem 278: 8869-8872

89. Holtje, J. V. 1998. Growth of the stress-bearing and shape-maintaining murein sacculus of Escherichia coli. Microbiol. Mol. Biol. Rev. 62: 181-203

148

90. Sinha, R. K., and R. S. Rosenthal. 1980. Release of soluble peptidoglycan from growing conococci: demonstration of anhydro-muramyl-containing fragments. Infect. Immun. 29: 914-925

91. Cookson, B. T., A. N. Tyler, and W. E. Goldman. 1989. Primary structure of the peptidoglycan-derived tracheal cytotoxin of Bordetella pertussis. Biochemistry 28: 1744-1749

92. Araki, Y., S. Fukuoka, S. Oba, and E. Ito. 1971. Enzymatic deacetylation of N-acetylglucosamine residues in peptidoglycan from Bacillus cereus cell walls. Biochem. Biophys. Res. Commun. 45: 751-758

93. Franchi, L., J. H. Park, M. H. Shaw, N. Marina-Garcia, G. Chen, Y. G. Kim, and G. Nunez. 2008. Intracellular NOD-like receptors in innate immunity, infection and disease. Cell Microbiol. 10: 1-8

94. Nembrini, C., J. Kisielow, A. T. Shamshiev, L. Tortola, A. J. Coyle, M. Kopf, and B. J. Marsland. 2009. The kinase activity of Rip2 determines its stability and consequently Nod1- and Nod2-mediated immune responses. J Biol. Chem. 284: 19183-19188

95. Chen, C. M., Y. Gong, M. Zhang, and J. J. Chen. 2004. Reciprocal cross-talk between Nod2 and TAK1 signaling pathways. J. Biol. Chem. 279: 25876-25882

96. Kim, J. Y., E. Omori, K. Matsumoto, G. Nunez, and J. Ninomiya-Tsuji. 2008. TAK1 is a central mediator of NOD2 signaling in epidermal cells. J. Biol. Chem. 283: 137-144

97. Hsu, L. C., S. R. Ali, S. McGillivray, P. H. Tseng, S. Mariathasan, E. W. Humke, L. Eckmann, J. J. Powell, V. Nizet, V. M. Dixit, and M. Karin. 2008. A NOD2-NALP1 complex mediates caspase-1-dependent IL-1beta secretion in response to Bacillus anthracis infection and muramyl dipeptide. Proc. Natl. Acad. Sci. U. S. A 105: 7803-7808

98. Rosenstiel, P., K. Huse, A. Till, J. Hampe, S. Hellmig, C. Sina, S. Billmann, K. O. von, G. H. Waetzig, M. Platzer, D. Seegert, and S. Schreiber. 2006. A short isoform of NOD2/CARD15, NOD2-S, is an endogenous inhibitor of NOD2/receptor-interacting protein kinase 2-induced signaling pathways. Proc. Natl. Acad. Sci. U. S. A 103: 3280-3285

99. Netea, M. G., R. Sutmuller, C. Hermann, C. A. Van der Graaf, J. W. Van der Meer, J. H. van Krieken, T. Hartung, G. Adema, and B. J. Kullberg. 2004. Toll-like receptor 2 suppresses immunity against Candida albicans through induction of IL-10 and regulatory T cells. J Immunol. 172: 3712-3718

149

100. Netea, M. G., G. Ferwerda, D. J. de Jong, T. Jansen, L. Jacobs, M. Kramer, T. H. Naber, J. P. Drenth, S. E. Girardin, B. J. Kullberg, G. J. Adema, and J. W. Van der Meer. 2005. Nucleotide-binding oligomerization domain-2 modulates specific TLR pathways for the induction of cytokine release. J Immunol. 174: 6518-6523

101. Uehara, A., Y. Sugawara, S. Kurata, Y. Fujimoto, K. Fukase, S. Kusumoto, Y. Satta, T. Sasano, S. Sugawara, and H. Takada. 2005. Chemically synthesized pathogen-associated molecular patterns increase the expression of peptidoglycan recognition proteins via toll-like receptors, NOD1 and NOD2 in human oral epithelial cells. Cell Microbiol. 7: 675-686

102. van Heel, D. A., S. Ghosh, K. A. Hunt, C. G. Mathew, A. Forbes, D. P. Jewell, and R. J. Playford. 2005. Synergy between TLR9 and NOD2 innate immune responses is lost in genetic Crohn's disease. Gut 54: 1553-1557

103. Watanabe, T., A. Kitani, P. J. Murray, and W. Strober. 2004. NOD2 is a negative regulator of Toll-like receptor 2-mediated T helper type 1 responses. Nat. Immunol. 5: 800-808

104. Watanabe, T., A. Kitani, P. J. Murray, Y. Wakatsuki, I. J. Fuss, and W. Strober. 2006. Nucleotide binding oligomerization domain 2 deficiency leads to dysregulated TLR2 signaling and induction of antigen-specific colitis. Immunity. 25: 473-485

105. Wolfert, M. A., T. F. Murray, G. J. Boons, and J. N. Moore. 2002. The origin of the synergistic effect of muramyl dipeptide with endotoxin and peptidoglycan. J Biol. Chem 277: 39179-39186

106. Yang, S., R. Tamai, S. Akashi, O. Takeuchi, S. Akira, S. Sugawara, and H. Takada. 2001. Synergistic effect of muramyldipeptide with lipopolysaccharide or lipoteichoic acid to induce inflammatory cytokines in human monocytic cells in culture. Infect. Immun. 69: 2045-2053

107. Abbott, D. W., A. Wilkins, J. M. Asara, and L. C. Cantley. 2004. The Crohn's disease protein, NOD2, requires RIP2 in order to induce ubiquitinylation of a novel site on NEMO. Curr. Biol. 14: 2217-2227

108. Abbott, D. W., Y. Yang, J. E. Hutti, S. Madhavarapu, M. A. Kelliher, and L. C. Cantley. 2007. Coordinated regulation of Toll-like receptor and NOD2 signaling by K63-linked polyubiquitin chains. Mol. Cell Biol. 27: 6012-6025

109. Rosenstiel, P., M. Fantini, K. Brautigam, T. Kuhbacher, G. H. Waetzig, D. Seegert, and S. Schreiber. 2003. TNF-alpha and IFN-gamma regulate the expression of the NOD2 (CARD15) gene in human intestinal epithelial cells. Gastroenterology 124: 1001-1009

150

110. Takahashi, Y., K. Isuzugawa, Y. Murase, M. Imai, S. Yamamoto, M. Iizuka, S. Akira, G. M. Bahr, E. Momotani, M. Hori, H. Ozaki, and K. Imakawa. 2006. Up-regulation of NOD1 and NOD2 through TLR4 and TNF-alpha in LPS-treated murine macrophages. J Vet. Med. Sci. 68: 471-478

111. Takada, H., S. Yokoyama, and S. Yang. 2002. Enhancement of endotoxin activity by muramyldipeptide. Journal of Endotoxin Research 8: 337-342

112. McDonald, C., F. F. Chen, V. Ollendorff, Y. Ogura, S. Marchetto, P. Lecine, J. P. Borg, and G. Nunez. 2005. A role for Erbin in the regulation of Nod2-dependent NF-kappaB signaling. J Biol. Chem 280: 40301-40309

113. Kufer, T. A., E. Kremmer, D. J. Banks, and D. J. Philpott. 2006. Role for erbin in bacterial activation of Nod2. Infect. Immun. 74: 3115-3124

114. Barnich, N., J. E. Aguirre, H. C. Reinecker, R. Xavier, and D. K. Podolsky. 2005. Membrane recruitment of NOD2 in intestinal epithelial cells is essential for nuclear factor-{kappa}B activation in muramyl dipeptide recognition. J. Cell Biol. 170: 21-26

115. Takada, H., and S. Kotani. 1995. In The Theory of Pratical Application of Adjuvants. Wiley: Chichester, ed., pp. 171.

116. Inamura, N., S. Sone, A. Okubo, E. Kunishige, M. Nakanishi, and T. Ogura. 1989. Tumor cytotoxicity of human monocyte membrane-bound interleukin-1 alpha induced by synergistic actions of interferon-gamma and synthetic acyltripeptide, FK-565. Cancer Immunol. Immunother. 28: 164-170

117. Hasan, S. I., K. Ahmed, and J. L. Turk. 1990. Effect of anticancer drugs on the release of tumour necrosis factor in vitro. Cancer Immunol. Immunother. 30: 363-366

118. Maeda, M., T. Asano, and E. S. Kleinerman. 1993. Anti-(tumor necrosis factor) alters the response of human monocytes to liposomal muramyl tripeptide. Cancer Immunol. Immunother. 37: 203-208

119. Morikage, T., Y. Mizushima, and S. Yano. 1991. Prevention of bacterial infection and LPS-induced lethality by interleukin-1 alpha in mice. Tohoku J Exp. Med. 163: 47-57

120. Morikage, T., Y. Mizushima, and S. Yano. 1991. Prevention of bacterial infection and LPS-induced lethality by interleukin-1 alpha in mice. Tohoku J Exp. Med. 163: 47-57

151

121. Asano, T., A. McWatters, T. An, K. Matsushima, and E. S. Kleinerman. 1994. Liposomal muramyl tripeptide up-regulates interleukin-1 alpha, interleukin-1 beta, tumor necrosis factor-alpha, interleukin-6 and interleukin-8 gene expression in human monocytes. J Pharmacol. Exp. Ther. 268: 1032-1039

122. Li, J., T. Moran, E. Swanson, C. Julian, J. Harris, D. K. Bonen, M. Hedl, D. L. Nicolae, C. Abraham, and J. H. Cho. 2004. Regulation of IL-8 and IL-1beta expression in Crohn's disease associated NOD2/CARD15 mutations. Hum. Mol. Genet. 13: 1715-1725

123. Nau, G. J., J. F. Richmond, A. Schlesinger, E. G. Jennings, E. S. Lander, and R. A. Young. 2002. Human macrophage activation programs induced by bacterial pathogens. Proc. Natl. Acad. Sci. U. S. A 99: 1503-1508

124. Todate, A., T. Suda, H. Kuwata, K. Chida, and H. Nakamura. 2001. Muramyl dipeptide-Lys stimulates the function of human dendritic cells. J Leukoc. Biol. 70: 723-729

125. Heinzelmann, M., H. C. Polk, Jr., A. Chernobelsky, T. P. Stites, and L. E. Gordon. 2000. Endotoxin and muramyl dipeptide modulate surface receptor expression on human mononuclear cells. Immunopharmacology 48: 117-128

126. Opitz, B., A. Puschel, B. Schmeck, A. C. Hocke, S. Rosseau, S. Hammerschmidt, R. R. Schumann, N. Suttorp, and S. Hippenstiel. 2004. Nucleotide-binding oligomerization domain proteins are innate immune receptors for internalized Streptococcus pneumoniae. J. Biol. Chem. 279: 36426-36432

127. Ferwerda, G., S. E. Girardin, B. J. Kullberg, L. Le Bourhis, D. J. de Jong, D. M. Langenberg, R. van Crevel, G. J. Adema, T. H. Ottenhoff, J. W. Van der Meer, and M. G. Netea. 2005. NOD2 and toll-like receptors are nonredundant recognition systems of Mycobacterium tuberculosis. PLoS. Pathog. 1: 279-285

128. Girardin, S. E., R. Tournebize, M. Mavris, A. L. Page, X. Li, G. R. Stark, J. Bertin, P. S. DiStefano, M. Yaniv, P. J. Sansonetti, and D. J. Philpott. 2001. CARD4/Nod1 mediates NF-kappaB and JNK activation by invasive Shigella flexneri. EMBO Rep. 2: 736-742

129. Kim, J. G., S. J. Lee, and M. F. Kagnoff. 2004. Nod1 is an essential signal transducer in intestinal epithelial cells infected with bacteria that avoid recognition by toll-like receptors. Infect. Immun. 72: 1487-1495

130. Marriott, I., D. M. Rati, S. H. McCall, and S. L. Tranguch. 2005. Induction of Nod1 and Nod2 intracellular pattern recognition receptors in murine osteoblasts following bacterial challenge. Infect. Immun. 73: 2967-2973

152

131. Opitz, B., S. Forster, A. C. Hocke, M. Maass, B. Schmeck, S. Hippenstiel, N. Suttorp, and M. Krull. 2005. Nod1-mediated endothelial cell activation by Chlamydophila pneumoniae. Circ. Res. 96: 319-326

132. Travassos, L. H., L. A. Carneiro, S. E. Girardin, I. G. Boneca, R. Lemos, M. T. Bozza, R. C. Domingues, A. J. Coyle, J. Bertin, D. J. Philpott, and M. C. Plotkowski. 2005. Nod1 participates in the innate immune response to Pseudomonas aeruginosa. J Biol Chem 280: 36714-36718

133. Viala, J., C. Chaput, I. G. Boneca, A. Cardona, S. E. Girardin, A. P. Moran, R. Athman, S. Memet, M. R. Huerre, A. J. Coyle, P. S. DiStefano, P. J. Sansonetti, A. Labigne, J. Bertin, D. J. Philpott, and R. L. Ferrero. 2004. Nod1 responds to peptidoglycan delivered by the Helicobacter pylori cag pathogenicity island. Nat. Immunol. 5: 1166-1174

134. Opitz, B., A. Puschel, W. Beermann, A. C. Hocke, S. Forster, B. Schmeck, L. van, V, T. Chakraborty, N. Suttorp, and S. Hippenstiel. 2006. Listeria monocytogenes activated p38 MAPK and induced IL-8 secretion in a nucleotide-binding oligomerization domain 1-dependent manner in endothelial cells. J. Immunol. 176: 484-490

135. Divangahi, M., S. Mostowy, F. Coulombe, R. Kozak, L. Guillot, F. Veyrier, K. S. Kobayashi, R. A. Flavell, P. Gros, and M. A. Behr. 2008. NOD2-deficient mice have impaired resistance to Mycobacterium tuberculosis infection through defective innate and adaptive immunity. J. Immunol. 181: 7157-7165

136. Deshmukh, H. S., J. B. Hamburger, S. H. Ahn, D. G. McCafferty, S. R. Yang, and V. G. Fowler, Jr. 2009. Critical role of NOD2 in regulating the immune response to Staphylococcus aureus. Infect. Immun. 77: 1376-1382

137. Amer, A., L. Franchi, T. D. Kanneganti, M. Body-Malapel, N. Ozoren, G. Brady, S. Meshinchi, R. Jagirdar, A. Gewirtz, S. Akira, and G. Nunez. 2006. Regulation of Legionella phagosome maturation and infection through flagellin and host Ipaf. J. Biol. Chem. 281: 35217-35223

138. Hugot, J. P., M. Chamaillard, H. Zouali, S. Lesage, J. P. Cezard, J. Belaiche, S. Almer, C. Tysk, C. A. O'Morain, M. Gassull, V. Binder, Y. Finkel, A. Cortot, R. Modigliani, P. Laurent-Puig, C. Gower-Rousseau, J. Macry, J. F. Colombel, M. Sahbatou, and G. Thomas. 2001. Association of NOD2 leucine-rich repeat variants with susceptibility to Crohn's disease. Nature 411: 599-603

139. Ogura, Y., D. K. Bonen, N. Inohara, D. L. Nicolae, F. F. Chen, R. Ramos, H. Britton, T. Moran, R. Karaliuskas, R. H. Duerr, J. P. Achkar, S. R. Brant, T. M. Bayless, B. S. Kirschner, S. B. Hanauer, G. Nunez, and J. H. Cho. 2001. A

153

frameshift mutation in NOD2 associated with susceptibility to Crohn's disease. Nature 411: 603-606

140. Miceli-Richard, C., S. Lesage, M. Rybojad, A. M. Prieur, S. Manouvrier-Hanu, R. Hafner, M. Chamaillard, H. Zouali, G. Thomas, and J. P. Hugot. 2001. CARD15 mutations in Blau syndrome. Nat. Genet. 29: 19-20

141. Kanazawa, N., I. Okafuji, N. Kambe, R. Nishikomori, M. Nakata-Hizume, S. Nagai, A. Fuji, T. Yuasa, A. Manki, Y. Sakurai, M. Nakajima, H. Kobayashi, I. Fujiwara, H. Tsutsumi, A. Utani, C. Nishigori, T. Heike, T. Nakahata, and Y. Miyachi. 2005. Early-onset sarcoidosis and CARD15 mutations with constitutive nuclear factor-kappaB activation: common genetic etiology with Blau syndrome. Blood 105: 1195-1197

142. Fritz, J. H., R. L. Ferrero, D. J. Philpott, and S. E. Girardin. 2006. Nod-like proteins in immunity, inflammation and disease. Nat. Immunol. 7: 1250-1257

143. Sanderson, J. D., M. T. Moss, M. L. V. Tizard, and J. Hermon-Taylor. 1992. Mycobacterium paratuberculosis DNA in Crohn's disease tissue. Gut 33: 890-896

144. Hermon-Taylor, J. 2009. Mycobacterium avium subspecies paratuberculosis, Crohn's disease and the Doomsday scenario. Gut Pathog. 1: 15

145. Maeda, S., L. C. Hsu, H. Liu, L. A. Bankston, M. Iimura, M. F. Kagnoff, L. Eckmann, and M. Karin. 2005. Nod2 mutation in Crohn's disease potentiates NF-kappaB activity and IL-1beta processing. Science 307: 734-738

146. Netea, M. G., G. Ferwerda, D. J. de Jong, S. E. Girardin, B. J. Kullberg, and J. W. Van der Meer. 2005. NOD2 3020insC mutation and the pathogenesis of Crohn's disease: impaired IL-1beta production points to a loss-of-function phenotype. Neth. J Med. 63: 305-308

147. Hu, C., L. Sun, Y. Hu, D. Lu, H. Wang, and S. Tang. 2010. Functional characterization of the NF-kappaB binding site in the human NOD2 promoter. Cell Mol. Immunol. 7: 288-295

148. WHO. 2009. 2009 Update Tuberculosis Facts.

149. Fine, P. E. M. 1995. Variation in protection by BCG: Implications of and for heterologous immunity. Lancet 346: 1339-1345

150. Brewer, T. F. 2000. Preventing tuberculosis with bacillus Calmette-Guerin vaccine: a meta-analysis of the literature. Clin. Infect. Dis. 31 Suppl 3: S64-S67

151. Udani, P. M. 1994. BCG vaccination in India and tuberculosis in children: newer facets. Indian J Pediatr. 61: 451-462

154

152. Smith, D., E. Wiegeshaus, and V. Balasubramanian. 2000. An analysis of some hypotheses related to the Chingelput bacille Calmette-Guerin trial. Clin. Infect. Dis. 31 Suppl 3: S77-S80

153. Gupta, U. D., V. M. Katoch, and D. N. McMurray. 2007. Current status of TB vaccines. Vaccine 25: 3742-3751

154. Guy, and Mallampalli. 2008. Review: Managing TB in the 21st century: existing and novel drug therapies. Therapeutic Advances in Respiratory Disease 2: 401-408

155. WHO. 2010. Multidrug and extensively drug-resistant TB (M/XDR-TB): 2010 global report on surveillance and response. Zignol and M, eds. WHO Press, Geneva, Switzerland, pp. 1-72.

156. Ma, Z., C. Lienhardt, H. McIlleron, A. J. Nunn, and X. Wang. 2010. Global tuberculosis drug development pipeline: the need and the reality. Lancet 375: 2100-2109

157. Sundaramurthy, V., and J. Pieters. 2007. Interactions of pathogenic mycobacteria with host macrophages. Microbes. Infect. 9: 1671-1679

158. Torrelles, J. B., A. K. Azad, L. N. Henning, T. K. Carlson, and L. S. Schlesinger. 2008. Role of C-type lectins in mycobacterial infections. Curr. Drug Targets. 9: 102-112

159. Kaufmann, S. H. 2001. How can immunology contribute to the control of tuberculosis? Nature Reviews Immunology 1: 20-30

160. Wright, J. R. 2005. Immunoregulatory functions of surfactant proteins. Nat. Rev. Immunol. 5: 58-68

161. Gaynor, C. D., F. X. McCormack, D. R. Voelker, S. E. McGowan, and L. S. Schlesinger. 1995. Pulmonary surfactant protein A mediates enhanced phagocytosis of Mycobacterium tuberculosis by a direct interaction with human macrophages. J. Immunol. 155: 5343-5351

162. Downing, J. F., R. Pasula, J. R. Wright, H. L. Twigg, III, and W. J. Martin, II. 1995. Surfactant protein A promotes attachment of Mycobacterium tuberculosis to alveolar macrophages during infection with human immunodeficiency virus. Proc. Natl. Acad. Sci. USA 92: 4848-4852

163. Pasula, R., J. R. Wright, D. L. Kachel, and W. J. Martin, III. 1999. Surfactant protein A suppresses reactive nitrogen intermediates by alveolar macrophages in response to Mycobacterium tuberculosis. J. Clin. Invest. 103: 483-490

155

164. Ferguson, J. S., J. L. Martin, A. K. Azad, T. R. McCarthy, P. B. Kang, D. R. Voelker, E. C. Crouch, and L. S. Schlesinger. 2006. Surfactant protein D increases fusion of Mycobacterium tuberculosis-containing phagosomes with lysosomes in human macrophages. Infect. Immun. 74: 7005-7009

165. Ferguson, J. S., D. R. Voelker, F. X. McCormack, and L. S. Schlesinger. 1999. Surfactant protein D binds to Mycobacterium tuberculosis bacili and lipoarrabinomannan via carbohydrate-lectin interactions resulting in reduced phagocytosis of the bacteria by macrophages. J. Immunol. 163: 312-321

166. Schlesinger, L. S. 1993. Macrophage phagocytosis of virulent but not attenuated strains of Mycobacterium tuberculosis is mediated by mannose receptors in addition to complement receptors. J. Immunol. 150: 2920-2930

167. Rooyakkers AW, and R. W. Stokes. 2005. Absence of complement receptor 3 results in reduced binding and ingestion of Mycobacterium tuberculosis but has no significant effect on the induction of reactive oxygen and nitrogen intermediates or on the survival of the bacteria in resident and interferon-gamma activated macrophages., pp. 57-67.

168. Hu, C., T. Mayadas-Norton, K. Tanaka, J. Chan, and P. Salgame. 2000. Mycobacterium tuberculosis infection in complement receptor 3-deficient mice. J. Immunol. 165: 2596-2602

169. Zimmerli, S., S. Edwards, and J. D. Ernst. 1996. Selective receptor blockade during phagocytosis does not alter the survival and growth of Mycobacterium tuberculosis in human macrophages. Am. J. Respir. Cell Mol. Biol. 15: 760-770

170. Peterson, P. K., G. Gekker, S. Hu, W. S. Sheng, W. R. Anderson, R. J. Ulevitch, P. S. Tobias, K. V. Gustafson, T. W. Molitor, and C. C. Chao. 1995. CD14 receptor-mediated uptake of nonopsonized Mycobacterium tuberculosis by human microglia. Infect. Immun. 63: 1598-1602

171. Yadav, M., and J. S. Schorey. 2006. The {beta}-glucan receptor Dectin-1 fucntions together with TLR2 to mediate macrophage activation by mycobacteria. Blood 108: 3168-3175

172. Schlesinger, L. S., C. G. Bellinger-Kawahara, N. R. Payne, and M. A. Horwitz. 1990. Phagocytosis of Mycobacterium tuberculosis is mediated by human monocyte complement receptors and complement component C3. J. Immunol. 144: 2771-2780

173. Armstrong, J. A., and P. D. Hart. 1975. Phagosome-lysosome interactions in cultured macrophages infected with virulent tubercle bacilli: Reversal of the usual nonfusion pattern and observations on bacterial survial. J. Exp. Med. 142: 1-16

156

174. Lopez, M., L. M. Sly, Y. Luu, D. Young, H. Cooper, and N. E. Reiner. 2003. The 19-kDa Mycobacterium tuberculosis protein induces macrophage apoptosis through Toll-like receptor-2. J. Immunol. 170: 2409-2416

175. Dao, D. N., L. Kremer, Y. Guerardel, A. Molano, W. R. Jacobs, Jr., S. A. Porcelli, and V. Briken. 2004. Mycobacterium tuberculosis lipomannan induces apoptosis and interleukin-12 production in macrophages. Infect. Immun. 72: 2067-2074

176. Vergne, I., R. A. Fratti, P. J. Hill, J. Chua, J. Belisle, and V. Deretic. 2004. Mycobacterium tuberculosis phagosome maturation arrest: mycobacterial phosphatidylinositol analog phosphatidylinositol mannoside stimulates early endosomal fusion. Mol. Biol. Cell 15: 751-760

177. Master, S. S., S. K. Rampini, A. S. Davis, C. Keller, S. Ehlers, B. Springer, G. S. Timmins, P. Sander, and V. Deretic. 2008. Mycobacterium tuberculosis prevents inflammasome activation. Cell Host. Microbe 3: 224-232

178. Mishra, B. B., P. Moura-Alves, A. Sonawane, N. Hacohen, G. Griffiths, L. F. Moita, and E. Anes. 2010. Mycobacterium tuberculosis protein ESAT-6 is a potent activator of the NLRP3/ASC inflammasome. Cell Microbiol. 12: 1046-1063

179. Gandotra, S., S. Jang, P. J. Murray, P. Salgame, and S. Ehrt. 2007. Nucleotide-binding oligomerization domain protein 2-deficient mice control infection with Mycobacterium tuberculosis. Infect. Immun. 75: 5127-5134

180. Yang, Y., C. Yin, A. Pandey, D. Abbott, C. Sassetti, and M. A. Kelliher. 2007. NOD2 pathway activation by MDP or Mycobacterium tuberculosis infection involves the stable polyubiquitination of Rip2. J. Biol. Chem. 282: 36223-36229

181. Pandey, A. K., Y. Yang, Z. Jiang, S. M. Fortune, F. Coulombe, M. A. Behr, K. A. Fitzgerald, C. M. Sassetti, and M. A. Kelliher. 2009. NOD2, RIP2 and IRF5 play a critical role in the type I interferon response to Mycobacterium tuberculosis. PLoS. Pathog. 5: e1000500

182. Deretic, V., S. Singh, S. Master, J. Harris, E. Roberts, G. Kyei, A. Davis, H. S. De, J. Naylor, H. H. Lee, and I. Vergne. 2006. Mycobacterium tuberculosis inhibition of phagolysosome biogenesis and autophagy as a host defence mechanism. Cell Microbiol. 8: 719-727

183. Russell, D. G. 2001. Mycobacterium tuberculosis: here today, and here tomorrow. Nature Reviews 2: 1-9

157

184. Vergne, I., J. Chua, and V. Deretic. 2003. Tuberculosis toxin blocking phagosome maturation inhibits a novel Ca2+/calmodulin-PI3K hVPS34 cascade. J. Exp. Med 198: 653-659

185. Malik, Z. A., C. R. Thompson, S. Hashimi, B. Porter, S. S. Iyer, and D. J. Kusner. 2003. Cutting edge: Mycobacterium tuberculosis blocks Ca2+ signaling and phagosome maturation in human macrophages via specific inhibition of sphingosine kinase. J Immunol. 170: 2811-2815

186. Schlesinger LS, Azad AK, Torrelles JB, E. Roberts, I. Vergne, and V. Deretic. 2008. Determinants of Phagocytosis, Phagosome Biogenesis and Autophagy for Mycobacterium tuberculosis. In Handbook of Tuberculosis. Immunology and Cell Biology. S. H. E. Kaufmann and W. J. Britton, eds. Wiley-VCH Verlag GmbH&Co. KGaA, Weinheim, Germany, pp. 1.

187. Aderem, A., and D. M. Underhill. 1999. Mechanisms of phagocytosis in macrophages. Annu. Rev. Immunol. 17: 593-623

188. Kang, B. K., A. K. Azad, J. B. Torrelles, T. M. Kaufman, A. A. Beharka, E. Tibesar, L. E. Desjardin, and L. S. Schlesinger. 2005. The human macrophage mannose receptor directs Mycobacterium tuberculosis lipoarabinomannan-mediated phagosome biogenesis. J Exp. Med 202: 987-999

189. Singh, C. R., R. A. Moulton, L. Y. Armitige, A. Bidani, M. Snuggs, S. Dhandayuthapani, R. L. Hunter, and C. Jagannath. 2006. Processing and presentation of a mycobacterial antigen 85B epitope by murine macrophages is dependent on the phagosomal acquisition of vacuolar proton ATPase and in situ activation of cathepsin D. J Immunol. 177: 3250-3259

190. Sturgill-Koszycki, S., P. H. Schlesinger, P. Chakraborty, P. L. Haddix, H. L. Collins, A. K. Fok, R. D. Allen, S. L. Gluck, J. Heuser, and D. G. Russell. 1994. Lack of acidification in Mycobacterium phagosomes produced by exclusion of the vesicular proton-ATPase. Science 263: 678-681

191. Abdallah, A. M., N. C. Gey van Pittius, P. A. Champion, J. Cox, J. Luirink, C. M. Vandenbroucke-Grauls, B. J. Appelmelk, and W. Bitter. 2007. Type VII secretion--mycobacteria show the way. Nat. Rev. Microbiol. 5: 883-891

192. Mahairas, G. G., P. J. Sabo, M. J. Hickey, D. C. Singh, and C. K. Stover. 1996. Molecular analysis of genetic differences between Mycobacterium bovis BCG and virulent M-bovis. J. Bacteriol. 178: 1274-1282

193. Gordon, S. V., R. Brosch, A. Billault, T. Garnier, K. Eiglmeier, and S. T. Cole. 1999. Identification of variable regions in the genomes of tubercle bacilli using bacterial artificial chromosome arrays. Mol. Microbiol. 32: 643-655

158

194. Tekaia, F., S. V. Gordon, T. Garnier, R. Brosch, B. G. Barrell, and S. T. Cole. 1999. Analysis of the proteome of Mycobacterium tuberculosis in silico. Tuber. Lung Dis. 79: 329-342

195. Gey van Pittius, N. C., J. Gamieldien, W. Hide, G. D. Brown, R. J. Siezen, and A. D. Beyers. 2001. The ESAT-6 gene cluster of Mycobacterium tuberculosis and other high G+C Gram-positive bacteria. Genome Biol. 2: Research0044

196. Pallen, M. J. 2002. The ESAT-6/WXG100 superfamily -- and a new Gram-positive secretion system? Trends Microbiol. 10: 209-212

197. MacGurn, J. A., S. Raghavan, S. A. Stanley, and J. S. Cox. 2005. A non-RD1 gene cluster is required for Snm secretion in Mycobacterium tuberculosis. Mol. Microbiol. 57: 1653-1663

198. Fortune, S. M., A. Jaeger, D. A. Sarracino, M. R. Chase, C. M. Sassetti, D. R. Sherman, B. R. Bloom, and E. J. Rubin. 2005. Mutually dependent secretion of proteins required for mycobacterial virulence. Proc. Natl. Acad. Sci. U. S. A 102: 10676-10681

199. Stanley, S. A., S. Raghavan, W. W. Hwang, and J. S. Cox. 2003. Acute infection and macrophage subversion by Mycobacterium tuberculosis require a specialized secretion system. Proc. Natl. Acad. Sci. U. S. A 100: 13001-13006

200. Renshaw, P. S., K. L. Lightbody, V. Veverka, F. W. Muskett, G. Kelly, T. A. Frenkiel, S. V. Gordon, R. G. Hewinson, B. Burke, J. Norman, R. A. Williamson, and M. D. Carr. 2005. Structure and function of the complex formed by the tuberculosis virulence factors CFP-10 and ESAT-6. EMBO J 24: 2491-2498

201. Brodin, P., M. I. de Jonge, L. Majlessi, C. Leclerc, M. Nilges, S. T. Cole, and R. Brosch. 2005. Functional analysis of early secreted antigenic target-6, the dominant T-cell antigen of Mycobacterium tuberculosis, reveals key residues involved in secretion, complex formation, virulence, and immunogenicity. J Biol. Chem 280: 33953-33959

202. Nagai, H., E. D. Cambronne, J. C. Kagan, J. C. Amor, R. A. Kahn, and C. R. Roy. 2005. A C-terminal translocation signal required for Dot/Icm-dependent delivery of the Legionella RalF protein to host cells. Proc. Natl. Acad. Sci. U. S. A 102: 826-831

203. Vergunst, A. C., M. C. van Lier, A. den Dulk-Ras, T. A. Stuve, A. Ouwehand, and P. J. Hooykaas. 2005. Positive charge is an important feature of the C-terminal transport signal of the VirB/D4-translocated proteins of Agrobacterium. Proc. Natl. Acad. Sci. U. S. A 102: 832-837

159

204. Christie, P. J., K. Atmakuri, V. Krishnamoorthy, S. Jakubowski, and E. Cascales. 2005. Biogenesis, architecture, and function of bacterial type IV secretion systems. Annu. Rev. Microbiol. 59: 451-485

205. Lenz, L. L., S. Mohammadi, A. Geissler, and D. A. Portnoy. 2003. SecA2-dependent secretion of autolytic enzymes promotes Listeria monocytogenes pathogenesis. Proc. Natl. Acad. Sci. U. S. A 100: 12432-12437

206. Kurtz, S., K. P. McKinnon, M. S. Runge, J. P. Ting, and M. Braunstein. 2006. The SecA2 secretion factor of Mycobacterium tuberculosis promotes growth in macrophages and inhibits the host immune response. Infect. Immun. 74: 6855-6864

207. McDonough, J. A., K. E. Hacker, A. R. Flores, M. S. Pavelka, Jr., and M. Braunstein. 2005. The twin-arginine translocation pathway of Mycobacterium smegmatis is functional and required for the export of mycobacterial beta-lactamases. J Bacteriol. 187: 7667-7679

208. Posey, J. E., T. M. Shinnick, and F. D. Quinn. 2006. Characterization of the twin-arginine translocase secretion system of Mycobacterium smegmatis. J Bacteriol. 188: 1332-1340

209. Raynaud, C., C. Guilhot, J. Rauzier, Y. Bordat, V. Pelicic, R. Manganelli, I. Smith, B. Gicquel, and M. Jackson. 2002. Phospholipases C are involved in the virulence of Mycobacterium tuberculosis. Mol. Microbiol. 45: 203-217

210. Molofsky, A. B., B. G. Byrne, N. N. Whitfield, C. A. Madigan, E. T. Fuse, K. Tateda, and M. S. Swanson. 2006. Cytosolic recognition of flagellin by mouse macrophages restricts Legionella pneumophila infection. J Exp. Med. 203: 1093-1104

211. Ren, T., D. S. Zamboni, C. R. Roy, W. F. Dietrich, and R. E. Vance. 2006. Flagellin-deficient Legionella mutants evade caspase-1- and Naip5-mediated macrophage immunity. PLoS Pathog. 2: e18

212. Myrvik, Q. N., E. S. Leake, and M. J. Wright. 1984. Disruption of phagosomal membranes of normal alveolar macrophages by the H37Rv strain of Mycobacterium tuberculosis. A correlate of virulence. Am. Rev. Respir. Dis. 129: 322-328

213. McDonough, K. A., Y. Kress, and B. R. Bloom. 1993. Pathogenesis of tuberculosis: Interaction of Mycobacterium tuberculosis with macrophages. Infect. Immun. 61: 2763-2773

160

214. Stahl, P. D. 1990. The macrophage mannose receptor: current status. Am. J. Respir. Cell Mol. Biol. 2: 317-318

215. van der Wel, N., D. Hava, D. Houben, D. Fluitsma, Z. M. van, J. Pierson, M. Brenner, and P. J. Peters. 2007. M. tuberculosis and M. leprae translocate from the phagolysosome to the cytosol in myeloid cells. Cell 129: 1287-1298

216. Zhang, F. R., W. Huang, S. M. Chen, L. D. Sun, H. Liu, Y. Li, Y. Cui, X. X. Yan, H. T. Yang, R. D. Yang, T. S. Chu, C. Zhang, L. Zhang, J. W. Han, G. Q. Yu, C. Quan, Y. X. Yu, Z. Zhang, B. Q. Shi, L. H. Zhang, H. Cheng, C. Y. Wang, Y. Lin, H. F. Zheng, X. A. Fu, X. B. Zuo, Q. Wang, H. Long, Y. P. Sun, Y. L. Cheng, H. Q. Tian, F. S. Zhou, H. X. Liu, W. S. Lu, S. M. He, W. L. Du, M. Shen, Q. Y. Jin, Y. Wang, H. Q. Low, T. Erwin, N. H. Yang, J. Y. Li, X. Zhao, Y. L. Jiao, L. G. Mao, G. Yin, Z. X. Jiang, X. D. Wang, J. P. Yu, Z. H. Hu, C. H. Gong, Y. Q. Liu, R. Y. Liu, D. M. Wang, D. Wei, J. X. Liu, W. K. Cao, H. Z. Cao, Y. P. Li, W. G. Yan, S. Y. Wei, K. J. Wang, M. L. Hibberd, S. Yang, X. J. Zhang, and J. J. Liu. 2009. Genomewide association study of leprosy. N. Engl. J Med. 361: 2609-2618

217. Austin, C. M., X. Ma, and E. A. Graviss. 2008. Common nonsynonymous polymorphisms in the NOD2 gene are associated with resistance or susceptibility to tuberculosis disease in African Americans. J Infect. Dis. 197: 1713-1716

218. Le, B. L., S. Benko, and S. E. Girardin. 2007. Nod1 and Nod2 in innate immunity and human inflammatory disorders. Biochem. Soc. Trans. 35: 1479-1484

219. Sartor, R. B. 2006. Mechanisms of disease: pathogenesis of Crohn's disease and ulcerative colitis. Nat. Clin. Pract. Gastroenterol. Hepatol. 3: 390-407

220. Lala, S., K. Dheda, J. S. Chang, J. F. Huggett, L. U. Kim, M. A. Johnson, G. A. Rook, S. Keshav, and A. Zumla. 2007. The pathogen recognition sensor, NOD2, is variably expressed in patients with pulmonary tuberculosis. BMC. Infect. Dis. 7: 96

221. Waterston, R. H., K. Lindblad-Toh, E. Birney, J. Rogers, J. F. Abril, P. Agarwal, R. Agarwala, R. Ainscough, M. Alexandersson, P. An, S. E. Antonarakis, J. Attwood, R. Baertsch, J. Bailey, K. Barlow, S. Beck, E. Berry, B. Birren, T. Bloom, P. Bork, M. Botcherby, N. Bray, M. R. Brent, D. G. Brown, S. D. Brown, C. Bult, J. Burton, J. Butler, R. D. Campbell, P. Carninci, S. Cawley, F. Chiaromonte, A. T. Chinwalla, D. M. Church, M. Clamp, C. Clee, F. S. Collins, L. L. Cook, R. R. Copley, A. Coulson, O. Couronne, J. Cuff, V. Curwen, T. Cutts, M. Daly, R. David, J. Davies, K. D. Delehaunty, J. Deri, E. T. Dermitzakis, C. Dewey, N. J. Dickens, M. Diekhans, S. Dodge, I. Dubchak, D. M. Dunn, S. R. Eddy, L. Elnitski, R. D. Emes, P. Eswara, E. Eyras, A. Felsenfeld, G. A. Fewell, P. Flicek, K. Foley, W. N. Frankel, L. A. Fulton, R. S. Fulton, T. S. Furey, D.

161

Gage, R. A. Gibbs, G. Glusman, S. Gnerre, N. Goldman, L. Goodstadt, D. Grafham, T. A. Graves, E. D. Green, S. Gregory, R. Guigo, M. Guyer, R. C. Hardison, D. Haussler, Y. Hayashizaki, L. W. Hillier, A. Hinrichs, W. Hlavina, T. Holzer, F. Hsu, A. Hua, T. Hubbard, A. Hunt, I. Jackson, D. B. Jaffe, L. S. Johnson, M. Jones, T. A. Jones, A. Joy, M. Kamal, E. K. Karlsson, D. Karolchik, A. Kasprzyk, J. Kawai, E. Keibler, C. Kells, W. J. Kent, A. Kirby, D. L. Kolbe, I. Korf, R. S. Kucherlapati, E. J. Kulbokas, D. Kulp, T. Landers, J. P. Leger, S. Leonard, I. Letunic, R. Levine, J. Li, M. Li, C. Lloyd, S. Lucas, B. Ma, D. R. Maglott, E. R. Mardis, L. Matthews, E. Mauceli, J. H. Mayer, M. McCarthy, W. R. McCombie, S. McLaren, K. McLay, J. D. McPherson, J. Meldrim, B. Meredith, J. P. Mesirov, W. Miller, T. L. Miner, E. Mongin, K. T. Montgomery, M. Morgan, R. Mott, J. C. Mullikin, D. M. Muzny, W. E. Nash, J. O. Nelson, M. N. Nhan, R. Nicol, Z. Ning, C. Nusbaum, M. J. O'Connor, Y. Okazaki, K. Oliver, E. Overton-Larty, L. Pachter, G. Parra, K. H. Pepin, J. Peterson, P. Pevzner, R. Plumb, C. S. Pohl, A. Poliakov, T. C. Ponce, C. P. Ponting, S. Potter, M. Quail, A. Reymond, B. A. Roe, K. M. Roskin, E. M. Rubin, A. G. Rust, R. Santos, V. Sapojnikov, B. Schultz, J. Schultz, M. S. Schwartz, S. Schwartz, C. Scott, S. Seaman, S. Searle, T. Sharpe, A. Sheridan, R. Shownkeen, S. Sims, J. B. Singer, G. Slater, A. Smit, D. R. Smith, B. SPENCER, A. Stabenau, N. Stange-Thomann, C. Sugnet, M. Suyama, G. Tesler, J. Thompson, D. Torrents, E. Trevaskis, J. Tromp, C. Ucla, A. Ureta-Vidal, J. P. Vinson, A. C. Von Niederhausern, C. M. Wade, M. Wall, R. J. Weber, R. B. Weiss, M. C. Wendl, A. P. West, K. Wetterstrand, R. Wheeler, S. Whelan, J. Wierzbowski, D. Willey, S. Williams, R. K. Wilson, E. Winter, K. C. Worley, D. Wyman, S. Yang, S. P. Yang, E. M. Zdobnov, M. C. Zody, and E. S. Lander. 2002. Initial sequencing and comparative analysis of the mouse genome. Nature 420: 520-562

222. Pabst, R., and I. Gehrke. 1990. Is the bronchus-associated lymphoid tissue (BALT) an integral structure of the lung in normal mammals, including humans? Am. J Respir. Cell Mol. Biol. 3: 131-135

223. Doeing, D. C., J. L. Borowicz, and E. T. Crockett. 2003. Gender dimorphism in differential peripheral blood leukocyte counts in mice using cardiac, tail, foot, and saphenous vein puncture methods. BMC Clin. Pathol. 3: 3

224. Risso, A. 2000. Leukocyte antimicrobial peptides: multifunctional effector molecules of innate immunity. J Leukoc. Biol. 68: 785-792

225. Bogdan, C. 2001. Nitric oxide and the immune response. Nat. Immunol. 2: 907-916

226. Koppel, E. A., K. P. van Gisbergen, T. B. Geijtenbeek, and Y. Van Kooyk. 2005. Distinct functions of DC-SIGN and its homologues L-SIGN (DC-SIGNR) and mSIGNR1 in pathogen recognition and immune regulation. Cell Microbiol. 7: 157-165

162

227. Svajger, U., M. Anderluh, M. Jeras, and N. Obermajer. 2010. C-type lectin DC-SIGN: an adhesion, signalling and antigen-uptake molecule that guides dendritic cells in immunity. Cell Signal. 22: 1397-1405

228. Monteiro, R. C., and J. G. van De Winkel. 2003. IgA Fc receptors. Annu. Rev. Immunol. 21: 177-204

229. Daeron, M. 1997. Fc receptor biology. Annu. Rev. Immunol. 15: 203-234

230. Snapper, C. M., and Finkelman F.D. 1999. Immunoglobulin class switching. In Fundamental Immunology. p. 831.

231. Minegishi, Y., J. Rohrer, E. Coustan-Smith, H. M. Lederman, R. Pappu, D. Campana, A. C. Chan, and M. E. Conley. 1999. An essential role for BLNK in human B cell development. Science 286: 1954-1957

232. Tokugawa, Y., M. Koyama, and J. Silver. 1997. A molecular basis for species differences in Thy-1 expression patterns. Mol. Immunol. 34: 1263-1272

233. Farrar, J. D., Smith J.D., Murphy T.L., Leung S., Stark G.R., and Murphy K.M. 2000. Selective loss of type I interferon-induced STAT4 activation caused by a minisatellite insertion in mouse Stat2. Nature Immunology

234. Bekpen, C., J. P. Hunn, C. Rohde, I. Parvanova, L. Guethlein, D. M. Dunn, E. Glowalla, M. Leptin, and J. C. Howard. 2005. The interferon-inducible p47 (IRG) GTPases in vertebrates: loss of the cell autonomous resistance mechanism in the human lineage. Genome Biol. 6: R92

235. Singh, S. B., A. S. Davis, G. A. Taylor, and V. Deretic. 2006. Human IRGM induces autophagy to eliminate intracellular mycobacteria. Science 313: 1438-1441

236. WHO. 2006. Clinical Aspects of HIV/AIDS: Pulmonary Manifestations.

237. Fenton, M. J., L. W. Riley, and L. S. Schlesinger. 2005. Receptor-Mediated Recognition of Mycobacterium tuberculosis by Host Cells. In Tuberculosis and the Tubercle Bacillus. S. T. Cole, K. D. Eisenach, D. N. McMurray, and W. R. Jacobs, Jr., eds. ASM Press, New York, pp. 405.

238. Kang, P. B., A. K. Azad, J. B. Torrelles, T. M. Kaufman, A. Beharka, E. Tibesar, L. E. Desjardin, and L. S. Schlesinger. 2005. The human macrophage mannose receptor directs Mycobacterium tuberculosis lipoarabinomannan-mediated phagosome biogenesis. J Exp. Med. 202: 987-999

239. Ting, J. P., J. A. Duncan, and Y. Lei. 2010. How the noninflammasome NLRs function in the innate immune system. Science 327: 286-290

163

240. Gutierrez, O., C. Pipaon, N. Inohara, A. Fontalba, Y. Ogura, F. Prosper, G. Nunez, and J. L. Fernandez-Luna. 2002. Induction of Nod2 in myelomonocytic and intestinal epithelial cells via nuclear factor-kappa B activation. J Biol Chem 277: 41701-41705

241. Inohara, N., and G. Nunez. 2003. NODs: intracellular proteins involved in inflammation and apoptosis. Nat. Rev. Immunol. 3: 371-382

242. Kobayashi, K. S., M. Chamaillard, Y. Ogura, O. Henegariu, N. Inohara, G. Nunez, and R. A. Flavell. 2005. Nod2-dependent regulation of innate and adaptive immunity in the intestinal tract. Science 307: 731-734

243. Herskovits, A. A., V. Auerbuch, and D. A. Portnoy. 2007. Bacterial ligands generated in a phagosome are targets of the cytosolic innate immune system. PLoS. Pathog. 3: e51

244. Kapetanovic, R., M. A. Nahori, V. Balloy, C. Fitting, D. J. Philpott, J. M. Cavaillon, and M. dib-Conquy. 2007. Contribution of phagocytosis and intracellular sensing for cytokine production by Staphylococcus aureus-activated macrophages. Infect. Immun. 75: 830-837

245. Kim, Y. G., J. H. Park, M. H. Shaw, L. Franchi, N. Inohara, and G. Nunez. 2008. The cytosolic sensors Nod1 and Nod2 are critical for bacterial recognition and host defense after exposure to Toll-like receptor ligands. Immunity. 28: 246-257

246. Till, A., P. Rosenstiel, K. Brautigam, C. Sina, G. Jacobs, H. H. Oberg, D. Seegert, T. Chakraborty, and S. Schreiber. 2008. A role for membrane-bound CD147 in NOD2-mediated recognition of bacterial cytoinvasion. J Cell Sci. 121: 487-495

247. Hruz, P., A. S. Zinkernagel, G. Jenikova, G. J. Botwin, J. P. Hugot, M. Karin, V. Nizet, and L. Eckmann. 2009. NOD2 contributes to cutaneous defense against Staphylococcus aureus through alpha-toxin-dependent innate immune activation. Proc. Natl. Acad. Sci. U. S. A 106: 12873-12878

248. Loving, C. L., M. Osorio, Y. G. Kim, G. Nunez, M. A. Hughes, and T. J. Merkel. 2009. Nod1/Nod2-mediated recognition plays a critical role in induction of adaptive immunity to anthrax after aerosol exposure. Infect. Immun. 77: 4529-4537

249. Strober, W., P. J. Murray, A. Kitani, and T. Watanabe. 2006. Signalling pathways and molecular interactions of NOD1 and NOD2. Nat. Rev. Immunol. 6: 9-20

250. Kufer, T. A. 2008. Signal transduction pathways used by NLR-type innate immune receptors. Mol. Biosyst. 4: 380-386

164

251. Ganguly, N., P. H. Giang, C. Gupta, S. K. Basu, I. Siddiqui, D. M. Salunke, and P. Sharma. 2008. Mycobacterium tuberculosis secretory proteins CFP-10, ESAT-6 and the CFP10:ESAT6 complex inhibit lipopolysaccharide-induced NF-kappaB transactivation by downregulation of reactive oxidative species (ROS) production. Immunol. Cell Biol. 86: 98-106

252. Simeone, R., D. Bottai, and R. Brosch. 2009. ESX/type VII secretion systems and their role in host-pathogen interaction. Curr. Opin. Microbiol. 12: 4-10

253. Schneemann, M., and G. Schoedon. 2002. Species differences in macrophage NO production are important. Nat. Immunol. 3: 102

254. Mestas, J., and C. C. Hughes. 2004. Of Mice and Not Men: Differences between Mouse and Human Immunology. J. Immunol. 172: 2731-2738

255. Powlesland, A. S., E. M. Ward, S. K. Sadhu, Y. Guo, M. E. Taylor, and K. Drickamer. 2006. Widely divergent biochemical properties of the complete set of mouse DC-SIGN-related proteins. J. Biol. Chem. 281: 20440-20449

256. Schneemann, M., and G. Schoeden. 2007. Macrophage biology and immunology: man is not a mouse. J. Leukoc. Biol. 81: 579

257. Horwitz, M. A., and S. C. Silverstein. 1980. Legionnaries' disease bacterium (Legionella pneumophila) multiplies intracellularly in human monocytes. J. Clin. Invest. 66: 441-450

258. Beharka, A. A., C. D. Gaynor, B. K. Kang, D. R. Voelker, F. X. McCormack, and L. S. Schlesinger. 2002. Pulmonary surfactant protein A up-regulates activity of the mannose receptor, a pattern recognition receptor expressed on human macrophages. J. Immunol. 169: 3565-3573

259. Brieland, J., P. Freeman, R. Kunkel, C. Chrisp, M. Hurley, J. Fantone, and C. Engleberg. 1994. Replicative Legionella pneumophila lung infection in intratracheally inoculated A/J mice. A murine model of human Legionnaires' disease. Am. J Pathol. 145: 1537-1546

260. Gavrilin, M. A., I. J. Bouakl, N. L. Knatz, M. D. Duncan, M. W. Hall, J. S. Gunn, and M. D. Wewers. 2006. Internalization and phagosome escape required for Francisella to induce human monocyte IL-1beta processing and release. Proc. Natl. Acad. Sci U. S. A 103: 141-146

261. Rajaram, M. V., L. P. Ganesan, K. V. Parsa, J. P. Butchar, J. S. Gunn, and S. Tridandapani. 2006. Akt/Protein kinase B modulates macrophage inflammatory response to Francisella infection and confers a survival advantage in mice. J Immunol. 177: 6317-6324

165

262. Olakanmi, O., B. E. Britigan, and L. S. Schlesinger. 2000. Gallium disrupts iron metabolism of mycobacteria residing within human macrophages. Infect. Immun. 68: 5619-5627

263. Nakagawara, A., and C. F. Nathan. 1983. A simple method for counting adherent cells: application to cultured human monocytes, macrophages and multi-nucleated giant cells. J. Immunol. Methods 56: 261-268

264. Karin, M., and Y. Ben-Neriah. 2000. Phosphorylation meets ubiquitination: the control of NF-[kappa]B activity. Annu. Rev. Immunol. 18: 621-663

265. Flesch, I. E. A., and S. H. E. Kaufmann. 1993. Role of cytokines in tuberculosis. Immunobiology 189: 316-339

266. Phiet, P. H., J. Wietzerbin, E. Zissman, J. F. Petit, and E. Lederer. 1976. Analysis of the cell wall of five strains of Myocbacterium tuberculosis BCG and of an attenuated human strain, W 115. Infect. Immun. 13: 677-681

267. Legrand-Poels, S., G. Kustermans, F. Bex, E. Kremmer, T. A. Kufer, and J. Piette. 2007. Modulation of Nod2-dependent NF-kappaB signaling by the actin cytoskeleton. J Cell Sci. 120: 1299-1310

268. Kleinnijenhuis, J., L. A. Joosten, D. Van, V, N. Savage, C. R. van, B. J. Kullberg, d. van, V, T. H. Ottenhoff, C. A. Dinarello, J. W. Van der Meer, and M. G. Netea. 2009. Transcriptional and inflammasome-mediated pathways for the induction of IL-1beta production by Mycobacterium tuberculosis. Eur. J Immunol. 39: 1914-1922

269. Ferwerda, G., M. Kramer, J. D. de, A. Piccini, L. A. Joosten, I. Devesaginer, S. E. Girardin, G. J. Adema, J. W. Van der Meer, B. J. Kullberg, A. Rubartelli, and M. G. Netea. 2008. Engagement of NOD2 has a dual effect on proIL-1beta mRNA transcription and secretion of bioactive IL-1beta. Eur. J Immunol. 38: 184-191

270. Pan, Q., J. Mathison, C. Fearns, V. V. Kravchenko, C. J. da Silva, H. M. Hoffman, K. S. Kobayashi, J. Bertin, E. P. Grant, A. J. Coyle, F. S. Sutterwala, Y. Ogura, R. A. Flavell, and R. J. Ulevitch. 2007. MDP-induced interleukin-1beta processing requires Nod2 and CIAS1/NALP3. J Leukoc. Biol. 82: 177-183

271. Boot, R. G., A. P. Bussink, M. Verhoek, P. A. de Boer, A. F. Moorman, and J. M. Aerts. 2005. Marked differences in tissue-specific expression of chitinases in mouse and man. J Histochem. Cytochem. 53: 1283-1292

272. Reginato, A. M., and B. R. Olsen. 2007. Genetics and experimental models of crystal-induced arthritis. Lessons learned from mice and men: is it crystal clear? Curr. Opin. Rheumatol. 19: 134-145

166

273. Rigamonti, E., G. Chinetti-Gbaguidi, and B. Staels. 2008. Regulation of macrophage functions by PPAR-alpha, PPAR-gamma, and LXRs in mice and men. Arterioscler. Thromb. Vasc. Biol. 28: 1050-1059

274. Amer, A. O. 2010. Modulation of caspases and their non-apoptotic functions by Legionella pneumophila. Cell Microbiol. 12: 140-147

275. Gavrilin, M. A., S. Mitra, S. Seshadri, J. Nateri, F. Berhe, M. W. Hall, and M. D. Wewers. 2009. Pyrin critical to macrophage IL-1beta response to Francisella challenge. J Immunol. 182: 7982-7989

276. Pauleau, A. L., and P. J. Murray. 2003. Role of nod2 in the response of macrophages to toll-like receptor agonists. Mol. Cell Biol. 23: 7531-7539

277. Leake, E. S., Q. N. Myrvik, and M. J. Wright. 1984. Phagosomal membranes of Mycobacterium bovis BCG-immune alveolar macrophages are resistant to disruption by Mycobacterium tuberculosis H37Rv. Infect. Immun. 45: 443-446

278. Teitelbaum, R., M. Cammer, M. L. Maitland, N. E. Freitag, J. Condeelis, and B. R. Bloom. 1999. Mycobacterial infection of macrophages results in membrane-permeable phagosomes. Proc. Natl. Acad. Sci USA 96: 15190-15195

279. Lecine, P., S. Esmiol, J. Y. Metais, C. Nicoletti, C. Nourry, C. McDonald, G. Nunez, J. P. Hugot, J. P. Borg, and V. Ollendorff. 2007. The NOD2-RICK complex signals from the plasma membrane. J Biol. Chem. 282: 15197-15207

280. Korhonen, R., A. Lahti, H. Kankaanranta, and E. Moilanen. 2005. Nitric oxide production and signaling in inflammation. Curr. Drug Targets. Inflamm. Allergy 4: 471-479

281. Moreno, L., S. K. McMaster, T. Gatheral, L. K. Bailey, L. S. Harrington, N. Cartwright, P. C. Armstrong, T. D. Warner, M. Paul-Clark, and J. A. Mitchell. 2010. Nucleotide oligomerization domain 1 is a dominant pathway for NOS2 induction in vascular smooth muscle cells: comparison with Toll-like receptor 4 responses in macrophages. Br. J Pharmacol. 160: 1997-2007

282. Totemeyer, S., M. Sheppard, A. Lloyd, D. Roper, C. Dowson, D. Underhill, P. Murray, D. Maskell, and C. Bryant. 2006. IFN-gamma enhances production of nitric oxide from macrophages via a mechanism that depends on nucleotide oligomerization domain-2. J Immunol. 176: 4804-4810

283. Bansal, K., and K. N. Balaji. 2011. Intracellular pathogen sensor NOD2 programs macrophages to trigger Notch1 activation. J Biol. Chem 286: 5823-5835

167

284. MacMicking, J. D., R. J. North, R. LaCourse, J. S. Mudgett, S. K. Shah, and C. F. Nathan. 1997. Identification of nitric oxide synthase as a protective locus against tuberculosis. Proc. Natl. Acad. Sci. USA 94: 5243-5248

285. Nims, R. W., J. C. Cook, M. C. Krishna, D. Christodoulou, C. M. Poore, A. M. Miles, M. B. Grisham, and D. A. Wink. 1996. Colorimetric assays for nitric oxide and nitrogen oxide species formed from nitric oxide stock solutions and donor compounds. Methods Enzymol. 268: 93-105

286. Cameron, M. L., D. L. Granger, J. B. Weinberg, W. J. Kozumbo, and H. S. Koren. 1990. Human alveolar and peritoneal macrophages mediate fungistasis independently of L-arginine oxidation to nitrite or nitrate. Am. Rev. Respir. Dis. 142: 1313-1319

287. Weinberg, J. B., M. A. Misukonis, P. J. Shami, S. N. Mason, D. L. Sauls, W. A. Dittman, E. R. Wood, G. K. Smith, B. McDonald, K. E. Bachus, A. F. Haney, and D. L. Granger. 1995. Human mononuclear phagocyte inducible nitric oxide synthase (iNOS): Analysis of iNOS mRNA, iNOS protein, biopterin, and nitric oxide production by blood monocytes and peritoneal macrophages. Blood 86: 1184-1195

288. Sakai, N., and S. Milstien. 1993. Availability of tetrahydrobiopterin is not a factor in the inability to detect nitric oxide production by human macrophages. Biochem. Biophys. Res. Commun. 193: 378-383

289. Jagannath, C., J. K. Actor, and R. L. Hunter, Jr. 1998. Induction of nitric oxide in human monocytes and monocyte cell lines by Mycobacterium tuberculosis. Nitric. Oxide. 2: 174-186

290. Choi, H. S., P. R. Rai, H. W. Chu, C. Cool, and E. D. Chan. 2002. Analysis of nitric oxide synthase and nitrotyrosine expression in human pulmonary tuberculosis. Am. J Respir. Crit Care Med. 166: 178-186

291. Nicholson, S., M. da Gloria, M. D. G. Bonecini-Almeida, J. R. L. E. Silva, C. Nathan, Q. W. Xie, R. Mumford, J. R. Weidner, J. Calaycay, J. Geng, N. Boechat, C. Linhares, W. Rom, and J. L. Ho. 1996. Inducible nitric oxide synthase in pulmonary alveolar macrophages from patients with tuberculosis. J. Exp. Med. 183: 2293-2302

292. Schlesinger, L. S., and M. A. Horwitz. 1991. Phagocytosis of Mycobacterium leprae by human monocyte-derived macrophages is mediated by complement receptors CR1(CD35), CR3(CD11b/CD18), and CR4(CD11c/CD18) and interferon gamma activation inhibits complement receptor function and phagocytosis of this bacterium. J. Immunol. 147: 1983-1994

168

293. Sheng, J. Z., D. Wang, and A. P. Braun. 2005. DAF-FM (4-amino-5-methylamino-2',7'-difluorofluorescein) diacetate detects impairment of agonist-stimulated nitric oxide synthesis by elevated glucose in human vascular endothelial cells: reversal by vitamin C and L-sepiapterin. J Pharmacol. Exp. Ther. 315: 931-940

294. Howard, M., J. Roux, H. Lee, B. Miyazawa, J. W. Lee, B. Gartland, A. J. Howard, M. A. Matthay, M. Carles, and J. F. Pittet. 2010. Activation of the stress protein response inhibits the STAT1 signalling pathway and iNOS function in alveolar macrophages: role of Hsp90 and Hsp70. Thorax 65: 346-353

295. Rich, E. A., M. Torres, E. Sada, C. K. Finegan, B. D. Hamilton, and Z. Toossi. 1997. Mycobacterium tuberculosis (MTB)-stimulated production of nitric oxide by human alveolar macrophages and relationship of nitric oxide production to growth inhibition of MTB. Tuber. Lung Dis. 78: 247-255

296. Cooney, R., J. Baker, O. Brain, B. Danis, T. Pichulik, P. Allan, D. J. Ferguson, B. J. Campbell, D. Jewell, and A. Simmons. 2010. NOD2 stimulation induces autophagy in dendritic cells influencing bacterial handling and antigen presentation. Nat. Med. 16: 90-97

297. Travassos, L. H., L. A. Carneiro, M. Ramjeet, S. Hussey, Y. G. Kim, J. G. Magalhaes, L. Yuan, F. Soares, E. Chea, B. L. Le, I. G. Boneca, A. Allaoui, N. L. Jones, G. Nunez, S. E. Girardin, and D. J. Philpott. 2010. Nod1 and Nod2 direct autophagy by recruiting ATG16L1 to the plasma membrane at the site of bacterial entry. Nat. Immunol. 11: 55-62

298. Homer, C. R., A. L. Richmond, N. A. Rebert, J. P. Achkar, and C. McDonald. 2010. ATG16L1 and NOD2 interact in an autophagy-dependent antibacterial pathway implicated in Crohn's disease pathogenesis. Gastroenterology 139: 1630-41, 1641

299. Tattoli, I., L. H. Travassos, L. A. Carneiro, J. G. Magalhaes, and S. E. Girardin. 2007. The Nodosome: Nod1 and Nod2 control bacterial infections and inflammation. Semin. Immunopathol. 29: 289-301

300. Chan, J., K. Tanaka, D. Carroll, J. Flynn, and B. R. Bloom. 1995. Effects of nitric oxide synthase inhibitors on murine infection with Mycobacterium tuberculosis. Infect. Immun. 63: 736-740

301. Misko, T. P., R. J. Schilling, D. Salvemini, W. M. Moore, and M. G. Currie. 1993. A fluorometric assay for the measurement of nitrite in biological samples. Anal. Biochem. 214: 11-16

169

302. Rhoades, E. R., and I. M. Orme. 1997. Susceptibility of a panel of virulent strains of Mycobacterium tuberculosis to reactive nitrogen intermediates. Infect. Immun. 65: 1189-1195

303. Yu, K., C. Mitchell, Y. Xing, R. S. Magliozzo, B. R. Bloom, and J. Chan. 1999. Toxicity of nitrogen oxides and related oxidants on mycobacteria: M. tuberculosis is resistant to peroxynitrite anion. Tuber. Lung Dis. 79: 191-198

304. Bryk, R., P. Griffin, and C. Nathan. 2000. Peroxynitrite reductase activity of bacterial peroxiredoxins. Nature 407: 211-215

305. Ehrt, S., M. U. Shiloh, J. Ruan, M. Choi, S. Gunzburg, C. Nathan, Q. Xie, and L. W. Riley. 1997. A novel antioxidant gene from Mycobacterium tuberculosis. J Exp. Med. 186: 1885-1896

306. Nathan, C., and M. U. Shiloh. 2000. Reactive oxygen and nitrogen intermediates in the relationship between mammalian hosts andmicrobial pathogens. Proc. Natl. Acad. Sci USA 97: 8841-8848

307. Kuo, H. P., C. H. Wang, K. S. Huang, H. C. Lin, C. T. Yu, C. Y. Liu, and L. C. Lu. 2000. Nitric oxide modulates interleukin-1beta and tumor necrosis factor-alpha synthesis by alveolar macrophages in pulmonary tuberculosis. Am. J Respir. Crit Care Med. 161: 192-199

308. Balligand, J. L., D. Ungureanu-Longrois, W. W. Simmons, D. Pimental, T. A. Malinski, M. Kapturczak, Z. Taha, C. J. Lowenstein, A. J. Davidoff, R. A. Kelly, and . 1994. Cytokine-inducible nitric oxide synthase (iNOS) expression in cardiac myocytes. Characterization and regulation of iNOS expression and detection of iNOS activity in single cardiac myocytes in vitro. J Biol. Chem 269: 27580-27588

309. Adams, V., B. Nehrhoff, U. Spate, A. Linke, P. C. Schulze, A. Baur, S. Gielen, R. Hambrecht, and G. Schuler. 2002. Induction of iNOS expression in skeletal muscle by IL-1beta and NFkappaB activation: an in vitro and in vivo study. Cardiovasc. Res. 54: 95-104

310. Chan, E. D., J. Chan, and N. W. Schluger. 2001. What is the role of nitric oxide in murine and human host defense against tuberculosis? Current knowledge. Am. J. Respir. Cell Mol. Biol. 25: 606-612

311. Nozaki, Y., Y. Hasegawa, S. Ichiyama, I. Nakashima, and K. Shimokata. 1997. Mechanism of nitric oxide-dependent killing of Mycobacterium bovis BCG in human alveolar macrophages. Infect. Immun. 65: 3644-3647

312. Toshchakov, V., B. W. Jones, P. Y. Perera, K. Thomas, M. J. Cody, S. Zhang, B. R. Williams, J. Major, T. A. Hamilton, M. J. Fenton, and S. N. Vogel. 2002.

170

TLR4, but not TLR2, mediates IFN-beta-induced STAT1alpha/beta-dependent gene expression in macrophages. Nat. Immunol. 3: 392-398

313. Thoma-Uszynski, S., S. Stenger, O. Takeuchi, M. Ochoa, M. Engele, P. Sieling, P. Barnes, M. Rollinghoff, P. Bolcskei, M. Wagner, S. Akira, M. Norgard, J. Belisle, P. Godowski, B. Bloom, and R. Modlin. 2001. Induction of direct antimicrobial activity through mammalian Toll-like receptors. Science 291: 1544-1547

314. Kimura, H., S. Miura, T. Shigematsu, N. Ohkubo, Y. Tsuzuki, I. Kurose, H. Higuchi, Y. Akiba, R. Hokari, M. Hirokawa, H. Serizawa, and H. Ishii. 1997. Increased nitric oxide production and inducible nitric oxide synthase activity in colonic mucosa of patients with active ulcerative colitis and Crohn's disease. Dig. Dis. Sci. 42: 1047-1054

315. Grisham, M. B., K. Ware, H. E. Gilleland, Jr., L. B. Gilleland, C. L. Abell, and T. Yamada. 1992. Neutrophil-mediated nitrosamine formation: role of nitric oxide in rats. Gastroenterology 103: 1260-1266

316. Grisham, M. B., R. D. Specian, and T. E. Zimmerman. 1994. Effects of nitric oxide synthase inhibition on the pathophysiology observed in a model of chronic granulomatous colitis. J Pharmacol. Exp. Ther. 271: 1114-1121

317. Shaul, P. W., A. J. North, L. C. Wu, L. B. Wells, T. S. Brannon, K. S. Lau, T. Michel, L. R. Margraf, and R. A. Star. 1994. Endothelial nitric oxide synthase is expressed in cultured human bronchiolar epithelium. J Clin. Invest 94: 2231-2236

318. Shames, B. D., C. H. Selzman, E. J. Pulido, X. Meng, D. R. Meldrum, R. C. McIntyre, Jr., A. H. Harken, and A. Banerjee. 1999. LPS-Induced NF-kappaB activation and TNF-alpha release in human monocytes are protein tyrosine kinase dependent and protein kinase C independent. J Surg. Res. 83: 69-74

319. Wadsworth, S. J., and H. Goldfine. 2002. Mobilization of protein kinase C in macrophages induced by Listeria monocytogenes affects its internalization and escape from the phagosome. Infect. Immun. 70: 4650-4660

320. Schwegmann, A., R. Guler, A. J. Cutler, B. Arendse, W. G. Horsnell, A. Flemming, A. H. Kottmann, G. Ryan, W. Hide, M. Leitges, C. Seoighe, and F. Brombacher. 2007. Protein kinase C delta is essential for optimal macrophage-mediated phagosomal containment of Listeria monocytogenes. Proc. Natl. Acad. Sci. U. S. A 104: 16251-16256

321. Brandt, S., S. Wessler, R. Hartig, and S. Backert. 2009. Helicobacter pylori activates protein kinase C delta to control Raf in MAP kinase signalling: role in

171

AGS epithelial cell scattering and elongation. Cell Motil. Cytoskeleton 66: 874-892

322. Sumaroka, M. V., I. S. Litvinov, S. V. Khaidukov, T. N. Golovina, M. V. Kamraz, R. L. Komal'eva, T. M. Andronova, E. A. Makarov, V. A. Nesmeyanov, and V. T. Ivanov. 1991. Muramyl peptide-binding sites are located inside target cells. FEBS Lett. 295: 48-50

323. von, K. O., S. Lipinski, A. Till, S. J. Martin, W. Nietfeld, H. Lehrach, S. Schreiber, and P. Rosenstiel. 2010. Caspase recruitment domain-containing protein 8 (CARD8) negatively regulates NOD2-mediated signaling. J Biol. Chem 285: 19921-19926

324. Lipinski, S., A. Till, C. Sina, A. Arlt, H. Grasberger, S. Schreiber, and P. Rosenstiel. 2009. DUOX2-derived reactive oxygen species are effectors of NOD2-mediated antibacterial responses. J Cell Sci. 122: 3522-3530

325. Rigutto, S., C. Hoste, H. Grasberger, M. Milenkovic, D. Communi, J. E. Dumont, B. Corvilain, F. Miot, and D. De, X. 2009. Activation of dual oxidases Duox1 and Duox2: differential regulation mediated by camp-dependent protein kinase and protein kinase C-dependent phosphorylation. J Biol. Chem 284: 6725-6734

326. Voss, O. H., S. Kim, M. D. Wewers, and A. I. Doseff. 2005. Regulation of monocyte apoptosis by the protein kinase Cdelta-dependent phosphorylation of caspase-3. J Biol. Chem 280: 17371-17379

327. Monick, M. M., A. B. Carter, G. Gudmundsson, L. J. Geist, and G. W. Hunninghake. 1998. Changes in PKC isoforms in human alveolar macrophages compared with blood monocytes. Am. J. Physiol. Lung Cell. Mol. Physiol. 19: L389-L397

328. Meller, N., Y. Elitzur, and N. Isakov. 1999. Protein kinase C-theta (PKCtheta) distribution analysis in hematopoietic cells: proliferating T cells exhibit high proportions of PKCtheta in the particulate fraction. Cell Immunol. 193: 185-193

329. Mellor, H., and P. J. Parker. 1998. The extended protein kinase C superfamily. Biochem. J 332 ( Pt 2): 281-292

330. Thome, M., K. Hofmann, K. Burns, F. Martinon, J. L. Bodmer, C. Mattmann, and J. Tschopp. 1998. Identification of CARDIAK, a RIP-like kinase that associates with caspase-1. Curr. Biol. 8: 885-888

331. McCarthy, J. V., J. Ni, and V. M. Dixit. 1998. RIP2 is a novel NF-kappaB-activating and cell death-inducing kinase. J Biol. Chem 273: 16968-16975

172

332. Sarkar, A., M. Duncan, J. Hart, E. Hertlein, D. C. Guttridge, and M. D. Wewers. 2006. ASC directs NF-kappaB activation by regulating receptor interacting protein-2 (RIP2) caspase-1 interactions. J Immunol. 176: 4979-4986

333. Inohara, N., P. L. del, T. Koseki, S. Chen, and G. Nunez. 1998. RICK, a novel protein kinase containing a caspase recruitment domain, interacts with CLARP and regulates CD95-mediated apoptosis. J Biol. Chem 273: 12296-12300

334. Meylan, E., F. Martinon, M. Thome, M. Gschwendt, and J. Tschopp. 2002. RIP4 (DIK/PKK), a novel member of the RIP kinase family, activates NF-kappa B and is processed during apoptosis. EMBO Rep. 3: 1201-1208

335. Chen, L., K. Haider, M. Ponda, A. Cariappa, D. Rowitch, and S. Pillai. 2001. Protein kinase C-associated kinase (PKK), a novel membrane-associated, ankyrin repeat-containing protein kinase. J Biol. Chem 276: 21737-21744

336. Dempsey, P. W., S. A. Vaidya, and G. Cheng. 2003. The art of war: Innate and adaptive immune responses. Cell Mol. Life Sci. 60: 2604-2621

337. Kobzik, L. 2009. Translating NO biology into clinical advances: still searching for the right dictionary? Am. J Respir. Cell Mol. Biol. 41: 9-13

338. Tigno-Aranjuez, J. T., J. M. Asara, and D. W. Abbott. 2010. Inhibition of RIP2's tyrosine kinase activity limits NOD2-driven cytokine responses. Genes Dev. 24: 2666-2677

339. Kelsall, B. 2005. Getting to the guts of NOD2. Nat. Med. 11: 383-384

340. Dzierzbicka, K., A. Wardowska, and P. Trzonkowski. 2011. Recent developments in the synthesis and biological activity of muramylpeptides. Curr. Med. Chem 18: 2438-2451

341. Wang, L. Z., X. L. Li, X. Y. Pang, and L. R. Sun. 2011. Muramyl Dipeptide modulates differentiation, maturity of dendritic cells and anti-tumor effect of DC-mediated T cell in acute leukemia children. Hum. Vaccin. 7: