transcriptional regulation and protective efficacy of

192
TRANSCRIPTIONAL REGULATION AND PROTECTIVE EFFICACY OF BORDETELLA COLONIZATION FACTOR A (BcfA) IN BORDETELLA INFECTIONS BY NEELIMA SUKUMAR A Dissertation Submitted to the Graduate Faculty of WAKE FOREST UNIVESITY GRADUATE SCHOOL OF ARTS AND SCIENCES in Partial Fulfillment of the Requirements for the Degree of DOCTOR OF PHILOSOPHY Microbiology and Immunology May 2009 Winston-Salem, North Carolina Approved By: Rajendar Deora, Ph.D., Advisor ....................................................... Examining Committee: Mark O. Lively, Ph.D., Chairman ....................................................... Griffith Parks, Ph.D. ....................................................... Purnima Dubey, Ph.D. ....................................................... Sean Reid, Ph.D. .......................................................

Upload: others

Post on 24-Mar-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

TRANSCRIPTIONAL REGULATION AND PROTECTIVE EFFICACY OF

BORDETELLA COLONIZATION FACTOR A (BcfA) IN BORDETELLA

INFECTIONS

BY

NEELIMA SUKUMAR

A Dissertation Submitted to the Graduate Faculty of WAKE FOREST UNIVESITY GRADUATE SCHOOL OF ARTS AND SCIENCES

in Partial Fulfillment of the Requirements

for the Degree of

DOCTOR OF PHILOSOPHY

Microbiology and Immunology

May 2009

Winston-Salem, North Carolina

Approved By: Rajendar Deora, Ph.D., Advisor ....................................................... Examining Committee: Mark O. Lively, Ph.D., Chairman ....................................................... Griffith Parks, Ph.D. ....................................................... Purnima Dubey, Ph.D. ....................................................... Sean Reid, Ph.D. .......................................................

ACKNOWLEDGEMENTS

Over the past four years, I have realized that the adage “If you love

what you do, then you don’t have to work a single day in your life” is absolutely

true. I enjoyed every single moment of my graduate school. And this is not only

because I loved what I did, but also due to the great people I could share this

part of my life’s journey with.

First of all, I would like to express my sincere gratitude to my

mentor Dr. Rajendar Deora for the painstaking guidance, unconditional support

and advices which have helped me through out my research. Above all I am

thankful for the confidence he had in me. It is an honor to have worked with him.

I express my sincere acknowledgements to my committee members; Drs.

Griffith Parks, Dan Wozniak, Mark Lively, Sean Reid and Purnima Dubey.

Despite their busy schedules, they always extended whole hearted help and

provided insightful suggestions.

None of my research projects would have been complete without the help

of my Lab mates- both past and current; Dr. Meenu Mishra, Dr. Gina Parise

Sloan, Cheraton Love and Matt Conover. They are great friends and I enjoyed

working with them and will cherish all the moments I spent with them both inside

and outside of the lab.

Without the love and blessings of my parents; R. Sukumarakurup & Dr.

Sujatha and Mohanan Nair & Padmini, I wouldn’t even be in a graduate school.

They exemplify hard work, patience and trust; features that aid one in graduate

school. Also I would like to thank my brother Dr. Krishna Kumar, sister-in-law Dr.

ii

Jyotsana Menon and my twin sister Poornima Sukumar whose unfailing

encouragement and motivation have given me the strength to face the trials and

tribulations of the graduate school.

Although, only the last two years of my graduate school was shared with

my soul mate, my husband, Dr. Sunish Mohanan, he taught me some of the

greatest lessons- to believe in myself and always aim for the best. Since life

presented me with the best partner, I was positive that science would also gift

me. I would like to thank him for his patience and motivation.

Above all, I thank God for making me take the right decisions and blessing

me with a wonderful graduate school experience.

iii

TABLE OF CONTENTS

LIST OF FIGURES ..............................................................................................vi

LIST OF TABLES..............................................................................................viii

LIST OF ABBREVIATIONS ................................................................................ ix

ABSTRACT ........................................................................................................xii

CHAPTER I: INTRODUCTION.............................................................................1

The Bordetella Genus. ......................................................................................1

Bordetella as a Pathogen. .................................................................................2

The BvgAS Signal Transduction System...........................................................7

Virulence Factors. ...........................................................................................12

Adhesins. .....................................................................................................12

Toxins. .........................................................................................................16

Animal Models.................................................................................................19

Immune Responses to B. pertussis and B. bronchiseptica. ............................21

Innate Immune Responses. .........................................................................22

B cell responses. .........................................................................................24

T cell responses...........................................................................................26

Currently Available Vaccines against B. pertussis...........................................27

Vaccines against B. bronchiseptica................................................................29

References. .....................................................................................................33

CHAPTER II ( Published in Journal of Bacteriology 2007 May; 189(10): 3695-

704).....................................................................................................................48

iv

Introduction .....................................................................................................49

Materials and Methods ....................................................................................52

Results ............................................................................................................63

Discussion.......................................................................................................91

References......................................................................................................96

CHAPTER III (Published in Infection and Immunity 2009 Feb;77(2);885-95)

..........................................................................................................................100

Introduction ...................................................................................................101

Materials and Methods ..................................................................................105

Results ..........................................................................................................113

Discussion.....................................................................................................140

References....................................................................................................147

CHAPTER IV: DISCUSSION............................................................................152

Potential roles of BipA and BcfA in mediating respiratory tract colonization........

by Bordetella. ................................................................................................153

BvgAS-mediated regulation of bcfA expression. ...........................................158

Differential phase specific expression profile of BipA and BcfA. ...................162

BcfA as a vaccine candidate against B. bronchiseptica. ...............................163

BcfA as a Th1 response inducing adjuvant. ..................................................165

BcfA as a vaccine candidate against B. pertussis. ........................................167

References. ...................................................................................................171

Curriculum Vitae............................................................................................174

v

LIST OF FIGURES

CHAPTER I

Figure 1 Clinical manifestation of whooping cough and the corresponding B. pertussis load in infected individuals..................... 4 Figure 2 The BvgAS two component system of Bordetella spp..................... 10

CHAPTER II

Figure 1 Schematic representation of the similarity of BcfA with BipA............65 Figure 2 Quantification of bcfA transcriptional activity by β-galactosidase

assays in the wt, Bvg+, Bvgi and Bvg- phase locked strains of B. bronchiseptica...................................................................................69

Figure 3 Determination of the phase-dependent expression profiles of different Bvg-regulated genes in B. bronchiseptica by real time RT-PCR analysis...............................................................................71 Figure 4 Kinetics of transcriptional activation of different Bvg-activated genes and the Bvg-independent gene recA......................................75 Figure 5 The putative promoter region of bcfA................................................78 Figure 6 Electrophoretic Mobility Shift Assay...................................................81

Figure 7 BcfA is localized to the outer-membrane...........................................84

Figure 8 BcfA is expressed during infection.....................................................86

Figure 9 Colonization of rat respiratory tract by wt and isogenic mutant derivatives RB25 (ΔbipA), RKD110 (ΔbcfA) and MM101 (ΔbipAΔbcfA)......................................................................................90

vi

CHAPTER III

Figure 1 Immunization with BcfA protects mice against B. bronchiseptica challenge.........................................................................................115

Figure 2 Immunization with BcfA reduces lung pathology in mice challenged with RB50......................................................................118 Figure 3 Anti-BcfA antibody titers in immunized mice....................................122 Figure 4 Effect of adoptive transfer of BcfA-specific sera on respiratory tract colonization..............................................................................127 Figure 5 Opsonization with anti-BcfA serum enhances the phagocytosis of RB50 by J774 murine macrophages............................................130 Figure 6 Neutrophils are required for anti-BcfA antibody-mediated clearance of B. bronchiseptica.........................................................133 Figure 7 BcfA-induced production of IFN-γ and IL-4 in splenocytes..............137 Figure 8 Expression of BcfA among clinical isolates of B. bronchiseptica.....139 Figure 9 Model for BcfA-mediated protective immunity.................................145

CHAPTER IV

Figure 1 Potential roles of BipA and BcfA in B. bronchiseptica pathogenesis...................................................................................157

Figure 2 Model illustrating relative occupancy of BvgA-P to putative bcfA

promoter..........................................................................................161 Figure 3 Expression of BcfA by B. pertussis strains......................................170

vii

LIST OF TABLES

CHAPTER II Table 1 Strains and plasmids used in this study..............................................53 Table 2 Oligonucleotide primers used in this study.........................................56 CHAPTER III Table 1 Strains used in this study..................................................................106 Table 2 BcfA immunization reduces lung pathology......................................120

viii

LIST OF ABBREVIATIONS

ADP………………………………………………...Adenine Diphosphate

AIDS………………………………………………..Acquired Immune Deficiency

Syndrome

ATP…………………………………………………AdenineTriphosphate

BcfA....................................................................Bordetella Colonization Factor A

BipA....................................................................Bordetella Intermediate Phase

Protein A

Bvg………………………………………………….Bordetella virulence gene

BrkA....................................................................Bordetella Resistance to Killing A

cAMP……………………………………………….Cyclic Adenine Monophosphate

CDC………………………………………………...Centers for Disease Control

CFU………………………………………………....Colony Forming Units

CR…………………………………………………..Complement Receptor

CyaA………………………………………………. Adenylate Cyclase Toxin

D…………………………………………………….Aspartic Acid

DNA....................................................................Deoxyribo Nucleic Acid

DNT…………………………………………………Dermonecrotic Toxin

DTP…………………………………………………Diptheria-Tetnus-Pertussis

EMSA………………………………………………Electrophoretic Mobility Shift Assay

FHA…………………………………………………Filamentous Hemagglutinin

Fim………………………………………………….Fimbriae

G........................................................................Guanine nucleotide

ix

GTP…………………………………………….......Guanosine Triphosphate

H…………………………………………………….Histidine

HPt………………………………………………….Histidine Phosphotransferase

IAP………………………………………………….Intergrin Associated Protein

ICAM……………………………………………….Intercellular Adhesion Molecule

ID.......................................................................Infectious Dose

IFN-γ..................................................................Interferon-gamma

Ig……………………………………………………Immunoglobulin

IL……………………………………………………Interleukin

ITB......................................................................Infectious Tracheobronchitis

LPS.....................................................................Lipopolysaccharide

LRI......................................................................Leukocyte Response Intergrin

MHC...................................................................Major Histocompactability Complex

MgSO4................................................................Magnesium Sulphate

NO......................................................................Nitric Oxide

ORF………………………………………………...Open Reading Frame

PBP....................................................................Periplasmic Binding Protein

Pc........................................................................Pertussis Acellular vaccine

PCR………………………………………………...Polymerase Chain Reaction

PMN………………………………………………..Polymorphonuclear Leukocytes

PRDC.................................................................Porcine Reproductive and

Respiratory Disease Complex

Prn………………………………………………….Pertactin

x

PT…………………………………………………..Pertussis Toxin

Pw.......................................................................Pertussis Whole cell vaccine

RGD...................................................................Arginine-Glycine-Aspartic Acid

RT-PCR……………………………………………Reverse Transcriptase-Polymerase

Chain reaction

RTX………………………………………………...Repeats in Toxin

SCID..................................................................Severe Combined Immune

Deficiency

Th......................................................................T-Helper

TCT..…………………………………………........Tracheal Cytotoxin

TNF....................................................................Tumor Necrosis Factor

TLR………………………………………………...Toll Like Receptors

Wt…………………………………………………..Wild type

xi

ABSTRACT

Sukumar, Neelima

TRANSCRIPTIONAL REGULATION AND PROTECTIVE EFFICACY OF

BORDETELLA COLONIZATION FACTOR A (BcfA) IN BORDETELLA INFECTIONS

To successfully colonize their mammalian hosts many bacteria produce

multiple virulence factors that play essential roles in disease processes and pathogenesis. Some of these molecules are adhesins that allow efficient attachment to host cells, a prerequisite for successful host colonization. Bordetella spp. express a number of proteins which either play a direct role in attachment to the respiratory epithelia or exhibit similarity to previously known bacterial adhesins. One such recently identified protein is BipA. Despite similarity to intimins and invasins, its deletion from B. bronchiseptica did not result in any significant defect in respiratory tract colonization. We hypothesized the existence of a paralogous protein that could complement the function of BipA. In the studies described here, we report the identification and characterization of an ORF in B. bronchiseptica, designated as bcfA (Bordetella colonization factor A) that is similar to bipA. We show that in contrast to maximal expression of bipA in the Bvgi phase, bcfA is expressed at high levels in both the Bvg+ and the Bvgi phases. We have identified multiple sequence elements resembling the consensus BvgA binding site in the bcfA promoter region. Direct binding of purified BvgA to the bcfA promoter revealed differences in the DNA binding profiles of BvgA and BvgA-P to the promoter region. Utilizing an antibody raised against BcfA, we show that BcfA is localized in the outer membrane. Finally, we demonstrate that simultaneous deletion of both bipA and bcfA results in a defect in colonization of the rat trachea and that BcfA is expressed during Bordetella infectious cycle.

Based on our findings that BcfA is an outer membrane immunogenic protein

and is critical for murine respiratory tract colonization, we examined its utility in inducing a protective immune response against B. bronchiseptica in a mouse model of intranasal infection. Mice vaccinated with BcfA demonstrated reduced pathology in the lungs and harbored lower bacterial burdens in the respiratory tract. Immunization with BcfA led to the generation of BcfA-specific antibodies in both the serum and the lungs and passive immunization led to the reduction of B. bronchiseptica in the trachea and the lungs. These results suggest that protection after immunization with BcfA is mediated in part by antibodies against BcfA. To further investigate the mechanism of BcfA-induced immune clearance, we examined the role of neutrophils and macrophages. Our results demonstrate that neutrophils are critical for anti-BcfA antibody-mediated clearance and that opsonization with anti-BcfA serum enhances phagocytosis of B. bronchiseptica by murine macrophages. We show that immunization with BcfA results in the production of IFN-γ and subclasses of IgG antibodies that are consistent with the induction of a Th1 type immune response. In

xii

combination, our findings suggest that mechanism of BcfA-mediated immunity involves humoral and cellular responses. Expression of BcfA is conserved among multiple clinical isolates of B. bronchiseptica. Our results demonstrate the striking protective efficacy of BcfA-mediated immunization thereby highlighting its utility as a potential vaccine candidate. These results also provide a model for the development of cell-free vaccines against B. bronchiseptica.

xiii

CHAPTER I: INTRODUCTION

The Bordetella Genus.

Bordetellae are Gram- negative aerobic cocobaccilli that preferentially attach

to the ciliated respiratory epithelium of mammals (95). Currently there are ten known

Bordetella species. Among these, B. bronchiseptica (44), B. pertussis (12) and B.

parapertussis (40) are the most well-studied and are known as the classical species.

Other known species include B. avium, B. hinzii, B. holmesii, B. trematum, B. ansorpii

and B. petrii. B. bronchiseptica causes respiratory infections in a wide range of four

legged animals such dogs, cats, pigs, horses, non human primates (49) and

occasionally humans (151). Unlike B. bronchiseptica, B. pertussis has a restricted

host range and only infects humans (17). B. parapertussis exists in two host adapted

subspecies. B. parapertussishu is associated with respiratory illness in humans while

B. parapertussisov is ovine adapted and causes chronic respiratory infections in

sheep (60, 116). Based on a combination of different phylogenetic analysis tools such

as multilocus enzyme electrophoresis, insertion sequence polymorphisms and

genome sequencing, it is predicted that B. bronchiseptica is the ancestral organism

from which B. pertussis and B. parapertussis evolved independently into host

restricted species (35, 46, 114, 142). B. avium causes infections in poultry and is also

an opportunistic human pathogen (128, 134). Although majority of the Bordetella spp

are associated with respiratory tract infections, there are multiple species that have

variant pathogenic and survival characteristics. B. trematum has never been

associated with the respiratory tract, but has been isolated from ear and skin wound

1

infections (143). On the other hand, another unique species, B. petrii was isolated

from the environment and is considered to be the progenitor from which other

pathogenic Bordetella species evolved (144, 145). The range of differences in host

adaptation, virulence traits and survival niches highlight the versatility of the

Bordetella genus as a pathogen.

Bordetella as a Pathogen.

B. pertussis is the etiological agent of the highly communicable disease called

pertussis or whooping cough. This illness is most severe in children and can result in

mortality (52). Despite the availability and widespread use of vaccines, World Health

Organization estimates 48.5 million cases and 294,000 deaths worldwide resulting

from this disease in 2002. According to the Centers of Disease Control (CDC) in the

United States, more than 25,000 cases of pertussis were reported in the year 2004,

which is the highest incidence since 1959. In addition to its impact on infants and

young children, there is increasing prevalence of pertussis in adults and adolescents

(64). Pertussis is circulating and is highly ubiquitous in adults and adolescents of all

ages, even in previously infected or vaccinated populations. In fact, in the United

States, multiple reports estimate 1 million pertussis cases each year in this age group

(21). Moreover, case studies conducted in different countries including United States

reveal that more than 20% of adolescents and adults having prolonged cough illness

are colonized by B. pertussis (47, 109, 125). Although this disease is not as severe in

this age group, the significance of increasing incidence lies in the fact that

2

FIG. 1. Clinical manifestation of whooping cough and the corresponding B. pertussis

load in infected individuals. Following the 7-10 incubation period, the catarrhal phase

is manifested and is characterized by common cold-like symptoms. It is during this

stage that the bacterial numbers reach a peak and infected persons are at the

highest risk of transmission to uninfected individuals. The paroxysmal stage is

distinguished by the characteristic whooping cough and can last from 2 to 6 weeks.

The bacterial load begins to decline in this stage making the clinical diagnosis

difficult. The convalescent phase is the recovery stage. This phase is characterized

by the reduced incidence and severity of paroxysms and can last from 3 to 4 weeks

or upto several months.

3

Phases

Catarrhal

Paroxsymal

Convalescent

Bac

teria

l Num

bers

Incubation

3-4 weeks

or longer

7-10 days

1-2 weeks

2-4 weeks

Duration

4

adolescents and adults act as source for B. pertussis infections in partially immune

and nonimmune children especially in household settings.

The clinical manifestation of this disease occurs in three phases- catarrhal,

paroxysmal and convalescent (19). Following the 7-10 days incubation period, the

first stage of infection, the catarrhal phase, is manifested with runny nose, sneezing,

low grade fever and occasional mild cough (Fig. 1). These symptoms are similar to

the common cold or minor respiratory tract infections and thus pertussis is often

ignored during the initial diagnosis. This stage lasts for 1-2 weeks after which the

cough gradually attains severity and develops into the distinctive whooping cough in

the paroxysmal stage. The paroxysmal cough is characterized by bursts or

paroxysms of rapid, numerous (5-10), forceful coughs in a single expiration followed

by a massive inhalation, producing the characteristic whoop sound. Concurrent with

the paroxysms, cyanosis, eye bulging and posttussive vomiting can occur. The

paroxysmal stage can last 1-6 weeks and can even persist upto 10 weeks. It is during

the paroxysmal stage that pertussis is usually suspected; however the bacterial loads

begin to decline making the clinical diagnosis difficult. Recovery occurs during the

convalescent phase and is characterized by the gradual reduction in the incidence

and severity of paroxysms. The cough may completely disappear by 2-3 weeks or it

can take upto several months. Complications are common and include pneumonia,

otitis media, seizures and encephalopathy (95). In adults and adolescents and

partially immunized children, pertussis is manifested as a milder disease or may be

asymptomatic. B. pertussis infections in these persons may result in mild or

persistent, severe cough without the distinguishing whoop (37, 87, 88).

5

B. parapertussishu infections can result in mild pertussis like illness or severe

classic pertussis (10, 41). Coexistence and contributions of B. parapertussis in

pertussis infections may be underestimated due to difficulties in distinguishing

between the two human adapted species (68, 76).

B. bronchiseptica is the primary etiological cause and/or pre disposing factor

in a variety of veterinary infections such as Porcine Reproductive and Respiratory

Disease complex (PRDC), pneumonia and atrophic rhinitis in swine, Infectious

tracheobronchitis (ITB, Kennel Cough) in dogs and bronchopneumonia in guinea

pigs, rats, mice, rabbits, cats and non human primates (16, 95). Upper respiratory

illness by B. bronchiseptica in infant pigs is characterized by coughing and sneezing

followed by deformation of the nose bony structures leading to atrophy. In swine

populations, this disease is widely prevalent and results in loss of 17 million dollars

annually in the United States. Moreover, infection with B. bronchiseptica predisposes

pigs to infection by other bacterial and viral pathogens thus contributing to a multi-

factorial disease condition PRDC, which has an annual economic impact of 40 million

dollars in the United States. In dogs, the highly contagious ITB can manifest as mild

illness or severe cough with pneumonia. Although vaccination can considerably

reduce the severity of disease, B. bronchiseptica is frequently isolated from nasal

cavities of immunized animals, suggesting persistence of this pathogen in animals.

Although mainly implicated as an animal pathogen, there have been many recent

reports of B. bronchiseptica infections in humans. The majority of human infections

by B. bronchiseptica occur in immuno-compromised individuals such as AIDS and

cystic fibrosis patients (39, 71, 135). However, it has been also isolated from an

6

immunocompetent individual (118). There are also several reports of zoonotic

transmission of this organism from farm or pet animals (151).

The BvgAS Signal Transduction System.

Bordetellae control the expression of majority of their known virulence genes

and other factors through the BvgAS signal transduction system. In general, bacteria

utilize complex signaling mechanisms for eliciting adaptive responses to enhance

survival in constantly changing environment. These signaling systems are devised to

detect fluctuations in the environment and to trigger subsequent changes in gene

expression. The most common regulatory mechanism utilized by bacteria is the two

component signal transduction system consisting of a sensor kinase and a response

regulator. Typically, extracellular signals are processed by transfer of phosphate

group from histidine residue in the sensor to aspartate residue in the receiver (53).

The bvgAS locus encodes for a polydomain transmembrane sensor, BvgS and a

cytoplasmic response regulator, BvgA (28). This virulence regulatory system is 96 %

identical at the nucleotide level among B. pertussis, B. parapertussis and B.

bronchiseptica and is functionally interchangeable (94). However, BvgAS belongs to

a class of bacterial two component signal transduction systems, which deviates from

the biphasic paradigm in that communication between the transmembrane sensor

kinase (BvgS) and the cytoplasmic response regulator (BvgA) occurs via a

sophisticated four step His-Asp-His-Asp phosphorylation cascade (Fig. 2) (141).

Similar to other regulatory proteins, both BvgS and BvgA function as dimers. BvgS at

its N terminus contains two periplasmic binding protein domains (PBP) followed by

transmembrane linker region and cytoplasmically localized autokinase, receiver and

7

the C terminal histidine phosphotansferase (HPt) domains. In response to changes in

environmental signals, the autokinase domain catalyzes ATP hydrolysis and

phosphorylation of histidine (H) 729. The γ-phosphate moiety is subsequently

transferred to aspartic acid (D) 1023 on the receiver domain and then to H1172 of the

HPt domain. These intramolecular phosphotransfers eventually result in the

phosphorylation of BvgA (Fig. 2). BvgA is a typical response regulator composed of

N terminal receiver and C terminal DNA binding domains. Phosphorylation causes

dimerization of BvgA and enhances its capacity to bind Bvg-regulated promoters and

regulate transcription (9, 14, 127). This can either result in transcriptional activation or

repression of cognate genes based on the affinity and location of BvgA binding sites

with respect to the transcription initiation site (38).

A hallmark feature of the BvgAS signal transduction system is that it can cause

phenotypic transition of Bordetella among three known phases-Bvg+, Bvgi (Bvg-

intermediate) and Bvg- and potentially multiple unknown states (28). Bvg+ is the

virulent phase and is characterized by the expression of adhesins such as

filamentous hemagglutinin, fimbriae and pertactin and toxins such as adenylate

cyclase, dermonecrotic toxin and pertussis toxin in the case of B. pertussis, all

contributing to attachment and subsequent invasion. Bvg+ phase is sufficient for

establishing respiratory tract colonization (29). In the Bvg- phase, the avirulent stage,

BvgAS system is inactive and there is suppression of Bvg-activated loci and

expression of Bvg-repressed genes such as flagella (2, 3) and urease (99, 100) in B.

bronchiseptica. Bvgi phase is characterized by the expression of some of the Bvg+

phase factors and the maximal expression of a set of antigens of which only one-

8

FIG. 2. The BvgAS two component system of Bordetella spp. BvgS is the

transmembrane polydomain sensor kinase consisting of two periplasmic binding

protein domains, autokinase, receiver and histidine phosophotransferase domains.

When active, autokinase domain catalyzes the hydrolysis of ATP and initiates a

series of intramolecular phosphotransfers that eventually leads to the phosphorylation

of the cytoplasmic response regulator BvgA. Phosphorylated BvgA regulates

activation or repression of transcription of genes. When the BvgAS system is active,

Bordetella is in a Bvg+ phase and expresses factors with known or hypothesized role

in virulence such as adhesins (FHA, Fim, Prn) and toxins (CyaA, PT). Inactivation of

BvgAS system renders Bordetella to Bvg- phase. Under these conditions, Bvg-

activated factors are not expressed, but Bvg-repressed factors such as flagella and

urease are expressed. Bordetella also exsist in a Bvgi phase. Although the

phosphorylation status of BvgAS system is unknown in this phase, it is postulated

that BvgA-P is in an intermediate levels as compared to the Bvg+ and Bvg- phases.

The only known Bvgi phase- specific protein is BipA. In the laboratory, Bordetella can

be modulated between the different phases using nicotinic acid, MgSO4 or growth at

different temperature conditions.

9

P

BvgA

BvgS

Sulfate <30 C Nicotinic acid ·

Phosphotransfers

MotilityBvg

UreaseBipA

Bvgi

Bvg + Adenylate cyclase toxin

Pertactin Fimbriae

FHA

10

BipA has been identified at the molecular level (30, 138). In the Bvgi phase,

Bordetella exhibits decreased virulence and improved survival under nutrient limiting

conditions. Although the signals required for regulation of BvgAS in vivo have not yet

been identified, under laboratory conditions, modulation among these phases can be

achieved by varying temperature, sulphate anion or nicotinic acid concentrations (83)

or through specific mutations of the BvgAS system.

The third component of the bvg locus encodes for a regulatory protein called

BvgR. bvgR is located directly downstream of bvgAS and is transcribed convergently

(104). BvgR activity is induced by transcriptional activation of bvgR by activated BvgA

(103). BvgR is involved in the transcriptional repression of Bvg-repressed genes in

the Bvg+ phase. Deletion analysis revealed that the absence of BvgR-mediated

regulation of Bvg-repressed genes is detrimental for colonization by B. pertussis in a

mouse model of infection (105). This and other studies indicate that ectopic

expression of Bvg repressed genes can impede the development of infection (1,

105). Although the exact roles of Bvg repressed genes are unknown, it is

hypothesized that these factors are involved in survival outside the host and under

nutrient deprived conditions. Previous studies have demonstrated that Bvg-activated

factors are necessary and sufficient for colonization of the respiratory tract. On the

other hand, Bvgi is postulated to be important for transmission of bacteria, although

there is no experimental evidence for this hypothesis (138). These data therefore

suggests that a possible role for Bvg regulon is to sense whether the bacteria is

inside a host or outside and accordingly regulate expression of vital genes (28).

11

Virulence Factors. Bordetella spp produce an array of virulence determinants which

can be categorized into two broad groups- adhesins and toxins. Expression of

majority of these virulence associated factors is coordinately regulated by the BvgAS

system. Efficient colonization of the respiratory tract by Bordetellae requires the

interplay of multiple factors. The current literature provides evidence for unique

properties as well as synergistic and antagonistic roles for these factors in virulence

functions such as attachment, invasion, modulation and suppression of immune

responses. Furthermore, the presence of many of these antigens in currently

available vaccines, exemplify the importance of these factors in Bordetella

pathogenesis. Therefore, a better understanding the roles of these virulence

components in Bordetella-host interactions will contribute to the development of more

efficient therapeutic techniques as well as vaccines.

Adhesins. This section is a brief overview of the major adhesins expressed by

Bordetella spp along with their functions in pathogenesis.

Filamentous hemagglutinin (FHA).

One of the dominant adherence factors of Bordetellae is FHA; a large highly

immunogenic protein which is both surface associated and secreted in the

extracellular milieu (32, 33). FHA is a component of the currently available B.

pertussis acellular vaccines. In vitro studies utilizing different mammalian cell lines

indicate that FHA contains four binding domains that aid in attachment of Bordetella

to respiratory epithelium or host immune cells. The Arginine-Glycine-Aspartic Acid

12

(RGD) motif stimulates adherence to macrophages and leukocytes potentially

through the Complement Receptor Type 3 (CR3) and Leukocyte Response

Intergrin/Intergrin-Associated Protein (LRI/IAP) complex (74, 119). The RGD domain

has also been shown to facilitate attachment of Bordetella to bronchial epithelial cells

through interactions with very late antigen 5 (VLA-5) (75). This interaction results in

the upregulation of epithelial intercellular adhesion molecule 1 (ICAM-1) via the NFκB

signaling pathway. The carbohydrate recognition domain of FHA mediates adherence

to ciliated epithelial cells of the respiratory tract and macrophages (117). In addition,

FHA also has a heparin binding domain, which promotes binding to sulfated

carbohydrates (102). Finally FHA also possess a CR3 recognition motif, function for

which has not yet been identified. Furthermore, inhibition of CD4+ T cell proliferation

and induction of apoptosis by B. pertussis has been demonstrated to be FHA-

dependent (13). Studies using purified FHA revealed that this protein induces

immunosuppressive effects on macrophages and dendritic cells by downregulating

production of Interleukin (IL)-12 in an IL-10 dependent manner (97, 98).

In vivo studies utilizing mutant strains having in frame deletion of fha revealed

that FHA is essential for tracheal colonization by B. bronchiseptica (31). A role for

FHA in B. pertussis colonization has been more difficult to discern due to inconsistent

results. Utilizing mouse models, McGuirk et al. showed that in the absence of FHA, B.

pertussis showed defective colonization of lungs, while others found no difference

between the wild type strain (wt) and FHA mutant derivatives in their ability to

colonize the respiratory tract (50, 148). These conflicting results may be due to the

fact that mice are not natural hosts for B. pertussis and these models fail to represent

13

natural course of infection. Comparison of genome sequences from B. bronchiseptica

and B. pertussis indicates that these two closely related species encode FHA which

is similar but not identical. In fact, studies utilizing B. bronchiseptica strains

ectopically expressing B. pertussis FHA (FHA ) fail to colonize the rat trachea.

FHA could mediate attachment to epithelial cells in vitro but failed to protect B.

bronchiseptica from inflammation-mediated clearance (73). These studies highlight

the importance of FHA in modulating the immune responses and effecting successful

colonization and persistence of B. bronchiseptica.

Bp

Bp

Pertactin (Prn).

Prn is a Bvg-regulated surface associated protein belonging to the

autotransporter secretion system. Typically, these proteins direct their own export

across the outermembrane. The highly conserved C terminal β- barrel domain of

these proteins facilitates the transport of the N terminal passenger domain, which

confers the effector function (36). Prn contains the tripeptide RGD domain and is

proposed to function as an adhesin (42). However, in vitro studies comparing wt type

parent strain and mutants lacking Prn revealed no significant role for this protein in

adherence or invasion of HEp 2 cells (122). Additionaly prn- mutants of both B.

pertussis and B. bronchiseptica do not differ from their parental wt strains in the

ability to colonize and persist within the respiratory tracts in vivo. In contrast to the in

vitro and in vivo studies described above that fail to demonstrate a precise function

for this protein in Bordetella pathogenesis, data from several vaccine efficacy trials

conducted during the early 1990s demonstrate that antibodies against Prn are the

14

most vital to confer protection, suggesting a significant role in protective immunity

(20, 139). These vaccine efficacy evaluation studies have also suggested that

inclusion of Prn in acellular vaccines containing FHA and Pertussis toxin, augment its

efficacy in preventing B. pertussis infection (108). Moreover, a role for anti-Prn

antibodies in efficient phagocytosis of B. pertussis has been demonstrated (63).

Fimbriae (Fim).

Many Gram negative bacteria express filamentous, polymeric structures

localized to the cell surface called fimbriae (Fim). Bordetellae synthesize four fimbrial

serotypes including the predominantly expressed Fim2 and Fim3 and FimX and FimN

which are expressed only at low levels (78, 85, 111, 121). Several studies have

demonstrated a role for Fim in mediating adherence of Bordetella to the respiratory

epithelium and monocytes (58, 59). In vivo studies revealed that Fim is involved in

the efficient establishment of tracheal colonization and persistence (96). Fim is also

critical for production of appropriate serum antibody responses. Specifically, Fim aids

in eliciting an immunoglobulin (Ig) M response early during the infection as well as in

inducing the IgG2a component of host humoral immunity in a rat model of infection

(96). Furthermore, studies in mice showed that Fim is vital for inducing an anti-

inflammatory response and preventing killing of Bordetella by alveolar macrophages

(95). Fim is also a component of the accelluar vaccines and vaccine trials in children

suggested that antibody to Fim contributes to protective immunity. Moreover,

inclusion of Fim2/3 in vaccines has been shown to significantly enhance efficacy

(113).

15

Toxins. A different category of virulence determinants expressed by Bordetella spp

are toxins. The following section briefly discusses the toxins secreted by Bordetella

and their roles in bacterial invasion and host inflammatory responses.

Adenylate cyclase (CyaA).

CyaA is a bifunctional adenylate cyclase and hemolysin expressed by all three

classical species of Bordetella. The short 400 amino acid N terminal region confers

the catalytic activity of CyaA while the longer 1300 amino acid C terminal region

mediates the translocation of the catalytic domain into mammalian cell cytosol as well

as its function as a hemolysin (65). This calmodulin- sensitive toxin belongs to the

repeats in toxin (RTX) family of calcium dependent, pore forming cytotoxins (120,

124, 126). The receptor for CyaA has been identified as CD11b, a cell surface

glycoprotein, expressed by myeloid cells such as macrophages, dendritic cells,

neutrophils and natural killer cells. On entry into a mammalian cell, calmodulin

activates CyaA, which then catalyzes the enhanced conversion of cellular ATP to

cyclic AMP (cAMP). The supraphysiologic quantities of cAMP thus synthesized

interfere with intracellular signaling and result in altered cell physiology (23, 24, 150).

In many of the host immune effector cells, intoxication by cAMP results in inhibition of

bacteoricidal functions. Purified CyaA inhibits super oxide generation and chemotaxis

of peripheral blood monocytes and neutrophils. In addition, CyaA induces apoptosis

in macrophages and inhibits phagocytosis of B. pertussis by neutrophils (146, 147).

Furthermore studies using bone marrow derived dendritic cells revealed that CyaA

promoted upregulation of MHC class II and costimulatory molecule (CD80, CD83 and

16

CD86) expression and suppressed production of proinflammatory cytokines such as

IL-12 and Tumor Necrosis Factor (TNF)α (132). In accordance with its

immunomodulatory role in vitro, in vivo studies show that CyaA deficient mutants are

defective in causing lethal infection in infant mice (54, 148). These data thus suggest

an anti-inflammatory and anti-phagocytic role for CyaA in Bordetella infections.

Although not a component of acellular vaccines, studies has shown that

primary infection with B. pertussis induces anti-CyaA antibodies in children (22).

Moreover, studies utilizing convalescent serum samples revealed that antibodies

against CyaA promoted phagocytosis of B. pertussis by human neutrophils, thus

suggesting a role in infection-induced immunity (110).

Tracheal Cytotoxin (TCT).

TCT is a peptidoglycan derived disaccharide–tetrapeptide monomer

synthesized commonly by all gram-negative bacteria during growth and cell division.

Typically this monomer is recycled by recovering it back into the cell cytoplasm via a

membrane protein called AmpG (25, 26, 77). Bordetella spp. lack functional AmpG

and thus release this peptidoglycan fragment into the environment. The potency of

this toxin is reflected by the fact that TCT alone is sufficient and necessary to induce

specific cytopathology characteristic to B. pertussis infections in ciliated cells of

tracheal explants (48). TCT causes ciliostasis, cell blebbing and mitochondrial

damage. Additionally, TCT induces IL-1 α production and consequently nitric oxide

(NO) production in hamster tracheal epithelial cells (61, 62). TCT triggered NO

production is proposed to mediate annihilation of ciliated cells. In vivo, it is

17

hypothesized that TCT induces IL-1 production in non ciliated mucus secreting cells.

The resultant NO diffuses through the neighboring ciliated cells and causes severe

damage to the epithelium (45).

Dermonecrotic Toxin (DNT).

DNT is a heat labile A-B toxin consisting of an N terminal receptor-binding

domain and C terminal catalytic domain (34). The receptor for DNT has not yet been

identified. When intradermally injected, DNT induces necrotic lesions in laboratory

animals such as mice, guinea pigs and rabbits and is lethal for mice when delivered

intravenously (11, 72). In vitro studies utilizing purified DNT has shown that this toxin

stimulates modification and activation of the small GTP-binding protein Rho leading

to induction of DNA and protein synthesis, alterations to cell cytoskeleton and

inhibition of cell division (69, 70). DNT has also been associated with turbinate

atrophy and bronchopneumonia in B. bronchiseptica infected pigs (90, 123).

Pertussis Toxin (PT).

Unlike other toxins discussed above, PT is preferentially expressed only by B.

pertussis. Although B. bronchiseptica and B. parapertussis genomes contain genes

that can potentially express PT, mutations present in the promoters cause the genes

to be transcriptionally silent (4). PT is an ADP- ribosylating AB toxin composed of six

polypeptides designated as S1 to S5. The A subunit of the toxin comprises of S1

polypeptide and the B subunit is pentameric and is composed of S2, S3, S5 and two

S4 subunits (86, 112). S2 to S5 polypeptides form a ring like structure with S1 subunit

18

atop (79, 129). The B subunit mediates binding to eukaryotic cell membrane and

transports the enzymatically active S1 subunit into the cytoplasm, where it catalyzes

the transfer of ADP ribose to Guanine nucleotide (G) binding proteins. The ADP

ribosylation of different isoforms of G proteins leads to disruption of signal

transduction pathways within the cell (79). Studies have attributed both

immunomodulatory as well as an attachment function for PT (140). PT has been

suggested to mediate adherence of B. pertussis to ciliated epithelium of the

respiratory tract as well as immune cells such as macrophages. On the other hand,

PT has also been shown to have immunosuppressive effects. Studies revealed that

PT inhibits chemotaxis, oxidative responses and lysosome release in macrophages

and neutrophils (15, 101). In another study, a mutant strain lacking PT displayed

higher serum anti-Bordetella antibody responses as compared to the wt strain (18).

Nevertheless, PT is proposed to be the principal virulence factor responsible for all

the major pertussis-associated typical disease symptoms such as leukocytosis and

lymphocytosis and is a component of all the currently available acellular vaccines.

Animal Models.

Multiple animal models have been developed to gain insights into the function

of various virulence determinants and to study the progression of Bordetella

pathogenesis. For B. bronchiseptica, pathogen free rabbits, rats and mice are

commonly utilized. The infectious dose (ID)50 dose of this species for intranasal

inoculation is less than 200 Colony Forming Units (CFU) for rabbits, 25 CFU for rats

and 5 CFU for mice (95). Previous studies have shown that these model systems can

19

accurately simulate the characteristics of a natural infection by B. bronchiseptica. For

intranasal challenge either a high volume (25-50 μl) or low volume (5-10 μl) can be

used to deliver bacteria and both of these inoculation regimen lead to efficient

establishment and persistence in all animal models. In rats and mice, the nasal cavity

gets colonized persistently, while the trachea and larynx get cleared by 50-60 days

post inoculation with B. bronchiseptica (133). Apart from the upper respiratory tract,

the high volume treatment also consistently delivers bacteria to the lungs, which

eventually gets cleared by 50-70 days post inoculation (57).

Even though humans are the only known host for B. pertussis, a number of

animal species are used to study immune responses to this organism. In contrast to

B. bronchiseptica, for B. pertussis, a large infectious dose in a large volume is

commonly used to infect animals. The most prevalently used is the murine model of

intranasal or aerosol infection. Intranasal infection of mice with 5 x 105 CFU of

respective strains in 50 μl volume consistently and reproducibly delivers bacteria to

the nasal cavity, larynx, trachea and lungs (57). Unlike B. bronchiseptica, B. pertussis

does not persist for the life in mice and gets cleared from the respiratory tract by 20-

70 days post inoculation. Although mice do not display overt symptoms of the human

disease, the intranasal model replicates many of the attributes of pertussis: i) bacteria

rapidly multiply and the infection is limited to the respiratory tract, ii) young animals

display comparatively severe infections, iii) various systemic physiological and

neurological changes characteristic of human infection are observed in mice. One of

the apparent drawbacks of the murine model, for which it is often criticized, is its

inability to display the characteristic paroxysmal cough. Additionally, a rat model of B.

20

pertussis that can reproduce human pertussis illness with respect to course of

infection and cough production has been developed. This involves intrabronchially

injecting rats with agarose beads coated with challenge strain (55, 56).

Animal models are also critical for studying the protective efficacy of vaccines

and its pathophysiological responses. The gold standard to test the potency of

vaccines against B. pertussis is the lethal intracerebral challenge model or the

Kendrick test. In this model, vaccines are assessed based on its ability to protect

immunized mice against a lethal intracerebral challenge. Surprisingly, although the

previously available B. pertussis whole cell vaccines protect mice against the lethal

intracerebral challenge, the currently available new generation acellular vaccines fail

to pass the Kendrick test (27).

Knock out mice and immunodeficient mice such as Severe Combined

Immunodeficiency (SCID) mice have been frequently utilized for investigating the

interaction of Bordetella spp with the host immune system as well as to discern the

immunomodulatory role of specific virulence factors (73).

Immune Responses to B. pertussis and B. bronchiseptica. Bordetella spp exploit

both extracellular and intracellular host niches and employ different pathogenic

strategies to subvert immune responses. The majority of studies exploring immune

responses to Bordetella infection focus on development of new vaccine approaches,

its immunological properties and effector mechanisms that contribute to protection

against the human pathogen B. pertussis. Studies also provide evidence of

21

complimentary roles for both innate and adaptive immunity in bacterial clearance

subsequent to infections as well as vaccinations (57, 108).

Innate Immune Responses.

Neutrophils (PMN) and Macrophages.

PMNs are a major component of the initial immune responses vital for clearing

B. bronchiseptica infections. Infection with B. bronchiseptica results in rapid

recruitment of neutrophils to the lungs of infected mice. Consistent with this

observation, neutropenic mice succumb to B. bronchiseptica infections within 1-4

days post inoculation. Furthermore in vitro cytotoxicity assays using murine

macrophage cell line J774 and tunnel assay of B. bronchiseptica infected mice lungs

suggested that this species may induce apoptosis of alveolar macrophages. In

contrast, B. pertussis infection leads to cellular infiltration in lungs consisting

predominantly of macrophages and to a lesser extent neutrophils. In fact, neutropenic

mice survive B. pertussis infection. These studies thus demonstrate the differences in

immune responses elicited to the animal adapted and the human adapted Bordetella

spp and that alveolar macrophages are critical for limiting B. pertussis infection while

neutrophils are required for B. bronchiseptica clearance (57).

Toll like receptors (TLR).

Several studies have revealed a role for Toll like receptors (TLR) in innate

immune responses to Bordetella spp in mice. TLR4 is critical for LPS- induced

22

proinflammatory cytokine production in response to B. bronchiseptica infection.

Furthermore, TLR4 deficient mice show very drastic pathology in lungs and succumb

to B. bronchiseptica infection as early as 3 days post inoculation (93). Studies

investigating the mechanism of antibody mediated clearance of B. bronchiseptica

revealed that TLR4 is critical for recruitment of neutrophils and subsequent clearance

of antibody and complement opsonized bacteria (82). B. pertussis also causes a

more severe disease in TLR4 deficient mice. Studies have revealed that TLR 4-

signaling- induced IL-10 production inhibits inflammation and pathology in response

to B. pertussis infection (67). TLR4 is also vital for B. pertussis vaccine- mediated

protective immunity through the induction of Th1 and Th17 cells (66).

In vitro studies also demonstrate that B. bronchiseptica flagellin elicits

chemokine and cytokine production through TLR5 signaling (89). Despite these

convincing in vitro data, the in vivo role of TLR5 in B. bronchiseptica- host cell

interaction is unknown as flagellin expression is repressed on infection and ectopic

expression of flagellin has been shown to be inhibitory to Bordetella colonization (1).

Complement.

Complement system is an integral part of the host immune responses involved

in promoting clearance of pathogens like bacteria through opsonization –mediated

phagocytosis, killing, or augmentation of inflammation. Complement activation can

occur via three pathways: the antibody mediated classical pathway, antibody

independent alternate pathway and the mannose binding lectin pathway. Bordetella

pertussis expresses an autotransporter protein; BrkA (Bordetella resistance to killing)

23

that imparts resistance to the bacteoricidal activity of complement and provides

protection against antimicrobial peptides (43). Studies have demonstrated that the

disruption of brkA gene leads to enhanced deposition of C4 protein and subsequently

increasing susceptibility of B. pertussis to complement mediated killing. These

studies therefore, suggest that BrkA inhibits the classical pathway of complement

activation through preventing accumulation of C4 protein on bacterial surface (8). In

vitro serum killing assays suggest that B. pertussis is resistant to the alternate

pathway at lower naïve serum concentrations. In contrast, B. bronchiseptica is

resistant to killing by naïve serum, even at high concentrations, thereby suggesting

that this animal adapted species is resistant to the alternate pathway of complement

activation (57). C3 deficient mice are as efficient as wt mice in controlling Bordetella

infection as well as in the generation of infection-induced protective immunity.

However, studies have also demonstrated that C3, but not C5 and CR3 are vital for

vaccine induce immunity and antibody –mediated clearance of Bordetella from the

lower respiratory tract (82).

B cell responses.

There is definitive evidence primarily based on passive transfer experiments

that antibodies are critical for resolving Bordetella infections as well as for vaccine

mediated immune responses. Antibodies function through different mechanisms- 1)

Inhibition of adherence to prevent colonization, 2) neutralization and clearance of

toxins, 3) opsonization to enhance phagocytosis and 4) complement activation to

promote lysis (106). B cell deficient MuMT mice fail to clear Bordetella infection from

24

the trachea and lungs (81). Adoptive transfer of immune serum from mice infected

with B. bronchiseptica can rapidly clear infection of this organism from the lower

respiratory tract as early as 7 days post challenge. In contrast, B. pertussis can resist

clearance by passively transferred immune serum upto 7 days, after which it gets

cleared (80). Numerous studies investigating the role of individual virulence factors in

imparting protective immunity have revealed that passive immunization with

antibodies against FHA, Prn, Fim, PT and LPS can confer different levels of

protection (108). Moreover, studies have also shown a direct evidence for the critical

role of antibodies in humans by demonstrating that adoptive transfer of anti-B.

pertussis sera reduced the severity of disease in infected patients (51).

The efficiency of different antibodies directed against a range of virulence

factors to inhibit adherence has also been tested in vitro. Antibodies to FHA, Prn,

Fim, PT and LPS inhibit attachment of B. pertussis to human bronchial epithelial cells

(106). Also anti-FHA antibodies prevent attachment of B. pertussis to neutrophils,

while anti- CyaA sera promotes adherence and phagocytosis by neutrophils. Similarly

antibodies to Prn have also been shown to be crucial for phagocytosis of B. pertussis

by neutrophils. Natural infection as well as immunization with the whole cell vaccines

induces predominantly IgG2a antibodies, a subclass which is considered to be critical

for opsonization and complement fixation. IgG2a is also associated with greater

infection or vaccine induced protection. B. pertussis infection also induces the

mucosal, secretory isotype; IgA, both in mice and humans (106). IgA in convalescent

serum inhibits adherence of B. pertussis to ciliated epithelial cells in vitro and is

suggested to be critical for clearance of primary infection in vivo (106). Furthermore,

25

studies using IgA deficient mice also demonstrate that this isotype is important for

reducing B. bronchiseptica colonization from the upper respiratory tract (149).

T cell responses.

Direct evidence of the critical role for T cells in immunity is provided by the fact

that B. bronchiseptica and certain strains of B. pertussis cause lethal infection in

athymic as well as SCID mice (57). In sub-lethally irradiated mice, in which the T cells

are incapable of responding to antigens, B. pertussis causes a protracted infection.

Adoptive transfer of immune spleen cells from B. pertussis infected convalescent

mice imparted the recipient irradiated mice the ability to clear the infection within 2-3

weeks. Furthermore, studies investigating the contributions of individual

subpopulations of immune cells revealed that adoptive transfer of immune CD4+ T

cells, but not CD8+ T cells conferred the ability to clear B. pertussis infections in

irradiated mice (107). The vital role of CD4+ T cells in protective immunity is further

established by the study which demonstrated that intranasal immunization with

inactivated bacteria did not protect CD4 knock out mice against B. pertussis

challenge. In contrast, the same inactivated whole cell vaccine protected CD8 knock

out mice (84). Collectively these data suggest that CD4+ T cells, but not CD8+ T cells,

are important in both containing the primary infection as well as in vaccine mediated

protection.

Respiratory tract infection and vaccination with whole cell vaccines induce Th1

type responses with IFN-γ production. In contrast, the currently available acellular

vaccines generate Th2 cells that secrete IL-5, IL-4 and low levels of IFN-γ (66).

26

Addition of IL-12 to acellular vaccine led to priming of Th1 response and augmented

its protective efficacy (91). Consistent with these data, IFN-γ knockout and IL-12

deficient mice show greater bacterial load and delayed clearance while IL-10 knock-

out mice show faster clearance of infection (115). Apart from Th1 cells, recent in vitro

studies also demonstrated that B. bronchiseptica pulsed macrophages when

cocultured with CD4+ T cells, stimulate IL-17 production and preferentially induce

Th17 cells. Similarly immunization with B. pertussis whole cell vaccines, but not

acellular vaccines induced IL -17 producing T cells in addition to Th1 cells. Consistent

with these data, neutralization of IL-17 in vivo immediately prior to and after B.

pertussis challenge abrogated the protective efficacy of whole cell vaccines (66).

Currently Available Vaccines against B. pertussis.

Introduction of vaccines have decreased the incidence of whooping cough by

more than 80% as compared to the pre-vaccine era. The whole cell pertussis vaccine

(PW) containing killed whole B. pertussis organisms, was developed in the 1930s and

was combined with Diptheria (D), Tetanus (T) vaccines in 1940s and widely used in

clinical practice henceforth. PW is prepared by different methods, some of which

includes formalin fixing, heat killing, fractionation and extraction of B. pertussis (95).

Efficacy studies revealed that a primary series of four doses have 70-90%

effectiveness in preventing a serious pertussis illness. However, PW imparted

immunity has been suggested to wane over time with modest to no protection 5-10

years following the last dose (21). Furthermore, reactogenicity evaluations revealed

common local and systemic reactions in recipients of PW which was mainly attributed

27

to LPS. Studies have demonstrated that common local reactions are redness,

swelling and pain at the injection site and these occurred in 37-50% of DTP recipients

as compared to DT recipients. Systemic reactions to PW include drowsiness, fever,

vomiting and more serious events such as neurological diseases and death (5-7).

Although PW is still routinely used in many developing countries, at present this

vaccine is replaced by acellular pertussis vaccines in the United States due to the

adverse reactogenicity concerns.

The pediatric formulations of acellular vaccines (PC) were approved to be used

in United States in 1996. PC are subunit vaccines containing purified inactivated

components in varying concentrations. There are different combinations of PC

available that can range from monocomponent vaccines containing only PT to four

component vaccines consisting of PT, FHA, Fim and Prn. Although individual

antigens impart a certain level of protection, both murine model studies and clinical

trials demonstrate that multicomponent vaccines have higher estimated efficacy as

compared to monocomponent vaccines (108). PC in combination with Diphtheria and

Tetanus vaccines (DTaP) is administered in a series of four doses at 2 , 4 , 6 and 15-

18 months of age. A fifth dose is recommended for children who received the

primary four doses before the age of four, before joining school. Despite having high

effectiveness in clinical trials, this new generation PC does not pass the Kendrick test

and is considered to be less effective than PW. Due to increasing cases of pertussis

incidence in adults and adolescents, adult formulations of acellular vaccines were

approved by the CDC in 2005. Similar to PW, PC- imparted immunity is also

28

suggested to last for only 5-10 years following the last dose. Therefore, adult

formulations are theorized to booster immunity and lessen the incidence of pertussis.

Vaccines against B. bronchiseptica.

Currently available vaccines against B. bronchiseptica use live attenuated or

heat killed bacteria. Vaccines against kennel cough available in United States include

either individual heat killed or avirulent whole cells or in combination with other

vaccines such as live attenuated adenovirus type 2. Vaccines are administrated at 6-

8 and 10-12 weeks of age and subsequently annually. Similarly numerous vaccines

against atrophic rhinitis in pigs are available. These consist of inactivated or

nonpathogenic whole cells and are typically administered in combination with various

preparations of another important pathogen of swine, Pasterulla multocida. The

recommended schedule for vaccine administration in swine includes 1 dose prior to

delivery, shortly after birth and at 3-4 months of age (95). However, there are only

limited data on the safety and effectiveness of these vaccines. Although, heat killed

whole cell vaccines elicit high antibody responses, they do not induce as effective

protective immunity as infection induced immunity. Moreover, the basis for

attenuation of the commercially available live attenuated vaccines is unknown. As a

result, there is always the possibility of reversion of these strains back to the virulent

wild type form (136, 137). Therefore, current studies have been focusing on the use

of attenuated strains containing stable genetically defined mutations as vaccines.

These include strains with metabolic defects (aroA), as well as defects in production

29

of toxins such as DNT. A temperature sensitive urease strain defective in growth

above 340C has also been evaluated in a guinea pig model (130, 131). Furthermore,

the protective efficacy of a double mutant strain lacking adenylate cyclase toxin and

type III secretion system has been characterized using the mouse model system of

infection (92). However, the use of genetically defined attenuated strains still does not

resolve the possibility of reversion back into virulent form. The possibility of reversion

might be greater under competitive environments in the host such as co-infections

with other pathogenic organisms. Moreover, use of live attenuated strains for

vaccination purposes may increase the chances of zoonotic transmission especially

in immuno-compromised patients. Thus, development of acellular vaccines that can

provide protective immunity against B. bronchiseptica must be a priority.

A principal impediment towards development or improvement of vaccines for

B. pertussis and B. bronchiseptica is a gap in our understanding of the identity and

function of gene products that are critical for efficient respiratory tract infection. We

believe that efficient colonization and subsequent development of disease by

Bordetella require the interplay of multiple factors. The majority of current studies are

focused on the previously identified and characterized Bordetella factors which are

mentioned above and there is only limited focus on identification of novel antigens

that are vital for pathogenesis and with the potential to elicit protective immunity.

While several published data support the importance of the major virulence factors

namely, FHA, Fim, Prn CyaA, and PT in mediating attachment, colonization and

persistence, our unpublished data (Sukumar et al; in preparation) reveal that

simultaneous deletion of all these five factors does not abrogate upper respiratory

30

tract colonization by Bordetella. In fact, our observations provide evidence that

intranasal administration of this mutant strain results in protection from subsequent

challenge with wt B. bronchiseptica and B. pertussis in a mouse model of infection.

These studies highlight the contribution of hitherto unknown factors in Bordetella

infectious cycle. With the goal of identifying novel virulence factors and /or vaccine

candidates, we examined the genomic content of B. bronchiseptica. A computational

search was employed to identify unknown Open reading Frames (ORF) that have

potential BvgA binding motifs in the putative promoter regions. Our rationale for

searching ORFs which may be regulated by the BvgAS system was that the majority

of the previously characterized virulence factors are positively regulated by this

master regulatory system. Thus, our prediction was that, if BvgAS locus controls the

expression of a gene, then it may be important for Bordetella pathogenesis. Although

the search revealed multiple genes, we focused our study on an ORF that was

designated as a “putative adhesin” in the annotated B. bronchiseptica genome data

base and harbored homology to the previously characterized Bordetella protein BipA.

In Chapter II, we describe studies characterizing this ORF, which we designated as

bcfA. Using RT-PCR assays and EMSA, we confirm that the BvgAS system positively

regulates bcfA expression. Moreover, our studies indicate that, although homologous,

BipA and BcfA have differential phase specific expression patterns. We also

demonstrated that BcfA is localized to the outermembrane and that it is expressed in

vivo during Bordetella infection in rats. We also evaluated the role of these two

paralogous proteins in B. bronchiseptica colonization in the rat model of infection.

While individual deletions of BipA or BcfA did not abrogate B. bronchiseptica

31

colonization, a mutant strain deficient in both these proteins was defective in tracheal

colonization as compared to its wt parent strain. Based on this finding, we

hypothesize that BipA and BcfA have overlapping function in mediating tracheal

colonization by B. bronchiseptica.

In the subsequent Chapter III we explore the efficacy of BcfA as a vaccine

candidate in eliciting protective immune responses against B. bronchiseptica

infection. Our justifications for investigating the vaccine potential of BcfA were i) BcfA

is an outermembrane protein ii) BcfA-specific antibodies are produced during a wild

type B. bronchiseptica infection in rats. These results show that BcfA has the

potential to interact with the host immune system and induce at least the humoral

responses. Our studies reveal that both active and passive immunization with BcfA

provides protection against subsequent challenge with B. bronchiseptica in a mouse

model of infection. We also show that immunization with BcfA induces specific

antibodies in mice, with Ig2a being the predominant isotype. Data from this section

also demonstrate that neutrophils are critical for anti-BcfA antibody mediated

clearance of B. bronchiseptica from the lower respiratory tract. Finally we show that

immunization with BcfA induces a Th1 type response in splenocytes leading to IFN-γ

production. The utility of this antigen as a vaccine candidate is highlighted by the data

showing the conservation of expression of BcfA among multiple clinical isolates of B.

bronchiseptica.

32

REFERENCES 1. Akerley, B. J., P. A. Cotter, and J. F. Miller. 1995. Ectopic expression of

the flagellar regulon alters development of the Bordetella-host interaction. Cell 80:611-20.

2. Akerley, B. J., and J. F. Miller. 1993. Flagellin gene transcription in

Bordetella bronchiseptica is regulated by the BvgAS virulence control system. J Bacteriol 175:3468-79.

3. Akerley, B. J., D. M. Monack, S. Falkow, and J. F. Miller. 1992. The

bvgAS locus negatively controls motility and synthesis of flagella in Bordetella bronchiseptica. J Bacteriol 174:980-90.

4. Arico, B., and R. Rappuoli. 1987. Bordetella parapertussis and

Bordetella bronchiseptica contain transcriptionally silent pertussis toxin genes. J Bacteriol 169:2847-53.

5. Baraff, L. J., W. J. Ablon, and R. C. Weiss. 1983. Possible temporal

association between diphtheria-tetanus toxoid-pertussis vaccination and sudden infant death syndrome. Pediatr Infect Dis 2:7-11.

6. Baraff, L. J., C. R. Manclark, J. D. Cherry, P. Christenson, and S. M.

Marcy. 1989. Analyses of adverse reactions to diphtheria and tetanus toxoids and pertussis vaccine by vaccine lot, endotoxin content, pertussis vaccine potency and percentage of mouse weight gain. Pediatr Infect Dis J 8:502-7.

7. Baraff, L. J., W. D. Shields, L. Beckwith, G. Strome, S. M. Marcy, J. D.

Cherry, and C. R. Manclark. 1988. Infants and children with convulsions and hypotonic-hyporesponsive episodes following diphtheria-tetanus-pertussis immunization: follow-up evaluation. Pediatrics 81:789-94.

8. Barnes, M. G., and A. A. Weiss. 2001. BrkA protein of Bordetella

pertussis inhibits the classical pathway of complement after C1 deposition. Infect Immun 69:3067-72.

9. Beier, D., B. Schwarz, T. M. Fuchs, and R. Gross. 1995. In vivo

characterization of the unorthodox BvgS two-component sensor protein of Bordetella pertussis. J Mol Biol 248:596-610.

10. Bergfors, E., B. Trollfors, J. Taranger, T. Lagergard, V. Sundh, and G.

Zackrisson. 1999. Parapertussis and pertussis: differences and similarities in incidence, clinical course, and antibody responses. Int J Infect Dis 3:140-6.

33

11. Bordet, J. a. O. G. 1909. L'endotoxine coquelucheuse. Ann. Inst. Pasteur 23:415-419.

12. Bordet, J. a. O. G. 1906. Le microbe de la coqueluche. Ann. Inst. Pasteur

20:48-68. 13. Boschwitz, J. S., J. W. Batanghari, H. Kedem, and D. A. Relman. 1997.

Bordetella pertussis infection of human monocytes inhibits antigen-dependent CD4 T cell proliferation. J Infect Dis 176:678-86.

14. Boucher, P. E., and S. Stibitz. 1995. Synergistic binding of RNA

polymerase and BvgA phosphate to the pertussis toxin promoter of Bordetella pertussis. J Bacteriol 177:6486-91.

15. Bradford, P. G., and R. P. Rubin. 1985. Pertussis toxin inhibits

chemotactic factor-induced phospholipase C stimulation and lysosomal enzyme secretion in rabbit neutrophils. FEBS Lett 183:317-20.

16. Buonavoglia, C., and V. Martella. 2007. Canine respiratory viruses. Vet

Res 38:355-73. 17. Carbonetti, N. H. 2007. Immunomodulation in the pathogenesis of

Bordetella pertussis infection and disease. Curr Opin Pharmacol 7:272-8. 18. Carbonetti, N. H., G. V. Artamonova, C. Andreasen, E. Dudley, R. M.

Mays, and Z. E. Worthington. 2004. Suppression of serum antibody responses by pertussis toxin after respiratory tract colonization by Bordetella pertussis and identification of an immunodominant lipoprotein. Infect Immun 72:3350-8.

19. Cherry, J. D. 1999. Pertussis in the preantibiotic and prevaccine era, with

emphasis on adult pertussis. Clin Infect Dis 28 Suppl 2:S107-11. 20. Cherry, J. D., J. Gornbein, U. Heininger, and K. Stehr. 1998. A search

for serologic correlates of immunity to Bordetella pertussis cough illnesses. Vaccine 16:1901-6.

21. Cherry, J. D., and P. Olin. 1999. The science and fiction of pertussis

vaccines. Pediatrics 104:1381-3. 22. Cherry, J. D., D. X. Xing, P. Newland, K. Patel, U. Heininger, and M. J.

Corbel. 2004. Determination of serum antibody to Bordetella pertussis adenylate cyclase toxin in vaccinated and unvaccinated children and in children and adults with pertussis. Clin Infect Dis 38:502-7.

34

23. Confer, D. L., and J. W. Eaton. 1982. Phagocyte impotence caused by an invasive bacterial adenylate cyclase. Science 217:948-50.

24. Confer, D. L., A. S. Slungaard, E. Graf, S. S. Panter, and J. W. Eaton.

1984. Bordetella adenylate cyclase toxin: entry of bacterial adenylate cyclase into mammalian cells. Adv Cyclic Nucleotide Protein Phosphorylation Res 17:183-7.

25. Cookson, B. T., H. L. Cho, L. A. Herwaldt, and W. E. Goldman. 1989.

Biological activities and chemical composition of purified tracheal cytotoxin of Bordetella pertussis. Infect Immun 57:2223-9.

26. Cookson, B. T., A. N. Tyler, and W. E. Goldman. 1989. Primary

structure of the peptidoglycan-derived tracheal cytotoxin of Bordetella pertussis. Biochemistry 28:1744-9.

27. Corbel, M. J., and D. K. Xing. 1997. A consideration of control

requirements for acellular pertussis vaccines. Dev Biol Stand 89:343-7. 28. Cotter, P. A., and A. M. Jones. 2003. Phosphorelay control of virulence

gene expression in Bordetella. Trends Microbiol 11:367-73. 29. Cotter, P. A., and J. F. Miller. 1994. BvgAS-mediated signal transduction:

analysis of phase-locked regulatory mutants of Bordetella bronchiseptica in a rabbit model. Infect Immun 62:3381-90.

30. Cotter, P. A., and J. F. Miller. 1997. A mutation in the Bordetella

bronchiseptica bvgS gene results in reduced virulence and increased resistance to starvation, and identifies a new class of Bvg-regulated antigens. Mol Microbiol 24:671-85.

31. Cotter, P. A., M. H. Yuk, S. Mattoo, B. J. Akerley, J. Boschwitz, D. A.

Relman, and J. F. Miller. 1998. Filamentous hemagglutinin of Bordetella bronchiseptica is required for efficient establishment of tracheal colonization. Infect Immun 66:5921-9.

32. Coutte, L., R. Antoine, H. Drobecq, C. Locht, and F. Jacob-

Dubuisson. 2001. Subtilisin-like autotransporter serves as maturation protease in a bacterial secretion pathway. EMBO J 20:5040-8.

33. Coutte, L., E. Willery, R. Antoine, H. Drobecq, C. Locht, and F. Jacob-

Dubuisson. 2003. Surface anchoring of bacterial subtilisin important for maturation function. Mol Microbiol 49:529-39.

35

34. Cowell, J. L., E. L. Hewlett, and C. R. Manclark. 1979. Intracellular localization of the dermonecrotic toxin of Bordetella pertussis. Infect Immun 25:896-901.

35. Cummings, C. A., M. M. Brinig, P. W. Lepp, S. van de Pas, and D. A.

Relman. 2004. Bordetella species are distinguished by patterns of substantial gene loss and host adaptation. J Bacteriol 186:1484-92.

36. Dautin, N., and H. D. Bernstein. 2007. Protein secretion in gram-

negative bacteria via the autotransporter pathway. Annu Rev Microbiol 61:89-112.

37. De Serres, G., R. Shadmani, B. Duval, N. Boulianne, P. Dery, M.

Douville Fradet, L. Rochette, and S. A. Halperin. 2000. Morbidity of pertussis in adolescents and adults. J Infect Dis 182:174-9.

38. Deora, R., H. J. Bootsma, J. F. Miller, and P. A. Cotter. 2001. Diversity

in the Bordetella virulence regulon: transcriptional control of a Bvg-intermediate phase gene. Mol Microbiol 40:669-83.

39. Dworkin, M. S., P. S. Sullivan, S. E. Buskin, R. D. Harrington, J.

Olliffe, R. D. MacArthur, and C. E. Lopez. 1999. Bordetella bronchiseptica infection in human immunodeficiency virus-infected patients. Clin Infect Dis 28:1095-9.

40. Eldering, G., and P. Kendrick. 1938. Bacillus Para-Pertussis: A Species

Resembling Both Bacillus Pertussis and Bacillus Bronchisepticus but Identical with Neither. J Bacteriol 35:561-72.

41. Eldering, G., and P. L. Kendrick. 1952. Incidence of parapertussis in the

Grand Rapids area as indicated by 16 years' experience with diagnostic cultures. Am J Public Health Nations Health 42:27-31.

42. Emsley, P., G. McDermott, I. G. Charles, N. F. Fairweather, and N. W.

Isaacs. 1994. Crystallographic characterization of pertactin, a membrane-associated protein from Bordetella pertussis. J Mol Biol 235:772-3.

43. Fernandez, R. C., and A. A. Weiss. 1994. Cloning and sequencing of a

Bordetella pertussis serum resistance locus. Infect Immun 62:4727-38. 44. Ferry, N. S., and H. C. Klix. 1918. Studies Relative to the Apparent Close

Relationship between Bact. pertussis and B. bronchisepticus: II. Complement Fixation Tests. J Bacteriol 3:309-12.

45. Flak, T. A., and W. E. Goldman. 1999. Signalling and cellular specificity

of airway nitric oxide production in pertussis. Cell Microbiol 1:51-60.

36

46. Gerlach, G., F. von Wintzingerode, B. Middendorf, and R. Gross. 2001. Evolutionary trends in the genus Bordetella. Microbes Infect 3:61-72.

47. Gilberg, S., E. Njamkepo, I. P. Du Chatelet, H. Partouche, P. Gueirard,

C. Ghasarossian, M. Schlumberger, and N. Guiso. 2002. Evidence of Bordetella pertussis infection in adults presenting with persistent cough in a french area with very high whole-cell vaccine coverage. J Infect Dis 186:415-8.

48. Goldman, W. E., D. G. Klapper, and J. B. Baseman. 1982. Detection,

isolation, and analysis of a released Bordetella pertussis product toxic to cultured tracheal cells. Infect Immun 36:782-94.

49. Goodnow, R. A. 1980. Biology of Bordetella bronchiseptica. Microbiol

Rev 44:722-38. 50. Goodwin, M. S., and A. A. Weiss. 1990. Adenylate cyclase toxin is

critical for colonization and pertussis toxin is critical for lethal infection by Bordetella pertussis in infant mice. Infect Immun 58:3445-7.

51. Granstrom, M., A. M. Olinder-Nielsen, P. Holmblad, A. Mark, and K.

Hanngren. 1991. Specific immunoglobulin for treatment of whooping cough. Lancet 338:1230-3.

52. Greenberg, D. P., C. H. von Konig, and U. Heininger. 2005. Health

burden of pertussis in infants and children. Pediatr Infect Dis J 24:S39-43. 53. Gross, R., B. Arico, and R. Rappuoli. 1989. Families of bacterial signal-

transducing proteins. Mol Microbiol 3:1661-7. 54. Gueirard, P., and N. Guiso. 1993. Virulence of Bordetella bronchiseptica:

role of adenylate cyclase-hemolysin. Infect Immun 61:4072-8. 55. Hall, E., R. Parton, and A. C. Wardlaw. 1998. Responses to acellular

pertussis vaccines and component antigens in a coughing-rat model of pertussis. Vaccine 16:1595-603.

56. Hall, E., R. Parton, and A. C. Wardlaw. 1999. Time-course of infection

and responses in a coughing rat model of pertussis. J Med Microbiol 48:95-8.

57. Harvill, E. T., P. A. Cotter, and J. F. Miller. 1999. Pregenomic

comparative analysis between bordetella bronchiseptica RB50 and Bordetella pertussis tohama I in murine models of respiratory tract infection. Infect Immun 67:6109-18.

37

58. Hazenbos, W. L., C. A. Geuijen, B. M. van den Berg, F. R. Mooi, and R. van Furth. 1995. Bordetella pertussis fimbriae bind to human monocytes via the minor fimbrial subunit FimD. J Infect Dis 171:924-9.

59. Hazenbos, W. L., B. M. van den Berg, C. W. Geuijen, F. R. Mooi, and

R. van Furth. 1995. Binding of FimD on Bordetella pertussis to very late antigen-5 on monocytes activates complement receptor type 3 via protein tyrosine kinases. J Immunol 155:3972-8.

60. Heininger, U., K. Stehr, S. Schmitt-Grohe, C. Lorenz, R. Rost, P. D.

Christenson, M. Uberall, and J. D. Cherry. 1994. Clinical characteristics of illness caused by Bordetella parapertussis compared with illness caused by Bordetella pertussis. Pediatr Infect Dis J 13:306-9.

61. Heiss, L. N., T. A. Flak, J. R. Lancaster, Jr., M. L. McDaniel, and W. E.

Goldman. 1993. Nitric oxide mediates Bordetella pertussis tracheal cytotoxin damage to the respiratory epithelium. Infect Agents Dis 2:173-7.

62. Heiss, L. N., S. A. Moser, E. R. Unanue, and W. E. Goldman. 1993.

Interleukin-1 is linked to the respiratory epithelial cytopathology of pertussis. Infect Immun 61:3123-8.

63. Hellwig, S. M., M. E. Rodriguez, G. A. Berbers, J. G. van de Winkel,

and F. R. Mooi. 2003. Crucial role of antibodies to pertactin in Bordetella pertussis immunity. J Infect Dis 188:738-42.

64. Hewlett, E. L., and K. M. Edwards. 2005. Clinical practice. Pertussis--not

just for kids. N Engl J Med 352:1215-22. 65. Hewlett, E. L., V. M. Gordon, J. D. McCaffery, W. M. Sutherland, and

M. C. Gray. 1989. Adenylate cyclase toxin from Bordetella pertussis. Identification and purification of the holotoxin molecule. J Biol Chem 264:19379-84.

66. Higgins, S. C., A. G. Jarnicki, E. C. Lavelle, and K. H. Mills. 2006.

TLR4 mediates vaccine-induced protective cellular immunity to Bordetella pertussis: role of IL-17-producing T cells. J Immunol 177:7980-9.

67. Higgins, S. C., E. C. Lavelle, C. McCann, B. Keogh, E. McNeela, P.

Byrne, B. O'Gorman, A. Jarnicki, P. McGuirk, and K. H. Mills. 2003. Toll-like receptor 4-mediated innate IL-10 activates antigen-specific regulatory T cells and confers resistance to Bordetella pertussis by inhibiting inflammatory pathology. J Immunol 171:3119-27.

68. Hoppe, J. E. 1999. Update on respiratory infection caused by Bordetella parapertussis. Pediatr Infect Dis J 18:375-81.

38

69. Horiguchi, Y., T. Nakai, and K. Kume. 1991. Effects of Bordetella bronchiseptica dermonecrotic toxin on the structure and function of osteoblastic clone MC3T3-e1 cells. Infect Immun 59:1112-6.

70. Horiguchi, Y., N. Sugimoto, and M. Matsuda. 1993. Stimulation of DNA

synthesis in osteoblast-like MC3T3-E1 cells by Bordetella bronchiseptica dermonecrotic toxin. Infect Immun 61:3611-5.

71. Huebner, E. S., B. Christman, S. Dummer, Y. W. Tang, and S.

Goodman. 2006. Hospital-acquired Bordetella bronchiseptica infection following hematopoietic stem cell transplantation. J Clin Microbiol 44:2581-3.

72. Iida, T., and T. Okonogi. 1971. Lienotoxicity of Bordetella pertussis in

mice. J Med Microbiol 4:51-61. 73. Inatsuka, C. S., S. M. Julio, and P. A. Cotter. 2005. Bordetella

filamentous hemagglutinin plays a critical role in immunomodulation, suggesting a mechanism for host specificity. Proc Natl Acad Sci U S A 102:18578-83.

74. Ishibashi, Y., S. Claus, and D. A. Relman. 1994. Bordetella pertussis

filamentous hemagglutinin interacts with a leukocyte signal transduction complex and stimulates bacterial adherence to monocyte CR3 (CD11b/CD18). J Exp Med 180:1225-33.

75. Ishibashi, Y., D. A. Relman, and A. Nishikawa. 2001. Invasion of human

respiratory epithelial cells by Bordetella pertussis: possible role for a filamentous hemagglutinin Arg-Gly-Asp sequence and alpha5beta1 integrin. Microb Pathog 30:279-88.

76. Iwata, S., T. Aoyama, A. Goto, H. Iwai, Y. Sato, H. Akita, Y. Murase, T.

Oikawa, T. Iwata, S. Kusano, and et al. 1991. Mixed outbreak of Bordetella pertussis and Bordetella parapertussis in an apartment house. Dev Biol Stand 73:333-41.

77. Jacobs, C., B. Joris, M. Jamin, K. Klarsov, J. Van Beeumen, D.

Mengin-Lecreulx, J. van Heijenoort, J. T. Park, S. Normark, and J. M. Frere. 1995. AmpD, essential for both beta-lactamase regulation and cell wall recycling, is a novel cytosolic N-acetylmuramyl-L-alanine amidase. Mol Microbiol 15:553-9.

78. Kania, S. A., S. Rajeev, E. H. Burns, Jr., T. F. Odom, S. M. Holloway,

and D. A. Bemis. 2000. Characterization of fimN, a new Bordetella bronchiseptica major fimbrial subunit gene. Gene 256:149-55.

39

79. Katada, T., M. Tamura, and M. Ui. 1983. The A protomer of islet-activating protein, pertussis toxin, as an active peptide catalyzing ADP-ribosylation of a membrane protein. Arch Biochem Biophys 224:290-8.

80. Kirimanjeswara, G. S., L. M. Agosto, M. J. Kennett, O. N. Bjornstad,

and E. T. Harvill. 2005. Pertussis toxin inhibits neutrophil recruitment to delay antibody-mediated clearance of Bordetella pertussis. J Clin Invest 115:3594-601.

81. Kirimanjeswara, G. S., P. B. Mann, and E. T. Harvill. 2003. Role of

antibodies in immunity to Bordetella infections. Infect Immun 71:1719-24. 82. Kirimanjeswara, G. S., P. B. Mann, M. Pilione, M. J. Kennett, and E. T.

Harvill. 2005. The complex mechanism of antibody-mediated clearance of Bordetella from the lungs requires TLR4. J Immunol 175:7504-11.

83. Lacey, B. W. 1960. Antigenic modulation of Bordetella pertussis. J Hyg

(Lond) 58:57-93. 84. Leef, M., K. L. Elkins, J. Barbic, and R. D. Shahin. 2000. Protective

immunity to Bordetella pertussis requires both B cells and CD4(+) T cells for key functions other than specific antibody production. J Exp Med 191:1841-52.

85. Livey, I., C. J. Duggleby, and A. Robinson. 1987. Cloning and

nucleotide sequence analysis of the serotype 2 fimbrial subunit gene of Bordetella pertussis. Mol Microbiol 1:203-9.

86. Locht, C., and J. M. Keith. 1986. Pertussis toxin gene: nucleotide

sequence and genetic organization. Science 232:1258-64. 87. Long, S. S., H. W. Lischner, A. Deforest, and J. L. Clark. 1990.

Serologic evidence of subclinical pertussis in immunized children. Pediatr Infect Dis J 9:700-5.

88. Long, S. S., C. J. Welkon, and J. L. Clark. 1990. Widespread silent

transmission of pertussis in families: antibody correlates of infection and symptomatology. J Infect Dis 161:480-6.

89. Lopez-Boado, Y. S., L. M. Cobb, and R. Deora. 2005. Bordetella

bronchiseptica flagellin is a proinflammatory determinant for airway epithelial cells. Infect Immun 73:7525-34.

90. Magyar, T., N. Chanter, A. J. Lax, J. M. Rutter, and G. A. Hall. 1988.

The pathogenesis of turbinate atrophy in pigs caused by Bordetella bronchiseptica. Vet Microbiol 18:135-46.

40

91. Mahon, B. P., M. S. Ryan, F. Griffin, and K. H. Mills. 1996. Interleukin-12 is produced by macrophages in response to live or killed Bordetella pertussis and enhances the efficacy of an acellular pertussis vaccine by promoting induction of Th1 cells. Infect Immun 64:5295-301.

92. Mann, P., E. Goebel, J. Barbarich, M. Pilione, M. Kennett, and E.

Harvill. 2007. Use of a genetically defined double mutant strain of Bordetella bronchiseptica lacking adenylate cyclase and type III secretion as a live vaccine. Infect Immun 75:3665-72.

93. Mann, P. B., K. D. Elder, M. J. Kennett, and E. T. Harvill. 2004. Toll-like

receptor 4-dependent early elicited tumor necrosis factor alpha expression is critical for innate host defense against Bordetella bronchiseptica. Infect Immun 72:6650-8.

94. Martinez de Tejada, G., J. F. Miller, and P. A. Cotter. 1996.

Comparative analysis of the virulence control systems of Bordetella pertussis and Bordetella bronchiseptica. Mol Microbiol 22:895-908.

95. Mattoo, S., and J. D. Cherry. 2005. Molecular pathogenesis,

epidemiology, and clinical manifestations of respiratory infections due to Bordetella pertussis and other Bordetella subspecies. Clin Microbiol Rev 18:326-82.

96. Mattoo, S., J. F. Miller, and P. A. Cotter. 2000. Role of Bordetella

bronchiseptica fimbriae in tracheal colonization and development of a humoral immune response. Infect Immun 68:2024-33.

97. McGuirk, P., P. A. Johnson, E. J. Ryan, and K. H. Mills. 2000.

Filamentous hemagglutinin and pertussis toxin from Bordetella pertussis modulate immune responses to unrelated antigens. J Infect Dis 182:1286-9.

98. McGuirk, P., and K. H. Mills. 2000. Direct anti-inflammatory effect of a

bacterial virulence factor: IL-10-dependent suppression of IL-12 production by filamentous hemagglutinin from Bordetella pertussis. Eur J Immunol 30:415-22.

99. McMillan, D. J., E. Medina, C. A. Guzman, and M. J. Walker. 1999.

Expression of urease does not affect the ability of Bordetella bronchiseptica to colonise and persist in the murine respiratory tract. FEMS Microbiol Lett 178:7-11.

100. McMillan, D. J., M. Shojaei, G. S. Chhatwal, C. A. Guzman, and M. J.

Walker. 1996. Molecular analysis of the bvg-repressed urease of Bordetella bronchiseptica. Microb Pathog 21:379-94.

41

101. Meade, B. D., P. D. Kind, J. B. Ewell, P. P. McGrath, and C. R. Manclark. 1984. In vitro inhibition of murine macrophage migration by Bordetella pertussis lymphocytosis-promoting factor. Infect Immun 45:718-25.

102. Menozzi, F. D., C. Gantiez, and C. Locht. 1991. Interaction of the

Bordetella pertussis filamentous hemagglutinin with heparin. FEMS Microbiol Lett 62:59-64.

103. Merkel, T. J., P. E. Boucher, S. Stibitz, and V. K. Grippe. 2003. Analysis

of bvgR expression in Bordetella pertussis. J Bacteriol 185:6902-12. 104. Merkel, T. J., and S. Stibitz. 1995. Identification of a locus required for

the regulation of bvg-repressed genes in Bordetella pertussis. J Bacteriol 177:2727-36.

105. Merkel, T. J., S. Stibitz, J. M. Keith, M. Leef, and R. Shahin. 1998.

Contribution of regulation by the bvg locus to respiratory infection of mice by Bordetella pertussis. Infect Immun 66:4367-73.

106. Mills, K. H. 2001. Immunity to Bordetella pertussis. Microbes Infect 3:655-

77. 107. Mills, K. H., A. Barnard, J. Watkins, and K. Redhead. 1993. Cell-

mediated immunity to Bordetella pertussis: role of Th1 cells in bacterial clearance in a murine respiratory infection model. Infect Immun 61:399-410.

108. Mills, K. H., M. Ryan, E. Ryan, and B. P. Mahon. 1998. A murine model

in which protection correlates with pertussis vaccine efficacy in children reveals complementary roles for humoral and cell-mediated immunity in protection against Bordetella pertussis. Infect Immun 66:594-602.

109. Mink, C. M., J. D. Cherry, P. Christenson, K. Lewis, E. Pineda, D.

Shlian, J. A. Dawson, and D. A. Blumberg. 1992. A search for Bordetella pertussis infection in university students. Clin Infect Dis 14:464-71.

110. Mobberley-Schuman, P. S., B. Connelly, and A. A. Weiss. 2003.

Phagocytosis of Bordetella pertussis incubated with convalescent serum. J Infect Dis 187:1646-53.

111. Mooi, F. R., H. G. van der Heide, A. R. ter Avest, K. G. Welinder, I.

Livey, B. A. van der Zeijst, and W. Gaastra. 1987. Characterization of fimbrial subunits from Bordetella species. Microb Pathog 2:473-84.

42

112. Nicosia, A., M. Perugini, C. Franzini, M. C. Casagli, M. G. Borri, G. Antoni, M. Almoni, P. Neri, G. Ratti, and R. Rappuoli. 1986. Cloning and sequencing of the pertussis toxin genes: operon structure and gene duplication. Proc Natl Acad Sci U S A 83:4631-5.

113. Olin, P., F. Rasmussen, L. Gustafsson, H. O. Hallander, and H.

Heijbel. 1997. Randomised controlled trial of two-component, three-component, and five-component acellular pertussis vaccines compared with whole-cell pertussis vaccine. Ad Hoc Group for the Study of Pertussis Vaccines. Lancet 350:1569-77.

114. Parkhill, J., M. Sebaihia, A. Preston, L. D. Murphy, N. Thomson, D. E.

Harris, M. T. Holden, C. M. Churcher, S. D. Bentley, K. L. Mungall, A. M. Cerdeno-Tarraga, L. Temple, K. James, B. Harris, M. A. Quail, M. Achtman, R. Atkin, S. Baker, D. Basham, N. Bason, I. Cherevach, T. Chillingworth, M. Collins, A. Cronin, P. Davis, J. Doggett, T. Feltwell, A. Goble, N. Hamlin, H. Hauser, S. Holroyd, K. Jagels, S. Leather, S. Moule, H. Norberczak, S. O'Neil, D. Ormond, C. Price, E. Rabbinowitsch, S. Rutter, M. Sanders, D. Saunders, K. Seeger, S. Sharp, M. Simmonds, J. Skelton, R. Squares, S. Squares, K. Stevens, L. Unwin, S. Whitehead, B. G. Barrell, and D. J. Maskell. 2003. Comparative analysis of the genome sequences of Bordetella pertussis, Bordetella parapertussis and Bordetella bronchiseptica. Nat Genet 35:32-40.

115. Pilione, M. R., and E. T. Harvill. 2006. The Bordetella bronchiseptica

type III secretion system inhibits gamma interferon production that is required for efficient antibody-mediated bacterial clearance. Infect Immun 74:1043-9.

116. Porter, J. F., K. Connor, and W. Donachie. 1994. Isolation and

characterization of Bordetella parapertussis-like bacteria from ovine lungs. Microbiology 140 ( Pt 2):255-61.

117. Prasad, S. M., Y. Yin, E. Rodzinski, E. I. Tuomanen, and H. R. Masure.

1993. Identification of a carbohydrate recognition domain in filamentous hemagglutinin from Bordetella pertussis. Infect Immun 61:2780-5.

118. Rath, B. A., K. B. Register, J. Wall, D. M. Sokol, and R. B. Van Dyke.

2008. Persistent Bordetella bronchiseptica pneumonia in an immunocompetent infant and genetic comparison of clinical isolates with kennel cough vaccine strains. Clin Infect Dis 46:905-8.

119. Relman, D., E. Tuomanen, S. Falkow, D. T. Golenbock, K. Saukkonen,

and S. D. Wright. 1990. Recognition of a bacterial adhesion by an

43

integrin: macrophage CR3 (alpha M beta 2, CD11b/CD18) binds filamentous hemagglutinin of Bordetella pertussis. Cell 61:1375-82.

120. Rhodes, C. R., M. C. Gray, J. M. Watson, T. L. Muratore, S. B. Kim, E.

L. Hewlett, and C. M. Grisham. 2001. Structural consequences of divalent metal binding by the adenylyl cyclase toxin of Bordetella pertussis. Arch Biochem Biophys 395:169-76.

121. Riboli, B., P. Pedroni, A. Cuzzoni, G. Grandi, and F. de Ferra. 1991.

Expression of Bordetella pertussis fimbrial (fim) genes in Bordetella bronchiseptica: fimX is expressed at a low level and vir-regulated. Microb Pathog 10:393-403.

122. Roberts, M., N. F. Fairweather, E. Leininger, D. Pickard, E. L. Hewlett,

A. Robinson, C. Hayward, G. Dougan, and I. G. Charles. 1991. Construction and characterization of Bordetella pertussis mutants lacking the vir-regulated P.69 outer membrane protein. Mol Microbiol 5:1393-404.

123. Roop, R. M., 2nd, H. P. Veit, R. J. Sinsky, S. P. Veit, E. L. Hewlett, and

E. T. Kornegay. 1987. Virulence factors of Bordetella bronchiseptica associated with the production of infectious atrophic rhinitis and pneumonia in experimentally infected neonatal swine. Infect Immun 55:217-22.

124. Rose, T., P. Sebo, J. Bellalou, and D. Ladant. 1995. Interaction of calcium with Bordetella pertussis adenylate cyclase toxin. Characterization of multiple calcium-binding sites and calcium-induced conformational changes. J Biol Chem 270:26370-6.

125. Rosenthal, S., P. Strebel, P. Cassiday, G. Sanden, K. Brusuelas, and

M. Wharton. 1995. Pertussis infection among adults during the 1993 outbreak in Chicago. J Infect Dis 171:1650-2.

126. Rossi, C., J. Homand, C. Bauche, H. Hamdi, D. Ladant, and J.

Chopineau. 2003. Differential mechanisms for calcium-dependent protein/membrane association as evidenced from SPR-binding studies on supported biomimetic membranes. Biochemistry 42:15273-83.

127. Scarlato, V., A. Prugnola, B. Arico, and R. Rappuoli. 1990. Positive

transcriptional feedback at the bvg locus controls expression of virulence factors in Bordetella pertussis. Proc Natl Acad Sci U S A 87:10067.

128. Sebaihia, M., A. Preston, D. J. Maskell, H. Kuzmiak, T. D. Connell, N.

D. King, P. E. Orndorff, D. M. Miyamoto, N. R. Thomson, D. Harris, A. Goble, A. Lord, L. Murphy, M. A. Quail, S. Rutter, R. Squares, S. Squares, J. Woodward, J. Parkhill, and L. M. Temple. 2006. Comparison of the genome sequence of the poultry pathogen Bordetella

44

avium with those of B. bronchiseptica, B. pertussis, and B. parapertussis reveals extensive diversity in surface structures associated with host interaction. J Bacteriol 188:6002-15.

129. Sekura, R. D., F. Fish, C. R. Manclark, B. Meade, and Y. L. Zhang.

1983. Pertussis toxin. Affinity purification of a new ADP-ribosyltransferase. J Biol Chem 258:14647-51.

130. Shimizu, T. 1978. Prophylaxis of Bordetella bronchiseptica infection in

guinea pigs by intranasal vaccination with live strain ts-S34. Infect Immun 22:318-21.

131. Shimizu, T., and H. Ishikawa. 1982. Some characteristics of a urease-

negative, temperature- sensitive strain of Bordetella bronchiseptica as a live, attenuated vaccine. Infect Immun 36:198-201.

132. Skinner, J. A., A. Reissinger, H. Shen, and M. H. Yuk. 2004. Bordetella

type III secretion and adenylate cyclase toxin synergize to drive dendritic cells into a semimature state. J Immunol 173:1934-40.

133. Sloan, G. P., C. F. Love, N. Sukumar, M. Mishra, and R. Deora. 2007.

The Bordetella Bps polysaccharide is critical for biofilm development in the mouse respiratory tract. J Bacteriol 189:8270-6.

134. Spears, P. A., L. M. Temple, D. M. Miyamoto, D. J. Maskell, and P. E. Orndorff. 2003. Unexpected similarities between Bordetella avium and other pathogenic Bordetellae. Infect Immun 71:2591-7.

135. Spilker, T., A. A. Liwienski, and J. J. LiPuma. 2008. Identification of

Bordetella spp. in respiratory specimens from individuals with cystic fibrosis. Clin Microbiol Infect 14:504-6.

136. Stevenson, A., and M. Roberts. 2002. Use of a rationally attenuated

Bordetella bronchiseptica as a live mucosal vaccine and vector for heterologous antigens. Vaccine 20:2325-35.

137. Stevenson, A., and M. Roberts. 2003. Use of Bordetella bronchiseptica

and Bordetella pertussis as live vaccines and vectors for heterologous antigens. FEMS Immunol Med Microbiol 37:121-8.

138. Stockbauer, K. E., B. Fuchslocher, J. F. Miller, and P. A. Cotter. 2001.

Identification and characterization of BipA, a Bordetella Bvg-intermediate phase protein. Mol Microbiol 39:65-78.

139. Storsaeter, J., H. O. Hallander, L. Gustafsson, and P. Olin. 1998.

Levels of anti-pertussis antibodies related to protection after household exposure to Bordetella pertussis. Vaccine 16:1907-16.

45

140. Tuomanen, E., and A. Weiss. 1985. Characterization of two adhesins of Bordetella pertussis for human ciliated respiratory-epithelial cells. J Infect Dis 152:118-25.

141. Uhl, M. A., and J. F. Miller. 1994. Autophosphorylation and

phosphotransfer in the Bordetella pertussis BvgAS signal transduction cascade. Proc Natl Acad Sci U S A 91:1163-7.

142. van der Zee, A., F. Mooi, J. Van Embden, and J. Musser. 1997.

Molecular evolution and host adaptation of Bordetella spp.: phylogenetic analysis using multilocus enzyme electrophoresis and typing with three insertion sequences. J Bacteriol 179:6609-17.

143. Vandamme, P., M. Heyndrickx, M. Vancanneyt, B. Hoste, P. De Vos,

E. Falsen, K. Kersters, and K. H. Hinz. 1996. Bordetella trematum sp. nov., isolated from wounds and ear infections in humans, and reassessment of Alcaligenes denitrificans Ruger and Tan 1983. Int J Syst Bacteriol 46:849-58.

144. von Wintzingerode, F., G. Gerlach, B. Schneider, and R. Gross. 2002.

Phylogenetic relationships and virulence evolution in the genus Bordetella. Curr Top Microbiol Immunol 264:177-99.

145. von Wintzingerode, F., A. Schattke, R. A. Siddiqui, U. Rosick, U. B.

Gobel, and R. Gross. 2001. Bordetella petrii sp. nov., isolated from an anaerobic bioreactor, and emended description of the genus Bordetella. Int J Syst Evol Microbiol 51:1257-65.

146. Weingart, C. L., P. S. Mobberley-Schuman, E. L. Hewlett, M. C. Gray,

and A. A. Weiss. 2000. Neutralizing antibodies to adenylate cyclase toxin promote phagocytosis of Bordetella pertussis by human neutrophils. Infect Immun 68:7152-5.

147. Weingart, C. L., and A. A. Weiss. 2000. Bordetella pertussis virulence

factors affect phagocytosis by human neutrophils. Infect Immun 68:1735-9.

148. Weiss, A. A., and M. S. Goodwin. 1989. Lethal infection by Bordetella

pertussis mutants in the infant mouse model. Infect Immun 57:3757-64. 149. Wolfe, D. N., G. S. Kirimanjeswara, E. M. Goebel, and E. T. Harvill.

2007. Comparative role of immunoglobulin A in protective immunity against the Bordetellae. Infect Immun 75:4416-22.

46

150. Wolff, J., G. H. Cook, A. R. Goldhammer, and S. A. Berkowitz. 1980. Calmodulin activates prokaryotic adenylate cyclase. Proc Natl Acad Sci U S A 77:3841-4.

151. Woolfrey, B. F., and J. A. Moody. 1991. Human infections associated

with Bordetella bronchiseptica. Clin Microbiol Rev 4:243-55.

47

Chapter II

Differential Bvg-Phase Dependent Regulation and Role in

Pathogenesis of Two Bordetella Paralogs

Neelima Sukumar, Meenu Mishra, Gina Parise, Tomoo Ogi and Rajendar Deora

This chapter has been published in Journal of Bacteriology 2007 May;189(10):3695-

704 and is reprinted with permission. Experiments for Figure 9 were done by Dr.

Meenu Mishra. All animal work were performed with the help of Drs. Meenu Mishra

and Gina Parise-Sloan.

48

INTRODUCTION

Bordetellae are small aerobic, gram-negative coccobacilli that colonize the

respiratory tracts of humans and animals (26). Of the three classical species, B.

pertussis infects only humans and causes the acute respiratory disease known as

whooping cough (38). B. parapertussis strains can be divided into two genetically

distinct types, those which infect humans causing a pertussis-like illness, and those

which cause respiratory infections in sheep (26, 36). In contrast to the former two

species, B. bronchiseptica has a broad host range infecting a variety of nonhuman

animals (17, 26). It typically establishes asymptomatic infections but can cause

atrophic rhinitis in pigs, kennel cough in dogs, snuffles in rabbits and

bronchopneumonia in guinea pigs (17).

Efficient and productive colonization of the respiratory tract by Bordetella

requires interplay of multiple factors that allow bacterial adherence to the respiratory

epithelium leading to the eventual development of disease. The majority of these

virulence determinants are regulated by a two component regulatory system known

as BvgAS. BvgA is a DNA-binding response regulator and BvgS is a transmembrane

sensor protein kinase (26, 37). On perception of a signal, BvgS gets

autophosphorylated at a histidine residue (34). Phosphorylation of BvgA at a

conserved aspartate residue then occurs through a series of phosphotransfer

reactions from BvgS (5, 35). Phosphorylated BvgA (BvgA-P) can bind to the cognate

promoter regions of the Bvg-regulated genes and mediates activation and/or

repression of transcription (4, 29). A striking feature of the BvgAS regulatory circuitry

is its ability to control the transition among multiple phenotypic states of which at least

49

three phenotypic phases, the Bvg+, Bvg- and Bvgi (Bvg-intermediate) are known.

Each of these phases is characterized by the differential expression of known Bvg-

regulated gene products (9, 12). For example, during the Bvg+ phase a variety of

Bvg-activated factors including adhesins and toxins are maximally expressed and the

Bvg-repressed genes are minimally expressed (9, 12). For both B. pertussis and B.

bronchiseptica, it has been demonstrated that the Bvg+ phase is necessary and

sufficient for respiratory tract colonization (1, 10)

Transition to the Bvg- phase occurs as a result of either mutational inactivation

of BvgAS or growth in the presence of modulating signals (sulfate anion, nicotinic

acid or growth at low temperature). This phase is characterized by expression of the

Bvg-repressed factors (e. g. flagella in B. bronchiseptica and outermembrane

proteins of unknown function in B. pertussis) and the repression of Bvg-activated

genes (26). It has been suggested that this phase may be responsible for survival of

B. bronchiseptica in the environment (9). The Bvgi phase is expressed either as a

result of specific genetic mutations in BvgS or by growth of wild type (wt) Bordetella

strains in the presence of semi-modulating concentrations of chemical signals (9).

The Bvgi phase is principally distinguished by the maximal expression of a set of

antigens of which BipA (Bordetella intermediate phase protein A) is the first to be

identified at the molecular level (12, 33). The role of the Bvgi phase in Bordetella

infectious cycle is presently unclear.

BipA shares at its amino terminus similarity with intimin proteins of

enteropathogenic (EPEC) and enterohemorrhagic (EHEC) E. coli and with invasins of

Yersinia species, thereby leading to suggestions that BipA plays a role in the

50

Bordetella infectious cycle (33). To date a significant demonstrable role for BipA in

Bordetella pathogenesis has been elusive.

In this study, we report the identification of a B. bronchiseptica ORF termed

bcfA, which is a paralog of bipA. We demonstrate that the Bvg-phase dependent

expression profile of bcfA is strikingly different from that of bipA. In contrast to

maximal expression of bipA in the Bvgi phase, bcfA is expressed at high levels both

in the Bvg+ and the Bvgi phases. We have identified DNA sequences similar to the

consensus BvgA binding site in the region upstream of the bcfA ORF. Utilizing

electrophoretic mobility shift assay (EMSA), we observed differences in the nature of

DNA binding between BvgA and BvgA-P to the bcfA promoter region. Higher order

BvgA-DNA complexes were observed in the presence of acetyl phosphate. Our

results also document that BcfA is an outer membrane protein and that it is

expressed during Bordetella infection of rats. By comparing strains, harboring single

and double deletion mutations in bipA and bcfA, in intranasally infected rats, we show

that BipA and BcfA have an overlapping function in mediating efficient colonization of

trachea.

51

MATERIALS AND METHODS

Bacterial strains, plasmids, media and growth conditions. The bacterial strains

and plasmids used in this study are listed in Table 1. Bordetella strains were

maintained on Bordet Gengou (BG) agar (Becton Dickinson Microbiology Systems,

MD, USA) supplemented with 7.5% defibrinated sheep blood. For RNA extraction

and β-galactosidase assays, cells were grown in Stainer Scholte (SS) broth at 37°C

with shaking (32). E. coli strains were grown in Luria-Bertani medium at 37°C with

shaking. The different growth media were supplemented with appropriate antibiotics

(μg/ml) as needed, ampicillin, 100; chloramphenicol, 50; gentamycin sulfate 25;

streptomycin, 50; kanamycin, 25.

LacZ transcriptional fusions and β-galactosidase assays. For constructing lacZ

transcriptional fusion, a 640 bp DNA fragment consisting of 445 bp upstream and 195

bp downstream of the bcfA translational site was amplified from B. bronchiseptica

using BcfA5 and BcfA6 primers. The PCR fragment was cloned as a blunt ended

fragment into the EcoRV site of linearized pSTBlue-1 vector (Novagen, USA)

resulting in plasmid pRKD22. This plasmid was subsequently digested with EcoR1

and the resultant fragment was then cloned into EcoR1 site of the suicide plasmid

pEGZ (13) leading to the creation of the fusion plasmid pRKD23. This places the bcfA

promoter fragment upstream of a promoterless lacZ gene. pRKD23 was then

integrated into the genome of the different Bordetella strains by a single crossover at

the bcfA locus as described previously (12). β-galactosidase assays were performed

as previously described (11, 12). For detecting the β-galactosidase units under

52

TABLE 1. Strains and plasmids used in this study Strain/Plasmid Description ReferencesRB50 RB53 RB53i RB54 SM10λpir pSTBlue-1 pRKD22 pEGZ pRKD23 pRE112 pRKD40 pET24(a) pNS101 BL21(DE3)pLysE RKD101(ΔbcfA) RB25 (ΔbipA) MM101 (ΔbipAΔbcfA)

Wt B. bronchiseptica strain Bvg+ phase locked derivative Bvgi phase locked derivative Bvg- phase locked, ΔbvgS th-1 thr leu tonA lacY supE, recA::RP4–2-Tc::Mu Km

R (λ pir) Blunt end cloning vector; Novagen pSTBlue-1 derivative containing the 640 bp promoter region of bcfA cloned into the EcoRV site of pSTBlue1 lacZ transcriptional fusion vector pEGZ derivative; EcoRI fragment from pRKD22 cloned into EcoRI site of pEGZ upstream of promoter less lacZ Allelic exchange vector; Cmr

pRE112 derivative, bcfA deletion plasmid T7 based expression plasmid, Novagen pET24(a) derivative, BcfA overexpression plasmid Competent cells, RB50 derivative with in frame chromosomal deletion of bcfA RB50 derivative with in frame chromosomal deletion of bipA ΔbipA derivative with in frame chromosomal deletion of bcfA

(9) (9) (9) (9) (31)

This study (13) This study (14) This study This study This study (33) This study

53

conditions where the BvgAS system is modulated, the wt strain was grown in the

presence of 40 mM MgSO4.

RNA isolation and Real Time RT-PCR. Total RNA was isolated using the RNeasy

kit from Qiagen (MD, USA), treated with RNase free DNase-1 (Invitrogen, CA, USA)

to degrade contaminating DNA. 2-3 μg of RNA was primed with random hexamers

(Invitrogen, CA, USA) and cDNA was prepared utilizing the reverse transcriptase

enzyme (RT) Superscript III (Invitrogen, CA, USA) following manufacturer’s protocol.

Specific primers (Table 2) for various genes were designed using the ABI PRISM

Primer Express software (PE Applied Biosystems) in order to obtain similar-sized

amplicons. Diluted (1:25) reverse-transcription products obtained as above and a

blank control without cDNA were used as template and amplified using the TaqMan

Universal PCR Master mix (Applied Biosystems). The Master mix contains dNTPs

with dUTP, AmpliTaq Gold DNA polymerase, Amperase UNG, optimized buffer and a

passive reference dye. For each PCR, a mixture (20 μl) containing template cDNA, 1x

Master Mix, 250 nM of each sense and antisense primer and 500 nM of the TaqMan

probe was placed in 96-well optically clear PCR plates (Greiner Bio-one, USA).

Amplification and detection of the specific products were performed on an ABI PRISM

7000 Sequence Detection System (PE Biosystems) with the following cycle protocol:

one cycle at 50°C for 2 min and one cycle at 95°C for 10 min followed by 40 cycles at

95°C for 15 s and 60°C for 1 min. The threshold value was set manually and kept

constant for all comparison groups. To control for variation in RNA quantity and

quality, the constitutively expressed Bordetella gene recA was used as the

54

endogenous reference control. The relative quantification of gene expression was

performed using the comparative CT (threshold cycle) method according to the

manufacturer's instructions (User Bulletin # 2: ABI PRISM Sequence Detection

System). The critical threshold cycle (CT) was defined as the cycle at which the

fluorescence became detectable above the background fluorescence, and was

inversely proportional to the logarithm of the initial number of template molecules. The

ΔCT value was determined by subtracting recA CT value from the gene-specific CT

values. The ΔΔ CT was calculated by subtracting the obtained ΔCT value with the

ΔCT-calibrator value. For measurement of the relative levels of Bvg-activated genes

bcfA, fhaB (encodes filamentous hemagglutinin), bipA and cyaA (encodes adenylate

cyclase toxin), the ΔCT values obtained from the Bvg- phase locked cells was

designated as the calibrator, since in this strain there is lowest expression of these

target genes. Similarly, for measurement of the relative levels of the Bvg-repressed

gene flaA (codes for flagellin), the ΔCT values obtained from the Bvg+ phase locked

cells was designated as the calibrator, since there is lowest expression of flaA in this

strain. We also performed a validation experiment to ensure that the efficiencies of

the target amplification and that of the reference amplification were similar by

determining the ΔCT values with template dilution. Data from at least three

measurements each carried out on at least two different batches of RNA were

plotted. Error bars represent standard deviation from independent values.

Time point analysis. To modulate bvg-activity, the wt B. bronchiseptica strain RB50

was grown in 50 ml of SS medium in the presence of 40 mM MgSO4 for 18 h. At time

55

TABLE 2. Oligonucleotide primers used in this study a,b

Primer Sequence/Reference BcfA1 BcfA2 BcfA3 BcfA4 BcfA5 BcfA6 BcfA7 BcfA8 BcfA9 BcfA10 BcfA11 BcfA12 BcfA13 FhaF FhaR BipAF BipAR PrnF PrnR RecF RecR RecA9 RecA10 RecA11 Fha9 Fha10 Fha11 Fla9 Fla10 Fla11 BipA109 BipA110 BipA111

5’-CTAGTCTAGACCTACATATCCGTAGGATTG -3’ 5’-TTGGCGCGCCGTCTGCGCGACCCGCAGCAT -3’ 5’-TTGGCGCGCCTGATCGGTGGCGAGGGCGGC-3’ 5’-CGGGGTACCCTTGAACAGCGGCAGCACGTCG -3’ 5’-CGCATGGGATTCTCCCGGGTA-3’ 5’-GCTGTCGGCGGCATCCTGGCGATC -3’ 5’-GGCGAGCGCGACCGCGTCTTGCT-3’ 5’-GGCTGAGATTGACGCCAAGCTGCA-3’ 5’-CGCCATGCCTTGAA-3’ 5’-[DFAM]-GGGCGCCAACGCATT-[DTAM]-3’ 5’-TGGTTCTTGCCGAA-3’ 5’-CGCGGATCCGTGAAGCAAGCCATCCACG-3’ 5’-CCCAAGCTTCCCAGCAGGCCGCCCTC-3 5’-ATCCGACCTACACCGAATGG-3’5’-GAGTGTGCGCCGATTTTCAG-3’5’-GGGTCTGCCCTTTCTGCGCAATCTG-3’5’-GCCGATCACCTTGGTCTGCTCCAG-3’5’-GCCGCTGCAGCCGGAAGACCTTC-3’ 5’-GCAGCTCCGTCGCGACGATGTCG-3’ (13) (13) 5’-ACGTGCAATACGCCTCCAA-3’ 5’- [DFAM]-TGGGCGTCAACCTGACCGACCT-[DTAM]-3’ 5’-TGTCCGGCTGGGAGATCA-3’ 5’-TGTCCGCCATGGAGTATTTCA-3’ 5’-[DFAM]-CCCGGTGAGCCTGACAGCCCT-[DTAM]-3’ 5’-CCAGCAGATAATCCAGGAGTTCAT-3’ 5’-GATCCAGCAGGAAGTCAACCA-3’ 5’- [DFAM]-AAATCAACCGCATCGCCGAGCA-[DTAM]-3’ 5’-GACCTGATGCCGTTGAAGTC-3’ 5’-GGCCCAGGTCAATGATGTCTT-3’ 5’-[DFAM]-AACCTGGCTCGGGAATCGGGTC-[DTAM]-3’ 5’-CCTTGCAGATTGCGCAGA-3’

a Sequences in bold represent restriction enzyme sites b [DFAM]- 6 carboxyfluorescein: Reporter fluorochrome. b [DTAM]- 6-carboxy-tetramethyl-rhodamine: Quencher fluorochrome.

56

zero, the culture was spun down and re-suspended in 100 ml of SS media lacking

MgSO4. At times 5 min, 30 min, 1 h and 4 h, 5 ml of the bacterial culture was utilized

for total RNA preparation as described above.

After reverse transcription of the RNA, an aliquot of cDNA (5%) was used as

template in subsequent RT-PCR. In order to eliminate possible interference by

genomic DNA, mock reactions without RT were also performed. RT-PCR was carried

out as previously described (11, 12) with gene-specific primers. Primer pairs used are

listed in Table 2. Genomic DNA, prepared from the wt strain RB50, was used as the

positive control for PCR. Aliquots of amplified products obtained, were

electrophoresed on 1% agarose gels. Images of the ethidium-bromide-stained gels

were captured by Alpha Innotech Gel Doc System (Alpha Innotech Corporation).

Electrophoretic mobility shift assays. The purified DNA fragment was end labeled

by T4 polynucleotide kinase (New England Biolabs, MA) with [γ-32P]ATP (Amersham

Biosciences, NJ). Unincorporated radioactivity was removed by passage through G-

50 quick spin columns (Amersham Biosciences, NJ). Each reaction (20 μl), contained

indicated amounts of purified BvgA or BvgA-P and the radiolabeled promoter DNA in

1x binding buffer (10 mM Tris-HCl (pH 7.8), 2 mM MgCl2 50 mM NaCl, 1 mM DTT,

0.5 μg of poly (dI-dC), 0.01% NP-40, 100 ng of BSA and 10% glycerol).

Phosphorylation of BvgA was carried out as described earlier by incubation of the

protein at room temperature for 15 min in 1x binding buffer containing 20 mM acetyl

phosphate (11, 29). The reaction mixtures were incubated at 37oC for 15 min to allow

57

binding of BvgA/BvgA-P to radiolabeled promoter. The samples were

electrophoresed and visualized by autoradiography as previously described (11, 29).

For non-specific competition, a DNA fragment corresponding to the internal

region (+453 to +978) of bcfA ORF was amplified using primers BcfA7 and BcfA8.

Indicated concentrations of this fragment were included in the binding reactions. For

specific competition, bcfA promoter fragment was amplified using primers BcfA5 and

BcfA6 and indicated concentrations of the unlabeled DNA fragment was included in

the reaction.

Overexpression, purification of BcfA and antibody production. The entire bcfA

ORF excluding the stop codon was cloned as a BamHI-HindIII fragment in similarly

digested pET24(a) plasmid (Novagen), thereby resulting in the BcfA overexpression

plasmid pNS101. Amplification was carried out using recombinant Pfu DNA

polymerase (Stratagene, CA). E. coli BL21(DE3)pLysE cells containing pNS101

were grown in 2XTY (16 g of Bacto tryptone, 10 g of yeast extract, and 5 g/liter NaCl

plus 0.4% glucose) in the presence of kanamycin and chloramphenicol at 37 °C. The

cells were grown to an O.D.600 of 0.8-1.0 and induced by IPTG for 2 h. The cells

were harvested by centrifugation, resuspended in 30 ml of TGED buffer (10 mM

Tris-HCl (pH 7.9), 0.1 mM EDTA (pH 8.0), 0.2 mM DTT, and 0.05% sodium

deoxycholate, 5 % glycerol and 2 mM PMSF), incubated on ice for 20 min, and lysed

by passing through a French pressure cell three times at 14,000-16,000 p.s.i. The

lysate was centrifuged and the soluble BcfA was purified from the clarified lysate

58

using the T7. Tag affinity purification kit (Novagen) according to the manufacturer’s

protocol.

For antibody production, the cells after induction, were resuspended in the

TGED buffer, lysed by passing through French pressure cell and centrifuged. The

pellet was washed once with TGED buffer containing 0.5M NaCl, 10 ml of 1x

Bugbuster protein extraction reagent (Novagen), 10 μl of Benzonase nuclease and 1

mg/ml of lysozyme. After centrifugation, the cell pellet was resuspended in 5 ml of

TGED buffer containing 6M guanidine hydrochloride and incubated on ice for 10

min. Another 5 ml of cold TGED buffer was added and the suspension was

incubated on ice for an additional 10 min. The lysate was spun down and the

supernatant was dialyzed in 2 l of cold TGED buffer at 4oC for 18 h. The dialyzate

was centrifuged and the pellet containing overproduced BcfA was resuspended in

protein loading buffer (50 mM Tris-HCl (pH 6.8), 2 % SDS, 10 % glycerol and 5 mM

DTT) and was electrophoresed on a standard SDS polyacrylamide protein gel. The

band corresponding to the BcfA protein was excised and utilized for production of

anti-BcfA antibodies in rats. Antibody production was carried out on a fee for service

basis by Covance.

Preparation of Cellular Fractions. Stationary phase cultures of the different strains

were centrifuged at 17,000 rpm, and the cell pellets were resuspended in the cell

disruption buffer (10mM Tris-HCl (pH 8.0), 20% sucrose, 1 mM EDTA and 0.1 mg/mL

lysozyme). After incubation on ice for 10 min, the samples were frozen in dry ice

followed by thawing in cold water. The bacterial cells were sonicated on ice,

59

centrifuged initially at 3,700 rpm for 10 min to pellet unlysed cells. The clarified

suspension obtained was spun at 17,000 rpm for 1 h and the pellet resuspended in

an appropriate volume of 1x PBS. For separation of inner and outer membrane

proteins, the membrane fractions were incubated with 2% Triton X-100 for 30 min on

ice followed by centrifugation at 17,000 rpm for 1 h. The pellet (Triton X-100-

insoluble) consisting of the outer membrane proteins was utilized for SDS-PAGE

analysis.

SDS-PAGE and immunoblot analysis. Proteins from outer membrane fractions or

purified recombinant BcfA were separated on SDS-polyacrylamide gels, transferred

to nitrocellulose membranes and probed with anti-BcfA (1:5,000) or anti-BipA

antibody (1:5,000 dilution) (33) or rat serum (1:1,000 dilution) harvested from rats

infected with the indicated strains 30 days post-inoculation. The secondary antibody

(1:2,000) used, was either the goat anti-rat IgG or the goat anti rabbit IgG conjugated

to horseradish peroxidase. For detection of proteins, the Amersham ECL Western

blotting system was utilized.

Construction of deletion strains. Based on the pre-annotated sequence

information of the wt B. bronchiseptica strain RB50 present in the Sanger Center

database, we had designed primers to delete a region encompassing amino-acids

10-882 of the then 903 amino acid long bcfA ORF. Subsequent to the construction of

the deletion strains and the testing of these strains in animal models, we found that in

the updated database, the bcfA ORF was extended to include an upstream in frame

60

stretch of nucleotides encoding 66 amino acids. Thus, based on the current

annotation, the in-frame deletion of bcfA encompasses a region corresponding to

amino-acids 76-958 of the 969 amino acid long ORF.

A XbaI-AscI fragment (407 bp) containing sequences corresponding to the 5’

end including the first 75 amino acids of the bcfA ORF was amplified from the

chromosome of RB50 using primers BcfA1 and BcfA2. A second AscI-KpnI fragment

(525 bp) containing sequences corresponding to the 3’ end of bcfA including the last

11 amino-acids of the bcfA ORF was also amplified using primers BcfA3 and BcfA4.

These fragments were digested with respective restriction enzymes and were used

for three way ligation with Xba1 and Kpn1 digested suicide vector pRE112 (Cmr) (14)

resulting in plasmid pRKD40. This plasmid was transformed into SM10λpir and

mobilized from this strain into RB50. After conjugation, cointegrants were selected on

BG agar containing chloramphenicol and streptomycin. Colonies arising from second

recombination events were selected on LB-agar containing 7.5% sucrose as

described previously (11, 14). The genotype of the deletion strain RKD110 (ΔbcfA)

was confirmed by PCR and subsequent DNA sequencing.

MM101, the ΔbipAΔbcfA strain was constructed essentially as described

above for the ΔbcfA strain except that the parental strain used was the previously

described ΔbipA strain, RB25 (33).

Rat colonization Experiments. Four-five week old female Wistar rats (Charles

River Laboratories) were lightly anesthetized with halothane and intranasally

inoculated with 5 μl of sterile PBS alone or 40-100 CFUs of either the wt or the

61

various mutant strains. The number of delivered CFUs was confirmed by plating on

BG agar containing streptomycin (50μg/ml). Colonization levels in the respiratory

tract were determined by sacrificing rats 12 or 30 days post inoculation and

removing the entire nasal septum and 1cm of trachea. These were then

homogenized in 200 μL of sterile PBS and various dilutions of the homogenate were

plated on BG agar containing streptomycin to determine the number of CFUs.

62

RESULTS

Identification of BcfA, a parolog of BipA in B. bronchiseptica. It was previously

reported that a BipA-deficient strain of B. bronchiseptica colonized the rabbit

respiratory tract as efficiently as the wt strain, suggesting that this protein is not

essential for colonization (33). Since Bordetellae express multiple proteins that have

either been demonstrated or been predicted to be involved in attachment to the

respiratory epithelium (26), we hypothesized that the absence of BipA in B.

bronchiseptica could be compensated by either a known or yet to be identified

alternative factor, thereby leading to no apparent effect on colonization. In particular,

we explored the possibility of a paralogous protein in B. bronchiseptica. We searched

the recently sequenced B. bronchiseptica genome database for ORFs homologous to

BipA. BLASTP (2) searches revealed the presence of an ORF (BB0110) that

displayed 49% identity in N-terminal 809 amino acids to the BipA protein. In addition,

these searches also revealed similarity of BB0110 to intimins from EPEC and EHEC

(8, 15), invasins from Yersinia species (21) and the putative E. coli adhesin EaeH

(accession number AAZ57201). In the annotated B. bronchiseptica genome

database, the protein encoded by BB0110 has been designated as a putative

adhesin and based on our finding that it plays a role in respiratory tract colonization

(see below), we have designated this ORF as bcfA, Bordetella colonization factor A.

The B. bronchiseptica BcfA protein. The bcfA ORF has the potential to code for a

969 amino acid protein with a predicted molecular weight of 102 kDa (Fig. 1). The

63

FIG. 1. Schematic representation of the similarity of BcfA with BipA. Alignment of

amino acid sequences of BipA and BcfA was performed using the Needleman-

Wunsch global alignment algorithm by utilizing EMBOSS Needle available at

European Bioinformatics institute website. The regions of homology between these

two proteins are indicated by the dashed regions. The putative signal sequences are

represented by the shaded region. The black rectangle represents the region of BipA

and BcfA that displays similarity to intimins and invasins.

64

969

1578

BcfA

1

461

491 671 1158

BipA

A B

1

65

BipA protein of B. bronchiseptica is comparatively a larger protein having 1578 amino

acids and a molecular weight of 164.5 kDa (Fig. 1). Analysis of the predicted BcfA

protein sequence by SignalP 3.0 method (3) revealed that it contains an unusually

long signal sequence of 44 amino acids (signal peptide probability of 0.942) (Fig. 1),

including the characteristic positively charged N-region, the hydrophobic core H-

region and the C-region with a consensus cleavage site of a sec-dependent leader

peptide. The predicted cleavage site of the signal sequence is located between Ala44

and Gln45 residues (cleavage site probability of 0.928). The presence of a canonical

signal sequence in BcfA suggests that this protein can traverse the inner membrane.

In contrast to BcfA, the BipA protein lacks a canonical signal sequence.

Alignment of the amino acid sequences of BcfA and BipA revealed two main

regions of homology (Fig. 1). The N-terminal region of BcfA (amino acid residues 11-

461) is similar to N-terminal region of BipA (residues 29-491). The N-terminal regions

of BcfA and BipA are also similar to intimins from enteropathogenic and

enterohemorrhagic E. coli and invasins from Yersinia spp (Fig. 1). This region is

predicted to anchor these proteins in the outer membrane (25). The second region of

significant similarity between BipA and BcfA extends from amino acid residues 473-

969 of BcfA and 671-1158 of BipA (Fig. 1). The BipA protein possesses a 291 amino

acid long C-terminal surface-exposed domain and it appears that this domain is

lacking in BcfA (33).

Expression of bcfA is regulated by BvgAS in B. bronchiseptica. The BvgAS

locus regulates the expression of the majority of virulence gene expression in

66

Bordetella (26). The homology of bcfA with bipA, which is transcriptionally activated

by BvgAS (12, 33), prompted us to determine whether BvgAS controlled bcfA

expression. We cloned a DNA fragment, encompassing regions both upstream and

downstream of the predicted bcfA start codon, upstream of the promoter-less lacZ

gene contained in the suicide plasmid pEGZ (13). The resultant bcfA-lacZ fusion was

integrated into different strains of B. bronchiseptica by homologous recombination.

Measurement of the β-galactosidase activities from log-phase cultures of these

strains indicated that while bcfA is expressed at high levels in the wt strain, it is

expressed at low levels in the Bvg- phase locked strain (Fig. 2). Growth of the wt

strain in the presence of MgSO4, a known modulator of BvgAS activity, resulted in

very low levels of β-galactosidase activity (Fig. 2). Taken together these results

suggest that expression of bcfA is positively regulated by BvgAS.

bcfA and bipA demonstrate distinct Bvg-phase-dependent expression profile.

Previously, we and others have shown that the various Bvg-regulated genes exhibit

differences in their phase-dependent expression profiles (9, 12). Specifically, we have

demonstrated that in contrast to other known Bvg-regulated genes, bipA is expressed

maximally in the Bvgi phase and at low levels in the Bvg+ phase (12). To compare the

phase-dependent expression profile of bcfA directly with that of bipA and other Bvg-

regulated genes, we performed real time RT-PCR assays. We isolated total RNA

from Bvg+, Bvgi and Bvg- phase locked strains and generated cDNA through reverse

transcription. Primers and probes were designed to anneal to internal regions of

67

FIG. 2. Quantification of bcfA transcriptional activity by β-galactosidase assays in the

wt, Bvg+, Bvgi and Bvg- phase locked strains of B. bronchiseptica. The bcfA-lacZ

fusion was integrated in the various strains and the β-galactosidase activity was

determined after growth to mid-log phase in SS broth at 37°C. For modulating the

BvgAS activity, wt strain carrying bcfA-lacZ fusion was grown in the presence of 40

mM MgSO4. Error bars represent standard deviation.

68

0

200

600

1000

1400

wt Bvg+ Bvg

iBvg

-+MgSO4

β-ga

lact

osid

ase

units

Strains

69

FIG. 3. Determination of the phase-dependent expression profiles of different Bvg-

regulated genes in B. bronchiseptica by real time RT-PCR analysis. bcfA (A); bipA

(B); fhaB (C); cyaA (D) and flaA (E). cDNA prepared from various strains was used

as template for real time RT-PCR as described in the Materials and Methods. The

ΔCT value was determined by subtracting recA CT value from the gene-specific CT

values. ΔΔCT values for each gene was obtained by subtracting the ΔCT value of

either the Bvg- phase locked strain (for bcfA, bipA, fhaB, cyaA) or the Bvg+ phase

locked strain (for flaA) from the gene-specific ΔCT value. The relative expression

levels of respective genes (y-axes) were then calculated by the formula 2-ΔΔCT. Thus,

for bcfA, bipA, fhaB and cyaA the relative level of expression in the Bvg- phase locked

strain is 20=1. Similarly, the expression of flaA in Bvg+ phase locked strain is also 1.

Bars indicate standard deviation of three independent values obtained from two

different RNA batches. Y axes represent the relative expression levels of the

indicated gene.

70

cya

02468

10 12 14 16 18

0123

4567

bipA

B

0102030405060708090

bcfA

A

fha

02040

6080

100120

140C

D

Bvg+

Bvgi

Bvg-

Rel

ativ

e ex

pres

sion

Rel

ativ

e ex

pres

sion

Rel

ativ

e ex

pres

sion

0

2

4

6

8

10

12 fla

ER

elat

ive

expr

essi

on

si

onve

es

expr

Rel

ati

71

respective genes (Materials and Methods and Table 2). Expression levels of recA, a

Bvg-independent gene, were used as normalization controls. As shown in Fig. 3A,

bcfA was expressed at high levels in both the Bvg+ and the Bvgi phases. bcfA specific

RNA was barely detectable from Bvg- phase locked samples thereby confirming the

results of the β-galactosidase assays above. Strikingly, these results demonstrate

that the two paralogs bcfA and bipA have differential Bvg-dependent expression

profiles, since bipA is maximally expressed in the Bvgi phase and at a low level in the

Bvg+ phase (Fig. 3B and (11, 12)). Our data further suggest that the expression

profile of bcfA resembles that of the previously known Bvg-activated gene fhaB (6, 9).

As shown in Fig. 3C and consistent with previously published data, fhaB is expressed

at a high level in both the Bvg+ and Bvgi phases (9). As expected from previous

results (9, 12), cyaA and flaA were expressed maximally only in the Bvg+ and the

Bvg- phases (Figs. 3D and Fig. 3E), respectively, thereby further corroborating the

validity of the real time RT-PCR assays in accurately measuring the expression

patterns of Bvg-regulated genes.

We independently confirmed the phase-dependent expression pattern of bcfA

by measuring the β-galactosidase activities from log-phase cultures of the Bvg+ and

the Bvgi phase locked strains. Our results (Fig. 2) indicate that bcfA is expressed at

high levels in both the Bvg+ and Bvgi phases.

Kinetics of transcriptional activation of bipA and bcfA. Previous studies have

revealed the differential kinetics of expression of bvg-regulated genes upon induction

of the BvgAS activity either due to a temperature-shift (25°C to 37°C) or by first

72

growing in the presence of chemical modulators followed by growth in their absence

(22, 24). We compared the transcriptional activation of bcfA in B. bronchiseptica over

time following the induction of the BvgAS system. RB50 cells were modulated first by

growing them in the presence of 40 mM MgSO4 for 18h. At time zero, cells were spun

down and re-suspended in media without MgSO4 for the induction of the BvgAS

system. After 0, 5, 30, 60 and 240 min. of induction, total RNA was isolated from cells

and kinetics of promoter activation was determined by RT-PCR using primers specific

for fha, prn, bipA and bcfA. We also utilized primers specific for a bvg-independent

gene recA, as a normalization standard. RT-PCR assays revealed that of all the bvg-

regulated genes tested, only fhaB was transcribed at very low levels, suggesting very

efficient modulation of the BvgAS activity by overnight growth in presence of MgS04.

It was also not possible previously, using a slightly different protocol to completely

inhibit the expression of fhaB. As shown in Fig. 4, fhaB and bipA were reproducibly

activated as early as 5 min after induction and prn was transcribed after 30 min of

induction (Fig. 4). In accordance with our results, it has been shown previously that

the expression of fhaB and bipA was activated almost immediately after the switch to

inducing conditions, whereas there was a delay in the expression of prn (22, 24). Our

results clearly demonstrate that transcriptional activation of bcfA was reproducibly

observed within 5 min of growth under inducing conditions, suggesting that its

activation kinetics parallels to that of fhaB and bipA (Fig. 4). Control reactions were

performed to confirm the specificity and the reliability of the RT-PCR assay. No

detectable PCR products were obtained in the absence of RT indicating that

contaminating DNA was absent from the RNA preparation (data not shown).

73

FIG.4. Kinetics of transcriptional activation of different Bvg-activated genes and the

Bvg-independent gene recA. Wt strain was grown in liquid cultures under modulating

conditions (in the presence of 40 mM MgSO4) for 18h, spun and re-suspended in

media without MgSO4. Total RNA was isolated at indicated times and cDNA was

prepared as described in the Material and Methods. RT-PCR was used to detect the

transcript levels of indicated genes.

74

fha

5 30 60 240

+MgS

O4

bcfA

bipA

recAg0

Time (min)

prn

fha

5 30 60 240

+MgS

O4

bcfAbcfA

bipA

recArecAg0

Time (min)

prn

75

Identification of BvgA box like sequences upstream of the bcfA ORFs. Based on

our results demonstrating BvgAS-mediated activation of bcfA expression (Figs. 2 and

3), we hypothesized that regions upstream of bcfA ORF will harbor BvgA binding

sites. Scanning of the sequences upstream of the predicted translational start site of

the bcfA ORF revealed the presence of two directly joined inverted repeats (IR2 and

IR3) that were similar to previously identified BvgA binding sites (Fig. 5) and (6, 12).

In addition to the conjoined inverted repeat, we also detected additional overlapping

inverted repeat sequences (IR1) where the two halves were separated by multiple

nucleotides (Fig. 5). For some Bvg-regulated promoters it has been shown that BvgA

binding repeats are not conjoined but are separated by multiple nucleotides-2 for

cyaA, 10 for ptx (encodes for pertussis toxin) and 27 and 37 nucleotides respectively

for bipA (7, 11, 12, 23). The presence of BvgA binding sites upstream of the ORFs

combined with our finding that bcfA is positively regulated by BvgAS suggests that

BvgAS control of bcfA expression is direct.

BvgA and BvgA-P bind to promoter region of bcfA. The presence of putative

BvgA binding elements upstream of bcfA ORF led us to speculate that BvgA will bind

the bcfA promoter. To demonstrate BvgA occupancy to the bcfA promoter, we

conducted EMSA by utilizing a PCR fragment that was essentially similar to the

fragment used for β-galactosidase assays (see above and Materials and Methods).

Purified recombinant BvgA and BvgA phosphorylated in vitro with acetyl phosphate

(BvgA-P) were used for this analysis. These reactions also contained poly(dI-dC) as

a nonspecific competitor DNA. The results of EMSA are shown in Fig. 6A. Both BvgA

76

FIG. 5. The putative promoter region of bcfA. Arrows on top of the individual DNA

sequences designate sequence elements (bold) that are similar to the consensus BvgA

binding site. The predicted translational start codon is italicized.

77

IR1

IR2

IR3

ACAGAAACCTGAATTTAATGGAGTTCCTGTCATCACGAAAGGTTCCATTTTTTTGTGCACTGGCGCTTTTGGTCTGGCTGCCTA

TCAAGGAAATATCCTACATATCCGTAGGATTGGGCTGGCATCAGGACCGCACGGATCCAAATAATCTTCCTACATCGATTCTCC

GATATGTCTGCATAGCTCACGGGTTGGCAGGTGTTGTCGGTCCAACCAAGGGCCGTATCCGCCGGAAGTCGAAATTTCCTCTCT

GCCATTTCCTCCGATCGCCACGCCGGCGACCGGCGAGTCTTCTTGCGCCCTGAATGGCGCATTCCAAGGGCCACGGCCCGCGGG

CATGGCGGATCGATCGTGAACTTGGGAAAGGAGTAATCCGTGAAGCAAGCCATCCACGCCGTTGCGTTCCGCCATGATGCGCTC

HS1

IR1

IR2

IR3

ACAGAAACCTGAATTTAATGGAGTTCCTGTCATCACGAAAGGTTCCATTTTTTTGTGCACTGGCGCTTTTGGTCTGGCTGCCTA

TCAAGGAAATATCCTACATATCCGTAGGATTGGGCTGGCATCAGGACCGCACGGATCCAAATAATCTTCCTACATCGATTCTCC

GATATGTCTGCATAGCTCACGGGTTGGCAGGTGTTGTCGGTCCAACCAAGGGCCGTATCCGCCGGAAGTCGAAATTTCCTCTCT

GCCATTTCCTCCGATCGCCACGCCGGCGACCGGCGAGTCTTCTTGCGCCCTGAATGGCGCATTCCAAGGGCCACGGCCCGCGGG

CATGGCGGATCGATCGTGAACTTGGGAAAGGAGTAATCCGTGAAGCAAGCCATCCACGCCGTTGCGTTCCGCCATGATGCGCTC

HS1

78

and BvgA-P retarded the mobility of the radiolabeled DNA fragment. As a result of

phosphorylation, occupancy of BvgA to the radiolabeled DNA fragment gave rise to

higher order protein-DNA complexes suggestive of multiple binding sites or

multimerization of BvgA as a result of phosphorylation (Fig. 6A, compare lanes 2-6

with lanes 8-12, respectively). In order to confirm the specificity of BvgA binding, we

also performed competition EMSAs with increasing concentrations of specific (Fig.

6B) and non specific competitors (Fig. 6C). For specific competition, we used

unlabeled bcfA promoter fragment and as a non specific competitor the DNA

fragment corresponding to +453 to +978 of the bcfA coding region was utilized. There

was a gradual loss of BvgA binding activity with increasing concentrations of the

specific competitor (Fig. 6B). In contrast, in the presence of the non-specific

competitor, there was no significant loss in the binding affinity of BvgA-P. Therefore,

these results suggest that interaction of BvgA-P to bcfA promoter is DNA sequence

specific.

BcfA is an outer membrane protein. Due to the homology of BcfA to BipA,

particularly in the N-terminal region (Fig. 1), we hypothesized that BcfA will also be

localized to the outer membrane. We constructed two isogenic strains, RKD110

(ΔbcfA), and a double deletion strain, MM101 (ΔbipAΔbcfA). These strains are in-

frame deletions of the respective genes and are non-polar. Notably, these strains are

comparable to the wt strain with respect to growth in laboratory cultures (data not

shown). We purified the outer membranes from the wt, Bvg- phase locked strain,

RKD110 (ΔbcfA), RB25 (ΔbipA) and MM101 (ΔbipAΔbcfA), as the Triton X-100

79

FIG. 6. Electrophoretic Mobility Shift Assay.

A. 32P-end labeled bcfA promoter fragment was incubated with varying

concentrations of either BvgA (lanes 2-6) or BvgA-P (lanes 8-12). The DNA-protein

complexes were separated by electrophoresis on 4% non-denaturing polyacrylamide

gel and visualized by autoradiography. Lane 1,7 DNA alone. Lanes 2-6, 0.12, 0.6,

1.2 1.8 and 2.4 μg of BvgA, respectively. Lanes 8-12, same as lanes 2-5 except that

20 mM acetyl phosphate was added to the reactions.

B. Specific competition (lanes 3-5) was carried out using 10, 50 and 100 fold excess

of unlabeled bcfA promoter fragment. Lane1, DNA alone. 2 – 7, BvgA-P- 1.8μg. The

reactions were carried out in the presence of 20 mM acetyl phosphate.

C. Non-specific competition (lanes 3-7) was performed with 10,100,150 and 450 fold

excess of a 525 bp fragment encompassing a region corresponding to the bcfA ORF.

Lane1, DNA alone. 2 – 7, BvgA-P- 1.8μg. The reactions were carried out in the

presence of 20 mM acetyl phosphate.

80

1

0 0BvgA

BvgA-P

2 3 4 5 6 7 8 9 10 11 12

B C

A

1 2 3 4 5 6

Sp. competitor

BvgA-P0

1 2 3 4 5 6 7

Non sp. competitor

BvgA-P0

7

1

0 0BvgA

BvgA-P0 0

BvgABvgA-P

2 3 4 5 6 7 8 9 10 11 12

B C

A

1 2 3 4 5 6

Sp. competitor

BvgA-P0

1 2 3 4 5 6 7

Non sp. competitor

BvgA-P0

7

81

insoluble fractions (see Material and Methods). To facilitate the detection of BcfA,

polyclonal sera were raised against a BcfA-T7-tagged fusion protein purified from E.

coli (see Materials and Methods). Western-blotting with anti-BcfA antibody detected a

polypeptide that migrated at a mobility corresponding to ≈ 100 kDa in the outer

membrane fractions isolated from the wt and the RB25 (ΔbipA) strains but not from

the Bvg- phase locked strain, the RKD110 (ΔbcfA) and the MM101 (ΔbipAΔbcfA)

strains. The absence of a cross-reactive polypeptide band in the Bvg- phase locked

strain, RKD110 (ΔbcfA) and MM101 (ΔbipAΔbcfA) fractions confirms that the

polyclonal serum is specifically recognizing BcfA. As a positive control for a known

outer membrane protein, we probed for BipA. A previously described anti-BipA

antibody (33) raised against the C-terminus of BipA recognized a polypeptide present

in the outer membrane fractions prepared from wt and RKD110 (ΔbcfA) strains but

not from that prepared from the Bvg- phase locked, the RB25 (ΔbipA) and MM101

(ΔbipAΔbcfA) strains (Fig. 7).

BcfA is expressed during the Bordetella infectious cycle. To evaluate whether

BcfA is expressed during infection, we used sera from rats infected for 30 days with

wt, RKD110 (ΔbcfA), RB25 (ΔbipA) and MM101 (ΔbipAΔbcfA) strains as a probe to

detect the purified BcfA protein in western blotting assays. While the serum from the

wt and the RB25 (ΔbipA)-infected rats recognized the purified BcfA protein, sera from

RKD110 (ΔbcfA) and MM101 (ΔbipAΔbcfA) infected rats (Fig. 8) showed no

significant reactivity. Western blot analysis consistently revealed two bands, one

around 100 kDa corresponding to approximately the predicted size of the BcfA

82

FIG.7. BcfA is localized to the outer-membrane. Outer membrane proteins were

prepared by cellular fractionation of the indicated strains and were subjected to

western blot analysis using polyclonal antibodies raised against purified B.

bronchiseptica BcfA and BipA. The molecular weight markers used are indicated.

83

wt

Δbc

fA

Δbi

pA

Δbi

pAΔ

bcfA

Anti-BipA antibody180 115

B

180115

82w

t

Δbc

fA

Δbi

pA

Δbi

pAΔ

bcfA

Rec

. Bcf

A

A

Anti-BcfA antibodyw

t

Δbc

fA

Δbi

pA

Δbi

pAΔ

bcfA

Anti-BipA antibody180 115

B

180115

82w

t

Δbc

fA

Δbi

pA

Δbi

pAΔ

bcfA

Rec

. Bcf

A

A

Anti-BcfA antibody

84

FIG. 8. BcfA is expressed during infection. The purified recombinant BcfA protein

was subjected to SDS-PAGE and western blot analysis using serum collected 30

days post inoculation from rats infected with the indicated strains.

85

wt

Δbc

fA

Δbi

pAΔ

bcfA

Δbi

pA

105230

wt

Δbc

fA

Δbi

pAΔ

bcfA

Δbi

pA

105230

86

protein and another higher molecular weight polypeptide which migrated at greater

than 230 kDa. This higher molecular weight peptide was not recognized by sera from

RKD110 (ΔbcfA) and MM110 (ΔbipAΔbcfA) infected rats, thereby suggesting that this

polypeptide might be a higher molecular weight (possibly a dimer) form of BcfA (Fig.

8). However, this higher molecular weight band was not seen with outer membrane

fractions (Fig. 7), suggesting that it might also be an artifact of the purification

procedure. In addition to sera from rats infected with RB50 for 30 days, sera collected

from rats infected with the wt strain for 12 days also recognized purified BcfA (data

not shown).

Overlapping function of BipA and BcfA in colonization of the lower respiratory

tract. The similarity of BcfA to BipA and the failure to detect a function for bipA in

colonization of the upper respiratory tract led us to hypothesize that BcfA either alone

or in combination with BipA will play a role in respiratory tract colonization. We tested

the effect of these deletions on respiratory tract colonization. Groups of six female

Wistar rats were inoculated intranasally with the wt, RB25 (ΔbipA), RKD110 (ΔbcfA)

and MM110 (ΔbipAΔbcfA) strains. For intranasal inoculation, we utilized the well

studied and frequently utilized low volume-low-inoculum protocol (1, 18). Twelve days

and thirty days post-inoculation, animals were sacrificed and colonization levels were

determined by removing the entire nasal septum and 1 cm of the trachea. Each

tissue was homogenized in 200 μl of PBS, aliquots were diluted and viable colonies

were enumerated by plating on BG agar containing blood.

87

Consistent with previously published results (1, 18), high numbers of bacteria

ranging from 104 to 106 CFUs were recovered from the nasal septums and tracheas

of the wt-inoculated animals at both time points (Fig. 9). Although some animals (one

each for nasal septum at 12 days and 30 days and one for trachea at 12 days) did

not get colonized by the RB25 (ΔbipA) strain, there was no statistically significant

difference in the mean colonization levels as a result of the deletion of bipA. Note that

the failure of the RB25 (ΔbipA) strain to elicit a significant colonization defect is

consistent with previously reported results (33). Similarly, the difference in mean

colonization levels for the wt strain and the RKD110 (ΔbcfA) strain was not

statistically significant at any of the time points tested. In contrast to the individual

deletion of bipA and bcfA, a combined deletion of these two genes resulted in a

drastic defect in the colonization of the trachea at both the early and late time points

(Fig. 9). Notably, for five rats at 12 days and for four rats at 30 days, the number of

CFUs recovered from the trachea was either at or below the lower limit of detection

(Fig. 9). At 12 days, three animals did not get colonized in the nasal septum with the

double deletion strain, but this defect in nasal colonization was not apparent at thirty

days when all of the six animals were colonized and the mean colonization level at

this time point was not significantly different when compared to the wt strain. These

results thus suggest that BipA and BcfA have overlapping roles in colonization of the

trachea.

88

FIG. 9. Colonization of rat respiratory tract by wt and isogenic mutant derivatives

RB25 (ΔbipA), RKD110 (ΔbcfA) and MM101 (ΔbipAΔbcfA). (A) 12 days and (B) 30

days post inoculation. 4-5 weeks old female Wistar rats were inoculated with the

respective strains. The entire nasal septum and 1cm of trachea were harvested at

indicated times, homogenized and aliquots were plated on BG agar plates containing

7.5 % defibrinated blood. The plates also contain 50 μg/ml of streptomycin. The

resultant CFUs were enumerated. Horizontal bars indicate mean of each group. The

dashed lines represent the lower limit of detection.

89

wt

ΔbipAΔbcfA

∆bipA∆bcfA

Nasal Septum Trachea

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+07

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

A

B Nasal Septum Trachea

wt

ΔbipAΔbcfA

∆bipA∆bcfA

Nasal Septum Trachea

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+07

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

1.0E+00

1.0E+01

1.0E+02

1.0E+03

1.0E+04

1.0E+05

1.0E+06

A

B Nasal Septum Trachea

90

DISCUSSION

Efficient colonization of the mammalian respiratory tract by Bordetella occurs by the

production of a collection of virulence factors that include a wide variety of adhesins

and several toxins. Because of the presence of multiple potential adhesin molecules,

it is easy to envision the existence of redundancy in the adherence mechanisms of

Bordetella to cultured epithelial cells in vitro or to the respiratory tract. One or more

adhesin molecules can mediate their action by acting either in a concerted manner or

synergistically. In this study, we report the identification of a Bordetella gene bcfA, a

paralog of the previously identified Bordetella gene bipA (39). We were encouraged

to search for a gene homologous to bipA because a previous report (39) failed to

identify a role for BipA in respiratory tract colonization. With the aid of BLAST

searches, we have detected an ORF in B. bronchiseptica that harbors significant

similarity to BipA. Comparison of BcfA amino acid sequence revealed most significant

sequence identity of the N-terminal region with the corresponding regions of BipA and

intimins of EPEC and EHEC and invasins of Yersinia species. The N-terminal region

in intimins and invasins is required for outer membrane localization and for the export

of the carboxyl termini of these proteins (25, 39). Western Blot analysis of the outer

membrane fractions demonstrated that BcfA, like BipA is localized in the outer

membrane.

The BvgAS signal transduction system is the principal regulator of virulence

gene expression in all the three Bordetella species. We investigated the Bvg-

mediated control of bcfA expression in B. bronchiseptica. Our data show that BvgAS

positively regulates the expression of bcfA. We and others have previously reported

91

that bipA is expressed maximally in the Bvg-intermediate phase (12, 16, 33). Using

real time RT-PCR and β-galctosidase assays to quantitate the expression of bcfA

under different phenotypic phases, we show that in contrast to bipA, bcfA is

expressed maximally both in the Bvg+ and Bvgi phases. In this respect, the

expression profile of bcfA resembles that of another well-studied Bordetella adhesin,

FHA (Fig. 3) and (9, 11). Our kinetic transcriptional activation results indicated that

bipA and bcfA are activated concurrently after the initiation of an inducing signal.

Although results from these two studies seem contradictory at first, it is to be noted

that these represent two different ways by which the transition between the different

activation states of BvgA can occur. In the phase-locked mutant strains, the BvgAS

system is locked in its different activation states (ranging from maximal activation

(Bvg+-phase), sub-maximal (Bvgi-phase) to a minimal (Bvg--phase). In contrast, when

Bordetella are grown overnight in the presence of chemical modulators followed by

subsequent growth in their absence, the activity of BvgA transitions from that of

minimal activation state at early time-points to a maximal state at late time-points.

Thus, it can be hypothesized that genes that have a requirement for a low

concentration of BvgA-P to be activated, will be turned on first. While there are

obvious differences in the arrangement and the affinity of various BvgA binding sites

of fhaB, bipA and bcfA, a common feature of the promoters of all these genes is the

presence of a high affinity BvgA binding site in the form of inverted heptanucleotide

repeats (centered at -88.5 for fha (6), IR1 for bipA (11, 12) and IR2 for bcfA, our

unpublished results). Indeed we and others have demonstrated that under in vitro

conditions using purified BvgA and RNA polymerase, activation of fhaB and bipA

92

occurs at a relatively low concentration of BvgA-P (22, 29). We hypothesize that

similar to fhaB and bipA, bcfA will require a low concentration of BvgA-P to be

activated in vitro and are currently in the process of testing this hypothesis.

Encouraged by our finding of sequence identity of BcfA with BipA and other

bacterial adhesins, we investigated whether BcfA singly or in unison with BipA will

play a role in respiratory tract colonization. We utilized both single as well as double

deletion mutants. Our results utilizing an intranasal rat model of colonization show

that there was no significant demonstrable phenotype in respiratory tract colonization

as a result of individually deleting either bipA or bcfA. Both BipA and BcfA deficient

strains are similar to their isogenic wild type parent in their ability to colonize the nose

and the trachea. However, a combined deficiency of BipA and BcfA practically

resulted in an absence of tracheal colonization, showing that the two proteins indeed

have a role in Bordetella pathogenesis.

These findings raise an important question concerning the function of these

two proteins in Bordetella infections. One of our first hypotheses was that the function

of BipA and BcfA will be different. Based on our inability to detect a defect as a result

of individual deletion of either BipA or BcfA, this scenario seems highly unlikely.

Rather our data indicate that BipA and BcfA have either collective and/or overlapping

functions in Bordetella lifecycle in the host. We believe that this hypothesis is more

likely since a defect in respiratory tract colonization is apparent only after the

disruption of the function of both genes.

What role do these two proteins precisely play in colonization of the respiratory

tract? One simplest explanation is that these proteins are acting as adhesins, i.e.

93

they are directly involved in attachment to the respiratory epithelium. By utilizing

epithelial cell lines and tracheal explants, we are currently testing whether BcfA in

directly involved in attachment. It is also possible that the defect in colonization

observed as a result of gene deletions is not dependent directly on adherence to

respiratory epithelia and that BipA and BcfA might be involved in modulating the

components of the innate and/or adaptive immune systems. In this context, it is

important to note that the major Bordetella adhesin FHA plays an accessory role in

down-regulating the innate immune responses resulting in increased bacterial

persistence (20, 27, 28).

In conclusion, continued studies directed towards deciphering the mechanistic

basis of the role of BipA and BcfA in Bordetella life cycle within its hosts will provide

important insights towards understanding the multipartite nature of Bordetella-host

interactions. Additionally, studies targeted towards dissecting the biochemical basis

of the BvgAS–mediated control of these two genes will allow a clear and detailed

understanding of how a single regulatory locus directs the observed variations in

signal-dependent gene expression patterns.

94

ACKNOWLEDGMENTS

We thank Dan Wozniak for critical reading of the manuscript. Research in the

laboratory of RD is supported by funds from Wake Forest University Health Sciences,

National Research Initiative of the USDA Cooperative State Research, Education and

Extension Service, grant number # 35604-16874." and by NIH grant # R21

AI071054.

95

REFERENCES

1. Akerley, B. J., P. A. Cotter, and J. F. Miller. 1995. Ectopic Expression of the Flagellar Regulon Alters Development of the Bordetella Host Interaction. Cell 80:611-620.

2. Altschul, S. F., W. Gish, W. Miller, E. W. Myers, and D. J. Lipman. 1990. Basic Local Alignment Search Tool3. Journal of Molecular Biology 215:403-410. 3. Bendtsen, J. D., H. Nielsen, G. von Heijne, and S. Brunak. 2004. Improved prediction of signal peptides: SignalP 3.04. Journal of Molecular Biology 340:783-795.

4. Boucher, P. E., A. E. Maris, M. S. Yang, and S. Stibitz. 2003. The response regulator BvgA and RNA polymerase a subunit C-terminal domain bind simultaneously to different faces of the same segment of promoter DNA. Molecular Cell 11:163-173.

5. Boucher, P. E., F. D. Menozzi, and C. Locht. 1994. The Modular Architecture of Bacterial Response Regulators - Insights Into the Activation Mechanism of the Bvga Transactivator of Bordetella-Pertussis . Journal of Molecular Biology 241:363-377.

6. Boucher, P. E., K. Murakami, A. Ishihama, and S. Stibitz. 1997. Nature of DNA binding and RNA polymerase interaction of the Bordetella pertussis BvgA transcriptional activator at the fha promoter. Journal of Bacteriology 179:1755-1763.

7. Boucher, P. E. and S. Stibitz. 1995. Synergistic Binding of Rna-Polymerase and Bvga Phosphate to the Pertussis Toxin Promoter of Bordetella-Pertussis. Journal of Bacteriology 177:6486-6491.

8. Celli, J., W. Y. Deng, and B. B. Finlay. 2000. Enteropathogenic Escherichia coli (EPEC) attachment to epithelial cells: exploiting the host cell cytoskeleton from the outside. Cellular Microbiology 2:1-9.

9. Cotter, P. A. and J. F. Miller. 1997. A mutation in the Bordetella bronchiseptica bvgS gene results in reduced virulence and increased resistance to starvation, and identifies a new class of Bvg-regulated antigens. Molecular Microbiology 24:671-685.

10. de Tejada, G. M., P. A. Cotter, U. Heininger, A. Camilli, B. J. Akerley, J. J. Mekalanos, and J. F. Miller. 1998. Neither the Bvg(-) phase nor the vrg6 locus of Bordetella pertussis is required for respiratory infection in mice. Infection and Immunity 66:2762-2768.

96

11. Deora, R. 2002. Differential regulation of the Bordetella bipA gene: Distinct roles for different BvgA binding sites. Journal of Bacteriology 184:6942-6951.

12. Deora, R., H. J. Bootsma, J. F. Miller, and P. A. Cotter. 2001. Diversity in the Bordetella virulence regulon: transcriptional control of a Bvg-intermediate phase gene. Molecular Microbiology 40:669-683.

13. deTejada, G. M., J. F. Miller, and P. A. Cotter. 1996. Comparative analysis of the virulence control systems of Bordetella pertussis and Bordetella bronchiseptica. Molecular Microbiology 22:895-908.

14. Edwards, R. A., L. H. Keller, and D. M. Schifferli. 1998. Improved allelic exchange vectors and their use to analyze 987P fimbria gene expression 2. Gene 207:149-157.

15. Frankel, G., A. D. Phillips, L. R. Trabulsi, S. Knutton, G. Dougan, and S. Matthews. 2001. Intimin and the host cell - is it bound to end in Tir(s)?. Trends in Microbiology 9:214-218. 16. Fuchslocher, B., L. L. Millar, and P. A. Cotter. 2003. Comparison of bipA Alleles within and across Bordetella species 1. Infection and Immunity 71:3043-3052.

17. Goodnow, R. A. 1980. Biology of Bordetella bronchiseptica. Microbiol.Rev. 44:722-738.

18. Harvill, E. T., P. A. Cotter, and J. F. Miller. 1999. Pregenomic comparative analysis between Bordetella bronchiseptica RB50 and Bordetella pertussis Tohama I in murine models of respiratory tract infection. Infection and Immunity 67:6109-6118.

19. Heininger, U., P. A. Cotter, H. W. Fescemyer, G. M. de Tejada, M. H. Yuk, J. F. Miller, and E. T. Harvill. 2002. Comparative phenotypic analysis of the Bordetella parapertussis isolate chosen for genomic sequencing. Infection and Immunity 70:3777-3784.

20. Inatsuka, C. S., S. M. Julio, and P. A. Cotter. 2005. Bordetella filamentous hemagglutinin plays a critical role in immunomodulation, suggesting a mechanism for host specificity. Proceedings of the National Academy of Sciences of the United States of America 102:18578-18583. 21. Isberg, R. R. and P. Barnes. 2001. Subversion of integrins by enteropathogenic Yersinia. Journal of Cell Science 114:21-28. 22. Jones, A. M., P. E. Boucher, C. L. Williams, S. Stibitz, and P. A. Cotter. 2005. Role of BvgA phosphorylation and DNA binding affinity in control of Bvg-mediated phenotypic phase transition in Bordetella pertussis. Molecular Microbiology 58:700-713.

97

23. Karimova, G. and A. Ullmann. 1997. Characterization of DNA binding sites for the BvgA protein of Bordetella pertussis. Journal of Bacteriology 179:3790-3792. 24. Kinnear, S. M., P. E. Boucher, S. Stibitz, and N. H. Carbonetti. 1999. Analysis of BvgA activation of the pertactin gene promoter in Bordetella pertussis. Journal of Bacteriology 181:5234-5241.

25. Leong, J. M., R. S. Fournier, and R. R. Isberg. 1990. Identification of the Integrin Binding Domain of the Yersinia-Pseudotuberculosis Invasin Protein. Embo Journal 9:1979-1989. 26. Mattoo, S. and J. D. Cherry. 2005. Molecular pathogenesis, epidemiology, and clinical manifestations of respiratory infections due to Bordetella pertussis and other Bordetella subspecies. Clinical Microbiology Reviews 18:326-+. 27. McGuirk, P., C. McCann, and K. H. G. Mills. 2002. Pathogen-specific T regulatory 1 cells induced in the respiratory tract by a bacterial molecule that stimulates interleukin 10 production by dendritic cells: A novel strategy for evasion of protective T helper type 1 responses by Bordetella pertussis. Journal of Experimental Medicine 195:221-231. 28. McGuirk, P. and K. H. G. Mills. 2000. Direct anti-inflammatory effect of a bacterial virulence factor: IL-10-dependent suppression of IL-12 production by filamentous hemagglutinin from Bordetella pertussis. European Journal of Immunology 30:415-422. 29. Mishra, M. and R. Deora. 2005. Mode of action of the Bordetella BvgA protein: Transcriptional activation and repression of the Bordetella bronchiseptica bipA promoter. Journal of Bacteriology 187:6290-6299. 30. Relman, D., E. Tuomanen, S. Falkow, D. T. Golenbock, K. Saukkonen, and S. D. Wright. 1990. Recognition of A Bacterial Adhesin by An Integrin - Macrophage Cr3 (Alpha-M-Beta-2, Cd11B Cd18) Binds Filamentous Hemagglutinin of Bordetella-Pertussis. Cell 61:1375-1382. 31. Simon, R., U. Priefer, and A. Puhler. 1983. A Broad Host Range Mobilization System for Invivo Genetic-Engineering - Transposon Mutagenesis in Gram-Negative Bacteria. Bio-Technology 1:784-791.

32. Stainer, D. a. S. M. 1971. A simple chemically defined medium for the production of phase I Bordetella pertussis. J.Gen.Microbiology 211-220.

33. Stockbauer, K. E., B. Fuchslocher, J. F. Miller, and P. A. Cotter. 2001. Identification and characterization of BipA, a Bordetella Bvg-intermediate phase protein. Molecular Microbiology 39:65-78.

98

34. Uhl, M. A. and J. F. Miller. 1996. Central role of the BvgS receiver as a phosphorylated intermediate in a complex two-component phosphorelay. Journal of Biological Chemistry 271:33176-33180.

35. Uhl, M. A. and J. F. Miller. 1996. Integration of multiple domains in a two-component sensor protein: The Bordetella pertussis BVgAS phosphorelay. Embo Journal 15:1028-1036.

36. Watanabe M and Nagai, M. Whooping cough due to Bordetella parapertussis: an unresolved problem. Expert Rev Anti Infect Ther. 2[3], 447-454. 2004. 37. West, A. H. and A. M. Stock. 2001. Histidine kinases and response regulator proteins in two-component signaling systems. Trends in Biochemical Sciences 26:369-376.

38. Yeh, S. H. 2003. Pertussis: persistent pathogen, imperfect vaccines. Expert.Rev.Vaccines. 2:113-127.

39. Yu, L., E. A. Frey, R. A. Pfuetzner, A. L. Creagh, D. G. Knoechel, C. A. Haynes, B. B. Finlay, and N. C. J. Strynadka. 2000. Crystal structure of enteropathogenic Escherichia coli intimin-receptor complex. Nature 405:1073-1077.

99

Chapter III

Active and Passive Immunization with Bordetella Colonization

Factor A (BcfA) Protects Mice against Respiratory Challenge with

Bordetella bronchiseptica.

Neelima Sukumar, Cheraton F. Love, Matt S. Conover, Nancy D. Kock,

Purnima Dubey and Rajendar Deora

This chapter has been published in Infection and Immunity, 2009 Feb;77(2):885-95

and is reprinted with permission. Neelima Sukumar performed all the experiments

described in this chapter. All animal work were done in collaboration with Cheraton

Love and Matt Conover. Pathology evaluations were carried out by Dr. Nancy Kock.

100

INTRODUCTION

Respiratory pathogens are a major cause of morbidity and mortality in

humans and animals, making the development of efficacious vaccines that protect

against these infections a top priority. Bordetellae are small aerobic, Gram-negative

coccobacilli that colonize the respiratory tracts of humans and animals (31).

Bordetella pertussis infects only humans and causes the acute respiratory disease

whooping cough (6). B. parapertussis strains can be divided into two genetically

distinct types, those which infect humans causing a pertussis-like illness, and those

which cause respiratory infections in sheep (22, 38). B. avium mainly infects

commercially grown turkeys and wild and domesticated birds (43, 45). In contrast,

B. bronchiseptica has a broader host range and is considered a co-contributor to a

number of respiratory syndromes in agriculturally important food producing animals,

pets and nonhuman primates (17). B. bronchiseptica is also a primary etiological

agent and/or a pre disposing factor that results in Porcine Reproductive and

Respiratory Disease complex (PRDC), pneumonia and atrophic rhinitis (AR) in

swine, infectious tracheobronchitis ((ITB) or kennel cough) in dogs and

bronchopneumonia in sheep, guinea pigs, rats, mice, rabbits, cats and non human

primates (5, 31). According to the 2000 National Animal Health Monitoring System

(NAHMS) survey, respiratory disease was the greatest cause of mortality in swine,

accounting for 28.9 percent of nursery deaths and 39.1 percent of deaths in

grower/finisher pigs. The annual economic impact of AR and PRDC in USA alone is

estimated to be about $17 and $40 million, respectively. B. bronchiseptica is also

capable of infecting humans, mostly immunocompromized individuals with AIDS or

101

cystic fibrosis (14, 26, 46, 52), although it was recently isolated from an

immunocompetent individual (39).

Currently available and proposed vaccines against this pathogen include

live, attenuated, heat killed or genetically modified bacteria (2, 30, 32, 48, 49).

Problems associated with these various whole-cell vaccination approaches include:

persistence of the vaccine strain in animals, poor induction of an antibody response

and/or protective immunity and retention of some of the virulence characteristics by

the vaccine strains (2, 30, 32, 48, 49). The genetic mutations that result in the

attenuation of many of the commercially available live attenuated vaccines is

unknown, making it likely that these strains may revert to virulent forms because of

survival pressures in the host, such as co-infections with other pathogenic

organisms. B. bronchiseptica can predispose animals to other infectious agents or

exacerbate the disease symptoms. For example, B. bronchiseptica colonization

leads to increased severity of Canine parainfluenza virus-2 infections and

predisposition of pigs and rabbits to subsequent Pasturella multocida colonization

(8, 12, 15). Infection of porcine tracheal rings with B. bronchiseptica has also been

shown to enhance the adherence of P. multocida (13).

Despite vaccination, animals continue to be carriers resulting in outbreak

among herds. For laboratory animals like rats, mice and rabbits, experimental

infection with B. bronchiseptica results in a chronic and asymptomatic colonization

of the upper respiratory tract. We have been able to isolate B. bronchiseptica from

the rat nasopharynx even 85 days after inoculation (our unpublished results) and

this bacterium has previously been reported to exist in this site for the life of the

102

infected animals (30). Theoretically, persistent colonization of the upper respiratory

tract of the animals vaccinated with live or attenuated strains can create a reservoir

of infectious bacteria from which animal-animal and zoonotic transmission can

occur. Although transmission of a vaccine strain to humans has not been

experimentally proven, a number of such human cases have occurred in individuals

exposed to either infected, sick or recently immunized farm and companion animals

(20).

We propose that an effective acellular vaccination regimen capable of

providing long lasting protective immunity will limit the spread of B. bronchiseptica

not only among animals in a herd but also from animals to humans. For B.

pertussis, there has been a shift to acellular vaccines because of high frequency of

side effects and multiple adverse reactions associated with the whole cell vaccines

(34). Similarly, development of acellular vaccines capable of protecting against B.

bronchiseptica should be given a priority.

BcfA (Bordetella colonization factor) is an outer membrane protein which is

positively regulated by the BvgAS signal transduction system (50). We were

encouraged to examine the role of BcfA in protective immunity because our

previously published research revealed its role in the colonization of the rat trachea.

In addition, sera from rats infected with B. bronchiseptica specifically recognized

BcfA (50). In the current report, we have evaluated the immunogenicity and

protective efficacy of BcfA in an intranasal mouse model of respiratory infection.

Both active and passive immunization with BcfA provided protection against

subsequent intranasal challenge with B. bronchiseptica, which significantly

103

correlated with the production of subclasses of IgG antibodies with high opsonic

activity. Our results also suggested a role for a Th1 type cellular response in BcfA-

mediated protection. Finally, we demonstrated that BcfA is expressed by multiple

clinical isolates of B. bronchiseptica. Data presented in the current study

underscore the potential utility of an acellular vaccine approach for B.

bronchiseptica and highlight the importance of BcfA as a critical protective antigen

against B. bronchiseptica infections.

104

MATERIALS AND METHODS

Bacterial strains, media and growth conditions. The bacterial strains used in

this study are listed in Table 1. B. bronchiseptica wild type (wt) strain RB50 and the

isogenic mutant strain RKD110 (ΔbcfA) has been described previously (50). The

clinical strains of B. bronchiseptica were a kind gift from Dr. T.L. Nicholson and Dr.

K. Register at USDA-ARS. All the strains were maintained on Bordet-Gengou agar

(Becton Dickinson Microbiology Systems) containing 7.5% defibrinated sheep blood

(BG-agar). For RB50 and RKD110 (ΔbcfA), BG-agar was supplemented with

50μg/ml of streptomycin (SM) (Research Products International Corp). For animal

inoculations, single colonies of RB50 were inoculated into Stainer-Scholte broth

(47), cultured overnight, followed by sub-culture to OD600 of ≈ 1 at 37oC.

Over-expression and purification of BcfA. For over-expression of BcfA, E. coli

BL21(DE3)/pLysE cells containing the previously described bcfA over-expression

plasmid, pNS101 were used (50). Bacterial growth conditions, protein induction and

cell extract preparation were carried out as described previously (50). BcfA was

purified using a T7.Tag affinity purification kit (Novagen) as described (50).

Mouse immunizations. Five-six week old female C57/BL6 mice were obtained

from Charles River Laboratories. All experimental procedures were performed in

compliance with institutional regulations and were approved by the animal care and

use committee of WFUHS .

105

TABLE 1. Strains used in this study.

B. bronchiseptica strain isolated from a dog in United States

MBORD685

RB50 derivative having in-frame chromosomal deletion of bcfA (50) RKD110

B. bronchiseptica strain isolated from a sea otter in California (1999) SO3287-99

B. bronchiseptica strain isolated from a pig in Hungary (1993) (3) KM22

B. bronchiseptica strain isolated from a seal in Scottish coast (1999) M584/99/1

B. bronchiseptica strain isolated from a Koala in Australia MBORD698

B. bronchiseptica strain isolated from a cat in United States MBORD631

B. bronchiseptica strain isolated from a horse in United States MBORD628

Wild type B. bronchiseptica strain (9) RB50

Description/ Reference Strain

106

Active immunizations. Groups of four C57/BL6 mice were intraperitoneally

injected with 10 or 30μg of purified BcfA adsorbed to 50μg of alum (Sigma). The

control group of mice received 50μg of alum only. Three weeks later, all the mice

received another dose of the respective immunogen. A week after the second

immunization, mice were sedated with isoflurane (Abbott Laboratories) and

challenged intranasally with 5 x 105 CFU of RB50 in 25μl of sterile PBS. The

number of CFU delivered was confirmed by plating an aliquot of the inoculum on

BG-agar containing SM. One and 6 days post-challenge, mice were sacrificed and

nasal septum, trachea and left lung were harvested. Colonization levels were

determined by homogenizing these tissues in sterile PBS and plating various

dilutions of the homogenates on BG-agar containing SM. Prior to the second

immunization (post-first dose) and at the time of challenge (post-second dose),

blood was drawn from the tail to collect serum for ELISAs.

Passive immunizations. BcfA hyperimmune serum was generated in rats by

Covance as described previously (50). Excised pieces of standard SDS-

polyacrylamide protein gels containing the band corresponding to the BcfA protein

were utilized to immunize rats for the generation of anti-BcfA antibodies. Pre-

immune serum was collected from these rats prior to immunization with BcfA. Since

we were not able to obtain enough pre-immune serum, naïve rat serum (Invitrogen)

was utilized for some of these experiments. To obtain anti-B. bronchiseptica sera,

rats were intranasally infected with RB50 and convalescent phase serum was

collected 30 days post-inoculation.

107

Groups of five C57/BL6 mice were intraperitoneally injected with 200μl of

either anti-BcfA sera, anti-B. bronchiseptica sera, pre-immune serum, or sterile

PBS. Three to four hours later, they were intranasally challenged with 5 x 105 CFU

of RB50 in 25μl of sterile PBS. Mice were sacrificed 3 and 7 days post-challenge

and colonization levels were determined as described above.

Lung histopathology. The right lung from all mice were immersed in 10% neutral

buffered formalin (EMD Chemicals Inc.) just after sacrifice for at least 24 h,

trimmed, embedded in paraffin, processed routinely for histology, cut at 4-6μm,

stained with haematoxylin and eosin (H & E), and examined by light microscopy.

The sections were scored qualitatively for inflammation and injury, degree of overall

cellularity, thickness of alveolar walls, bronchiolar and vascular degeneration,

edema, hemorrhage, and degree and type of inflammatory cellular infiltration by

NK, a board certified veterinary pathologist, in a blind manner.

Neutrophil depletion. RB6-8C5 hybridoma cells were a kind gift from Dr. G.

Huffnagle (University of Michigan) (23, 29). Growth of these and purification of RB6-

8C5 monoclonal antibodies were performed as described previously (23, 29). For in

vivo neutrophil depletion, groups of five C57/BL6 mice were intraperitoneally

injected with 1mg of RB6-8C5 monoclonal antibody. Previous studies have shown

that this treatment regimen is able to deplete neutrophils for 1-2 weeks in mice (28).

A separate group of mice was injected with 1ml of sterile PBS. To determine the

efficiency of neutrophil depletion, blood collected from the tails was analyzed by

108

IDEXX laboratories for complete blood cell counts. The treatment was found to be

98% effective in depleting neutrophils (data not shown). One day post-

administration of RB6-8C5 antibodies, mice were intraperitoneally injected with

200μl of either anti-BcfA sera or naïve rat serum. Three to four h afterwards, they

were intranasally challenged with 5 x 105 CFU of RB50 in 25μl of sterile PBS. Two

days post-challenge, they were sacrificed and colonization levels were determined

as described above. We were unable to extend the infection of neutropenic mice

with B. bronchiseptica beyond two days, because 4 of 5 mice succumbed to

infection by 3 days.

ELISAs. Serum and lung homogenate antibody responses to BcfA were quantified

by coating 96 well flat-bottomed immuno plates (Nalge Nunc International) with

purified BcfA protein. The plates were incubated at 4oC overnight in a humidified

chamber. The wells were then washed 3 times with phosphate buffered saline

containing 0.05% Tween 20 (PBST) (EMD Chemicals Inc.). Blocking for non-

specific interactions was carried out by the addition of 200μl of 5% milk in PBST per

well and plates were incubated at 37oC for 1 h. Lung homogenates or serum

(100μl) from immunized mice or BcfA hyperimmune serum at various dilutions were

added and plates were incubated at 37oC for 2 h. Wells were washed 3 times with

PBST and bound antibodies were detected using HRP-conjugated goat anti-mouse

(Bio-Rad Laboratories) or goat anti-rat IgG (Rockland Inc.) (1:2000) antibodies.

Plates were washed 5 times with PBST and 3,3’, 5,5’-Tetramethyl benzidine

(Sigma) was used as the substrate. Absorbance at OD450 was determined using the

109

Labsystems Multiskan Plus plate reader. Absorbance was plotted against dilution

and the end point titer was determined as the inverse of the highest dilution giving

an OD450 reading 4-5 times above the background. Negative titers were plotted as

zero.

Specific class and isotypes of antibodies present in BcfA hyperimmune

serum and serum from BcfA-immunized mice were determined using rat and mouse

Immunoglobulin Isotyping ELISA Kit (BD Pharmingen), respectively according to

manufacturer’s instructions. Titers of the respective isotypes in the pre-immune

serum from rats or serum from alum-immunized mice were assayed by ELISA and

the OD450 values from these were subtracted from values obtained for BcfA

hyperimmune serum and serum from BcfA-immunized mice, respectively. The

resulting values were then plotted against the dilutions and the end point titer was

determined as the inverse of the highest dilution that gave an OD450 reading 4-5

times above the background. Negative titers were plotted as zero.

Opsonophagocytosis assay. The murine macrophage cell line J774A.1 and

murine monocyte/macrophage cell line RAW 264.7 were cultured in Dulbecco’s

Modified Eagle’s Medium (DMEM) (Invitrogen) supplemented with 10% fetal bovine

serum (FBS) (HyClone) and 4mM L- glutamine (Invitrogen). Approximately 2 x 105

cells were seeded into wells of a 24 well cell culture plate (Corning Incorporated)

and incubated overnight with 5% CO2. For opsonophagocytosis assay, BcfA

hyperimmune serum and naïve rat serum were heat inactivated by incubation at

55oC for 15 min. RB50 or RKD110 (ΔbcfA) was grown to an OD600 of 1 and ≈ 2 x

110

106 cells were incubated with either heat inactivated BcfA hyperimmune serum (at

1% or 10%) or naïve rat serum (10%) or PBS in 100μl at 37oC for 30 min. The

assay was performed by incubating macrophage cells with the above mentioned

mixture of serum or PBS containing bacteria, for 1 h. This was followed by

gentamycin (100μg/ml) (Invitrogen) treatment for 1 h to kill the extra-cellular

bacteria, washed twice with PBS to remove adherent bacteria, lysed with water and

different dilutions were plated to enumerate the number of phagocytosed bacteria.

The fold CFU for each treatment was calculated by dividing the intracellular CFU

obtained from different serum groups by the CFU obtained from the PBS control.

The assay was performed in triplicates and repeated 2-3 times.

T cell cytokine assays. Groups of five C57/BL6 mice were immunized at 0 and 3

weeks with 2 doses of either 30μg of BcfA adsorbed to alum or 50μg of alum only,

as described above. Two and four weeks after administrations of the second dose,

mice were sacrificed and spleens were harvested. Spleens were homogenized in

RPMI (Invitrogen) and red blood cells were lysed using ACK lysing buffer (Lonza).

Splenocytes were counted and approximately 2 x 106 cells were added to each well

of a 96 well plate in RPMI media supplemented with gentamycin. The splenocytes

were re-stimulated with 10μg of purified BcfA protein or medium alone.

Supernatants were harvested after 72 h of incubation and analyzed for IFN-γ, IL-4,

IL-10, IL-12 and TNF-α production using respective ELISA kits (BD OptEIA)

according to manufacturer’s instructions. Using the standards provided with the kit,

111

a standard curve was plotted for each cytokine and the respective cytokine

concentrations in the samples were derived from the standard curve.

SDS-PAGE and immunoblot analyses. Membrane fractions were prepared from

different B. bronchiseptica strains as described previously (50), resolved using

SDS-PAGE, transferred to nitrocellulose membranes (Osmonics Inc.) and probed

with enriched anti-BcfA antibody (1:5000). Anti-BcfA serum was enriched by

repeated incubation with overnight cultures of RKD110 (ΔbcfA) for 3-4 h as

described previously (50). Goat anti-rat IgG (Rockland Inc.) conjugated to HRP

(1:2000) was used as the secondary antibody and proteins were detected using the

Amersham ECL system.

112

RESULTS

Active immunization with BcfA induces protective immunity against B.

bronchiseptica infections. We have previously demonstrated that the outer-

membrane protein BcfA promotes colonization of the rat trachea (50). Our results

also suggested that BcfA-specific antibodies were produced during infection of

animals with B. bronchiseptica (50). Therefore, we hypothesized that the immune

response elicited against BcfA will provide protective immunity against a

subsequent B. bronchiseptica infection. To examine this hypothesis, mice were

intraperitoneally injected with two doses of either 10 or 30μg of purified BcfA

adsorbed to alum. One week after the second immunization, mice were challenged

with RB50 utilizing an intranasal inoculation regimen that seeds and leads to

colonization of the entire respiratory tract (30, 44). In control mice receiving only

alum, high numbers of bacteria were recovered from the nose, the trachea and the

lungs at 1 and 6 days following challenge (Figs. 1A and B, respectively). Three out

of four control mice showed signs of Bordetellosis and one mouse succumbed to

the infection at 6 days post-challenge. In contrast, mice immunized with either

doses of BcfA did not display any signs of Bordetellosis and had remarkably lower

bacterial burdens at both time-points post-challenge (Fig. 1). The reduction of

colonization due to immunization with BcfA was most dramatic in the lungs and the

trachea and at six days post-inoculation. Half of the mice immunized with the lower

dose and all the mice immunized with the higher dose of BcfA had no detectable

bacteria in the lungs (Fig. 1B). Both the immunizing doses resulted in lowering of

the bacterial burdens in tracheas of all of the vaccinated mice (Fig. 1B) and two of

113

FIG. 1. Immunization with BcfA protects mice against B. bronchiseptica challenge.

Mice were immunized intraperitoneally at 0 and 3 weeks with either 10 or 30μg of

BcfA adsorbed to alum or alum only. One week after the second immunization,

mice were intranasally challenged with 5 x 105 CFU of RB50 in a 25μl volume. Mice

were sacrificed at 1 day (A) and 6 days (B) post-challenge and the number of CFU

was determined in the nasal septum, trachea and lungs. Individual symbols

represent a single mouse. The dashed line represents the lower limits of CFU

detection. Black bars represent mean colonization of respective groups. A statistical

analysis was carried out using an unpaired two-tailed Student t test to compare the

CFU obtained from the respective groups of BcfA-immunized mice to that of mice

receiving alum. The asterisks indicate the range of the different P values (one

asterisk, ≤0.05; two asterisks, ≤0.005 and three asterisks, ≤0.0005).

114

0

1

2

3

4

5

6

Log 1

0 CFU

BcfA 10μg

BcfA 30μg

Alum

Trachea Nasal Septum Lungs

0

1

2

3

4

5

6

7

Log 1

0 CFU

Trachea Nasal Septum Lungs

A

B

*

*

* *

*** *

***

**

**

*

1 day

6 days

115

the four mice were cleared of bacteria from this site. Although immunization with

BcfA conferred protection against nasal challenge, compared to the other

respiratory organs, there was only a modest reduction in the bacterial burden. This

observation is consistent with the fact that B. bronchiseptica is extremely difficult to

clear from the nose (18). Our results thus demonstrate that immunization with BcfA

elicits immunity that is protective against B. bronchiseptica infections.

Immunization with BcfA reduces lung pathology in B. bronchiseptica infected

mice. Infection of mice with B. bronchiseptica results in significant lung pathology

and inflammation (21). We next sought to determine whether the reduced morbidity

observed as a result of B. bronchiseptica infection in BcfA-vaccinated mice,

compared to that observed in the control mice, correlated with reduced pulmonary

injury. Lungs from both groups of mice were excised at six days following challenge

with RB50 and examined microscopically. Control mice had pneumonia,

characterized by extensive neutrophilic infiltration of the parenchyma (Fig. 2A) while

BcfA-immunized mice had only modest cellular infiltration (Figs. 2B and 2C). Injury

scores from both groups are given in Table 2. Those of the control mice averaged

10.0, while those immunized with 10 or 30μg of BcfA were markedly lower at 3.0

and 2.7, respectively. This attests that immunization with BcfA considerably

diminishes the pulmonary injury in mice infected with B. bronchiseptica.

Immunization with BcfA induces high antibody responses. Our next objective

was to evaluate whether immunization with BcfA was able to induce the production

116

FIG. 2. Immunization with BcfA reduces lung pathology in mice challenged with

RB50. Representative H & E stained lung sections harvested six days post-

challenge with RB50 from mice immunized with alum only (A) or with 10μg (B) and

30μg (C) of BcfA adsorbed to alum. The sections were examined and evaluated in

a blinded manner. Magnification x 10.

117

C

A

B

118

TABLE 2. BcfA immunization reduces lung pathology. Mice were immunized with

10 or 30μg of BcfA adsorbed to alum or alum only and challenged with RB50 as

described in the legend to Fig. 1. Six days post-challenge, mice were sacrificed and

the right lung were harvested and processed for H & E staining. The sections were

examined by NK blinded to the treatment groups. The total score for each group

and the average score for indicated parameters are represented with ± standard

deviations. Unpaired two-tailed Student t test was used to compare pathology

scores between BcfA immunized and alum immunized mice. The asterisks indicate

the range of the different P values (one asterisk, ≤0.05; two asterisks, ≤0.005).

119

0.4 ± 0.6 0.09 ± 0.3 0 ± 0 Perivascular/Peribronchiolar Lymphocytes

10 ± 5.4 2.7 ± 2 3 ± 1.2 Average Total Score

0.4 ± 0.6 0.2 ± 0.4 0 ± 0 Alveolar Macrophages

1 ± 0.7 0.09 ± 0.3 0 ± 0 Intrabronchial PMNs

2 ± 1.0 0.3 ± 0.5 0.7 ± 0.5 Alveolar/ Interstitial PMNs

0.6 ± 0.9 0.09 ± 0.3 0.1 ± 0.4 Hemorrhage

1.2 ± 0.8 0 ± 0 0.4 ± 0.5 Edema

1 ± 0.7 0.18 ± 0.4 0 ±0 Degeneration

1.8 ± 0.5 1.5 ±0.5 1.7 ± 0.5 Alveolar Walls

1.6 ± 1.1 0.27 ± 0.5 0 ± 0 Consolidation

Alum BcfA 30μg BcfA 10μg Pathology parameters

* **

120

FIG. 3. Anti-BcfA antibody titers in immunized mice.

Anti-BcfA antibody titers in sera (A and B) and lung homogenates (C) collected from

mice immunized with 10μg or 30μg of BcfA adsorbed to alum or alum only.

A. Mouse serum was collected three weeks subsequent to delivery of the first dose

(post-first dose) and immediately before challenge with RB50 (post-second dose).

Total anti-BcfA IgG titers were determined using BcfA as an antigen by ELISA.

Values represented are mean titers from 3-5 mice and bars represent ± standard

deviation. A statistical analysis was carried out using an unpaired two-tailed Student

t test. The asterisks indicate the range of the different P values (one asterisk,

≤0.05; two asterisks, ≤0.005).

B. Mouse serum was collected immediately before challenge with RB50. Titers of

anti-BcfA-specific IgA, IgM and the different IgG isotypes were measured utilizing

specific ELISA kits as described in the Materials and Methods. Each symbol

represents an individual mouse. Black bars represent mean titer.

C. Total anti-BcfA IgG titers was determined from the lung homogenates of

individual mice immunized with either 10μg or 30μg of BcfA adsorbed to alum, alum

only or PBS. Lung homogenates were prepared six days post-challenge with RB50

as described in the Materials and Methods. Values represented are mean titers

from 3-4 mice and bars represent ± standard deviation. A statistical analysis was

carried out using an unpaired two-tailed Student t test. The asterisks indicate the

range of the different P values (one asterisk, ≤0.05; three asterisks, ≤0.0005).

121

0

500

BcfA 10μg BcfA 30μgAlum

700

10700

20700

30700

40700

50700

60700

Post- First Dose Post- Second Dose

**

* *

*

A.

122

0

10

20

30

40

50

60

70

80

90

IgG1 IgG2a IgG2b IgG3 IgA IgM

Tite

r ( X

10,

000)

B.

C.

0

10

20

30

40

50

BcfA 10μg BcfA 30μg Alum

Tota

l IgG

Tite

r

*** *

PBS

123

of specific antibodies in mice. A single dose of either 10 or 30μg of BcfA elicited

high levels of anti-BcfA IgG in the sera (Fig. 3A). After administration of the second

dose of BcfA, IgG levels were much higher, indicating a booster effect. In control

mice, the anti-BcfA IgG titers were at undetectable or background levels (Fig.

3A).While immunization with BcfA resulted in the induction of specific IgM

antibodies in three of the immunized mice, IgA antibodies were not detected (Fig.

3B). We also determined the levels of the different subclasses of anti-BcfA IgG

antibodies. While IgG1, IgG2b and IgG3 anti-BcfA antibodies were detected in the

sera of some of the BcfA-immunized mice, IgG2a was detected in all of the

immunized mice.

We also assayed lung homogenates from immunized mice 6 days post-

challenge with RB50 for antibody production. Whereas BcfA-immunized mice

generated a dose dependent antibody response in the lungs after infection with

RB50, mice which received only alum or PBS had considerably lower levels of anti-

BcfA antibodies (Fig. 3C). Therefore, these results are consistent with the

conclusion that clearance of B. bronchiseptica from the lower respiratory tract

correlates with the presence of high levels of specific antibodies in both the sera

and the lung homogenate.

Passive transfer of anti-BcfA antibodies provides protection against B.

bronchiseptica challenge. A critical role for anti-Bordetella antibodies in both

vaccine and infection-induced immunity against B. bronchiseptica has been

demonstrated (27) (18). We hypothesized that passive immunization with anti-BcfA

124

antibodies would protect mice against B. bronchiseptica challenge. For production

of hyperimmune serum, rats were immunized with excised polyacrylamide gel

fragments containing the recombinant BcfA protein and pre-immune and immune

rat sera were generated as described in the Materials and Methods. We have

previously shown that this polyclonal serum specifically recognizes BcfA (50).

ELISA showed that while in immune rats antibody titers specific for BcfA were very

high (≈1:70,000), in pre-immune rats, the levels of these specific antibodies were

undetectable.

For passive immunization experiments, groups of 5 mice were

intraperitoneally injected with either the BcfA hyperimmune serum, pre-immune rat

serum, the convalescent serum (from rats infected with B. bronchiseptica for 30

days) or sterile PBS, 3-4 hours prior to challenge with 5 x 105 CFU of B.

bronchiseptica. Three and 7 days post-challenge, mice were sacrificed and

bacterial burdens in the different respiratory organs were enumerated as described

in the Materials and Methods. Mice that received the BcfA-specific serum, harbored

lower bacterial burdens at 3 days post-challenge in both the trachea and the lungs,

compared to mice that received only PBS (Fig. 4A). At 7 days post-challenge, the

numbers of bacteria recovered from these sites was further reduced and in some

animals bacteria were not detected (Fig. 4B). Consistent with previous reports (27),

convalescent phase serum from B. bronchiseptica infected rats also resulted in the

lowering of bacterial numbers in the lungs at three days post-challenge and the

infection being cleared from both the trachea and the lungs of some mice at 7 days

post-challenge (Figs. 4A and B). Control mice which received pre-immune rat

125

FIG. 4. Effect of adoptive transfer of BcfA-specific sera on respiratory tract

colonization. Mice were intraperitoneally injected with anti-BcfA hyperimmune

serum, convalescent phase anti-RB50 serum, pre-immune serum or sterile PBS.

Three-four hours later, mice were intranasally challenged with 5 x 105 CFU of RB50

in 25μl volume. Three (A) and seven (B) days post-challenge, mice were sacrificed

and bacterial colonization in the nasal septum, trachea and lungs was determined.

Dashed line represents lower limits of CFU detection. Individual symbols represent

a single mouse. Black bars represent mean colonization of respective group.

Unpaired two-tailed Student t test was used to determine statistical significance.

The groups of mice receiving anti-BcfA hyperimmune serum and convalescent

phase anti-RB50 serum were compared to that of mice receiving sterile PBS. The

asterisks indicate the range of the different P values (one asterisk, ≤0.05; two

asterisks, ≤0.005; three asterisks, ≤0.0005). Differences between the CFU obtained

from pre-immune treated mice and PBS treated mice were not statistically

significant.

126

0

1

2

3

4

5

6

7

0 2

Log 1

0 CFU

BcfA hyperimmune serum

RB50 serum

PBS

Pre-immune serum

Trachea Nasal Septum Lungs

Trachea Nasal Septum Lungs 0

1

2

3

4

5

6

7 Lo

g 10 C

FU

***

*

*

A.

B.

3 days

7 days

*

*

***

127

serum or PBS prior to challenge showed high bacterial burdens in the lower

respiratory tract. In contrast to that observed in the trachea and the lungs, adoptive

transfer of either the BcfA hyperimmune serum or anti-B. bronchiseptica serum had

no significant effect on nasal septum colonization. These results thus suggest that

the protection from B. bronchiseptica infections observed in the lungs and the

trachea was mediated at least partially by antibodies generated in response to

immunization with BcfA. Further, these results also demonstrate that passively

transferred BcfA antibodies are efficient in protecting mice against B. bronchiseptica

challenge similar to those generated during an infection with the wt bacteria.

Anti-BcfA sera are opsonic. To better understand the mechanism of passive

protection, we next determined the levels of antibodies other than IgG and the

various IgG isotypes present in the rat sera utilized for passive protection. While

significant levels of IgM were detected in the anti-BcfA rat sera, no IgA was

detected. Isotyping analysis also revealed that whereas IgG1 was the predominant

isotype, IgG2a was present at lower levels (Fig. 5A).

The presence of IgG1, IgG2a and IgM in the immune serum has been

correlated with high opsonic activity (42). We examined the efficiency of BcfA

hyperimmune rat serum to promote opsonization and phagocytosis of B.

bronchiseptica by J774 murine macrophage cells (Fig. 5B) and RAW 264.7, the

murine monocyte/macrophage cells. These cells are frequently utilized to study

Bordetella pathogenesis and for phagocytic assays (24, 25, 41, 51). Naïve rat

serum or PBS was used as negative control. Opsonization with BcfA-specific sera

128

FIG. 5. Opsonization with anti-BcfA serum enhances the phagocytosis of RB50 by

J774 murine macrophages.

(A) Characterization of anti-BcfA antibodies in BcfA-specific hyperimmune serum.

Titers of different antibody classes and IgG isotypes was measured using specific

ELISA kit as described in the Materials and Methods. The titers obtained were

normalized to the titers of the pre-immune serum. ND = not detectable.

(B) Approximately, 2 x 106 CFU of RB50 or RKD110 (ΔbcfA) was incubated with

either 1 or 10% heat inactivated anti-BcfA hyperimmune serum, 10% naïve rat

serum or sterile PBS at 37o C for 30 min followed by incubation with 2 x 105 J774

cells for 1 h. Extracellular bacteria were killed by treatment with 100μg/ml of

gentamycin for 1 h followed by washing twice with sterile PBS to remove adherent

bacteria. The cells were lysed and the CFU of phagocytosed bacteria were

determined by plating on BG-agar containing SM. Results are expressed as fold

CFU of intracellular bacteria over the PBS treatment and are representative of three

independent experiments performed in triplicates. Bars represent ± standard

deviation. Statistical analysis was carried out using an unpaired two-tailed Student t

test. Asterisk indicates the P value of ≤0.05.

129

010

30

50

70

90

IgG1 IgG2a IgG2b IgG2c IgA IgM

Tite

rs (x

100

)

ND ND ND

A.

0

1

2

3

4

5

6

7

RB50

10% 1% 10% 1% 10% 10%

ΔbcfA

BcfA hyperimmune serum Naïve rat serum

PBS

Fold

intr

acel

lula

r CFU

*

B.

130

increased the efficiency of uptake of RB50 by J774 macrophages when compared

to incubation with naïve rat serum or PBS (Fig. 5B). In contrast, there were no

significant differences in the uptake of the RKD 110 (ΔbcfA) strain on opsonization

with either the BcfA-serum or naïve rat serum and PBS, thereby confirming the

specificity of BcfA-antibody mediated opsonization. Similar results were obtained

with the RAW 264.7 cells (data not shown). These results therefore suggest that

one of the mechanisms for the observed BcfA-mediated protection is increased

opsonization of B. bronchiseptica for phagocytosis.

Neutrophils are critical for anti-BcfA antibody-mediated bacterial clearance.

Neutrophils are a vital component of the immune responses responsible for clearing

B. bronchiseptica infections, since neutropenic mice succumb to B. bronchiseptica

infections within 1-4 days post-inoculation (21, 28). We hypothesized that

neutrophils will be a critical component of the BcfA-induced protective immunity. To

evaluate this possibility, we rendered mice neutropenic by injection of the anti-Gr1

antibody (RB6-8C5) which specifically depletes neutrophils without affecting

resident and circulating macrophages and lymphocytes (11, 28). Groups of 4-5

mice were administered either PBS or the RB6-8C5 antibody, followed by adoptive

transfer of either naïve rat sera or the anti-BcfA sera from rats. Subsequently, these

mice were inoculated with RB50. We examined only the lungs of these mice since

the effect of passive transfer of serum on bacterial clearance is greatest in lower

respiratory tract (Fig. 4 and (27)). Consistent with results presented in Fig. 4A, PBS-

injected mice which received BcfA-specific sera had significantly lower numbers of

131

FIG. 6. Neutrophils are required for anti-BcfA antibody-mediated clearance of B.

bronchiseptica. Mice were intraperitoneally injected with 1mg of RB6-8C5 antibody

or PBS. One day later mice were intraperitoneally injected with 200μl of anti-BcfA

serum or naïve rat serum followed by intranasal challenge with 5 x 105 CFU of

RB50 in a 25μl volume. Mice were sacrificed 2 days post-challenge and bacterial

colonization in the lungs were determined. Dashed line represents lower limits of

CFU detection. Black bars represent mean colonization of respective group.

Unpaired two-tailed Student t test was used to determine statistical significance.

Asterisks indicate the P value of ≤0.05.

132

0

1

2

3

4

5

6

7

8

9Lo

g 10 C

FU

RB

6-8C

5

PBS Treatment

* BcfA hyperimmune serum

Naïve rat serum

133

bacteria in the lungs as compared to those that received the naïve rat serum (Fig.

6). In contrast, neutropenic mice (RB6-8C5-injected) that received BcfA-specific

sera harbored comparatively greater numbers of bacteria in the lungs. These mice

harbored greater than 1000-fold more CFU in the lungs than those observed in

PBS-treated control mice. As expected from previous results (Figs. 4A and B and

(28), neutropenic mice, which received naïve rat serum also harbored high bacterial

burdens in the lungs (Fig. 6). The failure of adoptively transferred BcfA-specific sera

to provide protection against B. bronchiseptica challenge in neutropenic mice

suggests that neutrophils are required for anti-BcfA antibody-mediated clearance of

B. bronchiseptica from the lower respiratory tract.

BcfA induces the production of high levels of IFN-γ in ex-vivo stimulated

splenocytes. In order to characterize the cellular responses stimulated by BcfA

vaccination, we examined BcfA-induced cytokine production by splenocytes ex

vivo. Groups of 5 mice received two successive (at three weeks interval) doses of

either 30μg of BcfA adsorbed to alum or alum alone. Two and 4 weeks following the

administration of the second dose, spleens were excised, processed as described

in the Materials and Methods followed by stimulation with 10μg of BcfA. The culture

supernatant was collected and analyzed for the production of IFN-γ (a prototype

Th1 cytokine), IL-4 (a prototype Th2 cytokine) IL-10, IL-12 and TNF-α. BcfA-

stimulated splenocytes, collected two weeks after immunization with BcfA, secreted

high amounts of IFN-γ and low levels of IL-4 (Fig. 7, left panel). Four weeks post-

immunization with BcfA, there was ≈ 3-fold increase in the production of IFN-γ by

134

splenocytes exposed to BcfA (Fig. 7B). However, IL-4 was not detected. Very little

IL-10, IL-12 or TNF-α were detected in the cultures of splenocytes from BcfA-

immunized mice incubated with BcfA. Re-stimulation of spleen cells from mice

infected with alum alone produced either very low or undetectable levels of all the

cytokines tested (Fig. 7). These data are consistent with the conclusion that

immunization with BcfA elicits a Th1-biased immune response with high levels of

IFN-γ production, which has been previously reported to contribute to efficient

clearance of B. bronchiseptica infection (19, 37).

BcfA expression is prevalent in clinical isolates of B. bronchiseptica.

The sequenced laboratory strains and clinical isolates of the three classical

Bordetella spp. vary in the expression of different virulence factors (4, 16). In order

to strengthen the utility of BcfA-based therapeutics for treatment of B.

bronchiseptica infections, it is critical to determine if BcfA is expressed by

circulating clinical isolates of B. bronchiseptica. We performed SDS-PAGE and

immunoblot analyses on the membrane fractions of a number of B. bronchiseptica

strains isolated from a variety of animal species. Using the affinity-enriched anti-

BcfA serum, we detected the presence of a protein band in all the B. bronchiseptica

strains tested here, which corresponded to the BcfA protein from RB50 (Fig. 8).

135

FIG. 7. BcfA-induced production of IFN-γ and IL-4 in splenocytes. Mice were

immunized at 0 and 3 weeks with either 30μg of BcfA adsorbed to alum or 50μg of

alum only. Two and four weeks post-immunization, splenocytes were harvested and

re-stimulated with 10μg of purified BcfA protein. Supernatants were harvested 3

days post-stimulation and analyzed for IFN-γ and IL-4 production using respective

ELISA kits. Re-stimulations were carried out in triplicates for each individual mouse

and mean of the value from individual mice is represented with ± standard

deviations. Unpaired two-tailed Student t test was performed to determine statistical

significance. Asterisks indicate the P value of ≤0.0005.

136

0

4000

8000

12000

16000

20000

Tota

l pg BcfA

Alum

IFN-γ IL-4 IFN-γ IL-4 2 weeks 4 weeks

***

***

137

FIG. 8. Expression of BcfA among clinical isolates of B. bronchiseptica. Membrane

fractions of respective strains were prepared as described in the Materials and

Methods, subjected to SDS-PAGE followed by immunoblot analyses utilizing anti-

BcfA serum. For MBORD628, the amount of protein loaded corresponds to 0.03X of

the amount loaded for RB50.

138

139

DISCUSSION

There is an urgent need to search for new less-virulent or less-aggressive

vaccines for B. bronchiseptica. Similar to the switch from whole cell vaccines to

cell-free vaccines for B. pertussis, there exists a strong rationale for the use of

acellular vaccines containing defined antigens for B. bronchiseptica. A prerequisite

towards development of these vaccines is a complete understanding of the

antigens which elicit a protective immune response. In this report, we have

examined the protective efficacy of the protein antigen, BcfA in a murine respiratory

challenge model of B. bronchiseptica. Immunization of mice with purified BcfA

resulted in lowering of bacterial burdens as early as 1 day post-challenge and

complete clearance of B. bronchiseptica infection from the lower respiratory tract 7

days subsequent to challenge.

The observed protection due to immunization with BcfA appeared to be

mediated by antibodies specific to BcfA. Active immunization with BcfA resulted in

the production of high titers of IgG and IgM anti-BcfA antibodies in the sera of mice,

suggesting the induction of a strong systemic response. In addition, IgG to BcfA

were also detected in the lungs after challenge with RB50. Although it is possible

that the lung IgG may be derived from systemic IgG reaching the lungs by passive

diffusion, the presence of BcfA-specific antibodies in the lungs is important, since it

shows the generation of a localized antibody response in the actual site of B.

bronchiseptica colonization. Moreover, the BcfA immune serum but not the pre-

immune serum protected mice in the passive immunization and challenge

experiments. While adoptive transfer of immune serum from mice infected or

140

vaccinated with B. bronchiseptica can clear bacteria from the lower respiratory tract

(18), there is limited data available on the efficacy of antibodies against individual

protein components in protection against B. bronchiseptica infections. Strikingly, we

observed that the hyperimmune serum generated against purified BcfA was as

efficient as anti-Bordetella antibodies in mediating this protection.

We found that anti-BcfA antibodies of IgG2a and IgG1 subclasses were

either produced in mice as a result of vaccination with BcfA or were the major

antibody isotypes present in the hyperimmune serum. Although the exact

mechanism by which different IgG subclasses protect against B. bronchiseptica is

not clear, in general these antibodies induce efficient opsonization and

phagocytosis of the bacteria. Indeed, we found that anti-BcfA serum led to greater

internalization of B. bronchiseptica cells by macrophages. These data thus suggest

that one of the mechanisms of BcfA-mediated protection is by induction of large

amounts of specific antibodies with high opsonic activity which leads to enhanced

phagocytosis.

In addition to the role of antibodies, we also investigated the contribution of

cell-mediated immunity in BcfA-induced protection. Immunization with BcfA

preferably induced a Th1 response with high IFN-γ production. Production of this

cytokine has previously been shown to be important for clearance of B.

bronchiseptica (19, 37). The dominant Th1 pattern observed due to vaccination with

BcfA was also consistent with the titers of major BcfA-specific IgG isotypes. In all of

the vaccinated mice, IgG2a was detected, whereas IgG1 was the major isotype in

the hyperimmune BcfA-specific rat serum. Although the correlation between

141

induction of different IgG subclasses and Th1/Th2 activity has not been extensively

characterized, IgG2a in the mouse and IgG1 in the rat are Th1 cytokine driven

antibody subclass (7, 10, 36). Very little is known about specific cellular immune

responses generated as a result of vaccination with B. bronchiseptica proteins.

However, our results demonstrating the induction of a Th1 response by B.

bronchiseptica BcfA is different from the Th2 response induced by component

vaccines against B. pertussis. In contrast to acellular vaccines, infection with B.

bronchiseptica and B. pertussis and immunization with the respective whole cell

vaccines induces a Th1 type response (1, 33, 35, 40). CD4+Th1 cells have been

demonstrated to result in more efficient immunity to B. pertussis compared to CD4+

Th2 cells (1). Induction of a Th1 type response and IFN-γ production as a

consequence of BcfA immunization might explain the superior protective efficacy of

this protein even when administered as a single antigen.

An important prerequisite towards the development of an individual protein-

based therapeutics for treatment of bacterial infections is the ubiquitous

demonstration of its expression among circulating isolates. We have demonstrated

that BcfA is expressed by a wide variety of B. bronchiseptica strains isolated from

multiple animal species. These data further strengthen the utility of BcfA as a

potential vaccine candidate. Previous studies demonstrated that individual

immunization with B. pertussis antigens such as Fha, Pertactin and Pertussis toxin

imparts reasonable protective efficacy. However, vaccines containing multiple

components lead to enhanced protection compared to monocomponent vaccines

(35). Similarly, we believe that despite the remarkable efficacy of BcfA, a

142

multivalent vaccine containing other known Bordetella antigens such as Fha,

Pertactin or Fimbriae will be more efficacious. Future studies will have to address

whether immunization with BcfA alone or with composite vaccines will protect

against large animals like pigs and dogs, in which B. bronchiseptica causes

disease. Finally, the broader potential utility of BcfA as a vaccine candidate is

highlighted by our preliminary results which demonstrate the presence of

polypeptides that cross-react with anti-BcfA serum in the closely related human

pathogens B. pertussis and B. parapertussis (our unpublished results). We are

currently investigating whether these bands correspond to orthologs of BcfA and

whether BcfA can confer protection against these human adapted species. A

positive outcome will be highly significant and will promote the inclusion of BcfA in

the current acellular vaccines for B. pertussis.

143

FIG. 9. Model for BcfA-mediated protective immunity.

Based on the in vivo and in vitro analyses described here, we propose a model for

BcfA-induced immune clearance of B. bronchiseptica. We believe that immunization

with BcfA elicits specific antibodies, which have the ability to enhance opsonization

and phagocytosis of B. bronchiseptica. Findings with neutropenic mice and

opsonophagocytosis assays using J774 macrophage cells reveal a role for

neutrophils and macrophages in the subsequent phagocytosis and clearance of

Bordetella. Also our data demonstrate the induction of IFN-γ and Th1 type responses

in BcfA primed splenocytes. Therefore, we propose that apart from antibodies, BcfA-

immunization also results in the induction of CD 4+ Th1 cells, which is vital for

protective immunity to Bordetella.

144

Challenge with Bordetella

Immunization with BcfA

T

IFNγ

B

Antibodies Phagocytosis

Ciliated Epithelial cells of Respiratory tract

B T

145

ACKNOWLEDGMENTS

We thank Drs. Dan Wozniak and Steven Mizel for critical reading of this manuscript.

We thank Drs. T. Nicholson and K. Register for the different clinical isolates and Dr.

G. Huffnagle for a gift of RB6-8C5 hybridoma cells. Research in the laboratory of

R.D. is supported by funds from National Research Initiative of the USDA

Cooperative State Research, Education and Extension Service (grant 35604-16874)

and the NIH (grant R21AI071054 and 1R01AI075081).

146

REFERENCES

1. Barnard, A., B. P. Mahon, J. Watkins, K. Redhead, and K. H. Mills. 1996. Th1/Th2 cell dichotomy in acquired immunity to Bordetella pertussis: variables in the in vivo priming and in vitro cytokine detection techniques affect the classification of T-cell subsets as Th1, Th2 or Th0. Immunology 87:372-80. 2. Bey, R. F., F. J. Shade, R. A. Goodnow, and R. C. Johnson. 1981. Intranasal vaccination of dogs with liver avirulent Bordetella bronchiseptica: correlation of serum agglutination titer and the formation of secretory IgA with protection against experimentally induced infectious tracheobronchitis. Am. J. Vet. Res. 42:1130-2. 3. Brockmeier, S. L., and K. B. Register. 2007. Expression of the dermonecrotic toxin by Bordetella bronchiseptica is not necessary for predisposing to infection with toxigenic Pasteurella multocida. Vet. Microbiol.125: 284-9 4. Buboltz, A. M., T. L. Nicholson, M. R. Parette, S. E. Hester, J. Parkhill, and E. T. Harvill. 2008. Replacement of adenylate cyclase toxin in a lineage of Bordetella bronchiseptica. J. Bacteriol. 190:5502-11. 5. Buonavoglia, C., and V. Martella. 2007. Canine respiratory viruses. Vet. Res. 38:355-73. 6. Carbonetti, N. H. 2007. Immunomodulation in the pathogenesis of Bordetella pertussis infection and disease. Curr. Opin. Pharmacol. 7:272-8. 7. Cazorla, S. I., P. D. Becker, F. M. Frank, T. Ebensen, M. J. Sartori, R. S. Corral, E. L. Malchiodi, and C. A. Guzman. 2008. Oral vaccination with Salmonella enterica as a cruzipain-DNA delivery system confers protective immunity against Trypanosoma cruzi. Infect. Immun. 76:324-33. 8. Chanter, N., T. Magyar, and J. M. Rutter. 1989. Interactions between Bordetella bronchiseptica and toxigenic Pasteurella multocida in atrophic rhinitis of pigs. Res. Vet. Sci. 47:48-53. 9. Cotter, P. A., and J. F. Miller. 1997. A mutation in the Bordetella bronchiseptica bvgS gene results in reduced virulence and increased resistance to starvation, and identifies a new class of Bvg-regulated antigens. Mol. Microbiol. 24:671-85. 10. Cuturi, M. C., R. Josien, D. Cantarovich, L. Bugeon, I. Anegon, S. Menoret, H. Smit, P. Douillard, and J. P. Soulillou. 1994. Decreased anti-donor major histocompatibility complex class I and increased class II alloantibody response in allograft tolerance in adult rats. Eur. J. Immunol. 24:1627-31.

147

11. Czuprynski, C. J., J. F. Brown, N. Maroushek, R. D. Wagner, and H. Steinberg. 1994. Administration of anti-granulocyte mAb RB6-8C5 impairs the resistance of mice to Listeria monocytogenes infection. J. Immunol. 152:1836-46. 12. Deeb, B. J., R. F. DiGiacomo, B. L. Bernard, and S. M. Silbernagel. 1990. Pasteurella multocida and Bordetella bronchiseptica infections in rabbits. J. Clin. Microbiol. 28:70-5. 13. Dugal, F., M. Belanger, and M. Jacques. 1992. Enhanced adherence of Pasteurella multocida to porcine tracheal rings preinfected with Bordetella bronchiseptica. Can. J. Vet.Res. 56:260-4. 14. Dworkin, M. S., P. S. Sullivan, S. E. Buskin, R. D. Harrington, J. Olliffe, R. D. MacArthur, and C. E. Lopez. 1999. Bordetella bronchiseptica infection in human immunodeficiency virus-infected patients. Clin. Infect. Dis. 28:1095-9. 15. Elias, B., M. Albert, S. Tuboly, and P. Rafai. 1992. Interaction between Bordetella bronchiseptica and toxigenic Pasteurella multocida on the nasal mucosa of SPF piglets. J. Vet. Med. Sci. 54:1105-10. 16. Fennelly, N. K., F. Sisti, S. C. Higgins, P. J. Ross, H. van der Heide, F. R. Mooi, A. Boyd, and K. H. Mills. 2008. Bordetella pertussis expresses a functional type III secretion system that subverts protective innate and adaptive immune responses. Infect. Immun. 76:1257-66. 17. Goodnow, R. A. 1980. Biology of Bordetella bronchiseptica. Microbiol. Rev. 44:722-38. 18. Gopinathan, L., G. S. Kirimanjeswara, D. N. Wolfe, M. L. Kelley, and E. T. Harvill. 2007. Different mechanisms of vaccine-induced and infection-induced immunity to Bordetella bronchiseptica. Microbes Infect. 9:442-8. 19. Gueirard, P., P. Minoprio, and N. Guiso. 1996. Intranasal inoculation of Bordetella bronchiseptica in mice induces long-lasting antibody and T-cell mediated immune responses. Scand. J. Immunol. 43:181-92. 20. Gueirard, P., C. Weber, A. Le Coustumier, and N. Guiso. 1995. Human Bordetella bronchiseptica infection related to contact with infected animals: persistence of bacteria in host. J. Clin. Microbiol. 33:2002-6. 21. Harvill, E. T., P. A. Cotter, and J. F. Miller. 1999. Pregenomic comparative analysis between Bordetella bronchiseptica RB50 and Bordetella pertussis tohama I in murine models of respiratory tract infection. Infect. Immun. 67:6109-18. 22. Heininger, U., K. Stehr, S. Schmitt-Grohe, C. Lorenz, R. Rost, P. D. Christenson, M. Uberall, and J. D. Cherry. 1994. Clinical characteristics of illness

148

caused by Bordetella parapertussis compared with illness caused by Bordetella pertussis. Pediatr. Infect. Dis. J. 13:306-9. 23. Herring, A. C., N. R. Falkowski, G. H. Chen, R. A. McDonald, G. B. Toews, and G. B. Huffnagle. 2005. Transient neutralization of tumor necrosis factor alpha can produce a chronic fungal infection in an immunocompetent host: potential role of immature dendritic cells. Infect. Immun. 73:39-49. 24. Hewlett, E. L., G. M. Donato, and M. C. Gray. 2006. Macrophage cytotoxicity produced by adenylate cyclase toxin from Bordetella pertussis: more than just making cyclic AMP! Mol. Microbiol. 59:447-59. 25. Huang, X. Z., and L. E. Lindler. 2004. The pH 6 antigen is an antiphagocytic factor produced by Yersinia pestis independent of Yersinia outer proteins and capsule antigen. Infect. Immun. 72:7212-9. 26. Huebner, E. S., B. Christman, S. Dummer, Y. W. Tang, and S. Goodman. 2006. Hospital-acquired Bordetella bronchiseptica infection following hematopoietic stem cell transplantation. J. Clin. Microbiol. 44:2581-3. 27. Kirimanjeswara, G. S., P. B. Mann, and E. T. Harvill. 2003. Role of antibodies in immunity to Bordetella infections. Infect. Immun. 71:1719-24. 28. Kirimanjeswara, G. S., P. B. Mann, M. Pilione, M. J. Kennett, and E. T. Harvill. 2005. The complex mechanism of antibody-mediated clearance of Bordetella from the lungs requires TLR4. J. Immunol. 175:7504-11. 29. Lopez, S., A. J. Marco, N. Prats, and C. J. Czuprynski. 2000. Critical role of neutrophils in eliminating Listeria monocytogenes from the central nervous system during experimental murine listeriosis. Infect. Immun. 68:4789-91. 30. Mann, P., E. Goebel, J. Barbarich, M. Pilione, M. Kennett, and E. Harvill. 2007. Use of a genetically defined double mutant strain of Bordetella bronchiseptica lacking adenylate cyclase and type III secretion as a live vaccine. Infect. Immun. 75:3665-72. 31. Mattoo, S., and J. D. Cherry. 2005. Molecular pathogenesis, epidemiology, and clinical manifestations of respiratory infections due to Bordetella pertussis and other Bordetella subspecies. Clin. Microbiol. Rev. 18:326-82. 32. McArthur, J. D., N. P. West, J. N. Cole, H. Jungnitz, C. A. Guzman, J. Chin, P. R. Lehrbach, S. P. Djordjevic, and M. J. Walker. 2003. An aromatic amino acid auxotrophic mutant of Bordetella bronchiseptica is attenuated and immunogenic in a mouse model of infection. FEMS Microbiol. Lett. 221:7-16.

149

33. McGuirk, P., and K. H. Mills. 2002. Pathogen-specific regulatory T cells provoke a shift in the Th1/Th2 paradigm in immunity to infectious diseases. Trends Immunol. 23:450-5. 34. Miller, D. L., E. M. Ross, R. Alderslade, M. H. Bellman, and N. S. Rawson. 1981. Pertussis immunisation and serious acute neurological illness in children. Br. Med. J. (Clin. Res. Ed.) 282:1595-9. 35. Mills, K. H., M. Ryan, E. Ryan, and B. P. Mahon. 1998. A murine model in which protection correlates with pertussis vaccine efficacy in children reveals complementary roles for humoral and cell-mediated immunity in protection against Bordetella pertussis. Infect. Immun. 66:594-602. 36. Philips, J. R., W. Brouwer, M. Edwards, S. Mahler, J. Ruhno, and A. M. Collins. 1999. The effectiveness of different rat IgG subclasses as IgE-blocking antibodies in the rat basophil leukaemia cell model. Immunol. Cell. Biol. 77:121-6. 37. Pilione, M. R., and E. T. Harvill. 2006. The Bordetella bronchiseptica type III secretion system inhibits gamma interferon production that is required for efficient antibody-mediated bacterial clearance. Infect. Immun. 74:1043-9. 38. Porter, J. F., K. Connor, and W. Donachie. 1994. Isolation and characterization of Bordetella parapertussis-like bacteria from ovine lungs. Microbiology 140 ( Pt 2):255-61. 39. Rath, B. A., K. B. Register, J. Wall, D. M. Sokol, and R. B. Van Dyke. 2008. Persistent Bordetella bronchiseptica pneumonia in an immunocompetent infant and genetic comparison of clinical isolates with kennel cough vaccine strains. Clin. Infect. Dis. 46:905-8. 40. Redhead, K., J. Watkins, A. Barnard, and K. H. Mills. 1993. Effective immunization against Bordetella pertussis respiratory infection in mice is dependent on induction of cell-mediated immunity. Infect. Immun. 61:3190-8. 41. Ross, P. J., E. C. Lavelle, K. H. Mills, and A. P. Boyd. 2004. Adenylate cyclase toxin from Bordetella pertussis synergizes with lipopolysaccharide to promote innate interleukin-10 production and enhances the induction of Th2 and regulatory T cells. Infect. Immun. 72:1568-79. 42. Schlageter, A. M., and T. R. Kozel. 1990. Opsonization of Cryptococcus neoformans by a family of isotype-switch variant antibodies specific for the capsular polysaccharide. Infect. Immun. 58:1914-8. 43. Sebaihia, M., A. Preston, D. J. Maskell, H. Kuzmiak, T. D. Connell, N. D. King, P. E. Orndorff, D. M. Miyamoto, N. R. Thomson, D. Harris, A. Goble, A. Lord, L. Murphy, M. A. Quail, S. Rutter, R. Squares, S. Squares, J. Woodward,

150

J. Parkhill, and L. M. Temple. 2006. Comparison of the genome sequence of the poultry pathogen Bordetella avium with those of B. bronchiseptica, B. pertussis, and B. parapertussis reveals extensive diversity in surface structures associated with host interaction. J. Bacteriol. 188:6002-15. 44. Sloan, G. P., C. F. Love, N. Sukumar, M. Mishra, and R. Deora. 2007. The Bordetella Bps Polysaccharide is Critical for Biofilm Development in the Mouse Respiratory Tract. J. Bacteriol.189: 750-760. 45. Spears, P. A., L. M. Temple, D. M. Miyamoto, D. J. Maskell, and P. E. Orndorff. 2003. Unexpected similarities between Bordetella avium and other pathogenic Bordetellae. Infect. Immun. 71:2591-7. 46. Spilker, T., A. A. Liwienski, and J. J. LiPuma. 2008. Identification of Bordetella spp. in respiratory specimens from individuals with cystic fibrosis. Clin. Microbiol. Infect. 14:504-6. 47. Stainer, D. W., and M. J. Scholte. 1970. A simple chemically defined medium for the production of phase I Bordetella pertussis. J.Gen. Microbiol. 63:211-20. 48. Stevenson, A., and M. Roberts. 2002. Use of a rationally attenuated Bordetella bronchiseptica as a live mucosal vaccine and vector for heterologous antigens. Vaccine 20:2325-35. 49. Stevenson, A., and M. Roberts. 2003. Use of Bordetella bronchiseptica and Bordetella pertussis as live vaccines and vectors for heterologous antigens. FEMS Immunol. Med. Microbiol. 37:121-8. 50. Sukumar, N., M. Mishra, G. P. Sloan, T. Ogi, and R. Deora. 2007. Differential Bvg phase-dependent regulation and combinatorial role in pathogenesis of two Bordetella paralogs, BipA and BcfA. J. Bacteriol. 189:3695-704. 51. West, N. P., H. Jungnitz, J. T. Fitter, J. D. McArthur, C. A. Guzman, and M. J. Walker. 2000. Role of phosphoglucomutase of Bordetella bronchiseptica in lipopolysaccharide biosynthesis and virulence. Infect. Immun. 68:4673-80. 52. Woolfrey, B. F., and J. A. Moody. 1991. Human infections associated with Bordetella bronchiseptica. Clin. Microbiol. Rev. 4:243-55.

151

CHAPTER IV: DISCUSSION

The mammalian respiratory tract is equipped with specialized mutltitiered

defense mechanisms to prevent the entry and infection by pathogens. The common

basic defenses at this site include mechanical barriers such as the mucociliatory

escalator consisting of the ciliated epithelium and innate factors such as defensins,

lysozyme and phagocytic cells. If these are compromised, specific adaptive immunity

becomes operational. Despite the presence of sophisticated pulmonary host defense

systems, the respiratory tract acts as the main port of entry for many infectious

agents. These include several pathogenic bacteria that have evolved mechanisms to

overcome the protective host responses present in the respiratory tract and

successfully colonize and exert pathological infections. Understanding the nature of

bacterial – host interactions in the respiratory tract is of interest not only to study the

mechanisms used by pathogens to evade host immune responses, but also for

designing therapeutic and prophylactic strategies. This is highly relevant because

pathogens of the respiratory tract are the foremost cause of mortality and morbidity in

humans and animals. One of the major respiratory pathogens in agriculturally

important, food producing and pet animals is the Gram-negative bacterium Bordetella

bronchiseptica. In the studies described here, we have identified and characterized a

novel virulence factor of Bordetella spp, Bordetella colonization factor A (BcfA). We

show that BcfA is a Bvg-regulated outer membrane protein and that a strain lacking

bcfA, along with its paralog bipA is defective in colonization of the rat trachea.

Additionally, we show that BcfA-specific antibodies are produced during a wild type

152

Bordetella bronchiseptica infection. Subsequently, we describe studies demonstrating

the efficacy of BcfA as a protective antigen against B. bronchiseptica infections. Our

results strongly suggest that both active and passive immunization with BcfA provides

protection against subsequent challenge with B. bronchiseptica in a mouse model of

infection. Also, through ELISAs, we show that BcfA specific antibodies are produced

after immunization, with Ig2a being the predominant isotype. Furthermore,

investigation of the mechanistic basis of BcfA-induced immunity revealed a role for

neutrophils and macrophages in phagocytosis. Finally we show that immunization

with BcfA induces a Th1 type response in splenocytes leading to IFN-γ production,

which has been previously shown to aid clearance of Bordetella infections.

Potential roles of BipA and BcfA in mediating respiratory tract colonization

by Bordetella.

Our results demonstrate that individual deletion of either bipA or bcfA does not

have any significant effect on respiratory tract colonization by Bordetella. Strikingly,

we have found that concomitant deletion of bipA and bcfA results in a defect in

colonization of the rat trachea, thereby suggesting a potential overlapping role for

these two proteins in Bordetella pathogenesis. Although our data demonstrate a

prerequisite for BipA and BcfA in B. bronchiseptica colonization, the exact function of

these two paralogs is unknown. A crucial and initial step in colonization is the

attachment of bacteria to host cells. Therefore, one possibility is that these proteins

function as adhesins and mediate attachment of Bordetella to epithelial cells of

respiratory tract (Fig. 1A). Future studies evaluating the attachment efficiency of the

153

individual deletion strains, the double deletion ΔbipA-ΔbcfA strain and the wt strain to

tissue culture cell lines and cells directly acquired from respiratory tissues of animals

will facilitate addressing this hypothesis.

The respiratory tract is equipped with a multitude of innate and adaptive

immune mechanisms to efficiently eliminate pathogens. The mucociliatory escalator

is an effective process to clear pathogenic organisms from the respiratory tract and

this is complemented by other innate defenses such as neutrophils and macrophages

and adaptive immune responses. Therefore, another significant prerequisite for

successful colonization by Bordetella, subsequent to adherence, is to evade these

host defenses. Numerous pathogens including Bordetella spp. induce anti-

inflammatory immune responses to enhance their persistence (25). Since the ΔbipA-

ΔbcfA strain demonstrated a defective colonization of the rat trachea, another

persuasive hypothesis is that BipA and BcfA are involved in host immune

suppression and thus deletion of these proteins might make the ΔbipA-ΔbcfA strain

susceptible to enhanced immunological clearance. We speculate that BipA and BcfA

may diminish clearance of B. bronchiseptica from the trachea by modulating either

innate or adaptive immune responses including adherence to and phagocytosis by

immune cells such as neutrophils and macrophages, inhibition of the complement

system, altered cytokine and chemokine secretion or antibody production (Fig. 1B).

The role of BipA and BcfA in altering cytokine production as well as adherence and

phagocytosis can be evaluated in vitro by utilizing the epithelial cell lines (L2, Hep2)

or immune cell-derived lines (J774, RAW or isolated macrophages, neutrophils)

infected with either wt or single or double mutant ΔbipA-ΔbcfA strains. To address the

154

roles of these proteins in affecting components of immunity in vivo, the well

characterized mouse model of infection as well as mice defective in specific immune

responses can be utilized (17). Differences between the mutant and the wt strains in

the degree of inflammation in the lungs, recruitment of inflammatory cells, cytokine

production profile and eliciting serum antibody responses can be evaluated using

these models.

In the low volume-low inoculum rat colonization model, the defect in

colonization of the ΔbipA-ΔbcfA strain was surprisingly restricted to only the lower

respiratory tract. The failure to colonize the trachea and not the nasal cavity is a

phenotype common to several Bordetella mutants including strains deficient in FHA,

Cya and Fim (11, 22). While these data may provide evidence for stricter

requirements to colonize trachea as compared to the nasal cavity, it is important to

note that the model of infection that we adopted, utilized low volume (5 μl) to deliver

the inoculum dosage. It has been previously demonstrated that in this model,

Bordetella initially colonizes the nasal cavity from which the bacteria seeds and

colonizes the lower respiratory tract (1). Therefore, the lack of tracheal colonization

in the absence of BipA and BcfA could also be due to failure of Bordetella to traffic

from the nasal cavity to the trachea (Fig. 1C). If this is true, then the wt and the

ΔbipA-ΔbcfA strains will be indistinguishable in their ability to colonize the trachea in

a high volume model of intranasal infection (50 μl) or when delivered through the

intratracheal inoculation route, both of which directly deposit bacteria into trachea.

155

FIG. 1. Potential roles of BipA and BcfA in B. bronchiseptica pathogenesis.

A. BipA and BcfA as adhesins mediating attachment of Bordetella to the respiratory

epithelium.

B. BipA and BcfA as immunomodulators causing the suppression of the host immune

responses that are responsible for the clearance of Bordetella.

C. The lack of colonization of the trachea and not the nasal cavity by the ΔbipA-ΔbcfA

strain in a low volume-low inoculum dosage model of respiratory colonization might

also be due to the inefficient seeding of this strain from the nasal cavity to the

trachea. Therefore, another possibility is that BipA and BcfA function as mediators of

trafficking Bordetella from the upper respiratory tract to the lower respiratory tract.

156

A B

Attachment

Ciliated Epithelial cells of trachea

Immunomodulation Trafficking

T B

T

B

Ciliated Epithelial cells of

Nasal Septum

Ciliated Epithelial cells of trachea

trachea

NeutrophilBordetella BipA B B-cell T

C

BcfA T cell

Macrophage

157

Previous studies have shown the differential expression of bacterial virulence

factors

Evidently, all the proposed functions of BipA and BcfA are not mutually

exclus

BvgAS-mediated regulation of bcfA expression.

signal transduction system that

in a host tissue specific and time dependent manner with the progression of

the infectious cycle. Therefore, we cannot exclude the possibility that BipA and BcfA

are expressed only in the lower respiratory tract in a tissue specific manner at the

time points that we investigated. Also B. bronchiseptica causes prolonged

colonization of the nasal cavity despite clearance from the lower respiratory tract. We

and others have observed that B. bronchiseptica can be isolated from the nasal

cavities of infected animals for at least 270 days (15). Our studies were restricted to

12 and 30 days post inoculation, consequently, the possibility of contribution of BipA

and BcfA to nasal cavity colonization at further stages of infection cannot be ruled

out.

ive; in fact it is highly likely that these proteins may have multiple effector

mechanisms. This hypothesis is in agreement with several other studies investigating

the role of Bordetella factors such as FHA, PT and CyaA. It has been demonstrated

that all these factors have multiple functions in vivo including adherence and host

immunomodulation.

BvgAS is the master regulatory two component

controls the expression of the majority of the virulence factors of Bordetella spp (8).

Given the importance of the BvgAS system, our finding that expression of bcfA is

regulated by this system correlates with its role in Bordetella pathogenicity. The

158

BvgAS – mediated regulation of bcfA is evident by RT-PCR assays as well as by the

presence of multiple sequence elements resembling the consensus BvgA binding site

in the regions upstream of the bcfA ORF. Furthermore, EMSA demonstrates the

direct binding of purified BvgA to the putative bcfA promoter. Although EMSAs clearly

demonstrate the specific binding of BvgA to the bcfA promoter region, they do not

reveal the relative occupancy of individual binding sites. Thus, we can postulate at

least two models to explain the formation of higher order complexes with increasing

concentration of BvgA-P - i) With increasing concentration there is more binding of

BvgA-P dimers to a single binding site or ii) At higher concentrations multiple binding

sites are occupied by BvgA-P (Fig. 2). One of the striking features common to all of

the characterized BvgA-activated promoters is the presence of multiple BvgA binding

sites (2, 13, 26). Hence, this fact along with analysis of the expression profile of bcfA

and identification of several inverted heptads, we speculate that the bcfA promoter

will harbor multiple BvgA binding sites. We are currently conducting DNase I foot

printing assays to further delineate the above mentioned scenarios. Knowledge of

the nature and organization of BvgA binding sites would also enable us to predict the

specific mechanism of Bvg-mediated regulation of phase- specific expression of bcfA

(12). We believe that our studies along with others will provide new insights into how

a single virulence regulatory system controls the expression of a multitude of

virulence factors. Furthermore, these studies will provide the foundation to devise

strategies geared towards disrupting this regulatory circuitry as a preventive measure

against Bordetella pathogenesis. We believe that the complexity in the gene

expression states exhibited by the Bvg regulon may not be

159

FIG. 2. Model illustrating relative occupancy of phosphorylated BvgA (BvgA-P) to the

putative bcfA promoter.

Higher order complex formation seen in Electrophoretic Mobility Shift Assays with

increasing concentrations of BvgA-P may be a resultant of multiple BvgA-P dimers

binding to a single binding site (A) or though occupation of multiple binding sites (B).

160

+1

BvgA-P

BvgA-P

+1

0 0BvgA BvgA-P

BvgA-P

A B

161

unique and other pathogenic bacteria will show similar patterns of gene regulation.

Therefore, results from our studies will serve as a model system for understanding

the dynamics of virulence gene expression in other pathogens.

Differential phase specific expression profile of BipA and BcfA.

Previous studies have shown that Bvg+ phase is sufficient and necessary for

establishing respiratory tract infections and that the Bvgi phase displays decreased

ability to colonize the host (9, 10). Both real-time RT-PCR and β-galactosidase

assays indicate that despite their homology, surprisingly BipA and BcfA have distinct

phase specific expression profiles. Our studies demonstrate that BipA is expressed

only in the Bvgi phase while BcfA is expressed both in the Bvg+ and Bvgi phases.

However, our rat colonization model provides evidence for overlapping functions for

these proteins, since only the ΔbipA-ΔbcfA strain is defective in colonization of the

trachea. These seemingly contrasting observations may be explained on the basis

that our phase specific expression data are restricted to the transcript levels under in

vitro conditions and do not define protein levels generated during an infection. It is

also possible that low level of expression of BipA in the Bvg+ phase is sufficient for

Bordetella colonization. Nevertheless, future studies involving promoter exchange

between BipA and BcfA would address the in vivo significance of the differential

expression pattern of these paralogs.

162

BcfA as a vaccine candidate against B. bronchiseptica.

Despite the availability of a natural mouse infection model as well as other well

characterized animal models, there is only limited focus on the development as well

as efficacy evaluations of vaccines against B. bronchiseptica. An effective vaccination

regimen which provides long lasting protective immunity can easily limit the spread of

this pathogen both among animals in a herd as well as from animals to humans (29).

Whole cell vaccines consisting of inactivated or attenuated strains are currently being

administered to animals, although there is limited information on the safety, efficacy

and stability of these strains. Genetic attenuation of some of the vaccine strains is

still unknown, and there is increasing focus on the use of genetically defined mutants

as live vaccines (28, 29). While these studies reveal a great potential for defined

attenuated vaccine strains in clearing and protecting animals against a lethal B.

bronchiseptica infection, they fail to report the reversion rate of these mutants.

Moreover, in the murine model of infection, in which these strains are tested, bacteria

colonize the upper respiratory tract for the life of the animals (21, 27). Persistent

colonization of the upper respiratory tract of the vaccinated animals by B.

bronchiseptica might create a reservoir of infectious bacteria from which other

animals and even zoonotic infections can occur. Lack of clearance from the

respiratory tract combined with the possibility of reversion of these strains to a more

virulent form emphasizes the need for the development of superior vaccines. In the

context of these observations, our studies demonstrating the remarkable

immunogenicity of a single protein, BcfA, raise the novel concept of employing

acellular vaccines containing defined antigens against B. bronchiseptica. We propose

163

that BcfA in combination with other protective antigens will be a safer and more

efficient vaccine.

Although active immunization with BcfA had striking effects on colonization of

B. bronchiseptica in the lower respiratory tract, it only slightly lowered the bacterial

numbers in the upper respiratory tract. Likewise, passive immunization also had

minimal effects on nasal cavity colonization. As discussed above, this may be due to

the differences in organ specific requirements for colonization and the immune

responses generated in the lower and upper respiratory tract (18). The majority of the

studies in the Bordetella field are restricted to the mechanism of pathogenesis and

bacterial elimination from the lower respiratory tract. Consequentially, the

characteristics of Bordetella or the host immune system that enable this bacterium to

persist in the upper respiratory tract are still unclear. However, it is it important to note

that in our studies we used the intraperitoneal route to administer the antigen, which

is widely accepted to be a poor inducer of mucosal responses. The failure to induce

mucosal responses is evident in the isotyping data which demonstrates a lack of IgA

production both in immunized mice and rats. Previous reports have demonstrated

that IgA is critical for decreasing B. bronchiseptica infection in the upper respiratory

tract (30). Therefore we believe that, by adopting a distinct route of BcfA

administration, namely intranasal immunization, it may be possible to elicit a local

secretory response in the mucosa resulting in either more drastic reduction or

clearance of Bordetella from the upper respiratory tract.

Our studies evaluating the immunogenicity of BcfA are restricted to only its

vaccine potential. Although our data confirm that immunization with BcfA confer

164

relatively high levels of protection against B. bronchiseptica, it does not address the

therapeutic potential of this protein in clearing an established infection. Future

studies can be devised to verify whether administration of BcfA antigen or anti-BcfA

antibodies 7-10 days post inoculation of B. bronchiseptica can accelerate clearance

in a mouse model of infection.

BcfA as a Th1 response inducing adjuvant.

Another significant new finding of our studies is the induction of Th1 response

by BcfA antigen immunization. Ex vivo restimulation of splenocytes harvested from

immunized mice with BcfA resulted in the induction of IFN-γ and Th1 type responses.

In agreement with these results, we also found that immunization with BcfA induces

predominantly IgG1 in rats and IgG2a in mice, both of which are IFN-γ driven

isotypes. Obviously, these findings are surprising considering other studies which

indicate that B. pertussis protein-based vaccines induce a more adjuvant mediated

Th2 response (16, 24). On the contrary, either infection or immunization with B.

pertussis whole cell vaccines induce Th1 responses (24). Consistent with the well

established role of Th1 cells in protective immunity against Bordetella spp, both

natural infection as well as whole cell vaccine induced immunity is considered to be

superior to acellular vaccine induced immunity (23). To further enhance the efficacy

of acellular vaccines, many studies have proposed the identification and use of

effective adjuvants, which can preferentially induce the more appropriate Th1 cell

subtype. Also of note are studies demonstrating that addition of IL-12 to acellular

vaccine augmented its protective efficacy mainly through priming a Th1 biased

165

response (20). At present we do not have an explanation for why BcfA induces a Th1

response. Further investigations into this property of BcfA will involve testing whether

immunization with BcfA alone in the absence of any adjuvant or in the presence of a

different adjuvant other than alum results in Th1 response and IFN-γ production will

substantiate our observations. We are currently testing a BcfA-flagellin fusion protein

and evaluating whether in combination with a different adjuvant, flagellin, BcfA can

still induce a Th1 response. Positive outcome from this experiment will confirm our

results with BcfA immunizations supplemented with alum as an adjuvant. If the ability

to preferentially induce Th1 responses is inherent to the BcfA antigen and is

independent of the adjuvant used, we propose the use of BcfA as a Th1 promoting

adjuvant for antigen vaccines. However, further investigation into this possibility is

essential.

Our assumption that BcfA induces preferentially a Th1 response is based on

differential cytokine secretion. Detection of high amounts of IFN-γ in ex vivo

restimulated splenocytes from BcfA-immunized mice as compared to alum-

immunized mice, is the basis of our theory. Since we used whole splenocyte

populations in these investigations, we cannot rule out the possible contributions of

other immune cells such as natural killer (NK) cells in IFN-γ production. Future

experiments utilizing purified CD4+ T cells will provide further evidence for induction

of Th1 phenotype on BcfA immunization.

166

BcfA as a vaccine candidate against B. pertussis.

There are a plethora of studies describing the immunogenicity of inactivated

whole cells, individual antigens or antigen combinations against B. pertussis

infections (5, 19, 24). A common premise for many of these studies is the fact that

although there are multiple side effects associated with the whole cells vaccines, they

provide superior protection against a B. pertussis infection as compared to the

currently available acellular vaccines (14). The greater efficacy of whole cell vaccines

may be attributed to the presence of additional antigens and factors that contribute to

adjuvanticity such as LPS, which at high levels has been previously shown to

enhance IL-12 production and elicit Th1 responses. In addition several studies have

demonstrated that five, four, three or even two component vaccines have higher

immunogenicity than monocomponent vaccines (24). Consistent with the lower

efficacy of acellular vaccines, in the past two decades, there has been an increase in

the number of whooping cough cases reported in many of countries including USA

where immunization with pertussis vaccine is routinely carried out (6, 7). Gene

profiling studies have revealed that the currently circulating isolates are constantly

mutating antigens such as Prn and PT which are major components of the acellular

vaccines (3, 4). Consequently, antigens included in the acellular vaccines may fail to

impart protection. These observations signify the requirement to identify new vaccine

candidates. Therefore our studies identifying a novel vaccine candidate is highly

encouraging. Currently we are examining the expression of BcfA in B. pertussis.

Although the open reading frame of bcfA gene is disrupted by an insertion sequence,

our current results from SDS-PAGE and western blot analyses utilizing the BcfA

167

polyclonal serum on whole cell lysates of B. pertussis reveal cross reactive bands

(Fig. 3). Our preliminary data also reveal the expression of BcfA in recent clinical

isolates of B. pertussis. This observation is significant as it shows the conservation of

BcfA among currently circulating host adapted strains. The relevance of this result is

further justified in the light of pathogenomic studies demonstrating that the current

circulating isolates of B. pertussis are losing genetic material continuously and the

lost genetic material appears to be unimportant for virulence and pathogenesis (3).

Further studies are required to investigate whether these bands correspond to

orthologs of BcfA and whether it can confer protection against this human adapted

species. We speculate that inclusion of new antigens such as BcfA may improve the

immunogenic activity of the currently available vaccines.

In conclusion, our studies describe a novel antigen of Bordetella spp, BcfA, its

role in B. bronchiseptica colonization and its utility as immunogen to elicit protective

immune responses. We believe that our studies will be a stimulus for identifying new

vaccine targets and for developing novel therapeutic interventions against Bordetella

infections.

168

FIG. 3. Expression of BcfA by B. pertussis strains. Whole cell lysates of B. pertussis

strains Bp536 (laboratory strain), M3984, S49560 (clinical strains) and B.

bronchiseptica strains RB50 and ΔbcfA were separated by SDS-PAGE and analyzed

by Western blotting with anti-BcfA hyperimmune serum.

169

RB

50

Δbc

fA

S495

60

M39

84

Bp5

36

115 kDa

Clinical Strains B. bronchiseptica

B. pertussis

170

REFERENCES 1. Akerley, B. J., P. A. Cotter, and J. F. Miller. 1995. Ectopic expression of

the flagellar regulon alters development of the Bordetella-host interaction. Cell 80:611-20.

2. Boucher, P. E., and S. Stibitz. 1995. Synergistic binding of RNA

polymerase and BvgA phosphate to the pertussis toxin promoter of Bordetella pertussis. J Bacteriol 177:6486-91.

3. Bouchez, V., V. Caro, E. Levillain, G. Guigon, and N. Guiso. 2008.

Genomic content of Bordetella pertussis clinical isolates circulating in areas of intensive children vaccination. PLoS ONE 3:e2437.

4. Buboltz, A. M., T. L. Nicholson, M. R. Parette, S. E. Hester, J. Parkhill,

and E. T. Harvill. 2008. Replacement of adenylate cyclase toxin in a lineage of Bordetella bronchiseptica. J Bacteriol 190:5502-11.

5. Cherry, J. D. 2007. Immunity to pertussis. Clin Infect Dis 44:1278-9. 6. Cherry, J. D., S. J. Chang, D. Klein, M. Lee, S. Barenkamp, D.

Bernstein, R. Edelman, M. D. Decker, D. P. Greenberg, W. Keitel, J. Treanor, and J. I. Ward. 2004. Prevalence of antibody to Bordetella pertussis antigens in serum specimens obtained from 1793 adolescents and adults. Clin Infect Dis 39:1715-8.

7. Cherry, J. D., and P. Olin. 1999. The science and fiction of pertussis

vaccines. Pediatrics 104:1381-3. 8. Cotter, P. A., and A. M. Jones. 2003. Phosphorelay control of virulence

gene expression in Bordetella. Trends Microbiol 11:367-73. 9. Cotter, P. A., and J. F. Miller. 1994. BvgAS-mediated signal transduction:

analysis of phase-locked regulatory mutants of Bordetella bronchiseptica in a rabbit model. Infect Immun 62:3381-90.

10. Cotter, P. A., and J. F. Miller. 1997. A mutation in the Bordetella

bronchiseptica bvgS gene results in reduced virulence and increased resistance to starvation, and identifies a new class of Bvg-regulated antigens. Mol Microbiol 24:671-85.

11. Cotter, P. A., M. H. Yuk, S. Mattoo, B. J. Akerley, J. Boschwitz, D. A.

Relman, and J. F. Miller. 1998. Filamentous hemagglutinin of Bordetella bronchiseptica is required for efficient establishment of tracheal colonization. Infect Immun 66:5921-9.

171

12. Deora, R. 2002. Differential regulation of the Bordetella bipA gene: distinct roles for different BvgA binding sites. J Bacteriol 184:6942-51.

13. Deora, R. 2004. Multiple mechanisms of bipA gene regulation by the

Bordetella BvgAS phosphorelay system. Trends Microbiol 12:63-5. 14. Gopinathan, L., G. S. Kirimanjeswara, D. N. Wolfe, M. L. Kelley, and E.

T. Harvill. 2007. Different mechanisms of vaccine-induced and infection-induced immunity to Bordetella bronchiseptica. Microbes Infect 9:442-8.

15. Harvill, E. T., P. A. Cotter, and J. F. Miller. 1999. Pregenomic

comparative analysis between bordetella bronchiseptica RB50 and Bordetella pertussis tohama I in murine models of respiratory tract infection. Infect Immun 67:6109-18.

16. Higgins, S. C., A. G. Jarnicki, E. C. Lavelle, and K. H. Mills. 2006.

TLR4 mediates vaccine-induced protective cellular immunity to Bordetella pertussis: role of IL-17-producing T cells. J Immunol 177:7980-9.

17. Inatsuka, C. S., S. M. Julio, and P. A. Cotter. 2005. Bordetella

filamentous hemagglutinin plays a critical role in immunomodulation, suggesting a mechanism for host specificity. Proc Natl Acad Sci U S A 102:18578-83.

18. Kirimanjeswara, G. S., P. B. Mann, and E. T. Harvill. 2003. Role of

antibodies in immunity to Bordetella infections. Infect Immun 71:1719-24. 19. Leef, M., K. L. Elkins, J. Barbic, and R. D. Shahin. 2000. Protective

immunity to Bordetella pertussis requires both B cells and CD4(+) T cells for key functions other than specific antibody production. J Exp Med 191:1841-52.

20. Mahon, B. P., M. S. Ryan, F. Griffin, and K. H. Mills. 1996. Interleukin-

12 is produced by macrophages in response to live or killed Bordetella pertussis and enhances the efficacy of an acellular pertussis vaccine by promoting induction of Th1 cells. Infect Immun 64:5295-301.

21. Mann, P., E. Goebel, J. Barbarich, M. Pilione, M. Kennett, and E.

Harvill. 2007. Use of a genetically defined double mutant strain of Bordetella bronchiseptica lacking adenylate cyclase and type III secretion as a live vaccine. Infect Immun 75:3665-72.

22. Mattoo, S., J. F. Miller, and P. A. Cotter. 2000. Role of Bordetella

bronchiseptica fimbriae in tracheal colonization and development of a humoral immune response. Infect Immun 68:2024-33.

172

23. Mills, K. H. 2001. Immunity to Bordetella pertussis. Microbes Infect 3:655-77.

24. Mills, K. H., M. Ryan, E. Ryan, and B. P. Mahon. 1998. A murine model

in which protection correlates with pertussis vaccine efficacy in children reveals complementary roles for humoral and cell-mediated immunity in protection against Bordetella pertussis. Infect Immun 66:594-602.

25. Pilione, M. R., and E. T. Harvill. 2006. The Bordetella bronchiseptica

type III secretion system inhibits gamma interferon production that is required for efficient antibody-mediated bacterial clearance. Infect Immun 74:1043-9.

26. Roy, C. R., J. F. Miller, and S. Falkow. 1990. Autogenous regulation of

the Bordetella pertussis bvgABC operon. Proc Natl Acad Sci U S A 87:3763-7.

27. Shimizu, T., and H. Ishikawa. 1982. Some characteristics of a urease-

negative, temperature- sensitive strain of Bordetella bronchiseptica as a live, attenuated vaccine. Infect Immun 36:198-201.

28. Stevenson, A., and M. Roberts. 2002. Use of a rationally attenuated

Bordetella bronchiseptica as a live mucosal vaccine and vector for heterologous antigens. Vaccine 20:2325-35.

29. Stevenson, A., and M. Roberts. 2003. Use of Bordetella bronchiseptica

and Bordetella pertussis as live vaccines and vectors for heterologous antigens. FEMS Immunol Med Microbiol 37:121-8.

30. Wolfe, D. N., G. S. Kirimanjeswara, E. M. Goebel, and E. T. Harvill.

2007. Comparative role of immunoglobulin A in protective immunity against the Bordetellae. Infect Immun 75:4416-22.

173

CURRICULUM VITAE

NAME: Neelima Sukumar ADDRESS:

Residence: 390 Sailway Road Winston-Salem, NC 27127 Phone: (336) 602-2541; Cell: (336) 406-6633 Business: Department of Microbiology & Immunology Wake Forest University School of Medicine Medical Center Boulevard Winston-Salem, North Carolina 27157 Phone: (336) 716-1211 FAX: (336) 716-9928 E-mail: [email protected]

EDUCATION

2004-Present Wake Forest University School of Medicine Winston-Salem, NC Advisor: Dr. Rajendar Deora Ph.D.: Projected May 2009 Major: Microbiology & Immunology

Dissertation Title: Bvg-dependent Regulation and Protective Efficacy of BcfA GPA: 4.0

1999-2004 Kerala Agricultural University Vellayani, Kerala, India

Bachelor of Science in Agriculture GPA: 9.0/10 (Ranked 1st in the University) FELLOWSHIPS 2004 Dean’s Fellowship, Wake Forest University Graduate School of Arts & Sciences

1999-2003 Indian Council of Agricultural Research Merit Scholarship

174

1999-2003 Kerala Agricultural University Merit

Scholarship 1999-2003 Cochin Refineries School Higher Education Scholarship

PATENTS:

Deora R, Mishra M, Sukumar N. Role of a Bordetella outer-membrane protein BcfA in protective immunity and as a vaccine candidate. Applied, 2008

PUBLICATIONS

JOURNAL ARTICLES:

Sukumar N, Love CF, Conover MS, Kock ND, Dubey P, Deora R. Active and Passive Immunization with Bordetella Colonization Factor A (BcfA) Protects Mice against respiratory challenge with Bordetella bronchiseptica. Infect Immun. 2009 Feb;77(2):885-95. Featured in the Spot light section Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Differential Bvg phase-dependent regulation and combinatorial role in pathogenesis of two Bordetella paralogs, BipA and BcfA. J Bacteriol. 2007 May;189(10):3695-704. Sloan GP, Love CF, Sukumar N, Mishra M, Deora R. The Bordetella Bps polysaccharide is critical for biofilm development in the mouse respiratory tract. J Bacteriol. 2007 Nov;189(22):8270-6.

ABSTRACTS: Sukumar N, Love CF, Conover MS, Kock ND, Dubey P, Deora R. BcfA as a vaccine candidate against Bordetella infections in humans and animals. Charlotte Biotechnology Conference, Barnhardt Student Activities Center, UNC Charlotte, October 28, 2008. Sukumar N, Love CF, Conover MS, Kock ND, Dubey P, Deora R. Immunization with BcfA induces protective immunity against B. bronchiseptica infections. Graduate Student Festival, National Institutes of Health, Bethesda, Maryland, September 11-12, 2008. Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Differential Bvg phase-dependent regulation and combinatorial role in pathogenesis

175

of two Bordetella paralogs, BipA and BcfA. American Society for Microbiology 107th General meeting, Metro Convention Center, Toronto, Canada, May 21-25, 2007. Sloan GP, Love CF, Sukumar N, Mishra M, Deora R. Biofilm Formation of Bordetella in the Murine Respiratory Tract. American Society for Microbiology 107th General meeting, Metro Convention Center, Toronto, Canada, May 21-25, 2007. Sloan GP, Love CF, Sukumar N, Mishra M, Deora R. Role of the Bordetella BPS Exopolysaccaride in Promoting Biofilm Development and Persistent Colonization of the Mammalian Respiratory Tract. American Society for Microbiology Conference on Biofilm, Quebec City, Canada, March 25-29 2007. Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Differential Bvg phase-dependent Regulation and Combinatorial Role in Pathogenesis of two Bordetella Paralogs, BipA and BcfA. Mid-Atlantic Microbial Pathogenesis Meeting, Wintergreen Resort, Wintergreen, Virginia, February 11-13, 2007. Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Synergistic Role of Two Homologous Proteins BipA and BcfA in Respiratory Tract Colonization by Bordetella. 2005 NC ASM Annual meeting The Jane S. McKimmon Conference Center 1101 Gorman Street, NC State University, Raleigh, NC,October 7, 2005

ORAL PRESENTATIONS:

Sukumar N, Love CF, Conover MS, Kock ND, Dubey P, Deora R. Immunization with Bordetella Colonization Factor A (BcfA) Protects Mice against respiratory challenge with Bordetella bronchiseptica. 2008 NC ASM Annual meeting, Frank Family Science Center Guilford College, Greensboro, NC, October 10, 2008 Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Differential Bvg phase-dependent Regulation and Combinatorial Role in Pathogenesis of two Bordetella Paralogs, BipA and BcfA. Mid-Atlantic Microbial Pathogenesis Meeting, Wintergreen Resort, Wintergreen, Virginia, February 11-13, 2007. Sukumar N, Mishra M, Sloan GP, Ogi T, Deora R. Synergistic Role of Two Homologous Proteins BipA and BcfA in Respiratory Tract Colonization by Bordetella. 2005 NC ASM Annual meeting The Jane S. McKimmon Conference Center 1101 Gorman Street, NC State University, Raleigh, NC, October 7, 2005

176

Sukumar N, Deora R. Role of two paralogous proteins BipA and BcfA in Bordetella pathogenesis. Fall Symposium, Department of Microbiology and Immunology, Wake Forest University Baptist Medical Center, Winston-Salem, NC, August, 2006. Sukumar N, Deora R. Role of Bvg regulated genes bcfA and bipA in respiratory tract colonization. Fall Symposium, Department of Microbiology and Immunology, Commons Conference Room, Wake Forest University Baptist Medical Center, Winston-Salem, NC, August 23, 2005.

AWARDS: Recipient of Second Prize, Annual Graduate Student Poster Competition, Charlotte Biotechnology Conference, Barnhardt Student Activities Center, UNC Charlotte, October 28, 2008. Selected to Participate in the Annual Graduate Student Poster Competition, Charlotte Biotechnology Conference, Barnhardt Student Activities Center, UNC Charlotte, October 28, 2008. Selected to Participate in the National Graduate Student Research Festival (Poster Presentation) – NIH, Bethesda, Maryland, September 11-12, 2008. Mid-Atlantic Microbial Pathogenesis Meeting Student Travel Award: Mid-Atlantic Microbial Pathogenesis Meeting, Wintergreen Resort, Wintergreen, Virginia, February 11-13, 2007.

Gold Medal from Kerala Agricultural University, Vellayani, India: For the student with highest GPA, 2003.

Bronze Medal, Duke of Edinburg: For Extracurricular activities, Kochi Refineries School, India, 1999.

RESEARCH SKILLS: Molecular Techniques: DNase-1 Footprinting, EMSA, Primer Extension, RACE-PCR, RNA and DNA isolation, Protein over-expression and purification, Gel electrophoresis, Western Blotting, RT-PCR, Reporter Assays, Cloning and Mutagenesis.

Immunological Assays: ELISA, Cytokine Assays, Splenocyte Re-stimulation Assays, Opsonophagocytosis Assays.

177

Animal Models: Worked with mice and rats, transgenic mice, Immunizations, Harvesting different organs. Adhoc review of manuscripts.

PROFESSIONAL APPOINTMENTS & ACTIVITIES

Graduate School Activities Member of Wake Forest University Graduate Student Association, 2004- Present

PROFESSIONAL MEMBERSHIPS

Member of American Society for Microbiology. Member of American Association for the Advancement of Science.

178

179