type iv pilins and identification of a novel d ......diversity of pseudomonas aeruginosa type iv...

257
DIVERSITY OF PSEUDOMONAS AERUGINOSA TYPE IV PILINS AND IDENTIFICATION OF A NOVEL D-ARABINOFURANOSE POST-TRANSLATIONAL MODIFICATION BY JULIANNE V. KUS A thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy Graduate Department of Dentistry University of Toronto © 2008 Julianne V. Kus

Upload: others

Post on 01-Feb-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

  • DIVERSITY OF PSEUDOMONAS AERUGINOSA TYPE IV PILINS

    AND IDENTIFICATION OF A NOVEL D-ARABINOFURANOSE POST-TRANSLATIONAL

    MODIFICATION

    BY

    JULIANNE V. KUS

    A thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy

    Graduate Department of Dentistry University of Toronto

    © 2008 Julianne V. Kus

  • ii

    DIVERSITY OF PSEUDOMONAS AERUGINOSA TYPE IV PILINS AND IDENTIFICATION OF A NOVEL D-ARABINOFURANOSE

    POST-TRANSLATIONAL MODIFICATION

    Julianne V. Kus

    Doctor of Philosophy, Faculty of Dentistry, University of Toronto

    2008

    ABSTRACT The opportunistic bacterial pathogen Pseudomonas aeruginosa uses type IV pili

    (T4P) for adherence to, and rapid colonization of, surfaces via twitching motility. T4P are

    formed from thousands of pilin (PilA) subunits. Two groups of P. aeruginosa pilins were

    described previously (I and II), distinguished by protein length and sequence. PilAI was

    glycosylated with an O-antigen subunit through the action of PilO/TfpO, encoded

    downstream of pilAI. To determine if additional pilin variants existed, analysis of the pilin

    locus of >300 P. aeruginosa strains from a variety of environments was conducted. Three

    additional pilin alleles were discovered, each of which was invariantly associated with a

    unique, previously unidentified, downstream gene(s): pilAIII+tfpY, pilAIV+tfpW+tfpX,

    pilAV+tfpZ. This survey also revealed that strains with group I T4P were more commonly

    associated with respiratory infections than strains with other pilins, suggesting that

    glycosylated T4P may confer a colonization advantage in this environment. The newly

    identified group IV pilin, represented by strain Pa5196, migrated aberrantly through SDS-PA

    gels, suggesting it was also glycosylated, a hypothesis confirmed by periodic acid-Schiff

    staining and mass spectrometry (MS) analyses. Disruption of Pa5196 O-antigen

    biosynthesis did not prevent the production of glycosylated pilins, demonstrating that these

    pilins were modified in a novel manner, unlike group I pilins. Using MS, nuclear magnetic

    resonance spectroscopy and site-directed mutagenesis, the Pa5196 pilins were shown to be

    uniquely modified with homo-oligosaccharides of mycobacterial-like α-1,5-D-

    arabinofuranose at multiple locations. Residues Thr64 and Thr66, located on the αβ-loop

    region of the protein, appear to be the preferred, but not exclusive sites of modification, each

    being modified with up to four D-Araf sugars. This region of the pilin is partially surface-

    exposed in the pilus, therefore modification of these sites may influence the surface

    chemistry of the fibre. Residues Ser81, Ser82, Ser85 and Ser89, located in the β-strand

  • iii

    region, were also modified, mainly with mono- and disaccharides. Bioinformatic analyses

    and mutagenesis of TfpW suggest that this novel protein is an arabinosyltransferase

    necessary for PilAIV modification. This research has increased our understanding of the

    complexity of this virulence factor, and may aid in development of new therapeutics for P.

    aeruginosa and mycobacterial infections.

  • iv

    ACKOWLEDGEMENTS

    I am grateful for the mentorship, support, patience, enthusiasm and friendship of my

    supervisors Dr. Lori Burrows and Dr. Dennis Cvitkovitch. I truly appreciate the

    opportunities, and challenges, that you have provided me with over the years. I

    would also like to thank Dr. John Kelly and his colleagues at the National Research

    Council for their help with my research; their contributions were invaluable. I am

    also thankful for the critiques, suggestions and guidance of my supervisory

    committee members, Dr. Richard Ellen and Dr. Cliff Lingwood.

    I also want to thank the many people that I’ve had the privilege to work with at

    Toronto General Hospital, SickKids and the Faculty of Dentistry. I’ve been lucky to

    have developed some truly great friendships during this degree. There were many

    times when it was these friendships that brought me into the lab, even when my

    experiments weren’t working. Thank you especially: Claude, Melissa, Craig, Poney,

    Marc, Selva, Kirsten, Kowthar, and Carlos.

    Beyond the lab, I need to thank my friends for their support and for trying to

    understand why I always seemed to be busy.

    Finally, I need to thank my family, especially my parents who have always

    encouraged me, and believed in me, even when I didn’t believe in myself. Thank

    you for all the opportunities that your love and support have made available to me.

  • v

    TABLE OF CONTENTS

    ABSTRACT ii ACKNOWLEDGEMENTS iv TABLE OF CONTENTS v LIST OF TABLES xi LIST OF FIGURES xii ABBREVIATIONS xiv PUBLICATIONS FROM THIS DEGREE xvi AWARDS HELD DURING THIS DEGREE xvii QUOTATION xviii CHAPTER 1: LITERATURE REVIEW _1 I. PSEUDOMONAS AERUGINOSA _2

    A. Pseudomonas aeruginosa _2

    B. The clinical significance of P.aeruginosa _3

    1. Nosocomial infections _3

    2. Cystic fibrosis and other pulmonary infections _3

    C. The biofilm mode of growth _6

    D. Attaching to a surface _7

    II. TYPE IV PILI 10

    A. History and classification 10

    B. Distribution 10

    C. Pilin to pilus 11

    1. N-terminal α-helix 14

    2. αβ-loop 14

    3. β-strands 15

    4. Disulphide-bonded-loop 15

    5. The pilus fibre 16

    6. Pilus assembly complex 20

    D. Functions 21

    1. Twitching motility 21

    2. Adherence 24

    E. Diversity of PilA 25

  • vi

    1. Group II 26

    2. Group I 27

    3. Group III 27

    4. Horizontal gene transfer 28

    5. Additional mechanisms of pilin diversity 29

    III. T4P GLYCOSYLATION SYSTEMS 30

    A. Introduction to Gram-negative glycoproteins 30

    B. Neisserial pilin glycosylation 32

    1. Site of modification 32

    2. Mechanism 33

    3. Role of glycosylation in T4P function 34

    C. P. aeruginosa pilin glycosylation: strain 1244 36

    1. Site of modification 36

    2. Mechanism 37

    3. Role of glycosylation in T4P function 39

    IV. RESEARCH AIMS 40

    A. Hypothesis 40

    B. The main objectives of this dissertation 40

    V. REFERENCES 42

    CHAPTER 2: SIGNIFICANT DIFFERENCES IN TYPE IV PILIN ALLELE DISTRIBUTION AMONG PSEUDOMONAS AERUGINOSA ISOLATES FROM CYSTIC FIBROSIS (CF) VERSUS NON-CF PATIENTS 58 I. ABSTRACT 59

    II. INTRODUCTION 59

    III. MATERIALS AND METHODS 63

    A. Bacterial strains and media 63

    B. Twitching assays 64

    C. Polymerase chain reaction (PCR), cloning and DNA sequencing 65

    D. Identification of novel accessory genes 66

    E. Restriction fragment length polymorphism typing of tfpOb strains 66

    F. Pulsed-field gel electrophoresis 67

    G. Analysis of pilin proteins by SDS-PAGE 68

    H. Phylogenetic analysis 69

  • vii

    I. Statistical analysis _69

    J. Accession numbers _69

    IV. RESULTS _70

    A. Twitching motility in CF versus non-CF strains _70

    B. PCR amplification analysis of pilA and flanking sequences _70

    C. Group II pilins _71

    D. Group III pilins _73

    E. Group IV pilins _75

    F. Group V pilins _76

    G. Group I pilin genes and their predominance in CF isolates _77

    H. Analysis of novel pilin proteins _79

    I. Phylogenetic analysis of PilA proteins from P. aeruginosa

    confirms that there are five distinct groups _80

    V. DISCUSSION _81

    VI. ACKNOWLEDGEMENTS _89

    VII. REFERENCES _97

    CHAPTER 3: THE PREDOMINANCE OF P. AERUGINOSA STRAINS WITH GROUP I T4P IN CF SPUTUM IS NOT CF-SPECIFIC 104 I. INTRODUCTION 105

    II. MATERIALS AND METHODS 107

    A. Bacterial Strains 107

    B. Expression of pilin variants in a common background strain 107

    III. RESULTS AND CONCLUSIONS 108

    A. Pulsed-field gel electrophoresis 108

    B. PCR amplification analysis of pilA and accessory genes 108

    C. Distribution of T4P groups 109

    D. Pilin variants in a common background strain 110

    IV. REFERENCES 119

    CHAPTER 4: GLYCOSYLATION OF PSEUDOMONAS AERUGINOSA STRAIN PA5196 TYPE IV PILINS WITH MYCOBACTERIAL-LIKE α-1,5 LINKED D-ARAf OLIGOSACCHARIDES 122 I. ABSTRACT 123

  • viii

    II. INTRODUCTION 124

    III. MATERIALS AND METHODS 126

    A. Bacterial strains and serotyping 126

    B. Generation of a wbpM mutant of Pa5196 127

    C. Pilin isolation 128

    D. Fluorescent glycoprotein detection 129

    E. Mass spectrometry analysis of the intact pilin 129

    F. Enzymatic digestion and nanoLC-MS/MS analysis of the pilin 129

    G. Fractionation of the enzymatic digests and analysis by

    nano-electrospray ionization – mass spectrometry (nESI-MS) 130

    H. Glycopeptide purification for NMR analysis 131

    I. NMR analysis of the pilin glycan 131

    J. Determination of the enantiomeric form of the glycan by GC-MS 132

    K. M. smegmatis whole cell lysate preparation and Western blot

    analysis 133

    IV. RESULTS 134

    A. Pa5196 pilins are modified with a novel glycan that is not the

    O-antigen 134

    B. Intact protein mass determination by ESI-MS 136

    C. Analysis of enzymatic digests by nanoLC-MS/MS 136

    D. Characterization of the T55-79 tryptic glycopeptide by

    nESI-feCID-MS/MS 140

    E. Elucidation of the O-linked pilin glycopeptide structure 141

    F. Determination of the chiral orientation of the arabinose glycan 143

    G. Antigenic identity of the Pa5196 pilin glycan with M. tuberculosis

    lipoarabinomannan 144

    V. DISCUSSION 146

    VI. ACKNOWLEDGMENTS 151

    VII. REFERENCES 152

  • ix

    CHAPTER 5: MODIFICATION OF THE TYPE IV PILINS IN PSEUDOMONAS AERUGINOSA PA5196 AT MULTIPLE SITES WITH 1,5-D-ARABINO FURANOSE AND IDENTIFICATION OF THE PUTATIVE D-ARABINOSYL-TRANSFERASE, TFPW 159 I. ABSTRACT 160

    II. INTRODUCTION 161

    III. EXPERIMENTAL PROCEDURES 164

    A. Bacterial strains and growth conditions 164

    B. Sequence analysis of PilA and TfpW 165

    C. Generation of pilA and tfpW mutants 168

    D. Complementation of mutants 169

    E. Site-directed mutagenesis 170

    F. Twitching motility assays 170

    G. Surface pilin isolation, SDS-PAGE and Western blots 171

    H. Mass spectrometry analysis of intact pilin 173

    I. Enzymatic digestion and LC-MS analysis of the pilA protein

    from Pa5196NP+pilA 174

    J. Electron-transfer dissociation mass spectrometry 175

    IV. RESULTS 175

    A. Antibodies to arabinosylated pilins recognize mycobacterial cell

    wall material 175

    B. Modification sites of Pa5196 PilA 179

    C. Electron-transfer dissociation analysis of Pa5196 pilins 186

    D. Glycosylation of other pilins in the Pa5196 background 188

    E. Characterization of the putative arabinosyltransferase, TfpW 189

    V. DISCUSSION 193

    A. TfpW: a novel glycosyltransferase? 198

    VI. ACKNOWLEDGEMENTS 203

    VII. REFERENCES 204

    CHAPTER 6: SUMMARY AND CONCLUSIONS 213 I. SUMMARY OF DISSERTATION AND DISCUSSION 214

    A. T4P variation and distribution 214

    B. A novel P. aeruginosa T4P modification 218

  • x

    C. Sites of PilA5196 glycosylation 220

    D. TfpW: A novel arabinosyltransferase 224

    E. Proposed mechanisms for PilA5196 glycosylation 226

    II. FUTURE WORK 229

    A. Identifcation of arabinose biosynthetic genes 229

    B. Determining the biological role of arabinose-modified pilins 230

    III. SIGNIFICANCE AND CONCLUSION 231

    IV. REFERENCES 233

  • xi

    LIST OF TABLES

    CHAPTER 2: SIGNIFICANT DIFFERENCES IN TYPE IV PILIN ALLELE DISTRIBUTION AMONG PSEUDOMONAS AERUGINOSA ISOLATES FROM CYSTIC FIBROSIS (CF) VERSUS NON-CF PATIENTS Table 2-1. Distribution of pilin alleles among isolates from various sources _72

    Supplemental Table 2-1.

    Pseudomonas aeruginosa strains analyzed for this study _90

    CHAPTER 3: THE PREDOMINANCE OF P. AERUGINOSA STRAINS WITH GROUP I T4P IN CF SPUTUM IS NOT CF-SPECIFIC Table 3-1. Primers to amplify pilA variants 115

    Table 3-2. Primer combinations and templates used to amplify pilin genes 115

    Table 3-3. Non-CF respiratory disease isolates of P. aeruginosa 116

    Table 3-4. Percentage of each T4P group by source of isolation 117

    CHAPTER 4: GLYCOSYLATION OF PSEUDOMONAS AERUGINOSA STRAIN PA5196 TYPE IV PILINS WITH MYCOBACTERIAL-LIKE α-1,5 LINKED D-ARAf OLIGOSACCHARIDES Table 4-1. 1H and 13C chemical shift assignments for the two αAraf

    monomers that give rise to the dominant peaks in the NMR spectra

    of the pilin-associated glycan from P. aeruginosa Pa5196 144

    CHAPTER 5: MODIFICATION OF THE TYPE IV PILINS IN PSEUDOMONAS AERUGINOSA PA5196 AT MULTIPLE SITES WITH 1,5-D-ARABINO- FURANOSE AND IDENTIFICATION OF THE PUTATIVE D-ARABINOSYL-TRANSFERASE, TFPW Table 5-1. List of bacteria strains and plasmids used in this study 166 Table 5-2. PCR primers used in this study 172 Table 5-3. Selected BLASTP results for TfpW5196 191

  • xii

    LIST OF FIGURES

    CHAPTER 1: LITERATURE REVIEW Figure 1-1. Crystal structures of the pilin structural subunits of

    N. gonorrhoeae and P. aeruginosa _13

    Figure 1-2. The pole of a P. aeruginosa cell displaying T4P and flagella _18

    Figure 1-3. Simplified cartoon of P. aeruginosa pilus assembly complex _20

    Figure 1-4. Twitching motility assay _22

    Figure 1-5. Genetic organization of the pilin locus of P. aeruginosa _26

    Figure 1-6. Structure and molecular mechanism of pilin glycosylation _35

    CHAPTER 2: SIGNIFICANT DIFFERENCES IN TYPE IV PILIN ALLELE DISTRIBUTION AMONG PSEUDOMONAS AERUGINOSA ISOLATES FROM CYSTIC FIBROSIS (CF) VERSUS NON-CF PATIENTS Figure 2-1. Five distinct pilin alleles in P. aeruginosa _71

    Figure 2-2. Amino acid sequence alignment of the pilins’ C-terminal

    DSL region _74

    Figure 2-3. PFGE typing of group V strains _77

    Figure 2-4. SDS-PAGE of sheared pilins _80

    Figure 2-5. Kyte-Doolittle hydropathy plots of TfpO and TfpW proteins _82

    Figure 2-6. Phylogenetic relationships between PilA and related proteins _83

    CHAPTER 3: THE PREDOMINANCE OF P. AERUGINOSA STRAINS WITH GROUP I T4P IN CF SPUTUM IS NOT CF-SPECIFIC Figure 3-1. Twitching assays of transgenic strains 112

    CHAPTER 4: GLYCOSYLATION OF PSEUDOMONAS AERUGINOSA STRAIN PA5196 TYPE IV PILINS WITH MYCOBACTERIAL-LIKE α-1,5 LINKED D-ARAf OLIGOSACCHARIDES Figure 4-1. Pa5196 pilins continue to be glycosylated in a wbpM mutant 135

    Figure 4-2. Analysis of the intact Pa5196 pilin by ESI-MS 137

    Figure 4-3. Identification of pilin glycopeptides by MS/MS 138

    Figure 4-4. feCID-MS/MS analysis of the T55-79 glycopeptide 141

  • xiii

    Figure 4-5. 1D 1H spectrum and 1H-13C HSQC spectrum of the O-linked

    pilin glycan from P. aeruginosa 143

    Figure 4-6. Recognition of P. aeruginosa Pa5196 pilins by antibodies

    to lipoarabinomannan 146

    CHAPTER 5: MODIFICATION OF THE TYPE IV PILINS IN PSEUDOMONAS AERUGINOSA PA5196 AT MULTIPLE SITES WITH 1,5-D-ARABINO- FURANOSE AND IDENTIFICATION OF THE PUTATIVE D-ARABINOSYL-TRANSFERASE, TFPW Figure 5-1. CLUSTALW amino acid alignment of mature pilins 177

    Figure 5-2. Pilins of strains Pa5196 and PA7 are modified with D-Araf 178

    Figure 5-3. Twitching motility of PilA5196 site-directed mutants 180

    Figure 5-4. SDS-PAGE and Westerns of SDM surface T4P 182

    Figure 5-5. Mass spectrometry of sheared intact pilins 183

    Figure 5-6. LC-MS analysis of the tryptic digest of pilin protein

    isolated from P. aeruginosa 5196NP + PilA5196 185

    Figure 5-7. ETD-MS analysis of glycosylated peptides of 5196NP+pilA5196 187

    Figure 5-8. Topology model of TfpW 190

    Figure 5-9. TfpW influences twitching motility 191

    Figure 5-10. TfpW participates in the glycosylation of PilA5196 194

    Figure 5-11. Model of truncated Pa5196 PilA illustrating sites, and potential

    sites, of D-Araf modification 196

  • xiv

    ABBREVIATIONS

    AA Amino acid

    AG Arabinogalactan

    Ala Alanine

    AP Alkaline phosphatase

    Asn Asparagine

    BAL Bronchoalveolar lavage

    CAD Collision assisted dissociation

    Carb Carbenicillin

    CF Cystic fibrosis

    CFTR Cystic fibrosis transmembrane regulator

    CID Collision induced dissociation

    COSY Correlation Spectroscopy

    Cys Cysteine

    Da Daltons

    D-Araf D-arabinofuranose

    DATDH 2,4-diacetimido-2,4,6-trideoxyhexopyranose

    DPB Diffuse panbronchiolitis

    DSL Disulphide-bonded loop

    ELISA Enzyme-linked immunosorbent assay

    ESI-MS Electrospray ionizing mass spectrometry

    ETEC Enterotoxigenic Escherichia coli

    ETD-MS Electron transfer dissociation mass spectrometry

    feCID Front-end collision induced dissociation

    GC-MS Gas-chromatography mass spectrometry

    Gg4 Gangliotetraosylceramide

    Gm Gentamicin

    asialo-GM1 Gangliotetraosylceramide

    HMBC Heteronuclear multiple bond correlation

    HPLC High proformance liquid cromatography

    HSQC Heteronuclear single quantum coherence

    IATS International antigenic typing scheme

    ICU Intensive care unit

  • xv

    LAM Lipoarabinomannan

    LB Luria-Bertani

    LC Liquid chromatography LPS Lipopolysaccharide

    LTQ Linear trap quadrupole

    mAb Monoclonal antibody

    MDCK Madin-Darby canine kidney min minutes

    MS Mass spectrometry

    NIAID National institute of allergy and infectious diseases

    NIH National institute of health

    NMR Nuclear magnetic resonance

    NOESY Nuclear overhauser effect spectroscopy

    NP No pili

    ODS Octadecyl silane

    O-GlcNAc O-linked N-acetylglucosamine

    ORF Open reading frame

    PAS Periodic acid-Schiff

    PCR Polymerase chain reaction

    PFGE Pulsed-field gel electrophoresis

    Pro Proline

    Q-TOF2 Quadrupole time-of-flight mass spectrometer

    rpHPLC-MS Reversed-phase high performance liquid chromatography mass

    spectroscopy

    s seconds

    SDM Site-directed mutagenesis

    SDS-PA Sodium dodecyl sulfate polyacrylamide

    SDS-PAGE Sodium dodecyl sulfate polyacrylamide gel electrophoresis

    Ser Serine

    T4P Type IV pili

    Thr Threonine

    TOCSY Total correlation spectroscopy

    WT Wild-type

  • xvi

    PUBLICATIONS FROM THIS DEGREE PUBLICATIONS FROM THE DISSERTATION CHAPTERS 1. Kus, J.V., Tullis, E., Cvitkovitch, D.G. and Burrows, L.L. (2004). Significant

    differences in type IV pilin distribution among Pseudomonas aeruginosa isolates

    from cystic fibrosis (CF) versus non-CF patients. Microbiology 150: 1315-26.

    2. Voisin, S.*, Kus, J.V.*, Houliston, S., St-Michael, F., Watson, D., Cvitkovitch, D.G., Kelly, J., Brisson, J-R., and Burrows, L.L. (2007). Glycosylation of Pseudomonas aeruginosa strain Pa5196 type IV pilins with mycobacterial-like α-

    1,5 linked D-Araf oligosaccharides. Journal of Bacteriology 189:151-9. *authors

    contributed equally to this work. 3. Kus, J.V., Kelly, J., Tessier, L., Harvey, H., Cvitkovitch, D.G., and Burrows, L.L.

    (2008). Modification of Pseudomonas aeruginosa Pa5196 type IV pilins at

    multiple sites with 1,5-D-arabinosfuranose and identification of the putative

    arabinosyltransferase, TfpW. Submitted to Journal of Biological Chemistry,

    February 2008.

    PUBLICATIONS NOT INCLUDED IN THIS DISSERTATION 1. Moraes, T.J.,Plumb, J., Martin, R., Vachon, E., Cherepanov, V., Koh, A., Loeve,

    C., Jongstra-Bilen, J., Kus, J.V., Burrows, L.L., Grinstein, S. and Downey, G.P. (2006) Abnormalities in the Pulmonary Innate Immune System in Cystic Fibrosis. American Journal of Respiratory Cell and Molecular Biology. 34: 364-74.

    2. Eman, A., Yu, A.R., Park, H-J, Mahfoud, R., Kus, J.V., Burrows, L.L. and

    Lingwood, C.A. (2006). Laboratory and clinical Pseudomonas aeruginosa strains do not bind glycosphingolipids in vitro or during type IV pili-mediated initial host

    cell attachment. Microbiology. 152: 2789-99.

  • xvii

    AWARDS HELD DURING THIS DEGREE

    2007 Constantine Maniatopoulos Graduate Scholarship, awarded for presenting the best student poster, Basic Sciences, Faculty of Dentistry’s Annual Research Day

    2006 Student Presentation Award/Travel Award, Federation of American

    Societies for Experimental Biology (FASEB) Conference on Microbial Polysaccharides of Medical, Agricultural and Industrial Importance

    2002-2007 Canadian Institute for Health Research (CIHR) Fellowship, “Cell

    Signaling in Mucosal Inflammation and Pain” STP-53877 2006-2007 Ontario Graduate Scholarship 2002-2007 Harron Scholarship, Faculty of Dentistry, University of Toronto 2005-2006 University of Toronto Open Fellowship 2004-2006 Canadian Cystic Fibrosis Foundation 2003-2004 Seymour Bresalier Ontario Graduate Scholarship in Science and

    Technology 2002 McMurrich Award (Best Basic Science Poster Presentation), 1st Annual

    University of Toronto, Department of Surgery, Gallie-Bateman Research Day

  • xviii

    "If we knew what it was we were doing,

    it would not be called research, would it?”

    Albert Einstein

  • 1

    CHAPTER 1:

    LITERATURE REVIEW

  • 2

    I. PSEUDOMONAS AERUGINOSA

    A. Pseudomonas aeruginosa

    Pseudomonas aeruginosa is a Gram-negative rod-shaped bacterium that is

    ubiquitous in nature, inhabiting diverse environments including soil, water, plants,

    animals and humans (Green et al., 1974; Lyczak et al., 2000; Ramos, 2004). They

    belong to the Gamma Proteobacteria, a group which includes several medically

    significant pathogenic bacteria including Escherichia coli, Vibrio cholerae,

    Salmonella spp., Haemophilus influenzae and Legionella pneumophila.

    P. aeruginosa is a metabolically versatile microorganism with meagre

    nutritional requirements and are capable of tolerating diverse physical conditions

    including a wide range of temperatures (>42°C), high salt concentrations, weak

    antiseptics, and many antimicrobial agents. These characteristics have contributed

    to the success of P. aeruginosa as an opportunistic pathogen of humans, and a

    dominant nosocomial pathogen capable of colonizing virtually any site within

    immunocompromised individuals (Bodey et al., 1983; Canadian Antimicrobial

    Resistance Alliance, 2007; Lyczak et al., 2000). P. aeruginosa’s ability to infect a

    variety of organisms, including the model hosts Mus musculus, Drosophila

    melanogaster, Caenorhabditis elegans and Arabidopsis thaliana has allowed for

    genetic studies of this bacterium’s mechanisms of pathogenicity (Mahajan-Miklos et

    al., 2000).

  • 3

    B. The clinical significance of P. aeruginosa

    1. Nosocomial infections. Nosocomial P. aeruginosa infections are most

    frequently associated with ventilator-associated acute pneumonia in intensive care

    unit (ICU) patients, however blood, wound, eye and urinary tract isolates are also

    common (Bodey et al., 1983; Canadian Antimicrobial Resistance Alliance, 2007;

    Lyczak et al., 2000). P. aeruginosa possesses a wide range of virulence factors that

    contribute to its success as an opportunistic pathogen, including a variety of

    extracellular enzymes and toxins which damage host cells and tissues, a number of

    different adhesins, and several mechanisms for antimicrobial resistance (Lyczak et

    al., 2000). It is currently the most commonly isolated pathogen in Ontario ICUs, and

    ranks in the top six pathogens isolated from ICUs throughout the rest of Canada,

    contributing significantly to patient morbidity and length of hospital stay (Canadian

    Antimicrobial Resistance Alliance, 2007; Carmeli et al., 1999). Clinical isolates are

    becoming increasingly resistant to multiple classes of antibiotics, further limiting

    treatment options for these infections (Pfaller et al., 2006). One study of patients in

    a North American hospital setting reported that patients infected with P. aeruginosa

    strains in which resistance had emerged had a 3-fold higher mortality rate than the

    overall hospital mortality rate, highlighting the severity of these infections (Carmeli et

    al., 1999).

    2. Cystic fibrosis and other pulmonary infections. P. aeruginosa

    pulmonary infections are the leading cause of morbidity and mortality amongst cystic

    fibrosis (CF) patients; it is reported that respiratory failure is the primary cause of

  • 4

    death in 90% of persons with CF (Cystic Fibrosis Foundation, 2006; Davies, 2002;

    Lyczak et al., 2002). Colonization of the respiratory tract by P. aeruginosa typically

    occurs early in the life of a CF patient and is never resolved (Burns et al., 2001;

    Murray et al., 2007). These individuals are not immunocompromised, but have a

    mutation in the gene encoding the CF transmembrane conductance regulator

    (CFTR) chloride ion channel which results in decreased salt transport across

    mucosal membranes, causing mucosal secretions to be thick and dehydrated rather

    than fluid. It is thought that mucociliary clearance of the lungs is inhibited by this

    viscous mucus and as a consequence, microorganisms become trapped and persist

    in the airways (Boucher, 2007a; Boucher, 2007b; Lyczak et al., 2002). This chronic

    infection eventually causes irreversible lung damage to which the majority of patients

    succumb.

    P. aeruginosa is not normally the pioneering microorganism of the CF lung,

    typically colonizing the lung after Staphylococcus aureus or Haemophilus influenzae

    infections; however, over time it becomes the dominant and prevailing pathogen in

    these patients (Tummler & Kiewitz, 1999). One theory of P. aeruginosa’s

    prevalence in the CF lung is that its type IV pili (T4P) specifically bind to the receptor

    gangliotetraosylceramide (asialo-GM1 or Gg4), which has been reported to be

    present in higher amounts on epithelial cells of CF patients homozygous for the

    common ∆F508 mutation of cftr (Imundo et al., 1995; Krivan et al., 1988; Saiman &

    Prince, 1993). However, this issue is currently debated, as there is evidence that P.

    aeruginosa does not require Gg4 to bind epithelial cells (see Section I, D) (Emam et

    al., 2006; Schroeder et al., 2001).

  • 5

    Evidence that P. aeruginosa is more of a generalist, capable of exploiting a

    stressed lung environment, as opposed to being a specific pathogen of CF patients,

    is the fact that it is a common pathogen of other chronic lung diseases such as

    diffuse panbronchiolitis (DPB) and non-CF related bronchiectasis (King et al., 2007;

    Martinez-Garcia et al., 2007; Poletti et al., 2006). Again, in these diseases P.

    aeruginosa colonization typically follows infection by other pathogens. As opposed

    to other pathogens, DPB and bronchiectasis patients colonized with P. aeruginosa

    had the worst clinical features, poorest lung function and the most extensive

    disease, often leading to death (King et al., 2007; Poletti et al., 2006).

    Although it remains unclear precisely why P. aeruginosa is the dominant

    pathogen within the lungs of patients with CF and other pulmonary disorders,

    several factors likely play significant roles in its success as a pathogen. These

    factors include its widespread presence in the environment (including food, medical

    instruments, sinks, drains, mops and dental water lines), the production of numerous

    virulence factors, its innate resistance to antimicrobials, as well as it ability to grow

    as a biofilm within the lung (Lyczak et al., 2000; Regnath et al., 2004; Singh et al.,

    2000). As an environmental organism, perhaps P. aeruginosa’s greatest asset is the

    ability to survive in a wide variety of niches and under continually changing and

    stressful conditions. One mechanism for survival under unfavourable conditions is

    the biofilm mode of growth.

  • 6

    C. The biofilm mode of growth.

    Biofilms are organized communities of microorganisms that are attached to

    biotic or abiotic surfaces encased in a self-produced polymeric matrix (Costerton et

    al., 1999). This mode of growth is clinically significant as it is predicted that biofilms

    account for >65% of all bacterial infections, including chronic and device-related

    infections (Lewis, 2001). Unlike free-swimming bacteria, when living as part of a

    biofilm bacteria are resistant to high levels of antibiotics and are able to evade the

    host immune system (Drenkard, 2003; Hoyle & Costerton, 1991; Leid et al., 2005;

    Lyczak et al., 2000). This resistance pattern is not due to a classic genetic

    mechanism for antibiotic resistance, but instead appears to be an intrinsic and

    complex property of the biofilm mode of growth (Drenkard, 2003). It is thought that

    in addition to being encased in a polymeric matrix which affords protection from the

    immune system, many of the bacteria within the biofilm enter a stationary state of

    growth and are therefore not susceptible to antibiotic agents that target actively

    dividing cells (del Pozo & Patel, 2007; Drenkard, 2003; Leid et al., 2005). These

    properties of the biofilm contribute to the persistence of bacteria in chronic

    infections.

    Biofilm development occurs in several stages: the first step is bacterial

    attachment to a surface, followed by microcolony formation, expansion, and finally

    detachment, where bacteria return to the planktonic phase of growth to leave the

    biofilm and establish a new community (O'Toole et al., 2000). Given that it is only

    after the bacteria switch from the planktonic to sessile mode of growth that traditional

  • 7

    treatment regimes are no longer effective, strategies aimed at preventing initial

    bacterial attachment, and subsequent biofilm formation, must be examined.

    D. Attaching to a surface

    In an effort to understand the processes involved in the initiation of biofilm

    development, O’Toole and Kolter (1998) examined the roles of flagella and T4P in

    the early stages of P. aeruginosa biofilm formation in a simple experiment examining

    a series of surface attachment defective (sad) mutants and their ability to colonize

    polyvinylchloride (PVC) plastic (O'Toole & Kolter, 1998). Flagellum-mediated

    swimming motility was found to be required for bacterial translocation to the surface

    for initial attachment to occur, however the role of T4P was less clear. Under their

    experimental conditions, T4P did not appear to be necessary for initial attachment to

    the surface as non-piliated (pilA-) strains became associated with the surface but

    failed to form microcolonies (O'Toole & Kolter, 1998). They speculated that T4P

    mediated twitching motility resulted in bacterial aggregation and that T4P

    participated in stabilizing interactions with the surface following attachment.

    However, work from our group and others has demonstrated that T4P function as

    important bacterial adhesins to both biotic and abiotic surfaces and are essential in

    the first step of biofilm formation when shear forces are present (Chi et al., 1991;

    Chiang & Burrows, 2003; Emam et al., 2006; Giltner et al., 2006; Hahn, 1997).

    Chiang and Burrows (2003) demonstrated that pilus-mediated initial

    attachment to glass was required for biofilm formation, as non-piliated (pilA-)

    bacteria failed to adhere to the surface. Interestingly, hyperpiliated, but twitching

  • 8

    deficient (pilT-), strains were capable of adhering to the surface and forming dense

    biofilms, indicating that twitching motility is not required for attachment or biofilm

    formation (Chiang & Burrows, 2003). Twitching motility however, did appear to be

    required for “normal” biofilm development as the pilT- strains failed to produce

    differentiated structures like those of the wild-type strains, pointing to T4P’s

    importance in both the establishment and development of P. aeruginosa biofilms

    (Chiang & Burrows, 2003). T4P clearly play an important role in bacterial

    attachment to abiotic surfaces, which can contribute to disease in the case of

    implanted medical devices, or act as a reservoir and potential source of infection.

    The role of T4P in attachment to eukaryotic cells has been widely reported

    (Chi et al., 1991; Emam et al., 2006; Lee et al., 1989; Ramphal et al., 1984; Sato et

    al., 1988; Tang et al., 1995; Woods et al., 1980). Significantly, P. aeruginosa T4P

    have been demonstrated to be essential virulence factors and adhesins to the Gg4

    (commonly termed asialoGM1) receptor on CF-epithelial cells (Comolli et al., 1999;

    Gupta et al., 1994; Hahn, 1997; Krivan et al., 1988; Lee et al., 1994; Saiman &

    Prince, 1993; Schweizer et al., 1998; Sheth et al., 1994). In one study Comolli et al.

    (1999) demonstrated in vitro that P. aeruginosa lab strain PA103 showed increased

    binding to Gg4-supplemented Madin-Darby canine kidney (MDCK) cells compared

    to those cells not treated with Gg4, and that this binding was inhibited when cells

    were treated with a commercially available antibody to Gg4. Additionally, they

    showed that this binding was T4P specific as a non-piliated strain showed little

    binding to the cells or to Gg4 (Comolli et al., 1999).

  • 9

    Despite this evidence supporting its role in attachment, there are

    contradictory reports on the specificity of Gg4 as a receptor for P. aeruginosa

    (Emam et al., 2006; Schroeder et al., 2001). One study examined the binding of

    PA103 as well as eight minimally-passaged clinical isolates to Gg4-supplemented

    MDCK cells and found that while PA103 appeared to demonstrate increased

    adherence to these cells, none of the clinical isolates displayed similarly enhanced

    binding (Schroeder et al., 2001). This study was also critical of the use of the

    commercially available Gg4 antibody in the cell-binding-inhibitory studies, as they

    determined that this antibody preparation was contaminated with high titres of

    antibodies to several P. aeruginosa antigens, including pili, thus bringing into

    question the role of Gg4 as a pilus, or P. aeruginosa, receptor (Schroeder et al.,

    2001). These findings were supported by work by members of our group that

    demonstrated failure of eleven clinical and laboratory strains to bind to Gg4 or other

    glycosphingolipids in a thin-layer chromatography overlay experiment, as well as a

    glycosphingolipid ELISA, while the control strain enterohaemorrhagic E. coli strain

    CL56 displayed binding (Emam et al., 2006). Additionally, P. aeruginosa were found

    to bind to cultured lung epithelial cells in a pilus-dependent manner and the

    frequency of binding was not altered when Gg4 was chemically depleted from

    epithelial cells indicating that Gg4 is not a specific P. aeruginosa receptor (Emam et

    al., 2006). It seems more likely that as a primarily environmental organism, P.

    aeruginosa has evolved to utilize T4P to attach to diverse surfaces in a receptor-

    independent manner.

  • 10

    II. TYPE IV PILI

    A. History and Classification

    Some of the earliest recorded observations of bacterial pili were micrographs

    taken at the turn of the 20th century by Hinterberger and Reitman who observed

    numerous fine threads on cultures of Pseudomonas pyocyanea (aeruginosa) on old

    media; flagella however were observed more frequently on bacteria on moist agar

    and in liquid culture (described in Houwink & van Iterson, 1950). Years later, while

    studying flagella of E. coli and P. pyocyanea, Houwink and van Iterson made the first

    detailed report of “non-flagellar appendages”, noting that these long thread-like

    structures were apparent on bacteria associated with surfaces, however were not

    observed on swimming bacteria (Houwink & van Iterson, 1950). In 1955 the terms

    “fimbriae” (Latin for threads or fibres) and “pili” (Latin for hair or fur) were introduced

    to describe these structures (Brinton, 1959; Duguid et al., 1955) thus establishing

    the divisive nomenclature of these bacterial structures. In 1975, a standardized

    classification system was proposed and the flexible pili that contribute to twitching

    motility were placed into Group IV (Ottow, 1975). With the advent of molecular

    biology and structural biochemistry, T4P have been further divided into two sub-

    types, T4Pa and T4Pb based on protein size, sequence and structure (described

    below) (Craig et al., 2004). The T4Pa of P. aeruginosa are the focus of this thesis.

    B. Distribution

    T4P are primarily found on Gram negative bacteria of the Beta, Gamma and

    Delta Proteobacteria, including Haemophilus influenzae, Moraxella spp., Neisseria

  • 11

    gonorrhoeae, N. meningitidis, Pseudomonas spp., Eikenella corrodens, Myxococcus

    xanthus, Dichelobacter nodosus, enteropathogenic E. coli and Vibrio chloerae

    (Bakaletz et al., 2005; Fullner & Mekalanos, 1999; Henrichsen, 1975; Mattick, J.,

    Hobbs, M., Cox, PT., Dalrymple, BP, 1993). As well, there is a report of twitching

    motility and polar fimbriae in the Gram positive bacterium Streptococcus sanguis

    (Henrichsen, 1975), and the recently published genome of this opportunistic

    pathogen contains PilB and PilT-like proteins, as well as an open reading frame with

    weak homology to the pilin structural subunit, suggesting that these structures may

    be more widely distributed than once thought (Xu et al., 2007).

    Interestingly, T4Pb seem to be strictly associated with human pathogens that

    colonize the intestines, including V. cholerae, S. enterica serovar Typhi, EPEC and

    ETEC (Donnenberg et al., 1992; Giron et al., 1991; Giron et al., 1994; Shaw &

    Taylor, 1990; Taniguchi et al., 1995; Zhang et al., 2000). In contrast, bacteria

    possessing T4Pa are found in a variety of different environments (Craig et al., 2004).

    Regardless, both types of T4P are considered important colonization factors.

    C. Pilin to Pilus

    The T4P of P. aeruginosa are composed of thousands of monomeric subunits

    of the protein PilA, or pilin, encoded by the gene pilA (Pasloske et al., 1985). The

    pilins are approximately 150 amino acids in length and ~15 kDa (Frost &

    Paranchych, 1977; Pasloske et al., 1985). PilA is synthesized as a pre-pilin with a

    short (5 - 6 amino acids), positively charged leader sequence which is highly

    conserved amongst species (Alm & Mattick, 1997; Mattick, 2002; Watson et al.,

  • 12

    1996). This sequence is followed by an N-terminal hydrophobic α helix, and a region

    of anti-parallel β-strands which end in a C-terminal disulphide-bonded loop (DSL),

    creating overall a ladle- or lollipop-like structure (Figure 1-1) (Craig et al., 2003;

    Craig et al., 2004; Craig et al., 2006; Dalrymple & Mattick, 1987; Hazes et al., 2000;

    Parge et al., 1995).

    Pseudopilins are pilin-like proteins which are thought to combine to form a

    short dynamic fibre (the pseudopilus) that acts as a piston in the type II secretion

    system of many Gram-negative bacteria (Durand et al., 2003; Filloux, 2004). In the

    event of over-expression of the major subunit of the pseudopilus, an extracellular

    pilin-like fibre is formed (Durand et al., 2003; Vignon et al., 2003). The pseudopilin

    subunits, like T4Pa subunits are composed of a similar and conserved positively

    charged N-terminal leader sequence followed by a 20 – 30 amino acid α-helix, and a

    globular head domain composed of four β-strands, however, the C-terminal DSL is

    not present (Kohler et al., 2004; Nunn & Lory, 1993). Through the study of the

    pseudopili of Klebsiella oxytoca and P. aeruginosa information on the early events of

    pseudopilus, and ostensibly pilus biogenesis, have been generated. The pre-

    pseudopilins are translated in the cytoplasm and translocated across the inner

    membrane in a Sec-dependent, SRP-dependent process, in which the N-terminal

    leader sequence acts as a Sec-signal sequence (Arts et al., 2007; Francetic et al.,

    2007). The hydrophobic α-helix portion of the protein becomes embedded in the

    membrane, with the C-terminus on the periplasmic side. Once in the membrane, the

    signal-sequence is cleaved by the bi-functional enzyme XcpA/PilD, which also N-

    methylates the new N-terminal residue, which is most often phenylalanine (Arts et

  • 13

    al., 2007; Strom et al., 1991; Strom & Lory, 1992; Strom et al., 1994). This

    processing step is essential to make the pilins competent for assembly, as mutants

    lacking XcpA/PilD are unable to express pili on the cell surface.

    A B

    Figure 1-1. Crystal structures of the pilin structural subunits of N. gonorrhoeae and P.

    aeruginosa. A. N. gonorrhoeae pilin PilE, displaying the different regions of the protein and the

    covalently attached glycan at Ser63 (orange) and a phosphate at Ser68 (red). B. P. aeruginosa

    group II PAK pilin, PilA. The α-helix and β- strands are in grey, the αβ-loop is in green, the D-region

    (disulphide-bonded loop region) is purple, the disulphide bond is cyan. Reproduced from Craig et

    al., 2004, with the permission of Nature Reviews Microbiology, Nature Publishing Group.

    A BA B

    Figure 1-1. Crystal structures of the pilin structural subunits of N. gonorrhoeae and P.

    aeruginosa. A. N. gonorrhoeae pilin PilE, displaying the different regions of the protein and the

    covalently attached glycan at Ser63 (orange) and a phosphate at Ser68 (red). B. P. aeruginosa

    group II PAK pilin, PilA. The α-helix and β- strands are in grey, the αβ-loop is in green, the D-region

    (disulphide-bonded loop region) is purple, the disulphide bond is cyan. Reproduced from Craig et

    al., 2004, with the permission of Nature Reviews Microbiology, Nature Publishing Group.

  • 14

    1. N-terminal α-helix. The N-terminal α-helix portion of the mature pilin is

    highly conserved among all T4P, including T4Pa and T4Pb and can be divided into

    two regions, α1-N (residues 1 - 28) and α1-C, (residues 29 - ~53) (Figure 1-1) (Craig

    et al., 2003; Craig et al., 2004). The α1-N portion of the α-helix is hydrophobic and

    retains the pilins in the inner membrane for additional processing (i.e. post

    translational modifications) and as a reservoir for rapid pilus assembly, and is also

    involved in mediating pilus assembly through hydrophobic interactions between

    subunits (Castric, 1995; Craig et al., 2004; Dalrymple & Mattick, 1987; Hegge et al.,

    2004; Stimson et al., 1995; Stimson et al., 1996). The α1-C portion of the α-helix is

    amphipathic in nature, is embedded in the C-terminal region of the protein and

    contributes to pilus stability (Craig et al., 2004). The α-helix, in part, drives pilus

    assembly; the α-helices of the pilin subunits form a strong but flexible hydrophobic

    helical bundle that composes the core of the pilus fibre, while the β-strands of the C-

    terminal region cover the surface of the fibre (Aas et al., 2007b; Craig et al., 2003;

    Craig et al., 2004; Craig et al., 2006; Parge et al., 1995).

    2. αβ-loop. Joining the α-helix and the β-strands of the pilin is a short region

    called the αβ-loop (Figure 1-1). Crystal structure studies have demonstrated that

    this region is structurally variable among all T4Pa examined to date; in P.

    aeruginosa strain PAK this region is composed of a minor β-sheet made up of three

    short β-strands, while in the related P. aeruginosa strain K122-4 this region is poorly

    organized, with a short α-helix (Craig et al., 2004; Hazes et al., 2000; Keizer et al.,

    2001). In N. gonorrhoeae the αβ-loop is extended and forms a ridge containing

  • 15

    residues which are post-translationally modified (Craig et al., 2006; Forest et al.,

    1999; Marceau et al., 1998; Parge et al., 1995). In all assembled T4P examined to

    date, this region is partially buried within the pilus fibre and therefore participates in

    subunit-subunit interactions, but it is also partially exposed on the pilus surface thus

    influencing the fibres’ interaction with the environment (Craig et al., 2003; Craig et

    al., 2004; Craig et al., 2006).

    3. β-strands. Following the αβ-loop, and comprising the bulk of the globular

    head domain, is a short series of anti-parallel β-strands which form the outer face of

    the pilus fibre (Figure 1-1) (Craig et al., 2003; Craig et al., 2004; Craig et al., 2006).

    Despite substantial sequence differences in this part of the protein, all T4P contain

    this β-sheet region which acts as the structural core of the globular head. Within the

    T4Pa this region is composed of 4 strands, while in T4Pb it is made up of 5 strands

    (Craig et al., 2003; Craig et al., 2004). Interestingly, this portion of the T4Pb protein

    contains a novel protein fold (Craig et al., 2003) which has led some researchers to

    suggest that the evolutionary relationship between T4Pa and T4Pb may be indirect

    (Craig et al., 2004).

    4. Disulphide-bonded-loop. At the C-terminal portion of the protein there

    are 2 conserved cysteine residues that form a disulphide-bonded loop (DSL) which

    has been shown to be essential for pilus assembly (Figure 1-1) (Harvey et al.,

    submitted 2007). However, there are studies in which the Cys residues were

    mutated or replaced, or in which the disulphide-bond isomerase DsbA was mutated,

  • 16

    which have shown that the DSL is not necessary for fibre assembly, but did appear

    to play a role in pilus function (Burrows, 2005; Farinha et al., 1994; Ha et al., 2003).

    The length of the DSL varies between species with the most dramatic

    differences occurring between those of T4Pa and T4Pb, with average lengths of 22

    and 55 amino acids respectively (Craig et al., 2004). These differences in size and

    sequence result in distinct folding patterns and influence subunit-subunit

    interactions. However, in all pilins, this region is predicted to form a protruding edge

    on the opposite side of the globular head from the αβ-loop ridge, thereby influencing

    the fibres’ surface chemistry (Craig et al., 2004; Craig et al., 2006). The P.

    aeruginosa pilins that have been crystallized (PAK and K122-4) have DSLs of 12

    amino acids which are predicted to be exposed only at the tip of the pilus, and buried

    along the length of the pilus fibre (Craig et al., 2003; Craig et al., 2004; Hazes et al.,

    2000; Keizer et al., 2001; Lee et al., 1994). This region is of particular interest in P.

    aeruginosa, as it is implicated as the adhesive part of the pilus that mediates

    attachment to host cells and abiotic surfaces, and in some strains has been shown

    to be glycosylated (Comer et al., 2002; Giltner et al., 2006; Lee et al., 1994; Sheth et

    al., 1994). The DSL is also considered a good target for the development of

    vaccines against P. aeruginosa infections (Audette et al., 2004; Cachia et al., 1998;

    Sheth et al., 1995; Suh et al., 2001).

    5. The pilus fibre. Mature pili are long (> 1 µm), thin (~ 6 – 8 nm) flexible

    fibres that are extremely strong, capable of withstanding > 100 pN of force (Figure 1-

    2) (Maier et al., 2002; Maier, 2005; Mattick, 2002; Parge et al., 1987; Strom & Lory,

  • 17

    1993). There is much interest in how the pilin subunits combine to form such

    remarkable filaments. Recently Craig et al. (2006) developed a high resolution

    model of assembled T4P from N. gonorrhoeae, using both cryo-electron microscopy

    and crystallography. The model predicts that the subunits pack in a twisted three-

    helix bundle with the N-terminal α helices pointed towards the pilus base (Craig et

    al., 2006). Like previous models, this current model places the first half of the N-

    terminal α helix (α1-N) of the subunits inside the central core of the fibre in a

    staggered helix of interlocking subunits, stabilized through hydrophobic interactions

    (Craig et al., 2004; Craig et al., 2006; Parge et al., 1995). Contrary to previous

    models that predicted the β-sheets of the globular heads were wrapped tightly

    around the fibre (Forest & Tainer, 1997; Parge et al., 1995), this newer model

    predicts that there are deep grooves separating the heads resulting in a “corrugated”

    surface (Craig et al., 2006). The αβ loop and the C-terminal domain are located at

    the base of the grooves and polar interactions are observed between the αβ loop of

    one subunit and the C-terminal region of the adjacent subunit. Post-translational

    modifications of the N. gonorrhoeae pilin are predicted to protrude from exterior of

    the filament towards the solvent (Craig et al., 2004; Craig et al., 2006). The outer

    diameter of the fibre is ~ 60 Å while the central core is of variable diameter (~6 – 11

    Å). The authors speculated that this channel is not solvent filled but is a

    compressible space which may allow for further filament flexibility (Craig et al.,

    2006).

  • 18

    This model helps to explain the strength and flexibility of the thin T4P filament: the

    tight hydrophobic interactions between the α-helices provide strength, while the gaps

    and grooves between the pilin heads provide flexibility and additional stability via the

    αβ loop and C-terminal interactions between subunits. This model also helps

    explain how pilin architecture can be conserved amongst bacteria, yet have

    functional and antigenic differences. The conserved structural core supports the

    diverse αβ loop and C-terminal regions line the grooves of the fibre which are

    Figure 1-2. The pole of a P. aeruginosa cell displaying T4P and flagella. This image (100,000x

    magnification) demonstrates the single polar flagellum of P. aeruginosa and multiple, thin, long type

    IV pili. This is a pilD- mutant strain, unable to retract the T4P. Courtesy of P. Chiang.

    Figure 1-2. The pole of a P. aeruginosa cell displaying T4P and flagella. This image (100,000x

    magnification) demonstrates the single polar flagellum of P. aeruginosa and multiple, thin, long type

    IV pili. This is a pilD- mutant strain, unable to retract the T4P. Courtesy of P. Chiang.

  • 19

    exposed to the environment and thereby subject to host-immune or other selective

    pressures.

    Although the N. gonorrhoeae pilus represents an excellent model, the distinct

    differences between Neisseria and P. aeruginosa pili suggest that other models

    must be considered in order to understand these structures. Although less detailed

    than that of N. gonorrhoeae, X-ray diffraction analysis of the PAK fibres reveals that

    these fibres have a 52 Å outer diameter and a 12 Å inner diameter, with the subunits

    arranged in either a right-handed one-start helix with four subunits per turn or a left-

    handed three-start helix with four subunits per turn (Craig et al., 2004). The most

    significant structural and biological difference between the pili of these two species

    involves the putative binding region. Neisseria pili are thought to be capped with a

    separate PilC protein, a component involved in pilus-mediated adhesion to host cells

    (Merz & So, 2000; Nassif et al., 1997; Scheuerpflug et al., 1999) whereas in P.

    aeruginosa, PilA is thought to be used as both the structural component of the fibre

    and the adhesive tip (Irvin et al., 1989; Lee et al., 1989; Lee et al., 1994). As

    mentioned above, the C-terminal DSL portion of PilA has been demonstrated to be

    the adhesive component of the pilus (Lee et al., 1994). While it appears to be at

    least partially buried along the length of the fibre, antibody studies have

    demonstrated its exposure at the pilus tip (Lee et al., 1994). Structural studies

    support the contention that the DSL is obscured along the length of the fibre and

    exposed only at its tip (Craig et al., 2004). Based on the N. gonorrhoeae pilin

    packing model, up to three DSL sub-domains may be available for binding at the tip

    of the P. aeruginosa T4P (Burrows, 2005; Craig et al., 2006)

  • 20

    PilC

    PilT PilU

    PilD/XcpA

    PilQ

    Figure 1-3. Simplified cartoon of P. aeruginosa pilus assembly complex. Pre-pilins (PilA) are

    produced in the cytoplasm, targeted to the inner membrane (IM) via the N-terminal signal peptide.

    PilD/XcpA cleaves off the signal sequence and N-methylates the N-terminal phenylalanine. The

    mature pilins are polymerized into the pilus fibre through the action of PilB. The pilus crosses the

    outer membrane (OM) through the PilQ pore. Retraction of the pilus occurs via disassembly via the

    depolymerase(s) PilT and/or PilU. See text for references and details. PG = peptidoglycan.

    IM

    OM

    PG

    PilA

    Pilus

    PilQ

    PilB

    PilC

    PilT PilU

    PilD/XcpA

    PilQ

    Figure 1-3. Simplified cartoon of P. aeruginosa pilus assembly complex. Pre-pilins (PilA) are

    produced in the cytoplasm, targeted to the inner membrane (IM) via the N-terminal signal peptide.

    PilD/XcpA cleaves off the signal sequence and N-methylates the N-terminal phenylalanine. The

    mature pilins are polymerized into the pilus fibre through the action of PilB. The pilus crosses the

    outer membrane (OM) through the PilQ pore. Retraction of the pilus occurs via disassembly via the

    depolymerase(s) PilT and/or PilU. See text for references and details. PG = peptidoglycan.

    IM

    OM

    PG

    PilA

    Pilus

    PilQ

    PilB

    6. Pilus assembly complex. The regulation, assembly and disassembly of

    the mature pilus involves the products of over 50 genes (Alm & Mattick, 1997;

    Mattick, 2002). A model of pilus assembly which highlights some of the proteins

    relevant to this thesis, including the motor proteins PilB, the pilin polymerase, and

    PilT the pilin depolymerase, is presented in Figure 1-3. PilB and PilT are predicted

  • 21

    to be peripherally associated with the inner membrane on the cytoplasmic face at

    the base of the pilus in close proximity to the multi-protein complex that forms the

    base from which the pilus extends (Chiang et al., 2005; Mattick, 2002).

    D. Functions

    T4P participate in many functions including bacteriophage infection (Bradley,

    1974; Whitchurch & Mattick, 1994), transformation (Wolfgang et al., 1998), DNA

    uptake (Aas et al., 2002; Fussenegger et al., 1997; Hamilton & Dillard, 2006;

    Mattick, 2002), adhesion and twitching motility (Burrows, 2005; Mattick, 2002).

    Unlike some T4P-possessing bacteria such as N. gonorrhoeae and P. stutzeri, P.

    aeruginosa are not naturally competent and there is little evidence that the T4P in

    this system are involved in DNA binding and uptake (Graupner et al., 2000; van

    Schaik et al., 2005; Wolfgang et al., 1998). For this reason, I will focus on the

    functions of twitching motility and adherence.

    1. Twitching motility. Twitching motility is the hallmark of T4P. It is a form

    of flagellar-independent motility that can occur on moist surfaces (Bradley, 1980;

    Henrichsen, 1983; Kaiser, 1979; Lautrop, 1961; Ottow, 1975). Twitching bacteria

    typically move as groups, or “rafts”, and use this motility as an important means of

    rapidly colonizing new environments, and for biofilm establishment and

    differentiation (Klausen et al., 2003a; Klausen et al., 2003b; Mattick, 2002; O'Toole &

    Kolter, 1998; Semmler et al., 1999). The term “twitching” refers to the jerky fashion

    in which bacteria move during this type of motility, which is achieved by the

  • 22

    extension of the T4P, attachment of the pilus tip to a surface and subsequent

    retraction of the fibre, effectively moving the bacterial cell body towards the point of

    pilus attachment (Bradley, 1980; Henrichsen, 1972; Mattick, 2002; Skerker & Berg,

    2001). Pilus extension is achieved through the polymerization of the pilin subunits

    via the activity of the ATPase PilB, while retraction is due to disassembly of the fibre

    by action of the ATPases PilT and/or the PilT homologue PilU (Hobbs & Mattick,

    1993; Kaiser, 2000; Merz et al., 2000; Whitchurch et al., 1991; Whitchurch & Mattick,

    1994).

    Figure 1-4. Twitching motility assay. A simple assay for twitching motility. A. Bacteria are stab

    inoculated to the bottom of the agar plat. B. Bacteria migrate away from the point of inoculation via

    twitching motility between agar and plastic of the plate forming a “halo” or twitching zone around the

    colony. C. The twitching zone can be more easily visualized by removing the agar and staining the

    bacteria attached to the plastic with crystal violet dye.

    A

    CB

    Figure 1-4. Twitching motility assay. A simple assay for twitching motility. A. Bacteria are stab

    inoculated to the bottom of the agar plat. B. Bacteria migrate away from the point of inoculation via

    twitching motility between agar and plastic of the plate forming a “halo” or twitching zone around the

    colony. C. The twitching zone can be more easily visualized by removing the agar and staining the

    bacteria attached to the plastic with crystal violet dye.

    A

    CB

  • 23

    Macroscopically, twitching motility can be observed as a thin opaque

    monolayer of cells surrounding the edge of a colony growing on an agar plate. This

    zone can be enhanced when bacteria are stab inoculated to the bottom of an agar

    plate where the bacteria grow at the interstitial surface between the agar and the

    plastic (Figure 1-4) (Semmler et al., 1999). Twitching zones can reach several

    centimetres in diameter after overnight growth. It is thought that the smooth surface

    of the agar and the moist conditions at the bottom of the agar are ideal for twitching

    motility, whereas on the surface of agar, minor surface variations and variations in

    moisture levels retard twitching motility (Semmler et al., 1999). At the microscopic

    level, twitching motility is seen at the outermost edge of the bacterial colony where a

    lattice-like pattern of cells moving away from the colony is formed (Mattick, 2002).

    The motive force of twitching motility is achieved by the retraction of the T4P;

    the first evidence for this was the inhibition of twitching by pilus-specific phage and

    antibodies which interfered with the retraction of the pili (Bradley, 1972a; Bradley,

    1972b). Additional evidence for this action was the lack of twitching motility in hyper-

    piliated strains, which were found to be retraction deficient (Merz et al., 2000;

    Skerker & Berg, 2001; Whitchurch et al., 1991). The hyper-piliated phenotype is due

    to mutations within PilT or PilU, the ATPases involved in pilus disassembly (Chiang

    & Burrows, 2003; Mattick, 2002; Whitchurch & Mattick, 1994). Direct observational

    evidence for pilus retraction and measurement of the forces generated by this action

    has been obtained using N. gonorrhoeae and P. aeruginosa (Merz et al., 2000; Merz

    & Forest, 2002; Skerker & Berg, 2001). These studies demonstrated that individual

  • 24

    pili retract independently of one another, and that retraction forces can exceed 140

    pN, two orders of magnitude greater than a eukaryotic kinesin motor.

    In order for twitching motility to occur, the tips of the T4P must be firmly

    attached to a surface in order to anchor the bacteria; without attachment, the

    bacterial cells would not be translocated upon pilus retraction.

    2. Adherence. In P. aeruginosa T4P are considered to be the dominant

    adhesins responsible for initiating attachment and subsequent infections (Hahn,

    1997; Irvin et al., 1989). As described earlier (Sections I, B2 & D) there is evidence

    that P. aeruginosa T4P bind to the epithelial cell receptor Gg4 (Comolli et al., 1999;

    Gupta et al., 1994; Hahn, 1997; Krivan et al., 1988; Lee et al., 1994; Saiman &

    Prince, 1993; Schweizer et al., 1998; Sheth et al., 1994). However, unlike obligate

    pathogens such as N. gonorrhoeae, P. aeruginosa’s ability to exploit many different

    niches, and to be an opportunistic pathogen of a variety of organisms is likely due in

    part to the general rather than specific adhesive properties of its T4P. P. aeruginosa

    is known to be associated with device-related infections, particularly nosocomial

    infections involving catheters and ventilators (Pierce, 2005; Rosenthal et al., 2006)

    and has long been known to colonize contact lenses, leading to microbial keratitis

    (Willcox, 2007). Also, P. aeruginosa has been shown to be a problem in industrial

    settings and on medical equipment, as it binds readily to stainless steel (Giltner et

    al., 2006; Vanhaecke et al., 1990). Giltner et al. (2006) have demonstrated a role for

    T4P in P. aeruginosa adherence to stainless steel, as wild-type and flagella-deficient

    strains were capable of binding to stainless steel while pilin-deficient strains did not

  • 25

    bind. Further, the DSL of the pilin was implicated as the adhesive domain since

    purified T4P were shown to bind to stainless steel, polystyrene and PVC, and this

    binding was inhibited with the addition of an antibody against the pilin DSL (Giltner et

    al., 2006). The authors point out that the sequence of the DSL varies greatly

    amongst P. aeruginosa strains, however all variants tested have retained the ability

    to bind to biotic and abiotic surfaces, albeit with different affinities.

    The strains examined by Giltner et al. (2006) varied in amino acid sequence

    in the C-terminal region, however all strains tested had DSLs of 12 amino acids.

    There have been reports of strains with DSLs of different lengths among P.

    aeruginosa T4P, for example the DSL of strain 1244 is 17 amino acids and that of

    strain G7 is 31 amino acids (Castric et al., 1989; Castric & Deal, 1994; Spangenberg

    et al., 1995). This range in size and sequence diversity of the adhesive region of the

    protein raises questions regarding binding specificities of the T4P of this organism.

    E. Diversity of PilA

    In the PAO1 chromosome, pilA is located at approximately 72 minutes on the

    genetic map, and lies in a region that contains several pilin-associated genes,

    (Farinha et al., 1993; Mattick et al., 1996). The pilA gene is divergently transcribed

    from the other pilin-associated genes in this region, with pilBCD positioned upstream

    and tRNA-Thr situated downstream (Figure 1-5) (Hobbs et al., 1988; Mattick et al.,

    1996; Stover et al., 2000). Genetic mapping and partial sequencing of the

    chromosomes of small groups of P. aeruginosa strains has revealed that this area is

    a region of extensive inter-strain variability (Pasloske et al., 1988; Romling et al.,

  • 26

    1995; Schmidt et al., 1996; Spangenberg et al., 1995). As mentioned above

    (Section II, C4) there have been reports of several pilin variants in P. aeruginosa,

    which form several distinct groups based on sequence, pilin length and DSL length

    (Figure 1-5) (Castric & Deal, 1994; Pasloske et al., 1988; Spangenberg et al., 1995;

    Spangenberg et al., 1997).

    1. Group II. Group II pilins are represented by those of the well-described

    laboratory strains PAK and PAO1, which have mature T4Ps of 143-144 amino acids

    long, DSLs of 12 amino acids, a pilin gene %G+C content of 48% (versus 67%

    pilB

    pilB

    pilB

    pilA

    pilA

    pilA

    ORF1

    pilO/tfpO

    tRNA-Thr

    tRNA-Thr

    tRNA-Thr

    Group II (PAO1, PAK)

    Group I (P1, 1244)

    Group III (G7)

    Figure 1-5. Genetic organization of the pilin locus of P. aeruginosa. Three different pilin

    variants have been described, two of which are associated with novel open-reading frames directly

    downstream. The group I ORFhas been identified as pilO/tfpO, which encodes a protein involved in

    the glycosylation of pilin (Castric, 1989; Castric & Deal, 1994; Spangenberg et al., 1995; Castric,

    1995). Figure adapted from Spangenberg et al. (1995). Not to scale.

    pilB

    pilB

    pilB

    pilA

    pilA

    pilA

    ORF1

    pilO/tfpO

    tRNA-Thr

    tRNA-Thr

    tRNA-Thr

    Group II (PAO1, PAK)

    Group I (P1, 1244)

    Group III (G7)

    pilB

    pilB

    pilB

    pilA

    pilA

    pilA

    ORF1

    pilO/tfpO

    tRNA-Thr

    tRNA-Thr

    tRNA-Thr

    Group II (PAO1, PAK)

    Group I (P1, 1244)

    Group III (G7)

    Figure 1-5. Genetic organization of the pilin locus of P. aeruginosa. Three different pilin

    variants have been described, two of which are associated with novel open-reading frames directly

    downstream. The group I ORFhas been identified as pilO/tfpO, which encodes a protein involved in

    the glycosylation of pilin (Castric, 1989; Castric & Deal, 1994; Spangenberg et al., 1995; Castric,

    1995). Figure adapted from Spangenberg et al. (1995). Not to scale.

  • 27

    overall for PAO1) and a tRNA-Thr gene immediately downstream of the 3’ end of

    pilA (Kiewitz & Tummler, 2000; Sastry et al., 1985; Spangenberg et al., 1995;

    Wolfgang et al., 2003). Pili from these strains have been shown to specifically

    interact with the Gg4 glycolipid in binding studies (Comolli et al., 1999; Gupta et al.,

    1994; Hahn, 1997; Krivan et al., 1988; Lee et al., 1994; Saiman & Prince, 1993;

    Schweizer et al., 1998; Sheth et al., 1994).

    2. Group I. Group I pilins are related to those from the P1 strain, a CF

    isolate, which was originally identified as having an unusual pilin sequence over the

    last three quarters of the protein, with the most notable features being its mature

    length of 148 amino acids, a 17 amino acid DSL, and a pilin gene with a %G+C

    content of 52% (Pasloske et al., 1988). Additional strains with group I pili, including

    strain 1244, were characterized, and were found to contain an additional 1.4 kb open

    reading frame between pilA and tRNA-Thr, later named pilO (Figure 1-5) (Castric,

    1995; Castric et al., 1989; Castric & Deal, 1994; Spangenberg et al., 1995).

    3. Group III. Group III T4P were first identified in a single report of a strain,

    G7, with another unusual pilin sequence. Group III pilins have a %G+C content of

    54.8%, and are the longest mature pilin, composed of of 173 amino acids, with a

    DSL of 31 residues (Spangenberg et al., 1995). Similar to group I strains, additional

    DNA was found between the pilA and tRNA-Thr genes; however it was only 798

    base pairs long, and failed to react with a probe made from the corresponding region

    from group I strains; it was designated ORF1 (Spangenberg et al., 1995).

  • 28

    Interestingly, this additional ORF had a %G+C content of 48.4%, different from both

    that of the pilin gene and average chromosomal content (Spangenberg et al., 1995).

    4. Horizontal gene transfer. There is less than 30% sequence identity

    between the P. aeruginosa pilin groups; in phylogenetic analyses, each group is

    more closely related to pilins from other species than to one another (Castric & Deal,

    1994; Hazes et al., 2000; Spangenberg et al., 1995; Spangenberg et al., 1997). A

    phylogenetic analysis of pilins from 19 P. aeruginosa strains, as well as pilins from

    other common T4P-possessing organisms, demonstrated that group I T4P form a

    tight independent cluster, while the group II pilins were more closely related to pilins

    from D. nodosus and Myxococcus xanthus. The single group III pilin grouped with

    E. corrodens, D. nodosus, N. gonorrhoeae and N. meningitidis, though not closely

    (Spangenberg et al., 1997). The pilins of P. aeruginosa, E. corrodens and D.

    nodosus vary greatly in their %G+C content relative to the rest of their genomes, in

    contrast to the pilins of Neisseria and Moraxella which form tight phylogenetic

    clusters and have pilins of the same %G+C content and codon usage as their

    chromosomes (Spangenberg et al., 1997). The pilins of P. aeruginosa, E. corrodens

    and D. nodosus have a similar %G+C content to Moraxella suggesting that this locus

    was acquired from other species by horizontal gene transfer (Castric & Deal, 1994;

    Hazes et al., 2000; Kiewitz & Tummler, 2000; Mattick, 2002; Spangenberg et al.,

    1995).

    Several factors likely contribute to the diversity of pilA, including its

    chromosomal organization (adjacent to a tRNA gene) and selective pressures due to

  • 29

    interaction of pili with the environment and host immune systems. tRNA genes are

    common sites of bacteriophage and plasmid integration, possibly due to their

    conserved structure and the presence of multiple copies on bacterial genomes that

    affords functional redundancy (Cheetham & Katz, 1995; Reiter et al., 1989). There

    is evidence in D. nodosus that virulence-associated regions are often adjacent to

    tRNA genes, and these regions contain genes with homology to N. gonorrhoeae

    plasmids and bacteriophages (Cheetham & Katz, 1995). Upstream of pilA are the

    conserved pilBCD genes; pilBC share similarity to genes involved in type II secretion

    (T2S) in P. aeruginosa, and pilD encodes PilD/XcpA that acts in both the pilin and

    T2S systems (Nunn & Lory, 1993; Strom et al., 1993). Because of the proximity to

    the known site of recombination tRNA-Thr and the conserved upstream genes

    pilBCD, this area of the P. aeruginosa chromosome is a likely region of genetic

    integration and recombination that has resulted in diversity within PilA. It is

    interesting to consider how the size and sequence variations of different pilins

    influence the functions and binding specificities of T4P.

    5. Additional mechanisms of pilin diversity. Group I and III pilins were

    found to be associated with additional genetic material inserted between pilA and

    tRNA-Thr (Castric & Deal, 1994). In the case of group I strains this ORF was named

    pilO and was demonstrated to encode a glycosyltransferase involved in the post-

    translational modification of PilA (Castric, 1995; DiGiandomenico et al., 2002).

    Castric’s group have demonstrated that PilO is a glycosyltransferase that mediates

    the transfer of the O-antigen to the terminal Ser148 residue of the strain 1244 pilin

  • 30

    (Comer et al., 2002; DiGiandomenico et al., 2002). This region is adjacent to the

    DSL, and possibly influences the T4P’s adhesive function. The unknown ORF

    downstream of the group III pilA has no known function, however due to its proximity

    to the pilin structural gene may encode a product that participates in pilin modulation

    or assembly.

    III. T4P GLYCOSYLATION SYSTEMS

    A. Introduction to Gram-negative glycoproteins

    Protein glycosylation is an additional means of modulating a protein’s

    structure and/or function and can contribute to diversity (Benz & Schmidt, 2002;

    Upreti et al., 2003). In recent years, many prokaryotes have been found to produce

    N- and O-linked glycosylated proteins, including the Gram-negative pathogens

    Campylobacter jejuni, N. meningitidis, N. gonorrhoeae and P. aeruginosa (Benz &

    Schmidt, 2002; Power & Jennings, 2003; Schmidt et al., 2003; Upreti et al., 2003).

    Many of the glycosylated proteins of pathogenic bacteria described to date are

    associated with the surface of the organism and include flagellins of P. aeruginosa,

    (Brimer & Montie, 1998), and C. jejuni (Doig et al., 1996), the E. coli adhesins TibA

    and AIDA-1 (Benz & Schmidt, 2002; Lindenthal & Elsinghorst, 2001), and the type IV

    pilins of some strains of P. aeruginosa (Castric et al., 2001), N. meningitidis and N.

    gonorrhoeae (Parge et al., 1995; Stimson et al., 1995).

    The glycosylation of bacterial proteins appears to be strain-dependent in

    many cases; however for such an energetically expensive process, there must be a

    fitness benefit for the strains that produce the modified proteins. The exact

  • 31

    biological function, or benefit, of glycosylation in many cases remains unclear.

    Some functions that have been described include the maintenance of protein

    structure or conformation, increased resistance to proteolytic enzymes, alteration of

    the physicochemical properties of the protein/bacterial surface (e.g. solubility,

    surface charge, hydrophilicity), alteration of immune detection and cell adhesion

    (Banerjee et al., 2002; Benz & Schmidt, 2002; Moens & Vanderleyden, 1997;

    Schmidt et al., 2003; Szymanski et al., 2002; Upreti et al., 2003).

    In several cases the glycosylation of surface proteins of pathogenic bacteria

    has been shown to increase adherence and colonization of host tissue and increase

    pathogen virulence; examples include C. jejuni, in which the general glycosylation

    pathway, which modifies several proteins, has been shown to increase bacterial

    adherence, the outer membrane protein TibA of ETEC E. coli, and the P. aeruginosa

    strain 1244 T4P (Elsinghorst & Weitz, 1994; Lindenthal & Elsinghorst, 2001;

    Smedley et al., 2005; Szymanski et al., 2002). However, the adhesive function of

    other surface proteins does not appear to be modulated by the presence/absence of

    a glycan; examples include the pili of N. meningitidis, and the flagellin of C. jejuni

    (Doig et al., 1996; Stimson et al., 1995).

    There is significant heterogeneity in bacterial glycosylation systems with

    respect to the nature of the target proteins and the sugars that modify them. In order

    to understand the function and significance of glycosylation of bacterial proteins,

    comparison between the wild-type strain and a non-glycosylated, or de-glycosylated,

    version of the same strain must be performed, as generalizations regarding function

    cannot readily be made.

  • 32

    Amongst Gram-negative bacteria, two systems involved in the biosynthesis of

    glycan used for protein glycosylation have been described. The first involves

    dedicated machinery that is specific for protein glycosylation, where most of the

    genes required for glycan synthesis and transfer are grouped together on a genetic

    island. Examples include the well-described systems for glycosylation of C. jejuni

    surface proteins, Neisseria pilins, and P. aeruginosa flagellins (Arora et al., 2001;

    Power et al., 2000; Power et al., 2003; Szymanski et al., 1999; Verma et al., 2006;

    Wacker et al., 2002). The second type of system appropriates glycans that have

    been synthesized and transported to the periplasm by the lipopolysaccharide (LPS)

    machinery; an example of this type is the P. aeruginosa 1244 pilin glycosylation

    system (DiGiandomenico et al., 2002).

    B. Neisseria pilin glycosylation

    1. Site of Modification. The T4P of N. meningitidis are essential virulence

    factors for this obligate pathogen, mediating attachment to host epithelial cells and

    subsequent colonization (Power & Jennings, 2003; Virji et al., 1992). Studies of the

    pilin structural subunit, PilE, of strain C311 have revealed that these proteins are

    glycosylated at Ser63 with the trisaccharide Gal(β1,4)Gal(α1-3)2,4-diacetamido-

    2,4,6-trideoxyhexose (DATDH) (Power & Jennings, 2003; Stimson et al., 1995). The

    N. gonorrhoeae MS11 pilin is also glycosylated at this site with a similar glycan, a

    disaccharide of hexose linked to 2,4-diacetamido-2,4,6-trideoxyhexose (HexDATDH)

    (Figure 1-1) (Parge et al., 1995). Recently, intact pili from N400 were analysed by

    mass spectrometry, showing that this pilin glycan is also O-acetylated, revealing a

  • 33

    complex post-translational modification (Aas et al., 2007a). Significantly, based on

    crystallographic models, Ser63 is located on the αβ-loop region of the mature pilin

    subunit, which is thought to be solvent exposed on the assembled fibre (Craig et al.,

    2006; Parge et al., 1995).

    2. Mechanism. The pilin glycosylation systems of N. meningitidis and N.

    gonorrhoeae are representative of those dedicated to this process. Many of the

    genes that encode the proteins involved in the biosynthesis, assembly and transfer

    of the glycans have been identified (Aas et al., 2007a; Power & Jennings, 2003;

    Power et al., 2006). The genes identified to date are found in several locations on

    the chromosome, some on “pilin glycosylation loci” (pgl); however, none of the

    genes found were linked to pilE, which encodes the pilin subunit (Power & Jennings,

    2003).

    The molecular mechanism for pilin glycosylation has been well established for

    N. meningitidis and a model is presented in Figure 1-6 (Power et al., 2006). On the

    cytoplasmic face of the inner membrane, PglB is involved in the biosynthesis of

    DATDH and is thought to transfer this sugar to an undecaprenol carrier, although

    there is currently no evidence that this precursor exists (Aas et al., 2007a; Kahler et

    al., 2001; Power et al., 2000). The trisaccharide is then assembled by the addition of

    two galactose residues by the galactosyltransferases PglA and PglE (Banerjee et al.,

    2002; Jennings et al., 1998; Power & Jennings, 2003). Once fully assembled, the

    mature glycan is translocated to the periplasmic face of the membrane by the

    flippase PglF (Kahler et al., 2001; Power et al., 2000). PglL in N. meningitidis and

  • 34

    PglO in N. gonorrhoeae have been identified as the oligosaccharyltransferases that

    transfer the completed glycan to the PilE subunits prior to their incorporation into the

    pilus fibre (Aas et al., 2007a; Power et al., 2006).

    PglLNM and PlgONG were identified as the pilin oligosaccharyltransferases

    based on their similarities to the P. aeruginosa 1244 pilin oligosaccharyltransferase

    PilO, and the O-antigen ligase WaaL, which transfers preformed O-antigen polymers

    from a lipid-linked precursor to the separately synthesized core-lipid A portion of LPS

    (Aas et al., 2007a; Abeyrathne & Lam, 2007; Power et al., 2006). PglLNM is

    predicted to have 13 transmembrane spanning domains, with a short cytoplasmic N-

    terminal segment and a longer periplasmic C-terminal domain. A Wzy_C (PFAM

    PF04932) domain, present in PilO1244 and WaaL, is predicted in the large

    periplasmic loop between transmembrane segments 9 and 10 and is thought to be

    the site at which ligation of the O-antigen to lipid A-core takes place (Mulford &

    Osborn, 1983; Power et al., 2006). Mutants that lacked PglL produced non-

    glycosylated T4P, demonstrating this protein has an essential role in pilin

    glycosylation (Power et al., 2006).

    3. Role of Glycosylation in T4P function. The role of glycosylation of

    Neisseria pili is still not well understood. Due to the location of Ser63 on the αβ-loop

    ridge of the pilin, the glycans are predicted to be exposed to the environment and

  • 35

    Figure. 1-6. Structure and molecular mechanism of pilin glycosylation. A. Structure of the N.

    meningitidis pilin glycan and enzymes responsible for its biosynthesis. B. Model of ligase-

    dependent pilin glycosylation. Beginning on the left-hand side the initial UDP-linked sugar is

    synthesized and transferred to an undecaprenol carrier via PglB. Additional sugars are added via

    the glycosyltransferases PglA and PglE. The completed undecaprenol-linked trisaccharide is

    translocated across the inner membrane via the ‘flippase’ PglF to the periplasmic face where the

    oligosaccaryltransferase PglL transfers it to the pilin subunit. Reproduced from Power et al. (2006),

    with the permission of Biochemical and Biophysical Research Communications, Elsevier Limited.

    A

    BGal(β1-4)Gal(α1-3) 2,4-diacetimido-2,4,6-trideoxyhexose-Ser

    Periplasm

    Cytoplasm

    Membrane

    Figure. 1-6. Structure and molecular mechanism of pilin glycosylation. A. Structure of the N.

    meningitidis pilin glycan and enzymes responsible for its biosynthesis. B. Model of ligase-

    dependent pilin glycosylation. Beginning on the left-hand side the initial UDP-linked sugar is

    synthesized and transferred to an undecaprenol carrier via PglB. Additional sugars are added via

    the glycosyltransferases PglA and PglE. The completed undecaprenol-linked trisaccharide is

    translocated across the inner membrane via the ‘flippase’ PglF to the periplasmic face where the

    oligosaccaryltransferase PglL transfers it to the pilin subunit. Reproduced from Power et al. (2006),

    with the permission of Biochemical and Biophysical Research Communications, Elsevier Limited.

    Figure. 1-6. Structure and molecular mechanism of pilin glycosylation. A. Structure of the N.

    meningitidis pilin glycan and enzymes responsible for its biosynthesis. B. Model of ligase-

    dependent pilin glycosylation. Beginning on the left-hand side the initial UDP-linked sugar is

    synthesized and transferred to an undecaprenol carrier via PglB. Additional sugars are added via

    the glycosyltransferases PglA and PglE. The completed undecaprenol-linked trisaccharide is

    translocated across the inner membrane via the ‘flippase’ PglF to the periplasmic face where the

    oligosaccaryltransferase PglL transfers it to the pilin subunit. Reproduced from Power et al. (2006),

    with the permission of Biochemical and Biophysical Research Communications, Elsevier Limited.

    A

    BGal(β1-4)Gal(α1-3) 2,4-diacetimido-2,4,6-trideoxyhexose-Ser

    Periplasm

    Cytoplasm

    Membrane

    therefore able to influence the pili’s surface chemistry and interaction with its

    surroundings (Craig et al., 2006; Parge et al., 1995). Experiments in which Ser63 of

    PilENG was altered to Ala revealed that non-glycosylated pili are more bundled than

  • 36

    the glycosylated pili, suggesting that the surface chemistry of the fibres is altered,

    influencing the fibres’ interactions with one another (Craig et al., 2006; Marceau et

    al., 1998; Stimson et al., 1995). The modification at Ser63 was found to be essential

    for truncated soluble pilin in N. meningitidis, but not N. gonorrhoeae, but did not

    appear to influence adhesion as there was no significant difference in binding to

    cells between strains expressing glycosylated or non-glycosylated T4P (Marceau &

    Nassif, 1999).

    Many of the pgl genes have been found to contain sequence repeats that

    mediate phase variation, resulting in production of different versions of the glycans,

    and leading to further diversity amongst these bacteria (Kahler et al., 2001; Power et

    al., 2003). Neisseria are capable of further modifying the pilins with

    phosphoethanolamine and/or phosphocholine at Ser68, and these modifications can

    also undergo phase variation (Craig et al., 2006; Hegge et al., 2004; Weiser et al.,

    1998). It is possible that the main function of T4P glycosylation in Neisseria is host

    immune evasion. The diversity of post-translational modifications observed in these

    systems is likely influenced by the fact that N. meningitidis and N. gonorrhoeae are

    obligate human pathogens which are subject to strong pressures from the host

    immune system, and must escape detection in order to survive.

    C. P. aeruginosa pilin glycosylation: strain 1244