169
Study of the physiological and molecular mechanisms underlying peptide-induced cell death and biofilm formation in Streptococcus mutans JULIE ANN PERRY A thesis submitted in conformity with the requirements for the Degree of Doctor of Philosophy Graduate Department of Dentistry University of Toronto Copyright by Julie A. Perry 2009 ©

Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

Study of the physiological and molecular mechanisms underlying

peptide-induced cell death and biofilm formation in

Streptococcus mutans

JULIE ANN PERRY

A thesis submitted in conformity with the requirements for the

Degree of Doctor of Philosophy

Graduate Department of Dentistry

University of Toronto

Copyright by Julie A. Perry 2009 ©

Page 2: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

ii

Study of the physiological and molecular mechanisms underlying peptide-

induced cell death and biofilm formation in Streptococcus mutans

Julie Ann Perry

Doctor of Philosophy, Faculty of Dentistry, University of Toronto

2009

Abstract

Biofilms are complex and highly adapted communities of microorganisms found attached to

surfaces. Among the best characterized infectious multi-cellular biofilms is the oral community

known as dental plaque. Streptococcus mutans resides in the oral biofilm, and is one of the

main causative agents of dental caries. Streptococci are known to monitor their population

density using a peptide pheromone (CSP)/two component signalling system (ComDE) in a

process classically known as quorum sensing (QS). Previous work in S. mutans has implicated

the QS system in genetic competence, the stress response, bacteriocin production and biofilm

formation. Our objective in this work was to thoroughly characterize the transcriptional and

phenotypic response to CSP in S. mutans, and determine its role in biofilm formation. We have

shown that the CSP pheromone is more than simply a QS signal, and is also an inducible

„alarmone‟ capable of communicating stress in the population. We have demonstrated that

elevated concentrations of CSP such as those that occur during stress trigger autolysis in a

fraction of the population. Importantly, we have shown that autolysis in S. mutans occurs via a

novel mechanism of action: intracellular accumulation of a self-acting bacteriocin. We have also

identified and characterized the autolysis immunity protein, which is differentially regulated from

the bacteriocin to allow survival at low cell density. A second regulatory system was shown to

govern expression of autolysis immunity in the absence of CSP signaling, and also contribute to

the oxidative stress response in the biofilm. Finally, we present evidence that autolysis is

involved in the release of DNA in the biofilm, which contributes to the architecture of the

extracellular matrix and may provide a mechanism for the dissemination of fitness-enhancing

genes under stress. Together, our data provides a mechanistic link between phenotypes

previously ascribed to the CSP pheromone in S. mutans.

Page 3: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

iii

Acknowledgements

I extend sincere thanks to both my supervisors, Dr. Celine Levesque and Dr. Dennis Cvitkovitch

for their patience, mentorship and support over the years. I am particularly grateful to Dr.

Levesque, who took me on as her first student, and trusted me with this project. I truly

appreciate the opportunities, challenges, advice and friendship you have given me. I would like

to thank our collaborators Dr. Marcus Jones and Dr. Scott Peterson at the J. Craig Venter

Institute for their help performing the DNA microarrays presented in Chapters two and four. I

am also grateful to Dr. Roslyn Devlin, who gave me my first taste of science and set me on this

path. I sincerely thank Dr. Stanley Holt for having such confidence in me, and for always

treating me like a fellow scientist. Finally, I am grateful to the members of my advisory

committee, Dr. Debora Foster, Dr. Chris McCulloch and Dr. Martin McGavin for their invaluable

critiques and expert guidance.

Thank you also to the members of the Levesque and Cvitkovitch labs, particularly Kirsten,

Timmy, Elena and Marie-Christine. Your friendship and support has been much appreciated.

Beyond the lab, I would like to thank my friends for putting up with my moments of „intensity‟,

and always listening when I needed it.

Finally, I thank my family for their unwavering support. To Mom, Dad, Ted, and of course to

Patrick, I am unendingly grateful for the opportunities your love has provided me, and for the

peace you bring to my life. Thank you.

Julie Ann Perry

2009

Page 4: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

iv

Table of Contents

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

Acknowledgements .................................................................................................................... iii

Table of Contents ....................................................................................................................... iv

List of Tables ........................................................................................................................... viii

List of Figures ............................................................................................................................ ix

Preface ...................................................................................................................................... xi

Publications reproduced as dissertation chapters ..................................................................... xii

Awards ..................................................................................................................................... xiii

Abbreviations ........................................................................................................................... xiv

Quote ....................................................................................................................................... xvi

Chapter 1: Literature Review ............................................................................................. 1

1.1. The biofilm mode of growth ................................................................................................. 2

1.1.1. The lifecycle of a biofilm ........................................................................................... 2

1.1.2. Structure of a mature biofilm..................................................................................... 4

1.1.3. Bacterial biofilm infections ........................................................................................ 6

1.2. Streptococcus mutans: a model for infectious biofilm formation ................................... 8

1.2.1. S. mutans in the oral biofilm ..................................................................................... 9

1.2.1.1. Attachment to the tooth: formation of the oral biofilm ......................................... 9

1.2.1.2. Acidogenicity and aciduricity ............................................................................13

1.2.1.3 Interspecies competition in the oral biofilm: production of mutacins ..................14

1.2.1.4. Two-component signal transduction systems ...................................................16

1.3. Quorum sensing.................................................................................................................19

1.3.1. A brief history of quorum sensing ............................................................................20

1.3.2. Quorum sensing mechanisms .................................................................................22

1.3.3. Quorum Sensing in streptococci ..............................................................................25

1.3.3.1. Overview of the CSP-ComDE system in S. pneumoniae ..................................25

1.3.3.2 CSP-ComDE in S. mutans ............................................................................28

Page 5: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

v

1.3.4 Phenotypes controlled by CSP-ComDE signalling in streptococci ........................31

1.3.4.1 Genetic competence ..........................................................................................31

1.3.4.2.1 Mechanism of DNA uptake in streptococci .................................................31

1.3.4.2.2 Purpose of DNA uptake .............................................................................33

1.3.4.2 Autolysis and pneumococcal fratricide ..........................................................35

1.3.4.2.1 Autolysins and autolysis ............................................................................36

1.3.4.2.2 The holin/antiholin system in S. aureus......................................................37

1.3.4.2.3 Streptococcal fratricide ..............................................................................38

1.3.4.3 Bacteriocin production ..................................................................................43

1.3.4.3.1 Classification of bacteriocins ......................................................................43

1.3.4.3.2 Biosynthesis and export of class IIa bacteriocins .......................................44

1.3.4.3.3 Mode of action ...........................................................................................46

1.3.4.4 CSP and the stress response .......................................................................48

1.3.4.5. Biofilm formation ...........................................................................................49

1.4. Statement of the problem ...................................................................................................51

General hypothesis: ...............................................................................................................52

Primary objective: ..................................................................................................................52

Rationale: ..............................................................................................................................52

1.5 References ........................................................................................................................54

Chapter 2: Peptide alarmone signalling triggers an auto-active bacteriocin

necessary for genetic competence ..................................................................................64

2.1 Abstract .............................................................................................................................65

2.2 Introduction ........................................................................................................................66

2.3 Experimental procedures ...................................................................................................68

Culture conditions ..................................................................................................................68

Gene expression analysis......................................................................................................70

DNA microarrays ...................................................................................................................70

Recombinant peptides ...........................................................................................................71

Page 6: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

vi

Bacteriocin overlay assays ....................................................................................................72

Transformation assays ..........................................................................................................72

2.4 Results ...............................................................................................................................72

2.4.1 Stress induces expression of the CSP pheromone ..................................................72

2.4.2 CSP pheromone triggers autolysis in a fraction of the population ............................75

2.4.3 Genome-wide expression response to CSP: identification of mutacin V ..................78

2.4.4 CipB bacteriocin likely acts intracellularly ................................................................84

2.4.5 The small protein CipI (SMU.925) confers immunity ................................................89

2.4.6 Role of CSP-induced lysis in genetic competence ...................................................91

2.5 Discussion .........................................................................................................................96

2.6 Acknowledgements ............................................................................................................99

2.7 References ...................................................................................................................... 100

Chapter 3: Cell Death in Streptococcus mutans Biofilms: a Link Between CSP and

Extracellular DNA ............................................................................................................... 104

3.1 Abstract ........................................................................................................................... 105

3.2 Introduction ...................................................................................................................... 106

3.3 Materials and methods ..................................................................................................... 107

3.4 Results and discussion .................................................................................................... 110

3.4.1 Low cell density-dependent expression of cipI is regulated by LiaR....................... 110

3.4.2 CipI is protective at low cell density, while CipB is lethal at high cell density .......... 111

3.4.3 Cell death participates in biofilm formation through eDNA release......................... 114

3.5 Conclusions ..................................................................................................................... 118

3.6 Acknowledgements .......................................................................................................... 119

3.7 References ...................................................................................................................... 120

Chapter 4: Involvement of Streptococcus mutans regulator RR11 in oxidative

stress response during biofilm growth and in the development of genetic

competence ......................................................................................................................... 123

4.1 Abstract ........................................................................................................................... 124

Page 7: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

vii

4.2 Introduction ...................................................................................................................... 125

4.3 Materials and methods ..................................................................................................... 126

Bacterial strains and growth conditions................................................................................ 126

In vitro model for growing biofilms ....................................................................................... 127

Scanning electron microscopy ............................................................................................. 127

DNA microarrays ................................................................................................................. 127

Transformation experiments ................................................................................................ 128

4.4 Results ............................................................................................................................. 128

4.4.1 Phenotypic characterization of Δrr11 defective mutant .......................................... 128

4.4.2 Microarray identification of RR11-regulated genes involved in the stress response.

129

4.4.3 SMRR11 biofilms under oxidative, osmotic and acid stresses ............................... 130

4.4.4 Regulatory role for RR11 in competence development .......................................... 132

4.5 Discussion ....................................................................................................................... 133

4.6 Acknowledgements .......................................................................................................... 136

4.7 References ...................................................................................................................... 137

Chapter 5: Summary and Conclusions .......................................................................... 138

5.1 Summary of Dissertation .................................................................................................. 139

5.2 General Discussion .......................................................................................................... 140

5.2.1 Peptide pheromone-induced cell death ................................................................. 140

5.2.2 Immunity to peptide induced cell death .................................................................. 141

5.2.3 Peptide-induced cell death in genetic competence ................................................ 142

5.3 Future Directions .............................................................................................................. 143

5.4 Significance ..................................................................................................................... 144

5.5 References ...................................................................................................................... 145

Appendix A: Supplementary Information ...................................................................... 146

Page 8: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

viii

List of Tables

Table 1.1: Common biofilm-mediated infections in humans........................................................ 7

Table 1.2: S. mutans two-component signal transduction systems ............................................18

Table 1.3: Comparison of the CSP-ComDE systems in S. pneumoniae and S. mutans ............29

Table 1.4: Autolysin and fratricidal effector genes and their homologs in S. mutans UA159 ......42

Table 1.5: Bacteriocins encoded by S. mutans..........................................................................45

Table 2.1: Bacterial strains used in this study ...........................................................................69

Table 2.2: Relative expression levels of highly-expressed CSP-induced S. mutans genes

encoding putative and known bacteriocins and their accessory genes ..................................79

Table 3.1: Bacterial strains used in this study ......................................................................... 106

Table 4.1: Bacterial strains used in this study ......................................................................... 125

Table 4.2: Genes potentially regulated by RR11 in S. mutans growing in biofilms ................... 130

Table S1: Genes showing a minimum ± 2-fold difference in expression when S. mutans UA159

cells were exposed to 2 μM sCSP ………………………………………………………………. 147

Table S2: S. mutans genes showing a minimum ± 2-fold difference in expression when

S. mutans ∆comX cells were exposed to 2 μM sCSP ………………………………………… 153

Page 9: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

ix

List of Figures

Figure 1.1: Summary of factors influencing the survival and virulence of S. mutans in the oral

biofilm. ......................................................................................................................................10

Figure 1.2: Formation of dental plaque ......................................................................................12

Figure 1.3: Simplified schematic representation of quorum sensing systems in bacteria ...........24

Figure 1.4: Mechanistic and genomic representation of the CSP-ComDE circuit controlling

competence development in S. pneumoniae and S. mutans. ..................................................30

Figure 1.5: Representation of the holin/anti-holin system in S. aureus ......................................39

Figure 1.6: Proposed mechanism of action of the CSP-induced fratricidal pathway in S.

pneumoniae ..............................................................................................................................41

Figure 1.7: Cartoon representation of the mechanism of action of bacteriocins .........................47

Figure 2.1. Recovery of S. mutans from stress. .........................................................................74

Figure 2.2. S. mutans growth kinetics .......................................................................................76

Figure 2.3. Effect of sCSP on culture density and cell lysis .......................................................77

Figure 2.4: A sub-population of cells is always resistant to sCSP-mediated cell lysis. ...............78

Figure 2.5. Growth of Streptococcus thermophilus and Streptococcus salivarius in increasing

concentrations of their species-specific signaling peptides ........................................................81

Figure 2.6. Effect of 2 µM sCSP pheromone on S. mutans wild-type and mutants defective in

the bacteriocin CipB (SMU.1914) and its putative immunity factors SMU.1913 and CipI

(SMU.925). ...............................................................................................................................82

Figure 2.7. Agar overlay assays showing extracellular bacteriocin ............................................83

Figure 2.8. Growth kinetics of S. mutans UA159 strain containing CipB under the control of a

raffinose-inducible promoter .....................................................................................................85

Figure 2.9. RT-PCR gene expression profiles of cipB and cipI ..................................................86

Figure 2.10. Quantitative real-time RT-PCR gene expression profiles of cipB, comC and cipI at

following dilution from an overnight culture ................................................................................86

Figure 2.11. CipB may act intracellularly ...................................................................................88

Figure 2.12. Growth of the wild-type and a wild-type strain over-expressing cipI .......................91

Figure 2.13. Transformation efficiency of S. mutans wild-type strain and its mutants deficient in

the CipB bacteriocin and CipI immunity protein .........................................................................93

Page 10: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

x

Figure 2.14. Competence and lysis in cultures of S. mutans .....................................................94

Figure 2.15. Summary of data ...................................................................................................95

Figure 3.1: Growth of S. mutans TCS mutants in the presence of 2 µM sCSP ........................ 111

Figure 3.2: Growth kinetics of S. mutans UA159 wild-type strain in the presence of 2µM sCSP

............................................................................................................................................... 113

Figure 3.3: Fold-change in quantity of eDNA in ΔCipI and ΔCipB mutant biofilms ................... 115

Figure 3.4: Biofilm biomass of S. mutans UA159 wild-type strain, ΔCipB and ΔCipI mutants .. 117

Figure 4.1. S. mutans UA159 and RR11- mutant biofilm formation ......................................... 129

Figure 4.2. Transformation efficiency of S. mutans wild-type and mutant strains. .................... 133

Figure 4.3 Model for competence development in S. mutans ................................................. 136

Page 11: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xi

Preface

Dissertation format

This dissertation is presented in the „Publishable Style‟. Chapter 1 presents a general

introduction to the subject, and serves to provide context for the following chapters. It includes

portions of an invited book chapter appearing in the ASM Press publication „Genomic Inquiries

Into Oral Bacterial Communities‟ (P.E. Kolenbrander, Ed.). Chapters 2 to 4 describe

experimental data that have either been published or submitted for publication. They are

presented in their published form, other than minor changes made to improve readability and

reduce repetition. Chapter 5 serves a brief discussion of all experimental data. Written

permission for reproduction of all publications has been obtained, and is held on file.

Page 12: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xii

Publications reproduced as dissertation chapters

Perry JA, Cvikovitch DG. 2009. Autoinducer-2-regulated genes in Streptococcus mutans and

impact on oral bacterial communities. Invited book chapter in Genomic Inquiries Into Oral

Bacterial Communities, Paul E. Kolenbrander (Ed.). ASM Press.

Perry JA, Jones MB, Peterson SN, Cvitkovitch DG, Lévesque CM. 2009. Peptide alarmone

signaling triggers an auto-active bacteriocin necessary for genetic competence. Mol Microbiol.

72: 905-917.

Perry JA, Cvitkovitch DG, Levesque CM. 2009. Cell death in Streptococcus mutansbiofilms: a

link between CSP and extracellular DNA. FEMS Microbiol. Lett. 299:261-6

Perry JA, Lévesque CM, Suntharaligam P, Mair RW, Bu M, Cline RT, Peterson SN, Cvitkovitch

DG. 2008. Involvement of Streptococcus mutans Regulator RR11 in Oxidative Stress Response

in the Biofilm and in the Development of Genetic Competence. Lett Appl Microbiol. 47: 439-44.

Additional publications

Kreth J, Merritt J, Huang D, Perry J, Zhu L, Goodman S, Cvitkovitch DG, Shi W, Qi F. 2007.

The response regulator ComE in Streptococcus mutans functions both as a transcription

activator of mutacin production and repressor of CSP biosynthesis. Microbiology. 153: 1799-

807.

Lévesque CM, Mair RW, Perry JA, Lau P, Li Y-H, Cvitkovitch DG. 2007. Systemic inactivation

and phenotypic characterization of two-component systems in expression of Streptococcus

mutans virulence properties. Lett Appl Microbiol. 45: 398-404.

Perry J, Ho M, Viero S, Zheng K, Jacobs R, Thorner P. 2007. The intermediate filament nestin

is highly expressed in normal human podocytes and podocytes in glomerular disease. Pediatr

Dev Pathol.10: 369-82.

Harvey S, Perry J, Zheng K, Chen D, Sado Y, Jefferson B, Ninomiya Y, Jacobs R, Hudson B,

Thorner, P. 2006. Sequential expression of type IV collagen networks: testis as a model and

relevance to spermatogenesis. Am J Pathol. 168: 1587-97

Perry J, Tam S, Zheng K, Harvey S, Sado Y, Jefferson B, Jacobs R, Thorner P. 2006. Type IV

collagen induces podocytic features in bone marrow stromal stem cells in vitro. J Am Soc

Nephrol. 17: 66-76.

Zheng K, Perry J, Harvey S, Sado Y, Ninomiya Y, Jefferson B, Jacobs R, Hudson B, Thorner P.

2005. Regulation of collagen type IV genes is organ-specific: evidence from a canine model of

Alport syndrome.Kidney Int 68: 2121-2130.

Page 13: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xiii

Awards

2007 2nd place poster award, Faculty of Dentistry Research Day

2007-2009 Ontario Graduate Scholarship

2006-2009 CIHR Strategic Training Program (“Cell Signals”) Fellowship

2005-2009 Harron Scholarship, Faculty of Dentistry, University of Toronto

2005 1st place poster award, Hospital for Sick Children Research Day

2003-2005 University of Toronto Open Fellowship

2002-2003 Ontario Thoracic Society Summer Studentship

1999 Univeristy of Guelph Entrance Scholarship

Page 14: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xiv

Abbreviations

aa amino acid

ABC ATP-binding casette

AHL acylhomoserine lactone

AI-2 auto-inducer 2

Ala alanine

ATP adenosine triphosphate

ATR acid tolerance response

BIP bacteriocin-inducing peptide

CSP competence stimulating peptide

DPD (S)-4,5-dihydroxy-2,3-pentanedione

ds double-stranded

eDNA extracellular DNA

EDTA ethylenediaminetetraacetic acid

EPS exopolymeric substances

FTF fructosyltransferase

GTF glucosytransferase

GUS β-glucuronidase

h hour

His histidine

HK histidine kinase

IPTG isopropyl-beta-D-thiogalactopyranoside

LB Luria-Bertani

min minute

MU Miller units

Page 15: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xv

PBS phosphate-buffered saline

PCR polymerase chain reaction

PMF proton motive force

PNPG p-nitrophenyl glucuronide

RR response regulator

s second

ss single-stranded

SAH S-adenosylhomocysteine

SAM S-adenosylmethionine

sCSP synthetic CSP

SDM semi-defined minimal medium

SEM scanning electron microscopy

TCS two-component signal transduction system

THYE Todd-Hewitt-Yeast extract

WT wild-type

Page 16: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

xvi

Quote

“It is not the strongest of the species that survives, nor the most intelligent. It is the one

that is the most adaptable to change.”

– Charles Darwin

Page 17: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

1

Chapter 1: Literature Review

Page 18: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

2

1.1. The biofilm mode of growth

Bacteria have long been studied as single-celled, primitive organisms, free-floating in

laboratory culture. However, bacterial species in nature have a strong tendency to colonize

surfaces and form complex, multi-species communities referred to as biofilms (Costerton et al.,

1994). This behaviour has been likened to the human condition as described by the

philosopher John Locke (1632-1704), who argued that the default state of human existence is

independent and solitary, but that individuals may voluntarily form communities for mutual

protection and economic prosperity. In nature, biofilms are found on rocks in streams, in

industrial bioreactors, and in animal host environments like the oropharyngeal, gastrointestinal

and vaginal tracts, and on medical prostheses (Costerton et al., 1994; Costerton et al., 1999).

The ubiquitous presence of biofilms in such diverse environments suggests a strong

evolutionary advantage for surface dwellers over their free-floating counterparts (Dunne, 2002).

The obvious explanation of the sessile lifestyle is that nutrients in an aqueous environment tend

to concentrate at solid surfaces. However, surface-associated community living offers many

more advantages to biofilm dwellers, including protection from antimicrobial agents and host

defences and division of the metabolic burden between neighbours.

1.1.1. The lifecycle of a biofilm

A biofilm in its simplest form is composed of a surface (or „substratum‟), surface attached

cells, and a surrounding extracellular matrix of biopolymers (Dunne, 2002). The lifecycle of a

biofilm begins with the attachment of free-floating (or „planktonic‟) cells to a living or an abiotic

substratum, followed by the formation of discreet microcolonies, the subsequent maturation of

Page 19: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

3

the biofilm including the development of a complex three dimensional architecture, and finally

production of a planktonic subpopulation for dispersal and re-colonization.

Primary attachment of a bacterium to a substratum is often considered to be a two step

process: an initial, reversible, non-specific interaction (e.g hydrophobic interaction), is followed

by a secondary, permanent interaction mediated by specific adhesins (Dunne, 2002). During

the initial attachment phase, the adherent cells are not considered „committed‟ to biofilm

differentiation, and may leave the surface to resume their planktonic lifestyle (reviewed in

(Stoodley et al., 2002)). In the well-characterized Gram-negative biofilm-forming organism

Pseudomonas aeruginosa, the reversible attachment phase ends when cells begin to up-

regulate the gene responsible for production of the exopolysaccharide alginate, within 15

minutes of initial contact with the substratum (Davies et al., 1993). Indeed, production of

exopolymeric substances (EPS) is considered to be the hallmark of the transition to biofilm

growth. Once biofilm formation is initiated by the primary colonizers, other bacteria may adhere

to the surface-attached primary colonizers through specific receptor-mediated interactions (for a

review of this process during the formation of the dental plaque biofilm, see section 1.2). The

receptor-mediated nature of these interactions results in an organized succession of bacterial

attachment in a multi-species biofilm. Often, metabolically complementary species are located

in spatial proximity in a multi-species biofilm, to benefit from metabolic substrate exchange and

waste removal within the microcolony. These types of reciprocal relationships are only possible

during dense, surface-associated growth. Once attached, cells begin to divide (and to die),

forming dense microcolonies within the larger biofilm. The lifecycle of a biofilm is complete once

the biofilm reaches a critical mass: at this point, the outermost layers of growth begin to

generate planktonic cells which depart the biofilm and colonize other surfaces (Dunne, 2002).

Page 20: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

4

Recent evidence has suggested that the stages in the maturation of a biofilm are controlled by

cell-density monitoring and cell-cell communication systems in a process known as quorum

sensing (reviewed in (Davies et al., 1998; Parsek and Greenberg, 2005) and discussed in

section 1.3), and that each stage in the biofilm lifecycle is distinct from the others (Stoodley et

al., 2002). At maturity, biofilms formed by P. aeruginosa show changes in as much as 50% of

the detectable proteome compared with their planktonic counterparts (Sauer et al., 2002).

1.1.2. Structure of a mature biofilm

While the biofilm lifestyle provides the opportunity for nutrient and waste-sharing, the obvious

disadvantage to permanent surface attachment is the inevitable over-crowding and nutrient and

oxygen shortages within each micro-colony. The three-dimensional structure of a biofilm is

therefore essential to its function. By microscopic analysis, a biofilm appears to be a highly

hydrated and open structure, composed mainly of non-cellular material including water channels

and EPS (Lawrence et al., 1991). The EPS is composed of a hydrated, anionic mesh of

bacterial exopolymers and trapped environmental molecules, and forms the outermost layer of

the biofilm (Branda et al., 2005). Despite the universal presence of EPS in all biofilms, there is

variation in the composition and timing of EPS production in biofilms from different species

(Branda et al., 2005). The most extensively studied components of the EPS are carbohydrate-

rich polymers (like alginate) and proteins, but extracellular DNA (eDNA) is now also widely

recognized as a major constituent of the matrix (Allesen-Holm et al., 2006; Hall-Stoodley et al.,

2008; Thomas et al., 2009; Whitchurch et al., 2002). Whatever its composition, the extracellular

matrix functions as a permeability barrier to limit both the diffusion of beneficial nutrients away

from the biofilm, and prevent or slow the diffusion of harmful substances like antibiotics and

Page 21: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

5

predatory cells of the immune system from accessing matrix-embedded cells (Costerton et al.,

1999).

Within the extracellular matrix, individual cells occupy a distinct niche environment that is

connected to the rest of the biofilm and to the external environment by a network of water-filled

channels, which behave like a primitive circulatory system (Costerton et al., 1994). These

channels have been shown to permit the penetration of large molecules (up to 2,000-kDa) into

the environment, and to transport dissolved oxygen into the biofilm by convective flow

(Costerton et al., 1994). However, nutrients and dissolved oxygen can only penetrate into the

outer-most cells of each microcolony, and the growth of the biofilm is limited by the availability of

nutrients to deeply embedded cells. The result of this limited diffusion is that the centre of each

microcolony is essentially anaerobic, and may experience nutrient deficiencies and waste build-

up. Likewise, oxygen decreases through the depth of the biofilm, such that the surface is more

aerobic while cells located close to the colonized surface are anaerobic (Costerton et al., 1994).

The spatial separation of sessile cells combined with nutrient/waste and oxygen gradients

within the biofilm results in a heterogeneous population of cells, distinct from their planktonic

counterparts in gene expression patterns and behaviours. The metabolic task sharing,

communication and heterogeneity within a biofilm lead to the success of the group where

individual bacteria would have failed (Parsek and Greenberg, 2005). In essence, the combined

metabolic repertoire of multiple bacterial genomes acting in concert is able to accomplish the

function of a multi-cellular organism. However, the advantage to sharing metabolic

responsibilities among many small prokaryotic genomes instead of one large eukaryotic

genome is the phenotypic plasticity possible in a prokaryotic multi-species biofilm (Stoodley et

al., 2002). The result is that biofilms may take advantage of periods of nutrient bounty (during

which they may even release fast-growing planktonic cells), but may also survive periods of

Page 22: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

6

nutrient shortage through the increased adaptation potential present in multiple genomes

located in close proximity (Stoodley et al., 2002). This ability to adapt and persist in the face of

adversity has led to the ubiquitous nature of biofilms in the environment and in infectious

disease.

1.1.3. Bacterial biofilm infections

What may be a highly beneficial lifestyle for bacteria has become highly problematic for their

human hosts. Although acute bacterial infections were historically caused by planktonic

populations of specialized pathogens like Vibrio cholerae and Yersinia pestis, more than 80% of

modern-day infections in the developed world are thought to involve biofilms (Costerton et al.,

1999; Fux et al., 2005) (Table 1.1). Infections caused by biofilms are often slow-growing, slow

to produce overt symptoms, and are rarely resolved by either the host immune system or

conventional antibiotic therapy. These characteristics are common to all biofilm-mediated

infections, no matter the location or the causative agent, and thought to result from the matrix-

enclosed and sequestered nature of the biofilm lifestyle. Antibodies and cells of the immune

system cannot access biofilm-embedded cells due to the extracellular matrix, often resulting in

immune complex damage to surrounding tissues (Stewart and Costerton, 2001). The increased

antibiotic tolerance of biofilms is not due to traditional mechanisms (efflux, modifying enzymes

or target mutations), but is instead due to the combined effects of diffusion limitation by the

extracellular matrix, the slow growth rate of matrix-enclosed cells and the formation of dormant

and highly resistant „persister‟ cells in the biofilm (for excellent reviews see (Fux et al., 2005;

Lewis, 2005; Stewart and Costerton, 2001)). As a result, antibiotic therapy often results in

short-term relief of symptoms through the elimination of planktonic cells released from the

biofilm, but fails to kill the surface-associated infectious centre. Biofilms are therefore thought to

Page 23: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

7

act as reservoirs for recurrent infections that persist until the colonised surface is removed from

the body by physical means (Stewart and Costerton, 2001). Due to their prevalence and

persistence, understanding the processes leading to the formation and maturation of a biofilm is

critical to both our understanding of ecology and of pathogenesis.

Table 1.1: Common bacterial biofilm-mediated infections in humans (adapted from (Costerton

et al., 1999).

Site Infection or disease Bacterial species involved

Oral cavity Dental caries

Acid-producing Gram-positive cocci (e.g. Streptococcus)

Periodontitis Gram-negative anaerobes

Ear, nose, throat Otitis media

Non-typeable Haemophilus influenzae

Chronic tonsillitis Various species

Lung Cystic fibrosis pneumoniae Pseudomonas aeruginosa, Burkholderia cepacia

Heart Endocarditis Viridans group streptococci, staphylococci

Bone Musculosketetal infections Gram-positive cocci

Osteomyelitis Various species

Genitourinary tract Bacterial prostatitis

Escherichia coli and other Gram-negative rods

Infectious kidney stones Gram-negative rods

Medical device-related

Contact lens P. aeruginosa, Gram-positive cocci

Ventilation-associated pneumoniae

Gram-negative rods

Mechanical heart valves Staphylococci

Urinary catheter infections E. coli, Gram-negative rods

Orthopaedic prosthesis Staphylococci

Page 24: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

8

1.2. Streptococcus mutans: a model for infectious biofilm formation

The genus Streptococcus comprises Gram positive, non-spore forming, non-motile cocci,

which characteristically grow as pairs or chains. Streptococci are carried in the nasopharynx of

man as members of the commensal microflora. However, streptococci may also cause disease

ranging in severity and prevalence, including pharyngitis, necrotizing faciitis, meningitis, and the

most common infectious diseases affecting humans: dental caries (Mitchell, 2003).

First isolated in 1924, Streptococcus mutans is the causative agent of one of the most benign

(although the most prevalent) streptococcal disease: human dental caries. In rare cases, S.

mutans has also been shown to cause infective endocarditis (Nomura et al., 2006). In both

cases, S. mutans colonizes the host as a biofilm. The transmission of S. mutans is thought to

occur from mother to child via salivary transfer, although there is typically a high degree of

homology between strains recovered from members of the same family, indicating both vertical

and horizontal routes of transmission (Napimoga et al., 2005). Studies of initial colonization by

S. mutans have shown that these bacteria require erupted tooth surfaces to become established

in the oral cavity, but are among the first colonizers of those surfaces. Interestingly, the same

initially colonizing strain may persist until young adulthood (Napimoga et al., 2005). S. mutans

has several virulence factors that contribute to its pathogenicity, foremost of which is its ability to

adhere to the tooth surface as part of the multi-species oral biofilm community.

Page 25: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

9

1.2.1. S. mutans in the oral biofilm

The oral biofilm comprises more than 700 other species (approximately 20% of which are

streptococci) (Aas et al., 2005). It is now generally recognized that the etiology of dental caries

is influenced by the metabolism of the entire biofilm, and that disease results from the combined

interactions of multiple acid-producing organisms in the oral biofilm (Kuramitsu et al., 2007).

However, S. mutans is the only organism that has been conclusively linked to the causation of

dental caries (Aas et al., 2008; Loesche, 1986). What characteristics of S. mutans separate it

from the other residents of dental plaque as particularly cariogenic? S. mutans has the ability to

attach tenaciously to the tooth, to both produce and withstand highly acidic conditions, to out-

compete its neighbors through the production of antimicrobial peptides called bacteriocins, and

to finely tune all these virulence factors through the coordinated efforts of fourteen two-

component signal transduction systems that monitor the environment and population density.

The factors influencing the survival and virulence of S. mutans in the oral biofilm are discussed

in detail below, and summarized in Figure 1.1.

1.2.1.1. Attachment to the tooth: formation of the oral biofilm

The formation of the oral biofilm known as dental plaque is a well characterized process,

involving a defined sequence of adhesion and maturation events. In the oral cavity, the tooth

surface is bathed in saliva, which rapidly coats the tooth in salivary proteins, glycoproteins and

polysaccharides like mucins, salivary agglutinins, amylases and statherin (reviewed in (Rickard,

2008a)). These polymers, in addition to bacterially-derived proteins, form the so-called

„acquired„ or „salivary-pellicle‟, which is bound by „early colonizing‟ bacteria such as oxygen-

tolerant Gram-positive bacteria like streptococci (Kolenbrander et al., 2006). Over time, an

Page 26: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

10

Figure 1.1: Summary of factors influencing the survival and virulence of S. mutans in the oral

biofilm. S. mutans is able to adhere robustly to the tooth surface via the production of glucans

and glucan binding proteins in the presence of sucrose. In addition to sucrose, the organism

can metabolize a wide variety of sugars, which result in the production of acid that desolves the

enamel of the tooth surface and causes dental caries. The ability of S. mutans to tolerate that

acidic environment allows it to persist in the oral biofilm. Additonal environmental conditions are

sensed using two-component signal transduction systems (TCSs). S. mutans is also able to

monitor its population density using the quorum sensing system composed of the CSP peptide

pheromone and the ComDE TCS. Phenotypes regulated by CSP-ComDE include the

production of bacteriocins (important to interspecies competition) and the development of

genetic competence, which allows S. mutans to take up DNA from the environment and further

adapt to its surroundings.

Page 27: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

11

increasingly complex population of bacteria develops through binding of „secondary-‟ or „late-

colonizers‟ to the early colonizers. The attachment of each progressive cell type presents a new

surface for the subsequent attachment of different species of bacteria, resulting in a progression

of nascent surfaces and changes in species diversity that culminates in non-random bacterial

succession in the oral biofilm (Figure 1.2). The adhesive interactions between members of the

same or different species can occur through non-specific physico-chemical interactions, or

through specific adhesion-receptor interactions known as auto-aggregation or co-aggregation

(reviewed in (Rickard, 2008a)). Auto-aggregation refers to the propensity for cells of the same

species to aggregate, while co-aggregations are non-random interactions between genetically

distinct microorganisms (Kinder and Holt, 1994; Rickard, 2008a). Both types of aggregation are

mediated by specific adhesion/receptor interactions, and imply that the formation of dental

plaque is a co-ordinated developmental process involving complex interactions between

bacteria and host, and between interacting bacterial species (Kolenbrander et al., 2006;

Kuramitsu et al., 2007).

S. mutans is among the early colonizers, and uses two methods to attach to the tooth surface

depending on the availability of sucrose in the growth environment. In the absence of sucrose,

the bacterium expresses several main adhesins including streptococcal protein antigen P

(SpaP, also known as antigen I/II), which mediates binding to the salivary component salivary

agglutinin glycoprotein (SAG) (Jenkinson and Demuth, 1997). SpaP is a multi-functional

adhesin that can also mediate binding of the bacterium to collagen and other host proteins

(Demuth and Irvine, 2002). S. mutans may also adhere to other bacteria, the extracellular

matrix and epithelial cell surface receptors in the absence of sucrose using ionic and lectin-like

interactions (Mitchell, 2003). The adhesion of S. mutans to the tooth surface is greatly improved

when sucrose is present. In the presence of sucrose, cell-wall-associated enzymes called

Page 28: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

12

Figure 1.2: Formation of dental plaque. Early colonizing streptococci (in blue) include

Streptococcus sanguinis (formerly S.sanguis), Streptococcus mitis, Streptococcus mutans,

Streptococcus oralis and Streptococcus gordonii. Other early colonizers shown are

Actinomyces naeslundii, Haemophilus parainfluenzae, Capnocytophaga ochracea, Eikenella

corrodens, Prevotella denticola, Prevotella loescheii, Propionibacterium acnes, and Veillonella

atypica. Fusobacterium nucleatum is considered the „bridge‟ between the early colonizers and

the late colonizers, which include Actinobacillus actinomycetemcomitans, Prevotella intermedia,

Treponema denticola and Porphyromonas gingivalis. Figure adapted from (Kolenbrander et al.,

2006).

Page 29: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

13

glucosyltransferases (GTFs) mediate the synthesis of D-glucose polysaccharides known as

glucans, which are in turn bound by the surface associated glucan-binding proteins to promote

cell-cell aggregation. At least three glucan binding proteins (Gbp) have been identified in S.

mutans:GbpA, GbpB and GbpC (reviewed in (Banas and Vickerman, 2003)). Mutation of the

genes encoding either the glucosyltransferases or glucan-binding proteins can alter dental

plaque structure and cariogenesis (Munro et al., 1991; Yamashita et al., 1993). In addition to

the GTFs, S. mutans also produces a fructosyltransferase (FTF) enzyme which synthesizes

fructan polymers (Birkhed et al., 1979). However, the role of fructan in S. mutans is believed to

be in nutrient storage, not in adhesion (Burne et al., 1996). Other adhesins produced by S.

mutans are wall-associated protein A (WapA), SloC, and a protein with similarity to the

fibronectin-binding protein PavA from Streptococcus pneumoniae (reviewed in (Mitchell, 2003).

Once adhered to the tooth surface, S. mutans causes dental caries through its ability to produce

acid from the fermentation of host dietary carbohydrates, and withstand the acidic conditions

better than many other inhabitants of the oral biofilm.

1.2.1.2. Acidogenicity and aciduricity

While other bacterial diseases are caused by the production of virulence factors like secreted

proteases or toxins, the main pathology of dental caries is due to the broad metabolic repertoire

of S. mutans (Lemos and Burne, 2008). Genome sequencing indicates that S. mutans has

genes for the transport and metabolism of glucose, fructose, sucrose, lactose, galactose,

mannose, cellobiose, -glucosides, trehalose, maltose, raffinose, ribulose, mellobiose, starch,

isomaltosaccharides and possibly sorbose (Ajdic et al., 2002)- more metabolic diversity than

any other Gram-positive organism sequenced thus far (Mitchell, 2003). The metabolism of

Page 30: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

14

sugars provides both the principal source of energy for S. mutans, and the acid that causes the

erosion of tooth enamel leading to dental caries. Sugar levels can increase up to 1,000-fold

following intake of heavily sweetened food, which results in a decrease in the pH of the oral

biofilm to values around pH 4 (Lemos et al., 2005). Sustained acidification of the biofilm can

lead to population shifts that favor acid tolerant species. The ability of S. mutans to tolerate the

acidic conditions caused by its own metabolism is central to its survival in the oral biofilm, and

means that stress tolerance by the bacterium is vital to its virulence (Lemos and Burne, 2008).

The production of acid via sugar fermentation lowers the pH of the oral biofilm such that

many species can no longer grow. However, S. mutans has developed a sophisticated series of

both constitutive and inducible mechanisms to cope with the acidification of the environment

resulting from its own metabolism. Central to the constitutive response is a membrane-bound,

acid-stable, proton-translocating F1F0 ATPase that operates to maintain a ΔpH of approximately

1.0 (Quivey et al., 2001). The inductive mechanisms S. mutans uses to adapt to low pH are

collectively referred to as the acid tolerance response (ATR). As part of its ATR, S. mutans

alters its catabolic pathways to attempt to alkalinize the cytoplasm, alters its cell envelope to

decrease proton permeability, and induces the expression of protein chaperones and DNA

repair pathways (for excellent reviews of these processes, see (Lemos et al., 2005; Lemos and

Burne, 2008)).

1.2.1.3 Interspecies competition in the oral biofilm: production of mutacins

The exceptional ability of S. mutans to withstand the acidic conditions in the oral biofilm

allows this microorganism to out-compete many of its co-inhabitants, and become the

Page 31: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

15

numerically dominant species in carious lesions (Loesche, 1986). However, the intense

selective pressures found in the tightly packed oral biofilm community (where nutritional and

oxidative stresses compound with acid stress) may have necessitated the development of

additional competitive strategies. As is often the case, disease causing microorganisms like S.

mutans are also present under healthy conditions, and shifts in the population which favor the

outgrowth of the pathogenic species are part of what lead to disease. Although the

mechanisms leading to changes in the species composition of the dental biofilm are not well

understood, S. mutans gains a competitive advantage in the oral biofilm via the production of

small antimicrobial peptides called bacteriocins.

S. mutans is known to produce a repertoire of ribosomally synthesized antimicrobial peptides

belonging to the bacteriocin family (known as mutacins in the case of S. mutans), which are

secreted into the environment and are active against closely related competitors in the dental

biofilm. Mutacins are important in this dissertation, and are discussed here only briefly from an

ecological perspective in the context of interspecies competition. A more comprehensive

discussion of the structure, regulation and mechanism of action of mutacins in found in section

1.3.4.3).

Mutacins are assumed to confer an ecological advantage to the producing strain in vivo by

eliminating closely related organisms. Mutacin production is therefore advantageous in

situations of high cell density, were closely related organisms are in direct competition for limited

(and metabolically overlapping) nutrients. While Gronroos et al. found a link between mutacin

production and the ability of S. mutans to establish itself in the nacent biofilm following

transmission from mother to child (Gronroos et al., 1998), others have found no correlation

between mutacin expression and transmission (Alaluusua et al., 1991; Longo et al., 2003).

Page 32: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

16

However, Kamiya and colleagues showed distinct mutacin production profiles between

individuals with active carious lesions and those without (Kamiya et al., 2005). The authors

postulated that mutacin production enabled S. mutans to prevail numerically in the biofilm of

caries-active individuals, resulting in more acid production and more tooth destruction. Although

mutacin production is potentially highly advantageous to S. mutans, the production of secreted

proteins is costly for the cell, particularly under conditions of nutrient stress. To ensure that

mutacins are produced only when competitors are present, their expression is linked to

situations of high population density via the quorum sensing system discussed in detail in

section 1.3.3. Quorum sensing in Gram-positive organisms occurs via specialized

sensor/receptor systems known as two-component signal transduction systems. Like the ATR

and the production of mutacins, the stress-responses mediated by two-component signal

transduction systems allow S. mutans to adapt to environmental shifts in the biofilm, and out-

compete its neighbors.

1.2.1.4. Two-component signal transduction systems

The S. mutans reference strain UA159 has only one alternative sigma factors in its genome

(Ajdic et al., 2002). As a result, this microorganism relies on regulatory systems to integrate

various chemical and physical signals to coordinate gene expression in response to stress.

Two Component Signal transduction systems (TCSs) are the systems used by bacteria,

archaea, protozoa, fungi and plants to sense and respond to environmental stimuli (Stock et al.,

2000). They are composed of a transmembrane sensor kinase that detects environmental

changes, and a cytosolic regulatory protein which binds to DNA when activated to modulate

gene expression. Activation of the TCS occurs when a stimulus triggers the

Page 33: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

17

autophosphorylation of the sensor on a conserved histidine residue (giving this class of

receptors the designation histidine kinases, or HKs), which subsequently transfers the

phosphoryl group to a conserved aspartate residue on the cognate response regulator (RR) to

activate it (Stock et al., 2000). Analysis of the UA159 genome sequence revealed the presence

of 13 TCSs (Ajdic et al., 2002), while a fourteenth was subsequently identified experimentally

(Biswas et al., 2008).

Since subverting the ability of S. mutans to sense and respond to stress could potentially

attenuate its ability to cause caries, TCSs are viewed as desirable targets for new antimicrobial

therapies (Lemos and Burne, 2008). Several studies have therefore been conducted to

evaluate the role of each TCS in the stress response (Bhagwat et al., 2001; Kawada-Matsuo et

al., 2009; Levesque et al., 2007) (Summarized in Table 1.2). We preformed one such broad-

ranging study, and found that the TCS-2 (CiaRH) is involved in biofilm formation and tolerance

to environmental stresses, the TCS-3 (ScnRK-like) participates in the survival of cells at acidic

pH, and the TCS-9 affects the acid tolerance response and the process of streptococcal

competence development. Further work was also done to determine the role of the response

regulatory component of the TCS LiaFSR (formerly known as HK/RR11) in biofilm formation

(Chapter 4). Other TCSs have been linked to the oxidative stress response (VicRK,

(Senadheera et al., 2006)), to the stringent response (HK/RR4,(Lemos et al., 2007)), and in

tolerance to acid and DNA damage-induced stress (CiaHR, (Ahn et al., 2006; Biswas et al.,

2008; Qi et al., 2004)) (Table 1.2). The most well studied of the S. mutans TCSs is the ComDE

system, which senses the competence stimulating peptide (CSP) pheromone in a form of

intraspecies communication known as quorum sensing (Li et al., 2001b).

Page 34: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

18

Table 1.2: S. mutans two-component signal transduction systems identified by bioinformatic

analysis of the UA159 genome sequence (modified from (Levesque et al., 2007)).

TCS ORF

(NCBI) Best match

a

(% aa identity) Known Phenotypes

TCS-1 (VicKR)

SMU.1516 VicK S. pyogenes (73%) Oxidative stress, biofilm

formation, competence SMU.1517 VicR S. pyogenes (84%)

TCS-2 (CiaHR)

SMU.1128 CiaH S. agalactiae (60%)

Mutacin production, genetic competence, biofilm formation, oxidative and osmotic stress tolerance, sensitivity to DNA

damage SMU.1129 CiaR S. pneumoniae (89%)

TCS-3 SMU.1145

ScnK S. pyogenes (26%) Acid stress

SMU.1146 ScnR S. pyogenes (41%)

TCS-4 SMU.928

Putative HK S. pyogenes (56%) (p)ppGpp regulation?

SMU.927 Putative RR S. agalactiae (55%)

TCS-5

SMU.1814 ScnK S. pyogenes (66%)

Oxidative stress tolerance, susceptibility to phagocyte-

mediated killing SMU.1815 ScnR S. pyogenes (83%)

TCS-6 SMU.660

SpaK S. thermophilus (44%) Uncharacterized

SMU.659 SpaR S. thermophilus (62%)

TCS-7 SMU.1037

Putative HK C. acetobutylicum (28%) Uncharacterized

SMU.1038 Putative RR L. casei (45%)

TCS-8 SMU.1009

Putative HK S. agalactiae (45%)

Uncharacterized SMU.1008 Putative RR S. suis (71%)

TCS-9

SMU.1965 Putative HK L. johnsonii (39%)

Acid stress, competence

SMU.1964 Putative RR C. acetobutylicum (67%)

Page 35: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

19

TCS-10 SMU.577

LytS S. aureus (39%) Uncharacterized

SMU.576 Putative RR S. agalactiae (77%)

TCS-11

SMU.486 Putative HK S. pyogenes (61%) Cell envelope stress; oxidative

stress in the biofilm, genetic competence

SMU.487 YvqC S. pyogenes (85%)

TCS-12 SMU.1548 Putative HK S. agalactiae (46%)

Uncharacterized

SMU.1547 Putative RR S. thermophilus (78%)

TCS-13 (ComDE)

SMU.1916 ComD S. pneumoniae (36%) Genetic competence, mutacin production, autolysis, biofilm

formation, acid stress tolerance SMU.1917 ComE S. pyogenes (52%)

TCS-14

SMU.45 BaeS Microscilla marina

Uncharacterized

SMU.46 LuxR

1.3. Quorum sensing

Communication amongst cells in higher order organisms like eukaryotes permits the

development and differentiation of highly specialized cell types and complex networks of

interacting systems. It is now widely recognized that prokaryotes also use cell-cell signalling

systems to regulate a diverse set of behavioural and virulence traits between individuals in a

clonal population, resulting in coordinated population-level behaviour (Keller and Surette, 2006).

For true communication to occur, one (or several) individuals must produce a signal that can be

detected by other individuals, and that signal must alter the behaviour of the perceivers (Keller

Page 36: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

20

and Surette, 2006). Most importantly, communication amongst microorganisms will only remain

stable across evolution if both parties benefit from the transfer of information conveyed by the

signal (Keller and Surette, 2006). Communication between different members of prokaryotic

multi-species biofilm community occurs either via physical cell-cell adhesion events mediated by

specific surface-attached protein adhesins and polysaccharide receptors on complementary cell

types, or by diffusible chemical signalling molecules produced by the cells expressly for the

purpose of communication (Rickard, 2008a). We refer to the regulation of bacterial gene

expression in response to cell population density as „quorum sensing‟. Quorum sensing occurs

when bacteria produce, sense, and respond to small extracellular molecules called

autoinducers, that are produced (usually) constitutively as the population grows. As a result, the

extracellular concentration of autoinducer mimics the population density of autoinducer-

producing bacteria. When a threshold autoinducer concentration is reached, it triggers a change

in gene expression of the population as a whole. As a result, bacteria can behave in a co-

ordinated manner akin to that observed with multi-cellular organisms.

1.3.1. A brief history of quorum sensing

The view that bacteria were more than independently acting single celled organisms was

first challenged in the 1960s and 1970s in work done by Tomasz and Hotchchkiss (Tomasz and

Hotchkiss, 1964) on the nasopharyngeal inhabitant S. pneumoniae, and by Nealson and

Hastings on the bioluminescent marine bacteria Vibrio fischeri and Vibrio harveyi ((Nealson et

al., 1970), and reviewed in (Bassler and Losick, 2006)). Tomasz and Hotchkiss reported that

the ability of cultures of S. pneumoniae to co-ordinate DNA uptake from the environment (a

process known as „genetic competence‟) was governed by an extracellular factor produced by

the bacteria themselves (Tomasz and Hotchkiss, 1964). Further work by Pakula and Walczak

Page 37: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

21

identified this competence „activator‟ as a protein-like macromolecule (Pakula and Walczak,

1963), which we now know as competence stimulating peptide, or CSP (Håvarstein et al.,

1995). Nealson‟s work on the bioluminescence of marine vibrios is perhaps the better known

example of the discovery of co-ordinated behaviour in prokaryotes. Nealson and Hastings‟ work

described the ability of V. fischeri and V. harveyi to produce light at high cell density or in cell-

free „conditioned medium‟ from dense cultures. This work was the first to use the term

„autoinduction‟, and suggest that co-ordinated behaviour might be beneficial to a bacterial

population (Nealson et al., 1970). The „autoinducer‟ responsible for biolumniscence was later

identified as an acyl-homoserine lactone molecule (Eberhard et al., 1981). These discoveries

were thought to be isolated incidences of species-specific behaviours, and were largely ignored

for the following 20 years. However, a virtual explosion has occurred in the field of bacterial

cell-cell communication since the early 1990s, during which over two thousand articles have

been published documenting the wide-spread nature of this phenomenon. Bacterial density

dependent cell-cell communication has now been linked to the control of processes including

virulence factor secretion, antibiotic production, sporulation, and biofilm formation (reviewed in

(Waters and Bassler, 2005)) in addition to genetic competence and bioluminescence, and exists

in both Gram-negative and Gram-positive organisms. Intriguingly, recent evidence indicates

bacteria may also be able to communicate reciprocally with their hosts through hormone and

hormone-like signals (reviewed in (Hughes and Sperandio, 2008)). Although beyond the scope

of this review, the possibility of inter-kingdom signalling adds a new level of complexity to the

story of host/microbe co-evolution.

Page 38: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

22

1.3.2. Quorum sensing mechanisms

The mechanics of chemical quorum sensing can be divided into three categories

(summarized in Figure 1.3), which range from broadly-ranging to highly-specific. The first

category (and least specific) is the interspecies quorum sensing system typically referred to as

LuxS/AI-2. The gene required for AI-2 production encodes the S-ribosylhomocysteinase LuxS

(Surette et al., 1999), which functions in the S-adenosylmethionine (SAM) utilization pathway.

SAM is an essential methyl donor in DNA, RNA and other methylation reactions. Its use as a

methyl donor yields the toxic intermediate S-adenosylhomocysteine (SAH), which is detoxified

through several intermediates before eventual cleavage by LuxS into homocystine and 4,5-

dihydroxy-2,3-pentanedione (DPD) (reviewed in (Vendeville et al., 2005)). DPD spontaneously

forms the cyclic pro-AI-2 molecule, which reacts with borate to form a stable cyclic furanosyl

borate diester in equilibrium with several compounds (Thiel et al., 2009), some (or perhaps all)

or which encode information. The biosynthetic pathways, chemical intermediates, and possibly

even the autoinducer signal AI-2 itself are identical in all AI-2 producing bacteria (Xavier and

Bassler, 2003). These similarities mean that the AI-2 signal may be recognized by any AI-2

producing bacteria, and may function in interspecies communication. However, few instances

of true AI-2 mediated communication have been documented in naturally occurring bacterial

pairings (with the exception of the oral biofilm (Rickard et al., 2006; Rickard et al., 2008b)),

leaving some debate as to whether the conservation of the LuxS enzyme across the bacterial

kingdom is due to its role in the activated methyl cycle and not in inter-species signalling.

The second category of quorum sensing is found in Gram-negative bacteria, which use N-

acylated homoserine lactone (AHL) molecules for signalling. Most AHLs in Gram-negative

bacteria are all synthesized via similar pathways, which are homologuous to the V. fisheri

Page 39: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

23

pathway composed of the AHL synthase LuxI and the sensor/response regulator LuxR (Miller

and Bassler, 2001). The binding of an AHL autoinducer by LuxR results in transcriptional

activation of target genes. AHL molecules are species-specific due to variations in acyl-chain

length and substitutions (typically oxo- or hydroxy- groups), although the LuxR sensor may

show some cross-reactivity with AHLs from other species (Keller and Surette, 2006). While the

prototypical interspecies quorum sensing system (AI-2) was initially discovered in V. harveyi

(Surette et al., 1999), this organism also possesses an intraspecies quorum sensing system

which operates via a slightly different mechanism from other Gram-negative organisms. Two

variants of the species-specific AHL signal are produced, called HA-1 and CAI-1. HAI-1 is

synthesized by LuxLM, and sensed by LuxN, while CAI-1 is synthesized by CqsA and sensed

by CqsS. The intraspecies pathway converges with the universal AI-2 pathway at the shared

LuxU phosphorelay protein, and LuxO activates transcription of target genes (Keller and

Surette, 2006) (Figure 1.3).

The most species-specific signals are represented by the peptide quorum sensing systems of

Gram-positive bacteria (Sturme et al., 2002). Typically, the peptide signal is synthesized as a

propeptide, and cleaved upon export by a dedicated ABC-type transporter to generate the

mature signal. This signal is sensed by a cognate receptor protein of a dedicated two-

component signal transduction system. These Gram-positive signals are so specific that

signalling can sometimes discriminate between different strains of the same species (Keller and

Surette, 2006). The best characterized streptococcal quorum sensing system belongs to the

human pathogen S. pneumoniae.

Page 40: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

24

Figure 1.3: Simplified schematic representation of quorum sensing systems in bacteria. A)

LuxI/R quorum sensing in Gram-negative bacteria. Acyl-homoserine lactone molecules are

synthesized by a homologue of the LuxI autoinducer synthase, and freely diffuse out of the cell.

At a specific concentration, AHL binds a homologue of the LuxR response regulator which then

activates transcription of target genes. (B) Peptide quorum sensing system of Gram-positive

bacteria. The quorum sensing signal is synthesized as a propeptide, which is processed upon

export by an ABC transporter to generate the mature signal. The peptide binds to its cognate

receptor protein, which autophosphorylates on a conserved histidine residue. The phosphoryl

group is subsequently transferred to a response regulator protein, which regulates transcription

of target genes. (C, D) Quorum sensing in V. harveyi. Two species-specific systems operate in

Vibrio species: HAI-1 is synthesized by LuxLM, and sensed by LuxN, while CAI-1 is synthesized

by CqsA and sensed by CqsS. The universal quorum sensing system operates via the AI-2

furanone signal (synthesized via LuxS), which is detected at the cell surface by the LuxP

periplasmic binding protein and the receptor LuxQ. Both pathways converge at the shared

LuxU phosphorelay protein, and LuxO activates transcription of target genes.

Page 41: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

25

1.3.3. Quorum Sensing in Streptococci

1.3.3.1. Overview of the CSP-ComDE system in S. pneumoniae

The pioneering work of Pakula and Walczak (described above) identified a secreted signal in

S. pneumoniae which induced genetic competence (Pakula and Walczak, 1963). Subsequent

studies by the Morrison and Håvarstein labs (Håvarstein et al., 1995; Håvarstein et al., 1996;

Pestova et al., 1996) identified this signal as an unmodified peptide of 17 amino acids (ComC,

also known as CSP), which is translated as a 41-amino acid precursor peptide and cleaved

upon export by the dedicated ABC-transporter ComAB. Activation of CSP was found to occur

concomitant to cleavage of the propeptide at a Gly-Gly residue in a cleavage reaction typical of

peptide pheromones and bacteriocins (reviewed in (Senadheera, 2005)). The mature form of

CSP is sensed by the histidine kinase receptor protein ComD, which consists of an N-terminal

membrane-spanning domain and a C-terminal kinase domain (Håvarstein et al., 1996; Stock et

al., 2000). In general, streptococcal species respond only to their own peptide signal and not to

heterologous signals. However, S. pneumoniae is known to harbour 6 distinct CSP variants,

although the majority of strains produce one of two variants: CSP-1 or CSP-2 (Pozzi et al.,

1996; Whatmore et al., 1999). This auto-dependency functions to induce competence in the

presence of genetically identical or highly similar bacteria (and bacterial DNA) are present,

optimizing the chance of donor DNA being similar enough for homologous recombination

(Winans and Zhu, 2000).

Upon CSP binding, ComD autophosphorylates and transfers the phosphoryl group to the

cognate response regulator ComE. Like other response regulators, ComE is composed of an

N-terminal regulatory domain that interacts with the ComD receptor, and a C-terminal DNA-

Page 42: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

26

binding domain (Stock et al., 2000). The phosphorylation of ComE enhances its binding to

promoters containing two 9 bp imperfect direct repeats separated by 12 bp (aCAtTTct(a/g)G ---

12 bp--- ACA(t/g)TtgAG) (Ween et al., 1999). Genes that are transcribed due to direct

ComE/promoter binding are known as „early genes‟, and include the genes encoding the CSP

signal and two-component signal transduction system themselves (transcribed as a single

transcript comCDE) and the ABC transporter responsible for CSP maturation and export

(comAB). This auto-catalytic expression of the CSP-ComDE circuit leads to a high degree of

synchrony in pneumococcal cultures, resulting in induction of competence in almost the entire

population simultaneously. However, expression of the early genes is brief (peaking at 5

minutes after CSP exposure and declining to baseline levels by 20 min post-exposure), and is

followed by a long period in which previously responsive cells become unresponsive to CSP

(Alloing et al., 1998). In total, 23 „early genes‟ have been identified, encoding transporters,

bacteriocin-related genes and regulators (Peterson et al., 2004).

Two early genes transcribed via direct ComE binding are absolutely required for competence

induction: comX (present in duplicate in the genome, and encoding the alternate sigma factor

ComX) (Lee and Morrison, 1999; Luo and Morrison, 2003) and comW, encoding a protein

whose function is less well characterized (Luo et al., 2004). ComX (also known as Sigma X or

σx) shows high homology to the σ70 family of RNA polymerase sigma subunits, which in complex

with RNA polymerase binds to the so-called „cin-box‟ (TACGAATA) upstream of target genes

(Luo and Morrison, 2003). ComX is essential for the transcription of the CSP-induced „late

genes‟, which are expressed maximally after 10-12 minutes (Lee and Morrison, 1999)

(Summarized in Figure 1.4). Eighty-one genes were found to belong to the late class, including

all genes previously identified as necessary for the uptake and processing of DNA for

transformation (Peterson et al., 2004). While ComX controls the competence transcriptome, S.

Page 43: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

27

pneumoniae also controls its competence cascade post-transcriptionally via ComW. ComW

both activates ComX by an unknown mechanism, and acts to prevent its proteolysis by the ClpE

and ClpP proteases (Sung and Morrison, 2005). It was once thought that the proteolytic

degradation of ComX lead to the eventual shut-off of competence (Lee and Morrison, 1999).

However, a rapid decrease of late competence gene transcripts has been observed even when

ComX is still present (Luo and Morrison, 2003). While the ClpE and ClpP proteases appear to

be responsible for the degradation of ComX after cells escape from the competent state, it has

recently been shown that termination of ComX activity is not due to proteolysis of the sigma

factor or of ComW (Piotrowski et al., 2009). Instead, it has been proposed that competence

shut-off could occur due to the presence of a ComE-specific phosphatase (Alloing et al., 1998),

through the modification/sequestration of ComX and/or ComW, through inhibition of translation

of late gene products, or through the accumulation of an inhibitor (Piotrowski et al., 2009).

However, evidence for a ComE-specific phosphatase or inhibitor has remained elusive. Given

the importance of the competence cascade to the physiology of S. pneumoniae, elucidating the

mechanism of its shut-off remains an important goal.

In addition to the well characterized „early‟ and „late‟ genes, a third group of 19 „delayed

genes‟ was identified in the CSP regulon. Expression of this cluster of genes continued to

accumulate after the peak in expression of both the early and late classes, and showed varying

dependence on comX for expression. Genes in this class include stress-reponse genes and

chaperones (Peterson et al., 2004). Interestingly, the competence cascade has been

subsequently linked to the stress reponse in S. pneumoniae (Prudhomme et al., 2006),

discussed in detail in section 1.3.4.4.

Page 44: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

28

1.3.3.2 CSP-ComDE in S. mutans

Pioneering work done in S. pneumoniae has led to the discovery of similar quorum-sensing

regulated competence systems in other streptococci, including those in the oral cavity (for

reviews, see (Martin et al., 2006; Senadheera, 2005)). In the pre-genome sequence era, many

of these competence systems were identified by PCR using primers complementary to the Arg-

and Glu-tRNA genes that flank the comCDE operon in S. pneumoniae (Håvarstein et al., 1997).

However, the genes encoding the competence system in S. mutans are not flanked by tRNA

genes. Instead, the signal peptide is transcribed divergently from the sensor/receptor gene pair

(Figure 1.4; (Li et al., 2001b; Li et al., 2002a)) like the pneumococcal CSP-ComDE paralog BIP-

BlpHR (de Saizieu et al., 2000). In S. mutans, the CSP precursor peptide is 46 amino acids

long, and is thought to be cleaved upon export to generate the mature 21 amino acid-long signal

peptide. Structure analysis of the mature S. mutans CSP indicated that the C-terminal portion

of the peptide is essential for receptor activation (Syvitski et al., 2007). In contrast to the

multiple pneumococcal CSP pherotypes, S. mutans produces only one CSP pherotype despite

the presence of minor genetic variation in sequence among strains (Allan et al., 2007). When

the peptide reaches a critical concentration (at early- to mid-exponential phase in S. mutans), it

activates the ComDE TCS leading to competence induction through a homologue of the

alternate sigma factor ComX. In contrast to the pneumococcal model, only a small percentage

(≤ 1%) of the S. mutans population ever becomes competent in the presence of CSP when

grown planktonically. In addition, a residual level of transformation is present in S. mutans

mutants lacking the com genes (reviewed in (Senadheera, 2005)). Moreover, while competence

develops between 5-7 minutes following exogenous CSP addition in S. pneumoniae, a ~2 hr

delay in competence is common following exogenous CSP addition in S. mutans. The highest

levels of transformation in S. mutans are observed during biofilm growth, during which

transformation efficiencies can approach levels 10- 600-fold higher than their planktonic

Page 45: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

29

counterparts (Li et al., 2001b). While this increase in transformation efficiency may simply

represent more efficient CSP signalling between cells in close spatial proximity, the known

phenotypic differences between biofilm and planktonic cells means that additional biofilm-

specific factors cannot be ruled out. Together, these results indicate that multiple inputs and/or

shut-off mechanisms many exist for the S. mutans competence cascade, and that significant

differences exist between the prototypical competence cascade of S. pneumoniae and that of S.

mutans (summarized in Table 1.3). However, most of the phenotypes regulated by this system

in pneumococcus appear to be also regulated by its homologue in S. mutans.

Table 1.3: Comparison of the CSP-ComDE systems in S. pneumoniae and S. mutans.

S. pneumoniae S. mutans

Peptide Pheromone Induction Signal CSP, 17 aa CSP, 21 aa

CSP Pherotypes? 6 known, 2 main 1

Other Induction Stress (antibiotics,

DNA damage) ??

Receptor/Effectors ComDE, ComX ComDE, ComX

Accessory proteins ComW ??

Paralogous system? BIP-BlpHR No

Timing of Induction after CSP addition 5-7 minutes ~2 hours

Percent of culture induced Approaching 100% ≤1%

Phenotypes controlled

Competence,fratricide, biofilm formation, stress

response

Competence, bacteriocin production, biofilm formation, autolysis, stress response.

Page 46: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

30

Figure 1.4: Mechanistic and genomic representation of the CSP-ComDE circuit controlling

competence development in S. pneumoniae and S. mutans. Entries in parentheses are unique

to S. pneumoniae.For simplicity‟s sake, only genes involved in competence development are

shown. The role of the pneumococcal CSP-ComDE circuit in stress tolerance and fratricide is

discussed in section 1.3.4, and a more detailed representation of the fratricidal pathway is

presented in Figure 1.5.

Page 47: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

31

1.3.4 Phenotypes controlled by CSP-ComDE signalling in streptococci

1.3.4.1 Genetic competence

Genetic competence is a transient physiological state during which bacteria take up DNA

from the environment and integrate it into their genomes (Cvitkovitch, 2001). For successful

gene transfer to occur, recipient cells must be in a metabolically active, competent state, and

DNA must be present in the environment. For recombination to occur, the foreign DNA

sequence must share between 70-100% identity with the sequence in the recipient strain‟s

genome (Dowson et al., 1997). Considerable energy is used during the transition to

competence, which also requires the complex genetic machinery described in the preceeding

section. Competence therefore plays an important role in the physiology of the organism to

remain evolutionarily conserved in such varying species. A brief overview of the processes

involved are discussed below.

1.3.4.2.1 Mechanism of DNA uptake in streptococci

In Gram-positive bacteria, DNA must pass through the cell wall and the cytoplasmic

membrane before integration into the genome for transformation. In streptococci, this process

is initiated when exogenous double-stranded DNA binds to the surface of a competent cell,

which has been induced to express DNA uptake and processing machinery by CSP. This

binding is sequence-independent in streptococci, meaning that both autologous and foreign

DNA can be incorporated (Berge et al., 2002). The first hurdle in DNA uptake is traversing the

cell wall. To take up DNA, the cell expresses a pseudopilus composed of the ComGC (major

pilin-like protein), ComGD, ComGE and ComGG (minor pilin-like) proteins (Dubnau, 1999). The

Page 48: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

32

retraction of the pseudopilus (by disassembly of the structure) allows the DNA to traverse the

cell wall peptidoglycan (Craig and Li, 2008). Once across the cell wall, double stranded DNA

binds to the C-terminal domain of the membrane-anchored DNA binding protein ComEA. Single

stranded breaks are then introduced into the bound DNA molecule by an unknown

endonuclease, and one strand of the double-stranded DNA molecule is degraded by the

endonuclease EndA concomitant with its import in a 3‟ to 5‟direction (Mejean and Claverys,

1993). This reaction has been shown to be energy dependent, and to occur through an

aqueous channel in the membrane composed of ComEC, ComFA and possibly EndA itself

(Dubnau, 1999). Following uptake, DNA recovered from transformed pneumococci is no longer

able to transform other cultures of competent cells. This transient loss of transforming activity is

called eclipse (Ephrussi-Taylor, 1960), and reflects the fact that ssDNA has less transforming

activity than dsDNA (Miao and Guild, 1970). Investigation of eclipse DNA revealed that ssDNA

isolated after transformation was bound to protein, forming the „eclipse complex‟ (Morrison,

1977, 1978). This protein component has been subsequently identified as SsbB (Morrison et

al., 2007), and renders ssDNA in eclipse complex 50-1000-fold more resistant to exogenous

nuclease activity (Morrison and Mannarelli, 1979). Incoming ssDNA is also protected via the

recombination mediator protein DprA and the recombinase RecA (Berge et al., 2003). DprA is

involved in recruiting RecA to the ssDNA, which can be then wholly or partially integrated into

the recipient‟s genome (depending on homology) via RecA-mediated recombination (Mortier-

Barriere et al., 2007). Although beyond the scope of this dissertation, a detailed description of

the steps involved in DNA uptake, processing and recombination can be found in (Claverys et

al., 2009).

Page 49: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

33

1.3.4.2.2 Purpose of DNA uptake

There are now over 70 known naturally transformable bacterial species in nature (Johnsborg

et al., 2007). Competence requires the expression of more than a dozen specialized proteins,

which are all highly controlled in terms of expression. The widespread and complicated nature

of the competence phenotype begs the question “What use is competence?” to a bacterial cell?

Three hypotheses are generally put forward to explain the evolutionary conservation of the

competence phenotype: DNA for food, DNA for repair, and DNA for genetic diversity.

The fact that competence induction is generally induced by „quorum sensing‟-type systems

implies that it is a phenotype associated with high cell density. Since nutritional starvation is a

feature of high cell density, it has been proposed that DNA uptake may serve as a source of

nucleotides for a starving cell. In fact, DNA may serve as the sole carbon and energy source for

Escherichia coli at stationary phase (Finkel and Kolter, 2001). While the DNA-for-food

hypothesis is certainly valid, several points argue against it as the sole role for the DNA uptake

machinery. Firstly, although the CSP-ComDE circuit has been shown to be necessary for

transformation in S. mutans, competence naturally develops in liquid cultures at very low cell

density (OD600 ~0.1) where nutrients are abundant and nucleotide scavenging is not necessary.

More generally, the degradation of one strand of the incoming DNA molecule suggests that the

uptake of nucleotides for food is not the main goal of transformation- why throw away half your

food if you are hungry? Moreover, B. subtilis is known to secrete a powerful nonspecific

nuclease, and express uptake systems for the nucleolytic products. This would seem to be a far

simpler mechanism for scavenging nucleotides from the environment (Dubnau, 1999). Finally,

the naturally competent organisms Haemophilus influenzae and Neisseria gonorrhoeae exhibit

sequence specificity in their DNA uptake systems to restrict incoming DNA to their own species.

Page 50: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

34

This presumably acts to increase the chances of homologous recombination, and strongly

suggests the DNA for repair or transformation hypotheses.

Evidence supports both the DNA-for-repair and the DNA-for-genetic diversity hypotheses,

and may indicate that the two are not mutually exclusive. It is now known that DNA damaging

agents can induce the competence regulon in S. pneumoniae (Prudhomme et al., 2006), and it

is widely recognized that DNA repair machinery is induced during competence in many Gram-

positive organisms (reviewed in (Dubnau, 1999)). However, supporting the DNA-for-repair

hypothesis, transformation of UV-irradiated H. influenzae with a cloned fragment had the same

effect on survival as transformation with total chromosomal DNA (while foreign DNA had no

effect), indicating that recombination is important, but not necessary at the site of repair

(Mongold, 1992). An alternative explanation to the induction of competence in stress is therefore

that cells are scavenging the environment for fitness-enhancing genes. It is widely recognized

that non-competent bacteria become hypermutable under stressful conditions, which has been

argued to be an attempt to generate genetic diversity under stress (Bjedov et al., 2003). While

this strategy is efficient enough to be evolutionarily conserved, it would be an obvious

advantage for competent bacteria to also co-ordinate the uptake of intact and functional DNA

from the environment during stress. It has been proposed that S. pneumoniae induces its

competence regulon as part of its general stress response for just such a purpose (Claverys et

al., 2006). Since instances of adaptation mediated by horizontal gene transfer are known to

exist in nature, this hypothesis has gained acceptance in the literature, and is widely recognized

as one of the principle functions for competence.

Page 51: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

35

If one accepts the DNA-for-genetic diversity hypothesis, however, the obvious next question

becomes “What is the source of transforming DNA?” Interestingly, recent studies have shown

that the majority of CSP-responsive genes in S. pneumoniae (~70 of 105-124 genes) are

dispensable for transformation (Dagkessamanskaia et al., 2004; Peterson et al., 2000; Peterson

et al., 2004). What is the function of these other CSP-induced genes? The answer may lie in

the newly discovered link between the competence cascade and autolysis.

1.3.4.2 Autolysis and pneumococcal fratricide

Nowhere is the concept of quorum sensing-mediated multi-cellular behaviour more apparent

than in the field of bacterial cell death and lysis. First observed in S. pneumoniae in the early

1900s, the seemingly counterproductive habit of some bacterial cells to commit „cellular suicide‟

and self-lyse has received considerable attention in the past decade. In the context of multi-

cellular biofilm growth, programmed cell death-like processes may benefit the community by

sacrificing individual cells damaged by toxic factors, viral infection or during nutritional

starvation. The lysis phenomenon has been documented in several developmental processes,

and is often linked to quorum sensing, high population density, or environmental stress. A

distinction is made between the phenomenon of autolysis, observed during biofilm formation in

Staphylococcus aureus and preceding sporulation in Bacillus subtilis and the fratricidal pathway

of S. pneumoniae, which involves the killing of non-competent sibling cells by CSP-induced but

(otherwise clonal) competent cells.

Page 52: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

36

1.3.4.2.1 Autolysins and autolysis

Autolysis is the self-digestion of the cell wall by bacterial peptidoglycan (or murein)

hydrolases called autolysins (Shockman et al., 1996). It is thought that these cell wall

degrading enzymes are involved in normal cell wall turnover, and have been shown to be

required for daughter cell separation after completion of the newly formed septum during cell

division (Rice and Bayles, 2008). Additional roles for autolysins have been shown in antibiotic

resistance (Groicher et al., 2000), cell-to-surface adhesion (Heilmann et al., 1997; Heilmann et

al., 2005), genetic competence (Moscoso and Claverys, 2004), protein secretion and

pathogenicity (Ahn and Burne, 2007). The activity of murein hydrolases has been studied

extensively in Gram-negative bacteria, and regulatory mechanisms such as sequestration within

lipid membranes, controlled transport across the cytoplasmic membrane, and topographical

control within the peptidoglycan have been reported (reviewed in (Holtje and Tuomanen, 1991)).

In Gram-positive organisms, the control of autolysin activity may also be regulated by the family

of carbohydrates referred to as teichoic acids. Two basic forms of teichoic acids are present in

the cell wall of Gram-positive bacteria: wall teichoic acids, which are covalently attached to the

peptidoglycan, and lipoteichoic acids, which are anchored in the cytoplasmic membrane. These

sugar molecules combine with the peptidoglycan to form a polyanionic gel surrounding the cell,

which acts as a semi-permeable buffer zone between the external environment and the

cytoplasmic membrane (Rice and Bayles, 2008). Since they are negatively charged, teichoic

acids are susceptible to cationic peptide antibiotics and bacteriocins. However their primary

function may be to control the activity of autolysins during normal cell growth.

The autolysins produced by many bacteria contain elements that target them to the cell wall.

In the case of S. pneumoniae, the primary autolysin LytA contains repeated choline-binding

Page 53: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

37

domains which associate with choline-substituted teichoic acids in the cell wall. The autolysin is

activated in vitro by this association in a process known as „conversion‟ (reviewed in (Rice and

Bayles, 2008)). Teichoic acid modification has also been implicated in regulation of murein

hydrolase activity. D-alanylation of teichoic acids plays a vital role in modulating surface charge,

which seems to have a negative impact on autolysin activity. The extent of cell wall modification

with D-ala is thought to be dependent on the charged state of the membrane, or its proton

motive force (PMF), since protonation is thought to stabilize D-alanyl ester linkages within the

teichoic acid. Therefore, in a metabolically active cell, the pH gradient that is established

moving away from the respiring membrane results in less D-ala modification farther away from

the cell membrane than close to it. As a result, cell-wall associated autolysins positioned farther

away from the membrane are most active, accounting for the increased peptidoglycan turnover

in the outer layers of the cell wall (Figure 1.5). This model also accounts for data showing that

autolysis is inhibited by growth at low pH (where D-ala substitution of the teichoic acids would

increase), and that PMF was critical in controlling this process. The processes described up to

this point are part of the normal control of cell wall re-modelling during growth. However,

induction of autolysin activity has also been described during nutrient starvation and biofilm

growth in the major Gram-positive human pathogen Staphylococcus aureus and in response to

CSP accumulation in S. pneumoniae.

1.3.4.2.2 The holin/antiholin system in S. aureus

The role of PMF in control of autolysis is best illustrated in the autolysis cascade of S.

aureus. Autolysis in this organism is mediated by the balance of gene expression between two

regulatory systems: 1) LytSR, which controls the expression of a bicistronic operon designated

lrgA and lrgB and 2) CidR, controlling expression of a second bicistronic operon designated

Page 54: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

38

cidA and cidB. The gene products of lrgAB and cidAB are predicted to be extremely

hydrophobic, and are likely membrane proteins. Interestingly, all four proteins showed high

homology with bacteriophage membrane permeabilizing proteins called holins. Holins function

by inserting into the cell membrane causing pores, leading to the gradual dissipation of the

proton gradient. They are also thought to control phage-encoded murein hydrolases by either

controlling their transport or mediating their activation (Bayles, 2007). In S. aureus, CidA is

predicted to function as the „holin‟ component, causing PMF dispersal and activation of

bacterially encoded murein hydrolases in the cell wall. The presence of LrgA is thought to

inhibit the action of CidA, earning it the designation „anti-holin‟ (Figure 1.5). The CidAB/LrgAB

cell death pathway has been implicated in biofilm development in S. aureus, due to the release

of genomic DNA for stabilization of the extracellular matrix. While cell death mediated by the

holin/antiholin system is induced by changes in carbohydrate metabolism, not by a CSP-like

peptide, the functional similarity between the membrane-bound pore-forming holin-like peptides

and the pore-forming membrane bound bacteriocins induced by CSP and involved in fratricide

deserves mention.

1.3.4.2.3 Streptococcal fratricide

While all cells in a population of S. aureus are potentially susceptible to autolysis by the

CidAB holin system, the process of streptococcal fratricide is characterized by lysis of the non-

competent fraction of the population by competent sister cells. One constitutively expressed

autolysin, encoded by lytC, and the products of six CSP-induced genes, cbpD, cibA, cibB, cibC,

comM and lytA have been implicated in CSP-induced cell death in S. pneumoniae (reviewed in

(Claverys et al., 2007)) (Figure 1.6). In liquid culture, the key component of the lytic pathway is

the late competence gene cbpD. CbpD is a putative amidase/peptidase consisting of a cysteine

Page 55: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

39

Figure 1.5: Representation of the holin/anti-holin system in S. aureus. In this model, the proton

gradient that naturally exists in a polarized, respiring cell causes the D-ala-D-ala linkages on

teichoic acids and lipoteichoic acids close to the cell surface to become protonated (red

crosses). This local area of protonation keeps murein hydrolases present in the cell wall in an

inactive form. Further away from the cell membrane, the localized area of acidity is reduced,

and the inhibition of murein hydrolases is lifted, allowing for peptidoglycan degradation in the

outermost leaflet of the cell wall. The holin/anti-holin system composed of CidAB and LrgAB

(respectively) maintains the polarized state of the membrane when both components are

present. If lysis is triggered, the LrgAB antiholin component is degraded, and the holin

component causes membrane depolarization, de-protonation of the teichoic acids, and

subsequent activation of the murein hydrolases in the cell wall (Figure adapted from (Rice and

Bayles, 2008)).

Page 56: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

40

protease (CHAP) domain, two 3(SH3b) domains with Src homology, and four choline-binding

domains (Guiral et al., 2005). CHAP domains are present in bacterial and phage autolysins,

and are known to act as murein hydrolases. SH3 domains in eukaryotes function in protein-

protein interactions involving proline-rich motifs, but in bacteria may mediate protein binding to

the cell wall (Guiral et al., 2005). The choline binding domains likely fulfill a similar function by

mediating binding of CbpD to the teichoic acid residues to position the CHAP domain relative to

its substrate. It has been suggested that CbpD is tethered to the surface of competent cells,

where it cleaves the peptide bonds within murein stem peptides of target cells via cell-to-cell

contact. This cleavage is thought to trigger the activation of the autolysins LytA and LytC,

resulting in cell lysis (Guiral et al., 2005). While CbpD is expressed only in competent predatory

cells, recent work has shown that LytA and LytC are more effective when expressed by the

target cells themselves (Eldholm et al., 2009).

While CbpD is absolutely required for cell lysis in liquid culture, the products of cibA and cibB

form a putative two- peptide bacteriocin that is necessary for allolysis in S. pneumoniae grown

on solid surfaces (Guiral et al., 2005; Kausmally et al., 2005). Although it has been proposed

that CibAB are also tethered to the cell surface of competent cells and function via cell contact-

dependent activation of LytA and LytC, the mechanism of action is not known (Claverys et al.,

2007; Guiral et al., 2005). Competent cells protect themselves from lysis through the action of

ComM and CibC, which are protective against the autolysins (CbpD, LytA and LytC) and the

bacteriocin (CibAB), respectively (Guiral et al., 2005; Håvarstein et al., 2006)(Figure 1.6).

Pneumococcal fratricide is thought to occur to provide DNA for uptake during natural

transformation, and has been suggested to represent a means for S. pneumoniae to increase its

genetic repertoire through DNA exchange under stress. CSP is also known to induce cell death

Page 57: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

41

at high concentrations in S. mutans(Qi et al., 2005). Homologues of several autolysin- and

fractridical effector-encoding genes exist in S. mutans (Table 1.4), although their involvement in

CSP-induced cell death in this organism is not known.

Figure 1.6: Proposed mechanism of action of the CSP-induced fratricidal pathway in S.

pneumoniae. A pre-competent cell may differentiate into either a competent lineage if it

responds to CSP, or remain non-competent if it expresses and responds to a different

pherotype than what is present in the environment. CSP-responsive cells express the CbpD

amidase or the CibAB bacteriocins on their cell surface (in liquid culture or on solid surfaces,

respectively), and are protected via the ComM or CibC immunity proteins. Non-competent

sibling cells are then killed via cell-to-cell contact, which requires the presence of the LytA and

LytC autolysins. Figure adapted from (Rice and Bayles, 2008)

Page 58: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

42

Table 1.4: Autolysin and fratricidal effector genes and their homologs in S. mutans UA159

Genea

(GenBank accession)

Description Homologue in S. mutans UA159 (% aa identity)b

S.

au

reu

s

lytR (L42945) Two-component response regulator

LytR, controls lrgAB SMU.576 (38%; 96/249)

lytS (L42945) Two-component sensor histidine

kinase LytS, controls lrgAB SMU.577 (42%; 245/582)

lrgA (U52961) Antiholin-like protein LrgA SMU.575 (38%; 46/118)

lrgB (U52961) LytSR-regulated gene SMU.574 (48%; 105/215)

cidA (AY581892) Holin-like protein CidA SMU.1701 (40%; 35/87)

cidB (AY581892) Hydrophobic protein CidB SMU.1700 (30%; 63/207)

cidC (AY581892) Pyruvate oxidase SMU.231 (29%; 165/560)

cidR (AY581892) LysR-type transcriptional regulator

(LTTR) SMU.2060 (22%; 61/273)

S.

pn

eu

mo

nia

e

lytA (SPD1737) autolysin/N-acetylmuramoyl-L-

alanine amidase SMU.704 (25%; 32/128)

cbpD (SPD2028) choline binding protein D SMU.609 (32%; 47/144)

lytC (SPD1403) 1,4-beta-N-acetylmuramidase,

putative SMU.689 (24%; 47/192)

cibA (SPG0129) competence induced bacteriocin A SMU.150 (35%; 20/57)

cibB (SPG0128) competence induced bacteriocin A no significant homologies

aNCBI annotation.

bGenomic BLAST Search (http://www.ncbi.nlm.nih.gov/sutils/genom_table.cgi) was used

for sequence comparison.

Page 59: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

43

1.3.4.3 Bacteriocin production

In addition to competence and fratricide, a third CSP-regulated phenotype in streptococci is

the production of bacteriocins. Bacteriocins are small ribosomally synthesized peptides or

proteins with antibacterial activity. As mentioned above in section 1.2.1.3, bacteriocins are

assumed to confer an ecological advantage to the producing strain in vivo by eliminating closely

related organisms. Mutacin production is therefore advantageous in situations of high cell

density, were closely related organisms are in direct competition for limited (and metabolically

overlapping) nutrients.The field of bacteriocin research has been directed mainly towards those

produced by lactic acid bacteria in the food industry, as a means of preventing food spoilage. In

the age of increasing antibiotic resistance, bacteriocins have been receiving increasing attention

as an alternative means of preventing infection by pathogens (reviewed in (Nes et al., 2007)). A

survey of 143 strains of S. mutans revealed that 70% produce one or more bacteriocins in vitro

(van der Ploeg, 2005), although their ecological significance in dental plaque is still a matter of

debate (see section 1.2.1.3, above). Studies by van der Ploeg and Kreth have shown that the

expression of mutacin IV is regulated via CSP-ComDE, and is co-ordinated with expression of

the competence system which can lead to exchange of DNA in co-culture with the mutacin-

susceptible strains (Kreth et al., 2005; van der Ploeg, 2005). Bacteriocins therefore appear to

play important roles in both the ecology and the genetics of S. mutans.

1.3.4.3.1 Classification of bacteriocins

Several classifications schemes have been proposed for bacteriocins from Gram-positive

bacteria, based on their structure and mode of action. For the purpose of this dissertation, we

Page 60: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

44

adopt the classification scheme proposed by Cotter et al. (2005) (Figure 1.7). Accordingly,

Class I bacteriocins, or lantibiotics, contain the amino acids lanthionine or β-methyllanthionine,

and are post-translationallly modified (Cotter et al., 2005). Up to 11 subclasses have been

proposed for the lantibiotics, according to structural features. Well known bacteriocins including

nisin, mersacidin and cytolysin are characterized as class I (lantibiotics). Class II bacteriocins

include small non-modified, heat-stable peptides that do not contain lanthionine modifications.

They are a heterogeneous class, and are divided into subclass IIa (pediocin-like bacteriocins ex.

leucocin A), subclass IIb (two-peptide bacteriocins ex. mutacin IV), subclass IIc (cyclic

bacteriocins ex. enterocin AS48), and subclass IId (non-pediocin single linear peptides ex.

lactococcin A) (Cotter et al., 2005). The third class of bacteriocins is represented by non-

bacteriocin lytic proteins called bacteriolysins. This class includes large, heat-labile proteins

(often murein hydrolases), like lysostaphin and enterolysin A (Cotter et al., 2005). Most

bacteriocins from streptococci belong to the lantibiotic class (Nes et al., 2007). However,

among the bacteriocins classified for S. mutans are both traditional class I bacteriocins (mutacin

I, mutacin II, mutacin III/mutacin 1140, mutacin N and mutacin B-Ny266), two-peptide class I

lantibiotics (SmbA and SmbB), a class IIb two-peptide bacteriocin (mutacin IV) and the

bacteriocin mutacin V, which has structural homology to class IIa pediocin-like bacteriocins.

1.3.4.3.2 Biosynthesis and export of class IIa bacteriocins

Mutacin V is a class IIa bacteriocin, and deserves specific mention since it is the focus of

Chapter 2 of this dissertation. Expression of class II bacteriocins like mutacin V typically

requires an inducer peptide pheromone and a TCS (van der Ploeg, 2005). Interestingly, the

majority of putative bacteriocins and their accessory genes identified in the UA159 genome are

located in a 13.5kb island which also harbours the comC, comD and comE genes (van der

Page 61: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

45

Ploeg, 2005) (Table 1.5). Some of these bacteriocins-encoding genes including mutacin V were

also found preceded by a putative ComE binding site, implicating this regulator in their

expression. Furthermore, the gene encoding mutacin V (SMU.1914, also known as nlmC) is

located immediately upstream of comC itself (Figure 1.4), although the two are transcribed

divergently.

Table 1.5: Putative and known bacteriocins encoded by S. mutans strain UA159 located in the

genomic island that also includes CSP-ComDE (adapted from (van der Ploeg, 2005))

Gene ID Size of pre- bacteriocin

(amino acids)

Putative ComE

binding site

Characteristics

SMU.1889 87 - SMU.1889 and SMU.1892 are

located adjacent to each other, may form a two-peptide bacateriocin SMU.1892 61 -

SMU.1895 53 - Separated from SMU.1889/1892 by

an insertion element. SMU.1895/1896 may also form a two

peptide bacteriocin. SMU.1896 83 -

SMU.1902 47 - Single peptide bacteriocin?

SMU.1905 62 + SMU.1905 and SMU.1906 are located adjacent to each other, may

form a two-peptide bacateriocin SMU.1906 70 +

SMU.1914 76 + Transcribed divergently from comC. Single peptide bacteriocin mutacin V

Page 62: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

46

At least four genes are required for production of class IIa bacteriocins like mutacin V: 1) the

structural gene, encoding the „prebacteriocin‟ with its leader peptide; 2) a gene encoding the

immunity protein, which is usually co-transcribed with the bacteriocin structural gene; 3) a gene

encoding an ABC transporter necessary for secretion of the bacteriocin; and 4) a gene encoding

an accessory protein of unknown function (Drider et al., 2006). These four required gene

elements are not necessarily found in a single operon, and may be found transcribed as three

separate units where one operon encodes the bacteriocin and immunity protein, a second

operon carries genes for secretion, and a third operon encodes genes involved in the regulation

of bacteriocin expression (Drider et al., 2006). Interestingly, with few exceptions (including

mutacin V), most class IIa bacteriocins are plasmid encoded (Drider et al., 2006).

The class IIa bacteriocins are translated as „prebacteriocins‟, having an N-terminal extension.

This presequence is removed during export by site-specific cleavage following a conserved

double glycine motif. This leader sequence may serve as an export signal to direct bacteriocins

to the correct ABC transporter, but is also thought to play a protective role in preventing the

insertion of the bacteriocin into the membrane of the producing cell (Drider et al., 2006).

1.3.4.3.3 Mode of action

The majority of bacteriocins have a net positive charge and contain sequences of

hydrophobic and/or amphiphilic nature, allowing them to insert into the negatively charged

Gram-positive cytoplasmic membrane, creating pores in target cells (Hechard and Sahl, 2002).

The result is a disruption of proton motive force, ATP depletion and leakage of nutrients and

Page 63: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

47

Figure 1.7: Cartoon representation of the mechanism of action of bacteriocins, divided

according to classification. Class I bacteriocins both inhibit cell wall synthesis by binding to lipid

II (to prevent translocation of peptidoglycan precursors across the cell membrane), and create

pores in target cell membranes. Class II bacteriocins act by pore formation and disruption of the

cell‟s PMF. Bacteriolysins actively degrade the cell wall, and are considered murein hydrolases

rather than true bacteriocins (figure adapted from (Cotter et al., 2005)).

Page 64: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

48

metabolites from the target cell, and eventual cell death (Hechard and Sahl, 2002). The killing

spectrum of bacteriocins is narrow, and typically includes only closely related target organisms

(Drider et al., 2006). However, with few exceptions, little is known about how bacteriocins

specifically recognize their target cells. The well characterized type I lantibiotic bacteriocin nisin

employs the cell-wall precursor lipid II as a docking molecule, and subsequently kills cells by

simultaneously inhibiting peptidoglycan biosynthesis and creating pores in the cytoplasmic

membrane (Linnett and Strominger, 1973). The receptor for lactococcin A (and similar pediocin-

like type IIa bacteriocins) is also known. This class of bacteriocins act by binding to the proteins

IIC and IID of the mannose phosphotransferase system and permeabilizing the cytoplasmic

membrane (Diep et al., 2007). Immunity is conferred to the producing cell via binding of the

cognate immunity protein to the bacteriocin-receptor complex, thereby preventing the further

action of the bacteriocin (Diep et al., 2007).

1.3.4.4 CSP and the stress response

Bacteriocins are often produced by bacteria to inhibit competitors at high cell density, during

which nutrient, oxidative, and acid end-product stresses abound. Importantly, recent work by

the Claverys lab on S. pneumoniae indicates that the fourth CSP-regulated phenotype in

streptococci may be the coordination of the general stress response. Prudhomme et al. (2006)

demonstrated that antibiotic and DNA-damage stress could induce competence in pneumococci

via up-regulation of expression of the CSP-ComDE circuit (Prudhomme et al., 2006). Induction

of competence in these stressed cells was proposed to be a survival strategy designed to

enhance the fitness of the organism by allowing it to scavenge the environment for potential

resistance genes. These authors suggested that CSP is not simply an indicator of cell density in

Page 65: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

49

pneumococci, but may also signal the presence of environmental stress as an inducible peptide

„alarmone‟. Furthermore, Pinas et al. (2008) found that acid stress could induce autolysis

independently from competence in S. pneumoniae in a CSP-independent ComE-mediated

pathway. These authors proposed that ComE is a principal player in a global stress response

that includes the uptake of fitness enhancing DNA via the competence cascade (Pinas et al.,

2008). Finally, competence has been shown to respond to the presence of alkaline conditions

also in a cell-density-independent manner. These results imply that competence is linked to the

stress in the environment more closely than to cell density.

In S. mutans, evidence has implicated the CSP-ComDE pathway in the acid tolerance

response (Li et al., 2001a), arguably one of the most important environmental stresses that this

organism encounters. Furthermore, the coordinated regulation of competence and bacteriocin

production through CSP-ComDE has been shown to result in DNA exchange between S.

mutans and its neighbour S. gordonii (Kreth et al., 2005). Even in the most traditional definition

of function for CSP, it can be argued that the CSP-ComDE circuit is involved in adaptation to the

significant competitive stresses that occur at high cell density. Since the oral biofilm

environment is rife with stresses, could the S. mutans CSP-ComDE system be functioning as a

mediator of the stress response in that environment? Biofilm formation may represent the

combined effects of all CSP-mediated phenotypes described thus far.

1.3.4.5. Biofilm formation

The formation of a densely packed biofilm provides an efficient milieu for chemical signalling,

which breeds cooperative „multi-cellular‟-type behaviours (described in detail in sections 1.1 and

Page 66: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

50

1.2). Environmental stresses accumulate when cells are located in such close proximity,

demanding efficient stress response mechanisms. The autolytic sacrifice of individual cells

damaged by toxic factors, viral infection or during nutritional starvation, the production of

secreted bacteriocins, and/or the scavenging of fitness-enhancing genes from the environment

through competence induction may provide solutions to the problems associated with growth at

high cell density. Given the nature of the high cell density biofilm lifestyle, it is not surprising

that biofilm formation represents the final phenotype we describe which has been linked to CSP

signalling in S. pneumoniae (Oggioni et al., 2006) and S. mutans (Li et al., 2001b; Li et al.,

2002b).

In vitro, S. pneumoniae forms biofilms only in the presence of exogenously added CSP, and

does not form biofilms in the absence of the ComD receptor (Oggioni et al., 2006). These

authors also reported that ComD-deficient mutants were less virulent in a murine model of

pneumoniae, which they describe as a biofilm-like infection. While the study of CSP‟s influence

on biofilm formation in S. pneumoniae has received less attention than its role in competence

development, a well established connection exists between CSP and biofilm formation in S.

mutans.

Evidence supporting a role for CSP signalling in S. mtuans biofilms was first presented by Li

et al., who found that a comC- mutant of S. mutans formed a biofilm with an altered structure,

which could be restored to wild-type architecture using exogenously added CSP or plasmid-

based complementation (Li et al., 2002b). Interestingly, biofilm formation by mutants defective

in the ComD and ComE TCS components formed biofilms with reduced biomass which could

not be fully complemented with exogenous CSP (Li et al., 2002b). The different biofilm

Page 67: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

51

phenotypes associated with inactivation of the signal and sensory components of the QS

system indicate that CSP may impact on more than one TCS in the biofilm. Alternatively,

signals other than CSP may be sensed by ComDE in the biofilm. The expression of ComX has

also been monitored in S. mutans biofilms using a green fluorescent protein (GFP)-promoter

fusion, revealing CSP signalling occurs in areas of high cell density (Aspiras et al., 2004).

Although similar results linking CSP signalling to biofilm formation and architecture in S. mtuans

have since been reported by others (Petersen et al., 2005; Zhang et al., 2009), no mechanism

has been proposed to explain the role of CSP-ComDE in biofilm formation.

This dissertation aims to address the role of the CSP peptide pheromone in the physiology of

S. mutans, with specific focus on its role in autolysis, competence, and biofilm formation. We

attempt to show that the overriding theme common to all CSP-responsive phenotypes in S.

mutans is their involvement in a global CSP-mediated stress response, which is necessary to

the proper formation and maintenance of the high density biofilm community.

1.4. Statement of the problem

As the major etiological agent of human dental caries, the naturally transformable oral

bacterium Streptococcus mutans is well studied at the genetic and physiological level. The

regulatory system that governs genetic competence in this species is similar to the system in

S. pneumoniae, and is composed of the CSP peptide pheromone, the ComDE two-component

signal transduction system, and the alternate sigma factor ComX. In S. mutans, the CSP

pheromone has been linked to the induction of genetic competence, autolysis, bacteriocin

Page 68: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

52

production and acid tolerance. While streptococcal CSP was originally thought to act as a

classical quorum sensing signal, recent data in S. pneumoniae has shown that comCDE

expression can be altered under certain environmental stress conditions. Differences exist in the

timing and regulation of competence in S. pneumoniae and S. mutans. However, direct and

indirect evidence has tied CSP to the S. mutans stress response in the past. Although many

phenotypes have been attributed to CSP signalling in S. mutans, little is known about the

genetic pathways downstream of ComDE, nor about how the seemingly diverse CSP-regulated

phenotypes are connected. The general aim of this dissertation was to examine the

physiological and molecular response to the CSP pheromone in S. mutans, and determine the

contribution of phenotypes regulated by this pathway to the evolutionary fitness of the organism.

General hypothesis: S. mutans co-ordinates genetic competence and autolysis with its stress

response through the CSP peptide pheromone to acquire fitness-enhancing genes under stress

and build a stronger biofilm.

Primary objective: To understand how and when CSP-induced autolysis occurs in S. mutans,

and what role this process plays in the growth and genetic adaptability of the organism.

Rationale: Recent studies have suggested that competence may play a role in the stress

response in S. pneumoniae, and that fratricide may contribute DNA for the exchange of fitness

enhancing genes under stress. Although the competence cascades of S. pneumoniae and S.

mutans are physiologically divergent, S. mutans has been shown to regulate autolysis through

CSP accumulation. Moreover, the tightly packed oral biofilm community provides an excellent

environment for gene exchange, since spatial proximity, a „multi-cellular‟ altruistic lifestyle, and

Page 69: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

53

constant selective pressure from environmental stress would appear to favour the evolution of

such a strategy. Alternatively, the release of DNA into the biofilm environment via autolysis may

contribute to stress tolerance by adding to the extracellular matrix of the biofilm. Given that

biofilm formation and stress tolerance are vital to the virulence of S. mutans, understanding the

contribution of the primary intracellular communication system to their regulation is of utmost

importance. To attempt to elucidate the physiological and molecular mechanisms underlying

the CSP response in S. mutans, the goal of this dissertation is to investigate the following

specific aims:

Specific Aim 1: Determine if and how the CSP peptide pheromone participates in the

S. mutans stress response.

Specific Aim 2: Characterize the physiological response to CSP and determine the

signaling pathways involved in this response.

Specific Aim 3: Determine the role of autolysis in the physiology of S. mutans, with

emphasis on the competence cascade and biofilm formation.

Page 70: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

54

1.5 References

Aas, J.A., Paster, B.J., Stokes, L.N., Olsen, I., and Dewhirst, F.E. (2005) Defining the normal bacterial flora of the oral cavity. J Clin Microbiol 43: 5721-5732.

Aas, J.A., Griffen, A.L., Dardis, S.R., Lee, A.M., Olsen, I., Dewhirst, F.E., Leys, E.J., and Paster, B.J. (2008) Bacteria of dental caries in primary and permanent teeth in children and young adults. J Clin Microbiol 46: 1407-1417.

Ahn, S.-J., and Burne, R.A. (2007) Effects of Oxygen on Biofilm Formation and the AtlA Autolysin of Streptococcus mutans 189: 6293-6302.

Ahn, S.J., Wen, Z.T., and Burne, R.A. (2006) Multilevel control of competence development and stress tolerance in Streptococcus mutans UA159. Infect Immun 74: 1631-1642.

Ajdic, D., McShan, W.M., McLaughlin, R.E., Savic, G., Chang, J., Carson, M.B., Primeaux, C., Tian, R., Kenton, S., Jia, H., Lin, S., Qian, Y., Li, S., Zhu, H., Najar, F., Lai, H., White, J., Roe, B.A., and Ferretti, J.J. (2002) Genome sequence of Streptococcus mutans UA159, a cariogenic dental pathogen. Proc Natl Acad Sci U S A 99: 14434-14439.

Alaluusua, S., Takei, T., Ooshima, T., and Hamada, S. (1991) Mutacin activity of strains isolated from children with varying levels of mutants streptococci and caries. Arch Oral Biol 36: 251-255.

Allan, E., Hussain, H.A., Crawford, K.R., Miah, S., Ascott, Z.K., Khwaja, M.H., and Hosie, A.H.F. (2007) Genetic variation in comC, the gene encoding competence-stimulating peptide (CSP) in Streptococcus mutans. FEMS Microbiology Letters 268: 47-51.

Allesen-Holm, M., Barken, K.B., Yang, L., Klausen, M., Webb, J.S., Kjelleberg, S., Molin, S., Givskov, M., and Tolker-Nielsen, T. (2006) A characterization of DNA release in Pseudomonas aeruginosa cultures and biofilms. Mol Microbiol 59: 1114-1128.

Alloing, G., Martin, B., Granadel, C., and Claverys, J.P. (1998) Development of competence in Streptococcus pneumonaie: pheromone autoinduction and control of quorum sensing by the oligopeptide permease. Mol Microbiol 29: 75-83.

Aspiras, M.B., Ellen, R.P., and Cvitkovitch, D.G. (2004) ComX activity of Streptococcus mutans growing in biofilms. FEMS Microbiol Lett 238: 167-174.

Banas, J.A., and Vickerman, M.M. (2003) Glucan-binding proteins of the oral streptococci. Crit Rev Oral Biol Med 14: 89-99.

Bayles, K.W. (2007) The biological role of death and lysis in biofilm development. Nat Rev Microbiol 5: 721-726.

Berge, M., Moscoso, M., Prudhomme, M., Martin, B., and Claverys, J.-P. (2002) Uptake of transforming DNA in Gram-positive bacteria: a view from Streptococcus pneumoniae. Molecular Microbiology 45: 411-421.

Page 71: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

55

Berge, M., Mortier-Barriere, I., Martin, B., and Claverys, J.P. (2003) Transformation of Streptococcus pneumoniae relies on DprA- and RecA-dependent protection of incoming DNA single strands. Mol Microbiol 50: 527-536.

Bhagwat, S.P., Nary, J., and Burne, R.A. (2001) Effects of mutating putative two-component systems on biofilm formation by Streptococcus mutans UA159. FEMS Microbiol Lett 205: 225-230.

Birkhed, D., Rosell, K.G., and Granath, K. (1979) Structure of extracellular water-soluble polysaccharides synthesized from sucrose by oral strains of Streptococcus mutans, Streptococcus salivarius, Streptococcus sanguis and Actinomyces viscosus. Arch Oral Biol 24: 53-61.

Biswas, I., Drake, L., Erkina, D., and Biswas, S. (2008) Involvement of sensor kinases in the stress tolerance response of Streptococcus mutans. J Bacteriol 190: 68-77.

Bjedov, I., Tenaillon, O., Gerard, B., Souza, V., Denamur, E., Radman, M., Taddei, F., and Matic, I. (2003) Stress-Induced Mutagenesis in Bacteria. Science 300: 1404-1409.

Branda, S.S., Vik, A., Friedman, L., and Kolter, R. (2005) Biofilms: the matrix revisited. Trends in Microbiology 13: 20-26.

Burne, R.A., Chen, Y.Y., Wexler, D.L., Kuramitsu, H., and Bowen, W.H. (1996) Cariogenicity of Streptococcus mutans strains with defects in fructan metabolism assessed in a program-fed specific-pathogen-free rat model. J Dent Res 75: 1572-1577.

Claverys, J.-P., Prudhomme, M., and Martin, B. (2006) Induction of Competence Regulons as a General Response to Stress in Gram-Positive Bacteria. Annual Review of Microbiology 60: 451-475.

Claverys, J.-P., Martin, B., and Havarstein, L.S. (2007) Competence-induced fratricide in streptococci. Molecular Microbiology 64: 1423-1433.

Claverys, J.P., and Havarstein, L.S. (2007) Cannibalism and fratricide: mechanisms and raisons d'etre. Nat Rev Microbiol 5: 219-229.

Claverys, J.P., Martin, B., and Polard, P. (2009) The genetic transformation machinery: composition, localization, and mechanism. FEMS Microbiol Rev 33: 643-656.

Costerton, J.W., Lewandowski, Z., DeBeer, D., Caldwell, D., Korber, D., and James, G. (1994) Biofilms, the customized microniche. J Bacteriol 176: 2137-2142.

Costerton, J.W., Stewart, P.S., and Greenberg, E.P. (1999) Bacterial biofilms: a common cause of persistent infections. Science 284: 1318-1322.

Cotter, P.D., Hill, C., and Ross, R.P. (2005) Bacteriocins: developing innate immunity for food. Nat Rev Microbiol 3: 777-788.

Craig, L., and Li, J. (2008) Type IV pili: paradoxes in form and function. Curr Opin Struct Biol 18: 267-277.

Cvitkovitch, D.G. (2001) Genetic competence and transformation in oral streptococci. Crit Rev Oral Biol Med 12: 217-243.

Page 72: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

56

Dagkessamanskaia, A., Moscoso, M., Henard, V., Guiral, S., Overweg, K., Reuter, M., Martin, B., Wells, J., and Claverys, J.P. (2004) Interconnection of competence, stress and CiaR regulons in Streptococcus pneumoniae: competence triggers stationary phase autolysis of ciaR mutant cells. Mol Microbiol 51: 1071-1086.

Davies, D.G., Chakrabarty, A.M., and Geesey, G.G. (1993) Exopolysaccharide production in biofilms: substratum activation of alginate gene expression by Pseudomonas aeruginosa. Appl Environ Microbiol 59: 1181-1186.

Davies, D.G., Parsek, M.R., Pearson, J.P., Iglewski, B.H., Costerton, J.W., and Greenberg, E.P. (1998) The involvement of cell-to-cell signals in the development of a bacterial biofilm. Science 280: 295-298.

de Saizieu, A., Gardes, C., Flint, N., Wagner, C., Kamber, M., Mitchell, T.J., Keck, W., Amrein, K.E., and Lange, R. (2000) Microarray-based identification of a novel Streptococcus pneumoniae regulon controlled by an autoinduced peptide. J Bacteriol 182: 4696-4703.

Demuth, D.R., and Irvine, D.C. (2002) Structural and functional variation within the alanine-rich repetitive domain of streptococcal antigen I/II. Infect Immun 70: 6389-6398.

Diep, D.B., Skaugen, M., Salehian, Z., Holo, H., and Nes, I.F. (2007) Common mechanisms of target cell recognition and immunity for class II bacteriocins. PNAS 104: 2384-2389.

Dowson, C.G., Barcus, V., King, S., Pickerill, P., Whatmore, A., and Yeo, M. (1997) Horizontal gene transfer and the evolution of resistance and virulence determinants in Streptococcus. Soc Appl Bacteriol Symp Ser 26: 42S-51S.

Drider, D., Fimland, G., Hechard, Y., McMullen, L.M., and Prevost, H. (2006) The Continuing Story of Class IIa Bacteriocins. Microbiol. Mol. Biol. Rev. 70: 564-582.

Dubnau, D. (1999) DNA uptake in bacteria. Annual Review of Microbiology 53: 217-244.

Dubnau, D., and Losick, R. (2006) Bistability in bacteria. Mol Microbiol 61: 564-572.

Dunne, W.M., Jr. (2002) Bacterial Adhesion: Seen Any Good Biofilms Lately? Clin Microbiol Rev 15: 155-166.

Eberhard, A., Burlingame, A.L., Eberhard, C., Kenyon, G.L., Nealson, K.H., and Oppenheimer, N.J. (1981) Structural identification of autoinducer of Photobacterium fischeri luciferase. Biochemistry 20: 2444-2449.

Eldholm, V., Johnsborg, O., Haugen, K., Ohnstad, H.S., and Havarstein, L.S. (2009) Fratricide in Streptococcus pneumoniae: contributions and role of the cell wall hydrolases CbpD, LytA and LytC. Microbiology 155: 2223-2234.

Ephrussi-Taylor, H. (1960) The status of the transforming DNA during the 1st phases of bacterial transformation. C R Seances Soc Biol Fil 154: 1951-1955.

Finkel, S.E., and Kolter, R. (2001) DNA as a nutrient: novel role for bacterial competence gene homologs. J Bacteriol 183: 6288-6293.

Fux, C.A., Costerton, J.W., Stewart, P.S., and Stoodley, P. (2005) Survival strategies of infectious biofilms. Trends in Microbiology 13: 34-40.

Page 73: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

57

Groicher, K.H., Firek, B.A., Fujimoto, D.F., and Bayles, K.W. (2000) The Staphylococcus aureus lrgAB operon modulates murein hydrolase activity and penicillin tolerance. J Bacteriol 182: 1794-1801.

Gronroos, L., Saarela, M., Matto, J., Tanner-Salo, U., Vuorela, A., and Alaluusua, S. (1998) Mutacin production by Streptococcus mutans may promote transmission of bacteria from mother to child. Infect Immun 66: 2595-2600.

Guiral, S., Mitchell, T.J., Martin, B., and Claverys, J.P. (2005) Competence-programmed predation of noncompetent cells in the human pathogen Streptococcus pneumoniae: genetic requirements. Proc Natl Acad Sci U S A 102: 8710-8715.

Hall-Stoodley, L., Nistico, L., Sambanthamoorthy, K., Dice, B., Nguyen, D., Mershon, W.J., Johnson, C., Hu, F.Z., Stoodley, P., Ehrlich, G.D., and Post, J.C. (2008) Characterization of biofilm matrix, degradation by DNase treatment and evidence of capsule downregulation in Streptococcus pneumoniae clinical isolates. BMC Microbiol 8: 173.

Håvarstein, L.S., Coomaraswamy, G., and Morrison, D.A. (1995) An unmodified heptadecapeptide pheromone induces competence for genetic transformation in Streptococcus pneumoniae. Proc Natl Acad Sci U S A 92: 11140-11144.

Håvarstein, L.S., Gaustad, P., Nes, I.F., and Morrison, D.A. (1996) Identification of the streptococcal competence-pheromone receptor. Mol Microbiol 21: 863-869.

Håvarstein, L.S., Hakenbeck, R., and Gaustad, P. (1997) Natural competence in the genus Streptococcus: evidence that streptococci can change pherotype by interspecies recombinational exchanges. J Bacteriol 179: 6589-6594.

Håvarstein, L.S., Martin, B., Johnsborg, O., Granadel, C., and Claverys, J.P. (2006) New insights into the pneumococcal fratricide: relationship to clumping and identification of a novel immunity factor. Mol Microbiol 59: 1297-1307.

Hechard, Y., and Sahl, H.G. (2002) Mode of action of modified and unmodified bacteriocins from Gram-positive bacteria. Biochimie 84: 545-557.

Heilmann, C., Hussain, M., Peters, G., and Gotz, F. (1997) Evidence for autolysin-mediated primary attachment of Staphylococcus epidermidis to a polystyrene surface. Mol Microbiol 24: 1013-1024.

Heilmann, C., Hartleib, J., Hussain, M.S., and Peters, G. (2005) The multifunctional Staphylococcus aureus autolysin aaa mediates adherence to immobilized fibrinogen and fibronectin. Infect Immun 73: 4793-4802.

Holtje, J.V., and Tuomanen, E.I. (1991) The murein hydrolases of Escherichia coli: properties, functions and impact on the course of infections in vivo. J Gen Microbiol 137: 441-454.

Hughes, D.T., and Sperandio, V. (2008) Inter-kingdom signalling: communication between bacteria and their hosts. Nat Rev Micro 6: 111-120.

Jenkinson, H.F., and Demuth, D.R. (1997) Structure, function and immunogenicity of streptococcal antigen I/II polypeptides. Mol Microbiol 23: 183-190.

Johnsborg, O., Eldholm, V., and Havarstein, L.S. (2007) Natural genetic transformation: prevalence, mechanisms and function. Res Microbiol 158: 767-778.

Page 74: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

58

Kamiya, R.U., Napimoga, M.H., Rosa, R.T., Hofling, J.F., and Goncalves, R.B. (2005) Mutacin production in Streptococcus mutans genotypes isolated from caries-affected and caries-free individuals. Oral Microbiol Immunol 20: 20-24.

Kausmally, L., Johnsborg, O., Lunde, M., Knutsen, E., and Havarstein, L.S. (2005) Choline-binding protein D (CbpD) in Streptococcus pneumoniae is essential for competence-induced cell lysis. J Bacteriol 187: 4338-4345.

Kawada-Matsuo, M., Shibata, Y., and Yamashita, Y. (2009) Role of two component signaling response regulators in acid tolerance of Streptococcus mutans. Oral Microbiol Immunol 24: 173-176.

Keller, L., and Surette, M.G. (2006) Communication in bacteria: an ecological and evolutionary perspective. Nat Rev Microbiol 4: 249-258.

Kinder, S.A., and Holt, S.C. (1994) Coaggregation between bacterial species. Methods Enzymol 236: 254-270.

Kolenbrander, P.E., Palmer, R.J., Jr., Rickard, A.H., Jakubovics, N.S., Chalmers, N.I., and Diaz, P.I. (2006) Bacterial interactions and successions during plaque development. Periodontol 2000 42: 47-79.

Kreth, J., Merritt, J., Shi, W., and Qi, F. (2005) Co-ordinated bacteriocin production and competence development: a possible mechanism for taking up DNA from neighbouring species. Mol Microbiol 57: 392-404.

Kreth, J., Merritt, J., Zhu, L., Shi, W., and Qi, F. (2006b) Cell density- and ComE-dependent expression of a group of mutacin and mutacin-like genes in Streptococcus mutans. FEMS Microbiology Letters 265: 11-17.

Kuramitsu, H.K., He, X., Lux, R., Anderson, M.H., and Shi, W. (2007) Interspecies interactions within oral microbial communities. Microbiol Mol Biol Rev 71: 653-670.

Lawrence, J.R., Korber, D.R., Hoyle, B.D., Costerton, J.W., and Caldwell, D.E. (1991) Optical sectioning of microbial biofilms. J Bacteriol 173: 6558-6567.

Lee, M.S., and Morrison, D.A. (1999) Identification of a new regulator in Streptococcus pneumoniae linking quorum sensing to competence for genetic transformation. J Bacteriol 181: 5004-5016.

Lemos, J.A., Abranches, J., and Burne, R.A. (2005) Responses of cariogenic streptococci to environmental stresses. Curr Issues Mol Biol 7: 95-107.

Lemos, J.A., Lin, V.K., Nascimento, M.M., Abranches, J., and Burne, R.A. (2007) Three gene products govern (p)ppGpp production by Streptococcus mutans. Mol Microbiol 65: 1568-1581.

Lemos, J.A., and Burne, R.A. (2008) A model of efficiency: stress tolerance by Streptococcus mutans. Microbiology 154: 3247-3255.

Levesque, C.M., Mair, R.W., Perry, J.A., Lau, P.C.Y., Li, Y.-H., and Cvitkovitch, D.G. (2007) Systemic inactivation and phenotypic characterization of two-component systems in expression of Streptococcus mutans virulence properties. Letters in Applied Microbiology 45: 398-404.

Page 75: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

59

Lewis, K. (2005) Persister cells and the riddle of biofilm survival. Biochemistry (Mosc) 70: 267-274.

Li, Y.H., Hanna, M.N., Svensater, G., Ellen, R.P., and Cvitkovitch, D.G. (2001a) Cell density modulates acid adaptation in Streptococcus mutans: implications for survival in biofilms. J Bacteriol 183: 6875-6884.

Li, Y.H., Lau, P.C., Lee, J.H., Ellen, R.P., and Cvitkovitch, D.G. (2001b) Natural genetic transformation of Streptococcus mutans growing in biofilms. J Bacteriol 183: 897-908.

Li, Y.H., Lau, P.C., Tang, N., Svensater, G., Ellen, R.P., and Cvitkovitch, D.G. (2002a) Novel two-component regulatory system involved in biofilm formation and acid resistance in Streptococcus mutans. J Bacteriol 184: 6333-6342.

Li, Y.H., Tang, N., Aspiras, M.B., Lau, P.C., Lee, J.H., Ellen, R.P., and Cvitkovitch, D.G. (2002b) A quorum-sensing signaling system essential for genetic competence in Streptococcus mutans is involved in biofilm formation. J Bacteriol 184: 2699-2708.

Linnett, P.E., and Strominger, J.L. (1973) Additional antibiotic inhibitors of peptidoglycan synthesis. Antimicrob Agents Chemother 4: 231-236.

Loesche, W.J. (1986) Role of Streptococcus mutans in human dental decay. Microbiol Rev 50: 353-380.

Longo, P.L., Mattos-Graner, R.O., and Mayer, M.P. (2003) Determination of mutacin activity and detection of mutA genes in Streptococcus mutans genotypes from caries-free and caries-active children. Oral Microbiol Immunol 18: 144-149.

Luo, P., and Morrison, D.A. (2003) Transient association of an alternative sigma factor, ComX, with RNA polymerase during the period of competence for genetic transformation in Streptococcus pneumoniae. J Bacteriol 185: 349-358.

Luo, P., Li, H., and Morrison, D.A. (2004) Identification of ComW as a new component in the regulation of genetic transformation in Streptococcus pneumoniae. Mol Microbiol 54: 172-183.

Martin, B., Quentin, Y., Fichant, G., and Claverys, J.-P. (2006) Independent evolution of competence regulatory cascades in streptococci? Trends in Microbiology 14: 339-345.

McLaughlin, R.E., and Ferretti, J.J. (1996) The multiple-sugar metabolism (msm) gene cluster of Streptococcus mutans is transcribed as a single operon. FEMS Microbiology Letters 140: 261-264.

Mejean, V., and Claverys, J.P. (1993) DNA processing during entry in transformation of Streptococcus pneumoniae. J Biol Chem 268: 5594-5599.

Miao, R., and Guild, W.R. (1970) Competent Diplococcus pneumoniae accept both single- and double-stranded deoxyribonucleic acid. J Bacteriol 101: 361-364.

Miller, M.B., and Bassler, B.L. (2001) Quorum sensing in bacteria. Annu Rev Microbiol 55: 165-199.

Mitchell, T.J. (2003) The pathogenesis of streptococcal infections: from tooth decay to meningitis. Nat Rev Microbiol 1: 219-230.

Page 76: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

60

Mongold, J.A. (1992) DNA repair and the evolution of transformation in Haemophilus influenzae. Genetics 132: 893-898.

Morrison, D.A. (1977) Transformation in pneumococcus: existence and properties of a complex involving donor deoxyribonucleate single strands in eclipse. J Bacteriol 132: 576-583.

Morrison, D.A. (1978) Transformation in pneumococcus: protein content of eclipse complex. J Bacteriol 136: 548-557.

Morrison, D.A., and Mannarelli, B. (1979) Transformation in pneumococcus: nuclease resistance of deoxyribonucleic acid in the eclipse complex. J Bacteriol 140: 655-665.

Morrison, D.A., Mortier-Barriere, I., Attaiech, L., and Claverys, J.P. (2007) Identification of the major protein component of the pneumococcal eclipse complex. J Bacteriol 189: 6497-6500.

Mortier-Barriere, I., Velten, M., Dupaigne, P., Mirouze, N., Pietrement, O., McGovern, S., Fichant, G., Martin, B., Noirot, P., Le Cam, E., Polard, P., and Claverys, J.P. (2007) A key presynaptic role in transformation for a widespread bacterial protein: DprA conveys incoming ssDNA to RecA. Cell 130: 824-836.

Moscoso, M., and Claverys, J.P. (2004) Release of DNA into the medium by competent Streptococcus pneumoniae: kinetics, mechanism and stability of the liberated DNA. Mol Microbiol 54: 783-794.

Munro, C., Michalek, S.M., and Macrina, F.L. (1991) Cariogenicity of Streptococcus mutans V403 glucosyltransferase and fructosyltransferase mutants constructed by allelic exchange. Infect Immun 59: 2316-2323.

Napimoga, M.H., Hofling, J.F., Klein, M.I., Kamiya, R.U., and Goncalves, R.B. (2005) Tansmission, diversity and virulence factors of Streptococcus mutans genotypes. J Oral Sci 47: 59-64.

Nealson, K.H., Platt, T., and Hastings, J.W. (1970) Cellular control of the synthesis and activity of the bacterial luminescent system. J Bacteriol 104: 313-322.

Nes, I.F., Diep, D.B., and Holo, H. (2007) Bacteriocin diversity in Streptococcus and Enterococcus. J Bacteriol 189: 1189-1198.

Nomura, R., Nakano, K., Nemoto, H., Fujita, K., Inagaki, S., Takahashi, T., Taniguchi, K., Takeda, M., Yoshioka, H., Amano, A., and Ooshima, T. (2006) Isolation and characterization of Streptococcus mutans in heart valve and dental plaque specimens from a patient with infective endocarditis. J Med Microbiol 55: 1135-1140.

Oggioni, M.R., Trappetti, C., Kadioglu, A., Cassone, M., Iannelli, F., Ricci, S., Andrew, P.W., and Pozzi, G. (2006) Switch from planktonic to sessile life: a major event in pneumococcal pathogenesis 61: 1196-1210.

Pakula, R., and Walczak, W. (1963) On the nature of competence of transformable streptococci. J Gen Microbiol 31: 125-133.

Parsek, M.R., and Greenberg, E.P. (2005) Sociomicrobiology: the connections between quorum sensing and biofilms. Trends in Microbiology 13: 27-33.

Page 77: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

61

Pestova, E.V., Havarstein, L.S., and Morrison, D.A. (1996) Regulation of competence for genetic transformation in Streptococcus pneumoniae by an auto-induced peptide pheromone and a two-component regulatory system. Mol Microbiol 21: 853-862.

Petersen, F.C., Tao, L., and Scheie, A.A. (2005) DNA binding-uptake system: a link between cell-to-cell communication and biofilm formation. J Bacteriol 187: 4392-4400.

Peterson, S., Cline, R.T., Tettelin, H., Sharov, V., and Morrison, D.A. (2000) Gene expression analysis of the Streptococcus pneumoniae competence regulons by use of DNA microarrays. J Bacteriol 182: 6192-6202.

Peterson, S.N., Sung, C.K., Cline, R., Desai, B.V., Snesrud, E.C., Luo, P., Walling, J., Li, H., Mintz, M., Tsegaye, G., Burr, P.C., Do, Y., Ahn, S., Gilbert, J., Fleischmann, R.D., and Morrison, D.A. (2004) Identification of competence pheromone responsive genes in Streptococcus pneumoniae by use of DNA microarrays. Mol Microbiol 51: 1051-1070.

Pinas, G.E., Cortes, P.R., Orio, A.G., and Echenique, J. (2008) Acidic stress induces autolysis by a CSP-independent ComE pathway in Streptococcus pneumoniae. Microbiology 154: 1300-1308.

Piotrowski, A., Luo, P., and Morrison, D.A. (2009) Competence for Genetic Transformation in Streptococcus pneumoniae: Termination of Activity of the Alternative Sigma Factor ComX Is Independent of Proteolysis of ComX and ComW 191: 3359-3366.

Pozzi, G., Masala, L., Iannelli, F., Manganelli, R., Havarstein, L.S., Piccoli, L., Simon, D., and Morrison, D.A. (1996) Competence for genetic transformation in encapsulated strains of Streptococcus pneumoniae: two allelic variants of the peptide pheromone. J Bacteriol 178: 6087-6090.

Prudhomme, M., Attaiech, L., Sanchez, G., Martin, B., and Claverys, J.-P. (2006) Antibiotic Stress Induces Genetic Transformability in the Human Pathogen Streptococcus pneumoniae. Science 313: 89-92.

Qi, F., Merritt, J., Lux, R., and Shi, W. (2004) Inactivation of the ciaH Gene in Streptococcus mutans diminishes mutacin production and competence development, alters sucrose-dependent biofilm formation, and reduces stress tolerance. Infect Immun 72: 4895-4899.

Qi, F., Kreth, J., Levesque, C.M., Kay, O., Mair, R.W., Shi, W., Cvitkovitch, D.G., and Goodman, S.D. (2005) Peptide pheromone induced cell death of Streptococcus mutans. FEMS Microbiol Lett 251: 321-326.

Quivey, R.G., Kuhnert, W.L., and Hahn, K. (2001) Genetics of acid adaptation in oral streptococci. Crit Rev Oral Biol Med 12: 301-314.

Rice, K.C., Firek, B.A., Nelson, J.B., Yang, S.J., Patton, T.G., and Bayles, K.W. (2003) The Staphylococcus aureus cidAB operon: evaluation of its role in regulation of murein hydrolase activity and penicillin tolerance. J Bacteriol 185: 2635-2643.

Rice, K.C., Mann, E.E., Endres, J.L., Weiss, E.C., Cassat, J.E., Smeltzer, M.S., and Bayles, K.W. (2007) The cidA murein hydrolase regulator contributes to DNA release and biofilm development in Staphylococcus aureus. Proc Natl Acad Sci 104: 8113-8118.

Page 78: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

62

Rice, K.C., and Bayles, K.W. (2008) Molecular Control of Bacterial Death and Lysis. Microbiol. Mol. Biol. Rev. 72: 85-109.

Rickard, A.H., Palmer, R.J., Jr., Blehert, D.S., Campagna, S.R., Semmelhack, M.F., Egland, P.G., Bassler, B.L., and Kolenbrander, P.E. (2006) Autoinducer 2: a concentration-dependent signal for mutualistic bacterial biofilm growth. Mol Microbiol 60: 1446-1456.

Rickard, A.H., Campagna, S.R., and Kolenbrander, P.E. (2008b) Autoinducer-2 is produced in saliva-fed flow conditions relevant to natural oral biofilms. J Appl Microbiol 105: 2096-2103.

Rickard, A.H., Bachrach, G., Davies, D. G. (2008a) Cell-Cell Communication in Oral Microbial Communities. In Molecular Oral Microbiology. Rogers, A. (ed): Caister Academic Press, pp. 87-103.

Ronda, C., Garcia, J.L., Garcia, E., Sanchez-Puelles, J.M., and Lopez, R. (1987) Biological role of the pneumococcal amidase. Cloning of the lytA gene in Streptococcus pneumoniae. Eur J Biochem 164: 621-624.

Sauer, K., Camper, A.K., Ehrlich, G.D., Costerton, J.W., and Davies, D.G. (2002) Pseudomonas aeruginosa displays multiple phenotypes during development as a biofilm. J Bacteriol 184: 1140-1154.

Senadheera, M.D., Lee, A.W., Hung, D.C., Spatafora, G.A., Goodman, S.D., and Cvitkovitch, D.G. (2006) The Streptococcus mutans vicX gene product modulates gtfB/C Expression, Biofilm Formation, Genetic Competence and Oxidative Stress Tolerance. J Bacteriol.

Senadheera, M.D., Levesque, C. M., Cvitkovitch, D. G. (2005) Cell-density-dependent regulation of streptococcal competence. In Bacterial Cell-to-Cell Communication: Role in Virulence and Pathogenesis. Demuth, D.R., Lamont, R. J. (ed). Cambridge: Cambridge University Press, pp. 233-267.

Shockman, G.D., Daneo-Moore, L., Kariyama, R., and Massidda, O. (1996) Bacterial walls, peptidoglycan hydrolases, autolysins, and autolysis. Microb Drug Resist 2: 95-98.

Stewart, P.S., and Costerton, W.J. (2001) Antibiotic resistance of bacteria in biofilms. The Lancet 358: 135-138.

Stock, A.M., Robinson, V.L., and Goudreau, P.N. (2000) Two-component signal transduction. Annu Rev Biochem 69: 183-215.

Stoodley, P., Sauer, K., Davies, D.G., and Costerton, J.W. (2002) Biofilms as complex differentiated communities. Annu Rev Microbiol 56: 187-209.

Sturme, M.H., Kleerebezem, M., Nakayama, J., Akkermans, A.D., Vaugha, E.E., and de Vos, W.M. (2002) Cell to cell communication by autoinducing peptides in gram-positive bacteria. Antonie Van Leeuwenhoek 81: 233-243.

Sung, C.K., and Morrison, D.A. (2005) Two distinct functions of ComW in stabilization and activation of the alternative sigma factor ComX in Streptococcus pneumoniae. J Bacteriol 187: 3052-3061.

Page 79: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

63

Surette, M.G., Miller, M.B., and Bassler, B.L. (1999) Quorum sensing in Escherichia coli, Salmonella typhimurium, and Vibrio harveyi: a new family of genes responsible for autoinducer production. Proc Natl Acad Sci U S A 96: 1639-1644.

Syvitski, R.T., Tian, X.L., Sampara, K., Salman, A., Lee, S.F., Jakeman, D.L., and Li, Y.H. (2007) Structure-activity analysis of quorum-sensing signaling peptides from Streptococcus mutans. J Bacteriol 189: 1441-1450.

Thiel, V., Vilchez, R., Sztajer, H., Wagner-Dobler, I., and Schulz, S. (2009) Identification, quantification, and determination of the absolute configuration of the bacterial quorum-sensing signal autoinducer-2 by gas chromatography-mass spectrometry. Chembiochem 10: 479-485.

Thomas, V.C., Hiromasa, Y., Harms, N., Thurlow, L., Tomich, J., and Hancock, L. (2009) A fratricidal mechanism is responsible for eDNA release and contributes to biofilm development of Enterococcus faecalis. Mol Microbiol.

Tomasz, A., and Hotchkiss, R.D. (1964) Regulation of the Transformability of Pheumococcal Cultures by Macromolecular Cell Products. Proc Natl Acad Sci U S A 51: 480-487.

van der Ploeg, J.R. (2005) Regulation of bacteriocin production in Streptococcus mutans by the quorum-sensing system required for development of genetic competence. J Bacteriol 187: 3980-3989.

Vendeville, A., Winzer, K., Heurlier, K., Tang, C.M., and Hardie, K.R. (2005) Making 'sense' of metabolism: autoinducer-2, LuxS and pathogenic bacteria. Nat Rev Microbiol 3: 383-396.

Waters, C.M., and Bassler, B.L. (2005) Quorum sensing: cell-to-cell communication in bacteria. Annu Rev Cell Dev Biol 21: 319-346.

Ween, O., Gaustad, P., and Havarstein, L.S. (1999) Identification of DNA binding sites for ComE, a key regulator of natural competence in Streptococcus pneumoniae. Mol Microbiol 33: 817-827.

Whatmore, A.M., Barcus, V.A., and Dowson, C.G. (1999) Genetic diversity of the streptococcal competence (com) gene locus. J Bacteriol 181: 3144-3154.

Whitchurch, C.B., Tolker-Nielsen, T., Ragas, P.C., and Mattick, J.S. (2002) Extracellular DNA required for bacterial biofilm formation. Science 295: 1487.

Winans, S.C., and Zhu, J. (2000) Roles of Cell-Cell Communication in Confronting the Limitations and Opportunities of High Population Density. In Bacterial Stress Responses. Storz, G. and Hengge-Aronis, R. (eds). Washington, D.C.: ASM Press, pp. 261-272.

Xavier, K.B., and Bassler, B.L. (2003) LuxS quorum sensing: more than just a numbers game. Curr Opin Microbiol 6: 191-197.

Yamashita, Y., Bowen, W.H., Burne, R.A., and Kuramitsu, H.K. (1993) Role of the Streptococcus mutans gtf genes in caries induction in the specific-pathogen-free rat model. Infect Immun 61: 3811-3817.

Zhang, K., Ou, M., Wang, W., and Ling, J. (2009) Effects of quorum sensing on cell viability in Streptococcus mutans biofilm formation. Biochem Biophys Res Commun 379: 933-938.

Page 80: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

64

Chapter 2: Peptide alarmone signalling triggers an auto-active

bacteriocin necessary for genetic competence

JA Perry, MB Jones, SN Peterson, DG Cvitkovitch, and CM Lévesque. 2009. Mol

Micro. 72: 905-917

Page 81: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

65

2.1 Abstract

The induction of genetic competence is a strategy used by bacteria to increase their genetic

repertoire under stressful environmental conditions. Recently, Streptococcus pneumoniae has

been shown to coordinate the uptake of transforming DNA with fratricide via increased

expression of the peptide pheromone responsible for competence induction. Here we document

that environmental stress induced expression of the peptide pheromone CSP in the oral

pathogen Streptococcus mutans. We showed that CSP is involved in the stress response, and

determined the CSP-induced regulon in S. mutans by microarray analysis. Contrary to

pneumococcus, S. mutans responds to increased concentrations of CSP by cell lysis in only a

fraction of the population. We have focused on the mechanism of cell lysis, and have identified

a novel bacteriocin as the „death effector‟. Most importantly, we showed that this bacteriocin

causes cell death via a novel mechanism of action: intracellular action against self. We have

also identified the cognate bacteriocin immunity protein, which resides in a separate unlinked

genetic locus to allow its differential regulation. The role of the lytic response in S. mutans

competence is also discussed. Together, these findings reveal a novel autolytic pathway in

S. mutans which may be involved in the dissemination of fitness-enhancing genes in the oral

biofilm.

Page 82: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

66

2.2 Introduction

Free-living bacteria are at the mercy of a variety of environmental stress conditions that

impose constant selective pressure on the microorganism. To compete or simply survive in their

ecological niche, bacteria must rely on an ability to sense and respond to stress. Often, the

response to stress involves the induction of a transient state of hyper-mutability, which is argued

to increase the probability of generating adaptive variants in the bacterial population (Bjedov et

al., 2003). Although some debate exists as to whether mutagenesis is an inductive strategy or

simply a by-product of the accumulation of DNA lesions, stress-induced mutations certainly

participate in the adaptive evolution of bacteria (Bjedov et al., 2003).

Although an increased mutation rate may lead to the chance development of a fitness-

enhancing phenotype, the probability of such an event occurring is limited to the available DNA

sequence in an organism‟s own genome. However, naturally transformable bacteria are able to

sample the DNA pool of an entire community during stress, and acquire fitness enhancing

genes across species barriers. The major human pathogen Streptococcus pneumoniae and the

soil-dweller Bacillus subtilis are the best-characterized naturally transformable Gram-positive

bacteria. Although they employ different mechanisms to achieve the competent state, both

organisms turn on their competence regulons in response to specific environmental stresses,

which may improve fitness by generating genetic diversity through natural transformation

(Claverys et al., 2006).

As the major etiological agent of human dental caries (Mitchell, 2003), the naturally

transformable oral bacterium Streptococcus mutans is well studied at the genetic and

physiological level. The regulatory system that governs genetic competence in this species is

homologous to the system in S. pneumoniae ((Håvarstein et al., 1996)), and is composed of a

peptide pheromone (competence stimulating peptide, or CSP), the ComDE two-component

Page 83: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

67

signal transduction system, and the alternate sigma factor ComX (Li et al., 2001b). Although

competence is regulated by similar signaling systems in both streptococcal species, important

differences separate the two species‟ response to the pheromone. Firstly, while competence

develops uniformly across a population of S. pneumoniae (Håvarstein et al., 2006), it is well

established that only a fraction of the S. mutans population (~1%) ever becomes CSP-

responsive (Aspiras et al., 2004; Li et al., 2001b; Qi et al., 2005). Moreover, the competence

cascade in S. mutans is known to incorporate inputs from additional two-component systems

(Ahn et al., 2006; Perry et al., 2008; Qi et al., 2004; Senadheera et al., 2005). Finally, while

S. pneumoniae controls expression of its bacteriocins through the dedicated BlpRH system (de

Saizieu et al., 2000), S. mutans controls the expression of many of its bacteriocins through

ComDE (Hale et al., 2005a; Kreth et al., 2005; Kreth et al., 2006a; Kreth et al., 2007; van der

Ploeg, 2005). The co-ordination of bacteriocin production and competence suggests that

S. mutans can generate DNA for uptake from lysis of neighboring species (Kreth et al., 2005), in

what may be an evolutionary adaptation to the multi-species oral biofilm environment.

Streptococcal CSP pheromone was originally thought to accumulate passively in proportion

to population density, and act as a classical quorum sensing signal to activate the competence

regulon at a specific cell density (Håvarstein et al., 1995; Håvarstein et al., 1996; Li et al.,

2001b). However, early work done in S. mutans (Li et al., 2001a; Li et al., 2002a) suggested an

intimate link between the competence cascade and the organism‟s response to acid stress. A

link between competence and oxidative stress has also been made in S. mutans (Senadheera

et al., 2006; Wen et al., 2005), but a mechanistic explanation for these phenotypes has

remained elusive. Evidence for stress-induced genetic plasticity has also accumulated in regard

to S. pneumoniae (Chastanet et al., 2001; Claverys et al., 2000; Prudhomme et al., 2006),

where it has been suggested that pneumococcal CSP may act as a secreted stress-induced

Page 84: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

68

pheromone (or „alarmone‟) that triggers expression of stress-responsive genes (Claverys et al.,

2006).

Here we present evidence that S. mutans integrates its response to specific environmental

stresses with its competence cascade via the CSP pheromone, and describe for the first time

the global transcriptome analyses of CSP-regulated genes in S. mutans. Our most important

finding was that in the presence of high concentrations of CSP pheromone, the unprocessed

form of mutacin V acted as an intracellular auto-active bacteriocin causing S. mutans autolysis.

To our knowledge, this is a completely novel mechanism of action for a bacteriocin. Moreover,

the impaired ability of S. mutans cells lacking mutacin V to become competent indicates that

stress-induced lysis in a subpopulation may be required for the acquisition of diversity through

genetic transformation in the surviving cells.

2.3 Experimental procedures

Culture conditions

The S. mutans strains used in this study are listed in Table 2.1. Mutants were constructed in

S. mutans UA159 wild-type as described previously (Lau et al., 2002). Strains were grown in

Todd-Hewitt–Yeast Extract (THYE) broth at 37ºC with 5% CO2. Growth was monitored using a

microbiology workstation (Bioscreen C Labsystems, Finland). Co-culture experiments were

conducted by adding equal volumes of each strain, and CFUs were enumerated by plating.

Viability staining was performed using the LIVE/DEAD BacLight kit (Invitrogen) according to the

manufacturer‟s directions. Lysis was assessed by harvesting the supernatant of cultures

expressing a -glucuronidase (GUS) reporter gene cloned into a theta-replicating plasmid

(Biswas et al., 2008) in the absence and presence of 2 μM sCSP. Supernatants were combined

in equal parts with 2 GUS buffer (100 mM Na2HPO4, 20 mM -mercaptoethanol, 2 mM EDTA,

Page 85: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

69

0.2% Triton-X, 1 mM PNPG substrate (Sigma)). Absorbance at 420 nm was measured after 15

min of color development. GUS activity was expressed as [1000 A420]/[time (min) OD600] in

Miller units (MU).

Table 2.1. Bacterial strains used in this study

Strain Description* Reference

S. mutans UA159 Wild-type ATCC ∆comC ∆smu.1915; Emr This work ∆comC complemented ∆smu.1915(pcomC); Emr, Spcr This work ∆comE ∆smu.1917; Emr This work ∆comDE ∆smu.1916-smu.1917; Emr This work ∆comX ∆smu.1997; Emr This work ∆mutacin IV ∆smu.150-smu.151; Emr This work ∆cipB ∆smu.1914; Emr This work ∆cipI ∆smu.925; Emr This work ∆1913 ∆smu.1913; Emr This work ∆423 ∆smu.423; Emr This work ∆1906 ∆smu.1906; Emr This work ∆nlmTE ∆smu.286-smu.287; Emr Hale et al., 2005 ∆luxS ∆smu.474; Emr Sztajer et al., 2008 UA159(pIB187) Plasmid with the gusA reporter gene under

the constitutive control Biswas et al., 2008

UA159(PcomX–gfp) PcomX–gfp fusion into pDL277; Spcr Aspiras et al., 2004 UA159(pDL277) pDL277; Spcr This work UA159(Pmsm–1914) Pmsm–1914 into pDL277; Spcr This work ∆cipI(Pmsm–1914) Pmsm–1914 into pDL277; Emr, Spcr This work UA159(p925) Smu.925 into pDL277; Spcr This work

E. coli BL21(pET28a(+)) T7 expression vector, non-expression

host; Kanr Novagen

BL213DE3 (pET28a(+)) T7 expression vector, expression host; Kanr

Novagen

BL21(His6-fullCipB) CipB precursor cloned into pET28a(+), non-expression host; Kanr

This work

BL21(His6-GGCipB) Mature form of CipB cloned into pET28a(+),non-expression host; Kanr

This work

BL21DE3(His6-fullCipB) CipB precursor cloned into pET28a(+), expression host; Kanr

This work

BL21DE3(His6-GGCipB) Mature form of CipB cloned into pET28a(+), expression host; Kanr

This work

L. lactis I6 Indicator strain, susceptible to CipB Hale et al., 2005 S. salivarius 25975 Wild-type M. Frenette, U. Laval S. thermophilus LMG18311 Wild-type S. Moineau, U. Laval

*Emr, erythromycin resistance; Spcr, spectinomycin resistance; Kanr, kanamycin resistance

Page 86: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

70

Gene expression analysis

Transcriptional analysis of comC in environmentally stress-induced UA159 cells was conducted

by real-time reverse transcription (RT)-PCR. Cells were grown in THYE until an OD600 of ~ 0.4

(mid-log phase) was reached, and an aliquot was reserved (pre-stress control). Cells were then

re-suspended in fresh THYE and exposed for 30 min at 37ºC to the following stresses: acid

shock (THYE acidified to pH 5.0 by the addition of HCl), amino acid starvation (100 µg ml-1

serine hydroxamate), DNA damage (50 ng ml-1 mitomycin C), and inhibition of RNA synthesis

(erythromycin and spectinomycin at sub-MIC of 0.5 µg ml-1 and 50 µg ml-1, respectively). UA159

cells were processed with the Bio101 Fast Prep System (Qbiogen), and total RNA was

extracted using Trizol reagent (Invitrogen). DNA-free RNA samples were subjected to reverse

transcription using the First-Strand cDNA Synthesis Kit (MBI Fermentas). Real-time RT-PCR

reactions were carried out using the QuantiTech SYBR Green PCR master mix (Qiagen) in an

MX3000P System (Stratagene). A standard curve was plotted with cycle threshold (Ct) values

obtained from amplification of known quantities of cDNAs. The standard curve was used to

determine the efficiency (E) of comC primer set binding and amplification: E=10-1/slope.

Comparison of the expression of comC gene between its control and stress was determined

using the formula: Ratio=(EcomC)∆Ct(control-stress)/(E16SrRNA)∆Ct(control-stress). The 16S rRNA gene was

used as internal reference as we found the expression of this gene to be stable under the test

conditions. All assays were performed in triplicate with RNA isolated from three independent

experiments and using a P <0.01.

DNA microarrays

S. mutans UA159 and ΔcomX cells were grown with 2 µM sCSP or without (uninduced control)

to mid-log phase. Total RNA was extracted as described above. The cDNAs were prepared for

hybridization using the PFGRC protocol (http://pfgrc.tigr.org/protocols/M007.pdf). Microarray

Page 87: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

71

chips were scanned using a Gene Pix 4000B (Axon) and analyzed using the TM4 Microarray

Software Suite (http://www.tm4.org/). Transcripts levels were measured by cDNA hybridized to

a fourthfold redundant S. mutans microarray and averaged for three replicated hybridization.

Differential gene expression was based on a post-normalization cutoff of ± >2-fold. Significance

was determined using a one-class t-test.

Recombinant peptides

Recombinant CipB fusion peptides (precursor and mature form) were generated using the T7

expression vector pET28a(+) (Novagen). The full-length coding region of CipB (His6-fullCipB)

and its mature form (His6-GGCipB) were PCR amplified using forward and reverse primers

containing an NheI and XhoI restriction site, respectively. The amplicons were cloned in-frame

downstream from the hexa-His sequence in the T7 expression vector pET-28a(+) precut by the

same enzymes and transformed into E. coli BL21 (non-expression host) competent cells. The

nucleotide sequences of the inserts were confirmed by DNA sequencing. Recombinant

plasmids were then transformed into E. coli BL21(DE3) competent cells. E. coli transformant

cells were incubated aerobically at 37ºC in 100-ml LB-kanamycin 30 µg ml-1 supplemented with

1% glucose until the culture reached an OD600 of 0.6. IPTG was then added at a final

concentration of 1 mM to induce expression of recombinant fusion proteins and the incubation

was continued for a further 3 h at 37ºC. The cells were collected by centrifugation and

resuspended in 1× binding buffer (Novagen) and disrupted on ice by sonication. The soluble

fractions of the disrupted cells were recovered by centrifugation and the hexa-His-tagged

recombinant CipB proteins, His6-fullCipB and His6-GGCipB, were then purified by affinity

chromatography on Ni2+-nitrilotriacetic acid (Ni-NTA) resin (Novagen) as described previously

(Levesque et al., 2004). A total cell protein extract was prepared from IPTG-induced vectorless

E. coli BL21(DE3) and used as a negative control.

Page 88: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

72

Bacteriocin overlay assays

Twenty microlitres of S. mutans cultures were spotted onto THYE-agar plates directly from

overnight cultures (control) or after growth to mid-log phase in the presence of 2 µM sCSP.

Spots were allowed to dry, and were then over-laid with a 1/100 dilution of the indicator strain

L. lactis I6 suspended in 3 ml of THYE top agar. Plates were allowed to set, and were incubated

overnight before analysis.

Transformation assays

S. mutans cells were exposed to stress at mid-log phase as described for gene expression

analysis. To test the role of cell lysis in competence, cells were grown to OD600 of ~ 0.1 and

divided into three aliquots: i) no sCSP; ii) 0.2 µM sCSP; iii) 2 µM sCSP. Ten micrograms of

streptococcal genomic DNA containing a spectinomycin resistance marker was added to the

cultures, which were grown for a further 2.5 h before differential plating.

2.4 Results

2.4.1 Stress induces expression of the CSP pheromone

We asked whether environmental stress could activate the S. mutans competence regulon

by monitoring the expression of the CSP pheromone-encoding gene (comC) under stress.

Levels of comC transcript were significantly induced by acidic conditions at pH 5.0 (4.8 ± 0.8-

fold) and in the presence of a sub-inhibitory concentration of the protein synthesis-inhibitor

antibiotic spectinomycin (8.9 ± 3.8-fold). Levels of comC transcript were, however, unchanged in

the presence of the DNA damaging agent mitomycin C, the anti-metabolite serine hydroximate,

Page 89: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

73

and the antibiotic erythromycin under the conditions assayed. These data suggest that up-

regulation of expression of the CSP pheromone may link the competence cascade and the

stress response in S. mutans under some conditions. To test the impact of CSP on the stress

response directly, cells of a mutant defective in the CSP pheromone-encoding gene (∆comC)

were exposed to antibiotic and acid stress, harvested, and re-suspended in fresh medium for

growth analysis. The same experiment was carried out with the ∆comC mutant complemented

with a functional comC gene in trans. When exposed to a sub-inhibitory concentration of

spectinomycin, ∆comC mutant showed a significant increase of lag phase before recovering

from the antibiotic stress, while the complemented strain showed better recovery (Figure 2.1A).

In keeping with the less pronounced induction of comC expression under low pH stress

condition, the lag phase of ∆comC mutant was significantly shorter in acid stress (Figure. 2.1B)

than in antibiotic stress (Figure 2.1A). As observed under spectinomycin-stressed condition,

complementation of the ∆comC mutant under acid stress resulted in an initial increase in growth

rate followed by an eventual decline in cellular yield compared with the wild-type UA159 strain.

These results suggested to us that the up-regulation of CSP expression that occurs under

stress actively contributes to the initial stages of recovery as indicated by resumed cell growth,

but is detrimental if it is allowed to continue accumulating in the culture and may even cause cell

death if over-produced.

Page 90: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

74

Figure 2.1. Recovery of S. mutans from stress. To test the contribution of endogenous CSP to

stress recovery, early-log phase cells of the UA159 wild-type, a ∆comC mutant (no CSP

production) and an ∆comC complemented strain (overproduction of CSP) were grown in THYE

(control), in THYE containing spectinomycin (A), or in THYE at pH 5.0 (B) for 2.5 h. Cells were

harvested, and re-suspended in fresh THYE at 1/100 dilution. Absorbance of the growing

cultures was then automatically recorded for 16 h. Cells that recovered from stress in the

absence of CSP showed a growth defect, while the CSP over-producing cells recovered from

stress more quickly but attained lower growth yields than the wild-type. These results imply that

the S. mutans CSP pheromone is important in the stress response to acid and the antibiotic

spectinomycin.

Page 91: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

75

2.4.2 CSP pheromone triggers autolysis in a fraction of the population

These results suggest a mechanistic link for competence induction under stress in

S. mutans, and broaden the implications of previous work done by Claverys‟ group

(Dagkessamanskaia et al., 2004; Prudhomme et al., 2006) in S. pneumoniae, by demonstrating

a new role for the CSP pheromone in the stress response of more than one streptococcal

species. However, since important differences exist between the CSP-induced competence

cascade in S. mutans and S. pneumoniae (Martin et al., 2006), we examined the phenotypic

effect of increasing the concentration of CSP in cultures of S. mutans. We used exogenously

added synthetic CSP (sCSP) to study the effect of the CSP pheromone reproducibly, without

secondary complications from the stress itself (e.g., protein synthesis inhibition from

spectinomycin). We routinely use 0.2 μM sCSP to induce competence in S. mutans (Li et al.,

2001b), and based our choice of sCSP concentrations on our stress experiments (which

showed up to a 15-fold increase in comC transcript levels). It is possible that even higher

localized endogenous CSP concentrations exist in the biofilm environment, but we attempted to

avoid potential artifacts associated with overloading a signaling system by restricting the amount

of exogenously added sCSP to concentrations close to those which induce competence in vitro.

By monitoring the growth kinetics of planktonic cultures of S. mutans in the presence of

increasing concentrations of sCSP, we found a ComX-dependent decrease in the growth rate

and a ComDE-dependent decrease in the final growth yield of S. mutans proportional to the

concentration of sCSP (Fig. 2. 2). This phenotype differs from autolysis in S. pneumoniae, in

which the whole population lyses in the presence of CSP (Ronda et al., 1987).

Page 92: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

76

Figure 2.2. S. mutans growth kinetics. (A) Growth of S. mutans in the presence of sCSP. To

artificially mimic the high CSP concentrations induced by stress, cells were diluted into media

supplemented with increasing concentrations of sCSP or without (control) and grown for 16 h.

Cultures responded to sCSP with a slower growth rate and lower growth yield proportional to

the increase in sCSP concentration. To ensure the specificity of the phenotype, a peptide was

synthesized with the same amino acid composition as S. mutans CSP but with the order

randomized. The randomized peptide did not have any effect on S. mutans growth (not shown).

(B) Growth of S. mutans wild-type (WT), ∆comDE mutant and ∆comX mutant in the absence (–)

and presence (+) of 2 µM CSP pheromone. While eliminating ComDE restores both the growth

rate and yield of the culture in the presence of sCSP, the ∆comX mutant shows a defect in

growth yield both in the absence and presence of sCSP. We infer from this data that the altered

growth yield of the culture in the presence of sCSP is due to ComE-controlled genes.

Our cultures responded to high sCSP concentrations with a decreased growth rate but

stable plateau, which we attributed to either the average of two distinct populations of CSP-

induced (lysing) and uninduced (resistant) cells (known as all-or-none induction, reviewed in

(Davidson and Surette, 2008)), or to a uniform state of bacteriostasis in the whole population.

To distinguish between these two phenomena, we monitored the supernatants of S. mutans

cultures constitutively expressing a -glucuronidase (GUS) reporter gene for release to the

culture medium of intracellular GUS by cell lysis. GUS activity was significantly increased in the

supernatant of cultures grown with sCSP, indicating cell lysis occured at high sCSP

Page 93: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

77

concentrations (Figure 2.3). We next quantified the extent of cell death in mid-log phase

cultures grown in the presence of 2 µM sCSP using fluorescent staining for viability, and found

that <10% of the S. mutans UA159 population stains with propidium iodide due to cell death

(Figure 2.4). This result is in agreement with previous work done on S. mutans strain UA140

using fluorescence viability staining, in which only a sub-population of cells lysed no matter what

the concentration of exogenous sCSP (Qi et al., 2005). Finally, we restored the growth of the

culture to wild-type levels by sub-culturing sCSP-exposed cells into fresh medium (Figure 2.4),

confirming the existence of a CSP-resistant subpopulation even at concentrations of sCSP up to

200 μM. We conclude from these results that the CSP pheromone induces a state of population-

level stasis in S. mutans and invokes lysis in a fraction of the population while sparing the

remainder.

Figure 2.3. Effect of sCSP on culture density and cell lysis. The culture-wide graded response

to sCSP could be due to an identical bacteriostatic response in the whole population, or to an

„all-or-none‟ induction of lysis in a fraction of the population. To distinguish between the two, the

release of the cytoplasmic enzyme -glucuronidase (GUS) into the supernatant of growing

cultures was quantified. The observed increase in GUS activity in sCSP indicates lysis in a

fraction of the population. „Control‟ cultures were grown in THYE without sCSP, while „+ sCSP‟

cultures were supplemented with 2 μM sCSP. GUS activity is normalized to the optical density

of the culture at each time point, and is expressed in Miller units (MU).

Page 94: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

78

Figure 2.4: A sub-set of the population undergoes cells lysis, while the majority remains sCSP-

unresponsive. Viability staining using propidium iodide indicates less than 10% of the population

is dead in the presence of 2 μM sCSP (left). Growth of cells when re-suspended in fresh THYE

after overnight exposure to sCSP at the concentrations indicated. Growth resumes after sCSP

exposure even at 200 μM, suggesting that a sub-population of cells is always resistant to sCSP-

mediated cell lysis (right).

2.4.3 Genome-wide expression response to CSP: identification of mutacin V

Microarray-based expression profiling showed that 2 μM of sCSP altered the expression

(± ≥2-fold) of 277 genes in the S. mutans UA159 genome (Appendix A; Supplementary

Information SI Table S1). Since CSP induces gene expression both directly through ComE

signaling and secondarily through ComX, we also determined the expression response to sCSP

in the absence of ComX by microarray (Appendix A; SI Table S2). We found that ComE controls

the expression of 37 genes, among which are the S. mutans bacteriocins and comX itself (Table

2.2). Like S. pneumoniae, S. mutans ComX is a competence-specific sigma factor responsible

for expression of the entire competence regulon, including the CSP-encoding gene itself

(Peterson et al., 2004).

Page 95: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

79

Table 2.2. Relative expression levels of highly-expressed CSP-induced S. mutans genes

encoding putative and known bacteriocins and their accessory genes.

Locus* Common name, putative function Fold†

SMU.150 NlmA mutacin IV +11.19

SMU.151 NlmB mutacin IV +12.40

SMU.423 Putative bacteriocin +14.55

SMU.925 Putative immunity factor +18.20

SMU.1897 ABC transporter, ATP-binding; ComA +9.19

SMU.1898 Putative ABC transporter, ATP-binding and permease +4.18

SMU.1899 Putative ABC transporter fragment +5.22

SMU.1900 ABC transporter; ComB +5.92

SMU.1902 GG-motif-containing peptide +9.76

SMU.1903 Hypothetical protein +15.97

SMU.1904 Hypothetical protein +12.25

SMU.1905 GG-motif-containing peptide +10.11

SMU.1906 Putative bacteriocin +11.36

SMU.1907 Hypothetical protein +8.93

SMU.1908 Hypothetical protein +18.27

SMU.1909 Hypothetical protein +19.61

SMU.1910 Hypothetical protein +18.26

SMU.1912 Hypothetical protein +22.23

SMU.1913 Putative immunity protein +15.23

SMU.1914 Bacteriocin, mutacin V +20.41

SMU.1915 Precursor of CSP pheromone +3.76

SMU.1916 Histidine kinase ComD +10.50

SMU.1917 Response regulator ComE +11.32

SMU.1997 Sigma factor ComX +14.27

*Results for selected genes were ordered based on their position in the UA159

chromosome. The grey highlighted area indicates the genes located in a 13.5-kb

bacteriocin-related genomic island. Putative bacteriocin-encoding genes are in bold.

†Transcripts levels were measured by cDNA hybridized to a fourfold redundant S. mutans

microarray and averaged for three replicates hybridization. Quantitative real-time RT-PCR

was performed on selected genes to confirm the results obtained using microarray.

Page 96: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

80

The products of six CSP-responsive genes (cbpD, lytA, comM, cibA, cibB, cibC) have been

directly implicated in S. pneumoniae autolysis (Guiral et al., 2005; Håvarstein et al., 2006). Our

search for homologous effectors among the CSP-induced genes identified by microarray yielded

no significant sequence homologies, but suggested the involvement of a bacteriocin in

S. mutans CSP-induced lysis. Our search was also guided by experiments performed in

Streptococcus thermophilus and Streptococcus salivarius, showing similar dose-dependent

growth inhibition in the presence of the species-specific CSP paralogue BIP (Figure 2.5), a

signaling peptide known to induce bacteriocin expression (Fontaine et al., 2007). S. mutans

isogenic mutants were generated to be defective in mutacin IV (SMU.150, SMU.151), mutacin V

(SMU.1914), and GG-motif-containing peptides (SMU.423, SMU.1906). Importantly, inactivation

of SMU.1914, encoding the nonlantibiotic peptide bacteriocin mutacin V (Hale et al., 2005a),

drastically attenuated the response to sCSP in S. mutans (Fig. 2.6). To ensure that the

observed phenotype was not due to comC repression, we performed transcriptional analysis by

quantitative real-time PCR and found that the expression of the CSP-encoding gene was not

affected in the ∆1914 mutant (data not shown).

To confirm the bacteriocin-like activity of SMU.1914, we performed a series of agar overlay

assays using the indicator strain Lactococcus lactis I6. As observed by Hale et al., 2005, the

∆1914 mutant had a reduced zone of inhibition that was not observed for the mutacin IV mutant

against L. lactis I6 (Figure 2.7). Since some bacteriocins produced by Gram-positive bacteria

require two peptides for activity, we also investigated whether mutacin V (SMU.1914) caused

cell lysis alone or in combination with CSP itself (SMU.1915), given that the two peptide-

encoding genes share an overlapping (but divergent) promoter region (Kreth et al., 2007). We

constructed a CSP-independent raffinose-inducible expression system to express mutacin V

Page 97: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

81

Figure 2.5. Growth of Streptococcus thermophilus (A) and Streptococcus salivarius (B) in

increasing concentrations of their species-specific signaling peptides. In these two non-

competence species of streptococci, the peptide signaling system is composed of CSP-ComDE

paralogues BIP (bacteriocin-inducing peptide) and BlpRH, known to induce expression of

bacteriocins. The similarity between these BIP-induced growth curves and the CSP-induced

growth curves in S. mutans strongly suggests the involvement of a bacteriocin in the S. mutans

cell death cascade.

Page 98: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

82

Figure 2.6. Effect of 2 µM sCSP pheromone („+‟) on S. mutans wild-type (WT) and mutants

defective in the bacteriocin CipB (SMU.1914) and its putative immunity factors SMU.1913 and

CipI (SMU.925). Growth of the wild-type strain in THYE alone provided a baseline for

comparison (control).

Page 99: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

83

Figure 2.7. Agar overlay assays showing extracellular bacteriocin concentration in the UA159

wild-type and mutants defective in the CipB bacteriocin, the CipI immunity protein, the

bacteriocin mutacin IV and the bacteriocin transporter NlmTE. Extracellular bacteriocin

concentration was monitored by spoting mid-log phase cultures of the test strain grown in the

absence (control) and presence of sCSP.on agar plates. The test strains were then

subsequently overlayed with a 1/100 dilution of the indicator strain Lactococcus lactis I6

suspended in 1% top agar. Plates were then grown overnight at 37°C, and the amount of

extracellular bacteriocin was assessed according to the width of the zone of inhibition of L. lactis

surrounding the test strain.

Page 100: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

84

alone in S. mutans from a multi-copy plasmid, bypassing the need for exogenous sCSP

addition. This system allowed us to rule out the contribution of sCSP itself to our experimental

system, given that excesses of signaling peptide have been documented to function as

bacteriocin-like peptides at extremely high concentrations (Anderssen et al., 1998). Using this

CSP-independent inducible expression system, we found that mutacin V gene expression was

induced 5.0 ± 1.6-fold in raffinose (inducer) vs. in glucose (repressor). Although this level falls

short of the 20-fold induction of mutacin V we observed with sCSP (Table 2.1), we were still

able to induce a growth defect similar to sCSP using the raffinose-inducible system (Fig. 2.8).

This result confirms that mutacin V is necessary for the observed autolytic phenotype in sCSP.

We confirmed that SMU.1914 was regulated directly by ComE (Fig. 2.9), and was transcribed in

direct proportion to the amount of sCSP (data not shown) or the concentration of naturally

accumulating CSP in a growing culture of S. mutans (2.10). Although (Hale et al., 2005a)

proposed that SMU.1914 be designated nlmC (gene product mutacin V), we instead suggest

the three-letter prefix cip for CSP-induced peptide, and propose the name CipB for this self-

acting bacteriocin.

2.4.4 CipB bacteriocin likely acts intracellularly

We reasoned that CipB‟s action against self would occur either from 1) a position tethered to

the cell wall via cell-to-cell contact, 2) the extracellular environment, or 3) intracellularly via

membrane insertion from the cytoplasmic side. The Cib bacteriocins of S. pneumoniae are

tethered to the cell wall after export, and able to kill neighboring cells via cell-to-cell contact

(Guiral et al., 2005). We co-cultured sCSP-induced wild-type cells with ∆cipB mutant cells to

provide cell contact between CipB-expressing wild-type cells and the CipB-deficient mutant, but

did not observe any functional complementation of the mutant (results not shown). This result

Page 101: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

85

Figure 2.8. Growth kinetics of S. mutans UA159 strain containing the shuttle plasmid pDL277

(Leblanc et al., 1992) harboring the CipB-encoding gene under the control of the raffinose-

inducible promoter (notation „pMSM‟) of the S. mutans multiple-sugar metabolism operon

(McLaughlin and Ferretti, 1996). Growth was monitored by following OD600 for 16 h in TYE

containing either 0.5% raffinose (inducer, notation „raf‟) or 0.5% glucose (repressor, notation

„glc‟). The vector alone (no insert) was used as control. Also plotted is the growth of a mutant

defective in the putative immunity gene SMU.925 carrying the pMSM construct following

induction with 0.5% raffinose.

Page 102: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

86

Figure 2.9. RT-PCR gene expression profiles of cipB and cipI, encoding the bacteriocin and

immunity protein respectively, in S. mutans wild-type strain and mutants defective in the

alternate sigma factor ComX and the response regulator ComE. In addition to the lack of

detectable transcript in the ΔcomE mutant, both cipB and cipI have putative ComE binding sites

in their promoter regions (van der Ploeg, 2005). The constitutively expressed 16S rRNA gene

served as a loading control.

Figure 2.10. Quantitative real-time RT-PCR gene expression profiles of SMU.1914 (encoding

CipB), comC (encoding CSP) and SMU.925 (encoding CipI) at 1-h, 2-h, and 3-h intervals

following 1:100 dilution from an overnight culture. Gene expression was normalized to the

expression of the constitutively expressed 16S rRNA gene.

Page 103: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

87

indicated that CipB is not likely to act from a position tethered to the cell wall via cell-cell contact,

reinforcing the notion that differences exist between the autolytic pathways of S. mutans and S.

pneumoniae. Since bacteriocins traditionally act from the extracellular environment, we

investigated extracellular activity by growing the ∆cipB mutant in cell-free supernatant from a

sCSP-induced wild-type culture, and in medium supplemented with purified recombinant CipB

peptides (precursor and mature form). We found no effect on the growth of the S. mutans ∆cipB

mutant in conditioned medium or at concentrations of recombinant CipB peptides that

completely inhibited the indicator strain L. lactis I6 (Figure 2.11), indicating that extracellular

concentrations of the bacteriocin had no effect on the producing strain. Importantly, we found

that the recombinant bacteriocin precursor peptide (which includes the leader peptide required

for export) was equally effective at killing L. lactis. The function of the leader peptide is thought

to be two-fold: its presence keeps the bacteriocin inactive during translation, and secondly

traffics the bacteriocin to the correct ABC transporter for export (Drider et al. 2006). Since the

full-length peptide was equally effective against L. lactis, export-dependent processing does not

appear to be required for activation of the CipB bacteriocin. This result suggests that an

intracellular accumulation of the unprocessed bacteriocin may be lethal to the producing cell. To

further test if intracellular accumulation of CipB could cause lysis, we inactivated the dedicated

ABC transporter NlmTE, required for the export of S. mutans nonlantibiotic bacteriocins

including CipB (Hale et al., 2005b). We confirmed that the ∆nlmTE mutant was unable to

secrete CipB using an agar overlay assay with L. lactis indicator cells (Figure 2.7). We found a

significant growth defect in the ∆nlmTE mutant compared to the wild-type strain at all sCSP

concentrations tested (Figure 2.11), which we attribute to an increased intracellular

accumulation of unprocessed CipB. Interestingly, we also observed a longer lag phase in the

∆nlmTE mutant in the absence of sCSP following dilution from an overnight culture. Since the

CipB bacteriocin was highly expressed in overnight cultures (Figure 2.10), this initial growth

Page 104: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

88

Figure 2.11. CipB may act intracellularly. (A) Growth kinetics of S. mutans (WT) and L. lactis I6

in the presence of recombinant CipB (precursor and mature form). The full-length precursor

peptide represents the intracellular form of the bacteriocin, including its leader sequence. The

mature peptide represents the extracellular form of the bacteriocin, having been cleaved at the

GG-motif. Importantly, the precursor peptide is equally effective against L. lactis, implying that

export-dependent processing is not necessary for activity. (B) Effect of 2 μM sCSP („+‟) on the

wild-type strain and a mutant defective in NlmTE, the ABC transporter responsible for export of

CipB. The ΔnlmTE mutant was assayed over a range of sCSP concentrations, and showed

growth defects compared to the wild-type strain at all concentrations assayed (not shown), likely

due to the intracellular accumulation of CipB. The ΔnlmTE mutant has a growth defect even in

the absence of sCSP, possibly due to the accumulation of intracellular CipB induced by the high

endogenous CSP concentration found in the overnight culture from which it was diluted.

Page 105: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

89

defect supports the hypothesis that intracellular CipB accumulation was lethal to the producing

cell.

Of final note, the expression of nlmTE decreased approximately 2-fold in the presence of

2 µM sCSP, which was reflected in decreased zones of inhibition by agar overlay (Figure 2.7).

Given that the expression of cipB was increased 20-fold in these cells (Table 2.1), it appears

that CipB accumulated to high intracellular levels in the presence of high concentrations of

sCSP. This result was important in ruling out the notion that excess sCSP could have simply

overwhelmed the normal CipB export pathway. Instead, an active decrease in expression of the

transporter coupled with increased expression of the bacteriocin suggested that intracellular

accumulation of CipB may be an active process. Together, these data strongly suggested a

completely novel mechanism of action for a bacteriocin: intracellular action against self.

2.4.5 The small protein CipI (SMU.925) confers immunity

Bacteriocin-encoding genes are usually co-transcribed with their cognate immunity genes,

which are protective against the action of the bacteriocin (Abee, 1995; Diep et al., 2007).

Synthetic CSP exposure induced two candidate immunity genes, SMU.925 and SMU.1913

(Table 2.1), which share 82% a.a. identity and have putative ComE binding sites in their

promoter regions (van der Ploeg, 2005). Unexpectedly, the growth of a ∆1913 mutant was

identical to the wild-type strain in the presence of sCSP (Figure 2.6), indicating that although

SMU.1913 was co-transcribed with cipB (data not shown), it did not prevent CSP-induced lysis.

In contrast, inactivation of SMU.925 resulted in almost complete growth arrest at high sCSP

concentrations (Figure 2.6). To confirm the role of SMU.925 in immunity to cell death, wild-type

cells expressing SMU.925 from its own native promoter on a multicopy plasmid were exposed to

sCSP. The cells over-expressing SMU.925 were significantly more resistant to sCSP than the

Page 106: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

90

wild-type (Figure 2.12). We also transformed the Δ925 mutant with the same CSP-independent

raffinose-inducible cipB expression construct used above. Cells deficient in the immunity gene

were more susceptible than the wild-type to high levels of cipB expressed in presence of

raffinose (Figure 2.8), consistent with the proposed role for CipI as the immunity gene. We thus

propose the designation cipI (CSP-induced peptide Immunity) for SMU.925.

As for cipB, cipI was found to be dependent on ComE for its transcription (Figure 2.9), and was

transcribed in proportion to the amount of sCSP added (data not shown). However, unlike cipB,

the expression of cipI showed an additional level of density-dependent control, since dilution

from an overnight culture (high to low cell density) caused an increase in its expression (2.10).

We considered that a second mechanism controlling cipI expression might exist to prevent cell

lysis at low cell density, and tested whether autoinducer-2 (AI-2) might be the signaling

molecule regulating this second system. No growth defect was observed when a ∆luxS mutant

deficient in the LuxI-type autoinducer synthetase (enzyme responsible for the synthesis of AI-2

(Bassler and Losick, 2006) was grown in the presence of increasing sCSP concentrations

(results not shown), indicating that AI-2 is likely not the second signal molecule.

Page 107: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

91

Figure 2.12. Growth of the wild-type and a wild-type strain over-expressing SMU.925 from a

multi-copy plasmid in the absence and presence of 2 μM sCSP. The sCSP-resistant CipB

mutant was used as a control.

2.4.6 Role of CSP-induced lysis in genetic competence

In addition to its known role in competence (Li et al., 2001b), we have shown that S. mutans

CSP is involved in the stress response and can trigger autolysis. Bacteria are unicellular

organisms, and would seem incapable of altruistic behaviors like cellular suicide. For an

altruistic trait like autolysis to be preserved by evolution, it must be linked to a behavior that

benefits genetically identical sibling cells (Keller and Surette, 2006). We reasoned that the

linkage of the stress response, competence induction and autolysis through the CSP

pheromone could serve to facilitate the exchange of fitness-enhancing DNA under stress, and

maintain the autolysis pathway through evolution. If our hypothesis is correct, S. mutans cells

must be equally transformable at the high sCSP concentrations in which we observe cell lysis.

Indeed, we found cells exposed to 2 μM sCSP to be transformed at frequencies similar to those

Page 108: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

92

obtained using 0.2 μM sCSP (Figure 2.13). We next examined whether lysis and competence

development could occur simultaneously in a culture using fluorescent staining for viability and a

GFP reporter fused to the promoter of the gene encoding the sigma factor responsible for

induction of the CSP-dependent competence regulon, comX. We observed sporadic expression

of PcomX–gfp throughout the culture, consistent with the hypothesis that only a fraction of the

S. mutans population is sCSP-induced (and competent) at any one time (Figure 2.14A).

However, when sCSP-treated cultures were counter-stained with propidium iodide, we observed

overlap between PcomX–gfp expression and cell death in the majority of cases (Figure 2.14B-

C). This result was not surprising, since both cipB and comX are regulated directly via CSP-

ComDE signaling and suggests that sCSP-induced competent cells continue to accumulate

CipB under continuous sCSP stimulation, resulting in cell death. We therefore tested the

transformation efficiency of the ∆cipB mutant in the presence of sCSP to determine the

transformation efficiency at high sCSP concentrations in the absence of cell death. When cell

death in the culture was prevented by inactivation of cipB, the transformation efficiency in the

presence of sCSP could not be induced beyond sCSP-independent levels (no sCSP added)

(Figure 2.13). Conversely, when cell death in the culture was promoted by removal of the

immunity protein, ΔcipI cultures reached levels of transformation comparable to wild-type sCSP-

induced levels in the absence of exogenously added sCSP (Figure 2.13). These results

indicated that the sCSP-induced competence cascade is connected to CipB-mediated autolysis

in S. mutans. We are currently investigating whether factors released via lysis of a sub-

population of cells contribute to the development of competence in the surviving population.

Page 109: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

93

Figure 2.13. Transformation efficiency of S. mutans wild-type strain and its mutants deficient in

the CipB bacteriocin and CipI immunity protein. Sheared genomic DNA carrying an antibiotic

resistance gene was added alone or with 0.2 μM or 2 μM sCSP to growing cultures. Cells were

grown for a further 2.5 h before differential plating. Results obtained for the wild-type strain

showed that transformation is possible and equally efficient at the concentrations of sCSP that

induce cell lysis. Mutants unable to undergo cell lysis (CipB–) showed no increase in

transformation frequency in the presence of sCSP, while mutants with increased lysis potential

(CipI–) showed increased transformation in the absence of sCSP. Transformation efficiencies

are expressed as the number of antibiotic-resistant CFUs divided by the total number of CFUs.

A ComE-deficient strain served as a transformation-deficient control.

Page 110: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

94

Figure 2.14. Competence and lysis in cultures of S. mutans. The alternate sigma factor ComX

is induced by ComE, and is responsible for induction of the CSP-dependent competence

regulon. A PcomX–gfp reporter fusion was used to monitor the development of competence in

the presence of 2 μM sCSP (A). Cultures were then counter-stained with propidium iodide (B) to

determine cell death. When images were merged (C), it was apparent that cells expressing

comX were also undergoing cell lysis.

The results obtained with PcomX-gfp suggest that the sCSP-dependent competence

pathway does not contribute to the uptake of fitness-enhancing genes in a stressed population,

since sCSP-induced cells are killed due to CipB accumulation. However, we examined whether

stress alone (in the absence of sCSP) could induce transformation in S. mutans using sub-

inhibitory concentrations of spectinomycin known to increase comC expression. We found a

4.0 ± 1.3-fold increase in transformation efficiency in stressed cultures vs. control cultures.

Importantly, the transformation efficiency of a ∆comC mutant strain was identical to the wild-type

in presence of sub-inhibitory concentration of spectinomycin, indicating that the increase in

transformation we observed was through the CSP-independent competence pathway. Together,

these data suggest that the increase in CSP pheromone production under stress causes

S. mutans autolysis through the ComDE signal transduction system, which releases DNA for

uptake via the CSP-independent competence pathway (Figure 2.15).

Page 111: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

95

Figure 2.15. Summary of data. The sensor histidine kinase ComD is activated either directly by

stress or by an accumulation of the CSP pheromone, and activates its cognate response

regulator ComE by phosphorylation. Activated ComE then directly regulates the expression of

37 genes, including the alternate sigma factor ComX, the CipB bacteriocin and its immunity

protein CipI. CipB causes cell lysis in a fraction of the population, which potentially contributes

DNA for uptake and other secondary signals to trigger genetic competence in the surviving

population. Expression of 240 genes, including the competence cascade and the CSP molecule

itself, are directly controlled by ComX. Open grey arrows: direct genetic regulation. Solid black

arrow: phenotype caused.

Page 112: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

96

2.5 Discussion

Experimental evidence has long supported a link between the competence cascade and

stress response in S. mutans (Ahn et al., 2006; Li et al., 2001a; Li et al., 2002a; Qi et al., 2004),

but a mechanistic explanation for these phenotypes has remained elusive. In this study, we

report the characterization of a novel self-acting intracellular bacteriocin which functions as a

mediator of autolysis in S. mutans. We have characterized this autolytic response in the

physiological context of an elevated concentration of the CSP pheromone under stress, which

provides the proverbial „missing link‟ between stress and competence in S. mutans. Moreover,

our results broaden the implications of previous work in pneumococci (Claverys et al., 2006) by

demonstrating the presence of a peptide „alarmone‟ in phylogenetic groups of streptococci

beyond the mitis group. However, like in pneumococci (Prudhomme et al., 2006), not all

stressful conditions assayed with S. mutans triggered CSP up-regulation. While we have

focused on elucidating the downstream pathways triggered by CSP up-regulation in S. mutans,

future work directed towards understanding why some stresses trigger CSP up-regulation while

others do not will complete our understanding of the stress response in this organism.

We focused on the effect of the stress-induced up-regulation of the CSP pheromone in the

absence of the stress itself using exogenously added sCSP. We demonstrated a slower growth

rate and a reduction in growth yield at high concentrations of sCSP due to a balance between

cell growth and autolysis in S. mutans. This response differs from the uniform autolytic response

mounted by S. pneumoniae in the presence of high CSP. S. pneumoniae achieves both a

death-susceptible and a death-resistant population by expressing two different CSP pherotypes

(Claverys et al., 2006). However, S. mutans has been shown to encode a single CSP pherotype

(Allan et al., 2007). The nature of cell lysis (and competence) responses in S. mutans is

therefore different than the response in S. pneumoniae, and may be closer to the other well-

characterized naturally transformable Gram-positive organism, B. subtilis, whose competence

Page 113: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

97

response occurs in only a fraction of the population due to bistability (Smits et al., 2005). The

bistable response occurs when positive feedback into a system occurs on a fast timescale and

negative feedback occurs on a slower timescale (Davidson and Surette, 2008). The result is an

amplification of transcriptional “noise” in a system, which manifests as two transient but distinct

populations of induced and uninduced cells (Davidson and Surette, 2008; Dubnau and Losick,

2006). We envision a situation in S. mutans in which stress-induced CSP upregulation serves to

„prime the pump‟ of bistability in the CSP-ComDE circuit, causing some cells to respond with a

rapid and extreme up-regulation of CSP-controlled genes (including comC itself), due to positive

feedback. What is unique about the genetic organization of comC in S. mutans is its divergent

regulation from mutacin V (Kreth et al., 2006b; Kreth et al., 2007). Kreth and colleagues showed

that this genetic organization results in repression of comC transcription when SMU.1914

expression is activated by ComE binding to their common intergenic region (Kreth et al., 2007).

When SMU.1914 expression was high, a built-in „safety mechanism‟ prevented the further

accumulation of CSP in the culture by repressing comC transcription, which eventually feeds

back into the loop to prevent autolysis of the whole population.

We identified the type II bacteriocin mutacin V (CipB) as a major factor in CSP-induced lysis

using genome-wide gene expression analysis and mutagenesis. Importantly, we have

demonstrated that its activity against self is due to intracellular accumulation in the producing

strain. We have presented evidence that CipB acts alone to cause cell death, that the

unprocessed form of the bacteriocin is active against an indicator strain, and that inactivation of

the NlmTE transporter causes cell death. This intracellular mechanism makes sense when

examined from an environmental context – a secreted or surface-located „death peptide‟ would

have the potential to wreak havoc in the tightly packed oral biofilm environment. Bacteriocins

generally act by creating channels or pores in the cell membrane that destroy the membrane

potential and cause cell death by cellular energy depletion (Nes et al., 2007). CipB likely causes

Page 114: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

98

cell death initially via a similar mechanism, and cell lysis secondarily by activating murein

hydrolases and/or lytic enzymes (Galvez et al., 1990). Alternatively, intracellular accumulation of

CipB may cause cell lysis in a manner analogous to the holin/anti-holin system characterized in

Staphylococcus aureus, in which a pore-forming peptide (the holin) inserts into the cytoplasmic

membrane to allow degradative enzymes to access the cell wall (Rice et al., 2003). The

antimicrobial activity of CipB is directed mainly against non-streptococcal species, and not

against traditional mutacin target organisms like Streptococcus gordonii and

Streptococcus salivarius (Hale et al., 2005a). Given our data showing that CipB acts

intracellularly, we suggest that the main function of this peptide is CSP-mediated autolysis, and

not as an exported bacteriocin. It is possible that the export of CipB outside S. mutans by

NlmTE transporter is part of a detoxification mechanism, and that the few bacterial species that

are susceptible to it are bystanders rather than primary targets. Data showing a significant

growth defect in S. mutans NlmTE-deficient cells compared with the parental strain supports this

argument. We also show the potential for a similar auto-active bacteriocin in members of the

salivarius group of streptococci.

Unexpectedly, inactivation of SMU.1913 excluded its involvement in CSP-mediated

autolysis. Instead, the product of the cipI gene encodes the immunity factor involved in

S. mutans autolysis. Although it is unexpected to find genes paired in function unlinked on the

genome, the need for redundant controls over this autolytic pathway may have necessitated this

duplication. Our view is supported by evidence showing that in addition to ComE regulation at

high CSP concentrations, an additional (and currently unknown) regulatory pathway triggered

activation of cipI expression at low cell density. This safety mechanism would be impossible if

the immunity protein was co-transcribed with the bacteriocin itself.

We hypothesized that competence and lysis would allow the exchange of fitness-enhancing

DNA under stress. However, when we monitored the induction of cell death and competence on

Page 115: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

99

a cell-by-cell basis using a transcriptional GFP reporter gene fusion, we found that the same

population became competent and lysed. This finding was not entirely surprising since both

competence and lysis are triggered by the CSP-ComDE circuit. Instead, we showed that

transformation can occur in a CSP-independent manner in spectinomycin stress, providing an

alternative pathway for the acquisition of fitness-enhancing genes. The lack of transformation in

the ∆cipB mutant is somewhat paradoxical, given that these cells are the survivors in a CSP-

induced population, and would be the expected recipients of transforming DNA. However, the

contrary result with ∆cipI implies that cell death in the CSP-responsive members of a population

may trigger genetic competence in the CSP-unresponsive survivors, and suggests that cellular

factors released via lysis could provide secondary signals to induce competence via a CSP-

ComDE-independent pathway. The ability to sense cell lysis would permit naturally competent

bacteria to turn on their uptake machinery when DNA is available in their environment. We are

currently exploring this fascinating possibility.

2.6 Acknowledgements

We thank John Tagg and Nicholas Heng for providing mutants related to mutacin V transport,

Indranil Biswas for providing shuttle expression plasmids for S. mutans, and Elena Voronejskaia

for assisting with the mutant constructions. This work was supported by CIHR-Priority

Announcement IMHA Grant FRN-90114 (to C.M.L.) and by NIDCR Grant R01 DE013230-08 (to

D.G.C.). DNA microarrays were supported through NIDCR via NIAID contract number N01-

AI15447 to JCVI. J.A.P. is the recipient of a CIHR Strategic Training Fellowship in Cell Signaling

in Mucosal Inflammation and Pain.

Page 116: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

100

2.7 References

Abee, T. (1995) Pore-forming bacteriocins of Gram-positive bacteria and self-protection

mechanisms of producer organisms. FEMS Microbiology Letters 129: 1-9.

Ahn, S.J., Wen, Z.T., and Burne, R.A. (2006) Multilevel control of competence development and

stress tolerance in Streptococcus mutans UA159. Infect Immun 74: 1631-1642.

Allan, E., Hussain, H.A., Crawford, K.R., Miah, S., Ascott, Z.K., Khwaja, M.H., and Hosie, A.H.F.

(2007) Genetic variation in comC, the gene encoding competence-stimulating peptide

(CSP) in Streptococcus mutans. FEMS Microbiology Letters 268: 47-51.

Anderssen, E.L., Diep, D.B., Nes, I.F., Eijsink, V.G., and Nissen-Meyer, J. (1998) Antagonistic

activity of Lactobacillus plantarum C11: two new two-peptide bacteriocins, plantaricins

EF and JK, and the induction factor plantaricin A. Appl Environ Microbiol 64: 2269-2272.

Aspiras, M.B., Ellen, R.P., and Cvitkovitch, D.G. (2004) ComX activity of Streptococcus mutans

growing in biofilms. FEMS Microbiol Lett 238: 167-174.

Bassler, B.L., and Losick, R. (2006) Bacterially speaking. Cell 125: 237-246.

Bjedov, I., Tenaillon, O., Gerard, B., Souza, V., Denamur, E., Radman, M., Taddei, F., and

Matic, I. (2003) Stress-Induced Mutagenesis in Bacteria. Science 300: 1404-1409.

Chastanet, A., Prudhomme, M., Claverys, J.P., and Msadek, T. (2001) Regulation of

Streptococcus pneumoniae clp genes and their role in competence development and

stress survival. J Bacteriol 183: 7295-7307.

Claverys, J.-P., Prudhomme, M., Mortier-Barriere, I., and Martin, B. (2000) Adaptation to the

environment: Streptococcus pneumoniae, a paradigm for recombination-mediated

genetic plasticity? Molecular Microbiology 35: 251-259.

Claverys, J.-P., Prudhomme, M., and Martin, B. (2006) Induction of Competence Regulons as a

General Response to Stress in Gram-Positive Bacteria. Annual Review of Microbiology

60: 451-475.

Dagkessamanskaia, A., Moscoso, M., Henard, V., Guiral, S., Overweg, K., Reuter, M., Martin,

B., Wells, J., and Claverys, J.P. (2004) Interconnection of competence, stress and CiaR

regulons in Streptococcus pneumoniae: competence triggers stationary phase autolysis

of ciaR mutant cells. Mol Microbiol 51: 1071-1086.

Davidson, C.J., and Surette, M.G. (2008) Individuality in Bacteria. Annual Review of Genetics

42: 253-268.

Page 117: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

101

de Saizieu, A., Gardes, C., Flint, N., Wagner, C., Kamber, M., Mitchell, T.J., Keck, W., Amrein,

K.E., and Lange, R. (2000) Microarray-based identification of a novel Streptococcus

pneumoniae regulon controlled by an autoinduced peptide. J Bacteriol 182: 4696-4703.

Diep, D.B., Skaugen, M., Salehian, Z., Holo, H., and Nes, I.F. (2007) Common mechanisms of

target cell recognition and immunity for class II bacteriocins. PNAS 104: 2384-2389.

Drider, D., Fimland, G., Hechard, Y., McMullen, L.M., and Prevost, H. (2006) The Continuing Story of Class IIa Bacteriocins. Microbiol. Mol. Biol. Rev. 70: 564-582.

Dubnau, D., and Losick, R. (2006) Bistability in bacteria. Mol Microbiol 61: 564-572.

Fontaine, L., Boutry, C., Guedon, E., Guillot, A., Ibrahim, M., Grossiord, B., and Hols, P. (2007)

Quorum-sensing regulation of the production of Blp bacteriocins in Streptococcus

thermophilus. J Bacteriol 189: 7195-7205.

Galvez, A., Valdivia, E., Martinez-Bueno, M., and Maqueda, M. (1990) Induction of autolysis in

Enterococcus faecalis S-47 by peptide AS-48. J Appl Bacteriol 69: 406-413.

Guiral, S., Mitchell, T.J., Martin, B., and Claverys, J.P. (2005) Competence-programmed

predation of noncompetent cells in the human pathogen Streptococcus pneumoniae:

genetic requirements. Proc Natl Acad Sci U S A 102: 8710-8715.

Hale, J.D., Ting, Y.T., Jack, R.W., Tagg, J.R., and Heng, N.C. (2005a) Bacteriocin (mutacin)

production by Streptococcus mutans genome sequence reference strain UA159:

elucidation of the antimicrobial repertoire by genetic dissection. Appl Environ Microbiol

71: 7613-7617.

Hale, J.D.F., Heng, N.C.K., Jack, R.W., and Tagg, J.R. (2005b) Identification of nlmTE, the

Locus Encoding the ABC Transport System Required for Export of Nonlantibiotic

Mutacins in Streptococcus mutans. J. Bacteriol. 187: 5036-5039.

Håvarstein, L.S., Coomaraswamy, G., and Morrison, D.A. (1995) An unmodified

heptadecapeptide pheromone induces competence for genetic transformation in

Streptococcus pneumoniae. Proc Natl Acad Sci U S A 92: 11140-11144.

Håvarstein, L.S., Gaustad, P., Nes, I.F., and Morrison, D.A. (1996) Identification of the

streptococcal competence-pheromone receptor. Mol Microbiol 21: 863-869.

Håvarstein, L.S., Martin, B., Johnsborg, O., Granadel, C., and Claverys, J.P. (2006) New

insights into the pneumococcal fratricide: relationship to clumping and identification of a

novel immunity factor. Mol Microbiol 59: 1297-1307.

Keller, L., and Surette, M.G. (2006) Communication in bacteria: an ecological and evolutionary

perspective. Nat Rev Microbiol 4: 249-258.

Kreth, J., Merritt, J., Shi, W., and Qi, F. (2005) Co-ordinated bacteriocin production and

competence development: a possible mechanism for taking up DNA from neighbouring

species. Mol Microbiol 57: 392-404.

Page 118: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

102

Kreth, J., Merritt, J., Zhu, L., Shi, W., and Qi, F. (2006a) Cell density- and ComE-dependent

expression of a group of mutacin and mutacin-like genes in Streptococcus mutans.

FEMS Microbiol Lett 265: 11-17.

Kreth, J., Merritt, J., Zhu, L., Shi, W., and Qi, F. (2006b) Cell density- and ComE-dependent

expression of a group of mutacin and mutacin-like genes in Streptococcus mutans.

FEMS Microbiology Letters 265: 11-17.

Kreth, J., Hung, D.C.I., Merritt, J., Perry, J., Zhu, L., Goodman, S.D., Cvitkovitch, D.G., Shi, W.,

and Qi, F. (2007) The response regulator ComE in Streptococcus mutans functions both

as a transcription activator of mutacin production and repressor of CSP biosynthesis.

Microbiology 153: 1799-1807.

Lau, P.C., Sung, C.K., Lee, J.H., Morrison, D.A., and Cvitkovitch, D.G. (2002) PCR ligation

mutagenesis in transformable streptococci: application and efficiency. J Microbiol

Methods 49: 193-205.

Levesque, C., Vadeboncoeur, C., and Frenette, M. (2004) The csp operon of Streptococcus

salivarius encodes two predicted cell-surface proteins, one of which, CspB, is associated

with the fimbriae. Microbiology 150: 189-198.

Li, Y.H., Hanna, M.N., Svensater, G., Ellen, R.P., and Cvitkovitch, D.G. (2001a) Cell density

modulates acid adaptation in Streptococcus mutans: implications for survival in biofilms.

J Bacteriol 183: 6875-6884.

Li, Y.H., Lau, P.C., Lee, J.H., Ellen, R.P., and Cvitkovitch, D.G. (2001b) Natural genetic

transformation of Streptococcus mutans growing in biofilms. J Bacteriol 183: 897-908.

Li, Y.H., Lau, P.C., Tang, N., Svensater, G., Ellen, R.P., and Cvitkovitch, D.G. (2002) Novel two-

component regulatory system involved in biofilm formation and acid resistance in

Streptococcus mutans. J Bacteriol 184: 6333-6342.

Martin, B., Quentin, Y., Fichant, G., and Claverys, J.-P. (2006) Independent evolution of

competence regulatory cascades in streptococci? Trends in Microbiology 14: 339-345.

McLaughlin, R.E., and Ferretti, J.J. (1996) The multiple-sugar metabolism (msm) gene cluster of

Streptococcus mutans is transcribed as a single operon. FEMS Microbiology Letters

140: 261-264.

Mitchell, T.J. (2003) The pathogenesis of streptococcal infections: from tooth decay to

meningitis. Nat Rev Microbiol 1: 219-230.

Nes, I.F., Diep, D.B., and Holo, H. (2007) Bacteriocin diversity in Streptococcus and

Enterococcus. J Bacteriol 189: 1189-1198.

Perry, J.A., Levesque, C.M., Suntharaligam, P., Mair, R.W., Bu, M., Cline, R.T., Peterson, S.N.,

and Cvitkovitch, D.G. (2008) Involvement of Streptococcus mutans regulator RR11 in

Page 119: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

103

oxidative stress response during biofilm growth and in the development of genetic

competence. Lett Appl Microbiol 47: 439-444.

Peterson, S.N., Sung, C.K., Cline, R., Desai, B.V., Snesrud, E.C., Luo, P., Walling, J., Li, H.,

Mintz, M., Tsegaye, G., Burr, P.C., Do, Y., Ahn, S., Gilbert, J., Fleischmann, R.D., and

Morrison, D.A. (2004) Identification of competence pheromone responsive genes in

Streptococcus pneumoniae by use of DNA microarrays. Mol Microbiol 51: 1051-1070.

Prudhomme, M., Attaiech, L., Sanchez, G., Martin, B., and Claverys, J.-P. (2006) Antibiotic

Stress Induces Genetic Transformability in the Human Pathogen Streptococcus

pneumoniae. Science 313: 89-92.

Qi, F., Merritt, J., Lux, R., and Shi, W. (2004) Inactivation of the ciaH Gene in Streptococcus

mutans diminishes mutacin production and competence development, alters sucrose-

dependent biofilm formation, and reduces stress tolerance. Infect Immun 72: 4895-4899.

Qi, F., Kreth, J., Levesque, C.M., Kay, O., Mair, R.W., Shi, W., Cvitkovitch, D.G., and Goodman,

S.D. (2005) Peptide pheromone induced cell death of Streptococcus mutans. FEMS

Microbiol Lett 251: 321-326.

Rice, K.C., Firek, B.A., Nelson, J.B., Yang, S.J., Patton, T.G., and Bayles, K.W. (2003) The

Staphylococcus aureus cidAB operon: evaluation of its role in regulation of murein

hydrolase activity and penicillin tolerance. J Bacteriol 185: 2635-2643.

Ronda, C., Garcia, J.L., Garcia, E., Sanchez-Puelles, J.M., and Lopez, R. (1987) Biological role

of the pneumococcal amidase. Cloning of the lytA gene in Streptococcus pneumoniae.

Eur J Biochem 164: 621-624.

Senadheera, M.D., Guggenheim, B., Spatafora, G.A., Huang, Y.C., Choi, J., Hung, D.C.,

Treglown, J.S., Goodman, S.D., Ellen, R.P., and Cvitkovitch, D.G. (2005) A VicRK signal

transduction system in Streptococcus mutans affects gtfBCD, gbpB, and ftf expression,

biofilm formation, and genetic competence development. J Bacteriol 187: 4064-4076.

Senadheera, M.D., Lee, A.W., Hung, D.C., Spatafora, G.A., Goodman, S.D., and Cvitkovitch,

D.G. (2006) The Streptococcus mutans vicX gene product modulates gtfB/C Expression,

Biofilm Formation, Genetic Competence and Oxidative Stress Tolerance. J Bacteriol.

Smits, W.K., Eschevins, C.C., Susanna, K.A., Bron, S., Kuipers, O.P., and Hamoen, L.W.

(2005) Stripping Bacillus: ComK auto-stimulation is responsible for the bistable response

in competence development. Molecular Microbiology 56: 604-614.

van der Ploeg, J.R. (2005) Regulation of bacteriocin production in Streptococcus mutans by the

quorum-sensing system required for development of genetic competence. J Bacteriol

187: 3980-3989.

Wen, Z.T., Suntharaligham, P., Cvitkovitch, D.G., and Burne, R.A. (2005) Trigger factor in

Streptococcus mutans is involved in stress tolerance, competence development, and

biofilm formation. Infect Immun 73: 219-225.

Page 120: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

104

Chapter 3: Cell Death in Streptococcus mutans Biofilms: a Link

Between CSP and Extracellular DNA

JA Perry, DG Cvitkovitch and CM Levesque. 2009. FEMS Micro Lett. 299:261-6

Page 121: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

105

3.1 Abstract

Streptococcal competence stimulating peptides were once thought to passively communicate

population density in a process known classically as quorum sensing. However, recent evidence

has shown that these peptides may also be inducible „alarmones‟, capable of conveying

sophisticated messages in a population including the induction of altruistic cellular suicide under

stressful conditions. We have previously characterized the alarmone response in

Streptococcus mutans, a cariogenic resident of the oral flora, in which a novel bacteriocin-like

peptide causes cell death in a sub-set of the population. Our objective in this work was to

characterize the mechanism of immunity to cell death in S. mutans. Towards this goal, we have

identified the conditions under which immunity is induced, and identified the regulatory system

responsible for differential (and protective) expression of immunity. We also showed that CSP-

induced death contributes to S. mutans biofilm formation through the release of chromosomal

DNA into the extracellular matrix, providing a long sought-after mechanistic explanation for the

role of CSP in S. mutans biofilm formation.

Page 122: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

106

3.2 Introduction

Bacteria have long been studied as single-celled, primitive organisms, free-floating in

laboratory culture. However, bacterial species in nature have a strong tendency to colonize

surfaces and form complex, multi-species communities referred to as biofilms (Costerton et al.,

1994). In nature, biofilms are found on rocks in streams, in industrial bioreactors, and in animal

host environments like the oropharyngeal, gastrointestinal and vaginal tracts, and on medical

devices. A biofilm in its simplest form is composed of a surface (or „substratum‟), surface-

attached cells, and a surrounding extracellular matrix of biopolymers (Dunne, 2002). By

microscopic analysis, a biofilm appears to be a highly hydrated and open structure, composed

mainly of non-cellular material including water channels and exopolymeric substances (EPS)

which form the extracellular matrix (Lawrence et al., 1991). The EPS forms the outermost layer

of the biofilm, and is composed of a hydrated, anionic mesh of bacterial exopolymers and

trapped environmental molecules (Branda et al., 2005). EPS comprise a wide variety of

polysaccharides, proteins, glycoproteins, glycolipids, and in some cases, large amounts of

extracellular DNA (eDNA). DNA was first shown to be present in the extracellular matrix of

biofilms formed by Pseudomonas aeruginosa (Whitchurch et al., 2002), and is now widely

recognized as a major constituent of the matrix (Flemming et al., 2007). The matrix functions as

a permeability barrier to limit both the diffusion of beneficial nutrients away from the biofilm, and

prevent or slow the diffusion of harmful substances like antibiotics and predatory cells of the

immune system from accessing matrix-embedded cells (Costerton et al., 1999).

The spatial separation of sessile cells combined with nutrient/waste and oxygen gradients

within the biofilm results in a heterogeneous population of cells, distinct from their planktonic

counterparts in gene expression patterns and „behaviours‟ (Stoodley et al., 2002; Beloin &

Ghigo, 2005). The metabolic task sharing, communication and phenotypic heterogeneity within

Page 123: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

107

a biofilm have led to their being likened to multicellular-type organisms, since cooperation

results in the success of the group (Shapiro, 1998; Parsek & Greenberg, 2005). The

coordinated behaviour of single-celled bacteria is accomplished using different classes of small

diffusible signalling molecules in a process called quorum sensing (reviewed in Waters &

Bassler, 2005).

Streptococcus mutans is a well-characterized resident of the oral biofilm, and is thought to be

the main causative agent of the most common human infectious disease, dental caries

(Loesche, 1986). The S. mutans quorum sensing system is composed of the competence

stimulating peptide (CSP) pheromone and the ComDE two-component signal transduction

system (TCS). The S. mutans CSP-ComDE system regulates several phenotypes including

genetic competence (Li et al., 2001a), biofilm formation (Li et al., 2002b), acid tolerance (Li et

al., 2001b), and bacteriocin production (Kreth et al., 2006). We have recently shown that the

CSP pheromone is also stress-inducible, and triggers autolysis in a fraction of the S. mutans

population at high concentrations (Perry et al., 2009). Autolysis in S. mutans occurs through the

intracellular accumulation of a self-acting bacteriocin, CipB, and is prevented by the action of

the bacteriocin immunity protein CipI. Previously, we showed that CipI was differentially

regulated from CipB at low cell density (Perry et al., 2009). Here we report the characterization

of the CipI immunity protein, detailing its expression and regulation, and propose a role for CSP-

induced autolysis in the release of eDNA in the S. mutans biofilm.

3.3 Materials and methods

Bacterial strains and culture conditions

Page 124: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

108

S. mutans UA159 wild-type strain and its mutants were grown in Todd-Hewitt–Yeast Extract

(THYE) broth at 37ºC with 5% CO2 without agitation. Growth kinetics were assessed using a

Bioscreen microbiology workstation (Bioscreen C Labsystems, Finland) as previously described

(Hasona et al., 2005). Results represent an average of five technical replicates, and two to four

independent experiments.

Table 3.1: Bacterial strains used in this study.

Strain Properties Reference

UA159 S. mutans wild-type ATCC

ΔcipI ∆smu.925; Emr (Perry et al., 2009)

ΔcipB ∆smu.1914; Emr (Perry et al., 2009)

ΔliaS ∆smu.486; Emr (Levesque et al., 2007)

ΔliaR ∆smu.487; Emr (Perry et al., 2008)

CipI expression studies

The expression of CipI-encoding gene in UA159 wild-type, ΔLiaS and ΔLiaR strains was

quantified by real-time reverse-transcription (RT)-PCR. Briefly, cells were harvested and total

RNA was extracted using the Bio101 Fast Prep System (Qbiogen) and Trizol reagent

(Invitrogen). DNA-free RNA samples were then reverse-transcribed using the First-Strand cDNA

Synthesis Kit (MBI Fermentas) in preparation for real-time RT-PCR using the QuantiTech SYBR

Green PCR kit in an Mx3000P QPCR system (Stratagene). Gene expression was determined

using the following formula: Ratio = (EcipI)∆Ct(control-test)/(E16SrRNA)∆Ct(control-test), where E = (10-1/slope)

Page 125: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

109

represents the efficiency of gene amplification. The 16S rRNA gene was used as internal

reference as we found the expression of this gene to be stable under the test conditions. All

assays were performed in triplicate with RNA isolated from three independent experiments and

using a P ≤0.01.

Biofilm assays

Static biofilms were developed in polystyrene microtiter plates at 37°C with 5% CO2 using semi-

defined minimal medium (SDM) containing 1% sucrose as previously described (Perry et al.,

2008). After 16 h of growth, the planktonic phase was carefully removed, and fresh SDM-

sucrose alone (control) or containing 2 µM synthetic CSP (sCSP) pheromone was overlayed

onto established biofilms and the plates were incubated for a further 5 h. The same experiment

was also repeated with SDM-sucrose or SDM-sucrose + sCSP supplemented with 50 U/ml

DNase I (Fermentas). The upper phase was then removed, and biofilms were allowed to dry

overnight before staining with crystal violet for biomass quantification. Purification and

quantification of eDNA in 16-h biofilms were performed according to (Rice et al., 2007). The

amount of eDNA was determined using real-time RT-PCR, using four sets of primers designed

to amplify genes randomly distributed across the S. mutans genome. CT expression values

were averaged, and normalized to the expression in the wild-type strain. All assays were

performed in triplicate from three independent experiments.

Page 126: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

110

3.4 Results and discussion

3.4.1 Low cell density-dependent expression of cipI is regulated by LiaR

In this work, we sought to characterize the regulation of the cell death immunity protein CipI.

Since previous findings indicated that CipI was differentially regulated from the death effector

protein CipB at low cell density, we first attempted to identify the regulator responsible for its

differential expression. We hypothesized that the regulator responsible for cipI expression at low

cell density would be one of the thirteen known TCSs in the S. mutans UA159 genome. TCSs

are typically composed of a membrane-bound histidine kinase sensor, which phosphorylates a

cytoplasmic response regulator when triggered by environmental stimuli (Stock et al., 2000). We

screened mutants defective in each TCS (Lévesque et al., 2007) for differential growth kinetics

under high concentrations of sCSP pheromone, conditions known to increase gene expression

of both the CipB bacteriocin and the CipI immunity protein (Perry et al., 2009). Inactivation of the

response regulator LiaR (originally referred to as RR11 (Li et al., 2002a; Perry et al., 2008;

Suntharalingam et al., 2009) resulted in an increased sensitivity to sCSP, while inactivation of its

cognate histidine kinase sensor LiaS (formerly known as HK11) showed an increased

resistance to sCSP compared to the wild-type (Fig. 3.1). These results implicate the LiaSR TCS

in regulation of cipI expression. To prove the direct regulation of cipI by LiaSR at low cell

density, we compared the expression of cipI between high and low cell density in the wild-type

strain, a ΔLiaS mutant and a ΔLiaR mutant. While the wild-type strain showed a highly

significant increase in cipI gene expression upon dilution from high to low cell density (120.6 ±

24.3-fold), little- to no-change in cipI expression was found in ΔLiaS and ΔLiaR mutant strains

(7.3 ± 7.6-fold and 17.3 ± 3.8-fold, respectively). Moreover, we found a 64.6 ± 4.4-fold increase

in liaR itself upon dilution of the wild-type strain from an overnight culture. Together, our results

Page 127: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

111

suggest that the LiaSR TCS is responsible for regulation of cipI gene expression at low cell

density.

Figure 3.1: Growth of S. mutans TCS mutants in the presence of 2 µM sCSP. All thirteen

TCSs in the UA159 genome were inactivated (Levesque et al., 2007) and assayed for their

growth kinetics in the presence of sCSP. The TCS composed of LiaS and LiaR showed

differential sensitivity to sCSP compared to the wild-type strain, with the histidine kinase sensor

deficient mutant (LiaS) showing increased resistance and the response regulator deficient

mutant (LiaR) showing increased sensitivity to sCSP.

3.4.2 CipI is protective at low cell density, while CipB is lethal at high cell density

Having now shown that cipI is differentially regulated at low cell density through the LiaSR TCS,

we next sought to determine if the up-regulation of cipI expression at low cell density protected

Page 128: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

112

S. mutans cells from CSP-induced autolysis via the action of CipB bacteriocin. To test this

hypothesis, we pre-grew aliquots of a UA159 overnight culture in THYE broth without sCSP for

up to 2.5 hr to induce expression of CipI at low cell density, before exposing them to 2 µM

sCSP. We found that inducing CipI expression before sCSP exposure had a protective effect on

the culture (Fig. 3.2).

Although CipI up-regulation at low cell density has a protective effect, cultures begin to

experience significant levels of autolysis as growth progresses in presence of sCSP,

culminating at stationary phase (Perry et al., 2009). How is the transition to autolysis

accomplished in these cultures? We hypothesized that death due to the CipB/CipI system would

occur at stationary phase if the CipB bacteriocin was expressed at higher levels than its CipI

immunity factor at high cell density. Indeed, the gene expression of CipB is induced 60.5 ± 1.2-

fold at stationary phase vs. early log phase, while the gene expression of CipI remains

unchanged (-0.6 ± 2.3-fold change in expression) as determined by real-time RT-PCR analysis.

It is tempting to speculate on the implications of these results in the physiological context of

cell death. Autolysis is often found linked to signals that convey high population density, since

the purpose of cell death in a unicellular organism is to provide benefits (e.g., provide nutrients,

space or transforming DNA; eliminate competition) to sibling cells (Claverys & Havarstein,

2007). Lysis of a sparse population serves no purpose since sibling cells are unlikely to benefit,

and is therefore counter-selected by evolution. The unlinked transcriptional units encoding CipB

and CipI may therefore serve to protect cells from lysis at low cell density to allow expansion of

the clonal population. Conversely, when nutrients become scarce at high cell density, autolysis

is triggered through CipB up-regulation to provide a competitive growth advantage to surviving

siblings. Nowhere is the above high-cell density-mediated altruistic process more likely to occur

than in a biofilm. In a previous report, we found that the expression of cipI is down-regulated

2.1-fold through LiaR in biofilm cells compared to their planktonic counterparts (Perry et al.,

Page 129: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

113

2008). In addition, others have reported the up-regulation of cipB expression in the biofilm

phase (Shemesh et al., 2007). We therefore hypothesized that the CipB/CipI autolysis cascade

is inhibited during rapid growth in planktonic cultures, but is induced during the high cell density

biofilm growth mode of S. mutans.

Figure 3.2: Growth kinetics of S. mutans UA159 wild-type strain in the presence of 2µM sCSP.

Overnight cultures were pre-grown for 2.5 hr before sCSP addition, to induce expression of the

CipI immunity protein-encoding gene at low cell density. Control cultures were exposed to sCSP

directly upon dilution from an overnight culture („no pre-growth‟), or were grown in the absence

of sCSP („control‟). The induction of cipI expression before sCSP addition confers a protective

effect on the culture, reflected in an increase in initial growth rate compared to the no pre-growth

curve.

Page 130: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

114

3.4.3 Cell death participates in biofilm formation through eDNA release

From the outset, both the CSP-ComDE and the LiaSR systems have been implicated in

S. mutans biofilm formation (Li et al., 2002a; Li et al., 2002b). Recently, Zhang and coll.

reported that exogenous CSP addition to S. mutans resulted in increased in both cell death and

biofilm biomass (Zhang et al., 2009). These authors also reported that scanning electron

microscopy of a mutant unable to synthesize the CSP signal molecule produced biofilms

composed of loosely attached, single cells, but that complementing the mutant with exogenous

sCSP resulted in formation of large aggregates with abundant extracellular matrix. Reports have

also suggested that autolysis is necessary for S. mutans biofilm formation (Wen & Burne, 2002),

and that CSP induces eDNA release (Petersen et al., 2005). However, none of these studies

has provided a mechanistic explanation for CSP-induced release of eDNA. We first sought to

link cell lysis in the biofilm to the CipB/CipI system by measuring the amount of eDNA in 16-hr

biofilms formed by the wild-type strain, ΔCipB and ΔCipI mutants. Using real-time RT-PCR, we

found a significant decrease in eDNA in the ΔCipB mutant biofilm and a significant increase in

eDNA in the ΔCipI mutant biofilm compared to the wild-type (Fig. 3.3). We conclude from this

result that death due to the CipB/CipI system can influence the amount of eDNA in the biofilm

matrix via autolysis.

To test whether the modulation of eDNA release via the CipB/CipI autolysis cascade affects

S. mutans biofilm biomass, 16-hr biofilms were treated with DNase I and the biomass quantified

by crystal violet staining. DNase treatment of biofilms formed by the wild-type strain decreased

their biomass by more than 20% (Fig. 4). In either case, this result confirms previous reports

that eDNA plays an important role in S. mutans biofilm formation (Petersen et al., 2005). We

next assayed the ΔCipB and ΔCipI mutant strains for their ability to form biofilms. Importantly,

ΔCipI formed biofilms with greater biomass than the wild-type strain even in the absence of

Page 131: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

115

Figure 3.3: Fold-change in quantity of eDNA in ΔCipI and ΔCipB mutant biofilms compared to

the UA159 wild-type control. Quantitative real-time RT-PCR was used to amplify four randomly

selected chromosomal genes from DNA extracted from the extracellular matrix of UA159, ΔCipI

and ΔCipB biofilms. Amplification values for UA159 were arbitrarily set at one, and results are

expressed as a fold-change relative to the UA159 wild-type. Standard deviation represents the

variation in amplification across the four chromosomal genes selected. Both CipI and CipB

results represent significant differences (P value ≤ 0.01) from the UA159 wild-type control.

Page 132: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

116

sCSP, as observed by an increased biomass of ~ 30% (Fig. 3.4). The ΔcipI mutant cells

possess an intact CSP-encoding gene, and are able to produce endogenous CSP pheromone

during biofilm formation. When grown planktonically, the ΔCipI mutant is approximately 10-times

more sensitive to sCSP than the wild-type (unpublished data), and is likely succumbing to the

accumulation of endogenous CSP in the biofilm. The increase in biomass is due solely to eDNA

release, since treatment with DNase I restored the biofilm biomass to wild-type DNase-treated

levels (Fig. 3.4). Interestingly, the ΔCipB mutant also formed more biofilm biomass than the

wild-type. However, when the ΔCipB mutant biofilms were treated with DNase I, the biofilm

biomass was unchanged (Fig. 3.4). This result indicates that ΔCipB has a growth advantage in

the biofilm due to its resistance to autolysis, and the increase in biofilm biomass appears to be

mostly due to increasing cell number.

Finally, we added sCSP to established biofilms to induce CipB-mediated autolysis. As

expected, adding sCSP to UA159 and ΔCipI mutant biofilms increased the biofilm biomass

beyond CSP-uninduced conditions (Fig. 3.4). The increase in biomass was again due to eDNA

in the extracellular matrix, since DNase I restored the biomass to wild-type levels. These results

suggest that cell death has a positive impact on biofilm biomass through the release of eDNA

into the extracellular matrix, through the endogenous CSP-induced CipB/CipI-mediated cell

death pathway.

Page 133: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

117

Figure 3.4: Biofilm biomass of S. mutans UA159 wild-type strain, ΔCipB and ΔCipI mutants.

Biofilms were allowed to develop for 16 h in SDM-sucrose. The planktonic phase was then

removed, and fresh SDM-sucrose alone (control) or supplemented with 2 µM sCSP was

overlayed onto the biofilms and incubated for a further 5 h. The same experiment was repeated

with SDM-sucrose supplemented with 50 U/ml DNase in the overlay. Biofilms were washed

once with sterile dH2O before quantification. Quantification of biomass was performed by optical

density (OD) of crystal violet-stained UA159, ΔCipB and ΔCipI biofilms. Results are expressed

as a % increase in biofilm biomass compared to the biofilm biomass of the UA159 wild-type

control, which was set at 100%. Significant (P <0.02) increase in biomass compared to the

UA159 wild-type control condition is denoted with a „*‟, while significant decrease in biomass

compared to the UA159 wild-type control is denoted with a „†‟.

Page 134: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

118

3.5 Conclusions

The lifecycle of a biofilm includes periods of both exponential growth and of nutrient limitation

(Stoodley et al., 2002). The cooperative nature of biofilm growth has been likened to the task-

sharing behavior common in higher-order multicellular organisms. As such, the biofilm lifestyle

may permit altruistic behaviors like autolysis in unicellular prokaryotic organisms, which can

contribute nutrients for the continued survival of siblings in a stressed population. Our recent

findings led us to propose a mechanistic explanation implicating CSP pheromone during the

development of S. mutans biofilm. At low cell density, S. mutans up-regulates expression of the

CipI immunity protein through the LiaSR TCS. Up-regulation of CipI has a protective effect on

the cell, and allows the culture to proliferate under favorable environmental conditions.

Conversely, in the high cell density biofilm environment, the high concentrations of CSP

pheromone signal up-regulation of the CipB autolysis effector through the ComDE TCS.

Altruistic autolysis in the S. mutans biofilm contributes nutrients for the continued survival of the

population as a whole, as well as eDNA to the extracellular matrix.

Page 135: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

119

3.6 Acknowledgements

This study was supported by CIHR-Priority Announcement IMHA Grant FRN-90114 (to

C.M.L.) and by National Institute of Dental and Craniofacial Research grant R01

DE013230-08 to D.G.C. DGC is supported by a Canada Research Chair. J.A.P is

supported by a CIHR Strategic Training Fellowship in Cell Signaling in Mucosal

Inflamation and Pain.

Page 136: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

120

3.7 References

Beloin C & Ghigo JM (2005) Finding gene-expression patterns in bacterial biofilms. Trends

Microbiol 13: 16-19.

Branda SS, Vik A, Friedman L & Kolter R (2005) Biofilms: the matrix revisited. Trends Microbiol

13: 20-26.

Claverys JP & Havarstein LS (2007) Cannibalism and fratricide: mechanisms and raisons d'etre.

Nat Rev Microbiol 5: 219-229.

Costerton JW, Stewart PS & Greenberg EP (1999) Bacterial biofilms: a common cause of

persistent infections. Science 284: 1318-1322.

Costerton JW, Lewandowski Z, DeBeer D, Caldwell D, Korber D & James G (1994) Biofilms, the

customized microniche. J Bacteriol 176: 2137-2142.

Dunne WM Jr (2002) Bacterial adhesion: seen any good biofilms lately? Clin Microbiol Rev 15:

155-166.

Flemming HC, Neu TR & Wozniak D (2007) The EPS matrix: the "house of biofilm cells". J

Bacteriol 189: 7945-7947.

Hasona A, Crowley PJ, Lévesque CM, Mair RW, Cvitkovitch DG, Bleiweis AS & Brady LJ (2005)

Streptococcal viability and diminished stress tolerance in mutants lacking the signal

recognition particle pathway or YidC2. Proc Natl Acad Sci USA 102: 17466-17471.

Kreth J, Merritt J, Zhu L, Shi W & Qi F (2006) Cell density- and ComE-dependent expression of

a group of mutacin and mutacin-like genes in Streptococcus mutans. FEMS Microbiol Lett

265: 11-17.

Lawrence JR, Korber DR, Hoyle BD, Costerton JW & Caldwell DE (1991) Optical sectioning of

microbial biofilms. J Bacteriol 173: 6558-6567.

Lévesque CM, Mair RW, Perry JA, Lau PCY, Li Y-H & Cvitkovitch DG (2007) Systemic

inactivation and phenotypic characterization of two-component systems in expression of

Streptococcus mutans virulence properties. Lett Appl Microbiol 45: 398-404.

Li YH, Lau PC, Lee JH, Ellen RP & Cvitkovitch DG (2001a) Natural genetic transformation of

Streptococcus mutans growing in biofilms. J Bacteriol 183: 897-908.

Li YH, Hanna MN, Svensater G, Ellen RP & Cvitkovitch DG (2001b) Cell density modulates acid

adaptation in Streptococcus mutans: implications for survival in biofilms. J Bacteriol 183:

6875-6884.

Page 137: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

121

Li YH, Lau PC, Tang N, Svensater G, Ellen RP & Cvitkovitch DG (2002a) Novel two-component

regulatory system involved in biofilm formation and acid resistance in

Streptococcus mutans. J Bacteriol 184: 6333-6342.

Li YH, Tang N, Aspiras MB, Lau PC, Lee JH, Ellen RP & Cvitkovitch DG (2002b) A quorum-

sensing signaling system essential for genetic competence in Streptococcus mutans is

involved in biofilm formation. J Bacteriol 184: 2699-2708.

Loesche WJ (1986) Role of Streptococcus mutans in human dental decay. Microbiol Rev 50:

353-380.

Parsek MR & Greenberg EP (2005) Sociomicrobiology: the connections between quorum

sensing and biofilms. Trends Microbiol 13: 27-33.

Perry JA, Jones MB, Peterson SN, Cvitkovitch DG & Lévesque CM (2009) Peptide alarmone

signalling triggers an auto-active bacteriocin necessary for genetic competence. Mol

Microbiol. 72: 905-17.

Perry JA, Lévesque CM, Suntharaligam P, Mair RW, Bu M, Cline RT, Peterson SN &

Cvitkovitch DG (2008) Involvement of Streptococcus mutans regulator RR11 in oxidative

stress response during biofilm growth and in the development of genetic competence. Lett

Appl Microbiol 47: 439-444.

Petersen FC, Tao L & Scheie AA (2005) DNA binding-uptake system: a link between cell-to-cell

communication and biofilm formation. J Bacteriol 187: 4392-4400.

Rice KC, Mann EE, Endres JL, Weiss EC, Cassat JE, Smeltzer MS & Bayles KW (2007) The

cidA murein hydrolase regulator contributes to DNA release and biofilm development in

Staphylococcus aureus. Proc Natl Acad Sci USA 104: 8113-8118.

Shapiro JA (1998) Thinking about bacterial populations as multicellular organisms. Annu Rev

Microbiol 52: 81-104.

Shemesh M, Tam A & Steinberg D (2007) Differential gene expression profiling of

Streptococcus mutans cultured under biofilm and planktonic conditions. Microbiology 153:

1307-1317.

Stock AM, Robinson VL & Goudreau PN (2000) Two-component signal transduction. Annu Rev

Biochem 69: 183-215.

Stoodley P, Sauer K, Davies DG & Costerton JW (2002) Biofilms as complex differentiated

communities. Annu Rev Microbiol 56: 187-209.

Suntharalingam P, Senadheera MD, Mair RW, Lévesque CM & Cvitkovitch DG (2009) The

LiaFSR system regulates the cell envelope stress response in Streptococcus mutans. J

Bacteriol 191: 2973-2984.

Page 138: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

122

Waters CM & Bassler BL (2005) Quorum sensing: cell-to-cell communication in bacteria. Annu

Rev Cell Dev Biol 21: 319-346.

Wen ZT & Burne RA (2002) Functional genomics approach to identifying genes required for

biofilm development by Streptococcus mutans. Appl Environ Microbiol 68: 1196-1203.

Whitchurch CB, Tolker-Nielsen T, Ragas PC & Mattick JS (2002) Extracellular DNA required for

bacterial biofilm formation. Science 295: 1487.

Zhang K, Ou M, Wang W & Ling J (2009) Effects of quorum sensing on cell viability in

Streptococcus mutans biofilm formation. Biochem Biophys Res Commun 379: 933-938.

Page 139: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

123

Chapter 4: Involvement of Streptococcus mutans regulator RR11 in

oxidative stress response during biofilm growth and in the

development of genetic competence

JA Perry, CM Lévesque, P Suntharaligam, RW Mair, M Bu, RT Cline, SN Peterson and DG Cvitkovitch. 2008. Lett Appl Microbiol 47: 439-44.

Page 140: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

124

4.1 Abstract

Aims: To identify the genes regulated by RR11, the regulator component of the

Streptococcus mutans HK/RR11 two-component signal transduction system.

Methods and Results: The S. mutans RR11-encoding gene was inactivated and the effect of

gene disruption on the cell‟s ability to form biofilms was tested. DNA microarray used to

decipher RR11-regulated genes during biofilm growth showed that ~5% of the UA159 genome

underwent a significant change in gene expression. RR11 was found to regulate 174 genes,

including genes involved in competence development, stress-response, and cell division.

Phenotypic assessment of biofilms showed a reduction in biomass in cells lacking RR11

following exposure to oxidative stress. RR11 defective cells showed ~20-fold reduction in

transformation efficiency.

Conclusions: Target genes controlled by RR11during biofilm growth were identified by a

comparison of transcriptional profiles between an RR11 mutant and the parental strain. The

results demonstrated that RR11 is involved in the control of processes such as the formation of

biofilm under oxidative stress and development of genetic competence.

Significance and Impact of Study: Mapping the RR11 signal transduction pathway in S.

mutans biofilms has determined its involvement in stress tolerance. Targeting this pathway may

lead to future antimicrobial therapies designed specifically for biofilm infections.

Page 141: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

125

4.2 Introduction

Streptococcus mutans causes dissolution of tooth enamel by producing acid from dietary

carbohydrates, and is considered one of the principal etiological agents of human dental

caries(Mitchell, 2003). It has evolved a biofilm lifestyle to survive and persist in the oral cavity,

where two-component systems are employed to sense and respond to environmental stimuli.

Changing environmental conditions are sensed by the membrane-bound histidine kinase (HK)

receptor, resulting in autophosphorylation. The resulting phosphoryl group is transferred to the

cognate cytosolic response regulator (RR), which acts as a transcriptional regulator (Stock et

al., 2000). One of the thirteen putative two-component systems found in S. mutans UA159

consists of the HK11 sensor and the regulator RR11(Ajdic et al., 2002; Levesque et al., 2007).

This system has a role in cell segregation, and in the response to thermal, oxidative, and acid

stresses (Biswas et al., 2007; Li et al., 2002a). Although HK/RR11 has been implicated in

S. mutans biofilm development, the genes controlled by RR11 during biofilm growth are still

unknown (Li et al., 2002a).

In the present study, DNA microarrays were used to determine gene expression under

biofilm and planktonic growth phases in wild-type S. mutans and its rr11– defective derivative to

determine the role of RR11 in biofilm development. Microarray data indicated that RR11 is

responsible for the induction of genes involved in general stress response in the biofilm. The

ability of an RR11-defective mutant to grow under stress in the biofilm and become competent

for genetic transformation was also investigated.

Page 142: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

126

4.3 Materials and methods

Bacterial strains and growth conditions

Bacterial strains are found in Table 4.1. Antibiotics were added when required as follows: 10 µg

ml-1 erythromycin, 1000 µg ml-1 spectinomycin, or 500 µg ml-1 kanamycin. Mutants were

constructed in strain UA159 as described previously (Lau et al., 2002). The erythromycin and

spectinomycin resistance cassettes were amplified from plasmid pALN122 (Macrina et al.,

1983) and pDL277 (Leblanc et al., 1992) using the primer pairs Erm19/Erm20 (Levesque et al.,

2005) and SpecF/SpecR (Aspiras et al., 2004), respectively. Primer sequences for mutant

construction are available upon request.

Table 4.1. Bacterial strains used in this study. All strains were grown in Todd Hewitt Yeast

Extract (THYE) broth at 37ºC in air with 5% CO2.

Strain

Relevant characteristicsa

Source

UA159 Wild-type strain; Ems Kms Sps J. Ferretti, U. of

Oklahoma

SMRR11 UA159∆(rr11); Emr This work

SMComDE UA159∆(comDE); Emr This work

SMCiaHR UA159∆(ciaHR); Emr This work

SMDE11 UA1591∆(comDE, rr11); Emr Spr This work

aEm: erythromycin; Km: kanamycin; Sp: spectinomycin.

Page 143: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

127

In vitro model for growing biofilms

Static biofilms were developed on polystyrene microtiter plates in semi-defined minimal medium

(SDM) (Li et al., 2002a) supplemented with 44 mmol l-1 glucose. Plates were incubated at 37°C

in air with 5% CO2 for 16 h (standard assay for biofilm quantification), or for 8 h before the

addition of SDM-glucose supplemented with either 0.5 mmol l-1 H2O2, 2.5% NaCl, or SDM-

glucose adjusted to pH5 with HCl for a further 8 h (stress assay). Planktonic cells were removed

after incubation, and the biofilms were air dried overnight before quantification as described

previously (Levesque et al., 2005).

Scanning electron microscopy

Scanning electron microscopy (SEM) was performed on 16 h biofilms grown on glass discs

according to the standard assay. Biofilms were washed with sterile phosphate-buffered saline

(PBS, pH 7.2), dehydrated through ethanol rinses, critical point dried with liquid CO2, mounted,

and sputter coated with gold. Samples were then examined using a scanning electron

microscope (model S-2500; Hitachi Instruments, San Jose, CA).

DNA microarrays

Biofilms of UA159 and SMRR11 were grown for 16 h in SDM-glucose as described in the

standard assay. Both planktonic and biofilm cell pools were harvested, washed once in PBS,

resuspended in Trizol reagent (Invitrogen), and processed with the Bio101 Fast Prep system

(Qbiogene). DNA-free RNA samples were labeled and prepared for hybridization according to

the PFGRC protocol (http://pfgrc.tigr.org/protocols/M007.pdf). Microarray chips were scanned

using a Gene Pix 4000B (Axon). The software package TM4 Microarray Software Suite

(http://www.tm4.org/) was used for data analysis. Microarray assays was performed on three

independent RNA isolations, and validated by quantitative real-time RT-PCR using the

Page 144: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

128

QuantiTech SYBR Green PCR kit in a Mx3000P QPCR system (Stratagene). Statistical

significance was determined using Student‟s t-test and a P value of <0.05.

Transformation experiments

One µg of plasmid pDL289 was added to growing cultures at an OD600 of 0.1 both in the

presence and absence of 0.4μM synthetic CSP, and incubated at 37°C for 2.5 h. Cultures were

then gently sonicated, and spread on THYE agar plates. Transformation efficiency was

expressed as the percentage of kanamycin resistant transformants over the total number of

recipient cells.

4.4 Results

4.4.1 Phenotypic characterization of Δrr11 defective mutant

The RR11-encoding gene of the S. mutans HK/RR11 TCS was successfully inactivated by

deletion-insertion mutagenesis. No significant difference in growth kinetics was observed during

planktonic growth. Li et al. (2002) reported that their S. mutans strain NG8 rr11–defective

mutant formed significantly less biofilm than the parent strain. In our experiments using

background strain UA159, SMRR11 mutant cells formed stable and reproducible biofilms with a

biomass of 12.2 %± 4.9% less than the wild-type. A closer examination by SEM revealed that

biofilms formed by the SMRR11 mutant exhibited an altered structure, with larger channels

visible at low magnifications and a difference in cell morphology observed higher magnifications

(Figure 4.1).

Page 145: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

129

Figure 4.1. S. mutans UA159 and RR11- mutant biofilm formation. Scanning electron

micrographs of UA159 and SMRR11 biofilms accumulated on the surface of glass

discs. Magnifications, 1 K (top panels), and 60 K

4.4.2 Microarray identification of RR11-regulated genes involved in the stress

response.

To investigate the morphological differences observed in SMRR11 biofilms, gene expression

profiles of the UA159 wild-type and SMRR11 were analyzed using DNA microarrays.

Expression data comparing biofilm and planktonic growth phases in the wild-type and mutant

strains suggested that RR11 directly and/or indirectly regulated 174 genes (~9% of the genome)

Page 146: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

130

in S. mutans biofilms. Of these, several genes encoding proteins involved in general stress

response were found differentially regulated (Table 4.2).

The microarray results also identified other TCSs whose expression was altered in SMRR11

mutant biofilms. Among these genes was comD, the receptor for the CSP pheromone. ComDE

was originally characterized for its role in the development of genetic competence in S.

pneumoniae (Håvarstein et al., 1995; Håvarstein et al., 1996) and S. mutans(Li et al., 2002a),

but has recently been implicated in the control of the stress-responsive autolysis pathway in

pneumococci (Guiral et al., 2005). The gene encoding the RR of the CiaHR TCS also showed

altered expression in our microarray. CiaHR is involved in stress tolerance and competence

development in S. mutans (Ahn et al., 2006). Finally, the gene encoding RR9 of the S. mutans

TCS HK/RR9 was also upregulated. This TCS has recently been shown to be involved in S.

mutans acid survival (Levesque et al., 2007).

4.4.3 SMRR11 biofilms under oxidative, osmotic and acid stresses

During the preparation of this manuscript, RR11 was shown to be involved in oxidative and

thermal stress responses in planktonic cultures (Biswas et al., 2007). Since changes in

expression of the above mentioned RR11-regulated stress response genes could impact the

formation of S. mutans biofilms through a reduced ability to respond to stress in the biofilm

environment, we examined the ability of SMRR11 biofilms to grow in the presence of oxidative,

osmotic, and acid stress conditions. A significant reduction in biofilm biomass (37.8 ± 17.1%,

P = 0.002) was observed when SMRR11 biofilms were grown in the presence of H2O2

compared to UA159 grown under the same conditions. We found no statistically significant

Page 147: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

131

Table 4.2. Genes potentially regulated by RR11 in S. mutans growing in biofilms

Locus

(NCBI)

Description/putative function Relative fold-

change

UA159 SMRR11

SMU.91 peptidyl-prolyl isomerase RopA

(trigger factor)

n

c

+2.2

SMU.228 alkaline-shock protein homolog nc +3.8

SMU.403 DNA-damage-inducible protein P nc +2.5

SMU.949 ATP-dependent protease Clp,

ATPase subunit ClpX

–2.2 nc

SMU.1063 ABC transporter, ATP-binding,

proline/glycine betaine

nc +2.5

SMU.1129 response regulator CiaR +2.6 nc

SMU.1672 ATP-dependent Clp protease,

proteolytic subunit

nc +2.0

SMU.1916 histidine kinase of the

competence regulon

–2.1 nc

SMU.1964 response regulator nc –2.0

SMU.2030 transcriptional regulator CtsR nc +2.0

SMU.2116 osmoprotectant amino acid ABC

transporter, ATP-binding

nc +5.1

nc= no change in gene expression

Page 148: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

132

difference between the growth of the wild-type and RR11- biofilms under osmotic and acid

stress. Under all three stress conditions assayed, planktonic cultures of UA159 and

SMRR11grew with identical kinetics, indicating that the defect in biofilm formation under

oxidative stress was due to true biofilm-specific growth impairments, and not simply slower

growth of the culture in both phases.

4.4.4 Regulatory role for RR11 in competence development

Our microarray demonstrated that the gene encoding the ComD receptor for the CSP

pheromone was likely regulated by RR11. Because competence and stress response in S.

pneumonie are linked through ComDE, we also examined the competence phenotype of the

SMRR11 mutant. Ahn et al. (2006) suggested that more than one TCS may be involved in

triggering competence induction, and proposed that CiaHR, and possibly some other

unidentified regulators, integrate CSP signals (Ahn et al., 2006). We hypothesized that RR11

may be one such additional regulator, that could be responsible for integrating stress signals in

the biofilm to trigger competence. We investigated the role of RR11 in the development of

genetic competence by evaluating the ability of comDE- (SMComDE), rr11- (SMRR11) and

rr11/comDE (SMDE11) mutants to be transformed with plasmid DNA (Fig. 4.2). Our results

demonstrated that SMRR11 had a ~20-fold reduction in transformation efficiency vs. the wild-

type strain in the absence of CSP. As expected, inactivation of comDE diminished the

transformation efficiency by several-fold. Surprisingly, the SMDE11 double mutant behaved like

the wild-type strain in the absence of CSP, regardless of whether CSP was added. This finding

led us to hypothesize that both ComE and RR11 may negatively regulate a third regulator in the

CSP-independent pathway. To test whether CiaR could be involved in CSP-independent

competence induction, transformation efficiency was measured in the ciaHR– deficient mutant.

However, inactivation of ciaHR had no impact on competence (results not shown).

Page 149: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

133

Figure 4.2. Transformation efficiency of S. mutans wild-type and mutant strains. The

transformation of S. mutans strains with plasmid pDL289 is plotted with and without the addition

of synthetic CSP. Transformation efficiency is expressed as the percentage of viable cells

transformed to kanamycin resistance. The results are expressed as the mean + standard

deviation of at least two independent experiments.

4.5 Discussion

Environmental conditions in the oral biofilm are highly variable with respect to pH, oxygen

and osmotic balance. Shifts from neutral pH to as low as 3.0 occur during host ingestion of

dietary carbohydrates, oxygen gradients occur in the oral biofilm, and salts may accumulate

from tooth demineralization. Thus, the ability of S. mutans to adapt to its environment is vital to

its fitness. The involvement of two-component signal transduction systems in environmental

stress response has been characterized during planktonic growth of S. mutans, but few studies

have examined the role of these systems in the stress response in the biofilm environment. Our

Page 150: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

134

aim was to characterize the response regulator component of the HK11/RR11 TCS with respect

to its role in biofilm development. Towards this goal, we examined RR11- biofilms for qualitative

differences by SEM, and for gene expression differences by DNA microarray. We discovered

phenotypic differences between the wild-type and SMRR11 via stress tolerance and

competence assays.

Although only a small difference in biofilm biomass resulted from deletion of RR11 in S.

mutans strain UA159, SEM data indicated that both biofilm structure and cellular morphology

were altered in its absence. To investigate the transcriptome underlying these changes in

morphology, a DNA microarray was performed. Several stress response genes were identified

in our microarray analysis, including ropA and clpP. RopA is a molecular chaperone that

functions in protein biogenesis and stress survival (Hesterkamp and Bukau, 1996). ClpP is the

proteolytic subunit of the ATP-dependent Clp protease, which performs protein reactivation and

degradation (Porankiewicz et al., 1999). Interestingly, Wen et al. (2005) found that an S. mutans

ropA mutant also showed longer chain length in broth and altered biofilm architecture, which

they attributed to alterations in protein trafficking. clpP null mutants are also defective in genetic

competence and biofilm formation, and are more susceptible to stress (Lemos and Burne, 2002;

Wen et al., 2005). These results suggested that the changes in ropA and/or clp gene expression

through RR11 may be responsible for the altered cell morphology and biofilm structure seen in

the SMRR11 mutant.

Specific stress response proteins like the osmoprotectant ABC transporters encoded by

SMU.2116 and SMU.1063 were also up-regulated in RR11- biofilms. SMU.2116 is highly

homologous to OpuCA of Streptococcus agalactiae, while SMU.1063 shares high identity with a

proline/glycine betaine transporter found in Lactococcus lactis. Both these transporters have

been shown to be upregulated under osmotic stress conditions in planktonically grown S.

mutans, and implicated in the survival of the organism under those stress conditions (Abranches

Page 151: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

135

et al., 2006) (Abranches et al., 2006). The possible regulatory role for RR11 in the general

stress response prompted us to examine the growth of RR11-deficient biofilms under stress.

Biswas et al. (2007) have recently shown that RR11 is involved in the oxidative stress response

in the planktonic phase, and our results indicated that RR11 may also be important in the

response to oxidative stress in the biofilm. Combining our microarray analysis with our

physiological data suggests that the increased susceptibility to oxidative stress may occur due

to RR11‟s role in regulating the turnover of damaged proteins via RopA and ClpP, which may

result in the abnormal biofilm architecture and cell morphology observed by SEM.

Exciting evidence has recently emerged linking the competence cascade to the general

stress response program in pneumococci (Claverys and Havarstein, 2007; Guiral et al., 2005).

These authors have shown that a population under antibiotic stress triggers the death of

damaged cells via the CSP signaling molecule and ComDE. A similar link between competence

and cell death exists in S. mutans (Leblanc et al., 1992; Perry et al., 2009). Due to microarray

evidence linking RR11 and ComD, and physiological evidence of a role for RR11 in stress

response, we hypothesized that the competence phenotype would also be affected in SMRR11.

Indeed, competence was reduced ~20-fold in the absence of CSP in SMRR11. Based on our

competence data, we have proposed a model for the S. mutans CSP-dependent and CSP-

independent competence regulatory networks (Figure 4.3). In this model, the ComDE TCS is

the primary circuit sensing CSP, and induces a high level of transformation at high levels of

CSP. At low levels of CSP, the major competence system remain inactive, and

unphosphorylated ComD may cross-regulate RR11 to induce a basal level of genetic

competence. Investigations are ongoing in our lab to elucidate the role of stress in the

development of genetic competence, including the role of RR11 in this pathway.

Page 152: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

136

Figure 4.3 Model for competence development in S. mutans. This model integrates the CSP-

dependent and CSP-independent pathways (see text for details).

4.6 Acknowledgements

This study was supported by National Institute of Dental and Craniofacial Research

grant R01 DE013230. DGC is supported by a Canada Research Chair. JAP and PS

are both supported by CIHR Cell Signals Fellowships. JAP, PS and MB performed

experiments; JAP, CL, RM, RC, SP and DGC contributed to experimental design, data

analysis and/or to the writing of this manuscript. The authors thank Robert Chernecky

for technical services.

Page 153: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

137

4.7 References

Abranches, J., J. A. Lemos, and R. A. Burne. 2006. Osmotic stress responses of Streptococcus mutans UA159. FEMS Microbiology Letters 255:240-246. Ahn, S. J., Z. T. Wen, and R. A. Burne. 2006. Multilevel control of competence development and stress tolerance in Streptococcus mutans UA159. Infect Immun 74:1631-1642. Biemans-Oldehinkel, E., M. K. Doeven, and B. Poolman. 2006. ABC transporter architecture and regulatory roles of accessory domains. FEBS Lett 580:1023-1035. Biswas, I., L. Drake, D. Erkina, and S. Biswas. 2007. Involvement of Sensor Kinases in the Stress Tolerance Response of Streptococcus mutans. J. Bacteriol.:JB.00990-00907. Claverys, J. P., and L. S. Havarstein. 2007. Cannibalism and fratricide: mechanisms and raisons d'etre. Nat Rev Microbiol 5:219-229. Guiral, S., T. J. Mitchell, B. Martin, and J. P. Claverys. 2005. Competence-programmed predation of noncompetent cells in the human pathogen Streptococcus pneumoniae: genetic requirements. Proc Natl Acad Sci U S A 102:8710-8715. Havarstein, L. S., G. Coomaraswamy, and D. A. Morrison. 1995. An unmodified heptadecapeptide pheromone induces competence for genetic transformation in Streptococcus pneumoniae. Proc Natl Acad Sci U S A 92:11140-11144. Havarstein, L. S., P. Gaustad, I. F. Nes, and D. A. Morrison. 1996. Identification of the streptococcal competence-pheromone receptor. Mol Microbiol 21:863-869. Hesterkamp, T., and B. Bukau. 1996. The Escherichia coli trigger factor. FEBS Lett 389:32-34. Lemos, J. A., and R. A. Burne. 2002. Regulation and Physiological Significance of ClpC and ClpP in Streptococcus mutans. J Bacteriol 184:6357-6366. Levesque, C. M., R. W. Mair, J. A. Perry, P. C. Y. Lau, Y.-H. Li, and D. G. Cvitkovitch. 2007. Systemic inactivation and phenotypic characterization of two-component systems in expression of Streptococcus mutans virulence properties. Letters in Applied Microbiology 45:398-404. Li, Y. H., N. Tang, M. B. Aspiras, P. C. Lau, J. H. Lee, R. P. Ellen, and D. G. Cvitkovitch. 2002. A quorum-sensing signaling system essential for genetic competence in Streptococcus mutans is involved in biofilm formation. J Bacteriol 184:2699-2708. Porankiewicz, J., J. Wang, and A. K. Clarke. 1999. New insights into the ATP-dependent Clp protease: Escherichia coli and beyond. Mol Microbiol 32:449-458. Prudhomme, M., L. Attaiech, G. Sanchez, B. Martin, and J.-P. Claverys. 2006. Antibiotic Stress Induces Genetic Transformability in the Human Pathogen Streptococcus pneumoniae. Science 313:89-92. Wen, Z. T., P. Suntharaligham, D. G. Cvitkovitch, and R. A. Burne. 2005. Trigger factor in Streptococcus mutans is involved in stress tolerance, competence development, and biofilm formation. Infect Immun 73:219-225.

Page 154: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

138

Chapter 5: Summary and Conclusions

Page 155: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

139

5.1 Summary of Dissertation

This dissertation examines the biological role of cell death and lysis in the biofilm-forming

organism Streptococcus mutans. We have shown that the CSP peptide pheromone is an

inducible signal in S. mutans, which may communicate stress in the population through ComDE.

The result of CSP upregulation is the induction of cell death and lysis in a fraction of the

population through intracellular accumulation of the auto-active bacteriocin CipB. The finding

that a bacteriocin may induce death through intracellular action is a novel finding in the field. We

have also characterized the mechanism of immunity to CSP-induced cell death, which occurs

through the differential regulation of the CipI immunity protein at low cell density via the LiaFSR

(formerly HK/RR11) regulatory system. This regulation is made possible due to the physical

separation of cipB and cipI on the chromosome, and allows for S. mutans survival at low cell

density. Further work also elucidated the LiaFSR regulon in the biofilm (which includes

regulation of cipI expression), and demonstrated a role for this signalling system in the

regulation of oxidative stress tolerance in the biofilm.

In the high cell density biofilm environment, the CipB/CipI cell death pathway contributes to

release of DNA into the extracellular matrix through cell lysis. This eDNA contributes to the

stability of the biofilm. Finally, we also provide evidence that the CipB/CipI death pathway is

involved in genetic competence, which we suggest may contribute to exchange of fitness

enhancing genes under stress and contribute to the evolutionary fitness of the organism.

Page 156: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

140

5.2 General Discussion

5.2.1 Peptide pheromone-induced cell death

Previous reports in both S. pneumoniae and S. mutans suggested that the CSP pheromone

was involved in inducing bacterial cell death (Dagkessamanskaia et al., 2004; Guiral et al.,

2005; Qi et al., 2005). We set out to characterize the response to CSP in S. mutans, framed in

the physiological context of elevated concentrations of the peptide induced by stress. Although

streptococcal CSP has traditionally been defined as a quorum sensing signal, our study and

those in S. pneumoniae have expanded that view by suggesting that CSP may also be an

inducible „alarmone‟-type molecule, capable of signalling stress in the population. However, we

suggest that in fact high cell density is a stress itself, and that quorum sensing is (and always

has been) a stress response system.

Our results showed that high concentrations of CSP induce lysis in a fraction of the S.

mutans population by intracellular accumulation of the auto-active bacteriocin CipB. This

mechanism is in contrast to that observed in S. pneumoniae, in which cell lysis is accomplished

in trans by competent cells expressing the CbpD protease or the two-peptide bacteriocin CibAB

on their cell surface. However, the intracellular action of CipB makes sense in the broader

context of biofilm growth, since expression of an auto-active death peptide on the cell surface

has destructive potential in the tightly packed biofilm environment. In fact, we suggest that

export of CipB is a detoxification mechanism, since no susceptibility to the exported form of the

bacteriocin has been noted for S. mutans or any of its normal target organisms in the oral

biofilm. The intracellular mechanism of action also prevents the lysis of the entire population,

Page 157: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

141

since S. mutans has only a single pherotype of CSP. While S. pneumoniae has the capacity to

differentially express immunity to cell death through the production of multiple CSP pherotypes,

S. mutans safe-guards against lysis of the whole population by triggering cell death

intracellularly.

We found that cell death in the biofilm may contribute extracellular DNA (eDNA) to

strengthen the biofilm matrix. Bayles has proposed a role for autolysis in biofilm formation by S.

aureus (Bayles, 2007). He argued that the biofilm lifestyle is akin to a multi-cellular organism in

its specialization of function, and that death of a sub-population is a natural extension of that

lifestyle. We propose a similar role for CSP-induced cell death in S. mutans. In the high density

oral biofilm where environmental stresses abound, the ability to form a biofilm and survive as a

sessile community is a strong evolutionary pressure. The act of cellular suicide under stress (or

high CSP concentration) provides both nutrients for continued growth and added protection in

the form of an enhanced extracellular matrix containing eDNA.

5.2.2 Immunity to peptide induced cell death

Although most bacteriocins are co-transcribed with their cognate immunity genes, it is not

without precedent to find an immunity gene un-linked on the chromosome (Diep et al., 2007).

We found that the CipI immunity protein-encoding gene was differentially regulated from the

CipB bacteriocin, and suggested that this differential expression necessitated the duplication of

immunity elsewhere in the genome. The up-regulation of CipI expression (by either plasmid-

based over-expression or through pre-growth of cultures prior to sCSP exposure) was shown to

Page 158: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

142

be protective in cultures at low cell density. In the context of stress and biofilm formation, this

result makes intuitive sense. If CSP (and autolysis) is high during stress and at high cell density

in the biofilm, low CSP (or density) should signal the absence of stress, and trigger survival and

proliferation. It is tempting to speculate that such conditions are met in the planktonic

population that departs the biofilm in the last stages of its lifecycle.

We found that the LiaFSR signalling system governs the up-regulation of CipI at low cell

density, and that LiaR (formerly RR11) is responsible for the down-regulation of cipI in biofilm

cells. We (and others (Li et al., 2002a)) found the LiaR mutant formed biofilms with aberrant

architecture. In the absence of LiaR in the biofilm, the cell death pathway would show

diminished activity due to an inability to down-regulate cipI. Could this imbalance lead to the

altered architecture we observed? Although it is impossible to say conclusively, the results of

our biofilm experiments suggest that the opposite is certainly true: a biofilm formed in the

absence of CipI has more biomass than the wild-type through the release of eDNA. Our

experiments with LiaR/RR11 biofilms also suggested potential cross-talk between ComD and

LiaR, albeit in the context of genetic competence. It has been previously suggested that these

two signalling systems can both respond to the CSP pheromone (Li et al., 2002a). Results

presented in both Chapters 3 and 4 of this dissertation support a link between these two

signalling systems.

5.2.3 Peptide-induced cell death in genetic competence

The release of DNA into the extracellular matrix has been shown to be essential for proper

biofilm architecture (Whitchurch et al., 2002). However, co-ordinated DNA release from the oral

Page 159: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

143

biofilm dweller S. gordonii and competence induction by S. mutans has been shown to allow the

transmission of usable genetic material from one species to the other (Kreth et al., 2005). Could

the eDNA released into biofilm by CSP-induced lysis serve both to physically strengthen the

biofilm and provide a reservoir of fitness-enhancing DNA under stress? Extensive horizontal

gene transfer has likely occurred between streptococcal species throughout evolution

(Cvitkovitch, 2001). However, the likelihood of productive recombination decreases with

evolutionary distance due to chromosomal divergence. Therefore, the exchange of DNA

between different strains within the same species offers the greatest opportunity for acquisition

of functional genes. We have shown that the induction of CipB-mediated cell death in a culture

is somehow tied to induction of competence, and suggested that the presence of cellular debris

may serve as an additional signal to induce CSP-independent DNA uptake. In the biofilm,

cellular debris released during CSP-induced lysis may trigger CSP-independent competence in

the surviving population, to allow for the uptake of fitness-enhancing genes under stress.

In summary, we have provided a detailed examination of the CSP-induced cell death

pathway in S. mutans, and have attempted to show a physiological role for this pathway in the

stress response, genetic competence and biofilm formation. Our data provides a mechanistic

link between phenotypes previously ascribed to CSP-ComDE signalling.

5.3 Future Directions

While significant progress has been made towards understanding the CSP-induced signalling

cascade, several important questions remain unanswered. Although the CipB-mediated cell

Page 160: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

144

death pathway appears to be able to induce genetic competence in a CSP-independent

manner, it is not clear how death in a population is able to trigger DNA uptake by the surviving

population. Moreover, although we suggest that CSP responsiveness vs. unresponsiveness in

a population of S. mutans is due to bi-stability, further investigation into these dual responses is

warranted. Questions also remain as to the role of CSP-induced genes not required for

competence, and surrounding the shut-off of the CSP response in S. mutans. Finally, although

our results further corroborate past evidence that the LiaFSR TCS may also respond to CSP,

definitive proof of this interaction remains elusive.

5.4 Significance

The ability of S. mutans to form biofilms and tolerate the fluctuating environmental conditions

within those environments is vital to its virulence. We have made significant progress towards

understanding the CSP-induced signalling pathway, which controls its ability to form biofilms

and is central to its stress response. S. mutans itself is of importance as one of the primary

causative agent of the most prevalent human infections, dental caries. However, it is also a

member of the medically important genus Streptococcus and a biofilm-forming organism. As

such, understanding the pathways involved in its stress response and biofilm formation may

provide clues to help combat infections beyond the oral cavity.

Page 161: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

145

5.5 References

Bayles, K.W. (2007) The biological role of death and lysis in biofilm development. Nat Rev Microbiol 5: 721-726.

Cvitkovitch, D.G. (2001) Genetic competence and transformation in oral streptococci. Crit Rev Oral Biol Med 12: 217-243.

Dagkessamanskaia, A., Moscoso, M., Henard, V., Guiral, S., Overweg, K., Reuter, M., Martin, B., Wells, J., and Claverys, J.P. (2004) Interconnection of competence, stress and CiaR regulons in Streptococcus pneumoniae: competence triggers stationary phase autolysis of ciaR mutant cells. Mol Microbiol 51: 1071-1086.

Diep, D.B., Skaugen, M., Salehian, Z., Holo, H., and Nes, I.F. (2007) Common mechanisms of target cell recognition and immunity for class II bacteriocins. Proc Natl Acad Sci 104: 2384-2389.

Guiral, S., Mitchell, T.J., Martin, B., and Claverys, J.P. (2005) Competence-programmed predation of noncompetent cells in the human pathogen Streptococcus pneumoniae: genetic requirements. Proc Natl Acad Sci 102: 8710-8715.

Kreth, J., Merritt, J., Shi, W., and Qi, F. (2005) Co-ordinated bacteriocin production and competence development: a possible mechanism for taking up DNA from neighbouring species. Mol Microbiol 57: 392-404.

Li, Y.H., Lau, P.C., Tang, N., Svensater, G., Ellen, R.P., and Cvitkovitch, D.G. (2002a) Novel two-component regulatory system involved in biofilm formation and acid resistance in Streptococcus mutans. J Bacteriol 184: 6333-6342.

Qi, F., Kreth, J., Levesque, C.M., Kay, O., Mair, R.W., Shi, W., Cvitkovitch, D.G., and Goodman, S.D. (2005) Peptide pheromone induced cell death of Streptococcus mutans. FEMS Microbiol Lett 251: 321-326.

Whitchurch, C.B., Tolker-Nielsen, T., Ragas, P.C., and Mattick, J.S. (2002) Extracellular DNA required for bacterial biofilm formation. Science 295: 1487.

Page 162: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

146

Appendix A: Supplementary Information

Page 163: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

147

SI Table S1. Genes showing a minimum ± 2-fold difference in expression when S. mutans

UA159 cells were exposed to 2 μM sCSP

Gene ID Putative or assigned function Fold Amino acid biosynthesis

SMU.1073 putative formyl-tetrahydrofolate synthetase -3.3

SMU.1265 putative phosphoribosyl formimino-5-aminoimidazole carboxamide ribonucleotide isomerase-2.0

SMU.1266 putative glutamine amidotransferase HisH -2.1

SMU.1268 putative imidazoleglycerol-phosphate dehydratase -2.2

SMU.1269 putative phosphoserine phosphatase -1.9

SMU.1270 putative histidinol dehydrogenase -2.3

SMU.1271 putative ATP phosphoribosyltransferase -2.1

SMU.1273 putative histidinol-phosphate aminotransferase -2.4

SMU.1877 putative PTS system, mannose-specific component IIAB -2.9

SMU.531 putative chorismate mutase 4.3

SMU.532 putative anthranilate synthase, alpha subunit 3.4

SMU.534 putative phosphoribosyl anthranilate transferase 3.4

SMU.535 putative indoleglycerol phosphate synthase 3.7

SMU.536 putative phosphoribosyl anthranilate isomerase 2.5

SMU.537 putative tryptophan synthase, beta subunit 2.5

SMU.538 putative tryptophan synthase, alpha subunit 2.8

Biosynthesis of cofactors, prosthetic groups, and carriers

SMU.1996 putative isopentenyl monophosphate kinase 3.3

SMU.353 conserved hypothetical protein 5.3

SMU.838 glutathione reductase 6.5

SMU.954 putative pyridoxal kinase 2.3

Cell envelope

SMU.109 conserved hypothetical protein; possible permease (efflux protein) 5.7

SMU.1196c conserved hypothetical protein 2.3

SMU.1677 putative UDP-N-acetylmuramoylananine-D-glutamate-2,6- diaminopimelate ligase; UDP-MurNac-tripeptide synthetase2.4

SMU.196c putative transfer protein 3.0

SMU.2075c conserved hypothetical protein 2.9

SMU.2081 hypothetical protein 7.5

SMU.539c signal peptidase type IV 22.4

SMU.610 cell surface antigen SpaP -2.5

SMU.627 conserved hypothetical protein 4.5

SMU.63c conserved hypothetical protein 5.3

SMU.67 putative acyltransferase 4.0

SMU.883 dextran glucosidase DexB -5.6

Cellular processes

SMU.1001 putative DNA processing Smf protein 19.0

SMU.1279c putative cell division protein (cell shape determining protein) 2.0

SMU.1343c putative polyketide synthase -3.4

SMU.1346 putative thioesterase BacT -2.5

SMU.150 hypothetical protein 11.2

SMU.1862 hypothetical protein 2.4

SMU.1897 putative ABC transporter, ATP-binding protein 9.2

SMU.1898 putative ABC transporter, ATP-binding and permease protein 4.2

SMU.1900 conserved hypothetical protein 5.9

SMU.1905c putative bacteriocin secretion protein 10.1

SMU.1906c hypothetical protein 11.4

Page 164: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

148

SMU.1914c hypothetical protein 20.4

SMU.1916 putative histidine kinase of the competence regulon, ComD 10.5

SMU.1917 putative response regulator of the competence regulon, ComE 11.3

SMU.1983 putative competence protein ComYD 26.6

SMU.1984 putative competence protein ComYC 26.2

SMU.1985 putative ABC transporter ComYB; probably part of the DNA transport machinery 23.9

SMU.1987 putative ABC transporter, ATP-binding protein ComYA; late competence gene 17.5

SMU.1997 16S ribosomal RNA 14.3

SMU.2084c conserved hypothetical protein 5.6

SMU.400 putative secreted esterase 2.8

SMU.423 hypothetical protein 14.6

SMU.426 copper-transporting ATPase; P-type ATPase 7.3

SMU.499 putative late competence protein 18.0

SMU.54 putative amino acid recemase -3.5

SMU.625 putative competence protein 14.1

SMU.626 putative competence protein 25.6

SMU.629 putative manganese-type superoxide dismutase, Fe/Mn-SOD -2.2

SMU.632 putative transcriptional regulator 2.4

SMU.644 putative competence protein/transcription factor 27.2

SMU.655 putative MutE 3.9

SMU.753 conserved hypothetical protein 4.0

Central intermediary metabolism

SMU.636 putative N-acetylglucosamine-6-phosphate isomerase 3.1

DNA metabolism

SMU.1002 putative DNA topoisomerase I 8.2

SMU.1034c putative integrase/recombinase; XerC-like 2.2

SMU.1055 putative DNA repair protein RadC 14.1

SMU.1967 putative single-stranded DNA-binding protein 16.4

SMU.2085 recombination protein RecA 6.7

SMU.2086 putative competence and damage inducible protein CinA 10.2

SMU.327 putative DNA repair protein 3.2

SMU.505 putative adenine-specific DNA methylase 11.8

SMU.506 putative type II restriction endonuclease 9.0

SMU.64 Holliday junction DNA helicase RuvB 8.0

Energy metabolism

SMU.1004 glucosyltransferase-I 6.0

SMU.127 putative acetoin dehydrogenase (TPP-dependent), E1 component alpha subunit -2.2

SMU.128 putative acetoin dehydrogenase (TPP-dependent), E1 component beta subunit -2.5

SMU.129 putative dihydrolipoamide acetyltransferase -2.6

SMU.130 putative dihydrolipoamide dehydrogenase -2.7

SMU.1327c conserved hypothetical protein; possible 4Fe-4S ferredoxin 3.0

SMU.1424 putative dihydrolipoamide dehydrogenase 2.9

SMU.148 putative alcohol-acetaldehyde dehydrogenase 3.4

SMU.1978 putative acetate kinase 9.0

SMU.2037 putative trehalose-6-phosphate hydrolase TreA 6.7

SMU.352 putative ribulose-phosphate-3-epimerase 6.0

SMU.402 pyruvate formate-lyase -2.2

SMU.772 glucan-binding protein D with lipase activity; BglB-like protein 10.0

SMU.79 fructan hydrolase; exo-beta-D-fructosidase; FruB -4.5

Page 165: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

149

SMU.877 alpha-galactosidase -3.3

SMU.881 sucrose phosphorylase, GtfA -5.8

SMU.886 galactokinase, GalK -2.6

SMU.887 galactose-1-P-uridyl transferase, GalT -2.6

SMU.888 UDP-galactose 4-epimerase, GalE -2.7

Fatty acid and phospholipid metabolism

SMU.1344c putative malonyl-CoA acyl-carrier-protein transacylase -2.9

SMU.1345c putative peptide synthetase similar to MycA -2.7

Hypothetical proteins

SMU.108 hypothetical protein 4.9

SMU.1197 tRNA-Arg 2.6

SMU.1267c hypothetical protein -2.1

SMU.1438c putative Zn-dependent protease 2.1

SMU.151 hypothetical protein 12.4

SMU.1651 putative arsenate reductase 2.4

SMU.167 hypothetical protein 10.0

SMU.1904c hypothetical protein 12.2

SMU.1915 competence stimulating peptide, precursor 3.8

SMU.1956c hypothetical protein -5.0

SMU.1979c conserved hypothetical protein 22.0

SMU.1980c conserved hypothetical protein 24.6

SMU.1982c conserved hypothetical protein 22.3

SMU.199c hypothetical protein 2.8

SMU.202c hypothetical protein 2.8

SMU.205c hypothetical protein 3.5

SMU.2076c hypothetical protein 10.1

SMU.2077c conserved hypothetical protein 2.7

SMU.2078c conserved hypothetical protein 2.4

SMU.2079c conserved hypothetical protein 2.8

SMU.2080 conserved hypothetical protein 3.3

SMU.209c hypothetical protein 3.8

SMU.212c hypothetical protein 3.8

SMU.217c hypothetical protein 2.5

SMU.326 conserved hypothetical protein 3.3

SMU.470 conserved hypothetical protein 2.4

SMU.503c hypothetical protein -2.2

SMU.53 conserved hypothetical protein -3.5

SMU.56 conserved hypothetical protein -3.1

SMU.649 conserved hypothetical protein -2.0

SMU.840c hypothetical protein 2.1

SMU.959c hypothetical protein -2.2

SMU.1047c hypothetical protein 6.2

SMU.1056 hypothetical protein 5.7

SMU.1069c hypothetical protein 3.7

SMU.1147c hypothetical protein 5.0

SMU.1250c hypothetical protein 1.9

SMU.152 hypothetical protein 16.1

SMU.153 hypothetical protein 12.2

SMU.166 hypothetical protein 9.1

SMU.1902c hypothetical protein 9.8

Page 166: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

150

SMU.1903c hypothetical protein 16.0

SMU.1907 hypothetical protein 8.9

SMU.1908c hypothetical protein 18.3

SMU.1909c hypothetical protein 19.6

SMU.1910c hypothetical protein 18.3

SMU.1912c hypothetical protein 22.2

SMU.1913c putative immunity protein, BLpL-like 15.2

SMU.1976c hypothetical protein 3.6

SMU.200c hypothetical protein 2.8

SMU.204c hypothetical protein 3.0

SMU.2083c hypothetical protein 4.0

SMU.210c hypothetical protein 3.2

SMU.215c hypothetical protein 3.2

SMU.216c hypothetical protein 2.5

SMU.378 hypothetical protein 2.5

SMU.41 hypothetical protein 2.7

SMU.49 hypothetical protein -2.9

SMU.55 hypothetical protein -3.6

SMU.58 hypothetical protein -3.1

SMU.637c hypothetical protein 2.9

SMU.68 hypothetical protein 2.9

SMU.735 hypothetical protein 2.1

SMU.771c hypothetical protein 8.0

SMU.925 hypothetical protein 18.2

Mobile and extrachromosomal element functions

SMU.149 putative transposase 4.1

SMU.195c hypothetical protein; similar to ORF 5 of bacteriophage SPP1 3.3

SMU.198c putative conjugative transposon protein 2.6

SMU.2027 putative transcriptional regulator 4.7

Protein fate

SMU.131 putative lipoate-protein ligase -2.4

SMU.645 putative oligopeptidase 10.4

Protein synthesis

SMU.1044c putative pseudouridylate synthase 2.5

SMU.1512 putative phenylalanyl-tRNA synthetase (alpha subunit) -2.3

SMU.154 30S ribosomal protein S15 2.2

SMU.1886 putative seryl-tRNA synthetase -2.1

SMU.2000 50S ribosomal protein L17 -2.3

SMU.2002 30S ribosomal protein S11 -2.1

SMU.2012 30S ribosomal protein S8 -2.1

SMU.2014 30S ribosomal protein S14 -2.0

SMU.2015 50S ribosomal protein L5 -2.5

SMU.2016 50S ribosomal protein L24 -2.1

SMU.2017 50S ribosomal protein L14 -2.1

SMU.2022 50S ribosomal protein L22 -2.3

SMU.500 putative ribosome-associated protein 2.7

SMU.558 isoleucine-tRNA synthetase -2.1

Purines, pyrimidines, nucleosides, and nucleotides

SMU.30 putative phosphoribosylformylglycinamidine synthase, (FGAM synthase) -2.6

SMU.325 putative dUTPase 4.1

SMU.356 purine operon repressor 3.8

Page 167: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

151

SMU.48 putative phosphoribosylamine-glycine ligase; phosphoribosyl glycinamide synthetase (GARS)-2.4

SMU.50 putative phosphoribosylaminoimidazole carboxylase, catalytic subunit -2.6

SMU.51 putative phosphoribosylaminoimidazole carboxylase, ATPase subunit -3.4

SMU.668c ribonucleotide reductase, large subunit -2.0

Regulatory functions

SMU.1048 conserved hypothetical protein 2.4

SMU.1145c putative histidine kinase; homolog of RumK and ScnK (HK3) 2.4

SMU.1193 putative transcriptional regulator 2.5

SMU.1409c putative transcriptional regulator 2.4

SMU.1509 putative transcriptional regulator 2.2

SMU.168 putative transcriptional regulator 8.7

SMU.1964c putative response regulator (RR9) 4.8

SMU.1977c putative transcriptional regulator 3.8

SMU.207c putative transposon protein 4.4

SMU.424 negative transcriptional regulator, CopY 8.4

SMU.507 putative transcriptional regulator (DeoR family) 6.0

SMU.61 putative transcriptional regulator 3.1

SMU.65 putative protein tyrosine-phosphatase 6.6

SMU.80 transcriptional regulator; repressor (HrcA) of class I heat shock genes 3.0

SMU.927 putative response regulator (RR4) 4.8

SMU.928 putative histidine kinase (HK4) 5.3

Signal transduction

SMU.1957 putative PTS system, mannose-specific IID component -5.1

SMU.1958c putative PTS system, mannose-specific IIC component -4.4

SMU.1960c putative PTS system, mannose-specific IIB component -3.4

SMU.1961c putative PTS system, sugar-specific enzyme IIA component -3.2

SMU.1965c putative histidine kinase (HK9) 5.5

Transcription

SMU.2001 DNA-directed RNA polymerase, alpha subunit -2.2

Transport and binding proteins

SMU.1006 putative ABC transporter, ATP-binding protein 2.9

SMU.1067c putative ABC transporter, permease protein 2.7

SMU.1068c putative ABC transporter, ATP-binding protein 3.0

SMU.1148 putative transporter, ATP-binding protein; bacteriocin immunity protein 2.4

SMU.1185 PTS system, mannitol-specific enzyme IIBC component 2.4

SMU.1194 putative ABC transporter, ATP-binding protein 2.2

SMU.1195 conserved hypothetical protein; possible permease 2.0

SMU.1848 hypothetical protein 2.1

SMU.1878 putative PTS system, mannose-specific component IIC -3.0

SMU.1879 putative PTS system, mannose-specific component IID -3.9

SMU.1899 putative ABC transporter, ATP-binding and permease protein (fragment) 5.2

SMU.1963c putative sugar-binding periplasmic protein 4.0

SMU.1966c putative periplasmic sugar-binding protein 7.1

SMU.2038 putative PTS system, trehalose-specific IIABC component 5.3

SMU.242c putative amino acid ABC transporter, permease protein, glutamine transport system -2.0

SMU.427 putative copper chaperone 6.5

SMU.862 conserved hypothetical protein; putative permease 2.1

SMU.863 putative ABC transporter, ATP-binding protein 1.9

SMU.864 putative ABC transporter, permease component 2.0

SMU.872 putative PTS system, fructose-specific enzyme IIABC component -2.1

Page 168: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

152

SMU.878 multiple sugar-binding ABC transporter, sugar-binding protein precursor MsmE -4.4

SMU.879 multiple sugar-binding ABC transporter, permease protein MsmF -5.4

SMU.880 multiple sugar-binding ABC transporter, permease protein MsmG -5.6

SMU.882 multiple sugar-binding ABC transporter, ATP-binding protein, MsmK -5.9

Unclassified

SMU.1111c conserved hypothetical protein 2.0

SMU.1342 putative bacitracin synthetase 1; BacA -2.9

SMU.1372c hypothetical protein 2.2

SMU.193c conserved hypothetical protein 2.9

SMU.1981c conserved hypothetical protein 25.9

SMU.1988c putative DNA binding protein 2.0

SMU.2057c putative cadmium-transporting ATPase; P-type ATPase 4.0

SMU.214c hypothetical protein 2.2

SMU.219 hypothetical protein 2.0

SMU.52 conserved hypothetical protein -3.6

SMU.73 conserved hypothetical protein -2.7

SMU.758c conserved hypothetical protein 3.3

SMU.769 conserved hypothetical protein 9.5

SMU.78 fructan hydrolase; exo-beta-D-fructosidase; fructanase, FruA -5.5

SMU.836 hypothetical protein 18.3

Unknown function

SMU.1003 putative glucose-inhibited division protein 7.1

SMU.1046c putative GTP pyrophosphokinase 2.4

SMU.1053 conserved hypothetical protein 6.4

SMU.1054 putative glutamine amidotransferase 6.0

SMU.1070c conserved hypothetical protein 3.4

SMU.1322 putative acetoin dehydrogenase 2.0

SMU.1323 conserved hypothetical protein; possible hydrolase 2.0

SMU.1340 putative surfactin synthetase -2.0

SMU.1341c putative gramicidin S synthetase -2.4

SMU.1400c conserved hypothetical protein 4.7

SMU.1975c conserved hypothetical protein; possible membrane protein 2.5

SMU.208c putative transposon protein; possible DNA segregation ATPase 3.3

SMU.328 putative carbonic anhydrase 2.2

SMU.354 conserved hypothetical protein 6.2

SMU.355 putative CMP-binding factor 6.8

SMU.399 conserved hypothetical protein 2.5

SMU.401c conserved hypothetical protein 2.5

SMU.498 putative late competence protein 22.4

SMU.508 conserved hypothetical protein 7.2

SMU.641 putative oxidoreductase 2.1

SMU.646 putative phosphatase 9.5

SMU.647 putative methyltransferase 2.7

SMU.66 conserved hypothetical protein 5.0

SMU.72 conserved hypothetical protein -2.4

SMU.807 putative membrane protein 2.1

SMU.837 putative reductase 14.1

SMU.890 conserved hypothetical protein 2.0

SMU.926 conserved hypothetical protein; possible GTP-pyrophosphokinase 5.4

Page 169: Study of the physiological and molecular mechanisms underlying peptide-induced … · 2013-12-13 · Study of the physiological and molecular mechanisms underlying peptide-induced

153

SI Table S2. S. mutans genes showing a minimum ± 2-fold difference in expression when

S. mutans ∆comX cells were exposed to 2 μM sCSP

Gene ID Putative or assigned function FoldSMU.925 hypothetical protein 2.7

SMU.150 hypothetical protein 5.3

SMU.151 hypothetical protein 4.9

SMU.152 hypothetical protein 5.4

SMU.153 hypothetical protein 5.2

SMU.1902c hypothetical protein 2.5

SMU.1903c hypothetical protein 5.0

SMU.1904c hypothetical protein 5.1

SMU.1905c putative bacteriocin secretion protein 5.1

SMU.1906c hypothetical protein 4.8

SMU.1908c hypothetical protein 4.9

SMU.1909c hypothetical protein 5.7

SMU.1910c hypothetical protein 5.6

SMU.1912c hypothetical protein 4.9

SMU.1913c putative immunity protein, BLpL-like 5.2

SMU.1914c hypothetical protein 4.5

SMU.2037 putative trehalose-6-phosphate hydrolase TreA 2.1

SMU.2038 putative PTS system, trehalose-specific IIABC component 2.3

SMU.41 hypothetical protein -3.3

SMU.423 hypothetical protein 5.8

SMU.424 negative transcriptional regulator, CopY 2.0

SMU.426 copper-transporting ATPase; P-type ATPase 2.5

SMU.427 putative copper chaperone 2.1

SMU.63c conserved hypothetical protein 2.7

SMU.64 Holliday junction DNA helicase RuvB 2.7

SMU.65 putative protein tyrosine-phosphatase 2.8

SMU.66 conserved hypothetical protein 2.1

SMU.78 fructan hydrolase; exo-beta-D-fructosidase; fructanase, FruA -2.0

SMU.799c conserved hypothetical protein 2.6

SMU.877 alpha-galactosidase -2.2

SMU.878 multiple sugar-binding ABC transporter, sugar-binding protein precursor MsmE -2.7

SMU.879 multiple sugar-binding ABC transporter, permease protein MsmF -2.6

SMU.880 multiple sugar-binding ABC transporter, permease protein MsmG -2.5

SMU.881 sucrose phosphorylase, GtfA -2.8

SMU.882 multiple sugar-binding ABC transporter, ATP-binding protein, MsmK -2.4

SMU.883 dextran glucosidase DexB -2.8

SMU.887 galactose-1-P-uridyl transferase, GalT -2.2