30.12.2012 Views

The Staphylococcus aureus secretome - TI Pharma

The Staphylococcus aureus secretome - TI Pharma

The Staphylococcus aureus secretome - TI Pharma

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>The</strong> <strong>Staphylococcus</strong> <strong>aureus</strong> <strong>secretome</strong>


ISBN<br />

978-90-902-5004-5 (printed version)<br />

978-90-367-4182-8 (digital version)<br />

Printing<br />

Printed by PrintPartners Ipskamp, Enschede, the Netherlands<br />

Copyright<br />

All rights reserved. No part of this publication may be reproduced or transmitted in any form<br />

or by any means without the permission of the author and the publisher holding the copyright<br />

of the published articles.<br />

Cover<br />

Cover designed by Feiko Beckers (www.feikobeckers.com). <strong>The</strong> cover illustrates the typical<br />

cluster of <strong>Staphylococcus</strong> <strong>aureus</strong> cells represented by ‘LEGO-tires’. On the back: a signpost<br />

with the different paths that a newly synthesized protein can take depending on the nature of<br />

its signal peptide.<br />

2


RIJKSUNIVERSITEIT GRONINGEN<br />

<strong>The</strong> <strong>Staphylococcus</strong> <strong>aureus</strong> <strong>secretome</strong><br />

Proefschrift<br />

ter verkrijging van het doctoraat in de<br />

Medische Wetenschappen<br />

aan de Rijksuniversiteit Groningen<br />

op gezag van de<br />

Rector Magnificus, dr. F. Zwarts,<br />

in het openbaar te verdedigen op<br />

woensdag 27 januari 2010<br />

om 13:15 uur<br />

door<br />

Mark Jan Jacobus Bernhard Sibbald<br />

geboren op 18 mei 1975<br />

te Bolsward<br />

3


Promotor : Prof. dr. J.M. van Dijl<br />

Co-promotor : Dr. J-.Y.F. Dubois<br />

Beoordelingscommissie : Prof. dr. Tarek Msadek<br />

: Prof. dr. Wim Quax<br />

: Prof. dr. Arnold Driessen<br />

4


Paranimfen: Thijs R.H.M. Kouwen<br />

Monika A. Chlebowicz<br />

Voor Regina<br />

Voor Pap<br />

“I’ll always remember the chill of November”<br />

“Carpe diem - seize the day”<br />

“Look around, hear the sounds<br />

Cherish your life while you’re still around”<br />

“Seize the day and don't you cry, now it's time to say goodbye<br />

Even though I'll be gone, I will live on”<br />

-Dream <strong>The</strong>ater – “A Change Of Seasons”-<br />

“Thank you for the inspiration, thank you for the smiles<br />

All the unconditional love that carried me for miles<br />

It carried me for miles<br />

But most of all: thank you for my life”<br />

-Dream <strong>The</strong>ater – “Best Of Times”-<br />

5


<strong>The</strong> work described in this thesis was performed in the laboratory of Molecular Bacteriology,<br />

Department of Medical Microbiology of the University Medical Center Groningen and University of<br />

Groningen, Groningen, the Netherlands.<br />

Printing of this thesis was financially supported by the Graduate School for Drug Exploration<br />

(GUIDE), the Juriaanse Stichting, Nederlandse Vereniging voor Medische Microbiologie<br />

(NVvM/NVMM), DSM Nutritional Products Ltd, and Biomade. <strong>The</strong>ir support is highly appreciated.<br />

6


Table of contents<br />

Chapter 1. General Introduction<br />

Chapter 2. Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

(Sibbald et al., MMBR, 2006)<br />

Chapter 3. Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation (Ziebandt et al., in<br />

revision)<br />

Chapter 4. Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants (Sibbald et al., to be<br />

submitted)<br />

Chapter 5. Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> (Sibbald et al., in revision)<br />

Chapter 6. Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

(Sibbald et al., submitted)<br />

Chapter 7. <strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to the<br />

bacteriocin sublancin 168 (Kouwen et al., Antimicrob Agents Chemother, 2009.)<br />

Chapter 8. <strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media and<br />

under different nutrient starvation conditions (Voigt et al., Proteomics, 2005)<br />

Chapter 9. General summary and discussion<br />

Chapter 10. Reference list<br />

Chapter 11. Nederlandse samenvatting<br />

Appendices:<br />

I. Dankwoord<br />

II. List of publications<br />

III. Supplemental tables<br />

7


"<strong>The</strong> greatest education in the world is watching the masters at work"<br />

- Michael J. Jackson-<br />

8


Chapter 1<br />

Introduction and scope of this thesis<br />

9


Chapter 1<br />

Introduction<br />

Bacteria are the oldest living organisms that inhabit this planet. During evolution some of<br />

these prokaryotic cells have been working together to evolve into multicellular organisms.<br />

This has resulted in the multidiversity of organisms that have existed and exist to this day.<br />

Like the vast majority of organisms in the animal kingdom, human beings carry a huge<br />

variety of microorganisms, including many bacterial species but also archaea and yeasts.<br />

Collectively these organisms are known as the human microbiota. Most bacteria amongst the<br />

human microbiota are commensal, but some of them are in fact opportunistic pathogens that<br />

can cause a wide range of diseases. On the other hand, some bacteria seem to help the human<br />

host by competing with opportunistic and primary pathogens. <strong>The</strong>se beneficial species thus<br />

prevent harmful bacteria to colonize and spread throughout the host. However, when such<br />

pathogens break through the human and bacterial defenses, they can cause a wide range of<br />

diseases which, in some cases, can be life-threatening. To conquer certain niches in the<br />

human host, pathogenic bacteria have to overcome many stressful conditions, as imposed by<br />

the human innate and adaptive immune systems. To accomplish this, bacteria preduce an<br />

arsenal of virulence factors. Although one proteinaceous virulence factor can already be<br />

sufficient to cause particular symptoms of disease, the synergistic actions of many other<br />

proteins are needed for bacterial survival in the host. Both groups of proteins are part of the<br />

arsenal of virulence factors that contribute to the disease-causing ability of pathogenic<br />

bacteria. <strong>The</strong> proteins that are actively involved in the processes of colonization, invasion,<br />

spreading, immune evasion and the triggering of excessive immune responses are all<br />

synthesized in the cytoplasm of the bacteria, and then transported across the bacterial<br />

membrane to an extracytoplasmic location, such as the bacterial cell wall or the host’s millieu.<br />

<strong>The</strong>se transport systems are complex systems that are embedded in the bacterial membrane<br />

and consist of a translocation motor and a channel through which the proteins are<br />

translocated. Like all living organisms, bacteria contain several transport systems that are<br />

used to transport proteins across the membrane. <strong>The</strong> best-studied transport system is the<br />

general secretion (Sec) pathway. Furthermore, several other “special purpose” transport<br />

pathways are under investigation to understand their contribution to the transport of virulence<br />

factors, such as the Twin-arginine translocase (Tat) and the ESX-1 or ESAT-6 secretion<br />

system (Ess) pathway.<br />

All proteins that are actively translocated via the Sec or Tat pathways are synthesized with Nterminal<br />

signal peptides that lead them to the respective transport system. <strong>The</strong> signal that<br />

directs proteins to the Ess pathway remains to be defined. During or shortly after passage of<br />

precursor proteins through the Sec or Tat translocation channels, their signal peptides are<br />

removed by a so-called signal peptidase. While the Tat channel has the potential to transport<br />

fully folded proteins acrosse the membrane, the Sec channel can only handle proteins in an<br />

unfolded state. <strong>The</strong>refore, proteins have to fold into their active and protease-resistant<br />

conformation after translocation through the Sec channel. This can occur spontaneously, or<br />

with the aid of chaperones and folding catalysts. Finally, the protein is either retained in an<br />

extracytoplasmic compartment of the cell or released into the extracellular environment. In<br />

the case of Gram-positive bacteria, three extracytoplasmic compartments can be<br />

distinguished: the membrane, the membrane-cell wall interface, and the cell wall. <strong>The</strong><br />

proteins that are exposed at the surface of bacterial cells are very important for the bacterial<br />

10


Introduction and scope of this thesis<br />

adherence to host tissues and evasion of the host defense systems. In addition, the virulence<br />

factors that are released into the host milieu may damage host cells, degrade<br />

biomacromolecules of the host, or cause strong inflammatory responses, thereby contributing<br />

to the symptoms of disease caused by a particular bacterium (For reviews see Tjalsma et al.,<br />

2000; Tjalsma et al., 2004; Sibbald et al., 2006; van Dijl et al., 2007; Sibbald and van Dijl,<br />

2009).<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> is a Gram-positive bacterium that is part of the human microbiota,<br />

residing mostly in the mucosal environment in the nose. In fact, ~20% of the human<br />

population is a persistant carrier of S. <strong>aureus</strong>, while 60% is an intermittent carrier.<br />

Unfortunately, S. <strong>aureus</strong> can transform from an apparently harmles commensal into a<br />

dangerous pathogen. Once the organism crosses the defense systems of the human host, it can<br />

spread to almost every organ and tissue, causing a wide range of diseases. <strong>The</strong>se can vary<br />

from superficial lesions, styes and furunculosis, to more serious infections such as<br />

pneumonia, urinary tract infections, endocarditis, and in rare cases even meningitis (Cheng et<br />

al., 2009; Dubrac et al., 2008; Fedtke et al., 2004; García-Lara et al., 2005; Novick, 2003;<br />

van Belkum A., 2006; Wardenburg et al., 2007). Moreover, S. <strong>aureus</strong> has an amazing ability<br />

to develop resistance against several antibiotics, which became evident already shortly after<br />

the introduction of penicillin for clinical applications. In the mean time, S. <strong>aureus</strong> strains<br />

exhibiting resistance to most antibiotics are known, the methicillin resistant S. <strong>aureus</strong><br />

(MRSA) being most notorious (annual Report EARSS 2007; http://www.rivm.nl/earss/<br />

results/Monitoring_reports). Up till now, most MRSA infections were nosocomial (i.e.<br />

hospital-acquired). However, recent reports indicate an increase in dangerous communityacquired<br />

MRSA infections (Centers for Disease Control and Prevention, 2003; Grundmann et<br />

al., 2002; Vandenesch et al., 2003). Vancomycin has been for long time a last resort antibiotic<br />

against MRSA, but in 1996 the first vancomycin intermediate resistant strain (VISA) was<br />

reported (Hiramatsu et al., 1997). Since then, several other cases of other VISA strains and<br />

even strains with complete resistance (VRSA) against vancomycin have been isolated (Cui et<br />

al., 2003; Weigel et al., 2003). Because of the anticipated rise of multiple antibiotic resistant<br />

S. <strong>aureus</strong> strains, innovative strategies for prevention and intervention of S. <strong>aureus</strong> infections<br />

are urgently needed. Recent studies are therefore focusing on the development of novel<br />

antibiotics, anti-staphylococcal vaccines and therapeutic protective antibodies (Arrecubieta et<br />

al., 2008; Glowalla et al., 2009; Middleton, 2008; Nanra et al., 2009; Otto, 2008; Schaffer<br />

and Lee, 2009; Zweers et al., 2009).<br />

11


Chapter 1<br />

Scope of this thesis<br />

Due to its large arsenal of virulence factors and the ability to adapt rapidly to externally<br />

imposed stresses and insults, S. <strong>aureus</strong> has become one of the most “successful” human<br />

pathogens. <strong>The</strong>se virulence factors are synthesized on cytoplasmic ribosomes and transported<br />

across the membrane to extracytoplasmic locations as introduced above. Especially for the<br />

Gram-negative bacterium Escherichia coli, the protein translocation machinery has been<br />

described in great detail. For studies on Gram-positive bacterial secretion systems, Bacillus<br />

subtilis has served as the main model organism. However, B. subtilis is non-pathogenic and<br />

relatively little is known about the precise roles of protein translocation systems in related<br />

pathogens, such as Bacillus anthracis, Listeria monocytogenes, Mycobacterium tuberculosis,<br />

Streptococcus pneumoniae and, last but not least, S. <strong>aureus</strong>. <strong>The</strong> present thesis studies were<br />

aimed at defining the <strong>secretome</strong> of S. <strong>aureus</strong>. <strong>The</strong> <strong>secretome</strong> includes all proteins that are<br />

involved in protein export processes from the cytoplasm to extracytoplasmic locations, as<br />

well as the proteins that are translocated across the bacterial membrane. Where appropriate,<br />

studies involved comparisons with Gram-positive bacteria that are related to S. <strong>aureus</strong>, like<br />

<strong>Staphylococcus</strong> epidermidis, Bacillus subtilis and Bacillus licheniformis.<br />

In the studies described in Chapter 2, the genomes of several S. <strong>aureus</strong> strains and one S.<br />

epidermidis strain were scanned with bioinformatic tools for the presence of genes encoding<br />

proteins that are involved in the export of extracytoplasmic proteins. Furthermore, proteins<br />

that carry N-terminal signal peptides were identified through bioinformatics. <strong>The</strong> obtained<br />

results were compared with each other to define the core and variant S. <strong>aureus</strong> exoproteomes.<br />

Chapter 3 reports on a first comprehensive survey of the composition and variability of the S.<br />

<strong>aureus</strong> exoproteome following a proteogenomics approach. Dissection of the exoproteomes<br />

of 25 clinical isolates revealed that only seven out of 63 identified secreted proteins were<br />

produced by all isolates, revealing a high exoproteome heterogeneity. <strong>The</strong> observed variations<br />

were caused by both genome plasticity and an unprecedented variation in gene expression.<br />

<strong>The</strong> data have important implications for future studies on staphylococcal virulence and the<br />

development of protective vaccines against this pathogen.<br />

Chapter 4 describes the construction and analysis of a collection of isogenic S. <strong>aureus</strong><br />

secretion mutants. <strong>The</strong> exproteomes of the mutant strains were analyzed by SDS-PAGE,<br />

proteomics, western blotting, zymogram analysis, spreading assays, hemolysin activity<br />

assays, electron microscopy, and a Caenorhabditis elegans killing assay. While no<br />

phenotypes were detectable for some of the mutants, strains with mutations in dsbA, lgt or<br />

secG genes did show clear secretion defects. Notably, in certain strains second site mutations<br />

were observed that led to the loss of RNAIII. While this seems to be a consequence of the<br />

natural adaptive capabilities of S. <strong>aureus</strong>, it is also an important warning for future studies on<br />

protein secretion in this organism and it underscores the need for genetically stable model<br />

strains.<br />

In Chapter 5 the roles of the non-essential Sec channel components SecG and SecY2 in the<br />

biogenesis of the extracellular proteome of S. <strong>aureus</strong> were investigated. <strong>The</strong> results show that<br />

SecG is of major importance for protein secretion by S. <strong>aureus</strong>. No secretion defects were<br />

observed for strains with a secY2 single mutation, but deletion of secY2 significantly<br />

exacerbated the secretion defects of secG mutants. Furthermore, the secG secY2 double<br />

mutant displayed a synthetic growth defect. <strong>The</strong>se findings suggest that SecY2 can interact<br />

with the Sec1 channel of S. <strong>aureus</strong>.<br />

12


Introduction and scope of this thesis<br />

Some of the staphylococcal virulence factors are proteins that are displayed at the cell wall<br />

surface. In S. <strong>aureus</strong> some of these surface proteins are linked to the cell wall by so-called<br />

sortases. In Chapter 6, the exoproteomes of S. <strong>aureus</strong> and S. epidermidis srtA mutants were<br />

investigated and compared to the respective parental strains. Several SrtA substrates were<br />

identified, and their final subcellular localization was found to be altered in the srtA mutants.<br />

Among these identified proteins were the S. <strong>aureus</strong> surface protein G (SasG) and the<br />

accumulation associated protein (Aap) from S. epidermidis. Biofilm formation was affected in<br />

the srtA mutants. Complementation studies were performed with SrtA from S. <strong>aureus</strong> and S.<br />

epidermidis as well as with SrtC from S. epidermidis and YhcS from B. subtilis. Only<br />

complementation with S. <strong>aureus</strong> or S. epidermidis SrtA resulted in restoration of the parental<br />

phenotype. In contrast, SrtC of S. epidermidis only partially seems to restore the parental<br />

phenotype, and YhcS of B. subtilis was not able to complement for the loss of SrtA at all.<br />

In Chapter 7, the susceptibility of bacteria towards the extremely stable and broad-spectrum<br />

lantibiotic sublancin 168 was investigated. S. <strong>aureus</strong> is one of several important pathogens<br />

that is susceptible towards sublancin 168. Growth inhibition and competition assays on plates<br />

and in liquid cultures revealed that the addition of NaCl lowered the sublancin 168<br />

susceptibility of S. <strong>aureus</strong> and B. subtilis. In addition, it was shown that the presence of the<br />

large mechanosensitive channel of conductance MscL is important for the susceptibility of<br />

these bacteria. Taken together, the results demonstrate that MscL is a critical and specific<br />

determinant in bacterial sublancin 168 susceptibility that may either serve as a direct target for<br />

this lantibiotic, or as a gate of entry to the cytoplasm.<br />

Chapter 8 describes the exoproteome of the commercially interesting organism B.<br />

licheniformis. With the annotated genome sequence of B. licheniformis DSM 13, a <strong>secretome</strong><br />

prediction analysis was performed. A total of 296 proteins were predicted to contain an Nterminal<br />

signal peptide directing most of them into the Sec pathway. From analysis of the<br />

extracellular proteomes of B. licheniformis it was concluded that higher amounts of protein<br />

are secreted when cell are grown in a complex medium compared to cell grown in a minimal<br />

medium. In addition, limitation of phosphate, carbon and nitrogen sources resulted in the<br />

secretion of specific proteins that may be involved in counteracting the imposed starvation<br />

conditions.<br />

Finally, in Chapter 9 the results described in this thesis are discussed and ideas for future<br />

research are presented.<br />

13


“Music is a moral law<br />

It gives soul to the universe, wings to the mind, flight to the imagination,<br />

and charm and gaiety to life and to everything”<br />

-Plato-<br />

14


Chapter 2<br />

Mapping the pathways to staphylococcal pathogenesis by<br />

comparative secretomics<br />

M.J.J.B. Sibbald, A.K. Ziebandt, S. Engelmann, M. Hecker, A. de Jong, H.J.M. Harmsen,<br />

G.C. Raangs, I. Stokroos, J.P. Arends, J.Y.F. Dubois, and J.M. van Dijl<br />

Published in Microbiology and Molecular Biology Reviews (2006) 70, 755-788<br />

15


Chapter 2<br />

Summary<br />

<strong>The</strong> Gram-positive bacterium <strong>Staphylococcus</strong> <strong>aureus</strong> is a frequent component of the<br />

human microbial flora that can turn into a dangerous pathogen. As such, this organism<br />

is capable of infecting almost every tissue and organ system in the human body. It does<br />

so by actively exporting a variety of virulence factors to the cell surface and<br />

extracellular milieu. Upon reaching their respective destinations, these virulence factors<br />

have pivotal roles in the colonization and subversion of the human host. It is therefore of<br />

major importance to obtain a clear understanding of the protein transport pathways<br />

that are active in S. <strong>aureus</strong>. <strong>The</strong> present review aims to provide a state-of-the-art<br />

roadmap of staphylococcal <strong>secretome</strong>s, which include both protein transport pathways<br />

and the extracytoplasmic proteins of these organisms. Specifically, an overview is<br />

presented of the exported virulence factors, pathways for protein transport, signals for<br />

cellular protein retention or secretion, and the exoproteomes of different S. <strong>aureus</strong><br />

isolates. <strong>The</strong> focus is on S. <strong>aureus</strong>, but comparisons with <strong>Staphylococcus</strong> epidermidis and<br />

other Gram-positive bacteria like Bacillus subtilis are included where appropriate.<br />

Importantly, the results of genomic and proteomic studies on S. <strong>aureus</strong> <strong>secretome</strong>s are<br />

integrated through a comparative "secretomics" approach, resulting in a first definition<br />

of the core and variant <strong>secretome</strong>s of this bacterium. While the core <strong>secretome</strong> seems to<br />

be largely employed for general house-keeping functions, necessary to thrive in<br />

particular niches provided by the human host, the variant <strong>secretome</strong> seems to contain<br />

the “gadgets” that S. <strong>aureus</strong> needs to conquer these well-protected niches.<br />

16


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

General introduction and scope of this review<br />

<strong>The</strong> Gram-positive bacterium <strong>Staphylococcus</strong> <strong>aureus</strong> is a frequent component of the human<br />

microbial flora that can turn into a dangerous pathogen. As such, this organism is capable of<br />

infecting almost every tissue and organ system in the human body. It does so by exporting a<br />

variety of virulence factors to the cell surface and extracellular milieu of the human host. As<br />

in all living organisms (Wickner and Schekman, 2005), S. <strong>aureus</strong> contains several protein<br />

transport pathways, of which the general secretory (Sec) pathway is the most well known and<br />

best described. Proteins that need to be transported to an extracytoplasmic location contain, in<br />

general, an N-terminal signal peptide that is needed to target the newly synthesized protein<br />

from the ribosome to the translocation machinery in the cytoplasmic membrane. Next, the<br />

protein is threaded through the Sec translocon in an unfolded state. During this translocation<br />

step, or shortly thereafter, the signal peptide is removed by a so-called signal peptidase. Upon<br />

complete membrane translocation, the protein has to fold into its correct conformation and<br />

will then be retained in an extracytoplasmic compartment of the cell, or secreted into the<br />

extracellular milieu. In the case of Gram-positive cocci, such as S. <strong>aureus</strong> (Figure 1), we<br />

distinguish three extracytoplasmic subcellular compartments: the membrane, the membranecell<br />

wall interface and the cell wall. Since surface-exposed and secreted proteins of S. <strong>aureus</strong><br />

play pivotal roles in the colonization and subversion of the human host, it is of major<br />

importance to obtain a clear understanding of the protein transport pathways that are active in<br />

this organism (Lee and Schneewind, 2001). Knowledge about the protein sorting mechanism<br />

has become all the more relevant with the upcoming of staphylococcal resistance against lastdefence<br />

antibiotics, such as vancomycin. <strong>The</strong> scope of this review is to provide a state-of-theart<br />

roadmap of staphylococcal <strong>secretome</strong>s, which include both protein transport pathways and<br />

the extracytoplasmic proteins of these organisms. <strong>The</strong> focus is on S. <strong>aureus</strong>, but comparisons<br />

with <strong>Staphylococcus</strong> epidermidis and the best characterized Gram-positive bacterium Bacillus<br />

subtilis are included where appropriate. Importantly, the present review aims to integrate the<br />

results of genomic and proteomic studies on S. <strong>aureus</strong> <strong>secretome</strong>s, representing the first<br />

documented “comparative secretomics” study. Specifically, this review deals with known and<br />

predicted exported virulence factors, pathways for protein transport, signals for subcellular<br />

protein sorting or secretion, and the exoproteomes of different S. <strong>aureus</strong> isolates as defined by<br />

two-dimensional polyacrylamide gel electrophoresis (2D-PAGE) and mass spectrometry<br />

(Figures 2 and 3). <strong>The</strong> exoproteome is defined by all S. <strong>aureus</strong> proteins that can be identified<br />

in the extracellular milieu of this organism and thus includes proteins actively secreted by<br />

living cells and the remains of dead cells. For a clear appreciation of the present review, it is<br />

important to bear in mind that the proteins exported from the cytoplasm could be directly<br />

involved in staphylococcal virulence, whereas the respective protein export systems represent<br />

the “pathways to pathogenesis”.<br />

17


Chapter 2<br />

18<br />

A B<br />

Figure 1. Imaging of S. <strong>aureus</strong> RN6390. (A) For scanning electron microscopy a drop of washed culture of<br />

bacteria was fixated for 30 min with 2% glutaraldehyde in 0.1 M Cacodylate buffer, pH 7.38. Next, the fixated<br />

bacteria were placed on a piece (1 cm 2 ) of cleaved 0.1% Poly-L Lysine coated mica sheet and washed in 0.1 M<br />

Cacodylate buffer. This specimen was dehydrated in ethanol series consisting of 30%, 50%, 70%, 96% and<br />

anhydrous 100% solution (3X) 10 min each, then Critical point dried with CO 2 and sputter-coated with 2-3 nm<br />

Au/Pd (Balzers coater). <strong>The</strong> specimen was fixed on a SEM-stub-holder and observed in a JEOL FE-SEM 6301F.<br />

(B) Micrograph of a cluster of S. <strong>aureus</strong> cells grown in blood culture medium. <strong>The</strong> cells were fixed with ethanol and<br />

hybridized with the fluorescein labelled Peptide Nucleid Acid (PNA) probe PNA-Stau. <strong>The</strong> image was generated by<br />

merging an epi-fluorescence image with the negative of a phase-contrast image.<br />

Figure 2. <strong>The</strong> extracellular proteomes of different S. <strong>aureus</strong> strains<br />

Proteins of the growth medium fractions of different staphylococcal isolates, grown in TSB medium (37 °C) to an<br />

optical density at 540 nm (OD 540) of 10, were separated by 2D-PAGE using immobilized pH gradient (IPG) strips<br />

in the pH range of 3 to 10 (Amersham <strong>Pharma</strong>cia Biotech, Piscataway, N. J.). Each gel was loaded with 350 µg<br />

protein extracts and, after electrophoresis, stained with Colloidal Coomassie. Proteins were identified by MALDI-<br />

TOF mass spectrometry. <strong>The</strong> corresponding protein spots are labelled with protein names according to the S.<br />

<strong>aureus</strong> N315 database or NCBI entries for proteins not present in N315. <strong>The</strong> S. <strong>aureus</strong> strains that were used in<br />

these experiments are RN6390 and COL, and four clinical isolates from the University Medical Center Groningen<br />

named MRSA693331, 035699y/bm, 0440579/rmo, CA-MRSA021708m/rmo.


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

Figure 3. Dynamics of the amount of extracellular proteins during growth of S. <strong>aureus</strong> RN6390 in TSB<br />

medium<br />

(A) Individual dual channel 2D patterns of extracellular proteins during the different phases of the growth curve<br />

of cells grown in TSB medium were assembled into a movie. <strong>The</strong> protein pattern at an OD 540 of 1 (labelled in<br />

green) was compared with the protein pattern at the respective higher optical densities (labelled in red). As a<br />

consequence of the dual channel labelling, spots of which the intensities do not differ in the compared gels will be<br />

yellow; spots of different intensities will be either green or red (Bernhardt et al., 1999). (B) Growth curve of S.<br />

<strong>aureus</strong> RN6390 grown in TSB medium as determined by OD 540 readings. <strong>The</strong> sampling points for proteomics<br />

analyses are indicated by arrows. (C) Proteomic signatures of selected proteins representing different regulatory<br />

groups as revealed by dual channel imaging. <strong>The</strong> relative amounts of the respective proteins at an OD 540 of 1<br />

(spots labelled in green) of cells grown in TSB medium were compared with the relative amounts of these proteins<br />

at higher optical densities (spots labelled in red). Proteins were stained with Sypro Ruby ® .<br />

Exported staphylococcal virulence factors<br />

S. <strong>aureus</strong> and S. epidermidis are organisms that occur naturally in and on the human body.<br />

While S. epidermidis is mostly present on the human skin, S. <strong>aureus</strong> can be found on mucosal<br />

surfaces. S. <strong>aureus</strong> is carried by 30-40% of the population (Peacock et al., 2001) and can<br />

readily be identified in the nose, but the organism can also be detected in other moist regions<br />

of the human body, such as axilla, perineum, vagina and rectum, thereby forming a major<br />

reservoir for infections. Although most staphylococcal infections are nosocomial (i.e.<br />

hospital-acquired), an increase in the number of cases of community-acquired antibiotic<br />

(methicillin) resistant infections is currently observed world-wide (Centers for Disease<br />

Control and Prevention, 2003; Grundmann et al., 2002; Vandenesch et al., 2003). <strong>The</strong> risk of<br />

intravascular and systemic infection by S. <strong>aureus</strong> rises when the epithelial barrier is disrupted<br />

by intravascular catheters, implants, mucosal damage or trauma. Interestingly, after infection,<br />

19


Chapter 2<br />

cells of S. <strong>aureus</strong> can persist unnoticed in the human body for long periods of time (years)<br />

after which they can suddenly cause another infection. S. <strong>aureus</strong> is primarily an extracellular<br />

pathogen whose colonization and invasion of human tissues and organs can lead to severe<br />

cytotoxic effects. Nevertheless, S. <strong>aureus</strong> can also be internalized by various cells, including<br />

non-phagocytic cells, which seems to induce apoptosis (Hauck and Ohlsen, 2006; da Silva et<br />

al., 2004; Mempel et al., 2002). Although S. <strong>aureus</strong> has the potential to form biofilms (Götz,<br />

2002), S. epidermidis infections are particularly notorious for the formation of thick<br />

multilayered biofilms on indwelling catheters and other implanted devices. <strong>The</strong> formation of<br />

such a biofilm takes place in several steps during which the bacteria first adhere rapidly to the<br />

surface of the polymer material that has been coated with a film of proteinaceous and nonproteinaceous<br />

organic host molecules (Escher and Characklis, 1990). Bacteria that adhere to<br />

this film produce extracellular polymeric substances, mostly polysaccharides and proteins, in<br />

turn resulting in a strong attachment to the polymer surface and other bacteria in the growing<br />

biofilm. Ultimately, the biofilm is composed of multiple layers of cells, cellular debris,<br />

polysaccharides and proteins. S. epidermidis proteins that are essential for biofilm formation<br />

are, for example, the polysaccharide intercellular adhesin (PIA) (Mack et al., 1996), the<br />

accumulation associated protein (AAP) (Rohde et al., 2005) and the biofilm-associated<br />

protein (Bap) (Tormo et al., 2005). PIA is most likely identical to the polysaccharide adhesion<br />

(PS/A).<br />

Virulence of S. <strong>aureus</strong><br />

<strong>The</strong> pathogenicity of S. <strong>aureus</strong> is caused by the expression of an arsenal of virulence factors<br />

(Table 1), which can lead to superficial skin lesions such as styes, furunculosis and<br />

paronychia, or to more serious infections such as pneumonia, mastitis, urinary tract infections,<br />

osteomyelitis, endocarditis and even sepsis. In very rare cases, S. <strong>aureus</strong> causes meningitis.<br />

<strong>The</strong> virulence factors that S. <strong>aureus</strong> employs to cause these diseases are displayed at the<br />

surface of the staphylococcal cell or secreted into the host milieu (Fedtke et al., 2004).<br />

Specifically, these virulence factors include (a) surface proteins that promote adhesion to and<br />

colonization of host tissues; (b) invasins that are exported to an extracytoplasmic location and<br />

promote bacterial spread in tissues (leukocidin, kinases, hyaluronidase); (c) surface factors<br />

that inhibit phagocytic engulfment (capsule, Protein A); (d) biochemical properties that<br />

enhance staphylococcal survival in phagocytes (carotenoids, catalase production); (e)<br />

immunological disguises (Protein A, coagulase, clotting factor); (f) membrane-damaging<br />

toxins that disrupt eukaryotic cell membranes (hemolysins, leukotoxin); (g) superantigens that<br />

contribute to the symptoms of septic shock (SEA-G, TSST, ET); and (h) determinants for<br />

inherent and acquired resistance to antimicrobial agents. Most virulence factors are expressed<br />

in a coordinated fashion during the growth cycle of S. <strong>aureus</strong>. <strong>The</strong> best characterized<br />

regulators of virulence factors are the accessory gene regulator (agr) (Morfeldt et al., 1988;<br />

Peng et al., 1988; Recsei et al., 1986) and the Staphylococcal accessory regulator (SarA)<br />

(Cheung et al., 1992; Cheung and Projan, 1994). Ziebandt et al. (Ziebandt et al., 2004)<br />

showed that extracellular proteins can be divided into two groups, based on the timing of their<br />

expression in cells grown in tryptic soy broth (TSB): proteins that are mainly expressed at low<br />

cell densities, or proteins exclusively expressed at high cell densities. Agr seems to be an<br />

important positive regulator of proteins that are expressed at higher optical densities (e.g.<br />

proteases, hemolysins and lipases) and a negative regulator for proteins that are expressed<br />

during the exponential growth phase (e.g. immunodominant antigen A, secretory antigen<br />

20


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

precursor and several proteins with unknown functions). In addition, Gill et al. (Gill et al.,<br />

2005) identified 15 other two-component regulatory systems in the genomes of S. <strong>aureus</strong> and<br />

S. epidermidis that are potentially involved in staphylococcal virulence. In this respect, it is<br />

interesting to note that the antibiotic cerulenin, which is known to inhibit protein secretion by<br />

S. <strong>aureus</strong> at sub-MIC levels, was recently reported to block transcriptional activation of at<br />

least two regulatory determinants, agr and sae. Thus, it seems that cerulenin inhibits the<br />

transcription of genes for secretory proteins rather than the secretion process of these proteins<br />

(Adhikari and Novick, 2005). In contrast, it was previously believed that cerulenin would<br />

interfere with membrane function through an inhibition of normal fatty acid synthesis.<br />

Table 1. Virulence factors of S. <strong>aureus</strong><br />

Pathogenic action Virulence factors Protein or other compound Functions<br />

Colonization of host<br />

tissues<br />

Lysis eukaryotic cell<br />

membranes and<br />

bacterial spread<br />

Inhibition phagocytic<br />

engulfment<br />

Survival in<br />

phagocytes<br />

Immunological<br />

disguise and<br />

modulation<br />

Contribution to<br />

symptoms of septic<br />

shock<br />

Acquired resistance to<br />

antimicrobial agents<br />

Surface proteins ClfA, ClfB, FnbA, FnbB, IsdA Adhesins, fibronectin and<br />

SdrC, SdrD, SdrE,<br />

fibrinogen-binding proteins<br />

Membrane- Geh, Hla, Hld, HlgA-C, HysA, Hemolysins, hyaluronidase,<br />

damaging toxins, Lip, LukD, LukE, LukF, LukS, leukocidin, leukotoxin,<br />

invasins<br />

Nuc<br />

lipases, nucleases<br />

Surface factors CapA-P, Efb, Spa Capsule, protein A<br />

Biochemical KatA, Staphyloxanthin Carotenoids, catalase<br />

compounds<br />

production<br />

Surface proteins ClfA, ClfB, Coa, Spa Clumping factor, coagulase,<br />

protein A<br />

Exotoxins Eta, Etb, SEA-G, TSST-1 Enterotoxins SEA-G,<br />

exfoliative toxin, toxic<br />

shock syndrome toxin<br />

TSST<br />

Resistance proteins BlaZ, MecA, VanA MRSA, VRSA<br />

Notably, to date relatively little information is available on the molecular nature of the stimuli<br />

that are perceived by the major regulators of the expression of virulence factors. Overall, it<br />

should be clear that strain-specific differences in gene regulation by agr, sae or other<br />

regulators may dramatically influence the repetoire of produced virulence factors, thereby<br />

having a profound impact on the disease-causing potential of different strains. Since the<br />

interplay of different regulators and cell-to-cell communication can impact differently on the<br />

expression of virulence factors, even the disease-causing potential of individual S. <strong>aureus</strong><br />

cells within a genetically identical population may vary.<br />

Resistance of S. <strong>aureus</strong> to antibiotics<br />

Resistance of S. <strong>aureus</strong> to antibiotics has been observed very soon after the introduction of<br />

penicillin about sixty years ago. In the following years, the amazing ability of staphylococci<br />

to develop resistance to antibiotics has resulted in the emergence of methicillin-resistant S.<br />

<strong>aureus</strong> (MRSA) and S. epidermidis (MRSE) strains. In fact, methicillin resistance was<br />

observed already in 1961 in nosocomial isolates of S. <strong>aureus</strong>, one year after the introduction<br />

of methicillin (Jevons, 1961). <strong>The</strong> resistance towards methicillin is a result of the production<br />

of an altered penicillin binding protein, PBP2a (or PBP2’), which has less affinity to most βlactam<br />

antibiotics. <strong>The</strong> PBP2a protein, which is located at the membrane-cell wall interface, is<br />

of major importance for cell wall biogenesis by mediating the cross linking of peptidoglycans.<br />

21


Chapter 2<br />

PBP2a is encoded by the mecA gene, which is located on a mobile genetic element, also<br />

known as the staphylococcal cassette chromosome (SCC) mec (Chambers, 1997; Ito et al.,<br />

2004). <strong>The</strong> SCCmec element is a basic mobile genetic element that serves as a vehicle for<br />

gene exchange among staphylococcal species (Dobrindt et al., 2004). In addition to the mecA<br />

gene, SCCmec carries the mecA regulatory genes mecI and mecR, an insertion sequence<br />

element (IS431mec) and a unique cassette of recombinase genes (ccr), which are responsible<br />

for SCCmec chromosomal integration and excision. Eight different types of SCCmec<br />

elements, type I-V, have been identified so far, based on the classes of mecA gene and ccr<br />

gene complexes (Ito et al., 2009). Notably, type II and III elements contain, besides mecA,<br />

multiple determinants for resistance against non-β-lactam antibiotics. Accordingly, type II and<br />

III SCCmec elements are responsible for multidrug resistance in nosocomial MRSA isolates.<br />

Some SCCmec elements (e.g. type IV SCCmec), contain no other resistance gene than mecA,<br />

and they are significantly smaller compared to for example the type II and III elements. This<br />

might serve as an evolutionary advantage, making it easier for these mobile genetic elements<br />

to spread across bacterial populations. Phylogenetic analyses of the genes encode by SCCmec<br />

elements showed distant relationships with homologues in other S. <strong>aureus</strong> genomes and<br />

suggest foreign origins for these genes.<br />

Vancomycin resistance has been first reported for Enterococcus faecium (Leclercq et al.,<br />

1989), and transfer of vancomycin resistance from enterococci, such as Enterococcus faecalis,<br />

to S. <strong>aureus</strong> has been shown to occur (Noble et al., 1992). Vancomycin has long been a last<br />

resort antibiotic for multiple resistant S. <strong>aureus</strong> strains, but already in 1996 a strain was<br />

isolated, which showed a reduced sensitivity towards vancomycin (Hiramatsu et al., 1997).<br />

Shortly afterwards, additional strains were isolated in different countries that were designated<br />

as vancomycin-intermediately resistant S. <strong>aureus</strong> (VISA). <strong>The</strong>se strains show a significantly<br />

thickened cell wall, which allows them to sequester more vancomycin than non-VISA strains,<br />

thereby preventing the detrimental effects of this antibiotic (Cui et al., 2003). A search for the<br />

genetic basis of the lowered vancomycin sensitivity of the S. <strong>aureus</strong> Mu50 strain revealed that<br />

important genes for cell wall biosynthesis and intermediary metabolism have mutations<br />

compared to MRSA strains, which might lead to altered expression of genes involved in the<br />

cell wall metabolism and a thickened cell wall (Avison et al., 2002). <strong>The</strong> first highly<br />

vancomycin resistant strain was isolated in 2002 (Weigel et al., 2003). This strain was shown<br />

to carry a plasmid, which contains, among other resistance genes, the vanA gene plus several<br />

additional genes required for vancomycin resistance. <strong>The</strong> proteins encoded by these genes are<br />

responsible for replacing the C-terminal D-alanyl-D-alanine (D-ala-D-ala) of the disaccharide<br />

pentapeptide cell wall precursor with a depsipeptide, D-alanyl-D-lactate (D-ala-D-lac),<br />

thereby lowering the cell wall affinity for vancomycin (Bugg et al., 1991).<br />

Export of virulence factors from the cytoplasm<br />

As most proteinaceous virulence factors are displayed at the surface of the staphylococcal cell<br />

or released into the medium, it is important for our understanding of the pathogenic potential<br />

of these organisms to map their pathways for protein transport. While specific questions<br />

relating to surface display or secretion of particular virulence factors have been addressed for<br />

several years, more holistic studies on the genomics and proteomics of these processes have<br />

been documented in the scientific literature only very recently. Moreover, no systematic<br />

analysis of pathways and cellular machinery for protein transport has thus far been performed<br />

for staphylococci. This review is aimed at filling this knowledge gap. To do so, we have taken<br />

22


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

full advantage of the availability of six completely sequenced and annotated S. <strong>aureus</strong><br />

genomes and one of the two sequenced S. epidermidis strains, as well as recently published<br />

data on the analysis of staphylococcal cell wall- and exoproteomes. Additionally, we have<br />

combined the published information with bioinformatics-derived data on all potential signals<br />

for protein export from the cytoplasm and secretion into the extracellular milieu, or retention<br />

in the membrane or cell wall. Since polytopic membrane proteins do not appear to have major<br />

direct roles in virulence other than causing drug resistance, such membrane proteins remain<br />

beyond the scope of this review. Furthermore, since the <strong>secretome</strong> of B. subtilis has been<br />

characterized extensively, both at the level of the protein export machinery and the<br />

exoproteome, we have compared the staphylococcal <strong>secretome</strong>s with that of B. subtilis. To<br />

our knowledge this has resulted in the first “comparative secretomics” study.<br />

S. <strong>aureus</strong> strains suitable for comparative secretomics<br />

Fourteen sequenced and fully annotated genomes of S. <strong>aureus</strong> are available in public<br />

databases (Table 2; http://www.ncbi.nlm.nih.gov/genomes/lproks.cgi) and thirteen of these<br />

genomes were used in the present study. <strong>The</strong>se include one of the first hospital-acquired<br />

MRSA isolates, S. <strong>aureus</strong> COL (Gill et al., 2005), which has been widely used in research on<br />

staphylococcal methicillin and vancomycin resistance. <strong>The</strong> sequenced MRSA252 strain<br />

(Holden et al., 2004) is a hospital-acquired epidemic strain, which was isolated from a patient<br />

who died as a consequence of septicemia. <strong>The</strong> sequenced MSSA476 strain (Holden et al.,<br />

2004) is a community-acquired invasive strain that is penicillin- and fusidic acid-resistant, but<br />

susceptible to most commonly used antibiotics. S. <strong>aureus</strong> Mu50 and N315 (Kuroda et al.,<br />

2001) are hospital-acquired MRSA strains isolated from Japanese patients. In addition, the<br />

Mu50 strain displays vancomycin intermediate sensitivity. <strong>The</strong> S. <strong>aureus</strong> Mu3 strain was the<br />

first isolated strain from a healthy carrier in Brazil that showed vancomycin-resistance<br />

(Hiramatsu et al., 1997; Neoh et al., 2008). <strong>The</strong> community-acquired S. <strong>aureus</strong> strains MW2<br />

(Baba et al., 2002), USA300 and USA300_TCH1516 (Diep et al., 2006) are highly virulent<br />

MRSA strains, isolated in the USA. Both S. <strong>aureus</strong> JH1 and JH9 strains are MRSA strains<br />

that were isolated from one patient undergoing vanconcomycin treatment on different time<br />

points. <strong>The</strong> JH1 strain was the earliest strain isolated from the patient. <strong>The</strong> JH9 strain was<br />

isolated at a later stage of the treatment and was diagnosed as a vancomycin-resistant strain<br />

(Mwangi et al., 2007). Comparison between these two strains would gain insight in the<br />

evolution of isogenic strains and the acquirement of vancomycin resistance under antibiotic<br />

pressure. <strong>The</strong> Newman strain (Baba et al., 2008) was isolated from a human infection (Duthie<br />

and Lorenz, 1952) and has been widely used as a research strain due to its robust virulence<br />

phenotypes. Finally, the NCTC 8325 strain (Gillaspy et al., 2006) is generally regarded as the<br />

prototypical strain for all genetic midifications in order to address specific gene regulatory<br />

and virulence traits. Furthermore, the sequence of S. <strong>aureus</strong> RF122, a strain that is associated<br />

with mastitis in cattle, is now also available in the NCBI database (Herron et al., 2002), but<br />

has not been included in the present review which is primarily focused on staphylococcal<br />

pathogenicity towards humans. Secretome predictions for this strain are presented in<br />

Appendix IIIH. Using Multilocus Sequence Typing with seven housekeeping genes of the<br />

different S. <strong>aureus</strong> strains, Holden et al. (Holden et al., 2004) showed that the MRSA252<br />

strain is phylogenetically most distantly related to the other sequenced strains, while the<br />

Mu50 and N315 strains are indistinguishable by MLST, and the same is true for the<br />

MSSA476 and MW2 strains. <strong>The</strong> COL and NCTC8325 strains are relatively closely related to<br />

23


Chapter 2<br />

each other. However, analysis of the two major pathogenicity islands present in all these<br />

strains shows that the distribution of these pathogenicity islands gives contradictory results on<br />

phylogenetic relationships of the sequenced S. <strong>aureus</strong> strains (Baba et al., 2008).<br />

Table 2. Sequenced and annotated genomes of S. <strong>aureus</strong> strains<br />

Genome size (kbp) Nr. Of protein encoding genes<br />

Strain Origin a Chromosome Plasmid Chromosome Plasmid<br />

COL HA- MRSA 2809 4 2615 3<br />

JH1 HA- MRSA 2907 30 2747 33<br />

JH9 HA- VISA 2907 3 2697 29<br />

MRSA252 HA- MRSA 2903 - 2656 -<br />

MSSA476 CA- MSSA 2800 21 2579 19<br />

Mu3 HA- VISA 2880 - 2698 -<br />

Mu50 HA- VISA 2879 25 2697 34<br />

MW2 CA- MRSA 2820 - 2632 -<br />

N315 HA- MRSA 2815 25 2588 31<br />

NCTC8325 HA- MSSA 2821 - 2892 -<br />

Newman HA- MRSA 2879 - 2614 -<br />

USA300 CA- MRSA 2873 3,4,37 2560 5,3,36<br />

USA300_TCH1516 CA- MRSA 2873 27,20 2657 26,20<br />

RF122 Bovine mastitis 2743 - 2515 -<br />

a HA-MRSA: Hospital-acquired MRSA; CA-MRSA: Community-acquired MRSA<br />

Sequenced and annotated genomes of other staphylococcal species, such as S. epidermidis,<br />

<strong>Staphylococcus</strong> haemolyticus and <strong>Staphylococcus</strong> carnosus, are also publicly available.<br />

However, with the exception of the S. epidermidis strain ATCC 12228 (Zhang et al., 2003),<br />

these are not included in the present review, which is focused primarily on S. <strong>aureus</strong>. A<br />

comparative genomic analysis of S. <strong>aureus</strong> COL, Mu50, MW2, N315 and the sequenced S.<br />

epidermidis strains RP62A and ATCC 12228 has revealed that these species and strains have<br />

a set of 1681 genes in common (Gill et al., 2005). In contrast, 454 genes are present in the S.<br />

<strong>aureus</strong> strains, but not in S. epidermidis, whereas 286 genes are present in S. epidermidis, but<br />

not in S. <strong>aureus</strong>. Most of the strain-specific and species-specific genes can be related to the<br />

presence or absence of particular prophages and genomic islands.<br />

Pathways for staphylococcal protein transport<br />

<strong>The</strong> bacterial machinery for protein transport is currently best-described for Escherichia coli<br />

(Gram-negative) and B. subtilis (Gram-positive) (for reviews see (de Keyzer et al., 2003;<br />

Tjalsma et al., 2000; Tjalsma et al., 2004). Many of the known components that are involved<br />

in the different routes for protein export from the cytoplasm and post-translocational<br />

modification of exported proteins in these organisms are also conserved in S. <strong>aureus</strong> and S.<br />

epidermidis (Table 3). In general, proteins that are exported are synthesized with an Nterminal<br />

signal peptide, which directs them into a particular transport pathway. Consequently,<br />

the presently known signal peptides are classified according to the export pathway into which<br />

they direct the corresponding proteins, or the type of signal peptidase that is responsible for<br />

their removal (processing) upon membrane translocation.<br />

24


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

Table 3. Secretion machinery of S. <strong>aureus</strong>, S. epidermidis and B. subtilis<br />

Sec-pathway S. <strong>aureus</strong> S. epidermidis B. subtilis<br />

Chaperone Ffh + + +<br />

FtsY + + +<br />

FlhF - - +<br />

CsaA - - +<br />

Translocation motor SecA1<br />

+<br />

+<br />

+<br />

SecA2<br />

+<br />

+<br />

-<br />

Translocation channel SecY1<br />

+<br />

+<br />

+<br />

SecY2<br />

+<br />

+<br />

-<br />

SecE + + +<br />

SecG + + +<br />

SecDF + + +<br />

YajC (YrbF) + + +<br />

Lipid modification Lgt + + +<br />

Sec-pathway S. <strong>aureus</strong> S. epidermidis B. subtilis<br />

Signal peptidase SpsA (inactive) + + -<br />

SpsB (SipSTUV) + + a +<br />

SipW (ER-type) - - +<br />

LspA + + b +<br />

Folding catalyst PrsA + c + +<br />

BdbC - - +<br />

DsbA (BdbD) + + +<br />

Cell wall anchoring SrtA + + -<br />

SrtB + - -<br />

SrtC - + d -<br />

SrtD - - +<br />

Tat-pathway<br />

Translocase TatA<br />

TatC<br />

Pseudopilin pathway<br />

Bacteriocins<br />

Holins<br />

Ess<br />

25<br />

+<br />

+<br />

ComGA + + +<br />

ComGB + + +<br />

ComC + + +<br />

Bacteriocin-specific<br />

ABC-transporters<br />

-<br />

-<br />

? ? +<br />

CidA (holin) + + +<br />

LrgA (anitholin) + + +<br />

EsaA + - +<br />

EsaB + - +<br />

EsaC + e - -<br />

EssA + - -<br />

EssB + - +<br />

EssC + - +<br />

Based on BLAST searches with the corresponding proteins of B. subtilis in the finished genome database<br />

(http://www.ncbi.nlm.nih.gov/sutils/genom_table.cgi).<br />

a<br />

Two potentially active type I SPases are present in this strain and share homology to B. subtilis SipS and SipU<br />

b<br />

Two LspA proteins present in this strain<br />

c<br />

This protein is truncated at the C-terminus in the S. <strong>aureus</strong> JH9 strain<br />

d<br />

<strong>The</strong> genome of S. epidermidis RP62A only contains a srtA gene, whereas the genome of S. epidermidis<br />

ATCC12228 also contains a srtC gene<br />

e<br />

This protein is missing in the S. <strong>aureus</strong> MRSA252 strain<br />

<strong>The</strong> staphylococcal protein export pathways that have been characterized experimentally or<br />

that can be deduced from sequenced genomes are schematically shown in Figure 4 and will be<br />

discussed in the following sections. Since these pathways are likely to be used for the export<br />

+<br />

+


Chapter 2<br />

of virulence factors to the cell surface and the milieu of the host, Figure 4 can be regarded as a<br />

subcellular road map to staphylococcal pathogenesis.<br />

Components of the general secretory (Sec) Pathway<br />

26<br />

Figure 4. <strong>The</strong> staphylococcal “pathways to<br />

pathogenesis”. Schematic representation of a<br />

staphylococcal cell with potential pathways for<br />

protein sorting and secretion. (A) Proteins<br />

without signal peptide reside in the cytoplasm.<br />

(B) Proteins with one or more transmembrane<br />

spanning domains can be inserted into the<br />

membrane via the Sec, Tat or Com pathways. (C)<br />

Lipoproteins are exported via the Sec pathway<br />

and after lipid-modification anchored to the<br />

membrane. (D) Proteins with cell wall retention<br />

signals are exported via the Sec, Tat or Com<br />

pathways and retained in the cell wall via<br />

covalent-, or high-affinity binding to cell wall<br />

components. (E) Exported proteins with a signal<br />

peptide and without a membrane or cell wall<br />

retention signal can be secreted into the<br />

extracellular milieu via the various indicated<br />

pathways.<br />

<strong>The</strong> most commonly used pathway for bacterial protein transport is the general secretory<br />

(Sec) pathway. Specifically this pathway is responsible for the secretion of the majority of the<br />

proteins found in the exoproteome of B. subtilis and this is probably also the case for most<br />

other Gram-positive bacteria, including S. <strong>aureus</strong> (Tjalsma et al., 2004). Unfortunately, there<br />

are only very few published data available concerning the Sec pathway of S. <strong>aureus</strong> and,<br />

therefore, we will fill in the current knowledge gaps with data obtained from studies in B.<br />

subtilis or E. coli. Proteins that are exported via the Sec-pathway contain signal peptides with<br />

recognition sites for so-called type I or type II signal peptidases (SPases). Notably, type II<br />

SPase recognition sites overlap with the recognition sites for the diacylglyceryl transferase<br />

Lgt. Precursor proteins with a type II SPase recognition sequence are lipid-modified prior to<br />

processing and the resulting mature proteins are retained as lipopoteins in the cytoplasmic<br />

membrane via their diacylglyceryl moiety. Furthermore, the Sec-dependent export of proteins<br />

can be divided into three stages: a) targeting to the membrane translocation machinery by<br />

export-specific or general chaperones, b) translocation across the membrane by the Sec<br />

machinery, and c) post-translocational folding and modification. If the translocated proteins<br />

of Gram-positive bacteria lack specific retention signals for the membrane or cell wall, they<br />

are secreted into the growth medium.<br />

Preprotein targeting to the membrane<br />

In B. subtilis the only known secretion-specific chaperone is the signal recognition particle<br />

(SRP), which consists of the small cytoplasmic RNA (scRNA), the histon-like protein HBsU<br />

and the Ffh protein. Ffh and HBsU bind to different moieties of the scRNA. Studies in E. coli<br />

have shown that, upon emergence from the ribosome, the signal peptide of a nascent secretory<br />

protein can be recognized by several cytoplasmic chaperones and/or targeting factors, such as<br />

Ffh or Trigger Factor (TF) (Eisner et al., 2003). In contrast to Ffh, which is required for co-


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

translational protein export in E. coli, the cytoplasmic chaperone SecB has mainly been<br />

implicated in post-translational protein targeting. Notably however, SecB is absent from the<br />

sequenced Gram-positive bacteria, including S. <strong>aureus</strong> and B. subtilis. Most likely, ribosomenascent<br />

chain complexes of S. <strong>aureus</strong> are thus targeted to the membrane by SRP, which, by<br />

analogy to B. subtilis and E. coli will probably involve the SRP receptor FtsY. At the<br />

membrane, the nascent preprotein will be directed to the translocation machinery. This<br />

process is likely to be stimulated by negatively charged phospholipids (De Leeuw et al.,<br />

2000), the Sec translocon (Bibi, 1998; De Leeuw et al., 2000) and/or the SecA protein (Bunai<br />

et al., 1999). In this respect SecA may not only function as the translocation motor (see<br />

below), but also as a chaperone for preprotein targeting (Herbort et al., 1999). While it has<br />

been shown that Ffh is essential for growth and viability in E. coli and B. subtilis, this does<br />

not seem to be the case in all bacteria. For example, Ffh, FtsY and scRNA are not essential in<br />

Streptococcus mutans. In this organism the SRP is merely required for growth under stressful<br />

conditions, such as low pH (


Chapter 2<br />

homologues are not essential for growth and viability. It is presently unknown whether SecA2<br />

and SecY2 transport specific proteins across the membrane of S. <strong>aureus</strong>. However, it has been<br />

shown for other pathogenic Gram-positive bacteria, which also possess a second set of SecA<br />

and SecY, that these proteins are required for the transport of certain proteins related to<br />

virulence. In Streptococcus gordonii, the export of GspB, a large cell-surface glycoprotein<br />

that contributes to platelet binding, seems to be dependent on the presence of SecA2 and<br />

SecY2 (Bensing and Sullam, 2002). This protein contains large serine-rich repeats, an LPxTG<br />

motif for cell wall anchoring (see below), and a very large signal peptide of 90 amino acids.<br />

In Streptococcus parasanguis two other proteins, FimA and Fap1, are known to be secreted<br />

via SecA2-dependent membrane translocation. FimA is a (predicted) lipoprotein, which is a<br />

major virulence factor implicated in streptococcal endocarditis. <strong>The</strong> FimA homologue in S.<br />

<strong>aureus</strong> is a manganese-binding lipoprotein (MntA), associated with an ATP-binding cassette<br />

(ABC) transporter. Fap1 of S. parasanguis is involved in adhesion to the surface of teeth.<br />

Like GspB of S. gordonii, Fap1 has a long signal peptide of 50 amino acids, serine-rich<br />

repeats and an LPxTG motif for cell wall anchoring. To date, it is not known what determines<br />

the difference in the specificity of SecA1/SecY1 and SecA2/SecY2 translocases. However,<br />

for S. gordonii it has been shown that Gly residues in the signal peptide are important for<br />

directing GpsB to the SecA2/SecY2 translocon (Bensing et al., 2007). It is also not known<br />

whether the SecA2/SecY2 shares SecE and/or SecG with the SecA1/SecY1 translocase, and<br />

whether these translocases function completely independently from each other or whether<br />

mixed translocases can occur. Clearly, the secE and secG genes are not duplicated in S.<br />

<strong>aureus</strong>.<br />

In E. coli, the heterotrimeric SecYEG complex is associated with another heterotrimeric<br />

complex that is composed of the SecD, SecF and YajC proteins (Nouwen et al., 2005). This<br />

complex has been shown to be involved in the cycling of SecA (Driessen et al., 1998) and<br />

release of the translocated protein from the translocation channel (Matsuyama et al., 1993).<br />

SecD and SecF are separate, but structurally related proteins in most bacteria, including E.<br />

coli. Interestingly, in B. subtilis and S. <strong>aureus</strong>, natural gene fusions between the secD and<br />

secF genes are observed. Accordingly, the corresponding SecDF proteins can be regarded as<br />

molecular “Siamese twins” (Bolhuis et al., 1998). Unlike SecA, SecY and SecE, the SecDF<br />

protein of B. subtilis is not essential for growth and viability and its role in protein secretion is<br />

presently poorly understood (Bolhuis et al., 1998). B. subtilis secDF mutants only showed a<br />

mild secretion defect under conditions of high-level synthesis of secretory proteins. <strong>The</strong><br />

known SecDF proteins have 12 (predicted) transmembrane domains with two large<br />

extracytoplasmic loops between the first and second transmembrane segments, and between<br />

the seventh and eighth transmembrane segments. For E. coli SecD it has been shown that<br />

small deletions in the large extracytoplasmic loop result in a malfunctioning of the protein,<br />

while the stability of the SecD/F-YajC complex is not affected (Nouwen et al., 2005). It has<br />

therefore been proposed that this loop in SecD might provide a protective structure in which<br />

translocated proteins can fold more efficiently. <strong>The</strong> large extracytoplasmic loop in SecF has<br />

been proposed to interact with SecY, thereby stabilizing the translocation channel formed by<br />

SecYEG. Homologues of the E. coli YajC protein are present in many bacteria, including S.<br />

<strong>aureus</strong> and B. subtilis (YrbF), but their role in protein secretion has not been established yet.<br />

It is presently not known whether the S. <strong>aureus</strong> SecDF-YajC complex associates specifically<br />

with the SecA1/SecY1 translocase, the SecA2/SecY2 translocase, or both translocases.<br />

28


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

Type I Signal peptidases<br />

Signal peptides of preproteins are cleaved during or shortly after translocation by SPase I or<br />

SPase II, depending on the nature of the signal peptide (Tjalsma et al., 2001; van Roosmalen<br />

et al., 2004). <strong>The</strong> B. subtilis chromosome encodes five type I SPases, named SipS, SipT,<br />

SipU, SipV and SipW (van Dijl et al., 1992; Tjalsma et al., 1997; Tjalsma et al., 1998). Two<br />

of these, SipS and SipT, are of major importance for the processing of secretory preproteins,<br />

growth and viability. In S. <strong>aureus</strong> only two SPase I homologues are present, SpsA and SpsB.<br />

<strong>The</strong> catalytically active SPase I in S. <strong>aureus</strong> is SpsB, which is probably essential for growth<br />

and viability (Cregg et al., 1996). This SPase can be used to complement an E. coli strain that<br />

is temperature-sensitive for preprotein processing. In general, type I SPases recognize<br />

residues at the -1 and -3 positions relative to the cleavage site (van Roosmalen et al., 2004).<br />

For B. subtilis it has been shown that all secretory proteins identified by proteomics have Ala<br />

at the -1 position, and 71% of these secretory proteins have Ala at the -3 position (Tjalsma et<br />

al., 2004). In contrast, various residues are tolerated at the -2 position, including Ser, Lys,<br />

Glu, His, Tyr, Gln, Gly, Phe, Leu, Ala, Asp, Asn, Trp and Pro. Interestingly, Bruton et al.<br />

(Bruton et al., 2003) studied the cleavage sites in substrates of SpsB of S. <strong>aureus</strong> and showed<br />

that this enzyme has a preference for basic residues at the -2 position and tolerance for<br />

hydrophobic residues at this position. However, an acidic residue at the -2 position resulted in<br />

a significantly reduced rate of processing. <strong>The</strong> second SPase I homologue of S. <strong>aureus</strong> (SpsA)<br />

appears to be inactive, since it lacks the catalytic Ser and Lys residues that are, respectively,<br />

replaced with Asp and Ser residues. <strong>The</strong> presence of an apparently catalytically inactive SpsA<br />

homologue is a conserved feature of all staphylococci with sequenced genomes. Notably, in<br />

addition to an inactive SpsA homologue, S. epidermidis contains two SpsB homologues that<br />

respectively show the greatest similarity to SipS and SipU of B. subtilis. To date, it is not<br />

known whether the inactive SpsA homologues contribute somehow to protein secretion in<br />

these organisms.<br />

Lipid-modification of lipoproteins<br />

In E. coli, lipid-modification of prolipoproteins involves three sequential steps that are<br />

catalyzed by cytoplasmic membrane-bound proteins. <strong>The</strong> first step involves the transfer of a<br />

diacylglyceryl group from phosphatidylglycerol to the sulfhydryl group of the invariant Cys<br />

residue that is present at the +1 position of the signal peptide cleavage site in lipoprotein<br />

precursors. This step is catalyzed by a phosphatidyl glycerol diacylglyceryl transferase (Lgt)<br />

as was shown for E. coli by Sankaran et al. (Sankaran and Wu, 1994). <strong>The</strong> recognition<br />

sequence for Lgt, which includes the Cys residue that becomes diacylglyceryl-modified, is<br />

known as the lipobox. <strong>The</strong> lipid-modification of the lipobox Cys residue is necessary for the<br />

lipoprotein-specific type II signal peptidase (LspA) to recognize and cleave the signal peptide<br />

of a prolipoprotein, which represents the second step in lipoprotein modification. <strong>The</strong> third<br />

step involves the transfer of an N-acyl group by an N-acyl transferase (Lnt), resulting in the<br />

formation of N-acyl diacylglycerylcysteine at the N-terminus of the mature lipoprotein.<br />

Although Lgt and LspA are present in most, if not all bacteria, Lnt is only present in Gramnegative<br />

bacteria (Tjalsma et al., 2001). As for other Gram-positive bacteria, no homologue<br />

of Lnt could be detected in the genomes of S. <strong>aureus</strong> or S. epidermidis (Stoll et al., 2005),<br />

which suggests that the lipoproteins of these organisms are not N-acylated.<br />

29


Chapter 2<br />

<strong>The</strong> S. <strong>aureus</strong> Lgt is a protein of 279 amino acids that contains a highly conserved HGGLIG<br />

motif (residues 97 to 102). Although the His residue in this motif was shown to be essential<br />

for catalytic activity of the E. coli Lgt (Sankaran et al., 1997), it is not strictly conserved in all<br />

known Lgt proteins. On the other hand, the strictly conserved Gly at position 103 of E. coli<br />

Lgt, which is equivalent to Gly98 of S. <strong>aureus</strong> Lgt, is required for activity of this protein.<br />

Stoll et al. (Stoll et al., 2005) showed that a S. <strong>aureus</strong> lgt mutation has no effect on growth in<br />

broth as was also observed for B. subtilis (Leskelä et al., 1999). Nevertheless, the absence of<br />

Lgt has a considerable effect on the induction of an inflammatory response. Importantly, lipid<br />

modification serves to retain exported proteins at the membrane-cell wall interface. This is<br />

particularly relevant for Gram-positive bacteria, which lack an outer membrane that<br />

represents a retention barrier for exported proteins. In the absence of Lgt, B. subtilis cells<br />

release a variety of lipoproteins into the extracellular milieu, both in the form of unmodified<br />

precursor proteins and alternatively processed mature proteins that lack the N-terminal Cys<br />

residue (Antelmann et al., 2001). Similarly, the S. <strong>aureus</strong> lgt mutation resulted in the<br />

shedding of certain abundant lipoproteins, such as OppA, PrsA and SitC, into the broth. <strong>The</strong>se<br />

lipoproteins are normally retained in the membrane or cell wall of S. <strong>aureus</strong>.<br />

Type II Signal Peptidase<br />

As described above, lipoprotein signal peptides of prolipoproteins are cleaved by type II<br />

SPases after the Cys residue in the lipobox is modified by Lgt. Although B. subtilis and many<br />

other bacteria contain only one copy of the lspA gene, some contain a second copy, such as S.<br />

epidermidis, Bacillus licheniformis and Listeria monocytogenes. LspA is a membrane protein<br />

that spans the membrane four times and both the N- and C-termini are facing the cytoplasmic<br />

side of the membrane (Tjalsma et al., 1997; van Roosmalen et al., 2001). Six amino acid<br />

residues are important for SPase II activity, of which two Asp residues form the active site<br />

(Tjalsma et al., 1997). While processing of lipoproteins by LspA is essential for growth and<br />

viability for E. coli and other Gram-negative bacteria (Wu, 1996), it is not essential for B.<br />

subtilis (Tjalsma et al., 1999) and other Gram-positive bacteria, such as Lactococcus lactis<br />

(Venema et al., 2003). This suggests that processing of prolipoproteins is not essential for<br />

their functionality. <strong>The</strong> latter view is supported by the fact that PrsA, a lipoprotein required<br />

for correct folding of translocated proteins, is essential for viability of B. subtilis (Kontinen<br />

and Sarvas, 1993). In the absence of LspA, some of the lipoproteins of B. subtilis are<br />

processed in an alternative way by yet unidentified proteases and activity of unprocessed<br />

lipoproteins in lspA mutants is reduced. Also, in B. subtilis the secretion of the nonlipoprotein<br />

AmyQ was severely reduced (Tjalsma et al., 1999). This reduction might be the<br />

consequence of a malfunction of non-modified PrsA in AmyQ folding. Although most lspA<br />

mutants have been studied in Gram-negative bacteria and a few non-pathogenic Grampositive<br />

bacteria (Tjalsma et al., 1999; Venema et al., 2003), Sander et al. (Sander et al.,<br />

2004) showed a severe attenuated phenotype of lspA mutants of the pathogen Mycobacterium<br />

tuberculosis, which implies an important role for lipoprotein-processing by LspA during<br />

infection of M. tuberculosis. In S. <strong>aureus</strong> both the lspA and lgt genes are present in single<br />

copy in the genomes of all six sequenced strains. Interestingly, one of the two LspA<br />

homologues in S. epidermidis (125 amino acids) is considerably shorter than other known<br />

LspA proteins, including its large paralogue (177 amino acids). This is mainly the result of an<br />

additional N-terminal transmembrane domain in the large LspA proteins. As a result the short<br />

S. epidermidis LspA protein is predicted to have three membrane spanning domains, with the<br />

30


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

N-terminus located on the outside of the cell, the C-terminus on the inside of the cell and the<br />

(putative) active site Asp residues located on the outer surface of the cytoplasmic membrane.<br />

Signal peptide peptidase<br />

After translocation and processing of the preproteins by signal peptidases, the signal peptides<br />

are rapidly degraded by signal peptide peptidases (SPPases). In B. subtilis two SPPases, TepA<br />

and SppA, are known to be involved in translocation and processing of preproteins (Bolhuis<br />

et al., 1999). While TepA is required for translocation and processing of preproteins, SppA is<br />

only required for efficient processing of preproteins. Remarkably, no homologues of SppA or<br />

TepA were detectable by BLAST searches in the sequenced genomes of S. <strong>aureus</strong> and S.<br />

epidermidis. As reported by Meima and van Dijl (Meima and van Dijl, 2003), L. lactis<br />

contains a protein that shows limited similarity to TepA of B. subtilis and ClpP of C. elegans,<br />

suggesting that this protein might be an SPPase-analog of L. lactis. In S. <strong>aureus</strong> and S.<br />

epidermidis this protein homologue also seems to be present and is predicted to be a<br />

cytoplasmic membrane protein (our unpublished observations).<br />

Folding catalysts (PrsA and BdbD)<br />

Proteins that are transported across the membrane in a Sec-dependent manner emerge at the<br />

extracytoplasmic membrane surface in an unfolded state. <strong>The</strong>se proteins need to be rapidly<br />

and correctly folded into their native and protease-resistant conformation, before they are<br />

degraded by proteases in the cell wall or extracellular environment (Sarvas et al., 2004). An<br />

important folding catalyst in B. subtilis is PrsA, which shows homology to peptidyl-prolyl<br />

cis/trans-isomerases. PrsA is a lipoprotein (see also section “Lipoproteins”) that is essential<br />

for efficient protein secretion and cell viability in B. subtilis (Sarvas et al., 2004; Kontinen<br />

and Sarvas, 1993). Studies on the effects of PrsA depletion showed that the relative amounts<br />

of extracellular proteins from PrsA-depleted cells, were significantly reduced (Vitikainen et<br />

al., 2004). <strong>The</strong> solution structure of PrsA has been solved (Heikkinen et al., 2009), but no<br />

data has been published on S. <strong>aureus</strong> mutants lacking PrsA and it will be interesting to<br />

investigate whether PrsA is also essential for viability and virulence of this organism. It has<br />

already been shown that S. <strong>aureus</strong> lacking Lgt, releases an increased amount of PrsA into the<br />

extracellular milieu (Stoll et al., 2005), which might indicate that (most) pre-PrsA is not fully<br />

functional, but sufficient for viability. <strong>The</strong> observation by Stoll et al. (Stoll et al., 2005) also<br />

shows that, like in B. subtilis (Antelmann et al., 2001), the unmodified pre-PrsA is not<br />

effectively retained in the cytoplasmic membrane.<br />

Other proteins that are involved in proper folding of extracellular proteins in B. subtilis are the<br />

membrane proteins BdbC and BdbD, which are involved in the formation of disulfide bonds.<br />

Both proteins have been shown to be necessary for the stabilization of the membrane- and cell<br />

wall-associated pseudopilin ComGC (Meima et al., 2002). This protein, which is required for<br />

DNA binding and uptake during natural competence, contains an intramolecular disulfide<br />

bond (Chung et al., 1998). Both BdbC and BdbD are also important for the folding of<br />

heterologously produced E. coli PhoA, which contains two disulfide bonds, into an active and<br />

protease-resistant conformation (Bolhuis et al., 1999; Meima et al., 2002). Though a<br />

homologue of BdbD (named DsbA) is present in S. <strong>aureus</strong>, there is no homologue of BdbC in<br />

this organism. <strong>The</strong> same appears to be true for S. epidermidis. Nevertheless, measurements of<br />

the redox potential of purified DsbA indicate that this protein can act as an oxidase, and this<br />

31


Chapter 2<br />

view is confirmed by complementation studies in a dsbA mutant strain of E. coli (Dumoulin et<br />

al., 2005). <strong>The</strong> absence of a BdbC homologue from staphylococci is remarkable, since B.<br />

subtilis BdbC and BdbD are jointly required in the folding of ComGC and E. coli PhoA.<br />

Notably, all sequenced S. <strong>aureus</strong> genomes encode homologues of ComGC, including the Cys<br />

residues that form the disulfide bond in B. subtilis ComGC. This raises the question whether<br />

ComGC of S. <strong>aureus</strong> does indeed contain a disulfide bond and, if so, which protein(s) are<br />

involved in the formation of this disulfide bond. DsbA would be a candidate for this task<br />

since it has been shown that this S. <strong>aureus</strong> protein can functionally replace BdbB, BdbC and<br />

BdbD in the production of ComGC, E. coli PhoA and the S-S bond-containing sublancin 168<br />

in B. subtilis (Kouwen et al., 2007). This idea is further supported by the findings of Heras et<br />

al. (Heras et al., 2008) that the oxidized and reduced states of DsbA are energetically<br />

equivalent, which suggests that this facilitates the reoxidation of DsbA, likely by extracellular<br />

oxidants. Notably, S. <strong>aureus</strong> DsbA was shown to be a lipoprotein that does not seem to<br />

contribute to the virulence of this organism as tested in mouse and Caenorhabditis elegans<br />

models (Dumoulin et al., 2005). Furthermore, DsbA was shown to be dispensable for βhemolysin<br />

activity, despite the fact that this protein contains a disulfide bond, which is<br />

required for activity (Dziewanowska et al., 1996). <strong>The</strong>refore, the biological function of DsbA<br />

in staphylococci remains to be elucidated.<br />

Twin-arginine translocation (Tat) pathway<br />

<strong>The</strong> Tat-pathway exists in many bacteria, archaea, and chloroplasts. This pathway has been<br />

named after the consensus double (twin) Arg residues that are present in the signal peptide.<br />

<strong>The</strong> twin Arg residues are part of a motif that directs proteins specifically into the Tat<br />

pathway. In contrast to the Sec-machinery where only unfolded proteins are translocated<br />

across the membrane, the Tat-machinery is capable of translocating folded proteins. In Gramnegative<br />

bacteria, streptomycetes, mycobacteria and chloroplasts, an active Tat-pathway<br />

seems to require three core components, named TatA, TatB and TatC (Berks et al., 2005;<br />

Dilks et al., 2003; Mori and Cline, 2001; Robinson and Bolhuis, 2001; Yen et al., 2002). In<br />

all Gram-positive bacteria except streptomycetes and Mycobacterium smegmatis, the Tat<br />

pathway involves only TatA and TatC (Dilks et al., 2003; Yen et al., 2002). Recent studies in<br />

E. coli and chloroplasts have resulted in a model that proposes a key role for TatB-TatC<br />

complexes in signal peptide reception and TatA-TatB-TatC complexes in preprotein<br />

translocation (Cline and Mori, 2001; Alami et al., 2003). Interestingly, certain mutations in E.<br />

coli TatA have been shown to allow this protein to compensate for the absence of TatB<br />

(Blaudeck et al., 2005). This demonstrated that TatA is intrinsicaIly bifunctional, which is<br />

consistent with the fact that most Gram-positive bacteria lack TatB, but have TatA (Jongbloed<br />

et al., 2005). In B. subtilis, two minimal TatA-TatC translocases with distinct specificities are<br />

active (Jongbloed et al., 2004). While the constitutively expressed TatAy-TatCy translocase<br />

of B. subtilis is required for secretion of the protein with unknown function YwbN, the<br />

TatAd-TatCd translocase seems to be expressed only under conditions of phosphate starvation<br />

for secretion of the phosphodiesterase PhoD (Tjalsma et al., 2000; van Roosmalen et al.,<br />

2001). Most other Gram-positive bacteria that have tatA and tatC genes, including S. <strong>aureus</strong>,<br />

appear to have only one TatA-TatC translocase. <strong>The</strong> functionality of the S. <strong>aureus</strong> Tat<br />

translocase was recently demonstrated (Biswas et al., 2009). In contrast to S. <strong>aureus</strong>, S.<br />

epidermidis seems to lack a Tat pathway.<br />

32


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

Pseudopilin export (Com) pathway<br />

In B. subtilis four proteins, ComGC, ComGD, ComGE and ComGG, have been identified<br />

with an N-terminal pseudopilin-like signal peptide (Tjalsma et al., 2004; Tjalsma et al.,<br />

2000). All four of these proteins are involved in DNA binding and uptake, and are localized in<br />

the membrane and cell wall. It is thought that these proteins form a pilus-like structure in the<br />

cell wall or modify the cell wall to provide a passage for DNA uptake. Translocation to the<br />

extracytoplasmic membrane surface is only possible when these proteins are processed by the<br />

pseudopilin-specific SPase ComC in B. subtilis (Dubnau, 1999). SPases of this type are<br />

bifunctional and do not only catalyze signal peptide cleavage, but also methylation of the Nterminus<br />

of the mature protein (Strom et al., 1993). Furthermore, export and functionality of<br />

the four ComG proteins depends on the integral membrane protein ComGB and the traffic<br />

ATPase ComGA, which is located at the cytoplasmic side of the membrane (Chung and<br />

Dubnau, 1998; Hahn et al., 2005). Homologues of ComC, ComGA, ComGB and ComGC, but<br />

not ComGD, ComGE and ComGG, are present in the six sequenced S. <strong>aureus</strong> strains. This<br />

suggests that the Com system of S. <strong>aureus</strong> is not involved in DNA uptake, but in another<br />

solute transport process.<br />

ABC transporters<br />

Bacteriocins are peptides or proteins that inhibit the growth of other bacteria. Most of the<br />

characterized bacteriocins can be divided into several classes, depending on specific<br />

posttranslational modifications, the presence and processing of particular leader peptides and<br />

the machinery for export from the cytoplasm. A well described class of bacteriocins is formed<br />

by the lantibiotics. Members of this class are composed of short peptides that contain posttranslationally<br />

modified amino acids, like lanthionine and β-methyllanthionine (McAuliffe et<br />

al., 2001). <strong>The</strong> production of bacteriocins in S. <strong>aureus</strong> has been described for various strains.<br />

S. <strong>aureus</strong> C55 produces the two lantibiotics C55α and C55β (Navaratna et al., 1998). <strong>The</strong>se<br />

lantibiotics are both encoded by a 32 kb plasmid, which is readily lost upon growth at<br />

elevated temperatures. C55α and C55β showed antimicrobial activity towards other S. <strong>aureus</strong><br />

strains and Micrococcus luteus, but not towards S. epidermidis. Furthermore, the nonlantibiotics<br />

BacR1 (Crupper et al., 1997), aureocin A53 (Netz et al., 2001) and aureocin A70<br />

(Netz et al., 2002a; Netz et al., 2002b) have been identified as bacteriocins with activity<br />

against a broad range of bacteria. <strong>The</strong> genes for both aureocins are located on a plasmid that is<br />

present in S. <strong>aureus</strong> strains that were isolated from milk. By analogy with well described<br />

bacteriocin export machinery in other organisms (Håvarstein et al., 1995; Peschel et al.,<br />

1997), it can be anticipated that all of the afore-mentioned bacteriocins are exported to the<br />

external staphylococcal milieu by dedicated ABC transporters. However, no experimental<br />

evidence for this assumption has been published for S. <strong>aureus</strong>. Notably, it has been<br />

demonstrated that the secretion of the lantibiotics epidermin and gallidermin of S. epidermidis<br />

Tü3298 and <strong>Staphylococcus</strong> gallinarum, respectively, is facilitated by so-called onecomponent<br />

ABC transporters. Specifically, the ABC-transporter GdmT has been implicated in<br />

the transport of these lantibiotics (Peschel et al., 1997).<br />

33


Chapter 2<br />

Holins<br />

Holins are dedicated export systems for peptidoglycan-degrading endolysins that have been<br />

implicated in the programmed cell death of bacteria. <strong>The</strong>se exporters, which are composed of<br />

homo-oligomeric complexes, can be subdivided into two classes, depending on their number<br />

of transmembrane segments. While class I holin subunits have three transmembrane<br />

segments, class II holin subunits have two transmembrane segments (Young and Bläsi, 1995).<br />

In S. <strong>aureus</strong> the lrg and cid operons are involved in murein hydrolase activity and antibiotic<br />

tolerance (Groicher et al., 2000; Rice et al., 2003). A disrupted lrg operon leads to an increase<br />

in murein hydrolase activity and a decrease in penicillin tolerance, and a disrupted cid operon<br />

leads to a decrease in murein hydrolase activity and an increase in penicillin tolerance. It is<br />

still unclear how the CidA and LrgA proteins are involved in these mechanisms, but these<br />

proteins display significant similarity to the bacteriaphage holin protein family, suggesting<br />

that they have a role in protein export. It has therefore been proposed that the CidA and LrgA<br />

proteins act on the murein hydrolase activity and antibiotic tolerance analogous to holins and<br />

antiholins, respectively (Bayles, 2000; Rice et al., 2003). Sequence similarity searches show<br />

that the genes for LrgA and CidA are conserved in the six sequenced S. <strong>aureus</strong> strains, as well<br />

as S. epidermidis and B. subtilis. Notably, none of the three holins of B. subtilis was shown to<br />

be involved in the secretion of proteins to the extracellular milieu (Westers et al., 2003;<br />

Tjalsma et al., 2004).<br />

Ess pathway<br />

<strong>The</strong> ESX-1 or ESAT-6 secretion system (Ess) pathway has first been described for M.<br />

tuberculosis. It has been proposed that at least two virulence factors, ESAT-6 (early secreted<br />

antigen target 6 kDa) and CFP-10 (culture filtrate protein 10 kDa), are secreted via this<br />

pathway in a Sec-independent manner (Berthet et al., 1998; Sørensen et al., 1995). As this<br />

pathway was discovered in mycobacteria, it is also known as the Snm pathway (Secretion in<br />

mycobacteria; (Converse and Cox, 2005)). <strong>The</strong> genes for ESAT-6 en CPF-10 are located in<br />

conserved gene clusters, which also encode proteins with domains that are conserved in FtsK-<br />

and SpoIIIE-like transporters. <strong>The</strong>se conserved FtsK/SpoIIIE domains have therefore been<br />

termed FSDs (Burts et al., 2005). In other Gram-positive bacteria including S. <strong>aureus</strong>, B.<br />

subtilis, Bacillus anthracis, Clostridium acetobutylicum and L. monocytogenes, homologues<br />

of ESAT-6 have been identified (Pallen, 2002). <strong>The</strong> genes for these ESAT-6 homologues are<br />

also found in gene clusters that contain at least one membrane protein with a FSD. In S.<br />

<strong>aureus</strong>, two proteins named EsxA and EsxB have been identified that seem to be secreted via<br />

the Ess pathway (Burts et al., 2005). <strong>The</strong> esxA and esxB genes are part of a cluster containing<br />

six other genes for proteins that have been implicated in the translocation of EsxA and EsxB.<br />

<strong>The</strong>se include the cytoplasmic protein EsaB and the secreted protein EsaC (Burts et al.,<br />

2008), as well as the predicted membrane proteins EsaA, EssA, EssB and EssC, of which<br />

EssC contains a FSD. Mutations in essA, essB or essC result in a loss of EsxA and EsxB<br />

production, which may relate to an inhibition of the synthesis of these proteins, or their<br />

folding into a protease-resistant conformation. EsaB is a negative regulator of EsaC and<br />

represses the production of EsaC in a post-transcriptional manner. EsaC, although secreted,<br />

does not contain the WxG-motif or any other signal peptide and it is still unclear how this<br />

protein is recognized by the Ess secretion pathway. All sequenced S. <strong>aureus</strong> strains contain<br />

this cluster of esa, ess and esx genes, but it seems to be absent from S. epidermidis.<br />

34


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

Interestingly, the genes for EsxB and EsaC appear to be absent from the S. <strong>aureus</strong> MRSA252<br />

strain. This implies that the Ess machinery of this strain may be required for the transport of<br />

only EsxA and perhaps a few other unidentified proteins. If so, EsaC would be dispensable<br />

for an active ESAT-6 pathway and might be specifically involved in the export of EsxB. This<br />

view is also suggested from the data published by Burts et al. (Burts et al., 2008), which show<br />

that a S. <strong>aureus</strong> Newman strain lacking esxB does not produce EsaC. Alternatively, the Ess<br />

pathway could be inactive in the S. <strong>aureus</strong> MRSA252 strain due to the absence of EsaC.<br />

Lysis<br />

Various studies have shown that certain proteins with typical cytoplasmic functions and<br />

without known signals for protein secretion can nevertheless be detected on the extracellular<br />

proteome of different bacteria (Tjalsma et al., 2004). Notably, many of these proteins, such as<br />

catalase, elongation factor G, enolase, glyceraldehyde-3-phosphate dehydrogenase, GroEL<br />

and superoxide dismutase, are amongst the most highly abundant cytoplasmic proteins. This<br />

makes it likely that they are detectable in the extracellular proteome due to cell lysis. Perhaps,<br />

such proteins are more resistant to extracytoplasmic degradation than other proteins that are<br />

simultaneously released by lysis. However, the possibility that the extracellular localization of<br />

typical cytoplasmic proteins is due to the activity of, as yet unidentified, export pathways<br />

cannot be excluded. Clearly, until recently this possibility did still apply for the EsxA, EsxB<br />

and EsaC proteins, which are now known to be exported via the Ess pathway. A clear<br />

indication that the presence of certain “cytoplasmic” proteins in the extracytoplasmic milieu<br />

of bacteria may relate to specific export processes was provided by Boël and co-workers<br />

(Boël et al., 2004), who showed that 2-phosphoglycerate-dependent automodification of<br />

enolase is necessary for its export from the cytoplasm.<br />

Properties of staphylococcal signal peptides and cell<br />

retention signals<br />

Signal peptides<br />

All proteins that have to be transported from the cytoplasm across the membrane to the<br />

extracytoplasmic compartments of the cell, or the extracellular milieu, need to contain a<br />

specific sorting signal for their distinction from resident proteins of the cytoplasm. <strong>The</strong> known<br />

bacterial sorting signals for protein export from the cytoplasm are signal peptides (von Heijne,<br />

1990). <strong>The</strong>se signal peptides can be classified by the transport and modification pathway into<br />

which they direct proteins. Presently, four different bacterial signal peptides are recognized<br />

that share a common architecture, but differ in details (Figure 5). Two of these direct proteins<br />

into the widely used Sec pathway, including the secretory (Sec type) signal peptides and the<br />

lipoprotein signal peptides. Proteins with Sec type or lipoprotein signal peptides are processed<br />

by different SPases (type I or type II SPases, respectively), and are targeted to different<br />

destinations. In S. <strong>aureus</strong> the proteins with Sec type signal peptides are processed by the type<br />

I SPase SpsB and targeted to the cell wall or extracellular milieu.<br />

35


Chapter 2<br />

Figure 5. General properties and classification of S. <strong>aureus</strong> signal peptides. Signal peptide properties are based<br />

on SPase cleavage sites and the export pathways via which the preproteins are exported. Predicted signal peptides<br />

(144) were divided into five distinct classes: secretory (Sec-type) signal peptides, twin-arginine (RR/KR) signal<br />

peptides, lipoprotein signal peptides, pseudopilin-like signal peptides, and bacteriocin leader peptides. Most of<br />

these signal peptides have a tripartite structure: a positively charged N-domain (N), containing lysine and/or<br />

arginine residues (indicated by +), a hydrophobic H-domain (H, indicated by a black box), and a C-domain (C)<br />

that specifies the cleavage site for a specific SPase. Where appropriate, the most frequently occurring amino acid<br />

residues at particular positions in the signal peptide or mature protein are indicated. Also, the numbers of signal<br />

peptides identified for each class and the respective SPase are indicated.<br />

<strong>The</strong> proteins with a lipoprotein signal peptide are lipid-modified by Lgt, prior to processing<br />

by the type II SPase LspA. In principle, these lipoproteins are retained at the membrane-cell<br />

wall interface, but they can be liberated from this compartment by proteolytic removal of the<br />

N-terminal Cys that contains the diacylglyceryl moiety (Antelmann et al., 2001). Proteins<br />

with twin-arginine (RR) signal peptides appear to be processed by type I SPases, at least in B.<br />

subtilis, and targeted to the cell wall or extracellular milieu (Tjalsma et al., 2004). <strong>The</strong><br />

proteins with a pseudopilin signal peptide are processed by the pseudopilin signal peptidase<br />

ComC and most likely localized in the cytoplasmic membrane and cell wall. Finally,<br />

bacteriocins contain a completely different sorting and modification signal that is usually<br />

called the leader peptide. <strong>The</strong> known leader peptides show no resemblance to the aforementioned<br />

signal peptides. <strong>The</strong> export of bacteriocins via ABC-transporters results in their<br />

secretion into the extracellular milieu (Michiels et al., 2001; Schnell et al., 1988).<br />

Sec type, lipoprotein and RR-signal peptides contain three distinguishable domains: the N-,<br />

H- and C-domains. <strong>The</strong> N-terminal domain contains positively charged amino acids, which<br />

are thought to interact with the secretion machinery and/or with negatively charged<br />

phospholipids in the membrane. <strong>The</strong> H-domain is formed by a stretch of hydrophobic amino<br />

acids which facilitate membrane insertion. Helix-breaking residues in the middle of the Hdomain<br />

may facilitate H-domain looping during membrane insertion and translocation of the<br />

precursor protein. <strong>The</strong> subsequent unlooping of the H-domain would display the SPase<br />

recognition and cleavage site at the extracytoplasmic membrane surface where the catalytic<br />

domains of type I and type II SPases are localized (van Roosmalen et al., 2004). Helixbreaking<br />

residues just before the SPase recognition and cleavage site would facilitate<br />

precursor processing by SPase I or II. In fact, these helix-breaking residues and the SPase<br />

cleavage site, respectively, define the beginning and the end of the C-domain. Notably, the Cdomain<br />

of pseudopilin signal peptides is located between the N- and H-domains (Chung and<br />

Dubnau, 1995; Pugsley, 1993; Chung and Dubnau, 1998; Lory, 1998). Accordingly, processing<br />

36


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

by pseudopilin-specific SPases, like ComC, takes place at the cytoplasmic side of the<br />

membrane and leaves the H-domain attached to the translocated protein.<br />

While many proteins that end up in the extracellular milieu or the cell wall of Gram-positive<br />

bacteria have signal peptides, proteins without known export signals can also be found on<br />

these locations. <strong>The</strong> relative numbers of proteins without known signal peptides seem to vary<br />

per organism. While these numbers are relatively low for B. subtilis and S. <strong>aureus</strong>, they are<br />

high for group A Streptococcus and M. tuberculosis (Tjalsma et al., 2004). As indicated<br />

above, some of the proteins without known export signals appear to be liberated from the cell<br />

by lysis, while others are actively exported, for example via the Ess pathway. Although the<br />

precise export signal in proteins secreted via the Ess pathway has not yet been defined, a<br />

WxG motif is shared by many of these proteins and may serve a function in protein targeting<br />

(Pallen, 2002). Furthermore, the signal for specific release of lysins via holins is presently not<br />

known.<br />

Signal peptide predictions<br />

Several prediction programs that are accessible through the world-wide web are useful tools<br />

to predict whether a given protein contains some sort of sorting signal or SPase cleavage site.<br />

<strong>The</strong> programs that we have used in this and other studies were: SignalP-NN and SignalP-<br />

HMM version 2.0 (Nielsen et al., 1997), LipoP version 1.0 (Juncker et al., 2003), PrediSi<br />

(Hiller et al., 2004) and Phobius (Kall et al., 2004). <strong>The</strong>se programs have been designed to<br />

identify Sec type signal peptides, N-terminal membrane anchors (Phobius), or lipoprotein<br />

signal peptides in Gram-negative bacteria (LipoP). <strong>The</strong> TMHMM-program version 2.0<br />

(Cserzö et al., 1997) was used to exclude proteins with (predicted) multiple membrane<br />

spanning domains. Predictions for proteins containing a signal peptide were performed with<br />

the SignalP program, using the Neural Network (NN) and Hidden Markov Model algorithms<br />

(HMM). Version 2.0 of the SignalP program was preferred above Version 3.0 (Bendtsen et<br />

al., 2004) for our signal peptide predictions in S. <strong>aureus</strong> and S. epidermidis, because the best<br />

overall prediction accuracy was obtained with Version 2.0 in a recent proteomics-based<br />

verification of predicted export and retention signals in B. subtilis (Tjalsma and van Dijl,<br />

2005). Specifically, the Hidden Markov Model in SignalP 2.0 assigns probability scores to<br />

each amino acid of a potential signal peptide and indicates whether it is likely to belong to the<br />

N-, H-, or C-domains. Proteins with no detectable N-, H-, and C-domain were excluded from<br />

the set. Searching for transmembrane domains was performed with the TMHMM program<br />

and proteins with more than one (predicted) transmembrane domain were excluded from the<br />

set, because they are most likely integral membrane proteins. All proteins with a predicted Cterminal<br />

transmembrane segment in addition to a signal peptide were screened for the<br />

presence of a conserved motif for covalent cell wall binding. It should be noted that this<br />

approach does not automatically result in the exclusion of potential membrane proteins with<br />

one N-terminal transmembrane domain. <strong>The</strong> LipoP-program was used to predict lipoproteins.<br />

<strong>The</strong> combined results of all these programs resulted in a list of proteins, which have: a) signal<br />

peptides with distinctive N-, H-, and C-domains, b) no additional transmembrane domains,<br />

and c) predicted extracytoplasmic localizations. <strong>The</strong>se proteins were scanned for the presence<br />

of proteomics-based consensus motifs for type I, type II or pseudopilin-specific SPase<br />

recognition and cleavage sites, twin-arginine motifs, and known leader peptides of<br />

bacteriocins by BLAST searches and by use of the PAT<strong>TI</strong>NPROT program (http://npsapbil.ibcp.fr)<br />

as previously described (Tjalsma and van Dijl, 2005). To define the core<br />

37


Chapter 2<br />

exoproteome and variant exoproteome of the S. <strong>aureus</strong> strains, the sets of proteins with<br />

predicted signal peptides were used in multiple blasts with the freeware BLASTall from the<br />

NCBI. <strong>The</strong> output was then filtered using Genome2D (Baerends et al., 2004).<br />

Secretory (Sec type) signal peptides<br />

Proteomics-based data sets of membrane, cell wall and extracellular proteins have been<br />

extremely valuable for a recent verification of signal peptide predictions in B. subtilis<br />

(Tjalsma and van Dijl, 2005). Such data sets are now becoming available for S. <strong>aureus</strong>, as<br />

exemplified by studies on the membrane plus cell wall proteomes of S. <strong>aureus</strong> Phillips<br />

(Nandakumar et al., 2005), and the extracellular proteomes of S. <strong>aureus</strong> strains that have been<br />

derived from the recently sequenced NCTC8325 and COL strains (Ziebandt et al., 2001;<br />

Ziebandt et al., 2004) (Figure 2). Additionally, the extracellular proteomes of several clinical<br />

S. <strong>aureus</strong> isolates have been analyzed (Figure 2). <strong>The</strong> membrane, cell wall and extracellular<br />

proteins of S. <strong>aureus</strong> that have been identified by proteomics (Ziebandt et al., 2001; Ziebandt<br />

et al., 2004; Nandakumar et al., 2005; Gatlin et al., 2006; Pocsfalvi et al., 2008; Ziebandt et<br />

al., submitted), involving 2D-PAGE and subsequent mass spectrometry, are listed in<br />

(Supplemental tables IIIa and IIIb). <strong>The</strong>se tables also show the -3 to +1 amino acid sequences<br />

of the respective signal peptidase cleavage sites, if present.<br />

Based on the proteomics data for membrane and extracellular proteins of B. subtilis, the<br />

optimized -3 to +1 pattern [AVS<strong>TI</strong>] - [SEKYHQFLDGPW] – A - [AQVEKDFHLNS] for<br />

signal peptide recognition and cleavage by type I SPases of this organism was identified<br />

(Tjalsma and van Dijl, 2005). SPase cleavage occurs C-terminally of the invariant Ala residue<br />

at the -1 position. <strong>The</strong> residues between square brackets in the pattern are listed in the order of<br />

their frequency, the most frequently identified residue at each position being placed in first<br />

position. By comparing the predicted SPase recognition and cleavage sites in signal peptides<br />

of proteomically identified extracellular proteins of S. <strong>aureus</strong> (Supplemental table IIIa) we<br />

defined the -3 to +1 pattern [AVST] - [KQNESDHYLFAGR] – A - [AESKDIFLQTY] for<br />

productive recognition and cleavage by the type I SPase SpsB. Compared to the equivalent<br />

pattern of B. subtilis it is interesting to note that the frequencies of certain residues at the -3, -<br />

2 and +1 positions differ, as reflected by the most-frequent-first order in which they are listed<br />

in the pattern. Moreover, Asn can be present at the -2 position, while Ile is accepted at the +1<br />

position. <strong>The</strong> latter residues are found in the -2 and +1 positions of certain serine proteases,<br />

hemolysins, immunoglobulin G binding protein A and aureolysin (Supplemental table IIIa). It<br />

should also be noted that, compared to the optimized SPase recognition pattern of B. subtilis,<br />

several residues are not found at the -3, -2 and +1 positions of potential SpsB recognition and<br />

cleavage sites in identified extracellular proteins of S. <strong>aureus</strong>. As such residues may be<br />

present in SPase recognition and cleavage sites of proteins that have escaped identification<br />

through proteomics, we have included them in the -3 to +1 search pattern (printed in<br />

lowercase) for the identification of potential secretory proteins of staphylococci: [AVSit] -<br />

[KHNDQSYEGLRAfpw] – A - [AESDIKLTYfhnqv]. This optimized S. <strong>aureus</strong> search<br />

pattern was used as an indicator for the quality of signal peptide predictions that were based<br />

on the SignalP-NN, SignalP-HMM, LipoP, PrediSi, Phobius and TMHMM programs.<br />

Proteins with potential signal peptides containing this pattern were assigned to have a high<br />

probability for an extracytoplasmic localization and a low probability for membrane retention<br />

(Supplemental tables IIIc-f). Proteins with potential signal peptides that do not contain this<br />

pattern were assigned to have a high probability to be retained in the membrane (data not<br />

38


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

shown). In this case, the uncleaved signal peptide could serve as an N-terminal membrane<br />

anchor. Following this approach, sets of 186-211 proteins (depending on the S. <strong>aureus</strong> strain)<br />

were identified that contain a potential signal peptide or N-terminal transmembrane segment.<br />

Scanning for the presence of the SpsB recognition and cleavage motif [AVSit] -<br />

[KHNDQSYEGLRAfpw] – A - [AESDIKLTYfhnqv] revealed that, depending on the S.<br />

<strong>aureus</strong> strain investigated, 86-106 proteins carry this motif. <strong>The</strong>se proteins are most likely<br />

processed by SPase, liberated from the membrane and secreted into the extracellular milieu,<br />

unless they contain a cell wall retention signal (see following sections). Most of the other<br />

proteins with signal peptides that do not conform to the SpsB recognition and cleavage motif<br />

lack the invariant Ala at the –1 position. Also, some of these preproteins contain different<br />

residues at the –3, -2 or +1 positions. For example, Asp, Glu, Phe and Lys are highly unlikely<br />

residues at the -3 position (van Roosmalen et al., 2004). On the other hand, some preproteins<br />

have a Gly at the -3 position (e.g. the exotoxins 4 and 5 from S. <strong>aureus</strong> COL) or a Leu<br />

(Supplemental tables IIIc and IIId). Since Gly and Leu residues at the -3 position of signal<br />

peptides are accepted by the E. coli SPase, it seems likely that they are also accepted at this<br />

position by SpsB. However, we have not included these residues in the current SpsB<br />

recognition and cleavage motif, since we could neither identify these proteins amongst the<br />

secreted proteins of S. <strong>aureus</strong> COL (Figure 2), nor find published evidence that these proteins<br />

are indeed secreted. Among the proteins with predicted cleavable Sec type signal peptides<br />

there are many known extracellular staphylococcal virulence factors, such as exotoxins,<br />

enterotoxins (SEM, SEN and SEO), hemolysins, toxic shock syndrome toxin-1 (TSST-1),<br />

leukotoxins (LukD, LukE), a secretory antigen SsaA homologue and the immunodominant<br />

antigen A (IsaA). Remarkably, the lists of identified extracellular proteins of S. <strong>aureus</strong> COL<br />

and RN6390 (Ziebandt et al., 2004) (Supplemental tables IIIa and IIIb) reveal that about 41%<br />

of these proteins lack known signal peptides. This percentage is substantially higher than the<br />

initial estimate of 10%, which was based on a limited proteomics-derived data set (Tjalsma et<br />

al., 2004). It is also interesting to note that the list of identified extracellular proteins without<br />

a signal peptide includes enolase, which may be actively exported by an unknown mechanism<br />

(Boël et al., 2004), but lacks EsxA and EsxB, which are exported by the Ess pathway (Burts<br />

et al., 2005).<br />

Twin-arginine (RR-)signal peptides<br />

<strong>The</strong> consensus RR-motif that directs proteins into the Tat pathway has previously been<br />

defined as [KR]-R-x-#-#, where # is a hydrophobic residue (Cristóbal et al., 1999; Jongbloed<br />

et al., 2000). Dilks et al. (Dilks et al., 2003) have used a genomic approach to identify<br />

possible Tat substrates for 84 diverse prokaryotes using the TATFIND 1.2 program. This<br />

study included S. <strong>aureus</strong> Mu50, MW2 and N315. Two potential Tat substrates of unknown<br />

function were predicted for S. <strong>aureus</strong> Mu50 and MW2 and one of these was also predicted for<br />

S. <strong>aureus</strong> N315 (Dilks et al., 2003). However, both proteins are conserved in all sequenced S.<br />

<strong>aureus</strong> strains, including the N315 strain. One of these two predicted Tat substrates has no<br />

known function, whereas the other was annotated as a hypothetical protein similar to a<br />

ferrichrome ABC transporter (permease). <strong>The</strong>se proteins, however, are not in our list of<br />

proteins that have a predicted (RR-)signal peptide. Although they have signal peptides<br />

according to the SignalP program, these proteins are localized in the cytoplasm or membrane,<br />

respectively, according to the LipoP, PrediSi and Phobius programs. It is therefore unlikely<br />

39


Chapter 2<br />

that these proteins are destined for secretion. Specifically, the hypothetical permease has eight<br />

predicted transmembrane helices.<br />

Our own pattern searches for proteins with a possible RR-motif resulted in 24-32 positive<br />

hits, depending on the S. <strong>aureus</strong> strain investigated. However, most of these proteins have no<br />

detectable N-, H- or C-domains and were, therefore, discarded from our data set. Also, some<br />

other proteins with a possible RR-motif are predicted to contain a lipoprotein signal peptide.<br />

<strong>The</strong>se predicted lipoproteins were also discarded from the list of potential S. <strong>aureus</strong> Tat<br />

substrates, firstly, because none of the identified lipoproteins of B. subtilis that have a RRmotif<br />

were shown to be secreted via the Tat pathway (Jongbloed et al., 2000; Jongbloed et al.,<br />

2002), and secondly, because there is limited published evidence for other bacteria that<br />

lipoproteins can be exported Tat-dependently (Widdick et al., 2006). Thus, it appears that<br />

only 5-7 proteins, depending on the S. <strong>aureus</strong> strain investigated, are potentially exported by<br />

the Tat pathway and cleaved by SpsB. However, it is noteworthy that none of the B. subtilis<br />

proteins with a KR-motif were so far shown to be secreted Tat-dependently, even though KRmotifs<br />

are capable of directing proteins into the Tat pathways of chloroplasts and Gramnegative<br />

bacteria, such as E. coli and Salmonella enterica (Stanley et al., 2000; Hinsley et al.,<br />

2001; Molik et al., 2001; Ignatova et al., 2002). If KR-motifs are also rejected by the S.<br />

<strong>aureus</strong> Tat pathway, there would not be a single protein in any sequenced S. <strong>aureus</strong> strain that<br />

is secreted Tat-dependently. This would be highly remarkable in view of the presence of tatA<br />

and tatC genes in all these strains. Notably, the only known strictly Tat-dependent<br />

extracellular proteins of B. subtilis are the phosphodiesterase PhoD (Tjalsma et al., 2000) and<br />

the protein of unknown function YwbN (Jongbloed et al., 2004). While a homologue of PhoD<br />

is not present in any of the six sequenced S. <strong>aureus</strong> strains, homologues of YwbN are present<br />

in all these strains. Close inspection of the YwbN homologues of S. <strong>aureus</strong> COL, JH1, JH9,<br />

MRSA252, MSSA476, NCTC8325, Newman, USA300 and USA300_TCH1516 revealed the<br />

presence of an N-terminal RR-motif, but a potential signal peptide was not identified as such<br />

by the SignalP program. In contrast, the YwbN homologues of S. <strong>aureus</strong> Mu3, Mu50, MW2<br />

and N315 appeared to lack this RR-motif. According to comparisons of the deduced amino<br />

acid sequences, the latter three YwbN homologues would miss the first 40 residues of B.<br />

subtilis YwbN. Most likely, this is not the case since the sequences upstream of the annotated<br />

S. <strong>aureus</strong> Mu3, Mu50, MW2 and N315 ywbN genes encode for a peptide with a RR-motif in<br />

the same open reading frame as the ywbN structural gene (Supplemental table IIIc). Thus, the<br />

RR-motif of S. <strong>aureus</strong> Mu3, Mu30, MW2 and N315 YwbN has escaped identification due to<br />

a systematic difference in sequence annotation. Recent studies by Biswas et al. (Biswas et al.,<br />

2009) have shown that the YwbN homologue of S. <strong>aureus</strong> is an iron-dependent peroxidase<br />

(now named FepB). Tat mutant S. <strong>aureus</strong> strains did no longer export active FepB, and the<br />

RR-signal peptide of FepB was able to direct Tat-dependent secretion of prolipase or protein<br />

A. Interestingly, FepB is needed for iron-acquisition in S. <strong>aureus</strong>, and in a mouse kidney<br />

abscess model the bacterial loads of tat or fepB mutant strains were substantially reduced.<br />

Pseudopilin signal peptides<br />

<strong>The</strong> signal peptides of pseudopilins differ from the Sec type signal peptides in the location of<br />

SPase cleavage sites. In pseudopilin signal peptides, this cleavage site is located between the<br />

N- and H-domains (Lory, 1994). <strong>The</strong> consensus recognition and cleavage motif for<br />

pseudopilin SPases, such as ComC, is K-G-F-x-x-x-E. Cleavage by pseudopilin SPases occurs<br />

within this motif, between the Gly and Phe residues. Upon cleavage the Phe residue is<br />

40


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

methylated. In all sequenced S. <strong>aureus</strong> strains, three proteins were found, which have the<br />

canonical pseudopilin SPase recognition and cleavage motif. <strong>The</strong>se proteins are homologues<br />

of the cold shock proteins CspB, CspC and CspD of B. subtilis. However, even though these<br />

proteins do contain the pseudopilin SPase recognition and cleavage pattern, they lack the Hdomain.<br />

Since the active site of pseudopilin SPases is located in the cytoplasm, cleavage of<br />

the CspBCD homologues of S. <strong>aureus</strong> would be possible, but their export via the Com<br />

pathway is unlikely. Nevertheless, it should be noted that one of the CspBCD homologues of<br />

S. <strong>aureus</strong>, which is known as CspA, was found in the extracellular proteome of a clinical<br />

isolate (Figure 2 and Supplemental table IIIb). To verify the absence or presence of<br />

pseudopilins in S. <strong>aureus</strong>, BLAST searches with the known ComGC, ComGD, ComGE or<br />

ComGG proteins of B. subtilis were performed. This revealed the presence of only one<br />

potential pseudopilin, which is a homologue of B. subtilis ComGC. Although the consensus<br />

pseudopilin SPase recognition and cleavage site is absent from S. <strong>aureus</strong> ComGC, a putative<br />

cleavage pattern (Q-A-F-T-L-I-E) is present at the position in the ComGC signal peptide<br />

where a pseudopilin SPase recognition and cleavage site would be expected. Further analyses<br />

revealed that similar observations can be made for ComGC homologues in other Grampositive<br />

bacteria, such as Bacillus cereus, B. anthracis, L. monocytogenes, S. haemolyticus<br />

and Oceanobacillus iheyensis. By comparing the ComGC homologues of these organisms, an<br />

expanded search pattern for Gram-positive bacterial pseudopilin SPase recognition and<br />

cleavage sites can be defined as [KEQRS]-[GA]-F-x-x-x-E. Interestingly, using this expanded<br />

search pattern, two additional potential pseudopilins of S. <strong>aureus</strong> were identified. <strong>The</strong>se<br />

potential pseudopilins show similarity to ComGD of B. cereus and B. anthracis, and ComGF<br />

of Bacillus halodurans and L. lactis. It remains to be shown whether the three identified<br />

potential pseudopilins of S. <strong>aureus</strong> are indeed able to assemble into pilin-like structures after<br />

processing by the ComC homologue. If so, it will be even more interesting to identify their<br />

biological function, for example in adhesion to surfaces, motility, or export of proteins. Such<br />

functions could play a role in virulence and have been attributed to type IV pili and<br />

pseudopilins of Gram-negative bacteria (Lory, 1998).<br />

Bacteriocin leader peptides<br />

Bacteriocins form a distinct group of proteins with cleavable N-terminal signal peptides,<br />

which are often called leader peptides. <strong>The</strong>se leader peptides only have N- and C-domains,<br />

and thus completely lack the hydrophobic H-domain. <strong>The</strong> bacteriocin leader peptides are<br />

invoked in posttranslational modification and prevention of premature antimicrobial activity,<br />

which would be deleterious to the producing organism. Of the sequenced S. <strong>aureus</strong><br />

bacteriocins, C55α and C55β contain a leader peptide (Navaratna et al., 1999), whereas<br />

leader peptides are absent from aureocin A53 (Netz et al., 2001) and aureocin A70 (Netz et<br />

al., 2001). Two potential lantibiotics with leader peptides were identified by sequencing the<br />

genomes of S. <strong>aureus</strong> MW2 (Baba et al., 2002) (GI numbers 49486642 and 49486641) and<br />

MSSA476. In both strains, the corresponding genes are located on the genomic island νSAβ.<br />

Both S. <strong>aureus</strong> proteins show similarity to the lantibiotic gallidermin precursor GdmA of<br />

<strong>Staphylococcus</strong> gallinarum, and to the lantibiotic epidermin precursor EpiA of S. epidermidis.<br />

Notably, the S. <strong>aureus</strong> COL strain contains only one of these two potential lantibiotics, which<br />

is most similar to the potential MW2 lantibiotic with the accession number 49486641. Two<br />

additional putative bacteriocins that were identified by genome sequencing seem to be<br />

homologous to L. lactis lactococcin 972. <strong>The</strong> hypothetical protein SAP019 (N315 annotation)<br />

41


Chapter 2<br />

is plasmid-encoded in S. <strong>aureus</strong> N315 and MSSA476, and chromosomally encoded in S.<br />

<strong>aureus</strong> MRSA252. <strong>The</strong> other hypothetical bacteriocin SAS029 is chromosomally encoded in<br />

all sequenced S. <strong>aureus</strong> strains. Recently, a program for detecting potential bacteriocins in<br />

bacterial genomes (BAGEL) has been released (de Jong et al., 2006). This program is based<br />

on the properties of known bacteriocins, including genes that lie close to these bacteriocin<br />

genes. Such genes may encode proteins involved in the processing or transport of the<br />

bacteriocin. Using the BAGEL program with standard settings to detect potential bacteriocins<br />

in the sequenced and annotated S. <strong>aureus</strong> strains, 2-6 proteins were identified as significant.<br />

No published data is presently available on the charateristics of these proteins, so it remains to<br />

be seen whether they are genuine bacteriocins.<br />

A potential Ess export signal?<br />

As described above, the EsxA, EsxB and EsaC proteins are secreted by S. <strong>aureus</strong> via the Ess<br />

route (Burts et al., 2005). All three proteins lack a known signal peptide, but are specifically<br />

transported across the membrane nonetheless. This implies that these two proteins must<br />

contain an export signal that is recognized by one or more Ess pathway components. <strong>The</strong><br />

nature of this signal is presently unknown. <strong>The</strong> most common feature of proteins that are<br />

known (or predicted) to be translocated across the membrane via the Ess pathway is a WxGmotif,<br />

which is located at ~100 amino acids from the N-terminus of the protein (Pallen, 2002).<br />

In particular since the WxG motif appears to be absent from EsaC, the involvement of the<br />

WxG-motif in Ess targeting remains to be demonstrated.<br />

Retention signals<br />

Lipoproteins<br />

Lipoproteins appear to be exported via the Sec-pathway. During, or shortly after<br />

translocation, the invariant Cys in the lipobox is diacylglyceryl-modified by Lgt. This results<br />

in signal peptide cleavage by SPase II and retention of the mature lipoprotein in the<br />

membrane. Based on the cleavage sites of lipoproteins that have been identified in various<br />

Gram-positive bacteria, Sutcliffe et al. (Sutcliffe and Harrington, 2002) have reported the -4<br />

to +2 lipobox pattern [LIVMFESTAG] - [LVIAMGT] - [IVMSTAFG] - [AG] – C -<br />

[SGANERQTL]. Furthermore, they reported that neither the charged residues Asp, Glu, Arg,<br />

or Lys, nor Gln are present in the region between six and twenty residues N-terminal of the<br />

lipobox. Searching the translated proteins encoded by the thirteen S. <strong>aureus</strong> genomes with the<br />

pattern shown above using the PAT<strong>TI</strong>NPROT program, revealed about 50 proteins with this<br />

motif (Supplemental table IIIe and IIIf). A comparison of the PAT<strong>TI</strong>NPROT results to the<br />

results obtained with the LipoP program shows that 12-18 more potential lipoproteins may be<br />

present in S. <strong>aureus</strong>. Most of these extra predicted lipoproteins contain an amino acid at the -<br />

1-position (mostly Ser) or the +2-position (mostly Asp) that differs from the lipobox pattern<br />

of Sutcliffe et al. (Sutcliffe and Harrington, 2002). Recently, Tjalsma and van Dijl (Tjalsma<br />

and van Dijl, 2005) have proposed the lipobox search pattern [LITAGMV] - [ASG<strong>TI</strong>MVF] -<br />

[AG] – C - [SGENTAQR] for potential lipoproteins of B. subtilis on the basis of published<br />

proteomics data. <strong>The</strong> only difference compared to the pattern by Sutcliffe et al. (Sutcliffe and<br />

Harrington, 2002) is that Leu absent from the +2 position, which is due to the fact that no<br />

42


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

potential B. subtilis lipoprotein with Leu at this position was identified by proteomics.<br />

Consistently, none of the predicted S. <strong>aureus</strong> lipoproteins has a Leu at the +2 position<br />

(Supplemental tables IIIe and IIIf). It is also noteworthy that some lipoproteins contain a<br />

[KR]-R-x-#-# motif in their signal peptides, although it has not been shown yet that<br />

lipoproteins can be transported via the Tat-pathway. Finally, the hypothetical protein Lpl2 of<br />

S. <strong>aureus</strong> N315 and Mu50 was excluded from our lipoprotein predictions, because Asp does<br />

not seem to occur at the +2 position of lipoproteins from Gram-positive bacteria (Juncker et<br />

al., 2003; Tjalsma and van Dijl, 2005). Nevertheless, the homologues of Lpl2 of the other<br />

sequenced S. <strong>aureus</strong> strains are classified as lipoproteins, because they have residues at the +2<br />

position that conform to the lipobox consensus.<br />

Lipoprotein Release Determinant<br />

Although lipoproteins were generally believed to be retained at the membrane-cell wall<br />

interface, the presence of lipoproteins in the growth medium of B. subtilis was documented by<br />

Antelmann et al. (Antelmann et al., 2001). This unexpected finding is correlated to the<br />

proteolytic removal of the amino-terminal, lipid-modified Cys, which suggests that the<br />

observed lipoprotein release into the growth medium is caused by proteolytic “shaving” after<br />

processing by LspA. In most of these lipoproteins Tjalsma and van Dijl (Tjalsma and van<br />

Dijl, 2005) identified the +1 to +10 consensus sequence C – G - [NSTF] – x - [SGN] – x -<br />

[SGKAE] – x – x - [SGA] that might represent the recognition site for a yet unidentified<br />

shaving protease. Most probably, a Gly at the +2 position is of major importance for<br />

lipoprotein release into the growth medium, while a Ser at this position seems to inhibit this<br />

proces. In the sequenced genomes of S. <strong>aureus</strong> only one lipoprotein could be found with the<br />

exact motif described above. By searching for patterns with 80% similarity to the consensus<br />

sequence (i.e. one different residue), five to seven additional lipoproteins can be found,<br />

depending on the S. <strong>aureus</strong> strain. In none of these proteins, Thr was found at the +3 position.<br />

Instead, a Lys was identified at this position in one or two predicted lipoproteins with a<br />

potential release motif, depending on the S. <strong>aureus</strong> strain. In other predicted lipoproteins with<br />

a potential release motif, no Gly or Ala residues were found at the +7 position. However, one<br />

of these lipoproteins contains a predicted release motif with a Gln at the +7 position. To date,<br />

a total number of four potential lipoproteins have been identified in the extracellular milieu of<br />

S. <strong>aureus</strong>. <strong>The</strong> first one was identified by Ziebandt et al. (Ziebandt et al., 2004). This protein<br />

has been annotated as a thioredoxin reductase (Supplemental table IIIa), but it shows no<br />

similarity to known thioredoxin reductases. Instead, it is highly similar to phosphate-binding<br />

lipoproteins, such as PstS of B. subtilis. It should be noted that this protein was not predicted<br />

as a lipoprotein, because the signal peptide contains a Gln residue in the N-domain.<br />

According to the search profile of Sutcliffe et al. (Sutcliffe and Harrington, 2002) lipoproteins<br />

would not contain Gln residues at this position. On the other hand, PstS of B. subtilis is a<br />

lipoprotein and it would seem quite likely that this is also true for its S. <strong>aureus</strong> homologue.<br />

<strong>The</strong> other three potential lipoproteins were identified in the growth medium of clinical<br />

isolates (Tabel 4). Remarkably, none of these four lipoproteins with an extracellular<br />

localization contain the complete lipoprotein release motif that was identified in extracellular<br />

lipoproteins of B. subtilis. However, they do contain a Gly residue at the +2 position, which<br />

strengthens the idea that this amino acid residue is probably important for lipoprotein release.<br />

It is interesting to note that in lipoproteins of Gram-negative bacteria, an Asp, Gly, Phe, or<br />

Trp residue at the +2 position prevents the transport of the mature lipoprotein to the outer<br />

43


Chapter 2<br />

membrane (Tokuda and Matsuyama, 2004; Narita and Tokuda, 2006). In Gram-positive<br />

bacteria no outer membrane is present and it is currently not known whether the residues at<br />

the +2 position have a role in subcellular protein sorting. However, a Gly at this position does<br />

seem to promote lipoprotein release into the extracellular milieu, not only in B. subtilis, but<br />

also in S. <strong>aureus</strong>.<br />

Cell wall binding domains<br />

Proteins that have to be displayed on the bacterial surface must be retained by non-covalent or<br />

covalent binding to the peptidoglycan moiety of the cell wall. In B. subtilis, several proteins<br />

involved in cell wall turnover contain repeated domains in the C-terminal part of the protein,<br />

which have affinity for cell wall components (Ghuysen et al., 1994; Margot and Karamata,<br />

1996; Rashid et al., 1995). Specifically, the B. subtilis proteins LytD, WapA, YocH, YvcE,<br />

and YwtD have been reported to bind to the cell wall (Ghuysen et al., 1994; Margot and<br />

Karamata, 1996; Rashid et al., 1995). While WapA is not conserved in staphylococci, various<br />

S. <strong>aureus</strong> proteins with regions that show amino acid sequence similarity to LytD, YocH,<br />

YvcE, and YwtD of B. subtilis can be found by BLAST searches. Accordingly, these S.<br />

<strong>aureus</strong> proteins may be cell wall-bound, but this remains to be shown.<br />

One of the domains that have affinity for cell wall components is the “Lysin Motif”, or LysM<br />

domain, which has first been described for bacterial lysins (Ponting et al., 1999). <strong>The</strong> number<br />

of LysM domains can differ for wall-bound proteins from different Gram-positive bacterial<br />

species (Steen, 2005). For example, XlyA of B. subtilis contains only one LysM domain,<br />

whereas three domains can be detected in AcmA of L. lactis, or even five or six domains in<br />

muramidases from Enterococcus species (Joris et al., 1992). Using the LysM domain of<br />

AcmA from L. lactis in BLAST searches against the six sequenced and annotated S. <strong>aureus</strong><br />

genomes, four proteins with one or more LysM domains were detected. <strong>The</strong>se proteins<br />

include a hypothetical protein similar to autolysins (SA0423), the secretory antigen SsaA<br />

homologue (SA0620), a conserved hypothetical protein (SA0710), and the LytN protein.<br />

A different domain that can facilitate protein binding to the cell wall is the GW domain. In L.<br />

monocytogenes, the surface-exposed InlB protein contains three C-terminal GW domains.<br />

Each domain consists of ~80 amino acids and starts with a Gly and Trp residue (Braun et al.,<br />

1997). This domain specifically binds to lipoteichoic acids in the cell wall (Jonquieres et al.,<br />

1999), thereby facilitating the interaction of L. monocytogenes with components of human<br />

host cells. <strong>The</strong> only protein found in the sequenced S. <strong>aureus</strong> strains with GW domains is the<br />

autolysin protein Atl (Baba and Schneewind, 1998; Baba and Schneewind, 1996). This<br />

bifunctional autolysin contains three GW repeats of ~97 amino acids. <strong>The</strong> protein is exported<br />

as a prepro-Atl precursor of 1256 amino acids. Subsequent processing steps result in the<br />

removal of the signal peptide and the propeptide, and the separation of the mature region into<br />

an amidase and a glucosaminidase (Oshida et al., 1995). A similar separation of the mature<br />

region into an amidase and glucosaminidase has been reported for the AtlE protein of S.<br />

epidermidis (Heilmann et al., 1997). <strong>The</strong> GW repeats are both necessary and sufficient to<br />

direct reporter proteins to the equatorial surface ring of S. <strong>aureus</strong> cells where cell division<br />

starts.<br />

Other S. <strong>aureus</strong> wall proteins that contain repeated domains with potential wall binding<br />

properties have been described. <strong>The</strong>se include the clumping factors A and B (ClfAB)<br />

(Hartford et al., 1997; Ní Eidhin et al., 1998), several serine aspartate repeat proteins<br />

(SdrCDE) (Josefsson et al., 1998), the homologue of S. gordonii GspB (SasA) (Siboo et al.,<br />

44


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

2005) (see also the section on covalent cell wall binding below), and the ECM-binding<br />

protein homologue (Clarke et al., 2002). Although not documented in the literature, additional<br />

proteins with Sec type signal peptides and potential cell wall binding repeats that can be<br />

recognized readily. <strong>The</strong>se are the cell wall surface anchor family protein SACOL2505, and<br />

the methicillin-resistant surface protein SACOL0050, which both contain C-terminal repeated<br />

regions of ~130 amino acids. <strong>The</strong> latter protein, which shows a high degree of sequence<br />

similarity to the SACOL2505 protein, is only found in S. <strong>aureus</strong> COL, but not in the five<br />

other sequenced strains. This is due to the fact that the gene for SACOL0050 is localized on<br />

the mec cassette 1 and therefore not present in the other strains. Notably, the SACOL2505<br />

homologues in S. <strong>aureus</strong> Mu50 and N315 seem to lack the C-terminal part of the protein with<br />

the repeats. A close inspection of the sequence of the corresponding genes in these strains<br />

revealed that there is a frameshift mutation or sequencing error in these genes, resulting in an<br />

apparent or real C-terminal truncation of the corresponding proteins. Thus, the C-terminal cell<br />

wall binding repeats are absent or appear to be absent. Interestingly, most of the aforementioned<br />

proteins with cell wall binding motifs also contain the motif (LPxTG) for covalent<br />

attachment to the cell wall by sortase A or sortase B (see below).<br />

It should be noted that a variety of known cell wall binding domains, such as the choline<br />

binding domain (Yother and White, 1994), the Cpl-7 cell wall binding domain (Garcia et al.,<br />

1990) and the fructosyltransferase cell wall binding domain (Huard et al., 2003; Milward and<br />

Jacques, 1990; Rathsam et al., 1993) appear to be absent from staphylococcal proteins.<br />

Covalent attachment to the cell wall<br />

Cell wall sorting proteins, known as sortases, exist in many Gram-positive bacteria and serve<br />

to anchor proteins that are destined for cell surface display to the cell wall (Dramsi et al.,<br />

2005; Ton-That et al., 2004a). In almost all Gram-positive bacteria there is at least one sortase<br />

present and often genes for more than one sortase-like protein can be detected in a single<br />

genome. <strong>The</strong>se transpeptidases catalyze the formation of an amidebond between the<br />

carboxylgroup of a Thr and the freed amino end of pentaglycine cross-bridges in<br />

peptidoglycan precursors. Subsequently, the peptidoglycan precursors with covalently bound<br />

proteins are incorporated into the cell wall. More recently, it has been shown that sortases can<br />

also be involved in protein polymerization leading to the assembly of pili on the surface of<br />

Gram-positive bacteria, such as Corynebacterium diphtheriae (Ton-That et al., 2004b; Gaspar<br />

and Ton-That, 2006). <strong>The</strong> 3D-structure of sortase A (SrtA) of S. <strong>aureus</strong> revealed that this<br />

protein has a unique β-barrel structure in which a catalytic Cys residue is positioned close to a<br />

His residue. This suggests that sortase A forms a thiolate-imidazolium ion pair for catalysis<br />

(Ilangovan et al., 2001; Ton-That et al., 2002). Furthermore, it has been shown, that a<br />

conserved Arg residue is needed for efficient catalysis (Marraffini et al., 2004). <strong>The</strong> catalytic<br />

cysteine is part of a conserved motif, TLxTC, which can be found in the C-terminal part of<br />

the protein (x is usually Val, Thr or Ile). Recently, a classification of sortases has been<br />

proposed by Dramsi et al. (Dramsi et al., 2005) based on phylogenetic analyses of 61 sortases<br />

that are encoded by the genomes of 22 Gram-positive bacteria. <strong>The</strong>se analyses showed that<br />

sortases can be grouped into four different classes (A-D). Class A consists of sortases from<br />

many low GC% Gram-positive bacteria, including L. monocytogenes, Streptococcus pyogenes<br />

and S. <strong>aureus</strong>. <strong>The</strong> second class (Class B) is present in only a few low GC% Gram-positive<br />

bacteria, including, L. monocytogenes, B. anthracis and S. <strong>aureus</strong>. Sortases of this class<br />

contain three amino acid segments that are not present in the sortases of class A. <strong>The</strong>se<br />

45


Chapter 2<br />

sortases recognize a different motif (NPQTN in S. <strong>aureus</strong>). <strong>The</strong> genes for substrates of class B<br />

sortases are often found at the same locus as the sortase gene. <strong>The</strong> largest class (Class C)<br />

consists of sortases from high GC% and low GC% Gram-positive bacteria. <strong>The</strong> genes for<br />

class C sortases are often present in multiple copies per genome. Characteristic for this class<br />

of sortases is a C-terminal hydrophobic domain that might serve as a membrane anchor, and a<br />

conserved proline residue behind the catalytic site. Finally, class D sortases are present in<br />

high and low GC% Gram-positive bacteria. This class can be divided into three subclusters,<br />

depending on whether they are present in bacilli, clostridia or actinomycetales. Since class C<br />

and D sortases are absent from S. <strong>aureus</strong>, the (potential) substrates of these enzymes will not<br />

be reviewed here.<br />

Sortase A recognition signal<br />

For interaction with host cells during infection, many proteins are anchored to the cell wall of<br />

staphylococcal cells, thereby enabling the cells to adhere and invade the host cells, or to evade<br />

the immune system. Many of these proteins contain an LPxTG-motif in their C-terminal part,<br />

which is recognized by the cell wall sorting protein sortase A. In each of the six sequenced S.<br />

<strong>aureus</strong> strains there is only one sortase gene present, which encodes a class A sortase. <strong>The</strong><br />

LPxTG motif of sortase A substrates is followed by a stretch of hydrophobic amino acids and<br />

at least one positively charged amino acid (Lys or Arg) at the C-terminus. After translocation<br />

across the membrane, the LPxTG motif is recognized by SrtA and subsequently cleaved<br />

between the Thr and Gly residues (Mazmanian et al., 2001; Navarre and Schneewind, 1994).<br />

A transpeptidation is then mediated by SrtA through amide-linkage of the C-terminal Thr of<br />

the protein to pentaglycine cross-bridges. It has been suggested that SrtA actually uses lipid II<br />

as a peptidoglycan substrate and that the proteins that are linked to lipid II are subsequently<br />

incorporated into the cell wall. In addition to the canonical LPxTG motif, a LPxAG motif can<br />

also be recognized and cleaved by SrtA (Roche et al., 2003). It has been reported that S.<br />

<strong>aureus</strong> has 19 proteins that carry the LPxTG motif and 2 proteins that carry a LPxAG motif at<br />

their C-termini (Roche et al., 2003) (Table 4). Many of these proteins have been shown to be<br />

expressed. <strong>The</strong>se include: protein A (Spa), two clumping factors (ClfA and ClfB; also contain<br />

potential wall binding repeats), a collagen-binding protein (Cna), three serine aspartate repeat<br />

proteins (SdrC, SdrD and SdrE; also contain potential wall binding repeats), two fibronectinbinding<br />

protein (FnbpA and FnbpB) (reviewed by Foster and Hook, (Foster and Hook,<br />

1998)), a plasmin-sensitive protein (Pls) (Savolainen et al., 2001), FmtB (Komatsuzawa et al.,<br />

2000), and several S. <strong>aureus</strong> surface (Sas) proteins (Roche et al., 2003). A recent study on the<br />

cell wall and membrane proteome by Nandakumar et al. (Nandakumar et al., 2005) resulted<br />

in the identification of two proteins with an LPxTG cell wall sorting signal. Many of the<br />

LPxTG proteins contain in their N-terminal signal peptide a conserved motif, [YF]-SIRK<br />

(with some variance) that has also been observed in other proteins that are substrates for SrtA<br />

in several Gram-positive bacteria (Bae and Schneewind, 2003). However, this sequence is not<br />

found in all of the SrtA substrates and it can also be found in non-cell wall proteins. This<br />

suggests that [YF]-SIRK is not a specific SrtA targeting sequence. Interestingly, proteins that<br />

contain an LPxTG-motif and a [YF]-SIRK motif in their signal peptides seem to be<br />

distributed along the staphyolococcal cell surface in a different manner than those proteins<br />

that have an LPxTG-motif, but lack the [YF]-SIRK motif. Those proteins that do contain the<br />

[YF]-SIRK motif seems to be positioned in a ring-like structure that forms during cell<br />

division for the separation of cells, while the proteins that lack this motif are directed to the<br />

cell pole (Dedent et al., 2008). One of the sas genes, sasA also known as sraP, is situated in<br />

46


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

the secA2/secY2 cluster and has an unusually long signal peptide (90 residues). Similar to<br />

what has been reported for the cell wall-bound GspB protein in S. gordonii (Bensing and<br />

Sullam, 2002), it was recently shown that the accessory SecA2/SecY2 system is needed for<br />

the transport of SasA/SraP across the membrane (Siboo et al., 2008). Depending on the<br />

sequenced S. <strong>aureus</strong> strain, 10-13 proteins with an LPxTG cell wall sorting signal, followed<br />

by a hydrophobic stretch of residues and a positively charged C-terminus can be found (Table<br />

4). Among these proteins are the fibrinogen-binding protein A (ClfA), immunoglobulin G<br />

binding protein A precursor (Spa), and the Ser-Asp rich fibrinogen-binding, bone<br />

sialoprotein-binding protein (SdrC).<br />

Table 4. Staphylococcal proteins with (potential) Sec type signal peptides and (potential)<br />

signals for covalent cell wall binding<br />

PID Protein Function Signal<br />

15925728 AsdA Adenosine synthase A LPKTG<br />

15925815 Spa Immunoglobulin G binding protein A precursor LPETG<br />

15925838 a SasD hypothetical protein LPAAG<br />

15926239 b SdrC Ser-Asp rich fibrinogen-binding, bone sialoprotein-binding protein LPETG<br />

15926240 c,d SdrD Ser-Asp rich fibrinogen-binding, bone sialoprotein-binding protein LPETG<br />

15926241 b,d,e SdrE Ser-Asp rich fibrinogen-binding, bone sialoprotein-binding protein LPETG<br />

15926464 ClfA fibrinogen-binding protein A, clumping factor LPDTG<br />

15926713 IsdB conserved hypothetical protein LPKTG<br />

15926714 IsdA cell surface protein LPKTG<br />

15926715 IsdC conserved hypothetical protein NPQTN<br />

15927308 c HarA hypothetical protein LPKTG<br />

15927333 d,f SasC hypothetical protein, similar to FmtB protein LPNTG<br />

15927741 b,g SasB FmtB protein LPDTG<br />

15928076 b,c,h Aap hypothetical protein, similar to accumulation-associated protein LPKTG<br />

15928081 c,d FnbB fibronectin-binding protein homolog LPETG<br />

15928082 d,i FnbA fibronectin-binding protein homolog LPETG<br />

15928174 j SasK hypothetical protein LPKTG<br />

15928216 ClfB Clumping factor B LPETG<br />

15928232 b SasF conserved hypothetical protein LPKAG<br />

15928240 k SasA hypothetical protein, similar to streptococcal hemagglutinin protein LPDTG<br />

57652419 l Pls methicillin-resistant surface protein LPDTG<br />

21284341 m Cna collagen adhesin precursor LPKTG<br />

27467746 n SE0828 Lipoprotein VsaC LPETG<br />

27468418 n SE1500 hypothetical protein LPKTG<br />

27468419 n SE1501 hypothetical protein LPNTG<br />

27468546 n SE1628 hypothetical protein LPETG<br />

27469070 n SE2152 hypothetical protein LPNTG<br />

a Truncated in S. <strong>aureus</strong> NCTC 8325, thereby missing the C-terminal part containing the LPxTG motif<br />

b Proteins that are also present in S. epidermidis<br />

c <strong>The</strong> genes encoding SdrC, HarA and FnbB are not present in S. <strong>aureus</strong> MRSA252<br />

d <strong>The</strong>se proteins have a lower SignalP score than our threshold score<br />

e <strong>The</strong> gene encoding SdrE is not present in S. <strong>aureus</strong> NCTC8325<br />

f Truncated in S. <strong>aureus</strong> Mu50, thereby missing the C-terminal part containing the LPxTG motif<br />

g <strong>The</strong> gene encoding SasB is not present in S. <strong>aureus</strong> MRSA252, MSSA476 and USA300_TCHC1516; truncated<br />

in S. <strong>aureus</strong> MW2<br />

h Truncated in S. <strong>aureus</strong> Mu50, N315 and Newman, thereby missing the C-terminal part containing the LPxTG<br />

motif<br />

i Truncated in S. <strong>aureus</strong> Newman, thereby missing the C-terminal part containing the LPxTG motif (Grundmeier<br />

et al., 2004)<br />

j <strong>The</strong> gene encoding SasK is not present in S. <strong>aureus</strong> COL, MRSA252, MSSA476, NCTC 8325, Newman, USA300<br />

and USA300_TCHC1516; truncated in S. <strong>aureus</strong> MW2, thereby missing the N-terminal signal peptide<br />

k This protein has an unusually long signal peptide (90 amino acids)<br />

l <strong>The</strong> gene encoding Pls is only present in S. <strong>aureus</strong> COL<br />

m <strong>The</strong> gene encoding Cna is not present in S. <strong>aureus</strong> COL, JH1, JH9, Mu50, N315, NCTC 8325, Newman,<br />

USA300 and USA300_TCHC1516<br />

n Only present in S. epidermidis<br />

47


Chapter 2<br />

Five additional proteins (SdrD, SdrE, SasC, FnbA and FnbB) with an LPxTG motif can be<br />

found among the S. <strong>aureus</strong> strains (Table 4). <strong>The</strong>se five proteins were excluded from our<br />

initial list, because the corresponding SignalP scores were lower than our (high) score criteria,<br />

or, as was shown for the FnbA and FnbB proteins of the S. <strong>aureus</strong> Newman strain, the LPxTG<br />

motif is missing due to a premature stop codon (Grundmeier et al., 2004). However, since<br />

some of the domains present in these proteins (besides the LPxTG motif) are conserved in<br />

well-described cell wall proteins, they have been included in Table 4. <strong>The</strong> remaining proteins<br />

with a cell wall sorting signal are either missing in one or more S. <strong>aureus</strong> strains, or have been<br />

annotated wrongly. Interestingly, S. epidermidis ATCC 12228 contains a gene for a class C<br />

sortase (srtC), which seems to be absent from other staphylococci. This SrtC is most closely<br />

related to sortases of L. lactis and Streptococcus suis. Two proteins with LPxTG motifs,<br />

which are encoded by the same genomic island as SrtC, also seem to be strain-specific (Gill et<br />

al., 2005).<br />

Sortase B recognition signal<br />

All sequenced S. <strong>aureus</strong> strains contain sortase B (SrtB) in addition to SrtA. <strong>The</strong> gene for<br />

SrtB is situated at a locus, which is involved in the uptake of haeme iron (Mazmanian et al.,<br />

2002). This locus also contains the gene for the cell wall protein IsdC, which contains the<br />

SrtB recognition sequence NPQTN. In addition, this locus contains the genes for the SrtA<br />

substrates IsdA and IsdB that both contain LPxTG motifs. Notably, IsdC is so far the only<br />

known protein known to be anchored to the cell wall by sortase B. IsdC is cleaved by SrtB<br />

between the Thr and Asn residues of the NPQTN motif. <strong>The</strong> only other S. <strong>aureus</strong> protein with<br />

a motif that resembles NPQTN is the DNA-binding protein II, but this protein is probably not<br />

cell wall bound, because it lacks a signal peptide for export from the cytoplasm.<br />

Comparative <strong>secretome</strong> analysis<br />

Comparing the predicted <strong>secretome</strong>s of S. <strong>aureus</strong> and S. epidermidis with those of B. subtilis<br />

and other Gram-positive bacteria revealed that most of the known components of the<br />

translocation machinery are present in S. <strong>aureus</strong>. <strong>The</strong> most notable differences are the second<br />

set of secA and secY genes in S. <strong>aureus</strong>, the absence of known signal peptide peptidases from<br />

S. <strong>aureus</strong> and S. epidermidis, the absence of a BdbC homologue from S. <strong>aureus</strong> and S.<br />

epidermidis, the presence of a second lspA gene in S. epidermidis, the absence of a Tat system<br />

from S. epidermidis, and the absence of two potential components in the Ess pathway from S.<br />

<strong>aureus</strong> MRSA252 (Table 3).. However, it has been shown that the second secA/secY set is<br />

involved in the export of virulence factors in other pathogens (Takamatsu et al., 2005).<br />

Though most known determinants for protein export, processing and post-translocational<br />

modification in other Gram-positive bacteria are also present in S. <strong>aureus</strong>, in many cases it<br />

remains to be investigated to what extent they are necessary for protein export in general and<br />

the export of virulence factors in particular.<br />

As shown by multiple BLAST comparisons, the core exoproteome of the sequenced S. <strong>aureus</strong><br />

strains consists of 68 proteins (Supplemental table IIIc). All of these proteins have a signal<br />

peptide with a potential SpsB recognition and cleavage site. 46 of these core exoproteins have<br />

already been identified in the extracellular milieu and/or membrane/cell wall proteome of<br />

different S. <strong>aureus</strong> isolates (Ziebandt et al., 2001; Ziebandt et al., 2004; Nandakumar et al.,<br />

2005; Gatlin et al., 2006; Pocsfalvi et al., 2008; Ziebandt et al., submitted). Interestingly, 31<br />

core exoproteins of S. <strong>aureus</strong> are also conserved in S. epidermidis, suggesting that they<br />

48


Mapping the pathways to staphylococcal pathogenesis by comparative secretomics<br />

belong to a core staphylococcal exoproteome, which is presently still poorly defined.<br />

Interestingly, the core exoproteome of S. <strong>aureus</strong> seems to be largely composed of enzymes,<br />

like proteases, and other factors, like fibrinogen- and IgG-binding proteins, that are required<br />

for life in the ecological niches provided by the human host (Supplemental table IIIc). This is<br />

particularly true also for the proteins that have the potential to be covalently bound to the cell<br />

wall (Table 4). In contrast, the variant exoproteome of S. <strong>aureus</strong> contains most of the known<br />

staphylococcal toxins and immunomodulating factors (Supplemental tables IIId and IIIg).<br />

This suggests that the components of the variant exoproteome should be regarded as specific<br />

gadgets of S. <strong>aureus</strong> that help this organism to conquer certain protected niches of the human<br />

host, thereby causing disease. If this idea is correct, proteins of unknown function that belong<br />

to the variant exoproteome should be regarded as potentially important virulence factors.<br />

<strong>The</strong> (predicted) extracellular toxins of S. <strong>aureus</strong> are not present in S. epidermidis. This is<br />

mainly due to the fact that these toxins are encoded by pathogenicity islands in the genomes<br />

of S. <strong>aureus</strong> strains that have, so far, not been observed in S. epidermidis genomes. Proteins<br />

with predicted signal peptides that are specific for S. epidermidis are listed in Supplemental<br />

table IIIi. Notably, the majority (i.e. 31 out of 37) of predicted S. epidermidis exoproteins that<br />

have homologues in S. <strong>aureus</strong> share this homology with components of the core exoproteome<br />

of S. <strong>aureus</strong> (Supplemental tables IIIc and IIId). This suggests that also in S. epidermidis the<br />

core exoproteome is involved in housekeeping functions. In contrast to the exoproteome, it is<br />

presently difficult to speculate about housekeeping- and disease-causing roles of the constant<br />

and variant lipoproteomes of S. <strong>aureus</strong>. This is due to the fact that the function of only few S.<br />

<strong>aureus</strong> lipoproteins is known (Supplemental tables IIIe, IIIf and IIIh). In general terms, it is<br />

presently not clear why S. epidermidis seems to export a lower number of different proteins<br />

(94 in total) than S. <strong>aureus</strong> (~165 in total). This difference is all the more remarkable since the<br />

total number of proteins encoded by the genomes of S. <strong>aureus</strong> (~2600) and S. epidermidis<br />

(~2500) are comparable.<br />

Compared to B. subtilis and B. licheniformis (Voigt et al., 2005), S. <strong>aureus</strong> is also predicted to<br />

export a relatively higher number of proteins from the cytoplasm to the membrane-cell wall<br />

interface, the cell wall and the extracellular milieu. <strong>The</strong> genomes of B. subtilis and B.<br />

licheniformis contain ~4100 protein-encoding genes, while the S. <strong>aureus</strong> genomes contain<br />

significantly less genes (~2600). Using the most recent prediction protocols (Tjalsma and van<br />

Dijl, 2005), B. subtilis is predicted to export 190 proteins to an extracytoplasmic location<br />

whereas, depending on the strain investigated, S. <strong>aureus</strong> is predicted to export 145-168<br />

proteins (this review). Accordingly, as judged by the relative numbers of protein-encoding<br />

genes, S. <strong>aureus</strong> strains appear to export 1-2% more proteins to an extracytoplasmic location<br />

than the afore-mentioned bacilli. Most probably, this is related to the fact that S. <strong>aureus</strong> needs<br />

an arsenal of virulence factors, such as toxins and surface proteins, for colonization of and<br />

survival in its preferred niches in the human host. Such proteins are of less importance for soil<br />

bacteria, such as B. subtilis and B. licheniformis, which thrive predominantly on dead organic<br />

matter.<br />

Perspectives<br />

<strong>The</strong> present review provides a survey of possible protein transport pathways to staphylococcal<br />

pathogenesis. In many cases the knowledge gathered from protein secretion studies in other<br />

organisms has been projected on S. <strong>aureus</strong>, assuming that similar pathways or pathway<br />

components have similar functions in different organisms. Clearly, this leaves room for<br />

49


Chapter 2<br />

surprises when such pathways are investigated thoroughly in S. <strong>aureus</strong>. <strong>The</strong> same was true for<br />

studies on protein secretion in B. subtilis. <strong>The</strong>se studies showed, for example, that the absence<br />

of SecDF has barely any consequences for protein secretion by B. subtilis, whereas SecD and<br />

SecF are of key importance for protein translocation in E. coli, the organism in which SecD/F<br />

was first discovered (Bolhuis et al., 1998). Likewise, LspA was shown to be dispensable in B.<br />

subtilis, but not in E. coli (Tjalsma et al., 1999; Prágai et al., 1997). Thus the relative<br />

importance of different secretion machinery components of S. <strong>aureus</strong> needs to be assessed in<br />

a systematic manner, preferably in an isogenic background. Such studies would need to<br />

address the importance of secretion machinery components for in vitro growth on different<br />

substrates (e.g. broth or blood), and virulence in in vivo model systems (e.g. C. elegans,<br />

Drosophila melanogaster, mice and rats) (Bae et al., 2004; García-Lara et al., 2005;<br />

Tarkowski et al., 2001). <strong>The</strong>se studies should be complemented with a proteomic verification<br />

of our present lipoproteome, wall proteome and exoproteome predictions. Such a verification<br />

could involve both gel-based proteomics approaches as outlined in this review, and more<br />

sophisticated gel-free proteomics approaches (Völker and Hecker, 2005). This would lead to<br />

an improved understanding of the contribution of each protein transport pathway and its<br />

substrate proteins to staphylococcal cell physiology and virulence. Since the virulence of<br />

different S. <strong>aureus</strong> strains will not only depend on the presence (or absence) of particular<br />

genes for virulence factors, but also on their expression, such proteomic studies should also<br />

include experiments on the impact of major regulatory systems, such as agr, sae and sarA. On<br />

this basis, it should be possible to identify the most promising candidate drug targets in the<br />

staphylococcal <strong>secretome</strong>. Alternatively, <strong>secretome</strong> components thus identified could<br />

represent promising candidates for novel vaccines. For all these efforts, comparative<br />

secretomics approaches will provide the essential information on those potential drug targets<br />

that are most important for both bacterial housekeeping functions and virulence. Novel drugs<br />

and vaccines designed against these targets are likely to have the highest impact (Götz, 2004).<br />

Acknowledgements<br />

We thank Harold Tjalsma and members of the Groningen and European Bacillus Secretion<br />

Groups and the BACELL Health, Tat machine and StaphDynamics consortia for stimulating<br />

discussions. M.J.J.B.S., A.K.Z., S.E., M.H., J.-Y.F.D., and J.M.v.D. were supported by Grants<br />

LSHC-CT-2004-503468, LSHG-CT-2004-005257 and LSHM-CT-2006-019064 from the<br />

European Union. M.H. was supported from grants of the “Deutsche Forschungsgemeinschaft”,<br />

the “Bundesministerium für Bildung, Wissenschaft, Forschung und Technologie”, and the<br />

“Fonds der Chemischen Industrie”.<br />

50


“Music is the medicine of the mind”<br />

-John A. Logan-<br />

52


Chapter 3<br />

Proteogenomics uncovers extreme heterogeneity in the<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> exoproteome due to genomic plasticity<br />

and variant gene regulation<br />

A.K. Ziebandt, M. Degner, M.J.J.B. Sibbald, J.P. Arends, M.A. Chlebowicz,<br />

H. Kusch, D. Albrecht, R. Pantuček, J. Doškar, W. Ziebuhr, B.M. Bröker, M. Hecker,<br />

J.M. van Dijl, and S. Engelmann<br />

Submitted for publication, in revision<br />

53


Chapter 3<br />

Summary<br />

Sequencing of at least thirteen S. <strong>aureus</strong> isolates has shown that genomic plasticity<br />

impacts significantly on the repertoire of virulence factors. However, genome<br />

sequencing does not reveal which genes are de facto expressed by individual isolates.<br />

Here, we have therefore performed a first comprehensive survey of the composition and<br />

variability of the S. <strong>aureus</strong> exoproteome following a proteogenomics approach. This<br />

involved multi locus sequence typing, virulence gene and prophage profiling by<br />

multiplex PCR, and proteomic analyses of secreted proteins using two-dimensional<br />

protein gel electrophoresis. Dissection of the exoproteomes of 25 clinical isolates revealed<br />

that only seven out of 63 identified secreted proteins were produced by all isolates,<br />

indicating a remarkably high exoproteome heterogeneity within one bacterial species.<br />

<strong>The</strong> observed variations were caused by both genome plasticity and an unprecedented<br />

variation in gene expression. Our data imply that genomic studies focussing on virulence<br />

gene conservation patterns need to be complemented by protein expression analyses to<br />

assess the full virulence potential of bacterial pathogens like S. <strong>aureus</strong>. Importantly, the<br />

extensive variability of secreted virulence factors in S. <strong>aureus</strong> also suggests that the<br />

development of protective vaccines against this pathogen requires a carefully selected<br />

combination of invariably expressed antigens.<br />

54


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

Introduction<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> causes a wide variety of human infections ranging from superficial<br />

lesions to severe systemic diseases. Up to one third of the human population carries S. <strong>aureus</strong><br />

as a commensal bacterium without developing any clinical symptoms. Nevertheless, the<br />

colonizing strain can serve as an endogenous reservoir for infection, the incidence of<br />

bacteremia being less than 0.02% (von Eiff et al., 2001). Individuals colonized with S. <strong>aureus</strong><br />

have thus a higher risk to develop an S. <strong>aureus</strong> infection, but they are also more likely to<br />

overcome the disease (Wertheim et al., 2004). It is undisputed that the incidence and outcome<br />

of invasive staphylococcal diseases strongly depend on host factors, in particular the immune<br />

status of a patient. Yet, the species S. <strong>aureus</strong> is highly diverse and up to 30% of the genomes<br />

of different isolates consist of variable regions, such as pathogenicity islands, lysogenic<br />

bacteriophages, and plasmids (Witney et al., 2005). Since these genetic elements encode the<br />

majority of virulence factors, it is generally thought that their presence critically determines<br />

the clinical symptoms and outcome of an S. <strong>aureus</strong> infection. However, virulence-associated<br />

genes may also be located in the core genome as illustrated by the spa, aur, hla, lip, clfAB,<br />

map/eap, fnbA, and coa genes (Holden et al., 2004; Peacock et al., 2002). It is generally<br />

accepted that certain strains are more virulent than others (Melles et al., 2004), although,<br />

under certain conditions, any S. <strong>aureus</strong> genotype may have the potential to induce lifethreatening<br />

infections. This suggests that there is no simple relationship between bacterial<br />

genotype and clinical outcome. Apart from several well-defined toxin-mediated diseases (e.g.<br />

toxic shock and scalded skin syndrome, necrotizing pneumonia) (Musser et al., 1990;<br />

Gemmell, 1995; Labandeira-Rey et al., 2007), it remains difficult to predict the onset and<br />

course of an S. <strong>aureus</strong> infection in a given patient solely on the basis of the virulence gene<br />

repertoire of the strain involved. In fact, the vast majority of severe S. <strong>aureus</strong> infections are<br />

probably caused by the concerted action of multiple virulence factors. Molecular typing and<br />

genome analyses of clinical isolates focussing on virulence gene distribution have been major<br />

steps towards a thorough understanding of these complex phenomena. Also, there is growing<br />

evidence that the presence and activities of certain regulators play a role in the variation of<br />

virulence gene expression in clinical isolates (Blevins et al., 2002; Karlsson and Arvidson,<br />

2002).<br />

Here, we complement the genetic identification of virulence factors in different S. <strong>aureus</strong><br />

isolates with proteome analyses of secreted proteins, which represent an important reservoir<br />

of virulence factors. More specifically, in this proteogenomics approach we have addressed<br />

the questions (i) of whether particular virulence genes are expressed in different isolates and,<br />

if so, (ii) in what quantities? <strong>The</strong> results reveal an unprecedented heterogeneity in the<br />

analyzed S. <strong>aureus</strong> exoproteomes, which is caused by both genomic plasticity and an amazing<br />

variability in the levels of gene expression.<br />

Materials and Methods<br />

Bacterial strains<br />

Twenty five S. <strong>aureus</strong> isolates derived from a variety of human infections were collected by the<br />

university hospital in Groningen. All strains were identified as S. <strong>aureus</strong> by plating and coagulase tests.<br />

Resistance against antibiotics was determined by routine disk diffusion assays. For RNAIII<br />

55


Chapter 3<br />

transcription analyses, S. <strong>aureus</strong> RN6390, COL, and Newman were used as reference strains (Shafer<br />

and Iandolo, 1979; Duthie and Lorenz, 1952; Novick, 1967).<br />

Multilocus sequence typing (MLST)<br />

MLST was performed according to the protocol described by Enright et al. (Enright et al., 2000). <strong>The</strong><br />

obtained sequences for each locus were submitted to the Internet database (www.mlst.net) and the<br />

resulting allelic profiles were assigned to particular sequence types (ST) for each isolate. <strong>The</strong> eBURST<br />

(Based upon related sequences) algorithm software was used to classify different sequence types into<br />

clusters of clonal complexes (CC).<br />

Detection of prophages and virulence genes using multiplex PCR<br />

Genomic DNA was extracted from S. <strong>aureus</strong> clinical isolates and used as a template in multiplex PCR<br />

assays targeting structural prophage genes of head-tail modules described previously (Pantucek et al.,<br />

2004). Three different phage types (A-like, B-like, and F-like) corresponding to putative phage species<br />

3A, 11, and 77, respectively, were thus identified. <strong>The</strong> F-like phages include clearly distinguishable<br />

subgroups Fa and Fb, while the B-like phages include 5 subgroups Ba – Be of different phages, in<br />

which the packaging, head, and tail genes belong to different modules. <strong>The</strong>refore, three PCR assays<br />

were used for phage identification (Table 1).<br />

Table 1. Oligonucleotides used in this study<br />

Target Primer Sequence 5´- 3´ Reference<br />

S. <strong>aureus</strong>, positive SAU1<br />

control<br />

SAU2<br />

A-like phage, tail SGA1<br />

SGA2<br />

B-like phage (all SGB1<br />

subgroups), tail SGB2<br />

F-like phage (both SGF1<br />

subgroups), tail SGF2<br />

B-like phage, SGBa1<br />

subgroup Ba portal SGBa2<br />

B-like phage, SGBb1<br />

subgroup Bb portal SGBb2<br />

B-like phage, SGBc1<br />

subgroup Bc portal SGBc2<br />

B-like phage, SGBd1<br />

subgroup Bd portal SGBd2<br />

B-like phage, SGBe1<br />

subgroup Be portal SGBe2<br />

F-like phage, SGFa1<br />

subgroup Fa portal SGFa2<br />

F-like phage, SGFb1<br />

subgroup Fb portal SGFb2<br />

Phage integrase phiSa1-F<br />

ФSa1 (phage ETA) phiSa1-R<br />

Phage integrase phiSa2-F<br />

ФSa2 (phage 12) phiSa2-R<br />

Phage integrase phiSa3-F<br />

ФSa3 (phage 13) phiSa3-R<br />

Phage integrase phiSa4-F<br />

ФSa4<br />

phiSa4-R<br />

Phage integrase phiSa5-F<br />

ФSa5 (phage 11) phiSa5-R<br />

Phage integrase phiSa6-F<br />

ФSa6 (phage L54a) phiSa6-R<br />

Phage integrase phiSa7-F<br />

ФSa7 (phage 96) phiSa7-R<br />

Phage integrase phiSa8-F<br />

ФSa8 (phage 53) phiSa8-R<br />

GACGGCTTTGATGGCTAGTGG<br />

AGTTAATTCACGCCCTAGTG<br />

TATCAGGCGAGAATTAAGGG<br />

CTTTGACATGACATCCGCTTGAC<br />

ACTTATCCAGGTGGYGTTATTG<br />

TGTATTTAATTTCGCCGTTAGTG<br />

CGATGGACGGCTACACAGA<br />

TTGTTCAGAAACTTCCCAACCTG<br />

AAGATGATAACTTTAGTGGCAC<br />

TCATTGATGTYTCTAGGGTC<br />

CTGATTATGTGTACGCAGAG<br />

TTCCGTTAAACTCGTCAGA<br />

TTGTTAAGGAACCYAAGCC<br />

GCCTCTAATTCTTCGTGCTC<br />

AAGTTACGTCGCTGGC<br />

GCTTGTTCTGCTGGCACTCT<br />

AAATGAAACTATTCCGTGTT<br />

AAYGCTATAAAYGGYACTCT<br />

TACGGGAAAATATTCGGAAG<br />

ATAATCCGCACCTCATTCCT<br />

AGACACATTAAGTCGCACGATAG<br />

TCTTCTCTGGCACGGTCTCTT<br />

AAGCTAAGTTCGGGCACA<br />

GTAATGTTTGGGAGCCAT<br />

TCAAGTAACCCGTCAACTC<br />

ATGTCTAAATGTGTGCGTG<br />

GAAAAACAAACGGTGCTAT<br />

TTATTGACTCTACAGGCTGA<br />

ATTGATATTAACGGAACTC<br />

TAAACTTATATGCGTGTGT<br />

AAAGATGCCAAACTAGCTG<br />

CTTGTGGTTTTGTTCTGG<br />

GCCATCAATTCAAGGATAG<br />

TCTGCAGCTGAGGACAAT<br />

AAACTAAAGCTGAGGCAAC<br />

TCATTAGTACGACCTCGAC<br />

GTCCGGTAGCTAGAGGTC<br />

GGCGTATGCTTGACTGTGT<br />

56<br />

(Pantucek et al., 2004)<br />

(Pantucek et al., 2004)<br />

(Pantucek et al., 2004)<br />

(Pantucek et al., 2004)<br />

this study<br />

this study<br />

this study<br />

this study<br />

this study<br />

(Pantucek et al., 2004)<br />

(Pantucek et al., 2004)<br />

this study<br />

this study<br />

this study<br />

this study<br />

this study<br />

this study<br />

this study<br />

this study


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

Target Primer Sequence 5´- 3´ Reference<br />

hlb hlb-2<br />

AGCTTCAAACTTAAATGTCA (Goerke et al., 2006)<br />

hlb-527<br />

CCGAGTACAGGTGTTTGGTA<br />

sea Sea-1<br />

AGATCATTCGTGGTATAACG (Van Wamel et al., 2006)<br />

Sea-2<br />

TTAACCGAAGGTTCTGTAGA<br />

sep Sep-1<br />

AATCATAACCAACCGAATCA (Van Wamel et al., 2006)<br />

Sep-2<br />

TCATAATGGAAGTGCTATAA<br />

sak Sak-1<br />

AAGGCGATGACGCGAGTTAT (Van Wamel et al., 2006)<br />

Sak-2<br />

GCGCTTGGATCTAATTCAAC<br />

chp Chip-up CACCATCATTCAGCGAAAG this study<br />

Chip-low GATATATAAAGGTTTGGCAAG<br />

scn Scin-up<br />

AGTCTTTTGACTTAAGAGC this study<br />

Scin-low GTTTTAGCATCACCACTAGTA<br />

geh gehSA-up TTCTTAATGGCTACAACGAGT this study<br />

gehSA-low GATAAAATCGACATGATCCC<br />

hlgA hlgA_F<br />

GTCAAGGTGCAGAAATCATC this study<br />

hlgA_R GAAATAGTCTCTTGCTGCTG<br />

splD splD_F<br />

CCAAAGCAGAAAATAGTGTG this study<br />

splD_R<br />

CATAACACCAATTGCTTCTC<br />

splE splE_F<br />

GTGTCGTTTCGATAGGATCT this study<br />

splE_F<br />

GTTGCCTTTAACTGACAGCA<br />

24636604 24636604_F GACAGCATCTCACTGTAAAG this study<br />

24636604_R TCACGCTTTTGACAGATAAG<br />

SAR0422 SAR0422_F CAAGTTTAGCATTGGGAATG this study<br />

SAR0422_R CGTCGATAGAATCACCCATAC<br />

SAR0435 SAR0435_F GTTTAGCACTAGGGATTTTG this study<br />

SAR0435_R GGCTCTGTCTTATAAAGTTC<br />

SAR1905 SAR1905_F CGATTGCAACTCTTTCATTC this study<br />

SAR1905_R CCATACAAAGGCATACTAAG<br />

SA1755 SA1755_F CTTTTGAACCGTTTCCTACA this study<br />

SA1755_R CAGTATTTTTTGTGCCTTTC<br />

SA0092 SA0092_F GTAAAGAAGCGGAAGTTAAG this study<br />

SA0092_R GCTTTATTCGTCGGTATATC<br />

SA1001 SA1001_F CTGCAGGTCTTTTAACTCAA this study<br />

SA1001_R GCTTTCTTCACATCACCTTG<br />

SACOL0478 SACOL0478_F GTTAGATCACAAGCTACTCA this study<br />

SACOL0478_R GACGTTGATGTAACACTATC<br />

RNAIII RNAIII_F AGGAAGGAGTGATTTCAATG this study<br />

RNAIII_R CTAATACGACTCACTATAGGGAG<br />

AACTCATCCCTTCTTCATTAC<br />

Notably, the nomenclature of the prophage genomes based only on the genes for virion proteins is not<br />

absolute, because of their mosaic structure. <strong>The</strong>refore, a novel PCR-based molecular assay for<br />

identification and classification of S. <strong>aureus</strong> phage integrase genes was also applied for prophage<br />

characterization. For designation of bacteriophage integrase gene families, the updated classification<br />

scheme was used which denotes the previously sequenced phages and S. <strong>aureus</strong> genomes ФSa1 - ФSa5<br />

(Lindsay and Holden, 2004). Since so far unclassified integrases were identified, the new designations<br />

ФSa6, ФSa7 and ФSa8 were introduced in this work for integrases of phages L54a (GenBank<br />

Accession no. M27965), Ф96 (NC_007057) and Ф53 (NC_007049), respectively. For multiplex PCR,<br />

the reaction mixtures (25 µl) consisted of 75 mM Tris-HCl pH 9.0, 50 mM KCl, 2 mM MgCl2, 250 µM<br />

each of dNTPs, primer pairs phiSa1 and phiSa2 (0.25 µM each), phiSa3 (0.2 µM each), phiSa4 and<br />

phiSa5 (0.2 µM each), phiSa6, phiSa7 and phiSa8 (0.1 µM each). Primer sequences are listed in Table<br />

1. To each reaction, 1.5 unit Taq DNA polymerase (Invitrogen) and template DNA (50 ng) in a 3 µl<br />

volume were added. PCR was performed using 30 cycles of amplification consisting of denaturation (1<br />

min, 94°C), annealing (1 min 30 s, 56°C) and DNA chain extension (1 min, 72°C). Phage associated<br />

virulence genes sea, sep, sak, chp, and scn as well as geh and hlb were detected using standard PCR<br />

with the primers listed in Table 1.<br />

For the analysis of genes encoding superantigens and other extracellular proteins as well as for agr<br />

typing, primer pairs were used as shown in Table 1 or described previously (Holtfreter et al., 2004;<br />

57


Chapter 3<br />

Lina et al., 2003). <strong>The</strong> amplifications were performed with Taq polymerase in a thermocycler with the<br />

following conditions: initial denaturation at 95°C for 5 min, followed by 30 stringent cycles (1 min of<br />

denaturation at 95°C, 1 min of annealing at the temperature indicated in Table 1, and 1 min of<br />

extension at 72°C), and a final extension step at 72°C for 5 min. <strong>The</strong> quality of the DNA extracts and<br />

the absence of PCR inhibitors were confirmed by amplification of glyceraldehyde-3-phosphate<br />

dehydrogenase or 16S rRNA. <strong>The</strong> PCR products were then analyzed by electrophoresis through a 1%<br />

agarose gel. At least two independent experiments were performed for each determination.<br />

Extracellular proteins<br />

For the preparation of extracellular protein extracts, bacteria were grown in Tryptic Soy Broth (TSB).<br />

At optical densities (OD540) of 10, the extracellular proteins from 100 ml supernatant were precipitated,<br />

washed, dried, and resolved as described previously (Ziebandt et al., 2004). <strong>The</strong> protein concentration<br />

was determined using Roti ® -Nanoquant according to manufacturer´s instructions (Carl Roth GmbH &<br />

Co, Karlsruhe, Germany).<br />

Analytical and preparative 2D polyacrylamide gel electrophoresis (PAGE)<br />

Protein extracts (350 µg) were loaded onto commercially available IPG strips (pH 3-10, GE-<br />

Healthcare, Uppsala, Sweden). 2D PAGE was performed as described previously (Bernhardt et al.,<br />

1999; Eymann et al., 2004). <strong>The</strong> resulting protein gels were stained with colloidal Coomassie Blue G-<br />

250 (Candiano et al., 2004) and scanned with the light scanner.<br />

Protein identification<br />

For identification of proteins by Matrix-assisted Laser Desorption Ionization-Time of Flight mass<br />

spectrometry (MALDI-TOF MS), Coomassie stained protein spots were excised from gels using a spot<br />

cutter (Proteome Work TM ) with a picker head of 2 mm and transferred into 96-well microtiter plates.<br />

Digestion with trypsin and subsequent spotting of peptide solutions onto the MALDI targets were<br />

performed automatically in an Ettan Spot Handling Workstation (GE-Healthcare, Little Chalfont,<br />

United Kingdom) using a modified standard protocol (Eymann et al., 2004). MALDI-TOF MS<br />

analyses of spotted peptide solutions were carried out on a Proteome-Analyzer 4700/4800 (Applied<br />

Biosystems, Foster City, CA as described previously (Eymann et al., 2004). MALDI-TOF-TOF<br />

analysis was performed for the three highest peaks of the TOF spectrum as described previously<br />

(Eymann et al., 2004; Wolf et al., 2008). Database searches were performed using the GPS explorer<br />

software version 3.6 (build 329) with the organism-specific databases.<br />

By using the MASCOT search engine version 2.1.0.4. (Matrix Science, London, UK) the combined MS<br />

and MS/MS peak lists for each protein spot were searched against a database containing protein<br />

sequences derived from the genome sequences of all sequenced S. <strong>aureus</strong> strains and, moreover, all<br />

additional protein sequences of S. <strong>aureus</strong> that have been found in publically available databases. Search<br />

parameters were as described previously (Wolf et al., 2008).<br />

RNA-isolation and dot blot analysis<br />

Total RNA from S. <strong>aureus</strong> was isolated using the acid-phenol method with some modifications (Fuchs<br />

et al., 2007). Digoxigenin-labeled RNA probe for RNAIII was prepared by in vitro transcription with<br />

T7 RNA polymerase by using a PCR fragment as template. <strong>The</strong> PCR fragment was generated by using<br />

chromosomal DNA of S. <strong>aureus</strong> N315 and the respective oligonucleotides (Table 1). Dot blot analyses<br />

were carried out by using serial dilutions of total RNA prepared from S. <strong>aureus</strong> isolates and reference<br />

strains grown in TSB medium to an optical density (OD540) of 10. <strong>The</strong> digoxigenin-labeled RNA probe<br />

was used for hybridization. <strong>The</strong> hybridization signals were detected using a Lumi-Imager and analyzed<br />

using the software package Lumi-Analyst (Roche Diagnostics, Mannheim, Germany).<br />

58


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

Results and Discussion<br />

Genetic characterization of clinical isolates<br />

A total of 25 clinical isolates, including eight nosocomial methicillin-resistant S. <strong>aureus</strong><br />

(MRSA) and three community-acquired MRSA (caMRSA) isolates, were used in this study.<br />

<strong>The</strong> isolates were obtained from 19 patients with different clinical symptoms during a 4.5<br />

year period (from February 2001 to September 2005). Strains were isolated from blood, nose,<br />

feet, perineum, fingers, throat, pleural fluid, and umbilicus. Five of these isolates were<br />

colonizing strains while 12 isolates came from septic patients, five isolates induced wound<br />

infections, four isolates arthritis, three isolates pneumonia, two isolates cholangitis, and three<br />

individual isolates were involved in abscess formation, panaritium, or endocarditis. On the<br />

basis of strain typing data obtained by Pulsed-Field-Gel-Electrophoresis (PFGE) and Multi-<br />

Locus-Sequence-Typing (MLST) the isolates were grouped into eleven different sequence<br />

types (Supplemental table IVa). Three new sequence types were detected, i.e. ST869, ST870,<br />

and ST903. Moreover, the agr types and the prevalence of enterotoxin genes and other<br />

clinically relevant genes (e.g. mecA, pvl, etd, eta) were determined for all isolates by using<br />

multiplex PCR (Table 2, Supplemental table IVb).<br />

<strong>The</strong> sea gene (72%) as well as genes belonging to the enterotoxin gene cluster (egc) (44%)<br />

were found to be the most prevalent enterotoxin genes, while none of the isolates carried tst-1,<br />

sec, see, seh, or sel. Moreover, pvl was identified in four and etd in two isolates, respectively.<br />

Typing of agr revealed that six isolates encoded agr1, thirteen agr2, three agr3 and none of<br />

the isolates encoded agr4. Additionally, we analyzed the prophage content of these isolates by<br />

a multiplex PCR assay. Altogether we identified 11 different prophages. While most of the<br />

isolates contained three prophages, no known prophage was detected in two isolates (N, W)<br />

(Table 3). <strong>The</strong> hlb converting phage Sa3 was the most prevalent phage: in 20 isolates we<br />

identified either the Fa-type (16 isolates) or the Fb-type of the Sa3 prophage. This prophage is<br />

common among human isolates and often encodes immune evasion molecules (SAK, SCIN,<br />

and CHIP) as well as enterotoxins (SEA or SEP) and has previously been detected in<br />

collections of clinical strains (Goerke et al., 2006). Moreover, we identified the Fa-type phage<br />

Sa1 (1 isolate), the A-type phages Sa2 (7 isolates), Sa4 (1 isolate), and Sa6 (6 isolates), as<br />

well as the B- type phages Sa1 (2 isolates), Sa5 (1 isolate), Sa6 (3 isolates), Sa8 (9 isolates),<br />

and Sa9 (2 isolates).<br />

According to our genetic characterization, the 25 isolates represent 17 clonally divergent<br />

strains. Five strains were isolated more than once, either in the same patient or in different<br />

patients (Supplemental table IVa). Interestingly, isolates C, D, G, H, I, J, and K might have<br />

evolved from just one clone (=G228) which was responsible for a hospital outbreak and<br />

caused wound infections in three of the patients included in this study.<br />

<strong>The</strong> exoproteomes of clinical isolates are highly heterogeneous<br />

Proteomics is an extremely powerful tool for analyzing virulence gene expression in multiple<br />

clinical isolates. Since the extracellular proteome of a pathogenic bacterium represents a key<br />

reservoir of virulence factors, the present study focused on this particular subproteome. All 25<br />

isolates were cultivated in TSB medium and extracellular proteins were prepared from the<br />

supernatants at an optical density (OD) at 540 nm of 10. Proteins were separated on 2D gels<br />

and protein spots were identified by MALDI-TOF MS/MS or N-terminal sequencing<br />

59


Chapter 3<br />

(Supplemental tables IVb). Comparison of extracellular protein patterns revealed an<br />

unanticipated degree of exoproteome heterogeneity among the 17 clonally different strains<br />

(Figure 1).<br />

Table 2. Virulence genes identified by multiplex PCR in different clinical S. <strong>aureus</strong> isolates<br />

Isolates<br />

Gene A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

eta<br />

etd + +<br />

hlgA + + + + + + + + + + + + + + + + + + + + + + + + +<br />

pvl + + + +<br />

SA0092 + + + + + + + + + + + + + + + + + + + + + + +<br />

SA1001 + + + + + + + + + + + + + + + + + + + + + + +<br />

SA1755 + + + + + + + + +<br />

SACOL0478 + + + + + +<br />

SAR0422 + + + + + + + + + + + + + + + + + + + + + + + + +<br />

SAR0435 +<br />

SAR1905 + + +<br />

sea + + + + + + + + + + + + + + + +<br />

seb + + + + + + + + +<br />

sec<br />

sed + +<br />

see<br />

seg + + + + + + + + + + +<br />

seh<br />

sei + + + + + + + + + + +<br />

sej + +<br />

sek + + + +<br />

sel<br />

sem + + + + + + + + + + +<br />

sen + + + + + + + + + + +<br />

seo + + + + + + + + + + +<br />

sep + + +<br />

seq + + + +<br />

ser + +<br />

seu +<br />

splD + + + + + + + + + + + + + + + + + + + + + + + + +<br />

splE + + + + + + + + + + +<br />

tst-1<br />

24636604 + +<br />

60


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong> exoproteome due to genomic plasticity and variant gene<br />

regulation<br />

Table 3. Prophages identified in different clinical S. <strong>aureus</strong> isolates<br />

Phage integrase Classification of prophages according to Innate immune evasion cluster # of<br />

Assumed lysogenic pattern<br />

genes for virion proteins<br />

pattern of F-like phages phages<br />

(type of capsid a /integrase class)<br />

Strain Sa1 Sa2 Sa3 Sa4 Sa5 Sa6 Sa7 Sa8 Alike<br />

B-<br />

like<br />

F-<br />

like Ba Bb Bc Bd Be Fa Fb hlb sea sep sak chp scn<br />

A + + + + + + + + 3 A-phage/Sa2; B-phage/Sa6; Fa-phage/Sa3<br />

B + + + + + + + + + + + 3 A-phage/Sa2; B-phage/Sa6; Fa-phage/Sa3<br />

C + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

D + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

E + + + + + + + + + + + 3 A-phage/Sa2; B-phage/Sa8; Fb-phage/Sa3<br />

F + + + + + + + + + + + 3 A-phage/Sa2; B-phage/Sa8; Fb-phage/Sa3<br />

G + + + + + + 2 A-phage/Sa6; B-phage/Sa8<br />

H + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

I + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

J + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

K + + + + + + + + + + + 3 A-phage/Sa6; B-phage/Sa8; Fa-phage/Sa3<br />

L + + + + + + + + 2 A-phage/Sa2;Fb-phage/Sa3<br />

M + + + + + + + 2 A-phage/Sa2;Fb-phage/Sa3<br />

N + + 0<br />

O + + + + + + + 1 Fa-phage/Sa3<br />

P + + + + + + + + + 2 B-phage/Sa1; Fa-phage/Sa3<br />

Q + + + + + + + + + + + + + 4 A-phage/Sa2; B-phage/Sa5; B-phage/Sa6; Faphage/Sa3<br />

R + + + + + 1 B-phage/Sa8<br />

S + + + + + + 1 Fa-phage/Sa3<br />

T + + + + + + 1 Fa-phage/Sa1<br />

U + + + + + + + + + + 3 A-phage/Sa4; B-phage/not identified; Fa-phage/Sa3<br />

V + + + + + + + + + + 2 B-phage/Sa1; Fa-phage/Sa3<br />

W + 0<br />

X + + + + + + 1 Fa-phage/Sa3<br />

Y + + + + + + 1 Fa-phage/Sa3<br />

a According to R. Pantuček et al. (Pantucek et al., 2004)<br />

61


Chapter 3<br />

Figure 1. Characterization of extracellular proteomes of different clinical S. <strong>aureus</strong> isolates. Cells were grown<br />

in TSB medium to an OD 540 of 10. Proteins in culture supernatants were collected by TCA precipitation. 350 µg of<br />

the protein extract of each strain was separated on 2D gels, using commercially available IPG strips (pH 3-10,<br />

GE-Healthcare, Sweden) for the first dimension. Protein spots were detected by staining with colloidal Coomassie<br />

Brillant Blue. For protein identification, individual protein spots were excised from the gel and digested with<br />

trypsin. <strong>The</strong> resulting peptide solution was analyzed by tandem mass spectrometry on a Proteome Analyzer<br />

4700/4800 (Applied Biosystems, USA). <strong>The</strong> respective strain is indicated in the upper left corner of each gel.<br />

Altogether 206 distinct proteins were identified (Supplemental table IVb). 107 of these<br />

proteins showed signal peptides typical for Sec-translocated proteins. Using PSORTb<br />

software (http://www.psort.org/psortb), 63 of the Sec-translocated proteins were predicted to<br />

62


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

be extracellular, 19 were predicted to be cell wall-bound proteins (Figure 2) and the<br />

localization of a further 25 proteins is currently unknown (Sibbald et al., 2006). Moreover, we<br />

identified 72 cytoplasmic proteins and 5 membrane proteins. <strong>The</strong> N-terminal sequences<br />

(AEANSSMVSKK and GTTPAAA) of two protein spots did not match any of the protein<br />

sequences present in the NCBI and other databases, indicating that our knowledge of<br />

extracellular proteins produced by S. <strong>aureus</strong> is not yet complete.<br />

Membrane; 5<br />

Unknown; 47<br />

Cell wall; 19<br />

Extracellular; 63<br />

Cytosolic; 72<br />

Cytosolic<br />

Extracellular<br />

Cell wall<br />

Membrane<br />

Unknown<br />

63<br />

Figure 2. Predicted localization of extracellular<br />

proteins of 25 clinical S. <strong>aureus</strong> isolates. A total of<br />

206 distinct proteins were identified in the growth<br />

medium fractions of 25 clinical isolates. For the<br />

prediction of protein localization PSORT software<br />

was used.<br />

Interestingly, only seven out of 63 predicted Sec-dependent extracellular proteins were found<br />

to be produced by all 17 clonally different strains (i.e. Aly, IsaA, Lip, LytM, Nuc, SA0620,<br />

and SA2097). A further nine proteins (i.e. Aur, Geh, GlpQ, Hla, HlgB, SA0570, SA1812,<br />

SspA, SspB) were identified in at least 80% of these strains (Supplemental table IVc).<br />

Interestingly, four of the invariant proteins (i.e. IsaA, LytM, SA0620, and SA2097) were<br />

recently shown to be regulated by the WalK/WalR two-component system which is essential<br />

for cell viability and cell wall metabolism (Dubrac et al., 2007). IsaA and LytM share a<br />

conserved transglycosylase/muramidase domain, and SA0620 and SA2097 contain a CHAP<br />

amidase domain. All these proteins have an N-terminal cell wall-binding domain in common,<br />

indicating that they are exported and possibly non-covalently associated to the cell wall. It is<br />

worth noting in this context that IsaA-specific antibodies have previously been detected in<br />

healthy adults, colonized patients and patients suffering from sepsis, suggesting that IsaA is<br />

also expressed in vivo (Lorenz et al., 2000; Clarke et al., 2006). Most strikingly, the amounts<br />

of the invariant proteins varied significantly between individual strains (Figure 3).<br />

Expression heterogeneity probably triggered by varying SigB and SarA activities has been<br />

observed before for extracellular proteases (Karlsson and Arvidson, 2002). Notably, our<br />

present data indicate that this phenomenon is by no means restricted to proteases, but applies<br />

to secreted virulence factors in general.<br />

31 proteins were found to be unique for one or two strains under the conditions tested. While<br />

the functions of some of these proteins, such as Hlb, LukE, and SEB, in S. <strong>aureus</strong>-associated<br />

virulence are well characterized, the functions of other proteins are less clear and, in many<br />

cases, remain to be elucidated. Why were those proteins missing from the exoproteome of the<br />

remaining strains? <strong>The</strong>re are at least three possible explanations for this phenomenon: the<br />

respective genes (i) are absent, (ii) represent pseudo genes or (iii) are not expressed or<br />

expressed in very low amounts. <strong>The</strong> lack of detection of SED, SEK, SEP, SEQ, SER, Etd,<br />

24636604, and SAR0435 correlated with the absence of their coding genes. By contrast, while<br />

HlgA, SplD, SAR0422, SA0092, and SA1001 were encoded by at least 80% of the strains,<br />

these proteins were synthesized in detectable amounts only in one or two strains (Table 4).


Chapter 3<br />

Figure 3. Relative amounts of extracellular proteins detected in at least 80% of investigated S. <strong>aureus</strong> isolates.<br />

<strong>The</strong> respective sector on the 2D gel of each isolate is shown for the proteins indicated. <strong>The</strong> proteins were stained<br />

with colloidal Coomassie Brillant Blue as described for Figure 1.<br />

Table 4. Identification of genes for which gene products were detected in 20% of the<br />

investigated S. <strong>aureus</strong> isolates a<br />

Gene A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

sea + + + + + + + + + + + + + + + +<br />

seb + + + + + + + + +<br />

sed + +<br />

sek + + + +<br />

sep + + +<br />

seq + + + +<br />

ser + +<br />

etd + +<br />

hlgA + + + + + + + + + + + + + + + + + + + + + + + + +<br />

splD + + + + + + + + + + + + + + + + + + + + + + + + +<br />

splE + + + + + + + + + + +<br />

24636604 + +<br />

SAR0422 + + + + + + + + + + + + + + + + + + + + + + + + +<br />

SAR0435 +<br />

SAR1905 + + +<br />

SA1755 + + + + + + + + +<br />

SA0092 + + + + + + + + + + + + + + + + + + + + + + +<br />

SACOL0478 + + + + + +<br />

SA1001 + + + + + + + + + + + + + + + + + + + + + + +<br />

a genes whose gene product was detected on 2D gels are shaded grey for the respective strains<br />

64


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

Variations in the expression levels of virulence genes may relate to differential activities of<br />

specific S. <strong>aureus</strong> gene regulators. Especially RNAIII has previously been identified as a<br />

main regulator of virulence gene expression (Novick, 2003), and the loss of RNAIII was<br />

shown to affect the extracellular protein pattern of S. <strong>aureus</strong> dramatically (Ziebandt et al.,<br />

2004). To analyse RNAIII levels, we performed dot blot analyses using RNA prepared from<br />

cells grown to an OD540 of 10. This revealed that RNAIII was not produced at detectable<br />

levels in 11 isolates (C, D, G, H, I, J, K, T, W, X, Y) (Figure 4). With the exception of isolate<br />

W this correlates very well with a diminished expression level of late virulence factors<br />

(Figure 1), confirming the general importance of RNAIII in virulence gene expression. At the<br />

same time this shows that exceptions are possible as was the case for isolate W. Interestingly,<br />

these 11 isolates were involved in wound infection, arthritis or cholangitis, respectively<br />

(Supplemental table IVa). In mice, agr mutants were shown to have a growth advantage<br />

within a mixed population of S. <strong>aureus</strong> residing in abscesses and wounds (Schwan et al.,<br />

2003). Possibly, reduced RNAIII levels and the resulting decreased expression of RNAIIIdependent<br />

virulence genes might be correlated to the induction of some of the observed<br />

clinical symptoms. In the 14 remaining isolates, RNAIII transcripts have been detected in<br />

varying amounts, which may, at least in part, account for the observed differences in<br />

virulence factor production.<br />

Similar exoprotein patterns in closely related isolates<br />

While virulence gene expression of clonally different isolates was highly variable, we<br />

observed very similar protein expression patterns in closely related isolates (i.e. B398, E8,<br />

L80, G228, and X8) (Supplemental table IVa, Figure 1). Isolates L and M (=L80), which<br />

belong to the clonal group ST80, displayed all the genetic characteristics (pvl, hlgA, etd,<br />

edinB) as well as characteristic resistance patterns typically found in caMRSA strains with<br />

epidemic spread in the European population (Supplemental table IVa, Tables 2 and 3)<br />

(Monecke et al., 2006; Vandenesch et al., 2003). <strong>The</strong> isolates were derived from two patients<br />

suffering from panaritium and abscesses, respectively. Interestingly, the extracellular protein<br />

pattern of isolates L and M was identical, suggesting that the virulence gene expression<br />

pattern may be very stable over extended periods of time (April 2002 - April 2004) (Figure 1).<br />

<strong>The</strong> protein expression profiles of three other isolate pairs, A/B (=B398), E/F (=E8), and X/Y<br />

(=X8) were also very similar. <strong>The</strong> strain pair A/B was identified by MLST as sequence type<br />

398. This clonal lineage of S. <strong>aureus</strong> is frequently found on pigs in the Netherlands and was<br />

recently described to spread from animals to humans (Armand-Lefevre et al., 2005; van<br />

Belkum A. et al., 2008; Witte et al., 2007). In the present study the isolates came from a<br />

mother and her daughter and were involved in pneumonia and phlegmone, respectively. <strong>The</strong>y<br />

encode pvl but none of the known superantigens. As described for other isolates of this<br />

lineage, strain typing by PFGE failed, which was possibly due to the activity of a<br />

restriction/methylation system typical for this lineage (Bens et al., 2006). Notably, the two<br />

isolates of strains B398 differed with respect to mecA gene conservation (Supplemental table<br />

IVa), suggesting that the acquisition or loss of methicillin resistance had no significant<br />

influence on virulence gene expression (Figure 1). This difference in mecA conservation was<br />

also observed for the two X8 strains.<br />

As indicated above, one strain (G228), which was responsible for a hospital outbreak, was<br />

identified seven times in four different patients. <strong>The</strong> first isolate (G) was obtained at the end<br />

of 2003 and the most recent isolate (I) was identified early in 2005 as the colonizing S. <strong>aureus</strong><br />

65


Chapter 3<br />

strain of a patient who suffered from a wound infection induced by this strain (isolate H) one<br />

year before. We found the hlb converting phage FaSa3 in all these isolates, except for isolate<br />

G, as might be expected from the phage dynamics during infection (Goerke et al., 2006;<br />

Goerke et al., 2004). In accordance with this, the phenotype of strain G288 changed from Hlb<br />

positive (G) to Hlb negative (C, D, H, I, J, K) (Table 3). However, the prophage did not<br />

significantly alter the expression of other virulence-associated genes (Figure 1). This might be<br />

mainly due to the fact that RNAIII was not expressed in these isolates under our experimental<br />

conditions (Figure 4) and, accordingly, only a few virulence factors were produced.<br />

Figure 4. Transcription of RNAIII in different clinical S. <strong>aureus</strong> isolates. RNA was prepared from cells grown in<br />

TSB to an optical density of 10. Serial dilutions of total RNA of clinical isolates and reference strains (RN6390,<br />

COL, Newman) were blotted and crosslinked onto the same positively charged nylon membrane. <strong>The</strong> membranebound<br />

RNA was hybridized with a digoxigenin labelled RNA probe complementary to RNAIII. Chemiluminescence<br />

signals were detected with a LumiImager (Roche Diagnostics, Mannheim, Germany).<br />

Concluding remarks<br />

In conclusion, our data show that in the species S. <strong>aureus</strong>, genome plasticity is only one of<br />

several factors involved in exoproteome profile heterogeneity. Expression regulation as well<br />

as protein secretion and modification processes add further dimensions to the heterogeneity of<br />

the virulence potential of S. <strong>aureus</strong>. Such effects might be further enhanced and fine-tuned by<br />

promoter mutations, differential activities of regulatory molecules and translation regulation<br />

mechanisms. Most probably, the profoundly heterogeneous expression pattern of virulence<br />

genes observed under identical in vitro conditions reflects a very high degree of variability in<br />

vivo. It seems reasonable to suggest that these different virulence protein patterns are linked to<br />

different clinical symptoms in the host. We are currently comparing virulence gene<br />

expression profiles of S. <strong>aureus</strong> isolates that induced very similar symptoms. In this way, we<br />

hope to identify symptom-/disease-related proteomic signatures that may help in elucidating<br />

specific pathogenic mechanisms. It has been shown that S. <strong>aureus</strong> carriers mount a very<br />

selective and protective antibody response against superantigens of their colonizing strains<br />

(Holtfreter et al., 2006). Similarly, specific adaptive immune responses might also be raised<br />

against other virulence factors, which vary between strains to a similar extent. <strong>The</strong>se and<br />

other immune mechanisms might explain why S. <strong>aureus</strong> carriers with bacteremia generally<br />

have a better outcome than non-carriers (Wertheim et al., 2004). Finally, given the extensive<br />

variability of virulence factors and mechanisms in S. <strong>aureus</strong>, our study has important<br />

implications for vaccine development. <strong>The</strong> data strongly suggest that the development of<br />

protective vaccines will require a very careful selection and combination of bacterial antigens.<br />

66


Proteogenomics uncovers extreme heterogeneity in the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

exoproteome due to genomic plasticity and variant gene regulation<br />

Acknowledgements<br />

We are indebted to J. Ziebuhr for critical comments on the manuscript and S. Holtfreter and<br />

S. Kozitskaya for assistance in some experiments. Financial support was provided by CEU<br />

(StaphDynamics, LSHM-CT-2006-019064; BaSysBio, LSHG-CT-2006-037469), BMBF<br />

(031U107A/-207A; 031U213B), DFG (GK212/3-00, SFB/TR34, FOR 585), and Top Institute<br />

<strong>Pharma</strong> (T4-213).<br />

67


“Music is my religion”<br />

-Johnny A. Hendrix-<br />

68


Chapter 4<br />

Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

M.J.J.B. Sibbald, T. Winter, M. ten Brinke, G. Buist, D.G.A.M. Koedijk, M.M. van der Kooi-<br />

Pol, T. Msadek, O. Poupel, V. Rühmling, I. Stokroos, E. Tsompanidou,<br />

H. Antelmann, M. Hecker, S. Engelmann, J.M. van Dijl<br />

To be submitted<br />

69


Chapter 4<br />

Summary<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> is a dangerous human pathogen that can cause a range of<br />

diseases. <strong>The</strong> genome of this bacterium encodes an arsenal of virulence factors that are<br />

exported to extracytoplasmic locations. Once translocated across the membrane, these<br />

proteins are either retained at the surface of the cell or released into the environment. S.<br />

<strong>aureus</strong> contains several protein secretion pathways of which the general Sec pathway<br />

seems to be most frequently used. Other potential secretion pathways include the Twinarginine<br />

translocation (Tat), Com and Ess pathways. This study reports on the analysis<br />

of S. <strong>aureus</strong> secretion mutants. Sixteen isogenic S. <strong>aureus</strong> secretion mutants were<br />

created, and the influence of the respective mutations on the exoproteome composition<br />

was analyzed. Furthermore, the strains were tested for hemolysin production, and<br />

Caenorhabditis elegans killing. Several S. <strong>aureus</strong> mutants showed significantly altered<br />

exoproteomes, such as the secG or lgt mutants. In contrast, no obvious secretion defects<br />

were observed for the other mutant strains lacking factors required for protein<br />

translocation, precursor processing, and attachment of proteins to the cell wall. Notably,<br />

in various mutant strains second-site mutations were observed that led to the loss of<br />

hemolysin production, suggesting the acquisition of agr mutations. While this seems to<br />

be a consequence of the natural adaptive capabilities of S. <strong>aureus</strong>, it is also an important<br />

caveat for future studies on protein secretion in this organism that underscores the need<br />

for genetically stable model strains.<br />

70


Introduction<br />

Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> is part of the normal human microbiota. In its ecological niches such<br />

as the skin and the nose it is no major threat to the human host. About 20% of the population<br />

carries S. <strong>aureus</strong> permanently, ~60% can be a transient carrier and ~20% is apparently free of<br />

S. <strong>aureus</strong> (Peacock et al., 2001). However, S. <strong>aureus</strong> can also turn into a dangerous pathogen<br />

that can cause many different diseases, ranging from superficial skin lesions such as styes and<br />

furunculosis, to more serious infections such as pneumonia, urinary tract infections and<br />

endocarditis, and in rare cases even meningitis. Soon after the introduction of penicillin in the<br />

1960s, resistant S. <strong>aureus</strong> strains were isolated in hospitals and ever since multidrug resistant<br />

strains have been a major problem in hospitals. Vancomycin has long been a last resort<br />

antibiotic. However, in 1996 the first vancomycin intermediate resistant strain (VISA) was<br />

reported and soon after this date vancomycin resistant strains (VRSA) were isolated from<br />

hospital patients (Hiramatsu et al., 1997;Weigel et al., 2003).<br />

One of the prime reasons why S. <strong>aureus</strong> is able to infect almost every organ and tissue in the<br />

human body is that S. <strong>aureus</strong> cells can produce a very diverse arsenal of virulence factors that<br />

are exported to the cell surface and host milieu. <strong>The</strong>se factors include proteins that are<br />

necessary for: 1, adherence to cells (e.g. clumping factors, fibronectin and fibrinogen binding<br />

proteins); 2, invasion and spreading throughout the host (e.g. hyaluronidase, leukocidin); 3,<br />

evasion of the immune system (e.g. protein A and other IgG-binding proteins, capsuleforming<br />

proteins); and 4, damaging host cells, thereby contributing to the symptoms of septic<br />

shock (e.g. exotoxins, exfoliative toxin, toxic shock syndrome toxin TSST).<br />

Figure 1 Protein secretion pathways in S. <strong>aureus</strong>. (A) Schematic representation of a S. <strong>aureus</strong> cell with Sec, Tat,<br />

Com, and Ess pathways for protein secretion. Furthermore, potential protein secretion pathways involving ABC<br />

transporters or holins are shown. Proteins synthesized with an appropriate secretion signal are directed to one of<br />

the pathways and translocated across the membrane. <strong>The</strong> final destination of a translocated protein depends on<br />

various other features, such as the presence of a membrane anchor, a lipobox, or a cell wall binding motif. (B)<br />

Characteristics of signal peptides. Secretory signal peptides consist of three domains known as the N-, H-, and Cregions<br />

(Sibbald et al., 2006). Signal peptide cleavage by signal peptidase occurs in the C-region. <strong>The</strong> signal or<br />

leader peptides of bacteriocins do not contain a specific H-domain. Ess-dependent proteins lack a cleavable Nterminal<br />

signal peptide, but often contain a WXG-motif in the middle of the protein that may serve as a targeting<br />

signal.<br />

Bacteria have acquired several pathways to transport proteins to extracytoplasmic locations.<br />

Proteins that are translocated across the membrane of Gram-positive bacteria are either<br />

retained at the surface of the cell via special mechanisms or, if no retention signal is present,<br />

these proteins are secreted into the environment. In S. <strong>aureus</strong>, at least six different protein<br />

71


Chapter 4<br />

secretion pathways have been identified (Figure 1; (Sibbald et al., 2006)). <strong>The</strong> majority of<br />

secreted proteins are synthesized with a signal peptide, which directs them to a specific<br />

secretion pathway. During or after translocation across the membrane the signal peptide is<br />

cleaved from the mature protein by a signal peptidase. <strong>The</strong> protein is then folded into the right<br />

conformation (with or without the help of folding catalysts) and either released into the<br />

environment or retained at the surface of the cell. Most proteins (and virulence factors) are<br />

translocated across the membrane via the (general) Sec pathway. <strong>The</strong>se proteins are<br />

synthesized in the cytoplasm with a typical Sec-type signal peptide and then directed to the<br />

membrane-embedded Sec machinery. <strong>The</strong> most important components of the Sec machinery<br />

are the cytoplasmic translocation motor SecA and the membrane-embedded translocation<br />

channel, which is composed of SecY, SecE, and SecG (Hanada et al., 1996;Veenendaal et al.,<br />

2004). SecA binds and hydrolyzes ATP, and through the accompanying conformational<br />

changes it pushes the preprotein through the SecYEG channel. It has been shown for S.<br />

<strong>aureus</strong> that both SecA, SecE and SecY are essential for growth and viability (Chaudhuri et<br />

al., 2009). In addition, S. <strong>aureus</strong> has a second set of secA and secY genes in its genome (here<br />

referred to as secA2 and secY2). <strong>The</strong>se genes lie in an operon together with the gene for the<br />

only known SecA2-SecY2 substrate, SraP, and several other genes that are probably involved<br />

in the glycosylation of SraP (Siboo et al., 2005;Siboo et al., 2008). <strong>The</strong> secA2 secY2 operon is<br />

also found in several other pathogenic Gram-positive bacteria, such as Streptococcus<br />

gordonnii (Bensing and Sullam, 2002). SecDF is a protein that is associated with the Sec<br />

channel, together with a protein named YajC in E. coli or YrbF in B. subtilis (Tjalsma et al.,<br />

2000;Sibbald et al., 2006). In S. <strong>aureus</strong> and several other Gram-positive bacteria, the SecDF<br />

protein seems to be a natural fusion of homologues of the Escherichia coli SecD and SecF<br />

proteins, and it is thus regarded as a “Siamese Twin protein” (Bolhuis et al., 1998). <strong>The</strong><br />

secDF gene of Bacillus subtilis is not required for cell growth and viability, and its deletion<br />

has a relatively mild negative effect on protein translocation (Bolhuis et al., 1998). By<br />

contrast, secDF seems to be essential in S. <strong>aureus</strong> (Chaudhuri et al., 2009). So far, the precise<br />

role of SecDF in the translocation process has remained unclear.<br />

Type I signal peptidases remove signal peptides from secretory precursor proteins to liberate<br />

these proteins from the membrane upon translocation (van Roosmalen et al., 2004). <strong>The</strong><br />

genome of S. <strong>aureus</strong> contains two genes encoding for the signal peptidases SpsA and SpsB of<br />

which the latter is essential in S. <strong>aureus</strong> (Chaudhuri et al., 2009;Cregg et al., 1996). SpsA<br />

seems to be an inactive signal peptidase since the catalytic Ser and Lys residues are replaced<br />

with Asp and Ser residues, respectively (van Roosmalen et al., 2004;Dalbey et al., 1997).<br />

SpsB recognizes an -3 A-X-A -1 motif (with a few other amino acids permitted at the -3<br />

position) and cleaves after the Ala residue at the -1 position (Sibbald et al., 2006).<br />

It should be noted that the Sec pathway can only facilitate membrane passage of proteins in an<br />

unfolded state (Driessen and Nouwen, 2008;Papanikou et al., 2007;Yuan et al., 2009). While<br />

most translocated proteins seem to have an intrinsic capability to fold into their native<br />

conformation, their post-translocational folding in vivo is usually catalyzed by chaperones and<br />

other folding catalysts, such as DsbA and PrsA (Tjalsma et al., 2000;Tjalsma et al.,<br />

2004;Sibbald et al., 2006). PrsA is a general folding catalyst with peptidyl-prolyl cis/trans<br />

isomerase activity that was shown to be involved in the correct folding of AmyQ in B. subtilis<br />

(Kontinen and Sarvas, 1993) and the protective antigen of B. anthracis (Williams et al.,<br />

2003). DsbA is a homologue of the E. coli DsbA and B. subtilis BdbD proteins, which are<br />

both oxidases involved in the formation of disulfide bonds in exported proteins (Kouwen and<br />

van Dijl, 2009a). However, unlike its homologues in E. coli and B. subtilis, the S. <strong>aureus</strong><br />

72


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

DsbA does not seem to require a membrane-embedded partner protein for its re-oxidation<br />

during catalysis. Instead, this protein is re-oxidized by components in the extracellular<br />

environment (Heras et al., 2008;Kouwen et al., 2007).<br />

In addition to the Sec pathway, Gram-positive bacteria can use several other special purpose<br />

pathways for protein export from the cytoplasm. <strong>The</strong>se include the YidC pathway, the Twinarginine<br />

translocation (Tat) pathway, the Com pathway, secretion via holins or ABC<br />

transporters, and the Ess pathway (Sibbald et al., 2006;Driessen and Nouwen, 2008;Yuan et<br />

al., 2009).<br />

YidC functions as a membrane insertase to mediate membrane protein insertion either by<br />

itself or in concert with SecYEG. In the Sec pathway, YidC is linked to the Sec translocase<br />

via SecDF (see (Yuan et al., 2009)). <strong>The</strong> YidC homologues in B. subtilis have been<br />

designated SpoIIIJ and YqjG (Murakami et al., 2002;Tjalsma et al., 2003;Saller et al., 2009).<br />

<strong>The</strong> Tat translocase consists of the TatA and TatC subunits. This translocase recognizes a<br />

twin-arginine motif that consists of two adjacent (twin) Arg residues followed by another<br />

amino acid and two hydrophobic amino acids (R-R-x-φ-φ, where φ is a hydrophobic amino<br />

acid) (Cristóbal et al., 1999;Jongbloed et al., 2000). In contrast to the Sec pathway, this<br />

translocase is able to transport folded proteins across the membrane, which seems of<br />

particular relevance for the export of proteins with bound co-factors. <strong>The</strong> Tat pathway has<br />

been studied very well in E. coli and B. subtilis (Dilks et al., 2003;Berks et al.,<br />

2005;Robinson and Bolhuis, 2001), but fairly little is known about the roles of this pathway in<br />

S. <strong>aureus</strong>. Recently, the Tat pathway of S. <strong>aureus</strong> has been studied using biochemical and<br />

proteomics approaches, but this has led to the identification of only one substrate so far,<br />

which is FepB (Yamada et al., 2007;Schmaler et al., 2009).<br />

<strong>The</strong> Com pathway of B. subtilis is used for the export and assembly of pseudopili that are<br />

necessary for DNA binding and uptake during natural genetic competence development<br />

(Tjalsma et al., 2004;Tjalsma et al., 2000). Homologues of some of the components and the<br />

secreted proteins are also present in S. <strong>aureus</strong>. So far, limited information is available about<br />

this pathway in S. <strong>aureus</strong>. Morikawa and colleagues have shown that the expression of several<br />

competence genes is under control of the SigH factor (Morikawa et al., 2003), but whether<br />

these genes are involved in DNA uptake has so far remained unclear.<br />

For S. <strong>aureus</strong> it has been shown that at least three proteins (i.e. EsxA, EsxB and EsaC) are<br />

secreted via the ESX-1 or ESAT-6 secretion system (Ess) pathway (Burts et al., 2005;Burts et<br />

al., 2008). First discovered in Mycobacterium tuberculosis, this pathway seems to be<br />

conserved in several Gram-positive pathogens (Pallen, 2002). <strong>The</strong> EssA, EssB, and EssC<br />

components are required for a functional Ess secretion machinery. Notably, proteins secreted<br />

via the Ess pathway do not contain a classical signal peptide, but comparison of many of these<br />

secreted proteins revealed a conserved WXG motif in the middle of these proteins (Pallen,<br />

2002). Whether this motif represents an Ess targeting signal remains to be elucidated.<br />

S. <strong>aureus</strong> and related Gram-positive bacteria have several mechanisms for the retention of<br />

proteins that have been exported from the cytoplasm. One of these mechanisms is to bind the<br />

protein to the membrane via a lipid anchor. Such lipoproteins are synthesized with a Sec type<br />

signal peptide containing the so-called lipobox (von Heijne, 1989). <strong>The</strong> lipobox contains an<br />

invariant Cys residue at the +1 position and the consensus sequence -3 L-x-x-C +1 . <strong>The</strong> lipobox<br />

is recognized by the diacylglyceryl transferase Lgt and the invariant Cys is then modified by<br />

transferring a diacylglyceryl group from phosphatidylglycerol to the sulfhydryl group of this<br />

Cys residue (Tjalsma et al., 2000). Upon cleavage by the lipoprotein-specific signal peptidase<br />

II, the lipid-modified Cys residue serves as the membrane anchor for the mature lipoprotein. It<br />

73


Chapter 4<br />

has been shown that S. <strong>aureus</strong> Lgt is important for virulence (Stoll et al., 2005;Bubeck et al.,<br />

2006;Hashimoto et al., 2006), and that a mutant lacking the lgt gene is attenuated in the<br />

induction of an inflammatory response (Stoll et al., 2005). Likewise, Lgt is important for the<br />

virulence in other Gram-positive pathogens, such as Listeria monocytogenes (Baumgärtner et<br />

al., 2007;Machata et al., 2008). Paradoxically, it has also been reported than an lgt mutant of<br />

S. <strong>aureus</strong> can be hyper-virulent (Stoll et al., 2005;Bubeck et al., 2006). <strong>The</strong> molecular basis<br />

for these different observations is presently not clear.<br />

Another mechanism employed for the specific retention of proteins in the cell envelope is<br />

their covalent attachment to the cell wall. In S. <strong>aureus</strong> and many other Gram-positive<br />

pathogens, this reaction is catalyzed by sortases (Dramsi et al., 2005;Ton-That et al., 2004).<br />

Based on phylogenetic analyses of 61 sortases, at least four classes of sortase (A-D) can be<br />

distinguished (Dramsi et al., 2005). In the genome of S. <strong>aureus</strong> two genes encoding sortases<br />

are present. <strong>The</strong>se enzymes recognize proteins with a C-terminal signal, which consists of the<br />

recognition motif (LPxTG), followed by a stretch of hydrophobic residues and several<br />

positively charged residues (Arg or Lys). SrtA recognizes proteins with an LPxTG or LPxAG<br />

motif and cleaves N-terminally of the Gly residue. At the same time, SrtA couples the mature<br />

protein through its new C-terminal Thr or Ala residue to the peptidoglycan in the cell wall<br />

(Mazmanian et al., 2001;Navarre and Schneewind, 1994). In S. <strong>aureus</strong> 21 proteins have been<br />

identified with SrtA signals. For several of these proteins (e.g. protein A, ClfA, SasG) it has<br />

been shown that they are indeed attached to the cell wall by SrtA, and that they are involved<br />

in virulence (Corrigan et al., 2007;Josefsson et al., 2001). <strong>The</strong> gene encoding the second<br />

sortase of S. <strong>aureus</strong>, SrtB, lies in an operon that also encodes the SrtB substrate IsdC and<br />

several other genes involved in iron acquisition (Mazmanian et al., 2003). <strong>The</strong> structure of<br />

this operon is conserved in other Gram-positive pathogens, such as Bacillus anthracis,<br />

Bacillus cereus and L. monocytogenes (Dramsi et al., 2005). IsdC contains a C-terminal<br />

NPQTN SrtB recognition motif instead of the SrtA recognition motifs LPxTG or LPxAG<br />

(Mazmanian et al., 2002). <strong>The</strong>se different substrate specificities of SrtA and SrtB relate to<br />

differences in their substrate binding pockets (Bentley et al., 2007).<br />

In addition to covalent binding of proteins to the cell wall, there are also several proteins that<br />

bind to components of the cell wall via non-covalent interactions. This non-covalent cell wall<br />

binding can involve several different domains with repeated motifs, such as the LysM domain<br />

(Buist et al., 2008), the GW domain (Braun et al., 1997), or other domains as encountered in<br />

the clumping factors A and B (Hartford et al., 1997;Ní Eidhin et al., 1998) and the serine<br />

aspartate repeat proteins SdrC, SdrD, and SdrE (Josefsson et al., 1998).<br />

Most studies in the area of protein secretion by S. <strong>aureus</strong> were focused on the organism’s<br />

virulence factors and their modes of action. Relatively little attention has been attributed to<br />

the export and folding mechanisms for these virulence factors and other exported proteins. In<br />

this chapter, the construction and analysis of S. <strong>aureus</strong> protein secretion mutants is described.<br />

As such it represents a first step to define the functions of secretion machinery components of<br />

this pathogen in a systematic manner. Genes for several secretion machinery components<br />

were deleted from the chromosome (Table 1) and the exoproteomes of the resulting mutant<br />

strains were analyzed with functional assays, molecular biological approaches and<br />

proteomics.<br />

74


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

Table 1 Relevant components of S. <strong>aureus</strong> secretion pathways a<br />

Protein Essential b<br />

Sec-pathway<br />

Chaperone Ffh Y n.a.<br />

FtsY Y n.a.<br />

Translocation motor SecA1<br />

Y<br />

n.a.<br />

SecA2<br />

N<br />

N<br />

Translocation channel SecY1<br />

Y<br />

n.a.<br />

SecY2<br />

N<br />

Y<br />

SecE Y n.a.<br />

SecG N Y<br />

SecDF Y N<br />

YrbF ? n.a.<br />

Membrane insertase SpoIIIJ/YqjG Y N<br />

Lipid modification Lgt N Y<br />

Signal peptidase SpsA (inactive)<br />

N<br />

SpsB<br />

Y c<br />

Y<br />

N<br />

LspA N Y<br />

Folding catalyst PrsA N Y<br />

DsbA N Y<br />

Cell wall attachment SrtA N Y<br />

SrtB N Y<br />

Tat-pathway<br />

Pseudopilin pathway<br />

Holins<br />

Ess<br />

TatA<br />

TatC<br />

ComGA<br />

N<br />

Y<br />

ComGC<br />

N<br />

Y<br />

ComC N Y<br />

CidA N Y<br />

LrgA N Y<br />

EsaA<br />

EssA<br />

EssB<br />

EssC<br />

75<br />

N<br />

N<br />

Deletion Mutant<br />

a Table adapted from Sibbald et al. (2006); b Chaudhuri et al. (2009); c Cregg et al. (1996); for proteins marked<br />

“n.a.” we did not attempt to delete the corresponding genes, because they were known to be essential (ffh,<br />

ftsY, secA1, secY1, secE), or of seemingly minor relevance (yrbF).<br />

Material & Methods<br />

Bacterial strains and growth<br />

Strains and plasmids used in this study are listed in Table 2. E. coli and S. <strong>aureus</strong> strains were grown at<br />

37°C under aerobic conditions. E. coli strains were grown in Luria-Bertani broth (LB). Unless stated<br />

otherwise, S. <strong>aureus</strong> strains were grown at 37°C in tryptic soy broth (TSB) under vigorous shaking or<br />

on trypic soy agar (TSA) plates. Antibiotics for E. coli strains were added in the following final<br />

concentrations: ampicillin 100 µg/ml, kanamycin 20 µg/ml, and erythromycin 100 µg/ml. For S. <strong>aureus</strong><br />

the following final concentrations were used: kanamycin 20 µg/ml, and erythromycin 5 µg/ml. To<br />

monitor β-galactosidase activity in S. <strong>aureus</strong>, 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside (Xgal)<br />

was added to the plates at a final concentration of 80 µg/ml.<br />

N<br />

N<br />

N<br />

N<br />

Y<br />

Y<br />

N<br />

N<br />

N<br />

N


Chapter 4<br />

Table 2 Plasmids and bacterial strains used<br />

Plasmids Relevant Properties Reference<br />

TOPO Cloning vector pCR®-Blunt II-TOPO® vector; Km R Invitrogen<br />

technologies<br />

Life<br />

pMAD E. coli / S. <strong>aureus</strong> shuttle vector that is temperature-sensitive<br />

in S. <strong>aureus</strong> and contains the bgaB gene, Ery R , Amp R<br />

(Arnaud et al.,<br />

2004)<br />

pDG783 1.5-kb kanamycin resistance cassette in pSB118; Amp R , Km R (Guérout-Fleury<br />

et al., 1995)<br />

“cidA”::kan-pMAD pMAD plasmid containing the flanking regions of S. <strong>aureus</strong><br />

cidA, also contains the kanamycin gene, Ery R , Kan R<br />

This work<br />

pMU<strong>TI</strong>N4 Insertion vector with Pspac promoter, Ery R (Vagner<br />

1998)<br />

et al.,<br />

“comC”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> comC, Ery R This work<br />

“comGA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> comGA, Ery R This work<br />

“comGC”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> comGC, Ery R This work<br />

“dsbA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> dsbA, Ery R This work<br />

“lgt”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> lgt, Ery R This work<br />

“lrgA”::kan-pMAD pMAD with flanking regions of S. <strong>aureus</strong> lrgA, also contains<br />

the kanamycin gene, Ery R , Kan R<br />

This work<br />

“lspA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> lspA, Ery R This work<br />

“prsA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> prsA, Ery R This work<br />

“secDF”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> secDF, Ery R This work<br />

“secA2”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> secA2, Ery R This work<br />

“secG”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> secG, Ery R This work<br />

“secY2”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> secY2, Ery R This work<br />

“spsA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> spsA, Ery R This work<br />

“spsB”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> spsB, Ery R This work<br />

“srtA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> srtA, Ery R This work<br />

“srtB”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> srtB, Ery R This work<br />

“tatA”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> tatA, Ery R This work<br />

“tatC”::kan-pMAD pMAD with flanking regions of S. <strong>aureus</strong> tatC, also contains<br />

the kanamycin gene, Ery R , Km R<br />

This work<br />

“tatAC”-pMAD pMAD with flanking regions of S. <strong>aureus</strong> tatAC, Ery R This work<br />

secA2-pMU<strong>TI</strong>N4 Controlled expression of secA2 This work<br />

Strains<br />

E. coli<br />

Genotype Reference<br />

DH5α supE44; hsdR17; recA1; gyrA96; thi-1; relA1 (Hanahan, 1983)<br />

TOP10 Cloning host for TOPO vector; F - mcrA ∆(mrr-hsdRMSmcrBC)<br />

Φ80lacZ∆M15 ∆lacX74 recA1 araD139 ∆(araleu)7697<br />

galU galK rpsL (Str R ) endA1 nupG<br />

S. <strong>aureus</strong> RN4220<br />

Parental strain Restriction-deficient derivative of NCTC 8325, cured of all<br />

known prophages, rsbU-, agr-<br />

76<br />

Invitrogen Life<br />

technologies<br />

∆cidA::kan<br />

(Kreiswirth et al.,<br />

1983)<br />

b<br />

cidA, Kan R This work<br />

∆comC b comC This work<br />

∆comGA b comGA This work<br />

∆comGC b comGC This work<br />

∆dsbA b dsbA This work<br />

∆lgt b lgt This work<br />

∆lrgA::kan lrgA, Km R This work<br />

∆lspA b lspA This work<br />

∆prsA b prsA This work<br />

∆secG b secG This work<br />

∆secY2 b secY2 This work<br />

∆spsA b spsA This work


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

Strains<br />

S. <strong>aureus</strong> RN4220<br />

Genotype Reference<br />

∆srtA b srtA This work<br />

∆srtB b srtB This work<br />

∆tatA tatA This work<br />

∆tatC::kan tatC, Km R This work<br />

∆tatAC b S. <strong>aureus</strong> SH1000<br />

tatA tatC This work<br />

Parental strain rsbU+ derivative of 8325-4 (rsbU+, agr+) (Horsburgh et al.,<br />

2002)<br />

sasG-pMU<strong>TI</strong>N4 Overexpression of SasG from the pMU<strong>TI</strong>N4 plasmid; Ery R (Corrigan et al.,<br />

2007)<br />

sasG-pMU<strong>TI</strong>N4 ∆srtA Overexpression of SasG from the pMU<strong>TI</strong>N4 plasmid; srtA;<br />

Ery R<br />

This Work<br />

Construction of mutant strains<br />

Mutants were constructed with the temperature-sensitive pMAD plasmid as described by Arnaud and<br />

colleagues (Arnaud et al., 2004) (Figure 2). Primers used to PCR amplify the flanking regions of genes<br />

of interest are listed in Table 3. <strong>The</strong> flanking regions of ~500 bp were obtained by PCR using the<br />

primer pairs F1/R1 and F2/R2.<br />

Figure 2 Schematic representation of the pMAD-based strategy for deletion of genes from the S. <strong>aureus</strong><br />

chromosome. <strong>The</strong> flanking regions upstream and downstream of a target gene that is to be deleted are represented<br />

as grey boxes marked “front” and “back”. <strong>The</strong>se front and back regions are PCR-amplified, merged by PCR, and<br />

cloned in the plasmid pMAD. Integration of pMAD with the merged front and back regions (step 1) can occur via<br />

single cross-over recombination upstream or downstream of the target gene (the diagram shows integration<br />

upstream of the target gene). A second recombination event (step 2), leading to pMAD excision from the<br />

chromosome, can take place either at the front or back region. Depending on the place where this second<br />

recombination event occurs, the target gene will either be excised from the chromosome (right), or remain intact<br />

on the chromosome (left).<br />

77


Chapter 4<br />

Table 3 Primers used in this study<br />

Primer Sequence (5’→3’)<br />

cidA-F1 AATAAAGACTTTTACTTGAAT<br />

cidA-R1 a TTAGAACTCCAATTCACCCATGGCCCCCGCGCCATCCCTTTCTAAATA<br />

cidA-F2 a CCGCAACTGTCCATACCATGGCCCCCTGATTACGTGCAAGCCTTATTAAT<br />

cidA-R2 GATAGAAGATTCAAATCTTCC<br />

comC-F1 CGAGATGGTCAAACATTTAAG<br />

comC-R1 TCACGTCAGTCAGTCACCATGGCAATGACAACCTCCTTATGTAAA<br />

comC-F2 TGCCATGGTGACTGACTGACGTGAAAATTAAAGAAATGGTAA<br />

comC-R2 AACTGCGATGATTGCATTGGC<br />

comGA-F1 ATATCGGAGCAGTCGATGATA<br />

comGA-R1 TTACGTCAGTCAGTCACCATGGCAAAAAACACCTCCTACATA<br />

comGA-F2 TGCCATGGTGACTGACTGACGTAAACTACATTCTAAGAAGCG<br />

comGA-R2 GAGCATTACTACAATTATAGT<br />

comGC-F1 GCTCAATAAGATAAACTTTGT<br />

comGC-R1 CTACGTCAGTCAGTCACCATGGCAATATTAACCTCCATTATTTTA<br />

comGC-F2 TGCCATGGTGACTGACTGACGTAGAAAGCAGTCAGCATTTAC<br />

comGC-R2 GATTCATCATTGGTATCAATA<br />

dsbA-F1 ATTTCTTTGGATATTTATATT<br />

dsbA-R1 CTACGTCAGTCAGTCACCATGGCAAATAACTCCTATTCATAT<br />

dsbA-F2 TGCCATGGTGACTGACTGACGTAGTCTTAATTGTTGAGATCA<br />

dsbA-R2 CTTTCGTTATAGTTTTCCCAC<br />

lgt-F1 GGTGTTGGTGTACTAATTACC<br />

lgt-R1 CTACGTCAGTCAGTCACCATGGCATCAACCTACTCCTCACTCTTA<br />

lgt-F2 TGCCATGGTGACTGACTGACGTAGTGATAGTTTGAGGAAATTTTT<br />

lgt-R2 ACATTATTATTCTTTTGCGCC<br />

lrgA-F1 TAAAGCCAAAGATGATAATAA<br />

lrgA-R1 a TTAGAACTCCAATTCACCCATGGCCCCCGCCTCCTACGTTTGATTTAA<br />

lrgA-F2 a CCGCAACTGTCCATACCATGGCCCCCTAACCACTTAGCACTAAACACACC<br />

lrgA-R2 GTAATTCGGAAAAGCTTTAAG<br />

lspA-F1 CCAATTAAGTGTAGACGATTC<br />

lspA-R1 TTACGTCAGTCAGTCACCATGGCATTTCGTTCCTCCAATCAATCG<br />

lspA-F2 TGCCATGGTGACTGACTGACGTAATGGAGACTTATGAATTTAACA<br />

lspA-R2 CGATATATTTTCTTTTAACAG<br />

prsA-F1 GAAAATGGCTTATATTCTATA<br />

prsA-R1 TTACGTCAGTCAGTCACCATGGCAAGTTGAAACTCCTTTGTAAGT<br />

prsA-F2 TGCCATGGTGACTGACTGACGTAACACAAAACCGAGCGACCGTGG<br />

prsA-R2 TTTGTTATATAGTGGTATTAT<br />

secA2-F1 GTATAAAAGCATGCGGGTGAC<br />

secA2-R1 a TTAGAACTCCAATTCACCCATGGCCCCCTTACTTCCCCACCATTCAGTT<br />

secA2-F2 a CGCAACTGTCCATACCATGGCCCCCTAAATGAAAAGGGGTAGCGCATGA<br />

secA2-R2 GTCGCATATATAATTTCGCTT<br />

secDF-F1 TTTTGCGGTTATGTATTTCTT<br />

secDF-R1 a TTAGAACTCCAATTCACCCATGGCCCCC TGAACACCTCATTATTTACG<br />

secDF-F2 a CCGCAACTGTCCATACCATGGCCCCCTAAAATGAATTAAGCGGTATGTGA<br />

secDF-R2 ATCACTAAAATTGTAGTTGCG<br />

secG-F1 TTAAAACAGGACGCTTTATTG<br />

secG-R1 TTACGTCAGTCAGTCACCATGGCAAAATTGTCCTCCGTTCCTTAT<br />

secG-F2 TGCCATGGTGACTGACTGACGTAAGGTCCGGCGATGTAAATGTCG-<br />

secG-R2 GCGTGCATATTCTAAAAAGCC<br />

secY2-F1 TGTCTGGTTCACAAAGCATTT<br />

secY2-R1 TTACGTCAGTCAGTCACCATGGCAGTTGCACCTCTTTTATATCAA<br />

secY2-F2 TGCCATGGTGACTGACTGACGTAAGGAGGTAATTATGAAATACTT<br />

secY2-R2 GCCTCTCCCTGATCATCAAAA<br />

78


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

Primer Sequence (5’→3’)<br />

spsA-F1 TAGAGCTATAATTCCAGTATT<br />

spsA-R1 TTACGTCAGTCAGTCACCATGGCAGATGTCACTCCTTTTTCGATC<br />

spsA-F2 TGCCATGGTGACTGACTGACGTAAAAAGAGGTGTCAAAATTGAAA<br />

spsA-R2 CCAACAATTTGGTCTTCATCA<br />

spsB-F1 ATTTGATTTCATTGATACTTG<br />

spsB-R1 a TTAGAACTCCAATTCACCCATGGCCCCCTCTTTTTAAGATTTGAACTG<br />

spsB-F2 a CCGCAACTGTCCATACCATGGCCCCCTAATATGAAACAAATACAACATCG<br />

spsB-R2 CCCATAATATTTTGCTTGTGA<br />

srtA-F1 AATGGTGTAGTAATTGACTAG<br />

srtA-R1 TTACGTCAGTCAGTCACCATGGCAACGTTAAGGCTCCTTTTATAC<br />

srtA-F2 TGCCATGGTGACTGACTGACGTAATCTATTACGCTAATGGATGAA-<br />

srtA-R2 CTCACATTACTTACTATTAAT<br />

srtB-F1 TGAAAATATGGAGCGACGTAT<br />

srtB-R1 TTACGTCAGTCAGTCACCATGGCAAAAAATCCTCTTTTATTAACG<br />

srtB-F2 TGCCATGGTGACTGACTGACGTAAACAGAAAAGAGGATAATTATG<br />

srtB-R2 ATCAAAATGATATAATTGATG<br />

tatA-F1 TTATGGCATTTACATTATCTG<br />

tatA-R1 CTACGTCAGTCAGTCACCATGGCAGATAATCAACCTCACTCATAA<br />

tatA-F2 TGCCATGGTGACTGACTGACGTAGCACTGACCACACCTTACTGGT<br />

tatA-R2 GACCCATAAATAATATTGGTA<br />

tatC-F1 TGATGAAATGGCTGAAGCTGG<br />

tatC-R1 a TTAGAACTCCAATTCACCCATGGCCCCCAAAATTTTTACTAACCGATG<br />

tatC-F2 a CCGCAACTGTCCATACCATGGCCCCCTAACCTTATACGAATCAATGCTGT<br />

tatC-R2 CGATTAGTAATGGTAATTTGG<br />

kan-F1 GGGGGCCATGGGTGAATTGGAGTTCGTCTTG<br />

kan-R1 GGGGGCCATGGTATGGACAGTTGCGGATGTA<br />

secA2-F3 CGGAATTCGAATCCAGTACGATTTTTAG<br />

secA2-R3 CGGGATCCTCCCGGTAACATACGACCTG<br />

RNAIII-F AGGAAGGAGTGATTTCAATG<br />

RNAIII-R CTAATACGACTCACTATAGGGAGAACTCATCCCTTCTTCATTAC<br />

Overlapping nucleotides are shown in bold; restriction sites in primers are underlined<br />

a <strong>The</strong>se primers have an overlap with the kanamycin resistance cassette from pDG783<br />

Primers R1 and F2 contained an overlap of 21 nucleotides (seven codons), which served to fuse the<br />

amplified “front” and “back” flanking regions by PCR. For the deletion of some genes (cidA and lrgA),<br />

a kanamycin resistance cassette was introduced between the respective amplified flanking regions. To<br />

this end, overlap was created between the 5’ and 3’sequences of the kanamycin resistance cassette and<br />

the front and back regions. <strong>The</strong> kanamycin resistance cassette was obtained by PCR using plasmid<br />

pDG783 as a template. In the case of the tatAC mutant, the upstream region of tatA was linked with the<br />

downstream region of tatC. <strong>The</strong> linked fragments were purified either from gel or using the High Pure<br />

PCR Product Purification Kit (Roche). Next, they were cloned in the TOPO vector. <strong>The</strong> constructs thus<br />

obtained were cut with EcoRI and the merged flanking regions were ligated to pMAD cut with EcoRI.<br />

Upon transformation of E. coli, cells containing pMAD plasmids with the appropriate merged flanking<br />

regions were identified through colony PCR with specific primers. Correct “gene”-pMAD constructs<br />

were used for electro-transformation of competent S. <strong>aureus</strong> RN4220 cells. Transformants were<br />

selected at 30°C on TSA plates containing erythromycin and X-gal. A blue colony was transferred to<br />

10 ml Brain Heart Infusion (BHI) and grown at 30°C without shaking. From the overnight culture 100<br />

µl was transferred to 10 ml fresh BHI, grown for one hour at 30°C and then transferred to 42°C and<br />

grown for six hours without shaking. Dilutions of the culture were transferred to plates containing<br />

erythromycin and X-GAL and grown for 48 hours at 42°C. A blue colony was transferred to 10 ml BHI<br />

broth and grown at 42°C. From the overnight culture 10 µl was transferred to 10 ml fresh BHI and<br />

growth was continued at 30°C for 6 hours. Dilutions of the culture were transferred to TSA plates<br />

79


Chapter 4<br />

containing X-gal and incubated for 48 hours at 30°C. White colonies were selected and tested for<br />

erythromycin sensitivity on TSA plates containing X-gal with or without erythromycin. Colonies were<br />

screened for particular gene deletions by colony PCR. Genomic DNA of seemingly correct mutants<br />

was isolated using the GenElute Bacterial Genomic DNA Kit (Sigma) for further verification of gene<br />

deletions by PCR. To delete particular genes from the S. <strong>aureus</strong> SH1000 genome, the respective pMAD<br />

constructs were transferred from the S. <strong>aureus</strong> RN4220 strain to the SH1000 strain by transduction with<br />

phage φ85 (Novick, 1991).<br />

To place the secA2 gene under control of the IPTG-inducible Pspac promoter, a 600 bp fragment,<br />

starting before the ribosome-binding site of secA2, was cloned into the pMU<strong>TI</strong>N4 plasmid using the<br />

EcoRI and BamHI restriction sites. This construct was then introduced into S. <strong>aureus</strong> RN4220 by<br />

electro-transformation.<br />

DNA sequence analyses were performed at ServiceXS, Leiden, the Netherlands.<br />

Scanning electron microscopy<br />

For scanning electron microscopy, bacteria were fixed for 30 min with 2% glutaraldehyde in 0.1 M<br />

Cacodylate buffer, pH 7.38. <strong>The</strong> fixated bacteria were placed on a piece (1 cm 2 ) of cleaved 0.1% Poly-<br />

L Lysine coated mica sheet and washed in 0.1 M Cacodylate buffer. This specimen was dehydrated in<br />

ethanol series consisting of 30%, 50%, 70%, 96% and anhydrous 100% solution (3X) 10 min each,<br />

then Critical point dried with CO2, and sputter-coated with 2-3 nm Au/Pd (Balzers coater). <strong>The</strong><br />

specimen was fixed on a SEM-stub-holder and observed in a JEOL FE-SEM 6301F microscope.<br />

RNA-isolation and dot blot analysis<br />

Total RNA from S. <strong>aureus</strong> was isolated using the acid-phenol method with some modifications (Fuchs<br />

et al., 2007). A digoxigenin-labeled RNA probe for RNAIII was prepared by in vitro transcription with<br />

T7 RNA polymerase by using a PCR fragment as template. <strong>The</strong> PCR fragment was generated by using<br />

chromosomal DNA of S. <strong>aureus</strong> and the respective oligonucleotides (Table 3). Dot blot analyses were<br />

carried out by using serial dilutions of total RNA prepared from S. <strong>aureus</strong> isolates and reference strains<br />

grown in TSB medium to an optical density (OD540) of 10. <strong>The</strong> digoxigenin-labeled RNA probe was<br />

used for hybridization. <strong>The</strong> hybridization signals were detected using a Lumi-Imager and analyzed<br />

using the Lumi-Analyst software package (Roche Diagnostics, Mannheim, Germany).<br />

Hemolysin activity<br />

To test the hemolysin activity, blood agar plates containing 5% sheep blood (Mediaproducts B.V.)<br />

were inoculated with overnight cultures and incubated for 24 hours at 37°C.<br />

Cell fractionation, SDS-PAGE, and Western blotting<br />

Overnight cultures of S. <strong>aureus</strong> strains were diluted in TSB to an OD540 of 0.05 and grown at 37°C<br />

under aerobic conditions. To monitor growth, samples were taken every hour and the OD540 was<br />

measured. Samples for subsequent experiments were taken after six hours. Cells were separated from<br />

the medium by centrifugation (2 min, 14.000 rpm). Proteins in the medium fraction were precipitated<br />

with 10% trichloroacetic acid (TCA), washed with acetone, and dissolved in 100 µl 1x Loading Buffer<br />

(Invitrogen). Cells were resuspended in 300 µl 1x Loading Buffer (Invitrogen) and disrupted with glass<br />

beads using the Precellys ® 24 bead beating homogenizer (Bertin Technologies). Non-covalently cell<br />

wall bound proteins were obtained using KSCN treatment. Cells from 20 ml cultures were collected by<br />

centrifugation (5 min, 6.000 rpm), washed with 5 ml PBS, and incubated for 10 min with 1,5 ml 1M<br />

KSCN on ice. After centrifugation (15 min, 8000 rpm) the non-covalently cell wall bound proteins<br />

were precipitated from the supernatant fraction with 10% TCA, washed with acetone and dissolved in<br />

100 µl 1x Loading Buffer (Invitrogen). Upon addition of Reducing Agent (Invitrogen), the samples<br />

were incubated at 95ºC for 5 min. Proteins were separated by SDS-PAGE using precast NuPage gels<br />

(Invitrogen). <strong>The</strong> amounts of protein used for SDS-PAGE were corrected for the OD540 of the<br />

respective culture. Coomassie staining of gels was performed using the SimplyBlue TM SafeStain<br />

80


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

(Invitrogen). For Western blotting, proteins separated by SDS-PAGE were blotted onto a nitrocellulose<br />

membrane (Protran®, Schleicher & Schuell). Immunodetection of particular proteins was performed<br />

with specific rabbit antibodies raised against TrxA (Miller, submitted) or DsbA (Kouwen et al., 2007),<br />

or mouse antibodies against SspB (Shaw et al., 2004). Bound primary antibodies were visualized using<br />

fluorescent IgG secondary antibodies (IRDye 680 or 800 CW goat anti-mouse/anti-rabbit from LiCor<br />

Biosciences). Membranes were scanned for fluorescence at 700 or 800 nm using the Odyssey Infrared<br />

Imaging System (LiCor Biosciences).<br />

Zymography<br />

Zymography was used to test whether the secretion of staphylococcal cell wall hydrolases was affected<br />

in mutant strains. Micrococcus luteus cell wall fragments were from Sigma. S. <strong>aureus</strong> RN4220 cell<br />

wall fragments were isolated as described previously (Steen et al., 2003) with minor adaptations. After<br />

breaking the cells by bead-beating, the cell walls were boiled in 4% SDS for 30 min and washed with<br />

PBS. This procedure was repeated three times. 12.5% PAA gels contained 10% cell wall fragments<br />

from either M. luteus or S. <strong>aureus</strong> RN4220. All sample amounts were corrected for OD540 of the<br />

respective cultures. Upon electrophoresis, the gels were incubated overnight at room temperature in a<br />

buffer containing 25 mM Tris (pH 8.0) and 1% Triton X-100. After washing with water, the gels were<br />

stained with a 1% methylene blue solution in potassium hydroxide (Valence and Lortal, 1995).<br />

Analytical and preparative two-dimensional (2-D) PAGE<br />

Extracellular proteins from 100 ml culture supernatant were precipitated, washed, dried, and resolved<br />

as described previously (Ziebandt et al., 2004). <strong>The</strong> protein concentration was determined using Roti®-<br />

Nanoquant (Carl Roth GmbH & Co, Karlsruhe, Germany). Preparative 2-D PAGE was performed by<br />

using the immobilized pH gradient technique (Bernhardt et al., 1999;Eymann et al., 2004). <strong>The</strong> protein<br />

samples (350 µg) were separated on immobilized pH gradient strips (Amersham <strong>Pharma</strong>cia Biotech,<br />

Piscataway, NJ) with a pH range of 3-10. <strong>The</strong> resulting protein gels were stained with colloidal<br />

Coomassie Blue G-250G (Candiano et al., 2004) and scanned with the light scanner. Each experiment<br />

was performed at least three times.<br />

For identification of proteins by MALDI-TOF MS, Coomassie-stained protein spots were excised from<br />

gels using a spot cutter (Proteome WorkTM) with a picker head of 2 mm and transferred into 96-well<br />

microtiter plates. Digestion with trypsin and subsequent spotting of peptide solutions onto the MALDI<br />

targets were performed automatically in an Ettan Spot Handling Workstation (GE-Healthcare, Little<br />

Chalfont, United Kingdom) using a modified standard protocol. MALDI-TOF MS analyses of spotted<br />

peptide solutions were carried out on a Proteome-Analyzer 4700/4800 (Applied Biosystems, Foster<br />

City, CA) as described previously (Eymann et al., 2004). MALDI-TOF-TOF analysis was performed<br />

for the three highest peaks of the TOF spectrum as described previously (Eymann et al., 2004;Wolff et<br />

al., 2007). Database searches were performed using the GPS explorer software version 3.6 (build 329)<br />

with the organism-specific databases.<br />

By using the MASCOT search engine version 2.1.0.4. (Matrix Science, London, UK) the combined MS<br />

and MS/MS peak lists for each protein spot were searched against a database containing protein<br />

sequences derived from the genome sequences of S. <strong>aureus</strong> NCTC8325. Search parameters were as<br />

described previously (Wolff et al., 2007). For comparison of protein spot volumes, the Delta 2D<br />

software package was used (Decodon GmbH Germany). <strong>The</strong> induction ratio of mutant to parental<br />

strain was calculated for each spot (normalized intensity of a spot on the mutant image/normalized<br />

intensity of the corresponding spot on the parental image). <strong>The</strong> significance of spot volume differences<br />

of two-fold or higher was assessed by the Student´s t test (α


Chapter 4<br />

assay, 25 L4-stage nematodes were transferred to the plate containing the S. <strong>aureus</strong> spots, and each<br />

assay was performed in triplicate. <strong>The</strong> plates were incubated at 25°C and the numbers of living<br />

nematodes were counted at 24 hour intervals. Nematodes that were not moving after plate tapping or<br />

gentle touching with a platinum wire were counted as dead. Statistical analysis of nematode survival<br />

was performed using the StatView program version 5.0 (SAS Institute Inc.) to create the cumulative<br />

survival plots by the Kaplan-Meier method.<br />

Results and Discussion<br />

Construction of an S. <strong>aureus</strong> secretion mutant collection<br />

Judged by previous studies in organisms like E. coli and B. subtilis, at least 30 proteins can<br />

fulfill potential roles in protein secretion by S. <strong>aureus</strong> (Table 1). Of these 30 proteins, the Ffh,<br />

FtsY, SecA1, SecY1, SecE and SpsB proteins are known to be essential for growth of E. coli,<br />

B. subtilis and most likely also S. <strong>aureus</strong> (Ji et al., 2001;Cregg et al., 1996;Sibbald et al.,<br />

2006). Furthermore, in these organisms the role of YrbF in protein secretion seemed so far of<br />

minor importance. We therefore focused our attention on the potential roles of the remaining<br />

23 proteins in protein secretion. To this purpose, we tried to delete the corresponding genes<br />

with the pMAD chromosomal integration-excision system (Figure 2). In this manner, we were<br />

able to completely delete the secY2, secG, lgt, spsA, lspA, prsA, dsbA, srtA, srtB, tatA, tatC,<br />

comGA, comGC, comC, cidA, or lrgA genes from the S. <strong>aureus</strong> RN4220 chromosome. In<br />

contrast, several attempts to delete the secA2, secDF, spoIIIJ/yqjG, esaA, essA, essB and essC<br />

genes remained unsuccessful, as was the case for a control experiment in which we tried to<br />

delete spsB. <strong>The</strong> observation that secDF and spoIIIJ/yqjG could not be deleted is consistent<br />

with the results of a recent Transposon-Mediated Differential Hybridisation (TMDH) analysis<br />

in which about a million transposon mutants were screened using a microarray approach<br />

(Chaudhuri et al., 2009). This analysis revealed 351 S. <strong>aureus</strong> genes that are of major<br />

importance for growth and cell viability, among which the secDF and spoIIIJ/yqjG genes.<br />

This finding points towards an important difference in the secretion machinery of B. subtilis<br />

and S. <strong>aureus</strong> since secDF is completely dispensable for growth, cell viability and protein<br />

secretion in B. subtilis (Bolhuis et al., 1998), and the same is true for the individual spoIIIJ<br />

and yqjG genes (Tjalsma et al., 2003). However, consistent with the situation in S. <strong>aureus</strong>, the<br />

B. subtilis spoIIIJ and yqjG genes cannot be deleted simultaneously, which shows that YidC<br />

function is essential for this organism (Tjalsma et al., 2003). Furthermore, the studies by<br />

Chaudhuri et al. confirmed the essentiality of the ffh, ftsY, secA1, secE, secY1 and spsB genes<br />

(Chaudhuri et al., 2009). Interestingly, the secA2, esaA, essA, essB and essC genes were not<br />

identified as potentially essential in the TMDH analysis. Studies by Burts et al. (Burts et al.,<br />

2005;Burts et al., 2008) have shown that the esaA, essA, essB and essC genes can be mutated,<br />

which suggests that our attempts to delete these genes were unsuccessful due to an unknown<br />

technical problem. However, we can presently not exclude the possibility that these genes are<br />

essential in S. <strong>aureus</strong> RN4220 while being dispensable in other strains. Direct evidence that<br />

secA2 is dispensable for S. <strong>aureus</strong> was provided by Siboo et al. (Siboo et al., 2008), who<br />

reported the deletion of the secA2 gene from S. <strong>aureus</strong> ISP479C. This suggests that secA2 is<br />

either essential in S. <strong>aureus</strong> RN4220 or that we were unlucky in our attempts to delete this<br />

gene. As a final approach, we therefore attempted to deplete the cells of SecA2 by replacing<br />

the original promoter sequences with the Pspac promoter through a single cross-over<br />

integration of plasmid pMU<strong>TI</strong>N4 in front of the secA2 gene. As can be expected for a non-<br />

82


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

essential gene, cells with the correctly integrated pMU<strong>TI</strong>N4 construct were able to grow in<br />

the absence of IPTG.<br />

Growth properties of S. <strong>aureus</strong> secretion mutants<br />

For all obtained S. <strong>aureus</strong> secretion mutants, the growth in TSB under vigorous shaking at<br />

37°C was monitored through optical density readings at 540 nm (OD540). All mutants derived<br />

from strain RN4220 displayed highly comparable growth rates, reaching OD540 values of ~15<br />

in the post-exponential growth phase. Consistent with this observation, the cells of these<br />

secretion mutants showed no detectable morphological differences (Figure 3).<br />

Figure 3 Scanning electron microscopy of S. <strong>aureus</strong> secretion mutants. S. <strong>aureus</strong> RN4220-derived secretion<br />

mutants were grown in TSB and fixated for scanning electron microscopy as described in the Materials and<br />

Methods section. (A) S. <strong>aureus</strong> RN4220 parental strain, (B) ∆comGA, (C) ∆lgt, (D) ∆secG, (E) ∆srtA, and (F)<br />

∆tatAC.<br />

Remarkably, the comC, comGA, comGC, and prsA mutants derived from S. <strong>aureus</strong> SH1000<br />

reached significantly higher OD540 values in the post-exponential growth phase (OD540 of<br />

~15-20) than the parental strain SH1000 or the corresponding dsbA, lgt, or lsp mutants (OD540<br />

of ~8-10; Figure 4A). Furthermore, some obtained secY2 mutants in S. <strong>aureus</strong> SH1000 were<br />

able to grow to high density, whereas others reached optical densities comparable to that of<br />

the parental strain (Figure 4A). Upon close inspection, we observed that also different isolates<br />

of S. <strong>aureus</strong> SH1000 were able to reach high OD540 values of ~15-20 (Figure 4A), whereas<br />

this was not the case for the strain as it was originally obtained from Dr. Simon Foster<br />

(University of Sheffield, UK).<br />

83


Chapter 4<br />

D<br />

O<br />

540<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 5 10 15 20 25<br />

Time (h)<br />

SH1000+<br />

SH1000-<br />

secY2+<br />

secY2-<br />

comGA<br />

dsbA<br />

lgt<br />

prsA<br />

Figure 4 Growth properties of mutant S. <strong>aureus</strong> strains and RNAIII production. (A) Growth curves of several<br />

different S. <strong>aureus</strong> SH1000 secretion mutants. Strains were grown in TSB at 37 o C under vigorous shaking and the<br />

OD 540 was measured at hourly intervals. <strong>The</strong> S. <strong>aureus</strong> SH1000+ and ∆secY2+ strains show hemolysin activity on<br />

blood agar plates, while the SH1000- and ∆secY2- strains do not show this feature. (B) RNA was prepared from S.<br />

<strong>aureus</strong> cells grown in TSB to optical densities at 540 nm of 1, 10 or 15. Serial dilutions of total RNA of S. <strong>aureus</strong><br />

RN4220 wild-type (WT) or ∆prsA (left panels), or S. <strong>aureus</strong> SH1000 wild-type (WT) or ∆prsA (right panels) were<br />

blotted and cross linked onto a positively charged nylon membrane. <strong>The</strong> membrane-bound RNA was hybridized<br />

with a digoxigenin labeled RNA probe complementary to RNAIII. Chemiluminescence signals were detected with a<br />

LumiImager (Roche Diagnostics, Mannheim, Germany).<br />

Since the growth to high density coincided with a non-hemolytic phenotype on blood agar<br />

plates, we tested whether these strains might have accumulated agr mutations, leading to the<br />

loss of the regulatory RNAIII. Indeed, dot blot analyses performed for the S. <strong>aureus</strong> SH1000<br />

prsA mutant strain, for which this high density growth phenotype was first observed, revealed<br />

the loss of RNAIII production (Figure 4B). In contrast, the S. <strong>aureus</strong> RN4220 prsA mutant<br />

produced RNAIII at levels comparable to those of the parental strain (Figure 4B).<br />

Figure 5. Agr-like phenotypes of S. <strong>aureus</strong> SH1000-derived secretion mutants. (A) <strong>The</strong> S. <strong>aureus</strong> SH1000<br />

parental strain (WT), an SH1000 ∆cidA strain, an SH1000 ∆comGC strain, an SH1000 ∆dsbA strain, and an<br />

SH1000 ∆prsA strain were grown on blood agar plates and incubated overnight at 37ºC. Hemolytic activity is<br />

detectable as a halo around the streaked cells. (B) S. <strong>aureus</strong> SH1000 (WT), S. <strong>aureus</strong> SH1000 ∆cidA, S. <strong>aureus</strong><br />

SH1000 ∆comC, S. <strong>aureus</strong> SH1000 ∆comGA, S. <strong>aureus</strong> SH1000 ∆comGC, and S. <strong>aureus</strong> SH1000 ∆dsbA were<br />

grown in TSB medium at 37 o C till the early stationary phase. Samples of extracellular proteins isolated from the<br />

growth medium (M), non-covalently cell wall-bound proteins (CW) and total cells (C) were used for Western<br />

blotting and immunodetection with specific antibodies against TrxA or SspB. Note that the antibodies against TrxA<br />

are also bound by the IgG-binding proteins protein A (Spa) and Sbi. Bands corresponding to Spa, Sbi, SspB and<br />

TrxA are marked with arrows.<br />

84


RN4220 OD 540 of 20<br />

RN4220 ∆comGAOD 540 of 20<br />

Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

RN4220 ∆comGA<br />

SH1000 OD 540 of 15<br />

SH1000 ∆srtA OD 540 of 15<br />

RN4220 ∆srtA<br />

Geh<br />

SAOUHSC_02241<br />

Hlb<br />

1<br />

HlgB 2<br />

HlgC<br />

3<br />

1 2<br />

1 2 4<br />

Sle1<br />

YfnI<br />

1 2<br />

2<br />

1<br />

3<br />

SssA<br />

HlY<br />

SAOUHSC_00094<br />

Plc<br />

RN4220 OD 540 of 15<br />

RN4220 ∆secG OD 540 of 15<br />

RN4220 OD 540 of 20<br />

RN4220 ∆tatAC OD 540 of 20<br />

RN4220 ∆secG<br />

4<br />

5<br />

6<br />

SAOUHSC_02979<br />

1 2 3 4<br />

1<br />

2 3<br />

IsaA<br />

RN4220 ∆tatAC<br />

85<br />

2 Spa<br />

1<br />

SdrD<br />

RN4220 OD 540 of 15<br />

RN4220 ∆prsA OD 540 of 15<br />

SH1000 OD 540 of 8<br />

SH1000 ∆prsA OD 540 of 15<br />

RN4220 ∆prsA<br />

SH1000 ∆prsA<br />

Figure 6. Extracellular proteomes of S. <strong>aureus</strong> secretion mutants. (A) Coomassie stained gel of extracellular<br />

proteins of S. <strong>aureus</strong> RN4220 secretion mutants and the corresponding parental strain. Arrows indicate the major<br />

changes in the banding patterns of extracellular proteins of particular secretion mutants. (B) False-colored dualchannel<br />

images of 2-D gels of extracellular proteins of S. <strong>aureus</strong> RN4220 ∆comGA, ∆prsA, ∆secG, ∆srtA or<br />

∆tatAC (labeled red) and S. <strong>aureus</strong> RN4220 (labeled green). Additionally, a false-colored dual-channel image of<br />

2-D gels of extracellular proteins of S. <strong>aureus</strong> SH1000 ∆prsA (labeled red) and S. <strong>aureus</strong> SH1000 (labeled green)<br />

is shown. Proteins (350 µg) isolated from the supernatants of S. <strong>aureus</strong> strains grown in TSB medium were<br />

separated on 2-D gels by using immobilized pH gradient strips in the pH range of 3-10. Proteins were stained with<br />

colloidal Coomassie Brilliant Blue. Spots of proteins present in equal amounts in the media of the mutant and<br />

parental strains appear in yellow, spots of proteins present in higher amounts in media of the mutant strains<br />

appear in red, and spots of proteins present in higher amounts in the medium of the parental strain appear in<br />

green.<br />

B<br />

A


Chapter 4<br />

<strong>The</strong>se findings suggested that several of the secretion mutants obtained in S. <strong>aureus</strong> SH1000<br />

have accumulated agr mutations, which would impact on the production of various secreted<br />

virulence factors (Novick, 2003). This view was confirmed by a systematic analysis of<br />

hemolysin secretion on blood agar plates (Figure 5A) and Western blotting analyses (Figure<br />

5B), which revealed that all mutants growing to high cell density no longer secreted<br />

hemolysin or the cysteine protease SspB. Instead, these strains secreted protein A at strongly<br />

elevated levels (Figure 5). As shown by Western blotting with antibodies against the<br />

cytoplasmic protein TrxA, the mutations causing the agr phenotype did not affect the<br />

secretion of the second IgG-binding protein Sbi, nor did they result in increased levels of lysis<br />

and subsequent release of TrxA from the cells. Since the secretion of many proteins was<br />

suppressed by the mutations leading to the agr phenotype, as was shown by proteomics<br />

analyses for the S. <strong>aureus</strong> SH1000 prsA mutant (Figure 6), mutants with such a phenotype<br />

were excluded from all further protein secretion studies.<br />

In addition to hemolysin, protein A and sspB, it has been shown that agr regulates several<br />

proteases, such as serine proteases SplA-F (Saïd-Salim et al., 2003), the metalloprotease Aur<br />

(Arvidson and Tegmark, 2001;Chan and Foster, 1998), and the cysteine protease SspB<br />

(Arvidson and Tegmark, 2001;Saïd-Salim et al., 2003). Furthermore, by zymography we<br />

observed various differences between wild-type strains and strains with an agr phenotype<br />

with respect to the cellular and extracellular accumulation of as yet unidentified cell wall<br />

hydrolases (data not shown). At present the precise nature of the mutations leading to the agr<br />

phenotype of S. <strong>aureus</strong> SH1000-derived secretion mutants is not clear. PCR on genomic DNA<br />

of these strains and subsequent sequencing revealed that there are no mutations present in the<br />

RNAIII region. This suggests that the mutations are located in other regions of the agr<br />

regulon that control the expression of RNAIII. Though the S. <strong>aureus</strong> strains used in this study<br />

seem to be suitable for mutagenesis studies, the present results underscore the view that<br />

caution must be taken when deciding on what S. <strong>aureus</strong> strain to use for mutagenesis studies.<br />

S. <strong>aureus</strong> SH1000 is clearly suitable for studying many processes, but for mutagenesis with<br />

pMAD-based plasmids, it may be not the best choice due to the high frequency at which the<br />

obtained mutants display an agr-like phenotype. We can only speculate about the possible<br />

reasons. Conceivably, the temperature changes that are needed to promote chromosomal<br />

integration and excision of pMAD may give a competitive growth advantage to SH1000<br />

derivatives with an agr phenotype. However, also the introduction of pMAD itself may<br />

represent a stressful event for the cells that could lead to the selection for mutations that cause<br />

the agr phenotype. It has already been shown that mutations in the agr locus can easily occur<br />

during re-culturing of S. <strong>aureus</strong> strains (Somerville et al., 2002), and this also seems to<br />

happen in patients (Traber et al., 2008). Our present results indicate that the same can occur<br />

with S. <strong>aureus</strong> SH1000, especially when cells are subjected to our pMAD-based mutagenesis<br />

regime.<br />

Exoproteome analysis by SDS-PAGE and proteomics<br />

To study the effects of the different deletions of secretion machinery genes, the exoproteomes<br />

of the mutant S. <strong>aureus</strong> strains were analyzed by SDS-PAGE or 2D-gelelectrophoresis. This<br />

revealed that the exoproteomes of some mutants were severely changed due to the respective<br />

gene deletions, while the exoproteomes of other deletion mutants had remained unaltered<br />

(Figure 6). Especially, the secG and lgt mutants displayed a substantially different<br />

exoproteome composition compared to the parental strain RN4220 (Figure 6A).<br />

86


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

<strong>The</strong> significant differences in the exoproteome of the secG mutant were highly unexpected<br />

since deletion of secG in E. coli and B. subtilis does not seem to have a strong impact on<br />

protein export in these bacteria (Nishiyama et al., 1994;Van Wely et al., 1999). As further<br />

detailed in Chapter 5, many proteins are present at increased or decreased levels in the<br />

exoproteome of the S. <strong>aureus</strong> RN4220 secG mutant and this effect was reversed by ectopic<br />

expression of secG. Proteins that were detected in reduced amounts include well-known<br />

secreted virulence factors, such as the lipase Geh, the alpha and beta hemolysins and the<br />

leukocidins F and S, as well as the cell surface proteins SdrC and SdrD. Conversely, the<br />

extracellular levels of other virulence factors, like protein A, the immunodominant antigen A<br />

(IsaA) and the secretory antigen precursor SsaA were significantly increased. <strong>The</strong>se<br />

observations suggest that the absence of SecG changes the translocation efficiency of<br />

different exoproteins via the SecYE channel to different extents. However, we cannot<br />

completely rule out indirect effects on the transcription of the genes for certain exoproteins,<br />

their translation, or their post-translocational folding or degradation. In contrast to the secG<br />

mutation, no secretion defect was observed upon the deletion of the gene for the accessory<br />

Sec channel component SecY2. This confirmed the results obtained by Siboo et al., who<br />

showed that a secY2 mutation does not alter the S. <strong>aureus</strong> exoproteome (Siboo et al., 2008).<br />

Till now, the only known SecY2-dependent protein of S. <strong>aureus</strong> is SraP, and it seems as if the<br />

SecA2/SecY2 translocon is devoted to the translocation of this protein alone. Whether SecG<br />

and/or SecE interact with this accessory SecA2/SecY2 translocon remains to be determined.<br />

Similar to the secY2 deletion strain, no secretion defects were observed for strains lacking the<br />

tatA and/or tatC genes, the comGA, comGC or comC genes, or the genes for the holin CidA<br />

and antiholin LrgA (Figure 6B). This suggests that the Tat, Comand holing-antiholin<br />

pathways serve very specific roles in the translocation of particular non-abundantly expressed<br />

proteins, or proteins that are not detectable by 2-D PAGE due to their particular pI and size<br />

properties. This view was recently confirmed for the S. <strong>aureus</strong> Tat pathway, which seems to<br />

be required for the translocation of only one protein, namely FepB (Biswas et al.,<br />

2009;Yamada et al., 2007). So far, we have not been able to detect FepB by our 2-D PAGE<br />

approach.<br />

<strong>The</strong> spsA gene encodes a type I signal peptidase that seems to lack the active site serine and<br />

lysine residues (van Roosmalen et al., 2004). Interestingly, the occurrence of this type of an<br />

apparently inactive type I signal peptidase is wide-spread amongst the Firmicutes.<br />

Unfortunately, our proteomics studies did not shed light on the biological function of SpsA as<br />

no differences were observed between the exoproteomes of our spsA deletion mutant and the<br />

parental strain RN4220 (Figure 6).<br />

<strong>The</strong> exoproteome of the lgt mutant showed the most pronounced differences compared to the<br />

parental strain (Figure 6A). This result was not completely unexpected as previous studies of<br />

Stoll et al. (Stoll et al., 2005) had shown that an S. <strong>aureus</strong> lgt mutant released high amounts of<br />

certain unmodified prelipoproteins, such as the oligopeptide-binding protein OppA, the<br />

peptidyl-prolyl cis-trans isomerase PrsA, and the staphylococcal iron transporter SitC, into the<br />

growth medium. As shown by Western blotting, the same was true for the unmodified pre-<br />

DsbA precursor, which was completely released into the growth medium of lgt mutant cells.<br />

In contrast, in the parental strain, the mature DsbA fractionated with cells and non-covalently<br />

cell wall-associated proteins (Figure 7). <strong>The</strong> latter finding seems to suggest that some mature<br />

DsbA is not retained at the membrane but in the cell wall. <strong>The</strong>se findings confirm earlier<br />

observations of pre-lipoprotein release by lgt mutant strains of B. subtilis (Antelmann et al.,<br />

2001;Tjalsma et al., 1999) and L. lactis (Venema et al., 2003). <strong>The</strong> release of lipoprotein<br />

87


Chapter 4<br />

precursors by lgt mutant cells indicates that lipoprotein signal peptides are too short or<br />

insufficiently hydrophobic to retain the unmodified lipoprotein precursors in the membrane.<br />

<strong>The</strong> fact that the exoproteome of the lspA mutant strain, lacking the lipoprotein-specific signal<br />

peptidase II, remained unaltered confirms the notion that correctly lipid-modified but<br />

immature lipoprotein precursors do remain anchored to the membrane (Figure 6A).<br />

Figure 7. Localization of the lipoprotein DsbA in different S. <strong>aureus</strong> secretion mutants. S. <strong>aureus</strong> SH1000 (WT),<br />

S. <strong>aureus</strong> SH1000 ∆dsbA, and S. <strong>aureus</strong> SH1000 ∆lgt were grown in TSB medium at 37 o C till the early stationary<br />

phase. Samples of extracellular proteins isolated from the growth medium (M), non-covalently cell wall-bound<br />

proteins (CW) and total cells (C) were used for Western blotting and immunodetection with specific antibodies<br />

against DsbA. Bands corresponding to the precursor (preDsbA) and mature forms of DsbA are marked with<br />

arrows.<br />

Remarkably, deletion of either of the two known extracytoplasmic catalysts for protein<br />

folding, namely DsbA and PrsA, had no detectable effects on the composition of the S. <strong>aureus</strong><br />

exoproteome (Figure. 6, A and B). This suggests that PrsA is required for the folding of only<br />

a very limited number of extracytoplasmic proteins of S. <strong>aureus</strong> that are not readily detectable<br />

by 2-D PAGE or that are not expressed under the tested experimental conditions. Likewise,<br />

our results indicate that DsbA has a very specific function, as was previously proposed on the<br />

basis of genetic and biochemical analyses (Kouwen et al., 2007;Dumoulin et al., 2005).<br />

Unexpectedly, 2-D PAGE revealed no clear differences in the exoproteomes of the srtA or<br />

srtB mutants compared to the parental strain (Figure 6). Proteins with an LPxTG, LPxAG or<br />

NPQTN motif are not covalently anchored to the cell wall in the srtA or srtB mutants.<br />

<strong>The</strong>refore, one might expect the release of these proteins into the growth medium of srtA or<br />

srtB mutant strains. However, it is well conceivable that some proteins released by srtA or<br />

srtB mutant cells escaped detection by 2-D PAGE, because of their physical properties,<br />

because they are expressed at very low levels, or because they are not expressed at all under<br />

the tested conditions. <strong>The</strong> latter is most likely true for the srtB mutant since the only known<br />

substrate, IsdC, is only expressed under iron-limiting conditions (Mazmanian et al., 2003).<br />

Furthermore, it has been shown that the sortase A substrate SasG is not detectable in wildtype<br />

strains of S. <strong>aureus</strong> (Corrigan et al., 2007). Another possible explanation for the fact that<br />

no LPxTG proteins were identified in the exoproteome of the srtA mutant is that these<br />

proteins may remain linked to the cell wall in other ways. For example, all LPxTG proteins<br />

have a C-terminal membrane anchor that is proteolytically removed during covalent cell wall<br />

attachment of these proteins by sortase A. This anchor has the potential to retain the Cterminally<br />

uncleaved proteins in the membrane. In addition, a range of LPxTG proteins also<br />

contain other cell wall binding motifs (repeats) that would maintain protein linkage to the cell<br />

wall even in absence of sortase A. <strong>The</strong>se issues have been addressed in more detail in<br />

Chapter 6.<br />

88


Characterization of <strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants<br />

Caenorhabditis elegans killing assay<br />

Many S. <strong>aureus</strong> proteins that are transported to the cell surface or extracellular milieu have a<br />

role in virulence (Sibbald et al., 2006). This raises the question to what extent the generated<br />

mutations influence staphylococcal virulence. In most published studies this is addressed by<br />

experiments with animal models such as mice or rats. However, such analyses are timeconsuming<br />

and require special facilities. <strong>The</strong>refore, we explored the use of a Caenorhabditis<br />

elegans killing assay to screen for possible virulence defects of our secretion mutants.<br />

Importantly, this assay was previously used by Sifri and colleagues to study the effects of<br />

several S. <strong>aureus</strong> strains on the survival of nematodes (Sifri et al., 2003), including strains<br />

that carried mutations in regulators of virulence factors, such as agr and sarA, or strains that<br />

did not produce α-hemolysin. <strong>The</strong> S. <strong>aureus</strong> strains that carried these mutations were<br />

attenuated in killing the nematodes and this showed that the assay could be used to probe the<br />

virulence of particular S. <strong>aureus</strong> strains. <strong>The</strong> results of the C. elegans killing assays for a<br />

selection of our S. <strong>aureus</strong> secretion mutants are shown in Figure 8.<br />

Figure 8. Kaplan-Meier survival plots of C. elegans infected with S. <strong>aureus</strong> mutants. S. <strong>aureus</strong> strains were<br />

spotted on TSA plates and nematodes were placed onto these plates. Live nematodes were counted every 24 hours<br />

and transferred to fresh TSA plates containing S. <strong>aureus</strong> spots. <strong>The</strong> assay for each strain was performed three<br />

times. <strong>The</strong> plots were drawn with the StatView version 5.0 program (SAS Institute, Inc.).<br />

<strong>The</strong> tested S. <strong>aureus</strong> dsbA, spsA srtA, secG, secY2 and secG-secY2 mutants displayed C.<br />

elegans killing rates, which were comparable to that of the parental strain. <strong>The</strong>re were,<br />

however, a few mutants that showed attenuated killing rates, such as the lspA and srtB<br />

mutants. Conversely, the results in Figure 8 imply that lgt and tatAC mutant strains are<br />

slightly more virulent. <strong>The</strong> effect of the lspA mutation indicates that uncleaved lipoproteins,<br />

retained at the membrane in the absence of signal peptidase II, are insufficiently active for full<br />

virulence of S. <strong>aureus</strong> towards C. elegans. On the other hand, the increased release of<br />

lipoproteins by the lgt mutant may make S. <strong>aureus</strong> more virulent towards C. elegans. It has<br />

been shown previously that lipoproteins are responsible for activation of TLR2 and that such<br />

proteins are an important factor for the pathogenesis of S. <strong>aureus</strong> (Hashimoto et al.,<br />

2006;Kurokawa et al., 2009;Schmaler et al., 2009). Conversely, it has been reported that S.<br />

89


Chapter 4<br />

<strong>aureus</strong> lgt mutants may be hypervirulent in vivo (Stoll et al., 2005;Bubeck et al., 2006), which<br />

would be in line with our present findings. <strong>The</strong> observation that the srtB mutant is attenuated<br />

in C. elegans killing suggests that the IsdC protein is required for C. elegans killing and that<br />

S. <strong>aureus</strong> cells are exposed to iron-limiting conditions during C. elegans infection<br />

(Mazmanian et al., 2003). This would in fact suggest that the reason why the lspA mutant is<br />

attenuated might be related to the malfunction of certain lipoproteins involved in iron uptake<br />

(Schmaler et al., 2009). Remarkably, the tatAC mutant showed a higher killing rate compared<br />

to the parental strain. This finding is somewhat difficult to reconcile with the observed C.<br />

elegans killing phenotypes of the srtB and lspA mutants, because the only confirmed substrate<br />

for the S. <strong>aureus</strong> Tat pathway is the FepB protein, which is an iron-dependent peroxidase<br />

needed for iron uptake (Biswas et al., 2009). Moreover, the bacterial loads of tatAC and tatfep<br />

mutant strains in a mouse kidney abscess model were decreased, suggesting a requirement<br />

of the Tat pathway for virulence. At present, it remains unfortunately unclear why the effects<br />

of the srtB and lspA mutations on the one hand and the tatAC mutation on the other hand are<br />

so different. Finally, it is important to note that no C. elegans killing phenotype was observed<br />

for the srtA mutant. Clearly, the proteins anchored by sortase A to the cell surface are very<br />

important for the virulence of S. <strong>aureus</strong> towards mammals as they are, for example, involved<br />

in the binding of fibrinogen (McDevitt et al., 1997), binding to nasal epithelial cells (Corrigan<br />

et al., 2009), evasion of the immune system (Sasso et al., 1991) or biofilm formation<br />

(Corrigan et al., 2007). Possibly, these traits are less relevant for S. <strong>aureus</strong> virulence towards<br />

much simpler host organisms such as nematodes.<br />

Outlook<br />

In conclusion, our present studies show that many of the non-essential determinants for<br />

protein secretion by S. <strong>aureus</strong> have only a limited impact on general protein secretion by this<br />

organism. This implies that these factors are mainly required for special purposes that may,<br />

however, be relevant for the virulence of S. <strong>aureus</strong>. Defining the precise roles of such<br />

secretion factors will require more in-depth studies, especially under infection mimicking<br />

conditions. An important lesson that was learned from the studies with the S. <strong>aureus</strong> strain<br />

SH1000 is that this strain quite rapidly accumulates mutations causing an agr-like phenotype.<br />

This underscores the need for genetically stable model strains for molecular genetics and<br />

proteomics approaches to fully define the roles of the <strong>secretome</strong> in staphylococcal virulence.<br />

Acknowledgements<br />

We like to thank W. Baas for technical assistance, and colleagues from the StaphDynamics<br />

and <strong>TI</strong><strong>Pharma</strong> programs for helpful discussions. Financial support was provided by CEU<br />

(StaphDynamics, LSHM-CT-2006-019064; DFG (GK212/3-00, SFB/TR34, FOR 585), and<br />

Top Institute <strong>Pharma</strong> (T4-213).<br />

90


“Besides being a guitar player, I'm a big fan of the guitar<br />

I love that damn instrument”<br />

-Steven S. Vai-<br />

92


Chapter 5<br />

Synthetic effects of secG and secY2 mutations on exoproteome<br />

biogenesis in <strong>Staphylococcus</strong> <strong>aureus</strong><br />

M.J.J.B. Sibbald # , T. Winter # , M.M. van der Kooi-Pol, T. Bosma, T. Schäfer, K. Ohlsen,<br />

M. Hecker, H. Antelmann, S. Engelmann, and J.M. van Dijl<br />

# both authors contributed equally to this work<br />

Submitted for publication, in revision<br />

93


Chapter 5<br />

Summary<br />

<strong>The</strong> Gram-positive pathogen <strong>Staphylococcus</strong> <strong>aureus</strong> secretes various proteins into its<br />

extracellular milieu. Bioinformatics analyses have indicated that most of these proteins<br />

are directed to the canonical Sec pathway, which consists of the translocation motor<br />

SecA and a membrane-embedded channel composed of the SecY, SecE and SecG<br />

proteins. In addition, S. <strong>aureus</strong> contains an accessory Sec2 pathway involving the SecA2<br />

and SecY2 proteins. Here we have addressed the roles of the non-essential channel<br />

components SecG and SecY2 in the biogenesis of the extracellular proteome of S. <strong>aureus</strong>.<br />

<strong>The</strong> results show that SecG is of major importance for protein secretion by S. <strong>aureus</strong>.<br />

Specifically, the extracellular accumulation of eight abundant exoproteins and seven cell<br />

wall-bound proteins was significantly affected in the secG mutant. No secretion defects<br />

were detected for strains with a secY2 single mutation. However, deletion of secY2<br />

exacerbated the secretion defects of secG mutants, affecting the extracellular<br />

accumulation of one additional exoprotein and one cell wall protein. Furthermore, the<br />

secG secY2 double mutant displayed a synthetic growth defect. <strong>The</strong>se findings suggest<br />

that SecY2 can interact with the Sec1 channel of S. <strong>aureus</strong>. Such an interaction would be<br />

consistent with the presence of a single set of secE and secG genes in S. <strong>aureus</strong>.<br />

94


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

Introduction<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> is a well-represented component of the human microbiota as nasal<br />

carriage of this Gram-positive bacterium has been shown for 30-40% of the population<br />

(Peacock et al., 2001). This organism can, however, turn into a dangerous pathogen that is<br />

able to infect almost every tissue in the human body. S. <strong>aureus</strong> has become particularly<br />

notorious for its high potential to develop resistance against commonly used antibiotics<br />

(Hiramatsu et al., 1997; Weigel et al., 2003). Accordingly, the S. <strong>aureus</strong> genome encodes an<br />

arsenal of virulence factors that can be expressed when needed at different stages of growth.<br />

<strong>The</strong>se include surface proteins and invasins that are necessary for colonization of host tissues,<br />

surface-exposed factors for evasion of the immune system, exotoxins for the subversion of<br />

protective host barriers, and resistance proteins for protection against antimicrobial agents<br />

(Schneewind et al., 1992).<br />

Most proteinaceous virulence factors of S. <strong>aureus</strong> are synthesized with an N-terminal<br />

signal peptide to direct their transport from the cytoplasm across the membrane to an<br />

extracytoplasmic location, such as the cell wall or the extracellular milieu (Schneewind et al.,<br />

1992; Sibbald et al., 2006). Based on signal peptide predictions using a variety of algorithms,<br />

it is believed that most exoproteins of S. <strong>aureus</strong> are exported to extracytoplasmic locations via<br />

the general Secretory (Sec) Pathway (Sibbald et al., 2006). <strong>The</strong>se pre-proteins are bound by<br />

the translocation motor protein SecA. Through repeated cycles of ATP binding and<br />

hydrolysis, SecA pushes the protein in an unfolded state through the membrane-embedded<br />

SecYEG translocation channel (Driessen and Nouwen, 2008; Papanikou et al., 2007). This<br />

channel is homologous to the eukaryotic Sec61αγβ channel and it can be found in all three<br />

kingdoms of life (Pohlschröder et al., 1997; Yuan et al., 2009). Upon initiation of the<br />

translocation process, the proton-motive force is thought to accelerate pre-protein<br />

translocation through the Sec channel (Nishiyama et al., 1993). Recently, the structure of the<br />

SecA/SecYEG complex from the Gram-negative bacterium <strong>The</strong>rmotoga maritima was solved<br />

at 4.5 A resolution (Zimmer et al., 2008). In this structure, one SecA molecule is bound to<br />

one set of SecYEG channel proteins. <strong>The</strong> core of the Sec translocon consists of the SecA,<br />

SecY and SecE proteins, which are essential for growth and viability of bacteria, such as<br />

Escherichia coli and Bacillus subtilis (Cabelli et al., 1988; Brundage et al., 1990; Kobayashi<br />

et al., 2000). In contrast, the channel component SecG is dispensable for growth, cell viability<br />

and protein translocation (Nishiyama et al., 1993). Nevertheless, SecG does enhance the<br />

efficiency of pre-protein translocation through the SecYE channel. This is of particular<br />

relevance at low temperatures and in the absence of a proton-motive force (Hanada et al.,<br />

1996). Several studies suggest that E. coli SecG undergoes topology inversion during preprotein<br />

translocation (Nishiyama et al., 1993; Nagamori et al., 2000; Sugai et al., 2007). Even<br />

so, van der Sluis et al. reported that SecG cross-linked to SecY is fully functional despite its<br />

fixed topology (van der Sluis et al., 2006). During or shortly after membrane translocation of<br />

a pre-protein through the Sec channel, the signal peptide is removed by signal peptidase. This<br />

is a prerequisite for the release of the translocated protein from the membrane (Antelmann et<br />

al., 2001; van Roosmalen et al., 2004).<br />

Several pathogens, including Streptococcus gordonii, Streptococcus pneumonia,<br />

Bacillus anthracis, Bacillus cereus, and S. <strong>aureus</strong> contain a second set of chromosomal secA<br />

and secY genes named secA2 and secY2, respectively (Sibbald and van Dijl, 2009).<br />

Comparison of the amino acid sequences of the SecY1 and SecY2 proteins shows that their<br />

similarity is low (about 20% identity), and that the conserved regions are mainly restricted to<br />

95


Chapter 5<br />

the membrane spanning domains. It has been shown for S. gordonii that the transport of at<br />

least one protein is dependent on the presence of SecA2 and SecY2. This protein, GspB, is a<br />

large cell-surface glycoprotein that is involved in platelet binding (Bensing and Sullam,<br />

2002). <strong>The</strong> protein contains an unusually long N-terminal signal peptide of 90 amino acids,<br />

large serine-rich repeats, and a C-terminal LPxTG motif for covalent cell wall binding. <strong>The</strong><br />

gspB gene is located in a gene cluster with the secA2 and secY2 genes. Two other genes in<br />

this cluster encode for the glycosylation proteins GftA and GftB, which seem to be necessary<br />

for stabilization of pre-GspB. Furthermore, the asp4 and asp5 genes in the secA2 secY2 gene<br />

cluster show similarity to secE and secG, and they are important for GspB export by S.<br />

gordonii (Takamatsu et al., 2005). Despite this similarity, SecE and SecG cannot complement<br />

for the absence of Asp4 and Asp5, respectively. <strong>The</strong> secA2/secY2 gene cluster is also present<br />

in S. <strong>aureus</strong>, but homologues of the asp4 and asp5 genes are lacking. This seems to suggest<br />

that SecA2 and SecY2 of S. <strong>aureus</strong> share the SecE and SecG proteins with SecA1 and SecY1.<br />

<strong>The</strong> sraP gene in the secA2/secY2 gene cluster of S. <strong>aureus</strong> encodes a protein with similar<br />

features as described for GspB. Siboo and colleagues (Siboo et al., 2005) have shown that<br />

SraP is glycosylated and capable of binding to platelets. Importantly, the disruption of sraP<br />

resulted in a decreased ability to initiate infective endocarditis in a rabbit model. Consistent<br />

with the findings in S. gordonii, SraP export was shown to depend on SecA2/SecY2 (Siboo et<br />

al., 2008). However, it has remained unclear whether other S. <strong>aureus</strong> proteins are also<br />

translocated across the membrane in a SecA2/SecY2-dependent manner.<br />

<strong>The</strong> present studies were aimed at defining the roles of two Sec channel components,<br />

SecG and SecY2, in protein secretion by S. <strong>aureus</strong>. <strong>The</strong> results show that secG and secY2 are<br />

not essential for growth and viability of S. <strong>aureus</strong>. While the absence of SecY2 by itself had<br />

no detectable effect, the absence of SecG had a profound impact on the composition of the<br />

exoproteome of a S. <strong>aureus</strong>. Various extracellular proteins were present in decreased amounts<br />

in the growth medium of secG mutant strains, which is consistent with impaired Sec channel<br />

function. However, a few proteins were present in increased amounts. Furthermore, the<br />

absence of secG caused a serious decrease in the amounts of the cell wall-bound Sbi protein.<br />

Most notable, a secG secY2 double mutant strain displayed synthetic growth and secretion<br />

defects.<br />

Material & Methods<br />

Bacterial strains and plasmids<br />

All strains used in this study are listed in Table 1. Unless stated otherwise, E. coli strains were grown in<br />

Luria-Bertani broth (LB). S. <strong>aureus</strong> strains were grown at 37°C in tryptic soy broth (TSB) under<br />

vigorous shaking or on trypic soy agar (TSA) plates. If appropriate, media for E. coli were<br />

supplemented with 100 µg/ml ampicillin or 100 µg/ml erythromycin, and media for S. <strong>aureus</strong> with 5<br />

µg/ml erythromycin, 5 µg/ml tetracyclin or 20 µg/ml kanamycin. To monitor β-galactosidase activity in<br />

cells of E. coli and S. <strong>aureus</strong>, 5-bromo-4-chloro-3-indolyl-β-D-galactopyranoside (X-gal) was added to<br />

the plates at a final concentration of 80 µg/ml.<br />

96


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

Table 1. Plasmids and bacterial strains used<br />

Plasmids Properties Reference<br />

TOPO pCR®-Blunt II-TOPO® vector; Km R Invitrogen Life<br />

technologies<br />

pCN51 E. coli / S. <strong>aureus</strong> shuttle vector that contains a cadmium-inducible (Charpentier et<br />

promoter<br />

al., 2004)<br />

pMAD E. coli / S. <strong>aureus</strong> shuttle vector that is temperature-sensitive in S.<br />

<strong>aureus</strong> and contains the bgaB gene, Ery R , Amp R<br />

(Arnaud et al.,<br />

2004)<br />

pUC18 Amp R , ColE1, F80dLacZ, lac promoter (Norrander<br />

al., 1983)<br />

et<br />

pDG783 1.5-kb kanamycin resistance cassette in pSB118; Amp R (Guérout-Fleury<br />

et al., 1995)<br />

secG-pCN51 pCN51 with S. <strong>aureus</strong> secG gene, Amp R ; Ery R This work<br />

secY2-pCN51 pCN51 with S. <strong>aureus</strong> secY2 gene, Amp R ; Ery R This work<br />

Strains<br />

E. coli<br />

Genotype Reference<br />

DH5α supE44; hsdR17; recA1; gyrA96; thi-1; relA1 (Hanahan,<br />

1983)<br />

TOP10 Cloning host for TOPO vector; F - mcrA ∆(mrr-hsdRMS-mcrBC)<br />

Φ80lacZ∆M15 ∆lacX74 recA1 araD139 ∆(ara-leu)7697 galU<br />

galK rpsL (Str R ) endA1 nupG<br />

S. <strong>aureus</strong> RN4220<br />

Parental strain Restriction-deficient derivative of NCTC 8325, cured of all known<br />

prophages, partial defect in agr<br />

97<br />

Invitrogen Life<br />

technologies<br />

(Kreiswirth et<br />

al., 1983)<br />

∆secG secG This work<br />

∆secY2 secY2 This work<br />

∆secG ∆secY2 secG secY2 This work<br />

S. <strong>aureus</strong> SH1000<br />

WT Functional rsbU+ derivative of 8325-4 rsbU+, agr+ (Horsburgh et<br />

al., 2002)<br />

∆secG rsbU+, agr+, secG This work<br />

∆secY2 rsbU+, agr+, secY2 This work<br />

∆secG ∆secY2 rsbU+, agr+, secG secY2 This work<br />

S. <strong>aureus</strong> Newman<br />

∆spa spa (Patel et al.,<br />

1987)<br />

∆spa ∆sbi spa sbi This work<br />

Construction of S. <strong>aureus</strong> mutant strains<br />

Mutants of S. <strong>aureus</strong> were constructed using the temperature-sensitive plasmid pMAD (Arnaud et al.,<br />

2004) and previously described procedures (Kouwen et al., 2009c). Primers (Table 2) were designed<br />

using the genome sequence of S. <strong>aureus</strong> NCTC8325 (http://www.ncbi.nlm.nih.gov/nuccore/<br />

NC_007795). All mutant strains were checked by isolation of genomic DNA using the GenElute<br />

Bacterial Genomic DNA Kit (Sigma) and PCR with specific primers.<br />

To delete the secG or secY2 genes, primer pairs with the designations F1/R1 and F2/R2 were<br />

used for PCR amplification of the respective upstream and downstream regions (each ~500 bp), and<br />

their fusion with a 21 bp linker. <strong>The</strong> fused flanking regions were cloned in pMAD, and the resulting<br />

plasmids were used to delete the chromosomal secG or secY2 genes of S. <strong>aureus</strong> RN4220. To delete the<br />

secG or secY2 genes from the S. <strong>aureus</strong> SH1000 genome, the respective pMAD constructs were<br />

transferred from the RN4220 strain to the SH1000 strain by transduction with phage φ85 (Novick,<br />

1991).


Chapter 5<br />

Table 2. Primers used in this study<br />

Primer Sequence (5’→3’)<br />

secG-F1 TTAAAACAGGACGCTTTATTG<br />

secG-R1 TTACGTCAGTCAGTCACCATGGCA AAATTGTCCTCCGTTCCTTAT<br />

secG-F2 TGCCATGGTGACTGACTGACGTAA GGTCCGGCGATGTAAATGTCG<br />

secG-R2 GCGTGCATATTCTAAAAAGCC<br />

secY2-F1 TGTCTGGTTCACAAAGCATTT<br />

secY2-R1 TTACGTCAGTCAGTCACCATGGCA GTTGCACCTCTTTTATATCAA<br />

secY2-F2 TGCCATGGTGACTGACTGACGTAA GGAGGTAATTATGAAATACTT<br />

secY2-R2 GCCTCTCCCTGATCATCAAAA<br />

sbi-F1 TGTGTTCCTTTATTTTCTGCG<br />

sbi-R1 GAACTCCAATTCACCCATGGCCCCC CCCCAACTAGCAACTTCGAG<br />

sbi-F2 CCGCAACTGTCCATACCATGGCCCCC GGAAATAATCAATCAAAAATATCTTCTC<br />

sbi-R2 CTATTAAACCAACTGCTAAAGTTGC<br />

kan-F1 GGGGGCCATGGGTGAATTGGAGTTCGTCTTG<br />

kan-R1 GGGGGCCATGGTATGGACAGTTGCGGATGTA<br />

secG-F3 GGGGGGTCGACGGGATATACTACTTGTCGTATATA<br />

secG-R3 GGGGGGAATTCCCTTACATACCAAGATAACTTATGCA<br />

secY2-F3 GGGGGGTCGACGTCTTTTTAATGTTTTTGATA<br />

secY2-R3 GGGGGGAATTCCCTTACCAATACTGGTTTAAAAATGG<br />

spa_for ACCTGCTGCAAATGCTGCGC<br />

spa_rev a CTAATACGACTCACTATAGGGAGA GGTTAGCACTTTGGCTTGGG<br />

geh_for CACATCAAATGCAGTCAGG<br />

geh_rev a CTAATACGACTCACTATAGGGAGA AATCGACATGATCCCATCC<br />

hlb_for ATCAAACACCTGTACTCGG<br />

hlb_rev a CTAATACGACTCACTATAGGGAGA CGTAGTAATATGGGAACGC<br />

Overlap in primers are in bold; restriction sites are underlined<br />

a Oligonucleotides containing the recognition sequence for T7 polymerase at the 5’ end (shown in italic)<br />

To create the spa sbi double mutant of S. <strong>aureus</strong> Newman, the sbi gene was deleted from a spa<br />

mutant strain kindly provided by T. Foster (Patel et al., 1987). For this purpose, the kanamycin<br />

resistance marker encoded by pDG783 was introduced between the sbi flanking regions via PCR with<br />

the primer pairs sbi-F1/sbi-R1, sbi-F2/sbi-R2 and kan-F1/kan-R1. <strong>The</strong> obtained ~1000 bp fragment<br />

was ligated into pMAD, and the resulting plasmid was used to transform competent S. <strong>aureus</strong> Newman<br />

spa cells. Blue colonies were selected on TSA plates with erythromycin and kanamycin, and the spa<br />

sbi double mutant was subsequently identified following the previously described protocol (Kouwen et<br />

al., 2009c).<br />

For complementation studies, the secG or secY2 genes were cloned into plasmid pCN51<br />

(Charpentier et al., 2004). Expression of genes cloned in this plasmid is directed by a cadmiuminducible<br />

promoter. Primer pairs with the F3/R3 designation (Table 2) were used to amplify the secG<br />

or secY2 genes. <strong>The</strong>se primers contain an EcoRI restriction site at the 5’ end and a SalI restriction site<br />

at the 3’end of the amplified gene. PCR products were purified using the PCR Purification Kit (Roche),<br />

and ligated into the TOPO-vector (Invitrogen). <strong>The</strong> resulting constructs were then cut with EcoRI and<br />

SalI, and the secG or secY2 genes (284 and 1233 bp, respectively) were isolated from an agarose gel<br />

and ligated into pCN51 cut with EcoRI and SalI. This resulted in the secG- and secY2-pCN51<br />

plasmids. Competent S. <strong>aureus</strong> RN4220 ∆secG, ∆secY2 or ∆secG ∆secY2 cells were transformed with<br />

these plasmids by electroporation and colonies were selected on TSA plates containing erythromycin.<br />

<strong>The</strong> plasmids were then transferred to S. <strong>aureus</strong> SH1000 by transduction as described above.<br />

Analytical and preparative two-dimensional (2-D) PAGE<br />

Extracellular proteins from 100 ml culture supernatant were precipitated, washed, dried, and resolved<br />

as described previously (Ziebandt et al., 2004). <strong>The</strong> protein concentration was determined using Roti ® -<br />

Nanoquant (Carl Roth GmbH & Co, Karlsruhe, Germany). Preparative 2-D PAGE was performed by<br />

98


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

using the immobilized pH gradient technique (Bernhardt et al., 1999; Eymann et al., 2004). <strong>The</strong> protein<br />

samples (350 µg) were separated on immobilized pH gradient strips (Amersham <strong>Pharma</strong>cia Biotech,<br />

Piscataway, NJ) with a pH range of 3-10. <strong>The</strong> resulting protein gels were stained with colloidal<br />

Coomassie Blue G-250G (Candiano et al., 2004) and scanned with the light scanner. Each experiment<br />

was performed at least three times.<br />

For identification of proteins by MALDI-TOF MS, Coomassie-stained protein spots were<br />

excised from gels using a spot cutter (Proteome Work TM ) with a picker head of 2 mm and transferred<br />

into 96-well microtiter plates. Digestion with trypsin and subsequent spotting of peptide solutions onto<br />

the MALDI targets were performed automatically in an Ettan Spot Handling Workstation (GE-<br />

Healthcare, Little Chalfont, United Kingdom) using a modified standard protocol. MALDI-TOF MS<br />

analyses of spotted peptide solutions were carried out on a Proteome-Analyzer 4700/4800 (Applied<br />

Biosystems, Foster City, CA) as described previously (Eymann et al., 2004). MALDI-TOF-TOF<br />

analysis was performed for the three highest peaks of the TOF spectrum as described previously<br />

(Eymann et al., 2004; Wolff et al., 2007). Database searches were performed using the GPS explorer<br />

software version 3.6 (build 329) with the organism-specific databases.<br />

By using the MASCOT search engine version 2.1.0.4. (Matrix Science, London, UK) the<br />

combined MS and MS/MS peak lists for each protein spot were searched against a database containing<br />

protein sequences derived from the genome sequences of S. <strong>aureus</strong> NCTC8325. Search parameters<br />

were as described previously (Wolff et al., 2007). For comparison of protein spot volumes, the Delta<br />

2D software package was used (Decodon GmbH Germany). <strong>The</strong> induction ratio of parental strain to<br />

mutant was calculated for each spot (normalized intensity of a spot on the parental image/normalized<br />

intensity of the corresponding spot on the mutant image). <strong>The</strong> significance of spot volume differences<br />

of two-fold or higher was assessed by the Student´s t test (α


Chapter 5<br />

incubated at 95ºC. Proteins were separated by SDS-PAGE using precast NuPage gels (Invitrogen) and<br />

subsequently blotted onto a nitrocellulose membrane (Protran ® , Schleicher & Schuell). <strong>The</strong> presence of<br />

IsaA, Aly or DsbA was monitored by immunodetection with specific polyclonal antibodies raised in<br />

mice (IsaA, Aly) or rabbits (DsbA (Kouwen et al., 2007)) at 1:10.000 dilution. Bound primary<br />

antibodies were visualized using fluorescent IgG secondary antibodies (IRDye 800 CW goat antimouse/anti-rabbit<br />

from LiCor Biosciences). Membranes were scanned for fluorescence at 800 nm using<br />

the Odyssey Infrared Imaging System (LiCor Biosciences).<br />

Results<br />

<strong>The</strong> exoproteomes of secG and secY2 mutant S. <strong>aureus</strong> strains<br />

To investigate the roles of SecG and SecY2 in the biogenesis of the S. <strong>aureus</strong> exoproteome,<br />

the respective genes were completely deleted from the chromosome of S. <strong>aureus</strong> strain<br />

RN4220. This resulted in the single mutant strains ∆secG and ∆secY2, and the double mutant<br />

∆secG ∆secY2. Next, cells of these mutants were grown in TSB medium until they reached<br />

the stationary phase (Figure 1; not shown for the ∆secY2 strain).<br />

Figure 1 Growth of S. <strong>aureus</strong> secG and secG secY2 mutants. <strong>The</strong> S. <strong>aureus</strong> strains RN4220 ∆secG (A), ∆secG<br />

∆secY2 (B), and the parental strain RN4220 were grown in 100 ml TSB medium under vigorous shaking at 37 o C.<br />

Sampling points for the preparation of extracellular proteins are indicated in the growth curve by arrows.<br />

All three mutants displayed similar exponential growth rates as the parental strain. However,<br />

the secG secY2 double mutant entered the stationary phase at a lower optical density<br />

(OD540=8) than the parental strain and the ∆secG mutant (OD540=15). Extracellular proteins<br />

were collected from the supernatant of the cell cultures that had reached stationary phase for<br />

analysis by 2-D PAGE (Figures 1 and 2). Comparison of the exoproteomes of the secG<br />

mutant and its parental strain revealed that eleven proteins with Sec-type signal peptides and<br />

type I signal peptidase cleavage sites (i.e. SAOUHSC-00094, SdrD, Sle1, Geh, Hlb, HlY,<br />

HlgB, HlgC, Plc, SAOUHC-02241 and SAOUHSC-02979) were present in significantly<br />

decreased amounts when SecG was absent from the cells. This was also true for the secreted<br />

moiety of the polytopic membrane protein YfnI, which is processed by signal peptidase I as<br />

was previously shown for the YfnI homologue of B. subtilis (Antelmann et al., 2001). In<br />

contrast, the amounts of three other exoproteins (i.e. IsaA, SsaA, and Spa) were considerably<br />

increased due to the secG deletion (Figure 2A; Table 3A). <strong>The</strong>se effects of the secG mutation<br />

were fully compensated when secG was ectopically expressed from plasmid secG-pCN51<br />

(Figure 2C). Northern blot analyses revealed similar transcript levels for geh, hlb and spa in<br />

the secG mutant and the parental strain RN4220. This shows that the changes in the amounts<br />

100


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

of the respective exoproteins in the secG mutant were not caused by a decreased transcription<br />

of the corresponding genes (Figure 3).<br />

Geh<br />

HlgB HlgC<br />

Sle1<br />

SAOUHSC_02241<br />

1<br />

2<br />

2<br />

1<br />

3<br />

1 2 4<br />

SssA<br />

3<br />

HlY<br />

Hlb<br />

SAOUHSC_00094<br />

Plc<br />

RN4220 OD 540 of 15<br />

RN4220 ∆secG OD 540 of 15<br />

4<br />

1 2<br />

5<br />

6<br />

SAOUHSC_02979<br />

1 2 3 4<br />

RN4220 OD 540 of 15<br />

RN4220 ∆secG secG-pCN51OD 540 of 15<br />

1<br />

2<br />

IsaA<br />

3<br />

2 Spa<br />

1<br />

SdrD<br />

A<br />

C<br />

HlgC<br />

101<br />

Geh<br />

1<br />

LipA<br />

SAOUHSC_02979<br />

1 2 3<br />

2<br />

2<br />

SAOUHSC_02241<br />

1<br />

2<br />

3<br />

HlY<br />

Hlb<br />

1 2<br />

1 2<br />

3 4<br />

SssA<br />

SAOUHSC_00094<br />

4<br />

Plc<br />

LytM<br />

RN4220 OD 540 of 15<br />

RN4220 ∆secG ∆secY2 OD 540 of 8<br />

Figure 2. <strong>The</strong> extracellular proteomes of S. <strong>aureus</strong> secG and secG secY2 mutants. (A) False-colored dualchannel<br />

image of 2D gels of extracellular proteins of S. <strong>aureus</strong> RN4220 (green) and S. <strong>aureus</strong> RN4220 ∆secG<br />

(red). Proteins (350 µg) isolated from the supernatant of S. <strong>aureus</strong> RN4220 and S. <strong>aureus</strong> RN4220 ∆secG grown in<br />

TSB medium to an OD 540 of 15 were separated on 2D gels by using immobilized pH gradient strips in the pH<br />

range of 3-10. Proteins were stained with colloidal Coomassie Brilliant Blue. Protein spots present in equal<br />

amounts in both strains appear in yellow, protein spots present in higher amounts in the secG mutant appear in<br />

red, and protein spots present in higher amounts in the parental strain appear in green. (B) False-colored dualchannel<br />

image of 2D gels of extracellular proteins of S. <strong>aureus</strong> RN4220 (green) and S. <strong>aureus</strong> RN4220 ∆secG<br />

∆secY2 (red). For experimental details see (A). Protein spots present in equal amounts in both strains appear in<br />

yellow, protein spots present in higher amounts in the secG secY2 mutant appear in red, and protein spots present<br />

in higher amounts in the parental strain appear in green. (C) False-colored dual-channel image of 2D gels of<br />

extracellular proteins of S. <strong>aureus</strong> RN4220 (green) and S. <strong>aureus</strong> RN4220 ∆secG secG-pCN51 (red). For<br />

experimental details see (A). All protein spots are yellow, indicating that both strains secreted the respective<br />

proteins in equal amounts.<br />

6<br />

1 2<br />

IsaA<br />

4<br />

3<br />

2<br />

1<br />

Spa<br />

B


Chapter 5<br />

Table 3A: Cell wall proteins with altered secretion patterns in S. <strong>aureus</strong> ∆secG and ∆secG∆secY2<br />

Protein a Function Mr/pI<br />

mature<br />

ORFID S. <strong>aureus</strong><br />

NCTC8325<br />

102<br />

Accession<br />

NCBI<br />

Relative level compared to parental strain c<br />

Predicted<br />

location b ∆secG/WT ∆secG ∆secY2/WT<br />

IsaA 1 immunodominant antigen A 24.2/6.6 SAOUHSC_02887 88196515 cell wall 2.0 4.2<br />

IsaA 2 immunodominant antigen A 24.2/6.6 SAOUHSC_02887 88196515 cell wall 2.4 8.6<br />

IsaA 3 immunodominant antigen A 24.2/6.6 SAOUHSC_02887 88196515 cell wall 2.9 4.8<br />

IsaA 4 immunodominant antigen A 24.2/6.6 SAOUHSC_02887 88196515 cell wall 12.4<br />

LytM peptidoglycan hydrolase, putative 34.3/6.7 SAOUHSC_00248 88194055 cell wall 4.3<br />

SAOUHSC_00094 1 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.2<br />

SAOUHSC_00094 2 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.3<br />

SAOUHSC_00094 3 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.4<br />

SAOUHSC_00094 4 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.3 0.3<br />

SAOUHSC_00094 5 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.5<br />

SAOUHSC_00094 6 hypothetical protein 21.8/9.4 SAOUHSC_00094 88193909 cell wall 0.2 0.3<br />

SdrD 1 SdrD protein, putative 14.6/3.9 SAOUHSC_00545 88194324 cell wall 0.3<br />

Sle1 (Aaa) autolysin precursor, putative 35.8/9.9 SAOUHSC_00427 88194219 cell wall 0.4<br />

Spa 1 protein A 55.6/5.4 SAOUHSC_00069 88193885 cell wall 3.7 3.3<br />

Spa 2 protein A 55.6/5.4 SAOUHSC_00069 88193885 cell wall 5.0 2.9<br />

SsaA secretory antigen precursor, putative 29.3/9.1 SAOUHSC_02571 88196215 cell wall 5.0 9.9<br />

a<br />

Several proteins are detectable as multiple spots. <strong>The</strong> spot numbers as marked in Figure 2 are indicated in superscript.<br />

b<br />

Protein localization was predicted as described in Sibbald et al. (Sibbald et al., 2006); SceD, SsaA and IsaA were shown to be bound ionically to the cell wall by Stapleton et al.<br />

(Stapleton et al., 2007); Sle1 (Aaa) has two LysM domains that can bind to peptidoglycan.<br />

c<br />

<strong>The</strong> induction ratio of mutant to parental strain was calculated for each spot (normalized intensity of a spot on the mutant image/normalized intensity of the corresponding spot on<br />

the parental image). <strong>The</strong> significance of spot volume differences of two-fold or higher was assessed by the Student´s t test (α


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in <strong>Staphylococcus</strong> <strong>aureus</strong><br />

Table 3B: Extracellular proteins with altered secretion patterns in S. <strong>aureus</strong> ∆secG and ∆secG∆secY2<br />

Protein a Function Mr/pI<br />

mature<br />

ORFID S. <strong>aureus</strong><br />

NCTC8325<br />

103<br />

Accession<br />

NCBI<br />

Relative level compared to parental strain c<br />

Predicted<br />

location b ∆secG/WT ∆secG ∆secY2/WT<br />

Geh lipase precursor 76.4/9.6 SAOUHSC_00300 88194101 extracellular 0.4 0.2<br />

Hlb 1 truncated β-hemolysin 31.3/8.2 SAOUHSC_02240 88195913 extracellular 0.1 0.2<br />

Hlb 2 truncated β-hemolysin 31.3/8.2 SAOUHSC_02240 88195913 extracellular 0.4 0.4<br />

HlY 1 α-hemolysin precursor 35.9/9.1 SAOUHSC_01121 88194865 extracellular 0.3 0.5<br />

HlY 2 α-hemolysin precursor 35.9/9.1 SAOUHSC_01121 88194865 extracellular 0.4 0.2<br />

HlY 3 α-hemolysin precursor 35.9/9.1 SAOUHSC_01121 88194865 extracellular 0.2<br />

HlY 4 α-hemolysin precursor 35.9/9.1 SAOUHSC_01121 88194865 extracellular 0.4 0.2<br />

HlgB leukocidin F subunit precursor 36.7/9.8 SAOUHSC_02710 88196350 extracellular 0.3<br />

HlgC leukocidin S subunit precursor, putative 35.7/9.7 SAOUHSC_02709 88196349 extracellular 0.3 0.4<br />

LipA 1 Lipase 76.7/7.7 SAOUHSC_03006 88196625 extracellular 0.5<br />

LipA 2 Lipase 76.7/7.7 SAOUHSC_03006 88196625 extracellular 0.3<br />

LipA 3 Lipase 76.7/7.7 SAOUHSC_03006 88196625 extracellular 0.4<br />

Plc 1-phosphatidylinositol phosphodiesterase<br />

precursor<br />

37.1/8.6 SAOUHSC_00051 88193871 extracellular<br />

0.3 0.3<br />

SAOUHSC_02241 1 hypothetical protein 38.7/9.1 SAOUHSC_02241 88195914 extracellular 0.1 0.2<br />

SAOUHSC_02241 2 hypothetical protein 38.7/9.1 SAOUHSC_02241 88195914 extracellular 0.3 0.4<br />

SAOUHSC_02241 3 hypothetical protein 38.7/9.1 SAOUHSC_02241 88195914 extracellular 0.2 0.4<br />

SAOUHSC_02979 1 hypothetical protein 69.3/6.3 SAOUHSC_02979 88196599 extracellular 0.4<br />

SAOUHSC_02979 2 hypothetical protein 69.3/6.3 SAOUHSC_02979 88196599 extracellular 0.3 0.3<br />

SAOUHSC_02979 3 hypothetical protein 69.3/6.3 SAOUHSC_02979 88196599 extracellular 0.3<br />

SAOUHSC_02979 4 hypothetical protein 69.3/6.3 SAOUHSC_02979 88196599 extracellular 0.5<br />

YfnI 1 polytopic membrane protein, signal<br />

peptidase I substrate<br />

74.4/9.5 SAOUHSC_00728 88194493 extracellular 0,4 0,2<br />

YfnI 2 polytopic membrane protein, signal<br />

peptidase I substrate<br />

74.4/9.5 SAOUHSC_00728 88194493 extracellular 0,4 0,2<br />

a<br />

Several proteins are detectable as multiple spots. <strong>The</strong> spot numbers as marked in Figure 2 are indicated in superscript.<br />

b<br />

Protein localization was predicted as described in Sibbald et al. (Sibbald et al., 2006); SceD, SsaA and IsaA were shown to be bound ionically to the cell wall by Stapleton et al.<br />

(Stapleton et al., 2007); Sle1 (Aaa) has two LysM domains that can bind to peptidoglycan.<br />

c<br />

<strong>The</strong> induction ratio of mutant to parental strain was calculated for each spot (normalized intensity of a spot on the mutant image/normalized intensity of the corresponding spot on<br />

the parental image). <strong>The</strong> significance of spot volume differences of two-fold or higher was assessed by the Student´s t test (α


Chapter 5<br />

geh<br />

RN4220 RN4220∆secG<br />

OD1 OD10 OD15 OD1 OD10 hlb<br />

OD15<br />

RN4220 RN4220∆secG<br />

OD 1 OD10 OD 15<br />

spa<br />

OD1 OD10 OD15<br />

RN4220 RN4220∆secG<br />

OD 1 OD10 OD15<br />

OD1 OD10<br />

OD 15<br />

OD 1<br />

OD1<br />

OD1<br />

OD10<br />

104<br />

OD 10<br />

OD 10<br />

OD 15<br />

Dual image<br />

RN4220<br />

RN4220 ∆secG<br />

OD 15<br />

OD 15<br />

Dual image<br />

RN4220<br />

RN4220 ∆secG<br />

Dual image<br />

RN4220<br />

RN4220 ∆secG<br />

Figure 3. Expression of SecG-dependent exoproteins. RNA and exoproteins were collected from S. <strong>aureus</strong><br />

RN4220 and S. <strong>aureus</strong> RN4220 ∆secG grown in TSB medium at 37°C. Samples were collected at three different<br />

points during growth (OD 540 of 1, 10 and 15). In the Northern blotting experiments, membranes were hybridized<br />

with digoxigenin-labeled RNA probes specific for geh, hlb or spa. Protein spots from 2-D PAGE analyses of the<br />

respective proteins collected at OD 540 of 1, 10, and 15 are shown for the secG mutant and its parental strain both<br />

separately and as dual-channel images.<br />

Deletion of the secY2 gene encoding a channel component of the accessory Sec<br />

system in S. <strong>aureus</strong>, did not affect the extracellular protein pattern (data not shown).<br />

However, the deletion of both secG and secY2 caused additional changes in the extracellular<br />

proteome compared to the secG single mutant (Figure 2B). Specifically, one additional<br />

exoprotein was identified in decreased amounts (i.e. LipA) and one additional exoprotein (i.e.<br />

LytM) was identified in increased amounts (Table 3A and 3B). Furthermore, proteins such as<br />

Spa, IsaA, and SsaA were secreted in higher amounts not only by the secG mutant, but also<br />

by the secG secY2 double mutant. This effect was significantly exacerbated for IsaA and<br />

SsaA in the secG secY2 double mutant. It is interesting to note that IsaA, LytM, Spa, and<br />

SsaA represent cell surface-associated proteins (Schneewind et al., 1992; Ramadurai et al.,<br />

1999; Stapleton et al., 2007).<br />

In contrast, most proteins that were secreted in reduced amounts in the secG or secG secY2<br />

mutants are secretory proteins without retention signals, except for SAOUHSC-00094 (Table<br />

3A and 3B). Importantly, also the secretion and growth defects of the secG secY2 mutant<br />

strain could be fully reversed by ectopic expression of secG from the plasmid secG-pCN51,<br />

and the synthetic effects of the secG and secY2 mutations could be reversed by plasmid<br />

secY2-pCN51 (data not shown).


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

Impaired export of cell wall-bound Sbi in secG mutant cells<br />

Western blotting experiments were performed to investigate whether particular protein export<br />

defects of the secG and secY2 mutants had remained unnoticed in the proteomic analyses.<br />

<strong>The</strong>se analyses included secreted proteins in the growth medium, non-covalently cell wall<br />

attached and cellular proteins of S. <strong>aureus</strong> strains RN4220 and S. <strong>aureus</strong> SH1000.<br />

Furthermore, we used specific antibodies against membrane proteins, lipoproteins, cell wall<br />

proteins and exoproteins. For most tested proteins no differences were detectable between the<br />

secG and/or secY2 mutant strains and their parental strain. However, these analyses showed<br />

that a band of ~50 kDa, which was cross-reactive with all tested sera, had disappeared from<br />

the fraction of non-covalently bound cell wall proteins of the secG mutant. It is known that<br />

proteins, such as protein A (Sasso et al., 1991) and Sbi (Zhang et al., 1998) have IgG-binding<br />

properties. To investigate whether the missing band would relate to protein A or Sbi, protein<br />

fractions from a spa mutant, and a spa sbi double mutant, were included in the Western<br />

blotting analyses.<br />

Figure 4. Sbi localization to the cell wall of S. <strong>aureus</strong> depends on SecG. (A) S. <strong>aureus</strong> SH1000 (WT), S. <strong>aureus</strong><br />

SH1000 ∆secG, and S. <strong>aureus</strong> SH1000 ∆secG secG-pCN51 were grown in TSB medium at 37 o C till the early<br />

stationary phase. Samples of extracellular proteins isolated from the growth medium (M), non-covalently cell wallbound<br />

proteins (CW) and total cells (C) were used for Western blotting and immunodetection with serum of mice<br />

immunized with IsaA. As a contol for Sbi production, the strains S. <strong>aureus</strong> Newman ∆spa and S. <strong>aureus</strong> Newman<br />

∆spa ∆sbi were included in the analyses. (B) Proteins of S. <strong>aureus</strong> SH1000 (wt), S. <strong>aureus</strong> SH1000 ∆secG, and S.<br />

<strong>aureus</strong> SH1000 ∆secY2 were used for Western blotting and immunodetection as in (A). <strong>The</strong> position of Sbi is<br />

marked with an arrow.<br />

As shown in Figure 4A, the band of ~50 kDa that was missing from the non-covalently bound<br />

cell wall proteins in the secG mutant was also missing from these proteins in the spa sbi<br />

double mutant, but not in the spa single mutant (only the results for S. <strong>aureus</strong> SH1000 are<br />

shown but essentially the same results were obtained for S. <strong>aureus</strong> RN4220). Taken together,<br />

these findings show that Sbi is non-covalently bound to the cell wall of S. <strong>aureus</strong> RN4220 and<br />

SH1000, and that SecG is required for export of Sbi from the cytoplasm to the cell wall. As<br />

was the case for the secreted S. <strong>aureus</strong> proteins detected by proteomics, Sbi export to the cell<br />

wall was not affected by the absence of SecY2 (Figure 4B). Finally, it is noteworthy that Sbi<br />

is only detectable amongst the non-covalently bound cell wall proteins of S. <strong>aureus</strong> RN4220<br />

and SH1000, whereas it is detectable both in a cell wall-bound and a secreted state in S.<br />

<strong>aureus</strong> Newman.<br />

105


Chapter 5<br />

Discussion<br />

<strong>The</strong> extracellular and surface-associated proteins of bacterial pathogens, such as S. <strong>aureus</strong>,<br />

represent an important reservoir of virulence factors. Accordingly, protein export mechanisms<br />

will contribute to the virulence of these organisms. While protein export has been well<br />

characterized in model organisms, such as E. coli and B. subtilis, relatively few functional<br />

studies have addressed the protein export pathways of S. <strong>aureus</strong>. Notably, the Sec pathway is<br />

generally regarded as the main pathway for protein export but, to date, this has not been<br />

verified experimentally in S. <strong>aureus</strong>. <strong>The</strong>refore, the present studies were aimed at assessing<br />

the role of the Sec pathway in establishing the extracellular proteome of S. <strong>aureus</strong>. We<br />

focused attention on the non-essential channel component SecG as this allowed a facile coassessment<br />

of the non-essential accessory Sec channel component SecY2. Our results show<br />

that the extracellular accumulation of proteins is affected to different extents by the absence<br />

of SecG: some proteins are present in reduced amounts, some are not affected and some are<br />

present in elevated amounts. Furthermore, the effects of the absence of SecG are exacerbated<br />

by deletion of SecY2, suggesting that SecY2 directly or indirectly influences the functionality<br />

of the general Sec pathway. This is all the more remarkable since the absence of SecY2 by<br />

itself had no detectable effects on the composition of the extracellular proteome of S. <strong>aureus</strong>.<br />

<strong>The</strong> observation that the secretion of a wide range of proteins was affected by the<br />

absence of SecG is consistent with the fact that all of these proteins contain Sec-type signal<br />

peptides.On the other hand, this finding is remarkable since studies in other organisms, such<br />

as E. coli (Nishiyama et al., 1993) and B. subtilis (Van Wely et al., 1999) have shown that<br />

deletion of secG had fairly moderate effects on protein secretion in vivo. In B. subtilis, a<br />

phenotype of the secG mutation was only observed under conditions of high overproduction<br />

of secretory proteins (Van Wely et al., 1999). Clearly, our present data show that SecG is<br />

more important for Sec-dependent protein secretion in S. <strong>aureus</strong> than in B. subtilis or E. coli.<br />

Importantly, the transcription of genes for three proteins (Geh, Hlb and Spa) that were<br />

affected in major ways by the absence of SecG was not changed, and all observed effects of<br />

the secG mutation could be reversed by ectopic expression of secG. This suggests that the<br />

observed changes in the exoproteome composition of the S. <strong>aureus</strong> secG mutant strain relate<br />

to changes in the translocation efficiency of proteins through the Sec channel rather than<br />

regulatory responses at the gene expression level. This could be due to altered recognition of<br />

the respective signal peptides or mature proteins by the SecG-less Sec channel, or<br />

combinations thereof. However, some indirect effects, for example at the level of translation<br />

of exported proteins or cell wall binding of proteins like IsaA, LytM, Spa and SsaA, can<br />

currently not be excluded especially since no proteins were found to accumulate inside the<br />

secG mutant cells (data not shown). It remains to be shown why the extracellular<br />

accumulation of particular proteins is affected by the absence of SecG, while that of other<br />

proteins remains unaffected.<br />

Unexpectedly, our studies revealed that export of the IgG-binding protein Sbi to the<br />

cell wall was almost completely blocked in secG mutant strains. <strong>The</strong> reason why this export<br />

defect was not detected by 2-D PAGE relates to the fact that Sbi is predominantly cell wallbound<br />

in the tested S. <strong>aureus</strong> strains under the experimental conditions used. It has been<br />

proposed previously that Sbi would remain cell wall-attached through a proline-rich wallbinding<br />

domain and electrostatic interactions (Zhang et al., 1998). Nevertheless, Burman and<br />

colleagues showed that Sbi is extracellular and they suggested that cell surface-bound Sbi<br />

might be disadvantageous for the bacterium due to its role in modulating the complement<br />

106


Synthetic effects of secG and secY2 mutations on exoproteome biogenesis in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong><br />

system (Burman et al., 2008). On the other hand, cell surface localization of Sbi would be<br />

appropriate for interference with the adaptive immune system through IgG binding (Atkins et<br />

al., 2008). Irrespective of these previously reported findings, our Western blotting analyses<br />

show that Sbi is non-covalently bound to the cell wall, not only in S. <strong>aureus</strong> SH1000 and S.<br />

<strong>aureus</strong> RN4220, but also in S. <strong>aureus</strong> Newman. However, consistent with the findings of<br />

Burman et al., Sbi was also detected in the growth medium of S. <strong>aureus</strong> Newman, which<br />

indicates that the location of Sbi in the cell wall or extracellular milieu may differ for different<br />

S. <strong>aureus</strong> strains. In case of the Newman strain, the release of Sbi into the growth medium<br />

could be due to the fact that this strain produces Sbi at increased levels compared to the<br />

RN4220 and SH1000 strains (Rogasch et al., 2006). Conceivably, this increased production<br />

might lead to a saturation of available cell wall binding sites for Sbi.<br />

Many of the proteins of which the extracellular amounts are changed due to the<br />

absence of SecG are considered to be important virulence factors of S. <strong>aureus</strong>. <strong>The</strong>se proteins<br />

are involved host colonization (e.g. the serine-aspartic acid repeat proteins SdrC and SdrD),<br />

invasion of host tissues (e.g. hemolysins and leukocidins), cell wall turnover (LytM), and<br />

evasion of the immune system (Spa, Sbi). <strong>The</strong> altered amounts of these proteins suggest that<br />

S. <strong>aureus</strong> strains depleted of SecG might perhaps be less virulent. However, in a mouse<br />

infection model no changes in virulence of the S. <strong>aureus</strong> secG mutant strain could be detected<br />

(data not shown). This implies that the presence or absence of SecG is not critical for S.<br />

<strong>aureus</strong> virulence, at least under the conditions tested.<br />

Since we were unable to detect secretion defects for secY2 single mutant strains, our<br />

studies confirm that only very few proteins are translocated across the membrane in a<br />

SecA2/SecY2-dependent manner as has previously been suggested by Siboo et al. (Siboo et<br />

al., 2008). Furthermore, we did not detect differences in the export of glycosylated proteins<br />

by the secY2 mutants (data not shown), which is in line with the suggestion that glycosylated<br />

proteins are not strictly dependent on the accessory Sec pathway for export (Siboo et al.,<br />

2008). It was therefore quite surprising that the secY2 mutation exacerbated the secretion<br />

defect of the S. <strong>aureus</strong> secG mutant. In fact, the secretion of two additional proteins was<br />

found to be affected in the secG secY2 double mutant. Moreover, a synthetic growth defect<br />

was observed for this double mutant. At this stage, it seems most likely that both the growth<br />

defect and the secretion defects are consequences of an impaired Sec channel function.<br />

However, it is also possible that the exacerbated secretion defects are, to some extent, a<br />

secondary consequence of the growth defect of the double mutant. Irrespective of their<br />

primary cause, these synthetic effects of the secG and secY2 mutations imply that the regular<br />

Sec channel can somehow interact with the Sec2 channel. Whether this means that mixed Sec<br />

channels with both SecY and SecY2 exist remains to be determined. However, this possibility<br />

would be consistent with the observation that S. <strong>aureus</strong> lacks a second set of secE and secG<br />

genes. It would thus be important to focus future research activities in this area on possible<br />

interactions between the regular Sec channel components and SecY2.<br />

107


Chapter 5<br />

Acknowledgements<br />

We like to thank W. Baas and M. ten Brinke for technical assistance, S. Dubrac for providing<br />

the pCN51 plasmid, T. Foster for the spa mutant of S. <strong>aureus</strong> Newman, Decodon GmbH<br />

(Greifswald, Germany) for providing Delta2D software, and T. Msadek and other colleagues<br />

from the StaphDynamics and AntiStaph programs for advice and stimulating discussions.<br />

M.J.J.B.S, T.W., M.M.v.d.K.-P., T.B., T.S., K.O., M.H., H.A., S.E. and J.M.vD. were in parts<br />

supported by the CEU projects LSHM-CT-2006-019064, LSHG-CT-2006-037469 and PITN-<br />

GA-2008-215524, the Top Institute <strong>Pharma</strong> project T4-213, and the DFG research grants<br />

GK212/3-00, SFB/TR34 and FOR585.<br />

108


109


“One good thing about music:when it hits you feel no pain”<br />

-Robert N. Marley-<br />

110


Chapter 6<br />

Partially overlapping substrate specificities of sortases A and C of<br />

staphylococci<br />

M.J.J.B. Sibbald # , X.M. Yang # , E. Tsompanidou, M. Hecker, D. Becher,<br />

G. Buist, J.M. van Dijl<br />

# both authors contributed equally to this work<br />

Submitted for publication<br />

111


Chapter 6<br />

Summary<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> and <strong>Staphylococcus</strong> epidermidis display many proteins on their<br />

cell surface that are covalently linked to the peptidoglycan moiety by sortases. <strong>The</strong> class<br />

A sortase (SrtA) of S. <strong>aureus</strong> and S. epidermidis attaches proteins with an LPxTG motif<br />

to the cell wall. Deletion of the srtA genes of these bacteria resulted in the dislocation of<br />

several LPxTG proteins, such as ClfA, SasG, SdrC, SdrD, and protein A of S. <strong>aureus</strong>,<br />

and Aap of S. epidermidis 1457, to the growth medium. Nevertheless, substantial<br />

amounts of these proteins remained cell wall-bound through non-covalent interactions.<br />

<strong>The</strong> protein dislocation phenotypes of srtA mutations in S. <strong>aureus</strong> and S. epidermidis<br />

could be fully reverted by ectopic expression of srtA genes of either species.<br />

Interestingly, the class C sortase (SrtC) from S. epidermidis 12228 was able to revert the<br />

dislocation of ClfA, SasG and Aap to significant extents, showing that the substrate<br />

specificities of SrtA and SrtC overlap at least partially. Interestingly, biofilm formation<br />

was affected in srtA mutants of S. <strong>aureus</strong>, but not in the S. epidermidis srtA mutant. This<br />

difference can be correlated to the expression levels of particular covalently cell wallbound<br />

proteins involved in protein-dependent biofilm formation, such as S. <strong>aureus</strong> SasG<br />

and its S. epidermidis homologue Aap. Remarkably, SrtA activity was a limiting<br />

determinant for protein-dependent biofilm formation in S. <strong>aureus</strong> and S. epidermidis,<br />

whereas SrtC expression interfered with biofilm formation in S. epidermidis 1457. Taken<br />

together, these findings imply that sortases can have modulating roles in staphylococcal<br />

biofilm formation.<br />

112


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

Introduction<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> and <strong>Staphylococcus</strong> epidermidis are both part of the normal human<br />

microbiota (Peacock et al., 2001;Wertheim et al., 2004;Bibel et al., 1976). However, both<br />

organisms have the potential to cause life-threatening infections, especially when they<br />

become invasive and reach the blood stream. In addition, S. <strong>aureus</strong> and especially S.<br />

epidermidis are notorious for their ability to form biofilms on medical devices and implants<br />

(Escher and Characklis, 1990). Cell surface-associated proteins play crucial roles in the<br />

colonization and invasion of host tissues by staphylococci, and such proteins also have<br />

important roles in biofilm formation. <strong>The</strong> surface-exposed proteins can either be covalently or<br />

non-covalently linked to the bacterial cell wall. Covalent protein linkage to the peptidoglycan<br />

moiety of the cell wall is catalyzed by specific enzymes known as sortases.<br />

Gram-positive bacteria, such as S. <strong>aureus</strong> and S. epidermidis, but also various bacilli,<br />

streptococci and corynebacteria have one or more sortase-encoding genes (Dramsi et al.,<br />

2005;Sibbald and van Dijl, 2009). For several pathogens, such as S. <strong>aureus</strong>, Listeria<br />

monocytogenes and Streptococcus pneumoniae, it has been shown that the deletion of sortase<br />

genes results in decreased virulence (Mazmanian et al., 2000;Garandeau et al., 2002;Bierne et<br />

al., 2002;Hava and Camilli, 2002). This underscores the importance of covalent cell wall<br />

attachment of particular proteins in the pathogenicity of Gram-positive bacteria. Based on<br />

structural and functional criteria, Dramsi and colleagues (Dramsi et al., 2005) classified<br />

sortases into four different groups. Sortase A (SrtA) enzymes link several proteins with<br />

LPxTG or LPxAG motifs to the cell wall. Such proteins are mainly involved in adhesion to<br />

specific organ tissue, survival during phagocytosis, and invasion of host cells. <strong>The</strong> LPxTG or<br />

LPxAG motifs are cleaved by sortase between the Thr/Ala and Gly residues. <strong>The</strong> free Thr<br />

residue is then linked to the active site Cys residue of sortase A (Marraffini et al., 2006;Scott<br />

and Barnett, 2006). Formation of an amide bond between the Thr residue of the surface<br />

protein and the pentaglycine cross-bridge of branched lipid II completes the covalent protein<br />

attachment to the cell wall. For S. <strong>aureus</strong> and other Gram-positive pathogens it has been<br />

shown that srtA mutants have a decreased ability to establish a successful infection in animal<br />

models (Bierne et al., 2002;Mazmanian et al., 2000). In contrast to sortase A, most sortase B<br />

(SrtB) enzymes are involved in iron metabolism. In S. <strong>aureus</strong> the srtB gene is part of the isd<br />

operon, which contains several genes (isdA-G) that are important for iron-acquisition<br />

(Mazmanian et al., 2003). SrtB recognizes the NPQTN motif in the C-terminus of the IsdC<br />

protein, and cleaves this motif C-terminally of the Thr residue thereby linking IsdC to the cell<br />

wall (Mazmanian et al., 2002). Class C sortases (SrtC) are found in several Gram-positive<br />

bacteria and these sortases seem to be involved in the formation of pili (Mandlik et al.,<br />

2008;Ton-That and Schneewind, 2004;Scott and Barnett, 2006). In S. pneumoniae three class<br />

C sortases and their substrates were shown to be important for pilus formation and virulence<br />

in animal models (Hava and Camilli, 2002). Notably, the genes encoding class C sortases and<br />

their substrates are not part of the core genomes of particular species. This suggests that these<br />

genes have been acquired through horizontal gene transfer thereby giving the cells an<br />

adaptive advantage to certain host niches. Several Gram-positive bacteria have a class D<br />

sortase (Dramsi et al., 2005) but, so far, relatively little is known about this class of sortases.<br />

<strong>The</strong> class D sortase of Streptomyces coelicolor links several proteins to the surface of the cell<br />

wall. <strong>The</strong>se proteins contain so-called chaplin domains and have been shown to coat the<br />

surfaces of aerial hyphae and spores (Claessen et al., 2003;Elliot et al., 2003). Bacillus<br />

subtilis 168 also has a gene for a SrtD homologue named yhcS. However, no clear phenotype<br />

113


Chapter 6<br />

for yhcS mutant strains could be identified, and the only two B. subtilis proteins with typical<br />

sortase recognition sites, YhcR and YfkN, were found to be secreted both by the yhcS mutant<br />

and the parental strain 168 (H. Westers, doctoral thesis, University of Groningen).<br />

<strong>The</strong> crystal structure of SrtA revealed that the active site Cys residue is in close proximity of a<br />

His residue (Ilangovan et al., 2001;Zong et al., 2004;Suree et al., 2009;Race et al., 2009).<br />

Furthermore, an Arg residue in the C-terminus of the protein also seems to be important for<br />

efficient catalysis (Marraffini et al., 2004;Frankel et al., 2007;Bentley et al., 2007), probably<br />

by stabilizing the transition state. <strong>The</strong> Cys residue is part of the TLxTC motif that is<br />

conserved in all sortases (Figure 1). While the Thr-180, Leu-181 and Ile-182 of SrtA seem to<br />

determine the conformation of the substrate binding pocket, Thr-183 is most likely involved<br />

in positioning the Cys-184 and Arg-197 residues. Other residues that might play a role in<br />

efficiently binding surface proteins to the cell wall include Val-168 and Leu-169, which are<br />

involved in substrate recognition. <strong>The</strong>se residues recognize the Leu-Pro residues in the Cterminal<br />

LPxTG motif by hydrophobic interactions (Bentley et al., 2007). Also, Glu-171<br />

seems to be important for the function of SrtA and it was proposed that this residue binds<br />

Ca 2+ , thereby stabilizing the β6/β7 loop region of SrtA (Naik et al., 2006).<br />

Figure 1. Alignment of staphylococcal sortases A, B and C. Sortases of S. <strong>aureus</strong> NCTC8325 and S. epidermidis<br />

ATCC12228 were aligned using the ClustalW (EMBL-EBI) algorithm with standard settings (Thompson et al.,<br />

1994). <strong>The</strong> database accession codes of the aligned sortase sequences are: S. <strong>aureus</strong> SrtA (gi 88196468); S. <strong>aureus</strong><br />

SrtB (gi 88194835); S. epidermidis SrtA (gi 27468994); and S. epidermidis SrtC (gi 27468398). Residues<br />

conserved in all sortases are marked with black shading, and residues present in three of the four sequences are<br />

marked with grey shading. Active site residues are marked with #.<br />

All sequenced <strong>Staphylococcus</strong> species contain multiple genes for proteins with cell wall<br />

sorting motifs. In total, twenty-one proteins with this motif have been identified in S. <strong>aureus</strong><br />

and twelve in S. epidermidis (Sibbald et al., 2006). One of the proteins with a C-terminal<br />

LPxTG motif is the <strong>Staphylococcus</strong> <strong>aureus</strong> surface protein G (SasG; (Roche et al., 2003)). It<br />

has been shown that this protein is involved in adhesion to nasal epithelial cells. Interestingly,<br />

SasG is also involved in protein-based biofilm formation, independent of the ica-encoded<br />

polysaccharide for intercellular adhesion (PIA) or poly-N-acetyl glucosamine (PNAG)<br />

(Corrigan et al., 2007). <strong>The</strong> homologue of SasG in S. epidermidis is the accumulation<br />

associated protein (Aap). This protein has been implicated in the formation of both<br />

polysaccharide- (Hussain et al., 1997) and protein-based biofilms (Rohde et al., 2005;Rohde<br />

et al., 2007). SasG and Aap contain several functional domains. Both SasG and Aap are<br />

synthesized with a classical signal peptide containing the YSIRK/GS motif in the N-terminal<br />

part of the signal peptide. Dedent et al. (Dedent et al., 2008) have provided evidence that the<br />

YSIRK/GS motif directs site-specific secretion at the cross wall, which is the peptidoglycan<br />

layer that is formed during cell division to separate new daughter cells. Proteins without this<br />

motif are addressed to the cell pole of S. <strong>aureus</strong>. <strong>The</strong> N-terminal A domain of SasG is<br />

114


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

unrelated to the equivalent domain in Aap, but it is followed by a stretch of amino acids that<br />

is 59% identical to the equivalent domain in Aap. Next, several repeats of 128 residues are<br />

present that are known as G5 domains. <strong>The</strong>se domains have been shown to bind N-acetyl<br />

glucosamine (Bateman et al., 2005). It has been shown that these domains are necessary for<br />

SasG- or Aap-dependent biofilm formation, and that they need to be proteolytically separated<br />

from the respective mature proteins to fulfill their roles in intercellular adherence (Rohde et<br />

al., 2005;Corrigan et al., 2007). Furthermore, it has been shown for S. <strong>aureus</strong> SasG that at<br />

least five G5 domains are necessary for biofilm formation (Corrigan et al., 2007). <strong>The</strong> G5<br />

domains of both SasG and Aap promote the intercellular interactions in a Zn 2+ -dependent<br />

manner (Conrady et al., 2008).<br />

In contrast to S. epidermidis RP62A, the S. epidermidis ATCC12228 strain contains a srtC<br />

gene (Comfort and Clubb, 2004). This srtC gene lies on a genomic island (νSe2) that also<br />

contains the genes for the surface proteins SesJ and SesK (Gill et al., 2005). Both proteins<br />

contain an LPxTG motif in their C-termini. Unfortunately, the biological roles of SesJ and<br />

SesK are presently unclear. Class C sortases have also been found in a few other Grampositive<br />

bacteria where they recognize proteins with a C-terminal LPxTG motif for covalent<br />

cell wall binding (Dramsi et al., 2005). In C. diphtheriae it has been shown that SrtC is<br />

necessary for the formation of pili (Ton-That and Schneewind, 2003). Notably, the three pilus<br />

subunits SpaA, SpaB and SpaC are synthesized with a C-terminal LPxTG motif. SrtC is<br />

responsible for the processing of these subunits and their linkage to a neighboring subunit.<br />

Homologues of these proteins have also been identified in other bacteria, such as S.<br />

pneumoniae, Streptococcus agalactiae and Enterococcus faecalis (Dramsi et al., 2005). It<br />

seems unlikely that SrtC fulfils a similar role in S. epidermidis as pilus formation has not been<br />

observed in this bacterium. Importantly, whereas SrtA seems to handle a wide range of<br />

different substrates with LPxTG motifs, SrtC seems to be dedicated to the processing of only<br />

a few specific substrates, in some cases even only one substrate (Barnett et al., 2004;Hava and<br />

Camilli, 2002;Hava et al., 2003;Ton-That and Schneewind, 2003).<br />

<strong>The</strong> present studies were aimed at addressing the question to what extent sortases of different<br />

classes and from different Firmicutes, like S. <strong>aureus</strong>, S. epidermidis and B. subtilis, can<br />

functionally replace each other. While the class D sortase of B. subtilis does not seem to<br />

complement for the absence of SrtA in S. <strong>aureus</strong> or S. epidermidis, the SrtA proteins from S.<br />

<strong>aureus</strong> and S. epidermidis can complement for each other. Remarkably, our results show that<br />

SrtC of S. epidermidis can partially complement for the absence of SrtA in both<br />

<strong>Staphylococcus</strong> species. This implies that class A and class C type sortases of staphylococci<br />

have partially overlapping substrate specificities.<br />

Material & Methods<br />

Bacterial strains<br />

All strains and plasmids used in this study are listed in Table 1. Escherichia coli strains were grown in<br />

Luria-Bertani broth (LB) at 37°C and under vigorous shaking. Unless stated otherwise, S. <strong>aureus</strong> and S.<br />

epidermidis strains were grown on tryptic soy broth (TSB), tryptic soy agar (TSA) or brain heart<br />

infusion (BHI) at 37°C and under vigorous shaking. Where necessary, antibiotics were added in the<br />

following concentrations: ampicillin (100 µg/ml), erythromycin for E. coli (100 µg/ml), erythromycin<br />

for S. <strong>aureus</strong> and S. epidermidis (5 µg/ml), kanamycin (20 µg/ml), chloramphenicol (15 µg/ml). To<br />

monitor β-galactosidase activity in cells of E. coli, S. <strong>aureus</strong> or S. epidermidis, 5-bromo-4-chloro-3indolyl-β-D-galactopyranoside<br />

(X-gal) was added to the plates at a final concentration of 80 µg/ml.<br />

115


Chapter 6<br />

Table 1. Bacterial strains and plasmids used<br />

Plasmids Properties Reference<br />

TOPO pCR®-Blunt II-TOPO® vector; Km R Invitrogen Life technologies<br />

“srtA”-TOPO TOPO plasmid containing the flanking regions of S.<br />

<strong>aureus</strong> srtA, Kan R<br />

This work<br />

srtC-pUC18 pUC18 plasmid with the S. epidermidis srtC gene, Amp R This work<br />

“srtA”SE-TOPO TOPO plasmid containing the flanking regions of S.<br />

epidermidis srtA, Kan R<br />

This work<br />

pCN51 E. coli - S. <strong>aureus</strong> shuttle vector that contains a cadmiuminducible<br />

promoter<br />

(Charpentier et al., 2004)<br />

pMAD E. coli - S. <strong>aureus</strong> shuttle vector that is temperaturesensitive<br />

in S. <strong>aureus</strong> and contains the bgaB gene, Ery R ,<br />

(Arnaud et al., 2004)<br />

Amp R<br />

pUC18 Amp R , ColE1, F80dLacZ, lac promoter (Norrander et al., 1983)<br />

Sa-srtA-pCN51 pCN51 with S. <strong>aureus</strong> srtA gene, Amp R ; Ery R This work<br />

Se-srtA-pCN51 pCN51 with S. epidermidis srtA gene, Amp R ; Ery R This work<br />

Se-srtC-pCN51 pCN51 with S. epidermidis ATCC12228 srtC gene,<br />

Amp R ; Ery R<br />

This work<br />

Bs-yhcS-pCN51 pCN51 with B. subtilis yhcS gene, Amp R ; Ery R This work<br />

Strains<br />

E. coli<br />

Genotype Reference<br />

DH5α supE44; hsdR17; recA1; gyrA96; thi-1; relA1 (Hanahan, 1983)<br />

TOP10 Cloning host for TOPO vector; F - mcrA ∆(mrr-hsdRMSmcrBC)<br />

Φ80lacZ∆M15 ∆lacX74 recA1 araD139 ∆(araleu)7697<br />

galU galK rpsL (Str R S. <strong>aureus</strong> RN4220<br />

) endA1 nupG<br />

Invitrogen Life technologies<br />

Parental strain Restriction-deficient derivative of NCTC 8325, cured of<br />

all known prophages<br />

(Kreiswirth et al., 1983)<br />

∆srtA srtA This work<br />

∆srtA ∆spa srtA spa; replacement of spa by kanamycin resistance<br />

marker; Kan R<br />

This work<br />

∆srtA-comp-Sa ∆srtA strain complemented with S. <strong>aureus</strong> srtA through<br />

Sa-srtA-pCN51<br />

This work<br />

∆srtA-comp-Se ∆srtA strain complemented with S. epidermidis srtA<br />

through Se-srtA-pCN51<br />

This work<br />

∆yhcS-comp-Bs ∆srtA strain complemented with B. subtilis yhcS through<br />

Bs-yhcS-pCN51<br />

This work<br />

∆srtA-comp-srtC<br />

S. <strong>aureus</strong> SH1000<br />

∆srtA strain complemented with S. epidermidis srtC<br />

throught Se-srtC-pCN51<br />

This work<br />

Parental strain Functional rsbU+ derivative of 8325-4 rsbU+, agr+ (Horsburgh et al., 2002)<br />

∆srtA rsbU+, agr+, srtA This work<br />

∆srtA ∆spa rsbU+, agr+, srtA spa; replacement of spa by kanamycin<br />

resistance marker; Kan R<br />

This work<br />

∆srtA-comp-Sa ∆srtA complemented with S. <strong>aureus</strong> srtA through SasrtA-pCN51<br />

This work<br />

∆srtA-comp-Se ∆srtA complemented with S. epidermidis srtA through<br />

Se-srtA-pCN51<br />

This work<br />

∆yhcS-comp-Bs ∆srtA complemented with B. subtilis yhcS through BsyhcS-pCN51<br />

This work<br />

∆srtA-comp-srtC ∆srtA complemented with S. epidermidis srtC through<br />

Se-srtC-pCN51<br />

This work<br />

sasG-pMU<strong>TI</strong>N4 Overexpression of SasG due to integrated pMU<strong>TI</strong>N4<br />

plasmid; Ery R<br />

(Corrigan et al., 2007)<br />

sasG-pMU<strong>TI</strong>N4<br />

∆srtA<br />

∆srtA; overexpression of SasG due to integrated<br />

pMU<strong>TI</strong>N4 plasmid; Ery R<br />

116<br />

This Work


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

Strains Genotype Reference<br />

sasG-pMU<strong>TI</strong>N4 ∆srtA; overexpression of SasG due to integrated<br />

∆srtA comp-srtA pMU<strong>TI</strong>N4 plasmid; complemented with S. <strong>aureus</strong> srtA<br />

through Sa-srtA-pRIT5H; Ery R , Cm R<br />

This Work<br />

sasG-pMU<strong>TI</strong>N4 ∆srtA; overexpression of SasG due to integrated<br />

∆srtA comp-srtC pMU<strong>TI</strong>N4 plasmid; complemented with S. epidermidis<br />

srtC through Se-srtC-pRIT5H; Ery R , Cm R<br />

This Work<br />

S. epidermidis 1457<br />

Parental Biofilm positive strain (Mack et al., 1992)<br />

∆srtA srtA This work<br />

∆srtA-comp-Sa ∆srtA; complemented with S. <strong>aureus</strong> srtA through SasrtA-pCN51<br />

This work<br />

∆srtA-comp-Se ∆srtA; complemented with S. epidermidis srtA through<br />

Se-srtA-pCN51<br />

This work<br />

∆yhcS-comp-Bs ∆srtA; complemented with B. subtilis yhcS through BsyhcS-pCN51<br />

This work<br />

∆srtA-comp-srtC ∆srtA; complemented with S. epidermidis srtC through<br />

Se-srtC-pCN51<br />

This work<br />

Construction of sortase mutants<br />

Mutants of S. <strong>aureus</strong> and S. epidermidis were constructed using the temperature-sensitive plasmid<br />

pMAD (Arnaud et al., 2004) and previously described procedures (Kouwen et al., 2009). Primers<br />

(Table 2) were designed using the genome sequences of S. <strong>aureus</strong> NCTC8325 and S. epidermidis<br />

RP62A (http://www.ncbi.nlm.nih.gov/genomes/lproks.cgi). All mutant strains were checked by<br />

isolation of genomic DNA using the GenElute Bacterial Genomic DNA Kit (Sigma) and PCR with<br />

specific primers.<br />

To delete the srtA genes, primer pairs with the designations F1/R1 and F2/R2 were used for<br />

PCR amplification of the respective upstream and downstream regions (each ~500 bp), and their fusion<br />

with a 21 bp linker. <strong>The</strong> fused flanking regions were cloned in pMAD, and the resulting plasmids were<br />

used to delete the chromosomal srtA genes of S. <strong>aureus</strong> RN4220 and S. epidermidis 1457. To delete the<br />

srtA gene from the S. <strong>aureus</strong> SH1000 genome, the respective pMAD constructs were transferred from<br />

the RN4220 strain to the SH1000 strain by transduction with phage φ85 (Novick, 1991).<br />

Complementation of srtA mutations<br />

Primer pairs for PCR amplification of srtA from S. <strong>aureus</strong>, srtA from S. epidermidis, yhcS from B.<br />

subtilis and srtC from S. epidermidis were based on the genome sequences of the S. <strong>aureus</strong><br />

NCTC8325, S. epidermidis RP62A and B. subtilis 168 strains (Table 2). For expression in S. <strong>aureus</strong> the<br />

RBS and start codon of S. <strong>aureus</strong> srtA was used and for expression in S. epidermidis the RBS and start<br />

codon of S. epidermidis srtA was used. PCR products were purified using the PCR Purification Kit<br />

(Roche), and ligated into the pUC18 plasmid (Norrander et al., 1983). <strong>The</strong> cloned sortase genes were<br />

then excised from the resulting constructs with restriction enzymes as specified in Table 2 and ligated<br />

into plasmids pCN51 or pRIT5H that were cut with the same enzymes. Competent S. <strong>aureus</strong> RN4220<br />

∆srtA cells were transformed with these plasmids by electro-transformation and colonies were selected<br />

on TSA plates containing erythromycin. Subsequently, the plasmids were transferred from S. <strong>aureus</strong><br />

RN4220 to S. <strong>aureus</strong> SH1000 strains via transduction as described above, or to S. epidermidis 1457<br />

strains via electro-transformation of competent cells with the purified plasmids.<br />

Under standard laboratory growing conditions, no SasG production is detectable in S. <strong>aureus</strong> (Roche et<br />

al., 2003). To study the localization of SasG, we used strains in which the expression of SasG is<br />

directed by the IPTG-inducible Pspac-promoter of plasmid pMU<strong>TI</strong>N4 (kindly provided by T. Foster<br />

(Corrigan et al., 2007)). <strong>The</strong> sasG-pMU<strong>TI</strong>N4 plasmid was transferred to the S. <strong>aureus</strong> SH1000 ∆srtA<br />

strain by transduction as described above. Since sasG-pMU<strong>TI</strong>N4 carries an erythromycin resistance<br />

marker, strains containing this plasmid cannot be transformed with derivatives of plasmid pCN51.<br />

<strong>The</strong>refore, the pRIT5H plasmid (Morikawa et al., 2003) was used for complementation experiments<br />

117


Chapter 6<br />

with S. <strong>aureus</strong> srtA or S. epidermidis srtC in the S. <strong>aureus</strong> ∆srtA sasG-pMU<strong>TI</strong>N4 strain. Genes cloned<br />

in pRIT5H are transcribed from the S. <strong>aureus</strong> spa promoter. <strong>The</strong> srtA and srtC genes were PCR<br />

amplified with primer pairs srtA-F4/R4 and srtC-F2/R2, respectively, and cloned into the pRIT5H<br />

plasmid using restriction enzymes specified in Table 2. <strong>The</strong> resulting plasmids were introduced into S.<br />

<strong>aureus</strong> RN4220 strains containing sasG-pMU<strong>TI</strong>N4 via electro-transformation, and they were<br />

subsequently transferred to sasG-pMU<strong>TI</strong>N4-containing S. <strong>aureus</strong> SH1000 strains via transduction.<br />

Transformants and transductants were selected on TSA plates with erythromycin and chloramphenicol.<br />

Table 2. Primers used in this study<br />

Primer Sequence (5’→3’)<br />

Construction of S. <strong>aureus</strong> srtA mutant<br />

srtA-F1 AATGGTGTAGTAATTGACTAG<br />

srtA-R1 TTACGTCAGTCAGTCACCATGGCAACGTTAAGGCTCCTTTTATAC<br />

srtA-F2 TGCCATGGTGACTGACTGACGTAATCTATTACGCTAATGGATGAA<br />

srtA-R2 CTCACATTACTTACTATTAAT<br />

Construction of S. epidermidis srtA mutant<br />

esrtA-F1 AACTTTGTTCTTTAGCGTAACGAAT<br />

esrtA-R1 TGCCATGGTGACTGACTGACGTAATTATGTTACTCCTTTATATTTATT<br />

esrtA-F2 TTACGTCAGTCAGTCACCATGGCATATTCTTATAAGTGAAAGATACGTA<br />

esrtA-R2 CTTTATAGATGACTGCTCCAT<br />

Complementation in S. <strong>aureus</strong><br />

srtA-F3 CAGCCGGATCCAATGTATAAAAGGAGCCTTAACGT (BamHI)<br />

srtA-R3 CGGAATTCTTATTTGACTTCTGTAGCTACAAA (EcoRI)<br />

srtA-F4 GGGGGGGATCCTTAACAGGCATTGTGAAATGT (BamHI)<br />

srtA-R4 GGGGGGTCGACCCTTATTTGACTTCTGTAGCT (SalI)<br />

esrtA-F3 CAGCCGGATCCAATGTATAAAAGGAGCCTTAACGTATGAAGCAGTGGATGAATAGA<br />

(BamHI)<br />

esrtA-R3 CG GAATTCTTAGTTAATTTGTGTAGCTATGAA (EcoRI)<br />

yhcS-F1 AAAACTGCAGAATGTATAAAAGGAGCCTTAACGTATGAAAAAAGTTATTCCACTA (PstI)<br />

yhcS-R1 CAGCCGGATCCTTAAGTCACTCGTTTTCCATATAT (BamHI)<br />

esrtC-F1 GGGGGGTCGACTGAGGAGGTACATATGAGTGC (SalI)<br />

esrtC-R1 GGGGGGGATCCATTTATAATTTGAAAATACCA (BamHI)<br />

esrtC-F2 GGGGGGGATCCTGAGGAGGTACATATGAGTGC (BamHI)<br />

esrtC-R2 GGGGGGTCGACATTTATAATTTGAAAATACCA (SalI)<br />

Complementation in S. epidermidis<br />

srtA-F5 CAGCCGGATCCAAATAAATATAAAGGAGTAACATAAATGAAAAAATGGACAAATCG (BamHI)<br />

srtA-R5 CGGAATTCTTATTTGACTTCTGTAGCTACAAA (EcoRI)<br />

esrtA-F4 GGGGGGGATCCAAATAAATATAAAGGAGTAACATAA (BamHI)<br />

esrtA-R4 GGGGGGAATTCTTAGTTAATTTGTGTAGCTATGA (EcoRI)<br />

yhcS-F2 AAAACTGCAG AAATAAATATAAAGGAGTAACATAAATGAAAAAAGTTATTCCACTA (PstI)<br />

yhcS-R2 CAGCCGGATCCTTAAGTCACTCGTTTTCCATATAT (BamHI)<br />

Overlapping parts are shown in bold<br />

Restriction sites are underlined and shown in parentheses<br />

Cell fractionation, SDS-PAGE, and Western blotting<br />

Overnight cultures were diluted to an OD540 of 0.05 and grown in 25 ml TSB under vigorous shaking.<br />

For complementation of mutant strains with pCN51-based plasmids, CdSO4 was added after three<br />

hours of growth to a final concentration of 0.25 µM. For complementation of mutant strains with<br />

pRIT5H-based plasmids, IPTG was added after three hours of growth to a final concentration of 1 mM.<br />

Samples were taken after six hours of growth and separated in growth medium, whole cell and noncovalently<br />

cell wall-bound protein fractions. Cells were separated from the growth medium by<br />

centrifugation of 1 ml of the culture. <strong>The</strong> proteins in the growth medium were precipitated with 250 µl<br />

50% trichloroacetic acid (TCA), washed with acetone and dissolved in 100 µl Loading Buffer<br />

118


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

(Invitrogen). Cells were resuspended in 300 µl Loading Buffer (Invitrogen) and disrupted with glass<br />

beads using the Precellys ® 24 bead beating homogenizer (Bertin Technologies). From the same culture<br />

20 ml was used for the extraction of non-covalently bound cell wall proteins using KSCN. Cells were<br />

collected by centrifugation, washed with PBS, and incubated for 10 min with 1M KSCN on ice. After<br />

centrifugation the non-covalently cell wall bound proteins were precipitated from the supernatant<br />

fraction with TCA, washed with acetone and dissolved in 100 µl Loading Buffer (Invitrogen). Upon<br />

addition of Reducing Agent (Invitrogen), the samples were incubated at 95ºC. Proteins were separated<br />

by SDS-PAGE using precast NuPage gels (Invitrogen) and subsequently blotted onto a nitrocellulose<br />

membrane (Protran ® , Schleicher & Schuell). <strong>The</strong> presence of Aap, SasG or Clumping factor A (ClfA)<br />

was monitored by immunodetection with specific polyclonal antibodies raised in rabbits at 1:10.000<br />

dilution. <strong>The</strong>se antibodies were kind gifts from D. Mack (Aap) (Rohde et al., 2005) and T. Foster<br />

(SasG and ClfA) (Roche et al., 2003;McDevitt et al., 1995). Bound primary antibodies were visualized<br />

using fluorescent IgG secondary antibodies (IRDye 800 CW goat anti-rabbit from LiCor Biosciences).<br />

Membranes were scanned for fluorescence at 800 nm using the Odyssey Infrared Imaging System<br />

(LiCor Biosciences).<br />

Protein identification by mass spectrometry<br />

Proteins were separated by SDS-PAGE as described before and gels were stained with SimplyBlue TM<br />

SafeStain (Invitrogen). Protein bands were cut from the gels and in-gel digestion of the proteins was<br />

performed as described by Eymann et al. 2004 (Eymann et al., 2004). All peptides obtained from an ingel<br />

digestion were separated by liquid chromatography and measured online by ESI mass spectrometry.<br />

LC-MS/MS analyses were performed using a nanoACQUITY UPLC system (Waters) coupled to an<br />

LTQ Orbitrap or LTQ-F<strong>TI</strong>CR mass spectrometer (<strong>The</strong>rmo Fisher Scientific,Waltham, MA) creating<br />

an electro spray by the application of 1.5 kV between Picotip Emitter (SilicaTip, FS360-20-10<br />

Coating P200P, New Objective) and transfer capillary. Peptides were loaded onto a trap column<br />

(nanoAcquity UPLC TM column, Symmetry® C18, 5 µm, 180 µm inner diameter x 20 mm, Waters) and<br />

washed 3 min with 99% buffer A (0.1% (v/v) acetic acid) with a flow rate of 10 µl/min. Elution was<br />

performed onto an analytical column (nanoAcquity UPLC TM column, BEH130 C18 1.7 µm, 100 µm<br />

inner diameter x 100 mm, Waters) by a binary gradient of buffer A and B (100% (v/v) acetonitrile,<br />

0.1% (v/v) acetic acid) over a period of 80 min with a flow rate of 400 nl/ min.<br />

For MS/MS analysis full survey scans were performed in the Orbitrap or F<strong>TI</strong>CR (m/z 300–2000) with<br />

resolutions of 30,000 in the Orbitrap or 50,000 for the F<strong>TI</strong>CR respectively. <strong>The</strong> full scan was followed<br />

by MS/MS experiments of the five most abundant precursor ions acquired in the LTQ via CID.<br />

Precursors were dynamically excluded for 30 sec, and unassigned charge states as well as singly<br />

charged ions were rejected.<br />

For protein identification tandem mass spectra were extracted using Sorcerer TM v3.5 (Sage-N Research,<br />

Inc. Milpitas, CA). Charge state deconvolution and deisotoping were not performed. All MS/MS<br />

samples were analyzed using Sequest (<strong>The</strong>rmoFinnigan, San Jose, CA; version v.27, rev. 11), applying<br />

the following search parameters: peptide tolerance, 10 ppm; tolerance for fragment ions, 1 amu; b- and<br />

y-ion series; an oxidation of methionine (15.99 Da) was considered as variable modification (maximal<br />

three modifications per peptide). Sequest was set up to search the S. epidermidis RP62A database<br />

(extracted from NCBI, including concatenated reverse database, 4600 entries) assuming the digestion<br />

enzyme trypsin. For S. <strong>aureus</strong> RN4220 and S. <strong>aureus</strong> SH1000 samples, the S. <strong>aureus</strong> NCTC 8325-4<br />

database (extracted from NCBI, including concatenated reverse database, 5784 entries) was used.<br />

Scaffold (version Scaffold_2_04_00, Proteome Software Inc., Portland, OR) was used to validate<br />

MS/MS based peptide and protein identifications. Peptide identifications were accepted if they<br />

exceeded specific database search engine thresholds. Sequest identifications required at least deltaCn<br />

scores of greater than 0.10 and XCorr scores of greater than 2.2, 3.3 and 3.8 for doubly, triply and<br />

quadruply charged peptides. Protein identifications were accepted if they contained at least 2 identified<br />

peptides. With these filter parameter no false positive hit was obtained.<br />

119


Chapter 6<br />

Biofilm formation<br />

Biofilm formation by S. <strong>aureus</strong> and S. epidermidis strains was monitored with crystal-violet<br />

(Christensen et al., 1985). S. <strong>aureus</strong> overnight cultures were diluted 100-fold in TSB containing 5%<br />

glucose and 100 µl aliquots of the diluted cultures were transferred to a 96-well microtiter plate. After<br />

four hours incubation at 37ºC, non-adhered cells were removed and the wells were rinsed with PBS<br />

(pH 7,0). Fresh medium was added to the wells and the cultures were incubated for 24 hours at 37ºC.<br />

<strong>The</strong> supernatant was removed from the wells and non-adhered cells were removed by rinsing with PBS.<br />

Biofilms were fixated with 100 µl 99% methanol for 15 min. <strong>The</strong> supernatant was removed and after<br />

air-drying for 20 min, 100 µl 0,4% crystal-violet solution was added to the wells. Upon 20 min<br />

incubation, the supernatant was removed and the wells were rinsed three times with MilliQ. Bound<br />

crystal-violet was released by adding 150 µl 33% acetic acid, and the absorbance of the released<br />

crystal-violet was measured at 590 nm with the BIO-TEK ® ELx800 TM Universal Microplate Reader<br />

(BioTek Instruments, Inc.). For each strain, the assay was repeated sixteen times.<br />

Results<br />

Protein export in srtA mutants of S. <strong>aureus</strong><br />

In the absence of functional SrtA, proteins with an LPxTG or LPxAG motifs will not be<br />

covalently anchored to the cell wall of S. <strong>aureus</strong>. <strong>The</strong>refore, one might expect at least a partial<br />

release of these proteins into the growth medium of srtA mutant cells, especially if they are<br />

subject to the “shaving activity” of exported proteases. To investigate how a srtA mutation<br />

impacts on the localization of exported proteins of S. <strong>aureus</strong> RN4220, the srtA gene of this<br />

strain was deleted. As shown by SDS-PAGE, the banding pattern of extracellular proteins of<br />

the srtA mutant strain displayed major differences compared to the parental strain since the<br />

intensity of several bands was strongly increased (Figure 2A; compare lanes 1 and 2). In<br />

contrast, the differences observed for the non-covalently cell wall-bound proteins of both<br />

strains were much less pronounced (Figure 2B, lanes 1 and 2). Importantly, deletion of srtA<br />

affected neither the growth rate of S. <strong>aureus</strong> RN4220 nor the optical density reached in the<br />

stationary phase, and no obvious morphological differences between cells of the srtA mutant<br />

and the parental strain were detectable by electron microscopy (chapter 4). To identify the<br />

proteins in bands of which the relative amounts were increased in the medium or cell wall<br />

fractions of the srtA mutant (Figure 2), these bands were cut from gels and analyzed by mass<br />

spectrometry. Of the seven identified proteins, only SdrC, SdrD and protein A contain an<br />

LPxTG motif (Table 3). Interestingly, all three of these proteins also contain a YSIRK motif<br />

in their signal peptides, as does the LipA protein, which was also secreted in higher amounts<br />

by the srtA mutant strain. A very prominent effect of the srtA deletion was observed for<br />

protein A (Figure 2A). Moreover, SdrC- and SdrD-specific bands were identified several<br />

times in the medium fraction of the S. <strong>aureus</strong> srtA mutant, whereas these proteins were not at<br />

all detectable in the growth medium of the parental strain (Figure 2A and data not shown).<br />

<strong>The</strong> absence of srtA also had some clear effects on the composition of the non-covalently cell<br />

wall-bound proteins, which contained increased amounts of the MHC class II analog protein<br />

Map, and decreased amounts of the 5’-nucleotidase SA0295 (Figure 2B).<br />

Since several well-studied LPxTG proteins like SasG (Corrigan et al., 2007) and the clumping<br />

factor A (ClfA; (Josefsson et al., 2001;Siboo et al., 2001;Sullam et al., 1996)) are not<br />

detectable in the Coomassie-stained gels, we studied the localization of these proteins by<br />

Western blotting and subsequent immunodetection with specific antibodies.<br />

120


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

Figure 2. SDS-PAGE analyses of proteins secreted by S. <strong>aureus</strong> and S. epidermidis. Sortase mutants of two S.<br />

<strong>aureus</strong> strains (RN4220 and SH1000) and one S. epidermidis strain (1457) were grown in TSB medium at 37 o C till<br />

the early stationary growth phase. Proteins in the growth medium (A) and non-covalently cell wall-bound proteins<br />

(B) were collected and separated by SDS-PAGE. Gels were stained with Coomassie. Samples were loaded as<br />

follows: lane 1, parental strain; lane 2, srtA mutant; lane 3, srtA mutant complemented with plasmid-borne S.<br />

<strong>aureus</strong> srtA; lane 4, srtA mutant complemented with plasmid-borne S. epidermidis srtA; lane 5, srtA mutant<br />

complemented with plasmid-borne B. subtilis yhcS; and lane 6, srtA mutant complemented with plasmid-borne S.<br />

epidermidis srtC. Protein bands marked with arrows were cut from the gel and identified by mass spectrometry.<br />

121


Chapter 6<br />

Table 3. Extracellular proteins identified in sortase mutants of S. <strong>aureus</strong> and S.<br />

epidermidis<br />

Protein GI Accession # Function Mw mature<br />

protein (kD)<br />

122<br />

Cell Wall<br />

binding domain<br />

S. <strong>aureus</strong><br />

SdrC (YSIRK) 88194324 SdrC protein, putative 98.8 LPETG<br />

SdrD (YSIRK) 88194325 SdrD protein, putative 136.9 LPETG<br />

LipA (YSIRK) 88194101 Lipase a 72.3 -<br />

Spa (YSIRK) 88193885 protein A 49.7 LPETG, LysM (1x)<br />

Hla 88194865 hemolysin A 33.3 -<br />

Map 88195840 also known as Eap or p70;<br />

MHC class II analog protein<br />

62.5<br />

b<br />

SA0295<br />

S. epidermidis<br />

88194087 5'-nucleotidase, lipoprotein<br />

e(P4) family<br />

30.2 -<br />

Aap (YSIRK) 57865793 accumulation associated<br />

protein<br />

246.3 LPDTG, G5 (7x)<br />

AtlE 57866522 bifunctional autolysin 145.2 GW-repeats (3x)<br />

GehC<br />

(YSIRK)<br />

57865740 lipase 73.7 -<br />

SERP0100 57866082 LysM domain protein c 32.5 LysM (3x)<br />

SERP0270 57866259 hypothetical protein 16.2 -<br />

a<br />

cell wall binding has been shown, but no particular wall-binding domain has been identified (Bowden et al.,<br />

2002).<br />

b<br />

Cell wall binding depends on acetylation of lipoteichoic acid (Harraghy et al., 2003)<br />

c<br />

also known as ScaA or AaE; autolysin with a C-terminal cysteine, histidine-dependent amidohydrolase/<br />

peptidase (CHAP) domain; binding to fibrinogen, fibronectin and vitronectin (Heilmann et al., 2003)<br />

To reduce background levels of IgG that was bound by protein A, we used S. <strong>aureus</strong> SH1000<br />

for all Western blotting experiments as this strain produces less protein A than the RN4220<br />

strain. Furthermore, to detect SasG, we used strains that expressed sasG from the pMU<strong>TI</strong>N4<br />

plasmid upon induction with IPTG (Corrigan et al., 2007). In cells of the parental strain<br />

SH1000, SasG was detected predominantly as a band of very high molecular weight (Figure<br />

3A). Furthermore, the growth medium of the parental strain contained multiple SasG-specific<br />

protein species that were detectable in the Western blots as a ladder of discrete protein bands<br />

with molecular weights that were much lower than the molecular weight of the SasG detected<br />

in cells (Figure 3A). In contrast, the high molecular weight band of SasG was not detectable<br />

in cells of the srtA mutant, and much higher amounts of the low molecular weight bands were<br />

detectable not only in the growth medium fraction, but also in the fraction of non-covalently<br />

cell wall-bound proteins (Figure 3A). Similar to what we observed for SasG, a high molecular<br />

mass species of the ClfA protein was detectable in cells of the parental strain SH1000,<br />

whereas the growth medium contained two low molecular ClfA-specific protein species<br />

(Figure 3B). <strong>The</strong> high molecular weight species of ClfA was not detectable in the srtA mutant<br />

cells, and substantial amounts of at least three low molecular weight species of ClfA were<br />

detectable in the cell wall-bound protein fractions of these cells (Figure 3B). Furthermore, the<br />

ratio between the two low molecular weight extracellular ClfA species was changed in the<br />

srtA mutant. Taken together, these findings show that SrtA is indispensable for covalent cell<br />

wall attachment of proteins with LPxTG motifs, and that these proteins are to greater or lesser<br />

extents released into the growth medium of cells lacking SrtA.


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

Figure 3. Localization of LPxTG proteins in srtA mutants of S. <strong>aureus</strong> and S. epidermidis complemented with<br />

different srtA or srtC genes. Sortase A mutants of S. <strong>aureus</strong> SH1000 and S. epidermidis 1457 were grown in TSB<br />

medium at 37 o C till the early stationary growth phase. Samples of extracellular proteins isolated from the growth<br />

medium (M), non-covalently cell wall-bound proteins (CW) and total cells (C) were used for Western blotting and<br />

immunodetection with antibodies against S. <strong>aureus</strong> SasG (panel A), S. <strong>aureus</strong> ClfA (panel B) and S. epidermidis<br />

Aap (panel C). <strong>The</strong> positions of SasG, ClfA, and Aap are marked with arrows. Constructs used for<br />

complementation of ∆srtA mutations are indicated on top of each gel.<br />

Protein export in a srtA mutant of S. epidermidis<br />

To investigate how a srtA mutation influences the localization of proteins in S. epidermidis,<br />

the extracellular and non-covalently cell wall-bound proteins of S. epidermidis 1457 and a<br />

srtA mutant derivative of this strain were analyzed by SDS-PAGE. Bands of different<br />

intensities were excised and analyzed by mass spectrometry. A major difference in<br />

localization was observed for the Aap protein, which was dislocated to the growth medium of<br />

the srtA mutant in much higher amounts than was the case for the parental strain 1457 (Figure<br />

2A). Furthermore, while two Aap-specific bands were detectable in the growth medium of the<br />

parental strain, which probably correspond to the 220 kD and 180 kD forms of Aap described<br />

by Rohde et al. (Rohde et al., 2005), a third Aap-specific band was detectable in the medium<br />

of the srtA mutant. This might be the 140 kD Aap-derived band that is necessary for biofilm<br />

formation. Also, in the non-covalently cell wall-bound protein fraction of the S. epidermidis<br />

srtA mutant (Figure 2B), three bands for Aap were identified, which seem to correspond to<br />

the previously described 220 kD, 180 kD and 140 kD isoforms. Notably, these three isoforms<br />

are not present amongst the non-covalently cell wall-bound proteins of the parental strain.<br />

Furthermore, the fraction of non-covalently cell wall-bound proteins of the srtA mutant<br />

contained somewhat lower amounts of the bifunctional autolysin AtlE. <strong>The</strong> subcellular<br />

localization of Aap was further examined by Western blotting and immunodetection with<br />

specific antibodies. As shown in Figure 3C, two dominant Aap-specific bands of ~220 kD and<br />

~180 kD, and a faint band of ~140 kDa, were detectable in growth medium samples of the<br />

parental strain 1457. <strong>The</strong>se bands probably correspond to the 220 kDa, 180 kDa and 140 kDa<br />

bands identified by Rohde et al (Rohde et al., 2005). Remarkably, no Aap was detectable in<br />

the total cell fraction. In this respect, the behavior of Aap seems to differ from that of its S.<br />

<strong>aureus</strong> homologue SasG, which was detectable as a high molecular weight band (Figure 3A).<br />

Due to the srtA mutation, the extracellular amounts of the 220 kDa, 180 kDa and 140 kDa<br />

forms of Aap were significantly increased, and substantial amounts of the 220 kDa and 180<br />

kDa forms were also detectable in the fraction of non-covalently cell wall-bound proteins. To<br />

a lesser extent the 220 kDa and 180 kDa forms were also detectable in the cellular fraction of<br />

the srtA mutant. <strong>The</strong>se findings imply that, in the absence of SrtA, substantial amounts of Aap<br />

remain bound to the cell wall, but in a non-covalent manner.<br />

123


Chapter 6<br />

Complementation analysis of srtA mutant strains with Class A, C and D sortases<br />

Deletion of the srtA genes of S. <strong>aureus</strong> and S. epidermidis resulted for both organisms in clear<br />

changes in the localization of several proteins to the cell wall and growth medium. To<br />

distinguish between the effects of the srtA mutation and possible unwanted second site<br />

mutations, a complementation analysis was performed with the homologous srtA gene<br />

expressed from a plasmid. As shown in Figures 2 and 3, full reversion of the observed protein<br />

localization phenotypes was achieved by ectopic expression of the S. <strong>aureus</strong> srtA gene in the<br />

S. <strong>aureus</strong> ∆srtA mutants, and by ectopic expression of the S. epidermidis srtA gene in the S.<br />

epidermidis ∆srtA mutant. Interestingly, full reversion of the observed phenotypes was also<br />

achieved when the S. <strong>aureus</strong> srtA gene was expressed in the S. epidermidis ∆srtA mutant, or<br />

when the S. epidermidis srtA gene was expressed in the S. <strong>aureus</strong> ∆srtA mutant (Figure 2,<br />

Figure 3C). This raised the question to what extents the phenotypes of srtA mutant strains<br />

could also be reverted by ectopic expression of sortases of other classes that also seem to<br />

recognize the LPxTG motif, like SrtC or SrtD. To address this question, the S. epidermidis<br />

srtC gene and the B. subtilis yhcS gene (encoding a class D sortase) were expressed from the<br />

same plasmid-borne promoters that were used for the complementation analyses with srtA<br />

genes. Furthermore, for expression in the S. <strong>aureus</strong> srtA mutant, all heterologous sortase<br />

genes were provided with the ribosome-binding site and start codon from S. <strong>aureus</strong> srtA to<br />

minimize any possible differences in translation of the different sortases. Conversely, for<br />

expression in the S. epidermidis srtA mutant, all heterologous sortase genes were provided<br />

with the ribosome-binding site and start codon from S. epidermidis srtA. As shown in Figure<br />

2 and Figure 3C, none of the phenotypes observed in S. <strong>aureus</strong> or S. epidermidis srtA mutant<br />

strains were reversed by expression of yhcS from B. subtilis. In contrast, expression of srtC<br />

did lead to the complementation of some, but not all phenotypes of srtA mutant strains. As<br />

shown in Figure 2, expression of srtC in the S. <strong>aureus</strong> srtA mutant did not result in lowered<br />

extracellular levels of the LPxTG proteins SdrC, sdrD and protein A as would be expected if<br />

srtC could complement for the absence of srtA. However, srtC expression did result in a<br />

partial complementation of the localization defect observed for SasG in the S. <strong>aureus</strong> srtA<br />

mutant, where clearly lowered amounts of non-covalently cell wall-bound forms of SasG<br />

were observed, as well as an increased amount of the high molecular weight cellular form of<br />

SasG (Figure 3A). Similarly, srtC expression in the S. <strong>aureus</strong> srtA mutant resulted in a partial<br />

restoration of the localization of ClfA, the most prominent effect being the re-appearance of<br />

the high molecular weight form of ClfA in the cellular fraction (Figure 4B). Furthermore, srtC<br />

expression substantially reduced the amounts of low molecular weight forms of ClfA in the<br />

fraction of non-covalently cell wall-bound proteins, but not to the extent that was observed<br />

when the srtA mutant was complemented with srtA of S. <strong>aureus</strong> (Figure 3B).<br />

Consistent with the observations for SasG in S. <strong>aureus</strong> (Figure 3A), expression of srtC in the<br />

S. epidermidis srtA mutant resulted in a significant reversion of the dislocation phenotype of<br />

the SasG homologue Aap (Figure 2 and Figure 3C). Upon srtC expression, the amounts of the<br />

~220 kDa and ~140 kDa forms were clearly reduced in the growth medium, as well as the<br />

non-covalently cell wall-bound protein fractions of S. epidermidis ∆srtA. In both fractions, the<br />

strongest effects of srtC expression were observed for the ~220 kDa species of Aap, which<br />

virtually disappeared. Furthermore, the expression of srtC resulted in increased amounts of<br />

AtlE in the fraction of non-covalently cell wall-bound proteins. Taken together, these findings<br />

show that SrtA and SrtC have at least partially overlapping substrate specificities.<br />

124


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

Biofilm formation by complemented srtA mutants of S. <strong>aureus</strong> and S. epidermidis<br />

Surface proteins like SasG and protein A of S. <strong>aureus</strong> (Corrigan et al., 2007;Merino et al.,<br />

2009) and Aap in S. epidermidis (Hussain et al., 1997;Rohde et al., 2005;Rohde et al., 2007)<br />

have been implicated in biofilm formation. <strong>The</strong>refore, we analyzed the biofilm-forming<br />

capacity of complemented srtA mutants of S. <strong>aureus</strong> and S. epidermidis using a crystal-violetbinding<br />

assay. <strong>The</strong> results are summarized in Figure 4.<br />

A 590<br />

2,5<br />

2<br />

1,5<br />

1<br />

0,5<br />

0<br />

1 2 3 4 5 6 1 2 3 4 5 6 1 2 3 4 5 6<br />

RN4220 SH1000 1457<br />

A<br />

125<br />

A 590<br />

2,5<br />

2<br />

1,5<br />

1<br />

0,5<br />

0<br />

WT srtA Sa.srtA-pRIT5H Se.srtC-pRIT5H<br />

Figure 4. Biofilm formation by srtA mutants of S. <strong>aureus</strong> and S. epidermidis complemented with different srtA<br />

or srtC genes. Biofilm formation in 96-well microtiter plates by cells grown for 24 hours at 37ºC in TSB medium<br />

with 5% glucose was measured using crystal-violet staining. <strong>The</strong> amounts of crystal-violet liberated from the<br />

biofilms upon treatment with 33% acetic acid were determined by measuring the absorbance at 590 nm (A 590). For<br />

each strain, biofilm formation was measured sixteen times. (A) A 590 measurements for ∆srtA mutants of S. <strong>aureus</strong><br />

RN4220, S. <strong>aureus</strong> SH1000 and S. epidermidis 1457. 1, parental strain; 2, ∆srtA mutant; 3, ∆srtA mutant<br />

complemented with srtA from S. <strong>aureus</strong>; 4, ∆srtA mutant complemented with srtA from S. epidermidis; 5, ∆srtA<br />

mutant complemented with yhcS of B. subtilis; 6. ∆srtA mutant complemented with srtC of S. epidermidis. (B) A 590<br />

measurements for the srtA mutant of S. <strong>aureus</strong> SH1000 overproducing SasG and complemented with srtA from S.<br />

<strong>aureus</strong>, or srtC of S. epidermidis.<br />

Interestingly, the biofilm-forming capacity of srtA mutants of the S. <strong>aureus</strong> strains RN4220<br />

and SH1000 was significantly reduced compared to the respective parental strains. Biofilm<br />

formation by the srtA mutant of S. <strong>aureus</strong> RN4220 was largely complemented by expression<br />

of the srtA genes from S. <strong>aureus</strong> or S. epidermidis, whereas biofilm formation by the srtA<br />

mutant of S. <strong>aureus</strong> SH1000 was complemented by the S. <strong>aureus</strong> srtA gene, but not by the S.<br />

epidermidis srtA gene. This suggests that at least one covalently cell wall-bound protein of S.<br />

<strong>aureus</strong> is not properly attached to the cell wall by the heterologously produced SrtA of S.<br />

epidermidis. Furthermore, neither yhcS of B. subtilis, nor srtC of S. epidermidis were able to<br />

restore biofilm formation by S. <strong>aureus</strong> srtA mutant strains. Interestingly, the negative effect of<br />

the srtA mutation on biofilm formation by the S. <strong>aureus</strong> SH1000 ∆srtA strain was largely<br />

suppressed by SasG overproduction (Figure 4B), which indicates that, despite the absence of<br />

SrtA, sufficient SasG was correctly localized to have a stimulating effect on biofilm<br />

formation. Notably, biofilm formation was enhanced to levels that exceeded the biofilm<br />

formation by the SasG-overproducing parental strain SH1000 when the S. <strong>aureus</strong> srtA gene or<br />

the S. epidermidis srtC gene were ectopically expressed. This shows that SrtA is a limiting<br />

factor for the correct localization and functionality of overproduced SasG and that this<br />

particular function of SrtA can also be fulfilled by SrtC.<br />

In contrast to what was observed in srtA mutants of S. <strong>aureus</strong> RN4220 and SH1000, but<br />

similar to the SasG overproducing S. <strong>aureus</strong> srtA mutant, the srtA mutant of S. epidermidis<br />

1457 did not display a clear defect in biofilm formation. Interestingly however, the ectopic<br />

expression of srtA from S. <strong>aureus</strong> or S. epidermidis in the S. epidermidis srtA mutant resulted<br />

in a significant increase in the biofilm-forming capacity of S. epidermidis, similar to what was<br />

B


Chapter 6<br />

observed for SasG-overproducing S. <strong>aureus</strong> SH1000 strains. In contrast, the ectopic<br />

expression of srtC interfered with biofilm formation by S. epidermidis (Figure 4A). Taken<br />

together, our present findings show that the activities of SrtA and SrtC can be a limiting<br />

factors in biofilm formation by S. <strong>aureus</strong> and S. epidermidis, and that this feature can be<br />

correlated with the level of SasG production in S. <strong>aureus</strong>.<br />

Discussion<br />

Surface proteins of the Gram-positive bacterial pathogens S. <strong>aureus</strong> and S. epidermidis serve<br />

important roles in virulence and biofilm formation. Several of these surface proteins are<br />

linked covalently to the cell wall by the transpeptidase SrtA, which recognizes a C-terminal<br />

LPxTG motif for cell wall attachment of its substrates. Previous studies have already shown<br />

that srtA mutants of S. <strong>aureus</strong> are attenuated in their ability to cause infections in animal<br />

models (Mazmanian et al., 2001). In the present studies we report for the first time the<br />

construction of a srtA mutant of S. epidermidis. This mutant and equivalent srtA mutants of S.<br />

<strong>aureus</strong> were used to address three questions. First, we investigated to what extent srtA<br />

mutations affect the localization of cell wall-associated proteins. <strong>The</strong> results show that the<br />

absence of SrtA causes substantial changes in the composition of the S. <strong>aureus</strong> and S.<br />

epidermidis exoproteomes, mainly due to the secretion of normally cell wall-attached<br />

proteins. Nevertheless, substantial amounts of the different LPxTG proteins remain attached<br />

to the cell wall in a non-covalent manner. Secondly, the srtA mutants were used to study<br />

whether there is any overlap in the substrate specificities of class A, C and D sortases, all of<br />

which recognize proteins with LPxTG motifs. Our results show that the substrate specificities<br />

of the staphylococcal class A and C sortases overlap only partially. Thirdly, we addressed the<br />

roles of sortases in biofilm formation by S. <strong>aureus</strong> and S. epidermidis, which revealed that<br />

sortase activity can be a limiting factor in this process.<br />

Three possible phenotypes can be expected when SrtA is not expressed. Firstly, the LPxTG<br />

proteins may remain anchored to the cell surface through their C-terminal transmembrane<br />

domain. Secondly, the LPxTG proteins may be released into the growth medium through<br />

proteolytic “shaving” by extracellular proteases, a phenomenon that was previously observed<br />

for many unprocessed lipoproteins (Antelmann et al 2001; Venema et al 2003; Stoll et al<br />

2005). Thirdly, the LPxTG proteins may remain attached to the cell surface via non-covalent<br />

interaction with components of the cell wall. Clearly, all LPxTG proteins investigated in the<br />

present studies were released into the growth medium of srtA mutant strains, which implies<br />

that they had lost their C-terminal transmembrane domains. Nevertheless, we cannot exclude<br />

the possibility that a subfraction of these proteins remained attached to the membrane via an<br />

uncleaved C-terminal transmembrane domain. Furthermore, substantial amounts of the<br />

LPxTG proteins remained attached to the cell wall in a non-covalent manner. This is not<br />

altogether surprising since several of these proteins have repeated cell wall-binding domains.<br />

For example, LysM domains for peptidoglycan-binding are present in the protein A of S.<br />

<strong>aureus</strong> (Buist et al., 2008), and G5 repeats for N-acetylglucosamine-binding are present in<br />

SasG of S. <strong>aureus</strong> and Aap of S. epidermidis (Bateman et al., 2005). Interestingly, high<br />

molecular weight species of SasG were observed in srtA-proficient cells of S. <strong>aureus</strong>. <strong>The</strong>se<br />

might represent SasG proteins interacting with each other through their G5 domains as was<br />

previously proposed for the homologous Aap protein of S. epidermidis (Conrady et al., 2008).<br />

However, the high molecular weight species may also represent SasG molecules covalently<br />

attached to the cell wall of S. <strong>aureus</strong>. Similar explanations can be entertained for the presence<br />

126


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

of a high molecular weight form of ClfA in srtA-proficient cells of S. <strong>aureus</strong>. Clearly, in<br />

absence of srtA these proteins are not linked covalently to the cell wall and, in agreement with<br />

this notion, no high molecular weight species of SasG and ClfA were detectable in srtA<br />

mutant cells. It is in this context remarkable that we did not detect a high molecular weight<br />

species of Aap in srtA-proficient S. epidermidis cells. Possibly, such a species remained<br />

undetectable due to a poor mobility in SDS-PAGE.<br />

Interestingly, all LPxTG proteins of S. <strong>aureus</strong> and S. epidermidis that were found to be<br />

dislocated in the respective srtA mutant strains contain a YSIRK/GS domain within their<br />

signal peptide. It has been shown that the proteins with this YSIRK/GS motif, such as ClfA,<br />

protein A, fibronectin-binding protein B (FnbpB), and the serine-aspartate repeat proteins<br />

SdrC and SdrD display a ring-like distribution on the S. <strong>aureus</strong> cell surface (Dedent et al.,<br />

2008). This has led to the proposal that proteins with the YSIRK/GS motif are sitespecifically<br />

translocated to the cross wall, which is the peptidoglycan layer that forms during<br />

cell division to separate new daughter cells. Our finding that in particular proteins with the<br />

YSIRK/GS motif in their signal peptides are dislocated to the growth medium when SrtA is<br />

absent could suggest that the release of these proteins from the cell wall is related to the site<br />

of secretion or surface display. This could also be a possible explanation for the observed<br />

increased release of the lipase LipA by the srtA mutant of S. <strong>aureus</strong>. However, it has to be<br />

noted that secretion of the lipase GehC of S. epidermidis, which also has the YSIRK/GC<br />

motif in its signal peptide, was not detectably influenced by the srtA mutation. Most likely,<br />

the observed effects of srtA mutations on the localization of non-LPxTG proteins, such as<br />

LipA, Hla and Map of S. <strong>aureus</strong>, or AtlE of S. epidermidis are indirectly caused by the<br />

absence of SrtA. This could relate to as yet unidentified alterations in the cell wall<br />

composition of srtA mutant strains, or perhaps even to precluded interactions with LPxTG<br />

proteins that are dislocated due to the srtA mutations.<br />

To date, little information is available on the class C and D sortases in Gram-positive bacteria.<br />

Since these sortases also recognize the LPxTG motif, we studied the complementation of the<br />

S. <strong>aureus</strong> and S. epidermidis srtA mutants with srtC from S. epidermidis ATCC12228 or yhcS<br />

(srtD) from B. subtilis 168. No complementation was observed upon introduction of yhcS in<br />

any of the srtA mutant strains tested. This may either mean that YhcS does not recognize the<br />

LPxTG proteins that we monitored in the present studies, or that the cells contained<br />

insufficient amounts of YhcS which could, for example, be due to inefficient translation or<br />

post-translational degradation. In contrast, partial complementation was observed for the<br />

localization of SasG and ClfA in srtA mutant strains of S. <strong>aureus</strong> expressing srtC, and for Aap<br />

and AtlE in the srtA mutant strain of S. epidermidis expressing srtC. <strong>The</strong> molecular basis for<br />

this partial overlap in the specificities of SrtA and SrtC is presently not completely clear.<br />

Firstly, the “LPxTG” sites of SasG (LPKTG), ClfA (LPDTG) and Aap (LPDTG) differ only<br />

in the non-conserved central “x residue” with the LPxTG sites of SdrC, SdrD and protein A<br />

(LPETG). This could mean that a Glu residue at the x position is not acceptable for SrtC,<br />

whereas Lys or Asp residues at this position are acceptable both for SrtA and SrtC. Based on<br />

bioinformatics analyses, Comfort and Club (Comfort and Clubb, 2004) have classified<br />

various LPxTG recognition sites for different sortases, which suggests that a central Lys<br />

residue in the LPxTG motif, as encountered in SasG, would be acceptable to several different<br />

groups of sortases. This would be consistent with our present findings that SasG is a substrate<br />

for SrtA and SrtC. In contrast, this bioinformatics-based classification did not indicate LPxTG<br />

motifs with an Asp residue at the central x position as SrtA substrates. Even so, our present<br />

analyses indicate that proteins, like ClfA and Aap, which have an LPDTG motif are SrtA<br />

127


Chapter 6<br />

substrates that are also recognized by SrtC. Clearly, at this stage we cannot rule out the<br />

possibility that other features of SasG, ClfA and Aap are probably responsible for the fact that<br />

these LPxTG proteins are substrates for SrtA and SrtC, while SdrC, SdrD and protein A are<br />

only substrates for SrtA. Furthermore, SrtC displays several structural differences to class A<br />

sortases (Figure 1). Specifically, SrtA of S. <strong>aureus</strong> and SrtC of S. epidermidis 12228 merely<br />

share 34% amino acid sequence identity. Importantly, the key residues involved in catalysis<br />

(His-120, Cys-184 and Arg-197) and substrate recognition (Val-168 and Leu-169) are<br />

conserved in both sortases, but the differences between both proteins are large enough to<br />

allow for specific differences in the geometry of their active sites. Similarly, a stretch of<br />

amino acids in the β6/β7 loop was shown to determine the substrate specificity of SrtB<br />

(Bentley et al., 2007). Intriguingly, Aap appears to be conserved also in S. epidermidis 12228<br />

from which the srtC gene was derived. This suggests that the SrtC of this S. epidermidis strain<br />

may not only be dedicated to cell wall attachment of the LPxTG proteins SesJ and SesK as<br />

was previously suggested (Gill et al., 2005), but it may also be involved in covalent cell wall<br />

attachment of Aap.<br />

<strong>The</strong> results of our comparative analyses on the roles of sortases in biofilm formation by S.<br />

<strong>aureus</strong> and S. epidermidis are intriguing. Clearly, SrtA is important for biofilm formation by<br />

S. <strong>aureus</strong>, whereas this sortase is dispensable for biofilm formation by S. epidermidis. At this<br />

stage, it is difficult to say which LPxTG protein is responsible for the SrtA-dependence of<br />

biofilm formation by S. <strong>aureus</strong>, but protein A is clearly an attractive candidate. Firstly,<br />

Merino et al. have recently shown the involvement of protein A in the formation of proteindependent<br />

biofilms in S. <strong>aureus</strong> (Merino et al., 2009) and, secondly, a major dislocating effect<br />

of the srtA mutation was observed for protein A in the present studies. However, also other<br />

LPxTG proteins may be involved in this phenomenon. Furthermore, SasG overexpression was<br />

able to compensate for the absence of SrtA, underpinning the importance of SasG for proteindependent<br />

biofilm formation. However, in this case, the levels of biofilm formation were even<br />

further increased upon ectopic expression of SrtA or SrtC, which indicates that sortase<br />

activity is a limiting factor for biofilm formation under the conditions tested. <strong>The</strong> finding that<br />

StrC production in SasG-overproducing cells did also stimulate biofilm formation is<br />

consistent with the finding that SrtC was able to revert the SasG dislocation phenotype of srtA<br />

mutant cells at least in part. In the light of these findings, the observation that S. epidermidis<br />

srtA mutant cells were not impaired in biofilm formation is not really surprising since these<br />

cells produced readily detectable amounts of the SasG homologue Aap. Also consistent with<br />

the findings in S. <strong>aureus</strong> cells producing SasG, the ectopic expression of SrtA in S.<br />

epidermidis had a strongly stimulating effect on biofilm formation. This shows that at least in<br />

S. epidermidis strain 1457, SrtA is a limiting determinant for biofilm formation. Whether this<br />

is also the case in other S. epidermidis strains remains to be shown. Quite unexpectedly, srtC<br />

expression resulted in impaired biofilm formation by S. epidermidis 1457. However, this<br />

observation can be reconciled with the fact that the S. epidermidis strain 12228 from which<br />

the srtC gene was isolated is unable to form biofilms (Zhang et al., 2003), despite the fact that<br />

this strain has an Aap-encoding gene. It is thus tempting to speculate that protein-dependent<br />

biofilm formation by S. epidermidis 12228 is suppressed by SrtC. Possibly, the SrtCdependent<br />

suppression of biofilm formation in S. epidermidis 1457 is correlated to the<br />

strongly suppressed extracellular accumulation of the ~220 kDa and ~140 kDa forms of Aap,<br />

which are actually present in lower amounts than observed in the growth medium of the<br />

parental strain 1457. However, this needs to be investigated in more detail in future studies<br />

128


Partially overlapping substrate specificities of sortases A and C of staphylococci<br />

with particular attention for the role of srtC in a possible suppression of biofilm formation in<br />

S. epidermidis 12228.<br />

Acknowledgements<br />

We like to thank W. Baas and M. ten Brinke for technical assistance, S. Dubrac for providing<br />

the pCN51 plasmid, and colleagues from the StaphDynamics and <strong>TI</strong><strong>Pharma</strong> programs for<br />

helpful discussions. Financial support was provided by the CEU (LSHM-CT-2006-019064,<br />

LSHG-CT-2006-037469 and PITN-GA-2008-215524), DFG (GK212/3-00, SFB/TR34, FOR<br />

585), and Top Institute <strong>Pharma</strong> (T4-213).<br />

129


“Without music, life would be a mistake”<br />

-Friedrich W. Nietzsche-<br />

130


Chapter 7<br />

<strong>The</strong> large mechanosensitive channel MscL determines bacterial<br />

susceptibility to the bacteriocin sublancin 168<br />

R.H.M. Kouwen, E.N. Trip, E.L. Denham, M.J.J.B. Sibbald,<br />

J.-Y.F. Dubois, and J.M. van Dijl #<br />

Published in Antimicrobial Agents and Chemotherapy (2009) 53, 4702-4711<br />

131


Chapter 7<br />

Summary<br />

Bacillus subtilis strain 168 produces the extremely stable and broad-spectrum lantibiotic<br />

sublancin 168. Known sublancin 168 susceptible organisms include important<br />

pathogens, such as <strong>Staphylococcus</strong> <strong>aureus</strong>. Nevertheless, since its discovery, the mode of<br />

action of sublancin 168 has remained elusive. <strong>The</strong> present studies were, therefore, aimed<br />

at the identification of cellular determinants for bacterial susceptibility towards<br />

sublancin 168. Growth inhibition and competition assays on plates and in liquid cultures<br />

revealed that sublancin 168-mediated growth inhibition of susceptible B. subtilis and S.<br />

<strong>aureus</strong> cells is affected by the NaCl concentration in the growth medium. Added NaCl<br />

did not influence the production, activity or stability of sublancin 168 but, instead,<br />

lowered the susceptibility of sensitive cells towards this lantibiotic. Importantly, the<br />

susceptibility of B. subtilis and S. <strong>aureus</strong> cells towards sublancin 168 was shown to<br />

depend on the presence of the large mechanosensitive channel of conductance MscL. In<br />

contrast, MscL was not involved in susceptibility towards the bacteriocins nisin or Pep5.<br />

Taken together, our unprecedented results demonstrate that MscL is a critical and<br />

specific determinant in bacterial sublancin 168 susceptibility that may either serve as a<br />

direct target for this lantibiotic, or as a gate of entry to the cytoplasm.<br />

132


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

Introduction<br />

Lantibiotics are small post-translationally modified peptides with antimicrobial activity that<br />

are produced by Gram-positive bacteria (Chatterjee et al., 2005; McAuliffe et al., 2001; Sahl<br />

and Bierbaum, 1998). <strong>The</strong> Bacillus subtilis strain 168 is known to produce an extremely<br />

stable lantibiotic named sublancin 168. This lantibiotic exhibits bactericidal activity against<br />

other Gram-positive bacteria, including important pathogens, such as Bacillus cereus,<br />

Streptococcus pyogenes, and <strong>Staphylococcus</strong> <strong>aureus</strong> (Paik et al., 1998; Stein, 2005).<br />

Sublancin 168 is encoded by the sunA gene, which is located within the SPβ prophage region<br />

of the B. subtilis 168 chromosome (Hemphill et al., 1980; Lazarevic et al., 1999; Paik et al.,<br />

1998). Newly synthesized sublancin 168 is exported from the cytoplasm by the ABC<br />

transporter SunT, the gene of which is located immediately downstream of sunA (Serizawa et<br />

al., 2005; Dorenbos et al., 2002). SunT also contains a proteolytic domain (McAuliffe et al.,<br />

2001) and is, therefore, thought to be required both for sublancin 168 export and concomitant<br />

removal of the leader peptide (Dorenbos et al., 2002; Paik et al., 1998). It is noteworthy that<br />

sublancin 168 displays the extraordinary characteristic for lantibiotics of having two disulfide<br />

bonds in addition to a β-methyllanthionine bridge (Paik et al., 1998). <strong>The</strong> thiol-disulfide<br />

oxidoreductase BdbB, which is encoded by a gene downstream of sunA and sunT, appears to<br />

be of major importance for the formation of the disulfide bonds (Kouwen et al., 2007;<br />

Dorenbos et al., 2002).<br />

<strong>The</strong> cellular target(s) of sublancin 168 and the determinant(s) for producer immunity<br />

against this lantibiotic have remained elusive for a long time. Very recently however, we have<br />

identified the SunI protein (also known as YolF) as the immunity protein that protects<br />

producer cells against sublancin 168 (Dubois et al., 2009). SunI was found to be both required<br />

and sufficient for sublancin 168 immunity, even when produced in a heterologous sublancinsensitive<br />

host organism, such as S. <strong>aureus</strong>. Interestingly, localization studies showed that the<br />

SunI protein is anchored to the membrane through a single N-terminal membrane-spanning<br />

domain with the bulk of the protein facing the cytoplasm. This is a topology that has not been<br />

reported before for other known bacteriocin immunity proteins (Dubois et al., 2009).<br />

<strong>The</strong> present studies were aimed at identifying bacterial determinants that confer<br />

susceptibility to sublancin 168. To date, two major mechanisms for bactericidal lantibiotic<br />

activity have been reported. Type A lantibiotics, such as nisin (Kuipers et al., 1993; Siegers et<br />

al., 1996), epidermin (Peschel and Götz, 1996), and Pep5 (Meyer et al., 1995) usually act by<br />

forming pores in the cytoplasmic membrane of sensitive target organisms in processes that<br />

may involve specific molecules, such as the cell wall precursor lipid II (Breukink et al., 1999;<br />

Wiedemann et al., 2001). In contrast, type B lantibiotics, such as cinnamycin (Fredenhagen et<br />

al., 1990) and mersacidin (Chatterjee et al., 1992), inhibit particular enzyme functions. On the<br />

basis of its leader peptide sequence, sublancin 168 was previously classified as a type A<br />

lantibiotic (Paik et al., 1998). Nevertheless, sublancin 168 does not display the usual flexible,<br />

elongated and amphipathic molecular shape that is so characteristic for other type A<br />

lantibiotics (Nagao et al., 2006), suggesting that sublancin 168 might have a specific mode of<br />

action. Consistent with this idea, our present results show that the sublancin 168 susceptibility<br />

of B. subtilis and S. <strong>aureus</strong> is determined by the presence of large-conductance<br />

mechanosensitive channels, which is an unprecedented finding.<br />

133


Chapter 7<br />

Materials and Methods<br />

Bacterial strains, plasmids and growth media<br />

<strong>The</strong> bacterial strains and plasmids used in this study are listed in Table 1.<br />

Table 1. Strains and plasmids used in this study<br />

Plasmids Relevant properties Reference<br />

pDG783 pSB118 derivative; contains the kanamycin resistance marker<br />

from Streptococcus faecalis; Amp R ; Km R<br />

(Guérout-Fleury et<br />

al., 1995)<br />

pUC18 Ap R , ColE1, φ80lacZ, lac promoter (Sambrook et al.,<br />

1989)<br />

pX Vector for the integration of genes in the amyE locus of B.<br />

subtilis; integrated genes will be transcribed from the xylose<br />

inducible xylA promoter; carries the xylR gene; Amp R ; Cm R<br />

(Kim et al., 1996)<br />

pXTC pX derivative in which the chloramphenicol resistance marker<br />

has been replaced with a tetracycline resistance marker; Amp R (Darmon et al.,<br />

; 2006)<br />

Tet R<br />

TOPO pCR®-Blunt II-TOPO® vector; Km R Invitrogen Life<br />

Technologies, Inc.<br />

pMAD shuttle vector for E. coli and S. <strong>aureus</strong> with thermosensitive ori<br />

in S. <strong>aureus</strong>; contains the bgaB gene; Em R ; Amp R<br />

(Arnaud et al., 2004)<br />

pMAD-mscL pMAD plasmid with the flanking regions of S. <strong>aureus</strong> mscL;<br />

Em R ; Amp R<br />

This work<br />

BaSysBio II Ligation-independent cloning vector based on pDG1727 for<br />

Promoter-GFP activity analysis; Amp R , Spec R<br />

Stéphane Aymerich<br />

et al., unpublished<br />

Strains<br />

E. coli<br />

Genotype Reference<br />

DH5α F - , Ф80dlacZ ∆M15 endA1 recA1 gyrA96 thi-1 hsdR17 (rk- Invitrogen Life<br />

mk+) supE44 relA1 deoR ∆(lacZYA-argF) U169<br />

Technologies, Inc.<br />

TOP10 Cloning host for TOPO vector; F - mcrA ∆(mrr-hsdRMS-mcrBC)<br />

Φ80lacZ∆M15 ∆lacX74 recA1 araD139 ∆(ara-leu)7697 galU<br />

galK rpsL (Str R Invitrogen Life<br />

B. subtilis<br />

) endA1 nupG<br />

Technologies, Inc.<br />

168 trpC2 (Kunst et al., 1997)<br />

168 Cm trpC2; amyE::pX; Cm R<br />

(Dubois et al., 2009)<br />

∆SPβ trpC2; ∆SPβ; sublancin 168 sensitive (Dorenbos et al.,<br />

2002)<br />

∆SPβ Tc trpC2; ∆SPβ; amyE::pXTC; sublancin 168 sensitive; Tet R<br />

(Dubois et al., 2009)<br />

∆mscL trpC2; pMU<strong>TI</strong>N4mcs::mscL; double crossover deletion of mscL;<br />

Em R<br />

BSFA collection<br />

strain BFS1257<br />

∆SPβ∆mscL trpC2; ∆SPβ; pMU<strong>TI</strong>N4mcs::mscL; Em R This work<br />

Strains<br />

B. subtilis<br />

Genotype Reference<br />

∆sunA-<br />

∆yolF<br />

trpC2; ∆yolF; ∆sunA; Km R (Dubois et al., 2009)<br />

∆sunA trpC2; ∆sunA; Km R S. <strong>aureus</strong><br />

(Dubois et al., 2009)<br />

RN4220 Restriction-deficient derivative of NCTC 8325, cured of all (Kreiswirth et al.,<br />

known prophages; rsbU-; agr-<br />

1983)<br />

RN4220 RN4220 that contains the pMAD vector for erythromycin<br />

Em resistance; Em R<br />

This work<br />

RN4220<br />

∆mscL<br />

RN4220 derivative; rsbU-; agr-; ∆mscL This work<br />

134


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

<strong>The</strong> standard LB medium consisted of 1% trypton, 0.5% yeast extract and 1% NaCl, pH 7.4. Where<br />

appropriate, the NaCl content of the LB medium was adjusted to concentrations ranging from 0 to 5%.<br />

S. <strong>aureus</strong> strains were grown in brain-heart infusion broth (BHI) or tryptone soya broth (TSB). Where<br />

necessary, media were supplemented with antibiotics at the following concentrations: ampicillin, 100<br />

µg/ml (Escherichia coli); kanamycin, 20 µg/ml (E. coli, B. subtilis, S. <strong>aureus</strong>); chloramphenicol, 5<br />

µg/ml (E. coli and B. subtilis); tetracycline, 10 µg/ml (E. coli and B. subtilis); erythromycin, 100 µg/ml<br />

(E. coli), 2 µg/ml (B. subtilis) or 5 µg/ml (S. <strong>aureus</strong>); spectinomycin, 100 µg/ml (B. subtilis). To<br />

visualize α-amylase activity (specified by the amyE gene), LB plates were supplemented with 1%<br />

starch. To visualize β-galactosidase activity, plates contained 5-bromo-4-chloro-3-indolyl-β-Dgalactopyranoside<br />

(X-GAL) at a final concentration of 80 µg/ml.<br />

DNA techniques<br />

Procedures for DNA amplification, restriction, ligation and transformation of E. coli DH5α were<br />

carried out according to standard laboratory procedures (Sambrook et al., 1989). Chromosomal DNA<br />

of B. subtilis was isolated according to Bron and Venema (Bron and Venema, 1972). B. subtilis was<br />

transformed as described by Kunst and Rapoport (Kunst and Rapoport, 1995). S. <strong>aureus</strong> was<br />

transformed by electroporation as described by Kreiswirth et al. (Kreiswirth et al., 1983). All primers<br />

used for PCR are listed in Table 2. PCR products were purified using the High Pure PCR Purification<br />

Kit (Roche Applied Science).<br />

Table 2. Primers used in this study<br />

Primer Sequence (5’→3’)<br />

pSU1 ATATATACCATCATTGAATCGAGA<br />

pSU2 CAAGACGAACTCCAATTCAC AAAACTATCGTCAATTCTGCAGA<br />

pYF1 TACATCCGCAACTGTCCATA GATTATCATAACTACATATTCCAT<br />

pYF2 GCTACTCAGTAAGCTTGCACT<br />

Kana1 GTGAATTGGAGTTCGTCTTG<br />

Kana2 TATGGACAGTTGCGGATGTA<br />

mscL-F1 CATAGCAAACCATGAAATTGT<br />

mscL-R1 TCACGTCAGTCAGTCACCATGGCATTACACTCAACCTCTCTTTTT<br />

mscL-F2 TGCCATGGTGACTGACTGACGTGATTTTTAAATAAAAAGAGATGG<br />

mscL-R2 GCAAATCTGAGATTAGCAACA<br />

sunA Fwd CCGCGGGCTTTCCCAGCCAAATAGTTAGTATTAAAGAGTCAGAC<br />

sunA Rev GTTCCTCCTTCCCACCGTTTGCAATCCGGATTACATT<br />

mscL Fwd CCGCGGGCTTTCCCAGCTTAAAGAAATTATTCCTCACC<br />

mscL Rev GTTCCTCCTTCCCACCTATATTTGTAAAGAAAAAAAGAC<br />

Note: <strong>The</strong> 5' sequences of primer pSU2 (printed in italics) is complementary to the Kana1 primer. <strong>The</strong> 5'<br />

sequences of primer pYF1 (printed in italics) are complementary to the Kana2 primer. Sequences for linkage of<br />

amino acids are shown in bold. Sequences for ligation-independent cloning are shown underlined.<br />

Construction of B. subtilis and S. <strong>aureus</strong> mscL mutant strains<br />

A B. subtilis 168 strain lacking the mscL gene (originally named ywpC) was obtained from the B.<br />

subtilis functional analysis (BSFA) program. This ∆mscL strain was constructed in the laboratory of<br />

Prof. G. Rapoport according to a protocol described in detail on the Micado website<br />

(http://genome.jouy.inra.fr/cgi-bin/micado/index.cgi). Briefly, a PCR fragment containing the 500 bp<br />

region upstream of the mscL gene was fused to a fragment containing the 500 bp region downstream of<br />

the mscL gene and cloned into a pMU<strong>TI</strong>N4mcs vector using the BamHI and HindIII sites. This plasmid<br />

was subsequently used to transform B. subtilis 168, thereby replacing the mscL gene with the pMU<strong>TI</strong>N<br />

vector in a double crossover recombination event, which yielded strain ∆mscL. Strain ∆SPβ∆mscL was<br />

constructed by transformation of strain ∆SPβ with genomic DNA of strain ∆mscL and selection for<br />

erythromycin resistant transformants. Correct deletion of mscL from the genomes of the ∆mscL and<br />

∆SPβ∆mscL strains was verified by PCR.<br />

135


Chapter 7<br />

Mutants of S. <strong>aureus</strong> were constructed using the chromosomal integration-excision approach<br />

described by Arnaud et al. (Arnaud et al., 2004). Primers for the downstream (mscL-F2/mscL-R2) and<br />

upstream (mscL-F1/mscL-R1) regions of mscL (Table 2) were designed to obtain PCR products of 524<br />

and 522 bp, respectively, including a 24 bp linker. <strong>The</strong>se PCR products were linked in 10 PCR cycles.<br />

<strong>The</strong> resulting 1025 bp product was directly ligated into the TOPO-vector (Invitrogen) according to the<br />

manufacturer’s protocol. This construct was digested with BamHI and the 1030 bp product was ligated<br />

into the chromosomal integration-excision vector pMAD, resulting in pMAD-mscL. S. <strong>aureus</strong> RN4220<br />

cells were transformed with the pMAD-mscL plasmid by electroporation, and grown on TSA-plates<br />

containing erythromycin and X-GAL for 48 hours at 30°C. To obtain cells with a chromosomally<br />

integrated copy of pMAD-mscL, blue colonies were used to inoculate overnight cultures in BHI<br />

medium. Next, 10 ml BHI was inoculated with 100 µl overnight culture, grown for 1 hour at 30°C and<br />

then transferred to 42°C for 6 hours. To select cells with a chromosomally integrated copy of pMADmscL,<br />

dilutions (1000x) of the culture were plated on TSA plates with erythromycin and X-GAL and<br />

incubated for 48 hours at 42°C. To subsequently obtain cells that had excised pMAD-mscL from the<br />

chromosome, blue colonies with integrated pMAD-mscL were used to inoculate overnight cultures in<br />

BHI medium at 42°C. Next, 10 ml BHI was inoculated with 10 µl of the overnight culture and growth<br />

was continued for 6 hours at 30°C. Dilutions (1000x) of the cultures were plated on TSA plates with X-<br />

GAL and incubated at 42°C for 48 hours. White colonies were tested for erythromycin sensitivity and<br />

checked for the presence or absence of mscL by colony PCR. <strong>The</strong> correct deletion of mscL was<br />

confirmed by PCR on isolated genomic DNA using the Bacterial Genomic DNA Isolation Kit (Sigma).<br />

Lantibiotic activity assays<br />

A sublancin 168-induced B. subtilis growth inhibition assay was performed on plates essentially as<br />

described by Dorenbos et al. (Dorenbos et al., 2002). Briefly, indicator strains and strains to be tested<br />

for sublancin 168 production were grown overnight in LB broth containing the appropriate<br />

antibiotic(s). Overnight cultures of the indicator strains were then diluted 100-fold in LB, and 100 µl<br />

aliquots of the diluted cultures were plated on LB agar. After drying of the plates, 2 µl aliquots of<br />

undiluted overnight cultures of strains to be tested for sublancin 168 production were spotted.<br />

Alternatively, aliquots of nisin, Pep5, or concentrated spent medium with sublancin 168 were spotted.<br />

<strong>The</strong> plates were then incubated overnight at 37 o C, and growth inhibition of the indicator strain was<br />

analyzed the next day. Nisin was obtained from Sigma, and Pep5 was kindly provided by Dr. Hans<br />

Georg Sahl.<br />

Lyophilization of spent medium<br />

<strong>The</strong> concentration of sublancin 168 in spent medium of B. subtilis 168 cells was increased by<br />

lyophilization. For this purpose, B. subtilis 168 or the negative control strain ∆sunA were grown in 25<br />

ml of LB broth containing 0% NaCl. At an OD600 of 2.5, cells were separated from the growth medium<br />

by centrifugation and medium fractions were frozen in liquid nitrogen. <strong>The</strong> frozen growth media were<br />

lyophilized for 4 days under vacuum using a freeze dryer. After this period, the lyophilized medium<br />

was resuspended in 500 µl demineralized water and filtered (0.22 µm) before use. 2 µl of the<br />

concentrate was spotted on agar plates.<br />

Co-culturing of B. subtilis and S. <strong>aureus</strong> strains<br />

B. subtilis 168 and S. <strong>aureus</strong> strains were grown separately overnight in LB medium. In the morning,<br />

cultures were diluted to an OD600 of 0.05 in fresh LB medium. Next, specific strains to be tested<br />

together were mixed in a 1:1 ratio, resulting in co-cultures consisting of 50% sublancin producing cells<br />

(B. subtilis 168 Cm) and 50% non-producing cells (either B. subtilis ∆SPβ Tc, B. subtilis ∆SPβ∆mscL,<br />

S. <strong>aureus</strong> RN4220 Em, or S. <strong>aureus</strong> RN4220 ∆mscL). Upon mixing, growth was continued for eight or<br />

nine hours. Samples for plating were taken at hourly intervals during growth. <strong>The</strong> samples thus<br />

obtained were diluted 10 4 or 10 6 fold, and plated on LB agar containing specific antibiotics that permit<br />

growth of only one of the two co-cultured strains. After overnight incubation at 37 o C, resistant colonies<br />

136


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

were counted, and numbers of colony forming units (CFU) per ml of culture of each strain at the time<br />

of sampling were calculated. In case of the S. <strong>aureus</strong> RN4220 ∆mscL strain, which does not contain an<br />

antibiotic resistance marker, no antibiotics were present in the plates used to determine the CFU<br />

numbers of this strain. Specifically, the CFU number of the ∆mscL strain was calculated by subtraction<br />

of the CFU number of the co-cultivated antibiotic resistant strain from the total CFU number of the two<br />

co-cultivated strains. <strong>The</strong> presence of B. subtilis and/or S. <strong>aureus</strong> in samples used for plating was<br />

inspected by light microscopy.<br />

Promoter activity assay<br />

<strong>The</strong> promoter regions of mscL and sunA were amplified by PCR from genomic DNA prepared from B.<br />

subtilis 168 using the primers described in Table 2. <strong>The</strong> resulting PCR fragments were prepared for<br />

ligation-independent cloning using T4 DNA Polymerase (Novagen) and dTTP (Aslanidis and de Jong,<br />

1990). <strong>The</strong> BaSysBioII cloning vector was digested with SmaI, purified from an agarose gel and<br />

prepared for ligation-independent cloning using T4 DNA polymerase (Novagen) and dATP. 1 µl of the<br />

treated vector and 3 µl of the treated insert were mixed and left to anneal at room temperature before<br />

transformation of E. coli. <strong>The</strong> resulting constructs PmscL-GFP and PsunA-GFP were used to transform<br />

B. subtilis 168 and transformants were selected on LB plates containing spectinomycin. Strains with<br />

mscL-GFP or sunA-GFP fusions were pre-cultured in LB containing 1% or 5% NaCl and diluted 1:100<br />

in the same medium three hours before the start of the growth experiments. Next, the cells were diluted<br />

in the respective medium to an optical density at 600 nm (OD600) of 0.01. Growth was continued in<br />

triplicate wells of a 96-well black optical bottom microtitre plate (Nunc) that was placed in a Biotek<br />

Synergy 2 plate reader (37°C, variable shaking). <strong>The</strong> OD600 and fluorescence (excitation 485/20,<br />

emission 528/20) of the strains were measured for 14 hours. Fluorescence measurements were<br />

processed essentially as described by Ronen et al. (Ronen et al., 2002). Before starting the experiment<br />

the OD’s of the wells at 977 nm and 900 nm were measured to allow light path correction to 1 cm. For<br />

each of the recorded fluorescence data points the blank of neat LB (average of 3 wells) was removed<br />

and each point was corrected for a light path of 1 cm using the following equation: (0.18/(OD977 -<br />

OD900)). <strong>The</strong> average of the three samples was then calculated. <strong>The</strong> fluorescence data was further<br />

processed by calculating the average of the three samples and subtracting the background fluorescence<br />

of the parental strain 168 (without GFP) for each data point. <strong>The</strong> promoter activity was calculated using<br />

the following equation: (GFP(t) – GFP (t-1))/O.D. (t). We acknowledge the collaborative effort with<br />

the teams of Stéphane Aymerich, Kevin Devine, Vincent Fromion, and Tony Wilkinson in<br />

standardising the fluorescence measurements. Each experiment was repeated at least three times.<br />

Results<br />

Sublancin 168 activity is salt dependent.<br />

To study the activity of sublancin 168 under different growth conditions we made use of a<br />

previously developed sublancin growth inhibition plate assay. <strong>The</strong> essence of this assay, in<br />

which strains potentially producing sublancin 168 were spotted onto a lawn of sensitive or<br />

immune indicator cells, is demonstrated in Figure 1A. This Figure shows that the ∆SPβ strain,<br />

that lacks all genes of the SPβ prophage (including those for sublancin production and<br />

immunity), is not able to grow in the vicinity of the sublancin 168-producing parental strain<br />

168. This growth inhibition is strictly dependent on the presence of an intact copy of the sunA<br />

gene for sublancin 168 since no zone of growth inhibition is formed around ∆sunA spotted<br />

cells on a plated ∆SPβ cell layer (Figure 1A). As previously demonstrated (Dubois et al.,<br />

2009), the sublancin 168 susceptibility of the ∆SPβ cells is solely due to the absence of the<br />

sunI (yolF) immunity gene, since full sublancin 168 immunity can be restored by ectopic<br />

expression of sunI (Figure 1A). To screen for possible sublancin 168 activity determinants,<br />

137


Chapter 7<br />

we deployed this assay to monitor the effects of growth medium composition on the activity<br />

of sublancin 168. It was thus noticed that the sublancin 168 activity was dependent on the<br />

type of growth medium used in the plate assays (data not shown). Upon inspection of the<br />

composition of the tested media, it was found that, in particular, the NaCl content seemed to<br />

differ.<br />

Figure 1. Sublancin 168-mediated growth inhibition of B. subtilis. (A) sublancin 168 growth inhibition assay.<br />

Strains to be tested for sublancin 168 production were spotted on a lawn of indicator cells. <strong>The</strong> names of strains<br />

spotted to test for sublancin production are listed above the plate images. Names of the strains plated as indicators<br />

for sublancin sensitivity/immunity are listed below the plate images. (B) NaCl-dependent growth inhibition by<br />

sublancin 168. Strain 168 was used as the spotted sublancin 168 producer, strain ∆SPβ was used as the plated<br />

indicator for sublancin activity. <strong>The</strong> percentage of NaCl listed below each image indicates the amount of NaCl that<br />

was present in the LB agar plate and in the liquid LB cultures from which the plated or spotted cells were derived.<br />

(C) Growth inhibition by sublancin 168 depends on the presence of MscL. <strong>The</strong> sublancin 168 producer B. subtilis<br />

168 was spotted on a lawn of plated ∆SPβ ∆mscL indicator cells as described for panel B. (D) Growth inhibition<br />

by nisin or Pep5 does not depend on the presence of MscL. B. subtilis 168 or 2µl aliquots of either 2.5 mg/ml nisin<br />

or 10 mg/ml Pep5 dissolved in water were spotted on a lawn of plated ∆SPβ or ∆SPβ∆mscL indicator cells on LB<br />

agar as in panel A.<br />

138


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

To investigate whether the NaCl concentration might influence the outcome of our sublancin<br />

growth inhibition assay, we performed a series of sublancin 168 activity assays in which the<br />

LB broth and agar media used for growth of the 168 and ∆SPβ cells contained increasing<br />

concentrations of NaCl, ranging from 0 to 5%. <strong>The</strong> results of these experiments showed that<br />

the size of the growth inhibition zone of the ∆SPβ cells is inversely correlated with the NaCl<br />

content of the growth medium (Figure 1B). Next, we investigated whether this effect of NaCl<br />

was related to the plate assay or whether this also occurred in liquid media. For this purpose,<br />

we performed co-culturing and competition experiments in liquid medium by inoculation of<br />

the sublancin 168 producing B. subtilis strain 168 amyE::pX (chloramphenicol R ) in growth<br />

medium in a 1:1 ratio with the non-producing B. subtilis strain ∆SPβ amyE::pXTC<br />

(tetracycline R ). This was done both in LB medium containing the standard concentration of<br />

1% NaCl and in LB medium containing 5% NaCl. <strong>The</strong> results of co-cultivation and<br />

subsequent transfer of samples to plates containing either chloramphenicol or tetracycline<br />

showed that the ∆SPβ strain was able to survive only for a few hours in the presence of the<br />

sublancin 168 producing strain when these strains were grown in standard LB medium<br />

(Figure 2A). In contrast, when the co-cultures were grown in LB containing 5% NaCl, the<br />

∆SPβ strain was not inhibited by the presence of B. subtilis 168 (Figure 2B). As expected, the<br />

deleterious effect of the 168 strain on the ∆SPβ strain was not observed when the sunA gene<br />

was deleted from the 168 strain (Figure 2C). <strong>The</strong>se findings were fully consistent with those<br />

obtained by the sublancin 168 growth inhibition assays on LB agar (Figure 1B). Taken<br />

together, the results show that the NaCl concentration in the growth medium influences the<br />

outcome of sublancin 168 growth inhibition assays. This suggests that either the production of<br />

sublancin 168, the activity of sublancin 168 or the susceptibility of the target cells are<br />

influenced by the concentration of NaCl in the growth medium.<br />

Figure 2. Assessment of salt-dependent sublancin 168 activity, and MscL-determined sublancin 168<br />

susceptibility, by co-cultivation in liquid cultures. Co-cultures of B. subtilis 168 Cm (white bars) together with B.<br />

subtilis ∆SPβ Tet or ∆SPβ∆mscL Tet (black bars) in LB containing 1% NaCl (A/D) or 5% NaCl (B/E). Likewise,<br />

B. subtilis 168 ∆sunA (white bars) was co-cultured together with B. subtilis ∆SPβ Tet (black bars) in normal LB<br />

containing 1% NaCl (C). <strong>The</strong> overnight grown strains 168 Cm, 168 ∆sunA, ∆SPβ Tet or ∆SPβ ∆mscL Tet were<br />

diluted to an OD 600 of 0.05 in fresh LB medium containing the required amounts of NaCl and mixed in a 1:1 ratio,<br />

resulting in co-cultures consisting of 50% B. subtilis 168 Cm or B. subtilis 168 ∆sunA plus 50% of B. subtilis<br />

∆SPβ Tet or B. subtilis ∆SPβ∆mscL Tet. Growth was continued and monitored by OD 600 measurements (depicted<br />

as a black line) and samples were plated at hourly intervals. Chloramphenicol, kanamycin and tetracycline<br />

resistant colonies were counted and used to calculate the number of colony forming units (CFU) per ml of culture<br />

for each strain at each time point of sampling.<br />

139


Chapter 7<br />

NaCl affects the sublancin 168 susceptibility of B. subtilis<br />

To explore the nature of the effect of NaCl, as observed in the sublancin 168 activity assays,<br />

we first investigated a possible influence of NaCl on the production of sublancin 168. As an<br />

indication for this, we measured the activity of the sunA promoter in the B. subtilis 168 strain<br />

grown in LB medium with either 1% or 5% NaCl, using a transcriptional fusion between the<br />

sunA promoter and the GFP gene (PsunA-GFP). <strong>The</strong> results of this analysis show that the<br />

sunA promoter displayed similar average promoter activities in both media, although the 168<br />

cells grew somewhat slower in LB medium with 5% NaCl than in LB medium with 1% NaCl<br />

(Figure 3A). This suggests that the different outcomes in the sublancin 168 growth inhibition<br />

assays in the presence of 1% or 5% NaCl are not due to differences in the levels of sublancin<br />

168 production by B. subtilis 168. Interestingly, expression of sunA is rather suppressed until<br />

late exponential phase, which seems to coincide with the onset of growth inhibition as<br />

observed in Figure 2A.<br />

A<br />

5000<br />

4000<br />

3000<br />

2000<br />

Promoter<br />

B<br />

Promoter<br />

Activity<br />

140<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

-500<br />

-1000<br />

-1500<br />

0<br />

1<br />

0 100 200 300 400 500 600<br />

Figure 3. Promoter activities of sunA and mscL. Promoter activities of sunA (A) and mscL (B) were measured<br />

with a B. subtilis 168 strain expressing GFP from the native sunA or mscL promoter. Promoter activity is<br />

calculated as (GFP(t)-GFP(t-1)) / OD 600. Cells were grown in LB containing 1% NaCl (black lines) or 5% NaCl<br />

(grey lines). Promoter activity is shown with continuous lines; OD 600 is shown with interrupted lines.<br />

As an alternative, we investigated the possible direct effects of NaCl on the stability or<br />

activity of sublancin 168. For this purpose, we concentrated sublancin 168 by lyophilization<br />

of spent medium (without NaCl) from the sublancin producer B. subtilis 168. By doing so, we<br />

increased the sublancin 168 concentration 50-fold. Figure 4A shows that a spotted concentrate<br />

of this spent medium inhibits the growth of plated ∆SPβ cells, whereas a control concentrate<br />

from the spent medium of a ∆sunA strain does not inhibit the growth of plated ∆SPβ cells.<br />

Next, we used the concentrate from the 168 spent medium to establish whether NaCl might<br />

directly affect the stability or activity of sublancin 168. For this purpose, the concentrated<br />

spent media were incubated overnight with or without 5% NaCl, prior to spotting. <strong>The</strong> very<br />

similar sizes of the resulting growth inhibition zones of ∆SPβ cells on plates without NaCl<br />

(Figure 4B) revealed that the activity of the concentrated sublancin 168 was not affected by<br />

overnight incubation with 5% NaCl. This indicated that the absence of sublancin 168-directed<br />

growth inhibition during growth on LB with 5% NaCl is not due to an irreversible<br />

inactivating effect of NaCl on sublancin 168.<br />

To investigate the possible effect of NaCl on the ∆SPβ indicator cells, we also spotted<br />

the concentrated 168 medium on ∆SPβ cells that were plated on LB agar plates containing 5%<br />

NaCl. <strong>The</strong> results show that ∆SPβ cells plated on LB agar containing 5% NaCl were no<br />

longer sensitive to the spotted sublancin 168, in contrast to ∆SPβ cells plated on LB agar<br />

containing 0% NaCl (Figure 4B). Taken together, these findings indicate that NaCl does not<br />

Time<br />

2,5<br />

2<br />

1,5<br />

0,5<br />

0<br />

OD 600nm


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

influence the production, activity or stability of sublancin 168. Instead NaCl seems to<br />

influence the sublancin 168 susceptibility of B. subtilis.<br />

Figure 4. NaCl influences the sublancin 168 susceptibility of B. subtilis. (A) Sublancin 168 growth inhibition<br />

assays were performed with concentrated spent medium fractions (“concentrate”) of B. subtilis 168 (contains<br />

sublancin 168) or B. subtilis 168 ∆sunA (no sublancin present), which were spotted on a lawn of plated ∆SPβ<br />

cells. <strong>The</strong> concentrates were obtained by lyophilization of spent media. (B) Before spotting, the concentrated<br />

sublancin 168 was incubated overnight with 0 or 5% salt, indicated by “NaCl in concentrate”. <strong>The</strong> NaCl content<br />

in the LB agar plates is indicated below the images.<br />

Sublancin 168 activity depends on the presence of MscL<br />

In search for possible cellular mechanisms that could be involved in the NaCl-dependent<br />

susceptibility of target cells towards sublancin 168, we explored the available literature for<br />

known effects of NaCl on bacterial growth and survival. This drew our attention to a possible<br />

role of the mechanosensitive channels of ion conductance. <strong>The</strong>se channels are located in the<br />

cytoplasmic membrane and catalyze the efflux of ions and osmolytes from the cytoplasm<br />

when cells encounter a downshift in the osmolarity of their environment (Blount et al., 1999;<br />

Martinac, 2001; Pivetti et al., 2003; Sukharev et al., 1997). <strong>The</strong> opening or closing of these<br />

channels is dependent on the osmolarity (salt content) of the media in which cells are<br />

growing. At high osmolarity of the medium the channels will mostly be closed, while at low<br />

osmolarity they may be open more frequently due to a constantly increasing turgor pressure in<br />

the cell. We therefore investigated a possible involvement of these channels in the observed<br />

NaCl-dependent sublancin 168 susceptibility of B. subtilis. For this purpose, the gene<br />

encoding the largest mechanosensitive channel, named MscL, was deleted from the genome<br />

of the sublancin 168 sensitive strain ∆SPβ. <strong>The</strong>n, sublancin 168 growth inhibition assays were<br />

performed with this ∆SPβ∆mscL strain to assess its susceptibility to sublancin 168 in the<br />

presence of different concentrations of NaCl. <strong>The</strong> results, shown in figure 1C, were striking<br />

141


Chapter 7<br />

since the growth of B. subtilis ∆SPβ∆mscL was not at all inhibited by the sublancin 168<br />

producer strain, irrespective of the absence or presence of NaCl (Figure 1C). <strong>The</strong>se results<br />

were confirmed by co-culturing experiments with the 168 and ∆SPβ∆mscL strains in liquid<br />

media, which showed that growth of the ∆SPβ∆mscL strain was not inhibited by the parental<br />

strain producing sublancin 168, irrespective of the NaCl concentration (Figure 2, D and E).<br />

This shows that the susceptibility of the ∆SPβ cells to sublancin 168 depends on the presence<br />

of the mscL gene. To investigate a possible effect of NaCl on the production level of MscL,<br />

we monitored the mscL promoter activity in B. subtilis cells during growth in LB medium<br />

containing 1% or 5% NaCl. This was done with a transcriptional PmscL-GFP fusion<br />

construct. <strong>The</strong> results of this analysis show that the mscL promoter activity profile of cells<br />

grown in LB medium with 5% NaCl is comparable to that of cells grown in LB with 1% NaCl<br />

(Figure 3B). Growth in LB medium with 5% NaCl does therefore not seem to prevent MscL<br />

production, which is consistent with our previously published results on the effects of hypoosmotic<br />

shock on MscL proficient and deficient cells (Kouwen et al., 2009b). Taken together,<br />

these observations imply that the susceptibility of sensitive B. subtilis cells toward sublancin<br />

168 is dependent on the production of MscL.<br />

To investigate whether MscL is also involved in the susceptibility of B. subtilis to<br />

lantibiotics other than sublancin 168, we tested the sensitivity of ∆SPβ∆mscL cells for the<br />

type A lantibiotics Nisin and Pep5 using a plate assay (Figure 1D). <strong>The</strong> results showed that<br />

∆SPβ∆mscL cells have no altered sensitivity towards 2.5 mg/ml Nisin or 10 mg/ml Pep5 as<br />

compared to the ∆SPβ strain. <strong>The</strong>se results therefore indicate that MscL is a specific<br />

determinant for sublancin 168 susceptibility in B. subtilis.<br />

MscL-dependent sublancin 168 susceptibility is conserved in <strong>Staphylococcus</strong> <strong>aureus</strong><br />

Sublancin 168 has antimicrobial activity against a broad range of Gram-positive bacteria. <strong>The</strong><br />

observed MscL-dependent sublancin 168 susceptibility of B. subtilis prompted us to<br />

investigate whether this phenomenon is specific for B. subtilis 168, or whether MscL is also<br />

involved in the sublancin 168 susceptibility of other bacteria. <strong>The</strong>refore, the sublancin 168<br />

activity against the pathogenic Gram-positive bacterium S. <strong>aureus</strong> was also investigated. As a<br />

first approach, the sublancin 168 producing B. subtilis strain 168 amyE::pX<br />

(chloramphenicol R ) was used to inoculate LB growth medium in a 1:1 ratio with the S. <strong>aureus</strong><br />

strain RN4220 Em (erythromycin R ). NaCl was present at concentrations of either 1% or 5%.<br />

<strong>The</strong> results of this co-cultivation and subsequent transfer of samples to plates (Figure 5, A and<br />

B) were comparable to those obtained with the B. subtilis ∆SPβ strain (Figure 2, A and B).<br />

When grown in normal LB containing 1% NaCl, the S. <strong>aureus</strong> strain was only able to survive<br />

for a few hours in the presence of the sublancin 168 producing B. subtilis strain (Figure 5A),<br />

whereas the S. <strong>aureus</strong> strain was hardly inhibited by the sublancin 168 producing B. subtilis<br />

strain when grown in LB medium containing 5% NaCl (Figure 5B). To confirm that this<br />

inhibitory effect was indeed due to the sublancin 168 produced by the B. subtilis 168 strain,<br />

we also co-cultured the ∆sunA strain with the S. <strong>aureus</strong> strain RN4220 Em. Indeed, the results<br />

show that the S. <strong>aureus</strong> strain was hardly inhibited by the ∆sunA strain, although a slight<br />

inhibitory effect of B. subtilis on the growth of S. <strong>aureus</strong> was still detectable (Figure 5,<br />

compare A and C). Notably, when we performed sublancin 168 growth inhibition assays on<br />

agar plates during which the ∆sunA strain was spotted on a lawn of plated RN4220 Em cells,<br />

a growth inhibition zone around the ∆sunA strain was still observed (data not shown).<br />

142


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

Figure 5. Sublancin 168 susceptibility of S. <strong>aureus</strong> is determined by MscL. (A) Co-cultivation of B. subtilis 168<br />

Cm (white bars) together with S. <strong>aureus</strong> RN4220 Em (black bars) in LB containing 1% NaCl. (B) Co-cultivation of<br />

B. subtilis 168 Cm (white bars) together with S. <strong>aureus</strong> RN4220 Em (black bars) in LB containing 5% NaCl. (C)<br />

Co-cultivation of B. subtilis 168 ∆sunA (white bars) together with S. <strong>aureus</strong> RN4220 Em (black bars) in LB<br />

containing 1% NaCl. (D) Co-cultivation of B. subtilis 168 Cm (white bars) together with S. <strong>aureus</strong> RN4220 ∆mscL<br />

Em (black bars) in LB containing 1% NaCl. All strains were grown overnight and diluted to an OD 600 of 0.05 in<br />

fresh LB medium containing the indicated amounts of NaCl and mixed at a 1:1 ratio. <strong>The</strong> number of colony<br />

forming units (CFU) per ml was calculated as described in the legend to Figure 2 and the “Materials and<br />

Methods” section. OD 600 measurements are depicted as a black line.<br />

This indicated that B. subtilis also produces other antimicrobial factors to which S. <strong>aureus</strong> is<br />

sensitive. However, sublancin 168 seems to represent the most effective anti-staphylococcal<br />

activity of B. subtilis, as can be concluded from the strong inhibitory effect on the growth of<br />

S. <strong>aureus</strong> shown in Figure 5. Finally, we tested whether sublancin 168 susceptibility of S.<br />

<strong>aureus</strong> was also dependent of MscL. For this purpose, we constructed an S. <strong>aureus</strong> RN4220<br />

strain lacking the mscL gene, and performed subsequent co-culturing experiments of this<br />

strain together with the sublancin producing strain B. subtilis 168. <strong>The</strong> results of this coculturing<br />

showed that growth of the RN4220 ∆mscL strain was no longer inhibited by the<br />

deleterious effects of B. subtilis 168 producing sublancin (Figure 5D). Also in this case, the<br />

RN4220 ∆mscL S. <strong>aureus</strong> strain was still slightly inhibited by the B. subtilis 168 strain, but<br />

this effect was comparable to the effect observed when the S. <strong>aureus</strong> RN4220 strain was cocultured<br />

with the B. subtilis ∆sunA strain. This clearly shows that the S. <strong>aureus</strong> RN4220<br />

∆mscL strain is not sensitive to sublancin 168. Taken together, these observations demonstrate<br />

that the susceptibility of S. <strong>aureus</strong> to sublancin 168 is dependent on the presence of an MscL<br />

channel and that this feature is conserved in B. subtilis and S. <strong>aureus</strong>.<br />

143


Chapter 7<br />

Discussion<br />

In the present manuscript we report on bacterial and environmental factors that determine<br />

bacterial susceptibility to the lantibiotic sublancin 168. Since its discovery in 1980, the<br />

antimicrobial mechanism of this broad-range and highly stable lantibiotic has never been so<br />

much as hinted at. We now show that the sublancin 168 susceptibility of cells lacking the<br />

immunity protein SunI is dependent on the NaCl content of the growth medium. By contrast,<br />

the production, stability and activity of the sublancin 168 seem to remain largely unaffected<br />

by NaCl. Furthermore, we show that MscL plays a critical and unprecedented role in the<br />

mode of action of sublancin 168, since this channel is indispensable for sublancin 168<br />

susceptibility. As MscL determines sublancin 168 susceptibility both in B. subtilis and S.<br />

<strong>aureus</strong>, it is well conceivable that the respective mechanism is conserved in many sublancin<br />

168 susceptible bacteria.<br />

Mechanosensitive ion channels are membrane-embedded channels that are present in<br />

all three domains of life, but are especially widespread among bacteria (Blount et al., 1999;<br />

Martinac, 2001; Pivetti et al., 2003; Sukharev et al., 1997). So far, three families of<br />

mechanosensitive channels have been distinguished, named MscM (Mini), MscS (Small), and<br />

MscL (Large). <strong>The</strong>se channels are activated at different levels of applied pressure (Berrier et<br />

al., 1996). <strong>The</strong> MscL channel opens at the highest applied pressure and also has the highest<br />

pore diameter (30 Å). Mechanosensitive channels catalyze the efflux of osmolytes or<br />

osmoprotectants upon hypo-osmotic shock (Berrier et al., 1992). When cells encounter a<br />

sudden downshift in osmolarity of their environment they respond by excreting ions and<br />

osmolytes from the cytoplasm in order to maintain an appropriate turgor pressure. <strong>The</strong> cells<br />

thereby protect themselves from death by lysis due to overpressure. Possession of effective<br />

osmo-regulatory protection mechanisms seems, therefore, of vital importance, in particular<br />

for soil-dwelling organisms that are frequently exposed to hypo-osmotic shocks, such as rain.<br />

In fact, it has been reported that B. subtilis cells lacking functional mechanosensitive<br />

channels, especially MscL, are severely compromised in their ability to survive a hypoosmotic<br />

shock (Hoffmann et al., 2008; Wahome and Setlow, 2008). Notably, in the present<br />

studies we did not expose the cells to hypo-osmotic shock, but merely grew them in LB media<br />

with varying concentrations of NaCl. Furthermore, we also did not observe any difference in<br />

viability between the B. subtilis or S. <strong>aureus</strong> parental strains and their ∆mscL derivative<br />

strains under the applied conditions as long as sublancin 168 was absent. It seems therefore<br />

unlikely that sublancin 168 exerts its bactericidal activity by blocking the MscL channels of<br />

sensitive organisms. Yet, the sublancin 168 susceptibility was clearly shown to depend on the<br />

presence of MscL. Although MscL channels are vital for lowering the turgor pressure during<br />

osmotic shock, it is not unlikely that these channels can also occasionally open during a<br />

constant environmental osmolarity, especially if cells are grown in media of low osmolarity.<br />

Since an open state of the MscL channel is potentially hazardous for cells due to ion loss, and<br />

since a closed state of the MscL channel is more or less equivalent to the absence of MscL, it<br />

is very well conceivable that sublancin 168 susceptibility relates to an open state of the MscL<br />

pore. This would fit with the observed protective effect of salt against the detrimental effects<br />

of sublancin 168. In fact, this protective effect of salt was the prime reason for us to<br />

investigate a possible involvement of MscL in bacterial susceptibility to sublancin 168. <strong>The</strong><br />

MscL channels open at a certain level of pressure on the membrane, which is usually the<br />

result of cell swelling due to a difference in the osmolarity (or salt content) between the<br />

cytoplasm and the environment. <strong>The</strong> higher the salt content of the media, the less likely it is<br />

144


<strong>The</strong> large mechanosensitive channel MscL determines bacterial susceptibility to<br />

the bacteriocin sublancin 168<br />

that mechanosensitive channels are open. <strong>The</strong>refore, the observation that addition of salt to<br />

the media lowers the susceptibility to sublancin 168, apparently without affecting sublancin<br />

168 production or activity, seems to indicate that the MscL channel needs to be in an open<br />

state to allow for bactericidal activity of sublancin 168. If so, sublancin 168 might prevent the<br />

MscL pore from closing, which would result in cell death by leakage of ions from the<br />

cytoplasm. This putative mechanism would in fact be consistent with the classification of<br />

sublancin 168 as a group A lantibiotic, since lantibiotics of this class are known to create<br />

pores in the membrane. In this case, however, sublancin 168 would keep an already existing<br />

pore in an open state. Clearly, several alternative hypotheses can still be entertained. For<br />

example, MscL might function as a gate for sublancin 168 import into the cell, allowing it to<br />

perform bactericidal activity by interacting with an unidentified essential cytoplasmic target.<br />

Furthermore, MscL might be involved in an indirect process that is required to promote the<br />

bactericidal activity of sublancin 168.<br />

Interestingly, even though MscL is conserved in all Gram-positive organisms, not all<br />

of these organisms are inhibited by sublancin 168. This might be due to the existence of<br />

systems for lantibiotic producer immunity or natural lantibiotic resistance. Recently, we have<br />

identified the B. subtilis sublancin 168 producer immunity protein. This protein, SunI (YolF),<br />

was also found to give S. <strong>aureus</strong> immunity against sublancin 168 when heterologously<br />

expressed in this organism. SunI was shown to be a membrane protein with an Nout-Cin<br />

topology, the bulk of the protein facing the cytoplasm (Dubois et al., 2009). This topology<br />

was unprecedented for known lantibiotic producer immunity proteins but, notably, it seems<br />

compatible with all above-mentioned mechanisms by which MscL might confer susceptibility<br />

to sublancin 168. For example, SunI might close an MscL pore that is kept in an open state by<br />

sublancin 168, it might prevent interactions between sublancin 168 and MscL (or other<br />

compounds) in the membrane, or it might block the entry of sublancin 168 into the cytoplasm.<br />

If producer immunity proteins analogous to SunI are active in other Gram-positive bacteria,<br />

they may perhaps also be able to counteract the bactericidal effects of sublancin 168. A<br />

second and perhaps more likely possibility is the existence of effective resistance<br />

mechanisms. For example, it was shown that certain genes of the σ W regulon confer sublancin<br />

168 resistance to B. subtilis (Butcher and Helmann, 2006). Furthermore, an altered membrane<br />

or cell wall with an increased net positive charge might protect against the bactericidal effects<br />

of cationic lantibiotics, like sublancin 168 (Peschel et al., 1999; Peschel and Collins, 2001).<br />

Such a resistance mechanism has been shown to exist in <strong>Staphylococcus</strong> epidermidis which,<br />

indeed, is resistant to sublancin 168 (Paik et al., 1998).<br />

In conclusion, the present studies have focused a general interest on mechanosensitive<br />

channels as potential determinants for bacterial susceptibility towards bacteriocins.<br />

Specifically, we have identified a critical role of the MscL channel in the susceptibility of B.<br />

subtilis and S. <strong>aureus</strong> towards the lantibiotic sublancin 168. Our findings suggest that this<br />

may be a conserved phenomenon in sublancin 168 sensitive organisms. <strong>The</strong>refore, ongoing<br />

and future studies, including electrophysiology and interaction studies, are aimed at<br />

identifying the precise mechanisms by which (i) sublancin 168 exerts its bactericidal activity<br />

and (ii) MscL confers susceptibility to sublancin 168. Such studies will also further increase<br />

our knowledge of mechanosensitive channels in general, which is important as these channels<br />

serve important functions in all domains of life, including humans.<br />

145


Chapter 7<br />

Acknowledgements<br />

We thank Magda van der Kooi-Pol for technical assistance, Hans-Georg Sahl for the gift of<br />

purified Pep5, Leslie Aïchaoui-Denève, Stéphane Aymerich, Eric Botella, Kevin Devine,<br />

Mark Fogg, Vincent Fromion, Annette Hansen, Matthieu Jules, Pascal Neveu, Sjouke<br />

Piersma, Patrick Veiga and Tony Wilkinson for their collaboration in developing tools and<br />

protocols for pBaSysBioII-based fluorescence measurements, and other colleagues from the<br />

BaSysBio program for helpful discussions. Funding for this project was provided by the CEU<br />

projects LSHG-CT-2004-503468, LSHG-CT-2004-005257, LSHM-CT-2006-019064 and<br />

LSHG-CT-2006-037469, the transnational SysMO initiative through project BACELL<br />

SysMO, the European Science Foundation under the EUROCORES Programme EuroSCOPE,<br />

and grant 04-EScope 01-011 from the Research Council for Earth and Life Sciences of the<br />

Netherlands Organization for Scientific Research.<br />

146


147


“Although one can get very clever at home, progress comes a lot quicker<br />

when you step into a room with other people and start playing”<br />

-Stephen J. Howe-<br />

148


Chapter 8<br />

<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in<br />

different media and under different nutrient starvation<br />

conditions<br />

B. Voigt, T. Schweder, M.J.J.B. Sibbald, D. Albrecht, A. Ehrenreich, J. Bernhardt,<br />

J. Feesche, K.H. Maurer, G. Gottschalk, J.M. van Dijl, M. Hecker<br />

Published in Proteomics (2005) 6, 268-281<br />

149


Chapter 8<br />

Summary<br />

<strong>The</strong> now finished genome sequence of Bacillus licheniformis DSM 13 allows the<br />

prediction of the genes involved in protein secretion into the extracellular environment<br />

as well as the prediction of the proteins which are translocated. From the sequence 296<br />

proteins were predicted to contain an N-terminal signal peptide directing most of them<br />

to the Sec system, the main transport system in Gram-positive bacteria. Using 2-DE the<br />

extracellular proteome of B. licheniformis grown in different media was studied. From<br />

the approximately 200 spots visible on the gels, 89 were identified that either contain an<br />

N-terminal signal sequence or are known to be secreted by other mechanisms than the<br />

Sec pathway. <strong>The</strong> extracellular proteome of B. licheniformis includes proteins from<br />

different functional classes, like enzymes for the degradation of various macromolecules,<br />

proteins involved in cell wall turnover, flagellum- and phage-related proteins and some<br />

proteins of yet unknown function. Protein secretion is highest during stationary growth<br />

phase. Furthermore, cells grown in complex medium secrete considerably higher protein<br />

amounts than cells grown in minimal medium. Limitation of phosphate, carbon and<br />

nitrogen sources results in the secretion of specific proteins that may be involved in<br />

counteracting the starvation.<br />

150


<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media<br />

and under different nutrient starvation conditions<br />

Introduction<br />

Export of proteins from the cytoplasm to the extracellular environment is a common<br />

phenomenon for all kinds of cells, including bacteria. Exported bacterial proteins can fulfil<br />

different functions, either in the cell wall, like proteins involved in cell wall synthesis and<br />

turnover and in folding and quality control of exported proteins, or in the surrounding<br />

environment, like proteins involved in the usage of nutrient sources, in the communication<br />

between cells, proteins with antimicrobial activity and virulence factors (Antelmann et al.,<br />

2001; Gohar et al., 2002; Nouwens et al., 2003; Ziebandt et al., 2004). <strong>The</strong> entirety of the<br />

secreted proteins and the components of the secretory protein machinery have been defined as<br />

the <strong>secretome</strong> (Tjalsma et al., 2004).<br />

To direct proteins to the transport machineries located in the cytoplasmic membrane, they are<br />

usually synthesised as precursors with an N-terminal signal peptide. <strong>The</strong>se signal peptides are<br />

recognised by targeting factors in the cytoplasm which target the proteins to the transport<br />

machinery in the membrane. <strong>The</strong> protein is then transported through the membrane via this<br />

translocation machinery in an energy dependent manner. Finally, the signal peptide is cleaved<br />

by specific signal peptidases. Bacteria have developed different pathways to secrete proteins.<br />

<strong>The</strong> most important one is the general secretory or Sec pathway. <strong>The</strong> twin-arginine<br />

translocation pathway (Tat), ABC transporters and the pseudopilin pathway are secretion<br />

pathways for special purposes and have been found to transport only few proteins in Bacillus<br />

subtilis (Tjalsma et al., 2004; Jongbloed et al., 2002; Jongbloed et al., 2004).<br />

<strong>The</strong> highest level of protein secretion in B. subtilis was observed when cells were grown in a<br />

complex medium (Antelmann et al., 2001). Cells grown in minimal medium (MM) secrete<br />

considerable lower amounts of protein (Hirose et al., 2000). Most extracellular proteins have<br />

been found after entry into the stationary growth phase.<br />

B. licheniformis is known to secrete a number of different proteins into the extracellular<br />

medium and this ability has been used in the fermentation industry for a long time, especially<br />

for the production of industrial enzymes. <strong>The</strong> availability of the genome sequence of B.<br />

licheniformis allows the prediction of the composition of its <strong>secretome</strong>. For example, all<br />

genes for the Sec machinery have been found in the B. licheniformis sequence (Veith et al.,<br />

2004; Veith et al., 2004). In addition, there are also two tatA genes (i.e., tatAd and tatAy) and<br />

two tatC genes (i.e., tatCd and tatCy), which encode subunits of the Tat secretion machinery.<br />

<strong>The</strong> nature of many of the extracellular enzymes which have been deduced from the sequence<br />

of B. licheniformis suggests that they have functions in utilisation of alternative nutrient<br />

sources that might be present in the environment. <strong>The</strong>re are numerous hydrolases for the<br />

degradation of various high molecular weight carbohydrates (e.g., α-amylase, cellulase,<br />

chitinase) as well as several extracellular proteases and other degradative enzymes (Veith et<br />

al., 2004; Rey et al., 2004). Despite of its biotechnological significance there has been no<br />

investigation of the protein secretion process in B. licheniformis at a proteome-wide scale so<br />

far. <strong>The</strong> recently finished genome sequence of B. licheniformis DSM 13 opened the chance to<br />

establish proteomics of this organism. <strong>The</strong> cytoplasmic proteome and the regulation of central<br />

carbon metabolic pathways were in the focus of our first study (Voigt et al., 2004). In the<br />

study presented here, we analysed the secretion of proteins of B. licheniformis cells grown in<br />

different media and under different nutrient starvation conditions by 2-DE.<br />

151


Chapter 8<br />

Materials and methods<br />

Strains and growth conditions<br />

B. licheniformis DSM 13 (equivalent to ATCC 14580, type strain from the German Collection of<br />

Microorganisms and Cell Cultures, DSMZ GmbH Braunschweig, Germany) was cultivated at 37ºC<br />

under vigorous agitation either in a complex medium (Luria Broth, LB) or in MM containing 15mM<br />

(NH4)2SO4, 8mM MgSO4 x 7 H2O, 27mM KCl, 7mM sodium citrate (2 H2O), 50mM Tris-HCl pH 7.5,<br />

supplemented with 0.6mM KH2PO4, 2mM CaCl2 (2 H2O), 1 mM FeSO4 x 7 H2O, 10 mM MnSO4 x 4<br />

H2O and 0.2% glucose. For the starvation experiments, the concentrations of glucose, phosphate and<br />

nitrogen were reduced to 0.08%, 0.15mM and 1.0mM, respectively. <strong>The</strong> concentration of the limiting<br />

nutrient was adjusted in such a way that the cultures grew to a maximum OD of about 1.0 (growth<br />

curve in LB and MM see Voigt et al. (Voigt et al., 2004)).<br />

Preparation of the extracellular protein fraction<br />

Bacteria were harvested during exponential growth (OD500 0.4 in MM, OD540 2.0 in LB), at the onset of<br />

the stationary growth phase (OD500 1.0 in MM, OD540 4.0 in LB) and during the stationary phase (MM<br />

1 h after transition into the stationary phase, OD540 6.0 in LB). PMSF (3mM) was added when the<br />

cultures were harvested to prevent proteolysis during sample preparation. <strong>The</strong> cells were removed by<br />

centrifugation at 4ºC (8000 rpm, 10 min). TCA was then added to the medium to a final concentration<br />

of 10%. <strong>The</strong> extracellular proteins were precipitated at 47C overnight and collected by centrifugation<br />

(4ºC, 10 000 rpm, 60 min). <strong>The</strong> protein pellet was washed eight times with 96% ethanol, dried and<br />

dissolved in a solution containing 8 M urea and 2 M thiourea. <strong>The</strong> protein concentration of the extract<br />

was determined with the RotiNanoquant Kit (Roth).<br />

2-DE<br />

For the IEF protein extracts (500 mg protein) were loaded onto commercially available IPG strips (pH<br />

3–10 NL, Amersham Biosciences) according to Büttner et al. (Büttner et al., 2001). In the second<br />

dimension, polyacrylamide gels of 12.5% acrylamide and 2.6% bisacrylamide were used. <strong>The</strong> resulting<br />

2-D gels were stained with colloidal CBB as described by Voigt et al. (Voigt et al., 2004).<br />

Protein identification<br />

Proteins were identified by MS. Protein spots were excised from stained gels with the Proteome<br />

Works Spot Cutter System (Bio-Rad). In-gel digestion with trypsin and the extraction of the peptides<br />

were done in the Ettan Spot HandlingWorkstation (Amersham Biosciences) according to a modified<br />

protocol of the manufacturer. Peptide masses were determined in the Proteomics Analyser 4700<br />

(Applied Biosystems). <strong>The</strong> 4700 Explorer Software V2.0 was used for spectrum calibration and<br />

analysis. After calibration, peak lists within a mass range of 900–3700 Da were created using the ’peak<br />

to mascot’ script of the 4700 Explorer Software. For the peak search, filters were set to a peak<br />

density of 20 peaks per 200 Da, a minimal peak area of 350 and maximal 60 peaks per spot. Peak lists<br />

were created with a signal to noise (S/N) ratio of 10 to 15. When necessary, TOF-TOF measurements<br />

for the three highest peaks in a spectrum were carried out. <strong>The</strong> internal calibration was automatically<br />

performed as one-point-calibration either with the mono- isotopic arginine (M+H) + peak, m/z at<br />

175.119, or with the lysine (M+H) + peak, m/z at 147.107, when the peaks reached at least an S/N ratio<br />

of 5. Peak lists were created as described above with the following settings: a mass range from 60 to<br />

precursor ion mass minus 20 Da, a peak density of five peaks per 200 Da, a minimal peak area of 100<br />

and maximal 20 peaks per precursor. Peak lists were created with an S/N ratio of 5. For database search<br />

the Mascot search engine (Matrix Science Ltd, London, UK) with a specific B. licheniformis sequence<br />

database was used. If this search did not result in the identification of the protein, PMF were reanalysed<br />

using the gpmaw software (Peri et al., 2001).<br />

152


<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media<br />

and under different nutrient starvation conditions<br />

Signal peptide prediction<br />

For prediction of signal peptides the SignalP 2.0 software (http://www.cbs.dtu.dk/services/SignalP-<br />

2.0/) (Nielsen et al., 1997) and the LipoP 1.0 software (http://www.cbs.dtu.dk/services/LipoP/)<br />

(Juncker et al., 2003) were used with the settings for Gram-positive organisms. Signal peptides were<br />

predicted by a Hidden Markov Model as well as by a neural networks method. Only proteins<br />

recognized by both methods to contain a signal peptide are classified as such. Proteins that are<br />

predicted to contain one or more membrane spanning domains in addition to an N-terminal signal<br />

peptide were excluded from the list (note that proteins with one membrane spanning domain in the Nterminal<br />

part of the protein might be wrongly identified as containing a signal peptide).<br />

Transmembrane helices were identified by the TMHMM v2.0 program (http://www.cbs.dtu.dk/<br />

services/TMHMM/). This program predicts transmembrane segments in proteins, using a Hidden<br />

Markov Model (Krogh et al., 2001).<br />

Denomination of the proteins<br />

Proteins with similarity to a B. subtilis protein were named accordingly. Proteins with no homolog in B.<br />

subtilis received the gene ID of the sequencing project.<br />

Dual channel imaging, quantification and colour coding<br />

<strong>The</strong> 2-D gel images were analyzed using the DECODON Delta2D software (Decodon GmbH<br />

Greifswald). Artificial coloring and warping of the gel images were done as described by Bernhardt et<br />

al. (Bernhardt et al., 1999). Spot quantities were also determined by the Delta2D software. <strong>The</strong><br />

numbers given represent the relative portion (% volume) of an individual spot of the total protein<br />

present on the gel. For comparison of the protein patterns during the different starvation conditions the<br />

“union fusion” approach of Delta2D for the generation of a proteome map was applied. For this<br />

purpose the gel images of the phosphate, glucose and nitrogen starvation were warped to the gel image<br />

of the control (exponential growth phase) and a fusion gel (union fusion) was created by combining the<br />

images. <strong>The</strong> spots were color coded according to their expression profile. Only spots induced more<br />

than 2.5-fold by the starvation conditions compared to the exponential phase were considered. Each<br />

spot subset, i.e., all spots induced in response to one starvation condition or a certain combination of<br />

starvation conditions, received a defined colour.<br />

Results and discussion<br />

Prediction of signal peptides<br />

<strong>The</strong> recently finished genome sequence of B. licheniformis allows the prediction of all<br />

proteins containing signals for secretion through one of the so-far-known protein secretion<br />

systems. From the B. licheniformis genome data there are 296 proteins predicted to have an<br />

N-terminal signal peptide for export from the cytoplasm. Most of these signal peptides would<br />

direct proteins to the Sec secretion machinery, but 19 have a twin-arginine motif (4 RR, 15<br />

KR) that may direct the corresponding precursor proteins into the Tat pathway, and 4 have the<br />

potential to direct proteins into a pseudopilin export pathway (Supplemental table Va;<br />

proteins with one or more membrane spanning domains in addition to the signal peptide were<br />

excluded). Of the 296 identified signal peptides, 220 have recognition sites for cleavage by a<br />

type I signal peptidase (e.g., SipS, SipT or SipV), 72 have recognition sites for cleavage by a<br />

lipoprotein-specific signal peptidase II (LspA), and 4 have recognition sites for a pseudopilin<br />

signal peptidase (ComC).<br />

153


Chapter 8<br />

<strong>The</strong> extracellular proteome of B. licheniformis and the comparison of the predicted to<br />

the real <strong>secretome</strong><br />

<strong>The</strong> highest level of protein secretion in B. licheniformis was observed in the stationary<br />

growth phase (Figs. 1, 2). Many of the degradative extracellular enzymes were secreted at low<br />

levels during the exponential growth phase and were induced in the stationary phase. <strong>The</strong><br />

induction of such enzymes in B. subtilis in the stationary growth phase is dependent on the<br />

DegSU two component system (Msadek et al., 1995; Ogura et al., 2001). As in B. subtilis,<br />

cells of B. licheniformis secrete much higher levels of protein when grown in a complex<br />

medium compared to cells grown in minimal medium (this is not apparent from our gels,<br />

because the same protein amount was loaded onto each gel).<br />

Figure 1. Dual channel images of the extracellular proteome of B. licheniformis. Cells were grown in LB and<br />

the images were created with the Delta 2D software (Decodon GmbH Greifswald). Proteins were prepared during<br />

exponential growth (OD 2), in the transition phase (OD 4) and in the stationary growth phase (OD 6). Proteins<br />

were separated in a pH gradient 3–10 and stained with colloidal CBB. Spots labeled in italics are presumably<br />

intracellular proteins; fr: fragment.<br />

From about 200 protein spots visible on the gels, altogether 143 proteins could be identified<br />

(Supplemental tables Vb and Vc). Some of the proteins, e.g., Vpr and YfnI, occur as multiple<br />

spots. Only 89 of the identified proteins were expected to be secreted because they either have<br />

a predicted N-terminal signal peptide sequence (79 proteins) or are known to be secreted by<br />

Sec independent pathways (10 proteins, Supplemental table Vb). <strong>The</strong>se extracellular proteins<br />

of B. licheniformis include carbohydrate degrading enzymes, several proteases and<br />

peptidases, enzymes involved in nucleic acid degradation, phosphodiesterases and<br />

phosphatases, enzymes involved in cell wall turnover, transport related proteins, flagellum-<br />

and phage-related proteins, proteins involved in sporulation and membrane bioenergetics and<br />

some proteins of yet unknown function.<br />

Of the 79 identified proteins containing an N-terminal signal peptide, 18 are involved in<br />

transport processes, among them several ABC transporter binding proteins. <strong>The</strong>se proteins<br />

would be expected to be lipid-anchored in the membrane. It is most likely that they are<br />

released from the membrane by “proteolytic shaving” (Antelmann et al., 2001).<br />

154


<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media<br />

and under different nutrient starvation conditions<br />

Figure 2. <strong>The</strong> extracellular proteome of B. licheniformis. Cells were grown in minimal medium under different<br />

starvation conditions. (A) Exponential growth. (B) Stationary phase phosphate starvation. (C) Stationary phase<br />

glucose starvation. (D) Stationary phase nitrogen starvation. Proteins were separated in a pH gradient 3–10 and<br />

stained with colloidal CBB. <strong>The</strong> graphs show the quantification of the ten most prominent spots in the gels given as<br />

% volume (representing the relative portion of an individual spot of the total protein present on the gel).<br />

Quantification was done with the Delta 2D software. Spots labelled in italics are presumably intracellular<br />

proteins; fr: fragment.<br />

155


Chapter 8<br />

<strong>The</strong>re were also some cell wall-related proteins present in the extracellular medium. <strong>The</strong>se<br />

cell wall proteins were present to a higher extent in the extracellular proteome of<br />

exponentially growing B. licheniformis cells (e.g., YvcE, YwtD; Supplemental table Vb). <strong>The</strong><br />

same was described for B. subtilis (Antelmann et al., 2002). In B. subtilis such proteins are<br />

retained in the cell wall because they contain specific wall-binding domains additionally to<br />

the signal peptide (Tjalsma et al., 2000). <strong>The</strong>y are probably released into the medium by the<br />

action of extracellular proteases as was shown by Antelmann et al. (Antelmann et al., 2002),<br />

since in multiple protease-deficient strains the cell wall proteins were stabilised.<br />

In the signal peptides of four secreted proteins found in this study, a potential twin-arginine<br />

signal peptide was identified (containing the canonical twin-arginine motif R-R-XH- H,<br />

where H is a hydrophobic amino acid, Supplemental table Vb) (Tjalsma et al., 2000;<br />

Jongbloed et al., 2000). One of these proteins is PhoD, one of the two proteins which have<br />

been actually shown to be translocated Tat dependently in B. subtilis (Jongbloed et al., 2000).<br />

Additional 15 proteins were predicted from the sequence to contain a K-R-X-H-H motif that<br />

may be functional in directing proteins to the Tat machinery (Tjalsma et al., 2004).<br />

<strong>The</strong> ten proteins known to be secreted by other pathways included seven flagellum-related<br />

proteins and three phage-related proteins, which typically lack signal peptides. <strong>The</strong> flagellum<br />

proteins are probably transported via a specific flagellum pathway. Hirose et al. (Hirose et al.,<br />

2000) found that Hag is not transported via the Sec pathway, because SecA depletion in a<br />

mutant strain did not affect Hag secretion. This was confirmed by studies of Jongbloed et al.<br />

(Jongbloed et al., 2002), who inhibited SecA activity by sodium azide and found that the<br />

secretion of the flagellum related proteins FliD and Hag was not affected. <strong>The</strong> proteins with<br />

phage-related functions may be secreted via prophage-encoded holins that can form<br />

membrane pores (Krogh et al., 1998).<br />

In addition, 55 proteins were identified in the extracellular proteome of B. licheniformis,<br />

which were not expected to be secreted because they have no predicted signal for secretion<br />

through one of the known systems, and which, taking into account their function, in most<br />

cases would be expected to have a cytoplasmic localisation (Supplemental table Vc). <strong>The</strong><br />

secretion of proteins without known export signals has been also described in other studies of<br />

the extracellular proteome (Antelmann et al., 2001; Hirose et al., 2000). It is not yet known<br />

whether these proteins are secreted via an unknown secretion mechanism, but more likely<br />

they have been released into the medium by cell lysis. Remarkably, the two type I signal<br />

peptidases SipS and SipT, which are membrane proteins with an N-terminal membrane<br />

anchor, could both be detected in the extracellular proteome of B. licheniformis cells grown in<br />

LB medium (Figure 1). <strong>The</strong> molecular mechanism underlying this unprecedented observation<br />

is presently not clear.<br />

Proteins secreted under all starvation conditions<br />

In response to phosphate, carbon and nitrogen starvation B. licheniformis cells secreted<br />

different proteins into the extracellular medium. <strong>The</strong>re were, however, also proteins that were<br />

secreted regardless of the growth conditions. Among these proteins was flagellin (Hag),<br />

which formed a prominent spot under all conditions tested. Furthermore, most of the<br />

proteases were secreted in response to all conditions, although not to the same level. <strong>The</strong> main<br />

protease secreted by B. licheniformis was Vpr. In B. subtilis the majority of the extracellular<br />

proteases are induced in the stationary growth phase (Antelmann et al., 2001). This induction<br />

is mainly dependent on the two component system DegSU and the regulator Hpr (Mäder et<br />

al., 2002; Antelmann et al., 2004).<br />

156


<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media<br />

and under different nutrient starvation conditions<br />

Phosphate starvation<br />

Surprisingly the main protein induced by phosphate starvation in B. licheniformis was the<br />

phytase (13.9% relative volume, Figure 2b, Tab. 1). <strong>The</strong> phytase hydrolyses phytate, the salt<br />

of phytic acid (myo-inositol 1,2,3,4,5,6-hexakis dihydrogen phosphate) (Tye et al., 2002).<br />

Phytate is the major storage form of phosphorus in plants and therefore present in soil, where<br />

it accounts for up to 50% of the organic phosphorus. B. licheniformis is able to grow on<br />

phytate as a sole phosphate source, and in this case phytase is already secreted during<br />

exponential growth (data not shown). In B. subtilis the phytase is expressed at a low level and<br />

the protein was not found in the extracellular proteome of phosphate-starving cells<br />

(Antelmann et al., 2004; Antelmann et al., 2000).<br />

Phosphate starvation also led to the specific secretion of the phosphatases PhoB (there is no<br />

homolog of the phoA gene encoded in the B. licheniformis genome) and PhoD. Secretion of<br />

phosphatases was also noticed in B. subtilis cells subjected to phosphate starvation, where<br />

they are among the main extracellular proteins. In B. licheniformis, however, these<br />

phosphatases belonged to the minor proteins secreted in response to phosphate starvation. In<br />

phosphate-starved B. licheniformis cells, YhdW, a protein that shows similarity to GlpQ, a<br />

glycerophosphodiester phosphodiesterase, was secreted to a high amount (GlpQ itself was<br />

only found in the proteome of cells grown in LB). Furthermore, there was a strong secretion<br />

of some proteins involved in the metabolism of nucleic acids (YfkN, YhcR, NucB and<br />

BLi03719). Induction of nucleic acid degrading enzymes during phosphate starvation has also<br />

been described for Corynebacterium glutamicum (Ishige et al., 2003). To be able to<br />

effectively take up the limiting amount of phosphate, B. licheniformis cells secreted PstS, the<br />

phosphate binding protein of an ABC transporter involved in the high-affinity phosphate<br />

uptake. In B. subtilis, PstS, probably attached to the membrane via a lipid anchor, is also an<br />

abundant protein in the <strong>secretome</strong> (Antelmann et al., 2000).<br />

Growth of B. subtilis under phosphate-starvation conditions results in the specific induction of<br />

genes belonging to the Pho regulon (Eymann et al., 1996; Prágai and Harwood, 2002). This<br />

regulon comprises several genes, which allow the cells to adapt to the limiting concentration<br />

of phosphate. <strong>The</strong> expression of these genes is under control of at least three two-component<br />

signal transduction systems (Hulett, 1996; Birkey et al., 1998). <strong>The</strong> main regulatory system is<br />

PhoPR consisting of the sensor kinase PhoR and the transcriptional activator PhoP<br />

(Antelmann et al., 2000; Liu and Hulett, 1998). Although there is not yet any experimental<br />

evidence, it can be suggested that the regulation in B. licheniformis is similar, because the<br />

genes for the two component system PhoPR are present in the genome.<br />

Glucose starvation<br />

We also investigated the extracellular protein pattern of B. licheniformis cells grown in a<br />

glucose limited medium (Figure 2c, Supplemental table Vb). <strong>The</strong> main protein secreted<br />

during glucose starvation was flagellin (Hag), which represents 10.3% (relative volume) of<br />

the total protein on the gel. To our surprise, we identified only two proteins involved in the<br />

degradation of carbohydrates in the extracellular proteome of glucose-starving cells (YvfO, an<br />

arabinogalactan endo-1,4-β-galactosidase, and YvpA, a pectate lyase), although many genes<br />

coding for such proteins have been identified in the genome. <strong>The</strong>re was, however, a strong<br />

secretion of some proteases (Vpr, Mpr and Bpr) and of the oligopeptide ABC transporter<br />

binding proteins AppA and OppA. Other proteases (Epr, Ggt, BLi01109, BLi01747) were<br />

157


Chapter 8<br />

secreted only at a low level. Beside this, some proteins of still unknown function (e.g., YusA,<br />

BLi02210, BLi01431) were strongly secreted.<br />

Figure 3. Colour-coded extracellular proteome map of B. licheniformis. Cells were grown in MM under<br />

phosphate, glucose and nitrogen starvation conditions. Proteome images of the three starvation conditions were<br />

fused with an exponential growth phase proteome image to create a proteome map, which contains all protein<br />

spots from the four individual images. Colour coding was done in such a way that all proteins belonging to a spot<br />

subset, i.e., all proteins induced in response to one starvation condition or a certain combination of starvation<br />

conditions received a defined, subset specific colour (colour scheme see upper left corner, P: phosphate<br />

starvation, G: glucose starvation, N: nitrogen starvation). Only proteins induced more than 2.5 fold compared to<br />

the exponential growth phase were included in the colour coding. Quantification, gel fusion and colour coding was<br />

done with the Delta 2D software.<br />

Perhaps B. licheniformis expresses some of the exoenzyme genes and secretes the<br />

corresponding hydrolyzing enzymes only when the carbon sources are present in the medium.<br />

Similar results were presented by Hirose et al. (Hirose et al., 2000) and Chu et al. (Chu et al.,<br />

2000) who found new proteins exported into the medium when they cultured Bacillus cells on<br />

different carbon sources, like cellobiose and xylan. Antelmann et al. (Antelmann et al., 2004)<br />

reported the secretion of SacB and Phy, enzymes not seen in earlier experiments, when B.<br />

subtilis was grown in a medium containing maltodextrin. Our results with cells grown in<br />

complex medium also point to such a mechanism, because these cells secrete a considerable<br />

158


<strong>The</strong> extracellular proteome of Bacillus licheniformis grown in different media<br />

and under different nutrient starvation conditions<br />

number of carbohydrate degrading exoenzymes (Supplemental table Vb, Figure 1), although<br />

not at a high level.<br />

Such a substrate induction mechanism is a known regulatory phenomenon. It has been<br />

described for instance for the usage of lichenan and levan by B. subtilis (Tobisch et al., 1997;<br />

Steinmetz et al., 1989). <strong>The</strong> genes for the usage of these carbohydrates are induced by the<br />

corresponding degradation products. Some other catabolic genes, like amyE and xynA,<br />

however, are not regulated by substrate induction but only by carbon catabolite repression<br />

(Stülke and Hillen, 2000). Carbon catabolite repression is the global regulatory mechanism to<br />

coordinate the carbon metabolism in Bacillus. In the presence of the preferred sugar, glucose,<br />

the regulator CcpA represses the catabolic genes whose expression is required for utilising<br />

alternative carbon sources. Most of the genes encoding extracellular carbohydrate degrading<br />

enzymes are under CcpA control in B. subtilis (Henkin et al., 1991; Martin-Verstraete et al.,<br />

1999; Yoshida et al., 2001).<br />

Nitrogen starvation<br />

<strong>The</strong> flagellin protein Hag was also the main protein secreted from nitrogen-starving cells<br />

(24.7% relative volume, Figure 2d). In addition, nitrogen starvation led to the secretion of<br />

some proteases and peptidases (Supplemental table Vb). <strong>The</strong> main protease, Vpr, accumulated<br />

during nitrogen starvation to the highest relative volume in all three conditions (14.5%<br />

compared to 7.1% in glucose starvation and 4.5% in phosphate starvation). This was also the<br />

case for some other proteases/ peptidases (e.g., Ggt, BLi01747, data not shown), although<br />

these proteins are not secreted at high levels. <strong>The</strong> aminopeptidase YwaD was secreted<br />

exclusively during nitrogen starvation conditions. In addition, the substrate binding proteins<br />

of several peptide transporters were found in the extracellular protein fraction of nitrogenstarving<br />

cells (DppE, OpuAC, BLi02811).<br />

<strong>The</strong> nitrogen metabolism of B. subtilis is controlled by different regulatory proteins in<br />

response to nutrient availability (Fisher, 1999). <strong>The</strong> regulatory proteins TnrA and GlnR<br />

control many genes that are expressed at high levels in nitrogen starving cells. TnrA induces<br />

the expression of genes involved in the usage of alternative nitrogen sources in nitrogenstarving<br />

cells. GlnR, on the other hand, represses such genes in cells growing with excess<br />

nitrogen. Although there are no data about the regulation of the nitrogen metabolism in B.<br />

licheniformis available, the regulation might be similar as in B. subtilis, because the genes<br />

tnrA and glnR are both present in the B. licheniformis genome.<br />

Comparison of the extracellular proteome under the different starvation conditions<br />

In Figure 3, an image fused from the images of the extracellular proteomes of the exponential<br />

growth phase and from all three starvation conditions is presented. To gain an overview over<br />

the differences in the proteins patterns, all proteins induced in response to the different<br />

starvation conditions more than 2.5 fold compared to the exponential growth phase were<br />

colour coded. Colour coding was done in such a way that all proteins belonging to a spot<br />

subset, i.e., all proteins induced by one starvation condition or a certain combination of<br />

starvation conditions, received a defined, subset specific colour. Seven protein subsets could<br />

be defined: (1) proteins induced only by phosphate starvation (red colour), (2) proteins<br />

induced only by glucose starvation (yellow), (3) proteins induced only by nitrogen starvation<br />

(blue), (4) proteins induced by phosphate as well as by glucose starvation (orange), (5)<br />

proteins induced by phosphate as well as by nitrogen starvation (green), (6) proteins induced<br />

159


Chapter 8<br />

by glucose as well as by nitrogen starvation (magenta), (7) proteins induced by all three<br />

conditions (cyan). This colour coding revealed the differences in the pattern of the<br />

extracellular proteins in response to phosphate, glucose and nitrogen starvation.<br />

Concluding remarks<br />

Prediction of N-terminal signal peptide sequences revealed 296 proteins in B. licheniformis,<br />

most of which are presumably Sec-dependently exported to the extracellular medium. This<br />

number is (almost) identical to the estimated number of 297 signal peptide-containing<br />

proteins of B. subtilis 168 (Tjalsma et al., 2004), which would be consistent with the fact that<br />

B. subtilis 168 and B. licheniformis DSM13 have about 4100 and 4200 protein-encoding<br />

genes, respectively. However, there appear to be some interesting differences with respect to<br />

the numbers of predicted secretory proteins and lipoproteins. While 179 predicted signal<br />

peptides of B. subtilis have a recognition site for a type I signal peptidase, and 114 for a type<br />

II signal peptidase, 220 predicted signal peptides of B. licheniformis have a recognition site<br />

for a type I signal peptidase and 72 for a type II signal peptidase. Thus, the B. licheniformis<br />

genome seems to encode a relatively higher number of secretory proteins, and a relatively<br />

lower number of lipoproteins, than the B. subtilis genome. <strong>The</strong> most effective technique to<br />

verify the localisation of proteins in certain cellular compartments is proteomics. Using this<br />

technique 89 proteins were identified in the extracellular proteome of B. licheniformis that<br />

either contain an N-terminal signal peptide or are known to be exported in a Sec-independent<br />

manner. <strong>The</strong>se proteins are involved in different physiological processes like cell wall<br />

turnover, mobility, transport of low molecular weight substrates and utilisation of alternative<br />

nutrient sources. <strong>The</strong> highest level of protein secretion was reached in the stationary growth<br />

phase in complex medium, when cells secreted a high number of degradative enzymes to<br />

make alternative nutrient sources accessible. Growth in defined medium under different<br />

nutrient starvation conditions led to the secretion of specific proteins, presumably involved in<br />

counteracting the starvation.<br />

Acknowledgements<br />

We thank Anne Krause for excellent technical assistance. We are grateful to the Decodon<br />

GmbH Greifswald for cooperation and the pre-release access to new software tools. This<br />

work received financial support from the Bundesministerium für Bildung und Forschung<br />

(0312707, 031U214B), the Henkel KGaA, the Bildungsministerium of the country<br />

Mecklenburg-Vorpommern, the Fonds of Chemical Industry of Germany and the EU (LSHC-<br />

CT- 2004–503468/005257).<br />

160


161


“I will choose free will”<br />

-Neil E. Peart-<br />

162


Chapter 9<br />

General summary and discussion<br />

163


Chapter 9<br />

General summary and discussion<br />

<strong>The</strong> human body functions as a host to many bacteria. Normally, we have no problem<br />

carrying all these bacteria, and some are even beneficial to us. One can think of the bacteria<br />

that are present in our colon that help us by synthesizing vitamins (Vitamine K, biotin), or<br />

bacteria that help to break down food components that we cannot digest, such as starch and<br />

various other fibers. Other bacteria are dangerous pathogens, causing diseases that can be<br />

lethal to human beings. Several of these primary pathogens cause well known diseases, like<br />

Clostridium tetani (tetanus), Corynebacterium diphtheria (diphtheria), Treponema pallidum<br />

(syphilis), Vibrio cholerae (cholera), Mycobacterium leprae (leprosy) and Mycobacterium<br />

tuberculosis (tuberculosis).<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> and <strong>Staphylococcus</strong> epidermdis are two of those bacteria that are<br />

living as commensals in the moist places and the skin of the human body. For healthy<br />

humans, S. <strong>aureus</strong> is not considered as a serious threat and ~20% of the population carries this<br />

organism in the nose. However, when the host defenses are breached and S. <strong>aureus</strong> is able to<br />

spread throughout the human body, this organism can turn into a dangerous pathogen. <strong>The</strong><br />

infective properties of S. <strong>aureus</strong> are caused by the arsenal of virulence factors that are<br />

expressed by this organism. As with all pathogens, these virulence factors are often<br />

proteinaceous compounds that are necessary for the bacterium to invade, colonize and spread<br />

throughout the (human) host and compounds that contribute to the symptoms of disease. All<br />

these virulence factors are synthesized in the cytoplasm of the cell and translocated across the<br />

bacterial membrane to an extracytoplasmic location, such as the cell surface or the host<br />

milieu. Well-characterized virulence factors include the many exotoxins of S. <strong>aureus</strong>.<br />

Most virulence factors are synthesized with an N-terminal signal peptide. This signal peptide<br />

is recognized by cytoplasmic chaperones that lead the protein to translocation pathways that<br />

are present in the cell. At the trans-side of the membrane, the translocated virulence factors<br />

are processed into their mature form and folded into the correct three-dimensional<br />

conformation with or without the help of other proteins. While several proteins are released<br />

into the environment (e.g. exotoxins), other proteins are retained at the surface of the cell. In<br />

Gram-positive bacteria, at least four extracytoplasmic cellular locations are distinguished: the<br />

membrane, the membrane-cell wall interface, the cell wall and the cell surface. All these<br />

locations can contain important virulence factors.<br />

Soon after the introduction of penicillin to eradicate the threat of S. <strong>aureus</strong> infections,<br />

resistant strains emerged and today the multidrug resistant S. <strong>aureus</strong> strains are imposing both<br />

therapeutic and financial challenges to our health care. <strong>The</strong> last resort antibiotic is<br />

vancomycin, but in the last 10-15 years S. <strong>aureus</strong> strains that are less susceptible or even fully<br />

resistant to vancomycin have been isolated. <strong>The</strong> need to search for alternatives to antibiotics<br />

such as vaccines and target-directed drugs is therefore increasing. <strong>The</strong>se alternative drugs or<br />

vaccines should be directed against a specific protein or process which decreases the chance<br />

of survival of the pathogenic bacteria in the human host without interfering with the host’s<br />

health.<br />

This thesis describes investigations on the <strong>secretome</strong> of S. <strong>aureus</strong>. By definition, the<br />

<strong>secretome</strong> includes all proteins that are translocated across the bacterial membrane to<br />

extracytoplasmic locations plus all the proteins that make up the protein translocation<br />

machinery.<br />

164


General summary and discussion<br />

Prediction of the <strong>secretome</strong> of Gram-positive bacteria<br />

In Chapter 2, the <strong>secretome</strong> of S. <strong>aureus</strong> is investigated with the help of bioinformatics. <strong>The</strong><br />

genomes of thirteen sequenced and annotated S. <strong>aureus</strong> strains were used to search for<br />

components of known translocation pathways, such as the general Sec pathway, the Twinarginine<br />

translocon, the Com pathway, ABC-transporters, holins and the Ess pathway. <strong>The</strong><br />

proteins that make up these pathways are all present in S. <strong>aureus</strong>, including proteins that are<br />

involved in the processing (e.g. signal peptidases, sortases) and folding (e.g. PrsA, DsbA)<br />

after translocation of these proteins. In addition, the deduced protein sequences were searched<br />

for the presence of N-terminal signal peptides. <strong>The</strong> search patterns developed for Bacillus<br />

subtilis were used to identify S. <strong>aureus</strong> signal peptides and the available proteomic data for<br />

extracellular proteins of S. <strong>aureus</strong> were used to further refine the respecitive staphylococcal<br />

consensus motifs. Five different signal peptides were classified according to the translocation<br />

pathway that is used and to the signal peptidase that processes the protein. In total, 145-168<br />

proteins, depending on the S. <strong>aureus</strong> strain investigated, were found that carry an N-terminal<br />

signal peptide for membrane translocation. This number is relatively high compared to other<br />

Gram-positive bacteria, such as Bacillus licheniformis (Chapter 10), B. subtilis, and S.<br />

epidermidis. In fact, B. subtilis has been claimed to be a secretion factory and various<br />

biotechnological industries have been using this organism for the production of commercially<br />

interesting enzymes. <strong>The</strong> high numbers of secreted proteins in S. <strong>aureus</strong> are partly related to<br />

the many exotoxins that are produced. From the proteins that have (predicted) signal peptides<br />

two groups were defined. <strong>The</strong> core exoproteome consists mainly of housekeeping proteins<br />

that seem to be necessary to maintain the position of S. <strong>aureus</strong> in its ecological niche in the<br />

(human) host. This is especially true for the proteins that are retained at the cell wall by<br />

sortase A. It is known that many of these proteins have adhesive properties and are involved<br />

in colonization of the host tissue. Some of these proteins (e.g. protein A) are also involved in<br />

protection of the staphylococcal cell against the immune system of the host. <strong>The</strong> proteins of<br />

the core exoproteome are interesting candidates for future vaccine development as is indicated<br />

in Chapter 3. <strong>The</strong> second group of proteins is defined as the variable exoproteome, which<br />

mainly consists of the exotoxins that are produced by S. <strong>aureus</strong>. <strong>The</strong>se proteins can be<br />

regarded as special S. <strong>aureus</strong> “gadgets” and the proteins with a yet unknown function should<br />

be regarded as potentially relevant virulence factors. Whether these proteins of unknown<br />

function are all important for virulence remains to be assessed. This is especially true for the<br />

lipoproteome, where many lipid-modified proteins (lipoproteins) have unknown functions. In<br />

recent years it has become clear that lipoproteins play very important roles in the process of<br />

infection.<br />

Using a proteomics approach, the exoproteomes of 25 clinical isolates that have been<br />

collected at the University Medical Center Groningen were analysed (Chapter 3). It is<br />

surprising that from the 69 core proteins only seven proteins were identified in all<br />

exoproteomes of these isolates. This difference can partly be explained by the retention of<br />

certain proteins in the cell wall, such as the covalently attached proteins (e.g. protein A,<br />

clumping factors, fibronectin and fibrinogen binding proteins, SasG) and proteins with other<br />

cell wall binding properties (e.g. LytN, SA0620, Atl). Other proteins are expressed only under<br />

specific growth conditions, such as iron-limiting conditions, oxygen stress, and other stress<br />

conditions, and these will therefore escape detection under the tested laboratory conditions.<br />

Nevertheless, these proteins are potential candidates for vaccine development. It will<br />

therefore be necessary to determine the exoproteomes of clinical S. <strong>aureus</strong> strains that are<br />

grown under infection mimicking conditions.<br />

165


Chapter 9<br />

Systematic analysis of translocation machineries of S. <strong>aureus</strong><br />

To investigate the importance of translocation pathways in the secretion of proteins, a set of<br />

isogenic S. <strong>aureus</strong> mutants were created and the exoproteomes of these mutants were<br />

analyzed. A summary of all mutants that were analyzed has been presented in Chapter 4, and<br />

a few of selected mutants were analyzed in more detail (Chapters 5, 6 and 7). To select<br />

proteins that are important in proteins translocation, the thirteen sequenced and annotated S.<br />

<strong>aureus</strong> genomes that are publicly available were scanned for these proteins. This in silico<br />

analysis of S. <strong>aureus</strong> genomes revealed that the components for the Sec pathway, Twinarginine<br />

pathway, Com pathway, ABC transporters for the translocation of bacteriocins,<br />

holins and Ess pathway are present in all strains. Differences were only detected for the Ess<br />

pathway in the S. <strong>aureus</strong> MRSA252 strain, where one of the substrates (EsxB) is not present<br />

and one other component (EsaC) is absent as well. This suggests either that EsaC is<br />

dispensable and that EsxB is dependent on the presence of EsaC, or that the Ess pathway may<br />

be not functional in this particular S. <strong>aureus</strong> strain. In this case, EsaC could still be required<br />

for the proper function of this pathway.<br />

Most proteins are translocated across the membrane via the general Sec pathway.<br />

Translocation via this pathway occurs in three distinguishable stages: chaperoning of the<br />

newly synthesized protein to the translocon, translocation, and modification and processing of<br />

the translocated proteins. In this thesis, several components that are involved in the latter two<br />

stages have been analyzed in more detail. In general, a translocon consists of a translocation<br />

motor and a channel. <strong>The</strong> translocation motor in the Sec pathway of S. <strong>aureus</strong> is the SecA<br />

protein that drives the translocation through repeated cycles of ATP binding and hydrolysis.<br />

<strong>The</strong> core channel is composed of the SecYEG proteins. Both SecA and SecY are known to be<br />

essential for growth and viability, and it was therefore decided not to attempt making the<br />

respective mutant strains. As has been found for other Gram-positive pathogens, the genome<br />

of S. <strong>aureus</strong> contains an additional set of secA and secY genes, encoding the SecA2 and<br />

SecY2 proteins. Usually, this accessory translocon is used only for a few selective proteins.<br />

Since the substrate and genes encoding glycosylation proteins are present in the same operon<br />

as the secA2 and secY2 genes it is conceivable that this accessory translocon only transports<br />

glycosylated proteins as was shown for Streptococcus gordonnii. In Chapter 5, the analyses<br />

of the isogenic secG and secY2 mutants were investigated. While the deletion of secY2 had no<br />

detectable effect on protein secretion, the importance of SecG became clearly evident upon<br />

analysis of the exoproteom of a secG mutant by 2-D PAGE. <strong>The</strong> extracellular accumulation<br />

of nine abundant exoproteins and seven cell wall-bound proteins was significantly affected in<br />

the secG mutant. Among these proteins are some known virulence factors, such as protein A,<br />

immunodominant antigen A, lipase and α-hemolysin. Interestingly, deletion of secY2<br />

exacerbated the secretion defects of secG mutants, affecting the extracellular accumulation of<br />

one additional exoprotein and one cell wall protein. Furthermore, the secG secY2 double<br />

mutant displayed a synthetic growth defect. <strong>The</strong>se findings suggest that SecY2 can interact<br />

with the main Sec channel of S. <strong>aureus</strong>. This was in fact already predicted on the basis of<br />

genome analyes, because all sequenced S. <strong>aureus</strong> strains have only a single set of the secE<br />

and secG genes in S. <strong>aureus</strong> (Chapter 2). Another interesting finding was that the second<br />

IgG-binding protein Sbi was almost completely absent from the cell wall of the secG mutant<br />

strain. Despite these interesting phenotypes of the secG mutant, it is questionable whether<br />

SecG might be a good target for novel antibiotics, since this protein is homologous to SecG of<br />

other bacteria and the Sec61β unit of the mammalian Sec61p complex. Moreover, infection<br />

experiments in a mouse model did not reveal any attenuation of the secG mutant strain,<br />

166


General summary and discussion<br />

suggesting that this component of the secretion machinery is dispensable for host subversion<br />

by S. <strong>aureus</strong>.<br />

Chapter 6 of this thesis deals with the modification and processing of certain translocated<br />

proteins by the class A sortase (SrtA) of S. <strong>aureus</strong> and S. epidermidis. SrtA attaches proteins<br />

with an LPxTG motif to the cell wall, by cleaving this motif between the conserved Thr and<br />

Gly residues and simultaneously attaching the Thr residue to peptidoglycan. In this study, the<br />

extracellular proteins of sortase A mutants from S. <strong>aureus</strong> and S. epidermidis were analyzed.<br />

Deletion of the srtA genes of these bacteria resulted in the dislocation of several LPxTG<br />

proteins, such as ClfA, SasG, SdrC, SdrD, and protein A of S. <strong>aureus</strong>, and the Aap protein of<br />

S. epidermidis 1457, from the cell wall to the growth medium. Nevertheless, substantial<br />

amounts of these proteins remained cell wall-bound through non-covalent interactions<br />

facilitated by cell wall-binding domains. <strong>The</strong> protein dislocation phenotypes of srtA mutations<br />

in S. <strong>aureus</strong> and S. epidermidis were fully reversed by ectopic expression of srtA genes of<br />

either species. Interestingly, the class C sortase (SrtC) from S. epidermidis 12228 was capable<br />

of reversing the dislocation of ClfA, SasG and Aap to significant extents, showing that the<br />

substrate specificities of SrtA and SrtC overlap at least partially. By contrast, SrtC was unable<br />

to restore the covalent cell wall attachment of protein A and the SdrC and SdrD proteins in<br />

the S. <strong>aureus</strong> srtA mutant. Interestingly, biofilm formation was affected in srtA mutants of S.<br />

<strong>aureus</strong>, but not in the S. epidermidis srtA mutant. This difference can be correlated to the<br />

expression levels of particular covalently cell wall-bound proteins involved in proteindependent<br />

biofilm formation, such as S. <strong>aureus</strong> SasG and its S. epidermidis homologue Aap.<br />

Remarkably, SrtA activity was a limiting determinant for protein-dependent biofilm<br />

formation in S. <strong>aureus</strong> and S. epidermidis, whereas SrtC expression interfered with biofilm<br />

formation in S. epidermidis 1457. Taken together, these findings imply that sortases can have<br />

modulating roles in staphylococcal biofilm formation.<br />

While the large mechano-sensitive membrane channel MscL can form pores that are large<br />

enough to allow the passage of small proteins across the membrane, no evidence was obtained<br />

that an S. <strong>aureus</strong> mscL mutant would be blocked in the release of certain proteins. In chapter<br />

7, an interesting phenotype of the mscL mutant strain is however described. This relates to the<br />

production of the extremely stable and broad-spectrum lantibiotic sublancin 168 by Bacillus<br />

subtilis strain 168. Known sublancin 168 susceptible organisms include important pathogens,<br />

such as S. <strong>aureus</strong>. Nevertheless, since its discovery, the mode of action of sublancin 168 has<br />

remained elusive. <strong>The</strong> studies in chapter 7 were, therefore, aimed at the identification of<br />

cellular determinants for bacterial susceptibility towards sublancin 168. Growth inhibition and<br />

competition assays on plates and in liquid cultures revealed that sublancin 168-mediated<br />

growth inhibition of susceptible B. subtilis and S. <strong>aureus</strong> cells is affected by the NaCl<br />

concentration in the growth medium. Added NaCl did not influence the production, activity or<br />

stability of sublancin 168 but, instead, lowered the susceptibility of sensitive cells towards<br />

this lantibiotic. Importantly, the susceptibility of B. subtilis and S. <strong>aureus</strong> cells towards<br />

sublancin 168 was shown to depend on the presence of MscL. <strong>The</strong>se findings demonstrate<br />

that MscL is a critical and specific determinant in bacterial sublancin 168 susceptibility that<br />

may either serve as a direct target for this lantibiotic, or as a gate of entry to the cytoplasm<br />

both in B. subtilis and S. <strong>aureus</strong>.<br />

<strong>The</strong> combination of bioinformatics and proteomics to investigate the <strong>secretome</strong> of S. <strong>aureus</strong><br />

has been proven to be a powerful approach for dissecting the functions of <strong>secretome</strong><br />

components. That this approach is useful not only for the analyses of pathogens, but also for<br />

industrial workhorses like Bacillus licheniformis is shown in chapter 8. From the genome<br />

167


Chapter 9<br />

sequence of B. licheniformis DSM 13, 296 proteins were predicted to contain an N-terminal<br />

signal peptide for secretion via the Sec system. Using 2-D PAGE, the extracellular proteome<br />

of B. licheniformis grown in different media was studied. From the approximately 200 spots<br />

visible on the gels, 89 were identified that either contain an N-terminal signal sequence or are<br />

known to be secreted by other mechanisms than the Sec pathway. <strong>The</strong> extracellular proteome<br />

of B. licheniformis was shown to include proteins from different functional classes, like<br />

enzymes for the degradation of macromolecules, proteins involved in cell wall turnover,<br />

flagellum- and phage-related proteins and some proteins of yet unknown function. Protein<br />

secretion was shown to be highest during the stationary growth phase. Furthermore, cells<br />

grown in a complex medium were found to secrete considerably higher protein amounts than<br />

cells grown in a minimal medium. Limitation of phosphate, carbon and nitrogen sources<br />

resulted in the secretion of specific proteins that may be involved in counteracting the<br />

potentially negative effects of the respective starvation.<br />

In conclusion, comparative secretomics approaches are applicable to the functional dissection<br />

of <strong>secretome</strong>s of bacterial pathogens, such as S. <strong>aureus</strong>, and biotechnologically relevant cell<br />

factories, such as B. subtilis and B. licheniformis.<br />

168


169


“Good music is good music and everything else can go to hell”<br />

-David J. Matthews-<br />

170


Chapter 10<br />

Reference list<br />

171


Chapter 10<br />

Adhikari, R.P., and Novick, R.P. (2005) Subinhibitory cerulenin inhibits staphylococcal exoprotein production<br />

by blocking transcription rather than by blocking secretion. Microbiology 151: 3059-3069.<br />

Alami, M., Lüke, I., Deitermann, S., Eisner, G. et al. (2003) Differential interactions between a twin-arginine<br />

signal peptide and its translocase in Escherichia coli. Mol Cell 12: 937-946.<br />

Antelmann, H., Sapolsky, R., Miller, B., Ferrari, E. et al. (2004) Quantitative proteome profiling during the<br />

fermentation process of pleiotropic Bacillus subtilis mutants. Proteomics 4: 2408-2424.<br />

Antelmann, H., Scharf, C., and Hecker, M. (2000) Phosphate starvation-inducible proteins of Bacillus subtilis:<br />

proteomics and transcriptional analysis. J Bacteriol 182: 4478-4490.<br />

Antelmann, H., Tjalsma, H., Voigt, B., Ohlmeier, S. et al. (2001) A proteomic view on genome-based signal<br />

peptide predictions. Genome Res 11: 1484-1502.<br />

Antelmann, H., Yamamoto, H., Sekiguchi, J., and Hecker, M. (2002) Stabilization of cell wall proteins in<br />

Bacillus subtilis: a proteomic approach. Proteomics 2: 591-602.<br />

Armand-Lefevre, L., Ruimy, R., and Andremont, A. (2005) Clonal comparison of <strong>Staphylococcus</strong> <strong>aureus</strong><br />

isolates from healthy pig farmers, human controls, and pigs. Emerg Infect Dis 11: 711-714.<br />

Arnaud, M., Chastanet, A., and Debarbouille, M. (2004) New Vector for Efficient Allelic Replacement in<br />

Naturally Nontransformable, Low-GC-Content, Gram-Positive Bacteria. Appl Environ Microbiol 70: 6887-<br />

6891.<br />

Arrecubieta, C., Matsunaga, I., Asai, T., Naka, Y. et al. (2008) Vaccination with clumping factor A and<br />

fibronectin binding protein A to prevent <strong>Staphylococcus</strong> <strong>aureus</strong> infection of an aortic patch in mice. J Infect<br />

Dis 198: 571-575.<br />

Arvidson, S., and Tegmark, K. (2001) Regulation of virulence determinants in <strong>Staphylococcus</strong> <strong>aureus</strong>. Int J Med<br />

Microbiol 291: 159-170.<br />

Aslanidis, C., and de Jong, P.J. (1990) Ligation-independent cloning of PCR products (LIC-PCR). Nucleic Acids<br />

Res 18: 6069-6074.<br />

Atkins, K.L., Burman, J.D., Chamberlain, E.S., Cooper, J.E. et al. (2008) S. <strong>aureus</strong> IgG-binding proteins SpA<br />

and Sbi: host specificity and mechanisms of immune complex formation. Mol Immunol 45: 1600-1611.<br />

Avison, M.B., Bennett, P.M., Howe, R.A., and Walsh, T.R. (2002) Preliminary analysis of the genetic basis for<br />

vancomycin resistance in <strong>Staphylococcus</strong> <strong>aureus</strong> strain Mu50. J Antimicrob Chemother 49: 255-260.<br />

Baba, T., Bae, T., Schneewind, O., Takeuchi, F. et al. (2008) Genome sequence of <strong>Staphylococcus</strong> <strong>aureus</strong> strain<br />

Newman and comparative analysis of staphylococcal genomes: polymorphism and evolution of two major<br />

pathogenicity islands. J Bacteriol 190: 300-310.<br />

Baba, T., and Schneewind, O. (1996) Target cell specificity of a bacteriocin molecule: a C-terminal signal directs<br />

lysostaphin to the cell wall of <strong>Staphylococcus</strong> <strong>aureus</strong>. EMBO J 15: 4789-4797.<br />

Baba, T., and Schneewind, O. (1998) Targeting of muralytic enzymes to the cell division site of Gram-positive<br />

bacteria: repeat domains direct autolysin to the equatorial surface ring of <strong>Staphylococcus</strong> <strong>aureus</strong>. EMBO J 17:<br />

4639-4646.<br />

Baba, T., Takeuchi, F., Kuroda, M., Yuzawa, H. et al. (2002) Genome and virulence determinants of high<br />

virulence community-acquired MRSA. Lancet 359: 1819-1827.<br />

Bae, T., Banger, A.K., Wallace, A., Glass, E.M. et al. (2004) <strong>Staphylococcus</strong> <strong>aureus</strong> virulence genes identified<br />

by Bursa aurealis mutagenesis and nematode killing. Proc Natl Acad Sci U S A 101: 12312-12317.<br />

Bae, T., and Schneewind, O. (2003) <strong>The</strong> YSIRK-G/S motif of staphylococcal protein A and its role in efficiency<br />

of signal peptide processing. J Bacteriol 185: 2910-2919.<br />

Baerends, R.J., Smits, W.K., de Jong, A., Hamoen, L.W. et al. (2004) Genome2D: a visualization tool for the<br />

rapid analysis of bacterial transcriptome data. Genome Biol 5: R37.<br />

Barnett, T.C., Patel, A.R., and Scott, J.R. (2004) A novel sortase, SrtC2, from Streptococcus pyogenes anchors a<br />

surface protein containing a QVPTGV motif to the cell wall. J Bacteriol 186: 5865-5875.<br />

Bateman, A., Holden, M.T., and Yeats, C. (2005) <strong>The</strong> G5 domain: a potential N-acetylglucosamine recognition<br />

domain involved in biofilm formation. Bioinformatics 21: 1301-1303.<br />

Baumgärtner, M., Kärst, U., Gerstel, B., Loessner, M. et al. (2007) Inactivation of Lgt allows systematic<br />

characterization of lipoproteins from Listeria monocytogenes. J Bacteriol 189: 313-324.<br />

Bayles, K.W. (2000) <strong>The</strong> bactericidal action of penicillin: new clues to an unsolved mystery. Trends Microbiol 8:<br />

274-278.<br />

Bendtsen, J.D., Nielsen, H., von Heijne, G., and Brunak, S. (2004) Improved prediction of signal peptides:<br />

SignalP 3.0. J Mol Biol 340: 783-795.<br />

Bens, C.C., Voss, A., and Klaassen, C.H. (2006) Presence of a novel DNA methylation enzyme in methicillinresistant<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> isolates associated with pig farming leads to uninterpretable results in standard<br />

pulsed-field gel electrophoresis analysis. J Clin Microbiol 44: 1875-1876.<br />

Bensing, B.A., and Sullam, P.M. (2002) An accessory sec locus of Streptococcus gordonii is required for export<br />

of the surface protein GspB and for normal levels of binding to human platelets. Mol Microbiol 44: 1081-1094.<br />

Bensing, B.A., Siboo, I.R., and Sullam, P.M. (2007) Glycine residues in the hydrophobic core of the GspB signal<br />

sequence route export toward the accessory Sec pathway. J Bacteriol 189: 3846-3854.<br />

Bentley, M.L., Gaweska, H., Kielec, J.M., and McCafferty, D.G. (2007) Engineering the substrate specificity of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> sortase A. <strong>The</strong> β6/β7 loop from SrtB confers NPQTN recognition to SrtA. J Biol Chem<br />

282: 6571-6581.<br />

Bergman, S., Selig, M., Collins, M.D., Farrow, J.A. et al. (1995) "Streptococcus milleri" strains displaying a<br />

gliding type of motility. Int J Syst Bacteriol 45: 235-239.<br />

172


Reference list<br />

Berks, B.C., Palmer, T., and Sargent, F. (2005) Protein targeting by the bacterial twin-arginine translocation<br />

(Tat) pathway. Curr Opin Microbiol 8: 174-181.<br />

Bernhardt, J., Büttner, K., Scharf, C., and Hecker, M. (1999) Dual channel imaging of two-dimensional<br />

electropherograms in Bacillus subtilis. Electrophoresis 20: 2225-2240.<br />

Berrier, C., Besnard, M., Ajouz, B., Coulombe, A. et al. (1996) Multiple mechanosensitive ion channels from<br />

Escherichia coli, activated at different thresholds of applied pressure. J Membr Biol 151: 175-187.<br />

Berrier, C., Coulombe, A., Szabo, I., Zoratti, M. et al. (1992) Gadolinium ion inhibits loss of metabolites<br />

induced by osmotic shock and large stretch-activated channels in bacteria. Eur J Biochem 206: 559-565.<br />

Berthet, F.X., Rasmussen, P.B., Rosenkrands, I., Andersen, P. et al. (1998) A Mycobacterium tuberculosis<br />

operon encoding ESAT-6 and a novel low-molecular-mass culture filtrate protein (CFP-10). Microbiology 144<br />

( Pt 11): 3195-3203.<br />

Bibel, D.J., Lovell, D.J., and Smiljanic, R.J. (1976) Effects of occlusion upon population dynamics of skin<br />

bacteria. Br J Dermatol 95: 607-612.<br />

Bibi, E. (1998) <strong>The</strong> role of the ribosome-translocon complex in translation and assembly of polytopic membrane<br />

proteins. Trends Biochem Sci 23: 51-55.<br />

Bierne, H., Mazmanian, S.K., Trost, M., Pucciarelli, M.G. et al. (2002) Inactivation of the srtA gene in Listeria<br />

monocytogenes inhibits anchoring of surface proteins and affects virulence. Mol Microbiol 43: 869-881.<br />

Birkey, S.M., Liu, W., Zhang, X., Duggan, M.F. et al. (1998) Pho signal transduction network reveals direct<br />

transcriptional regulation of one two-component system by another two-component regulator: Bacillus subtilis<br />

PhoP directly regulates production of ResD. Mol Microbiol 30: 943-953.<br />

Biswas, L., Biswas, R., Nerz, C., Ohlsen, K. et al. (2009) Role of the twin-arginine translocation pathway in<br />

<strong>Staphylococcus</strong>. J Bacteriol 191: 5921-5929.<br />

Blaudeck, N., Kreutzenbeck, P., Müller, M., Sprenger, G.A. et al. (2005) Isolation and characterization of<br />

bifunctional Escherichia coli TatA mutant proteins that allow efficient tat-dependent protein translocation in<br />

the absence of TatB. J Biol Chem 280: 3426-3432.<br />

Blevins, J.S., Beenken, K.E., Elasri, M.O., Hurlburt, B.K. et al. (2002) Strain-dependent differences in the<br />

regulatory roles of sarA and agr in <strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 70: 470-480.<br />

Blount, P., Sukharev, S.I., Moe, P.C., Martinac, B. et al. (1999) Mechanosensitive channels of bacteria.<br />

Methods Enzymol 294: 458-482.<br />

Boël, G., Pichereau, V., Mijakovic, I., Mazé, A. et al. (2004) Is 2-phosphoglycerate-dependent automodification<br />

of bacterial enolases implicated in their export? J Mol Biol 337: 485-496.<br />

Bolhuis, A., Broekhuizen, C.P., Sorokin, A., van Roosmalen, M.L. et al. (1998) SecDF of Bacillus subtilis, a<br />

molecular Siamese twin required for the efficient secretion of proteins. J Biol Chem 273: 21217-21224.<br />

Bolhuis, A., Matzen, A., Hyyryläinen, H.L., Kontinen, V.P. et al. (1999) Signal peptide peptidase- and ClpPlike<br />

proteins of Bacillus subtilis required for efficient translocation and processing of secretory proteins. J Biol<br />

Chem 274: 24585-24592.<br />

Bouma, B., de Groot, P.G., van den Elsen, J.M., Ravelli, R.B. et al. (1999) Adhesion mechanism of human β 2glycoprotein<br />

I to phospholipids based on its crystal structure. EMBO J 18: 5166-5174.<br />

Bowden, M.G., Visai, L., Longshaw, C.M., Holland, K.T. et al. (2002) Is the GehD lipase from <strong>Staphylococcus</strong><br />

epidermidis a collagen binding adhesin? J Biol Chem 277: 43017-43023.<br />

Braun, L., Dramsi, S., Dehoux, P., Bierne, H. et al. (1997) InlB: an invasion protein of Listeria monocytogenes<br />

with a novel type of surface association. Mol Microbiol 25: 285-294.<br />

Breukink, E., Wiedemann, I., van, K.C., Kuipers, O.P. et al. (1999) Use of the cell wall precursor lipid II by a<br />

pore-forming peptide antibiotic. Science 286: 2361-2364.<br />

Bron, S., and Venema, G. (1972) Ultraviolet inactivation and excision-repair in Bacillus subtilis. I. Construction<br />

and characterization of a transformable eightfold auxotrophic strain and two ultraviolet-sensitive derivatives.<br />

Mutat Res 15: 1-10.<br />

Brundage, L., Hendrick, J.P., Schiebel, E., Driessen, A.J. et al. (1990) <strong>The</strong> purified E. coli integral membrane<br />

protein SecY/E is sufficient for reconstitution of SecA-dependent precursor protein translocation. Cell 62: 649-<br />

657.<br />

Bruton, G., Huxley, A., O'Hanlon, P., Orlek, B. et al. (2003) Lipopeptide substrates for SpsB, the<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> type I signal peptidase: design, conformation and conversion to α-ketoamide inhibitors.<br />

Eur J Med Chem 38: 351-356.<br />

Bubeck, W.J., Williams, W.A., and Missiakas, D. (2006) Host defenses against <strong>Staphylococcus</strong> <strong>aureus</strong> infection<br />

require recognition of bacterial lipoproteins. Proc Natl Acad Sci U S A 103: 13831-13836.<br />

Bugg, T.D., Wright, G.D., Dutka-Malen, S., Arthur, M. et al. (1991) Molecular basis for vancomycin resistance<br />

in Enterococcus faecium BM4147: biosynthesis of a depsipeptide peptidoglycan precursor by vancomycin<br />

resistance proteins VanH and VanA. Biochemistry 30: 10408-10415.<br />

Buist, G., Steen, A., Kok, J., and Kuipers, O.P. (2008) LysM, a widely distributed protein motif for binding to<br />

(peptido)glycans. Mol Microbiol 68: 838-847.<br />

Bunai, K., Yamada, K., Hayashi, K., Nakamura, K. et al. (1999) Enhancing effect of Bacillus subtilis Ffh, a<br />

homologue of the SRP54 subunit of the mammalian signal recognition particle, on the binding of SecA to<br />

precursors of secretory proteins in vitro. J Biochem (Tokyo) 125: 151-159.<br />

Burman, J.D., Leung, E., Atkins, K.L., O'Seaghdha, M.N. et al. (2008) Interaction of human complement with<br />

Sbi, a staphylococcal immunoglobulin-binding protein: indications of a novel mechanism of complement<br />

evasion by <strong>Staphylococcus</strong> <strong>aureus</strong>. J Biol Chem 283: 17579-17593.<br />

173


Chapter 10<br />

Burts, M.L., Dent, A.C., and Missiakas, D.M. (2008) EsaC substrate for the ESAT-6 secretion pathway and its<br />

role in persistent infections of <strong>Staphylococcus</strong> <strong>aureus</strong>. Mol Microbiol 69: 736-746.<br />

Burts, M.L., Williams, W.A., DeBord, K., and Missiakas, D.M. (2005) EsxA and EsxB are secreted by an<br />

ESAT-6-like system that is required for the pathogenesis of <strong>Staphylococcus</strong> <strong>aureus</strong> infections. Proc Natl Acad<br />

Sci U S A 102: 1169-1174.<br />

Butcher, B.G., and Helmann, J.D. (2006) Identification of Bacillus subtilis σ-dependent genes that provide<br />

intrinsic resistance to antimicrobial compounds produced by Bacilli. Mol Microbiol 60: 765-782.<br />

Büttner, K., Bernhardt, J., Scharf, C., Schmid, R. et al. (2001) A comprehensive two-dimensional map of<br />

cytosolic proteins of Bacillus subtilis. Electrophoresis 22: 2908-2935.<br />

Cabelli, R.J., Chen, L., Tai, P.C., and Oliver, D.B. (1988) SecA protein is required for secretory protein<br />

translocation into E. coli membrane vesicles. Cell 55: 683-692.<br />

Candiano, G., Bruschi, M., Musante, L., Santucci, L. et al. (2004) Blue silver: a very sensitive colloidal<br />

Coomassie G-250 staining for proteome analysis. Electrophoresis 25: 1327-1333.<br />

Centers for Disease Control and Prevention (2003) Methicillin-Resistant <strong>Staphylococcus</strong> <strong>aureus</strong> Infections in<br />

Correctional Facilities - Georgia, California, and Texas, 2001–2003. MMWR 52: 992-995.<br />

Chambers, H.F. (1997) Methicillin resistance in staphylococci: molecular and biochemical basis and clinical<br />

implications. Clin Microbiol Rev 10: 781-791.<br />

Chan, P.F., and Foster, S.J. (1998) Role of SarA in virulence determinant production and environmental signal<br />

transduction in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 180: 6232-6241.<br />

Charpentier, E., Anton, A.I., Barry, P., Alfonso, B. et al. (2004) Novel cassette-based shuttle vector system for<br />

gram-positive bacteria. Appl Environ Microbiol 70: 6076-6085.<br />

Chatterjee, C., Paul, M., Xie, L., and van der Donk, W.A. (2005) Biosynthesis and mode of action of<br />

lantibiotics. Chem Rev 105: 633-684.<br />

Chatterjee, S., Chatterjee, S., Lad, S.J., Phansalkar, M.S. et al. (1992) Mersacidin, a new antibiotic from<br />

Bacillus. Fermentation, isolation, purification and chemical characterization. J Antibiot (Tokyo) 45: 832-838.<br />

Chaudhuri, R.R., Allen, A.G., Owen, P.J., Shalom, G. et al. (2009) Comprehensive identification of essential<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> genes using Transposon-Mediated Differential Hybridisation (TMDH). BMC Genomics<br />

10: 291.<br />

Cheng, A.G., Kim, H.K., Burts, M.L., Krausz, T. et al. (2009) Genetic requirements for <strong>Staphylococcus</strong> <strong>aureus</strong><br />

abscess formation and persistence in host tissues. FASEB J 23: 3393-3404.<br />

Cheung, A.L., Koomey, J.M., Butler, C.A., Projan, S.J. et al. (1992) Regulation of exoprotein expression in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> by a locus (sar) distinct from agr. Proc Natl Acad Sci U S A 89: 6462-6466.<br />

Cheung, A.L., and Projan, S.J. (1994) Cloning and sequencing of sarA of <strong>Staphylococcus</strong> <strong>aureus</strong>, a gene<br />

required for the expression of agr. J Bacteriol 176: 4168-4172.<br />

Christensen, G.D., Simpson, W.A., Younger, J.J., Baddour, L.M. et al. (1985) Adherence of coagulasenegative<br />

staphylococci to plastic tissue culture plates: a quantitative model for the adherence of staphylococci<br />

to medical devices. J Clin Microbiol 22: 996-1006.<br />

Chu, P.W., Yap, M.N., Wu, C.Y., Huang, C.M. et al. (2000) A proteomic analysis of secreted proteins from<br />

xylan-induced Bacillus sp. strain K-1. Electrophoresis 21: 1740-1745.<br />

Chung, Y.S., Breidt, F., and Dubnau, D. (1998) Cell surface localization and processing of the ComG proteins,<br />

required for DNA binding during transformation of Bacillus subtilis. Mol Microbiol 29: 905-913.<br />

Chung, Y.S., and Dubnau, D. (1995) ComC is required for the processing and translocation of comGC, a pilinlike<br />

competence protein of Bacillus subtilis. Mol Microbiol 15: 543-551.<br />

Chung, Y.S., and Dubnau, D. (1998) All seven comG open reading frames are required for DNA binding during<br />

transformation of competent Bacillus subtilis. J Bacteriol 180: 41-45.<br />

Claessen, D., Rink, R., de, J.W., Siebring, J. et al. (2003) A novel class of secreted hydrophobic proteins is<br />

involved in aerial hyphae formation in Streptomyces coelicolor by forming amyloid-like fibrils. Genes Dev 17:<br />

1714-1726.<br />

Clarke, S.R., Brummell, K.J., Horsburgh, M.J., McDowell, P.W. et al. (2006) Identification of in vivoexpressed<br />

antigens of <strong>Staphylococcus</strong> <strong>aureus</strong> and their use in vaccinations for protection against nasal carriage.<br />

J Infect Dis 193: 1098-1108.<br />

Clarke, S.R., Harris, L.G., Richards, R.G., and Foster, S.J. (2002) Analysis of Ebh, a 1.1-megadalton cell wallassociated<br />

fibronectin-binding protein of <strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 70: 6680-6687.<br />

Cline, K., and Mori, H. (2001) Thylakoid ∆pH-dependent precursor proteins bind to a cpTatC-Hcf106 complex<br />

before Tha4-dependent transport. J Cell Biol 154: 719-729.<br />

Comfort, D. and Clubb, R.T. (2004) A comparitive genome analysis identifies distinct sorting pathways in<br />

Gram-positive bacteria. Infect Immun 72: 2710-2722.<br />

Conrady, D.G., Brescia, C.C., Horii, K., Weiss, A.A. et al. (2008) A zinc-dependent adhesion module is<br />

responsible for intercellular adhesion in staphylococcal biofilms. Proc Natl Acad Sci U S A 105: 19456-19461.<br />

Converse, S.E., and Cox, J.S. (2005) A protein secretion pathway critical for Mycobacterium tuberculosis<br />

virulence is conserved and functional in Mycobacterium smegmatis. J Bacteriol 187: 1238-1245.<br />

Corrigan, R.M., Miajlovic, H., and Foster, T.J. (2009) Surface proteins that promote adherence of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> to human desquamated nasal epithelial cells. BMC Microbiol 9: 22.<br />

Corrigan, R.M., Rigby, D., Handley, P., and Foster, T.J. (2007) <strong>The</strong> role of <strong>Staphylococcus</strong> <strong>aureus</strong> surface<br />

protein SasG in adherence and biofilm formation. Microbiology 153: 2435-2446.<br />

174


Reference list<br />

Cregg, K.M., Wilding, I., and Black, M.T. (1996) Molecular cloning and expression of the spsB gene encoding<br />

an essential type I signal peptidase from <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 178: 5712-5718.<br />

Cristóbal, S., de Gier, J.W., Nielsen, H., and von Heijne, G. (1999) Competition between Sec- and TATdependent<br />

protein translocation in Escherichia coli. EMBO J 18: 2982-2990.<br />

Crupper, S.S., Gies, A.J., and Iandolo, J.J. (1997) Purification and characterization of staphylococcin BacR1, a<br />

broad-spectrum bacteriocin. Appl Environ Microbiol 63: 4185-4190.<br />

Cserzö, M., Wallin, E., Simon, I., von Heijne, G. et al. (1997) Prediction of transmembrane alpha-helices in<br />

prokaryotic membrane proteins: the dense alignment surface method. Protein Eng 10: 673-676.<br />

Cui, L., Ma, X., Sato, K., Okuma, K. et al. (2003) Cell wall thickening is a common feature of vancomycin<br />

resistance in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Clin Microbiol 41: 5-14.<br />

da Silva, M.C., Zahm, J.M., Gras, D., Bajolet, O. et al. (2004) Dynamic interaction between airway epithelial<br />

cells and <strong>Staphylococcus</strong> <strong>aureus</strong>. Am J Physiol Lung Cell Mol Physiol 287: L543-L551.<br />

Dalbey, R.E., Lively, M.O., Bron, S., and van Dijl, J.M. (1997) <strong>The</strong> chemistry and enzymology of the type I<br />

signal peptidases. Protein Sci 6: 1129-1138.<br />

Darmon, E., Dorenbos, R., Meens, J., Freudl, R. et al. (2006) A disulfide bond-containing alkaline phosphatase<br />

triggers a BdbC-dependent secretion stress response in Bacillus subtilis. Appl Environ Microbiol 72: 6876-<br />

6885.<br />

de Jong, A., van Hijum, S.A., Bijlsma, J.J., Kok, J. et al. (2006) BAGEL: a web-based bacteriocin genome<br />

mining tool. Nucleic Acids Res 34: W273-W279.<br />

de Keyzer, J., van der Does, C., and Driessen, A.J. (2003) <strong>The</strong> bacterial translocase: a dynamic protein channel<br />

complex. Cell Mol Life Sci 60: 2034-2052.<br />

De Leeuw, E., te Kaat, K., Moser, C., Menestrina, G. et al. (2000) Anionic phospholipids are involved in<br />

membrane association of FtsY and stimulate its GTPase activity. EMBO J 19: 531-541.<br />

Dedent, A., Bae, T., Missiakas, D.M., and Schneewind, O. (2008) Signal peptides direct surface proteins to two<br />

distinct envelope locations of <strong>Staphylococcus</strong> <strong>aureus</strong>. EMBO J.<br />

Diep, B.A., Gill, S.R., Chang, R.F., Phan, T.H. et al. (2006) Complete genome sequence of USA300, an<br />

epidemic clone of community-acquired meticillin-resistant <strong>Staphylococcus</strong> <strong>aureus</strong>. Lancet 367: 731-739.<br />

Dilks, K., Rose, R.W., Hartmann, E., and Pohlschröder, M. (2003) Prokaryotic utilization of the twin-arginine<br />

translocation pathway: a genomic survey. J Bacteriol 185: 1478-1483.<br />

Dobrindt, U., Hochhut, B., Hentschel, U., and Hacker, J. (2004) Genomic islands in pathogenic and<br />

environmental microorganisms. Nat Rev Microbiol 2: 414-424.<br />

Dorenbos, R., Stein, T., Kabel, J., Bruand, C. et al. (2002) Thiol-disulfide oxidoreductases are essential for the<br />

production of the lantibiotic sublancin 168. J Biol Chem 277: 16682-16688.<br />

Dramsi, S., Trieu-Cuot, P., and Bierne, H. (2005) Sorting sortases: a nomenclature proposal for the various<br />

sortases of Gram-positive bacteria. Res Microbiol 156: 289-297.<br />

Driessen, A.J., Fekkes, P., and van der Wolk, J.P. (1998) <strong>The</strong> Sec system. Curr Opin Microbiol 1: 216-222.<br />

Driessen, A.J., and Nouwen, N. (2008) Protein translocation across the bacterial cytoplasmic membrane. Annu<br />

Rev Biochem 77: 643-667.<br />

Dubnau, D. (1999) DNA uptake in bacteria. Annu Rev Microbiol 53: 217-244.<br />

Dubois, J.Y., Kouwen, T.R., Schurich, A.K., Reis, C.R. et al. (2009) Immunity to the bacteriocin sublancin 168<br />

Is determined by the SunI (YolF) protein of Bacillus subtilis. Antimicrob Agents Chemother 53: 651-661.<br />

Dubrac, S., Bisicchia, P., Devine, K.M., and Msadek, T. (2008) A matter of life and death: cell wall homeostasis<br />

and the WalKR (YycGF) essential signal transduction pathway. Mol Microbiol 70: 1307-1322.<br />

Dubrac, S., Boneca, I.G., Poupel, O., and Msadek, T. (2007) New insights into the WalK/WalR (YycG/YycF)<br />

essential signal transduction pathway reveal a major role in controlling cell wall metabolism and biofilm<br />

formation in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 189: 8257-8269.<br />

Dumoulin, A., Grauschopf, U., Bischoff, M., Thöny-Meyer, L. et al. (2005) <strong>Staphylococcus</strong> <strong>aureus</strong> DsbA is a<br />

membrane-bound lipoprotein with thiol-disulfide oxidoreductase activity. Arch Microbiol 184: 117-128.<br />

Duthie, E.S., and Lorenz, L.L. (1952) Staphylococcal coagulase; mode of action and antigenicity. J Gen<br />

Microbiol 6: 95-107.<br />

Dziewanowska, K., Edwards, V.M., Deringer, J.R., Bohach, G.A. et al. (1996) Comparison of the β-toxins<br />

from <strong>Staphylococcus</strong> <strong>aureus</strong> and <strong>Staphylococcus</strong> intermedius. Arch Biochem Biophys 335: 102-108.<br />

Eisner, G., Koch, H.G., Beck, K., Brunner, J. et al. (2003) Ligand crowding at a nascent signal sequence. J Cell<br />

Biol 163: 35-44.<br />

Elliot, M.A., Karoonuthaisiri, N., Huang, J., Bibb, M.J. et al. (2003) <strong>The</strong> chaplins: a family of hydrophobic<br />

cell-surface proteins involved in aerial mycelium formation in Streptomyces coelicolor. Genes Dev 17: 1727-<br />

1740.<br />

Enright, M.C., Day, N.P., Davies, C.E., Peacock, S.J. et al. (2000) Multilocus sequence typing for<br />

characterization of methicillin-resistant and methicillin-susceptible clones of <strong>Staphylococcus</strong> <strong>aureus</strong>. J Clin<br />

Microbiol 38: 1008-1015.<br />

Escher, A., and Characklis, W.G. (1990) Modeling the initial events in biofilm accumulation. In Biofilms.<br />

Characklis,W.G. (ed). New York, NY, USA: Wiley Press, pp. 445-486.<br />

Eymann, C., Dreisbach, A., Albrecht, D., Bernhardt, J. et al. (2004) A comprehensive proteome map of<br />

growing Bacillus subtilis cells. Proteomics 4: 2849-2876.<br />

Eymann, C., Mach, H., Harwood, C.R., and Hecker, M. (1996) Phosphate-starvation-inducible proteins in<br />

Bacillus subtilis: a two-dimensional gel electrophoresis study. Microbiology 142 ( Pt 11): 3163-3170.<br />

175


Chapter 10<br />

Fedtke, I., Götz, F., and Peschel, A. (2004) Bacterial evasion of innate host defenses--the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

lesson. Int J Med Microbiol 294: 189-194.<br />

Fisher, S.H. (1999) Regulation of nitrogen metabolism in Bacillus subtilis: vive la difference! Mol Microbiol 32:<br />

223-232.<br />

Foster, T.J., and Hook, M. (1998) Surface protein adhesins of <strong>Staphylococcus</strong> <strong>aureus</strong>. Trends Microbiol 6: 484-<br />

488.<br />

Frankel, B.A., Bentley, M., Kruger, R.G., and McCafferty, D.G. (2004) Vinyl sulfones: inhibitors of SrtA, a<br />

transpeptidase required for cell wall protein anchoring and virulence in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Am Chem<br />

Soc 126: 3404-3405.<br />

Frankel, B.A., Tong, Y., Bentley, M.L., Fitzgerald, M.C. et al. (2007) Mutational analysis of active site residues<br />

in the <strong>Staphylococcus</strong> <strong>aureus</strong> transpeptidase SrtA. Biochemistry 46: 7269-7278.<br />

Fredenhagen, A., Fendrich, G., Marki, F., Marki, W. et al. (1990) Duramycins B and C, two new lanthionine<br />

containing antibiotics as inhibitors of phospholipase A2. Structural revision of duramycin and cinnamycin. J<br />

Antibiot (Tokyo) 43: 1403-1412.<br />

Fuchs, S., Pane-Farre, J., Kohler, C., Hecker, M. et al. (2007) Anaerobic gene expression in <strong>Staphylococcus</strong><br />

<strong>aureus</strong>. J Bacteriol 189: 4275-4289.<br />

Garandeau, C., Réglier-Poupet, H., Dubail, I., Beretti, J.L. et al. (2002) <strong>The</strong> sortase SrtA of Listeria<br />

monocytogenes is involved in processing of internalin and in virulence. Infect Immun 70: 1382-1390.<br />

Garcia, P., Garcia, J.L., Garcia, E., Sanchez-Puelles, J.M. et al. (1990) Modular organization of the lytic<br />

enzymes of Streptococcus pneumoniae and its bacteriophages. Gene 86: 81-88.<br />

García-Lara, J., Needham, A.J., and Foster, S.J. (2005) Invertebrates as animal models for <strong>Staphylococcus</strong><br />

<strong>aureus</strong> pathogenesis: a window into host-pathogen interaction. FEMS Immunol Med Microbiol 43: 311-323.<br />

Gaspar, A.H., and Ton-That, H. (2006) Assembly of distinct pilus structures on the surface of Corynebacterium<br />

diphtheriae. J Bacteriol 188: 1526-1533.<br />

Gatlin, C.L., Pieper, R., Huang, S.T., Mongodin, E. et al. (2006) Proteomic profiling of cell envelope-associated<br />

proteins from <strong>Staphylococcus</strong> <strong>aureus</strong>. Proteomics.<br />

Gemmell, C.G. (1995) Staphylococcal scalded skin syndrome. J Med Microbiol 43: 318-327.<br />

Gertz, S., Engelmann, S., Schmid, R., Ohlsen, K. et al. (1999) Regulation of σ B -dependent transcription of sigB<br />

and asp23 in two different <strong>Staphylococcus</strong> <strong>aureus</strong> strains. Mol Gen Genet 261: 558-566.<br />

Ghuysen, J.M., Lamotte-Brasseur, J., Joris, B., and Shockman, G.D. (1994) Binding site-shaped repeated<br />

sequences of bacterial wall peptidoglycan hydrolases. FEBS Lett 342: 23-28.<br />

Gill, S.R., Fouts, D.E., Archer, G.L., Mongodin, E.F. et al. (2005) Insights on evolution of virulence and<br />

resistance from the complete genome analysis of an early methicillin-resistant <strong>Staphylococcus</strong> <strong>aureus</strong> strain<br />

and a biofilm-producing methicillin-resistant <strong>Staphylococcus</strong> epidermidis strain. J Bacteriol 187: 2426-2438.<br />

Gillaspy, A.F., Worrell, V., Orvis, J., Roe, B.A. et al. (2006) <strong>The</strong> <strong>Staphylococcus</strong> <strong>aureus</strong> NCTC8325 Genome.<br />

In Gram-positive Pathogens. Fischetti,V., Novick,R., Ferretti,J., Portnoy,D., and Rood,J. (eds). Washington<br />

DC: American Society for Microbiology, pp. 381-412.<br />

Glowalla, E., Tosetti, B., Krönke, M., and Krut, O. (2009) Proteomics-based identification of anchorless cell<br />

wall proteins as vaccine candidates against <strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 77: 2719-2729.<br />

Goerke, C., Papenberg, S., Dasbach, S., Dietz, K. et al. (2004) Increased frequency of genomic alterations in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> during chronic infection is in part due to phage mobilization. J Infect Dis 189: 724-734.<br />

Goerke, C., Wirtz, C., Fluckiger, U., and Wolz, C. (2006) Extensive phage dynamics in <strong>Staphylococcus</strong> <strong>aureus</strong><br />

contributes to adaptation to the human host during infection. Mol Microbiol 61: 1673-1685.<br />

Gohar, M., Gilois, N., Graveline, R., Garreau, C. et al. (2005) A comparative study of Bacillus cereus, Bacillus<br />

thuringiensis and Bacillus anthracis extracellular proteomes. Proteomics 5: 3696-3711.<br />

Gohar, M., Økstad, O.A., Gilois, N., Sanchis, V. et al. (2002) Two-dimensional electrophoresis analysis of the<br />

extracellular proteome of Bacillus cereus reveals the importance of the PlcR regulon. Proteomics 2: 784-791.<br />

Götz, F. (2002) <strong>Staphylococcus</strong> and biofilms. Mol Microbiol 43: 1367-1378.<br />

Götz, F. (2004) Staphylococci in colonization and disease: prospective targets for drugs and vaccines. Curr Opin<br />

Microbiol 7: 477-487.<br />

Groicher, K.H., Firek, B.A., Fujimoto, D.F., and Bayles, K.W. (2000) <strong>The</strong> <strong>Staphylococcus</strong> <strong>aureus</strong> lrgAB<br />

operon modulates murein hydrolase activity and penicillin tolerance. J Bacteriol 182: 1794-1801.<br />

Grundmann, H., Tami, A., Hori, S., Halwani, M. et al. (2002) Nottingham <strong>Staphylococcus</strong> <strong>aureus</strong> population<br />

study: prevalence of MRSA among elderly people in the community. BMJ 324: 1365-1366.<br />

Grundmeier, M., Hussain, M., Becker, P., Heilmann, C. et al. (2004) Truncation of fibronectin-binding proteins<br />

in <strong>Staphylococcus</strong> <strong>aureus</strong> strain Newman leads to deficient adherence and host cell invasion due to loss of the<br />

cell wall anchor function. Infect Immun 72: 7155-7163.<br />

Guérout-Fleury, A.M., Shazand, K., Frandsen, N., and Stragier, P. (1995) Antibiotic-resistance cassettes for<br />

Bacillus subtilis. Gene 167: 335-336.<br />

Hahn, J., Maier, B., Haijema, B.J., Sheetz, M. et al. (2005) Transformation proteins and DNA uptake localize to<br />

the cell poles in Bacillus subtilis. Cell 122: 59-71.<br />

Hanada, M., Nishiyama, K., and Tokuda, H. (1996) SecG plays a critical role in protein translocation in the<br />

absence of the proton motive force as well as at low temperature. FEBS Lett 381: 25-28.<br />

Hanahan, D. (1983) Studies on transformation of Escherichia coli with plasmids. J Mol Biol 166: 557-580.<br />

Harraghy, N., Hussain, M., Haggar, A., Chavakis, T. et al. (2003) <strong>The</strong> adhesive and immunomodulating<br />

properties of the multifunctional <strong>Staphylococcus</strong> <strong>aureus</strong> protein Eap. Microbiology 149: 2701-2707.<br />

176


Reference list<br />

Hartford, O., Francois, P., Vaudaux, P., and Foster, T.J. (1997) <strong>The</strong> dipeptide repeat region of the fibrinogenbinding<br />

protein (clumping factor) is required for functional expression of the fibrinogen-binding domain on the<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> cell surface. Mol Microbiol 25: 1065-1076.<br />

Hashimoto, M., Tawaratsumida, K., Kariya, H., Kiyohara, A. et al. (2006) Not lipoteichoic acid but<br />

lipoproteins appear to be the dominant immunobiologically active compounds in <strong>Staphylococcus</strong> <strong>aureus</strong>. J<br />

Immunol 177: 3162-3169.<br />

Hasona, A., Crowley, P.J., Levesque, C.M., Mair, R.W. et al. (2005) Streptococcal viability and diminished<br />

stress tolerance in mutants lacking the signal recognition particle pathway or YidC2. Proc Natl Acad Sci U S A<br />

102: 17466-17471.<br />

Hauck, C.R., and Ohlsen, K. (2006) Sticky connections: extracellular matrix protein recognition and integrinmediated<br />

cellular invasion by <strong>Staphylococcus</strong> <strong>aureus</strong>. Curr Opin Microbiol 9: 5-11.<br />

Hava, D.L., and Camilli, A. (2002) Large-scale identification of serotype 4 Streptococcus pneumoniae virulence<br />

factors. Mol Microbiol 45: 1389-1406.<br />

Hava, D.L., Hemsley, C.J., and Camilli, A. (2003) Transcriptional regulation in the Streptococcus pneumoniae<br />

rlrA pathogenicity islet by RlrA. J Bacteriol 185: 413-421.<br />

Håvarstein, L.S., Diep, D.B., and Nes, I.F. (1995) A family of bacteriocin ABC transporters carry out proteolytic<br />

processing of their substrates concomitant with export. Mol Microbiol 16: 229-240.<br />

Heikkinen, O., Seppala, R., Tossavainen, H., Heikkinen, S. et al. (2009) Solution structure of the parvulin-type<br />

PPIase domain of <strong>Staphylococcus</strong> <strong>aureus</strong> PrsA--implications for the catalytic mechanism of parvulins. BMC<br />

Struct Biol 9: 17.<br />

Heilmann, C., Hartleib, J., Hussain, M.S., and Peters, G. (2005) <strong>The</strong> multifunctional <strong>Staphylococcus</strong> <strong>aureus</strong><br />

autolysin Aaa mediates adherence to immobilized fibrinogen and fibronectin. Infect Immun 73: 4793-4802.<br />

Heilmann, C., Hussain, M., Peters, G., and Götz, F. (1997) Evidence for autolysin-mediated primary attachment<br />

of <strong>Staphylococcus</strong> epidermidis to a polystyrene surface. Mol Microbiol 24: 1013-1024.<br />

Heilmann, C., Thumm, G., Chhatwal, G.S., Hartleib, J. et al. (2003) Identification and characterization of a<br />

novel autolysin (Aae) with adhesive properties from <strong>Staphylococcus</strong> epidermidis. Microbiology 149: 2769-<br />

2778.<br />

Hemphill, H.E., Gage, I., Zahler, S.A., and Korman, R.Z. (1980) Prophage-mediated production of a<br />

bacteriocinlike substance by SPβ lysogens of Bacillus subtilis. Can J Microbiol 26: 1328-1333.<br />

Henkin, T.M., Grundy, F.J., Nicholson, W.L., and Chambliss, G.H. (1991) Catabolite repression of α-amylase<br />

gene expression in Bacillus subtilis involves a trans-acting gene product homologous to the Escherichia coli<br />

lacl and galR repressors. Mol Microbiol 5: 575-584.<br />

Henrichsen, J. (1972) Bacterial surface translocation: a survey and a classification. Bacteriol Rev 36: 478-503.<br />

Henrichsen, J. (1997) "Streptococcus milleri" strains exhibit not gliding motility but sliding. Int J Syst Bacteriol<br />

47: 604.<br />

Heras, B., Kurz, M., Jarrott, R., Shouldice, S.R. et al. (2008) <strong>Staphylococcus</strong> <strong>aureus</strong> DsbA does not have a<br />

destabilizing disulfide. A new paradigm for bacterial oxidative folding. J Biol Chem 283: 4261-4271.<br />

Herbort, M., Klein, M., Manting, E.H., Driessen, A.J. et al. (1999) Temporal expression of the Bacillus subtilis<br />

secA gene, encoding a central component of the preprotein translocase. J Bacteriol 181: 493-500.<br />

Herron, L.L., Chakravarty, R., Dwan, C., Fitzgerald, J.R. et al. (2002) Genome sequence survey identifies<br />

unique sequences and key virulence genes with unusual rates of amino acid substitution in bovine<br />

<strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 70: 3978-3981.<br />

Hiller, K., Grote, A., Scheer, M., Münch, R. et al. (2004) PrediSi: prediction of signal peptides and their<br />

cleavage positions. Nucleic Acids Res 32: W375-W379.<br />

Hinsley, A.P., Stanley, N.R., Palmer, T., and Berks, B.C. (2001) A naturally occurring bacterial Tat signal<br />

peptide lacking one of the 'invariant' arginine residues of the consensus targeting motif. FEBS Lett 497: 45-49.<br />

Hiramatsu, K., Hanaki, H., Ino, T., Yabuta, K. et al. (1997) Methicillin-resistant <strong>Staphylococcus</strong> <strong>aureus</strong> clinical<br />

strain with reduced vancomycin susceptibility. J Antimicrob Chemother 40: 135-136.<br />

Hirose, I., Sano, K., Shioda, I., Kumano, M. et al. (2000) Proteome analysis of Bacillus subtilis extracellular<br />

proteins: a two-dimensional protein electrophoretic study. Microbiology 146 ( Pt 1): 65-75.<br />

Hoffmann, T., Boiangiu, C., Moses, S., and Bremer, E. (2008) Responses of Bacillus subtilis to hypotonic<br />

challenges: physiological contributions of mechanosensitive channels to cellular survival. Appl Environ<br />

Microbiol 74: 2454-2460.<br />

Holden, M.T., Feil, E.J., Lindsay, J.A., Peacock, S.J. et al. (2004) Complete genomes of two clinical<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> strains: evidence for the rapid evolution of virulence and drug resistance. Proc Natl<br />

Acad Sci U S A 101: 9786-9791.<br />

Holtfreter, S., Bauer, K., Thomas, D., Feig, C. et al. (2004) egc-Encoded superantigens from <strong>Staphylococcus</strong><br />

<strong>aureus</strong> are neutralized by human sera much less efficiently than are classical staphylococcal enterotoxins or<br />

toxic shock syndrome toxin. Infect Immun 72: 4061-4071.<br />

Holtfreter, S., Roschack, K., Eichler, P., Eske, K. et al. (2006) <strong>Staphylococcus</strong> <strong>aureus</strong> carriers neutralize<br />

superantigens by antibodies specific for their colonizing strain: a potential explanation for their improved<br />

prognosis in severe sepsis. J Infect Dis 193: 1275-1278.<br />

Horsburgh, M.J., Aish, J.L., White, I.J., Shaw, L. et al. (2002) σ B modulates virulence determinant expression<br />

and stress resistance: characterization of a functional rsbU strain derived from <strong>Staphylococcus</strong> <strong>aureus</strong> 8325-4.<br />

J Bacteriol 184: 5457-5467.<br />

177


Chapter 10<br />

Huard, C., Miranda, G., Wessner, F., Bolotin, A. et al. (2003) Characterization of AcmB, an Nacetylglucosaminidase<br />

autolysin from Lactococcus lactis. Microbiology 149: 695-705.<br />

Hulett, F.M. (1996) <strong>The</strong> signal-transduction network for Pho regulation in Bacillus subtilis. Mol Microbiol 19:<br />

933-939.<br />

Hussain, M., Herrmann, M., von, E.C., Perdreau-Remington, F. et al. (1997) A 140-kilodalton extracellular<br />

protein is essential for the accumulation of <strong>Staphylococcus</strong> epidermidis strains on surfaces. Infect Immun 65:<br />

519-524.<br />

Ignatova, Z., Hornle, C., Nurk, A., and Kasche, V. (2002) Unusual signal peptide directs penicillin amidase<br />

from Escherichia coli to the Tat translocation machinery. Biochem Biophys Res Commun 291: 146-149.<br />

Ilangovan, U., Ton-That, H., Iwahara, J., Schneewind, O. et al. (2001) Structure of sortase, the transpeptidase<br />

that anchors proteins to the cell wall of <strong>Staphylococcus</strong> <strong>aureus</strong>. Proc Natl Acad Sci U S A 98: 6056-6061.<br />

Ishige, T., Krause, M., Bott, M., Wendisch, V.F. et al. (2003) <strong>The</strong> phosphate starvation stimulon of<br />

Corynebacterium glutamicum determined by DNA microarray analyses. J Bacteriol 185: 4519-4529.<br />

Ito, T., Hiramatsu, K., Oliveira, D.C., de Lencastre, H. et al. (2009) Classification of Staphylococcal Cassette<br />

Chromosome mec (SCCmec): Guidelines for Reporting Novel SCCmec Elements. Antimicrob Agents<br />

Chemother.<br />

Ito, T., Ma, X.X., Takeuchi, F., Okuma, K. et al. (2004) Novel type V staphylococcal cassette chromosome mec<br />

driven by a novel cassette chromosome recombinase, ccrC. Antimicrob Agents Chemother 48: 2637-2651.<br />

Jevons, M.P. (1961) "Celbenin"-resistant staphylococci. BMJ 1: 124-126.<br />

Ji, Y., Zhang, B., Van, S.F., Horn et al. (2001) Identification of critical staphylococcal genes using conditional<br />

phenotypes generated by antisense RNA. Science 293: 2266-2269.<br />

Jongbloed, J.D., Antelmann, H., Hecker, M., Nijland, R. et al. (2002) Selective contribution of the twinarginine<br />

translocation pathway to protein secretion in Bacillus subtilis. J Biol Chem 277: 44068-44078.<br />

Jongbloed, J.D., Grieger, U., Antelmann, H., Hecker, M. et al. (2004) Two minimal Tat translocases in<br />

Bacillus. Mol Microbiol 54: 1319-1325.<br />

Jongbloed, J.D., Martin, U., Antelmann, H., Hecker, M. et al. (2000) TatC is a specificity determinant for<br />

protein secretion via the twin-arginine translocation pathway. J Biol Chem 275: 41350-41357.<br />

Jongbloed, J.D., van der Ploeg, R., and van Dijl, J.M. (2005) Bifunctional TatA subunits in minimal Tat protein<br />

translocases. Trends Microbiol.<br />

Jonquieres, R., Bierne, H., Fiedler, F., Gounon, P. et al. (1999) Interaction between the protein InlB of Listeria<br />

monocytogenes and lipoteichoic acid: a novel mechanism of protein association at the surface of gram-positive<br />

bacteria. Mol Microbiol 34: 902-914.<br />

Joris, B., Englebert, S., Chu, C.P., Kariyama, R. et al. (1992) Modular design of the Enterococcus hirae<br />

muramidase-2 and Streptococcus faecalis autolysin. FEMS Microbiol Lett 70: 257-264.<br />

Josefsson, E., Hartford, O., O'Brien, L., Patti, J.M. et al. (2001) Protection against experimental<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> arthritis by vaccination with clumping factor A, a novel virulence determinant. J Infect<br />

Dis 184: 1572-1580.<br />

Josefsson, E., McCrea, K.W., Ni Eidhin, D., O'Connell, D. et al. (1998) Three new members of the serineaspartate<br />

repeat protein multigene family of <strong>Staphylococcus</strong> <strong>aureus</strong>. Microbiology 144 ( Pt 12): 3387-3395.<br />

Juncker, A.S., Willenbrock, H., von Heijne, G., Brunak, S. et al. (2003) Prediction of lipoprotein signal<br />

peptides in Gram-negative bacteria. Protein Sci 12: 1652-1662.<br />

Kaito, C., Omae, Y., Matsumoto, Y., Nagata, M. et al. (2008) A novel gene, fudoh, in the SCCmec region<br />

suppresses the colony spreading ability and virulence of <strong>Staphylococcus</strong> <strong>aureus</strong>. PLoS ONE 3: e3921.<br />

Kaito, C., and Sekimizu, K. (2007) Colony spreading in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 189: 2553-2557.<br />

Kall, L., Krogh, A., and Sonnhammer, E.L. (2004) A combined transmembrane topology and signal peptide<br />

prediction method. J Mol Biol 338: 1027-1036.<br />

Kang, S.S., Kim, J.G., Lee, T.H., and Oh, K.B. (2006) Flavonols inhibit sortases and sortase-mediated<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> clumping to fibrinogen. Biol Pharm Bull 29: 1751-1755.<br />

Karlsson, A., and Arvidson, S. (2002) Variation in extracellular protease production among clinical isolates of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> due to different levels of expression of the protease repressor sarA. Infect Immun 70:<br />

4239-4246.<br />

Kim, L., Mogk, A., and Schumann, W. (1996) A xylose-inducible Bacillus subtilis integration vector and its<br />

application. Gene 181: 71-76.<br />

Kinsinger, R.F., Kearns, D.B., Hale, M., and Fall, R. (2005) Genetic requirements for potassium ion-dependent<br />

colony spreading in Bacillus subtilis. J Bacteriol 187: 8462-8469.<br />

Kobayashi, H., Ohashi, Y., Nanamiya, H., Asai, K. et al. (2000) Genetic analysis of SecA-SecY interaction<br />

required for spore development in Bacillus subtilis. FEMS Microbiol Lett 184: 285-289.<br />

Komatsuzawa, H., Ohta, K., Sugai, M., Fujiwara, T. et al. (2000) Tn551-mediated insertional inactivation of<br />

the fmtB gene encoding a cell wall-associated protein abolishes methicillin resistance in <strong>Staphylococcus</strong><br />

<strong>aureus</strong>. J Antimicrob Chemother 45: 421-431.<br />

Kontinen, V.P., and Sarvas, M. (1993) <strong>The</strong> PrsA lipoprotein is essential for protein secretion in Bacillus subtilis<br />

and sets a limit for high-level secretion. Mol Microbiol 8: 727-737.<br />

Kouwen, T.R., and van Dijl, J.M. (2009a) Interchangeable modules in bacterial thiol-disulfide exchange<br />

pathways. Trends Microbiol 17: 6-12.<br />

Kouwen, T.R., Antelmann, H., van der Ploeg, R., Denham, E.L. et al. (2009b) MscL of Bacillus subtilis<br />

prevents selective release of cytoplasmic proteins in a hypotonic environment. Proteomics 9: 1033-1043.<br />

178


Reference list<br />

Kouwen, T.R., Trip, E.N., Denham, E.L., Sibbald, M.J. et al. (2009c) <strong>The</strong> large mechanosensitive channel<br />

MscL determines bacterial susceptibility to the bacteriocin sublancin 168. Antimicrob Agents Chemother 53:<br />

4702-4711.<br />

Kouwen, T.R., van der, G.A., Dorenbos, R., Winter, T. et al. (2007) Thiol-disulphide oxidoreductase modules<br />

in the low-GC Gram-positive bacteria. Mol Microbiol 64: 984-999.<br />

Kreiswirth, B.N., Löfdahl, S., Betley, M.J., O'Reilly, M. et al. (1983) <strong>The</strong> toxic shock syndrome exotoxin<br />

structural gene is not detectably transmitted by a prophage. Nature 305: 709-712.<br />

Krogh, A., Larsson, B., von, H.G., and Sonnhammer, E.L. (2001) Predicting transmembrane protein topology<br />

with a hidden Markov model: application to complete genomes. J Mol Biol 305: 567-580.<br />

Krogh, S., Jorgensen, S.T., and Devine, K.M. (1998) Lysis genes of the Bacillus subtilis defective prophage<br />

PBSX. J Bacteriol 180: 2110-2117.<br />

Kruger, R.G., Otvos, B., Frankel, B.A., Bentley, M. et al. (2004) Analysis of the substrate specificity of the<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> sortase transpeptidase SrtA. Biochemistry 43: 1541-1551.<br />

Kuipers, O.P., Beerthuyzen, M.M., Siezen, R.J., and de Vos, W.M. (1993) Characterization of the nisin gene<br />

cluster nisABTCIPR of Lactococcus lactis. Requirement of expression of the nisA and nisI genes for<br />

development of immunity. Eur J Biochem 216: 281-291.<br />

Kunst, F., Ogasawara, N., Moszer, I., Albertini, A.M. et al. (1997) <strong>The</strong> complete genome sequence of the grampositive<br />

bacterium Bacillus subtilis. Nature 390: 249-256.<br />

Kunst, F., and Rapoport, G. (1995) Salt stress is an environmental signal affecting degradative enzyme synthesis<br />

in Bacillus subtilis. J Bacteriol 177: 2403-2407.<br />

Kuroda, M., Ohta, T., Uchiyama, I., Baba, T. et al. (2001) Whole genome sequencing of meticillin-resistant<br />

<strong>Staphylococcus</strong> <strong>aureus</strong>. Lancet 357: 1225-1240.<br />

Kurokawa, K., Lee, H., Roh, K.B., Asanuma, M. et al. (2009) <strong>The</strong> Triacylated ATP Binding Cluster Transporter<br />

Substrate-binding Lipoprotein of <strong>Staphylococcus</strong> <strong>aureus</strong> Functions as a Native Ligand for Toll-like Receptor 2.<br />

J Biol Chem 284: 8406-8411.<br />

Labandeira-Rey, M., Couzon, F., Boisset, S., Brown, E.L. et al. (2007) <strong>Staphylococcus</strong> <strong>aureus</strong> Panton-<br />

Valentine leukocidin causes necrotizing pneumonia. Science 315: 1130-1133.<br />

Lazarevic, V., Düsterhöft, A., Soldo, B., Hilbert, H. et al. (1999) Nucleotide sequence of the Bacillus subtilis<br />

temperate bacteriophage SPβc2. Microbiology 145 ( Pt 5): 1055-1067.<br />

Leclercq, R., Derlot, E., Weber, M., Duval, J. et al. (1989) Transferable vancomycin and teicoplanin resistance<br />

in Enterococcus faecium. Antimicrob Agents Chemother 33: 10-15.<br />

Lee, V.T., and Schneewind, O. (2001) Protein secretion and the pathogenesis of bacterial infections. Genes Dev<br />

15: 1725-1752.<br />

Leskelä, S., Wahlström, E., Kontinen, V.P., and Sarvas, M. (1999) Lipid modification of prelipoproteins is<br />

dispensable for growth but essential for efficient protein secretion in Bacillus subtilis: characterization of the<br />

Lgt gene. Mol Microbiol 31: 1075-1085.<br />

Lina, G., Boutite, F., Tristan, A., Bes, M. et al. (2003) Bacterial competition for human nasal cavity<br />

colonization: role of Staphylococcal agr alleles. Appl Environ Microbiol 69: 18-23.<br />

Lindsay, J.A., and Holden, M.T. (2004) <strong>Staphylococcus</strong> <strong>aureus</strong>: superbug, super genome? Trends Microbiol 12:<br />

378-385.<br />

Liu, W., and Hulett, F.M. (1998) Comparison of PhoP binding to the tuaA promoter with PhoP binding to other<br />

Pho-regulon promoters establishes a Bacillus subtilis Pho core binding site. Microbiology 144 ( Pt 5): 1443-<br />

1450.<br />

Lorenz, U., Ohlsen, K., Karch, H., Hecker, M. et al. (2000) Human antibody response during sepsis against<br />

targets expressed by methicillin resistant <strong>Staphylococcus</strong> <strong>aureus</strong>. FEMS Immunol Med Microbiol 29: 145-153.<br />

Lory, S. (1994) Leader peptidase of type IV prepilins and related proteins. In Signal Peptidases. von Heijne,G.<br />

(ed). Austin, TX, USA: R.G. Landes Company, pp. 17-29.<br />

Lory, S. (1998) Secretion of proteins and assembly of bacterial surface organelles: shared pathways of<br />

extracellular protein targeting. Curr Opin Microbiol 1: 27-35.<br />

Machata, S., Tchatalbachev, S., Mohamed, W., Jänsch, L. et al. (2008) Lipoproteins of Listeria monocytogenes<br />

are critical for virulence and TLR2-mediated immune activation. J Immunol 181: 2028-2035.<br />

Mack, D., Fischer, W., Krokotsch, A., Leopold, K. et al. (1996) <strong>The</strong> intercellular adhesin involved in biofilm<br />

accumulation of <strong>Staphylococcus</strong> epidermidis is a linear β-1,6-linked glucosaminoglycan: purification and<br />

structural analysis. J Bacteriol 178: 175-183.<br />

Mack, D., Siemssen, N., and Laufs, R. (1992) Parallel induction by glucose of adherence and a polysaccharide<br />

antigen specific for plastic-adherent <strong>Staphylococcus</strong> epidermidis: evidence for functional relation to<br />

intercellular adhesion. Infect Immun 60: 2048-2057.<br />

Mäder, U., Antelmann, H., Buder, T., Dahl, M.K. et al. (2002) Bacillus subtilis functional genomics: genomewide<br />

analysis of the DegS-DegU regulon by transcriptomics and proteomics. Mol Genet Genomics 268: 455-<br />

467.<br />

Mandlik, A., Swierczynski, A., Das, A., and Ton-That, H. (2008) Pili in Gram-positive bacteria: assembly,<br />

involvement in colonization and biofilm development. Trends Microbiol 16: 33-40.<br />

Margot, P., and Karamata, D. (1996) <strong>The</strong> wprA gene of Bacillus subtilis 168, expressed during exponential<br />

growth, encodes a cell-wall-associated protease. Microbiology 142 ( Pt 12): 3437-3444.<br />

Marraffini, L.A., Dedent, A.C., and Schneewind, O. (2006) Sortases and the art of anchoring proteins to the<br />

envelopes of gram-positive bacteria. Microbiol Mol Biol Rev 70: 192-221.<br />

179


Chapter 10<br />

Marraffini, L.A., Ton-That, H., Zong, Y., Narayana, S.V. et al. (2004) Anchoring of surface proteins to the cell<br />

wall of <strong>Staphylococcus</strong> <strong>aureus</strong>. A conserved arginine residue is required for efficient catalysis of sortase A. J<br />

Biol Chem 279: 37763-37770.<br />

Martin-Verstraete, I., Deutscher, J., and Galinier, A. (1999) Phosphorylation of HPr and Crh by HprK, early<br />

steps in the catabolite repression signalling pathway for the Bacillus subtilis levanase operon. J Bacteriol 181:<br />

2966-2969.<br />

Martinac, B. (2001) Mechanosensitive channels in prokaryotes. Cell Physiol Biochem 11: 61-76.<br />

Martínez, A., Torello, S., and Kolter, R. (1999) Sliding motility in mycobacteria. J Bacteriol 181: 7331-7338.<br />

Matsuyama, S., Fujita, Y., and Mizushima, S. (1993) SecD is involved in the release of translocated secretory<br />

proteins from the cytoplasmic membrane of Escherichia coli. EMBO J 12: 265-270.<br />

Mazmanian, S.K., Liu, G., Jensen, E.R., Lenoy, E. et al. (2000) <strong>Staphylococcus</strong> <strong>aureus</strong> sortase mutants<br />

defective in the display of surface proteins and in the pathogenesis of animal infections. Proc Natl Acad Sci U<br />

S A 97: 5510-5515.<br />

Mazmanian, S.K., Skaar, E.P., Gaspar, A.H., Humayun, M. et al. (2003) Passage of heme-iron across the<br />

envelope of <strong>Staphylococcus</strong> <strong>aureus</strong>. Science 299: 906-909.<br />

Mazmanian, S.K., Ton-That, H., and Schneewind, O. (2001) Sortase-catalysed anchoring of surface proteins to<br />

the cell wall of <strong>Staphylococcus</strong> <strong>aureus</strong>. Mol Microbiol 40: 1049-1057.<br />

Mazmanian, S.K., Ton-That, H., Su, K., and Schneewind, O. (2002) An iron-regulated sortase anchors a class<br />

of surface protein during <strong>Staphylococcus</strong> <strong>aureus</strong> pathogenesis. Proc Natl Acad Sci U S A 99: 2293-2298.<br />

McAuliffe, O., Ross, R.P., and Hill, C. (2001) Lantibiotics: structure, biosynthesis and mode of action. FEMS<br />

Microbiol Rev 25: 285-308.<br />

McCarter, L.L. (2006) Regulation of flagella. Curr Opin Microbiol 9: 180-186.<br />

McDevitt, D., Francois, P., Vaudaux, P., and Foster, T.J. (1995) Identification of the ligand-binding domain of<br />

the surface-located fibrinogen receptor (clumping factor) of <strong>Staphylococcus</strong> <strong>aureus</strong>. Mol Microbiol 16: 895-<br />

907.<br />

McDevitt, D., Nanavaty, T., House-Pompeo, K., Bell, E. et al. (1997) Characterization of the interaction<br />

between the <strong>Staphylococcus</strong> <strong>aureus</strong> clumping factor (ClfA) and fibrinogen. Eur J Biochem 247: 416-424.<br />

Meima, R., Eschevins, C., Fillinger, S., Bolhuis, A. et al. (2002) <strong>The</strong> bdbDC operon of Bacillus subtilis encodes<br />

thiol-disulfide oxidoreductases required for competence development. J Biol Chem 277: 6994-7001.<br />

Meima, R., and van Dijl, J.M. (2003) Protein secretion in Gram-positive bacteria. In Protein secretion pathways<br />

in bacteria. Oudega,B. (ed). Dordrecht, the Netherlands: Kluwer Academic Publishers, pp. 271-296.<br />

Melles, D.C., Gorkink, R.F., Boelens, H.A., Snijders, S.V. et al. (2004) Natural population dynamics and<br />

expansion of pathogenic clones of <strong>Staphylococcus</strong> <strong>aureus</strong>. J Clin Invest 114: 1732-1740.<br />

Mempel, M., Schnopp, C., Hojka, M., Fesq, H. et al. (2002) Invasion of human keratinocytes by <strong>Staphylococcus</strong><br />

<strong>aureus</strong> and intracellular bacterial persistence represent haemolysin-independent virulence mechanisms that are<br />

followed by features of necrotic and apoptotic keratinocyte cell death. Br J Dermatol 146: 943-951.<br />

Merino, N., Toledo-Arana, A., Vergara-Irigaray, M., Valle, J. et al. (2009) Protein A-mediated multicellular<br />

behavior in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 191: 832-843.<br />

Meyer, C., Bierbaum, G., Heidrich, C., Reis, M. et al. (1995) Nucleotide sequence of the lantibiotic Pep5<br />

biosynthetic gene cluster and functional analysis of PepP and PepC. Evidence for a role of PepC in thioether<br />

formation. Eur J Biochem 232: 478-489.<br />

Michiels, J., Dirix, G., Vanderleyden, J., and Xi, C. (2001) Processing and export of peptide pheromones and<br />

bacteriocins in Gram-negative bacteria. Trends Microbiol 9: 164-168.<br />

Middleton, J.R. (2008) <strong>Staphylococcus</strong> <strong>aureus</strong> antigens and challenges in vaccine development. Expert Rev<br />

Vaccines 7: 805-815.<br />

Milward, C.P., and Jacques, N.A. (1990) Secretion of fructosyltransferase by Streptococcus salivarius involves<br />

the sucrose-dependent release of the cell-bound form. J Gen Microbiol 136: 165-169.<br />

Molik, S., Karnauchov, I., Weidlich, C., Herrmann, R.G. et al. (2001) <strong>The</strong> Rieske Fe/S protein of the<br />

cytochrome b6/f complex in chloroplasts: missing link in the evolution of protein transport pathways in<br />

chloroplasts? J Biol Chem 276: 42761-42766.<br />

Monecke, S., Slickers, P., Hotzel, H., Richter-Huhn, G. et al. (2006) Microarray-based characterisation of a<br />

Panton-Valentine leukocidin-positive community-acquired strain of methicillin-resistant <strong>Staphylococcus</strong><br />

<strong>aureus</strong>. Clin Microbiol Infect 12: 718-728.<br />

Morfeldt, E., Janzon, L., Arvidson, S., and Löfdahl, S. (1988) Cloning of a chromosomal locus (exp) which<br />

regulates the expression of several exoprotein genes in <strong>Staphylococcus</strong> <strong>aureus</strong>. Mol Gen Genet 211: 435-440.<br />

Mori, H., and Cline, K. (2001) Post-translational protein translocation into thylakoids by the Sec and ∆pHdependent<br />

pathways. Biochim Biophys Acta 1541: 80-90.<br />

Morikawa, K., Inose, Y., Okamura, H., Maruyama, A. et al. (2003) A new staphylococcal sigma factor in the<br />

conserved gene cassette: functional significance and implication for the evolutionary processes. Genes Cells 8:<br />

699-712.<br />

Msadek,T., Kunst, F., and Rapoport, G. (1995) Two-component signal transduction Washington DC: pp. 447-<br />

471.<br />

Müller, J.P., Ozegowski, J., Vettermann, S., Swaving, J. et al. (2000) Interaction of Bacillus subtilis CsaA with<br />

SecA and precursor proteins. Biochem J 348 Pt 2: 367-373.<br />

Murakami, T., Haga, K., Takeuchi, M., and Sato, T. (2002) Analysis of the Bacillus subtilis spoIIIJ gene and<br />

its paralogue gene, yqjG. J Bacteriol 184: 1998-2004.<br />

180


Reference list<br />

Musser, J.M., Schlievert, P.M., Chow, A.W., Ewan, P. et al. (1990) A single clone of <strong>Staphylococcus</strong> <strong>aureus</strong><br />

causes the majority of cases of toxic shock syndrome. Proc Natl Acad Sci U S A 87: 225-229.<br />

Mwangi, M.M., Wu, S.W., Zhou, Y., Sieradzki, K. et al. (2007) Tracking the in vivo evolution of multidrug<br />

resistance in <strong>Staphylococcus</strong> <strong>aureus</strong> by whole-genome sequencing. Proc Natl Acad Sci U S A 104: 9451-9456.<br />

Nagamori, S., Nishiyama, K., and Tokuda, H. (2000) Two SecG molecules present in a single protein<br />

translocation machinery are functional even after crosslinking. J Biochem (Tokyo) 128: 129-137.<br />

Nagao, J., Asaduzzaman, S.M., Aso, Y., Okuda, K. et al. (2006) Lantibiotics: insight and foresight for new<br />

paradigm. J Biosci Bioeng 102: 139-149.<br />

Naik, M.T., Suree, N., Ilangovan, U., Liew, C.K. et al. (2006) <strong>Staphylococcus</strong> <strong>aureus</strong> Sortase A transpeptidase.<br />

Calcium promotes sorting signal binding by altering the mobility and structure of an active site loop. J Biol<br />

Chem 281: 1817-1826.<br />

Nandakumar, R., Nandakumar, M.P., Marten, M.R., and Ross, J.M. (2005) Proteome Analysis of Membrane<br />

and Cell Wall Associated Proteins from <strong>Staphylococcus</strong> <strong>aureus</strong>. J Proteome Res 4: 250-257.<br />

Nanra, J.S., Timofeyeva, Y., Buitrago, S.M., Sellman, B.R. et al. (2009) Heterogeneous in vivo expression of<br />

clumping factor A and capsular polysaccharide by <strong>Staphylococcus</strong> <strong>aureus</strong>: implications for vaccine design.<br />

Vaccine 27: 3276-3280.<br />

Narita, S., and Tokuda, H. (2006) An ABC transporter mediating the membrane detachment of bacterial<br />

lipoproteins depending on their sorting signals. FEBS Lett 580: 1164-1170.<br />

Navaratna, M.A., Sahl, H.G., and Tagg, J.R. (1998) Two-component anti-<strong>Staphylococcus</strong> <strong>aureus</strong> lantibiotic<br />

activity produced by <strong>Staphylococcus</strong> <strong>aureus</strong> C55. Appl Environ Microbiol 64: 4803-4808.<br />

Navaratna, M.A., Sahl, H.G., and Tagg, J.R. (1999) Identification of genes encoding two-component lantibiotic<br />

production in <strong>Staphylococcus</strong> <strong>aureus</strong> C55 and other phage group II S. <strong>aureus</strong> strains and demonstration of an<br />

association with the exfoliative toxin B gene. Infect Immun 67: 4268-4271.<br />

Navarre, W.W., and Schneewind, O. (1994) Proteolytic cleavage and cell wall anchoring at the LPXTG motif of<br />

surface proteins in gram-positive bacteria. Mol Microbiol 14: 115-121.<br />

Neoh, H.M., Cui, L., Yuzawa, H., Takeuchi, F. et al. (2008) Mutated response regulator graR is responsible for<br />

phenotypic conversion of <strong>Staphylococcus</strong> <strong>aureus</strong> from heterogeneous vancomycin-intermediate resistance to<br />

vancomycin-intermediate resistance. Antimicrob Agents Chemother 52: 45-53.<br />

Netz, D.J., Bastos Mdo, C., and Sahl, H.G. (2002a) Mode of action of the antimicrobial peptide aureocin A53<br />

from <strong>Staphylococcus</strong> <strong>aureus</strong>. Appl Environ Microbiol 68: 5274-5280.<br />

Netz, D.J., Pohl, R., Beck-Sickinger, A.G., Selmer, T. et al. (2002b) Biochemical characterisation and genetic<br />

analysis of aureocin A53, a new, atypical bacteriocin from <strong>Staphylococcus</strong> <strong>aureus</strong>. J Mol Biol 319: 745-756.<br />

Netz, D.J., Sahl, H.G., Marcelino, R., dos Santos Nascimento, J. et al. (2001) Molecular characterisation of<br />

aureocin A70, a multi-peptide bacteriocin isolated from <strong>Staphylococcus</strong> <strong>aureus</strong>. J Mol Biol 311: 939-949.<br />

Ní Eidhin, D., Perkins, S., Francois, P., Vaudaux, P. et al. (1998) Clumping factor B (ClfB), a new surfacelocated<br />

fibrinogen-binding adhesin of <strong>Staphylococcus</strong> <strong>aureus</strong>. Mol Microbiol 30: 245-257.<br />

Nielsen, H., Engelbrecht, J., Brunak, S., and von Heijne, G. (1997) Identification of prokaryotic and eukaryotic<br />

signal peptides and prediction of their cleavage sites. Protein Eng 10: 1-6.<br />

Nishiyama, K., Hanada, M., and Tokuda, H. (1994) Disruption of the gene encoding p12 (SecG) reveals the<br />

direct involvement and important function of SecG in the protein translocation of Escherichia coli at low<br />

temperature. EMBO J 13: 3272-3277.<br />

Nishiyama, K., Mizushima, S., and Tokuda, H. (1993) A novel membrane protein involved in protein<br />

translocation across the cytoplasmic membrane of Escherichia coli. EMBO J 12: 3409-3415.<br />

Noble, W.C., Virani, Z., and Cree, R.G. (1992) Co-transfer of vancomycin and other resistance genes from<br />

Enterococcus faecalis NCTC 12201 to <strong>Staphylococcus</strong> <strong>aureus</strong>. FEMS Microbiol Lett 72: 195-198.<br />

Norrander, J., Kempe, T., and Messing, J. (1983) Construction of improved M13 vectors using<br />

oligodeoxynucleotide-directed mutagenesis. Gene 26: 101-106.<br />

Nouwen, N., Piwowarek, M., Berrelkamp, G., and Driessen, A.J. (2005) <strong>The</strong> large first periplasmic loop of<br />

SecD and SecF plays an important role in SecDF functioning. J Bacteriol 187: 5857-5860.<br />

Nouwens, A.S., Beatson, S.A., Whitchurch, C.B., Walsh, B.J. et al. (2003) Proteome analysis of extracellular<br />

proteins regulated by the las and rhl quorum sensing systems in Pseudomonas aeruginosa PAO1.<br />

Microbiology 149: 1311-1322.<br />

Novick, R. (1967) Properties of a cryptic high-frequency transducing phage in <strong>Staphylococcus</strong> <strong>aureus</strong>. Virology<br />

33: 155-166.<br />

Novick, R.P. (1991) Genetic systems in staphylococci. Methods Enzymol 204: 587-636.<br />

Novick, R.P. (2003) Autoinduction and signal transduction in the regulation of staphylococcal virulence. Mol<br />

Microbiol 48: 1429-1449.<br />

Ogura, M., Yamaguchi, H., Yoshida, K., Fujita, Y. et al. (2001) DNA microarray analysis of Bacillus subtilis<br />

DegU, ComA and PhoP regulons: an approach to comprehensive analysis of B. subtilis two-component<br />

regulatory systems. Nucleic Acids Res 29: 3804-3813.<br />

Oh, K.B., Oh, M.N., Kim, J.G., Shin, D.S. et al. (2006) Inhibition of sortase-mediated <strong>Staphylococcus</strong> <strong>aureus</strong><br />

adhesion to fibronectin via fibronectin-binding protein by sortase inhibitors. Appl Microbiol Biotechnol 70:<br />

102-106.<br />

Oshida, T., Sugai, M., Komatsuzawa, H., Hong, Y.M. et al. (1995) A <strong>Staphylococcus</strong> <strong>aureus</strong> autolysin that has<br />

an N-acetylmuramoyl-L-alanine amidase domain and an endo-β-N-acetylglucosaminidase domain: cloning,<br />

sequence analysis, and characterization. Proc Natl Acad Sci U S A 92: 285-289.<br />

181


Chapter 10<br />

Otto, M. (2008) Targeted immunotherapy for staphylococcal infections: focus on anti-MSCRAMM antibodies.<br />

BioDrugs 22: 27-36.<br />

Paik, S.H., Chakicherla, A., and Hansen, J.N. (1998) Identification and characterization of the structural and<br />

transporter genes for, and the chemical and biological properties of, sublancin 168, a novel lantibiotic produced<br />

by Bacillus subtilis 168. J Biol Chem 273: 23134-23142.<br />

Pallen, M.J. (2002) <strong>The</strong> ESAT-6/WXG100 superfamily -- and a new Gram-positive secretion system? Trends<br />

Microbiol 10: 209-212.<br />

Pantucek, R., Doskar, J., Ruzicková, V., Kaspárek, P. et al. (2004) Identification of bacteriophage types and<br />

their carriage in <strong>Staphylococcus</strong> <strong>aureus</strong>. Arch Virol 149: 1689-1703.<br />

Papanikou, E., Karamanou, S., and Economou, A. (2007) Bacterial protein secretion through the translocase<br />

nanomachine. Nat Rev Microbiol 5: 839-851.<br />

Patel, A.H., Nowlan, P., Weavers, E.D., and Foster, T. (1987) Virulence of protein A-deficient and α-toxindeficient<br />

mutants of <strong>Staphylococcus</strong> <strong>aureus</strong> isolated by allele replacement. Infect Immun 55: 3103-3110.<br />

Peacock, S.J., de Silva, I., and Lowy, F.D. (2001) What determines nasal carriage of <strong>Staphylococcus</strong> <strong>aureus</strong>?<br />

Trends Microbiol 9: 605-610.<br />

Peacock, S.J., Moore, C.E., Justice, A., Kantzanou, M. et al. (2002) Virulent combinations of adhesin and toxin<br />

genes in natural populations of <strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 70: 4987-4996.<br />

Peng, H.L., Novick, R.P., Kreiswirth, B., Kornblum, J. et al. (1988) Cloning, characterization, and sequencing<br />

of an accessory gene regulator (agr) in <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 170: 4365-4372.<br />

Peri, S., Steen, H., and Pandey, A. (2001) GPMAW--a software tool for analyzing proteins and peptides. Trends<br />

Biochem Sci 26: 687-689.<br />

Peschel, A., and Collins, L.V. (2001) Staphylococcal resistance to antimicrobial peptides of mammalian and<br />

bacterial origin. Peptides 22: 1651-1659.<br />

Peschel, A., and Götz, F. (1996) Analysis of the <strong>Staphylococcus</strong> epidermidis genes epiF, -E, and -G involved in<br />

epidermin immunity. J Bacteriol 178: 531-536.<br />

Peschel, A., Otto, M., Jack, R.W., Kalbacher, H. et al. (1999) Inactivation of the dlt operon in <strong>Staphylococcus</strong><br />

<strong>aureus</strong> confers sensitivity to defensins, protegrins, and other antimicrobial peptides. J Biol Chem 274: 8405-<br />

8410.<br />

Peschel, A., Schnell, N., Hille, M., Entian, K.D. et al. (1997) Secretion of the lantibiotics epidermin and<br />

gallidermin: sequence analysis of the genes gdmT and gdmH, their influence on epidermin production and their<br />

regulation by EpiQ. Mol Gen Genet 254: 312-318.<br />

Pivetti, C.D., Yen, M.R., Miller, S., Busch, W. et al. (2003) Two families of mechanosensitive channel proteins.<br />

Microbiol Mol Biol Rev 67: 66-85, table.<br />

Pocsfalvi, G., Cacace, G., Cuccurullo, M., Serluca, G. et al. (2008) Proteomic analysis of exoproteins expressed<br />

by enterotoxigenic <strong>Staphylococcus</strong> <strong>aureus</strong> strains. Proteomics 8: 2462-2476.<br />

Pohlschröder, M., Prinz, W.A., Hartmann, E., and Beckwith, J. (1997) Protein translocation in the three<br />

domains of life: variations on a theme. Cell 91: 563-566.<br />

Ponting, C.P., Aravind, L., Schultz, J., Bork, P. et al. (1999) Eukaryotic signalling domain homologues in<br />

archaea and bacteria. Ancient ancestry and horizontal gene transfer. J Mol Biol 289: 729-745.<br />

Prágai, Z., and Harwood, C.R. (2002) Regulatory interactions between the Pho and σ B -dependent general stress<br />

regulons of Bacillus subtilis. Microbiology 148: 1593-1602.<br />

Prágai, Z., Tjalsma, H., Bolhuis, A., van Dijl, J.M. et al. (1997) <strong>The</strong> signal peptidase II (Isp) gene of Bacillus<br />

subtilis. Microbiology 143 ( Pt 4): 1327-1333.<br />

Pugsley, A.P. (1993) Processing and methylation of PuIG, a pilin-like component of the general secretory pathway<br />

of Klebsiella oxytoca. Mol Microbiol 9: 295-308.<br />

Race, P.R., Bentley, M.L., Melvin, J.A., Crow, A. et al. (2009) Crystal structure of Streptococcus pyogenes<br />

sortase A: implications for sortase mechanism. J Biol Chem 284: 6924-6933.<br />

Ramadurai, L., Lockwood, K.J., Nadakavukaren, M.J., and Jayaswal, R.K. (1999) Characterization of a<br />

chromosomally encoded glycylglycine endopeptidase of <strong>Staphylococcus</strong> <strong>aureus</strong>. Microbiology 145 ( Pt 4):<br />

801-808.<br />

Rashid, M.H., Mori, M., and Sekiguchi, J. (1995) Glucosaminidase of Bacillus subtilis: cloning, regulation,<br />

primary structure and biochemical characterization. Microbiology 141 ( Pt 10): 2391-2404.<br />

Rashid, M.H., Rao, N.N., and Kornberg, A. (2000) Inorganic polyphosphate is required for motility of bacterial<br />

pathogens. J Bacteriol 182: 225-227.<br />

Rathsam, C., Giffard, P.M., and Jacques, N.A. (1993) <strong>The</strong> cell-bound fructosyltransferase of Streptococcus<br />

salivarius: the carboxyl terminus specifies attachment in a Streptococcus gordonii model system. J Bacteriol<br />

175: 4520-4527.<br />

Recsei, P., Kreiswirth, B., O'Reilly, M., Schlievert, P. et al. (1986) Regulation of exoprotein gene expression in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> by agar. Mol Gen Genet 202: 58-61.<br />

Rey, M.W., Ramaiya, P., Nelson, B.A., Brody-Karpin, S.D. et al. (2004) Complete genome sequence of the<br />

industrial bacterium Bacillus licheniformis and comparisons with closely related Bacillus species. Genome Biol<br />

5: R77.<br />

Rice, K.C., Firek, B.A., Nelson, J.B., Yang, S.J. et al. (2003) <strong>The</strong> <strong>Staphylococcus</strong> <strong>aureus</strong> cidAB operon:<br />

evaluation of its role in regulation of murein hydrolase activity and penicillin tolerance. J Bacteriol 185: 2635-<br />

2643.<br />

182


Reference list<br />

Robinson, C., and Bolhuis, A. (2001) Protein targeting by the twin-arginine translocation pathway. Nat Rev Mol<br />

Cell Biol 2: 350-356.<br />

Roche, F.M., Massey, R., Peacock, S.J., Day, N.P. et al. (2003) Characterization of novel LPXTG-containing<br />

proteins of <strong>Staphylococcus</strong> <strong>aureus</strong> identified from genome sequences. Microbiology 149: 643-654.<br />

Rogasch, K., Rühmling, V., Pané-Farré, J., Höper, D. et al. (2006) Influence of the two-component system<br />

SaeRS on global gene expression in two different <strong>Staphylococcus</strong> <strong>aureus</strong> strains. J Bacteriol 188: 7742-7758.<br />

Rohde, H., Burandt, E.C., Siemssen, N., Frommelt, L. et al. (2007) Polysaccharide intercellular adhesin or<br />

protein factors in biofilm accumulation of <strong>Staphylococcus</strong> epidermidis and <strong>Staphylococcus</strong> <strong>aureus</strong> isolated<br />

from prosthetic hip and knee joint infections. Biomaterials 28: 1711-1720.<br />

Rohde, H., Burdelski, C., Bartscht, K., Hussain, M. et al. (2005) Induction of <strong>Staphylococcus</strong> epidermidis<br />

biofilm formation via proteolytic processing of the accumulation-associated protein by staphylococcal and host<br />

proteases. Mol Microbiol 55: 1883-1895.<br />

Ronen, M., Rosenberg, R., Shraiman, B.I., and Alon, U. (2002) Assigning numbers to the arrows:<br />

parameterizing a gene regulation network by using accurate expression kinetics. Proc Natl Acad Sci U S A 99:<br />

10555-10560.<br />

Sahl, H.G., and Bierbaum, G. (1998) Lantibiotics: biosynthesis and biological activities of uniquely modified<br />

peptides from gram-positive bacteria. Annu Rev Microbiol 52: 41-79.<br />

Saïd-Salim, B., Dunman, P.M., McAleese, F.M., Macapagal, D. et al. (2003) Global regulation of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> genes by Rot. J Bacteriol 185: 610-619.<br />

Saller, M.J., Fusetti, F., and Driessen, A.J. (2009) Bacillus subtilis SpoIIIJ and YqjG function in membrane<br />

protein biogenesis. J Bacteriol 191: 6749-6757.<br />

Sambrook,J., Fritsch, E.F., and Maniatis, T. (1989) Molecular cloning: a laboratory manual New York: Cold<br />

Spring Harbor Laboratory Press.<br />

Sander, P., Rezwan, M., Walker, B., Rampini, S.K. et al. (2004) Lipoprotein processing is required for<br />

virulence of Mycobacterium tuberculosis. Mol Microbiol 52: 1543-1552.<br />

Sankaran, K., Gan, K., Rash, B., Qi, H.Y. et al. (1997) Roles of histidine-103 and tyrosine-235 in the function of<br />

the prolipoprotein diacylglyceryl transferase of Escherichia coli. J Bacteriol 179: 2944-2948.<br />

Sankaran, K., and Wu, H.C. (1994) Lipid modification of bacterial prolipoprotein. Transfer of diacylglyceryl<br />

moiety from phosphatidylglycerol. J Biol Chem 269: 19701-19706.<br />

Sarvas, M., Harwood, C.R., Bron, S., and van Dijl, J.M. (2004) Post-translocational folding of secretory<br />

proteins in Gram-positive bacteria. Biochim Biophys Acta 1694: 311-327.<br />

Sasso, E.H., Silverman, G.J., and Mannik, M. (1991) Human IgA and IgG F(ab')2 that bind to staphylococcal<br />

protein A belong to the VHIII subgroup. J Immunol 147: 1877-1883.<br />

Savolainen, K., Paulin, L., Westerlund-Wikström, B., Foster, T.J. et al. (2001) Expression of pls, a gene<br />

closely associated with the mecA gene of methicillin-resistant <strong>Staphylococcus</strong> <strong>aureus</strong>, prevents bacterial<br />

adhesion in vitro. Infect Immun 69: 3013-3020.<br />

Schaffer, A.C., and Lee, J.C. (2009) Staphylococcal vaccines and immunotherapies. Infect Dis Clin North Am 23:<br />

153-171.<br />

Schmaler, M., Jann, N.J., Ferracin, F., Landolt, L.Z. et al. (2009) Lipoproteins in <strong>Staphylococcus</strong> <strong>aureus</strong><br />

mediate inflammation by TLR2 and iron-dependent growth in vivo. J Immunol 182: 7110-7118.<br />

Schneewind, O., Model, P., and Fischetti, V.A. (1992) Sorting of protein A to the staphylococcal cell wall. Cell<br />

70: 267-281.<br />

Schnell, N., Entian, K.D., Schneider, U., Götz, F. et al. (1988) Prepeptide sequence of epidermin, a ribosomally<br />

synthesized antibiotic with four sulphide-rings. Nature 333: 276-278.<br />

Schwan, W.R., Langhorne, M.H., Ritchie, H.D., and Stover, C.K. (2003) Loss of hemolysin expression in<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> agr mutants correlates with selective survival during mixed infections in murine<br />

abscesses and wounds. FEMS Immunol Med Microbiol 38: 23-28.<br />

Scott, J.R., and Barnett, T.C. (2006) Surface proteins of gram-positive bacteria and how they get there. Annu Rev<br />

Microbiol 60: 397-423.<br />

Serizawa, M., Kodama, K., Yamamoto, H., Kobayashi, K. et al. (2005) Functional analysis of the YvrGHb twocomponent<br />

system of Bacillus subtilis: identification of the regulated genes by DNA microarray and northern<br />

blot analyses. Biosci Biotechnol Biochem 69: 2155-2169.<br />

Shafer, W.M., and Iandolo, J.J. (1979) Genetics of staphylococcal enterotoxin B in methicillin-resistant isolates<br />

of <strong>Staphylococcus</strong> <strong>aureus</strong>. Infect Immun 25: 902-911.<br />

Sibbald, M.J., Ziebandt, A.K., Engelmann, S., Hecker, M. et al. (2006) Mapping the pathways to<br />

staphylococcal pathogenesis by comparative secretomics. Microbiol Mol Biol Rev 70: 755-788.<br />

Sibbald, M.J.J.B., and van Dijl, J.M. (2009) Secretome mapping in Gram-positive pathogens. In Bacterial<br />

protein secretion systems. Wooldridge,K. (ed). Norwich, UK: Horizon Scientific Press.<br />

Siboo, I.R., Chaffin, D.O., Rubens, C.E., and Sullam, P.M. (2008) Characterization of the accessory Sec system<br />

of <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 190: 6188-6196.<br />

Siboo, I.R., Chambers, H.F., and Sullam, P.M. (2005) Role of SraP, a Serine-Rich Surface Protein of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong>, in binding to human platelets. Infect Immun 73: 2273-2280.<br />

Siboo, I.R., Cheung, A.L., Bayer, A.S., and Sullam, P.M. (2001) Clumping factor A mediates binding of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> to human platelets. Infect Immun 69: 3120-3127.<br />

183


Chapter 10<br />

Siegers, K., Heinzmann, S., and Entian, K.D. (1996) Biosynthesis of lantibiotic nisin. Posttranslational<br />

modification of its prepeptide occurs at a multimeric membrane-associated lanthionine synthetase complex. J<br />

Biol Chem 271: 12294-12301.<br />

Sifri, C.D., Begun, J., Ausubel, F.M., and Calderwood, S.B. (2003) Caenorhabditis elegans as a model host for<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> pathogenesis. Infect Immun 71: 2208-2217.<br />

Somerville, G.A., Beres, S.B., Fitzgerald, J.R., DeLeo, F.R. et al. (2002) In vitro serial passage of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong>: changes in physiology, virulence factor production, and agr nucleotide sequence. J<br />

Bacteriol 184: 1430-1437.<br />

Sørensen, A.L., Nagai, S., Houen, G., Andersen, P. et al. (1995) Purification and characterization of a lowmolecular-mass<br />

T-cell antigen secreted by Mycobacterium tuberculosis. Infect Immun 63: 1710-1717.<br />

Stanley, N.R., Palmer, T., and Berks, B.C. (2000) <strong>The</strong> twin arginine consensus motif of Tat signal peptides is<br />

involved in Sec-independent protein targeting in Escherichia coli. J Biol Chem 275: 11591-11596.<br />

Stapleton, M.R., Horsburgh, M.J., Hayhurst, E.J., Wright, L. et al. (2007) Characterization of IsaA and SceD,<br />

two putative lytic transglycosylases of <strong>Staphylococcus</strong> <strong>aureus</strong>. J Bacteriol 189: 7316-7325.<br />

Steen, A. Functional characterization and cell wall interactions of peptidoglycan hydrolysases of Lactococcus<br />

lactis. 25-30. 23-12-2005. University of Groningen, <strong>The</strong> Netherlands.<br />

Steen, A., Buist, G., Leenhouts, K.J., El, K.M. et al. (2003) Cell wall attachment of a widely distributed<br />

peptidoglycan binding domain is hindered by cell wall constituents. J Biol Chem 278: 23874-23881.<br />

Stein, T. (2005) Bacillus subtilis antibiotics: structures, syntheses and specific functions. Mol Microbiol 56: 845-<br />

857.<br />

Steinmetz, M., Le, C.D., and Aymerich, S. (1989) Induction of saccharolytic enzymes by sucrose in Bacillus<br />

subtilis: evidence for two partially interchangeable regulatory pathways. J Bacteriol 171: 1519-1523.<br />

Stoll, H., Dengjel, J., Nerz, C., and Götz, F. (2005) <strong>Staphylococcus</strong> <strong>aureus</strong> deficient in lipidation of<br />

prelipoproteins is attenuated in growth and immune activation. Infect Immun 73: 2411-2423.<br />

Strom, M.S., Nunn, D.N., and Lory, S. (1993) A single bifunctional enzyme, PilD, catalyzes cleavage and Nmethylation<br />

of proteins belonging to the type IV pilin family. Proc Natl Acad Sci U S A 90: 2404-2408.<br />

Stülke, J., and Hillen, W. (2000) Regulation of carbon catabolism in Bacillus species. Annu Rev Microbiol 54:<br />

849-880.<br />

Sugai, R., Takemae, K., Tokuda, H., and Nishiyama, K. (2007) Topology inversion of SecG is essential for<br />

cytosolic SecA-dependent stimulation of protein translocation. J Biol Chem 282: 29540-29548.<br />

Sukharev, S.I., Blount, P., Martinac, B., and Kung, C. (1997) Mechanosensitive channels of Escherichia coli:<br />

the MscL gene, protein, and activities. Annu Rev Physiol 59: 633-657.<br />

Sullam, P.M., Bayer, A.S., Foss, W.M., and Cheung, A.L. (1996) Diminished platelet binding in vitro by<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> is associated with reduced virulence in a rabbit model of infective endocarditis. Infect<br />

Immun 64: 4915-4921.<br />

Suree, N., Liew, C.K., Villareal, V.A., Thieu, W. et al. (2009) <strong>The</strong> structure of the <strong>Staphylococcus</strong> <strong>aureus</strong><br />

sortase-substrate complex reveals how the universally conserved LPXTG sorting signal is recognized. J Biol<br />

Chem 284: 24465-24477.<br />

Sutcliffe, I.C., and Harrington, D.J. (2002) Pattern searches for the identification of putative lipoprotein genes in<br />

Gram-positive bacterial genomes. Microbiology 148: 2065-2077.<br />

Sutcliffe, I.C., and Russell, R.R. (1995) Lipoproteins of gram-positive bacteria. J Bacteriol 177: 1123-1128.<br />

Takamatsu, D., Bensing, B.A., and Sullam, P.M. (2005) Two additional components of the accessory sec system<br />

mediating export of the Streptococcus gordonii platelet-binding protein GspB. J Bacteriol 187: 3878-3883.<br />

Tarkowski, A., Collins, L.V., Gjertsson, I., Hultgren, O.H. et al. (2001) Model systems: modeling human<br />

staphylococcal arthritis and sepsis in the mouse. Trends Microbiol 9: 321-326.<br />

Thompson, J.D., Higgins, D.G., and Gibson, T.J. (1994) CLUSTAL W: improving the sensitivity of progressive<br />

multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix<br />

choice. Nucleic Acids Res 22: 4673-4680.<br />

Tjalsma, H., Antelmann, H., Jongbloed, J.D., Braun, P.G. et al. (2004) Proteomics of protein secretion by<br />

Bacillus subtilis: separating the "secrets" of the <strong>secretome</strong>. Microbiol Mol Biol Rev 68: 207-233.<br />

Tjalsma, H., Bolhuis, A., Jongbloed, J.D., Bron, S. et al. (2000) Signal peptide-dependent protein transport in<br />

Bacillus subtilis: a genome-based survey of the <strong>secretome</strong>. Microbiol Mol Biol Rev 64: 515-547.<br />

Tjalsma, H., Bolhuis, A., van Roosmalen, M.L., Wiegert, T. et al. (1998) Functional analysis of the secretory<br />

precursor processing machinery of Bacillus subtilis: identification of a eubacterial homolog of archaeal and<br />

eukaryotic signal peptidases. Genes Dev 12: 2318-2331.<br />

Tjalsma, H., Bron, S., and van Dijl, J.M. (2003) Complementary impact of paralogous Oxa1-like proteins of<br />

Bacillus subtilis on post-translocational stages in protein secretion. J Biol Chem 278: 15622-15632.<br />

Tjalsma, H., Kontinen, V.P., Prágai, Z., Wu, H. et al. (1999) <strong>The</strong> role of lipoprotein processing by signal<br />

peptidase II in the Gram-positive eubacterium Bacillus subtilis. Signal peptidase II is required for the efficient<br />

secretion of α-amylase, a non-lipoprotein. J Biol Chem 274: 1698-1707.<br />

Tjalsma, H., Noback, M.A., Bron, S., Venema, G. et al. (1997) Bacillus subtilis contains four closely related<br />

type I signal peptidases with overlapping substrate specificities. Constitutive and temporally controlled<br />

expression of different sip genes. J Biol Chem 272: 25983-25992.<br />

Tjalsma, H., and van Dijl, J.M. (2005) Proteomics-based consensus prediction of protein retention in a bacterial<br />

membrane. Proteomics.<br />

184


Reference list<br />

Tjalsma, H., Zanen, G., Bron, S., and van Dijl, J.M. (2001) <strong>The</strong> eubacterial lipoprotein-specific (Type II) signal<br />

peptidase. In <strong>The</strong> Enzymes Volume XXII, Co- and post translational proteolysis of proteins. Dalbey,R.E., and<br />

Sigman,D.S. (eds). San Diego, CA, USA: Academic Press, pp. 3-23.<br />

Tobisch, S., Glaser, P., Krüger, S., and Hecker, M. (1997) Identification and characterization of a new βglucoside<br />

utilization system in Bacillus subtilis. J Bacteriol 179: 496-506.<br />

Tokuda, H., and Matsuyama, S. (2004) Sorting of lipoproteins to the outer membrane in E. coli. Biochim<br />

Biophys Acta 1694: IN1-IN9.<br />

Ton-That, H., Marraffini, L.A., and Schneewind, O. (2004a) Protein sorting to the cell wall envelope of Grampositive<br />

bacteria. Biochim Biophys Acta 1694: 269-278.<br />

Ton-That, H., Marraffini, L.A., and Schneewind, O. (2004b) Sortases and pilin elements involved in pilus<br />

assembly of Corynebacterium diphtheriae. Mol Microbiol 53: 251-261.<br />

Ton-That, H., Mazmanian, S.K., Alksne, L., and Schneewind, O. (2002) Anchoring of surface proteins to the<br />

cell wall of <strong>Staphylococcus</strong> <strong>aureus</strong>. Cysteine 184 and histidine 120 of sortase form a thiolate-imidazolium ion<br />

pair for catalysis. J Biol Chem 277: 7447-7452.<br />

Ton-That, H., and Schneewind, O. (2003) Assembly of pili on the surface of Corynebacterium diphtheriae. Mol<br />

Microbiol 50: 1429-1438.<br />

Tormo, M.A., Knecht, E., Götz, F., Lasa, I. et al. (2005) Bap-dependent biofilm formation by pathogenic species<br />

of <strong>Staphylococcus</strong>: evidence of horizontal gene transfer? Microbiology 151: 2465-2475.<br />

Traber, K.E., Lee, E., Benson, S., Corrigan, R. et al. (2008) agr function in clinical <strong>Staphylococcus</strong> <strong>aureus</strong><br />

isolates. Microbiology 154: 2265-2274.<br />

Tye, A.J., Siu, F.K., Leung, T.Y., and Lim, B.L. (2002) Molecular cloning and the biochemical characterization<br />

of two novel phytases from B. subtilis 168 and B. licheniformis. Appl Microbiol Biotechnol 59: 190-197.<br />

Vagner, V., Dervyn, E., and Ehrlich, S.D. (1998) A vector for systematic gene inactivation in Bacillus subtilis.<br />

Microbiology 144 ( Pt 11): 3097-3104.<br />

Valence, F., and Lortal, S. (1995) Zymogram and preliminary characterization of Lactobacillus helveticus<br />

autolysins. Appl Environ Microbiol 61: 3391-3399.<br />

van Belkum A. (2006) Staphylococcal colonization and infection: homeostasis versus disbalance of human<br />

(innate) immunity and bacterial virulence. Curr Opin Infect Dis 19: 339-344.<br />

van Belkum A., Melles, D.C., Peeters, J.K., van Leeuwen, W.B. et al. (2008) Methicillin-resistant and -<br />

susceptible <strong>Staphylococcus</strong> <strong>aureus</strong> sequence type 398 in pigs and humans. Emerg Infect Dis 14: 479-483.<br />

van der Sluis, E.O., van der Vries, E., Berrelkamp, G., Nouwen, N. et al. (2006) Topologically fixed SecG is<br />

fully functional. J Bacteriol 188: 1188-1190.<br />

van Dijl, J.M., de Jong, A., Vehmaanperä, J., Venema, G. et al. (1992) Signal peptidase I of Bacillus subtilis:<br />

patterns of conserved amino acids in prokaryotic and eukaryotic type I signal peptidases. EMBO J 11: 2819-<br />

2828.<br />

van Roosmalen, M.L., Geukens, N., Jongbloed, J.D., Tjalsma, H. et al. (2004) Type I signal peptidases of<br />

Gram-positive bacteria. Biochim Biophys Acta 1694: 279-297.<br />

van Roosmalen, M.L., Jongbloed, J.D., Dubois, J.Y., Venema, G. et al. (2001) Distinction between major and<br />

minor Bacillus signal peptidases based on phylogenetic and structural criteria. J Biol Chem 276: 25230-25235.<br />

Van Wamel, W.J., Rooijakkers, S.H., Ruyken, M., van Kessel, K.P. et al. (2006) <strong>The</strong> innate immune<br />

modulators staphylococcal complement inhibitor and chemotaxis inhibitory protein of <strong>Staphylococcus</strong> <strong>aureus</strong><br />

are located on β-hemolysin-converting bacteriophages. J Bacteriol 188: 1310-1315.<br />

Van Wely, K.H., Swaving, J., Broekhuizen, C.P., Rose, M. et al. (1999) Functional identification of the product<br />

of the Bacillus subtilis yvaL gene as a SecG homologue. J Bacteriol 181: 1786-1792.<br />

Vandenesch, F., Naimi, T., Enright, M.C., Lina, G. et al. (2003) Community-acquired methicillin-resistant<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> carrying Panton-Valentine leukocidin genes: worldwide emergence. Emerg Infect Dis<br />

9: 978-984.<br />

Veenendaal, A.K., van der Does, C., and Driessen, A.J. (2004) <strong>The</strong> protein-conducting channel SecYEG.<br />

Biochim Biophys Acta 1694: 81-95.<br />

Veith, B., Herzberg, C., Steckel, S., Feesche, J. et al. (2004) <strong>The</strong> complete genome sequence of Bacillus<br />

licheniformis DSM13, an organism with great industrial potential. J Mol Microbiol Biotechnol 7: 204-211.<br />

Venema, R., Tjalsma, H., van Dijl, J.M., de Jong, A. et al. (2003) Active lipoprotein precursors in the Grampositive<br />

eubacterium Lactococcus lactis. J Biol Chem 278: 14739-14746.<br />

Vitikainen, M., Lappalainen, I., Seppala, R., Antelmann, H. et al. (2004) Structure-function analysis of PrsA<br />

reveals roles for the parvulin-like and flanking N- and C-terminal domains in protein folding and secretion in<br />

Bacillus subtilis. J Biol Chem 279: 19302-19314.<br />

Voigt, B., Schweder, T., Becher, D., Ehrenreich, A. et al. (2004) A proteomic view of cell physiology of<br />

Bacillus licheniformis. Proteomics 4: 1465-1490.<br />

Voigt, B., Schweder, T., Sibbald, M.J., Albrecht, D. et al. (2005) <strong>The</strong> extracellular proteome of Bacillus<br />

licheniformis grown in different media and under different nutrient starvation conditions. Proteomics.<br />

Völker, U., and Hecker, M. (2005) From genomics via proteomics to cellular physiology of the Gram-positive<br />

model organism Bacillus subtilis. Cell Microbiol 7: 1077-1085.<br />

von Eiff, C., Becker, K., Machka, K., Stammer, H. et al. (2001) Nasal carriage as a source of <strong>Staphylococcus</strong><br />

<strong>aureus</strong> bacteremia. Study Group. N Engl J Med 344: 11-16.<br />

von Heijne, G. (1989) <strong>The</strong> structure of signal peptides from bacterial lipoproteins. Protein Eng 2: 531-534.<br />

von Heijne, G. (1990) <strong>The</strong> signal peptide. J Membr Biol 115: 195-201.<br />

185


Chapter 10<br />

Vrontou, E., and Economou, A. (2004) Structure and function of SecA, the preprotein translocase nanomotor.<br />

Biochim Biophys Acta 1694: 67-80.<br />

Wahome, P.G., and Setlow, P. (2008) Growth, osmotic downshock resistance and differentiation of Bacillus<br />

subtilis strains lacking mechanosensitive channels. Arch Microbiol 189: 49-58.<br />

Wardenburg, J.B., Patel, R.J., and Schneewind, O. (2007) Surface proteins and exotoxins are required for the<br />

pathogenesis of <strong>Staphylococcus</strong> <strong>aureus</strong> pneumonia. Infect Immun 75: 1040-1044.<br />

Weigel, L.M., Clewell, D.B., Gill, S.R., Clark, N.C. et al. (2003) Genetic analysis of a high-level vancomycinresistant<br />

isolate of <strong>Staphylococcus</strong> <strong>aureus</strong>. Science 302: 1569-1571.<br />

Wertheim, H.F., Vos, M.C., Ott, A., van Belkum, A. et al. (2004) Risk and outcome of nosocomial<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> bacteraemia in nasal carriers versus non-carriers. Lancet 364: 703-705.<br />

Westers, H., Dorenbos, R., van Dijl, J.M., Kabel, J. et al. (2003) Genome engineering reveals large dispensable<br />

regions in Bacillus subtilis. Mol Biol Evol 20: 2076-2090.<br />

Wetzstein, M., Völker, U., Dedio, J., Löbau, S. et al. (1992) Cloning, sequencing, and molecular analysis of the<br />

dnaK locus from Bacillus subtilis. J Bacteriol 174: 3300-3310.<br />

Wickner, W., and Schekman, R. (2005) Protein translocation across biological membranes. Science 310: 1452-<br />

1456.<br />

Widdick, D.A., Dilks, K., Chandra, G., Bottrill, A. et al. (2006) <strong>The</strong> twin-arginine translocation pathway is a<br />

major route of protein export in Streptomyces coelicolor. Proceedings of the National Academy of Sciences<br />

103: 17927-17932.<br />

Wiedemann, I., Breukink, E., van, K.C., Kuipers, O.P. et al. (2001) Specific binding of nisin to the<br />

peptidoglycan precursor lipid II combines pore formation and inhibition of cell wall biosynthesis for potent<br />

antibiotic activity. J Biol Chem 276: 1772-1779.<br />

Williams, R.C., Rees, M.L., Jacobs, M.F., Pragai, Z. et al. (2003) Production of Bacillus anthracis protective<br />

antigen is dependent on the extracellular chaperone, PrsA. J Biol Chem 278: 18056-18062.<br />

Witney, A.A., Marsden, G.L., Holden, M.T., Stabler, R.A. et al. (2005) Design, validation, and application of a<br />

seven-strain <strong>Staphylococcus</strong> <strong>aureus</strong> PCR product microarray for comparative genomics. Appl Environ<br />

Microbiol 71: 7504-7514.<br />

Witte, W., Strommenger, B., Stanek, C., and Cuny, C. (2007) Methicillin-resistant <strong>Staphylococcus</strong> <strong>aureus</strong><br />

ST398 in humans and animals, Central Europe. Emerg Infect Dis 13: 255-258.<br />

Wolf, C., Hochgräfe, F., Kusch, H., Albrecht, D. et al. (2008) Proteomic analysis of antioxidant strategies of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong>: diverse responses to different oxidants. Proteomics 8: 3139-3153.<br />

Wolff, S., Antelmann, H., Albrecht, D., Becher, D. et al. (2007) Towards the entire proteome of the model<br />

bacterium Bacillus subtilis by gel-based and gel-free approaches. J Chromatogr B Analyt Technol Biomed Life<br />

Sci 849: 129-140.<br />

Wu, H.C. (1996) Biosynthesis of lipoproteins. In Escherichia coli and Salmonella: Cellular and Molecular<br />

Biology. Neidhardt,F.C., Curtis III,R., Ingraham,J.L., Low,K.B., Magasanik,B., Reznikoff,W.S. et al. (eds).<br />

Washington, D.C., USA: American Society for Microbiology, pp. 1005-1014.<br />

Yamada, K., Sanzen, I., Ohkura, T., Okamoto, A. et al. (2007) Analysis of twin-arginine translocation pathway<br />

homologue in <strong>Staphylococcus</strong> <strong>aureus</strong>. Curr Microbiol 55: 14-19.<br />

Yen, M.R., Tseng, Y.H., Nguyen, E.H., Wu, L.F. et al. (2002) Sequence and phylogenetic analyses of the twinarginine<br />

targeting (Tat) protein export system. Arch Microbiol 177: 441-450.<br />

Yoshida, K., Kobayashi, K., Miwa, Y., Kang, C.M. et al. (2001) Combined transcriptome and proteome analysis<br />

as a powerful approach to study genes under glucose repression in Bacillus subtilis. Nucleic Acids Res 29: 683-<br />

692.<br />

Yother, J., and White, J.M. (1994) Novel surface attachment mechanism of the Streptococcus pneumoniae<br />

protein PspA. J Bacteriol 176: 2976-2985.<br />

Young, R., and Bläsi, U. (1995) Holins: form and function in bacteriophage lysis. FEMS Microbiol Rev 17: 191-<br />

205.<br />

Yuan, J., Zweers, J.C., van Dijl, J.M., and Dalbey, R.E. (2009) Protein transport across and into cell<br />

membranes in bacteria and archaea. Cell Mol Life Sci.<br />

Zhang, L., Jacobsson, K., Vasi, J., Lindberg, M. et al. (1998) A second IgG-binding protein in <strong>Staphylococcus</strong><br />

<strong>aureus</strong>. Microbiology 144 ( Pt 4): 985-991.<br />

Zhang, Y.Q., Ren, S.X., Li, H.L., Wang, Y.X. et al. (2003) Genome-based analysis of virulence genes in a nonbiofilm-forming<br />

<strong>Staphylococcus</strong> epidermidis strain (ATCC 12228). Mol Microbiol 49: 1577-1593.<br />

Ziebandt, A.K., Becher, D., Ohlsen, K., Hacker, J. et al. (2004) <strong>The</strong> influence of agr and σ B in growth phase<br />

dependent regulation of virulence factors in <strong>Staphylococcus</strong> <strong>aureus</strong>. Proteomics 4: 3034-3047.<br />

Ziebandt, A.K., Weber, H., Rudolph, J., Schmid, R. et al. (2001) Extracellular proteins of <strong>Staphylococcus</strong><br />

<strong>aureus</strong> and the role of SarA and σ B . Proteomics 1: 480-493.<br />

Zimmer, J., Nam, Y., and Rapoport, T.A. (2008) Structure of a complex of the ATPase SecA and the proteintranslocation<br />

channel. Nature 455: 936-943.<br />

Zong, Y., Bice, T.W., Ton-That, H., Schneewind, O. et al. (2004) Crystal structures of <strong>Staphylococcus</strong> <strong>aureus</strong><br />

sortase A and its substrate complex. J Biol Chem 279: 31383-31389.<br />

Zweers, J.C., Wiegert, T., and van Dijl, J.M. (2009) Stress responsive systems set specific limits to the<br />

overproduction of membrane proteins in Bacillus subtilis. Appl Environ Microbiol.<br />

186


187


“I only got seventh-grade education, but I have a doctorate in funk<br />

and I like to put that to good use”<br />

-James J. Brown-<br />

188


Chapter 11<br />

Nederlandse samenvatting voor de leek<br />

189


Chapter 11<br />

Nederlandse samenvatting voor de leek<br />

Bacteriën<br />

Bacteriën behoren tot de oudste levensvormen op aarde en ze worden beschouwd als de<br />

voorgangers van al het andere leven op aarde. De oudste bacteriën zijn zo’n 4 miljard jaar<br />

geleden ontstaan, terwijl de meest recente “moderne” bacteriesoorten zo’n 3 miljard jaar<br />

geleden zijn ontstaan. Rond deze tijd waren bacteriën en archaea, die beide waarschijnlijk<br />

ontstaan zijn uit een gezamenlijke voorouder, de dominante levensvormen op aarde. Bacteriën<br />

en archaea zijn eencellig en bevatten in deze ene cel alle informatie, die ze nodig hebben om<br />

te overleven onder de meest verschillende omstandigheden. Zo kunnen we deze organismen<br />

vinden op plaatsen waar geen enkel ander organisme zou kunnen overleven, zoals op de<br />

bodem van de oceaan (hoge druk) en nabij vulkanen in de oceaan (zeer giftig en hoge<br />

temperatuur) tot onder het ijs van antarctica (extreem lage temperatuur). De eerste eukaryote<br />

cellen (cellen waaruit alle dieren en planten bestaan) zijn waarschijnlijk ontstaan door de<br />

opname van bacteriën door de voorouders van de eukaryote cellen. Deze endosymbionten<br />

hebben zich vervolgens ontwikkeld tot organellen zoals mitochondriën (de “energiecentrales”<br />

van eukaryote cellen) en chloroplasten (de organellen die nodig zijn voor fotosynthese in<br />

groene algen en planten). De Nederlander Antoni van Leeuwenhoek was in 1676 de eerste die<br />

met behulp van een miscroscoop bacteriën kon waarnemen.<br />

Bacteriën zijn afgesloten compartimenten (cellen), die bestaan uit een waterige oplossing<br />

(cytoplasma), waarin het DNA, eiwitten en alle andere componenten aanwezig zijn, die nodig<br />

zijn voor de bacterie om te kunnen groeien en zich te vermenigvuldigen, ofwel om te leven.<br />

Het cytoplasma is omgeven door een membraan, die het cytoplasma volledig scheidt van de<br />

buitenwereld. Om deze cytoplasmamembraan zit meestal een stevige celwand en, bij<br />

sommige bacteriesoorten een tweede membraan, die de bacterie weerbaar maken tegen<br />

fysieke stress. Er zijn vele verschillende vormen van bacteriën bekend, maar de meest<br />

voorkomende vormen zijn de cocci (bolletjes) en bacilli (staafjes). Verder wordt er vaak nog<br />

onderscheid gemaakt tussen Gram-positieve bacteriën (bacteriën met slechts één membraan<br />

en een dikke celwand) en de Gram-negatieve bacteriën (bacteriën met een dubbele membraan<br />

waartussen een marginale celwand ligt).<br />

Bacteriën hebben een grote invloed op het welbevinden van de mens. Aangenomen wordt dat<br />

er tussen de 500 en 1000 verschillende soorten bacteriën in het menselijke darmkanaal leven<br />

en ongeveer een even groot aantal op de huid. Onder normale omstandigheden vormen deze<br />

bacteriën geen bedreiging voor ons en een aantal soorten zijn zelfs heel nuttig. Denk hierbij<br />

bijvoorbeeld aan bacteriën in ons darmkanaal die vitamines maken (Vitamine K, biotine) of<br />

helpen zetmeel en andere vezels af te breken (wat de mens uit zichzelf niet goed kan). Ook<br />

gebruiken we al eeuwen bacteriën bij het bereiden van verschillende voedingsmiddelen (vaak<br />

Lactobacillus en Lactococcus), zoals kaas, yoghurt en worst. Daarnaast zijn er helaas ook een<br />

groot aantal bacteriën, die schadelijk kunnen zijn. Ze kunnen ziektes veroorzaken die in<br />

sommige gevallen uiteindelijk tot de dood van hun gastheer kunnen leiden. Enkele<br />

voorbeelden hiervan zijn de bacteriën die tetanus (Clostridium tetani), difterie<br />

(Corynebacterium diphtheriae), syfilis (Treponema pallidum), cholera (Vibrio cholerae),<br />

lepra (Mycobacterium leprae) of tuberculose (Mycobacterium tuberculosis) veroorzaken.<br />

Sommige voor de mens gevaarlijke bacteriën (pathogenen) maken deel uit van de normale<br />

microbiële flora (microbiota) van de mens. Zo zitten <strong>Staphylococcus</strong> epidermidis en<br />

190


Nederlandse samenvatting<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> bijvoorbeeld op de huid en in de neus waar deze organismen normaal<br />

gesproken geen problemen veroorzaken, terwijl ze toch ook vele onaangename of zelfs<br />

levensbedreigende infecties kunnen veroorzaken. In dit proefschrift is onderzocht over welke<br />

ziekmakende factoren bacteriën zoals S. <strong>aureus</strong> kunnen beschikken. Daarnaast is de<br />

verworven kennis ook gebruikt om belangrijke eigenschappen van bacteriën zoals Bacillus<br />

subtilis en Bacillus licheniformis te bestuderen. Dit laatste was interessant omdat deze twee<br />

bacilli veel gebruikt worden voor de productie van nuttige en waardevolle eiwitten in de<br />

biotechnologische industrie.<br />

DNA en Eiwitten<br />

Al het leven op aarde is mogelijk door het bestaan van DNA. Alle informatie die nodig is om<br />

te (over)leven is opgeslagen in dit molecuul. DNA is opgebouwd uit vier verschillende<br />

bouwstenen, die we nucleotiden noemen. Deze vier zijn Adenine (A), Guanine (G), Thymine<br />

(T) and Cytosine (C) en de combinatie van deze vier nucleotiden vormt de juiste informatie.<br />

Genen zijn opgebouwd uit deze vier bouwstenen en elk gen heeft een unieke combinatie van<br />

deze bouwstenen. Chromosomen zijn DNA strengen die vele genen bevatten. De meeste<br />

bacteriën hebben maar één chromosoom. DNA kan naast het chromosoom op andere<br />

manieren meegedragen worden, bijvoorbeeld in de vorm van plasmiden. Dit zijn meestal<br />

ronde DNA moleculen waar extra informatie op kan zitten, die voordelig voor een bacterie<br />

zou kunnen zijn. In een heel enkel geval kunnen deze plasmiden (of gedeeltes daarvan) ook<br />

worden ovegebracht naar een andere bacteriesoort. De genen die zijn opgeslagen in deze<br />

chromosomen en plasmiden coderen voor eiwitten. Deze eiwitten zijn vaak enzymen, die een<br />

bepaalde chemische reactie versnellen. Elk eiwit heeft zo zijn eigen specifieke functie. Zo zijn<br />

er enzymen die DNA kunnen kopiëren of kunnen afbreken, enzymen die schadelijke stoffen<br />

kunnen omzetten in ongevaarlijke stoffen, en enzymen die kleine, maar vaak cruciale<br />

veranderingen aan andere eiwitten aanbrengen. Sommige eiwitten zijn alleen nodig voor het<br />

organisme om bepaalde processen in de cel te kunnen uitvoeren, terwijl andere eiwitten ter<br />

verdediging zijn of juist worden gebruikt om de omliggende cellen, zowel prokaryoot als<br />

eukaryoot, aan te vallen. Voor een goede werking van eiwitten is de juiste vorm (conformatie)<br />

nodig en daarvoor moeten ze op de juiste manier worden opgevouwen. Eiwitten worden in het<br />

cytoplasma gesynthetiseerd om vervolgens naar de juiste plek te worden gebracht om hun<br />

functie te kunnen vervullen. Zo kunnen eiwitten naar verschillende locaties worden geleid<br />

waar hun functie nodig is. Eiwitten die buiten de cel hun functie moeten vervullen worden<br />

met een signaalpeptide gesynthetiseerd. Dit signaalpeptide zorgt ervoor dat het eiwit naar een<br />

bepaald transportsysteem wordt geleid en vervolgens door het membraan naar het<br />

celoppervlak of buiten de cel wordt gebracht. Tijdens of kort na de passage van het membraan<br />

wordt het signaal peptide afgeknipt en het eiwit in de juiste vorm gevouwen. Sommige<br />

eiwitten zijn juist nodig aan het opervlak van de cel en deze eiwitten kunnen op verschillende<br />

manieren aan de cel gehecht worden. Ook hiervoor zijn er bepaalde signalen in het eiwit<br />

aanwezig, die ervoor zorgen dat ze aan het oppervlak van de cel blijven vastzitten en daar hun<br />

functie kunnen vervullen. Wanneer er helemaal geen signalen zijn om aan het oppervlak te<br />

hechten, zullen de eiwitten vrij in de omgeving van de bacterie terecht komen.<br />

191


Chapter 11<br />

Inleiding – Hoofdstuk 1<br />

S. <strong>aureus</strong> is een bacterie, die als commensaal bij de mens voorkomt. Deze bacterie komt men<br />

vaak tegen op de slijmvliezen in de neus en op andere vochtige plekken in het lichaam. Bij<br />

30-40% van de populatie kan men S. <strong>aureus</strong> aantreffen. Een deel van deze groep draagt S.<br />

<strong>aureus</strong> continu bij zich, terwijl het andere deel deze bacterie ook weer spontaan kwijtraakt.<br />

De meeste mensen die S. <strong>aureus</strong> bij zich dragen zullen hier nooit last van hebben. Echter,<br />

wanneer S. <strong>aureus</strong> in de bloedbaan terecht komt, dan kan dit leiden tot ernstige infecties.<br />

Infecties worden onder andere opgelopen in het ziekenhuis wanneer invasieve medische<br />

hulpmiddelen zoals katheters en implantaten gebruikt worden. Samen met S. epidermidis is S.<br />

<strong>aureus</strong> berucht om het vermogen zich op oppervlakken van katheters en implantaten te<br />

hechten en op de plaatsen in het lichaam waar deze hulpmiddelen aangebracht zijn infecties te<br />

veroorzaken. Een van de grote problemen bij het bestrijden van S. <strong>aureus</strong> is het vermogen van<br />

deze bacterie om resistent te worden tegen antibiotica. Al vrij snel na de eerste inzet van<br />

penicilline als antibioticum tegen S. <strong>aureus</strong> infecties, werden penicilline-resistente S. <strong>aureus</strong><br />

stammen geïsoleerd. Sinds een aantal jaren wordt vancomycine gebruikt als laatste redmiddel,<br />

maar vancomycine-resistente stammen zijn nu ook al gevonden. Een goed voorbeeld hiervan<br />

zijn de S. <strong>aureus</strong> JH1 en JH9 stammen waarvan de genomen gesequenced zijn. De<br />

vancomycine-sensitieve S. <strong>aureus</strong> stam JH1 werd geïsoleerd uit een patient die behandeld<br />

werd met vancomycine. Een aantal dagen later werd de vancomycine-intermediair sensitieve<br />

S. <strong>aureus</strong> stam JH9 uit dezelfde patiënt geïsoleerd. JH9 stamt af van de JH1 stam, waaruit<br />

duidelijk blijkt hoe snel S. <strong>aureus</strong> ongevoelig kan worden voor antibiotica. Het vermogen van<br />

S. <strong>aureus</strong> om bijna alle weefsels en organen van de mens te infecteren is gebaseerd op het feit<br />

dat deze bacterie een heel arsenaal aan virulentiefactoren kan produceren. De meeste van deze<br />

virulentiefactoren zijn eiwitten die in het cytoplasma van de cel gemaakt worden, om<br />

vervolgens over het membraan te worden getransporteerd naar het oppervlak of het externe<br />

milieu van de bacterie.<br />

Elke bacterie heeft transportsystemen die bedoeld zijn om eiwitten over het membraan te<br />

transporteren. In dit proeschrift wordt ingegaan op de verschillende eiwittransportsystemen<br />

die aanwezig zijn in S. <strong>aureus</strong> en S. epidermidis. Door te onderzoeken welke<br />

transportsystemen aanwezig kunnen zijn en wat voor eiwitten via deze systemen naar buiten<br />

worden getransporteerd kan inzicht verkregen worden in het belang van deze systemen voor<br />

de virulentie van S. <strong>aureus</strong> en S. epidermidis. Door de individuele componenten van elk<br />

systeem te bestuderen, kan tevens inzicht worden verkregen in het belang van deze<br />

componenten voor het transport van specifieke eiwitten. De verkregen kennis kan in de<br />

toekomst wellicht gebruikt worden voor de ontwikkeling van nieuwe diagnostische of<br />

therapeutische middelen om <strong>Staphylococcus</strong> infecties te voorkomen of te behandelen.<br />

Hoofdstuk 2<br />

In Hoofdstuk 2 wordt gebruik gemaakt van de genoomsequenties van verschillende S. <strong>aureus</strong><br />

isolaten om te onderzoeken welke (potentiële) eiwittransportsystemen aanwezig zijn en om<br />

een voorspelling te doen hoeveel eiwitten er worden gesecreteerd via de aanwezige<br />

transportsystemen. Alle eiwitten die worden getransporteerd over het membraan, worden<br />

gesynthetiseerd met een signaalpeptide. Met behulp van programma’s die zoeken naar<br />

verschillende signaalpeptidemotieven werd een voorspelling gedaan hoeveel eiwitten er<br />

mogelijk over het membraan getransporteerd kunnen worden. In de veertien S. <strong>aureus</strong><br />

192


Nederlandse samenvatting<br />

genomen die inmiddels zijn gesequenced werden zes verschillende eiwittransportsystemen<br />

gevonden, waarvan het meest gebruikte transport systeem het Sec-systeem is. Dit systeem<br />

transporteert ongevouwen eiwitten via het SecYEG kanaal over het membraan. Het Tat<br />

transportsysteem kan gevouwen eiwitten transporteren, maar er zijn veel minder eiwitten<br />

gevonden die mogelijk via dit systeem worden getransporteerd. Daarnaast zijn er nog een<br />

aantal specifieke transportsystemen (Ess, Com, holins en ABC transporters) die maar enkele<br />

eiwitten transporteren. In totaal zijn ~150 eiwitten gevonden, die via deze systemen over het<br />

membraan worden getransporteerd en vervolgens vastgehouden worden in de celwand of<br />

losgelaten in het externe milieu van de bacterie. Dit betreft daarom allemaal potentiële<br />

virulentiefactoren. Uit een vergelijking van de datasets die voor de verschillende S. <strong>aureus</strong><br />

stammen verkregen zijn blijkt dat 117 eiwitten in alle stammen voorkomen. Deze eiwitten<br />

vormen samen het zogenaamde kern-exoproteoom. Veel van de eiwitten die covalent aan de<br />

celwand worden gekoppeld vallen onder deze categorie. De eiwitten uit het kern-exoproteoom<br />

lijken belangrijke “huishoud”-functies te vervullen. Het betreft bijvoorbeeld proteases,<br />

fibrinogeen- en IgG-bindende eiwitten die S. <strong>aureus</strong> nodig heeft om zich in de gastheer te<br />

handhaven en het immuunsysteem te omzeilen. De eiwitten die niet in alle stammen aanwezig<br />

zijn vormen samen het zogenaamde variabele exoproteoom. Deze groep van eiwitten omvat<br />

voornamelijk de toxines en andere eiwitten die nodig zijn om nieuwe plekken in het menselijk<br />

lichaam te veroveren. Deze eiwitten zijn vaak de veroorzakers van ziekteverschijnselen.<br />

Hoofdstuk 3<br />

Ten tijde van het schrijven van dit proefschrift waren de genomen van dertien humane S.<br />

<strong>aureus</strong> isolaten gesequenced. Doordat de genomen van deze isolaten in grote mate<br />

overeenkomen zou men verwachten dat dit ook het geval is voor de virulentiefactoren die<br />

door de verschillende S. <strong>aureus</strong> stammen geproduceerd worden. Dat dit niet altijd het geval is,<br />

wordt zichtbaar gemaakt in Hoofdstuk 3. Hier zijn de exoproteomen van 25 klinische isolaten<br />

uit het Universitair Medisch Centrum Groningen onderzocht. Deze isolaten zijn geïsoleerd uit<br />

verschillende patiënten met verschillende soorten infecties. In deze analyses werden 63<br />

verschillende extracellulaire eiwitten geïdentificeerd, waarvan er slechts zeven door alle<br />

isolaten worden gesecreteerd. Hieruit blijkt dat er een enorme heterogeniteit bestaat in de<br />

gesecreteerde eiwitten van deze ene soort bacteriën. Deze heterogeniteit wordt veroorzaakt<br />

door verschillen in de genomen van de verschillende S. <strong>aureus</strong> isolaten en door variaties in de<br />

expressie van de gesecreteerde eiwitten. De verworven kennis over het constante<br />

exoproteoom is van groot belang voor de ontwikkeling van een geschikt vaccin ter<br />

voorkoming van (liefst alle) S. <strong>aureus</strong> infecties. Dit onderzoek dient bij voorkeur in de<br />

toekomst uitgebreid te worden tot de eiwitten die bij alle S. <strong>aureus</strong> stammen aan het<br />

celoppervlak gehecht zijn.<br />

Hoofdstuk 4<br />

De virulentiefactoren van S. <strong>aureus</strong> worden in het cytoplasma gesynthetiseerd en vervolgens<br />

over het membraan getransloceerd. Daarna worden deze eiwitten ofwel met een<br />

retentiesignaal in het membraan of de celwand vastgehouden ofwel, als retentiesignalen<br />

ontbreken, gesecreteerd in het externe milieu. Bij het proces van eiwittranslocatie en retentie<br />

zijn verschillende eiwitten specifiek betrokken. In Hoofdstuk 4 worden de effecten van het<br />

193


Chapter 11<br />

verwijderen van deze componenten van eiwittransportystemen op de secretie van eiwitten<br />

beschreven. Het betreft hierbij componenten van de Sec-, Com-, Tat- en holin-systemen. Deze<br />

componenten worden verwijderd door de desbetreffende genen van het chromosoom van S.<br />

<strong>aureus</strong> te deleteren hetgeen leidt tot mutante S. <strong>aureus</strong> stammen met defecten in de<br />

machinerie voor eiwittransport. De verwijderde componenten die bij het Sec-systeem<br />

betrokken zijn vervullen verschillende rollen tijdens de verschillende stadia van het<br />

eiwittransportproces. SecG en SecY2 zijn betrokken bij de vorming van transportkanalen in<br />

het membraan waardoor eiwitten het membraan kunnen passeren, Lgt en LspA zijn betrokken<br />

bij het modificeren en processen van eiwitten met een vetzuurmodificatie (lipoproteinen) voor<br />

hun retentie aan de buitenkant van het membraan, DsbA en PrsA hebben een functie bij het<br />

goed vouwen van getransporteerde eiwitten en SrtA en SrtB verankeren bepaalde eiwitten<br />

covalent aan de celwand. Uit de analyses van de exoproteomen van de geconstrueerde<br />

mutanten blijkt, dat het merendeel weinig invloed heeft op de secretie van eiwitten. De deletie<br />

van secG heeft een sterke invloed op eiwitsecretie en dit wordt in meer detail beschreven in<br />

Hoofdstuk 5. Ook de lgt mutant vertoont een sterk veranderd exoproteoom hetgeen te maken<br />

heeft met het feit dat een aantal lipoproteinen niet meer goed aan het membraan hechten en<br />

daardoor in het exoproteoom van de mutante cellen terecht komen.<br />

Hoofdstuk 5<br />

De meeste eiwitten (en virulentiefactoren) worden via het Sec-systeem gesecreteerd. De kern<br />

van het Sec-systeem bestaat uit de translocatiemotor SecA en het kanaal dat door de SecY,<br />

SecE en SecG eiwitten wordt gevormd. Het is bekend dat de genen voor SecY en SecA<br />

essentieel zijn voor de bacterie en daardoor niet kunnen worden uitgeschakeld. S. <strong>aureus</strong> en<br />

een aantal andere Gram-positieve pathogenen, waaronder S. epidermidis, Bacillus anthracis<br />

en Listeria monocytogenes, hebben nog een tweede set secA en secY genen. De SecA2 en<br />

SecY2 eiwitten lijken een zeer beperkt aantal specifieke eiwitten te transporteren. Voor S.<br />

<strong>aureus</strong> is in de vakliteratuur gesuggereerd dat zelfs maar één eiwit SecA2/SecY2-afhankelijk<br />

wordt gesecreteerd. In Hoofdstuk 5 wordt beschreven wat de gevolgen zijn van de deletie<br />

van de niet-essentiële secG en secY2 genen op het transport van eiwitten naar het externe<br />

milieu van S. <strong>aureus</strong> en naar de celwand van dit organisme. De extracellulaire accumulatie<br />

van acht gesecreteerde eiwitten en zeven celwand eiwitten blijkt aanzienlijk te zijn veranderd<br />

in de secG mutant. Een aantal van deze eiwitten zijn bekende en belangrijke<br />

virulentiefactoren, zoals Spa, IsaA en Hla. Een eiwit, waarvan het transport naar de celwand<br />

zeer sterk is verminderd is het antilichaam-bindende eiwit Sbi. De secY2 mutant vertoonde<br />

geen veranderingen in het exoproteoom. Echter, de secretie van twee extra eiwitten was<br />

veranderd in een dubbelmutant waarin de secY2 mutatie gecombineerd was met de secG<br />

mutatie. Daarnaast vertoonde de secG secY2 dubbelmutant een duidelijk groeidefect wat erop<br />

duidt dat SecY2 interacties aan kan gaan met het reguliere Sec kanaal.<br />

Hoofdstuk 6<br />

Een aantal eiwitten in S. <strong>aureus</strong> vervullen hun functie aan het oppervlak van de cel. Deze<br />

eiwitten zijn bijvoorbeeld nodig voor de hechting van bacteriecellen aan weefsels van de<br />

gastheer of aan elkaar. Ook kunnen dergelijke geëxponeerde eiwitten op een dusdanige<br />

manier antilichamen binden, dat de bacteriecel niet meer door het menselijke immuunsysteem<br />

194


Nederlandse samenvatting<br />

als lichaamsvreemd herkend wordt. De enzymen die voor de covalente koppeling van eiwitten<br />

aan de celwand zorgen heten sortases. Ze zijn ingedeeld in vier verschillende klassen: sortase<br />

A-D. S. <strong>aureus</strong> heeft twee sortases: SrtA dat meerdere verschillende eiwitten met een LPxTG<br />

of LPxAG motief aan de celwand bindt, en SrtB dat slechts één substraat met een NPQTN<br />

motief aan de celwand bindt. Afhankelijk van de stam zijn er in S. epidermidis één of twee<br />

sortases te vinden: SrtA dat dezelfde functie heeft als SrtA in S. <strong>aureus</strong> en SrtC. Op basis van<br />

genomische informatie werd aangenomen dat SrtC ook eiwitten met een LPxTG motief<br />

herkent, maar de exacte activiteit van dit eiwit was nog niet onderzocht. Hoofdstuk 6<br />

beschrijft onderzoek naar de functies van SrtA en SrtC. Allereerst werden de effecten van<br />

srtA mutaties op de binding van LPxTG-eiwitten aan de celwand in S. <strong>aureus</strong> en S.<br />

epidermidis onderzocht. Het blijkt dat een aantal eiwitten zoals ClfA, SasG, SdrC, SdrD en<br />

Spa in de S. <strong>aureus</strong> srtA mutant voor een groot deel in het externe milieu terechtkomen.<br />

Echter, een substantieel deel van deze eiwitten blijft ook aan de celwand gebonden via nietcovalente<br />

interacties. Expressie vanaf een plasmide van S. <strong>aureus</strong> of S. epidermidis SrtA<br />

resulteerde in herstel van het fenotype van de niet-mutante (wild-type) cellen wat betreft de<br />

locatie van eerder genoemde eiwitten. Ook expressie vanaf een plasmide van SrtC van S.<br />

epidermidis zorgde ervoor dat een aantal eiwitten, te weten ClfA, SasG en Aap, weer covalent<br />

aan de celwand werden gebonden, net als in wild-type cellen. Dit betekent dat SrtA en SrtC<br />

een gedeeltelijk overlappende substraatspecificiteit hebben. Daarnaast werd de betrokkenheid<br />

van sortases bij de vorming van een biofilm onderzocht. Het vermogen van de S. <strong>aureus</strong> srtA<br />

mutant om een biofilm te vormen was beduidend minder dan dat van wildtype cellen. Deletie<br />

van srtA in S. epidermidis had daarentegen geen gevolg voor de vorming van een biofilm. Dit<br />

verschil kan gerelateerd worden aan de expressieniveaus van een beperkt aantal specifieke<br />

eiwitten, die betrokken zijn bij biofilm vorming, zoals Spa en SasG in S. <strong>aureus</strong> en Aap in S.<br />

epidermidis. Een interessante waarneming was dat de activiteit van SrtA bepalend bleek te<br />

zijn voor de vorming van biofilms door S. <strong>aureus</strong> en S. epidermidis. De activiteit van SrtC<br />

had daarentegen een remmende werking op de vorming van een biofilm door S. epidermidis.<br />

Hoofdstuk 7<br />

Het eiwit sublancin 168, dat van nature door de B. subtilis stam 168 geproduceerd wordt,<br />

vertoont antibacteriële activiteiten tegen een groot aantal (pathogene) Gram-positieve<br />

bacteriën, zoals S. <strong>aureus</strong>, Streptococcus pyogenes en Bacillus cereus. Hoofdstuk 7 beschrijft<br />

een studie naar de activiteit van sublancin 168 onder verschillende groeicondities. Verhoging<br />

van de zoutconcentratie in het groeimedium zorgde voor een verlaging van de gevoeligheid<br />

van cellen voor sublancin 168, maar had geen effect op de productie, activiteit of stabiliteit<br />

van dit antimicrobiële eiwit. Tevens bleek dat het eiwit MscL, dat mechanosensitieve kanalen<br />

in de cytoplasmamembraan vormt, een bepalende rol heeft in de gevoeligheid van de cellen<br />

voor sublancin 168. De resultaten van dit onderzoek suggereren dat MscL ofwel het<br />

aangrijpingspunt voor sublancin 168 is, ofwel een porie in het membraan die sublancin 168<br />

toegang geeft tot een vooralsnog onbekend aangrijpingspunt binnen in de cel.<br />

Hoofdstuk 8<br />

Één van de bacteriën die in de biotechnologische industrie gebruikt worden voor de productie<br />

van gesecreteerde eiwitten is Bacillus licheniformis. Al lang wordt van deze bacterie gebruik<br />

195


Chapter 11<br />

gemaakt, maar er was nog nooit onderzocht welke eiwitten er nu precies worden gesecreteerd.<br />

Met het beschikbaar komen van de genoomsequentie van B. licheniformis DSM 13 werd het<br />

mogelijk om te voorspellen welke eiwitten in het medium terecht zouden komen. Hoofdstuk<br />

8 beschrijft onderzoek naar de transportsystemen die aanwezig zijn in B. licheniformis. Met<br />

behulp van de genoomsequentie werden voorspellingen gedaan over de eiwitten die een<br />

signaal peptide hebben en via welke transportsystemen deze eiwitten naar buiten worden<br />

getransporteerd. In totaal werden 298 eiwitten met signaalpeptiden voorspeld, waarvan de<br />

meeste naar het Sec-systeem worden gestuurd voor secretie. Met behulp van 2-D<br />

gelelectroforese werd het exoproteoom van B. licheniformis bestudeerd. Dit heeft geleid tot<br />

identificatie van 89 daadwerkelijk gesecreteerde eiwitten, die gesynthetiseerd worden met een<br />

duidelijk herkenbaar signaalpeptide. Tot deze groep van eiwitten behoren enzymen, die nodig<br />

zijn voor de afbraak van macromoleculen, enzymen die nodig zijn voor verandering en<br />

afbraak van de celwand, bacteriofaag- en flagel-gerelateerde eiwitten en meerdere eiwitten<br />

waarvan de functie nog niet bekend is. Onder verschillende groeicondities worden<br />

verschillende eiwitten in het exoproteoom gevonden. Zo worden er meer verschillende en<br />

grotere hoeveelheden extracellulaire eiwitten gevonden wanneer de bacterie in een rijk en<br />

complex medium wordt gekweekt dan wanneer de bacterie in een minimaal medium wordt<br />

gekweekt.<br />

Hoofdstuk 9<br />

In Hoofdstuk 9 worden de onderwerpen die in dit proefschrift zijn beschreven besproken en<br />

in een breder perspectief geplaatst.<br />

196


Appendix I<br />

Dankwoord<br />

197


Appendix I<br />

Dankwoord<br />

En nu je eindelijk het hele proefschrift hebt doorgelezen (yeah, right!!!), het meest gelezen<br />

“Dankwoord”. Laat ik eerst maar beginnen met het bedanken van mijn ouders, zodat ze met trots<br />

bovenaan staan; en terecht wat mij betreft.<br />

Pap en mam, ik wil jullie hier in de eerste plaats heel erg bedanken voor de steun in de afgelopen vijf<br />

jaar. Het ging in de eerste twee jaar fantastisch goed en jullie hebben gedurende mijn AIO-tijd altijd<br />

interesse getoond naar de vorderingen op en naast het werk en volgens mij zijn jullie een van de<br />

weinige ouders die (doordeweeks) op het lab hebben rondgekeken. Ook (of misschien wel juist) tijdens<br />

en na de persoonlijke tegenslag met het verlies van Regina in het laatste jaar, heb ik ontzettend veel<br />

steun aan jullie gehad. Daarnaast wil ik jullie bedanken dat jullie me altijd vrij hebben gelaten in mijn<br />

keuzes (soms met een klein beetje bijsturen).<br />

Pap, ondanks dat je dit niet meer mee mocht maken, weet ik zeker dat je nu apetrots zou zijn geweest.<br />

Je bent in een aantal opzichten een groot voorbeeld geweest, waar ik tegenop kon kijken. Je interesse<br />

voor muziek heb je aan mij doorgegeven, en muziek is altijd een goede uitlaatklep voor mij geweest.<br />

Dat heeft me zeker door de vele AIO-dipjes geholpen. Bedankt voor de “unconditional love” and het<br />

feit dat je zo blij met mijn Eleni was. We vinden het erg jammer dat jij en Regina er volgend jaar niet<br />

bij kunnen zijn op onze trouwerij in Griekenland. Ik mis je.<br />

Mam, voor mij ben je altijd de sterkste vrouw geweest die ik ken. Ik ben altijd een mama’s kindje<br />

geweest en ik weet dat ik altijd naar jou kon gaan om over problemen te praten (ook al deed ik dat erg<br />

weinig). In 2007 en 2008 werd (te) veel van jou en ons gevraagd, maar ondanks dit heb jij je staande<br />

weten te houden. Hierdoor is mijn respect voor jou, die al groot was, nog groter geworden. Ook hier<br />

komen we doorheen! Bedankt voor al die bezoekjes in Groningen samen met Pap, Elvira en Lisa,<br />

Tamara en Patrick, of Regina en Feiko. Samen met Pap heb je de basis gelegd voor mijn interesse in<br />

muziek. Ook dank voor de “muzikale opvoeding” thuis en in de auto naar Brabant, waar ik met veel<br />

plezier aan terug denk.<br />

Mijn drie zusjes Elvira, Tamara en Regina wil ik bedanken voor de jaren thuis op de Skarren in<br />

Bolsward en de jaren daarna in Groningen/Sneek/Leeuwarden. Ik ben altijd trots op alledrie geweest en<br />

ik hoop dat jullie nu ook trots op mij kunnen zijn met het behalen van de Doctors titel.<br />

Lieve Elvira (Zus 1), vroeger waren we al twee handen op één buik, maar daar is volgens mij helemaal<br />

niks mis mee. Juist daarom weet ik dat je er altijd voor me zult zijn wanneer dat nodig is en daar heb ik<br />

de laatste 2 jaar heel veel aan gehad. Ook de komst van jullie dochter Lisa en het mogen oppassen op<br />

haar heeft mij door het laatste jaar geholpen. Elvira en Jan, heel veel plezier met jullie kleine aliens en<br />

stuur ze maar langs bij oom Ma’k en tante ‘Nena als ze vervelend beginnen te worden. Ik zal het even<br />

langs wippen voor een lekker hapje eten of gewoon een praatje bij jullie in het restaurant missen.<br />

Lieve Tamara (Zus 2), heel erg bedankt voor het meehelpen opknappen van het huisje op de<br />

Goudsbloemstraat. Samen hebben we er toch iets heel moois van gemaakt (al heb ik de trap nog een<br />

keer moeten verven) en ik heb er de afgelopen jaren met veel plezier gewoond. Ook jouw steun in de<br />

laatste 2 jaar is heel belangrijk voor mij geweest. Veel plezier samen met Patrick in jullie nieuwe<br />

huisje!<br />

Lieve Regina (Zus 3), de avonden met Feiko bij mij of bij jullie waren voor mij altijd dikke pret.<br />

Lekker lullen over niks tijdens een spelletje doen of een filmpje kijken. Bedankt dat ik bij jou helemaal<br />

en totaal mezelf kon zijn. Je hebt de betekenis van “Carpe Diem” wel een beetje op een rare manier<br />

duidelijk gemaakt. Ik mis je.<br />

Hoe vaak ik er ook om gezeurd heb, een broertje heb ik nooit gekregen. Gelukkig was er die<br />

buurjongen van nummer 8 die een aantal jaar ouder is en gemakkelijk ingezet kon worden als “grote<br />

broer”. Marco, hartelijk dank voor de vele uurtjes ontspanning in de vorm van een filmpje of een potje<br />

Mario Kart of Burnout. Het heeft misschien wel wat schrijftijd gekost, maar het heeft ook veel uurtjes<br />

plezier (bij winst) en mateloze irritatie (bij verlies) opgeleverd. Ook de vakanties waren van<br />

198


Dankwoord<br />

onschatbare waarde (Montpellier, Italië en 2x USA) en ik heb daar veel mooie foto’s en een<br />

fantastische tattoo van een “Injun” aan overgehouden. De potjes MarioKart moeten we nu dan maar<br />

online doen of de keren dat ik in Nederland ben, en als je goed oefent kun je misschien net zo goed<br />

worden als ik (………).<br />

Jan Maarten, promoter, full-time optimist en vertrouwenspersoon. De eerste keer dat we elkaar<br />

ontmoetten was tijdens een kennismakingsgesprek bij jou in je kamer toen je nog bij Farmaceutische<br />

Biologie werkte. Ik wilde toen als vrijwilliger aan de slag om mijn labvaardigheden een beetje op peil<br />

te houden. Wat me toen al opviel, was dat je altijd open staat voor dingen. Toen je vertelde dat je een<br />

eigen groep ging beginnen, wilde ik graag meegaan. Je hebt me gelukkig niet heel lang hoeven laten<br />

wachten, want je belde al vrij snel dat ik bij je kon beginnen. Door mij zo snel mogelijk heel veel<br />

dingen te laten doen, heb je meer uit me gehaald dan ik zelf voor mogelijk hield. Het presenteren, waar<br />

ik altijd een hekel aan heb gehad, werd minder vervelend. De vele congressen waar je ons naar<br />

toestuurde, waren altijd nuttig, heel erg gezellig en natuurlijk vol met mooie uitstapjes (met als<br />

hoogtepunten Kreta en Australië). Ook heel erg bedankt voor het feit je altijd wilde luisteren en klaar<br />

stond op de momenten wanneer dat nodig was (half twee ’s nachts) en voor de gesprekken die we<br />

hebben gehad in de tijden dat ik het moeilijk had. Je altijd positieve instelling heeft mij veel geholpen<br />

wanneer het even wat minder ging. Ook ben ik heel blij dat ik je heb kunnen overtuigen dat Eleni een<br />

goede AIO zou zijn en prima bij ons zou kunnen werken. Succes met de groep en wie weet, komen we<br />

elkaar in de toekomst weer tegen.<br />

I would like to thank the reading committee, Tarek Msadek, Wim Quax and Arnold Driessen for<br />

reading this thesis and for their critical comments. A special thanks goes out to Tarek for his<br />

enthusiasm at my first presentation at the BACELL meeting at the Pasteur Institute in 2005 in Paris.<br />

Also many thanks for the possibility to stay for a month and play with worms in your lab. Too bad we<br />

never got to play some music together, but maybe we can arrange something for the party!!<br />

Also many thanks to the people who have helped one way or another to make this thesis as it is now.<br />

Anne, Birgit, Haike, Susanne, <strong>The</strong>resa, Vanessa, and Michael, thank you very much for the pictures of<br />

the 2D gels and the nice discussions we had on meetings or via mail. Anne en Harold, hartelijk dank<br />

voor het begin met de predicties. Dit hoofdstuk heeft een mooie basis gelegd voor de rest van het<br />

proefschrift. Dörte also thanks for your contribution to Chapter 6 and the identification of the many<br />

protein samples I have sent to you. Ietse, bedankt voor de mooie EM-plaatjes van de <strong>Staphylococcus</strong><br />

<strong>aureus</strong> stammen. To the people from Tarek’s group, and especially Olivier: thank you for your warm<br />

welcome in Paris and introducing me to the worms. Many thanks to the BACELL-HEALTH,<br />

StaphDynamics, and <strong>TI</strong><strong>Pharma</strong> AntiStaph communities for giving me the chance to present my work<br />

and the many helpful discussions.<br />

Beste paranimfen, Thijs en Monika; de afgelopen 5 jaar (en met name de laatste twee jaar) heb ik<br />

ontzettend veel steun aan jullie gehad.<br />

Thijs, als collega, goede vriend en metalhead heb ik je beter leren kennen. Op het lab en tijdens de<br />

werkbesprekingen had je vaak wel antwoord op mijn vragen en/of respons op wat ik te vertellen had. Je<br />

bent als vriend een enorme steun geweest na het overlijden van Regina en de zware tijden die daarop<br />

volgden. Altijd klaar om ’s avonds even langs te komen als het niet ging en dan maar even te praten<br />

en/of met de PS2 of Gamecube de gedachten op iets anders te zetten. En dan zijn er natuurlijk de vele<br />

concerten waar we met Esther, Astrid, Ralph en anderen naartoe zijn geweest (Iron Maiden (2x), Judas<br />

Priest, Dragonforce, Fields of Rock en Apocalyptica). Ik hoop dat we er nog een aantal leuke concerten<br />

aan kunnen toevoegen. We hebben veel plezier gehad om jouw cabaret in elkaar te zetten, dus doe je<br />

best met die van mij! Veel succes met je nieuwe baan bij DSM.<br />

Monika, (Zus 4), de vele uurtjes die we tijdens en na het werk hebben benut om je Nederlands te<br />

verbeteren, onze frustraties te kunnen uiten, zijn erg waardevol geweest. Frustraties van het werk of<br />

prive heb ik aan jou kunnen vertellen. Ik was wel blij met iemand op het lab die ook wel van<br />

199


Appendix I<br />

regelmatig schoonmaken hield. Het tripje naar Polen wat je samen met Gosia hebt georganiseerd, was<br />

zeker een van leukste vakanties die ik heb gehad. Vooral de wandelingen door de Poolse bossen naar<br />

de bergtoppen met de soundtrack van Lord of the Rings op mijn iPod zijn herinneringen die me altijd<br />

bij zullen blijven. Ik kijk ernaar uit wanneer je jouw proefschrift af hebt, zodat ik eindelijk je cabaret in<br />

elkaar kan zetten (materiaal genoeg!!).<br />

En dan er natuurlijk al die andere collega’s die op een of andere manier invloed hebben gehad op het<br />

tot stand komen van dit proefschrift. René, less talking, more eating!! Verbazingwekkend, dat het toch<br />

goed gegaan is met een Sneker en een Bolswarder op één lab. Samen met Thijs waren we “<strong>The</strong> Three<br />

Musketeers” (of “<strong>The</strong> Three Stooges”) die het samen met Jan Maarten en Jean-Yves allemaal zijn<br />

begonnen. Ik vond het leuk om op vrijdag ochtend te gaan zwemmen voordat we weer aan het werk<br />

moesten. Veel succes met je onderzoek in Amsterdam.<br />

En ik kan natuurlijk niet al die andere collega’s vergeten waar ik samen mee heb gewerkt en samen een<br />

aantal leuke dingen hebben gedaan; Jean-Yves, Gosia, Girbe, Jessica, May, Dennis, Rense, Matty,<br />

Willy, Sjouke, Emma, Jetta, Jolanda, Magda, Arthur, Lakshmi, Henrik, Vahid, Carmine, Danai,<br />

Marcus, Hermie, Vivianne, Eleni en alle studenten die bij ons hebben rond gelopen, allemaal hartelijk<br />

dank voor jullie hulp, gezelligheid, alle borrels die we samen hebben gedaan en natuurlijk het labuitje<br />

naar Franeker en het godvergeten Bolsward (hoe verzin je het!?!?!). Ook de UMCG volleybal<br />

toernooien waarin we twee keer hebben gewonnen waren voor mij leuke sportieve uitjes.<br />

Natuurlijk kan ik de studenten die mee hebben geholpen in het onderzoek niet vergeten. Marloes,<br />

Magda, Jessica, Vincent, Eleni en Diana, jullie ook hartelijk dank voor jullie inzet en voor de<br />

mogelijkheid voor mij om onderwijs te geven. Het kan geen toeval zijn dat drie van mijn studenten<br />

uiteindelijk ook bij Jan Maarten gaan promoveren (zei hij met een knipoog).<br />

Vier jaar bezig zijn met een proefschrift kan niet zonder enige vorm van ontspanning. Gelukkig waren<br />

er genoeg mensen om te klaverjassen. Samen met Rik, Rene en Thijs zijn we begonnen met<br />

klaverjassen en zijn uiteindelijk verder gegaan met pokeren. Het klaverjassen hebben we daarna weer<br />

opgepikt met Thijs, Esther en Annechien of met Thijs, Esther en Chris. Bedankt voor de vele avonden<br />

met goed eten, bier, chips en een hoop onzinnig geklets. Mattijs, Ykelien, Almer, en alle anderen ook<br />

hartelijk dank voor het ontzettend mooie jaar bij FarmBio, met de vele borrels, filmpjes en Kolonisten<br />

avonden.<br />

Because limited exercise is being performed during board- and cardgames, there was need for some<br />

physical training to get rid of all the calories obtained from the chips and beers. Many thanks to Tao<br />

and Yoshitaka, who convinced me to join them for playing badminton every Tuesday. Within the<br />

department, several other people were interested in playing badminton, with the one condition that after<br />

playing we would go out for dinner. Tao and May “ ”, Yoshitaka “ ”,Ank, Rudi, and<br />

Erwin “bedankt” for these very nice moments of physical exhaustion and the very nice dinners<br />

afterwards.<br />

Dennis en Grietje (en kinderen), Folkert en Klaske, Ebbo en Ingrid (en kinderen), hartelijk dank voor<br />

de bezoekjes en de etentjes die we hebben georganiseerd. Ik hoop dat we hiermee nog door kunnen<br />

gaan. Many thanks to the Greek community in Groningen (Basilis and Xrysa, Dimitris, Evi, Evelina<br />

and Danai) for your warm welcome and introduction to the Greek people and their food.<br />

Ανέστη, Βέτα και Θράσο σας ευχαριστώ για το θερµό καλοσώρισµα στην οικογένεια Τσοµπανίδη.<br />

Τώρα που τελείωσα το διδακτορικό µου θα έχω επιτέλους τον χρόνο να µάθω ελληνικά ώστε να<br />

µπορούµε να µιλάµε άµεσα.<br />

Danai, you thought you were going to share a house with Eleni, but you were getting another guest as<br />

well. Thank you for your understanding that you accepted me as a housemate and slowly took Eleni to<br />

be my housemate.<br />

200


Dankwoord<br />

And last but not least: Eleni, zouzounaki mou, “euxaristo para para polu” for your help as a student, for<br />

being there at the right time, for your love and understanding and accepting me the way I am, for you<br />

making me laugh again in difficult times, for sharing our passion for music (I can’t wait to go to<br />

Jamaica to visit Bob) and the concerts we went to and will go to, for showing me the nice places in the<br />

Netherlands where, without you, I probably would not have gone, for showing me the nice parts of<br />

Greece, for going out with me for dinner every 24 th of the month, for introducing me to your family,<br />

and for being part of my family. I don’t know where it will go from here, but I want you to know that I<br />

will always love you. Sagapo!<br />

Mark<br />

201


202


Appendix II<br />

Publication list<br />

203


Appendix II Publication list<br />

Publication list<br />

1. Huchon, D., Madsen, O., Sibbald, M.J., Ament, K., Stanhope, M.J., Catzeflis, F., de<br />

Jong, W.W., Douzery, E.J. (2002) Rodent phylogeny and a timescale for the evolution of<br />

Glires: evidence from an extensive taxon sampling using three nuclear genes. Mol. Biol.<br />

Evol. 19, 1053-1065.<br />

2. Bokma, E., Rozeboom, H.J., Sibbald, M., Dijkstra, B.W., Beintema, J.J. (2002)<br />

Expression and characterization of active site mutants of hevamine, a chitinase from the<br />

rubber tree Hevea brasiliensis. Eur. J. Biochem. 269, 893-901.<br />

3. Voigt, B., Schweder, T., Sibbald, M.J.J.B., Albrecht, D., Ehrenreich, A., Feesche, J.,<br />

Maurer, K.-H., Gottschalk, G., van Dijl, J.M., Hecker, M. (2005) <strong>The</strong> extracellular<br />

proteome of Bacillus licheniformis grown in different media and under different nutrient<br />

starvation conditions. Proteomics. 6, 268-281.<br />

4. Sibbald, M.J.J.B., Ziebandt, A.-K., Engelmann, S., Hecker, M., de Jong, A., Harmsen,<br />

H.J.M., Raangs, G.C., Arends, J., Dubois, J.-Y.F., van Dijl, J.M. (2006) Mapping the<br />

pathways to staphylococcal pathogenesis by comparative secretomics. Microbiol. Mol.<br />

Biol. Rev. 70, 755-788.<br />

5. van Dijl, J.M., Buist, G., Sibbald, M.J.J.B., Zweers, J.C., Dubois, J.-Y.F., and Tjalsma,<br />

H. (2006) In's and out's of the Bacillus subtilis membrane proteome. In: Bacillus: Cellular<br />

and Molecular Biology (P. Graumann ed.), Horizon Scientific Press, Hethersett, Norwich,<br />

UK.<br />

6. Sibbald, M.J.J.B., van Dijl, J.M. (2008) Secretome mapping in Gram-positive pathogens.<br />

In: Bacterial secreted proteins: secretory mechanisms and role in pathogenesis (K.<br />

Wooldridge ed.). Horizon Scientific Press, Hethersett, Norwich, UK.<br />

7. Ziebandt, A.-K., Degner, M., Sibbald, M.J.J.B., Arends, J.P., Chlebowicz, M.A., Kusch,<br />

H., Albrecht, D., Pantuček, R., Doškar, J., Ziebuhr, W., Bröker, B.M., Hecker, M., van<br />

Dijl, J.M., Engelmann1, S. Proteogenomics uncovers extreme heterogeneity of the<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> exoproteome. Proteomics. In revision.<br />

8. Kouwen, T.R.H.M., Trip, E.N., Denham, E.L., Sibbald, M.J.J.B., Dubois, J.-Y.F., van<br />

Dijl, J.M. (2009) <strong>The</strong> large mechanosensitive channel (MscL) determines bacterial<br />

susceptibility to the bacteriocin sublancin 168. Antimicrob. Agents Chemother. 53, 4702-<br />

4711.<br />

9. Sibbald, M.J.J.B. # , Winter, T. # , van der Kooi-Pol, M.M., Bosma, T., Schäfer, T., Ohlsen,<br />

K., Hecker, M., Antelmann, H., Engelmann, S., van Dijl, J.M. SecG of <strong>Staphylococcus</strong><br />

<strong>aureus</strong> plays a crucial role in the translocation of virulence factors. In revision.<br />

10. Sibbald, M.J.J.B., Winter, T., ten Brinke M., Buist, G., Koedijk, D.G.A.M., van der<br />

Kooi-Pol, M.M., Msadek, T., Poupel, O., Rühmling, V., Stokroos, I., Tsompanidou, E.,<br />

Antelmann, H., Hecker, M., Engelmann, S., van Dijl, J.M. Characterization of<br />

<strong>Staphylococcus</strong> <strong>aureus</strong> secretion mutants. To be submitted.<br />

11. Sibbald, M.J.J.B. # , Yang, X.M. # , Tsompanidou, E., Hecker, M., Becher, D., Buist, G.,<br />

van Dijl, J.M. Partially overlapping substrate specificities of sortases A and C of<br />

staphylococci. Submitted.<br />

# both authors contributed equally to this work<br />

204


Appendices III-V<br />

Supplemental tables<br />

205


Appendix III: Supplemental tables to Chapter 2<br />

a. Identified proteins in extracellular proteomes of various S. <strong>aureus</strong> strains with a<br />

known signal peptide<br />

b. Identified proteins in the extracellular proteomes of various S. <strong>aureus</strong> strains without<br />

a known signal peptide<br />

c. <strong>The</strong> core exoproteome of S. <strong>aureus</strong>; proteins with predicted Sec type signal peptides<br />

present in all sequenced strains<br />

d. <strong>The</strong> variant exoproteome of S. <strong>aureus</strong>; proteins with predicted Sec type signal<br />

peptides present in at least one sequenced strain<br />

e. <strong>The</strong> core lipoproteome of S. <strong>aureus</strong>; proteins with predicted lipoprotein signal<br />

peptides present in all sequenced strains<br />

f. <strong>The</strong> variant lipoproteome of S. <strong>aureus</strong>; proteins with predicted lipoprotein signal<br />

peptides present in at least one sequenced strain<br />

g. Composition of the variant exoproteome of sequenced S. <strong>aureus</strong> strains<br />

h. Composition of the variant lipoproteome of sequenced S. <strong>aureus</strong> strains<br />

i. Specific S. epidermidis proteins with a predicted signal peptide<br />

j. Proteins with predicted Sec type signal peptides present in S. <strong>aureus</strong> RF122<br />

Appendix IV: Supplemental tables to Chapter 3<br />

a. Characteristics of the clinical S. <strong>aureus</strong> isolates used in this study<br />

b. Identified proteins on 2-D gels of the S. <strong>aureus</strong> isolates<br />

c. Extracellular proteins in different S. <strong>aureus</strong> isolates<br />

Appendix V: Supplemental tables to Chapter 9<br />

a. Proteins of B. licheniformis DSM 13 with a predicted signal peptide<br />

b. Extracellular proteins of B. licheniformis which either contain an N-terminal signal<br />

peptide or which are known to be secreted by other pathways<br />

c. Proteins detected in the extracellular proteome lacking known export signals<br />

206


Appendix III Supplemental Table IIIa<br />

Supplemental table IIIa. Identified proteins in extracellular proteomes of various S. <strong>aureus</strong><br />

strains a with a known signal peptide<br />

PID Protein Function -3 -2 -1 +1 Localization Motif<br />

15925728 b SA0022 hypothetical protein S N A A Extracellular LPKTG<br />

15925799 Plc 1-phosphatidylinositol<br />

phosphodiesterase precurosr<br />

A H A S Extracellular<br />

15925800 SA0092 hypothetical protein T A G C Extracellular Lipo<br />

15925815 b Spa immunoglobulin G binding A N A A Membrane / LPETG<br />

protein A precursor<br />

Cell Wall /<br />

Extracellular<br />

15925838 SasD hypothetical protein A H A D Extracellular LPAAG<br />

15925848 SA0139 hypothetical protein S L A I Extracellular<br />

15925933 Coa staphylocoagulase precursor A D A I Extracellular<br />

15925978 LytM peptidoglycan hydrolase A D A A Extracellular<br />

15925983 SA0270 hypothetical protein A Q A Y Extracellular<br />

15926008 SA0295 hypothetical protein A F A K Extracellular<br />

15926022 Geh glycerol ester hydrolase A Q A S Extracellular<br />

15926073 SA0359 hypothetical protein L T A C Extracellular Lipo<br />

15926099 Set6 exotoxin 6 V Q A K Extracellular<br />

15926104 Set11 exotoxin 11 V H A K Extracellular<br />

15926111 Set15 exotoxin 15 V K A S Extracellular<br />

15926112 SA0394 hypothetical protein A E A S Extracellular<br />

15926142 SA0423 hypothetical protein A N A A Extracellular LysM<br />

15926239 b SdrC Ser-Asp rich fibrinogen- A K A A Membrane / LPETG<br />

binding, bone sialoproteinbinding<br />

protein<br />

Cell Wall<br />

15926240 SdrD Ser-Asp rich fibrinogen- A K A A Membrane / LPETG<br />

binding, bone sialoproteinbinding<br />

protein<br />

Cell Wall<br />

15926241 SdrE Ser-Asp rich fibrinogenbinding,<br />

bone sialoproteinbinding<br />

protein<br />

A K A A Extracellular LPETG<br />

15926291 SA0570 hypothetical protein A E A A Extracellular<br />

15926342 b SA0620 hypothetical protein A Q A S Extracellular<br />

15926373 SA0651 hypothetical protein A L A K Extracellular<br />

15926385 SA0663 hypothetical protein L G A C Extracellular Lipo<br />

15926417 SA0695 hypothetical protein I S A C Extracellular Lipo<br />

15926464 b ClfA fibrinogen-binding protein A A D A S Membrane /<br />

Cell Wall<br />

LPQTG<br />

15926548 GlpQ glycerophosphoryl diester<br />

phosphodiesterase<br />

A G A E Extracellular<br />

15926570 SA0841 hypothetical protein V S A A Extracellular<br />

15926634 SspB cysteine protease precursor A K A D Extracellular<br />

15926635 SspA serine protease, V8 protease,<br />

glutamyl endopeptidase<br />

A N A L Extracellular<br />

15926639 Atl autolysin, N-acetylmuramyl-Lalanine<br />

amidase and endo-β-Nacetylglucosaminidase<br />

V Q A A Extracellular GW<br />

15926648 SA0914 hypothetical protein A D A T Extracellular<br />

15926713 IsdB iron-regulated heme-iron<br />

binding protein<br />

A Q A A Extracellular LPQTG<br />

15926714 IsdA cell surface protein V N A A Extracellular LPKTG<br />

15926715 IsdC hypothetical protein A N A A Extracellular NPQTN<br />

15926739 SA1001 hypothetical protein A K A F Extracellular<br />

15926746 SA1007 α-hemolysin precursor A N A A Extracellular<br />

15926796 SA1056 hypothetical protein V A G C Extracellular Lipo<br />

207


Appendix III Supplemental Table IIIa<br />

PID Protein Function -3 -2 -1 +1 Localization Motif<br />

15926830 b LytN hypothetical protein A Y A D Membrane /<br />

Cell Wall<br />

15926969 c SA1221 thioredoxin reductase L G A C Extracellular Lipo<br />

15927068 SA1318 hypothetical protein L S G C Extracellular Lipo<br />

15927308 SA1552 hypothetical protein A Q A A Extracellular LPKTG<br />

15927383 SplF serine protease SplF A K A E Extracellular 2 TM<br />

15927384 SplD serine protease SplD A K A E Membrane /<br />

Cell Wall<br />

2 TM<br />

15927385 SplC serine protease SplC A N A E Extracellular<br />

15927386 SplB serine protease SplB A K A E Extracellular<br />

15927387 SplA serine protease SplA A K A E Membrane /<br />

Cell Wall<br />

15927389 SA1633 probable β-lactamase A K A E Extracellular<br />

15927393 LukD leukotoxin, LukD V D A A Extracellular<br />

15927394 LukE leukotoxin, LukE S R A N Extracellular<br />

15927483 SA1725 staphopain, cysteine proteinase A N A E Extracellular<br />

15927512 Map truncated map-w protein A S A A Extracellular<br />

15927517 SA1755 hypothetical protein A K A F Extracellular<br />

15927520 Sak staphylokinase precursor V S A S Membrane /<br />

Cell Wall<br />

15927579 Hlb truncated β-hemolysin A K A E Extracellular<br />

15927580 SA1812 hypothetical protein S Y A K Extracellular<br />

15927581 SA1813 hypothetical protein T Q A N Extracellular<br />

15927586 SA1818 hypothetical protein A K A E Extracellular<br />

15927607 SA1839 hypothetical protein V E A K Extracellular<br />

15927670 SceD hypothetical protein A H A S Extracellular<br />

15927741 b FmtB FmtB protein A S A D Membrane /<br />

Cell Wall<br />

LPDTG<br />

15927757 SA1979 hypothetical protein V A A C Extracellular Lipo<br />

15927785 SA2006 hypothetical protein A S A D Extracellular<br />

15927879 SsaA hypothetical protein A H A S Extracellular<br />

15927884 SA2097 hypothetical protein A D A A Extracellular<br />

15927890 SA2103 hypothetical protein V A A K Extracellular<br />

15927988 SA2198 hypothetical protein L T A C Membrane /<br />

Cell Wall<br />

Lipo<br />

15927996 Sbi IgG-binding protein SBI A K A S Extracellular<br />

15927997 HlgA γ-hemolysin chain II precursor S K A E Extracellular<br />

15927998 HlgC γ-hemolysin component C A K A A Extracellular<br />

15927999 HlgB γ-hemolysin component B A N A E Extracellular<br />

15928076 SA2285 hypothetical protein A E A A Extracellular LPKTG<br />

15928148 IsaA immunodominant antigen A A H A A Extracellular<br />

15928216 ClfB clumping factor B A Q A S Extracellular LPETG<br />

15928223 Aur zinc metalloproteinase A L A I Membrane /<br />

aureolysin<br />

Cell Wall<br />

15928230 SA2437 N-acetylmuramoyl-L-alanine<br />

amidase<br />

A Y A D Extracellular<br />

15928232 b SasF hypothetical protein A Q A A Membrane /<br />

Cell Wall<br />

LPKAG<br />

15928254 IcaB intercellular adhesion protein A N A D Membrane /<br />

B<br />

Cell Wall<br />

15928257 Lip triacylglycerol lipase precursor A Q A A Extracellular<br />

49482650 d Set16 exotoxin 16 V Q A K Extracellular<br />

49482656 d Set11 exotoxin 11 V Q A A Extracellular<br />

49482662 d Set26 exotoxin 26 V K A S Extracellular<br />

49482663 d SAR0436 putative exported protein A E A S Extracellular<br />

49483672 SAR1494 hypothetical protein L S G C Extracellular Lipo<br />

208


Appendix III Supplemental Table IIIa<br />

PID Protein Function -3 -2 -1 +1 Localization Motif<br />

49484059 d SAR1905 serine protease A K A E Extracellular<br />

48494887 d Cna collagen adhesin precursor A F A A Extracellular LPKTG<br />

49484898 d SAR2788 putative exported protein A E A S Extracellular<br />

57652419 e Pls methicillin-resistant surface<br />

protein<br />

A E A A Extracellular LPDTG<br />

57651309 e SACOL0468 exotoxin 3, putative V Q A K Extracellular<br />

57651319 e SACOL0478 exotoxin 3, putative V K A S Extracellular<br />

57651320 e SACOL0479 surface protein, putative A E A S Extracellular<br />

57650159 e Sek staphylococcal enterotoxin A S A Q Extracellular<br />

57650160 e Sei staphylococcal enterotoxin<br />

type I<br />

A Y A D Extracellular<br />

57651597 e Seb staphylococcal enterotoxin B V L A E Extracellular<br />

57651598 e SACOL0908 hypothetical protein A K A S Extracellular<br />

57650600 e SACOL1865 serine protease SplE, putative A K A E Extracellular<br />

57650605 e SACOL1870 hypothetical protein A K A E Extracellular<br />

57650692 e Hlb β-hemolysin/phospholipase C A K A E Extracellular<br />

57651004 e SACOL2505 cell wall surface anchor family<br />

protein<br />

A E A A Extracellular LPKTG<br />

24636603 f Etd exfoliative toxin D S H A E Extracellular<br />

24636604 f probable glutamylendopeptidase<br />

V S A S Extracellular<br />

37196678 f Ser enterotoxin R V S A K Extracellular<br />

a<br />

<strong>The</strong> annotation of proteins is based on that of S. <strong>aureus</strong> N315, except for those proteins that are not encoded by<br />

the N315 genome<br />

b<br />

Identified in membrane/cell wall fraction by Nandakumar et al. (2005) and/or Gatlin et al. (2006)<br />

c<br />

Note that SA1221 is probably not a thioredoxin reductase, but a phosphate binding protein<br />

d<br />

Proteins encoded by S. <strong>aureus</strong> MRSA252 and absent from N315<br />

e<br />

Proteins encoded by S. <strong>aureus</strong> COL and absent from N315<br />

f<br />

Note that the corresponding gene is not present in the sequenced S. <strong>aureus</strong> genomes<br />

209


Appendix III Supplemental Table IIIb<br />

Supplemental table IIIb. Identified proteins in the extracellular proteomes of various S.<br />

<strong>aureus</strong> strains without a known signal peptide<br />

PID Protein Function<br />

15925714 SerS seryl-tRNA synthetase<br />

15925746 a MecR1 methicillin resistance protein<br />

15925747 a MecI methicillin resistance regulatory protein<br />

15925748 XylR hypothetical protein<br />

15925892 SA0182 hypothetical protein, similar to indole-3-pyruvate decarboxylas<br />

15925929 PflB formate acetyltransferase<br />

15925944 a LctE L-lactate dehydrogenase<br />

15925982 SA0269 hypothetical protein<br />

15925985 SA0272 hypothetical protein<br />

15926071 SA0357 hypothetical protein<br />

15926081 AhpF alkyl hydroperoxide reductase subunit F<br />

15926082 AhpC alkyl hydroperoxide reductase subunit C<br />

15926088 SA0372 hypothetical protein<br />

15926091 GuaB inositol-monophosphate dehydrogenase<br />

15926092 GuaA GMP synthase<br />

15926167 MetS methionyl-tRNA synthetase<br />

15926175 a SpoVG stage V sporulation protein G homologue<br />

15926178 RplY 50S ribosomal protein L25<br />

15926188 a FtsH cell-division protein<br />

15926190 CysK cysteine synthase (o-acetylserine sulfhydrylase) homologue<br />

15926194 LysS lysyl-tRNA synthetase<br />

15926202 ClpC HSP100/Clp ATPase<br />

15926205 GltX glutamyl-tRNA synthetase<br />

15926216 RplA 50S ribosomal protein L1<br />

15926225 Fus translational elongation factor G<br />

15926226 TufA translational elongation factor TU<br />

15926229 SA0509 chaperone protein HchA<br />

15926232 IlvE branched-chain amino acid aminotransferase<br />

15926236 SA0516 hypothetical protein<br />

15926244 SA0524 hypothetical protein<br />

15926266 Pta phosphotransacetylase<br />

15926283 Adh1 alcohol dehydrogenase<br />

15926294 SarA staphylococcal accessory regulator A<br />

15926363 SA0641 transcriptional regulator<br />

15926396 YfnI hypothetical protein with 5 transmembrane segments and a potential SPase I<br />

cleavage site<br />

15926420 a PepT aminotripeptidase<br />

15926429 SA0707 hypothetical protein<br />

15926441 TrxB thioredoxine reductase<br />

15926445 ClpP peptidase<br />

15926449 Gap glyceraldehyde-3-phosphate dehydrogenase<br />

15926450 Pgk phosphoglycerate kinase<br />

15926451 Tpi triosephosphate isomerase<br />

15926452 Pgm 2, 3-diphosphoglycerate-independentphosphoglycerate mutase<br />

15926453 Eno enolase<br />

15926469 a CspC cold-shock protein C<br />

15926497 a SA0769 ABC transporter ATP-binding protein homologue<br />

15926503 SA0775 hypothetical protein<br />

15926543 SA0806 hypothetical protein<br />

15926546 RocD ornithine-oxo-acid transaminase<br />

15926551 Pgi glucose-6-phosphate isomerase A<br />

15926554 a SpsB type-1 signal peptidase 1B<br />

15926559 Cdr coenzyme A disulfide reductase<br />

210


Appendix III Supplemental Table IIIb<br />

PID Protein Function<br />

15926572 FabH 3-oxoacyl-(acyl-carrier protein) synthase homologue<br />

15926589 SA0859 hypothetical protein<br />

15926603 SA0873 hypothetical protein<br />

15926632 MenB naphthoate synthase<br />

15926642 SA0908 hypothetical protein; predicted to have an uncleaved signal peptide<br />

15926657 PurM phosphoribosylformylglycinamidine cyclo-ligase PurM<br />

15926669 PtsH phophocarrier protein Hpr phosphohistidin-containing protein<br />

15926670 PtsI phosphoenolpyruvate-protein phosphatase<br />

15926679 PdhB pyruvate dehydrogenase E1 component beta subunit<br />

15926680 PdhC dihydrolipoamide S-acetyltransferase component of pyruvate dehydrogenase<br />

complex E2<br />

15926681 PdhD dihydrolipoamide dehydrogenase component of pyruvate dehydrogenase E3<br />

15926699 PycA pyruvate carboxylase<br />

15926716 IsdD hypothetical protein<br />

15926723 PheT Phe-tRNA synthetase β chain<br />

15926735 SA0998 hypothetical protein<br />

15926759 SA1019 hypothetical protein<br />

15926764 a PbpA penicillin-binding protein 1<br />

15926776 IleS Ile-tRNA synthetase<br />

15926779 a LspA lipoprotein signal peptidase<br />

15926817 a Smc chromosome segregation SMC protein<br />

15926828 SucD succinyl-CoA synthetase, β subunit<br />

15926829 SucC succinyl-CoA synthetase, α subunit<br />

15926838 CodY transcriptional repressor CodY<br />

15926839 RpsB 30S ribosomal protein S2<br />

15926840 Tsf elongation factor TS<br />

15926857 PnpA polyribonucleotide nucleotidyltransferase<br />

15926883 GlpK glycerol kinase<br />

15926884 GlpD aerobic glycerol-3-phosphate dehydrogenase<br />

15926892 GlnA glutamine-ammonia ligase<br />

15926901 a SA1157 hypothetical protein, similar to ABC transporter integral membrane protein<br />

15926915 KatA catalase<br />

15926919 SA1173 hypothetical protein<br />

15926923 Tkt transketolase<br />

15926930 CitB aconitate hydratase<br />

15926936 a AlsT amino acid carrier protein (sodium/alanine symporter)<br />

15926982 b CspA major cold shock protein CspA<br />

15926992 OdhB dihydrolipoamide succinyltransferase<br />

15926993 OdhA 2-oxoglutarate dehydrogenase E1<br />

15927003 SA1255 PTS system, glucose-specific enzyme II, A component<br />

15927005 SA1257 peptide methionine sulfoxide reductase<br />

15927015 SA1267 hypothetical protein, similar to streptococcal adhesin emb<br />

15927031 Pbp2 penicillin-binding protein 2<br />

15927035 AsnC asparaginyl-tRNA synthetase<br />

15927054 SA1305 DNA-binding protein II<br />

15927057 SA1308 30S ribosomal protein S1<br />

15927062 EbpS elastin binding protein<br />

15927086 SA1336 glucose-6-phosphate 1-dehydrogenase<br />

15927092 Gnd phosphogluconate dehydrogenase<br />

15927107 a AccC acetyl-CoA carboxylase AccC, biotin carboxylase subunit<br />

15927109 SA1359 translation elongation factor EF-P<br />

15927115 SA1365 glycine dehydrogenase subunit 2<br />

15927132 Pbp3 penicillin-binding protein 3<br />

15927133 SodA superoxide dismutase SodA<br />

15927145 GlyS glycyl-tRNA synthetase<br />

15927160 DnaK DnaK protein<br />

211


Appendix III Supplemental Table IIIb<br />

PID Protein Function<br />

15927161 GrpE GrpE protein<br />

15927190 GreA transcription elongation factor<br />

15927206 SA1453 hypothetical protein<br />

15927229 SA1475 hypothetical protein<br />

15927242 ValS valine t-RNA Ligase<br />

15927254 Tig trigger factor (prolyl isomerase)<br />

15927257 RplT 50S ribosomal protein L20<br />

15927261 ThrS threonyl-tRNA synthetase<br />

15927265 GapB glyceraldehyde 3-phosphate dehydrogenase 2<br />

15927273 CitZ citrate synthase II<br />

15927286 Ald alanine dehydrogenase<br />

15927287 SA1532 hypothetical protein<br />

15927288 AckA hypothetical protein<br />

15927309 Fhs formyltetrahydrofolate synthetase<br />

15927327 SA1571 D-alanine aminotransferase<br />

15927328 SA1572 hypothetical protein<br />

15927365 PckA phosphoenolpyruvate carboxykinase<br />

15927403 a Sem enterotoxin SEM<br />

15927409 TRAP signal transduction protein TRAP<br />

15927412 SA1656 Hit-like protein involved in cell-cycle regulation<br />

15927425 FumC fumarate hydratase<br />

15927428 SA1671 hypothetical protein<br />

15927495 SA1737 hypothetical protein<br />

15927503 SA1743 hypothetical protein<br />

15927524 Sep enterotoxin P<br />

15927539 SA1774 hypothetical protein<br />

15927584 Sel enterotoxin L<br />

15927604 GroEL GroEL protein<br />

15927605 GroES GroES protein<br />

15927638 SA1868 hypothetical protein<br />

15927681 AtpF ATP synthase subunit B<br />

15927687 GlyA serine hydroxymethyl transferase<br />

15927699 FbaA fructose-bisphosphate aldolase<br />

15927712 DeoD purine nucleoside phosphorylase<br />

15927753 SAS074 hypothetical protein<br />

15927762 Asp23 alkaline shock protein 23, Asp23<br />

15927782 a HysA hyaluronate lyase precursor<br />

15927798 RplM 50S ribosomal protein L13<br />

15927803 RplQ 50S ribosomal protein L17<br />

15927804 RpoA DNA-directed RNA polymerase α-subunit<br />

15927825 RplV 50S ribosomal protein L22<br />

15927911 HutU urocanate hydratase<br />

15927994 SA2204 phosphoglycerate mutase, Pgm homolog<br />

15928070 SA2279 hypothetical protein<br />

15928081 FnbB fibronectin binding protein B<br />

15928082 FnbA fibronectin binding protein A<br />

15928103 Ddh D-specific D-2-hydroxyacid dehydrogenase<br />

15928107 SrtA sortase A<br />

15928125 MvaS 3-hydroxy-3-methylglutaryl CoA synthase<br />

15928133 RocA 1-pyrroline-5-carboxylate dehydrogenase<br />

15928160 SA2367 hypothetical protein<br />

15928185 PanB 3-methyl-2-oxobutanoate hydroxymethyltransferase<br />

15928188 a SA2395 L-lactate dehydrogenase<br />

15928192 SA2399 fructose-bisphosphate aldolase<br />

15928221 ArcA arginine deiminase<br />

15928284 SA2490 hypothetical protein<br />

212


Appendix III Supplemental Table IIIb<br />

PID Protein Function<br />

16119211 BlaR1 bla regulator protein BlaR1<br />

15924938 c Sep enterotoxin P<br />

15925528 c PtsG PTS system, glucose-specific II ABC component<br />

18920604 d 18920604 phi ETA orf 18-like protein<br />

24636605 d Edin-B epidermal cell differentation inhibitor B<br />

49484190 e Sea enterotoxin type A precursor<br />

57650441 f SACOL1528 hypothetical protein<br />

57651702 f PdhA pyruvate dehydrogenase complex E1 component, α subunit<br />

a<br />

Identified in membrane/cell wall fraction by Nandakumar et al. (2005) and/or Gatlin et al. (2007)<br />

b<br />

Contains a pseudopilin SPase recognition and cleavage site<br />

c<br />

Proteins encoded by S. <strong>aureus</strong> Mu50 and absent from N315<br />

d<br />

Note that the corresponding gene is not present in the sequenced S. <strong>aureus</strong> genomes<br />

e<br />

Proteins encoded by S. <strong>aureus</strong> MRSA252 and absent from N315<br />

f Proteins encoded by S. <strong>aureus</strong> COL and absent from N315<br />

213


Appendix III Supplemental Table IIIc<br />

Supplemental table IIIc. <strong>The</strong> core exoproteome of S. <strong>aureus</strong>; proteins with predicted Sec<br />

type signal peptides present in all sequenced strains<br />

PID Protein -3 -2 -1 +1 Function<br />

15925728 a,b SA0022 S N A A hypothetical protein<br />

15925815 a Spa A N A A Immunoglobulin G binding protein A precursor<br />

15925838 c,d SasD A H A D hypothetical protein<br />

15925848 e SA0139 S L A I hypothetical protein<br />

15925933 Coa A D A I staphylocoagulase precursor<br />

15925978 f LytM A D A A peptidoglycan hydrolase<br />

15925983 SA0270 A Q A Y hypothetical protein, similar to precursor SsaA<br />

15926008 e SA0295 A F A K hypothetical protein<br />

15926022 e Geh A Q A S glycerol ester hydrolase<br />

15926106 g Set13 V H A K exotoxin 13<br />

15926107 h Set 14 G H A K exotoxin 14<br />

15926113 SA0395 A D A K hypothetical protein<br />

15926142 e,i Aaa A N A A hypothetical protein<br />

15926239 a,e SdrC A K A A Ser-Asp rich fibrinogen-binding<br />

15926291 e SA0570 A E A A hypothetical protein<br />

15926319 e,j Pbp4 A Q A T penicillin binding protein 4<br />

15926342 e,i SA0620 A Q A S secretory antigen SsaA homologue<br />

15926373 e SA0651 A L A K hypothetical protein<br />

15926432 e,i SA0710 A H A Q hypothetical protein<br />

15926464 a ClfA A D A S fibrinogen-binding protein A<br />

15926466 Ssp A K A A extracellular ECM and plasma binding protein<br />

15926467 SA0745 A N A L hypothetical protein<br />

15926468 Nuc A N A S staphylococcal nuclease<br />

15926548 e,j GlpQ A G A E glycerophosphoryl diester phosphodiesterase<br />

15926570 SA0841 V S A A hypothetical protein<br />

15926634 SspB A K A D cysteine protease precursor<br />

15926635 e SspA A N A L serine protease<br />

15926639 e,i Atl V Q A A autolysin<br />

15926648 e SA0914 A D A T hypothetical protein<br />

15926713 a IsdB A Q A A hypothetical protein<br />

15926714 a IsdA V N A A cell surface protein<br />

15926715 e,k IsdC A N A A hypothetical protein<br />

15926738 SA1000 S H A Q hypothetical protein<br />

15926741 Efb A D A S hypothetical protein<br />

15926742 SA1004 A D A S hypothetical protein<br />

15926749 SA1009 A K A Y hypothetical protein<br />

15926750 SA1010 A K A Y hypothetical protein<br />

15926751 SA1011 A K A Y hypothetical protein<br />

15926830 i LytN A Y A D LytN protein<br />

15927055 e GpsA V L A E glycerol-3-phosphate dehydrogenase<br />

15927385 SplC A N A E serine protease SplC<br />

15927456 e SA1698 S L A D hypothetical protein<br />

15927483 e SspB2 A N A E staphopain, cysteine proteinase<br />

15927498 e SAS056 A F A Y hypothetical protein<br />

15927580 l SA1812 S Y A K hypothetical protein<br />

15927581 SA1813 T Q A N hypothetical protein<br />

15927607 e SA1839† V E A K hypothetical protein<br />

15927670 e SA1898 A H A S hypothetical protein<br />

15927785 SA2006 A S A D hypothetical protein<br />

15927879 e SsaA A H A S hypothetical protein<br />

15927884 e SA2097 A D A A hypothetical protein<br />

15927890 e SA2103 V A A K hypothetical protein<br />

15927996 Sbi A K A S IgG-binding protein Sbi<br />

214


Appendix III Supplemental Table IIIb<br />

PID Protein -3 -2 -1 +1 Function<br />

15927997 HlgA S K A E γ-hemolysin chain II precursor<br />

15927998 HlgC A K A A γ-hemolysin component C<br />

15927999 HlgB A N A E γ-hemolysin component B<br />

15928114 SA2323 A Y A H hypothetical protein<br />

15928123 e SA2332 S H A A hypothetical protein<br />

15928145 e SA2353 A Q A A hypothetical protein<br />

15928148 e IsaA A H A A immunodominant antigen A<br />

15928216 a ClfB A Q A S clumping factor B<br />

15928223 e,m Aur A L A I zinc metalloproteinase aureolysin<br />

15928224 e,n IsaB A Q A A immunodominant antigen B<br />

15928225 SA2432 I Y A A hypothetical protein<br />

15928230 e SA2437 A Y A D hypothetical protein<br />

15928232 c,e SasF A Q A A conserved hypothetical protein<br />

15928254 IcaB A N A D intercellular adhesion protein B<br />

15928257 e Lip A Q A A triacylglycerol lipase precursor<br />

Signal peptide predictions were performed with SignalP-NN and SignalP-HMM version 2.0 (Nielsen et al., 1997;<br />

http://www.cbs.dtu.dk/services/SignalP-2.0/), PrediSi (Hiller et al, 2004; http://www.predisi.de/), Phobius (Kall et<br />

al., 2004; http://phobius.cgb.ki.se/) and LipoP version 1.0 (Juncker et al., 2003; (http://www.cbs.dtu.dk /services<br />

/LipoP/). <strong>The</strong>se programs are designed to identify Sec-type signal peptides, amino-terminal membrane anchors<br />

(Phobius), or lipoprotein signal peptides in Gram-negative bacteria (LipoP). <strong>The</strong> TMHMM-program version 2.0<br />

(Cserzö et al., 1997; http://www.cbs.dtu.dk/services/TMHMM/) was used to identify transmembrane segments in<br />

proteins.<br />

a<br />

Proteins with an LPxTG-motif<br />

b<br />

Protein homologue of B. subtilis is found in extracellular proteome(Tjalsma et al., 2004)<br />

c<br />

Proteins with an LPxAG-motif<br />

d<br />

Only the S. <strong>aureus</strong> NCTC 8325 protein is truncated at the C-terminus, thereby missing the LPxTG-motif<br />

e<br />

Proteins that are also present in S. epidermidis<br />

f<br />

Only the S. <strong>aureus</strong> Newman protein is truncated at the N-terminus, thereby missing the signal peptide<br />

g<br />

<strong>The</strong> S. <strong>aureus</strong> COL, NCTC8325, Newman, USA300 and USA300_TCHC1516 have a Gly on the -3 position,<br />

which is not included in the recognition and cleavage pattern<br />

h<br />

<strong>The</strong> homologue in S. <strong>aureus</strong> RF122 contains a recognition and cleavage site, but the proteins in all other strains<br />

contain a Gly on the -3 position, which is not included in the recognition and cleavage pattern<br />

i<br />

Proteins with a LysM or GW domain motif<br />

j<br />

Protein homologues of B. subtilis are classified as Sec-attached membrane protein (Tjalsma and van Dijl, 2005)<br />

k<br />

Proteins with an NPQTN-motif<br />

l<br />

Only the S. <strong>aureus</strong> COL protein is truncated at the N-terminus, thereby missing the signal peptide. All other S.<br />

<strong>aureus</strong> strains conform the proposed pattern<br />

m<br />

Protein homologue of B. subtilis is classified as secretory protein (Tjalsma and van Dijl, 2005)<br />

n<br />

Only for the S. <strong>aureus</strong> MRSA252 protein; protein homologues in other strains are predicted to have two<br />

transmembrane domains and were excluded from the list<br />

215


Appendix III Supplemental Table IIId<br />

Supplemental table IIId. <strong>The</strong> variant exoproteome of S. <strong>aureus</strong>; proteins with predicted<br />

Sec type signal peptides present in at least one sequenced strain<br />

PID Protein -3 -2 -1 +1 Function<br />

15925799 Plc A H A S 1-phosphatidylinositol phosphodiesterase<br />

precursor<br />

15926099 a Set6 V Q A K exotoxin 6<br />

15926100 a Set7 V H A E exotoxin 7<br />

15926101 a Set8 V K A E exotoxin 8<br />

15926102 a Set9 A N A T exotoxin 9<br />

15926103 a Set10 V N A S exotoxin 10<br />

15926104 a Set11 V H A K exotoxin 11<br />

15926105 a Set12 V N A K exotoxin 12<br />

15926111 a Set15 V K A S exotoxin 15<br />

15926112 SA0394 A E A S hypothetical protein<br />

15926240 b SdrD A K A A Ser-Asp rich fibrinogen-binding, bone<br />

sialoprotein-binding protein; low Smax score<br />

15926241 b SdrE A K A A Ser-Asp rich fibrinogen-binding, bone<br />

sialoprotein-binding protein; low Smax score<br />

15926465 SA0743 A S A V hypothetical protein<br />

15926739 SA1001 A K A F hypothetical protein<br />

15926746 Hly A N A A α-hemolysin precursor<br />

15927016 c EbhB A H A A hypothetical protein<br />

15927120 SA1370 I D A S hypothetical protein<br />

15927308 b Fhs A Q A A hypothetical protein<br />

15927384 d SplD A K A E serine protease SplD<br />

15927386 SplB A K A E serine protease SplB<br />

15927387 SplA A K A E serine protease SplA<br />

15927389 SA1633 A K A E probable β-lactamase<br />

15927393 LukD V D A A leukotoxin, LukD<br />

15927394 e LukE S R A N leukotoxin, LukE<br />

15927398 SEG V N A Q extracellular enterotoxin type G precursor<br />

15927399 SEN V N A E enterotoxin SeN<br />

15927402 SEI T Y A Q extracellular enterotoxin type I precursor<br />

15927404 SEO A Y A N enterotoxin SeO<br />

15927512 Map A S A A truncated Map-W protein<br />

15927513 Hlb A K A E truncated β-hemolysin<br />

15927516 SA1754 A Q A S hypothetical protein<br />

15927517 SA1755 A K A F hypothetical protein<br />

15927520 Sak V S A S staphylokinase precursor<br />

15927522 SA1760 A K A I hypothetical protein<br />

15927579 c,f SA1811 A K A E truncated β-hemolysin<br />

15927585 SEC3 V L A E enterotoxin type C3<br />

15927586 SA1818 A K A E hypothetical protein<br />

15927587 TSST-1 A K A S toxic shock syndrome toxin-1<br />

15927741 b,g FmtB A S A A FmtB protein<br />

15928076 b,c,h SA2285 A E A A hypothetical protein<br />

15928174 b,i SA2381 A N A E hypothetical protein<br />

15928182 SA2389 V L A D hypothetical protein<br />

16119203 j SAP003 A Y A N hypothetical protein<br />

16119219 k SAP019 A E A A hypothetical protein<br />

15923851 SAV0861 S D A I hypothetical protein<br />

14141830 j SAVP008 A N A E hypothetical protein<br />

21281780 SEH A K A E enterotoxin H<br />

21282111 Set16 V Q A K hypothetical protein<br />

21282116 Set21 V K A A hypothetical protein<br />

21282123 Set26 V K A I hypothetical protein<br />

216


Appendix III Supplemental Table IIIb<br />

PID Protein -3 -2 -1 +1 Function<br />

21283107 LukF V D A A Panton-Valentine leukocidin chain F precursor<br />

21283108 LukS S K A D Panton-Valentine leukocidin chain S precursor<br />

21283486 MW1757 A K A E hypothetical protein<br />

21283490 EpiP A S A S epidermin leader peptide processing serine<br />

protease EpiP<br />

21283666 SEG2 A Y A D staphylococcal enterotoxin G<br />

21283667 SEK A S A Q staphylococcal enterotoxin K<br />

49482651 SAR0423 V H A E exotoxin<br />

49482652 SAR0424 A N A E exotoxin<br />

49482653 SAR0425 A N A E exotoxin<br />

49482925 c SAR0721 T F A E multicopper oxidase protein<br />

49484047 SAR1886 A Y A F putative exported protein<br />

49484058 SplE A K A E serine protease<br />

49484059 c SAR1905 A K A E serine protease<br />

49484887 b Cna A L A A collagen adhesin precursor<br />

49484898 SAR2788 A E A S putative exported protein<br />

57651309 SACOL0468 V Q A K exotoxin 3, putative<br />

57651319 SACOL0478 V K A S exotoxin 3, putative<br />

57651320 SACOL0479 A E A S putative surface protein<br />

57651597 SEB V L A E staphylococcal enterotoxin B<br />

57652419 b Pls A E A A methicillin-resistant surface protein<br />

87159841 pUSA010004 A Q A Q hypothetical protein<br />

a<br />

Although these proteins share homology with exotoxins from other S. <strong>aureus</strong> strains these proteins are highly<br />

variable (Holtfreter et al., 2004)<br />

b<br />

Proteins with an LPxTG motif<br />

c<br />

Proteins that are also present in S. epidermidis<br />

d<br />

Only for the S. <strong>aureus</strong> COL protein; protein homologues in other strains are predicted to have two<br />

transmembrane domains and were excluded from the list<br />

e<br />

Only the S. <strong>aureus</strong> NCTC 8325 protein is truncated in the N-terminus and therefore missing the signal peptide<br />

f<br />

Truncated in all S. <strong>aureus</strong> strains, except COL, RF122 and S. epidermidis, thereby missing the signal peptide<br />

g<br />

S. <strong>aureus</strong> NCTC8325 misses the LPxTG retention signal; annotated as two proteins in S. <strong>aureus</strong> USA300<br />

h<br />

<strong>The</strong> S. <strong>aureus</strong> Mu50, N315 and USA300 proteins are truncated in their C-termini, therefore missing the LPxTG<br />

motif<br />

i<br />

Only the S. <strong>aureus</strong> MW2 protein is truncated in the N-terminus, therefore missing the signal peptide<br />

j Protein encoding gene lies on a plasmid<br />

k Protein encoding gene lies on a plasmid, except for S. <strong>aureus</strong> MRSA252<br />

217


Appendix III Supplemental Table IIIe<br />

Supplemental table IIIe. <strong>The</strong> core lipoproteome of S. <strong>aureus</strong>; proteins with predicted<br />

lipoprotein signal peptides present in all sequenced strains<br />

PID Gene -3 -2 -1 +1 +2 Function<br />

15925800 SA0092 T A G C G hypothetical protein<br />

15925819 SirA L A G C S lipoprotein<br />

15925912 Slp L S G C G RGD-containing lipoprotein<br />

15925918 SA0207 V T A C G hypothetical protein, similar to maltose/<br />

maltodextrin-binding protein<br />

15925928 a,b SA0217 L S S C A hypothetical protein, similar to periplasmic-ironbinding<br />

protein BitC<br />

15925940 c,d SA0229 L S G C G hypothetical protein, similar to nickel ABC<br />

transporter nickel-binding protein<br />

15926044 SA0331 I A A C G conserved hypothetical protein<br />

15926073 SA0359 L T A C G conserved hypothetical protein<br />

15926079 a SA0363 L T G C A hypothetical protein<br />

15926141 a,e SA0422 L A A C G hypothetical protein, similar to lactococcal<br />

lipoprotein<br />

15926287 a,f SA0566 L S G C G hypothetical protein, similar to iron-binding<br />

protein<br />

15926308 a SA0587 V A A C G lipoprotein, Streptococcal adhesin PsaA<br />

homologue<br />

15926354 a SA0632 L T G C G conserved hypothetical protein<br />

15926385 a,g SA0663 L G A C G hypothetical protein<br />

15926413 a,h SA0691 L A A C G lipoprotein, similar to ferrichrome ABC<br />

transporter<br />

15926417 a SA0695 I S A C G hypothetical protein<br />

15926461 SA0739 L G A C G conserved hypothetical protein<br />

15926499 a,I SA0771 L A A C G conserved hypothetical protein<br />

15926579 a SA0849 L S G C A hypothetical protein, similar to peptide binding<br />

protein OppA<br />

15926625 f SA0891 V A G C G hypothetical protein, similar to ferrichrome ABC<br />

transporter<br />

15926678 a SA0943 L A G C T conserved hypothetical protein<br />

15926717 j IsdE L T S C Q hypothetical protein<br />

15926796 a SA1056 V A G C S hypothetical protein<br />

15926969 j SA1221 L G A C G thioredoxin reductase<br />

15927111 a SA1361 L A G C G hypothetical protein<br />

15927372 j SA1616 L S S C G hypothetical protein<br />

15927373 k SA1617 L S A C S hypothetical protein<br />

15927375 a,l SA1619 L T A C G hypothetical protein<br />

15927415 a PrsA L G A C G peptidyl-prolyl cis/trans isomerase homolog<br />

15927477 a SA1719 L A A C G conserved hypothetical protein<br />

15927757 a,m SA1979 V A A C G hypothetical protein, similar toferrichrome ABC<br />

transporter (binding protein)<br />

15927859 a ModA L A G C S probable molybdate-binding protein<br />

15927864 i SA2079 L A A C G hypothetical protein, similar to ferrichrome ABC<br />

transporter FhuD precursor<br />

15927948 a SA2158 L A A C G hypothetical protein, similar to TpgX protein<br />

15927961 a,j SA2171 L I V C I hypothetical protein<br />

15927984 a DsbA L A A C G DsbA; hypothetical protein, similar to Zn-binding<br />

lipoprotein AdcA<br />

15927987 c SA2197 L T A C G conserved hypothetical protein<br />

15927988 c,f SA2198 I S G C G hypothetical protein<br />

15927992 a,i SA2202 L A A C G hypothetical protein, similar to ABC transporter,<br />

periplasmic amino acid-binding protein<br />

218


Appendix III Supplemental Table IIIb<br />

PID Gene -3 -2 -1 +1 +2 Function<br />

15928025 a OpuCC L S G C S glycine betaine/carnitine/choline ABC transporter<br />

OpuC<br />

15928037 a SA2247 L S A C G conserved hypothetical protein<br />

15928046 a Opp-1A L T G C G oligopeptide transporter putative substrate<br />

binding domain<br />

15928066 a,m SA2275 I G A C G hypothetical protein<br />

15928267 j SA2473 L Y S C S hypothetical protein<br />

a<br />

Proteins that are also present in S. epidermidis<br />

b<br />

Excluded for S. <strong>aureus</strong> JH1, JH9, Mu3, Mu50 and N315 because of the motif proposed by Sutcliffe and<br />

Harrington (2002)<br />

c<br />

Excluded for S. epidermidis because of the motif proposed by Sutcliffe and Harrington (2002)<br />

d<br />

<strong>The</strong> S. <strong>aureus</strong> JH1, JH9 and NCTC 8325 proteins are truncated at the N-terminus and thereby missing the signal<br />

peptide<br />

e<br />

Contains the lipoprotein release motif (Tjalsma and van Dijl, 2005) with one amino acid changes, except for S.<br />

<strong>aureus</strong> COL, NCTC 8325, Newman, USA300 and S. epidermidis<br />

f<br />

<strong>The</strong> S. <strong>aureus</strong> NCTC 8325 and USA300 proteins are truncated at the N-terminus and thereby missing the signal<br />

peptide<br />

g<br />

Contains the lipoprotein release motif (Tjalsma and van Dijl, 2005) with one amino acid change for all<br />

staphylococcal strains<br />

h<br />

Contains the exact lipoprotein release motif (Tjalsma and van Dijl, 2005), except in S. epidermidis<br />

i<br />

Contains the lipoprotein release motif (Tjalsma and van Dijl, 2005) with one amino acid change, except in S.<br />

epidermidis<br />

j<br />

Excluded for all S. <strong>aureus</strong> strains because of the motif proposed by Sutcliffe and Harrington (2002)<br />

k<br />

Excluded for all S. <strong>aureus</strong> strains because of the motif proposed by Sutcliffe and Harrington (2002), except for S.<br />

<strong>aureus</strong> MRSA252<br />

l<br />

Annotated as two proteins in the S. <strong>aureus</strong> USA300 strain<br />

m<br />

Contains the lipoprotein release motif (Tjalsma and van Dijl, 2005) with one amino acid change, only for S.<br />

epidermidis<br />

219


Appendix III Supplemental Table IIIf<br />

Supplemental table IIIf. <strong>The</strong> variant lipoproteome of S. <strong>aureus</strong>; proteins with predicted<br />

lipoprotein signal peptides present in at least one sequenced strain<br />

PID Gene -3 -2 -1 +1 +2 Function<br />

15925801 a SA0093 F A G C G hypothetical protein<br />

15925802 SA0094 V A G C G hypothetical protein<br />

15925803 SA0095 T A G C G hypothetical protein<br />

15925804 SA0096 T A G C G hypothetical protein<br />

15925847 b SA0138 A A A C G hypothetical protein, similar to<br />

alkylphosphonate ABC tranporter<br />

15925877 a SA0167 I T G C D hypothetical protein, similar to membrane<br />

lipoprotein SrpL<br />

15926004 SA0291 L A G C S hypothetical protein<br />

15926114 c Lpl1 I A G C G hypothetical protein<br />

15926115 a Lpl2 I I G C D hypothetical protein<br />

15926116 Lpl3 I A G C G hypothetical protein<br />

15926118 Lpl4 I I G C G hypothetical protein<br />

15926119 Lpl5 V A G C G hypothetical protein<br />

15926120 a Lpl6 I I G C D hypothetical protein<br />

15926121 a Lpl7 I I G C N hypothetical protein<br />

15926122 a Lpl8 A T S C G hypothetical protein<br />

15926123 Lpl9 I G G C G hypothetical protein<br />

15926465 a SA0743 G A L C V hypothetical protein<br />

15926580 d SA0850 L S A C G hypothetical protein, similar to<br />

oligopeptide ABC transporter<br />

oligopeptide-binding protein<br />

15927067 SA1317 L S G C S hypothetical protein<br />

15927068 b SA1318 L S G C S hypothetical protein<br />

15927069 SA1319 L S G C S hypothetical protein<br />

15927071 e SA1321 L G G C S hypothetical protein<br />

15927396 c SA1640 L V A C G conserved hypothetical protein<br />

15928064 a SA2273 I G G C I hypothetical protein<br />

16119210 b,f BlaZ L S A C N β-lactamase precursor<br />

14141829 a SAVP006 L V S C N hypothetical protein<br />

15923793 SAV0803 L T A C S hypothetical protein<br />

15924991 b SAV2001 L S A C G hypothetical protein<br />

21281801 MW0072 T A G C G <strong>Staphylococcus</strong> tandem lipoprotein<br />

21282126 Lpl10 I A G C G hypothetical protein<br />

21282127 a Lpl11 V T S C G hypothetical protein<br />

21282129 a Lpl13 I I G C D hypothetical protein<br />

21283103 MW1374 L S G C S conserved hypothetical protein<br />

21283167 MW1438 L T A C G hypothetical protein<br />

21283173 g MW1444 L G G C S hypothetical protein<br />

21284135 MW2406 I G A C G hypothetical protein<br />

21284306 MW2577 V S G C S hypothetical protein<br />

49482670 SAR0445 I G G C G putative lipoprotein<br />

49483474 SAR1288 L S A C G putative lipoprotein<br />

49483672 SAR1494 L S G C S hypothetical protein<br />

49484287 a SAR2149 L I V C G hypothetical protein<br />

49484977 b,h SAS0074 I G G C G putative lipoprotein<br />

57650161 SACOL0888 L G A C G pathogenicity island, putative lipoprotein<br />

57650444 SACOL1531 L S G C S hypothetical protein<br />

57650485 SACOL1574 L S A C G hypothetical protein<br />

57650996 a SACOL2497 I G G C V staphylococcus tandem lipoprotein<br />

57651323 a SACOL0482 I M G C D staphylococcus tandem lipoprotein<br />

57651324 SACOL0483 M A G C E staphylococcus tandem lipoprotein<br />

57651327 i SACOL0486 I G G C G staphylococcus tandem lipoprotein<br />

220


Appendix III Supplemental Table IIIb<br />

PID Gene -3 -2 -1 +1 +2 Function<br />

57652445 SACOL0081 T A G C G hypothetical protein<br />

150392787 SaurJH1_0313 L S A C G hypothetical protein<br />

150393510 SaurJH1_1042 L G A C G hypothetical protein<br />

87159856 pUSA03_0017 L A G C G transfer complex protein TraH<br />

87160691 a SAUSA300_0411 I G G C D staphylococcus tandem lipoprotein<br />

87160733 b,j SAUSA300_0079 L S A C S putative lipoprotein<br />

87161538 SAUSA300_0073 L G A C G peptide ABC transporter, peptide-binding<br />

protein<br />

151220617 a NWMN_0405 I G G C D truncated staphylococcal tandem<br />

lipoprotein<br />

a<br />

Excluded for all S. <strong>aureus</strong> strains because of the motif proposed by Sutcliffe and Harrington (2002)<br />

b<br />

Proteins that are also present in S. epidermidis<br />

c<br />

Excluded for S. <strong>aureus</strong> COL, Mu3, Mu50, N315, NCTC 8325, Newman and USA300 because of the motif<br />

proposed by Sutcliffe and Harrington (2002)<br />

d<br />

Contains the lipoprotein release motif (Tjalsma and van Dijl, 2005) with one amino acid change<br />

e<br />

<strong>The</strong> S. <strong>aureus</strong> COL, Mu3, Mu50, N315, NCTC 8325, Newman and USA300 proteins are truncated at the Nterminus,<br />

thereby missing the signal peptide<br />

f<br />

<strong>The</strong> S. <strong>aureus</strong> JH1, JH9, MSSA476 and N315 proteins are encoded by genes that lie on a plasmid<br />

g<br />

<strong>The</strong> S. <strong>aureus</strong> COL, NCTC 8325 and USA300 proteins are truncated at the N-terminus, thereby missing the<br />

signal peptide<br />

h<br />

Contains the lipoprotein release motif for S. epidermidis (Tjalsma and van Dijl, 2005) with one amino acid<br />

change<br />

i<br />

<strong>The</strong> S. <strong>aureus</strong> NCTC 8325 protein is truncated at the N-terminus, thereby missing the signal peptide<br />

j<br />

<strong>The</strong> S. epidermidis ATCC 12228 protein is truncated at the N-terminus, thereby missing the signal peptide<br />

221


Appendix III Supplemental Table IIIg<br />

Supplemental table IIIg. Composition of the variant exoproteome of sequenced S. <strong>aureus</strong> strains<br />

PID Protein Function COL JH1 JH9 MRSA 252 MSSA 476 Mu3 Mu50<br />

15925799 Plc 1-phosphatidylinositol phosphodiesterase precursor Y Y Y Y N Y Y<br />

15926099 Set6 exotoxin 6 N Y Y N N Y Y<br />

15926100 Set7 exotoxin 7 Y Y Y N Y Y Y<br />

15926101 Set8 exotoxin 8 N Y Y N Y Y Y<br />

15926102 Set9 exotoxin 9 Y Y Y N Y Y N<br />

15926103 Set10 exotoxin 10 N Y Y Y Y Y Y<br />

15926104 Set11 exotoxin 11 N Y Y Y Y Y Y<br />

15926105 Set12 exotoxin 12 N Y Y N Y Y Y<br />

15926111 Set15 exotoxin 15 N Y Y N N Y Y<br />

15926112 SA0394 hypothetical protein N Y Y Y Y Y Y<br />

15926465 SA0743 hypothetical protein Y Y Y N Y Y Y<br />

15926739 SA1001 hypothetical protein Y Y Y N Y Y Y<br />

15926746 HlY α-hemolysin precursor Y Y Y N Y Y Y<br />

15927016 EbhB hypothetical protein Y Y Y Y N Y Y<br />

15927120 SA1370 hypothetical protein N Y Y Y Y Y Y<br />

15927308 Fhs hypothetical protein Y Y Y N Y Y Y<br />

15927384 SplD serine protease SplD Y N N N N Y Y<br />

15927386 SplB serine protease SplB Y Y Y N Y Y Y<br />

15927387 SplA serine protease SplA Y Y Y N Y Y Y<br />

15927389 SA1633 probable β-lactamase N Y Y N N Y Y<br />

15927393 LukD leukotoxin, LukD Y Y Y N Y Y Y<br />

15927394 LukE leukotoxin, LukE Y Y Y N Y Y Y<br />

15927398 SEG extracellular enterotoxin type G precursor N Y Y Y N Y Y<br />

15927399 SEN enterotoxin SEN N Y Y Y N Y Y<br />

15927402 SEI extracellular enterotoxin type I precursor N Y Y Y N Y Y<br />

15927404 SEO enterotoxin SEO N Y Y Y N Y Y<br />

15927512 Map truncated map-w protein Y N Y Y N Y Y<br />

15927513 Hlb β-hemolysin/ phospholipase C Y Y Y N N Y Y<br />

15927516 SA1754 hypothetical protein N Y Y Y Y Y Y<br />

15927517 SA1755 hypothetical protein N Y Y Y N N N<br />

15927520 Sak staphylokinase precursor N Y Y Y Y Y Y<br />

15927522 SA1760 hypothetical protein N Y Y Y Y Y Y<br />

15927579 SA1811 phospholipase C Y Y Y N N Y Y<br />

15927585 SEC3 enterotoxin typeC3 Y N N Y N Y Y<br />

15927586 SA1818 hypothetical protein Y Y Y N N Y Y<br />

222


Appendix III Supplemental Table IIIg<br />

PID Protein Function COL JH1 JH9 MRSA 252 MSSA 476 Mu3 Mu50<br />

15927587 TSST-1 toxic shock syndrome toxin-1 N N N N N Y Y<br />

15927741 FmtB FmtB protein Y Y Y N N Y Y<br />

15928076 SA228 hypothetical protein Y Y Y N Y Y Y<br />

15928174 SA2381 hypothetical protein N Y Y N N Y Y<br />

15928182 SA2389 hypothetical protein N Y Y N N Y Y<br />

16119203 SAP003 hypothetical protein N Y Y N N N N<br />

16119219 SAP019 hypothetical protein N N N Y Y N N<br />

15923851 SAV0861 hypothetical protein N N N N N Y Y<br />

14141830 SAVP008 hypothetical protein N N N N N N Y<br />

21281780 SEH enterotoxin H N N N N Y N N<br />

21282111 Set16 hypothetical protein N N N Y Y N N<br />

21282116 Set21 hypothetical protein N N N N Y N N<br />

21282123 Set26 hypothetical protein N N N Y Y N N<br />

21283107 LukF Panton-Valentine leukocidin chain F precursor N N N N N N N<br />

21283108 LukS Panton-Valentine leukocidin chain S precursor N N N N N N N<br />

21283486 MW1757 hypothetical protein Y N N N Y N N<br />

21283490 EpiP epidermin leader peptide processing serine protease EpiP Y N N N Y N N<br />

21283666 SEG2 staphylococcal enterotoxin G Y N N N Y N N<br />

21283667 SEK staphylococcal enterotoxin K Y N N N Y N N<br />

21284341 Cna collagen adhesin precursor N N N Y Y N N<br />

49482651 SAR0423 exotoxin N N N Y N N N<br />

49482652 SAR0424 exotoxin N N N Y N N N<br />

49482653 SAR0425 exotoxin N N N Y N N N<br />

49482925 SAR0721 multicopper oxidase protein N N N Y N N N<br />

49483328 SAR1139 exotoxin N N N Y N N N<br />

49484047 SAR1886 putative exported protein N N N Y N N N<br />

49484058 SplE serine protease Y N N Y N N N<br />

49484059 SAR1905 serine protease N N N Y N N N<br />

49484898 SAR2788 putative exported protein N N N Y N N N<br />

57651309 SACOL0468 exotoxin 3, putative Y N N N N N N<br />

57651319 SACOL0478 exotoxin 3, putative Y N N N N N N<br />

57651320 SACOL0479 hypothetical protein Y N N N N N N<br />

57651597 SEB staphylococcal enterotoxin B Y N N N N N N<br />

57652419 Pls methicillin-resistant surface protein Y N N N N N N<br />

87159841 pUSA010004 hypothetical protein N N N N N N N<br />

223


Appendix III Supplemental Table IIIg<br />

Supplemental Table IIIg. Continued<br />

PID Protein Function MW2 N315 NCTC 8325 Newman USA300 USA300 TCHC1516<br />

15925799 Plc 1-phosphatidylinositol phosphodiesterase precursor Y Y Y Y Y Y<br />

15926099 Set6 exotoxin 6 N Y N N N N<br />

15926100 Set7 exotoxin 7 Y Y Y Y Y Y<br />

15926101 Set8 exotoxin 8 Y Y Y Y Y Y<br />

15926103 Set10 exotoxin 10 Y Y Y Y Y Y<br />

15926104 Set11 exotoxin 11 Y Y Y Y Y Y<br />

15926105 Set12 exotoxin 12 Y Y Y Y Y Y<br />

15926106 Set13 exotoxin 13 Y Y N N N N<br />

15926111 Set15 exotoxin 15 N Y N N N N<br />

15926465 SA0743 hypothetical protein Y Y Y Y Y Y<br />

15926739 SA1001 hypothetical protein Y Y Y Y Y Y<br />

15926746 HlY α-hemolysin precursor Y Y Y Y Y Y<br />

15927016 EbhB hypothetical protein Y Y Y Y Y Y<br />

15927120 SA1370 hypothetical protein Y Y Y Y N Y<br />

15927308 Fhs hypothetical protein Y Y Y Y Y Y<br />

15927384 SplD serine protease SplD N Y Y Y Y Y<br />

15927386 SplB serine protease SplB Y Y Y Y Y Y<br />

15927387 SplA serine protease SplA Y Y Y Y Y Y<br />

15927389 SA1633 probable β-lactamase N Y N N N N<br />

15927393 LukD leukotoxin, LukD Y Y Y Y Y Y<br />

15927394 LukE leukotoxin, LukE Y Y Y Y Y Y<br />

15927398 SEG extracellular enterotoxin type G precursor N Y N N N N<br />

15927399 SEN enterotoxin SEN N Y N N N N<br />

15927402 SEI extracellular enterotoxin type I precursor N Y N N N N<br />

15927404 SEO enterotoxin SEO N Y N N N N<br />

15927512 Map truncated map-w protein Y Y Y Y N Y<br />

15927513 Hlb β-hemolysin/ phospholipase C Y Y Y Y Y Y<br />

15927516 SA1754 hypothetical protein Y Y Y Y Y Y<br />

15927517 SA1755 hypothetical protein N Y Y Y Y Y<br />

15927520 Sak staphylokinase precursor Y Y Y Y Y Y<br />

15927522 SA1760 hypothetical protein Y Y Y Y Y Y<br />

15927579 SA1811 phospholipase C Y Y Y Y Y Y<br />

15927585 SEC3 enterotoxin typeC3 Y Y N N N N<br />

15927586 SA1818 hypothetical protein Y Y N N Y Y<br />

15927587 TSST-1 toxic shock syndrome toxin-1 N Y N N N N<br />

224


Appendix III Supplemental Table IIIg<br />

PID Protein Function MW2 N315 NCTC 8325 Newman USA300 USA300 TCHC1516<br />

15927741 FmtB FmtB protein Y Y Y Y Y N<br />

15928076 SA228 hypothetical protein Y Y Y Y Y N<br />

15928174 SA2381 hypothetical protein Y Y N N N N<br />

15928182 SA2389 hypothetical protein N Y N N N N<br />

16119203 SAP003 hypothetical protein N Y N N N N<br />

16119219 SAP019 hypothetical protein N Y N N N N<br />

15923851 SAV0861 hypothetical protein N N N N N N<br />

14141830 SAVP008 hypothetical protein N N N N N N<br />

21281780 SEH enterotoxin H Y N N N N N<br />

21282111 Set16 hypothetical protein Y N N N N N<br />

21282116 Set21 hypothetical protein Y N Y Y Y Y<br />

21282123 Set26 hypothetical protein Y N N N N N<br />

21283107 LukF Panton-Valentine leukocidin chain F precursor Y N N N Y Y<br />

21283108 LukS Panton-Valentine leukocidin chain S precursor Y N N N Y Y<br />

21283486 MW1757 hypothetical protein Y N Y Y Y Y<br />

21283490 EpiP epidermin leader peptide processing serine protease<br />

EpiP<br />

Y N Y Y Y Y<br />

21282666 SEG2 staphylococcal enterotoxin G Y N N N Y Y<br />

21282667 SEK staphylococcal enterotoxin K Y N N N Y Y<br />

21284341 Cna collagen adhesin precursor Y N N N N N<br />

49482651 SAR0423 exotoxin N N N N N N<br />

49482652 SAR0424 exotoxin N N N N N N<br />

49482653 SAR0425 exotoxin N N N N N N<br />

49482925 SAR0721 multicopper oxidase protein N N N N N N<br />

49484047 SAR1886 putative exported protein N N N N N N<br />

49484058 SplE serine protease N N Y Y Y Y<br />

49484059 SAR1905 serine protease N N N N N N<br />

49484898 SAR2788 putative exported protein N N N N N N<br />

57651309 SACOL0468 exotoxin 3, putative N N Y Y Y Y<br />

57651319 SACOL0478 exotoxin 3, putative N N Y Y Y Y<br />

57651597 SEB staphylococcal enterotoxin B N N N N N N<br />

57652419 Pls methicillin-resistant surface protein N N N N N N<br />

87159841 pUSA010004 hypothetical protein N N N N Y N<br />

225


Appendix III Supplemental Table IIIh<br />

Supplemental table IIIh. Composition of the variant lipoproteome of sequenced S. <strong>aureus</strong> strains<br />

PID Protein Function COL JH1 JH9 MRSA252 MSSA476 Mu3 Mu50<br />

15925801 SA0093 hypothetical protein N Y Y N N Y Y<br />

15925802 SA0094 hypothetical protein N Y Y N N Y Y<br />

15925803 SA0095 hypothetical protein N Y Y N N Y Y<br />

15925804 SA0096 hypothetical protein Y Y Y N N Y Y<br />

15925847 SA0138 hypothetical protein, similar to alkylphosphonate ABC<br />

tranporter<br />

Y Y Y Y Y Y Y<br />

15925877 SA0167 hypothetical protein N Y Y Y Y Y Y<br />

15926004 SA0291 hypothetical protein Y Y Y N Y Y Y<br />

15926114 Lpl1 hypothetical protein Y Y Y Y N Y Y<br />

19526115 Lpl2 hypothetical protein N Y Y N N Y Y<br />

15926116 Lpl3 hypothetical protein Y Y Y N N Y Y<br />

15926118 Lpl4 hypothetical protein N Y Y N N Y Y<br />

15926119 Lpl5 hypothetical protein N Y Y N N Y Y<br />

15926120 Lpl6 hypothetical protein N Y Y N Y Y Y<br />

15926122 Lpl8 hypothetical protein Y Y Y Y N Y Y<br />

15926123 Lpl9 hypothetical protein Y Y Y Y Y Y Y<br />

15926465 SA0743 hypothetical protein, similar to staphylocoagulase precursor Y Y Y N Y Y Y<br />

15926580 SA0850 hypothetical protein, similar to oligopeptide ABC transporter<br />

oligopeptide-binding protein<br />

Y Y Y N Y Y Y<br />

15927067 SA1317 hypothetical protein N Y Y N N Y Y<br />

15927068 SA1318 hypothetical protein N Y Y N Y Y Y<br />

15927069 SA1319 hypothetical protein N Y Y Y Y Y Y<br />

15927071 SA1321 hypothetical protein Y Y Y N N Y Y<br />

15927396 SA1640 conserved hypothetical protein N Y Y N N Y Y<br />

15928064 SA2273 hypothetical protein Y Y Y N N Y Y<br />

16119210 BlaZ β-lactamase precursor N Y Y Y Y N N<br />

14141829 SAVP006 hypothetical protein N N N N N N Y<br />

15923793 SAV0803 hypothetical protein Y N N N Y Y Y<br />

15924491 SAV2001 hypothetical protein N N N Y N Y Y<br />

21281801 MW0072 hypothetical protein Y N N N Y N N<br />

21282126 Lpl10 hypothetical protein N N N Y Y N N<br />

21282127 Lpl11 hypothetical protein N N N Y Y N N<br />

21282129 Lpl13 hypothetical protein N N N N Y N N<br />

226


Appendix III Supplemental Table IIIh<br />

PID Protein Function COL JH1 JH9 MRSA252 MSSA476 Mu3 Mu50<br />

21283103 MW1374 conserved hypothetical protein N N N N Y N N<br />

21283167 MW1438 hypothetical protein N N N Y Y N N<br />

21283173 MW1444 hypothetical protein Y N N N Y N N<br />

21284135 MW2406 hypothetical protein N N N N Y N N<br />

21284306 MW2577 hypothetical protein N N N N Y N N<br />

49482670 SAR0445 putative lipoprotein N N N Y N N N<br />

49483474 SAR1288 putative lipoprotein N N N Y N N N<br />

49483672 SAR1494 hypothetical protein Y N N Y N N N<br />

49484287 SAR2149 putative exported protein N N N Y N N N<br />

49484977 SAS0074 putative lipoprotein N N N N Y N N<br />

57650161 SACOL0888 pathogenicity island, putative lipoprotein Y N N N N N N<br />

57650444 SACOL1531 hypothetical protein Y N N N N N N<br />

57650485 SACOL1574 hypothetical protein SA1574 Y N N Y N N N<br />

57650996 SACOL2497 staphylococcus tandem lipoprotein Y N N N N N N<br />

57651323 SACOL0482 <strong>Staphylococcus</strong> tandem lipoprotein Y N N N N N N<br />

57651324 SACOL0483 staphylococcus tandem lipoprotein Y N N N N N N<br />

57651327 SACOL0486 staphylococcus tandem lipoprotein Y N N N N N N<br />

57652445 SACOL0081 hypothetical protein SA0081 Y N N N N N N<br />

150392787 SaurJH1_0313 hypothetical protein N Y Y N N N N<br />

150393510 SaurJH1_1042 hypothetical protein N Y Y N N N N<br />

87159856 pUSA03_0017 transfer complex protein TraH N N N N N N N<br />

87160691 SAUSA300_0411 staphylococcus tandem lipoprotein N N N N N N N<br />

87160733 SAUSA300_0079 putative lipoprotein N N N N N N N<br />

87161538 SAUSA300_0073 peptide ABC transporter, peptide-binding protein N N N N N N N<br />

151220617 NWMN_0405 truncated staphylococcal tandem lipoprotein N N N N N N N<br />

227


Appendix III Supplemental Table IIIh<br />

Supplemental Table IIIh. Continued<br />

PID Protein Function MW2 N315 NCTC 8325 Newman USA300 USA300 TCHC1516<br />

15925801 SA0093 hypothetical protein N Y N N N N<br />

15925802 SA0094 hypothetical protein N Y N N N N<br />

15925803 SA0095 hypothetical protein N Y Y Y Y Y<br />

15925804 SA0096 hypothetical protein N Y Y Y Y Y<br />

15925847 SA0138 hypothetical protein, similar to<br />

alkylphosphonate ABC tranporter<br />

Y Y Y N Y Y<br />

15925877 SA0167 hypothetical protein Y Y Y Y Y Y<br />

15926004 SA0291 hypothetical protein Y Y Y Y Y Y<br />

15926114 Lpl1 hypothetical protein N Y Y Y Y Y<br />

19526115 Lpl2 hypothetical protein N Y N N N N<br />

15926116 Lpl3 hypothetical protein N Y N N N N<br />

15926118 Lpl4 hypothetical protein N Y N N N N<br />

15926119 Lpl5 hypothetical protein N Y N N N N<br />

15926120 Lpl6 hypothetical protein Y Y N N N N<br />

15926122 Lpl8 hypothetical protein N Y Y Y Y Y<br />

15926123 Lpl9 hypothetical protein N Y N Y Y Y<br />

15926465 SA0743 hypothetical protein, similar to<br />

staphylocoagulase precursor<br />

Y Y Y Y Y Y<br />

15926580 SA0850 hypothetical protein, similar to<br />

oligopeptide ABC transporter<br />

oligopeptide-binding protein<br />

Y Y Y Y Y Y<br />

15927067 SA1317 hypothetical protein N Y N N N N<br />

15927068 SA1318 hypothetical protein Y Y N N N N<br />

15927069 SA1319 hypothetical protein Y Y N N N N<br />

15927071 SA1321 hypothetical protein N Y Y Y Y Y<br />

15927396 SA1640 conserved hypothetical protein N Y N N N N<br />

15928064 SA2273 hypothetical protein N Y N N N N<br />

16119210 BlaZ β-lactamase precursor N Y N N N Y<br />

14141829 SAVP006 hypothetical protein N N N N N N<br />

15923793 SAV0803 hypothetical protein Y N Y Y Y Y<br />

15924491 SAV2001 hypothetical protein N N N Y N N<br />

21281801 MW0072 hypothetical protein Y N Y Y Y Y<br />

228


Appendix III Supplemental Table IIIh<br />

PID Protein Function MW2 N315 NCTC 8325 Newman USA300 USA300 TCHC1516<br />

21282126 Lpl10 hypothetical protein Y N N N N N<br />

21282127 Lpl11 hypothetical protein Y N N Y Y Y<br />

21282129 Lpl13 hypothetical protein Y N N N N N<br />

21283103 MW1374 conserved hypothetical protein Y N N N N N<br />

21283167 MW1438 hypothetical protein Y N N N Y Y<br />

21283173 MW1444 hypothetical protein Y N Y Y Y Y<br />

21284135 MW2406 hypothetical protein Y N N N N N<br />

21284306 MW2577 hypothetical protein Y N N N N N<br />

49482670 SAR0445 putative lipoprotein N N N N N N<br />

49483474 SAR1288 putative lipoprotein N N N N Y Y<br />

49483672 SAR1494 hypothetical protein N N Y Y Y Y<br />

49484287 SAR2149 putative exported protein N N N N N N<br />

49484977 SAS0074 putative lipoprotein N N N N N N<br />

57650161 SACOL0888 pathogenicity island, putative lipoprotein N N N N N N<br />

57650444 SACOL1531 hypothetical protein N N Y Y Y Y<br />

57650485 SACOL1574 hypothetical protein SA1574 N N N N N N<br />

57650996 SACOL2497 staphylococcus tandem lipoprotein N N Y Y Y Y<br />

57651323 SACOL0482 <strong>Staphylococcus</strong> tandem lipoprotein N N N Y Y Y<br />

57651324 SACOL0483 staphylococcus tandem lipoprotein N N N Y Y Y<br />

57651327 SACOL0486 staphylococcus tandem lipoprotein N N Y Y Y Y<br />

57652445 SACOL0081 hypothetical protein SA0081 N N N N N N<br />

150392787 SaurJH1_0313 hypothetical protein N N N N N N<br />

150393510 SaurJH1_1042 hypothetical protein N N N N N N<br />

87159856 pUSA03_0017 transfer complex protein TraH N N N N Y N<br />

87160691 SAUSA300_0411 staphylococcus tandem lipoprotein N N N Y Y Y<br />

87160733 SAUSA300_0079 putative lipoprotein N N N N Y Y<br />

87161538 SAUSA300_0073 peptide ABC transporter, peptide-binding<br />

protein<br />

N N N N Y Y<br />

151220617 NWMN_0405 truncated staphylococcal tandem<br />

lipoprotein<br />

N N N Y N N<br />

229


Appendix III Supplemental Table IIIi<br />

Supplemental table IIIi. Specific S. epidermidis proteins with a predicted signal peptide<br />

Sec type signal peptide<br />

PID Protein -3 -2 -1 +1 Function<br />

27467163 SE0245 V H A A triacylglycerol lipase precursor<br />

27467176 SE0258 A K A Q immunodominant antigen B<br />

27467249 a,b SE0331 A K A E Ser-Asp rich fibrinogen-binding, bone sialoprotein-binding protein<br />

27467501 c SE0583 S F A N hypothetical protein<br />

27467545 c SE0627 T L A D poly D-alanine transfer protein<br />

27467746 a SE0828 A H A E lipoprotein VsaC<br />

27467938 c SE1020 I K A Q hypothetical protein<br />

27468418 a SE1500 S Y A Q hypothetical protein<br />

27468493 SE1575 A Q A H immunodominant antigen B<br />

27468546 a SE1628 V Y A D hypothetical protein<br />

27468838 SE1920 V Y A Q hypothetical protein<br />

27468875 SE1957 A H A T copper export proteins<br />

27469060 c SE2142 T L A F 2-dehydropantoate 2-reductase<br />

27469070 a SE2152 T H A A hypothetical protein<br />

27469115 c SE2197 S Y A S alkaline phosphatase III precursor<br />

27469119 SE2201 A D A Z phage-related protein<br />

27469291 SE2373 A Q A S 1,4-β-N-acetylmuramidase<br />

27469316 SE2398 S S A S hypothetical protein<br />

32470521 P601 A S A S hypothetical protein<br />

32470527 P607 I N A D hypothetical protein<br />

32470549 a P517 A K A E hypothetical protein<br />

32470583 P202 T F A L hypothetical protein<br />

Lipoprotein signal peptide<br />

PID Protein -3 -2 -1 +1 +2 Function<br />

27466952 SE0034 L S A C S hypothetical protein<br />

27467000 SE0082 L T A C G hypothetical protein<br />

27467062 SE0144 V S G C G hypothetical protein<br />

27467063 SE0145 V S G C G hypothetical protein<br />

27467067 d SE0149 L A G C D hypothetical protein<br />

27467309 d SE0391 L T T C S hypothetical protein<br />

27468024 SE1106 V T A C S ABC transporter<br />

27468425 d SE1507 L Y G C G hypothetical protein<br />

27469012 d SE2094 L I I C S hypothetical protein<br />

27469069 d SE2151 L A G C G hypothetical protein<br />

27469130 SE2212 V S G C S hypothetical protein<br />

27469141 d SE2223 L G S C S hypothetical protein<br />

27469317 SE2399 L S A C G hypothetical protein<br />

32470551 SE_p519 L A G C S hypothetical protein<br />

a Proteins with an LPxTG-motif<br />

b This protein has a lower SignalP score than our threshold score for all S. <strong>aureus</strong> strains<br />

c All S. <strong>aureus</strong> strains have one or more residues in the cleavage site, which are not included in the search pattern<br />

d Excluded for S. epidermidis because of the motif proposed by Sutcliffe and Harrington (2002)<br />

230


Appendix III Supplemental Table IIIj<br />

Supplemental table IIIj. Proteins with predicted Sec type signal peptides present in S.<br />

<strong>aureus</strong> RF122<br />

PID Protein -3 -2 -1 +1 Function<br />

82749800 a,b SAB0023 A R A E 5' nucleotidase<br />

82749803 c SAB0026 A H A S enterotoxin protein<br />

82749815 Plc S L A I 1-phosphatidylinositol phosphodiesterase<br />

82749847 a,d SasD A D A I surface protein<br />

82749858 a SAB0085 A D A A hypothetical protein<br />

82749938 a Coa A Q A S staphylocoagulase precursor<br />

82749981 a LytM A K A S peptidoglycan hydrolase<br />

82750007 a,e SAB0244 A I A K hypothetical protein<br />

82750020 a Geh V L A E glycerol ester hydrolase<br />

82750121 TSST-1 V Q A K toxic shock syndrome toxin-1<br />

82750124 SEC3 V H A E staphylococcal enterotoxin C-bovine<br />

82750136 Set6 V K A E superantigen-like protein<br />

82750137 Set7 V N A S superantigen-like protein<br />

82750138 Set8 V Q A K superantigen-like protein<br />

82750139 Set10 V N A K superantigen-like protein 5<br />

82750140 Set11 V H A K superantigen-like protein 7<br />

82750141 Set12 S H A K superantigen-like protein<br />

82750142 a Set13 V K A D superantigen-like protein<br />

82750143 a Set14 A E A S superantigen-like protein<br />

82750145 Set26 A D A K superantigen-like protein<br />

82750146 SAB0387 A E A A hypothetical protein<br />

82750147 a SAB0388c A Q A A hypothetical protein<br />

82750172 a,f Aaa A Q A S autolysin<br />

82750320 a SAB0566 A L A K hypothetical protein<br />

82750345 a Pbp4 A N A E penicillin binding protein 4<br />

82750367 a SAB0614c A S A V secretory antigen SsaA-like protein<br />

82750398 a SAB0645 A N A S hypothetical protein<br />

82750460 a,f SAB0708 A N A L hypothetical protein<br />

82750491 c SAB0739 A N A S hypothetical protein<br />

82750496 a,b,g ClfA A D A S truncated clumping factor<br />

82750497 SAB0745 A K A I secreted von Willebrand factor-binding protein precursor<br />

82750498 a Ssp T N A E extracellular matrix and plasma binding protein precursor<br />

82750499 a SAB0747 V D A A hypothetical protein<br />

82750500 a Nuc A G A E staphylococcal thermonuclease precursor<br />

82750530 SAB0780 V S A A phage-associated holin<br />

82750532 LukS A K A D leukocidin chain lukM precursor<br />

82750533 LukF A N A L Panton-Valentine leukocidin LukF-PV chain precursor<br />

82750575 a GlpQ V Q A A glycerophosphoryl diester phosphodiesterase<br />

82750593 a SAB0846 A D A T hypothetical protein<br />

82750659 a SspB S H A Q cysteine protease precursor<br />

82750660 a SspA A K A F glutamyl endopeptidase serine protease<br />

82750663 a,f Atl A D A S autolysin<br />

82750672 a SAB0928c A D A S hypothetical protein<br />

82750736 a,b IsdB A N A A iron-regulated cell wall-anchored protein<br />

82750737 a,b IsdA A K A Y cell surface transferrin-binding protein<br />

82750738 a,h IsbC A K A Y iron-regulated cell surface protein<br />

82750761 a SAB1018 A K A Y hypothetical protein<br />

82750762 SAB1019c A H A A formyl peptide receptor-like 1 inhibitory protein<br />

82750764 a Efb V L A E fibrinogen-binding protein<br />

82750765 a SAB1022 I D A S fibrinogen-binding protein precursor<br />

82750770 Hly A Q A A α-hemolysin precursor<br />

82750773 a SAB1030c A N A E superantigen-like protein<br />

82750774 a SAB1031c A K A D superantigen-like protein<br />

82750775 a SAB1032c A S A S superantigen-like protein<br />

231


Appendix III Supplemental Table IIIj<br />

PID Protein -3 -2 -1 +1 Function<br />

82751035 EbhB V D A A truncated cell surface fibronectin-binding protein<br />

82751071 a GpsA S R A N glycerol-3-phosphate dehydrogenase<br />

82751144 SAB1412c V N A E hypothetical protein<br />

82751320 i Fhs S Y A Q surface-anchored iron-regulated surface protein<br />

82751393 j SplE D K A E serine proteinase<br />

82751394 a SplC A Y A N serine proteinase<br />

82751395 k SplB A K A E serine proteinase<br />

82751397 SAB1675 S L A D hypothetical protein<br />

82751401 EpiP A N A E serine protease precursor<br />

82751408 LukD A F A Y leukotoxin D subunit<br />

82751409 LukE A S A A leukotoxin E subunit<br />

82751419 SEN A K A E enterotoxin N<br />

82751421 SEI S Y A K enterotoxin I<br />

82751422 SEO T Q A N enterotoxin O<br />

82751535 a SAB1814 V E A K hypothetical protein<br />

82751566 a SspB2 A H A S staphopain cysteine proteinase<br />

82751581 a SAB1860c A S A D hypothetical protein<br />

82751594 Map A H A S cell surface protein<br />

82751595 SAB1874 A D A A β-hemolysin<br />

82751596 a SAB1875c V A A K leukocidin F subunit<br />

82751597 a SAB1876c A K A S leukocidin S subunit<br />

82751634 a SAB1916c S K A E membrane anchored Ser-Asp rich fibrinogen-binding protein<br />

82751697 a SAB1980c A K A A hypothetical protein<br />

82751757 b FmtB A N A E truncated methicillin resistance-related surface protein<br />

82751802 a SAB2085 A Y A H hypothetical protein<br />

82751888 a SsaA S H A A secretory antigen precursor<br />

82751892 a SAB2176 V K A K exported secretory antigen precursor<br />

82751903 a SAB2187c A Q A A transcriptional regulator<br />

82752013 a Sbi A H A A immunoglobulin G-binding protein<br />

82752015 a HlgA A L A I γ-hemolysin component A precursor<br />

82752016 a HlgC I Y A A γ-hemolysin component C<br />

82752017 a HlgB A Y A D γ-hemolysin component B<br />

82752118 a SAB2409c A N A D hypothetical protein<br />

82752126 a SAB2418 A Q A T secretory antigen SsaA-like protein<br />

82752129 c SAB2421c A N A A hypothetical protein<br />

82752147 a SAB2439c A H A Q secretory antigen precursor<br />

82752151 a IsaA S N A A immunodominant antigen A<br />

82752211 a,b ClfB A Q A A clumping factor B<br />

82752218 a Aur V N A A zinc metalloproteinase aureolysin<br />

82752220 a SAB2514 A S A A hypothetical protein<br />

82752225 a SAB2519 A Q A S N-acetylmuramoyl-L-alanine amidase<br />

82752227 a,c SasF A Q A A surface anchored protein<br />

82752246 a IcaB A H A D intercellular adhesion protein B<br />

82752249 a Lip A N A A triacylglycerol lipase precursor<br />

a<br />

Proteins belonging to the S. <strong>aureus</strong> core proteome<br />

b<br />

Proteins with an LPxTG-motif<br />

c<br />

Proteins that are only present in S. <strong>aureus</strong> RF122<br />

d<br />

Proteins with an LPxAG-motif<br />

e<br />

<strong>The</strong> S. <strong>aureus</strong> RF122 protein has an Ile on the -2 position, which is not included in the recognition and<br />

cleavage pattern; all other S. <strong>aureus</strong> strains are conform the proposed pattern<br />

f<br />

Proteins with a LysM or GW domain motif<br />

g<br />

Only the S. <strong>aureus</strong> RF122 protein is truncated at the N-terminus, thereby missing part of the signal peptide<br />

h<br />

<strong>The</strong> S. <strong>aureus</strong> RF122 protein is truncated at the C-terminus, therefor missing the LPxTG motif<br />

i<br />

<strong>The</strong> S. <strong>aureus</strong> RF122 protein has an Asp on the -3 position , which is not included in the<br />

j<br />

<strong>The</strong> S. <strong>aureus</strong> RF122 protein has an Asp on the -3 position , which is not included in the recognition and<br />

cleavage pattern; the MRSA252 protein is conform the proposed pattern<br />

k<br />

Only the S. <strong>aureus</strong> RF122 protein is predicted to have two membrane domains<br />

232


Appendix III Supplemental Table IIIj<br />

Supplemental table IVa. Characteristics of the clinical S. <strong>aureus</strong> isolates used in this study<br />

Patient Age Isolate Related Acquired<br />

isolates<br />

b Date of Origin Infection / Antibiotic resistance others MLST CC agrisolation<br />

Colonization<br />

Type<br />

1 2 A B398 C 04.07.2003 umbilicus phlegmone ciprofloxacin, erythromycin, MSSA, pvl- 398 398 1<br />

penicillin<br />

positive<br />

2 35 B B398 C 17.06.2003 pleural pneumonia ciprofloxacin, erythromycin, caMRSA, 398 398 1<br />

fluid<br />

oxacillin, penicillin<br />

pvl-positive<br />

3 76 C G228 a H 23.01.2004 nose colonization ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin<br />

MRSA 228 5 2<br />

D G228 a H 23.12.2004 nose colonization amoxicillin-clavulanate,<br />

ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin<br />

MRSA 228 5 2<br />

E E8 a H 13.11.2003 feet chronic wound ciprofloxacin, co-trimoxazole, MSSA 8 8 1<br />

infection<br />

penicillin<br />

4 62 F E8 a H 16.12.2003 blood postoperative wound ciprofloxacin, co-trimoxazole, MSSA 8 8 1<br />

infection, sepsis erythromycin, penicillin<br />

G G228 a H 17.12.2003 blood postoperative wound amoxicillin-clavulanate,<br />

MRSA 228 5 2<br />

infection, sepsis ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin<br />

5 74 H G228 a H 06.01.2004 blood postoperative wound amoxicillin-clavulanate,<br />

MRSA 228 5 2<br />

infection, sepsis ciprofloxacin, co-trimoxazole,<br />

erythromycin, gentamicin, oxacillin,<br />

penicillin,<br />

I G228 a H 12.01.2005 perineum colonization ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin<br />

MRSA 228 5 2<br />

6 62 J G228 a H 12.01.2004 perineum colonization amoxicillin-clavulanate,<br />

ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin,<br />

rifampin<br />

MRSA 228 5 2<br />

K G228 a H 28.09.2004 feet chronic wound amoxicillin-clavulanate,<br />

MRSA 228 5 2<br />

infection<br />

ciprofloxacin, erythromycin,<br />

gentamicin, oxacillin, penicillin<br />

7 32 L L80 a C 13.04.2002 buttock abscess oxacillin, penicillin, tetracycline caMRSA,<br />

pvl-positive<br />

80 3<br />

8 2 M L80 a C 13.04.2004 finger panaritium oxacillin, penicillin, tetracycline caMRSA,<br />

pvl-positive<br />

80 3<br />

9 57 N C 06.01.2004 blood pneumonia, sepsis no resistance MSSA 15 15 2<br />

10 65 O C 09.02.2004 blood pneumonia, sepsis ciprofloxacin, erythromycin,<br />

penicillin<br />

233<br />

MSSA 5 5 2


Appendix IV Supplemental Table IVa<br />

Patient Age Isolate Related Acquired<br />

isolates<br />

b Date of Origin Infection / Antibiotic resistance others MLST CC agrisolation<br />

Colonization<br />

Type<br />

11 72 P C 29.09.2004 blood sepsis Penicillin MSSA 7 1<br />

12 47 Q C 15.01.2004 blood sepsis, arthritis Penicillin MSSA 30 30 3<br />

13 47 R C 16.07.2005 blood sepsis penicillin, tetracyclin MSSA 15 15 2<br />

14 22 S C 03.07.2005 blood sepsis Erythromycin MSSA 8 8 1<br />

15 54 T C 06.02.2001 blood sepsis, rheumatoid Penicillin MSSA 870 8 1<br />

arthritis<br />

16 72 U C 19.03.2005 blood sepsis, meningitis,<br />

probably<br />

endocarditis<br />

Penicillin MSSA 5 5 2<br />

17 13 V C 12.06.2005 blood septic arthritis Penicillin MSSA 903 5 2<br />

18 4 W C 25.09.2005 blood septic arthritis Ciprofloxacin, penicillin MSSA 869 2<br />

19 34 X X8 H 15.09.2005 throat colonization, ciprofloxacin, erythromycine, MSSA 8 8 1<br />

cholangitis<br />

Y X8 H 15.09.2005 nose colonization,<br />

cholangitis<br />

a strains involved in hospital outbreaks<br />

b Community (C); Hospital (H)<br />

234<br />

penicillin, tetracycline<br />

ciprofloxacin, erythromycine,<br />

oxacillin, penicillin, tetracycline<br />

MRSA 8 8 1


Appendix IV Supplemental Table IVb<br />

Supplemental table IVb. Identified proteins on 2D gels of the S. <strong>aureus</strong> isolates<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

AckA 15927288 44,60 5,70 cytosolic + +<br />

AhpC 13700295 20,96 4,88 cytosolic + + + + + + + + + + + + + + + + + + + +<br />

AhpF 13700294 54,70 6,70 cytosolic +<br />

Ald 13701504 40,05 5,58 cytosolic +<br />

Aly 13702602 69,19 5,96 66,3 5,8 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

Asp23 13701982 19,18 5,13 unknown + + + + + + + +<br />

AtlE 13700854 136,67 9,60 133,7 9,6 + extracellular + + + + + + + + + + + + + + + +<br />

Aur 13702595 56,34 5,14 33,4 4,8 + extracellular + + + + + + + + + + + + + + + + +<br />

Bbp 49482792 123,30 4,30 117,6 4,2 + cellwall + +<br />

BlaR1 16119211 69,00 9,50 membrane + + + + +<br />

CitB 13701147 98,91 4,83 cytosolic + + + + + + + + + + + + + + + + +<br />

ClfB 15925620 93,60 3,92 88,7 3,8 + cellwall + + + + + + +<br />

ClpC 13700415 90,98 5,51 cytosolic + +<br />

ClpP 13700659 21,50 5,13 cytosolic +<br />

Cna 49484887 133,00 5,90 169,7 5,8 + cellwall +<br />

Coa 13700145 74,50 8,25 129,7 5,8 + extracelular + + + + + + + +<br />

CspA 13701199 7,32 4,52 cytosolic + + + + + + + +<br />

CysK 13700403 33,00 5,20 cytosolic +<br />

DeoD 13701932 25,89 4,85 unknown + + + + + + + + + + + +<br />

Ddh 15928103 39,30 5,40 cytosolic +<br />

Dnak 13701378 66,32 4,65 cytosolic + + + + + + + + + +<br />

Ear 57650605 20,40 7,70 16 5,2 + unknown +<br />

EbpS 13701280 53,15 5,97 membrane + + + + + + + + + + + + + + + + + +<br />

Edin-B 24636605 27,50 9,38 23,64 9,3 + extracellular + +<br />

EF-G 13700438 76,56 4,80 cytosolic + + + + + + + + + + + + +<br />

EF-TS 13701057 32,50 5,15 cytosolic + + + + + + +<br />

EF-TU 13700439 43,08 4,74 cytosolic + + + + + + + + + + + + + + +<br />

Eno 13700667 47,09 4,55 cytosolic + + + + + + + + + + + + + + + + + + +<br />

Etd 24636603 30,80 8,90 27,9 8 + extracellular + +<br />

FabH 13700787 33,90 4,90 unknown + +<br />

FbaA 13701919 30,82 5,01 cytosolic + + + + + + + + + + + + + + + + +<br />

Fhs 13701527 59,83 5,69 cytosolic + + + + + + + + + + + + + + + +<br />

Fnb 15928082 113,60 4,60 109,7 4,5 + cellwall + +<br />

FnbA 57651010 109,10 4,56 107,8 5,6 + cellwall + + + + +<br />

235


Appendix IV Supplemental Table IVb<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

FnbB 57651008 101,00 4,60 99,7 4,6 + cellwall + + +<br />

Gap 13700663 36,26 4,89 cytosolic + + + + + + + + + + + + + + + + + + + +<br />

Geh 13700235 76,50 8,99 72,4 8,9 + extracellular + + + + + + + + + + + + + + + + + +<br />

GlnA 13701109 50,80 5,00 cytosolic + + + +<br />

GlpQ 13700763 35,29 8,67 32,2 8 + extracellular/wall + + + + + + + + + + + + + + + + + + +<br />

GltX 13700418 56,25 5,21 cytosolic + + +<br />

GlyA 13701907 45,14 5,75 cytosolic + + + +<br />

Gnd 13701310 51,80 4,80 cytosolic + + +<br />

GreA 13701408 17,70 4,50 cytosolic +<br />

GroEL 13701823 57,54 4,56 cytosolic + + + + + + + + +<br />

GroES 13701824 10,40 4,60 cytosolic + +<br />

GrpE 13701379 24,00 4,10 cytosolic +<br />

GuaA 13700305 58,17 5,03 cytosolic + + + + +<br />

GuaB 13700304 52,82 5,61 cytosolic + + + + + + + + + +<br />

Hla 13700962 35,95 8,70 33 7,9 + extracellular + + + + + + + + + + + + + + + + + + + + +<br />

Hlb 13701798 33,04 8,56 33,7 7,3 + extracellular + +<br />

HlgA 13702368 32,00 9,50 31,9 9,3 + extracellular + +<br />

HlgB 13702370 36,69 9,35 33,9 9,3 + extracellular + + + + + + + + + + + + + + + + +<br />

HlgC 13702369 35,56 9,29 32,5 9,1 + extracellular + + + + + + + + + + + + + +<br />

Hpr 13700884 9,49 4,50 cytosolic + +<br />

IleS 13700992 105,00 5,20 cytosolic + + + + + + + +<br />

IsaA 13702519 24,19 5,90 21,5 5,3 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

IsdA 15926714 38,70 9,60 33,7 9,6 + cellwall + +<br />

IsdB 49483291 73,00 9,00 68,6 8,8 + cellwall + +<br />

IsdC 15926715 24,90 8,90 21,8 8,5 + unknown + + + +<br />

IsdD 15926716 41,50 8,80 37,9 8,3 + unknown +<br />

KatA 13701132 58,58 5,27 cytosolic + + + + + + + + +<br />

Lip 13702629 76,62 6,58 21,5 5,3 + extracellular + + + + + + + + + + + + + + + + + + + + + + +<br />

LukD 13701612 36,88 9,21 34,2 9 + extracellular + + + + + +<br />

LukE 13701613 34,80 9,48 31,9 9,3 + extracellular + + +<br />

LytM 13700191 34,30 6,20 31,7 6 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

MetS 13700380 74,80 4,95 cytosolic + +<br />

MvaS 13702496 43,00 4,90 cytosolic +<br />

Nuc 13700682 25,07 9,21 18,8 9,3 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

OdhA 13701210 103,05 5,47 cytosolic + + + + +<br />

PanB 13702556 29,20 5,80 unknown +<br />

236


Appendix IV Supplemental Table IVb<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

Pbp2 57650405 80,40 8,70 74,4 6,6 + extracellular + +<br />

Pbp3 13701350 77,18 9,22 membrane + + + + + + + + + + + + + + + + +<br />

PdhA 57651702 41,36 4,90 cytosolic + + + + + + +<br />

PdhB 13700894 35,22 4,65 cytosolic + + + + + + + + + + + + + + + + + +<br />

PdhC 13700895 46,30 4,90 cytosolic + + + + +<br />

PdhD 13700896 49,40 4,95 cytosolic + + + + + + + + + + + + + + + + + + + + +<br />

Pgi 13700766 49,76 4,82 cytosolic + + + + + + + + + + + + + + + + +<br />

Pgk 13700664 42,58 5,17 cytosolic + + + + + + + + + + + + +<br />

Pgm 13700666 56,42 4,74 cytosolic + +<br />

PheT 13700939 88,88 4,66 cytosolic + +<br />

Plc 13700011 37,06 7,71 34,2 6,4 + extracellular + + + + + + + + + + + + + + + + +<br />

Pls 12644358 174,50 4,06 159,6 4 + cellwall + + + + + + +<br />

PnpA 13701074 77,31 4,89 cytosolic + +<br />

Pta 13700480 34,93 4,72 cytosolic + + +<br />

PtsI 13700885 63,18 4,62 cytosolic + + + + +<br />

PurM 13700872 37,00 4,50 unknown + + +<br />

PycA 13700915 128,60 5,00 cytosolic + + + + + + +<br />

RplA 15926216 24,70 9,00 cytosolic + + + + +<br />

RplM 13702018 16,32 9,30 unknown + + + + +<br />

RplY 15926178 23,80 4,40 cytosolic + + + + +<br />

RpsA 13701275 43,30 4,40 cytosolic + + + + + + + + + + + + + + +<br />

RpsB 15926839 29,10 5,40 cytosolic +<br />

SA0022 13699940 83,45 9,24 80,5 9,2 + cellwall + + + + + +<br />

SA0092 13700012 29,72 8,24 27,1 6,3 + extracellular + +<br />

SA0129 15925838 26,50 9,60 23,6 9,3 + cellwall + +<br />

SA0139 13700060 56,20 7,00 53,4 6 + unknown + + +<br />

SA0182 13700104 60,49 5,10 unknown + +<br />

SA0269 15925982 57,80 7,50 58 7,5 + unknown + +<br />

SA0295 13700221 33,33 9,49 30,2 9,4 + unknown + +<br />

SA0357 15926071 23,20 9,50 19,3 9,4 + unknown +<br />

SA0270 13700196 33,07 5,85 30,5 5,6 + extracellular + + + + + + + + + + +<br />

SA0272 15925985 114,70 6,20 membrane + + + + +<br />

SA0423 13700355 35,81 9,67 33,4 9,6 + extracellular + + + + + + + + + + + + +<br />

SACOL0468 57651309 25,60 8,50 22,4 6,3 + extracellular + +<br />

SA0359 13700286 21,26 5,70 unknown + + + + + + +<br />

SA0394 13700325 55,46 5,07 51,9 5 + unknown + +<br />

237


Appendix IV Supplemental Table IVb<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

SACOL0478 57651319 25,40 9,10 22,3 6 + extracellular + + + +<br />

SACOL0479 57651320 56,40 4,80 52,9 4,7 + unknown + + + + + + + + + +<br />

SA0516 15926236 17,10 8,30 cytosolic +<br />

SA0570 13700505 18,60 9,17 15,9 9,2 + unknown + + + + + + + + + + + + + + + + + + + + + + +<br />

SACOL0613 57651442 32,64 4,85 cytosolic +<br />

SA0620 13700556 28,17 6,13 25,6 5,6 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

SA0651 13700587 16,91 8,48 13,5 9 + unknown + + +<br />

SA0663 13700599 16,04 9,15 unknown + +<br />

SA0695 13700631 34,06 9,14 31,2 9,1 + extracellular + +<br />

SA0707 13700643 22,20 5,15 cytosolic +<br />

SA0775 13700717 48,49 5,44 cytosolic + +<br />

SA0815 13700758 21,60 4,30 unknown +<br />

SA0841 13700785 15,89 9,28 12,9 9,2 + unknown + + + + + + + + + +<br />

SA0859 13700804 69,80 5,00 cytosolic + + +<br />

SA0873 13700818 19,30 4,70 cytosolic +<br />

SACOL0908 57651598 20,20 9,18 16,2 6,6 + unknown + +<br />

SA0908 13700857 45,68 6,02 41,8 5,8 + unknown + + + + + + + + + + + + +<br />

SA1056 15926796 35,80 7,60 unknown +<br />

SA1173 13701136 39,50 9,40 unknown +<br />

SA1255 13701220 17,95 4,52 cytosolic +<br />

SA1257 13701222 20,60 6,37 cytosolic +<br />

SA1336 13701304 56,90 5,30 cytosolic + + + +<br />

SA0914 13700863 11,30 7,50 8,7 6,5 + unknown +<br />

SA0998 13700951 21,40 4,91 cytosolic +<br />

SA1001 13700955 15,20 9,12 12,3 8 + extracellular +<br />

SA1475 13701447 30,99 9,04 unknown + + + + + + + + + + + + +<br />

SACOL1528 57650441 34,60 6,40 32,7 5,7 unknown + +<br />

SA1532 13701505 13,90 5,15 cytosolic + + +<br />

SA1737 13701714 38,50 4,90 cytosolic + +<br />

SA1743 15927503 22,00 5,20 unknown +<br />

SA1755 13701736 17,06 9,45 14,1 9,1 + extracellular +<br />

SA1774 15927539 42,20 5,30 unknown + + + +<br />

SA1812 13701799 38,64 8,58 38,7 8,6 + extracellular + + + + + + + + + + + + + + + + + + +<br />

SA1572 13701547 52,79 4,58 49,1 4,6 + extracellular + + + +<br />

SA1633 13701608 20,86 8,48 16,5 5,3 + unknown + + + + + + + + +<br />

SA1813 13701800 40,44 9,43 37,4 9,3 + extracellular + + + + + + + + +<br />

238


Appendix IV Supplemental Table IVb<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

SA1552 15927308 104,00 5,10 96 4,9 + cellwall +<br />

SA1839 13701826 46,20 8,70 42,7 7,8 + cellwall + + + + + + + +<br />

SACOL1870 57650605 20,30 6,80 20,4 6,8 + unknown + +<br />

SA1979 15927757 36,60 9,40 29,3 9,4 + unknown +<br />

SA2006 13702005 15,42 9,22 12,5 9,1 + unknown + + + + + + + +<br />

SA2097 13702104 17,39 5,77 14,7 5,1 + extracellular + + + + + + + + + + + + + + + + + + + + + + + + +<br />

SA2367 15928160 31,00 4,70 cytosolic + + +<br />

SA2399 13702563 33,00 4,90 unknown + + + + +<br />

SA2490 13702656 30,70 6,40 unknown +<br />

SACOL2505 57651004 136,30 5,70 130,7 5,5 + cellwall + + + +<br />

SAR0106 49482346 29,70 9,10 26,8 8,8 + unknown +<br />

SAR0422 49482650 25,70 8,90 22,7 7,9 + extracellular +<br />

SA2103 13702110 34,70 9,70 30,8 6,8 + unknown + +<br />

SA2204 13702365 26,70 5,10 unknown + +<br />

SA2279 13702441 66,10 6,50 cytosolic +<br />

SA2285 15928076 48,90 8,60 43,4 6,4 + cellwall + + +<br />

SAR0435 49482662 26,40 8,70 23,3 6,5 + extracellular + +<br />

SAR0436 49482663 55,50 4,90 51,9 4,8 + extracellular + +<br />

SAR1905 49484059 25,60 8,80 21,9 8,1 + extracellular +<br />

SAR2788 49484898 27,70 9,70 25 9,7 + unknown +<br />

SAS074 13701973 10,00 6,00 unknown +<br />

SAS2383 49487275 149,90 5,70 144,4 5,5 + cellwall +<br />

Sak 13701739 18,51 6,75 15,5 6,2 + extracellular + + + + + + + + + + + + + + + + + +<br />

SasH 38259869 17,10 7,90 unknown + + +<br />

Sbi 13702367 50,00 9,38 50,2 9,4 + unknown + + + + + + + + + + +<br />

SceD 13701890 24,06 5,53 21,5 5,1 + extracellular + + + + + + + + + + + + +<br />

Sed 758691 29,70 8,60 26,9 7,2 + extracellular + +<br />

Set1 49482656 26,00 6,90 23 6,2 + extracellular + + + +<br />

Set6 13700312 25,67 9,05 22,7 8,5 + extracellular + + + + + + +<br />

Set11 15926104 26,20 9,20 23,2 9,1 + extracellular + + +<br />

Set15 13700324 25,42 8,79 22,4 6,5 + extracellular + + + +<br />

SodA 13701351 22,70 5,08 cytosolic + + + + + + + + + + + +<br />

Sea 49484190 29,70 8,30 26,9 7,2 + extracellular + + + + +<br />

Seb 57651597 31,40 8,90 28,4 8,3 + extracellular + + + +<br />

Spa 13700027 48,84 5,60 45,3 5,2 + cellwall + + + + + + + + + + + + + + + + + + + + + +<br />

SplA 13701606 25,40 9,60 21,7 8,8 + extracellular + +<br />

239


Appendix IV Supplemental Table IVb<br />

precursor mature<br />

Protein Accession # MW (kD) pI MW (kD) pI SP localization A B C D E F G H I J K L M N O P Q R S T U V W X Y<br />

SdrD 57651438 149,40 4,10 143,8 4,1 + cellwall + + +<br />

SdrE 15926241 123,95 4,24 118,3 4,2 + cellwall + + + + + + + + + +<br />

Sek 57650159 27,71 8,30 25,3 6,6 + extracellular + + +<br />

Seq 15625528 28,17 8,33 25,1 6,6 + extracellular + + +<br />

Sep 13701743 29,70 7,60 26,3 6,2 extracellular + + +<br />

Ser 37196678 30,00 8,90 27,1 8,8 + extracellular + +<br />

SerS 15925714 48,60 5,00 cytosolic + +<br />

SplB 13701605 26,13 9,11 22,4 9 + extracellular + + + + + + + + + +<br />

SplC 13701604 26,08 6,32 22,4 6,4 + extracellular + + + + + + + +<br />

SplD 13701603 25,70 9,12 22 8,9 + extracellular + + +<br />

SplE 57650600 25,70 9,80 22 9,2 + extracellular + + +<br />

SplF 13701602 25,63 9,16 21,9 8,9 + extracellular + + + + + + + + + + +<br />

SsaA 13702099 29,31 8,96 26,7 8,7 + extracellular + + + + + + + + + + + + + + + + +<br />

SspA 13700850 36,95 5,00 33,4 4,8 + extracellular + + + + + + + + + + + + + + + + + + + + + + +<br />

SspB 13700849 44,57 5,68 40,7 5,3 + extracellular + + + + + + + + + + + + + + + + +<br />

Stp 13701702 44,18 9,64 41,5 9,6 + extracellular + + + + + + + + + + + + + +<br />

SucD 13701046 31,52 5,47 cytosolic +<br />

Tig 13701472 48,58 4,34 cytosolic + + + + + + + + + + +<br />

Tkt 13701140 72,21 4,97 unknown + + + + + + + + + + + + + + + + + + +<br />

Tpi 13700665 27,24 4,80 unknown + + + + +<br />

Trap 13701628 19,54 6,13 cytosolic + + +<br />

TrxB 13700655 33,60 5,20 cytosolic + + + + + + + + + +<br />

ValS 13701460 102,00 4,80 cytosolic + + +<br />

YfnI 13700610 74,35 9,04 71/49.3 8.9/ + membrane + + + + + + + + + + + + + + + + + + + + +<br />

8.4<br />

18920604 18920604 21,30 5,30 unknown +<br />

24636604 24636604 25,40 7,00 22,8 6,6 + extracellular + +<br />

90585739 90585739 20,30 9,10 16,3 6,8 + extracellular +<br />

240


Appendix IV Supplemental Table IVb<br />

Supplemental table IVc. Extracellular proteins in different S. <strong>aureus</strong> isolates<br />

Name Function<br />

Extracellular proteins identified in at least 80% of the strains<br />

Aly hypothetical protein, similar to autolysin precursor<br />

Aur zinc metalloproteinase aureolysin<br />

Geh glycerol ester hydrolase<br />

GlpQ glycerophosphoryl diester phosphodiesterase<br />

Hla α-hemolysin precursor<br />

HlgB γ-hemolysin component B<br />

IsaA immunodominant antigen A<br />

Lip triacylglycerol lipase precursor<br />

LytM peptidoglycan hydrolase<br />

Nuc staphylococcal nuclease<br />

SA0570 hypothetical protein<br />

SA0620 secretory antigen SsaA homologue<br />

SA1812 hypothetical protein, similar to synergohymenotropic toxin precursor S. intermedius<br />

SA2097 hypothetical protein, similar to secretory antigen precursor SsaA<br />

SspA serine protease; V8 protease; glutamyl endopeptidase<br />

SspB cysteine protease precursor<br />

Extracellular proteins identified in less than 80% of the strains<br />

Atl autolysin<br />

Coa staphylocoagulase precursor<br />

HlgC γ-hemolysin component C<br />

LukD leukotoxin, LukD<br />

Plc 1-phosphatidylinositol phosphodiesterase precurosr<br />

SA0270 hypothetical protein, similar to secretory antigen precursor SsaA<br />

SA0423 hypothetical protein, similar to autolysin (N-acetylmuramoyl-L-alanine amidase)<br />

SACOL0479 surface protein, putative<br />

SA0841 hypothetical protein, similar to cell surface protein Map-w<br />

SA0908 conserved hypothetical protein<br />

SA1633 probable β-lactamase<br />

SA1813 hypothetical protein, similar to leukocidin chain lukM precursor<br />

SA2006 hypothetical protein, similar to MHC class II analog<br />

Sak staphylokinase precursor<br />

Sbi IgG-binding protein SBI<br />

SceD hypothetical protein, similar to SceD precursor<br />

Sea enterotoxin type A precursor<br />

Set1 exotoxin 1<br />

Set6 exotoxin 6<br />

SplB serine protease SplB<br />

SplC serine protease SplC<br />

SplF serine protease SplF<br />

SsaA secretory antigen precursor SsaA homolog<br />

Stp staphopain, cysteine proteinase<br />

Extracellular proteins identified in less than 20% of the strains<br />

Ear β-lactamase, putative<br />

Edin-B epidermal cell differentiation inhibitor B<br />

Etd exfoliative toxin D<br />

Hlb truncated β-hemolysin<br />

HlgA γ-hemolysin component A<br />

IsdC hypothetical protein SA0978<br />

IsdD isdD; hypothetical protein SA0979<br />

LukE leukotoxin LukE<br />

SA0092 hypothetical protein<br />

SA0139 conserved hypothetical protein<br />

SA0269 hypothetical protein SA0269<br />

SA0295 hypothetical protein, similar to outer membrane protein precursor<br />

241


Appendix IV Supplemental Table IVc<br />

Name Function<br />

Extracellular proteins identified in less than 20% of the strains<br />

SA0357 hypothetical protein SA0357<br />

SA0394 hypothetical protein<br />

SA0651 hypothetical protein<br />

SA0695 hypothetical protein<br />

SA0914 hypothetical protein, similar to chitinase B<br />

SA1001 hypothetical protein<br />

SA1755 hypothetical protein<br />

SA2103 hypothetical protein, similar to lyt divergon expression<br />

SACOL0468 exotoxin 3, putative<br />

SACOL0478 exotoxin 3, putative<br />

SACOL0908 hypothetical protein<br />

SACOL1870 hypothetical protein SACOL1870<br />

SAR0106 putative lipoprotein<br />

SAR0422 exotoxin<br />

SAR0435 exotoxin<br />

SAR0436 hypothetical protein SAR0436<br />

SAR1905 serine protease<br />

SAR2788 hypothetical protein SAR2788<br />

SEB staphylococcal enterotoxin B<br />

SED enterotoxin D precursor<br />

SEK staphylococcal enterotoxin<br />

SEP enterotoxin P<br />

SEQ staphylococcal enterotoxin type I<br />

SER enterotoxin R<br />

Set11 exotoxin 11<br />

Set15 exotoxin 15<br />

SplA serine protease SplA<br />

SplD serine protease SplD<br />

SplE serine protease SplE, putative<br />

24636604 probable glutamyl-endopeptidase<br />

90585739 conserved hypothetical protein<br />

242


Appendix V Supplemental Table Va<br />

Supplemental table Va. Proteins of B. licheniformis DSM 13 with a predicted signal<br />

peptide<br />

ID number Name Cleavage Site Sequence Function<br />

Secretory proteins<br />

BLi00015 DacA 31 32 AKA AN D-alanyl-D-alanine carboxypeptidase<br />

(penicillin-binding protein 5)<br />

BLi00171 CwlD 27 28 FNN DD Germination-specific N-acetylmuramoyl-Lalanine<br />

amidase (EC 3.5.1.28)<br />

BLi00181 PbpX 42 43 GMR DH penicillin-binding protein<br />

BLi00186 YbbC 23 24 AAA FP unknown<br />

BLi00187 YbbD 26 27 REA EA unknown; similar to β-hexosaminidase<br />

BLi00188 YbbE 23 24 AQT AI unknown; similar to β-lactamase<br />

BLi00223 YflP 38 39 VPA EP unknown<br />

BLi00238 YrkA1 25 26 EFA IV unknown; similar to hemolysin-like<br />

BLi00255 33 34 LSE LT unknown<br />

BLi00281 PhoD 46 47 VNA AP phosphodiesterase/alkaline phosphatase<br />

(EC3.1.3.1)<br />

BLi00302 YbdN 21 22 AFS AS unknown<br />

BLi00321 YcdA 30 31 ASG EK unknown<br />

BLi00338 32 33 AKA DS putative chitinase (EC 3.2.1.14)<br />

BLi00339 26 27 ISA ET putative chitinase (EC 3.2.1.14)<br />

BLi00340 Mpr 30 31 AQA AP glutamyl endopeptidase precursor (EC<br />

3.4.21.19)(glutamate specific endopeptidase)<br />

BLi00347 YvcE 30 31 ASA ET unknown; similar to cell wall-binding protein<br />

BLi00411 24 25 IHA QE unknown<br />

BLi00439 30 31 AGS AE putative sugar ABC transporter,<br />

periplasmicsugar-binding protein<br />

BLi00448 Phy 29 30 AEA SA 3-phytase (EC 3.1.3.8) / 6-phytase (EC 3.1.3.26)<br />

BLi00478 25 26 AEE QT unknown<br />

BLi00514 33 34 IAA AG putative transcriptional regulator, LytR family<br />

BLi00628 YoaR 19 20 GHS DS unknown<br />

BLi00654 30 31 INA SQ unknown<br />

BLi00656 29 30 AAA NL α-amylase precursor (EC 3.2.1.1)<br />

BLi00668 36 37 VKA SS hypothetical protein<br />

BLi00669 27 28 SSA SD unknown<br />

BLi00670 YdjM 27 28 ASA KT unknown<br />

BLi00671 YdjN 20 21 AFA AV unknown<br />

BLi00702 PurN 17 18 FEA IE phosphoribosylglycinamide formyltransferase<br />

BLi00712 YerB 27 28 EQQ EK unknown<br />

BLi00735 YdhT 24 25 SYA HT unknown; similar to mannan endo-1,4-βmannosidase<br />

BLi00784 28 29 VYA AE unknown<br />

BLi00824 YfkD 25 26 ADA AK unknown<br />

BLi00827 YfjS 24 25 AEA IS unknown; similar to polysaccharide deacetylase<br />

BLi00837 28 29 FNG NP unknown<br />

BLi00840 YckD 26 27 AYG ET unknown<br />

BLi00866 24 25 AST EE unknown<br />

BLi00967 27 28 ASS QD unknown<br />

BLi00976 25 26 AFS PE unknown<br />

BLi00979 YhcP 29 30 VFS QE unknown<br />

BLi00982 YhcR 33 34 THA SE unknown; similar to 5'-nucleotidase<br />

BLi00983 YhcS 19 20 TFA YG unknown<br />

BLi01008 LytE 25 26 ASA QT cell wall hydrolase (major autolysin)<br />

BLi01039 YheN 40 41 SSA AT unknown; similar to endo-1,4-β-xylanase<br />

BLi01079 YhaH 30 31 TSG KN unknown<br />

BLi01109 29 30 ASA AQ subtilisin carlsberg precursor (EC 3.4.21.62)<br />

243


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi01123 Epr 26 27 IQA ES minor extracellular serine protease (EC 3.4.21.-)<br />

BLi01138 23 24 VSA AE unknown<br />

BLi01150 27 28 GSG NT unknown<br />

BLi01154 YvgL 28 29 GGS AK unknown; similar to molybdate-binding protein<br />

BLi01295 AbnA 32 33 SAA EP arabinan-endo 1,5-α-L-arabinase<br />

BLi01299 32 33 VQA QE unknown<br />

BLi01308 YoeB 22 23 SLA AK unknown<br />

BLi01309 25 26 ASA KE putative cell wall-binding protein<br />

BLi01364 Ggt 25 26 SKA EG γ-glutamyltranspeptidase<br />

BLi01372 YesS 43 44 LFS AA unknown; similar to transcriptional regulator<br />

(AraC/XylS family)<br />

BLi01376 YesW 32 33 AEA DG unknown<br />

BLi01404 Pel 24 25 IEA AD pectate lyase (EC 4.2.2.2)<br />

BLi01455 27 28 AAA VW unknown<br />

BLi01536 27 28 DVK KE hypothetical protein<br />

BLi01539 25 26 IYA AK unknown<br />

BLi01566 26 27 VDA TT putative phosphodiesterase<br />

BLi01585 28 29 NNR EK unknown<br />

BLi01590 YkvT 24 25 EHA QA unknown; similar to spore cortex-lytic enzyme<br />

BLi01592 YkvV 27 28 ASA KQ unknown; similar to thioredoxin<br />

BLi01595 15 16 AEA KV hypothetical protein<br />

BLi01607 YkwD 26 27 ADA KE unknown<br />

BLi01622 27 28 AKA GE unknown<br />

BLi01644 MoaD 16 17 AGA QS molybdopterin converting factor (subunit 1)<br />

BLi01697 YlaJ 26 27 ARN EA unknown<br />

BLi01722 YlbL 35 36 GEA TE unknown<br />

BLi01733 PbpB 41 42 VNG EV penicillin-binding protein 2B (cell-division<br />

septum)<br />

BLi01742 YlxW 28 29 ARE NK unknown; similar to proteins<br />

BLi01743 YlxX 29 30 SLK AP unknown<br />

BLi01747 27 28 VQA DT putative bacillopeptidase F<br />

BLi01748 Bpr 30 31 SDA AA bacillopeptidase F (EC 3.4.21.-)<br />

BLi01851 FliL 31 32 GSA SE flagellar protein required for flagellar formation<br />

BLi01880 33 34 TRA AS putative endo-1,4-glucanase<br />

BLi01882 33 34 ASG TS putative cellulase (EC 3.2.1.4)<br />

BLi01883 31 32 ALA AS putative endo-1,4-β-mannosidase<br />

BLi02014 YoaW 25 26 AEA AV unknown<br />

BLi02027 37 38 IEK IP unknown<br />

BLi02030 NucB 32 33 AEG AA sporulation-specific extracellular nuclease<br />

BLi02033 YneA 28 29 AGK IE unknown<br />

BLi02048 YneN 21 22 VWN FT unknown; similar to thioldisulfide interchange<br />

protein<br />

BLi02088 BglC 49 50 AAA AS endo-1,4-β-glucanase<br />

BLi02095 YwoF 27 28 TGA KE unknown<br />

BLi02100 26 27 VFG GN unknown<br />

BLi02101 45 46 ANA AS unknown<br />

BLi02120 DctB 28 29 VIG VD possible C4-dicarboxylate binding protein<br />

BLi02165 33 34 EYA YM unknown<br />

BLi02166 24 25 LNA GD unknown<br />

BLi02178 YndF 27 28 SHE IE unknown; similar to spore germination protein<br />

BLi02210 23 24 SYA AA unknown<br />

BLi02213 YocH 25 26 ASA KE unknown; similar to cell wall-binding protein<br />

BLi02237 YoqH 23 24 AYA QV unknown<br />

BLi02255 YvgO 24 25 SEA KE unknown<br />

BLi02264 YojL 26 27 VEA QT unknown; similar to cell wall-binding protein<br />

244


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi02271 YoaJ 25 26 ASA AY unknown; similar to extracellular endoglucanase<br />

precursor<br />

BLi02281 CtpA 36 37 VYS AS carboxy-terminal processing protease<br />

BLi02310 YpmS 32 33 GGQ KE unknown<br />

BLi02312 YpmQ 24 25 TSK ID unknown<br />

BLi02321 YpjP 29 30 LMA DK unknown<br />

BLi02340 YpcP 22 23 ATA VH unknown; similar to DNA polymerase I<br />

BLi02367 PonA 62 63 VMV AD penicillin-binding proteins 1A/1B<br />

BLi02372 AspB 22 23 AKA KE aspartate aminotransferase (EC 2.6.1.1)<br />

BLi02373 YpmB 24 25 AGA NV unknown<br />

BLi02387 YpjB 22 23 LKA KE unknown; similar to proteins<br />

BLi02391 QcrA 39 40 RFA LD menaquinolcytochrome c oxidoreductase (ironsulfur<br />

subunit)<br />

BLi02420 GpsA 21 22 VLA DN NAD(P)H-dependent glycerol-3-phosphate<br />

dehydrogenase<br />

BLi02431 SleB 33 34 AFS EQ spore cortex-lytic enzyme<br />

BLi02447 30 31 SEA SE unknown<br />

BLi02450 22 23 SYG IY close homolog to LytR attenuator role for<br />

lytABC and lytR expression<br />

BLi02451 50 51 NSA AS putative peptidoglycan GlcNAc deacetylase<br />

BLi02461 ResA 35 36 ESV AV essential protein similar to cytochrome c<br />

biogenesis protein<br />

BLi02465 DacB 27 28 AQA QP penicillin-binding protein 5* (D-alanyl-Dalaninecarboxypeptidase)<br />

(EC 3.4.16.4)<br />

BLi02476 YpuD 43 44 VSS EE unknown<br />

BLi02479 32 33 VKV AE hypothetical protein<br />

BLi02498 DacF 27 28 ESA KK penicillin-binding protein (putative D-alanyl-Dalanine<br />

carboxypeptidase)<br />

BLi02506 27 28 AEA LN putative PTS cellobiose-specific enzyme IIB<br />

BLi02525 Lip 30 31 ASA AS extracellular lipase (EC 3.1.1.3)<br />

BLi02527 31 32 AKG EE putative ABC transporter<br />

BLi02543 23 24 AAA AG unknown<br />

BLi02544 28 29 VSA DT unknown<br />

BLi02564 YdhM 25 26 EYA HS unknown; similar to<br />

cellobiosephosphotransferase system enzyme II<br />

BLi02565 PhoB 34 35 AKK KE alkaline phosphatase III (EC 3.1.3.1)<br />

BLi02590 YqiI 24 25 AFA AE unknown; similar to N-acetylmuramoyl-Lalanine<br />

amidase<br />

BLi02607 SpoIII<br />

AH<br />

31 32 EGE NV mutants block sporulation after engulfment<br />

BLi02613 SpoIII<br />

AB<br />

22 23 EMA KP mutants block sporulation after engulfment<br />

BLi02637 TasA 27 28 TWA AF translocation-dependent antimicrobial spore<br />

component<br />

BLi02639 YqxM 39 40 LSQ HT unknown<br />

BLi02640 YqzG 23 24 AHA AA unknown<br />

BLi02671 YqzC 32 33 GKA EA unknown; similar to proteins from B. subtilis<br />

BLi02682 YqfZ 39 40 AAE AP unknown<br />

BLi02744 YqxA 28 29 ANN GM unknown<br />

BLi02820 31 32 SDA SP putative phosphatase<br />

BLi02825 YndA 27 28 AAG AG unknown<br />

BLi02826 YvaG 25 26 AIA AS unknown; similar to 3-oxoacyl- acylcarrierprotein<br />

reductase<br />

BLi02827 SacC 22 23 FSA AA levanase (EC 3.2.1.65)<br />

BLi02833 24 25 ALT FE unknown<br />

BLi02844 32 33 AAS EK hypothetical protein<br />

245


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi02850 32 33 AAE DS putative cell wall-associated protease<br />

precursor(EC 3.4.21.-)<br />

BLi02859 YrrR 26 27 RLA EI unknown; similar to penicillin-binding protein<br />

BLi02865 YrrL 43 44 VKS AL unknown; similar to folate metabolism<br />

BLi02884 YrvJ1 28 29 ASA AI unknown; similar to N-acetylmuramoyl-Lalanine<br />

amidase<br />

BLi02902 BofC 29 30 ALA EK forespore regulator of the sigma-K checkpoint<br />

BLi02914 NadB 36 37 ASV KD L-aspartate oxidase<br />

BLi02979 27 28 GKA EF hypothetical protein<br />

BLi03010 32 33 TEA SE unknown<br />

BLi03024 AraN 31 32 DQA DG L-arabinose transport (sugar-binding protein)<br />

BLi03029 27 28 ESG KA close homolog to AbnA arabinan-endo 1,5-α-Larabinase<br />

BLi03053 PelB 30 31 ASA AN pectate lyase (EC 4.2.2.10)<br />

BLi03092 SppA 27 28 LLA VF putative signal peptide peptidase required for<br />

efficient processing of pre-proteins (EC 3.4.21.-)<br />

BLi03095 25 26 VCG YY unknown<br />

BLi03138 YtzB 17 18 AAA VV unknown<br />

BLi03164 YteS 21 22 SCG KD unknown<br />

BLi03168 YtcQ 30 31 DQA SS unknown; similar to lipoprotein<br />

BLi03201 YtlA 22 23 SCG GQ unknown<br />

BLi03262 27 28 EAA SQ unknown<br />

BLi03304 29 30 AFA AS putative sugar hydrolase<br />

BLi03331 PbpD 28 29 REA QN penicillin-binding protein 4<br />

BLi03343 29 30 TRE QT unknown<br />

BLi03371 20 21 VSK AG putative lipase precursor<br />

BLi03389 YuiC 27 28 VEA QD unknown; similar to proteins<br />

BLi03405 29 30 GDA EF hypothetical protein<br />

BLi03421 YutC 23 24 ALN DT unknown; similar to proteins<br />

BLi03423 YunA 21 22 ALA KE unknown<br />

BLi03433 24 25 SQA AD unknown<br />

BLi03441 YurI 25 26 AEA FQ unknown; similar to ribonuclease<br />

BLi03490 GerA<br />

C<br />

23 24 DSR QI germination response to L-alanine<br />

BLi03538 BdbD 35 36 TQN AS thiol-disulfide oxidoreductase<br />

BLi03540 18 19 AFS AG putative ABC transporter sugar binding protein<br />

BLi03544 34 35 SQA DE putative sugar hydrolase<br />

BLi03547 MntA 27 28 SSS EE manganese ABC transporter (membrane protein)<br />

BLi03670 25 26 SFA KD unknown<br />

BLi03706 SacB 29 30 TFA KE levansucrase (EC 2.4.1.10)<br />

BLi03707 YveB 32 33 EKK GE unknown; similar to levanase<br />

BLi03719 30 31 AEG PQ putative ribonuclease (EC 3.1.27.-)<br />

BLi03739 38 39 LHS SH unknown<br />

BLi03741 YvpA 29 30 ALA AE unknown; similar to pectate lyase<br />

BLi03749 YvnB 28 29 SSA SG unknown<br />

BLi03765 YvjB 36 37 ASA AE unknown; similar to carboxy-terminal<br />

processing protease<br />

BLi03767 25 26 LKS NH putative cell wall-binding protein<br />

BLi03796 YvhJ 48 49 ISA AD unknown; similar to transcriptional regulator<br />

BLi03808 LytC 24 25 VFA AN N-acetylmuramoyl-L-alanine amidase<br />

(EC3.5.1.28)<br />

BLi03809 LytB 25 26 AQA AD modifier protein of major autolysin LytC<br />

BLi03811 LytR 31 32 SYA YY attenuator role for lytABC and lytR expression<br />

BLi03821 LytD 28 29 ALA AY N-acetylglucosaminidase (major autolysin)<br />

(EC3.2.1.96)<br />

BLi03833 YwtF 44 45 ANA SK unknown; similar to transcriptional regulator<br />

246


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi03835 YwtD 33 34 VRA DT unknown; similar to murein hydrolase<br />

BLi03836 YwtC 25 26 FKY SD unknown<br />

BLi03837 YwtB 46 47 GSA KT unknown; similar to capsular polyglutamate<br />

biosynthesis<br />

BLi03892 SpoII<br />

Q<br />

48 49 ASN ND required for completion of engulfment<br />

BLi03919 Ywm<br />

D<br />

17 18 AFA AE unknown<br />

BLi03920 Ywm<br />

C<br />

23 24 AFA AE unknown<br />

BLi03923 Ywm<br />

B<br />

29 30 EAA GN unknown<br />

BLi03942 SpoII<br />

R<br />

31 32 ETA QS required for processing of pro-σ-E<br />

BLi03981 24 25 AGA AK unknown<br />

BLi04019 Vpr 28 29 VQA TS minor extracellular serine protease (EC 3.4.21.-)<br />

BLi04029 25 26 ALL KE unknown<br />

BLi04074 YwaD 30 31 AQA AP unknown; similar to aminopeptidase<br />

BLi04089 LicB 23 24 EKS AE PTS lichenan-specific enzyme IIB component<br />

BLi04102 YweA 30 31 EEA SA unknown; similar to proteins from B. subtilis<br />

BLi04124 24 25 AQA KE unknown<br />

BLi04129 27 28 AEA AS putative pectate lyase (EC 2.1.3.3)<br />

BLi04148 YxeA 24 25 IHN EV unknown<br />

BLi04156 29 30 AGA QE unknown; similar to glycerophosphoryl diester<br />

phosphodiesterase<br />

BLi04157 YhjA 27 28 AEA KT unknown<br />

BLi04166 22 23 ASA EA carbamate kinase (EC 2.7.2.2)<br />

BLi04182 28 29 ANS QD putative sugar ABC transporter sugar<br />

bindingprotein<br />

BLi04185 33 34 AHA AN unknown<br />

BLi04206 25 26 RPA KT hypothetical protein<br />

BLi04220 YxiA 28 29 ASA QT unknown; similar to arabinan endo-1,5-α-Larabinosidase<br />

BLi04232 YdaJ 28 29 IKA ED unknown<br />

BLi04236 YdaN 23 24 AAA KD putative cellulose synthase<br />

BLi04254 GlpQ 28 29 AEA AS glycerophosphoryl diester phosphodiesterase<br />

(EC 3.1.4.46)<br />

BLi04272 24 25 LQK AT unknown<br />

BLi04276 YvfO 26 27 AEA AR unknown; similar to arabinogalactan endo-1,4-βgalactosidase<br />

BLi04294 27 28 AYA QS hypothetical<br />

BLi04306 33 34 VSR DT unknown<br />

BLi04308 41 42 SFA WV unknown<br />

BLi04333 YycH 27 28 IWG FQ unknown<br />

Lipoproteins<br />

BLi00173 GerD 19 20 VTA CA germination response to L-alanine and to the<br />

combination of glucose, fructose, L-asparagine,<br />

and KCl<br />

BLi00260 24 25 VAG CS putative sugar ABC transporter, periplasmic-<br />

binding protein<br />

BLi00261 19 20 LGA CS unknown<br />

BLi00280 PenP 26 27 LAG CG β-lactamase precursor (EC 3.5.2.6)<br />

(penicillinase)<br />

BLi00301 19 20 LAG CG putative serine protease<br />

BLi00384 YckB 25 26 TAA CS unknown; similar to amino acid ABC transporter<br />

(binding protein)<br />

247


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi00397 21 22 LSG CG putative spermidine/putrescine-binding<br />

periplasmic protein 2 precursor<br />

BLi00420 YckK 19 20 MAA CG unknown; similar to glutamine ABC transporter<br />

(glutamine-binding protein)<br />

BLi00442 19 20 VTA CS putative sugar ABC transporter, periplasmicbinding<br />

protein<br />

BLi00466 YclQ 19 20 VAA CG unknown; similar to ferrichrome ABC<br />

transporter (binding protein)<br />

BLi00550 YdcC 20 21 LSA CG unknown<br />

BLi00659 YvdG 22 23 LAA CS unknown; similar to maltose/maltodextrinbinding<br />

protein<br />

BLi00708 YybP 18 19 AGG CG unknown<br />

BLi00717 YerH 18 19 LSA CA unknown<br />

BLi00894 22 23 LMG CS putative oligopeptide ABC transporter (binding<br />

protein) (initiation of sporulation, competence<br />

development)<br />

BLi00942 SsuA 17 18 LAG CS aliphatic sulfonate ABC transporter (binding<br />

lipoprotein)<br />

BLi00974 YhcJ 19 20 IAG CA unknown; similar to ABC transporter (binding<br />

lipoprotein)<br />

BLi00978 YhcN 21 22 TAG CG unknown<br />

BLi01011 16 17 LGA CT putative oxidoreductase<br />

BLi01072 PrsA 19 20 LSA CS protein secretion (post-translocation molecular<br />

chaperone)<br />

BLi01111 YhfQ 19 20 MTA CS unknown; similar to iron(III) dicitrate-binding<br />

protein<br />

BLi01140 MsmE 20 21 LAG CS multiple sugar-binding protein<br />

BLi01199 Ipi 15 16 VSG CG intracellular proteinase inhibitor<br />

BLi01218 Med 17 18 LSG CG positive regulator of comK<br />

BLi01226 AppA 23 24 LTA CN oligopeptide ABC transporter (oligopeptidebinding<br />

protein)<br />

BLi01232 OppA 20 21 LSA CG oligopeptide ABC transporter (binding protein)<br />

(initiation of sporulation, competence<br />

development)<br />

BLi01241 22 23 LGG CG unknown<br />

BLi01368 YesO 21 22 LFG CS unknown; similar to sugar-binding protein<br />

BLi01384 LplA 23 24 LIG CS lipoprotein<br />

BLi01396 DppE 19 20 LFG CT dipeptide ABC transporter (dipeptide-binding<br />

protein) (sporulation)<br />

BLi01417 21 22 LAG CG unknown<br />

BLi01431 19 20 LAA CS unknown<br />

BLi01505 PbpC 20 21 AGA CS penicillin-binding protein 3<br />

BLi01570 18 19 LTA CN unknown<br />

BLi01673 YkyA 19 20 LTG CL unknown<br />

BLi01680 Slp 19 20 TSG CS small peptidoglycan-associated lipoprotein<br />

BLi01936 18 19 AAA CL unknown<br />

BLi02013 20 21 LSG CN unknown<br />

BLi02263 YojM 19 20 AAA CT unknown; similar to superoxide dismutase<br />

BLi02284 YodJ 23 24 GTG CT unknown; similar to D-alanyl-D-alanine<br />

carboxypeptidase<br />

BLi02311 YpmR 18 19 LSA CT unknown<br />

BLi02417 YphF 19 20 LSG CL unknown<br />

BLi02557 RocC 22 23 AAA CL amino acid permease<br />

BLi02558 22 23 LAA CG unknown<br />

BLi02577 YqiX 19 20 LTA CG unknown; similar to amino acid ABC transporter<br />

(binding protein)<br />

248


Appendix V Supplemental Table Va<br />

ID number Name Cleavage Site Sequence Function<br />

BLi02591 YqiH 17 18 LAG CG unknown; similar to lipoprotein<br />

BLi02626 OpuAC 20 21 LAA CG glycine betaine ABC transporter (glycine<br />

betaine-binding protein)<br />

BLi02656 YqgU 21 22 AAG CT unknown<br />

BLi02658 FhuD 23 24 LTA CG ferrichrome ABC transporter (ferrichromebinding<br />

protein)<br />

BLi02676 PstS 22 23 AAA CG phosphate ABC transporter (binding protein)<br />

BLi02763 YqeF 17 18 LSG CG unknown<br />

BLi02811 26 27 LAG CT putative oligopeptide transporter putative<br />

substrate binding domain<br />

BLi02821 19 20 LFA CT putative lipase/esterase<br />

BLi02910 CoxA 20 21 LSA CG spore cortex protein<br />

BLi02986 GerM 22 23 LSG CG germination (cortex hydrolysis) and sporulation<br />

(stage II, multiple polar septa)<br />

BLi03178 22 23 LAA CG unknown<br />

BLi03208 YtkA 18 19 LSA CS unknown<br />

BLi03213 YcdH 20 21 TAG CS unknown; similar to ABC transporter (binding<br />

protein)<br />

BLi03455 YusA 19 20 LAA CG unknown<br />

BLi03475 YfiY 20 21 LAA CG unknown; similar to iron(III) dicitrate transport<br />

permease<br />

BLi03476 YusW 18 19 MTG CG unknown<br />

BLi03506 YvrC 20 21 LSG CG unknown; similar to iron-binding protein<br />

BLi03623 19 20 AAG CE unknown<br />

BLi03649 OpuCC 22 23 ISG CA glycine betaine/carnitine/choline ABC<br />

transporter(osmoprotectant-binding protein)<br />

BLi03657 19 20 LAA CG putative iron(III) transporter binding protein<br />

BLi03770 CccB 18 19 LAA CG cytochrome c551<br />

BLi03810 17 18 LSA CG unknown<br />

BLi03845 RbsB 18 19 LSA CS ribose ABC transporter (ribose-binding protein)<br />

BLi03906 FeuA 19 20 AAG CG iron-uptake system (binding protein)<br />

BLi04191 22 23 TTG CG unknown<br />

BLi04262 YxeB 20 21 VSA CG unknown; similar to ABC transporter (binding<br />

protein)<br />

BLi04280 YvfK 22 23 LTA CG unknown; similar to maltose/maltodextrinbinding<br />

protein<br />

Pseudopilins<br />

BLi02642 ComGG 6 7 KG FIYPA pseudopilin<br />

BLi02644 ComGE 7 8 KG FTTVE pseudopilin<br />

BLi02645 ComGD 5 6 KG FTLLE pseudopilin<br />

BLi02646 ComGC 5 6 KG FTLIE pseudopilin<br />

Signal peptide prediction was done with the SignalP 2.0 software for proteins that contain a signal sequence for<br />

one of the type I signal peptidases (http://www.cbs.dtu.dk/services/SignalP-2.0/). Proteins with more than one<br />

membrane spanning domain were excluded from the list. <strong>The</strong> cleavage site is the position of the amino acids in the<br />

protein between which the signal peptidase is predicted to cleave. Note that signal peptides with a proline at the<br />

+1 position are probably not cleaved by signal peptidase (Tjalsma et al., 2000). Lipoproteins were predicted by<br />

using the search pattern described by Sutcliffe and Harrington (2002) with the PAT<strong>TI</strong>NPROT software<br />

(http://npsa-pbil.ibcp.fr/cgi-bin/npsa_automat.pl?page=/NPSA /npsa_server.html). <strong>The</strong> pseudopilins were also<br />

predicted by using the PAT<strong>TI</strong>NPROT software and by BLAST searches with ComG proteins of B. subtilis.<br />

249


Appendix V Supplemental Table Vb<br />

Supplemental table Vb: Extracellular proteins of B. licheniformis which either contain an N-terminal signal peptide or which are known to be secreted<br />

by other pathways<br />

Protein Function pI Mr ID number md signal peptide sequence P G N LB ex LB st<br />

Metabolism of carbohydrates<br />

Pel pectate lyase (EC 4.2.2.2) 5.85 48.7 BLi01404 MKRFFSVIILGALLLLGTSAPIEA AD x<br />

SacB levansucrase (EC 2.4.1.10) 8.73 53.7 BLi03706 MNIKNIAKKASALTVAAALLAGGAPQTFA KE x<br />

SacC levanase (EC 3.2.1.65) 7.18 75.5 BLi02827 MKKRMIQMGIIGAMMFPEAFSA AA x x<br />

YheN similar to endo-1,4-β-xylanase 9.94 34.6 BLi01039 MRQVSKKTAPSVAYLLTKAACFFVLCLILLYV<br />

WDLSQSSA AT<br />

x<br />

YvfO similar to arabinogalactan endo-1,4β-galactosidase<br />

5.87 46.2 BLi04276 MKNVLAVFVVLIFVLGAFGTSGPAEA AR x x<br />

YvpA similar to pectate lyase 9.00 23.7 BLi03741 MMKRLAGTVILSGLLVCGFGQALPEKALA AE x<br />

YxiA similar to arabinan endo-1,5-α-Larabinosidase<br />

6.63 53.1 BLi04220 MNMRKCFIQVLALLFIIAACFAPNQASA QT x<br />

BLi00338 putative chitinase (EC 3.2.1.14) 4.97 65.8 BLi00338 MLINKSKKFFVFSFIFVMMLSLSFVNGEVAKA<br />

DS<br />

x<br />

BLi03029 a close homolog to AbnA arabinanendo<br />

1,5-α-L-arabinase<br />

8.62 35.6 BLi03029 MKNVLRKMSLAALIFGLLLSFSMPESGKA AF x a)<br />

BLi04129 putative pectate lyase (EC 2.1.3.3) 9.32 37.4 BLi04129 MKKLISIIFIFVLGVVGSLTAAVSAEA AS x x<br />

Metabolism of proteins and peptides<br />

Bpr bacillopeptidase F (EC 3.4.21.-) 5.04 155.0 BLi01748 MKRKLRKKAFS<strong>TI</strong>LSGLLIGSLFMPAVSDAA<br />

AK<br />

x x x x<br />

Epr minor extracellular serine protease<br />

(EC 3.4.21.-)<br />

10.2 63.0 BLi01123 MKKLWKIAVSAAMFVGFFANSPRIQA ES x x x<br />

Ggt γ-glutamyltranspeptidase (EC<br />

2.3.2.2)<br />

4.81 64.0 BLi01364 1 MRRLAFLVVAFCLAVGCFFSPVSKA EG x x x x<br />

Mpr glutamyl endopeptidase precursor<br />

(EC 3.4.21.19)<br />

9.78 33.7 BLi00340 MVSKKSVKRGLITGLIGISIYSLGMHPAQA AP x x x<br />

Vpr extracellular serine protease (EC<br />

3.4.21.-)<br />

8.87 85.6 BLi04019 MRKSIVRYFVMAFILLFALSTFLTGVQA TS x x x x<br />

YwaD similar to aminopeptidase 9.09 48.2 BLi04074 MKRKMMMFGLALSIIAGGVVADGTGNAAQA<br />

AP<br />

x x x<br />

BLi00301 putative serine protease 6.65 45.0 BLi00301 MKSKWSAMVVIAGLLLLAG CGA x<br />

250


Appendix V Supplemental Table Vb<br />

Protein Function pI Mr ID number md signal peptide sequence P G N LB ex LB st<br />

BLi01109 subtilisin carlsberg precursor (EC<br />

3.4.21.62)<br />

8.63 38.9 BLi01109 MMRKKSFWLGMLTALMLVFTMAFS DS x x x<br />

BLi01747 putative bacillopeptidase F 9.83 54.2 BLi01747 MKKKPLFRTFMCAALIGSLLAPVAVQA DT x x x<br />

Metabolism of nucleotides and nucleic acids<br />

NucB sporulation-specific extracellular<br />

nuclease (EC 3.-.-.-)<br />

9.05 15.3 BLi02030 MIKKWAVHLLFSALVLLGLSGGAA YS x<br />

YfkN similar to 2',3'-cyclic-nucleotide 2'- 5.11 161.7 BLi00814 2 MVGIQKRRFSRKNILRILLTSVMILSLLMPNTQT x x x<br />

phosphodiesterase (EC 3.1.4.16)<br />

YA EE<br />

YhcR similar to 5'-nucleotidase 4.68 130.6 BLi00982 MVNVVKSRFMAGL<strong>TI</strong>TFMMIASFLTPFADVTH<br />

A SE<br />

x x x x<br />

YurI similar to ribonuclease 5.14 30.5 BLi03441 MNRKCVIPFILMLSAMCAPAQNAEA FQ x x<br />

BLi03719 putative ribonuclease (EC 3.1.27.-) 9.58 16.6 BLi03719 MKKILSTLALGFVLALGFLAGNLFTSSA EG x x x x<br />

Metabolism of lipids<br />

GlpQ glycerophosphoryl diester<br />

phosphodiesterase (EC 3.1.4.46)<br />

YhdW similar to glycerophosphodiester<br />

phosphodiesterase (EC 3.1.4.46)<br />

9.17 33.8 BLi04254 MKRLVRSIFLITAAIAAFGFGFSGHAEA AS x x<br />

9.43 31.5 BLi04156 MSALFKKLMLSSLIGVSIGSALFAPNAGA QE x x<br />

Metabolism of phosphate<br />

PhoB alkaline phosphatase III (EC 3.1.3.1) 9.51 50.3 BLi02565 MGFLRNRIVGITLAGAVALGSAGTGSA AM x<br />

PhoD phosphodiesterase/alkaline<br />

7.25 65.8 BLi00281 MKKLSEESLKDNTFDRRRFIQGAGKIAGLSLGL x<br />

phosphatase (EC 3.1.3.1)<br />

AIAQSMGAMEVNA AP<br />

Phy 3-phytase (EC 3.1.3.8) / 6-phytase<br />

(EC 3.1.3.26)<br />

4.66 42.0 BLi00448 MNFYKTLALSTLAASLLSPSWSILPRAEA SA x x<br />

Metabolism of the cell wall<br />

LytD N-acetylglucosaminidase (major<br />

autolysin) (EC 3.2.1.96)<br />

9.60 97.0 BLi03821 MKNIRKTVIFAAIILLVHTAVPA IP x<br />

PbpB penicillin-binding protein 2B (celldivision<br />

septum)<br />

9.70 78.8 BLi01733 MPKKNKFMNRGAAILSICFALFFFVIVGRFA FI x<br />

YodJ similar to D-alanyl-D-alanine<br />

carboxypeptidase (EC 3.4.16.4)<br />

6.03 30.9 BLi02284 MNGKYKYV<strong>TI</strong>ASLLSAAVLLGTG CTM x<br />

YvcE similar to cell wall-binding protein 9.66 49.1 BLi00347 MKKKVYTFGLASILGTASLFTPFMNNTASA ET x<br />

251


Appendix V Supplemental Table Vb<br />

Protein Function pI Mr ID number md signal peptide sequence P G N LB ex LB st<br />

YrvJ1 similar to N-acetylmuramoyl-Lalanine<br />

amidase<br />

? ? BLi02884 MKKRAVLILSMMMLAAQAAFYTSSNTASA AI x x x x x<br />

YwtD similar to murein hydrolase 9.61 45.6 BLi03835 1 MIKKAANKKLVLFCGIAVLWMSLFLTNHNDVR<br />

A DT<br />

x<br />

BLi01309 putative cell wall-binding protein 9.13 25.9 BLi01309 MKK<strong>TI</strong>MSLAAAAAMSATAFGATASA KE x<br />

BLi03767 putative cell wall-binding protein 9.52 46.3 BLi03767 MKRKLMTLGLTAVLGSSAVLIPLKSNHALA YE x x<br />

Transport/binding proteins and lipoproteins<br />

AppA oligopeptide ABC transporter<br />

(oligopeptide-binding protein)<br />

4.77 62.7 BLi01226 MNKRKTGFSILSLLLILSIFLTA CNS x x x<br />

DppE dipeptide ABC transporter<br />

(dipeptide-binding protein)<br />

(sporulation)<br />

4.85 61.2 BLi01396 MKRLTSVLASAFAVILLFG CTA x<br />

FeuA iron-uptake system (binding protein) 6.46 34.9 BLi03906 MRKISIFLFILLLALGAAG CGN x x<br />

MntA manganese ABC transporter<br />

(membrane protein)<br />

5.21 34.3 BLi03547 MKWKQTLAIAAALILILAAGCSSKSSS EE x<br />

OppA oligopeptide ABC transporter<br />

(binding protein)<br />

5.52 60.8 BLi01232 MKKRLSFISLMLIFTLVLSA CGF x x x x<br />

OpuAC glycine betaine ABC transporter<br />

(glycine betaine-binding protein)<br />

6.00 32.0 BLi02626 MWKKIAGIGTAAVLTLGLAA CGS x<br />

PstS phosphate ABC transporter (binding<br />

protein)<br />

4.56 32.6 BLi02676 MKPFKKITLMFIMSVLVVFAAA CGS x<br />

YcdH similar to ABC transporter (binding<br />

protein)<br />

5.02 38.4 BLi03213 MKKTFGIASAFILAAGLTAG CSS x<br />

YclQ similar to ferrichrome ABC<br />

transporter (binding protein)<br />

5.41 34.7 BLi00466 MKKLSLLIMALITVLVVAA CGN x x x x<br />

YesO similar to sugar-binding protein 4.89 48.7 BLi01368 MLMRRFVFVSLCILLTLGLFG CSS x<br />

YfiY similar to iron(III) dicitrate transport<br />

permease<br />

5.56 36.7 BLi03475 MKRWSIVGFIALLAISILAA CGG x<br />

YflE similar to anion-binding protein 5.67 74.3 BLi00793 5 MKRFIKERGLAFFLIAAILLWLKTYA AY x x x x<br />

YfnI similar to anion-binding protein 5.95 72.6 BLi04159 5 MKKIFSHKLSFFVLAVVFVWAKTYASYFLEFN<br />

LGVKGSTQHMLLFINPLSF<strong>TI</strong>AALGLALFAKGR<br />

RSA IW<br />

x x x x<br />

252


Appendix V Supplemental Table Vb<br />

Protein Function pI Mr ID number md signal peptide sequence P G N LB ex LB st<br />

YhcJ similar to ABC transporter (binding<br />

lipoprotein)<br />

4.67 29.4 BLi00974 MKKFACVVIFLLLAAVIAG CAA x<br />

YqgS similar to putative molybdate binding<br />

protein<br />

5.65 73.2 BLi02659 5 MRKSFFSKISFLLIATLLMWLKTYVVYK TS x<br />

BLi02527 putative ABC transporter 6.49 57.0 BLi02527 MAYIAKRMIIPIIFLFILASCSAGGAGSAKG EE x x x<br />

BLi02811 putative oligopeptide transporter<br />

(putative substrate binding domain)<br />

5.83 61.1 BLi02811 MMSGRISLKIKIIFILMLAFSILLAG CTT x x<br />

BLi03657 putative iron(III) transporter binding<br />

protein<br />

Membrane bioenergetics<br />

QcrA menaquinolcytochrome c<br />

oxidoreductase (iron-sulfur subunit)<br />

5.81 34.9 BLi03657 MKRFKWFALFAALILLLAA CGN x x<br />

7.67 19.4 BLi02391 MKMSEKRHRVSRRQFLNYTLTGVGGFMAAG<br />

MLMPMVRFA LD<br />

Mobility and chemotaxis<br />

FlgE flagellar hook protein 4.63 28.0 BLi01849 x x<br />

FlgK flagellar hook-associated protein 1<br />

(HAP1)<br />

4.44 54.3 BLi03785 x x x<br />

FlhO flagellar basal-body rod protein 4.74 30.7 BLi03875 x<br />

FlhP flagellar hook-basal body protein 6.55 30.3 BLi03874 x<br />

FliD flagellar hook-associated protein 2<br />

(HAP2)<br />

5.19 54.2 BLi03778 x x<br />

FliK flagellar hook-length control 4.74 51.2 BLi01847<br />

FliL flagellar protein required for flagellar<br />

formation<br />

5.06 16.0 BLi01851 MNKKLLGIMM<strong>TI</strong>ILAIAVLGTAAFFVIKGSA SE x<br />

Hag flagellin protein 5.28 33.2 BLi03780 x x x x x<br />

Sporulation<br />

TasA translocation-dependent<br />

antimicrobial spore component<br />

5.27 28.8 BLi02637 MGTKKKLGLGVASAALGLALVGGGTWA AF x x x x x<br />

RNA synthesis and regulation<br />

YwtF similar to transcriptional regulator 9.64 35.8 BLi03833 1 MLRSQRTKKKRLRKWVKYSLFFIALILTATA<br />

AA<br />

253<br />

x<br />

x x


Appendix V Supplemental Table Vb<br />

Protein Function pI Mr ID number md signal peptide sequence P G N LB ex LB st<br />

Phage-related functions<br />

XkdG PBSX prophage 5.14 34.3 BLi01331 x x x<br />

XkdK PBSX prophage 4.54 48.5 BLi01337 x x x x<br />

XkdM PBSX prophage 4.70 16.4 BLi01338 x x x x<br />

Unknown<br />

YbdN 5.35 30.7 BLi00302 MKKSLFLFVFSLFLMAIPAFS AS x x x x<br />

YdaJ 5.04 41.3 BLi04232 MKAPVRYIWIGMILCFLSVSLAVGCIKA ED x<br />

YkwD 10.16 26.6 BLi01607 MKKAFLLSAAAAATLFTFSGVQHA DA x<br />

YpjP 5.38 23.4 BLi02321 MKMWMRKALVALF<strong>TI</strong>ATFGLVSPPAALMA DK x<br />

YqgU 5.24 41.4 BLi02656 MAGLRVSLLIIAALMAVAAAG CTP x x<br />

YusA 6.45 30.5 BLi03455 MKKGLLTALFIIFAGVLAA CGS x x x x<br />

YusW 4.46 16.6 BLi03476 MNQFRMAVIALVLILMTG CGS x x<br />

YwoF 4.57 51.5 BLi02095 MRKWYFILSACILVSVIIAFA YD x x x x<br />

BLi00654 6.10 25.9 BLi00654 MKKAMAILGFLVLTASLLFIINKGTSQINA SQ x x x<br />

Unknown<br />

BLi00784 9.06 60.6 BLi00784 MKIQKRVQALLATSAMFAGLMLSDAVYA AE x<br />

BLi01431 4.56 18.6 BLi01431 MKKFLVLFLSFGLALALAA CSS x x x<br />

BLi02210 4.63 34.9 BLi02210 MLMSAFVLVLAACQQADPKGSYA AA x x x<br />

BLi02558 5.12 34.9 BLi02558 MKKWKSLSWMVLLLTLVMGLAA CGS x x x x<br />

BLi03010 9.50 34.9 BLi03010 MSKIKWIITL<strong>TI</strong>CTALAFSLFIFFNKA NF x x<br />

BLi03178 5.89 22.9 BLi03178 MKKLLKWTSALGLALSLALLAA CGN x<br />

BLi03260 4.68 37.4 BLi03260 2 MKKRLMSLLVCILVLVPAAGAFA AP x<br />

BLi03670 4.06 17.6 BLi03670 MKKLLFMVILSVLTLVFGSSSVSFA KD x<br />

BLi04124 6.80 48.5 BLi04124 MKRIYIFLLCFAVLLPVGGKTAQA KE x x x<br />

BLi04294 9.46 37.5 BLi04294 MVFKKPKVFIAAVILALSSFAGTAAYA QS x x x x x<br />

BLi04308 10.4 14.1 BLi04308 MKNHLYEKKKRKPLTR<strong>TI</strong>KATLAVLTMSIALV<br />

GGATVPSFA WV<br />

x x<br />

Signal peptides were predicted as described in the section on signal peptide prediction. <strong>The</strong> signal peptidase cleavage site is indicated with a gap in the amino acid sequence and the twinarginine<br />

motif in the signal peptides is highlighted in bold letters.<br />

md: number of additional predicted membrane spanning domains, P: phosphate starvation, G: glucose starvation, N: nitrogen starvation, LB ex : cells grown in LB, exponential growth phase,<br />

LB st : cells grown in LB, stationary growth phase.<br />

a this protein was found only in the extracellular proteome after 72 h cultivation (gel not shown).<br />

254


Appendix V Supplemental Table Vb<br />

Supplemental table Vc. Proteins detected in the extracellular proteome lacking known<br />

export signals<br />

Protein Function P G N LB<br />

Cell envelope related functions<br />

AtpD ATP synthase (subunit β) (EC 3.6.1.34) x<br />

MntB manganese ABC transporter (ATP-binding protein) x<br />

YkuA similar to penicillin-binding protein x<br />

Protein secretion<br />

SipS signal peptidase I (EC 3.4.21.89) x<br />

SipT signal peptidase I x<br />

Sporulation<br />

SpoVG required for spore cortex synthesis x<br />

Carbohydrate metabolism<br />

CitB aconitate hydratase (aconitase) (EC 4.2.1.3) x x x x<br />

Eno enolase (EC 4.2.1.11) x x x<br />

GapA glyceraldehyde-3-phosphate dehydrogenase (EC 1.2.1.12) x x<br />

Icd isocitrate dehydrogenase (EC 1.1.1.42) x x<br />

IolS myo-inositol catabolism x<br />

Mdh malate dehydrogenase (EC 1.1.1.37) x<br />

PdhA pyruvate dehydrogenase (E1 α subunit) (EC 1.2.4.1) x x<br />

PdhB pyruvate dehydrogenase (E1 β subunit) (EC 1.2.4.1) x<br />

PdhC pyruvate dehydrogenase (dihydrolipoamide acetyltransferase E2<br />

subunit) (EC 2.3.1.12)<br />

x<br />

PdhD pyruvate dehydrogenase / 2-oxoglutarate dehydrogenase<br />

(dihydrolipoamide dehydrogenase E3 subunit) (EC 1.8.1.4)<br />

x x x x<br />

Pgk phosphoglycerate kinase (EC 2.7.2.3) x x x x<br />

PtsH histidine-containing phosphocarrier protein of the PTS (HPr protein) x<br />

YjeA similar to chitooligosaccharide deacetylase x<br />

YwjH similar to transaldolase (pentose phosphate) (EC 2.2.1.2) x<br />

Amino acid metabolism<br />

GlnA glutamine synthetase (EC 6.3.1.2) x x x<br />

GlyA serine hydroxymethyltransferase (EC 2.1.2.1) x<br />

IlvC ketol-acid reductoisomerase (EC 1.1.1.86) x x x<br />

MetE cobalamin-independent methionine synthase (EC 2.1.1.14) x<br />

YjbG similar to oligoendopeptidase (EC 3.4.24.-) x<br />

BLi04164 ornithine carbamoyltransferase, catabolic (EC 2.1.3.3) x<br />

BLi04275 close homolog to Ald L-alanine dehydrogenase x<br />

Lipid metabolism<br />

Bcd leucine dehydrogenase S81735 (EC 1.4.1.9) x<br />

Nucleotide and nucleic acid metabolism<br />

GuaB inosine-monophosphate dehydrogenase (EC 1.1.1.205) x<br />

Ndk nucleoside diphosphate kinase (EC 2.7.4.6) x<br />

DNA packaging and segregation<br />

Hbs non-specific DNA-binding protein Hbsu x x<br />

RNA synthesis and modification<br />

TenA transcriptional regulator of extracellular enzyme genes x<br />

YugI similar to polyribonucleotide nucleotidyltransferase x<br />

Translation<br />

FusA elongation factor G x<br />

Frr ribosome recycling factor x<br />

RplJ ribosomal protein L10 (BL5) x<br />

RplL ribosomal protein L12 (BL9) x<br />

RpsF ribosomal protein S6 (BS9) x<br />

255


Appendix V Supplemental Table Vc<br />

Protein Function P G N LB<br />

Protein modification and folding<br />

GroES class I heat-shock protein (chaperonin) x<br />

PpiB peptidyl-prolyl isomerase (EC 5.2.1.8) x x<br />

Tig trigger factor (prolyl isomerase)(EC 5.2.1.8) x<br />

Detoxification and adaptation to atypical conditions<br />

AhpC alkyl hydroperoxide reductase (small subunit) (EC 1.6.4.-) x<br />

ClpP ATP-dependent Clp protease proteolytic subunit (class III heat-shock<br />

protein) (EC 3.4.21.92)<br />

x<br />

SodA superoxide dismutase (EC 1.15.1.1) x x x x<br />

PspA phage shock protein A homolog x<br />

YceD similar to tellurium resistance protein x<br />

YvtA similar to HtrA-like serine protease x<br />

Unknown<br />

YdhD Unknown x<br />

YkrZ Unknown x x x<br />

YkuV Unknown x<br />

YqbG Unknown x<br />

YxjG Unknown x<br />

BLi00303 Unknown x x x x<br />

BLi03379 Unknown x<br />

P: phosphate starvation, G: glucose starvation, N: nitrogen starvation, LB: Luria Broth.<br />

256

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!