187
Institute for Microbiology Department of Infectious Diseases University of Veterinary Medicine Hannover Virulence mechanisms of Streptococcus suis: Molecular characterisation of the biological activities of suilysin THESIS Submitted in partial fulfilment of the requirements for the degree DOCTOR OF PHILOSOPHY (PhD) awarded by the University of Veterinary Medicine Hannover by Maren Seitz (Hannover) Hannover 2011

Virulence mechanisms of Streptococcus suis

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Institute for Microbiology

Department of Infectious Diseases

University of Veterinary Medicine Hannover

Virulence mechanisms of Streptococcus suis:

Molecular characterisation of the

biological activities of suilysin

THESIS

Submitted in partial fulfilment of the requirements for the degree

DOCTOR OF PHILOSOPHY (PhD)

awarded by the University of Veterinary Medicine Hannover

by

Maren Seitz

(Hannover)

Hannover 2011

Supervisor: Prof. Dr. Peter Valentin-Weigand

Supervision Group: Prof. Dr. Silke Rautenschlein PD Dr. Manfred Rohde

1st Evaluation: Prof. Dr. Peter Valentin-Weigand

Institute for Microbiology

Department of Infectious Diseases

University of Veterinary Medicine Hannover

Prof. Dr. Silke Rautenschlein Clinic for Poultry

University of Veterinary Medicine Hannover

PD Dr. Manfred Rohde Department of Medical Microbiology

Helmholtz Centre for Infectious Research, Braunschweig

2nd Evaluation: Prof. Dr. Barbara Spellerberg

Institute of Medical Microbiology and Hygiene

University of Ulm

Date of final exam: 9th November 2011

This work was financially supported by the Deutsche Forschungsgemeinschaft

(DFG), Bonn, Germany (SFB 587).

Meiner Familie

Table of contents Chapter 1 General introduction ..........................................................................13

1. Streptococcus (S.) suis ................................................................15 1.1. S. suis infections .......................................................................................15

1.2. Pathogenesis and virulence mechanisms ................................................17

2. Cholesterol-dependent pore-forming cytolysins (CDC) ............19 2.1. The structure of CDC................................................................................19

2.2. The mechanism of pore-formation by CDC ..............................................21

2.3. The functional role of cholesterol and membrane recognition..................22

2.4. Suilysin......................................................................................................23

2.4.1. Prevalence, diversity, and regulation of the sly gene ..................23

2.4.2. Role of suilysin in host-pathogen interaction ...............................25

2.4.3. Role of suilysin in virulence and pathogenesis............................27

3. Outline of the present study.........................................................28

Chapter 2 Material and methods.........................................................................39 1. Bacterial strains and growth conditions.....................................41 2. Molecular biology and protein biochemical methods................42

2.1. Construction of mutated suilysin W461F and SVD...................................42

2.2. Expression of recombinant proteins .........................................................43

2.3. Heterologous expression of SLY, W461F and SVD in L. lactis ................43

2.4. Immunoblot analysis .................................................................................44

3. Cell culture methods.....................................................................44 3.1. Epithelial cells ...........................................................................................44

3.2. Antibiotic protection assay ........................................................................45

3.3. Labelling of bacteria for flow cytometry experiments................................46

3.4. Bacteria-cell association ...........................................................................46

3.5. Immunofluorescence microscopy .............................................................47

3.6. Double immunofluorescence microscopy (DIF)........................................48

3.7. Colocalisation experiments.......................................................................48

3.8. Field emission scanning electron microscopy (FESEM) ..........................49

3.9. Cell-permeability and macropore-formation assay ...................................49

3.10. Detection of α5β1 integrin expression on HEp-2 cells ...............................50

3.11. Haemolysis assay .....................................................................................50

3.12. Cytotoxicity assay .....................................................................................50

3.13. Pull down experiments..............................................................................51

3.14. GLISA .......................................................................................................52

4. Mouse infection model .................................................................52 4.1. Preparation of infection culture .................................................................52

4.2. Intranasal infection of mice .......................................................................52

4.3. Intravenous infection of mice ....................................................................53

4.4. Clinical score.............................................................................................53

4.5. Histopathological screening......................................................................54

4.6. Reisolation of S. suis from tissue and tracheo-nasal lavage (TNL)..........54

5. Statistical analyses .......................................................................55 Chapter 3 Results, part I: ....................................................................................57

Subcytolytic suilysin promotes invasion of Streptococcus suis in HEp-2

epithelial cells by Rac1-dependent activation of the actin cytoskeleton

Chapter 4 Results, part II:....................................................................................91

Identification of a RGD-motif in suilysin possibly involved in host cell

binding, Rac1-activation and macropore-formation

Chapter 5 Results, part III: ................................................................................115

Establishment of an intranasal CD1 mouse infection model for

colonization and invasion of Streptococcus suis serotype 2

Chapter 6 General discussion ..........................................................................153 Chapter 7 Summary ...........................................................................................163 Chapter 8 Zusammenfassung...........................................................................167 Chapter 9 Literature...........................................................................................171

List of abbreviations Δ Delta % Percent A. bidest. Aqua bidestillata A. dest. Aqua destillata ANOVA ANalysis Of VAriance ATCC American Type Culture Collection, Manassas, USA BALB/c Bagg ALBino; color locus c/c BBB Blood Brain Barrier BMEC Brain Microvascular Endothelial Cells bp Base pair(s) BSA Bovine Serum Albumin °C Degree Celsius CcpA Catabolite control protein A CDC Cholesterol-Dependent pore-forming Cytolysins Cdc42 Cell division control protein 42 homolog cf. Conferre CFSE CarboxyFluorescein Succinimidyl Ester CFU Colony Forming Units cm Centimetre CNS Central Nervous System CO2 Carbon dioxide CPS Capsule PolySaccharide Crl:CD1 (ICR) Charles River:CD1 (Institute for Cancer Research) CRM Cholesterol Recognition Motif C3 C3 toxin of Clostridium limosum Da Dalton DAPI 4',6-DiAmidino-2-PhenylIndole CD Cluster of Differentiation DMEM Dulbecco’s Modified Eagle’s Medium DNA DeoxyriboNucleic Acid DTT DiThioThreitol EDTA EthyleneDiamineTetraacetic Acid EF Extracellular Factor e.g. Exampli gratia ECM ExtraCellular Matrix et al. Et alii etc. Et cetera FBPS Fibronectin and Fibrinogen Binding Protein of S. suis FCS Fetal Calf Serum FESEM Field Emission Scanning Electron Microscopy Fig. Figure FITC Fluorescein IsoThioCyanate FL FLuorescent FSC Forward SCatter g Gravitational constant

g Gram(s) GLISA G-Protein-Linked ImmunoSorbent Assay GM17 M17 medium supplemented with 5% Glucose G-protein Guanine nucleotide-binding protein GST Glutathione S-Transferase h Hour(s) HE Hematoxylin and Eosin stain Hsd:ICR (CD1®) Harlan:Institute for Cancer Research (CD1®) HU Haemolytic Unit IgG Immunoglobulin G IPTG IsoPropylThioGalactoside IL InterLeukin ILY Intermedilysin of Streptococcus intermedius k Kilo kb Kilo base pair(s) kDa KiloDalton l Litre LB Luria Bertani LDH Lactate-Dehydrogenase LLO Listeriolysin O of Listeria monocytogenes LTA LipoTeichonic Acid M Molar (mol/l) MCP-1 Monocyte Chemotractic Protein-1 mF Milli Farad MFI Mean Fluorescent Intensity mg Milligram ml Millilitre mM Millimolar μg Microgram μl Microlitre µM Micromolar min Minute(s) MOI Multiplicity Of Infection MRP Muramidase-Released Protein ng Nano gram(s) nm Nanometer Ω Ohm O2 Oxygen ODXXX Optical Density at xxx nanometres OFS Opacity Factor of S. suis ORF Open Reading Frame PAGE PolyAcrylamide Gel Electrophoresis PBMEC Porcine Brain Microvascular Endothelial Cells PBS Phosphate Buffered Saline PCPEC Porcine Choroid Plexus Epithelial Cells PCR Polymerase Chain Reaction PFO Perfringolysin O of Clostridium perfringens

pH Power of Hydrogen PLY Pneumolysin of Streptococcus pneumoniae PMN Polymorphonuclear Neutrophil PTS PhosphoTransferase System PVDF PolyVinyliDene Fluoride Rac Ras-related C3 botulinum toxin substrate Rho Ras homolog gene family RNA RiboNucleic Acid rSLY Recombinant suilysin rSVD Recombinant RGD-SVD mutant of suilysin rW461F Recombinant W461F mutant of suilysin rpm Rounds per minute RT Room Temperature RT-PCR Reverse Transcriptase PCR ® Registered trademark s Second(s) SDS Sodium Dodecyl Sulphate SLY Suilysin SSC Sideward SCatter ST Sequence Type STSS Streptococcal Toxic Shock like Syndrome TACY Thiol-Activated CYtolysin TcdB Toxin B of Clostridium difficile TcdB-F Toxin B of Clostridium difficile Serotype F strain 1470 THB Todd Hewitt Broth Tig Trigger factor from S. suis TLR Toll-Like Receptor TM Trade mark TMH TransMembrane Hairpin TNL Tracheo-Nasal Lavage TNF-α Tumor Necrosis Factor alpha Tris Tris-(hydroxymethyl-) aminomethane TRITC TetramethylRhodamine-IsoThioCyanate U Unit V Volt wt wild type × Multiply Abbreviations of bacterial strains mentioned C. limosum Clostridium limosum C. defficile Clostridium difficile E. coli Escherischia coli L. lactis Lactococcus lactis St. aureus Staphylococcus aureus S. suis Streptococcus suis

List of figures and tables Figure 1-1: Crystal structure of suilysin and mechanism of pore-formation by CDC. ....................21

Figure 3-1: Suilysin promotes invasion of S. suis in HEp-2 epithelial cells. ...................................64

Figure 3-2: Suilysin-mediated invasion of S. suis in epithelial cells involves the actin

cytoskeleton and Rho-GTPases. .................................................................................66

Figure 3-3: Treatment of HEp-2 cells with recombinant suilysin (rSLY) leads to activation

of Rac1, but not RhoA. ................................................................................................67

Figure 3-4: Recombinant suilysin (rSLY) binds to HEp-2 cell membrane and is located in

association with F-actin and Rac1. ..............................................................................68

Figure 3-5: Substitution of an amino acid in the Trp-rich undecapeptide of domain 4 of

recombinant suilysin leads to abolishment of its haemolytic and cytolytic

activity. .........................................................................................................................70

Figure 3-6: Substitution of an amino acid in the Trp-rich undecapeptide of domain 4 of

recombinant suilysin does not affect its ability to promote HEp-2 cell

association of S. suis or to activate Rac1. ...................................................................72

Figure 4-1: Alignment of amino acid sequences of different CDC. ................................................98

Figure 4-2: Substitution of the RGD-motif of recombinant suilysin leads to abolishment

of its haemolytic and cytotoxic activity. .........................................................................99

Figure 4-3: Heterologous expression of SLY, but not W461F or SVD in L. lactis, confers

a haemolytic and cytotoxic phenotype. ..................................................................... 100

Figure 4-4: Substitution of the RGD-motif of recombinant suilysin leads to a loss in

membrane binding. . .................................................................................................. 104

Figure 4-5: Substitution of the RGD-motif leads to loss in Rac1-activation and

macropore-formation. ............................................................................................... 105

Figure 5-1: Histological findings of the nose and lung of mice infected intranasally with

5 x 109 CFU of S. suis after predisposition with 1% acetic acid. .............................. 122

Figure 6-1: Hypothetical model of suilysin-promoted interaction of S. suis with epithelial

cells............................................................................................................................ 159

Table 5-1: Clinical score sheet for health monitoring of mice after infection with S. suis. ......... 123

Table 5-2: Clinical course of CD1 mice intranasally infected with 5 x 109 CFU of S. suis

strain 10 (mrp+ epf+ sly+ cps2). ............................................................................... 124

Table 5-3: Scoring of fibrinosuppurative and purulent necrotizing lesions of mice

intranasally infected with 5 x 109 CFU of S. suis (mrp+ epf+ sly+ cps2). .................. 125

Table 5-4: Reisolation of the challenge strain S. suis strain 10 (mrp+ epf+ sly+ cps2) in

intranasally infected mice. ........................................................................................ 126

Table 5-5: Clinical course of Hsd:ICR (CD-1®) mice intranasally infected with 5 x 109

CFU of S. suis strain 10 (sly+, cps2), 10Δsly (sly-, cps2), 10cpsΔEF (sly+,

cps2-), and 10cpsΔEFΔsly (sly-, cps2-) after predisposition with 1% acetic

acid. ........................................................................................................................... 128

Table 5-6: Scoring of fibrinosuppurative and purulent necrotizing lesions of Hsd:ICR

(CD-1®) mice intranasally infected with 5 x 109 CFU of S. suis strain 10 (sly+,

cps2), 10Δsly (sly-, cps2), 10cpsΔEF (sly+, cps2-), and 10cpsΔEFΔsly (sly-,

cps2-) after predisposition with 1% acetic acid. ........................................................ 129

Table 5-7: Reisolation of the challenge strains S. suis strain 10 (sly+, cps2), 10Δsly

(sly-, cps2), 10cpsΔEF (sly+, cps2-), and 10cpsΔEFΔsly (sly-, cps2-) in

intranasally infected Hsd:ICR (CD-1®) after predisposition with 1% acetic. . ............ 130

Table 5-S1: Register of all infected mice including the clinical course of infection, scoring

of fibrinosuppurative and purulent necrotizing lesions and reisolation of the

challenge strain (table includes intravenously infected mice). ................................. 140

Chapter 1

General introduction

General introduction Chapter 1

15

1. Streptococcus (S.) suis S. suis is a gram-positive, facultative anaerobic bacterium causing invasive diseases

in swine worldwide, associated with meningitis, septicaemia, arthritis, endocarditis

and bronchopneumonia. In the last years S. suis has been considered as an

important human pathogen leading to bacterial meningitis and the life-threatening

streptococcal toxic shock like syndrome (STSS) in humans (Gottschalk et al., 2010;

Tang et al., 2006). The bacterium shows α-haemolysis on sheep blood agar plates

and α- and β-haemolysis on horse blood agar plates.

On the basis of the capsule polysaccharides, 33 different serotypes have been

described so far, of which serotype 2 is worldwide most frequently isolated from

diseased pigs and humans in Europe and Asia (Gottschalk et al., 2010; Silva et al.,

2006; Wei et al., 2009). Distribution of serotypes differs between geographical

regions. Serotype 9 has emerged as the most common pig isolate in Germany and

The Netherlands, serotype 7 is most prevalent in Scandinavia and Germany, and

serotype 1 and 14 in the United Kingdom (Baums and Valentin-Weigand, 2009;

Perch et al., 1983; Tian et al., 2004; Wisselink et al., 2000). In contrast, in Canada

and the USA serotypes 2, 1/2 and 3 are most frequently associated with disease

(Messier et al., 2008). Noteworthy, a specific sequence type (ST), namely ST7,

evolved from the highly pathogenic ST1 type of a serotype 2 strain, which was found

to be responsible for human outbreaks in china and directly associated with the

STSS (Ye et al., 2006; Ye et al., 2009). The ST7 carries a putative pathogenicity

island (designated 89K), possibly involved in development of STSS (Chen et al.,

2007; Zhao et al., 2011).

1.1. S. suis infections S. suis can infect pigs of each age group, including suckling and weaning piglets as

well as growers. The natural habitat of S. suis is the upper respiratory tract. S. suis

colonizes the nasopharynx, in particular the tonsils and the nasal cavities, as well as

other mucosal surfaces like the intestinal and genital tract asymptomatically, resulting

in a high carrier rate of healthy pigs of up to 100% (Arends et al., 1984; Clifton-

Hadley et al., 1986; Higgins and Gottschalk, 1990; Lowe et al., 2011; O'Sullivan et

General introduction Chapter 1

16

al., 2011). Such carrier-pigs are the most important source of infection (Clifton-

Hadley and Alexander, 1980). Horizontal transmission of disease occurs most

frequently via the direct 'nose-to-nose' contact between healthy and infected pigs,

besides airborne transmission of S. suis has also been shown in experimental

infected piglets (Berthelot-Herault et al., 2001). Oral infection (feed) and transmission

via skin wounds is also possible. Vertical transmission of S. suis via the navel or the

genital tract is another relevant route of infection (Amass et al., 1997; Robertson et

al., 1991; Staats et al., 1997). Furthermore, insects (houseflies) as potential vectors

have been discussed (Enright et al., 1987).

Although the morbidity rate of pigs is less than 5% due to prophylaxis with antibiotics

(Clifton-Hadley et al., 1986), in the case of disease the mortality rate can reach 20%

in the absence of treatment. Several forms of streptococcal disease in pigs are

known. The occurrence of sudden death due to a peracute septicaemia as well as

the development of severe meningitis, polyarthritis and bronchopneumonia can be

observed. Acute infections are indicated by high fever (>40˚C), persistent anorexia

followed by lameness and central nervous failure (Sanford and Tilker, 1982; Windsor

and Elliott, 1975). Typical histopathological lesions are characterized by an acute

fibrinosuppurative inflammation of the respective tissue (Beineke et al., 2008;

Williams and Blakemore, 1990).

As an important emerging zoonotic agent S. suis has gained public interest due to

increased reports on human infections (Gottschalk et al., 2007; Trottier et al., 1991;

Wertheim et al., 2009). To date human cases of S. suis infections are mainly

reported in Asia. In Vietnam S. suis is considered as the most common causative

pathogen of bacterial meningitis (second most common in Thailand and third most

frequent in Hong Kong) (Mai et al., 2008; Petersen et al., 2011; Sriskandan and

Slater, 2006; Wangkaew et al., 2006). Particularly humans exposed to infected pigs

or contaminated pig-products are at risk (Arends and Zanen, 1988). An additional

potential source of infection for humans are wild boars (Baums et al., 2007).

Therefore, S. suis is an occupational disease of people in close contact with swine,

like farmers, butcher, hunters and veterinarians (high-risk group). Hence, the

nasopharyngeal carrier rate of S. suis serotype 2 strains of persons belonging to the

General introduction Chapter 1

17

high-risk group in Germany was 5.3% (Strangmann et al., 2002). For humans without

swine contact oral transmission via raw pork or contaminated pig-products is

possible, whereas a 'human-to-human' infection has not been proven so far

(Wertheim et al., 2009). After an incubation period ranging from a few hours to five

days post infectionem, purulent meningitis, septicaemia and arthritis associated with

leukocytosis and neutrophilia are the most common manifestations in humans

(Arends and Zanen, 1988; Fongcom et al., 2001). A serious consequence following

S. suis meningitis is chronic deafness (Navacharoen et al., 2009). Two human

outbreaks in China in 1998 and 2005 were associated with increased severeness of

clinical symptoms. A noticeable high incidence of the STSS, which is characterized

by high fever, erythoderma and multi organ failure (liver, heart, kidney, CNS) was

observed, resulting in a high mortality rate of more than 20% (Tang et al., 2006; Yu et

al., 2006).

1.2. Pathogenesis and virulence mechanisms The mechanisms underlying pathogenesis of S. suis infections are only poorly

known. A hypothetical model of pathogenesis includes three main steps. Firstly, after

colonisation the mucosal surface of the upper respiratory tract, S. suis invades into

deeper tissues of the epithelium. Secondly, the bacterium disseminates within the

bloodstream to finally cross the endothelium of target tissues, such as the blood brain

barrier (BBB) of the central nervous system (CNS) to cause meningitis (Chanter et

al., 1993; Gottschalk and Segura, 2000). S. suis produces a wide array of virulence

and virulence-associated factors, either secreted or surface-associated, involved in

this process. A comprehensive review of bacterial factors expressed by S. suis was

recently published (Baums and Valentin-Weigand, 2009).

Bacterial factors such as the fibronectin and fibrinogen binding protein (FBPS) (de

Greeff et al., 2002) or the cell wall component lipoteichonic acid (LTA) (Fittipaldi et

al., 2007; Vanier et al., 2007) mediate adherence of bacteria to target cells for initial

colonisation. Another bacterial mechanism for sufficient colonisation is the formation

of biofilms probably enhancing bacterial resistance to innate and adaptive host

defence mechanisms and treatment with antibiotics (Bonifait et al., 2008). To get

General introduction Chapter 1

18

access into deeper tissues bacteria might invade the respiratory epithelium. Suilysin,

the haemolysin of S. suis, is discussed to play a role in interaction of S. suis with

epithelial cells and disruption of these cells due to its cytolytic function (cf. 2.4.2.).

Furthermore, the capsule is assumed to be involved in host cell interaction. Since its

main function is protection against phagocytosis after entering the bloodstream

(Benga et al., 2008; Chabot-Roy et al., 2006; Charland et al., 1998; Segura and

Gottschalk, 2002), it has been proposed that the capsule is down-regulated during

colonisation of the mucosal epithelium to allow adherence and invasion of the

bacterium to overcome this first barrier within the host (Gottschalk and Segura, 2000;

Okamoto et al., 2004; Willenborg et al., 2011). In accordance, unencapsulated

S. suis stains showed higher adhesion and invasion rates, indicating a negative

correlation between encapsulation and interaction with host cells (Benga et al., 2004;

Gottschalk et al., 1991). A possible explanation for this phenotype is the masking

effect of the capsule, whereby potentially involved surface associated proteins or cell

wall components might be hidden (Lalonde et al., 2000; Tenenbaum et al., 2008;

Vanier et al., 2007). Moreover, a direct uptake of S. suis by monocytes for crossing

the epithelium as well as for entering the bloodstream within circulating cells, known

as the 'Trojan horse theory', is controversially discussed due to the protecting effect

of the capsule. A 'travelling' of either free bacteria or monocyte-associated bacteria

('modified Trojan horse theory') is more likely (Gottschalk and Segura, 2000).

Circulation of bacteria within the bloodstream may lead to onset of acute bacteraemia

or septicaemia and the release of several pro-inflammatory cytokines by cells of the

innate immune system to control acute infection or to contribute to immunopathology

(Segura et al., 1999; Segura et al., 2002; Segura et al., 2006).

Nevertheless, for induction of meningitis S. suis has to penetrate the BBB to reach

the CNS. S. suis has the ability to adhere to and invade into brain microvascular

endothelial cells (BMEC) and porcine choroid plexus epithelial cells (PCPEC), the

main components of the BBB (Benga et al., 2005; Charland et al., 2000; Tenenbaum

et al., 2005; Tenenbaum et al., 2008; Vanier et al., 2004). Moreover, an increase in

tight junction permeability and loss of barrier function is ascribed to direct cytotoxic

effects of suilysin (Charland et al., 1998; Vanier et al., 2004). Apart from suilysin

General introduction Chapter 1

19

S. suis can stimulate the production of pro-inflammatory cytokines like interleukin-6

(IL-6), IL-8 and monocyte chemotactic protein-1 (MCP-1) by BMEC, which in turn

alters BBB permeability (Vadeboncoeur and Pelletier, 1997). However, Tenenbaum

et al. (2009) described the entry S. suis into the CNS as a transcellular translocation

without destruction of PCPEC lining of the BBB.

2. Cholesterol-dependent pore-forming cytolysins (CDC) Suilysin, the secreted haemolysin of S. suis, belongs to the family of cholesterol-

dependent pore-forming cytolysins (CDC) and was considered as a virulence-

associated factor, which contributes to pathogenesis of S. suis. CDC are a large

family of membrane-damaging toxins produced by more than 20 gram-positive

bacteria (Tweten, 2005), including the genera Streptococcus, Listeria, Bacillus,

Clostridum, Paenibacillus, Arcanobacterium and most recently Gardnerella (Gelber et

al., 2008) and Lactobacillus (Rampersaud et al., 2011).

2.1. The structure of CDC All common CDC show a high similarity in primary amino acid sequence varying

between 40-70% identity. CDC are single stranded peptides showing an elongated

rod-like three-dimensional structure, which was first described for the prototype of

CDC, perfringolysin O of Clostrium perfringens.

CDC were previously named thiol-activated cytolysins (TACY) due to the fact, that

most members carry a single cysteine residue within the undecapeptide at the

carboxy-terminus of the molecule. This cysteine residue is believed to be essential

for membrane binding and the cytolytic function. It is known that oxidation of carbon-

bonded sulfohydryl groups (formation of disulfide bonds) inhibits CDC. The name

'thiol-activated' derives from the properties of thiols, which are able to reactivate the

toxin via the cleavage of disulfide bonds of the cysteine residue. However, two

members of the toxin family, pyolysin (Arcanobacterium pyogenes) (Billington et al.,

1997) and intermedilysin (Streptococcus intermedius) carry an alanin residue instead

of the cysteine without any deficiency in lysis activity (Herbert and Todd, 1941).

Moreover, mutation studies targeting the cysteine residue did not alter the cytolysis

General introduction Chapter 1

20

process (Michel et al., 1990; Pinkney et al., 1989; Saunders et al., 1989; Stachowiak

et al., 2009). In contrast to the cytolytic activity, cysteine substitution of listeriolysin O

decreased invasion of Listieria monocytogenes (Stachowiak et al., 2009).

Nevertheless, these findings clearly show, that the cysteine residue is not required

for cytolytic function of the toxin, therefore the name 'thiol-activated' is no longer

appropriate.

The monomeric protein contains four domains (D1-D4, Figure 1-1A): D1 and D3 are

located at the N-terminus of the protein, linked to D4 via the connecting domain D2

(Rossjohn et al., 1997). The main part of the protein is formed by β-strands (43%),

whereas in particular D3 also contains a couple of α-helixes (14% α-helices of the

whole structure), playing a special role in pore-formation (Xu et al., 2010). The most

conserved regions are the C-terminal tryptophan-rich undecapeptide (Figure 1-1A),

consisting of eleven amino acids, (ECTGLAWEWWR; c.f. Chapter 4, Figure 4-2) of

D4 and the hydrophobic core of D1. The undecapeptide is considered to be

responsible for the initial membrane binding and triggers the formation of functional

pores (Rossjohn et al., 1997). Supportively, it has been shown, that recombinant D4

alone is still able to bind to membranes of erythrocytes (Weis and Palmer, 2001). In a

recent study the crystal structure and D4-folding of CDC, including suilysin,

intermedilysin, perfringolysin O and anthrolysin (Bacillus anthracis), was analyzed in

more detail. Bending-degree of the angle between D1 and D4 as well as the

conformations of the tryptophan-rich loop on the tip of the undecapeptide were

compared. Predictions of intermedilysin and suilysin revealed a more extended

structure (Figure 1-1A), whereas the undecapeptide of perfringolysin O and

anthrolysin was folded back into a hydrophobic pocket (Xu et al., 2010).

General introduction Chapter 1

21

Figure 1-1: Crystal structure of suilysin and the mechanism of pore-formation by CDC. (A) Ribbon diagram of suilysin crystal structure modified from Xu et al., 2010. Each domain (D1-D4) is coloured individually and the location of the tryptophan-rich undecapeptide, the two transmembrane β-hairpins (THM1 and THM2) and the three loops (L1-3) are indicated by arrows. (B) The putative CDC pore-forming mechanism modified from Tweten et al., 2005. The CDC molecule is secreted as a soluble monomer consisting of 4 domains (D1-D4) and binds to cholesterol-containing target membranes via D4 (1). Several monomers interact to form the arc-shaped prepore complex triggered by structural changes within D3 (2), followed by formation of the preinsertion transmembrane β-barrel structure (TMH1 and 2) (3). Upon collapse of D2 the oligomer inserts into the membrane generating a permeable pore (4). (C) 3D-reconstruction of electronmicroscopy images of liposomes with the prepore complex and the permeable pore of pneumolysin modified from Tilley et al., 2005.

2.2. The mechanism of pore-formation by CDC All members of the CDC family are secreted as water soluble monomers due to a

signal sequence at the N-terminus of the protein, with the exception of pneumolysin

(Streptococcus pneumoniae), which is not released until lysis of the bacterial cell

(Walker et al., 1987). A common, but still putative, mechanism of pore-formation was

described by Tweten et al., 2005. Generation of channels is a 4-step process (Figure

1-1B) (Tweten, 2005). (1) Initially several monomers bind vertically to the target cell

membrane. Interaction between the CDC molecule and the cell is mediated via D4,

forming a hydrophobic 'danger' consisting of the undecapeptide, which superficially

General introduction Chapter 1

22

inserts into the membrane (Rossjohn et al., 1998). Binding of D4 to the membrane is

necessary for subsequent structural changes within the not directly connected D3,

called 'conformational coupling' (Rossjohn et al., 2007). Subsequently, cell-bound

monomers undergo a lateral shift to form an oligomer. (2) Further structural changes

within D3 lead to the formation of the oligomeric ring- or arc-shaped prepore complex

(Figure 1-1C). (3) The N-terminal D3 builds the inner structure of the transmembrane

pore. Therefore, six α-helices form two β-hairpins (Figure 1-1A; TMH1 and TMH2).

This ultimate β-barrel structure consists of several β-hairpins of agminated

monomers, which are connected through an intermolecular interaction of the

β1-strand and β4-strand of two individual molecules (Ramachandran et al., 2004). (4)

In the final step, the connecting domain D2 severs as a 'hinge-joint' to bring D3 in

close contact with the cell surface. After the 'collapse' of D2 the β-barrel structure

inserts into the membrane and forms the transmembrane pore. This pore, 30 nm in

size, consists of 35-50 monomers and is permeable for ions and macro molecules

(Figure 1-1C).

2.3. The functional role of cholesterol and membrane recognition The role of cholesterol in CDC function is not finally clarified (Hotze and Tweten,

2011). Nevertheless, a general observation is that cholesterol is required for pore-

formation and free cholesterol inhibits haemolysis of CDC (Alouf, 2000; Jacobs et al.,

1994; Watson et al., 1972). Furthermore, cholesterol-depletion studies using

cholesterol-containing liposomes showed that >30mol% of total membrane lipids has

to be cholesterol for efficient membrane binding (Flanagan et al., 2009; Heuck et al.,

2000). However, not all CDC use cholesterol as a membrane receptor. It has been

shown that the human complement regulator molecule CD59 is required for

membrane binding of intermedilysin and vaginolysin, thereby conferring a species

specificity. However, cell lysis activity of intermedilysin and vaginolysin depends on

cholesterol (Gelber et al., 2008; Giddings et al., 2004).

The cholesterol recognition motif (CRM) is most likely located within D4. Several

studies concerning the CRM defined the undecapeptide to be responsible for

cholesterol binding (Rossjohn et al., 1997; Rossjohn et al., 2007). In contrast, more

General introduction Chapter 1

23

recent studies revealed three highly conserved loops (Figure 1-1A; L1-3) located next

to the undecapeptide to mediate recognition of cholesterol (Soltani et al., 2007;

Soltani et al., 2007b). Furthermore, it has been suggested that CDC present only a

single binding side due to the fact that cholesterol binding activity is linear to CDC

concentration (Johnson et al., 1980).

2.4. Suilysin Suilysin was identified as a haemolysin of S. suis several years ago (Jacobs et al.,

1994) and confirmed to be a member of CDC by sequence analysis. The protein has

a molecular weight of 54 kDa and possesses an N-terminal signal sequence, thus

considered as a secreted exotoxin. Suilysin is most related to pneumolysin, sharing

52% amino acid identity (Segers et al., 1998).

2.4.1. Prevalence, diversity, and regulation of the sly gene Suilysin is expressed by many but not all S. suis strains. The sly gene has been

detected in 95% of European and Asian invasive serotype 2 strains, but only in 7% of

the North American strains (Segers et al., 1998). In various studies concerning the

prevalence of the sly gene in isolates, belonging to different S. suis serotypes,

obtained from diseased pigs in European countries, including Germany, Italy, Spain,

Poland, France, The Netherlands and The United Kingdom, the sly gene was

detectable in a range between 33.7% and 100% (Berthelot-Herault et al., 2000;

Blume et al., 2009; de Greeff et al., 2011; Fabisiak et al., 2005; King et al., 2001;

Princivalli et al., 2009; Silva et al., 2006; Tarradas et al., 2001). Approximately 80%

of these clinical isolates were obtained from porcine cases of invasive disease

associated with meningitis, septicaemia, and arthritis. Isolates derived from lung

samples (pneumonia) were positive tested for suilysin to a lesser degree of

approximately 50% (King et al., 2001; Silva et al., 2006). Furthermore, the sly gene is

prevalent in up to 90% of colonising S. suis strains isolated from the tonsils of healthy

pigs (Fabisiak et al., 2005; King et al., 2001). Besides, suilysin is present in wild

boars and domestic swine S. suis strains of Northwestern Germany in 39% and 66%,

respectively (Baums et al., 2007). In Asian S. suis strains, isolated from healthy

General introduction Chapter 1

24

(slaughtered) as well as from diseased swine, the prevalence of suilysin reaches a

maximum of 100% (Hoa et al., 2011; Kim et al., 2010; Padungtod et al., 2010; Wei et

al., 2009; Xiong et al., 2007). Equally, almost all isolates from human cases of a

S. suis infection carry a sly gene (de Greeff et al., 2011; King et al., 2001; Princivalli

et al., 2009). The situation is different for North American isolates, in which suilysin is

less frequently present (Fittipaldi et al., 2009; Gottschalk et al., 1998).

However, presence of the sly gene has been used for characterisation and

differentiation of S. suis strains (Gottschalk et al., 2007; Vecht et al., 1991), but

presence of the sly gene does not necessarily result in protein expression (de Greeff

et al., 2011).

Suilysin was detected as a highly conserved single copy gene within the S. suis

genome (Okwumabua et al., 1999). Furthermore, the genetic diversity as well as the

allelic variance of the sly gene appears to be low (King et al., 2001). The flanking

genes (2 open reading frames (ORF) upstream, orf100 and orf101; and 2 ORF

downstream, nanE and pstG) are also highly conserved. Sly-negative strains exhibit

an alternative gene (orf102) at the same position instead of the sly gene. Therefore,

genetic exchange via homologous recombination between different S. suis strain is

most likely (Takamatsu et al., 2002). Moreover, it is conceivable that the sly gene is

not organised as an operon and under the control of its own promoter, because of

large non coding region upstream and downstream of the sly gene.

Little is known about the regulation of the sly gene. The protein is expressed at late

logarithmic growth phase, possibly dependent on nutrient availability, pH and

bacterial density (Gottschalk et al., 1995). A hyper-haemolytic phenotype is

described for a manN-negative S. suis mutant strain, suggesting that the mannose

phosphotransferase system affects the suilysin promoter activity (Lun et al., 2003).

The global orphon response regulator CovR (control of virulence regulator) controls

the expression of about 200 genes, including the capsule biosynthesis and suilysin.

Inactivation of covR led to the production of a thicker capsule and slightly higher

haemolytic activity associated with increased adhesion to epithelial cells and reduced

phagocytosis and killing by polymorphonuclear neutrophils (PMN) (Pan et al., 2009).

In contrast to other CDC, like intermedilysin, which is under the transcriptional

General introduction Chapter 1

25

catabolite control protein A (CcpA), the homologous global regulator of S. suis is

most likely not involved in regulation of suilysin (Tomoyasu et al., 2010; Willenborg et

al., 2011). Besides, expression of the sly gene is influenced by the trigger factor from

S. suis (Tig) and the orphan response regulator RevSC21. Both, deletion of the tig

gene and the RevSC21 gene degraded expression of suilysin and resulted in a lack

of haemolytic activity of the respective S. suis strain (Wu et al., 2009; Wu et al.,

2011).

2.4.2. Role of suilysin in host-pathogen interaction Similar to other members of the CDC family suilysin can damage different types of

host cells by its cytolytic activity (Benga et al., 2004; Charland et al., 2000; Jacobs et

al., 1994; Lalonde et al., 2000; Norton et al., 1999; Segura and Gottschalk, 2002;

Tenenbaum et al., 2005; Tenenbaum et al., 2006; Vanier et al., 2004). A suilysin-

induced cell injury is associated with loss of cytoplasmic density, disruption of

cytoplasmic membranes and the release of cellular contents (Allen et al., 2001;

Segura and Gottschalk, 2002). Haemolysis of erythrocytes was first described by

Jacobs et al. (1994), whereas human red blood cells were the most susceptible,

followed by horse, sheep, and cow erythrocytes (Gottschalk et al., 1995).

A multifunctional role in pathogenesis was described for other members of the CDC-

family. These biological effects can be observed even at subcytolytic concentration of

the respective toxins (Billington et al., 2000). For pneumolysin, listeriolysin O, and

intermedilysin was reported that they may contribute to bacterial adherence and

invasion (Cockeran et al., 2002; Krawczyk-Balska and Bielecki, 2005; Rubins et al.,

1998; Sukeno et al., 2005). Likewise, suilysin has been described to be involved in

the modulation of S. suis host cell interaction, including endothelial cells, epithelial

cells, PMN and macrophages (Baums and Valentin-Weigand, 2009; Gottschalk and

Segura, 2000). In particular, it has been suggested that suilysin plays a role in

pathogenesis of S. suis (Norton et al., 1999) such as the crossing of the BBB by

interruption of intracellular junctions for entering the CNS (Charland et al., 2000;

Tenenbaum et al., 2005; Tenenbaum et al., 2008). However, the ability of several

General introduction Chapter 1

26

S. suis strains to interact with PBMEC does not correlate with suilysin production

(Vanier et al., 2004; Vanier et al., 2007).

More recently, it has been found that the toxin may be involved in cytokine release

and protection against opsonophagocytosis. Thus, a reduced survival time of a

sly-deficient mutant strain after co-incubation with polymorphonuclear neutrophils

(PMN) was observed by Benga et al. (2008). This was confirmed by using antisera

raised against purified suilysin, which increased the uptake of the wild type strain by

PMN. In addition, adding recombinant suilysin at subcytolytic concentration increased

the growth capacity of the sly-deficient mutant (Benga et al., 2008). Furthermore,

suilysin contributes to resistance of complement-dependent killing of S. suis by

neutrophils (Chabot-Roy et al., 2006), perhaps by impairing complement factor C3

deposition on the surface of S. suis (Lecours et al., 2011). For pneumolysin,

reduction of serum complement levels and decreased opsonisation of pneumococci

has been described as well (Alcantara et al., 2001). Although, S. suis was found to

be resistant to phagocytosis by murine astrocytes, suilysin was mainly responsible

for pro-inflammatory cytokine production and partially involved in toll-like receptor 2

(TLR2) expression of these cells (Zheng et al., 2011). Similarly, recognition of

pneumolysin via the TLR4 is critically involved in the innate immune response to

pneumococci (Dessing et al., 2009; Malley et al., 2003).

In general, interference of suilysin with different types of immune cells and induction

of cytokine release suggests the important role of suilysin in host innate defence

response. Suilysin is responsible for the release of IL-6 and IL-8 by BMEC

(Vadeboncoeur et al., 2003) and PBMEC (Vanier et al., 2008) as well as for the

production of tumour necrosis factor α (TNF-α) by human monocytes and IL-6 by

porcine pig pulmonary alveolar macrophages and monocytes (Lun et al., 2003). In

contrast, suilysin did not have a major impact on phagocytosis as well as on TNF-α

and MCP-1 production by murine microglia (Dominguez-Punaro et al., 2010).

Accordingly, suilysin failed to induce TNF-α and IL-6 in murine macrophage line J774

(Segura et al., 1999) and plays a limited role in modulation of cytokines and

chemokine response in a whole-blood system (Segura et al., 2006). An involvement

in cytokine production by bone marrow–derived dendritic cells was recently described

General introduction Chapter 1

27

by (Lecours et al., 2011). In inflammatory events increased recruitment of leucocytes

is linked to adhesion molecules. Stimulation of THP-1 monocytes with suilysin led to

an up-regulation of CD11a/CD18 and CD11c/CD18 (Al Numani et al., 2003). Most of

the other CDC have also been shown to display a role in modulation of immune

response mechanisms (Cockeran et al., 2002; Ratner et al., 2006; Tsuchiya et al.,

2005).

2.4.3 Role of suilysin in virulence and pathogenesis The importance of suilysin in the pathogenesis of S. suis is not finally clarified. While

suilysin has been shown to be associated with virulent strains, there are also virulent

strains that do not produce suilysin (Staats et al., 1999). Only few experimental infections of mice and piglets addressing the role of suilysin

were performed so far. Intraperitonally infection of BALB/c mice demonstrated

attenuation of a sly-deficient mutant in comparison to the highly virulent S. suis

serotype 2 strain P1/7. In contrast, the sly knock-out strain was only slightly

attenuated (reduced severeness of clinical signs and pathological findings) in an

intravenous porcine infection model (Allen et al., 2001), indicating that suilysin does

not play a major role in disease development after systemic administration. Similar

results were obtained by Lun et al. (2003) using three different sly-deficient strains in

an intrapharyngeal piglet infection model. All swine developed clinical symptoms

associated with septicaemia, arthritis, and meningitis regardless of the challenge

strain. However, vaccination containing purified suilysin was capable to induce

protection in BALB/c mice and piglets after challenge with a homologous strain

(S. suis P1/7). Furthermore, immunisation led to an increase in haemolysin

neutralisation antibody titre (Jacobs et al., 1994; Jacobs et al., 1996). Additionally,

intranasal immunisation of piglets with a S. suis live vaccine elicited most prominently

serum immunoglobulin G responses against suilysin (Kock et al., 2009). Likewise,

the highly related pneumolysin protects mice against homologues challenge and

moreover a pneumolysin-deficient mutant was attenuated in a BALB/c infection

model (Alexander et al., 1994; Orihuela et al., 2004).

General introduction Chapter 1

28

3. Outline of the present study As pointed out in the general introduction, knowledge on pathogenesis of S. suis

infection and potentially involved bacterial factors is still limited. The aim of this study

was to evaluate suilysin-mediated effects on S. suis-epithelial cell interaction

including investigations on the underlying mechanisms and the phenotypical

characterisation of putatively involved functional domains of the suilysin molecule.

Furthermore, the study focussed on the development of an intranasal mouse

infection model to further elucidate the role of suilysin in colonisation and invasion of

the upper respiratory tract in vivo.

According to these objectives, the results are divided into 3 chapters. In chapter 3

the role of suilysin in S. suis invasion in epithelial cells is investigated, revealing a

subcytolytic activity of suilysin in a Rac-dependent activation of the actin cytoskeleton

promoting invasion of S. suis. Chapter 4 comprises the characterisation of two

(functional) domains within the suilysin molecule. Site-directed amino acid

substitution and comparative functional analysis of the tryptophan-rich undecapeptide

and a putative integrin-binding RGD-motif indicate that both domains are required for

cytolytic function. Additionally, a functional RGD-motif is essential for membrane-

binding and activation of Rac. Finally, in chapter 5 the establishment of an intranasal

S. suis mouse infection model is described. The results are generally discussed in

chapter 6 with regard to relevance for S. suis pathogenesis. A short summary of this

thesis is provided in chapter 7 (English) and chapter 8 (German).

General introduction Chapter 1

29

References Al Numani, D., Segura, M., Dore, M., and Gottschalk, M. (2003) Up-regulation of ICAM-1, CD11a/CD18 and CD11c/CD18 on human THP-1 monocytes stimulated by Streptococcus suis serotype 2. Clin Exp Immunol 133: 67-77. Alcantara, R. B., Preheim, L. C., and Gentry-Nielsen, M. J. (2001) Pneumolysin-induced complement depletion during experimental pneumococcal bacteremia. Infect Immun 69: 3569-3575.

Alexander, J. E., Lock, R. A., Peeters, C. C., Poolman, J. T., Andrew, P. W., Mitchell, T. J., Hansman, D., and Paton, J. C. (1994) Immunization of mice with pneumolysin toxoid confers a significant degree of protection against at least nine serotypes of Streptococcus pneumoniae. Infect Immun 62: 5683-5688.

Allen, A. G., Bolitho, S., Lindsay, H., Khan, S., Bryant, C., Norton, P., Ward, P., Leigh, J., Morgan, J., Riches, H., Eastty, S., and Maskell, D. (2001) Generation and characterization of a defined mutant of Streptococcus suis lacking suilysin. Infect Immun 69: 2732-2735.

Alouf, J. E. (2000) Cholesterol-binding cytolytic protein toxins. Int J Med Microbiol 290: 351-356.

Amass, S. F., SanMiguel, P., and Clark, L. K. (1997) Demonstration of vertical transmission of Streptococcus suis in swine by genomic fingerprinting. J Clin Microbiol 35: 1595-1596.

Arends, J. P., Hartwig, N., Rudolphy, M., and Zanen, H. C. (1984) Carrier Rate of Streptococcus-Suis Capsular Type-2 in Palatine Tonsils of Slaughtered Pigs. Journal of Clinical Microbiology 20: 945-947.

Arends, J. P. and Zanen, H. C. (1988) Meningitis Caused by Streptococcus Suis in Humans. Reviews of Infectious Diseases 10: 131-137.

Baums, C. G. and Valentin-Weigand, P. (2009) Surface-associated and secreted factors of Streptococcus suis in epidemiology, pathogenesis and vaccine development. Anim Health Res Rev 10: 65-83.

Baums, C. G., Verkuhlen, G. J., Rehm, T., Silva, L. M., Beyerbach, M., Pohlmeyer, K., and Valentin- Weigand, P. (2007) Prevalence of Streptococcus suis genotypes in wild boars of Northwestern Germany. Appl Environ Microbiol 73: 711-717.

Beineke, A., Bennecke, K., Neis, C., Schroder, C., Waldmann, K. H., Baumgartner, W., Valentin- Weigand, P., and Baums, C. G. (2008) Comparative evaluation of virulence and pathology of Streptococcus suis serotypes 2 and 9 in experimentally infected growers. Vet Microbiol 128: 423-430.

Benga, L., Friedl, P., and Valentin-Weigand, P. (2005) Adherence of Streptococcus suis to porcine endothelial cells. J Vet Med B Infect Dis Vet Public Health 52: 392-395.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cellular Microbiology 6: 867-881.

Berthelot-Herault, F., Gottschalk, M., Labbe, A., Cariolet, R., and Kobisch, M. (2001) Experimental airborne transmission of Streptococcus suis capsular type 2 in pigs. Vet Microbiol 82: 69-80.

General introduction Chapter 1

30

Berthelot-Herault, F., Morvan, H., Keribin, A. M., Gottschalk, M., and Kobisch, M. (2000) Production of Muraminidase-Released Protein (MRP), Extracellular Factor (EF) and Suilysin by field isolates of Streptococcus suis capsular types 2, 1/2, 9, 7 and 3 isolated from swine in France. Veterinary Research 31: 473-479.

Billington, S. J., Jost, B. H., Cuevas, W. A., Bright, K. R., and Songer, J. G. (1997) The Arcanobacterium (Actinomyces) pyogenes hemolysin, pyolysin, is a novel member of the thiol- activated cytolysin family. J Bacteriol 179: 6100-6106.

Billington, S. J., Jost, B. H., and Songer, J. G. (2000) Thiol-activated cytolysins: structure, function and role in pathogenesis. FEMS Microbiol Lett 182: 197-205.

Blume, V., Luque, I., Vela, A. I., Borge, C., Maldonado, A., Dominguez, L., Tarradas, C., and Fernandez-Garayzabal, J. F. (2009) Genetic and virulence-phenotype characterization of serotypes 2 and 9 of Streptococcus suis swine isolates. Int Microbiol 12: 161-166.

Bonifait, L., Grignon, L., and Grenier, D. (2008) Fibrinogen induces biofilm formation by Streptococcus suis and enhances its antibiotic resistance. Appl Environ Microbiol 74: 4969-4972.

Chabot-Roy, G., Willson, P., Segura, M., Lacouture, S., and Gottschalk, M. (2006) Phagocytosis and killing of Streptococcus suis by porcine neutrophils. Microb Pathog 41: 21-32.

Chanter, N., Jones, P. W., and Alexander, T. J. (1993) Meningitis in pigs caused by Streptococcus suis--a speculative review. Vet Microbiol 36: 39-55.

Charland, N., Harel, J., Kobisch, M., Lacasse, S., and Gottschalk, M. (1998) Streptococcus suis serotype 2 mutants deficient in capsular expression. Microbiology-Sgm 144: 325-332. Charland, N., Nizet, V., Rubens, C. E., Kim, K. S., Lacouture, S., and Gottschalk, M. (2000) Streptococcus suis serotype 2 interactions with human brain microvascular endothelial cells. Infect Immun 68: 637-643.

Chen, C., Tang, J., Dong, W., Wang, C., Feng, Y., Wang, J., Zheng, F., Pan, X., Liu, D., Li, M., Song, Y., Zhu, X., Sun, H., Feng, T., Guo, Z., Ju, A., Ge, J., Dong, Y., Sun, W., Jiang, Y., Wang, J., Yan, J., Yang, H., Wang, X., Gao, G. F., Yang, R., Wang, J., and Yu, J. (2007) A glimpse of streptococcal toxic shock syndrome from comparative genomics of S. suis 2 Chinese isolates. PLoS ONE 2: e315.

Clifton-Hadley, F. A. and Alexander, T. J. (1980) The carrier site and carrier rate of Streptococcus suis type II in pigs. Vet Rec 107: 40-41.

Clifton-Hadley, F. A., Alexander, T. J., and Enright, M. R. (1986) Monitoring herds for Streptococcus suis type 2: chance contamination of slaughter pigs. Vet Rec 118: 274.

Cockeran, R., Durandt, C., Feldman, C., Mitchell, T. J., and Anderson, R. (2002) Pneumolysin activates the synthesis and release of interleukin-8 by human neutrophils in vitro. J Infect Dis 186: 562-565.

de Greeff, A., Buys, H., Verhaar, R., Dijkstra, J., van Alphen, L., and Smith, H. E. (2002) Contribution of fibronectin-binding protein to pathogenesis of Streptococcus suis serotype 2. Infect Immun 70: 1319-1325.

de Greeff, A., Wisselink, H. J., de Bree, F. M., Schultsz, C., Baums, C. G., Thi, H. N., Stockhofe- Zurwieden, N., and Smith, H. E. (2011) Genetic diversity of Streptococcus suis isolates as determined by comparative genome hybridization. BMC Microbiol 11: 161.

General introduction Chapter 1

31

Dessing, M. C., Hirst, R. A., de Vos, A. F., and van der, Poll T. (2009) Role of Toll-like receptors 2 and 4 in pulmonary inflammation and injury induced by pneumolysin in mice. PLoS One 4: e7993.

Dominguez-Punaro, Mde L., Segura, M., Contreras, I., Lachance, C., Houde, M., Lecours, M. P., Olivier, M., and Gottschalk, M. (2010) In vitro characterization of the microglial inflammatory response to Streptococcus suis, an important emerging zoonotic agent of meningitis. Infect Immun 78: 5074-5085.

Enright, M. R., Alexander, T. J., and Clifton-Hadley, F. A. (1987) Role of houseflies (Musca domestica) in the epidemiology of Streptococcus suis type 2. Vet Rec 121: 132-133.

Fabisiak, M., Kita, J., Jedryczko, R., and Binek, M. (2005) Prevalence of the suilysin gene in Streptococcus suis strains isolated from diseased and healthy carrier pigs. Pol J Vet Sci 8: 141-145.

Feldman, C., Mitchell, T. J., Andrew, P. W., Boulnois, G. J., Read, R. C., Todd, H. C., Cole, P. J., and Wilson, R. (1990) The effect of Streptococcus pneumoniae pneumolysin on human respiratory epithelium in vitro. Microb Pathog 9: 275-284.

Fittipaldi, N., Fuller, T. E., Teel, J. F., Wilson, T. L., Wolfram, T. J., Lowery, D. E., and Gottschalk, M. (2009) Serotype distribution and production of muramidase-released protein, extracellular factor and suilysin by field strains of Streptococcus suis isolated in the United States. Vet Microbiol 139: 310-317.

Fittipaldi, N., Gottschalk, M., Vanier, G., Daigle, F., and Harel, J. (2007) Use of selective capture of transcribed sequences to identify genes preferentially expressed by Streptococcus suis upon interaction with porcine brain microvascular endothelial cells. Appl Environ Microbiol 73: 4359- 4364.

Flanagan, J. J., Tweten, R. K., Johnson, A. E., and Heuck, A. P. (2009) Cholesterol exposure at the membrane surface is necessary and sufficient to trigger perfringolysin O binding. Biochemistry 48: 3977-3987.

Fongcom, A., Pruksakorn, S., Mongkol, R., Tharavichitkul, P., and Yoonim, N. (2001) Streptococcus suis infection in northern Thailand. J Med Assoc Thai 84: 1502-1508.

Gelber, S. E., Aguilar, J. L., Lewis, K. L., and Ratner, A. J. (2008) Functional and phylogenetic characterization of Vaginolysin, the human-specific cytolysin from Gardnerella vaginalis. J Bacteriol 190: 3896-3903.

Giddings, K. S., Johnson, A. E., and Tweten, R. K. (2003) Redefining cholesterol's role in the mechanism of the cholesterol-dependent cytolysins. Proc Natl Acad Sci U S A 100: 11315- 11320.

Giddings, K. S., Zhao, J., Sims, P. J., and Tweten, R. K. (2004) Human CD59 is a receptor for the cholesterol-dependent cytolysin intermedilysin. Nat Struct Mol Biol 11: 1173-1178.

Gottschalk, M., Lebrun, A., Wisselink, H., Dubreuil, J. D., Smith, H., and Vecht, U. (1998) Production of virulence-related proteins by Canadian strains of Streptococcus suis capsular type 2. Canadian Journal of Veterinary Research-Revue Canadienne de Recherche Veterinaire 62: 75-79.

Gottschalk, M., Petitbois, S., Higgins, R., and Jacques, M. (1991) Adherence of Streptococcus suis capsular type 2 to porcine lung sections. Can J Vet Res 55: 302-304.

General introduction Chapter 1

32

Gottschalk, M. and Segura, M. (2000) The pathogenesis of the meningitis caused by Streptococcus suis: the unresolved questions. Veterinary Microbiology 76: 259-272.

Gottschalk, M., Segura, M., and Xu, J. (2007) Streptococcus suis infections in humans: the Chinese experience and the situation in North America. Anim Health Res Rev 8: 29-45.

Gottschalk, M., Xu, J., Calzas, C., and Segura, M. (2010) Streptococcus suis: a new emerging or an old neglected zoonotic pathogen? 25. Future Microbiol 5: 371-391. Gottschalk, M. G., Lacouture, S., and Dubreuil, J. D. (1995) Characterization of Streptococcus suis capsular type 2 haemolysin. Microbiology 141 ( Pt 1): 189-195.

Herbert, D. and Todd, E. W. (1941) Purification and properties of a haemolysin produced by group A haemolytic streptococci (streptolysin O). Biochem J 35: 1124-1139.

Heuck, A. P., Hotze, E. M., Tweten, R. K., and Johnson, A. E. (2000) Mechanism of membrane insertion of a multimeric beta-barrel protein: perfringolysin O creates a pore using ordered and coupled conformational changes. Mol Cell 6: 1233-1242.

Higgins, R. and Gottschalk, M. (1990) An update on Streptococcus suis identification. J Vet Diagn Invest 2: 249-252.

Hoa, N. T., Chieu, T. T., Nghia, H. D., Mai, N. T., Anh, P. H., Wolbers, M., Baker, S., Campbell, J. I., Chau, N. V., Hien, T. T., Farrar, J., and Schultsz, C. (2011) The antimicrobial resistance patterns and associated determinants in Streptococcus suis isolated from humans in southern Vietnam, 1997-2008. BMC Infect Dis 11: 6.

Hotze, E. M. and Tweten, R. K. (31-7-2011a) Membrane assembly of the cholesterol-dependent cytolysin pore complex. Biochim Biophys Acta.

Jacobs, A. A. C., Loeffen, P. L. W., vandenBerg, A. J. G., and Storm, P. K. (1994) Identification, Purification, and Characterization of A Thiol-Activated Hemolysin (Suilysin) of Streptococcus Suis. Infection and Immunity 62: 1742-1748.

Jacobs, A. A. C., vandenBerg, A. J. G., and Loeffen, P. L. W. (1996) Protection of experimentally infected pigs by suilysin, the thiol-activated haemolysin of Streptococcus suis. Veterinary Record 139: 225-228.

Johnson, M. K., Geoffroy, C., and Alouf, J. E. (1980) Binding of cholesterol by sulfhydryl-activated cytolysins. Infect Immun 27: 97-101.

Kim, D., Han, K., Oh, Y., Kim, C. H., Kang, I., Lee, J., Gottschalk, M., and Chae, C. (2010) Distribution of capsular serotypes and virulence markers of Streptococcus suis isolated from pigs with polyserositis in Korea. Can J Vet Res 74: 314-316.

King, S. J., Heath, P. J., Luque, I., Tarradas, C., Dowson, C. G., and Whatmore, A. M. (2001) Distribution and genetic diversity of suilysin in Streptococcus suis isolated from different diseases of pigs and characterization of the genetic basis of suilysin absence. Infection and Immunity 69: 7572-7582.

Kock, C., Beineke, A., Seitz, M., Ganter, M., Waldmann, K. H., Valentin-Weigand, P., and Baums, C. G. (2009) Intranasal immunization with a live Streptococcus suis isogenic ofs mutant elicited suilysin-neutralization titers but failed to induce opsonizing antibodies and protection. Veterinary Immunology and Immunopathology 132: 135-145.

General introduction Chapter 1

33

Krawczyk-Balska, A. and Bielecki, J. (2005) Listeria monocytogenes listeriolysin O and phosphatidylinositol-specific phospholipase C affect adherence to epithelial cells. Can J Microbiol 51: 745-751.

Lalonde, M., Segura, M., Lacouture, S., and Gottschalk, M. (2000) Interactions between Streptococcus suis serotype 2 and different epithelial cell lines. Microbiology 146 ( Pt 8): 1913- 1921.

Lecours, M. P., Gottschalk, M., Houde, M., Lemire, P., Fittipaldi, N., and Segura, M. (2011) Critical Role for Streptococcus suis Cell Wall Modifications and Suilysin in Resistance to Complement-Dependent Killing by Dendritic Cells. J Infect Dis 204: 919-929.

Lowe, B. A., Marsh, T. L., Isaacs-Cosgrove, N., Kirkwood, R. N., Kiupel, M., and Mulks, M. H. (2011) Microbial communities in the tonsils of healthy pigs. Vet Microbiol 147: 346-357.

Lun, S. C., Perez-Casal, J., Connor, W., and Willson, P. J. (2003) Role of suilysin in pathogenesis of Streptococcus suis capsular serotype 2. Microbial Pathogenesis 34: 27-37.

Mai, N. T., Hoa, N. T., Nga, T. V., Linh, le D., Chau, T. T., Sinh, D. X., Phu, N. H., Chuong, L. V., Diep, T. S., Campbell, J., Nghia, H. D., Minh, T. N., Chau, N. V., de Jong, M. D., Chinh, N. T., Hien, T. T., Farrar, J., and Schultsz, C. (2008) Streptococcus suis meningitis in adults in Vietnam. Clin Infect Dis 46: 659-667.

Malley, R., Henneke, P., Morse, S. C., Cieslewicz, M. J., Lipsitch, M., Thompson, C. M., Kurt-Jones, E., Paton, J. C., Wessels, M. R., and Golenbock, D. T. (2003) Recognition of pneumolysin by Toll-like receptor 4 confers resistance to pneumococcal infection. Proc Natl Acad Sci U S A 100: 1966-1971.

Messier, S., Lacouture, S., and Gottschalk, M. (2008) Distribution of Streptococcus suis capsular types from 2001 to 2007. Can Vet J 49: 461-462.

Michel, E., Reich, K. A., Favier, R., Berche, P., and Cossart, P. (1990) Attenuated mutants of the intracellular bacterium Listeria monocytogenes obtained by single amino acid substitutions in listeriolysin O. Mol Microbiol 4: 2167-2178.

Navacharoen, N., Chantharochavong, V., Hanprasertpong, C., Kangsanarak, J., and Lekagul, S. (2009) Hearing and vestibular loss in Streptococcus suis infection from swine and traditional raw pork exposure in northern Thailand. J Laryngol Otol 123: 857-862.

Norton, P. M., Rolph, C., Ward, P. N., Bentley, R. W., and Leigh, J. A. (1999) Epithelial invasion and cell lysis by virulent strains of Streptococcus suis is enhanced by the presence of suilysin. Fems Immunology and Medical Microbiology 26: 25-35.

O'Sullivan, T., Friendship, R., Blackwell, T., Pearl, D., McEwen, B., Carman, S., Slavic, D., and Dewey, C. (2011) Microbiological identification and analysis of swine tonsils collected from carcasses at slaughter. Can J Vet Res 75: 106-111.

Okamoto, S., Kawabata, S., Terao, Y., Fujitaka, H., Okuno, Y., and Hamada, S. (2004) The Streptococcus pyogenes capsule is required for adhesion of bacteria to virus-infected alveolar epithelial cells and lethal bacterial-viral superinfection. Infect Immun 72: 6068-6075.

Okwumabua, O., Abdelmagid, O., and Chengappa, M. M. (1999) Hybridization analysis of the gene encoding a hemolysin (suilysin) of Streptococcus suis type 2: evidence for the absence of the gene in some isolates. FEMS Microbiol Lett 181: 113-121.

General introduction Chapter 1

34

Orihuela, C. J., Gao, G., Francis, K. P., Yu, J., and Tuomanen, E. I. (2004) Tissue-specific contributions of pneumococcal virulence factors to pathogenesis. J Infect Dis 190: 1661-1669.

Padungtod, P., Tharavichitkul, P., Junya, S., Chaisowong, W., Kadohira, M., Makino, S., and Sthitmatee, N. (2010) Incidence and presence of virulence factors of Streptococcus suis infection in slaughtered pigs from Chiang Mai, Thailand. Southeast Asian J Trop Med Public Health 41: 1454-1461.

Pan, X. Z., Ge, J. C., Li, M., Wu, B., Wang, C. J., Wang, J., Feng, Y. J., Yin, Z. M., Zheng, F., Cheng, G., Sun, W., Ji, H. F., Hu, D., Shi, P. J., Feng, X. D., Hao, X. N., Dong, R. P., Hu, F. Q., and Tang, J. Q. (2009) The Orphan Response Regulator CovR: a Globally Negative Modulator of Virulence in Streptococcus suis Serotype 2. Journal of Bacteriology 191: 2601-2612. Perch, B., Pedersen, K. B., and Henrichsen, J. (1983) Serology of capsulated streptococci pathogenic for pigs: six new serotypes of Streptococcus suis. J Clin Microbiol 17: 993-996.

Petersen, R., Hannerz, H., Tuchsen, F., and Egerton, J. R. (2011) Meningitis, sepsis and endocarditis among workers occupationally exposed to pigs. Occup Med (Lond).

Pinkney, M., Beachey, E., and Kehoe, M. (1989) The thiol-activated toxin streptolysin O does not require a thiol group for cytolytic activity. Infect Immun 57: 2553-2558.

Princivalli, M. S., Palmieri, C., Magi, G., Vignaroli, C., Manzin, A., Camporese, A., Barocci, S., Magistrali, C., and Facinelli, B. (2009) Genetic diversity of Streptococcus suis clinical isolates from pigs and humans in Italy (2003-2007). Euro Surveill 14.

Ramachandran, R., Tweten, R. K., and Johnson, A. E. (2004) Membrane-dependent conformational changes initiate cholesterol-dependent cytolysin oligomerization and intersubunit beta-strand alignment. Nat Struct Mol Biol 11: 697-705.

Rampersaud, R., Planet, P. J., Randis, T. M., Kulkarni, R., Aguilar, J. L., Lehrer, R. I., and Ratner, A. J. (2011) Inerolysin, a cholesterol-dependent cytolysin produced by Lactobacillus iners. J Bacteriol 193: 1034-1041.

Ratner, A. J., Hippe, K. R., Aguilar, J. L., Bender, M. H., Nelson, A. L., and Weiser, J. N. (2006) Epithelial cells are sensitive detectors of bacterial pore-forming toxins. J Biol Chem 281: 12994-12998.

Robertson, I. D., Blackmore, D. K., Hampson, D. J., and Fu, Z. F. (1991) A longitudinal study of natural Infection of piglets with Streptococcus suis types 1 and 2. Epidemiol Infect 107: 119- 126.

Rossjohn, J., Feil, S. C., McKinstry, W. J., Tweten, R. K., and Parker, M. W. (1997) Structure of a cholesterol-binding, thiol-activated cytolysin and a model of its membrane form. Cell 89: 685- 692.

Rossjohn, J., Gilbert, R. J., Crane, D., Morgan, P. J., Mitchell, T. J., Rowe, A. J., Andrew, P. W., Paton, J. C., Tweten, R. K., and Parker, M. W. (1998) The molecular mechanism of pneumolysin, a virulence factor from Streptococcus pneumoniae. J Mol Biol 284: 449-461.

Rossjohn, J., Polekhina, G., Feil, S. C., Morton, C. J., Tweten, R. K., and Parker, M. W. (2007) Structures of perfringolysin O suggest a pathway for activation of cholesterol-dependent cytolysins. J Mol Biol 367: 1227-1236.

Rubins, J. B., Paddock, A. H., Charboneau, D., Berry, A. M., Paton, J. C., and Janoff, E. N. (1998) Pneumolysin in pneumococcal adherence and colonization. Microb Pathog 25: 337-342.

General introduction Chapter 1

35

Sanford, S. E. and Tilker, M. E. (1982) Streptococcus suis type II-associated diseases in swine: observations of a one-year study. J Am Vet Med Assoc 181: 673-676.

Saunders, F. K., Mitchell, T. J., Walker, J. A., Andrew, P. W., and Boulnois, G. J. (1989) Pneumolysin, the thiol-activated toxin of Streptococcus pneumoniae, does not require a thiol group for in vitro activity. Infect Immun 57: 2547-2552.

Segers, R. P., Kenter, T., de Haan, L. A., and Jacobs, A. A. (1998) Characterisation of the gene encoding suilysin from Streptococcus suis and expression in field strains. FEMS Microbiol Lett 167: 255-261.

Segura, M. and Gottschalk, M. (2002) Streptococcus suis interactions with the murine macrophage cell line J774: adhesion and cytotoxicity. Infect Immun 70: 4312-4322.

Segura, M., Stankova, J., and Gottschalk, M. (1999) Heat-killed Streptococcus suis capsular type 2 strains stimulate tumor necrosis factor alpha and interleukin-6 production by murine macrophages. Infect Immun 67: 4646-4654.

Segura, M., Vadeboncoeur, N., and Gottschalk, M. (2002) CD14-dependent and -independent cytokine and chemokine production by human THP-1 monocytes stimulated by Streptococcus suis capsular type 2. Clin Exp Immunol 127: 243-254.

Segura, M., Vanier, G., Al Numani, D., Lacouture, S., Olivier, M., and Gottschalk, M. (2006) Proinflammatory cytokine and chemokine modulation by Streptococcus suis in a whole-blood culture system. FEMS Immunol Med Microbiol 47: 92-106.

Silva, L. M., Baums, C. G., Rehm, T., Wisselink, H. J., Goethe, R., and Valentin-Weigand, P. (2006) Virulence-associated gene profiling of Streptococcus suis isolates by PCR. Vet Microbiol 115: 117-127.

Soltani, C. E., Hotze, E. M., Johnson, A. E., and Tweten, R. K. (2007) Specific protein-membrane contacts are required for prepore and pore assembly by a cholesterol-dependent cytolysin. J Biol Chem 282: 15709-15716.

Soltani, C. E., Hotze, E. M., Johnson, A. E., and Tweten, R. K. (2007b) Structural elements of the cholesterol-dependent cytolysins that are responsible for their cholesterol-sensitive membrane interactions. Proc Natl Acad Sci U S A 104: 20226-20231.

Sriskandan, S. and Slater, J. D. (2006) Invasive disease and toxic shock due to zoonotic Streptococcus suis: an emerging infection in the East? PLoS Med 3: e187.

Staats, J. J., Feder, I., Okwumabua, O., and Chengappa, M. M. (1997) Streptococcus suis: past and present. Vet Res Commun 21: 381-407.

Staats, J. J., Plattner, B. L., Stewart, G. C., and Chengappa, M. M. (1999) Presence of the Streptococcus suis suilysin gene and expression of MRP and EF correlates with high virulence in Streptococcus suis type 2 isolates. Veterinary Microbiology 70: 201-211.

Stachowiak, R., Wisniewski, J., Osinska, O., and Bielecki, J. (2009) Contribution of cysteine residue to the properties of Listeria monocytogenes listeriolysin O. Can J Microbiol 55: 1153-1159.

Strangmann, E., Froleke, H., and Kohse, K. P. (2002) Septic shock caused by Streptococcus suis: case report and investigation of a risk group. Int J Hyg Environ Health 205: 385-392.

General introduction Chapter 1

36

Sukeno, A., Nagamune, H., Whiley, R. A., Jafar, S. I., Aduse-Opoku, J., Ohkura, K., Maeda, T., Hirota, K., Miyake, Y., and Kourai, H. (2005) Intermedilysin is essential for the invasion of hepatoma HepG2 cells by Streptococcus intermedius. Microbiol Immunol 49: 681-694.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2002) Evidence for lateral transfer of the Suilysin gene region of Streptococcus suis. J Bacteriol 184: 2050-2057.

Tang, J., Wang, C., Feng, Y., Yang, W., Song, H., Chen, Z., Yu, H., Pan, X., Zhou, X., Wang, H., Wu, B., Wang, H., Zhao, H., Lin, Y., Yue, J., Wu, Z., He, X., Gao, F., Khan, A. H., Wang, J., Zhao, G. P., Wang, Y., Wang, X., Chen, Z., and Gao, G. F. (2006) Streptococcal toxic shock syndrome caused by Streptococcus suis serotype 2. PLoS Med 3: e151.

Tarradas, C., Borge, C., Arenas, A., Maldonado, A., Astorga, R., Miranda, A., and Luque, I. (2001) Suilysin production by Streptococcus suis strains isolated from diseased and healthy carrier pigs in Spain. Vet Rec 148: 183-184.

Tenenbaum, T., Adam, R., Eggelnpohler, I., Matalon, D., Seibt, A., GE, K. Novotny, Galla, H. J., and Schroten, H. (2005) Strain-dependent disruption of blood-cerebrospinal fluid barrier by Streptoccocus suis in vitro. FEMS Immunol Med Microbiol 44: 25-34.

Tenenbaum, T., Essmann, F., Adam, R., Seibt, A., Janicke, R. U., Novotny, G. E., Galla, H. J., and Schroten, H. (2006) Cell death, caspase activation, and HMGB1 release of porcine choroid plexus epithelial cells during Streptococcus suis infection in vitro. Brain Res 1100: 1-12.

Tenenbaum, T., Papandreou, T., Gellrich, D., Friedrichs, U., Seibt, A., Adam, R., Wewer, C., Galla, H. J., Schwerk, C., and Schroten, H. (2008) Polar bacterial invasion and translocation of Streptococcus suis across the blood-cerebrospinal fluid barrier in vitro. Cell Microbiol.

Tian, Y., Aarestrup, F. M., and Lu, C. P. (2004) Characterization of Streptococcus suis serotype 7 isolates from diseased pigs in Denmark. Vet Microbiol 103: 55-62.

Tomoyasu, T., Tabata, A., Hiroshima, R., Imaki, H., Masuda, S., Whiley, R. A., Aduse-Opoku, J., Kikuchi, K., Hiramatsu, K., and Nagamune, H. (2010) Role of catabolite control protein A in the regulation of intermedilysin production by Streptococcus intermedius. Infect Immun 78: 4012- 4021.

Trottier, S., Higgins, R., Brochu, G., and Gottschalk, M. (1991) A Case of Human Endocarditis Due to Streptococcus Suis in North-America. Reviews of Infectious Diseases 13: 1251-1252.

Tsuchiya, K., Kawamura, I., Takahashi, A., Nomura, T., Kohda, C., and Mitsuyama, M. (2005) Listeriolysin O-induced membrane permeation mediates persistent interleukin-6 production in Caco-2 cells during Listeria monocytogenes infection in vitro. Infect Immun 73: 3869-3877.

Tweten, R. K. (2005) Cholesterol-dependent cytolysins, a family of versatile pore-forming toxins. Infection and Immunity 73: 6199-6209.

Vadeboncoeur, C. and Pelletier, M. (1997) The phosphoenolpyruvate:sugar phosphotransferase system of oral streptococci and its role in the control of sugar metabolism. Fems Microbiology Reviews 19: 187-207. Vadeboncoeur, N., Segura, M., Al Numani, D., Vanier, G., and Gottschalk, M. (2003) Pro-inflammatory cytokine and chemokine release by human brain microvascular endothelial cells stimulated by Streptococcus suis serotype 2. FEMS Immunol Med Microbiol 35: 49-58.

Vanier, G., Segura, M., Friedl, P., Lacouture, S., and Gottschalk, M. (2004) Invasion of porcine brain microvascular endothelial cells by Streptococcus suis serotype 2. Infection and Immunity 72: 1441-1449.

General introduction Chapter 1

37

Vanier, G., Segura, M., and Gottschalk, M. (2007) Characterization of the invasion of porcine endothelial cells by Streptococcus suis serotype 2. Can J Vet Res 71: 81-89.

Vanier, G., Segura, M., Lecours, M. P., Grenier, D., and Gottschalk, M. (2008) Porcine brain microvascular endothelial cell-derived interleukin-8 is first induced and then degraded by Streptococcus suis. Microb Pathog.

Vecht, U., Wisselink, H. J., Jellema, M. L., and Smith, H. E. (1991) Identification of 2 Proteins Associated with Virulence of Streptococcus Suis Type-2. Infection and Immunity 59: 3156- 3162.

Walker, J. A., Allen, R. L., Falmagne, P., Johnson, M. K., and Boulnois, G. J. (1987) Molecular cloning, characterization, and complete nucleotide sequence of the gene for pneumolysin, the sulfhydryl-activated toxin of Streptococcus pneumoniae. Infect Immun 55: 1184-1189.

Wangkaew, S., Chaiwarith, R., Tharavichitkul, P., and Supparatpinyo, K. (2006) Streptococcus suis infection: a series of 41 cases from Chiang Mai University Hospital. J Infect 52: 455-460.

Watson, K. C., Rose, T. P., and Kerr, E. J. (1972) Some factors influencing the effect of cholesterol on streptolysin O activity. J Clin Pathol 25: 885-891.

Wei, Z., Li, R., Zhang, A., He, H., Hua, Y., Xia, J., Cai, X., Chen, H., and Jin, M. (2009) Characterization of Streptococcus suis isolates from the diseased pigs in China between 2003 and 2007. Vet Microbiol 137: 196-201.

Weis, S. and Palmer, M. (2001) Streptolysin O: the C-terminal, tryptophan-rich domain carries functional sites for both membrane binding and self-interaction but not for stable oligomerization. Biochim Biophys Acta 1510: 292-299.

Wertheim, H. F., Nghia, H. D., Taylor, W., and Schultsz, C. (2009) Streptococcus suis: an emerging human pathogen. Clin Infect Dis 48: 617-625.

Willenborg, J., Fulde, M., de Greeff, A., Rohde, M., Smith, H. E., Valentin-Weigand, P., and Goethe, R. (2011) Role of glucose and CcpA in capsule expression and virulence of Streptococcus suis. Microbiology 157: 1823-1833.

Williams, A. E. and Blakemore, W. F. (1990) Pathogenesis of meningitis caused by Streptococcus suis type 2. J Infect Dis 162: 474-481.

Windsor, R. S. and Elliott, S. D. (1975) Streptococcal infection in young pigs. IV. An outbreak of streptococcal meningitis in weaned pigs. J Hyg (Lond) 75: 69-78.

Wisselink, H. J., Smith, H. E., Stockhofe-Zurwieden, N., Peperkamp, K., and Vecht, U. (2000) Distribution of capsular types and production of muramidase-released protein (MRP) and extracellular factor (EF) of Streptococcus suis strains isolated from diseased pigs in seven European countries. Veterinary Microbiology 74: 237-248.

Wu, T., Chang, H., Tan, C., Bei, W., and Chen, H. (2009) The orphan response regulator RevSC21 controls the attachment of Streptococcus suis serotype-2 to human laryngeal epithelial cells and the expression of virulence genes. FEMS Microbiol Lett 292: 170-181.

Wu, T., Zhao, Z., Zhang, L., Ma, H., Lu, K., Ren, W., Liu, Z., Chang, H., Bei, W., Qiu, Y., and Chen, H. (2011) Trigger factor of Streptococcus suis is involved in stress tolerance and virulence. Microb Pathog 51: 69-76.

General introduction Chapter 1

38

Xiong, Y., Liu, Q., Qin, F. Y., Bai, Y., Zhu, W., Li, H. M., Guo, J. G., Qin, L., Pan, J., Long, J. M., and Chen, L. (2007) [Study on the molecular epidemiology of Streptococcus suis type 2 from healthy pigs in Guangxi]. Zhonghua Liu Xing Bing Xue Za Zhi 28: 593-596.

Xu, L., Huang, B., Du, H., Zhang, X. C., Xu, J., Li, X., and Rao, Z. (2010) Crystal structure of cytotoxin protein suilysin from Streptococcus suis. Protein Cell 1: 96-105.

Ye, C., Zheng, H., Zhang, J., Jing, H., Wang, L., Xiong, Y., Wang, W., Zhou, Z., Sun, Q., Luo, X., Du, H., Gottschalk, M., and Xu, J. (2009) Clinical, Experimental, and Genomic Differences between Intermediately Pathogenic, Highly Pathogenic, and Epidemic Streptococcus suis. J Infect Dis 199: 97-107.

Ye, C., Zhu, X., Jing, H., Du, H., Segura, M., Zheng, H., Kan, B., Wang, L., Bai, X., Zhou, Y., Cui, Z., Zhang, S., Jin, D., Sun, N., Luo, X., Zhang, J., Gong, Z., Wang, X., Wang, L., Sun, H., Li, Z., Sun, Q., Liu, H., Dong, B., Ke, C., Yuan, H., Wang, H., Tian, K., Wang, Y., Gottschalk, M., and Xu, J. (2006) Streptococcus suis sequence type 7 outbreak, Sichuan, China. Emerg Infect Dis 12: 1203-1208.

Yu, H., Jing, H., Chen, Z., Zheng, H., Zhu, X., Wang, H., Wang, S., Liu, L., Zu, R., Luo, L., Xiang, N., Liu, H., Liu, X., Shu, Y., Lee, S. S., Chuang, S. K., Wang, Y., Xu, J., and Yang, W. (2006) Human Streptococcus suis outbreak, Sichuan, China. Emerg Infect Dis 12: 914-920.

Zhao, Y., Liu, G., Li, S., Wang, M., Song, J., Wang, J., Tang, J., Li, M., and Hu, F. (2011) Role of a type IV-like secretion system of Streptococcus suis 2 in the development of streptococcal toxic shock syndrome. J Infect Dis 204: 274-281.

Zheng, H., Punaro, M. C., Segura, M., Lachance, C., Rivest, S., Xu, J., Houde, M., and Gottschalk, M. (2011) Toll-like receptor 2 is partially involved in the activation of murine astrocytes by Streptococcus suis, an important zoonotic agent of meningitis. J Neuroimmunol 234: 71-83.

Chapter 2

Material and methods

Material and methods Chapter 2

41

If not stated otherwise all materials were purchased from Sigma (Muenchen,

Germany).

1. Bacterial strains and growth conditions In this study Streptococcus (S.) suis serotype 2 wild type strain 10, kindly provided by

H. Smith (Lelystad, NL), and its isogenic mutants were used. This strain expresses

EF, MRP, SLY, FBPS and OFS. It has been used by different groups successfully for

mutagenesis and experimental intranasal infections of pigs (Baums et al., 2006;

Smith et al., 1999; Vecht et al., 1997). The unencapsulated isogenic mutant strain

10cpsΔEF of the virulent Streptococcus (S.) suis serotype 2 wild type strain 10

(designated strain 10cpsΔEF) was kindly provided by H. Smith (Lelystad, NL). Strain

10cpsΔEF has been generated by insertional mutagenesis of the genes cps2E and

cps2F involved in biosynthesis of the capsule and has been demonstrated to be

severely attenuated in virulence, most likely due to increased opsonophagocytosis

(Smith et al., 1999). The corresponding suilysin deficient mutants of wild type strain

10 (designed 10Δsly) and 10cpsΔEF (designed 10cpsΔEFΔsly) were constructed by

insertion of an erythromycin cassette in the sly-gene of the unencapsulated strain

10cpsΔEF using the plasmid pBlue/sly/erm as previously described for generation of

strain 10Δsly (Benga et al., 2008). Mutagenesis was confirmed by PCR and Southern

Blot analysis. Streptococci were grown on Columbia agar supplemented with 7%

sheep blood (Oxoid, Wesel, Germany) or in Bacto Todd Hewitt broth (THB) overnight

under aerobic conditions at 37°C. If required, antibiotics were added to the media at

the following concentrations: spectinomycin 100 µg/ml for S. suis; erythromycin 1

µg/ml for S. suis.

Lactococcus lactis subspecies cremoris MG1363 (L. lactis), L. lactis heterologously

expressing SLY (L. lactis pORI-SLY, L. lactis pORI-W461F, L. lactis pORI-SVD) and

L. lactis carrying the shuttle vector pORI23 (L. lactis pORI23) were kindly provided by

D. Reinscheid (Fachhochschule Bonn-Rhein-Sieg, Germany). All L. lactis strains

were cultured on M17 agar plates (Oxoid) supplemented with 5% glucose (GM17)

overnight under aerobic conditions at 30°C. If required, 5 μg/ml of erythromycin was

added to the GM17 medium or plates.

Material and methods Chapter 2

42

Escherichia (E.) coli strains BL21, BL21 (DE3) and DH5α were used for molecular

cloning and protein expression experiments. E. coli were grown in Luria Bertani (LB)

medium overnight under aerobic conditions at 37°C with vigorous shaking. In

appropriate cases, 100 μg/ml of ampicillin was added to the LB medium or plates.

For infection of HEp-2 cells, streptococci were grown in Todd-Hewitt broth (THB,

Difco, Detroit, USA) overnight at 37°C under aerobic conditions, adjusted to an

optical density (OD600) of 0.02 in pre-warmed media the next day and grown to late

exponential growth phase (OD600 0.8).

2. Molecular biology and protein biochemical methods Routine molecular biology techniques including restriction endonuclease digestion,

DNA ligations, agarose gel electrophoresis, transformation of E. coli and plasmid

isolation were performed according to standard procedures (Sambrook et al., 1989).

Restriction enzymes were purchased from New England Biolabs (Frankfurt am Main,

Germany). Plasmid preparations were performed with kits from Machery-Nagel

(Dueren, Germany).

2.1. Construction of mutated suilysin W461F und SVD W461F and RGD-SVD substitutions were constructed using site-directed-

mutagenesis according to the instruction manual of QuikChange® Site-Directed

Mutagenesis Kit (Stratagene, Ja Jolla, CA). As template DNA pET45bslynew was

used (Kock et al., 2009). Oligonucleotide primers encoding the W461 to F461

substitution, namely slyTrp-Phefornew (TACAGGATTAGCATTTGAGTGGTGGAG

AAC) and slyTrp-Pherevnew (GTTCTCCACCACTCAAATGCTAATCCTGTAC), and

oligonucleotide primers encoding R124-G125-D126 to S124-V125-D126, namely

slyRGDMutfor (CAGTATTGCGTCGGTAGATCTGACGCTTAG) and slyRGDMutrev

(CTAAGCGTCAGATCTACCGACGCAATACTG) were used. After digestion of the

methylated non-mutated parental DNA template with DpnI, mutated plasmid

pET45bslynewW461F was electroporated (2.5 kV, 200 Ω, 25 mF) into E. coli BL21

(DE3) and then purified. Mutation sites of each suilysin derivate were sequenced

Material and methods Chapter 2

43

using the primers slyseqfor (GGATCATTCAGGTGCTTATG) and pET45seqrev

(TGCTGGCGTTCAAATTTCGC).

2.2. Expression of recombinant proteins Recombinant His-tagged suilysin (rSLY) was expressed in E. coli BL21

pET45bslynew (Kock et al., 2009). In accordance, recombinant point-mutated

suilysin W461F (rW461F) was expressed in E. coli BL21 pET45bslynewW461F and

recombinant RGD-SVD (rSVD) in E. coli BL21 pET45bslynewRGD-SVD. Proteins

were expressed after induction with 0.4 mM isopropyl-β-D-thiogalactoside for 4 h.

His-tagged proteins were purified under native conditions with Protino® Ni-TED 2000

packed columns as recommended by the manufacturer (Macherey-Nagel) with some

modification. rSLY and rW461F were eluted with a 1:10 dilution of the elution buffer

(containing 10 mM imidazole). The eluted proteins were dialysed against 100 mM

TrisHCl [pH 7.0] with 1 mM DTT. Protein concentrations were measured according to

microplate assay protocol (Bio-Rad, Muenchen, Germany) and the haemolytic

activity was tested as described by Takamatsu et al. (2001). Purification of proteins

were verified by immunoblot analysis with anti-Penta His antibodies (Qiagen, Hilden,

Germany) as recommended. Silver staining was used to control the purity grade. The

protein was stored at -20°C.

Suilysin was completely sequenced with standard oligonucleotide primers T7

(TAATACGACTCACTATAGGG) and T7 term (CTAGTTATTGCTCAGCGGT;

Eurofins MWG operon, Ebersberg, Germany).

2.3. Heterologous expression of SLY, W461F and SVD in L. lactis

The sly gene was amplified from chromosomal DNA of S. suis by PCR with

oligonucleotide primers slyPstI (TAGTCTGCAGCTCCTAGCCTCTGTGGCTAA) and

slyBamHIoptRBS (CAGAGGATCCAGGAGAAAACTTATGAGAAAAAG). After

digestion of the PCR product with BamHI and PstI the fragment containing the whole

sly-gene was cloned into the BamHI/PstI-digested shuttle vector pOri23 (Que et al.,

2000). The resulting plasmid was named pOri23-SLY and transferred in L. lactis as

described previously (Holo and Nes, 1989). Transformants were screened by plating

Material and methods Chapter 2

44

on GM17 agar plates containing 5 µg/ml erythromycin. For heterologous expression

of W461F and SVD site-directed mutagenesis was performed as described above

using purified plasmid DNA pOri23sly and the respective oligonucleotide primers

slyTrp-Phefornew and slyTrp-Pherewnew as well as slyRGDMutfor and

slyRGDMutrev. Resulting constructs pOri23-W461F and pOri23-SVD were

electroporated in L. lactis.

2.4. Immunoblot analysis Recombinant proteins or supernatants of infected cells were separated by SDS-

polyacrylamid gel electrophoresis with a 4% stacking and a 10% separating gel

under denaturing conditions and transferred to a PVDF-membrane (Serva,

Heidelberg, Germany). For suilysin immunoblot, membrane-blocking was performed

overnight with 3% milk powder in TBS with 0.5% Tween. Polyclonal antiserum raised

against rSLY (Benga et al., 2008) diluted 1:1300 in 1% milk powder was used to

detect either rSLY, rW461F, rSVD or suilysin and its mutated derivates expressed by

the respective L. lactis strains (L. lactis pORI-SLY, L. lactis pORI-W461F and L. lactis

pORI-SVD) present in culture supernatants. For detection of His-tagged recombinant

proteins membranes were blocked with 3% BSA in TBS with 0.5% Tween overnight

and incubated with a monoclonal anti-Penta His antibody diluted 1:2000 in 3% BSA.

Membranes were developed with horseradish peroxidase-conjugated anti-rabbit or

anti-mouse IgG antiserum diluted 1:10000 in 1% milk powder (Amersham, Freiburg,

Germany) followed by using SuperSignal® West Pico Chemiluminescent Substrate

(Pierce, Rockford, USA) according to the manufactures protocol.

3. Cell culture methods 3.1. Epithelial cells The human laryngeal epithelial cell line HEp-2 (ATCC CCL 23) was used. Cells were

maintained in Dulbecco’s modified Eagle’s medium (DMEM, Gibco-Invitrogen,

Groningen, The Netherlands) supplemented with 10% fetal calf serum (FCS) and 5

mM glutamine at 37°C and 8% CO2. The cells were subcultured every 2-3 days after

detachment with 0.25% trypsin and 1 mM Na-EDTA (trypsin-EDTA, Gibco-

Material and methods Chapter 2

45

Invitrogen). For antibiotic protection assays, approximately 1.8 x 105 cells per well

were seeded on 24 well tissue culture plates, for immunofluorescence and

electronmicroscopy 0.5 x 105 and 1.5 x 105 cells per well were seeded on a 12 mm

diameter glass cover slips placed in 24 well plates. For flow cytometry analysis

approximately 6 x 105 cells per well were seeded on a 6 well plate and for cellular

cytotoxicity assay 0.2 x 105 cells were seeded on a 96 well plate. The cells were

grown over-night and then used for experiments.

3.2. Antibiotic protection assay The number of adherent and invasive streptococci was quantified using a gentamicin

protection assay as described with the following modifications (Valentin-Weigand et

al., 1996). Adhesion and invasion assays were performed with fresh grown

streptococci in THB harvested at an OD600 of 0.8 by centrifugation. Subsequently,

bacteria were resuspended in PBS containing 0.9 mM CaCl2/0.5 mM MgCl26H2O

[pH 7.3] (PBS+; Gibco-Invitrogen) and adjusted to an OD600 0.6. Then, diluted 1:10 in

PBS+ and used for adherence and invasion assay. Confluent HEp-2 cell monolayers

were inoculated with 100 bacteria per cell. The number of colony forming units (CFU)

inoculated per well was determined by serial platings on Columbia agar

supplemented with 7% sheep blood. The cells were incubated for 2 h at 37°C with

8% CO2 in order to allow adherence and invasion of the bacteria. Supernatants of

infected cells were collected 2 hours post infection for immunoblot analysis of

suilysin. To determine adherence rates cells were washed three times with PBS

immediately after supernatant collection, and 100 μl trypsin-EDTA solution was

added to each well. After 5 min, cells were lysed by adding 400 μl 0.025%

Triton X-100. Serial dilutions of these lysates were plated in triplicates on Columbia

agar and incubated at 37°C for 24 h. For invasion assay, cells were washed 3 times

with PBS 2 h post infection and incubated with 500 μl per well of DMEM containing

100 μg gentamicin and 5 μg penicillin per ml for 2 additional hours at 37°C with 8%

CO2 to kill extracellular bacteria. The cells were washed and lysed as described for

adherence and lysates were plated for viable counts of streptococci. Results for

Material and methods Chapter 2

46

adherence and invasion were expressed as percentage of the inoculum used for

infection.

In some adherence and invasion assays, HEp-2 cells were treated before infection

with the following inhibitors: 1 µg/ml Latrunculin B for 30 min, 10 ng/ml toxin B of

Clostridium (C.) difficile (TcdB), 10 ng/ml toxin B of C. difficile serotype F strain 1470

(TcdB-F) and 1 µg/ml C3 toxin of C. limosum for 3 h. C3 toxin was always

reconstituted freshly by mixing the binding domain (C2IIa) with the catalytic domain

(FT) in a ratio of 1:1 as described (Just et al., 1995). After preincubation with

inhibitors cells were washed and adherence and invasion assays were performed as

described above. Latrunculin B was left on to the cells during the whole infection

experiment because its effect is reversible up to 1 h.

3.3. Labelling of bacteria for flow cytometry experiments CellTrace™ CFSE Cell Proliferation Kit (Invitrogen) was used for fluorescent staining

of bacteria. Labelling of bacteria with CFSE was done as described by Logan et al.

(1998) with some modifications. Streptococcal strains 10cpsΔEF and 10cpsΔEFΔsly

were grown to an OD600 of 0.8, pelleted by centrifugation, washed once with PBS

(Gibco-Invitrogen) and resuspended in PBS containing 5 µM carboxyfluorescein

succinimidyl ester (CFSE)-reagent (prepared from a CFSE stock solution as

recommended by the manufacturer). Staining procedures were carried out at 37°C

for 20 min on an end-over-end rotator. After washing (removal of unbound

fluorescent dye) culture stocks were prepared as described above. Labelling

efficiency (>90%) was confirmed by fluorescent microscopy and measurement of

fluorescent intensity by flow cytometry (side-ward-scatter [ssc] versus fluorescence

channel FL-1) in comparison to unlabeled streptococci (data not shown).

3.4. Bacteria-cell association HEp-2 cell monolayers were infected with CFSE-labelled streptococci at a multiplicity

of infection [moi] of 125:1 bacteria per cell, followed by centrifugation (250 x g for

5 min) and incubation at 37°C for 2 h. After incubation supernatants of infected cell

were collected for determination of suilysin expression by immunoblot analysis. Cells

Material and methods Chapter 2

47

were detached with trypsin-EDTA solution, washed with PBS and fixed by

resuspension in PBS with 0.375% formaldehyde in PBS. In complementation

experiments, rSLY and rW461F were added to strain 10cpsΔEFΔsly in a protein

concentration of 100 ng/ml during the experiment. Bacteria-cell association was

measured using FACScan® (Becton Dickinson, 488 nm Argon laser). Further analysis

was performed with the software WinMDI (version 2.9.). For each determination at

least 10,000 events were acquired and initial analysis of infected cells was carried

out by dot plot analysis (forward scatter [fsc] versus sideward scatter [ssc]) to define

the cell population of interest (data not shown). Subsequently, fluorescent cells were

detected at channel FL-1. The numerical results were expressed as mean

fluorescent intensity [mfi] values of cell population. Noninfected cells served as

background control. Means and standard deviations refer to eight independent

experiments. To exclude cytotoxic effects of SLY during the experiment cells were

stained afterwards with propidium iodid (4 µg/ml, DNA-intercalating dye) for 2 min at

room temperature to discriminate between viable and nonviable cells. Red

fluorescent cells (dead cells) were determined by dot plot analysis (ssc versus FL-2).

In some assays, HEp-2 cells were treated before infection with 1µg/ml latrunculin B

for 30 min and latrunculin B was also added to the cells during the whole experiment

because its effect is reversible during 1 h. After preincubation with latrunculin B cells

were washed and flow cytometry was performed as described above.

3.5. Immunofluorescence microscopy Membrane binding of rSLY, rW461F and rSVD was determined by

immunofluorescent microscopy. HEp-2 cells were incubated with the recombinant

suilysin derivates at subcytolytic (100 ng/ml) or lytic (400 ng/ml) concentrations for 30

min at 37°C. Cells were washed three times to remove unbound suilysin, fixed with

2% parafolmaldehyde and permeabilised with 0.1% Triton X-100 in PBS (Bio-Rad,

Muenchen, Germany). Blocking was performed for 30 min at room temperature with

PBS containing 10% FCS. Suilysin was stained using a polyclonal antiserum raised

against rSLY (Benga et al., 2008) diluted 1:1,300 in 1% milk powder for 1 h at room

temperature followed by an incubation with FITC-conjugated goat anti-rabbit antibody

Material and methods Chapter 2

48

(1:1,000 in 1% FCS in PBS, Dianova, Hamburg, Germany). F-actin was labelled with

TRITC-conjugated Phalloidin (20 µg/ml, Invitrogen) for 40 min at room temperature.

After final washing DAPI (Invitrogen) was used for staining the nuclei. Antifading

reagent DABCO (Sigma) was used for sealing of the samples. Mounted samples

were examined using inverted immunofluorescence microscope Nikon Eclipse Ti-S

equipped with a 40×, 0.6 S Plan Fluor objective (Nikon, Duesseldorf, Germany)

driven by NIS Elements software BR 3.2..

3.6. Double immunofluorescence microscopy (DIF) Determination of adherence and invasion of streptococci by DIF was done as

described by Benga et al. (2004) with the following modifications. Semiconfluent

HEp-2 cells were incubated with streptococci at a moi of 100 bacteria per epithelial

cell. ProLong® Gold antifade reagent with DAPI (Invitrogen) was used for sealing and

staining of the nucleus. Mounted samples were examined using inverted

immunofluorescence microscope Nikon Eclipse Ti-S. equipped with a 40×, 0.6 S Plan

Fluor objective (Nikon, Duesseldorf, Germany) driven by NIS Elements software BR

3.2..

3.7. Colocalisation experiments For suilysin localization experiments GFP-Rac1-expressing HEp-2 cells were fixed

after incubation with rSLY (protein concentration of 100 ng/ml) for 30 min at 37°C

with 3% PFA in modified cytoskeleton buffer (10 mM MES [pH 7.0], 150 mM NaCl, 5

mM EGTA, 5 mM MgCl2, 5 mM sucrose) for 20 min and quenched with 10 mM glycin

in PBS and permeabelised with 0.1% Triton X-100 in PBS. Permeabilised cells were

blocked in PBS with 5% horse serum and 1% BSA (blocking buffer) for 30 min at

room temperature and subsequently incubated with polyclonal antiserum raised

against rSLY (Benga et al., 2008) diluted 1:1,200 in blocking buffer for 45 min at

room temperature. Samples were washed twice in PBS and incubated with Cy5-

conjugated goat anti-rabbit IgG (Millipore, Schwalbach/Ts, Germany) for 30 min at

room temperature. F-actin was labelled with Alexa® Fluor 568-conjugated Phalloidin

(Invitrogen) for 20 min at room temperature. Coverslips were washed three times in

Material and methods Chapter 2

49

PBS and then mounted using ProLong® Gold antifade reagent with DAPI (Invitrogen).

Mounted samples were examined using a LSM 510 Meta confocal microscope

equipped with a 63×, 1.2NA Plan-NEOFLUAR oil immersion objective (Zeiss, Jena,

Germany) driven by LSM software v3.2. For 3D z-stack acquisition the pinholes were

set to one Airy unit and confocal planes were acquired every 340 nm. Images were

deconvolved and maximum intensity projection were rendered using Huygens®

Essential (Hilversum, The Netherlands).

3.8. Field emission scanning electron microscopy (FESEM) Association of HEp-2 cell monolayers with either 10cpsΔEF or 10cpsΔEFΔsly at 100

bacteria per cell were investigated with FESEM as described previously (Benga et

al., 2004).

3.9. Cell-permeability and macropore-formation assay Experiments were performed in PBS containing 0.9 mM CaCl2/0.5 mM MgCl26H2O

[pH 7.3] (Gibco-Invitrogen). Cell permeability was determined after co-incubation of

HEp-2 cells (4 x 105 cells/ml) with recombinant proteins (rSLY, rW461F, and rSVD) at

subcytolytic (100 ng/ml) or lytic (400 ng/ml) concentrations and the fluorescent

marker calcein (2 µg/ml). Incubation was performed for 2 h at 37°C on an end-over-

end rotator. After two washing steps with PBS to remove extracellular calcein, cells

were fixed with 0.37% formaldehyde (for 10 min at room temperature and afterwards

resuspended in 500 µl PBS for measurement of fluorescent cells using FACScan®.

Further analysis was performed with the software WinMDI. For each determination at

least 10,000 events were measured. Initial analysis of fluorescent cells was carried

out by dot plot analysis (forward scatter [fsc] versus sideward scatter [ssc]) to define

the cell population of interest (data not shown). Subsequently, fluorescent cells were

detected at channel FL-1. Data were expressed in comparison to cells only treated

with calcein (background control).

Material and methods Chapter 2

50

3.10. Detection of α5β1 integrin expression on HEp-2 cells HEp-2 cell-suspension was incubated with either a monoclonal anti-human CD29

antibody or a monoclonal anti-human CD49e (Bio Legend) both used at a

concentration of 500 ng per 1.5 x 106 cells (Bio Legend, Fell, Germany). All

antibodies were diluted in blocking buffer (PBS [pH 7.3] containing 3% BSA) and

incubation was performed for 30 min at room temperature with gently agitation. After

three washing steps with blocking buffer, cells were incubated with a TRITC-

conjugated goat anti-mouse antibody (1:1,000, Dianova, Hamburg, Germany).

Stained cells were washed three times, fixed with 0.37% formaldehyde and

resuspended in blocking buffer for subsequent measurement. Flow cytometry

measurement was performed as described above. Fluorescent cells were counted at

channel FL-1 and results were expressed in comparison to unstained cells

(background control).

3.11. Haemolysis assay The haemolytic activities of rSLY, rW461F or rSVD were determined by a haemolysis

assay, performed essentially as described by Takamatsu et al. (2001). Titration of

haemolytic activity was performed in a 96-well vee bottom microplate. Twofold

dilutions of the toxins were incubated with a 2% sheep erythrocyte suspension in

0.9% NaCl for 2 h at 37°C mixed gently. Unlysed red blood cells were allowed to

pellet by centrifugation and 100 µl of the supernatant was transferred into a new flat

bottom microplate. Subsequently, absorption was measured at 550 nm in a

microplate reader (GENios Pro, TECAN AUSTRIA GMBH). Determination of

haemolytic activity was done in three independent experiments. One haemolytic unit

(HU) was defined as the concentration of toxin causing 50% haemolysis of a 2%

sheep blood suspension.

3.12. Cytotoxicity assays Cytotoxicity was detected by an LDH-release assay as described previously (Benga

et al., 2004). HEp-2 cells were either incubated with rSLY or rW461F. Supernatants

were removed 2 h after treatment and their LDH activities were determined using the

Material and methods Chapter 2

51

Cytotox® 96 assay kit (Promega, Mannheim, Germany). The experiments were

performed in triplicates and repeated at least three times. To quantify relative cellular

damage, results were expressed relative to LDH release observed after Triton X-100

(Biorad, Muenchen, Germany) lysis of non-infected cells.

To determine haemolytic and cytotoxic activity of heterologous expressed SLY,

W461F or SVD in L. lactis, respective lactococci (L. lactis pORI-SLY, -W461F and -

SVD) were grown in GM17 medium overnight, adjusted to an optical density (OD600)

of 0.02 in pre-warmed media the next day and then grown to late exponential growth

phase (OD600 0.8). Subsequently bacteria were removed by centrifugation and

culture supernatants were concentrated and purified using Amicon centrifHugal filter

devices 30 kDa (Millipore, Schwalbach/Ts., Germany) to remove remaining medium

components. Concentrated culture supernatants diluted 1:5 in 0.9% NaCl were used

as test samples.

3.13. Pull down experiments Experiments were done as described previously with some modifications (Reid et al.,

1996). HEp-2 cells were incubated with rSLY or rW461F (100 ng/ml) for indicated

time periods. Then cells were rinsed twice with ice cold PBS to stop incubation. Lysis

of cells was done by addition of 1 ml of ice cold lysis buffer (50 mM NaCl, 20 mM

Tris-HCl [pH 7.4], 3 mM MgCl2, 1% Nonidet-P 40, 0.25% Triton X-100, 5 mM

dithiothreitol, 100 µM PMSF). Cells were scraped off, and the lysates were

centrifuged at 14,000 rpm. The supernatant was split and used for pull down

experiments of activated RhoA and Rac1 in parallel. For this, 20 µl of beads slurry

bearing approximately 20 µg protein of the GST-fusion protein of either the Rho-

binding domain C21 or the PAK-GBD, respectively, were added to 500 µl sample and

incubated on a rotator at 4°C for 30 min. Beads were collected by centrifugation at

10,000 rpm, washed twice with lysis buffer and subjected to immunoblot analysis

using monoclonal antibodies against Rac-1 (cloe 102; BD Pharmingen, Heidelberg,

Germany) and against RhoA (clone 26C4; Santa Cruz, Heidelberg, Germany),

respectively. Amounts of (activated and total) GTPases were quantified by

densitometric measurement of three independent immunoblot analyses.

Material and methods Chapter 2

52

3.14. GLISA HEp-2 cells were treated with rSLY, rW461F or rSVD at subcytolytic concentration

(100 ng/ml) for 15 min and rinsed twice with ice cold PBS to stop incubation. Lysis of

cells was achieved by addition of 1 ml of ice cold lysis buffer (50 mM NaCl, 20 mM

Tris-HCl, pH 7.4, 3 mM MgCl2, 1% Nonidet-P 40, 0.25% Triton X-100, 5 mM

dithiothreitol, 100 µM PMSF). Cells were scraped off, and the lysates were

centrifuged at 14,000 rpm. The supernatant was used for a pull down-based GLISA

(Cytoskeleton, CO, USA) to determine activation of Rac1, RhoA, and Cdc42

according to the protocol by the supplier.

4. Mouse infection model 4.1. Preparation of infection culture For infection of mice, streptococci were grown in Todd-Hewitt broth (THB, Difco,

Detroit, USA) overnight at 37°C under aerobic conditions, adjusted to an optical

density (OD600) of 0.02 in pre-warmed media the next day and then grown to late

exponential growth phase (OD600 0.8). Streptococci were harvested by centrifugation,

resuspended in sterile PBS (pH 7.4) and adjusted to the final concentration of

5 x 1011 CFU/ml for intranasal infection. Inoculum concentrations were verified by

platting 10-fold serial dilutions on Colombia agar plates with 7% sheep blood after

infection.

4.2. Intranasal infection of mice Four (expt. 1 to 4) or six (expt. 5) week old specific-pathogen-free female mice of the

outbreed strain Crl:CD1 (ICR) were obtained from Charles River Laboratories

(Sulzfeld, Germany). Mice of the Staphylococcus (St.) aureus free outbred strain

Hsd:ICR (CD1®) were purchased from Harlan Laboratories (AN Venray, The

Netherlands). Animals were randomly divided into groups consisting of five to six

animals each. Mice were allowed to acclimate for one week and cared for in

accordance with the principles outlined in the European Convention for the Protection

of Vertebrate Animal Used for Experimental and Other Scientific Purposes [European

Treaty Series, no. 123: http://conventions.coe.int/treaty/EN/Menuprincipal.htm; permit

Material and methods Chapter 2

53

no.33.9-42502-04-08/1589]. Before infection mice were anesthetized via inhalation of

isofluran (IsoFlo®, Albrecht, Germany). In experiment 2 to 5 mice were pre-treated

with 12.5 µl 1% acetic acid [pH 4.0] placed in each nostril 1 h prior intranasal

infection. After a controlled recovery phase and re-anaesthesia per isofluran

inhalation, mice were infected with 5 x 109 CFU of either S. suis wild type strain 10,

10Δsly, 10cpsΔEF or 10cpsΔEFΔsly. The dose was applied in two drops of 12.5 µl

volume placed in front of the nostrils.

4.3. Intravenous infection of mice Five- to seven-week-old specific-pathogen-free female inbreed strain C57/BL6 mice

and outbreed strain Crl:CD1 (ICR) mice were purchased from Charles River

(Sulzfeld, Germany), randomly divided into groups (five to six mice each) and kept in

conformity with the European conventions. After an adjustment period of one week,

mice were infected via the lateral tail vein with ~6 x 105 CFU, 1x108 CFU, 5 x 108

CFU, and 1 x 109 CFU per individual (total volume of 100 µl per ventral tail vein).

4.4. Clinical score Animals were clinically scored every 8 h. The health status was rated using a clinical

score sheet (Table 5-1), including weight development, clinical signs of general

sickness (rough coat, rapid breathing, dehydration), clinical signs indicating

meningitis (apathy, apraxia), and septicaemia (swollen eyes, depression), developed

on the basis of mouse clinical monitoring score described by the Research office of

the Australian National University [http://www.anu.edu.au/ro/ORI/Animal/AEEC001_

MouseMonitoringSOP.doc]. A cumulative score of 3 to 4 indicated mild clinical

symptoms, a score of 5 to 6 moderate clinical symptoms and a score greater than 6

severe clinical symptoms, in particular persistent anorexia, apathy, and/or neural

disorder. Mice with a cumulative score equal or greater than 3 were classified as

diseased (calculation of morbidity). In the case of severe weight loss (> 20%) and/or

enduring severe clinical signs, mice were euthanized for reasons of animal welfare

by inhalation of CO2 and cervical dislocation.

Material and methods Chapter 2

54

4.5. Histological screening Immediately after euthanasia, necropsy was conducted and the organs were

aseptically removed and split for histological and bacteriological screenings, including

spleen, liver, kidney, heart, lung, brain, spinal cord, and nose. The histological

screenings were carried out as blind experiments. Findings were scored as described

for piglets (Baums et al., 2006). In contrast to piglets, in addition to fibrinosuppurative

lesions, purulent necrotizing lesions associated with the challenge strain were scored

as well (see results). Rhinitis was not included in the general score ω designed to

reflect lesions caused by invading streptococci.

4.6. Reisolation of S. suis strains from tissue and tracheo-nasal lavage (TNL) One half of each organ was suspended in 5 ml PBS [pH 7.4] and weighed. All organs

were homogenized with an Ultra Turrax (IKA, Staufen, Germany). Ten-fold serial

dilutions of samples were plated on blood agar plates. Colony forming units (CFU)

were counted the next day after incubation at 37°C for 24 h and CFU per mg organ

was determined.

For sampling TNL the trachea was opened and a retrograde irrigation of the nasal

cavity with 300 µl PBS was collected. Number of typical α-haemolytic streptococci

per µl TNL was determined by serial plating on blood agar. Isolated α-haemolytic

streptococci were investigated in a S. suis multiplex PCR for the detection of mrp,

epf, sly, arcA, gdh, cps1, cps2, cps7, and cps9 (Silva et al., 2006). Isolates received

from mice challenged either with strain 10Δsly, 10cpsΔEF or 10cpsΔsly were

additionally tested in a cps2E-specific PCR (with oligonucleotide primers cps2Efor

(TTTCGCACTTTCAAGACGTG) and cps2Erev (GGACGGGTACCGACTAGACTC)

and sly-specific PCR using primers slyAgefor (TGTACCGGTGATTCCAAACAAGA

TATTAA) and slyAge3new (TTAACCGGTTACTCTATCACCTCATCCG). Based on

CFU per mg organ or per µl TNL, bacterial loads were classified as mild (+; <100),

moderate (++; >100 - <1000) or severe (+++; >1000).

Material and methods Chapter 2

55

5. Statistical analyses If not stated otherwise, experiments were performed at least three times and results

were expressed as means and standard deviations. Data were analyzed by t-test.

Pull down analysis was carried out using an unpaired t-test with no assumption of

Gaussian distribution (Mann-Whitney). A P-value <0.05 was considered significant.

Material and methods Chapter 2

56

References

Baums, C. G., Kaim, U., Fulde, M., Ramachandran, G., Goethe, R., and Valentin-Weigand, P. (2006) Identification of a novel virulence determinant with serum opacification activity in Streptococcus suis. Infect Immun 74: 6154-6162.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cellular Microbiology 6: 867-881.

Holo, H. and Nes, I. F. (1989) High-Frequency Transformation, by Electroporation, of Lactococcus lactis subsp. cremoris Grown with Glycine in Osmotically Stabilized Media. Appl Environ Microbiol 55: 3119-3123.

Just, I., Selzer, J., Wilm, M., Eichel-Streiber, C., Mann, M., and Aktories, K. (1995) Glucosylation of Rho proteins by Clostridium difficile toxin B. Nature 375: 500-503.

Kock, C., Beineke, A., Seitz, M., Ganter, M., Waldmann, K. H., Valentin-Weigand, P., and Baums, C. G. (2009) Intranasal immunization with a live Streptococcus suis isogenic ofs mutant elicited suilysin-neutralization titers but failed to induce opsonizing antibodies and protection. Veterinary Immunology and Immunopathology 132: 135-145.

Logan, R. P., Robins, A., Turner, G. A., Cockayne, A., Borriello, S. P., and Hawkey, C. J. (1998) A novel flow cytometric assay for quantitating adherence of Helicobacter pylori to gastric epithelial cells. J Immunol Methods 213: 19-30.

Que, Y. A., Haefliger, J. A., Francioli, P., and Moreillon, P. (2000) Expression of Staphylococcus aureus clumping factor A in Lactococcus lactis subsp. cremoris using a new shuttle vector. Infect Immun 68: 3516-3522.

Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N Y.

Silva, L. M., Baums, C. G., Rehm, T., Wisselink, H. J., Goethe, R., and Valentin-Weigand, P. (2006) Virulence-associated gene profiling of Streptococcus suis isolates by PCR. Vet Microbiol 115: 117-127.

Smith, H. E., Damman, M., van der Velde, J., Wagenaar, F., Wisselink, H. J., Stockhofe-Zurwieden, N., and Smits, M. A. (1999) Identification and characterization of the cps locus of Streptococcus suis serotype 2: the capsule protects against phagocytosis and is an important virulence factor. Infect Immun 67: 1750-1756.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2001) Thermosensitive suicide vectors for gene replacement in Streptococcus suis. Plasmid 46: 140-148.

Vecht, U., StockhofeZurwieden, N., Tetenburg, B. J., Wisselink, H. J., and Smith, H. E. (1997) Virulence of Streptococcus suis type 2 for mice and pigs appeared host-specific. Veterinary Microbiology 58: 53-60.

Chapter 3

Results, part I:

Subcytolytic suilysin promotes invasion of Streptococcus suis in HEp-2

epithelial cells by Rac1- dependent activation of the actin cytoskeleton

M. Seitz, C. G. Baums, R. Gerhard, I. Just, C. Neis, L. Benga, M. Fulde, A. Nerlich,

M. Rohde, R. Goethe and P. Valentin-Weigand

Running title: Subcytoytic suilysin promotes invasion of S. suis

(Manuscript submitted)

Results, part I Chapter 3

59

Abstract Suilysin is a cholesterol-dependent pore-forming cytolysin secreted by Streptococcus

suis, an important swine and zoonotic pathogen. The role of suilysin in S. suis-host

cell interaction is still unclear. In the present study we found a higher invasion rate in

epithelial cells of an unencapsulated sly-positive strain compared to its sly-negative

mutant. Electron microscopy revealed that uptake-accompanying membrane ruffling

was abolished in the sly-negative mutant. Since invasion of S. suis depended on

actin polymerization, we used specific inhibitors of Rho-GTPases and could show

that Rac, but not Rho was involved in suilysin-mediated uptake. Accordingly, pull

down analysis with specific ligands demonstrated activation of Rac, but not Rho in

suilysin-treated cells. Furthermore, confocal microscopy revealed colocalisation of

suilysin with F-actin and Rac1. To further dissect the relevance of cytolytic pore-

formation we produced recombinant suilysin in which the protein domain responsible

for membrane binding and pore-formation was inactivated via substitution of a

tryptophan by phenylalanine. The mutated suilysin had lost haemolytic and cytolytic

activity but retained its ability to mediate S. suis invasion and activation of Rac.

Concluding, our results indicate that subcytolytic suilysin promotes S. suis invasion in

epithelial cells by Rac-dependent activation of the actin cytoskeleton.

Results, part I Chapter 3

60

Introduction Streptococcus (S.) suis is one of the most important swine pathogens worldwide

causing meningitis, arthritis, septicaemia, bronchopneumonia, and other pathologies.

Furthermore, S. suis colonizes the nasopharynx and other mucosal surfaces,

resulting in high carrier rates of healthy pigs (Arends et al., 1984; O'Sullivan et al.,

2011). S. suis is also an important zoonotic agent (Trottier et al., 1991; Gottschalk et

al., 2007). Meningitis and septicaemia are the most common manifestations in

humans, occurring particularly in people exposed to pigs or pig-products (Arends and

Zanen, 1988). Two human outbreaks in China in 1998 and 2005 were associated

with increased severeness of clinical symptoms, a high rate of mortality, and

streptococcal toxic shock-like syndrome (Tang et al., 2006). Among the 33 serotypes

of S. suis, serotype 2 is worldwide most frequently isolated from diseased pigs

(Wisselink et al., 2000) and humans (Gottschalk et al., 2010). Mechanisms

underlying pathogenesis of S. suis infections, however, are only poorly known.

Suilysin was identified as haemolysin of S. suis several years ago (Jacobs et al.,

1994). It is a member of the cholesterol-dependent pore-forming cytolysins (CDC)

family and expressed by many S. suis strains. The sly gene has been detected in

95% of European and Asian invasive serotype 2 strains (Segers et al., 1998), and it

was found in 69.4% of isolates from 10 different serotypes. Though these isolates

were mainly obtained from porcine cases of meningitis and septicaemia (King et al.,

2001), experimental infections demonstrated attenuation of a sly knock-out mutant

only in mice, but not in piglets (Allen et al., 2001; Lun et al., 2003).

Similar to other members of the CDC family suilysin can damage host cells by its

cytolytic activity (Norton et al., 1999; Charland et al., 2000; Lalonde et al., 2000;

Segura and Gottschalk, 2002; Vanier et al., 2004; Tenenbaum et al., 2005;

Tenenbaum et al., 2006). It has also been suggested that suilysin plays a role in

invasion and pathogenesis of S. suis (Norton et al., 1999). More recently, it has been

found that the toxin may be involved in cytokine release and protection against

opsonophagocytosis (Benga et al., 2008; Lecours et al., 2011). Some of the other

CDC have been shown to display biological effects at sublytic concentrations, e. g.

Results, part I Chapter 3

61

phosphorylation of p38 mitogen-activated protein kinase (MAPK) in epithelial cells,

which is crucial for local production of IL-8 and subsequent recruitment of neutrophils

to the site of infection (Ratner et al., 2006). For the CDC of S. pneumoniae

(pneumolysin), Listeria monocytogenes (listeriolysin O), and S. intermedius

(intermedilysin) it was suggested that they contribute to bacterial adherence and

invasion (Rubins et al., 1998; Sukeno et al., 2005; Krawczyk-Balska and Bielecki,

2005). These findings indicate that CDC express subcytolytic activities that may

modify host cell responses to infection. For suilysin, however, such activities and

their possible biological relevance still await to be elucidated in detail.

The objective of this study was to evaluate the possible role of suilysin in host cell

interaction of S. suis, in particular its impact on association of S. suis with respiratory

epithelial cells, i. e. adherence and invasion. Using respective mutant strains and

HEp-2 epithelial cells we identified a suilysin-dependent invasive phenotype of an

unencapsulated S. suis serotype 2 strain and determined host cell GTPases involved

in suilysin-mediated effects. Furthermore, we showed that these effects did not

require formation of a cytolytic pore.

Results, part I Chapter 3

62

Results Suilysin promotes adherence and invasion of S. suis. The present study was

undertaken to investigate the possible role of the cytolytic toxin suilysin, since its role

in virulence of S. suis is poorly understood. For this, we constructed a suilysin-

negative mutant of the unencapsulated strain 10cpsΔEF (kindly provided by H. E.

Smith, Lelystad, NL). We used this strain because it originated from a highly virulent

serotype 2 strain and displays a much better epithelial cell adherence and invasion

than encapsulated strains, as observed in our previous studies (Benga et al., 2004).

We compared the sly-positive strain with its sly-negative mutant (10cpsΔEFΔsly) in

adherence and invasion using the respiratory epithelial cell line HEp-2. Under our

experimental conditions we could exclude cytotoxic effects of suilysin, as confirmed

by viability staining of HEp-2 cells with propidium iodide (data not shown). First we

determined bacterial association with HEp-2 cells by FACS of CFSE-labelled

streptococci, which had been incubated with epithelial cells for 2 h. The

unencapsulated mutant showed a high association with HEp-2 cells (mean

fluorescence intensity [mfi] of 10.4 as compared to 3.1 of uninfected cells). The

sly-negative mutant of this strain demonstrated significantly lower association (mfi of

7.2, P = 0.0092; Figure 3-1A). Secondly, we performed antibiotic protection assays to

differentiate adherence and invasion of both S. suis strains. Results revealed that the

sly-negative strain adhered to a much lower degree and showed a significantly

reduced invasion as compared to the sly-positive strain (Figure 3-1B and C). Notably,

differences in adherence were not significant due to high standard deviations, but

were significant with respect to invasion. Thus, we assume that differences seen in

FACS analyses (Figure 3-1A) reflect differences in invasion capacity rather than in

adherence. Detection of suilysin by immunoblot analysis confirmed that suilysin was

expressed under these conditions only by strain 10cpsΔEF, but not by its sly-

negative mutant (Figure 3-1C, inlay). Thirdly, we compared both strains by double

immunofluorescent microscopy (DIF) for differentiation of extra- and intracellular

streptococci. Results confirmed that the sly-positive strain adhered and invaded HEp-

2 cells much better than the sly-negative mutant, which could be detected only

Results, part I Chapter 3

63

extracellularly (Figure 3-1D and E). Quantification of microscopic examinations by

counting of intracellular bacteria per HEp-2 cell confirmed that these differences were

significant (P = 0.042; Figure 3-1F). Finally, we analysed adherence and invasion

process of both strains by field emission scanning electron microscopy (FESEM).

Similar to what we have found in previous studies (Benga et al., 2004), uptake of

unencapsulated sly-positive S. suis was accompanied by formation of membrane

ruffles in close vicinity to the streptococci (Figure 3-1G and H). Interestingly, this was

observed only for the sly-positive strain, but not for the sly-negative strain, which

showed adherence to cells but no association with ruffles or other cell morphologies

reminiscent of invasion (Figure 3-1I). These findings suggest that suilysin contributes

to S. suis induced membrane ruffling and uptake by epithelial cells.

Results, part I Chapter 3

64

Figure 3-1: Suilysin promotes invasion of S. suis in HEp-2 epithelial cells. Comparison of HEp-2 cell association, adherence and invasion of S. suis strain 10Δcps and its sly-negative mutant 10ΔcpsΔsly by microscopic analyses. (A) FACS cytometry analysis of HEp-2-cell association of S. suis strains. Cells were incubated with CFSE-labelled bacteria for 2 h and then fluorescence intensity of epithelial cells was measured by FACS. Results are expressed as mean fluorescent intensity (mfi) values. Uninfected cells served as background control. Mean and SD of eight independent experiments are shown. Significance is indicated by ** (P-value < 0.01). (B, C). Determination of adherence (B) and invasion (C) of S. suis strains by antibiotic protection assay after their incubation with HEp-2 cells for 2 h. Results are expressed as percentage of CFU recovered as compared to CFU used for infection. Mean and SD of three independent experiments are shown. Inlay in C shows immunoblot analysis for detection of suilysin expression in the supernatant of infected cells. The respective Coomassie stained SDS gel shown below indicates an unrelated protein used for loading control. Significance is indicated by * (P-value < 0.05).

Results, part I Chapter 3

65

(D-F). Determination of extra- and intracellular bacteria by double immunofluorescence (DIF) microscopy after incubation of S. suis strains with HEp-2 cells. In (D) cells were infected with strain 10Δcps, in (E) strain 10ΔcpsΔsly was used for infection. Extracellular bacteria are shown in orange-red and intracellular bacteria are stained in green. The nucleus was labelled by DAPI (blue). Bars represent 15 µm. (F) shows results of quantification of intracellular streptococci per cell by counting green (intracellular) bacteria in 50 epithelial cells (F). Mean and SD of two independent experiments are shown. Significance is indicated by * (P-value < 0.05). (G-I). Analysis of association of S. suis strains with HEp-2 cells by field emission scanning electron microscopy (FESEM). In (G) and (H) strain 10Δcps is shown to be associated with the formation of membrane ruffles. Strain 10ΔcpsΔsly adhered to the cell surface, but did not induce membrane ruffling or uptake by HEp-2 cells (I). Bars represent 1 µm. Invasion-promoting activity of suilysin involves Rac-dependent activation of the actin cytoskeleton. Based on our results described above we hypothesized that

suilysin promotes invasion of S. suis by manipulating the host cell cytoskeleton.

Since actin is the major component involved in formation of lammellipodia and

membrane ruffles, we first performed experiments with latrunculin B, a specific

inhibitor of actin polymerisation (Coue et al., 1987; Greenwood et al., 2006). As

shown in Figure 3-2A, FACS analysis revealed that bacteria-cell association of the

sly-positive strain was significantly reduced in the presence of latrunculin B as

compared to control cells without inhibitor. Reduced cell-association reached a level

which was almost similar to that of the sly-negative mutant (Figure 3-2A).

Next we studied the role of the small GTPases Rho, Rac, and Cdc42. These belong

to the family of Rho-GTPases which are molecular switches signalling to the actin

cytoskeleton, thereby translating environmental signals into cellular morphological

responses (Hall, 1998). We performed bacterial invasion experiments with HEp-2

cells, which were pretreated with different inhibitors of Rho-GTPases. First we used

toxin B of Clostridium (C.) difficile (TcdB), which inhibits all Rho-GTPases by

monoglycosylation (Just et al., 1995). The treatment of cells with TcdB resulted in a

significant reduction of invasion of strain 10cpsΔEF as determined by gentamicin

protection assays (Figure 3-2B). In contrast, the invasion of the isogenic sly-negative

strain 10cpsΔEFΔsly was not affected. Next we tested invasion of HEp-2 cells

pretreated with C3 exoenzyme, an ADP-ribosyltransferase of C. limosum, which

specifically inactivates Rho-, but not Rac- and Cdc42-GTPase. This treatment had no

effect on invasion of either of the two S. suis strains (Figure 3-2C). Thirdly, we used a

specific inhibitor of Rac1-GTPase, TcdB toxin of C. difficile serotype F strain 1470

(TcdB-F). TcdB-F pretreatment of epithelial cells resulted in a significant inhibition of

Results, part I Chapter 3

66

invasion of the sly-positive S. suis strain, which was as low as that of the isogenic

sly-mutant (Figure 3-2D). Taken together, these findings suggest that the actin

cytoskeleton and Rac-GTPases are involved in suilysin-mediated invasion of S. suis.

Figure 3-2: Suilysin-mediated invasion of S. suis in epithelial cells involves the actin cytoskeleton and Rho-GTPases. Effects of specific inhibitors of the actin cytoskeleton and small Rho-GTPases on HEp-2 cell association and invasion of S. suis strain 10cps and its sly-negative mutant 10ΔcpsΔsly. (A) FACS analysis of association of S. suis strains 10Δcps and 10ΔcpsΔsly with HEp-2 cells that were either incubated in the presence of latrunculin A (Lat A, 1 µg/ml) or left untreated. Results are expressed as means and SD of three independent experiments. Significance is indicated by * (P-value < 0.05). (C-D). Determination of invasion of S. suis strains by antibiotic protection assay after their incubation with HEp-2 cells that were pre-treated either with 10 ng/ml TcdB (B), 1 µg/ml C3 (C) or 10 ng/ml TcdB-F (D). Untreated cells served as controls (shown in black or white bars, respectively). Results are expressed as means and SD of three independent experiments. Significance is indicated by * (P-value < 0,05) or ** (P-value < 0.01). To further prove that suilysin can lead to activation of Rac-GTPase, we performed

pull down assays with specific ligands of activated Rac1 and RhoA. HEp-2 cells were

treated with recombinant suilysin (rSLY) at subcytolytic concentration (100 ng/ml),

Results, part I Chapter 3

67

lysed, and supernatants were precipitated with specific ligands coupled to agarose

beads. Amounts of bound GTPases were then detected by immunoblot analysis and

densitometric quantification of detected GTPase. Four independent analyses were

performed and results expressed as the relative quotient of active and total GTPase.

Results revealed a time-dependent increase of the amount of activated Rac1,

whereas level of activated RhoA remained unchanged (Figure 3-3A and B). The

relative amount of active Rac1 increased significantly (1.4 fold) reaching a maximum

after 15 min of HEp-2 cell treatment with rSLY, whereas activated RhoA remained

unchanged, even at longer times of rSLY treatment (data not shown).

Figure 3-3: Treatment of HEp-2 cells with recombinant suilysin (rSLY) leads to activation of Rac1, but not RhoA. HEp-2 cells were treated with rSLY (100 ng/ml) for indicated time periods and were then analysed by pull down assay using specific ligands for activated Rac1 and RhoA, respectively, as described in Experimental procedures. A representative immunoblot is shown in (A). Amounts of (activated and total) GTPases were quantified by immunoblot analysis and densitometry (B). Results are expressed as ratio of activated compared to total GTPase. Mean and SD of three independent experiments are shown. Significance is indicated by * (P-value < 0.05).

Results, part I Chapter 3

68

Assuming that actin and Rac seemed to play crucial roles in suilysin-mediated effects

we then performed colocalisation studies of actin, Rac1 and suilysin by confocal

immunofluorescence microscopy. For this, we used transiently Rac1-eGFP

expressing HEp-2 cells and treated them with rSLY at subcytolytic concentration

(100 ng/ml). Results revealed that rSLY bound to the cell plasma membrane, as

indicated by blue spots shown in Figure 3-4. The spots varied in size suggesting

different number of suilysin-molecules and stages of oligomerisation. Bound rSLY

appeared to be scattered over the entire surface of the HEp-2 cell and was frequently

located in close vicinity to or colocalised with F-actin (stained in red) and Rac1

(stained in green) (Figure 3-4).

Figure 3-4: Recombinant suilysin (rSLY) binds to HEp-2 cell membrane and is located in association with F-actin and Rac1. Confocal laser scanning micrograph of GFP-Rac1-expressing HEp-2 cells which were treated with rSLY for 30 min and then processed for confocal microscopy as described in Experimental procedures. F-actin is stained in red, Rac1 in green, and rSLY in blue. The optical slice of the region indicated by two vertical white lines in (A) is shown in (B). The right part shows an enlargement of (B). Arrows indicate colocalisation of rSLY, F-actin and Rac1. Bar represents 15 µm.

Suilysin-mediated promotion of bacteria-cell association and activation of Rac does not require formation of a cytolytic pore. The suilysin-mediated effects

observed in our study occurred at subcytolytic conditions, and they seemed to be

associated with binding and oligomerisation of suilysin, as suggested by

colocalisation experiments (see Figure 3-4). Therefore, we were interested to find out

Results, part I Chapter 3

69

whether formation of a cytolytic pore was required for suilysin-mediated bacteria-cell

association and Rac-activation. In some CDC it has been shown by structural or

functional analyses that the tryptophan (Trp)-rich motif at the tip of the C-terminal of

domain 4 is crucial for cytolytic activity (Billington et al., 2000). Hence, we introduced

a point mutation in the sly gene by site-directed mutagenesis, resulting in a single

amino acid substitution of a conserved Trp residue at position 461 (W was replaced

by F). An alignment of the homologous Trp-rich motif of suilysin, the mutated suilysin

derivative W461F, and pneumolysin (PLY) is shown in Figure 3-5A. We first

compared haemolytic and cytolytic activities of rSLY and rW461F. Determination of

haemolytic activities showed that the point-mutated toxin rW461F almost completely

lost its capacity to lyse sheep erythrocytes. Compared to rSLY, of which 256 (28)

ng/ml were sufficient for 50 % haemolysis, rW461F had to be applied at 4096 (212)

ng/ml to cause 50 % haemolysis (Figure 3-5B). Calculation of haemolytic units (HU)

revealed 0.53 x 105 HU per mg of rSLY as compared to 0.25 x 104 HU/mg for

rW461F. In good agreement, cytotoxic activities determined by LDH release assay

using HEp-2 cells were abolished in rW461F, reaching only 2.1% of that of rSLY.

Figure 3-5C shows the respective dose-response curve and estimation of

concentrations needed for 50% cytotoxicity. Based on these findings, for the

following experiments we used rSLY and rW461F at 100 ng/ml, which was clearly

below the (rSLY) protein concentrations causing 50% cytotoxicity. These data

indicate that substitution of one amino acid in the conserved Trp-rich motif of domain

4 completely inactivates haemolytic and cytolytic activity of suilysin. To our

knowledge, this is the first experimental proof that the Trp-rich motif of suilysin is

crucial for formation of a functional (i. e. cytolytic) pore, which corresponds well to the

high similarities of this region within the CDC family (as shown for pneumolysin in

Figure 3-5A) as well as to predictions based on the crystal structure (Xu et al., 2010).

Results, part I Chapter 3

70

Figure 3-5: Substitution of an amino acid in the Trp-rich undecapeptide of domain 4 of recombinant suilysin leads to abolishment of its haemolytic and cytolytic activity. (A) Primary sequence alignment of suilysin (SLY), suilysin with a point mutation resulting in replacement of a tryptophan by phenylalanine at position 461 (W461F) and pneumolysin (PLY). The highly conserved N-terminal undecapeptide of domain 4 within the CDC family is highlighted in grey. (B) Determination of haemolytic activity of rSLY and rW461F proteins by standard haemolysis assay using sheep erythrocytes. Proteins were added to erythrocytes in microtiter plates at indicated concentrations and haemolysis was determined by measuring the absorbance of cell-free supernatants at 550 nm. Results are expressed as % haemolysis compared to H2O-lysed erythrocytes. The horizontal line indicates 50% haemolysis. The inlay shows representative effects of rSLY (left, haemolysis) and rW461F (right, no haemolysis). (C) Determination of cytotoxic activity of rSLY and rW461F proteins by standard LDH release assay using HEp-2 cells. Proteins were added to cells in microtiter plates at indicated concentrations and LDH release as an indicator of cytotoxicity was determined in the supernatants as described in Experimental procedures. Results are expressed as % cytotoxicity compared to Triton X-100-lysed cells. The horizontal line indicates 50% cytotoxicity.

Results, part I Chapter 3

71

Next we tested whether addition of both recombinant proteins to the sly-negative

S. suis mutant could reconstitute the phenotype of the sly-positive parental strain. For

this, we determined HEp-2 cell-association of S. suis strains by FACS in the

presence and absence of both recombinant proteins. Results revealed that addition

of rSLY and rW461F both resulted in regain of function of the sly-negative strain, i. e.

the mutant reached cell-association levels comparable to that of the sly-positive

strain when proteins were present (Figure 3-6A). Immunoblot analysis for detection of

recombinant suilysin protein in the supernatants of infected cells showed that rSLY

and rW461F were present in equal amounts and were comparable to suilysin

expressed by the sly-positive strain (Figure 3-6A, lower part). Finally, we tested

whether rW461F was still able to activate host cell GTPase Rac. For this, we

performed pull down experiments as described above. Results clearly showed that

rW461F treatment of HEp-2 cells led to an increase of the amount of activated Rac1

comparable to that induced by rSLY (Figure 3-6B).

Concluding, our results revealed that suilysin can promote invasion of S. suis in

epithelial cells by Rac-dependent activation of the actin cytoskeleton. Furthermore,

we showed that these effects occur at subcytolytic concentrations and do not require

formation of a functional (cytolytic) pore. Whether this can be considered as a

common mechanism of CDC producing streptococci remains to be an interesting

question for future studies.

Results, part I Chapter 3

72

Figure 3-6: Substitution of an amino acid in the Trp-rich undecapeptide of domain 4 of recombinant suilysin does not affect its ability to promote HEp-2 cell association of S. suis or to activate Rac1. (A) FACS determination of HEp-2 cell association of S. suis strains in the presence of rSLY or rW461F. Cells were incubated with CFSE-labelled bacteria for 2 h and then fluorescence intensity of epithelial cells was measured by FACS. Results are expressed as mean fluorescent intensity (mfi) values. Uninfected cells served as background control. Mean and SD of three independent experiments are shown. Significance is indicated by * (P-value < 0.05) or ** (P-value < 0.01). Detection of suilysin in the respective supernatants by immunoblot analysis is shown underneath the graph. The respective Coomassie stained SDS gel shown below indicates an unrelated protein used for loading control. (B, C) Pull down analysis of Rac1-activation in HEp-2 cells treated with rSLY or rW461F for 15 min. (B) Representative immunoblot, (C) quantification of amounts of activated Rac1 as described in Figure 3-3.

Results, part I Chapter 3

73

Discussion Suilysin has been identified as a secreted toxin produced by many virulent S. suis

strains (Jacobs et al., 1995; Segers et al., 1998). It belongs to the pore-forming

cholesterol-dependent cytolysin (CDC) family, as recently confirmed by determination

of its crystal structure (Xu et al., 2010). In vitro and in vivo experiments suggest that

suilysin, though it is not essential for virulence of S. suis, most likely contributes to

pathogenesis by modification of host-pathogen interactions (Norton et al., 1999; Allen

et al., 2001; Vadeboncoeur et al., 2003; Lun et al., 2003; Benga et al., 2008). The

current concept of the biological role of suilysin is a matter of ongoing discussions. A

possible role of suilysin in host cell invasion of S. suis has been suggested by Norton

et al. (1999). Furthermore, we and others have observed that suilysin may be

involved in protection against opsonophagocytosis and cytokine release by host cells

(Benga et al., 2008; Lecours et al., 2011). However, the molecular mechanisms of

these effects are poorly understood. Considering that suilysin may in vivo often not

reach concentrations sufficient for lysis of host cell membranes, it is plausible to

speculate that the toxin expresses biological activities below cytotoxicity. Subcytolytic

effects of suilysin have not been investigated in detail, but for some other CDC such

activities have been shown to modulate host cell responses (Tsuchiya et al., 2005;

Krawczyk-Balska and Bielecki, 2005; Ratner et al., 2006; Iliev et al., 2009).

This prompted us to analyse effects of suilysin on epithelial cell-interactions of S. suis

at subcytolytic conditions. Based on preliminary studies to establish such conditions,

we then used 100 ng/ml (recombinant) suilysin in all experiments, which are clearly

below cytolytic effects (see also Figure 3-5). Accordingly, in studies on other CDC

nanomolar concentrations were also found to be far beyond that needed for cytolytic

pore-formation (Ratner et al., 2006).

An unencapsulated S. suis serotype 2 strain was used since in our previous studies

unencapsulated S. suis adhered to and invaded epithelial cells much better than

encapsulated strains (Benga et al., 2004). One might argue that an unencapsulated

strain does not represent a virulent phenotype. However, capsule expression is

highly regulated, most likely in response to environmental signals such as glucose

Results, part I Chapter 3

74

availability (Willenborg et al., 2011). Thus, by using this strain we mimic an in vivo

situation of low capsule expression which is relevant mainly during adherence and

invasion of host cells. Our comparison of this strain with its sly-negative mutant

revealed that inactivation of the sly gene resulted in a significant reduction of

bacteria-cell association, as determined by FACS analysis, and of invasion, as

determined by gentamicin-protection assay and microscopy. Furthermore, we

observed that cholesterol-depletion of epithelial cells significantly reduced uptake of

streptococci (data not shown). These results suggest that suilysin contributes to

invasion of S. suis.

Our findings in this and a previous study (Benga et al., 2004) revealed that S. suis

invasion of epithelial cells was accompanied by formation of membrane ruffles. A

major component of ruffles is actin. Thus, we hypothesized that the invasion-

mediating effects of suilysin may be due to activation of the actin host cell

cytoskeleton after binding of the (secreted) toxin to the host cell membrane.

Involvement of the actin cytoskeleton in invasion process could be shown using

latrunculin B as a specific inhibitor. It is known to complex G-actin and prevents actin

polymerisation (Coue et al., 1987), and, in our study, caused a significantly reduced

HEp-2 cell association of S. suis. Membrane ruffling and involvement of actin is also

seen in the trigger-like uptake mechanism described for other pathogenic bacteria,

e. g. Salmonella (Finlay and Cossart, 1997). Noteworthy, our results suggest that in

S. suis a toxin can promote membrane ruffling and uptake of bacteria.

It has been shown for other pathogens that activation of GTPases is involved in

bacterial invasion of host cells (Finlay, 2005). The underlying mechanisms seem to

be rather complex and differ between pathogenic bacteria. For instance,

Rho-GTPases Rac, Rho and Cdc42 are necessary for invasion of group A

streptococci (Burnham et al., 2007). In contrast, invasion of Salmonella Typhimurium

requires only Rac1 and Cdc42 (Criss and Casanova, 2003). Hence, to analyse the

molecular basis of suilysin-mediated effects on actin remodelling and streptococcal

uptake in more depth, we focussed on involvement of small Rho-GTPases. These

proteins are known to signal to the actin cytoskeleton, thereby regulating

morphological responses of the cell to environmental signals (Mackay and Hall,

Results, part I Chapter 3

75

1998). Using clostridial toxins as specific inhibitors we could show that the

recruitment of Rac-GTPase, but not Rho is crucial for invasion-mediating effects of

suilysin. This corresponds to the well-established role of Rac in polymerization of

actin fibres involved in formation of lamellipodia and membrane ruffles (Ridley et al.,

1992), whereas Rho is considered a regulator mainly of stress fibre formation.

Accordingly, our colocalisation experiments revealed that suilysin bound to the

epithelial cell membranes and was spatially associated with Rac1 and F-actin. This

supports our assumption that suilysin can bind to epithelial cells and activate Rac1,

which leads to activation of the actin cytoskeleton and formation of membrane ruffles.

To prove that activated GTP-ases were involved we treated HEp-2 cells with suilysin

and then detected activated Rac1 and RhoA by pull down analysis using specific

ligands coupled to agarose beads. Results showed that Rac1 was activated within

15 min after addition of suilysin, whereas activation of RhoA seemed not to be

affected. Involvement of Cdc42 in streptococcal internalisation is unlikely, because

activation of Cdc42 triggers mainly the formation of filopodia and microspikes

(Mackay and Hall, 1998). Accordingly, in preliminary experiments we found that

Cdc42 is not significantly activated by suilysin (data not shown).

Among other members of CDC an interaction with G-proteins has been described for

pneumolysin. Iliev et al. (2007) reported formation of stress fibres, filopodia, and

lamellipodia as a consequence of RhoA and Rac1 recruitment after treatment of

human neuroblastoma cells with pneumolysin. Strikingly, suilysin seems to activate

only Rac1, but not RhoA. However, it has to be considered that we used respiratory

epithelial cells, and that both host cells play different roles during streptococcal

infection process.

The CDC share two basic functions, which is reflected by more than 40% sequence

identity over their common domain regions 1-4 (Xu et al., 2010). First, CDC bind to

host cell membrane via specific cholesterol-dependent interaction between the

C-terminal region of domain 4 and the cell membrane. Second, after binding toxin

monomers oligomerise and change conformation in the N-terminal region, which

finally leads to formation of transmembrane pores and lysis of the target cell (Tweten,

2005). Structural-functional analyses revealed that domain 4 carries a highly

Results, part I Chapter 3

76

conserved tryptophan-rich motif consisting of 11 amino acids, which is crucial for

cytolytic activity of CDC (Billington et al., 2000). An alignment of the homologous

region of pneumolysin and suilysin is shown in Figure 3-5A. Point mutations within

this undecapeptide of pneumolysin (Feldman et al., 1990), perfringolysin O of

Clostridium perfringens (Polekhina et al., 2005) or listeriolysin O of Listeria

monocytogenes (Michel et al., 1990) lead to an attenuated ability to cause membrane

damage. It is known that such mutated toxins are still able to bind to the cell plasma

membrane, but the formation of a functional permeable transmembrane pore is

inhibited (Korchev et al., 1998). For suilysin, predictions based on its crystal structure

suggest that it shares common features with other CDC and, concerning the

undecapeptide, particularly with intermedilysin (Xu et al., 2010). However, these

predictions have yet to been proven by functional experiments.

Interactions of pneumolysin with Rho-GTPases (see above) seem to occur

independent of (macro) pore-formation (Iliev et al., 2007). This prompted us to further

dissect the relevance of pore-formation in suilysin-mediated effects on epithelial cells.

For this, we took advantage of the high similarities in primary amino acid sequence

and three-dimensional secondary structure among members of CDC family (Segers

et al., 1998b; Xu et al., 2010) as described above. Hence we produced recombinant

suilysin in which the undecapeptide was inactivated by substitution of the first

tryptophan (W461) by phenylalanine (see also alignment in Figure 3-5A). Haemolytic

and cytolytic activity of the mutated recombinant suilysin (rW461F) were almost

completely abolished in comparison with the suilysin (rSLY), indicating a loss in

transmembrane pore-forming activity. This was further confirmed in experiments

using fluorescent calcein to prove macropore-formation (data not shown).

Remarkably, the mutated suilysin had retained its ability to promote epithelial cell

invasion of S. suis and to activate Rac1 comparable to rSLY.

To our knowledge, this is the first experimental evidence for the function of this

domain in suilysin-mediated cytolysis according to the predictions based on the

crystal structure (Segers et al., 1998b; Xu et al., 2010). Furthermore, our results

indicate that a functional undecapeptide motif in domain 4 is not required for

CDC-mediated GTPase-activation and invasion of bacteria.

Results, part I Chapter 3

77

Concluding, this study revealed that suilysin can promote invasion of S. suis in

epithelial cells by Rac-dependent activation of the actin cytoskeleton at subcytolytic

conditions. Thus, we propose that subcytolytic activities of suilysin play a role in host-

pathogen interactions by modulating different host cell functions, such as

morphological responses, as shown here, or release of cytokines, as shown by

others (Lun et al., 2003; Chen et al., 2007; Vanier et al., 2009; Zheng et al., 2011;

Lecours et al., 2011). Further studies will show which domains of suilysin are

involved in this process and which significance this might have for virulence and

pathogenesis.

Results, part I Chapter 3

78

Experimental procedures If not stated otherwise all materials were purchased from Sigma (Muenchen,

Germany).

Bacterial strains and growth conditions. The unencapsulated isogenic mutant

strain 10cpsΔEF was generated by insertional mutagenesis of the virulent

Streptococcus (S.) suis serotype 2 strain 10 (both strains were kindly provided by H.

Smith, Lelystad, NL) as described previously (Smith et al., 1999). The corresponding

suilysin deficient mutant of 10cpsΔEF, strain 10cpsΔEFΔsly, was constructed as

described below. Escherichia (E.) coli strains BL21, BL21 (DE3) and DH5α were

used for molecular cloning and protein expression experiments. Streptococci were

grown on Columbia agar supplemented with 7% sheep blood (Oxoid, Wesel,

Germany) overnight under aerobic conditions at 37°C. For infection of HEp-2 cells,

streptococci were grown in Todd-Hewitt broth (THB, Difco, Detroit, USA) overnight at

37°C under aerobic conditions, adjusted to an optical density (OD600) of 0.02 in pre-

warmed media the next day and grown to late exponential growth phase (OD600 0.8).

E. coli strains were cultured on Luria Bertani (LB) agar overnight at 37°C under

aerobic conditions. When necessary, antibiotics were added to the media at the

following concentrations: spectinomycin 100 µg/ml for S. suis; erythromycin 1 µg/ml

for S. suis; ampicillin 100 µg/ml for E. coli.

DNA techniques. Routine molecular biology techniques including restriction

endonuclease digestion, DNA ligations, agarose gel electrophoresis, Southern Blot

analysis, transformation of E. coli and plasmid isolation were performed according to

standard procedures (Sambrook et al., 1989). Restriction enzymes were purchased

from New England Biolabs (Frankfurt am Main, Germany). Plasmid preparations

were performed with kits from Machery-Nagel (Dueren, Germany).

Construction of the suilysin deficient mutant strain 10cpsΔEFΔsly. The mutant

10cpsΔEFΔsly was constructed by insertion of an erythromycin cassette in the sly

Results, part I Chapter 3

79

gene of the unencapsulated strain 10cpsΔEF using the plasmid pBlue/sly/erm as

previously described (Benga et al., 2008). Mutants were controlled by PCR and

Southern Blot analysis.

Expression of recombinant wild type suilysin proteins. Recombinant His-tagged

suilysin (rSLY) and point-mutated suilysin W461F (rW461F) were expressed in BL21

pET45bslynew (Kock et al., 2009) and BL21 pET45bslynewW461F, respectively. The

plasmid pET45bslynewW461F was constructed by site-directed-mutagenesis

according to the instruction manual of QuikChange® Site-Directed Mutagenesis Kit

(Stratagene, Ja Jolla, CA) with some modifications. Purified plasmid DNA carrying

the sly gene (pET45bslynew) was amplified using oligonucleotide primers encoding

the W461 to F461 substitution (slyTrp-Phefornew [TACAGGATTAGCAT

TTGAGTGGTGGAGAAC], slyTrp-Pherevnew [GTTCTCCACCACTCAAATGCTAATC

CTGTAC]). The resulting plasmid pET45bslynewW461F was electroporated into E.

coli BL21 (DE3), purified and controlled by sequencing. Expression and purification

of rSLY and rW461F were performed as described previously (Willenborg et al.,

2011). The quality of purified proteins was controlled by separation on SDS-

polyacrylamide gels and immunoblot analysis using with anti-Penta His antibodies

(Qiagen, Hilden, Germany).

Immunoblot analysis. Recombinant proteins or supernatants of infected cells were

separated by SDS-polyacrylamide gel electrophoresis with a 4% stacking and a 10%

separating gel under denaturing conditions and transferred to a PVDF-membrane

(Serva, Heidelberg, Germany). Membrane-blocking was performed overnight with 1%

milk powder in TBS with 0.5% Tween. For detection of His-tagged recombinant

proteins membranes were blocked with 3% BSA and incubated with a monoclonal

anti-Penta His antibody (Qiagen) diluted 1:2,000 in 3% BSA. Polyclonal antiserum

raised against rSLY (Benga et al., 2008) diluted 1:1,300 in 1% milk powder was used

to detect either rSLY, rW461F or secreted suilysin in supernatants of infected cells.

Membranes were developed with horseradish peroxidase-conjugated anti-mouse or

anti-rabbit IgG antiserum diluted 1:10,000 in 1% milk powder (Amersham, Freiburg,

Results, part I Chapter 3

80

Germany) followed by using SuperSignal® West Pico Chemiluminescent Substrate

(Pierce, Rockford, USA) according to the manufactures protocol.

HEp-2 epithelial cell culture. The human laryngeal epithelial cell line HEp-2 (ATCC

CCL 23) was used as described previously (Benga et al., 2004). For adherence and

invasion assays, approximately 1.8 x 105 cells per well were seeded on 24 well tissue

culture plates. For immunofluorescence and electron microscopy 0.5 x 105 and

1.5 x 105 cells per well were seeded on 12 mm diameter glass cover slips placed in

24 well plates. For FACS analysis approximately 6 x 105 cells per well were seeded

on a 6 well plate, and for cellular cytotoxicity assay 0.2 x 105 cells per well were

seeded on a 96 well plate. The cells were grown overnight and then used for

experiments. For colocalisation experiments HEp-2 cells were transfected with

GFP-tagged wild type Rac1 (GFP-Rac1) using the Amaxa nucleofection system

(Amaxa, Cologne, Germany) applying the standard protocol as specified by the

manufacturer. Plasmids were purified with the EndoFreeTM plasmid purification kit

from Qiagen.

In some assays, HEp-2 cells were treated before infection with the following

inhibitors: 1 µg/ml latrunculin B for 30 min, 10 ng/ml toxin B of Clostridium (C.) difficile

(TcdB), 10 ng/ml toxin B of C. difficile serotype F strain 1470 (TcdB-F) and 1 µg/ml

C3 toxin of C. limosum for 3 h. Toxins were purified as described previously

(Huelsenbeck et al., 2007). C3 toxin was always reconstituted freshly by mixing the

binding domain (C2IIa) with the catalytic domain (FT) in a ratio of 1:1 as described

(Just et al., 1995). After preincubation with inhibitors cells were washed and used for

further experiments as described below, except for latrunculin B, which was left on to

the cells during the whole infection experiment because its effect is reversible up to

1 h.

Determination of haemolytic activity and cytotoxicity. Haemolytic activities of

rSLY and rW461F were determined in 96-well vee bottom microplates as described

by Takamatsu et al. (2001) with some modifications. Briefly, twofold dilutions of the

toxins were incubated with a 2% sheep erythrocyte suspension in 0.9% NaCl for 2 h

Results, part I Chapter 3

81

at 37°C. Unlysed red blood cells were allowed to pellet by centrifugation, and 100 µl

of the supernatant was transferred into a new flat bottom microplate. Subsequently,

absorption was measured at 550 nm in a microplate reader (GENios Pro, TECAN

AUSTRIA GMBH). Determination of haemolytic activity was done in three

independent experiments. One haemolytic unit (HU) was defined as the

concentration of toxin causing 50% haemolysis of a 2% sheep blood suspension

(Takamatsu et al., 2001).

Cytotoxicity was detected by LDH-release assay as described previously (Benga et

al., 2004). Briefly, HEp-2 cells were either incubated with rSLY or rW461F at

indicated concentrations. Supernatants were removed 2 h after treatment and LDH

activities were determined using the Cytotox® 96 assay kit (Promega, Mannheim,

Germany). All experiments were performed in triplicates and repeated at least three

times. Results were expressed as % LDH-release compared that of Triton X-100

lysed non-infected cells.

Antibiotic protection assay. The number of adherent and invasive streptococci was

quantified using a gentamicin protection assay as described earlier (Valentin-

Weigand et al., 1996) with the following modifications. Adherence and invasion

assays were performed with streptococci grown in THB to an OD600 of 0.8.

Subsequently, bacteria were resuspended in Dulbecco’s PBS containing 0.9 mM

CaCl2 and 0.5 mM MgCl26H2O, pH 7.3 (PBS+; Gibco-Invitrogen) and adjusted to an

OD600 0.6. Then, streptococcal suspensions were diluted 1:10 in PBS+ and used for

adherence and invasion assay. Confluent HEp-2 cell monolayers were inoculated

with 100 bacteria per cell. The number of colony forming units (CFU) inoculated per

well was determined by serial platings. After incubation infected cells for 2 h the

supernatants were collected for detection of suilysin expression by immunoblot

analysis. To determine adherence cells were washed three times with PBS

immediately after supernatant collection and lysed by 0.025% Triton X-100. Serial

dilutions of these lysates were plated in triplicates on Columbia agar for

determination of CFU. For invasion assay, after PBS washes cells were incubated for

2 h with 100 μg gentamicin and 5 μg penicillin per ml to kill extracellular bacteria.

Results, part I Chapter 3

82

Then lysates were prepared and plated as described for adherence. Results were

expressed as % adherence/invasion compared to the inoculum used for infection.

FACS analysis of bacteria-cell association. For this, streptococcal strains were

labelled with the carboxyfluorescein succinimidyl ester (CFSE) using CellTrace™

CFSE Cell Proliferation Kit (Invitrogen). Labelling was performed as described by

(Logan et al., 1998) with some modifications. Briefly, S. suis strains were grown to an

OD600 of 0.8, pelleted by centrifugation, washed once with PBS+ and resuspended in

PBS+ containing 5 µM CFSE-reagent (prepared from a CFSE stock solution as

recommended by the manufacturer). Staining procedures were carried out at 37°C

for 20 min on an end-over-end rotator. After removal of unbound fluorescent dye by

washing labelling efficiency was confirmed to be > 90%.

HEp-2 cell monolayers were infected with CFSE-labelled streptococci at a multiplicity

of infection [moi] of 125:1 bacteria per cell. For this, streptococci were centrifuged on

the epithelial cells (250 x g for 5 min) and then incubated with the cells at 37°C for

2 h. Then supernatants were collected for determination of suilysin expression by

immunoblot analysis. Epithelial cells were then detached with trypsin-EDTA, washed

with PBS and fixed with 0.375% formaldehyde in PBS. In some experiments, rSLY or

rW461F were added to epithelial cells together with streptococcal strains (100 ng/ml,

final concentration). To exclude cytotoxic effects of suilysin during the experiment

cells were stained afterwards with the DNA-intercalating dye propidium iodide

(4 µg/ml) for 2 min at room temperature to discriminate between viable and nonviable

cells. Bacteria-cell association was measured using FACScan® (Becton Dickinson,

488 nm Argon laser) and analysed using the software WinMDI (version 2.9.). For

each determination at least 10,000 events were measured. Results were expressed

as mean fluorescent intensity [mfi] values of the cell population. Uninfected cells

served as background control. Means and standard deviations were determined from

eight independent experiments. Fluorescence and electron microscopy. Determination of adherence and invasion

of streptococci by double immunofluorescence microscopy (DIF) was performed as

Results, part I Chapter 3

83

described previously (Benga et al., 2004) with the following modifications.

Semiconfluent HEp-2 cells were incubated with streptococci at a moi of 100 bacteria

per epithelial cell. ProLong® Gold antifade reagent with DAPI (Invitrogen) was used

for sealing and staining of the nucleus. Mounted samples were examined using

inverted immunofluorescence microscope Nikon Eclipse Ti-S equipped with a 40×,

0.6 S Plan Fluor objective (Nikon, Duesseldorf, Germany) driven by NIS Elements

software BR 3.2.. Intracellular streptococci were quantified by counting green

(intracellular) bacteria in at least 50 cells, and results were expressed as intracellular

bacteria per epithelial cell.

For colocalisation experiments HEp-2 cells transiently expressing GFP-Rac1 were

fixed after incubation with rSLY (100 ng/ml) for 30 min at 37°C with 3%

paraformaldehyde in modified cytoskeleton buffer (10 mM MES [pH 7.0], 150 mM

NaCl, 5 mM EGTA, 5 mM MgCl2, 5 mM sucrose) for 20 min, quenched with 10 mM

glycin in PBS and then permeabelised with 0.1% Triton X-100 in PBS. Permeabilised

cells were blocked in PBS with 5% horse serum and 1% BSA (blocking buffer) for

30 min at room temperature and subsequently incubated with polyclonal antiserum

raised against rSLY (Benga et al., 2008) diluted 1:1,200 in blocking buffer for 45 min

at room temperature. Samples were washed twice with PBS and incubated with

Cy5-conjugated goat anti-rabbit IgG (Millipore, Schwalbach/Ts, Germany) for 30 min

at room temperature. F-actin was labelled with Alexa® Fluor 568-conjugated

Phalloidin (Invitrogen) for 20 min at room temperature. Coverslips were washed three

times in PBS and then mounted using ProLong® Gold antifade reagent with DAPI

(Invitrogen). Mounted samples were examined using a LSM 510 Meta confocal

microscope equipped with a 63×, 1.2NA Plan-NEOFLUAR oil immersion objective

(Zeiss, Jena, Germany) driven by LSM software v3.2. For 3D z-stack acquisition the

pinholes were set to one Airy unit and confocal planes were acquired every 340 nm.

Images were deconvolved and maximum intensity projection were rendered using

Huygens® Essential (Hilversum, The Netherlands).

Field emission scanning electron microscopy (FESEM) was performed as described

previously (Benga et al., 2004) using HEp-2 cells infected as described for DIF.

Results, part I Chapter 3

84

Pull down experiments. Experiments were done as described previously with some

modifications (Reid et al., 1996). HEp-2 cells were incubated with rSLY or rW461F

(100 ng/ml) for indicated time periods. Then cells were rinsed twice with ice cold PBS

to stop incubation. Lysis of cells was done by addition of 1 ml of ice cold lysis buffer

(50 mM NaCl, 20 mM Tris-HCl [pH 7.4], 3 mM MgCl2, 1% Nonidet-P 40,

0.25% Triton X-100, 5 mM dithiothreitol, 100 µM PMSF). Cells were scraped off, and

the lysates were centrifuged at 14,000 rpm. The supernatant was split and used for

pull down experiments of activated RhoA and Rac1 in parallel. For this, 20 µl of

beads slurry bearing approximately 20 µg protein of the GST-fusion protein of either

the Rho-binding domain C21 or the PAK-GBD, respectively, were added to 500 µl

sample and incubated on a rotator at 4°C for 30 min. Beads were collected by

centrifugation at 10,000 rpm, washed twice with lysis buffer and subjected to

immunoblot analysis using monoclonal antibodies against Rac-1 (cloe 102; BD

Pharmingen, Heidelberg, Germany) and against RhoA (clone 26C4; Santa Cruz,

Heidelberg, Germany), respectively. Amounts of (activated and total) GTPases were

quantified by densitometric measurement of three independent immunoblot analyses.

Statistical analyses. If not stated otherwise, experiments were performed at least

three times and results were expressed as means and standard deviations. Data

were analyzed by t-test. Pull down analysis was carried out using an unpaired t-test

with no assumption of Gaussian distribution (Mann-Whitney). A P-value <0.05 was

considered significant.

Acknowledgements We thank Hilde E. Smith (Lelystad, NL) for providing S. suis strain 10cpsΔEF. This

work was financially supported by the Deutsche Forschungsgemeinschaft (DFG),

Bonn, Germany (SFB587, GRK745).

Results, part I Chapter 3

85

References Allen, A. G., Bolitho, S., Lindsay, H., Khan, S., Bryant, C., Norton, P., Ward, P., Leigh, J., Morgan, J., Riches, H., Eastty, S., and Maskell, D. (2001) Generation and characterization of a defined mutant of Streptococcus suis lacking suilysin. Infect Immun 69: 2732-2735.

Arends, J. P., Hartwig, N., Rudolphy, M., and Zanen, H. C. (1984) Carrier Rate of Streptococcus suis Capsular Type-2 in Palatine Tonsils of Slaughtered Pigs. J Clin Microbiol 20: 945-947.

Arends, J. P. and Zanen, H. C. (1988) Meningitis Caused by Streptococcus suis in Humans. Rev Infect Dis 10: 131-137.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cell Microbiol 6: 867-881.

Billington, S. J., Jost, B. H., and Songer, J. G. (2000) Thiol-activated cytolysins: structure, function and role in pathogenesis. FEMS Microbiol Lett 182: 197-205.

Burnham, C. A., Shokoples, S. E., and Tyrrell, G. J. (2007) Rac1, RhoA, and Cdc42 participate in HeLa cell invasion by group B streptococcus. FEMS Microbiol Lett 272: 8-14.

Charland, N., Nizet, V., Rubens, C. E., Kim, K. S., Lacouture, S., and Gottschalk, M. (2000) Streptococcus suis serotype 2 interactions with human brain microvascular endothelial cells. Infect Immun 68: 637-643.

Chen, C., Tang, J., Dong, W., Wang, C., Feng, Y., Wang, J., Zheng, F., Pan, X., Liu, D., Li, M., Song, Y., Zhu, X., Sun, H., Feng, T., Guo, Z., Ju, A., Ge, J., Dong, Y., Sun, W., Jiang, Y., Wang, J., Yan, J., Yang, H., Wang, X., Gao, G. F., Yang, R., Wang, J., and Yu, J. (2007) A glimpse of streptococcal toxic shock syndrome from comparative genomics of S. suis serotype 2 Chinese isolates. PLoS ONE 2: e315.

Coue, M., Brenner, S. L., Spector, I., and Korn, E. D. (1987) Inhibition of actin polymerization by latrunculin A. FEBS Lett 213: 316-318.

Criss, A. K. and Casanova, J. E. (2003) Coordinate regulation of Salmonella enterica serovar Typhimurium invasion of epithelial cells by the Arp2/3 complex and Rho GTPases. Infect Immun 71: 2885-2891.

Feldman, C., Mitchell, T. J., Andrew, P. W., Boulnois, G. J., Read, R. C., Todd, H. C., Cole, P. J., and Wilson, R. (1990) The effect of Streptococcus pneumoniae pneumolysin on human respiratory epithelium in vitro. Microb Pathog 9: 275-284.

Finlay, B. B. (2005) Bacterial virulence strategies that utilize Rho GTPases. Curr Top Microbiol Immunol 291: 1-10.

Finlay, B. B. and Cossart, P. (1997) Exploitation of mammalian host cell functions by bacterial pathogens. Science 276: 718-725

Gottschalk, M., Segura, M., and Xu, J. (2007) Streptococcus suis infections in humans: the Chinese experience and the situation in North America. Anim Health Res Rev 8: 29-45.

Results, part I Chapter 3

86

Gottschalk, M., Xu, J., Calzas, C., and Segura, M. (2010) Streptococcus suis: a new emerging or an old neglected zoonotic pathogen? Fut Microbiol 5: 371-391. Greenwood, J., Steinman, L., and Zamvil, S. S. (2006) Statin therapy and autoimmune disease: from protein prenylation to immunomodulation. Nat Rev Immunol 6: 358-370.

Hall, A. (1998) Rho GTPases and the actin cytoskeleton. Science 279: 509-514.

Huelsenbeck, J., Dreger, S., Gerhard, R., Barth, H., Just, I., Genth, H. (2007) Difference in the cytotoxic effects of toxin B from Clostridium difficile strain VPI 10463 and toxin B from variant Clostridium difficile strain 1470. Infect Immun 75(2): 801-9.

Iliev, A. I., Djannatian, J. R., Nau, R., Mitchell, T. J., and Wouters, F. S. (2007) Cholesterol-dependent actin remodeling via RhoA and Rac1 activation by the Streptococcus pneumoniae toxin pneumolysin. Proc Natl Acad Sci U S A 104: 2897-2902.

Iliev, A. I., Djannatian, J. R., Opazo, F., Gerber, J., Nau, R., Mitchell, T. J., and Wouters, F. S. (2009) Rapid microtubule bundling and stabilization by the Streptococcus pneumoniae neurotoxin pneumolysin in a cholesterol-dependent, non-lytic and Src-kinase dependent manner inhibits intracellular trafficking. Mol Microbiol 71: 461-477.

Jacobs, A. A. C., Loeffen, P. L. W., vandenBerg, A. J. G., and Storm, P. K. (1994) Identification, Purification, and Characterization of A Thiol-Activated Hemolysin (Suilysin) of Streptococcus suis. Infect Immun 62: 1742-1748.

Jacobs, A. A. C., vandenBerg, A. J. G., and Loeffen, P. L. W. (1996) Protection of experimentally infected pigs by suilysin, the thiol-activated haemolysin of Streptococcus suis. Veterinary Record 139: 225-228.

Just, I., Selzer, J., Wilm, M., Eichel-Streiber, C., Mann, M., and Aktories, K. (1995) Glucosylation of Rho proteins by Clostridium difficile toxin B. Nature 375: 500-503.

King, S. J., Heath, P. J., Luque, I., Tarradas, C., Dowson, C. G., and Whatmore, A. M. (2001) Distribution and genetic diversity of suilysin in Streptococcus suis isolated from different diseases of pigs and characterization of the genetic basis of suilysin absence. Infect Immun 69: 7572-7582.

Kock, C., Beineke, A., Seitz, M., Ganter, M., Waldmann, K. H., Valentin-Weigand, P., and Baums, C. G. (2009) Intranasal immunization with a live Streptococcus suis isogenic ofs mutant elicited suilysin-neutralization titers but failed to induce opsonizing antibodies and protection. Vet Immuno Immunopathol 132: 135-145.

Korchev, Y. E., Bashford, C. L., Pederzolli, C., Pasternak, C. A., Morgan, P. J., Andrew, P. W., and Mitchell, T. J. (1998) A conserved tryptophan in pneumolysin is a determinant of the characteristics of channels formed by pneumolysin in cells and planar lipid bilayers. Biochem J 329 ( Pt 3): 571-577.

Krawczyk-Balska, A. and Bielecki, J. (2005) Listeria monocytogenes listeriolysin O and phosphatidylinositol-specific phospholipase C affect adherence to epithelial cells. Can J Microbiol 51: 745-751.

Lalonde, M., Segura, M., Lacouture, S., and Gottschalk, M. (2000) Interactions between Streptococcus suis serotype 2 and different epithelial cell lines. Microbiology 146 ( Pt 8): 1913- 1921.

Results, part I Chapter 3

87

Lecours, M. P., Gottschalk, M., Houde, M., Lemire, P., Fittipaldi, N., and Segura, M. (2011) Critical Role for Streptococcus suis Cell Wall Modifications and Suilysin in Resistance to Complement-Dependent Killing by Dendritic Cells. J Infect Dis 204: 919-929.

Logan, R. P., Robins, A., Turner, G. A., Cockayne, A., Borriello, S. P., and Hawkey, C. J. (1998) A novel flow cytometric assay for quantitating adherence of Helicobacter pylori to gastric epithelial cells. J Immunol Methods 213: 19-30.

Lun, S. C., Perez-Casal, J., Connor, W., and Willson, P. J. (2003) Role of suilysin in pathogenesis of Streptococcus suis capsular serotype 2. Microb Pathogen 34: 27-37.

Mackay, D. J. and Hall, A. (1998) Rho GTPases. J Biol Chem 273: 20685-20688.

Michel, E., Reich, K. A., Favier, R., Berche, P., and Cossart, P. (1990) Attenuated mutants of the intracellular bacterium Listeria monocytogenes obtained by single amino acid substitutions in listeriolysin O. Mol Microbiol 4: 2167-2178.

Norton, P. M., Rolph, C., Ward, P. N., Bentley, R. W., and Leigh, J. A. (1999) Epithelial invasion and cell lysis by virulent strains of Streptococcus suis is enhanced by the presence of suilysin. FEMS Immunol Med Microbiol 26: 25-35.

O'Sullivan, T., Friendship, R., Blackwell, T., Pearl, D., McEwen, B., Carman, S., Slavic, D., and Dewey, C. (2011) Microbiological identification and analysis of swine tonsils collected from carcasses at slaughter. Can J Vet Res 75: 106-111.

Polekhina, G., Giddings, K. S., Tweten, R. K., and Parker, M. W. (2005) Insights into the action of the superfamily of cholesterol-dependent cytolysins from studies of intermedilysin. Proc Natl Acad Sci U S A 102: 600-605.

Ratner, A. J., Hippe, K. R., Aguilar, J. L., Bender, M. H., Nelson, A. L., and Weiser, J. N. (2006) Epithelial cells are sensitive detectors of bacterial pore-forming toxins. J Biol Chem 281: 12994-12998.

Reid, T., Furuyashiki, T., Ishizaki, T., Watanabe, G., Watanabe, N., Fujisawa, K., Morii, N., Madaule, P., and Narumiya, S. (1996) Rhotekin, a new putative target for Rho bearing homology to a serine/threonine kinase, PKN, and rhophilin in the rho-binding domain. J Biol Chem 271: 13556-13560.

Ridley, A. J., Paterson, H. F., Johnston, C. L., Diekmann, D., and Hall, A. (1992) The small GTP- binding protein rac regulates growth factor-induced membrane ruffling. Cell 70: 401-410.

Rubins, J. B., Paddock, A. H., Charboneau, D., Berry, A. M., Paton, J. C., and Janoff, E. N. (1998) Pneumolysin in pneumococcal adherence and colonization. Microb Pathog 25: 337-342.

Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N Y.

Segers, R. P., Kenter, T., de Haan, L. A., and Jacobs, A. A. (1998) Characterisation of the gene encoding suilysin from Streptococcus suis and expression in field strains. FEMS Microbiol Lett 167: 255-261.

Segura, M. and Gottschalk, M. (2002) Streptococcus suis interactions with the murine macrophage cell line J774: adhesion and cytotoxicity. Infect Immun 70: 4312-4322.

Results, part I Chapter 3

88

Smith, H. E., Damman, M., van der Velde, J., Wagenaar, F., Wisselink, H. J., Stockhofe-Zurwieden, N., and Smits, M. A. (1999) Identification and characterization of the cps locus of Streptococcus suis serotype 2: the capsule protects against phagocytosis and is an important virulence factor. Infect Immun 67: 1750-1756.

Sukeno, A., Nagamune, H., Whiley, R. A., Jafar, S. I., Aduse-Opoku, J., Ohkura, K., Maeda, T., Hirota, K., Miyake, Y., and Kourai, H. (2005) Intermedilysin is essential for the invasion of hepatoma HepG2 cells by Streptococcus intermedius. Microbiol Immunol 49: 681-694.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2001) Thermosensitive suicide vectors for gene replacement in Streptococcus suis. Plasmid 46: 140-148.

Tang, J., Wang, C., Feng, Y., Yang, W., Song, H., Chen, Z., Yu, H., Pan, X., Zhou, X., Wang, H., Wu, B., Wang, H., Zhao, H., Lin, Y., Yue, J., Wu, Z., He, X., Gao, F., Khan, A. H., Wang, J., Zhao, G. P., Wang, Y., Wang, X., Chen, Z., and Gao, G. F. (2006) Streptococcal toxic shock syndrome caused by Streptococcus suis serotype 2. PLoS Med 3: e151.

Tenenbaum, T., Adam, R., Eggelnpohler, I., Matalon, D., Seibt, A., GE, K. Novotny, Galla, H. J., and Schroten, H. (2005) Strain-dependent disruption of blood-cerebrospinal fluid barrier by Streptoccocus suis in vitro. FEMS Immunol Med Microbiol 44: 25-34.

Tenenbaum, T., Essmann, F., Adam, R., Seibt, A., Janicke, R. U., Novotny, G. E., Galla, H. J., and Schroten, H. (2006) Cell death, caspase activation, and HMGB1 release of porcine choroid plexus epithelial cells during Streptococcus suis infection in vitro. Brain Res 1100: 1-12.

Trottier, S., Higgins, R., Brochu, G., and Gottschalk, M. (1991) A Case of Human Endocarditis Due to Streptococcus Suis in North-America. Reviews of Infectious Diseases 13: 1251-1252.

Tsuchiya, K., Kawamura, I., Takahashi, A., Nomura, T., Kohda, C., and Mitsuyama, M. (2005) Listeriolysin O-induced membrane permeation mediates persistent interleukin-6 production in Caco-2 cells during Listeria monocytogenes infection in vitro. Infect Immun 73: 3869-3877.

Tweten, R. K. (2005) Cholesterol-dependent cytolysins, a family of versatile pore-forming toxins. Infection and Immunity 73: 6199-6209.

Vadeboncoeur, N., Segura, M., Al Numani, D., Vanier, G., and Gottschalk, M. (2003) Pro-inflammatory cytokine and chemokine release by human brain microvascular endothelial cells stimulated by Streptococcus suis serotype 2. FEMS Immunol Med Microbiol 35: 49-58.

Valentin-Weigand, P., Benkel, P., Rohde, M., and Chhatwal, G. S. (1996) Entry and intracellular survival of group B streptococci in J774 macrophages. Infect Immun 64: 2467-2473.

Vanier, G., Fittipaldi, N., Slater, J. D., Dominguez-Punaro, Mde L., Rycroft, A. N., Segura, M., Maskell, D. J., and Gottschalk, M. (2009) New putative virulence factors of Streptococcus suis involved in invasion of porcine brain microvascular endothelial cells. Microb Pathog 46: 13-20.

Vanier, G., Segura, M., Friedl, P., Lacouture, S., and Gottschalk, M. (2004) Invasion of porcine brain microvascular endothelial cells by Streptococcus suis serotype 2. Infection and Immunity 72: 1441-1449.

Willenborg, J., Fulde, M., de Greeff, A., Rohde, M., Smith, H. E., Valentin-Weigand, P., and Goethe, R. (2011) Role of glucose and CcpA in capsule expression and virulence of Streptococcus suis. Microbiology 157: 1823-1833.

Results, part I Chapter 3

89

Wisselink, H. J., Smith, H. E., Stockhofe-Zurwieden, N., Peperkamp, K., and Vecht, U. (2000) Distribution of capsular types and production of muramidase-released protein (MRP) and extracellular factor (EF) of Streptococcus suis strains isolated from diseased pigs in seven European countries. Veterinary Microbiology 74: 237-248.

Xu, L., Huang, B., Du, H., Zhang, X. C., Xu, J., Li, X., and Rao, Z. (2010) Crystal structure of cytotoxin protein suilysin from Streptococcus suis. Protein Cell 1: 96-105.

Zheng, H., Punaro, M. C., Segura, M., Lachance, C., Rivest, S., Xu, J., Houde, M., and Gottschalk, M. (2011) Toll-like receptor 2 is partially involved in the activation of murine astrocytes by Streptococcus suis, an important zoonotic agent of meningitis. J Neuroimmunol 234: 71-83.

Chapter 4

Results, part II:

Identification of a RGD-motif in suilysin possibly involved in host-cell binding,

Rac1-activation and macropore-formation

M. Seitz, R. Gerhard, R. Goethe and P. Valentin-Weigand

Running title: RGD-motif in suilysin

(Manuscript in preparation)

Results, part II Chapter 4

93

Abstract Suilysin is a cholesterol-dependent pore-forming cytolysin secreted by Streptococcus

suis, an important causative agent of meningitis in swine and human. The role of

suilysin in host cell interaction is still unclear. Here, we showed that suilysin contains

structural elements involved in specific functions of the toxin. Using oligonucleotide

directed mutagenesis we produced recombinant suilysin in which either the highly

conserved tryptophan-rich undecapeptide of the fourth domain or the RGD-motif of

domain one was inactivated. Measurement of haemolytic and cytotoxic activity

revealed that both domains are required for full lytic function of suilysin. In contrast to

the tryptophan-mutant, the RGD-mutant showed also a deficiency in membrane

binding to HEp-2 cells. To further resolve if subsequent cellular signalling in host cells

is affected by amino acid substitution, we used a pull down based GLISA to study the

activation of small Rho-GTPases. As anticipated from our previous studies,

unmodified suilysin and the tryptophan-mutant activated Rac1. In contrast, mutation

of the RGD-motif abolished ability to induce G-proteins. Furthermore, studies with the

fluorescent dye calcein revealed that both mutated derivates of suilysin had lost their

ability to form lytic macropores in host cells. Thus, we suggest that the RGD-motif is

required for membrane binding of suilysin as well as for an outside-in signalling

mediated by the G-protein kinase Rac1 and the induction of lytic macropores in

HEp-2 cells.

Results, part II Chapter 4

94

Introduction Streptococcus (S.) suis is one of the most important swine pathogens worldwide

causing high economical losses in pig husbandry due to invasive diseases

associated with meningitis, arthritis, septicaemia, and bronchopneumonia.

Furthermore, S. suis is a zoonotic bacterium causing meningitis and the life-

threatening streptococcal toxic shock like syndrome in humans (Gottschalk et al.,

2007; Gottschalk et al., 2010; Tang et al., 2006; Trottier et al., 1991).

Suilysin, the secreted haemolysin of S. suis, belongs to the family of cholesterol-

dependent pore-forming cytolysins (CDC) (Jacobs et al., 1994). This large family of

toxins generate complex channels in target cells, causing membrane damage of host

cells and lysis of erythrocytes (Lalonde et al., 2000; Norton et al., 1999; Tenenbaum

et al., 2006; Vanier et al., 2004). The primary and secondary structure of CDC is

highly conserved. They share the same monomeric structure, comprised of four

domains (Rossjohn et al., 1997; Xu et al., 2010). A mechanism of pore-formation has

been proposed by Tweten (2005). Several monomers bind vertically to the

cholesterol-containing membrane of target cells via the fourth domain, followed by an

oligomerisation step resulting in formation of the prepore complex. The second

domain builds a 'hinge-like' region allowing the insertion of the oligomer into the

membrane, whereas the third domain mainly serves as a transmembrane domain,

representing the internal structure of the permeable pore (Czajkowsky et al., 2004;

Ramachandran et al., 2002; Shepard et al., 2000). The fourth domain contains the

tryptophan-rich undecapeptide, which is highly conserved and known to be

responsible for formation of a functional pore (Feldman et al., 1990; Korchev et al.,

1998; Michel et al., 1990; Polekhina et al., 2005). We recently confirmed the

predicted 'pore-forming' function of the undecapeptide for suilysin using a point-

mutated derivate of suilysin W461F (Seitz et al., submitted; cf. chapter 3).

Furthermore, we showed that suilysin promotes invasion of S. suis independent of

the formation of a functional pore. However, the specific molecular mechanism

underlying suilysin-induced activation of host cells is still unknown.

Results, part II Chapter 4

95

The objective of this study was to further characterize possible functional domains of

suilysin to understand the molecular mechanism underlying suilysin-induced

activation of host cells. We identified a RGD (Arg-Gly-Asp)-motif within the sequence

of the first domain of the molecule. This is a special feature of suilysin, since none of

the other well-characterized CDC posses this sequence. Bacterial proteins containing

a RGD-sequence are known to mediate adherence and entry of pathogens into host

cells (Scibelli et al., 2007), most likely by binding host cell surface proteins

recognizing RGD-sequences, known as integrins. Using target mutagenesis we

constructed recombinant suilysin in which the putative integrin-binding side

(RGD-motif) is inactivated by substitution. This abolished cytolytic and haemolytic

function of suilysin. Detailed analysis revealed that the RGD-motif is most likely

required for membrane binding of suilysin, subsequent activation of Rac1 as well as

for the formation of lytic macropores.

Results, part II Chapter 4

96

Results and Discussion

Inactivation of the RGD-motif leads to almost complete loss of haemolytic and cytotoxic activity of suilysin. The purpose of this study was to characterise a

second putative functional region, a RGD-motif identified within the first domain of

suilysin. Hence, we introduced a point-mutation in the sly gene by site-directed

mutagenesis, resulting in amino substitution of RGD by SVD.

Full length suilysin lacking the signal sequence (rSLY) and the point-mutated

derivates (rW461F and rSVD) were overexpressed in E. coli and purified. Silver

straining of respective recombinant proteins showed no difference in size and a high

purity grade, as indicated by a single clear band at approximately 58 kDa (Figure

4-1). To exclude frameshifts, rSLY was sequenced in total. Moreover, both point-

mutations (rW461F and rSVD) were confirmed by sequencing of the respective

regions within the molecule. Alignment of amino acid sequences of suilysin (NCBI

accession # 253753945), sequenced rSLY, in silico analyzed rW461F and rSVD and

four other members of the CDC-family, including pneumolysin (PLY), perfringolysin O

(PFO), intermedilysin (ILY), and listeriolysin O (LLO), is shown in Figure 4-1.

Sequence alignment confirmed the high similarity of the undecapeptide of the fourth

domain, termed tryptophan-rich motif (coloured in light green; blue similarity bar

below sequences indicates 100% homology). The RGD-motif within domain one

(coloured in light red) is shown to be only present within the suilysin molecule. The

substitution of the undecapeptide at position W461 (rW461F, highlighted in white)

and the RGD-sequence (rSVD, highlighted in white), respectively, was confirmed.

Results, part II Chapter 4

97

Results, part II Chapter 4

98

Figure 4-1: Alignment of amino acid sequences of different CDC. Suilysin (NCBI accession # 253753945), sequenced rSLY, in silico analyzed rW461F and rSVD and four other members of the CDC family, including pneumolysin (PLY, # 294652455), perfringolysin O (PFO, # 144884), intermedilysin (ILY, # 6729344) and listeriolysin O (LLO, # 44112) are shown. Subunits of the suilysin molecule are highlighted, including domain 1 (red; RGD-motif in light red), domain 2 (blue), domain 3 (yellow) and domain 4 (green; tryptophan-rich undecapetide in light green). The same colour code was used for the crystal structure of the suilysin molecule modified from (Xu et al., 2010a) (cf. chapter 1). Substitutions of the RGD-motif and W461 of the undecapeptide are marked in white. The signal-sequence of SLY is marked in orange. Similarity bars (light blue) below sequences indicate a similarity of 100% of all analyzed CDC. Alignment was performed with the program Clone Manager 9. The inlay shows a silver staining of recombinant suilysin rSLY, rW461F, and rSVD (~58kDa) separated by SDS-polyacrylamid gel electrophoresis.

We have previously shown that substitution of the tryptophan-rich undecapeptide

resulted in loss of lytic function (Seitz et al., submitted; cf. Chapter 3). Therefore,

haemolytic and cytotoxic activity tests of the mutated suilysin rSVD were carried out

in comparison to rSLY and rW461F. Determination of haemolytic capacity showed

that the point-mutated toxin rSVD almost completely lost its capacity to lyse sheep

erythrocytes. Compared to rSLY, of which 256 (28) ng/ml were sufficient for 50%

haemolysis, rSVD had to applied at 2048 (211) ng/ml and rW461F at 4096 (212) ng/ml

to cause 50% haemolysis (Figure 4-2A). Calculation of haemolytic units revealed

0.53 x 105 HU/mg for rSLY as compared to 0.3 x 104 HU/mg for rSVD and 0.25 x 104

HU/mg for rW461F. In addition, membrane damage of HEp-2 cells by suilysin and its

mutated derivates was determined using a LDH release test (Figure 4-2B). Testing

the specific cytotoxic activity of rSVD revealed a loss of cytotoxic function (2.1% of

the rSLY), similar to the rW461F (8.1% of rSLY).

Results, part II Chapter 4

99

Figure 4-2: Substitution of the RGD-motif of recombinant suilysin leads to abolishment of its haemolytic and cytolytic activity. (A) Determination of haemolytic activity of rSLY, rW461F and rSVD proteins by standard haemolysis assay using sheep erythrocytes. Proteins were added to erythrocytes in microtiter plates at indicated concentrations and haemolysis was determined by measuring the absorbance of cell-free supernatants at 550 nm. Results are expressed as % haemolysis compared to H2O-lysed erythrocytes. The horizontal line indicates 50% haemolysis. The inlay shows representative effects of rSLY (left, haemolysis), rW461F (middle, no haemolysis) and rSVD (right, no haemolysis). (B) Determination of cytotoxic activity of rSLY, rW461F and rSVD proteins by standard LDH release assay using HEp-2 cells. Proteins were added to cells in microtiter plates at indicated concentrations and LDH release as an indicator of cytotoxicity was determined in the supernatants as described in Experimental procedures. Results are expressed as % cytotoxicity compared to Triton X-100-lysed cells. The horizontal line indicates 50% cytotoxicity.

In order to analyze whether the different suilysin derivates conferred a respective

phenotype a heterologous expression system was used. Suilysin and its mutated

derivates, all containing the signal sequence, were transformed into the

gram-positive, non-pathogenic bacterium Lactococcus (L.) lactis. Protein processing

and expression in L. lactis is very similar to S. suis. Due to the signal sequence,

overexpressed proteins are secreted, thus present in the culture supernatant of the

Results, part II Chapter 4

100

respective lactococci. To verify suilysin expression and to determine haemolytic and

cytotoxic activity, culture supernatants of L. lactis pORI23 WT, pORI23-SLY, pORI23-

W461F, and pORI23-SVD obtained form late logarithmic growth phase were tested in

immunoblot analysis and cytotoxic assays. By immunoblot analysis using a

polyclonal antiserum raised against rSLY, we could clearly show that SLY, W461F,

and SVD were present in the supernatant of respective lactococci, except for L. lactis

pORI23 WT carrying the control plasmid (Figure 4-3A, inlay). In accordance to our

findings observed using recombinant proteins (rSLY, W461F, and rSVD), only

L. lactis pORI23-SLY was able to cause lysis of sheep red blood cells (53.3%

haemolysis; Figure 4-3A) and epithelial cells (142.0% cytotoxicity; Figure 4-3B),

respectively. Lactococci expressing either W461F or SVD showed no haemolytic

activity and only low levels of cytotoxicity (W461F: 4.0% and SVD: 37.7%).

Taken together these data indicate that the RGD-motif is required for haemolytic and

cytotoxic activity of suilysin.

Figure 4-3: Heterologous expression of SLY, but not W461F or SVD in L. lactis, confers a haemolytic and cytotoxic phenotype. (A) Titration of haemolytic activity of SLY, W461F, and SVD present in culture supernatants of respective lactococci. Haemoglobin present in the supernatant caused by haemolysis was measured by absorbance at 550 nm. (B) Dose-dependent epithelial cell injury. Cytotoxicity was measured by LDH-release test. L. lactis pORI23 (carrying plasmid control) severed as background control for both tests. Lytic effects of distilled water (haemolysis test) or 1% Triton X-100 in PBS (cytotoxicity test) were set as 100% and depicted as controls.

Results, part II Chapter 4

101

The RGD-motif is required for membrane binding, Rac1-activation and macropore-formation in HEp-2 cells. Next we analysed in more detail, in which

step the RGD-motif might be involved. In general, RGD-containing proteins can act

as connecting molecules between pathogens and host cells to allow adherence and

uptake of pathogens or in order to induce cell signalling pathways (Scibelli et al.,

2005). Therefore, we first analysed the initial membrane binding capacity of suilysin

and its mutated derivates using immunofluorescent staining. HEp-2 cells were

treated with rSLY, rW461F or rSVD at subcytolytic (100 ng/ml; Figure 4-4B-D) or lytic

(400 ng/ml; Figure 4-4E-G) concentration. Cell-bound suilysin was detected using a

polyclonal antiserum raised against rSLY. Unmodified rSLY was able to bind to the

cell surface at subcytolytic and lytic concentration (green spots in Figure 4-4B and E).

The size of individual spots (most likely due to different numbers of molecules or

oligomerisation stages) increased at higher concentration. Likewise, rW461F bound

to cell membranes at lytic concentration (green spots in Figure 4-4G). This

phenotype was described before for homologous tryptophan-mutants of the highly

related pneumolysin as well as for intermedilysin (Polekhina et al., 2005). It has been

shown, that loss of lytic function did not include absence of oligomerisation and

formation of the prepore complex (Boulnois et al., 1991; Korchev et al., 1998;

Polekhina et al., 2005). In contrast, we showed that substitution of the RGD-motif

affected membrane binding capacity drastically, so that rSVD was unable to bind to

HEp-2 cells (Figure 4-4C and F).

Possible interaction partners of RGD-containing proteins are the heterodimeric

transmembrane integrins. The integrin family is subdivided into groups, whereas one

subgroup is specified on recognizing RGD-containing proteins (RGD-receptors), such

as α5β1 integrins (Scibelli et al., 2005; Scibelli et al., 2007). Two mechanisms of

integrin-bacteria interactions are known (Scibelli et al., 2007). An indirect interaction

is mediated by extracellular matrix components (ECM) containing a RGD-sequence,

such as fibronectin or vitronectin. These 'bridging molecules' connect bacterial

ECM-binding proteins with integrins (Cue et al., 2000; Jonsson et al., 1991). An

alternative way is the direct linking of microbial RGD-exhibiting proteins to integrins

(Hamzaoui et al., 2004; Tran et al., 1999). We hypothesized that the latter is

Results, part II Chapter 4

102

plausible for suilysin-integrin interaction. Therefore, we analyzed by flow cytometry, if

α5β1 integrins were expressed on the cellular surface of unstimulated HEp-2 cells.

The results clearly showed that α5β1 integrins were present on HEp-2 cells (Figure

4-4H), thus available as putative interaction partners for suilysin. Further studies will

have to proof whether they in fact function as receptors.

In principle, CDC bind preferentially to cholesterol-rich microdomains (lipid rafts)

(Gekara and Weiss, 2004; Shimada et al., 2002; Waheed et al., 2001), but

membrane components required for CDC binding have not been characterized in

detail. The impact of cholesterol on membrane binding of CDC and CDC function is

still controversially discussed (Hotze and Tweten, 2011). First assumed as the

cellular receptor, recognized by the fourth domain of the CDC monomer (Sekino-

Suzuki et al., 1996), a recent study pointed out that cholesterol plays a role

downstream of membrane binding by triggering the conversion of the prepore

complex to a functional transmembrane pore (Giddings et al., 2003). Nevertheless,

lipid rafts are known to function as a 'platform' for bacteria-host cell interaction,

enhancing bacterial adherence and invasion (Lafont and van der Goot, 2005). It is

proposed that this 'triggering effect' is mediated by clustering of cell surface receptors

within cholesterol enriched microdomains. Receptors associated with lipid rafts are

among other integrins. We have previously demonstrated that cholesterol is essential

for cytolytic activity. Future studies will have to show how cholesterol contributes to

cell-activation by suilysin.

Integrins can also transmit extracellular signals into the cell. The small family of Rho-

GTPases, namely Rac, Rho and Cdc42, play a key role in integrin-mediated outside-

in signalling. Integrin-ligand interaction triggers either a direct activation of GTPases

or regulates GTPase effector proteins. Furthermore, activation of G-proteins in turn

induces integrin-clustering and a higher ligand-affinity. This bidirectional interaction is

known to regulate cytoskeletal rearrangements (Schwartz and Shattil, 2000)

putatively involved in bacteria-cell interaction.

As described in our previous study suilysin promotes invasion of S. suis by a

Rac-dependent activation of the actin cytoskeleton (Seitz et al., submitted; cf. chapter

3). We showed that rSLY specifically activated Rac1, a member of small

Results, part II Chapter 4

103

Rho-GTPases, in HEp-2 cells at subcytolytic concentration (100 ng/ml), thus

suggesting a pore-independent mechanism. Moreover, the mutated suilysin rW461F,

deficient in functional pore-formation, still retained its ability to activate Rac1. To

elucidate, if the RGD-motif is required for activation of Rho signalling G-proteins, we

performed pull down based GLISA analysis. HEp-2 cells were treated with rSVD in

comparison to rSLY and rW461F used at subcytolytic concentration (100 ng/ml) for

15 min. Results revealed an activation of Rac1 after stimulation of HEp-2 cells with

either rSLY or W461F (Figure 4-5A), confirming previous results (Seitz et al.,

submitted; cf. chapter 3). As anticipated, neither RhoA nor Cdc42 were inducible.

Interestingly, in contrast to rSLY and W461F, rSVD failed to activate any

Rho-GTPase (Figure 4-5A), indicating the requirement of the RGD-motif for

suilysin-induced host cell signalling.

Since macropore-formation is the final step in cytolysis, we speculated that the point-

mutations of suilysin might abolish macropore-formation. Therefore, we used the

fluorescent marker calcein to test cell membrane integrity. HEp-2 cells were

incubated with recombinant proteins (rSLY, rW461F or rSVD) at subcytolytic (100

ng/ml, Figure 4-5B) or lytic concentrations (400 ng/ml, Figure 4-5C) to induce large

channel formation in lipid bilayers. As intact cell membranes are impermeable for

calcein due to its size (diameter of 1.3 nm) macropore-formation (20 - 30 nm in size)

is required for incorporation of calcein by cells. As anticipated, rSLY was capable to

induce macropores in epithelial cells at lytic concentrations (Figure 4-5C), resulting in

a respective shift of the curve (cell population) on the x-axis. Both rW461F and rSVD

failed to form large pores. In accordance to these results, a homologous tryptophan-

mutant of pneumolysin formed large pores to significant lower degree and less

frequently compared to wild type pneumolysin analyzed by conductance

measurement of patch clamps (El Rachkidy et al., 2008; Korchev et al., 1998).

Furthermore, it has been found that rapid activation of RhoA and Rac1 in neuronal

cells by pneumolysin preceded the formation of lytic macropores (Iliev et al., 2007).

Likewise, the tryptophan-mutant of suilysin still retained Rac1-activation, but failed to

induce large channels. This suggests that cell-activation occurs before

Results, part II Chapter 4

104

macropore-formation, as previously reported for the mechanism underlying suilysin-

promoted invasion of S. suis (Seitz et al., submitted; cf. chapter 3).

Figure 4-4: Substitution of the RGD-motif of recombinant suilysin leads to a loss in membrane binding. Membrane binding was analyzed by immunofluorescent microscopy. HEp-2 cells were treated with rSLY (B + E), rW461F (C, F) or rSVD (D, G) at subcytolytic (100 ng/ml; B-D) or lytic concentrations (400 ng/ml; E-G). Untreated cells are shown in (A). The actin cytoskeleton was stained in red and suilysin in green. Bar represent 10 µm. (H) Detection of surface associated α5β1 integrins on HEp-2 cells via flow cytometry analysis. Shift of the curve on the x-axis represents fluorescent cells, indicating expression of respective integrin subunit.

Results, part II Chapter 4

105

Figure 4-5 Substitution of the RGD-motif leads to loss in Rac1-activation and macropore-formation. (A) Activation of Rac1, RhoA, and Cdc42 after a 15 min stimulation of HEp-2 cells with rSLY and its mutated derivates rW461F and rSVD at subcytolytic concentrations (100 ng/ml), expressed as percentage of control. (B, C) Flow cytometry analysis of HEp-2 cells treated with the fluorescent marker calcein and recombinant proteins (rSLY, rW461F or rSVD) at subcytolytic (B) or lytic (C) concentrations, respectively. Cells treated with calcein alone served as background control. Right shift of the curve on the x-axis represents fluorescent cells, indicting macropore-formation (cell permeable for calcein).

Concluding, our results showed that the pore-forming region as well as the

RGD-motif are both needed for the cytolytic activity of suilysin. Interestingly,

formation of a functional channel is not required for additional effects of suilysin, such

as activation of morphological host cell response. However, the RGD-motif identified

here seems to be essential for such activities since substitution of the RGD sequence

led to loss of membrane binding and Rac1-activation. It is plausible to speculate that

the RGD-motif of suilysin interacts with α5β1 integrins on the host cell surface to

induce a Rac1-dependent remodelling of the actin cytoskeleton to allow invasion of

S. suis. Further studies will have to be performed to elucidate the precise role of the

RGD-motif in S. suis-host cell interaction and pathogenesis.

Results, part II Chapter 4

106

Experimental procedures

If not stated otherwise all materials were purchased from Sigma (Muenchen,

Germany).

Bacterial strain and growth conditions. Escherichia (E.) coli strains BL21, BL21

(DE3) and DH5α were used for molecular cloning and protein expression

experiments. E. coli were grown in Luria Bertani (LB) medium overnight under

aerobic conditions at 37°C with vigorous shaking. For heterologously expression of

suilysin Lactococcus lactis subsp. cremoris MG1363 (L. lactis) was used as

described recently (Baums et al., 2006).

DNA techniques. Routine molecular biology techniques including restriction

endonuclease digestion, DNA ligations, agarose gel electrophoresis, transformation

of E. coli and plasmid isolation were performed according to standard procedures

(Sambrook et al., 1989). Restriction enzymes were purchased from New England

Biolabs (Frankfurt am Main, Germany). Plasmid preparations were performed with

kits from Machery-Nagel (Dueren, Germany).

Construction and expression of recombinant wild type suilysin and its point-mutated derivates W461F and RGD-SVD. W461F and RGD-SVD substitutions

were constructed using site-directed-mutagenesis according to the instruction

manual of QuikChange® Site-Directed Mutagenesis Kit (Stratagene, Ja Jolla, CA) as

described in a previous study (Seitz et al., submitted; cf. chapter 3). As template DNA

pET45bslynew was used (Kock et al., 2009). For RGD replacement the

oligonucleotide primers encoding R124-G125-D126 to S124-V125-D126

slyRGDMutfor (CAGTATTGCGTCGGTAGATCTGACGCTTAG) and slyRGDMutrev

(CTAAGCGTCAGATCTACCGACGCAATACTG) were used. Mutation sites of each

suilysin derivate were sequenced using the primers slyseqfor

(GGATCATTCAGGTGCTTATG) and pET45seqrev (TGCTGGCGTTCAAATTTCGC).

Recombinant His-tagged suilysin (rSLY) was expressed in E. coli BL21

Results, part II Chapter 4

107

pET45bslynew (Kock et al., 2009). In accordance, recombinant point-mutated

suilysin W461F (rW461F) was expressed in E. coli BL21 pET45bslynewW461F as

described in a previous study (Seitz et al., submitted; cf. chapter 3) and recombinant

RGD-SVD (rSVD) in E. coli BL21 pET45bslynewRGD-SVD. Purification of all

recombinant proteins was verified by immunoblot analysis with anti-Penta His

antibodies (Qiagen, Hilden, Germany) as recommended and silver staining was used

to control the purity grade. The proteins were stored at -20°C. The wild type suilysin

was completely sequenced with standard oligonucleotid primers T7

(TAATACGACTCACTATAGGG) and T7 term (CTAGTTATTGCTCAGCGGT;

Eurofins MWG operon, Ebersberg, Germany).

Heterologous expression of SLY, W461F and SVD in L. lactis. Heterologous

expression in L. lactis was performed as described previously (Baums et al., 2006).

The sly-gene was amplified from chromosomal DNA of Streptococcus suis by PCR

with oligonucleotide primers slyPstI (TAGTCTGCAGCTCCTAGCCTCTGTGGCT AA)

and slyBamHIoptRBS (CAGAGGATCCAGGAGAAAACTTATGAGAAAAAG). After

digestion of the PCR product with BamHI and PstI the fragment containing the whole

sly-gene was cloned into the BamHI/PstI-digested shuttle vector pORI23 (Que et al.,

2000). The resulting plasmid was named pORI23-SLY and transferred in L. lactis as

described previously (Holo and Nes, 1989b). Transformants were screen by plating

on GM17 agar plates containing 5 µg/ml erythromycin. For heterologous expression

of W461F and SVD site-directed mutagenesis was performed as described above

using purified plasmid DNA pORI23sly and the respective oligonucleotid primers

slyTrp-Phefornew and slyTrp-Pherewnew as well as slyRGDMutfor and

slyRGDMutrev. Resulting constructs pORI23-W461F and pORI23-SVD were

electroporated in L. lactis.

Immunoblot analysis. Recombinant proteins were separated by SDS-polyacrylamid

gel electrophoresis with a 4% stacking and a 10% separating gel under denaturing

conditions and transferred to a PVDF-membrane (Serva, Heidelberg, Germany). For

Immunoblot analysis, membrane-blocking was performed overnight with 3% milk

powder in TBS with 0.5% Tween. Polyclonal antiserum raised against rSLY (Benga

Results, part II Chapter 4

108

et al., 2008) diluted 1:1,300 in 1% milk powder was used to detect either rSLY,

rW461F, rSVD or suilysin and its mutated derivates in culture supernatants of the

respective L. lactis strains (L. lactis pORI-SLY, L. lactis pORI-W461F and L. lactis

pORI-SVD). For detection of His-tagged recombinant proteins membranes were

blocked with 3% BSA in TBS with 0.5% Tween overnight and incubated with a

monoclonal anti-Penta His antibody (Qiagen) diluted 1:2,000 in 3% BSA. Membranes

were developed with horseradish peroxidase-conjugated anti-rabbit or anti-mouse

IgG antiserum diluted 1:10000 in 1% milk powder (Amersham, Freiburg, Germany)

followed by addition of SuperSignal® West Pico Chemiluminescent Substrate (Pierce,

Rockford, USA) according to the manufactures protocol.

HEp-2 epithelial cell culture. The human laryngeal epithelial cell line HEp-2 (ATCC

CCL 23) was used. Cells were maintained in Dulbecco’s modified Eagle’s medium

(DMEM, Gibco-Invitrogen, Groningen, The Netherlands) supplemented with 10%

fetal calf serum (FCS) and 5 mM glutamine at 37°C and 8% CO2. The cells were

subcultured every 2-3 days after detachment with 0.25% trypsin and 1 mM Na-EDTA

(trypsin-EDTA, Gibco-Invitrogen). For cellular cytotoxicity assay 0.2 x 105 cells were

seeded on a 96 well plate. For immunofluorescence 0.8 x 105 cells per well were

seeded on a 12 mm diameter glass cover slips placed in 24 well plates. The cells

were grown over-night and then used for experiments.

Determination of haemolytic activity and cytotoxicity. The haemolytic activity of

rSLY, rW461F and rSVD was determined by a haemolysis assay as described in a

previously (Seitz et al., submitted; cf. chapter 3; Takamatsu et al., 2001). Cytotoxicity

of recombinant proteins was detected by an LDH-release assay as described

previously (Seitz et al.; submitted; cf. chapter 3; Benga et al., 2004).

Briefly, L. lactis expressing either SLY, W461F or SVD (L. lactis pORI-SLY, -W461F

and –SVD), respectively, were grown in GM17 medium overnight, adjusted to an

optical density (OD600) of 0.02 in pre-warmed media and then grown to late

exponential growth phase (OD600 0.8). Subsequently bacteria were removed by

centrifugation and culture supernatants were concentrated and purified using Amicon

centrifugal filter devices 30 kDa (Millipore, Schwalbach/Ts., Germany) to remove

Results, part II Chapter 4

109

remaining medium components. Concentrated culture supernatants diluted 1:5 in

0.9% NaCl were used as test samples.

Immunofluorescent microscopy. Membrane binding of rSLY, rW461F and rSVD

was determined by immunofluorescent microscopy. HEp-2 cells were incubated with

the recombinant suilysin derivates at subcytolytic (100 ng/ml) or lytic (400 ng/ml)

concentrations for 30 min at 37°C. Cells were washed three times to remove

unbound suilysin, fixed with 2% parafolmaldehyde and permeabilised with

0.1% Triton X-100 in PBS (Bio-Rad, Muenchen, Germany). Blocking was performed

for 30 min at room temperature with PBS containing 10% FCS. Suilysin was stained

using a polyclonal antiserum raised against rSLY (Benga et al., 2008) diluted 1:1,300

in 1% milk powder for 1 h at room temperature followed by an incubation with

FITC-conjugated goat anti-rabbit antibody (1:1,000 in 1% FCS in PBS, Dianova,

Hamburg, Germany). F-actin was labelled with TRITC-conjugated Phalloidin (20

µg/ml, Invitrogen) for 40 min at room temperature. After final washing DAPI

(Invitrogen) was used for staining the nuclei. Antifading reagent DABCO (Sigma) was

used for sealing of the samples. Mounted samples were examined using inverted

immunofluorescence microscope Nikon Eclipse Ti-S equipped with a 40×, 0.6 S Plan

Fluor objective (Nikon, Duesseldorf, Germany) driven by NIS Elements software BR

3.2..

Detection of α5β1 integrin expression on HEp-2 cells. HEp-2 cell-suspension was

incubated with either a monoclonal anti-human CD29 antibody or a monoclonal anti-

human CD49e (Bio Legend) both used at a concentration of 500 ng per

1.5 x 106 cells (Bio Legend, Fell, Germany). All antibodies were diluted in blocking

buffer (PBS [pH 7.3] containing 3% BSA) and incubation was performed for 30 min at

room temperature with gently agitation. After three washing steps with blocking

buffer, cells were incubated with a TRITC-conjugated goat anti-mouse antibody

(1:1000, Dianova, Hamburg, Germany). Stained cells were washed three times, fixed

with 0.37% formaldehyde and resuspended in blocking buffer for subsequent

measurement. Flow cytometry measurement was performed as described above.

Results, part II Chapter 4

110

Fluorescent cells were counted at channel FL-1 and unstained cells severed as

background control. GLISA. HEp-2 cells were treated with rSLY, rW461F or rSVD at subcytolytic

concentration (100 ng/ml) for 15 min and rinsed twice with ice cold PBS to stop

incubation. Lysis of cells was achieved by addition of 1 ml of ice cold lysis buffer

(50 mM NaCl, 20 mM Tris-HCl, pH 7.4, 3 mM MgCl2, 1% Nonidet-P 40,

0.25% Triton X-100, 5 mM dithiothreitol, 100 µM PMSF). Cells were scraped off, and

the lysates were centrifuged at 14,000 rpm. The supernatant was used for a pull

down-based G-LISA (Cytoskeleton, CO, USA) to determine activation of Rac1,

RhoA, and Cdc42 according to the protocol by the supplier.

Cell-permeability and macropore-formation assay. Experiments were performed

in PBS containing 0.9 mM CaCl2/0.5 mM MgCl26H2O [pH 7.3] (Gibco-Invitrogen). Cell

permeability was determined after co-incubation of HEp-2 cells (4 x 105 cells/ml) with

recombinant proteins (rSLY, rW461F, and rSVD) at subcytolytic (100 ng/ml) or lytic

(400 ng/ml) concentrations (cf. cytotoxicity test) and the fluorescent marker calcein

(2 µg/ml, Sigma, Taufkirchen, Germany). Incubation was performed for 2 h at 37°C

on an end-over-end rotator. After two washing steps with PBS to remove extracellular

calcein, cells were fixed with 0.37% formaldehyde (for 10 min at room temperature

and then resuspended in 500 µl PBS for measurement of fluorescent cells using

FACScan® (Becton Dickinson, 488 nm Argon laser). Further analysis was performed

with the software WinMDI (version 2.9.). For each determination at least

10,000 events were measured. Initial analysis of fluorescent cells was carried out by

dot plot analysis (forward scatter [fsc] versus sideward scatter [ssc]) to define the cell

population of interest (data not shown). Subsequently, fluorescent cells were

detected at channel FL-1. Cells treated only with calcein served as background

control.

Acknowledgements This work was financially supported by the Deutsche Forschungsgemeinschaft

(DFG), Bonn, Germany (SFB 587).

Results, part II Chapter 4

111

References Baums, C. G., Kaim, U., Fulde, M., Ramachandran, G., Goethe, R., and Valentin-Weigand, P. (2006) Identification of a novel virulence determinant with serum opacification activity in Streptococcus suis. Infect Immun 74: 6154-6162.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cellular Microbiology 6: 867-881.

Boulnois, G. J., Paton, J. C., Mitchell, T. J., and Andrew, P. W. (1991) Structure and function of pneumolysin, the multifunctional, thiol-activated toxin of Streptococcus pneumoniae. Mol Microbiol 5: 2611-2616.

Cue, D., Southern, S. O., Southern, P. J., Prabhakar, J., Lorelli, W., Smallheer, J. M., Mousa, S. A., and Cleary, P. P. (2000) A nonpeptide integrin antagonist can inhibit epithelial cell ingestion of Streptococcus pyogenes by blocking formation of integrin alpha 5beta 1-fibronectin-M1 protein complexes. Proc Natl Acad Sci U S A 97: 2858-2863.

Czajkowsky, D. M., Hotze, E. M., Shao, Z., and Tweten, R. K. (2004) Vertical collapse of a cytolysin prepore moves its transmembrane beta-hairpins to the membrane. EMBO J 23: 3206-3215.

El Rachkidy, R. G., Davies, N. W., and Andrew, P. W. (2008) Pneumolysin generates multiple conductance pores in the membrane of nucleated cells. Biochem Biophys Res Commun 368: 786-792.

Feldman, C., Mitchell, T. J., Andrew, P. W., Boulnois, G. J., Read, R. C., Todd, H. C., Cole, P. J., and Wilson, R. (1990) The effect of Streptococcus pneumoniae pneumolysin on human respiratory epithelium in vitro. Microb Pathog 9: 275-284.

Gekara, N. O. and Weiss, S. (2004) Lipid rafts clustering and signalling by listeriolysin O. Biochem Soc Trans 32: 712-714.

Giddings, K. S., Johnson, A. E., and Tweten, R. K. (2003) Redefining cholesterol's role in the mechanism of the cholesterol-dependent cytolysins. Proc Natl Acad Sci U S A 100: 11315- 11320.

Gottschalk, M., Segura, M., and Xu, J. (2007) Streptococcus suis infections in humans: the Chinese experience and the situation in North America. Anim Health Res Rev 8: 29-45.

Gottschalk, M., Xu, J., Calzas, C., and Segura, M. (2010) Streptococcus suis: a new emerging or an old neglected zoonotic pathogen? 25. Future Microbiol 5: 371-391. Hamzaoui, N., Kerneis, S., Caliot, E., and Pringault, E. (2004) Expression and distribution of beta1 integrins in in vitro-induced M cells: implications for Yersinia adhesion to Peyer's patch epithelium. Cell Microbiol 6: 817-828.

Holo, H. and Nes, I. F. (1989) High-Frequency Transformation, by Electroporation, of Lactococcus lactis subsp. cremoris Grown with Glycine in Osmotically Stabilized Media. Appl Environ Microbiol 55: 3119-3123.

Results, part II Chapter 4

112

Hotze, E. M. and Tweten, R. K. (2011) Membrane assembly of the cholesterol-dependent cytolysin pore complex. Biochim Biophys Acta.

Iliev, A. I., Djannatian, J. R., Nau, R., Mitchell, T. J., and Wouters, F. S. (2007) Cholesterol-dependent actin remodeling via RhoA and Rac1 activation by the Streptococcus pneumoniae toxin pneumolysin. Proc Natl Acad Sci U S A 104: 2897-2902.

Jacobs, A. A. C., Loeffen, P. L. W., vandenBerg, A. J. G., and Storm, P. K. (1994) Identification, Purification, and Characterization of A Thiol-Activated Hemolysin (Suilysin) of Streptococcus Suis. Infection and Immunity 62: 1742-1748.

Jonsson, K., Signas, C., Muller, H. P., and Lindberg, M. (1991) Two different genes encode fibronectin binding proteins in Staphylococcus aureus. The complete nucleotide sequence and characterization of the second gene. Eur J Biochem 202: 1041-1048.

Kock, C., Beineke, A., Seitz, M., Ganter, M., Waldmann, K. H., Valentin-Weigand, P., and Baums, C. G. (2009) Intranasal immunization with a live Streptococcus suis isogenic ofs mutant elicited suilysin-neutralization titers but failed to induce opsonizing antibodies and protection. Veterinary Immunology and Immunopathology 132: 135-145.

Korchev, Y. E., Bashford, C. L., Pederzolli, C., Pasternak, C. A., Morgan, P. J., Andrew, P. W., and Mitchell, T. J. (1998) A conserved tryptophan in pneumolysin is a determinant of the characteristics of channels formed by pneumolysin in cells and planar lipid bilayers. Biochem J 329 ( Pt 3): 571-577.

Lafont, F. and van der Goot, F. G. (2005) Bacterial invasion via lipid rafts. Cell Microbiol 7: 613-620.

Lalonde, M., Segura, M., Lacouture, S., and Gottschalk, M. (2000) Interactions between Streptococcus suis serotype 2 and different epithelial cell lines. Microbiology 146 ( Pt 8): 1913- 1921.

Michel, E., Reich, K. A., Favier, R., Berche, P., and Cossart, P. (1990) Attenuated mutants of the intracellular bacterium Listeria monocytogenes obtained by single amino acid substitutions in listeriolysin O. Mol Microbiol 4: 2167-2178.

Norton, P. M., Rolph, C., Ward, P. N., Bentley, R. W., and Leigh, J. A. (1999) Epithelial invasion and cell lysis by virulent strains of Streptococcus suis is enhanced by the presence of suilysin. Fems Immunology and Medical Microbiology 26: 25-35.

Polekhina, G., Giddings, K. S., Tweten, R. K., and Parker, M. W. (2005) Insights into the action of the superfamily of cholesterol-dependent cytolysins from studies of intermedilysin. Proc Natl Acad Sci U S A 102: 600-605.

Que, Y. A., Haefliger, J. A., Francioli, P., and Moreillon, P. (2000) Expression of Staphylococcus aureus clumping factor A in Lactococcus lactis subsp. cremoris using a new shuttle vector. Infect Immun 68: 3516-3522.

Ramachandran, R., Heuck, A. P., Tweten, R. K., and Johnson, A. E. (2002) Structural insights into the membrane-anchoring mechanism of a cholesterol-dependent cytolysin. Nat Struct Biol 9: 823- 827.

Rossjohn, J., Feil, S. C., McKinstry, W. J., Tweten, R. K., and Parker, M. W. (1997) Structure of a cholesterol-binding, thiol-activated cytolysin and a model of its membrane form. Cell 89: 685- 692.

Results, part II Chapter 4

113

Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N Y.

Schwartz, M. A. and Shattil, S. J. (2000) Signaling networks linking integrins and rho family GTPases. Trends Biochem Sci 25: 388-391.

Scibelli, A., Matteoli, G., Roperto, S., Alimenti, E., Dipineto, L., Pavone, L. M., Della, Morte R., Menna, L. F., Fioretti, A., and Staiano, N. (2005) Flavoridin inhibits Yersinia enterocolitica uptake into fibronectin-adherent HeLa cells. FEMS Microbiol Lett 247: 51-57.

Scibelli, A., Roperto, S., Manna, L., Pavone, L. M., Tafuri, S., Della, Morte R., and Staiano, N. (2007) Engagement of integrins as a cellular route of invasion by bacterial pathogens. Vet J 173: 482- 491.

Seitz, M., Baums, CG, Gerhard, R, Just, I, Neis, C., Benga, L., Fulde, M., Nerlich, A, Rohde, M., Goethe, R., and Valentin-Weigand, P. Subcytolytic activity of suilysin promotes invasion of Streptococcus suis in HEp-2 epithelial cells by Rac-dependent activation of the actin cytoskeleton. Ref Type: submitted to Cellular Microbiology Sekino-Suzuki, N., Nakamura, M., Mitsui, K. I., and Ohno-Iwashita, Y. (1996) Contribution of individual tryptophan residues to the structure and activity of theta-toxin (perfringolysin O), a cholesterol- binding cytolysin. Eur J Biochem 241: 941-947.

Shepard, L. A., Shatursky, O., Johnson, A. E., and Tweten, R. K. (2000) The mechanism of pore assembly for a cholesterol-dependent cytolysin: formation of a large prepore complex precedes the insertion of the transmembrane beta-hairpins. Biochemistry 39: 10284-10293.

Shimada, Y., Maruya, M., Iwashita, S., and Ohno-Iwashita, Y. (2002) The C-terminal domain of perfringolysin O is an essential cholesterol-binding unit targeting to cholesterol-rich microdomains. Eur J Biochem 269: 6195-6203.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2001) Thermosensitive suicide vectors for gene replacement in Streptococcus suis. Plasmid 46: 140-148.

Tang, J., Wang, C., Feng, Y., Yang, W., Song, H., Chen, Z., Yu, H., Pan, X., Zhou, X., Wang, H., Wu, B., Wang, H., Zhao, H., Lin, Y., Yue, J., Wu, Z., He, X., Gao, F., Khan, A. H., Wang, J., Zhao, G. P., Wang, Y., Wang, X., Chen, Z., and Gao, G. F. (2006) Streptococcal toxic shock syndrome caused by Streptococcus suis serotype 2. PLoS Med 3: e151.

Tenenbaum, T., Essmann, F., Adam, R., Seibt, A., Janicke, R. U., Novotny, G. E., Galla, H. J., and Schroten, H. (2006) Cell death, caspase activation, and HMGB1 release of porcine choroid plexus epithelial cells during Streptococcus suis infection in vitro. Brain Res 1100: 1-12.

Tran, Van Nhieu, Caron, E., Hall, A., and Sansonetti, P. J. (1999) IpaC induces actin polymerization and filopodia formation during Shigella entry into epithelial cells. EMBO J 18: 3249-3262.

Trottier, S., Higgins, R., Brochu, G., and Gottschalk, M. (1991) A Case of Human Endocarditis Due to Streptococcus Suis in North-America. Reviews of Infectious Diseases 13: 1251-1252.

Tweten, R. K. (2005) Cholesterol-dependent cytolysins, a family of versatile pore-forming toxins. Infection and Immunity 73: 6199-6209.

Vanier, G., Segura, M., Friedl, P., Lacouture, S., and Gottschalk, M. (2004) Invasion of porcine brain microvascular endothelial cells by Streptococcus suis serotype 2. Infection and Immunity 72: 1441-1449.

Results, part II Chapter 4

114

Waheed, A. A., Shimada, Y., Heijnen, H. F., Nakamura, M., Inomata, M., Hayashi, M., Iwashita, S., Slot, J. W., and Ohno-Iwashita, Y. (2001) Selective binding of perfringolysin O derivative to cholesterol-rich membrane microdomains (rafts). Proc Natl Acad Sci U S A 98: 4926-4931.

Xu, L., Huang, B., Du, H., Zhang, X. C., Xu, J., Li, X., and Rao, Z. (2010a) Crystal structure of cytotoxin protein suilysin from Streptococcus suis. Protein Cell 1: 96-105.

Chapter 5

Results, part III:

Establishment of an intranasal CD1 mouse infection model for colonization and invasion of Streptococcus suis serotype 2

M. Seitz, A. Beineke, J. Seele, M. Fulde, P. Valentin-Weigand and C. G. Baums

Running title: Intranasal mouse infection model for S. suis

(Manuscript in preparation)

Results, part III Chapter 5

117

Abstract Streptococcus (S.) suis causes meningitis and various other diseases in pigs and

humans. Healthy piglets carrying virulent S. suis strains on their mucosal surfaces

are epidemiologically very important. The objective of this study was to establish an

intranasal mouse model for colonisation and invasion of the respiratory tract. CD1

mice were intranasally infected with a highly virulent S. suis serotype 2 strain under

different conditions of predisposition. Clinical, histological and bacteriological

examination revealed that invasion of host tissues occurred only in mice predisposed

intranasally with 1% acetic acid. This model is characterized by efficient colonization

as all mice carried S. suis on their respiratory mucosa 7 days post infection.

Furthermore, severe fibrinosuppurative or purulent necrotizing pneumonia associated

with S. suis was a common manifestation in this model. The intranasal S. suis model

was applied to investigate the relevance of suilysin in colonisation and invasion by

comparison of the wild type with its isogenic sly-mutant. Results suggested

attenuation in virulence but not in colonization of the suilysin mutant. In conclusion,

this study revealed the first intranasal mouse model to study colonization and

invasion of the respiratory tract by a highly virulent S. suis pathotype.

Results, part III Chapter 5

118

Introduction Streptococcus (S.) suis is a major swine pathogen worldwide, causing severe

diseases such as meningitis, septicaemia and bronchopneumonia (Higgins and

Gottschalk, 2005). It is also an important zoonotic agent. Humans might be infected

following contact with pigs or pork. Meningitis, septicaemia and the life-threatening

streptococcal toxic shock-like syndrome (STSS) are important manifestations of

S. suis infections in humans (Gottschalk et al., 2007; Tang et al., 2006).

Pigs and wild boars are considered the natural reservoir of S. suis (Baums et al.,

2007; Clifton-Hadley and Alexander, 1980; Higgins and Gottschalk, 2005). Different

mucosal surfaces might be colonized by S. suis. In weaning piglets, S. suis is among

other bacteria the most abundant colonizers of the upper respiratory and alimentary

tract (Baele et al., 2001; Lowe et al., 2011; O'Sullivan et al., 2011a; Su et al., 2008).

Healthy carriers of virulent S. suis strains play an important role in the epidemiology

of S. suis diseases in pig and humans (Arends et al., 1984; Ngo et al., 2011).

S. suis is characterized by a high diversity as reflected by the presence of at least 33

serotypes. Serotype 2 is worldwide the most prevalent among invasive isolates of

pigs and humans (Wei et al., 2009; Wisselink et al., 2000). The serotype is

determined by the polysaccharide capsule. The capsule protects the bacteria against

opsonophagocytosis and functions as an important virulence factor (Charland et al.,

1998; Smith et al., 1999). A number of other surface-associated factors have also

been demonstrated to contribute to pathogenicity of S. suis (Baums and Valentin-

Weigand, 2009). The pathogenesis of S. suis meningitis is not well understood, and

even less is known about the mechanisms employed by S. suis to colonize mucosa.

As a matter of fact, not a single factor of S. suis has been demonstrated to be crucial

for colonization.

Recently, a S. suis meningitis model was described in CD1 mouse, in which typical

histopathological lesions and inflammatory responses were recorded. Sudden death

of 20% of infected animals associated with high levels of systemic proinflammatory

cytokines was observed. Infected mice that survived developed clinical signs of

meningitis characterized by infiltrates of neutrophils and bacterial emboli

Results, part III Chapter 5

119

(Dominguez-Punaro et al., 2007). Similar to many other experimental infections of

mice, S. suis was applied intraperitoneally. As the upper respiratory tract and, in

particular, the tonsils are considered to be the port of entry for S. suis (Williams and

Blakemore, 1990), early steps in the pathogenesis of S. suis diseases cannot be

studied in the CD1 mouse model described by Dominguez-Punaro et al. (2007).

Thus, investigation of colonization of the respiratory mucosa requires a new murine

intranasal S. suis model. Here, we describe for the first time that S. suis serotype 2

colonizes and invades mice tissue efficiently after intranasal application following

predisposition. This work includes a description of an intranasal colonization and

invasion model for S. suis using CD1 mice. Furthermore, the model was applied to

identify suilysin-dependent phenotypes in vivo using encapsulated as well as

unencapsulated strains.

Results, part III Chapter 5

120

Results and Discussion To our knowledge, a mucosal infection model for S. suis in mice has not been

described. In this study we investigated whether a virulent S. suis serotype 2 strain

can infect CD1 mice after intranasal application. As healthy carrier animals play a

crucial role in the epidemiology of S. suis diseases, there is also a urgent need to

study colonization of mucosal surfaces. A major problem in swine is the restricted

availability of S. suis free piglets for such investigations. Therefore, the second

objective of this study was to introduce a mice model for S. suis colonization of

mucosal surfaces. For this, we compared different conditions of experimental

infection without and after predisposition with 1% acetic acid prior intranasal

infection.

Intranasal infection without predisposition. A dose of 5 x 109 CFU of S. suis strain

10 was applied intranasally to Crl:CD1 (ICR) mice, as this mouse strain has been

demonstrated to develop typical lesions of meningitis after intraperitoneal application

of S. suis (Dominguez-Punaro et al., 2007). Nine of 10 Crl:CD1 (ICR) mice did not

show any clinical signs of infection over an observation period of 5 or 12 days

(5 mice each, Table 5-2). The low morbidity in mice infected intranasally with a high

dose of S. suis serotype 2 is in accordance with the findings of Williams et al. (1988).

Apathy, continuing anorexia and a weight loss of more than 20% was registered in

one mouse starting on the 4th dpi (Table 5-2; Table 5-S1). Histophathological

screening revealed severe purulent meningitis and encephalitis associated with a

bacterial load of 4,000 CFU per mg brain tissue (Table 5-3 and Table 5-4).

Furthermore, the challenge strain was also reisolated in 3 of 10 mice in the TNL and

in 2 of 10 mice in the lungs on the 5th day post intranasal infection. The bacterial

load of the respiratory tract of these mice was below 100 CFU per μl TNL or mg

tissue, respectively. These results suggest that S. suis serotype 2 colonized the

respiratory tract and even caused meningitis, though only with low morbidity.

Results, part III Chapter 5

121

Intranasal infection model after predisposition with acetic acid. In piglets, S. suis

mucosal infection models have been described which include experimental

predisposition, such as infection with Bordetella bronchiseptica (Smith et al., 1996;

Smith et al., 1999) or local application of 1% acetic acid (Baums et al., 2006; Pallares

et al., 2003). Therefore we predisposed mice by nasal application of 12.5 µl 1%

acetic acid per nostril 1 h pre intranasal infectionem with S. suis. In this experiment,

morbidity was significantly higher in comparison to the previous experiment without

predisposition (P = 0.003; Table 5-2). Sixty-seven percent of mice received a

cumulative clinical index equal or above 3 and were, thus, classified as diseased

(Table 5-1 and Table 5-2). Six (Table 5-S1) out of 24 infected mice developed severe

clinical symptoms (anorexia, scrubby coat, rapid abdominal breathing, and apathy)

and were euthanized one day post infection for reasons of animal welfare (mortality

rate of 25%). Severe fibrinosuppurative or purulent necrotizing pneumonia as well as

severe infiltration of the splenic red pulp with neutrophils was diagnosed in these six

mice post mortem. S. suis was detected in the lungs of these six mice in high number

(up to 23,000 CFU per mg lung tissue; Table 5-4). Purulent meningitis and/or

encephalitis was found only in one of 24 infected mice. However, the challenge strain

was reisolated from the brain in 46% of these animals. The bacterial load in the brain

of these mice was below 150 CFU per mg brain tissue except in one mouse

(34,000 CFU per mg brain tissue). Severe fibrinosuppurative and even purulent

necrotizing rhinitis was a common finding in surviving mice. The challenge strain was

frequently detected in the lungs but also in other inner organs (Table 5-4).

Unexpectedly, in eight of 24 mice St. aureus was also detected in inner organs (see

number in brackets in Table 5-4 and Table 5-S1). A likely explanation for this finding

was that St. aureus colonizing the respiratory tract in mice might have invaded the

host after local application of 1% acetic acid. However, infection with St. aureus was

not a prerequisite for S. suis infection, since more than 50% of mice positive for

S. suis in the brain, lung, heart or any other organ were free of St. aureus in these

tissues. In this experiment, mice were sacrificed on the 3rd, 5th and 7th days post

infection to monitor the time course of early colonization and invasion as shown in

Table 5-4. The challenge strain was isolated from the upper respiratory tract (TNL

Results, part III Chapter 5

122

was positive tested for the challenge strain in 92% of all infected mice) and inner

organs from numerous mice on the 5th day post infection, but in mice sacrificed on

the 7th day detection of S. suis was restricted to the TNL (with the exception of one

positive lung; Table 5-S1). In conclusion, intranasal application of 1% acetic acid 1 h

pre infection significantly increased morbidity, mainly due to severe pneumonia

associated with S. suis. Invasion of inner organs by S. suis was also enhanced in

mice after acetic acid predisposition. Furthermore, S. suis colonized the respiratory

epithelium during the whole period of observation (up to seven days post

infectionem).

To exclude coinfection with commensal St. aureus, further experiments were

conducted with the St. aureus-free CD1 mice strain Hsd:ICR (CD-1®). The challenge

strain was detectable in almost all inner organs of six mice on the 3rd day post

intranasal infection (Table 5-S1). Severe pneumonia was a common finding as in the

2nd experiment (Figure 5-1B). Accordingly, the pathoscore ω of experiments 2 and 3

were very similar (2.0 and 1.9, respectively). Whereas moderate to severe purulent

necrotizing or fibrinosuppurative rhinitis was detectable in all infected animals (Figure

5-1A), meningitis and encephalitis were not observed in any of the infected mice. The

challenge strain colonized the respiratory epithelium most efficiently as indicated by a

reisolation rate of 100% of the challenge strain from TNL. As expected St. aureus

coinfection was not a problem in this experiment (Table 5-4).

Figure 5-1 Histological findings of the nose and lung of mice infected intranasally with 5 x 109 CFU of S. suis after predisposition with 1% acetic acid (A) Severe multifocal fibrinosuppurative rhinitis, HE 20x. (B) Severe multifocal fibrinous-necrotizing (arrow) pneumonia, HE 20x.

Results, part III Chapter 5

123

The results indicated that application of acetic acid prior to experimental S. suis

infection predisposed mice to rhinitis and pneumonia rather than to meningitis, the

most common pathology of S. suis infection in swine and humans. This is in contrast

to the effect of intranasal acetic acid predisposition in swine, leading mainly to

S. suis-associated pleuritis, peritonitis and meningitis (Baums et al., 2006; Pallares et

al., 2003) and the intraperitoneal model described by Dominguez-Punaro et al.

(2007) associated with a prevalence of meningitis in 40% of mice. However, as

demonstrated by bacteriology results, S. suis colonized the mucosal epithelium up to

seven days post infectionem and invaded into tissue after intranasal application and

acetic acid predisposition. Therefore, this model might be used to study colonization

and invasion of S. suis in mice, using bacteriological screening of different inner

organs as important read out parameter.

Table 5-1: Clinical score sheet for health monitoring of mice after infection with S. suis. Parameter 0 1 2

Body weight Constant or gain > 5% weight los > 20% weight loss

Coat Flat, glossy Rough, reduced grooming Scrubby, failure of grooming

Breathing Adequate, rhythmic Rapid, shallow Rapid, abdominal

Dehydrationa Normal skin elasticity

Moderately reduced skin elasticity Persisting skin fold, sunken eyes

Bearing Normal Moderately curved back Cowering, highly curved back

Eyes Normal Moderately protruding Highly swollen

Activity Normal active Depression Apathy, social isolation

Locomotion No anomaly Moderate incoordination Apraxia, stagger

a Elapsing of a drawn up dorsal skin fold

Results, part III Chapter 5

124

Results, part III Chapter 5

125

Results, part III Chapter 5

126

Results, part III Chapter 5

127

Evaluation of colonization and invasiveness of isogenic sly-mutants. The

intranasal mouse model with acetic acid predisposition was used to investigate the

contribution of suilysin to the invasiveness of S. suis. Six mice each were infected

either with the wild type strain 10 or its isogenic suilysin-deficient mutant (10Δsly).

None of the mice infected with 10Δsly developed clinical symptoms except for a

weight loss >5% one day post infectionem. In comparison, all mice infected with the

wild type strain 10 showed mild clinical symptoms of general sickness including

rough coat, moderate depression and persistent weight loss >5% (morbidity 100%).

No case of severe sickness with obvious neural dysfunction was seen (Table 5-5).

However, histopathological screening revealed one case of severe purulent

encephalitis located in the olfactory bulb in a mouse infected with the wild type strain

10 (Table 5-S1). In accordance with differences in morbidity between wild type and

10Δsly infected mice, the general pathoscore ω was substantially lower in mice

infected with the isogenic sly-mutant in comparison to the score of the wild type

infected group (0.5 versus 1.7, Table 5-6). Attenuation in virulence of the isogenic

sly-mutant is in accordance with results of experimental mouse infections reported by

Allen et al. (2001), who observed substantial suilysin-dependent differences in

mortality at high intraperitoneal doses. However, experimental infections of pigs with

suilysin-negative and wild type strains did not reveal an attenuated phenotype of the

sly-mutant (Allen et al., 2001; Lun et al., 2003).

In contrast to the two previous experiments with acetic acid predisposition cases of

pneumonia were not observed in wild type infected mice (Table 5-6). Furthermore,

the wild type strain was detectable only in the TNL but not in inner organs (Table

5-7). These differences might be related to differences in ages: Mice of experiments

2 and 3 were 4 weeks old at the time of infection, whereas mice of the 4th

experiment were 6 to 7 weeks old. Northworthy, Williams et al. (1988) also observed

that young mice were more susceptible to S. suis infection than old mice.

Results, part III Chapter 5

128

Results, part III Chapter 5

129

Results, part III Chapter 5

130

Results, part III Chapter 5

131

In conclusion, comparative evaluation of virulence of the wild type and the isogenic

sly-mutant in the intranasal infection model suggested attenuation of the sly-mutant.

Further experiments are necessary to clear why the wild type challenge strain was

not detectable in inner organs and why pathohistological lesions of inner organs were

relatively low, except for one case of severe encephalitis.

In vitro experiments with the respiratory epithelial cell line HEp-2 revealed a suilysin-

dependent invasive phenotype of the isogenic unencapsulated mutant 10cpsΔEF

(Seitz et al., submitted; cf. chapter 3). It is, however, not clear what role suilysin-

dependent invasion of host cells might play for virulence. Unencapsulated S. suis

strains are generally regarded as avirulent because of enhanced

opsonophagocytosis (Charland et al., 1998; Smith et al., 1999). However, Smith et al.

(1999) isolated the isogenic unencapsulated mutants 10cpsΔEF and 10cpsΔB from

CNS, serosae and joints after intranasal infections of gnotobiotic piglets, though

these piglets did not develop clinical signs of disease. Furthermore, unencapsulated

non-typeable stains have been isolated from diseased pigs (Silva et al., 2006).

Therefore, we used the intranasal mouse model to further investigate whether the

unencapsulated strain 10cpsΔEF also showed a suilysin-dependent invasive

phenotype by comparison with the double mutant 10cpsΔEFΔsly. In contrast to the

experiment with the encapsulated strains, mice were sacrificed on the 3rd and not on

the 5th day post infection to increase the probability of reisolation. However, neither

the unencapsulated mutant 10cpsΔEF nor the isogenic suilysin double mutant

(10cpsΔEFΔsly) was detectable in inner organs except for one reisolation for

10cpsΔEF from lung tissue (Table 5-6). Surprisingly, 50% morbidity (mild clinical

symptoms) were observed in mice infected with 10cpsΔEF, of which one mouse

(Table 5-S1) developed severe clinical symptoms (rapid abdominal breathing,

kyphosis, and showed a weight loss >20%; Table 5-5) associated with a mild

histiocytic interstitial pneumonia. The type of the latter lesion is atypical for S. suis

infections. In comparison, only one out of six mice infected with 10cpsΔEFΔsly

developed mild clinical symptoms, like shallow breathing and rough coat (16.7%

morbidity). Though the comparison of morbidity suggested a further suilysin-

dependent attenuation of the 10cpsΔEFΔsly double mutant, results have to be

Results, part III Chapter 5

132

interpreted with caution as the challenge strains were undetectable in inner organs.

As discussed for the wild type strain, further experiments with younger animals might

reveal different results.

The intranasal infection of mice with all different S. suis strains after acetic acid

predisposition resulted in severe rhinitis in at least 50% of each group (Table 5-6). To

avoid possible non-specific histopathological findings of the nose, rhinitis was

excluded from the general pathoscore ω and, therefore, differences in ω between

different strains are not due to rhinitis lesions (Table 5-5 and Table 5-6).

In all 4 groups at least 50% of mice were positive in the TNL for the respective

challenge strain. Quantification of bacteria obtained from TNL revealed no significant

difference between mice infected either with S. suis wild type strain 10 or its isogenic

sly-negative mutant strain 10Δsly (Table 5-7). Comparably, infection with either

10cpsΔEF or 10cpsΔEFΔsly showed no clear difference of the recovered bacterial

load of the upper respiratory tract, suggesting no deficiency in colonization of the

sly-negative strains 10Δsly and 10cpsΔEFΔsly.

In conclusion, an intranasal murine model for S. suis colonization and invasion was

established in this work. Severe purulent necrotizing pneumonia and rhinitis, rather

than meningitis, were common findings among infected mice. Furthermore,

preliminary results suggested attenuation in virulence of an isogenic sly-mutant. But

a suilysin-dependent phenotype with regard to early colonization was not recorded in

these experiments. However, this murine model is suitable to study invasion and

colonization of mucosal surfaces in mice in the future.

Results, part III Chapter 5

133

Material and Methods Bacterial strains and culture conditions. Streptococcus (S.) suis serotype 2 wild

type strain 10 was kindly provided by H. Smith (Lelystad, NL). This strain expresses

EF, MRP, SLY, FBPS and OFS. It has been used by different groups successfully for

mutagenesis and experimental intranasal infections of pigs (Baums et al., 2006;

Smith et al., 1999; Vecht et al., 1997). Strain 10cpsΔEF has been generated by

insertional mutagenesis of the genes cps2E and cps2F involved in biosynthesis of

the capsule and has been demonstrated to be severely attenuated in virulence, most

likely due to increased opsonophagocytosis (Smith et al., 1999). The corresponding

suilysin deficient mutants of wild type strain 10 (designed 10Δsly) and 10cpsΔEF

(designed 10cpsΔEFΔsly) were constructed as described previously (Benga et al.,

2008; Seitz et al., submitted; cf. chapter 3). Streptococci were grown on Columbia

agar supplemented with 7% sheep blood (Oxoid, Wesel, Germany) overnight under

aerobic conditions at 37°C.

Preparation of infection culture. For infection of mice, streptococci were grown in

Todd-Hewitt broth (THB, Difco, Detroit, USA) overnight at 37°C under aerobic

conditions, adjusted to an optical density (OD600) of 0.02 in pre-warmed media the

next day and then grown to late exponential growth phase (OD600 0.8). Streptococci

were harvested by centrifugation, resuspended in sterile PBS (pH 7.4) and adjusted

to the final concentration of 5 x 1011 CFU/ml for intranasal infection. Inoculum

concentrations were verified by platting 10-fold serial dilutions on Colombia agar

plates with 7% sheep blood after infection.

Intranasal infection of mice. Four (expt. 1 to 4) or six (expt. 5) week old specific-

pathogen-free female mice of the outbreed strain Crl:CD1 (ICR) were obtained from

Charles River Laboratories (Sulzfeld, Germany). Mice of the Staphylococcus (St.)

aureus free outbred strain Hsd:ICR (CD1®) were purchased from Harlan Laboratories

(AN Venray, The Netherlands). Animals were randomly divided into groups

consisting of five to six animals each. Mice were allowed to acclimate for one week

Results, part III Chapter 5

134

and cared for in accordance with the principles outlined in the European Convention

for the Protection of Vertebrate Animal Used for Experimental and Other Scientific

Purposes [European Treaty Series, no. 123:

http://conventions.coe.int/treaty/EN/Menuprincipal.htm; permit no.33.9-42502-04-

08/1589]. Before infection mice were anesthetized via inhalation of isofluran (IsoFlo®,

Albrecht, Germany). In experiment 2 to 5 mice were pre-treated with 12.5 µl 1%

acetic acid [pH 4.0] placed in each nostril 1 h prior intranasal infection. After a

controlled recovery phase and re-anaesthesia per isofluran inhalation, mice were

infected with 5 x 109 CFU of either S. suis wild type strain 10, 10Δsly, 10cpsΔEF or

10cpsΔEFΔsly. The dose was applied in two drops of 12.5 µl volume placed in front

of the nostrils.

Clinical score. Animals were clinically scored every 8 h. The health status was rated

using a clinical score sheet (Table 5-1), including weight development, clinical signs

of general sickness (rough coat, rapid breathing, dehydration), clinical signs

indicating meningitis (apathy, apraxia), and septicaemia (swollen eyes, depression),

developed on the basis of mouse clinical monitoring score described by the

Research office of the Australian National University

[http://www.anu.edu.au/ro/ORI/Animal/ AEEC001_MouseMonitoringSOP.doc]. A

cumulative score of 3 to 4 indicated mild clinical symptoms, a score of 5 to 6

moderate clinical symptoms and a score greater than 6 severe clinical symptoms, in

particular persistent anorexia, apathy, and/or neural disorder. Mice with a cumulative

score equal or greater than 3 were classified as diseased (calculation of morbidity).

In the case of severe weight loss (> 20%) and/or enduring severe clinical signs, mice

were euthanized for reasons of animal welfare by inhalation of CO2 and cervical

dislocation.

Histological screening. Immediately after euthanasia, necropsy was conducted and

the organs were aseptically removed and split for histological and bacteriological

screenings, including spleen, liver, kidney, heart, lung, brain, spinal cord, and nose.

The histological screenings were carried out as blind experiments. Findings were

Results, part III Chapter 5

135

scored as described for piglets (Baums et al., 2006). In contrast to piglets, in addition

to fibrinosuppurative lesions, purulent necrotizing lesions associated with the

challenge strain were scored as well (see results). Rhinitis was not included in the

general score ω designed to reflect lesions caused by invading streptococci.

Reisolation of S. suis strains from tissue and tracheo-nasal lavage (TNL). One

half of each organ was suspended in 5 ml PBS [pH 7.4] and weighed. All organs

were homogenized with an Ultra Turrax (IKA, Staufen, Germany). Ten-fold serial

dilutions of samples were plated on blood agar plates. Colony forming units (CFU)

were counted the next day after incubation at 37°C for 24 h and CFU per mg organ

was determined.

For sampling TNL the trachea was opened and a retrograde irrigation of the nasal

cavity with 300 µl PBS was collected. Number of typical α-haemolytic streptococci

per µl TNL was determined by serial plating on blood agar. Isolated α-haemolytic

streptococci were investigated in a S. suis multiplex PCR for the detection of mrp,

epf, sly, arcA, gdh, cps1, cps2, cps7, and cps9 (Silva et al., 2006). Isolates received

from mice challenged either with strain 10Δsly, 10cpsΔEF or 10cpsΔsly were

additionally tested in a cps2E-specific PCR (with oligonucleotide primers cps2Efor

(TTTCGCACTTTCAAGACGTG) and cps2Erev (GGACGGGTACCGACTAGACTC)

and sly-specific PCR using primers slyAgefor (TGTACCGGTGATTCCAAACAAGA

TATTAA) and slyAge3new (TTAACCGGTTACTCTATCACCTCATCCG). Based on

CFU per mg organ or per µl TNL, bacterial loads were classified as mild (+; <100),

moderate (++; >100 - <1000) or severe (+++; >1000).

Results, part III Chapter 5

136

Acknowledgements We thank Hilde Smith (DLO-Institute for Animal Science and Health, The

Netherlands) for providing strain 10 and the non-encapsulated isogenic mutant

10cpsΔEF. Oliver Goldmann (Helmholtz Centre for Infection Research,

Braunschweig, Germany) kindly introduced us to intranasal application techniques

and anaesthesia of mice.

This study was financially supported by the Deutsche Forschungsgemeinschaft

(DFG), Bonn, Germany (SFB587).

Results, part III Chapter 5

137

References Allen, A. G., Bolitho, S., Lindsay, H., Khan, S., Bryant, C., Norton, P., Ward, P., Leigh, J., Morgan, J., Riches, H., Eastty, S., and Maskell, D. (2001) Generation and characterization of a defined mutant of Streptococcus suis lacking suilysin. Infect Immun 69: 2732-2735.

Arends, J. P., Hartwig, N., Rudolphy, M., and Zanen, H. C. (1984) Carrier Rate of Streptococcus Suis Capsular Type-2 in Palatine Tonsils of Slaughtered Pigs. Journal of Clinical Microbiology 20: 945-947.

Baele, M., Chiers, K., Devriese, L. A., Smith, H. E., Wisselink, H. J., Vaneechoutte, M., and Haesebrouck, F. (2001) The gram-positive tonsillar and nasal flora of piglets before and after weaning. J Appl Microbiol 91: 997-1003.

Baums, C. G., Kaim, U., Fulde, M., Ramachandran, G., Goethe, R., and Valentin-Weigand, P. (2006) Identification of a novel virulence determinant with serum opacification activity in Streptococcus suis. Infect Immun 74: 6154-6162.

Baums, C. G. and Valentin-Weigand, P. (2009) Surface-associated and secreted factors of Streptococcus suis in epidemiology, pathogenesis and vaccine development. Anim Health Res Rev 10: 65-83.

Baums, C. G., Verkuhlen, G. J., Rehm, T., Silva, L. M., Beyerbach, M., Pohlmeyer, K., and Valentin- Weigand, P. (2007) Prevalence of Streptococcus suis genotypes in wild boars of Northwestern Germany. Appl Environ Microbiol 73: 711-717.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Charland, N., Harel, J., Kobisch, M., Lacasse, S., and Gottschalk, M. (1998) Streptococcus suis serotype 2 mutants deficient in capsular expression 9. Microbiology-Sgm 144: 325-332.

Clifton-Hadley, F. A. and Alexander, T. J. (12-7-1980) The carrier site and carrier rate of Streptococcus suis type II in pigs. Vet Rec 107: 40-41.

Dominguez-Punaro, M. C., Segura, M., Plante, M. M., Lacouture, S., Rivest, S., and Gottschalk, M. (2007) Streptococcus suis serotype 2, an important swine and human pathogen, induces strong systemic and cerebral inflammatory responses in a mouse model of infection. J Immunol 179: 1842-1854.

Gottschalk, M., Segura, M., and Xu, J. (2007) Streptococcus suis infections in humans: the Chinese experience and the situation in North America. Anim Health Res Rev 8: 29-45.

Higgins,R. and Gottschalk,M. (2005) Streptococcal diseases. In Diseases of Swine. pp. 769-783.

Lowe, B. A., Marsh, T. L., Isaacs-Cosgrove, N., Kirkwood, R. N., Kiupel, M., and Mulks, M. H. (2011) Microbial communities in the tonsils of healthy pigs. Vet Microbiol 147: 346-357.

Lun, S. C., Perez-Casal, J., Connor, W., and Willson, P. J. (2003) Role of suilysin in pathogenesis of Streptococcus suis capsular serotype 2. Microbial Pathogenesis 34: 27-37.

Results, part III Chapter 5

138

Ngo, T. H., Tran, T. B., Tran, T. T., Nguyen, V. D., Campbell, J., Pham, H. A., Huynh, H. T., Nguyen, V. V., Bryant, J. E., Tran, T. H., Farrar, J., and Schultsz, C. (2011) Slaughterhouse pigs are a major reservoir of Streptococcus suis serotype 2 capable of causing human infection in southern Vietnam. PLoS One 6: e17943.

O'Sullivan, T., Friendship, R., Blackwell, T., Pearl, D., McEwen, B., Carman, S., Slavic, D., and Dewey, C. (2011) Microbiological identification and analysis of swine tonsils collected from carcasses at slaughter. Can J Vet Res 75: 106-111.

Pallares, F. J., Halbur, P. G., Schmitt, C. S., Roth, J. A., Opriessnig, T., Thomas, P. J., Kinyon, J. M., Murphy, D., Frank, D. E., and Hoffman, L. J. (2003) Comparison of experimental models for Streptococcus suis infection of conventional pigs. Can J Vet Res 67: 225-228.

Seitz, M., Baums, CG, Gerhard, R, Just, I, Neis, C., Benga, L., Fulde, M., Nerlich, A, Rohde, M., Goethe, R., and Valentin-Weigand, P. Subcytolytic activity of suilysin promotes invasion of Streptococcus suis in HEp-2 epithelial cells by Rac-dependent activation of the actin cytoskeleton. Ref Type: submitted to Cellular Microbiology Silva, L. M., Baums, C. G., Rehm, T., Wisselink, H. J., Goethe, R., and Valentin-Weigand, P. (2006) Virulence-associated gene profiling of Streptococcus suis isolates by PCR. Vet Microbiol 115: 117-127.

Smith, H. E., Damman, M., van der Velde, J., Wagenaar, F., Wisselink, H. J., Stockhofe-Zurwieden, N., and Smits, M. A. (1999) Identification and characterization of the cps locus of Streptococcus suis serotype 2: the capsule protects against phagocytosis and is an important virulence factor. Infection and Immunity 67: 1750-1756.

Smith, H. E., Vecht, U., Wisselink, H. J., StockhofeZurwieden, N., Biermann, Y., and Smits, M. A. (1996) Mutants of Streptococcus suis types 1 and 2 impaired in expression of muramidase- released protein and extracellular protein induce disease in newborn germfree pigs. Infection and Immunity 64: 4409-4412.

Su, Y., Yao, W., Perez-Gutierrez, O. N., Smidt, H., and Zhu, W. Y. (2008) Changes in abundance of Lactobacillus spp. and Streptococcus suis in the stomach, jejunum and ileum of piglets after weaning. FEMS Microbiol Ecol 66: 546-555.

Tang, J., Wang, C., Feng, Y., Yang, W., Song, H., Chen, Z., Yu, H., Pan, X., Zhou, X., Wang, H., Wu, B., Wang, H., Zhao, H., Lin, Y., Yue, J., Wu, Z., He, X., Gao, F., Khan, A. H., Wang, J., Zhao, G. P., Wang, Y., Wang, X., Chen, Z., and Gao, G. F. (2006) Streptococcal toxic shock syndrome caused by Streptococcus suis serotype 2. PLoS Med 3: e151.

Vecht, U., StockhofeZurwieden, N., Tetenburg, B. J., Wisselink, H. J., and Smith, H. E. (1997) Virulence of Streptococcus suis type 2 for mice and pigs appeared host-specific. Veterinary Microbiology 58: 53-60.

Wei, Z., Li, R., Zhang, A., He, H., Hua, Y., Xia, J., Cai, X., Chen, H., and Jin, M. (2009) Characterization of Streptococcus suis isolates from the diseased pigs in China between 2003 and 2007. Vet Microbiol 137: 196-201.

Williams, A. E. and Blakemore, W. F. (1990) Pathogenesis of meningitis caused by Streptococcus suis type 2. J Infect Dis 162: 474-481.

Williams, A. E., Blakemore, W. F., and Alexander, T. J. (1988) A murine model of Streptococcus suis type 2 meningitis in the pig. Res Vet Sci 45: 394-399.

Results, part III Chapter 5

139

Wisselink, H. J., Smith, H. E., Stockhofe-Zurwieden, N., Peperkamp, K., and Vecht, U. (2000) Distribution of capsular types and production of muramidase-released protein (MRP) and extracellular factor (EF) of Streptococcus suis strains isolated from diseased pigs in seven European countries. Vet Microbiol 74: 237-248.

Results, part III Chapter 5

140

Table 5-S1: Register of all infected mice including the clinical course of infection, scoring of fibrinosuppurative and purulent necrotizing lesions and reisolation of the challenge strain (table includes intravenously infected mice).

Maximum weight loss (%)

Mouse # Mouse strain

Challenge strain Appa Acetic

acidb CFUc dpid Morbiditye Mortality Severe clinical

symptomsf >5 >20

1 A1E1 6 (0) - - - - 2 A2E1 6 (0) - - - - 3 A3E1 6 (0) - - - - 4 A4E1 6 (0) - - - - 5 A5E1

C57BL/6NCrl

S. suis strain 10 i. n. - 6 x 106

6 (0) - - - -

6 B1E2 5 (0) - - - - 7 B2E2 5 (1) - - + - 8 B3E2 5 6 - - - + 9 B4E2 5 (0) - - - -

10 B5E2 5 (0) - - - - 11 A1E2 12 (1) - - + - 12 A2E2 12 (0) - - - - 13 A3E2 12 (0) - - - - 14 A4E2 12 (0) - - - - 15 A5E2

Crl:CD1 (ICR)

S. suis strain 10 i. n. - 5 x 109

12 (0) - - - -

16 A5E3

1 8 + + + - 17 A6E3

1 8 + + + -

18 B5E3

1 7 + + + - 19 C2E3

1 7 + + + -

20 C5E3

1 8 + + + - 21 D5E3

1 9 + + + -

22 C4E3 3 4 - - + - 23 D1E3 3 3 - - + - 24 D2E3 3 4 - - + - 25 D3E3 3 3 - - + - 26 D4E3 3 4 - - + - 27 D6E3 3 5 - - + - 28 B3E3 5 3 - - + - 29 B4E3 5 3 - - + - 30 B6E3 5 (2) - - + - 31 C1E3

5 3 - - + -

32 C3E3

5 9 - + + - 33 C6E3 5 (2) - - + - 34 A1E3 7 (1) - - + - 35 A2E3

Crl:CD1 (ICR)

S. suis strain 10 i. n. 12.5 µl 5 x 109

7 (1) - - + -

Results, part III Chapter 5

141

Maximum weight loss (%)

Mouse # Mouse strain

Challenge strain Appa Acetic

acidb CFUc dpid Morbiditye Mortality Severe clinical

symptomsf >5 >20

36 A3E3 7 (1) - - + - 37 A4E3 7 (0) - - - - 38 B1E3 7 (1) - - - - 39 B2E3

Crl:CD1 (ICR)

S. suis strain 10 i. n. 12.5 µl 5 x 109

7 (2) - - + -

40 A3E4 2 7 + + + - 41 A5E4

Hsd:ICR (CD-1®)

S. suis strain 10 i. n. 12.5 µl 5 x 109

2 8 + + - + 42 A1E4 3 (2) - - + - 43 A2E4 3 3 - - + - 44 A4E4 3 (2) - - + - 45 B1E4 3 (1) - - - - 46 B2E4 3 4 - - + - 47 B3E4 3 (2) - - + - 48 B4E4 3 (2) - - + - 49 B5E4

Hsd:ICR (CD-1®)

S. suis strain 10 i. n. 12.5 µl 5 x 109

3 (0) - - - -

50 C1E4 7 (2) - - + - 51 C2E4 7 (0) - - - - 52 C3E4 7 3 - - - - 53 C4E4 7 (1) - - - - 54 C5E4 7 (0) - - - - 55 D1E4 7 (0) - - - - 56 D2E4 7 (1) - - - - 57 D3E4 7 (0) - - - - 58 D4E4 7 (0) - - - - 59 D5E4

Hsd:ICR (CD-1®)

S. suis strain 10 i. n. 5.0 µl 5 x 109

7 (1) - - + -

60 B1E1 6 (0) - - - - 61 B2E1 6 (0) - - - - 62 B3E1 6 (0) - - - - 63 B4E1 6 (0) - - - - 64 B5E1

C57BL/6NCrl

S. suis strain 10 i. v. - 6 x 105

6 (0) - - - -

65 C1E1 10 (0) - - - - 66 C2E1 6 7 + + + - 67 C3E1 10 (0) - - - - 68 C4E1 10 (0) - - - - 69 C5E1

C57BL/6NCrl

S. suis strain 10 i. v. - 1 x 108

3 7 + + + -

Results, part III Chapter 5

142

Maximum weight loss (%)

Mouse # Mouse strain

Challenge strain Appa Acetic

acidb CFUc dpid Morbiditye Mortality Severe clinical

symptomsf >5 >20

70 C1E2 9 (0) - - - - 71 C2E2 9 (0) - - - - 72 C3E2 9 (0) - - - - 73 C4E2 9 (0) - - - - 74 C5E2

C57BL/6NClr

S. suis strain 10 i. v. - 1 x 108

9 (1) - - + -

75 D1E2 7 (1) - - - - 76 D2E2 7 (1) - - + - 77 D3E2 7 (1) - - + - 78 D4E2 7 6 - - + - 79 D5E2

C57BL/6NClr

S. suis strain 10 i. v. - 5 x 108

7 (1) - - + -

80 B2E5 2 8 + + + - 81 B3E5 2 9 + + + - 82 B1E5 15 (2) - - + - 83 B4E5 15 (2) - - + - 84 B5E5 15 (1) - - + - 85 B6E5

C57BL/6NClr

S. suis strain 10 i. v. - 1 x 109

15 (1) - - + -

86 A1E5 5 9 + + + - 87 A2E5 15 4.5 - - + - 88 A3E5 15 4 - - + - 89 A4E5 15 (2.5) - - - - 90 A5E5 15 (2) - - - - 91 A6E5

Crl:CD1 (ICR)

S. suis strain 10 i. v. - 1 x 109

15 (2) - - - -

92 C1E6 5 5.5 - - + - 93 C2E6 5 4 - - + - 94 C3E6 5 4 - - + - 95 C4E6 5 4 - - + - 96 C5E6 5 5 - - + - 97 C6E6

Hsd:ICR (CD-1®)

S. suis strain 10 i. n. 12.5 µl 5 x 109

5 4 - - + -

98 D1E6 5 (1.5) - - + - 99 D2E6 5 (1.5) - - + -

100 D3E6 5 (1) - - + - 101 D4E6 5 (2) - - + - 102 D5E6 5 (2) - - + - 103 D6E6

Hsd:ICR (CD-1®)

S. suis 10Δsly i. n. 12.5 µl 5 x 109

5 (2) - - + -

Results, part III Chapter 5

143

Maximum weight loss (%)

Mouse # Mouse strain

Challenge strain Appa Acetic

acidb CFUc dpid Morbiditye Mortality Severe clinical

symptomsf >5 >20

104 A1E6 3 (2) - - - - 105 A2E6 3 3 - - + - 106 A3E6 3 3 - - + - 107 A4E6 3 9 - + - + 108 A5E6 3 (2) - - - - 109 A6E6

Hsd:ICR (CD-1®)

S. suis 10cpsΔE

F

i. n.

12.5 µl

5 x 109

3 (2.5) - - - -

110 B1E6 3 3 - - + - 111 B2E6 3 (1) - - + - 112 B3E6 3 (0) - - - - 113 B4E6 3 (2) - - + - 114 B5E6 3 (0) - - - - 115 B6E6

Hsd:ICR (CD-1®)

S. suis 10cpsΔE

FΔsly

i. n.

12.5 µl

5 x 109

3 (1) - - + -

a Intranasal (i. n.) or intravenous (i. v.) application of bacterial inoculum b Volume of 1% acetic acid applied to each nostril 1 h pre infectionem c Infection dose

d Days post infection, on which mice were sacrificed or killed for reasons of animal welfare e Mice with a cumulative clinical score greater >3 were regarded as diseased f In particular persistent anorexia, apathy, and/or neural disorder leading to a cumulative clinical score >6

Results, part III Chapter 5

144

Nose Spleen and Liver Kidney and

Heart Lung Brain and Spinal cord

Rhinitis Splenitish Hepatitis Nephritis,

Myocarditis Pneumonia Meningitis, Encephalitis, Ependymitis, Ventriculitis

Mouse #

4i 2j 1k 4i 2j 1k 4i 2j 1k 4i 2j 1k 5i 3j 1k 1 A1E1 - - + - - - - - - - - - - - - 2 A2E1 - - + - - - - - - - - - - - - 3 A3E1 - - + - - - - - - - - - - - - 4 A4E1 - - - - - - - - - - - - - - - 5 A5E1 - - - - - - - - - - - - - - -

6 B1E2 - - - - - - - - - - - - - - - 7 B2E2 - - - - - - - - - - - + - - - 8 B3E2 - + - + - - - - - + - - + - - 9 B4E2 - - - - - - - - - - - - - - -

10 B5E2 - - - - - - - - - - - - - - - 11 A1E2 - - - - - - - - - - - + - - - 12 A2E2 - - - - - - - - - - - - - - - 13 A3E2 - - - - - - - - - - - - - - - 14 A4E2 - - - - - - - - - - - - - - - 15 A5E2 - - - - - - - - - - - - - - -

16 A5E3g + - - + - - - - - + - - - - - 17 A6E3g + - - - - - - - - + - - - - - 18 B5E3g + - - + - - - - - + - - - - - 19 C2E3g + - - + - - - - - + - - - - - 20 C5E3g + - - + - - - - - + - - - - - 21 D5E3g + - - - + - - - - - + - - - - 22 C4E3 + - - + - - - - - + - - - - - 23 D1E3 + - - - + - - - - + - - - - - 24 D2E3 - + - - + - - - - - - - - - - 25 D3E3 + - - - - - - - - - - - - - - 26 D4E3 + - - - + - - - - - - - - - - 27 D6E3 + - - + - - - - - - - - - - - 28 B3E3 + - - - - - - - - - - - - - - 29 B4E3 - + - - + - - - - - - - - - - 30 B6E3 - - - - - - - - - - - - - - - 31 C1E3g - - - - - - - - - - - - - - - 32 C3E3g + - - + - - - - - - - - + - - 33 C6E3 + - - - + - - - - - - - - - - 34 A1E3 - - - - - - - - - - - - - - - 35 A2E3 - - + - - - - - - - - - - - -

Results, part III Chapter 5

145

Nose Spleen and Liver Kidney and

Heart Lung Brain and Spinal cord

Rhinitis Splenitish Hepatitis Nephritis,

Myocarditis Pneumonia Meningitis, Encephalitis, Ependymitis, Ventriculitis

Mouse #

4i 2j 1k 4i 2j 1k 4i 2j 1k 4i 2j 1k 5i 3j 1k 36 A3E3 - - - - + - - - - - - - - - - 37 A4E3 - - - - - - - - - - - - - - - 38 B1E3 - - - - - - - - - - - - - - - 39 B2E3 + - - - - - - - - - - - - - -

40 A3E4 + - - - - + - - - - - - - - - 41 A5E4 + - - - - - - - - - - - - - - 42 A1E4 + - - - + - - - - + - - - - - 43 A2E4 - + - - - - - - - - - - - - - 44 A4E4 + - - - - + - - - - - - - - - 45 B1E4 + - - - - - - - - - - - - - - 46 B2E4 + - - - - + - - - + - - - - - 47 B3E4 + - - - - + - - - + - - - - - 48 B4E4 - + - - - + - - - - - - - - - 49 B5E4 - + - - - - - - - + - - - - -

50 C1E4 - - - - - + - - - - - - - - - 51 C2E4 - - - - - + - - - - - - - - - 52 C3E4 - + - - - - - - - - - - - - - 53 C4E4 - - + - - - - - - - - - - - - 54 C5E4 - - - - - - - - - - - - - - - 55 D1E4 - - - - - - - - - - - - - - - 56 D2E4 + - - - - - - - - - - - - - - 57 D3E4 - - - - - - - - - - - - - - - 58 D4E4 - - - - - - - - - - - - - - - 59 D5E4 - - - - - - - - - - - - - - -

60 B1E1 - - - - - - - - - - - - 61 B2E1 - - - - - - - - - - - - 62 B3E1 - - - - - - - - - - - - 63 B4E1 - - - - - - - - - - - - 64 B5E1 - - - - - - - - - + - -

65 C1E1 - - - - - - - - - - - - 66 C2E1 - - - + - - - - - - + - 67 C3E1 - - - - - - - - - - - - 68 C4E1 - - - - - - - - - - - - 69 C5E1 + - - - - - + - - + - -

Results, part III Chapter 5

146

Nose Spleen and Liver Kidney and

Heart Lung Brain and Spinal cord

Rhinitis Splenitish Hepatitis Nephritis,

Myocarditis Pneumonia Meningitis, Encephalitis, Ependymitis, Ventriculitis

Mouse #

4i 2j 1k 4i 2j 1k 4i 2j 1k 4i 2j 1k 5i 3j 1k 70 C1E2 - - - - - - - - - - - - 71 C2E2 - - - - - - - - - - - - 72 C3E2 - - - + - - - - - - - - 73 C4E2 - - - + - - - - - - - - 74 C5E2 - - - - - - - - - - - -

75 D1E2 - - - - - - - - - - - - 76 D2E2 - - - - - - - - - - - - 77 D3E2 - - - - - - - - - - - - 78 D4E2 - + - - - - - + - - + - 79 D5E2 - - - - - - - + - - - -

80 B2E5 - - - - + - - - - + - - 81 B3E5 +l - - + - - - - - + - - 82 B1E5 - - - - - - - - - - - - 83 B4E5 - - - - - - - - - - - - 84 B5E5 - - - - - - - - - - - - 85 B6E5 - - - - - - - - - + - -

86 A1E5 - - - - - - - - - + - - 87 A2E5 - - - - - - + - - - - - 88 A3E5 - - + - - - - - - - - - 89 A4E5 + + - - - - - - - - - - 90 A5E5 - - - - - - - - - - - - 91 A6E5 - - + - - - - - - - - -

92 C1E6 + - - - - + - - - - - - - - - 93 C2E6 + - - - - + - - - - - - - - - 94 C3E6 + - - - - + - - - - - - + - - 95 C4E6 + - - - - + - - - - - - - - + 96 C5E6 - + - - - + - - - - - - - - - 97 C6E6 + - - - - + - - - - - - - - -

98 D1E6 + - - - - - - - - - - - - - - 99 D2E6 + - - - - + - - - - - - - - -

100 D3E6 - - - - - - - - - - - - - - - 101 D4E6 - - + - - - - - - - - - - - - 102 D5E6 - + - - + - - - - - - - - - - 103 D6E6 + - - - - - - - - - - - - - -

Results, part III Chapter 5

147

Nose Spleen and Liver Kidney and

Heart Lung Brain and Spinal cord

Rhinitis Splenitish Hepatitis Nephritis,

Myocarditis Pneumonia Meningitis, Encephalitis, Ependymitis, Ventriculitis

Mouse #

4i 2j 1k 4i 2j 1k 4i 2j 1k 4i 2j 1k 5i 3j 1k 104 A1E6 - - - - - - - - - - - - - - - 105 A2E6 - - + - - - - - - - - - - - - 106 A3E6 + - - - - - - - - - - - - - - 107 A4E6 + - - - + - - - - - +m - - - - 108 A5E6 - + - - - - - - - - - - - - - 109 A6E6 + - - + - - - - - - - - - - -

110 B1E6 + - - - + - - - - - - - - - - 111 B2E6 - + - - - + - - - - - - - - - 112 B3E6 - - - - - - - - - - - - - - - 113 B4E6 - + - - - + - - - - - - - - - 114 B5E6 + - - - + - - - - - - - - - - 115 B6E6 + - - - - + - - - - - - - - -

g Coinfection with St. aureus (isolation of S. suis and St. aureus from different affected tissues) h Infiltration of splenic red pulp with neutrophilic granulocytes i Scoring of 4 and 5 indicates moderate to severe diffuse or multifocal fibrinosuppurative or purulent necrotizing

inflammations of the indicated tissue j Scoring of 2 and 3 indicates mild focal fibrinosuppurative or purulent necrotizing inflammation of the respective organs

k Individual single perivascular immune cells received a score of 1 l Severe necrotizing hepatitis associated with the challenge strain m Mild histiocytic interstitial pneumonia associated with the challenge strain

Results, part III Chapter 5

148

Organ (CFU/mg)n

Mouse # TNLn, o (CFU/µl)n

Spleen Liver Kidney Heart Lung Brain

1 A1E1 - - - - - - - - - - - - - - 2 A2E1 - - - - - - - - - - - - - - 3 A3E1 - - - - - - - - - - - - - - 4 A4E1 - - - - - - - - - - - - - - 5 A5E1 - - - - - - - - - - - - - -

6 B1E2 0.5 + - - - - - - - - - - - - 7 B2E2 - - - - - - - - - - 1.2 + - - 8 B3E2 - - - - - - - - - - 3.3 + 3979.6 +++ 9 B4E2 - - - - - - - - - - - - - -

10 B5E2 2.4 + - - - - - - - - - - - - 11 A1E2 1.4 + - - - - - - - - - - - - 12 A2E2 - - - - - - - - - - - - - - 13 A3E2 - - - - - - - - - - - - - - 14 A4E2 - - - - - - - - - - - - - - 15 A5E2 - - - - - - - - - - - - - -

16 A5E3g 10.0 + 23.3 + 7.7g + 9.7 + 9.0 + 2139.6

+++ 8.6 g +

17 A6E3g 10.0 + 10.0 + 2.2 + 4.9 + 11.8g + 22727.3

+++ 26.8g + 18 B5E3g 10.0 + - - 20.7 + - - 67.8 + 562.6g +++ 11.1g + 19 C2E3g 10.0 + 2.3 + 29.3g + - - 2.0 + 1207.3 +++ 8.2 + 20 C5E3g 10.0 + - - - - - - 104.2 ++ 1320.1

+++ 146.7g ++

21 D5E3g 10.0 + 16.0 + 27.5g + 47.5g + 45.2g + 207.0g ++ 13.5g + 22 C4E3 10.0 + 35.6 + 2.0 + 1.8 + 8.3 + 9.1 + 0.8 + 23 D1E3 6.7 + 6.4 + 0.4 + - - 4.0 + - - - - 24 D2E3 6.7 + - - - - - - - - - - - - 25 D3E3 4.5 + - - 0.3 + - - - - 0.9 + - - 26 D4E3 13.3 + - - - - - - - - - - 10.1 + 27 D6E3 12.6 + 2.5 + - - - - - - 4.9 + - - 28 B3E3 136.7 ++ - - - - 20.8 + 0.7 + - - - - 29 B4E3 380.0 ++ 5.8 + 4.5 + 5.8 + - - - - 1.6 + 30 B6E3 0.3 + - - - - - - - - 0.7 + - - 31 C1E3g 420.0 ++ - - - - - - - - 1.8g + - - 32 C3E3g - - 1.0 + 2.3g + 1385.0 +++ - - 216.0 ++ 33735.

9 +++

33 C6E3 4.3 + - - - - - - - - - - 0.6 + 34 A1E3 186.7 ++ - - - - - - - - 1.4 + - - 35 A2E3 310.0 ++ - - - - - - - - - - - -

Results, part III Chapter 5

149

Organ (CFU/mg)n

Mouse # TNLn, o (CFU/µl)n

Spleen Liver Kidney Heart Lung Brain

36 A3E3 53.3 + - - - - - - - - - - - - 37 A4E3 3.1 + - - - - - - - - - - - - 38 B1E3 - - - - - - - - - - - - - - 39 B2E3 0.7 + - - - - - - - - - - - -

40 A3E4 16.0 + 6.5 + 25.7 + 0.8 + 2.5 + 4.0 + 0.5 + 41 A5E4 2753.3 +++ 373.7 ++ 69.7 + 35.8 + 849.2 ++ 85.3 + 3.5 + 42 A1E4 56.7 + 16.7 + 1.0 + 1.9 + 24.2 + 20.3 + - - 43 A2E4 0.2 + 12.8 + 5.2 + 3.8 + 17.0 + 6.1 + - - 44 A4E4 2.3 + - - - - - - 2.1 + - - - - 45 B1E4 533.3 ++ - - - - - - - - - - - - 46 B2E4 1680.0 +++ 463.0 ++ 381.0 ++ 106.4 ++ - - 25.4 + 2.4 + 47 B3E4 16.0 + 54.7 + 10.2 + 10.2 + 8.4 + 202.3 ++ 0.6 + 48 B4E4 56.7 + - - - - - - - - - - - - 49 B5E4 120.0 ++ - - - - - - - - - - - -

50 C1E4 24.0 + - - - - - - - - - - - - 51 C2E4 2.7 + - - - - - - - - - - - - 52 C3E4 6.3 + - - - - - - - - 3.5 + - - 53 C4E4 173.3 ++ - - - - - - - - 64.6 + - - 54 C5E4 566.7 ++ - - - - - - - - 49.5 + - - 55 D1E4 3.9 + - - - - - - - - - - - - 56 D2E4 433.3 ++ - - - - - - - - - - - - 57 D3E4 2056.7 +++ - - - - - - - - 318.5 ++ - - 58 D4E4 0.9 + - - - - - - - - 14.3 + - - 59 D5E4 116.7 ++ - - - - - - - - 5.8 + - -

60 B1E1 - - - - - - - - - - - - 61 B2E1 - - - - - - - - - - - - 62 B3E1 - - - - - - - - - - - - 63 B4E1 - - - - - - - - - - - - 64 B5E1 - - - - - - - - - - - -

65 C1E1 - - - - - - - - - - - - 66 C2E1 +/-p +/- +/-p +/- +/-p +/- +/-p +/- +/-p +/- +/-p +/- 67 C3E1 - - <100 + - - - - - - - - 68 C4E1 - - <100 + ~500 ++ <100 + - - - - 69 C5E1 - - <100 + <100 + <100 + - - >1000 +++

Results, part III Chapter 5

150

Organ (CFU/mg)n

Mouse # TNLn, o (CFU/µl)n

Spleen Liver Kidney Heart Lung Brain

70 C1E2 2.1 + 0.2 + 1412.3 +++ 88.9 + - - - - 71 C2E2 - - - - - - - - - - - - 72 C3E2 - - - - - - - - - - - - 73 C4E2 - - - - 1.6 + - - - - - - 74 C5E2 - - - - - - - - - - - -

75 D1E2 - - - - - - - - - - - - 76 D2E2 - - - - - - - - - - - - 77 D3E2 - - - - - - 1.5 + 4.9 + 361.1 ++ 78 D4E2 4.5 + 5.2 + 7.8 + - - - - - - 79 D5E2 - - - - - - - - - - - -

80 B2E5 37.4 + 27.4 + 555.6 ++ 1052.1 ++

1867.8 +++ 1278.3 +++

81 B3E5 >1000p +++ >1000p +++ >1000p +++ >1000p ++

>1000p +++ >1000p +++ 82 B1E5 - - - - - - - - - - - - 83 B4E5 - - - - - - - - - - - - 84 B5E5 - - - - - - - - - - - - 85 B6E5 - - - - - - - - - - - -

86 A1E5 - - 0.6 + - + 2.2 + - + 5.0 + 87 A2E5 - - - - - - - - - - - - 88 A3E5 38.6 + 5.6 + 9.2 + 19.0 + 17.5 + - - 89 A4E5 - - - - - - - - - - - - 90 A5E5 - - - - - - - - - - - - 91 A6E5 - - - - - - - - - - - -

92 C1E6 71.7 + - - - - - - - - - - - - 93 C2E6 6.3 + - - - - - - - - - - - - 94 C3E6 373.3 ++ - - - - - - - - - - - - 95 C4E6 436.7 ++ - - - - - - - - - - - - 96 C5E6 18.3 + - - - - - - - - - - - - 97 C6E6 19.3 + - - - - - - - - - - - -

98 D1E6 50.0 + - - - - - - - - - - - - 99 D2E6 300.0 ++ - - - - - - - - - - - -

100 D3E6 376.7 ++ - - - - - - - - - - - - 101 D4E6 183.3 ++ - - - - - - - - - - - - 102 D5E6 1050.0 +++ - - - - - - - - - - - - 103 D6E6 226.7 ++ - - - - - - - - - - - -

Results, part III Chapter 5

151

Organ (CFU/mg)n

Mouse # TNLn, o (CFU/µl)n

Spleen Liver Kidney Heart Lung Brain

104 A1E6 - - - - - - - - - - - - - - 105 A2E6 - - - - - - - - - - - - - - 106 A3E6 - - - - - - - - - - - - - - 107 A4E6 0.1 + - - - - - - - - 2.4 + - - 108 A5E6 0.1 + - - - - - - - - - - - - 109 A6E6 0.1 + - - - - - - - - - - - -

110 B1E6 - - - - - - - - - - - - - - 111 B2E6 0.2 + - - - - - - - - - - - - 112 B3E6 0.3 + - - - - - - - - - - - - 113 B4E6 18.0 + - - - - - - - - - - - - 114 B5E6 1.1 + - - - - - - - - - - - - 115 B6E6 - - - - - - - - - - - - - -

n Isolation of the challenge strain was confirmed in a multiplex PCR for detection of mrp, sly, epf, arcA, cps1, cps2, cps7,

and cps 9 (Silva et al., 2006) a cps2E specific PCR, and a sly specific PCR; Based on CFU per mg organ or per µl TNL, bacterial loads were classified as mild (+ <100), moderate (100 ≤ ++ <1000) or severe (+++ >1000).

o Tracheo-nasal lavage; p Departed; estimation of bacterial load (<8 h after death) or determination not applicable

Chapter 6

General discussion

General discussion Chapter 6

155

General discussion S. suis is considered as a major problem worldwide due to high economical losses in

swine husbandry. To reduce infection threat preventive antimicrobial medication and

early treatment with antibiotics in the case of clinical signs are the only treatment

options. Moreover, most currently available vaccines only provide homologous

protection. Therefore, understanding of pathogenesis and identification of bacterial

virulence (-associated) factors involved in colonisation, persistence, invasion and

immune evasion are urgently needed for better control and prevention measures.

Suilysin, the haemolysin of S. suis, has been identified many years ago (Jacobs et

al., 1994), but is still controversially discussed as a virulence factor of S. suis. In vivo

experiments using suilysin knock-out mutants demonstrated attenuation in virulence

in mice, but only slight attenuation was seen in piglets (Allen et al., 2001; Lun et al.,

2003). Furthermore, immunisation studies with a vaccine containing purified suilysin

completely protected mice against challenge with a highly virulent S. suis serotype 2

strain, but failed to fully protected piglets (Jacobs et al., 1994; Jacobs et al., 1996).

Nevertheless, the sly gene is prevalent in almost all highly virulent S. suis strains

isolated from diseased pigs and humans in Europe and Asia. Taken together suilysin

is currently considered to contribute to virulence, though the toxin seems not to be

essential for pathogenesis. At present, it remains unclear which biological activities

are relevant for virulence of S. suis. Suilysin belongs to the large family of

cholesterol-dependent pore-forming cytolysins (CDC), thus membrane damage

followed by cell lysis (Figure 6-1A) is the obvious but possibly not the only biological

effect which contributes to pathogenesis (Billigton et al., 2000).

As described in the general introduction, pathogenesis of S. suis includes three

crucial steps. Briefly, colonisation and invasion, dissemination within the bloodstream

and finally translocation into target tissues (Gottschalk and Segura, 2000). In

chapter 3 biological activities of suilysin which play a role in early stages of S. suis

pathogenesis were investigated. The first barrier of defence after the intranasal route

of infection is the mucosal epithelium lining of the upper respiratory tract. Although

S. suis is considered an extracellular pathogen, interaction with epithelial cells to

General discussion Chapter 6

156

breach the mucosal barrier is a crucial step. It is largely unknown which bacterial

factors and host cell components are involved in this process. The present study

showed that suilysin promotes invasion of an unencapsulated S. suis serotype 2

strain in vitro in a trigger-like uptake mechanism associated with membrane ruffles

(Figure 3-1). The polysaccharide capsule is assumed to interfere with the ability of

S. suis to interact with host tissue (Benga et al., 2004; Vanier et al., 2004). Since its

main function is protection against phagocytosis after entering the bloodstream, it

has been proposed that the capsule is down-regulated during colonisation of the

mucosal epithelium to allow adherence and invasion (Gottschalk and Segura, 2000;

Okamoto et al., 2004; Willenborg et al., 2011). In accordance, unencapsulated

S. suis stains showed a higher adherence and invasion rate in comparison to the well

encapsulated wild type strains (Benga et al., 2004). Furthermore, we found evidence

that the signal protein kinase Rac, belonging to the small family of Rho-GTPases, is

a key regulator in suilysin promoted uptake of S. suis (Figure 3-2). Recruitment of the

Rho family guanosine triphosphatases, namely Rac1, Rho, Cdc42, for bacterial

internalisation has been previously reported for several bacterial species (Agarwal

and Hammerschmidt, 2009; Burnham et al., 2007; Criss and Casanova, 2003; Finlay,

2005). Rho-GTPases can act as targets of bacterial toxins (Aktories et al., 2000) like

the cytotoxic necrotizing factors CNF1 and CNF2 from Escherichia coli (Schmidt and

Aktories, 2000) or the Salmonella outer protein E (Hardt et al., 1998). An interaction

of CDC with G-proteins has only been described for pneumolysin so far (Iliev et al.,

2007).

It still remains open whether these in vitro results reflect the in vivo situation of

S. suis pathogenesis. Therefore, an intranasal mouse infection model for colonisation

and invasion for S. suis was established (chapter 5). As asymptomatically colonized

carrier pigs play a key role in epidemiology of S. suis, a important step in

pathogenesis is sufficient colonisation of (respiratory) mucosal surfaces. Using an

intranasal instillation of a highly virulent S. suis serotype 2 strain after nasal

predisposition with acetic acid we were able to induce colonisation of the upper

respiratory tract (Table 5-4). We observed that invasive disease associated with

rhinitis, pneumonia and/or meningitis was only seen after pretreatment with acetic

General discussion Chapter 6

157

acid. This corresponds well to the situation of natural infection, where coinfections

with other pathogenic agents, such as PRRSV or Bordetella bronchiseptica, mucosal

irrigation or stress are predisposing factors for S. suis infections (Pallares et al.,

2003). In contrast to typical histopathological lesions for S. suis infection in swine,

which are characterized by fibrinosuppurative inflammation of the respective tissue

(Beineke et al., 2008; Williams and Blakemore, 1990), we observed purulent

necrotizing alteration of the respiratory apparatus (nose and lung) and the central

nervous system (CNS) in mice. Similar pathologies of the CNS were described for

mice after intraperitoneal challenge (Dominguez-Punaro et al., 2007; Vecht et al.,

1997). Presumably, induction of necrosis is directly linked to cytotoxic effects of

suilysin as observed for different cell types in vitro (Charland et al., 2000; Gottschalk

and Segura, 2000; Lalonde et al., 2000; Norton et al., 1999; Tenenbaum et al., 2005;

Vanier et al., 2004). Tenenbaum et al. (2006) found necrosis of porcine choroid

plexus epithelial cells after incubation with a sly-positive S. suis strain. A similar

distribution of S. suis and suilysin in affected brain tissues of experimental infected

pigs associated with massive infiltration of neutrophils and acute necrosis of neurons

was described by Zheng et al. (2009).

To further investigate the role of suilysin in early stages of S. suis pathogenesis in

vivo we applied this model to perform a comparative study using a highly virulent

serotype 2 strain, its isogenic sly-negative strain (10Δsly), as well as the respective

isogenic unencapsulated strains (10cpsΔEF and 10cpsΔEFΔsly). Preliminary results

suggested attenuation of sly-deficient strains indicated by reduced morbidity and less

histopathological changes (Table 5-5 and Table 5-6). In contrast, colonisation of the

respiratory epithelium by suilysin-negative strains was unaffected. Thus, suilysin

seems to be not critical for colonisation, but contributes to invasion of S. suis in vivo.

Nevertheless, the results are not conclusive because induction of invasive disease

such as meningitis or pneumonia failed in this trial except for two cases of meningitis

in the wild type infected group and rhinitis in all infected animals. For the closely

related pneumococci proven intranasal pneumonia and bacteraemia models in

mouse have been described previously (Berry et al., 1989; Medina, 2010; Saeland et

al., 2000). Intranasal challenge studies showed that pneumolysin deficient mutant

General discussion Chapter 6

158

strains were less virulent in mice indicated by a significant longer median survival

time, reduced apathetical alterations in the lung and diminished onset of bacteraemia

(Alexander et al., 1998; Berry et al., 1989; Jounblat et al., 2003). In contrast,

pneumolysin production was not required for successful nasopharyngeal colonization

(Rubins et al., 1998).

As described in chapter 3, we observed that suilysin effects did not require cytolytic

activities. Therefore, to further dissect functional domains involved in biological

activities of suilysin, we introduced point-mutations of the tryptophan-rich pore-

forming region (W461F) and the RGD-motif (SVD) within the sly gene (chapter 4).

Both mutations resulted in a loss of haemolytic and cytolytic activity (Figure 4-2). The

tryptophan-mutant had retained its ability to bind to the host-cell membrane and to

activate Rac. In contrast, an unmodified RGD-motif seems to be required for efficient

membrane binding (Figure 4-6) and more strikingly for activation of Rac1 (Figure

4-7). Therefore we hypothesized that the RGD-motif of suilysin interacts with

heterodimeric transmembrane cell surface receptors, namely α5β1 integrins, to

mediate suilysin-promoted activation of Rac1 and subsequent invasion of S. suis in

target cells. In addition, the proven cholesterol dependency of suilysin would be

explainable since integrins are most frequently expressed on cholesterol enriched

microdomains (lipid rafts) (Lafont and van der Goot, 2005). Nevertheless, it remains

hypothetical and the role of the first domain of suilysin and integrins in S. suis-host

cell interaction should be addressed in future studies.

Based on our results we propose the following model of suilysin-promoted invasion of

the upper respiratory tract by S. suis after colonisation of the mucosal surface as

illustrated by Figure 6-1B. Secreted suilysin binds to cholesterol enriched micro

domains (lipid rafts?) of the host cell membrane either via domain 4 and/or via an

interaction between α5β1 integrins and the RGD-motif of domain 1 (1). If applicable,

the prepore complex is formed and a downstream suilysin-mediated activation of

Rac1 (2) leads to induction of F-actin (3) and the formation of membrane ruffles.

Finally, S. suis is taken up by the cell (4). Suilysin-promoted invasion into host tissue

and intracellular persistence might be crucial for bacterial pathogenesis, since they

allow pathogens to escape the immune system for further dissemination and invasive

General discussion Chapter 6

159

disease development (Finlay and Falkow, 1997; Garzoni and Kelley, 2009). On the

other hand suilysin might be beneficial for the host, since subcytolytic concentrations

of suilysin may be sensed by epithelial cells to commence an early onset of innate

immune response in the host as proposed by Ratner et al. (2006) for other bacterial

toxins.

This study contributed to the knowledge on virulence mechanisms of S. suis, but

many open questions remain to be addressed in future studies to further dissect

mechanisms underlying pathogenesis and to evaluate the role of suilysin in virulence

of S. suis.

Figure 6-1: Hypothetical model of suilysin-promoted interaction of S. suis with epithelial cells. (A) Direct cytotoxic effects of suilysin: Pore-formation followed by cell lysis modified from Gottschalk et al. 2010; Willenborg (unpublished). (B) Subcytolytic effects of suilysin: Suilysin-induced invasion of S. suis by a Rac1-dependent remodelling of the actin cytoskeleton. Secreted suilysin binds to the membrane, forms the prepore complex (1) and activates Rac1 (2). Activation of Rac1 leads to induction of F-actin (3) and the formation of membrane ruffles which results in an uptake of S. suis by the cell (4).

General discussion Chapter 6

160

References Agarwal, V. and Hammerschmidt, S. (2009) Cdc42 and the phosphatidylinositol 3-kinase-Akt pathway are essential for PspC-mediated internalization of pneumococci by respiratory epithelial cells. J Biol Chem 284: 19427-19436.

Aktories, K., Schmidt, G., and Just, I. (2000) Rho GTPases as targets of bacterial protein toxins. Biol Chem 381: 421-426.

Alexander, J. E., Berry, A. M., Paton, J. C., Rubins, J. B., Andrew, P. W., and Mitchell, T. J. (1998) Amino acid changes affecting the activity of pneumolysin alter the behaviour of pneumococci in pneumonia. Microb Pathog 24: 167-174.

Allen, A. G., Bolitho, S., Lindsay, H., Khan, S., Bryant, C., Norton, P., Ward, P., Leigh, J., Morgan, J., Riches, H., Eastty, S., and Maskell, D. (2001) Generation and characterization of a defined mutant of Streptococcus suis lacking suilysin. Infect Immun 69: 2732-2735.

Beineke, A., Bennecke, K., Neis, C., Schroder, C., Waldmann, K. H., Baumgartner, W., Valentin- Weigand, P., and Baums, C. G. (2008) Comparative evaluation of virulence and pathology of Streptococcus suis serotypes 2 and 9 in experimentally infected growers. Vet Microbiol 128: 423-430.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cellular Microbiology 6: 867-881.

Berry, A. M., Yother, J., Briles, D. E., Hansman, D., and Paton, J. C. (1989) Reduced virulence of a defined pneumolysin-negative mutant of Streptococcus pneumoniae. Infect Immun 57: 2037- 2042.

Billington, S. J., Jost, B. H., and Songer, J. G. (2000) Thiol-activated cytolysins: structure, function and role in pathogenesis. FEMS Microbiol Lett 182: 197-205.

Burnham, C. A., Shokoples, S. E., and Tyrrell, G. J. (2007) Rac1, RhoA, and Cdc42 participate in HeLa cell invasion by group B streptococcus. FEMS Microbiol Lett 272: 8-14.

Charland, N., Nizet, V., Rubens, C. E., Kim, K. S., Lacouture, S., and Gottschalk, M. (2000) Streptococcus suis serotype 2 interactions with human brain microvascular endothelial cells. Infect Immun 68: 637-643.

Clifton-Hadley, F. A. and Alexander, T. J. (1980) The carrier site and carrier rate of Streptococcus suis type II in pigs. Vet Rec 107: 40-41.

Criss, A. K. and Casanova, J. E. (2003) Coordinate regulation of Salmonella enterica serovar Typhimurium invasion of epithelial cells by the Arp2/3 complex and Rho GTPases. Infect Immun 71: 2885-2891.

Dominguez-Punaro, M. C., Segura, M., Plante, M. M., Lacouture, S., Rivest, S., and Gottschalk, M. (2007) Streptococcus suis serotype 2, an important swine and human pathogen, induces strong systemic and cerebral inflammatory responses in a mouse model of infection. J Immunol 179: 1842-1854.

Finlay, B. B. (2005) Bacterial virulence strategies that utilize Rho GTPases. Curr Top Microbiol Immunol 291: 1-10.

General discussion Chapter 6

161

Finlay, B. B. and Falkow, S. (1997) Common themes in microbial pathogenicity revisited. Microbiol Mol Biol Rev 61: 136-169.

Garzoni, C. and Kelley, W. L. (2009) Staphylococcus aureus: new evidence for intracellular persistence. Trends Microbiol 17: 59-65.

Gottschalk, M. and Segura, M. (2000) The pathogenesis of the meningitis caused by Streptococcus suis: the unresolved questions. Veterinary Microbiology 76: 259-272.

Hardt, W. D., Chen, L. M., Schuebel, K. E., Bustelo, X. R., and Galan, J. E. (1998) S. typhimurium encodes an activator of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell 93: 815-826.

Iliev, A. I., Djannatian, J. R., Nau, R., Mitchell, T. J., and Wouters, F. S. (2007) Cholesterol-dependent actin remodeling via RhoA and Rac1 activation by the Streptococcus pneumoniae toxin pneumolysin. Proc Natl Acad Sci U S A 104: 2897-2902.

Jacobs, A. A. C., Loeffen, P. L. W., vandenBerg, A. J. G., and Storm, P. K. (1994) Identification, Purification, and Characterization of A Thiol-Activated Hemolysin (Suilysin) of Streptococcus Suis. Infection and Immunity 62: 1742-1748.

Jacobs, A. A. C., vandenBerg, A. J. G., and Loeffen, P. L. W. (1996) Protection of experimentally infected pigs by suilysin, the thiol-activated haemolysin of Streptococcus suis. Veterinary Record 139: 225-228.

Jounblat, R., Kadioglu, A., Mitchell, T. J., and Andrew, P. W. (2003) Pneumococcal behavior and host responses during bronchopneumonia are affected differently by the cytolytic and complement- activating activities of pneumolysin. Infect Immun 71: 1813-1819.

Lafont, F. and van der Goot, F. G. (2005) Bacterial invasion via lipid rafts. Cell Microbiol 7: 613-620.

Lalonde, M., Segura, M., Lacouture, S., and Gottschalk, M. (2000) Interactions between Streptococcus suis serotype 2 and different epithelial cell lines. Microbiology 146 ( Pt 8): 1913- 1921.

Lun, S. C., Perez-Casal, J., Connor, W., and Willson, P. J. (2003) Role of suilysin in pathogenesis of Streptococcus suis capsular serotype 2. Microbial Pathogenesis 34: 27-37.

Medina, E. (2010) Murine model of pneumococcal pneumonia. Methods Mol Biol 602: 405-410.

Norton, P. M., Rolph, C., Ward, P. N., Bentley, R. W., and Leigh, J. A. (1999) Epithelial invasion and cell lysis by virulent strains of Streptococcus suis is enhanced by the presence of suilysin. Fems Immunology and Medical Microbiology 26: 25-35.

Okamoto, S., Kawabata, S., Terao, Y., Fujitaka, H., Okuno, Y., and Hamada, S. (2004) The Streptococcus pyogenes capsule is required for adhesion of bacteria to virus-infected alveolar epithelial cells and lethal bacterial-viral superinfection. Infect Immun 72: 6068-6075.

Pallares, F. J., Halbur, P. G., Schmitt, C. S., Roth, J. A., Opriessnig, T., Thomas, P. J., Kinyon, J. M., Murphy, D., Frank, D. E., and Hoffman, L. J. (2003) Comparison of experimental models for Streptococcus suis infection of conventional pigs. Can J Vet Res 67: 225-228.

Ratner, A. J., Hippe, K. R., Aguilar, J. L., Bender, M. H., Nelson, A. L., and Weiser, J. N. (2006) Epithelial cells are sensitive detectors of bacterial pore-forming toxins. J Biol Chem 281: 12994-12998.

General discussion Chapter 6

162

Rubins, J. B., Paddock, A. H., Charboneau, D., Berry, A. M., Paton, J. C., and Janoff, E. N. (1998) Pneumolysin in pneumococcal adherence and colonization. Microb Pathog 25: 337-342.

Saeland, E., Vidarsson, G., and Jonsdottir, I. (2000) Pneumococcal pneumonia and bacteremia model in mice for the analysis of protective antibodies. Microb Pathog 29: 81-91.

Schmidt, G. and Aktories, K. (2000) Rho GTPase-activating toxins: cytotoxic necrotizing factors and dermonecrotic toxin. Methods Enzymol 325: 125-136.

Tenenbaum, T., Adam, R., Eggelnpohler, I., Matalon, D., Seibt, A., GE, K. Novotny, Galla, H. J., and Schroten, H. (2005) Strain-dependent disruption of blood-cerebrospinal fluid barrier by Streptoccocus suis in vitro. FEMS Immunol Med Microbiol 44: 25-34.

Tenenbaum, T., Essmann, F., Adam, R., Seibt, A., Janicke, R. U., Novotny, G. E., Galla, H. J., and Schroten, H. (2006) Cell death, caspase activation, and HMGB1 release of porcine choroid plexus epithelial cells during Streptococcus suis infection in vitro. Brain Res 1100: 1-12.

Vanier, G., Segura, M., Friedl, P., Lacouture, S., and Gottschalk, M. (2004) Invasion of porcine brain microvascular endothelial cells by Streptococcus suis serotype 2. Infection and Immunity 72: 1441-1449.

Vecht, U., StockhofeZurwieden, N., Tetenburg, B. J., Wisselink, H. J., and Smith, H. E. (1997) Virulence of Streptococcus suis type 2 for mice and pigs appeared host-specific. Veterinary Microbiology 58: 53-60.

Willenborg, J., Fulde, M., de Greeff, A., Rohde, M., Smith, H. E., Valentin-Weigand, P., and Goethe, R. (2011) Role of glucose and CcpA in capsule expression and virulence of Streptococcus suis. Microbiology 157: 1823-1833.

Williams, A. E. and Blakemore, W. F. (1990) Pathogenesis of meningitis caused by Streptococcus suis type 2. J Infect Dis 162: 474-481.

Zheng, H., Punaro, M. C., Segura, M., Lachance, C., Rivest, S., Xu, J., Houde, M., and Gottschalk, M. (2011) Toll-like receptor 2 is partially involved in the activation of murine astrocytes by Streptococcus suis, an important zoonotic agent of meningitis. J Neuroimmunol 234: 71-83.

Chapter 7

Summary

Summary Chapter 7

165

Summary

Streptococcus (S.) suis is a swine pathogen of the upper respiratory tract causing

meningitis, arthritis and septicaemia. Due to high economical losses S. suis is

considered as an important pathogen in pig husbandry worldwide. Asymptomatically

colonized carrier-pigs are the most important source of infection. Moreover, S. suis is

a zoonotic agent associated with meningitis and the streptococcal toxic shock like

syndrome (STSS) in humans.

Suilysin is a secreted virulence-associated haemolysin of S. suis. It is present in

many S. suis strains and belongs to the family of Cholesterol-dependent pore-

forming cytolysins (CDC). At high concentrations these multifunctional toxins lyse

host cells. Additional biological activities of CDC such as the activation of host cells

or immune modulation, occur at subcytolytic concentrations. The role of suilysin in

pathogenesis of S. suis is still unclear.

The aim of this study was to analyse the role of suilysin in pathogen-host cell

interaction and to characterise the molecular mechanisms underling suilysin-

mediated biological effects.

Using an isogenic suilysin-negative mutant strain and the human respiratory

epithelial cell line HEp-2 a suilysin-dependent invasive phenotype of an

unencapsulated S. suis serotype 2 strain was identified in vitro. Subcytolytic

concentration of suilysin activated the small Rho-GTPase Rac1, which resulted in

formation of membrane ruffles and finally led to the uptake of S. suis by the cell.

Furthermore, point-mutation of the putative pore-forming region of the suilysin gene

abolished cytolytic activity, but activation of the G-protein Rac1 was unaffected. A

second putative functional region, a RGD-motif, was identified and also point-

mutated, resulting in loss of lytic activity, membrane binding and the ability to activate

Rac1. Thus, the RGD-motif is proposed to be involved in suilysin-mediated host cell-

activation. Integrins as possible interaction partners were identified on the surface of

HEp-2 cells, but further studies are needed to proof the precise role of integrins in

suilysin-mediated host cell-interaction.

Summary Chapter 7

166

After characterization of suilysin-dependent effects in vitro, an intranasal mouse

infection model for S. suis was established to investigate the role of suilysin in

colonisation and invasion of S. suis in vivo. After predisposition with acetic acid

followed by mucosal application of a S. suis serotype 2 strain we were able to induce

colonisation of the upper respiratory tract as well as invasive disease associated with

rhinitis, pneumonia and/or meningitis in mice. The suilysin-negative mutant strain

colonised the respiratory epithelium, but was slightly attenuated in virulence in

comparison to the wild type strain.

Taken together, these results show that subcytolytic activity of suilysin plays a role in

pathogen-host cell interaction. The established mouse model is suitable to study the

relevance of subcytolytic activities of suilysin in vivo in the future. The results of this

work improved our understanding of S. suis pathogenicity mechanisms.

Chapter 8

Zusammenfasung

Zusammenfassung Chapter 8

169

Zusammenfassung

Streptococcus (S.) suis ist ein Erreger im oberen Respirationstrakt des Schweins, der

vor allem Meningitiden, Arthritiden und Septikämien verursacht. Aufgrund von hohen

Verlusten in der Schweineproduktion sind S. suis-bedingte Erkrankungen weltweit

von großer Bedeutung. Epidemiologisch bedeutsam sind vor allem klinisch

inapparente Trägertiere, die als wichtigstes Reservoir für Neuinfektionen gelten.

Zudem besitzt S. suis ein zoonotisches Potential und spielt beim Menschen sowohl

eine wichtige Rolle als Meningitiserreger als auch als Verursacher des streptococcal

toxic shock like syndrom (STSS).

Zu den virulenzassoziierten Faktoren von S. suis gehört das Suilysin, welches ein

sezerniertes Hämolysin ist. Dieses tritt bei vielen Stämmen auf und zählt zur Familie

der cholesterol-dependent pore-forming cytolysins (CDC). Diese multifunktionalen

Toxine führen in hoher Konzentration zur Lyse von Wirtszellen. Neben diesen

lytischen Effekten sind auch biologische Aktivitäten im subzytoloytischen Bereich,

wie die Aktivierung von Wirtszellen oder immunmodulatorische Effekte, von

Bedeutung. Die genaue Rolle des Suilysins in der Pathogenese von S. suis ist bisher

nicht hinreichend geklärt.

Ziel dieser Arbeit war es daher, die Bedeutung von Suilysin in der Erreger-Wirtszell-

Interaktion und die molekularen Mechanismen Suilysin-vermittelter biologischer

Effekte näher zu untersuchen.

Mit Hilfe einer isogenen Suilysin-negativen Mutante konnte in in vitro

Untersuchungen an der humanen respiratorischen Epithelzelllinie HEp-2 ein Suilysin-

abhängiger invasiver Phänotyp eines unbekapselten S. suis-Serotyp-2-Stammes

gezeigt werden. Die durch subzytolytische Suilysin-Konzentration induzierte

Aktivierung der kleinen Rho-GTPase Rac1 vermittelte hierbei die Ausbildung von

membrane ruffles, wodurch S. suis in die Zelle aufgenommen wurde. Des Weiteren

konnte durch gezielte Punktmutagenese im Suilysin-Gen der putative Poren-bildende

Bereich ausgeschaltet werden, was zu einem Verlust der zytolytischen Aktivität

führte, nicht aber die G-Protein-aktivierende Eigenschaft des Suilysins beeinflusste.

Neben dem für die Porenbildung notwendigen Bereich im Suilysin-Gen, wurde eine

Zusammenfassung Chapter 8

170

weitere putative funktionale Domäne, ein RGD-Motiv, identifiziert und ebenfalls

punktmutiert. Der dadurch bedingte Verlust der zytolytischen Aktivität, der Bindung

des Suilysins an die Wirtszell-Membran und der Rac1-Aktivierung, weist auf eine

Beteiligung des RGD-Motivs an der Suilysin-vermittelten Interaktion mit Wirtszellen

hin. Als mögliche Interaktionspartner konnten Integrine auf der Zelloberfläche von

HEp-2-Zellen identifiziert werden, dennoch sind weitere funktionelle Studien

erforderlich, um die genaue Rolle von Integrinen in der Suilysin-vermittelten Zell-

Aktivierung aufzuklären.

Nach der Charakterisierung Suilysin-abhängiger Effekte im Zellkultursystem wurde

ein intranasales Mausinfektionsmodells für S. suis etabliert, um die Bedeutung des

Suilysins für die Kolonisation und Invasion von S. suis in vivo zu untersuchen. Nach

vorheriger Prädisposition mit Essigsäure war es möglich durch mukosale Applikation

des S. suis-Serotyp-2-Stammes eine Besiedlung des oberen Respirationstraktes

sowie das Auftreten von Rhinitis, Pneumonie und/oder Meningitis in der Maus zu

induzieren. Im Vergleich zum Wildtyp-Stamm erwies sich die Suilysin-negative

Mutante als geringgradig virulenzattenuiert, wohingegen keine Unterschiede in der

Kolonisation beobachtet wurden.

Zusammenfassend zeigen die Ergebnisse, dass Suilysin nicht nur zytolytische

Aktivitäten besitzt, sondern auch unter subzytolytischen Bedingungen eine

Bedeutung in der Erreger-Wirtszell-Interaktion hat. Die weitere Klärung der Relevanz

subzytolytischer Aktivitäten des Suilysins in vivo kann zukünftig durch das im

Rahmen dieses Projektes etablierten Mausmodels näher geklärt werden. Damit

wurde ein Beitrag zum besseren Verständnis von Pathogenitätsmechanismen von S.

suis geleistet.

Chapter 9

Literature

Literature Chapter 9

173

Literature

Agarwal, V. and Hammerschmidt, S. (2009) Cdc42 and the phosphatidylinositol 3-kinase-Akt pathway are essential for PspC-mediated internalization of pneumococci by respiratory epithelial cells. J Biol Chem 284: 19427-19436.

Aktories, K., Schmidt, G., and Just, I. (2000) Rho GTPases as targets of bacterial protein toxins. Biol Chem 381: 421-426.

Al Numani, D., Segura, M., Dore, M., and Gottschalk, M. (2003) Up-regulation of ICAM-1, CD11a/CD18 and CD11c/CD18 on human THP-1 monocytes stimulated by Streptococcus suis serotype 2. Clin Exp Immunol 133: 67-77.

Alcantara, R. B., Preheim, L. C., and Gentry-Nielsen, M. J. (2001) Pneumolysin-induced complement depletion during experimental pneumococcal bacteremia. Infect Immun 69: 3569-3575.

Alexander, J. E., Lock, R. A., Peeters, C. C., Poolman, J. T., Andrew, P. W., Mitchell, T. J., Hansman, D., and Paton, J. C. (1994) Immunization of mice with pneumolysin toxoid confers a significant degree of protection against at least nine serotypes of Streptococcus pneumoniae. Infect Immun 62: 5683-5688.

Allen, A. G., Bolitho, S., Lindsay, H., Khan, S., Bryant, C., Norton, P., Ward, P., Leigh, J., Morgan, J., Riches, H., Eastty, S., and Maskell, D. (2001) Generation and characterization of a defined mutant of Streptococcus suis lacking suilysin. Infect Immun 69: 2732-2735.

Alouf, J. E. (2000) Cholesterol-binding cytolytic protein toxins. Int J Med Microbiol 290: 351-356.

Amass, S. F., SanMiguel, P., and Clark, L. K. (1997) Demonstration of vertical transmission of Streptococcus suis in swine by genomic fingerprinting. J Clin Microbiol 35: 1595-1596.

Arends, J. P., Hartwig, N., Rudolphy, M., and Zanen, H. C. (1984) Carrier Rate of Streptococcus-Suis Capsular Type-2 in Palatine Tonsils of Slaughtered Pigs. Journal of Clinical Microbiology 20: 945-947.

Arends, J. P. and Zanen, H. C. (1988) Meningitis Caused by Streptococcus Suis in Humans. Reviews of Infectious Diseases 10: 131-137.

Baele, M., Chiers, K., Devriese, L. A., Smith, H. E., Wisselink, H. J., Vaneechoutte, M., and Haesebrouck, F. (2001) The gram-positive tonsillar and nasal flora of piglets before and after weaning. J Appl Microbiol 91: 997-1003.

Baums, C. G. and Valentin-Weigand, P. (2009) Surface-associated and secreted factors of Streptococcus suis in epidemiology, pathogenesis and vaccine development. Anim Health Res Rev 10: 65-83.

Baums, C. G., Verkuhlen, G. J., Rehm, T., Silva, L. M., Beyerbach, M., Pohlmeyer, K., and Valentin- Weigand, P. (2007) Prevalence of Streptococcus suis genotypes in wild boars of Northwestern Germany. Appl Environ Microbiol 73: 711-717.

Baums, C. G., Kaim, U., Fulde, M., Ramachandran, G., Goethe, R., and Valentin-Weigand, P. (2006) Identification of a novel virulence determinant with serum opacification activity in Streptococcus suis. Infect Immun 74: 6154-6162.

Literature Chapter 9

174

Beineke, A., Bennecke, K., Neis, C., Schroder, C., Waldmann, K. H., Baumgartner, W., Valentin- Weigand, P., and Baums, C. G. (2008) Comparative evaluation of virulence and pathology of Streptococcus suis serotypes 2 and 9 in experimentally infected growers. Vet Microbiol 128: 423-430.

Benga, L., Friedl, P., and Valentin-Weigand, P. (2005) Adherence of Streptococcus suis to porcine endothelial cells. J Vet Med B Infect Dis Vet Public Health 52: 392-395.

Benga, L., Fulde, M., Neis, C., Goethe, R., and Valentin-Weigand, P. (2008) Polysaccharide capsule and suilysin contribute to extracellular survival of Streptococcus suis co-cultivated with primary porcine phagocytes. Vet Microbiol 132: 211-219.

Benga, L., Goethe, R., Rohde, M., and Valentin-Weigand, P. (2004) Non-encapsulated strains reveal novel insights in invasion and survival of Streptococcus suis in epithelial cells. Cellular Microbiology 6: 867-881.

Berry, A. M., Yother, J., Briles, D. E., Hansman, D., and Paton, J. C. (1989) Reduced virulence of a defined pneumolysin-negative mutant of Streptococcus pneumoniae. Infect Immun 57: 2037- 2042.

Berthelot-Herault, F., Gottschalk, M., Labbe, A., Cariolet, R., and Kobisch, M. (2001) Experimental airborne transmission of Streptococcus suis capsular type 2 in pigs. Vet Microbiol 82: 69-80.

Berthelot-Herault, F., Morvan, H., Keribin, A. M., Gottschalk, M., and Kobisch, M. (2000) Production of Muraminidase-Released Protein (MRP), Extracellular Factor (EF) and Suilysin by field isolates of Streptococcus suis capsular types 2, 1/2, 9, 7 and 3 isolated from swine in France. Veterinary Research 31: 473-479.

Billington, S. J., Jost, B. H., Cuevas, W. A., Bright, K. R., and Songer, J. G. (1997) The Arcanobacterium (Actinomyces) pyogenes hemolysin, pyolysin, is a novel member of the thiol- activated cytolysin family. J Bacteriol 179: 6100-6106.

Billington, S. J., Jost, B. H., and Songer, J. G. (2000) Thiol-activated cytolysins: structure, function and role in pathogenesis. FEMS Microbiol Lett 182: 197-205.

Blume, V., Luque, I., Vela, A. I., Borge, C., Maldonado, A., Dominguez, L., Tarradas, C., and Fernandez-Garayzabal, J. F. (2009) Genetic and virulence-phenotype characterization of serotypes 2 and 9 of Streptococcus suis swine isolates. Int Microbiol 12: 161-166.

Bonifait, L., Grignon, L., and Grenier, D. (2008) Fibrinogen induces biofilm formation by Streptococcus suis and enhances its antibiotic resistance. Appl Environ Microbiol 74: 4969-4972.

Boulnois, G. J., Paton, J. C., Mitchell, T. J., and Andrew, P. W. (1991) Structure and function of pneumolysin, the multifunctional, thiol-activated toxin of Streptococcus pneumoniae. Mol Microbiol 5: 2611-2616.

Burnham, C. A., Shokoples, S. E., and Tyrrell, G. J. (2007) Rac1, RhoA, and Cdc42 participate in HeLa cell invasion by group B streptococcus. FEMS Microbiol Lett 272: 8-14.

Chabot-Roy, G., Willson, P., Segura, M., Lacouture, S., and Gottschalk, M. (2006) Phagocytosis and killing of Streptococcus suis by porcine neutrophils. Microb Pathog 41: 21-32.

Chanter, N., Jones, P. W., and Alexander, T. J. (1993) Meningitis in pigs caused by Streptococcus suis--a speculative review. Vet Microbiol 36: 39-55.

Literature Chapter 9

175

Charland, N., Harel, J., Kobisch, M., Lacasse, S., and Gottschalk, M. (1998) Streptococcus suis serotype 2 mutants deficient in capsular expression. Microbiology-Sgm 144: 325-332. Charland, N., Nizet, V., Rubens, C. E., Kim, K. S., Lacouture, S., and Gottschalk, M. (2000) Streptococcus suis serotype 2 interactions with human brain microvascular endothelial cells. Infect Immun 68: 637-643.

Chen, C., Tang, J., Dong, W., Wang, C., Feng, Y., Wang, J., Zheng, F., Pan, X., Liu, D., Li, M., Song, Y., Zhu, X., Sun, H., Feng, T., Guo, Z., Ju, A., Ge, J., Dong, Y., Sun, W., Jiang, Y., Wang, J., Yan, J., Yang, H., Wang, X., Gao, G. F., Yang, R., Wang, J., and Yu, J. (2007) A glimpse of streptococcal toxic shock syndrome from comparative genomics of S. suis 2 Chinese isolates. PLoS ONE 2: e315.

Clifton-Hadley, F. A. and Alexander, T. J. (1980) The carrier site and carrier rate of Streptococcus suis type II in pigs. Vet Rec 107: 40-41.

Clifton-Hadley, F. A., Alexander, T. J., and Enright, M. R. (1986) Monitoring herds for Streptococcus suis type 2: chance contamination of slaughter pigs. Vet Rec 118: 274.

Cockeran, R., Durandt, C., Feldman, C., Mitchell, T. J., and Anderson, R. (2002) Pneumolysin activates the synthesis and release of interleukin-8 by human neutrophils in vitro. J Infect Dis 186: 562-565.

Coue, M., Brenner, S. L., Spector, I., and Korn, E. D. (1987) Inhibition of actin polymerization by latrunculin A. FEBS Lett 213: 316-318.

Criss, A. K. and Casanova, J. E. (2003) Coordinate regulation of Salmonella enterica serovar Typhimurium invasion of epithelial cells by the Arp2/3 complex and Rho GTPases. Infect Immun 71: 2885-2891.

Cue, D., Southern, S. O., Southern, P. J., Prabhakar, J., Lorelli, W., Smallheer, J. M., Mousa, S. A., and Cleary, P. P. (2000) A nonpeptide integrin antagonist can inhibit epithelial cell ingestion of Streptococcus pyogenes by blocking formation of integrin alpha 5beta 1-fibronectin-M1 protein complexes. Proc Natl Acad Sci U S A 97: 2858-2863.

Czajkowsky, D. M., Hotze, E. M., Shao, Z., and Tweten, R. K. (2004) Vertical collapse of a cytolysin prepore moves its transmembrane beta-hairpins to the membrane. EMBO J 23: 3206-3215.

de Greeff, A., Buys, H., Verhaar, R., Dijkstra, J., van Alphen, L., and Smith, H. E. (2002) Contribution of fibronectin-binding protein to pathogenesis of Streptococcus suis serotype 2. Infect Immun 70: 1319-1325.

de Greeff, A., Wisselink, H. J., de Bree, F. M., Schultsz, C., Baums, C. G., Thi, H. N., Stockhofe- Zurwieden, N., and Smith, H. E. (2011) Genetic diversity of Streptococcus suis isolates as determined by comparative genome hybridization. BMC Microbiol 11: 161.

Dessing, M. C., Hirst, R. A., de Vos, A. F., and van der, Poll T. (2009) Role of Toll-like receptors 2 and 4 in pulmonary inflammation and injury induced by pneumolysin in mice. PLoS One 4: e7993.

Dominguez-Punaro, Mde L., Segura, M., Contreras, I., Lachance, C., Houde, M., Lecours, M. P., Olivier, M., and Gottschalk, M. (2010) In vitro characterization of the microglial inflammatory response to Streptococcus suis, an important emerging zoonotic agent of meningitis. Infect Immun 78: 5074-5085.

Literature Chapter 9

176

Dominguez-Punaro, M. C., Segura, M., Plante, M. M., Lacouture, S., Rivest, S., and Gottschalk, M. (2007) Streptococcus suis serotype 2, an important swine and human pathogen, induces strong systemic and cerebral inflammatory responses in a mouse model of infection. J Immunol 179: 1842-1854.

El Rachkidy, R. G., Davies, N. W., and Andrew, P. W. (2008) Pneumolysin generates multiple conductance pores in the membrane of nucleated cells. Biochem Biophys Res Commun 368: 786-792.

Enright, M. R., Alexander, T. J., and Clifton-Hadley, F. A. (1987) Role of houseflies (Musca domestica) in the epidemiology of Streptococcus suis type 2. Vet Rec 121: 132-133.

Fabisiak, M., Kita, J., Jedryczko, R., and Binek, M. (2005) Prevalence of the suilysin gene in Streptococcus suis strains isolated from diseased and healthy carrier pigs. Pol J Vet Sci 8: 141-145.

Feldman, C., Mitchell, T. J., Andrew, P. W., Boulnois, G. J., Read, R. C., Todd, H. C., Cole, P. J., and Wilson, R. (1990) The effect of Streptococcus pneumoniae pneumolysin on human respiratory epithelium in vitro. Microb Pathog 9: 275-284.

Finlay, B. B. (2005) Bacterial virulence strategies that utilize Rho GTPases. Curr Top Microbiol Immunol 291: 1-10.

Finlay, B. B. and Cossart, P. (1997) Exploitation of mammalian host cell functions by bacterial pathogens. Science 276: 718-725

Fittipaldi, N., Fuller, T. E., Teel, J. F., Wilson, T. L., Wolfram, T. J., Lowery, D. E., and Gottschalk, M. (2009) Serotype distribution and production of muramidase-released protein, extracellular factor and suilysin by field strains of Streptococcus suis isolated in the United States. Vet Microbiol 139: 310-317.

Fittipaldi, N., Gottschalk, M., Vanier, G., Daigle, F., and Harel, J. (2007) Use of selective capture of transcribed sequences to identify genes preferentially expressed by Streptococcus suis upon interaction with porcine brain microvascular endothelial cells. Appl Environ Microbiol 73: 4359- 4364.

Flanagan, J. J., Tweten, R. K., Johnson, A. E., and Heuck, A. P. (2009) Cholesterol exposure at the membrane surface is necessary and sufficient to trigger perfringolysin O binding. Biochemistry 48: 3977-3987.

Fongcom, A., Pruksakorn, S., Mongkol, R., Tharavichitkul, P., and Yoonim, N. (2001) Streptococcus suis infection in northern Thailand. J Med Assoc Thai 84: 1502-1508.

Garzoni, C. and Kelley, W. L. (2009) Staphylococcus aureus: new evidence for intracellular persistence. Trends Microbiol 17: 59-65.

Gekara, N. O. and Weiss, S. (2004) Lipid rafts clustering and signalling by listeriolysin O. Biochem Soc Trans 32: 712-714.

Gelber, S. E., Aguilar, J. L., Lewis, K. L., and Ratner, A. J. (2008) Functional and phylogenetic characterization of Vaginolysin, the human-specific cytolysin from Gardnerella vaginalis. J Bacteriol 190: 3896-3903.

Giddings, K. S., Johnson, A. E., and Tweten, R. K. (2003) Redefining cholesterol's role in the mechanism of the cholesterol-dependent cytolysins. Proc Natl Acad Sci U S A 100: 11315- 11320.

Literature Chapter 9

177

Giddings, K. S., Zhao, J., Sims, P. J., and Tweten, R. K. (2004) Human CD59 is a receptor for the cholesterol-dependent cytolysin intermedilysin. Nat Struct Mol Biol 11: 1173-1178.

Gottschalk, M., Lebrun, A., Wisselink, H., Dubreuil, J. D., Smith, H., and Vecht, U. (1998) Production of virulence-related proteins by Canadian strains of Streptococcus suis capsular type 2. Canadian Journal of Veterinary Research-Revue Canadienne de Recherche Veterinaire 62: 75-79.

Gottschalk, M., Petitbois, S., Higgins, R., and Jacques, M. (1991) Adherence of Streptococcus suis capsular type 2 to porcine lung sections. Can J Vet Res 55: 302-304.

Gottschalk, M. and Segura, M. (2000) The pathogenesis of the meningitis caused by Streptococcus suis: the unresolved questions. Veterinary Microbiology 76: 259-272.

Gottschalk, M., Segura, M., and Xu, J. (2007) Streptococcus suis infections in humans: the Chinese experience and the situation in North America. Anim Health Res Rev 8: 29-45.

Gottschalk, M., Xu, J., Calzas, C., and Segura, M. (2010) Streptococcus suis: a new emerging or an old neglected zoonotic pathogen? 25. Future Microbiol 5: 371-391.

Gottschalk, M. G., Lacouture, S., and Dubreuil, J. D. (1995) Characterization of Streptococcus suis capsular type 2 haemolysin. Microbiology 141 ( Pt 1): 189-195.

Greenwood, J., Steinman, L., and Zamvil, S. S. (2006) Statin therapy and autoimmune disease: from protein prenylation to immunomodulation. Nat Rev Immunol 6: 358-370.

Hamzaoui, N., Kerneis, S., Caliot, E., and Pringault, E. (2004) Expression and distribution of beta1 integrins in in vitro-induced M cells: implications for Yersinia adhesion to Peyer's patch epithelium. Cell Microbiol 6: 817-828.

Hall, A. (1998) Rho GTPases and the actin cytoskeleton. Science 279: 509-514.

Hardt, W. D., Chen, L. M., Schuebel, K. E., Bustelo, X. R., and Galan, J. E. (1998) S. typhimurium encodes an activator of Rho GTPases that induces membrane ruffling and nuclear responses in host cells. Cell 93: 815-826.

Herbert, D. and Todd, E. W. (1941) Purification and properties of a haemolysin produced by group A haemolytic streptococci (streptolysin O). Biochem J 35: 1124-1139.

Heuck, A. P., Hotze, E. M., Tweten, R. K., and Johnson, A. E. (2000) Mechanism of membrane insertion of a multimeric beta-barrel protein: perfringolysin O creates a pore using ordered and coupled conformational changes. Mol Cell 6: 1233-1242.

Higgins, R. and Gottschalk, M. (1990) An update on Streptococcus suis identification. J Vet Diagn Invest 2: 249-252.

Higgins,R. and Gottschalk,M. (2005) Streptococcal diseases. In Diseases of Swine. pp. 769-783.

Hoa, N. T., Chieu, T. T., Nghia, H. D., Mai, N. T., Anh, P. H., Wolbers, M., Baker, S., Campbell, J. I., Chau, N. V., Hien, T. T., Farrar, J., and Schultsz, C. (2011) The antimicrobial resistance patterns and associated determinants in Streptococcus suis isolated from humans in southern Vietnam, 1997-2008. BMC Infect Dis 11: 6.

Literature Chapter 9

178

Holo, H. and Nes, I. F. (1989) High-Frequency Transformation, by Electroporation, of Lactococcus lactis subsp. cremoris Grown with Glycine in Osmotically Stabilized Media. Appl Environ Microbiol 55: 3119-3123.

Hotze, E. M. and Tweten, R. K. (2011) Membrane assembly of the cholesterol-dependent cytolysin pore complex. Biochim Biophys Acta.

Huelsenbeck, J., Dreger, S., Gerhard, R., Barth, H., Just, I., Genth, H. (2007) Difference in the cytotoxic effects of toxin B from Clostridium difficile strain VPI 10463 and toxin B from variant Clostridium difficile strain 1470. Infect Immun 75(2): 801-9.

Iliev, A. I., Djannatian, J. R., Nau, R., Mitchell, T. J., and Wouters, F. S. (2007) Cholesterol-dependent actin remodeling via RhoA and Rac1 activation by the Streptococcus pneumoniae toxin pneumolysin. Proc Natl Acad Sci U S A 104: 2897-2902.

Iliev, A. I., Djannatian, J. R., Opazo, F., Gerber, J., Nau, R., Mitchell, T. J., and Wouters, F. S. (2009) Rapid microtubule bundling and stabilization by the Streptococcus pneumoniae neurotoxin pneumolysin in a cholesterol-dependent, non-lytic and Src-kinase dependent manner inhibits intracellular trafficking. Mol Microbiol 71: 461-477.

Jacobs, A. A. C., Loeffen, P. L. W., vandenBerg, A. J. G., and Storm, P. K. (1994) Identification, Purification, and Characterization of A Thiol-Activated Hemolysin (Suilysin) of Streptococcus Suis. Infection and Immunity 62: 1742-1748.

Jacobs, A. A. C., vandenBerg, A. J. G., and Loeffen, P. L. W. (1996) Protection of experimentally infected pigs by suilysin, the thiol-activated haemolysin of Streptococcus suis. Veterinary Record 139: 225-228.

Johnson, M. K., Geoffroy, C., and Alouf, J. E. (1980) Binding of cholesterol by sulfhydryl-activated cytolysins. Infect Immun 27: 97-101.

Jonsson, K., Signas, C., Muller, H. P., and Lindberg, M. (1991) Two different genes encode fibronectin binding proteins in Staphylococcus aureus. The complete nucleotide sequence and characterization of the second gene. Eur J Biochem 202: 1041-1048.

Jounblat, R., Kadioglu, A., Mitchell, T. J., and Andrew, P. W. (2003) Pneumococcal behavior and host responses during bronchopneumonia are affected differently by the cytolytic and complement- activating activities of pneumolysin. Infect Immun 71: 1813-1819.

Just, I., Selzer, J., Wilm, M., Eichel-Streiber, C., Mann, M., and Aktories, K. (1995) Glucosylation of Rho proteins by Clostridium difficile toxin B. Nature 375: 500-503.

Kim, D., Han, K., Oh, Y., Kim, C. H., Kang, I., Lee, J., Gottschalk, M., and Chae, C. (2010) Distribution of capsular serotypes and virulence markers of Streptococcus suis isolated from pigs with polyserositis in Korea. Can J Vet Res 74: 314-316.

King, S. J., Heath, P. J., Luque, I., Tarradas, C., Dowson, C. G., and Whatmore, A. M. (2001) Distribution and genetic diversity of suilysin in Streptococcus suis isolated from different diseases of pigs and characterization of the genetic basis of suilysin absence. Infection and Immunity 69: 7572-7582.

Kock, C., Beineke, A., Seitz, M., Ganter, M., Waldmann, K. H., Valentin-Weigand, P., and Baums, C. G. (2009) Intranasal immunization with a live Streptococcus suis isogenic ofs mutant elicited suilysin-neutralization titers but failed to induce opsonizing antibodies and protection. Veterinary Immunology and Immunopathology 132: 135-145.

Literature Chapter 9

179

Korchev, Y. E., Bashford, C. L., Pederzolli, C., Pasternak, C. A., Morgan, P. J., Andrew, P. W., and Mitchell, T. J. (1998) A conserved tryptophan in pneumolysin is a determinant of the characteristics of channels formed by pneumolysin in cells and planar lipid bilayers. Biochem J 329 ( Pt 3): 571-577.

Krawczyk-Balska, A. and Bielecki, J. (2005) Listeria monocytogenes listeriolysin O and phosphatidylinositol-specific phospholipase C affect adherence to epithelial cells. Can J Microbiol 51: 745-751.

Lafont, F. and van der Goot, F. G. (2005) Bacterial invasion via lipid rafts. Cell Microbiol 7: 613-620.

Lalonde, M., Segura, M., Lacouture, S., and Gottschalk, M. (2000) Interactions between Streptococcussuis serotype 2 and different epithelial cell lines. Microbiology 146 ( Pt 8): 1913- 1921.

Lecours, M. P., Gottschalk, M., Houde, M., Lemire, P., Fittipaldi, N., and Segura, M. (2011) Critical Role for Streptococcus suis Cell Wall Modifications and Suilysin in Resistance to Complement-Dependent Killing by Dendritic Cells. J Infect Dis 204: 919-929.

Logan, R. P., Robins, A., Turner, G. A., Cockayne, A., Borriello, S. P., and Hawkey, C. J. (1998) A novel flow cytometric assay for quantitating adherence of Helicobacter pylori to gastric epithelial cells. J Immunol Methods 213: 19-30.

Lowe, B. A., Marsh, T. L., Isaacs-Cosgrove, N., Kirkwood, R. N., Kiupel, M., and Mulks, M. H. (2011) Microbial communities in the tonsils of healthy pigs. Vet Microbiol 147: 346-357.

Lun, S. C., Perez-Casal, J., Connor, W., and Willson, P. J. (2003) Role of suilysin in pathogenesis of Streptococcus suis capsular serotype 2. Microbial Pathogenesis 34: 27-37.

Medina, E. (2010) Murine model of pneumococcal pneumonia. Methods Mol Biol 602: 405-410.

Mai, N. T., Hoa, N. T., Nga, T. V., Linh, le D., Chau, T. T., Sinh, D. X., Phu, N. H., Chuong, L. V., Diep, T. S., Campbell, J., Nghia, H. D., Minh, T. N., Chau, N. V., de Jong, M. D., Chinh, N. T., Hien, T. T., Farrar, J., and Schultsz, C. (2008) Streptococcus suis meningitis in adults in Vietnam. Clin Infect Dis 46: 659-667.

Malley, R., Henneke, P., Morse, S. C., Cieslewicz, M. J., Lipsitch, M., Thompson, C. M., Kurt-Jones, E., Paton, J. C., Wessels, M. R., and Golenbock, D. T. (2003) Recognition of pneumolysin by Toll-like receptor 4 confers resistance to pneumococcal infection. Proc Natl Acad Sci U S A 100: 1966-1971.

Mackay, D. J. and Hall, A. (1998) Rho GTPases. J Biol Chem 273: 20685-20688.

Messier, S., Lacouture, S., and Gottschalk, M. (2008) Distribution of Streptococcus suis capsular types from 2001 to 2007. Can Vet J 49: 461-462.

Michel, E., Reich, K. A., Favier, R., Berche, P., and Cossart, P. (1990) Attenuated mutants of the intracellular bacterium Listeria monocytogenes obtained by single amino acid substitutions in listeriolysin O. Mol Microbiol 4: 2167-2178.

Navacharoen, N., Chantharochavong, V., Hanprasertpong, C., Kangsanarak, J., and Lekagul, S. (2009) Hearing and vestibular loss in Streptococcus suis infection from swine and traditional raw pork exposure in northern Thailand. J Laryngol Otol 123: 857-862.

Literature Chapter 9

180

Ngo, T. H., Tran, T. B., Tran, T. T., Nguyen, V. D., Campbell, J., Pham, H. A., Huynh, H. T., Nguyen, V. V., Bryant, J. E., Tran, T. H., Farrar, J., and Schultsz, C. (2011) Slaughterhouse pigs are a major reservoir of Streptococcus suis serotype 2 capable of causing human infection in southern Vietnam. PLoS One 6: e17943.

Norton, P. M., Rolph, C., Ward, P. N., Bentley, R. W., and Leigh, J. A. (1999) Epithelial invasion and cell lysis by virulent strains of Streptococcus suis is enhanced by the presence of suilysin. Fems Immunology and Medical Microbiology 26: 25-35.

O'Sullivan, T., Friendship, R., Blackwell, T., Pearl, D., McEwen, B., Carman, S., Slavic, D., and Dewey, C. (2011) Microbiological identification and analysis of swine tonsils collected from carcasses at slaughter. Can J Vet Res 75: 106-111.

Okamoto, S., Kawabata, S., Terao, Y., Fujitaka, H., Okuno, Y., and Hamada, S. (2004) The Streptococcus pyogenes capsule is required for adhesion of bacteria to virus-infected alveolar epithelial cells and lethal bacterial-viral superinfection. Infect Immun 72: 6068-6075.

Okwumabua, O., Abdelmagid, O., and Chengappa, M. M. (1999) Hybridization analysis of the gene encoding a hemolysin (suilysin) of Streptococcus suis type 2: evidence for the absence of the gene in some isolates. FEMS Microbiol Lett 181: 113-121.

Orihuela, C. J., Gao, G., Francis, K. P., Yu, J., and Tuomanen, E. I. (2004) Tissue-specific contributions of pneumococcal virulence factors to pathogenesis. J Infect Dis 190: 1661-1669.

Padungtod, P., Tharavichitkul, P., Junya, S., Chaisowong, W., Kadohira, M., Makino, S., and Sthitmatee, N. (2010) Incidence and presence of virulence factors of Streptococcus suis infection in slaughtered pigs from Chiang Mai, Thailand. Southeast Asian J Trop Med Public Health 41: 1454-1461.

Pallares, F. J., Halbur, P. G., Schmitt, C. S., Roth, J. A., Opriessnig, T., Thomas, P. J., Kinyon, J. M., Murphy, D., Frank, D. E., and Hoffman, L. J. (2003) Comparison of experimental models for Streptococcus suis infection of conventional pigs. Can J Vet Res 67: 225-228.

Pan, X. Z., Ge, J. C., Li, M., Wu, B., Wang, C. J., Wang, J., Feng, Y. J., Yin, Z. M., Zheng, F., Cheng, G., Sun, W., Ji, H. F., Hu, D., Shi, P. J., Feng, X. D., Hao, X. N., Dong, R. P., Hu, F. Q., and Tang, J. Q. (2009) The Orphan Response Regulator CovR: a Globally Negative Modulator of Virulence in Streptococcus suis Serotype 2. Journal of Bacteriology 191: 2601-2612. Perch, B., Pedersen, K. B., and Henrichsen, J. (1983) Serology of capsulated streptococci pathogenic for pigs: six new serotypes of Streptococcus suis. J Clin Microbiol 17: 993-996.

Petersen, R., Hannerz, H., Tuchsen, F., and Egerton, J. R. (2011) Meningitis, sepsis and endocarditis among workers occupationally exposed to pigs. Occup Med (Lond).

Pinkney, M., Beachey, E., and Kehoe, M. (1989) The thiol-activated toxin streptolysin O does not require a thiol group for cytolytic activity. Infect Immun 57: 2553-2558.

Polekhina, G., Giddings, K. S., Tweten, R. K., and Parker, M. W. (2005) Insights into the action of the superfamily of cholesterol-dependent cytolysins from studies of intermedilysin. Proc Natl Acad Sci U S A 102: 600-605.

Princivalli, M. S., Palmieri, C., Magi, G., Vignaroli, C., Manzin, A., Camporese, A., Barocci, S., Magistrali, C., and Facinelli, B. (2009) Genetic diversity of Streptococcus suis clinical isolates from pigs and humans in Italy (2003-2007). Euro Surveill 14.

Literature Chapter 9

181

Que, Y. A., Haefliger, J. A., Francioli, P., and Moreillon, P. (2000) Expression of Staphylococcus aureus clumping factor A in Lactococcus lactis subsp. cremoris using a new shuttle vector. Infect Immun 68: 3516-3522.

Ramachandran, R., Tweten, R. K., and Johnson, A. E. (2004) Membrane-dependent conformational changes initiate cholesterol-dependent cytolysin oligomerization and intersubunit beta-strand alignment. Nat Struct Mol Biol 11: 697-705.

Ramachandran, R., Heuck, A. P., Tweten, R. K., and Johnson, A. E. (2002) Structural insights into the membrane-anchoring mechanism of a cholesterol-dependent cytolysin. Nat Struct Biol 9: 823- 827.

Rampersaud, R., Planet, P. J., Randis, T. M., Kulkarni, R., Aguilar, J. L., Lehrer, R. I., and Ratner, A. J. (2011) Inerolysin, a cholesterol-dependent cytolysin produced by Lactobacillus iners. J Bacteriol 193: 1034-1041.

Ratner, A. J., Hippe, K. R., Aguilar, J. L., Bender, M. H., Nelson, A. L., and Weiser, J. N. (2006) Epithelial cells are sensitive detectors of bacterial pore-forming toxins. J Biol Chem 281: 12994-12998.

Reid, T., Furuyashiki, T., Ishizaki, T., Watanabe, G., Watanabe, N., Fujisawa, K., Morii, N., Madaule, P., and Narumiya, S. (1996) Rhotekin, a new putative target for Rho bearing homology to a serine/threonine kinase, PKN, and rhophilin in the rho-binding domain. J Biol Chem 271: 13556-13560.

Ridley, A. J., Paterson, H. F., Johnston, C. L., Diekmann, D., and Hall, A. (1992) The small GTP- binding protein rac regulates growth factor-induced membrane ruffling. Cell 70: 401-410.

Robertson, I. D., Blackmore, D. K., Hampson, D. J., and Fu, Z. F. (1991) A longitudinal study of natural infection of piglets with Streptococcus suis types 1 and 2. Epidemiol Infect 107: 119- 126.

Rossjohn, J., Feil, S. C., McKinstry, W. J., Tweten, R. K., and Parker, M. W. (1997) Structure of a cholesterol-binding, thiol-activated cytolysin and a model of its membrane form. Cell 89: 685- 692.

Rossjohn, J., Gilbert, R. J., Crane, D., Morgan, P. J., Mitchell, T. J., Rowe, A. J., Andrew, P. W., Paton, J. C., Tweten, R. K., and Parker, M. W. (1998) The molecular mechanism of pneumolysin, a virulence factor from Streptococcus pneumoniae. J Mol Biol 284: 449-461.

Rossjohn, J., Polekhina, G., Feil, S. C., Morton, C. J., Tweten, R. K., and Parker, M. W. (2007) Structures of perfringolysin O suggest a pathway for activation of cholesterol-dependent cytolysins. J Mol Biol 367: 1227-1236.

Rubins, J. B., Paddock, A. H., Charboneau, D., Berry, A. M., Paton, J. C., and Janoff, E. N. (1998) Pneumolysin in pneumococcal adherence and colonization. Microb Pathog 25: 337-342.

Saeland, E., Vidarsson, G., and Jonsdottir, I. (2000) Pneumococcal pneumonia and bacteremia model in mice for the analysis of protective antibodies. Microb Pathog 29: 81-91.

Sambrook, J., Fritsch, E. F., and Maniatis, T. (1989) Molecular cloning: a laboratory manual, 2nd ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, N Y.

Sanford, S. E. and Tilker, M. E. (1982) Streptococcus suis type II-associated diseases in swine: observations of a one-year study. J Am Vet Med Assoc 181: 673-676.

Literature Chapter 9

182

Saunders, F. K., Mitchell, T. J., Walker, J. A., Andrew, P. W., and Boulnois, G. J. (1989) Pneumolysin, the thiol-activated toxin of Streptococcus pneumoniae, does not require a thiol group for in vitro activity. Infect Immun 57: 2547-2552.

Schmidt, G. and Aktories, K. (2000) Rho GTPase-activating toxins: cytotoxic necrotizing factors and dermonecrotic toxin. Methods Enzymol 325: 125-136.

Schwartz, M. A. and Shattil, S. J. (2000) Signaling networks linking integrins and rho family GTPases. Trends Biochem Sci 25: 388-391.

Scibelli, A., Matteoli, G., Roperto, S., Alimenti, E., Dipineto, L., Pavone, L. M., Della, Morte R., Menna, L. F., Fioretti, A., and Staiano, N. (2005) Flavoridin inhibits Yersinia enterocolitica uptake into fibronectin-adherent HeLa cells. FEMS Microbiol Lett 247: 51-57.

Scibelli, A., Roperto, S., Manna, L., Pavone, L. M., Tafuri, S., Della, Morte R., and Staiano, N. (2007) Engagement of integrins as a cellular route of invasion by bacterial pathogens. Vet J 173: 482- 491.

Segers, R. P., Kenter, T., de Haan, L. A., and Jacobs, A. A. (1998) Characterisation of the gene encoding suilysin from Streptococcus suis and expression in field strains. FEMS Microbiol Lett 167: 255-261.

Segura, M. and Gottschalk, M. (2002) Streptococcus suis interactions with the murine macrophage cell line J774: adhesion and cytotoxicity. Infect Immun 70: 4312-4322.

Segura, M., Stankova, J., and Gottschalk, M. (1999) Heat-killed Streptococcus suis capsular type 2 strains stimulate tumor necrosis factor alpha and interleukin-6 production by murine macrophages. Infect Immun 67: 4646-4654.

Segura, M., Vadeboncoeur, N., and Gottschalk, M. (2002) CD14-dependent and -independent cytokine and chemokine production by human THP-1 monocytes stimulated by Streptococcus suis capsular type 2. Clin Exp Immunol 127: 243-254.

Segura, M., Vanier, G., Al Numani, D., Lacouture, S., Olivier, M., and Gottschalk, M. (2006) Proinflammatory cytokine and chemokine modulation by Streptococcus suis in a whole-blood culture system. FEMS Immunol Med Microbiol 47: 92-106.

Seitz, M., Baums, CG, Gerhard, R, Just, I, Neis, C., Benga, L., Fulde, M., Nerlich, A, Rohde, M., Goethe, R., and Valentin-Weigand, P. Subcytolytic activity of suilysin promotes invasion of Streptococcus suis in HEp-2 epithelial cells by Rac-dependent activation of the actin cytoskeleton. Ref Type: submitted to Cellular Microbiology Sekino-Suzuki, N., Nakamura, M., Mitsui, K. I., and Ohno-Iwashita, Y. (1996) Contribution of individual tryptophan residues to the structure and activity of theta-toxin (perfringolysin O), a cholesterol- binding cytolysin. Eur J Biochem 241: 941-947.

Shepard, L. A., Shatursky, O., Johnson, A. E., and Tweten, R. K. (2000) The mechanism of pore assembly for a cholesterol-dependent cytolysin: formation of a large prepore complex precedes the insertion of the transmembrane beta-hairpins. Biochemistry 39: 10284-10293.

Shimada, Y., Maruya, M., Iwashita, S., and Ohno-Iwashita, Y. (2002) The C-terminal domain of perfringolysin O is an essential cholesterol-binding unit targeting to cholesterol-rich microdomains. Eur J Biochem 269: 6195-6203.

Literature Chapter 9

183

Silva, L. M., Baums, C. G., Rehm, T., Wisselink, H. J., Goethe, R., and Valentin-Weigand, P. (2006) Virulence-associated gene profiling of Streptococcus suis isolates by PCR. Vet Microbiol 115: 117-127.

Smith, H. E., Damman, M., van der Velde, J., Wagenaar, F., Wisselink, H. J., Stockhofe-Zurwieden, N., and Smits, M. A. (1999) Identification and characterization of the cps locus of Streptococcus suis serotype 2: the capsule protects against phagocytosis and is an important virulence factor. Infect Immun 67: 1750-1756.

Smith, H. E., Vecht, U., Wisselink, H. J., StockhofeZurwieden, N., Biermann, Y., and Smits, M. A. (1996) Mutants of Streptococcus suis types 1 and 2 impaired in expression of muramidase- released protein and extracellular protein induce disease in newborn germfree pigs. Infection and Immunity 64: 4409-4412.

Su, Y., Yao, W., Perez-Gutierrez, O. N., Smidt, H., and Zhu, W. Y. (2008) Changes in abundance of Lactobacillus spp. and Streptococcus suis in the stomach, jejunum and ileum of piglets after weaning. FEMS Microbiol Ecol 66: 546-555.

Soltani, C. E., Hotze, E. M., Johnson, A. E., and Tweten, R. K. (2007) Specific protein-membrane contacts are required for prepore and pore assembly by a cholesterol-dependent cytolysin. J Biol Chem 282: 15709-15716.

Soltani, C. E., Hotze, E. M., Johnson, A. E., and Tweten, R. K. (2007b) Structural elements of the cholesterol-dependent cytolysins that are responsible for their cholesterol-sensitive membrane interactions. Proc Natl Acad Sci U S A 104: 20226-20231.

Sriskandan, S. and Slater, J. D. (2006) Invasive disease and toxic shock due to zoonotic Streptococcus suis: an emerging infection in the East? PLoS Med 3: e187.

Staats, J. J., Feder, I., Okwumabua, O., and Chengappa, M. M. (1997) Streptococcus suis: past and present. Vet Res Commun 21: 381-407.

Staats, J. J., Plattner, B. L., Stewart, G. C., and Chengappa, M. M. (1999) Presence of the Streptococcus suis suilysin gene and expression of MRP and EF correlates with high virulence in Streptococcus suis type 2 isolates. Veterinary Microbiology 70: 201-211.

Stachowiak, R., Wisniewski, J., Osinska, O., and Bielecki, J. (2009) Contribution of cysteine residue to the properties of Listeria monocytogenes listeriolysin O. Can J Microbiol 55: 1153-1159.

Strangmann, E., Froleke, H., and Kohse, K. P. (2002) Septic shock caused by Streptococcus suis: case report and investigation of a risk group. Int J Hyg Environ Health 205: 385-392.

Sukeno, A., Nagamune, H., Whiley, R. A., Jafar, S. I., Aduse-Opoku, J., Ohkura, K., Maeda, T., Hirota, K., Miyake, Y., and Kourai, H. (2005) Intermedilysin is essential for the invasion of hepatoma HepG2 cells by Streptococcus intermedius. Microbiol Immunol 49: 681-694.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2002) Evidence for lateral transfer of the Suilysin gene region of Streptococcus suis. J Bacteriol 184: 2050-2057.

Takamatsu, D., Osaki, M., and Sekizaki, T. (2001) Thermosensitive suicide vectors for gene replacement in Streptococcus suis. Plasmid 46: 140-148.

Tang, J., Wang, C., Feng, Y., Yang, W., Song, H., Chen, Z., Yu, H., Pan, X., Zhou, X., Wang, H., Wu, B., Wang, H., Zhao, H., Lin, Y., Yue, J., Wu, Z., He, X., Gao, F., Khan, A. H., Wang, J., Zhao, G. P., Wang, Y., Wang, X., Chen, Z., and Gao, G. F. (2006) Streptococcal toxic shock syndrome caused by Streptococcus suis serotype 2. PLoS Med 3: e151.

Literature Chapter 9

184

Tarradas, C., Borge, C., Arenas, A., Maldonado, A., Astorga, R., Miranda, A., and Luque, I. (2001) Suilysin production by Streptococcus suis strains isolated from diseased and healthy carrier pigs in Spain. Vet Rec 148: 183-184.

Tenenbaum, T., Adam, R., Eggelnpohler, I., Matalon, D., Seibt, A., GE, K. Novotny, Galla, H. J., and Schroten, H. (2005) Strain-dependent disruption of blood-cerebrospinal fluid barrier by Streptoccocus suis in vitro. FEMS Immunol Med Microbiol 44: 25-34.

Tenenbaum, T., Essmann, F., Adam, R., Seibt, A., Janicke, R. U., Novotny, G. E., Galla, H. J., and Schroten, H. (2006) Cell death, caspase activation, and HMGB1 release of porcine choroid plexus epithelial cells during Streptococcus suis infection in vitro. Brain Res 1100: 1-12.

Tenenbaum, T., Papandreou, T., Gellrich, D., Friedrichs, U., Seibt, A., Adam, R., Wewer, C., Galla, H. J., Schwerk, C., and Schroten, H. (2008) Polar bacterial invasion and translocation of Streptococcus suis across the blood-cerebrospinal fluid barrier in vitro. Cell Microbiol.

Tian, Y., Aarestrup, F. M., and Lu, C. P. (2004) Characterization of Streptococcus suis serotype 7 isolates from diseased pigs in Denmark. Vet Microbiol 103: 55-62.

Tomoyasu, T., Tabata, A., Hiroshima, R., Imaki, H., Masuda, S., Whiley, R. A., Aduse-Opoku, J., Kikuchi, K., Hiramatsu, K., and Nagamune, H. (2010) Role of catabolite control protein A in the regulation of intermedilysin production by Streptococcus intermedius. Infect Immun 78: 4012- 4021.

Tran, Van Nhieu, Caron, E., Hall, A., and Sansonetti, P. J. (1999) IpaC induces actin polymerization and filopodia formation during Shigella entry into epithelial cells. EMBO J 18: 3249-3262.

Trottier, S., Higgins, R., Brochu, G., and Gottschalk, M. (1991) A Case of Human Endocarditis Due to Streptococcus Suis in North-America. Reviews of Infectious Diseases 13: 1251-1252.

Tsuchiya, K., Kawamura, I., Takahashi, A., Nomura, T., Kohda, C., and Mitsuyama, M. (2005) Listeriolysin O-induced membrane permeation mediates persistent interleukin-6 production in Caco-2 cells during Listeria monocytogenes infection in vitro. Infect Immun 73: 3869-3877.

Tweten, R. K. (2005) Cholesterol-dependent cytolysins, a family of versatile pore-forming toxins. Infection and Immunity 73: 6199-6209.

Vadeboncoeur, C. and Pelletier, M. (1997) The phosphoenolpyruvate:sugar phosphotransferase system of oral streptococci and its role in the control of sugar metabolism. Fems Microbiology Reviews 19: 187-207. Vadeboncoeur, N., Segura, M., Al Numani, D., Vanier, G., and Gottschalk, M. (2003) Pro-inflammatory cytokine and chemokine release by human brain microvascular endothelial cells stimulated by Streptococcus suis serotype 2. FEMS Immunol Med Microbiol 35: 49-58.

Valentin-Weigand, P., Benkel, P., Rohde, M., and Chhatwal, G. S. (1996) Entry and intracellular survival of group B streptococci in J774 macrophages. Infect Immun 64: 2467-2473.

Vanier, G., Segura, M., Friedl, P., Lacouture, S., and Gottschalk, M. (2004) Invasion of porcine brain microvascular endothelial cells by Streptococcus suis serotype 2. Infection and Immunity 72: 1441-1449.

Vanier, G., Segura, M., and Gottschalk, M. (2007) Characterization of the invasion of porcine endothelial cells by Streptococcus suis serotype 2. Can J Vet Res 71: 81-89.

Literature Chapter 9

185

Vanier, G., Segura, M., Lecours, M. P., Grenier, D., and Gottschalk, M. (2008) Porcine brain microvascular endothelial cell-derived interleukin-8 is first induced and then degraded by Streptococcus suis. Microb Pathog.

Vanier, G., Fittipaldi, N., Slater, J. D., Dominguez-Punaro, Mde L., Rycroft, A. N., Segura, M., Maskell, D. J., and Gottschalk, M. (2009) New putative virulence factors of Streptococcus suis involved in invasion of porcine brain microvascular endothelial cells. Microb Pathog 46: 13-20.

Vecht, U., Wisselink, H. J., Jellema, M. L., and Smith, H. E. (1991) Identification of 2 Proteins Associated with Virulence of Streptococcus Suis Type-2. Infection and Immunity 59: 3156- 3162.

Vecht, U., StockhofeZurwieden, N., Tetenburg, B. J., Wisselink, H. J., and Smith, H. E. (1997) Virulence of Streptococcus suis type 2 for mice and pigs appeared host-specific. Veterinary Microbiology 58: 53-60.

Waheed, A. A., Shimada, Y., Heijnen, H. F., Nakamura, M., Inomata, M., Hayashi, M., Iwashita, S., Slot, J. W., and Ohno-Iwashita, Y. (2001) Selective binding of perfringolysin O derivative to cholesterol-rich membrane microdomains (rafts). Proc Natl Acad Sci U S A 98: 4926-4931.

Walker, J. A., Allen, R. L., Falmagne, P., Johnson, M. K., and Boulnois, G. J. (1987) Molecular cloning, characterization, and complete nucleotide sequence of the gene for pneumolysin, the sulfhydryl-activated toxin of Streptococcus pneumoniae. Infect Immun 55: 1184-1189.

Wangkaew, S., Chaiwarith, R., Tharavichitkul, P., and Supparatpinyo, K. (2006) Streptococcus suis infection: a series of 41 cases from Chiang Mai University Hospital. J Infect 52: 455-460.

Watson, K. C., Rose, T. P., and Kerr, E. J. (1972) Some factors influencing the effect of cholesterol on streptolysin O activity. J Clin Pathol 25: 885-891.

Wei, Z., Li, R., Zhang, A., He, H., Hua, Y., Xia, J., Cai, X., Chen, H., and Jin, M. (2009) Characterization of Streptococcus suis isolates from the diseased pigs in China between 2003 and 2007. Vet Microbiol 137: 196-201.

Weis, S. and Palmer, M. (2001) Streptolysin O: the C-terminal, tryptophan-rich domain carries functional sites for both membrane binding and self-interaction but not for stable oligomerization. Biochim Biophys Acta 1510: 292-299.

Wertheim, H. F., Nghia, H. D., Taylor, W., and Schultsz, C. (2009) Streptococcus suis: an emerging human pathogen. Clin Infect Dis 48: 617-625.

Willenborg, J., Fulde, M., de Greeff, A., Rohde, M., Smith, H. E., Valentin-Weigand, P., and Goethe, R. (2011) Role of glucose and CcpA in capsule expression and virulence of Streptococcus suis. Microbiology 157: 1823-1833.

Williams, A. E. and Blakemore, W. F. (1990) Pathogenesis of meningitis caused by Streptococcus suis type 2. J Infect Dis 162: 474-481.

Williams, A. E., Blakemore, W. F., and Alexander, T. J. (1988) A murine model of Streptococcus suis type 2 meningitis in the pig. Res Vet Sci 45: 394-399. Windsor, R. S. and Elliott, S. D. (1975) Streptococcal infection in young pigs. IV. An outbreak of streptococcal meningitis in weaned pigs. J Hyg (Lond) 75: 69-78.

Literature Chapter 9

186

Wisselink, H. J., Smith, H. E., Stockhofe-Zurwieden, N., Peperkamp, K., and Vecht, U. (2000) Distribution of capsular types and production of muramidase-released protein (MRP) and extracellular factor (EF) of Streptococcus suis strains isolated from diseased pigs in seven European countries. Veterinary Microbiology 74: 237-248.

Wu, T., Chang, H., Tan, C., Bei, W., and Chen, H. (2009) The orphan response regulator RevSC21 controls the attachment of Streptococcus suis serotype-2 to human laryngeal epithelial cells and the expression of virulence genes. FEMS Microbiol Lett 292: 170-181.

Wu, T., Zhao, Z., Zhang, L., Ma, H., Lu, K., Ren, W., Liu, Z., Chang, H., Bei, W., Qiu, Y., and Chen, H. (2011) Trigger factor of Streptococcus suis is involved in stress tolerance and virulence. Microb Pathog 51: 69-76.

Xiong, Y., Liu, Q., Qin, F. Y., Bai, Y., Zhu, W., Li, H. M., Guo, J. G., Qin, L., Pan, J., Long, J. M., and Chen, L. (2007) [Study on the molecular epidemiology of Streptococcus suis type 2 from healthy pigs in Guangxi]. Zhonghua Liu Xing Bing Xue Za Zhi 28: 593-596.

Xu, L., Huang, B., Du, H., Zhang, X. C., Xu, J., Li, X., and Rao, Z. (2010) Crystal structure of cytotoxin protein suilysin from Streptococcus suis. Protein Cell 1: 96-105.

Ye, C., Zheng, H., Zhang, J., Jing, H., Wang, L., Xiong, Y., Wang, W., Zhou, Z., Sun, Q., Luo, X., Du, H., Gottschalk, M., and Xu, J. (2009) Clinical, Experimental, and Genomic Differences between Intermediately Pathogenic, Highly Pathogenic, and Epidemic Streptococcus suis. J Infect Dis 199: 97-107.

Ye, C., Zhu, X., Jing, H., Du, H., Segura, M., Zheng, H., Kan, B., Wang, L., Bai, X., Zhou, Y., Cui, Z., Zhang, S., Jin, D., Sun, N., Luo, X., Zhang, J., Gong, Z., Wang, X., Wang, L., Sun, H., Li, Z., Sun, Q., Liu, H., Dong, B., Ke, C., Yuan, H., Wang, H., Tian, K., Wang, Y., Gottschalk, M., and Xu, J. (2006) Streptococcus suis sequence type 7 outbreak, Sichuan, China. Emerg Infect Dis 12: 1203-1208.

Yu, H., Jing, H., Chen, Z., Zheng, H., Zhu, X., Wang, H., Wang, S., Liu, L., Zu, R., Luo, L., Xiang, N., Liu, H., Liu, X., Shu, Y., Lee, S. S., Chuang, S. K., Wang, Y., Xu, J., and Yang, W. (2006) Human Streptococcus suis outbreak, Sichuan, China. Emerg Infect Dis 12: 914-920.

Zhao, Y., Liu, G., Li, S., Wang, M., Song, J., Wang, J., Tang, J., Li, M., and Hu, F. (2011) Role of a type IV-like secretion system of Streptococcus suis 2 in the development of streptococcal toxic shock syndrome. J Infect Dis 204: 274-281.

Zheng, H., Punaro, M. C., Segura, M., Lachance, C., Rivest, S., Xu, J., Houde, M., and Gottschalk, M. (2011) Toll-like receptor 2 is partially involved in the activation of murine astrocytes by Streptococcus suis, an important zoonotic agent of meningitis. J Neuroimmunol 234: 71-83.

Acknowledgments First and foremost, I would like to sincerely thank my supervisor, Prof. Dr. Peter Valentin-Weigand, for offering me the opportunity of performing this work, for his support and invaluable advice. I would like to thank my co-supervisors Prof. Dr. Sike Rautenschlein and PD Dr. Manfred Rohde for their interest in my work, their ideas and productive discussions. I would like to thank PD Dr. Christoph Baums for his competence and precious advises and Prof. Dr. Ralf Goethe for his excellent support and everyday supervision. Many thanks, to Prof. Dr. Ingo Just and PD Dr. Ralf Gerhard for good cooperation and expertise in G-proteins, PD Dr. Manfred Rohde for providing the wonderful electron micrographs, Prof. Dr. Andreas Beineke for the histopathological examination of endless mice organs and Dr. Oliver Goldmann for his assistance with the intranasal application techniques and anaesthesia of mice. I would like to thank all the past and present members of the Institute for Microbiology, for an excellent collegial and productive working atmosphere and for never getting tired of laughing with and at me: Jens Abel, Tina Basler, Tanja Bauer, Sabine Baumert, Laurentiu Benga, Nadine Büttner, Sabrina Diehl, Elke Eckelt, Marcus Fulde, Sabine Göbel, Nina Janze, Christoph Kock, Jochen Meens, Kristin Laarmann, Thorsten Meißner, Andreas Mietze, Franziska Müller, Christina Neis, Andreas Nerlich, Girish Ramachandran, Nantaporn Ruangkiattikul, Julian Sander, Franziska Schill, Jana Seele, Mathias Weigoldt, Yenehiwot Berhanu Weldearegay, Joerg Willenborg and Daniela Willms. Many thanks, to Jörg Merkel for helping me with all technical questions concerning computer and statistic trouble. Finally and most important, I would like to thank my parents, grandma and Jörg, for their constant support in all respects, for their love and encouragement all over the years, and allowing me to pursue my dreams.