• No results found

Structural studies of the surface adhesin SspB from Streptococcus gordonii

N/A
N/A
Protected

Academic year: 2022

Share "Structural studies of the surface adhesin SspB from Streptococcus gordonii"

Copied!
43
0
0

Loading.... (view fulltext now)

Full text

(1)
(2)
(3)

Structural Studies of the Surface Adhesin SspB from Streptococcus gordonii

Nina Forsgren

Department of Odontology, Cariology 901 87 Umeå

Umeå 2010

(4)

Copyright©Nina Forsgren ISBN: 978-91-7264-955-2 New series no. 111 ISSN: 0345-7532 Printed by: Print & Media Umeå, Sweden 2010

(5)
(6)

players in the formation of the biofilm. Antigen I/II proteins are surface adhesins found on virtually all oral streptococci and share a conserved multi- domain architecture. These adhesins bind surface components on other bacteria and on host cells. Thus, they are crucial for the development of the biofilm.

The objective of this thesis work is the structural characterization of the large multi-domain Antigen I/II protein SspB from the primary colonizing commensal bacterium Streptococcus gordonii.

The crystal structure of the variable domain of SspB was determined to 2.3 Å resolution. The domain comprises a β-supersandwich and a putative binding cleft stabilized by a calcium ion. Despite high similarity in the overall structure, the cleft within SspB is significantly smaller than the cleft within the homologous protein from Streptococcus mutans, indicating that different substrates may bind in the clefts. A screen for carbohydrate binding resulted in no hits for interaction with the SspB variable domain suggesting that the cleft may not be suitable for binding sugars.

This thesis also presents the high resolution 1.5 Å structure of a truncated C-terminal domain of SspB, the first of an Antigen I/II C-domain. The structure contains two structurally related domains, each containing one calcium ion and one intramolecular isopeptide bond. The SspB protein shares the feature of intramoleular isopeptide bonds with other surface proteins from Gram positive bacteria, such as pili from Streptococcus pyogenes and Corynebacterium diphtheriae. Intramolecular isopeptide bonds are suggested to be a common feature for retaining stability in a harsh environment. The SspB adherence region, shown to be the recognition motif for Porphyromonas gingivalis attachment to S. gordonii, protrudes from the core protein as a handle available for recognition.

In conclusion, this thesis work has provided new knowledge about the SspB protein and increased the understanding of the common structure of AgI/II proteins.

(7)

Table of contents

LIST OF PUBLICATIONS ... 7

INTRODUCTION ... 8

BIOFILMS, DENTAL PLAQUE AND MICROBIAL ECOLOGY ... 8

Biofilms ... 8

Dental plaque ... 8

Shift from health to disease ... 9

Adhesion and coaggregation in the oral biofilm ... 10

Streptococci ... 11

Streptococcus gordonii ... 11

Stabilization of surface proteins from Gram-positive bacteria ... 12

Antigen I/II adhesins ... 13

Interaction between Porphyromonas gingivalis and Streptococcus gordonii ... 15

STRUCTURAL BIOLOGY AND GENERAL METHODS ... 15

Why structural biology? ... 15

Cloning, expression and protein purification ... 16

Crystallization ... 16

Structure determination ... 18

AIMS ... 20

RESULTS AND DISCUSSION ... 21

THE VARIABLE DOMAIN OF SSPB FROM STREPTOCOCCUS GORDONII (PAPER I) 21

The structure of the SspB V-domain ... 21

The putative binding pocket ... 21

SspB-V in comparison with SpaP-V ... 23

THE C-TERMINAL DOMAIN OF THE SURFACE ADHESIN SSPB FROM STREPTOCOCCUS GORDONII (PAPER II AND III) ... 24

The structure of the proteolytically stable SspB C-terminal domain ... 24

Intramolecular isopeptide bonds ... 24

Similar structures ... 26

P. gingivalis interaction motifs ... 27

AGI/II DOMAIN INTERACTION WITH GP340(UNPUBLISHED DATA) ... 27

The gp340 overlay assay ... 28

AgI/II interaction with gp340 ... 28

CONCLUSIONS ... 30

ACKNOWLEDGEMENTS ... 31

REFERENCES ... 33

(8)
(9)

List of Publications

The original papers that comprise this thesis are listed below and will be referred to by Roman numerals in the text.

I. Forsgren, N., Lamont, R. J., Persson, K. Crystal Structure of the Variable Domain of the Streptococcus gordonii Surface Protein SspB. (2009). Prot. Science 18, 1896-1905.

II. Forsgren, N., Lamont, R. J., Persson, K. A Crystallizable Form of the C-terminal Domain of Streptococcus gordonii Surface Protein SspB Obtained by Limited Proteolysis. (2009). Acta. Cryst. F65, 712- 714.

III. Forsgren, N., Lamont, R.J., Persson, K. Two Intramolecular Isopeptide Bonds Are Identified in the Crystal Structure of the Streptococcus gordonii SspB C-terminal Domain. (2010). J. Mol.

Biol. 397, 740-751.

Structure coordinates from this thesis can be found at the RCSB protein data bank, accession codes: 2WD6, 2WOY, 2WQS and 2WZA.

Reprints were made with permission from the publishers.

(10)

Introduction

Bacterial infection within the oral cavity is a fundamental factor for the development of dental diseases, such as caries, gingivitis and periodontitis, that affect a major part of the population. Bacteria that are accountable for the development of bacterial infections are found within the dental plaque (oral biofilm) present on tooth and mucosal surfaces in the oral cavity.

Biofilms, dental plaque and microbial ecology

Biofilms

Biofilms can be defined as highly structured microbial communities consisting of single or multiple species, embedded in an extracellular polymer matrix. In natural environments, 95-99% of microorganisms exist as biofilms (Nikolaev and Plakunov, 2007). Biofilms are found on many of the epithelial cells in the human body, such as within the oral cavity, upper- respiratory tract and gastro-intestinal tract. Biofilms can also coat superficial surfaces introduced into the body, such as implants and catheters (O‟Toole et al., 2000), causing medical complications.

In general, biofilm development starts by planktonic (free-floating) cells being reversibly fastened to a surface through non-specific physiochemical forces (such as van der Waals‟, hydrophobic and electrostatic forces) (Stanley and Lazazzera, 2004; Nikolaev and Plakunov, 2007). Next, adsorbed cells attach to the surface through higher affinity interactions involving specific adhesins and their associated interaction partners. Then, cells adhere to each other by aggregation and additional planktonic cells are recruited, leading to the formation of microcolonies. Simultaneously, extracellular polymeric matrix begins to form, embedding the cells. Microcolonies proliferate and secondary colonizers arrive, building up an architecturally distinctive mature biofilm with channels, cavities and pores. Environmental factors influence the characteristics of the biofilm. Among the most important factors are pH, oxygen availability, osmolarity, nutrient sources and the force of liquid motion (Nikolaev and Plakunov, 2007).

Dental plaque

More than 700 species of micro-organisms have been found within oral biofilms (Jenkinson and Lamont, 2005). Since the 1960s, several reports have provided knowledge on how the dental biofilm matures over time and how the composition changes (Ritz, 1967; Theilade et al., 1966; Listgarten, 1966).

Several studies have, by using different methods, identified Gram- positive streptococci as comprising the majority of the bacterial population

(11)

in initial supragingival oral bacterial communities (Figure 1). Nyvad and Kilian (1987) reported that a variety of other bacteria, such as Veillonellae and Actinomyces, were also present in the culturable microflora colonizing the enamel. By molecular characterization of the microflora, very low levels of periodontal pathogens, such as Tanerella forsythia, Porphyromonas gingivalis and Prevotella spp. were also detected in early plaque (Diaz et al., 2006; Li et al., 2004).

Subgingival plaque is formed from the migration of biofilm growing on the tooth surface into the gingival sulcus (Theilade and Theilade, 1985).

Gram-negative bacteria and spirochetes dominate the subgingival biofilm, and among species found are the periodontal pathogens P. gingivalis, T.

forsythia and Treponema denticola (Socransky and Haffajee, 2005).

Shift from health to disease

The healthy oral cavity mainly harbours bacteria such as Actinomyces, Streptococcus and Veillonella (Moore et al., 1987; Aas et al., 2005). These can be regarded as commensal bacteria and can serve as a protective film.

Streptococci and Actinomyces are pioneer organisms in the initiation of biofilm development as they colonize and attach to the tooth surface via salivary proteins and peptides, such as proline-rich proteins, statherin, salivary mucins and agglutinins, coating the tooth (Gibbons, 1989; Cisar et al., 1988; Jakubovics et al., 2005a; Loimaranta et al., 2005). Through co- adhesive, signalling, nutritional and metabolic interactions, these pioneer species provide a new environment and substrates for late colonizers, including Gram-negative anaerobes considered to be etiological agents of periodontal diseases (Rickard et al., 2003)(Figure 2). Changes in the oral cavity environment, such as lowering of the pH, changes in nutritional composition, reduced oxygen tension and availability of adhesion molecules, can cause an ecological shift in the community composition, from a healthy to a more pathogenic biofilm (Lamont and Jenkinson, 2005; Marsh, 2006).

Thus, bacteria with pathogenic potential that are present in the commensal biofilm, though at very low numbers, may benefit from the environmental changes and proliferate. The infectious diseases in the mouth, such as caries

Tooth

Supragingival plaque

Subgingival plaque Gingival margin

Figure 1. A simplified view of dental plaque close to the tooth.

(12)

and periodontitis, are infections characterized by such an ecological shift in the biofilm.

The disease risk associated with the presence of certain organisms and the absence of others is well established (Socransky and Hajaffee, 2005).

Although single species have been shown to be implicated in infection, such as Streptococcus mutans in caries and P. gingivalis in periodontitis, it is important to consider the oral infections as mixed-species infections rather than focusing only on specific species (Filoche et al., 2010).

Figure 2. Simplified view of the oral biofilm. Actinomyces (A), Streptococcus (S), S. gordonii (S.g.) and S. mutans (S.m.) interact with salivary and host molecules.

Secondary colonizers such as F. nucleatum (F.n.) and P. gingivalis (P.g.) can attach.

Adhesion and coaggregation in the oral biofilm

Bacteria must attach to a surface or they will be swallowed or expectorated and removed from the oral cavity. Adherence is a dynamic process; following initial adherence bacteria will grow and survive if the physical and chemical environment, e.g. pH and oxygen levels, are favourable. If the growth conditions become less favourable, bacteria may detach from the surface.

Bacteria interact with each other through co-adhesive, signalling, nutritional and metabolic interactions. The partnerships between the bacteria are highly specific. The primary colonizers can co-aggregate with each other, but usually not with late colonizers. Fusobacterium nucleatum has receptors for both early and late colonizers, and can be regarded as a co-aggregation bridge organism (Kolenbrander et al., 2006). In the absence of F. nucleatum, many secondary colonizers cannot become part of the biofilm community (Bradshaw et al., 1998). One exception, however, is the late colonizer P.

gingivalis which is not dependent upon F. nucleatum to interact with the primary colonizer Streptococcus gordonii (Lamont et al., 2002).

A S

S.m.

A

S S.g.

F.n.

S

P.g.

Fimbria AgI/II

Salivary proteins

Salivary and host molecules

(13)

Streptococci

Streptococcus comes from the Greek words strepto (twisted) and coccus (spherical). Today, over 100 species of Streptococcus are recognized. Most species of streptococci are considered commensals, and are usually found at mucosal surfaces, such as those within the oral cavity, upper-respiratory tract and gastro-intestinal tract. Commensal streptococci coat the mucosa and protect the host from adherence and metabolism of pathogens.

Based on 16S rRNA gene sequence comparisons, streptococci can be divided into six groups (Kilian, 2005). The pyogenic group includes for example, Streptococcus pyogenes and Streptococcus agalactiae. The anginosus and salivarius groups include mainly human and animal oral cavity microbes. The bovis group include Streptococcus infantarius (previously Streptococcus bovis) and Streptococcus gallonlyticus, which are cow and sheep oral microbes. The mitis group includes for example, S.

gordonii, Streptococcus mitis and Streptococcus pneumoniae. S. mutans and Streptococcus sobrinus are members of the mutans group, containing bacteria colonizing the oral cavity of humans.

Though many are considered commensals, streptococci can, during appropriate conditions, initiate local or systemic infections/diseases. Among the oral streptococci, S. mutans is implicated in the infection causing caries (Loesche, 1986), while S. gordonii has pathogenic potential as an etiological agent in endocarditis (Douglas et al., 1993). Other streptococci of more pathogenic character are for example, S. pyogenes (a group A Streptococcus) that can cause infection leading to rheumatic fever and glomerulonephritis (Cunningham, 2008). One of the most frequent microbial causes of death in humans is S. pneumoniae, a pathogen associated with diseases such as bronchitis, meningitis and pneumonia (Fedson et al., 1999).

Streptococci are versatile when it comes to occupying adherence sites because they present an array of surface proteins that enable them to interact with multiple host components. Among the surface adhesins are serine-rich repeat proteins, pili and fimbriae, salivary protein binders, glucan-binding proteins, collagen-binding proteins, and fibronectin-binding proteins (Nobbs et al., 2009).

Streptococcus gordonii

S. gordonii is a commensal bacterium which colonizes multiple sites in the human oral cavity. It was found to inhibit the release of IL-6 and IL-8 from epithelial cells (Hasegawa et al., 2007), as well as to suppress the release of IL-8 normally induced by the periodontopathogen F. nucleatum (Zhang et al., 2008). This suppression of cytokine release provides insights into how this bacterium can exist as a commensal and how streptococci might promote a “healthy” oral environment (Jenkinson and Lamont, 2005; Nobbs et al., 2009).

(14)

S. gordonii expresses a number of cell-surface associated proteins with which it interacts with a variety of host proteins and other bacteria. CshA and CshB bind fibronectin (McNab et al., 1996; McNab et al., 1994), AbpA binds amylase (Rogers et al., 1998), serine-rich Hsa (strain Challis (DL1)) and GspB (strain M99) bind to cell surface glycoproteins on leukocytes (Urano-Tashiro et al., 2008) and platelets (Bensing and Sullam, 2002;

Takamatsu et al., 2005). SspA and SspB adhesins, members of the Antigen I/II (AgI/II) family of proteins and products of tandemly arranged genes (Demuth et al., 1996), are also found attached to the surface of S. gordonii.

The interaction properties of S. gordonii are important for initial adhesion to saliva-coated surfaces, co-aggregation and colonization, processes which can contribute to a protective biofilm. However, the streptococcal biofilm is used as an attachment surface by other organisms, such as F. nucleatum and the secondary colonizer P. gingivalis (Lamont et al., 1992). During favourable conditions, these secondary colonizers can proliferate and initiate infection that can lead to various dental diseases (Rosan and Lamont, 2000).

Hence, S. gordonii biofilms play an important role in the development of biofilms associated with health, but can also contribute indirectly to the development of pathogenic biofilms.

Stabilization of surface proteins from Gram-positive bacteria

Surface proteins need to have a stable structure to withstand the challenges of adhesion to host cells and the environment in which they perform their tasks. Recent studies of pili (elongated flexible appendages) and their associated adhesins have provided evidence of a stabilization mechanism. It is accepted that pili in Gram-positive bacteria are formed through covalent associations of major pilin subunits by the action of transpeptidase enzymes called sortases (Budzik et al., 2008; Mandlik et al., 2008; Mora et al., 2005;

Ton-That et al., 2004; Ton-That and Schneewind, 2003). Recently, Kang et al. (2007) solved the crystal structure of the backbone pilin subunit Spy0128 from S. pyogenes. The protein has an extended all-β two-domain structure.

Interestingly, they observed one intramolecular covalent bond in each domain tying together the first and last β-strand of the domains. The covalent bonds are formed between lysine and asparagines side chains, and stabilized by a glutamate residue which hydrogen-bonds with the isopeptide.

The authors concluded that the intramolecular bonds appear to be self- generated. A sequence-comparison search identified intramolecular isopeptide bonds in other cell surface proteins, all from Gram-positive bacteria. Unrecognized isopeptide bonds were also found in two previously crystallized proteins, the minor pilin subunit GBS52 from S. agalactiae (Krishnan et al., 2007), and the collagen-binding adhesin Cna from Staphylococcus aureus (Deivanayagam et al., 2000; Zong et al., 2005). Since

(15)

the first identification of intramolecular isopeptide bonds in cell surface proteins, the newly solved crystal structures of Corynebacterium diphtheriae shaft pilin SpaA (Kang et al., 2009), the pilus adhesin RrgA from S. pneumoniae (Izoré et al., 2010), the BcpA major pilin subunit from Bacillus cereus (Budzik et al., 2008; Budzik et al., 2009) and the S. gordonii SspB C-terminal domain (Paper III) have added to the list of proteins containing these bonds. Kang et al. (2007) hypothesized that the intramolecular isopeptide bonds are a mode of stabilization of surface proteins involved in host pathogenesis. By studying mutants of Spy0128 that lacked either one or both isopeptide bonds, it was revealed that the loss of isopeptide bonds compromised both thermal stability and proteolytic stability (Kang and Baker, 2009). In addition, mutants of the S. pneumoniae pilus adhesin RrgA lacking isopeptide bonds were also shown to be compromised in thermal stability (Izoré et al., 2010). The existence of intramolecular isopeptide bonds in several surface proteins confirms that these crosslinks are a common feature of Gram-positive surface proteins.

Antigen I/II adhesins

Surface proteins are crucial for colonization and for the interaction with the surrounding environment. Thus, surface proteins are important in determining the success of survival for a bacterial strain in its competition with others.

Surface adhesins of the Antigen I/II family are expressed by virtually all oral (viridans) streptococci. Through their AgI/II adhesins, streptococci interact with a variety of different human host molecules, such as collagen, fibronectin and gp340/SAG, and with other bacteria, such as Actinomyces naeslundii, Streptococcus oralis, P. gingivalis, and Candida albicans (Jacubovics et al., 2005; Holmes et al., 1996; Jenkinson and Demuth, 1997;

Prakobphol et al., 2000; Brady et al., 1992; Jenkinson and Lamont, 2005).

This multifunctionality of AgI/II proteins indicates that these proteins may contribute to both initial adherence and community development.

Figure 3. AgI/II domain architecture. The discrete domains (separated by shading) are; a signal peptide, an N-terminal region, an alanine-rich repeat (A-) domain, a variable (V-) domain, a proline-rich repeat (P-) domain, a C-terminal (C-) domain and a cell wall anchoring region (CWA).

Variable domain

C-terminal domain Proline-rich

domain CWA

Alanine-rich domain

N-terminal region Signal peptide

(16)

All AgI/II proteins share a well-conserved common primary sequence structure that consists of 1310-1653 amino acids divided into seven discrete regions (Figure 3). The N-terminal leader (signal) peptide region directs the protein to the surface by the general secretion (Sec) pathway. Following the signal peptide, are an N-terminal (N-) region, an alanine-rich repeat (A-) domain, a variable (V-) domain, a proline-rich repeat (P-) region, a C- terminal domain, and a cell wall anchoring segment.

The three-dimensional structure of the full-length AgI/II proteins and the function of the separate domains are unknown. The V-domain is composed of a central β-supersandwich and some helices (Paper I; Troffer-Charlier et al., 2002). The P-domain is rich in prolines and was shown to be important for the stability and secretion of AgI/II (Seifert et al., 2004). The A- and P- domains form tight interactions as they intertwine as a long supercoiled structure (Larson et al., 2010) and are thought to form a “stalk” presenting the V-domain. The C-terminal domain consists of three domains (C1, C2 and C3). The C2 and C3 domains are each composed of a (central) β-sandwich and form an elongated structure, and the C1 domain is predicted to have the same fold (Paper III).

Despite the high overall sequence similarity, differences in functional properties of members of the AgI/II family have been identified (Egland et al., 2001; Holmes et al., 1998; Brooks et al., 1997), suggesting that AgI/II proteins are involved in defining the ability of each species to interact with its surroundings.

Among the most extensively studied AgI/II proteins are those of S.

gordonii (SspA and SspB) and S. mutans (SpaP/Pac). Despite a high overall sequence similarity between S. gordonii SspA and SspB, differences in functional properties have been identified. For example, substrate-binding properties of S. gordonii SspA and SspB expressed on the surface of Lactococcus lactis differ in affinity for collagen and SAG (Holmes et al., 1998). In addition, SspA and SspB have independent functions in co- aggregation with different Actinomyces naeslundii co-aggregation groups (Egland et al., 2001; Jakubovics et al., 2005). Differences in functional properties are also seen when comparing SspA and SspB with other AgI/II family members. For example, SspA and SspB of S. gordonii are recognized by the P. gingivalis fimbrial protein Mfa1, whereas the related SpaP/Pac from S. mutans is not (Brooks et al., 1997). It has been shown that saliva- induced aggregation of S. mutans is inhibited by purified SpaP and that various recombinant fragments incorporating the A, V and P regions competitively inhibit bacterial binding to saliva-coated hydroxyapatite beads (Munro et al., 1993).

AgI/II polypeptides are also found on group A Streptococcus (GAS), a human pathogen responsible for infections such as pharyngitis and life-

(17)

threatening diseases such as streptococcal toxic-shock-like syndrome and necrotizing fasciitis (Zhang et al., 2006).

Interaction between Porphyromonas gingivalis and Streptococcus gordonii

P. gingivalis is a Gram-negative anaerobe which is one of the primary periodontopathogens involved in adult periodontitis (Lamont and Jenkinson, 1998). The bacterium is primarily found in subgingival plaque, although initial colonization by P. gingivalis likely occurs in the supragingival areas where established biofilm communities provide physiological support and reduced oxygen tension (Bradshaw et al., 1998;

Kuboniwa et al., 2009). P. gingivalis is a secondary colonizer which can interact directly with the early colonizer S. gordonii, and the interaction between the two bacteria are believed to be important in the development of bacterial communities that are associated with periodontal diseases (Rosan and Lamont, 2000). The interaction is multimodal; the P. gingivalis major fimbria, FimA, interacts with streptococcal surface-GAPDH and the minor fimbria, Mfa1, interacts with the streptococcal SspA/B (Maeda et al., 2004;

Park et al., 2005). Lamont et al. (1992) showed that Mfa1-SspB interaction is essential for P. gingivalis colonization on streptococcal substrates. In contrast, P. gingivalis does not bind S. mutans AgI/II protein SpaP. Specific amino acids in SspB that are not conserved in SpaP are thought to be responsible for this species specificity of interaction (Demuth et al., 2001).

Structural biology and general methods

Why structural biology?

Proteins are involved in all of the diverse processes within cells, such as enzymes catalyzing biochemical reactions, receptors in the immune system, and signaling and structural molecules (such as actin and myosin in muscle and cytoskeleton). This diversity is reflected in drug design, where proteins are the most common targets for developing medical drugs. To know the structure of a protein is of great interest, even if not intended to be used as a direct target for drug design. Knowing the structure can help us better understand the function of the protein. A great advantage is to have a structure of a complex between two interacting proteins, a protein and DNA/RNA (if it is a DNA or RNA binding protein), or a protein and its

“ligand”.

Proteins are too small to be studied with conventional microscopic techniques. Instead, NMR spectroscopy and X-ray crystallography make it possible to study the three-dimensional (3D) structure of proteins at atomic

(18)

resolution. In order to be able to obtain a 3D structure of a protein by the use of X-ray crystallography, one needs to obtain diffracting crystals. A prerequisite for obtaining a crystal is a relatively pure and homogenous protein, usually a recombinant protein obtained by cloning, overexpression and purification.

This section describes general methods used to obtain the 3D crystal structure of a protein by the use of X-ray crystallography. A schematic view of the procedure is given in Figure 4.

Cloning, expression and protein purification

The gene of interest is amplified by polymerase-chain reaction (PCR) by the use of specific primers. The amplified gene is cloned into an expression vector, which usually generates a construct with the gene fused to a tag (for purification purposes) and which confers antibiotic resistance to the vector containing the correct insert.

The vector can be expressed in several different expression systems, usually in E. coli expression strains. Several approaches can be used in order to successfully express and obtain a soluble protein. Parameters commonly varied during expression include expression temperature, growth medium and inducing agent.

There are various ways to purify a protein. The most commonly used chromatography techniques are affinity chromatography, ion exchange and size exclusion chromatography. If you are lucky, you will obtain a homogenous and pure protein, which can be used in crystallization experiments. If not, you may have to change your approach. For example you may need to re-clone, change expression conditions, change purification buffers or method.

Crystallization

Obtaining protein crystals is the major bottleneck in determining a 3D structure by X-ray crystallography. All proteins have their own chemical properties and behave differently, which means that there is no universal

“optimal” buffer condition that promotes crystal growth. Searching for specific conditions that force the protein molecules to from a crystal can be very time-consuming. In the search for the “right” condition, there are commercial screening kits available to facilitate the search. Diffraction quality crystals may arise from initial screening, but usually only “hits” are identified and thus, further optimization of crystallizing conditions will be needed.

(19)

Figure 4. Schematic view: from gene to structure.

There are many parameters that can affect the crystallization process, for example pH, temperature, ionic-strength, concentration and precipitant used. When attempting to optimize the conditions, these are common variables to vary and fine tune.

Protein crystals can be grown in different set-ups. The most commonly used method is vapor-diffusion by the hanging drop or sitting drop method.

In short, a drop of protein is mixed with well solution (precipitant) either on a cover-slip which is put upside down over a well (hanging drop), or on a ledge (sitting drop). The set-ups are left to equilibrate and since the

(20)

precipitant concentration in the drop is lower than in the well, water from the drop will diffuse down to the bottom of the well until the vapor pressures are at equilibrium. At suitable conditions nucleation occurs. Nucleation is the starting point where the protein molecules arrange into an ordered 3D pattern and a crystal starts to grow. Once crystals are obtained, they can be pre-soaked in suitable cryo-protectant, flash-cooled and stored in liquid nitrogen.

If no crystals are obtained, further screening may be needed. Adding a co-factor, a ligand or an additive may help. In addition, flexible parts of the protein or flexible surface residues (such as lysines) may interfere with crystal packing. Re-cloning or in some cases, reductive methylation (Walter et al., 2006) of surface-lysines on the protein may help.

Structure determination

The only way of knowing if a crystal is of good quality is to subject it to X- rays. If the crystal diffracts to reasonable resolution, it can be used to solve the crystal structure. In order to solve the structure, data need to be collected from the crystal. This is usually performed at a synchrotron, a huge facility where electrons are accelerated to extreme speed, creating an X-ray beam.

The crystal is mounted in a stream of nitrogen at 100K and subjected to the X-ray beam. If the crystal is well-packed with highly ordered protein molecules, the electrons of the atoms within the crystal will scatter the X- rays and a diffraction pattern is recorded on a detector. In order to get a complete picture of contents of the crystal, the crystal is rotated in increments of degrees (commonly 0.5 - 1°) and after each rotation another image of the diffraction pattern is collected. The more images collected, the more information is obtained, but as the crystals are eventually destroyed by the X-ray beam, a compromise needs to be made between the number of images collected (the completeness of the data) and the accuracy of the data obtained.

The X-rays that are scattered during data-collection hit the detector as a wave, with amplitude and phase. To find the position of the atoms within the protein, both the amplitude and the phase of the wave that resulted in each spot of the diffraction pattern need to be known. The amplitude can easily be obtained from measured intensities of each spot, but the phase information is lost. This is often referred to as “the phase problem” in crystallography.

The most common methods used to obtain the phases are multiple isomorphous replacement (MIR), multiple-wavelength anomalous dispersion (MAD), single-wavelength anomalous dispersion (SAD) or molecular replacement (MR). To use the MIR method, data sets from native and heavy-atom derivative crystals are collected. The differences in scattered intensities between the data sets are attributed to the heavy-atoms, and the locations of these can be determined. Once the locations of the heavy atoms

(21)

are known, these can be used to determine the phase for the entire protein.

In both MAD and SAD, data is collected at specific wavelengths to make use of the anomalous scattering of heavy atoms introduced into the protein. The anomalous scattering is used to determine the locations of the heavy atoms and subsequently the phase can be resolved. In MR, a known structure of a homologous protein sharing enough sequence identity (usually more than 30%) to the unknown protein structure is used as a model to obtain estimates of the phases. Once the estimated phases are obtained and the amplitudes are measured from the intensities, the electron distribution (i.e., the electron density) can be determined.

All models, whether obtained from MIR, MAD/SAD or MR, contains errors in the calculated electron density and need to be refined to get an as accurate as possible model of the protein from the resolution of the data.

This is done through repetitive iterations of model building and refinement resulting in improvement of the electron density map and the model.

At the end of the refinement, the quality of the model is validated.

Validated factors are for example, whether the atomic positions are reliable and how well the bonds and angles within the protein (stereochemistry) fit the ideal values.

(22)

Aims

The objective of this doctoral thesis was to obtain structural information about bacterial surface proteins involved in the formation of the oral biofilm by using X-ray crystallographic techniques. Furthermore, the purpose was to investigate biochemical aspects of surface proteins and their interaction with the surroundings.

Specific aims of this thesis were:

1. To determine the crystal structure of the Variable-domain of Streptococcus gordonii surface protein SspB and to compare it with the structure of the homolog SpaP V-domain of Streptococcus mutans. The aim was to examine similarities and differences between the two structures and relate them to possible differences in terms of interaction specificity of the two homologues.

2. To determine the crystal structure of the C-terminal domain of Streptococcus gordonii surface protein SspB and to identify the structure of known motifs within the protein as well as to examine putative surfaces for interaction.

3. To investigate biological and biochemical properties of domains of the SspB protein.

(23)

Results and Discussion

The variable domain of SspB from Streptococcus gordonii (Paper I)

The variable (V-) domain of Antigen I/II proteins is, as implied by the name, the most variable region within the proteins in the family. The two most distant V-domains within the oral streptococci are the S. gordonii SspB V- domain and the S. mutans SpaP V-domain. As the structure of the SpaP V- domain was already known (Troffer-Charlier et al., 2002), it was of interest to determine the structure of the SspB V-domain. A comparison of the structures of the two homologous proteins would give insights as to whether or not there are any structural differences that could suggest any distinction in substrate specificity.

The structure of the SspB V-domain

The crystal structure of the SspB V-domain (SspB-V) was solved by SAD- phasing and refined to 2.3 Å resolution (Paper I). The structure was solved from a crystal of methylated SspB-V protein, as methylation of surface lysines was needed to obtain diffracting crystals. The SspB-V structure is built up by a central β-sandwich flanked by two subdomains A (SDA) and B (SDB) on each side of the sandwich (Figure 5a). A putative binding pocket containing a metal ion is found between the subdomains and the top and bottom β-sheet. In addition, the N-terminal α-helix and a C-terminal coil run anti-parallel to each other and protrude from the core of the protein as a

„stalk‟ connecting the V-domain to the A- and P-domains.

The putative binding pocket

The inner part of the putative binding pocket contains a metal ion. It is coordinated by four protein atoms: Ser648 O and OG, Asn650 OD1 and Glu664 OE1, as well as three water molecules (Figure 5b). We believe that the metal is a calcium ion since the metal-protein distances refines to an average of 2.36±0.12, values consistent with calcium (2.36-2.39 Å) (Harding, 2006). Also, thermal induced melting analysis showed that the protein was most stable when supplemented with calcium. In addition, there is biological evidence that calcium is needed for the adherence of SspA and SspB to β1- integrin (Nobbs et al., 2007).

Three aromatic residues (tryptophan 709, 744 and 613) are found at the entrance of the pocket and in close connection with the metal ion, and these residues may stack with the bound ligand. Notably, mostly hydrophobic and

(24)

Figure 5. The structure of the SspB-V domain. (a) The domain is organized in a central immunoglobulin-like β-sandwich flanked by two subdomains, subdomain A (SDA) and subdomain B (SDB). The calcium ion is depicted as a yellow sphere.(b) The putative binding pocket. The calcium ion (brown sphere) is coordinated by four protein atoms and three waters (blue spheres). The coiled region comprising the major part of SDB is colored in green. The three tryptophans in the opening of the binding pocket and the amino acids of SDB are represented as stick models.

Figure 6. Comparison of SspB-V and SpaP-V. (a) Cα representation of SspB-V in black and SpaP-V in blue. The metal ion is represented as a gold sphere. (b) Electrostatic surface of SspB-V (left) and Spa-P (right), colored red and blue according to negative and positive electrostatic potential. The metal ion is depicted in yellow.

(a) (b)

(b) (a)

SDA

SDB

(25)

aromatic residues cluster in a long coiled region of SDB (Figure 5b); a region that may interact with or form a lid over a bound ligand.

SspB-V in comparison with SpaP-V

The overall structure of SspB-V is very similar to the SpaP-V structure except for some differences in length and conformation of connecting loops and helices mainly in the lower part (SDA & SDB) of the structure (Figure 6a).

Superimposing the structures results in a root mean square deviation of 2.9 Å for 281 aligned Cα atoms. Among the most similar regions are, besides the β-sandwich, the connection between the N-terminal helix and the C-terminal coil, along with the area around the putative binding pocket. The preserved position of the N-terminal helix and the C-terminal coil is not surprising as they are connected to the A- and P-domains respectively, which are intertwined into a left-handed supercoiled structure in a recombinant A3VP1-domain from SpaP (Larson et al., 2010). The two intertwined domains show a highly extended fibrillar form generated by a high affinity interaction that is believed to form a “stalk” presenting the V-domain to the surroundings. The tight interaction between the A- and P-domains is also supported by the conformational epitope recognized by a monoclonal antibody (van Dolleweerd et al., 2003).

The metal ion within the putative ligand binding pocket is found in approximately the same position and with similar coordination in the two proteins, indicating that it is a structurally conserved metal binding site.

Interestingly, the putative pocket of SspB-V is five times smaller than the one in SpaP-V, allowing for the speculation that the two proteins may bind different substrates (Figure 6b).

Previously, the putative binding pocket has been suggested to be designed for binding carbohydrates, due to the negative electrostatic potential and the aromatic sidechains close to the metal ion that may stack with a carbohydrate moiety (Troffer-Charlier et al., 2002). However, a glycan screen performed on the SspB-V protein in order to test binding of glycoconjugates to the SspB-V domain, detected no binding to any of the 400 tested common sugars (Paper I). This does not completely rule out the possibility of carbohydrates as a suitable substrate for binding, however, from these studies it does not seem likely that carbohydrates bind in the SspB-V pocket.

The SspB-V and the SpaP-V domains are very similar in their overall composition, and are suggested to be presented to the surrounding environment on the tip of the extended fibrillar structure formed by the A- and P-domains. The difference in volume of the putative pocket suggests that they may recognize different substrates, although their specificity remains unknown.

(26)

The C-terminal domain of the surface adhesin SspB from Streptococcus gordonii (Paper II and III)

Within the multi domain proteins belonging to the AgI/II family, the C- terminal domain is one of the most conserved domains based on primary sequence. For example, the C-terminal domains of SspB from commensal S.

gordonii (M5) and SpaP from caries associated S. mutans (UA159) share 68% primary sequence identity. This thesis presents the first structure of a C-terminal domain of an AgI/II protein and adds another piece to the puzzle of the structure of the full-length multi domain protein.

The structure of the proteolytically stable SspB C-terminal domain

The original recombinant C-terminal domain protein used in this work comprised residues 913-1413 of the full length SspB protein. However, despite repeated attempts no homogenous protein could be obtained.

Subjecting the protein to limited proteolysis resulted in a proteolytically stable protein comprising residues 1061-1413, named SspB-C1061-1413. We crystallized two truncated forms of the C-domain, SspB-C1061-1413 and SspB-C1083-1413, the latter of which produced more reproducible crystals.

The structures were solved at 1.7 Å with MR and at 1.5 Å by SAD-phasing, respectively. The two high-resolution structures were almost identical. The truncated C-domain consists of two structurally related domains, C2 and C3, each built up by a central β-sandwich containing one structural calcium ion and one isopeptide bond (Figure 7a). The C1 domain corresponds to the part of the protein that was cleaved off during limited proteolysis and is predicted to have the same fold as C2 and C3. The C2 and C3 domains have a similar topology, with the calcium ion and the isopeptide bond in identical positions. The most evident difference is a stretch of 21 amino acids including a helix located between β8 and β9 in C2 where the corresponding strands in C3 are connected by a short turn.

Intramolecular isopeptide bonds

Interestingly, we found that each domain contains a self-generated intramolecular isopeptide bond clearly visible in the electron density (Figure 7b). The isopeptide bond formed between Lys1082 and Asn1232 links the two sheets in C2, and the bond between Lys1259 and Asn1393 stabilizes the fold in C3. Each isopeptide bond is stabilized by an aspartate residue, in the immediate vicinity of Lys and Asn, that hydrogen bonds with the isopeptide C=O and NH groups. The reaction is facilitated by a local hydrophobic and aromatic environment, which increases the pKa of the Asp

(27)

and decreases the pKa of the Lys amino group. A plausible mechanism is that protonation of the Asp side chain causes polarization of the C=O bond of the Asn side chain. This allows the unprotonated Lys to perform a nucleophilic attack on the CG carbon of Asn, resulting in a covalent bond and the release of ammonium (Kang et al., 2007).

Figure 7. The structure of SspB-C1061-1413. (a)The C2 and C3 domains are indicated.

The calcium ions are depicted as black spheres and the isopeptide bonds are represented as cylinders in blue. The BAR region, with the KKVQDLLKK motif shown in yellow and the NITVK motif in red, is indicated. (b) The isopeptide bonds in the C2 and C3 domain. The isopeptide bonds in C2 (left) and C3 (right) are represented as ball-and-stick models in a simulated annealing omit FO-FC map. Hydrogen bonds with the aspartate residues are shown as broken lines. Hydrophobic residues surrounding the isopeptide are shown in green.

(b) (a)

2.82 Å 2.54 Å

3.09 Å 2.45 Å

Asp Lys Asn

Asp Lys Asn

C2

C3

(28)

At the time we solved the crystal structure of the C-domain, there was one crystal structure already solved in which similar self-generated intramoleular isopeptide bonds were identified, that being the crystal structure of the S.

pyogenes pili subunit Spy0128 (Kang et al., 2007). However, isopeptide bonds have now been reported in several other Gram-positive surface proteins (Kang et al., 2009; Izoré et al., 2010; Krishnan et al., 2007;

Deivanayagam et al., 2000; Zong et al., 2005; Budzik et al., 2009), confirming the common use of this means of stabilization of surface proteins from Gram-positive bacteria. By reporting decreased thermal stability and proteolytic stability of mutants that lacked either one or both isopeptides found in native Spy0128, Kang and Baker (2009) confirmed their hypothesis that intramolecular isopeptide bonds are important for the stability of surface proteins. To see whether this was the case with our protein, we created isopeptide mutants and tested them for proteolytic stability. In contrast to the Spy0128 protein, the C-domain isopeptide mutants were equally resistant to proteolysis as the native protein (Paper III). Therefore, we hypothesize that the isopeptide bonds in SspB have evolved for other purposes, such as to withstand mechanical stress from for example, the flow of saliva and tongue movement.

One hypothesis is that the elongated structure of the C-domain (which is the part of the AgI/II protein that is connected to the cell wall) is the first part of the “stalk” that together with the A- and P-domain structure, projects the V-domain away from the cell surface. In that sense, it may be of extreme importance that the C-domain structure is intrinsically stable to withstand the stress of the harsh environment of the oral cavity.

Similar structures

Several distant relatives to the SspB-C1061-1413 protein, all surface proteins, were identified in a homology search. Similar hits resulted from separate searches of the C2 and C3 domains. The closest relative was the middle domain of the C. diphtheriae pilin SpaA (Kang et al., 2009), which shares the same β-sheet topology as C2 and C3, as well as an isopeptide bond in the identical position. Other close relatives are the collagen-binding domain of CnaA from S. aureus (Zong et al., 2005) and the collagen-binding subdomain from Enterococcus faecalis (Ponnuraj and Narayana, 2007).

Since AgI/II proteins have been shown to interact with collagen, we specifically looked into the position and composition of the collagen-binding surfaces of these domains in relation to the SspB-C1061-1413 protein. The collagen-binding surfaces are rich in aromatic residues that can stack with the prolines in collagen. Though the collagen-binding surface superposes well with a symmetric groove in both C2 and C3, the grooves do not contain many aromatic residues, making interaction with collagen unlikely at these surfaces of SspB-C.

(29)

P. gingivalis interaction motifs

As P. gingivalis minor fimbrial subunit Mfa1 and its interaction with a specific region within SspB-C has been shown to be essential for P.

gingivalis colonization of a streptococcal biofilm (Demuth et al., 2001), we were interested in localizing it in the SspB-C structure. The specific region, named the SspB adherence region (BAR), comprises residues 1167-1193 in the full-length SspB protein, and is identified as the region that protrudes from the C2 domain (Figure 7a). The position of the BAR region is stabilized by the calcium ion in C2. In addition, two motifs (KKVQDLLKK and NITVK) have been identified within the BAR region that are important for the specific interaction between SspB and Mfa1 (Daep et al., 2006). The KKVQDLLKK motif forms an amphipathic α-helix which is followed by an extended region comprising the NITVK motif. Despite the high sequence similarity between the two streptococcal C-domains, P. gingivalis interacts with S. gordonii SspB-C but not with S. mutans SpaP-C. This interaction specificity is attributed to the NITVK motif as the single mutations N1182G and V1185P creates a protein more like S. mutans SpaP and results in the loss of adherence (Demuth et al., 2001). The specificity of P. gingivalis interaction with SspB is interesting, as it opens up the possibility for the use of the BAR handle structure in future medical drug design to interfere with detrimental interactions.

AgI/II domain interaction with gp340 (unpublished data)

The gp340 protein is an innate defense molecule produced at mucosal surfaces. It interacts with oral surface proteins such as AgI/II, various non- oral pathogens such as Helicobacter pylori and S. pyogenes, as well as viruses (Prakobphol et al., 2000; Edwards et al., 2008; Hartshorn et al., 2003; Wu et al., 2003). Fluid phase salivary gp340 promotes bacterial clearance but when surface immobilized it functions as a receptor for adherence of oral streptococci (Brady et al., 1992). AgI/II proteins interact with the peptide backbone and with gp340 oligosaccharides (Jakubovics et al., 2005; Bikker et al., 2002). In addition, SspA and SspB have been shown to be involved in aggregation of S. gordonii cells by fluid-phase gp340 (Loimaranta et al., 2005). To investigate which region of the SspB protein interacts with gp340, we purified several domains separately and performed overlay assays.

(30)

The gp340 overlay assay

gp340-binding of SspB domains was tested by a gp340-overlay assay. The AgI/II protein domains were run on an SDS-PAGE gel after which they were transferred to a membrane. gp340 was allowed to bind the protein domains.

Protein-bound gp340 was detected using Western blotting. In short, an anti- gp340 antibody was allowed to bind protein-bound gp340. Following incubation with a secondary antibody carrying a flourophore, the blot was developed and examined for protein-gp340 binding. The protein domains tested were; SspB-NA, SspB-V, SspB-C1081-1413, SspB-C913-1413 and SspB- VC. A P. gingivalis serine phosphatase (PG1170) was used as control. Two forms of gp340 were used: gp340 and non-glycosylated gp340.

From the overlay assays performed, we could conclude that all SspB protein domains tested were able to bind both glycosylated and non- glycosylated gp340. No binding could be detected for the control. A representative SDS-PAGE gel and the corresponding overlay membrane are showed in Figure 8.

Figure 8. The gp340 overlay. (a) The SDS-PAGE gel of the SspB domains tested.

Lane 1, moleular mass marker; Lane 2, SspB-NA; Lane 3, SspB-V; Lane 4, SspB- C1081-1413; Lane 5 SspB-C913-1413; Lane 6, SspB-VC; and Lane 7, PG1170 (control). (b) The overlay membrane showing SspB domains able to bind gp340.

Proteins in each lane are the same as in (a).

AgI/II interaction with gp340

The common binding of gp340 to all domains tested is not surprising, since gp340 is shown to interact with several sequences identified within the A-, V- and C-regions (Brady et al., 1992; Moisset et al., 1994; Kelly et al., 1995;

Senpuku et al., 1995). All of the SspB domains tested could bind non- glycosylated gp340, which suggests that the glycoconjugates that decorate the surface of gp340 are not critical ligands for AgI/II interaction with gp340. Interestingly, it has been observed that L. lactis expressing SspB on its surface is aggregated by the synthetic peptide SRCRP2 derived from the peptide backbone of gp340 (Jakubovics et al., 2005). However, further

7 1 2 3 4

1 2 3 4 5 6 5 6 7

(31)

overlay studies on neuraminidase treated gp340 need to be performed to investigate the contribution of sialic acid to the binding.

The evidence that all SspB regions tested interact with salivary glycoprotein further strengthens the concept that there are multiple contact points for AgI/II binding to gp340.

(32)

Conclusions

 We have solved the crystal structure of the variable domain of the AgI/II adhesin SspB from S. gordonii. We conclude that the overall structure of the SspB variable domain is very similar to the corresponding domain in SpaP from S. mutans. The putative binding pocket is significantly smaller in the SspB protein compared to SpaP, indicating a possible difference in substrate specificity, although the nature of the substrate remains to be elucidated.

 We have solved the first crystal structure of an AgI/II adhesin C- terminal domain: the truncated form of the C-terminal domain of the S. gordonii SspB adhesin. The C-terminal domain structure contains the C2 and the C3 domains, each composed of a central β- sandwich. Both domains contain an intramolecular isopeptide bond, believed to be a common stabilizing mechanism in Gram-positive surface proteins. In addition, the recognition region important for P.

gingivalis interaction with S. gordonii protrudes from the C2 domain as a handle for attachment.

 The interaction between gp340 and the SspB adhesin is multimodal, as shown by the fact that all discrete domains of the protein that were tested were able to bind gp340. Furthermore, the protein may interact with the peptide backbone of the gp340 protein.

 The crystal structures of AgI/II domains solved during the last years have improved the understanding of the function and composition of these multi domain proteins. However, there are still a lot of pieces missing, structurally as well as functionally, before we can figure out how these proteins interact with their surroundings.

(33)

Acknowledgements

Det är många som hjälpt och stöttat mig på den resa som resulterat i denna avhandling. Tack till er alla! Ett särskilt tack till:

Min handledare, Karina Persson, för ditt engagemang, din vägledning och för att du gett mig möjlighet att delta i diverse kurser och konferenser, både i och utanför Sverige. Ett särskilt tack för att du alltid funnits till hands när jag behövt din hjälp!

Min biträdande handledare, Elisabeth Sauer-Eriksson, för att jag fått tillgång till ert labb och er utrustning och särskilt för möjligheten att delta i era gruppmöten och diskussioner. Det har varit väldigt givande!

Mina arbetskamrater under tiden på forskarvåningen Odontologi:

Medlemmar av professor Sirkka Asikainens grupp som stått ut med att lyssna på prat om strukturer och kristallografi under åren med journal clubs och projektpresentationer. Det har varit nyttigt för mig att höra om er favoritbakterie A.a.

Ulla, Elisabeth, Pernilla, Inger, Maria, Liza, Rolf, Emma, Jan, Anders och alla ni andra för trevligt sällskap på fika- och lunchraster.

Arbetskamraterna i X-ray gruppen:

Liz, Uwe, Christin, Tobias, Anders O, Anders Ö, Kristoffer, ShengHua, Ulrika, Malin, Gitte, Åsa, Lixiao och Aaron för all input och idéer som hjälpt mig i min forskning. Tack också för att ni bidragit till en trevlig arbetsmiljö.

F.d. X-ray medlemmen Erik L för assistans och råd vid datainsamlingar på Max-lab. Guld värt!

Vill också passa på att tacka min familj och alla mina vänner för att ni förgyllt mitt liv både till vardags och till fest. Särskilt tack till:

Maria S, vi har varit vänner länge, ca 26 år om jag inte räknat alldeles galet :-). Gemensamma minnen har vi ju många, men oförglömlig är ju den där kalla vinterkvällen vid Millennium-skiftet… Tack för att du alltid funnits där i vått och torrt! Snart är det din tur… Lycka till med avhandlingen!

(34)

Maria L, det har varit många trevliga år med resor, partyn och sköna promenader. Du är verkligen någon jag kan dela mina innersta tankar med.

Tack!

Caroline, utan dig kanske det inte blivit nån avhandling i strukturbiologi.

Det var ju trots allt ditt tips som ledde mig till Karina :-). Du är dessutom en bra vän. Tack!

Cecilia, Anna och Sofie, tack för er vänskap och alla trevliga stunder! Det har hunnit slinka ner några räktoasts under våra tjejträffar… Mums!

Självklart ska även party-gänget nämnas: Maria S, Cecilia, Maria L, Anna, Caroline och Jenny. Det är tur att man haft ett sådant trofast gäng att hjälpa till med att göra livet ännu roligare!

Anna och Roger, tack för alla middagar och trevliga TP-kvällar! Anna, de är alldeles för kaxiga, vi måste se till att styra upp det här nu! Ett extra litet tack till Roger för alla timmar jag fått sitta på pass i skogen i väntan på att du äntligen ska hitta fram till vägen så man får avbryta nån gång.

Lars-Göran och Rosmari, tack för att ni välkomnat mig till er familj! Er hjälp är ovärderlig!

Ett särskilt stort tack till Mor och Far för allt stöd genom åren. Hur ska jag kunna beskriva vad ni betyder för mig? Tack för att ni finns! Tack också till bror och syster!

Slutligen vill jag tacka Erik, min underbara pojkvän. Tack för all kärlek och allt stöd. Jag älskar dig!!

(35)

References

Aas, J. A., Paster, B. J., Stokes, L. N., Olsen, I., and Dewhirst, F. E.(2005).

Defining the normal bacterial flora of the oral cavity. J. Clin. Microbiol. 43, 5721-5732.

Bensing, B. A., and Sullam, P. M. (2002). An accessory sec locus of Streptococcus gordonii is required for export of the surface protein GspB and for normal levels of binding to human platelets. Mol. Microbiol. 44, 1081-1094.

Bikker, F. J., Ligtenberg, A. J., Nazami, K., Veerman, E. C., van‟t Hoft, W., Bolscher, J. G., Poustka, A., Amerongen, V. N., and Mollenhauer, J. (2002).

Identifcation of the bacteria-binding peptide domain on salivary agglutinin (gp340/DMBT1), a member of the scavenger receptor cysteine-rich superfamily. J. Biol. Chem. 277, 32109-32115.

Bradshaw, D. J., Marsh, P. D., Watson, G. K., and Allison, C. (1998). Role of Fusobacterium nucleatum and coaggregation in anaerobe survival in planktonic and biofilm oral microbial communities during aeration. Infect.

Immun. 66, 4429-4432.

Brady, L. J., Piacentini, D.A., Crowley, P. J., Oyston, P. C., and Bleiweis, A. S.

(1992). Differentiation of salivary agglutinin-mediated adherence and aggregation of mutans streptococci by use of monoclonal antibodies against the major surface adhesin P1. Infect. Immun. 60, 1008-1017.

Brooks, W., Demuth, D. R., Gil, S. and Lamont, R. J. (1997). Identification of a Streptococcus gordonii SspB domain that mediates adhesion to Porphyromonas gingivalis. Infect. Immun. 65, 3753-3758.

Budzik, J. M., Marraffini, L. A., Souda, P., Whitelegge, J. P., Faull, K. F., and Schneewind, O. (2008). Amide bonds assemble pili on the surface of bacilli.

Proc. Natl. Acad. Sci. USA 105, 10215-10220.

Budzik, J. M., Poor, C. B., Faull, K. F., Whitelegge, J. P., He, C., and Schneewind, O. (2009). Intramolecular amide bonds stabilize pili on the surface of bacilli. Proc. Natl. Acad. Sci. USA 106, 19992-19997.

Cisar, J. O., Vatter, A. E., Clark, W. B., Curl, S. H., Hurst-Calderone, S., and Sandberg, A. L. (1988). Mutants of Antinomyces viscosus T14V lacking type 1, type 2, or both types of fimbriae. Infect. Immun. 21, 41-47.

(36)

Cunningham, M. W. (2008). Pathogenesis of group A streptococcal infections and their sequelae. Adv. Exp. Med. Biol. 609, 29-42.

Deivanayagam, C. C., Rich, R. L., Carson, M., Owens, R. T., Danthuluri, S., Bice, T., Höök, M., and Narayana, S. V. (2000). Novel fold and assembly of the repetitive B region of the Staphylococcus aureus collagen-binding surface protein. Structure 8, 67-78.

Demuth, D. R., Duan, Y., Brooks, W., Holmes, A. R., McNab, R., and Jenkinson, H. F. (1996). Tandem genes encode cell-surface polypeptides SspA and SspB which mediates adhesion of the oral bacterium Streptococcus gordonii to human and bacterial receptors. Mol. Microbiol. 20, 403-413.

Demuth, D. R., Irvine, D. C., Costerton, J. W., Cook, G. S., and Lamont, R. J.

(2001). Discrete protein determinants directs the species-specific adherence of Porphyromonas gingivalis to oral streptococci. Infect. Immun. 69, 5736- 3741.

Diaz, P. I., Chalmers, N. I., Rickard, A. H., Kong, C., Milburn, C. L., Palmer, R. J. Jr., and Kolenbrander P. E. (2006). Molecular characterization of subject-specific oral microflora during initial colonization of enamel. Appl.

Environ. Microbiol. 72, 2837-2842.

Douglas, C.W., Heath, J., Hampton, K. K., and Preston, F. E. (1993). Identity of viridians streptococci isolated from cases of infective endocarditis. J. Med.

Microbiol. 39, 179-182.

Edwards, A. M., Manetti, A. G. O., Falugi, F., Zingaretti, C., Capo, S., Buccato, S., Bensi, G., Telford, J. L., Margarit, I., and Grandi, G. (2008).

Scavenger receptor gp340 aggregates group A streptococci by binding pili.

Mol. Microbiol. 68, 1378-1394.

Egland, P. G., Du, L. D., and Kolenbrander, P.E. (2001). Identification of independent Streptococcus gordonii SspA and SspB functions in coaggregation with Actinomyces naeslundii. Infect. Immun. 69, 7512-7516.

Fedson, D. S., Anthony, J., and Scotts, G. (1999). The burden of pneumococcal disease among adults in developed and developing countries:

what is known and what is not known. Vaccine 17, S11-S18.

Filoche, S., Wong, L., and Sissons, C. H. (2010). Oral biofilm: Emerging concepts in microbial ecology. J. Dent. Res. 89, 8-18.

(37)

Gibbons, R.J. (1989). Bacterial adhesion to oral tissues: a model for infectious diseases. J. Dent. Res. 68, 750-760.

Harding, M. M. (2006). Small revisions to predicted distances around metal sites in proteins. Acta Crystallogr. D Biol. Crystallogr. 62, 678-682.

Hartshorn, K. L., White, M. R., Mogues, T., Ligtenberg, T., Crouch, E., and Holmskov, U. (2003). Lung and salivary scavenger receptor glycoprotein- 340 contribute to the host defense against influenza A viruses. Am. J.

Physiol. Lung Cell Mol. Physiol. 285, L1066–L1076.

Hasegawa, Y., Mans, J. J., Mao, S., Lopez, M. C., Baker, H. V., Handfield, M., and Lamont, R. J. (2007). Gingival epithelial cell transcriptional responses to commensal and opportunistic oral microbial species. Infect. Immun. 75, 2540-2547.

Holmes, A. R., McNab, R., and Jenkinson, H. F. (1996). Candida albicans binding to the oral bacterium Streptococcus gordonii involves multiple adhesin-receptor interactions. Infect. Immun. 64, 4680-4685.

Izoré, T., ontreras-Martel, C., El Mortaji, L., Manzano, C., Terrasse, R., Vernet, T., Di Guilmi, A. M., and Dessen, A. (2010). Structural basis of host cell recognition by the pilus adhesin from Streptococcus pneumoniae.

Structure 18, 106-115.

Jakubovics, N. S., Strömberg, N., van Dolleweerd,C. J., Kelly, C. G., and Jenkinson, H. F. (2005). Differential binding specificities of oral streptococcal antigen I/II family adhesins for human or bacterial ligands.

Mol. Microbiol. 55, 1591-1605.

Jenkinson, H. F., and Demuth, D. R. (1997). Structure, function and immunogenicity of streptococcal antigen I/II polypeptides. Mol Microbiol.

23, 183-190.

Jenkinson, H. F., and Lamont, R. J. (2005). Oral microbial communities in sickness and in health. Trends Microbiol. 13, 589-595.

Kang, H. J., Coulibaly, F., Clow, F., Proft, T., and Baker, E. N. (2007).

Stabilizing isopeptide bonds revealed in Gram-positive pilus structure.

Science 313, 1625-1628.

References

Related documents

Stöden omfattar statliga lån och kreditgarantier; anstånd med skatter och avgifter; tillfälligt sänkta arbetsgivaravgifter under pandemins första fas; ökat statligt ansvar

Byggstarten i maj 2020 av Lalandia och 440 nya fritidshus i Søndervig är således resultatet av 14 års ansträngningar från en lång rad lokala och nationella aktörer och ett

Omvendt er projektet ikke blevet forsinket af klager mv., som det potentielt kunne have været, fordi det danske plan- og reguleringssystem er indrettet til at afværge

I Team Finlands nätverksliknande struktur betonas strävan till samarbete mellan den nationella och lokala nivån och sektorexpertis för att locka investeringar till Finland.. För

Both Brazil and Sweden have made bilateral cooperation in areas of technology and innovation a top priority. It has been formalized in a series of agreements and made explicit

Domain architecture of the AgI/II proteins: a signal peptide, an N-terminal region, an alanine-rich repeat domain, a variable domain, a proline-rich repeat domain, a C-terminal

The crystal structures of several do- mains of the AgI/II proteins from the oral streptococci Streptococcus mutans and Streptococcus gordonii revealed that the AgI/II proteins fold

A recent crystal structure of the inactive (unphosphorylated) Btk kinase domain [68] has shown that although this tyrosine is located in the activation loop, the loop conformation is