Skip to main content
Wiley Open Access Collection logoLink to Wiley Open Access Collection
. 2011 Mar 25;319(2):97–105. doi: 10.1111/j.1574-6968.2011.02260.x

Location, synthesis and function of glycolipids and polyglycerolphosphate lipoteichoic acid in Gram-positive bacteria of the phylum Firmicutes

Nathalie T Reichmann 1, Angelika Gründling 1
Editor: Michael Mourez
PMCID: PMC3089915  PMID: 21388439

Abstract

Lipoteichoic acid (LTA) is a zwitterionic polymer found in the cell wall of many Gram-positive bacteria. A widespread and one of the best-studied forms of LTA consists of a polyglycerolphosphate (PGP) chain that is tethered to the membrane via a glycolipid anchor. In this review, we will summarize our current understanding of the enzymes involved in glycolipid and PGP backbone synthesis in a variety of different Gram-positive bacteria. The recent identification of key LTA synthesis proteins allowed the construction and analysis of mutant strains with defined defects in glycolipid or backbone synthesis. Using these strains, new information on the functions of LTA for bacterial growth, physiology and during developmental processes was gained and will be discussed. Furthermore, we will reintroduce the idea that LTA remains in close proximity to the bacterial membrane for its function during bacterial growth rather than as a surface-exposed structure.

Keywords: LTA, LtaS, cell wall, Gram-positive periplasm, protein localization

Introduction

The typical cell wall of low G+C Gram-positive bacteria, which include the well-studied model organism Bacillus subtilis and many important pathogens such as Staphylococcus aureus, Enterococcus faecalis, Streptococcus agalactiae, Streptococcus pyogenes and Listeria monocytogenes, is composed of proteins, peptidoglycan and teichoic acids (TAs). TAs can be further grouped into wall teichoic acid (WTA) and lipoteichoic acid (LTA), the latter of which is the focus of this review. The chemical structure of these cell wall components and the enzymes involved in peptidoglycan and WTA synthesis have been known for some time. However, our understanding of the synthesis and function of LTA has recently come into focus with the identification of key enzymes involved in its production. This has allowed the construction of bacterial strains that lack LTA, which in turn could be used to study the function of this polymer. These recent studies corroborate earlier suggestions of important roles for LTA in bacterial growth, physiology and during developmental processes, which will be discussed.

Both WTA and LTA are zwitterionic cell wall polymers; however, their synthesis proceeds through separate pathways. WTAs are usually composed of ribitol phosphate, glycerolphosphate (GroP) or more complex sugar-containing polymers that are polymerized on an undecaprenyl-phosphate precursor within the cytoplasm and transported across the membrane before being covalently linked to the peptidoglycan (Xia & Peschel, 2008). LTA has a simpler structure that typically consists of a polyglycerolphosphate (PGP) chain that is linked via a glycolipid anchor to the bacterial membrane (Fischer, 1990) (Fig. 1). The GroP backbone chain of both WTA and LTA is modified with alanine residues and in many bacteria with additional glycosyl groups (Fig. 1). The PGP chain is polymerized on the outside of the cell and the GroP subunits are derived from the head group of the membrane lipid phosphatidylglycerol (PG) and most likely added to the distal end of the growing chain (Taron et al., 1983; Koch et al., 1984).

Fig. 1.

Fig. 1

Chemical structure of PGP-type LTA. Schematic representation of Staphylococcus aureus and Bacillus subtilis LTA, which consists of a PGP chain that is linked to a diglucosyldiacylglycerol lipid anchor (boxed in orange). The hydroxyl groups at position C2 of the GroP subunits are substituted with d-alanine residues in both organisms and additional N-acetylglucosamine modifications are found in B. subtilis (shown in blue).

Here, we will summarize our current understanding of glycolipid and PGP-LTA backbone synthesis, and highlight enzymes that are involved in these synthesis processes in several well-studied Gram-positive organisms. We will discuss the function of glycolipids and LTA in terms of their importance for cell growth, cell shape and virulence, which emerged based on phenotypes observed in defined mutants. LTA backbone chain modifications will not be discussed further as the incorporation and function of d-alanine modification has only recently been reviewed (Neuhaus & Baddiley, 2003) and the process and function of glycosyl modifications is largely unknown. In addition, we will emphasize and reintroduce a theory for the localization of LTA in close proximity to the bacterial membrane (Fischer, 1994), which is important for bacterial growth and cell division, rather than LTA as a surface-exposed structure.

Glycolipid production: the LTA anchor

LTA production begins in the cytoplasm of the cell with the synthesis of the glycolipid anchor, which tethers the PGP chain to the cell membrane. In most cases, the LTA anchor is a glycolipid consisting of a disaccharide linked to diacylglycerol (DAG) (Fischer, 1988) that is produced by either a single processive glycosyltransferase or two separate enzymes. No uniform nomenclature is used to describe these glycosyltransferases that use nucleotide-activated sugars for the production of glycolipids, and in most cases, these enzymes are named based on observed phenotypes in deletion strains.

The LTA anchor in S. aureus and B. subtilis is composed of a glucosyl (β1-6) glucosyl (β1-3) DAG (Glc2-DAG) that is generated by the transfer of two glucose moieties from uridine diphosphate glucose (UDP-Glc) to DAG (Jorasch et al., 1998, 2000; Kiriukhin et al., 2001). As shown in Fig. 2a and b, UDP-Glc itself is produced by conversion of glucose-6-phosphate to glucose-1-phosphate by the α-phosphoglucomutase PgcA (also named Pgm), followed by activation by the UTP:α-glucose-1-phosphate uridyltransferase GtaB (sometimes named GalU) (Pooley et al., 1987; Lu & Kleckner, 1994; Lazarevic et al., 2005; Gründling & Schneewind, 2007a). A processive glycosyltransferase (YpfP in S. aureus and UgtP in B. subtilis) transfers two glucose moieties from UDP-Glc to DAG, leading to the formation of the glycolipid Glc2-DAG (Jorasch et al., 1998; Kiriukhin et al., 2001). In the absence of pgcA, gtaB or ypfP, no glycolipids are produced, but the LTA backbone is still synthesized. In these mutants, LTA is believed to be linked directly to DAG, a simpler lipid anchor also naturally used by some bacteria (Iwasaki et al., 1986; Lazarevic et al., 2005; Gründling & Schneewind, 2007a).

Fig. 2.

Fig. 2

Enzymes required for glycolipid and LTA backbone synthesis in different Gram-positive bacteria. Schematic representation of enzymes involved in glycolipid precursor, glycolipid and/or LTA backbone synthesis in (a) Staphylococcus aureus, (b) Bacillus subtilis, (c) Listeria monocytogenes, (d) Streptococcus pneumoniae, (e) Streptococcus agalactiae and (f) Enterococcus faecalis. Enzymes with experimental evidence for their activity are shown in color and their function is described in the text, while enzymes identified by blast searches based on homology are shown in gray. Predicted but still unidentified proteins are indicated with question marks (?). Glycolipid synthesis precursors are UDP nucleotide-activated sugars that are produced in the cytoplasm of the cell. These precursors are used by glycosyltransferases to synthesize the LTA glycolipid anchor by the transfer of glucose (red circles) or galactose (green circle) moieties onto the membrane lipid DAG. After transport of the glycolipid from the inside to the outside of the membrane, GroP subunits (yellow circles) derived from the membrane lipid phosphatidylglycerol are added to the anchor and the chain is extended by the repeated addition of GroP subunits at the distal end. This leads to the formation of the PGP-LTA backbone chain, which is schematically depicted by a string of yellow circles. In contrast, the non-PGP-type LTA in S. pneumoniae (d) is depicted by different symbols. Proteins are either named according to the information available in the literature or based on L. monocytogenes EGD-e, S. pneumoniae R6, S. agalactiae NEM316, and E. faecalis V583 gene numbers. Abbreviations used are as follows: Glc-6-P, glucose-6-phosphate; Glc-1-P, glucose-1-phosphate; Glc2-DAG, diglucosyldiacylglcerol; GalGlc-DAG, galactosylglucosyldiacylglycerol.

Listeria monocytogenes uses galactosyl (α1-2) glucosyl (α1-3) DAG (GalGlc-DAG) as the glycolipid anchor of LTA (Uchikawa et al., 1986) and therefore requires two distinct glycosyltransferases, LafA (LTA anchor formation protein A; Lmo2555) and LafB (LTA anchor formation protein B; Lmo2554), to produce monoglucosyldiacylglycerol (Glc-DAG) and GalGlc-DAG, respectively (Webb et al., 2009). Neither of the enzymes involved in UDP-Glc or UDP-galactose (UDP-Gal) production have been characterized in any detail, but homologues of S. aureus and B. subtilis PgcA and GtaB can be identified using blast searches (Fig. 2c). An additional phosphatidyl lipid can be linked to the glucose moiety of the glycolipid in L. monocytogenes, although the enzyme required for its formation is unknown (Uchikawa et al., 1986). Recently, it was found that at a higher temperature (37 °C vs. 30 °C), the fraction of LTA that contains this modified glycolipid anchor is increased (Dehus et al., 2011). It was further shown that LTA with this additional lipid modification less efficiently activates the human complement system through the lectin pathway, and it was speculated that this modification could help bacteria survive during an infection (Dehus et al., 2011).

LTA anchor formation has also been investigated in S. agalactiae and E. faecalis, where the α1-2-linked Glc2-DAG glycolipid is produced by a two-enzyme system similar to L. monocytogenes (Doran et al., 2005; Theilacker et al., 2009). The S. agalactiae glycosyltransferase IagA (invasion-associated gene A) and the E. faecalis glycosyltransferase BgsA (biofilm-associated glycolipid synthesis gene A) are responsible for the addition of the second glucose moiety and strains deleted for the respective genes accumulate Glc-DAG in the membrane (Doran et al., 2005; Theilacker et al., 2009). Both IagA and BgsA are homologous to the L. monocytogenes LafB and are encoded in an operon with gbs0683 and ef2890, respectively, that displays similarity to lafA, presumably encoding the glycosyltransferase responsible for the production of Glc-DAG (Fig. 2e and f).

Following synthesis in the cytoplasm, a fraction of the glycolipids are transported to the outer leaflet of the membrane, where they are used as the LTA anchor. Staphylococcus aureus LtaA, a member of a major facilitator superfamily clan that is encoded in an operon with ypfP, has been suggested to be involved in this process (Gründling & Schneewind, 2007a). This was based on the observation that in the absence of LtaA, the majority of LTA is linked directly to DAG despite the presence of wild-type levels of Glc2-DAG (Gründling & Schneewind, 2007a). In L. monocytogenes, an additional membrane protein, LafC, is encoded in an operon with lafA and lafB. However, LafC does not show any homology to the S. aureus LtaA protein and while slight changes in glycolipid and the LTA structure could be detected in its absence, the observed alterations were not necessarily consistent with a function in glycolipid transport (Webb et al., 2009). Furthermore, no apparent homologues to LafC or LtaA are present in B. subtilis. It is commonly accepted that a protein component is involved in the membrane transfer of glycolipids. However, it remains to be determined whether indeed, as it appears, a variety of nonhomologous membrane proteins can perform this reaction.

PGP production: the LTA backbone

Once the glycolipid reaches the outer leaflet of the membrane, in S. aureus, a single LTA synthase enzyme (LtaS) polymerizes the GroP to generate the PGP chain (Gründling & Schneewind, 2007b) (Fig. 2a). Staphylococcus aureus LtaS and its homologues in other Gram-positive bacteria are predicted membrane proteins with a large extracellular enzymatic domain, which is consistent with LTA synthesis occurring on the outer side of the membrane (Gründling & Schneewind, 2007b; Lu et al., 2009). Based on pulse-chase experiments, it has been suggested that the GroP subunits are derived from the membrane lipid PG (Koch et al., 1984), and recently, it has been shown that the purified enzymatic domain of LtaS can cleave the GroP head group of fluorescently labeled PG to form DAG, providing further evidence for this notion (Karatsa-Dodgson et al., 2010). In contrast, two LtaS-type enzymes produce the PGP backbone in L. monocytogenes. The LTA primase LtaP (Lmo0644) adds the first GroP subunit to the GalGlc-DAG lipid anchor and LtaS (Lmo0927) polymerizes the LTA backbone chain (Fig. 2c) (Webb et al., 2009). In the absence of the primase, LTA is still produced, but the deletion of ltaS abolishes LTA synthesis.

The situation is ever more complex in B. subtilis, which encodes four LtaS-type enzymes: LtaS (originally YflE), YqgS, YvgJ and YfnI (Gründling & Schneewind, 2007b; Schirner et al., 2009). All four proteins can hydrolyze fluorescently labeled PG in vitro and the expression of LtaS, YqgS or YfnI as the sole LTA synthesis enzyme is sufficient for PGP-LTA production (Wörmann et al., 2011) (Fig. 2b). Expression of YvgJ leads to the accumulation of GroP-Glc2-DAG, indicating that it functions as an LTA primase, similar to L. monocytogenes LtaP (Wörmann et al., 2011). LtaS appears to be the most important enzyme of the four proteins and the other homologues are thought to play overlapping roles in response to changing environmental conditions. YfnI expression is controlled by the stress σ-factor σM, while YqgS function is important during the sporulation process (Thackray & Moir, 2003; Jervis et al., 2007; Eiamphungporn & Helmann, 2008; Schirner et al., 2009).

Two different models for LTA synthesis have been proposed that differ in the enzymatic activity that is required for the linkage of the PGP chain to the glycolipid anchor and these have been reviewed recently (Rahman et al., 2009b; Sutcliffe, 2011). In one model, it was proposed that an ‘LTA transferase’ moves fully synthesized PGP polymers from a DAG lipid anchor onto the glycolipid anchor. The second model suggests that an LTA primase adds the first GroP subunit to the glycolipid to form a GroP-glycolipid intermediate and subsequently an LTA synthase extends the GroP chain to produce the LTA backbone. The discovery of enzymes with LTA primase activity in L. monocytogenes and B. subtilis strongly favors the second model (Webb et al., 2009; Sutcliffe, 2011; Wörmann et al., 2011). However, in addition to this two-enzyme system, a slightly altered version of the latter model, in which a single enzyme can directly start and extend the GroP chain on a glycolipid anchor, is also used in some bacteria such as S. aureus. The reason why some bacteria use one enzyme while others use multiple enzymes to synthesize the LTA backbone is currently not understood, but it is thought that multiple enzymes allow bacteria to fine-tune LTA synthesis in changing environments and during developmental processes such as sporulation in B. subtilis.

It is interesting to note that recent literature has highlighted the presence of a PGP-type LTA in various Actinobacteria (Rahman et al., 2009a,c). However, these bacteria appear to lack LtaS homologues and, therefore, it has been suggested that the polymer is synthesized by an alternative pathway (Rahman et al., 2009c).

Streptococcus pneumoniae LTA: an exception to the PGP-type LTA

In contrast to the Firmicutes discussed above, other members of this phylum including Lactococcus garvieae, Clostridium innocuum and S. pneumoniae are known to produce LTA with more complex structures (Fischer, 1990; Greenberg et al., 1996). Recent progress has been made regarding LTA synthesis in the extensively studied Gram-positive pathogen S. pneumoniae, where LTA is known to consist of N-acetylgalactosamines, ribitol phosphate, glucose and acetamido-4-amino-2,4,6-trideoxy-d-galactose (AATGal) repeating units (Behr et al., 1992; Seo et al., 2008). Consistent with the lack of a PGP-type LTA, no LtaS-type enzyme is present in S. pneumoniae. Despite this, it is still assumed that the LTA backbone is linked to a membrane glycolipid. Initially, it was proposed that this lipid anchor was Glc-AATGal-Glc-DAG (Behr et al., 1992), but this glycolipid has never been detected in the membrane. More recently, Seo et al. (2008) proposed that the anchor is Glc-DAG, which constitutes a minor fraction of the membrane lipids in S. pneumoniae, while GalGlc-DAG is the predominant glycolipid (Brundish et al., 1965). This lipid is likely synthesized in the same manner as in L. monocytogenes, requiring the production of UDP-Glc by the phosphoglucomutase Pgm (Spr1351) and uridyltransferase GalU (Spr1903) (Mollerach et al., 1998; Hardy et al., 2001) (Fig. 2d). For GalGlc-DAG production, the transfer of the second sugar moiety using UDP-Gal as the substrate is presumably performed by the glycosyltransferase CpoA, a LafB homologue (Edman et al., 2003), which is encoded in an operon with a second glycosyltransferase (Spr0982) that shows homology to LafA and presumably synthesizes Glc-DAG (Fig. 2d).

While the key enzymes in PGP-LTA synthesis have now been characterized, many more enzymes remain to be identified to decipher the LTA synthesis pathway in S. pneumoniae. Because the chain structures of LTA and WTA in S. pneumoniae are identical, it is possible that these polymers are, as suggested recently, synthesized by overlapping pathways and deciphering one may also shed light on the other (Seo et al., 2008; Rahman et al., 2009b).

Phenotypes of mutants defective in glycolipid synthesis

Bacteria with deletions in genes required for glycolipid synthesis are viable, but display diverse phenotypes. Staphylococcus aureusΔypfP cells, which lack the glycolipid anchor and produce LTA bound to DAG, are enlarged, misshaped cocci (Kiriukhin et al., 2001), while B. subtilisΔugtP mutants are shorter (Price et al., 1997; Weart et al., 2007; Salzberg & Helmann, 2008), exhibit a reduction in swarming motility and increased susceptibility to cationic antimicrobial peptides (Salzberg & Helmann, 2008). Depending on the strain background, an S. aureus ypfP mutant shows either a slight increase in PGP-LTA production and release into the supernatant or a drastic reduction in LTA production to 10% of the parental strain (Kiriukhin et al., 2001; Fedtke et al., 2007; Gründling & Schneewind, 2007a). The reason for these opposing phenotypes is currently not understood. The ypfP mutant strain with reduced LTA amounts shows a defect in biofilm formation (Fedtke et al., 2007) and has a reduced ability to penetrate the blood–brain barrier (Sheen et al., 2010); the latter phenotype was also observed for a group B Streptococcus iagA mutant (Doran et al., 2005). The E. faecalis bgsA mutant was named because of its defect in biofilm formation and is also defective in adhesion to Caco-2 cells (Theilacker et al., 2009). Subsequently, it was shown that Glc2-DAG, but not LTA, was able to inhibit the binding of E. faecalis to Caco-2 cells (Sava et al., 2009). This led the authors to propose a model in which LTA released into the supernatant binds to glucosaminoglycans on the eukaryotic membrane via the glycolipid anchor. The PGP chain, which is now tethered to the host cell, is then capable of rebinding to the bacterial cell wall, allowing for enhanced adhesion.

While the observed defects and phenotypes between different Gram-positive glycolipid mutants may vary, some general conclusions can be made. LTA backbone synthesis still proceeds in the absence of glycolipids, indicating that LtaS-type enzymes can initiate LTA synthesis on a different lipid. Mutants show some morphological defects, often a reduced ability to form biofilms, reduced ability to adhere to eukaryotic host cells and cross host cell barriers, which also leads to reduced virulence.

Morphological defects in bacterial strains lacking PGP-LTA

LtaS-type enzymes need to be inactivated to completely abolish LTA synthesis. So far, strains lacking LTA have been constructed in S. aureus, B. subtilis and L. monocytogenes (Gründling & Schneewind, 2007b; Schirner et al., 2009; Webb et al., 2009), and very recently, the construction of an LTA-deficient Lactobacillus acidophilus strain has been described (Mohamadzadeh et al., 2011). With the exception of L. acidophilus, these bacteria display growth and severe morphological defects. However, two LtaS homologues appear to be encoded in the L. acidophilus genome and only one of them was inactivated. It will be interesting to further investigate in future studies the contribution of both proteins to LTA synthesis and growth in this bacterium.

Upon deletion of the single ltaS gene in S. aureus, bacteria increase in size, display aberrant placement of division septa and can only grow in the presence of high sucrose or NaCl concentrations, conditions that presumably provide osmoprotection (Gründling & Schneewind, 2007b; Oku et al., 2009). An L. monocytogenes ltaS mutant lacking the PGP polymer shows a temperature-sensitive growth phenotype and forms long filaments that eventually lyse at 37 °C, the nonpermissive growth temperature (Webb et al., 2009). A similar filamentation phenotype was observed in a B. subtilis ltaS mutant, where it was further observed that assembly of the FtsZ ring, one of the key initial steps in the cell division process, does not proceed properly (Schirner et al., 2009). However, it should be noted that even in the absence of LtaS, PGP-LTA is still produced in this mutant. Only the combined absence of the three LTA synthases LtaS, YqgS and YfnI leads to a complete lack of the polymer (Wörmann et al., 2011). A quadruple mutant, which is unable to synthesize LTA, shows more severe morphological defects and bacteria form filaments that spiral along their long axes (Schirner et al., 2009). Lastly, it was found that an ltaS/yqgS double mutant cannot sporulate, highlighting the importance of LTA during this developmental process (Schirner et al., 2009).

Microscopic analysis of bacteria lacking LTA implicates a function for this polymer in the cell division process. This is especially obvious in rod-shape bacteria, where the lack of LTA leads to a filamentation phenotype without drastically affecting the diameter of the cell. Our current understanding of the molecular mechanism that links cell division and LTA synthesis is limited, but will be discussed below.

Localization of LTA synthesis proteins and their link with the cell division machinery

The protein localization pattern of three B. subtilis LTA synthesis enzymes, LtaS, YqgS and UgtP, was investigated and all three proteins were detected at septa, suggesting that glycolipid and LTA synthesis occurs at the site of cell division (Nishibori et al., 2005; Weart et al., 2007; Schirner et al., 2009). Interestingly, the localization pattern of UgtP correlated specifically with the location of the invaginating septum, reminiscent of the localization of FtsZ (Nishibori et al., 2005). Furthermore, Weart et al. (2007) found that purified UgtP protein could inhibit FtsZ polymerization in an in vitro assay, suggesting a direct physical interaction between these proteins. UgtP localization to the septum was dependent on the presence of UDP-Glc and was abolished in PgcA- and GtaB-deficient strains. It was suggested that B. subtilis uses the intracellular UDP-Glc concentration to sense the metabolic state of the cell and fine-tune FtsZ ring assembly, and hence the shorter cell size of B. subtilis ugtP, pgcA or gtaB mutants is due to the absence of UgtP-dependent inhibition of FtsZ polymerization (Weart et al., 2007). A direct interaction between FtsZ and UgtP in living cells remains to be determined and it is still an outstanding question whether other Gram-positive bacteria use glycosyltransferases to regulate cell division by influencing FtsZ function. While such a mechanism can be used to fine-tune cell division, the block/retardation in FtsZ assembly in the complete absence of LTA must be caused by a different mechanism. How the absence of LTA on the outside of the cell influences FtsZ assembly on the inside of the cell awaits further investigation.

Location of LTA in the Gram-positive periplasm and its release

More than 20 years ago, the existence of a periplasmic space as part of the Gram-positive cell wall was proposed (Takade et al., 1988; Umeda et al., 1992). Recently, this idea was reintroduced by Beveridge and colleagues based on cryo-transmission electron microscopy images of unstained, ultrarapid frozen and hydrated sections of B. subtilis and S. aureus cells (Matias & Beveridge, 2005, 2006). Two clearly distinct cell wall zones were observed outside the membrane: a low-density inner wall zone (IWZ), which is thought to be the Gram-positive periplasm, and a high-density outer wall zone (OWZ), which consists of the cross-linked peptidoglycan layer (Fig. 3). Based on these images, the thickness of the B. subtilis cell wall was estimated to be 55 nm (22.3±4.8 IWZ and 33.3±4.7 OWZ) and that of the S. aureus cell wall to be 36 nm (15.8±2.5 IWZ and 19±4.3 OWZ) (Matias & Beveridge, 2005, 2006). Labeling experiments with positively charged gold nanoparticles showed that these particles accumulate uniformly across the IWZ and OWZ, suggesting that the periplasmic space is filled with negatively charged components (Matias & Beveridge, 2008). After removing the OWZ layer with cell wall hydrolytic enzymes, a diffuse layer of approximately 30 nm extended from the outer leaflet of the membrane in B. subtilis. This layer was not significantly removed by protease treatment, providing evidence for components of a nonproteinaceous nature, and the fact that it could be labeled with an LTA-specific monoclonal antibody indicated that LTA might be a major component of the Gram-positive periplasm (Matias & Beveridge, 2008). Previous electron microscopy studies and immunoperoxidase labeling experiments using S. aureus protoplasts revealed LTA-specific labeling of the bacterial membrane (Aasjord & Grov, 1980). However, surface-exposed LTA could only be detected in whole cells of S. aureus strain Cowan I, but not in S. aureus strain Wood 46, providing evidence that LTA is not necessarily a cell surface-exposed molecule (Aasjord & Grov, 1980). Indeed, the 30 nm size of the diffuse layer in B. subtilis and the estimate that an extended LTA chain of 25 repeating units, as for instance present in S. aureus, is <20 nm in length (Labischinski et al., 1991; Fischer, 1994) does not necessarily agree with the commonly depicted schematic representation of the Gram-positive cell wall, in which the membrane-linked LTA chain extends through the cell wall to the surface. In addition, the extension of the PGP chain by the membrane protein LtaS, which uses PG as a substrate and presumably occurs at the distal end (Taron et al., 1983), necessitates that the polymer remains closely associated with the bacterial membrane, at least during its synthesis. The introduction of d-alanines, which can also occur on fully synthesized LTA, suggests that LTA chains remain in contact with the cell membrane even after their synthesis (Fig. 3). Therefore, it seems reasonable to assume that the physiological function of LTA in bacterial growth and cell division might predominantly take place in close proximity to the cell membrane and not at the bacterial cell surface. However, the transfer of d-alanines from LTA to WTA indicates an apparent flexibility, allowing the chain to also extend towards the peptidoglycan layer (Haas et al., 1984) (Fig. 3).

Fig. 3.

Fig. 3

Cell wall of Gram-positive bacteria. This representation of the Gram-positive cell wall is based on cryo-electron microscopy images, in which two clearly distinct cell wall zones were observed outside the membrane: a low-density IWZ, the Gram-positive periplasm, and a high-density OWZ, composed of the peptidoglycan-layer (PG-layer) (Matias & Beveridge, 2005, 2006, 2008). Besides the PG-layer, other components of the Gram-positive cell wall include the peptidoglycan-linked WTA and the membrane-linked LTA polymers, indicated in this figure. LTA, predicted to be a major component of the IWZ, is closely associated with the bacterial membrane during its production and may also remain closely associated with the lipid bilayer once fully synthesized. Also, fully extended membrane-linked LTA may not be able to completely penetrate the peptidoglycan layer and only reach the bacterial surface once released from the membrane.

Despite this, it is well documented that LTA is released from the bacterial cell and is frequently found in the culture supernatant and possibly rebinds to the bacterial cell surface (see review; Fischer, 1988). The mechanism by which LTA is released is not understood, but it is thought that it does not depend on actual cell lysis. In all likelihood, this fraction of LTA is not required for bacterial growth and cell division, but may explain how the polymer can play a role in processes such as adhesion and biofilm formation as suggested by others (Sava et al., 2009).

Concluding remarks

An original suggestion for the function of TAs 40 years ago was an involvement in the retention of ions in the cell wall and controlling ion access to the inner region of the wall (Archibald et al., 1961). Recently, it was shown that a B. subtilis ltaS mutant can grow at a low Mg2+ concentration, but becomes very sensitive to Mn2+ ions, corroborating this early suggestion (Schirner et al., 2009). Recently, the mechanism of ion binding to WTA, which is still linked to peptidoglycan, was reinvestigated using solid-state nuclear magnetic resonance (Kern et al., 2010). Based on these experiments, a model was proposed in which the metal ion binds cooperatively between a peptidoglycan peptide strand and the phosphate group of WTA (Kern et al., 2010). In analogy, one might expect a complex metal-binding mechanism between LTA and the negatively charged phospholipids in the bacterial membrane. While both WTA and LTA may influence the metal-binding ability of the bacterial cell wall, the construction and analysis of defined WTA- and LTA-deficient mutants showed that these polymers have clearly distinct functions in the cell (Schirner et al., 2009). Furthermore, as highlighted in this review, LTA-deficient cells have morphological defects that are consistent with a function of this polymer in the cell division process. For this function, LTA should not be seen as an actual cell surface structure, but more as a polymer in a unique location in the cell wall, in close proximity to the bacterial membrane.

Acknowledgments

Work in the Gründling laboratory is funded by the Wellcome Trust grant WT084483, the Medical Research Council grant G0701212 and the European Research Council grant 260371.

Statement

Re-use of this article is permitted in accordance with the Term and Conditions set out at http://wileyonlinelibrary.com/onlineopen#OnlineOpen-Terms

References

  1. Aasjord P, Grov A. Immunoperoxidase and electron microscopy studies of staphylococcal lipoteichoic acid. Acta Path Micro Im B. 1980;88:47–52. doi: 10.1111/j.1699-0463.1980.tb02602.x. [DOI] [PubMed] [Google Scholar]
  2. Archibald AR, Armstrong JJ, Baddiley J, Hay JB. Teichoic acids and the structure of bacterial walls. Nature. 1961;191:570–572. doi: 10.1038/191570a0. [DOI] [PubMed] [Google Scholar]
  3. Behr T, Fischer W, Peter-Katalinic J, Egge H. The structure of pneumococcal lipoteichoic acid. Improved preparation, chemical and mass spectrometric studies. Eur J Biochem. 1992;207:1063–1075. doi: 10.1111/j.1432-1033.1992.tb17143.x. [DOI] [PubMed] [Google Scholar]
  4. Brundish DE, Shaw N, Baddiley J. The glycolipids from the non-capsulated strain of Pneumococcus I-192R, A.T.C.C. 12213. Biochem J. 1965;97:158–165. doi: 10.1042/bj0970158. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Dehus O, Pfitzenmaier M, Stuebs G, Fischer N, Schwaeble W, Morath S, Hartung T, Geyer A, Hermann C. Growth temperature-dependent expression of structural variants of Listeria monocytogenes lipoteichoic acid. Immunobiology. 2011;216:24–31. doi: 10.1016/j.imbio.2010.03.008. [DOI] [PubMed] [Google Scholar]
  6. Doran KS, Engelson EJ, Khosravi A, Maisey HC, Fedtke I, Equils O, Michelsen KS, Arditi M, Peschel A, Nizet V. Blood–brain barrier invasion by group B Streptococcus depends upon proper cell-surface anchoring of lipoteichoic acid. J Clin Invest. 2005;115:2499–2507. doi: 10.1172/JCI23829. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Edman M, Berg S, Storm P, Wikström M, Vikström S, Öhman A, Wieslander A. Structural features of glycosyltransferases synthesizing major bilayer and nonbilayer-prone membrane lipids in Acholeplasma laidlawii and Streptococcus pneumoniae. J Biol Chem. 2003;278:8420–8428. doi: 10.1074/jbc.M211492200. [DOI] [PubMed] [Google Scholar]
  8. Eiamphungporn W, Helmann JD. The Bacillus subtilis sigma(M) regulon and its contribution to cell envelope stress responses. Mol Microbiol. 2008;67:830–848. doi: 10.1111/j.1365-2958.2007.06090.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Fedtke I, Mader D, Kohler T, Moll H, Nicholson G, Biswas R, Henseler K, Götz F, Zahringer U, Peschel A. A Staphylococcus aureus ypfP mutant with strongly reduced lipoteichoic acid (LTA) content: LTA governs bacterial surface properties and autolysin activity. Mol Microbiol. 2007;65:1078–1091. doi: 10.1111/j.1365-2958.2007.05854.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Fischer W. Physiology of lipoteichoic acids in bacteria. Adv Microb Physiol. 1988;29:233–302. doi: 10.1016/s0065-2911(08)60349-5. [DOI] [PubMed] [Google Scholar]
  11. Fischer W. Bacterial phosphoglycolipids and lipoteichoic acids. In: Kates M, editor. Handbook of Lipid Research. Vol. 6. New York: Plenum Press; 1990. pp. 123–234. [Google Scholar]
  12. Fischer W. Lipoteichoic acid and lipids in the membrane of Staphylococcus aureus. Med Microbiol Immun. 1994;183:61–76. doi: 10.1007/BF00277157. [DOI] [PubMed] [Google Scholar]
  13. Greenberg JW, Fischer W, Joiner KA. Influence of lipoteichoic acid structure on recognition by the macrophage scavenger receptor. Infect Immun. 1996;64:3318–3325. doi: 10.1128/iai.64.8.3318-3325.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Gründling A, Schneewind O. Genes required for glycolipid synthesis and lipoteichoic acid anchoring in Staphylococcus aureus. J Bacteriol. 2007a;189:2521–2530. doi: 10.1128/JB.01683-06. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Gründling A, Schneewind O. Synthesis of glycerol phosphate lipoteichoic acid in Staphylococcus aureus. P Natl Acad Sci USA. 2007b;104:8478–8483. doi: 10.1073/pnas.0701821104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Haas H, Koch U, Fischer W. Alanyl turnover from lipoteichoic acid to teichoic acid in Staphylococcus aureus. FEMS Microbiol Lett. 1984;21:27–31. [Google Scholar]
  17. Hardy GG, Magee AD, Ventura CL, Caimano MJ, Yother J. Essential role for cellular phosphoglucomutase in virulence of type 3 Streptococcus pneumoniae. Infect Immun. 2001;69:2309–2317. doi: 10.1128/IAI.69.4.2309-2317.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  18. Iwasaki H, Shimada A, Ito E. Comparative studies of lipoteichoic acids from several Bacillus strains. J Bacteriol. 1986;167:508–516. doi: 10.1128/jb.167.2.508-516.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  19. Jervis AJ, Thackray PD, Houston CW, Horsburgh MJ, Moir A. SigM-responsive genes of Bacillus subtilis and their promoters. J Bacteriol. 2007;189:4534–4538. doi: 10.1128/JB.00130-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Jorasch P, Wolter FP, Zahringer U, Heinz E. A UDP glucosyltransferase from Bacillus subtilis successively transfers up to four glucose residues to 1,2-diacylglycerol: expression of ypfP in Escherichia coli and structural analysis of its reaction products. Mol Microbiol. 1998;29:419–430. doi: 10.1046/j.1365-2958.1998.00930.x. [DOI] [PubMed] [Google Scholar]
  21. Jorasch P, Warnecke DC, Lindner B, Zahringer U, Heinz E. Novel processive and nonprocessive glycosyltransferases from Staphylococcus aureus and Arabidopsis thaliana synthesize glycoglycerolipids, glycophospholipids, glycosphingolipids and glycosylsterols. Eur J Biochem. 2000;267:3770–3783. doi: 10.1046/j.1432-1327.2000.01414.x. [DOI] [PubMed] [Google Scholar]
  22. Karatsa-Dodgson M, Wörmann ME, Gründling A. In vitro analysis of the Staphylococcus aureus lipoteichoic acid synthase enzyme using fluorescently labeled lipids. J Bacteriol. 2010;192:5341–5349. doi: 10.1128/JB.00453-10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Kern T, Giffard M, Hediger S, Amoroso A, Giustini C, Bui NK, Joris B, Bougault C, Vollmer W, Simorre JP. Dynamics characterization of fully hydrated bacterial cell walls by solid-state NMR: evidence for cooperative binding of metal ions. J Am Chem Soc. 2010;132:10911–10919. doi: 10.1021/ja104533w. [DOI] [PubMed] [Google Scholar]
  24. Kiriukhin MY, Debabov DV, Shinabarger DL, Neuhaus FC. Biosynthesis of the glycolipid anchor in lipoteichoic acid of Staphylococcus aureus RN4220: role of YpfP, the diglucosyldiacylglycerol synthase. J Bacteriol. 2001;183:3506–3514. doi: 10.1128/JB.183.11.3506-3514.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Koch HU, Haas R, Fischer W. The role of lipoteichoic acid biosynthesis in membrane lipid metabolism of growing Staphylococcus aureus. Eur J Biochem. 1984;138:357–363. doi: 10.1111/j.1432-1033.1984.tb07923.x. [DOI] [PubMed] [Google Scholar]
  26. Labischinski H, Naumann D, Fischer W. Small and medium-angle X-ray analysis of bacterial lipoteichoic acid phase structure. Eur J Biochem. 1991;202:1269–1274. doi: 10.1111/j.1432-1033.1991.tb16499.x. [DOI] [PubMed] [Google Scholar]
  27. Lazarevic V, Soldo B, Medico N, Pooley H, Bron S, Karamata D. Bacillus subtilisα-phosphoglucomutase is required for normal cell morphology and biofilm formation. Appl Environ Microb. 2005;71:39–45. doi: 10.1128/AEM.71.1.39-45.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Lu D, Wörmann ME, Zhang X, Schneewind O, Gründling A, Freemont PS. Structure-based mechanism of lipoteichoic acid synthesis by Staphylococcus aureus LtaS. P Natl Acad Sci USA. 2009;106:1584–1589. doi: 10.1073/pnas.0809020106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Lu M, Kleckner N. Molecular cloning and characterization of the pgm gene encoding phosphoglucomutase of Escherichia coli. J Bacteriol. 1994;176:5847–5851. doi: 10.1128/jb.176.18.5847-5851.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Matias VR, Beveridge TJ. Cryo-electron microscopy reveals native polymeric cell wall structure in Bacillus subtilis 168 and the existence of a periplasmic space. Mol Microbiol. 2005;56:240–251. doi: 10.1111/j.1365-2958.2005.04535.x. [DOI] [PubMed] [Google Scholar]
  31. Matias VR, Beveridge TJ. Native cell wall organization shown by cryo-electron microscopy confirms the existence of a periplasmic space in Staphylococcus aureus. J Bacteriol. 2006;188:1011–1021. doi: 10.1128/JB.188.3.1011-1021.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Matias VR, Beveridge TJ. Lipoteichoic acid is a major component of the Bacillus subtilis periplasm. J Bacteriol. 2008;190:7414–7418. doi: 10.1128/JB.00581-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Mohamadzadeh M, Pfeiler EA, Brown JB, et al. Regulation of induced colonic inflammation by Lactobacillus acidophilus deficient in lipoteichoic acid. P Natl Acad Sci USA. 2011;108:4623–4630. doi: 10.1073/pnas.1005066107. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Mollerach M, Lopez R, Garcia E. Characterization of the galU gene of Streptococcus pneumoniae encoding a uridine diphosphoglucose pyrophosphorylase: a gene essential for capsular polysaccharide biosynthesis. J Exp Med. 1998;188:2047–2056. doi: 10.1084/jem.188.11.2047. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Neuhaus FC, Baddiley J. A continuum of anionic charge: structures and functions of d-alanyl-teichoic acids in Gram-positive bacteria. Microbiol Mol Biol R. 2003;67:686–723. doi: 10.1128/MMBR.67.4.686-723.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Nishibori A, Kusaka J, Hara H, Umeda M, Matsumoto K. Phosphatidylethanolamine domains and localization of phospholipid synthases in Bacillus subtilis membranes. J Bacteriol. 2005;187:2163–2174. doi: 10.1128/JB.187.6.2163-2174.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Oku Y, Kurokawa K, Matsuo M, Yamada S, Lee BL, Sekimizu K. Pleiotropic roles of polyglycerolphosphate synthase of lipoteichoic acid in growth of Staphylococcus aureus cells. J Bacteriol. 2009;191:141–151. doi: 10.1128/JB.01221-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Pooley HM, Paschoud D, Karamata D. The gtaB marker in Bacillus subtilis 168 is associated with a deficiency in UDP-glucose pyrophosphorylase. J Gen Microbiol. 1987;133:3481–3493. doi: 10.1099/00221287-133-12-3481. [DOI] [PubMed] [Google Scholar]
  39. Price KD, Roels S, Losick R. A Bacillus subtilis gene encoding a protein similar to nucleotide sugar transferases influences cell shape and viability. J Bacteriol. 1997;179:4959–4961. doi: 10.1128/jb.179.15.4959-4961.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Rahman O, Cummings SP, Sutcliffe IC. Phenotypic variation in Streptomyces sp. DSM 40537, a lipoteichoic acid producing actinomycete. Lett Appl Microbiol. 2009a;48:226–229. doi: 10.1111/j.1472-765X.2008.02514.x. [DOI] [PubMed] [Google Scholar]
  41. Rahman O, Dover LG, Sutcliffe IC. Lipoteichoic acid biosynthesis: two steps forwards, one step sideways? Trends Microbiol. 2009b;17:219–225. doi: 10.1016/j.tim.2009.03.003. [DOI] [PubMed] [Google Scholar]
  42. Rahman O, Pfitzenmaier M, Pester O, Morath S, Cummings SP, Hartung T, Sutcliffe IC. Macroamphiphilic components of thermophilic actinomycetes: identification of lipoteichoic acid in Thermobifida fusca. J Bacteriol. 2009c;191:152–160. doi: 10.1128/JB.01105-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Salzberg LI, Helmann JD. Phenotypic and transcriptomic characterization of Bacillus subtilis mutants with grossly altered membrane composition. J Bacteriol. 2008;190:7797–7807. doi: 10.1128/JB.00720-08. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Sava IG, Zhang F, Toma I, Theilacker C, Li B, Baumert TF, Holst O, Linhardt RJ, Huebner J. Novel interactions of glycosaminoglycans and bacterial glycolipids mediate binding of enterococci to human cells. J Biol Chem. 2009;284:18194–18201. doi: 10.1074/jbc.M901460200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Schirner K, Marles-Wright J, Lewis RJ, Errington J. Distinct and essential morphogenic functions for wall- and lipo-teichoic acids in Bacillus subtilis. EMBO J. 2009;28:830–842. doi: 10.1038/emboj.2009.25. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Seo HS, Cartee RT, Pritchard DG, Nahm MH. A new model of pneumococcal lipoteichoic acid structure resolves biochemical, biosynthetic, and serologic inconsistencies of the current model. J Bacteriol. 2008;190:2379–2387. doi: 10.1128/JB.01795-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Sheen TR, Ebrahimi CM, Hiemstra IH, Barlow SB, Peschel A, Doran KS. Penetration of the blood–brain barrier by Staphylococcus aureus: contribution of membrane-anchored lipoteichoic acid. J Mol Med. 2010;88:633–639. doi: 10.1007/s00109-010-0630-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Sutcliffe IC. Priming and elongation: dissection of the lipoteichoic acid biosynthetic pathway in Gram-positive bacteria. Mol Microbiol. 2011;79:553–556. doi: 10.1111/j.1365-2958.2010.07480.x. [DOI] [PubMed] [Google Scholar]
  49. Takade A, Umeda A, Yokoyama S, Amako K. The substitution-fixation of Staphylococcus aureus. J Electron Microsc. 1988;37:215–217. [PubMed] [Google Scholar]
  50. Taron DJ, Childs WC, III, Neuhaus FC. Biosynthesis of d-alanyl-lipoteichoic acid: role of diglyceride kinase in the synthesis of phosphatidylglycerol for chain elongation. J Bacteriol. 1983;154:1110–1116. doi: 10.1128/jb.154.3.1110-1116.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Thackray PD, Moir A. SigM, an extracytoplasmic function sigma factor of Bacillus subtilis, is activated in response to cell wall antibiotics, ethanol, heat, acid, and superoxide stress. J Bacteriol. 2003;185:3491–3498. doi: 10.1128/JB.185.12.3491-3498.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Theilacker C, Sanchez-Carballo P, Toma I, Fabretti F, Sava I, Kropec A, Holst O, Huebner J. Glycolipids are involved in biofilm accumulation and prolonged bacteraemia in Enterococcus faecalis. Mol Microbiol. 2009;71:1055–1069. doi: 10.1111/j.1365-2958.2008.06587.x. [DOI] [PubMed] [Google Scholar]
  53. Uchikawa K, Sekikawa I, Azuma I. Structural studies on lipoteichoic acids from four Listeria strains. J Bacteriol. 1986;168:115–122. doi: 10.1128/jb.168.1.115-122.1986. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Umeda A, Yokoyama S, Arizono T, Amako K. Location of peptidoglycan and teichoic acid on the cell wall surface of Staphylococcus aureus as determined by immunoelectron microscopy. J Electron Microsc. 1992;41:46–52. [PubMed] [Google Scholar]
  55. Weart RB, Lee AH, Chien AC, Haeusser DP, Hill NS, Levin PA. A metabolic sensor governing cell size in bacteria. Cell. 2007;130:335–347. doi: 10.1016/j.cell.2007.05.043. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Webb AJ, Karatsa-Dodgson M, Gründling A. Two-enzyme systems for glycolipid and polyglycerolphosphate lipoteichoic acid synthesis in Listeria monocytogenes. Mol Microbiol. 2009;74:299–314. doi: 10.1111/j.1365-2958.2009.06829.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Wörmann ME, Corrigan RM, Simpson PJ, Matthews SJ, Gründling A. Enzymatic activities and functional interdependencies of Bacillus subtilis lipoteichoic acid synthesis enzymes. Mol Microbiol. 2011;74:566–583. doi: 10.1111/j.1365-2958.2010.07472.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Xia G, Peschel A. Toward the pathway of Staphylococcus aureus WTA biosynthesis. Chem Biol. 2008;15:95–96. doi: 10.1016/j.chembiol.2008.02.005. [DOI] [PubMed] [Google Scholar]

Articles from Fems Microbiology Letters are provided here courtesy of Wiley

RESOURCES