Abstract

Most bacteria have multiple peptidoglycan hydrolases capable of cleaving covalent bonds in peptidoglycan sacculi or its fragments. An overview of the different classes of peptidoglycan hydrolases and their cleavage sites is provided. The physiological functions of these enzymes include the regulation of cell wall growth, the turnover of peptidoglycan during growth, the separation of daughter cells during cell division and autolysis. Specialized hydrolases enlarge the pores in the peptidoglycan for the assembly of large trans-envelope complexes (pili, flagella, secretion systems), or they specifically cleave peptidoglycan during sporulation or spore germination. Moreover, peptidoglycan hydrolases are involved in lysis phenomena such as fratricide or developmental lysis occurring in bacterial populations. We will also review the current view on the regulation of autolysins and on the role of cytoplasm hydrolases in peptidoglycan recycling and induction of β-lactamase.

Introduction

Bacterial peptidoglycan hydrolases form a vast and highly diverse group of enzymes capable of cleaving bonds in polymeric peptidoglycan (sacculi) and/or its soluble fragments (Shockman & Höltje, 1994; Shockman et al., 1996). They participate in bacterial cell wall growth and its regulation and also in different lysis phenomena. It is often difficult to assign a distinct function to a peptidoglycan hydrolase for several reasons. First, many bacteria possess a high number of hydrolases and they appear to have redundant roles (Höltje & Tuomanen, 1991; Smith et al., 2000). Second, a hydrolase may have more than one function. For example, Escherichia coli has five N-acetylmuramyl-l-alanine amidases, six membrane-bound lytic transglycosylases and three peptidoglycan endopeptidases, all of which appear to contribute (at variable extent) to cleavage of the septum during cell division to allow separation of daughter cells (Heidrich et al., 2001, 2002). In addition, these hydrolases are responsible for the release of turnover products from peptidoglycan during cell growth (Goodell & Schwarz, 1985). Turnover products serve as signalling molecules for recognition of bacteria by other organisms and, in some bacteria, for the induction of β-lactamase (Jacobs et al., 1997). Moreover, these hydrolases are also responsible for autolysis induced under certain conditions (Tomasz, 1974, 1979a, b, 1980; Shockman, 1992; Shockman & Höltje, 1994; Shockman et al., 1996). There are examples of hydrolases for every glycosidic and amide bond in peptidoglycan, although not every specificity is found in every species. It is impossible to cover all knowledge gathered on peptidoglycan hydrolases over the last decades in a single review. Here, the focus will be on bacterial peptidoglycan hydrolases. An overview on the different specificities will be presented, and the various functions of hydrolases and their regulation will then be described.

Specificities of peptidoglycan hydrolases

The cleavage sites of the different hydrolases are shown in Figs 1 and 2 (Höltje, 1995). N-Acetylmuramyl-l-alanine amidases hydrolyse the amide bond between MurNAc and l-alanine separating the glycan strand from the peptide, whereas carboxy- and endopeptidases cleave the various ld- and dd-bonds in the stem peptides (Fig. 1). There are three types of glycan strand-cleaving enzymes (glycosidases), N-acetylglucosaminidases, lysozymes and lytic transglycosylases, the latter two cleaving the same glycosidic bond (Fig. 2). Lysozymes and lytic transglycosylases are also collectively known as N-acetyl-β-d-muramidases (muramidases). It should be noted that the activity of a hydrolase is often specific for a certain peptidoglycan type (Vollmer et al., 2008), for the presence or the absence of (a) secondary modification(s) (Vollmer et al., 2007), or for either high-molecular-weight (MW) peptidoglycan or small fragments.

Hydrolysis of amide and peptide bonds in peptidoglycan. The peptidoglycan structure shown is from E. coli or B. subtilis, with an m-A2pm residue at position 3 of the peptide (m-A2pm, meso-diaminopimelic acid, the free amino group is indicated). N-acetylmuramyl-l-alanine amidases (Ami) hydrolyse the amide bonds between the lactyl group of MurNAc and the l-alanine of the stem peptide. Some amidases (anhAmi) specifically cleave at 1,6-anhydroMurNAc residues, which are the hallmarks of peptidoglycan turnover products in many species. Endopeptidases (dd-EPase, ld-EPase, dl-EPase) cleave amide bonds in the peptides. The cleavage sites for ld- or dl-endopeptidases are indicated in a dimeric (cross-linked) peptide, but there are also cleavage sites for these enzymes in monomeric peptides. Carboxypeptidases (dd-CPase, ld-CPase, dl-CPase) hydrolyse peptide bonds to remove C-terminal d- or l-amino acids. See also: Höltje (1995, 1998), Smith (2000).
Figure 1

Hydrolysis of amide and peptide bonds in peptidoglycan. The peptidoglycan structure shown is from E. coli or B. subtilis, with an m-A2pm residue at position 3 of the peptide (m-A2pm, meso-diaminopimelic acid, the free amino group is indicated). N-acetylmuramyl-l-alanine amidases (Ami) hydrolyse the amide bonds between the lactyl group of MurNAc and the l-alanine of the stem peptide. Some amidases (anhAmi) specifically cleave at 1,6-anhydroMurNAc residues, which are the hallmarks of peptidoglycan turnover products in many species. Endopeptidases (dd-EPase, ld-EPase, dl-EPase) cleave amide bonds in the peptides. The cleavage sites for ld- or dl-endopeptidases are indicated in a dimeric (cross-linked) peptide, but there are also cleavage sites for these enzymes in monomeric peptides. Carboxypeptidases (dd-CPase, ld-CPase, dl-CPase) hydrolyse peptide bonds to remove C-terminal d- or l-amino acids. See also: Höltje (1995, 1998), Smith (2000).

Cleavage of glycosidic bonds in peptidoglycan by N-acetylglucosaminidases (1), lysozymes (2) or lytic transglycosylases (3). The latter cleave the glycosidic bond between MurNAc and GlcNAc with the concomitant formation of a 1,6-anhydro ring at MurNAc by an intramolecular transglycosylation reaction. R, peptide attached to the lactyl residue of MurNAc.
Figure 2

Cleavage of glycosidic bonds in peptidoglycan by N-acetylglucosaminidases (1), lysozymes (2) or lytic transglycosylases (3). The latter cleave the glycosidic bond between MurNAc and GlcNAc with the concomitant formation of a 1,6-anhydro ring at MurNAc by an intramolecular transglycosylation reaction. R, peptide attached to the lactyl residue of MurNAc.

N-Acetylmuramyl-l-alanine amidases

N-Acetylmuramyl-l-Ala amidases (MurNAc-LAAs) cleave the amide bond between MurNAc and the N-terminal l-alanine residue of the stem peptide (Fig. 1). These enzymes, also referred to in the literature as peptidoglycan amidases or amidases, are found in bacterial and bacteriophage or prophage genomes (Lopez et al., 1981; Young et al., 1992; Shockman & Höltje, 1994). In most cases, bacterial MurNAc-LAAs are members of the bacterial autolytic system and carry a signal peptide in their N-termini that allow their transport across the cytoplasmic membrane. However, the bacteriophage peptidoglycan amidases are endolysins. As opposed to autolysins, almost all endolysins have no signal peptides and their translocation through the cytoplasmic membrane is thought to proceed with the help of phage-encoded holin proteins (Young, 1992; Loessner, 2005).

Most often, several different MurNAc-LAAs are present in a bacterial proteome. For example, five MurNAc-LAAs are present in E. coli: AmiA, AmiB, AmiC (Heidrich et al., 2001) and AmiD (Uehara & Park, 2007; A. Pennartz, C. Généreux, & B. Joris, unpublished data) have a periplasmic localization, whereas AmpD is cytoplasmic (Jacobs et al., 1995). After cleavage of their signal peptide, AmiA, AmiB and AmiC are soluble in the periplasm whereas AmiD is synthesized as prelipoprotein and matured to become a lipoprotein anchored in the external membrane (Uehara & Park, 2007; A. Pennartz, C. Généreux, & B. Joris, unpublished data). Amino acid sequence comparisons show that the five E. coli amidases can be subdivided into two structurally distinct clusters. The first group comprises the homologous AmiA, AmiB and AmiC enzymes, and the second compreses AmiD and AmpD. Among the three enzymes forming the first group, AmiA possesses the shortest sequence for the amidase module. The sequence of this amidase module is conserved in AmiB and AmiC but it is fused at its N-terminal end to a module of unknown function different in the two amidases. AmiA, AmiB and AmiC play an important role in cleaving the septum to release daughter cells after cell division (see ‘Separation of daughter cells’). Functional fusion proteins of AmiA or AmiC and GFP show different subcellular localization patterns in living cells. AmiA-GFP is dispersed throughout the periplasm whereas AmiC-GFP appears throughout the periplasm in small cells and is concentrated almost exclusively at the septal ring in dividing cells. AmiC is the first entirely periplasmic component of the divisome to be localized, and its recruitment to the division ring is mediated by its N-terminal nonamidase domain (Bernhardt & de Boer, 2003). Interestingly, AmiA and AmiC are exported by the Twin-arginine transport (Tat) system instead of using the Sec machinery, which presumably exports AmiB (Bernhardt & de Boer, 2003). The Tat system is dedicated to the translocation of folded proteins containing cofactor(s) across the bacterial cytoplasmic membrane. AmiA and C are not predicted to contain a cofactor. However, amino sequence comparisons with the MurNAc-LAA PlyA from the Bacillus polymyxa bacteriophage PSA (Korndorfer et al., 2006) suggest that AmiA, AmiB and AmiC are Zn2+-containing metallo-enzymes. This could explain their secretion by the Tat pathway. The reason why AmiB is not transported by the Tat system still remains unexplained. The cytoplasmic AmpD amidase is among the enzymes that degrade the GlcNAc-1,6-anhydroMurNAc-tetra- and -tripeptide resulting from peptidoglycan turnover (Höltje, 1994; Jacobs et al., 1995) (see ‘Cytoplasmic hydrolases for peptidoglycan recycling and induction of β-lactamases’). The Citrobacter freundii AmpD amidase is a Zn2+-metallo-enzyme and its structure in solution has been determined by NMR (Liepinsh et al., 2003). In contrast to AmpD, AmiD lipoprotein cleaves in muropeptides with and without a 1,6-anhydro group at MurNAc, and it can utilize high-MW murein sacculi as a substrate (Uehara & Park, 2007; A. Pennartz, C. Généreux, & B. Joris, unpublished data). AmiD is also a Zn2+-metallo-enzyme and its recently elucidated structure closely resembles the structure of AmpD (Liepinsh, 2003; S. Petrella, R. Herman, C. Genereux, A. Pennartz, E. Sauvage, B. Joris, P. Charlier et al., unpublished data). Their folds are related to the fold of bacteriophage T7 lysozyme that is also a MurNAc-LAA (Cheng et al., 1994).

Unexpectedly, these structures are similar to those of the eukaryotic peptidoglycan-recognizing proteins (PGRPs). PGRPs are found in most animals, including insects, echinoderms, molluscs and vertebrates, but not in lower metazoa or plants and are a part of the innate immune system (Dziarski & Gupta, 2006). In insects, PGRPs activate antimicrobial pathways in the haemolymph and/or cells, or are lytic MurNAc-LAAs. In mammals, some PGRPs are MurNAc-LAAs and reduce the proinflammatory activity or kill bacteria. Others are noncatalytic, bind to peptidoglycan and are probably involved in signalling pathways.

The redundancy and the diversity of E. coli MurNAc-LAAs reflect well the plethora of peptidoglycan amidases described in the literature or predicted after bioinformatic analysis of known bacterial genomes. Even though it is impossible to describe all these enzymes in this review, it is, however, possible to extract some specific features that characterize this type of amidases. Generally, the amidase catalytic module is fused, by its N- or C-terminal end, to another functional module that is responsible for binding of the protein to the cell wall. Based on their amino-acid sequences, two different types of catalytic modules can be identified. The first one corresponds to the AmiA/B/C catalytic domain and the second one to that of AmiD/AmpD. Six structures are known. The bacteriophage PSA PlyA endolysin (Korndorfer et al., 2006) and the Paenibacillus polymyxa CwlV catalytic domain (T. Yamane, Y. Koyama, Y. Nojiri et al. unpublished data) are representative of the AmiA/B/C catalytic modules. On the other hand, the T7 lysozyme from T7 bacteriophage (Cheng et al., 1994), the C. freundii AmpD amidase (Liepinsh et al., 2003), the Bacillus anthracis prophage PlyL endolysin (Low et al., 2005) and the E. coli AmiD (S. Petrella, R. Herman, C. Genereux, A. Pennartz, E. Sauvage, B. Joris, P. Charlier et al., unpublished data) are members of the AmiD/AmpD catalytic domain. Greater diversity than in the catalytic domain lies in the noncatalytic domain fused to the amidase module. Usually, they are repeat domains named cell wall-binding domains (CBDs) that bind to peptidoglycan, lipoteichoic acid or pneumococcal teichoic acids containing choline (see ‘CBDs associated with peptidoglycan hydrolases’).

The presence of peptidoglycan amidase-sequence signatures in known translated genomes highlights that the catalytic and/or the repeated domains of the functional modules are present in several eukaryotic cells. These features are probably another example of horizontal gene transfer between prokaryotes and eukaryotes (Ponting et al., 1999).

Endopeptidases and carboxypeptidases

These enzymes cleave amide bonds between amino acids in peptidoglycan or its soluble fragments (Shockman & Höltje, 1994; Höltje, 1995). A large number of peptidases are known for the cleavage of the different amide bonds in the various types of peptidoglycan. According to the strict definition for a peptide bond as being a bond between the α-carboxylic group of one amino acid and the α-amino group of another, some bonds in the peptides of peptidoglycan would not be designated as a peptide bond. For example, the amide bond between d-Glu and meso-A2pm involves the γ-carboxylic group of d-Glu, and the dd-cross-bridge involves the ɛ-amino group of meso-A2pm (meso-2,6-diaminopimelic acid) or l-Lys. For the sake of clarity and consistency with the previous literature (e.g. Shockman & Höltje, 1994; Smith et al., 2000), in this review the terms ‘endopeptidase’ and ‘carboxypeptidase’ are used for enzymes hydrolysing the amide bond between two amino acids in peptidoglycan, and the term ‘amidase’ is reserved for enzymes cleaving between MurNAc and l-Ala. Depending on their specificity, peptidoglycan peptidases are classified as carboxypeptidases (removal of a C-terminal amino acid) or endopeptidases (cleavage within the peptide). dd-Peptidases cleave between two d-amino acids, whereas ld- or dl-peptidases cleave between an l- and a d-amino acid (Shockman & Höltje, 1995; Smith et al., 2000).

dd-Endopeptidases hydrolyse the d-Ala-meso-A2pm cross-bridges, which had been formed by dd-transpeptidation in nascent peptidoglycan by the class A and B high-MW penicillin-binding proteins (HMW-PBPs) (Sauvage et al., 2008). The side chain ammonium-carboxylate group at diaminopimelic acid is in the d-configuration and is structurally equivalent to the leaving d-alanine in dd-trans- or carboxypeptidation reactions. The dd-endopeptidase activity can thus be considered to be as the reverse facet of dd-trans- or carboxypeptidase activities. In fact, many dd-endopeptidases belong to the low-MW PBPs that share the typical sequence motifs of PBPs and are inhibited by β-lactams, which are structural analogues of the terminal d-Ala-d-Ala moiety in the pentapeptide donor for a peptidoglycan transpeptidation reaction (Goffin & Ghuysen, 1998; Sauvage et al., 2008).

Two types of PBPs have been identified as endopeptidases. The first one is a multi-modular enzyme described as type-4 PBP (Sauvage et al., 2008) and is not anchored to the cytoplasmic membrane via a transmembrane helix (Gittins et al., 1994; Harris et al., 2002; Stefanova et al., 2003). The second one is a mono-modular enzyme devoid of a C-terminal domain, but with an amphipathic helix that anchors the protein to the cytoplasmic membrane. These enzymes belong to the type-7 PBPs (Sauvage et al., 2008).

Three X-ray structures of type-4 PBPs are available: the Actinomadura R39 PBP, the E. coli PBP4 (Fig. 3) and the Bacillus subtilis PBP4a (Sauvage et al., 2005, 2007; Kishida et al., 2006). They share the same overall fold that is composed of three domains, one of which (domain I) possesses the typical structure and active site of the PBPs. The penicillin-binding/endopeptidase domain is associated with two other domains (II and III), which are not in N- or C-terminal positions but are inserted into the penicillin-binding/endopeptidase domain between the conserved motif 1 (with the SXXK signature, with S being the active serine) and motif 2 (SNN). In all type-4 PBPs, the positioning of the active site residues is nearly identical, with only minor differences affecting the very fine structure of the binding cavity. The bottom of the active site is characterized by the presence of a hydrophobic residue at the C-terminus of the β3 strand and several residues belonging to the additional domain II. These residues form a pocket that accommodates the terminal H3N+–CH–COO– group of the diaminopimelic acid, the antepenultimate amino acid of the peptidoglycan stem peptide, as observed in the complex of B. subtilis PBP4a with α-aminopimelyl-ɛ-d-alanyl-d-alanine (Sauvage et al., 2007). Unlike in some other PBPs, the binding of a β-lactam does not result in a significant movement of the active site residues. Different X-ray structures show that the antibiotics covalently linked to the active site serine adopt the common standard positioning as described for the acyl-enzyme PBP4a-α-aminopimelyl-ɛ-d-alanyl (Sauvage et al., 2005, 2007; Kishida et al., 2006). Type-4 PBPs are very loosely associated with the cytoplasmic membrane. A positively charged surface appears in domain II of the dd-peptidase of Actinomadura R39 and B. subtilis PBP4a but not in E. coli PBP4. It has been suggested that this positive surface could interact with the teichoic acid present in Bacillus and Actinomycetales (Sauvage et al., 2007). Type-4 PBPs are generally considered as being indirectly involved in cell morphology (Meberg et al., 2004), in daughter cell separation (Priyadarshini et al., 2006) and could be implicated in biofilm formation (Gallant et al., 2005) (see ‘Physiological functions of extracytoplasmic peptidoglycan hydrolases’).

Structures of penicillin-sensitive dd-peptidases. (a) dd-Endopeptidase PBP4 (E. coli), PDB code 2EX2, (b) dd-carboxypeptidase PBP5 (E. coli), PDB code 1NZO, and (c) dd-endopeptidase PBP7 (Mycobacterium tuberculosis), PDB code 2BCF. The catalytic serine residue of the penicillin binding/dd-peptidase-domain is shown as a green sphere.
Figure 3

Structures of penicillin-sensitive dd-peptidases. (a) dd-Endopeptidase PBP4 (E. coli), PDB code 2EX2, (b) dd-carboxypeptidase PBP5 (E. coli), PDB code 1NZO, and (c) dd-endopeptidase PBP7 (Mycobacterium tuberculosis), PDB code 2BCF. The catalytic serine residue of the penicillin binding/dd-peptidase-domain is shown as a green sphere.

PBP7 is the second penicillin-sensitive dd-endopeptidase in E. coli and is encoded by the pbpG gene (Henderson et al., 1995). PBP7 lacks dd-carboxypeptidase activity and is membrane-associated (Romeis & Höltje, 1994a). It hydrolyses the d-Ala-meso-A2pm cross-bridge in high-molecular mass sacculi and not in water-soluble low-molecular mass muropeptides. In vitro, PBP7 as well as its proteolytic degradation product PBP8 stabilize and stimulate the activity of another peptidoglycan hydrolase, the soluble lytic transglycosylase Slt70, by direct protein–protein interaction (Romeis & Höltje, 1994b). The crystal structure of Mycobacterium tuberculosis PBP7 has been determined (Krieger, unpublished data, Fig. 3). Its global fold is similar to the penicillin-binding domain of the E. colidd-carboxypeptidase PBP5 (Fig. 3). PBP7 lacks the N-terminal helix and exhibits on the top of the active site a β-hairpin protuberance, which, in other type-7 PBPs, was hypothesized to anchor the protein to the plasma membrane (Fonze et al., 1999).

There are also dd-endopeptidases unrelated to PBPs, which are not inhibited by β-lactams. The penicillin-insensitive endopeptidase MepA from E. coli hydrolyses both dd-(meso-A2pm–d-Ala) and ld-(meso-A2pm–meso-A2pm) peptide cross-links (Keck & Schwarz, 1979; Keck et al., 1990; Engel et al., 1992). MepA has crystallized as a dimer and its structure has been solved (Marcyjaniak et al., 2004) (Fig. 4). The enzyme is metal-dependent and contains a conserved His(113)-Asp(120)-His(211) Zn2+-binding triad essential for its activity. Two other histidine residues (at positions 206 and 209) were implicated in catalysis although the precise mechanism is not yet clear (Firczuk & Bochtler, 2007). MepA shares the active-site architecture with a large group of metallopeptidases including, for example, the peptidoglycan glycyl–glycyl endopeptidases LytM (Odintsov et al., 2004; Firczuk et al., 2005), lysostaphin (Heinrich et al., 1987) and ALE-1 (Lu et al., 2006), all of which cleave in the pentaglycine interpeptide bridge in the peptidoglycan of Staphylococcus aureus, the d-Ala-d-Ala carboxypeptidase from Streptomyces albus G (Charlier et al., 2003) and the d-Ala-d-Ala carboxypeptidase VanX involved in vancomycin resistance in enterococci (Bussiere et al., 1998). This group of enzymes has been termed ‘LAS-metallopeptidases’ because it includes lysostaphin-like enzymes, d-Ala-d-Ala metallopeptidases and the N-terminal peptidase domain of sonic hedgehog (Bochtler et al., 2004).

Structures of penicillin-insensitive ld-peptidases. (a) ld-carboxypeptidase LdcA (Pseudomonas aeruginosa), PDB code 1ZRS, with the catalytic triad Ser115–His285–Glu217. (b) ld-Endopeptidase MepA (E. coli), PDB code 1U10, with the catalytic zinc ion represented as a blue sphere and its three ligands (His113–Asp120–His211) with the two putative general base residues His206 and His209 represented as sticks.
Figure 4

Structures of penicillin-insensitive ld-peptidases. (a) ld-carboxypeptidase LdcA (Pseudomonas aeruginosa), PDB code 1ZRS, with the catalytic triad Ser115–His285–Glu217. (b) ld-Endopeptidase MepA (E. coli), PDB code 1U10, with the catalytic zinc ion represented as a blue sphere and its three ligands (His113–Asp120–His211) with the two putative general base residues His206 and His209 represented as sticks.

Several ld-endopeptidases have been identified in B. subtilis, two of which, CwlK and LytH, hydrolyse the bond between l-Ala (position 1) and d-Glu (position 2) in the stem peptide (Fukushima et al., 2007). The amino acid sequence of CwlK is similar to that of the AepA (Ply500) endopeptidase of the Listeria monocytogenes phage A500 (Loessner et al., 1995) and the enterococcal d-Ala-d-Ala carboxypeptidase VanY, which is required for the expression of vancomycin resistance (Arthur et al., 1998). CwlK is unrelated to a second B. subtilisld-endopeptidase, LytH, which presumably cleaves the same peptide bond (between l-Ala and d-Glu), and that is responsible for the high extent of MurNAc residues carrying a single l-Ala in the spore peptidoglycan of this species (Horsburgh et al., 2003a).

Bacillus subtilis has various dl-endopeptidases cleaving the bond between d-Glu (position 2) and meso-A2pm (position 3) of the peptide. These enzymes are members of two different families of dl-endopeptidases. CwlS (Fukushima et al., 2006), CwlO (Yamaguchi et al., 2004), LytE (Ohnishi et al., 1999) and LytF (Margot et al., 1999; Ohnishi et al., 1999) are related to the family II dl-endopeptidase from Bacillus sphaericus (Hourdou et al., 1992) and the p60 autolysin from L. monocytogenes (Wuenscher et al., 1993). The homologous LytE, CwlS and LytF proteins differ mainly in the number of their LysM peptidoglycan-binding repeats (three in LytE, four in CwlS and five in LytF) (see ‘CBDs associated with peptidoglycan hydrolases’) (Fukushima et al., 2006). Bacillus subtilis has also a hypothetical member (YqgT) of the family I dl-endopeptidases (Smith et al., 2000) that are Zn2+-dependent enzymes with sequence similarity to endopeptidase I from B. sphaericus (Hourdou et al., 1993). This enzyme has both dl-endopeptidase and dl-carboxypeptidase activities and cleaves the peptide bond between d-Glu and m-A2pm in tripeptides (l-Ala-d-Glu(γ)-meso-A2pm) or tetrapeptides (l-Ala-d-Glu(γ)-meso-A2pm-d-Ala) with or without attached GlcNAc-MurNAc moiety (at l-Ala). Another member of this family is the cytoplasmic dl-endopeptidase MpaA from E. coli (Uehara & Park, 2003).

ld-carboxypeptidases cleave between m-A2pm (position 3) and d-Ala (position 4) in a tetrapeptide to remove the terminal d-Ala residue. Although such an activity has been detected in many Gram-negative and Gram-positive species, there are only a few enzymes known to have this specificity. One of them might be encoded by the dacB gene of Lactococcus lactis as, unlike the wild type, a dacB mutant lacks tripeptides in its peptidoglycan and has an increased percentage of tetrapeptides (Courtin et al., 2006). Like several of the ld-specific endopeptidases, DacB has sequence similarity to the VanY d-Ala-d-Ala carboxypeptidase but the enzyme has not yet been characterized. LdcA from E. coli is a cytoplasmic ld-carboxypeptidase involved in peptidoglycan recycling and is essential at the onset of the stationary phase of growth (Templin et al., 1999). The substrates of LdcA are free tetrapeptides (l-Ala-d-Glu(γ)-meso-A2pm-d-Ala), or tetrapeptides linked to MurNAc, GlcNAc-MurNAc, GlcNAc-1,6-anhydroMurNAc or UDP-MurNAc, but not peptidoglycan sacculi or cross-linked fragments. The crystal structure of the homologous ld-carboxypeptidase from Pseudomonas aeruginosa has been solved (Korza & Bochtler, 2005) (Fig. 4). The enzyme has a two-domain architecture with a conserved catalytic Ser–His–Glu triad. This triad is very rare among the known serine peptidases, most of which contain a Ser–His–Asp triad and are evolutionarily unrelated to LdcA (Korza & Bochtler, 2005).

Usually, dd-carboxypeptidases belonging to type-5 PBPs are the most abundant low-MW PBPs. Three crystal structures of type-5 PBPs are available, PBP5 from E. coli (Davies et al., 2001; Nicholas et al., 2003) (Fig. 3), PBP3 from Streptococcus pneumoniae (Morlot et al., 2005) and PBP4 from S. aureus (http://www.pdb.org/pdb/explore.do?structureId=1TVF). They are bimodular enzymes with a penicillin-binding domain of classical topology and a β-strand-rich C-terminal domain that was shown in E. coli to be essential for the correct functioning of PBP5 (Nelson & Young, 2001). This terminal domain is characterized by an amphipathic helix at its end, which is responsible for anchoring the enzyme to the cytoplasmic membrane. Truncated PBP5 lacking the C-terminal domain is soluble (like PBP7) and its overproduction leads to cell lysis (Nelson & Young, 2001). Structures of dd-carboxypeptidases are reviewed in more detail in another article in this issue (Sauvage et al., 2008).

N-Acetyl-β-d-muramidases

N-acetyl-β-d-muramidases (termed N-acetylmuramidases in this review) are ubiquitous enzymes cleaving the β1,4-glycosidic bond between MurNAc and GlcNAc residues in peptidoglycan. This bond can be cleaved in two different ways (Fig. 2): lysozymes hydrolyse the glycosidic bond, resulting in a product with a terminal reducing MurNAc residue. In contrast, lytic transglycosylases cleave the glycosidic bond with a concomitant intramolecular transglycosylation reaction, resulting in the formation of the 1,6-anhydro ring at the MurNAc residue of the product.

Lysozymes

Lysozymes are produced in phages, bacteria, fungi, vertebrates and invertebrates, and they are among the most and best-studied enzymes in biology and biochemistry (Jolles et al., 1996). There are nine classes of proteins with a proven or a predicted ‘lysozyme-like’ fold (Pei & Grishin, 2005). Although all of them cleave β1,4-glycosidic bonds, some of these enzymes hydrolyse chitin [β1,4(GlcNAc)-polymer] or chitosan (partly deacetylated chitin), or are lytic transglycosylases. Four classes contain members with proven hydrolytic activity against peptidoglycan. The prototypes of these four classes are hen egg-white lysozyme (HEWL), goose egg-white lysozyme (GEWL), bacteriophage T4 lysozyme (T4L) and Chalaropsis lysozyme. Possibly HEWL, GEWL and phage T4L are the result of divergent evolution from a common ancestor. Despite the fact that they do not share any statistically significant sequence identity, their three-dimensional structures show some intriguing, albeit distant, similarities (Strynadka & James, 1996). The active site is located in a crevice between two domains, which are connected by a long α-helix. A glutamic acid residue, proposed to be essential for general acid catalysis, is located at the C-terminal end of an α-helix in the N-terminal half of each of the proteins. In contrast, the structure of cellosyl, which belongs to the Chalaropsis type of lysozymes, comprises a single domain, with the shape of a flattened ellipsoid. Its β/α-barrel fold is unrelated to the tertiary structures of HEWL, GEWL and T4L (Rau et al., 2001). There is a prominent, long groove, very likely the substrate-binding site, located on the C-terminal face of the cellosyl β-barrel. This groove culminates in a deep hole of a highly negative electrostatic potential forming the catalytic site of the enzyme.

As compared with other peptidoglycan hydrolases, there are relatively few well-characterized bacterial lysozymes with an experimentally proven specificity (Höltje, 1996b). Two autolytic lysozymes have been identified in Enterococcus faecium ATCC9790 (now Enterococcus hirae ATCC9790) (Kawamura & Shockman, 1983; Barrett et al., 1984). One of them, termed SF muramidase, has been shown to be an exoenzyme processively degrading glycan strands from their GlcNAc end. In contrast, pesticin is an endo-specific N-acetylmuramidase like HEWL (Vollmer et al., 1997). Pesticin is encoded by the pst gene, which is tandemly arranged with its immunity gene pim on a small Yersinia pestis virulence plasmid pPCP1. Such a gene arrangement is typical for bacteriocin (colicin)-encoding plasmids. Another bacterial lysozyme, cellosyl, is produced by Streptomyces coelicolor and belongs to the class of Chalaropsis lysozymes. Unlike HEWL and T4L, cellosyl exhibits a β1,4-N,6-O-diacetylmuramidase activity and is able to degrade O-acetylated peptidoglycan present in S. aureus and other pathogens. Cellosyl is also active on peptidoglycan-carrying (N-deacetylated) glucosamine and/or muramic acid residues in their glycan strands (Vollmer, 2007). Next to cellosyl, the autolysin LytC from S. pneumoniae is another bacterial N-acetylmuramidase belonging to the class of Chalaropsis lysozymes (Garcia et al., 1999a). Like other pneumococcal peptidoglycan hydrolases, LytC contains a choline-binding module mediating its attachment to cell wall teichoic acid, which is essential for its activity. The structures of the catalytic and choline-binding domains of LytC have been modelled (Monterroso et al., 2005).

Lytic transglycosylases

Lytic transglycosylases cleave the β1,4-glycosidic bond between MurNAc and GlcNAc with the concomitant formation of a 1,6-anhydro ring at the MurNAc residue (Höltje, 1975). The proposed intramolecular transglycosylation reaction involves a single catalytic acidic residue, often glutamate, which donates its proton to the glycosidic oxygen between MurNAc and GlcNAc (Thunnissen et al., 1994). The resulting oxocarbonium presumably is stabilized by the N-acetamido group at position 2 of MurNAc under formation of an intermediate oxazoline ring (Scheurwater et al., 2007). Deprotonation of the hydroxyl group at C-6 of MurNAc by the catalytic glutamate allows nucleophilic attack at C1 to form the 1,6-anhydro ring and resolve the oxazoline intermediate. This mechanism is in accordance with the observed inhibition of the membrane-bound lytic transglycosylase MltB from P. aeruginosa by N-acetlyglucosamine thiazoline, which is a structural analogue of the oxazoline intermediate (Reid et al., 2004).

Many catalytic domains of lytic transglycosylases, for example that of the soluble lytic transglycosylase Slt70 from E. coli, possess a similar fold as GEWL, which also has an acid glutamate residue in the active site (Thunnissen et al., 1995a). However, there are prominent structural differences between GEWL and lytic transglycosylases. The environment of the catalytic glutamate residue is much more hydrophobic in lytic transglycosylases, which also lack the second catalytic residue (an aspartate) present in lysozymes, explaining the different reaction mechanisms (hydrolysis vs. intramolecular ring formation) of both types of enzymes. Also, the substrate specificities of lysozymes and lytic transglycosylases differ. Lysozymes are endo-N-acetylmuramidases whereas most of the characterized lytic transglycosylases exhibit exo-lytic activity releasing 1,6-anhydroMurNAc containing disaccharide peptide units from one of the glycan strand ends. This is consistent with crystal structures showing differences in the arrangements of the substrate-binding pockets. Whereas lysozymes accommodate a hexasaccharide in the substrate-binding site, different lytic transglycosylases have binding sites for three to six N-acetylaminosugar residues. In the case of Slt70, exo-enzymatic activity is also determined by the overall doughnut shape of the enzyme (Thunnissen et al., 1995a). Furthermore, N-acetylmuramidases differ in their substrate specificities with respect to the requirement of peptide substitution for activity. For example, Slt70 and bacteriophage T4 lysozymes only cleave peptidoglycan containing the peptides linked to the glycan strands, and they are inactive against unsubstituted glycan strands (lacking the peptides). On the other hand, the membrane-bound lytic transglycosylase MltA from E. coli and HEWL can cleave equally well peptide-substituted and unsubstituted glycan strands (Romeis et al., 1993; Ursinus & Höltje, 1994). These differences in substrate requirement indicate the presence of binding sites for peptides in some but not all N-acetylmuramidases.

Lytic transglycosylases have been grouped into four families according to their amino acid sequences and the presence of conserved sequence motifs (Blackburn & Clarke, 2001). Family 1 was divided into five subfamilies (termed 1A–1E). Family 4 contains primarily lytic transglycosylases from bacteriophages. With the exception of a family 4 enzyme, E. coli has one member of each (sub-)family of lytic transglycosylases: Slt70 (family 1A), MltC (family 1B), EmtA (family 1C), MltD (family 1D), YfhD (family 1E), MltA (family 2) and MltB (family 3). All these enzymes either have a periplasmic localization (Slt70) or are lipoproteins attached to the inner leaflet of the outer membrane (all Mlts and EmtA). Proteolytic degradation of membrane-bound MltB yields the soluble Slt35. One of the lytic transglycosylases of E. coli, EmtA, shows endo-specific activity (Kraft et al., 1998). The crystal structures of the following lytic transglycosylases have been solved: E. coli Slt70 alone (Thunnissen et al., 1994) and with bound 1,6-anhydromuropeptide (van Asselt, 1999a) or bound inhibitor bulgecin A (Thunnissen et al., 1995b) (Fig. 5); E. coli MltA alone (van Straaten, 2005) or with bound chitohexaose (van Straaten, 2007) (Fig. 5); Neisseria gonorrhoeae MltA (Powell et al., 2006); E. coli Slt35 alone (van Asselt, 1999b) and in complex with two different peptidoglycan fragments or bulgecin A (van Asselt, 2000) (Fig. 5); and the phage λ lytic transglycosylase with bound hexa-2N-acetylchitohexaose (Leung et al., 2001).

Structures of lytic transglycosylases. (a) Membrane-bound lytic transglycosylase MltA (E. coli), PDB code 2PI8; left side, the chitohexaose-bound MltA-D308A mutant with the chitohexaose shown as spheres and the catalytic Asp308Ala as a green sphere; right side, zoomed view of the catalytic cleft, with the hexasaccharide and the catalytic Asp308Ala represented as sticks. (b) Soluble lytic transglycosylase Slt35 (E. coli), PDB code 1D0K; left side, Slt35 in complex with two muropeptides GlcNAc-MurNAc-l-Ala-d-iGlu, represented in spheres, with the catalytic Glu162 as a green sphere; right side, zoomed side-on view on the large catalytic cleft, with the two muropeptides and the catalytic Glu162 shown as sticks. (c) Soluble lytic transglycosylase Slt70 (E. coli), PDB code 1QTE; left side, view down the large central hole of the doughnut, with a 1,6-anhydromuropeptide and the catalytic Glu478 (in green) represented as spheres; right side, zoomed side-on view of the catalytic cleft, with 1,6-anhydromuropeptide and the catalytic Glu478 represented as sticks.
Figure 5

Structures of lytic transglycosylases. (a) Membrane-bound lytic transglycosylase MltA (E. coli), PDB code 2PI8; left side, the chitohexaose-bound MltA-D308A mutant with the chitohexaose shown as spheres and the catalytic Asp308Ala as a green sphere; right side, zoomed view of the catalytic cleft, with the hexasaccharide and the catalytic Asp308Ala represented as sticks. (b) Soluble lytic transglycosylase Slt35 (E. coli), PDB code 1D0K; left side, Slt35 in complex with two muropeptides GlcNAc-MurNAc-l-Ala-d-iGlu, represented in spheres, with the catalytic Glu162 as a green sphere; right side, zoomed side-on view on the large catalytic cleft, with the two muropeptides and the catalytic Glu162 shown as sticks. (c) Soluble lytic transglycosylase Slt70 (E. coli), PDB code 1QTE; left side, view down the large central hole of the doughnut, with a 1,6-anhydromuropeptide and the catalytic Glu478 (in green) represented as spheres; right side, zoomed side-on view of the catalytic cleft, with 1,6-anhydromuropeptide and the catalytic Glu478 represented as sticks.

N-Acetyl-β-d-glucosaminidases

Endo-N-acetyl-β-d-glucosaminidases (termed ‘N-acetylglucosaminidases’ in this review) are widespread in bacteria. They hydrolyse the glycosidic bond between N-acetyl-β-d-glucosamine residues and adjacent monosaccharides in different oligosaccharide substrates including peptidoglycan, chitin and N-glycans (Karamanos, 1997). Most of these enzymes have a catalytic domain belonging to the protein family (Pfam) 01832. N-acetylglucosaminidases cleaving in peptidoglycan often have one or several CBDs (e.g. the LysM domain, see ‘CBDs associated with peptidoglycan hydrolases’). Lactococcus lactis has three known (AcmA, AcmB and AcmC) and one hypothetical (AcmD) N-acetylglucosaminidase, two of which (AcmA and AcmD) having three LysM peptidoglycan-binding domains and one (AcmB) having another, putative CBD (Huard et al., 2003, 2004; Steen et al., 2005). Interestingly, the presence of three LysM domains was shown to be optimal for the activity of AcmA because protein variants with less or more LysM domains were less active (Steen et al., 2005). Similarly, the N-acetylglucosaminidase AtlA from Enterococcus faecalis has six LysM domains required for optimal enzymatic activity (Eckert et al., 2006), whereas LytB from S. pneumoniae has 18 imperfect repeats mediating its binding to teichoic acid phosphocholine residues (De Las Rivas, 2002). Other recently identified and characterized peptidoglycan N-acetylglucosaminidases are AtlA from S. aureus (a bifunctional enzyme also containing an amidase domain) (Oshida et al., 1995; Biswas et al., 2006), Acd from Clostridium difficile (Dhalluin et al., 2005) and LytD and LytG from B. subtilis (Margot et al., 1994; Rashid et al., 1995; Horsburgh et al., 2003b).

In contrast to the endo-N-acetylglucosaminidases mentioned above, which hydrolyse glycosidic bonds in high-MW peptidoglycan, the cytoplasmic N-acetylglucosaminidase NagZ from E. coli utilizes a disaccharide substrate, GlcNAc-1,6-anhydroMurNAc(peptide), which is an intracellular intermediate in the peptidoglycan-recycling pathway (see ‘Cytoplasmic hydrolases for peptidoglycan recycling and induction of β-lactamases’) (Vötsch & Templin, 2000). The crystal structure of NagZ from Vibrio cholerae has been solved recently with a bound small-molecule inhibitor mimicking the putative transition state of the catalytic reaction (Stubbs et al., 2007).

CBDs associated with peptidoglycan hydrolases

Hydrolytic enzymes degrading a high-MW substrate (such as cellulose or chitin) often have a substrate-binding domain, whose presence greatly enhances enzymatic activity (Gilbert et al., 2007). The same is true for many peptidoglycan hydrolases. At least seven different proven or hypothetical CBDs are reported in the literature: (1) the cysteine and histidine-dependent amidohydrolases/peptidases (CHAP) domain (Bateman & Rawlings, 2003); (2) the GW domain for noncovalent attachment to lipoteichoic acid (Scott & Barnett, 2006); (3) the SH3 domain (Whisstock & Lesk, 1999); (4) the PlyPSA CBD (Korndorfer et al., 2006); (5) the FtsN-peptidoglycan-binding domain (Ursinus et al., 2004; Yang et al., 2004); (6) the choline-binding domain (Fernandez-Tornero et al., 2002); and (7) the LysM CBD (Bateman & Bycroft, 2000). The structures of the last four CBDs are known and they can be fused to cell wall hydrolase domains of different specificity.

For example, the autolytic amidase (Ami) in L. monocytogenes is bound to cell wall lipoteichoic acid via a domain containing GW modules (Milohanic et al., 2001). The GW domain also has a role in interaction with host cells and is present in peptidoglycan hydrolases from other species.

The LysM–CBD has been described initially for the E. hirae extracellular muramidase 2 and consists of about 40 amino acid residues (Joris et al., 1992). This domain has been found in over 1500 proteins including many peptidoglycan hydrolases from all types of bacteria (Desvaux et al., 2006). LysM motifs are located generally at one end of a protein and vary in number from one to six. They bind directly to peptidoglycan and the exact number of motifs affects the efficiency of substrate binding as well as function (Steen et al., 2005). It is also present in many other bacterial proteins involved in pathogenicity such as staphylococcal IgG-binding proteins and E. coli intimin (Bateman & Bycroft, 2000). LysM-CBDs are also found in eukaryotic proteins. For example, in Medicago truncatula, a small legume, specific recognition of Sinorhizobium meliloti and its nodulation (Nod) factors requires the Nod factor perception gene (NFP), which encodes a putative receptor in which the extracellular sensor domain contains three tandem LysM CBDs (Mulder et al., 2006). The structure of the LysM–CBD of E. coli membrane-bound lytic peptidoglycan transglycosylase D (MltD) has been elucidated. The LysM-CBD consists of a βααβ secondary structure with the two helices packing on the same side (Bateman & Bycroft, 2000).

A choline-binding domain is present in several proteins from a range of species to anchor proteins to the choline residues present in the wall-associated polymers lipoteichoic and teichoic acid (Desvaux et al., 2006). The structure of the choline-binding domain of S. pneumoniae LytA has been solved and the molecular basis for interaction has been determined (Fernandez-Tornero et al., 2001).

Several peptidoglycan hydrolases contain direct repeat domains (often three) unrelated to the above-mentioned CBDs and required for cell wall binding (Margot & Karamata, 1992; Baba & Schneewind, 1998). In the case of Atl from S. aureus, the repeats target the various processed forms of the enzyme to the septum by binding to an unknown receptor (Baba & Schneewind, 1998; Biswas et al., 2006).

Physiological functions of extracytoplasmic peptidoglycan hydrolases

Extracytoplasmic bacterial peptidoglycan hydrolases have a variety of physiological functions for the producing cell and for the cell population (Table 1). In the following sections, the various functions of peptidoglycan hydrolases are listed and examples for each function are included.

Table 1

Functions of bacterial peptidoglycan hydrolases

Functions of peptidoglycan hydrolasesRemarksExamples
Functions in bacterial cell physiology
Regulation of cell wall growthRemoval of excess of pentapeptides in new peptidoglycan by dd-carboxypeptidasesPBP5 (E. coli); PBP3 (S. pneumoniae)
Enlargement of the peptidoglycan sacculusBreaking of bonds to allow expansion of the sacculus during growthLytic transglycosylases (E. coli); LytE (B. subtilis)
Peptidoglycan turnoverRelease of soluble peptidoglycan fragments from the sacculus during growthLytic transglycosylases (E. coli)
Production of signalling moleculesInduction of β-lactamase by peptidoglycan turnover productsLytic transglycosylases (E. coli)
Recycling of peptidoglycan turnover productsCleavage of turnover products to allow reuse in peptidoglycan synthesisAmpD, LdcA, NagZ (E. coli)
Cell separation during/after cell divisionCleavage of the septum during cell division in Gram-negative speciesAmiA, AmiB, AmiC, Slt70 (E. coli)
Cleavage of the cross-wall after division in Gram-positive speciesLytB (S. pneumoniae); Atl (S. aureus)
Roles in sporulation and germinationCleavage of the asymmetric septumSpoIID, SpoIIP (B. subtilis)
Spore cortex maturationLytH (B. subtilis)
Digestion of the peptidoglycan of the mother cell to release the endosporeLytC, CwlC, CwlH (B. subtilis)
Digestion of the spore peptidoglycan during germinationSleB, CwlJ (B. subtilis)
Assembly of secretion systemsSpecialized peptidoglycan hydrolases (lytic transglycosylases) for localized peptidoglycan degradation associated with type II, type III and type IV secretion systemsVirB1 (A. tumefaciens), TraB (E. coli plasmid R721)
Pilus assembly (type IV)Specialized peptidoglycan hydrolases for pilus assemblyPilT (E. coli EPEC strains)
Flagellum assemblySpecialized peptidoglycan hydrolases for flagellum assemblyFlgJ (E. coli)
Resuscitation of dormant cellsStimulation of cell division to exit dormant stateRpfA (M. luteus), Rpf proteins (M. tuberculosis)
Functions in bacterial populations
Autolysis in genetic transformationFratricide in S. pneumoniae: induced lysis (allolysis) of noncompetent cellsLytA, LytC (S. pneumoniae)
Developmental lysisLysis of cells during fruiting body formation in M. xanthus
Cannibalism in B. subtilis
Lysis of prey cellsSecretion of peptidoglycan hydrolases to digest peptidoglycan of prey cellsExoenzymes of M. xanthus
Lysis of nonimmune cellsPlasmid-encoded bacteriocin/immunity factorPesticin (Yersinia pestis)
Biofilm formationPeptidoglycan hydrolases are required for initial attachment of cells to hydrophobic surfacesAtlE (S. epidermidis)
Pathogen–host interactionRelease of peptidoglycan fragments recognized by the host
Functions of peptidoglycan hydrolasesRemarksExamples
Functions in bacterial cell physiology
Regulation of cell wall growthRemoval of excess of pentapeptides in new peptidoglycan by dd-carboxypeptidasesPBP5 (E. coli); PBP3 (S. pneumoniae)
Enlargement of the peptidoglycan sacculusBreaking of bonds to allow expansion of the sacculus during growthLytic transglycosylases (E. coli); LytE (B. subtilis)
Peptidoglycan turnoverRelease of soluble peptidoglycan fragments from the sacculus during growthLytic transglycosylases (E. coli)
Production of signalling moleculesInduction of β-lactamase by peptidoglycan turnover productsLytic transglycosylases (E. coli)
Recycling of peptidoglycan turnover productsCleavage of turnover products to allow reuse in peptidoglycan synthesisAmpD, LdcA, NagZ (E. coli)
Cell separation during/after cell divisionCleavage of the septum during cell division in Gram-negative speciesAmiA, AmiB, AmiC, Slt70 (E. coli)
Cleavage of the cross-wall after division in Gram-positive speciesLytB (S. pneumoniae); Atl (S. aureus)
Roles in sporulation and germinationCleavage of the asymmetric septumSpoIID, SpoIIP (B. subtilis)
Spore cortex maturationLytH (B. subtilis)
Digestion of the peptidoglycan of the mother cell to release the endosporeLytC, CwlC, CwlH (B. subtilis)
Digestion of the spore peptidoglycan during germinationSleB, CwlJ (B. subtilis)
Assembly of secretion systemsSpecialized peptidoglycan hydrolases (lytic transglycosylases) for localized peptidoglycan degradation associated with type II, type III and type IV secretion systemsVirB1 (A. tumefaciens), TraB (E. coli plasmid R721)
Pilus assembly (type IV)Specialized peptidoglycan hydrolases for pilus assemblyPilT (E. coli EPEC strains)
Flagellum assemblySpecialized peptidoglycan hydrolases for flagellum assemblyFlgJ (E. coli)
Resuscitation of dormant cellsStimulation of cell division to exit dormant stateRpfA (M. luteus), Rpf proteins (M. tuberculosis)
Functions in bacterial populations
Autolysis in genetic transformationFratricide in S. pneumoniae: induced lysis (allolysis) of noncompetent cellsLytA, LytC (S. pneumoniae)
Developmental lysisLysis of cells during fruiting body formation in M. xanthus
Cannibalism in B. subtilis
Lysis of prey cellsSecretion of peptidoglycan hydrolases to digest peptidoglycan of prey cellsExoenzymes of M. xanthus
Lysis of nonimmune cellsPlasmid-encoded bacteriocin/immunity factorPesticin (Yersinia pestis)
Biofilm formationPeptidoglycan hydrolases are required for initial attachment of cells to hydrophobic surfacesAtlE (S. epidermidis)
Pathogen–host interactionRelease of peptidoglycan fragments recognized by the host
*

Only examples are given, see text for references.

Table 1

Functions of bacterial peptidoglycan hydrolases

Functions of peptidoglycan hydrolasesRemarksExamples
Functions in bacterial cell physiology
Regulation of cell wall growthRemoval of excess of pentapeptides in new peptidoglycan by dd-carboxypeptidasesPBP5 (E. coli); PBP3 (S. pneumoniae)
Enlargement of the peptidoglycan sacculusBreaking of bonds to allow expansion of the sacculus during growthLytic transglycosylases (E. coli); LytE (B. subtilis)
Peptidoglycan turnoverRelease of soluble peptidoglycan fragments from the sacculus during growthLytic transglycosylases (E. coli)
Production of signalling moleculesInduction of β-lactamase by peptidoglycan turnover productsLytic transglycosylases (E. coli)
Recycling of peptidoglycan turnover productsCleavage of turnover products to allow reuse in peptidoglycan synthesisAmpD, LdcA, NagZ (E. coli)
Cell separation during/after cell divisionCleavage of the septum during cell division in Gram-negative speciesAmiA, AmiB, AmiC, Slt70 (E. coli)
Cleavage of the cross-wall after division in Gram-positive speciesLytB (S. pneumoniae); Atl (S. aureus)
Roles in sporulation and germinationCleavage of the asymmetric septumSpoIID, SpoIIP (B. subtilis)
Spore cortex maturationLytH (B. subtilis)
Digestion of the peptidoglycan of the mother cell to release the endosporeLytC, CwlC, CwlH (B. subtilis)
Digestion of the spore peptidoglycan during germinationSleB, CwlJ (B. subtilis)
Assembly of secretion systemsSpecialized peptidoglycan hydrolases (lytic transglycosylases) for localized peptidoglycan degradation associated with type II, type III and type IV secretion systemsVirB1 (A. tumefaciens), TraB (E. coli plasmid R721)
Pilus assembly (type IV)Specialized peptidoglycan hydrolases for pilus assemblyPilT (E. coli EPEC strains)
Flagellum assemblySpecialized peptidoglycan hydrolases for flagellum assemblyFlgJ (E. coli)
Resuscitation of dormant cellsStimulation of cell division to exit dormant stateRpfA (M. luteus), Rpf proteins (M. tuberculosis)
Functions in bacterial populations
Autolysis in genetic transformationFratricide in S. pneumoniae: induced lysis (allolysis) of noncompetent cellsLytA, LytC (S. pneumoniae)
Developmental lysisLysis of cells during fruiting body formation in M. xanthus
Cannibalism in B. subtilis
Lysis of prey cellsSecretion of peptidoglycan hydrolases to digest peptidoglycan of prey cellsExoenzymes of M. xanthus
Lysis of nonimmune cellsPlasmid-encoded bacteriocin/immunity factorPesticin (Yersinia pestis)
Biofilm formationPeptidoglycan hydrolases are required for initial attachment of cells to hydrophobic surfacesAtlE (S. epidermidis)
Pathogen–host interactionRelease of peptidoglycan fragments recognized by the host
Functions of peptidoglycan hydrolasesRemarksExamples
Functions in bacterial cell physiology
Regulation of cell wall growthRemoval of excess of pentapeptides in new peptidoglycan by dd-carboxypeptidasesPBP5 (E. coli); PBP3 (S. pneumoniae)
Enlargement of the peptidoglycan sacculusBreaking of bonds to allow expansion of the sacculus during growthLytic transglycosylases (E. coli); LytE (B. subtilis)
Peptidoglycan turnoverRelease of soluble peptidoglycan fragments from the sacculus during growthLytic transglycosylases (E. coli)
Production of signalling moleculesInduction of β-lactamase by peptidoglycan turnover productsLytic transglycosylases (E. coli)
Recycling of peptidoglycan turnover productsCleavage of turnover products to allow reuse in peptidoglycan synthesisAmpD, LdcA, NagZ (E. coli)
Cell separation during/after cell divisionCleavage of the septum during cell division in Gram-negative speciesAmiA, AmiB, AmiC, Slt70 (E. coli)
Cleavage of the cross-wall after division in Gram-positive speciesLytB (S. pneumoniae); Atl (S. aureus)
Roles in sporulation and germinationCleavage of the asymmetric septumSpoIID, SpoIIP (B. subtilis)
Spore cortex maturationLytH (B. subtilis)
Digestion of the peptidoglycan of the mother cell to release the endosporeLytC, CwlC, CwlH (B. subtilis)
Digestion of the spore peptidoglycan during germinationSleB, CwlJ (B. subtilis)
Assembly of secretion systemsSpecialized peptidoglycan hydrolases (lytic transglycosylases) for localized peptidoglycan degradation associated with type II, type III and type IV secretion systemsVirB1 (A. tumefaciens), TraB (E. coli plasmid R721)
Pilus assembly (type IV)Specialized peptidoglycan hydrolases for pilus assemblyPilT (E. coli EPEC strains)
Flagellum assemblySpecialized peptidoglycan hydrolases for flagellum assemblyFlgJ (E. coli)
Resuscitation of dormant cellsStimulation of cell division to exit dormant stateRpfA (M. luteus), Rpf proteins (M. tuberculosis)
Functions in bacterial populations
Autolysis in genetic transformationFratricide in S. pneumoniae: induced lysis (allolysis) of noncompetent cellsLytA, LytC (S. pneumoniae)
Developmental lysisLysis of cells during fruiting body formation in M. xanthus
Cannibalism in B. subtilis
Lysis of prey cellsSecretion of peptidoglycan hydrolases to digest peptidoglycan of prey cellsExoenzymes of M. xanthus
Lysis of nonimmune cellsPlasmid-encoded bacteriocin/immunity factorPesticin (Yersinia pestis)
Biofilm formationPeptidoglycan hydrolases are required for initial attachment of cells to hydrophobic surfacesAtlE (S. epidermidis)
Pathogen–host interactionRelease of peptidoglycan fragments recognized by the host
*

Only examples are given, see text for references.

Regulation of cell wall growth by dd-carboxypeptidases

dd-carboxypeptidases remove terminal d-alanine residues at position 5 of pentapeptides in the peptidoglycan to form tetrapeptides; this alters the substrate properties of the peptide. While pentapeptides can act as donors and acceptors in transpeptidation (cross-linking) reactions, the tetrapeptides can only serve as acceptors. Apparently, the controlled enlargement of the sacculus requires the trimming of an excess of pentapeptides to tetrapeptides by dd-carboxypeptidases in some but not all bacteria. For example, S. pneumoniae mutants lacking the dd-carboxypeptidase PBP3 form aberrant septa and have a thickened cell wall (Schuster et al., 1990; Morlot et al., 2004), and E. coli mutants lacking the dd-carboxypeptidase PBP5 and additional PBPs may have bended or even branched cell shapes (Denome et al., 1999; Nelson & Young, 2001), indicating that these enzymes are required for correct positioning of the division plane and for correct cell shape. PBP5 (dacA) is the major dd-carboxypeptidase found in vegetative cells of B subtilis (Lawrence & Strominger, 1970). PBP5 has no role in spore peptidoglycan synthesis but two other dd-carboxypeptidases, PBP5* (dacB) and DacF (dacF), function in regulating the degree of cross-linking of spore peptidoglycan (Popham et al., 1999). DacB is expressed only in the mother cell compartment of the developing sporangium whereas dacF is expressed in the forespore compartment and they can act differently on the nascent spore peptidoglycan (Popham et al., 1999). The proportion of pentapeptides was largely increased in a dacA (PBP5) mutant of B. subtilis (Atrih et al., 1999). In L. monocytogenes, cells lacking PBP5 displayed an irregular morphology and shape (Guinane et al., 2006) and a thicker cell wall (Korsak et al., 2005a). Moreover, the ratio of pentapeptides to tripeptides was increased in cells lacking PBP5. Other species such as Caulobacter crescentus appear to lack dd-carboxypeptidase activity and have a high content of pentapeptides in their peptidoglycan (Markiewicz et al., 1983).

Enlargement of the peptidoglycan sacculus

The peptidoglycan sacculus is a giant bag-shaped and net-like molecule. In order to expand such a network, new subunits have to be inserted by peptidoglycan synthases. It is impossible to envision a mechanism of expansion of such a net that does not include the rupture of covalent bonds by hydrolases (Weidel & Pelzer, 1964; Shockman & Höltje, 1994). Thus, deletion of peptidoglycan hydrolases genes required for this process should result in growth arrest due to failure to expand the sacculus. So far, there is no example of such a phenotype of a single or a multiple deletion (or depletion) mutant of any species. This is probably due to the occurrence of multiple hydrolases in most species examined, making it difficult if not impossible to delete all corresponding genes. For example, E. coli has as many as 12 known peptidoglycan hydrolases with a periplasmic location (Vollmer & Bertsche, 2007). The following observations provide indirect evidence for the participation of hydrolases in cell wall growth and for a link between peptidoglycan synthesis and hydrolysis during growth.

There is no doubt that hydrolases are active during wall growth in cell elongation because there is a release of peptidoglycan fragments from the wall during growth, a process termed peptidoglycan turnover in both Gram-negative (Goodell & Schwarz, 1983, 1985; Goodell et al., 1985) and Gram-positive species (Pooley et al., 1976; de Boer, 1981). Escherichia coli loses about 40% of the existing peptidoglycan in one generation, and the structure of the turnover products implies that they are products of lytic transglycosylases, endopeptidases and amidases. Furthermore, the amount of turnover products is reduced in hydrolase mutants of E. coli (Kraft et al., 1999). It has been proposed by Höltje that synthetic and hydrolytic reactions are coordinated temporarily and spatially for a safe enlargement of the sacculus by the formation of multi-enzyme complexes combining different murein synthases and hydrolases. This would ensure that the hydrolases are active only at sites of new synthesis (Höltje, 1996a, c, 1998). A block in peptidoglycan biosynthesis, either at early steps of precursor synthesis or the late periplasmic steps (e.g. by inhibition of the PBPs), results in an uncontrolled activity of hydrolases (autolysins) and cell lysis. This observation is consistent with multienzyme complexes, which are fully active in the hydrolytic subcomplex but are (partly) blocked in their synthetic subcomplex. This view has been backed up by indirect genetic evidence. Cells lyse if catalytically inactive variants of the synthase PBP1B (harbouring single amino acid exchanges) are overproduced in wild-type cells, but not in mutants lacking the activities of the lytic transglycosylases (Meisel et al., 2003). It was proposed that PBP1B is active in a complex with lytic transglycosylase(s), and that the overproduced inactive enzyme replaced active PBP1B from the complexes, resulting in lysis due to the activity of lytic transglycosylases. Indeed, affinity chromatography experiments with immobilized murein enzymes identified several direct and/or indirect interactions between lytic transglycosylases and peptidoglycan synthases. Immobilized MltA retains the murein synthases PBP1B, PBP1C, PBP2 and PBP3 as well as non-PBP proteins (Vollmer et al., 1999). One of these non-PBP-proteins, MipA, was identified as a structural protein mediating the interaction between MltA and PBP1B. Immobilized Slt70 retains the synthases PBP1B, PBP1C, PBP2, PBP3 as well as the hydrolase PBP7/8 from crude membrane extract, whereas immobilized MltB specifically binds PBP1B, PBP1C and PBP3 (Romeis & Höltje, 1994b;,von Rechenberg, 1996).

Rod-shaped Gram-positive bacteria grow according to an inside-to-outside growth mechanism (Pooley & Shockman, 1970; Archibald et al., 1976; Koch & Doyle, 1985). Accordingly, new peptidoglycan is attached underneath the existing wall. Autolysins cleave in the stress-bearing layer(s), which results in stretching the new material that will eventually bear the stress. The outermost wall layers are degraded and released into the growth medium as turnover material. Cell wall turnover has been studied extensively in B. subtilis (Hughes et al., 1970; Pooley et al., 1976). Importantly, there is a correlation between the growth rate and the rate of cell wall turnover (de Boer, 1981, 1982; Cheung et al., 1983), indicating that the hydrolases are more active at high growth rates. Because of the requirement to cleave covalent bonds to allow cell wall enlargement, it has been speculated that hydrolases are pacemakers for cell growth (Höltje, 1995). A direct interaction between peptidoglycan synthases and hydrolases, as has been observed in E. coli, appears to be impossible in Gram-positive species, because they are physically separated. The synthases form the inner, new layers whereas the hydrolases act, well apart from the synthases, on the outer layers. In most rod-shaped species, including E. coli and B. subtilis, peptidoglycan synthesis is controlled by the actin-like, cytoplasmic MreB proteins. These cytoskeletal elements form helical filaments at the cytoplasmic membrane and direct peptidoglycan synthesis presumably by localizing the cell wall synthesis complexes (Carballido-Lopez & Errington, 2003; Daniel & Errington, 2003). In B. subtilis, the peptidoglycan hydrolase LytE has been shown to interact with the MreB-homologue MreBH, which is required for correct localization of LytE on the cylindrical part of the cell wall (Carballido-Lopez et al., 2006). Consequently, lytE and mreBH mutants show similar morphological, cell wall-related defects. Thus, peptidoglycan synthases and hydrolases required for cell elongation might be colocalized in rod-shaped, Gram-positive bacteria by the intracellular cytoskeleton.

Separation of daughter cells

It is well established that peptidoglycan hydrolases are required to cleave the septum to allow separation of daughter cells. In most Gram-positive species, a septum is formed and then cleaved along a middle line. In contrast, in most Gram-negative species, cleavage of the septum occurs almost simultaneously with cell division, leading to the typical V-shaped constriction rather than septation. There are many examples of mutants lacking one or several hydrolases and having a defect in cell separation. For example, a coccoid S. aureus atl null mutant forms clusters of nonseparated cells (Takahashi et al., 2002). A lytB mutant of S. pneumoniae forms long chains of non-separated cells that could be dispersed by the addition of purified LytB enzyme (Garcia et al., 1999b). As expected, both hydrolases, Atl and LytB, localized at the site of cell division (Yamada et al., 1996; Garcia et al., 1999b).

Multiple hydrolases cleave the septum during cell division in E. coli. Cells of the amiC deletion mutant separate poorly during division, and a fraction of 20–30% of the cell population exists as a chain of three to six unseparated cells. In these chains, the cells are separated by a septum made of peptidoglycan sandwiched between the cytoplasmic membranes of neighbouring cells. If amiA is inactivated, only 5–10% of the population grows as chains of three to four cells. Deletion of amiB produces no chaining phenotype at all. Deletion of three amidase genes (encoding for AmiA, AmiB and AmiC) or of six lytic transglycosylase genes (Slt70, MltA, MltB, MltC, MltD and EmtA) results in the formation of chains with up to c. 20 cells per chain (Heidrich et al., 2001, 2002). Also the endopeptidases PBP4, PBP7 and MepA contribute to cleavage of the septum. If deletions in amidases are combined with deletions in lytic transglycosylases or endopeptidases, the chains become very long and may contain up to 100 cells (Heidrich et al., 2002; Priyadarshini et al., 2006). These cells have additional defects including the presence of aberrant intracellular membrane structures and increased permeability of the outer membrane, making them susceptible to antibiotics that do not kill wild-type E. coli cells (e.g. vancomycin) (Heidrich et al., 2002; Korsak et al., 2005b). In N. gonorrhoeae, deletion of a single amidase gene, amiC, results in a defect in cell separation and in the formation of cell clumps (Garcia & Dillard, 2006).

Peptidoglycan hydrolases active in sporulation and germination

Under conditions of nutrient deprivation, Bacillus and Clostridium species undergo a differentiation process that leads to the production of highly resistant endospores (Fig. 6) (Errington et al., 2003; Piggot & Hilbert, 2004). Endospores have two juxtaposed peptidoglycan layers of different structure: an inner primordial cell wall, which will become the basis of the new vegetative cell wall after germination, and an outer spore-specific cortex with a unique structure (Atrih et al., 1996). The spore retains an alert sensory mechanism that can respond to specific germinants, which eventually leads to outgrowth to form a new vegetative cell (Atrih et al., 1998). The morphological changes during differentiation indicate a number of stages where peptidoglycan hydrolases are involved (Fig. 6; Foster & Popham, 2002).

Peptidoglycan hydrolases in the life cycle of B. subtilis. Vegetative growth, symmetric division, cell wall expansion and turnover involve multiple peptidoglycan hydrolases. Progress of differentiation requires the activities of several specific enzymes. Note that B. subtilis has more than 30 known or hypothetical peptidoglycan hydrolase genes (Smith et al., 2000).
Figure 6

Peptidoglycan hydrolases in the life cycle of B. subtilis. Vegetative growth, symmetric division, cell wall expansion and turnover involve multiple peptidoglycan hydrolases. Progress of differentiation requires the activities of several specific enzymes. Note that B. subtilis has more than 30 known or hypothetical peptidoglycan hydrolase genes (Smith et al., 2000).

Digestion of the asymmetric septum

The first stage at which autolysin activity appears necessary for sporulation is hydrolysis of the asymmetric septum to allow prespore engulfment. In B. subtilis, this involves a complex of two autolysins, SpoIID and SpoIIP, both of which are required for septal hydrolysis and have peptidoglycan hydrolase activity (Abanes-De Mello, 2002; Chastanet & Losick, 2007). SpoIID is homologous to LytB, the modifier protein that enhances the activity of the major vegetative cell amidase LytC (Abanes-De Mello, 2002). SpoIIP has similarity to the catalytic domain of CwlV (an amidase of P. polymyxa) (Chastanet & Losick, 2007). SpoIID and SpoIIP form a complex with SpoIIM (an integral membrane protein). SpoIIM localizes the two autolysins at the septal membrane. Digestion of the asymmetric septal peptidoglycan requires both the activities (Chastanet & Losick, 2007).

Cortex maturation

The spore cortex, which surrounds the spore core, has a unique and highly conserved structure, which is essential for the maintenance of heat resistance and dormancy. The basic structure of cortex peptidoglycan is similar to that of B. subtilis vegetative peptidoglycan. The cortical peptidoglycan is subject to sporulation-specific modifications either during or after its assembly. About 50% of the disaccharide subunits in the cortex have the muramic acid δ-lactam structure, which is found only in bacterial endospores. The δ-lactams are not randomly distributed, but occur predominantly at every alternate disaccharide. The formation of δ-lactam residues requires two proteins acting in concert (Gilmore et al., 2004). CwlD is a sporulation-specific amidase, which cleaves off the peptide side-chain (Atrih et al., 1996; Popham et al., 1996). Then, PdaA deacetylates the N-acetyl muramic acid product and forms the lactam ring via a cyclization event (Fukushima et al., 2002; Gilmore et al., 2004).

The spore cortex is the least cross-linked of any peptidoglycan, at only 2.9% of muramic acid residues (Atrih et al., 1996), compared with that seen in vegetative cell walls, 29–33% (Atrih et al., 1999). A gradient of cross-linking across the cortex may allow inward mechanical pressure, maintaining a decreased volume and water content of the spore core and thus dormancy (Popham et al., 1999; Meador-Parton & Popham, 2000).

About 25% of muramic acid residues in cortex peptidoglycan have single l-alanine side chains (Atrih et al., 1996). These are formed by the action of an l-alanyl-d-glutamate endopeptidase encoded by lytH (Horsburgh et al., 2003a). LytH is required for full spore heat resistance but its loss does not affect spore water content or cortex cross-linking.

Mother cell lysis

The final morphologically observable event during sporulation is hydrolysis of the mother cell wall peptidoglycan to release the mature endospore. Two autolysins are present in large amounts at the time of mother cell lysis: the major vegetative amidase, LytC, and a 30 kDa amidase, CwlC, which is expressed only late in sporulation (Foster et al., 1992; Kuroda et al., 1993; Smith & Foster, 1995). Single inactivations of either lytC or cwlC have no effect on mother cell lysis. However, a mutant inactivated in both genes is blocked in mother cell lysis, and so the two amidases have a mutually compensatory role in the hydrolysis of the mother cell wall (Smith & Foster, 1995). A third sporulation-specific amidase (CwlH) has also been shown to be required for mother cell lysis in which it acts in a compensatory fashion with CwlC (Nugroho et al., 1999). It is possible that, as for asymmetric septum hydrolysis, a complex of autolytic amidases act in concert to degrade the mother cell wall and allow endospore release.

Germination

As the cortex is essential for the maintenance of spore dormancy, its specific hydrolysis is a requirement for germination, allowing subsequent uptake of water, core expansion and outgrowth (Foster & Johnstone, 1990). Biochemical analysis of spore peptidoglycan dynamics during germination of B. subtilis shows that a probable stereochemical change, a likely epimerase, occurs at or near the muramic δ-lactam (Atrih et al., 1998). Accompanying this, hydrolytic events, primarily lytic transglycosylase and to a lesser extent glucosaminidase, cause release of the cortical peptidoglycan as muropeptide fragments (Atrih et al., 1998). The muramic δ-lactam is crucial for substrate recognition by the germination-specific lytic enzymes (GSLEs), as a cwlD mutant that lacks the δ-lactam, produces resistant endospores, which cannot outgrow as the cortex is refractile to GSLE activity (Sekiguchi et al., 1995; Atrih et al., 1996, 1998; Popham et al., 1996). The primordial cell wall, which does not have δ-lactam, is not hydrolysed and remains to develop into the cell wall of the outgrowing spore.

Bacillus subtilis has 2 GSLEs (SleB and CwlJ) required for the later stages of germination, as a double mutant is totally blocked in cortex hydrolysis and outgrowth (Ishikawa et al., 1998). Loss of SleB results in slow germination and coincides with a total lack of lytic transglycosylase activity during germination (Boland et al., 2000). SleB is present in the spore integument and inner membrane and its presence in the dormant spore depends on a cotranscribed gene, ypeB (Moriyama et al., 1999; Chirakkal et al., 2002). YpeB may act as an anchor to give correct SleB localization. How SleB is activated as part of the germination process is unknown. CwlJ is present in the spore coats and can be activated during germination by the release of dipicolinate, as an early event during spore germination (Paidhungat et al., 2001; Bagyan & Setlow, 2002). The rationale for the possession of apparently functionally redundant GSLEs is justified by their different activity optima and resistance to deleterious treatments, thus allowing the germination mechanism to be active under a broad range of conditions and after environmental assault (Atrih & Foster, 2001).

Assembly of large trans-envelope structures

The pores in the stretched peptidoglycan of E. coli are quite uniform in size and they allow penetration of globular proteins of up to 50–100 kDa (Demchick & Koch, 1996; Vazquez-Laslop et al., 2001). However, these pores appear to be too small for the assembly of large trans-envelope complexes such as secretion systems of types II, III and IV, type IV pili and flagella. The large gene clusters encoding for proteins for the regulation and assembly of these structures typically contain a peptidoglycan hydrolase gene, most often encoding for a lytic transglycosylase (Koraimann, 2003). These are specialized hydrolases required to enlarge gaps locally in the peptidoglycan to allow penetration of trans-envelope structures. These hydrolases include, for example, the VirB1 enzyme encoded from the plasmid Ti of Agrobacterium tumefaciens, which is required for a type IV secretion system (DNA transport into plant cells) (Zahrl et al., 2005), the PilT enzyme involved in type IV pilus synthesis in E. coli (Koraimann, 2003) or the FlgJ enzyme for flagellum synthesis in E. coli (Nambu et al., 1999).

Resuscitation of dormant cells

More than 60 bacterial species including many pathogens were reported to be able to enter a viable-but-nonculturable (VBNC) state in which they do not divide and have very low metabolic activity (Oliver, 2005). For example, Micrococcus luteus can enter such a dormant state by prolonged stationary-phase culturing. Resting cells of M. luteus can be stimulated to divide (resuscitated) by addition of the RpfA protein (rpf, resuscitation-promoting factor) to the culture medium (Mukamolova et al., 1998). Mycobacterium tuberculosis has five RpfA-like proteins, which presumably have redundant growth-promoting activities on resting cells. Single and double deletions of Rpf proteins have no effect but different combinations of triple deletions generated mutants that cannot resuscitate spontaneously in culture (Downing et al., 2005). Strikingly, Rpf proteins have weak similarity in amino acid sequence to lysozymes and lytic transglycosylases. This has been confirmed by analysing the crystal structure of RpfB from M. tuberculosis and by demonstration of a muramidase activity of Rpf from M. luteus (Mukamolova et al., 2006). Either these peptidoglycan hydrolases are required for daughter cell separation during the first cell division following the resting period, or the released peptidoglycan fragments serve as signalling molecules initiating regrowth of the cells (Keep et al., 2006).

Lysis of prey cells and nonimmune cells

The Gram-negative myxobacteria are predatory bacteria living in the soil (Reichenbach, 1999). They have several unusual features making them unique among prokaryotes by virtue of their intercellular communication systems and their complex morphogenesis. Myxococcus xanthus became a model bacterium to study these aspects (Kaiser, 2003, 2004; Sogaard-Andersen, 2004). Swarms of thousands of myxobacterial cells glide on surfaces and they attack and lyse other microorganisms, especially many kinds of bacteria on which they feed (Singh, 1947). For this, the myxobacteria produce a plethora of extracellular lytic agents including antibiotics, bacteriocins, lipases, proteases and cell wall hydrolases. For example, lysozyme activity was shown in the culture supernatant of Myxococcus fulvus (Draft et al., 1985), and an endo-β-N-acetylglucosaminidase has been purified from the culture medium of Stigmatella aurantiaca (Bourgerie et al., 1994). The formation of large swarms of cells strongly enhances the efficiency of the secreted lytic agents because their local concentrations are higher when they are produced simultaneously from many cells. The production of bacteriolytic exo-enzymes is not limited to myxobacteria and is also a property of many species of actinomycetes, in particular of streptomycetes (Ghuysen, 1968; Heymer & Schmidt, 1975). In fact, several peptidoglycan hydrolases from myxobacteria and streptomycetes are valuable tools to produce peptidoglycan fragments for chemical analysis (Ghuysen, 1968; Sudo & Dworkin, 1972).

Many bacteria secrete one or more lytic bacteriocin together with an immunity factor for self-protection against its activity. There are different modes of action of bacteriocins, which are often small peptides, for example disruption of the cytoplasmic membrane or inhibition of peptidoglycan synthesis by binding to lipid II of the target cells. In contrast, pesticin is a plasmid-encoded muramidase (lysozyme) of wild-type strains of Y. pestis and is active against certain strains of E. coli, Yersinia pseudotuberculosis and Yersinia enterocolitica (Vollmer et al., 1997).

Developmental lysis

Upon starvation, myxobacteria aggregate to form a fruiting body with myxospores in a highly ordered developmental process. Developmental lysis in M. xanthus describes the autolysis of most of the cells – 80–90% of the initial population – within the first 72 h of development of the fruiting body. It is assumed that lysis of these cells is a regulated phenomenon of ‘programmed cell death’ and is of benefit for the subset of cells determined to form the myxospores. According to this view, the nutrients released by the lysed cells would feed the sporulating cells (Rosenbluh & Rosenberg, 1993; Lewis et al., 2000). Developmental lysis is induced by auticides, a mixture of at least five lipid compounds and by glucosamine. The mechanism leading from induction to lysis as well as the mechanism for regulation and activation of the autolysins are as yet unknown.

A similar lysis phenomenon has been observed in the early sporulation stage of B. subtilis although this species does not form a fruiting body. A fraction of starved cells secrete factors to kill and lyse sibling cells that have not developed immunity to these compounds. This behaviour has been termed ‘cannibalism’ and it has been speculated that its biological function for the producing cell is to get nutrients to delay or avoid sporulation (Gonzalez-Pastor et al., 2003; Ellermeier et al., 2006).

Autolysis in genetic transformation (allolysis)

In S. pneumoniae, competence for genetic transformation is induced in the early exponential phase by a quorum-sensing mechanism involving a peptide hormone, the competence-stimulating factor. Cells in the competent state are capable of taking up exogenous DNA, which might then be integrated into the chromosome. Competent cells are more prone to autolyse (Seto & Tomasz, 1975), and the two autolysins of this bacterium, LytA and LytC, contribute to the release of DNA in a culture of competent cells (Moscoso & Claverys, 2004). Results from micro-array experiments indicate that there is a connection between competence, stress responses and autolysis (Dagkessamanskaia et al., 2004), and competent cells were shown to be able to lyse competence-deficient cells during cocultivation (Steinmoen et al., 2003). A careful examination of the genetic requirements for the lysis of noncompetent cells revealed a new phenomenon that was termed ‘allolysis’ (Guiral et al., 2005). Accordingly, competent cells are able to trigger lysis of cocultivated noncompetent cells, resulting in the release of DNA from the lysed cells. The mechanism involves the two-peptide bacteriocin CibAB, its immunity factor CibC, the two autolysins LytA and LytC and the putative murein hydrolase CbpD. A model was proposed, according to which the bacteriocin (and the immunity factor) is produced by competent cells. It is active and kills only noncompetent cells, because they do not produce the immunity factor. The bacteriocin activates the autolysins, which disrupt the cell wall of the predated, noncompetent cells resulting in lysis. Induced lysis of genetically identical cells has been termed ‘fratricide’. The biological role of fratricide for S. pneumoniae in the context of its particular ecological niche (the nasopharynx of humans) and its population dynamics is yet to be defined. Cannibalism (see ‘Developmental lysis’) and fratricide in bacterial populations are discussed in a recent review (Claverys & Havarstein, 2007).

Biofilm formation

Most bacteria in their natural environment are associated with solid surfaces, be this simply attached or in more complex bacterial community structures as biofilms (Costerton et al., 1987). Biofilms are of particular importance for pathogens when associated with host tissues or indwelling medical devices as the bacteria are often refractory to antibiotic treatment. Staphylococcus epidermidis is an opportunist pathogen very frequently found on infected catheters as a biofilm. The major autolysin AtlE is required for biofilm formation on polymer surfaces as an atlE mutant is biofilm negative and cannot mediate the initial attachment process (Heilmann et al., 1997). AtlE is also able to bind human ligands and so may be involved in host–pathogen interaction. AtlE is the equivalent of Atl from S. aureus, and loss of either results in a cell division defect and cluster formation (Foster, 1995; Oshida et al., 1995; Heilmann et al., 1997).

In L. lactis, the major autolysin AcmA is also required for biofilm formation whereby an acmA mutant forms nonadherent long chains of cells (Mercier et al., 2002). Addition of lysozyme disrupts the chains and results in a simultaneous increase in adherence. The role of the peptidoglycan hydrolase activity may be to alter surface charge or to expose other cellular adhesins. Also in Streptococcus mutans, the autolysin AtlA is necessary for the production of effective biofilms (Ahn & Burne, 2006).

Host–pathogen interaction

Recently, autolysins have been implicated in the interaction between several pathogens and their hosts. This occurs due to their ability to release bioactive products, act as adhesins, affect the mammalian immune system, be required for biofilm formation, allowing pathogen separation and dispersal and likely many other mechanisms. Such functions are covered in detail in recent review articles (Dziarski et al., 2003; Dziarski & Gupta, 2005, 2006; Mengin-Lecreulx & Lemaitre, 2005).

Regulation of extracytoplasmic hydrolase activity

Peptidoglycan hydrolases are potentially lethal enzymes able, in many cases, to degrade the cell wall of producing organisms. Thus, they must be highly regulated to prevent adventitious cell lysis. This regulation occurs at multiple levels from the transcriptional to the posttranslational. For many bacteria, cell lysis is a natural part of the lifecycle and in some scenarios it is in fact advantageous to remove damaged cells from the population (Rice & Bayles, 2003).

Genetic regulation

Control of production of peptidoglycan hydrolases is important but in many cases is overridden by biochemical regulation as many of the enzymes are found to be associated with their substrate but controlled at the level of activity. Most of the enzymes involved in sporulation and germination are controlled at the transcriptional level by the sporulation-specific cascade of sigma factors that allows their production only in particular sporangial compartments and at particular times for their exact functions. During vegetative growth, the control of peptidoglycan hydrolase expression has been particularly illuminating in the determination of the role of some enzymes (with a few particular examples discussed below).

Coregulation with flagellation and motility

In B. subtilis, most autolysin activity is produced as the cells enter the stationary phase (Foster, 1992). The two major enzymes are LytC and LytD, an amidase and N-acetylglucosaminidase, respectively (Margot & Karamata, 1992; Margot et al., 1994). Their production coincides with major morphological and physiological changes to the cells. The short chains break up into single cells, they become motile, secrete large amounts of extracellular enzymes and develop natural competence. Many of these processes are coregulated by the alternative sigma factor SigD (Serizawa et al., 2004). SigD controls expression of the ‘flagellar, chemotaxis and motility regulon’.

Transcription of the lytABC operon proceeds from two promoters. One is controlled by σD, which accounts for 70–90% of the transcription during growth (Lazarevic et al., 1992; Kuroda & Sekiguchi, 1993). Ninety-five percent of lytD transcription is controlled by σD (Margot et al., 1994). The physiological rationale for the coregulation of the autolysins with motility lies in the phenotype of their respective mutants. The enzymes are required for de-chaining as cells linked together are unable to chemotax effectively, the so-called pushmi-pullyu effect (Blackman et al., 1998).

In S. aureus, perturbation of cell wall synthesis, either by the addition of a cell wall-targeting antibiotic at sub-inhibitory concentration or by lowering the transcription of a cell wall synthesis gene, results in the formation of a weaker cell wall (with less cross-links) that is more rapidly digested in vitro by peptidoglycan hydrolases. Surprisingly, cells grown under this condition show reduced autolysis rates, which was found to be due to reduced expression of autolysin genes. Thus, there appears to be an as yet unknown regulatory mechanism by which expression of autolysin genes is down-regulated when cell wall synthesis is disturbed (Antignac et al., 2007).

Elusive essential enzyme

For many years, it has been proposed that hydrolysis of existing bonds within peptidoglycan is necessary to allow wall expansion, cell growth and division (Shockman & Höltje, 1994). In a number of Gram-positive species, a homologous essential sensor regulator has been characterized (originally called YycGF) (Fabret & Hoch, 1998; Martin et al., 1999; Clausen et al., 2003). Analysis of its regulon has revealed the control of a number of putative and proven peptidoglycan hydrolases (Howell et al., 2003; Dubrac & Msadek, 2004). This alluded to their individual or combined essentiality for growth. Only in S. pneumoniae has an essential enzyme been identified. PcsB is essential and controlled by YycGF (VikSR) (Ng et al., 2003, 2004). The hydrolytic bond specificity of PcsB is unknown but it contains a CHAP domain present in several hydrolases that may have amidase function (Kajimura et al., 2005). Depletion of PcsB results in deregulated cell wall synthesis and bacteriostasis (Ng et al., 2004). Whether the essentiality of PcsB is due directly to its hydrolytic properties, and the molecular basis for its role, are currently unknown.

Programmed cell death

The LytSR sensor regulator of S. aureus has been proposed to respond to a decline in proton motive force across the cell membrane (Patton et al., 2006). LytSR controls the expression of lrgAB, which encode a postulated antiholin system (Brunskill & Bayles, 1996). LrgAB may interact with their cognate holins (CidAB) and prevent cell lysis (Rice & Bayles, 2003). Under conditions of stress (such as the presence of penicillin), CidAB may collapse the PMF and allow access of autolysins to their substrate, or deregulation of activity, resulting in cellular lysis (Rice & Bayles, 2003).

Posttranslational regulation

Subcellular localization of peptidoglycan hydrolases

In many species, defects in autolysin production lead to aberrant cell separation after septation. Thus, hydrolysis of the nascent septa is required for cell division. In S. aureus, loss of the major autolysin, Atl, results in cell cluster formation and microscopy studies have revealed a ring of Atl at the septum (Yamada et al., 1996). Amidase activity is required for septal hydrolysis in E. coli and the periplasmic AmiC has been found to localize specifically at the septum (Heidrich et al., 2001; Bernhardt & de Boer, 2003). In contrast, AmiA is found throughout the periplasm. More recently, the putative periplasmic endopeptidase EnvC has also been found to be septum associated (Bernhardt & De Boer, 2004). The Mre proteins of B. subtilis are required for maintenance of rod morphology and form a helical pattern along the cylindrical axis (Jones et al., 2001). MreBH interacts with the autolysin LytE (a putative endopeptidase), leading to its localization in a helical pattern and so coordinates cell wall hydrolysis with cylinder elongation (Carballido-Lopez et al., 2006). The subcellular localization of autolysins must reflect specific targeting mechanisms.

Peptidoglycan hydrolase targeting

Secretion of peptidoglycan hydrolases occurs primarily by the well-established Sec dependent pathway but can also use the TAT system and in some cases there is no signal sequence (Smith & Foster, 1995; Bernhardt & de Boer, 2003). The interaction between the enzymes and the cell wall is crucial for their activity. Several conserved mechanisms have been found for a number of enzymes. Covalent binding of autolysins to their substrate, which may render the enzymes inactive until they are further processed, is rare but has been reported for the GSLE of B. megaterium (Foster & Johnstone, 1988) and a sortase-linked enzyme from S. mutans (Catt & Gregory, 2005). Much more common are the use of conserved, often repeated, motifs or domains for ionic attachment to peptidoglycan or other cell wall components (see ‘CBDs associated with peptidoglycan hydrolases’). Many enzymes have an overall basic charge at neutral pH, which may enhance their binding to negatively charged cell wall components.

In E. coli, five lytic transglycosylases are made as lipoproteins and they are targeted to the outer membrane (Lommatzsch et al., 1997). It has been postulated that this specific orientation in respect of the peptidoglycan opposite to the biosynthetic machinery would allow cell wall expansion and growth (Lommatzsch et al., 1997).

Peptidoglycan hydrolase processing

Proteolytic processing of enzymes is common not only in their activation but also in their stability in response to a changing environment. The major autolysin of S. aureus called Atl is produced as a 138 kDa proenzyme that undergoes proteolytic processing to generate the two extracellular lytic enzymes found in the supernatant of S. aureus cultures identified as the 51 kDa endo-β-acetylglucosaminidase and a 62 kDa N-acetylmuramyl-l-alanine amidase (Foster et al., 1995; Oshida et al., 1995). Pro-Atl, amidase and glucosaminidase display peptidoglycan hydrolase activity in zymogram analysis. The signal peptide of 138 kDa pro-Atl is first cleaved, followed by further processing through intermediate forms of 115 and 85 kDa to result in the mature active 51 and 62 kDa cell surface-associated, and secreted, glucosaminidase and amidase (Komatsuzawa et al., 1997). Proteases have also been postulated to maintain a balance between production and degradation of peptidoglycan hydrolases during growth of B. subtilis to maintain appropriate activity levels (Yamamoto et al., 2003). Proteolytic cleavage of the lipoprotein precursor of a membrane-bound lytic transglycosylase of E. coli, MltB, results in the formation of a soluble form, the soluble lytic transglycosylase (Slt35). It is not known whether this ‘periplasmic solubilization’ of MltB has a particular physiological role (Ehlert et al., 1995).

Control of activity

The activity of peptidoglycan hydrolases must be highly regulated and a number of mechanisms have been proposed. These include physical or chemical alterations to the peptidoglycan substrate including conformational changes (Koch et al., 1985) and covalent modification (Clarke & Dupont, 1992; Atrih et al., 1998; Bera et al., 2005). The presence, absence or exact chemical nature of other wall polymers has also been shown to affect activity. Teichoic acids modify the activity of the amidase of B. subtilis (Herbold & Glaser, 1975a, b), and other amphiphiles such as cardiolipin and lipoteichoic acids, etc. are known to inhibit peptidoglycan hydrolase activities of many Gram-positive species (Höltje & Tomasz, 1975; Cleveland et al., 1976; Fischer et al., 1981). Growth at a low pH has also been shown to inhibit bacteriolysis (Goodell et al., 1976). This may be due to a change in the wall ionic environment affecting autolysin activity (Cheung & Freese, 1985) as the energized state of the membrane of B. subtilis is also important in regulation (Jolliffe et al., 1981; Calamita et al., 2001). It has also been proposed that bonds near the surface of Gram-positive cell walls, and the bonds at the middle of a division septum, are under stress and may be recognized by ‘smart autolysins’ (Koch et al., 1985; Koch et al., 1990).

For the enlargement of the thin sacculus of Gram-negative bacteria, Höltje has proposed that different peptidoglycan synthases and hydrolases form multi-enzyme complexes. This would allow the enlargement of the peptidoglycan layer by a defined mechanism and would restrict the activities of the potentially dangerous hydrolases to the sites of synthesis. As indicated above (‘Enlargement of the peptidoglycan sacculus’), there is evidence for the existence of protein–protein interactions between lytic transglycosylases and peptidoglycan synthases.

Cytoplasmic hydrolases for peptidoglycan recycling and induction of β-lactamase

The activity of exocytoplasmic peptidoglycan results in a massive peptidoglycan turnover in E. coli (Vollmer & Höltje, 2001). In one generation, about 40–50% of the peptidoglycan is released from the sacculus by lytic transglycosylases, dd-endopeptidases and amidases (Goodell & Schwarz, 1983, 1985; Goodell et al., 1985). The muropeptides released contain a 1,6-anhydroMurNAc residue and are subject to an efficient recycling pathway (Park et al., 1995, 1996). In E. coli and presumably in most other Gram-negative bacteria, the 1,6-anhydro-disaccharide-peptides are transported into the cytoplasm by the AmpG permease. Several cytoplasmic enzymes are involved in recycling of these peptidoglycan fragments. The AmpD amidase liberates the tripeptide l-Ala-γ-d-Glu-meso-A2pm (Höltje, 1994; Jacobs et al., 1995, 1997), which can then be attached to UDP-MurNAc by the Mpl ligase (Mengin-Lecreulx et al., 1996). Interestingly, AmpD has a strict specificity for 1,6-anhydroMurNAc-containing muropeptides and therefore does not hydrolyse the murein precursor molecules such as UDP-MurNAc-pentapeptide also present in the cytoplasm. The tripeptide can be further processed by the dl-carboxypeptidase MpaA, which releases the terminal meso-A2pm residue (Uehara & Park, 2003). The resulting l-Ala-d-Glu dipeptide can be degraded to the l-amino acids. For this, the epimerase YcjI forms l-Ala-l-Glu, which is then hydrolysed by the peptidase PepD (Schmidt et al., 2001; Uehara et al., 2005).

The GlcNAc-1,6-anhydroMurNAc disaccharide is hydrolysed by the N-acetylglucosaminidase NagZ to GlcNAc and 1,6-anhydroMurNAc (Vötsch & Templin, 2000). GlcNAc is converted in two steps by NagK and NagA to GlcN-6-phosphate, which is the substrate for both the peptidoglycan precursor UDP-GlcNAc and the central metabolite fructose-6-phosphate (Uehara & Park, 2004). 1,6-AnhydroMurNAc is converted by AnmK to MurNAc-6-phosphate, which is hydrolysed by the etherase MurQ to GlcNAc-6-phosphate (Schmidt et al., 2001; Uehara et al., 2005, 2006).

Interestingly, in Gram-negative bacteria having an inducible AmpC β-lactamase, the induction mechanism is directly linked to the relative cytoplasmic levels of 1,6-anhydromuropeptides and UDP-MurNAc-pentapeptide, both of which can bind to the transcriptional activator AmpR (Jacobs et al., 1997). Inactivation of the ampD gene leads to accumulation of 1,6-anhydromuropeptides and therefore to constitutive β-lactamase overexpression in C. freundii (Jacobs et al., 1997).

Concluding remarks

Since the discovery of lysozyme by Alexander Fleming, enormous knowledge has been gathered on the vast group of peptidoglycan hydrolases identified so far. Today, a comprehensive collection of all relevant literature on peptidoglycan hydrolases would fill many bookshelves, and it is impossible to summarize all the data in a single article. In this review, an overview has been presented on the known specificities of these hydrolases and a few examples of structures has been shown that have been solved recently. The knowledge is still incomplete, and it is likely that new enzyme families of peptidoglycan hydrolases will be discovered in the future. We have also exemplified the different roles of these enzymes in the physiology of a bacterial cell as well as at the bacterial population level. Perhaps one of the most important aspects is the regulation of the activity of peptidoglycan hydrolases. In many cases, knowledge of this is very limited. For example, all 12 known periplasmic peptidoglycan hydrolases from E. coli have the potential to degrade the sacculus and cause lysis of the cell. The question as to why this does not occur under normal growth conditions still remains to be answered. We will understand the roles of peptidoglycan hydrolases only if we know how their activities are regulated in the cell.

Acknowledgements

This work was supported in part by the European Commission through the EUR-INTAFAR (LSHM-CT-2004-512138, to WV, BJ and PC) and COBRA (LSMH-CT-2003-503335, to BJ and PC) projects, by the ‘Deutsche Forschungsgemeinschaft (DFG)’ within the ‘Forschergruppe Bakterielle Zellhülle’ (FOR 449, to WV), by the Belgian Program on Interuniversity Poles of Attraction initiated by the Belgian State, Prime Minister's Office, Science Policy programming (IAP no. P6/19, to BJ and PC), by the Actions de Recherche Concertées (grant 03/08-297. to BJ and PC), by the Fonds National de la Recherche Scientifique (IISN 4.4505.00, FRFC 9.45/9.99, FRFC 2.4.508.01.F, FRFC 9.4.538.03.F, FRFC2.4.524.03, to BJ and PC), the University of Liège (Fonds spéciaux, Crédit classique, 1999, to B.J and P.C.), the BBSRC (SJF) and the MRC (SJF).

References

Abanes-De Mello
A
Sun
YL
Aung
S
Pogliano
K
(
2002
)
A cytoskeleton-like role for the bacterial cell wall during engulfment of the Bacillus subtilis forespore
.
Genes Dev
16
:
3253
3264
.

Ahn
SJ
Burne
RA
(
2006
)
The atlA operon of Streptococcus mutans: role in autolysin maturation and cell surface biogenesis
.
J Bacteriol
188
:
6877
6888
.

Antignac
A
Sieradzki
K
Tomasz
A
(
2007
)
Perturbation of cell wall synthesis suppresses autolysis in Staphylococcus aureus: evidence for co-regulation of cell wall synthetic and hydrolytic enzymes
.
J Bacteriol
189
:
7573
7580
.

Archibald
AR
(
1976
)
Cell wall assembly in Bacillus subtilis: development of bacteriophage-binding properties as a result of the pulsed incorporation of teichoic acid
.
J Bacteriol
127
:
956
960
.

Arthur
M
Depardieu
F
Cabanie
L
Reynolds
P
Courvalin
P
(
1998
)
Requirement of the VanY and VanX d,d-peptidases for glycopeptide resistance in enterococci
.
Mol Microbiol
30
:
819
830
.

Atrih
A
Foster
SJ
(
2001
)
In vivo roles of the germination-specific lytic enzymes of Bacillus subtilis 168
.
Microbiology
147
:
2925
2932
.

Atrih
A
Zollner
P
Allmaier
G
Foster
SJ
(
1996
)
Structural analysis of Bacillus subtilis 168 endospore peptidoglycan and its role during differentiation
.
J Bacteriol
178
:
6173
6183
.

Atrih
A
Zollner
P
Allmaier
G
Williamson
MP
Foster
SJ
(
1998
)
Peptidoglycan structural dynamics during germination of Bacillus subtilis 168 endospores
.
J Bacteriol
180
:
4603
4612
.

Atrih
A
Bacher
G
Allmaier
G
Williamson
MP
Foster
SJ
(
1999
)
Analysis of peptidoglycan structure from vegetative cells of Bacillus subtilis 168 and role of PBP 5 in peptidoglycan maturation
.
J Bacteriol
181
:
3956
3966
.

Baba
T
Schneewind
O
(
1998
)
Targeting of muralytic enzymes to the cell division site of gram-positive bacteria: repeat domains direct autolysin to the equatorial surface ring of Staphylococcus aureus
.
Embo J
17
:
4639
4646
.

Bagyan
I
Setlow
P
(
2002
)
Localization of the cortex lytic enzyme CwlJ in spores of Bacillus subtilis
.
J Bacteriol
184
:
1219
1224
.

Barrett
JF
Dolinger
DL
Schramm
VL
Shockman
GD
(
1984
)
The mechanism of soluble peptidoglycan hydrolysis by an autolytic muramidase. A processive exodisaccharidase
.
J Biol Chem
259
:
11818
11827
.

Bateman
A
Bycroft
M
(
2000
)
The structure of a LysM domain from E. coli membrane-bound lytic murein transglycosylase D (MltD)
.
J Mol Biol
299
:
1113
1119
.

Bateman
A
Rawlings
ND
(
2003
)
The CHAP domain: a large family of amidases including GSP amidase and peptidoglycan hydrolases
.
Trends Biochem Sci
28
:
234
237
.

Bera
A
Herbert
S
Jakob
A
Vollmer
W
Götz
F
(
2005
)
Why are pathogenic staphylococci so lysozyme resistant? The peptidoglycan O-acetyltransferase OatA is the major determinant for lysozyme resistance of Staphylococcus aureus
.
Mol Microbiol
55
:
778
787
.

Bernhardt
TG
De Boer
PA
(
2003
)
The Escherichia coli amidase AmiC is a periplasmic septal ring component exported via the twin-arginine transport pathway
.
Mol Microbiol
48
:
1171
1182
.

Bernhardt
TG
De Boer
PA
(
2004
)
Screening for synthetic lethal mutants in Escherichia coli and identification of EnvC (YibP) as a periplasmic septal ring factor with murein hydrolase activity
.
Mol Microbiol
52
:
1255
1269
.

Biswas
R
Voggu
L
Simon
UK
Hentschel
P
Thumm
G
Götz
F
(
2006
)
Activity of the major staphylococcal autolysin Atl
.
FEMS Microbiol Lett
259
:
260
268
.

Blackburn
NT
Clarke
AJ
(
2001
)
Identification of four families of peptidoglycan lytic transglycosylases
.
J Mol Evol
52
:
78
84
.

Blackman
SA
Smith
TJ
Foster
SJ
(
1998
)
The role of autolysins during vegetative growth of Bacillus subtilis 168
.
Microbiology
144
:
73
82
.

Bochtler
M
Odintsov
SG
Marcyjaniak
M
Sabala
I
(
2004
)
Similar active sites in lysostaphins and d-Ala-d-Ala metallopeptidases
.
Protein Sci
13
:
854
861
.

Boland
FM
Atrih
A
Chirakkal
H
Foster
SJ
Moir
A
(
2000
)
Complete spore-cortex hydrolysis during germination of Bacillus subtilis 168 requires SleB and YpeB
.
Microbiology
146
:
57
64
.

Bourgerie
S
Karamanos
Y
Grard
T
Julien
R
(
1994
)
Purification and characterization of an endo-N-acetyl-beta-d-glucosaminidase from the culture medium of Stigmatella aurantiaca DW4
.
J Bacteriol
176
:
6170
6174
.

Brunskill
EW
Bayles
KW
(
1996
)
Identification of LytSR-regulated genes from Staphylococcus aureus
.
J Bacteriol
178
:
5810
5812
.

Bussiere
DE
Pratt
SD
Katz
L
Severin
JM
Holzman
T
Park
CH
(
1998
)
The structure of VanX reveals a novel amino-dipeptidase involved in mediating transposon-based vancomycin resistance
.
Mol Cell
2
:
75
84
.

Calamita
HG
Ehringer
WD
Koch
AL
Doyle
RJ
(
2001
)
Evidence that the cell wall of Bacillus subtilis is protonated during respiration
.
Proc Natl Acad Sci USA
98
:
15260
15263
.

Carballido-Lopez
R
Errington
J
(
2003
)
A dynamic bacterial cytoskeleton
.
Trends Cell Biol
13
:
577
583
.

Carballido-Lopez
R
Formstone
A
Li
Y
Ehrlich
SD
Noirot
P
Errington
J
(
2006
)
Actin homolog MreBH governs cell morphogenesis by localization of the cell wall hydrolase LytE
.
Dev Cell
11
:
399
409
.

Catt
DM
Gregory
RL
(
2005
)
Streptococcus mutans murein hydrolase
.
J Bacteriol
187
:
7863
7865
.

Charlier
P
Wery
JP
Dideberg
O
Frere
J-M
(
2003
)
Streptomyces albus G d-Ala-d-Ala carboxypeptidase
.
Handbook of Metalloproteins
, Vol.
3
(
Messerschmidt
A
Bode
W
&
Cygler
M
, eds), pp.
164
175
.
Wiley
, Chichester.

Chastanet
A
Losick
R
(
2007
)
Engulfment during sporulation in Bacillus subtilis is governed by a multi-protein complex containing tandemly acting autolysins
.
Mol Microbiol
64
:
139
152
.

Cheng
X
Zhang
X
Pflugrath
JW
Studier
FW
(
1994
)
The structure of bacteriophage T7 lysozyme, a zinc amidase and an inhibitor of T7 RNA polymerase
.
Proc Natl Acad Sci USA
91
:
4034
4038
.

Cheung
HY
Freese
E
(
1985
)
Monovalent cations enable cell wall turnover of the turnover-deficient lyt-15 mutant of Bacillus subtilis
.
J Bacteriol
161
:
1222
1225
.

Cheung
HY
Vitkovic
L
Freese
E
(
1983
)
Rates of peptidoglycan turnover and cell growth of Bacillus subtilis are correlated
.
J Bacteriol
156
:
1099
1106
.

Chirakkal
H
O'Rourke
M
Atrih
A
Foster
SJ
Moir
A
(
2002
)
Analysis of spore cortex lytic enzymes and related proteins in Bacillus subtilis endospore germination
.
Microbiology
148
:
2383
2392
.

Clarke
AJ
Dupont
C
(
1992
)
O-acetylated peptidoglycan: its occurrence, pathobiological significance, and biosynthesis
.
Can J Microbiol
38
:
85
91
.

Clausen
VA
Bae
W
Throup
J
Burnham
MK
Rosenberg
M
Wallis
NG
(
2003
)
Biochemical characterization of the first essential two-component signal transduction system from Staphylococcus aureus and Streptococcus pneumoniae
.
J Mol Microbiol Biotechnol
5
:
252
260
.

Claverys
JP
Havarstein
LS
(
2007
)
Cannibalism and fratricide: mechanisms and raisons d'etre
.
Nat Rev Microbiol
5
:
219
229
.

Cleveland
RF
Wicken
AJ
Daneo-Moore
L
Shockman
GD
(
1976
)
Inhibition of wall autolysis in Streptococcus faecalis by lipoteichoic acid and lipids
.
J Bacteriol
126
:
192
197
.

Costerton
JW
Cheng
KJ
Geesey
GG
Ladd
TI
Nickel
JC
Dasgupta
M
Marrie
TJ
(
1987
)
Bacterial biofilms in nature and disease
.
Annu Rev Microbiol
41
:
435
464
.

Courtin
P
Miranda
G
Guillot
A
et al. (
2006
)
Peptidoglycan structure analysis of Lactococcus lactis reveals the presence of an l,d-carboxypeptidase involved in peptidoglycan maturation
.
J Bacteriol
188
:
5293
5298
.

Dagkessamanskaia
A
Moscoso
M
Henard
V
et al. (
2004
)
Interconnection of competence, stress and CiaR regulons in Streptococcus pneumoniae: competence triggers stationary phase autolysis of ciaR mutant cells
.
Mol Microbiol
51
:
1071
1086
.

Daniel
RA
Errington
J
(
2003
)
Control of cell morphogenesis in bacteria: two distinct ways to make a rod-shaped cell
.
Cell
113
:
767
776
.

Davies
C
White
SW
Nicholas
RA
(
2001
)
Crystal structure of a deacylation-defective mutant of penicillin-binding protein 5 at 2.3-A resolution
.
J Biol Chem
276
:
616
623
.

De Boer
WR
Kruyssen
FJ
Wouters
JT
(
1981
)
Cell wall turnover in batch and chemostat cultures of Bacillus subtilis
.
J Bacteriol
145
:
50
60
.

De Boer
WR
Meyer
PD
Jordens
CG
Kruyssen
FJ
Wouters
JT
(
1982
)
Cell wall turnover in growing and nongrowing cultures of Bacillus subtilis
.
J Bacteriol
149
:
977
984
.

De Las Rivas
B
Garcia
JL
Lopez
R
Garcia
P
(
2002
)
Purification and polar localization of pneumococcal LytB, a putative endo-beta-N-acetylglucosaminidase: the chain-dispersing murein hydrolase
.
J Bacteriol
184
:
4988
5000
.

Demchick
P
Koch
AL
(
1996
)
The permeability of the wall fabric of Escherichia coli and Bacillus subtilis
.
J Bacteriol
178
:
768
773
.

Denome
SA
Elf
PK
Henderson
TA
Nelson
DE
Young
KD
(
1999
)
Escherichia coli mutants lacking all possible combinations of eight penicillin binding proteins: viability, characteristics, and implications for peptidoglycan synthesis
.
J Bacteriol
181
:
3981
3993
.

Desvaux
M
Dumas
E
Chafsey
I
Hebraud
M
(
2006
)
Protein cell surface display in gram-positive bacteria: from single protein to macromolecular protein structure
.
FEMS Microbiol Lett
256
:
1
15
.

Dhalluin
A
Bourgeois
I
Pestel-Caron
M
et al. (
2005
)
Acd, a peptidoglycan hydrolase of Clostridium difficile with N-acetylglucosaminidase activity
.
Microbiology
151
:
2343
2351
.

Downing
KJ
Mischenko
VV
Shleeva
MO
et al. (
2005
)
Mutants of Mycobacterium tuberculosis lacking three of the five rpf-like genes are defective for growth in vivo and for resuscitation in vitro
.
Infect Immun
73
:
3038
3043
.

Draft
MJ
Burnham
JC
Yamamoto
Y
(
1985
)
Lysis of Phormidium iridum by Myxococcus fulvus in continuous flow cultures
.
J Appl Bacteriol
59
:
73
80
.

Dubrac
S
Msadek
T
(
2004
)
Identification of genes controlled by the essential YycG/YycF two-component system of Staphylococcus aureus
.
J Bacteriol
186
:
1175
1181
.

Dziarski
R
(
2003
)
Recognition of bacterial peptidoglycan by the innate immune system
.
Cell Mol Life Sci
60
:
1793
1804
.

Dziarski
R
Gupta
D
(
2005
)
Peptidoglycan recognition in innate immunity
.
J Endotoxin Res
11
:
304
310
.

Dziarski
R
Gupta
D
(
2006
)
The peptidoglycan recognition proteins (PGRPs)
.
Genome Biol
7
:
232
.

Eckert
C
Lecerf
M
Dubost
L
Arthur
M
Mesnage
S
(
2006
)
Functional analysis of AtlA, the major N-acetylglucosaminidase of Enterococcus faecalis
.
J Bacteriol
188
:
8513
8519
.

Ehlert
K
Höltje
J-V
Templin
MF
(
1995
)
Cloning and expression of a murein hydrolase lipoprotein from Escherichia coli
.
Mol Microbiol
16
:
761
768
.

Ellermeier
CD
Hobbs
EC
Gonzalez-Pastor
JE
Losick
R
(
2006
)
A three-protein signaling pathway governing immunity to a bacterial cannibalism toxin
.
Cell
124
:
549
559
.

Engel
H
Van Leeuwen
A
Dijkstra
A
Keck
W
(
1992
)
Enzymatic preparation of 1,6-anhydro-muropeptides by immobilized murein hydrolases from Escherichia coli fused to staphylococcal protein A
.
Appl Microbiol Biotechnol
37
:
772
783
.

Errington
J
(
2003
)
Regulation of endospore formation in Bacillus subtilis
.
Nat Rev Microbiol
1
:
117
126
.

Fabret
C
Hoch
JA
(
1998
)
A two-component signal transduction system essential for growth of Bacillus subtilis: implications for anti-infective therapy
.
J Bacteriol
180
:
6375
6383
.

Fernandez-Tornero
C
Lopez
R
Garcia
E
Gimenez-Gallego
G
Romero
A
(
2001
)
A novel solenoid fold in the cell wall anchoring domain of the pneumococcal virulence factor LytA
.
Nat Struct Biol
8
:
1020
1024
.

Fernandez-Tornero
C
Garcia
E
Lopez
R
Gimenez-Gallego
G
Romero
A
(
2002
)
Two new crystal forms of the choline-binding domain of the major pneumococcal autolysin: insights into the dynamics of the active homodimer
.
J Mol Biol
321
:
163
173
.

Firczuk
M
Bochtler
M
(
2007
)
Mutational analysis of peptidoglycan amidase MepA
.
Biochemistry
46
:
120
128
.

Firczuk
M
Mucha
A
Bochtler
M
(
2005
)
Crystal structures of active LytM
.
J Mol Biol
354
:
578
590
.

Fischer
W
Rosel
P
Koch
HU
(
1981
)
Effect of alanine ester substitution and other structural features of lipoteichoic acids on their inhibitory activity against autolysins of Staphylococcus aureus
.
J Bacteriol
146
:
467
475
.

Fonze
E
Vermeire
M
Nguyen-Disteche
M
Brasseur
R
Charlier
P
(
1999
)
The crystal structure of a penicilloyl-serine transferase of intermediate penicillin sensitivity. The DD-transpeptidase of Streptomyces K15
.
J Biol Chem
274
:
21853
21860
.

Foster
SJ
(
1992
)
Analysis of the autolysins of Bacillus subtilis 168 during vegetative growth and differentiation by using renaturing polyacrylamide gel electrophoresis
.
J Bacteriol
174
:
464
470
.

Foster
SJ
(
1995
)
Molecular characterization and functional analysis of the major autolysin of Staphylococcus aureus 8325/4
.
J Bacteriol
177
:
5723
5725
.

Foster
SJ
Johnstone
K
(
1988
)
Germination-specific cortex-lytic enzyme is activated during triggering of Bacillus megaterium KM spore germination
.
Mol Microbiol
2
:
727
733
.

Foster
SJ
Johnstone
K
(
1990
)
Pulling the trigger: the mechanism of bacterial spore germination
.
Mol Microbiol
4
:
137
141
.

Foster
SJ
Popham
DL
(
2002
)
Structure and synthesis of cell wall, spore cortex, teichoic acids, S-layers and capsules
.
Bacillus subtilis and its Closest Relatives: from Genes to Cells
(
Sonenshein
AL
Hoch
JA
&
Losick
R
, eds), pp.
21
41
.
ASM Press
, Washington, DC.

Fukushima
T
Yamamoto
H
Atrih
A
Foster
SJ
Sekiguchi
J
(
2002
)
A polysaccharide deacetylase gene (pdaA) is required for germination and for production of muramic delta-lactam residues in the spore cortex of Bacillus subtilis
.
J Bacteriol
184
:
6007
6015
.

Fukushima
T
Afkham
A
Kurosawa
S
Tanabe
T
Yamamoto
H
Sekiguchi
J
(
2006
)
A new d,l-endopeptidase gene product, YojL (renamed CwlS), plays a role in cell separation with LytE and LytF in Bacillus subtilis
.
J Bacteriol
188
:
5541
5550
.

Fukushima
T
Yao
Y
Kitajima
T
Yamamoto
H
Sekiguchi
J
(
2007
)
Characterization of new l,d-endopeptidase gene product CwlK (previous YcdD) that hydrolyzes peptidoglycan in Bacillus subtilis
.
Mol Genet Genomics
278
:
371
383
.

Gallant
CV
Daniels
C
Leung
JM
Ghosh
AS
Young
KD
Kotra
LP
Burrows
LL
(
2005
)
Common beta-lactamases inhibit bacterial biofilm formation
.
Mol Microbiol
58
:
1012
1024
.

Garcia
DL
Dillard
JP
(
2006
)
AmiC functions as an N-acetylmuramyl-l-alanine amidase necessary for cell separation and can promote autolysis in Neisseria gonorrhoeae
.
J Bacteriol
188
:
7211
7221
.

Garcia
P
Paz Gonzalez
M
Garcia
E
Garcia
JL
Lopez
R
(
1999a
)
The molecular characterization of the first autolytic lysozyme of Streptococcus pneumoniae reveals evolutionary mobile domains
.
Mol Microbiol
33
:
128
138
.

Garcia
P
Gonzalez
MP
Garcia
E
Lopez
R
Garcia
JL
(
1999b
)
LytB, a novel pneumococcal murein hydrolase essential for cell separation
.
Mol Microbiol
31
:
1275
1277
.

Ghuysen
JM
(
1968
)
Use of bacteriolytic enzymes in determination of wall structure and their role in cell metabolism
.
Bacteriol Rev
32
(suppl)
:
425
464
.

Gilbert
HJ
(
2007
)
Cellulosomes: microbial nanomachines that display plasticity in quaternary structure
.
Mol Microbiol
63
:
1568
1576
.

Gilmore
ME
Bandyopadhyay
D
Dean
AM
Linnstaedt
SD
Popham
DL
(
2004
)
Production of muramic δ-lactam in Bacillus subtilis spore peptidoglycan
.
J Bacteriol
186
:
80
89
.

Gittins
JR
Phoenix
DA
Pratt
JM
(
1994
)
Multiple mechanisms of membrane anchoring of Escherichia coli penicillin-binding proteins
.
FEMS Microbiol Rev
13
:
1
12
.

Goffin
C
Ghuysen
JM
(
1998
)
Multimodular penicillin-binding proteins: an enigmatic family of orthologs and paralogs
.
Microbiol Mol Biol Rev
62
:
1079
1093
.

Gonzalez-Pastor
JE
Hobbs
EC
Losick
R
(
2003
)
Cannibalism by sporulating bacteria
.
Science
301
:
510
513
.

Goodell
EW
(
1985
)
Recycling of murein by Escherichia coli
.
J Bacteriol
163
:
305
310
.

Goodell
EW
Schwarz
U
(
1983
)
Cleavage and resynthesis of peptide cross bridges in Escherichia coli murein
.
J Bacteriol
156
:
136
140
.

Goodell
EW
Schwarz
U
(
1985
)
Release of cell wall peptides into culture medium by exponentially growing Escherichia coli
.
J Bacteriol
162
:
391
397
.

Goodell
EW
Lopez
R
Tomasz
A
(
1976
)
Suppression of lytic effect of beta lactams on Escherichia coli and other bacteria
.
Proc Natl Acad Sci USA
73
:
3293
3297
.

Guinane
CM
Cotter
PD
Ross
RP
Hill
C
(
2006
)
Contribution of penicillin-binding protein homologs to antibiotic resistance, cell morphology, and virulence of Listeria monocytogenes EGDe
.
Antimicrob Agents Chemother
50
:
2824
2828
.

Guiral
S
Mitchell
TJ
Martin
B
Claverys
J-P
(
2005
)
Competence-programmed predation of noncompetent cells in the human pathogen Streptococcus pneumoniae: genetic requirements
.
Proc Natl Acad Sci USA
102
:
8710
8715
.

Harris
F
Brandenburg
K
Seydel
U
Phoenix
D
(
2002
)
Investigations into the mechanisms used by the C-terminal anchors of Escherichia coli penicillin-binding proteins 4, 5, 6 and 6b for membrane interaction
.
Eur J Biochem
269
:
5821
5829
.

Heidrich
C
Templin
MF
Ursinus
A
et al. (
2001
)
Involvement of N-acetylmuramyl-l-alanine amidases in cell separation and antibiotic-induced autolysis of Escherichia coli
.
Mol Microbiol
41
:
167
178
.

Heidrich
C
Ursinus
A
Berger
J
Schwarz
H
Höltje
J-V
(
2002
)
Effects of multiple deletions of murein hydrolases on viability, septum cleavage, and sensitivity to large toxic molecules in Escherichia coli
.
J Bacteriol
184
:
6093
6099
.

Heilmann
C
Hussain
M
Peters
G
Götz
F
(
1997
)
Evidence for autolysin-mediated primary attachment of Staphylococcus epidermidis to a polystyrene surface
.
Mol Microbiol
24
:
1013
1024
.

Heinrich
P
Rosenstein
R
Bohmer
M
Sonner
P
Götz
F
(
1987
)
The molecular organization of the lysostaphin gene and its sequences repeated in tandem
.
Mol Gen Genet
209
:
563
569
.

Henderson
TA
Templin
M
Young
KD
(
1995
)
Identification and cloning of the gene encoding penicillin-binding protein 7 of Escherichia coli
.
J Bacteriol
177
:
2074
2079
.

Herbold
DR
Glaser
L
(
1975a
)
Interaction of N-acetylmuramic acid l-alanine amidase with cell wall polymers
.
J Biol Chem
250
:
7231
7238
.

Herbold
DR
Glaser
L
(
1975b
)
Bacillus subtilis N-acetylmuramic acid l-alanine amidase
.
J Biol Chem
250
:
1676
1682
.

Heymer
B
Schmidt
WC
(
1975
)
Purification and characterization of a Streptomyces albus endo-N-acetylmuramidase lytic for group A and other beta haemolytic streptococci
.
Microbios
12
:
51
66
.

Höltje
J-V
(
1995
)
From growth to autolysis: the murein hydrolases in Escherichia coli
.
Arch Microbiol
164
:
243
254
.

Höltje
J-V
(
1996a
)
A hypothetical holoenzyme involved in the replication of the murein sacculus of Escherichia coli
.
Microbiology
142
:
1911
1918
.

Höltje
J-V
(
1996b
)
Bacterial lysozymes
.
Lysozymes: Model Enzymes in Biochemistry and Biology
(
Jolles
P
, ed), pp.
65
74
.
Birkhäuser Verlag
, Basel, Switzerland.

Höltje
J-V
(
1996c
)
Molecular interplay of murein synthases and murein hydrolases in Escherichia coli
.
Microb Drug Resist
2
:
99
103
.

Höltje
J-V
(
1998
)
Growth of the stress-bearing and shape-maintaining murein sacculus of Escherichia coli
.
Microbiol Mol Biol Rev
62
:
181
203
.

Höltje
J-V
Tomasz
A
(
1975
)
Lipoteichoic acid: a specific inhibitor of autolysin activity in pneumococcus
.
Proc Natl Acad Sci USA
72
:
1690
1694
.

Höltje
J-V
Tuomanen
EI
(
1991
)
The murein hydrolases of Escherichia coli: properties, functions and impact on the course of infections in vivo
.
J Gen Microbiol
137
:
441
454
.

Höltje
J-V
Mirelman
D
Sharon
N
Schwarz
U
(
1975
)
Novel type of murein transglycosylase in Escherichia coli
.
J Bacteriol
124
:
1067
1076
.

Höltje
J-V
Kopp
U
Ursinus
A
Wiedemann
B
(
1994
)
The negative regulator of beta-lactamase induction AmpD is a N-acetyl-anhydromuramyl-l-alanine amidase
.
FEMS Microbiol Lett
122
:
159
164
.

Horsburgh
GJ
Atrih
A
Foster
SJ
(
2003a
)
Characterization of LytH, a differentiation-associated peptidoglycan hydrolase of Bacillus subtilis involved in endospore cortex maturation
.
J Bacteriol
185
:
3813
3820
.

Horsburgh
GJ
Atrih
A
Williamson
MP
Foster
SJ
(
2003b
)
LytG of Bacillus subtilis is a novel peptidoglycan hydrolase: the major active glucosaminidase
.
Biochemistry
42
:
257
264
.

Hourdou
ML
Duez
C
Joris
B
Vacheron
MJ
Guinand
M
Michel
G
Ghuysen
JM
(
1992
)
Cloning and nucleotide sequence of the gene encoding the gamma-d-glutamyl-l-diamino acid endopeptidase II of Bacillus sphaericus
.
FEMS Microbiol Lett
70
:
165
170
.

Hourdou
ML
Guinand
M
Vacheron
MJ
et al. (
1993
)
Characterization of the sporulation-related gamma-d-glutamyl-(l)meso-diaminopimelic-acid-hydrolysing peptidase I of Bacillus sphaericus NCTC 9602 as a member of the metallo(zinc) carboxypeptidase A family. Modular design of the protein
.
Biochem J
292
:
563
570
.

Howell
A
Dubrac
S
Andersen
KK
Noone
D
Fert
J
Msadek
T
Devine
K
(
2003
)
Genes controlled by the essential YycG/YycF two-component system of Bacillus subtilis revealed through a novel hybrid regulator approach
.
Mol Microbiol
49
:
1639
1655
.

Huard
C
Miranda
G
Wessner
F
Bolotin
A
Hansen
J
Foster
SJ
Chapot-Chartier
MP
(
2003
)
Characterization of AcmB, an N-acetylglucosaminidase autolysin from Lactococcus lactis
.
Microbiology
149
:
695
705
.

Huard
C
Miranda
G
Redko
Y
Wessner
F
Foster
SJ
Chapot-Chartier
MP
(
2004
)
Analysis of the peptidoglycan hydrolase complement of Lactococcus lactis: identification of a third N-acetylglucosaminidase, AcmC
.
Appl Environ Microbiol
70
:
3493
3499
.

Hughes
RC
Tanner
PJ
Stokes
E
(
1970
)
Cell-wall thickening in Bacillus subtilis. Comparison of thickened and normal walls
.
Biochem J
120
:
159
170
.

Ishikawa
S
Yamane
K
Sekiguchi
J
(
1998
)
Regulation and characterization of a newly deduced cell wall hydrolase gene (cwlJ) which affects germination of Bacillus subtilis spores
.
J Bacteriol
180
:
1375
1380
.

Jacobs
C
Joris
B
Jamin
M
et al. (
1995
)
AmpD, essential for both beta-lactamase regulation and cell wall recycling, is a novel cytosolic N-acetylmuramyl-l-alanine amidase
.
Mol Microbiol
15
:
553
559
.

Jacobs
C
Frere
JM
Normark
S
(
1997
)
Cytosolic intermediates for cell wall biosynthesis and degradation control inducible beta-lactam resistance in gram-negative bacteria
.
Cell
88
:
823
832
.

Jolles
P
(
1996
)
Lysozymes: Model Enzymes in Biochemistry and Biology
.
Birkhäuser
, Basel.

Jolliffe
LK
Doyle
RJ
Streips
UN
(
1981
)
The energized membrane and cellular autolysis in Bacillus subtilis
.
Cell
25
:
753
763
.

Jones
LJ
Carballido-Lopez
R
Errington
J
(
2001
)
Control of cell shape in bacteria: helical, actin-like filaments in Bacillus subtilis
.
Cell
104
:
913
922
.

Joris
B
Englebert
S
Chu
CP
Kariyama
R
Daneo-Moore
L
Shockman
GD
Ghuysen
JM
(
1992
)
Modular design of the Enterococcus hirae muramidase-2 and Streptococcus faecalis autolysin
.
FEMS Microbiol Lett
70
:
257
264
.

Kaiser
D
(
2003
)
Coupling cell movement to multicellular development in myxobacteria
.
Nat Rev Microbiol
1
:
45
54
.

Kaiser
D
(
2004
)
Signaling in myxobacteria
.
Annu Rev Microbiol
58
:
75
98
.

Kajimura
J
Fujiwara
T
Yamada
S
et al. (
2005
)
Identification and molecular characterization of an N-acetylmuramyl-l-alanine amidase Sle1 involved in cell separation of Staphylococcus aureus
.
Mol Microbiol
58
:
1087
1101
.

Karamanos
Y
(
1997
)
Endo-N-acetyl-beta-d-glucosaminidases and their potential substrates: structure/function relationships
.
Res Microbiol
148
:
661
671
.

Kawamura
T
Shockman
GD
(
1983
)
Purification and some properties of the endogenous, autolytic N-acetylmuramoylhydrolase of Streptococcus faecium, a bacterial glycoenzyme
.
J Biol Chem
258
:
9514
9521
.

Keck
W
Schwarz
U
(
1979
)
Escherichia coli murein-dd-endopeptidase insensitive to beta-lactam antibiotics
.
J Bacteriol
139
:
770
774
.

Keck
W
Van Leeuwen
AM
Huber
M
Goodell
EW
(
1990
)
Cloning and characterization of mepA, the structural gene of the penicillin-insensitive murein endopeptidase from Escherichia coli
.
Mol Microbiol
4
:
209
219
.

Keep
NH
Ward
JM
Cohen-Gonsaud
M
Henderson
B
(
2006
)
Wake up! Peptidoglycan lysis and bacterial non-growth states
.
Trends Microbiol
14
:
271
276
.

Kishida
H
Unzai
S
Roper
DI
Lloyd
A
Park
SY
Tame
JR
(
2006
)
Crystal structure of penicillin binding protein 4 (dacB) from Escherichia coli, both in the native form and covalently linked to various antibiotics
.
Biochemistry
45
:
783
792
.

Koch
AL
(
1990
)
Additional arguments for the key role of “smart” autolysins in the enlargement of the wall of gram-negative bacteria
.
Res Microbiol
141
:
529
541
.

Koch
AL
Doyle
RJ
(
1985
)
Inside-to-outside growth and turnover of the wall of gram-positive rods
.
J Theor Biol
117
:
137
157
.

Koch
AL
Kirchner
G
Doyle
RJ
Burdett
ID
(
1985
)
How does a Bacillus split its septum right down the middle?
Ann Inst Pasteur Microbiol
136A
:
91
98
.

Komatsuzawa
H
Sugai
M
Nakashima
S
Yamada
S
Matsumoto
A
Oshida
T
Suginaka
H
(
1997
)
Subcellular localization of the major autolysin, ATL and its processed proteins in Staphylococcus aureus
.
Microbiol Immunol
41
:
469
479
.

Koraimann
G
(
2003
)
Lytic transglycosylases in macromolecular transport systems of gram-negative bacteria
.
Cell Mol Life Sci
60
:
2371
2388
.

Korndorfer
IP
Danzer
J
Schmelcher
M
Zimmer
M
Skerra
A
Loessner
MJ
(
2006
)
The crystal structure of the bacteriophage PSA endolysin reveals a unique fold responsible for specific recognition of Listeria cell walls
.
J Mol Biol
364
:
678
689
.

Korsak
D
Vollmer
W
Markiewicz
Z
(
2005a
)
Listeria monocytogenes EGD lacking penicillin-binding protein 5 (PBP5) produces a thicker cell wall
.
FEMS Microbiol Lett
251
:
281
288
.

Korsak
D
Liebscher
S
Vollmer
W
(
2005b
)
Susceptibility to antibiotics and beta-lactamase induction in murein hydrolase mutants of Escherichia coli
.
Antimicrob Agents Chemother
49
:
1404
1409
.

Korza
HJ
Bochtler
M
(
2005
)
Pseudomonas aeruginosald-carboxypeptidase, a serine peptidase with a Ser-His-Glu triad and a nucleophilic elbow
.
J Biol Chem
280
:
40802
40812
.

Kraft
AR
Templin
MF
Höltje
J-V
(
1998
)
Membrane-bound lytic endotransglycosylase in Escherichia coli
.
J Bacteriol
180
:
3441
3447
.

Kraft
AR
Prabhu
J
Ursinus
A
Höltje
J-V
(
1999
)
Interference with murein turnover has no effect on growth but reduces beta-lactamase induction in Escherichia coli
.
J Bacteriol
181
:
7192
7198
.

Kuroda
A
Sekiguchi
J
(
1993
)
High-level transcription of the major Bacillus subtilis autolysin operon depends on expression of the sigma D gene and is affected by a sin (flaD) mutation
.
J Bacteriol
175
:
795
801
.

Kuroda
A
Asami
Y
Sekiguchi
J
(
1993
)
Molecular cloning of a sporulation-specific cell wall hydrolase gene of Bacillus subtilis
.
J Bacteriol
175
:
6260
6268
.

Lawrence
PJ
Strominger
JL
(
1970
)
Biosynthesis of the peptidoglycan of bacterial cell walls XVI. The reversible fixation of radioactive penicillin G to the d-alanine carboxypeptidase of Bacillus subtilis
.
J Biol Chem
245
:
3660
3666
.

Lazarevic
V
Margot
P
Soldo
B
Karamata
D
(
1992
)
Sequencing and analysis of the Bacillus subtilis lytRABC divergon: a regulatory unit encompassing the structural genes of the N-acetylmuramoyl-l-alanine amidase and its modifier
.
J Gen Microbiol
138
(Part 9)
:
1949
1961
.

Leung
AK
Duewel
HS
Honek
JF
Berghuis
AM
(
2001
)
Crystal structure of the lytic transglycosylase from bacteriophage lambda in complex with hexa-N-acetylchitohexaose
.
Biochemistry
40
:
5665
5673
.

Lewis
K
(
2000
)
Programmed death in bacteria
.
Microbiol Mol Biol Rev
64
:
503
514
.

Liepinsh
E
Genereux
C
Dehareng
D
Joris
B
Otting
G
(
2003
)
NMR structure of Citrobacter freundii AmpD, comparison with bacteriophage T7 lysozyme and homology with PGRP domains
.
J Mol Biol
327
:
833
842
.

Loessner
MJ
(
2005
)
Bacteriophage endolysins – current state of research and applications
.
Curr Opin Microbiol
8
:
480
487
.

Loessner
MJ
Wendlinger
G
Scherer
S
(
1995
)
Heterogeneous endolysins in Listeria monocytogenes bacteriophages: a new class of enzymes and evidence for conserved holin genes within the siphoviral lysis cassettes
.
Mol Microbiol
16
:
1231
1241
.

Lommatzsch
J
Templin
MF
Kraft
AR
Vollmer
W
Höltje
J-V
(
1997
)
Outer membrane localization of murein hydrolases: MltA, a third lipoprotein lytic transglycosylase in Escherichia coli
.
J Bacteriol
179
:
5465
5470
.

Lopez
R
Garcia
E
Ronda
C
(
1981
)
Bacteriophages of Streptococcus pneumoniae
.
Rev Infect Dis
3
:
212
223
.

Low
LY
Yang
C
Perego
M
Osterman
A
Liddington
RC
(
2005
)
Structure and lytic activity of a Bacillus anthracis prophage endolysin
.
J Biol Chem
280
:
35433
35439
.

Lu
JZ
Fujiwara
T
Komatsuzawa
H
Sugai
M
Sakon
J
(
2006
)
Cell wall-targeting domain of glycylglycine endopeptidase distinguishes among peptidoglycan cross-bridges
.
J Biol Chem
281
:
549
558
.

Marcyjaniak
M
Odintsov
SG
Sabala
I
Bochtler
M
(
2004
)
Peptidoglycan amidase MepA is a LAS metallopeptidase
.
J Biol Chem
279
:
43982
43989
.

Margot
P
Karamata
D
(
1992
)
Identification of the structural genes for N-acetylmuramoyl-l-alanine amidase and its modifier in Bacillus subtilis 168: inactivation of these genes by insertional mutagenesis has no effect on growth or cell separation
.
Mol Gen Genet
232
:
359
366
.

Margot
P
Mauel
C
Karamata
D
(
1994
)
The gene of the N-acetylglucosaminidase, a Bacillus subtilis 168 cell wall hydrolase not involved in vegetative cell autolysis
.
Mol Microbiol
12
:
535
545
.

Margot
P
Pagni
M
Karamata
D
(
1999
)
Bacillus subtilis 168 gene lytF encodes a gamma-d-glutamate-meso-diaminopimelate muropeptidase expressed by the alternative vegetative sigma factor, sigmaD
.
Microbiology
145
:
57
65
.

Markiewicz
Z
Glauner
B
Schwarz
U
(
1983
)
Murein structure and lack of dd- and ld-carboxypeptidase activities in Caulobacter crescentus
.
J Bacteriol
156
:
649
655
.

Martin
PK
Li
T
Sun
D
Biek
DP
Schmid
MB
(
1999
)
Role in cell permeability of an essential two-component system in Staphylococcus aureus
.
J Bacteriol
181
:
3666
3673
.

Meador-Parton
J
Popham
DL
(
2000
)
Structural analysis of Bacillus subtilis spore peptidoglycan during sporulation
.
J Bacteriol
182
:
4491
4499
.

Meberg
BM
Paulson
AL
Priyadarshini
R
Young
KD
(
2004
)
Endopeptidase penicillin-binding proteins 4 and 7 play auxiliary roles in determining uniform morphology of Escherichia coli
.
J Bacteriol
186
:
8326
8336
.

Meisel
U
Höltje
J-V
Vollmer
W
(
2003
)
Overproduction of inactive variants of the murein synthase PBP1B causes lysis in Escherichia coli
.
J Bacteriol
185
:
5342
5348
.

Mengin-Lecreulx
D
Lemaitre
B
(
2005
)
Structure and metabolism of peptidoglycan and molecular requirements allowing its detection by the Drosophila innate immune system
.
J Endotoxin Res
11
:
105
111
.

Mengin-Lecreulx
D
Van Heijenoort
J
Park
JT
(
1996
)
Identification of the mpl gene encoding UDP-N-acetylmuramate: l-alanyl-gamma-d-glutamyl-meso-diaminopimelate ligase in Escherichia coli and its role in recycling of cell wall peptidoglycan
.
J Bacteriol
178
:
5347
5352
.

Mercier
C
Durrieu
C
Briandet
R
Domakova
E
Tremblay
J
Buist
G
Kulakauskas
S
(
2002
)
Positive role of peptidoglycan breaks in lactococcal biofilm formation
.
Mol Microbiol
46
:
235
243
.

Milohanic
E
Jonquieres
R
Cossart
P
Berche
P
Gaillard
JL
(
2001
)
The autolysin Ami contributes to the adhesion of Listeria monocytogenes to eukaryotic cells via its cell wall anchor
.
Mol Microbiol
39
:
1212
1224
.

Monterroso
B
Lopez-Zumel
C
Garcia
JL
Saiz
JL
Garcia
P
Campillo
NE
Menendez
M
(
2005
)
Unravelling the structure of the pneumococcal autolytic lysozyme
.
Biochem J
391
:
41
49
.

Moriyama
R
Fukuoka
H
Miyata
S
et al. (
1999
)
Expression of a germination-specific amidase, SleB, of Bacilli in the forespore compartment of sporulating cells and its localization on the exterior side of the cortex in dormant spores
.
J Bacteriol
181
:
2373
2378
.

Morlot
C
Noirclerc-Savoye
M
Zapun
A
Dideberg
O
Vernet
T
(
2004
)
The d,d-carboxypeptidase PBP3 organizes the division process of Streptococcus pneumoniae
.
Mol Microbiol
51
:
1641
1648
.

Morlot
C
Pernot
L
Le Gouellec
A
Di Guilmi
AM
Vernet
T
Dideberg
O
Dessen
A
(
2005
)
Crystal structure of a peptidoglycan synthesis regulatory factor (PBP3) from Streptococcus pneumoniae
.
J Biol Chem
280
:
15984
15991
.

Moscoso
M
Claverys
JP
(
2004
)
Release of DNA into the medium by competent Streptococcus pneumoniae: kinetics, mechanism and stability of the liberated DNA
.
Mol Microbiol
54
:
783
794
.

Mukamolova
GV
Kaprelyants
AS
Young
DI
Young
M
Kell
DB
(
1998
)
A bacterial cytokine
.
Proc Natl Acad Sci USA
95
:
8916
8921
.

Mukamolova
GV
Murzin
AG
Salina
EG
Demina
GR
Kell
DB
Kaprelyants
AS
Young
M
(
2006
)
Muralytic activity of Micrococcus luteus Rpf and its relationship to physiological activity in promoting bacterial growth and resuscitation
.
Mol Microbiol
59
:
84
98
.

Mulder
L
Lefebvre
B
Cullimore
J
Imberty
A
(
2006
)
LysM domains of Medicago truncatula NFP protein involved in Nod factor perception. Glycosylation state, molecular modeling and docking of chitooligosaccharides and Nod factors
.
Glycobiology
16
:
801
809
.

Nambu
T
Minamino
T
Macnab
RM
Kutsukake
K
(
1999
)
Peptidoglycan-hydrolyzing activity of the FlgJ protein, essential for flagellar rod formation in Salmonella typhimurium
.
J Bacteriol
181
:
1555
1561
.

Nelson
DE
Young
KD
(
2001
)
Contributions of PBP 5 and dd-carboxypeptidase penicillin binding proteins to maintenance of cell shape in Escherichia coli
.
J Bacteriol
183
:
3055
3064
.

Ng
WL
Robertson
GT
Kazmierczak
KM
Zhao
J
Gilmour
R
Winkler
ME
(
2003
)
Constitutive expression of PcsB suppresses the requirement for the essential VicR (YycF) response regulator in Streptococcus pneumoniae R6
.
Mol Microbiol
50
:
1647
1663
.

Ng
WL
Kazmierczak
KM
Winkler
ME
(
2004
)
Defective cell wall synthesis in Streptococcus pneumoniae R6 depleted for the essential PcsB putative murein hydrolase or the VicR (YycF) response regulator
.
Mol Microbiol
53
:
1161
1175
.

Nicholas
RA
Krings
S
Tomberg
J
Nicola
G
Davies
C
(
2003
)
Crystal structure of wild-type penicillin-binding protein 5 from Escherichia coli: implications for deacylation of the acyl-enzyme complex
.
J Biol Chem
278
:
52826
52833
.

Nugroho
FA
Yamamoto
H
Kobayashi
Y
Sekiguchi
J
(
1999
)
Characterization of a new sigma-K-dependent peptidoglycan hydrolase gene that plays a role in Bacillus subtilis mother cell lysis
.
J Bacteriol
181
:
6230
6237
.

Odintsov
SG
Sabala
I
Marcyjaniak
M
Bochtler
M
(
2004
)
Latent LytM at 1.3A resolution
.
J Mol Biol
335
:
775
785
.

Ohnishi
R
Ishikawa
S
Sekiguchi
J
(
1999
)
Peptidoglycan hydrolase LytF plays a role in cell separation with CwlF during vegetative growth of Bacillus subtilis
.
J Bacteriol
181
:
3178
3184
.

Oliver
JD
(
2005
)
The viable but nonculturable state in bacteria
.
J Microbiol
43
:
93
100
.

Oshida
T
Sugai
M
Komatsuzawa
H
Hong
YM
Suginaka
H
Tomasz
A
(
1995
)
A Staphylococcus aureus autolysin that has an N-acetylmuramoyl-l-alanine amidase domain and an endo-beta-N-acetylglucosaminidase domain: cloning, sequence analysis, and characterization
.
Proc Natl Acad Sci USA
92
:
285
289
.

Paidhungat
M
Ragkousi
K
Setlow
P
(
2001
)
Genetic requirements for induction of germination of spores of Bacillus subtilis by Ca(2+)-dipicolinate
.
J Bacteriol
183
:
4886
4893
.

Park
JT
(
1995
)
Why does Escherichia coli recycle its cell wall peptides?
Mol Microbiol
17
:
421
426
.

Park
JT
(
1996
)
The convergence of murein recycling research with beta-lactamase research
.
Microb Drug Resist
2
:
105
112
.

Patton
TG
Yang
SJ
Bayles
KW
(
2006
)
The role of proton motive force in expression of the Staphylococcus aureus cid and lrg operons
.
Mol Microbiol
59
:
1395
1404
.

Pei
J
Grishin
NV
(
2005
)
COG3926 and COG5526: a tale of two new lysozyme-like protein families
.
Protein Sci
14
:
2574
2581
.

Piggot
PJ
Hilbert
DW
(
2004
)
Sporulation of Bacillus subtilis
.
Curr Opin Microbiol
7
:
579
586
.

Ponting
CP
Aravind
L
Schultz
J
Bork
P
Koonin
EV
(
1999
)
Eukaryotic signalling domain homologues in archaea and bacteria. Ancient ancestry and horizontal gene transfer
.
J Mol Biol
289
:
729
745
.

Pooley
HM
(
1976
)
Turnover and spreading of old wall during surface growth of Bacillus subtilis
.
J Bacteriol
125
:
1127
1138
.

Pooley
HM
Shockman
GD
(
1970
)
Relationship between the location of autolysin, cell wall synthesis, and the development of resistance to cellular autolysis in Streptococcus faecalis after inhibition of protein synthesis
.
J Bacteriol
103
:
457
466
.

Popham
DL
Helin
J
Costello
CE
Setlow
P
(
1996
)
Muramic lactam in peptidoglycan of Bacillus subtilis spores is required for spore outgrowth but not for spore dehydration or heat resistance
.
Proc Natl Acad Sci USA
93
:
15405
15410
.

Popham
DL
Gilmore
ME
Setlow
P
(
1999
)
Roles of low-molecular-weight penicillin-binding proteins in Bacillus subtilis spore peptidoglycan synthesis and spore properties
.
J Bacteriol
181
:
126
132
.

Powell
AJ
Liu
ZJ
Nicholas
RA
Davies
C
(
2006
)
Crystal structures of the lytic transglycosylase MltA from N. gonorrhoeae and E. coli: insights into interdomain movements and substrate binding
.
J Mol Biol
359
:
122
136
.

Priyadarshini
R
Popham
DL
Young
KD
(
2006
)
Daughter cell separation by penicillin-binding proteins and peptidoglycan amidases in Escherichia coli
.
J Bacteriol
188
:
5345
5355
.

Rashid
MH
Mori
M
Sekiguchi
J
(
1995
)
Glucosaminidase of Bacillus subtilis: cloning, regulation, primary structure and biochemical characterization
.
Microbiology
141
:
2391
2404
.

Rau
A
Hogg
T
Marquardt
R
Hilgenfeld
R
(
2001
)
A new lysozyme fold. Crystal structure of the muramidase from Streptomyces coelicolor at 1.65 A resolution
.
J Biol Chem
276
:
31994
31999
.

Reichenbach
H
(
1999
)
The ecology of the myxobacteria
.
Environ Microbiol
1
:
15
21
.

Reid
CW
Blackburn
NT
Legaree
BA
Auzanneau
FI
Clarke
AJ
(
2004
)
Inhibition of membrane-bound lytic transglycosylase B by NAG-thiazoline
.
FEBS Lett
574
:
73
79
.

Rice
KC
Bayles
KW
(
2003
)
Death's toolbox: examining the molecular components of bacterial programmed cell death
.
Mol Microbiol
50
:
729
738
.

Romeis
T
Höltje
J-V
(
1994a
)
Penicillin-binding protein 7/8 of Escherichia coli is a DD-endopeptidase
.
Eur J Biochem
224
:
597
604
.

Romeis
T
Höltje
J-V
(
1994b
)
Specific interaction of penicillin-binding proteins 3 and 7/8 with soluble lytic transglycosylase in Escherichia coli
.
J Biol Chem
269
:
21603
21607
.

Romeis
T
Vollmer
W
Höltje
J-V
(
1993
)
Characterization of three different lytic transglycosylases in Escherichia coli
.
FEMS Microbiol Lett
111
:
141
146
.

Rosenbluh
A
Rosenberg
E
(
1993
)
Developmental lysis and auticides
.
Myxobacteria II
(
Dworkin
M
&
Kaiser
D
, eds), pp.
213
233
.
American Society for Microbioloy
, Washington, DC.

Sauvage
E
Herman
R
Petrella
S
Duez
C
Bouillenne
F
Frere
JM
Charlier
P
(
2005
)
Crystal structure of the Actinomadura R39 DD-peptidase reveals new domains in penicillin-binding proteins
.
J Biol Chem
280
:
31249
31256
.

Sauvage
E
Duez
C
Herman
R
et al. (
2007
)
Crystal structure of the Bacillus subtilis penicillin-binding protein 4a, and its complex with a peptidoglycan mimetic peptide
.
J Mol Biol
371
:
528
539
.

Sauvage
E
Charlier
P
Terrak
M
Ayala
JA
(
2008
)
The penicillin-binding proteins: structure and role in peptidoglycan biosynthesis
.
FEMS Microbiol Rev
32
:
234
258
.

Scheurwater
E
Reid
CW
Clarke
AJ
(
2007
)
Lytic transglycosylases: bacterial space-making autolysins
.
Int J Biochem Cell Biol
, doi: DOI:

Schmidt
DM
Hubbard
BK
Gerlt
JA
(
2001
)
Evolution of enzymatic activities in the enolase superfamily: functional assignment of unknown proteins in Bacillus subtilis and Escherichia coli as l-Ala-d/l-Glu epimerases
.
Biochemistry
40
:
15707
15715
.

Schuster
C
Dobrinski
B
Hakenbeck
R
(
1990
)
Unusual septum formation in Streptococcus pneumoniae mutants with an alteration in the d,d-carboxypeptidase penicillin-binding protein 3
.
J Bacteriol
172
:
6499
6505
.

Scott
JR
Barnett
TC
(
2006
)
Surface proteins of gram-positive bacteria and how they get there
.
Annu Rev Microbiol
60
:
397
423
.

Sekiguchi
J
Akeo
K
Yamamoto
H
Khasanov
FK
Alonso
JC
Kuroda
A
(
1995
)
Nucleotide sequence and regulation of a new putative cell wall hydrolase gene, cwlD, which affects germination in Bacillus subtilis
.
J Bacteriol
177
:
5582
5589
.

Serizawa
M
Yamamoto
H
Yamaguchi
H
Fujita
Y
Kobayashi
K
Ogasawara
N
Sekiguchi
J
(
2004
)
Systematic analysis of SigD-regulated genes in Bacillus subtilis by DNA microarray and Northern blotting analyses
.
Gene
329
:
125
136
.

Seto
H
Tomasz
A
(
1975
)
Protoplast formation and leakage of intramembrane cell components: induction by the competence activator substance of pneumococci
.
J Bacteriol
121
:
344
353
.

Shockman
GD
(
1992
)
The autolytic (‘suicidase’) system of Enterococcus hirae: from lysine depletion autolysis to biochemical and molecular studies of the two muramidases of Enterococcus hirae ATCC 9790
.
FEMS Microbiol Lett
79
:
261
267
.

Shockman
GD
Höltje
J-V
(
1994
)
Microbial peptidoglycan (murein) hydrolases
.
Bacterial Cell Wall
(
Ghuysen
J-M
&
Hakenbeck
R
, eds), pp.
131
166
.
Elsevier
, Amsterdam.

Shockman
GD
Daneo-Moore
L
Kariyama
R
Massidda
O
(
1996
)
Bacterial walls, peptidoglycan hydrolases, autolysins, and autolysis
.
Microb Drug Resist
2
:
95
98
.

Singh
BN
(
1947
)
Myxobacteria in soils and compost; their distribution, number and lytic action on bacteria
.
J Gen Microbiol
1
:
1
10
.

Smith
TJ
Foster
SJ
(
1995
)
Characterization of the involvement of two compensatory autolysins in mother cell lysis during sporulation of Bacillus subtilis 168
.
J Bacteriol
177
:
3855
3862
.

Smith
TJ
Blackman
SA
Foster
SJ
(
2000
)
Autolysins of Bacillus subtilis: multiple enzymes with multiple functions
.
Microbiology
146
:
249
262
.

Sogaard-Andersen
L
(
2004
)
Cell polarity, intercellular signalling and morphogenetic cell movements in Myxococcus xanthus
.
Curr Opin Microbiol
7
:
587
593
.

Steen
A
Buist
G
Horsburgh
GJ
Venema
G
Kuipers
OP
Foster
SJ
Kok
J
(
2005
)
AcmA of Lactococcus lactis is an N-acetylglucosaminidase with an optimal number of LysM domains for proper functioning
.
Febs J
272
:
2854
2868
.

Stefanova
ME
Tomberg
J
Olesky
M
Holtje
JV
Gutheil
WG
Nicholas
RA
(
2003
)
Neisseria gonorrhoeae penicillin-binding protein 3 exhibits exceptionally high carboxypeptidase and beta-lactam binding activities
.
Biochemistry
42
:
14614
14625
.

Steinmoen
H
Teigen
A
Havarstein
LS
(
2003
)
Competence-induced cells of Streptococcus pneumoniae lyse competence-deficient cells of the same strain during cocultivation
.
J Bacteriol
185
:
7176
7183
.

Strynadka
NCJ
James
MNG
(
1996
)
Lysozyme
: a model enzyme in protein crystallography.
Lysozymes: Model Enzymes in Biochemistry and Biology
(
Jolles
P
, ed), pp.
185
222
.
Birkhäuser Verlag
, Basel.

Stubbs
KA
Balcewich
M
Mark
BL
Vocadlo
DJ
(
2007
)
Small molecule inhibitors of a glycoside hydrolase attenuate inducible AmpC mediated beta-lactam resistance
.
J Biol Chem
282
:
21382
21391
.

Sudo
S
Dworkin
M
(
1972
)
Bacteriolytic enzymes produced by Myxococcus xanthus
.
J Bacteriol
110
:
236
245
.

Takahashi
J
Komatsuzawa
H
Yamada
S
et al. (
2002
)
Molecular characterization of an atl null mutant of Staphylococcus aureus
.
Microbiol Immunol
46
:
601
612
.

Templin
MF
Ursinus
A
Höltje
J-V
(
1999
)
A defect in cell wall recycling triggers autolysis during the stationary growth phase of Escherichia coli
.
EMBO J
18
:
4108
4117
.

Thunnissen
AM
Dijkstra
AJ
Kalk
KH
Rozeboom
HJ
Engel
H
Keck
W
Dijkstra
BW
(
1994
)
Doughnut-shaped structure of a bacterial muramidase revealed by X-ray crystallography
.
Nature
367
:
750
753
.

Thunnissen
AM
Isaacs
NW
Dijkstra
BW
(
1995a
)
The catalytic domain of a bacterial lytic transglycosylase defines a novel class of lysozymes
.
Proteins
22
:
245
258
.

Thunnissen
AM
Rozeboom
HJ
Kalk
KH
Dijkstra
BW
(
1995b
)
Structure of the 70-kDa soluble lytic transglycosylase complexed with bulgecin A. Implications for the enzymatic mechanism
.
Biochemistry
34
:
12729
12737
.

Tomasz
A
(
1974
)
The role of autolysins in cell death
.
Ann N Y Acad Sci
235
:
439
447
.

Tomasz
A
(
1979a
)
From penicillin-binding proteins to the lysis and death of bacteria: a 1979 view
.
Rev Infect Dis
1
:
434
467
.

Tomasz
A
(
1979b
)
The mechanism of the irreversible antimicrobial effects of penicillins: how the beta-lactam antibiotics kill and lyse bacteria
.
Annu Rev Microbiol
33
:
113
137
.

Tomasz
A
(
1980
)
On the mechanism of the irreversible antimicrobial effects of beta-lactams
.
Philos Trans R Soc Lond B Biol Sci
289
:
303
308
.

Uehara
T
Park
JT
(
2003
)
Identification of MpaA, an amidase in Escherichia coli that hydrolyzes the gamma-d-glutamyl-meso-diaminopimelate bond in murein peptides
.
J Bacteriol
185
:
679
682
.

Uehara
T
Park
JT
(
2004
)
The N-acetyl-d-glucosamine kinase of Escherichia coli and its role in murein recycling
.
J Bacteriol
186
:
7273
7279
.

Uehara
T
Park
JT
(
2007
)
An anhydro-N-acetylmuramyl-l-alanine amidase with broad specificity tethered to the outer membrane of Escherichia coli
.
J Bacteriol
189
:
5634
5641
.

Uehara
T
Suefuji
K
Valbuena
N
Meehan
B
Donegan
M
Park
JT
(
2005
)
Recycling of the anhydro-N-acetylmuramic acid derived from cell wall murein involves a two-step conversion to N-acetylglucosamine-phosphate
.
J Bacteriol
187
:
3643
3649
.

Uehara
T
Suefuji
K
Jaeger
T
Mayer
C
Park
JT
(
2006
)
MurQ Etherase is required by Escherichia coli in order to metabolize anhydro-N-acetylmuramic acid obtained either from the environment or from its own cell wall
.
J Bacteriol
188
:
1660
1662
.

Ursinus
A
Höltje
J-V
(
1994
)
Purification and properties of a membrane-bound lytic transglycosylase from Escherichia coli
.
J Bacteriol
176
:
338
343
.

Ursinus
A
Van Den Ent
F
Brechtel
S
De Pedro
M
Höltje
J-V
Löwe
J
Vollmer
W
(
2004
)
Murein (peptidoglycan) binding property of the essential cell division protein FtsN from Escherichia coli
.
J Bacteriol
186
:
6728
6737
.

Van Asselt
EJ
Thunnissen
AM
Dijkstra
BW
(
1999a
)
High resolution crystal structures of the Escherichia coli lytic transglycosylase Slt70 and its complex with a peptidoglycan fragment
.
J Mol Biol
291
:
877
898
.

Van Asselt
EJ
Dijkstra
AJ
Kalk
KH
Takacs
B
Keck
W
Dijkstra
BW
(
1999b
)
Crystal structure of Escherichia coli lytic transglycosylase Slt35 reveals a lysozyme-like catalytic domain with an EF-hand
.
Structure Fold Des
7
:
1167
1180
.

Van Asselt
EJ
Kalk
KH
Dijkstra
BW
(
2000
)
Crystallographic studies of the interactions of Escherichia coli lytic transglycosylase Slt35 with peptidoglycan
.
Biochemistry
39
:
1924
1934
.

Van Straaten
KE
Dijkstra
BW
Vollmer
W
Thunnissen
AM
(
2005
)
Crystal structure of MltA from Escherichia coli reveals a unique lytic transglycosylase fold
.
J Mol Biol
352
:
1068
1080
.

Van Straaten
KE
Barends
TR
Dijkstra
BW
Thunnissen
AM
(
2007
)
Structure of Escherichia coli Lytic Transglycosylase MltA with bound Chitohexaose: implications for peptidoglycan-binding and cleavage
.
J Biol Chem
282
:
21197
21205
.

Vazquez-Laslop
N
Lee
H
Hu
R
Neyfakh
AA
(
2001
)
Molecular sieve mechanism of selective release of cytoplasmic proteins by osmotically shocked Escherichia coli
.
J Bacteriol
183
:
2399
2404
.

Vollmer
W
(
2007
)
Structural variation in the glycan strands of bacterial peptidoglycan
.
FEMS Microbiol Rev
32
:
287
306
.

Vollmer
W
Bertsche
U
(
2007
)
Murein (peptidoglycan) structure, architecture and biosynthesis in Escherichia coli
.
Biochimica et Biophysica Acta (BBA) – Biomembranes
, doi: DOI:

Vollmer
W
Höltje
J-V
(
2001
)
Morphogenesis of Escherichia coli
.
Curr Opin Microbiol
4
:
625
633
.

Vollmer
W
Pilsl
H
Hantke
K
Höltje
J-V
Braun
V
(
1997
)
Pesticin displays muramidase activity
.
J Bacteriol
179
:
1580
1583
.

Vollmer
W
Von Rechenberg
M
Höltje
J-V
(
1999
)
Demonstration of molecular interactions between the murein polymerase PBP1B, the lytic transglycosylase MltA, and the scaffolding protein MipA of Escherichia coli
.
J Biol Chem
274
:
6726
6734
.

Vollmer
W
Blanot
D
De Pedro
MA
(
2008
)
Peptidoglycan structure and architecture
.
FEMS Microbiol Rev
32
:
149
167
.

Von Rechenberg
M
Ursinus
A
Höltje
J-V
(
1996
)
Affinity chromatography as a means to study multienzyme complexes involved in murein synthesis
.
Microb Drug Resist
2
:
155
157
.

Vötsch
W
Templin
MF
(
2000
)
Characterization of a beta-N-acetylglucosaminidase of Escherichia coli and elucidation of its role in muropeptide recycling and beta-lactamase induction
.
J Biol Chem
275
:
39032
39038
.

Weidel
W
Pelzer
H
(
1964
)
Bagshaped macromolecules – a new outlook on bacterial cell walls
.
Advances in Enzymology
26
:
193
232
.

Whisstock
JC
Lesk
AM
(
1999
)
SH3 domains in prokaryotes
.
Trends Biochem Sci
24
:
132
133
.

Wuenscher
MD
Kohler
S
Bubert
A
Gerike
U
Goebel
W
(
1993
)
The iap gene of Listeria monocytogenes is essential for cell viability, and its gene product, p60, has bacteriolytic activity
.
J Bacteriol
175
:
3491
3501
.

Yamada
S
Sugai
M
Komatsuzawa
H
Nakashima
S
Oshida
T
Matsumoto
A
Suginaka
H
(
1996
)
An autolysin ring associated with cell separation of Staphylococcus aureus
.
J Bacteriol
178
:
1565
1571
.

Yamaguchi
H
Furuhata
K
Fukushima
T
Yamamoto
H
Sekiguchi
J
(
2004
)
Characterization of a new Bacillus subtilis peptidoglycan hydrolase gene, yvcE (named cwlO), and the enzymatic properties of its encoded protein
.
J Biosci Bioeng
98
:
174
181
.

Yamamoto
H
Kurosawa
S
Sekiguchi
J
(
2003
)
Localization of the vegetative cell wall hydrolases LytC, LytE, and LytF on the Bacillus subtilis cell surface and stability of these enzymes to cell wall-bound or extracellular proteases
.
J Bacteriol
185
:
6666
6677
.

Yang
JC
Van Den Ent
F
Neuhaus
D
Brevier
J
Löwe
J
(
2004
)
Solution structure and domain architecture of the divisome protein FtsN
.
Mol Microbiol
52
:
651
660
.

Young
R
(
1992
)
Bacteriophage lysis: mechanism and regulation
.
Microbiol Rev
56
:
430
481
.

Zahrl
D
Wagner
M
Bischof
K
et al. (
2005
)
Peptidoglycan degradation by specialized lytic transglycosylases associated with type III and type IV secretion systems
.
Microbiology
151
:
3455
3467
.

Editor: Jacques Coyette