Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter October 29, 2013

MurD enzymes: some recent developments

  • Roman Šink

    Roman Šink is a post-doctoral researcher at the Faculty of Pharmacy of the University of Ljubljana. He was trained in organic and medicinal chemistry, enzymology and X-ray crystallography. He obtained his PhD degree in pharmaceutical sciences from the University of Ljubljana in 2010. His main research topics have been the design, synthesis, and study of the binding modes of peptidoglycan biosynthetic pathway inhibitors. Currently his research activity focuses on the design and synthesis of new antituberculotics.

    , Hélène Barreteau

    Hélène Barreteau is assistant professor at the University Paris-Sud, Orsay, and a member of the Laboratory of Bacterial Envelopes and Antibiotics. She was trained in biochemistry and microbiology, and received her PhD degree in biochemistry from the University Picardie Jules-Verne, Amiens, in 2004. She is especially interested in the search for antimicrobial compounds and has been working on peptidoglycan as an antibacterial target for 8 years.

    , Delphine Patin

    Delphine Patin was trained in genetics and molecular biology at the University Paris Diderot, Paris. She received a Diplôme d’Etudes Approfondies in human genetics from the University Paris Diderot in 1999. She joined the Laboratory of Bacterial Envelopes and Antibiotics as an engineer in 2007. Her main research interests have been the study of Mur ligases from different bacterial species and the search for inhibitors of these enzymes.

    , Dominique Mengin-Lecreulx

    Dominique Mengin-Lecreulx was trained in biochemistry, enzymology, molecular biology and organic chemistry and obtained his PhD degree at the University Paris-Sud in 1987. He is director of research at CNRS and, since 2001, is the head of the Laboratory of Bacterial Envelopes and Antibiotics which, for more than 30 years, has been studying the biochemistry, genetics and physiology of the bacterial cell-wall peptidoglycan biosynthesis pathway as a whole. His main research interests concern the identification of genes and enzymes involved in this pathway, the regulation of this metabolism, the mode of action of antibiotics and bacteriocins interfering with this pathway and the corresponding resistance mechanisms, as well as the molecular bases for the recognition of bacterial cell-wall peptidoglycan by host innate immunity mechanisms.

    , Stanislav Gobec

    Stanislav Gobec is professor of medicinal chemistry at the Faculty of Pharmacy, University of Ljubljana. He graduated in pharmacy and received his PhD degree in medicinal chemistry from the University of Ljubljana in 1999. His main research interests are the structure-based design, synthesis and evaluation of small-molecule enzyme inhibitors and receptor ligands. For more than 15 years he has been involved in discovery of inhibitors of bacterial cell-wall biosynthesis as potential antibacterial agents.

    and Didier Blanot

    Didier Blanot is director of research at CNRS and a member of the Laboratory of Bacterial Envelopes and Antibiotics, University Paris-Sud, Orsay. He was trained in organic chemistry, biochemistry and enzymology, and received his PhD degree in natural sciences from the University Paris-Sud in 1979. For more than 30 years, his main research interests have been the functional and structural study of the enzymes of peptidoglycan biosynthesis, the determination of the structure of peptidoglycan from certain bacterial species, and the search for peptidoglycan biosynthesis inhibitors that could act as antibacterial agents.

    EMAIL logo
From the journal BioMolecular Concepts

Abstract

The synthesis of the peptide stem of bacterial peptidoglycan involves four enzymes, the Mur ligases (MurC, D, E and F). Among them, MurD is responsible for the ATP-dependent addition of d-glutamic acid to UDP-MurNAc-l-Ala, a reaction which involves acyl-phosphate and tetrahedral intermediates. Like most enzymes of peptidoglycan biosynthesis, MurD constitutes an attractive target for the design and synthesis of new antibacterial agents. Escherichia coli MurD has been the first Mur ligase for which the tridimensional (3D) structure was solved. Thereafter, several co-crystal structures with different ligands or inhibitors were released. In the present review, we will deal with work performed on substrate specificity, reaction mechanism and 3D structure of E. coli MurD. Then, a part of the review will be devoted to recent work on MurD orthologs from species other than E. coli and to cellular organization of Mur ligases and in vivo regulation of the MurD activity. Finally, we will review the different classes of MurD inhibitors that have been designed and assayed to date with the hope of obtaining new antibacterial compounds.

Introduction

Peptidoglycan (murein) is an essential component of the bacterial cell wall (1). It is found on the outside of the cytoplasmic membrane of almost all bacteria. It consists of linear glycan chains alternating N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc) residues linked by β-1→4 bonds. The d-lactoyl group of each MurNAc residue is substituted by a peptide stem whose composition is often l-Ala-γ-d-Glu-meso-A2pm (or l-Lys)-d-Ala-d-Ala (A2pm, 2,6-diaminopimelic acid) in nascent peptidoglycan, the last d-Ala residue being removed in the mature macromolecule. Cross-linking of the glycan strands generally occurs between the carboxyl group of d-Ala at position 4 and the amino group of the diamino acid at position 3, either directly in most Gram-negative bacteria or through a short peptide bridge in most Gram-positive organisms. This results in a rigid tridimensional (3D) network surrounding the entire cell.

The biosynthesis of peptidoglycan is a complex process that involves ca. 20 reactions occurring in the cytoplasm (synthesis of the nucleotide precursors) and on the inner side (synthesis of lipid-linked intermediates) and outer side (polymerization reactions) of the cytoplasmic membrane (2). The cytoplasmic steps lead to the formation of UDP-MurNAc-pentapeptide from UDP-GlcNAc. After the formation of UDP-MurNAc (enzymes MurA and MurB), the successive additions of l-Ala (MurC), d-Glu (MurD), diamino acid (MurE) and d-Ala-d-Ala (MurF) are catalyzed by four enzymes, the Mur ligases (3). Thereafter, the phospho-MurNAc-pentapeptide moiety is transferred (enzyme MraY) to a membrane polyprenol, undecaprenyl phosphate, yielding lipid I. The addition of a GlcNAc residue from UDP-GlcNAc (catalyzed by MurG) leads to the formation of lipid II, which contains the whole monomer motif (4). The disaccharide-pentapeptide motif then crosses the membrane to reach the outer face; recent work has demonstrated that this process is performed by a specific protein, FtsW (5). Finally, the polymerization reactions (transglycosylation and transpeptidation), catalyzed by the penicillin-binding proteins, take place (6).

The main function of peptidoglycan is to preserve cell integrity by withstanding the inner osmotic pressure. Indeed, inhibition of its biosynthesis (mutation, antibiotic) or its specific degradation (e.g., by lysozyme) during bacterial growth will generally result in cell lysis. Owing to the different forms of resistance that pathogenic bacteria are developing, there is an urgent need to search for new inhibitors (7). In this context, peptidoglycan is an attractive target (8). As a matter of fact, most antibiotics known to interfere with peptidoglycan biosynthesis (e.g., β-lactams, vancomycin, bacitracin) target the polymerization reactions or the synthesis of lipid-linked intermediates. This is the reason why the cytoplasmic steps, and in particular the Mur ligases, which were underexploited in the past, have recently gained more attention.

Among the Mur ligases, MurD from Escherichia coli has been the most studied enzyme from biochemical, mechanistic and structural standpoints; perhaps as a consequence of this, many inhibitors have been designed and synthesized. In this paper, we will review the knowledge recently gathered about MurD ligase orthologs from E. coli and other species.

Generalities on Mur ligases

As mentioned in the Introduction, four enzymes, MurC, MurD, MurE and MurF, are involved in the de novo synthesis of the peptide stem of peptidoglycan (3). A fifth enzyme, Mpl, directly adds tripeptide l-Ala-γ-d-Glu-meso-A2pm to UDP-MurNAc during peptidoglycan recycling in some Gram-negative bacteria (9, 10). The Mur ligases catalyze the formation of an amide or peptide bond with the simultaneous formation of ADP and Pi from ATP. A divalent cation, Mg2+ or Mn2+, is necessary. The Mur ligases share the following three characteristics.

  1. They have a common reaction mechanism, which consists first in the activation of the carboxyl group of the UDP-containing precursor by ATP to yield an acyl-phosphate and ADP; the acyl-phosphate then undergoes a nucleophilic attack by the amino group of the amino acid (or dipeptide) substrate to form a high-energy tetrahedral intermediate, which eventually breaks down into amide or peptide and Pi [(11) and references therein].

  2. They possess a series of invariant residues in addition to an ATP-binding consensus sequence. This finding led to the definition of the Mur ligases as a new family of enzymes (12–14). The family also includes MurT, an enzyme involved in d-isoglutamine formation in lipid II in Staphylococcus aureus (15, 16), and three enzymes not related to peptidoglycan synthesis, folylpoly-γ-l-glutamate synthetase FolC (17), cyanophycin synthetase CphA (18) and poly-γ-glutamate synthetase CapB (19).

  3. They display similar 3D structures in three domains, the N-terminal, central and C-terminal domains, being involved in the binding of the nucleotide precursor, ATP, and the amino acid or dipeptide, respectively. Whereas the topologies of the central and the C-terminal domains are similar among the Mur ligases, that of the N-terminal domain shows differences related to the length of the UDP-precursor substrates, with MurC and MurD on the one hand, and MurE and MurF on the other hand. Moreover, Mur ligases exist as ‘open’ and ‘closed’ conformations, the closure being thought to result from ligand binding (20).

Properties of the MurD enzyme from E. coli

Substrate specificity

MurD is the enzyme adding the second amino acid on the peptide stem, and in all of the bacteria studied to date, this amino acid turns out to be d-glutamic acid (1, 21). The minor variations that can be encountered in mature peptidoglycan (d-isoglutamine, d-threo-3-hydroxyglutamate) are the result of modifications posterior to the action of Mur ligases (15, 16, 22).

The E. coli MurD enzyme has been purified and characterized in the 1990s in its native (23–25) and His-tagged (14) forms. Optimal pH value, optimal Mg2+ concentration and kinetic parameters can be seen in Table 1. Its substrate specificity has been studied for the addition of various amino acids or derivatives to the radiolabeled precursor UDP-MurNAc-l-[14C]Ala (26). It appeared that (i) an α-amino acid moiety of d configuration, (ii) an anionic group on carbon 4 and (iii) a proper distance between the two acid groups were essential requirements for a substrate. As a matter of fact, apart from d-Glu, very few compounds turned out to be good substrates: dl-homocysteic acid, d-erythro-3- and -4-methylglutamic acids, and two cyclic analogs of Glu (five- or six-membered ring). Moreover, l-Glu is absolutely not a substrate of the enzyme (26), even though some N-sulfonyl- or N-acyl-derivatives of l-Glu were shown to inhibit E. coli MurD (27–29).

While MurD is very specific with respect to its amino acid substrate, it differs from the other Mur ligases by its less strict specificity towards the nucleotide precursor. Indeed, it has been shown that the removal of the UMP moiety leads to a substrate (1-phospho-MurNAc-l-Ala) having a fairly good affinity for the enzyme (Km=0.2 mmol/l) (30).

Table 1

Optimal pH, optimal Mg2+ concentration, and kinetic parameters of MurD orthologs.

SpeciesPeptide extensionOptimal pHOptimal [Mg]2+ (mmol/l)Km ATP (mmol/l)Km UMAa (mmol/l)Kmd-Glu (mmol/l)Vmax(μmol/[min mg])References
E. coliNo8.9–9.250.1380.00750.0558.4(23, 24)
No8b5c0.135±0.0330.007±0.0030.042±0.0114.78±0.42(25)
His-tag9.4b50.0570.0050.0648.4(14)
S. aureusNo8b20–250.084±0.0100.29±0.0340.534±0.02532.1±3.7(25)
His-tag8b5cn.d.dn.d.n.d.5e(105)
His-tag8.4–9.6155.4±1.90.041±0.020f0.13±0.044f35±6(51)
S. pneumoniaeHis-tag7.720n.d.0.030.105n.d.(106)
His-tag8.2–9.65∼2g0.096±0.039h0.19±0.026h15±3(52)
P. aeruginosaHis-tag8b5cn.d.n.d.n.d.n.d.(107)
His-tagn.s.i5cn.d.n.d.n.d.3.12j(108)
E. faecalisNo8b10–250.047±0.0040.036±0.0070.118±0.01429.8±5.9(25)
H. influenzaeNo8b5 or 10c0.102±0.0060.008±0.0040.169±0.02013.1±2.2(25)
B. burgdorferiHis-tag8.2–9.050.053±0.0120.0063±0.00300.11±0.0291.6±0.3(52)
M. tuberculosisHis-tag8.8–9.8100.71±0.290.34±0.010.70±0.180.49±0.11(52)
Yesk8.550.106±0.00010.053±0.00010.085±0.00020.8±0.1(63)

Unless noted otherwise, Km values were determined at saturating concentrations of the other substrates. Detailed conditions can be found in the references. Partially purified enzymes, or purified enzymes whose activity was not checked, were not considered.

aUMA, UDP-MurNAc-l-Ala.

bpH of the assay, no optimal value specified.

cConcentration used in the assay, no optimal value specified.

dn.d., not determined.

eSpecific activity at 1 mmol/l ATP, 0.12 mmol/l UMA and 1 mmol/l d-Glu.

fDetermined at subsaturating ATP concentration (2 mmol/l) owing to high Km ATP value.

gMarked inhibition by excess ATP over 1.5 mmol/l.

hDetermined at subsaturating ATP concentration (1.5 mmol/l) owing to inhibition by excess ATP.

in.s., not specified.

jSpecific activity at 1 mmol/l ATP, 0.1 mmol/l UMA and 0.5 mmol/l d-Glu.

k21-amino acid N-terminal extension.

Reaction mechanism and order of binding

As mentioned above, the Mur ligases share the same reaction mechanism involving two intermediates, an acyl-phosphate and a high-energy tetrahedral intermediate (Scheme 1). The formation of acyl-phosphate was first suggested indirectly, mainly with MurC [see references in (11)]. It was then firmly established for MurC and MurD through the use of the borohydride reduction method, introduced by Degani and Boyer (31) and Todhunter and Purich (32). This method allowed to chemically trap the acyl-phosphate intermediates into stable derivatives (2-amino-2-deoxy-3-O-[(S)-(1-hydroxymethyl)ethyl]-d-glucose for MurC, l-alaninol for MurD) that were clearly identified (Scheme 1) (11, 33). The existence of the tetrahedral intermediate was suggested by the strong inhibition of the MurD activity with phosphinate analogues (Scheme 1) (34–36).

Scheme 1 Reaction mechanism of MurD showing the acyl-phosphate and tetrahedral intermediates (black). The acyl-phosphate intermediate can be trapped by reduction with sodium borohydride followed by acid hydrolysis, yielding l-alaninol (red). The existence of the high-energy tetrahedral intermediate is suggested by the strong inhibition of the MurD activity by a phosphinate analog of tetrahedral geometry (purple). Stereochemistry was omitted for the sake of simplicity. R, UDP-MurNAc.
Scheme 1

Reaction mechanism of MurD showing the acyl-phosphate and tetrahedral intermediates (black). The acyl-phosphate intermediate can be trapped by reduction with sodium borohydride followed by acid hydrolysis, yielding l-alaninol (red). The existence of the high-energy tetrahedral intermediate is suggested by the strong inhibition of the MurD activity by a phosphinate analog of tetrahedral geometry (purple). Stereochemistry was omitted for the sake of simplicity. R, UDP-MurNAc.

A particular reaction was observed with MurD: the synthesis of adenosine tetraphosphate (33). This compound was formed upon incubation of the enzyme with ATP and UDP-MurNAc-l-Ala (without d-Glu). This synthesis was explained by a reaction of ATP with the acyl-phosphate intermediate. No adenosine tetraphosphate formation could be detected with MurC (33), but this molecule was present in the crystal structure of FolC (37).

The kinetic mechanism was established for MurC (38) and MurF (39). Data were unambiguously consistent with a sequential ordered kinetic mechanism, with ATP binding first, followed by the nucleotide precursor and, finally, amino acid (or dipeptide). Although such experiments were never performed with MurD, it is reasonable to assume that it follows the same type of mechanism. However, the fact that the molecular isotope exchange reaction was not strictly ADP-dependent (35), contrary to MurC (40), suggested some randomness in the kinetic mechanism.

Tridimensional structure

MurD from E. coli was the first Mur ligase for which a crystal structure, in complex with substrate UDP-MurNAc-l-Ala, was solved (41). Thereafter, other structures containing ADP, Mg2+ or Mn2+, the product UDP-MurNAc-l-Ala-d-Glu, or inhibitors, were reported (27, 42), as well as ‘open’ conformations (43). The 3D structure of MurD from E. coli consists of three domains (Figure 1A). The N-terminal domain is responsible for the binding of UDP-MurNAc-l-Ala, the central domain binds ATP and the C-terminal domain binds d-glutamic acid. The most important residues for UDP-MurNAc-l-Ala binding are Leu15, Thr16, Asp35, Thr36, Arg37, Gly73, Asn138 and His183, the last one via Mg2+ (Figure 1B) (41–43). ADP is held in place thanks to the interactions with Gly114, Lys115, Ser116, Thr117, Asn271, Arg302 and Asp317 (Figure 1C) (27, 42), while d-glutamic acid interacts with Thr321, Lys348, Ser415, Leu416 and Phe422 (Figure 1D) (42). Interestingly, the topologies of the central and C-terminal domains of E. coli MurD were found to be similar to those of FolC (17, 43).

Figure 1 Tridimensional structure of, and interactions of ligands with, E. coli MurD.(A) Tridimensional structure of E. coli MurD (PDB entry: 4UAG). Blue: N-terminal domain; red: central domain; magenta: C-terminal domain. (B) Interactions of UDP-MurNAc-l-Ala with MurD. (C) Interactions of ADP with MurD. (D) Interactions of d-glutamic acid with MurD. (E) Superposition of closed conformation (green, PDB entry: 2UAG) with two open conformations (blue, PDB entry: 1E0D; red, PDB entry: 1EEH) of E. coli MurD. N-terminal and central domains were aligned.
Figure 1

Tridimensional structure of, and interactions of ligands with, E. coli MurD.

(A) Tridimensional structure of E. coli MurD (PDB entry: 4UAG). Blue: N-terminal domain; red: central domain; magenta: C-terminal domain. (B) Interactions of UDP-MurNAc-l-Ala with MurD. (C) Interactions of ADP with MurD. (D) Interactions of d-glutamic acid with MurD. (E) Superposition of closed conformation (green, PDB entry: 2UAG) with two open conformations (blue, PDB entry: 1E0D; red, PDB entry: 1EEH) of E. coli MurD. N-terminal and central domains were aligned.

Tridimensional structures were also determined for MurD enzymes from Streptococcus agalactiae (44) and Thermotoga maritima (45). The domain arrangement is similar to that of the E. coli ortholog. An important feature that was observed in E. coli MurD was the rigid body C-terminal domain rotation. The position of the C-terminal domain dictates the conformation of MurD: the closed conformation (C-terminal domain rotated towards the central and N-terminal domains) is considered as the active conformation, while open conformations (C-terminal domain rotated away from the N-terminal and central domains) are considered as inactive (43, 46). Binding of substrates (UDP-MurNAc-l-Ala-d-Glu and ATP, or only ATP) provokes C-terminal domain rotation to the closed conformation, while binding of UDP-MurNAc-l-Ala-d-Glu provokes its rotation either to the closed conformation or to the distinct open conformation (27, 41–43). Computational simulations revealed that a small energy barrier exists between the open conformation without ligands and the closed conformation (47). Different C-terminal domain conformations are presented in Figure 1E. C-terminal domain rotation was also observed in T. maritima MurD (45).

A catalytic mechanism for MurD at the molecular level was proposed by analyzing the biochemical data together with the 3D structural data of MurD complexes with ligands. The two ADP-bound crystal structures of MurD revealed the ATP-binding mode and two important divalent cation (Mg2+ or Mn2+) binding sites (site 1 and site 2), which seem to be very important for the phosphorylation of UDP-MurNAc-l-Ala in the presence of ATP (Scheme 2). Although the electron density for the γ-phosphate of ATP was never visible, it was presumed that one of the γ-phosphoryl oxygen atoms interacts with Mg2+ in site 1. This approximation suggests that the γ-phosphate is transferred to UDP-MurNAc-l-Ala via a bimolecular nucleophilic substitution type mechanism (Sn2). After the transfer of the γ-phosphate, the acyl-phosphate intermediate is stabilized with interactions with Lys115 and the two Mg2+ ions. Then, the properly oriented incoming d-Glu is added to the acyl-phosphate intermediate, resulting in the formation of the tetrahedral intermediate. A catalytic base is required to withdraw a proton from the amino group of d-Glu. In the disruption of the tetrahedral intermediate, another catalytic base (probably His183) is essential for the withdrawal of the second proton (42) and thus the formation of UDP-MurNAc-l-Ala-d-Glu. Enzymatic reaction pathways were recently studied using a hybrid quantum mechanical/molecular mechanical replica path method and found to be in line with the proposed sequential mechanism (48). However, it was discovered that proton transfer from d-Glu to either the γ-phosphate of ATP or the nitrogen of His183 was energetically too demanding to fit in the normal energy scope of enzymatic reactions. This suggested that d-Glu most likely enters the enzyme reaction in its deprotonated form obtained by an as-yet unidentified mechanism (48).

Scheme 2 Schematic representation of the γ-phosphate transfer from ATP to UDP-MurNAc-l-Ala, and important residues that are involved in the reaction. Mg2+ ions in sites 1 and 2 are depicted as green and cyan spheres, respectively.
Scheme 2

Schematic representation of the γ-phosphate transfer from ATP to UDP-MurNAc-l-Ala, and important residues that are involved in the reaction. Mg2+ ions in sites 1 and 2 are depicted as green and cyan spheres, respectively.

An important chemically modified residue in the proximity of the active site is carbamylated lysine 198 (KCX198), which interacts via a water bridge with the Mg2+ ion in site 2, thus stabilizing its position (Scheme 2) (42). KCX198 was only observed in the ligand-bound closed E. coli MurD crystal structures (27, 41, 42), while in open structures (43) Lys198 was not carbamylated. From an enzymatic point of view, the importance of the carbamyl group, which is also present in MurE and MurF, was investigated by chemical rescue experiments (49). It was shown that mutant proteins MurD Lys198Ala/Phe/Glu/Cys, which displayed low enzymatic activity, were rescued upon incubation with short-chain carboxylic acids, but not amines. These observations are in favor of a functional role for the carbamate in MurD.

MurD enzymes from other bacteria

After MurD from E. coli, several orthologs of the MurD enzyme from various bacteria were isolated and characterized. A list, along with optimal pH values, optimal Mg2+ concentrations and kinetic parameters, is presented in Table 1. In almost all cases, Km values are in the micromoles per liter range. A few discrepancies between data from different laboratories were encountered (e.g., for S. aureus or Mycobacterium tuberculosis), presumably owing to different assay conditions. For two MurD orthologs (S. aureus and Streptococcus pneumoniae), high Km values for ATP (≥2 mmol/l) were observed. Such unusual values have also been found for MurC from S. aureus and MurE from T. maritima (50, 51). Maximum velocities of enzymes from Gram-positive bacteria are higher than those from Gram-negative bacteria. In two cases (Borrelia burgdorferi and M. tuberculosis), Vmax values are quite low. This might be related to the slow-growth properties of the bacterial species in question (52).

It is interesting to mention that d-isoglutamine was not a substrate for the enzymes originating from S. aureus, S. pneumoniae and M. tuberculosis [(51); H. Barreteau and D. Blanot, unpublished results], species in which this amino acid is found at position 2 of the peptide stem of mature peptidoglycan. This is in keeping with amidation occurring at a later stage of peptidoglycan synthesis, namely, lipid II (15, 16, 53).

Genetic aspects, cellular organization and regulation

Conditional-lethal (thermosensitive) mutants of E. coli altered at different levels of the peptidoglycan pathway were isolated in the early 1970s, and the mutations were mapped in several chromosomal regions (54–56). One region of particular interest, located at 2 min of the E. coli map, was named mra (murein region A) or dcw (division cell wall) because it contained a large cluster of genes, from mraZ to envA, that coded for proteins involved in peptidoglycan biosynthesis and cell division. The murD gene encoding the d-glutamic acid-adding enzyme was identified within this cluster (57, 58), together with the murC, murE and murF genes encoding the other three Mur ligases, and the mraY and murG genes encoding lipids I and II synthesizing activities. These different genes, tightly packed and overlapping in some cases, were transcribed in the same direction, and their expression was mainly dependent on a promoter, Pmra, identified upstream of the cluster (59). Repression of Pmra resulted in a dramatic depletion of MurD ligase in vivo and was followed by the arrest of peptidoglycan synthesis and cell lysis (60). Demonstration that the E. coli murD gene is essential for growth was recently provided by the construction of a conditional mutant strain carrying the murD gene on a plasmid bearing a thermosensitive replicon (51). A similar cluster organization with interspersed mur (murein synthesis) and fts (division) genes was found in many other bacterial species. As discussed by Mingorance et al. (61), such a ‘genomic channeling’ could favor the formation of Mur and Fts protein complexes which then contribute together to the channeling of peptidoglycan precursors from the cytoplasm to the cell wall-synthesizing machinery operating at the division site. However, this is not a general rule as Mur ligase-encoding genes of many species, for instance, murC in Bacillus subtilis or murE in Thermus thermophilus, appear variably dispersed in other chromosomal locations. No evidence that Mur ligases are assembled into functional complexes in vivo has been provided to date, but interactions between these enzymes and other proteins involved in peptidoglycan synthesis or cell division were reported in Caulobacter crescentus (62), M. tuberculosis (63) and T.maritima (45). It has been speculated, but not demonstrated, that a concerted action of Mur ligases in this pathway, achieved through the formation of a multi-enzyme complex and allowing the channeling of intermediates, may explain that the lack of efficiency of some potent Mur ligase inhibitors observed in vivo is due to a restricted accessibility of these compounds to the active sites of these enzymes (64, 65).

It was earlier shown that E. coli Mur ligases were expressed constitutively at a relatively high level as compared to the cell requirements for peptidoglycan synthesis (66), suggesting that there was no rate-limiting step in this part of the pathway. This was consistent with the low pool levels observed for these different UDP-MurNAc-peptide intermediates in growing cells. The unrestricted accumulation of some of these intermediates that is observed in wild-type cells treated with antibiotics [UDP-MurNAc-tripeptide following treatment with d-cycloserine (67), a d-Ala:d-Ala ligase inhibitor] or in conditional mutants grown in restrictive conditions [UDP-MurNAc-l-Ala in a d-glutamate-requiring mutant starved for this amino acid (68)] further suggested that these pools were not tightly regulated and that the intermediates can be released from the enzyme active sites and escape a putative channeling process. However, this does not exclude that the activity of Mur ligases could be regulated in some way. Feedback inhibition of the MurD activity by its reaction product, UDP-MurNAc-l-Ala-d-Glu, as well as by the final nucleotide precursor of the pathway, UDP-MurNAc-pentapeptide, has been reported earlier, which could contribute to the regulation of this enzyme activity (68, 69). Moreover, the fact that MurD enzymes from E.coli and Haemophilus influenzae are (i) inhibited by moderate (15–30 μmol/l) concentrations of the substrate UDP-MurNAc-l-Ala and (ii) require monocations such as NH4+ and/or K+ for optimal activity, contrary to S. aureus and Enterococcus faecalis orthologs, has led Walsh et al. (25) to postulate a key role for MurD in the regulation of peptidoglycan synthesis in Gram-negative bacteria. More recently, Mur ligases of some species were identified as targets of serine-threonine protein kinases (STPKs) and shown to be phosphorylated by these activities both in vitro and in vivo. This was in particular reported for MurC from Corynebacterium glutamicum (70) and S. pneumoniae (71), and for MurD from M. tuberculosis (72). The phosphorylation of C. glutamicum MurC protein by the PknA kinase, which occurs on six different threonine residues, was shown to negatively modulate its ligase activity in vitro (70). Protein phosphorylation was also demonstrated for other enzymes catalyzing earlier cytoplasmic steps or the final polymerization steps of this pathway: GlmM (73), GlmU (74) and penicillin-binding proteins (75). Deletion of S. aureus STP phosphatase and pknB serine/threonine kinase genes brought about alterations of the pool levels of peptidoglycan nucleotide precursors that also suggested a regulation of Mur ligase activities by phosphorylation in this bacterial species; according to the quantitative changes in the relevant metabolites, the effect was stronger for MurC/F than for MurD/E (76).

MurD inhibitors

Most of the efforts in designing MurD inhibitors were concentrated on MurD from E. coli, for which several classes and generations of inhibitors have been developed. MurD inhibitors can be divided into several distinct structural classes as presented in Figure 2.

Figure 2 Different classes of MurD inhibitors.
Figure 2

Different classes of MurD inhibitors.

The most fundamental distinction of the classification is between peptide and non-peptide inhibitors. Among the former, two cyclic nonapeptides are worth mentioning, with amino acid sequences Cys-Pro-Ala-His-Trp-Pro-His-Pro-Cys and Cys-Ser-Ala-Trp-Ser-Asn-Lys-Phe-Cys. They inhibit E. coli MurD with IC50 values of 1.5 and 0.62 mmol/l, respectively (77). Linear peptide Arg-Pro-Thr-His-Ser-Pro-Ile, which inhibits Pseudomonas aeruginosa MurD in the low micromoles per liter range (IC50, 4 μmol/l), was also discovered (78).

Among the non-peptide inhibitors, the non-glutamic acid-based inhibitors constitute a distinct class (Scheme 3). Macrocyclic inhibitors of MurD from E. coli were discovered using computer-based molecular design. The most potent inhibitor 1 had an IC50 value of 0.7 μmol/l (79). 2-Phenyl-5,6-dihydro-2H-thieno[3,2-c]pyrazol-3-ol derivatives were found to inhibit MurD from S. aureus (but also MurB and MurC) in the micromoles per liter range. The most potent compound 2 inhibited MurD with an IC50 value of 33 μmol/l (80). An important property of these compounds is their antibacterial activity: compound 2 inhibited the growth of some Gram-positive strains, while the antibacterial activity against Gram-negative strains was much lower (80). Pulvinones inhibited the MurA-D enzymes from E. coli and S. aureus. Representative inhibitor 3 also expressed antibacterial activity against some Gram-positive strains (81). Structurally diverse 9H-xanthene derivatives and polycyclic inhibitors were discovered with structure-based virtual screening. The most potent representative from this campaign (compound 4) inhibited E. coli MurD with an IC50 value of 10 μmol/l (82). N-Acyl-hydrazones were discovered by screening an in-house bank of compounds on MurD (83). One promising inhibitor (compound 5) can be highlighted in the synthesized series. It inhibited E. coli MurD and MurC with IC50 values of 230 and 123 μmol/l, respectively. Another feature of this compound is a weak antibacterial activity of 128 μg/ml against S. aureus, E. coli and the efflux pump AcrAB-deficient E. coli strain (83). Sulfonohydrazides were designed and synthesized as analogs that mimic a diphosphate part of the MurD substrate UDP-MurNAc-l-Ala. In the presence of a 500 μmol/l concentration of the most potent inhibitor (compound 6), E. coli MurD still retained 56% residual activity (84). The sulfonohydrazone moiety was incorporated into previously known MurD inhibitors, thus forming a new class. The most potent compound of this series (compound 7) inhibited E. coli MurC and MurD (IC50, 30 μmol/l for both enzymes) (85). 5-Benzylidenethiazolidin-4-ones were discovered by combining rhodanine, which had previously been employed as a diphosphate surrogate or a phosphate mimetic, with the 2,3,4-trihydroxyphenyl group, which had been incorporated in earlier MurD inhibitors. The best compound (8) inhibited E. coli MurD, S. aureus MurE and E. coli MurF in the low micromoles per liter range, thus appearing as a multi-target inhibitor. Unfortunately, it displayed only weak antibacterial activity against P. aeruginosa (MIC, 128 μg/ml) (86). Benzene-1,3-dicarboxylic acid derivatives were discovered in a virtual screening campaign. The most potent compound (9) had an IC50 value of 270 μmol/l with E. coli MurD. It was also a fairly good MurE inhibitor (IC50, 32 μmol/l) (87).

Scheme 3 Representative non-glutamic acid-based inhibitors of MurD. Multi-target inhibitor 8 is depicted in red.
Scheme 3

Representative non-glutamic acid-based inhibitors of MurD. Multi-target inhibitor 8 is depicted in red.

The largest class of E. coli MurD inhibitors consists of glutamic acid-based inhibitors. As the reaction catalyzed by MurD proceeds via a tetrahedral intermediate, the phosphinate transition-state analogs were the first potent inhibitors of MurD that were described in the literature. Compound 10 (Scheme 4), in which the MurNAc moiety was replaced by a hydrophobic linker of appropriate length, inhibited MurD from E. coli with an IC50 value of 680 nmol/l (34). The potency of phosphinate inhibitors was improved with the inclusion of the MurNAc moiety, which led to the best inhibitor reported to date (compound 11, IC50 <1 nmol/l, Scheme 4) (36). The next generation of phosphinate transition-state analogs featured simplified compounds with the key phosphinodipeptide Ala-Ψ(PO2-CH2)-Glu structural motif. The inhibitory activities of such compounds (i.e., 12) were in the micromoles per liter range (88), and some of them also inhibited MurE (89).

Scheme 4 Representative phosphinate inhibitors of MurD. The most potent inhibitor reported to date (11) is depicted in red.
Scheme 4

Representative phosphinate inhibitors of MurD. The most potent inhibitor reported to date (11) is depicted in red.

Several 5-benzylidenethiazolidin-2,4-dione and 5-benzylidenerhodanine inhibitors that contained d-(or l-)glutamic acid in their structures were developed. Starting with virtual screening and taking into consideration the fact that rhodanine-containing compounds possess in vitro antimicrobial activity, a first generation was developed. An interesting finding of this research was the observation that the most potent MurD inhibitor did not lose significantly its activity when d-glutamic acid (compound 13, Scheme 5) was replaced with l-glutamic acid (compound 14) (28).

Scheme 5 5-Benzylidenerhodanines containing d-glutamic acid (13) and l-glutamic acid (14). IC50 values were determined with E. coli MurD.
Scheme 5

5-Benzylidenerhodanines containing d-glutamic acid (13) and l-glutamic acid (14). IC50 values were determined with E. coli MurD.

In the next generation of 5-benzylidenethiazolidin-2,4-dione- and 5-benzylidenerhodanine-substituted glutamic acid, the substitution of the aromatic rings was altered, which resulted in E. coli MurD inhibitors (compounds 15 and 16, Scheme 6) with IC50 values of 45 and 85 μmol/l, respectively (90). These compounds were further improved with minor alterations of the ring substituents and variations in the linker regions between the two phenyl rings. Such efforts produced inhibitor 17, which was the best of the series (IC50, 3 μmol/l). Another interesting inhibitor, compound 18, contained a sulfonamide function instead of the carboxamide one (IC50, 45 μmol/l) (91, 92). The next step in the development of this class of inhibitors featured a shorter linker between the glutamic acid part and the 4-thiazolidinone part of the molecule (29). Compound 19 inhibited E. coli MurD with an IC50 value of 10 μmol/l (Scheme 6). An interesting fact here is that the compound containing l-glutamic acid was the most potent one; the d-Glu-containing derivative 20 displayed a lower potency (IC50, 45 μmol/l) (29). Finally, an important progress in this series was the design and synthesis of compound 21, in which the NH-CH2 group connecting the two phenyl rings had been reversed (Scheme 6). This compound turned out to be a dual MurD and MurE inhibitor. It was endowed with antibacterial activity against S. aureus and its methicillin-resistant strain (MRSA) (MIC, 8 μg/ml) (93).

Scheme 6 5-Benzylidenethiazolidin-2,4-dione and 5-benzylidenerhodanine inhibitors. Unless stated otherwise, IC50 values were determined with E. coli MurD. Dual MurD and MurE inhibitor 21 is depicted in red.
Scheme 6

5-Benzylidenethiazolidin-2,4-dione and 5-benzylidenerhodanine inhibitors. Unless stated otherwise, IC50 values were determined with E. coli MurD. Dual MurD and MurE inhibitor 21 is depicted in red.

A highlight in the development of 4-thiazolidinone inhibitors was the determination of their binding mode by solving several X-ray structures of MurD/inhibitor complexes (90–93). The binding modes of these compounds are very similar, as can be seen from the structures of complexes with typical inhibitors 16 and 21 (Figure 3). A superposition of the crystal structures of inhibitor 16 [Protein Data Bank (PDB) entry: 2X5O] (90) and UDP-MurNAc-dipeptide (PDB entry: 4UAG) (42) within E. coli MurD reveals that the inhibitor occupies the binding site of the product (Figure 3A). The thiazolidin-2,4-dione ring occupies the uracil-binding site with interactions with Asp35, Thr36 and Arg37. The carboxylic groups of the d-glutamic acid part of the inhibitor occupy exactly the same position as those of UDP-MurNAc-dipeptide. The γ-carboxyl group interacts with Ser415, Leu416 and Phe422, while the α-carboxyl group is held in position through interactions with Thr321 and Lys348 directly. In the MurD/21 complex (PDB entry: 2Y1O), the α-carboxyl group interacts with Lys115 and Lys348 via water molecules (Figure 3B), which is a unique property of the binding mode of inhibitor 21 (93). No polar contacts between the protein and the central linker part of the inhibitors exist. Here, only π-π stacking of the aminophenyl ring with Phe161 and hydrophobic interactions of the benzylidene ring with Gly73 are worth mentioning.

Figure 3 Binding modes of 5-benzylidenethiazolidin-2,4-dione and 5-benzylidenerhodanine inhibitors.(A) Binding mode of MurD inhibitor 16 (magenta; PDB entry: 2X5O) superimposed to UDP-MurNAc-l-Ala-d-Glu (orange; PDB entry: 4UAG). (B) Binding mode of MurD inhibitor 21 (PDB entry: 2Y1O).
Figure 3

Binding modes of 5-benzylidenethiazolidin-2,4-dione and 5-benzylidenerhodanine inhibitors.

(A) Binding mode of MurD inhibitor 16 (magenta; PDB entry: 2X5O) superimposed to UDP-MurNAc-l-Ala-d-Glu (orange; PDB entry: 4UAG). (B) Binding mode of MurD inhibitor 21 (PDB entry: 2Y1O).

Sulfonamides represent the last class of important MurD inhibitors reported to date. The sulfonamide function was, similarly to the phosphinate one, incorporated into the target compounds in order to mimic the MurD tetrahedral intermediate (27, 94, 95). The first generation of naphthalene sulfonamides was N-substituted glutamic acid derivatives (27, 95) The inhibitory properties of 2-naphthalene sulfonamides 22 and 23 (Scheme 7), which differ only in the configuration of the glutamic acid residue, revealed that their mechanism of inhibition was competitive towards d-glutamic acid and non-competitive towards UDP-MurNAc-l-Ala or ATP. However, the potency of the l-enantiomer was significantly lower than that of the D-enantiomer (27). An important achievement of the thorough investigation of the inhibitory mechanism was the solution of the X-ray co-crystal structures of both enantiomers in the active site of MurD. Crystal structures revealed almost exactly the same binding mode of the two enantiomers, as both the l- and d-glutamic acid parts of the inhibitors occupy the same position as the glutamic acid part of UDP-MurNAc-l-Ala-d-Glu, while the naphthalene-sulfonamido groups are displaced by approximately 1.5 Å (27). Naphthalene sulfonamides were later further modified and biochemically and structurally evaluated (95). The outcome was a series of new inhibitors with the most potent compound (24), which inhibits MurD with an IC50 value of 85 μmol/l. The binding energies of the naphthalene sulfonamides were calculated using the linear interaction energy method; the calculated energies correlated very well with the experimentally obtained free energies (96). Further analysis of the interactions between naphthalene sulfonamides and MurD was performed by nuclear magnetic resonance followed by molecular dynamics, which takes into account the ligand flexibility and its effect on particular ligand-enzyme contact, thus offering potential explanations for moderate inhibitory activities (97).

Scheme 7 Representative naphthalene sulfonamides containing glutamic acid or rigid mimetics thereof. IC50 values were determined with E. coli MurD. The best inhibitor 26 is depicted in red.
Scheme 7

Representative naphthalene sulfonamides containing glutamic acid or rigid mimetics thereof. IC50 values were determined with E. coli MurD. The best inhibitor 26 is depicted in red.

Armed with a considerable knowledge about the binding modes of naphthalene sulfonamides, a second generation of sulfonamide inhibitors was developed. Instead of the d-glutamic acid moiety, they possessed cyclic, rigidified mimetics, as for example in compounds 25 and 26 (Scheme 7) (98). Incorporation of rigid mimetics at the position of d-glutamic acid improved the inhibitory activities against MurD when compared to the parent compounds, thereby confirming the advantage of conformational restriction. The co-crystal structure of MurD with rigidified inhibitor 25 revealed that both carboxyl groups occupied exactly the same binding sites as the carboxyl groups of d-glutamic acid (Figure 4) (98). They formed typical interactions with Ser415, Leu416 and Phe422 on one hand, and Lys348 and His183 on the other hand. The 2-cyano-4-fluoro-phenyl ring occupied the uracil-binding pocket of the nucleotide substrate with a hydrogen bond between the cyano group and Thr36. The central naphthalene moiety did not interact with the protein at all, and this represented a major opportunity for the improvement in the potency of future generations of sulfonamide inhibitors, which should interact with MurD also in this region.

Figure 4 Binding mode of naphthalene sulfonamide 25 containing a rigid glutamic acid mimetic (magenta; PDB entry: 2XPC) superimposed to its d-glutamic acid-containing counterpart 24 (orange; PDB entry: 2VTD).
Figure 4

Binding mode of naphthalene sulfonamide 25 containing a rigid glutamic acid mimetic (magenta; PDB entry: 2XPC) superimposed to its d-glutamic acid-containing counterpart 24 (orange; PDB entry: 2VTD).

The second generation of naphthalene sulfonamide inhibitors of MurD was investigated by NMR using 1H/13C heteronuclear single quantum correlation. Conformational and dynamic properties of the bound ligands and their binding interactions were examined using the transferred nuclear Overhauser effect and saturation transfer difference, while the binding mode was examined using unrestrained molecular dynamics simulations. The discoveries in this research shed light on the dynamic behavior of MurD/ligand complexes and its influence on contacts between both partners (99).

Amino acid sequence alignment of MurD enzymes from different bacterial species revealed that they do share conserved residues that are essential for the catalytic activity. However, the overall similarity is small. Considering this, it was not quite surprising that compounds that had been designed as inhibitors of MurD from E. coli turned out to be weaker inhibitors of other MurD orthologs (52). For example, compounds 15,16, 20 (Scheme 6) and 22, 24, 25, 26 (Scheme 7) either inhibited MurD enzymes from S. aureus, S. pneumoniae, B. burgdorferi and M. tuberculosis in the 0.1–1 mmol/l range or were not inhibitors at all. This divergent result could be explained by differences in amino acid sequences and topologies of the active sites of the MurD ligases in question (52).

Recently, computational efforts were devoted to finding new MurD inhibitors (100–103), which resulted in some new and potentially interesting scaffolds (104).

Expert opinion

In part because of its propensity to yield high-quality crystal structures, MurD from E. coli is the Mur ligase for which a great deal of mechanistic and structural data has been gathered. For the same reason, many MurD inhibitors belonging to different classes have been described, some being endowed with antibacterial activity. However, the fact that MurD orthologs from other pathogenic bacteria display less sensitivity to certain classes of inhibitors suggests that differences in active site topologies may be crucial for inhibitor recognition. Moreover, it remains to be demonstrated that the antibacterial activities already observed are not due to off-target activities. Therefore, we think that efforts should be devoted to the structural studies of MurD orthologs in order to find more powerful, panactive MurD inhibitors that could be developed into broad-spectrum antibacterial agents. Moreover, the best inhibitors should now be assayed for MurD inhibition in vivo to demonstrate that they indeed provoked an arrest of peptidoglycan synthesis in the susceptible species, which resulted from the depletion of the intracellular pool of UDP-MurNAc-pentapeptide. From cellular and physiological standpoints, the studies of protein complexes in which Mur ligases participate as well as the search for regulation mechanisms, which are still in their infancy, should be developed.

Outlook

The last two decades have seen major advances in the discovery of mechanistic and structural properties of MurD from E. coli. The study of MurD orthologs from important pathogenic organisms has started and should be pursued. Several classes of MurD inhibitors have been reported. Among the enzymes of peptidoglycan synthesis, the large amount of data now available makes MurD one of the best candidates as a target for new antibacterial agents. One of the most significant problems in the development of MurD inhibitors has been the lack of antibacterial activity of the majority of compounds synthesized so far. We believe that this is due to their poor penetration through the bacterial cell wall and/or to the rapid extrusion from the cell interior to the external environment via efflux pumps. A possible solution to this problem could be the design and synthesis of conjugates with siderophores, which would improve the penetration of inhibitors and enable them to accumulate within bacteria. The involvement of MurD (and other Mur ligases) in protein complexes and the deciphering of mechanisms of regulation are more fundamental topics that will undoubtedly be further developed.

Highlights

  • MurD catalyzes the ATP-dependent addition of d-glutamic acid to UDP-MurNAc-l-Ala, a reaction which involves acyl-phosphate and tetrahedral intermediates.

  • Several co-crystal structures of E. coli MurD with different ligands or inhibitors are available.

  • Several MurD orthologs from pathogenic bacteria have been purified and characterized.

  • Different classes of MurD inhibitors have been reported, some being endowed with antibacterial activity.

  • The search for panactive MurD inhibitors, for which the mode of action is clearly demonstrated, has to be performed now.

  • The study of cellular organization of Mur ligases and in vivo regulation of the MurD activity should be further developed.


Corresponding author: Didier Blanot, Laboratoire des Enveloppes Bactériennes et Antibiotiques, Institut de Biochimie et Biophysique Moléculaire et Cellulaire, UMR 8619 CNRS, Université Paris-Sud, F-91405 Orsay, France, e-mail:

About the authors

Roman Šink

Roman Šink is a post-doctoral researcher at the Faculty of Pharmacy of the University of Ljubljana. He was trained in organic and medicinal chemistry, enzymology and X-ray crystallography. He obtained his PhD degree in pharmaceutical sciences from the University of Ljubljana in 2010. His main research topics have been the design, synthesis, and study of the binding modes of peptidoglycan biosynthetic pathway inhibitors. Currently his research activity focuses on the design and synthesis of new antituberculotics.

Hélène Barreteau

Hélène Barreteau is assistant professor at the University Paris-Sud, Orsay, and a member of the Laboratory of Bacterial Envelopes and Antibiotics. She was trained in biochemistry and microbiology, and received her PhD degree in biochemistry from the University Picardie Jules-Verne, Amiens, in 2004. She is especially interested in the search for antimicrobial compounds and has been working on peptidoglycan as an antibacterial target for 8 years.

Delphine Patin

Delphine Patin was trained in genetics and molecular biology at the University Paris Diderot, Paris. She received a Diplôme d’Etudes Approfondies in human genetics from the University Paris Diderot in 1999. She joined the Laboratory of Bacterial Envelopes and Antibiotics as an engineer in 2007. Her main research interests have been the study of Mur ligases from different bacterial species and the search for inhibitors of these enzymes.

Dominique Mengin-Lecreulx

Dominique Mengin-Lecreulx was trained in biochemistry, enzymology, molecular biology and organic chemistry and obtained his PhD degree at the University Paris-Sud in 1987. He is director of research at CNRS and, since 2001, is the head of the Laboratory of Bacterial Envelopes and Antibiotics which, for more than 30 years, has been studying the biochemistry, genetics and physiology of the bacterial cell-wall peptidoglycan biosynthesis pathway as a whole. His main research interests concern the identification of genes and enzymes involved in this pathway, the regulation of this metabolism, the mode of action of antibiotics and bacteriocins interfering with this pathway and the corresponding resistance mechanisms, as well as the molecular bases for the recognition of bacterial cell-wall peptidoglycan by host innate immunity mechanisms.

Stanislav Gobec

Stanislav Gobec is professor of medicinal chemistry at the Faculty of Pharmacy, University of Ljubljana. He graduated in pharmacy and received his PhD degree in medicinal chemistry from the University of Ljubljana in 1999. His main research interests are the structure-based design, synthesis and evaluation of small-molecule enzyme inhibitors and receptor ligands. For more than 15 years he has been involved in discovery of inhibitors of bacterial cell-wall biosynthesis as potential antibacterial agents.

Didier Blanot

Didier Blanot is director of research at CNRS and a member of the Laboratory of Bacterial Envelopes and Antibiotics, University Paris-Sud, Orsay. He was trained in organic chemistry, biochemistry and enzymology, and received his PhD degree in natural sciences from the University Paris-Sud in 1979. For more than 30 years, his main research interests have been the functional and structural study of the enzymes of peptidoglycan biosynthesis, the determination of the structure of peptidoglycan from certain bacterial species, and the search for peptidoglycan biosynthesis inhibitors that could act as antibacterial agents.

This work was supported by the European Commission through the EUR-INTAFAR project (LSHM-CT-2004-512138); the Centre National de la Recherche Scientifique (UMR 8619 and PICS 3729); the Délégation Générale pour l’Armement (Contrats Jeune Chercheur 036000104 and 056000030); the Ministry of Education, Science and Sports of the Republic of Slovenia; the Franco-Slovene Proteus program; and the Institut Charles Nodier (Ljubljana). The authors would like to thank all of their colleagues, whose names are found in the references, for experimental work.

References

1. Vollmer W, Blanot D, de Pedro MA. Peptidoglycan structure and architecture. FEMS Microbiol Rev 2008; 32: 149–67.10.1111/j.1574-6976.2007.00094.xSearch in Google Scholar

2. Lovering AL, Safadi SS, Strynadka NC. Structural perspective of peptidoglycan biosynthesis and assembly. Annu Rev Biochem 2012; 81: 451–78.10.1146/annurev-biochem-061809-112742Search in Google Scholar

3. Barreteau H, Kovač A, Boniface A, Sova M, Gobec S, Blanot D. Cytoplasmic steps of peptidoglycan biosynthesis. FEMS Microbiol Rev 2008; 32: 168–207.10.1111/j.1574-6976.2008.00104.xSearch in Google Scholar

4. Bouhss A, Trunkfield AE, Bugg TD, Mengin-Lecreulx D. The biosynthesis of peptidoglycan lipid-linked intermediates. FEMS Microbiol Rev 2008; 32: 208–33.10.1111/j.1574-6976.2007.00089.xSearch in Google Scholar

5. Mohammadi T, van Dam V, Sijbrandi R, Vernet T, Zapun A, Bouhss A, Diepeveen-de Bruin M, Nguyen-Distèche M, de Kruijff B, Breukink E. Identification of FtsW as a transporter of lipid-linked cell wall precursors across the membrane. EMBO J 2011; 30: 1425–32.10.1038/emboj.2011.61Search in Google Scholar

6. Sauvage E, Kerff F, Terrak M, Ayala JA, Charlier P. The penicillin-binding proteins: structure and role in peptidoglycan biosynthesis. FEMS Microbiol Rev 2008; 32: 234–58.10.1111/j.1574-6976.2008.00105.xSearch in Google Scholar

7. Chopra I, Schofield C, Everett M, O’Neill A, Miller K, Wilcox M, Frère JM, Dawson M, Czaplewski L, Urleb U, Courvalin P. Treatment of health-care-associated infections caused by Gram-negative bacteria: a consensus statement. Lancet Infect Dis 2008; 8: 133–9.10.1016/S1473-3099(08)70018-5Search in Google Scholar

8. Bugg TD, Braddick D, Dowson CG, Roper DI. Bacterial cell wall assembly: still an attractive antibacterial target. Trends Biotechnol 2011; 29: 167–73.10.1016/j.tibtech.2010.12.006Search in Google Scholar PubMed

9. Hervé M, Boniface A, Gobec S, Blanot D, Mengin-Lecreulx D. Biochemical characterization and physiological properties of Escherichia coli UDP-N-acetylmuramate:l-alanyl-γ-d-glutamyl-meso-diaminopimelate ligase. J Bacteriol 2007; 189: 3987–95.10.1128/JB.00087-07Search in Google Scholar PubMed PubMed Central

10. Mengin-Lecreulx D, van Heijenoort J, Park JT. Identification of the mpl gene encoding UDP-N-acetylmuramate:l-alanyl-γ-d-glutamyl-meso-diaminopimelate ligase in Escherichia coli and its role in recycling of cell wall peptidoglycan. J Bacteriol 1996; 178: 5347–52.10.1128/jb.178.18.5347-5352.1996Search in Google Scholar PubMed PubMed Central

11. Bouhss A, Dementin S, van Heijenoort J, Parquet C, Blanot D. MurC and MurD synthetases of peptidoglycan biosynthesis: borohydride trapping of acyl-phosphate intermediates. Methods Enzymol 2002; 354: 189–96.10.1016/S0076-6879(02)54015-5Search in Google Scholar

12. Eveland SS, Pompliano DL, Anderson MS. Conditionally lethal Escherichia coli murein mutants contain point defects that map to regions conserved among murein and folyl poly-γ-glutamate ligases: identification of a ligase superfamily. Biochemistry 1997; 36: 6223–9.10.1021/bi9701078Search in Google Scholar PubMed

13. Bouhss A, Mengin-Lecreulx D, Blanot D, van Heijenoort J, Parquet C. Invariant amino acids in the Mur peptide synthetases of bacterial peptidoglycan synthesis and their modification by site-directed mutagenesis in the UDP-MurNAc:l-alanine ligase from Escherichia coli. Biochemistry 1997; 36: 11556–63.10.1021/bi970797fSearch in Google Scholar PubMed

14. Bouhss A, Dementin S, Parquet C, Mengin-Lecreulx D, Bertrand JA, Le Beller D, Dideberg O, van Heijenoort J, Blanot D. Role of the ortholog and paralog amino acid invariants in the active site of the UDP-MurNAc-l-alanine:d-glutamate ligase (MurD). Biochemistry 1999; 38: 12240–7.10.1021/bi990517rSearch in Google Scholar PubMed

15. Münch D, Roemer T, Lee SH, Engeser M, Sahl HG, Schneider T. Identification and invitro analysis of the GatD/MurT enzyme-complex catalyzing lipid II amidation in Staphylococcus aureus. PLoS Pathog 2012; 8: e1002509.10.1371/journal.ppat.1002509Search in Google Scholar PubMed PubMed Central

16. Figueiredo TA, Sobral RG, Ludovice AM, Almeida JM, Bui NK, Vollmer W, de Lencastre H, Tomasz A. Identification of genetic determinants and enzymes involved with the amidation of glutamic acid residues in the peptidoglycan of Staphylococcus aureus. PLoS Pathog 2012; 8: e1002508.10.1371/journal.ppat.1002508Search in Google Scholar PubMed PubMed Central

17. Sheng Y, Sun X, Shen Y, Bognar AL, Baker EN, Smith CA. Structural and functional similarities in the ADP-forming amide bond ligase superfamily: implications for a substrate-induced conformational change in folylpolyglutamate synthetase. J Mol Biol 2000; 302: 427–40.10.1006/jmbi.2000.3987Search in Google Scholar PubMed

18. Ziegler K, Diener A, Herpin C, Richter R, Deutzmann R, Lockau W. Molecular characterization of cyanophycin synthetase, the enzyme catalyzing the biosynthesis of the cyanobacterial reserve material multi-l-arginyl-poly-l-aspartate (cyanophycin). Eur J Biochem 1998; 254: 154–9.10.1046/j.1432-1327.1998.2540154.xSearch in Google Scholar PubMed

19. Candela T, Fouet A. Poly-gamma-glutamate in bacteria. Mol Microbiol 2006; 60: 1091–8.10.1111/j.1365-2958.2006.05179.xSearch in Google Scholar PubMed

20. Smith CA. Structure, function and dynamics in the mur family of bacterial cell wall ligases. J Mol Biol 2006; 362: 640–55.10.1016/j.jmb.2006.07.066Search in Google Scholar PubMed

21. Schleifer KH, Kandler O. Peptidoglycan types of bacterial cell walls and their taxonomic implications. Bacteriol Rev 1972; 36: 407–77.10.1128/br.36.4.407-477.1972Search in Google Scholar

22. Schleifer KH, Plapp R, Kandler O. Identification of threo-3-hydroxyglutamic acid in the cell wall of Microbacterium lacticum. Biochem Biophys Res Commun 1967; 28: 566–70.10.1016/0006-291X(67)90351-8Search in Google Scholar

23. Pratviel-Sosa F, Mengin-Lecreulx D, van Heijenoort J. Over-production, purification and properties of the uridine diphosphate N-acetylmuramoyl-l-alanine:d-glutamate ligase from Escherichia coli. Eur J Biochem 1991; 202: 1169–76.10.1111/j.1432-1033.1991.tb16486.xSearch in Google Scholar

24. Auger G, Martin L, Bertrand J, Ferrari P, Fanchon E, Vaganay S, Pétillot Y, van Heijenoort J, Blanot D, Dideberg O. Large-scale preparation, purification, and crystallization of UDP-N-acetylmuramoyl-l-alanine:d-glutamate ligase from Escherichia coli. Prot Express Purif 1998; 13: 23–9.10.1006/prep.1997.0850Search in Google Scholar

25. Walsh AW, Falk PJ, Thanassi J, Discotto L, Pucci MJ, Ho H-T. Comparison of the d-glutamate-adding enzymes from selected Gram-positive and Gram-negative bacteria. J Bacteriol 1999; 181: 5395–401.10.1128/JB.181.17.5395-5401.1999Search in Google Scholar

26. Pratviel-Sosa F, Acher F, Trigalo F, Blanot D, Azerad R, van Heijenoort J. Effect of various analogues of d-glutamic acid on the d-glutamate-adding enzyme from Escherichia coli. FEMS Microbiol Lett 1994; 115: 223–8.Search in Google Scholar

27. Kotnik M, Humljan J, Contreras-Martel C, Oblak M, Kristan K, Hervé M, Blanot D, Urleb U, Gobec S, Dessen A, Šolmajer T. Structural and functional characterization of enantiomeric glutamic acid derivatives as potential transition state analogue inhibitors of MurD ligase. J Mol Biol 2007; 370: 107–15.10.1016/j.jmb.2007.04.048Search in Google Scholar

28. Tomašić T, Zidar N, Rupnik V, Kovač A, Blanot D, Gobec S, Kikelj D, Peterlin Mašič L. Synthesis and biological evaluation of new glutamic acid-based inhibitors of MurD ligase. Bioorg Med Chem Lett 2009; 19: 153–7.10.1016/j.bmcl.2008.10.129Search in Google Scholar

29. Tomašić T, Kovač A, Simčič M, Blanot D, Golič Grdadolnik S, Gobec S, Kikelj D, Peterlin Mašič L. Novel 2-thioxothiazolidin-4-one inhibitors of bacterial MurD ligase targeting d-Glu- and diphosphate-binding sites. Eur J Med Chem 2011; 46: 3964–75.10.1016/j.ejmech.2011.05.070Search in Google Scholar

30. Michaud C, Blanot D, Flouret B, van Heijenoort J. Partial purification and specificity studies of the d-glutamate-adding and d-alanyl-d-alanine-adding enzymes from Escherichia coli K12. Eur J Biochem 1987; 166: 631–7.10.1111/j.1432-1033.1987.tb13560.xSearch in Google Scholar

31. Degani C, Boyer PD. A borohydride reduction method for characterization of the acyl phosphate linkage in proteins and its application to sarcoplasmic reticulum adenosine triphosphatase. J Biol Chem 1973; 248: 8222–6.10.1016/S0021-9258(19)43217-1Search in Google Scholar

32. Todhunter JA, Purich DL. Use of the sodium borohydride reduction technique to identify a γ-glutamyl phosphate intermediary in the Escherichia coli glutamine synthetase reaction. J Biol Chem 1975; 250: 3505–9.10.1016/S0021-9258(19)41543-3Search in Google Scholar

33. Bouhss A, Dementin S, van Heijenoort J, Parquet C, Blanot D. Formation of adenosine 5′-tetraphosphate from the acyl phosphate intermediate: a difference between the MurC and MurD synthetases of Escherichia coli. FEBS Lett 1999; 453: 15–9.10.1016/S0014-5793(99)00684-5Search in Google Scholar

34. Tanner MC, Vaganay S, van Heijenoort J, Blanot D. Phosphinate inhibitors of the d-glutamic acid-adding enzyme of peptidoglycan biosynthesis. J Org Chem 1996; 61: 1756–60.10.1021/jo951780aSearch in Google Scholar

35. Vaganay S, Tanner MC, van Heijenoort J, Blanot D. Study of the reaction mechanism of the d-glutamic acid-adding enzyme from Escherichia coli. Microb Drug Resist 1996; 2: 51–4.10.1089/mdr.1996.2.51Search in Google Scholar

36. Gegnas LD, Waddell ST, Chabin RM, Reddy S, Wong KK. Inhibitors of the bacterial cell wall biosynthesis enzyme MurD. Bioorg Med Chem Lett 1998; 8: 1643–8.10.1016/S0960-894X(98)00285-6Search in Google Scholar

37. Sun X, Cross JA, Bognar AL, Baker EN, Smith CA. Folate-binding triggers the activation of folylpolyglutamate synthetase. J Mol Biol 2001; 310: 1067–78.10.1006/jmbi.2001.4815Search in Google Scholar PubMed

38. Emanuele JJ, Jr., Jin H, Yanchunas J, Villafranca JJ. Evaluation of the kinetic mechanism of Escherichia coli uridine diphosphate-N-acetylmuramate:l-alanine ligase. Biochemistry 1997; 36: 7264–71.10.1021/bi970266rSearch in Google Scholar PubMed

39. Anderson MS, Eveland SS, Onishi HR, Pompliano DL. Kinetic mechanism of the Escherichia coli UDPMurNAc-tripeptide d-alanyl-d-alanine-adding enzyme: use of a glutathione S-transferase fusion. Biochemistry 1996; 35: 16264–9.10.1021/bi961872+Search in Google Scholar PubMed

40. Liger D, Masson A, Blanot D, van Heijenoort J, Parquet C. Study of the overproduced uridine-diphosphate-N-acetylmuramate:l-alanine ligase from Escherichia coli. Microb Drug Resist 1996; 2: 25–7.10.1089/mdr.1996.2.25Search in Google Scholar PubMed

41. Bertrand JA, Auger G, Fanchon E, Martin L, Blanot D, van Heijenoort J, Dideberg O. Crystal structure of UDP-N-acetylmuramoyl-l-alanine:d-glutamate ligase from Escherichia coli. EMBO J 1997; 16: 3416–25.10.1093/emboj/16.12.3416Search in Google Scholar PubMed PubMed Central

42. Bertrand JA, Auger G, Martin L, Fanchon E, Blanot D, Le Beller D, van Heijenoort J, Dideberg O. Determination of the MurD mechanism through crystallographic analysis of enzyme complexes. J Mol Biol 1999; 289: 579–90.10.1006/jmbi.1999.2800Search in Google Scholar PubMed

43. Bertrand JA, Fanchon E, Martin L, Chantalat L, Auger G, Blanot D, van Heijenoort J, Dideberg O. “Open” structures of MurD: domain movements and structural similarities with folylpolyglutamate synthetase. J Mol Biol 2000; 301: 1257–66.10.1006/jmbi.2000.3994Search in Google Scholar PubMed

44. Stein AJ, Sather A, Shakelford G, Joachimiak A. The crystal structure of UDP-N-acetylmuramoylalanine-d-glutamate (MurD) ligase from Streptococcus agalactiae to 1.5 Å. 2010; PDB entry: 3LK7 (http://www.rcsb.org/pdb/explore/explore.do?structureId=3LK7).10.2210/pdb3lk7/pdbSearch in Google Scholar

45. Favini-Stabile S, Contreras-Martel C, Thielens N, Dessen A. MreB and MurG as scaffolds for the cytoplasmic steps of peptidoglycan biosynthesis. Environ Microbiol 2013; doi:http://dx/doi.org/10.1111/1462–2920.12171.10.1111/1462-2920.12171Search in Google Scholar PubMed

46. Perdih A, Kotnik M, Hodošček M, Šolmajer T. Targeted molecular dynamics simulation studies of binding and conformational changes in E. coli MurD. Proteins 2007; 68: 243–54.10.1002/prot.21374Search in Google Scholar PubMed

47. Perdih A, Šolmajer T. MurD ligase from Escherichia coli: C-terminal domain closing motion. Comput Theor Chem 2012; 979: 73–81.10.1016/j.comptc.2011.10.018Search in Google Scholar

48. Perdih A, Hodošček M, Šolmajer T. MurD ligase from E. coli: tetrahedral intermediate formation study by hybrid quantum mechanical/molecular mechanical replica path method. Proteins 2009; 74: 744–59.10.1002/prot.22188Search in Google Scholar PubMed

49. Dementin S, Bouhss A, Auger G, Parquet C, Mengin-Lecreulx D, Dideberg O, van Heijenoort J, Blanot D. Evidence of a functional requirement for a carbamoylated lysine residue in MurD, MurE and MurF synthetases as established by chemical rescue experiments. Eur J Biochem 2001; 268: 5800–7.10.1046/j.0014-2956.2001.02524.xSearch in Google Scholar PubMed

50. Boniface A, Bouhss A, Mengin-Lecreulx D, Blanot D. The MurE synthetase from Thermotoga maritima is endowed with an unusual d-lysine adding activity. J Biol Chem 2006; 281: 15680–6.10.1074/jbc.M506311200Search in Google Scholar PubMed

51. Patin D, Boniface A, Kovač A, Hervé M, Dementin S, Barreteau H, Mengin-Lecreulx D, Blanot D. Purification and biochemical characterization of Mur ligases from Staphylococcus aureus. Biochimie 2010; 92: 1793–800.10.1016/j.biochi.2010.07.009Search in Google Scholar PubMed

52. Barreteau H, Sosič I, Turk S, Humljan J, Tomašić T, Zidar N, Hervé M, Boniface A, Peterlin-Mašič L, Kikelj D, Mengin-Lecreulx D, Gobec S, Blanot D. MurD enzymes from different bacteria: evaluation of inhibitors. Biochem Pharmacol 2012; 84: 625–32.10.1016/j.bcp.2012.06.006Search in Google Scholar PubMed

53. Mahapatra S, Yagi T, Belisle JT, Espinosa BJ, Hill PJ, McNeil MR, Brennan PJ, Crick DC. Mycobacterial lipid II is composed of a complex mixture of modified muramyl and peptide moieties linked to decaprenyl phosphate. J Bacteriol 2005; 187: 2747–57.10.1128/JB.187.8.2747-2757.2005Search in Google Scholar PubMed PubMed Central

54. Fletcher G, Irwin CA, Henson JM, Fillingim C, Malone MM, Walker JR. Identification of the Escherichia coli cell division gene sep and organization of the cell division-cell envelope genes in the sep-mur-ftsA-envA cluster as determined with specialized transducing lambda bacteriophages. J Bacteriol 1978; 133: 91–100.10.1128/jb.133.1.91-100.1978Search in Google Scholar PubMed PubMed Central

55. Miyakawa T, Matsuzawa H, Matsuhashi M, Sugino Y. Cell wall peptidoglycan mutants of Escherichia coli K-12: existence of two clusters of genes, mra and mrb, for cell wall peptidoglycan biosynthesis. J Bacteriol 1972; 112: 950–8.10.1128/jb.112.2.950-958.1972Search in Google Scholar PubMed PubMed Central

56. Wijsman HJ. A genetic map of several mutations affecting the mucopeptide layer of Escherichia coli. Genet Res 1972; 20: 65–74.10.1017/S0016672300013598Search in Google Scholar PubMed

57. Mengin-Lecreulx D, Parquet C, Desviat LR, Plá J, Flouret B, Ayala JA, van Heijenoort J. Organization of the murE-murG region of Escherichia coli: identification of the murD gene encoding the d-glutamic-acid-adding enzyme. J Bacteriol 1989; 171: 6126–34.10.1128/jb.171.11.6126-6134.1989Search in Google Scholar PubMed PubMed Central

58. Mengin-Lecreulx D, van Heijenoort J. Nucleotide sequence of the murD gene encoding the UDP-MurNAc-l-Ala-d-Glu synthetase of Escherichia coli. Nucleic Acids Res 1990; 18: 183.10.1093/nar/18.1.183Search in Google Scholar PubMed PubMed Central

59. Hara H, Yasuda S, Horiuchi K, Park JT. A promoter for the first nine genes of the Escherichia coli mra cluster of cell division and cell envelope biosynthesis genes, including ftsI and ftsW. J Bacteriol 1997; 179: 5802–11.10.1128/jb.179.18.5802-5811.1997Search in Google Scholar PubMed PubMed Central

60. Mengin-Lecreulx D, Ayala J, Bouhss A, van Heijenoort J, Parquet C, Hara H. Contribution of the Pmra promoter to expression of genes in the Escherichia coli mra cluster of cell envelope biosynthesis and cell division genes. J Bacteriol 1998; 180: 4406–12.10.1128/JB.180.17.4406-4412.1998Search in Google Scholar PubMed PubMed Central

61. Mingorance J, Tamames J, Vicente M. Genomic channeling in bacterial cell division. J Mol Recognit 2004; 17: 481–7.10.1002/jmr.718Search in Google Scholar PubMed

62. White CL, Kitich A, Gober JW. Positioning cell wall synthetic complexes by the bacterial morphogenetic proteins MreB and MreD. Mol Microbiol 2010; 76: 616–33.10.1111/j.1365-2958.2010.07108.xSearch in Google Scholar PubMed

63. Munshi T, Gupta A, Evangelopoulos D, Guzman JD, Gibbons S, Keep NH, Bhakta S. Characterisation of ATP-dependent Mur ligases involved in the biogenesis of cell wall peptidoglycan in Mycobacterium tuberculosis. PLoS One 2013; 8: e60143.10.1371/journal.pone.0060143Search in Google Scholar PubMed PubMed Central

64. Silver LL. Challenges of antibacterial discovery. Clin Microbiol Rev 2011; 24: 71–109.10.1128/CMR.00030-10Search in Google Scholar PubMed PubMed Central

65. Chopra I. The 2012 Garrod lecture: discovery of antibacterial drugs in the 21st century. J Antimicrob Chemother 2013; 68: 496–505.10.1093/jac/dks436Search in Google Scholar PubMed

66. Mengin-Lecreulx D, van Heijenoort J. Effect of growth conditions on peptidoglycan content and cytoplasmic steps of its biosynthesis in Escherichia coli. J Bacteriol 1985; 163: 208–12.10.1128/jb.163.1.208-212.1985Search in Google Scholar PubMed PubMed Central

67. de Roubin MR, Mengin-Lecreulx D, van Heijenoort J. Peptidoglycan biosynthesis in Escherichia coli: variations in the metabolism of alanine and d-alanyl-d-alanine. J Gen Microbiol 1992; 138: 1751–7.10.1099/00221287-138-8-1751Search in Google Scholar PubMed

68. Doublet P, van Heijenoort J, Mengin-Lecreulx D. Identification of the Escherichia coli murI gene, which is required for the biosynthesis of d-glutamic acid, a specific component of bacterial peptidoglycan. J Bacteriol 1992; 174: 5772–9.10.1128/jb.174.18.5772-5779.1992Search in Google Scholar PubMed PubMed Central

69. Mengin-Lecreulx D, Flouret B, van Heijenoort J. Cytoplasmic steps of peptidoglycan synthesis in Escherichia coli. J Bacteriol 1982; 151: 1109–17.10.1128/jb.151.3.1109-1117.1982Search in Google Scholar PubMed PubMed Central

70. Fiuza M, Canova MJ, Patin D, Letek M, Zanella-Cléon I, Becchi M, Mateos LM, Mengin-Lecreulx D, Molle V, Gil JA. The MurC ligase essential for peptidoglycan biosynthesis is regulated by the serine/threonine protein kinase PknA in Corynebacterium glutamicum. J Biol Chem 2008; 283: 36553–63.10.1074/jbc.M807175200Search in Google Scholar PubMed PubMed Central

71. Falk SP, Weisblum B. Phosphorylation of the Streptococcus pneumoniae cell wall biosynthesis enzyme MurC by a eukaryotic-like Ser/Thr kinase. FEMS Microbiol Lett 2013; 340: 19–23.10.1111/1574-6968.12067Search in Google Scholar PubMed

72. Thakur M, Chakraborti PK. Ability of PknA, a mycobacterial eukaryotic-type serine/threonine kinase, to transphosphorylate MurD, a ligase involved in the process of peptidoglycan biosynthesis. Biochem J 2008; 415: 27–33.10.1042/BJ20080234Search in Google Scholar

73. Nováková L, Sasková L, Pallová P, Janeček J, Novotná J, Ulrych A, Echenique J, Trombe MC, Branny P. Characterization of a eukaryotic type serine/threonine protein kinase and protein phosphatase of Streptococcus pneumoniae and identification of kinase substrates. FEBS J 2005; 272: 1243–54.10.1111/j.1742-4658.2005.04560.xSearch in Google Scholar

74. Parikh A, Verma SK, Khan S, Prakash B, Nandicoori VK. PknB-mediated phosphorylation of a novel substrate, N-acetylglucosamine-1-phosphate uridyltransferase, modulates its acetyltransferase activity. J Mol Biol 2009; 386: 451–64.10.1016/j.jmb.2008.12.031Search in Google Scholar

75. Dasgupta A, Datta P, Kundu M, Basu J. The serine/threonine kinase PknB of Mycobacterium tuberculosis phosphorylates PBPA, a penicillin-binding protein required for cell division. Microbiology 2006; 152: 493–504.10.1099/mic.0.28630-0Search in Google Scholar

76. Liebeke M, Meyer H, Donat S, Ohlsen K, Lalk M. A metabolomic view of Staphylococcus aureus and its Ser/Thr kinase and phosphatase deletion mutants: involvement in cell wall biosynthesis. Chem Biol 2010; 17: 820–30.10.1016/j.chembiol.2010.06.012Search in Google Scholar

77. Bratkovič T, Lunder M, Urleb U, Štrukelj B. Peptide inhibitors of MurD and MurE, essential enzymes of bacterial cell wall biosynthesis. J Basic Microbiol 2008; 48: 202–6.10.1002/jobm.200700133Search in Google Scholar

78. Paradis-Bleau C, Beaumont M, Boudreault L, Lloyd A, Sanschagrin F, Bugg TD, Levesque RC. Selection of peptide inhibitors against the Pseudomonas aeruginosa MurD cell wall enzyme. Peptides 2006; 27: 1693–700.10.1016/j.peptides.2006.01.017Search in Google Scholar

79. Horton JR, Bostock JM, Chopra I, Hesse L, Phillips SE, Adams DJ, Johnson AP, Fishwick CW. Macrocyclic inhibitors of the bacterial cell wall biosynthesis enzyme MurD. Bioorg Med Chem Lett 2003; 13: 1557–60.10.1016/S0960-894X(03)00176-8Search in Google Scholar

80. Li Z, Francisco GD, Hu W, Labthavikul P, Petersen PJ, Severin A, Singh G, Yang Y, Rasmussen BA, Lin YI, Skotnicki JS, Mansour TS. 2-Phenyl-5,6-dihydro-2H-thieno[3,2-c]pyrazol-3-ol derivatives as new inhibitors of bacterial cell wall biosynthesis. Bioorg Med Chem Lett 2003; 13: 2591–4.10.1016/S0960-894X(03)00471-2Search in Google Scholar

81. Antane S, Caufield CE, Hu W, Keeney D, Labthavikul P, Morris K, Naughton SM, Petersen PJ, Rasmussen BA, Singh G, Yang Y. Pulvinones as bacterial cell wall biosynthesis inhibitors. Bioorg Med Chem Lett 2006; 16: 176–80.10.1016/j.bmcl.2005.09.021Search in Google Scholar PubMed

82. Turk S, Kovač A, Boniface A, Bostock JM, Chopra I, Blanot D, Gobec S. Discovery of new inhibitors of the bacterial peptidoglycan biosynthesis enzymes MurD and MurF by structure-based virtual screening. Bioorg Med Chem Lett 2009; 17: 1884–9.10.1016/j.bmc.2009.01.052Search in Google Scholar PubMed

83. Šink R, Kovač A, Tomašić T, Rupnik V, Boniface A, Bostock J, Chopra I, Blanot D, Peterlin Mašič L, Gobec S, Zega A. Synthesis and biological evaluation of N-acylhydrazones as inhibitors of MurC and MurD ligases. ChemMedChem 2008; 3: 1362–70.10.1002/cmdc.200800087Search in Google Scholar PubMed

84. Frlan R, Kovač A, Blanot D, Gobec S, Pečar S, Obreza A. Design, synthesis and in vitro biochemical activity of novel amino acid sulfonohydrazide inhibitors of MurC. Acta Chim Slov 2011; 58: 295–310.Search in Google Scholar

85. Frlan R, Kovač A, Blanot D, Gobec S, Pečar S, Obreza A. Design and synthesis of novel N-benzylidenesulfonohydrazide inhibitors of MurC and MurD as potential antibacterial agents. Molecules 2008; 13: 11–30.10.3390/molecules13010011Search in Google Scholar PubMed PubMed Central

86. Tomašić T, Zidar N, Kovač A, Turk S, Simčič M, Blanot D, Müller-Premru M, Filipič M, Golič Grdadolnik S, Zega A, Anderluh M, Gobec S, Kikelj D, Peterlin Mašič L. 5-Benzylidenethiazolidin-4-ones as multitarget inhibitors of bacterial Mur ligases. ChemMedChem 2010; 5: 286–95.10.1002/cmdc.200900449Search in Google Scholar PubMed

87. Perdih A, Kovač A, Wolber G, Blanot D, Gobec S, Šolmajer T. Discovery of novel benzene 1,3-dicarboxylic acid inhibitors of bacterial MurD and MurE ligases by structure-based virtual screening approach. Bioorg Med Chem Lett 2009; 19: 2668–73.10.1016/j.bmcl.2009.03.141Search in Google Scholar PubMed

88. Štrancar K, Blanot D, Gobec S. Design, synthesis and structure-activity relationships of new phosphinate inhibitors of MurD. Bioorg Med Chem Lett 2006; 16: 343–8.10.1016/j.bmcl.2005.09.086Search in Google Scholar PubMed

89. Štrancar K, Boniface A, Blanot D, Gobec S. Phosphinate inhibitors of UDP-N-acetylmuramoyl-l-alanyl-d-glutamate:l-lysine ligase (MurE). Arch Pharm Chem Life Sci 2007; 340: 127–34.10.1002/ardp.200600191Search in Google Scholar PubMed

90. Zidar N, Tomašić T, Šink R, Rupnik V, Kovač A, Turk S, Patin D, Blanot D, Contreras Martel C, Dessen A, Müller Premru M, Zega A, Gobec S, Peterlin Mašič L, Kikelj D. Discovery of novel 5-benzylidenerhodanine and 5-benzylidenethiazolidine-2,4-dione inhibitors of MurD ligase. J Med Chem 2010; 53: 6584–94.10.1021/jm100285gSearch in Google Scholar PubMed

91. Tomašić T, Zidar N, Šink R, Kovač A, Blanot D, Contreras-Martel C, Dessen A, Müller-Premru M, Zega A, Gobec S, Kikelj D, Peterlin Mašič L. Structure-based design of a new series of d-glutamic acid based inhibitors of bacterial UDP-N-acetylmuramoyl-l-alanine:d-glutamate ligase (MurD). J Med Chem 2011; 54: 4600–10.10.1021/jm2002525Search in Google Scholar PubMed

92. Zidar N, Tomašić T, Šink R, Kovač A, Patin D, Blanot D, Contreras-Martel C, Dessen A, Müller Premru M, Zega A, Gobec S, Peterlin Mašič L, Kikelj D. New 5-benzylidenethiazolidin-4-one inhibitors of bacterial MurD ligase: design, synthesis, crystal structures, and biological evaluation. Eur J Med Chem 2011; 46: 5512–23.10.1016/j.ejmech.2011.09.017Search in Google Scholar PubMed

93. Tomašić T, Šink R, Zidar N, Fic A, Contreras Martel C, Dessen A, Patin D, Blanot D, Müller-Premru M, Gobec S, Zega A, Kikelj D, Peterlin Mašič L. Dual inhibitor of MurD and MurE ligases from Escherichia coli and Staphylococcus aureus. ACS Med Chem Lett 2012; 3: 626–30.10.1021/ml300047hSearch in Google Scholar PubMed PubMed Central

94. Humljan J, Kotnik M, Boniface A, Šolmajer T, Urleb U, Blanot D, Gobec S. A new approach towards peptidosulfonamides: synthesis of potential inhibitors of bacterial peptidoglycan biosynthesis enzymes MurD and MurE. Tetrahedron 2006; 62: 10980–8.10.1016/j.tet.2006.08.030Search in Google Scholar

95. Humljan J, Kotnik M, Contreras-Martel C, Blanot D, Urleb U, Dessen A, Šolmajer T, Gobec S. Novel naphthalene-N-sulfonyl-d-glutamic acid derivatives as inhibitors of MurD, a key peptidoglycan biosynthesis enzyme. J Med Chem 2008; 51: 7486–94.10.1021/jm800762uSearch in Google Scholar PubMed

96. Perdih A, Bren U, Šolmajer T. Binding free energy calculations of N-sulphonyl-glutamic acid inhibitors of MurD ligase. J Mol Model 2009; 15: 983–96.10.1007/s00894-009-0455-8Search in Google Scholar PubMed

97. Simčič M, Hodošček M, Humljan J, Kristan K, Urleb U, Kocjan D, Golič Grdadolnik S. NMR and molecular dynamics study of the binding mode of naphthalene-N-sulfonyl-d-glutamic acid derivatives: novel MurD ligase inhibitors. J Med Chem 2009; 52: 2899–908.10.1021/jm900117nSearch in Google Scholar PubMed

98. Sosič I, Barreteau H, Simčič M, Šink R, Cesar J, Zega A, Golič Grdadolnik S, Contreras-Martel C, Dessen A, Amoroso A, Joris B, Blanot D, Gobec S. Second-generation sulfonamide inhibitors of d-glutamic acid-adding enzyme: activity optimisation with conformationally rigid analogues of d-glutamic acid. Eur J Med Chem 2011; 46: 2880–94.10.1016/j.ejmech.2011.04.011Search in Google Scholar PubMed

99. Simčič M, Sosič I, Hodošček M, Barreteau H, Blanot D, Gobec S, Golič Grdadolnik S. The binding mode of second-generation sulfonamide inhibitors of MurD: clues for rational design of potent MurD inhibitors. PLoS One 2012; 7: e52817.10.1371/journal.pone.0052817Search in Google Scholar PubMed PubMed Central

100. Arvind A, Kumar V, Saravanan P, Mohan CG. Homology modeling, molecular dynamics and inhibitor binding study on MurD ligase of Mycobacterium tuberculosis. Interdiscip Sci Comput Life Sci 2012; 4: 223–38.10.1007/s12539-012-0133-xSearch in Google Scholar PubMed

101. Shanmugam A, Anbazhagan V, Natarajan J. Virtual screening of phenylsulfonamido-3-morpholonopropan-2-yl dihydrogen phosphate derivatives as novel inhibitors of MurC-MurF ligases from Mycobacterium leprae. Med Chem Res 2012; 21: 4341–51.10.1007/s00044-011-9958-9Search in Google Scholar

102. Shanmugam A, Natarajan J. Homology modeling and docking analyses of M. leprae Mur ligases reveals the common binding residues for structure based drug designing to eradicate leprosy. J Mol Model 2012; 18: 2659–72.10.1007/s00894-011-1285-zSearch in Google Scholar PubMed

103. Umamaheswari A, Pradhan D, Hemanthkumar M. Virtual screening for potential inhibitors of homology modeled Leptospira interrogans MurD ligase. J Chem Biol 2010; 3: 175–87.10.1007/s12154-010-0040-8Search in Google Scholar

104. Tomašić T, Kovač A, Klebe G, Blanot D, Gobec S, Kikelj D, Peterlin Mašič L. Virtual screening for potential inhibitors of bacterial MurC and MurD ligases. J Mol Model 2012; 18: 1063–72.10.1007/s00894-011-1139-8Search in Google Scholar

105. El-Sherbeini M, Geissler WM, Pittman J, Yuan X, Wong KK, Pompliano DL. Cloning and expression of Staphylococcus aureus and Streptococcus pyogenes murD genes encoding uridine diphosphate N-acetylmuramoyl-l-alanine:d-glutamate ligases. Gene 1998; 210: 117–25.10.1016/S0378-1119(98)00059-6Search in Google Scholar

106. Smith MC, Cook JA, Birch GM, Hitchcock SA, Peery RB, Hoskins J, Skatrud PL, Yao RC, Cox KL. Development of a high-throughput screen for Streptococcus pneumoniae UDP-N-acetylmuramoyl-alanine:d-glutamate ligase (MurD) for the identification of MurD inhibitors. In: Kirst HA, Yeh WK, Zmijewski MJ, Jr., editors. Enzyme Technologies for Pharmaceutical and Biotechnological Applications. New York, USA: Marcel Dekker, 2001: 298–306.Search in Google Scholar

107. El Zoeiby A, Sanschagrin F, Havugimana PC, Garnier A, Levesque RC. In vitro reconstruction of the biosynthetic pathway of peptidoglycan cytoplasmic precursor in Pseudomonas aeruginosa. FEMS Microbiol Lett 2001; 201: 229–35.10.1111/j.1574-6968.2001.tb10761.xSearch in Google Scholar PubMed

108. Azzolina BA, Yuan X, Anderson MS, El-Sherbeini M. The cell wall and cell division gene cluster in the mra operon of Pseudomonas aeruginosa: cloning, production, and purification of active enzymes. Protein Expr Purif 2001; 21: 393–400.10.1006/prep.2001.1390Search in Google Scholar PubMed

Received: 2013-8-6
Accepted: 2013-9-10
Published Online: 2013-10-29
Published in Print: 2013-12-01

©2013 by Walter de Gruyter Berlin Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 10.12.2023 from https://www.degruyter.com/document/doi/10.1515/bmc-2013-0024/html
Scroll to top button