首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
2.
A lysine racemase (lyr) gene was isolated from a soil metagenome by functional complementation for the first time by using Escherichia coli BCRC 51734 cells as the host and d-lysine as the selection agent. The lyr gene consisted of a 1,182-bp nucleotide sequence encoding a protein of 393 amino acids with a molecular mass of about 42.7 kDa. The enzyme exhibited higher specific activity toward lysine in the l-lysine-to-d-lysine direction than in the reverse reaction.Amino acids are the building blocks of proteins and play an important role in the regulation of the metabolism of living organisms. Among two enantiomers of naturally occurring amino acids, l-amino acids are predominant in living organisms, while d-amino acids are found in both free and bound states in various organisms like bacteria (36), yeasts (35), plants (47), insects (11), mammals (17), bivalves (39), and fish (28). The d-amino acids are mostly endogenous and produced by racemization from their counterparts by the action of a racemase. Thus, the amino acid racemases are involved in d-amino acid metabolism (29, 46). Since the discovery of alanine racemase in 1951 (42), several racemases toward amino acids, such as those for glutamate, threonine, serine, aspartate, methionine, proline, arginine, and phenylalanine, have been reported in bacteria, archaea, and eukaryotes, including mammals (1, 2, 15, 30, 31, 44). They are classified into two groups: pyridoxal 5′-phosphate (PLP)-dependent and PLP-independent enzymes (9, 36).Lysine racemase (Lyr, EC 5.1.1.5) was first reported in Proteus vulgaris ATCC 4669 (19) and proposed to be involved in the lysine degradation of bacterial cells (5, 19). Catabolism of lysine occurs via two parallel pathways. In one of the pathways, δ-aminovalerate is the key metabolite, whereas in the other l-lysine is racemized to d-lysine, and l-pipecolate and α-aminoadipate (AMA) are the key metabolites (5). d-Lysine catabolism proceeds through a series of cyclized intermediates which are necessary to regenerate an α-amino acid and comprise the following metabolites (AMA pathway): d-lysine→α-keto-ɛ-amino caproate→Δ1-piperideine-2-carboxylate→pipecolate→Δ1-piperideine-6-carboxylate→α-amino-δ-formylcaproate→α-AMA→α-ketoadipate (6, 7, 12, 27). The final product is converted to α-ketoglutarate via a series of coenzyme A derivatives and subsequently participates as an intermediate in the Krebs cycle. This pathway suggests that the biological function of d-lysine in the bacteria is that of a carbon or nitrogen source. Racemization of added l-lysine to d-lysine by whole cells of Proteus spp. and Escherichia spp. (19) and by the cell extract of Pseudomonas putida ATCC 15070 (5, 20) has been found. However, the enzyme has not been purified to homogeneity, and thus, its molecular and catalytic characteristics, including its gene structure, have not been elucidated. In this study, we explored a metagenomic library constructed from a garden soil to isolate a novel Lyr enzyme. After expression in Escherichia coli, the purified enzyme was characterized in terms of optimal pH and temperature, thermal stability, and racemization activity.  相似文献   

3.
4.
Prolyl dipeptide synthesis by S9 aminopeptidase from Streptomyces thermocyaneoviolaceus (S9AP-St) has been demonstrated. In the synthesis, S9AP-St preferentially used l-Pro-OBzl as the acyl donor, yielding synthesized dipeptides having an l-Pro-Xaa structure. In addition, S9AP-St showed broad specificity toward the acyl acceptor. Furthermore, S9AP-St produced cyclo (l-Pro-l-His) with a conversion ratio of substrate to cyclo (l-Pro-l-His) higher than 40%.Some proline-containing dipeptides and their cyclic analogs exhibit biological activity. For example, cyclo (l-arginyl-d-proline) [c(lR-dP)] is known to act as a specific inhibitor of family 18 chitinase (4, 10). A cyclic peptide, c(lP-lH), produced by the cleavage of thyrotropin-releasing hormone protects against oxidative stress, promotes cytoprotection (6, 7), and exhibits antihyperglycemic activity (11).Some serine peptidases exhibit peptide bond formation (i.e., aminolysis of esters, thioesters, and amides) in accordance with their hydrolytic activity (2, 14). The exchange of catalytic Ser for Cys to engineer the serine endopeptidase into “transpeptidase” for peptide bond formation has been well characterized (3, 5). Our recent approach confirmed the wide distribution of family S9 aminopeptidases that have catalytic Ser in actinomycetes (12). Of them, we obtained S9 aminopeptidase from Streptomyces thermocyaneoviolaceus NBRC14271 (S9AP-St). The enzyme was engineered into “transaminopeptidase” by exchange of catalytic Ser for Cys, and its aminolytic activity was evaluated (13). The engineered enzyme, designated as aminolysin-S, can synthesize hydrophobic dipeptides through an aminolysis reaction. However, aminolysin-S was unable to synthesize peptides containing proline. Although the report of aminolysin-S demonstrated that S9AP-St shows no aminolysis reaction toward limited substrates, details of its characteristics remain unknown. This study verified the peptide synthetic activity of S9AP-St, demonstrating that S9AP-St can synthesize widely varied prolyl dipeptides through an aminolysis reaction. The report also shows that S9AP-St is applicable to the synthesis of a biologically active peptide—c(lP-lH).  相似文献   

5.
The HIV gp41 N-trimer pocket region is an ideal viral target because it is extracellular, highly conserved, and essential for viral entry. Here, we report on the design of a pocket-specific d-peptide, PIE12-trimer, that is extraordinarily elusive to resistance and characterize its inhibitory and structural properties. d-Peptides (peptides composed of d-amino acids) are promising therapeutic agents due to their insensitivity to protease degradation. PIE12-trimer was designed using structure-guided mirror-image phage display and linker optimization and is the first d-peptide HIV entry inhibitor with the breadth and potency required for clinical use. PIE12-trimer has an ultrahigh affinity for the gp41 pocket, providing it with a reserve of binding energy (resistance capacitor) that yields a dramatically improved resistance profile compared to those of other fusion inhibitors. These results demonstrate that the gp41 pocket is an ideal drug target and establish PIE12-trimer as a leading anti-HIV antiviral candidate.The HIV envelope protein (Env) mediates viral entry into cells (11). Env is cleaved into surface (gp120) and transmembrane (gp41) subunits that remain noncovalently associated to form trimeric spikes on the virion surface (16). gp120 recognizes target cells by interacting with cellular receptors, while gp41 mediates membrane fusion. Peptides derived from heptad repeats near the N and C termini of the gp41 ectodomain (N and C peptides) interact in solution to form a six-helix bundle, representing the postfusion structure (3, 55, 56). In this structure, N peptides form a central trimeric coiled coil (N trimer), creating grooves into which C peptides bind. This structure, in conjunction with the dominant-negative inhibitory properties of exogenous N and C peptides, suggests a mechanism for Env-mediated entry (10, 22, 58-60).During entry, gp41 forms an extended prehairpin intermediate that leaves the exposed N-trimer region vulnerable to inhibition for several minutes (18, 35). This intermediate ultimately collapses as the C-peptide regions bind to the N-trimer grooves to form a trimer of hairpins (six-helix bundle), juxtaposing viral and cellular membranes and inducing fusion. Enfuvirtide (Fuzeon), the only clinically approved HIV fusion inhibitor, is a C peptide that binds to part of the N-trimer groove and prevents six-helix bundle formation in a dominant-negative manner (61). Enfuvirtide is active in patients with multidrug resistance to other classes of inhibitors and is a life-prolonging option for these patients (30, 31). However, enfuvirtide use is restricted to salvage therapy due to several limitations, including (i) high dosing requirements (90 mg, twice-daily injections), (ii) high cost (∼$30,000/year/patient in the United States), and (iii) the rapid emergence of resistant strains (21, 47).A deep hydrophobic pocket at the base of the N-trimer groove is an especially attractive inhibitory target because of its high degree of conservation (3, 12, 48), poor tolerance to substitution (4, 34), and critical role in membrane fusion (2). Indeed, this region is conserved at both the amino acid level (for gp41 function in membrane fusion) and the nucleotide level (for the structured RNA region of the Rev-responsive element). Enfuvirtide binds to the N-trimer groove just N terminal to the pocket and is significantly more susceptible to resistance mutations than 2nd-generation C-peptide inhibitors, such as T-1249, that also bind to the pocket (8, 13, 29, 44, 46, 47, 58).Peptide design, molecular modeling, and small-molecule screening have produced a diverse set of compounds that interact with the gp41 pocket and inhibit HIV-1 entry with modest potency, but often with significant cytotoxicity (7, 14, 15, 17, 23, 24, 26, 34, 51, 54). The first direct evidence that pocket-specific binders are sufficient to inhibit HIV entry came with the discovery of protease-resistant d-peptides identified using mirror-image phage display (12). In this technique, a phage library is screened against a mirror-image version of the target protein (synthesized using d-amino acids) (50). By symmetry, mirror images (d-peptides) of the discovered sequences will bind to the natural l-peptide target. As the mirror images of naturally occurring l-peptides, d-peptides cannot be digested by natural proteases. Protease resistance provides d-peptides theoretical treatment advantages of extended survival in the body and possible oral bioavailability (41, 42, 49).These 1st-generation d-peptide entry inhibitors possess potency against a laboratory-adapted isolate (HXB2) at low to mid-μM concentrations (12). We previously reported an affinity-matured 2nd-generation d-peptide called PIE7, pocket-specific inhibitor of entry 7 (57). A trimeric version of PIE7 is the first high-affinity pocket-specific HIV-1 inhibitor and has potency against X4-tropic (HXB2) and R5-tropic (BaL) strains at sub-nM concentrations. However, significant further optimization is required to create a robust clinical candidate for two reasons. First, this d-peptide is much less potent (requiring high nM concentrations) against JRFL, a primary R5-tropic strain. Therefore, improved PIE potency is necessary to combat diverse primary strains. Second, by improving the affinity of our inhibitors for the pocket target, we hope to provide a reserve of binding energy that will delay the emergence of drug resistance, as described below.We and others have reported a potency plateau for some gp41-based fusion inhibitors that is likely imposed by the transient exposure of the prehairpin intermediate (9, 27, 53, 57). For very high-affinity inhibitors, association kinetics (rather than affinity) limits potency so that two inhibitors with significantly different affinities for the prehairpin intermediate can have similar antiviral potencies. We proposed that overengineering our d-peptides with substantial affinity beyond this potency plateau would provide a reserve of binding energy that would combat affinity-disrupting resistance mutations (57). Such a resistance capacitor should also prevent the stepwise accumulation of subtle resistance mutations in Env by eliminating the selective advantage that such mutants would otherwise confer.Here, we report on the design and characterization of a 3rd-generation pocket-specific d-peptide, PIE12-trimer, with ∼100,000-fold improved target binding compared to that of the best previous d-peptide, significantly broadened inhibitory potency, and an enhanced resistance capacitor that provides a strong barrier to viral resistance. We achieved this increased potency via structure-guided phage display and crosslinker optimization. PIE12-trimer has a dramatically improved resistance profile compared to the profiles of earlier d-peptides, as well as those of enfuvirtide and T-1249. These results validate the resistance capacitor hypothesis and establish PIE12-trimer as a leading anti-HIV therapeutic candidate.  相似文献   

6.
The uncharacterized gene previously proposed as a mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in Escherichia coli. The maximal activity of the recombinant enzyme was observed at pH 7.5 and 40°C in the presence of 0.5 mM Co2+. The isomerization activity was specific for aldose substrates possessing hydroxyl groups oriented in the same direction at the C-2 and C-3 positions, such as the d and l forms of ribose, lyxose, talose, mannose, and allose. The enzyme exhibited the highest activity for l-ribulose among all pentoses and hexoses. Thus, l-ribose, as a potential starting material for many l-nucleoside-based pharmaceutical compounds, was produced at 213 g/liter from 300-g/liter l-ribulose by mannose-6-phosphate isomerase at 40°C for 3 h, with a conversion yield of 71% and a volumetric productivity of 71 g liter−1 h−1.l-Ribose is a potential starting material for the synthesis of many l-nucleoside-based pharmaceutical compounds, and it is not abundant in nature (5, 19). l-Ribose has been produced mainly by chemical synthesis from l-arabinose, l-xylose, d-glucose, d-galactose, d-ribose, or d-mannono-1,4-lactone (2, 17, 23). Biological l-ribose manufacture has been investigated using ribitol or l-ribulose. Recently, l-ribose was produced from ribitol by a recombinant Escherichia coli containing an NAD-dependent mannitol-1-dehydrogenase (MDH) with a 55% conversion yield when 100 g/liter ribitol was used in a 72-h fermentation (18). However, the volumetric productivity of l-ribose in the fermentation is 28-fold lower than that of the chemical method synthesized from l-arabinose (8). l-Ribulose has been biochemically converted from l-ribose using an l-ribose isomerase from an Acinetobacter sp. (9), an l-arabinose isomerase mutant from Escherichia coli (4), a d-xylose isomerase mutant from Actinoplanes missouriensis (14), and a d-lyxose isomerase from Cohnella laeviribosi (3), indicating that l-ribose can be produced from l-ribulose by these enzymes. However, the enzymatic production of l-ribulose is slow, and the enzymatic production of l-ribose from l-ribulose has been not reported.Sugar phosphate isomerases, such as ribose-5-phosphate isomerase, glucose-6-phosphate isomerase, and galactose-6-phosphate isomerase, work as general aldose-ketose isomerases and are useful tools for producing rare sugars, because they convert the substrate sugar phosphates and the substrate sugars without phosphate to have a similar configuration (11, 12, 21, 22). l-Ribose isomerase from an Acinetobacter sp. (9) and d-lyxose isomerase from C. laeviribosi (3) had activity with l-ribose, d-lyxose, and d-mannose. Thus, we can apply mannose-6-phosphate (EC 5.3.1.8) isomerase to the production of l-ribose, because there are no sugar phosphate isomerases relating to l-ribose and d-lyxose. The production of the expensive sugar l-ribose (bulk price, $1,000/kg) from the rare sugar l-ribulose by mannose-6-phosphate isomerase may prove to be a valuable industrial process, because we have produced l-ribulose from the cheap sugar l-arabinose (bulk price, $50/kg) using the l-arabinose isomerase from Geobacillus thermodenitrificans (20) (Fig. (Fig.11).Open in a separate windowFIG. 1.Schematic representation for the production of l-ribulose from l-arabinose by G. thermodenitrificans l-arabinose isomerase and the production of l-ribose from l-ribulose by B. subtilis mannose-6-phosphate isomerase.In this study, the gene encoding mannose-6-phosphate isomerase from Bacillus subtilis was cloned and expressed in E. coli. The substrate specificity of the recombinant enzyme for various aldoses and ketoses was investigated, and l-ribulose exhibited the highest activity among all pentoses and hexoses. Therefore, mannose-6-phosphate isomerase was applied to the production of l-ribose from l-ribulose.  相似文献   

7.
8.
Chlamydiae are unusual obligately intracellular bacteria that do not synthesize detectable peptidoglycan. However, they possess genes that appear to encode products with peptidoglycan biosynthetic activity. Bioinformatic analysis predicts that chlamydial MurE possesses UDP-MurNAc-l-Ala-d-Glu:meso-diaminopimelic acid (UDP-MurNAc-l-Ala-d-Glu:meso-A2pm) ligase activity. Nevertheless, there are no experimental data to confirm this hypothesis. In this paper we demonstrate that the murE gene from Chlamydia trachomatis is capable of complementing a conditional Escherichia coli mutant impaired in UDP-MurNAc-l-Ala-d-Glu:meso-A2pm ligase activity. Recombinant MurE from C. trachomatis (MurECt) was overproduced in and purified from E. coli in order to investigate its kinetic parameters in vitro. By use of UDP-MurNAc-l-Ala-d-Glu as the nucleotide substrate, MurECt demonstrated ATP-dependent meso-A2pm ligase activity with pH and magnesium ion optima of 8.6 and 30 mM, respectively. Other amino acids (meso-lanthionine, the ll and dd isomers of A2pm, d-lysine) were also recognized by MurECt. However, the activities for these amino acid substrates were weaker than that for meso-A2pm. The specificity of MurECt for three possible C. trachomatis peptidoglycan nucleotide substrates was also determined in order to deduce which amino acid might be present at the first position of the UDP-MurNAc-pentapeptide. Relative kcat/Km ratios for UDP-MurNAc-l-Ala-d-Glu, UDP-MurNAc-l-Ser-d-Glu, and UDP-MurNAc-Gly-d-Glu were 100, 115, and 27, respectively. Our results are consistent with the synthesis in chlamydiae of a UDP-MurNAc-pentapeptide in which the third amino acid is meso-A2pm. However, due to the lack of specificity of MurECt for nucleotide substrates in vitro, it is not obvious which amino acid is present at the first position of the pentapeptide.Chlamydiae cause serious respiratory tract and genital infections in humans (9). They are obligately intracellular gram-negative bacteria, with a unique biphasic development cycle. Elementary bodies (EBs) are the infectious form of the organism and invade susceptible host cells. Once internalized, EBs differentiate into reticulate bodies (RBs), which have the capacity to divide (39, 40). The RBs are fragile and pleomorphic, whereas EBs are comparatively rigid and stable (19, 39). After repeated cycles of binary fission, the RBs differentiate into EBs, and the host cell lyses, releasing infectious EBs (1).In contrast to the vast majority of eubacteria, chlamydiae lack detectable amounts of peptidoglycan (PG), an essential polymer. PG is a giant macromolecule composed of alternating N-acetylglucosamine (GlcNAc) and N-acetylmuramic acid (MurNAc) residues cross-linked by short peptides. It determines the shape of bacteria, protects the cell from lysis due to internal osmotic pressure, and also plays a role in cell division. However, PG has not been detected in EBs (14, 20) or RBs (5).Although chlamydiae appear to lack PG, they contain penicillin-binding proteins and are sensitive to antibiotics that inhibit PG synthesis (5, 40). The Chlamydia trachomatis genome contains most of the genes coding for proteins involved in, or associated with, PG synthesis (54). Chlamydial MurA, MurC-Ddl, and the MurC domain of the latter fusion protein are active in vitro and complement Escherichia coli mutants deficient in the respective enzymes (23, 32, 33). Furthermore, proteomic analysis reveals that the murE gene product, which was assigned as UDP-MurNAc-l-Ala-d-Glu:meso-diaminopimelic acid ligase (UDP-MurNAc-l-Ala-d-Glu:meso-A2pm ligase), is expressed in RBs (52).MurE ligases catalyze the addition of the third amino acid residue to the peptide chain of PG. This residue, generally a diamino acid, is usually meso-A2pm for gram-negative bacteria and bacilli, and l-lysine for gram-positive bacteria, although other amino acids (for example, l-ornithine, meso-lanthionine, ll-A2pm, l-diaminobutyric acid, or l-homoserine) occur in certain species (6, 50, 57). In many organisms, the third residue of the peptide chain participates in PG cross-linking; consequently, the MurE enzyme is highly specific for the relevant amino acid so as to avoid incorporation of incorrect amino acids into the macromolecule, which could result in deleterious morphological changes and cell lysis (35). Crystallization of MurE from E. coli (MurEEc) has permitted analysis of the structural basis for this high specificity (22). Sequence alignments of different MurE orthologues have also revealed the specific consensus sequences DNPR and D(D/N)P(N/A) located in the binding pockets for meso-A2pm and l-Lys, respectively (11, 17). Chlamydia trachomatis MurE (MurECt) possesses the DNPR motif, which suggests that it adds meso-A2pm (17). However, there are no experimental data to confirm this prediction.In this paper we report for the first time the overproduction and purification of MurECt, as well as a detailed investigation of its in vivo and in vitro biochemical properties. These studies contribute to our understanding of the nature and properties of the PG biosynthetic enzymes in chlamydiae and do indeed suggest that MurECt has meso-A2pm ligase activity.  相似文献   

9.
10.
A major question about cytokinesis concerns the role of the septin proteins, which localize to the division site in all animal and fungal cells but are essential for cytokinesis only in some cell types. For example, in Schizosaccharomyces pombe, four septins localize to the division site, but deletion of the four genes produces only a modest delay in cell separation. To ask if the S. pombe septins function redundantly in cytokinesis, we conducted a synthetic-lethal screen in a septin-deficient strain and identified seven mutations. One mutation affects Cdc4, a myosin light chain that is an essential component of the cytokinetic actomyosin ring. Five others cause frequent cell lysis during cell separation and map to two loci. These mutations and their dosage suppressors define a signaling pathway (including Rho1 and a novel arrestin) for repairing cell-wall damage. The seventh mutation affects the poorly understood RNA-binding protein Scw1 and severely delays cell separation when combined either with a septin mutation or with a mutation affecting the septin-interacting, anillin-like protein Mid2, suggesting that Scw1 functions in a pathway parallel to that of the septins. Taken together, our results suggest that the S. pombe septins participate redundantly in one or more pathways that cooperate with the actomyosin ring during cytokinesis and that a septin defect causes septum defects that can be repaired effectively only when the cell-integrity pathway is intact.THE fission yeast Schizosaccharomyces pombe provides an outstanding model system for studies of cytokinesis (McCollum and Gould 2001; Balasubramanian et al. 2004; Pollard and Wu 2010). As in most animal cells, successful cytokinesis in S. pombe requires an actomyosin ring (AMR). The AMR begins to assemble at the G2/M transition and involves the type II myosin heavy chains Myo2 and Myp2 and the light chains Cdc4 and Rlc1 (Wu et al. 2003). Myo2 and Cdc4 are essential for cytokinesis under all known conditions, Rlc1 is important at all temperatures but essential only at low temperatures, and Myp2 is essential only under stress conditions. As the AMR constricts, a septum of cell wall is formed between the daughter cells. The primary septum is sandwiched by secondary septa and subsequently digested to allow cell separation (Humbel et al. 2001; Sipiczki 2007). Because of the internal turgor pressure of the cells, the proper assembly and structural integrity of the septal layers are essential for cell survival.Septum formation involves the β-glucan synthases Bgs1/Cps1/Drc1, Bgs3, and Bgs4 (Ishiguro et al. 1997; Le Goff et al. 1999; Liu et al. 1999, 2002; Martín et al. 2003; Cortés et al. 2005) and the α-glucan synthase Ags1/Mok1 (Hochstenbach et al. 1998; Katayama et al. 1999). These synthases are regulated by the Rho GTPases Rho1 and Rho2 and the protein kinase C isoforms Pck1 and Pck2 (Arellano et al. 1996, 1997, 1999; Nakano et al. 1997; Hirata et al. 1998; Calonge et al. 2000; Sayers et al. 2000; Ma et al. 2006; Barba et al. 2008; García et al. 2009b). The Rho GTPases themselves appear to be regulated by both GTPase-activating proteins (GAPs) and guanine-nucleotide-exchange factors (GEFs) (Nakano et al. 2001; Calonge et al. 2003; Iwaki et al. 2003; Tajadura et al. 2004; Morrell-Falvey et al. 2005; Mutoh et al. 2005; García et al. 2006, 2009a,b). In addition, septum formation and AMR function appear to be interdependent. In the absence of a normal AMR, cells form aberrant septa and/or deposit septal materials at random locations, whereas a mutant defective in septum formation (bgs1) is also defective in AMR constriction (Gould and Simanis 1997; Le Goff et al. 1999; Liu et al. 1999, 2000). Both AMR constriction and septum formation also depend on the septation initiation network involving the small GTPase Spg1 (McCollum and Gould 2001; Krapp and Simanis 2008). Despite this considerable progress, many questions remain about the mechanisms and regulation of septum formation and its relationships to the function of the AMR.One major question concerns the role(s) of the septins. Proteins of this family are ubiquitous in fungal and animal cells and typically localize to the cell cortex, where they appear to serve as scaffolds and diffusion barriers for other proteins that participate in a wide variety of cellular processes (Longtine et al. 1996; Gladfelter et al. 2001; Hall et al. 2008; Caudron and Barral 2009). Despite the recent progress in elucidating the mechanisms of septin assembly (John et al. 2007; Sirajuddin et al. 2007; Bertin et al. 2008; McMurray and Thorner 2008), the details of septin function remain obscure. However, one prominent role of the septins and associated proteins is in cytokinesis. Septins concentrate at the division site in every cell type that has been examined, and in Saccharomyces cerevisiae (Hartwell 1971; Longtine et al. 1996; Lippincott et al. 2001; Dobbelaere and Barral 2004) and at least some Drosophila (Neufeld and Rubin 1994; Adam et al. 2000) and mammalian (Kinoshita et al. 1997; Surka et al. 2002) cell types, the septins are essential for cytokinesis. In S. cerevisiae, the septins are required for formation of the AMR (Bi et al. 1998; Lippincott and Li 1998). However, this cannot be their only role, because the AMR itself is not essential for cytokinesis in this organism (Bi et al. 1998; Korinek et al. 2000; Schmidt et al. 2002). Moreover, there is no evidence that the septins are necessary for AMR formation or function in any other organism. A further complication is that in some cell types, including most Caenorhabditis elegans cells (Nguyen et al. 2000; Maddox et al. 2007) and some Drosophila cells (Adam et al. 2000; Field et al. 2008), the septins do not appear to be essential for cytokinesis even though they localize to the division site.S. pombe has seven septins, four of which (Spn1, Spn2, Spn3, and Spn4) are expressed in vegetative cells and localize to the division site shortly before AMR constriction and septum formation (Longtine et al. 1996; Berlin et al. 2003; Tasto et al. 2003; Wu et al. 2003; An et al. 2004; Petit et al. 2005; Pan et al. 2007; Onishi et al. 2010). Spn1 and Spn4 appear to be the core members of the septin complex (An et al. 2004; McMurray and Thorner 2008), and mutants lacking either of these proteins do not assemble the others at the division site. Assembly of a normal septin ring also depends on the anillin-like protein Mid2, which colocalizes with the septins (Berlin et al. 2003; Tasto et al. 2003). Surprisingly, mutants lacking the septins are viable and form seemingly complete septa with approximately normal timing. These mutants do, however, display a variable delay in separation of the daughter cells, suggesting that the septins play some role(s) in the proper completion of the septum or in subsequent processes necessary for cell separation (Longtine et al. 1996; An et al. 2004; Martín-Cuadrado et al. 2005).It is possible that the septins localize to the division site and yet are nonessential for division in some cell types because their role is redundant with that of some other protein(s) or pathway(s). To explore this possibility in S. pombe, we screened for mutations that were lethal in combination with a lack of septins. The results suggest that the septins cooperate with the AMR during cytokinesis and that, in the absence of septin function, the septum is not formed properly, so that an intact system for recognizing and repairing cell-wall damage becomes critical for cell survival.  相似文献   

11.
12.
Halophilic archaea were found to contain in their cytoplasm millimolar concentrations of γ-glutamylcysteine (γGC) instead of glutathione. Previous analysis of the genome sequence of the archaeon Halobacterium sp. strain NRC-1 has indicated the presence of a sequence homologous to sequences known to encode the glutamate-cysteine ligase GshA. We report here the identification of the gshA gene in the extremely halophilic archaeon Haloferax volcanii and show that H. volcanii gshA directs in vivo the synthesis and accumulation of γGC. We also show that the H. volcanii gene when expressed in an Escherichia coli strain lacking functional GshA is able to restore synthesis of glutathione.Many organisms contain millimolar concentrations of low-molecular-weight thiol compounds that participate in a number of important biological functions involving thiol-disulfide exchanges (7). In particular, they serve to maintain an intracellular reducing environment, to provide reducing power for key reductive enzymes, to combat the effects of oxidative and disulfide stress, and to detoxify xenobiotic compounds (7). Glutathione (GSH), a cysteine-containing tripeptide, l-γ-glutamyl-l-cysteinylglycine, is the best-characterized low-molecular-weight thiol (7, 19, 21). GSH is made in a highly conserved two-step ATP-dependent process by two unrelated peptide bond-forming enzymes (3, 21). The γ-carboxyl group of l-glutamate and the amino group of l-cysteine are ligated by the enzyme glutamylcysteine (GC) ligase EC 6.3.2.2 (GshA, encoded by gshA), which is then condensed with glycine in a reaction catalyzed by GSH synthetase (GshB, encoded by gshB) to form GSH (10, 38). GSH is found primarily in gram-negative bacteria and eukaryotes and only rarely in gram-positive bacteria (26). Fahey and coworkers showed that GSH is absent from the high-GC gram-positive actinomycetes which produce, as the major low-molecular-weight thiol, mycothiol, 1-d-myo-inosityl-2-(N-acetyl-l-cysteinyl)-amido-2-deoxy-α-d-glucopyranoside (13, 26-28, 35). GSH is also absent in Archaea. In Pyrococcus furiosus, coenzyme A SH (CoASH) is the main thiol (11), whereas in Halobacterium salinarum, γGC is the predominant thiol and the organism possesses bis-γGC reductase activity (30, 36). Similarly, Leuconostoc kimchi and Leuconostoc mesenteroides, gram-positive lactic acid bacterial species, were recently found to contain γGC rather than GSH (15). To date, these are the sole procaryotic species reported to naturally produce γGC but not GSH (6, 30). In this report, we describe the identification of the gshA gene in the extremely halophilic archaeon Haloferax volcanii. Copley and Dhillon (6) previously identified, using bioinformatic tools, an open reading frame (ORF) (gene VNG1397C) in Halobacterium sp. strain NRC-1 with limited sequence relatedness to known GshA proteins (6). However, no genetic or biochemical evidence was presented to substantiate their conclusion. Here, we show that Haloferax volcanii strain DS2 (1, 25) contains an ORF that directs in vivo the synthesis and accumulation of γGC. We also show that the H. volcanii ORF, when expressed in Escherichia coli lacking functional GshA, is able to restore synthesis of GSH.  相似文献   

13.
The vancomycin-resistant Staphylococcus aureus VRSA-9 clinical isolate was partially dependent on glycopeptide for growth. The responsible vanA operon had the same organization as that of Tn1546 and was located on a plasmid. The chromosomal d-Ala:d-Ala ligase (ddl) gene had two point mutations that led to Q260K and A283E substitutions, resulting in a 200-fold decrease in enzymatic activity compared to that of the wild-type strain VRSA-6. To gain insight into the mechanism of enzyme impairment, we determined the crystal structure of VRSA-9 Ddl and showed that the A283E mutation induces new ion pair/hydrogen bond interactions, leading to an asymmetric rearrangement of side chains in the dimer interface. The Q260K substitution is located in an exposed external loop and did not induce any significant conformational change. The VRSA-9 strain was susceptible to oxacillin due to synthesis of pentadepsipeptide precursors ending in d-alanyl-d-lactate which are not substrates for the β-lactam-resistant penicillin binding protein PBP2′. Comparison with the partially vancomycin-dependent VRSA-7, whose Ddl is 5-fold less efficient than that of VRSA-9, indicated that the levels of vancomycin dependence and susceptibility to β-lactams correlate with the degree of Ddl impairment. Ddl drug targeting could therefore be an effective strategy against vancomycin-resistant S. aureus.Methicillin-resistant Staphylococcus aureus (MRSA) bacteria that have acquired the vancomycin resistance vanA operon from glycopeptide-resistant enterococci are designated vancomycin-resistant S. aureus (VRSA) (29). Vancomycin acts by binding to the C-terminal acyl-d-Ala-d-Ala of the undecaprenol-diphosphate MurNAc-pentapeptide intermediate and inhibits transglycosylation and transpeptidation reactions in cell wall peptidoglycan polymerization and cross-linking (30). d-Ala-d-Ala is synthesized by the ATP-dependent d-Ala:d-Ala ligase (Ddl) (EC 6.3.2.4) before its incorporation in peptidoglycan precursors (26, 35). VanA-type vancomycin resistance results from the incorporation into peptidoglycan intermediates of a d-alanyl-d-lactate (d-Ala-d-Lac) depsipeptide, synthesized by a d-Ala:d-Lac ligase, which is responsible for diminished binding affinity of glycopeptides for their target. Kinetic analyses of Ddls have established two subsites in the active site for d-Ala binding (24, 27). The reaction mechanism culminates in the transfer of the γ-phosphoryl of ATP to the carboxyl group of d-Ala1 to produce an acylphosphate and ADP. The acyl carbon atom of the acylphosphate then reacts with the amino group of d-Ala2 to yield a tetrahedral intermediate. Finally, the intermediate releases phosphate to yield d-Ala-d-Ala.Mutants of Enterococcus faecium (8, 14), Enterococcus faecalis (34), and S. aureus (23) with an impaired Ddl are able to grow because they use the vancomycin resistance pathway for cell wall synthesis. Since resistance is inducible by the drug, these bacteria require the presence of vancomycin in the culture medium for growth. Ddls from vancomycin-dependent enterococci (14) have mutations affecting amino acids highly conserved in the d-Ala:d-Ala ligase superfamily (10). Molecular modeling based on the X-ray structure of Escherichia coli DdlB (11) revealed that all the mutated residues interact directly with one of the substrates of the enzymatic reaction or stabilize the position of critical residues in the active site. However, the degree of enzyme impairment was not evaluated biochemically. Recently, we reported the mechanism of vancomycin dependence in VanA-type S. aureus VRSA-7 and showed that the chromosomal Ddl had the single mutation N308K, which probably affects the binding of the transition-state intermediate, leading to a 1,000-fold decrease in activity relative to that of the wild-type enzyme (23). Glycopeptide-dependent mutants could therefore be considered useful tools to explore structure-activity relationships of the Ddl, which represents an attractive target for designing new drugs. Here we describe the partially vancomycin-dependent VanA-type S. aureus strain VRSA-9 and report the biochemical and structural characterization of its mutated Ddl.  相似文献   

14.
In order to achieve direct and efficient fermentation of optically pure d-lactic acid from raw corn starch, we constructed l-lactate dehydrogenase gene (ldhL1)-deficient Lactobacillus plantarum and introduced a plasmid encoding Streptococcus bovis 148 α-amylase (AmyA). The resulting strain produced only d-lactic acid from glucose and successfully expressed amyA. With the aid of secreting AmyA, direct d-lactic acid fermentation from raw corn starch was accomplished. After 48 h of fermentation, 73.2 g/liter of lactic acid was produced with a high yield (0.85 g per g of consumed sugar) and an optical purity of 99.6%. Moreover, a strain replacing the ldhL1 gene with an amyA-secreting expression cassette was constructed. Using this strain, direct d-lactic acid fermentation from raw corn starch was accomplished in the absence of selective pressure by antibiotics. This is the first report of direct d-lactic acid fermentation from raw starch.Poly-lactic acid (PLA) is an important agro-based plastic that can be produced from inexpensive, renewable, and abundantly available biomass resources, including starchy materials. These resources have advantages over limited oil- and fossil-based sources, as they do not result in any net carbon dioxide release to the atmosphere (7). Recently, stereocomplex PLA, which is composed of both poly-l- and -d-lactic acid, has been attracting much attention due to its high thermostability. Stereocomplex-type polymers show a melting point (ca. 230°C) that is approximately 50°C higher than that of the respective single polymers (8). Therefore, d-lactic acid, in addition to l-lactic acid, which has been the focus of production to date, is of significant importance.Lactic acid bacteria (LAB) are promising microorganisms for the efficient production of lactic acid from various sugars, such as glucose, sucrose, and lactose. However, when starchy materials are used as a carbon source, they must be saccharified by physicochemical and enzymatic treatment because most LAB cannot utilize starchy materials directly (13). This makes the whole process less economically viable. Therefore, many researchers have examined the direct production of lactic acid from starchy materials by using wild amylolytic LAB (ALAB) (6, 24, 25) or genetically modified amylase-producing LAB (15, 16). Although d-lactic acid has been produced by fermentation from pretreated substrates such as rice starch (5) and by simultaneous saccharification and fermentation from cellulose (23), there have been no reports on the direct production of d-lactic acid from starchy materials. This is due to a lack of d-lactic acid-producing ALAB and difficulties in gene manipulation of d-lactic acid-producing LAB, such as Lactobacillus delbrueckii (22).We focused on Lactobacillus plantarum, which is an industrially important strain due to its environmental flexibility and its ability to assimilate a wide range of carbohydrates (9). In recent years, several gene manipulation methods for Lactobacillus plantarum have been established (18, 19). Moreover, the complete genome sequence has been decoded for L. plantarum NCIMB 8826 (9). Based on whole-genome analysis, L. plantarum possesses two types of lactate dehydrogenase (LDH), l-LDH and d-LDH, which convert pyruvate into l- and d-lactic acid, respectively. Ferain et al. (4) reported that chromosomal deletion in the ldhL1 gene of L. plantarum NCIMB 8826 provoked an absence of l-LDH activity and produced d-lactic acid from glucose.In the present study, to produce d-lactic acid directly from starch, we constructed an l-LDH-deficient, α-amylase-secreting L. plantarum strain. The engineered strain expressed α-amylase from Streptococcus bovis 148 (AmyA) (20) and efficiently degraded raw starch with the aid of a C-terminal starch-binding domain (11). Using this strain, we achieved the direct and efficient fermentation of optically pure d-lactic acid from raw corn starch.  相似文献   

15.
Clostridium sordellii is a spore-forming, obligately anaerobic, Gram-positive bacterium that can cause toxic shock syndrome after gynecological procedures. Although the incidence of C. sordellii infection is low, it is fatal in most cases. Since spore germination is believed to be the first step in the establishment of Bacilli and Clostridia infections, we analyzed the requirements for C. sordellii spore germination in vitro. Our data showed that C. sordellii spores require three structurally different amino acids and bicarbonate for maximum germination. Unlike the case for Bacilli species, d-alanine had no effect on C. sordellii spore germination. C. sordellii spores germinated only in a narrow pH range between 5.7 and 6.5. In contrast, C. sordellii spore germination was significantly less sensitive to temperature changes than that of the Bacilli. The analysis of the kinetics of C. sordellii spore germination showed strong allosteric behavior in the binding of l-phenylalanine and l-alanine but not in that of bicarbonate or l-arginine. By comparing germinant apparent binding affinities to their known in vivo concentrations, we postulated a mechanism for differential C. sordellii spore activation in the female reproductive tract.Clostridium sordellii is an anaerobic, Gram-positive, spore-forming bacterium that is commonly found in soil and in the intestines of animals (4). Many C. sordellii strains are nonpathogenic; however, virulent strains cause lethal infections in several animal species, such as hemorrhagic enteritis in foals, sheep, and cattle (5, 10, 16, 28), omphalitis in foals (43), and wound infection in humans (4, 35).C. sordellii also can cause life-threatening necrotizing infections after gynecological procedures (4). In addition, fatal cases of C. sordellii endometritis following medical abortion with a mifepristone-misoprostol combination have been reported recently (13, 19, 56). The increased use of mifepristone-misoprostol for medical abortion may result in larger numbers of C. sordellii infections (38, 40).Although C. sordellii rarely has been identified in the genital tract, a correlation between gynecological procedures and C. sordellii-mediated toxic shock syndrome is apparent (19). Pregnancy, childbirth, or abortion may predispose some women to acquire C. sordellii in the vaginal tract (19). Under these conditions, C. sordellii infections result in an almost 100% mortality rate.Since there is no national system for tracking and reporting complications associated with gynecological procedures, the identification of the true rates of reproductive tract infections in women is not readily available (8). Therefore, the number of known C. sordellii-associated infections, although low, may be underreported (19, 29). Furthermore, unsafe abortion practices in developing countries cause large mortality rates due to complicating infections (24, 34). In many cases, however, the causative agent of the abortion-associated sepsis have not been characterized (24). Thus, the worldwide morbidity and mortality associated with C. sordellii infections is not currently known.C. sordellii produces several virulence factors. The two major toxins are the lethal toxin (TcsL) and the hemorrhagic toxin (37, 46). The lethal toxin produced by C. sordellii is causally involved in enteritis of domestic animals and in systemic toxicity following infections of humans (46). Furthermore, TcsL is associated with rapid mortality in C. sordellii endometritis rodent models (26). Interestingly, TcsL cytopathic effects are increased at low pH, a characteristic found in the vaginal tract (48). The hemorrhagic toxin is not well characterized, but it has been reported to cause dermal and intestinal necrosis in guinea pigs (6, 52).C. sordellii, like other Bacilli and Clostridia species, has the ability to form metabolically dormant spores that are extremely resistant to environmental stresses, such as heat, radiation, and toxic chemicals (42, 55). Upon encountering a suitable environment, spores germinate into vegetative cells, the form that is responsible for toxin production and disease onset (39, 54).In most cases, the germination process initially is triggered by the detection of low-molecular-weight germinants by a sensitive biosensor (39, 54). This sensor consists of a proteinaceous germination (Ger) receptor encoded, in general, by a tricistronic operon. Spore germination requirements have been studied most extensively for Bacilli and can be initiated by a variety of factors, including amino acids, sugars, and nucleosides (20, 30).Spore germination in the Clostridia generally requires combinations of multiple germinants. The germination of spores of proteolytic Clostridium botulinum types A and B was triggered by a defined three-component mixture comprised of l-alanine (or l-cysteine), l-lactate (or sodium thioglycolate), and sodium bicarbonate (3). In contrast, the optimum germination of spores of nonproteolytic C. botulinum types B, E, and F required binary combinations of l-alanine-l-lactate, l-cysteine-l-lactate, and l-serine-l-lactate (45).Clostridium difficile is a human pathogen that can cause fulminant colitis (11). Interestingly, C. difficile does not encode any known Ger receptors (53). However, it is likely that germination receptors exist, because C. difficile spores must germinate in order to complete their life cycle. While C. difficile germination receptors remain elusive, the spores of C. difficile germinate in rich medium supplemented with bile salts (62). More recently, taurocholate (a bile salt) and glycine (an amino acid) were shown to act as cogerminants for C. difficile spore germination (57, 61).Clostridium bifermentans is a close relative of C. sordellii (14). The minimum requirement for C. bifermentans spore germination was the presence of l-alanine, l-phenylalanine, and l-lactate (59). In addition, an unknown factor present in yeast extract was suggested to enhance germination (59). However, the Ger receptors involved in C. bifermentans spore germination are not known.Even though many Bacilli and Clostridia species use similar metabolites as germinants, the mechanisms of germinant recognition remain to be elucidated. Unfortunately, the multimeric interactions of Ger receptor complexes and the hydrophobic nature of the Ger receptor subunits have hindered our understanding of the mechanism of germinant recognition.To understand the molecular determinants of germinant recognition, we recently applied kinetic methods to study bacterial spore germination (1, 2, 18). Spore germination can be analyzed quantitatively by fitting optical density (OD) decreases to the Michaelis-Menten equation (2). The kinetic parameters obtained allow the determination of the apparent binding affinity (Km) of spores for the different cogerminants and the maximum rate of spore germination (Vmax). In these instances, Km refers to the concentration of substrate required to reach half of the maximal germination rate. These parameters can, in turn, be used to determine the mechanism of germination and potential interactions between germination receptors. Furthermore, by comparing apparent Km values to germinant concentrations in vivo, models for spore-germinant complex distribution can be proposed, and rate-limiting steps for the germination process can be derived. Thus, kinetic analysis can yield information on spore activation even if the identities of the germination receptors are not known.Using this procedure, we were able to determine the mechanism for Bacillus anthracis germination with inosine and l-alanine. In turn, this information was used to design nucleoside analogs that inhibit B. anthracis spore germination in vitro and protect macrophages from anthrax cytotoxicity (2).Since C. sordellii germination receptors have not been identified, we used chemical probes and kinetic methods to investigate the conditions necessary for spore germination. We found that C. sordellii spores germinate better at slightly acidic pH. Furthermore, germination rates varied slightly from 25 to 40°C. We also found that C. sordellii spores have an absolute requirement for a small amino acid, a basic amino acid, an aromatic amino acid, and bicarbonate (NaHCO3) for efficient germination. Kinetic analysis showed allosteric interaction for the putative l-phenylalanine and l-alanine germination receptors. In contrast, l-arginine or bicarbonate recognition followed typical Michaelis-Menten kinetics. The implication of germinant recognition and host environment is discussed.  相似文献   

16.
Class A penicillin-binding proteins (PBPs) are large, bifunctional proteins that are responsible for glycan chain assembly and peptide cross-linking of bacterial peptidoglycan. Bacteria in the genus Mycobacterium have been reported to have only two class A PBPs, PonA1 and PonA2, that are encoded in their genomes. We report here that the genomes of Mycobacterium smegmatis and other soil mycobacteria contain an additional gene encoding a third class A penicillin-binding protein, PonA3, which is a paralog of PonA2. Both the PonA2 and PonA3 proteins contain a penicillin-binding protein and serine/threonine protein kinase-associated (PASTA) domain that we propose may be involved in sensing the cell cycle and a C-terminal proline-rich region (PRR) that may have a role in protein-protein or protein-carbohydrate interactions. We show here that an M. smegmatis ΔponA2 mutant has an unusual antibiotic susceptibility profile, exhibits a spherical morphology and an altered cell surface in stationary phase, and is defective for stationary-phase survival and recovery from anaerobic culture. In contrast, a ΔponA3 mutant has no discernible phenotype under laboratory conditions. We demonstrate that PonA2 and PonA3 can bind penicillin and that PonA3 can partially substitute for PonA2 when ponA3 is expressed from a constitutive promoter on a multicopy plasmid. Our studies suggest that PonA2 is involved in adaptation to periods of nonreplication in response to starvation or anaerobiosis and that PonA3 may have a similar role. However, the regulation of PonA3 is likely different, suggesting that its importance could be related to stresses encountered in the environmental niches occupied by M. smegmatis and other soil-dwelling mycobacteria.The cell envelope of mycobacteria is a complex carbohydrate- and lipid-rich entity and is a major factor contributing to the success of these organisms as saprophytic and pathogenic bacteria (7, 8, 29, 35). The innermost layer of the cell envelope is a peptidoglycan (PG) composed of N-acylmuramic acid and N-acetylglucosamine with l-alanyl (or glycyl in the case of Mycobacterium leprae)-d-isoglutaminyl-meso-diaminopimelyl-d-alanyl-d-alanine pentapeptides attached to the muramic acid residues (13, 16, 54). While some of the muramyl residues are N acetylated, as they are in most other bacteria, a majority of the muramyl residues are N glycolylated (2, 37, 48, 49), a modification that confers lysozyme resistance (53) and also influences the innate immune response to mycobacterial cell walls (10). The pentapeptide chains of the mycobacterial PG can be modified by amidation, glycylation, or methylation, but the functional significance of these modifications is unknown (28, 31, 32, 38, 54).Approximately 80% of the pentapeptides in mycobacterial PG are cross-linked, and a majority of the links are between the carboxyl group of a penultimate d-Ala residue in a pentapeptide precursor and the amino group of the side chain d center of a meso-diaminopimelic acid (DAP) residue from an adjacent peptide (referred to as a 4-3 cross-link), while approximately one-third of the links are between the carboxyl group of the l center of a DAP residue of one peptide and the amino group of the side chain d center of the DAP residue in an adjacent peptide (referred to as a 3-3 cross-link) (17, 65). The 4-3 linkage is considered the “standard” linkage and is catalyzed by classical, penicillin-sensitive dd-transpeptidases, while the novel 3-3 linkage is thought to be catalyzed by the concerted action of dd-carboxypeptidases and novel ld-transpeptidases (31, 34, 39-41). The reasons why bacteria produce both 4-3 and 3-3 linkages are unknown. Some workers have suggested that the 3-3 linkages might reinforce the wall during times of stress and under nonreplicating conditions or stabilize complex cell envelopes (17, 50, 51, 55). In this regard, the high percentage of 3-3 linkages found in the PG of mycobacteria and their predominance in stationary-phase M. tuberculosis cells (31) suggest that these linkages may have an important role in maintaining cell envelope integrity during periods of growth and under nonreplicating conditions.The enzymes involved in peptidoglycan assembly, the penicillin-binding proteins (PBPs), have a triad sequence motif that forms the transpeptidation active site ([SxxK]——[S/YxN/C]——[K/H][T/S]G), which is the target of the β-lactam class of antibiotics (for a review, see reference 18). The PBPs have been grouped into several classes based on this motif, surrounding sequences, and other structural features (18). Of interest here are the class A PBPs, which are high-molecular-weight (HMW) proteins with both a transglycosylase domain (also called a non-penicillin-binding module [n-PB]) and a transpeptidase domain (also called a penicillin-binding module [PB]) (18). These proteins are tethered to the cytoplasmic membrane by a transmembrane helix with the catalytic domains facing the outside of the cell. Mycobacteria have been reported to have only two genes that encode class A PBPs, ponA1 and ponA2, which are annotated Rv0051 and Rv3682 in the sequence genome of M. tuberculosis H37Rv (9, 17). Previous studies that analyzed collections of transposon mutants to obtain clones with various phenotypes identified strains with insertions in these two genes. The phenotypes of these mutants have clearly shown that these PBPs play a complex role in mycobacterial physiology. One group of workers found a ponA1 mutant of M. smegmatis in a search for mutants with an altered dye-binding phenotype (an indicator of changes in the cell envelope) and showed that this slowly growing mutant was hypersusceptible to β-lactam antibiotics and had altered permeability (6). A ponA2 mutant of M. smegmatis was discovered in a screen for mutants defective for survival during long-term culture (25), while other workers isolated an M. tuberculosis ponA2 mutant in a screen for mutants sensitive to low pH (61). The same group of workers also showed that the M. tuberculosis mutant was more sensitive to heat, H2O2, and NO and was attenuated for persistence in the mouse model of inhalation tuberculosis (62). We previously identified an M. tuberculosis ponA2 mutant in a screen for mutants hypersusceptible to β-lactam antibiotics (14). All of these studies identified transposon mutants in searches for mutants with specific phenotypes, but there have been no direct genetic studies that have specifically examined the function of these PBPs in peptidoglycan metabolism.In this study we demonstrated that M. smegmatis has three class A PBPs. We show here that a newly recognized protein, which we designated PonA3, is a paralog of the PonA2 protein and is found only in certain environmental species of mycobacteria. We analyzed the phenotypes of M. smegmatis mutants with in-frame deletions of ponA2 and ponA3 singly and in combination to increase our understanding of the role that these PBPs play in mycobacterial peptidoglycan biology.  相似文献   

17.
The recombinant industrial Saccharomyces cerevisiae strain MA-R5 was engineered to express NADP+-dependent xylitol dehydrogenase using the flocculent yeast strain IR-2, which has high xylulose-fermenting ability, and both xylose consumption and ethanol production remarkably increased. Furthermore, the MA-R5 strain produced the highest ethanol yield (0.48 g/g) from nonsulfuric acid hydrolysate of wood chips.Successful fermentation of lignocellulosic biomass to ethanol is dependent on efficient utilization of d-xylose, which is the second most common fermentable sugar in the hydrolysate. Although the well-known fermentative yeast Saccharomyces cerevisiae is one of the most effective ethanol-producing organisms for hexose sugars, it is not able to ferment d-xylose. However, S. cerevisiae can metabolize an isomerization product of d-xylose, d-xylulose, which is phosphorylated to d-xylulose 5-phosphate, channeled through the pentose phosphate pathway and glycolysis.S. cerevisiae transformed with the XYL1 and XYL2 genes encoding xylose reductase (XR) and xylitol dehydrogenase (XDH) from Pichia stipitis (referred to as PsXR and PsXDH, respectively) acquires the ability to ferment d-xylose to ethanol (2, 5, 6, 9, 10, 12, 22). Furthermore, overexpression of the XKS1 gene encoding xylulokinase (XK) from S. cerevisiae (ScXK) has been shown to aid d-xylose utilization (4, 7, 11, 16, 23), with xylitol still a major by-product. Kuyper et al. (14) also demonstrated the successful fermentation of d-xylose to ethanol using recombinant S. cerevisiae strains expressing fungal xylose isomerase. However, these approaches are insufficient for industrial bioprocesses, mainly due to the low rate of d-xylose fermentation.Xylitol excretion has been ascribed mainly to the difference in coenzyme specificities between PsXR (with NADPH) and PsXDH (with NAD+), which creates an intracellular redox imbalance (1). Therefore, modifying the coenzyme specificity of XR and/or XDH by protein engineering is an attractive approach for achieving efficient fermentation of ethanol from d-xylose using recombinant S. cerevisiae. Watanabe et al. (24) previously succeeded in generating several PsXDH mutants (e.g., quadruple ARSdR mutant) with a complete reversal of coenzyme specificity toward NADP+ by multiple site-directed mutagenesis on amino acids from the coenzyme-binding domain. The ARSdR mutant (D207A/I208R/F209S/N211R) has more that 4,500-fold-higher catalytic efficiency (kcat/Km) with NADP+ than the wild-type PsXDH. In addition, we recently found that several laboratory recombinant S. cerevisiae strains, in which the ARSdR mutant, along with PsXR and ScXK, is expressed through a strong constitutive promoter, increased ethanol production from d-xylose at the expense of xylitol excretion (17, 18), probably by maintaining the intracellular redox balance. However, commercialization of lignocellulosic hydrolysate fermentation requires industrial strains that are more robust than laboratory strains (5, 19, 21).A potential host for developing d-xylose-fermenting strains requires an active and efficient pentose phosphate pathway linking the d-xylose-to-d-xylulose pathway to glycolysis. In the case of S. cerevisiae, strains have different d-xylulose fermentation abilities (3, 25), indicating inherent differences in the capacities of these strains to ferment pentose sugars. Furthermore, anaerobic d-xylulose fermentation was investigated to identify genetic backgrounds potentially beneficial to anaerobic d-xylose fermentation in strains not exhibiting product formation related to the redox imbalance generated by the first two steps of the eukaryotic d-xylose metabolism (3), although the physiological purpose of the different d-xylulose fermentation abilities of S. cerevisiae is not yet clear. Therefore, we selected an efficient industrial d-xylulose-fermenting S. cerevisiae strain as a host for constructing a recombinant strain through chromosomal integration of the NADP+-dependent XDH gene and the XR and endogenous XK genes. Using this recombinant strain, we characterized the enzyme activity and ability to ferment both d-xylose and lignocellulosic hydrolysate.  相似文献   

18.
19.
Producer cell immunity to the streptococcolytic enzyme zoocin A, which is a d-alanyl-l-alanine endopeptidase, is due to Zif, the zoocin A immunity factor. Zif has high degrees of similarity to MurM and MurN (members of the FemABX family of proteins), which are responsible for the addition of amino acids to cross bridges during peptidoglycan synthesis in streptococci. In this study, purified peptidoglycans from strains with and without zif were compared to determine how Zif modifies the peptidoglycan layer to cause resistance to zoocin A. The peptidoglycan from each strain was hydrolyzed using the streptococcolytic phage lysin B30, and the resulting muropeptides were separated by reverse-phase high-pressure liquid chromatography, labeled with 4-sulfophenyl isothiocyanate, and analyzed by tandem mass spectrometry in the negative-ion mode. It was determined that Zif alters the peptidoglycan by increasing the proportion of cross bridges containing three l-alanines instead of two. This modification decreased binding of the recombinant target recognition domain of zoocin A to peptidoglycan. Zif-modified peptidoglycan also was less susceptible to hydrolysis by the recombinant catalytic domain of zoocin A. Thus, Zif is a novel FemABX-like immunity factor because it provides resistance to a bacteriolytic endopeptidase by lengthening the peptidoglycan cross bridge rather than by causing an amino acid substitution.During streptococcal peptidoglycan synthesis, monomer subunits are generated inside the cell, with nonribosomal peptidyl transferases responsible for the addition of amino acids onto the epsilon amino group of lysine in the subunits. These nonribosomal peptidyl transferases are part of the FemABX family of proteins, some of which have been implicated in penicillin resistance (5, 26). In Streptococcus pneumoniae peptidoglycan synthesis, MurM attaches either an l-alanine or an l-serine to the epsilon amino group of lysine, and MurN then adds an l-alanine (11, 26).Zoocin A is a d-alanyl-l-alanine endopeptidase produced by Streptococcus equi subsp. zooepidemicus 4881 that hydrolyzes peptidoglycan cross bridges of susceptible streptococci (12). Zoocin A has two functional domains (18). The N-terminal catalytic domain (CAT) has high degrees of similarity to several other bacteriolytic endopeptidases, including the staphylolytic enzyme lysostaphin. The C-terminal target recognition domain (TRD), which facilitates binding of the enzyme to peptidoglycan (1), has very little similarity to any characterized conserved domain.Producer cell immunity to zoocin A is due to zif (zoocin A immunity factor), which is adjacent to zooA on the chromosome and is transcribed divergently (4). Zif has high degrees of similarity to MurM and MurN and also to the lysostaphin resistance protein and other FemABX-like immunity proteins (23). Previously characterized FemABX-like immunity proteins provide resistance to peptidoglycan cross-bridge hydrolases by inserting an amino acid different from those specified by the normal FemABX-like proteins (6, 9, 15, 25), whereas Zif does not (4). It has been shown previously that Zif-specified resistance to zoocin A is an intrinsic characteristic of the peptidoglycan layer (12). Therefore, Zif must modify the peptidoglycan layer in a novel way that provides resistance to zoocin A. In the present study, Zif was shown to insert an additional l-alanine into the peptidoglycan cross bridges, which inhibited both binding of the zoocin A TRD and the ability of the zoocin A CAT to hydrolyze the cross bridge.  相似文献   

20.
The human oropharyngeal pathogen Aggregatibacter actinomycetemcomitans synthesizes multiple adhesins, including the nonfimbrial extracellular matrix protein adhesin A (EmaA). EmaA monomers trimerize to form antennae-like structures on the surface of the bacterium, which are required for collagen binding. Two forms of the protein have been identified, which are suggested to be linked with the type of O-polysaccharide (O-PS) of the lipopolysaccharide (LPS) synthesized (G. Tang et al., Microbiology 153:2447-2457, 2007). This association was investigated by generating individual mutants for a rhamnose sugar biosynthetic enzyme (rmlC; TDP-4-keto-6-deoxy-d-glucose 3,5-epimerase), the ATP binding cassette (ABC) sugar transport protein (wzt), and the O-antigen ligase (waaL). All three mutants produced reduced amounts of O-PS, and the EmaA monomers in these mutants displayed a change in their electrophoretic mobility and aggregation state, as observed in sodium dodecyl sulfate (SDS)-polyacrylamide gels. The modification of EmaA with O-PS sugars was suggested by lectin blots, using the fucose-specific Lens culinaris agglutinin (LCA). Fucose is one of the glycan components of serotype b O-PS. The rmlC mutant strain expressing the modified EmaA protein demonstrated reduced collagen adhesion using an in vitro rabbit heart valve model, suggesting a role for the glycoconjugant in collagen binding. These data provide experimental evidence for the glycosylation of an oligomeric, coiled-coil adhesin and for the dependence of the posttranslational modification of EmaA on the LPS biosynthetic machinery in A. actinomycetemcomitans.The Gram-negative, nonmotile, microaerophilic, and oropharyngeal bacterium Aggregatibacter actinomycetemcomitans preferentially colonizes the subgingival region of the human oral cavity. This microorganism is implicated as the etiological agent of localized aggressive periodontitis (9, 13) and causes extraoral infections, including pneumonia, osteitis (30), and infective endocarditis (6). Recent studies also link this periodontal pathogen to cardiovascular diseases, such as atherosclerosis (20).Typical of Gram-negative bacteria, the outer membrane of A. actinomycetemcomitans possesses an asymmetric lipid-protein bilayer. The inner leaflet of the outer membrane is mainly phospholipids, and the outer leaflet consists of lipopolysaccharide (LPS), phospholipids, and proteins (4). LPS molecules are ubiquitously distributed on the outer membrane and are essential for maintaining the membrane integrity (3). Intact LPS molecules are also required for the assembly of some large outer membrane proteins (3, 18, 41). A typical LPS molecule is composed of hydrophobic lipid A, a nonrepeat core oligosaccharide, and a repeating O-antigen or O-polysaccharide (O-PS). The distal O-PS is a major antigen, stimulating the host immune response, and the basis for serotyping Gram-negative bacteria (36), including A. actinomycetemcomitans (32, 50).Six different serotypes (a to f) and the corresponding genetic loci have been identified for A. actinomycetemcomitans (19, 22, 27, 44, 50, 54, 55). Serotype b remains one of the common serotypes found in the human oral cavity (9, 13, 51). The serotype b O-PS of A. actinomycetemcomitans is encoded by an operon composed of 21 genes, which are responsible for the biosynthesis of the repeating trisaccharide unit of this particular serotype (53, 55). Each O-PS unit of serotype b contains a disaccharide backbone composed of d-fucose (d-Fuc) and l-rhamnose (l-Rha), linked by a nonreducing d-N-acetylgalactosamine (d-GalNAc) at the O-3 position of l-Rha (33) (Fig. (Fig.1A1A).Open in a separate windowFIG. 1.(A) O-PS structure of serotype b A. actinomycetemcomitans. (B) Silver-stained 5 to 15% polyacrylamide-SDS gel of serotype b LPS. A total of 1.0 ml of mid-logarithmic-phase cells were collected and lysed. Three lysates from each strain were combined and treated with proteinase K at 60°C for 60 min before electrophoresis, followed by silver staining. C, control: whole-cell lysate without proteinase K digestion; WT, wild type (VT1169); emaA, extracellular matrix protein adhesin A mutant; rmlC, rhamnose epimerase mutant; wzt, ATP-binding cassette sugar transport mutant. The dark brown staining of the high molecular weight (75,000 to 250,000) corresponds to polymerized O-PS.The assembly of LPS molecules in Gram-negative bacteria involve diverse enzymes and pathways due to the variation of the O-PS structures among different bacteria (36). RmlC (previously RfbD), Wzt (previously AbcA or RfbB), and WaaL are three enzymes involved in different stages of the LPS synthesis of some Gram-negative bacteria (7, 36, 37). A homologue of RmlC, TDP-4-keto-6-deoxy-d-glucose 3,5-epimerase, which is required for l-Rha synthesis, has been identified in A. actinomycetemcomitans (53, 55). Wzt is an ATP binding cassette (ABC) transporter that exports saccharide polymers from the cytoplasm to the periplasmic space (7, 36). A homologue of wzt was originally identified from a serotype b strain of A. actinomycetemcomitans, based on protein sequence identity with Aeromonas salmonicida (55). Kaplan et al. (19) later showed that a serotype f wzt mutant strain of A. actinomycetemcomitans produces less O-PS. WaaL, an O-antigen ligase found in Escherichia coli and Pseudomonas aeruginosa, ligates an undecaprenol pyrophosphate-linked oligo- or polysaccharide onto the lipid A-core oligosaccharide in the periplasm (1, 36). A putative O-antigen ligase is located in the chromosome of a serotype b A. actinomycetemcomitans strain (HK1651), based on sequence homology (Oralgen, Los Alamos, NM).Our earlier work suggested a correlation between the type of LPS molecule and the form of EmaA synthesized by A. actinomycetemcomitans (46). The EmaA of serotype b A. actinomycetemcomitans is a 202-kDa protein that forms the antennae-like appendages found on the surface of A. actinomycetemcomitans and is required for collagen binding (40). The appendages are composed of three EmaA monomers that oligomerize to form an ellipsoidal structure required for the collagen binding activity (56, 57). The ellipsoidal structure corresponds to the amino termini of the proteins and is located at the distal end of a long stalk domain that is attached to the outer membrane by the carboxyl termini (57). The carboxyl termini of the proteins assume β-barrel structures required for pore formation and translocation of the molecules through the outer membrane, similar to those of other type Vc autotransporter proteins (14). Recently, we have demonstrated that EmaA is important in the initiation of infective endocarditis in a rabbit model of infectious endocarditis (45).Two transposon mutant strains (rmlC and wzt) and a waaL mutant strain generated by site-directed insertional mutagenesis have been developed and characterized in this study. The rmlC mutant did not synthesize l-Rha and did not produce detectable O-PS. The wzt and waaL mutant strains synthesized less O-PS than the wild-type strain. Complementation of the mutant strains restored the production of the serotype b O-PS to wild-type levels. An increase in the electrophoretic mobility of the EmaA monomer was observed in all three mutants, which suggests the presence of carbohydrate. The EmaA mobility reverted to wild-type mobility upon complementation. The presence of carbohydrate associated with EmaA was confirmed by lectin blotting, and in vitro collagen binding assessment demonstrated that the glycoconjugant is important for the full function of this adhesin. The experimental data suggest that EmaA contains carbohydrate similar to that present in O-PS and is a substrate for the O-antigen ligase of the LPS biosynthetic pathway of A. actinomycetemcomitans.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号