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 93 毫秒
1.
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   

2.
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   

3.
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   

4.
An open reading frame located in the bisC-cspA intergenic region, or at 80.1 min on the Escherichia coli chromosome, encodes a hypothetical 2-hydroxyacid dehydrogenase, which was identified as a result of the E. coli Genome Sequencing Project. We report here that the product of the gene (yiaE) is a 2-ketoaldonate reductase (2KR). The gene was cloned and expressed with a C-terminal His tag in E. coli, and the protein was purified by metal-chelate affinity chromatography. The determination of the NH2-terminal amino acid sequence of the protein defined the translational start site of this gene. The enzyme was found to be a 2KR catalyzing the reduction of 2,5-diketo-d-gluconate to 5-keto-d-gluconate, 2-keto-d-gluconate (2KDG) to d-gluconate, 2-keto-l-gulonate to l-idonate. The reductase was optimally active at pH 7.5, with NADPH as a preferred electron donor. The deduced amino acid sequence showed 69.4% identity with that of 2KR from Erwinia herbicola. Disruption of this gene on the chromosome resulted in the loss of 2KR activity in E. coli. E. coli W3110 was found to grow on 2KDG, whereas the mutant deficient in 2KR activity was unable to grow on 2KDG as the carbon source, suggesting that 2KR is responsible for the catabolism of 2KDG in E. coli and the diminishment of produced 2KDG from d-gluconate in the cultivation of E. coli harboring a cloned gluconate dehydrogenase gene.

We previously reported the cloning and expression of a gene cluster encoding three subunits of membrane-bound gluconate dehydrogenase (GADH) from Erwinia cypripedii in Escherichia coli (26). In the course of further study on the conversion of d-gluconate to 2-keto-d-gluconate (2KDG) with a recombinant E. coli strain, we observed that the level of 2KDG produced in the medium gradually decreased after the exhaustion of d-gluconate in the medium (see Fig. Fig.1).1). In an effort to find the reason, the NADPH-dependent reductase activity catalyzing the conversion of 2KDG to d-gluconate was detected in extracts of E. coli cells. This result suggested the existence of enzymes involved in ketogluconate metabolism in E. coli, as reported for several species of the genera Corynebacterium, Brevibacterium, Erwinia, Acetobacter, Gluconobacter, Serratia, and Pseudomonas (20, 23, 25). In Erwinia, Acetobacter, Gluconobacter, Serratia, and Pseudomonas, oxidation of glucose to ketogluconates such as 2KDG, 5-keto-d-gluconate (5KDG), and 2,5-diketo-d-gluconate (25DKG) has been shown to proceed via membrane-bound dehydrogenases, which are linked to the electron transport chain (2, 21). The ketogluconates or their phosphorylated forms are unique substrates in that they enter into central metabolism only after they are reduced by NADPH-dependent reductases (20, 23). NADPH-dependent 2-ketoaldonate reductase (2KR), which catalyzes the reduction of 2KDG to d-gluconate, 25DKG to 5KDG, and 2-keto-l-gulonate (2KLG) to l-idonate (IA), has been purified and characterized from Brevibacterium ketosoreductum (25) and Erwinia herbicola (23). Even if the substrate specificity has not been examined with 25DKG as a substrate, 2KDG reductases from acetic acid bacteria also catalyze the reduction of 2KLG to IA as well as of 2KDG to d-gluconate (1).Open in a separate windowFIG. 1Time course of bioconversion of d-gluconate to 2KDG by E. coli harboring the cloned GADH gene. E. coli W3110(pGA313) was grown in a 2-liter fermentor at 37°C with aeration at 1 vvm and agitation at 500 rpm.Until now, no ketoaldonate reductase has been reported for E. coli. We report here that the product of the yiaE gene, located in the bisC-cspA intergenic region at 80.1 min on the E. coli chromosome, is a 2KR; in addition, the diminishment of produced 2KDG from d-gluconate in the cultivation of recombinant E. coli harboring a cloned membrane-bound GADH gene is due to 2KR as the cytosolic enzyme responsible for conversion of 2KDG to d-gluconate. We found also that E. coli W3110 grows on 2KDG as the sole carbon source.  相似文献   

5.
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   

6.
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   

7.
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   

8.
9.
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   

10.
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   

11.
The oxidative d-xylose catabolic pathway of Caulobacter crescentus, encoded by the xylXABCD operon, was expressed in the gram-negative bacterium Pseudomonas putida S12. This engineered transformant strain was able to grow on d-xylose as a sole carbon source with a biomass yield of 53% (based on g [dry weight] g d-xylose−1) and a maximum growth rate of 0.21 h−1. Remarkably, most of the genes of the xylXABCD operon appeared to be dispensable for growth on d-xylose. Only the xylD gene, encoding d-xylonate dehydratase, proved to be essential for establishing an oxidative d-xylose catabolic pathway in P. putida S12. The growth performance on d-xylose was, however, greatly improved by coexpression of xylXA, encoding 2-keto-3-deoxy-d-xylonate dehydratase and α-ketoglutaric semialdehyde dehydrogenase, respectively. The endogenous periplasmic glucose dehydrogenase (Gcd) of P. putida S12 was found to play a key role in efficient oxidative d-xylose utilization. Gcd activity not only contributes to d-xylose oxidation but also prevents the intracellular accumulation of toxic catabolic intermediates which delays or even eliminates growth on d-xylose.The requirement for renewable alternatives to replace oil-based chemicals and fuels necessitates development of novel technologies. Lignocellulose provides a promising alternative feedstock. However, since the pentose sugar fraction may account for up to 25% of lignocellulosic biomass (12), it is essential that this fraction is utilized efficiently to obtain cost-effective biochemical production. In a previous study, the solvent-tolerant bacterium Pseudomonas putida S12, known for its use as a platform host for the production of aromatic compounds (15, 16, 19, 22), was engineered to use d-xylose as a sole carbon source. This was achieved by introducing genes encoding the phosphorylative d-xylose metabolic pathway of Escherichia coli, followed by laboratory evolution (14). Prior to evolutionary improvement, extensive oxidation of d-xylose to d-xylonate occurred, resulting in a very low biomass-for-substrate yield as d-xylonate is a metabolic dead-end product in P. putida. The evolution approach resulted in elimination of the activity of periplasmic glucose dehydrogenase (Gcd), the enzyme responsible for d-xylose oxidation, which turned out to be a critical step in optimizing phosphorylative d-xylose utilization in P. putida S12.Instead of prevention of endogenous oxidation of d-xylose, this oxidation may be used to our advantage when it is combined with an oxidative d-xylose metabolic pathway, such as the pathways described for several Pseudomonas species, Caulobacter crescentus, and Haloarcula marismortui (7, 11, 18, 20). In these pathways, d-xylonate is dehydrated to 2-keto-3-deoxy-d-xylonate. This intermediate either can be cleaved into pyruvate and glycolaldehyde (7) or is further dehydrated to α-ketoglutaric semialdehyde (α-KGSA). In the final step of the latter pathway, α-KGSA is oxidized to the tricarboxylic acid (TCA) cycle intermediate α-ketoglutarate (18, 20).In addition to Gcd (PP1444), some of the enzymes required for oxidative d-xylose metabolism are expected to be endogenous in P. putida S12. Transport of d-xylonate into the cytoplasm likely occurs through the gluconate transporter (encoded by gntP [PP3417]). The enzyme catalyzing the final step of the pathway, α-KGSA dehydrogenase, is also likely to be present (presumably PP1256 and/or PP3602) because of the requirement for metabolism of 4-hydroxyproline (1), a compound that is efficiently utilized by P. putida S12. In view of these properties, the most obvious approach for constructing d-xylose-utilizing P. putida S12 is reconstruction of a complete oxidative d-xylose metabolic pathway by introducing the parts of such a pathway that complement the endogenous activities. Recently, the genetic information for one such oxidative d-xylose pathway has become available (18), enabling the approach used in the present study, i.e., expression of the oxidative d-xylose metabolic pathway of C. crescentus in P. putida S12 and investigation of the contribution of endogenous enzyme activities.  相似文献   

12.
α-l-Arabinofuranosidases I and II were purified from the culture filtrate of Aspergillus awamori IFO 4033 and had molecular weights of 81,000 and 62,000 and pIs of 3.3 and 3.6, respectively. Both enzymes had an optimum pH of 4.0 and an optimum temperature of 60°C and exhibited stability at pH values from 3 to 7 and at temperatures up to 60°C. The enzymes released arabinose from p-nitrophenyl-α-l-arabinofuranoside, O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose, and arabinose-containing polysaccharides but not from O-β-d-xylopyranosyl-(1→2)-O-α-l-arabinofuranosyl-(1→3)-O-β-d-xylopyranosyl-(1→4)-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. α-l-Arabinofuranosidase I also released arabinose from O-β-d-xylopy-ranosyl-(1→4)-[O-α-l-arabinofuranosyl-(1→3)]-O-β-d-xylopyranosyl-(1→4)-d-xylopyranose. However, α-l-arabinofuranosidase II did not readily catalyze this hydrolysis reaction. α-l-Arabinofuranosidase I hydrolyzed all linkages that can occur between two α-l-arabinofuranosyl residues in the following order: (1→5) linkage > (1→3) linkage > (1→2) linkage. α-l-Arabinofuranosidase II hydrolyzed the linkages in the following order: (1→5) linkage > (1→2) linkage > (1→3) linkage. α-l-Arabinofuranosidase I preferentially hydrolyzed the (1→5) linkage of branched arabinotrisaccharide. On the other hand, α-l-arabinofuranosidase II preferentially hydrolyzed the (1→3) linkage in the same substrate. α-l-Arabinofuranosidase I released arabinose from the nonreducing terminus of arabinan, whereas α-l-arabinofuranosidase II preferentially hydrolyzed the arabinosyl side chain linkage of arabinan.Recently, it has been proven that l-arabinose selectively inhibits intestinal sucrase in a noncompetitive manner and reduces the glycemic response after sucrose ingestion in animals (33). Based on this observation, l-arabinose can be used as a physiologically functional sugar that inhibits sucrose digestion. Effective l-arabinose production is therefore important in the food industry. l-Arabinosyl residues are widely distributed in hemicelluloses, such as arabinan, arabinoxylan, gum arabic, and arabinogalactan, and the α-l-arabinofuranosidases (α-l-AFases) (EC 3.2.1.55) have proven to be essential tools for enzymatic degradation of hemicelluloses and structural studies of these compounds.α-l-AFases have been classified into two families of glycanases (families 51 and 54) on the basis of amino acid sequence similarities (11). The two families of α-l-AFases also differ in substrate specificity for arabinose-containing polysaccharides. Beldman et al. summarized the α-l-AFase classification based on substrate specificities (3). One group contains the Arafur A (family 51) enzymes, which exhibit very little or no activity with arabinose-containing polysaccharides. The other group contains the Arafur B (family 54) enzymes, which cleave arabinosyl side chains from polymers. However, this classification is too broad to define the substrate specificities of α-l-AFases. There have been many studies of the α-l-AFases (3, 12), especially the α-l-AFases of Aspergillus species (28, 1215, 17, 22, 23, 2832, 3639, 4143, 46). However, there have been only a few studies of the precise specificities of these α-l-AFases. In previous work, we elucidated the substrate specificities of α-l-AFases from Aspergillus niger 5-16 (17) and Bacillus subtilis 3-6 (16, 18), which should be classified in the Arafur A group and exhibit activity with arabinoxylooligosaccharides, synthetic methyl 2-O-, 3-O-, and 5-O-arabinofuranosyl-α-l-arabinofuranosides (arabinofuranobiosides) (20), and methyl 3,5-di-O-α-l-arabinofuranosyl-α-l-arabinofuranoside (arabinofuranotrioside) (19).In the present work, we purified two α-l-AFases from a culture filtrate of Aspergillus awamori IFO 4033 and determined the substrate specificities of these α-l-AFases by using arabinose-containing polysaccharides and the core oligosaccharides of arabinoxylan and arabinan.  相似文献   

13.
d-Galacturonic acid can be obtained by hydrolyzing pectin, which is an abundant and low value raw material. By means of metabolic engineering, we constructed fungal strains for the conversion of d-galacturonate to meso-galactarate (mucate). Galactarate has applications in food, cosmetics, and pharmaceuticals and as a platform chemical. In fungi d-galacturonate is catabolized through a reductive pathway with a d-galacturonate reductase as the first enzyme. Deleting the corresponding gene in the fungi Hypocrea jecorina and Aspergillus niger resulted in strains unable to grow on d-galacturonate. The genes of the pathway for d-galacturonate catabolism were upregulated in the presence of d-galacturonate in A. niger, even when the gene for d-galacturonate reductase was deleted, indicating that d-galacturonate itself is an inducer for the pathway. A bacterial gene coding for a d-galacturonate dehydrogenase catalyzing the NAD-dependent oxidation of d-galacturonate to galactarate was introduced to both strains with disrupted d-galacturonate catabolism. Both strains converted d-galacturonate to galactarate. The resulting H. jecorina strain produced galactarate at high yield. The A. niger strain regained the ability to grow on d-galacturonate when the d-galacturonate dehydrogenase was introduced, suggesting that it has a pathway for galactarate catabolism.d-Galacturonate is the main component of pectin, an abundant and cheap raw material. Sugar beet pulp and citrus peel are both rich in pectin residues. At present, these residues are mainly used as cattle feed. However, since energy-consuming drying and pelletizing of the residues is required to prevent them from rotting, it is not always economical to process the residues, and it is desirable to find alternative uses.Various microbes which live on decaying plant material have the ability to catabolize d-galacturonate using various, completely different pathways (19). Eukaryotic microorganisms use a reductive pathway in which d-galacturonate is first reduced to l-galactonate by an NAD(P)H-dependent reductase (12, 17). In the following steps a dehydratase, aldolase, and reductase convert the l-galactonate to pyruvate and glycerol (9, 11, 14).In Hypocrea jecorina (anamorph Trichoderma reesei) the gar1 gene codes for a strictly NADPH-dependent d-galacturonate reductase. In Aspergillus niger a homologue gene sequence, gar2, exists; however, a different gene, gaaA, is upregulated during growth on d-galacturonate containing medium (16). The gaaA codes for a d-galacturonate reductase with different kinetic properties than the H. jecorina enzyme, having a higher affinity toward d-galacturonate and using either NADH or NADPH as cofactor. It is not known whether gar2 codes for an active protein.Some bacteria, such as Agrobacterium tumefaciens or Pseudomonas syringae, have an oxidative pathway for d-galacturonate catabolism. In this pathway d-galacturonate is first oxidized to meso-galactarate (mucate) by an NAD-utilizing d-galacturonate dehydrogenase. Galactarate is then converted in the following steps to α-ketoglutarate. This route is sometimes called the α-ketoglutarate pathway (20). Galactarate can also be catabolized through the glycerate pathway (20). The products of this pathway are pyruvate and d-glycerate. These pathways have been described in prokaryotes, and it is not certain whether similar pathways also exist in fungi, some of which are able to metabolize galactarate.d-Galacturonate dehydrogenase (EC 1.1.1.203) has been described in Agrobacterium tumefaciens and in Pseudomonas syringae, and the enzymes from these organisms have been purified and characterized (3, 6, 22). Recently, the corresponding genes were also identified (4, 24). Both enzymes are specific for NAD as a cofactor but are not specific for the substrate. They oxidize d-galacturonate and d-glucuronate to meso-galactarate (mucate) and d-glucarate (saccharate), respectively. The reaction product is probably the hexaro-lactone which spontaneously hydrolyzes. The reverse reaction can only be observed at acidic pH where some of the galactarate is in the lactone form (22).We describe here strains of filamentous fungi that have been genetically engineered to produce galactarate by disruption of d-galacturonate reductase and expression of d-galacturonate dehydrogenase (Fig. (Fig.1).1). Galactarate is currently commercially produced from d-galactose by oxidation with nitric acid (1) or from d-galacturonate by electrolytic oxidation (8). Oxidation with nitric acid is expensive and produces toxic wastes. Galactarate is used as a chelator and in skin care products. It was formerly used as a leavening agent in self-rising flour (2) and has potential applications in polymer synthesis (10) and as a platform chemical (for a review, see reference 13).Open in a separate windowFIG. 1.Engineering the d-galacturonic acid pathway in fungi. Deletion of the gene encoding d-galacturonate reductase resulted in strains unable to utilize d-galacturonic acid as a carbon source. The expression of a bacterial udh gene, encoding an NAD-dependent d-galacturonate dehydrogenase, resulted in fungal strains, which were able to oxidize d-galacturonic acid to meso-galactaric acid (mucic acid). d-Galacturonate dehydrogenase forms a galactaro-lactone which spontaneously hydrolyzes.  相似文献   

14.
15.
Hydroxyacid dehydrogenases of lactic acid bacteria, which catalyze the stereospecific reduction of branched-chain 2-keto acids to 2-hydroxyacids, are of interest in a variety of fields, including cheese flavor formation via amino acid catabolism. In this study, we used both targeted and random mutagenesis to identify the genes responsible for the reduction of 2-keto acids derived from amino acids in Lactococcus lactis. The gene panE, whose inactivation suppressed hydroxyisocaproate dehydrogenase activity, was cloned and overexpressed in Escherichia coli, and the recombinant His-tagged fusion protein was purified and characterized. The gene annotated panE was the sole gene responsible for the reduction of the 2-keto acids derived from leucine, isoleucine, and valine, while ldh, encoding l-lactate dehydrogenase, was responsible for the reduction of the 2-keto acids derived from phenylalanine and methionine. The kinetic parameters of the His-tagged PanE showed the highest catalytic efficiencies with 2-ketoisocaproate, 2-ketomethylvalerate, 2-ketoisovalerate, and benzoylformate (Vmax/Km ratios of 6,640, 4,180, 3,300, and 2,050 U/mg/mM, respectively), with NADH as the exclusive coenzyme. For the reverse reaction, the enzyme accepted d-2-hydroxyacids but not l-2-hydroxyacids. Although PanE showed the highest degrees of identity to putative NADP-dependent 2-ketopantoate reductases (KPRs), it did not exhibit KPR activity. Sequence homology analysis revealed that, together with the d-mandelate dehydrogenase of Enterococcus faecium and probably other putative KPRs, PanE belongs to a new family of d-2-hydroxyacid dehydrogenases which is unrelated to the well-described d-2-hydroxyisocaproate dehydrogenase family. Its probable physiological role is to regenerate the NAD+ necessary to catabolize branched-chain amino acids, leading to the production of ATP and aroma compounds.Hydroxyacid dehydrogenases catalyze the stereospecific and reversible reduction of 2-keto acids to 2-hydroxyacids. These NAD(H)-dependent oxidoreductases are of interest in a variety of fields. Firstly, they are valuable catalysts for the production of the stereospecific isomers of 2-hydroxyacids that are used in the production of semisynthetic antibiotics or pharmaceuticals (32). Secondly, in lactic acid bacteria, hydroxyacid dehydrogenases are believed to be negatively involved in flavor production, since they compete with other enzymes generating flavor compounds from 2-keto acids derived from amino acids (66), while 2-hydroxyacids are not aroma compounds or precursors of flavor compounds. Indeed, overexpression of the d-2-hydroxyisocaproate dehydrogenase (d-2-HicDH) of Lactobacillus casei has been shown previously to decrease the production of aroma compounds and to delay flavor formation in low-fat cheddar cheese (12). Additionally, hydroxyacid dehydrogenases are involved in the biopreservation properties of lactic acid bacteria, because certain 2-hydroxyacids exhibit antifungal and antilisterial activities (15, 37, 64). Several hydroxyacid dehydrogenases in lactic acid bacteria have been characterized previously (21, 27). Lactate dehydrogenases (LDH), responsible for the specific reduction of pyruvate to lactic acid, have been studied extensively (2, 7, 13, 16, 21, 40, 49, 54). HicDHs and mandelate dehydrogenases (manDHs), active toward a broad range of 2-keto acids, including straight-chain aliphatic 2-keto acids, branched-chain 2-keto acids, and 2-keto acids with aromatic side chains, in several lactic acid bacteria have also been characterized previously (5, 6, 27, 28, 29, 38, 39, 50). Although manDHs and HicDHs of lactic acid bacteria prefer 2-ketoisocaproate (KIC) among 2-keto acid substrates, they differ in their activities toward C-3-branched substrates. In particular, manDHs exhibit high levels of activity toward 2-ketoisovalerate (KIV) and benzoylformate, unlike HicDHs. LDHs, HicDHs, and manDHs are divided into two groups, the l and d groups, depending on the stereoisomer produced. d-LDHs and d-HicDHs are members of the same family of 2-hydroxyacid dehydrogenases, which is distinct from the l-LDH family (13, 54). In general, each lactic acid bacterium contains several hydroxyacid dehydrogenases.In Lactococcus lactis, which is widely employed as a starter in cheese production, the main LDH is an l-LDH activated by fructose 1,6-bisphosphate (FBP) (23). In addition to the ldh gene that encodes the l-LDH (40), L. lactis contains other genes showing significant levels of similarity to the hydroxyacid dehydrogenases of other lactic acid bacteria (9), but their functions remain unknown, except for that of the recently identified ldhB gene. ldhB is a silent gene that can be activated in ldh-deficient strains via an IS981 element-specific insertion to produce a functional LDH (10). However L. lactis produces 2-hydroxyacids from the 2-keto acids derived from amino acids not only in vitro, using resting cells, but also in cheese (65, 66). In L. lactis, the catabolism of the aromatic and branched-chain amino acids that are precursors of aroma compounds is initiated by aminotransferases, producing 2-keto acids. These 2-keto acids can be further catabolized into carboxylic acids via a 2-keto acid dehydrogenase, a transacetylase, and a kinase or reduced to 2-hydroxyacids by a hydroxyacid dehydrogenase (66) (Fig. (Fig.1).1). In some strains, branched-chain 2-keto acids can also be decarboxylated to aldehydes that are potent aroma compounds. The aim of the present study was to identify and characterize the enzyme(s) involved in the production of 2-hydroxyacids from the 2-keto acids derived from phenylalanine and branched-chain amino acids and to evaluate its impact on amino acid catabolism.Open in a separate windowFIG. 1.Leucine catabolism pathway in L. lactis TIL46. AT, aminotransferase; HADH, hydroxyacid dehydrogenase; TA, transacetylase; KDH, keto acid dehydrogenase; HIC, 2-hydroxyisocaproate.  相似文献   

16.
Given its availability and low price, glycerol has become an ideal feedstock for the production of fuels and chemicals. We recently reported the pathways mediating the metabolism of glycerol in Escherichia coli under anaerobic and microaerobic conditions. In this work, we engineer E. coli for the efficient conversion of glycerol to d-lactic acid (d-lactate), a negligible product of glycerol metabolism in wild-type strains. A homofermentative route for d-lactate production was engineered by overexpressing pathways involved in the conversion of glycerol to this product and blocking those leading to the synthesis of competing by-products. The former included the overexpression of the enzymes involved in the conversion of glycerol to glycolytic intermediates (GlpK-GlpD and GldA-DHAK pathways) and the synthesis of d-lactate from pyruvate (d-lactate dehydrogenase). On the other hand, the synthesis of succinate, acetate, and ethanol was minimized through two strategies: (i) inactivation of pyruvate-formate lyase (ΔpflB) and fumarate reductase (ΔfrdA) (strain LA01) and (ii) inactivation of fumarate reductase (ΔfrdA), phosphate acetyltransferase (Δpta), and alcohol/acetaldehyde dehydrogenase (ΔadhE) (strain LA02). A mutation that blocked the aerobic d-lactate dehydrogenase (Δdld) also was introduced in both LA01 and LA02 to prevent the utilization of d-lactate. The most efficient strain (LA02Δdld, with GlpK-GlpD overexpressed) produced 32 g/liter of d-lactate from 40 g/liter of glycerol at a yield of 85% of the theoretical maximum and with a chiral purity higher than 99.9%. This strain exhibited maximum volumetric and specific productivities for d-lactate production of 1.5 g/liter/h and 1.25 g/g cell mass/h, respectively. The engineered homolactic route generates 1 to 2 mol of ATP per mol of d-lactate and is redox balanced, thus representing a viable metabolic pathway.Lactic acid (lactate) and its derivatives have many applications in the food, pharmaceutical, and polymer industries (13, 30). An example is polylactic acid, a renewable, biodegradable, and environmentally friendly polymer produced from d- and l-lactate (19). In this context, biological processes have the advantage of being able to produce chirally pure lactate from inexpensive media containing only the carbon source and mineral salts (43). While lactic acid bacteria traditionally have been used in the production of d-lactate from carbohydrate-rich feedstocks, several laboratories recently have reported alternative biocatalysts (13, 30), many of which are engineered Escherichia coli strains that produce d- or l-lactate (4, 8, 50, 51, 52).Unlike the aforementioned reports, which have dealt with the use of carbohydrates, our work focuses on the use of glycerol as a carbon source for the production of d-lactate. Glycerol has become an inexpensive and abundant substrate due to its generation in large amounts as a by-product of biodiesel and bioethanol production (18, 32, 47). The conversion of glycerol to higher-value products has been proposed as a path to economic viability for the biofuels industry (47). One such product is lactate, whose production could be readily integrated into existing biodiesel and bioethanol facilities, thus establishing true biorefineries.Although many microorganisms are able to metabolize glycerol (25), the use of industrial microbes such as E. coli could greatly accelerate the development of platforms to produce fuels and chemicals from this carbon source. We recently reported on the ability of E. coli to metabolize glycerol under either anaerobic or microaerobic conditions and identified the environmental and metabolic determinants of these processes (9, 11, 28). In one of the studies, the pathways involved in the microaerobic utilization of glycerol were elucidated, and they are shown in Fig. Fig.11 (9). A common characteristic of glycerol metabolism under either anaerobic or microaerobic conditions is the generation of ethanol as the primary product and the negligible production of lactate (6, 9, 11, 28). In the work reported here, the knowledge base created by the aforementioned studies was used to engineer E. coli for the efficient conversion of glycerol to d-lactate in minimal medium. The engineered strains hold great promise as potential biocatalysts for the conversion of low-value glycerol streams to a higher-value product like d-lactate.Open in a separate windowFIG. 1.Pathways involved in the microaerobic utilization of glycerol in E. coli (9). Genetic modifications supporting the metabolic engineering strategies employed in this work are illustrated by thicker lines (overexpression of gldA-dhaKLM, glpK-glpD, and ldhA) or cross bars (disruption of pflB, pta, adhE, frdA, and dld). Broken lines illustrate multiple steps. Relevant reactions are represented by the names of the gene(s) coding for the enzymes: aceEF-lpdA, pyruvate dehydrogenase complex; adhE, acetaldehyde/alcohol dehydrogenase; ackA, acetate kinase; dhaKLM, dihydroxyacetone kinase; dld, respiratory d-lactate dehydrogenase; fdhF, formate dehydrogenase, part of the formate hydrogenlyase complex; frdABCD, fumarate reductase; gldA, glycerol dehydrogenase; glpD, aerobic glycerol-3-phosphate dehydrogenase; glpK, glycerol kinase; hycB-I, hydrogenase 3, part of the formate hydrogenlyase complex; ldhA, fermentative d-lactate dehydrogenase; pflB, pyruvate formate-lyase; pta, phosphate acetyltransferase; pykF, pyruvate kinase. Abbreviations: DHA, dihydroxyacetone; DHAP, DHA phosphate; G-3-P, glycerol-3-phosphate; PEP, phosphoenolpyruvate; PYR, pyruvate; P/O, amount of ATP produced in the oxidative phosphorylation per pair of electrons transferred through the electron transport system; QH2, reduced quinones.  相似文献   

17.
A trehalose synthase (TSase) that catalyzes the synthesis of trehalose from d-glucose and α-d-glucose 1-phosphate (α-d-glucose 1-P) was detected in a basidiomycete, Grifola frondosa. TSase was purified 106-fold to homogeneity with 36% recovery by ammonium sulfate precipitation and several steps of column chromatography. The native enzyme appears to be a dimer since it has apparent molecular masses of 120 kDa, as determined by gel filtration column chromatography, and 60 kDa, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Although TSase catalyzed the phosphorolysis of trehalose to d-glucose and α-d-glucose 1-P, in addition to the synthesis of trehalose from the two substrates, the TSase equilibrium strongly favors trehalose synthesis. The optimum temperatures for phosphorolysis and synthesis of trehalose were 32.5 to 35°C and 35 to 37.5°C, respectively. The optimum pHs for these reactions were 6.5 and 6.5 to 6.8, respectively. The substrate specificity of TSase was very strict: among eight disaccharides examined, only trehalose was phosphorolyzed, and only α-d-glucose 1-P served as a donor substrate with d-glucose as the acceptor in trehalose synthesis. Two efficient enzymatic systems for the synthesis of trehalose from sucrose were identified. In system I, the α-d-glucose 1-P liberated by 1.05 U of sucrose phosphorylase was linked with d-glucose by 1.05 U of TSase, generating trehalose at the initial synthesis rate of 18 mmol/h in a final yield of 90 mol% under optimum conditions (300 mM each sucrose and glucose, 20 mM inorganic phosphate, 37.5°C, and pH 6.5). In system II, we added 1.05 U of glucose isomerase and 20 mM MgSO4 to the reaction mixture of system I to convert fructose, a by-product of the sucrose phosphorylase reaction, into glucose. This system generated trehalose at the synthesis rate of 4.5 mmol/h in the same final yield.Trehalose (1-α-d-glucopyranosyl-α-d-glucopyranoside) is a nonreducing disaccharide with an α,α-1,1 glycosidic linkage and is widely distributed in plants, insects, fungi, yeast, and bacteria (7). Due to the absence of reducing ends in trehalose, it is highly resistant to heat, pH, and Maillard’s reaction (24). In trehalose-producing organisms, this compound may serve as an energy reserve, a buffer against stresses such as desiccation and freezing, and a protein stabilizer (5, 7, 26, 31, 32). If trehalose can be produced economically, then it has potential commercial applications as a sweetener, a food stabilizer, and an additive in cosmetics and pharmaceuticals (6, 25). Recently, trehalose production through fermentation of yeast (17) and Corynebacterium (30), enzymatic processes from starch (18, 34) and maltose (19, 22, 23, 33), and extraction from transformed plants (10) has been reported.Our approach to trehalose production is to use an enzymatic process to produce trehalose from sucrose, one of the least expensive sugars. Since sucrose is efficiently converted to α-d-glucose 1-phosphate (α-d-glucose 1-P) and fructose by sucrose phosphorylase (SPase), we screened microorganisms for an enzyme that converts α-d-glucose 1-P to trehalose on the assumption that the combination of the putative trehalose synthase (TSase) and SPase would convert sucrose into trehalose. Although similar enzyme activities have been reported in the basidiomycete Flammulina velutipes (11) and in the yeast Pichia fermentans (27), these enzymes have not been well characterized.Our objectives were (i) to screen microorganisms, primarily fungi, for TSase activity; (ii) to purify and characterize the TSase; (iii) to identify the enzymatic process by which trehalose is produced from sucrose; and (iv) to identify an enzymatic process for production of trehalose from sucrose in which the fructose component is also converted to trehalose.  相似文献   

18.
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   

19.
20.
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   

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