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 15 毫秒
1.
The N-succinyl-ll-diaminopimelate desuccinylase gene (dapE) in the four-step succinylase branch of the l-lysine biosynthetic pathway of Corynebacterium glutamicum was disrupted via marker-exchange mutagenesis to create a mutant strain that uses only the one-step meso-diaminopimelate dehydrogenase branch to overproduce lysine. This mutant strain grew and utilized glucose from minimal medium at the same rate as the parental strain. In addition, the dapE  strain produced lysine at the same rate as its parent strain. Transformation of the parental and dapE  strains with the amplified meso-diaminopimelate dehydrogenase gene (ddh) on a plasmid did not affect lysine production in either strain, despite an eightfold amplification of the activity of the enzyme. These results indicate that the four-step succinylase pathway is dispensable for lysine overproduction in shake-flask culture. In addition, the one-step meso-diaminopimelate dehydrogenase pathway does not limit lysine flux in Corynebacterium under these conditions. Received: 20 May 1998 / Received revision: 12 August 1998 / Accepted: 3 September 1998  相似文献   

2.
Species-specific antimicrobial therapy has the potential to combat the increasing threat of antibiotic resistance and alteration of the human microbiome. We therefore set out to demonstrate the beginning of a pathogen-selective drug discovery method using the periodontal pathogen Porphyromonas gingivalis as a model. Through our knowledge of metabolic networks and essential genes we identified a “druggable” essential target, meso-diaminopimelate dehydrogenase, which is found in a limited number of species. We adopted a high-throughput virtual screen method on the ZINC chemical library to select a group of potential small-molecule inhibitors. Meso-diaminopimelate dehydrogenase from P. gingivalis was first expressed and purified in Escherichia coli then characterized for enzymatic inhibitor screening studies. Several inhibitors with similar structural scaffolds containing a sulfonamide core and aromatic substituents showed dose-dependent inhibition. These compounds were further assayed showing reasonable whole-cell activity and the inhibition mechanism was determined. We conclude that the establishment of this target and screening strategy provides a model for the future development of new antimicrobials.  相似文献   

3.
meso-Diaminopimelate dehydrogenase (EC 1.4.1.16) was purified to homogeneity from Corynebacterium glutamicum ATCC 13032. The enzyme had a molecular weight of about 70,000 and consisted of two subunits identical in molecular weight. The enzyme was highly specific for meso-2,6-diaminopimelate. The pH optima for deamination and amination were about 9.8 and 7.9, respectively. The Michaelis constants were 3.1mm for meso-2,6-diaminopimelate, 0.12mm for NADP+, 0.28 mm for l-2-amino-6-ketopimelate, 36 mm for ammonia, and 0.13 mm for NADPH. d and l isomers of 2,6-diaminopimelate competitively inhibited the oxidative deamination of meso-2,6-diaminopimelate. The enzyme was distributed in a wider range of bacterial species than reported previously [Misono et al., J. Bacteriol., 137, 22 (1979)] when assayed by a sensitive formazan formation method.  相似文献   

4.
In order to enlarge the substrate binding pocket of the meso-diaminopimelate dehydrogenase from Symbiobacterium thermophilum to accommodate larger 2-keto acids, four amino acid residues (Phe146, Thr171, Arg181, and His227) were targeted for site saturation mutagenesis. Among all mutants, the single mutant H227V had a specific activity of 2.39 ± 0.06 U · mg−1, which was 35.1-fold enhancement over the wild-type enzyme.  相似文献   

5.
meso-Diaminopimelate dehydrogenase (meso-DAPDH) catalyzes the reversible NADP+-dependent oxidative deamination of meso-2,6-diaminopimelate (meso-DAP) to produce l-2-amino-6-oxopimelate. meso-DAPDH is divided into two major clusters, types I and II, based on substrate specificity and structural characteristic. Here, we describe a novel type II meso-DAPDH from Thermosyntropha lipolytica (TlDAPDH). The gene encoding a putative TlDAPDH was expressed in Escherichia coli cells, and then the enzyme was purified 7.3-fold to homogeneity from the crude cell extract. The molecule of TlDAPDH seemed to form a hexamer, which is the typical structural characteristic of type II meso-DAPDHs. The purified enzyme exhibited oxidative deamination activity toward meso-DAP with both NADP+ and NAD+ as coenzymes. TlDAPDH exhibited reductive amination activity of corresponding 2-oxo acid to produce d-amino acid. In particular, the productivities for d-aspartate and d-glutamate have not been reported in the type II enzymes. The optimum pH and temperature for oxidative deamination of meso-DAP were 10.5 and 55°C, respectively. TlDAPDH retained more than 80% of its activity after incubation for 30 min at temperatures between 50°C and 65°C and in the pH range of 4.5–9.5. Moreover, the coenzyme and substrate recognition mechanisms of TlDAPDH were elucidated based on a multiple sequence alignment and the homology model. The results of these analyses suggested that the molecular mechanisms for coenzyme and substrate recognition of TlDAPDH were similar to those of meso-DAPDH from S. thermophilum, which is the representative type II enzyme. Based on the kinetic characteristics and structural comparison, TlDAPDH was considered to be a novel type II meso-DAPDH.  相似文献   

6.
A simple and sensitive spectrophotometric method for determining meso-α,ε-diaminopime-late with meso-α,ε-diaminopimelate d-dehydrogenase (EC class 1.4.1) is described. meso-α,ε-Diaminopimelate was determined spectrophotometrically with the enzyme by measuring the NADPH formed (Procedure A) or the formazan produced by NADPH (Procedure B). A linear relationship was established between absorbance and the amount of amino acid (0.02-0.20 μmol). This method can be used to assay diaminopimelate epimerase (EC 5.1.1.7) and is applicable for determining meso-α,ε-diaminopimelate specifically in hydrolyzates of bacterial cell wall.  相似文献   

7.
The gene encoding the meso-diaminopimelate dehydrogenase (DAPDH) from Corynebacterium glutamicum was overexpressed and purified to homogeneity. Crystals of the binary DAPDH-NADP+ complex were obtained from solutions of polyethylene glycol 8000, 100 mM sodium cacodylate, pH 6.5, and 150–300 mM Mg(OAc)2. The crystals diffract to 2.2 Å, belong to the orthorhombic space group P21, and contain two molecules per asymmetric unit. © 1996 Wiley-Liss, Inc.  相似文献   

8.
Verrucomicrobium spinosum is a Gram-negative bacterium that is related to bacteria from the genus Chlamydia. The bacterium is pathogenic towards Drosophila melanogaster and Caenorhabditis elegans, using a type III secretion system to facilitate pathogenicity. V. spinosum employs the recently discovered l,l-diaminopimelate aminotransferase biosynthetic pathway to generate the bacterial cell wall and protein precursors diaminopimelate and lysine. A survey of the V. spinosum genome provides evidence that the bacterium should be able to synthesize peptidoglycan de novo, since all of the necessary genes are present. The enzyme UDP-N-acetylmuramoyl-l-alanyl-d-glutamate: meso-2,6-diaminopimelate ligase (MurE) (E.C. 6.3.2.15) catalyzes a reaction in the cytoplasmic step of peptidoglycan biosynthesis by adding the third amino acid residue to the peptide stem. The murE ortholog from V. spinosum (murE Vs) was cloned and was shown to possess UDP-MurNAc-l-Ala-d-Glu:meso-2,6-diaminopimelate ligase activity in vivo using functional complementation. In vitro analysis using the purified recombinant enzyme demonstrated that MurEVs has a pH optimum of 9.6 and a magnesium optimum of 30 mM. meso-Diaminopimelate was the preferred substrate with a K m of 17 µM, when compared to other substrates that are structurally related. Sequence alignment and structural analysis using homology modeling suggest that key residues that make up the active site of the enzyme are conserved in MurEVs. Our kinetic analysis and structural model of MurEVs is consistent with other MurE enzymes from Gram-negative bacteria that have been characterized. To verify that V. spinosum incorporates diaminopimelate into its cell wall, we purified peptidoglycan from a V. spinosum culture; analysis revealed the presence of diaminopimelate, consistent with that of a bona fide peptidoglycan from Gram-negative bacteria.  相似文献   

9.
Two diastereomeric analogs (1 and 2) of diaminopimelic acid (DAP) bearing an isoxazoline moiety were synthesized and evaluated for their inhibitory activities against meso-diaminopimelate dehydrogenase (m-Ddh) from the periodontal pathogen, Porphyromonas gingivalis. Compound 2 showed promising inhibitory activity against m-Ddh with an IC50 value of 14.9 µM at pH 7.8. The two compounds were further tested for their antibacterial activities against a panel of periodontal pathogens, and compound 2 was shown to be selectively potent to P. gingivalis strains W83 and ATCC 33277 with minimum inhibitory concentration (MIC) values of 773 µM and 1.875 mM, respectively. Molecular modeling studies revealed that the inversion of chirality at the C-5 position of these compounds was the primary reason for their different biological profiles. Based on these preliminary results, we believe that compound 2 has properties consistent with it being a lead compound for developing novel pathogen selective antibiotics to treat periodontal diseases.  相似文献   

10.
Murein peptide ligase (Mpl) is an enzyme found in Gram-negative bacteria. It catalyses the addition of tripeptide l-Ala-γ-d-Glu-meso-diaminopimelate to nucleotide precursor UDP-N-acetylmuramic acid during the recycling of peptidoglycan. Although not essential, this enzyme represents an interesting target for antibacterial compounds through the synthesis of alternate substrates whose incorporation into peptidoglycan might be deleterious for the bacterial cell. Therefore, we have synthesised 10 tripeptides l-Ala-γ-d-Glu-Xaa in which Xaa represents amino acids different from diaminopimelic acid. Tripeptide with Xaa = ε-d-Lys proved to be an excellent substrate of Escherichia coli Mpl in vitro. Tripeptides with Xaa = p-amino- or p-nitro-l-phenylalanine were poor substrates, while tripeptides with Xaa = d- or l-2-aminopimelate, dl-2-aminoheptanoic acid, l-Glu, l-norleucine, l-norvaline, l-2-aminobutyric acid or l-Ala were not substrates at all. Although a good Mpl substrate, the d-Lys-containing tripeptide was devoid of antibacterial activity against E. coli, presumably owing to poor uptake.  相似文献   

11.
Reimers, J. M., Huang, Q., Albe, K. R., and Wright, B. E. 1993. Purification and kinetic characterization of glucose-6-phosphate dehydrogenase from Dictyostelium discoideum. Experimental Mycology 17, 1-6. Glucose-6-phosphate dehydrogenase from Dictyostelium discoideum was purified 650-fold and kinetically characterized. The enzyme catalyzed the conversion of G6P + NADP to 6PG + NADPH stoichiometrically and irreversibly in vitro . The purified enzyme is specific for NADP. Michaelis constants for G6P and NADP were 0.040 and 0.011 mM, respectively. NADPH was found to be a competitive inhibitor with respect to NADP with a Ki of 0.006 mM and a noncompetitive inhibitor with respect to G6P. The data from initial velocity and product inhibition studies were consistent with a sequential mechanism.  相似文献   

12.
The meso-2,3-butanediol dehydrogenase (meso-BDH) from S. marcescens H30 is responsible for converting acetoin into 2,3-butanediol during sugar fermentation. Inactivation of the meso-BDH encoded by budC gene does not completely abolish 2,3-butanediol production, which suggests that another similar enzyme involved in 2,3-butanediol formation exists in S. marcescens H30. In the present study, a glycerol dehydrogenase (GDH) encoded by gldA gene from S. marcescens H30 was expressed in Escherichia coli BL21(DE3), purified and characterized for its properties. In vitro conversion indicated that the purified GDH could catalyze the interconversion of (3S)-acetoin/meso-2,3-butanediol and (3R)-acetoin/(2R,3R)-2,3-butanediol. (2S,3S)-2,3-Butanediol was not a substrate for the GDH at all. Kinetic parameters of the GDH enzyme showed lower K m value and higher catalytic efficiency for (3S/3R)-acetoin in comparison to those for (2R,3R)-2,3-butanediol and meso-2,3-butanediol, implying its physiological role in favor of 2,3-butanediol formation. Maximum activity for reduction of (3S/3R)-acetoin and oxidations of meso-2,3-butanediol and glycerol was observed at pH 8.0, while it was pH 7.0 for diacetyl reduction. The enzyme exhibited relative high thermotolerance with optimum temperature of 60 °C in the oxidation–reduction reactions. Over 60 % of maximum activity was retained at 70 °C. Additionally, the GDH activity was significantly enhanced for meso-2,3-BD oxidation in the presence of Fe2+ and for (3S/3R)-acetoin reduction in the presence of Mn2+, while several cations inhibited its activity, particularly Fe2+ and Fe3+ for (3S/3R)-acetoin reduction. The properties provided potential application for single configuration production of acetoin and 2,3-butanediol .  相似文献   

13.
Cell-free extracts of Methanobacterium thermoautotrophicum were found to contain high activities of the following oxidoreductases (at 60°C): pyruvate dehydrogenase (coenzyme A acetylating), 275 nmol/min per mg of protein; α-ketoglutarate dehydrogenase (coenzyme A acylating), 100 nmol/min per mg; fumarate reductase, 360 nmol/min per mg; malate dehydrogenase, 240 nmol/min per mg; and glyceraldehyde-3-phosphate dehydrogenase, 100 nmol/min per mg. The kinetic properties (apparent Vmax and KM values), pH optimum, temperature dependence of the rate, and specificity for electron acceptors/donors of the different oxidoreductases were examined. Pyruvate dehydrogenase and α-ketoglutarate dehydrogenase were shown to be two separate enzymes specific for factor 420 rather than for nicotinamide adenine dinucleotide (NAD), NADP, or ferredoxin as the electron acceptor. Both activities catalyzed the reduction of methyl viologen with the respective α-ketoacid and a coenzyme A-dependent exchange between the carboxyl group of the α-ketoacid and CO2. The data indicate that the two enzymes are similar to pyruvate synthase and α-ketoglutarate synthase, respectively. Fumarate reductase was found in the soluble cell fraction. This enzyme activity coupled with reduced benzyl viologen as the electron donor, but reduced factor 420, NADH, or NADPH was not effective. The cells did not contain menaquinone, thus excluding this compound as the physiological electron donor for fumarate reduction. NAD was the preferred coenzyme for malate dehydrogenase, whereas NADP was preferred for glyceraldehyde-3-phosphate dehydrogenase. The organism also possessed a factor 420-dependent hydrogenase and a factor 420-linked NADP reductase. The involvement of the described oxidoreductases in cell carbon synthesis is discussed.  相似文献   

14.
Klebsiella pneumoniae is known to produce meso-2,3-butanediol and 2S,3S-butanediol, whereas 2R,3R-butanediol was detected in the culture broth of K. pneumoniae CGMCC 1.6366. The ratio of 2R,3R-butanediol to all isomers obtained using glycerol as the carbon source was higher than that obtained using glucose as the carbon source. Therefore, enzymes involved in glycerol metabolism are likely related to 2R,3R-butanediol formation. In vitro reactions show that glycerol dehydrogenase catalyzes the stereospecific conversion of R-acetoin to 2R,3R-butanediol and S-acetoin to meso-2,3-butanediol. Butanediol dehydrogenase exhibits high (S)-enantioselectivity in ketone reduction. Genes encoding glycerol dehydrogenase, α-acetolactate decarboxylase, and butanediol dehydrogenase were individually disrupted in K. pneumoniae CGMCC 1.6366, and the 2,3-butanediol synthesis characteristics of these mutants were investigated. K. pneumoniae ΔdhaD lost the ability to synthesize 2R,3R-butanediol. K. pneumoniae ΔbudA showed reduced 2R,3R-butanediol synthesis. However, K. pneumoniae ΔbudC produced a high level of 2R,3R-butanediol, and R-acetoin was accumulated in the broth. The metabolic characteristics of these mutants and in vitro experiment results demonstrated the mechanism of the 2,3-butanediol stereoisomer synthesis pathway. Glycerol dehydrogenase, encoded by dhaD, exhibited 2R,3R-butanediol dehydrogenase activity and was responsible for 2R,3R-butanediol synthesis from R-acetoin. This enzyme also contributed to meso-2,3-butanediol synthesis from S-acetoin. Butanediol dehydrogenase, encoded by budC, was the only enzyme that catalyzed the conversion of diacetyl to S-acetoin and further to 2S,3S-butanediol.  相似文献   

15.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

16.
ω-Hydroxyfatty acid dehydrogenase (ω-hydroxyfatty acid:NADP oxidoreductase) catalyzes the reaction ω-hydroxyfatty acid + NADP ? ω-oxofatty acid + NADPH +H+. In wound-healing potato tuber disks, the ω-oxofatty acid generated by this enzyme is further oxidized to the corresponding dicarboxylic acid by a separate enzyme, ω-oxofatty acid dehydrogenase. ω-Hydroxy acid dehydrogenase, but not ω-oxo acid dehydrogenase, was found to be induced by wounding potato tubers. ω-Hydroxy acid dehydrogenase has been purified 600-fold to near homogeneity from wound-healing potato tuber disks by a combination of gel filtration, anion-exchange, and hydroxylapatite chromatography followed by NADP-Sepharose affinity chromatography, in about 1% yield. The molecular weight and Stokes radius of this enzyme as determined by gel exclusion chromatography are 60,000 and 31 Å, respectively. Sodium dodecyl sulfate-gel electrophoresis gave a molecular weight of 31,000, indicating that the deydrogenase is a dimer with subunits of similar molecular weight. The pH optima for the reaction in the forward and reverse directions are 9.5 and 8.5, respectively, and V in the forward and reverse directions are 140 and 3200 nmol/min/mg, respectively. Apparent Km values for NADP, 16-hydroxyhexadecanoic acid, NADPH, and 16-oxohexadecanoic acid are 100, 20, 5, and 7 μm respectively. The equilibrium constant of the reaction at pH 9.5 and 30 °C is 1.4 × 10?9m. The enzyme preparation did not show any stereospecificity for hydride transfer from NADPH to 16-oxohexadecanoic acid.  相似文献   

17.
A simple method for the amide group location in monoamidated peptides of bacterial cell wall peptidoglycans is presented. It is based on the high substrate specificities of two Bacillus sphaericus endopeptidases I and II. These enzymes catalyze the hydrolysis of γ-d-glutamyl-(l)meso-diaminopimelate linkages in tetrapeptides such as l-Ala-d-iso-Gln-(l)ms-A2pm(l)-d-Ala or l-Ala-γ-d-Glu-(l)ms-A2pm(ωNH2)-d-Ala. After hydrolysis the resulting dipeptides, l-Ala-d-iso-Gln and ms-A2pm(l)-d-Ala or l-Ala-d-Glu and ms-A2pm(ωNH2)-d-Ala, are resolved by thin-layer chromatography. Thus, the dipeptides can be identified and the amide group subsequently located in the original tetrapeptide. This method should be useful for the identification of the monoamidated tetrapeptide moiety of bacterial cell wall peptidoglycans.  相似文献   

18.
Practical uses of a novel alcohol dehydrogenase from Thermoanaerobium brockii have been examined in crude and purified form. Stoichiometric reduction of NADP (50 mg) was demonstrated with agarose-immobilized enzyme and 0.3 (v/v) 2-propanol solution as reductant. A coenzyme recycle number of 20000 was achieved in enzymatic reactions that employed the alcohol dehydrogenase for NADPH/NADP regeneration. Gram-scale synthesis of chiral R(+) 2-pentanol was shown in a system composed of enzyme, 2-pentanone and 2-propanol as reductant. The effect of temperature, reaction time and substrate concentration on alcohol optical purity was examined. An optical purity of 80% was achieved in the enzymatic synthesis of R(+) 2-pentanol. The enzyme was easily immobilized and stable on an enzyme electrode for analytical detection of alcohols and carbonyls. T. brockii enzyme has potential applications as a commercial alcohol dehydrogenase because of broad substrate specificity and activity at high temperature or high solvent concentration, rare carbonyl si-face stereo-specificity in hydrogen transfer, and high stability and activation of immobilized enzyme.  相似文献   

19.
SYNOPSIS. Nicotinamide adenine dinucleotide phosphate (NADP) and nicotinamide adenine dinucleotide (NAD) linked isocitrate dehydrogenase and NADP linked alcohol dehydrogenase have been detected in Euglena gracilis var. bacillaris. The NADP isocitrate dehydrogenase showed half-maximal activity at a concentration of 3 × 10?5 M DL-isocitrate, but did not follow simple Michaelis-Menten kinetics with respect to substrate concentration. The optimal NADP concentration was about 0.06 mM, and activity fell off sharply on either side of this optimum. Fresh preparations of the enzyme migrated as single bands in disc electrophoresis, but two enzymatically active bands were present after frozen storage. The NAD isocitrate dehydrogenase followed Michaelis-Menten kinetics with respect to substrate. In crude extracts, no requirement for adenosine monophosphate, adenosine diphosphate, or sulfhydryl compounds could be found. NADP alcohol dehydrogenase activity could be found with either ethanol or propanol as substrate. Low concentrations of coenzyme A were moderately inhibitory. In tris(hydroxymethyl) aminomethane buffer (tris buffer), Euglena extracts reduced NAD slowly in the absence of exogenous substrate. In the absence of tris, no such reduction occurred. A similar phenomenon was observed with NADP.  相似文献   

20.
Spinach chloroplast glyceraldehyde phosphate dehydrogenase (d-glyceraldehyde-3-phosphate: NADP oxidoreductase, phosphorylating; EC 1.2.1.13) is an equilibrium mixture of aggregates of a basic protomer (Mr about 145,000) and is active with both NADP and NAD. The enzyme is primarily “tetrameric” (Mr about 600,000), although minor amounts of smaller and larger oligomers are also found. Gel chromatography in buffer containing 30 μm NADP results in depolymerization of the enzyme, mainly to protomers. NAD does not dissociate and counteracts this effect of NADP.The apparent Km values of the protomers are 7 μm (NADP) and 8 μm (NAD). The aggregates with a Mr > 106 have properties similar to the protomers. The tetramer as first isolated has higher Mm values for NADP (380 μm) and NAD (48 μm), but its apparent affinity for NADP is further decreased by repeated gel filtrations in buffer or by a single one in buffer containing NAD. Such preparations display nonlinear kinetics when NADP is the varied substrate and have a Km (NADP) of about 1.5–3.3 μm. All these effects are reversible.V values are apparently the same in all enzyme forms and the V (NADP)V (NAD) ratio always approaches 2. Since, however, the enzyme is presumably dissociated by the NADP concentrations required for a “saturating” assay, the significance of V (NADP) seems questionable.  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