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 62 毫秒
1.
Ko J  Kim I  Yoo S  Min B  Kim K  Park C 《Journal of bacteriology》2005,187(16):5782-5789
Methylglyoxal (MG) is a toxic metabolite known to accumulate in various cell types. We detected in vivo conversion of MG to acetol in MG-accumulating Escherichia coli cells by use of (1)H nuclear magnetic resonance ((1)H-NMR) spectroscopy. A search for homologs of the mammalian aldo-keto reductases (AKRs), which are known to exhibit activity to MG, revealed nine open reading frames from the E. coli genome. Based on both sequence similarities and preliminary characterization with (1)H-NMR for crude extracts of the corresponding mutant strains, we chose five genes, yafB, yqhE, yeaE, yghZ, and yajO, for further study. Quantitative assessment of the metabolites produced in vitro from the crude extracts of these mutants and biochemical study with purified AKRs indicated that the yafB, yqhE, yeaE, and yghZ genes are involved in the conversion of MG to acetol in the presence of NADPH. When we assessed their in vivo catalytic activities by creating double mutants, all of these genes except for yqhE exhibited further sensitivities to MG in a glyoxalase-deficient strain. The results imply that the glutathione-independent detoxification of MG can occur through multiple pathways, consisting of yafB, yqhE, yeaE, and yghZ genes, leading to the generation of acetol.  相似文献   

2.
An NADPH-dependent enzyme that reduces ethyl 2-methylacetoacetate stereoselectively to ethyl (2R)-methyl-(3S)-hydroxybutanoate was purified 730-fold from Escherichia coli. The N-terminal amino acid sequence data obtained from the purified reductase were used to search the E. coli genome, and a single match was found at the start of the yqhE open reading frame. The YqhE protein had been identified previously by Yum et al. as a 2,5-diketo-D-gluconate reductase on the basis of sequence similarity to other bacterial homologues [Yum, D.-Y.; Lee, B.-Y.; Pan, J.-G. Appl.Environ. Microbiol. 1999, 65, 3341-3346]; however, it had not been examined for beta-keto ester reductions. Our results thus link a key enzyme in the microbial production of ascorbate with stereoselective beta-keto ester reductions, two important fields in biocatalysis. The purified YqhE reductase accepts ethyl acetoacetate and a variety of 2-substituted derivatives, and its sequence is similar to other aldose reductase superfamily members that also reduce alpha-substituted beta-keto esters to syn-(2R,3S) alcohols.  相似文献   

3.
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   

4.
2,5-Diketo-D-gluconate reductase, a novel enzyme that catalyzes the stereospecific NADPH-dependent reduction of 2,5-diketo-D-gluconate to 2-keto-L-gulonate, has been purified to homogeneity by sequential anion exchange, Cibacron blue F3GA affinity, and gel permeation chromatography from Corynebacterium sp. ATCC 31090. Molecular weight of the native form, determined by gel permeation chromatography, is 35,000 +/- 2,000. The subunit molecular weight, determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis is 34,000; thus, the enzyme is active as a monomer. A pI value of 4.4 is measured for the enzyme. Amino- and carboxyl-terminal sequences are consistent with that predicted by the DNA sequence of the reductase gene. At 25 degrees C, pH 6.4, the turnover number is 500 min-1, and the apparent Km values for 2,5-diketo-D-gluconate and NADPH are 26 mM and 10 microM, respectively. The enzyme is specific for NADPH, but the sugar binding site will also accept 5-keto-D-fructose and dihydroxyacetone as substrates. The enzyme is active over a broad pH range (pH 5-8) for the reduction of 2,5-diketo-D-gluconate; a sharp optimum at pH 9.2 is observed for the oxidation of 2-keto-L-gulonate. A Keq value of 5.6 X 10(-13) M indicates that reduction of substrate by NADPH is highly preferred. An activation energy of 12.3 kcal mol-1 is measured. Enzyme turnover is slow relative to dehydration of the gem-diol at C-5 of the substrate.  相似文献   

5.
2,5-diketo-D-gluconic acid reductase (2,5-DKGR; E.C. 1.1.1.-) catalyzes the Nicotinamide adenine dinucleotide phosphate (NADPH)-dependent stereo-specific reduction of 2, 5-diketo-D-gluconate (2,5-DKG) to 2-keto-L-gulonate (2-KLG), a precursor in the industrial production of vitamin C (L-ascorbate). Microorganisms that naturally ferment D-glucose to 2,5-DKG can be genetically modified to express the gene for 2,5-DKGR, and thus directly produce vitamin C from D-glucose. Two naturally occurring variants of DKGR (DKGR A and DKGR B) have been reported. DKGR B exhibits higher specific activity toward 2,5-DKG than DKGR A; however, DKGR A exhibits a greater selectivity for this substrate and significantly higher thermal stability. Thus, a modified form of DKGR, combining desirable properties from both enzymes, would be of substantial commercial interest. In the present study we use a molecular dynamics-based approach to understand the conformational changes in DKGR A as the active site is mutated to include two active site residue changes that occur in the B form. The results indicate that the enhanced kinetic properties of the B form are due, in part, to residue substitutions in the binding pocket. These substitutions augment interactions with the substrate or alter the alignment with respect to the putative proton donor group. Proteins 2000;39:68-75.  相似文献   

6.
The gene for ribonucleotide reductase from Anabaena sp. strain PCC 7120 was identified and expressed in Escherichia coli. This gene codes for a 1,172-amino-acid protein that contains a 407-amino-acid intein. The intein splices itself from the protein when it is expressed in E. coli, yielding an active ribonucleotide reductase of 765 residues. The mature enzyme was purified to homogeneity from E. coli extracts. Anabaena ribonucleotide reductase is a monomer with a molecular weight of approximately 88,000, as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and Superose 12 column chromatography. The enzyme reduces ribonucleotides at the triphosphate level and requires a divalent cation and a deoxyribonucleoside triphosphate effector. The enzyme is absolutely dependent on the addition of the cofactor, 5'-adenosylcobalamin. These properties are characteristic of the class II-type reductases. The cyanobacterial enzyme has limited sequence homology to other class II reductases; the greatest similarity (38%) is to the reductase from Lactobacillus leichmannii. In contrast, the Anabaena reductase shows over 90% sequence similarity to putative reductases found in genome sequences of other cyanobacteria, such as Nostoc punctiforme, Synechococcus sp. strain WH8102, and Prochlorococcus marinus MED4, suggesting that the cyanobacterial reductases form a closely related subset of the class II enzymes.  相似文献   

7.
8.
维生素C(Vitamin C,简称Vc),又称L-抗坏血酸(L-Ascorbic acid)是人体必需的维生素,生理作用广泛,在医药和食品工业上均有重要地位。目前国内厂家多以我国发明的“二步发酵法”进行生产,即以D-山梨醇为原料生产2-酮基-L-古龙酸(以下简称2-KLG),然后制备维生素C。而近年来引起国内外普遍关注的是从D-葡萄糖串联发酵生产2-KLG的新工艺,以及采用基因工程技术,构建直接由D-葡萄糖转化生成2-KLG的基因工程菌的研究(图1)。1987年以来我国学者尹光琳等人采用了欧文氏菌(Erwinia sp.)和棒状杆菌(Corynebacterium sp.)进行串联发酵产生维生素C前体——2-酮基-L-古龙酸,并开展了一系列的研究。  相似文献   

9.
Adapting metabolic enzymes of microorganisms to low temperature environments may require a difficult compromise between velocity and affinity. We have investigated catalytic efficiency in a key metabolic enzyme (dihydrofolate reductase) of Moritella profunda sp. nov., a strictly psychrophilic bacterium with a maximal growth rate at 2 degrees C or less. The enzyme is monomeric (Mr=18,291), 55% identical to its Escherichia coli counterpart, and displays Tm and denaturation enthalpy changes much lower than E. coli and Thermotoga maritima homologues. Its stability curve indicates a maximum stability above the temperature range of the organism, and predicts cold denaturation below 0 degrees C. At mesophilic temperatures the apparent Km value for dihydrofolate is 50- to 80-fold higher than for E. coli, Lactobacillus casei, and T. maritima dihydrofolate reductases, whereas the apparent Km value for NADPH, though higher, remains in the same order of magnitude. At 5 degrees C these values are not significantly modified. The enzyme is also much less sensitive than its E. coli counterpart to the inhibitors methotrexate and trimethoprim. The catalytic efficiency (kcat/Km) with respect to dihydrofolate is thus much lower than in the other three bacteria. The higher affinity for NADPH could have been maintained by selection since NADPH assists the release of the product tetrahydrofolate. Dihydrofolate reductase adaptation to low temperature thus appears to have entailed a pronounced trade-off between affinity and catalytic velocity. The kinetic features of this psychrophilic protein suggest that enzyme adaptation to low temperature may be constrained by natural limits to optimization of catalytic efficiency.  相似文献   

10.
2,4-Dienoyl-CoA reductases, enzymes of the beta-oxidation of unsaturated fatty acids which were purified from bovine liver and oleate-induced cells of Escherichia coli, revealed very similar substrate specificities but distinctly different molecular properties. The subunit molecular weights, estimated by sodium dodecyl sulfate-polyacrylamide gel electrophoresis were 32,000 and 73,000 for the mammalian and the bacterial enzyme, respectively. The native molecular weights, calculated from sedimentation coefficients and Stokes radii yielded 124,000 for the bovine liver and 70,000 for the bacterial enzyme. Thus, bovine liver 2,4-dienoyl-CoA reductase is a tetramer consisting of four identical subunits. The E. coli 2,4-dienoyl-CoA reductase, however, possesses a monomeric structure. The latter enzyme contains 1 mol of FAD/mol of enzyme, whereas the former reductase is not a flavoprotein. The bovine liver reductase reduced 2-trans, 4-cis- and 2-trans,4-trans-decadienoyl-CoA to 3-trans-decenoyl-CoA. The E. coli reductase catalyzed the reduction of the same two substrates but in contrast yielded 2-trans-decenoyl-CoA as reaction product. Certain other properties of the two 2,4-dienoyl-CoA reductases are also presented. The localization of the reductase step within the degradation pathway of 4-cis-decenoyl-CoA, a metabolite of linoleic acid, is discussed.  相似文献   

11.
A cytosolic aldo-keto reductase was purified from Saccharomyces cerevisiae ATCC 26602 to homogeneity by affinity chromatography, chromatofocusing, and hydroxylapatite chromatography. The relative molecular weights of the aldo-keto reductase as determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and size exclusion chromatography were 36,800 and 35,000, respectively, indicating that the enzyme is monomeric. Amino acid composition and N-terminal sequence analysis revealed that the enzyme is closely related to the aldose reductases of xylose-fermenting yeasts and mammalian tissues. The enzyme was apparently immunologically unrelated to the aldose reductases of other xylose-fermenting yeasts. The aldo-keto reductase is NADPH specific and catalyzes the reduction of a variety of aldehydes. The best substrate for the enzyme is the aromatic aldehyde p-nitrobenzaldehyde (Km = 46 microM; kcat/Km = 52,100 s-1 M-1), whereas among the aldoses, DL-glyceraldehyde was the preferred substrate (Km = 1.44 mM; kcat/Km = 1,790 s-1 M-1). The enzyme failed to catalyze the reduction of menadione and p-benzoquinone, substrates for carbonyl reductase. The enzyme was inhibited only slightly by 2 mM sodium valproate and was activated by pyridoxal 5'-phosphate. The optimum pH of the enzyme is 5. These data indicate that the S. cerevisiae aldo-keto reductase is a monomeric NADPH-specific reductase with strong similarities to the aldose reductases.  相似文献   

12.
A flavin reductase, which is naturally part of the ribonucleotide reductase complex of Escherichia coli, acted in cell extracts of recombinant E. coli strains under aerobic and anaerobic conditions as an "azo reductase." The transfer of the recombinant plasmid, which resulted in the constitutive expression of high levels of activity of the flavin reductase, increased the reduction rate for different industrially relevant sulfonated azo dyes in vitro almost 100-fold. The flavin reductase gene (fre) was transferred to Sphingomonas sp. strain BN6, a bacterial strain able to degrade naphthalenesulfonates under aerobic conditions. The flavin reductase was also synthesized in significant amounts in the Sphingomonas strain. The reduction rates for the sulfonated azo compound amaranth were compared for whole cells and cell extracts from both recombinant strains, E. coli, and wild-type Sphingomonas sp. strain BN6. The whole cells showed less than 2% of the specific activities found with cell extracts. These results suggested that the cytoplasmic anaerobic "azo reductases," which have been described repeatedly in in vitro systems, are presumably flavin reductases and that in vivo they have insignificant importance in the reduction of sulfonated azo compounds.  相似文献   

13.
Two sequences with homology to a thioredoxin oligonucleotide probe were detected by Southern blot analysis of Anabaena sp. strain PCC 7120 genomic DNA. One of the sequences was shown to code for a protein with 37% amino acid identity to thioredoxins from Escherichia coli and Anabaena sp. strain PCC 7119. This is in contrast to the usual 50% homology observed among most procaryotic thioredoxins. One gene was identified in a library and was subcloned into a pUC vector and used to transform E. coli strains lacking functional thioredoxin. The Anabaena strain 7120 thioredoxin gene did not complement the trxA mutation in E. coli. Transformed cells were not able to use methionine sulfoxide as a methionine source or support replication of T7 bacteriophage or the filamentous viruses M13 and f1. Sequence analysis of a 720-base-pair TaqI fragment indicated an open reading frame of 115 amino acids. The Anabaena strain 7120 thioredoxin gene was expressed in E. coli, and the protein was purified by assaying for protein disulfide reductase activity, using insulin as a substrate. The Anabaena strain 7120 thioredoxin exhibited the properties of a conventional thioredoxin. It is a small heat-stable redox protein and an efficient protein disulfide reductase. It is not a substrate for E. coli thioredoxin reductase. Chemically reduced Anabaena strain 7120 thioredoxin was able to serve as reducing agent for both E. coli and Anabaena strain 7119 ribonucleotide reductases, although with less efficiency than the homologous counterparts. The Anabaena strain 7120 thioredoxin cross-reacted with polyclonal antibodies to Anabaena strain 7119 thioredoxin. However, this unusual thioredoxin was not detected in extracts of Anabaena strain 7120, and its physiological function is unknown.  相似文献   

14.
2-Keto-L-gulonic acid was produced from gluconic acid using co-immobilized cells of Gluconobacter oxydans and Corynebacterium sp. with 2,5-diketo-D-gluconic acid. Gluconobacter oxydans and Corynebacterium sp. were entrapped together with polyvinylalcohol and alginate. 50 g/l glucose, 50 g/l gluconic acid, and the mixture of equal volume of 50 g/l glucose and 50 g/l gluconic acid were used as substrates. When the ratio of two cells was 1 to 1 with 100 mg cells/ml, the conversion of 2-KLG from gluconic acid was 38% (g/g). © Rapid Science Ltd. 1998  相似文献   

15.
We isolated Gluconobacter oxydans T-100 that had an activity to produce 2-KLGA from D-sorbitol; however, the yield of 2-KLGA was quite insufficient. Therefore, enzymes involved in the biosynthesis of L-sorbosone and 2-KLGA, L-sorbose dehydrogenase (SDH) and L-sorbosone dehydrogenase (SNDH), respectively, were purified from G. oxydans T-100. A genomic library of G. oxydans T-100 was screened to clone both genes for SDH and SNDH based on their amino acid sequences. SNDH and SDH were encoded in sequential open reading frames with 1497 and 1596 nucleotides, respectively, which were verified by the expression in Escherichia coli. The amino acid sequence of SDH and SNDH showed close similarity with E. coli choline dehydrogenase (CDH) and betaine-aldehyde dehydrogenase (BADH), respectively, which cooperatively play a key role for conferring osmotic tolerance. Because the yield of 2-KLGA by G. oxydans introduced with the genes for SDH and SNDH were insufficient, replacement of the promoter with that of Escherichia coli tufB1 in combination with chemical mutagenesis by N-methyl-N'-nitro-N-nitrosoguanidine resulted in improvement of the production level.  相似文献   

16.
Glutathione synthetase from Escherichia coli B showed amino acid sequence homology with mammalian and bacterial dihydrofolate reductases over 40 residues, although these two enzymes are different in their reaction mechanisms and ligand requirements. The effects of ligands of dihydrofolate reductase on the reaction of E. coli B glutathione synthetase were examined to find resemblances in catalytic function to dihydrofolate reductase. The E. coli B enzyme was potently inhibited by 7,8-dihydrofolate, methotrexate, and trimethoprim. Methotrexate was studied in detail and proved to bind to an ATP binding site of the E. coli B enzyme with K1 value of 0.1 mM. The homologous portion of the amino acid sequence in dihydrofolate reductases, which corresponds to the portion coded by exon 3 of mammalian dihydrofolate reductase genes, provided a binding site of the adenosine diphosphate moiety of NADPH in the crystal structure of dihydrofolate reductase. These analyses would indicate that the homologous portion of the amino acid sequence of the E. coli B enzyme provides the ATP binding site. This report gives experimental evidence that amino acid sequences related by sequence homology conserve functional similarity even in enzymes which differ in their catalytic mechanisms.  相似文献   

17.
In addition to the ability of Penicillium notatum to grow on sucrose, glucose, fructose and gluconate, substantial growth occurred on 2-ketogluconate and 5-ketogluconate thereby indicating a diverse sugar metabolism. Cell-free extracts contained all the enzymes of the Embden-Meyerhof-Parnas pathway and for both oxidative and non-oxidative pentose phosphate metabolism. Despite inconsistencies in results between different assay methods for the conventional Entner-Doudoroff (ED) enzymes, the data indicated the route was enzymatically possible. Demonstrations of the activities of the enzymes of the non-phosphorylative equivalent of the ED pathway were achieved. No evidence was found of a phosphorylative linking enzyme between the two pathways. Both 2- and 5-ketogluconate reductases were detected along with gluconate dehydrogenase which suggested interconvertibility between the ketogluconates and gluconate. However, ketogluconokinase, responsible for the conversion of ketogluconate to 2-keto-6-phosphogluconate, was not detected. A scheme for the inter-relationships of routes of gluconate metabolism is discussed.  相似文献   

18.
Two 2,5-diketo-d-gluconate reductases, I and II, were purified respectively 918-fold and 28-fold from a mutant strain derived from Corynebacterium sp. SHS 0007. The enzymes appeared to be homogeneous on polyacrylamide gel electrophoresis. Both reductases converted 2,5-diketo-d-gluconate to 2-keto-l-gulonate in the presence of NADPH and seemed to be active only for reduction. The molecular weights of reductases I and II were estimated to be 29,000 and 34,000, respectively; and both were monomeric. Their isoelectric points were respectively pH 4.3 and pH 4.1. The optimum pH was 6.0 to 7.0 for reductase I, and 6.0 to 7.5 for reductase II. The Km values (pH 7.0, 30°C) of reductase I for 2,5-diketo-d-gluconate and for NADPH were 1.8 mM and 12 μM, respectively; and the corresponding values of reductase II were 13.5 mM and 13 μM. Both reductases converted 5-keto-d-fructose to l-sorbose in the presence of NADPH.  相似文献   

19.
Nitrate reductase from the haloalkalophilic denitrifying bacterium Halomonas sp. strain AGJ 1-3 was isolated and purified to homogeneity. The isolated enzyme belongs to a novel family of molybdenum-free nitrate reductases. It presents as a 130-140 kD monomeric protein with specific activity of 250 micromol/min per mg protein. The enzyme reduces not only nitrate, but also other anions, thus showing polyoxoanion reductase activity. Enzyme activity was maximal at pH 7.0 and 70-80 degrees C.  相似文献   

20.
研究了在10L发酵罐中D-葡萄糖串联发酵生产维生素C前体——2-酮基-L-古龙酸的发酵工艺条件。第一步发酵采用欧文氏菌(Erwinia sp.)的突变株SCB247,培养36小时,可将D-葡萄糖转化成中间体2,5-二酮基-D-葡萄糖酸,在发酵液中约累积180mg/ml。第二步发酵采用棒状杆菌(Corynebacterium sp.)SCB3058,可将2,5-二酮基-D-葡萄糖酸专一性地还原生成2-酮基-L-古龙酸。在细胞生长进入对数生长期后期时,加入经十二烷基硫酸钠处理的第一  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