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 gene encoding (6R)-2,2,6-trimethyl-1,4-cyclohexanedione (levodione) reductase was cloned from the genomic DNA of the soil isolate bacterium Corynebacterium aquaticum M-13. The gene contained an open reading frame consisting of 801 nucleotides corresponding to 267 amino acid residues. The deduced amino acid sequence showed approximately 35% identity with other short chain alcohol dehydrogenase/reductase (SDR) superfamily enzymes. The probable NADH-binding site and three catalytic residues (Ser-Tyr-Lys) were conserved. The enzyme was sufficiently produced in recombinant Escherichia coli cells using an expression vector pKK223-3, and purified to homogeneity by two-column chromatography steps. The enzyme purified from E. coli catalyzed stereo- and regio-selective reduction of levodione, and was strongly activated by monovalent cations, such as K+, Na+, and NH4+, as was the case of that from C. aquaticum M-13. To our knowledge, this is the first sequencing report of a monovalent cation-activated SDR enzyme.  相似文献   

2.
(6R)-2,2,6-Trimethyl-1,4-cyclohexanedione (levodione) reductase was isolated from a cell extract of the soil isolate Corynebacterium aquaticum M-13. This enzyme catalyzed regio- and stereoselective reduction of levodione to (4R,6R)-4-hydroxy-2,2, 6-trimethylcyclohexanone (actinol). The relative molecular mass of the enzyme was estimated to be 142,000 Da by high-performance gel permeation chromatography and 36,000 Da by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme required NAD(+) or NADH as a cofactor, and it catalyzed reversible oxidoreduction between actinol and levodione. The enzyme was highly activated by monovalent cations, such as K(+), Na(+), and NH(4)(+). The NH(2)-terminal and partial amino acid sequences of the enzyme showed that it belongs to the short-chain alcohol dehydrogenase/reductase family. This is the first report of levodione reductase.  相似文献   

3.
(6R)-2,2,6-Trimethyl-1,4-cyclohexanedione (levodione) reductase was isolated from a cell extract of the soil isolate Corynebacterium aquaticum M-13. This enzyme catalyzed regio- and stereoselective reduction of levodione to (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone (actinol). The relative molecular mass of the enzyme was estimated to be 142,000 Da by high-performance gel permeation chromatography and 36,000 Da by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The enzyme required NAD+ or NADH as a cofactor, and it catalyzed reversible oxidoreduction between actinol and levodione. The enzyme was highly activated by monovalent cations, such as K+, Na+, and NH4+. The NH2-terminal and partial amino acid sequences of the enzyme showed that it belongs to the short-chain alcohol dehydrogenase/reductase family. This is the first report of levodione reductase.  相似文献   

4.
A practical enzymatic synthesis of a doubly chiral key compound, (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone, starting from the readily available 2,6,6-trimethyl-2-cyclohexen-1,4-dione is described. Chirality is first introduced at the C-6 position by a stereoselective enzymatic hydrogenation of the double bond using old yellow enzyme 2 of Saccharomyces cerevisiae, expressed in Escherichia coli, as a biocatalyst. Thereafter, the carbonyl group at the C-4 position is reduced selectively and stereospecifically by levodione reductase of Corynebacterium aquaticum M-13, expressed in E. coli, to the corresponding alcohol. Commercially available glucose dehydrogenase was also used for cofactor regeneration in both steps. Using this two-step enzymatic asymmetric reduction system, 9.5 mg of (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone/ml was produced almost stoichiometrically, with 94% enantiomeric excess in the presence of glucose, NAD(+), and glucose dehydrogenase. To our knowledge, this is the first report of the application of S. cerevisiae old yellow enzyme for the production of a useful compound.  相似文献   

5.
The (6R)-2,2,6-trimethyl-1,4-cyclohexanedione (levodione) reductase (LVR) of the soil isolate bacterium Corynebacterium aquaticum M-13 is a NAD(H)-linked enzyme that catalyzes reversible oxidoreduction between (4R)-hydroxy-(6R)-2,2,6-trimethylcyclohexanone (actinol) and levodione. Here the crystal structure of a ternary complex of LVR with NADH and its inhibitor 2-methyl-2,4-pentanediol has been determined by molecular replacement and refined at 1.6-A resolution with a crystallographic R factor of 0.199. The overall structure is similar to those of other short-chain alcohol dehydrogenase/reductase enzymes. The positions of NADH and 2-methyl-2,4-pentanediol indicate the binding site of the substrate and identify residues that are likely to be important in the catalytic reaction. Modeling of the substrate binding in the active site suggests that the specificity of LVR is determined by electrostatic interactions between the negatively charged surface of Glu-103 of LVR and the positively charged surface on the re side of levodione. Mutant LVR enzymes in which Glu-103 is substituted with alanine (E103A), glutamine (E103Q), asparagines (E103N), or aspartic acid (E103D) show a 2-6-fold increase in Km values as compared with wild-type LVR and a much lower enantiomeric excess of the reaction products (60%) than the wild-type enzyme (95%). Together, these data indicate that Glu-103 has an important role in determining the stereospecificity of LVR.  相似文献   

6.
A practical enzymatic synthesis of a doubly chiral key compound, (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone, starting from the readily available 2,6,6-trimethyl-2-cyclohexen-1,4-dione is described. Chirality is first introduced at the C-6 position by a stereoselective enzymatic hydrogenation of the double bond using old yellow enzyme 2 of Saccharomyces cerevisiae, expressed in Escherichia coli, as a biocatalyst. Thereafter, the carbonyl group at the C-4 position is reduced selectively and stereospecifically by levodione reductase of Corynebacterium aquaticum M-13, expressed in E. coli, to the corresponding alcohol. Commercially available glucose dehydrogenase was also used for cofactor regeneration in both steps. Using this two-step enzymatic asymmetric reduction system, 9.5 mg of (4R,6R)-4-hydroxy-2,2,6-trimethylcyclohexanone/ml was produced almost stoichiometrically, with 94% enantiomeric excess in the presence of glucose, NAD+, and glucose dehydrogenase. To our knowledge, this is the first report of the application of S. cerevisiae old yellow enzyme for the production of a useful compound.  相似文献   

7.
Flavin reductase plays an important biological role in catalyzing the reduction of flavin by NAD(P)H oxidation. The gene that codes for flavin reductase from Citrobacter freundii A1 was cloned and expressed in Escherichia coli BL21(DE3)pLysS. In this study, we aimed to characterize the purified recombinant flavin reductase of C. freundii A1. The recombinant enzyme was purified to homogeneity and the biochemical profiles, including the effect of pH, temperature, metal ions and anions on flavin reductase activity and stability, were determined. This enzyme exhibited optimum activity at 45 °C in a 10-min reaction at pH 7.5 and was stable at temperatures up to 30 °C. At 0.1 mM concentration of metal ions, flavin reductase activity was stimulated by divalent cations including Mn2+, Sr2+, Ni2+, Sn2+, Ba2+, Co2+, Mg2+, Ca2+ and Pb2+. Ag+ was noticeably the strongest inhibitor of recombinant flavin reductase of C. freundii A1. This enzyme should not be defined as a standard flavoprotein. This is the first attempt to characterize flavin reductase of C. freundii origin.  相似文献   

8.
The gene encoding phenylacetaldehyde reductase (PAR), a useful biocatalyst for producing chiral alcohols, was cloned from the genomic DNA of the styrene-assimilating Corynebacterium sp. strain ST-10. The gene contained an opening reading frame consisting of 1,158 nucleotides corresponding to 385 amino acid residues. The subunit molecular weight was calculated to be 40,299, which was in agreement with that determined by polyacrylamide gel electrophoresis. The enzyme was sufficiently expressed in recombinant Escherichia coli cells for practical use and purified to homogeneity by three-column chromatography steps. The predicted amino acid sequence displayed only 20–29% identity with zinc-containing, NAD+-dependent, long-chain alcohol dehydrogenases. Nevertheless, the probable NAD+- and zinc-binding sites are conserved although one of the three catalytic zinc-binding residues of the zinc-containing, long-chain alcohol dehydrogenases was substituted by Asp in PAR. The protein contains 7.6 mol zinc/mol tetramer. Therefore, the enzyme was considered as a new member of zinc-containing, long-chain alcohol dehydrogenases with a particular and broad substrate specificity. Received: 5 March 1999 / Received last revision: 10 May 1999 / Accepted: 16 May 1999  相似文献   

9.
Ketopantoic acid (KPA) reductase catalyzes the stereospecific reduction of ketopantoic acid to d-pantoic acid. Based on the N-terminal amino acid sequence of KPA reductase from Stenotrophomonas maltophilia 845, the KPA reductase gene was cloned from S. maltophilia NBRC14161 and sequenced. This gene contains an open reading frame of 777 bp encoding 258 amino acid residues, and the deduced amino acid sequence showed high similarity to the SDR superfamily proteins. An expression vector, pETSmKPR, containing the full KPA reductase gene was constructed and introduced into Escherichia coli BL21 (DE3) to overexpress the enzyme. Bioreduction of KPA using E. coli transformant cells coexpressing KPA reductase together with cofactor regeneration enzyme gene was also performed. The conversion yield of KPA to d-pantoic acid reached over 88% with a substrate concentration up to 1.17 M.  相似文献   

10.
The alginate-assimilating bacterium, Sphingomonas sp. strain A1, degrades the polysaccharides to monosaccharides through four alginate lyase reactions. The resultant monosaccharide, which is nonenzymatically converted to 4-deoxy-l-erythro-5-hexoseulose uronate (DEH), is further metabolized to 2-keto-3-deoxy-d-gluconate by NADPH-dependent reductase A1-R in the short-chain dehydrogenase/reductase (SDR) family. A1-R-deficient cells produced another DEH reductase, designated A1-R′, with a preference for NADH. Here, we show the identification of a novel NADH-dependent DEH reductase A1-R′ in strain A1, structural determination of A1-R′ by x-ray crystallography, and structure-based conversion of a coenzyme requirement in SDR enzymes, A1-R and A1-R′. A1-R′ was purified from strain A1 cells and enzymatically characterized. Except for the coenzyme requirement, there was no significant difference in enzyme characteristics between A1-R and A1-R′. Crystal structures of A1-R′ and A1-R′·NAD+ complex were determined at 1.8 and 2.7 Å resolutions, respectively. Because of a 64% sequence identity, overall structures of A1-R′ and A1-R were similar, although a difference in the coenzyme-binding site (particularly the nucleoside ribose 2′ region) was observed. Distinct from A1-R, A1-R′ included a negatively charged, shallower binding site. These differences were caused by amino acid residues on the two loops around the site. The A1-R′ mutant with the two A1-R-typed loops maintained potent enzyme activity with specificity for NADPH rather than NADH, demonstrating that the two loops determine the coenzyme requirement, and loop exchange is a promising method for conversion of coenzyme requirement in the SDR family.  相似文献   

11.
Drosophila alcohol dehydrogenase (ADH) is an NAD(H)-dependent oxidoreductase that catalyzes the oxidation of alcohols and aldehydes. Structurally and biochemically distinct from all the reported ADHs (typically, the mammalian medium-chain dehydrogenase/reductase–ethanol-metabolizing enzyme), it stands as the only small-alcohol transforming system that has originated from a short-chain dehydrogenase/reductase (SDR) ancestor. The crystal structures of the apo, binary (E·NAD+) and three ternary (E·NAD+·acetone, E·NAD+·3-pentanone and E·NAD+·cyclohexanone) forms of Drosophila lebanonensis ADH have allowed us to infer the structural and kinetic features accounting for the generation of the ADH activity within the SDR lineage.  相似文献   

12.
The chloroplastic isoform of monodehydroascorbate (MDA) radical reductase was purified from spinach chloroplasts and leaves. The cDNA of chloroplastic MDA reductase was cloned, and its deduced amino acid sequence, consisting of 497 residues, showed high homology with those of putative organellar MDA reductases deduced from cDNAs of several plants. The amino acid sequence of the amino terminal of the purified enzyme suggested that the chloroplastic enzyme has a transit peptide consisting of 53 residues. A southern blot analysis suggested the occurrence of a gene encoding another isoform homologous to the chloroplastic isoform in spinach. The recombinant enzyme was highly expressed in Eschericia coli using the cDNA, and purified to a homogeneous state with high specific activity. The enzyme properties of the chloroplastic isoform are presented in comparison with those of the cytosolic form.  相似文献   

13.
Summary Escherichia coli rnh mutants deficient in ribonuclease H (RNase H) are capable of DNA replication in the absence of protein synthesis. This constitutive stable DNA replication (SDR) is dependent upon the recA + gene product. The requirement of SDR for recA + can be suppressed by rin mutations (for recA+-independent), or by lexA(Def) mutations which inactivate the LexA repressor. Thus, there are at least three genetically distinct types of SDR in rnh mutants: recA +-dependent SDR seen in rnh - rin+ lexA+ strains, recA +-independent in rnh - rin- lexA+, and recA +-independent in rnh - rin+ lexA(Def). The expression of SDR in rin - and lexA(Def) mutants demonstrated a requirement for RNA synthesis and for the absence of RNase H. The suppression of the recA + requirement by rin mutations was shown to depend on some new function of the recF + gene product. In contrast, the suppression by lexA-(Def) mutations was not dependent on recF +. The lexA3 mutation inhibited recA +-dependent SDR via reducing the amount of recA + activity available, and was suppressed by the recAo254 mutation. The SDR in rnh - rin- cells was also inhibited by the lexA3 mutation, but the inhibition was not reversed by the recAo254 mutation, indicating a requirement for some other lexA +-regulated gene product in the recA +-independent SDR process. A model is presented for the regulation of the expression of these three types of SDR by the products of the lexA +, rin+ and recF + genes.  相似文献   

14.
The gene encoding dimethyl sulfoxide (DMSO) reductase, which contains a molybdenum cofactor, of the phototrophic bacterium Rhodobacter sphaeroides f. sp. denitrificans was isolated using an oligonucleotide probe, which was synthesized based on a internal amino acid sequence of the purified enzyme. The DMSO reductase gene coded for 822 amino acids (2466 base pairs, Mr = 89,206) as a precursor form having a signal peptide of 42 amino acids. The deduced amino acid sequence had high homology with those of some enzymes containing a molybdenum cofactor: trim ethyl amine N-oxide reductase (48%), biotin sulfoxide reductase (44%), and DMSO reductase (29%) of Escherichia coli.  相似文献   

15.
A novel nicotinamide adenine dinucleotide phosphate-dependent carbonyl reductase, 3-quinuclidinone reductase, was isolated from Rhodotorula rubra JCM3782. The enzyme catalyzes the asymmetric reduction of 3-quinuclidinone to (R)-3-quinuclidinol. The gene encoding the enzyme was also cloned and sequenced. A 819-bp nucleotide fragment was confirmed to be the gene encoding the 3-quinuclidinone reductase by agreement of the internal amino acid sequences of the purified enzyme. The gene encodes a total of 272 amino acid residues, and the deduced amino acid sequence shows similarity to those of several short-chain dehydrogenase/reductase family proteins. An expression vector, pWKLQ, which contains the full length 3-quinuclidinone reductase gene was constructed. Using Escherichia coli cells coexpressing the 3-quinuclidinone reductase and glucose dehydrogenase (cofactor regeneration enzyme) genes, 618 mM 3-quinuclidinone was almost stiochiometrically converted to (R)-3-quinuclidinol with an >99.9% enantiomeric excess within 21 h of reaction.  相似文献   

16.
In the present investigation, four macrophytes, namely Typha latifolia (L.), Lemna minor (L.), Eichhornia crassipes (Mart.) Solms-Laubach, and Myriophyllum aquaticum (Vell.) Verdc, were evaluated for their heavy metal (Cu, Pb, Hg, and Zn) hyperaccumulation potential under laboratory conditions. Tolerance analyses were performed for 7 days of exposure at five different treatments of the metals mixture (Cu+2, Hg+2, Pb+2, and Zn+2). The production of chlorophyll and carotenoids was determined at the end of each treatment. L. minor revealed to be sensitive, because it did not survive in all the tested concentrations after 72 hours of exposure. E. crassipes and M. aquaticum displayed the highest tolerance to the metals mixture. For the most tolerant species of aquatic macrophytes, The removal kinetics of E. crassipes and M. aquaticum was carried out, using the following mixture of metals: Cu (0.5 mg/L) and Hg, Pb, and Zn 0.25 mg/L. The obtained results revealed that E. crassipes can remove 99.80% of Cu, 97.88% of Pb, 99.53% of Hg, and 94.37% of Zn. M. aquaticum withdraws 95.2% of Cu, 94.28% of Pb, 99.19% of Hg, and 91.91% of Zn. The obtained results suggest that these two species of macrophytes could be used for the phytoremediation of this mixture of heavy metals from the polluted water bodies.  相似文献   

17.
The gene encoding l-lactate dehydrogenase (LDH) was cloned from an industrial dairy strain of Streptococcus thermophilus M-192 using a synthetic oligonucleotide probe based on the N-terminal amino acid sequence of the purified enzyme, and its nucleotide sequence was determined. The enzyme was deduced to have 328 amino acid residues with a molecular weight of 35,428 and found to have high sequence similarity to LDHs from other lactic acid bacteria (89.0% to Streptococcus mutans, 76.3% to Lactococcus lactis subsp. lactis, 67% to Lactobacillus casei, and 60% to Lactobacillus plantarum). The gene contained a promoter-like sequence similar to the Escherichia coli promoter consensus, and expression of the S. thermophilus LDH gene was observed in E. coli cells.  相似文献   

18.
3-Methylaspartase was purified 24-fold and crystallized from the crude extract of the cells of a facultative anaerobic bacterium from soil, strain YG-1002. The molecular mass of the native enzyme was about 84 kDa and that of the subunit was about 42 kDa. The pH optimum for the deamination reaction of (2S, 3S)-3-methylaspartic acid and those for the amination reaction of mesaconic acid were 9.7 and 8.5; its optimum temperature was 50°C. The enzyme was stable at pH 5.5–11.0 and up to 50°C. The enzyme required both divalent and monovalent cations such as Mg2+ and K+. The enzyme was inhibited by sulfhydryl reagents, metal-chelating reagents and some divalent cations. The enzyme catalyzed the reversible amination/deamination reactions between several 3-substituted (S)-aspartic acids and their corresponding fumaric acid derivatives. The enzyme preferentially acted on (2S, 3S)-3-methylaspartic acid and mesaconic acid in the deamination and the amination reactions respectively. The enzyme showed high similarities in several enzymological properties and N-terminal amino acid sequence with 3-methylaspartase from an obligate anaerobic bacteriumClostridium tetanomorphum.  相似文献   

19.
The general aminopeptidase PepN from Streptococcus thermophilus A was purified to protein homogeneity by hydroxyapatite, anion-exchange, and gel filtration chromatographies. The PepN enzyme was estimated to be a monomer of 95 kDa, with maximal activity on N-Lys–7-amino-4-methylcoumarin at pH 7 and 37°C. It was strongly inhibited by metal chelating agents, suggesting that it is a metallopeptidase. The activity was greatly restored by the bivalent cations Co2+, Zn2+, and Mn2+. Except for proline, glycine, and acidic amino acid residues, PepN has a broad specificity on the N-terminal amino acid of small peptides, but no significant endopeptidase activity has been detected. The N-terminal and short internal amino acid sequences of purified PepN were determined. By using synthetic primers and a battery of PCR techniques, the pepN gene was amplified, subcloned, and further sequenced, revealing an open reading frame of 2,541 nucleotides encoding a protein of 847 amino acids with a molecular weight of 96,252. Amino acid sequence analysis of the pepN gene translation product shows high homology with other PepN enzymes from lactic acid bacteria and exhibits the signature sequence of the zinc metallopeptidase family. The pepN gene was cloned in a T7 promoter-based expression plasmid and the 452-fold overproduced PepN enzyme was purified to homogeneity from the periplasmic extract of the host Escherichia coli strain. The overproduced enzyme showed the same catalytic characteristics as the wild-type enzyme.  相似文献   

20.
《Insect Biochemistry》1985,15(2):299-307
NADPH-cytochrome c reductase was purified by affinity chromatographic techniques from microsomes prepared from the abdomens of insecticide-resistant (R) and -susceptible (S) house flies Musca domestica and of the black blow fly Phormia regina. Data are presented which describe (1) the ability of the purified enzymes to support an in vitro reconstitution of mono-oxygenase activity, (2) the changes in activity of these preparations observed in buffers of varying ionic strength, (3) comparative kinetic behaviour between microsomal and purified forms of the enzymes, (4) the immunochemical characteristics of these preparations, and (5) their amino acid composition. The reductases from the three sources were found to be very similar in all of these tests. Substrate binding constants were 5 μM for NADPH, 12 μM for cytochrome c, and the catalytic mechanism was interpreted as ordered Bi Bi. The inhibitory constant of the reductase from the resistant fly for 2′-AMP, an analogue of NADP+, was 187 μM; whereas no assciation of the inhibitor was observed below concentrations of 400 μM for the enzyme of either the susceptible house fly or the blow fly. However, the data are insufficient to suggest that the reductase is a significant factor in insecticide resistance. Compared to the same enzymes from rat and rabbit liver, the insect reductases show a different ionic strength optimum (0.14), have distinct antigenic determinants, and have different levels of acidic and basic amino acids in the membrane-binding peptide.  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