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.
(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   

2.
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   

3.
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   

4.
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   

5.
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   

6.
R plasmid dihydrofolate reductase with a dimeric subunit structure   总被引:5,自引:0,他引:5  
Dihydrofolate reductase specified by plasmid R483 from a trimethoprim-resistant strain of Escherichia coli has been purified 2,000-fold to homogeneity using dye-ligand chromatography, gel filtration, and polyacrylamide gel electrophoresis. The protein migrated as a single band on nondenaturing polyacrylamide gel electrophoresis and had a specific activity of 250 mumol/mg min(-1). The molecular weight was estimated to be 32,000 by gel filtration and 39,000 by Ferguson analysis of polyacrylamide gel electrophoresis. When subjected to electrophoresis in the presence of sodium dodecyl sulfate, the protein migrated as a single 19,000-molecular weight species, a fact that suggests that the native enzyme is a dimer of similar or identical subunits. Antibody specific for R483-encoded dihydrofolate reductase did not cross-react with dihydrofolate reductase encoded by plasmid R67, T4 phage, E. coli RT500, or mouse L1210 leukemia cells. The amino acid sequence of the first 34 NH2-terminal residues suggests that the R483 plasmid dihydrofolate reductase is more closely related to the chromosomal dihydrofolate reductase than is the enzyme coded by plasmid R67.  相似文献   

7.
Activation of rat liver phenylalanine hydroxylase by limited proteolysis catalyzed by chymotrypsin was investigated with the use of sodium dodecyl sulfate-polyacrylamide gel electrophoresis and high pressure gel filtration. Both activation and proteolysis were decreased by the addition of the natural cofactor, (6R)-tetrahydrobiopterin. From chymotryptic digests of the hydroxylase carried out in the presence and absence of (6R)-tetrahydrobiopterin, several different enzyme species were isolated by high pressure gel filtration. One species (subunit Mr = 47,000) with unchanged hydroxylase activity was isolated from the chymotryptic digest in the presence of (6R)-tetrahydrobiopterin; it was derived from the native enzyme (Mr = 52,000) by cleavage of the COOH-terminal Mr = 5,000 portion of the native enzyme. In the absence of (6R)-tetrahydrobiopterin, another species (subunit Mr = 36,000) was isolated. In addition to modification at the COOH-terminal end of the molecule, this species also had lost a Mr = 11,000 fragment from the NH2-terminal end of the hydroxylase. The Mr = 11,000 fragment was shown to include the phosphorylation site of the enzyme. This Mr = 36,000 species was 30-fold more active than the native phenylalanine hydroxylase when assayed in the presence of tetrahydrobiopterin. These results suggest that the regulatory domain that inhibits hydroxylase activity in the basal state may be located at the NH2 terminus of the phenylalanine hydroxylase subunit.  相似文献   

8.
The NAD+ kinase (EC 2.7.1.23) of the filamentous fungus N. crassa is localized in cytosol. The activity in the dialyzed cell free extract has a pH optimum 8.3; it utilizes only ATP but not inorganic polyphosphates as a phosphoryl donor. A method for 200-fold purification of NAD+ kinase with a 20% yield has been developed. The procedure includes 105000 g centrifugation, fractionation with (NH4)2SO4, isoelectrofocusing in a Ultrodex layer and preparative electrophoresis in polyacrylamide gel. The molecular heterogeneity of NAD+ kinase was demonstrated by polyacrylamide gradient electrophoresis and by gel filtration through Sephadex G-200. The molecular weights of four individual forms of the enzyme are: 330000-338000, 305000-306000, 215000-229000 and 203000 Da. The Km values for the reaction catalyzed by purified NAD+ kinase for NAD+ and ATP are 3.0 X 10(-4) M and 0.9 X 10(-3) M, respectively.  相似文献   

9.
An NAD(+)-dependent alcohol dehydrogenase was purified to homogeneity from Nocardia fusca AKU 2123. The enzyme catalyzed (S)-specific oxidation of 3-pentyn-2-ol (PYOH), i.e., part of the stereoinversion reaction for the production of (R)-PYOH, which is a valuable chiral building block for pharmaceuticals, from the racemate. The enzyme used a broad variety of secondary alcohols including alkyl alcohols, alkenyl alcohols, acetylenic alcohols, and aromatic alcohols as substrates. The oxidation was (S)-isomer specific in every case. The K(m) and Vmax for (S)-PYOH and (S)-2-hexanol oxidation were 1.6 mM and 53 mumol/min/mg, and 0.33 mM and 130 mumol/min/mg, respectively. The enzyme also catalyzed stereoselective reduction of carbonyl compounds. (S)-2-Hexanol and ethyl (R)-4-chloro-3-hydroxybutanoate in high optical purity were produced from 2-hexanone and ethyl 4-chloro-3-oxobutanoate by the purified enzyme, respectively. The K(m) and Vmax for 2-hexanone reduction were 2.5 mM and 260 mumol/min/mg. The enzyme has a relative molecular mass of 150,000 and consists of four identical subunits. The NH2-terminal amino acid sequence of the enzyme shows similarity with those of the carbonyl reductase from Rhodococcus erythropolis and phenylacetaldehyde reductase from Corynebacterium sp.  相似文献   

10.
Fungal metabolism of biphenyl.   总被引:9,自引:0,他引:9       下载免费PDF全文
gamma-Glutamyl phosphate reductase, the second enzyme of proline biosynthesis, catalyses the formation of l-glutamic acid 5-semialdehyde from gamma-glutamyl phosphate with NAD(P)H as cofactor. It was purified 150-fold from crude extracts of Pseudomonas aeruginosa PAO 1 by DEAE-cellulose chromatography and hydroxyapatite adsorption chromatography. The partially purified preparation, when assayed in the reverse of the biosynthetic direction, utilized l-1-pyrroline-5-carboxylic acid as substrate and reduced NAD(P)(+). The apparent K(m) values were: NAD(+), 0.36mm; NADP(+), 0.31mm; l-1-pyrroline-5-carboxylic acid, 4mm with NADP(+) and 8mm with NAD(+); P(i), 28mm. 3-(Phosphonoacetylamido)-l-alanine, a structural analogue of gamma-glutamyl phosphate, inhibited this enzyme competitively (K(i)=7mm). 1-Pyrroline-5-carboxylate reductase (EC 1.5.1.2), the third enzyme of proline biosynthesis, was purified 56-fold by (NH(4))(2)SO(4) fractionation, Sephadex G-150 gel filtration and DEAE-cellulose chromatography. It reduced l-1-pyrroline-5-carboxylate with NAD(P)H as a cofactor to l-proline. NADH (K(m)=0.05mm) was a better substrate than NADPH (K(m)=0.02mm). The apparent K(m) values for l-1-pyrroline-5-carboxylate were 0.12mm with NADPH and 0.09mm with NADH. The 3-acetylpyridine analogue of NAD(+) at 2mm caused 95% inhibition of the enzyme, which was also inhibited by thio-NAD(P)(+), heavy-metal ions and thiol-blocking reagents. In cells of strain PAO 1 grown on a proline-medium the activity of gamma-glutamyl kinase and gamma-glutamyl phosphate reductase was about 40% lower than in cells grown on a glutamate medium. No repressive effect of proline on 1-pyrroline-5-carboxylate reductase was observed.  相似文献   

11.
The influence of growth conditions on assimilatory and respiratory nitrate reduction in Aerobacter aerogenes was studied. The level of nitrate reductase activity in cells, growing in minimal medium with nitrate as the sole nitrogen source, was much lower under aerobic than anaerobic conditions. Further, the enzyme of the aerobic cultures was very sensitive to sonic disintegration, as distinct from the enzyme of anaerobic cultures. When a culture of A. aerogenes was shifted from anaerobic growth in minimal medium with nitrate and NH(4) (+) to aerobiosis in the same medium, but without NH(4) (+), the production of nitrite stopped instantaneously and the total activity of nitrate reductase decreased sharply. Moreover, there was a lag in growth of about 3 hr after such a shift. After resumption of growth, the total enzymatic activity increased again slowly and simultaneously became gradually sensitive to sonic disintegration. These findings show that oxygen inactivates the anaerobic nitrate reductase and represses its further formation; only after a de novo synthesis of nitrate reductase with an assimilatory function will growth be resumed. The enzyme in aerobic cultures was not significantly inactivated by air, only by pure oxygen. The formation of the assimilatory enzyme complex was repressed, however, by NH(4) (+), under both aerobic and anaerobic conditions. The results indicate that the formation of the assimilatory enzyme complex and that of the respiratory enzyme complex are regulated differently. We suggest that both complexes have a different composition, but that the nitrate reductase in both cases is the same protein.  相似文献   

12.
In situ product crystallization was investigated for solid product crystals that were obtained during fermentation. The model reaction was the asymmetric reduction of 4-oxoisophorone (OIP) using baker's yeast (S. cerevisiae) as a biocatalyst. The target product was 6R-dihydro-oxoisophorone (DOIP), also known as levodione, a key intermediate in carotenoid synthesis. DOIP was degraded by baker's yeast mainly to (4S,6R)-actinol, an unwanted byproduct in the process. Actinol formation reached up to 12.5% of the initial amount of OIP in the reactor during a batch process. However, better results were obtained when the dissolved DOIP concentration was controlled using an integrated fermentation-crystallization process because: (a) actinol formation was reduced to 4%; and (b) DOIP crystal formation in the reactor was avoided. DOIP productivity was improved by 50% and its selectivity was raised from 87% to 96% relative to the batch process. In the integrated process, most of the product was recovered as pure crystals; this may already minimize, if not eliminate, the need for organic solvents in the final purification steps. An almost sixfold reduction in biocatalyst consumption per kilogram product was achieved, which also can contribute to the minimization of waste streams.  相似文献   

13.
2,5-Diamino-4-oxy-6-ribosylaminopyrimidine-5'-phosphate reductase has been isolated from cells of Pichia guilliermondii and subjected to 20-fold purification by treating extracts with streptomycin sulphate, frationating proteins (NH4)2SO4 at 45-75% of saturation and chromatography on blue sepharose CL-6B. The use of gel filtration through Sephadex G-150 and chromatography on DEAE-cellulose proved to be less effective for the enzyme purification. It has been established that it is 2,5-diamino-4-oxy-6-ribosylaminopyrimidine-5-phosphate but not its dephosphorylated form that is the substrate of the given reductase; Km is equal to 7.10(-5) M. The reaction proceeds in the presence of NADPH or NADH. The enzyme affinity to NADPH (Km = 4.7.10(-5) M) is approximately one order higher than that to NADPH (Km = 5.5.10(-4) M). The enzyme manifests the optimum of action at pH 7.2 and the temperature of 37 degrees C; the molecular weight is 140 kD. EDTA as well as flavins in the concentration of 1.10(-3) M exert no effect on the reductase activity. The enzyme is labile at 4 degrees C and is inactivated in the frozen state at -15 degrees C. The 2.5-diamino-4-oxy-6-ribosylaminopyrimidine-5'-phosphate reductase has been also revealed in Torulopsis candida, Debaryomyces kl?ckeri, Schwanniomyces occidentalis, Eremothecium ashbyii (flavinogenic species) and Candida utilis. Aspergillus nidulans, Neurospora crassa (nonflavinogenic species). The synthesis of this enzyme contrary to other enzymes of the riboflavin biosynthesis is not regulated in flavinogenic yeast by iron ions.  相似文献   

14.
It was found that the cytoplasm of light-grown cells of Rhodospirillum rubrum could catalyze the reduction of methyl viologen (MV) (Em, 7 = -0.44 V) by NADH and NADPH. In the present study, the enzyme capable of catalyzing MV reduction by NADH (NADH-MV reductase) was purified 1,500-fold from an extract of cells with a yield of 4.4%. The purification procedure comprised (NH4)2SO4 fractionation, and chromatographies on Sepharose CL-6B, DEAE-Sepharose CL-6B, phenyl-Sepharose CL-4B, Blue-Cellulofine, and TSK-Gel G3000SW. Two NADPH-MV reductases were separated during the purification. The NADH-MV reductase obtained was nearly homogeneous, as judged on polyacrylamide gel electrophoresis both in the presence and absence of sodium dodecyl sulfate. The enzyme has a molecular weight of 220,000 and an isoelectric point of 4.8; it is composed of four subunits with a molecular weight of 57,000, and is bound with about 1 mol FAD/mol subunit. The activity is optimum at pH 8. The Km values for NADH and MV are 115 microM and 1.3 mM, respectively, with a molecular activity of 13,000 min-1. The activity was stimulated 2.4-fold in the presence of 20-100 mM ammonium ions. The enzyme also catalyzed the reduction of benzyl viologen, methylene blue and 2,6-dichlorophenol-indophenol (Em, 7 = -0.36, +0.011, and +0.217 V, respectively) at comparable rates. The ratios of the activity with NADH to that with NADPH were 80, 133, 41, and 5.5 with MV, benzyl viologen, methylene blue and 2,6-dichlorophenolindophenol, respectively. The enzyme was significantly stable in the presence of both 5mM 2-mercaptoethanol and 20% (w/v) glycerol. The activity was not appreciably influenced by the presence of 2 M urea, although the reagent caused dissociation to the subunits.  相似文献   

15.
We used the baculovirus/Sf9 expression system to gain new information on the mechanistic properties of the rat non-gastric H,K-ATPase, an enzyme that is implicated in potassium homeostasis. The alpha2-subunit of this enzyme (HKalpha2) required a beta-subunit for ATPase activity thereby showing a clear preference for NaKbeta1 over NaKbeta3 and gastric HKbeta. NH4(+), K+, and Na+ maximally increased the activity of HKalpha2-NaKbeta1 to 24.0, 14.2, and 5.0 micromol P(i) x mg(-1) protein x h(-1), respectively. The enzyme was inhibited by relatively high concentrations of ouabain and SCH 28080, whereas it was potently inhibited by oligomycin. From the phosphorylation level in the presence of oligomycin and the maximal NH4(+)-stimulated ATPase activity, a turnover number of 20,000 min(-1) was determined. All three cations decreased the steady-state phosphorylation level and enhanced the dephosphorylation rate, disfavoring the hypothesis that Na+ can replace H+ as the activating cation. The potency with which vanadate inhibited the cation-activated enzyme decreased in the order K+ > NH4(+) > Na+, indicating that K+ is a stronger E2 promoter than NH4(+), whereas in the presence of Na+ the enzyme is in the E1 form. For K+ and NH4(+), the E2 to E1 conformational equilibrium correlated with their efficacy in the ATPase reaction, indicating that here the transition from E2 to E1 is rate-limiting. Conversely, the low maximal ATPase activity with Na+ is explained by a poor stimulatory effect on the dephosphorylation rate. These data show that NH4(+) can replace K+ with similar affinity but higher efficacy as an extracellular activating cation in rat nongastric H,K-ATPase.  相似文献   

16.
Individual flowers from inflorescences of Dahlia variabilis (cv Scarlet Star) in young developmental stages contained relatively high activity of (+)-dihydroflavonol (DHF) 4-reductase. The DHF reductase was purified from such flowers to apparent homogeneity by a five-step procedure. This included affinity adsorption on Blue Sepharose and elution of the enzyme with NADP+. By gel filtration and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis it was shown that DHF reductase contains only one polypeptide chain with a Mr of about 41,000. The reductase required NADPH as cofactor and catalyzed transfer of the pro-S hydrogen of NADPH to the substrate. Flavanones and dihydroflavonols (3-hydroxyflavanones) were substrates for DHF reductase with pH optima of about 6.0 for flavanones and of about 6.8 for dihydroflavonols. Flavanones were reduced to the corresponding flavan-4-ols and (+)-dihydroflavonols to flavan-3,4-cis-diols. Apparent Michaelis constants determined for (2S)-naringenin, (2S)-eriodicytol, (+)-dihydrokaempferol, (+)-dihydroquercetin, and NADPH were, respectively, 2.3, 2, 10, 15, and 42 microM. V/Km values were higher for dihydroflavonols than for flavanones. Conversion of dihydromyricetin to leucodelphinidin was also catalyzed by the enzyme at a low rate, whereas flavones and flavonols were not accepted as substrates. DHF reductase was not inhibited by metal chelators.  相似文献   

17.
Microorganisms were screened for ones that reduced 3,5,5-trimethyl-2-cyclohexene-1,4-dione (ketoisophorone; KIP), and several strains were found to produce (6R)-2,2,6-trimethylcyclohexane-1,4-dione (levodione). The enzyme catalyzing the reduction of the C=C bond of KIP to yield (6R)-levodione was isolated from Candida macedoniensis AKU4588. The results of primary structural analysis and its enzymatic properties suggested that the enzyme might be an Old Yellow Enzyme family protein.  相似文献   

18.
The effect of inhibiting lysosomal protein degradation on the activity of 3-hydroxy-3-methylglutaryl coenzyme A (HMG-CoA) reductase was determined using a mouse mammary cell line (TS-85) which expresses a temperature-sensitive mutation in the ubiquitin degradative pathway. Incubating cells for 18 hr in medium containing 20 mM NH4Cl did not alter total protein synthesis or cell growth, but it did inhibit the rate of total protein degradation by 19%, which is consistent with the known inhibitory effect of NH4Cl on lysosomal protein degradation. NH4Cl treatment also resulted in an increase (81% +/- 20) in HMG-CoA reductase activity. The increase in reductase activity was not correlated with changes in the phosphorylation state of the enzyme or with alteration in the relative rate of reductase synthesis. However, the basal degradation rate of the reductase was significantly inhibited, and after NH4Cl treatment, the half-life of the enzyme increased from 4.0 +/- 0.4 hr to 8.3 +/- 0.8 hr. The change in the rate of reductase degradation can account completely for the increase in reductase activity observed in NH4Cl-treated cells. The accelerated degradation of HMG-CoA reductase induced by 25-hydroxycholesterol treatment was not affected by either NH4Cl or by inactivation of the ubiquitin degradative pathway. Therefore, two different mechanisms may be responsible for the accelerated degradation and basal degradation of HMG-CoA reductase. The latter can be inhibited by NH4Cl and may imply that under basal conditions the enzyme may be degraded in lysosomes.  相似文献   

19.
The mechanism for "NH4+ switch-off/on" of nitrogenase activity in Azospirillum brasilense and A. lipoferum was investigated. A correlation was established between the in vivo regulation of nitrogenase activity by NH4Cl or glutamine and the reversible covalent modification of dinitrogenase reductase. Dinitrogenase reductase ADP-ribosyltransferase (DRAT) activity was detected in extracts of A. brasilense with NAD as the donor molecule. Dinitrogenase reductase-activating glycohydrolase (DRAG) activity was present in extracts of both A. brasilense and A. lipoferum. The DRAG activity in A. lipoferum was membrane associated, and it catalyzed the activation of inactive nitrogenase (by covalent modification of dinitrogenase reductase) from both A. lipoferum and Rhodospirillum rubrum. A region homologous to R. rubrum draT and draG was identified in the genomic DNA of A. brasilense as a 12-kilobase EcoRI fragment and in A. lipoferum as a 7-kilobase EcoRI fragment. It is concluded that a posttranslational regulatory system for nitrogenase activity is present in A. brasilense and A. lipoferum and that it operates via ADP-ribosylation of dinitrogenase reductase as it does in R. rubrum.  相似文献   

20.
The enzyme 6-oxocamphor hydrolase, which catalyzes the desymmetrization of 6-oxocamphor to yield (2R,4S)-alpha-campholinic acid, has been purified with a factor of 35.7 from a wild type strain of Rhodococcus sp. NCIMB 9784 grown on (1R)-(+)-camphor as the sole carbon source. The enzyme has a subunit molecular mass of 28,488 Da by electrospray mass spectrometry and a native molecular mass of approximately 83,000 Da indicating that the active protein is trimeric. The specific activity was determined to be 357.5 units mg(-)1, and the K(m) was determined to be 0.05 mm for the natural substrate. The N-terminal amino acid sequence was obtained from the purified protein, and using this information, the gene encoding the enzyme was cloned. The translation of the gene was found to bear significant homology to the crotonase superfamily of enzymes. The gene is closely associated with an open reading frame encoding a ferredoxin reductase that may be involved in the initial step in the biodegradation of camphor. A mechanism for 6-oxocamphor hydrolase based on sequence homology and the known mechanism of the crotonase enzymes is proposed.  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