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 46 毫秒
1.
Styrene oxide and 2-phenylethanol metabolism in the styrene-degrading Xanthobacter sp. strain 124X was shown to proceed via phenylacetaldehyde and phenylacetic acid. In cell extracts 2-phenylethanol was oxidized by a phenazine methosulfate-dependent enzyme, probably a pyrroloquinoline quinone enzyme. Xanthobacter sp. strain 124X also contains a novel enzymatic activity designated as styrene oxide isomerase. Styrene oxide isomerase catalyzes the isomerization of styrene oxide to phenylacetaldehyde. The enzyme was partially purified and shown to have a very high substrate specificity. Of the epoxides tested, styrene oxide was the only substrate transformed. The initial step in styrene metabolism in Xanthobacter sp. strain 124X is oxygen dependent and probably involves oxidation of the aromatic nucleus.  相似文献   

2.
We investigated the biosynthetic pathway for 2-phenylethanol, the dominant floral scent compound in roses, using enzyme assays. L-[(2)H8] Phenylalanine was converted to [(2)H8] phenylacetaldehyde and [(2)H8]-2-phenylethanol by two enzymes derived from the flower petals of R. 'Hoh-Jun,' these being identified as pyridoxal-5'-phosphate-dependent L-aromatic amino acid decarboxylase (AADC) and phenylacetaldehyde reductase (PAR). The activity of rose petal AADC to yield phenylacetaldehyde was nine times higher toward L-phenylalanine than toward its D-isomer, and this conversion was not inhibited by iproniazid, a specific inhibitor of monoamine oxidase. Under aerobic conditions, rose petal AADC stoichiometrically produced NH3 together with phenylacetaldehyde during the course of decarboxylation and oxidation, followed by the hydrolysis of L-phenylalanine. Phenylacetaldehyde was subsequently converted to 2-phenylethanol by the action of PAR. PAR showed specificity toward several volatile aldehydes.  相似文献   

3.
The nucleotide sequence of the 4,377-bp chromosomal region of Pseudomonas fluorescens ST that codes for the oxidation of styrene to phenylacetic acid was determined. Four open reading frames, named styA, styB, styC, and styD, were identified in this region. Sequence analysis and biotransformation assays, performed with batch and continuous cultures, allowed us to identify the functions of the sequenced genes. styA and styB encode a styrene monooxygenase responsible for the transformation of styrene to epoxystyrene; styC codes for the second enzyme of the pathway, an epoxystyrene isomerase that converts epoxystyrene to phenylacetaldehyde; and the styD gene produces a phenylacetaldehyde dehydrogenase that oxidizes phenylacetaldehyde to phenylacetic acid. StyA, 415-amino-acids long, was found to be weakly homologous to p-hydroxybenzoate hydroxylase from both P. fluorescens and P. aeruginosa and to salicylate hydroxylase from P. putida, suggesting that it might be a flavin adenine dinucleotide-binding monooxygenase. StyB was found to be partially homologous to the carboxyterminal part of the 2,4-dichlorophenol-6-monooxygenase encoded by plasmid pJP4, while the styC product did not share significant homology with any known proteins. The fourth open reading frame, styD, could encode a protein of 502 amino acids and was strongly homologous to several eukaryotic and prokaryotic aldehyde dehydrogenases. The order of the genes corresponds to that of the catabolic steps. The previously suggested presence of the gene for epoxystyrene reductase, which directly converts epoxystyrene to 2-phenylethanol (A.M. Marconi, F. Beltrametti, G. Bestetti, F. Solinas, M. Ruzzi, E. Galli, and E. Zennaro, Appl. Environ. Microbiol. 61:121-127, 1996), has not been confirmed by sequencing and by biotransformation assays performed in continuous cultures. A copy of the insertion sequence ISI162, belonging to the IS21-like family of elements, was identified immediately downstream of the styrene catabolic genes.  相似文献   

4.
We investigated the biosynthetic pathway for 2-phenylethanol, the dominant floral scent compound in roses, using enzyme assays. L-[2H8] Phenylalanine was converted to [2H8] phenylacetaldehyde and [2H8]-2-phenylethanol by two enzymes derived from the flower petals of R. ‘Hoh-Jun,’ these being identified as pyridoxal-5′-phosphate-dependent L-aromatic amino acid decarboxylase (AADC) and phenylacetaldehyde reductase (PAR). The activity of rose petal AADC to yield phenylacetaldehyde was nine times higher toward L-phenylalanine than toward its D-isomer, and this conversion was not inhibited by iproniazid, a specific inhibitor of monoamine oxidase. Under aerobic conditions, rose petal AADC stoichiometrically produced NH3 together with phenylacetaldehyde during the course of decarboxylation and oxidation, followed by the hydrolysis of L-phenylalanine. Phenylacetaldehyde was subsequently converted to 2-phenylethanol by the action of PAR. PAR showed specificity toward several volatile aldehydes.  相似文献   

5.
The initial enzyme of ethylbenzene metabolism in denitrifying Azoarcus strain EbN1, ethylbenzene dehydrogenase, was purified and characterized. The soluble periplasmic enzyme is the first known enzyme oxidizing a nonactivated hydrocarbon without molecular oxygen as cosubstrate. It is a novel molybdenum/iron-sulfur/heme protein of 155 kDa, which consists of three subunits (96, 43, and 23 kDa) in an alphabetagamma structure. The N-terminal amino acid sequence of the alpha subunit is similar to that of other molybdenum proteins such as selenate reductase from the related species Thauera selenatis. Ethylbenzene dehydrogenase is unique in that it oxidizes the hydrocarbon ethylbenzene, a compound without functional groups, to (S)-1-phenylethanol. Formation of the product was evident by coupling to an enantiomer-specific (S)-1-phenylethanol dehydrogenase from the same organism. The apparent K(m) of the enzyme for ethylbenzene is very low at <2 microm. Oxygen does not affect ethylbenzene dehydrogenase activity in extracts but inactivates the purified enzyme, if the heme b cofactor is in the reduced state. A variant of ethylbenzene dehydrogenase exhibiting significant activity also with the homolog n-propylbenzene was detected in a related Azoarcus strain (PbN1).  相似文献   

6.
A gene bank from Pseudomonas fluorescens ST was constructed in the broad-host-range cosmid pLAFR3 and mobilized into Pseudomonas putida PaW340. Identification of recombinant cosmids containing the styrene catabolism genes was performed by screening transconjugants for growth on styrene and epoxystyrene. Transposon mutagenesis and subcloning of one of the selected genome fragments have led to the identification of three enzymatic activities: a monooxygenase activity encoded by a 3-kb PstI-EcoRI fragment and an epoxystyrene isomerase activity and an epoxystyrene reductase activity encoded by a 2.3-kb BamHI fragment. Escherichia coli clones containing the 3-kb PstI-EcoRI fragment were able to transform styrene into epoxystyrene, and those containing the 2.3-kb BamHI fragment converted epoxystyrene into phenylacetaldehyde or, only in the presence of glucose, into 2-phenylethanol. The three genes appear to be clustered and are probably encoded by the same DNA strand. In E. coli, expression of the epoxystyrene reductase gene was under the control of its own promoter, whereas the expression of the other two genes was dependent on the presence of an external vector promoter.  相似文献   

7.
8.
The molecular weights of Ehrlich tumor cell ribonucleotide reductase and its individual components were determined by sedimentation equilibrium in the Beckman Airfuge. The distribution of enzyme after sedimentation equilibrium was determined by measurement of the CDP reductase and ADP reductase activities associated with ribonucleotide reductase. The apparent molecular weight of the intact enzyme was 304,000 when assayed for CDP reductase and 254,000 when assayed for ADP reductase. This difference in apparent molecular weights was statistically significant with a P value of 0.0002. The molecular weights of the individual components of ribonucleotide reductase were determined in a similar fashion by assaying in the presence of an excess of the complementary component. The non-heme iron component had a molecular weight of 81,000 when assayed for either CDP or ADP reductase activity. The effector-binding component had an apparent molecular weight of 127,000 when assayed for CDP reductase and 95,000 when assayed for ADP reductase. This difference in apparent molecular weights was statistically significant with a P value of 0.004. The effectors ATP and dGTP altered the apparent molecular weights of the intact enzyme and individual components. In the presence of ATP the molecular weight of intact CDP reductase was 481,000 while the apparent molecular weight of the effector-binding component of CDP reductase alone was 418,000. In the presence of dGTP, the molecular weight of intact ADP reductase was 293,000 while the apparent molecular weight of the effector-binding component of ADP reductase alone was 154,000. These results indicate that the proportion of the non-heme iron component and the effector-binding component is not equimolar and that the composition of the enzyme is not constant but is altered by the presence of effectors. Our data also suggest that CDP reduction and ADP reduction are catalyzed by different molecular species of the enzyme which apparently have different effector-binding components.  相似文献   

9.
Optically active styrene oxide derivatives are versatile chiral building blocks. Stereoselective reduction of phenacyl halide to chiral 2-halo-1-phenylethanol is the key reaction of the most economical synthetic route. Rhodotorula glutinis var. dairenensis IFO415 was discovered on screening as a potent microorganism reducing a phenacyl halide to the (R)-form of the corresponding alcohol. An NADPH-dependent carbonyl reductase was purified to homogeneity through four steps from this strain. The relative molecular mass of the enzyme was estimated to be 40,000 on gel filtration and 30,000 on SDS-polyacrylamide gel electrophoresis. This enzyme reduced a broad range of carbonyl compounds in addition to phenacyl halides. Some properties of the enzyme and preparation of a chiral styrene oxide using the crude enzyme are reported herein.  相似文献   

10.
Cobaltous chloride induced in rat liver an enzyme which converted biliverdin reductase molecular form 1 into the molecular form 3. This conversion involves the oxidation of two sulfhydryl groups of form 1 giving rise to a disulfide bond in form 3. The converting enzyme was isolated from the liver peroxisomal fraction (which was devoid of biliverdin reductase activity), and was absent in liver peroxisomes of control rats. The enzyme was solubilized by treatment of the peroxisomes with 0.1% sodium deoxycholate, and partially purified by DEAE-cellulose and Sephadex G-100 filtration. It is a NAD+ dependent enzyme which was inactivated by trypsing and heat treatments. It did not oxidize either reduced glutathione or cysteine. The converting enzyme had a molecular weight of about 54,000 daltons. The oxidation of biliverdin reductase molecular form 1 mediated by the converting enzyme did not affect the latter's molecular weight or activity.  相似文献   

11.
A new type of dihydropteridine reductase [EC 1.6.99.10], which is specific for NADPH as the substrate in the reduction of quinonoid-dihydropterin to tetrahydropterin, was purified to homogeneity from bovine liver and human liver. The molecular weight of the enzyme was determined to be 65,000-70,000. The enzyme was composed of two subunits with identical molecular weight of 35,000; the amino terminal residue was determined to be valine. The isoelectric point of the enzyme was 7.05. The physicochemical properties of this enzyme were quite different from those of bovine liver NADH-specific dihydropteridine reductase [EC 1.6.99.7]. NADPH-specific dihydropteridine reductase did not cross-react with an antiserum raised against the NADH-specific dihydropteridine reductase, nor did the latter enzyme react with an antiserum to the former enzyme, indicating that the two enzymes have no common antigenic determinants. NADPH-specific dihydropteridine reductase from human liver was shown to have properties similar to those of the bovine liver enzyme.  相似文献   

12.
Glutathione reductase was purified from iron-grown Thiobacillus ferrooxidas AP19-3 to an electrophoretically homogeneous state. The enzyme had an apparent molecular weight of 100,000 and was composed of two identical subunits of molecular weight (Mrs, 52,000) as estimated by sodium dodecyl sulfate–polyacrylamide gel electrophoresis. A purified enzyme reduced one mole of the oxidized form of glutathione (GSSG) with one mole of NADPH to produce two moles of the reduced form of glutathione (GSH) and one mole of NADP+. The glutathione reductase was most active at pH 6.5 and 40°C, and had an isoelectric point at 5.1. The Michaelis constants of glutathione reductase for GSSG, NADPH, and NADH were 300, 26, and 125 μM, respectively.  相似文献   

13.
Styrene oxide isomerase (SOI) is involved in peripheral styrene catabolism of bacteria and converts styrene oxide to phenylacetaldehyde. Here, we report on the identification, enrichment, and biochemical characterization of a novel representative from the actinobacterium Rhodococcus opacus 1CP. The enzyme, which is strongly induced during growth on styrene, was shown to be membrane integrated, and a convenient procedure was developed to highly enrich the protein in active form from the wild-type host. A specific activity of about 370 U mg(-1) represents the highest activity reported for this enzyme class so far. This, in combination with a wide pH and temperature tolerance, the independence from cofactors, and the ability to convert a spectrum of substituted styrene oxides, makes a biocatalytic application imaginable. First, semipreparative conversions were performed from which up to 760 μmol of the pure phenylacetaldehyde could be obtained from 130 U of enriched SOI. Product concentrations of up to 76 mM were achieved. However, due to the high chemical reactivity of the aldehyde function, SOI was shown to be the subject of an irreversible product inhibition. A half-life of 15 min was determined at a phenylacetaldehyde concentration of about 55 mM, indicating substantial limitations of applicability and the need to modify the process.  相似文献   

14.
To find microorganisms that could reduce phenyl trifluoromethyl ketone (PTK) to (S)-1-phenyltrifluoroethanol [(S)-PTE], styrene-assimilating bacteria (ca. 900 strains) isolated from soil samples were screened. We found that Leifsonia sp. strain S749 was the most suitable strain for the conversion of PTK to (S)-PTE in the presence of 2-propanol as a hydrogen donor. The enzyme corresponding to the reaction was purified homogeneity, characterized and designated Leifsonia alcohol dehydrogenase (LSADH). The purified enzyme had a molecular weight of 110,000 and was composed of four identical subunits (molecular weight, 26,000). LSADH required NADH as a cofactor, showed little activity with NADPH, and reduced a wide variety of aldehydes and ketones. LSADH catalyzed the enantioselective reduction of some ketones with high enantiomeric excesses (e.e.): PTK to (S)-PTE (>99% e.e.), acetophenone to (R)-1-phenylethanol (99% e.e.), and 2-heptanone to (R)-2-heptanol (>99% e.e.) in the presence of 2-propanol without an additional NADH regeneration system. Therefore, it would be a useful biocatalyst.  相似文献   

15.
A tomato short-chain dehydrogenase-reductase (SlscADH1) is preferentially expressed in fruit with a maximum expression at the breaker stage while expression in roots, stems, leaves and flowers is very weak. It represents a potential candidate for the formation of aroma volatiles by interconverting alcohols and aldehydes. The SlscADH1 recombinant protein produced in Escherichia coli exhibited dehydrogenase-reductase activity towards several volatile compounds present in tomato flavour with a strong preference for the NAD/NADH co-factors. The strongest activity was observed for the reduction of hexanal (K(m)=0.175mM) and phenylacetaldehyde (K(m)=0.375mM) in the presence of NADH. The oxidation process of hexanol and 1-phenylethanol was much less efficient (K(m)s of 2.9 and 23.0mM, respectively), indicating that the enzyme preferentially acts as a reductase. However activity was observed only for hexanal, phenylacetaldehyde, (E)-2-hexenal and acetaldehyde and the corresponding alcohols. No activity could be detected for other aroma volatiles important for tomato flavour, such as methyl-butanol/methyl-butanal, 5-methyl-6-hepten-2-one/5-methyl-6-hepten-2-ol, citronellal/citronellol, neral/nerol, geraniol. In order to assess the function of the SlscADH1 gene, transgenic plants have been generated using the technique of RNA interference (RNAi). Constitutive down-regulation using the 35S promoter resulted in the generation of dwarf plants, indicating that the SlscADH1 gene, although weakly expressed in vegetative tissues, had a function in regulating plant development. Fruit-specific down-regulation using the 2A11 promoter had no morphogenetic effect and did not alter the aldehyde/alcohol balance of the volatiles compounds produced by the fruit. Nevertheless, SlscADH1-inhibited fruit unexpectedly accumulated higher concentrations of C5 and C6 volatile compounds of the lipoxygenase pathway, possibly as an indirect effect of the suppression of SlscADH1 on the catabolism of phospholipids and/or integrity of membranes.  相似文献   

16.
A novel (R)-1-phenylethanol dehydrogenase was successfully purified from Lysinibacillus sp. NUST506 by preparative polyacrylamide gel electrophoresis. The enzyme is a NAD+-dependent oxidoreductase. The molecular weight of the (R)-1-phenylethanol dehydrogenase measured by SDS-PAGE was about 28 kDa. Furthermore, the optimal reaction conditions for the oxidative reaction were 70°C and pH 9.5 and for the reductive reaction were 65°C and pH 6.5. Under the optimal conditions, the KM and kcat values with (R)-1-phenylethanol as a substrate were found to be 0.78 mM and 123 s–1 and with acetophenone they were 0.56 mM and 125 s–1, respectively. The (R)-1-phenylethanol dehydrogenase became more stable at pH 9.5 in comparison with pH 5.0 and high stability was noticed at 4 and 37°C. Properties of the enzyme place it as a promising candidate for industrial applications.  相似文献   

17.
Styrene oxide isomerase (SOI) [EC 5.3.99.7], most probably located in the cell wall, was partially purified from Coiynebacterium sp. AC-5 cells grown in a styrene gas atmospheres. The enzyme catalyzed the isomerization reaction to give phenylacetaldehyde, but did not catalyze its reverse reaction. The optimum pH of the reaction was around 7.0, and the enzyme was unstable below pH 6.0. The Km toward styrene oxide was very low (7.7 × 10?5 m), indicating its high affinity for styrene oxide. The enzyme showed strict substrate specificity, and epoxide compounds other than styrene oxide did not serve as substrates. (S)-Styrene oxide was preferentially converted by the enzyme, compared with the (R)-isomer. The possible application of SOI as a biocatalyst is also discussed.  相似文献   

18.
An NADPH-dependent aldehyde reductase was purified from rat brain microsomes to electrophoretic homogeneity. The purified enzyme had a molecular weight of 75,000 and reduced long chain fatty aldehydes such as octanal and hexadecanal with higher affinity (Km values of 0.21 mM and 0.03 mM, respectively) than for various artificial carbonyl compounds such as p-nitrobenzaldehyde and p-nitroacetophenone (Km values of 0.31 mM and 1.4 mM, respectively). The purified microsomal aldehyde reductase also showed NADPH-cytochrome c reductase activity, and it could not be distinguished from NADPH-cytochrome c reductase in molecular weight (75,000), chromatographic behavior, electrophoretic mobility, or immunological properties. The solubilized microsomal fraction treated with steapsin lost the reductase activity for hexadecanal but not that for cytochrome c. These results suggest that the aldehyde reductase in brain microsomes is identical to NADPH-cytochrome c reductase and that a hydrophobic portion of the NADPH-cytochrome c reductase is required for the reduction of hexadecanal.  相似文献   

19.
A soluble nitrate reductase from the bacterium Acinetobacter calcoaceticus grown on nitrate has been characterized. The reduction of nitrate to nitrite is mediated by an enzyme of 96000 molecular weight that can use as electron donors either viologen dyes chemically reduced with dithionite or enzymatically reduced with NAD(P)H, through specific diaphorases which utilize viologens as electron acceptors. Nitrate reductase activity is molybdenum-dependent as shown by tungstate antagonistic experiments and is sensitive to -SH reagents and metal chelators such as KCN.The enzyme synthesis is repressed by ammonia. Moreover, nitrate reductase activity undergoes a quick inactivation either by dithionite and temperature or by dithionite in the presence of small amounts of nitrate. Cyanate prevents this inactivating process and can restore the activity once the inactivation had occurred, thus suggesting that an interconversion mechanism may participate in the regulation of Acinetobacter nitrate reductase.Abbreviations EDTA ethylenediaminetetraacetate - BV benzyl viologen - MV methyl viologen - MW molecular weight - NEM N-ethylmaleimide - p-HMB p-hydroxymercuribenzoate - DCPIP 2,6-dichlorophenol-indophenol - FMN flavin mononucleotide - FAD flavin adenine dinucleotide - KCNO potassium cyanate  相似文献   

20.
Three kinds of NADPH-dependent aldehyde reducing enzymes were present in the dog kidney. Aldose reductase was located in the inner medulla region and aldehyde reductase in all regions of the renal cortex, outer medulla and inner medulla. In addition, a new reductase designated tentatively as high-Km aldose reductase, which was converted into an aldose reductase-like enzyme, was present in the inner medulla region of the kidney. Aldose reductase, aldehyde reductase and high-Km aldose reductase were purified to homogeneity from each region of the dog kidney. The molecular weight of aldose reductase was estimated to be 38,500 by SDS-polyacrylamide gel electrophoresis and the isoelectric point was found to be 5.7 by chromatofocusing. Aldose reductase had activity for aldo-sugars such as D-xylose, D-glucose and D-galactose as substrates and utilized both NADPH and NADH as coenzymes. Sulfate ions resulted in over 2-fold activation of aldose reductase. All aldehyde reductases from the three regions had the same properties. The molecular weights and isoelectric points of aldehyde reductases were 40,000 and 6.1, respectively. The aldehyde reductases were inactive for D-hexose, utilized only NADPH as coenzyme and were not affected by sulfate ions. High-Km aldose reductase had a molecular weight of 38,500 and an isoelectric point of 5.4. It had activity for aldo-sugars, but showed much higher Km and lower kcat/Km values than aldose reductase. Sulfate ions inhibited high-Km aldose reductase. It was converted into an aldose reductase-like enzyme by incubation in phosphate buffer at pH 7.0. The three kinds of enzymes were strongly inhibited by the known aldose reductase inhibitors. However, aldehyde reductase and high-Km aldose reductase were, in general, less susceptible than aldose reductase.  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