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.
We developed a biocatalyst by cloning the styrene monooxygenase genes (styA and styB) from Pseudomonas fluorescens ST responsible for the oxidation of styrene to its corresponding epoxide. Recombinant Escherichia coli was able to oxidize different aryl vinyl and aryl ethenyl compounds to their corresponding optically pure epoxides. The results of bioconversions indicate the broad substrate preference of styrene monooxygenase and its potential for the production of several fine chemicals.  相似文献   

2.
Nuclear enzymes were found to develop earlier than the corresponding microsomal activities. In fact styrene monooxygenase enzymatic activity at 18 days gestational age reached about half the values of adult animals, whereas fetal microsomal activity was only about 120 the adult level at the same age. In microsomes and nuclei the ontogenic development of epoxide hydrolase is slightly slower than styrene monooxygenase. This suggests that fetuses and newborn animals are exposed to higher risk of accumulation of styrene-7,8-oxide, a toxic and possibly teratogenic product of styrene monooxygenase.  相似文献   

3.
Solubilized cytochrome P-450 monooxygenase and epoxide hydrase activities from rat liver microsomes have been separated by column chromatography. The highly active epoxide hydrase fraction is still contaminated with cytochrome P-450, which has very low monooxygenase activity. The highly purified cytochrome P-450 fraction possesses high monooxygenase activity and is essentially devoid of epoxide hydrase activity. Purification factors for the epoxide hydrase through four purification steps are similar with [3H]styrene oxide, [3H]naphthalene oxide, [3H]cyclohexene oxide, and benzene oxide as substrates. Failure of benzene oxide to inhibit hydration of styrene or naphthalene oxide in the most purified preparations in indicative of the presence of at least two hydrases. These purified cytochrome monooxygenase and hydrase preparations represent valuable tools for the study of the intermediacy of arene oxides in drug metabolism. Thus, with naphthalene, only naphthol is formed with the monooxygenase, while both naphthol and the dihydrodiol are formed in the presence of monooxygenase and hydrase. A convenient radiochemical synthesis of [3H]naphthalene 1,2-oxide and assays for the measurement of the hydration of [3H]naphthalene oxide and benzene oxide, based on differential extractions and high-pressure liquid chromatography, respectively, are described.  相似文献   

4.
Yu Sun 《Inorganica chimica acta》2006,359(15):4807-4810
3(5)-Pyrazolyl substituted triphenylphosphines have been investigated as ligands for the palladium catalyzed Heck reaction of aryl halides with styrene. Catalysts formed in situ from those phosphines and PdII(OAc)2 are comparable in activity and selectivity with the corresponding pre-synthesized Pd(II) complexes, while Pd2(dba)3 has turned out to be a less suitable palladium source. Among the ligands investigated, the bidentate P,N-ligand 2-[3(5)-pyrazolylphenyl]diphenylphosphine has shown the highest activities for the coupling of bromobenzene with styrene in the presence of PdII(OAc)2. In the presence of 1 equiv. of nBu4NI as the additive, unreactive 4-chloroacetophenone also undergoes Heck coupling with styrene.  相似文献   

5.
A novel styrene monooxygenase (SMO) was isolated from Pseudomonas sp. LQ26, a styrene degrader from activated sludge. Sequence alignment demonstrated that it was the most distant member of all SMOs originating from the genus of Pseudomonas. The substrate spectrum of this enzyme extended beyond typical SMO substrates to 1-allylbenzene analogues, previously reported as non-substrates for the SMO from Pseudomonas fluorescens ST. The results demonstrate for the first time the asymmetric epoxidation of both conjugated and unconjugated alkenes catalyzed by SMO and suggest that a much broader substrate spectrum is expected for SMOs.  相似文献   

6.
Transformations of Aromatic Compounds by Nitrosomonas europaea   总被引:6,自引:2,他引:4       下载免费PDF全文
Benzene and a variety of substituted benzenes inhibited ammonia oxidation by intact cells of Nitrosomonas europaea. In most cases, the inhibition was accompanied by transformation of the aromatic compound to a more oxidized product or products. All products detected were aromatic, and substituents were often oxidized but were not separated from the benzene ring. Most transformations were enhanced by (NH4)2SO4 (12.5 mM) and were prevented by C2H2, a mechanism-based inactivator of ammonia monooxygenase (AMO). AMO catalyzed alkyl substituent hydroxylations, styrene epoxidation, ethylbenzene desaturation to styrene, and aniline oxidation to nitrobenzene (and unidentified products). Alkyl substituents were preferred oxidation sites, but the ring was also oxidized to produce phenolic compounds from benzene, ethylbenzene, halobenzenes, phenol, and nitrobenzene. No carboxylic acids were identified. Ethylbenzene was oxidized via styrene to two products common also to oxidation of styrene; production of styrene is suggestive of an electron transfer mechanism for AMO. Iodobenzene and 1,2-dichlorobenzene were oxidized slowly to halophenols; 1,4-dichlorobenzene was not transformed. No 2-halophenols were detected as products. Several hydroxymethyl (-CH2OH)-substituted aromatics and p-cresol were oxidized by C2H2-treated cells to the corresponding aldehydes, benzaldehyde was reduced to benzyl alcohol, and o-cresol and 2,5-dimethylphenol were not depleted.  相似文献   

7.
The 4-hydroxyacetophenone monooxygenase (HAPMO) from Pseudomonas fluorescens ACB catalyzes NADPH- and oxygen-dependent Baeyer-Villiger oxidation of 4-hydroxyacetophenone to the corresponding acetate ester. Using the purified enzyme from recombinant Escherichia coli, we found that a broad range of carbonylic compounds that are structurally more or less similar to 4-hydroxyacetophenone are also substrates for this flavin-containing monooxygenase. On the other hand, several carbonyl compounds that are substrates for other Baeyer-Villiger monooxygenases (BVMOs) are not converted by HAPMO. In addition to performing Baeyer-Villiger reactions with aromatic ketones and aldehydes, the enzyme was also able to catalyze sulfoxidation reactions by using aromatic sulfides. Furthermore, several heterocyclic and aliphatic carbonyl compounds were also readily converted by this BVMO. To probe the enantioselectivity of HAPMO, the conversion of bicyclohept-2-en-6-one and two aryl alkyl sulfides was studied. The monooxygenase preferably converted (1R,5S)-bicyclohept-2-en-6-one, with an enantiomeric ratio (E) of 20, thus enabling kinetic resolution to obtain the (1S,5R) enantiomer. Complete conversion of both enantiomers resulted in the accumulation of two regioisomeric lactones with moderate enantiomeric excess (ee) for the two lactones obtained [77% ee for (1S,5R)-2 and 34% ee for (1R,5S)-3]. Using methyl 4-tolyl sulfide and methylphenyl sulfide, we found that HAPMO is efficient and highly selective in the asymmetric formation of the corresponding (S)-sulfoxides (ee >99%). The biocatalytic properties of HAPMO described here show the potential of this enzyme for biotechnological applications.  相似文献   

8.
The styAB genes from Pseudomonas putida CA-3, which encode styrene monooxygenase, were subjected to three rounds of in vitro evolution using error-prone polymerase chain reaction with a view to improving the rate of styrene oxide and indene oxide formation. Improvements in styrene monooxygenase activity were monitored using an indole to indigo conversion assay. Each round of random mutagenesis generated variants improved in indigo formation with third round variants improved nine- to 12-fold over the wild type enzyme. Each round of in vitro evolution resulted in two to three amino acid substitutions in styrene monooxygenase. While the majority of mutations occurred in styA (oxygenase), mutations were also observed in styB (reductase). A mutation resulting in the substitution of valine with isoleucine at amino acid residue 303 occurred near the styrene and flavin adenine dinucleotide binding site of styrene monooxygenase. One mutation caused a shift in the reading frame in styA and resulted in a StyA variant that is 19 amino acids longer than the wild-type protein. Whole cells expressing the best styrene monooxygenase variants (round 3) exhibited eight- and 12-fold improvements in styrene and indene oxidation rates compared to the wild-type enzyme. In all cases, a single enantiomer, (S)-styrene oxide, was formed from styrene while (1S,2R)-indene oxide was the predominant enantiomer (e.e. 97%) formed from indene. The average yield of styrene oxide and indene oxide from their respective alkene substrates was 65% and 90%, respectively.  相似文献   

9.
Aspergillus flavus isolates produce only aflatoxins B1 and B2, while Aspergillus parasiticus and Aspergillus nomius produce aflatoxins B1, B2, G1, and G2. Sequence comparison of the aflatoxin biosynthesis pathway gene cluster upstream from the polyketide synthase gene, pksA, revealed that A. flavus isolates are missing portions of genes (cypA and norB) predicted to encode, respectively, a cytochrome P450 monooxygenase and an aryl alcohol dehydrogenase. Insertional disruption of cypA in A. parasiticus yielded transformants that lack the ability to produce G aflatoxins but not B aflatoxins. The enzyme encoded by cypA has highest amino acid identity to Gibberella zeae Tri4 (38%), a P450 monooxygenase previously shown to be involved in trichodiene epoxidation. The substrate for CypA may be an intermediate formed by oxidative cleavage of the A ring of O-methylsterigmatocystin by OrdA, the P450 monooxygenase required for formation of aflatoxins B1 and B2.  相似文献   

10.
We describe the development of biocatalysis for producing optically pure straight-chain (S)-epoxyalkanes using styrene monooxygenase of Rhodococcus sp. strain ST-10 (RhSMO). RhSMO was expressed in the organic solvent-tolerant microorganism Kocuria rhizophila DC2201, and the bioconversion reaction was performed in an organic solvent-water biphasic reaction system. The biocatalytic process enantioselectively converted linear terminal alkenes to their corresponding (S)-epoxyalkanes using glucose and molecular oxygen. When 1-heptene and 6-chloro-1-hexene were used as substrates (400 mM) under optimized conditions, 88.3 mM (S)-1,2-epoxyheptane and 246.5 mM (S)-1,2-epoxy-6-chlorohexane, respectively, accumulated in the organic phase with good enantiomeric excess (ee; 84.2 and 95.5%). The biocatalysis showed broad substrate specificity toward various aliphatic alkenes, including functionalized and unfunctionalized alkenes, with good to excellent ee. Here, we demonstrate that this biocatalytic system is environmentally friendly and useful for producing various enantiopure (S)-epoxyalkanes.  相似文献   

11.
Styrene can efficiently be oxidized to (S)-styrene oxide by recombinant Escherichia coli expressing the styrene monooxygenase genes styAB from Pseudomonas sp. strain VLB120. Targeting microbial physiology during whole-cell redox biocatalysis, we investigated the interdependency of styrene epoxidation, growth, and carbon metabolism on the basis of mass balances obtained from continuous two-liquid-phase cultures. Full induction of styAB expression led to growth inhibition, which could be attenuated by reducing expression levels. Operation at subtoxic substrate and product concentrations and variation of the epoxidation rate via the styrene feed concentration allowed a detailed analysis of carbon metabolism and bioconversion kinetics. Fine-tuned styAB expression and increasing specific epoxidation rates resulted in decreasing biomass yields, increasing specific rates for glucose uptake and the tricarboxylic acid (TCA) cycle, and finally saturation of the TCA cycle and acetate formation. Interestingly, the biocatalysis-related NAD(P)H consumption was 3.2 to 3.7 times higher than expected from the epoxidation stoichiometry. Possible reasons include uncoupling of styrene epoxidation and NADH oxidation and increased maintenance requirements during redox biocatalysis. At epoxidation rates of above 21 μmol per min per g cells (dry weight), the absence of limitations by O2 and styrene and stagnating NAD(P)H regeneration rates indicated that NADH availability limited styrene epoxidation. During glucose-limited growth, oxygenase catalysis might induce regulatory stress responses, which attenuate excessive glucose catabolism and thus limit NADH regeneration. Optimizing metabolic and/or regulatory networks for efficient redox biocatalysis instead of growth (yield) is likely to be the key for maintaining high oxygenase activities in recombinant E. coli.  相似文献   

12.
Apparent Km- and Vmax-values of nuclear styrene 7,8-oxide hydrolase were determined at different protein concentrations. In the protein concentrations range used no significant differences in the apparent Km-values were observed. The influence of the incubation with different modifiers (i.e. SKF-525A, metyrapone, 1,2-epoxy-3,3,3 trichloropropane, cyclohexene oxide) at two different concentrations on this enzyme activity was also determined. Cyclohexene oxide and 1,2-epoxy-3,3,3-trichloropropane, two well known inhibitors of the microsomal epoxide hydrolase(s) caused a marked inhibition, metyrapone had a strong activating effect whereas SKF-525A had no effect. In vivo pretreatment with phenobarbital significantly induced the nuclear epoxide hydrolase whereas β-naphthoflavone caused a lower degree of induction. This pattern is quantitatively different but qualitatively very similar to the microsomal one. Moreover a toxifying to detoxifying enzymatic activity balance is attempted for the metabolization of the alkenic double bond of styrene, taking into account the ratio between the styrene monooxygenase (toxifying enzyme) and the styrene 7,8-oxide hydrolase (detoxifying enzyme) after the above mentioned pretreatments, both in the microsomal and nuclear fractions.  相似文献   

13.
The microsomal FAD-containing monooxygenase (EC 1.14.13.8, dimethylaniline monooxygenase) purified to homogeneity from hog liver catalyzes NADPH- and oxygen-dependent S-oxygenation of phenylthiourea, ethylenethiourea, thiocarbanilide, N-methylthiourea, and thiourea to their corresponding formamidine sulfinic acids. The sulfinic acids are formed by sequential enzymic oxidation of the thioureas through intermediate sulfenic acids. The reaction sequence was established by separating intermediate and final oxygenated metabolites of phenylthiourea and ethylenethiourea. The sulfenic and sulfinic acids of these two thioureas, produced enzymically, were chromatographically and spectrally identical with chemically synthesized reference compounds. Phenylformamidine and ethyleneformamidine sulfinic acids are slowly converted to their sulfonic acids upon prolonged incubation. While N-substituted formamidine sulfinic acids oxidize spontaneously to formamidine sulfonic acids at 37 °C, the further oxidation of ethyleneformamidine sulfinic acid may be, at least in part, enzyme catalyzed. The purified monooxygenase also catalyzes rapid oxygenation of mercaptoimidazoles to the corresponding imidazole sulfinic acids. The instability of S-oxygenated mercaptoimidazoles prevented their isolation and positive identification, but analysis of kinetic data obtained with sulfenic acid trapping agents suggests that these compounds are oxygenated by the same reaction sequence established for N-substituted thioureas. The NADPH- and oxygen-dependent oxidation of thiocarbamates and of 2-mercaptoimidazoles catalyzed by hog or hamster liver microsomes correlates with dimethylaniline N-oxidase activity and appears completely independent from cytochrome P-450. The S-oxidation of thiourea and its derivatives is not inhibited by n-octylamine, a known inhibitor of cytochrome P-450 dependent oxygenations. Furthermore, differential thermal inactivation of the flavin-containing monooxygenase totally abolishes phenylthiourea S-oxidase activity of hamster liver microsomes.  相似文献   

14.
Mycobacterium strain NBB4 is an ethene-oxidising micro-organism isolated from estuarine sediments. In pursuit of new systems for biocatalytic epoxidation, we report the capacity of strain NBB4 to convert a diverse range of alkene substrates to epoxides. A colorimetric assay based on 4-(4-nitrobenzyl)pyridine) has been developed to allow the rapid characterisation and quantification of biocatalytic epoxide synthesis. Using this assay, we have demonstrated that ethene-grown NBB4 cells epoxidise a wide range of alkenes, including terminal (propene, 1-butene, 1-hexene, 1-octene and 1-decene), cyclic (cyclopentene, cyclohexene), aromatic (styrene, indene) and functionalised substrates (allyl alcohol, dihydropyran and isoprene). Apparent specific activities have been determined and range from 2.5 to 12.0 nmol min?1 per milligram of cell protein. The enantioselectivity of epoxidation by Mycobacterium strain NBB4 has been established using styrene as a test substrate; (R)-styrene oxide is produced in enantiomeric excesses greater than 95%. Thus, the ethene monooxygenase of Mycobacterium NBB4 has a broad substrate range and promising enantioselectivity, confirming its potential as a biocatalyst for alkene epoxidation.  相似文献   

15.
The effect of aryl substituents on the rate at which epoxide hydrase catalyzes the addition of water to styrene and cis-stilbene oxides has been examined. Plots of log Vm for each substrate versus the Hammett σ constants for the substituent suggest that a nucleophilic attack occurs and that a free carbonium ion form of the substrate is not involved at the rate-determining step in the mechanism. For the stilbene oxides, high selectivity for attack by water at the carbon atom with [S] absolute stereochemistry was observed.  相似文献   

16.
The toluene/o-xylene monooxygenase cloned from Pseudomonas stutzeri OX1 displays a very broad range of substrates and a very peculiar regioselectivity, because it is able to hydroxylate more than one position on the aromatic ring of several hydrocarbons and phenols. The nucleotide sequence of the gene cluster coding for this enzymatic system has been determined. The sequence analysis revealed the presence of six open reading frames (ORFs) homologous to other genes clustered in operons coding for multicomponent monooxygenases found in benzene- and toluene-degradative pathways cloned from Pseudomonas strains. Significant similarities were also found with multicomponent monooxygenase systems for phenol, methane, alkene, and dimethyl sulfide cloned from different bacterial strains. The knockout of each ORF and complementation with the wild-type allele indicated that all six ORFs are essential for the full activity of the toluene/o-xylene monooxygenase in Escherichia coli. This analysis also shows that despite its activity on both hydrocarbons and phenols, toluene/ o-xylene monooxygenase belongs to a toluene multicomponent monooxygenase subfamily rather than to the monooxygenases active on phenols.  相似文献   

17.
Xylene monooxygenase of Pseudomonas putida mt-2 catalyzes multistep oxidations of one methyl group of toluene and xylenes. Recombinant Escherichia coli expressing the monooxygenase genes xylM and xylA catalyzes the oxygenation of toluene, pseudocumene, the corresponding alcohols, and the corresponding aldehydes, all by a monooxygenation type of reaction (B. Bühler, A. Schmid, B. Hauer, and B. Witholt, J. Biol. Chem. 275:10085-10092, 2000). Using E. coli expressing xylMA, we investigated the kinetics of this one-enzyme three-step biotransformation. We found that unoxidized substrates like toluene and pseudocumene inhibit the second and third oxygenation steps and that the corresponding alcohols inhibit the third oxygenation step. These inhibitions might promote the energetically more favorable alcohol and aldehyde dehydrogenations in the wild type. Growth of E. coli was strongly affected by low concentrations of pseudocumene and its products. Toxicity and solubility problems were overcome by the use of a two-liquid-phase system with bis(2-ethylhexyl)phthalate as the carrier solvent, allowing high overall substrate and product concentrations. In a fed-batch-based two-liquid-phase process with pseudocumene as the substrate, we observed the consecutive accumulation of aldehyde, acid, and alcohol. Our results indicate that, depending on the reaction conditions, product formation could be directed to one specific product.  相似文献   

18.
Biological oxidation of cyclic ketones normally results in formation of the corresponding dicarboxylic acids, which are further metabolized in the cell. Rhodococcus ruber strain SC1 was isolated from an industrial wastewater bioreactor that was able to utilize cyclododecanone as the sole carbon source. A reverse genetic approach was used to isolate a 10-kb gene cluster containing all genes required for oxidative conversion of cyclododecanone to 1,12-dodecanedioic acid (DDDA). The genes required for cyclododecanone oxidation were only marginally similar to the analogous genes for cyclohexanone oxidation. The biochemical function of the enzymes encoded on the 10-kb gene cluster, the flavin monooxygenase, the lactone hydrolase, the alcohol dehydrogenase, and the aldehyde dehydrogenase, was determined in Escherichia coli based on the ability to convert cyclododecanone. Recombinant E. coli strains grown in the presence of cyclododecanone accumulated lauryl lactone, 12-hydroxylauric acid, and/or DDDA depending on the genes cloned. The cyclododecanone monooxygenase is a type 1 Baeyer-Villiger flavin monooxygenase (FAD as cofactor) and exhibited substrate specificity towards long-chain cyclic ketones (C11 to C15), which is different from the specificity of cyclohexanone monooxygenase favoring short-chain cyclic compounds (C5 to C7).  相似文献   

19.
The white-rot basidiomycetes efficiently degrade all wood cell wall polymers. Generally, these fungi simultaneously degrade cellulose and lignin, but certain organisms, such as Ceriporiopsis subvermispora, selectively remove lignin in advance of cellulose degradation. However, relatively little is known about the mechanism of selective ligninolysis. To address this issue, C. subvermispora was grown in liquid medium containing ball-milled aspen, and nano-liquid chromatography-tandem mass spectrometry was used to identify and estimate extracellular protein abundance over time. Several manganese peroxidases and an aryl alcohol oxidase, both associated with lignin degradation, were identified after 3 days of incubation. A glycoside hydrolase (GH) family 51 arabinofuranosidase was also identified after 3 days but then successively decreased in later samples. Several enzymes related to cellulose and xylan degradation, such as GH10 endoxylanase, GH5_5 endoglucanase, and GH7 cellobiohydrolase, were detected after 5 days. Peptides corresponding to potential cellulose-degrading enzymes GH12, GH45, lytic polysaccharide monooxygenase, and cellobiose dehydrogenase were most abundant after 7 days. This sequential production of enzymes provides a mechanism consistent with selective ligninolysis by C. subvermispora.  相似文献   

20.
Epoxide hydrase activity, measured with [3H]styrene oxide as substrate, is present in mammalian liver, kidney, lung, intestine and skin. The hepatic level of the enzyme, measured in vitro with [3H]styrene oxide, benzene oxide or naphthalene-1,2-oxide, is elevated substantially by pretreatment of rats with phenobarbital and to a lesser extent by pretreatment with 3-methylcholanthrene. Metyrapone and 1-(2-isopropylphenyl)-imidazole, two monooxygenase inhibitors, activate epoxide hydrase in vitro, but have no demonstrable effect on the enzyme in vivo. 3,3,3-Trichloropropene oxide, a potent in vitro inhibitor of epoxide hydrase, has no effect on monooxygenase activity measured in vitro with [3H]benzenesulfonanilide. Trichloropropene oxide is extremely toxic. In sub-lethal dosages, it does not significantly inhibit epoxide hydrase activity in vivo, although it and several other epoxides do react with and thereby reduce hepatic levels of glutathione. Cyclohexane oxide, another potent in vitro inhibitor of epoxide hydrase, reduces hepatic glutathione levels to 10% of control values. This relatively non-toxic substance should potentiate the hepatotoxicity of chlorobenzene by inhibiting further metabolism of the toxic chlorobenzene oxide intermediate through either hydration or conjugation with glutathione. Instead, co-administration of cyclohexene oxide and chlorobenzene significantly reduces the rate of metabolism of [14C]chlorobenzene and prevents the hepatic centrilobular necrosis caused by chlorobenzene in rats. Arene oxide-mediated hepatotoxicity apparently is dependent upon a variety of factors including both rates of formation and degradation of arene oxides in tissue. The presently known hydrase inhibitors are not sufficiently selective in their effects on liver cells to permit a quantitative assessment of the relative importance of these factors.  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