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 marine cyanobacterium Oscillatoria sp. strain JCM oxidized naphthalene predominantly to 1-naphthol. Experiments with [1-2H]naphthalene and [2-2H]naphthalene indicated that 1-naphthol was formed with 68 and 74% retention of deuterium, respectively. No significant isotope effect was observed when the organism was incubated with a 1:1 mixture of naphthalene and [2H8]naphthalene. The results indicate that 1-naphthol is formed through a naphthalene 1,2-oxide intermediate, which rearranges spontaneously via an NIH shift mechanism.  相似文献   

2.
Bacillus cereus ATCC 14579 transformed naphthalene predominately to 1-naphthol. Experiments with [14C]naphthalene showed that over a 24 h period, B. cereus oxidized 5.2% of the added naphthalene. 1-Naphthol accounted for approximately 80% of the total metabolites. B. cereus incubated with naphthalene under the presence of 18O2 led to the isolation of 1-naphthol that contained 94% 18O. The metabolism of [1-2H]-and [2-2H]-naphthalene by B. cereus yielded 1-naphthol which retained 95% and 94% deuterium, respectively, as determined by mass spectral analysis. NMR spectroscopic analysis of the deuterated 1-naphthol formed from [1-2H]-naphthalene indicated an NIH shift mechanism in which 19% of the deuterium migrated from the C-1 to the C-2 position. The 18O2 and NIH shift experiments implicate naphthalene-1,2-oxide as an intermediate in the formation of 1-naphthol from naphthalene by B. cereus.Abbreviations HPLC High performance liquid chromatography - NMR nuclear magnetic resonance  相似文献   

3.
Recent studies of a number of volatile aromatic hydrocarbons have suggested that the formation of covalently bound metabolites arises solely through the intermediate formation of phenols. This study further examines the involvement of 1-naphthol in the in vivo and in vitro formation of covalently bound metabolites and pulmonary bronchiolar necrosis by naphthalene. Marked differences were observed in the rate of 1-naphthol formation in lung and liver microsomal incubations without correspondingly large differences between the rates of formation of covalently bound metabolites from naphthalene and 1-naphthol. Glutathione decreased covalent binding in hepatic microsomal incubations containing 14[C]1-naphthol but did not result in the formation of any of the glutathione adducts isolated from identical incubations containing 14[C]naphthalene. Tissue levels of covalently bound radioactivity in mice treated with 14[C]1-naphthol or 14[C]naphthalene were similar; however, in contrast to studies with naphthalene, 1-naphthol administration did not deplete tissue glutathione nor result in detectable tissue injury. These studies indicate that 1-naphthol is not an obligate intermediate in the formation of covalently bound metabolites from naphthalene nor does it appear to be a more proximate lung toxic metabolite.  相似文献   

4.
Cunninghamella elegans oxidized naphthalene to ethyl acetate-soluble and water-soluble metabolites. Experiments with [14C]-naphthalene indicated that 21% of the substrate was converted into metabolites. The ratio of organic-soluble metabolites to water-soluble metabolites was 76:24. The major ethyl acetate-soluble naphthalene metabolites were trans-1,2-dihydroxy-1,2-dihydro-naphthalene, 4-hydroxy-1-tetralone, and 1-naphthol. Enzymatic treatment of the aqueous phase with either arylsulfatase or beta-glucuronidase released metabolites of naphthalene that were extractable with ethyl acetate. In both cases, the major metabolite was 1-naphthol. The ratio of water-soluble sulfate conjugates to water-soluble glucuronide conjugates was 1:1. Direct analysis of the aqueous phase by high-pressure liquid and thin-layer chromatographic and mass spectrometric techniques indicated that 1-naphthyl sulfate and 1-naphthyl glucuronic acid were major water-soluble metabolites formed from the fungal metabolism of naphthalene. C. elegans oxidized biphenyl primarily to 4-hydroxy biphenyl. Deconjugation experiments with biphenyl water-soluble metabolites indicated that the glucuronide and sulfate ester of 4-hydroxy biphenyl were metabolites. The data demonstrate that sulfation and glucuronidation are major pathways in the metabolism of aromatic hydrocarbons by fungi.  相似文献   

5.
The regiospecific oxidation of the polycyclic aromatic hydrocarbons naphthalene and fluorene was examined with Escherichia coli strains expressing wildtype toluene 4-monooxygenase (T4MO) from Pseudomonas mendocina KR1, toluene para-monooxygenase (TpMO) from Ralstonia pickettii PKO1, toluene ortho-monooxygenase (TOM) from Burkholderia cepacia G4, and toluene/ortho-xylene monooxygenase (ToMO) from P. stutzeri OX1. T4MO oxidized toluene (12.1+/-0.8 nmol/min/mg protein at 109 microM), naphthalene (7.7+/-1.5 nmol/min/mg protein at 5 mM), and fluorene (0.68+/-0.04 nmol/min/mg protein at 0.2 mM) faster than the other wildtype enzymes (2-22-fold) and produced a mixture of 1-naphthol (52%) and 2-naphthol (48%) from naphthalene, which was successively transformed to a mixture of 2,3-, 2,7-, 1,7-, and 2,6-dihydroxynaphthalenes (7%, 10%, 20%, and 63%, respectively). TOM and ToMO made 1,7-dihydroxynaphthalene from 1-naphthol, and ToMO made a mixture of 2,3-, 2,6-, 2,7-, and 1,7-dihydroxynaphthalene (26%, 22%, 1%, and 44%, respectively) from 2-naphthol. TOM had no activity on 2-naphthol, and T4MO had no activity on 1-naphthol. To take advantage of the high activity of wildtype T4MO but to increase its regiospecificity on naphthalene, seven engineered enzymes containing mutations in T4MO alpha hydroxylase TmoA were examined; the selectivity for 2-naphthol by T4MO I100A, I100S, and I100G was enhanced to 88-95%, and the selectivity for 1-naphthol was enhanced to 87% and 99% by T4MO I100L and G103S/A107G, respectively, while high oxidation rates were maintained except for G103S/A107G. Therefore, the regiospecificity for naphthalene oxidation was altered to practically pure 1-naphthol or 2-naphthol. All four wildtype monooxygenases were able to oxidize fluorene to different monohydroxylated products; T4MO oxidized fluorene successively to 3-hydroxyfluorene and 3,6-dihydroxyfluorene, which was confirmed by gas chromatography-mass spectrometry and 1H nuclear magnetic resonance analysis. TOM and its variant TomA3 V106A oxidize fluorene to a mixture of 1-, 2-, 3-, and 4-hydroxyfluorene. This is the first report of using enzymes to synthesize 1-, 3-, and 4-hydroxyfluorene, and 3,6-dihydroxyfluorene from fluorene as well as 2-naphthol and 2,6-dihydroxynaphthalene from naphthalene.  相似文献   

6.
The ability of naphthalene 1,2-oxide to diffuse across intact cellular membranes, the subsequent biotransformation of this epoxide and its potential to produce losses in cellular viability have been examined in incubations of isolated hepatocytes. Addition of 1R,2S- or 1S,2R-naphthalene oxide enantiomers (15, 30 and 60 microM) to isolated hepatocytes resulted in a rapid depletion of intracellular glutathione. Depletion of glutathione was concentration dependent and maximal at 5-15 min. Addition of either of the enantiomeric oxides at 60 microM resulted in the loss of more than 20 nmol glutathione/10(6) cells (1 ml cells); thus more than a third of the added epoxide was available for conjugation with intracellular glutathione. The time course and concentration dependence of glutathione depletion corresponded to the rapid, concentration-dependent formation of naphthalene oxide glutathione conjugates. The levels of glutathione adduct were highest 1 min after addition of naphthalene oxide and declined to 25% of this level after 30 min. Loss of glutathione conjugates from incubations correlated with the formation of N-acetylcysteine adducts. In contrast, the levels of glutathione adducts added exogenously to hepatocytes were relatively stable over a 120-min incubation suggesting that although further metabolism of naphthalene oxide glutathione adducts formed intracellularly is possible, extracellular glutathione adducts cannot penetrate the hepatocellular membrane. Small amounts of radiolabel from [3H]naphthalene 1,2-oxide were bound covalently to macromolecules in hepatocytes; the rate of this binding slowed rapidly after the first minute of incubation. Severe blebbing of the surface of the hepatocytes was noted in cells incubated for 30 min with 480 microM naphthalene oxide. Many of the cells were vacuolated at 60 min and progressed to frank necrosis with pyknotic nuclei and inability to exclude trypan blue. Cells incubated with 1-naphthol responded in a qualitatively similar fashion to those cells incubated with epoxide; however, hepatocytes incubated with 1-naphthol progressed to frank cellular necrosis at a slower rate. In hepatocytes partially depleted of glutathione by pretreatment with buthionine sulfoximine, addition of 1S,2R-naphthalene oxide at a rate of 1 nmol/min/10(6) cells resulted in significant losses in cell viability. In contrast, no losses in cell viability were observed with the enantiomer, 1R,2S-naphthalene oxide. Both epoxides produced similar losses in cellular glutathione levels.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

7.
Anaerobic naphthalene degradation by a sulfate-reducing enrichment culture   总被引:5,自引:0,他引:5  
Anaerobic naphthalene degradation by a sulfate-reducing enrichment culture was studied by substrate utilization tests and identification of metabolites by gas chromatography-mass spectrometry. In substrate utilization tests, the culture was able to oxidize naphthalene, 2-methylnaphthalene, 1- and 2-naphthoic acids, phenylacetic acid, benzoic acid, cyclohexanecarboxylic acid, and cyclohex-1-ene-carboxylic acid with sulfate as the electron acceptor. Neither hydroxylated 1- or 2-naphthoic acid derivatives and 1- or 2-naphthol nor the monoaromatic compounds ortho-phthalic acid, 2-carboxy-1-phenylacetic acid, and salicylic acid were utilized by the culture within 100 days. 2-Naphthoic acid accumulated in all naphthalene-grown cultures. Reduced 2-naphthoic acid derivatives could be identified by comparison of mass spectra and coelution with commercial reference compounds such as 1,2,3, 4-tetrahydro-2-naphthoic acid and chemically synthesized decahydro-2-naphthoic acid. 5,6,7,8-Tetrahydro-2-naphthoic acid and octahydro-2-naphthoic acid were tentatively identified by their mass spectra. The metabolites identified suggest a stepwise reduction of the aromatic ring system before ring cleavage. In degradation experiments with [1-(13)C]naphthalene or deuterated D(8)-naphthalene, all metabolites mentioned derived from the introduced labeled naphthalene. When a [(13)C]bicarbonate-buffered growth medium was used in conjunction with unlabeled naphthalene, (13)C incorporation into the carboxylic group of 2-naphthoic acid was shown, indicating that activation of naphthalene by carboxylation was the initial degradation step. No ring fission products were identified.  相似文献   

8.
A soil Pseudomonas sp. grew with 1-naphthol as sole organic carbon source and produced a 3,4-dihydro-dihydroxy-1(2H)-naphthalenone as the main early intermediary metabolite. Washed 1-naphthol-grown organisms oxidized naphthalene, 1- or 2-naphthol, salicylic acid and, to some extent, 2,3-dihydroxybenzoic acid.  相似文献   

9.
Metabolism of naphthalene by cell extracts of Cunninghamella elegans.   总被引:3,自引:0,他引:3  
Microsomal preparations of Cunninghamella elegans oxidized naphthalene to trans-1,2-dihydroxy-1,2-dihydronaphthalene, 1-naphthol, and 2-naphthol. Enzymatic activity was dependent on the presence of reduced nicotinamide adenine dinucleotide phosphate and oxygen. Reduced microsomal preparations, when treated with carbon monoxide, showed absorption maxima at 450 and 420 nm. The inhibitor 1,2-epoxy-3,3,3-trichloropropane suppressed the formation of trans-1,2-dihydroxy-1,2-dihydronaphthalene and enhanced 1-naphthol formation. The results suggest that the metabolism of naphthalene by fungal microsomes may be analogous to the cytochrome P-450-dependent monooxygenase activity that is associated with mammalian liver microsomes.  相似文献   

10.
The regulation of naphthalene and 1-naphthol metabolism in a Rhodococcus sp. (NCIMB 12038) has been investigated. The microorganism utilizes separate pathways for the degradation of these compounds, and they are regulated independently. Naphthalene metabolism was inducible, but not by salicylate, and 1-naphthol metabolism, although constitutive, was also repressed during growth on salicylate. The biochemistry of naphthalene degradation in this strain was otherwise identical to that found in Pseudomonas putida, with salicylate as a central metabolite and naphthalene initially being oxidized via a naphthalene dioxygenase enzyme to cis-(1R,2S)-1,2-dihydroxy-1,2-dihydronaphthalene (naphthalene cis-diol). A dioxygenase enzyme was not expressed under growth conditions which facilitate 1-naphthol degradation. However, biotransformations with indene as a substrate suggested that a monooxygenase enzyme may be involved in the degradation of this compound. Indole was transformed to indigo by both naphthalene-grown NCIMB 12038 and by cells grown in the absence of an inducer. Therefore, the presence of a naphthalene dioxygenase enzyme activity was not necessary for this reaction. Thus, the biotransformation of indole to indigo may be facilitated by another type of enzyme (possibly a monooxygenase) in this organism.  相似文献   

11.
Eighty-six species of fungi belonging to sixty-four genera were examined for their ability to metabolize naphthalene. Analysis by thin-layer and high pressure liquid chromatography revealed that naphthalene metabolism occurred in forty-seven species belonging to thirty-four genera from the major fungal taxa. All organisms tested from the order Mucorales oxidized naphthalene with species of Cunninghamella, Syncephalastrum and Mucor showing the greatest activity. Significant metabolism was also observed with Neurospora crassa, Claviceps paspali and four species of Psilocybe. The predominant metabolite formed by most organisms was 1-naphthol. Other products identified were, 4-hydroxy-1-tetralone, trans-1,2-dihydroxy-1,2-dihydronaphthalene, 2-naphthol, 1,2-and 1,4-naphthoquinone.  相似文献   

12.
Six yeasts were examined for their ability to metabolize naphthalene, biphenyl and benzo(a)pyrene. All of the organisms tested oxidized these aromatic hydrocarbons. Candida lipolytica oxidized naphthalene to 1-naphthol, 2-naphthol, 4-hydroxy-1-tetralone and trans-1,2-dihydroxy-1,2-dihydronaphthalene. The major metabolite was 1-naphthol. C. lipolytica oxidized biphenyl to produce 2-, 3-, and 4-hydroxybiphenyl, 4,4′-dihydroxybiphenyl and 3-methoxy-4-hydroxybiphenyl. 4-Hydroxybiphenyl was the predominant metabolite formed. C. lipolytica oxidized benzo(a)pyrene to 3-hydroxybenzo(a)pyrene and 9-hydroxybenzo(a)pyrene. Metabolites were isolated and identified by absorption spectrophotometry, mass spectrometry and thin-layer, gasliquid and high-pressure liquid chromatography. Where possible the structures of these metabolites were confirmed by comparison with authenic compounds.  相似文献   

13.
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   

14.
Ethanol oxidation causes redox effects. The coupling of this oxidation via NADH to intermediary metabolism was studied in order to reveal the underlying mechanisms. Isolated rat hepatocytes were incubated with [1,1-2H2]-, (1R)-[1-2H]- and (1S)-[1-2H]-ethanol and the 2H incorporation was measured in lactate, beta-hydroxybutyrate, fumarate, malate, succinate, alpha-oxoglutarate and citrate. The results differed in the following ways from results obtained in intact rats. Lactate became labelled to an increasing extent, and in more than one position, indicating labelling of pyruvate. A small and constant fraction of malate and fumarate was formed without access to [2H]coenzyme. Addition of aspartate increased this fraction, which was concluded to be formed in the mitochondria. Citrate was essentially unlabelled. The 2H from (1R)-[1-2H]ethanol contributed to malate to a larger extent and to beta-hydroxybutyrate to a smaller extent, and 2H from (1S)-[1-2H]ethanol contributed to lactate to a smaller extent. These results indicate that the exchange via shuttle system was less efficient in isolated hepatocytes than in intact rats. The 2H incorporation was independent of concentration of [1,1-2H2]ethanol when this was above 5mM. Additions known to increase ethanol elimination, and cyanamide, which decreases it, had no marked effect on the 2H incorporation. This indicates equilibration of the NADH bound to alcohol dehydrogenase with free NADH. Disulfiram and cyanamide caused a decrease in the relative incorporation from (1S)-[1-2H]ethanol into malate in liver of intact rats. Addition of cyanamide to incubations with hepatocytes resulted in a decrease of the contribution of 2H from (1S)-[1-2H]ethanol in lactate, beta-hydroxybutyrate and malate. This indicates that acetaldehyde was only oxidized in the mitochondrial compartment.  相似文献   

15.
The mechanism of 3-dehydroquinate synthase was explored by incubating partially purified enzyme with mixtures of [1-14C]3-deoxy-D-arabino-heptulosonic acid 7-phosphate (DAHP) and one of the specifically tritiated substrates [4-3H]DAHP, [5-3H]DAHP, [6-3H]DAHP, (7RS)-[7-3H]DAHP, (7R)-[7-3H]DAHP, or (7S)-[7-3H]DAHP. Kinetic and secondary 3H isotope effects were calculated from 3H:14C ratios obtained in unreacted DAHP, 3-dehydroquinate, and 3-dehydroshikimate. 3H was not incorporated from the medium into 3-dehydroquinate, indicating that a carbanion (or methyl group) at C-7 is not formed. A kinetic isotope effect kH/k3H of 1.7 was observed at C-5, and afforded support for a mechanism involving oxidation of C-5 with NAD. A similar kinetic isotope effect was found at C-6 owing to removal of a proton in elimination of phosphate, which is reasonably assumed to be the next step in 3-dehydroquinate synthase. Hydrogen at C-7 of DAHP was not lost in the cyclization step of the reaction, indicating that the enol formed in phosphate elimination participated directly in an aldolase-type reaction with the carbonyl at C-2. In the dehydration of 3-dehydroquinate to 3-dehydroshikimate the (7R) proton from (7RS)- or (7R)-[7-3H]DAHP is lost, indicating that the 7R proton occupies the 2R position in dehydroquinate. Hence the cyclization step occurs with inversion of configuration at C-7. A kinetic isotope effect kH/k3H = 2.3 was observed in the conversion of (2R)-[2-3H]dehydroquinate to dehydroshikimate. Hence loss of a proton from the enzyme-dehydroquinate imine contributed to rate limitation in the reaction.  相似文献   

16.
1. R-Acetate was generated enzymically from R-acetate in the sequence acetate leads to malate leads to oxaloacetate leads to acetate, and S-acetate likewise from S-acetate. It was concluded that the formation of malate on malate synthase involves the operation of a normal isotopic effect combined with inversion of configuration. The malate synthase kH/k2H was determined as 3.7 +/- 0.5 by a method which yields results independently of the stereochemical purity of the chiral acetates used initially. 2. R-Acetate was also generated from R-acetate in the sequence acetate leads to citrate leads to malate leads to oxaloacetate leads to acetate, and S-acetate likewise from S-acetate. The conclusion is the same as given above, but refers to the formation of citrate on the re-synthase. 3. 2S,3R-[2-2H1,3-2H1,3H1]Malate and 2S,3S-[2-2H1,3-2H1]malate were prepared from 2S-[2,3-2H3]malate by treatment with fumarase in tritiated water and normal water, respectively. It was assumed that these malate specimens were pure with respect to chirality as generated by isotopic labelling. 4. These two malate specimens were partially converted (about 9%) to acetates in conditions where no racemization at the level of transiently formed oxaloacetate occurred. That no racemization took place was demonstrated experimentally. Oxidative enzymic hydrolysis of 2S,3R-[2-2H1,3-2H1,3H1]malate in normal water and of 2S,3S-[2-2H1,3-2H1]malate in tritiated water produced S-[2H1,3H1]acetate and R-[2H1,3H1]acetate, respectively. 5. The isolated R-[2H1,3H1]acetate and S-[2H1,3H1]acetate on configurational analysis yielded malates which in the presence of fumarase retained 79.7 +/- 0.7% and 20.3 +/- 0.9%, respectively, of their total tritium content. The symmetric deviation from the 50% value found with [3H1]acetate strengthens the conclusion that stereochemically pure chiral acetates were analyzed. The malate synthase kH/k2H was determined from the data of this study as 3.9 +/- 0.2. 6. The average of the values given under paragraphs 1 and 5 for the isotopic discrimination on malate synthase corresponds to kH/k2H=3.8 +/- 0.1. It was concluded that the configurational analysis of stereochemically pure R-[2H1,3H1]acetate and S-[2H1,3H1]acetate yields malates which in the presence of fumarase retain 79 +/- 2% and 21 +/- 2%, respectively, of their total tritium content. Hence, a deviation of 29 +/- 2% from the 50% value represents the actual amplitude of the configurational assay. 7. Outlines are given for an enzymic generation of chiral acetates in preparative scale.  相似文献   

17.
The metabolism of D-[U-14C]glucose and the incorporation of [methyl-3H]thymidine by aerobic and anaerobic marine sediment microbes exposed to 1 to 1,000 ppm anthracene, naphthalene, p,p'-dichlorodiphenyltrichloroethane, and pentachlorophenol were examined. Cell-specific rates of [14C]glucose metabolism averaged 1.7 X 10(-21) and 0.5 X 10(-21) mol/min per cell for aerobic and anaerobic sediment slurries, respectively; [3H]thymidine incorporation rates averaged 43 X 10(-24) and 9 X 10(-24) mol/min per cell for aerobic and anaerobic slurries, respectively. Aerobic sediments exposed to three of the organic pollutants for 2 to 7 days showed recovery of both activities. Anaerobic sediments showed little recovery after 2 days of pre-exposure to the pollutants. We conclude that (i) anaerobic sediments are more sensitive than aerobic sediments to pollutant additions; (ii) [3H]thymidine incorporation is more sensitive to pollutant additions than is [14C]glucose metabolism; and (iii) the toxicity of the pollutants increased in the following order: anthracene, p,p'-dichlorodiphenyltrichloroethane, naphthalene, and pentachlorophenol.  相似文献   

18.
Placental aldose reductase (EC 1.1.1.21) was incubated with glucose in the presence of [4A-2H] NADPH prepared in the oxidation of [2-2H] isocitrate by isocitrate dehydrogenase (EC 1.1.1.42) or [4B-2H] NADPH prepared in the oxidation of [1-2H] glucose-6-phosphate dehydrogenase (EC 1.1.1.49). The sorbitol formed from [4A-2H] NADPH contained deuterium and from [4B-2H] NADPH it did not. Therefore, aldose reductase in an A-type enzyme.  相似文献   

19.
A two-phase system is developed here for converting: (1) benzene to phenol and (2) naphthalene to 2-naphthol, using whole cells expressing wild-type toluene 4-monooxygenase (T4MO) and the alpha subunit variant TmoA I100A from Pseudomonas mendocina KR1. Using the T4MO TmoA I100A variant, the solubility of naphthalene was enhanced and the toxicity of the naphthols was prevented by the use of a water/dioctyl phthalate (80:20, vol%) system which yielded 21-fold more 2-naphthol. More than 99% 2-naphthol was extracted to the dioctyl phthalate phase, dihydroxynaphthalene formation was prevented, 92% 2-naphthol was formed, and 12% naphthalene was converted. Similarly, using 50 vol% dioctyl phthalate, an initial concentration of 3.0 g l−1 (39 mM), and wild-type T4MO, a 51±9% conversion of benzene was obtained and phenol was produced at a purity of 97%. Relative to the one-phase system, there was a 12-fold reduction in the formation of the byproduct catechol.  相似文献   

20.
The stereochemical course of the reduction of acetaldehyde to ethanol was investigated by evaluating, with the enzymic system yeast alcohol dehydrogenase/diaphorase and g.c.-m.s., the configuration of [1-2H]ethanol obtained from [1-2H]acetaldehyde with different micro-organisms. Although only S-[1-2H]ethanol was formed, all the micro-organisms showed evidence of the existence of alcohol dehydrogenases with opposite stereospecificity.  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