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 343 毫秒
1.
The substrate specificity of purified rat liver glutathione S-transferases (GSTs) for a series of gamma-glutamyl-modified GSH analogues was investigated. GST isoenzyme 3-3 catalysed the conjugation of 1-chloro-2,4-dinitrobenzene with six out of the nine analogues. alpha-L-Glu-L-Cys-Gly and alpha-D-Glu-L-Cys-Gly showed catalytic efficiencies of 40% and 130% that of GSH respectively. The GSH analogue with an alpha-D-glutamyl moiety appeared to be a highly isoenzyme-3-3-specific co-substrate: kcat./Km with GST isoenzyme 4-4 was only about 5% that with GST isoenzyme 3-3, and no enzymic activity was detectable with GST isoenzymes 1-1 and 2-2. GST isoenzyme 4-4 showed some resemblance to GST 3-3: five out of nine co-substrate analogues were accepted by this second isoenzyme of the Mu multigene family. Isoenzymes 1-1 and 2-2, of the Alpha multigene family, accepted only two alternative co-substrates, which indicates that their GSH-binding site is much more specific.  相似文献   

2.
The spontaneous and glutathione (GSH) transferase catalyzed reactions of GSH and N-acetyl-p-benzoquinonimine (NABQI) have been studied by stopped-flow kinetics. The spontaneous reaction was shown to be first order in NABQI, GSH and inversely proportional to the H+ concentration; e.g., at pH 7.0 and 25 degrees C the second-order rate constant was 3.2 X 10(4) M-1 s-1. Data for the enzymatic reaction gave values for Km of 27, 1.3, 7, and 7 microM and values for kappa cat of 90, 37, 5.1, and 165 s-1 for rat liver GSH transferases 1-1, 2-2, 3-3, and 7-7, respectively. Over a wide range of reactant concentrations and pH, the spontaneous reaction yields three products, namely a GSH conjugate, 3-(glutathion-S-yl)acetaminophen; a reduction product, acetaminophen; and an oxidation product, glutathione disulfide in the proportions 2:1:1. Analysis of products formed after enzymatic reaction showed that both GSH conjugation and the reduction of NABQI to acetaminophen were catalyzed to an extent characteristic of each isoenzyme. With respect to GSH conjugation, GSH transferase isoenzymes were effective in the order 7-7 greater than 2-2 greater than 1-1 greater than 3-3 greater than 4-4, and with respect to NABQI reduction these isoenzymes were effective in the order 1-1 greater than 2-2 greater than 7-7 the position of isoenzymes 3-3 and 4-4 being uncertain. Human GSH transferases delta, mu, and pi behave similarly to the homologous rat enzymes, i.e., toward conjugation in the order pi greater than delta greater than mu and the reduction delta greater than mu greater than pi (for nomenclature see W. B. Jakoby, B. Ketterer, and B. Mannervik, (1984) Biochem. Pharmacol. 33, 2539-2540). Possible mechanisms of the reaction and its effect on the toxicity of NABQI are discussed.  相似文献   

3.
Glutathione transferase (GST) activity towards racemic as well as the resolved enantiomers of 7 beta,8 alpha-dihydroxy-9 alpha,10 alpha-epoxy-7,8,9,10-tetrahydrobenzo[a] pyrene (anti-BPDE) and 1-chloro-2,4-dinitrobenzene (CDNB) was measured in post-microsomal supernatants (PMS) obtained from eight human skin samples. All preparations showed significant activity towards anti-BPDE and an almost exclusive preference for the more tumourigenic (+)-enantiomer. The specific activity towards (+)-anti-BPDE varied about five-fold between different PMS (range 147-781 pmol/min per mg protein) whereas the variation in specific activities towards CDNB was about two-fold (range 30-71 nmol/min per mg protein). The activities obtained with PMS at saturating concentrations of racemic anti-BPDE were about half of the activity towards the (+)-enantiomer indicating that (-)-anti-BPDE competitively inhibits conjugation of the (+)-form. No correlation was evident between the activities towards (+)-anti-BPDE and CDNB implying that different classes of GST isoenzymes participated in the two different reactions. Immunoblot analysis revealed the presence of Class Alpha and Pi isoenzymes whereas Class Mu isoenzymes seemed to be absent in the human skin samples analyzed. Quantitatively, the Class Pi isoenzyme(s) predominated in all skin samples and the amount of enzyme was about 1-3 micrograms GST Pi/mg PMS protein. The almost exclusive conjugation of (+)-anti-BPDE by PMS and previous results with GST Pi enzymes from human placenta suggested that this type of enzymes catalysed the conjugation reaction. The five-fold variation in specific activity towards (+)-anti-BPDE observed among the different PMS may be explained by individual differences in GST Pi content or by the presence of endogenous modifiers of GST activity towards the diol-epoxide.  相似文献   

4.
Hypoxia is known to stimulate reactive oxygen species (ROS) generation. Because reduced glutathione (GSH) is compartmentalized in cytosol and mitochondria, we examined the specific role of mitochondrial GSH (mGSH) in the survival of hepatocytes during hypoxia (5% O2). 5% O2 stimulated ROS in HepG2 cells and cultured rat hepatocytes. Mitochondrial complex I and II inhibitors prevented this effect, whereas inhibition of nitric oxide synthesis with Nomega-nitro-L-arginine methyl ester hydrochloride or the peroxynitrite scavenger uric acid did not. Depletion of GSH stores in both cytosol and mitochondria enhanced the susceptibility of HepG2 cells or primary rat hepatocytes to 5% O2 exposure. However, this sensitization was abrogated by preventing mitochondrial ROS generation by complex I and II inhibition. Moreover, selective mGSH depletion by (R,S)-3-hydroxy-4-pentenoate that spared cytosol GSH levels sensitized rat hepatocytes to hypoxia because of enhanced ROS generation. GSH restoration by GSH ethyl ester or by blocking mitochondrial electron flow at complex I and II rescued (R,S)-3-hydroxy-4-pentenoate-treated hepatocytes to hypoxia-induced cell death. Thus, mGSH controls the survival of hepatocytes during hypoxia through the regulation of mitochondrial generation of oxidative stress.  相似文献   

5.
It has been proposed that the chiral inversion of the 2-arylpropionic acids is due to the stereospecific formation of the (-)-R-profenyl-CoA thioesters which are putative intermediates in the inversion. Accordingly, amino acid conjugation, for which the CoA thioesters are obligate intermediates, should be restricted to those optical forms which give rise to the (-)-R-profenyl-CoA, i.e., the racemates and the (-)-(R)-isomers. We have examined this problem in dogs with respect to 2-phenylpropionic acid(2-PPA). Regardless of the optical configuration of 2-phenylpropionic acid administered, the glycine conjugate was the major urinary metabolite and this was shown to be exclusively the (+)-(S)-enantiomer by chiral HPLC. Both (-)-(R)- and (+)-(S)-2-phenylpropionic acid were present in plasma after the administration of either antipode, and further evidence of the chiral inversion of both enantiomers was provided by the presence of some 25% of the opposite enantiomer in the free 2-phenylpropionic acid and its glucuronide excreted in urine after administration of (-)-(R)- and (+)-(S)-2-phenylpropionic acid. The (+)-(S)-enantiomer underwent chiral inversion to the (-)-(R)-antipode when incubated with dog hepatocytes. These data suggests that both enantiomers of 2-phenylpropionic acid are substrates for canine hepatic acyl CoA ligase(s) and thus undergo chiral inversion, but that the CoA thioester of only (+)-(S)-2-phenylpropionic acid is a substrate for the glycine N-acyl transferase. These studies are presently being extended to the structure and species specificity of the reverse inversion and amino acid conjugation of profen NSAIDs.  相似文献   

6.
A reduction of cellular glutathione (GSH) content was observed when isolated rat hepatocytes were incubated with a stereoisomer of a uricosuric diuretic (S-8666) at a high concentration. Subsequent studies have revealed it was due to conjugation of GSH and S-8666 (-)-enantiomer in the liver cytosol. The (+)-enantiomer strongly inhibited the conjugation reaction, therefore, GSH depletion did not take place when a racemic form of S-8666 was incubated with the liver cells. A possible chemical structure of the GSH-conjugate is tentatively proposed.  相似文献   

7.
During the investigation of chemical properties of the dicyclic system of insect juvenile hormone analogues, biotransformation of 2-(4-methoxybenzyl)-1-cyclohexanone (1) by plant cell cultures was studied. Among other components, the cis-(1S, 2S)- and cis-(1R, 2R)-2-(4-methoxybenzyl)-1-cyclohexanol enantiomers 2a and 2b were found in the reaction mixture. Higher stereoselectivity of the biotransformation was observed for trans-(1S, 2R)-enantiomer 3a formation, which occurred in at least 60% of calculated enantiomeric excess.  相似文献   

8.
Isoelectric focusing of a cytosol fraction from human foetal liver revealed the existence of an acidic and a basic isoenzyme of GSH transferase. The acidic and basic forms of GSH transferase were purified in good yield by use of ion-exchange chromatography on DEAE-cellulose followed by affinity chromatography on S-hexyl-GSH coupled to epoxy-activated Sepharose 6B. The content of the acidic and the basic isoenzymes of GSH transferase together was calculated to constitute 1-2% of the soluble proteins in the hepatic cytoplasm. Physical, catalytic and immunological analyses of the acidic and the basic isoenzymes from foetal liver demonstrated unambiguously that the two forms are different structures with distinct properties. On the other hand, the results show clearly extensive similarities between the foetal acidic transferase and transferase pi from human placenta as well as between the foetal basic form and the basic isoenzymes isolated from adult liver. An exception is that both foetal enzymes seem to be considerably more efficient in catalysing the conjugation of GSH with styrene 7,8-epoxide than the corresponding adult forms of GSH transferase.  相似文献   

9.
The stereoselectivity of the peroxisome proliferation potency of 2-ethylhexanoic acid (2-EHA), a metabolite of the plasticizer di-(2-ethylhexyl) adipate, was investigated in vitro. The enantiomers of 2-EHA were prepared via the semipreparative HPLC resolution of their diastereoisomeric (+)-(R)-1-phenylethylamine derivatives and the subsequent hydrolytic cleavage. Monolayers of hepatocytes were incubated 3 days with solution of (-)-(R), (+)-(S), and (+/-)-2-EHA. The peroxisome proliferation potency was measured by means of determination of the peroxisomal palmitoyl coenzyme A oxidation. The theoretical induction component due to each enantiomer were calculated from the experimental data considering the enantiomeric purities of the acids. The (+)-(S)-enantiomer was found to be the most potent inducer e.g., the eutomer, while the (-)-(R) was the distomer. The eudismic ratio was about 1.6 and the racemic mixture exhibited an intermediary potency. These results, obtained in vitro in conditions avoiding confounding factors such as pharmacokinetics, suggest that the peroxisome proliferation induced by 2-ethylhexanoic acid is a stereoselective phenomenon.  相似文献   

10.
Rat cytosolic glutathione S-transferases catalyzed the conjugation of phenethyl chloride and phenethyl bromide with glutathione. The reaction proceeded with a high degree of stereoselectivity. The glutathione conjugate possessing the (R,S,S)- absolute configuration was formed in major quantities from the racemic substrates. The use of the enantiomers of the phenethyl chloride substrates indicated that the (S)-phenethyl chloride was conjugated in preference to the (R)-enantiomer. The conjugation proceeded with inversion of configuration at the benzylic carbon consistent with an SN2-type mechanism. The stereoselectivity was greater for phenethyl chloride than for phenethyl bromide. Varying the substrate or enzyme concentration had no effect upon the observed stereoselectivity. The diastereomeric glutathione conjugates were separated by high performance liquid chromatography. These findings represent the first demonstration of the substrate stereoselectivity of the glutathione S-transferases.  相似文献   

11.
Glutathione (GSH) conjugation of the chiral compound 2-bromo-3-phenylpropionic acid (BPP) was studied in vitro and in the rat in vivo. GSH conjugation of BPP, catalyzed by a mixture of glutathione-S-transferases (GST's) from rat liver cytosol in vitro, was stereoselective: at a substrate concentration of 250 microM, (R)-BPP was more rapidly conjugated than (S)-BPP (R/S-ratio = 2.6). The blood elimination kinetics of the separate BPP enantiomers and the biliary excretion kinetics of the corresponding GSH conjugates were studied in the rat in vivo after administration of (R)- or (S)-BPP at a dose level of 50 mumol/kg. Elimination of (R)-BPP from blood was faster than that of (S)-BPP: half lives were 9 +/- 2 min for (R)-BPP and 13 +/- 1 min for (S)-BPP. The biliary excretion rate of the GSH conjugate of (R)-BPP declined monoexponentially, while that of the GSH conjugate of (S)-BPP displayed a biphasic profile. Half lives of excretion were 13 +/- 1 for the GSH conjugate of (R)-BPP, and 11 +/- 2 for the GSH conjugate of (S)-BPP (second phase). The first phase in the biliary excretion of the GSH conjugate of (S)-BPP could not be attributed to capacity limitation of biliary transport carriers as higher excretion rates were attained upon administration of higher doses (100 and 200 mumol/kg) of (S)-BPP). The blood elimination profiles of (R)- and (S)-BPP differed greatly from the biliary excretion profiles of the corresponding GSH conjugates. This suggests that the kinetics of BPP conjugate excretion are determined by other processes than hepatic GSH conjugation.  相似文献   

12.
This study examines the ability of P450cam to catalyze the formation of 2-ethylhexanoic acid from 2-ethylhexanol relative to its activity on the natural substrate camphor. As is the case for camphor, the P450cam exhibits stereoselectivity for binding (R)- and (S)-2-ethylhexanol. Kinetic studies indicate (R)-2-ethylhexanoic acid is produced 3.5 times as fast as the (S)-enantiomer. In a racemic mixture of 2-ethylhexanol, P450cam produces 50% more (R)-2-ethylhexanoic acid than (S)-2-ethylhexanoic acid. The reason for stereoselective 2-ethylhexanoic acid production is seen in regioselectivity assays, where (R)-2-ethylhexanoic acid comprises 50% of total products while (S)-2-ethylhexanoic acid comprises only 13%. (R)- and (S)-2-ethylhexanol exhibit similar characteristics with respect to the amount of oxygen and reducing equivalents consumed, however, with (S)-2-ethylhexanol turnover producing more water than the (R)-enantiomer. Crystallographic studies of P450cam with (R)- or (S)-2-ethylhexanoic acid suggest that the (R)-enantiomer binds in a more ordered state. These results indicate that wild-type P450cam displays stereoselectivity toward 2-ethylhexanoic acid synthesis, providing a platform for rational active site design.  相似文献   

13.
K H Tan  D J Meyer  B Coles  B Ketterer 《FEBS letters》1986,207(2):231-233
The thymine hydroperoxide, 5-hydroperoxymethyluracil, is a substrate for Se-dependent glutathione (GSH) peroxidase and the Se-independent GSH peroxidase activity associated with the GSH transferase fraction. These enzymes may contribute to repair mechanisms for damage caused by oxygen radicals. GSH transferases 1-1, 2-2, 3-3, 4-4, 6-6, and 7-7 [(1984) Biochem. Pharmacol. 33, 2539-2540] are shown to differ considerably in their ability to utilize this substrate. For example, high activity is found in GSH transferase 6-6 which is the major isoenzyme in spermatogenic tubules where DNA synthesis is so active and faithful DNA replication so important. The activity of the purified GSH transferase isoenzymes towards 5-hydroperoxymethyluracil is comparable with their activity towards other endogenous substrates related to cellular peroxidation such as linoleate hydroperoxide and 4-hydroxynon-2-enal or biologically important xenobiotic metabolites such as benzo(a)pyrene-7,8-diol-9,10-oxide.  相似文献   

14.
Stereoselectivity in the hydrolysis of racemic ethyl 2-phenylacetate derivatives by cultured cells of noncancerous cell lines from rat liver (BRL, BRL 3A, Clone 9, and ARLJ301–3), spontaneously or oncogene transformed rat liver cell lines (ARLJ301–3TR1, Anr4, Anr9–1, and Anr13–1), and cancer cell lines from rat hepatoma (H4-II-E, McA-RH7777, and MH1C1) and sarcoma (XC) was studied. A strong (R)-enantiomer preference was found in the hydrolysis of ethyl 2-hydroxy- ( 2c ) and 2-methoxy-2-phenylacetate ( 3c ) by the noncancerous and oncogene-transformed cells and an (S)-enantiomer preference for ethyl N-acylphenylalaninates with all the present cell lines. These inclinations were, however, not recognized with ethyl 2-methoxy-2-phenylpropanoate and ethyl N-difluoroacetyl-or N-trifluoroacetylphenylalaninate. Moreover, the R preference for 3c was reversed in the reaction by hepatoma cells. Thus, the stereoselectivity is influenced by both structure of acyl group and species of cell lines. The hepatoma cells were considerably different from the noncancerous or oncogene-transformed cells in stereoselectivity. This fact was consistent with the order of colony formation in soft agar cultures (index of malignancy) and the resemblance in actively stained esterase patterns in gel electrophoresis. The stereoselective hydrolysis leads to cell-specific activation of anticancer prodrugs. This has been confirmed for the first time by the stereoselectivity of Anr4 and H4-II-E cells in the hydrolysis of a chiral mustard ester, bis(2-chloroethyl)aminophenyl 2-methoxy-2-phenylacetate ( 14 ) and by the difference of IC50 values of (R)- and (S)- 14 against the two cell lines. © 1995 Wiley-Liss, Inc.  相似文献   

15.
S I Rao  M W Duffel 《Chirality》1991,3(2):104-111
Aryl sulfotransferase IV catalyzes the 3'-phosphoadenosine-5'-phosphosulfate (PAPS)-dependent formation of sulfuric acid esters of benzylic alcohols. Since the benzylic carbon bearing the hydroxyl group can be asymmetric, the possibility of stereochemical control of substrate specificity of the sulfotransferase was investigated with benzylic alcohols. Benzylic alcohols of known stereochemistry were examined as potential substrates and inhibitors for the homogeneous enzyme purified from rat liver. For 1-phenylethanol, both the (+)-(R)- and (-)-(S)-enantiomers were substrates for the enzyme, and the kcat/Km value for the (-)-(S)-enantiomer was twice that of the (+)-(R)-enantiomer. The enzyme displayed an absolute stereospecificity with ephedrine and pseudoephedrine, and with 2-methyl-1-phenyl-1-propanol; that is, only (-)-(1R,2S)-ephedrine, (-)-(1R,2R)-pseudoephedrine, and (-)-(S)-2-methyl-1-phenyl-1-propanol were substrates for the sulfotransferase. In the case of 1,2,3,4-tetrahydro-1-naphthol, only the (-)-(R)-enantiomer was a substrate for the enzyme. Both (+)-(R)-2-methyl-1-phenyl-1-propanol and (+)-(S)-1,2,3,4-tetrahydro-1-naphthol were competitive inhibitors of the aryl sulfotransferase-catalyzed sulfation of 1-naphthalenemethanol. Thus, the configuration of the benzylic carbon bearing the hydroxyl group determined whether these benzylic alcohols were substrates or inhibitors of the rat hepatic aryl sulfotransferase IV. Furthermore, benzylic alcohols such as (+)-(S)-1,2,3,4-tetrahydro-1-naphthol represent a new class of inhibitors for the aryl sulfotransferase.  相似文献   

16.
A protective Mr28K antigen of Schistosoma mansoni, expressed from its cDNA, has been purified in a single step and shown to possess glutathione (GSH) transferase activity as predicted from sequence homologies with two mammalian GSH transferase multigene families. It is notable for its high 1-chloro-2,4-dinitrobenzene GSH transferase and linoleic acid hydroperoxide GSH peroxidase activities. The major GSH transferase of S. mansoni has been purified and its subunit is identical to this Mr28K antigen by criteria of Mr, immunochemistry, substrate specificity and peptide sequence analysis. In the parasite, the antigen is present in the tegument, protonephridial cells and subtegumental parenchymal cells. No significant immunological cross-reactivity between the S.mansoni and mammalian (human and rat) GSH transferases was observed.  相似文献   

17.
HA-966 (1-hydroxy-3-aminopyrrolidone-2) is an antagonist at the glycine allosteric site of the N-methyl-D-aspartate receptor ionophore complex. Unlike presently known glycine antagonists, HA-966 is chiral. We report stereoselectivity for the (R)-enantiomer at the glycine antagonist site. In [3H]glycine binding, the (R)-enantiomer has an IC50 of 4.1 +/- 0.6 microM. The racemic mixture has an IC50 of 11.2 +/- 0.5 microM, whereas (S)-HA-966 has an IC50 greater than 900 microM. In glycine-stimulated [3H]1-[1-(2- thienyl)cyclohexyl]piperidine binding, the (R)-enantiomer inhibits with an IC50 of 121 +/- 61 microM, whereas the racemic mixture has an IC50 of 216 +/- 113 microM and (S)-HA-966 is inactive. The inhibition by (R)-HA-966 can be prevented by the addition of glycine. (R)-HA-966 and racemic HA-966, but not (S)-HA-966, also prevent N-methyl-D-aspartate cytotoxicity in cortical cultures. The (R)-enantiomer and, less potently, the (S)-enantiomer inhibit N-methyl-D-aspartate-evoked [3H]norepinephrine release from rat hippocampal slices (IC50 values of about 0.3 mM and 1.6 mM, respectively), but only the inhibition by (R)-HA-966 is reversed by added glycine. In glutamate-evoked contractions of the guinea pig ileum, (R)-HA-966 causes a glycine-reversible inhibition (IC50 of about 150 microM), whereas (S)-HA-966 is much less potent (IC50 of greater than 1 mM). These results demonstrate stereoselectivity of the glycine antagonist site of the N-methyl-D-aspartate receptor complex in a variety of tissues and assays. The stereoselectivity also confirms the specificity of N-methyl-D-aspartate receptors in glutamate-evoked contractions of the guinea pig ileum, and supports their similarity to central N-methyl-D-aspartate receptors.  相似文献   

18.
RS-8359, (+/-)-4-(4-cyanoanilino)-5,6-dihydro-7-hydroxy-7H-cyclopenta[d]-pyrimidine is a racemic compound with a selective and reversible monoamine oxidase A (MAO-A) inhibition activity. The substrate and product enantioselectivity with respect to 2-hydroxylation of RS-8359 enantiomers was studied using mouse and rat liver microsomes. In mice, the (S)-enantiomer was transformed to the cis-diol metabolite, whereas the (R)-enantiomer to the trans-diol metabolite. The Vmax/Km value for the formation of the cis-diol metabolite from the (S)-enantiomer was sevenfold greater than that for the formation of the trans-diol metabolite from the (R)-enantiomer. The greater Vmax/Km value for the (S)-enantiomer was due to the tenfold smaller Km value compared to that for the (R)-enantiomer. The results were in fair agreement with the previously reported low plasma concentrations of the (S)-enantiomer and the high recovery of the cis-diol metabolite derived from the (S)-enantiomer in urine after oral administration of RS-8359 to mice. Similarly to mice, in rats the (R)-enantiomer was transformed to the trans-diol metabolite, whereas the (S)-enantiomer yielded the cis-diol and trans-diol metabolites. The Vmax/Km value for the (R)-enantiomer was larger than that for the (S)-enantiomer in rats, indicating that the low plasma concentration of the (S)-enantiomer in rats might be caused by a metabolic reaction other than P450-dependent hydroxylation. CYP3A was shown to be responsible for the trans-diol formation from the (R)-enantiomer.  相似文献   

19.
Glutathione transferase reaches 0.5–0.8 mM concentration in the cell so it works in vivo under the unusual conditions of, [S] ? [E]. As glutathione transferase lowers the pKa of glutathione (GSH) bound to the active site, it increases the cytosolic concentration of deprotonated GSH about five times and speeds its conjugation with toxic compounds that are non-typical substrates of this enzyme. This acceleration becomes more efficient in case of GSH depletion and/or cell acidification. Interestingly, the enzymatic conjugation of GSH to these toxic compounds does not require the assumption of a substrate–enzyme complex; it can be explained by a simple bimolecular collision between enzyme and substrate. Even with typical substrates, the astonishing concentration of glutathione transferase present in hepatocytes, causes an unusual “inverted” kinetics whereby the classical trends of v versus E and v versus S are reversed.  相似文献   

20.
Summary An ethene-utilising bacterium was isolated in pure culture from soil and was tentatively identified as aMicrococcus sp. The organism accumulated epoxyalkanes (0.2–13 mM) from internal, terminal, cyclic and aryl-substituted olefins and exhibited a substrate specificity which was different from that expected on the basis of the chemical reactivity pattern in peracid epoxidations. Epoxyalkanes were hydrolysed at a much slower rate than the epoxidation step which allowed them to accumulate. Ethene-grown cells catalysed the stereospecific formation of R-1,2-epoxypropane (enantiomeric excess: e.e.=96%), R-1,2-epoxybutane (e.e.=94%) andtrans-(2R,3R)-epoxybutane (e.e.=84%). An ethene monooxygenase was implicated in the production of chiral epoxides in cell-free extracts of the bacterium. The (2S,3S)-enantiomer of racemictrans-2,3-epoxybutane was stereoselectively hydrolysed to completion resulting in an enrichment in the (2R,3R)-enantiomer. Further hydrolysis of 1,2-epoxyalkanes (C3-C4), however, occurred via complete destruction of both stereoisomers.  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