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 234 毫秒
1.
Characterization of pulmonary carbonyl reductase of mouse and guinea pig   总被引:2,自引:0,他引:2  
Carbonyl reductases were purified from mouse and guinea pig lung. The mouse enzyme exhibited structural and catalytic similarity to the guinea pig enzyme: tetrameric structure consisting of an identical 23 kDa subunit; basicity (pI of 8.8); low substrate specificity for aliphatic and aromatic carbonyl compounds; dual cofactor specificity for NADPH and NADH; stereospecific transfer of the 4-pro S hydrogen of NADPH; and sensitivity to pyrazole, 2-mercaptoethanol and ferrous ion. Although 3-ketosteroids were extensively reduced by the mouse enzyme but not by the guinea pig enzyme in the forward reaction, the two enzymes similarly oxidized some alicyclic alcohols such as acenaphthenol, cyclohex-2-en-1-ol and benzenedihydrodiol in the presence of NADP+ and NAD+. A partial similarity between the two enzymes was observed immunologically, using antibodies against the purified guinea pig enzyme. The lung enzymes differ in several aspects from other oxidoreductases from extrapulmonary tissues. The immunoreactive protein was detected only in lung of the tissues of the two species.  相似文献   

2.
A carbonyl reductase has been extracted into 0.5 M KCl from dog liver and purified to apparent homogeneity by a three-step procedure consisting of chromatography on CM-Sephadex, Matrex green A, and Sephadex G-100 in high-ionic-strength buffers. The enzyme is a dimer composed of two identical subunits of molecular weight 27,000. The pH optimum is 5.5 and the isoelectric point of the enzyme is 9.3. The enzyme reduces aromatic ketones and aldehydes; the aromatic ketones with adjacent medium alkyl chains are the best substrates. Quinones, ketosteroids, prostaglandins, and aliphatic carbonyl compounds are poor or inactive substrates for the enzyme. As a cofactor the enzyme utilizes NADPH, the pro-S hydrogen atom of which is transferred to the substrate. Two moles of NADPH bind to one mole of the enzyme molecule, causing a blue shift and enhancement of the cofactor fluorescence. The reductase reaction is reversible and the equilibrium constant determined at pH 7.0 is 12.8. Steady-state kinetic measurements in both directions suggest that the reaction proceeds through a di-iso ordered bi-bi mechanism.  相似文献   

3.
The NADPH-linked reductase activity of pig lung carbonyl reductase was activated two- to fivefold by fatty acids with a carbon chain length greater than nine at pH 7.0. cis-Unsaturated fatty acids of C:18 and C:20 were potent activators, showing Ka values of 2-14 microM which were lower than the values of 21-125 microM for saturated fatty acids (C:9 to C:16). Of the fatty acids arachidonic acid (C20:4) gave the highest activation. No significant stimulatory effect was observed with acyl CoAs, fatty alcohols, phospholipids, and nonionic detergents. Anionic detergents (sodium dodecyl sulfate and sarkosyl) stimulated the enzyme activity more than ninefold, but the Ka values for them were much higher than those for the cis-unsaturated fatty acids. Although no change in molecular weight or in subunit composition was observed in the enzyme activated by C20:4, the activation led to a decrease in thermal stability of the enzyme. The binding of C20:4 to the enzyme was instantaneous and reversible, shifted the pH optimum of the activity from 5.8 to 6.5, and changed the inhibitor sensitivity. In addition, C20:4 acted as an allosteric effector abolishing the negative interaction of the enzyme with carbonyl substrates which was seen without the fatty acid, but the activation increased both Vmax and [S]0.5 values for the substrates. Kinetic analysis with respect to NADPH concentration, in which no cooperativity was detected with or without C20:4, indicated that C20:4 was a nonessential activator of mixed type showing a binding constant of 10 microM. These results suggest that cis-unsaturated fatty acids may be potential modulators of pulmonary carbonyl reductase.  相似文献   

4.
Dimeric and monomeric proteins containing dihydrodiol dehydrogenase and aldehyde reductase activities were purified from pig lens. The dimeric enzyme of Mr 65,000 specifically oxidized the trans-dihydrodiols of naphthalene and benzene with NADP+ as a strict cofactor, and reduced alpha-diketones, aromatic aldehydes and glyceraldehyde with NADPH as a cofactor. The monomeric enzyme of Mr 35,000, although identical with aldose reductase, oxidized the trans-dihydrodiol of naphthalene at a pH optimum of 7.6. These results suggest that the two enzymes are involved in the pathogenesis of naphthalene cataract.  相似文献   

5.
Individual flowers from inflorescences of Dahlia variabilis (cv Scarlet Star) in young developmental stages contained relatively high activity of (+)-dihydroflavonol (DHF) 4-reductase. The DHF reductase was purified from such flowers to apparent homogeneity by a five-step procedure. This included affinity adsorption on Blue Sepharose and elution of the enzyme with NADP+. By gel filtration and by sodium dodecyl sulfate-polyacrylamide gel electrophoresis it was shown that DHF reductase contains only one polypeptide chain with a Mr of about 41,000. The reductase required NADPH as cofactor and catalyzed transfer of the pro-S hydrogen of NADPH to the substrate. Flavanones and dihydroflavonols (3-hydroxyflavanones) were substrates for DHF reductase with pH optima of about 6.0 for flavanones and of about 6.8 for dihydroflavonols. Flavanones were reduced to the corresponding flavan-4-ols and (+)-dihydroflavonols to flavan-3,4-cis-diols. Apparent Michaelis constants determined for (2S)-naringenin, (2S)-eriodicytol, (+)-dihydrokaempferol, (+)-dihydroquercetin, and NADPH were, respectively, 2.3, 2, 10, 15, and 42 microM. V/Km values were higher for dihydroflavonols than for flavanones. Conversion of dihydromyricetin to leucodelphinidin was also catalyzed by the enzyme at a low rate, whereas flavones and flavonols were not accepted as substrates. DHF reductase was not inhibited by metal chelators.  相似文献   

6.
Pyridoxine dehydrogenase (1.1.1.65) (pyridoxal reductase), purified to homogeneity from baker's yeast, is a monomer of Mr approximately 33,000. It catalyzes the reversible oxidation of pyridoxine by NADP to yield pyridoxal and NADPH; equilibrium lies far in the direction of pyridoxine formation (Keq approximately 1.4 X 10(11) l/mol at 25 degrees C). Reduction of pyridoxal occurs most rapidly at pH 6.0-7.0; oxidation of pyridoxine is optimal at pH 8.6. NAD and NADH do not replace NADP and NADPH as substrates; pyridoxine, pyridoxal and pyridoxal 5'-phosphate are the only naturally occurring cosubstrates found. Several other aromatic aldehydes also are reduced, but substrate specificity and other properties of the enzyme distinguish it clearly from other alcohol dehydrogenases or aldehyde reductases. Between pH 6.3 and 7.1 (the intracellular pH of yeast), V/Km with pyridoxal and NADPH as substrates is greater than 600 times that observed with pyridoxine and NADPH as substrates is greater than 600 times that observed with pyridoxine and NADP as substrates. These and other considerations strongly indicate that the dehydrogenase functions in vivo to reduce pyridoxal to pyridoxine, which is the preferred substrate for pyridoxal (pyridoxine) kinase in yeast.  相似文献   

7.
Prostaglandin 9-ketoreductase (PG-9-KR) was purified from pig kidney to homogeneity, as judged by SDS/PAGE using an improved procedure. The enzyme is pro-S stereoselective with regard to hydrogen transfer from NADPH with prostaglandin E2 as substrate and reduces its 9-keto group with approximately 90% stereoselectivity to form prostaglandin F2 alpha. Approximately 8% of the prostaglandin F formed has the beta-configuration. In addition to catalyzing the interconversion of prostaglandin E2 to F2 alpha, PG-9-KR also oxidizes prostaglandin E2, F2 alpha and D2 to their corresponding, biologically inactive, 15-keto metabolites. Incubation of PG-9-KR with prostaglandin F2 alpha and NAD+ leads to the preferential formation of 15-keto prostaglandin F2 alpha rather than prostaglandin E2. This suggests that the prostaglandin E2/prostaglandin F2 alpha ratio is not determined by the NADP+/NADPH redox couple. The enzyme also reduces various other carbonyl compounds (e.g. 9,10-phenanthrenequinone) with high efficiency. The catalytic properties measured for PG-9-KR suggest that its in vivo function is unlikely to be to catalyze formation of prostaglandin F2 alpha. The monomeric enzyme has a molecular mass of 32 kDa and exists as four isoforms, as judged by isoelectric focusing. PG-9-KR contains 1.9 mol Zn2+/mol enzyme and no other cofactors. Human kidney PG-9-KR was also purified to homogeneity. The human enzyme has a molecular mass of 34 kDa and also exists as four isoforms. Polyclonal antibodies raised against pig kidney PG-9-KR cross-react with human kidney PG-9-KR and also with human brain carbonyl reductase, as demonstrated by Western blot analysis. Sequence data of tryptic peptides from pig kidney PG-9-KR show greater than 90% identity with human placenta carbonyl reductase. From comparison of several properties (catalytical, structural and immunological properties), it is concluded that PG-9-KR and carbonyl reductase are identical enzymes.  相似文献   

8.
Prostaglandin D2 11-ketoreductase activity of bovine liver was purified 340-fold to apparent homogeneity. The purified enzyme was a monomeric protein with a molecular weight of about 36 kDa, and had a broad substrate specificity for porstaglandins D1, D2, D3, and H2, and various carbonyl compounds (e.g., phenanthrenequinone and nitrobenzaldehyde, etc.). Prostaglandin D2 was reduced to 9 alpha,11 beta-prostaglandin F2 and prostaglandin H2 to prostaglandin F2 alpha with NADPH as a cofactor. Phenanthrenequinone competitively inhibited the reduction of prostaglandin D2, while it did not inhibit that of prostaglandin H2. Moreover, chloride ion stimulated the reduction of prostaglandin D2 and carbonyl compounds, while it had no effect on that of prostaglandin H2. Besides, the enzyme was inhibited by flavonoids (e.g., quercetin) that inhibit carbonyl reductase, but was not inhibited by barbital and sorbinil, which are the inhibitors of aldehyde and aldose reductases, respectively. These results indicate that the bovine liver enzyme has two different active sites, i.e., one for prostaglandin D2 and carbonyl compounds and the other for prostaglandin H2, and appears to be a kind of carbonyl reductase like bovine lung prostaglandin F synthase (Watanabe, K., Yoshida, R., Shimizu, T., and Hayaishi, O., 1985, J. Biol. Chem. 260, 7035-7041). However, the bovine liver enzyme was different from prostaglandin F synthase of bovine lung with regard to the Km value for prostaglandin D2 (10 microM for the liver enzyme and 120 microM for the lung enzyme), the sensitivity to chloride ion (threefold greater activation for the liver enzyme) and the inhibition by CuSO4 and HgCl2 (two orders of magnitude more resistant in the case of the liver enzyme). These results suggest that the bovine liver enzyme is a subtype of bovine lung prostaglandin F synthase.  相似文献   

9.
We studied the localization of carbonyl reductase (E.C. 1.1.1.184) in guinea pig and mouse lung by enzyme histochemistry and immunohistochemistry, using antibodies against the guinea pig lung enzyme which crossreacted with the lung enzymes of both animals. Carbonyl reductase activity was detectable in the bronchiolar epithelial cells of small airways and in alveolar cells. In the immunohistochemical staining for carbonyl reductase, the reaction was strongest in the non-ciliated bronchiolar cells (Clara cells) and was weak in the ciliated cells and type II alveolar pneumocytes. Injection of a single dose of naphthalene led to significant impairment of carbonyl reductase activity and of microsomal mixed-function oxidase activities in mouse lung, with a marked decrease in both activity and immunoreactive staining in the bronchiolar epithelial cells. The results indicate that carbonyl reductase is localized primarily in the Clara cells, which are known to be sites of pulmonary drug metabolism.  相似文献   

10.
Two NADPH-dependent oxidoreductases catalyzing the enantioselective reduction of 3-oxo esters to (S)- and (R)-3-hydroxy acid esters, [hereafter called (S)- and (R)-enzymes] have been purified 121- and 332-fold, respectively, from cell extracts of Saccharomyces cerevisiae by means of streptomycin sulfate treatment, Sephadex G-25 filtration, DEAE-Sepharose CL-6B chromatography, Sephadex G-150 filtration, Sepharose 6B filtration and hydroxyapatite chromatography. The relative molecular mass Mr, of the (S)-enzyme was estimated to be 48,000-50,000 on Sephadex G-150 column chromatography and 48,000 on sodium dodecyl sulfate/polyacrylamide gel electrophoresis. The enzyme was most active at pH 6.9 and reduced 3-oxo esters, 4-oxo and 5-oxo acids and esters enantioselectively to (S)- hydroxy compounds in the presence of NADPH. The Km values for ethyl 3-oxobutyrate, ethyl 3-oxohexanoate, 4-oxopentanoic and 5-oxohexanoic acid were determined as 0.9 mM, 5.3 mM, 17.1 mM and 13.1 mM, respectively. The Mr of the (R)-enzyme, estimated by means of column chromatography on Sepharose 6B, was 800,000. Under dissociating conditions of SDS/polyacrylamide gel electrophoresis the enzyme resolved into subunits of Mr 200,000 and 210,000, respectively. The enzyme is optimally active at pH 6.1, catalyzing specifically the reduction of 3-oxo esters to (R)-hydroxy esters, using NADPH for coenzyme. Km values for ethyl 3-oxobutyrate and ethyl 3-oxohexanoate were determined as 17.0 mM and 2.0 mM, respectively. Investigations with purified fatty acid synthase of baker's yeast revealed that the (R)-enzyme was identical with a subunit of this multifunctional complex; intact fatty acid synthase (Mr 2.4 X 10(6)) showed no activity in catalyzing the reduction of 3-oxo esters.  相似文献   

11.
Using the cDNA of bovine lung prostaglandin F synthase (EC 1.1.1.2) as a probe, we isolated a clone from a bovine liver cDNA library which differed in only eleven nucleotides from the probe. The corresponding protein contained three amino acid substitutions, including a leucine residue which is conserved throughout all aldo-keto reductases. We inserted the liver cDNA into expression vector pUC19 and expressed the recombinant liver enzyme in E.coli. The purified liver enzyme reduced prostaglandin H2 as well as prostaglandin D2 and various carbonyl compounds. The high relative activity against prostaglandin H2 in combination with a high Km value for prostaglandin D2 identified this liver enzyme as a lung type prostaglandin F synthase. However, the binding constant for NADPH of the liver enzyme was 3.5 fold higher than that of lung prostaglandin F synthase.  相似文献   

12.
酵母3-脱氧葡糖醛酮代谢酶的分离纯化及部分性质   总被引:1,自引:0,他引:1  
3-脱氧葡糖醛酮 ( 3- deoxyglucosone)是美拉德反应的主要中间产物 ,对生物体具有毒性作用 .用硫酸铵分部沉淀、DEAE- cellulose52、Hydroxyapatite、DEAE- Sepharose CL- 6B柱层析从酿酒酵母 YBr-M( S.cerevisiae YBr-M)抽提液中分离纯化了 3-脱氧葡糖醛酮代谢酶 (以 NADPH为辅酶 ) .该酶是单一的分子 ,分子量为 44k D,反应最适 p H为 7.0 ,p H6.0~ 8.0之间酶活性相对稳定 ,以 3-脱氧葡糖醛酮为底物的米氏常数 Km 为 2 .2 5mmol/ L.在 35℃以下保温 30 min酶活不变 ,50℃保温 30 min后酶活损失 50 % .该酶对二羰基化合物的活性较高 ,对单羰基化合物则较低 ,其催化作用受碘乙酸、N-乙基顺丁烯二酰亚胺的抑制 ,而被β-巯基乙醇、二硫苏糖醇激活 ,催化作用必须以 NADPH为专一辅酶 ,当用 NADH代替 NADPH时 ,活力只有 5.3% .  相似文献   

13.
During the purification of pig kidney aldehyde reductase by an established procedure [Flynn, Cromlish & Davidson (1982) Methods Enzymol. 89, 501-506] a second enzyme with aldehyde reductase activity may be purified. When the procedure was performed in the presence of 5 mM-EDTA, only traces of the second reductase, pig kidney aldehyde reductase (minor form), were present. By the criterion of sodium dodecyl sulphate/polyacrylamide-gel electrophoresis, pig kidney aldehyde reductase (minor form) had Mr 35 000, in comparison with Mr 40 200 found for pig kidney aldehyde reductase. Amino acid analysis of both enzymes and tryptic-peptide-map comparisons indicated differences in primary structure. The N-terminus of pig kidney aldehyde reductase (minor form) had the sequence Lys-Val-Leu, in contrast with the blocked (acetylated) N-terminus of pig kidney aldehyde reductase. The C-terminal sequence of both enzymes was the same. Both reductases were immunologically identical by double immunodiffusion and rocket immunoelectrophoresis. Pig kidney aldehyde reductase (minor form) had 50% of the specific activity of pig kidney aldehyde reductase when tested with a variety of aldehyde substrates. Michaelis constants of both enzymes for these substrates and for NADPH were similar, but values for kcat. and kcat./Km indicated that catalytically pig kidney aldehyde reductase was the more efficient enzyme. Typical aldehyde reductase inhibitors, such as phenobarbital and sodium valproate, had the same effect on both enzymes. It was concluded that pig kidney aldehyde reductase (minor form) is an enzymically active cleavage product of pig kidney aldehyde reductase which is formed when the latter is purified in the absence of the metalloproteinase inhibitor EDTA.  相似文献   

14.
NADPH-cytochrome c reductase was purified to electrophoretic homogeneity from detergent solubilized sheep lung microsomes. The specific activity of the purified enzyme ranged from 56 to 67 mumol cytochrome c reduced/min/mg protein and the yield was 48-52% of the initial activity in lung microsomes. The reductase had Mr of 78,000 and contained 1 mol each of FAD and FMN. Km values obtained in 0.3 M phosphate buffer, pH 7.8 at 37 degrees C for NADPH and cytochrome c were 11.1 +/- 0.70 microM and 20.0 +/- 2.15 microM. Lung reductase was inhibited by its substrate, cytochrome c when its concentration was above 160 microM. The lung reductase exhibited a ping-pong type kinetic mechanism for NADPH mediated cytochrome c reduction. Purified lung reductase was biocatalytically active in supporting benzo(a)pyrene hydroxylation reaction when coupled with lung cytochrome P-450 and lipid.  相似文献   

15.
Xenobiotic reductase A (XenA) from Pseudomonas putida 86 catalyzes the NAD(P)H-dependent reduction of various α,β-unsaturated carbonyl compounds and is a member of the old yellow enzyme family. The reaction of XenA follows a ping-pong mechanism, implying that its active site has to accommodate and correctly position the various substrates to be oxidized (NADH/NADPH) and to be reduced (different α,β-unsaturated carbonyl compounds) to enable formal hydride transfers between the substrate and the isoalloxazine ring.The active site of XenA is lined by two tyrosine (Tyr27, Tyr183) and two tryptophan (Trp302, Trp358) residues, which were proposed to contribute to substrate binding. We analyzed the individual contributions of the four residues, using site-directed mutagenesis, steady-state and transient kinetics, redox potentiometry and crystal structure analysis. The Y183F substitution decreases the affinity of XenA for NADPH and reduces the rate of the oxidative half-reaction by two to three orders of magnitude, the latter being in agreement with its function as a proton donor in the oxidative half-reaction. Upon reduction of the flavin, Trp302 swings into the active site of XenA (in-conformation) and decreases the extent of the substrate-binding pocket. Its exchange against alanine induces substrate inhibition at elevated NADPH concentrations, indicating that the in-conformation of Trp302 helps to disfavor the nonproductive NADPH binding in the reduced state of XenA.Our analysis shows that while the principal catalytic mechanism of XenA, for example, type of proton donor, is analogous to that of other members of the old yellow enzyme family, its strategy to correctly position and accommodate different substrates is unprecedented.  相似文献   

16.
In atypical reaction, incubation of purified rabbit pulmonary flavin-containing monooxygenase with certain primary alkylamines results in the oxidation of NADPH and the formation of hydrogen peroxide. In addition, significant amounts of N-hydroxylated primary amine are also generated, as determined by colorimetric assay and GC/MS analysis of n-octylamine metabolites. Similar reactions appear to be catalyzed by the mouse pulmonary enzyme. In contrast, incubation of primary alkylamines with hepatic flavin-containing monooxygenases from rabbit, mouse, or pig does not result in NADPH oxidation or metabolism. Another effect of primary alkylamines is marked activation of the mouse pulmonary and pig hepatic flavin-containing monooxygenases with some substrates. The structural requirements for primary alkylamines to elicit NADPH oxidation by the rabbit pulmonary enzyme or to activate the mouse pulmonary and pig hepatic enzymes are identical. This indicates that different flavin-containing monooxygenases probably have a conserved alkylamine-binding site of defined specificity. In the case of the rabbit pulmonary enzyme, this binding may occur very close to or at the catalytic site resulting in some N-hydroxylation of the alkylamine.  相似文献   

17.
A monomeric 3 alpha-hydroxysteroid dehydrogenase with a molecular weight of 34,000 was purified to apparent homogeneity from mouse liver cytosol. The enzyme catalyzed the reversible oxidation of the 3 alpha-hydroxy group of C19-, C21-, and C24-steroids, reduced a variety of carbonyl compounds, and was inhibited by SH-reagents, synthetic estrogens, anti-inflammatory drugs, prostaglandins, and delta 4-3-ketosteroids. Although these properties are similar to those of the enzyme from rat liver cytosol, the mouse enzyme exhibited low dehydrogenase activity toward benzene dihydrodiol and some alicyclic alcohols, it showed a strict cofactor specificity for NADP(H), and high substrate inhibition was observed in the reverse reaction. In addition, dexamethasone, deoxycorticosterone, and medroxyprogesterone acetate inhibited the mouse enzyme competitively at low concentrations and noncompetitively at high concentrations, whereas hexestrol, indomethacin, and prostaglandin A1 were competitive inhibitors. Steady-state kinetic measurements in both directions indicated that the reaction proceeds through an ordered bi bi mechanism with the cofactors binding to the free enzyme. The 3-ketosteroid substrates inhibited the enzyme uncompetitively at elevated concentrations, suggesting that the substrates bind to the enzyme.NADPH complex and to the enzyme NADP+ complex.  相似文献   

18.
Ligninase-I (Mr 42,000-43,000; carbohydrate, 21%) and peroxidase-M2 (Mr 45,000-47,000; carbohydrate, 17%), two representative, hydrogen peroxide-dependent extracellular enzymes produced by ligninolytic cultures of the white-rot fungus Phanerochaete chrysosporium BKM-F-1767, were purified and their properties compared. Spectroscopic studies showed that both native enzymes are heme proteins containing protoporphyrin IX. EPR spectroscopy indicated that iron ions are coordinated with the enzymes' prosthetic groups as high-spin ferriheme complexes. We confirmed reports of others that the ligninase-hydrogen peroxide complex (activated enzyme) reverts to its native state on addition of dithionite or one of the enzyme's substrates (e.g., veratryl alcohol); however, we found that the peroxidase-M2-hydrogen peroxide complex required Mn2+ ions to accomplish a similar cycle. The peroxidase oxidized Mn2+ to a higher oxidation state, and the oxidized Mn acted as a diffusible catalyst able to oxidize numerous organic substrates. Unlike ligninase-I which is found free extracellularly, peroxidase-M2 appears to be associated closely with the fungal mycelium. In its peroxidatic reactions, ligninase-I oxidizes a variety of nonphenolic and phenolic lignin model compounds. In the presence of Mn2+, peroxidase-M2 oxidizes numerous phenolic compounds, especially syringyl (3,5-dimethoxy-4-hydroxyphenyl) and vinyl side-chain substituted substrates. Also, the peroxidase-Mn2+ system (without hydrogen peroxide) expresses oxidase activity against NADPH, GSH, dithiothreitol, and dihydroxymaleic acid, forming hydrogen peroxide at the expense of oxygen. Both enzymes were believed to play roles in lignin degradation, and these are discussed.  相似文献   

19.
In the screening of 11 E. coli strains overexpressing recombinant oxidoreductases from Bacillus sp. ECU0013, an NADPH-dependent aldo-keto reductase (YtbE) was identified with capability of producing chiral alcohols. The protein (YtbE) was overexpressed, purified to homogeneity, and characterized of biocatalytic properties. The purified enzyme exhibited the highest activity at 50°C and optimal pH at 6.5. YtbE served as a versatile reductase showing a broad substrate spectrum towards different aromatic ketones and keto esters. Furthermore, a variety of carbonyl substrates were asymmetrically reduced by the purified enzyme with an additionally coupled NADPH regeneration system. The reduction system exhibited excellent enantioselectivity (>99% ee) in the reduction of all the aromatic ketones and high to moderate enantioselectivity in the reduction of α- and β-keto esters. Among the ketones tested, ethyl 4,4,4-trifluoroacetoacetate was found to be reduced to ethyl (R)-4,4,4-trifluoro-3-hydroxy butanoate, an important pharmaceutical intermediate, in excellent optical purity. To the best of our knowledge, this is the first report of ytbE gene-encoding recombinant aldo-keto reductase from Bacillus sp. used as biocatalyst for stereoselective reduction of carbonyl compounds. This study provides a useful guidance for further application of this enzyme in the asymmetric synthesis of chiral alcohol enantiomers.  相似文献   

20.
The response of the hexose monophosphate shunt (HMS) in organ-cultured guinea pig lens to 1,2-naphthoquinone and 5-hydroxy-1,4-naphthoquinone (juglone) has been investigated. Both these compounds, which are substrates of guinea pig lens zeta-crystallin (NADPH:quinone oxidoreductase), were found to cause increases in the rate of 14CO2 production from 1-14C-labelled glucose. Exposure of lenses to 15 microM 1,2-naphthoquinone or 20 microM juglone yielded 5.9- and 7-fold stimulation of HMS activity, respectively. Unlike hydrogen peroxide-induced stimulation of HMS activity, these effects were not abolished by preincubation with the glutathione reductase inhibitor, 1,3-bis(2-chloroethyl)-1 nitrosourea (BCNU). While hydrogen peroxide produced substantial decrements in lens glutathione (GSH) levels, incubation with quinones was not associated with a similar reduction in GSH concentration. Protein-bound NADPH content in quinone-exposed guinea pig lenses was decreased, with a concomitant increase in the amounts of free NADP+. This finding supported the involvement of zeta-crystallin bound NADPH in the in vivo enzymic reduction of quinones. Hydrogen peroxide, on the other hand, caused decreases in the level of free NADPH alone, serving to confirm our earlier inference that quinone stimulated increases in the guinea pig lens HMS could be mediated through zeta-crystallin NADPH:quinone oxidoreductase activity.  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