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 31 毫秒
1.
Initial-rate measurements were made of the oxidations of pyridine-3-methanol and glycerol by NADP+ and of the reduction of the corresponding aldehydes by NADPH catalysed by pig kidney aldehyde reductase. In addition, a brief survey of the specificity of the enzyme towards aldehyde substrates and its sensitivity to the inhibitors ethacrynic acid, sodium barbitone and warfarin was made. The detailed kinetic work indicates a compulsory mechanism for aldehyde reduction, with NADPH binding before aldehyde. For alcohol oxidation, however, it is necessary to postulate the formation of kinetically significant amounts of binary complexes of the type enzyme-alcohol to explain the results. Thus, for alcohol oxidation random-order addition of substrates may occur. Inhibition studies of the kinetics of aldehyde reduction in the presence of the corresponding alcohol product provide further evidence for the existence of enzyme-alcohol complexes. Finally, detailed kinetic studies were made of the inhibition of pyridine-3-aldehyde reduction by sodium barbitone. The mechanism of the inhibition is discussed.  相似文献   

2.
Kinetic studies with the low-Km aldehyde reductase from ox brain.   总被引:1,自引:1,他引:0       下载免费PDF全文
Initial-rate studies of the low-Km aldehyde reductase-catalysed reduction of pyridine-3-aldehyde by NADPH gave families of parallel double-reciprocal plots, consistent with a double-displacement mechanism being obeyed. Studies on the variation of the initial velocity with the concentration of a mixture of the two substrates were also consistent with a double-displacement mechanism. In contrast, the initial-rate data indicated that a sequential mechanism was followed when NADH was used as the coenzyme. Product-inhibition studies, however, indicated that a compulsory-order mechanism was followed in which NADPH bound before pyridine-3-aldehyde with a ternary complex being formed and the release of pyrid-3-ylcarbinol before NADP+. The apparently parallel double-reciprocal plots obtained in the initial-rate studies with NADPH and pyridine-3-aldehyde were thus attributed to the apparent dissociation constant for the binary complex between the enzyme and coenzyme being finite but very low.  相似文献   

3.
The kinetic mechanism of the major sheep liver aldehyde reductase (ALR1) was studied with three aldehyde substrates: p-nitrobenzaldehyde, pyridine-3-aldehyde and D-glucuronate. In each case the enzyme mechanism was sequential and product-inhibition studies were consistent with an ordered Bi Bi mechanism, with the coenzymes binding to the free enzyme. Binding studies were used to investigate the interactions of substrates, products and inhibitors with the free enzyme. These provided evidence for the binding of D-glucuronate, L-gulonate and valproate, as well as NADP+ and NADPH. The enzyme was inhibited by high concentrations of D-glucuronate in a non-competitive manner, indicating that this substrate was able to bind to the free enzyme and to the E X NADP+ complex at elevated concentrations. Although the enzyme was inhibited by high pyridine-3-aldehyde concentrations, there was no evidence for the binding of this substrate to the free enzyme. Sheep liver ALR1 was inhibited by the ionized forms of alrestatin, sorbinil, valproate, 2-ethylhexanoate and phenobarbitone, indicating the presence of an anion-binding site similar to that described for the pig liver enzyme, which interacts with inhibitors and substrates containing a carboxy group. Sorbinil, valproate and 2-ethylhexanoate inhibited the enzyme uncompetitively at low concentrations and non-competitively at high concentrations, whereas phenobarbitone and alrestatin were non-competitive and uncompetitive inhibitors respectively. The significance of these results with respect to inhibitor and substrate binding is discussed.  相似文献   

4.
Aldehyde reductases (alcohol: NADP+-oxidoreductases, EC 1.1.1.2) I and II from human placenta have been purified to homogeneity. Aldehyde reductase I, molecular weight about 74 000, is a dimer of two nonidentical subunits of molecular weigths of about 32 500 and 39 000, whereas aldehyde erductase II is a monomer of about 32 500. Aldehyde reductase I can be dissociated into subunits under high ionic concentrations. The isoelectric pH for aldehyde reductases I and II are 5.76 and 5.20, respectively. Amino acid compositions of the two enzymes are significantly different. Placenta aldehyde reductase I can utilize glucose with a lower affinity, whereas aldehyde reductase II is not capable to reducing aldo-sugars. Similarly, aldehyde reductase I does not catalyse the reduction of glucuronate while aldehyde reductase II has a high affinity for glucuronate. Both enzymes, however, exhibit strong affinity towards various other aldehydes such as glyceraldehyde, propionaldehyde, and pyridine-3-aldehyde. The pH optima for aldehyde reductases I and II are 6.0 and 7.0, respectively. Aldehyde reductaase I can use both NADH and NADPH as cofactors, whereas aldehyde reductase II activity is dependent on NADPH only. Both enzymes are susceptible to inhibition by sulfhydryl group reagents, aldose reductase inhibitors, lithium sulfate, and sodium chloride to varying degrees.  相似文献   

5.
The distribution of microbial aldo-keto reductases was examined and their immunochemical characterization was performed. p-Nitrobenzaldehyde, pyridine-3-aldehyde and ethyl 4-chloro-3-oxobutanoate reductase activities were found to be widely distributed in a variety of microorganisms. In immunodiffusion studies, most yeasts belonging to the genera Sporobolomyces, Sporidiobolus and Rhodotorula formed precipitin bands with anti-Sporobolomyces salmonicolor aldehyde reductase serum. Furthermore, the results of immunotitration experiments suggested that Sporobolomyces salmonicolor AKU 4429 contains other enzyme(s) which can reduce p-nitrobenzaldehyde, pyridine-3-aldehyde and/or ethyl 4-chloro-3-oxobutanoate, and which are inactivated by anti-Sporobolomyces salmonicolor aldehyde reductase serum.  相似文献   

6.
Aldose reductase (ALR2) has been purified to homogeneity from human psoas muscle. From sodium dodecyl sulfate-polyacrylamide electrophoresis the enzyme is monomeric and has a molecular weight of 37,000. ALR2 catalyzes the primarily NADPH-dependent reduction of a wide variety of aldehydes, although the enzyme can also utilize NADH. The best substrates for ALR2 are aromatic aldehydes (e.g. pyridine-3-aldehyde; Km = 9 microM; kcat/Km = 150,000 s-1 M-1), while among aldoses DL-glyceraldehyde is the preferred substrate (Km = 72 microM; kcat/Km = 17,250). Low (100 microM) concentrations of CaCl2 and CaSO4 cause a marked inhibition (90%) of ALR2 as do higher concentrations (0.2 M) of MgCl2. (NH4)2SO4 caused a 2-fold activation of ALR2. The enzyme is also inhibited by quercetin and the commercially developed aldose reductase inhibitors alrestatin and sorbinil. ALR2 is inhibited only very slightly by sodium valproate and barbiturates. ALR2 cross-reacts immunologically with human brain and human placental aldose reductase and with ALR2 from monkey tissue. There is no precipitin cross-reaction of ALR2 with aldose reductases from other species nor with human aldehyde reductase 1 (ALR1) or with ALR1 from other species. The data show that human muscle is a new and relatively rich source of a monomeric NADPH/NADH reductase which is clearly identifiable as aldose reductase.  相似文献   

7.
Abstract NADPH-dependent aldehyde reductase (EC 1.1.1.2) was purified 23-fold with an overall yield of 11% from Sporobolomyces salmonicolor AKU 4429, in 4 steps and, by adding ammonium sulfate, the enzyme was crystallized. The enzyme has a strict requirement for NADPH and irrversibly reduces a number of aldehydes, such as p -nitrobenzaldehyde, pyridine-3-aldehyde and d -glyceraldehyde. Furthermore, it was found that the enzyme catalyses stereospecific reduction of 4-halo-3-oxobutanoate esters to the corresponding ( R )-4-halo-3-hydroxybutanoate esters, which are promising chiral compounds for the chemical synthesis of l -carnitine.  相似文献   

8.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid–base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR–Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

9.
In order to understand more fully the structural features of aldo-keto reductases (AKRs) that determine their substrate specificities it would be desirable to obtain crystal structures of an AKR with a substrate at the active site. Unfortunately the reaction mechanism does not allow a binary complex between enzyme and substrate and to date ternary complexes of enzyme, NADP(H) and substrate or product have not been achieved. Previous crystal structures, in conjunction with numerous kinetic and theoretical analyses, have led to the general acceptance of the active site tyrosine as the general acid-base catalytic residue in the enzyme. This view is supported by the generation of an enzymatically inactive site-directed mutant (tyrosine-48 to phenylalanine) in human aldose reductase [AKR1B1]. However, crystallization of this mutant was unsuccessful. We have attempted to generate a trapped cofactor/substrate complex in pig aldehyde reductase [AKR1A2] using a tyrosine 50 to phenylalanine site-directed mutant. We have been successful in the generation of the first high resolution binary AKR-Y50F:NADP(H) crystal structure, but we were unable to generate any ternary complexes. The binary complex was refined to 2.2A and shows a clear lack of density due to the missing hydroxyl group. Other residues in the active site are not significantly perturbed when compared to other available reductase structures. The mutant binds cofactor (both oxidized and reduced) more tightly but shows a complete lack of binding of the aldehyde reductase inhibitor barbitone as determined by fluorescence titrations. Attempts at substrate addition to the active site, either by cocrystallization or by soaking, were all unsuccessful using pyridine-3-aldehyde, 4-carboxybenzaldehyde, succinic semialdehyde, methylglyoxal, and other substrates. The lack of ternary complex formation, combined with the significant differences in the binding of barbitone provides some experimental proof of the proposal that the hydroxyl group on the active site tyrosine is essential for substrate binding in addition to its major role in catalysis. We propose that the initial event in catalysis is the binding of the oxygen moiety of the carbonyl-group of the substrate through hydrogen bonding to the tyrosine hydroxyl group.  相似文献   

10.
Aldose reductase (EC 1.1.1.21) and aldehyde reductase II (L-hexonate dehydrogenase, EC 1.1.1.2) have been purified to homogeneity from human erythrocytes by using ion-exchange chromatography, chromatofocusing, affinity chromatography, and Sephadex gel filtration. Both enzymes are monomeric, Mr 32,500, by the criteria of the Sephadex gel filtration and polyacrylamide slab gel electrophoresis under denaturing conditions. The isoelectric pH's for aldose reductase and aldehyde reductase II were determined to be 5.47 and 5.06, respectively. Substrate specificity studies showed that aldose reductase, besides catalyzing the reduction of various aldehydes such as propionaldehyde, pyridine-3-aldehyde and glyceraldehyde, utilizes aldo-sugars such as glucose and galactose. Aldehyde reductase II, however, did not use aldo-sugars as substrate. Aldose reductase activity is expressed with either NADH or NADPH as cofactors, whereas aldehyde reductase II can utilize only NADPH. The pH optima for aldose reductase and aldehyde reductase II are 6.2 and 7.0, respectively. Both enzymes are susceptible to the inhibition by p-hydroxymercuribenzoate and N-ethylmaleimide. They are also inhibited to varying degrees by aldose reductase inhibitors such as sorbinil, alrestatin, quercetrin, tetramethylene glutaric acid, and sodium phenobarbital. The presence of 0.4 M lithium sulfate in the assay mixture is essential for the full expression of aldose reductase activity whereas it completely inhibits aldehyde reductase II. Amino acid compositions and immunological studies further show that erythrocyte aldose reductase is similar to human and bovine lens aldose reductase, and that aldehyde reductase II is similar to human liver and brain aldehyde reductase II.  相似文献   

11.
The neuromodulator gamma-hydroxybutyrate is synthesized in vivo from gamma-aminobutyrate by transamination to succinic semialdehyde and subsequent reduction of the aldehyde group. In human brain, succinic semialdehyde reductase is thought to be responsible for the conversion of succinic semialdehyde to gamma-hydroxybutyrate. In the present work, we cloned the cDNA coding for succinic semialdehyde reductase and expressed it in Escherichia coli. A data bank search indicated that the enzyme is identical with aflatoxin B1-aldehyde reductase, an enzyme implicated in the detoxification of xenobiotic carbonyl compounds. Structurally, succinic semialdehyde reductase thus belongs to the aldo-keto reductase superfamily. The recombinant protein was indistinguishable from native human brain succinic semialdehyde reductase by SDS/PAGE. In addition to succinic semialdehyde, it readily catalyzed the reduction 9,10-phenanthrene quinone, phenylglyoxal and 4-nitrobenzaldehyde, typical substrates of aflatoxin B1 aldehyde reductase. The results suggest multiple functions of succinic semialdehyde reductase/aflatoxin B1 aldehyde reductase in the biosynthesis of gamma-hydroxybutyrate and the detoxification of xenobiotic carbonyl compounds, respectively.  相似文献   

12.
13.
The pH dependence of steady-state parameters for aldehyde reduction and alcohol oxidation were determined in the human liver aldehyde reductase reaction. The maximum velocity of aldehyde reduction with NADPH or 3-acetyl pyridine adenine dinucleotide phosphate (3-APADPH) was pH independent at low pH but decreased at high pH with a pK of 8.9-9.6. The V/K for both nucleotides decreased below a pK of 5.7-6.2, as did the pKi of competitive inhibitors NADP and ATP-ribose, suggesting that the 2'-phosphate of the nucleotide has to be deprotonated for binding to the enzyme. The pK of the 2'-phosphate of NADPH appears to be perturbed in the ternary complexes to 5.2-5.4. The V/K for NADPH, the V/K for 3-APADPH, and the pKi of ATP-ribose also decreased above a pK of 9-10, suggesting interaction of the 2'-phosphate of the nucleotide with a protonated base, perhaps lysine. Since protonation of a residue with a pK of 8 (evident in V/K for DL-glyceraldehyde and V/K for L-gulonate versus pH profiles) appears to be essential for aldehyde reduction, and deprotonation for alcohol oxidation, this residue appears to act as a general acid-base catalyst. An additional anion binding site with a pK of 9.94 facilitates the binding of carboxylic substrates such as D-glucuronate. With NADPH as the coenzyme the primary deuterium isotope effects on V and V/K for NADPH were close to unity and pH independent, suggesting that the hydride transfer step is not rate determining over the experimental pH range. With 3-APADPH as the coenzyme, the maximum velocity, relative to NADPH was three- to four-fold lower. Isotope effects on V, V/K for 3-APADPH, and V/K for D-glucuronate were pH independent and equal to 2.2-2.8, indicating that the chemical step of the reaction is relatively insensitive to pH. These data suggest that substrates bind to both the protonated and the deprotonated forms of the enzyme, though only the protonated enzyme catalyzes aldehyde reduction and the deprotonated enzyme catalyzes alcohol oxidation. On the basis of these results a scheme for the chemical mechanism of aldehyde reductase is postulated.  相似文献   

14.
Bovine kidney aldose reductase (ALR2) displays substrate inhibition by aldehyde substrates that is uncompetitive versus NADPH when allowance is made for nonenzymic reaction of the aldehyde with the adenine moiety of NADPH. A time-dependent increase in substrate inhibition observed in product versus time plots for reduction of short-chain aldoses containing an enolizable alpha-proton, but not for p-nitrobenzaldehyde, is shown to be consistent with a model in which rapidly reversible inhibition due to formation of the dead-end E-NADP-glycolaldehyde complex is combined with the formation at the enzyme active site of a tightly-bound covalent NADP-glycolaldehyde adduct. Quantitative analysis of reaction time courses for ALR2-catalyzed reduction of glycolaldehyde using NADPH or the 3-acetylpyridine analogue, (AP)ADPH, yields values of the forward and reverse rate constants for ALR2-mediated adduct formation that agree with the values determined in the absence of glycolaldehyde turnover. Substrate inhibition is only partial, indicating that reaction can occur via an alternate pathway at high [glycolaldehyde]. Kinetic evidence for a slow isomerization of the E-NADP complex at pH 8.0 is used to explain the similar V/Et values observed for glycolaldehyde reduction at pH 7.0 using NADPH, (AP)ADPH, and the hypoxanthine analogue N(Hx)DPH. The practical implications of these results for kinetics studies of aldose reductase are discussed.  相似文献   

15.
16.
Aldose reductase is an NADPH-dependent enzyme which catalyzes the reduction of glucose to sorbitol. Specific potent inhibitors of aldose reductase are of potential pharmacological use because elevated levels of sorbitol produced by this enzyme in lens, peripheral nerve, retina, and renal glomeruli may be responsible for the pathogenesis associated with chronic diabetes. These inhibitors could also serve as probes of the mechanism of action of aldose reductase. anti-Oximes of aromatic aldehydes (e.g., benzaldoxime and 4-fluorobenzaldoxime) have proved to be effective inhibitors of aldose reductase rivaling pharmacological agents currently used to inhibit this enzyme in vivo. The kinetic patterns of inhibition in which benzyl alcohol is used as the oxidizable substrate suggest that the inhibition is due to the formation of a stable ternary complex composed of aldose reductase, NADP+, and the anti-oxime. Analogus ternary complexes are formed at the active site of horse liver alcohol dehydrogenase which is also inhibited by anti-oximes of efficient substrates.  相似文献   

17.
Pig kidney aldehyde reductase is inactivated by 2,3-butanedione, phenylglyoxal, methylglyoxal, and 1,2-cyclohexanedione. 2,3-Butanedione caused the most rapid loss in enzyme activity, the rate of loss being proportional to the concentration of 2,3-butanedione. Neither D-glyceraldehyde nor pyridine 3-aldehyde, both substrates for this broadly specific enzyme, protected the enzyme from inactivation but 1 mM NADPH or NADP completely prevented the loss of activity by 2,3-butanedione suggesting the involvement of arginine in the binding of cofactor. Nicotinamide mononucleotide (NMN) (reduced form) offered no protection to inactivation whereas ADP-ribose phosphate gave complete protection indicating that it is the latter portion of NADPH which interacts with the essential arginine. Both NMN and ADP-ribose phosphate are competitive inhibitors of aldehyde reductase with respect to NADPH. Butanedione-modified aldehyde reductase could still bind to a blue dextran-Sepharose 4B column suggesting that the modified arginine did not bind NADPH. This was confirmed by fluorescence spectra which showed that chemically modified aldehyde reductase caused the same blue shift of NADPH fluorescence as did native aldehyde reductase. Of additional interest was the quenching of NADPH fluorescence by aldehyde reductase which, with one exception, is in contrast to the fluorescence behavior of all other oxidoreductases.  相似文献   

18.
The kinetic mechanism of NADPH-dependent aldehyde reductase II and aldose reductase, purified from human placenta, has been studied using L-glucuronate and DL-glyceraldehyde as their respective substrates. For aldehyde reductase II, the initial velocity and product inhibition studies (using NADP and gulonate) indicate that the enzyme reaction sequence is ordered with NADPH binding to the free enzyme and NADP being the last product to be released. Inhibition patterns using menadione (an analog of the aldehydic substrate) and ATP-ribose (an analog of NADPH) are also consistent with a compulsory ordered reaction sequence. Isotope effects of deuterium-substituted NADPH (NADPD) also corroborate the above reaction scheme and indicate that hydride transfer is not the sole rate-limiting step in the reaction sequence. For aldose reductase, initial velocity patterns, product, and dead-end inhibition studies indicate a random binding pattern of the substrates and an ordered release of product; the coenzyme is released last. A steady-state random mechanism is also consistent with deuterium isotope effects of NADPD on the reaction sequence catalyzed by this enzyme. However, the hydride transfer step seems to be more rate determining for aldose reductase than for aldehyde reductase II.  相似文献   

19.
The effects of ligands of active and inhibitory centers of homogeneous aldose reductase from cattle eye lens on glucose reduction were studied. Using spectrophotometric titration and equilibrium gel filtration, the interaction of the enzyme active center with substrates was investigated. It was shown that the reaction kinetics obeys a mechanism with a quasi-equilibrium non-ordered attachment of substrates and isomerization of enzyme complexes with nicotinamide dinucleotide phosphates in the course of the catalytic act. It was found that the NADPH in equilibrium NADP equilibrium in the enzyme active center is shifted to the right; however, NADP dissociation may occur only as a result of the aldehyde reduction. The mechanisms of regulation of the enzyme activity by NADP, ADP and alpha-glycerol phosphate were proposed. It was shown that the binding of catalin and morine to the enzyme results in the inhibition of the enzymatic reaction and in the isomerization blocking. It was found that the inhibitory site of the isomeric form of aldose reductase displays a lower affinity for morine.  相似文献   

20.
1. Kinetic studies of the reductive amination of 2-oxoglutarate catalysed by glutamate dehydrogenase with NADH and NADPH as coenzyme were made at pH7.0 and pH 8.0. The concentrations of both substrates and coenzymes were simultaneously varied over wide ranges. Lineweaver-Burk plots with respect to each substrate and coenzyme were linear, except that with high concentrations of 2-oxoglutarate or coenzyme inhibition occurred. There was no evidence of the negative homotropic interactions between the enzyme subunits that were revealed in previous kinetic studies of the reverse reaction. 2. The initial-rate results are shown to be inconsistent with any of the six possible compulsory-order mechanisms for this three-substrate reaction, and it is concluded that a random-order mechanism is the most likely one. On the basis of this mechanism, the dissociation constants of all the binary, ternary and quaternary complexes of the enzyme and substrates are calculated from initial-rate parameters. 3. The results are discussed in relation to those of earlier workers who concluded that the mechanism is of the compulsory-order type.  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