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 inhibition of the major form of ox kidney aldehyde reductase (AR 1) by sodium barbitone revealed linear mixed kinetics. This behaviour is distinct from the non-linear intercept effect we reported for valproate [Daly and Mantle (1982) Biochem. J. 205, 381]. 4-Carboxybenzaldehyde exhibits partial uncompetitive substrate inhibition. These results are discussed in terms of a model that involves nucleotide-induced isomerization and an additional flux (with some substrates and inhibitors) through an enzyme.nucleotide.substrate/inhibitor ternary complex.  相似文献   

2.
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   

3.
A mutant of Arabidopsis thaliana (L.) Heyn. (a small plant in the crucifer family) that lacks glycine decarboxylase activity owing to a recessive nuclear mutation has been isolated on the basis of a growth requirement for high concentrations of atmospheric CO2. Mitochondria isolated from leaves of the mutant did not exhibit glycine-dependent O2 consumption, did not release 14CO2 from [14C]glycine, and did not catalyse the glycine-bicarbonate exchange reaction that is considered to be the first partial reaction associated with glycine cleavage. Photosynthesis in the mutant was decreased after illumination under atmospheric conditions that promote partitioning of carbon into intermediates of the photorespiratory pathway, but was not impaired under non-photorespiratory conditions. Thus glycine decarboxylase activity is not required for any essential function unrelated to photorespiration. The photosynthetic response of the mutant in photorespiratory conditions is probably caused by an increased rate of glyoxylate oxidation, which results from the sequestering of all readily transferable amino groups in a metabolically inactive glycine pool, and by a depletion of intermediates from the photosynthesis cycle. The rate of release of 14CO2 from exogenously applied [14C]glycollate was 14-fold lower in the mutant than in the wild type, suggesting that glycine decarboxylation is the only significant source of photorespiratory CO2.  相似文献   

4.
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   

5.
The distribution of the two principal isoenzymes of aldehyde reductase (EC 1.1.1.2) has been studied in ox brain. The more active of these, which has been termed the high-Km enzyme, has been shown to be located in the cytosol and the less abundant low-Km form has a similar localization. p-Nitrobenzaldehyde, which has been used as a substrate in previous studies, caused the reduction of NADH in the presence of the mitochondrial fraction, but mixed substrate experiments with 1,3-dinitrobenzene and the effects of pH on the activity indicate that this is due to the presence of a nitro reductase activity which has been recently described (Köchli, Wermuth & von Wartburg (1980) Biochim. Biophys. Acta 616, 133-142] rather than to the low-Km aldehyde reductase activity. Fractionation of the mitochondria indicated this activity to be largely confined to the mitochondrial inner membrane.  相似文献   

6.
Purification and properties of low-Km aldehyde reductase from ox brain   总被引:1,自引:0,他引:1  
A low-Km aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2), which may be identical with aldose reductase (alditol:NADP+ 1-oxidoreductase, EC 1.1.1.21), has been purified from ox brain to homogeneity. It was shown to be a monomer with Mr values of 31 000 and 35 100 being obtained by gel filtration and polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulphate, respectively. The enzyme catalyses the NADPH-dependent reduction of a number of aromatic and sugar aldehydes. The activity of the enzyme with 133 microM NADH was about one-third of that with 120 microM NADPH. Activity with both these coenzymes was optimum at pH 6.2 and was inhibited by increasing the ionic strength with KCl, NaCl or NaNO3. In contrast, the activity was stimulated by sodium phosphate. The activity with NADH as the coenzyme was more sensitive to stimulation by phosphate and to inhibition by increasing ionic strength than that determined with NADPH.  相似文献   

7.
8.
An improved procedure for purifying aldehyde reductase is described. Utilization of Blue Dextran--Sepharose 4B and elimination of hydroxyapatite chromatography greatly improves the yield and ease of purification. Starting with 340 g of kidney tissue (two pig kidneys) approx. 50 mg of purified reductase may be routinely and reproducibly obtained. The purified reductase was used to establish the kinetic reaction mechanism of the enzyme. Initial-velocity analysis and product-inhibition data revealed that pig kidney aldehyde reductase follows an Ordered Bi Bi reaction mechanism in which NADPH binds first before D-glyceraldehyde. The limiting Michaelis constants for D-glyceraldehyde and NADPH were 4.8 +/- 0.7 mM and 9.1 +/- 2.1 micrometer respectively. The mechanism is similar to that of another monomeric oxidoreductase, octopine dehydrogenase, towards which aldehyde reductase exhibits several similarities, but differs from that of other aldehyde reductases. Phenobarbital is a potent inhibitor of aldehyde reductase, inhibiting both substrate and cofactor non-competitively (Ki = 80.4 +/- 10.5 micrometer and 66.9 +/- 1.6 micrometer respectively). Barbiturate inhibition seems to be a common property of NADPH-dependent aldehyde reductases.  相似文献   

9.
The liver, kidney and spleen of the mouse and rat and the kidney and spleen of the ox express a monomeric form of biliverdin reductase (Mr 34,000), which in the case of the ox kidney enzyme exists in two forms (pI 5.4 and 5.2) that are probably charge isomers. The livers of the mouse and rats express, in addition, a protein (Mr 46,000) that cross-reacts with antibodies raised against the ox kidney enzyme and may be related to form 2 described by Frydman, Tomaro, Awruch & Frydman [(1983) Biochim. Biophys. Acta 759, 257-263]. Higher-Mr forms appear to exist in the guinea pig and hamster. The ox kidney enzyme has three thiol groups, of which two are accessible to 5,5'-dithiobis-(2-nitrobenzoate) in the native enzyme. Immunocytochemical analysis reveals that biliverdin reductase is localized in proximal tubules of the inner cortex of the rat kidney. Biliverdin reductase antiserum also stains proximal tubules in human and ox kidney. The staining of podocytes in glomeruli of ox kidney with antiserum to aldose reductase is particularly prominent. The localization of biliverdin reductase in the inner cortical zone of rat kidney is similar to that described for glutathione S-transferase YfYf, and it is suggested that one function of this 'intracellular binding protein' may be to maintain a low free concentration of biliverdin to allow biliverdin reductase to operate efficiently.  相似文献   

10.
A procedure is described that yields an apparently homogeneous preparation of the high-Km aldehyde reductase from rat brain. This procedure is also applicable to the purification of this enzyme from rat liver and ox brain. In the latter case, however, the purified preparation could be resolved into two protein bands, both of which had enzyme activity, by polyacrylamide-gel electrophoresis. Since a sample of the ox brain enzyme from an earlier step in the purification procedure only showed the presence of a single band of activity after electrophoresis, this apparent multiplicity probably results from modification of the enzyme, possibly by oxidation, during the final step of the purification. A number of properties of the rat brain enzyme were determined and these were compared with those of the enzyme from rat liver. The two preparations were similar in their stabilities, behaviour during purification, kinetic properties, electrophoretic mobilities and amino acid compositions. Antibodies to the rat liver enzyme cross-reacted with that from brain and the inhibition of both these preparations by the antiserum was similar, further supporting the view that the enzymes from these two sources were closely similar if not identical.  相似文献   

11.
The reaction mechanism of bovine kidney biliverdin reductase   总被引:1,自引:0,他引:1  
The steady-state kinetics of biliverdin reductase can be studied in detail at pH 9 as under these conditions the Km for biliverdin is high enough to obtain reliable measurements of the initial rate in the absence of any biliverdin binding proteins. The initial rate kinetics and the product-inhibition studies are consistent with an ordered sequential mechanism provided the biliverdin concentration was below 20 microM. Above this concentration significant flux occurs through a substrate inhibition pathway involving an enzyme-NAD(P)-biliverdin complex. Chloride is shown to cause a significant activation of the enzyme under certain conditions and this is shown to result from an inability of biliverdin to bind to an enzyme-NAD-chloride complex.  相似文献   

12.
Aldose reductase and aldehyde reductases have been purified to homogeneity from human kidney and have molecular weights of 32,000 and 40,000 and isoelectric pH 5.8 and 5.3, respectively. Aldose reductase, beside catalyzing the reduction of various aldehydes, reduces aldo-sugars, whereas aldehyde reductase, does not reduce aldo-sugars. Aldose reductase activity is expressed with either NADH or NADPH as cofactor, whereas aldehyde reductase utilizes only NADPH. Both enzymes are inhibited to varying degrees by aldose reductase inhibitors. Antibodies against bovine lens aldose reductase precipitated aldose reductase but not aldehyde reductase. The sequence of addition of the substrates to aldehyde reductase is ordered and to aldose reductase is random, whereas for both the enzymes the release of product is ordered with NADP released last.  相似文献   

13.
The mechanism of action of the flavoprotein D-aspartate oxidase (EC 1.4.3.1) has been investigated by steady-state and stopped flow kinetic studies using D-aspartate and O2 as substrates in 50 mM KPi, 0.3 mM EDTA, pH 7.4, 4 degrees C. Steady-state results indicate that a ternary complex containing enzyme, O2, and substrate (or product) is an obligatory intermediate in catalysis. The kinetic parameters are turnover number = 11.1 s-1, Km(D-Asp) = 2.2 x 10(-3) M, Km(O2) = 1.7 x 10(-4) M. Rapid reaction studies show that 1) the reductive half reaction is essentially irreversible with a maximum rate of reduction of 180 s-1; 2) the free reduced enzyme cannot be the species which is reoxidized during turnover since its reoxidation by oxygen (second order rate constant equal to 5.3 x 10(2) M-1 s-1) is too slow to be of relevance in catalysis; 3) reduced enzyme can bind a ligand rapidly and be reoxidized as a complex at a rate faster than that observed for the free reduced enzyme; 4) the rate of reoxidation of reduced enzyme by oxygen during turnover is dependent on both O2 and D-aspartate concentrations (second order rate constant of reaction between O2 and reduced enzyme-substrate complex equal to 6.2 x 10(4) M-1 s-1); and 5) the rate-limiting step in catalysis occurs after reoxidation of the enzyme and before its reduction in the following turnover. A mechanism involving reduction of enzyme by substrate, dissociation of product from reduced enzyme, binding of a second molecule of substrate to the reduced enzyme, and reoxidation of the reduced enzyme-substrate complex is proposed for the enzyme-catalyzed oxidation of D-aspartate.  相似文献   

14.
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   

15.
16.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

17.
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   

18.
A detailed kinetic investigation was made of the binding mechanism of gamma-glutamylcysteine synthetase purified from rat kidney. The results of initial rate and inhibition studies are consistent with a partially random mechanism in which ATP is the obligatory first substrate and both amino acids bind in a random order to the enzyme-ATP complex. Formation of the enzyme-substrate quaternary complex is necessary prior to release of products. This mechanism is consistent with previous binding studies with the enzyme and while it does not rule out participation of enzyme-bound gamma-glutamyl phosphate as an intermediate in catalysis, such an intermediate cannot be a discrete covalent complex.  相似文献   

19.
Rat kidney was shown to contain two NADPH-linked aldehyde reductases (alcohol:NADP+) oxidoreductase, EC 1.1.1.2) with different substrate affinities. The high-Km aldehyde reductase, which was purified to apparent homogeneity, had a molecular weight of 32 000 as determined by Sephadex G-100 gel filtration, and of 37 000 by polyacrylamide gel electrophoresis in the presence of sodium dodecyl sulfate. The purified enzyme reduced various aliphatic aldehydes of different carbon-chain lengths besides many chemicals containing aldehyde groups. The Km values for n-hexadecanal and n-octadecanal were 8 microM and 4 microM, respectively. Bovine serum albumin (1.8 mM) stimulated the reduction of n-hexadecanal and n-octadecanal, and increased the Vmax values by about 15-fold without changing the Km values. The kidney enzyme was not distinguishable from the brain and liver high-Km aldehyde reductases in mobility on polyacrylamide gel electrophoresis, immunological properties, peptide maps or substrate specificity.  相似文献   

20.
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