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.
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   

2.
The reduction of p-nitroso-N-dimethylaniline, p-nitroso-N-diethylaniline, p-nitrosophenol and p-nitroso-N-phenylaniline with NADPH in the presence of aldehyde reductases 1 and 2 is described. The reactivity of these nitroso substrates is increased by hydrophobic substituents and those promoting OH- elimination from the molecule of the reduced substrate. NN-Dimethylbenzoquinonedi-iminium cation was proved to be the reaction product formed from p-nitroso-N-dimethylaniline. The kinetics of the reduction of p-nitroso-N-dimethylaniline catalysed with aldehyde reductase 1 are rather complex at pH 7, and the preferred-pathway mechanism is probably involved. The reaction sequence approaches the ordered pattern at pH 8.5. It was shown that NADPH in equilibrium NADP+ recyclization proceeds in the presence of NADP+, p-nitroso-N-dimethylaniline, cyclohexanol and aldehyde reductase 1, the alcohol oxidation being the slowest step in this reaction. However, the rate of cyclohexanol oxidation surpasses that of the dissociation of NADPH from the enzyme.  相似文献   

3.
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   

4.
Steady state kinetic analysis at pH 7.0 of the reduction of DL-glyceraldehyde by pig muscle aldose reductase showed that the enzyme follows a sequential ordered mechanism with NADPH binding first. However, the "off constant" for NADP+ in the forward direction was 1 order of magnitude less than the kcat. Analysis of this anomaly by pre-steady state kinetics using stopped-flow fluorescence spectroscopy showed that this could be accounted for by isomerization of the enzyme-NADP+ complex and that the rate of isomerization is the rate-limiting step. The rate constant for this step was of the same order of magnitude as the kcat for the forward reaction. Fluorescence emission spectra of free and NADP(H)-bound enzyme suggested a conformational change upon binding of coenzyme. In the reverse direction (oxidation of glycerol) pre-steady state and steady state kinetic analyses were consistent with the rate-limiting step occurring before isomerization of the enzyme-NADPH complex. We conclude, therefore, that during the kinetic mechanism of the reduction of aldehydes by aldose reductase, a slow (kinetically detectable) conformational change in the enzyme occurs upon coenzyme binding. Since NADPH and NADP+ bind to the enzyme very tightly, this has implications for the targeting and binding of drugs that are aldose reductase inhibitors.  相似文献   

5.
Characterization of aldose reductase and aldehyde reductase from rat testis   总被引:4,自引:0,他引:4  
Aldose reductase (alditol:NAD(P)+ 1-oxidoreductase, EC 1.1.1.21) and aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2) were purified to a homogeneity from rat testis. The molecular weights of aldose reductase and aldehyde reductase were estimated to be 38,000 and 41,000 by SDS-polyacrylamide gel electrophoresis, and the pI values of these enzymes were found to be 5.3 and 6.1 by chromatofocusing, respectively. Aldose reductase had activity for aldo-sugars such as xylose, glucose and galactose, whereas aldehyde reductase was virtually inactive for these aldo-sugars. The Km values of aldose reductase for aldo-sugars were relatively high. When a correction was made for the fraction of aldo-sugar present as the aldehyde form, which is the real substrate of the enzyme, the Km values were much lower. Aldose reductase utilized both NADPH and NADH as coenzyme, whereas aldehyde reductase utilized only NADPH. Aldose reductase was activated significantly by sulfate ion, while aldehyde reductase was little affected. Both enzymes were inhibited strongly by the known aldose reductase inhibitors. However, aldehyde reductase was in general less susceptible to these inhibitors when compared to aldose reductase. Both aldose reductase and aldehyde reductase treated with pyridoxal 5-phosphate have lost the susceptibility to aldose reductase inhibitor, suggesting that in these two enzymes aldose reductase inhibitor interacts with a lysine residue.  相似文献   

6.
Human aldose reductase and aldehyde reductase are members of the aldo-keto reductase superfamily that share three domains of homology and a nonhomologous COOH-terminal region. The two enzymes catalyze the NADPH-dependent reduction of a wide variety of carbonyl compounds. To probe the function of the domains and investigate the basis for substrate specificity, we interchanged cDNA fragments encoding the NH2-terminal domains of aldose and aldehyde reductase. A chimeric enzyme (CH1, 317 residues) was constructed in which the first 71 residues of aldose reductase were replaced with first 73 residues of aldehyde reductase. Catalytic effectiveness (kcat/Km) of CH1 for the reduction of various substrates remained virtually identical to wild-type aldose reductase, changing a maximal 4-fold. Deletion of the 13-residue COOH-terminal end of aldose reductase, yielded a mutant enzyme (AR delta 303-315) with markedly decreased catalytic effectiveness for uncharged substrates ranging from 80- to more than 600-fold (average 300-fold). The KmNADPH of CH1 and AR delta 303-315 were nearly identical to that of the wild-type enzyme indicating that cofactor binding is unaffected. The truncated AR delta 303-315 displayed a NADPH/D isotope effect in kcat and an increased D(kcat/Km) value for DL-glyceraldehyde, suggesting that hydride transfer has become partially rate-limiting for the overall reaction. We conclude that the COOH-terminal domain of aldose reductase is crucial to the proper orientation of substrates in the active site.  相似文献   

7.
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   

8.
The kinetic mechanism of NADP(+)-dependent 3 alpha-hydroxysteroid dehydrogenase and NAD(+)-dependent 3 alpha(17 beta)-hydroxysteroid dehydrogenase, purified from hamster liver cytosol, was studied in both directions. For 3 alpha-hydroxysteroid dehydrogenase, the initial velocity and product inhibition studies indicated that the enzyme reaction sequence is ordered with NADP+ binding to the free enzyme and NADPH being the last product to be released. Inhibition patterns by Cibacron blue and hexestrol, and binding studies of coenzyme and substrate are also consistent with an ordered bi bi mechanism. For 3 alpha(17 beta)-hydroxysteroid dehydrogenase, the steady-state kinetic measurements and substrate binding studies suggest a random binding pattern of the substrates and an ordered release of product; NADH is released last. However, the two enzymes transferred the pro-R-hydrogen atom of NAD(P)H to the carbonyl substrate.  相似文献   

9.
Erythritol biosynthesis is catalyzed by erythrose reductase, which converts erythrose to erythritol. Erythrose reductase, however, has never been characterized in terms of amino acid sequence and kinetics. In this study, NAD(P)H-dependent erythrose reductase was purified to homogeneity from Candida magnoliae KFCC 11023 by ion exchange, gel filtration, affinity chromatography, and preparative electrophoresis. The molecular weights of erythrose reductase determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis and gel filtration chromatography were 38,800 and 79,000, respectively, suggesting that the enzyme is homodimeric. Partial amino acid sequence analysis indicates that the enzyme is closely related to other yeast aldose reductases. C. magnoliae erythrose reductase catalyzes the reduction of various aldehydes. Among aldoses, erythrose was the preferred substrate (K(m) = 7.9 mM; k(cat)/K(m) = 0.73 mM(-1) s(-1)). This enzyme had a dual coenzyme specificity with greater catalytic efficiency with NADH (k(cat)/K(m) = 450 mM(-1) s(-1)) than with NADPH (k(cat)/K(m) = 5.5 mM(-1) s(-1)), unlike previously characterized aldose reductases, and is specific for transferring the 4-pro-R hydrogen of NADH, which is typical of members of the aldo/keto reductase superfamily. Initial velocity and product inhibition studies are consistent with the hypothesis that the reduction proceeds via a sequential ordered mechanism. The enzyme required sulfhydryl compounds for optimal activity and was strongly inhibited by Cu(2+) and quercetin, a strong aldose reductase inhibitor, but was not inhibited by aldehyde reductase inhibitors and did not catalyze the reduction of the substrates for carbonyl reductase. These data indicate that the C. magnoliae erythrose reductase is an NAD(P)H-dependent homodimeric aldose reductase with an unusual dual coenzyme specificity.  相似文献   

10.
Carboxymethylation of bovine lens aldose reductase with 10 mM iodoacetate for 1 h at 25 degrees C led to a more than 4-fold increase in kcat. Carboxymethylation led to a 3- to 5-fold increase in Km NADPH and Km D-glyceraldehyde, whereas Km L-glyceraldehyde increased approx. 30-fold. Activation of the enzyme on carboxymethylation was accompanied by a decrease in the sensitivity of the enzyme to inhibition by 5,5'-dithiobis(2-nitrobenzoic acid) (DTNB), sorbinil (Kii increased from 0.4 to 109 microM) and NADP (Kis increased from 0.01 to 0.03 mM), but not tolrestat. Activation of the enzyme was almost completely prevented by NADPH and to a lesser extent by DL-glyceraldehyde. Carboxymethylation of the enzyme did not result in the generation of several partially oxidized enzyme species, indicating the absence of partially carboxymethylated forms. Primary deuterium isotope effects on the reduced enzyme were consistent with a preferred ordered kinetic reaction scheme, in which hydride transfer is not rate limiting. The hydride transfer step does not seem to be significantly affected by carboxymethylation, nor do changes in the substrate binding steps seem to contribute to the observed rate enhancement. Increase in the turnover number of the enzyme on carboxymethylation appears to be due to facilitation of the isomerization of the E:NADP binary complex. The differential effect of carboxymethylation on sorbinil and tolrestat suggests distinct inhibitor sites on the enzyme, an S-site that binds sorbinil and a T-site that binds tolrestat.  相似文献   

11.
Xylose reductase from the xylose-fermenting yeast Pichia stipitis was purified to electrophoretic and spectral homogeneity via ion-exchange, affinity and high-performance gel chromatography. The enzyme was active with various aldose substrates, such as DL-glyceraldehyde, L-arabinose, D-xylose, D-ribose, D-galactose and D-glucose. Hence the xylose reductase of Pichia stipitis is an aldose reductase (EC 1.1.1.21). Unlike all aldose reductases characterized so far, the enzyme from this yeast was active with both NADPH and NADH as coenzyme. The activity with NADH was approx. 70% of that with NADPH for the various aldose substrates. NADP+ was a potent inhibitor of both the NADPH- and NADH-linked xylose reduction, whereas NAD+ showed strong inhibition only with the NADH-linked reaction. These results are discussed in the context of the possible use of Pichia stipitis and similar yeasts for the anaerobic conversion of xylose into ethanol.  相似文献   

12.
Recent evidence of a pyruvate malate shuttle capable of transporting a large amount of NADPH equivalents out of mitochondria in pancreatic islets suggests that cytosolic NADP(H) plays a role in beta cell metabolism. To obtain clues about these processes the activities of several NADPH-utilizing enzymes were estimated in pancreatic islets. Low levels of pyrroquinolone quinone (PQQ) and low levels of enzyme activity that reduce PQQ were found in islets. Low activities of palmitoyl-CoA and stearoyl-CoA desaturases were also detected. Significant activities of glutathione reductase, aldose reductase (EC.1.1.1.21) and aldehyde reductase (EC.1.1.1.2) were present in islets. Potent inhibitors of aldehyde and aldose reductases inhibited neither glucose-induced insulin release nor glucose metabolism in islets indicating that these reductases are not directly involved in glucose-induced insulin reaction. Over 90% of aldose reductase plus aldehyde reductase enzyme activity was present in the cytosol. Kinetic and chromatographic studies indicated that 60-70% of this activity in cytosol was due to aldehyde reductase and the remainder due to aldose reductase. Aldehyde reductase-like enzyme activity, as well as aldose reductase immunoreactivity, was detected in rat islet plasma membrane fractions purified by a polyethylene glycol-Dextran gradient or by a sucrose gradient. This is interesting in view of the fact that voltage-gated potassium channel beta subunits that contain aldehyde and aldose reductase-like NADPH-binding motifs have been detected in plasma membrane fractions of islets [Receptors and Channels 7: 237-243, 2000] and suggests that NADPH might have a yet unknown function in regulating activity of these potassium channels. Reductases may be present in cytosol to protect the insulin cell from molecules that cause oxidative injury.  相似文献   

13.
The mechanism of action of yeast beta-hydroxy-beta-methylglutaryl-coenzyme A reductase has been investigated through kinetic studies on the oxidation of mevaldate by nicotinamide adeninine dinucleotide phosphate (NADP) in the presence of coenzyme A (CoA) and on the reduction of mevaldate by reduced NADP (NADPH) in the absence of presence of CoA or acetyl-CoA. NADP and mevalonate were also used as product inhibitors of the reduction of mevaldate. In the reduction of mevaldate to mevalonate, coenzyme A and acetyl-CoA decreased the Km for mevaldate 30- and 3-fold, respectively. Both compounds increased the Vmax 1.5-fold. These results suggest that CoA is an allosteric activator for the second reductive step and that it acts by enhancing the binding of mevaldate. The intersecting patterns obtained from initial velocities and the patterns produced by product inhibitions suggest the following features of the mechanism. The binding of substrates and release of products proceeds sequentially in both reductive steps, and is ordered throughout or random with respect to the binding of the beta-hydroxy-beta-methylglutaryl-coenzymeA and the first NADPH. The binding of NADPH enhances the binding of the beta-hydroxy-beta-methylglutaryl portion of the CoA ester and the binding of free mevaldate, whereas the binding of NADP leads to an increased affinity of the enzyme for the hemithioacetal (of mevaldate and CoA) and for mevalonate. Thus, the replacement of NADP by NADPH after the first reductive step promotes the conversion of the hemithioacetal to the free carbonyl form, which is then rapidly reduced. The products, CoA and mevalonic acid, of the second reductive step leave the enzyme before the release of the second NADP. This release of the last product is probably the rate-limiting step for the overall process.  相似文献   

14.
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   

15.
Guanosine monophosphate (GMP) reductase catalyzes the reductive deamination of GMP to inosine monophosphate (IMP). GMP reductase plays an important role in the conversion of nucleoside and nucleotide derivatives of guanine to adenine nucleotides. In addition, as a member of the purine salvage pathway, it also participates in the reutilization of free intracellular bases. Here we present cloning, expression and purification of Escherichia coli guaC-encoded GMP reductase to determine its kinetic mechanism, as well as chemical and thermodynamic features of this reaction. Initial velocity studies and isothermal titration calorimetry demonstrated that GMP reductase follows an ordered bi-bi kinetic mechanism, in which GMP binds first to the enzyme followed by NADPH binding, and NADP(+) dissociates first followed by IMP release. The isothermal titration calorimetry also showed that GMP and IMP binding are thermodynamically favorable processes. The pH-rate profiles showed groups with apparent pK values of 6.6 and 9.6 involved in catalysis, and pK values of 7.1 and 8.6 important to GMP binding, and a pK value of 6.2 important for NADPH binding. Primary deuterium kinetic isotope effects demonstrated that hydride transfer contributes to the rate-limiting step, whereas solvent kinetic isotope effects arise from a single protonic site that plays a modest role in catalysis. Multiple isotope effects suggest that protonation and hydride transfer steps take place in the same transition state, lending support to a concerted mechanism. Pre-steady-state kinetic data suggest that product release does not contribute to the rate-limiting step of the reaction catalyzed by E. coli GMP reductase.  相似文献   

16.
Steady-state kinetic studies including initial velocity, NADPH product inhibition, dead-end inhibition, and combined dead-end and product inhibition measurements with purified rat liver glucose-6-phosphate dehydrogenase indicate a sequential and obligatory addition of substrates in the order of NADP+, glucose-6-P for the catalytic pathway at pH 8.0. Although instability of 6-phosphoglucono-delta-lactone precluded product inhibition experiments which might directly exclude an enzyme-6-phosphoglucono-delta-lactone complex, the absence of an enzyme-glucose-6-P complex suggests that the enzyme-lactone product is unlikely and the release of products is also ordered, with NADPH released last. Consideration of the kinetic constants (Ka = 2.0 muM, Kiq = 13 muM) and cellular concentration of the substrates and products suggests extensive inhibition of the enzyme in vivo and control by the NADPH/NADP+ ratios. Circular dichroism spectra of the enzyme in 20 mM phosphate buffer at pH 7.0 and 25 degrees C indicate 51% helix and 33% pleated sheet structures which is considerably different from results (14% helix) with yeast enzymes.  相似文献   

17.
A reaction mechanism for aldose reductase from lens   总被引:4,自引:0,他引:4  
Sheys and Doughty, (Sheys, G.H. and Doughty, C.C. (1979) Biochim. Biophys. Acta 242, 523-531) suggested a model for Rhodotorula (yeast) aldose reductase (alditol:NADP+ 1-oxidoreductase, EC 1.1.1.21) which offered a unified explanation for changes in reversibility, reaction mechanism, and effects of multivalent anions as well as substrate activation. The present paper extends this model to lens aldose reductase, explaining its similarities to the reverse reaction in Rhodotorula in regard to its reaction mechanism, as well as multivalent anion effects of sulfate, pyrophosphate and NADPH (above 20 micro M) and also substrate activation with glyceraldehyde involving formation of an abortive complex (above 50 micro M). Activation of lens aldose reductase resulted with multivalent anions, due to increased V max and apparent Km values with increasing concentration of multivalent anions. The lens enzyme mechanism is similar to the reverse reaction mechanism for the Rhodotorula enzyme, being partially random in character, based on NADP+ inhibitor studies presented here. The binding of NADPH appears to occur at a basic center containing arginine and possibly histidine. Evidence of the participation of these residues at the active center is based on time-course inactivation protection studies using reagents specific for these residues.  相似文献   

18.
Some properties of aldose reductase isolated from various sources and possible ways of regulation of the enzyme catalytic activity are reviewed. Mammalian aldose reductases are monomeric enzymes with M(r) of 30-40 kDa and a broad substrate specificity towards aldoses. The physiological role of this enzyme consists, apparently, in providing an additional pathway for utilization of glucose and removing toxic compounds carrying an aldehyde group from the cell. Aldose reductase is thought to play a key role in various hyperglycemic states, including diabetic cataract. The kinetics of the aldose reductase reaction is hyperbolic with NADPH and nonhyperbolic with glucose. The rate of the enzyme-catalyzed reaction is determined by the effector binding in the active of inhibitory center of the enzyme. Incubation with substrates leads to the activation of the enzyme which is accompanied by a decrease of the effector binding in the enzyme inhibitory center with a sharp decrease in the sensitivity of the activated enzyme to NADPH concentration changes in the presence of glucose excess. A mechanism underlying the catalytic effect of both native and activated forms of the enzyme is proposed.  相似文献   

19.
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   

20.
Bohren KM  Grimshaw CE 《Biochemistry》2000,39(32):9967-9974
Kinetic and crystallographic studies have demonstrated that negatively charged aldose reductase inhibitors act primarily by binding to the enzyme complexed with oxidized nicotinamide dinucleotide phosphate (E.NADP(+)) to form a ternary dead-end complex that prevents turnover in the steady state. A recent fluorescence study [Nakano and Petrash (1996) Biochemistry 35, 11196-11202], however, has concluded that inhibition by sorbinil, a classic negatively charged aldose reductase inhibitor, results from binding to the enzyme complexed with reduced cofactor (E.NADPH) and not binding to E.NADP(+). To resolve this controversy, we present transient kinetic data which show unequivocally that sorbinil binds to E.NADP(+) to produce a dead-end complex, the so-called sorbinil trap, which prevents steady-state turnover in the presence of a saturating concentration of aldehyde substrate. The reported fluorescence binding results, which we have confirmed independently, are further shown to be fully consistent with the proposed sorbinil trap mechanism. Our conclusions are supported by KINSIM simulations of both pre-steady-state and steady-state reaction time courses in the presence and absence of sorbinil. Thus, while sorbinil binding indeed occurs to both E.NADPH and E.NADP(+), only the latter dead-end complex shows significant inhibition of the steady-state turnover rate. The effect of tight-binding kinetics on the inhibition patterns observed for zopolrestat, another negatively charged inhibitor, is further examined both experimentally and with KINSIM, with the conclusion that all reported aldose reductase inhibition can be rationalized in terms of binding of an alrestatin-like inhibitor at the active site, with no need to postulate a second inhibitor binding site.  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