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.
Two structurally different forms of bovine lens aldose reductase have been identified. Freshly prepared lens extracts contain an unactivated "b form" (ARb) which is sensitive to inhibition by Sorbinil. Upon incubation of the extracts with oxygen radical generating systems, ARb is converted to a more active "a form" (ARa), which is not inhibited by Sorbinil. ARa and ARb were purified to electrophoretic homogeneity.  相似文献   

2.
Aldose reductase was purified from human skeletal and heart muscle by a rapid and efficient scheme involving Red Sepharose chromatography, chromatofocusing on Pharmacia PBE 94, and hydroxylapatite high pressure liquid chromatography. The scheme afforded homogeneous enzyme, 65% recovery, in 2 days. All muscle samples express aldose reductase but not the closely related aldehyde reductase. Aldose reductase is isolated in one of two forms that are distinguishable by their kinetic patterns with glyceraldehyde as substrate and which are interconvertible by treatment with dithiothreitol. Both forms are capable of catalyzing the reduction of glucose (Km = 68 mM), and both are highly sensitive to inhibition by aldose reductase inhibitors. The reduction of glucose was shown to be nearly stoichiometric with production of sorbitol (92 +/- 2%). Dialysis of aldose reductase in the absence of thiols or NADP converts it into a form that shows markedly different kinetic properties, including very weak catalytic activity toward glucose and insensitivity to aldose reductase inhibitors. This modified form can be converted back into the native form by dithiothreitol. Thiol titration of the two forms of aldose reductase with Ellman's reagent indicated that two thiol groups were lost when the enzyme was dialyzed in the absence of dithiothreitol or NADP.  相似文献   

3.
使用DEAE纤维素柱层析、PBE-94层析聚焦、NADP~+-Sepharose 4B亲合层析及SephadexG-100凝胶过滤分离纯化了人脑醛糖还原酶。在DEAE层析中,用咪唑-HCI缓冲液替代了磷酸缓冲液,改善了分离效果。在聚丙烯酰胺及SDS聚丙烯酰胺凝胶电泳中,纯化的人脑醛糖还原酶均呈一条区带。它的pI为5.6,最适pH为6.5,分子量为36,000,底物特异性和氨基酸组成与其它哺乳动物的醛糖还原酶有相似性。开链式醛糖是醛糖还原酶的真正底物,它在开链式和半缩醛的平衡体系中占比例极小,因而推知醛糖还原酶对此底物有很高的K_(cat)和K_(cat)/K_m值,能有效地将它们还原成相应的醇。  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.
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   

6.
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   

7.
Bovine lens aldose reductase (alditol:NADP+ oxidoreductase, EC 1.1.1.21) undergoes an oxidative modification, greatly stimulated by high ionic strength, upon incubation in the presence of oxygen radical generating systems (Del Corso, A., Camici, M., and Mura, U. (1987) Biochem. Biophys. Res. Commun. 148, 369-375). The enzyme modification is accompanied by a change in stereospecificity toward the two enantiomers of glyceraldehyde. In particular, the Km for L-glyceraldehyde of the native form increased over 150 times after the enzyme modification, with a decrease in the catalytic efficiency of over 200 times. By contrast, for the D-enantiomer the Km increased only 7 times with respect to the native form, with a concomitant decrease in the catalytic efficiency of only approximately 3 times. This dramatic change in stereospecificity may account for the reported apparent cooperative behavior exhibited also by highly purified electrophoretically homogeneous preparations of aldose reductase.  相似文献   

8.
Ye Q  Hyndman D  Li X  Flynn TG  Jia Z 《Proteins》2000,38(1):41-48
Chinese hamster ovary (CHO) reductase is an enzyme belonging to the aldo-keto reductase (AKR) superfamily that is induced by the aldehyde-containing protease inhibitor ALLN (Inoue, Sharma, Schimke, et al., J Biol Chem 1993;268: 5894). It shows 70% sequence identity to human aldose reductase (Hyndman, Takenoshita, Vera, et al., J Biol Chem 1997;272:13286), which is a target for drug design because of its implication in diabetic complications. We have determined the crystal structure of CHO reductase complexed with nicotinamide adenine dinucleotide phosphate (NADP)+ to 2.4 A resolution. Similar to aldose reductase and other AKRs, CHO reductase is an alpha/beta TIM barrel enzyme with cofactor bound in an extended conformation. All key residues involved in cofactor binding are conserved with respect to other AKR members. CHO reductase shows a high degree of sequence identity (91%) with another AKR member, FR-1 (mouse fibroblast growth factor-regulated protein), especially around the variable C-terminal end of the protein and has a similar substrate binding pocket that is larger than that of aldose reductase. However, there are distinct differences that can account for differences in substrate specificity. Trp111, which lies horizontal to the substrate pocket in all other AKR members is perpendicular in CHO reductase and is accompanied by movement of Leu300. This coupled with movement of loops A, B, and C away from the active site region accounts for the ability of CHO reductase to bind larger substrates. The position of Trp219 is significantly altered with respect to aldose reductase and appears to release Cys298 from steric constraints. These studies show that AKRs such as CHO reductase are excellent models for examining the effects of subtle changes in amino acid sequence and alignment on binding and catalysis.  相似文献   

9.
Chromatography of bovine kidney aldose reductase using Matrex Orange A affinity gel results in the separation of the unactivated and activated enzyme forms. The former washes through the column, while the latter is eluted with an NADPH step-gradient. The separated enzyme forms display Vmax and Km glycolaldehyde values, and relative sensitivities to inhibition by the aldose reductase inhibitor AL-1576 (spiro[2,7-difluorofluorene-9,4'-imidazolidine]-2',5'- dione), that are similar to those reported previously for the individual forms. However, because Vmax is 17-fold lower for the unactivated enzyme, the purification of aldose reductase via NADP(H) elution from a dye-ligand affinity matrix can result in the selective purification of only the activated enzyme form. These results have direct implications for the study of potential aldose reductase inhibitors, and may explain why linear double-reciprocal plots are commonly observed for enzyme prepared in this manner, while nonlinear plots are seen in other cases.  相似文献   

10.
Large quantities of recombinant human aldose reductase were produced using Spodoptera frugiperda cells and properties of the enzyme were characterized. Direct purification of the recombinant aldose reductase by affinity column chromatography using Matrex gel orange A yielded a single 36 kDa band, similar in size to the purified human muscle aldose reductase, on a sodium dodecyl sulfate-polyacrylamide gel after silver staining. The isoelectric point of the recombinant enzyme was 5.85 which is identical to the human muscle aldose reductase. Following the treatment with an acylamino-acid releasing enzyme, the blocked NH2-terminal amino acid was identified to be acetylalanine. The successive NH2-terminal sequence and that of the COOH-terminal peptide concurred with the expected translated sequence. Kinetic analyses of the recombinant enzyme activity for various substrates and the cofactor, NADPH, demonstrated a good agreement with the previously reported kinetic data on the purified human aldose reductase. A high concentration of (NH4)2SO4 elicited a significant increase in both Km and Kcat for DL-glyceraldehyde as well as D-glucose. Although IC50 values for most of the aldose reductase inhibitors with recombinant enzyme were found to fall within the comparable range of those obtained with nonhuman mammalian enzymes, the IC50 value for epalrestat was more than 10-fold higher in the recombinant enzyme. These results indicate that the recombinant human aldose reductase expressed in the baculovirus system possesses structurally and enzymatically similar properties as those reported for the native human enzyme and should serve as a superior enzyme preparation to nonhuman mammalian enzymes for the screening of the efficacy and potency of newly developed aldose reductase inhibitors.  相似文献   

11.
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   

12.
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   

13.
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   

14.
The first enzyme of the lysine degradation pathway in maize (Zea mays L.), lysine-ketoglutarate reductase, condenses lysine and [alpha]-ketoglutarate into saccharopine using NADPH as a cofactor, whereas the second, saccharopine dehydrogenase, converts saccharopine to [alpha]-aminoadipic-[delta]-semialdehyde and glutamic acid using NAD+ or NADP+ as a cofactor. The reductase and dehydrogenase activities are optimal at pH 7.0 and 9.0, respectively. Both enzyme activities, co-purified on diethylaminoethyl-cellulose and gel filtration columns, were detected on nondenaturing polyacrylamide gels as single bands with identical electrophoretic mobilities and share tissue specificity for the endosperm. The highly purified preparation containing the reductase and dehydrogenase activities showed a single polypeptide band of 125 kD on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. The native form of the enzyme is a dimer of 260 kD. Limited proteolysis with elastase indicated that lysine-ketoglutarate reductase and saccharopine dehydrogenase from maize endosperm are located in two functionally independent domains of a bifunctional polypeptide.  相似文献   

15.
Aldose reductase (AR) catalyzes the NADPH-dependent reduction of glucose and other sugars to their respective sugar alcohols. The NADP+/NADPH exchange is the rate-limiting step for this enzyme and contributes in varying degrees to the catalytic rates of other aldo-keto reductase superfamily enzymes. The mutation of Arg268 to alanine in human recombinant AR removes one of the ligands of the C2-phosphate of NADP+ and markedly reduces the interaction of the apoenzyme with the nucleotide. The crystal structure of human R268A apo-aldose reductase determined to a resolution of 2.1 A is described. The R268A mutant enzyme has similar kinetic parameters to the wild-type enzyme for aldehyde substrates, yet has greatly reduced affinity for the nucleotide substrate which greatly facilitates its crystallization in the apoenzyme form. The apo-structure shows that a high temperature factor loop (between residues 214 and 226) is displaced by as much as 17 A in a rigid body fashion about Gly213 and Ser226 in the absence of the nucleotide cofactor as compared to the wild-type holoenzyme structure. Several factors act to stabilize the NADPH-holding loop in either the 'open' or 'closed' conformations: (1) the presence and interactions of the nucleotide cofactor, (2) the residues surrounding the Gly213 and Ser226 hinges which form unique hydrogen bonds in the 'open' or 'closed' structure, and (3) the Trp219 "latch" residue which interacts with an arginine residue, Arg293, in the 'open' conformation or with a cysteine residue, Cys298, in the 'closed' conformation. Several mutations in and around the high temperature factor loop are examined to elucidate the role of the loop in the mechanism by which aldose reductase binds and releases its nucleotide substrate.  相似文献   

16.
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   

17.
The primary structure of the aldose xylose reductase from Candida tenuis (CtAR) is shown to be 39% identical to that of human aldose reductase (hAR). The catalytic tetrad of hAR is completely conserved in CtAR (Tyr51, Lys80, Asp46, His113). The amino acid residues involved in binding of NADPH by hAR (D.K. Wilson, et al., Science 257 (1992) 81-84) are 64% identical in CtAR. Like hAR the yeast enzyme is specific for transferring the 4-pro-R hydrogen of the coenzyme. These properties suggest that CtAR is a member of the aldo/keto reductase superfamily. Unlike hAR the enzyme from C. tenuis has a dual coenzyme specificity and shows similar specificity constants for NADPH and NADH. It binds NADP(+) approximately 250 times less tightly than hAR. Typical turnover numbers for aldehyde reduction by CtAR (15-20 s(-1)) are up to 100-fold higher than corresponding values for hAR, probably reflecting an overall faster dissociation of NAD(P)(+) in the reaction catalyzed by the yeast enzyme.  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.
Aldehyde reductase (alcohol:NADP+ oxidoreductase, EC 1.1.1.2), aldose reductase (alditol:NAD(P)+ 1-oxidoreductase, EC 1.1.1.21) and carbonyl reductase (secondary-alcohol:NADP+ oxidoreductase, EC 1.1.1.184) constitute the enzyme family of the aldo-keto reductases, a classification based on similar physicochemical properties and substrate specificities. The present study was undertaken in order to obtain information about the structural relationships between the three enzymes. Treatment of human aldehyde and carbonyl reductase with phenylglyoxal and 2,3-butanedione caused a complete and irreversible loss of enzyme activity, the rate of loss being proportional to the concentration of the dicarbonyl reagents. The inactivation of aldehyde reductase followed pseudo-first-order kinetics, whereas carbonyl reductase showed a more complex behavior, consistent with protein modification cooperativity. NADP+ partially prevented the loss of activity of both enzymes, and an even better protection of aldehyde reductase was afforded by the combination of coenzyme and substrate. Aldose reductase was partially inactivated by phenylglyoxal, but insensitive to 2,3-butanedione. The degree of inactivation with respect to the phenylglyoxal concentration showed saturation behavior. NADP+ partially protected the enzyme at low phenylglyoxal concentrations (0.5 mM), but showed no effect at high concentrations (5 mM). These findings suggest the presence of an essential arginine residue in the substrate-binding domain of aldehyde reductase and the coenzyme-binding site of carbonyl reductase. The effect of phenylglyoxal on aldose reductase may be explained by the modification of a reactive thiol or lysine rather than an arginine residue.  相似文献   

20.
NADP+-dependent dihydrodiol dehydrogenase (trans-1,2-dihydrobenzene-1,2-diol: NADP+ oxidoreductase, EC 1.3.1.20) activity in the cytosol of guinea-pig testis was separated into two major and two minor peaks by Q-Sepharose chromatography; one minor form was immunologically cross-reacted with hepatic aldehyde reductase. The two major enzyme forms were purified to homogeneity. One form, which had the highest amount in the tissue, was a monomeric protein with a molecular weight of 32,000 and isoelectric point of 4.2, showed strict specificity for benzene dihydrodiol and NADP+, and reduced pyridine aldehydes, glyceraldehyde and diacetyl at low rates. Another form, with a molecular weight of 36,000 and isoelectric point of 5.0, oxidized n-butanol, glycerol and sorbitol as well as benzene dihydrodiol in the presence of NADP+ or NAD+, and exhibited much higher reductase activity towards various aldehydes, aldoses and diacetyl. The pI 5.0 form was more sensitive to inhibition by sorbinil and p-chloromercuriphenyl sulfonate than the pI 4.2 form and was activated by sulfate ion. The two enzymes did not catalyze the oxidation of hydroxysteroids and xenobiotic alicyclic alcohols and were immunologically different from hepatic 17 beta-hydroxysteroid-dihydrodiol dehydrogenase. The results indicate that guinea-pig testis contains at least two dihydrodiol dehydrogenases distinct from the hepatic enzymes, one of which, the pI 5.0 enzyme form, may be identical to aldose reductase.  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