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.
Human brain contains multiple forms of aldehyde-reducing enzymes. One major form (AR3), as previously shown, has properties that indicate its identity with NADPH-dependent aldehyde reductase isolated from brain and other organs of various species; i.e., low molecular weight, use of NADPH as the preferred cofactor, and sensitivity to inhibition by barbiturates. A second form of aldehyde reductase ("SSA reductase") specifically reduces succinic semialdehyde (SSA) to produce gamma-hydroxybutyrate. This enzyme form has a higher molecular weight than AR3, and uses NADH as well as NADPH as cofactor. SSA reductase was not inhibited by pyrazole, oxalate, or barbiturates, and the only effective inhibitor found was the flavonoid quercetine. Although AR3 can also reduce SSA, the relative specificity of SSA reductase may enhance its in vivo role. A third form of human brain aldehyde reductase, AR2, appears to be comparable to aldose reductases characterized in several species, on the basis of its activity pattern with various sugar aldehydes and its response to characteristic inhibitors and activators, as well as kinetic parameters. This enzyme is also the most active in reducing the aldehyde derivatives of biogenic amines. These studies suggest that the various forms of human brain aldehyde reductases may have specific physiological functions.  相似文献   

2.
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   

3.
NADH- and NAD(P)H-Nitrate Reductases in Rice Seedlings   总被引:7,自引:4,他引:3       下载免费PDF全文
By use of affinity chromatography on blue dextran-Sepharose, two nitrate reductases from rice (Oryza sativa L.) seedlings, specifically, NADH:nitrate oxidoreductase (EC 1.6.6.1) and NAD(P)-H:nitrate oxidoreductase (EC 1.6.6.2), have been partially separated. Nitrate-induced seedlings contained more NADH-nitrate reductase than NAD(P)H-nitrate reductase, whereas chloramphenicol-induced seedlings contained primarily NAD(P)H-nitrate reductase. NAD(P)H-nitrate reductase was shown to utilize NADPH directly as reductant. This enzyme has a preference for NADPH, but reacts about half as well with NADH.  相似文献   

4.
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   

5.
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   

6.
Monodehydroascorbate reductase (EC 1.6.5.4) was purified from cucumber fruit to a homogeneous state as judged by polyacrylamide gel electrophoresis. The cucumber monodehydroascorbate reductase was a monomer with a molecular weight of 47,000. It contained 1 mol of FAD/mol of enzyme which was reduced by NAD(P)H and reoxidized by monodehydroascorbate. The enzyme had an exposed thiol group whose blockage with thiol reagents inhibited the electron transfer from NAD(P)H to the enzyme FAD. Both NADH and NADPH served as electron donors with Km values of 4.6 and 23 microM, respectively, and Vmax of 200 mol of NADH and 150 mol of NADPH oxidized mol of enzyme-1 s-1. The Km for monodehydroascorbate was 1.4 microM. The amino acid composition of the enzyme is presented. In addition to monodehydroascorbate, the enzyme catalyzed the reduction of ferricyanide and 2,6-dichloroindophenol but showed little reactivity with calf liver cytochrome b5 and horse heart cytochrome c. The kinetic data suggested a ping-pong mechanism for the monodehydroascorbate reductase-catalyzed reaction. Cucumber monodehydroascorbate reductase occurs in soluble form and can be distinguished from NADPH dehydrogenase, NADH dehydrogenase, DT diaphorase, microsome-bound NADH-cytochrome b5 reductase, and NADPH-cytochrome c reductase by its molecular weight, amino acid composition, and specificity of electron acceptors and donors.  相似文献   

7.
Three kinds of NADPH-dependent aldehyde reducing enzymes were present in the dog kidney. Aldose reductase was located in the inner medulla region and aldehyde reductase in all regions of the renal cortex, outer medulla and inner medulla. In addition, a new reductase designated tentatively as high-Km aldose reductase, which was converted into an aldose reductase-like enzyme, was present in the inner medulla region of the kidney. Aldose reductase, aldehyde reductase and high-Km aldose reductase were purified to homogeneity from each region of the dog kidney. The molecular weight of aldose reductase was estimated to be 38,500 by SDS-polyacrylamide gel electrophoresis and the isoelectric point was found to be 5.7 by chromatofocusing. Aldose reductase had activity for aldo-sugars such as D-xylose, D-glucose and D-galactose as substrates and utilized both NADPH and NADH as coenzymes. Sulfate ions resulted in over 2-fold activation of aldose reductase. All aldehyde reductases from the three regions had the same properties. The molecular weights and isoelectric points of aldehyde reductases were 40,000 and 6.1, respectively. The aldehyde reductases were inactive for D-hexose, utilized only NADPH as coenzyme and were not affected by sulfate ions. High-Km aldose reductase had a molecular weight of 38,500 and an isoelectric point of 5.4. It had activity for aldo-sugars, but showed much higher Km and lower kcat/Km values than aldose reductase. Sulfate ions inhibited high-Km aldose reductase. It was converted into an aldose reductase-like enzyme by incubation in phosphate buffer at pH 7.0. The three kinds of enzymes were strongly inhibited by the known aldose reductase inhibitors. However, aldehyde reductase and high-Km aldose reductase were, in general, less susceptible than aldose reductase.  相似文献   

8.
Abstract— The presence of a nonspecific NADH-linked aldehyde reductase was demonstrated in various regions of bovine brain in vitro. With m-nitrobenzaldehyde as substrate, the rate of NADH oxidation was approximately 4 nmol.min-1.(mg of protein)-1 in the cerebellum, pons and medulla; but somewhat lower rates [2–3 nmol.min-1.(mg of protein)-l] were obtained in the other areas of the brain examined. The enzyme was localized primarily in the soluble, supernatant fraction of rat brain homogenates. The enzyme from the supernatant fluid fraction of bovine brain was purified approximately 350-fold by ammonium sulphate fractionation and chromatography on calcium phosphate-gel, DEAE-cellulose and Sephadex G200 columns. The partially purified enzyme catalysed the reduction of a number of aldehydes, including substituted benzaldehydes and aliphatic aldehydes of intermediate chain lengths. Short chain aliphatic aldehydes, such as acetaldehyde, were not reduced by the enzyme and butyraldehyde was a poor substrate. With m-nitrobenzaldehyde as substrate, NADH was oxidized at an approximately 10-fold faster rate than NADPH. The pH optimum for the enzyme was 6.75 for aldehyde reduction, whereas the rate of oxidation of m-nitrobenzylalcohol was optimal at pH 10.0 with NAD as the co-substrate. Km and K3 values ranged from 10 μM to 10 mM for various aldehydes and from 10 to 30 μM for the cofactors. Oxidation of NADH by the partially purified enzyme was not inhibited by 10m pyrazole or by 1 mM phenobarbital. However, the enzyme activity was inhibited by approximately 60 percent by 1 mM chlorpromazine or by 5 mM 1,10-orthophenanthroline. Our data demonstrate that the enzyme is not only separable from the NADPH-linked aldehyde reductase described previously by TABAKOFF and ERWIN, but also is quite different in substrate specificity and inhibitor sensitivity from the ‘classical’, pyrazole-sensitive, NAD- linked alcohol dehydrogenase (EC 1.1.1.1).  相似文献   

9.
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   

10.
Pyridine nucleotide specificity of barley nitrate reductase   总被引:6,自引:4,他引:2       下载免费PDF全文
Dailey FA  Kuo T  Warner RL 《Plant physiology》1982,69(5):1196-1199
NADPH nitrate reductase activity in higher plants has been attributed to the presence of NAD(P)H bispecific nitrate reductases and to the presence of phosphatases capable of hydrolyzing NADPH to NADH. To determine which of these conditions exist in barley (Hordeum vulgare L. cv. Steptoe), we characterized the NADH and NADPH nitrate reductase activities in crude and affinity-chromatography-purified enzyme preparations. The pH optima were 7.5 for NADH and 6 to 6.5 for the NADPH nitrate reductase activities. The ratio of NADPH to NADH nitrate reductase activities was much greater in crude extracts than it was in a purified enzyme preparation. However, this difference was eliminated when the NADPH assays were conducted in the presence of lactate dehydrogenase and pyruvate to eliminate NADH competitively. The addition of lactate dehydrogenase and pyruvate to NADPH nitrate reductase assay media eliminated 80 to 95% of the NADPH nitrate reductase activity in crude extracts. These results suggest that a substantial portion of the NADPH nitrate reductase activity in barley crude extracts results from enzyme(s) capable of converting NADPH to NADH. This conversion may be due to a phosphatase, since phosphate and fluoride inhibited NADPH nitrate reductase activity to a greater extent than the NADH activity. The NADPH activity of the purified nitrate reductase appears to be an inherent property of the barley enzyme, because it was not affected by lactate dehydrogenase and pyruvate. Furthermore, inorganic phosphate did not accumulate in the assay media, indicating that NADPH was not converted to NADH. The wild type barley nitrate reductase is a NADH-specific enzyme with a slight capacity to use NADPH.  相似文献   

11.
The catalytic properties of a new type of dihydropteridine reductase, NADPH-specific dihydropteridine reductase [EC 1.6.99.10], from bovine liver, were studied and compared with those of the previously characterized enzyme, NADH-specific dihydropteridine reductase [EC 1.6.99.7]. With quinonoid-dihydro-6-methylpterin, approximate Km values of NADPH-specific dihydropteridine reductase for NADPH and NADH were estimated to be 1.4 micron and 2,900 microns, respectively. The Vmax values were 1.34 mumol/min/mg with NADPH and 1.02 mumol/min/mg with NADPH. With NADPH, the Km values of the enzyme for the quinonoid-dihydro forms of 6-methylpterin and biopterin were 1.4 micron and 6.8 microns, respectively. The enzyme was inhibited by its reaction product, NADP+, in a competitive manner, and the inhibition constant was determined to be 3.2 microns. The enzyme was severely inhibited by L-thyroxine and by 2,6-dichlorophenolindophenol.  相似文献   

12.
Coenzyme specificity of mammalian liver D-glycerate dehydrogenase   总被引:1,自引:0,他引:1  
D-Glycerate dehydrogenase (glyoxylate reductase) was partially purified from rat liver by anion- and cation-exchange chromatography. When assayed in the direction of D-glycerate or glycolate formation, the enzyme was inhibited by high (greater than or equal to 0.5 mM), unphysiological concentrations of hydroxypyruvate or glyoxylate much more potently in the presence of NADPH than in the presence of NADH. However, the dehydrogenase displayed a much greater affinity for NADPH (Km less than 1 microM) than for NADH (Km = 48-153 microM). Furthermore, NADP was over 1000-fold more potent than NAD in inhibiting the enzyme competitively with respect to NADH. NADP also inhibited the reaction competitively with respect to NADPH whereas NAD, at concentrations of up to 10 mM had no inhibitory effect. When measured by the formation of hydroxypyruvate from D-glycerate, the enzyme also displayed a much greater affinity for NADP than for NAD. These properties indicate that liver D-glycerate dehydrogenase functions physiologically as an NADPH-specific reductase. In agreement with this conclusion, the addition of hydroxypyruvate or glyoxylate to suspensions of rat hepatocytes stimulated the pentose-phosphate pathway. The coenzyme specificity of D-glycerate dehydrogenase is discussed in relation to the biochemical findings made in D-glyceric aciduria and in primary hyperoxaluria type II (L-glyceric aciduria).  相似文献   

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

14.
Summary Three different NAD(P)H-FMN reductases were extracted from Beneckea harveyi MB-20 cells and separated by DEAE-Sephadex A50 column chromatography. Further purification was achieved by affinity chromatography. In determinations of Km values for NADH, NADPH, and FMN, these three reductases exhibited different specificities and kinetic parameters. One reductase utilizes NADH, whereas a second one utilizes NADPH as the preferred substrate. The third, a newly described reductase species, exhibits about the same reaction rates with NADH and NADPH. The reaction mechanisms of the three enzyme forms have been deduced by steady state kinetic analysis. The highly pure (based on gel electrophoresis) NADPH-FMN reductase still exhibited a low (approximately 2%) activity for NADH, which activity was increased upon storage at 4° but suppressed completely by the replacement of the phosphate buffer with sodium citrate buffer. This high specificity of NADPH-FMN reductase for NADPH under these conditions is useful for the assay of NADPH, notably in systems coupled to bacterial luciferase.  相似文献   

15.
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   

16.
NAD (P) H-dependent reduction of nicotinamide N-oxide was investigated with rabbit liver preparations. Microsomes, microsomal NADPH-cytochrome c reductase or cytosolic aldehyde oxidase alone exhibited no nicotinamide N-oxide reductase activity in the presence of NADPH or NADH. However, when the microsomal preparations were combined with the cytosolic enzyme, a significant N-oxide reductase activity was observed in the presence of the reduced pyridine nucleotide. The activity was enhanced by FAD or methyl viologen. Cytosol alone supplemented with NADPH or NADH exhibited only a slight, but when combined with microsomes, a significant N-oxide reductase activity. Based on these facts, we propose a new electron transfer system consisting of NADPH-cytochrome c reductase and aldehyde oxidase, which exhibits nicotinamide N-oxide reductase activity in the presence of the reduced pyridine nucleotide.  相似文献   

17.
A newly discovered enzyme, α-ketoadipate reductase, has been purified 1000-fold from human placenta. This enzyme catalyzes the following reaction: α-ketoadipate + NADH + H+ → α-hydroxyadipate + NAD. The enzyme has an estimated molecular weight of 95,000 on gel filtration and an isoelectric point at pH 7.0 on electrofocusing. Several forms of the enzyme were isolated during purification. The pH optimum for the major form was 6.3. The reaction product of α-ketoadipate reductase was identified as α-hydroxyadipate by comparison of the enzyme product with chemically prepared α-hydroxyadipate. Studies of the reaction stoichiometry indicated that equimolar quantities of NADH and α-ketoadipate were used in the synthesis of an equivalent quantity of α-hydroxyadipate. Under conditions where the remaining lactate dehydrogenase and malate dehydrogenase were completely inhibited without affecting the α-ketoadipate reductase activity, it was found that α-ketoadipate reductase was highly specific for α-ketoadipate as substrate. NADPH could not substitute for NADH. Initial velocity experiments showed that NADH was an uncompetitive substrate inhibitor.  相似文献   

18.
Glyoxylate and hydroxypyruvate are metabolites involved in the pathway of carbon in photorespiration. The chief glyoxylate-reducing enzyme in leaves is now known to be a cytosolic glyoxylate reductase that uses NADPH as the preferred cofactor but can also use NADH. Glyoxylate reductase has been isolated from spinach leaves, purified to homogeneity, and characterized kinetically and structurally. Chloroplasts contain lower levels of glyoxylate reductase activity supported by both NADPH and NADH, but it is not yet known whether a single chloroplastic enzyme catalyzes glyoxylate reduction with both cofactors. The major hydroxypyruvate reductase activity of leaves has long been known to be a highly active enzyme located in peroxisomes; it uses NADH as the preferred cofactor. To a lesser extent, NADPH can also be used by the peroxisomal enzyme. A second hydroxypyruvate reductase enzyme is located in the cytosol; it preferentially uses NADPH but can also use NADH as cofactor. In a barley mutant deficient in peroxisomal hydroxypyruvate reductase, the NADPH-preferring cytosolic form of the enzyme permits sufficient rates of hydroxypyruvate reduction to support continued substrate flow through the terminal stages of the photosynthetic carbon oxidation (glycolate/glycerate) pathway. The properties and metabolic significance of the cytosolic and organelle-localized glyoxylate and hydroxypyruvate reductase enzymes are discussed.  相似文献   

19.
The NADPH-dependent 2,5-diketo-D-gluconic acid (2,5-DKG) reductase enzyme is a required component in some novel biosynthetic vitamin C production processes. This enzyme catalyzes the conversion of 2,5-DKG to 2-keto-L-gulonic acid, which is an immediate precursor to L-ascorbic acid. Forty unique site-directed mutations were made at five residues in the cofactor-binding pocket of 2,5-DKG reductase A in an attempt to improve its ability to use NADH as a cofactor. NADH is more stable, less expensive and more prevalent in the cell than is NADPH. To the best of our knowledge, this is the first focused attempt to alter the cofactor specificity of a member of the aldo-keto reductase superfamily by engineering improved activity with NADH into the enzyme. Activity of the mutants with NADH or NADPH was assayed using activity-stained native polyacrylamide gels. Eight of the mutants at three different sites were identified as having improved activity with NADH. These mutants were purified and subjected to a kinetic characterization with NADH as a cofactor. The best mutant obtained, R238H, produced an almost 7-fold improvement in catalysis with NADH compared with the wild-type enzyme. Surprisingly, most of this catalytic improvement appeared to be due to an improvement in the apparent kcat for the reaction rather than a large improvement in the affinity of the enzyme for NADH.  相似文献   

20.
Pyrroline-5-carboxylate reductase catalyzes the final step in proline synthesis by NAD(P)H-dependent reduction of pyrroline-5-carboxylate. We have purified and characterized this enzyme from human erythrocytes. Purification to homogeneity (approximately 600,000-fold) was accomplished by sonication, ultracentrifugation, 2',5'-ADP-Sepharose affinity chromatography, and DEAE-Sephacel ion exchange chromatography. The enzyme runs as a single band of 30,000 Mr on sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Sizing chromatography under nondenaturating conditions demonstrates activity in the 300,000-350,000 Mr range, suggesting that the native enzyme exists as a 10- to 12-mer. The purified enzyme exhibits kinetic characteristics similar to those previously described for whole red cell homogenates. The Vmax is 10-fold higher and the Km for pyrroline-5-carboxylate is 7-fold higher with NADH versus NADPH as cofactor. The affinity for NADPH is 15-fold higher than that for NADH. Erythrocyte pyrroline-5-carboxylate reductase is competitively inhibited by NADP+. Unlike the enzyme from some other sources, erythrocyte pyrroline-5-carboxylate reductase is not inhibited by proline or ATP. Double label studies using [14C]pyrroline-5-carboxylate and [3H]exNADPH in the presence of both NADH and NADPH were performed to determine the preferred source of reducing equivalents. In the presence of physiologic concentrations of pyrroline-5-carboxylate and both pyridine nucleotides, all of the reducing equivalents came from NADPH. We suggest that, in some cell types including human erythrocytes, a physiologic function of pyrroline-5-carboxylate reductase is the generation of NADP+.  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