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.
An NAD+-linked 17 beta-hydroxysteroid dehydrogenase was purified to homogeneity from a fungus, Cylindrocarpon radicicola ATCC 11011 by ion exchange, gel filtration, and hydrophobic chromatographies. The purified preparation of the dehydrogenase showed an apparent molecular weight of 58,600 by gel filtration and polyacrylamide gel electrophoresis. SDS-gel electrophoresis gave Mr = 26,000 for the identical subunits of the protein. The amino-terminal residue of the enzyme protein was determined to be glycine. The enzyme catalyzed the oxidation of 17 beta-hydroxysteroids to the ketosteroids with the reduction of NAD+, which was a specific hydrogen acceptor, and also catalyzed the reduction of 17-ketosteroids with the consumption of NADH. The optimum pH of the dehydrogenase reaction was 10 and that of the reductase reaction was 7.0. The enzyme had a high specific activity for the oxidation of testosterone (Vmax = 85 mumol/min/mg; Km for the steroid = 9.5 microM; Km for NAD+ = 198 microM at pH 10.0) and for the reduction of androstenedione (Vmax = 1.8 mumol/min/mg; Km for the steroid = 24 microM; Km for NADH = 6.8 microM at pH 7.0). In the purified enzyme preparation, no activity of 3 alpha-hydroxysteroid dehydrogenase, 3 beta-hydroxysteroid dehydrogenase, delta 5-3-ketosteroid-4,5-isomerase, or steroid ring A-delta-dehydrogenase was detected. Among several steroids tested, only 17 beta-hydroxysteroids such as testosterone, estradiol-17 beta, and 11 beta-hydroxytestosterone, were oxidized, indicating that the enzyme has a high specificity for the substrate steroid. The stereospecificity of hydrogen transfer by the enzyme in dehydrogenation was examined with [17 alpha-3H]testosterone.  相似文献   

2.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

3.
The purified respiratory chain NADH dehydrogenase of Escherichia coli oxidizes NADH with either dichlorophenolindophenol (DCIP). ferricyanide, or menadione as electron acceptors, with values for NADH are similar with the three electron acceptors (approximately 50 muM). The purified enzyme contains no flavin and has an absolute requirement for FAD, with Km values around 4 muM. The pH optimum of the enzyme appears to be between 6.5 and 7; the optimum is difficult to establish because of nonenzymatic reduction of DCIP at the lower pH values. Potassium cyanide stimulates the DCIP reductase activity about 2-fold, but has no effect on ferricyanide reductase. The enzyme exhibits hyperbolic kinetics with respect to NADH concentration in both the ferricyanide and DCIP reductase assays, but cooperatively is seen in the menadione reductase reaction. NAD+ is an effective competitive inhibitor of the reaction (Ki congruent to 20 muM); in the presence of NAD+, the NADH saturation curve becomes cooperative, even in the DCIP reductase assay. Many adenine containing nucleotides are competitive inhibitors of the enzyme. The apparent Ki values for these nucleotides as inhibitors of the purified enzyme, the membrane-bound NADH dehydrogenase, and the NADH oxidase are equivalent. An examination of inhibitory effects of a series of adenine nucleotides suggests that the inhibitors act as analogues of NAD+, which is the true physiological inhibitor. The results suggest that the enzyme in situ is always partially inhibited by the levels of NAD- in the E coli cell, and thus behaves in a cooperative fashion to changes in the NAD+/NADH ratio. An antibody has been elicited against the purified NADH dehydrogenase. Immunodiffusion and crossed immunoelectrophoresis show that the antibody is directed principally against the NADH dehydrogenase, with some activity against minor contaminants in the purified preparation. The antibody inhibits NADH dehydrogenase activity 50% at saturating levels. When this antibody preparation is used to examine solubilized membrane preparations, two major immunoprecipitates are found. A parallel inhibition of the membrane-bound NADH dehydrogenase and NADH oxidase activities is seen, supporting the hypothesis that the purified enzyme is indeed a component of the respiratory chain-dependent NADH oxidase pathway.  相似文献   

4.
Dihydrouracil dehydrogenase (NADP+) (EC 1.3.1.2) was partially purified from the cytosol fraction of rat liver and fractionated by disc gel electrophoresis. A major and minor band were visualized by staining for enzyme activity. The substrate specificity of these bands was investigated. It was found that both bands were two to three times more active with dihydrothymine as substrate than with dihydrouracil in the presence of NADP+ and the optimum pH of 7.4. Mitochondrial fractions containing most of the NADH-dependent uracil reductase of rat liver cells were fractionated by centrifugation in sucrose density gradients. Two procedures involving linear or discontinuous gradients were used. By both, good separation of NADH- and NADPH- dependent reductases was achieved. Marker enzyme studies supported the view that the NADH-dependent enzyme is located principally in mitochondria whereas the NADPH-dependent enzyme is mainly in plasma and endoplasmic reticulum membranes. For the NADH-dependent reductase the apparent Km for thymine at pH 7.4 was 1.39 times that found for uracil whereas for the NADPH-dependent enzyme the apparent Km values were similar for the two substrates at this pH. Dihydrouracil was the principal product isolated by paper chromatography from the reaction mixture containing a partially purified fraction of mitochondria, uracil and NADH at pH 7.4. This fraction also catalyzed the formation of radioactive carbon dioxide from [2-14C]uracil. The proportion of CO2 formed by the mitochondria was about 10% of that formed by the original homogenate.  相似文献   

5.
Initial velocity studies and product inhibition studies were conducted for the forward and reverse reactions of formaldehyde dehydrogenase (formaldehyde: NAD oxidoreductase, EC 1.2.1.1) isolated from a methanol-utilizing yeast Candida boidinii. The data were consistent with an ordered Bi-Bi mechanism for this reaction in which NAD+ is bound first to the enzyme and NADH released last. Kinetic studies indicated that the nucleoside phosphates ATP, ADP and AMP are competitive inhibitors with respect to NAD and noncompetitive inhibitors with respect to S-hydroxymethylglutathione. The inhibitions of the enzyme activity by ATP and ADP are greater at pH 6.0 and 6.5 than at neutral or alkaline pH values. The kinetic studies of formate dehydrogenase (formate:NAD oxidoreductase, EC 1.2.1.2) from the methanol grown C. boidinii suggested also an ordered Bi-Bi mechanism with NAD being the first substrate and NADH the last product. Formate dehydrogenase the last enzyme of the dissimilatory pathway of the methanol metabolism is also inhibited by adenosine phosphates. Since the intracellular concentrations of NADH and ATP are in the range of the Ki values for formaldehyde dehydrogenase and formate dehydrogenase the activities of these main enzymes of the dissimilatory pathway of methanol metabolism in this yeast may be regulated by these compounds.  相似文献   

6.
l-Glycerol 3-phosphate dehydrogenase has been isolated and partially purified from the endosperm of developing castor beans. The enzyme is entirely cytosolic and is not found in the plastid fraction. No activity was found in germinating castor beans. The pH optimum for the reduction of dihydroxyacetone phosphate is 8.1 and is 9.6 for the reverse reaction. The molecular weight determined by gel filtration chromatography is between 71,000 and 83,000. Both substrates show substrate inhibition at concentrations about 13 μm for NADH and 400 μm for dihydroxyacetone phosphate. Substrate interaction kinetics gave limiting Km values of 2.7 and 35.5 μm for NADH and dihydroxyacetone phosphate, respectively. Substrate interaction and product inhibition kinetics were consistent with an ordered sequential mechanism with NADH being the first substrate to bind and NAD+ being the last product to dissociate.  相似文献   

7.
Intact glyoxysomes were isolated from castor bean endosperm on isometric Percoll gradients. The matrix enzyme, malate dehydrogenase, was 80% latent in the intact glyoxysomes. NADH:ferricyanide and NADH:cytochrome c reductase activities were measured in intact and deliberately broken organelles. The latencies of these redox activities were found to be about half the malate dehydrogenase latency. Incubation of intact organelles with trypsin eliminated NADH:cytochrome c reductase activity, but did not affect NADH:ferricyanide reductase activity. NADH oxidase and transhydrogenase activities were negligible in isolated glyoxysomes. Mersalyl and Cibacron blue 3GA were potent inhibitors of NADH:cytochrome c reductase. Quinacrine, Ca2+ and Mg2+ stimulated NADH:cytochrome c reductase activity in intact glyoxysomes. The data suggest that some electron donor sites are on the matrix side and some electron acceptor sites are on the cytosolic side of the membrane.  相似文献   

8.
1. The mechanisms of the reduction of oxaloacetate and of 3-fluoro-oxaloacetate by NADH catalysed by cytoplasmic pig heart malate dehydrogenase (MDH) were investigated. 2. One mol of dimeric enzyme produces 1.7+/-0.4 mol of enzyme-bound NADH when mixed with saturating NAD+ and L-malate at a rate much higher than the subsequent turnover at pH 7.5. 3. Transient measurements of protein and nucleotide fluorescence show that the steady-state complex in the forward direction is MDH-NADH and in the reverse direction MDH-NADH-oxaloacetate. 4. The rate of dissociation of MDH-NADH was measured and is the same as Vmax. in the forward direction at pH 7.5. Both NADH-binding sites are kinetically equivalent. The rate of dissociation varies with pH, as does the equilibrium binding constant for NADH. 5. 3-Fluoro-oxaloacetate is composed of three forms (F1, F2 and S) of which F1 and F2 are immediately substrates for the enzyme. The third form, S, is not a substrate, but when the F forms are used up form S slowly and non-enzymically equilibrates to yield the active substrate forms. S is 2,2-dihydroxy-3-fluorosuccinate. 6. The steady-state compound during the reduction of form F1 is an enzyme form that does not contain NADH, probably MDH-NAD+-fluoromalate. The steady-state compound for form F2 is an enzyme form containing NADH, probably MDH-NADH-fluoro-oxaloacetate. 7. The rate-limiting reaction in the reduction of form F2 shows a deuterium isotope rate ratio of 4 when NADH is replaced by its deuterium analogue, and the rate-limiting reaction is concluded to be hydride transfer. 8. A novel titration was used to show that dimeric cytoplasmic malate dehydrogenase contains two sites that can rapidly reduce the F1 form of 3-fluoro-oxaloacetate. The enzyme shows 'all-of-the-sites' behaviour. 9. Partial mechanisms are proposed to explain the enzyme-catalysed transformations of the natural and the fluoro substrates. These mechanisms are similar to the mechanism of pig heart lactate dehydrogenase and this, and the structural results of others, can be explained if the two enzymes are a product of divergent evolution.  相似文献   

9.
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   

10.
Enzymatic properties, renaturation and metabolic role of mannitol-1-phosphate dehydrogenase from Escherichia coli. D-mannitol-1-phosphate dehydrogenase was purified to homogeneity from Escherichia coli, and its physicochemical and enzymatic properties were investigated. The molecular weight of the polypeptide chain is 45,000 as determined by polyacrylamide gel electrophoresis in denaturing conditions. High performance size exclusion chromatography gives an apparent molecular weight of 47,000 for the native enzyme, showing that D-mannitol-1-phosphate dehydrogenase is a monomeric NAD-dependent dehydrogenase. D-mannitol-1-phosphate dehydrogenase is rapidly denatured by 6 M guanidine hydrochloride. Non-superimposable transition curves for the loss of activity and the changes in fluorescence suggest the existence of a partially folded inactive intermediate. The protein can be fully renatured after complete unfolding, and the regain of both native fluorescence and activity occurs rapidly within a few seconds at pH 7.5 and 20 degrees C. Such a high rate of reactivation is unusual for a protein of this size. D-mannitol-1-phosphate dehydrogenase is specific for mannitol-1-phosphate (or fructose-6-phosphate) as a substrate and NAD+ (or NADH) as a cofactor. Zinc is not required for the activity. The affinity of D-mannitol-1-phosphate dehydrogenase for the reduced or oxidized form of its substrate or cofactor remains constant with pH. The affinity for NADH is 20-fold higher than for NAD+. The forward and reverse catalytic rate constants of the reaction: mannitol-1-phosphate + NAD+ in equilibrium fructose-6-phosphate + NADH have different pH dependences. The oxidation of mannitol-1-phosphate has an optimum pH of 9.5, while the reduction of fructose-6-phosphate has its maximum rate at pH 7.0. At pH values around neutrality the maximum rate of reduction of fructose-6-phosphate is much higher than that of oxidation of mannitol-1-phosphate. The enzymatic properties of isolated D-mannitol-1-phosphate dehydrogenase are discussed in relation to the role of this enzyme in the intracellular metabolism.  相似文献   

11.
Irmgard Ziegler 《Phytochemistry》1974,13(11):2403-2410
Sephadex G-200 gel filtration of an ammonium sulfate fraction, containing the bulk of NAD-dependent malate dehydrogenase, yields forms of differing MW. Both Mg2+ and NADH stabilize the 127000 daltons MW form. K+, or incubation with dithioerythritol, cause splitting and partial reaggregation, resulting in MWs ranging between 35000 and 180000 daltons. Chromatography in the presence of dithioerythritol and NADH results in an enzyme with a non-linear reaction rate at low substrate concentrations. Plots of initial velocity vs substrate and cofactor concentration respectively are characterized by two slopes of positive cooperativity separated by an intermediary plateau of negative cooperativity. Gel chromatography in the presence of Mg2+ or K+ or drastic dilution of the enzyme results in an enzyme with linear reaction rates also at low substrate concentration. Its kinetics are consistent with the view that the enzyme undergoes conformational changes when the substrate concentration is varied.  相似文献   

12.
Lactate dehydrogenase (EC 1.1.1.27) catalyzes the NAD-dependent oxidation to (oxalate) and reduction (to glycollate) of glyoxylate. The kinetics of this disproportionation are in accord with the usual reaction pathway of lactate dehydrogenase:substrate inhibition with appropriate pH dependence occurs; a steady state in the ratio of NADH to NAD+ is set up during the reaction, has the expected dependence on pH, and is independent of the initial glyoxylate, coenzyme, and enzyme concentration. At pH 7 the lactate dehydrogenase-NADH complex is about fivefold more likely to react with and reduce glyoxylate (at a concentration of 100 mm) than to dissociate to produce free NADH, and the ratio of the fraction of the enzyme-NADH complex which dissociates to the fraction which reacts with and reduces glyoxylate varies with glyoxylate concentration and with pH in a manner in agreement with the normal reaction pathway of the enzyme. With all concentrations of glyoxylate and over the pH range 7–9.6 both free (not enzyme bound) NAD+ and free NADH are formed in the steady state of the disproportionation. From these results it is apparent that lactate dehydrogenase, like alcohol dehydrogenase (EC 1.1.1.1), catalyzes a disproportionation within the bounds of its normal kinetic reaction pathway.  相似文献   

13.
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   

14.
Deamino-NADH/ubiquinone 1 oxidoreductase activity in membrane preparations from Escherichia coli GR19N is 20-50% of NADH/ubiquinone 1 oxidoreductase activity. In comparison, membranes from E. coli IY91, which contain amplified levels of NADH dehydrogenase, exhibit about 100-fold higher NADH/ubiquinone 1 reductase activity but about 20-fold less deamino-NADH/ubiquinone 1 reductase activity. Deamino-NADH/ubiquinone 1 reductase is more sensitive than NADH/ubiquinone 1 reductase activity to inhibition by 3-undecyl-2-hydroxyl-1,4-naphthoquinone, piericidin A, or myxothiazol. Furthermore, GR19N membranes exhibit two apparent Kms for NADH but only one for deamino-NADH. Inside-out membrane vesicles from E. coli GR19N generate a H+ electrochemical gradient (interior positive and acid) during electron transfer from deamino-NADH to ubiquinone 1 that is large and stable relative to that observed with NADH as substrate. Generation of the H+ electrochemical gradient in the presence of deamino-NADH is inhibited by 3-undecyl-2-hydroxy-1,4-naphthoquinone and is not observed in IY91 membrane vesicles or in vesicles from GR19N that are deficient in deamino-NADH/ubiquinone 1 reductase activity. The data provide a strong indication that the E. coli aerobic respiratory chain contains two species of NADH dehydrogenases: (i) an enzyme (NADH dh I) that reacts with deamino-NADH or NADH whose turnover leads to generation of a H+ electrochemical gradient at a site between the primary dehydrogenase and ubiquinone and (ii) an enzyme (NADH dh II) that reacts with NADH exclusively whose turnover does not lead to generation of a H+ electrochemical gradient between the primary dehydrogenase and ubiquinone 1.  相似文献   

15.
NADH, (3-acetylpyridine)ADH, (3-thionicotinamide)ADH and (pyridine-3-aldehyde)ADH were labeled with deuterium by the B specific enzyme lipoyl dehydrogenase. The primary acid rearrangement products, α-O2′-6B cyclotetra-hydropyridine adenine dinucleotides, of the respective reduced coenzyme analogs were synthesized by incubation at pH 2–2.5. Analysis of the spin-spin coupling constants of the 220 MHz proton magnetic resonance spectra establishes that the structure of the ribose-tetrahydropyridine linkage of these analogs is the same as that of the primary acid product of NADH. All four specifically labeled dinucleotides possess the identical absolute configuration at the C4 position with respect to the deuterium label, establishing that the three coenzyme analogs are reduced by lipoyl dehydrogenase with the same absolute stereospecificity as NAD+. The primary acid reaction is described as a general method for the determination of the absolute configuration at the C4 position of specifically labeled β-1,3-substituted, 1,4 dihydropyridine nucleotides.  相似文献   

16.
Cell free extracts of Hansenula miso IFO 0146 contained an enzyme which catalyzed acyloin condensation of acetaldehyde and α-ketoglutarate to form 5~hydroxy-4-ketohexanoic acid (HKH). The enzyme was specific for acetaldehyde and α-ketoglutarate. Condensation could not be demonstrated between α-ketoglutarate and other aldehydes tested (formaldehyde, propionaldehyde or butyraldehyde). No reaction occurred when boiled enzyme was used. The apparent Km values (at pH 7.5) for acetaldehyde and α-ketoglutarate are 24.4 mM and 3.2 mM, respectively. TPP and Mg2+ were not required for the reaction. The optimum pH of the reaction was 7.5~8.5. The reaction was inhibited by EDTA, PCMB and PMS. The enzyme forming HKH was different from that forming acetoin because the latter required TPP and was repressed when cells were grown in lactate medium while the former did not require TPP and was formed independently of its substrate. The product of this condensing reaction was isolated and identified as HKH from its chemical properties.  相似文献   

17.
The steady-state kinetic mechanism of the reaction catalyzed by octopine dehydrogenase [N2-(1-carboxyethyl)-L-arginine:NAD+ oxidoreductase] was investigated at pH 6.9 and 9.2 by studies of substrate inhibition, analogue inhibition, and product inhibition. In the direction of octopine synthesis, the inhibition patterns in the presence of delta- guanidinovalerate and propionate show that NADH binds to the enzyme first followed by L-arginine and pyruvate which bind randomly. In the direction of octopine oxidation, the substrate patterns show that NAD binds to the enzyme before octopine in a rapid equilibrium fashion, and the product inhibition patterns show that the products L-arginine and pyruvate are released in a random fashion. Double, synergistic, substrate inhibition by L-arginine and pyruvate was shown to be due to binding (hypothetically of the imine) to the free enzyme and the enzyme-NAD complex. Furthermore, an alternate minor pathway was demonstrated which includes an enzyme-NADH-octopine complex and an enzyme-octopine complex.  相似文献   

18.
Saccharopine dehydrogenase was previously purified 380-fold from human placenta. The enzyme was shown to catalyze the formation of α-aminoadipic-δ-semialdehyde and glutamate from saccharopine, to have a molecular weight of 480,000 on gel filtration, and not to be separable from l-lysine-α-ketoglutarate reductase. Additional properties of the saccharopine dehydrogenase are now described. The pH optimum for the conversion of saccharopine to glutamate and α-aminoadipic-δ-semialdehyde is 8.5 in Tris-HCl buffer and 8.9 in 2-amino-2-methyl-1,3-propanediol buffer. The specificity of the enzyme for Saccharopine and NAD and the inhibition by glutamate and product analogs were tested. It was found the NADP was the only cofactor that could replace NAD in the enzyme reaction and that several NAD analogs were reaction inhibitors. Glutamate was found to be only moderately effective as an inhibitor. Initial velocity studies revealed that the enzyme has an ordered reaction mechanism. The true Km values for saccharopine and NAD are 1.15 mm and 0.0645 mm, respectively.  相似文献   

19.
Steady state initial velocity studies were carried out to determine the kinetic mechanism of human liver aldehyde dehydrogenase. Intersecting double reciprocal plots obtained in the absence of inhibitors demonstrated that the dehydrogenase reaction proceeded by sequential addition of both substrates prior to release of products. Dead end inhibition patterns obtained with coenzyme and substrate analogues (e.g. thionicotinamide-AD+ and chloral hydrate) indicated that NAD+ and aldehyde can bind in random fashion. The patterns of inhibition by the product NADH and of substrate inhibition by glyceraldehyde were also consistent with this mechanism. However, comparisons between kinetic constants associated with the dehydrogenase and esterase activities of this enzyme suggested that most of the dehydrogenase reaction flux proceeds via formation of an initial binary NAD+-enzyme complex over a wide range of substrate and coenzyme concentrations.  相似文献   

20.
The reaction of NAD(P)H:flavin oxidoreductase (flavin reductase) from Photobacterium fischeri is proposed to follow a ping-pong bisubstrate-biproduct mechanism. This is based on a steady-state kinetic analysis of initial velocities and patterns of inhibition by NAD+ and AMP. The double reciprocal plots of initial velocities versus concentrations of FMN or NADH show, in both cases, a series of parallel lines. The Michaelis constants for NADH (FMN saturating) and FMN (NADH saturating) are 2.2 and 1.2 × 10?4m, respectively. The product NAD+ has been found to be an inhibitor competitive with FMN but non-competitive with NADH. Using AMP as an inhibitor, noncompetitive inhibition patterns were observed with respect to both NADH and FMN as the varied substrate. In addition, the reductase was not inactivated by treatment with N-ethylmaleimide either alone or in the presence of FMN, but the enzyme was inactivated by N-ethylmaleimide in the presence of NADH. These findings suggest that flavin reductase shuttles between disulfide- and sulfhydryl-containing forms during catalysis.  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