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 359 毫秒
1.
The dual nucleotide cofactor-specific enzyme, 3alpha-hydroxysteroid dehydrogenase (3alpha-HSD) from Pseudomonas sp. B-0831, is a member of the short-chain dehydrogenase/reductase (SDR) superfamily. Transient-phase kinetic studies using the fluorescence stopped-flow method were conducted with 3alpha-HSD to characterize the nucleotide binding mechanism. The binding of oxidized nucleotides, NAD(+), NADP(+) and nicotinic acid adenine dinucleotide (NAAD(+)), agreed well with a one-step mechanism, while that of reduced nucleotide, NADH, showed a two-step mechanism. This difference draws attention to previous characteristic findings on rat liver 3alpha-HSD, which is a member of the aldo-keto reductase (AKR) superfamily. Although functionally similar, AKRs are structurally different from SDRs. The dissociation rate constants associated with the enzyme-nucleotide complex formation were larger than the k(cat) values for either oxidation or reduction of substrates, indicating that the release of cofactors is not rate-limiting overall. It should also be noted that k(cat) for a substrate, cholic acid, with NADP(+) was only 6% of that with NAD(+), and no catalytic activity was detectable with NAAD(+), despite the similar binding affinities of nucleotides. These results suggest that a certain type of nucleotide can modulate nucleotide-binding mode and further the catalytic function of the enzyme.  相似文献   

2.
Pyridine Nucleotide Transhydrogenase from Azotobacter vinelandii   总被引:5,自引:0,他引:5       下载免费PDF全文
A method is described for the partial purification of pyridine nucleotide transhydrogenase from Azotobacter vinelandii (ATCC 9104) cells. The most highly purified preparation catalyzes the reduction of 300 mumoles of nicotinamide adenine dinucleotide (NAD(+)) per min per mg of protein under the assay conditions employed. The enzyme catalyzes the reduction of NAD(+), deamino-NAD(+), and thio-NAD(+) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) as hydrogen donor, and the reduction of nicotinamide adenine dinucleotide phosphate (NADP(+)) and thio-NAD(+) with reduced NAD (NADH) as hydrogen donor. The reduction of acetylpyridine AD(+), pyridinealdehyde AD(+), acetylpyridine deamino AD(+), and pyridinealdehydedeamino AD(+) with NADPH as hydrogen donor was not catalyzed. The enzyme catalyzes the transfer of hydrogen more readily from NADPH than from NADH with different hydrogen acceptors. The transfer of hydrogen from NADH to NADP(+) and thio-NAD(+) was markedly stimulated by 2'-adenosine monophosphate (2'-AMP) and inhibited by adenosine diphosphate (ADP), adenosine triphosphate (ATP), and phosphate ions. The transfer of hydrogen from NADPH to NAD(+) was only slightly affected by phosphate ions and 2'-AMP, except at very high concentrations of the latter reagent. In addition, the transfer of hydrogen from NADPH to thio-NAD(+) was only slightly influenced by 2'-AMP, ADP, ATP, and other nucleotides. The kinetics of the transhydrogenase reactions which utilized thio-NAD(+) as hydrogen acceptor and NADH or NADPH as hydrogen donor were studied in some detail. The results suggest that there are distinct binding sites for NADH and NAD(+) and perhaps a third regulator site for NADP(+) or 2'-AMP. The heats of activation for the transhydrogenase reactions were determined. The properties of this enzyme are compared with those of other partially purified transhydrogenases with respect to the regulatory functions of 2'-AMP and other nucleotides on the direction of flow of hydrogen between NAD(+) and NADP(+).  相似文献   

3.
The pH dependence of the kinetic parameters and the primary deuterium isotope effects with nicotinamide adenine dinucleotide (NAD) and also thionicotinamide adenine dinucleotide (thio-NAD) as the nucleotide substrates were determined in order to obtain information about the chemical mechanism and location of rate-determining steps for the Ascaris suum NAD-malic enzyme reaction. The maximum velocity with thio-NAD as the nucleotide is pH-independent from pH 4.2 to 9.6, while with NAD, V decreases below a pK of 4.8. V/K for both nucleotides decreases below a pK of 5.6 and above a pK of 8.9. Both the tartronate pKi and V/Kmalate decrease below a pK of 4.8 and above a pK of 8.9. Oxalate is competitive vs. malate above pH 7 and noncompetitive below pH 7 with NAD as the nucleotide. The oxalate Kis increases from a constant value above a pK of 4.9 to another constant value above a pK of 6.7. The oxalate Kii also increases above a pK of 4.9, and this inhibition is enhanced by NADH. In the presence of thio-NAD the inhibition by oxalate is competitive vs. malate below pH 7. For thio-NAD, both DV and D(V/K) are pH-independent and equal to 1.7. With NAD as the nucleotide, DV decreases to 1.0 below a pK of 4.9, while D(V/KNAD) and D(V/Kmalate) are pH-independent. Above pH 7 the isotope effects on V and the V/K values for NAD and malate are equal to 1.45, the pH-independent value of DV above pH 7. From the above data, the following conclusions can be made concerning the mechanism for this enzyme. Substrates bind to only the correctly protonated form of the enzyme. Two enzyme groups are necessary for binding of substrates and catalysis. Both NAD and malate are released from the Michaelis complex at equal rates which are equal to the rate of NADH release from E-NADH above pH 7. Below pH 7 NADH release becomes more rate-determining as the pH decreases until at pH 4.0 it completely limits the overall rate of the reaction.  相似文献   

4.
The Na(+)-pumping NADH:quinone oxidoreductase (Na(+)-NQR) is a fundamental enzyme of the oxidative phosphorylation metabolism and ionic homeostasis in several pathogenic and marine bacteria. To understand the mechanism that couples electron transfer with sodium translocation in Na(+)-NQR, the ion dependence of the redox potential of the individual cofactors was studied using a spectroelectrochemical approach. The redox potential of one of the FMN cofactors increased 90 mV in the presence of Na(+) or Li(+), compared to the redox potentials measured in the presence of other cations that are not transported by the enzyme, such as K(+), Rb(+), and NH(4)(+). This shift in redox potential of one FMN confirms the crucial role of the FMN anionic radicals in the Na(+) pumping mechanism and demonstrates that the control of the electron transfer rate has both kinetic (via conformational changes) and thermodynamic components.  相似文献   

5.
Phenylalanine dehydrogenase catalyzes the reversible, pyridine nucleotide-dependent oxidative deamination of L-phenylalanine to form phenylpyruvate and ammonia. We have characterized the steady-state kinetic behavior of the enzyme from Rhodococcus sp. M4 and determined the X-ray crystal structures of the recombinant enzyme in the complexes, E.NADH.L-phenylalanine and E.NAD(+). L-3-phenyllactate, to 1.25 and 1.4 A resolution, respectively. Initial velocity, product inhibition, and dead-end inhibition studies indicate the kinetic mechanism is ordered, with NAD(+) binding prior to phenylalanine and the products' being released in the order of ammonia, phenylpyruvate, and NADH. The enzyme shows no activity with NADPH or other 2'-phosphorylated pyridine nucleotides but has broad activity with NADH analogues. Our initial structural analyses of the E.NAD(+).phenylpyruvate and E.NAD(+). 3-phenylpropionate complexes established that Lys78 and Asp118 function as the catalytic residues in the active site [Vanhooke et al. (1999) Biochemistry 38, 2326-2339]. We have studied the ionization behavior of these residues in steady-state turnover and use these findings in conjunction with the structural data described both here and in our first report to modify our previously proposed mechanism for the enzymatic reaction. The structural characterizations also illuminate the mechanism of the redox specificity that precludes alpha-amino acid dehydrogenases from functioning as alpha-hydroxy acid dehydrogenases.  相似文献   

6.
The kinetic mechanism of betaine aldehyde dehydrogenase from leaves of the plant Amaranthus hypochondriacus is ordered with NAD(+) adding first. NADH is a noncompetitive inhibitor against NAD(+), which was interpreted before as evidence of an iso mechanism, in which NAD(+) and NADH binds to different forms of free enzyme. With the aim of testing the proposed kinetic mechanism, we have now investigated the ability of NADH to form different complexes with the enzyme. By initial velocity and equilibrium binding studies, we found that the steady-state levels of E.glycine betaine are negligible, ruling out binding of NADH to this complex. However, NADH readily bind to E.betaine aldehyde, whose levels most likely are kinetically significant given its low dissociation constant. Also, NADH combined with E.NADH and E.NAD(+). Finally, NADH was not able to revert the hydride transfer step, what suggest that there is no acyl-enzyme intermediate, i.e. the release of the reduced dinucleotide takes place after the deacylation step. Although formation of the complex E.NAD(+).NADH would produce an uncompetitive effect in the inhibition of NADH against NAD(+), the iso mechanism cannot be conclusively discarded.  相似文献   

7.
The stereospecificity of hydrogen transfer between steroid (17-hydroxyprogesterone) and both natural cofactors by bovine testicular 20 alpha-hydroxysteroid dehydrogenase (20 alpha-HSD) has been determined. Cofactors used in these studies, [4-pro-S-3H]NADH ([4B-3H]NADH) and [4-pro-S-3H]NADPH ([4B-3H]NADPH) were generated with human placental estradiol 17 beta-dehydrogenase (EC 1.1.1.62) utilizing [17 alpha-3H]estradiol-17 beta and NAD+ or NADP+, respectively. The resulting [4B-3H]NADH and [4B-3H]NADPH were purified by ion-exchange chromatography and separately incubated with molar excess of 17-hydroxyprogesterone as substrate in the presence of 20 alpha-HSD. Following incubation, steroid reactant and product were extracted, separated by HPLC and quantitated as to mass and content of tritium. The oxidized and reduced cofactors were separated by ion-exchange chromatography and quantitated as to mass and tritium content. In all incubations, equimolar amounts of 17,20 alpha-dihydroxy-4-pregnen-3-one and oxidized cofactor were obtained. Further, all recovered radioactivity remained with cofactor and none was found in the steroid product. In additional experiments, both reduced cofactors were separately incubated with glutamate dehydrogenase, an enzyme known to transfer from the B-side of the nicotinamide ring. Here radioactivity was present only in the unreacted cofactor fractions and in the product, glutamic acid. The results indicate that bovine testicular 20 alpha-HSD catalyzes transfer of the 4A-hydrogen from the dihydronicotinamide moiety of the reduced cofactor. Finally, this work described modifications that represent considerable improvement in the purification and assay of bovine 20 alpha-HSD as originally described.  相似文献   

8.
Nidetzky B  Klimacek M  Mayr P 《Biochemistry》2001,40(34):10371-10381
Microbial xylose reductase, a representative aldo-keto reductase of primary sugar metabolism, catalyzes the NAD(P)H-dependent reduction of D-xylose with a turnover number approximately 100 times that of human aldose reductase for the same reaction. To determine the mechanistic basis for that physiologically relevant difference and pinpoint features that are unique to the microbial enzyme among other aldo/keto reductases, we carried out stopped-flow studies with wild-type xylose reductase from the yeast Candida tenuis. Analysis of transient kinetic data for binding of NAD(+) and NADH, and reduction of D-xylose and oxidation of xylitol at pH 7.0 and 25 degrees C provided estimates of rate constants for the following mechanism: E + NADH right arrow over left arrow E.NADH right arrow over left arrow E.NADH + D-xylose right arrow over left arrow E.NADH.D-xylose right arrow over left arrow E.NAD(+).xylitol right arrow over left arrow E.NAD(+) right arrow over left arrow E.NAD(+) right arrow over left arrow E + NAD(+). The net rate constant of dissociation of NAD(+) is approximately 90% rate limiting for k(cat) of D-xylose reduction. It is controlled by the conformational change which precedes nucleotide release and whose rate constant of 40 s(-)(1) is 200 times that of completely rate-limiting E.NADP(+) --> E.NADP(+) step in aldehyde reduction catalyzed by human aldose reductase [Grimshaw, C. E., et al. (1995) Biochemistry 34, 14356-14365]. Hydride transfer from NADH occurs with a rate constant of approximately 170 s(-1). In reverse reaction, the E.NADH --> E.NADH step takes place with a rate constant of 15 s(-1), and the rate constant of ternary-complex interconversion (3.8 s(-1)) largely determines xylitol turnover (0.9 s(-1)). The bound-state equilibrium constant for C. tenuis xylose reductase is estimated to be approximately 45 (=170/3.8), thus greatly favoring aldehyde reduction. Formation of productive complexes, E.NAD(+) and E.NADH, leads to a 7- and 9-fold decrease of dissociation constants of initial binary complexes, respectively, demonstrating that 12-fold differential binding of NADH (K(i) = 16 microM) vs NAD(+) (K(i) = 195 microM) chiefly reflects difference in stabilities of E.NADH and E.NAD(+). Primary deuterium isotope effects on k(cat) and k(cat)/K(xylose) were, respectively, 1.55 +/- 0.09 and 2.09 +/- 0.31 in H(2)O, and 1.26 +/- 0.06 and 1.58 +/- 0.17 in D(2)O. No deuterium solvent isotope effect on k(cat)/K(xylose) was observed. When deuteration of coenzyme selectively slowed the hydride transfer step, (D)()2(O)(k(cat)/K(xylose)) was inverse (0.89 +/- 0.14). The isotope effect data suggest a chemical mechanism of carbonyl reduction by xylose reductase in which transfer of hydride ion is a partially rate-limiting step and precedes the proton-transfer step.  相似文献   

9.
For redox reactions catalyzed by microbial cells the analysis of involved cofactors is of special interest since the availability of cofactors such as NADH or NADPH is often limiting and crucial for the biotransformation efficiency. The measurement of these cofactors has usually been carried out using spectrophotometric cycling assays. Today LC‐MS/MS methods have become a valuable tool for the identification and quantification of intracellular metabolites. This technology has been adapted to measure all four nicotinamide cofactors (NAD, NADP, NADH, and NADPH) during a whole cell biotransformation process catalyzed by recombinant Escherichia coli cells. The cells overexpressing an alcohol dehydrogenase from Lactobacillus brevis were used for the reduction of methyl acetoacetate (MAA) with substrate‐coupled cofactor regeneration by oxidation of 2‐propanol. To test the reliability of the measurement the data were evaluated using a process model. This model was derived using the measured concentrations of reactants and cofactors for initiation as well as the kinetic constants from in vitro measurements of the isolated enzyme. This model proves to be highly effective in the process development for a whole cell redox biotransformation in predicting both the right concentrations of cofactors and reactants in a batch and in a CSTR process as well as the right in vivo expression level of the enzyme. Moreover, a sensitivity analysis identifies the cofactor regeneration reaction as the limiting step in case for the reduction of MAA to the corresponding product (R)‐methyl 3‐hydroxybutyrate. Using the combination of in vitro enzyme kinetic measurements, measurements of cofactors and reactants and an adequate model initiated by intracellular concentrations of all involved reactants and cofactors the whole cell biotransformation process can be understood quantitatively. Biotechnol. Bioeng. 2009; 104: 251–260 © 2009 Wiley Periodicals, Inc.  相似文献   

10.
The kinetics of the NAD: artificial acceptor-oxidoreductase and delta mu H(+)-dependent succinate: NAD(+)-oxidoreductase reactions (reverse electron transfer) reactions catalyzed by the membrane-bound complex I was studied. The values of apparent rate constants of dissociation of complexes of the oxidized and reduced enzyme with NAD+ and NADH were determined. It was shown that the apparent affinity of NADH for the oxidized complex I is by nearly three orders of magnitude as high as that of the reduced one; a reverse correlation is found for NAD+. A kinetic scheme of complex I functioning in the forward and reverse reactions, according to which the free reduced enzyme is not an intermediate of the forward (NADH-oxidase) reaction and the free oxidized enzyme is not an intermediate of the reverse (NAD(+)-reductase) reaction, is proposed.  相似文献   

11.
Shimada H  Miura K  Imamura Y 《Life sciences》2006,78(25):2931-2936
Progesterone was stereoselectively reduced to a metabolite 20alpha-hydroxy-4-pregnen-3-one in the cytosolic fraction from the liver of male mice, indicating that the reduction of progesterone is catalyzed by 20alpha-hydroxysteroid dehydrogenase (20alpha-HSD). The cytosolic 20alpha-HSD activity was observed not only in the liver, but also in the kidney and lung. In liver cytosol, both NADPH and NADH were effective as cofactors for 20alpha-HSD activity, although NADPH was better than NADH for the enzyme activity. On the other hand, 20alpha-HSD activity in kidney cytosol required only NADPH as a cofactor. No significant sex-related difference of 20alpha-HSD activity was observed in liver and kidney cytosols. Flavonoids have been reported to inhibit the biosynthesis and metabolism of steroids. However, little is known about inhibitory effects of flavonoids on 20alpha-HSD activity. Thus, the effects of 16 flavonoids on 20alpha-HSD activity were examined, using liver cytosol of male mice. Among flavonoids tested, fisetin, apigenin, naringenin, luteolin, quercetin and kaempferol exhibited high inhibitory potencies for the 20alpha-HSD activity. We propose the possibility that these flavonoids augment progesterone signaling by inhibiting potently 20alpha-HSD activity in non-reproductive tissues.  相似文献   

12.
Kinetic studies of formate dehydrogenase   总被引:4,自引:1,他引:3       下载免费PDF全文
1. The kinetic mechanism of formate dehydrogenase is a sequential pathway. 2. The binding of the substrates proceeds in an obligatory order, NAD(+) binding first, followed by formate. 3. It seems most likely that the interconversion of the central ternary complex is extremely rapid, and that the rate-limiting step is the formation or possible isomerization of the enzyme-coenzyme complexes. 4. The secondary plots of the inhibitions with HCO(3) (-) and NO(3) (-) are non-linear, which suggests that more than one molecule of each species is able to bind to the same enzyme form. 5. The rate of the reverse reaction with carbon dioxide at pH6.0 is 20 times that with bicarbonate at pH8.0, although no product inhibition could be detected with carbon dioxide. The low rate of the reverse reaction precluded any steady-state analysis as the enzyme concentrations needed to obtain a measurable rate are of the same order as the K(m) values for NAD(+) and NADH.  相似文献   

13.
2-Hydroxybiphenyl 3-monooxygenase (EC 1.14.13.44) from Pseudomonas azelaica HBP1 is an FAD-dependent aromatic hydroxylase that catalyzes the conversion of 2-hydroxybiphenyl to 2, 3-dihydroxybiphenyl in the presence of NADH and oxygen. The catalytic mechanism of this three-substrate reaction was investigated at 7 degrees C by stopped-flow absorption spectroscopy. Various individual steps associated with catalysis were readily observed at pH 7.5, the optimum pH for enzyme turnover. Anaerobic reduction of the free enzyme by NADH is a biphasic process, most likely reflecting the presence of two distinct enzyme forms. Binding of 2-hydroxybiphenyl stimulated the rate of enzyme reduction by NADH by 2 orders of magnitude. The anaerobic reduction of the enzyme-substrate complex involved the formation of a transient charge-transfer complex between the reduced flavin and NAD(+). A similar transient intermediate was formed when the enzyme was complexed with the substrate analog 2-sec-butylphenol or with the non-substrate effector 2,3-dihydroxybiphenyl. Excess NAD(+) strongly stabilized the charge-transfer complexes but did not give rise to the appearance of any intermediate during the reduction of uncomplexed enzyme. Free reduced 2-hydroxybiphenyl 3-monooxygenase reacted rapidly with oxygen to form oxidized enzyme with no appearance of intermediates during this reaction. In the presence of 2-hydroxybiphenyl, two consecutive spectral intermediates were observed which were assigned to the flavin C(4a)-hydroperoxide and the flavin C(4a)-hydroxide, respectively. No oxygenated flavin intermediates were observed when the enzyme was in complex with 2, 3-dihydroxybiphenyl. Monovalent anions retarded the dehydration of the flavin C(4a)-hydroxide without stabilization of additional intermediates. The kinetic data for 2-hydroxybiphenyl 3-monooxygenase are consistent with a ternary complex mechanism in which the aromatic substrate has strict control in both the reductive and oxidative half-reaction in a way that reactions leading to substrate hydroxylation are favored over those leading to the futile formation of hydrogen peroxide. NAD(+) release from the reduced enzyme-substrate complex is the slowest step in catalysis.  相似文献   

14.
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   

15.
The kinetics of malate dehydrogenase (MDH) catalyzed oxidation/reduction of L-malate/oxaloacetate is pH-dependent due to the proton generated/taken up during the reaction. Previous kinetic studies on the mitochondrial MDH did not yield a consensus kinetic model that explains both substrate and pH dependency of the initial velocity. In this study, we propose, to our knowledge, a new kinetic mechanism to explain kinetic data acquired over a range of pH and substrate concentrations. Progress curves in the forward and reverse reaction directions were obtained under a variety of reactant concentrations to identify associated kinetic parameters. Experiments were conducted at physiologically relevant ionic strength of 0.17 M, pH ranging between 6.5 and 9.0, and at 25°C. The developed model was built on the prior observation of proton uptake upon binding of NADH to MDH, and that the MDH-catalyzed oxidation of NADH may follow an ordered bi-bi mechanism with NADH/NAD binding to the enzyme first, followed by the binding of oxaloacetate/L-malate. This basic mechanism was expanded to account for additional ionic states to explain the pH dependency of the kinetic behavior, resulting in what we believe to be the first kinetic model explaining both substrate and pH dependency of the reaction velocity.  相似文献   

16.
Initial rate kinetic studies of lactate dehydrogenase with ketomalonate and NADH as substrates suggest that this enzymatic system is adapted to a rapid equilibrium ordered bi-bi ternary complex mechanism. The application of the reaction product inhibition method reveals the existence of the enzyme-NADH-hydroxymalonate and enzyme-NAD(+)-ketomalonate abortive complexes. This kinetic behaviour is confirmed by the differential inhibition induced by several alternate products on the pyruvate-lactate dehydrogenase-NADH and ketomalonate-lactate dehydrogenase-NADH systems.  相似文献   

17.
1. The kinetics of oxidation of ethanol, propan-1-ol, butan-1-ol and propan-2-ol by NAD(+) and of reduction of acetaldehyde and butyraldehyde by NADH catalysed by yeast alcohol dehydrogenase were studied. 2. Results for the aldehyde-NADH reactions are consistent with a compulsory-order mechanism with the rate-limiting step being the dissociation of the product enzyme-NAD(+) complex. In contrast the results for the alcohol-NAD(+) reactions indicate that some dissociation of coenzyme from the active enzyme-NAD(+)-alcohol ternary complexes must occur and that the mechanism is not strictly compulsory-order. The rate-limiting step in ethanol oxidation is the dissociation of the product enzyme-NADH complex but with the other alcohols it is probably the catalytic interconversion of ternary complexes. 3. The rate constants describing the combination of NAD(+) and NADH with the enzyme and the dissociations of these coenzymes from binary complexes with the enzyme were measured.  相似文献   

18.
Tally JF  Maniscalco SJ  Saha SK  Fisher HF 《Biochemistry》2002,41(37):11284-11293
We present a multiwavelength, transient-state kinetic study of the oxidative deamination reaction catalyzed by Clostridium symbiosum glutamate dehydrogenase (csGDH) producing the real-time reaction courses of spectroscopically resolved kinetically competent intermediate complexes. The results show striking differences from a corresponding transient-state study of the same reaction by the structurally homologous enzyme from beef liver (blGDH). In addition to the highly blue-shifted alpha-iminoglutarate and highly red-shifted carbinolamine complexes observed in both reactions, the csGDH reaction appeared to show the release of free NADH at a very early and mechanistically unlikely point in the reaction. Using lactic acid dehydrogenase as a "reporter" for free NADH, we show that the early portion of this signal reflects previously unobserved spectrally unshifted enzyme-bound NADH complexes. We provide experimental evidence to show that such spectrally anomalous complexes must represent forms of the known alpha-imino and alpha-carbinolamine complexes in which the active site cleft is open. This evidence includes isothermal calorimetric measurements and pH-jump experiments that show the existence of differing two-state transitions in blGDH and csGDH and locate active site domain motions at differing points in the transient-state time courses of the two enzyme reactions. We prove the kinetic competence of a new and more highly detailed mechanism for the csGDH reaction that involves the alternation of open and closed enzyme complexes as integral steps. These findings, supported by the available X-ray crystal structure data, suggest the existence of a programmed time course of protein domain motions coordinated with the classically considered chemical time course. This new viewpoint may be presumed to be applicable to enzyme reactions other than those of the alpha-amino acid dehydrogenases.  相似文献   

19.
The kinetic scheme of octopine dehydrogenase of Pecten maximus L., a monomeric enzyme obeying a bi-ter sequential mechanism, was completed, essentially in the forward reaction, by steady-state studies over a wide range of substrate concentration at pH 7.0. Deviation from the Michaelis-Menten behavior with respect to NAD+ and other significant kinetic data led us to ascribe for octopine dehydrogenase mechanism the mnemonical enzyme concept. In addition, another regulatory behavior can be envisaged involving the formation of two dead-end complexes enzyme.NADH.D-octopine and enzyme.NAD+.pyruvate.L-arginine.  相似文献   

20.
Chicken liver lactate dehydrogenase L-lactate : NAD+ oxidoreductase, EC1.1.1.27) reversibly catalyses the conversion of hydroxypyruvate to L-glycerate. The variation of the initial reaction rate with the substrate or coenzyme (NADH) concentration together with the inhibition caused by the reaction products and excess substrates, reveal that the kinetic mechanism of the reaction, with hydroxypyruvate as substrate, is of the rapid-equilibrium, ordered-ternary-complex type; NADH is the first substrate in the reaction sequence. Rate equations have been developed for the hydroxypyruvate.E.NADH system without inhibitors, with excess substrates, and with reaction products. Comparison of the rate equations obtained with those calculated theoretically from an ordered-ternary-complex mechanism reveals the existence of E.NAD.NADH,E.NAD-hydroxypyruvate and E.hydroxypyruvate complexes.  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