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.
2.
The kinetics of the reductive amination reaction of lupine-nodule glutamate dehydrogenase (l-glutamate:NAD oxidoreductase (deaminating), EC 1.4.1.2) were found to vary with the identity of the ammonium salt which was used as a substrate. Normal Michaelis-Menten kinetics were obtained with (NH4)2SO4 but when NH4Cl or NH4-acetate was varied apparent substrate inhibition was observed. Linear double-reciprocal plots were obtained with NH4Cl and NH4-acetate, however, if the concentration of Cl? or acetate was maintained constant by adding KCl or K-acetate. Chloride and acetate were subsequently found to cause linear noncompetitive inhibition with respect to NH4+ and the apparent substrate inhibition by NH4Cl and NH4-acetate can be explained as the result varying a substrate and a noncompetitive inhibitor in constant ratio. Other anions were also found to be inhibitors of the glutamate dehydrogenase reaction; I? caused parabolic noncompetitive inhibition with respect to NH4+ and NO3? caused slope-parabolic noncompetitive inhibition with respect to all three substrates of the reductive amination reaction. For the oxidation deamination reaction, Cl? was a linear competitive inhibitor with respect to both NAD and l-glutamate whereas NO3? caused parabolic competitive inhibition with respect to these reactants. To explain the results, it is proposed that anions bind to an allosteric site and cause a change in some of the rate constants of the reaction. Specifically, the results are consistent with anions causing decreases in the rates of association of NADH and 2-oxoglutarate with the enzyme and an increase in the rate of dissociation of NAD.  相似文献   

3.
The data concerning the chemical and kinetic mechanisms of the glutamate dehydrogenase reaction have been reviewed. Based on the differences between two catalytically active glutamate dehydrogenase conformations induced by the substrates as well as on some other evidence, it has been proposed that the amino groups of lysine residues 27 and 126 in the beef liver enzyme are interchangeable depending on the direction of the glutamate dehydrogenase reaction.  相似文献   

4.
The enthalpy change for the oxidative deamination of glutamate by NADP+ catalyzed by bovine liver glutamate dehydrogenase has been determined calorimetrically. The ΔHo values are 64.6 ± 1.2 kJmol and 70.3 ± 1.2 kJmol at 25 and 35°C respectively. The equilibrium constants for the reaction at the two temperatures were determined spectrophotometrically. This enabled the determination of ΔGo and ΔSo of the reaction as well. ΔHovalues were also determined for the reaction using an alternative coenzyme and the deuterated substrate.  相似文献   

5.
Dewanti AR  Mitra B 《Biochemistry》2003,42(44):12893-12901
(S)-Mandelate dehydrogenase from Pseudomonas putida is a member of a FMN-dependent enzyme family that oxidizes (S)-alpha-hydroxyacids to alpha-ketoacids. The reductive half-reaction consists of the steps involved in substrate oxidation and FMN reduction. In this study, we investigated the mechanism of this half-reaction in detail. At low temperatures, a transient intermediate was formed in the course of the FMN reduction reaction. This intermediate is characteristic of a charge-transfer complex of oxidized FMN and an electron-rich donor and is formed prior to full reduction of the flavin. The intermediate was not due to binding of anionic substrates or inhibitors. It was only observed with efficient substrates that have high k(cat) values. At higher temperatures, it was formed within the dead time of the stopped-flow instrument. The rate of formation of the intermediate was 3-4-fold faster than its rate of disappearance; the former had a larger isotope effect. This suggests that the charge-transfer donor is an electron-rich carbanion/enolate intermediate that is generated by the base-catalyzed abstraction of the substrate alpha-proton. This is consistent with the observation that the intermediate was not observed with the R277K and R277G mutants, which have been shown to destabilize the carbanion intermediate (Lehoux, I. E., and Mitra, B. (2000) Biochemistry 39, 10055-10065). Thus, the MDH reaction has two rate-limiting steps of similar activation energies: the formation and breakdown of a distinct intermediate, with the latter step being slightly more rate limiting. We also show that MDH is capable of catalyzing the reverse reaction, the reoxidation of reduced MDH by the product ketoacid, benzoylformate. The transient intermediate was observed during the reverse reaction as well, confirming that it is indeed a true intermediate in the MDH reaction pathway.  相似文献   

6.
Competitive inhibition of glutamate dehydrogenase reaction   总被引:1,自引:0,他引:1  
Choudhury R  Punekar NS 《FEBS letters》2007,581(14):2733-2736
Irrespective of their pyridine nucleotide specificity, all glutamate dehydrogenases share a common chemical mechanism that involves an enzyme bound 'iminoglutarate' intermediate. Three compounds, structurally related to this intermediate, were tested for the inhibition of purified NADP-glutamate dehydrogenases from two Aspergilli, as also the bovine liver NAD(P)-glutamate dehydrogenase. 2-Methyleneglutarate, closely resembling iminoglutarate, was a potent competitive inhibitor of the glutamate dehydrogenase reaction. This is the first report of a non-aromatic structure with a better glutamate dehydrogenase inhibitory potency than aryl carboxylic acids such as isophthalate. A suitably located 2-methylene group to mimic the iminium ion could be exploited to design inhibitors of other amino acid dehydrogenases.  相似文献   

7.
8.
9.
Fluorinated substrate analogs were synthesized and incubated with rat liver 3-hydroxyacyl-CoA dehydrogenase, which reveals that the formation of an enolate intermediate is required for the reaction catalyzed by the enzyme.  相似文献   

10.
11.
Stopped flow studies of D2O kinetic solvent isotope effects on the reaction catalyzed by L-glutamate dehydrogenase reveal, in addition to several effects apparently attributable simply to pKa shifts, a 2-fold pH-independent effect on the velocity of the steady state oxidative deamination of L-glutamate by enzyme and NADP. Comparable pH-independent D2O kinetic solvent isotope effects are seen both in a transient phase of the reaction in which alpha-ketoglutarate is displaced by L-glutamate from an enzyme-NADPH-alpha-ketoglutarate (product) complex and in an analogous model reaction in which alpha-ketoglutarate is displaced by D-glutamate. These results suggest that alpha-ketoglutarate dissociation from an enzyme-NADPH-alpha-ketoglutarate complex is rate-limiting in the steady state.  相似文献   

12.
13.
H W Chih  E N Marsh 《Biochemistry》1999,38(41):13684-13691
Glutamate mutase catalyzes the reversible isomerization of L-glutamate to L-threo-3-methylaspartate. Rapid quench experiments have been performed to measure apparent rate constants for several chemical steps in the reaction. The formation of substrate radicals when the enzyme was reacted with either glutamate or methylaspartate was examined by measuring the rate at which 5'-deoxyadenosine was formed, and shown to be sufficiently fast for this step to be kinetically competent. Furthermore, the apparent rate constant for 5'-deoxyadenosine formation was very similar to that measured previously for cleavage of the cobalt-carbon bond of adenosylcobalamin by the enzyme, providing further support for a mechanism in which homolysis of the coenzyme is coupled to hydrogen abstraction from the substrate. The pre-steady-state rates of methylaspartate and glutamate formation were also investigated. No burst phase was observed with either substrate, indicating that product release does not limit the rate of catalysis in either direction. For the conversion of glutamate to methylaspartate, a single chemical step appeared to dominate the overall rate, whereas in the reverse direction a lag phase was observed, suggesting the accumulation of an intermediate, tentatively ascribed to glycyl radical and acrylate. The rates of formation and decay of this intermediate were also sufficiently rapid for it to be kinetically competent. When combined with information from previous mechanistic studies, these results allow a qualitative free energy profile to constructed for the reaction catalyzed by glutamate mutase.  相似文献   

14.
Aluminium(III) complexes are essential for understanding the toxicity, bioavailability and transport mechanisms of aluminium in environmental and biological systems. Since elucidation of the exact structures of these weakly coordinated systems is very difficult, the structures of Al(III) complexes in glutamate dehydrogenase reactions system were investigated recently from the following four aspects: (1) Constitutional studies: The keto-enol tautomerism of the complexes between aluminium(III) ion and alpha-ketoglutarate ligands in acidic aqueous solutions was studied. It is clearly demonstrated that Al(III) can promote the keto-enol tautomerization of alpha-ketoglutarate. (2) Configurational studies: Compared with L-Glu, the complex stability of D-Glu-Al is stronger, especially for the tridentate species. The result was further supported by computational results in the molecular mechanics model with the UFF forcefield. It is implied that Al(III) complexation may favor the racemization from L- to D-amino acids. (3) Conformational studies: At biologically relevant pH and concentrations of Al(III) and NADH, Al(III) was found to increase the percentage of folded forms of NADH, which results in reducing the activity of the coenzyme NADH in the hollow-dehydrogenase reactions system. However, the conformations of NAD(+) and Al-NAD(+) are dependent upon the solvents and other ligands in the complexes. (4) Biological effects: The effects of Al(III) on the activity of the glutamate dehydrogenase-catalyzed reactions were studied by monitoring the differential-pulse polarography reduction current of NAD(+). At the physiologically relevant pH values (pH 6.5 and 7.5), the activity of the GDH enzyme was strongly dependent on the concentration of the Al(III) in the assayed mixture solutions.  相似文献   

15.
Yeast alcohol dehydrogenase (EC 1.1.1.1) catalyzed reduction of N,N-dimethyl-4-nitrosoaniline by NADH. The stoichiometry of reaction, steady-state kinetic parameters, and the pH-profile for this reaction were estimated. On that basis, the minimal mechanism of the above reaction was postulated.  相似文献   

16.
Glutamate dehydrogenase catalyzes the reversible oxidation of L-proline and L-pipecolic acid to the corresponding cyclic alpha-imino acids. The active substrates are the amino acid anion in one direction and the iminium ion in the other. The oxidation of the ester, amide, and N-methyl derivatives of L-proline by enzyme-NADP+ and the reduction of N-methyl-delta 1-tetrahydropyridinium ion by enzyme-NADPH do not proceed to a detectable extent under the experimental conditions. The methyliminium ion, however, undergoes facile nonenzymatic reduction by NADPH. If it is assumed that the nonenzymatic reaction reflects the structural requirements of the redox step of the enzymatic reaction, then the lack of reactivity of the tetrahydropyridinium ion toward enzyme-NADPH must be due to the instability of the central complex. It appears that the alpha-carboxylate and NH groups in the amino acid anion and in the alpha-imino acid are involved in binding the substrates to the enzyme-coenzyme complexes. We conclude that each of these active substrates binds to its appropriate enzyme-coenzyme complex through a hydrogen bond between its NH group and a basic enzyme group; there is also an ionic bond in the central complex between the alpha-carboxylate group of the active substrate and a positively charged enzyme group. The five-membered amino acid anions are more reactive toward enzyme-NADP+ than the six-membered ones. The same reactivity order is seen for the reduction of imino acids by enzyme-NADPH. Since these effects are also present in the nonenzymatic reduction by NADPH we ascribe the ring size effects on V/Ksubstrate primarily to those on the hydride-transfer step.  相似文献   

17.
Delta 1-Piperidine 2-carboxylic acid, an alpha-imino acid, is reduced by 1,4-dihydropyridines to pipecolic acid, an alpha-amino acid, and the corresponding pyridinium ions. This nonenzymatic reaction occurs only in the direction of pipecolic acid production. Glutamate dehydrogenase catalyzes this reaction when the reductant is NADPH and gives as products L-pipecolic acid and NADP+. The reaction velocity for the enzyme-catalyzed reaction is measurable in either direction. The pH-independent equilibrium constant, Keq, for the reduction of the imino acid by NADPH to give pipecolic acid anion and NADP+ was determined from the equilibrium conditions and the pKa values of pipecolic acid (10.72) and of the cyclic imino acid (8.10). The value of Keq was found to be 175 +/- 30; the values of delta G0, delta H0 and delta S0 are -3.1 +/- 0.1 kcal/mol, 5 +/- 1 kcal/mol and 27 +/- 4 e.u., respectively. The data indicate that the reactants are far more solvated than the products and that there must be a large degree of solvent reorganization during the course of the reaction. If these thermodynamic parameters apply to the redox step of the enzyme-catalyzed glutamate reaction, then the burst phase which results upon mixing the enzyme, L-glutamate and NADP+ in stoichiometric amounts must contain a hidden nonredox step of large delta H0 value to account for the curved Arrhenius plot observed for this phase (A.H. Colen, R.T. Medary and H.F. Fisher, Biopolymers 20 (1981) 879).  相似文献   

18.
We have found previously (Fahien, L.A., Kmiotek, E.H., MacDonald, M. J., Fibich, B., and Mandic, M. (1988) J. Biol. Chem. 263, 10687-10697) that glutamate-malate oxidation can be enhanced by cooperative binding of mitochondrial aspartate aminotransferase and malate dehydrogenase to the alpha-ketoglutarate dehydrogenase complex. The present results demonstrate that glutamate dehydrogenase, which forms binary complexes with these enzymes, adds to this ternary complex and thereby increases binding of the other enzymes. Kinetic evidence for direct transfer of alpha-ketoglutarate and NADH, within these complexes, has been obtained by measuring steady-state rates of E2 when most of the substrate or coenzyme is bound to the aminotransferase or glutamate dehydrogenase (E1). Rates significantly greater than those which can be accounted for by the concentration of free ligand, calculated from the measured values of the E1-ligand dissociation constants, require that the E1-ligand complex serve as a substrate for E2 (Srivastava, D. K., and Bernhard, S. A. (1986) Curr. Tops. Cell Regul. 28, 1-68). By this criterion, NADH is transferred directly from glutamate dehydrogenase to malate dehydrogenase and alpha-ketoglutarate is channeled from the aminotransferase to both glutamate dehydrogenase and the alpha-ketoglutarate dehydrogenase complex. Similar evidence indicates that GTP bound to an allosteric site on glutamate dehydrogenase functions as a substrate for succinic thiokinase. The potential physiological advantages to channeling of activators and inhibitors as well as substrates within multienzyme complexes organized around the alpha-ketoglutarate dehydrogenase complex are discussed.  相似文献   

19.
20.
In previous studies it was found that: (a) aspartate aminotransferase increases the aspartate dehydrogenase activity of glutamate dehydrogenase; (b) the pyridoxamine-P form of this aminotransferase can form an enzyme-enzyme complex with glutamate dehydrogenase; and (c) the pyridoxamine-P form can be dehydrogenated to the pyridoxal-P form by glutamate dehydrogenase. It was therefore concluded (Fahien, L.A., and Smith, S.E. (1974) J. Biol. Chem 249, 2696-2703) that in the aspartate dehydrogenase reaction, aspartate converts the aminotransferase into the pyridoxamine-P form which is then dehydrogenated by glutamate dehydrogenase. The present results support this mechanism and essentially exclude the possibility that aspartate actually reacts with glutamate dehydrogenase and the aminotransferase is an allosteric activator. Indeed, it was found that aspartate is actually an activator of the reaction between glutamate dehydrogenase and the pyridoxamine-P form of the aminotransferase. Aspartate also markedly activated the alanine dehydrogenase reaction catalyzed by glutamate dehydrogenase plus alanine aminotransferase and the ornithine dehydrogenase reaction catalyzed by ornithine aminotransferase plus glutamate dehydrogenase. In these latter two reactions, there is no significant conversion of aspartate to oxalecetate and other compounds tested (including oxalacetate) would not substitute for aspartate. Thus aspartate is apparently bound to glutamate dehydrogenase and this increases the reactivity of this enzyme with the pyridoxamine-P form of aminotransferases. This could be of physiological importance because aspartate enables the aspartate and ornithine dehydrogenase reactions to be catalyzed almost as rapidly by complexes between glutamate dehydrogenase and the appropriate mitochondrial aminotransferase in the absence of alpha-ketoglutarate as they are in the presence of this substrate. Furthermore, in the presence of aspartate, alpha-ketoglutarate can have little or no affect on these reactions. Consequently, in the mitochondria of some organs these reactions could be catalyzed exclusively by enzyme-enzyme complexes even in the presence of alpha-ketoglutarate. Rat liver glutamate dehydrogenase is essentially as active as thebovine liver enzyme with aminotransferases. Since the rat liver enzyme does not polymerize, this unambiguously demonstrates that monomeric forms of glutamate dehydrogenase can react with aminotransferases.  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