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.
  • 1.1. A NAD+-dependent glutamate dehydrogenase (EC 1.4.1.2.) was purified 126-fold from Halobacterium halobium.
  • 2.2. Activity and stability of the enzyme were affected by salt concentration. Maximum activity of the NADH-dependent reductive amination of 2-oxoglutarate occurs at 3.2 M NaCl and 0.8 M KCl, and the NAD+-dependent oxidative deamination of l-glutamate occurs at 0.9 M NaCl and 0.4 M KCl. The maximum activity is higher with Na+ than with K+ in the amination reaction while the reverse is true in the deamination reaction.
  • 3.3. The apparent Km values of the various substrates and coenzymes under optimal conditions were: 2-oxoglutarate, 20.2 mM; ammonium, 0.45 M; NADH, 0.07 mM; l-glutamate, 4.0 mM; NAD+, 0.30 mM.
  • 4.4. No effect of ADP or GTP on the enzyme activity was found. The purified enzyme was activated by some l-amino acids.
  相似文献   

2.
Glutamate dehydrogenase, GDH (l-glutamate: NAD+ oxidoreductase (deaminating) EC 1.4.1.2) was purified from the plant fraction of lupin nodules and the purity of the preparation established by gel electrophoresis and electrofocusing. The purified enzyme existed as 4 charge isozymes with a MW of 270000. The subunit MW, as determined by dodecyl sulphate electrophoresis, was 45 000. On the basis of the results of the MW determinations a hexameric structure is proposed for lupin-nodule GDH. The pH optima for the enzyme were pH 8.2 for the amination reaction and pH 8.8 for the deamination reaction. GDH from lupin nodules showed a marked preference for NADH over NADPH in the amination reaction and used only NAD+ for the deamination reaction. Pyridoxal-5′-P and EDTA inhibited activity. The enzyme displayed Michaelis-Menten kinetics with respect to all substrates except NAD+. When NAD+ was the varied substrate, there was a deviation from Michaelis-Menten behaviour towards higher activity at high concentrations of NAD+.  相似文献   

3.
The inhibition of urocanase from Pseudomonas putida by O-methylhydroxylamine has been characterized as being due to the formation of an adduct between CH3ONH2 and NAD+, the latter of which has been recently shown to be a tightly bound coenzyme for this urocanase. Inhibition is maximal at pH 8.5 and is blocked by the presence of the substrate analog imidazole propionate. Loss of catalytic activity corresponds directly with the binding of 1 mol of 14CH3ONH2 per mole of enzyme, and partial reversibility of the modification, achieved by dialysis at pH 7.5, is accompanied by concomitant restoration of enzymatic activity. No incorporation of 14CH3ONH2 into urocanase is seen when enzyme-bound NAD+ is first converted to NADH or when NAD+ is removed by prior treatment of urocanase with 8 m urea. Stability and spectral properties of the CH3ONH · NAD adduct are consistent with previous data reported for the product of the hydroxylamine reaction with NAD+. It is concluded that other urocanases which exhibit inhibition by hydroxylamine may likewise contain NAD+ as an essential coenzyme and that the use of 14CH3ONH2 as a reversible modification reagent for NAD+ should prove helpful for studies on the role of NAD+ in the urocanase catalytic process.  相似文献   

4.
Procedures are described for isolating highly purified porcine liver pyruvate and α-ketoglutarate dehydrogenase complexes. Rabbit serum stabilized these enzyme complexes in mitochondrial extracts, apparently by inhibiting lysosomal proteases. The complexes were purified by a three-step procedure involving fractionation with polyethylene glycol, pelleting through 12.5% sucrose, and a second fractionation under altered conditions with polyethylene glycol. Sedimentation equilibrium studies gave a molecular weight of 7.2 × 106 for the liver pyruvate dehydrogenase complex. Kinetic parameters are presented for the reaction catalyzed by the pyruvate dehydrogenase complex and for the regulatory reactions catalyzed by the pyruvate dehydrogenase kinase and pyruvate dehydrogenase phosphatase. For the overall catalytic reaction, the competitive Ki to Km ratio for NADH versus NAD+ and acetyl CoA versus CoA were 4.7 and 5.2, respectively. Near maximal stimulations of pyruvate dehydrogenase kinase by NADH and acetyl CoA were observed at NADH:NAD+ and acetyl CoA:CoA ratios of 0.15 and 0.5, respectively. The much lower ratios required for enhanced inactivation of the complex by pyruvate dehydrogenase kinase than for product inhibition indicate that the level of activity of the regulatory enzyme is not directly determined by the relative affinity of substrates and products of catalytic sites in the pyruvate dehydrogenase complex. In the pyruvate dehydrogenase kinase reaction, K+ and NH+4 decreased the Km for ATP and the competitive inhibition constants for ADP and (β,γ-methylene)adenosine triphosphate. Thiamine pyrophosphate strongly inhibited kinase activity. A high concentration of ADP did not alter the degree of inhibition by thiamine pyrophosphate nor did it increase the concentration of thiamine pyrophosphate required for half-maximal inhibition.  相似文献   

5.
Sirtuins are key regulators of many cellular functions including cell growth, apoptosis, metabolism, and genetic control of age-related diseases. Sirtuins are themselves regulated by their cofactor nicotinamide adenine dinucleotide (NAD+) as well as their reaction product nicotinamide (NAM), the physiological concentrations of which vary during the process of aging. Nicotinamide inhibits sirtuins through the so-called base exchange pathway, wherein rebinding of the reaction product to the enzyme accelerates the reverse reaction. We investigated the mechanism of nicotinamide inhibition of human SIRT3, the major mitochondrial sirtuin deacetylase, in vitro and in silico using experimental kinetic analysis and Molecular Mechanics-Poisson Boltzmann/Generalized Born Surface Area (MM-PB(GB)SA) binding affinity calculations with molecular dynamics sampling. Through experimental kinetic studies, we demonstrate that NAM inhibition of SIRT3 involves apparent competition between the inhibitor and the enzyme cofactor NAD+, contrary to the traditional characterization of base exchange as noncompetitive inhibition. We report a model for base exchange inhibition that relates such kinetic properties to physicochemical properties, including the free energies of enzyme-ligand binding, and estimate the latter through the first reported computational binding affinity calculations for SIRT3:NAD+, SIRT3:NAM, and analogous complexes for Sir2. The computational results support our kinetic model, establishing foundations for quantitative modeling of NAD+/NAM regulation of mammalian sirtuins during aging and the computational design of sirtuin activators that operate through alleviation of base exchange inhibition.  相似文献   

6.
The perfused rat liver responds in several ways to NAD+ infusion (20–100 μM). Increases in portal perfusion pressure and glycogenolysis and transient inhibition of oxygen consumption and gluconeogenesis are some of the effects that were observed. Extracellular NAD+ is also extensively transformed in the liver. The purpose of the present work was to determine the main products of extracellular NAD+ transformation under various conditions and to investigate the possible contribution of these products for the metabolic effects of the parent compound. The experiments were done with the isolated perfused rat liver. The NAD+ transformation was monitored by HPLC. Confirming previous findings, the single-pass transformation of 100 μM NAD+ ranged between 75% at 1.5 min after starting infusion to 95% at 8 min. The most important products of single-pass NAD+ transformation appearing in the outflowing perfusate were nicotinamide, ADP-ribose, uric acid, and inosine. The relative proportions of these products presented some variations with the time after initiation of NAD+ infusion and the perfusion conditions, but ADP-ribose was always more abundant than uric acid and inosine. Cyclic ADP-ribose (cADP-ribose) as well as adenosine were not detected in the outflowing perfusate. The metabolic effects of ADP-ribose were essentially those already described for NAD+. These effects were sensitive to suramin (P2XY purinergic receptor antagonist) and insensitive to 3,7-dimethyl-1-(2-propargyl)-xanthine (A2 purinergic receptor antagonist). Inosine, a known purinergic A3 agonist, was also active on metabolism, but uric acid and nicotinamide were inactive. It was concluded that the metabolic and hemodynamic effects of extracellular NAD+ are caused mainly by interactions with purinergic receptors with a highly significant participation of its main transformation product ADP-ribose.  相似文献   

7.
Peter Jurtshuk  Linda McManus 《BBA》1974,368(2):158-172
l-(+)-Glutamate oxidation that is non-pyridine nucleotide dependent is readily carried out by a membrane-bound enzyme in Azotobacter vinelandii strain O. Enzyme activity concentrates in a membranous fraction that is associated with the Azotobacter electron transport system. This l-glutamate oxidation is not dependent on externally added NAD+, NADP+, FAD, or FMN for activity. O2, phenazine methosulfate and ferricyanide all served as relatively good electron acceptors for this reaction; while cytochrome c and nitrotetrazolium blue function poorly in this capacity. Paper chromatographic analyses revealed that the 2,4-dinitrophenylhydrazine derivative formed from the enzymatic oxidation of l-glutamate was α-ketoglutarate, while microdiffusion studies indicated that ammonia was also a key end product. These findings suggest that the overall reaction is an oxidative deamination. Ammonia formation was found to be stoichiometric with the amount of oxygen consumed (2 : 1 respectively, on a molar basis). The oxidation of glutamate was limited to the l-(+)-enantiomer indicating that this reaction is not the generalized type carried out by the l-amino acid oxidase. This oxidoreductase is functionally related to the Azotobacter electron transport system: (a) the activity concentrates almost exclusively in the electron transport fraction; (b) the l-glutamate oxidase activity is markedly sensitive to electron transport inhibitors, i.e. 2-n-heptyl-4-hydroxyquinoline-N-oxide, cyanide, and 4,4,4-trifluoro-1-(2-thienyl)-1,3-butanedione; and (c) spectral studies on the Azotobacter R3 fraction revealed that a substantial amount of the flavoprotein (non-heme iron) and cytochrome (a2, a1, b1, c4 and c5) are reduced by the addition of l-glutamate.  相似文献   

8.
Two methods are described for determining protein by measurement of amino acids released by acid or base hydrolysis. One achieves sufficient sensitivity to measure as little as 3 ng of protein by reaction of the amino acids with orthophthalaldehyde and mercaptoethanol in a small volume, followed by dilution in 0.5 n NaOH for fluorometric measurement of the product. The other method is somewhat more sensitive. Released l-glutamate is determined enzymatically by reaction with glutamic dehydrogenase (EC 1.4.1.2) and NAD+; the NADH produced is amplified by enzymatic cycling. To avoid the use of sealed tubes and minimize dangers of contamination, hydrolysis is carried out in small droplets in “oil wells.”  相似文献   

9.
Glutamate dehydrogenase [l-glutamate:NAD+ oxidoreductase (deaminating) EC 1.4.1.2]has been purified 487-fold from pea stem mitochondria. The enzyme has a specific activity in the presence of 1 mm CaCl2 of 54 Enzyme Commission (EC) units. Calcium, manganese, and zinc ions activate the reductive amination reaction. The [Ca2+]0.5 for activation by calcium is 9 μm. The extent of activation by calcium changed during purification and storage. The oxidative deamination was slightly inhibited by calcium. The pH optimum for the reductive amination reaction was 8.0 and for the oxidative deamination was 9.2. At pH 8.0 and in the presence of 1 mm CaCl2 with the ionic strength held constant the enzyme showed normal kinetics for the reductive amination reaction. Under identical conditions except for the absence of CaCl2 the oxidative deamination reaction showed normal kinetics for glutamate. There was substrate activation at high NAD+ concentrations and these concentrations were avoided in the kinetic analysis. A steady-state kinetic analysis showed that a simple mechanism could not be in effect and a partially random mechanism is proposed.  相似文献   

10.
Malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37) has been purified about 480-fold from crude extract of the facultative phototrophic bacterium, Rhodopseudomonas capsulata by only two purification steps, involving Red-Sepharose affinity chromatography. The enzyme has a molecular mass of about 80 kDa and consists of two subunits with identical molecular mass (35 kDa). The enzyme is susceptible to heat inactivation and loses its activity completely upon incubation at 40°C for 10 min. Addition of NAD+, NADH and oxaloacetate, but not l-malate, to the enzyme solution stabilized the enzyme. The enzyme catalyzes exclusively the oxidation of l-malate, and the reduction of oxaloacetate and ketomalonate in the presence of NAD+ and NADH, respectively, as the coenzyme. The pH optima are around 9.5 for the l-malate oxidation, and 7.75–8.5 and 4.3–7.0 for the reduction of oxaloacetate and ketomalonate, respectively. The Km values were determined to be 2.1 mM for l-malate, 48 μM for NAD+, 85 μM for oxaloacetate, 25 μM for NADH and 2.2 mM for ketomalonate. Initial velocity and product inhibition patterns of the enzyme reactions indicate a random binding of the substrates, NAD+ and l-malate, to the enzyme and a sequential release of the products: NADH is the last product released from the enzyme in the l-malate oxidation.  相似文献   

11.
Vanadate in the polymeric form of decavanadate, but not other forms, stimulated oxidation of NADH to NAD+ NADPH was also oxidized with comparable rates. This oxidation of NADH was accompanied by uptake of oxygen and generated hydrogen peroxide with the following stoichiometry: NADH + H+ + O2 → NAD+ + H2O2. The reaction followed second-order kinetics. The rate was dependent on the concentration of both NADH and vanadate and increased with decreasing pH. The reaction had an obligatory requirement for phosphate ions. Esr studies in the presence of the spin trap dimethyl pyrroline N oxide indicated the involvement of Superoxide anion as an intermediate. The reaction was sensitive to Superoxide dismutase and other scavengers of superoxide anions.  相似文献   

12.
We previously demonstrated inhibition of Na+-dependent 32Pi transport in canine renal brush-border membranes in association with NAD+-induced ADP ribosylation of membrane protein(s) and postulated that NAD+ inhibits Pi transport across the brush-border membrane via ADP ribosylation. Recently it was shown that incubation of rat brush-border membrane with NAD+ resulted in release of Pi which was prevented by EDTA. It was proposed that NAD+-mediated inhibition of 32Pi transport might occur through this mechanism. To determine whether NAD+ inhibited 32Pi transport by a mechanism other than or in addition to release of Pi, we compared Na+-dependent 32Pi counterflow in brush-border membrane equilibrated with Pi or with Pi generated from NAD+. Release of Pi from NAD+ incubated with brush-border membrane was confirmed. The increased uptake of 32Pi which was demonstrated in brush-border membrane equilibrated with Pi was not measured when intravesicular Pi was generated from a concentration of NAD+ which effected ADP-ribosylation of brush border membranes (100 μM NAD+). In contrast, increased uptake of 32Pi was demonstrated when intravesicular Pi was generated from 1 μM NAD+ which did not effect ADP ribosylation. Mg2+-dependent ADP ribosylation of brush-border membrane incubated with NAD+ was demonstrated which persisted during the time interval of 32Pi uptake measurements. Our findings are compatible with the hypothesis that NAD+-induced ADP ribosylation of brush-border membrane protein(s) results in inhibition of Pi transport across the membrane in vivo. EDTA may act to prevent this inhibition in brush-border membrane by chelation of Mg2+ and decreased ADP ribosylation.  相似文献   

13.
At physiological pH values the thermodynamic equilibrium constant was determined to be 6,9,10−11 (mol/l). Product inhibition studies are reported which clearly show that the kinetic mechanism of the NAD+-xylitol dehydrogenase is “Ordered-bi-bi”. It can be seen from experimental results that a cosubstrate inhibition with a dead end EA2-complex occurs at elevated NAD+-concentrations. Simulations were carried out which indicate, that under intracellular conditions the NAD+-xylitol dehydrogenase is regulated by the catabolic reduction charge and not by the total concentrations of NAD+ and NADH.  相似文献   

14.
Michel Neuburger  Roland Douce 《BBA》1980,589(2):176-189
Mitochondria isolated from spinach leaves oxidized malate by both a NAD+-linked malic enzyme and malate dehydrogenase. In the presence of sodium arsenite the accumulation of oxaloacetate and pyruvate during malate oxidation was strongly dependent on the malate concentration, the pH in the reaction medium and the metabolic state condition.Bicarbonate, especially at alkaline pH, inhibited the decarboxylation of malate by the NAD+-linked malic enzyme in vitro and in vivo. Analysis of the reaction products showed that with 15 mM bicarbonate, spinach leaf mitochondria excreted almost exclusively oxaloacetate.The inhibition by oxaloacetate of malate oxidation by spinach leaf mitochondria was strongly dependent on malate concentration, the pH in the reaction medium and on the metabolic state condition.The data were interpreted as indicating that: (a) the concentration of oxaloacetate on both sides of the inner mitochondrial membrane governed the efflux and influx of oxaloacetate; (b) the NAD+/NADH ratio played an important role in regulating malate oxidation in plant mitochondria; (c) both enzymes (malate dehydrogenase and NAD+-linked malic enzyme) were competing at the level of the pyridine nucleotide pool, and (d) the NAD+-linked malic enzyme provided NADH for the reversal of the reaction catalyzed by the malate dehydrogenase.  相似文献   

15.
In the rat liver NAD+ infusion produces increases in portal perfusion pressure and glycogenolysis and transient inhibition of oxygen consumption. The aim of the present work was to investigate the possible action of this agent on gluconeogenesis using lactate as a gluconeogenic precursor. Hemoglobin-free rat liver perfusion in antegrade and retrograde modes was used with enzymatic determination of glucose production and polarographic assay of oxygen uptake. NAD+ infusion into the portal vein (antegrade perfusion) produced a concentration-dependent (25–100 μM) transient inhibition of oxygen uptake and gluconeogenesis. For both parameters inhibition was followed by stimulation. NAD+ infusion into the hepatic vein (retrograde perfusion) produced only transient stimulations. During Ca2+-free perfusion the action of NAD+ was restricted to small transient stimulations. Inhibitors of eicosanoid synthesis with different specificities (indo-methacin, nordihydroguaiaretic acid, bromophenacyl bromide) either inhibited or changed the action of NAD+. The action of NAD+ on gluconeogenesis is probably mediated by eicosanoids synthesized in non-parenchymal cells. As in the fed state, in the fasted condition extracellular NAD+ is also able to exert two opposite effects, inhibition and stimulation. Since inhibition did not manifest significantly in retrograde perfusion it is likely that the generating signal is located in pre-sinusoidal regions.  相似文献   

16.
Oxidation of NADH in submitochondrial particles, with O2 or ferricyanide as electron acceptor, was inhibited by micromolar concentrations of NAD+ when measured in 240 mM sucrose or, in a lesser extent, in 120 mM NaCl or LiCl. In 120 mM solutions of either KCl, RbCl, CsCl or NH4Cl the inhibition by up to 100 μM concentrations of NAD+ did not occur. The inhibition observed in the sucrose medium disappeared after solubilization of the particles with detergents and re-appeared when the membranes were reconstituted. The inhibitory effect was potentiated by palmitoyl-CoA. The possibility is discussed that the inhibition of NADH oxidation by low concentrations of NAD+ and its release by K+, Rb+, Cs+ and NH4+ depend on the interaction between NAD+ and the negatively charged mitochondrial membrane.  相似文献   

17.
Sirtuin1 (SIRT1) deacetylase and poly(ADP-ribose)-polymerase-1 (PARP-1) respond to environmental cues, and both require NAD+ cofactor for their enzymatic activities. However, the functional link between environmental/oxidative stress-mediated activation of PARP-1 and SIRT1 through NAD+ cofactor availability is not known. We investigated whether NAD+ depletion by PARP-1 activation plays a role in environmental stimuli/oxidant-induced reduction in SIRT1 activity. Both H2O2 and cigarette smoke (CS) decreased intracellular NAD+ levels in vitro in lung epithelial cells and in vivo in lungs of mice exposed to CS. Pharmacological PARP-1 inhibition prevented oxidant-induced NAD+ loss and attenuated loss of SIRT1 activity. Oxidants decreased SIRT1 activity in lung epithelial cells; however increasing cellular NAD+ cofactor levels by PARP-1 inhibition or NAD+ precursors was unable to restore SIRT1 activity. SIRT1 was found to be carbonylated by CS, which was not reversed by PARP-1 inhibition or selective SIRT1 activator. Overall, these data suggest that environmental/oxidant stress-induced SIRT1 down-regulation and PARP-1 activation are independent events despite both enzymes sharing the same cofactor.  相似文献   

18.
Kinetic studies of oxidation reaction of (S)-1-phenyl-1,2-ethanediol (PED) catalyzed by a NAD+-dependent alcohol dehydrogenase from Candida parapsilosis CCTCC M203011 obtained from China Center for Type Culture Collection (CPADH) were observed for getting insight into the deracemization redox reaction. The data of initial velocity experiments in the absence of product, product (β-hydroxy-hypnone) inhibition experiments and dead-end (pyrazole) inhibition experiments strongly suggest that the reaction follows Theorell-Chance BiBi mechanism in which the coenzymes bind to the free form of the enzyme firstly. The kinetic parameters of this model were estimated by using non-linear regression analysis software.  相似文献   

19.
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   

20.
The pyruvate dehydrogenase complex was isolated from the mitochondria of broccoli florets and shown to be similar in its reaction mechanism to the complexes from other sources. Three families of parallel lines were obtained for the initial velocity patterns, indicating a multisite ping-pong mechanism. The apparent Km values obtained were 321 ± 18, 148 ± 13, and 7.2 ± 0.51 μm for pyruvate, NAD+, and CoA, respectively. Product inhibition studies using acetyl-CoA and NADH yielded results which were in agreement with those predicted by the multisite ping-pong mechanism. Acetyl-CoA and NADH were found to be competitive inhibitors versus CoA and NAD+, respectively. All other substrate-product combinations showed uncompetitive inhibition patterns, except for acetyl-CoA versus NAD+. Among various metabolites tested, only hydroxypyruvate (Ki = 0.11 mM) and glyoxylate (Ki = 3.27 mM) were found to be capable of inhibiting the broccoli enzyme to a significant degree. Initial velocity patterns using Mg2+? or Ca2+-thiamine pyrophosphate and pyruvate as the variable substrate were found to be consistent with an equilibrium ordered mechanism where Mg? or Ca-thiamine pyrophosphate bind first, with dissociation constants of 33.8 and 3 μm, respectively. The Mg- or Ca-thiamine pyrophosphate complexes also dissociated rapidly from the enzyme complex.  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