首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 31 毫秒
1.
Argyrou A  Blanchard JS  Palfey BA 《Biochemistry》2002,41(49):14580-14590
Lipoamide dehydrogenase catalyses the NAD(+)-dependent oxidation of the dihydrolipoyl cofactors that are covalently attached to the acyltransferase components of the pyruvate dehydrogenase, alpha-ketoglutarate dehydrogenase, and glycine reductase multienzyme complexes. It contains a tightly, but noncovalently, bound FAD and a redox-active disulfide, which cycle between the oxidized and reduced forms during catalysis. The mechanism of reduction of the Mycobacterium tuberculosis lipoamide dehydrogenase by NADH and [4S-(2)H]-NADH was studied anaerobically at 4 degrees C and pH 7.5 by stopped-flow spectrophotometry. Three phases of enzyme reduction were observed. The first phase, characterized by a decrease in absorbance at 400-500 nm and an increase in absorbance at 550-700 nm, was fast (k(for) = 1260 s(-)(1), k(rev) = 590 s(-)(1)) and represents the formation of FADH(2).NAD(+), an intermediate that has never been observed before in any wild-type lipoamide dehydrogenase. A primary deuterium kinetic isotope effect [(D)(k(for) + k(rev)) approximately 4.2] was observed on this phase. The second phase, characterized by regain of the absorbance at 400-500 nm, loss of the 550-700 nm absorbance, and gain of 500-550 nm absorbance, was slower (k(obs) = 200 s(-)(1)). This phase represents the intramolecular transfer of electrons from FADH(2) to the redox-active disulfide to generate the anaerobically stable two-electron reduced enzyme, EH(2). The third phase, characterized by a decrease in absorbance at 400-550 nm, represents the formation of the four-electron reduced form of the enzyme, EH(4). The observed rate constant for this phase showed a decreasing NADH concentration dependence, and results from the slow (k(for) = 57 s(-)(1), k(rev) = 128 s(-)(1)) isomerization of EH(2) or slow release of NAD(+) before rapid NADH binding and reaction to form EH(4). The mechanism of oxidation of EH(2) by NAD(+) was also investigated under the same conditions. The 530 nm charge-transfer absorbance of EH(2) shifted to 600 nm upon NAD(+) binding in the dead time of mixing of the stopped-flow instrument and represents formation of the EH(2).NAD(+) complex. This was followed by two phases. The first phase (k(obs) = 750 s(-)(1)), characterized by a small decrease in absorbance at 435 and 458 nm, probably represents limited accumulation of FADH(2).NAD(+). The second phase was characterized by an increase in absorbance at 435 and 458 nm and a decrease in absorbance at 530 and 670 nm. The observed rate constant that describes this phase of approximately 115 s(-)(1) probably represents the overall rate of formation of E(ox) and NADH from EH(2) and NAD(+), and is largely determined by the slower rates of the coupled sequence of reactions preceding flavin oxidation.  相似文献   

2.
Isomerization of 5-pregnene-3,20-dione to progesterone by human placental microsomes was stimulated by NAD and NADH. Concomitant oxidation or reduction of nucleotide was not detected based on absorbance at 340 nm. Concentrations giving half-maximum activity were 0.76 microM for NADH and 24.0 microM for NAD. Vmax values with 9.28 microM 5-pregnene-3,20-dione were 22.0 nmol/min/mg protein with NADH and 65.8 nmol/min/mg protein with NAD. When isomerase was assayed as a function of 5-pregnene-3,20-dione concentration, NAD increased Vmax but had no effect on the Km value for steroid. NADP, NADPH, acetylpyridine NAD and deamino NAD did not activate nor did they compete with NAD. Exposure of microsomes to trypsin, phospholipase A2 or phospholipase C resulted in the loss of isomerase activity. Approximately 30% of the initial activity was recovered after detergent solubilization of microsomes. Hydrogen peroxide did not affect activation by NAD. The data are consistent with nucleotide enhancement of a step in the isomerization reaction other than substrate binding.  相似文献   

3.
Soybean nodule xanthine dehydrogenase: a kinetic study   总被引:1,自引:0,他引:1  
Xanthine dehydrogenase was purified from soybean nodules and the kinetic properties were studied at pH 7.5. Km values of 5.0 +/- 0.6 and 12.5 +/- 2.5 microM were obtained for xanthine and NAD+, respectively. The pattern of substrate dependence suggested a Ping-Pong mechanism. Reaction with hypoxanthine gave Km's of 52 +/- 3 and 20 +/- 2.5 microM for hypoxanthine and NAD+, respectively. The Vmax for this reaction was twice that for the xanthine-dependent reaction. The pH dependence of Vmax gave a pKa of 7.6 +/- 0.1 for either xanthine or hypoxanthine oxidation. In addition the Km for xanthine had a pKa of 7.5 consistent with the protonated form of xanthine being the true substrate. Km for hypoxanthine varied only 2.5-fold between pH 6 and 10.7. Product inhibition studies were carried out with urate and NADH. Both products gave mixed inhibition with respect to both substrates. Xanthine dehydrogenase was able to use APAD+ as an electron acceptor for xanthine oxidation, with a Km at pH 7.5 of 21.2 +/- 2.5 microM and Vmax the same as that obtained with NAD+. Reduction of APAD+ by NADH was also catalyzed by xanthine dehydrogenase with a Km of 102 +/- 15 microM; Vmax was approximately 2.5 times that for the xanthine-dependent reaction, and was independent of pH between 6 and 9. Reaction with group-specific reagents indicated the possibility of an essential histidyl group. A thiol-modifying reagent did not cause inactivation of the enzyme. A role for the histidyl side chain in catalysis is proposed.  相似文献   

4.
The light absorption of ozone in an air stream allowed the monitoring of reactions of ozone with unsaturated fatty acids in solution. The kinetics for the reaction of ozone with linolenic acid was found to be of a pseudo-first-order after the first few minutes and did not vary with the concentration of ozone introduced into the solution. The reaction of ozone with linolenic acid in solution was found to be exceedingly rapid.When various combinations of polyunsaturated fatty acids were injected simultaneously, they reacted independently. The stoichiometry of ozone reacted to number of double bonds present in the fatty acid was one for mono- and diunsaturated; however, for triunsaturated fatty acid the stoichiometry was about 0.70.Malondialdehyde was produced upon the reaction of ozone with di- and triunsaturated fatty acids, as shown by both the thiobarbituric acid test and the characteristic UV absorption of malondialdehyde in solution. The true yield of malondialdehyde for the reaction of polyunsaturated fatty acids with ozone was found to be about 2%. In addition, other species, absorbing at 290–300 nm, were formed in solution during ozonolysis.  相似文献   

5.
Synopsis Serial transverse sections of porcine longissimus dorsi muscle (18 pigs, 50 to 178 kg live weight) were reacted for NAD tetrazolium reductase and ATPase at pH9.4, and for glycogen with the periodic acid-Schiff (PAS) reaction. Three histochemical types of muscle fibre were identified; (1) strong ATPase and weak NADH oxidative activity; (2) strong ATPase and intermediate NADH oxidative activity; and (3) weak ATPase and strong NADH oxidative activity. Immediatepost mortem samples from one side of each animal were compared with a laterpost mortem sample from the other side by measuring the absorbance of PAS-stained glycogen at 570 nm with a microscope photometer. Laterpost mortem absorbance was expressed as a percentage of immediatepost mortem absorbance in each category of muscle fibre in order to compensate for distributional error and different starting levels of glycogen. Muscle fibres with weak ATPase and strong NADH oxidative activity showed a progressive decrease in absorbance of PAS-stained glycogenpost mortem. In some animals, fibres with strong ATPase and intermediate or weak NADH oxidative activity showed an initialpost mortem increase in absorbance of PAS-stained glycogen which was then followed by a progressive decrease. The maximum rates of decrease in absorbance in the three fibre types did not differ to any great extent.  相似文献   

6.
The formation of binary complexes between sturgeon apoglyceralddhyde-3-phosphate dehydrogenase, coenzymes (NAD+ and NADH) and substrates (phosphate, glyceraldehyde 3-phosphate and 1,3-bisphosphoglycerate) has been studied spectrophotometrically and spectrofluorometrica-ly. Coenzyme binding to the apoenzyme can be characterized by several distinct spectroscopic properties: (a) the low intensity absorption band centered at 360 nm which is specific of NAD+ binding (Racker band); (b) the quenching of the enzyme fluorescence upon coenzyme binding; (c) the quenching of the fluorescence of the dihydronicotinamide moiety of the reduced coenzyme (NADH); (D) the hypochromicity and the red shift of the absorption band of NADH centered at 338 nm; (e) the coenzyme-induced difference spectra in the enzyme absorbance region. The analysis of these spectroscopic properties shows that up to four molecules of coenzyme are bound per molecule of enzyme tetramer. In every case, each successively bound coenzyme molecule contributes identically to the total observed change. Two classes of binding sites are apparent at lower temperatures for NAD+ Binding. Similarly, the binding of NADH seems to involve two distinct classes of binding sites. The excitation fluorescence spectra of NADH in the binary complex shows a component centered at 260 nm as in aqueous solution. This is consistent with a "folded" conformation of the reduced coenzyme in the binary complex, contradictory to crystallographic results. Possible reasons for this discrepancy are discussed. Binding of phosphorylated substrates and orthophosphate induce similar difference spectra in the enzyme absorbance region. No anticooperativity is detectable in the binding of glyceraldehyde 3-phosphate. These results are discussed in light of recent crystallographic studies on glyceraldehyde-3-phosphate dehydrogenases.  相似文献   

7.
The vinylogue of NAD, 3-pyridylacryloamide adenine dinucleotide, was prepared from NAD and 3-pyridylacryloamide through the snake venom NADase-catalyzed transglycosidation reaction. The analog, purified by ion-exchange chromatography, was obtained in a 55% yield. The cyanide adduct and reduced form of the analog exhibited absorbance maxima at 358 nm and 378 nm, respectively, with extinction coefficients in each case being 2.3-times higher than those reported for the corresponding NAD derivatives. 3-Pyridylacryloamide adenine dinucleotide served as a coenzyme with bovine liver glutamic dehydrogenase and to a lesser extent with malate and lactate dehydrogenases. The analog was not reduced in reactions catalyzed by yeast and horse liver alcohol dehydrogenases, sheep liver sorbitol dehydrogenase, and rabbit muscle glycerophosphate dehydrogenase. Substitution of the pyridylacryloamide analogs for NAD and NADH in the assay of substrates for glutamic dehydrogenase was demonstrated.  相似文献   

8.
In the usual reaction catalyzed by D-amino acid transaminase, cleavage of the alpha-H bond is followed by the reversible transfer of the alpha-NH2 to a keto acid cosubstrate in a two-step reaction mediated by the two vitamin B6 forms pyridoxal 5'-phosphate (PLP) and pyridoxamine 5'-phosphate (PMP). We report here a reaction not on the main pathway, i.e., beta-decarboxylation of D-aspartate to D-alanine, which occurs at 0.01% the rate of the major transaminase reaction. In this reaction, beta-C-C bond cleavage of the single substrate D-aspartate occurs rather than the usual alpha-bond cleavage in the transaminase reaction. The D-alanine produced from D-aspartate slowly inhibits both transaminase and decarboxylase activities, but NADH or NADPH instantaneously prevent D-aspartate turnover and D-alanine formation, thereby protecting the enzyme against inhibition. NADH has no effect on the enzyme spectrum itself in the absence of substrates, but it acts on the enzyme.D-aspartate complex with an apparent dissociation constant of 16 microM. Equivalent concentrations of NAD or thiols have no such effect. The suppression of beta-decarboxylase activity by NADH occurs concomitant with a reduction in the 415-nm absorbance due to the PLP form of the enzyme and an increase at 330 nm due to the PMP form of the enzyme. alpha-Ketoglutarate reverses the spectral changes caused by NADH and regenerates the active PLP form of the enzyme from the PMP form with an equilibrium constant of 10 microM. In addition to its known role in shuttling electrons in oxidation-reduction reactions, the niacin derivative NADH may also function by preventing aberrant damaging reactions for some enzyme-substrate intermediates. The D-aspartate-induced effect of NADH may indicate a slow transition between protein conformational studies if the reaction catalyzed is also slow.  相似文献   

9.
Human liver aldehyde dehydrogenase has been found to be capable of hydrolyzing p-nitrophenyl esters. Esterase and dehydrogenase activities exhibited identical ion exchange and affinity properties, indicating that the same protein catalyzes both reactions. Competitive inhibition of esterase activity by glyceraldehyde and chloral hydrate furnished evidence that p-nitrophenyl acetate was hydrolyzed at the aldehyde binding site for dehydrogenase activity. Pyridine nucleotides modified esterase activity; NAD+ accelerated the rate of p-nitrophenyl acetate hydrolysis more that 5-fold, whereas NADH increased activity by a factor of 2. Activation constants of 117 muM for NAD+ and 3.5 muM for NADH were obtained from double reciprocal plots of initial rates as a function of modifier concentration at pH 7. The kinetics of activation of ester hydrolysis were consistent with random addition of pyridine nucleotide modifier and ester substrate to this enzyme.  相似文献   

10.
The kinetic mechanisms of the 2-oxoglutarate and pyruvate dehydrogenease complexes from pig heart mitochondria were studied at pH 7.5 and 25 degrees. A three-site ping-pong mechanism for the actin of both complexes was proposed on the basis of the parallel lines obtained when 1/v was plotted against 2-oxoglutarate or pyruvate concentration for various levels of CoA and a level of NAD+ near its Michaelis constant value. Rate equations were derived from the proposed mechanism. Michaelis constants for the reactants of the 2-oxoglutarate dehydrogenase complex reaction are: 2-oxoglutarate, 0.220 mM; CoA, 0.025 mM; NAD+, 0.050 mM. Those of the pyruvate dehydrogenase complex are: pyruvate, 0.015 mM; CoA, 0.021 mM; NAD+, 0.079 mM. Product inhibition studies showed that succinyl-CoA or acetyl-CoA was competitive with respect to CoA, and NADH was competitive with respect to NAD+ in both overall reactions, and that succinyl-CoA or acetyl-CoA and NADH were uncompetitive with respect to 2-oxoglutarate or pyruvate, respectively. However, noncompetitive (rather than uncompetitive) inhibition patterns were observed for succinyl-CoA or acetyl-CoA versus NAD+ and for NADH versus CoA. These results are consistent with the proposed mechanisms.  相似文献   

11.
Our previous rapid-scanning stopped-flow studies of the reaction of substrate-free cytochrome P450cam with peracids [T. Spolitak, J.H. Dawson, D.P. Ballou, J. Biol. Chem. 280 (2005) 20300-20309; J. Inorg. Biochem. 100 (2006) 2034-2044; J. Biol. Inorg. Chem. 13 (2008) 599-611] spectrally characterized compound I (ferryl iron plus a porphyrin π-cation radical (FeIVO/Por+)), Cpd ES, and Cpd II (FeIVO/Tyr or FeIVO). We now report that reactions of CYP153A6 with peracids yield all these intermediates, with kinetic profiles allowing better resolution of all forms at pH 8.0 compared to similar reactions with WT P450cam. Properties of the reactions of these higher oxidation state intermediates were determined in double-mixing experiments in which intermediates are pre-formed and ascorbate is then added. Reactions of heptane-bound CYP153A6 (pH 7.4) with mCPBA resulted in conversion of P450 to the low-spin ferric form, presumably as heptanol was formed, suggesting that CYP 153A6 is a potential biocatalyst that can use peracids with no added NAD(P)H or reducing systems for bioremediation and other industrial applications.  相似文献   

12.
A decrease in absorbance at 340 nm, at rates similar to those obtained with alcohol dehydrogenases in routine assays, occurred when NADH or NADPH was mixed with acetaldehyde that had been exposed to air for various durations. NAD(P)H was apparently oxidized by interfering substance(s) present in acetaldehyde. Reagent-grade acetaldehyde from newly opened bottles as well as acetaldehyde redistilled under strictly O2-free conditions contained minimal amounts of NAD(P)H-reacting substance(s). Redistillation under poor anaerobic conditions or in air increased the amount of NAD(P)H-reacting substance(s) in redistilled acetaldehyde. NADPH reacted at a higher rate than NADH with the interfering substance(s) in Tris-Cl buffer at pH 7.5. Also, the reaction was faster in Tris buffer than in phosphate buffer at pH 7.5. The NAD(P)H-oxidizing reaction may not be apparent when the nominal concentration of acetaldehyde used was below 5 mM, but the measured ethanol dehydrogenase activity could be significantly lower with acetaldehyde containing a measurable level of interfering substance(s). This study suggests that acetaldehyde is most easily tested with NADPH for the presence of a significant level of interfering substance(s) and that redistillation, if necessary, must be performed under strictly O2-free conditions.  相似文献   

13.
The ozone-mediated oxidation of 2'-deoxycytidine (dCyd) was investigated on the basis of final product identification. The oxidation reaction gave rise to five major modified nucleosides which were isolated and characterized on the basis of extensive 1H NMR and mass spectrometry measurements. The comparison with the current knowledge of the hydroxyl radical mediated oxidation reactions of 2'-deoxycytidine in aerated aqueous solution, indicates that the formation of ozone oxidation products may be mostly explained by the opening of the pyrimidine C5-C6 double bond. Thus, the formation of the identified products obtained by ozonolysis of 2'-deoxycytidine is accounted for by the initial generation of an ozonide.  相似文献   

14.
Catalase-peroxidases (KatG) produced by Burkholderia pseudomallei, Escherichia coli, and Mycobacterium tuberculosis catalyze the oxidation of NADH to form NAD+ and either H2O2 or superoxide radical depending on pH. The NADH oxidase reaction requires molecular oxygen, does not require hydrogen peroxide, is not inhibited by superoxide dismutase or catalase, and has a pH optimum of 8.75, clearly differentiating it from the peroxidase and catalase reactions with pH optima of 5.5 and 6.5, respectively, and from the NADH peroxidase-oxidase reaction of horseradish peroxidase. B. pseudomallei KatG has a relatively high affinity for NADH (Km=12 microm), but the oxidase reaction is slow (kcat=0.54 min(-1)) compared with the peroxidase and catalase reactions. The catalase-peroxidases also catalyze the hydrazinolysis of isonicotinic acid hydrazide (INH) in an oxygen- and H2O2-independent reaction, and KatG-dependent radical generation from a mixture of NADH and INH is two to three times faster than the combined rates of separate reactions with NADH and INH alone. The major products from the coupled reaction, identified by high pressure liquid chromatography fractionation and mass spectrometry, are NAD+ and isonicotinoyl-NAD, the activated form of isoniazid that inhibits mycolic acid synthesis in M. tuberculosis. Isonicotinoyl-NAD synthesis from a mixture of NAD+ and INH is KatG-dependent and is activated by manganese ion. M. tuberculosis KatG catalyzes isonicotinoyl-NAD formation from NAD+ and INH more efficiently than B. pseudomallei KatG.  相似文献   

15.
A conjugated enzyme system, alanine dehydrogenase (AIDH) for stereospecific reduction of pyruvate to l-alanine and glucose dehydrogenase (GDH) for regeneration of NADH, were coimmobilized in a nanofiltration membrane bioreactor (NFMBR) for the continuous production of l-alanine from pyruvate with NADH regeneration. Since pyruvate was proved to be unstable at neutral pH, it was kept under acidic conditions and supplied to NFMBR separately from the other substrates. As 0.2 m pyruvate in HCl solution (pH 4), 10 mm NAD, 0.2 m glucose, and 0.2 m NH4Cl in 0.5 m Tris buffer (pH 8) were continuously supplied to NFMBR with immobilized AIDH (100 U/ml) and GDH (140 U/ml) at the retention time of 80 min, the maximum conversion, reactor productivity, and NAD regeneration number were 100%, 320 g/liter/d, and 20,000, respectively. To avoid the effect of pyruvate instability, a consecutive reaction system, lactate dehydrogenase (l-LDH) and AIDH, was also used. In this system, the l-LDH provides pyruvate, the substrate for the AIDH reaction, from l-lactate regenerating NADH simultaneously, so the pyruvate could be consumed as soon as it was produced. As 0.2 m l-lactate, 10 mm NAD, 0.2 m NH4Cl in 0.5 m Tris buffer (pH 8) were continuously supplied to NFMBR with immobilized l-LDH (100 U/ml) and AIDH (100 U/ml) at the retention time of 160 min, the maximum conversion, reactor productivity, and the NAD regeneration number were 100%, 160 g/Iiter/d, and 20,000, respectively.  相似文献   

16.
N Latruffe  Y Gaudemer 《Biochimie》1975,57(8):849-857
1. Rat liver mitochondrial D(-)-beta-hydroxybutyrate dehydrogenase (submitochondrial particles and partially purified preparation) is inhibited by some dicarboxylates, especially by malonate and succinate. The inhibition is reversible and competitive with beta-hydroxybutyrate while uncompetitive with acetoacetate, NAD and NADH: the inhibition is maximal at pH 6 and decrease with increasing pH. 2. Diethylpyrocarbonate (which reacts preferentially with histidyl residues at pH 6.6) inactivates the dehydrogenase at pH 6.1, beta-hydroxybutyrate protects against inactivation, this inactivation being almost completely released by hydroxylamine. The diethylpyrocarbonate-treated enzyme shows an absorbance increase at 242 nm which is characterisitic of reaction between diethylpyrocarbonate and histidyl residue. 3. The optimum pH of the enzyme for beta-hydroxybutyrate oxidation is around 8.2, while for acetoacetate reduction, the optimum pH is around 7. 4. All these results favour the existence of a histidyl residue in the catalytic center and taking into account previous results concerning the effect of thiol reagents on the same enzyme and especially, the protective effect of NAD+ and NADH against these reagents [11] we discuss the possible occurrence of, at least, one histidyl and one cysteyl residue on the catalytic center.  相似文献   

17.
Treatment of submitochondrial particles (ETP) with trypsin at 0 degrees destroyed NADPH leads to NAD (or 3-acetylpyridine adenine dinucleotide, AcPyAD) transhydrogenase activity. NADH oxidase activity was unaffected; NADPH oxidase and NADH leads to AcPyAD transhydrogenase activities were diminished by less than 10%. When ETP was incubated with trypsin at 30 degrees, NADPH leads to NAD transhydrogenase activity was rapidly lost, NADPH oxidase activity was slowly destroyed, but NADH oxidase activity remained intact. The reduction pattern by NADPH, NADPH + NAD, and NADH of chromophores absorbing at 475 minus 510 nm (flavin and iron-sulfur centers) in complex I (NADH-ubiquinone reductase) or ETP treated with trypsin at 0 degrees also indicated specific destruction of transhydrogenase activity. The sensitivity of the NADPH leads to NAD transhydrogenase reaction to trypsin suggested the involvement of susceptible arginyl residues in the enzyme. Arginyl residues are considered to be positively charged binding sites for anionic substrates and ligands in many enzymes. Treatment of ETP with the specific arginine-binding reagent, butanedione, inhibited transhydrogenation from NADPH leads to NAD (or AcPyAD). It had no effect on NADH oxidation, and inhibited NADPH oxidation and NADH leads to AcPyAD transhydrogenation by only 10 to 15% even after 30 to 60 min incubation of ETP with butanedione. The inhibition of NADPH leads to NAD transhydrogenation was diminished considerably when butanedione was added to ETP in the presence of NAD or NADP. When both NAD and NADP were present, the butanedione effect was completely abolished, thus suggesting the possible presence of arginyl residues at the nucleotide binding site of the NADPH leads to NAD transhydrogenase enzyme. Under conditions that transhydrogenation from NADPH to NAD was completely inhibited by trypsin or butanedione, NADPH oxidation rate was larger than or equal to 220 nmol min-1 mg-1 ETP protein at pH 6.0 and 30 degrees. The above results establish that in the respiratory chain of beef-heart mitochondria NADH oxidation, NADPH oxidation, and NADPH leads to NAD transhydrogenation are independent reactions.  相似文献   

18.
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   

19.
R E Johnson  J A Rupley 《Biochemistry》1979,18(16):3611-3616
The association reactions of NADH and NAD+ with dimeric pig heart supernatant malate dehydrogenase (s-MDH) have been measured at pH 6 and 8 by calorimetric and fluorescence methods, and the thermodynamic parameters describing these reactions have been evaluated. Coenzyme binding is associated with the uptake of 0.55 mol of H+/mol of NADH at pH 8 and 0.19 mol of H+ at pH 6. No significant effect of NAD+ binding on proton binding was observed. Increase in ionic strength strongly affects the free energies of binding of NAD+ and NADH. No cooperativity was observed in the enthalpy or free energy changes for binding of NAD+ or NADH. The differences in free energy of binding of NAD+ and NADH and the effect of pH on binding of NADH are entropy based. These effects are interpreted as reflecting a small number of interactions within the active site that are predominantly ionic.  相似文献   

20.
A simple rate equation for alcohol dehydrogenase was obtained by assuming independent binding sites for ethanol and NAD+ and fully competitive inhibition by the products of the reaction, acetaldehyde and NADH. A random binding order was also assumed. The rate equation is described by six parameters: four association constants (two for the substrates and two for the products of the reaction), Vf for the forward direction, and the equilibrium constant of the reaction. The six parameters were determined at pH 7.4 by numerical analysis of progress curves of reactions started with different concentrations of ethanol and NAD+. The parameters for alcohol dehydrogenase partially purified from rat liver were: Km for ethanol = 0.746 mM, Km for NAD+ = 0.0563 mM, Km for acetaldehyde = 7.07 microM, Km for NADH = 4.77 microM and Keq = 2.36 X 10(-4). The computed values allowed a very good simulation of the experimental progress curves and little variation was observed in the kinetic parameters when the reactions were started in the presence of either NADH or acetaldehyde.  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