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 0 毫秒
1.
K Yokota  I Yamazaki 《Biochemistry》1977,16(9):1913-1920
Under suitable experimental conditions the aerobic oxidation of NADH catalyzed by horseradish peroxidase occurred in four characteristic phases: initial burst, induction phase, steady state, and termination. A trace amount of H2O2 present in the NADH solution brought about initial burst in the formation of oxyperoxidase. About 2 mol of oxyperoxidase was formed per mol of H2O2. When a considerable amount of the ferric enzyme still remained, the initial burst was followed by an induction phase. In this phase the rate of oxyperoxidase formation from the ferric enzyme increased with the decrease of the ferric enzyme and an approximately exponential increase of oxyperoxidase was observed. A rapid oxidation of NADH suddenly began at the end of the induction phase and the oxidation continued at a relatively constant rate. In the steady state, oxygen was consumed and H2O2 accumulated. A drastic terminating reaction suddenly set in when the oxygen concentration decreased under a certain level. During the reaction, H2O2 disappeared accompanying an accelerated oxidation of NADH and the enzyme returned to the ferric form after a transient increase of peroxidase compound II. Time courses of NADH oxidation, O2 consumption, H2O2 accumulation, and formation of enzyme intermediates could be simulated with an electronic computer using 11 elementary reactions and 9 rate equations. The results were also discussed in relation to the mechanism for oscillatory responses of the reaction that appeared in an open system with a continuous supply of oxygen.  相似文献   

2.
3.
The kinetics of p-aminobenzoic acid oxidation catalyzed by horseradish peroxidase Compounds I and II was investigated intensively as a function of pH at 25 degrees in aqueous solutions of ionic strength 0.11. All of the rate data were collected from single turnover experiments involving reactions of a single enzyme compound. In reactions of both compounds, deviations from first order behavior with respect to the enzyme were observed at high pH values which were explained in terms of a free radical interaction of product with the enzyme. The effect could be eliminated with sufficient excess of substrate. Kinetic behavior which deviated from first order in substrate, observed at low pH, was explained by a mechanism involving an enzyme-substrate complex which reacted with an additional molecule of substrate but at a slower rate. The pH dependence of the second order rate constants for the reaction of p-aminobenzoic acid with free Compounds I and II is similar to results obtained for the comparable reactions of ferrocyanide, suggesting similar proton-transfer mechanisms for both reducing substrates. The reduction of Compound II by p-aminobenzoic acid appeared to be influenced by two ionizable groups on the enzyme which affect the electronic environment of the heme. The lack of influence of substrate ionizable groups on the rate of the Compound II reaction indicated that potential differences in reactivities of NH2C6H4COO- and NH2C6H4COOH were levelled by the diffusion-controlled limit in the acid region of pH. The reduction of Compound I by p-aminobenzoic acid was not diffusion-controlled and the rate-pH profile could be explained in terms of three acid ionizations, two on the substrate and one on Compound I.  相似文献   

4.
5.
Flagyl and reduced nicotinamide adenine dinucleotide   总被引:1,自引:0,他引:1  
G H Coombs  B R Rabin 《FEBS letters》1974,42(1):105-107
  相似文献   

6.
Phototransformation of phytochrome in lettuce seeds (Lactuca sativa L. var. Grand Rapids) was examined by testing germination responses of seeds irradiated at various temperatures. Temperature variations from 0 to 50 C had no influence on the germination of partially hydrated seeds (about 15% water content) irradiated with either red or far red light prior to imbibition. At −15 C far red light more effectively retarded germination than red light promoted it. No effective phototransformation was detected at −79 C or −196 C.  相似文献   

7.
Oxidation of reduced nicotinamide adenine dinucleotide by melanin   总被引:1,自引:0,他引:1  
M H Van Woert 《Life sciences》1967,6(24):2605-2612
  相似文献   

8.
Exogenous NADH oxidation by cauliflower (Brassica oleracea L.) bud mitochondria was sensitive to antimycin A and gave ADP/O ratios of 1.4 to 1.9. In intact mitochondria, NADH-cytochrome c reductase activity was only slightly inhibited by antimycin A. The antimycin-insensitive activity was associated with the outer membrane. Malate oxidation was sensitive to both rotenone and antimycin A and gave ADP/O values of 2.4 to 2.9. However in the presence of added NAD+, malate oxidation displayed similar properties to exogenous NADH oxidation. In both the presence and absence of added NAD+, malate oxidation was dependent on inorganic phosphate and inhibited by 2-n-butyl malonate.  相似文献   

9.
Formation of reduced nicotinamide adenine dinucleotide peroxide   总被引:1,自引:0,他引:1  
Incubation of NADH at neutral and slightly alkaline pH leads to the gradual absorption of 1 mol of H+. This uptake of acid requires oxygen and mainly yields anomerized NAD+ (NAD+), with only minimal formation od acid-modified NADH. The overall stoichiometry of the reaction is: NADH + H+ + 1/2O2 leads to H2O + NAD+, with NADH peroxide (HO2-NADH+) serving as the intermediate that anomerizes and breaks down to give NAD+ and H2O2. The final reaction reaction mixture contains less than 0.1% of the generated H2O2, which is nonenzymically reduced by NADH. The latter reaction is inhibited by catalase, leading to a decrease in the overall rate of acid absorption, and stimulated by peroxidase, leading to an increase in the overall rate of acid absorption. Although oxygen can attack NADH at either N-1 or C-5 of the dihydropyridine ring, the attack appears to occur primarily at N-1. This assignment is based on the inability of the C-5 peroxide to anomerize, whereas the N-1 peroxide, being a quaternary pyridinium compound, can anomerize via reversible dissociation of H2O2. The peroxidase-catalyzed oxidation of NADH by H2O2 does not lead to anomerization, indicating that anomerization occurs prior to the release of H2O2. Chromatography of reaction mixtures on Dowex 1 formate shows the presence of two major and several minor neutral and cationic degradation products. One of the major products is nicotinamide, which possibly arises from breakdown of nicotinamide-1-peroxide. The other products have not been identified, but may be derived from other isomeric nicotinamide peroxides.  相似文献   

10.
11.
NAD+ had a biphasic effect on the NADH oxidase activity in electron transport particles from Mycobacterium phlei. The oxidase was inhibited competitively by NAD+ at concentrations above 0.05 mM. NAD+ in concentrations from 0.02 to 0.05 mM resulted in maximum stimulation of both NADH oxidation and oxygen uptake with concentrations of substrate both above and below the apparent K-M. Oxygen uptake and cyanide sensitivity indicated that the NAD+ stimulatory effect was linked to the terminal respiratory chain. The stimulatory effect was specific for NAD+. NAD+ was also specific in protecting the oxidase during heating at 50 degrees and against inactivation during storage at 0 degrees. NAD+ glycohydrolase did not affect stimulation nor heat protection of the NADH oxidase activity if the particles were previously preincubated with NAD+. Binding studies revealed that the particles bound approximately 3.6 pmol of [14C1NAD+ per mg of electron transport particle protein. Although bound NAD+ represented only a small fraction of the total added NAD+ necessary for maximal stimulation, removal of the apparently unbound NAD+ by Sephadex chromatography revealed that particles retained the stimulated state for at least 48 hours. Further addition of NAD+ to stimulated washed particles resulted in competitive inhibition of oxidase activity. Desensitization of the oxidase to the stimulatory effect of NAD+ was achieved by heating the particles at 50 degrees for 2 min without appreciable loss of enzymatic activity. Kinetic studies indicated that addition of NADH to electron transport particles prior to preincubation with NAD+ inhibited stimulation. In addition, NADH inhibited binding of [14C]NAD+. The utilization of artificial electron acceptors, which act as a shunt of the respiratory chain at or near the flavoprotein component, indicated that NAD+ acts as at the level of the NADH dehydrogenase at a site other than the catalytic one resulting in a conformational change which causes restoration as well as protection of oxidase activity.  相似文献   

12.
13.
14.
15.
The enzyme rhodanese (thiosulfate sulfurtransferase; EC 2.8.1.1) is inactivated with a half-time of approximately 3 min when incubated with 50 mM NADH. NAD+, however, has virtually no effect on the activity. Inactivation can be prevented by the inclusion of the substrate thiosulfate. The concentration of thiosulfate giving half-protection is 0.038 mM. In addition, NADH, but not NAD+, is a competitive inhibitor with respect to thiosulfate in the catalyzed reaction (Ki = 8.3 mM). Fluorescence studies are consistent with a time-dependent oxidation of NADH in the presence of rhodanese. The sulfur-free form of rhodanese is more rapidly inactivated than the sulfur-containing form. Spectrophotometric titrations show that inactivation is accompanied by the loss of two free SH groups per enzyme molecule. Inactivation is prevented by the exclusion of air and the inclusion of EDTA (1 mM), and the enzyme activity can be largely protected by incubation with superoxide dismutase or catalase. Rhodanese, inactivated with NADH, can be reactivated by incubation with the substrate thiosulfate (75 mM) for 48 h or more rapidly, but only partially, by incubating with 180 mM dithiothreitol. It is concluded that, in the presence of rhodanese, NADH can be oxidized by molecular oxygen and produce intermediates of oxygen reduction, such as superoxide and/or hydrogen peroxide, that can inactivate the enzyme with consequent formation of an intraprotein disulfide. In addition, NADH, but not NAD+, can reversibly bind to the active site region in competition with thiosulfate. These data are of interest in view of x-ray studies that show structural similarities between rhodanese and nucleotide binding proteins.  相似文献   

16.
17.
18.
The binding of NADH to bull semen NAD nucleosidase was observed to be accompanied by a considerable enhancement of the fluorescence of NADH. The fluorescence enhancement observed in the binding of NADH to the enzyme was utilized to study the stoichiometry of binding of this compound to the enzyme. Results obtained from the fluorescence titration of the enzyme with NADH indicated the binding of one mole of NADH per mole of enzyme (36,000 g). The dissociation constant for the enzyme-NADH complex was determined to be 2.52 × 10?6m. NADH was also found to be a very effective competitive inhibitor of the NADase-catalyzed hydrolysis of NAD, and the inhibitor dissociation constant (KI) for the enzyme-NADH complex was determined to be 2.99 × 10?6m which was in good agreement with the value obtained from the fluorescence titration experiments.  相似文献   

19.
M A Vanoni  R G Matthews 《Biochemistry》1984,23(22):5272-5279
Previous work from this laboratory has established that the NADPH-menadione oxidoreductase reaction catalyzed by methylenetetrahydrofolate reductase from pig liver proceeds by Ping Pong Bi Bi kinetics and that the reductive half-reaction is rate limiting in steady-state turnover. We have now shown that methylenetetrahydrofolate reductase stereo-specifically removes the pro-S hydrogen from the 4-position of NADPH. During the oxidation of [4(S)-3H]NADPH, we observed a kinetic isotope on V/KNADPH of 10.8 +/- 0.4. When comparing the rates of oxidation of [4(S)-2H]NADPH and [4(S)-1H]NADPH, we measure kinetic isotope effects on V of 4.78 +/- 0.15 and on V/KNADPH of 4.54 +/- 0.59. When oxidation of [4(R)-2H]NADPH and [4(R)-1H]NADPH is compared, the secondary kinetic isotope effect on V is 1.04 +/- 0.01. When the NADPH-menadione oxidoreductase reaction is catalyzed in tritiated water, no incorporation of solvent tritium into residual NADPH is observed. We conclude from these observations that the oxidation of NADPH is largely or entirely rate limiting in the reductive half-reaction and, hence, in NADPH-menadione oxidoreductase turnover at saturating menadione concentration. In the presence of saturating NADPH, the flavin reduction proceeds with a rate constant of 160 S-1, which is at least 29-fold slower than estimates of the lower limit for the diffusion-limited rate constant characterizing NADPH binding to the enzyme under physiological conditions. Albery & Knowles have defined criteria for perfection in enzyme catalysis [Albery, W. J., & Knowles, J.R. (1976) Biochemistry 15, 5631-5640].(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
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