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 312 毫秒
1.
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   

2.
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   

3.
4.
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   

5.
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   

6.
7.
The liver microsomal enzyme system that catalyzes the oxidation of NADPH by organic hydroperoxides has been solubilized and resolved by the use of detergents into fractions containing NADPH-cytochrome c reductase, cytochrome P-450 (or P-448), and microsomal lipid. Partially purified cytochromes P-450 and P-448, free of the reductase and of cytochrome b5, were prepared from liver microsomes of rats pretreated with phenobarbital (PB) and 3-methylcholanthrene (3-MC), respectively, and reconstituted separately with the reductase and lipid fractions prepared from PB-treated animals to yield enzymically active preparations functional in cumene hydroperoxide-dependent NADPH oxidation. The reductase, cytochrome P-450 (or P-448), and lipid fractions were all required for maximal catalytic activity. Detergent-purified cytochrome b5 when added to the complete system did not enhance the reaction rate. However, the partially purified cytochrome P-450 (or P-448) preparation was by itself capable of supporting the NADPH-peroxidase reaction but at a lower rate (25% of the maximal velocity) than the complete system. Other heme compounds such as hematin, methemoglobin, metmyoglobin, and ferricytochrome c could also act as comparable catalysts for the peroxidation of NADPH by cumene hydroperoxide and in these reactions, NADH was able to substitute for NADPH. The microsomal NADH-dependent peroxidase activity was also reconstituted from solubilized components of liver microsomes and was found to require NADH-cytochrome b5 reductase, cytochrome P-450 (or P-448), lipid, and cytochrome b5 for maximal catalytic activity. These results lend support to our earlier hypothesis that two distinct electron transport pathways operate in NADPH- and NADH-dependent hydroperoxide decomposition in liver microsomes.  相似文献   

8.
The proton magnetic resonance spectra of the dihydronicotinamide ring of αNADH3 and the nicotinamide ring of αNAD+ are reported and the proton absorptions assigned. The absolute assignment of the C4 methylene protons of αNADH is based on the generation of specifically deuterium-labeled (pro-S) B-deuterio-αNADH from enzymatically prepared B-deuterio-βNADH. The C4 proton absorption of αNAD+ is assigned by oxidation of B-deuterio-αNADH by the A specific, yeast alcohol dehydrogenase to yield 4-deuterio-αNAD+.The epimerization of either αNADH or βNADH yields an equilibrium ratio of approximately 9:1 βNADH to αNADH. The rate of epimerization of αNADH to βNADH at 38 °C in 0.05, pH 7.5, phosphate buffer is 3.1 × 10?3 min?1, corresponding to a half-life of 4 hr. Four related dehydrogenases, yeast and horse liver alcohol dehydrogenase and chicken M4 and H4 lactate dehydrogenase, are shown to oxidize αNADH to αNAD+ at rates three to four orders of magnitude slower than for βNADH. By using specifically labeled B-deuterio-αNADH the enzymatic oxidation by yeast alcohol dehydrogenase has been shown to occur with the identical stereospecificity as the oxidation of βNADH. The nonenzymatic epimerization of αNADH to βNADH and the enzymatic oxidation αNADH are discussed as a possible source of αNAD+in vivo.  相似文献   

9.
10.
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   

11.
S L Johnson  P T Tuazon 《Biochemistry》1977,16(6):1175-1183
The rate of the primary acid modification reaction of 1,4-dihydronicotinamide adenine dinucleotide (NADH) and 1,4-dihydro-3-acetylpyridine adenine dinucleotide (APADH) and their analogues has been studied over a wide pH range (pH 1-7) with a variety of general acid catalysts. The rate depends on [H+] at moderate pH and becomes independent of [H+] at low pH. This behavior is attributed to substrate protonation at the carbonyl group (pK of NADH = 0.6). The reaction is general acid catalyzed; large solvent deuterium isotope effects are observed for the general acid and lyonium ion terms. Most buffers cause a linear rate increase with increasing buffer concentration, but certain buffers cause a hyperbolic rate increase. The nonlinear buffer effects are due to complexation of the buffer with the substrate, rather than to a change in rate-limiting step. The rate-limiting step is a proton transfer from the general acid species to the C5 position of the substrate. Anomerization is not a necessary first step in the case of the primary acid modification reaction of beta-NADH, in which beta to alpha anomerization takes place.  相似文献   

12.
Mitochondria isolated from various plant tissues (leaves, etiolated shoots and hypocotyls, and stem tubers) oxidize exogenous NADPH with respiratory control values and ADP:O ratios similar to those obtained with exogenous NADH as substrate. In all the mitochondria investigated, the electron-transfer inhibitors rotenone and amytal each had the same effect on the oxidation of NADPH as they had on the oxidation of NADH. The oxidation of exogenous NADPH by white potato tuber mitochondria was much more sensitive to inhibition by citrate or ethylene glycol bis-(beta-aminoethyl ether)-N,N-tetraacetic acid than was the oxidation of NADH. Mitochondria isolated from aged beetroot slices showed an increased capacity for the oxidation of exogenous NADH (compared with mitochondria from fresh tissue) but no such increase in the capacity to oxidize exogenous NADPH. These results suggest that exogenous NADPH and NADH are oxidized via different flavoproteins in plant mitochondria.  相似文献   

13.
14.
15.
We have measured the Raman spectra of oxidized nicotinamide adenine dinucleotide, NAD+, and its reduced form, NADH, as well as a series of fragments and analogues of NAD+ and NADH. In addition, we have studied the effects of pH as well as deuteration of the exchangeable protons on the Raman spectra of these molecules. In comparing the positions and intensities of the peaks in the fragment and analogue spectra with those of NADH and NAD+, we find that it is useful to consider these large molecules as consisting of component parts, namely, adenine, two ribose groups, two phosphate groups, and nicotinamide, for the purpose of assigning their spectral features. The Raman bands of NADH and NAD+ are found generally to arise from molecular motions in one or another of these molecular moieties, although some peaks are not quite so easily identified in this way. This type of assignment is the first step in a detailed understanding of the Raman spectra of NAD+ and NADH. This is needed to understand the binding properties of NADH and NAD+ acting as coenzymes with the NAD-linked dehydrogenases as deduced recently by using Raman spectroscopy.  相似文献   

16.
The kinetics of alpha-NADH-dichlorophenolindophenol (DCPIP) and alpha-NADH-cytochrome c reductase reactions of rat liver microsomes showed that the reactio ns proceeded by a ping-pong mechanism, and that the oxidation of alpha-NADH was the rate-determining reaction. The DCPIP-reducing activity with alpha-NADH in the presence of ADP was about 1% of that with beta-NADH. ADP inhibited the alpha-NADH-DCPIP reductase reaction in a competitive manner with respect to alpha-NADH and a value of 1.2 mM for the inhibition constant was obtained. ADP also inhibited cytochrome b5 reduction with alpha-NADH. More than 90% of cytochrome b5 was reduced under conditions where 90% of the alpha-NADH-DCPIP reductase activity was suppressed with ADP. The reduction of DCPIP with alpha-NADH preceded that of cytochrome b5, but the reductions partly overlapped. From these results, a diversed electron flow from alpha-NADH to cytochrome b5 and electron sharing between cytochrome b5 and DCPIP were indicated. alpha-NAD+ also inhibited the alpha-NADH-DCPIP reductase reaction. Analyses of the inhibition indicated that two types of alpha-NADH-DCPIP reductase reaction existed, one of which was resistant to alpha-NAD+ inhibition. In contrast to the reoxidation of beta-NADH-reduced cytochrome b5, the process was largely monophasic when cytochrome b5 was reduced with alpha-NADH.  相似文献   

17.
L Kato  M Ishaque 《Cytobios》1975,12(45):31-43
Particles from Mycobacterium lepraemurium catalysed the oxidation of NADH with oxygen as the terminal electron acceptor. The preparations contained cytochromes of the a + a3'b and c types, as well as CO-binding pigments. The NADH oxidase activity was sensitive to inhibitors of the flavoprotein system as well as to HQNO and antimycin A. In addition, a cytochrome oxidase sensitive to cyanide was also present. The system was inhibited by the thiol-binding agent, PCMB, and thus indicated the involvement of sulphydryl group in the enzymatic oxidation of NADH. The sensitivity of the NADH oxidase system to all the inhibitors of the respiratory chain and the effect of these inhibitors on the absorption spectra suggested that cytochromes of the b, c, a + a3 types are involved in the transfer of electrons in NADH oxidation.  相似文献   

18.
Circular dichroism (CD) spectra are reported for various concentrations of glutamate dehydrogenase in order to determine any role of protein aggregation on NADH-binding spectra. These CD spectra do appear to be sensitive to enzyme aggregation. These spectra raise some doubt about previous interpretation of CD spectra as direct evidence for a second NADH binding-site.  相似文献   

19.
NADH-cytochrome b5 reductase is an amphiphilic protein consisting of a hydrophilic (FAD-containing) moiety and a hydrophobic (membrane-binding) segment and exists in aqueous media as an oligomeric aggregate. Circular dichroism studies have shown that denaturation of the reductase by guanidine hydrochloride in the presence of Emulgen 109P, a nonionic detergent, is a two-stage process as a function of the denaturant concentration. The first transition occurs at about 1 m guanidine hydrochloride and the second one at much higher concentrations. The guanidine hydrochloride concentration causing the second-stage unfolding depends on the concentration of Emulgen 109P. A hydrophilic fragment of the reductase lacking the hydrophobic segment undergoes one-stage denaturation at about 1 m guandine hydrochloride regardless of the presence and absence of Emulgen 109P. Both the reductase as well as the hydrophilic fragment lose their NADH-ferricyanide reductase activity and FAD also at about 1 m guanidine hydrochloride in the presence of the detergent. These findings suggest that the first-stage denaturation of the reductase represents the unfolding of the hydrophilic moiety and the second one that of the hydrophobic segment. Gel chromatography experiments have suggested that in the presence of Emulgen 109P the reductase exists as a mixed micelle with the detergent and this aggregation state persists even after the first-stage denaturation (unfolding of the hydrophilic moiety). The dissociation of the mixed micelle seems to take place concomitant with the second-stage denaturation. It is concluded that the two moieties of the reductase molecule, though linked to each other covalently, exist as independent domains undergoing unfolding separately at least in the presence of Emulgen 109P. This structural feature of the reductase is similar to that of cytochrome b5 reported by us. The reductase is, therefore, another example of amphiphilic membrane proteins having two structurally independent domains in the molecule.  相似文献   

20.
R Koren  G G Hammes 《Biochemistry》1975,14(5):1021-1025
The interaction of NADH with s-malate dehydrogenase isolated from beef heart was studied in 20 mM potassium phosphate (pH 6.9)-1 mM EDTA, with forced dialysis, fluorescence, and temperature-jump techniques. Measurements of the change in fluorescence of NADH when it is titrated with enzyme indicate NADH bound to monomeric and dimeric enzyme have different fluorescence yields. These data and the results of direct binding studies can be explained in terms of a model in which the NADH binding sites on dimeric enzyme are equivalent or nearly equivalent, and NADH binding to monomeric enzyme occurs with an affinity very similar to that of the dimer. However, the fluorescence enhancement of NADH on binding to the enzyme is different for the monomer and for each of the two dimer sites.  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