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.
Progress in understanding the role of NAD(P)H oxidation in plant respiration is restricted by the lack of access to specific inhibitors of each of the unknown number of NAD(P)H dehydrogenases in the inner mitochondrial membrane. Platanetin (3,5,7,8-tetrahydroxy-6-isoprenyl flavone) is known to be an inhibitor of extermal NADH oxidation by plant mitochondria, while 7-iodo-acridone-4-carboxylic acid (IACA) is an inhibitor of an internal, rotenone-insensitive NAD(P)H dehydrogenase isolated from yeast mitochondria.
Here we show that platanetin inhibits external NAD(P)H oxidation by intact potato ( Solanum tuberosum L. cv. Bintje) tuber mitochondria, deamino-NADH oxidation by Complex I assayed using inside-out submitochondrial particles from these mitochondria, and rotenone-insensitive NAD(P)H oxidation by these submitochondrial particles. IACA was found to inhibit the oxidation of external NADH and succinate by intact mitochondria with similar efficiency. However, IACA also inhibited NADPH and duroquinol oxidation by intact mitochondria as well as deamino-NADH and NAD(P)H oxidation by inside-out submitochondrial particles. This indicates that IACA has several sites of inhibition in the electron transport chain. The lack of specificity of both platanetin and IACA prevents these inhibitors from being used to shed more light on the identity of the NAD(P)H dehydrogenases in plant mitochondria.  相似文献   

2.
Submitochondrial particles (SMP) were isolated from potato ( Solanum tuberosum L. cv. Bintje) tubers. The SMP were 91% inside-out and they were able to form a membrane potential, as monitored by oxonol VI, with succinate, NADH and NADPH. The pH dependence and kinetics of NADH and NADPH oxidation by these SMP was studied using three different electron acceptors – O2, duroquinone and ferricyanide. In addition, the SMP were solubilized, fractionated by non-denaturing polyacrylamide gel electrophoresis, and the gels were stained for NAD(P)H dehydrogenase activity and specificity at different pH using Nitro Blue Tetrazolium. From the results we conclude that there are at least two distinct NAD(P)H dehydrogenases on the inner surface of the inner membrane: (1) Complex 1 which oxidizes NADH and deamino-NADH in a rotenone-sensitive manner, (O2 as acceptor) with optimum activity at pH 8 and a very low Km(NADH) of 3 μ M . It also oxidizes NADPH and deamino-NADPH in a rotenone-sensitive manner, but with a pH optimum at pH 5.8 and a very high Km(NADPH) of more than 1 m M . This complex is found as a broad, diffuse band at the top of the gels. (2) A second dehydrogenase which oxidizes NADH in a rotenone-insensitive manner with optimum activity at pH 6.2 and a higher Km(NADH) of 14 μ M . It also oxidizes NADPH in a rotenone-insensitive manner with an activity optimum at pH 6.8 and low Km(NADPH) of 25 μ M . This dehydrogenase does not oxidize deamino-NAD(P)H. One of the sharp bands around the middle of the native gels may be caused by this dehydrogenase indicating that it has a relatively low molecular mass compared to Complex I. Several other NAD(P)H dehydrogenase bands were observed on the gels which we cannot yet assign.  相似文献   

3.
The addition of ubiquinone-1 (UQ-1) induced Ca2+-independent oxidation of deamino-NADH and NADH by intact potato (Solanum tuberosum L. cv Bintje) tuber mitochondria. The induced oxidation was coupled to the generation of a membrane potential. Measurements of NAD+-malate dehydrogenase activity indicated that the permeability of the inner mitochondrial membrane to NADH and deamino-NADH was not altered by the addition of UQ-1. We conclude that UQ-1-induced external deamino-NADH oxidation is due to a change in specificity of the external rotenone-insensitive NADH dehydrogenase. The addition of UQ-1 also induced rotenone-insensitive oxidation of deamino-NADH by inside-out submitochondrial particles, but whether this was due to a change in the specificity of the internal rotenone-insensitive NAD(P)H dehydrogenase or to a bypass in complex I could not be determined.  相似文献   

4.
Hepatocyte cytotoxicity caused by substituted benzoquinones was associated with increased cytosolic Ca2+ concentration. p-Benzoquinone-induced hepatotoxicity was enhanced when the hepatocytes were loaded with Ca2+ by preincubation with ATP. A similar order of potency of the substituted benzoquinones in releasing Ca2+ from isolated mitochondria and inducing hepatocyte cytotoxicity was found; in decreasing order, this was 2-Br-, unsubstituted-, 2-CH3-, 2,6-(CH3O)2-, 2,6-(CH3)2-, 2,5-(CH3)2-, 2,3,5-(CH3)3-, and 2,3,5,6-(CH3)4-benzoquinones (duroquinone). The cellular products of quinone metabolism, hydroquinones and glutathione conjugates, did not cause mitochondrial Ca2+ release. Benzoquinone-induced mitochondrial Ca2+ release was preceded by GSH conjugate formation and NAD(P)H oxidation but followed by mitochondrial swelling. With duroquinone, a slow GSH and NADPH oxidation preceded Ca2+ release, but GSH oxidation did not occur with Se-deficient mitochondria lacking glutathione peroxidase activity. Cyanide-insensitive respiration was also observed with duroquinone but not with benzoquinone, suggesting that duroquinone undergoes redox cycling. GSH was depleted by both arylation and oxidation with 2,6-(CH3O)2-, 2,6-(CH3)2-, 2,5(CH3)2-, and 2,3,5-(CH3)3-benzoquinones. Benzoquinone concentrations that totally depleted GSH did not cause Ca2+ release until intramitochondrial NAD(P)H was oxidized. Ca2+ release was also prevented when NAD(P)H generation was stimulated by the presence of isocitrate or 3-hydroxybutyrate. This suggests that mitochondrial Ca2+ release is associated with NAD(P)H oxidation catalyzed by NADH dehydrogenase with benzoquinone or by the glutathione peroxidase-glutathione reductase system with duroquinone.  相似文献   

5.
The short-term incubation of HeLa cells in the presence of diferric transferrin or ferricyanide, which are reduced externally by the transplasma membrane reductase, produces a stoichiometric decrease in NADH and increase in NAD+, which is stimulated by insulin. The NADP/NADPH ratio does not change during 15 min incubation with the oxidants. The total pyridine nucleotide pool of HeLa cells is not affected. Incubation with apotransferrin and ferrocyanide, which cannot act as oxidants for transmembrane electron transport, does not change the pyridine nucleotide concentrations in the cells. Our results show that NADH can act as the internal electron donor for the reduction of external oxidants by the transmembrane reductase. It appears that oxidation of NADH by the transmembrane electron transport using ferricyanide or iron transferrin as external electron acceptors is sufficient to stimulate growth in HeLa cells.  相似文献   

6.
The goal was to determine whether endogenous cytosolic NAD(P)H:quinone oxidoreductase 1 (NQO1) preferentially uses NADPH or NADH in intact pulmonary arterial endothelial cells in culture. The approach was to manipulate the redox status of the NADH/NAD(+) and NADPH/NADP(+) redox pairs in the cytosolic compartment using treatment conditions targeting glycolysis and the pentose phosphate pathway alone or with lactate, and to evaluate the impact on the intact cell NQO1 activity. Cells were treated with 2-deoxyglucose, iodoacetate, or epiandrosterone in the absence or presence of lactate, NQO1 activity was measured in intact cells using duroquinone as the electron acceptor, and pyridine nucleotide redox status was measured in total cell KOH extracts by high-performance liquid chromatography. 2-Deoxyglucose decreased NADH/NAD(+) and NADPH/NADP(+) ratios by 59 and 50%, respectively, and intact cell NQO1 activity by 74%; lactate restored NADH/NAD(+), but not NADPH/NADP(+) or NQO1 activity. Iodoacetate decreased NADH/NAD(+) but had no detectable effect on NADPH/NADP(+) or NQO1 activity. Epiandrosterone decreased NQO1 activity by 67%, and although epiandrosterone alone did not alter the NADPH/NADP(+) or NADH/NAD(+) ratio, when the NQO1 electron acceptor duroquinone was also present, NADPH/NADP(+) decreased by 84% with no impact on NADH/NAD(+). Duroquinone alone also decreased NADPH/NADP(+) but not NADH/NAD(+). The results suggest that NQO1 activity is more tightly coupled to the redox status of the NADPH/NADP(+) than NADH/NAD(+) redox pair, and that NADPH is the endogenous NQO1 electron donor. Parallel studies of pulmonary endothelial transplasma membrane electron transport (TPMET), another redox process that draws reducing equivalents from the cytosol, confirmed previous observations of a correlation with the NADH/NAD(+) ratio.  相似文献   

7.
Intrinsic NADPH diaphorase activity is a component of the membrane-bound NAD(P)H:O2 oxidoreductase of human neutrophils. NADH-specific diaphorase activity is also present in membrane fractions rich in oxidoreductase activity. Studies were undertaken to determine whether the NADH diaphorase might also be intrinsic to the oxidoreductase. The latter diaphorase was freed from the membrane by detergent extraction and partially purified approximately 80-fold. Its apparent molecular weight following solubilization in deoxycholate and Tween-20 was 204 000 +/- 10 000. The specific activity of the partially purified diaphorase with ferricyanide as electron acceptor was 7.6 X 10(3) mU/mg protein, its pH optimum was 7.0, and its Km for NADH was 13 microM. It is completely devoid of NADPH diaphorase activity, lacks the capacity to reduce molecular oxygen, yet readily reduces ferricyanide, 2,6-dichlorophenolindophenol and ferricytochrome c. Whereas the NADH diaphorase was freed from the particulate fraction of cell lysates by extraction in 10 mM Tris-HCl buffer (pH 8.6) made up in 15% glycerol and 0.5% Tween-20, NADPH-dependent diaphorase and superoxide-generating activities also present in the membrane were not. These observations make it unlikely that the principal membrane-bound NADH diaphorase found in human neutrophils is a component of the NAD(P)H:O2 oxidoreductase, despite its common association in the same particulate fraction of cell lysates.  相似文献   

8.
Glutathione reductase has been found to catalyze an NAD(P)H-dependent electron transfer to 2,4,6-trinitrobenzenesulfonate (TNBS). In the presence of oxygen TNBS is not consumed in the reaction, but is rapidly reoxidized with concomitant production of hydrogen peroxide. Cytochrome c can replace oxygen as the final electron acceptor, indicating that a one-electron transfer takes place. The rate is slightly higher in the absence than in the presence of oxygen, ruling out superoxide anion as an obligatory intermediate in cytochrome c reduction. In the absence of oxygen (or cytochrome c), TNBS limits the reaction and accepts a total of four electrons. The TNBS-dependent NADPH (or NADH) oxidation is markedly stimulated by NADP+, and to a smaller extent also by NAD+. The TNBS-dependent reactions are inhibited by excess of NADPH but not by NADH. The kinetics of these reactions are consistent with a branching reaction mechanism in which a pathway including a ternary complex between the two-electron reduced enzyme and NADP+ has the highest turnover. NADPH-dependent reductions of ferricyanide or 2,6-dichloroindophenol catalyzed by glutathione reductase are also markedly influenced by NADP+. Evidently NADP+ facilitates a shift of the catalyzed reaction from the normal two-electron reduction of glutathione disulfide to a more unspecific one-electron reduction of other acceptors. Spectral as well as kinetic data suggest that the rate of radical formation limits the reactions with the artificial electron acceptors and that NADP+ promotes this rate-limiting step.  相似文献   

9.
The present study compares the exogenous NAD(P)H oxidation and the membrane potential ([delta][psi]) generated in mitochondria isolated from different tissues of an important agricultural crop, sugar beet (Beta vulgaris}. We observed that mitochondria from taproots, cold-stored taproots, and in vitro-grown tissue cultures contain a functional NADH dehydrogenase, whereas only those isolated from tissue cultures displayed a functional NAD(P)H dehydrogenase. It is interesting that the NADH-dependent [delta][psi] of mitochondria from cold-stored taproots and from tissue cultures was not affected by free Ca2+ ions, whereas free Ca2+ was required for the mitochondrial NADPH oxidation by in vitro-grown cells and cytosolic NADH oxidation by mitochondria from fresh taproots. A tentative model accounting for the different response to Ca2+ ions of the NADH dehydrogenase in mitochondria from cold-stored taproots and tissue cultures of B. vulgaris is discussed.  相似文献   

10.
Electron transfer activities and steady state reduction levels of Fe-S centers of NADH-Q oxidoreductase were measured in mitochondria, submitochondrial particles (ETPH), and complex I after treatment with various reagents. p-Chloromercuribenzenesulfonate destroyed the signal from center N-4 (gx = 1.88) in ETPH but not in mitochondria, showing that N-4 is accessible only from the matrix side of the inner membrane. N-Bromosuccinimide also destroyed the signal from N-4 but without inhibiting rotenone-sensitive electron transfer to quinone, suggesting a branched pathway for electron transfer. Diethylpyrocarbonate caused oxidation of N-3 and N-4 in the steady state without changing N-1, suggesting N-1 is before N-3 and N-4. Difluorodinitrobenzene and dicyclohexylcarbodiimide inhibited oxidation of all Fe-S centers and tetranitromethane inhibited reduction of all Fe-S centers. Titrations of the rate of superoxide (O2-) generation in rotenone-treated submitochondrial particles were similar with the ratio [NADH]/[NAD] and that of 3-acetyl pyridine adenine nucleotide in spite of different midpoint potentials of the two couples. On reaction with inhibitors the inhibition of O2- formation was similar to that of ferricyanide reductase rather than quinone reductase. The rate of O2- formation during ATP-driven reverse electron transfer was 16% of the rate observed with NADH. The presence of NAD increased the rate to 83%. The results suggest that bound, reduced nucleotide, probably E-NAD., is the main source of O2- in NADH dehydrogenase. The effect of ATP on the reduction levels of Fe-S centers in well-coupled ETPH was measured by equilibrating with either NADH/NAD or succinate/fumarate redox couples. With NADH/NAD none of the Fe-S centers showed ATP induced changes, but with succinate/fumarate all centers showed ATP-driven reduction with or without NAD present. The effect on N-2 was smaller than that on N-1, N-3, and N-4. These observations indicate that the major coupling interaction is between N-2 and the low potential centers, N-1, N-3, and N-4. Possible schemes of coupling in this segment are discussed.  相似文献   

11.
Low concentrations of NAD+ inhibit the NADH: acceptor reductase reactions catalyzed by soluble NADH dehydrogenase from bovine heart mitochondria. The degree of incomplete inhibition of the enzyme depends on the nature and concentration of artificial electron acceptors and is manifested only at low concentrations of the latter. Marked inhibition was demonstrated for the 2.6-dichlorophenolindophenol-, ferricyanide- and O2-reductase reactions, being weakly pronounced during the measurement of the NADH: cytochrome c reductase activity. The inhibition of the above reactions by oxidized NAD+ isn't competitive towards NADH. A kinetic scheme is proposed, which postulates NADH: acceptor reductase reactions occurrence via two mechanisms, namely, a ping-pong mechanism and oxidation of the product-enzyme complex by the acceptor. It was shown that low concentrations of NAD+ also inhibit the NADH oxidase reaction catalyzed by complex I.  相似文献   

12.
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   

13.
Two enzyme systems carrying out the oxidation of NAD(P)H in the presence of various electron acceptors have been isolated and partially characterized from the supernatant of frozen-thawed mitochondria from Arum maculatum spadices. The two systems contain flavoproteins and differ by their ability to oxidize NADH or NADPH, optimum pH and pI values, sensitivity to Ca2+ and EGTA, denaturation by 4 molar urea, molecular mass, and number of subunits. These properties, together with methodological considerations, are compatible with the location of these enzyme activities on the outer surface of the inner mitochondrial membrane, and support the hypothesis of the existence of two separate dehydrogenases responsible for the mitochondrial oxidation of cytosolic NADH and NADPH.  相似文献   

14.
Characteristics of External NADH Oxidation by Beetroot Mitochondria   总被引:6,自引:6,他引:0       下载免费PDF全文
Mitochondria isolated from fresh red beetroot (Beta vulgaris L.) tissue do not oxidize external NADH with O2 as the electron acceptor. These mitochondria have a rotenone- and antimycin-insensitive pathway of NADH oxidation associated with the outer membrane and are capable of reducing cytochrome c or potassium ferricyanide. They are also capable of oxidizing internal NADH via the inner membrane electron transport chain with normal rotenone and antimycin sensitivity and ADP/O ratios. They differ from other plant mitochondria in the apparent lack of the NADH dehydrogenase located on the outer surface of the inner membrane. It is shown that this activity develops during the aging of red beetroot slices in aerated dilute CaSO4 solutions, and is present in the mitochondria isolated from aged tissue.  相似文献   

15.
1. Oxidation of NADPH by various acceptors catalyzed by submitochondrial particles and a partially purified NADH dehydrogenase from beef heart was investigated. Submitochondrial particles devoid of nicotinamide nucleotide transhydrogenase activity catalyze an oxidation of NADPH by oxygen. The partially purified NADH dehydrogenase prepared from these particles catalyzes an oxidation of NADPH by acetylpyridine-NAD. In both cases the rates of oxidation are about two orders of magnitude lower than those obtained with NADH as electron donor. 2. The kinetic characteristics of the NADPH oxidase reaction and reduction of acetylpyridine-NAD by NADPH are similar with regard to pH dependences and affinities for NADPH, indicating that both reactions involve the same binding site for NADPH. The binding of NADPH to this site appears to be rate limiting for the overall reactions. 3. At redox equilibrium NADPH and NADH reduce FMN and iron-sulphur center 1 of NADH dehydrogenase to the same extents. The rate of reduction of FMN by NADPH is at least two orders of magnitude lower than with NADH. 4. It is concluded that NADPH is a substrate of NADH dehydrogenase and that the nicotinamide nucleotide is oxidized by submitochondrial particles via the NADH--binding site of the enzyme.  相似文献   

16.
The involvement of pyridine nucleotides in the reduction of extracytoplasmatic electron acceptors by iron-deficient Plantago lanceolata L. roots has been examined by measuring the changes in NAD(P)H and NAD(P) induced by various external acceptors. Exposure of the plants to FeEDTA, ferricyanide, ferric citrate or hexachloroiri-date resulted in a transient decrease in NADPH and an increase in NAD. No major differences in this pattern were observed between acceptors which were assumed to be reduced by different enzymes. The application of the membrane-permeable oxidant nitro blue tetrazolium led to similar changes in reduced and oxidized pyridine nucleotides and decreased the reduction of external acceptors. The amino acid analog p -fluorophenylalanine caused a transient decline in both NADPH level and NADPH/ NADP ratio and a decrease in the ratio of NADH to NAD without affecting the level of NADH. Exposure of the plants to the translation inhibitor cycloheximide increased both NADH and NADPH concentrations. A comparison of the redox activities and pyridine nucleotide fractions after inhibitor treatment revealed that the constitutive, but not iron stress-induced redox activity correlates with NADPH levels. These results are interpreted as confirming that the redox systems on the root plasma membrane are separately regulated. Possible metabolic reactions during the reduction processes are discussed.  相似文献   

17.
Supermitochondrial liquid (SL) of rat and guinea-pig liver increases the activity of 2, 3, 5 triphenyltetrasolium chloride (TPC) and tetrasolium violet (TV) reduction at succinate, NADH and NADPH oxidation by mitochondria (MC). SL contains an activating factor A, being evidently of a protein nature and factor B, increasing the activating activity of factor A. NAD, NADP, NADH and NADP at 5 x 10(-5)-1 x 10(-4) M concentration activate the TPC and TV reduction at succinate oxidation by mitochondria. TPC and TV reduction at succinate and NADP oxidation by mitochondria makes antimicin and cyanide sensitive. SL does not influence succinate dehydrogenase activity, when used as electron acceptors of ferricyanide, blue Vurster, cytochrome C, blue and violet nitrotetrasolium. Activation of electron transfer chair between cytochrome C and oxygen is supposed to be responsible for such an effect.  相似文献   

18.
The results presented in this paper reveal the existence of three distinct menadione (2-methyl-1,4-naphthoquinone) reductases in mitochondria: NAD(P)H:(quinone-acceptor) oxidoreductase (D,T-diaphorase), NADPH:(quinone-acceptor) oxidoreductase, and NADH:(quinone-acceptor) oxidoreductase. All three enzymes reduce menadione in a two-electron step directly to the hydroquinone form. NADH-ubiquinone oxidoreductase (NADH dehydrogenase) and NAD(P)H azoreductase do not participate significantly in menadione reduction. In mitochondrial extracts, the menadione-induced NAD(P)H oxidation occurs beyond stoichiometric reduction of the quinone and is accompanied by O2 consumption. Benzoquinone is reduced more rapidly than menadione but does not undergo redox cycling. In intact mitochondria, menadione triggers oxidation of intramitochondrial pyridine nucleotides, cyanide-insensitive O2 consumption, and a transient decrease of delta psi. In the presence of intramitochondrial Ca2+, the menadione-induced oxidation of pyridine nucleotides is accompanied by their hydrolysis, and Ca2+ is released from mitochondria. The menadione-induced Ca2+ release leaves mitochondria intact, provided excessive Ca2+ cycling is prevented. In both selenium-deficient and selenium-adequate mitochondria, menadione is equally effective in inducing oxidation of pyridine nucleotides and Ca2+ release. Thus, menadione-induced Ca2+ release is mediated predominantly by enzymatic two-electron reduction of menadione, and not by H2O2 generated by menadione-dependent redox cycling. Our findings argue against D,T-diaphorase being a control device that prevents quinone-dependent oxygen toxicity in mitochondria.  相似文献   

19.
Rugolo M  Zannoni D 《Plant physiology》1992,99(3):1037-1043
The functional interaction between the externally located NAD(P)H dehydrogenase and the Q-pool acceptor site(s) in Percoll-purified mitochondria from Jerusalem artichoke (Helianthus tuberosus L. cv OB1) mitochondria has been investigated. Oxidation of exogenous NADH is stimulated by ubiquinone (UQ1) with a parallel decrease of the apparent Km for NADH. In the presence of saturating amounts of UQ1 as electron acceptor, the Km (NADH) is not affected by variations of the ionic strength. Conversely, the Km for UQ1 is decreased by the screening effect of negative charges on the outer membrane surface. Under low-ionic strength, the hydroxyflavone platanetin progressively inhibits NADH oxidation with a mean inhibition dose of approximately 3 nanomoles of inhibitor per milligram of protein. Interestingly, under high-ionic strength, oxidation of NADH proceeds through two platanetin binding sites, one of which has a lower affinity for the inhibitor (mean inhibition dose = 20 nanomoles per milligram protein), because it is located near the outer surface of the membrane. This latter site is the one involved in the oxidation of external NADPH and, possibly, also affected by spermine and spermidine. Similarly to NADH, oxidation of NADPH is fully sensitive to micromolar concentrations of free Ca2+ ions; in addition, similar concentrations of the sulfhydryl reagent mersalyl are required to inhibit both NADH and NADPH oxidative activities. The results are interpreted as evidence for the presence of a single nonspecific NAD(P)H dehydrogenase.  相似文献   

20.
1. An NADH-ferricyanide reductase activity has been isolated from the respiratory chain of Torulopsis utilis by using detergents. The isolated enzyme contains non-haem iron, acid-labile sulphide and FMN in the molar proportions 27.5:28.4:1. The preparation is free of FAD and largely free of cytochrome. 2. The enzyme catalyses ferricyanide reduction by NADPH at about 1% of the rate with NADH, and reacts poorly with acceptors other than ferricyanide. The rates of reduction of some acceptors are, as percentages of the rate with ferricyanide: menadione, 0.35%; lipoate, 0.01%; cytochrome c, 0.065%; dichlorophenolindophenol, 0.35%; ubiquinone-1, 0.08%. 3. Several properties of submitochondrial particles of T. utilis (non-haem iron, acid-labile sulphide, FMN and an NADH-reducible electron-paramagnetic-resonance signal) were found to co-purify with the NADH-ferricyanide reductase activity. Thus about 70% of the FMN and, within the limits of accuracy of the experiments, 100% of the non-haem iron and acid-labile sulphide of submitochondrial particles derived from T. utilis cells grown under conditions of glycerol limitation (but relatively low iron availability) can be attributed to the NADH-ferricyanide reductase. 4. It was also shown that the component of submitochondrial particles specifically bleached at 460nm by NADH [species 1 of Ragan & Garland (1971)] co-purifies with the NADH-ferricyanide reductase. 5. This successful purification of an NADH dehydrogenase from T. utilis forms a starting point for investigating the molecular properties of phenotypically modified mitochondrial NADH oxidation pathways that lack energy conservation between NADH and the cytochromes.  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