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 46 毫秒
1.
Kinetic measurements indicate that the energy-independent transhydrogenation of 3-acetylpyridine-NAD+ by NADPH in membranes of Escherichia coli follows a rapid equilibrium random bireactant mechanism. Each substrate, although reacting preferentially with its own binding site, is able to interact with the binding site of the other substrate to cause inhibition of enzyme activity. 5'-AMP (and ADP) and 2'-AMP interact with the NAD+- and NADP+-binding sites, respectively. Phenylglyoxal and 2,3-butanedione in borate buffer inhibit transhydrogenase activity presumably by reacting with arginyl residues. Protection against inhibition by 2,3-butanedione is afforded by NADP+, NAD+, and high concentrations of NADPH and NADH. Low concentrations of NADPH and NADH increase the rate of inhibition by 2,3-butanedione. Similar effects are observed for the inactivation of the transhydrogenase by tryptic digestion in the presence of these coenzymes. It is concluded that there are at least two conformations of the active site of the transhydrogenase which differ in the extent to which arginyl residues are accessible to exogenous agents such as trypsin and 2,3-butanedione. One conformation is induced by low concentrations of NADH and NADPH. Under these conditions the coenzymes could be reacting at the active site or at an allosteric site. The stimulation of transhydrogenase activity by low concentrations of the NADH is consistent with the latter possibility.  相似文献   

2.
Bovine heart mitochondrial transhydrogenase, a redox-linked proton pump, can be functionally and asymmetrically inserted into liposomes by a cholate-dialysis procedure such that the active site faces the external medium. N-(4-Azido-2-nitrophenyl)-2-aminoethylsulfonate (NAP-taurine), a membrane-impermeant photoprobe, when encapsulated in the vesicles, covalently modified the enzyme and inhibited transhydrogenation between NADPH and the 3-acetylpyridine analog of NAD+ (AcPyAD+) in a light-dependent manner. External AcPyAD+ increased the rate of inactivation several fold, whereas NADPH, NADP+, and NADH were without effect. Labeling of the enzyme by intravesicular [35S]NAP-taurine was enhanced by AcPyAD+ and NADP+, decreased by NADH, and not significantly affected by NADPH. These results indicate that transhydrogenase spans the membrane and that substrate binding alters the conformation of that domain of the enzyme protruding from the inner surface of the membrane.  相似文献   

3.
The yeast Candida parapsilosis possesses two routes of electron transfer from exogenous NAD(P)H to oxygen. Electrons are transferred either to the classical cytochrome pathway at the level of ubiquinone through an NAD(P)H dehydrogenase, or to an alternative pathway at the level of cytochrome c through another NAD(P)H dehydrogenase which is insensitive to antimycin A. Analyses of mitoplasts obtained by digitonin/osmotic shock treatment of mitochondria purified on a sucrose gradient indicated that the NADH and NADPH dehydrogenases serving the alternative route were located on the mitochondrial inner membrane. The dehydrogenases could be differentiated by their pH optima and their sensitivity to amytal, butanedione and mersalyl. No transhydrogenase activity occurred between the dehydrogenases, although NADH oxidation was inhibited by NADP+ and butanedione. Studies of the effect of NADP+ on NADH oxidation showed that the NADH:ubiquinone oxidoreductase had Michaelis-Menten kinetics and was inhibited by NADP+, whereas the alternative NADH dehydrogenase had allosteric properties (NADH is a negative effector and is displaced from its regulatory site by NAD+ or NADP+).  相似文献   

4.
Pyridine dinucleotide transhydrogenase of the Rhodospirillum rubrum chromatophore membrane was readily resolved by a washing procedure into two inactive components, a soluble transhydrogenase factor protein and an insoluble membrane-bound factor. Transhydrogenation was reconstituted on reassociation of these components. The capacity of the membrane factor to reconstitute enzymatic activity was lost after proteolysis of soluble transhydrogenase factor-depleted membranes with trypsin. NADP+ or NADPH, but neither NAD+ nor NADH, stimulated by several fold the rate of trypsin-dependent inactivation of the membrane factor. Substantial protection of the membrane factor from proteolytic inactivation was observed in the presence of Mg2+ ions, an inhibitor of transhydrogenation, or when the soluble transhydrogenase factor was bound to the membrane. Coincident with the loss of enzymatic reconstitutive capacity of the membrane factor was a loss in the ability of the membranes to bind the soluble transhydrogenase factor in a stable complex. The membrane component was inactivated by preincubating soluble transhydrogenase factor-depleted membranes at temperatures above 45 degrees. NADP+, NADPH, or Mg2+ ions, but neither NAD+ nor NADH, protected against inactivation. These studies indicate that (a) the binding of NADP+ or NADPH to the membrane factor promotes a conformational alteration in the protein such that its themostability and susceptibility to proteolysis are increased, and (b) the inhibitory Mg2+ ion-binding site resides in the membrane component.  相似文献   

5.
Incubation of Complex I (NADH-CoQ reductase) of ox heart mitochondria at 4 degrees C in the presence of 0.5 M NaClO4 followed by ammonium sulfate fractionation of the solubilized proteins results in the isolation of a resolved preparation still capable of catalyzing NADH-NAD+ transhydrogenation but having only low levels of NADH dehydrogenase activity. A number of NAD(H) analogues, including the photoaffinity probes, arylazido-beta-alanyl NAD+ (A3'-O-[3-[N-(4-azido-2-nitrophenyl)amino]propionyl]NAD+ and arylazido-beta-alanyl AcPyAD+ (A3'-O-[3-[N-(4-azido-2-nitrophenyl)amino]propionyl]AcPyAD+ can be utilized as substrates for transhydrogenation in this preparation. A further incubation (10 min) of the resolved NADH-NAD+ transhydrogenase in the presence of 0.5 M NaClO4, but now at 30 degrees C, results in the complete loss of this transhydrogenase activity. Photoaffinity labeling experiments utilizing arylazido-[3-3H]beta-alanyl NAD+ and arylazido-[3-3H]beta-alanyl AcPyAD+ with the resolved NADH-NAD+ transhydrogenase preparation prior to and following NaClO4 (30 degrees C) treatment indicates that the 42,000 molecular weight component of Complex I is the pyridine nucleotide binding site responsible for the major NADH-NAD+ (DD) transhydrogenase activity of Complex I.  相似文献   

6.
The mitochondrial nicotinamide nucleotide transhydrogenase catalyzes hydride ion transfer between NAD(H) and NADP(H) in a reaction that is coupled to proton translocation across the inner mitochondrial membrane. The enzyme (1043 residues) is composed of an N-terminal hydrophilic segment (approximately 400 residues long) which binds NAD(H), a C-terminal hydrophilic segment (approximately 200 residues long) which binds NADP(H), and a central hydrophobic segment (approximately 400 residues long) which appears to form about 14 membrane-intercalating clusters of approximately 20 residues each. Substrate modulation of transhydrogenase conformation appears to be intimately associated with its mechanism of proton translocation. Using trypsin as a probe of enzyme conformation change, we have shown that NADPH (and to a much lesser extent NADP) binding alters transhydrogenase conformation, resulting in increased susceptibility of several bonds to tryptic hydrolysis. NADH and NAD had little or no effect, and the NADPH concentration for half-maximal enhancement of trypsin sensitivity of transhydrogenase activity (35 microM) was close to the Km of the enzyme for NADPH. The NADPH-promoted trypsin cleavage sites were located 200-400 residues distant from the NADP(H) binding domain near the C-terminus. For example, NADPH binding greatly increased the trypsin sensitivity of the K410-T411 bond, which is separated from the NADP(H) binding domain by the 400-residue-long membrane-intercalating segment. It also enhanced the tryptic cleavage of the R602-L603 bond, which is located within the central hydrophobic segment. These results, which suggest a protein conformation change as a result of NADPH binding, have been discussed in relation to the mechanism of proton translocation by the transhydrogenase.  相似文献   

7.
Midgut mitochondria from fifth larval instar Manduca sexta exhibited a transhydrogenase that catalyzes the following reversible reaction: NADPH + NAD(+) <--> NADP(+) + NADH. The NADPH-forming transhydrogenation occurred as a nonenergy- and energy-linked activity. Energy for the latter was derived from the electron transport-dependent utilization of NADH or succinate, or from Mg++-dependent ATP hydrolysis by ATPase. The NADH-forming and all of the NADPH-forming reactions appeared optimal at pH 7.5, were stable to prolonged dialysis, and displayed thermal lability. N,N'-dicyclohexylcarbodiimide (DCCD) inhibited the NADPH --> NAD(+) and energy-linked NADH --> NADP(+) transhydrogenations, but not the nonenergy-linked NADH --> NADP(+) reaction. Oligomycin only inhibited the ATP-dependent energy-linked activity. The NADH-forming, nonenergy-linked NADPH-forming, and the energy-linked NADPH-forming activities were membrane-associated in M. sexta mitochondria. This is the first demonstration of the reversibility of the M. sexta mitochondrial transhydrogenase and, more importantly, the occurrence of nonenergy-linked and energy-linked NADH --> NADP(+) transhydrogenations. The potential relationship of the transhydrogenase to the mitochondrial, NADPH-utilizing ecdysone-20 monooxygenase of M. sexta is considered.  相似文献   

8.
Transhydrogenase couples the redox reaction between NADH and NADP+ to proton translocation across a membrane. The enzyme comprises three components; dI binds NAD(H), dIII binds NADP(H), and dII spans the membrane. The 1,4,5,6-tetrahydro analogue of NADH (designated H2NADH) bound to isolated dI from Rhodospirillum rubrum transhydrogenase with similar affinity to the physiological nucleotide. Binding of either NADH or H2NADH led to closure of the dI mobile loop. The 1,4,5,6-tetrahydro analogue of NADPH (H2NADPH) bound very tightly to isolated R. rubrum dIII, but the rate constant for dissociation was greater than that for NADPH. The replacement of NADP+ on dIII either with H2NADPH or with NADPH caused a similar set of chemical shift alterations, signifying an equivalent conformational change. Despite similar binding properties to the natural nucleotides, neither H2NADH nor H2NADPH could serve as a hydride donor in transhydrogenation reactions. Mixtures of dI and dIII form dI2dIII1 complexes. The nucleotide charge distribution of complexes loaded either with H2NADH and NADP+ or with NAD+ and H2NADPH should more closely mimic the ground states for forward and reverse hydride transfer, respectively, than previously studied dead-end species. Crystal structures of such complexes at 2.6 and 2.3 A resolution are described. A transition state for hydride transfer between dihydronicotinamide and nicotinamide derivatives determined in ab initio quantum mechanical calculations resembles the organization of nucleotides in the transhydrogenase active site in the crystal structure. Molecular dynamics simulations of the enzyme indicate that the (dihydro)nicotinamide rings remain close to a ground state for hydride transfer throughout a 1.4 ns trajectory.  相似文献   

9.
Acetylpyridine NADP replaced NADP in promoting the Mn2+ ion-requiring mitochondrial "malic" enzyme of Hymenolepis diminuta. Disrupted mitochondria displayed low levels of an apparent oxaloacetate-forming malate dehydrogenase activity when NAD or acetylpyridine NAD served as the coenzyme. Significant malate-dependent reduction of acetylpyridine NAD by H. diminuta mitochondria required Mn2+ ion and NADP, thereby indicating the tandem operation of "malic" enzyme and NADPH:NAD transhydrogenase. Incubation of mitochondrial preparations with oxaloacetate resulted in a non-enzymatic decarboxylation reaction. Coupling of malate oxidation with electron transport via the "malic" enzyme and transhydrogenase was demonstrated by polarographic assessment of mitochondrial reduced pyridine nucleotide oxidase activity.  相似文献   

10.
S R Earle  S G O'Neal  R R Fisher 《Biochemistry》1978,17(22):4683-4690
Chemical-modification studies on submitochondrial particle pyridine dinucleotide transhydrogenase (EC 1.6.1.1) demonstrate the presence of one class of sulfhydryl group in the nicotinamide adenine dinucleotide phosphate (NADP) site and another peripheral to the active site. Reaction of the peripheral sulfhydryl group with N-ethylmaleimide, or both classes with 5,5'-dithiobis(2-nitrobenzoic acid), completely inactivated transhydrogenase. NADP+ or NADPH nearly completely protected against 5,5'-dithiobis(2-nitrobenzoic acid) inactivation and modification of both classes of sulfhydryl groups, while NADP+ only partially protected against and NADPH substantially stimulated N-ethylmaleimide inactivation. Methyl methanethiolsulfonate treatment resulted in methanethiolation at both classes of sulfhydryl groups, and either NADP+ or NADPH protected only the NADP site group. S-Methanethio and S-cyano transhydrogenases were active derivatives with pH optima shifted about 1 unit lower than that of the native enzyme. These experiments indicate that neither class of sulfhydryl group is essential for transhydrogenation. Lack of involvement of either sulfhydryl group in energy coupling to transhydrogenation is suggested by the observations that S-methanethio transhydrogenase is functional in (a) energy-linked transhydrogenation promoted by phenazine methosulfate mediated ascorbate oxidation and (b) the generation of a membrane potential during the reduction of NAD+ by reduced nicotinamide adenine dinucleotide phosphate (NADPH).  相似文献   

11.
The reversible, membrane-associated transhydrogenase that catalyzes hydride-ion transfer between NADP(H) and NAD(H) was evaluated and compared to the corresponding NADH oxidase and succinate dehydrogenase activities in midgut and fat body mitochondria from fifth larval instar Manduca sexta. The developmentally significant NADPH-forming transhydrogenation occurs as a nonenergy- or energy-linked activity with energy for the latter derived from either electron transport-dependent NADH or succinate utilization, or ATP hydrolysis by Mg++-dependent ATPase. In general, the plant flavonoids examined (chyrsin, juglone, morine, quercetin, and myricetin) affected all reactions in a dose-dependent fashion. Differences in the responses to the flavonoids were apparent, with the most notable being inhibition of midgut, but stimulation of fat body transhydrogenase by morin, and myricetin as also noted for NADH oxidase and succinate dehydrogenase. Although quercetin inhibited or stimulated transhydrogenase activity depending on the origin of mitochondria, it was without effect on either midgut or fat body NADH oxidase or succinate dehydrogenase. Observed sonication-dependent increases in flavonoid inhibition may well reflect an alteration in membrane configuration, resulting in increased exposure of the enzyme systems to the flavonoids. The effects of flavonoids on the transhydrogenation, NADH oxidase, and succinate dehydrogenase reactions suggest that compounds of this nature may prove valuable in the control of insect populations by affecting these mitochondrial enzyme components.  相似文献   

12.
A unique Trp residue in the recombinant dIII component of transhydrogenase from human heart mitochondria (hsdIII), and an equivalent Trp engineered into the dIII component of Rhodospirillum rubrum transhydrogenase (rrdIII.D155W), are more fluorescent when NADP(+) is bound to the proteins, than when NADPH is bound. We have used this to determine the occupancy of the binding site during transhydrogenation reactions catalysed by mixtures of recombinant dI from the R. rubrum enzyme and either hsdIII or rrdIII.D155W. The standard redox potential of NADP(+)/NADPH bound to the dIII proteins is some 60-70 mV higher than that in free solution. This results in favoured reduction of NADP(+) by NADH at the catalytic site, and supports the view that changes in affinity at the nucleotide-binding site of dIII are central to the mechanism by which transhydrogenase is coupled to proton translocation across the membrane.  相似文献   

13.
NAD+ reduction catalyzed by transhydrogenase (EC 1.6.1.1) from E. coli membrane particles at the expense of NADPH oxidation is coupled with phenyldicarbaundecaborate (PCB-) absorption by the particles. This process is inhibited by oxidative phosphorylation protonophorous uncouplers and by equilibration of concentrations of the substrates and products of the transhydrogenase reaction. Elimination of the water-soluble part of membrane ATPase results in the inhibition of PCB- absorption at the expense of the transhydrogenase reaction energy. Treatment of the particles by dicyclohexyl carbodiimide increases the transhydrogenase-coupled absorption of PCB-. The transhydrogenase-induced increase of pPCB in the suspension of particles is directly correlated with the ratio of ([NADPH].[NAD+])/([NADP+].[NADH]). When this value is equal to 1, no energy-dependent increase of pPCB was observed. NADP+ reduction at the expense of NADH oxidation leads to a decrease in the amount of PCB- absorbed by the particles at the expense of ATP hydrolysis energy. The experimental data suggest that NADPH oxidation in the course of the transhydrogenase reaction is coupled with the formation of a membrane potential with a positive charge localized inside the particles.  相似文献   

14.
Energization of the pyridine nucleotide transhydrogenase in everted membrane vesicles from Escherichia coli JM83 was compared with the process in vesicles of the same strain transformed with the plasmid pDC21 overexpressing this enzyme. Proton translocation was assayed by the quenching of the fluorescence of the probe quinacrine. Agents able to discharge transmembrane proton gradients such as nigericin and the uncouplers 3,3',4',5-tetrachlorosalicylanilide and carbonyl cyanide m-chlorophenylhydrazone inhibited ATP-dependent transhydrogenation of NADP by NADH and discharged transmembrane proton gradients generated by transhydrogenation of AcNAD by NADPH, by oxidation of NADH, and by hydrolysis of ATP. This was observed in everted membrane vesicles of both strains JM83 and JM83pDC21. These strains differed significantly in the response of the NADH oxidation-dependent transhydrogenase. This reaction was inhibited by nigericin and uncouplers in membrane vesicles of JM83 but there was little inhibition or the reaction was stimulated in JM83pDC21, in spite of the discharge of the NADH oxidation-generated proton gradient measured by quinacrine fluorescence in the latter strain. It is proposed that the transhydrogenase is energized by direct or local (nonbulk phase) proton translocation in membranes of this strain. Uncouplers might facilitate these routes but would not discharge them. The generality of these observations was shown using other strains. NADH oxidase activity was severalfold lower in membrane vesicles of JM83pDC21 compared with JM83. The levels of ubiquinone and cytochromes, and the activities of NADH dehydrogenases I and II, and of cytochrome oxidase, were similar in the two strains. It is concluded that the NADH oxidase activity of JM83pDC21 is low because of the reduced rate of collision between electron-transferring complexes of the respiratory chain due to the large amount of transhydrogenase protein in the membranes of this strain. The large amount of transhydrogenase favors direct, nonbulk phase proton transfer. Transhydrogenase activity was stimulated by Ca2+, Mg2+, or Mn2+.  相似文献   

15.
We studied the physiological effect of the interconversion between the NAD(H) and NADP(H) coenzyme systems in recombinant Saccharomyces cerevisiae expressing the membrane-bound transhydrogenase from Escherichia coli. Our objective was to determine if the membrane-bound transhydrogenase could work in reoxidation of NADH to NAD+ in S. cerevisiae and thereby reduce glycerol formation during anaerobic fermentation. Membranes isolated from the recombinant strains exhibited reduction of 3-acetylpyridine-NAD+ by NADPH and by NADH in the presence of NADP+, which demonstrated that an active enzyme was present. Unlike the situation in E. coli, however, most of the transhydrogenase activity was not present in the yeast plasma membrane; rather, the enzyme appeared to remain localized in the membrane of the endoplasmic reticulum. During anaerobic glucose fermentation we observed an increase in the formation of 2-oxoglutarate, glycerol, and acetic acid in a strain expressing a high level of transhydrogenase, which indicated that increased NADPH consumption and NADH production occurred. The intracellular concentrations of NADH, NAD+, NADPH, and NADP+ were measured in cells expressing transhydrogenase. The reduction of the NADPH pool indicated that the transhydrogenase transferred reducing equivalents from NADPH to NAD+.  相似文献   

16.
The actions of Dexon on the NADH-ferricyanide oxidoreductase and the NADPH oxidase system of electron transfer particles (ETP) from beef heart as well as on the NADPH-cytochrome c oxidoreductase from brewer's yeast (Saccharomyces carlsbergensis Hansen) were investigated. The inhibition of the NADH dehydrogenase activity of ETP and that of the yeast enzyme correspond with respect to the following characteristics: 1) increase in the inhibition, 2) enhancement of the Dexon sensitivity by one order of magnitude after preincubation in the presence of NAD(P)H, 3) irreversibility of the inhibition, 4) no detectable changes in the spectral properties and in coenzyme activity of FMN after acid extraction from Dexon-treated enzyme. The inhibition of the NADH dehydrogenase activity of ETP is diminished by both NAD+ and FMN. However, no interaction of Dexon with NAD(P)H or FMN could be detected in the absence of enzyme or apoenzyme. The concentration of half-inhibition by Dexon for the yeast enzyme corresponds with its FMN concentration. It is proposed that both apoenzyme, NAD(P)H and FMN are involved in the interaction with Dexon. Possible mechanisms of binding are both complanar complexations of the ring systems and a triazene formation between FMNH2 and Dexon. The NADPH oxidase activity of the ETP is partly inhibited; the share inhibited by Dexon may represent the pathway via the transhydrogenase reaction.  相似文献   

17.
1. At 21 degrees C incubation of NADH-ubiquinone-1 reductase (Complex 1) with trypsin caused selective inhibition of nicotinamide nucleotide transhydrogenase activity. The reduction of K3Fe(CN)6 by NADH or NADPH was unaffected, but a slow decrease in the rate of reduction of ubiquinone-1 by NADH was observed. 2. The pH-dependence of nicotinamide nucleotide transhydrogenase activity differed in Complex I and trypsin-treated Complex I. The trypsin-labile activity had a pH optimum of approx. 6.5, whereas the trypsin-resistant activity had a pH optimum of approx. 5.5 or less. 3. The trypsinlabile transhydrogenase activity was specifically inhibited by butanedione or phenylglyoxal and was identified with the enzyme catalysing energy-linked transhydrogenase activity in submitochondrial particles. 4. Polyacrylamide-gel electrophoresis in the presence of sodium dodecyl sulphate revealed that trypsin caused degradation of a polypeptide of mol.wt 20500 in parallel with the loss of transhydrogenase activity. 5. At 30 degrees C and higher trypsin concentrations, the rate of reduction of K3Fe(CN)6 by NADH or NADPH slowly decreased. Increased lability of NADH-K3Fe(CN)6 reductase activity to trypsin was observed when the endogenous phospholipid of Complex I was depleted by detergent or phospholipase A treatment. 6. Polyacrylamide-gel electrophoresis indicated that removal of phospholipid allowed much more extensive degradation of constituent polypeptides by trypsin. The subunits of the low-molecular-weight (type II) dehydrogenase (53000 and 26000 mol.wt.) were, however, relatively resistant to trypsin even in phospholipid-depleted preparations.  相似文献   

18.
Pyridine Nucleotide Transhydrogenase from Azotobacter vinelandii   总被引:5,自引:0,他引:5       下载免费PDF全文
A method is described for the partial purification of pyridine nucleotide transhydrogenase from Azotobacter vinelandii (ATCC 9104) cells. The most highly purified preparation catalyzes the reduction of 300 mumoles of nicotinamide adenine dinucleotide (NAD(+)) per min per mg of protein under the assay conditions employed. The enzyme catalyzes the reduction of NAD(+), deamino-NAD(+), and thio-NAD(+) with reduced nicotinamide adenine dinucleotide phosphate (NADPH) as hydrogen donor, and the reduction of nicotinamide adenine dinucleotide phosphate (NADP(+)) and thio-NAD(+) with reduced NAD (NADH) as hydrogen donor. The reduction of acetylpyridine AD(+), pyridinealdehyde AD(+), acetylpyridine deamino AD(+), and pyridinealdehydedeamino AD(+) with NADPH as hydrogen donor was not catalyzed. The enzyme catalyzes the transfer of hydrogen more readily from NADPH than from NADH with different hydrogen acceptors. The transfer of hydrogen from NADH to NADP(+) and thio-NAD(+) was markedly stimulated by 2'-adenosine monophosphate (2'-AMP) and inhibited by adenosine diphosphate (ADP), adenosine triphosphate (ATP), and phosphate ions. The transfer of hydrogen from NADPH to NAD(+) was only slightly affected by phosphate ions and 2'-AMP, except at very high concentrations of the latter reagent. In addition, the transfer of hydrogen from NADPH to thio-NAD(+) was only slightly influenced by 2'-AMP, ADP, ATP, and other nucleotides. The kinetics of the transhydrogenase reactions which utilized thio-NAD(+) as hydrogen acceptor and NADH or NADPH as hydrogen donor were studied in some detail. The results suggest that there are distinct binding sites for NADH and NAD(+) and perhaps a third regulator site for NADP(+) or 2'-AMP. The heats of activation for the transhydrogenase reactions were determined. The properties of this enzyme are compared with those of other partially purified transhydrogenases with respect to the regulatory functions of 2'-AMP and other nucleotides on the direction of flow of hydrogen between NAD(+) and NADP(+).  相似文献   

19.
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   

20.
Ian M. Mller  John M. Palmer 《BBA》1981,638(2):225-233
(1) The optimum pH for the oxidation of exogenous NADH by mitochondria from both Jerusalem artichoke (Helianthus tuberosus) tubers and Arum maculatum spadices was 7.0–7.1. NADPH oxidation had a lower optimum pH of 6.6 in Arum and 6.0 in Jerusalem artichoke mitochondria. In both types of mitochondria the rates of NADH and NADPH oxidation were identical below pH 6.0–5.5. (2) It is shown conclusively that neither a phosphatase converting NADPH to NADH nor a nicotinamide nucleotide transhydrogenase was involved in the oxidation of NADPH by these mitochondria. (3) Palmitoyl-CoA, an inhibitor of transhydrogenase activity in mammalian mitochondria, inhibits both NADH and NADPH oxidation by plant mitochondria with a Ki of about 10 μM. (4) It is concluded that the known properties of NAD(P)H oxidation are best explained by assuming the presence of a second dehydrogenase specific for NADPH. At low pH, electron flow from the two dehydrogenases to oxygen shares a common rate-limiting step.  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