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.
Tania Bizouarn  Tina Bhakta 《BBA》2005,1708(3):404-410
Transhydrogenase (E.C. 1.6.1.1) couples the redox reaction between NAD(H) and NADP(H) to the transport of protons across a membrane. The enzyme is composed of three components. The dI and dIII components, which house the binding site for NAD(H) and NADP(H), respectively, are peripheral to the membrane, and dII spans the membrane. We have estimated dissociation constants (Kd values) for NADPH (0.87 μM), NADP+ (16 μM), NADH (50 μM), and NAD+ (100-500 μM) for intact, detergent-dispersed transhydrogenase from Escherichia coli using micro-calorimetry. This is the first complete set of dissociation constants of the physiological nucleotides for any intact transhydrogenase. The Kd values for NAD+ and NADH are similar to those previously reported with isolated dI, but the Kd values for NADP+ and NADPH are much larger than those previously reported with isolated dIII. There is negative co-operativity between the binding sites of the intact, detergent-dispersed transhydrogenase when both nucleotides are reduced or both are oxidised.  相似文献   

2.
NAD+ and NADP+, chemically similar and with almost identical standard oxidation–reduction potentials, nevertheless have distinct roles, NAD+ serving catabolism and ATP generation whereas NADPH is the biosynthetic reductant. Separating these roles requires strict specificity for one or the other coenzyme for most dehydrogenases. In many organisms this holds also for glutamate dehydrogenases (GDH), NAD+-dependent for glutamate oxidation, NADP+-dependent for fixing ammonia. In higher animals, however, GDH has dual specificity. It has been suggested that GDH in mitochondria reacts only with NADP(H), the NAD+ reaction being an in vitro artefact. However, contrary evidence suggests mitochondrial GDH not only reacts with NAD+ but maintains equilibrium using the same pool as accessed by β-hydroxybutyrate dehydrogenase. Another complication is the presence of an energy-linked dehydrogenase driving NADP+ reduction by NADH, maintaining the coenzyme pools at different oxidation–reduction potentials. Its coexistence with GDH makes possible a futile cycle, control of which is not yet properly explained. Structural studies show NAD+-dependent, NADP+-dependent and dual-specificity GDHs are closely related and a few site-directed mutations can reverse specificity. Specificity for NAD+ or for NADP+ has probably emerged repeatedly during evolution, using different structural solutions on different occasions. In various GDHs the P7 position in the coenzyme-binding domain plays a key role. However, whereas in other dehydrogenases an acidic P7 residue usually hydrogen bonds to the 2′- and 3′-hydroxyls, dictating NAD+ specificity, among GDHs, depending on detailed conformation of surrounding residues, an acidic P7 may permit binding of NAD+ only, NADP+ only, or in higher animals both.  相似文献   

3.
NADP-Utilizing Enzymes in the Matrix of Plant Mitochondria   总被引:9,自引:4,他引:5       下载免费PDF全文
Purified potato tuber (Solanum tuberosum L. cv Bintie) mitochondria contain soluble, highly latent NAD+- and NADP+-isocitrate dehydrogenases, NAD+- and NADP+-malate dehydrogenases, as well as an NADPH-specific glutathione reductase (160, 25, 7200, 160, and 16 nanomoles NAD(P)H per minute and milligram protein, respectively). The two isocitrate dehydrogenase activities, but not the two malate dehydrogenase activities, could be separated by ammonium sulfate precipitation. Thus, the NADP+-isocitrate dehydrogenase activity is due to a separate matrix enzyme, whereas the NADP+-malate dehydrogenase activity is probably due to unspecificity of the NAD+-malate dehydrogenase. NADP+-specific isocitrate dehydrogenase had much lower Kms for NADP+ and isocitrate (5.1 and 10.7 micromolar, respectively) than the NAD+-specific enzyme (101 micromolar for NAD+ and 184 micromolar for isocitrate). A broad activity optimum at pH 7.4 to 9.0 was found for the NADP+-specific isocitrate dehydrogenase whereas the NAD+-specific enzyme had a sharp optimum at pH 7.8. Externally added NADP+ stimulated both isocitrate and malate oxidation by intact mitochondria under conditions where external NADPH oxidation was inhibited. This shows that (a) NADP+ is taken up by the mitochondria across the inner membrane and into the matrix, and (b) NADP+-reducing activities of malate dehydrogenase and the NADP+-specific isocitrate dehydrogenase in the matrix can contribute to electron transport in intact plant mitochondria. The physiological relevance of mitochondrial NADP(H) and soluble NADP(H)-consuming enzymes is discussed in relation to other known mitochondrial NADP(H)-utilizing enzymes.  相似文献   

4.
It has been shown that direct excitation of NADH (or NADPH) in aqueous medium at 254 nm, or at wavelengths longer than 320 nm (where only the reduced nicotinamide moiety absorbs), leads to generation of NAD+ (or NADP+). The reaction proceeds both in the presence and absence of oxygen. Under aerobic conditions the reaction is accompanied by formation of H2O2 at a level equimolar with that of the NADH present in solution. On irradiation at wavelengths longer than 320 nm, conversion of NADH to enzymatically active NAD+ is about 75%. Under analogous irradiation conditions, the dimers (NAD)2 and (NADP)2 undergo disproportionation to NAD+ and NADP+, respectively, to the extent of 90%. Both physicochemical and enzymatic criteria were employed to formulate mechanisms for the photooxidation of NADH and the photodisproportionation of the dimer (NAD)2.  相似文献   

5.
The unicellular green algae Chlorella ellipsoidea was used tostudy transient changes in the energy state of adenylates andthe redox states of pyridine nucleotides induced by environmentalchanges. The transition from anaerobic to aerobic conditionsin the dark induced a sharp rise in the ATP ratio [ATP/(ATP+ADP+AMP)],a sudden decrease in the NADH ratio [NADH/(NAD++NADPH)] anda transient drop in the NADPH ratio [NADPH/(NADP++NADPH)]. Illuminationafter a dark period under anaerobic, CO2-free conditions inducedsharp increases in the ATP and NADPH ratios and a slower decreasein the NADH ratio. Illumination under aerobic conditions, ineither the presence or absence of CO2, caused a sharp increasein the NADPH ratio, a small increase in the ATP ratio and aslower increase in the NADH ratio. In the presence of CO2, asubsequent large drop in the NADPH ratio occurred. Darkeningunder anaerobic, CO2-free conditions induced a sudden decreasein the ATP ratio, a temporary fall in the NADPH ratio and aslow increase in the NADH ratio. Darkening under aerobic conditionsinduced transient drops in the ATP and NADPH ratios and a suddendrop in the NADH ratio. The addition of CO2 to the atmospherewith illumination produced a decrease in all three parameters. These results are discussed in relation to current theoriesof the interaction between photosynthesis and respiration. Ourobservations indicate that the energy and reducing potentialsgenerated by photochemical processes are used for and controlother processes besides CO2 fixation in photosynthetic cells. (Received December 3, 1981; Accepted May 4, 1982)  相似文献   

6.
A sonicate of Achromobacter parvulus IFO-13182 produced NADPH from NADP+by an NADP+-linked malic enzyme [l-malate: NAD(P)+oxidoreductase, EC 1.1.1.39–40] reaction in the presence of l-malic acid and divalent metal ions. Malic enzyme of A. parvulus was stabilized by 5% l-malic acid, and activity was maintained at 60°C for 1 h. Contaminating phosphatase (orthophosphoricmonoester phosphohydrolase, EC 3.1.3.1–2) was completely inactivated by this treatment. Among the conditions tested, the optimum NADPH production was done using 36 μmol NADP+, 67 μmol l-malic acid, 63 μmol MgCl2 and 1 unit of the malic enzyme in 3 ml of 55 mm phosphate buffer (pH 7.8). Conversion ratio of NADPH from NADP+ reached 100% after 4 h incubation at 30°C and the amount of NADPH accumulated was ~12 μmol ml?1of the reaction mixture. No dephosphorylation of NADP+to NAD+or of NADPH to NADH was found by high performance liquid chromatography. The NADPH produced by such enzymatic reduction was purified by ethanol precipitation and dried in vacuo in powdered form with 97% purity, judged from the ratio of the absorbances at 340 and 260 nm. The purity of the NADPH produced was determined to be 95% from its coenzyme activity with NAD(P)+-linked glutathione reductase [NAD(P)H: oxidized-glutathione oxidoreductase, EC 1.6.4.2].  相似文献   

7.
Mitochondria from the parasitic helminth, Hymenolepis diminuta, catalyzed both NADPH:NAD+ and NADH:NADP+ transhydrogenase reactions which were demonstrable employing the appropriate acetylpyridine nucleotide derivative as the hydride ion acceptor. Thionicotinamide NAD+ would not serve as the oxidant in the former reaction. Under the assay conditions employed, neither reaction was energy linked, and the NADPH:NAD+ system was approximately five times more active than the NADH:NADP+ system. The NADH:NADP+ reaction was inhibited by phosphate and imidazole buffers, EDTA, and adenyl nucleotides, while the NADPH:NAD+ reaction was inhibited only slightly by imidazole and unaffected by EDTA and adenyl nucleotides. Enzyme coupling techniques revealed that both transhydrogenase systems functioned when the appropriate physiological pyridine nucleotide was the hydride ion acceptor. An NADH:NAD+ transhydrogenase system, which was unaffected by EDTA, or adenyl nucleotides, also was demonstrable in the mitochondria of H. diminuta. Saturation kinetics indicated that the NADH:NAD+ reaction was the product of an independent enzyme system. Mitochondria derived from another parasitic helminth, Ascaris lumbricoides, catalyzed only a single transhydrogenase reaction, i.e., the NADH:NAD+ activity. Transhydrogenase systems from both parasites were essentially membrane bound and localized on the inner mitochondrial membrane. Physiologically, the NADPH:NAD+ transhydrogenase of H. diminuta may serve to couple the intramitochondrial metabolism of malate (via an NADP linked “malic” enzyme) to the anaerobic NADH-dependent ATP-generating fumarate reductase system. In A. lumbricoides, where the intramitochondrial metabolism of malate depends on an NAD-linked “malic” enzyme which is localized primarily in the intermembrane space, the NADH:NAD+ transhydrogenase activity may serve physiologically in the translocation of hydride ions across the inner membrane to the anaerobic energy-generating fumarate reductase system.  相似文献   

8.
The NAD+/NADH ratio was 12 in whole soybean nodules tissue,but only 2 in bacteroids, as a result of the high concentrationof NADH. By contrast, NADP+/NADPH ratios were less than unityin both nodules and bacteroids, being 0.28 and 0.37, respectively. The adenylate energy charge values in bacteroids and nodules,0.37 and 0.39, respectively, were remarkably low, and were insharp contrast to the normal value of 0.83 in root tissue. (Received July 19, 1988; Accepted March 9, 1989)  相似文献   

9.
Mitochond$$$a isolated from Jerusalem artichoke tubers oxidizedendogenous NADH by both a piericidin A-sensitive and -resistantdehydrogenase if the level of oxaloacetate was kept low. Inwashed mitochondria the addition of NAD+ stimulated respirationin the presence of a variety of NAD+ -linked substrates. Inmitochondria purified through a sucrose density gradient exogenousNAD+ caused a substantial stimulation of respiration only inthe presence of malate. The apparent Km for malate was 20 mMin the absence of NAD+ and 2 mM in the presence of exogenousNAD+ The products of malate oxidation were altered by the additionof exogenous NAD+. Oxaloacetate and pyruvate were produced inequal amounts in the absence of added NAD+, but in the presenceof exogenous NAD+ more pyruvate was formed. The rapid oxidationof malate in the absence of added NAD+ required phosphate whilethe NAD+-stimulated component was not affected by the absenceof phosphate. Phenylsuccinate inhibited the reduction of exogenousNAD+ by malate; this compound was found to inhibit isolatedma!ate dehydrogenase and mahe enzyme. Several lines of evidencesuggest that electron flux through one of the NADH dehydrogenasesystems may directly affect the flow through the other dehydrogenases.  相似文献   

10.
Inside-out submitochondrial particles from both potato (Solanum tuberosum L. cv. Bintje) tubers and pea (Pisum sativum L. cv. Oregon) leaves possess three distinct dehydrogenase activities: Complex I catalyzes the rotenone-sensitive oxidation of deamino-NADH, NDin(NADPH) catalyzes the rotenone-insensitive and Ca2+-dependent oxidation of NADPH and NDin(NADH) catalyzes the rotenone-insensitive and Ca2+-independent oxidation of NADH. Diphenylene iodonium (DPI) inhibits complex I, NDin(NADPH) and NDin (NADH) activity with a Ki of 3.7, 0.17 and 63 µM, respectively, and the 400-fold difference in Ki between the two NDin made possible the use of DPI inhibition to estimate NDin (NADPH) contribution to malate oxidation by intact mitochondria. The oxidation of malate in the presence of rotenone by intact mitochondria from both species was inhibited by 5 µM DPI. The maximum decrease in rate was 10–20 nmol O2 mg?1 min?1. The reduction level of NAD(P) was manipulated by measuring malate oxidation in state 3 at pH 7.2 and 6.8 and in the presence and absence of an oxaloacetate-removing system. The inhibition by DPI was largest under conditions of high NAD(P) reduction. Control experiments showed that 125 µM DPI had no effect on the activities of malate dehydrogenase (with NADH or NADPH) or malic enzyme (with NAD+ or NADP+) in a matrix extract from either species. Malate dehydrogenase was unable to use NADP+ in the forward reaction. DPI at 125 µM did not have any effect on succinate oxidation by intact mitochondria of either species. We conclude that the inhibition caused by DPI in the presence of rotenone in plant mitochondria oxidizing malate is due to inhibition of NDin(NADPH) oxidizing NADPH. Thus, NADP turnover contributes to malate oxidation by plant mitochondria.  相似文献   

11.
The light-driven, thiosulfate-dependent reduction of nicotinamideadenine dinucleotides under acrobic conditions in whole cellsof Chromatium vinosum was investigated. The total concentration of pyridine nucleotides in whole cellswas about 50 nmoles per µmole of bacteriochlorophyll.Under dark aerobic conditions, the majority of the nucleotidespresent was NAD+ with about 20% as NADP+. About 40% of the total NAD was reduced under continuous illumination.Thiosulfate or sulfide was needed for the photoreduction, whileorganic acids such as succinate or malate were not. The initialrate of NAD+ photoreduction in the presence of thiosulfate wasapproximately 100 nmoles per µmole of bacteriochlorophyllper min. The NAD+ photoreduction was strongly inhibited by uncouplersand electron transfer inhibitors. In contrast, an energy transferinhibitor, N, N'-dicyclohexylcarbodiimide, did not affect NAD+photoreduction at a concentration at which the light-inducedATP formation was inhibited. A transmembrane electrochemicalH+ gradient generated by cyclic electron transfer may be theenergy source for reduction of NAD+ in Chromatium vinosum. (Received April 2, 1980; )  相似文献   

12.
Here, we report that the smooth muscle and endothelium of the pig coronary artery differ in the profiles of energy metabolism nucleotides. ATP levels in the freshly isolated smooth muscle (1490 ± 93, all the values are in pmol/mg protein) were significantly greater than in the endothelium (418 ± 68). In contrast, endothelium contained higher levels of NADH (328 ± 21), NAD+ (1210 ± 28), NADPH (87 ± 2), and NADP+ (77 ± 4) than smooth muscle (17 ± 2, 96 ±14, 7 ± 1, and 8 ± 1, respectively). However, smooth muscle and endothelium do not differ from each other in the ratios of NADH/NAD+ or NADPH/NADP+. Cells cultured from smooth muscle and endothelium contained less ATP (93 ± 2, 141 ± 6) and had lower ratios of NADH/NAD+ than the freshly isolated tissues but the NADPH/NADP+ ratios remained similar. We conclude that (a) freshly isolated smooth muscle and endothelium differ in their profiles of the energy metabolism nucleotides, and (b) culturing the cells alters the profile.  相似文献   

13.
The steady-state kinetics of alcohol dehydrogenases (alcohol:NAD+ oxidoreductase, EC 1.1.1.1 and alcohol:NADP+ oxidoreductase, EC 1.1.1.2), lactate dehydrogenases (l-lactate:NAD+ oxidoreductase, EC 1.1.1.27 and d-lactate:NAD+ oxidoreductase, EC 1.1.1.28), malate dehydrogenase (l-malate:NAD+ oxidoreductase, EC 1.1.1.37), and glyceraldehyde-3-phosphate dehydrogenases [d-glyceraldehyde-3-phosphate:NAD+ oxidoreductase (phosphorylating), EC 1.2.1.12] from different sources (prokaryote and eukaryote, mesophilic and thermophilic organisms) have been studied using NAD(H), N6-(2-carboxyethyl)-NAD(H), and poly(ethylene glycol)-bound NAD(H) as coenzymes. The kinetic constants for NAD(H) were changed by carboxyethylation of the 6-amino group of the adenine ring and by conversion to macromolecular form. Enzymes from thermophilic bacteria showed especially high activities for the derivatives. The relative values of the maximum velocity (NAD = 1) of Thermus thermophilus malate dehydrogenase for N6-(2-carboxyethyl)-NAD and poly(ethylene glycol)-bound NAD were 5.7 and 1.9, respectively, and that of Bacillus stearothermophilus glyceraldehyde-3-phosphate dehydrogenase for poly(ethylene glycol)-bound NAD was 1.9.  相似文献   

14.
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   

15.
The activity of pure calf-liver and Escherichia coli thioredoxin reductases decreased drastically in the presence of NADPH or NADH, while NADP+, NAD+ and oxidized E. coli thioredoxin activated both enzymes significantly, particularly the bacterial one. The loss of activity under reducing conditions was time-dependent, thus suggesting an inactivation process: in the presence of 0.24 mM NADPH the half-lives for the E. coli and calf-liver enzymes were 13.5 and 2 min, respectively. Oxidized E. coli thioredoxin fully protected both enzymes from inactivation, and also promoted their complete reactivation after only 30 min incubation at 30° C. Lower but significant protection and reactivation was also observed with NADP+ and NAD+. EDTA protected thioredoxin reductase from NADPH inactivation to a great degree, thus indicating the participation of metals in the process; EGTA did not protect the enzyme from redox inactivation. Thioredoxin reductase was extensively inactivated by NADPH under aerobic and anaerobic conditions, thus excluding the participation of O2 or oxygen active species in redox inactivation. The loss of thioredoxin reductase activity promoted by NADPH was much faster and complete in the presence of NAD+ glycohydrolase, thus suggesting that inactivation was related to full reduction of the redox-active disulfide. Those results indicate that thioredoxin reductase activity can be modulated in bacteria and mammals by the redox status of NADP(H) and thioredoxin pools, in a similar way to glutathione reductase. This would considerably expand the regulatory potential of the thioredoxin-thioredoxin reductase system with the enzyme being self-regulated by its own substrate, a regulatory protein.Abbreviations DTNB 5,5-dithiobis(2-nitrobenzoate) - EGTA Ethylenglycoltetraacetic Acid - TNB 5-thio-2-nitrobenzoate - Trx Thioredoxin - Trx(SH)2 Reduced Thioredoxin - Trx-S2 Oxidized Thioredoxin  相似文献   

16.
Nicotinamide adenine dinucleotide (NAD+/NADH) along with its phosphorylated form (NADP+/NADPH) are two molecules ubiquitously present in all organisms, and they play key roles as cofactors in fundamental catabolic and anabolic processes, respectively. The oxidation of NADPH to NADP+ initiates a cascade of reactions, where a network of molecules is implicated. The molecules of this cascade form a network with eminent translational potential in redox metabolism. A special point of interest is that spectrophotometric assays have been developed both for NADH/NADPH and the molecules directly regulated by them. Therefore, crucial molecules of the NADPH-dependent redox network can be measured, and the results can be used to assess the bioenergetic and/or oxidative stress status. The main aim of this review is to collectively present the NADPH-related molecules, namely NADPH, NADH, NAD+ kinase, NADPH oxidase, peroxiredoxin, thioredoxin, thioredoxin reductase, and nitric oxide synthase, that can be measured in blood and tissues with the use of a spectrophotometer, which is probably the most simple, inexpensive and widely used tool in biochemistry. We are providing the researchers with reliable and valid spectrophotometric assays for the measurement of the most important biomarkers of the NADPH network in blood and other tissues, thus allowing the opportunity to follow the redox changes in response to a stimulus.  相似文献   

17.
A commonly used enzymatic recycling assay for pyridine nucleotides has been adapted to directly measure the NAD+/NADH redox ratio in Drosophila melanogaster. This method is also suitable for quantification of NADP+ and NADPH. The addition of a coupling reaction removing acetaldehyde produced from the alcohol dehydrogenase (ADH) reaction was shown to improve the linearity of NAD(H) assay. The advantages of this assay method are that it allows the determination of both NAD+ and NADH simultaneously while keeping enzymatic degradation of pyridine nucleotides minimal and also achieving better sensitivity. This method was used to determine the redox ratio of D. melanogaster and validated substantial decrease of redox ratio during starvation.  相似文献   

18.
We measured both pyridine nucleotide levels and ribonucleotide reductase-specific activity in Yoshida ascites hepatoma cells as a function of growth in vivo and during recruitment from non-cycling to cycling state in vitro. Oxidized nicotinamide adenine dinucleotide (NAD+) and reduced nicotinamide adenine dinucleotide (NADP) levels remained unchanged during tumour growth, while NADP+ and reduced nicotinamide adenine dinucleotide phosphate (NADPH) levels were very high in exponentially growing cells and markedly decreased in the resting phase. Ribonucleotide reductase activity paralleled NADP(H) (NADP+ plus NADPH) intracellular content. The concomitant increase in both NADP(H) levels and ribonucleotide reductase activity was also observed during G1-S transition in vitro. Cells treated with hydroxyurea showed a comparable correlation between the pool size of NADP(H) and ribonucleotide reductase activity. On the basis of these findings, we suggest that fluctuations in NADP(H) levels and ribonucleotide reductase activity might play a critical role in cell cycle regulation.  相似文献   

19.
Pyridine nucleotides were determined in lateral roots of iron-deficient and iron-sufficient Phaseolus vulgaris L. cv Prelude. In iron-deficient plants, total NADP per gram fresh weight and the NADPH/NADP+ ratio were twice the values found in iron-sufficient plants. The NADPH/NADP+ ratio in iron-deficient plants was considerably lowered after a 2 minute incubation in 1 millimolar ferricyanide. Total NAD was not influenced by growth conditions and was mainly present in oxidized form.

These results indicate that NADPH is the electron donor for the high FeIII reduction activity found in iron-deficient roots, a process that is part of the Fe-uptake mechanism.

  相似文献   

20.
Redox interconversion of glutathione reductase was studiedin situ withS. cerevisiae. The enzyme was more sensitive to redox inactivation in 24 hour-starved cells than in freshly-grown ones. While 5 μM NADPH or 100 μM NADH caused 50% inactivation in normal cells in 30 min, 0.75 μM NADPH or 50 μM NADH promoted a similar effect in starved cells. GSSG reactivated the enzyme previously inactivated by NADPH, ascertaining that the enzyme was subjected to redox interconversion. Low EDTA concentrations fully protected the enzyme from NADPH inactivation, thus confirming the participation of metals in such a process. Extensive inactivation was obtained in permeabilized cells incubated with glucose-6-phosphate or 6-phosphogluconate, in agreement with the very high specific activities of the corresponding dehydrogenases. Some inactivation was also observed with malate, L-lactate, gluconate or isocitrate in the presence of low NADP+ concentrations. The inactivation of yeast glutathione reductase has also been studiedin vivo. The activity decreased to 75% after 2 hours of growth with glucono-δ-lactone as carbon source, while NADPH rose to 144% and NADP+ fell to 86% of their initial values. Greater changes were observed in the presence of 1.5 μM rotenone: enzymatic activity descended to 23% of the control value, while the NADH/NAD+ and NADPH/NADP+ ratios rose to 171% and 262% of their initial values, respectively. Such results indicate that the lowered redox potential of the pyridine nucleotide pool existing when glucono-δ-lactone is oxidized promotesin vivo inactivation of glutathione reductase.  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