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 796 毫秒
1.
The effects of t-butyl hydroperoxide on glutathione and NADPH and the respiratory burst (an NADPH-dependent function) in rat alveolar macrophages was investigated. Alveolar macrophages were exposed for 15 min to t-butyl hydroperoxide in the presence or absence of added glucose. Cells were then assayed for concanavalin A-stimulated O2 production or for NADPH, NADP, reduced glutathione, glutathione disulfide, glutathione released into the medium and glutathione mixed disulfides. Exposure of rat alveolar macrophages to 1 X 10(-5) M t-butyl hydroperoxide causes a loss of concanavalin A-stimulated superoxide production (the respiratory burst) that can be prevented or reversed by added glucose. Cells incubated without glucose had a higher oxidation state of the NADPH/NADP couple than cells incubated with glucose. With t-butyl hydroperoxide, NADP rose to almost 100% of the NADP + NADPH pool; however, addition of glucose prevented this alteration of the NADPH oxidation state. Cells exposed to 1 X 10(-5) M t-butyl hydroperoxide in the absence of glucose showed a significant increase in the percentage GSSG in the GSH + GSSG pool and increased glutathione mixed disulfides. These changes in glutathione distribution could also be prevented or reversed by glucose. With 1 X 10(-4) M t-butyl hydroperoxide, changes in glutathione oxidation were not prevented by glucose and cells were irreversibly damaged. We conclude that drastic alteration of the NADPH/NADP ratio does not itself reflect toxicity and that significant alteration of glutathione distribution can also be tolerated; however, when oxidative stress exceeds the ability of glucose to prevent alterations in oxidation state, irreversible damage to cell function and structure may occur.  相似文献   

2.
The steady-state levels and redox states of pyridine nucleotide pools have been studied in yeast as a function of external growth conditions. Yeast grown aerobically on 0.8% glucose show two distinct phases of logarithmic growth, a first phase utilizing glucose with ethanol accumulation, and a second phase utilizing ethanol. During growth on glucose, the size of the NADP pool (NADP+ + NADPH) is maintained at approximately 12% the size of the NAD pool (NAD+ + NADH). Upon exhaustion of glucose, the mechanism(s) that maintain the levels of NADP relative to NAD are altered, resulting in a rapid 2- to 2.5-fold decrease in the size of the NADP pool relative to the size of the NAD pool. The lower levels of NADP are maintained during growth on ethanol. The NAD pool is approximately 50% NADH during both the glucose and ethanol phases of growth, while the NADP pool is approximately 67 and 48% NADPH during the glucose and ethanol phases of growth, respectively. Rapid media transfer experiments show that the decrease in NADP is reversible, that it does not require the net synthesis of pyridine nucleotide or protein, and that changes in the size of the NADP pool relative to the total pyridine nucleotide pool are correlated with changes in the redox state of the NADP pool.  相似文献   

3.
An isocratic reverse-phase high-performance liquid chromatography method for the separation and quantitation of total pyridine dinucleotides in hepatocyte cultures is described. Cells are extracted with cold 3 M perchloric acid or 0.5 N sodium hydroxide containing 50% (v/v) ethanol and 35% cesium chloride for the determination of the oxidized or reduced pyridine dinucleotides, respectively. Pyridine dinucleotides in the neutralized extracts were separated on an Excellopak ODS C18 (4.6 X 150 mm) column with 0.1 M potassium phosphate, pH 6.0, containing 3.75% methanol as the mobile phase. NAD+ and NADP+ were detected spectrophotometrically at 254 nm. The response was linear from 5 to 4000 pmol with recoveries of NAD+ and NADP+ of 98 and 101.1%, respectively. NADH and NADPH were monitored fluorometrically by activation at 370 nm and emission in the 400-700 nm range. The reduced pyridine dinucleotides had a linear response from 7.5 to 60 pmol with recoveries of NADH and NADPH of 99.4 and 101.3%, respectively. The coefficients of variation for all of the pyridine dinucleotide standards were less than 3.5%.  相似文献   

4.
The interactions of calcium with NAD+, NADH, NADP+ and NADPH in a 50% (by volume) methanol/water mixture (pH 7, 25 degrees C) were studied by calorimetry. The association constants for 1:1 complex formation were found to be 6.6 +/- 0.2, 270 +/- 76, 18 +/- 3 and 98 +/- 10 for NAD+, NADH, NADP+ and NADPH, respectively. Comparing these to the association constants for an aqueous system reveals that as the polarity of the solvent system is decreased the interactions involving NAD+, NADP+ and NADPH are all decreased. In contrast, the interaction involving NADH is markedly increased. All the interactions were found to be endothermic.  相似文献   

5.
In recent years, pyridine nucleotides NAD(H) and NADP(H) have been established as an important molecules in physiological and pathophysiological signaling and cell injury pathways. Protein modification is catalyzed by ADP-ribosyl transferases that attach the ADP-ribose moiety of NAD+ to specific aminoacid residues of the acceptor proteins, with significant changes in the function of these acceptors. Mono(ADP-ribosyl)ation reactions have been implicated to play a role both in physiological responses and in cellular responses to bacterial toxins. Cyclic ADP-ribose formation also utilizes NAD+ and primarily serves as physiological, signal transduction mechanisms regulating intracellular calcium homeostasis. In pathophysiological conditions associated with oxidative stress (such as various forms of inflammation and reperfusion injury), activation of the nuclear enzyme poly(ADP-ribose) polymerase (PARP) occurs, with subsequent, substantial fall in cellular NAD+ and ATP levels, which can determine the viability and function of the affected cells. In addition, NADPH oxidases can significantly affect the balance and fate of NAD+ and NADP in oxidatively stressed cells and can facilitate the generation of various positive feedback cycles of injury. Under severe oxidant conditions, direct oxidative damage to NAD+ has also been reported. The current review focuses on PARP and on NADPH oxidases, as pathophysiologically relevant factors in creating disturbances in the cellular pyridine nucleotide balance. A separate section describes how these mechanisms apply to the pathogenesis of endothelial cell injury in selected cardiovascular pathophysiological conditions.  相似文献   

6.
Xylose reductase has been purified to apparent homogeneity from cell extracts of the fungus Cryptococcus flavus grown on D-xylose as carbon source. The enzyme, the first of its kind from the phylum Basidiomycota, is a functional dimer composed of identical subunits of 35.3 kDa mass and requires NADP(H) for activity. Steady-state kinetic parameters for the reaction, D-xylose + NADPH + H(+)<--> xylitol + NADP(+), have been obtained at pH 7.0 and 25 degrees C. The catalytic efficiency for reduction of D-xylose is 150 times that for oxidation of xylitol. This and the 3-fold tighter binding of NADPH than NADP(+) indicate that the enzyme is primed for unidirectional metabolic function in microbial physiology. Kinetic analysis of enzymic reduction of aldehyde substrates differing in hydrophobic and hydrogen bonding capabilities with binary enzyme-NADPH complex has been used to characterize the substrate-binding pocket of xylose reductase. Total transition state stabilization energy derived from bonding with non-reacting sugar hydroxyls is approximately 15 kJ/mol, with a major contribution of 5-8 kJ/mol made by interactions with the C-2(R) hydroxy group. The aldehyde binding site is approximately 1.2 times more hydrophobic than n-octanol and can accommodate linear alkyl chains of 相似文献   

7.
Akira Kusai  Tateo Yamanaka 《BBA》1973,292(3):621-633
A highly purified preparation of an NAD(P) reductase was obtained from Chlorobium thiosulfatophilum and some of its properties were studied. The enzyme possesses FAD as the prosthetic group, and reduces benzyl viologen, 2,6-dichloro-phenolindophenol and cytochromes c, including cytochrome c-555 (C. thiosulfato-philum), with NADPH or NADH as the electron donor. It reduces NADP+ or NAD+ photosynthetically with spinach chloroplasts in the presence of added spinach ferredoxin. It reduces the pyridine nucleotides with reduced benzyl viologen. The enzyme also shows a pyridine nucleotide transhydrogenase activity. In these reactions, the type of pyridine nucleotide (NADP or NAD) which functions more efficiently with the enzyme varies with the concentration of the nucleotide used; at concentrations lower than approx. 1.0 mM, NADPH (or NADP+) is better electron donor (or acceptor), while NADH (or NAD+) is a better electron donor (or acceptor) at concentrations higher than approx. 1.0 mM. Reduction of dyes or cytochromes c catalysed by the enzyme is strongly inhibited by NADP+, 2′-AMP and and atebrin.  相似文献   

8.
Carbon tetrachloride metabolism was examined in hepatic microsomes isolated from control and phenobarbital-treated Sprague-Dawley rats to determine the mechanism of pyridine nucleotide synergism. An NADPH generator increased metabolism two fold as determined by lipid peroxidation. Addition of NADP to the reaction system did not alter the maximum velocity, but did decrease the Km for NADPH from 61 μM to 7.6 μM in control and from 21 μM to 6.3 mM PB microsomes. Addition of NAD+ produced an increase in metabolism similar to NADH. Substrates and competitive inhibitors of nucleotide pyrophosphatase also enhanced CCl4 metabolism. A high correlation (r=0.947) was indicated between the percent inhibition of nucleotide pyrophosphatase and the percent synergism of NADPH-catalyzed CCl4 metabolism. Thus, pyridine nucleotiode synergism in CCl4 metabolism appears to result from the increased availability of NADPH produced by a decreased degradation of the NADPH by the nucleotide pyrophosphatase.  相似文献   

9.
Interaction of ferredoxin-NADP+ reductase from Anabaena with its substrates   总被引:1,自引:0,他引:1  
The interaction of ferredoxin-NADP+ reductase from the cyanobacterium Anabaena variabilis with its substrates, NADP+ and ferredoxin, has been studied by difference absorption spectroscopy. Several structural analogs of NADP+ have been shown to form complexes the stabilities of which are strongly dependent on the ionic strength of the medium. In most cases the binding energy of these complexes and their difference absorption spectra are similar to those reported for the spinach enzyme. However, NADP+ perturbs the absorption spectra of the Anabaena and spinach enzymes in a different way. This difference has been shown to be related to the binding of the nicotinamide ring of NADP+ to the enzymes. These results are interpreted as being due to a different nicotinamide binding site in the two reductases. The enthalpic and entropic components of the Gibbs energy of formation of the NADP+ complex have been estimated. An increase in entropy on NADP+ binding seems to be the main source of stability for the complex. A shift of approximately 40 mV in the redox potential of the couple NADP+/NADPH has been observed to occur upon binding of NADP+ to the oxidized enzyme. This allows us to calculate the binding energy between the reductase and NADPH. The ability of the reductase, ferredoxin, and NADP+ to form a ternary complex indicates that the protein carrier binds to the reductase through a different site than that of the pyridine nucleotide.  相似文献   

10.
Circular dichroism has been used to monitor the binding of pyridine nucleotide cofactors to enzyme-folate analog complexes of dihydrofolate reductase from Escherichia coli B (MB 1428). The enzyme binds one molar equivalent of many folate analogs and two molar equivalents of several pyridine nucleotide cofactors. The apo-enzyme has very low optical activity. The binding of folate analogs including folate, dihydrofolate, methotrexate, trimethoprim and pyrimethamine induce large Cotton effects. Pyridine nucleotides when bound to the enzyme-folate analog complexes also induce new optically active bands; all the effects being due to the first molar equivalent of cofactor bound. NADPH and NADP+ induce very similar bands when bound to the enzyme-methotrexate complex suggesting that the geometry of the complexes formed are very similar. The oxidized and reduced cofactor likewise have similar effects on the enzyme-folate complex. However, NADPH and NADP+ addition to both the enzyme-trimethoprim and enzyme-pyrimethamine complexes have significantly different effects on the circular dichroism spectra, suggesting that the inhibitors which are less homologous to the natural dihydrofolate substrate allow more conformational freedom in the enzyme-inhibitor-cofactor complex. In most cases the prior binding of the folate analog greatly increases the binding of the first molar equivalent of cofactor so that at concentrations of approx. 5-20 muM the binding appears stoichiometric. Pyrimethamine is an exception in that it apparently has no effect on the binding of NADPH to the enzyme.  相似文献   

11.
Incubation of malic enzyme (L-malate:NADP+ oxidoreductase (oxaloacetate-decarboxylating), EC 1.1.1.40) with ethoxyformic anhydride caused the time-dependent loss of its ability to catalyze reactions requiring the nucleotide cofactor NADP+ or NADPH, such as the oxidative decarboxylase, the NADP+ - stimualted oxalacetate decarboxylase, the pyruvate reductase, and the pyruvate-medium proton exchange activities. Similar loss of oxidative decarboxylase and pyruvate reductase activities was affected by photo-oxidation in the presence of rose bengal. The inactivation of oxidative decarboxylase activity by ethoxyformic anhydride was accompanied by the reaction of greater than or equal to 2.3 histidyl residues per enzyme site and was strongly inhibited by NADP+. Ethoxyformylation also impaired the ability of malic enzyme to bind NADP+ or NADPH. These results support the involvement of histidyl residue(s) at the nucleotide binding site of malic enzyme.  相似文献   

12.
Anaerobic reduction of the flavoprotein adrenodoxin reductase with NADPH yields a spectrum with long wavelength absorbance, 750 nm and higher. No EPR signal is observed. This spectrum is produced by titration of oxidized adrenodoxin reductase with NADPH, or of dithionite-reduced adrenodoxin reductase with NADP+. Both titrations yield a sharp endpoint at 1 NADP(H) added per flavin. Reduction with other reductants, including dithionite, excess NADH, and catalytic NADP+ with an NADPH generating system, yields a typical fully reduced flavin spectrum, without long wavelength absorbance. The species formed on NADPH reduction appears to be a two-electron-containing complex, with a low dissociation constant, between reduced adrenodoxin reductase and NADP+, designated ARH2-NADP+. Titration of dithionite-reduced adrenodoxin reductase with NADPH also produces a distinctive spectrum, with a sharp endpoint at 1 NADPH added per reduced flavin, indicating formation of a four-electron-containing complex between reduced adrenodoxin reductase and NADPH. Titration of adrenodoxin reductase with NADH, instead of NADPH, provides a curved titration plot rather than the sharp break seen with NADPH, and permits calculation of a potential for the AR/ARH2 couple of -0.291 V, close to that of NAD(P)H (-0.316 V). Oxidized adrenodoxin reductase binds NADP+ much more weakly (Kdiss=1.4 X 10(-5) M) than does reduced adrenodoxin reductase, with a single binding site. The preferential binding of NADP+ to reduced enzyme permits prediction of a more positive oxidation-reduction potential of the flavoprotein in the presence of NADP+; a change of about + 0.1 V has been demonstrated by titration with safranine T. From this alteration in potential, a Kdiss of 1.0 X 10(-8) M for binding of NADP+ to reduced adrenodoxin reductase is calculated. It is concluded that the strong binding of NADP+ to reduced adrenodoxin reductase provides the thermodynamic driving force for formation of a fully reduced flavoprotein form under conditions wherein incomplete reduction would otherwise be expected. Stopped flow studies demonstrate that reduction of adrenodoxin reductase by equimolar NADPH to form the ARH2-NADP+ complex is first order (k=28 s-1). When a large excess of NADPH is used, a second apparently first order process is observed (k=4.25 s-1), which is interpreted as replacement of NADPH for NADP+ in the ARH2-NADP+ complex. Comparison of these rate constants to catalytic flavin turnover numbers for reduction of various oxidants by NADPH, suggests an ordered sequential mechanism in which reduction of oxidant is accomplished by the ARH2-NADP+ complex, followed by dissociation of NADP+. The absolute dependence of NADPH-cytochrome c reduction on both adrenodoxin reductase and adrenodoxin is confirmed...  相似文献   

13.
NADP+, NADPH and glucose 6-phosphate dehydrogenase were determined in the cerebral cortex of mice exposed to high O2 pressure for 0, 8 and 16 min. These time intervals corresponded to 0, 50 and 100% of the CT50 (the time taken for 50% of the mice to convulse). Cerebral NADP+, NADPH and glucose 6-phosphate dehydrogenase also were determined in O2-exposed mice exhibiting hyperactivity, convulsions, and in mice killed 10s after convulsions. Similar increases in cortical NADP+ and decreases in NADPH were found in mice exposed to 610kPa (6 atm.) of 100% O2 for 0, 50 and 100% of the CT50, during hyperactivity, onset of seizure and 10s after convulsions. The NADP+/NADPH ratio increased approx. 25% at 0% of the CT50, and remained at this increased value at all O2-exposure periods including the hyperactive state, onset of seizure and 10s after convulsions. Identical changes in cerebral NADP+ , NADPH and the NADP+/NADPH ratio were found in mice exposed for 16min to 100% O2 at 100, 350 or 610kPa. No change in cerebral glucose 6-phosphate dehydrogenase was found in mice exposed to 610kPa of 100% O2 during the various stages of O2 toxicity. Only in the 10s post-convulsive group was a statistically significant decrease in glucose 6-phosphate dehydrogenase observed. Disulfiram [bis(diethylthiocarbamoyl) disulphide], an effective O2-protective agent, did not prevent the O2-induced increase in cerebral NADP+ and the NADP+/NADPH ratio, or decrease in NADPH.  相似文献   

14.
A Aulabaugh  J V Schloss 《Biochemistry》1990,29(11):2824-2830
N-Hydroxy-N-isopropyloxamate (IpOHA) is an exceptionally potent inhibitor of the Escherichia coli ketol-acid reductoisomerase. In the presence of Mg2+ or Mn2+, IpOHA inhibits the enzyme in a time-dependent manner, forming a nearly irreversible complex. Nucleotide, which is essential for catalysis, greatly enhances the binding of IpOHA by the reductoisomerase, with NADPH (normally present during the enzyme's rearrangement step, i.e., conversion of a beta-keto acid into an alpha-keto acid, in either the forward or reverse physiological reactions) being more effective than NADP. In the presence of Mg2+ and NADPH, IpOHA appears to bind to the enzyme in a two-step mechanism, with an initial inhibition constant of 160 nM and a maximum rate of formation of the tight, slowly reversible complex of 0.57 min-1 (values that give an association rate of IpOHA, at low concentration, of 5.9 X 10(4) M-1 s-1). The rate of exchange of [14C]IpOHA from an enzyme-[14C]IpOHA-Mg2(+)-NADPH complex with exogenous, unlabeled IpOHA has a half-time of 6 days (150 h). This dissociation rate (1.3 X 10(-6) s-1) and the association rate determined by inactivation kinetics define an overall dissociation constant of 22 pM. By contrast, in the presence of Mn2+ and NADPH, the corresponding association and dissociation rates for IpOHA are 8.2 X 10(4) M-1 s-1 and 3.2 X 10(-6) s-1 (half-time = 2.5 days), respectively, which define an overall dissociation constant of 38 pM. In the presence of NADP or in the absence of nucleotide (both in the presence of Mg2+), the enzyme-IpOHA complex is far more labile, with dissociation half-times of 28 and 2 h, respectively. In the absence of Mg2+ or Mn2+, IpOHA does not exhibit time-dependent inhibition of the reductoisomerase.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

15.
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   

16.
Dormant spores of Bacillus megaterium contained no detectable reduced nicotinamide adenine dinucleotide (NADH) or reduced nicotinamide adenine dinucleotide phosphate (NADPH) despite significant levels of the oxidized forms of these nucleotides (NAD and NADP). During the first minutes of spore germination there was rapid accumulation of NADH and NADPH. However, this accumulation followed the fall in optical density that is characteristic of the initiation of spore germination. Accumulation of NADH and NADPH early in germination was not blocked by fluoride or cyanide, and it occurred even when germination was carried out in the absence of an exogenous source of reducing power. In addition to pyridine nucleotide reduction, de novo synthesis also began early in germination as the pyridine nucleotide levels increased to those found in growing cells. Midlog-phase cells grown in several different media had 20 to 35 times as much total pyridine nucleotide as did dormant spores. However, as growth and sporulation proceeded, the NADH plus NAD level fell four- to fivefold whereas the NADPH plus NADP level fell by a lesser amount. From min 10 of spore germination until midway through sporulation the value for the ratio of NADH/NAD is about 0.1 (0.03 to 0.18) while the ratio of NADPH/ANDP is about 1.4 (0.3 to 2.4). Comparison of these ratios in log-phase versus stationary phase (sporulation) growth in all three growth media tested did not reveal any common pattern of changes.  相似文献   

17.
The pyridine nucleotides NAD and NADP play vital roles in metabolic conversions as signal transducers and in cellular defence systems. Both coenzymes participate as electron carriers in energy transduction and biosynthetic processes. Their oxidized forms, NAD+ and NADP+, have been identified as important elements of regulatory pathways. In particular, NAD+ serves as a substrate for ADP-ribosylation reactions and for the Sir2 family of NAD+-dependent protein deacetylases as well as a precursor of the calcium mobilizing molecule cADPr (cyclic ADP-ribose). The conversions of NADP+ into the 2'-phosphorylated form of cADPr or to its nicotinic acid derivative, NAADP, also result in the formation of potent intracellular calcium-signalling agents. Perhaps, the most critical function of NADP is in the maintenance of a pool of reducing equivalents which is essential to counteract oxidative damage and for other detoxifying reactions. It is well known that the NADPH/NADP+ ratio is usually kept high, in favour of the reduced form. Research within the past few years has revealed important insights into how the NADPH pool is generated and maintained in different subcellular compartments. Moreover, tremendous progress in the molecular characterization of NAD kinases has established these enzymes as vital factors for cell survival. In the present review, we summarize recent advances in the understanding of the biosynthesis and signalling functions of NAD(P) and highlight the new insights into the molecular mechanisms of NADPH generation and their roles in cell physiology.  相似文献   

18.
The substrate analogue 3-bromo-2-ketoglutarate reacts with pig heart NADP+-dependent isocitrate dehydrogenase to yield partially inactive enzyme. Following 65% inactivation, no further inactivation was observed. Concomitant with this inactivation, incorporation of 1 mol of reagent/mol of enzyme dimer was measured. The dependence of the inactivation rate on bromoketoglutarate concentration is consistent with reversible binding of reagent (KI = 360 microM) prior to irreversible reaction. Manganous isocitrate reduces the rate of inactivation by 80% but does not provide complete protection even at saturating concentrations. Complete protection is obtained with NADP+ or the NADP+-alpha-ketoglutarate adduct. By modification with [14C]bromoketoglutarate or by NaB3H4 reduction of modified enzyme, a single major radiolabeled tryptic peptide was obtained by high performance liquid chromatography with the sequence: Asp-Leu-Ala-Gly-X-Ile-His-Gly-Leu-Ser-Asn-Val-Lys. Evidence in the following paper (Bailey, J.M., Colman, R.F. (1987) J. Biol. Chem. 262, 12620-12626) indicates that X is glutamic acid. Enzyme modified at the coenzyme site by 2-(bromo-2,3-dioxobutylthio)-1,N(6)-ethenoadenosine 2',5'-biphosphate in the presence of manganous isocitrate is not further inactivated by bromoketoglutarate. Bromoketoglutarate-modified enzyme exhibits a stoichiometry of binding isocitrate and NADPH equal to 1 mol/mol of enzyme dimer, half that of native enzyme. These results indicate that bromoketoglutarate modifies a residue in the nicotinamide region of the coenzyme site proximal to the substrate site and that reaction at one catalytic site of the enzyme dimer decreases the activity of the other site.  相似文献   

19.
The reaction process of adrenodoxin reductase with NADPH and NADH were investigated. The appearance of new intermediate with a broad absorption band at around 520 nm has been detected by rapid-scan stopped-flow spectrophotometry. Although the formation of this intermediate is more rapid with NADPH than with NADH, the rates of the subsequent decay to the fully reduced state are almost identical (Kobs values were 20.5 and 16.0s-1). These results indicate that the new intermediate is the complex formed between the oxidized enzyme and reduced pyridine nucleotide (enzyme-substrate complex), and that subsequent decay of the intermidiate is caused by a two-electron transfer process from the reduced pyridine nucleotide to the enzyme flavin. On the other hand, spectral and kinetic properties in the steady state of the reoxidation reaction of the enzyme reduced with NADPH and NADH were somewhat different. The rate of reoxidation of the enzyme under aerobic conditions from the reduced state to the oxidized state was 6.5 times faster when a 10-fold molar excess of NADH was used than when NADPH of the same concentration was used. This result is consistent with the fact that the NADH-dependent oxidase activity was 6.4 times greater than that dependent on NADPH. During reoxidation of the reduced enzyme under aerobic conditions in the presence of an excess of NADPH or NADH, the EPR spectra indicated the formation of the flavin semiquinone radical species. Similarly, the formation of semiquinone was observed in the absorption spectrum with either NADPH or NADH under the same conditions as in the EPR measurement. The intensity of the semiquinone signal on EPR was considerably smaller with NADH than with NADPH. These results suggest that NADP+ complex with the enzyme semiquinone protects the radical from oxidation by oxygen to a greater extent than NAD+, and consequently the semiquinone is easier to detect with NADPH than with NADH.  相似文献   

20.
Concentrations of adenine and pyridine nucleotides and the associated charge values were examined in extracts of mycelium of Penicillium notatum during vegetative growth and reproductive development promoted by the addition of Ca2+ (10 mmol dm-3). The significant increase in adenylate energy charge promoted by Ca2+ was due to a fall in intracellular AMP and a concomitant rise in ATP concentration. Intracellular concentrations of NADH and NAD fell within 1 h of the addition of Ca2+. The catabolic reduction charge was unchanged by Ca2+ whilst the anabolic reduction charge increased in Ca2+-induced mycelium due to lowered intracellular NADP concentration. Reduced concentration of NADPH in Ca2+-induced mycelium, relative to the vegetative controls, lowered the phosphorylated nucleotide fraction. The results are discussed in relation to metabolic economy during morphogenesis in P. notatum.  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