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.
The concentration of NADP + NADPH in resting human neutrophils has been measured to be 24.0 +/- 2.7 X 10(-18) mol/cell. Upon activation with opsonized zymosan A, phorbol myristic acetate or N-formylmethionyl-leucylphenylalanine, neutrophil NADP + NADPH pools increase to 80.5, 84.0, and 54.0 X 10(-18) mol/cell, respectively. These increases in pyridine nucleotide concentration are blocked by the addition of the calcium antagonist 8-(N,N-dimethylamino)-octyl-3,4,5-trimethoxybenzoate hydrochloride, while calcium ionophore A23187, in the presence of calcium, will trigger the increase in the absence of other stimuli. Calmodulin antagonists trifluoperazine and N-(6-aminohexyl)-5-chloro-1-naphthalene sulfonamide also inhibit stimulus-induced increases in the NADP + NADPH pool. These studies are interpreted as suggesting a role for calcium and calmodulin, and possibly protein kinase C in the regulation of pyridine nucleotide concentration in the activated neutrophil.  相似文献   

2.
An analysis of the kinetic mechanism of the microsomal NADPH-linked progesterone 5 alpha-reductase obtained from female rat anterior pituitaries was performed. Initial velocity, product inhibition and dead-end inhibition studies indicate that the kinetic mechanism for the progesterone 5 alpha-reductase is equilibrium ordered sequential. Analysis of the initial velocity data resulted in intersecting double reciprocal plots suggesting a sequential mechanism [apparent Km(progesterone) = 88.2 +/- 8.2 nM; apparent Kia(NADPH) = 7.7 +/- 1.1 microM]. Furthermore, the plot of 1/v vs 1/progesterone intersected on the ordinate which is indicative of an equilibrium ordered mechanism. Additional support for ordered substrate binding was provided by the product inhibition studies with NADPH versus NADP and progesterone versus NADP. NADP is a competitive inhibitor versus NADPH (apparent Kis = 7.8 +/- 1.0 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 9.85 +/- 2.1 microM and apparent Kii = 63.2 +/- 12.5 microM). These inhibition patterns suggest that NADPH binds prior to progesterone. In sum, these kinetic studies indicate that NADPH binds to the microsomal enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  相似文献   

3.
The kinetic mechanism of the hypothalamic NADPH-linked progesterone 5 alpha-reductase from female rats was determined to be equilibrium ordered sequential by initial velocity, product inhibition and dead-end inhibition studies. Analysis of the initial velocity data resulted in intersecting double reciprocal plots indicating a sequential mechanism (apparent Km (progesterone) = 95.4 +/- 4.5 nM; apparent Kia(NADPH) = 9.9 +/- 0.7 microM). The plot of 1/v vs 1/progesterone intersected on the ordinate which is consistent with an equilibrium ordered mechanism. Ordered addition of the substrates was also supported by product inhibition studies with NADP versus NADPH and NADP versus progesterone. NADP is a competitive inhibitor versus NADPH (apparent Kis = 4.3 +/- 1.3 microM) and a noncompetitive inhibitor versus progesterone (apparent Kis = 31.9 +/- 1.4 microM and apparent Kii = 145.4 +/- 15.5 microM). These inhibition patterns show that NADPH binds prior to progesterone. Taken together, these analyses indicate that the cofactor, NADPH, binds to the enzyme in rapid equilibrium and preferentially precedes the binding of progesterone.  相似文献   

4.
1. NADPH-cytochrome c reductase was solubilized with bromelain and purified about 400-fold from sucrose/pyrophosphate-washed microsomal fractions from southern armyworm (Spodoptera eridania) larval midguts. 2. The enzyme has a mol.wt. of 70 035 +/- 1300 and contained 2 mol of flavin/mol of enzyme consisting of almost equimolar amounts of FMN and FAD. 3. Aerobic titration of the enzyme with NADPH caused the formation of a stable half-reduced state at 0.5 mol of NADPH/mol of flavin. 4. Kinetic analysis showed that the reduction of cytochrome c proceeded by a Bi Bi Ping Pong mechanism. 5. Apparent Km values for NADPH and cytochrome c and Ki values for NADP+ and 2'-AMP were considerably higher for the insect reductase than for the mammalian liver enzyme. 6. These are discussed in relation to possible differences in the active sites of the enzymes.  相似文献   

5.
NADPH and NADP+ levels were measured in rat lens from normal controls, from galactose-fed and diabetic rats during the first week of cataract formation. The level of NADPH in normal rat lens was determined to be 12.3 +/- 0.4 nmol/g wet weight, and that of NADP+ 4.6 +/- 0.2 nmol/g wet weight. In early cataract formation NADPH levels decreased rapidly during the first 2 days and then remained stable at 76% of control for galactose-fed and 84% for diabetic rats. NADP+ levels increased by 38% of control for galactose-fed and 54% for diabetic rats. Calculated NADPH/NADP+ ratios dropped from 3.36 +/- 0.21 to 1.86 +/- 0.16 in galactose fed rats, and from 2.81 +/- 0.15 to 1.61 +/- 0.16 in diabetic rats (P less than 0.001 for both experimental groups). These data are consistent with rapid NADPH oxidation during onset of lens cataracts. No significant changes in aldose reductase enzymatic activity levels were observed in either the galactosemic or the diabetic rats during the times measured.  相似文献   

6.
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   

7.
The thionicotinamide analogues of NAD+ and NADP+ were shown to be good alternative coenzymes for bovine glutamate dehydrogenase, with similar affinity and approx. 40% of the maximum velocity obtained with the natural coenzymes. Both thionicotinamide analogues show non-linear Lineweaver-Burk plots, which with the natural coenzymes have been attributed to negative co-operativity. Since the reduced thionicotinamide analogues have an isosbestic point at 340nm and have an absorption maximum at 400nm, it is possible to monitor reduction of natural coenzyme and thionicotinamide analogue simultaneously by dual-wavelength spectroscopy. When glutamate dehydrogenase is presented with NADP+ and thio-NADP+ simultaneously, the enzyme oligomer senses saturation of its coenzyme-binding sites irrespective of the exact nature of the coenzyme and locks the oligomer into its highly saturated form even when low saturation of the monitored coenzyme is present. These experiments substantiate the suggestion that glutamate dehydrogenase shows negative co-operativity in its catalytically active form.  相似文献   

8.
The binding of NADP+ to dihydrofolate reductase (EC 1.5.1.3) in the presence and absence of substrate analogs has been studied using 1H and 13C nuclear magnetic resonance (NMR). NADP+ binds strongly to the enzyme alone and in the presence of folate, aminopterin, and methotrexate with a stoichiometry of 1 mol of NADP+/mol of enzyme. In the 13C spectra of the binary and ternary complexes, separate signals were observed for the carboxamide carbon of free and bound [13CO]NADP+ (enriched 90% in 13C). The 13C signal of the NADP+-reductase complex is much broader than that in the ternary complex with methotrexate because of exchange line broadening on the binary complex signal. From the difference in line widths (17.5 +/- 3.0 Hz) an estimate of the dissociation rate constant of the binary complex has been obtained (55 +/- 10 sec-1). The dissociation rate of the NADP+-reductase complex is not the rate-limiting step in the overall reaction. In the various complexes studied large 13C chemical shifts were measured for bound [13CO]NADP+ relative to free NADP+ (upfield shifts of 1.6-4.3 ppm). The most likely origin of the bound shifts lies in the effects on the shieldings of electric fields from nearby charged groups. For the NADP+-reductase-folate system two 13C signals from bound NADP+ are observed indicating the presence of more than one form of the ternary complex. The IH spectra of the binary and ternary complexes confirm both the stoichiometry and the value of the dissociation rate constant obtained from the 13C experiments. Substantial changes in the IH spectrum of the protein were observed in the different complexes and these are distinct from those seen in the presence of NADPH.  相似文献   

9.
A divalent cation electrode was used to measure the stability constants (association constants) for the magnesium and manganese complexes of the substrates for the NADP+-specific isocitrate dehydrogenase (EC 1.1.1.42) from pea stems. At an ionic strength of 26.5 mM and at pH 7.4 the stability constants for the Mg2+-isocitrate and Mg2+-NADP+ complexes were 0.85 +/- 0.2 and 0.43 +/- 0.04 mM-1 respectively and for the Mn2+-isocitrate and Mn2+-NADP+ complexes they were 1.25 +/- 0.07 and 0.75 +/- 0.09 mM-1 respectively. At the same ionic strength but at pH 6.0 the Mg2+-NADPH and Mn2+-NADPH complexes had stability constants of 0.95 +/- 0.23 and 1.79 +/- 0.34 mM-1 respectively. Oxalosuccinate and alpha-ketoglutarate do not form measureable complexes under these conditions. Saturation kinetics of the enzyme with respect to isocitrate and metal ions are consistent with the metal-isocitrate complex being the substrate for the enzyme. NADP+ binds to the enzyme in the free form. Saturation kinetics of NADPH and Mn2+ indicate that the metal-NADPH complex is the substrate in the reverse reaction. In contrast the pig heart enzyme appears to bind free NADPH and Mn2+. A scheme for the reaction mechanism is presented and the difference between the reversibility of the NAD+ and NADP+ enzyme is discussed in relation to the stability of the NADH and NADPH metal complexes.  相似文献   

10.
NAD kinase was purified 93-fold from Escherichia coli. The enzyme was found to have a pH optimum of 7.2 and an apparent Km for NAD+, ATP, and Mg2+ of 1.9, 2.1, and 4.1 mM, respectively. Several compounds including quinolinic acid, nicotinic acid, nicotinamide, nicotinamide mononucleotide, AMP, ADP, and NADP+ did not affect NAD kinase activity. The enzyme was not affected by changes in the adenylate energy charge. In contrast, both NADH and NADPH were potent negative modulators of the enzyme, since their presence at micromolar concentrations resulted in a pronounced sigmoidal NAD+ saturation curve. In addition, the presence of a range of concentrations of the reduced nucleotides resulted in an increase of the Hill slope (nH) to 1.7 to 2.0 with NADH and to 1.8 to 2.1 with NADPH, suggesting that NAD kinase is an allosteric enzyme. These results indicate that NAD kinase activity is regulated by the availability of ATP, NAD+, and Mg2+ and, more significantly, by changes in the NADP+/NADPH and NAD+/NADH ratios. Thus, NAD kinase probably plays a role in the regulation of NADP turnover and pool size in E. coli.  相似文献   

11.
12.
NADPH-cytochrome P-450 reductase was purified to apparent homogeneity from detergent-solubilized guinea pig liver microsomes. The reductase had a mol. wt of 78,000 and contained one mole each of FAD and FMN. Electron transfer activity to cytochrome c was optimal at a pH of 8.0 and an ionic strength of 0.43. The results of kinetic experiments were consistent with a ternary-complex mechanism for the interaction of the reductase with cytochrome c and NADPH. Km values for NADPH and cytochrome c were 3.1 and 26.7 microM, respectively. Inhibition by NADP+ and 2'-AMP was competitive with respect to NADPH; Ki values were 12.1 microM for NADP+ and 46.7 microM for 2'-AMP.  相似文献   

13.
1. The stoicheiometries and affinities of ligand binding to isocitrate dehydrogenase were studied at pH 7.0, mainly by measuring changes in NADPH and protein fluorescence. 2. The affinity of the enzyme for NADPH is about 100-fold greater than it is for NADP+ in various buffer/salt solutions, and the affinities for both coenzymes are decreased by Mg2+, phosphate and increase in ionic strength. 3. The maximum binding capacity of the dimeric enzyme for NADPH, from coenzyme fluorescence and protein-fluorescence measurements, and also for NADP+, by ultrafiltration, is 2 mol/mol of enzyme. Protein-fluorescence titrations of the enzyme with NADP+ are apparently inconsistent with this conclusion, indicating that the increase in protein fluorescence caused by NADP+ binding is not proportional to fractional saturation of the binding sites. 4. Changes in protein fluorescence caused by changes in ionic strength and by the binding of substrates, Mg2+ or NADP+ (but not NADPH) are relatively slow, suggesting conformation changes. 5. In the presence of Mg2+, the enzyme binds isocitrate very strongly, and 2-oxoglutarate rather weakly. 6. Evidence is presented for the formation of an abortive complex of enzyme-Mg2+-isocitrate-NADPH in which isocitrate and NADPH are bound much more weakly than in their complexes with enzyme and Mg2+ alone. 7. The results are discussed in relation to the interpretation of the kinetic properties of the enzyme and its behaviour in the mitochondrion.  相似文献   

14.
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   

15.
The kinetic mechanism of the reaction catalyzed by glucose-6-phosphate dehydrogenase (EC 1.1.1.49) from Dicentrarchus labrax liver was examined using initial velocity studies, NADPH and glucosamine 6-phosphate inhibition and alternate coenzyme experiments. The results are consistent with a steady-state ordered sequential mechanism in which NADP+ binds first to the enzyme and NADPH is released last. Replots of NADPH inhibition show an uncommon parabolic pattern for this enzyme that has not been previously described. A kinetic model is proposed in agreement with our kinetic results and with previously published structural studies (Bautista et al. (1988) Biochem. Soc. Trans. 16, 903-904). The kinetic mechanism presented provides a possible explanation for the regulation of the enzyme by the [NADPH]/[NADP+] ratio.  相似文献   

16.
6-Phosphogluconate dehydrogenase is the pivotal enzyme that links the gluconate route and the oxidative phase of the pentose phosphate pathway in Schizosaccharomyces pombe. The enzyme differs from the known 6-phosphogluconate dehydrogenases of other sources in that the Schizosaccharomyces enzyme is tetrameric having a subunit mass of 38 kDa, that it requires NADP+ obligatorily for activity, and that it can be activated by divalent metal ions such as Co2+ and Mn2+. Steady-state kinetic studies were undertaken. Initial rate and product inhibition results suggest that 6-phosphogluconate dehydrogenase from Schizosaccharomyces pombe catalyzes NADP(+)-linked oxidative decarboxylation of 6-phosphogluconate by an equilibrium random mechanism with two independent binding sites, namely one site for the nicotinamide coenzyme, NADP+/NADPH, and another site for 6-phosphogluconate-D-ribulose-5-phosphate and for CO2. Studies of pH dependence implicated a basic residue with a pK value of 7.4 in the binding of 6-phosphogluconate and an acidic residue with a pK value of 6.7 in the cation-mediated interaction of NADP+ with the enzyme.  相似文献   

17.
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   

18.
Delta 1-Piperidine 2-carboxylic acid, an alpha-imino acid, is reduced by 1,4-dihydropyridines to pipecolic acid, an alpha-amino acid, and the corresponding pyridinium ions. This nonenzymatic reaction occurs only in the direction of pipecolic acid production. Glutamate dehydrogenase catalyzes this reaction when the reductant is NADPH and gives as products L-pipecolic acid and NADP+. The reaction velocity for the enzyme-catalyzed reaction is measurable in either direction. The pH-independent equilibrium constant, Keq, for the reduction of the imino acid by NADPH to give pipecolic acid anion and NADP+ was determined from the equilibrium conditions and the pKa values of pipecolic acid (10.72) and of the cyclic imino acid (8.10). The value of Keq was found to be 175 +/- 30; the values of delta G0, delta H0 and delta S0 are -3.1 +/- 0.1 kcal/mol, 5 +/- 1 kcal/mol and 27 +/- 4 e.u., respectively. The data indicate that the reactants are far more solvated than the products and that there must be a large degree of solvent reorganization during the course of the reaction. If these thermodynamic parameters apply to the redox step of the enzyme-catalyzed glutamate reaction, then the burst phase which results upon mixing the enzyme, L-glutamate and NADP+ in stoichiometric amounts must contain a hidden nonredox step of large delta H0 value to account for the curved Arrhenius plot observed for this phase (A.H. Colen, R.T. Medary and H.F. Fisher, Biopolymers 20 (1981) 879).  相似文献   

19.
Dismutation of dihydrofolate by dihydrofolate reductase   总被引:1,自引:0,他引:1  
R L Blakley  L Cocco 《Biochemistry》1984,23(11):2377-2383
Degradation of 7,8-dihydrofolate (H2folate) in the presence of dihydrofolate reductase (DHFR) has been shown due not to an oxygenase activity of the reductase as previously reported but to dismutation of H2folate to folate and 5,6,7,8-tetrahydrofolate (H4folate). The reaction can be followed spectrophotometrically or by analysis of the reaction mixture by high-performance liquid chromatography (HPLC). The products have also been isolated and characterized. Oxygen uptake during the reaction is much less than stoichiometric with H2folate disappearance and is attributed to autoxidation of the H4folate formed. The dismutation activity is a property of highly purified Streptococcus faecium DHFR isoenzyme 2 (but not isoenzyme 1) and of Lactobacillus casei DHFR, but not of bovine liver DHFR. The activity is dependent on tightly bound NADP+ and/or NADPH. Removal of the nucleotide results in loss of dismutation activity, which is restored by adding NADP+ or NADPH. Maximum activity is obtained when approximately 1 mol equiv of nucleotide is added per mol of DHFR. It is proposed that in the dismutation reaction bound NADP(H) is alternately reduced and oxidized by incoming molecules of H2folate with release of folate and H4folate, respectively. The relatively slow rate of folate formation presumably limits the rate of the overall reaction. The equilibrium constant for the dismutation reaction is 19.4 +/- 7.4 at 22 degrees C and pH 7.0. Calculation of standard oxidation-reduction potentials at pH 7 gave values of -0.230 V for the H2folate/H4 folate pair and -0.268 V for the folate/H2folate pair. The mechanism by which NADP+ is retained by the enzyme from some sources during purification procedures is unclear.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
One- and two-electron reduction of quinones by glutathione reductase   总被引:1,自引:0,他引:1  
Yeast glutathione reductase (E.C. 1.6.4.2) catalyzes the oxidation of NADPH by p-quinones and ferricyanide with a maximal turnover number (TNmax) of 4-5 s-1.NADP+ stimulates the reaction and the TNmax/Km value of acceptors is reached at NADP+/NADPH greater than or equal to 100. TNmax is increased up to 30-33 s-1. The stimulatory effect of NADP+ may be associated with its complexation with the NADPH-binding site in the reduced enzyme (Kd = 40-60 microM). It is suggested that NADP+ shifts the electron density towards FAD in the two-electron-reduced enzyme and, evidently, changes its one-electron-reduction potentials, while quinones oxidize an equilibrium form of glutathione reductase containing reduced FAD. In the absence of NADP+ the reduction of quinones by glutathione reductase proceeds mainly in a two-electron manner. At NADP+/NADPH = 100 a one-electron reduction makes up 44% of the total process. At pH 6.0-7.0 the reduced forms of naphthoquinones undergo cyclic redox conversions. A hyperbolic dependence exists of the log TN/Km of quinones on their one-electron-reduction potentials.  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