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.
The relationship between the NADH:lipoamide reductase and NADH:quinone reductase reactions of pig heart lipoamide dehydrogenase (EC 1.6.4.3) was investigated. At pH 7.0 the catalytic constant of the quinone reductase reaction (kcat.) is 70 s-1 and the rate constant of the active-centre reduction by NADH (kcat./Km) is 9.2 x 10(5) M-1.s-1. These constants are almost an order lower than those for the lipoamide reductase reaction. The maximal quinone reductase activity is observed at pH 6.0-5.5. The use of [4(S)-2H]NADH as substrate decreases kcat./Km for the lipoamide reductase reaction and both kcat. and kcat./Km for the quinone reductase reaction. The kcat./Km values for quinones in this case are decreased 1.85-3.0-fold. NAD+ is a more effective inhibitor in the quinone reductase reaction than in the lipoamide reductase reaction. The pattern of inhibition reflects the shift of the reaction equilibrium. Various forms of the four-electron-reduced enzyme are believed to reduce quinones. Simple and 'hybrid ping-pong' mechanisms of this reaction are discussed. The logarithms of kcat./Km for quinones are hyperbolically dependent on their single-electron reduction potentials (E1(7]. A three-step mechanism for a mixed one-electron and two-electron reduction of quinones by lipoamide dehydrogenase is proposed.  相似文献   

2.
Glutathione reductase from S. cerevisiae (EC 1.6.4.2) catalyzes the NADPH oxidation by glutathione in accordance with a "ping-pong" scheme. The catalytic constant kcat) is 240 s-1 (pH 7.0, 25 degrees C); kcat for the diaphorase reaction is 4-5 s-1. The enzyme activity does not change markedly at pH 5.5-8.0. At pH less than or equal to 7.0, NADP+ acts as a competitive inhibitor towards NADPH and as a noncompetitive inhibitor towards glutathione. NADP+ increases the diaphorase activity of the enzyme. The maximal activity is observed, when the NADP+/NADPH ratio exceeds 100. At pH 8.0, NADP+ acts as a mixed type inhibitor during the reduction of glutathione. High concentrations of NADP+ also inhibit the diaphorase activity due to the reoxidation of the reduced enzyme by NADP+ at pH 8.0. The redox potential of glutathione reductase calculated from the inhibition data is--306 mV (pH 8.0). Glutathione reductase reduces quinoidal compounds in an one-electron way. The hyperbolic dependence of the logarithm of the oxidation constant on the one electron reduction potential of quinone is observed. It is assumed that quinones oxidize the equilibtium fraction of the two-electron reduced enzyme containing reduced FAD.  相似文献   

3.
Lipoamide dehydrogenase, a component of the bovine adrenal ketoglutarate dehydrogenase complex, catalyzes the oxidation of NADH by p-quinones and ferricyanide. The kinetics of oxidation obey the ping-pong mechanism. At pH 7.0, the constants for the active center oxidation by quinones (kox) are equal to 1.1 X 10(4)-5.3 X 10(5) M-1s-1 and increase as the acceptor potential rises. The values of kox for quinones change insignificantly within the pH range of 7.7-5.0, whereas that for ferricyanide increases 10-fold with a decrease of pH from 7.0 to 5.0. The value of the catalytic constant for the enzyme (kcat) reaches its maximum at pH 5.5. The quinones interact with the thiol groups of lipoamide dehydrogenase by inhibiting the fluorescence of FAD and diaphorease activity. The reaction is catalyzed by a basic amino acid (pK 6.7) within the composition of the enzyme.  相似文献   

4.
The pyruvate-dehydrogenase complex from Azotobacter vinelandii.   总被引:1,自引:0,他引:1  
The pyruvate dehydrogenase complex from Axotobacter vinelandii was isolated in a five-step procedure. The minimum molecular weight of the pure complex is 600,000, as based on an FAD content of 1.6 nmol-mg protein-1. The molecular weight is 1.0-1.2 X 10(6), indicating 1 mole of lipoamide dehydrogenase dimer per complex molecule. Sodium dodecylsulphate gel electrophoretical patterns show that apart from pyruvate dehydrogenase (Mr89,000) and lipoamide dehydrogenase (Mrmonomer 56,000) two active transacetylase isoenzymes are present with molecular weight on the gel 82,000 and 59,000 but probably actually lower. The pure complex has a specific activity of the pyruvate-NAD+ reductase (overall) reaction of 10 units-mg protein-1 at 25 degrees C. The partial reactions have the following specific activities in units-mg protein-1 at 25 degrees C under standard conditions: pyruvate-K3Fe(CN)6 reductase 0.14, transacetylase 3.6 and lipoamide dehydrogenase 2.9. The properties of this complex are compared with those from other sources. NADPH reduced the FAD of lipoamide dehydrogenase as well in the complex as in the free form. NADP+ cannot be used as electron acceptor. Under aerobic conditios pyruvate oxidase reaction, dependent on Mg2+ and thiamine pyrophosphate, converts pyruvate into CO2 and acetate; V is 0.2 mumol 02-min-1-mg-1, Km(pyruvate)0.3 mM. The kinetics of this reaction shows a linear 1/velocity-1/[pyruvate] plot. K3Fe(CN)6 competes with the oxidase reaction. The oxidase activity is stimulated by AMP and sulphate and is inhibited by acetyl-CoA. The partially purified enzyme contains considerable phosphotransacetylase activity. The pure complex does not contain this activity. The physiological significance of this activity is discussed.  相似文献   

5.
The reactivity of cuprous stellacyanin as a quinone and semiquinone reductase has been examined. Rate constants (25.0 degrees C) measured for the oxidation of stellacyanin by 1,4-benzoquinone and benzosemiquinone are 2.3 X 10(4) M-1 s-1 (delta H not equal to = 4.4 kcal/mol, delta S not equal to = -24 eu) and 5.1 X 10(6) M-1 s-1, respectively [pH 7.0, I = 0.1 M (phosphate)]. The agreement of these rate constants with those calculated on the basis of relative Marcus theory is discussed. Stellacyanin is more effective than laccase in quenching benzosemiquinone, suggesting that the physiological role of this metalloprotein is to regulate the concentration of free radicals generated through the laccase-catalyzed oxidation of phenols.  相似文献   

6.
In reaction centers from Rhodopseudomonas viridis that contain a single quinone, the decay of the photo-oxidized primary donor, P+, was found to be biphasic when the bound, donor cytochromes were chemically oxidized by ferricyanide. The ratio of the two phases was dependent on pH with an apparent pK of 7.6. A fast phase, which dominated at high pH (t1/2 = 1 ms at pH 9.5), corresponded to the expected charge recombination of P+ and the primary acceptor QA-. A much slower phase dominated at low pH and was shown to arise from a slow reduction of P+ by ferrocyanide in reaction centers where QA- has been rapidly oxidized by ferricyanide. The rate of QA- oxidation was linear with respect to ferricyanide activity and was strongly pH-dependent. The second-order rate constant, corrected for the activity coefficient of ferricyanide, approached a maximum of 2 X 10(8) M-1 X s-1 at low pH, but decreased steadily as the pH was raised above a pK of 5.8, indicating that a protonated state of the reaction center was involved. The slow reduction of P+ by ferrocyanide was also second-order, with a maximum rate constant at low pH of 8 X 10(5) M-1 X s-1 corrected for the activity coefficient of ferrocyanide. This rate also decreased at higher pH, with a pK of 7.4, indicating that ferrocyanide also was most reactive with a protonated form of the reaction center. The oxidation of QA- by ferricyanide was unaffected by the presence of o-phenanthroline, implying that access to QA- was not via the QB-binding site. In reaction centers supplemented with ubiquinone, oxidation of reduced secondary quinone, QB-, by ferricyanide was observed but was substantially slower than that for QA-. It is suggested that Q-B may be oxidized via QA so that the rate is modulated by the equilibrium constant for QA-QB in equilibrium with QAQB-.  相似文献   

7.
Ligand specificity of the type I steroid receptor is apparently conferred by the activity of 11 beta-hydroxysteroid dehydrogenase. To determine the kinetic properties of this enzyme, rat liver cDNA was expressed in cultured cells using recombinant vaccinia virus. Although this enzyme catalyzes only dehydrogenation when purified from rat liver, the recombinant enzyme obtained from cell lysates catalyzed both 11 beta-dehydrogenation of corticosterone to 11-dehydrocorticosterone and the reverse 11-oxoreduction reaction. At pH 8.5, the first order rate constant Kcat/Km for dehydrogenase activity exceeded that for reductase (63 vs. 38 min-1 x 10(-4], whereas the rate constants for the two reactions were nearly equal (48 vs. 47 min-1 x 10(-4] at pH 7.0. These results are consistent with the previously determined pH optima for these activities in liver microsomes. Removal (with glucose-6-phosphate dehydrogenase) of NADP+ produced by the reductase reaction significantly increased reductase activity. Glycyrrhetinic acid, a known inhibitor of the dehydrogenase reaction, also inhibited the reductase reaction at slightly higher concentrations (50% inhibitory concentration, less than 5 nM for dehydrogenase, 10-20 nM for reductase). Partial inhibition of glycosylation with A1-tunicamycin decreased dehydrogenase activity 50% without affecting reductase activity. The data demonstrate that a single polypeptide catalyzes both dehydrogenation and reduction, although the presence of additional enzyme forms catalyzing one or the other activity has not been ruled out.  相似文献   

8.
In addition to steady-state properties of calcium binding to parvalbumins, kinetic studies are required for adequate evaluation of the physiological roles of parvalbumins. By using a dual-wavelength spectrophotometer equipped with a stopped-flow accessory, the transient kinetics of calcium binding to parvalbumins (PA-1 and 2) from bullfrog skeletal muscle was examined at 20 degrees C in medium containing 20 mM MOPS-KOH, pH 6.80, 0.13 mM tetramethylmurexide, 25 microM CaCl2, metal-deprived PA-1 or PA-2, various concentrations of Mg2+, and KCl to adjust the ionic strength of the medium to 0.106. The results can be explained in terms of the following rate constants under the conditions mentioned above when a second-order kinetic scheme is assumed. For PA-1, the association and apparent dissociation rate constants for Ca2+ are 1.5 X 10(7) M-1 X s-1 and 1.5 s-1, respectively, or more. The rate constants for Mg2+ are 7,500 M-1 X s-1 and 5-6 s-1, respectively. For PA-2, the rate constants for Ca2+ are 7 X 10(6) M-1 X s-1 and 1.16 s-1, respectively, and those for Mg2+ are 3,500 M-1 X s-1 and 3.5-4 s-1, respectively. Increased affinities for Ca2+ and Mg2+ at 10 degrees C are largely due to decreased apparent dissociation rate constants for these divalent cations, because no significant change in the association rate constants was found.  相似文献   

9.
Dissociation constants of cytokinins, derivatives of purine which form complexes with cupric ion, were determined by spectrophotometry and the stability constants of their copper complexes by pH titration. The values found for kinetin were 3.76, 9.96, 7.8, and 15.3 for pK1, pK2, logk1, and log beta 2, respectively, and those for 6-benzylaminopurine were, in the same order, 3.90, 9.84, 8.3, and 15.9. The copper(II) complexes with kinetin and 6-benzylaminopurine had superoxide dismutase mimetic activity, and the reaction rate constants with superoxide, which were determined by polarography, were 2.3 X 10(-7) M-1 s-1 for kinetin and 1.5 X 10(-7) M-1 s-1 for 6-benzylaminopurine at pH 9.8 and 25 degrees C.  相似文献   

10.
T C Lehman  C Thorpe 《Biochemistry》1990,29(47):10594-10602
Medium-chain acyl-CoA dehydrogenase reduced with octanoyl-CoA is reoxidized in two one-electron steps by two molecules of the physiological oxidant, electron transferring flavoprotein (ETF). The organometallic oxidant ferricenium hexafluorophosphate (Fc+PF6-) is an excellent alternative oxidant of the dehydrogenase and mimics a number of the features shown by ETF. Reoxidation of octanoyl-CoA-reduced enzyme (200 microM Fc+PF6- in 100 mM Hepes buffer, pH 7.6, 1 degree C) occurs in two one-electron steps with pseudo-first-order rate constants of 40 s-1 and about 200 s-1 for k1 and k2, respectively. The reaction is comparatively insensitive to ionic strength, and evidence of rate saturation is encountered at high ferricenium ion concentration. As observed with ETF, the free two-electron-reduced dehydrogenase is a much poorer kinetic reductant of Fc+PF6-, with rate constants of 3 s-1 and 0.3 s-1 (for k1 and k2, respectively) using 200 microM Fc+PF6-. In addition to the enoyl-CoA product formed during the dehydrogenation of octanoyl-CoA, binding a number of redox-inert acyl-CoA analogues (notably 3-thia- and 3-oxaoctanoyl-CoA) significantly accelerates electron transfer from the dehydrogenase to Fc+PF6-. Those ligands most effective at accelerating electron transfer favor deprotonation of reduced flavin species in the acyl-CoA dehydrogenase. Thus this rate enhancement may reflect the anticipated kinetic superiority of anionic flavin forms as reductants in outer-sphere electron-transfer processes. Evidence consistent with the presence of two distinct loci for redox communication with the bound flavin in the acyl-CoA dehydrogenase is presented.  相似文献   

11.
Lipoamide dehydrogenase (E.C. 1.6.4.3) was found in Trypanosoma cruzi, Tulahuen strain, stocks Tul-2 and Q501, and CA-1 strain. After differential centrifugation of epimastigote homogenates, ammonium sulfate fractionation of the 105,000 g supernatant yielded a partially purified preparation which precipitated between 0.40 and 0.80 ammonium sulfate saturation. The enzyme (a) catalyzed the oxidation of dihydrolipoamide by NAD+ and the reduction of lipoamide by NADH, the forward reaction being 2.5-fold faster than the reverse reaction; (b) exhibited hyperbolic dependence on substrate concentration and (c) possessed diaphorase activity which was less than 5% of the lipoamide reductase activity. The NADH-reduced enzyme was inhibited by arsenite, cadmium and p-chloromercuribenzoate in a concentration-dependent manner. Substrate specificity allowed lipoamide dehydrogenase to be differentiated from T. cruzi trypanothione reductase and other NADPH-dependent flavoenzymes. After cell disruption, lipoamide dehydrogenase was found mostly in the cytosolic fraction and no evidence for association with the plasma membrane was obtained.  相似文献   

12.
NADH oxidation by quinone electron acceptors   总被引:1,自引:0,他引:1  
The rate constants of NADH oxidation by quinones are increased with the oxidation potential increase: log kox (M-1 X s-1) = -0.25 + 12.2 E0(7) (V) for o-quinones and log kox (M-1 X s-1) = -3.06 + 13.5 E0(7) (V) for p-quinones (pH 7.0, 25 degrees C). It is assumed that the oxidation proceeds via the hydride-ion transfer. The rate constants of NADH oxidation by single-electron quinone acceptors are also increased with the oxidizer potential increase; log kox (M-1 X s-1) = -0.64 + 9.34 E0(7) (V) and correlate with the constants of NADH oxidation by quinone radicals obtained earlier (Grodkowski, J., Neta, P., Carlson, B.W. and Miller, L. (1983) J. Phys. Chem. 87, 3135-3138). Single-electron transfer is the limiting stage of the process.  相似文献   

13.
H Ruf 《Biophysical chemistry》1987,26(2-3):313-320
The kinetics of adsorption of the proton carrier o-methyl red to the surface of unilamellar spherical phospholipid vesicles have been investigated by means of the temperature-jump relaxation technique with absorbance detection. Single-exponential relaxation curves were observed with time constants in the range 30-130 microseconds. o-Methyl red binds in both its anionic form A- and protonated form AH. Adsorption-desorption of the two species is coupled by two fast protolytic reactions, occurring in the aqueous bulk phase and in the surface region of the membrane. The rate constants for adsorption and desorption of the two species were obtained from the dependences of the relaxation time on lipid concentration at different pH values. The analysis yielded apparent adsorption rate constants of kasAH = 9.8 X 10(6) M-1 s-1 and kasA = 1.3 X 10(6) M-1 s-1 (expressed in terms of monomeric lipid), and kasAH = 1.2 X 10(11) M-1 s-1 and kasA = 1.6 X 10(10) M-1 s-1 (expressed in terms of vesicle concentration). From the order of these rate constants it is concluded that adsorption of both species is actually diffusion-controlled. The peculiar pH dependence of the relaxation time is a consequence of the protolytic reaction in the surface region of the membrane. Its implication for the kinetics of adsorption-desorption processes are discussed.  相似文献   

14.
The primary kinetic isotope effect of the reaction catalyzed by NAD+-dependent formate dehydrogenase (EC 1.2.1.2.) from the methylotrophic bacterium Pseudomonas sp. 101 has been studied. Analysis of the ratios HVm/DVm and H(Vm/KM)/D(Vm/KM) in the pH range 6.1-7.9 showed that the transfer of hydride ion in ternary enzyme-substrate complex is a limiting step of the reaction, and the formate binding to the binary complex (formate dehydrogenase + NAD+) reached equilibrium when the pH of the medium was increased. An approach has been developed to determine the elementary constants of substrate association (kon) and dissociation (koff) at the stages of the binary--ternary enzyme-substrate complexes for the random equilibrium 2-substrate kinetic mechanism. The kon and koff values obtained for the bacterial formate dehydrogenase by using the proposed approach for NAD+ were (4.8 +/- 0.8)*10(5)M-1s-1 and (90 +/- 10) s-1, and for formate (2.0 +/- 1.0)*10(4) M-1s-1 and (60 +/- 20) s-1, respectively.  相似文献   

15.
Complex formation and the kinetics of electron transfer between ferredoxin-NADP+ reductase (FNR) and two structurally homologous acidic 4Fe-4S high-potential ferredoxins (HiPIP's) from Ectothiorhodospira halophila (HP1 and HP2) and two structurally homologous cytochromes c2 from Paracoccus denitrificans and Rhodospirillum rubrum (PC2, and RC2, respectively) have been investigated by gel filtration and laser flash photolysis techniques. Gel filtration studies indicated that complex formation occurred between FNRox and HP1ox or HP2ox at low ionic strength (10 mM) and that the complexes were completely dissociated at high ionic strength (310 mM). Laser flash photolysis using lumiflavin as the reductant demonstrated that both free HP1ox and HP2ox reacted primarily with the anionic form of fully reduced lumiflavin (LFH-), whereas FNR was unreactive. Second-order rate constants of 1 X 10(6) and 0.8 X 10(6) M-1 s-1 were obtained for these reactions at 10 mM ionic strength. Increasing the ionic strength to 310 mM resulted in an approximately 1.5-fold increase in the rate constant. Inclusion of stoichiometric amounts of FNRox into the reaction mixture at low ionic strength led to a 2.5-fold increase in the rate constants. The reaction of 5-deazariboflavin semiquinone (5-dRf.) with the oxidized HiPIP's was also investigated by laser flash photolysis. Second-order rate constants of 3.0 X 10(8) M-1 s-1 (HP1) and 2.5 X 10(8) M-1 s-1 (HP2) were obtained for the free proteins at 10 mM ionic strength. Under the same conditions, 5-dRf. reacted with free FNRox, resulting in the formation of the neutral protein-bound semiquinone (FNR.), with a second-order rate constant of 6 X 10(8) M-1 s-1.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

16.
The reactions of NADPH oxidation by quinones and inorganic complexes catalyzed by NADPH: adrenodoxin reductase were studied. The catalytic constant for the enzyme at pH 7.0 is 20-25 s-1; the oxidative constants for the quinones vary from 5 X 10(5) to 1.1 X 10(3) M-1 s-1 and show an increase with a rise in the one-electron acceptor reduction potential. The mode of adrenodoxin reductase interaction with oxyquinones differs from that of the enzyme interaction with alkyl-substituted quinones and inorganic complexes. NADPH competitively inhibits electron acceptors, whereas NADP+ is a competitive inhibitor of NADPH and a uncompetitive inhibitor of electron acceptors. (Ki = 25 microM). The depth of FAD incorporation into the enzyme molecule as calculated according to the outer sphere electron transfer theory is 6.1 A.  相似文献   

17.
The lipoamide dehydrogenase of the glycine decarboxylase complex was purified to homogeneity (8 U/mg) from cells of the anaerobe Eubacterium acidaminophilum that were grown on glycine. In cell extracts four radioactive protein fractions labeled with D-[2-14C]riboflavin could be detected after gel filtration, one of which coeluted with lipoamide dehydrogenase activity. The molecular mass of the native enzyme could be determined by several methods to be 68 kilodaltons, and an enzyme with a molecular mass of 34.5 kilodaltons was obtained by sodium dodecyl sulfate-polyacrylamide gel electrophoresis. Immunoblot analysis of cell extracts separated by sodium dodecyl sulfate-polyacrylamide or linear polyacrylamide gel electrophoresis resulted in a single fluorescent band. NADPH instead of NADH was the preferred electron donor of this lipoamide dehydrogenase. This was also indicated by Michaelis constants of 0.085 mM for NADPH and 1.1 mM for NADH at constant lipoamide and enzyme concentrations. The enzyme exhibited no thioredoxin reductase, glutathione reductase, or mercuric reductase activity. Immunological cross-reactions were obtained with cell extracts of Clostridium cylindrosporum, Clostridium sporogenes, Clostridium sticklandii, and bacterium W6, but not with extracts of other glycine- or purine-utilizing anaerobic or aerobic bacteria, for which the lipoamide dehydrogenase has already been characterized.  相似文献   

18.
Energetics of the calcium-transporting ATPase   总被引:11,自引:0,他引:11  
A thermodynamic cycle for catalysis of calcium transport by the sarcoplasmic reticulum ATPase is described, based on equilibrium constants for the microscopic steps of the reaction shown in Equation 1 under a single set of experimental (formula; see text) conditions (pH 7.0, 25 degrees C, 100 mM KCl, 5 mM MgSO4): KCa = 5.9 X 10(-12) M2, K alpha ATP = 15 microM, Kint = 0.47, K alpha ADP = 0.73 mM, K'int = 1.7, K"Ca = 2.2 X 10(-6) M2, and Kp = 37 mM. The value of K"Ca was calculated by difference, from the free energy of hydrolysis of ATP. The spontaneous formation of an acylphosphate from Pi and E is made possible by the expression of 12.5 kcal mol-1 of noncovalent binding energy in E-P. Only 1.9 kcal mol-1 of binding energy is expressed in E X Pi. There is a mutual destabilization of bound phosphate and calcium in E-P X Ca2, with delta GD = 7.6 kcal mol-1, that permits transfer of phosphate to ADP and transfer of calcium to a concentrated calcium pool inside the vesicle. It is suggested that the ordered kinetic mechanism for the dissociation of E-P X Ca2, with phosphate transfer to ADP before calcium dissociation outside and phosphate transfer to water after calcium dissociation inside, preserves the Gibbs energies of these ligands and makes a major contribution to the coupling in the transport process. A lag (approximately 5 ms) before the appearance of E-P after mixing E and Pi at pH 6 is diminished by ATP and by increased [Pi]. This suggests that ATP accelerates the binding of Pi. The weak inhibition by ATP of E-P formation at equilibrium also suggests that ATP and phosphate can bind simultaneously to the enzyme at pH 6. Rate constants are greater than or equal to 115 s-1 for all the steps in the reaction sequence to form E-32P X Ca2 from E-P, Ca2+ and [32P]ATP at pH 7. E-P X Ca2 decomposes with kappa = 17 s-1, which shows that it is a kinetically competent intermediate. The value of kappa decreases to 4 s-1 if the intermediate is formed in the presence of 2 mM Ca2+. This decrease and inhibition of turnover by greater than 0.1 mM Ca2+ may result from slow decomposition of E-P X Ca3.  相似文献   

19.
D Y Cai  M Tien 《Biochemistry》1990,29(8):2085-2091
The oxycomplexes (compound III, oxyperoxidase) of two lignin peroxidase isozymes, H1 (pI = 4.7) and H8 (pI = 3.5), were characterized in the present study. After generation of the ferroperoxidase by photochemical reduction with deazoflavin in the presence of EDTA, the oxycomplex is formed by mixing ferroperoxidase with O2. The oxycomplex of isozyme H8 is very stable, with an autoxidation rate at 25 degrees C too slow to measure at pH 3.5 or 7.0. In contrast, the oxycomplex of isozyme H1 has a half-life of 52 min at pH 4.5 and 29 min at pH 7.5 at 25 degrees C. The decay of isozyme H1 oxycomplex follows a single exponential. The half-lives of lignin peroxidase oxycomplexes are much longer than those observed with other peroxidases. The binding of O2 to ferroperoxidase to form the oxycomplex was studied by stopped-flow methods. At 20 degrees C, the second-order rate constants for O2 binding are 2.3 X 10(5) and 8.9 X 10(5) M-1 s-1 for isozyme H1 and 6.2 X 10(4) and 3.5 X 10(5) M-1 s-1 for isozyme H8 at pH 3.6 and pH 6.8, respectively. The dissociation rate constants for the oxycomplex of isozyme H1 (3.8 Z 10(-3) s-1) and isozyme H8 (1.0 X 10(-3) s-1) were measured at pH 3.6 by CO trapping. Thus, the equilibrium constants (K, calculated from kon/koff) for both isozymes H1 (7.0 X 10(7) M-1) and H8 (6.2 X 10(7) M-1) are higher than that of myoglobin (1.9 Z 10(6) M-1).(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

20.
J Fisher  B R Abdella  K E McLane 《Biochemistry》1985,24(14):3562-3571
Spinach NADPH:ferredoxin oxidoreductase (EC 1.6.7.1) catalyzes the NADPH-dependent reduction of the anthracyclines daunomycin, aclacinomycin A, and nogalamycin and their respective 7-deoxyanthracyclinones. Under anaerobic conditions, the endogenous rate of O2 reduction by NADPH catalyzed by ferredoxin reductase (0.12 s-1 at pH 7.4) is augmented by the anthracyclines and 7-deoxyanthracyclinones. The catalytic constants are approximately equivalent for this augmentation for all substrates (approximate V of 2 s-1 and KM of 75 microM). Both O2- and H2O2 are made. Under anaerobic conditions, anthracycline reduction catalyzed by ferredoxin reductase results in the elimination of the C-7 substituent to provide a quinone methide intermediate. Following tautomerization by C-7 protonation, 7-deoxyanthracyclinones are obtained. Under appropriate conditions these may be further reduced to the 7-deoxyanthracyclinone hydroquinones. For daunomycin, the quinone methide is formed rapidly after reduction and is easily monitored at 600 nm. It may react with electrophiles other than H+, as demonstrated by its competitive trapping by p-carboxybenzaldehyde. It may also react with nucleophiles, as demonstrated by its competitive trapping by N-acetylcysteine. For aclacinomycin, quinone methide formation is also rapid although no distinct transient near 600 nm occurs. In addition to protonation, it reacts with itself providing the 7,7'-dimer. With ethyl xanthate as a thiolate nucleophile, adducts derived from addition to C-7 are obtained. For nogalamycin, quinone methide formation is not rapid. Nogalamycin is reduced to its hydroquinone, which slowly converts in a first-order process [k = (1.2 +/- 0.2) X 10(-3) s-1, pH 8.0, 30 degrees C] to the quinone methide, which is then quenched by protonation. Spinach ferredoxin in its reduced form is chemically competent for anthracycline reduction. Its effect on both the aerobic and anaerobic reactions catalyzed by ferredoxin reductase is to increase severalfold the overall velocity for anthracycline reduction. In conclusion, the aerobic reaction pathways for the anthracyclines as mediated by ferredoxin reductase are remarkably similar, while the anaerobic reactions are remarkably different. If these anthracyclines exert their antitumor activity by a common anaerobic pathway, it is most likely that the pathway is determined by the properties of the anthracycline as complexed to its in vivo target. The behavior of ferredoxin further suggests that not only low-potential flavin centers but also iron-sulfur centers should be regarded as important loci for anthracycline reductive activation.  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