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 oxidation of indole-3-acetic acid (IAA) catalyzed by horseradish peroxidase (HRP) in the absence of added H2O2 was studied at pH 7.4 using spectral and kinetic approaches. Upon addition of a hundred-fold excess of IAA to HRP the native enzyme was rapidly transformed to compound II (HRP-II). HRP-II was the predominant catalytic enzyme species during the steady state. No compound III was observed. HRP-II was slowly transformed to the stable inactive verdohemo-protein, P-670. A precursor of P-670, so-called P-940 was not detected. After the cessation of IAA oxidation there was neither oxygen consumption nor P-670 formation; the remaining HRP-II was spontaneously reduced to native enzyme. Single exponential kinetics were observed in the steady state for IAA oxidation, oxygen consumption and P-670 formation yielding identical first order rate constants of about 6 . 10(4) s(-1). A comparison of the rate of IAA oxidation by HRP-II in the steady state and in the transient state indicated that more than 1 3 of the IAA was oxidized non-enzymatically during the steady state, confirming that a free radical chain reaction is involved in the peroxidase-catalyzed oxidation of IAA. IAA oxidation stopped before IAA was completely consumed, which cannot be ascribed to enzyme inactivation because 30-50% of the enzyme was still active after the end of the reaction. Instead, incomplete IAA oxidation is explained in terms of termination of the free radical chain reaction. Bimolecular rate constants of IAA oxidation by HRP-I and HRP-II determined under transient state conditions were (2.2 +/- 0.1) x 10(3) M(-1) s(-1) and (2.3 +/- 0.2) x 10(2) M(-1) s(-1).  相似文献   

2.
The transient state kinetics of the oxidation of reduced nicotinamide adenine dinucleotide (NADH) by horseradish peroxidase compound I and II (HRP-I and HRP-II) was investigated as a function of pH at 25.0 degrees C in aqueous solutions of ionic strength 0.11 using both a stopped-flow apparatus and a conventional spectrophotometer. In agreement with studies using many other substrates, the pH dependence of the HRP-I-NADH reaction can be explained in terms of a single ionization of pKa = 4.7 +/- 0.5 at the active site of HRP-I. Contrary to studies with other substrates, the pH dependence of the HRP-II-NADH reaction can be interpreted in terms of a single ionization with pKa of 4.2 +/- 1.4 at the active site of HRP-II. An apparent reversibility of the HRP-II-NADH reaction was observed. Over the pH range of 4-10 the rate constant for the reaction of HRP-I with NADH varied from 2.6 X 10(5) to 5.6 X 10(2) M-1 s-1 and of HRP-II with NADH varied from 4.4 X 10(4) to 4.1 M-1 s-1. These rate constants must be taken into consideration to explain quantitatively the oxidase reaction of horseradish peroxidase with NADH.  相似文献   

3.
Reactions between horseradish peroxidase (HRP) compound I and II and some natural phenolic antioxidants were studied at pH 7. The bimolecular rate constants for these reactions were determined using a sequential mixing stopped-flow spectrometer. The rate constants for the reactions of compound I were found to be two orders of magnitude higher than those for compound II. The phenols under study showed a significant difference in their one-electron reduction potential values. As the rate constants also changed systematically with their one-electron potentials, the Marcus theory of electron transfer was applied to the above determined rate constants and the thermodynamic driving force (deltaG(o)), from which the reorganization energy (lambda) for the electron transfer from phenols to both compound I and II was estimated.  相似文献   

4.
Lignin and Mn peroxidase-catalyzed oxidation of phenolic lignin oligomers   总被引:2,自引:0,他引:2  
Banci L  Ciofi-Baffoni S  Tien M 《Biochemistry》1999,38(10):3205-3210
The oxidation of phenolic oligomers by lignin and manganese peroxidases was studied by transient-state kinetic methods. The reactivity of peroxidase intermediates compound I and compound II was studied with the phenol guaiacol along with a beta-O-4 phenolic dimer, trimer, and tetramer. Compound I of both peroxidases is much more reactive than compound II. The rate constants for these substrates with Mn peroxidase compound I range from 1.0 x 10(5) M-1 s-1 for guaiacol to 1.1 x 10(3) M-1 s-1 for the tetramer. Reactivity is much higher with lignin peroxidase compound I with rate constants ranging from 1.2 x 10(6) M-1s-1 for guaiacol to 3.6 x 10(5) M-1 s-1 for the tetramer. Rate constants with compound II are much lower with Mn peroxidase exhibiting very little reactivity. The rate constants dramatically decreased with both peroxidases as the size of the substrate increased. The extent of the decrease was much more dramatic with Mn peroxidase, leading us to conclude that, despite its ability to oxidize phenols, Mn2+ is the only physiologically significant substrate. The rate decrease associated with increasing substrate size was more gradual with lignin peroxidase. These data indicate that whereas Mn peroxidase cannot efficiently directly oxidize the lignin polymer, lignin peroxidase is well suited for direct oxidation of polymeric lignin.  相似文献   

5.
J Sakurada  R Sekiguchi  K Sato  T Hosoya 《Biochemistry》1990,29(17):4093-4098
The second-order rate constant (k4) for the oxidation of a series of aromatic donor molecules (monosubstituted phenols and anilines) by horseradish peroxidase (HRP) compound II was examined with a stopped-flow apparatus. The electronic states of these substrates were calculated by an ab initio molecular orbital method. It was found that in both phenols and anilines log k4 values correlate well with the highest occupied molecular orbital (HOMO) energy level and the lowest unoccupied molecular orbital (LUMO) energy level, but not with the net charge or frontier electron density on atoms of these molecules. The HOMO and LUMO energy levels of phenols and anilines further showed linear relationships with Hammett's sigma values with negative slopes. Similar results were obtained in the oxidation of substrates by HRP compound I, except that the rate of reaction was much higher than in the case of HRP compound II. In addition, the rates of oxidation of phenols by compound I or II were found to be about 1000 times higher than those of anilines with similar HOMO energy levels. On the basis of these results, the mechanism of electron transfer from the substrate to the heme iron of HRP compound II is discussed.  相似文献   

6.
The oxidation of serotonin (5-hydroxytryptamine) by the myeloperoxidase intermediates compounds I and II was investigated by using transient-state spectral and kinetic measurements at 25.0 +/- 0.1 degrees C. Rapid scan spectra demonstrated that both compound I and compound II oxidize serotonin via one-electron processes. Rate constants for these reactions were determined using both sequential-mixing and single-mixing stopped-flow techniques. The second order rate constant obtained for the one-electron reduction of compound I to compound II by serotonin is (1.7 +/- 0.1) x 10(7) M(-1) x s(-1), and that for compound II reduction to native enzyme is (1.4 +/- 0.1) x 10(6) M(-1) x s(-1) at pH 7.0. The maximum pH of the compound I reaction with serotonin occurs in the pH range 7.0-7.5. At neutral pH, the rate constant for myeloperoxidase compound I reacting with serotonin is an order of magnitude larger than for its reaction with chloride, (2.2 +/- 0.2) x 10(6) M(-1) x s(-1). A direct competition of serotonin with chloride for myeloperoxidase compound I oxidation was observed. Our results suggest that serotonin may have a role to protect lipoproteins from oxidation and to prevent enzymes from inactivation caused by the potent oxidants HOCl and active oxygen species.  相似文献   

7.
The catalytic cycle of heme peroxidases involves three processes: the formation of compound I, its conversion to compound II and regeneration of the native enzyme. Each of the processes consists of a reversible binding stage followed by an irreversible transformation stage. Our group has proposed a continuous, sensitive and reliable chronometric method for measuring the steady-state rate of peroxidase activity. Furthermore, we have derived an analytical expression for the steady-state rate and simplified it, taking into consideration the experimental values of the rate constants of some stages previously determined by other authors in stopped-flow assays. We determined the value of the constant for the transformation of a series of phenols and anilines by compound II, and found that it involves a deprotonation step and an electron transfer step. Study of the solvent deuterium isotope effect on the oxidation of phenol revealed the non-rate-limiting character of the deprotonation step in a proton inventory study. Usage of the Marcus equation showed that the electronic transfer step is rate-limiting in both cases, while phenols and anilines were oxidised at different rates for the same potentials. This can be attributed to the shorter electron-tunnelling distance for electron transfer to the iron ion in the phenols than in the anilines.  相似文献   

8.
A series of p- and m-substituted phenols were examined for their effect on the cyclooxygenase activity of prostaglandin H synthase in 0.1 M phosphate buffer at pH 8.0 and 25.0 +/- 0.1 degrees C. A biphasic response was observed. At low concentrations phenols stimulate, but at higher concentrations inhibit, cyclooxygenase activity. Both enhancement and inhibition are increased by phenolic substituents which are electron-donating, quantified by Hammett sigma constants, and hydrophobic, quantified by Hantsch tau constants. The same series of substituted phenols was also reacted with compound II of prostaglandin H synthase at 4.0 +/- 0.5 degrees C. The compound II data fit the Hammett rho sigma equation; no hydrophobicity factors are required. Phenols inhibit cyclooxygenase activity by interfering with the binding of arachidonic acid to compound I and by competing directly with arachidonic acid as reducing substrates for compound I. Phenols stimulate cyclooxygenase activity by acting as reducing substrates for compound II, thereby accelerating the peroxidatic cycle. Phenols also protect the enzyme from self-catalyzed inactivation, most likely by removing the free radical of prostaglandin G2 by reducing it to prostaglandin G2. Kinetic parameters Km and kcat for cyclooxygenase activity were determined in the presence of phenols. Identical values of Km (15.3 +/- 0.5 mM) and kcat (89 +/- 2 s-1) were obtained regardless of which phenol was employed. Therefore these represent the true Km and kcat values for cyclooxygenase activity.  相似文献   

9.
A high-level expression in Escherichia coli of a fully active recombinant form of a catalase-peroxidase (KatG) from the cyanobacterium Synechocystis PCC 6803 is reported. Since both physical and kinetic characterization revealed its identity with the wild-type protein, the large quantities of recombinant KatG allowed the first examination of second-order rate constants for the oxidation of a series of aromatic donor molecules (monosubstituted phenols and anilines) by a bifunctional catalase-peroxidase compound I using the sequential-mixing stopped-flow technique. Because of the overwhelming catalase activity, peroxoacetic acid has been used for compound I formation. A >/=50-fold excess of peroxoacetic acid is required to obtain a spectrum of relatively pure and stable compound I which is characterized by about 40% hypochromicity, a Soret maximum at 406 nm, and isosbestic points between the native enzyme and compound I at 357 and 430 nm. The apparent second-order rate constant for formation of compound I from ferric enzyme and peroxoacetic acid is (8.74 +/- 0.26) x 10(3) M(-)(1) s(-)(1) at pH 7. 0. Reduction of compound I by aromatic donor molecules is dependent upon the substituent effect on the benzene ring. The apparent second-order rate constants varied from (3.6 +/- 0.1) x 10(6) M(-)(1) s(-)(1) for p-hydroxyaniline to (5.0 +/- 0.1) x 10(2) M(-)(1) s(-)(1) for p-hydroxybenzenesulfonic acid. They are shown to correlate with the substituent constants in the Hammett equation, which suggests that in bifunctional catalase-peroxidases the aromatic donor molecule donates an electron to compound I and loses a proton simultaneously. The value of rho, the susceptibility factor in the Hammett equation, is -3.4 +/- 0.4 for the phenols and -5.1 +/- 0.8 for the anilines. The pH dependence of compound I reduction by aniline exhibits a relatively sharp maximum at pH 5. The redox intermediate formed upon reduction of compound I has spectral features which indicate that the single oxidizing equivalent in KatG compound II is contained on an amino acid which is not electronically coupled to the heme.  相似文献   

10.
 Kinetics of the steady-state oxidation of n–alkylferrocenes (alkyl = H, Me, Et, Bu and C5H11) by H2O2 to form the corresponding ferricenium cations catalyzed by horseradish peroxidase has been studied in micellar systems of Triton X-100, CTAB, and SDS, mostly at pH 6.0 and 25  °C. The rate of oxidation of ferrocenes with longer alkyl radicals is too slow to be measured. The reaction obeying the [RFc]:[H2O2] = 2 : 1 stoichiometry is strictly first-order in both HRP and RFc in a wide concentration range. The corresponding observed second-order rate constants k, which refer to the interaction of the peroxidase compound II (HRP-II) with RFc, decrease with the elongation of the alkyl substituent R, and this in turn is accompanied by an increase in the formal redox potentials E°′ in the same medium. Increasing the surfactant concentration lowers the rate constants k, the effect being due to the nonproductive binding of RFc to micelles rather than to enzyme inactivation. The micellar effects are accounted for in terms of the Berezin pseudo-phase model of micellar catalysis applied to the interaction of enzyme with organometallic substrates. The oxidation was found to occur primarily in the aqueous pseudo-phase and the calculated intrinsic second-order rate constants k w are (1.9 ± 0.5)×105, (2.7 ± 0.1)×104, and (5.9 ± 0.6)×103 M–1 s–1 for HFc, EtFc, and n–BuFc, respectively. The data obtained were used for estimating the self-exchange rate constants for the HRP-II/HRP couple in terms of the Marcus formalism. Received: 15 July 1996 / Accepted: 15 November 1996  相似文献   

11.
The rates of reduction of horseradish peroxidase compound II by p-methoxyphenol (4-hydroxyanisole) have been studied from pH 6.0 to 10.5. The kinetics are influenced by an acid group of pKa 8.7 on compound II. The acidic form of compound II is reactive; the basic form is not. Only the electrically neutral, unionized form of p-methoxyphenol is reactive. Fifteen different phenols were reacted with compound II at either pH 7.6 or pH 7.0 (three of them at both pH's). Rate constants varied from zero for p-nitrophenol to 3.2 X 10(7) M-1 for p-aminophenol. The reactive m- and p-substituted phenols yield a rho value of -4.6 +/- 0.5 when plotted according to the Hammett relation. This compares to the rho value of -6.9 obtained for horseradish peroxidase compound I reactions with phenols (1976, D. Job and H. B. Dunford, Eur. J. Biochem. 66, 607). The difference in sensitivity of compounds I and II to electron donating substituents on the phenols can be explained in terms of the relative simplicity of the reactions. Electron donation occurs to the electron-deficient porphyrin pi-cation radical of compound I accompanied by single proton addition to the protein. For compound II the electron is fed to the ferryl group at the center of the porphyrin in a reaction accompanied by two proton additions to the ferryl oxygen atom, one from the protein and the other from the substrate or solvent. This is followed by loss of water from the inner coordination sphere of the ferric ion. The relative reactivities of three o-substituted phenols can be explained in terms of steric hindrance which is minimal for a single o-substituent.  相似文献   

12.
The catalytic cycle of horseradish peroxidase (HRP; donor:hydrogen peroxide oxidoreductase; EC 1.11.1.7) is initiated by a rapid oxidation of it by hydrogen peroxide to give an enzyme intermediate, compound I, which reverts to the resting state via two successive single electron transfer reactions from reducing substrate molecules, the first yielding a second enzyme intermediate, compound II. To investigate the mechanism of action of horseradish peroxidase on catechol substrates we have studied the oxidation of both 4-tert-butylcatechol and dopamine catalysed by this enzyme. The different polarity of the side chains of both o-diphenol substrates could help in the understanding of the nature of the rate-limiting step in the oxidation of these substrates by the enzyme. The procedure used is based on the experimental data to the corresponding steady-state equations and permitted evaluation of the more significant individual rate constants involved in the corresponding reaction mechanism. The values obtained for the rate constants for each of the two substrates allow us to conclude that the reaction of horseradish peroxidase compound II with o-diphenols can be visualised as a two-step mechanism in which the first step corresponds to the formation of an enzyme-substrate complex, and the second to the electron transfer from the substrate to the iron atom. The size and hydrophobicity of the substrates control their access to the hydrophobic binding site of horseradish peroxidase, but electron density in the hydroxyl group of C-4 is the most important feature for the electron transfer step.  相似文献   

13.
The catalytic constant (k(cat)) and the second-order association constant of compound II with reducing substrate (k(5)) of horseradish peroxidase C (HRPC) acting on phenols and anilines have been determined from studies of the steady-state reaction velocities (V(0) vs. [S(0)]). Since k(cat)=k(2)k(6)/k(2)+k(6), and k(2) (the first-order rate constant for heterolytic cleavage of the oxygen-oxygen bond of hydrogen peroxide during compound I formation) is known, it has been possible to calculate the first-order rate constant for the transformation of each phenol or aniline by HRPC compound II (k(6)). The values of k(6) are quantitatively correlated to the sigma values (Hammett equation) and can be rationalized by an aromatic substrate oxidation mechanism in which the substrate donates an electron to the oxyferryl group in HRPC compound II, accompanied by two proton additions to the ferryl oxygen atom, one from the substrate and the other the protein or solvent. k(6) is also quantitatively correlated to the experimentally determined (13)C-NMR chemical shifts (delta(1)) and the calculated ionization potentials, E (HOMO), of the substrates. Similar dependencies were observed for k(cat) and k(5). From the kinetic analysis, the absolute values of the Michaelis constants for hydrogen peroxide and the reducing substrates (K(M)(H(2)O(2)) and K(M)(S)), respectively, were obtained.  相似文献   

14.
Capsaicinoids are the pungent compounds in Capsicum fruits (i.e., "hot" peppers). Peroxidases catalyze capsaicinoid oxidation and may play a central role in their metabolism. However, key kinetic aspects of peroxidase-catalyzed capsaicinoid oxidation remain unresolved. Using transient-state methods, we evaluated horseradish peroxidase compound I and II reduction by two prominent capsaicinoids (25 degrees C, pH 7.0). We determined rate constants approaching 2 x 10(7) and 5 x 10(5)M(-1)s(-1) for compound I and compound II reduction, respectively. We also determined k(app) values for steady-state capsaicinoid oxidation approaching 8 x 10(5)M(-1)s(-1) (25 degrees C, pH 7.0). Accounting for stoichiometry, these are in excellent agreement with constants for compound II reduction, suggesting that this reaction governs capsaicinoid-dependent peroxidase turnover. Ascorbate rapidly reduced capsaicinoid radicals, assisting in the determination of the kinetic constants reported. Because ascorbate accumulates in Capsicum fruits, it may also be an important determinant for capsaicinoid content and preservation in Capsicum fruits and related products.  相似文献   

15.
Numerical and Monte Carlo simulations of horseradish peroxidase-catalyzed phenolic polymerizations have been performed. Kinetic constants for the simulations were fit to data from the oxidation and polymerization of bisphenol A. Simulations of peroxidase-catalyzed phenolic polymerization were run as a function of enzyme concentration and radical transfer and radical coupling rate constants. Predictions were performed with respect to conversion vs. time and number average molecular weight and polydispersity vs. conversion. It is shown that the enzymatic polymerization of phenols can be optimized with respect to high molecular weights by employing low enzyme concentrations and phenols with low radical coupling rate constants coupled with relatively high radical transfer rate constants. Such phenols may be identified by using linear free energy relationships that relate radical reactivity to electron donating/withdrawing potential of the phenolic substituent. (c) 1993 John Wiley & Sons, Inc.  相似文献   

16.
Myeloperoxidase is the most abundant protein in neutrophils and catalyzes the production of hypochlorous acid. This potent oxidant plays a central role in microbial killing and inflammatory tissue damage. 4-Aminobenzoic acid hydrazide (ABAH) is a mechanism-based inhibitor of myeloperoxidase that is oxidized to radical intermediates that cause enzyme inactivation. We have investigated the mechanism by which benzoic acid hydrazides (BAH) are oxidized by myeloperoxidase, and we have determined the features that enable them to inactivate the enzyme. BAHs readily reduced compound I of myeloperoxidase. The rate constants for these reactions ranged from 1 to 3 x 10(6) M-1 s-1 (15 degrees C, pH 7.0) and were relatively insensitive to the substituents on the aromatic ring. Rate constants for reduction of compound II varied between 6.5 x 10(5) M-1 s-1 for ABAH and 1.3 x 10(3) M-1 s-1 for 4-nitrobenzoic acid hydrazide (15 degrees C, pH 7.0). Reduction of both compound I and compound II by BAHs adhered to the Hammett rule, and there were significant correlations with Brown-Okamoto substituent constants. This indicates that the rates of these reactions were simply determined by the ease of oxidation of the substrates and that the incipient free radical carried a positive charge. ABAH was oxidized by myeloperoxidase without added hydrogen peroxide because it underwent auto-oxidation. Although BAHs generally reacted rapidly with compound II, they should be poor peroxidase substrates because the free radicals formed during peroxidation converted myeloperoxidase to compound III. We found that the reduction of ferric myeloperoxidase by BAH radicals was strongly influenced by Hansch's hydrophobicity constants. BAHs containing more hydrophilic substituents were more effective at converting the enzyme to compound III. This implies that BAH radicals must hydrogen bond to residues in the distal heme pocket before they can reduce the ferric enzyme. Inactivation of myeloperoxidase by BAHs was related to how readily they were oxidized, but there was no correlation with their rate constants for reduction of compounds I or II. We propose that BAHs destroy the heme prosthetic groups of the enzyme by reducing a ferrous myeloperoxidase-hydrogen peroxide complex.  相似文献   

17.
The pre-steady-state kinetics of the prostaglandin endoperoxide synthase oxygenase reaction with eicosadienoic acids and the cyclooxygenase reaction with arachidonic acid were investigated by stopped-flow spectrophotometry at 426 nm, an isosbestic point between native enzyme and compound I. A similar reaction mechanism for both types of catalysis is defined from combined kinetic experiments and numerical simulations. In the first step a fatty acid hydroperoxide reacts with the native enzyme to form compound I and the fatty acid hydroxide. In the second step the fatty acid reduces compound I to compound II and a fatty acid carbon radical is formed. This is followed by two fast steps: (1) the addition of either one molecule of oxygen (the oxygenase reaction) or two molecules of oxygen (the cyclooxygenase reaction) to the fatty acid carbon radical to form the corresponding hydroperoxyl radical, and (2) the reaction of the hydroperoxyl radical with compound II to form the fatty acid hydroperoxide and a compound I-protein radical. A unimolecular reaction of the compound I-protein radical to reform the native enzyme is assumed for the last step in the cycle. This is a slow reaction not significantly affecting steps 1 and 2 under pre-steady-state conditions. A linear dependence of the observed pseudo-first-order rate constant, k(obs), on fatty acid concentration is quantitatively reproduced by the model for both the oxygenase and cyclooxygenase reactions. The simulated second order rate constants for the conversion of native enzyme to compound I with arachidonic or eicosadienoic acids hydroperoxides as a substrate are 8 x 10(7) and 4 x 10(7) M(-1) s(-1), respectively. The simulated and experimentally obtained second-order rate constants for the conversion of compound I to compound II with arachidonic and eicosadienoic acids as a substrate are 1.2 x 10(5) and 3.0 x 10(5) M(-1) s(-1), respectively.  相似文献   

18.
Rate constants for the reaction between horseradish peroxidase compound I and p-cresol have been determined at several values of pH between 2.98 and 10.81. These rate constants were used to construct a log (rate) versus pH profile from which it is readily seen that the most reactive form of the enzyme is its most basic form within this pH range so that base catalysis is occurring. At the maximum rate a second order rate constant of (5.1 +/- 0.3) x 10(-7) M-1 s-1 at 25 degrees is obtained. The activation energy of the reaction at the maximum rate was determined from an Arrhenius plot to be 5.0 +/- 0.5 kcal/mol. Evidence for an exception to the generally accepted enzymatic cycle of horseradish peroxidase is presented. One-half molar equivalent of p-cresol can convert compound I quantitatively to compound II at high pH, whereas usually this step requires 1 molar equivalent of reductant. The stoichiometry of this reaction is pH-dependent.  相似文献   

19.
 The second-order rate constants for the oxidation of a series of phenol derivatives by horseradish peroxidase compound II were compared to computer-calculated chemical parameters characteristic for this reaction step. The phenol derivatives studied were phenol, 4-chlorophenol, 3-hydroxyphenol, 3-methylphenol, 4-methylphenol, 4-hydroxybenzoate, 4-methoxyphenol and 4-hydroxybenzaldehyde. Assuming a reaction of the phenolic substrates in their non-dissociated, uncharged forms, clear correlations (r = 0.977 and r = 0.905) were obtained between the natural logarithm of the second-order rate constants (ln k app and ln k 2 respectively) for their oxidation by compound II and their calculated ionisation potential, i.e. minus the energy of their highest occupied molecular orbital [E(HOMO)]. In addition to this first approach in which the quantitative structure-activity relationship (QSAR) was based on a calculated frontier orbital parameter of the substrate, in a second and third approach the relative heat of formation (ΔΔHF) calculated for the process of one-electron abstraction and H abstraction from the phenol derivatives was used as a parameter. Plots of the natural logarithms of the second-order rate constants (k app and k 2) for the reaction and the calculated ΔΔHF values for the process of one-electron abstraction also provide clear QSARs with correlation coefficients of –0.968 and –0.926 respectively. Plots of the natural logarithms of the second-order rate constants (k app and k 2) for the reaction and the calculated ΔΔHF values for the process of H abstraction provide QSARs with correlation coefficients of –0.989 and –0.922 respectively. Since both mechanisms considered, i.e. initial electron abstraction versus initial H abstraction, provided clear QSARs, the results could not be used to discriminate between these two possible mechanisms for phenol oxidation by horseradish peroxidase compound II. The computer calculation-based QSARs thus obtained for the oxidation of the various phenol derivatives by compound II from horseradish peroxidase indicate the validity of the approaches investigated, i.e. both the frontier orbital approach and the approach in which the process is described by calculated relative heats of formation. The results also indicate that outcomes from computer calculations on relatively unrelated phenol derivatives can be reliably compared to one another. Furthermore, as the actual oxidation of peroxidase substrates by compound II is known to be the rate-limiting step in the overall catalysis by horseradish peroxidase, the QSARs of the present study may have implications for the differences in the overall rate of substrate oxidation of the phenol derivatives by horseradish peroxidase. Received: 29 March 1996 / Accepted: 17 July 1996  相似文献   

20.
Reductions of four platinum(IV) amine complexes, cis-diamminetetrachloroplatinum(IV), tetraammine-cis-dichloroplatinum(IV), cis,cis,trans-diamminedichlorodihydroxoplatinum(IV), and cis,trans,cis-dichloro-dihydroxo-bis(isopropylamine)platinum(IV) by ascorbic acid were catalyzed by platinum(II) at pH 7.3 and 22 degrees C. Except for the first mentioned compound, initial slow uncatalyzed reductions yielded platinum(II) products which served as catalyst as revealed by the presence of induction periods and their disappearance by the addition of the platinum(II) products. The platinum(II) catalysis generated ascorbate bound platinum(IV) intermediates. An internal electron transfer process within these intermediates led to the formation of platinum(II) complexes. Although the rate constants for the uncatalyzed reductions vary greatly depending on the nature of the ligands and their spatial arrangements, the magnitudes of the platinum(II) catalyzed rate constants fall in the narrow range, 100 to 300 M(-2) s(-1). The values of the uncatalyzed reductions lie in the range 5 x 10(-2) to 15 M(-1) s(-1), the tetrachloroplatinum(IV) complex suffered the faster reduction. The reduction of iproplatin with two hydroxide ligands in trans configuration was the slowest. The internal electron transfer rate constants span two orders of magnitude, from 0.15 to 4 x 10(-3) s(-1). These reactions were accompanied by the formation of the ascorbate radical which persists throughout the entire reaction. Although the tetrachloro species exhibited simple second order reduction, first order in each of the reactants, the rate of reduction was also accelerated by the addition of cis-diamminedichoroplatinum(II) indicating the presence of catalysis in this reaction as well.  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