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.
Catalytic oxidation of p-cresol by ascorbate peroxidase   总被引:2,自引:0,他引:2  
Transient and steady state kinetics, together with a range of chromatographic and spectroscopic techniques, have been used to establish the mechanism and the products of the H(2)O(2)-dependent oxidation of p-cresol by ascorbate peroxidase (APX). HPLC, GC-MS, and NMR analyses are consistent with the formation of 2, 2'-dihydroxy-5,5'-dimethylbiphenyl (II) and 4alpha,9beta-dihydro-8, 9beta-dimethyl-3(4H)-dibenzofuranone (Pummerer's ketone, III) as the major products of the reaction. In the presence of cumene hydroperoxide, two additional products were observed which, from GC and MS analyses, were shown to be 1,1-dimethylbenzylalcohol (IV) and bis-(1-methyl-1-phenyl-ethyl)-peroxide (V). The product ratio II:III was dependent on enzyme concentration: at low concentrations Pummerer's ketone (III) predominates and at high concentrations formation of the biphenyl compound (II) is favored. Steady-state data showed a sigmoidal dependence on [p-cresol] that was consistent with the presence of 2.01 +/- 0.15 binding sites for the substrate (25.0 degrees C, sodium phosphate, pH 7.0, mu = 2.2 mM) and independent of ionic strength in the range 2.2-500 mM. Single turnover kinetic experiments (pH 7.0, 5.0 degrees C, mu = 0.10 M) yielded second-order rate constants for Compound I reduction by p-cresol, k(2), of 5.42 +/- 0.10 x 10(5) M(-1) s(-1), respectively. Rate-limiting reduction of Compound II by p-cresol, k(3), showed saturation kinetics, giving values for K(d) = 1.54 +/- 0.12 x 10(-3) M and k(3) = 18.5 +/- 0.7 s(-1). The results are discussed in the more general context of APX-catalyzed aromatic oxidations.  相似文献   

2.
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   

3.
Guaiacol peroxidase from spinach catalyzes the oxidation of p-aminophenol to produce the aminophenoxy radical as the primary product which is converted further into a stable oxidation product with an absorption peak at 470 nm. The p-aminophenol radicals oxidize ascorbate (AsA) to produce monodehydroascorbate radicals. Kinetic analysis indicates that p-aminophenol radicals also oxidize monodehydroascorbate to dehydroascorbate. Incubation of AsA peroxidase from tea leaves and hydrogen peroxide with p-aminophenol, p-cresol, hydroxyurea, or hydroxylamine results in the inactivation of the enzyme. No inactivation of the enzyme was found upon incubation of the enzyme with these compounds either in the absence of hydrogen peroxide or with the stable oxidized products of these compounds. The enzyme was protected from inactivation by the inclusion of AsA in the incubation mixture. The radicals of p-aminophenol and hydroxyurea were produced by AsA peroxidase as detected by their ESR signals. These signals disappeared upon the addition of AsA, and the signal characteristic of monodehydroascorbate was found. Thus, AsA peroxidase is inactivated by the radicals of p-aminophenol, p-cresol, hydroxyurea, and hydroxylamine which are produced by the peroxidase reaction, and it is protected from inactivation by AsA via the scavenging of the radicals. Thus, these compounds are the suicide inhibitors for AsA peroxidase. Isozyme II of AsA peroxidase, which is localized in chloroplasts, is more sensitive to these compounds than isozyme I. In contrast to AsA peroxidase, guaiacol peroxidase was not affected by these various compounds, even though each was oxidized by it and the corresponding radicals were produced.  相似文献   

4.
The oxidation of yeast cytochrome c peroxidase by hydrogen peroxide produces a unique enzyme intermediate, cytochrome c peroxidase Compound I, in which the ferric heme iron has been oxidized to an oxyferryl state, Fe(IV), and an amino acid residue has been oxidized to a radical state. The reduction of cytochrome c peroxidase Compound I by horse heart ferrocytochrome c is biphasic in the presence of excess ferrocytochrome c as cytochrome c peroxidase Compound I is reduced to the native enzyme via a second enzyme intermediate, cytochrome c peroxidase Compound II. In the first phase of the reaction, the oxyferryl heme iron in Compound I is reduced to the ferric state producing Compound II which retains the amino acid free radical. The pseudo-first order rate constant for reduction of Compound I to Compound II increases with increasing cytochrome c concentration in a hyperbolic fashion. The limiting value at infinite cytochrome c concentration, which is attributed to the intracomplex electron transfer rate from ferrocytochrome c to the heme site in Compound I, is 450 +/- 20 s-1 at pH 7.5 and 25 degrees C. Ferricytochrome c inhibits the reaction in a competitive manner. The reduction of the free radical in Compound II is complex. At low cytochrome c peroxidase concentrations, the reduction rate is 5 +/- 3 s-1, independent of the ferrocytochrome c concentration. At higher peroxidase concentrations, a term proportional to the square of the Compound II concentration is involved in the reduction of the free radical. Reduction of Compound II is not inhibited by ferricytochrome c. The rates and equilibrium constant for the interconversion of the free radical and oxyferryl forms of Compound II have also been determined.  相似文献   

5.
Stopped-flow rapid scan techniques were used to obtain a spectrum of nearly homogeneous lignin peroxidase compound I (LiPI) under pseudo-first order conditions at the unusually low pH optimum (3.0) for the enzyme. The LiPI spectrum had a Soret band at 407 nm with approximately 60% reduced intensity and a visible maximum at 650 nm. Under steady-state conditions a Soret spectrum for lignin peroxidase compound II (LiPII) was also obtained. The Soret maximum of LiPII at 420 nm was only approximately 15% reduced in intensity compared to native LiP. Transient state kinetic results confirmed the pH independence of LiPI formation over the pH range 3.06-7.39. The rate constant was (6.5 +/- 0.2) x 10(5) M-1 S-1. Addition of excess veratryl alcohol to LiPI resulted in its reduction to LiPII with subsequent reduction of LiPII to the native enzyme. Reactions of LiPI and LiPII with veratryl alcohol exhibited marked pH dependencies. For the LiPI reaction the rate constants ranged from 2.5 x 10(6) M-1 S-1 at pH 3.06 to 4.1 x 10(3) M-1 S-1 at pH 7.39; for the LiPII reaction, 1.6 x 10(5) M-1 S-1 (pH 3.06) to 2.3 x 10(3) M-1 S-1 (pH 5.16). These single turnover experiments demonstrate directly that the pH dependence of these reactions dictates the overall pH dependence of this novel enzyme. These results are consistent with the one-electron oxidation of veratryl alcohol to an aryl cation radical by LiPI and by LiPII.  相似文献   

6.
Eosinophil peroxidase, the major granule protein in eosinophils, is the least studied human peroxidase. Here, we have performed spectral and kinetic measurements to study the nature of eosinophil peroxidase intermediates, compounds I and II, and their reduction by the endogenous one-electron donors ascorbate and tyrosine using the sequential-mixing stopped-flow technique. We demonstrate that the peroxidase cycle of eosinophil peroxidase involves a ferryl/porphyrin radical compound I and a ferryl compound II. In the absence of electron donors, compound I is shown to be transformed to a species with a compound II-like spectrum. In the presence of ascorbate or tyrosine compound I is reduced to compound II with a second-order rate constant of (1.0+/-0.2)x10(6) M(-1) s(-1) and (3.5+/-0.2)x10(5) M(-1) s(-1), respectively (pH 7.0, 15 degrees C). Compound II is then reduced by ascorbate and tyrosine to native enzyme with a second-order rate constant of (6.7+/-0.06)x10(3) M(-1) s(-1) and (2.7+/-0.06)x10(4) M(-1) s(-1), respectively. This study revealed that eosinophil peroxidase compounds I and II are able to react with tyrosine and ascorbate via one-electron oxidations and therefore generate monodehydroascorbate and tyrosyl radicals. The relatively fast rates of the compound I reduction demonstrate that these reactions may take place in vivo and are physiologically relevant.  相似文献   

7.
Summary p-Cresol was oxidized by hydrogen peroxide in a reaction catalysed by horseradish peroxidase and the low molecular weight products were investigated. In aqueous media Pummerer's ketone (I) was the dominating product but in organic media the product distribution was quite different; 2,2'-dihydroxy-5,5'-dimethyldiphenyl (II) was the main low molecular weight product. Similar product distributions were obtained with peroxidase adsorbed on a solid support and suspended in toluene and with peroxidase solubilized in a microemulsion containing the same solvent. The best selectivity for the formation of (II) was obtained when the enzyme was adsorbed on Celite and suspended in water-saturated chloroform with 0.5% (v/v) extra water added. The yield of low molecular weight products in this case was 28%; of this fraction, 95% was (II). Offprint requests to: P. Adlercreutz  相似文献   

8.
Elementary reactions have been studied quantitatively in the complex overall process catalyzed by horseradish peroxidase whereby isobutyraldehyde and molecular oxygen react to form triplet state acetone and formic acid. The rate constant for the reaction of the enol form of isobutyraldehyde with compound I of peroxidase is (8 +/- 1) X 10(6) M-1 s-1 and with compound II (1.3 +/- 0.3) X 10(6) M-1 s-1. Neither the enolate anion nor the keto form is reactive. The reactivity of enols with peroxidase parallels that of unionized phenols and a common mechanism is proposed. The overall catalyzed reaction of isobutyraldehyde and oxygen consists of an initial burst followed by a steady state phase. The burst is caused by the following sequence: 1) an initial high yield of compound I is formed from reaction of native enzyme with the autoxidation product of isobutyraldehyde, a peracid and 2) compound I rapidly depletes the equilibrium pool of enol which is present. After this burst a steady state phase is observed in which the rate-limiting step is the conversion of the keto to the enol form of the aldehyde catalyzed by phosphate buffer. The rate constant for the keto form reacting with phosphate is (8.7 +/- 0.6) X 10(-5) M-1 s-1. All constants were measured in dilute aqueous ethanol at 35 degrees C, pH 7.4, and ionic strength 0.67 M. Both the initial burst of light and the steady state emission from triplet acetone can be observed with the naked eye. Since the magnitude of the burst is a measure of the equilibrium amount of enol, the keto-enol equilibrium constant is readily calculated and hence also the rate constant for conversion of enol to keto. The keto-enol equilibrium constant is unaffected by phosphate which therefore acts as a true catalyst.  相似文献   

9.
Using both rapid-scan and conventional spectrophotometry, oxygenation of p-substituted thioanisoles by horseradish peroxidase compounds I and II was investigated at pH 5, 7 and 9. The pH-jump technique was applied to the compound II reactions at acidic and neutral pH. The rate of oxidation of the sulfides is dependent on pH, concentration of substrate and on the different substituents in the para position of the benzene ring. Our results, based on transient state observations of the enzyme intermediates, are in agreement with the results of Kobayashi, S., Minoru, N., Kimura, T. and Schaap, A.P. (Biochemistry (1987) 26, 5019-5022), obtained using 18O-labelling and studies of product formation, in which formation of a sulfur cation radical from compound I is proposed. We consider two reaction mechanisms for the compound II reaction: one a one-electron oxidation of the thioanisole, analogous to the compound I reaction, and the other, the attack of the hydroxyl radical originating from compound II on the sulfur-cation radical.  相似文献   

10.
Titration of guaiacol by hydrogen peroxide in the presence of a catalytic amount of horseradish peroxidase shows that the reduction of hydrogen peroxide proceeds by the abstraction of two electrons from a guaiacol molecule. In the same way, it can be demonstrated that 0.5 mol of guaiacol can reduce, at low temperature, 1 mol of peroxidase compound I to compound II. Moreover, the reaction between equal amounts of compound I and guaiacol at low temperature produces the native enzyme. A reaction scheme is proposed which postulates that two electrons are transferred from guaiacol to compound I giving ferriperoxidase and oxidized guaiacol with the intermediary formation of compound II. The direct two-electron transfer from guaiacol to compound I without a dismutation of product free radicals must be considered as an exception to the general mechanism involving a single-electron transfer.  相似文献   

11.
Proteus mirabilis catalase (PMC) belongs to the family of NADPH binding catalases. The function of NADPH in these enzymes is still a matter of debate. This study presents the effects of two independent phenylalanine mutations (F194 and F215), located between NADPH and heme in the PMC structure. The phenylalanines were replaced with tyrosines which we predicted could carry radicals in a NADPH-heme electron transfer. The X-ray crystal structures of the two mutants indicated that neither the binding site of NADPH nor the immediate environment of the residues was affected by the mutations. Measurements using H2O2 as a substrate confirmed that the variants were as active as the native enzyme. With equivalent amounts of peroxoacetic acid, wild-type PMC, F215Y PMC, and beef liver catalase (BLC) formed a stable compound I, while the F194Y PMC variant produced a compound I which was rapidly transformed into compound II and a tyrosyl radical. EPR studies showed that this radical, generated by the oxidation of Y194, was not related to the previously observed radical in BLC, located on Y369. In the presence of excess NADPH, compound I was reduced to a resting enzyme (k(obs) = 1.7 min(-1)) in a two-electron process. This was independent of the enzyme's origin and did not require any thus far identified tyrosyl radicals. Conversely, the presence of a tyrosyl radical in F194Y PMC greatly enhanced the oxidation of reduced beta-nicotinamide mononucleotide under a steady-state H2O2 flow with observable compound II. This process could involve a one-electron reduction of compound I via Y194.  相似文献   

12.
The reactions of the NAD radical (NAD.) with ferric horseradish peroxidase and with compounds I and II were investigated by pulse radiolysis. NAD. reacted with the ferric enzyme and with compound I to form the ferrous enzyme and compound II with second-order rate constants of 8 X 10(8) and 1.5 X 10(8) M-1 s-1, respectively, at pH 7.0. In contrast, no reaction of NAD. with native compound II at pH 10.0 nor with diacetyldeutero-compound II at pH 5.0-8.0 could be detected. Other reducing species generated by pulse radiolysis, such as hydrated electron (eaq-), superoxide anion (O2-), and benzoate anion radical, could not reduce compound II of the enzyme to the ferric state, although the methylviologen radical reduced it. The results are discussed in relation to the mechanism of catalysis of the one-electron oxidation of substrates by peroxidase.  相似文献   

13.
The kinetics of p-aminobenzoic acid oxidation catalyzed by horseradish peroxidase Compounds I and II was investigated intensively as a function of pH at 25 degrees in aqueous solutions of ionic strength 0.11. All of the rate data were collected from single turnover experiments involving reactions of a single enzyme compound. In reactions of both compounds, deviations from first order behavior with respect to the enzyme were observed at high pH values which were explained in terms of a free radical interaction of product with the enzyme. The effect could be eliminated with sufficient excess of substrate. Kinetic behavior which deviated from first order in substrate, observed at low pH, was explained by a mechanism involving an enzyme-substrate complex which reacted with an additional molecule of substrate but at a slower rate. The pH dependence of the second order rate constants for the reaction of p-aminobenzoic acid with free Compounds I and II is similar to results obtained for the comparable reactions of ferrocyanide, suggesting similar proton-transfer mechanisms for both reducing substrates. The reduction of Compound II by p-aminobenzoic acid appeared to be influenced by two ionizable groups on the enzyme which affect the electronic environment of the heme. The lack of influence of substrate ionizable groups on the rate of the Compound II reaction indicated that potential differences in reactivities of NH2C6H4COO- and NH2C6H4COOH were levelled by the diffusion-controlled limit in the acid region of pH. The reduction of Compound I by p-aminobenzoic acid was not diffusion-controlled and the rate-pH profile could be explained in terms of three acid ionizations, two on the substrate and one on Compound I.  相似文献   

14.
Stopped-flow techniques were used to investigate the kinetics of the formation of manganese peroxidase compound I (MnPI) and of the reactions of MnPI and manganese peroxidase compound II (MnPII) with p-cresol and MnII. All of the rate data were obtained from single turnover experiments under pseudo-first order conditions. In the presence of H2O2 the formation of MnPI is independent of pH over the range 3.12-8.29 with a second-order rate constant of (2.0 +/- 0.1) x 10(6) M-1 s-1. The activation energy for MnPI formation is 20 kJ mol-1. MnPI formation also occurs with organic peroxides such as peracetic acid, m-chloroperoxybenzoic acid, and p-nitroperoxybenzoic acid with second-order rate constants of 9.7 x 10(5), 9.5 x 10(4), and 5.9 x 10(4) M-1 s-1, respectively. The reactions of MnPI and MnPII with p-cresol strictly obeyed second-order kinetics. The second-order rate constant for the reaction of MnPII with p-cresol is extremely low, (9.5 +/- 0.5) M-1 s-1. Kinetic analysis of the reaction of MnII with MnPI and MnPII showed a binding interaction with the oxidized enzymes which led to saturation kinetics. The first-order dissociation rate constants for the reaction of MnII with MnPI and MnPII are (0.7 +/- 0.1) and (0.14 +/- 0.01) s-1, respectively, when the reaction is conducted in lactate buffer. Rate constants are considerably lower when the reactions are conducted in succinate buffer. Single turnover experiments confirmed that MnII serves as an obligatory substrate for MnPII and that both oxidized forms of the enzyme form productive complexes with MnII. Finally, these results suggest the alpha-hydroxy acids such as lactate facilitate the dissociation of MnIII from the enzyme.  相似文献   

15.
The reaction of Euphorbia characias latex peroxidase (ELP) with hydrogen peroxide as the sole substrate was studied by conventional and stopped-flow spectrophotometry. The reaction mechanism occurs via three distinct pathways. In the first (pathway I), ELP shows catalase-like activity: H2O2 oxidizes the native enzyme to compound I and subsequently acts as a reducing substrate, again converting compound I to the resting ferric enzyme. In the presence of an excess of hydrogen peroxide, compound I is still formed and further reacts in two other pathways. In pathway II, compound I initiates a series of cyclic reactions leading to the formation of compound II and compound III, and then returns to the native resting state. In pathway III, the enzyme is inactivated and compound I is converted into a bleached inactive species; this reaction proceeds faster in samples illuminated with bright white light, demonstrating that at least one of the intermediates is photosensitive. Calcium ions decrease the rate of pathway I and accelerate the rate of pathways II and III. Moreover, in the presence of calcium the inactive stable verdohemochrome P670 species accumulates. Thus, Ca2+ ions seem to be the key for all catalytic pathways of Euphorbia peroxidase.  相似文献   

16.
A Hillar  P Nicholls 《FEBS letters》1992,314(2):179-182
Catalase-bound NADPH both prevents and reverses the accumulation of inactive bovine liver catalase peroxide compound II generated by 'endogenous' donors under conditions of steady H2O2 formation without reacting rapidly with either compound I or compound II. It thus differs both from classical 2-electron donors of the ethanol type, and from 1-electron donors of the ferrocyanide/phenol type. NADPH also inhibits compound II formation induced by the exogenous one-electron donor ferrocyanide. A catalase reaction scheme is proposed in which the initial formation of compound II from compound I involves production of a neighbouring radical species. NADPH blocks the final formation of stable compound II by reacting as a 2-electron donor to compound II and to this free radical. The proposed behaviour resembles that of labile free radicals formed in cytochrome c peroxidase and myoglobin. Such radical migration patterns within haem enzymes are increasingly common motifs.  相似文献   

17.
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   

18.
The reactivity of recombinant pea cytosolic ascorbate peroxidase (rAPX) towards H2O2, the nature of the intermediates and the products of the reaction have been examined using UV/visible and EPR spectroscopies together with HPLC. Compound I of rAPX, generated by reaction of rAPX with 1 molar equivalent of H2O2, contains a porphyrin pi-cation radical. This species is unstable and, in the absence of reducing substrate, decays within 60 s to a second species, compound I*, that has a UV/visible spectrum [lambda(max) (nm) = 414, 527, 558 and 350 (sh)] similar, but not identical, to those of both horseradish peroxidase compound II and cytochrome c peroxidase compound I. Small but systematic differences were observed in the UV/visible spectra of compound I* and authentic rAPX compound II, generated by reaction of rAPX with 1 molar equivalent H2O2 in the presence of 1 molar equivalent of ascorbate [lambda(max) (nm) = 416, 527, 554, 350 (sh) and 628 (sh)]. Compound I* decays to give a 'ferric-like' species (lambda(max) = 406 nm) that is not spectroscopically identical to ferric rAPX (lambda(max) = 403 nm) with a first order rate constant, k(decay)' = (2.7 +/- 0.3) x 10(-4) s(-1). Authentic samples of compound II evolve to ferric rAPX [k(decay) = (1.1 +/- 0.2) x 10(-3) s(-1)]. Low temperature (10 K) EPR spectra are consistent with the formation of a protein-based radical, with g values for compound I* (g parallel = 2.038, g perpendicular = 2.008) close to those previously reported for the Trp191 radical in cytochrome c peroxidase (g parallel = 2.037, g perpendicular = 2.005). The EPR spectrum of rAPX compound II was essentially silent in the g = 2 region. Tryptic digestion of the 'ferric-like' rAPX followed by RP-HPLC revealed a fragment with a new absorption peak near 330 nm, consistent with the formation of a hydroxylated tryptophan residue. The results show, for the first time, that rAPX can, under certain conditions, form a protein-based radical analogous to that found in cytochrome c peroxidase. The implications of these data are discussed in the wider context of both APX catalysis and radical formation and stability in haem peroxidases.  相似文献   

19.
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   

20.
The spectral behavior of the enzyme prostaglandin H synthase was studied in the Soret region under conditions that permitted comparison of enzyme intermediates involved in peroxidase and cyclooxygenase activities. First, the peroxidase activity was examined. The enzyme's spectral behavior upon reacting with 5-phenyl-pent-4-enyl-1-hydroperoxide was different depending on the presence or absence of the reducing substrate, phenol. In the reaction of prostaglandin H synthase with the peroxide in the absence of phenol, formation of the enzyme intermediate compound I is observed followed by partial conversion to compound II and then by enzyme bleaching. In the reaction with both peroxide and phenol the absorbance decreases and a steady-state spectrum is observed which is a mixture of native enzyme and compound II. The steady state is followed by an increase in absorbance back to that of the native enzyme with no bleaching. The difference can be explained by the reactivity of phenol as a reducing substrate with the prostaglandin H synthase intermediate compounds. Cyclooxygenase activity with arachidonic acid could not be examined in the absence of diethyldithiocarbamate because extensive bleaching occurred. In the presence of diethyldithiocarbamate, enzyme spectral behavior similar to that seen in the reaction of the peroxide and phenol was observed. The similarity of the spectra strongly suggests that the enzyme intermediates involved in both the peroxidase and cyclooxygenase reactions are the same.  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