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.
In the presence of chloroperoxidase, indole was oxidized by H2O2 to give oxindole as the major product. Under most conditions oxindole was the only product formed, and under optimal conditions the conversion was quantitative. This reaction displayed maximal activity at pH 4.6, although appreciable activity was observed throughout the entire pH range investigated, namely pH 2.5-6.0. Enzyme saturation by indole could not be demonstrated, up to the limit of indole solubility in the buffer. The oxidation kinetics were first-order with respect to indole up to 8 mM, which was the highest concentration of indole that could be investigated. On the other hand, 2-methylindole was not affected by H2O2 and chloroperoxidase, but was a strong inhibitor of indole oxidation. The isomer 1-methylindole was a poor substrate for chloroperoxidase oxidation, and a weak inhibitor of indole oxidation. These results suggest the possibility that chloroperoxidase oxidation of the carbon atom adjacent to the nitrogen atom in part results from hydrogen-bonding of the substrate N-H group to the enzyme active site.  相似文献   

2.
A radical species of monochlorodimedone has been characterized by its high reactivity with molecular O2. Horseradish peroxidase greatly accelerated O2 uptake by acidic solutions of this substrate; the enzymatic reaction required exogenous H2O2 only with freshly prepared substrate solutions, and the total substrate oxidized was equal to the sum of H2O2 added and O2 consumed. However, with excess Br- and horseradish peroxidase, or high Br- or Cl- and chloroperoxidase, a 1:1 stoichiometry between H2O2 and substrate was observed. In the absence of halide, the stoichiometry of the chloroperoxidase-catalyzed oxidation of monochlorodimedone changed to two molecules of the organic donor per H2O2. Moreover, in the absence of halide, at substrate:H2O2 ratios greater than 2.0, chloroperoxidase catalyzed significant O2 uptake; this enzyme-dependent autoxidation of monochlorodimedone also occurred in the presence of Cl- or Br-, when H2O2 was limiting. These data, and recent evidence from this laboratory for free hypohalous acid as the first product of chloroperoxidase-catalyzed halide oxidation [B. W. Griffin (1983) Biochem. Biophys. Res. Commun. 116, 873-879], strongly support a mixed enzymatic/nonenzymatic radical chain process as the mechanism for halogenation of monochlorodimedone by chloroperoxidase. Both horseradish peroxidase and chloroperoxidase can catalyze either bromination or oxidation of this substrate, depending on the experimental conditions. Implications of these results for the mechanism of HOCl formation catalyzed by chloroperoxidase are considered.  相似文献   

3.
Fungal peroxygenases have recently been shown to catalyze remarkable oxidation reactions. The present study addresses the mechanism of benzylic oxygenations catalyzed by the extracellular peroxygenase of the agaric basidiomycete Agrocybe aegerita. The peroxygenase oxidized toluene and 4-nitrotoluene via the corresponding alcohols and aldehydes to give benzoic acids. The reactions proceeded stepwise with total conversions of 93% for toluene and 12% for 4-nitrotoluene. Using H218O2 as the co-substrate, we show here that H2O2 is the source of the oxygen introduced at each reaction step. A. aegerita peroxygenase resembles cytochromes P450 and heme chloroperoxidase in catalyzing benzylic hydroxylations.  相似文献   

4.
The incubation of 4-chloroaniline with H2O2 and myeloperoxidase results in the formation of at least 10 products. Possibly some structures with high complexity, like 4,4′-dichloroazobenzene, are present; however, no 4-chloronitrosobenzene is detectable. This result contrasts with the oxidation of 4-chloroaniline catalyzed by chloroperoxidase, which only yields 4-chloronitrosobenzene.  相似文献   

5.
A sigmoidal kinetic behavior of chloroperoxidase for the oxidation of 4,6-dimethyldibenzothiophene (4,6-DMDBT) in water-miscible organic solvent is for the first time reported. Kinetics of 4,6-DMDBT oxidation showed a cooperative profile probably due to the capacity of chloroperoxidase to recognize a substrate dimer (pi-pi dimer) in its active site. Experimental evidence is given for dimer formation and its presence in the active site of chloroperoxidase. The kinetic data were adjusted for a binding site able to interact with either monomer or dimer substrates, producing a cooperative model describing a one-site binding of two related species. Determination of kinetics constants by iterative calculations of possible oxidation paths of 4,6-DMDBT suggests that kinetics oxidation of dimer substrate is preferred when compared to monomer oxidation. Steady-state fluorometry of substrate in the absence and presence of chloroperoxidase, described by the spectral center of mass, supports this last conclusion.  相似文献   

6.
The chloroperoxidase catalyzed oxidation of methyl phenyl sulfide to (R)-methyl phenyl sulfoxide was investigated, both in batch and membrane reactors, using as oxidant H2O2, or O2 in the presence of either dihydroxyfumaric acid or ascorbic acid. The effects of pH and nature and concentration of the oxidants on the selectivity, stability, and productivity of the enzyme were evaluated. The highest selectivity was displayed by ascorbic acid/O2, even though the activity of chloroperoxidase with this system was lower than that obtained with the others. When the reaction was carried out in a membrane reactor, it was possible to reuse the enzyme for several conversion cycles. The results obtained with ascorbic acid/O2 and dihydroxyfumaric acid/O2 as oxidants do not seem to be compatible with either a mechanism involving hydroxyl radicals as the active species or with the hypothesis that oxidation occurs through the initial formation of H2O2. Copyright 1999 John Wiley & Sons, Inc.  相似文献   

7.
An analysis of the hydroxamic acid oxidation reaction by H2O2 and horseradish peroxidase (HRP) was made with three pairs of hydroxamic acids. Each pair consisted of the aceto- and glycolhydroxamic acid derivatives from one of three different arylhydroxylamines. The parent arylhydroxylamines were the known carcinogens, N-hydroxy-2-aminofluorene and N-hydroxy-4-aminobiphenyl and the noncarcinogen 4-chlorophenyl-hydroxylamine. All the hydroxamic acids appeared to be converted to products that were expected on the basis of the previously-proposed mechanism of this peroxidative reaction. Each acetohydroxamic acid gave the corresponding nitroso compound and O-acetyl ester of the starting material in approximately equal amounts. The glycolhydroxamic acids gave the corresponding nitroso compound and a relatively unstable product that was proposed, by analogy, to be the O-glycolyl ester of the starting material. A comparison of the initial rates of reaction of each hydroxamic acid pair showed that the glycolhydroxamic acid was much more susceptible to the peroxidation reaction than was the corresponding acetohydroxamic acid. The initial rate of the reaction was also highly dependent upon the nature of the aromatic ring in the order fluorene greater than biphenyl greater than 4-chlorophenyl. The relative degree of HRP-catalyzed covalent binding to DNA of the aceto- and glycolhydroxamic acids in the fluorene series was studied and found to parallel the relative rates of reaction of these substrates in the H2O2/HRP system. It was proposed that glycolhydroxamic acids are likely to be more genotoxic than are acetohydroxamic acids when subjected to peroxidative bioactivation conditions.  相似文献   

8.
NADPH-dependent oxidation of 1,3-butadiene by mouse liver microsomes or H2O2-dependent oxidation by chloroperoxidase produced both butadiene monoxide and crotonaldehyde; methyl vinyl ketone and 2,3- and 2,5- dihydrofuran were not detected. The crotonaldehyde to butadiene monoxide ratio remained constant over time in both the microsomal and the chloroperoxidase reactions; however, much more crotonaldehyde was produced by chloroperoxidase than microsomes; crotonaldehyde was not detected when reference samples of butadiene monoxide were used in control incubations containing NADPH and microsomes or H2O2 and chloroperoxidase. Moreover, incubations of 1,3-butadiene with horseradish peroxidase and H2O2, or microsomes and H2O2 or arachidonic acid did not result in the oxidation of 1,3-butadiene. In microsomes, metabolite formation was dependent on incubation time, NADPH, and protein concentrations and did not change when the 1,3-butadiene pressure was varied between 24 and 52 cm Hg. Inclusion of the cytochrome P450 inhibitor 1-benzylimidazole inhibited 1,3-butadiene metabolism, but inclusion of KCN, catalase, or superoxide dismutase had no effect. These results support the role of cytochrome P450 in 1,3-butadiene oxidation by mouse liver microsomes. The formation of crotonaldehyde but not methyl vinyl ketone by cytochrome P450 or chloroperoxidase indicates regioselectivity in the oxygen transfer from the hemoproteins to 1,3-butadiene. The intermediates formed may undergo either ring closure to form butadiene monoxide or a hydrogen shift to form 3-butenal which tautomerizes to produce crotonaldehyde. Evidence for this tautomerization was obtained by the finding that 3-buten-1-ol, an alternative precursor of 3-butenal, was oxidized to crotonaldehyde under incubation conditions similar to that used for 1,3-butadiene.  相似文献   

9.
Asymmetric sulfoxidation by means of a chloroperoxidase from Caldariomyces fumago and hydrogen peroxide as the oxygen source was studied for a series of sterically well-defined substrates. The stereochemistry of the sulfoxidation was the same for all substrates studied. While 2,3-dihydrobenzo[b]thiophene (1) is an excellent substrate (giving 99.5% yield and 99% e.e. of the (R)-sulfoxide), replacement of a methylene group by either a more sterically demanding group or a heteroatom caused a substantial decrease in reactivity or in reactivity as well as enantioselectivity. A further investigation of the lowered catalytic efficiency of chloroperoxidase with these substrates has been carried out in a series of competitive reactions. Thus, benzo[1,3]oxathiole (5) acted as a competitive inhibitor of the enzyme, whereas 1-thiochroman (2) and 1-thiochroman-4-one (3) were shown to be too sterically demanding to significantly compete for the active site. For the oxidation of 2, 3, and 5, it was found that in the low CPO concentration range the chemical yield after 60 min reaction time increased almost linearly with the amount of CPO used. The products from 2 and 3 could be obtained in over 80% yield with an e.e. exceeding 96%. Chloroperoxidase was also found to be an effective catalyst in the oxidation of labile episulfides, yielding the corresponding anti-sulfoxides quantitatively and giving 12% e.e. of (1R, 2R)-sulfoxide in the oxidation of propylene sulfide. Chirality 10:246252, 1998. © 1998 Wiley-Liss, Inc.  相似文献   

10.
In this report, we have examined the origin of the electrophilic chlorine formed during the microsomal metabolism of carbon tetrachloride and the possibility that liver microsomal proteins catalyze chloroperoxidase or myeloperoxidase halogenation reactions. Studies with stable isotopes of chlorine show that at least 99% of the trapped chlorine originated from carbon tetrachloride. When hydrogen peroxide or cumene hydroperoxide was added to liver microsomes in the presence of chloride ion, no trapped chlorine was observed. Thus, cytochrome P-450 does not catalyze chloroperoxidase type chloride ion oxidation but instead catalyzes a reaction leading to cleavage of a carbon-chlorine bond with concomitant chlorine atom oxidation.  相似文献   

11.
Caldariomyces fumago chloroperoxidase degrades monofluorophenols at both pH 3 and pH 6. 4-Fluorophenol is most readily degraded and its oxidation is most efficient at pH 6. GC-MS analyses of the reaction products revealed compounds relating to the reaction of fluorophenol radical. The degradation of fluorinated compounds is of significant environmental interest and this versatile enzyme may by employed to treat contaminated soil or water prior to discharge.  相似文献   

12.
The reaction of the hydrated electron with a ferrous oxygenated form of modified cytochrome P-450cam, containing 2,4-diacetyldeuteroheme, was investigated by the use of pulse radiolysis. The ferrous oxygenated form of this enzyme was reduced by hydrated electrons to form the product, which exhibits absorption maximum at 470 and 370 nm. From the spectrum obtained, the oxidation state of the product is discussed in relation to the higher oxidation states of chloroperoxidase.  相似文献   

13.
Several 2,3-unsaturated carboxylic acids, such as trans-cinnamic acid and its derivatives, were found to be halogenated by chloroperoxidase of Caldariomyces fumago in the presence of hydrogen peroxide and either Cl- or Br-. Cinnamic acid, 4-hydroxycinnamic acid, 4-methoxycinnamic acid, and 3,4-dimethoxycinnamic acid were suitable substrates of chloroperoxidase, and were converted to 2-halo-3-hydroxycarboxylic acid, 2,3-dihydroxycarboxylic acid, decarboxylated halohydrin, or decarboxylated halocompound. However, 4-nitrocinnamic acid and 4-chlorocinnamic acid having electron-attracting groups did not serve as a substrate of the enzyme. The enzyme also did not act on acrylic acid, acrylamide, crotonic acid, fumaric acid, etc. From these data, the enzymatic reactions of chloroperoxidase, concerning the substrate specificity, stereoselectivity, and the reaction mechanism, are discussed on the basis of current knowledge regarding the reaction mechanism of the enzyme. Also they are compared with the chemical reactions of molecular halogen and hypohalous acid.  相似文献   

14.
Bromination of the thiazole ring by the enzyme chloroperoxidase, produced by the fungus Caldariomyces fumago, was demonstrated. Both 2-acetoacetamido-4-methylthiazole and 2-acetamidothiazole were brominated on C-5 of the thiazole ring in the presence of chloroperoxidase, bromide, and hydrogen peroxide in 0.06 M Phosphate solution at pH 3.0. No reaction occurred in the absence of enzyme.  相似文献   

15.
Both the kinetics of ferric chloroperoxidase reduction by dithionite and the binding of molecular oxygen to ferrous chloroperoxidase have been studied. The oxyferrous chloroperoxidase decays spontaneously to the ferric enzyme. In addition the corresponding rapid-scan spectra have been recorded. The reduction reaction is caused by SO-.2 with a rate constant of (7.7 +/- 1.0) X 10(4) M-1 S-1. Oxygen binding occurs with a rate constant of (5.5 +/- 1.0) X 10(5) M-1 S-1 over the pH range 3.5-6. Oxyferrous chloroperoxidase has a Soret absorption peak at 428 nm and two partially resolved peaks at 555 nm and 588 nm. Isosbestic points occur at the following wavelengths: between ferrous and oxyferrous chloroperoxidase at 419, 545, 555 and 580 nm; between oxyferrous and ferric chloroperoxidase at 419, 487, 540, 609 and 682 nm.  相似文献   

16.
The first complete mechanistic analysis of halide ion oxidation by a peroxidase was that of iodide oxidation by horseradish peroxidase. It was shown conclusively that a two-electron oxidation of iodide by compound I was occurring. This implied that oxygen atom transfer was occurring from compound I to iodide, forming hypoiodous acid, HOI. Searches were conducted for other two-electron oxidations. It was found that sulfite was oxidized by a two-electron mechanism. Nitrite and sulfoxides were not. If a competing substrate reduces some compound I to compound II by the usual one-electron route, then compound II will compete for available halide. Thus compound II oxidizes iodide to an iodine atom, I*, although at a slower rate than oxidation of I by compound I. An early hint that mammalian peroxidases were designed for halide ion oxidation was obtained in the reaction of lactoperoxidase compound II with iodide. The reaction was accelerated by excess iodide, indicating a co-operative effect. Among the heme peroxidases, only chloroperoxidase (for example from Caldariomyces fumago) and mammalian myeloperoxidase are able to oxidize chloride ion. There is not yet a consensus as to whether the chlorinating agent produced in a peroxidase-catalyzed reaction is hypochlorous acid (HOCl), enzyme-bound hypochlorous acid (either Fe-HOCl or X-HOCl where X is an amino acid residue), or molecular chlorine Cl2. A study of the nonenzymatic iodination of tyrosine showed that the iodinating reagent was either HOI or I2. It was impossible to tell which species because of the equilibria: [reaction: see text] The same considerations apply to product analysis of an enzyme-catalyzed reaction. Detection of molecular chlorine Cl2 does not prove it is the chlorinating species. If Cl2 is in equilibrium with HOCl then one cannot tell which (if either) is the chlorinating reagent. Examples will be shown of evidence that peroxidase-bound hypochlorous acid is the chlorinating agent. Also a recent clarification of the mechanism of reaction of myeloperoxidase with hydrogen peroxide and chloride along with accurate determination of the elementary rate constants will be discussed.  相似文献   

17.
A novel reaction system was developed to maximize the catalytic efficiency of chloroperoxidase (CPO, from Caldariomyces fumago) toward the oxidation of hydrocarbons. The reaction system consisted of an organic/aqueous emulsion comprising pure substrate and aqueous buffer supplemented with the surfactant dioctyl sulfosuccinate. The emulsion system attenuated not only the destabilizing effects of the substrate and product on the enzyme by emulsifying the compounds, but also oxidant toxicity (oxidative stress) by increasing substrate availability. As a result, CPO exhibited total turnover numbers (TTNs, defined as the amount of product produced over the catalytic lifetime of the enzyme) of ca. 20,000 mol product/mol enzyme for the oxidation of styrene, toluene, and o-, m-, p-xylenes. The TTNs are over 10-fold higher than those previously reported for the oxidation of benzylic hydrocarbons by CPO. This study represents a significant step toward the development of CPO as a practical catalyst for large-scale organic syntheses.  相似文献   

18.
The effect of D2O on the kinetic parameters for the hydroperoxide-supported N-demethylation of N,N-dimethylaniline catalyzed by chloroperoxidase and horseradish peroxidase was investigated in order to assess the roles of exchangeable hydrogens in the demethylation reaction. The initial rate of the chloroperoxidase-catalyzed N-demethylation of N,N-dimethylaniline supported by ethyl hydroperoxide exhibited a pL optimum (where L denotes H or D) of 4.5 in both H2O and D2O. The solvent isotope effect on the initial rate of the chloroperoxidase-catalyzed demethylation reaction was independent of pL, suggesting that the solvent isotope effect is not due to a change in the pK of a rate-controlling ionization in D2O. The solvent isotope effect on the Vmax for the chloroperoxidase-catalyzed demethylation reaction was 3.66 +/- 0.62. In contrast, the solvent isotope effect on the Vmax for the horseradish peroxidase catalyzed demethylation reaction was approximately 1.5 with either ethyl hydroperoxide or hydrogen peroxide as the oxidant, indicating that the exchange of hydrogens in the enzyme and hydroperoxide for deuterium in D2O has little effect on the rate of the demethylation reaction. The solvent isotope effect on the Vmax/KM for ethyl hydroperoxide in the chloroperoxidase-catalyzed demethylation reaction was 8.82 +/- 1.57, indicating that the rate of chloroperoxidase compound I formation is substantially decreased in D2O. This isotope effect is suggested to arise from deuterium exchange of the hydroperoxide hydrogen and of active-site residues involved in compound I formation. A solvent isotope effect of 2.96 +/- 0.57 was observed on the Vmax/KM for N,N-dimethylaniline in the chloroperoxidase-catalyzed reaction.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

19.
The chloroperoxidase (CPO)-catalyzed oxidation of the methyl (2E)-2,4-pentadienoate gives the terminal double bond epoxide (25%) and a cyclodimerization compound (63%) as the major products.  相似文献   

20.
The vanadate cofactor in vanadium chloroperoxidase has been studied using UV-VIS absorption spectroscopy. A band is present in the near-UV that is red-shifted as compared to free vanadate and shifts in both position and intensity upon change in pH. Mutation of vanadate binding residues has a clear effect on the spectrum. Substrate-induced spectral effects allow direct measurement of separate kinetics steps for the first time for vanadium haloperoxidases. A peroxo intermediate is formed upon addition of H(2)O(2), which causes a decrease in the absorption spectrum at 315 nm, as well as an increase at 384 nm. This peroxo form is very stable at pH 8.3, whereas it is less stable at pH 5.0, which is the optimal pH for activity. Upon addition of halides to the peroxo form, the native spectrum is re-formed as a result of halide oxidation. Stopped-flow experiments show that H(2)O(2) binding and Cl(-) oxidation occur on the millisecond to second time scale. These data suggest that the oxidation of Cl(-) to HOCl occurs in at least two steps. In the presence of H(2)O(2), the affinity for the vanadate cofactor was found to be much higher than previously reported for vanadate in the absence of H(2)O(2). This is attributed to the uptake of pervanadate by the apo-enzyme. Human glucose-6-phosphatase, which is evolutionarily related to vanadium chloroperoxidase, is also likely to have a higher affinity for pervanadate than vanadate. This could explain the enhanced insulin mimetic effect of pervanadate as compared to vanadate.  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