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 objective of the present study is to delineate the role of active site arginine and histidine residues of horseradish peroxidase (HRP) in controlling iodide oxidation using chemical modification technique. The arginine specific reagent, phenylglyoxal (PGO) irreversibly blocks iodide oxidation following pseudofirst order kinetics with second order rate constant of 25.12 min-1 M-1. Radiolabelled PGO incorporation studies indicate an essential role of a single arginine residue in enzyme inactivation. The enzyme can be protected both by iodide and an aromatic donor such as guaiacol. Moreover, guaiacol-protected enzyme can oxidise iodide and iodide-protected enzyme can oxidise guaiacol suggesting the regulatory role of the same active site arginine residue in both iodide and guaiacol binding. The protection constant (Kp) for iodide and guaiacol are 500 and 10 M respectively indicating higher affinity of guaiacol than iodide at this site. Donor binding studies indicate that guaiacol competitively inhibits iodide binding suggesting their interaction at the same binding site. Arginine-modified enzyme shows significant loss of iodide binding as shown by increased Kd value to 571 mM from the native enzyme (Kd = 150 mM). Although arginine-modified enzyme reacts with H2O2 to form compound II presumably at a slow rate, the latter is not reduced by iodide presumably due to low affinity binding.The role of the active site histidine residue in iodide oxidation was also studied after disubstitution reaction of the histidine imidazole nitrogens with diethylpyrocarbonate (DEPC), a histidine specific reagent. DEPC blocks iodide oxidation following pseudofirst order kinetics with second order rate constant of 0.66 min-1 M-1. Both the nitrogens (, ) of histidine imidazole were modified as evidenced by the characteristic peak at 222 nm. The enzyme is not protected by iodide suggesting that imidazolium ion is not involved in iodide binding. Moreover, DEPC-modified enzyme binds iodide similar to the native enzyme. However, the modified enzyme does not form compound II but forms compound I only with higher concentration of H2O2 suggesting the catalytic role of this histidine in the formation and autoreduction of compound I. Interestingly, compound I thus formed is not reduced by iodide indicating block of electron transport from the donor to the compound I. We suggest that an active site arginine residue regulates iodide binding while the histidine residue controls the electron transfer to the heme ferryl group during oxidation.  相似文献   

2.
The pH rate profile for the hydrolysis of diethyl-p-nitrophenyl phosphate catalyzed by the phosphotriesterase from Pseudomonas diminuta shows a requirement for the deprotonation of an ionizable group for full catalytic activity. This functional group has an apparent pKa of 6.1 +/- 0.1 at 25 degrees C, delta Hion of 7.9 kcal/mol, and delta Sion of -1.4 cal/K.mol. The enzyme is not inactivated in the presence of the chemical modification reagents dithiobis-(2-nitrobenzoate), methyl methane thiosulfonate, carbodiimide, pyridoxal, butanedione, or iodoacetic acid and thus cysteine, asparate, glutamate, lysine, and arginine do not appear to be critical for catalytic activity. However, the phosphotriesterase is inactivated completely with methylene blue, Rose Bengal, or diethyl pyrocarbonate. The enzyme is not inactivated by diethyl pyrocarbonate in the presence of bound substrate analogs, and inactivation with diethyl pyrocarbonate is reversible upon addition of neutralized hydroxylamine. The modification of a single histidine residue by diethyl pyrocarbonate, as shown by spectrophotometric analysis, is responsible for the loss of catalytic activity. The pKinact for diethyl pyrocarbonate modification is 6.1 +/- 0.1 at 25 degrees C. These results have been interpreted to suggest that a histidine residue at the active site of phosphotriesterase is facilitating the reaction by general base catalysis.  相似文献   

3.
M Santimone 《Biochimie》1975,57(3):265-270
The kinetics of compound II formation, obtained upon mixing a highly purified horseradish peroxidase and hydrogen peroxide, was spectrophotometrically studied at three wavelengths in the absence of an added reducing agent. Our experiments confirm George's finding that more than one mole of compound II is formed per mole of hydrogen peroxide added. The new mechanism that we propose, contrary to the mechanism of George, is only valid when compound II is obtained in the absence of an added donor. Moreover, it is not inconsistent with the classical Chance mechanism of oxidation of an added donor by the system peroxidase -- hydrogen peroxide. According to this new mechanism, in the absence of an added donor, compound II formation involved two pathways. The first pathway is the monomolecular reduction of compound I by the endogenous donor, and the second pathway is the formation of two moles of compound II through the oxidoreduction reaction between one mole of peroxidase and one mole of compound I.  相似文献   

4.
R Roman  H B Dumbord 《Biochemistry》1972,11(11):2076-2082
  相似文献   

5.
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   

6.
Transient kinetic analysis of biphasic, single turnover data for the reaction of 2,2'-azino-bis[3-ethylbenzthiazoline-6-sulfonic acid] (ABTS) with horseradish peroxidase (HRPC) compound II demonstrated preequilibrium binding of ABTS (k(+5) = 7.82 x 10(4) M(-)(1) s(-)(1)) prior to rate-limiting electron transfer (k(+6) = 42.1 s(-)(1)). These data were obtained using a stopped-flow method, which included ascorbate in the reaction medium to maintain a low steady-state concentration of ABTS (pseudo-first-order conditions) and to minimize absorbance changes in the Soret region due to the accumulation of ABTS cation radicals. A steady-state kinetic analysis of the reaction confirmed that the reduction of HRPC compound II by this substrate is rate-limiting in the complete peroxidase cycle. The reaction of HRPC with o-diphenols has been investigated using a chronometric method that also included ascorbate in the assay medium to minimize the effects of nonenzymic reactions involving phenol-derived radical products. This enabled the initial rates of o-diphenol oxidation at different hydrogen peroxide and o-diphenol concentrations to be determined from the lag period induced by the presence of ascorbate. The kinetic analysis resolved the reaction of HRPC compound II with o-diphenols into two steps, initial formation of an enzyme-substrate complex followed by electron transfer from the substrate to the heme. With o-diphenols that are rapidly oxidized, the heterolytic cleavage of the O-O bond of the heme-bound hydrogen peroxide (k(+2) = 2.17 x 10(3) s(-)(1)) is rate-limiting. The size and hydrophobicity of the o-diphenol substrates are correlated with their rate of binding to HRPC, while the electron density at the C-4 hydroxyl group predominantly influences the rate of electron transfer to the heme.  相似文献   

7.
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   

8.
Lignin peroxidase oxidizes non-phenolic substrates by one electron to give aryl-cation-radical intermediates, which react further to give a variety of products. The present study investigated the possibility that other peroxidative and oxidative enzymes known to catalyse one-electron oxidations may also oxidize non-phenolics to cation-radical intermediates and that this ability is related to the redox potential of the substrate. Lignin peroxidase from the fungus Phanerochaete chrysosporium, horseradish peroxidase (HRP) and laccase from the fungus Trametes versicolor were chosen for investigation with methoxybenzenes as a homologous series of substrates. The twelve methoxybenzene congeners have known half-wave potentials that differ by as much as approximately 1 V. Lignin peroxidase oxidized the ten with the lowest half-wave potentials, whereas HRP oxidized the four lowest and laccase oxidized only 1,2,4,5-tetramethoxybenzene, the lowest. E.s.r. spectroscopy showed that this congener is oxidized to its cation radical by all three enzymes. Oxidation in each case gave the same products: 2,5-dimethoxy-p-benzoquinone and 4,5-dimethoxy-o-benzoquinone, in a 4:1 ratio, plus 2 mol of methanol for each 1 mol of substrate. Using HRP-catalysed oxidation, we showed that the quinone oxygen atoms are derived from water. We conclude that the three enzymes affect their substrates similarly, and that whether an aromatic compound is a substrate depends in large part on its redox potential. Furthermore, oxidized lignin peroxidase is clearly a stronger oxidant than oxidized HRP or laccase. Determination of the enzyme kinetic parameters for the methoxybenzene oxidations demonstrated further differences among the enzymes.  相似文献   

9.
The decarboxylation of retinoic acid by horseradish peroxidase was investigated. A marked increase in the yield of products was obtained. However, the data indicated the reaction was a nonenzymatic, heme catalyzed peroxidation. Previously reported requirements for phosphate, oxygen and ferrous ion were eliminated when hydrogen peroxide was provided. Peroxide also eliminated the EDTA and cyanide induced inhibition of the phosphate dependent system. In the presence of hydrogen peroxide, horseradish peroxidase was not essential to the reaction; heme equivalent amounts of hemoglobin decarboxylated retinoic acid with equal facility. However, hemoglobin was ineffective in the absence of hydrogen peroxide. Attainment of 50--60% decarboxylation represented complete utilization of the available retinoic acid. Thus the products of the reaction can be divided into two groups, products of retinoic acid oxidation and products of an oxidative decarboxylation of retinoic acid.  相似文献   

10.
Horseradish peroxidase was reacted with glutaraldehyde under various reaction conditions. The reaction product was, in a second step, bound covalently to aminohexyl groups attached to Sepharose particles. The influence of pH, time and the concentration ratio of enzyme:glutaraldehyde on the reaction was evaluated. A first step reaction with 100-fold molar excess of glutaraldehyde to horseradish peroxidase at pH 9.5 for 2 hr at room temperature results in a high yield of conjugated enzyme with well preserved enzymatic activity.  相似文献   

11.
12.
Histidine residues in horseradish peroxidase (HRP) were modified chemically with diethyl pyrocarbonate, 4,omega-dibromoacetophenone or diallylpyrocarbonate. Histidines were chosen as His-170, the fifth ligand of the heme iron atom, forms part of the active site of this enzyme. Good yields of hemoprotein were obtained in all cases. Analysis by HPLC of peptides obtained after tryptic digestion showed that His-170 of HRP was in fact modified. The specific activity remained satisfactory after chemical modification of the histidine residues, and so the active site of HRP can thus be altered without a dramatic loss of hemoprotein or peroxidase activity. This may open routes to the preparation of novel biocatalysts.  相似文献   

13.
Previous studies on the chlorination reaction catalyzed by horseradish peroxidase using chlorite as the source of chlorine detected the formation of a chlorinating intermediate that was termed Compound X (Shahangian, S., and Hager, L.P. (1982) J. Biol. Chem. 257, 11529-11533). These studies indicated that at pH 10.7, the optical absorption spectrum of Compound X was similar to the spectrum of horseradish peroxidase Compound II. Compound X was shown to be quite stable at alkaline pH values. This study was undertaken to examine the relationship between the oxidation state of the iron protoporphyrin IX heme prosthetic group in Compound X and the chemistry of the halogenating intermediate. The experimental results show that the optical absorption properties and the oxidation state of the heme prosthetic group in horseradish peroxidase are not directly related to the presence of the activated chlorine atom in the intermediate. The oxyferryl porphyrin heme group in alkaline Compound X can be reduced to a ferric heme species that still retains the activated chlorine atom. Furthermore, the reaction of chlorite with horseradish peroxidase at acidic pH leads to the secondary formation of a green intermediate that has the spectral properties of horseradish peroxidase Compound I (Theorell, H. (1941) Enzymologia 10, 250-252). The green intermediate also retains the activated chlorine atom. By analogy to peroxidase Compound I chemistry, the heme prosthetic group in the green chlorinating intermediate must be an oxyferryl porphyrin pi-cation radical species (Roberts, J. E., Hoffman, B. M., Rutter, R. J., and Hager, L. P. (1981) J. Am. Chem. Soc. 103, 7654-7656). To be consistent with traditional peroxidase nomenclature, the red alkaline form of Compound X has been renamed Compound XII, and the green acidic form has been named Compound XI. The transfer of chlorine from the chlorinating intermediate to an acceptor molecule follows an electrophilic (rather than a free radical) path. A mechanism for the reaction is proposed in which the activated chlorine atom is bonded to a heteroatom on an active-site amino acid side chain. Transient state kinetic studies show that the initial intermediate, Compound XII, is formed in a very fast reaction. The second-order rate constant for the formation of Compound XII is approximately 1.1 x 10(7) M-1 s-1. The rate of formation of Compound XII is strongly pH-dependent. At pH 9, the second-order rate constant for the formation of Compound XII drops to 1.5 M-1 s-1. At acidic pH values, Compound XII undergoes a spontaneous first-order decay to yield Compound XI.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

14.
Chloroperoxidase from Caldariomyces fumago is well documented as an extremely versatile catalyst, and studies are currently being conducted to delineate the fine structural features that allow the enzyme to possess chemical and physical similarities to the peroxidases, catalases, and P-450 cytochromes. Earlier investigations of ligand binding to the heme iron of chloroperoxidase, along with the presence of an invariant distal histidine residue in the active site of peroxidases and catalases, have led to the hypothesis that chloroperoxidase also possesses an essential histidine residue that may participate in catalysis. To address this in a more direct fashion, chemical modification studies were initiated with diethylpyrocarbonate. Incubation of chloroperoxidase with this reagent resulted in a time-dependent inactivation of enzyme. Kinetic analysis revealed that the inactivation was due to a simple bimolecular reaction. The rate of inactivation exhibited a pH dependence, indicating that modification of a titratable residue with a pKa value of 6.91 was responsible for inactivation; this data provided strong evidence for histidine derivatization by diethylpyrocarbonate. To further support these results, inactivation due to cysteine, tyrosine, or lysine modification was ruled out. The stoichiometry of histidine modification was estimated by the increase in absorption at 246 nm, and it was found that more than 1 histidine residue was derivatized when chloroperoxidase was inactivated with diethylpyrocarbonate. However, it was shown that the rates of modification and inactivation were not equivalent. This was interpreted to reflect that both essential and nonessential histidine residues were modified by diethylpyrocarbonate. Kinetic analysis indicated that modification of a single essential histidine residue was responsible for inactivation of the enzyme. Studies with [14C]diethylpyrocarbonate provided stoichiometric support that derivatization of a single histidine inactivated chloroperoxidase. Based on sequence homology with cytochrome c peroxidase, histidine 38 was identified as a likely candidate for the distal residue. Molecular modeling, based on secondary structure predictions, allows for the construction of an active site peptide, and implicates a number of other residues that may participate in catalysis.  相似文献   

15.
Sheep liver 6-phosphogluconate dehydrogenase is shown to be inactivated by diethylpyrocarbonate in a biphasic manner at pH 6.0, 25 degrees C. After allowing for the hydrolysis of the reagent, rate constants of 56 M-1 s-1 and 11.0 M-1 s-1 were estimated for the two processes. The complete reactivation of partially inactivated enzyme by neutral hydroxylamine, the elimination of the possibility that modification of cysteine or tyrosine residues are responsible for inactivation, and the magnitudes of the rate constants for inactivation relative to the experimentally determined value for the reaction of diethylpyrocarbonate with N alpha-acetylhistidine (2.2 M-1 s-1), all suggested that enzyme inactivation occurs solely by modification of histidine residues. Comparison of the experimental plot of residual fractional activity versus the number of modified histidine residues per subunit with simulated plots for three hypothetical models, each predicting biphasic kinetics, indicated that inactivation results from the modification of at most one essential histidine residue per subunit, although it appears that other (non-essential) histidines react independently. This histidine is thought to be His-242 and is present in the active site. Evidence in support of its role in catalysis is briefly discussed. Both 6-phosphogluconate and organic phosphate protect against inactivation, and a kinetic analysis of the protection indicated a dissociation constant of 2.1 X 10(-6) M for the enzyme--6-phosphogluconate complex. NADP+ also protected, but this might be due, at least in part, to a reduction in the effective concentration of diethylpyrocarbonate.  相似文献   

16.
Salicylic acid and phenylethylamine are putative substrates for naturally occurring reactions for generation of reactive oxygen species, which are catalyzed by plant peroxidases. Here, we used commercially available highly purified horseradish peroxidase-C (HRP-C) as a model enzyme for spectroscopic analysis, and obtained data suggesting that the Compound II form of HRP-C does not utilize phenylethylamine as substrate. In contrast, addition of salicylic acid to Compound II resulted in rapid conversion of Compound II to the native form.  相似文献   

17.
A kinetic study has been carried out over the pH range of 2.63-9.37 for the reaction of horseradish peroxidase with hydrogen peroxide to form compound I of th;e enzyme. Analysis of the results, indicates that there are two kinetic influencing, ionizable groups on the enzyme with pKa values of 3.2 and 3.9. Protonation of these groups results in a decrease in the rate of reaction of the enzyme with H2O2. A previous study of the kinetics of cyanide binding to horseradish peroxidase (Ellis, W.D. & Dunford, H.B.: Biochemistry 7, 2054-2062 (1968)) has been extended to down to pH 2.55, and analysis of these results also indicates the presence of two kinetically important ionizable groups on the enzyme with pKa values of 2.9 and 3.9.  相似文献   

18.
19.
Single-proton, exchangeable resonances have been detected in the high spin ferric hemoproteins, met-aquo myoglobin and horseradish peroxidase, which can be assigned to the proximal histidyl imidazole by virtue of their very large hyperfine shifts. While this proton is relatively labile in myoglobin, it is exchangeable in HRP only at extreme pH values, indicating a buried heme pocket. The insensitivity of the imidazole peak of HRP to substrate binding argues against direct interaction of imidazole and substrate.  相似文献   

20.
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