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 828 毫秒
1.
Three newly discovered non-heme bromoperoxidases isolated from marine algae were found to catalyze the production of singlet oxygen in reactions composed of the bromoperoxidase, hydrogen peroxide, and bromide. The bromoperoxidases studied were vanadium bromoperoxidase (V-BrPO) from Ascophyllum nodosum, native non-heme bromoperoxidase from Corallina vancouveriensis (which contains vanadium and iron), and the vanadium-reconstituted bromoperoxidase derivative from C. vancouveriensis. These enzyme systems generated near infrared emission, characteristic of singlet oxygen. The emission had a peak intensity near 1268 nm, was greatly increased in 2H2O-containing buffers, and was greatly decreased by the singlet oxygen quenchers, histidine and azide. The yield of singlet oxygen was approximately 80% of the theoretical yield. A unique feature of the non-heme bromoperoxidases distinct from the iron heme haloperoxidases, was the remarkable stability of the non-heme enzymes in the presence of singlet oxygen and oxidized bromine species. V-BrPO turned over multiple aliquots of 2 mM hydrogen peroxide without losing efficiency. In contrast, iron heme lactoperoxidase was completely inactivated after turnover of the first aliquot of 2 mM hydrogen peroxide, and iron heme chloroperoxidase was 50% deactivated. The profile of singlet oxygen formation by V-BrPO and the near stoichiometric yield of singlet oxygen suggest that the mechanism of singlet oxygen formation is the same as the mechanism of dioxygen formation determined by oxygen probe measurements.  相似文献   

2.
Coding DNA of the tobacco anionic peroxidase gene was cloned in pET40b vector. The problem of 11 arginine codons, rare in procaryotes, in the tobacco peroxidase gene was solved using E. coli BL21(DE3) Codon Plus strain. The expression level of the tobacco apo-peroxidase in the above strain was approximately 40% of the total E. coli protein. The tobacco peroxidase refolding was optimized based on the earlier developed protocol for horseradish peroxidase. The reactivation yield of recombinant tobacco enzyme was about 7% with the specific activity of 1100-1200 U/mg towards 2,2;-azino-bis(3-ethylbenzothiazoline-6-sulfonate) (ABTS). It was shown that the reaction of ABTS oxidation by hydrogen peroxide catalyzed by recombinant tobacco peroxidase proceeds via the ping-pong kinetic mechanism as for the native enzyme. In the presence of calcium ions, the recombinant peroxidase exhibits a 2.5-fold decrease in the second order rate constant for hydrogen peroxide and 1.5-fold decrease for ABTS. Thus, calcium ions have an inhibitory effect on the recombinant enzyme like that observed earlier for the native tobacco peroxidase. The data demonstrate that the oligosaccharide part of the enzyme has no effect on the kinetic properties and calcium inhibition of tobacco peroxidase.  相似文献   

3.
Inactivation mechanisms of chloroperoxidase (CPO) from Caldariomyces fumago have been investigated with the aim of improving the practical utility of CPO for hydrocarbon oxidation. Deactivation studies in the presence of oxidants (i.e., hydrogen peroxide and t-butyl hydroperoxide) indicated that CPO lost oxidation activity toward hydrocarbon substrates during dismutation of hydrogen peroxide. The loss of enzyme activity was accompanied by the apparent destruction of the heme rather than aggregation or denaturation of the apo-protein. The decrease of enzyme activity was significantly retarded by adding the radical scavenger t-butyl alcohol at pH 4.1, or by optimizing the reaction pH. CPO retained greatest oxidation activity at pH 5-6, which may produce a more favorable ionization state of the key amino acid (Glu-183) and thus reduce radical formation. As a result of higher activity at pH 5-6, the total turnover numbers (TTN, defined as the amount of product produced over the catalytic lifetime of the enzyme) for the oxidation of toluene and o-, m-, p-xylenes in substrate/aqueous emulsion systems ranged from ca. 10% to 110% higher at pH 5.5 (20,000 to 45,000 mol product/mol enzyme) compared to pH 4.1. Furthermore, TTNs of CPO increased with increasing turnover frequencies, indicating that higher activity toward reducing substrates reduces radical formation and stabilizes CPO toward inactivation by H(2)O(2). These findings demonstrate the important relationship between CPO stability and activity, and illustrate that large improvements in CPO activity and stability can be achieved through solvent engineering.  相似文献   

4.
Seed coat soybean peroxidase (SBP) belongs to class III of the plant peroxidase superfamily that includes the classical peroxidase, namely horseradish peroxidase (HRP). We have measured the catalytic activity (k(cat)) and catalytic efficiency (k(cat)/K(M)) of SBP and that of HRP-C for the oxidation of ABTS [2,2'-azino-bis-(3-ethylbenzthiazoline-6-sulphonate)] by hydrogen peroxide at 25 degrees C. We observed that the k(cat) and k(cat)/K(M) values for SBP are much higher than those for HRP-C at all pH values, rendering SBP a more potent peroxidase. This is attributed to the relatively more solvent exposed delta-meso heme edge in SBP. We observed that the maximum catalytic activity and conformational stability of SBP is at pH approximately 5.5. A pH maximum of 5.0 for the catalytic activity of SBP has recently been reported. Estimation of secondary structural elements at various pH values indicated that there is a maximal reduction of beta-strands and beta-turns at pH 5.5 causing the heme to be further exposed to the solvent and increasing the overall conformational flexibility of the protein.  相似文献   

5.
Site-directed mutagenesis has been used to investigate the role of glutamic acid 183 in chloroperoxidase catalysis. Based on the x-ray crystallographic structure of chloroperoxidase, Glu-183 is postulated to function on distal side of the heme prosthetic group as an acid-base catalyst in facilitating the reaction between the peroxidase and hydrogen peroxide with the formation of Compound I. In contrast, the other members of the heme peroxidase family use a histidine residue in this role. Plasmids have now been constructed in which the codon for Glu-183 is replaced with a histidine codon. The mutant recombinant gene has been expressed in Aspergillus niger. An analysis of the produced mutant gene shows that the substitution of Glu-183 with a His residue is detrimental to the chlorination and dismutation activity of chloroperoxidase. The activity is reduced by 85 and 50% of wild type activity, respectively. However, quite unexpectedly, the epoxidation activity of the mutant enzyme is significantly enhanced approximately 2.5-fold. These results show that Glu-183 is important but not essential for the chlorination activity of chloroperoxidase. It is possible that the increased epoxidation of the mutant enzyme is based on an increase in the hydrophobicity of the active site.  相似文献   

6.
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   

7.
Agrocybe aegerita, a bark mulch- and wood-colonizing basidiomycete, was found to produce a peroxidase (AaP) that oxidizes aryl alcohols, such as veratryl and benzyl alcohols, into the corresponding aldehydes and then into benzoic acids. The enzyme also catalyzed the oxidation of typical peroxidase substrates, such as 2,6-dimethoxyphenol (DMP) or 2,2′-azinobis-(3-ethylbenzothiazoline-6-sulfonate) (ABTS). A. aegerita peroxidase production depended on the concentration of organic nitrogen in the medium, and highest enzyme levels were detected in the presence of soybean meal. Two fractions of the enzyme, AaP I and AaP II, which had identical molecular masses (46 kDa) and isoelectric points of 4.6 to 5.4 and 4.9 to 5.6, respectively (corresponding to six different isoforms), were identified after several steps of purification, including anion- and cation-exchange chromatography. The optimum pH for the oxidation of aryl alcohols was found to be around 7, and the enzyme required relatively high concentrations of H2O2 (2 mM) for optimum activity. The apparent Km values for ABTS, DMP, benzyl alcohol, veratryl alcohol, and H2O2 were 37, 298, 1,001, 2,367 and 1,313 μM, respectively. The N-terminal amino acid sequences of the main AaP II spots blotted after two-dimensional gel electrophoresis were almost identical and exhibited almost no homology to the sequences of other peroxidases from basidiomycetes, but they shared the first three amino acids, as well as two additional amino acids, with the heme chloroperoxidase (CPO) from the ascomycete Caldariomyces fumago. This finding is consistent with the fact that AaP halogenates monochlorodimedone, the specific substrate of CPO. The existence of haloperoxidases in basidiomycetous fungi may be of general significance for the natural formation of chlorinated organic compounds in forest soils.  相似文献   

8.
The crystallographic structures of both the vanadium chloroperoxidase and bromoperoxidase enzymes have been determined with either vanadium or phosphate bound at their active site. The amino acids that are involved in phosphate binding in the acid phosphatase enzymes and those that are coordinated to vanadium in the haloperoxidases appear to be conserved between the two classes of enzyme. The detailed active site architecture for enzymes that recognize and use either vanadium or phosphate will be discussed in relation to their proposed enzymatic mechanism.  相似文献   

9.
We report the first characterization and classification of Orf13 (S. refuineus) as a heme-dependent peroxidase catalyzing the ortho-hydroxylation of L-tyrosine to L-DOPA. The putative tyrosine hydroxylase coded by orf13 of the anthramycin biosynthesis gene cluster has been expressed and purified. Heme b has been identified as the required cofactor for catalysis, and maximal L-tyrosine conversion to L-DOPA is observed in the presence of hydrogen peroxide. Preincubation of L-tyrosine with Orf13 prior to the addition of hydrogen peroxide is required for L-DOPA production. However, the enzyme becomes inactivated by hydrogen peroxide during catalysis. Steady-state kinetic analysis of L-tyrosine hydroxylation revealed similar catalytic efficiency for both L-tyrosine and hydrogen peroxide. Spectroscopic data from a reduced-CO(g) UV-vis spectrum of Orf13 and electron paramagnetic resonance of ferric heme Orf13 are consistent with heme peroxidases that have a histidyl-ligated heme iron. Contrary to the classical heme peroxidase oxidation reaction with hydrogen peroxide that produces coupled aromatic products such as o,o'-dityrosine, Orf13 is novel in its ability to catalyze aromatic amino acid hydroxylation with hydrogen peroxide, in the substrate addition order and for its substrate specificity for L-tyrosine. Peroxygenase activity of Orf13 for the ortho-hydroxylation of L-tyrosine to L-DOPA by a molecular oxygen dependent pathway in the presence of dihydroxyfumaric acid is also observed. This reaction behavior is consistent with peroxygenase activity reported with horseradish peroxidase for the hydroxylation of phenol. Overall, the putative function of Orf13 as a tyrosine hydroxylase has been confirmed and establishes the first bacterial class of tyrosine hydroxylases.  相似文献   

10.
Agrocybe aegerita, a bark mulch- and wood-colonizing basidiomycete, was found to produce a peroxidase (AaP) that oxidizes aryl alcohols, such as veratryl and benzyl alcohols, into the corresponding aldehydes and then into benzoic acids. The enzyme also catalyzed the oxidation of typical peroxidase substrates, such as 2,6-dimethoxyphenol (DMP) or 2,2'-azinobis-(3-ethylbenzothiazoline-6-sulfonate) (ABTS). A. aegerita peroxidase production depended on the concentration of organic nitrogen in the medium, and highest enzyme levels were detected in the presence of soybean meal. Two fractions of the enzyme, AaP I and AaP II, which had identical molecular masses (46 kDa) and isoelectric points of 4.6 to 5.4 and 4.9 to 5.6, respectively (corresponding to six different isoforms), were identified after several steps of purification, including anion- and cation-exchange chromatography. The optimum pH for the oxidation of aryl alcohols was found to be around 7, and the enzyme required relatively high concentrations of H(2)O(2) (2 mM) for optimum activity. The apparent K(m) values for ABTS, DMP, benzyl alcohol, veratryl alcohol, and H(2)O(2) were 37, 298, 1,001, 2,367 and 1,313 microM, respectively. The N-terminal amino acid sequences of the main AaP II spots blotted after two-dimensional gel electrophoresis were almost identical and exhibited almost no homology to the sequences of other peroxidases from basidiomycetes, but they shared the first three amino acids, as well as two additional amino acids, with the heme chloroperoxidase (CPO) from the ascomycete Caldariomyces fumago. This finding is consistent with the fact that AaP halogenates monochlorodimedone, the specific substrate of CPO. The existence of haloperoxidases in basidiomycetous fungi may be of general significance for the natural formation of chlorinated organic compounds in forest soils.  相似文献   

11.
A modified cytochrome c peroxidase was prepared by reconstituting apocytochrome c peroxidase with protoheme in which both heme propionic acid groups were converted to the methyl ester derivatives. The modified enzyme reacted with hydrogen peroxide with a rate constant of (1.3 +/- 0.2) x 10(7) M-1 s-1, which is 28% that of the native enzyme. The reaction between the modified enzyme and hydrogen peroxide was pH-dependent with an apparent pK of 5.1 +/- 0.1 compared to a value of 5.4 +/- 0.1 for the native enzyme. These observations support the conclusion that the apparent ionization near pH 5.4, which influences the hydrogen peroxide-cytochrome c peroxidase reaction is not due to the ionization of the propionate side chains of the heme group in the native enzyme. A second apparent ionization, with pK of 6.1 +/- 0.1, influences the spectrum of the modified enzyme which changes from a high spin type at low pH to a low spin type at high pH.  相似文献   

12.
Chloroperoxidase,a peroxidase with potential   总被引:1,自引:0,他引:1  
Summary Chloroperoxidase is an extracellular heme glycoprotein produced by the imperfect fungusCaldariomyces fumago. The enzyme can catalyse chlorination reactions as well as act as a catalase or a peroxidase. As a peroxidase, it has a wide substrate specificity and we are interested in some applied aspects of this activity, requiring the production and purification of moderate quantities of the enzyme. High levels of chloroperoxidase are produced in a fructose synthetic medium, and highest enzyme production occurs in a low-shear environment. fungal pellets produce enzyme continuously at low medium replacement rates and at up to 0.6 g enzyme per 1: chloroperoxidase is essentially the only extracellular enzyme produced. Enzyme purification is uncomplicated and gives good yields of high purity. Pure enzyme is stable for weeks at room temperature and under pH control. Chloroperoxidase can be ionically bound to aminopropyl glass, then covalently immobilized by glutaraldehyde crosslinking. Immobilized preparations have been washed and re-used five times, and are most stable at pH 5.5-6. Like many peroxidases, chloroperoxidase will oxidize phenols and phenolics, often causing a precipitate, and can totally remove phenols at low aqueous concentrations. Chloroperoxidase incubation with the petroporphyrin component of crude oil asphaltene (fraction 5) causes a reduction or removal of the Soret band (410 nm) and the -peak (573 nm). This petroporphyrin fraction is enriched with vanadium which poisons the chemical catalyst used in cracking crude oil.  相似文献   

13.
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   

14.
Nagababu E  Rifkind JM 《Biochemistry》2000,39(40):12503-12511
The reaction of Fe(II) hemoglobin (Hb) but not Fe(III) hemoglobin (metHb) with hydrogen peroxide results in degradation of the heme moiety. The observation that heme degradation was inhibited by compounds, which react with ferrylHb such as sodium sulfide, and peroxidase substrates (ABTS and o-dianisidine), demonstrates that ferrylHb formation is required for heme degradation. A reaction involving hydrogen peroxide and ferrylHb was demonstrated by the finding that heme degradation was inihibited by the addition of catalase which removed hydrogen peroxide even after the maximal level of ferrylHb was reached. The reaction of hydrogen peroxide with ferrylHb to produce heme degradation products was shown by electron paramagnetic resonance to involve the one-electron oxidation of hydrogen peroxide to the oxygen free radical, superoxide. The inhibition by sodium sulfide of both superoxide production and the formation of fluorescent heme degradation products links superoxide production with heme degradation. The inability to produce heme degradation products by the reaction of metHb with hydrogen peroxide was explained by the fact that hydrogen peroxide reacting with oxoferrylHb undergoes a two-electron oxidation, producing oxygen instead of superoxide. This reaction does not produce heme degradation, but is responsible for the catalytic removal of hydrogen peroxide. The rapid consumption of hydrogen peroxide as a result of the metHb formed as an intermediate during the reaction of reduced hemoglobin with hydrogen peroxide was shown to limit the extent of heme degradation.  相似文献   

15.
Human recombinant myeloperoxidase (recMPO), purified from an engineered Chinese hamster ovary (CHO) cell line, has been characterized and compared to the mature enzyme isolated from polymorphonuclear leukocytes. Both molecules appear essentially similar in physicochemical enzymatic terms according to the following observations. 1. The unprocessed recombinant protein displays the characteristic light absorption spectra of ferric mature MPO and exhibits its typical spectral changes in the presence of dithionite or hydrogen peroxide. 2. The addition of 14C-labeled 5-aminolevulinic acid, a heme precursor, to the culture medium of recombinant CHO cells yields labeled recMPO, indicating the presence of a heme-like structure in the molecule. 3. Like mature MPO, recMPO has a peroxidatic activity and catalyzes the oxidation of chloride ions in the presence of hydrogen peroxide, producing hypochlorous acid as measured by the monochlorodimedon assay. For both enzymes, the chlorinating activity optimally occurs around pH 5.0 at about 100 microM of hydrogen peroxide and is strongly inhibited by methimazole. 4. Diethylpyrocarbonate significantly reduces the enzymatic activity of both molecules, suggesting that histidine residues may be of prime importance in the active site of the enzymes. 5. According to infrared spectroscopy data, both enzymes present a very similar secondary structure organization. In conclusion, the data suggest that the processing of the precursor enzyme (recMPO) into the mature form occurs without major structural and functional consequences.  相似文献   

16.
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   

17.
Directed evolution was performed on vanadium chloroperoxidase from the fungus Curvularia inaequalis to increase its brominating activity at a mildly alkaline pH for industrial and synthetic applications and to further understand its mechanism. After successful expression of the enzyme in Escherichia coli, two rounds of screening and selection, saturation mutagenesis of a "hot spot," and rational recombination, a triple mutant (P395D/L241V/T343A) was obtained that showed a 100-fold increase in activity at pH 8 (k(cat) = 100 s(-1)). The increased K(m) values for Br(-) (3.1 mm) and H(2)O(2) (16 microm) are smaller than those found for vanadium bromoperoxidases that are reasonably active at this pH. In addition the brominating activity at pH 5 was increased by a factor of 6 (k(cat) = 575 s(-1)), and the chlorinating activity at pH 5 was increased by a factor of 2 (k(cat) = 36 s(-1)), yielding the "best" vanadium haloperoxidase known thus far. The mutations are in the first and second coordination sphere of the vanadate cofactor, and the catalytic effects suggest that fine tuning of residues Lys-353 and Phe-397, along with addition of negative charge or removal of positive charge near one of the vanadate oxygens, is very important. Lys-353 and Phe-397 were previously assigned to be essential in peroxide activation and halide binding. Analysis of the catalytic parameters of the mutant vanadium bromoperoxidase from the seaweed Ascophyllum nodosum also adds fuel to the discussion regarding factors governing the halide specificity of vanadium haloperoxidases. This study presents the first example of directed evolution of a vanadium enzyme.  相似文献   

18.
In the present study, we investigated how cytochrome c catalyzed the nitration of tyrosine at various pHs. The cytochrome c-catalyzed nitration of tyrosine occurred in proportion to the concentration of hydrogen peroxide, nitrite or cytochrome c. The cytochromec-catalyzed nitration of tyrosine was inhibited by catalase, sodium azide, cystein, and uric acid. These results show that the cytochrome c-catalyzed nitrotyrosine formation was due to peroxidase activity. The rate constant between cytochrome c and hydrogen peroxide within the pH range of 3-8 was the largest at pH 6 (37 degrees C). The amount of nitrotyrosine formed was the greatest at pH 5. At pH 3, only cytochromec-independent nitration of tyrosine occurred in the presence of nitrite. At this pH, the UV as well as visible spectrum of cytochrome c was changed by nitrite, even in the presence of hydrogen peroxide, probably via the formation of a heme iron-nitric oxide complex. Due to this change, the peroxidase activity of cytochrome c was lost.  相似文献   

19.
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   

20.
The reaction of nitrite (NO2-) with horseradish peroxidase and lactoperoxidase was studied. Sequential mixing stopped-flow measurements gave the following values for the rate constants of the reaction of nitrite with compounds II (oxoferryl heme intermediates) of horseradish peroxidase and lactoperoxidase at pH 7.0, 13.3 +/- 0.07 mol(-1) dm3 s(-1) and 3.5 +/- 0.05 x 10(4) mol(-1) dm3 s(-1), respectively. Nitrite, at neutral pH, influenced measurements of activity of lactoperoxidase with typical substrates like 2,2'-azino-bis[ethyl-benzothiazoline-(6)-sulphonic acid] (ABTS), guaiacol or thiocyanate (SCN-). The rate of ABTS and guaiacol oxidation increased linearly with nitrite concentration up to 2.5-5 mmol dm(-3). On the other hand, two-electron SCN- oxidation was inhibited in the presence of nitrite. Thus, nitrite competed with the investigated substrates of lactoperoxidase. The intermediate, most probably nitrogen dioxide (*NO2), reacted more rapidly with ABTS or guaiacol than did lactoperoxidase compound II. It did not, however, effectively oxidize SCN- to OSCN-. NO2- did not influence the activity measurements of horseradish peroxidase by ABTS or guaiacol method.  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