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.
U Pérez  H B Dunford 《Biochemistry》1990,29(11):2757-2763
Transient-state reactions of horseradish peroxidase compounds I and II with 1-methoxy-4-(methylthio)benzene (a para-substituted thioanisole) were studied over the pH range from 3.4 to 10.5. The pH-jump technique was applied to the compound II reactions at pH values below 8.6. The reactions of both compound I and compound II with the para-substituted thioanisole consisted predominantly of an initial burst. The burst was followed by a steady-state phase that became more obvious at lower concentrations of the thioanisoles. The burst phase for both compounds I and II can be explained in terms of two independent transient-state reactions with 1-methoxy-4-(methylthio)benzene as follows: (i) a single reaction of compound I (or compound II) with the substrate and (ii) the formation of a complex between compound I or II and the substrate followed by reaction of the productive complex with another molecule of sulfide. The overall rate of reaction path ii is faster than that of path i. The preference for path i or ii is highly dependent upon the concentration of sulfide with step ii favored at higher sulfide concentrations. The experimental results obtained on the overall reaction under both pseudo-first-order and single-turnover conditions indicate that compound II reacts competitively with both the organic sulfide substrate and the sulfur cation radical produced from compound I oxidation of sulfide.  相似文献   

2.
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   

3.
The oxidation of serotonin (5-hydroxytryptamine) by the myeloperoxidase intermediates compounds I and II was investigated by using transient-state spectral and kinetic measurements at 25.0 +/- 0.1 degrees C. Rapid scan spectra demonstrated that both compound I and compound II oxidize serotonin via one-electron processes. Rate constants for these reactions were determined using both sequential-mixing and single-mixing stopped-flow techniques. The second order rate constant obtained for the one-electron reduction of compound I to compound II by serotonin is (1.7 +/- 0.1) x 10(7) M(-1) x s(-1), and that for compound II reduction to native enzyme is (1.4 +/- 0.1) x 10(6) M(-1) x s(-1) at pH 7.0. The maximum pH of the compound I reaction with serotonin occurs in the pH range 7.0-7.5. At neutral pH, the rate constant for myeloperoxidase compound I reacting with serotonin is an order of magnitude larger than for its reaction with chloride, (2.2 +/- 0.2) x 10(6) M(-1) x s(-1). A direct competition of serotonin with chloride for myeloperoxidase compound I oxidation was observed. Our results suggest that serotonin may have a role to protect lipoproteins from oxidation and to prevent enzymes from inactivation caused by the potent oxidants HOCl and active oxygen species.  相似文献   

4.
Stoichiometry of the reaction between horseradish peroxidase and p-cresol.   总被引:4,自引:0,他引:4  
Over a wide range of pH horseradish peroxidase compound I can be reduced quantitatively via compound II to the native enzyme by only 1 molar equivalent of p-cresol. Since 2 molar equivalents of electrons are required for the single turnover of the enzymatic cycle, p-cresol behaves as a 2-electron reductant. With p-cresol and compound I in a 1:1 ratio compound II and p-methylphenoxy radicals are obtained in the transient state. Compound II is then reduced to the native enzyme. A possible explanation for the facile reduction of compound II involves reaction with the dimerization product of these radicals, 1/2 molar equivalent of 2,2'-dihydroxy-5,5'-dimethylbiphenyl. If only 1/2 molar equivalent of p-cresol is present, than at high pH the reduction stops at compound II. The major steady state peroxidase oxidation product of p-cresol (with p-cresol in large excess compared to the enzyme concentration) is Pummerer's ketone. Pummerer's ketone is only reactive at pH values greater than about 9 where significant amounts of the enol can be formed via the enolate anion. Therefore, in alkaline solution it is reactive with compound I, but not with compound II, which is converted into an unreactive basic form. These results indicate that Pummerer's ketone cannot be the intermediate free radical product responsible for reducing compound II in the single turnover experiments. It is postulated that Pummerer's ketone is formed only in the steady state by the reaction of the p-methylphenoxy radical with excess p-cresol.  相似文献   

5.
Faller P  Rutherford AW  Debus RJ 《Biochemistry》2002,41(43):12914-12920
Photosystem II (PSII) contains two redox-active tyrosines (TyrZ and TyrD) possessing very different properties. TyrD is stable in its oxidized form, while TyrZ is kinetically competent in the water oxidizing reaction. The temperature dependence of the formation of light-induced tyrosyl radical was studied as a function of pH in Mn-depleted PS II from cyanobacteria and plants. Tyrosyl radical formation was observed in the majority of the centers at pH 8.5 and 15 K. The use of two Synechocystis mutants where either TyrZ or TyrD was replaced by a phenylalanine residue allowed a clear demonstration that only TyrD and not TyrZ was oxidized under these conditions. By lowering the pH, the fraction of centers undergoing TyrD oxidation greatly diminished. This pH dependence could be fitted by assuming a single protonable group with a pK(a) of approximately 7.6. This pH dependence correlates with the unexpectedly rapid oxidation of TyrD at room temperature [Faller, P., Debus, R. J., Brettel, K., Sugiura, M., Rutherford, A. W., and Boussac, A. (2001) Proc. Nat. Acad. Sci. U.S.A. 98, 14368]. The equally unexpectedly ability to undergo oxidation at cryogenic temperatures reported here, puts limitations on the nature of the protonation reactions associated with electron transfer in this system, raising the possibility of proton tunneling along the hydrogen bond or alternatively the presence of deprotonated TyrD (tyrosinate) prior to oxidation.  相似文献   

6.
We reported evidence that horseradish peroxidase (HRP) and chloroperoxidase (CPO) catalyze oxygen transfer from H2O2 to thioanisoles [Kobayashi, S., Nakano, M., Goto, T., Kimura, T., & Schaap, A. P. (1986) Biochem. Biophys. Res. Commun. 135, 166-171]. In the present paper, the reaction mechanism of this oxygen transfer is discussed. The oxidation of para-substituted thioanisoles by HRP compound II showed a large negative rho value of -1.46 vs. the sigma + parameter in a Hammett plot. These results are in accord with the formation of a cation radical intermediate in the rate-determining step. Hammett treatments for HRP- and CPO-dependent S-oxygenations did not provide unequivocal proofs to judge the reaction mechanism, because of the poor correlations for sigma + and sigma p parameters. Different behavior was found in kinetics and stereoselectivity between the two enzymes. Results in the present study and recent studies strongly suggested the formation of a cation radical intermediate. The oxygen atom would transfer by reaction of compound II and the cation radical intermediate. Although involvement of the cation radical was not confirmed in the CPO system, a similar mechanism was proposed for CPO.  相似文献   

7.
Despite previous detection of hydroxyl radical formation during iron deposition into ferritin, no reports exist in the literature concerning how it might affect ferritin function. In the present study, hydroxyl radical formation during Fe(II) oxidation by apoferritin was found to be contingent on the "ferroxidase" activity (i.e., H subunit composition) exhibited by apoferritin. Hydroxyl radical formation was found to affect both the stoichiometry and kinetics of Fe(II) oxidation by apoferritin. The stoichiometry of Fe(II) oxidation by apoferritin in an unbuffered solution of 50 mM NaCl, pH 7.0, was approximately 3.1 Fe(II)/O(2) at all iron-to-protein ratios tested. The addition of HEPES as an alternate reactant for the hydroxyl radical resulted in a stoichiometry of about 2 Fe(II)/O(2) at all iron-to-protein ratios. HEPES functioned to protect apoferritin from oxidative modification, for its omission from reaction mixtures containing Fe(II) and apoferritin resulted in alterations to the ferritin consistent with oxidative damage. The kinetic parameters for the reaction of recombinant human H apoferritin with Fe(II) in HEPES buffer (100 mM) were: K(m) = 60 microM, k(cat) = 10 s(-1), and k(cat)/K(m) = 1.7 x 10(5) M(-1) x (-1). Collectively, these results contradict the "crystal growth model" for iron deposition into ferritin and, while our data would seem to imply that the ferroxidase activity of ferritin is adequate in facilitating Fe(II) oxidation at all stages of iron deposition into ferritin, it is important to note that these data were obtained in vitro using nonphysiologic conditions. The possibility that these findings may have physiological significance is discussed.  相似文献   

8.
The reaction of dioxygen with the ferrous forms of the cloned cytochrome c peroxidase [CCP(MI)] and mutants of CCP(MI) prepared by site-directed mutagenesis was studied by photolysis of the respective ferrous-CO complexes in the presence of dioxygen. Reaction of ferrous CCP(MI) with dioxygen transiently formed a FeII-O2 complex (bimolecular rate constant = (3.8 +/- 0.3) x 10(4) M-1 s-1 at pH 6.0; 23 degrees C) that reacted further (first-order rate constant = 4 +/- 1 s-1) to form a product with an absorption spectrum and an EPR radical signal at g = 2.00 that were identical to those of compound I formed by the reaction of CCP(MI)III with peroxide. Thus, the product of the reaction of CCP(MI)II with dioxygen retained three of the four oxidizing equivalents of dioxygen. Gel electrophoresis of the CCP(MI)II + dioxygen reaction products showed that covalent dimeric and trimeric forms of CCP(MI) were produced by the reaction of CCP(MI)II with dioxygen. Photolysis of the CCP(MI)II-CO complex in the presence of ferrous cytochrome c prevented the appearance of the cross-linked forms and resulted in the oxidation of 3 mol of cytochrome c/mol of CCP(MI)II-CO added. The results provide evidence that reaction of CCP(MI)II with dioxygen causes transient oxidation of the enzyme by 1 equiv above the normal compound I oxidation state. Mutations that eliminate the broad EPR signal at g = 2.00 characteristic of the compound I radical also prevented the rapid oxidation of the ferrous enzyme by dioxygen.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

9.
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   

10.
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   

11.
A novel anti-5,5-dimethyl-1-pyrroline N-oxide (DMPO) polyclonal antiserum that specifically recognizes protein radical-derived DMPO nitrone adducts has been developed. In this study, we employed this new approach, which combines the specificity of spin trapping and the sensitivity of antigen-antibody interactions, to investigate protein radical formation from lactoperoxidase (LPO). When LPO reacted with GSH in the presence of DMPO, we detected an LPO radical-derived DMPO nitrone adduct using enzyme-linked immunosorbent assay and Western blotting. The formation of this nitrone adduct depended on the concentrations of GSH, LPO, and DMPO as well as pH values, and GSH could not be replaced by H(2)O(2). The level of this nitrone adduct was decreased significantly by azide, catalase, ascorbate, iodide, thiocyanate, phenol, or nitrite. However, its formation was unaffected by chemical modification of free cysteine, tyrosine, and tryptophan residues on LPO. ESR spectra showed that a glutathiyl radical was formed from the LPO/GSH/DMPO system, but no protein radical adduct could be detected by ESR. Its formation was decreased by azide, catalase, ascorbate, iodide, or thiocyanate, whereas phenol or nitrite increased it. GSH caused marked changes in the spectrum of compound II of LPO, indicating that GSH binds to the heme of compound II, whereas phenol or nitrite prevented these changes and reduced compound II back to the native enzyme. GSH also dose-dependently inhibited the peroxidase activity of LPO as determined by measuring 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid) oxidation. Taken together, these results demonstrate that the GSH-dependent LPO radical formation is mediated by the glutathiyl radical, possibly via the reaction of the glutathiyl radical with the heme of compound II to form a heme-centered radical trapped by DMPO.  相似文献   

12.
The reaction of prostaglandin H synthase with prostaglandin G2, the physiological substrate for the peroxidase reaction, was examined by rapid reaction techniques at 1 degree C. Two spectral intermediates were observed and assigned to higher oxidation states of the enzymes. Intermediate I was formed within 20 ms in a bimolecular reaction between the enzyme and prostaglandin G2 with k1 = 1.4 x 10(7) M-1 s-1. From the resemblance to compound I of horseradish peroxidase, the structure of intermediate I was assigned to [(protoporphyrin IX)+.FeIVO]. Between 10 ms and 170 ms intermediate II was formed from intermediate I in a monomolecular reaction with k2 = 65 s-1. Intermediate II, spectrally very similar to compound II of horseradish peroxidase or complex ES of cytochrome-c peroxidase, was assigned to a two-electron oxidized state [(protoporphyrin IX)FeIVO] Tyr+. which was formed by an intramolecular electron transfer from tyrosine to the porphyrin-pi-cation radical of intermediate I. A reaction scheme for prostaglandin H synthase is proposed where the tyrosyl radical of intermediate II activates the cyclooxygenase reaction.  相似文献   

13.
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   

14.
The kinetics and spectra of the reactions of clozapine with compounds I and II of myeloperoxidase were investigated using both single- and sequential-mixing stopped-flow techniques, steady-state kinetics, and spectrophotometric measurements. The results show conclusively that both compounds I and II are reduced in one-electron reactions with clozapine. At pH 7.0 the rate constant for compound I reacting with clozapine is (1.5 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (4.8 +/- 0.1) x 10(4) M(-1) s(-1). The physiological pH of 7.4 was found to be optimal for the oxidation of clozapine by compound I. The rate constant for compound I reacting with ascorbate is (1.1 +/- 0.1) x 10(6) M(-1) s(-1) and for compound II (1.1 +/- 0.2) x 10(4) M(-1) s(-1), both obtained at pH 7.0. Experiments with both clozapine and ascorbate present showed that ascorbate acts both as a competitive inhibitor and free radical scavenger.  相似文献   

15.
Ascorbate peroxidase from L. Major (LmAPX) is a functional hybrid between cytochrome c peroxidase (CCP) and ascorbate peroxidase (APX). We utilized point mutagenesis to investigate if a conserved proximal tryptophan residue (Trp208) among Class I peroxidase helps in controlling catalysis. The mutant W208F enzyme had no effect on both apparent dissociation constant of the enzyme-cytochrome c complex and K(m) value for cytochrome c indicating that cytochrome c binding affinity to the enzyme did not alter after mutation. Surprisingly, the mutant was 1000 times less active than the wild type in cytochrome c oxidation without affecting the second order rate constant of compound I formation. Our diode array stopped-flow spectral studies showed that the substrate unbound wild type enzyme reacts with H(2)O(2) to form compound I (compound II type spectrum), which was quite different from that of compound I in W208F mutant as well as horseradish peroxidase (HRP). The spectrum of the compound I in wild type LmAPX showed a red shift from 409 nm to 420 nm with equal intensity, which was broadly similar to those of known Trp radical. In case of compound I for W208F mutant, the peak in the Soret region was decreased in heme intensity at 409 nm and was not shifted to 420 nm suggesting this type of spectrum was similar to that of the known porphyrin pi-cation radical. In case of an enzyme-H(2)O(2)-ascorbate system, the kinetic for formation and decay of compound I and II of a mutant enzyme was almost identical to that of a wild type enzyme. Thus, the results of cytochrome c binding, compound I formation rate and activity assay suggested that Trp208 in LmAPX was essential for electron transfer from cytochrome c to heme ferryl but was not indispensable for ascorbate or guaiacol oxidation.  相似文献   

16.
The formation of ferryl heme (Fe(IV) = O) species, i.e., compound I and compound II, has been identified as the main intermediates in heme protein peroxidative reactions. We report stopped-flow kinetic measurements which illustrate that the reaction of hemoglobin I (HbI) from Lucina pectinata with hydrogen peroxide produce ferryl intermediates compound I and compound II. Compound I appears relatively stable displaying an absorption at 648 nm. The rate constant value (k'(2)) for the conversion of compound I to compound II is 3.0 x 10(-2) s(-1), more than 100 times smaller than that reported for myoglobin. The rate constant value for the oxidation of the ferric heme (k'(12) + k'(13)) is 2.0 x 10(2) M(-1) s(-1). These values suggest an alternate route for the formation of compound II (by k'(13)) avoiding the step from compound I to compound II (k'(2)). In HbI from L. pectinata the stabilization of compound I is attribute to the unusual collection of amino acids residues (Q64, F29, F43, F68) in the heme pocket active site of the protein.  相似文献   

17.
The hydroxylation of phenylalanine by the Fenton reaction and gamma-radiolysis yields 2-hydroxy-, 3-hydroxy-, and 4-hydroxyphenylalanine (tyrosine), while the hydroxylation of tyrosine results in 2,3- and 3,4-dihydroxyphenylalanine (dopa). Yields are determined as a function of pH and the presence or absence of oxidants. During gamma-radiolysis and the Fenton reaction the same hydroxylated products are formed. The final product distribution depends on the rate of the oxidation of the hydroxyl radical adducts (hydroxycyclohexadiene radicals) relative to the competing dimerization reactions. The pH profiles for the hydroxylations of phenylalanine and tyrosine show a maximum at pH 5.5 and a minimum around pH 8. The lack of hydroxylated products around near pH 8 is due to the rapid oxidation of dopa to melanin. The relative abilities of iron chelates (HLFe(II) and HLFe(III) to promote hydroxyl radical formation from hydrogen peroxide are nitrilotriacetate (nta) greater than ethylenediaminediacetate (edda) much greater than hydroxyethylethylenediaminetriacetate greater than citrate greater than ethylenediaminetetraacetate greater than diethylenetriaminepentaacetate greater than adenosine 5'-triphosphate greater than pyrophosphate greater than adenosine 5'-diphosphate greater than adenosine 5'-monophosphate. The high activity of iron-nta and -edda chelates is explained by postulating the formation of a ternary Fe(III)-L-dopa complex in which dopa reduces Fe(III). The hydroxylations of phenylalanine and tyrosine are similar to that of salicylate (Z. Maskos, J. D. Rush, and W. H. Koppenol, 1990, Free Radical Biol. Med. 8, 153-162) and tryptophan (preceding paper) in that oxidants augment the formation of hydroxylated products by catalyzing the dismutation of hydroxyl radical adducts to the parent compound and a stable hydroxylated product. A comparison of salicylate and the amino acids tryptophan, phenylalanine, and tyrosine clearly shows that salicylate is the best indicator of hydroxyl radical production.  相似文献   

18.
Kinetic parameters for each reaction step of the peroxidase-catalyzed reaction were determined by the stopped-flow technique on three distinct isoenzymes: acidic A2, neutral C1, and basic E5. The pH dependence of the reaction for the formation of compound I with hydrogen peroxide was examined. The three isoenzymes had a common ionizing group at about pK 4 which affects the rate constant for the formation of compound I. The heat of ionization determined from the temperature dependence of the dissociation constant of the group strongly suggested that it is of carboxyl nature. The rate constant for isoenzyme A2 was a tenth of those for the other two isoenzymes over the whole range of pH. Furthermore, the thermodynamic parameters of isoenzyme A2 were found to be different from those for the other two isoenzymes. These difference as well as the different behavior in alkaline transition of the isoenzymes are discussed in relation to the sixth ligand of the heme. The rate constant of the reduction of compound I and compound II by ferrocyanide were also determined. In both reduction steps, the pH profiles of the apparent rate constant for isoenzyme A2 and E5 were similar, but they were different from that of C1. The ionization with pK 5.29, which was detected only in isoenzyme C1, may be attributed to a group near the porphyrin ring as a stabilizer for the pi-cation radical.  相似文献   

19.
The use of antioxidants to prevent intracellular free radical damage is an area currently attracting considerable research interest. The compound 2',7'-dichlorofluorescin diacetate (DCFH-DA) is a probe for intracellular peroxide formation commonly used in such studies. During our studies we unexpectedly found that incubation of Trolox, a water soluble vitamin E analog, with DCFH-DA in cell-free physiological buffers resulted in the deacetylation and oxidation of DCFH-DA to form the fluorescent compound, 2',7'-dichlorofluororescein (DCF). The reaction was time-, temperature-, and pH-dependent. Fluorescence intensity increased with an increase in either Trolox or DCFH-DA concentration. These results indicate that even at physiological pH, DCFH-DA can be deacetylated to form 2',7'-dichlorofluorescin (DCFH). DCFH can then be oxidized to DCF by abstraction of a hydrogen atom by the phenoxyl radical of Trolox. Exposure of the reaction mixture to 10 Gy of 60Co gamma radiation greatly increased production of DCF. Antioxidant compounds reported to “repair” the Trolox phenoxyl radical (e.g., ascorbic acid, salicylate) can also prevent the Trolox-induced DCFH-DA fluorescence. However, compounds that cannot repair the Trolox phenoxyl radical (e.g., catechin) or can themselves form a radical (e.g., uric acid, TEMPOL) either have no effect or can increase levels of DCF. These results demonstrate that experimental design must be carefully considered when using DCFH-DA to measure peroxide formation in combination with certain antioxidants.  相似文献   

20.
In the reaction between equimolar amounts of horseradish peroxidase and chlorite, the native enzyme is oxidized directly to Compound II (Hewson, W.D., and Hager, L.P. (1979) J. Biol. Chem. 254, 3175-3181). At acidic pH but not at alkaline values, this initial reaction is followed by oxidation of Compound II to Compound I. The highly pH-dependent chemistry of Compound II can be readily demonstrated by the reduction of Compound I, with ferrocyanide at acidic, neutral, and alkaline pH values. Titration at low pH yields very little Compound II, whereas at high pH, the yield is quantitative. Similarly, the reaction of horseradish peroxidase and chlorite at low pH yields Compound I while only Compound II is formed at high pH. At intermediate pH values both the ferrocyanide reduction and the chlorite reaction produce intermediate yields of Compound II. This behavior is explained in terms of acidic and basic forms of Compound II. The acidic form is reactive and unstable relative to the basic form. Compound II can be readily oxidized to Compound I by either chloride or chlorine dioxide in acidic solution. The oxidation does not occur in alkaline solution, nor will hydrogen peroxide cause the oxidation of Compound II, even at low pH.  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