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 93 毫秒
1.
It has been found that superoxide radicals formed at autooxidation of Fe(II) ions or of their complexes with EDTA in the presence of NAD.H cause reduction of met-Hb into deoxy or oxy-form. Under the effect of hydroxyl radicals generated in the ultrasonic field in the presence of NADH reduction of met-Hb into the ferro-form (carboxy-Hb) was observed in Co atmosphere and not in O2 atmosphere. It was induced by a high oxidation rate of oxy-Hb by hydroxyl radicals into met-Hb as compared to carboxy-Hb. Reduction of met-Hb was shown to be accompanied with the formation of NAD+. The role of NAD. radicals in the reduction of met-Hb which acts as an antioxidant in respect to organic free radicals was discussed.  相似文献   

2.
Direct and spin-trapping electron spin resonance methods have been used to study the reactivity of semiquinone radicals from the anthracycline antibiotics daunorubicin and adriamycin towards peroxides (hydrogen peroxide, t-butyl hydroperoxide and cumene hydroperoxide). Semiquinone radicals were generated by one-electron reduction of anthracyclines, using xanthine/xanthine oxidase. It is shown that the semiquinones are effective reducing agents for all the peroxides. From spin-trapping experiments it is inferred that the radical product is either OH (from H2O2) or an alkoxyl radical (from the hydroperoxides) which undergoes beta-scission to give the methyl radical. The rate constant for reaction of semiquinone with H2O2 is estimated to be approx. 10(4)-10(5) M-1 X s-1. The reduction does not appear to require catalysis by metal ions.  相似文献   

3.
The reduction of D-amino acid oxidase (DAAO) by hydrated electrons (eaq-) has been studied in the absence and presence of benzoate by pulse radiolysis. The eaq-did not reduce the flavin moiety in DAAO and reacted with the amino acid residues in the protein. In the presence of benzoate, eaq- first reacted with benzoate to yield benzoate anion radical. Subsequently, the benzoate anion radical transferred an electron to the complex of DAAO-benzoate to form the red semiquinone of the enzyme with a second-order rate constant of 1.2 X 10(9) M-1 s-1 at pH 8.3. After the first phase of the reduction, conversion of the red semiquinone to the blue semiquinone was observed in the presence of high concentration of benzoate. This process obeyed first-order kinetics, and the rate increased with an increase of the concentration of benzoate. In addition, the rate was found to be identical with that of the formation of the complex between benzoate and the red semiquinone of DAAO as measured by a stopped-flow method. This suggests that bound benzoate dissociates after the reduction of the benzoate-DAAO complex by benzoate anion radical and that free benzoate subsequently recombines with the red semiquinone of the enzyme to form the blue semiquinone.  相似文献   

4.
Daunorubicin, an anthracycline antitumor antibiotic, can be complexed in egg white apo-riboflavin binding protein. The reduction of this complex was studied by gamma-radiolysis and pulse radiolysis using COO.- free radicals as reductants. The final products are 7-deoxydaunomycinone intercalated in the protein and thiol groups coming from the reduction of disulfide bonds of the protein, in the respective proportions of 90% and 10%. One-electron reduction of the complex gives daunorubicin semiquinone radical and a disulfide anion. The rate constants of the reactions of COO.- ions with the complex and with the disulfide bond in the protein alone are respectively equal to 2.4 x 10(8) mol-1.L.s-1 and 6.4 x 10(7) mol-1.L.s-1. Daunorubicin semiquinone decays by a first-order process, the rate constant of which is independent of the initial protein and radical concentrations. Without protein, daunorubicin semiquinone undergoes a disproportionation-comproportionation equilibrium [Houée-Levin, C., Gardès-Albert, M., Ferradini, C., Faraggi, M., & Klapper, M. (1985) FEBS Lett. 179, 46-50]. We propose that a protein residue reduces semiquinone by an intramolecular path. This creates an electron hole in the protein which may alter its function. This reduction process is very different from the reduction mechanism of riboflavin binding protein by the same reductant [Faraggi, M., Steiner, J.P., & Klapper, M.H. (1985) Biochemistry 24, 3273-3279]. These results suggest a new deleterious pathway to explain the antitumor and/or cytotoxic effect of this drug.  相似文献   

5.
The oxidation-reduction potential values for the two electron transfers to glucose oxidase were obtained at pH 5.3, where the neutral radical is the stable form, and at pH 9.3, where the anion radical is the stable form. The midpoint potentials at 25 degrees were: pH 5.3 EFl1ox + e- H+ equilibrium EFlH. Em1 = -0.063 +/- 0.011 V EFlH. + e- + H+ equilibrium EFlredH2 Em2 = -0.065 +/- 0.007 V pH 9.3 EFlox + e- EFi- Em1 = -0.200 +/- 0.010 V EFi- + e- + H+ equilibrium EFlredH- Em2 = -0.240 +/- 0.005 V All potentials were measured versus the standard hydrogen electrode (SHE). The potentials indicated that glucose oxidase radicals are stabilized by kinetic factors and not by thermodynamic energy barriers. The pK for the glucose oxidase radical was 7.28 from dead time stopped flow measurements and the extinction coefficient of the neutral semiquinone was 4140 M-1 cm-1 at 570 nm. Both radical forms reacted with oxygen in a second order fashion. The rate at 25 degrees for the neutral semiquinone was 1.4 X 10(4) M-1 s-1; that for the anion radical was 3.5 X 10(4) M-1 s-1. The rate of oxidation of the neutral radical changed by a factor of 9 for a temperature difference of 22 degrees. For the anion radical, the oxidation rate changed by a factor of 6 for a 22 degrees change in temperature. We studied the oxygen reactivity of the 2-electron reduced form of the enzyme over a wide wavelength range and failed to detect either oxygenated flavin derivatives or semiquinoid forms as intermediates. The rate of reoxidation of fully reduced glucose oxidase at pH 9.3 was dependent on ionic strength.  相似文献   

6.
Reduced "leuco" dyes such as dichlorodihydrofluorescein (DCFH(2)) are widely used as profluorescent probes for oxidative stress, although they require a catalyst to be oxidized by hydrogen peroxide and react indiscriminately with oxidizing radicals and the fluorescent product (DCF) is a potential photosensitizer of superoxide generation. In this study, key properties of the radical intermediate in oxidation ("semiquinone," DCFH(.-)/DCF(.)(-)) were measured, to help understand the reactions that can occur in biological systems. The intermediate was generated by oxidizing DCFH(2) or reducing DCF by radiolytically generated radicals and monitoring the reactions using kinetic spectrophotometry. The semiquinone showed pH-sensitive absorption spectral changes, decay kinetics (both in the absence and in the presence of oxygen), and reduction potential, all corresponding to prototropic dissociations with pK(a)'s of approximately 7.1 and 9.0. DCFH(2) has pK(a)'s in a similar region (8-9) and hence pH variations are potentially important in the use of this probe. The rate constant for reaction of the semiquinone with oxygen at pH 7.4 is 5.3 x 10(8) M(-1) s(-1): this reaction, rather than disproportionation of DCFH(.-)/DCF(.)(-), generates DCF in biological systems, concomitantly forming superoxide and hence H(2)O(2) to cycle the catalyst. The midpoint reduction potential of the couple DCF,H(+)/DCFH() is approximately -0.75 V vs. NHE at pH 7.4; DCF is unlikely to be reduced rapidly by common flavoprotein reductases.  相似文献   

7.
Kinetic analysis and mechanistic aspects of autoxidation of catechins   总被引:3,自引:0,他引:3  
A peroxidase-based bioelectrochemical sensor of hydrogen peroxide (H(2)O(2)) and a Clark-type oxygen electrode were applied to continuous monitoring and kinetic analysis of the autoxidation of catechins. Four major catechins in green tea, (-)-epicatechin, (-)-epicatechin gallate, (-)-epigallocatechin, and (-)-epigallocatechin gallate, were used as model compounds. It was found that dioxygen (O(2)) is quantitatively reduced to H(2)O(2). The initial rate of autoxidation is suppressed by superoxide dismutase and H(+), but is independent of buffer capacity. Based on these results, a mechanism of autoxidation is proposed; the initial step is the one-electron oxidation of the B ring of catechins by O(2) to generate a superoxide anion (O(2)(*-)) and a semiquinone radical, as supported in part by electron spin resonance measurements. O(2)(*-) works as a stronger one-electron oxidant than O(2) against catechins and is reduced to H(2)O(2). The semiquinone radical is more susceptible to oxidation with O(2) than fully reduced catechins. The autoxidation rate increases with pH. This behavior can be interpreted in terms of the increase in the stability of O(2)(*-) and the semiquinone radical with increasing pH, rather than the acid dissociation of phenolic groups. Cupric ion enhances autoxidation; most probably it functions as a catalyst of the initial oxidation step of catechins. The product cuprous ion can trigger a Fenton reaction to generate hydroxyl radical. On the other hand, borate ion suppresses autoxidation drastically, due to the strong complex formation with catechins. The biological significance of autoxidation and its effectors are also discussed.  相似文献   

8.
The reduction of flavin in hepatic NADH-cytochrome b5 reductase by the hydrated electron (eaq-) was investigated by pulse radiolysis. The eaq- reduced the flavin of NADH-cytochrome b5 reductase to form the red semiquinone between pH 5 and 9. The spectrum of the red semiquinone differs from that of enzyme reduced by dithionite in the presence of NAD+. After the first phase of the reduction, conversion of the red to blue semiquinone was observed at acidic pH. Resulting products are the blue (neutral) or red (anionic) semiquinone or a mixture of the two forms. The pK value for this flavin radical was approximately 6.3. Subsequently, the semiquinone form reacted by dismutation to form the oxidized and the fully reduced forms of the enzyme with a rate constant of 1 x 10(3) M-1 s-1 at pH 7.1. In the presence of NAD+, eaq- reacted with NAD+ to yield NAD(.). Subsequently, NAD. transferred an electron to NAD+-bound oxidized enzyme to form the blue and red semiquinone or mixture of the two forms of the enzyme, where pK value of this flavin radical was approximately 6.3. The blue semiquinone obtained at acidic pH was found to convert to the red semiquinone with a first order rate constant of 90 s-1, where the rates were not affected by pH or the concentration of NAD+. The final product is NAD+-bound red semiquinone of the enzyme.  相似文献   

9.
A model (NADH-phenazine methosulfate-O2) formally similar to pyridine nucleotide-dependent flavoprotein hydroxylases catalyzed the hydroxylation of several aromatic compounds. The hydroxylation was maximal at acid pH and was inhibited by ovine Superoxide dismutase, suggesting that perhydroxyl radicals might be intermediates in this process. The stoichiometry of the reaction indicated that a univalent reduction of oxygen was occurring. The correlation between the concentration of semiquinone and hydroxylation, and the inhibition of hydroxylation by ethanol which inhibited semiquinone oxidation, suggested the involvement of phenazine methosulfate-semiquinone. Activation of hydroxylation by Fe3+ and Cu2+ supported the contention that univalently reduced species of oxygen was involved in hydroxylation. Catalase was without effect on the hydroxylation by the model, ruling out H2O2 as an intermediate. A reaction sequence, involving a two-electron reduction of phenazine methosulfate to reduced phenazine methosulfate followed by disproportionation with phenazine methosulfate to generate the semiquinone, was proposed. The semiquinone could donate an electron to O2 to generate O2 which could be subsequently protonated to form the perhydroxyl radical.  相似文献   

10.
Flavocytochrome P450 BM3 FMN domain is unique among the family of flavodoxins and homologues, in not forming a stable neutral blue FMN semiquinone radical. Anaerobic, one-electron reduction of the isolated domain over the pH 7-9.5 range showed that it forms an anionic red semiquinone that disproportionates slowly (0.014s(-1) at pH 7). The rate of disproportionation decreased at higher pH, indicating that protonation of the anionic semiquinone is an important feature of the mechanism. The reduction potential for the oxidised-semiquinone couple was determined to be -240mV and was largely independent of pH. The semiquinone appears, therefore, to be kinetically trapped by a slow protonation event, enabling it to act as a low-potential electron donor to the P450 heme.  相似文献   

11.
Anaerobic reduction of hydrogen peroxide in a xanthine/xanthine oxidase system by adriamycin semiquinone in the presence of chelators and radical scavengers was investigated by direct electron paramagnetic resonance and spin trapping techniques. Under these conditions, adriamycin semiquinone appears to react with hydrogen peroxide forming the hydroxyl radical in the presence of chelators such as ethylenediaminetetraacetic acid and diethylenetriaminepentaacetic acid. In the absence of chelators, a related, but unknown oxidant is formed. In the presence of desferrioxamine, adriamycin semiquinone does not disappear in the presence of hydrogen peroxide at a detectable rate. The presence of adventitious iron is therefore implicated during adriamycin semiquinone-catalyzed reduction of hydrogen peroxide. Formation of alpha-hydroxyethyl radical and carbon dioxide radical anion from ethanol and formate, respectively, was detected by spin trapping. Both the hydroxyl radical and the related oxidant react with these scavengers, forming the corresponding radical. In the presence of scavengers from which reducing radicals are formed, the rate of consumption of hydrogen peroxide in this system is increased. This result can be explained by a radical-driven Fenton reaction.  相似文献   

12.
The formation of semiquinone free radicals from antitumor drugs has been studied by pulse radiolysis. The semiquinone free radicals are reactive and have short half-lives in aqueous media under anaerobic conditions. The half-lives of the radicals formed from adriamycin, mitomycin C, and 2,5-diaziridinyl-3,6-bis(carboethoxy)amine-1,4-benzoquinone (AZQ) are 50,100, and 200 μs, respectively. The mean diffusion distance of the semiquinone free radical is less than 0.6 μm. In the presence of molecular oxygen the half-life of the semiquinone free radical is shortened. Adriamycin semiquinone reacts rapidly with oxygen, k = 4.4 × 107m?1s?1. In air-saturated buffer the half-life of adriamycin semiquinone radical can be calculated to be 8 μs with a mean diffusion distance of less than 0.1 μm. If the half-lives in buffer are comparable to those within a cell, semiquinone free radicals must be generated close to the site at which they produce a biological effect. One-electron reduction potentials (E71) were determined and were AZQ, ?168 mV, adrenochrome, ?253 mV, mitomycin C, ?271 mV, adriamycin, ?292 mV, daunomycin, ?305 mV, and anthracenedione, ?348 mV. Enzymatic one-electron reduction of these antitumor quinones by NADPH-cytochrome P-450 reductase increased at more positive values of quinone E71.  相似文献   

13.
p-Cresol methylhydroxylase, a heterodimer consisting of one flavoprotein subunit and one cytochrome c subunit, may be resolved into its subunits, and the holoenzyme may then be fully reconstituted from the pure subunits. In the present study we have characterized the reduction kinetics of the intact enzyme and its subunits, by using exogenous 5-deazariboflavin semiquinone radical generated in the presence of EDTA by the laser-flash-photolysis technique. Under anaerobic conditions the 5-deazariboflavin semiquinone radical reacts rapidly with the native enzyme with a rate constant approaching that of a diffusion-controlled reaction (k = 2.8 X 10(9) M-1 X s-1). Time-resolved difference spectra at pH 7.6 indicate that both flavin and haem are reduced initially by the deazariboflavin semiquinone radical, followed by an additional slower intramolecular electron transfer (k = 220 s-1) from the endogenous neutral flavin semiquinone radical to the oxidized haem moiety of the native enzyme. During the steady-state photochemical titration of the native enzyme at pH 7.6 with deazariboflavin semiquinone radical generated by light-irradiation the haem appeared to be reduced before the protein-bound flavin and was followed by the formation of the protein-bound anionic flavin radical. This result suggests that the redox potential of the haem is higher than that of the flavin, and that deprotonation of the flavin neutral radical occurred during the photochemical titration. Reduction kinetics of the flavoprotein and cytochrome subunits were also investigated by laser-flash photolysis. The protein-bound flavin of the isolated flavin subunit was reduced rapidly by the deazariboflavin semiquinone radical (k = 2.2 X 10(9) M-1 X s-1), as was the haem of the pure cytochrome c subunit (k = 3.7 X 10(9) M-1 X s-1). Flash-induced difference spectra obtained for the flavoprotein and cytochrome subunits at pH 7.6 were consistent with the formation of neutral flavin semiquinone radical and reduced haem, respectively. Investigation of the kinetic properties of the neutral flavin semiquinone radical of the flavoprotein subunit at pH 7.6 and at longer times (up to 5s) were consistent with a slow first-order deprotonation reaction (k = 1 s-1) of the neutral radical to its anionic form.  相似文献   

14.
Hellwig P  Yano T  Ohnishi T  Gennis RB 《Biochemistry》2002,41(34):10675-10679
During turnover of cytochrome bo(3) from Escherichia coli, a semiquinone radical is stabilized in a high-affinity binding site. To identify binding partners of this radical, site-directed mutants have been designed on the basis of a recently modeled quinone binding site (Abramson et al., 2000). The R71H, H98F, D75H, and I102W mutant enzymes were found to show very little or no quinol oxidase activity. The thermodynamic and EPR spectroscopic properties of semiquinone radicals in these mutants were characterized. For the H98F and the R71H mutants, no EPR signal of the semiquinone radical was observed in the redox potential range from -100 to 250 mV. During potentiometric titration of the D75H mutant enzyme, a semiquinone signal was detected in the same potential range as that of the wild-type enzyme. However, the EPR spectrum of the D75H mutant lacks the characteristic hyperfine structure of the semiquinone radical signal observed in the wild-type oxidase, indicating that D75 or the introduced His, interacts with the semiquinone radical. For the I102W mutant, a free radical signal was observed with a redox midpoint potential downshifted by about 200 mV. On the basis of these observations, it is suggested that R71, D75, and H98 residues are involved in the stabilization of the semiquinone state in the high-affinity binding site. Details of the possible binding motif and mechanistic implications are discussed.  相似文献   

15.
One-electron oxidation of dopamine by ferricyanide generates a highly reactive free radical intermediate that inactivates the V-type H(+)-ATPase proton pump in catecholamine storage vesicles, i.e., the driving force in both the vesicular uptake and the storage of catecholamines, in a cell-free in vitro model system at pH 7.0. Electron paramagnetic resonance spectroscopy revealed that a radical with g=2.0045, formed by this oxidation, was relatively long-lived (t(1/2) obs=79 s at pH 6.5 and 25 degrees C). Experimental evidence is presented that the observed radical most likely represents dopamine semiquinone free radical, although an o-quinone free radical cannot be ruled out. Oxidation of noradrenaline and adrenaline by ferricyanide generated similar isotropic radicals, but of shorter half-lives (i.e., 43 and 5.3 s, respectively), and the efficacy of inactivation of the H(+)-ATPase correlated with the half-life of the respective catecholamine free radical (i.e., dopamine >noradrenaline>adrenaline). Thus, the generation of relatively long-lived semiquinone free radicals, although at low concentrations, in dopaminergic and noradrenergic neurons may represent a common mechanism of cytotoxicity linked to neurodegeneration of the respective neurons related to Parkinson disease.  相似文献   

16.
R.W. Miller  F.D.H. Macdowall 《BBA》1975,387(1):176-187
Wheat chloroplasts photochemically reduced molecular oxygen, as a Hill oxidant in the Mehler reaction, to superoxide anion which then oxidized added 1,2-dihydroxybenzene-3,5-disulfonate to its semiquinone, a comparatively stable free radical at pH 7. The last mentioned reaction was rapid in aqueous solution, but the rate of formation of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone by the chloroplast system was calculated as a T12 of 0.6 s. The Mehler reaction, or more specifically the univalent reduction of oxygen by Photosystem I, was rate-limiting so that the 1,2-dihydroxybenzene-3,5-disulfonate semiquinone was a useful spin probe for superoxide anion production at room temperature. The ESR signal of 1,2-dihydroxybenzene-3,5-disulfonate semiquinone was proportional to its steady state concentration and decayed in the dark with a T12 of 5–6 s. This oxygen-dependent signal was enhanced by mediation of chloroplastic oxygen reduction through methyl viologen. The superoxide anion scavengers ascorbate and l-epinephrine competitively obscured 1,2-dihydroxybenzene-3,5-disulfonate semiquinone formation, but added superoxide dismutase was not as effective in this role. Partial inhibition by superoxide dismutase was achieved only by preincubation of Photosystem I enriched particles with ten times the endogenous concentration of superoxide dismutase. This and the persistence of a small amount of a 1,2-dihydroxybenzene-3,5-disulfonate (Tiron) oxidizing species in the dark supports the concept of Tiron accessibility but not the superoxide dismutase accessibility of superoxide anion bound in its formative enzyme complex. Benzoquinone and naphthoquinone disulfonate also reacted with superoxide anion, and supported both the Hill reaction and the Mehler reaction as final oxidants of both water and superoxide anion.  相似文献   

17.
The source of superoxide anion radical (O2-.) in aerobic mixtures consisting of NAD[P]H, 5-methylphenazinium methyl sulfate (or its 1-methoxy derivative) and tetrazolium salt was investigated using superoxide dismutase (SOD), Mn(II), ferricytochrome-C, and epinephrine as probes. NAD[P]H + phenazine + O2 was found to reduce nitroblue tetrazolium, iodonitrotetrazolium, and thiazolyl blue in a manner sensitive to agents that dismutase O2-., viz., SOD and Mn(II). It also mediated the reduction of ferricytochrome-C, and augmented the autooxidation of epinephrine to the adrenochrome, without a tetrazolium salt present in the medium. The autooxidation of epinephrine, but not the reduction of ferricytochrome-C, was found to be sensitive to SOD. Nitroblue tetrazolium, either singly or in combination with SOD, did not stimulate the reduction of ferricytochrome-C. The oxidation of NADH, mediated by a catalytically low concentration of phenazine(+O2), was augmented two-fold by SOD. These observations are consistent with, and lend support to, a scheme of redox events (Scheme-3) wherein it is proposed that the source of O2-. in the NAD[P]H + phenazine + tetrazolium(+O2) system is the reduced phenazine, that the tetrazoinyl radical (a one-electron reduction product of tetrazolium) may not reduce O2 to O2-., that the redox reaction between semiquinone radicals of phenazine and O2 is reversible, and that the disproportionation of semiquinone radicals constitutes an important rate-limiting reaction in the expression of phenazine redox couple.  相似文献   

18.
We have analyzed the mechanism of one-electron reduction of adriamycin (Adr) using recombinant full-length human neuronal nitric-oxide synthase and its flavin domains. Both enzymes catalyzed aerobic NADPH oxidation in the presence of Adr. Calcium/calmodulin (Ca(2+)/CaM) stimulated the NADPH oxidation of Adr. In the presence or absence of Ca(2+)/CaM, the flavin semiquinone radical species were major intermediates observed during the oxidation of the reduced enzyme by Adr. The FAD-NADPH binding domain did not significantly catalyze the reduction of Adr. Neither the FAD semiquinone (FADH*) nor the air-stable semiquinone (FAD-FMNH*) reacted rapidly with Adr. These data indicate that the fully reduced species of FMN (FMNH(2)) donates one electron to Adr, and that the rate of Adr reduction is stimulated by a rapid electron exchange between the two flavins in the presence of Ca(2+)/CaM. Based on these findings, we propose a role for the FAD-FMN pair in the one-electron reduction of Adr.  相似文献   

19.
During autoxidation of 1,4-hydroquinone (H2Q, less than 1 mM) at pH 7.4 and 37 degrees C, stoichiometric amounts of 1,4-benzoquinone (Q) and hydrogen peroxide were formed during the initial reaction. The reaction kinetics showed a significant induction period which was abolished by minute amounts of Q. Hydrogen peroxide and catalase were without effect on the autoxidation process. Transition metals apparently were not involved, since chelators like EDTA, DETAPAC, and desferrioxamine or FeSO4 had no influence on the autoxidation kinetics. Superoxide dismutase (SOD) did not abolish the induction period but dramatically enhanced the autoxidation rate by more than two orders of magnitude. The stimulatory effect was first-order in SOD concentration but showed saturation kinetics. The dependence of Q and hydrogen peroxide formation rates on H2Q concentration shows a biphasic behaviour: dependence on the square at low H2Q, but on the square root at high H2Q concentration. As revealed by calculatory simulations the results can be adequately described by the known reaction rate constants. The reaction starts with the comproportionation of H2Q and Q to yield two semiquinone molecules which autoxidize to give two superoxide radicals and two molecules of Q which enter into a new cycle of comproportionation. Because of unfavourable equilibria the autocatalytic reaction soon comes to steady state, and the further reaction is governed by the rate of superoxide removal. At excess SOD, the comproportionation reaction is rate-limiting, thus explaining the saturation effects of SOD. The experiments do not allow a decision between the two functions of SOD; the conventional action as a superoxide:superoxide oxidoreductase or as a semiquinone:superoxide oxidoreductase. In the latter reaction SOD is thought to be reduced by semiquinone with Q formation. In the second step the reduced enzyme would be re-oxidized by a superoxide radical which is formed during autoxidation of the second semiquinone molecule generated in the comproportionation reaction. From thermodynamic considerations, the latter function of SOD appears to be plausible.  相似文献   

20.
The redox cycle of 2,5-dimethoxybenzoquinone (2,5-DMBQ) is proposed as a source of reducing equivalent for the regeneration of Fe2+ and H2O2 in brown rot fungal decay of wood. Oxalate has also been proposed to be the physiological iron reductant. We characterized the effect of pH and oxalate on the 2,5-DMBQ-driven Fenton chemistry and on Fe3+ reduction and oxidation. Hydroxyl radical formation was assessed by lipid peroxidation. We found that hydroquinone (2,5-DMHQ) is very stable in the absence of iron at pH 2 to 4, the pH of degraded wood. 2,5-DMHQ readily reduces Fe3+ at a rate constant of 4.5 x 10(3) M(-1)s(-1) at pH 4.0. Fe2+ is also very stable at a low pH. H2O2 generation results from the autoxidation of the semiquinone radical and was observed only when 2,5-DMHQ was incubated with Fe3+. Consistent with this conclusion, lipid peroxidation occurred only in incubation mixtures containing both 2,5-DMHQ and Fe3+. Catalase and hydroxyl radical scavengers were effective inhibitors of lipid peroxidation, whereas superoxide dismutase caused no inhibition. At a low concentration of oxalate (50 micro M), ferric ion reduction and lipid peroxidation are enhanced. Thus, the enhancement of both ferric ion reduction and lipid peroxidation may be due to oxalate increasing the solubility of the ferric ion. Increasing the oxalate concentration such that the oxalate/ferric ion ratio favored formation of the 2:1 and 3:1 complexes resulted in inhibition of iron reduction and lipid peroxidation. Our results confirm that hydroxyl radical formation occurs via the 2,5-DMBQ redox cycle.  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