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 31 毫秒
1.
One-electron reduction of quinones (Q) by ascorbate (AscH ); (1) AscH + Q --> Q*- + Asc*- + H+, followed by the oxidation of semiquinone (Q*-) by molecular oxygen; (2) Q*- + O2 --> Q + O2*-, results in the catalytic oxidation of ascorbate (with Q as a catalyst) and formation of active forms of oxygen. Along with enzymatic redox cycling of Q. this process may be related to Q cytotoxicity and underlie an antitumor activity of some Qs. In this work, the kinetics of oxygen consumption accompanied the interaction of ascorbate with 55 Qs including substituted 1,4- and 1,2-benzoquinones, naphthoquinones and other quinoid compounds were studied in 50 mM sodium phosphate buffer, pH 7.40, at 37 degrees C by using the Clark electrode technique. The capability of Q to catalyze ascorbate oxidation was characterized by the effective value of kEFF calculated from the initial rate of oxygen consumption (R(OX)) by the equation R(OX) = kEFF[Q][AscH-] as well as by a temporary change in R(OX). The correlation of kEFF with one-electron reduction potential, E(Q/Q*-), showed a sigma-like plot, the same for different kinds of Qs. Only the Qs which reduction potential E(Q/Q*-) ranged from nearly -250 to + 50 mV displayed a pronounced catalytic activity, kEFF increased with shifting E(Q/Q*-) to positive values. The following linear correlation between kEFF (in M (-1) s(-1)) and E(Q/Q*-) (in mV) might be suggested for these Qs: lg(kEFF)= 3.91 + 0.0143E(Q/Q*-). In contrast, Qs with E(Q/Q*-) < - 250 mV and E(Q/Q*-) > + 50 mV showed no measurable catalytic activity. The Qs studied displayed a wide variety in the kinetic regularities of oxygen consumption. When E(Q/Q*-) was more negative than - 100 mV, Q displayed a simple ('standard') kinetic behavior--R(OX) was proportional to [AscH-][Q] independently of concentration of individual reagents, [AscH-] and [Q]; R(OX) did not decrease with time if [AscH-] was held constant: Q recycling was almost reversible. Meanwhile, Qs with E(Q/Q*-) > - 100 mV demonstrated a dramatic deviation from the 'standard' behavior that was manifested by the fast decrease in R(OX) with time and non-linear dependence of even starting values of R(OX) on [Q] and [AscH-]. These deviations were caused basically by the participation of Q*- in side reactions different from (2). The above findings were confirmed by kinetic computer simulations. Some biological implications of Q-AscH- interaction were discussed.  相似文献   

2.
The photoreduction of 1,4-benzoquinone (BQ), 1,4-naphthoquinone (NQ), 9,10-anthraquinone (AQ) and several derivatives, e.g. dimethylBQ, trimethylBQ, duroquinone, bromoNQ, methoxyNQ, methylAQ and dimethylAQ in acetonitrile-water by ascorbate was studied by time-resolved UV-vis spectroscopy using 20 ns laser pulses at 308 nm and continuous 254 nm irradiation. The semiquinone radical (*QH/Q*(-)) is formed after H-atom transfer from ascorbate to the quinone triplet state. The rate constant for quenching is k(q)=(2-9) x 10(9) M(-1) s(-1). Termination of the radicals takes place in the micros-ms range. The results are compared with those initiated by electron transfer from DABCO under similar conditions, where the k(q) values are similar, but the termination of Q*(-) takes place by electron back transfer not yielding hydroquinones. Specific properties of the quinone triplet state, e.g. self-quenching, nucleophilic water addition and the effects of structure are discussed.  相似文献   

3.
A kinetic study of the regeneration reaction of vitamin E (tocopherol) with eight biological hydroquinones (HQs) (ubiquinol-10 (Q10H2 1); ubiquinol-0 (Q0H2 2); vitamin K1 HQ (VK1H2 3); vitamin K3 HQ (VK3H2 4); alpha-, beta-, and gamma-tocopherol-HQs (alpha-, beta-, and gamma-TQH2 5-7); and 2,3,5-trimethyl-1,4-HQ (TMQH2 8)) in solution was performed. The second-order rate constants (k4) for the reaction of HQs 1-8 with alpha-tocopheroxyl and 5,7-diisopropyltocopheroxyl radicals in ethanol, benzene, and isopropyl alcohol/water (5:1, v/v) solutions were measured with a stopped-flow spectrophotometer. The order of magnitude of k4 values obtained for HQs is VK1H2 > VK3H2 > alpha-TQH2 > beta-TQH2 approximately gamma-TQH2 approximately TMQH2 > Q10H2 > Q0H2, being independent of the kinds of tocopheroxyl radicals and the polarity of the solvents. The log of the k4 values obtained for HQs was found to correlate with their peak oxidation potentials. Comparing the k2 value (2.68 x 10(6) M-1 s-1 obtained for the reaction of alpha-tocopheroxyl with vitamin C (sodium ascorbate) with those (k4 = 2.54 x 10(5) and 8.15 x 10(5) M-1 s-1) obtained for the reaction of alpha-tocopheroxyl with Q10H2 and alpha-TQH2 in isopropyl alcohol/water mixtures, the former is approximately 11 and 3 times as reactive as the latter, respectively. On the other hand, the k2 value obtained for sodium ascorbate is smaller than the k4 values obtained for VK1H2 and VK3H2. These results suggest that mixtures of vitamin E and these HQs (as well as those of vitamins E and C) may function synergistically as antioxidants in various tissues and mitochondria.  相似文献   

4.
The reduction of ferrate(VI) by ascorbate has been studied under anaerobic conditions in the pH range between 6.8 and 11.5 at 24 degrees C. A mechanism is proposed that is consistent with the observed rate constants k11 (HFeO4- + AH-) = (5.6 +/- 0.6) x 10(6) M-1 s-1, k12(FeO4(2-) + AH-) = (1.3 +/- 0.1) x 10(6) M-1 s-1 and the pK(HFeO4- in equilibrium with H(+) + FeO4(2-) = 7.9. Stoichiometric studies show that at high ratios of [AH-]/[FeO4(2-)], one ferrate(VI) oxidizes three molecules of ascorbate to the corresponding ascorbyl (A-) radicals.  相似文献   

5.
Lad L  Mewies M  Raven EL 《Biochemistry》2002,41(46):13774-13781
The catalytic mechanism of recombinant soybean cytosolic ascorbate peroxidase (rsAPX) and a derivative of rsAPX in which a cysteine residue (Cys32) located close to the substrate (L-ascorbic acid) binding site has been modified to preclude binding of ascorbate [Mandelman, D., Jamal, J., and Poulos, T. L. (1998) Biochemistry 37, 17610-17617] has been examined using pre-steady-state and steady-state kinetic techniques. Formation (k1 = 3.3 +/- 0.1 x 10(7) M(-1) s(-1)) of Compound I and reduction (k(2) = 5.2 +/- 0.3 x 10(6) M(-1) s(-1)) of Compound I by substrate are fast. Wavelength maxima for Compound I of rsAPX (lambda(max) (nm) = 409, 530, 569, 655) are consistent with a porphyrin pi-cation radical. Reduction of Compound II by L-ascorbate is rate-limiting: at low substrate concentration (0-500 microM), kinetic traces were monophasic but above approximately 500 microM were biphasic. Observed rate constants for the fast phase overlaid with observed rate constants extracted from the (monophasic) dependence observed below 500 microM and showed saturation kinetics; rate constants for the slow phase were linearly dependent on substrate concentration (k(3-slow)) = 3.1 +/- 0.1 x 10(3) M(-1) s(-1)). Kinetic transients for reduction of Compound II by L-ascorbic acid for Cys32-modified rsAPX are monophasic at all substrate concentrations, and the second-order rate constant (k(3) = 0.9 +/- 0.1 x 10(3) M(-1) s(-1)) is similar to that obtained from the slow phase of Compound II reduction for unmodified rsAPX. Steady-state oxidation of L-ascorbate by rsAPX showed a sigmoidal dependence on substrate concentration and data were satisfactorily rationalized using the Hill equation; oxidation of L-ascorbic acid by Cys32-modified rsAPX showed no evidence of sigmoidal behavior. The data are consistent with the presence of two kinetically competent binding sites for ascorbate in APX.  相似文献   

6.
Biological applications of stable nitroxyl radicals, NR, include their use as contrast agents for magnetic resonance imaging, spin labels, superoxide dismutase mimics, and antioxidants. The rapid reduction of NR in biological samples into hydroxylamines (HA) significantly limits their application. In turn, reoxidation of HA back to the NR has been used for detection of reactive oxygen species (ROS). In this work comparative studies of the reduction of pyrrolidine, imidazoline, and imidazolidine NR by ascorbate were performed taking advantage of recently synthesized tetraethyl-substituted NR with much higher stability toward reduction both in vitro and in vivo. Surprisingly, these NR kept 10-50% of initial intensity of electron paramagnetic resonance signal for about 1 h in the presence of 100-fold excess of ascorbate. To explain these data, reoxidation of the corresponding HA by ascorbate radical and dehydroascorbic acid back to the NR was proposed. This hypothesis was supported by direct measurement of the NR appearance from the HA on ascorbate radical generation by ascorbate oxidase, or in the presence of the dehydroascorbic acid. The reversible reaction between NR and ascorbate was observed for the various types of NR, and the rate constants for direct and reverse reactions were determined. The equilibrium constants for one-electron reduction of the tetraethyl-substituted NR by ascorbate were found to be in the range from 2.65x10(-6) to 10(-5) which is significantly lower than corresponding values for the tetramethyl-substituted NR (more or about 10(-4)). This explains the establishment of an EPR-detectable quasi-equilibrium level of tetraethyl-substituted NR in the presence of an excess of ascorbate. The redox reactions of the NR-HA couple in ascorbate-containing media were found to be significantly affected by glutathione (GSH). This effect was attributed to the reduction of ascorbate radicals by GSH, and the rate constant of this reaction was found to be equal to 10 M-1 s-1. In summary, the data provide new insight into the redox chemistry of NR and HA, and significantly affect interpretation and strategy of their use as redox- and ROS-sensitive probes, or as antioxidants.  相似文献   

7.
Protein radicals were selectively generated by reaction with azide radicals on Trp and Tyr residues in insulin, β-lactoglobulin, pepsin, chymotrypsin, and bovine serum albumin at rate constants in the range (2.9–19) × 108 M? 1 s? 1. Monohydrogen ascorbate reduced tryptophanyl radicals in chymotrypsin and pepsin with rate constants in the narrow range of (1.6–1.8) × 108 M? 1 s? 1, whereas β-lactoglobulin tryptophanyl radicals reacted almost 10 times slower. The corresponding values for the protein tyrosyl radicals were about an order of magnitude smaller. Comparison of the rate constants of reactions of free and protein-bound tryptophanyl and tyrosyl radicals showed that, in most cases, the location of the radicals in the protein chain did not constitute a major barrier to the reaction with monohydrogen ascorbate. The results suggest that, under physiological concentrations of dioxygen, monohydrogen ascorbate is likely to be a significant target of protein radicals. It seems likely, therefore, that reaction with protein radicals may be responsible for much of the well-documented loss of ascorbate in living organisms subjected to oxidative stress.  相似文献   

8.
The technique based on monitoring oxygen consumption was applied to study 12 alkyl- and methoxy-substituted p-hydroquinones (QH(2)) as a chain-breaking antioxidant during the oxidation of styrene and methyl linoleate (ML) in bulk as well as ML oxidation in micellar solution of sodium dodecyl sulfate (SDS) at 37 degrees C. The antioxidant activities of QH(2) were characterized by two parameters: the rate constant k(1) for reaction of QH(2) with the peroxy radical LO(2)*: QH(2)+LO(2)*-->QH*+LOOH and the stoichiometric factor of inhibition, f, which shows how many kinetic chains may be terminated by one molecule of QH(2). In the case of styrene and ML oxidation in bulk, f values never exceed two; for the majority of QH(2), f was found to be significantly less than two due to the interaction of QH* with molecular oxygen. In the absence of superoxide dismutase (SOD), all the studied QH(2) displayed a very moderate if any antioxidant capability during ML oxidation in SDS micelles. When 20U/ml SOD was added, the majority of QH(2) showed a pronounced ability to inhibit ML oxidation, f parameter being ca. one. The features of QH(2) as an antioxidant in aqueous environment are suggested to associate with the reactivity of semiquinone (Q*(-)). Q*(-) reacts readily with molecular oxygen with formation of superoxide (O(2)*(-)); further reactions of O(2)*(-) result in fast depleting QH(2) and chain propagation. The addition of SOD results in purging a reaction mixture from O(2)*(-) and, as a corollary, in depressing undesirable reactions with the participation of O(2)*(-). With all the oxidation models, QH(2) were found to be very reactive to LO(2)*. The rate constants k(1) decreased progressively when going from the oxidation of styrene to ML oxidation in bulk and further to ML oxidation in SDS micelles.  相似文献   

9.
Capsaicinoids are the pungent compounds in Capsicum fruits (i.e., "hot" peppers). Peroxidases catalyze capsaicinoid oxidation and may play a central role in their metabolism. However, key kinetic aspects of peroxidase-catalyzed capsaicinoid oxidation remain unresolved. Using transient-state methods, we evaluated horseradish peroxidase compound I and II reduction by two prominent capsaicinoids (25 degrees C, pH 7.0). We determined rate constants approaching 2 x 10(7) and 5 x 10(5)M(-1)s(-1) for compound I and compound II reduction, respectively. We also determined k(app) values for steady-state capsaicinoid oxidation approaching 8 x 10(5)M(-1)s(-1) (25 degrees C, pH 7.0). Accounting for stoichiometry, these are in excellent agreement with constants for compound II reduction, suggesting that this reaction governs capsaicinoid-dependent peroxidase turnover. Ascorbate rapidly reduced capsaicinoid radicals, assisting in the determination of the kinetic constants reported. Because ascorbate accumulates in Capsicum fruits, it may also be an important determinant for capsaicinoid content and preservation in Capsicum fruits and related products.  相似文献   

10.
The one- or two-electron oxidation of thyroglobulin by the thyroid peroxidase system was found to be regulated by the iodine content of thyroglobulin. The catalytic intermediate of thyroid peroxidase observed at steady state of the reaction was Compound I and II when the iodine content in thyroglobulin was 0.2 and 0.7%, respectively, apparent rate constants for the rate-limiting steps being estimated at 4.7 x 10(7) and 4.8 x 10(4) M-1 S-1. The thyroglobulin-mediated oxidation of GSH occurred by way of two-electron transfer at 0.2% iodine content and by way of one-electron transfer at 0.7% iodine content. The spin-trapping experiment with 5,5-dimethyl-1-pyrroline-N-oxide showed that glutathione radicals were formed in the latter reaction but not in the former. In the reactions of thyroid peroxidase, the one- and two-electron oxidations of ascorbate were also mediated by 0.2 and 0.7% iodine thyroglobulins, respectively. The reactions were analyzed and mimicked with the use of p-cresol and p-acetaminophenol as a mediator in the reactions of lactoperoxidase and thyroid peroxidase.  相似文献   

11.
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   

12.
To assess antioxidative protection by carnosic acid (CA) in combination with that of other low-molecular weight (M(r)) antioxidants (alpha-tocopherol [alpha-T] and ascorbate [Asc]) in chloroplasts, we measured endogenous concentrations of these antioxidants, their redox states, and other indicators of oxidative stress in chloroplasts of three Labiatae species, differing in their CA contents, exposed to drought stress in the field. Damage to the photosynthetic apparatus was observed neither in CA-containing species (rosemary [Rosmarinus officinalis]) and sage [Salvia officinalis]) nor in CA-free species (lemon balm [Melissa officinalis]) at relative leaf water contents between 86% and 58%, as indicated by constant maximum efficiency of photosystem II photochemistry ratios and malondialdehyde levels in chloroplasts. The three species showed significant increases in alpha-T, a shift of the redox state of alpha-T toward its reduced state, and increased Asc levels in chloroplasts under stress. Lemon balm showed the highest increases in alpha-T and Asc in chloroplasts under stress, which might compensate for the lack of CA. Besides, whereas in rosemary and sage, the redox state of CA was shifted toward its oxidized state and the redox state of Asc was kept constant, lemon balm displayed a shift of the redox state of Asc toward its oxidized state under stress. In vitro experiments showed that both CA and Asc protect alpha-T and photosynthetic membranes against oxidative damage. These results are consistent with the contention that CA, in combination with other low-M(r) antioxidants, helps to prevent oxidative damage in chloroplasts of water-stressed plants, and they show functional interdependence among different low-M(r) antioxidants in chloroplasts.  相似文献   

13.
Eosinophil peroxidase, the major granule protein in eosinophils, is the least studied human peroxidase. Here, we have performed spectral and kinetic measurements to study the nature of eosinophil peroxidase intermediates, compounds I and II, and their reduction by the endogenous one-electron donors ascorbate and tyrosine using the sequential-mixing stopped-flow technique. We demonstrate that the peroxidase cycle of eosinophil peroxidase involves a ferryl/porphyrin radical compound I and a ferryl compound II. In the absence of electron donors, compound I is shown to be transformed to a species with a compound II-like spectrum. In the presence of ascorbate or tyrosine compound I is reduced to compound II with a second-order rate constant of (1.0+/-0.2)x10(6) M(-1) s(-1) and (3.5+/-0.2)x10(5) M(-1) s(-1), respectively (pH 7.0, 15 degrees C). Compound II is then reduced by ascorbate and tyrosine to native enzyme with a second-order rate constant of (6.7+/-0.06)x10(3) M(-1) s(-1) and (2.7+/-0.06)x10(4) M(-1) s(-1), respectively. This study revealed that eosinophil peroxidase compounds I and II are able to react with tyrosine and ascorbate via one-electron oxidations and therefore generate monodehydroascorbate and tyrosyl radicals. The relatively fast rates of the compound I reduction demonstrate that these reactions may take place in vivo and are physiologically relevant.  相似文献   

14.
Catalytic oxidation of p-cresol by ascorbate peroxidase   总被引:2,自引:0,他引:2  
Transient and steady state kinetics, together with a range of chromatographic and spectroscopic techniques, have been used to establish the mechanism and the products of the H(2)O(2)-dependent oxidation of p-cresol by ascorbate peroxidase (APX). HPLC, GC-MS, and NMR analyses are consistent with the formation of 2, 2'-dihydroxy-5,5'-dimethylbiphenyl (II) and 4alpha,9beta-dihydro-8, 9beta-dimethyl-3(4H)-dibenzofuranone (Pummerer's ketone, III) as the major products of the reaction. In the presence of cumene hydroperoxide, two additional products were observed which, from GC and MS analyses, were shown to be 1,1-dimethylbenzylalcohol (IV) and bis-(1-methyl-1-phenyl-ethyl)-peroxide (V). The product ratio II:III was dependent on enzyme concentration: at low concentrations Pummerer's ketone (III) predominates and at high concentrations formation of the biphenyl compound (II) is favored. Steady-state data showed a sigmoidal dependence on [p-cresol] that was consistent with the presence of 2.01 +/- 0.15 binding sites for the substrate (25.0 degrees C, sodium phosphate, pH 7.0, mu = 2.2 mM) and independent of ionic strength in the range 2.2-500 mM. Single turnover kinetic experiments (pH 7.0, 5.0 degrees C, mu = 0.10 M) yielded second-order rate constants for Compound I reduction by p-cresol, k(2), of 5.42 +/- 0.10 x 10(5) M(-1) s(-1), respectively. Rate-limiting reduction of Compound II by p-cresol, k(3), showed saturation kinetics, giving values for K(d) = 1.54 +/- 0.12 x 10(-3) M and k(3) = 18.5 +/- 0.7 s(-1). The results are discussed in the more general context of APX-catalyzed aromatic oxidations.  相似文献   

15.
One-electron oxidation of (6R)-5,6,7,8-tetrahydrobiopterin (H(4)B) by the azide radical generates the radical cation (H(4)B(*)(+)) which rapidly deprotonates at physiological pH to give the neutral trihydrobiopterin radical (H(3)B(*)); pK(a) (H(4)B(*)(+) <==> H(3)B(*) + H(+)) = (5.2 +/- 0.1). In the absence of ascorbate both the H(4)B(*)(+) and H(3)B(*) radicals undergo disproportionation to form quinonoid dihydrobiopterin (qH(2)B) and the parent H(4)B with rate constants k(H(4)B(*)(+) + H(4)B(*)(+)) = 6.5 x 10(3) M(-1) s(-1) and k(H(3)B(*) + H(3)B(*)) = 9.3 x 10(4) M(-1) s(-1), respectively. The H(3)B(*) radical is scavenged by ascorbate (AscH(-)) with an estimated rate constant of k(H(3)B(*) + AscH(-)) similar 1.7 x 10(5) M(-1) s(-1). At physiological pH the pterin rapidly scavenges a range of biological oxidants often associated with cellular oxidative stress and nitric oxide synthase (NOS) dysfunction including hydroxyl ((*)OH), nitrogen dioxide (NO(2)(*)), glutathione thiyl (GS(*)), and carbonate (CO(3)(*-)) radicals. Without exception these radicals react appreciably faster with H(4)B than with AscH(-) with k(*OH + H(4)B) = 8.8 x 10(9) M(-1) s(-1), k(NO(2)(*) + H(4)B) = 9.4 x 10(8) M(-1) s(-1), k(CO(3)(*-) + H(4)B) = 4.6 x 10(9) M(-1) s(-1), and k(GS(*) + H(4)B) = 1.1 x 10(9) M(-1) s(-1), respectively. The glutathione disulfide radical anion (GSSG(*-)) rapidly reduces the pterin to the tetrahydrobiopterin radical anion (H(4)B(*-)) with a rate constant of k(GSSG(*-) + H(4)B) similar 4.5 x 10(8) M(-1) s(-1). The results are discussed in the context of the general antioxidant properties of the pterin and the redox role played by H(4)B in NOS catalysis.  相似文献   

16.
A methyl viologen (MV)* mediated Mehler reaction was studied using Type C and D chloroplasts (thylakoids) from spinach. The extent of photooxidative reactions were measured as (a) rate of ethylene formation from methional oxidation indicating the production of oxygen radicals, and (b) rate of malondialdehyde (MDA) formation as a measure of lipid peroxidation. Without added ascorbate, 1 M FerricEDTA increased ethylene formation by greater than 4-fold, but had no effect on MDA production. Ascorbate (1 mM) produced a tripling of ethylene while it reduced MDA formation in the presence of iron. Radical scavengers diethyldithiocarbamate (DDTC), formate, 1,4-diazabicyclo (2.2.2octane) (DABCO), inhibited ethylene formation. Using 0,4 M mannitol to scavenge hydroxyl radicals, the rates of ethylene formation were reduced 40 to 60% with or without 1 M Fe(III) EDTA. The strong oxidant(s) not scavenged by mannitol are hypothesized to be either alkoxyl radicals from lipid peroxidation, or site specific formation of hydroxyl radicals in a lipophillic environment not exposed to mannitol. Singlet oxygen does not appear to be a significant factor in this system. Catalase strongly inhibited both ethylene and MDA synthesis under all conditions; 1 mM ascorbate did not reverse this inhibition. However, the strong superoxide dismutase (SOD) inhibition of ethylene and MDA formation was completely reversed by 1 mM ascorbate. This suggests that superoxide was functioning as an iron reducing agent and that in its absence, ascorbate was similarly promoting oxidations. Therefore, these oxidative processes were dependent on the presence of H2O2 and a reducing agent, suggesting the involvement of a Fenton-type reaction.Abbreviation DABCO 1,4-diazabicyclo(2.2.2.octane) - DCMU 3-(3,4 Dichlorophenyl). 1,1-dimethyl urea - DDTC diethyldithiocarbamate - EDTA ethylenediamine-tetraacetic acid - MDA malondialdehyde - MV methyl viologen - SOD superoxide dismutase - TBA thiobarbituric acid - TCA trichloroacetic acid Scientific contribution number 1315 from the New Hampshire Agriculture Experiment Station.  相似文献   

17.
Distribution of ascorbate into tissues is an essential process in ascorbate antioxidant defense. Hibernating animals are studied as a model of tolerance to ischemia-reperfusion because of their tolerance to fluctuations in blood flow associated with prolonged torpor and periodic arousal episodes. Throughout hibernation, plasma ascorbate concentration ([Asc](p)) repetitively increases during torpor, then falls during periodic arousal bouts. We previously proposed that high [Asc](p) provides a ready source of antioxidant protection for distribution to the central nervous system and peripheral tissues during arousal. Here we tested whether deliberate oxidation of plasma ascorbate by intravenous administration of ascorbate oxidase (AO), prior to arousal, compromised tissue levels of ascorbate or the other water-soluble antioxidants, glutathione (GSH) and urate. Although AO decreased [Asc](p) to below the level of detection during torpor and after arousal, ascorbate oxidation did not decrease post-arousal tissue levels of reduced ascorbate, glutathione, or urate in any tissue examined, except liver. The data imply that ascorbate is taken up equally well into brain and other tissues as either ascorbate or its oxidized product dehydroascorbate, with subsequent intracellular reduction of dehydroascorbate. Lack of effect of ascorbate oxidation on tissue levels of GSH or urate indicates that dehydroascorbate uptake and reduction do not compromise tissue concentrations of these other water-soluble antioxidants. Thus, we show equal availability of reduced and oxidized plasma ascorbate during metabolically demanding thermogenesis and reperfusion associated with arousal from hibernation.  相似文献   

18.
Direct reactions of peroxidases with Trolox C (a vitamin E analogue) and vitamin E were observed in 50% (v/v) methanol. The kinetic results revealed that the reaction of horseradish peroxidase intermediate Compound II with Trolox C and vitamin E was the rate-determining step, and the rate constants were estimated to be 1.7 x 10(3) and 5.1 x 10(2) M-1.s-1, respectively. Peroxidases catalyzed the one-electron oxidation of Trolox C and vitamin E, and the vitamin E phenoxyl radicals resulting from the peroxidase reactions were detected by continuous-flow ESR spectroscopy.  相似文献   

19.
Previously, we showed that ascorbate (Asc), by donating electrons to photosystem II (PSII), supports a sustained electron transport activity in leaves in which the oxygen-evolving complexes were inactivated with a heat pulse (49°C, 40 s). Here, by using wild-type, Asc-overproducing, and -deficient Arabidopsis (Arabidopsis thaliana) mutants (miox4 and vtc2-3, respectively), we investigated the physiological role of Asc as PSII electron donor in heat-stressed leaves (40°C, 15 min), lacking active oxygen-evolving complexes. Chlorophyll-a fluorescence transients show that in leaves excited with trains of saturating single-turnover flashes spaced 200 ms apart, allowing continual electron donation from Asc to PSII, the reaction centers remained functional even after thousands of turnovers. Higher flash frequencies or continuous illumination (300 μmol photons m(-2) s(-1)) gradually inactivated them, a process that appeared to be initiated by a dramatic deceleration of the electron transfer from Tyr(Z) to P680(+), followed by the complete loss of charge separation activity. These processes occurred with half-times of 1.2 and 10 min, 2.8 and 23 min, and 4.1 and 51 min in vtc2-3, the wild type, and miox4, respectively, indicating that the rate of inactivation strongly depended on the Asc content of the leaves. The recovery of PSII activity, following the degradation of PSII proteins (D1, CP43, and PsbO), in moderate light (100 μmol photons m(-2) s(-1), comparable to growth light), was also retarded in the Asc-deficient mutant. These data show that high Asc content of leaves contributes significantly to the ability of plants to withstand heat-stress conditions.  相似文献   

20.
Using pulse radiolysis and laser flash photolysis, we have investigated the reactions of the deleterious species, e(-)(aq), HO&z.rad;, O(2)(*)(-) and O(2)((1)Delta(g)) with 10 water-soluble cyclopropyl-fused C(60) derivatives including a mono-adduct dendro[60]fullerene (d) and C(60) derivatives based on C(60)[C(COOH)(2)](n=2-6), some of which are known to be neuroprotective in vivo. The rate constants for reactions of e(-)(aq) and HO&z.rad; lie in the range 0.5-3.3 x 10(10) M(-1) s(-1). The d and bis-adduct monoanion radicals display sharp absorption peaks around 1000 nm (epsilon = 7 000-11 500 M(-1) cm(-1)); the anions of the tris-, tetra-, and penta-adduct derivatives have broader, weaker absorptions. The monohydroxylated radicals have their most intense absorption maxima around 390-440 nm (epsilon = 1000-3000 M(-1) cm(-1)). The anion and hydroxylated radical absorption spectra display a blue-shift as the number of addends increases. The radical anions react with oxygen (k approximately 10(7)-10(9) M(-1) s(-1)). The reaction of O(2)(*)(-) with the C(60) derivatives does not occur via an electron transfer. The rate constants for singlet oxygen reaction with the dendrofullerene and eee-derivative in D(2)O at pH 7.4 are k approximately 7 x 10(7) and approximately 2 x 10(7) M(-1) s(-1) respectively, in contrast to approximately 1.2 x 10(5) M(-1) s(-1) for the reaction with C(60) in C(6)D(6). The large acceleration of the rates for electron reduction and singlet oxygen reactions in water is due to a solvophobic process.  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