首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
Prostaglandin H synthase isoforms 1 and 2 (PGHS-1 and -2) each have a peroxidase activity and also a cyclooxygenase activity that requires initiation by hydroperoxide. The hydroperoxide initiator requirement for PGHS-2 cyclooxygenase is about 10-fold lower than for PGHS-1 cyclooxygenase, and this difference may contribute to the distinct control of cellular prostanoid synthesis by the two isoforms. We compared the kinetics of the initial peroxidase steps in PGHS-1 and -2 to quantify mechanistic differences between the isoforms that might contribute to the difference in cyclooxygenase initiation efficiency. The kinetics of formation of Intermediate I (an Fe(IV) species with a porphyrin free radical) and Intermediate II (an Fe(IV) species with a tyrosyl free radical, thought to be the crucial oxidant in cyclooxygenase catalysis) were monitored at 4 degrees c by stopped flow spectrophotometry with several hydroperoxides as substrate. With 15-hydroperoxyeicosatetraenoic acid, the rate constant for Intermediate I formation (k1) was 2.3 x 10(7) M-1 s-1 for PGHS-1 and 2.5 x 10(7) M-1 s-1 for PGHS-2, indicating that the isoforms have similar initial reactivity with this lipid hydroperoxide. For PGHS-1, the rate of conversion of Intermediate I to Intermediate II (k2) became the limiting factor when the hydroperoxide level was increased, indicating a rate constant of 10(2)-10(3) s-1 for the generation of the active cyclooxygenase species. For PGHS-2, however, the transition between Intermediates I and II was not rate-limiting even at the highest hydroperoxide concentrations tested, indicating that the k2 value for PGHS-2 was much greater than that for PGHS-1. Computer modelling predicted that faster formation of the active cyclooxygenase species (Intermediate II) or increased stability of the active species increases the resistance of the cyclooxygenase to inhibition by the intracellular hydroperoxide scavenger, glutathione peroxidase. Kinetic differences between the PGHS isoforms in forming or stabilizing the active cyclooxygenase species can thus contribute to the difference in the regulation of their cellular activities.  相似文献   

2.
The peroxidase reaction of prostaglandin endoperoxide synthase was investigated by transient state kinetics using stoichiometric amounts of substrates. The rate constants for the conversion of compound I to intermediate II determined with a stoichiometric amount of hydroperoxide were found to be lower by an order of magnitude than when an excess of hydroperoxide was used. The difference was attributed to ability of the compound I of prostaglandin endoperoxide synthase to be reduced by the excess of hydroperoxide. This suggests that the true rate constant of unimolecular conversion compound I to intermediate II at 3 degrees C is 5-10 s-1 instead of 50-200 s-1 as reported before. The latter value rather characterizes the combined process of spontaneous and hydroperoxide-dependent transformation of compound I. Stoichiometric amounts of reducing substrates significantly stimulated transformation of compound I. This effect could not be entirely explained by their reducing action, which was measured by following the oxidation kinetics. The results of the global fit of the experimental data suggest that reducing substrates, in addition to their direct action in reducing compound I to compound II, indirectly stimulate transformation of compound I to the tyrosyl radical form of intermediate II, thereby stimulating the cyclooxygenase reaction.  相似文献   

3.
The reactivity of rat liver glutathione (GSH) peroxidase with two hydroperoxides was determined using integrated rate equations. The bimolecular rate constant for the reaction of GSH peroxidase with linoleic acid hydroperoxide is approximately four times the rate constant with cumene hydroperoxide. The reactivity toward reduced glutathione is not altered by different hydroperoxides. The t12 for lipid hydroperoxide in rat liver is approximated at 9.5 × 10?5 min.  相似文献   

4.
9-Methylfluorene was metabolized by rat liver microsomes to 9-hydroperoxy-9-methylfluorene and 9-hydroxy-9-methylfluorene. The results were confirmed by using a reconstituted cytochrome P-450 oxygenase system purified from phenobarbital-induced rat liver which established its involvement. SKF-525A strongly inhibited the formation of both oxygenation products. Cytochrome P-450 alone brought about the conversion of the hydroperoxide to its alcohol. NADPH augmented the peroxidative reaction, but the presence of NADPH-cytochrome P-450 reductase was without effect. Certain microsomal preparations and reconstituted enzyme yielded little or no detectable amounts of hydroperoxide. This was due to a too rapid conversion of the hydroperoxide to its alcohol. The addition of metyrapone, a compound that inhibited such conversion, resulted in accumulation of 9-hydroperoxy-9-methylfluorene for positive identification. Incubation of 9-methylfluorene with microsomes and NADPH resulted in covalent binding of its metabolite to microsomal proteins. Incubation of 14C-labeled 9-hydroperoxy-9-methylfluorene caused covalent binding of label to proteins, RNA, and DNA.  相似文献   

5.
The pre-steady-state kinetics of the prostaglandin endoperoxide synthase oxygenase reaction with eicosadienoic acids and the cyclooxygenase reaction with arachidonic acid were investigated by stopped-flow spectrophotometry at 426 nm, an isosbestic point between native enzyme and compound I. A similar reaction mechanism for both types of catalysis is defined from combined kinetic experiments and numerical simulations. In the first step a fatty acid hydroperoxide reacts with the native enzyme to form compound I and the fatty acid hydroxide. In the second step the fatty acid reduces compound I to compound II and a fatty acid carbon radical is formed. This is followed by two fast steps: (1) the addition of either one molecule of oxygen (the oxygenase reaction) or two molecules of oxygen (the cyclooxygenase reaction) to the fatty acid carbon radical to form the corresponding hydroperoxyl radical, and (2) the reaction of the hydroperoxyl radical with compound II to form the fatty acid hydroperoxide and a compound I-protein radical. A unimolecular reaction of the compound I-protein radical to reform the native enzyme is assumed for the last step in the cycle. This is a slow reaction not significantly affecting steps 1 and 2 under pre-steady-state conditions. A linear dependence of the observed pseudo-first-order rate constant, k(obs), on fatty acid concentration is quantitatively reproduced by the model for both the oxygenase and cyclooxygenase reactions. The simulated second order rate constants for the conversion of native enzyme to compound I with arachidonic or eicosadienoic acids hydroperoxides as a substrate are 8 x 10(7) and 4 x 10(7) M(-1) s(-1), respectively. The simulated and experimentally obtained second-order rate constants for the conversion of compound I to compound II with arachidonic and eicosadienoic acids as a substrate are 1.2 x 10(5) and 3.0 x 10(5) M(-1) s(-1), respectively.  相似文献   

6.
The singlet oxygen trap diphenylfuran was rapidly oxidized to cis dibenzoylethylene during the decomposition of linoleic acid hydroperoxide catalyzed by ceric ions, methemoglobin or hematin. This conversion was enhanced in a deuterated medium and inhibited by other singlet oxygen quenchers or traps. The chemiluminescence accompanying the decomposition of the linoleic acid hydroperoxide was also markedly enhanced in a deuterated medium and inhibited by other singlet oxygen quenchers or traps. Antioxidants markedly inhibited these reactions. It is concluded that singlet oxygen is formed in substantial quantities during the metal catalyzed decomposition of linoleic acid hydroperoxide.  相似文献   

7.
The reaction of native myeloperoxidase (MPO) and its redox intermediate compound I with hydrogen peroxide, ethyl hydroperoxide, peroxyacetic acid, t-butyl hydroperoxide, 3-chloroperoxybenzoic acid and cumene hydroperoxide was studied by multi-mixing stopped-flow techniques. Hydroperoxides are decomposed by MPO by two mechanisms. Firstly, the hydroperoxide undergoes a two-electron reduction to its corresponding alcohol and heme iron is oxidized to compound I. At pH 7 and 15 degrees C, the rate constant of the reaction between 3-chloroperoxybenzoic acid and ferric MPO was similar to that with hydrogen peroxide (1.8x10(7) M(-1) s(-1) and 1.4x10(7) M(-1) s(-1), respectively). With the exception of t-butyl hydroperoxide, the rates of compound I formation varied between 5.2x10(5) M(-1) s(-1) and 2.7x10(6) M(-1) s(-1). Secondly, compound I can abstract hydrogen from these peroxides, producing peroxyl radicals and compound II. Compound I reduction is shown to be more than two orders of magnitude slower than compound I formation. Again, with 3-chloroperoxybenzoic acid this reaction is most effective (6. 6x10(4) M(-1) s(-1) at pH 7 and 15 degrees C). Both reactions are controlled by the same ionizable group (average pK(a) of about 4.0) which has to be in its conjugated base form for reaction.  相似文献   

8.
Yeasts lacking cytoplasmic superoxide dismutase (Cu,Zn-SOD) activity are permanently subjected to oxidative stress. We used two-dimensional PAGE to examine the proteome pattern of Saccharomyces cerevisiae strains lacking Cu,Zn-SOD. We found a new stable form of alkyl hydroperoxide reductase 1 (Ahp1) with a lower isoelectric point. This form was also present in wild type strains after treatment with tert-butyl hydroperoxide. In vitro enzyme assays showed that Ahp1p had lower specific activity in strains lacking Cu,Zn-SOD. We studied three mutants presenting a reduced production of the low pI variant under oxidative stress conditions. Two of the mutants (C62S and S59D) were totally inactive, thus suggesting that the acidic form of Ahp1p may only appear when the enzyme is functional. The other mutant (S59A) was active in vitro and was more resistant to inactivation by tert-butyl hydroperoxide than the wild type enzyme. Furthermore, the inactivation of Ahp1p in vitro is correlated with its conversion to the low pI form. These results suggest that in vivo during some particular oxidative stress (alkyl hydroperoxide treatment or lack of Cu,Zn-SOD activity but not hydrogen peroxide treatment), the catalytic cysteine of Ahp1p is more oxidized than cysteine-sulfenic acid (a natural occurring intermediate of the enzymatic reaction) and that cysteine-sulfinic acid or cysteine-sulfonic acid variant may be inactive.  相似文献   

9.
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   

10.
Our results demonstrate that the organic hydroperoxide t-butyl-hydroperoxide (TBHP) influences the synthesis of prostaglandins in the human embryo lung fibroblast. TBHP inhibits or stimulates prostaglandin synthesis as a function of its concentration. Regardless of the concentration employed in these experiments however, TBHP stimulated the release of arachidonate from lipid stores. When the arachidonate release step in prostaglandin (PG) synthesis was bypassed by the addition of free arachidonate to the cell cultures, t-butyl hydroperoxide further stimulated PG synthesis, indicating that the hydroperoxide activates arachidonate conversion to prostaglandins. 4-Hydroxy-2,2,6,6-tetramethylpiperidinoxy radical, a scavenger of oxygen radicals, when added to cell cultures alone had no measurable effect on either arachidonate release or prostaglandin synthesis. When 4-hydroxy-2,2,6,6-tetramethylpiperidinoxy radical was administered to cell cultures in combination with t-butyl hydroperoxide, it increased prostaglandin synthesis while inhibiting arachidonate release by the hydroperoxide. The participation of hydroperoxides and trappers of oxygen radicals in the regulation of PG synthesis is not unique to lung fibroblasts. Endothelial cells from the vasculature as well as fibroblasts from the cornea also appear to be affected by these compounds with respect to prostaglandin synthesis.  相似文献   

11.
R Labeque  L J Marnett 《Biochemistry》1988,27(18):7060-7070
Reaction of 10-hydroperoxyoctadec-8-enoic acid (10-OOH-18:1) (50 microM) with hematin (0.5 microM) in sodium phosphate buffer containing Tween 20 (200 microM) generates 10-oxooctadec-8-enoic acid, 10-oxodec-8-enoic acid (10-oxo-10:1), and 10-hydroxyoctadec-8-enoic acid in relative yields of 79, 4, and 17%, respectively. The product profile and relative distribution are unaffected by 1 mM butylated hydroxyanisole. Approximately 5% of the hydroperoxide isomerizes from the 10- to the 8-position. 10-Oxo-10:1 most likely arises via beta-scission of an intermediate alkoxyl radical to the aldehyde and the n-octyl radical. To test this, 10-hydroperoxyoctadeca-8,12-dienoic acid was reacted with hematin under identical conditions. 10-Oxooctadeca-8,12-dienoic acid, 10-oxodec-8-enoic acid, and 10-hydroxyoctadeca-8,12-dienoic acid are formed in relative yields of 50, 45, and 5%, respectively. The product ratios are constant with time and hydroperoxide to catalyst ratio and unaffected by inclusion of phenolic antioxidants. The higher yield of 10-oxo-10:1 from 10-OOH-18:2 compared to 10-OOH-18:1 is due to the higher rate of beta-scission of the intermediate alkoxyl radical from the former to the resonance-stabilized octenyl radical. Two products of reaction of the 2-octenyl radical with O2, octenal and octenol, were detected in 10% yield relative to 10-oxo-10:1. Inclusion of 7,8-dihydroxy-7,8-dihydrobenzo[a]pyrene (BP-7,8-diol) led to epoxidation by both 10-OOH-18:1 and 10-OOH-18:2. Studies with isotopically labeled hydroperoxide or O2 indicated approximately 65% of the epoxide oxygen was derived from O2 and 35% from hydroperoxide oxygen, consistent with the involvement of peroxyl free radicals as the oxidizing agents. The available evidence indicates that hematin reduces the fatty acid hydroperoxides homolytically to alkoxyl radicals that are oxidized to ketones, reduced to alcohols, or undergo beta-scission to aldehydes. Carbon radicals generated during these reactions couple to O2, generating peroxyl free radicals that epoxidize BP-7,8-diol. The smaller percentage of epoxidation that results from hydroperoxide oxygen may arise from oxidation of the hydroperoxide group to peroxyl radicals or from heterolytic cleavage of the hydroperoxide to alcohol and an iron-oxo complex.  相似文献   

12.
We studied the mechanism of formation of oxygen radicals during ferrous ion-induced decomposition of linoleic acid hydroperoxide using the spin trapping and chemiluminescence methods. The formation of the superoxide anion (O2*-) was verified in the present study. The hydroxyl radical is also generated through Fenton type decomposition of hydrogen peroxide produced on disproportionation of O2*-. A carbon-centered radical was detected using 5-(diethoxyphosphoryl)-5-methyl-1-pyrroline N-oxide (DEPMPO) as a spin trap. Alkoxyl radical formation is essential for the conversion of linoleic acid hydroperoxide into the peroxyl radical by ferrous ion. It is likely that the alkoxyl radical [R1CH(O*)R2] is converted into the hydroxylcarbon radical [R1C*(OH)R2] in water, and that this carbon radical reacts with oxygen to give the alpha-hydroxyperoxyl radical [R1R2C(OH)OO*], which decomposes into the carbocation [R1C+(OH)R2] and O2*-.  相似文献   

13.
BackgroundExtracellular surface protein disulfide isomerase-A1 (PDI) is involved in platelet aggregation, thrombus formation and vascular remodeling. PDI performs redox exchange with client proteins and, hence, its oxidation by extracellular molecules might alter protein function and cell response. In this study, we investigated PDI oxidation by urate hydroperoxide, a newly-described oxidant that is generated through uric acid oxidation by peroxidases, with a putative role in vascular inflammation.MethodsAmino acids specificity and kinetics of PDI oxidation by urate hydroperoxide was evaluated by LC-MS/MS and by stopped-flow. Oxidation of cell surface PDI and other thiol-proteins from HUVECs was identified using impermeable alkylating reagents. Oxidation of intracellular GSH and GSSG was evaluated with specific LC-MS/MS techniques. Cell adherence, detachment and viability were assessed using crystal violet staining, cellular microscopy and LDH activity, respectively.ResultsUrate hydroperoxide specifically oxidized cysteine residues from catalytic sites of recombinant PDI with a rate constant of 6 × 103 M−1 s−1. Incubation of HUVECs with urate hydroperoxide led to oxidation of cell surface PDI and other unidentified cell surface thiol-proteins. Cell adherence to fibronectin coated plates was impaired by urate hydroperoxide, as well as by other oxidants, thiol alkylating agents and PDI inhibitors. Urate hydroperoxide did not affect cell viability but significantly decreased GSH/GSSG ratio.ConclusionsOur results demonstrated that urate hydroperoxide affects thiol-oxidation of PDI and other cell surface proteins, impairing cellular adherence.General significanceThese findings could contribute to a better understanding of the mechanism by which uric acid affects endothelial cell function and vascular homeostasis.  相似文献   

14.
Cells of Saccharomycescerevisiae, harvested from log phase cultures, contain cytochrome P-450 and are capable of activating promutagens to products that are genetically active in the same cell. The effect of cumene hydroperoxide, a compound known to support cytochrome P-450-mediated reactions, on the activation of a variety of the promutagens was investigated. In all cases the genetic activity of the promutagens was increased. With dimethyl-nitrosamine as the promutagen, the increased rate of gene conversion was linear for at least 1 hr. Yeast cytochrome P-450 was stable in intact cells in the presence of cumene hydroperoxide. However, in microsomal preparations the cytochrome was rapidly destroyed. When cumene hydroperoxide was added to a suspension of intact yeast cells, a spectrum with a Soret maximum at 455 nm — indicative of an interaction with cytochrome P-450 — was observed.  相似文献   

15.
Isoelectric focusing in thin-layer polyacrylamide gels has been applied to the analysis of the enzymes involved in the formation and destruction of peroxides in soybeans [Glycine max (L.)], lipoxygenases and peroxidases, respectively. As a result of differences in pH optima for catalytic activity, lipoxygenases were selectively detected by adjusting the pH employed for activity-specific staining. Type-1 lipoxygenase was revealed not only by staining based on the conversion of linoleic acid to hydroperoxide but also by two stains based on the reduction of the hydroperoxide. These methods were found to be suitable for the analysis and characterization of isoenzyme patterns in different soybean cultivars. A substantial difference in the distribution of lipoxygenases maximally active near pH 7 was observed for cultivars Provar and Vickery. A similar degree of separation of the isoenzymes was achieved on a larger scale using chromato-focusing in the pH range 7.4-5.0.  相似文献   

16.
Deuterium isotope effects on the kinetic parameters for the hydroperoxide-supported N-demethylation of N,N-dimethylaniline catalyzed by chloroperoxidase and horseradish peroxidase were determined using N,N-di-(trideuteromethyl)aniline. The isotope effect on the Vmax for the chloroperoxidase-catalyzed demethylation reaction supported by ethyl hydroperoxide was 1.42 +/- 0.31. The isotope effects on the Vmax for the horseradish peroxidase-catalyzed reaction supported by ethyl hydroperoxide and hydrogen peroxide were 1.99 +/- 0.39 and 4.09 +/- 0.27, respectively. Isotope effects ranging from 1.76 to 5.10 were observed on the Vmax/Km for the hydroperoxide substrate (i.e. the second order rate constant for the reaction of the hydroperoxide with the peroxidase to form compound I) in both enzyme systems when the N-methyl groups of N,N-dimethylaniline were deuterated. These results are not predicted by the simple ping-pong kinetic model for peroxidase-catalyzed N-demethylation reactions. The data are most simply explained by a mechanism involving the transfer of deuterium (or hydrogen) from N,N-dimethylaniline to the enzyme during catalysis. The deuterium must subsequently be displaced from the enzyme by the hydroperoxide, causing the observed isotope effects.  相似文献   

17.
9-Methylfluorene was found to be oxygenated to 9-hydroperoxy-9-methylfluorene and 9-hydroxy-9-methylfluorene by cytochrome P-450 in the presence of cumene hydroperoxide. Molecular oxygen is required and carbon monoxide is inhibitory. The reaction is inhibited by SKF-525A and metyrapone. Metyrapone and cumene hydroperoxide also retard the conversion of 9-hydroperoxy-9-methylfluorene to 9-hydroxy-9-methylfluorene. The reaction is different from hydroperoxide-supported oxygenation, since the cumene hydroperoxide appears to act as an effector of the enzyme rather than oxygen donor. It is suggested that substrates with stable radicals can be dioxygenated in this manner.  相似文献   

18.
Cholesteryl ester hydroperoxide (CE-OOH) and phosphatidylcholine hydroperoxide (PC-OOH) are the major primary oxidation products of lipoproteins. CE-OOH is present in human and rat plasmas while PC-OOH is undetectable. This is likely due to the enzymatic (plasma glutathione peroxidase) and the nonenzymatic (apolipoproteins A and B-100) reducing activities of PC-OOH in plasma, and to the enzymatic conversion of PC-OOH to CE-OOH by lecithin:cholesterol acyltransferase in high density lipoproteins. The regioisomeric distribution of CE-O(O)H in human plasma indicates that free radical-mediated chain oxidation is an ongoing process, even in healthy young individuals.  相似文献   

19.
We studied the relationship between the site of production of oxygen radicals and their effect on a rat thymocyte functional activity, the glucose transport, measured using a radioactive analogue of glucose, 2-deoxy-glucose. We compared the effects of a hydrophilic thermolabile azo compound, mimicking a radical attack outside the cell, with the lipid-soluble cumene hydroperoxide, which initiates lipid peroxidation in cell membranes. Our results show that a low grade oxidative stress stimulated glucose uptake rapidly, independently of the site of radical generation. In the presence of the azocompound, glucose uptake increased smoothly, attaining its maximum extent within 1 h. In thymocytes treated with cumene hydroperoxide the rate of glucose transport increased suddenly and remained constant over 1 h. The effects of the radical donors on TBARS production and protein sulfhydryl groups content were also evaluated. In thymocytes treated with the azo derivative no lipid peroxidation was observed, but a slow decrease of protein thiol groups occurred; after the addition of cumene hydroperoxide sulfhydryl groups did not change and TBARS increased significantly. The water-soluble antioxidant Trolox was able to remove the glucose uptake increase induced by the hydrophilic initiator and to delay the loss of membrane integrity.  相似文献   

20.
[1-14C]Arachidonic acid was incubated with homogenates of the fungus, Saprolegnia parasitica. The products consisted of comparable amounts of two epoxy alcohols, 15-Ls-hydroxy-11,12-epoxy-5cis,8cis,13trans- eicosatrienoic acid and 15-hydroxy-13,14-epoxy-5cis,8cis,11cis-eicosatrienoic acid. Results of incubations carried out in the presence of nordihydroguaiaretic acid, 5,8,11,14-eicosatetraynoic acid, p-hydroxymercuribenzoate as well as glutathione peroxidase plus reduced glutathione demonstrated that transformation of arachidonic acid into epoxy alcohols occurred with the formation of 15-Ls-hydroperoxy-5cis,8cis,11cis,13trans- eicosatetraenoic acid (15-HPETE) as an intermediate. The pathway involved a lipoxygenase catalyzing the oxygenation of arachidonic acid at the 15L position to produce 15-HPETE, and a hydroperoxide isomerase activity which catalyzed conversion of 15-HPETE into the two epoxy alcohols. Studies with 15-[18O2]HPETE demonstrated that both oxygens of 15-HPETE were retained in the epoxy alcohols. Furthermore, experiments with mixtures of 15-[18O2]-and 15-[16O2]HPETE showed that conversion of 15-HPETE into epoxy alcohols occurred by an intramolecular transfer of hydroperoxide oxygen.  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