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.
1,3-Butadiene was oxidized by human myeloperoxidase in the absence of KCl to yield butadiene monoxide (BM) and crotonaldehyde (CA), but at KCl concentrations higher than 50 mM, 1-chloro-2-hydroxy-3-butene (CHB) was the major metabolite detected; metabolite formation was dependent on incubation time, pH, KCl, 1,3-butadiene, and H2O2 concentrations. The data are best explained by 1,3-butadiene being oxidized by myeloperoxidase by two different mechanisms. First, oxygen transfer from the hemoprotein would occur to either C-1 or C-4 of 1,3-butadiene to form an intermediate which may cyclize to form BM or undergo a hydrogen shift to form 3-butenal, an unstable precursor of CA. Further evidence for this mechanism was provided by the inability to detect methyl vinyl ketone, a possible product of an oxygen transfer reaction to C-2 or C-3 of 1,3-butadiene, and by the finding that CA was not simply a decomposition product of BM under assay conditions. In the second mechanism, however, chloride ion is oxidized by myeloperoxidase to HOCl which reacts with 1,3-butadiene to yield CHB. Further evidence for this mechanism was provided by the finding that CHB was readily formed when 1,3-butadiene was added to the filtrate of a myeloperoxidase/H2O2/KCl incubation and when 1,3-butadiene was allowed to react with authentic HOCl. In addition, CHB was not detected when BM or CA was incubated with myeloperoxidase, H2O2, and KCl for up to 60 min, or when 1,3-butadiene and KCl were incubated with chloroperoxidase and H2O2 or with mouse liver microsomes and NADPH, enzyme systems which catalyze 1,3-butadiene oxidation to BM and CA, but unlike myeloperoxidase, do not catalyze chloride ion oxidation to HOCl. These results provide clear evidence for novel olefinic oxidation reactions by myeloperoxidase.  相似文献   

2.
The heme in rat liver microsomal cytochrome P-450 was labeled with 14C or 3H and the microsomes were fractionated after in vitro incubations with a variety of agents known to destroy cytochrome P-450 heme. A major fraction of the heme label was irreversibly bound to apoprotein in all cases, including incubations with fluroxene, 1-octene, vinyl bromide, trichloroethylene, vinyl chloride, parathion, cumene hydroperoxide, NaN3, or iron-ADP complex. Label was also extensively bound to apoprotein when purified and reconstituted cytochrome P-450 was incubated with NADPH and vinyl chloride. This process appears to be widespread and involved to a significant extent in the cytochrome P-450 heme destruction observed with many compounds.  相似文献   

3.
Chloroperoxidase and H2O2 oxidize styrene to styrene oxide and phenylacetaldehyde but not benzaldehyde. The epoxide oxygen is shown by studies with H2(18)O2 to derive quantitatively from the peroxide. The epoxidation of trans-[1-2H]styrene by chloroperoxidase proceeds without detectable loss of stereochemistry, as does the epoxidation of styrene by rat liver cytochrome P-450, although much more phenylacetaldehyde is produced by chloroperoxidase than cytochrome P-450. Chloroperoxidase and cytochrome P-450 thus oxidize styrene by closely related oxygen-transfer mechanisms. Horseradish peroxidase does not oxidize styrene but does oxidize 2,4,6-trimethylphenol to 2,6-dimethyl-4-hydroxymethylphenol. The new hydroxyl group is partially labeled in incubations with H2(18)O but not H2(18)O2. The hydroxyl group thus appears to be introduced by addition of oxygen to the benzylic radical and water to the quinone methide intermediate but not by a cytochrome P-450-like oxene transfer mechanism. The results support the thesis that substrates primarily or exclusively react with the heme edge of horseradish peroxidase but are able to react with the ferryl oxygen of chloroperoxidase.  相似文献   

4.
Glycerol can be oxidized by rat liver microsomes to formaldehyde in a reaction that requires the production of reactive oxygen intermediates. Studies with inhibitors, antibodies, and reconstituted systems with purified cytochrome P4502E1 were carried out to evaluate whether P450 was required for glycerol oxidation. A purified system containing phospholipid, NADPH-cytochrome P450 reductase, P4502E1, and NADPH oxidized glycerol to formaldehyde. Formaldehyde production was dependent on NADPH, reductase, and P450, but not phospholipid. Formaldehyde production was inhibited by substrates and ligands for P4502E1, as well as by anti-pyrazole P4502E1 IgG. The oxidation of glycerol by the reconstituted system was sensitive to catalase, desferrioxamine, and EDTA but not to superoxide dismutase or mannitol, indicating a role for H2O2 plus non-heme iron, but not superoxide or hydroxyl radical in the overall glycerol oxidation pathway. The requirement for reactive oxygen intermediates for glycerol oxidation is in contrast to the oxidation of typical substrates for P450. In microsomes from pyrazole-treated, but not phenobarbital-treated rats, glycerol oxidation was inhibited by anti-pyrazole P450 IgG, anti-hamster ethanol-induced P450 IgG, and monoclonal antibody to ethanol-induced P450, although to a lesser extent than inhibition of dimethylnitrosamine oxidation. Anti-rabbit P4503a IgG did not inhibit glycerol oxidation at concentrations that inhibited oxidation of dimethylnitrosamine. Inhibition of glycerol oxidation by antibodies and by aminotriazole and miconazole was closely associated with inhibition of H2O2 production. These results indicate that P450 is required for glycerol oxidation to formaldehyde; however, glycerol is not a direct substrate for oxidation to formaldehyde by P450 but is a substrate for an oxidant derived from interaction of iron with H2O2 generated by cytochrome P450.  相似文献   

5.
In the presence of hepatic microsomes, vinyl chloride produces a ‘type I’ difference spectrum and stimulates carbon monoxide inhibitable NADPH consumption. A comparison of the binding and Michaelis parameters for the interaction of vinyl chloride with uninduced, phenobarbital and 3-methylcholanthrene induced microsomes indicates that the binding and metabolism of vinyl chloride is catalyzed by more than one type P-450 cytochrome, but predominantly by cytochrome P-450. Metabolites of vinyl chloride from this enzyme system decrease the levels of cytochrome P-450 and microsomal heme, but not cytochrome b5 or NADPH-cytochrome c reductase in vitro.  相似文献   

6.
The stoichiometry of NADPH oxidation in rabbit liver microsomes was studied. It was shown that in uncoupled reactions cytochrome P-450, besides O2- generation catalyzes direct two- and four-electron reduction of O2 to produce H2O2 and water, respectively. With an increase in pH and ionic strength, the amount of O2 reduced via an one-electron route increases at the expense of the two-electron reaction. In parallel, with a rise in pH the steady-state concentration of the oxy-complex of cytochrome P-450 increases, while the synergism of NADPH and NADH action in the H2O2 formation reaction is replaced by competition. The four-electron reduction is markedly accelerated and becomes the main pathway of O2 reduction in the presence of a pseudo-substrate--perfluorohexane. Treatment of rabbit with phenobarbital, which induces the cytochrome P-450 isozyme specific to benzphetamine results in a 2-fold increase in the degree of coupling of NADPH and benzphetamine oxidation. The experimental results suggest that the ratio of reactions of one- and two-electron reduction of O2 is controlled by the ratio of rates of one- and two-electron reduction of cytochrome P-450. In the presence of pseudo-substrates cytochrome P-450 acts predominantly as a four-electron oxidase; one of possible reasons for the uncoupling of microsomal monooxygenase reactions is the multiplicity of cytochrome P-450 isozymes.  相似文献   

7.
The cytochrome P450 enzyme systems catalyze the metabolism of a wide variety of naturally occurring and foreign compounds by reactions requiring NADPH and O2. Cytochrome P450 also catalyzes peroxide-dependent hydroxylation of substrates in the absence of NADPH and O2. Peroxidases such as chloroperoxidase and horseradish peroxidase catalyze peroxide-dependent reactions similar to those catalyzed by cytochrome P450. The kinetic and chemical mechanisms of the NADPH and O2-supported dealkylation reactions catalyzed by P450 have been investigated and compared with those catalyzed by P450 and peroxidases when the reactions are supported by peroxides. Detailed kinetic studies demonstrated that chloroperoxidase- and horseradish peroxidase-catalyzed N-demethylations proceed by a Ping Pong Bi Bi mechanism whereas P450-catalyzed O-dealkylations proceed by sequential mechanisms. Intramolecular isotope effect studies demonstrated that N-demethylations catalyzed by P450s and peroxidases proceed by different mechanisms. Most hemeproteins investigated catalyzed these reactions via abstraction of an alpha-carbon hydrogen whereas reactions catalyzed by P-450 and chloroperoxidase proceeded via an initial one-electron oxidation followed by alpha-carbon deprotonation. 18O-Labeling studies of the metabolism of NMC also demonstrated differences between the peroxidases and P450s. Because the hemeprotein prosthetic groups of P450, chloroperoxidase, and horseradish peroxidase are identical, the differences in the catalytic mechanisms result from differences in the environments provided by the proteins for the heme active site. It is suggested that the axial heme-iron thiolate moiety in P450 and chloroperoxidase may play a critical role in determining the mechanism of N-demethylation reactions catalyzed by these proteins.  相似文献   

8.
Furanoterpenes are widely distributed in the plant kingdom. In this study we have carried out enzymatic synthesis of simple furan compounds from the molecules containing an alpha-isopropylidene ketone unit and the role of cytochrome P450 in this biotransformation has been conclusively established. Eight model compounds (acyclic, monocyclic, and bicyclic, 1-8), having an alpha-isopropylidene ketone unit, were synthesized and incubated with PB-induced rat liver microsomes in the presence of NADPH and O(2). GC-MS and NMR analyses of the product(s) indicated the formation of corresponding furano derivatives (11-18). Cytochrome P450 inhibitors, metyrapone, SKF-525, and carbon monoxide, inhibited the formation of furan (8) to 76, 62, and 97%, respectively. Incubation of dehydrofukinone (7), a naturally occurring terpene, with purified cytochrome P450, NADPH-cytochrome P450 reductase, and dilaurylphosphatidylcholine in the presence of NADPH and O(2) resulted in the formation of 10 and furanodehydrofukinone (19). Based on these observations, we propose one of the probable biosynthetic routes for lower furanoterpenoids in higher plants.  相似文献   

9.
Vanadate-dependent NAD(P)H oxidation, catalyzed by rat liver microsomes and microsomal NADPH-cytochrome P450 reductase (P450 reductase) and NADH-cytochrome b5 reductase (b5 reductase), was investigated. These enzymes and intact microsomes catalyzed NAD(P)H oxidation in the presence of either ortho- or polyvanadate. Antibody to P450 reductase inhibited orthovanadate-dependent NADPH oxidation catalyzed by either purified P450 reductase or rat liver microsomes and had no effect on the rates of NADH oxidation catalyzed by b5 reductase. NADPH-cytochrome P450 reductase catalyzed orthovanadate-dependent NADPH oxidation five times faster than NADH-cytochrome b5 reductase catalyzed NADH oxidation. Orthovanadate-dependent oxidation of either NADPH or NADH, catalyzed by purified reductases or rat liver microsomes, occurred in an anaerobic system, which indicated that superoxide is not an obligate intermediate in this process. Superoxide dismutase (SOD) inhibited orthovanadate, but not polyvanadate-mediated, enzyme-dependent NAD(P)H oxidation. SOD also inhibited when pyridine nucleotide oxidation was conducted anaerobically, suggesting that SOD inhibits vanadate-dependent NAD(P)H oxidation by a mechanism independent of scavenging of O2-.  相似文献   

10.
Both rabbit liver microsomes and reconstituted system with purified cytochrome P-450 and cofactors enzymatically oxidized o-dithiane (1, 2-dithiane), 3-methyl-o-dithiane, thiane and 2-methylthiane to the corresponding mono-oxygenated products; sulfides or disulfides were oxidized to the corresponding sulfoxides or thiosulfinates, while thiosulfinate was oxidized to thiolsulfonate. The reconstituted systems required oxygen and NADPH and were not affected by the catalase which decomposes H2O2, or by 1,4-diazabicyclo-[2,2,2]octane (DABCO), which is a good quencher of singlet oxygen. The differences in the binding of substrates such as sulfides and disulfides with the enzyme system are discussed in connection with differences in the spectra of the substrates in the reconstituted system with pure cytochrome P-450. A correlation was found between the rates of oxidation of the substrates and the rates of oxidation of NADPH.  相似文献   

11.
The enzymatic oxidation of tetrachloro-1,4-hydroquinone (1,4-TCHQ), resulting in covalent binding to protein of tetrachloro-1,4-benzoquinone (1,4-TCBQ), was investigated, with special attention to the involvement of cytochrome P-450 and reactive oxygen species. 1,4-TCBQ itself reacted very rapidly and extensively with protein (58% of the 10 nmol added to 2 mg of protein, in a 5-min incubation). Ascorbic acid and glutathione prevented covalent binding of 1,4-TCBQ to protein, both when added directly and when formed from 1,4-TCHQ by microsomes. In microsomal incubations as well as in a reconstituted system containing purified cytochrome P-450b, 1,4-TCHQ oxidation and subsequent protein binding was shown to be completely dependent on NADPH. The reaction was to a large extent, but not completely, dependent on oxygen (83% decrease in binding under anaerobic conditions). Inhibition of cytochrome P-450 by metyrapone, which is also known to block the P-450-mediated formation of reactive oxygen species, gave a 80% decrease in binding, while the addition of superoxide dismutase prevented 75% of the covalent binding, almost the same amount as found in anerobic incubations. A large part of the conversion of 1,4-TCHQ to 1,4-TCBQ is apparently not catalyzed by cytochrome P-450 itself, but is mediated by superoxide anion formed by this enzyme. The involvement of this radical anion is also demonstrated by microsomal incubations without NADPH but including the xantine/xantine oxidase superoxide anion generating system. These incubations resulted in a 1.6-fold binding as compared to the binding in incubations with NADPH but without xantine/xantine oxidase. 1,4-TCHQ was shown to stimulate the oxidase activity of microsomal cytochrome P-450. It is thus not unlikely that 1,4-TCHQ enhances its own microsomal oxidation.  相似文献   

12.
Horseradish peroxidase rapidly catalyzed the H2O2-dependent polymerization of acetaminophen. Acetaminophen polymerization was decreased and formation of GSSG and minor amounts of GSH-acetaminophen conjugates were detected in reaction mixtures containing GSH. These data suggest that horseradish peroxidase catalyzed the 1-electron oxidation of acetaminophen and that GSH decreased polymerization by reducing the product, N-acetyl-p-benzosemiquinone imine, back to acetaminophen. Analyses of reaction mixtures that did not contain GSH showed N-acetyl-p-benzoquinone imine formation shortly after initiation of reactions. When GSH was added to similar reaction mixtures at various times, 3-(glutathion-S-yl)-acetaminophen was formed. The formation and disappearance of this product were very similar to N-acetyl-p-benzoquinone imine formation and were consistent with the disproportionation of 2 mol of N-acetyl-p-benzosemiquinone imine to 1 mol of N-acetyl-p-benzoquinone imine and 1 mol of acetaminophen followed by the rapid reaction of N-acetyl-p-benzoquinone imine with GSH to form 3-(glutathion-S-yl)acetaminophen. When acetaminophen was incubated with NADPH, oxygen and hepatic microsomes from phenobarbital-pretreated rats, 1.2 nmol 3-(glutathion-S-yl)acetaminophen/nmol cytochrome P-450/10 min was formed. Formation of polymers was not observed indicating that N-acetyl-p-benzoquinone imine was formed via an overall 2-electron oxidation rather than a disproportionation reaction. However, when cumene hydroperoxide was replaced by NADPH in microsomal incubations, polymerization was observed suggesting that cytochrome P-450 might also catalyze the 1-electron oxidation of acetaminophen.  相似文献   

13.
The interaction of heme nonapeptide (a proteolytic product of cytochrome c) with purified NADH:cytochrome b5 (EC 1.6.2.2) and NADPH:cytochrome P-450 (EC 1.6.2.4) reductases was investigated. In the presence of heme nonapeptide, NADH or NADPH were enzymatically oxidized to NAD+ and NADP+, respectively. NAD(P)H consumption was coupled to oxygen uptake in both enzyme reactions. In the presence of carbon monoxide the spectrum of a carboxyheme complex was observed during NAD(P)H oxidation, indicating the existence of a transient ferroheme peptide. NAD(P)H oxidation could be partially inhibited by cyanide, superoxide dismutase and catalase. Superoxide and peroxide ions (generated by enzymic xanthine oxidation) only oxidized NAD(P)H in the presence of heme nonapeptide. Oxidation of NAD(P)H was more rapid with O2- than O2-2. We suggest that a ferroheme-O2 and various heme-oxy radical complexes (mainly ferroheme-O-2 complex) play a crucial role in NAD(P)H oxidation.  相似文献   

14.
The kinetics of chromate reduction by liver microsomes isolated from rats pretreated with phenobarbital or 3-methylcholanthrene with NADPH or NADH cofactor have been followed. Induction of cytochrome P-450 and NADPH-cytochrome P-450 reductase activity in microsomes by phenobarbital pretreatment caused a decrease in the apparent chromate-enzyme dissociation constant, Km, and an increase in the apparent second-order rate constant, kcat/Km, but did not affect the kcat of NADPH-mediated microsomal metabolism of chromate. Induction of cytochrome P-448 in microsomes by 3-methylcholanthrene pretreatment did not affect the kinetics of NADPH-mediated reduction of chromate by microsomes. The kinetics of NADH-mediated microsomal chromate reduction were unaffected by the drug treatments. The effects of specific enzyme inhibitors on the kinetics of microsomal chromate reduction have been determined. 2'-AMP and 3-pyridinealdehyde-NAD, inhibitors of NADPH-cytochrome P-450 reductase and NADH-cytochrome b5 reductase, inhibited the rate of microsomal reduction of chromate with NADPH and NADH. Metyrapone and carbon monoxide, specific inhibitors of cytochrome P-450, inhibited the rate of NADPH-mediated microsomal reduction of chromate, whereas high concentrations of dimethyl-sulfoxide (0.5 M) enhanced the rate. These results suggest that the electron-transport cytochrome P-450 system is involved in the reduction of chromate by microsomal systems. The NADPH and NADH cofactors supply reducing equivalents ultimately to cytochrome P-450 which functions as a reductase in chromate metabolism. The lower oxidation state(s) produced upon chromate reduction may represent the ultimate carcinogenic form(s) of chromium. These studies provide evidence for the role of cytochrome P-450 in the activation of inorganic carcinogens.  相似文献   

15.
The nitrosoureas, CCNU (1-(2-chloroethyl)-3-(cyclohexyl)-1-nitrosourea) and BCNU (1,3-bis(2-chloroethyl)-1-nitrosourea) are representatives of a class of N-nitroso compounds which undergo denitrosation in the presence of NAD(P)H and deoxygenated hepatic microsomes from rats to yield nitric oxide (NO) and the denitrosated parent compound. Formation of NO during microsomal denitrosation of CCNU and BCNU was determined by three methods. With one procedure, NO was measured and concentration shown to increase over time in the head gas above microsomal incubations with BCNU. Two additional methods utilized NO binding to either ferrous cytochrome P-450 or hemoglobin to form distinct Soret maxima at 444 and 415 nm, respectively. Incubation of either BCNU or CCNU in the presence of NAD(P)H and deoxygenated microsomes resulted in the formation of identical cytochrome P-450 ferrous · NO optical difference spectra. Determination of the P-450 ferrous · NO extinction coefficient by the change in absorbance at 444 minus 500 nm allowed measurement of rates of denitrosation by monitoring the increase in absorbance at 444 nm. The rates of BCNU and CCNU denitrosation were determined to be 4.8 and 2.0 nmol NO/min/mg protein, respectively, for phenobarbital (PB) induced microsomes. For the purpose of comparison, the rate of [14C]CCNU (1-(2-[14C]chloroethyl)-3-(cyclohexyl)-1-nitrosourea turnover was examined by the isolation of [14C]CCU (1-(2-[14C] chloroethyl)-3-(cyclohexyl)-1-urea) from incubations that contained NADPH and deoxygenated PB-induced microsomes. These analyses showed stoichiometric amounts of NO and [14C]CCU being formed at a rate of 2.0 nmol/min/mg protein. Denitrosation catalysis by microsomes was enhanced by phenobarbital pretreatment and partially decreased by cytochrome P-450 inhibitors, SKF-525A, α-naphthoflavone (ANF), metyrapone, and CO, suggesting a cytochrome P-450-dependent denitrosation. However, in the presence of NADPH and purified NADPH cytochrome P-450 reductase reconstituted in dilauroylphosphatidylcholine, [14C]CCNU was shown to undergo denitrosation to [14C]CCU. Thus, NADPH cytochrome P-450 reductase could support denitrosation in the absence of cytochrome P-450.  相似文献   

16.
We have isolated a protein factor from rat liver which stimulates anthranilamide hydroxylation by the microsomes in the presence of NADPH and oxygen and showed this factor to contain Cu and Zn and to have superoxide dismutase activity [Biochim. Biophys. Acta 365, 148-157 (1974)]. In the present study, this protein factor was confirmed to be a superoxide dismutase (SOD) by comparison of the recovery of SOD activity with that of anthranilamide hydroxylation-stimulating activity at each step of its purification, by inhibition of SOD activity with NaCN and hydrogen peroxide (H2O2), and by recovery of the SOD activity of the protein factor after reconstitution with Cu2+ and/or Zn2+. At a given SOD activity level, there was no difference among the rat liver SOD, Cu,Zn-SOD from bovine erythrocytes, and Mn-SOD from Serratia marcescens in their ability to stimulate anthranilamide hydroxylation not only by rat liver microsomes, but also by the reconstituted cytochrome P-450-containing monooxygenase system. Rat liver SOD stimulated anthranilamide hydroxylation by the reconstituted system in proportion to its amount below a protein concentration of 1 microgram/ml. In anthranilamide hydroxylation by the reconstituted system without SOD, only a slight hydroxylase activity was found at the initial stage of the reaction and a marked increase in the amounts of NADPH oxidized and H2O2 formed was observed after a lag time. In the presence of rat liver SOD, however, the hydroxylase activity was markedly and continuously increased almost proportionally to reaction time with a concomitant decrease in the amounts of NADPH oxidized and H2O2 formed. In addition, a trace of 3-OH anthranilamide, one of the products, not only stimulated NADPH-dependent H2O2 formation in the reconstituted system, but also inhibited the apparent reduction of cytochrome P-450 by NADPH in the reconstituted system. These effects of 3-OH anthranilamide were diminished by rat liver SOD. When a trace of 3-OH anthranilamide were added to a system composed of NADPH-cytochrome c (P-450) reductase and NADPH, H2O2 formation and NADPH oxidation were markedly stimulated. However, on addition of 3-OH anthranilamide to the system containing rat liver SOD, no stimulation on either H2O2 formation or NADPH oxidation was found.(ABSTRACT TRUNCATED AT 400 WORDS)  相似文献   

17.
2-Hydroxyestradiol, 2-hydroxyestrone and 2-hydroxy-17α-ethynylestradiol, oxidation products of naturally occurring estrogens and synthetic estrogens in some oral contraceptives were found to be converted by rat liver microsomes to reactive metabolites that become irreversibly bound to microsomal protein. The irreversible binding required microsomes, oxygen and NADPH. The NADPH could be replaced by a xanthine-xanthine oxidase system which is known to generate superoxide anions. The irreversible binding was substantially inhibited by superoxide dismutase, 30% in those incubations containing NADPH and 98% in those incubations containing the xanthine-xanthine oxidase system. Further studies with 2-hydroxyestradiol showed that microsomal cytochrome P-450 was rate limiting in the NADPH-dependent irreversible binding, because the binding was inhibited 62% by an antibody against NADPH-cytochrome c reductase and 70% in an atmosphere of CO:O2 (9:1) when compared to an atmosphere of N2:O2 (9:1). Phenobarbital, a known inducer of cytochrome P-450, had no effect on the irreversible binding of 2-hydroxyestradiol, whereas another inducer of P-450, pregnenolone-16α-carbonitrile, markedly increased the irreversible binding. In contrast, cobaltous chloride, an inhibitor of the synthesis of cytochrome P-450, decreased both P-450 and the irreversible binding. These results are consistent with a mechanism for irreversible binding of estrogens and 2-hydroxyestrogens to microsomes that requires oxidation of the catechol nucleus by cytochrome P-450-generated superoxide anion.  相似文献   

18.
Glycerol can be oxidized to formaldehyde by rat liver microsomes and by cytochrome P450. The ability of other alcohols to be oxidized to formaldehyde was determined to evaluate the structural determinants of the alcohol which eventually lead to this production of formaldehyde. Monohydroxylated alcohols such as 1- or 2-propanol did not produce formaldehyde when incubated with NADPH and microsomes. Geminal diols such as 1,3-propanediol, 1,3-butanediol, or 1,4-butanediol also did not yield formaldehyde. However, vicinal diols such as 1,2-propanediol or 1,2-butanediol produced formaldehyde. With 1,2-propanediol, the residual two-carbon fragment was found to be acetaldehyde, while with 1,2-butanediol, the residual three-carbon fragment was propionaldehyde. Oxidation of 1,2-propanediol to formaldehyde plus acetaldehyde involved interaction with an oxidant derived from H2O2 plus nonheme iron, since production of the two aldehydic products was completely prevented by catalase or glutathione plus glutathione peroxidase and by chelators such as desferrioxamine or EDTA. The oxidant was not superoxide or hydroxyl radical. Product formation was fivefold lower when NADH replaced NADPH, and was inhibited by substrates, ligands, and inhibitors of cytochrome P450. A charged glycol such as alpha-glycerophosphate (but not the geminal beta-glycerophosphate) was readily oxidized to formaldehyde, suggesting that interaction of the glycol with the oxidant was occurring in solution and not in a hydrophobic environment. These results indicate that the carbon-carbon bond between 1,2-glycols can be cleaved by an oxidant derived from microsomal generated H2O2 and reduction of non-heme iron, with the subsequent production of formaldehyde plus an aldehyde with one less carbon than the initial glycol substrate.  相似文献   

19.
Superoxide generation, assessed as the rate of acetylated cytochrome c reduction inhibited by superoxide dismutase, by purified NADPH cytochrome P-450 reductase or intact rat liver microsomes was found to account for only a small fraction of their respective NADPH oxidase activities. DTPA-Fe3+ and EDTA-FE3+ greatly stimulated NADPH oxidation, acetylated cytochrome c reduction, and O(2) production by the reductase and intact microsomes. In contrast, all ferric chelates tested caused modest inhibition of acetylated cytochrome c reduction and O(2) generation by xanthine oxidase. Although both EDTA-Fe3+ and DTPA-Fe3+ were directly reduced by the reductase under anaerobic conditions, ADP-Fe3+ was not reduced by the reductase under aerobic or anaerobic conditions. Desferrioxamine-Fe3+ was unique among the chelates tested in that it was a relatively inert iron chelate in these assays, having only minor effects on NADPH oxidation and/or O(2) generation by the purified reductase, intact microsomes, or xanthine oxidase. Desferrioxamine inhibited microsomal lipid peroxidation promoted by ADP-Fe3+ in a concentration-dependent fashion, with complete inhibition occurring at a concentration equal to that of exogenously added ferric iron. The participation of O(2) generated by the reductase in NADPH-dependent lipid peroxidation was also investigated and compared with results obtained with a xanthine oxidase-dependent lipid peroxidation system. NADPH-dependent peroxidation of either phospholipid liposomes or rat liver microsomes in the presence of ADP-Fe3+ was demonstrated to be independent of O(2) generation by the reductase.  相似文献   

20.
Studying liver microsomes from 2,3,7,8-tetrachlorodibenzo-p-dioxin (TCDD)-induced or vehicle-treated (noninduced) mice, we evaluated the in vitro effects of added chemicals on the production of reactive oxygen due to substrate/P450-mediated uncoupling. The catalase-inhibited NADPH-dependent H(2)O(2) production (luminol assay) was lower in induced than noninduced microsomes. The effects of adding chemicals (2.5 microM) in vitro could be divided into three categories: Group 1, highly halogenated and coplanar compounds that increased H(2)O(2) production at least 5-fold in induced, but not in noninduced, microsomes; Group 2, non-coplanar halogenated biphenyls that did not affect H(2)O(2) production; Group 3, minimally halogenated biphenyls and benzo[a]pyrene that decreased H(2)O(2) production. Molar consumption of NADPH and O(2) and molar H(2)O(2) production (o-dianisidine oxidation) revealed that Group 1 compounds mostly increased, Group 2 had no effect, and Group 3 decreased the H(2)O(2)/O(2) and H(2)O(2)/NADPH ratios. Microsomal lipid peroxidation (thiobarbituric acid-reactive substances) was proportional to H(2)O(2) production. Although TCDD induction decreased microsomal production of H(2)O(2), addition of Group 1 compounds to TCDD-induced microsomes in vitro stimulated the second-electron reduction of cytochrome P450 and subsequent release of H(2)O(2) production. This pathway is likely to contribute to the oxidative stress response and associated toxicity produced by many of these environmental chemicals.  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