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.
NADPH-dependent oxidation of 1,3-butadiene by mouse liver microsomes or H2O2-dependent oxidation by chloroperoxidase produced both butadiene monoxide and crotonaldehyde; methyl vinyl ketone and 2,3- and 2,5- dihydrofuran were not detected. The crotonaldehyde to butadiene monoxide ratio remained constant over time in both the microsomal and the chloroperoxidase reactions; however, much more crotonaldehyde was produced by chloroperoxidase than microsomes; crotonaldehyde was not detected when reference samples of butadiene monoxide were used in control incubations containing NADPH and microsomes or H2O2 and chloroperoxidase. Moreover, incubations of 1,3-butadiene with horseradish peroxidase and H2O2, or microsomes and H2O2 or arachidonic acid did not result in the oxidation of 1,3-butadiene. In microsomes, metabolite formation was dependent on incubation time, NADPH, and protein concentrations and did not change when the 1,3-butadiene pressure was varied between 24 and 52 cm Hg. Inclusion of the cytochrome P450 inhibitor 1-benzylimidazole inhibited 1,3-butadiene metabolism, but inclusion of KCN, catalase, or superoxide dismutase had no effect. These results support the role of cytochrome P450 in 1,3-butadiene oxidation by mouse liver microsomes. The formation of crotonaldehyde but not methyl vinyl ketone by cytochrome P450 or chloroperoxidase indicates regioselectivity in the oxygen transfer from the hemoproteins to 1,3-butadiene. The intermediates formed may undergo either ring closure to form butadiene monoxide or a hydrogen shift to form 3-butenal which tautomerizes to produce crotonaldehyde. Evidence for this tautomerization was obtained by the finding that 3-buten-1-ol, an alternative precursor of 3-butenal, was oxidized to crotonaldehyde under incubation conditions similar to that used for 1,3-butadiene.  相似文献   

2.
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   

3.
Methemoglobin and metmyoglobin catalyze the H2O2-dependent oxidation of styrene to styrene oxide and benzaldehyde. The formation of styrene oxide requires molecular oxygen as well as H2O2 but does not, as shown by inhibitor studies, involve the superoxide or hydroxyl radicals. Approximately 38, 67, and 78% of the oxygen in styrene oxide derives from 18O2 in the reactions catalyzed, respectively, by bovine hemoglobin, sperm whale myoglobin, and equine heart myoglobin, whereas 70, 55, and 35% of the oxygen can be shown to be derived from [18O]H2O2. However, a larger fraction of the epoxide oxygen than suggested by the labeling data (perhaps all) derives from molecular oxygen rather than H2O2 because the hemoproteins produce molecular oxygen from the peroxide. The epoxidation of styrene by methemoglobin gives equal amounts of the R and S enantiomers and, as shown by studies with trans-[1-2H]styrene, proceeds with partial (33%) loss of the olefin stereochemistry. The results are rationalized by H2O2-dependent formation of a protein radical that combines with molecular oxygen to give a protein-peroxy radical that oxidizes styrene.  相似文献   

4.
Identification of the fifth axial heme ligand of chloroperoxidase   总被引:1,自引:0,他引:1  
Chloroperoxidase from Caldariomyces fumago catalyzes the peroxidative chlorination of organic acceptor molecules. From a variety of spectroscopic data, it had long been thought that chloroperoxidase possessed an active site structure similar to that of cytochrome P-450cam. Resonance Raman studies conducted with isotopically substituted enzyme proved conclusively that the fifth axial ligand to the ferriprotoporphyrin IX moiety of chloroperoxidase is indeed a cysteine thiolate (Bangcharoenpaurpong, O., Champion, P. M., Hall, K. S., and Hager, L. P. (1986) Biochemistry 25, 2374-2378). In this study, Ellman's reagent, 5,5'-dithiobis(2-nitrobenzoic acid), was used to ascertain which of the 3 cysteine residues in the primary structure of chloroperoxidase serves as the fifth axial heme ligand; two of the cysteine residues were earlier shown to be involved in a disulfide linkage. Apoprotein was labeled under denaturing conditions with 5,5'-dithiobis(2-nitrobenzoic acid). A unique peptide, containing the thionitrobenzoate adduct, was isolated via reverse phase HPLC following digestion with endoproteinase Glu-C. Amino acid and Edman sequence analysis revealed the fifth axial ligand in chloroperoxidase to be cysteine 29. Under reducing and denaturing conditions, incubation of apochloroperoxidase with Ellman's reagent resulted in 3 labeled residues. Proteolysis and isolation of the labeled peptides using reverse phase HPLC and subsequent Edman sequence analysis detected and identified the thionitrobenzoate adducts of each of the three cysteinyl peptides of chloroperoxidase.  相似文献   

5.
Direct infusion electrospray ionization triple quadrupole precursor scanning for three oxidized fatty acyl anions revealed 86 mass spectral peaks representing polar membrane lipids in extracts from Arabidopsis (Arabidopsis thaliana) infected with Pseudomonas syringae pv tomato DC3000 expressing AvrRpt2 (PstAvr). Quadrupole time-of-flight and Fourier transform ion cyclotron resonance mass spectrometry provided evidence for the presence of membrane lipids containing one or more oxidized acyl chains. The membrane lipids included molecular species of phosphatidylcholine, phosphatidylethanolamine, phosphatidylglycerol, digalactosyldiacylglycerol, monogalactosyldiacylglycerol, and acylated monogalactosyldiacylglycerol. The oxidized chains were identified at the level of chemical formula and included C(18)H(27)O(3) (abbreviated 18:4-O, to indicate four double bond equivalents and one oxygen beyond the carbonyl group), C(18)H(29)O(3) (18:3-O), C(18)H(31)O(3) (18:2-O), C(18)H(29)O(4) (18:3-2O), C(18)H(31)O(4) (18:2-2O), and C(16)H(23)O(3) (16:4-O). Mass spectral signals from the polar oxidized lipid (ox-lipid) species were quantified in extracts of Arabidopsis leaves subjected to wounding, infection by PstAvr, infection by a virulent strain of P. syringae, and low temperature. Ox-lipids produced low amounts of mass spectral signal, 0.1% to 3.2% as much as obtained in typical direct infusion profiling of normal-chain membrane lipids of the same classes. Analysis of the oxidized membrane lipid species and normal-chain phosphatidic acids indicated that stress-induced ox-lipid composition differs from the basal ox-lipid composition. Additionally, different stresses result in the production of varied amounts, different timing, and different compositional patterns of stress-induced membrane lipids. These data form the basis for a working hypothesis that the stress-specific signatures of ox-lipids, like those of oxylipins, are indicative of their functions.  相似文献   

6.
The mechanism by which the hepatic cytochrome P-450 (Cyt. P-450) containing mixed-function oxidase system oxidizes the analgesic drug paracetamol (PAR) to a hepatotoxic metabolite was studied. Since previous studies excluded the possibility of oxygenation of PAR, three other mechanisms, namely direct 1-electron oxidation by a Cyt. P-450-ferrous-dioxygen complex under concomitant formation of H2O2 to N-acetyl-p-semiquinone imine (NAPSQI), direct 2-electron oxidation by a Cyt. P-450-ferric-oxene complex to N-acetyl-p-benzoquinone imine (NAPQI) and indirect oxidation by active oxygen species released from Cyt. P-450, were considered. Indirect oxidation by active oxygen species was not involved, as active oxygen scavengers such as superoxide dismutase, catalase and DMSO did not affect the oxidation of PAR in hepatic microsomes. No reaction products characteristic for a direct 1-electron oxidation of PAR by Cyt. P-450 were observed: neither NAPSQI radical formation was detectable by ESR, nor PAR-dimer formation, nor stimulation of the microsomal H2O2 production was found to occur. In fact, PAR inhibited the spontaneous microsomal H2O2 formation. Studies on the reactions of NAPSQI with glutathione (GSH) revealed that NAPSQI hardly conjugated with GSH to a 3-glutathionyl-paracetamol conjugate (PAR-GSH) conjugate. The reactions of the elusive reactive metabolite formed during microsomal oxidation of PAR in the presence of GSH closely resembled those of synthetic NAPQI: both PAR-GSH and oxidized glutathione (GSSG) formation occurred. Furthermore, in agreement with a 2-electron oxidation hypothesis, iodosobenzene-dependent oxidation of PAR by cyt. P-450 in the presence of GSH resulted in the formation of the PAR-GSH conjugate. It is concluded that bioactivation of PAR by the Cyt. P-450 containing mixed-function oxidase system consists of a direct 2-electron oxidation to NAPQI.  相似文献   

7.
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   

8.
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   

9.
Cytochrome P450 enzymes promote a number of oxidative biotransformations including the hydroxylation of unactivated hydrocarbons. Whereas the long-standing consensus view of the P450 mechanism implicates a high-valent iron-oxene species as the predominant oxidant in the radicalar hydrogen abstraction/oxygen rebound pathway, more recent studies on isotope partitioning, product rearrangements with 'radical clocks', and the impact of threonine mutagenesis in P450s on hydroxylation rates support the notion of the nucleophilic and/or electrophilic (hydro)peroxo-iron intermediate(s) to be operative in P450 catalysis in addition to the electrophilic oxenoid-iron entity; this may contribute to the remarkable versatility of P450s in substrate modification. Precedent to this mechanistic concept is given by studies with natural and synthetic P450 biomimics. While the concept of an alternative electrophilic oxidant necessitates C-H hydroxylation to be brought about by a cationic insertion process, recent calculations employing density functional theory favour a 'two-state reactivity' scenario, implicating the usual ferryl-dependent oxygen rebound pathway to proceed via two spin states (doublet and quartet); state crossing is thought to be associated with either an insertion or a radicalar mechanism. Hence, challenge to future strategies should be to fold the disparate and sometimes contradictory data into a harmonized overall picture.  相似文献   

10.
A radical species of monochlorodimedone has been characterized by its high reactivity with molecular O2. Horseradish peroxidase greatly accelerated O2 uptake by acidic solutions of this substrate; the enzymatic reaction required exogenous H2O2 only with freshly prepared substrate solutions, and the total substrate oxidized was equal to the sum of H2O2 added and O2 consumed. However, with excess Br- and horseradish peroxidase, or high Br- or Cl- and chloroperoxidase, a 1:1 stoichiometry between H2O2 and substrate was observed. In the absence of halide, the stoichiometry of the chloroperoxidase-catalyzed oxidation of monochlorodimedone changed to two molecules of the organic donor per H2O2. Moreover, in the absence of halide, at substrate:H2O2 ratios greater than 2.0, chloroperoxidase catalyzed significant O2 uptake; this enzyme-dependent autoxidation of monochlorodimedone also occurred in the presence of Cl- or Br-, when H2O2 was limiting. These data, and recent evidence from this laboratory for free hypohalous acid as the first product of chloroperoxidase-catalyzed halide oxidation [B. W. Griffin (1983) Biochem. Biophys. Res. Commun. 116, 873-879], strongly support a mixed enzymatic/nonenzymatic radical chain process as the mechanism for halogenation of monochlorodimedone by chloroperoxidase. Both horseradish peroxidase and chloroperoxidase can catalyze either bromination or oxidation of this substrate, depending on the experimental conditions. Implications of these results for the mechanism of HOCl formation catalyzed by chloroperoxidase are considered.  相似文献   

11.
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   

12.
The use of 5-deazaFAD T491V cytochrome P450 reductase has made it possible to directly measure the rate of electron transfer to microsomal oxyferrous cytochrome (cyt) P450 2B4. In this reductase the FMN moiety can be reduced to the hydroquinone, FMNH(2), while the 5-deazaFAD moiety remains oxidized [Zhang, H., et al. (2003) Biochemistry 42, 6804-6813]. The rate of electron transfer from 5-deazaFAD cyt P450 reductase to oxyferrous cyt P450 was determined by rapidly mixing the ferrous cyt P450-2-electron-reduced 5-deazaFAD T491V reductase complex with oxygen in the presence of substrate. The 5-deazaFAD T491V reductase which can only donate a single electron reduces the oxyferrous cyt P450 and oxidizes to the air-stable semiquinone, with rate constants of 8.4 and 0.37 s(-1) at 15 degrees C. Surprisingly, oxyferrous cyt P450 turns over more slowly with a rate constant of 0.09 s(-1), which is the rate of catalysis under steady-state conditions at 15 degrees C (k(cat) = 0.08 s(-1)). In contrast, the rate constant for electron transfer from ferrous cyt b(5) to oxyferrous cyt P450 is 10 s(-1) with oxyferrous cyt P450 and cyt b(5) simultaneously undergoing spectral changes. Quantitative analyses by LC-MS/MS revealed that the product, norbenzphetamine, was formed with a coupling efficiency of 52% with cyt b(5) and 32% with 5-deazaFAD T491V reductase. Collectively, these results suggest that during catalysis a relatively stable reduced oxyferrous intermediate of cyt P450 is formed in the presence of cyt P450 reductase but not cyt b(5) and that the rate-limiting step in catalysis follows introduction of the second electron.  相似文献   

13.
A unique feature of P450 enzymes is in the presence of a thiolate ligand heme but its exact function in catalysis is a matter of debate. For P450 dependent monooxygenases the "active oxygen" complex seems to exist only as a transition state in which the thiolate ligand provides electron density in order to prevent pi-backbonding of the oxygen to the iron (-S-Fe-O(z.rad;)). The corresponding ground state (Compound I) would be a ferryl species (Fe(IV)z.dbnd6;O) with an electron hole either at the porphyrin or at the sulfur. Apart from this role we postulate that a second function is related to the electronic structure of Compound II as an electron acceptor and this property is shared among monooxygenases, thromboxane synthase, prostacyclin synthase, allene oxide synthase, P450(NOR(-)) and chloroperoxidase. As a common step in all P450 enzymes an extremely rapid electron uptake by Compound II allows that the primary substrate radicals are oxidized to cations which immediately combine with a neighbouring nucleophile. Thus "electron transfer" may substitute for "oxygen rebound" as the final step leading to product formation. The same principle also applies methane monooxygenases in which the role of the thiyl sulfur is replaced by a ferryl-oxyl entity.  相似文献   

14.
Effects and mechanisms of H(2)O(2) on production of dicarboxylic acid.   总被引:5,自引:0,他引:5  
The system of producing long chain dicarboxylic acid (DCA) by Candida tropicalis is an aerobic and viscous fermentation system. A method to overcome the gas-liquid transport resistance and to increase oxygen supply is by adding hydrogen peroxide (H(2)O(2)) to the fermentation system. Here we report that the H(2)O(2) not only can enhance the oxygen supply but also change the metabolism by inducing cytochrome P450, the key enzyme of a, o-oxidation. When C. tropicalis was cultivated in a 3-L bioreactor using the combination of aeration and H(2)O(2) feeding, DCA production rates increased by about 10% after a short period of decrease at the beginning. Furthermore, the experiments showed that the maximum activities of P450 could be induced at 2 mM H(2)O(2), and the inducible mechanisms are also discussed. Moreover, we suggest that alkane might be oxidized through the "peroxide shunt pathway" when H(2)O(2) is present. By adding H(2)O(2), the DCA yield in a 22-L bioreactor could increase by 25.3% and reach 153.9 g/L.  相似文献   

15.
The chloroperoxidase-catalyzed reactions of NAD(P)H with H2O2 in the presence of Cl- or Br- have been characterized. With 1 mol H2O2 per mol of NADH, one atom of 36Cl was incorporated into the 264-nm-absorbing intermediate product. This species was oxidized enzymatically by a second mole of H2O2 to a species distinct from NAD+, which retained one Cl atom. Spectroscopically identical species were also produced by reaction of NADH with one and two molar ratios of HOCl, respectively. These data indicate that, with respect to halogenation activities, chloroperoxidase functions similarly to myeloperoxidase, i.e., produces HOCl as the first product of Cl- oxidation by H2O2. Moreover, rapid chlorination of NAD(P)H followed by oxidation may be an important and highly lethal microbicidal effect of HOCl produced by myeloperoxidase in activated neutrophils.  相似文献   

16.
Normal values of Bohr effect of oxygenation of partially oxidized hemoglobin A with ferrihemes liganded either with H2O and OH or with CN have been found in the range of pH values from 6.8 to 7.6 in 45 micrometer (Fe)-hemoglobin containing 36--38% of ferrihemes. As the changes of oxygen affinity of Hb A induced by changes of pH are due to the modifications of R state, this quaternary conformation is thought to be unchanged in the studied of R state, this quaternary conformation is thought to be unchanged in the studied forms of partially oxidized hemoglobin. It is suggested that interactions between ferric and ferrous hemes leading to the increased affinity of ferrous hemes to oxygen occur in deoxygenated form of partially oxidized hemoglobin. In partially oxidized hemoglobin with ferric hemes liganded with H2O asymmetry of oxygen binding curves has been noted, which is not observed in forms with ferric hemes liganded with OH ot CN. This shows the effect of ligands of ferric hemes on interactions between chains containing ferric and ferrous hemes.  相似文献   

17.
The manganese peroxidase (MnP), from the lignin-degrading fungus Phanerochaete chrysosporium, an H2O2-dependent heme enzyme, oxidizes a variety of organic compounds but only in the presence of Mn(II). The homogeneous enzyme rapidly oxidizes Mn(II) to Mn(III) with a pH optimum of 5.0; the latter was detected by the characteristic spectrum of its lactate complex. In the presence of H2O2 the enzyme oxidizes Mn(II) significantly faster than it oxidizes all other substrates. Addition of 1 M equivalent of H2O2 to the native enzyme in 20 mM Na-succinate, pH 4.5, yields MnP compound II, characterized by a Soret maximum at 416 nm. Subsequent addition of 1 M equivalent of Mn(II) to the compound II form of the enzyme results in its rapid reduction to the native Fe3+ species. Mn(III)-lactate oxidizes all of the compounds which are oxidized by the enzymatic system. The relative rates of oxidation of various substrates by the enzymatic and chemical systems are similar. In addition, when separated from the polymeric dye Poly B by a semipermeable membrane, the enzyme in the presence of Mn(II)-lactate and H2O2 oxidizes the substrate. All of these results indicate that the enzyme oxidizes Mn(II) to Mn(III) and that the Mn(III) complexed to lactate or other alpha-hydroxy acids acts as an obligatory oxidation intermediate in the oxidation of various dyes and lignin model compounds. In the absence of exogenous H2O2, the Mn-peroxidase oxidized NADH to NAD+, generating H2O2 in the process. The H2O2 generated by the oxidation of NADH could be utilized by the enzyme to oxidize a variety of other substrates.  相似文献   

18.
Incubation of stimulated neutrophils with sulfhydryl (RSH) compounds or ascorbic acid (ascorbate) results in rapid superoxide (O2-)-dependent oxidation of these reducing agents. Oxidation of RSH compounds to disulfides (RSSR) is faster than the rate of O2- production by the neutrophil NADPH-oxidase, whereas about one ascorbate is oxidized per O2-. Ascorbate is oxidized to dehydroascorbate, which is also oxidized but at a slower rate. Oxidation is accompanied by a large increase in oxygen (O2) uptake that is blocked by superoxide dismutase. Lactoferrin does not inhibit, indicating that ferric (Fe3+) ions are not required, and Fe3+-lactoferrin does not catalyze RSH or ascorbate oxidation. Two mechanisms contribute to oxidation: 1) O2- oxidizes ascorbate or reduced glutathione and is reduced to hydrogen peroxide (H2O2), which also oxidizes the reductants. O2- reacts directly with ascorbate, but reduced glutathione oxidation is mediated by the reaction of O2- with manganese (Mn2+). The H2O2-dependent portion of oxidation is mediated by myeloperoxidase-catalyzed oxidation of chloride to hypochlorous acid (HOCl) and oxidation of the reductants by HOCl. 2) O2- initiates Mn2+-dependent auto-oxidation reactions in which RSH compounds are oxidized and O2 is reduced. Part of this oxidation is due to the RSH-oxidase activity of myeloperoxidase. This activity is blocked by superoxide dismutase but does not require O2- production by the NADPH-oxidase, indicating that myeloperoxidase produces O2- when incubated with RSH compounds. It is proposed that an important role for O2- in the cytotoxic activities of phagocytic leukocytes is to participate in oxidation of reducing agents in phagolysosomes and the extracellular medium. Elimination of these protective agents allows H2O2 and products of peroxidase/H2O2/halide systems to exert cytotoxic effects.  相似文献   

19.
Cyclic voltammetry data were obtained for eight phenazines and phenazine-N-oxides, and eleven quinoxalines and quinoxaline-N-oxides: 1,6-phenazine-diol-5,10-dioxide (iodinin), iodinin copper complex, 6-methoxy-1-phenazinol-5,10-dioxide 1,6-dimethoxyphenazine-5-oxide, 1,6-phenazinediol, 1,6-dimethoxyphenazine, quinoxaline-1,4-dioxide, 2-methylquinoxaline-1,4-dioxide, 2,3-diphenylquinoxaline-1,4-dioxide, 2-carboxyquinoxaline-1,4-dioxide, 5-hydroxyquinoxaline-1,4-dioxide, 5-hydroxy-8-methoxyquinoxaline-1,4-dioxide, 2-methylquinoxaline, 2,3-diphenylquinoxaline, 5-hydroxyquinoxaline, 5-hydroxy-8-methoxyquinoxaline and 2-(2-quinoxalinylmethylene)hydrazine carboxylic acid methyl ester-1,4-dioxide (Carbadox). The di-N-oxides exhibit the most positive E1/2 values within each class. Reversible first wave reductions were observed for iodinin, iodinin copper complex, 1,6-dimethoxyphenazine-5-oxide, 1,6-dimethoxyphenazine, quinoxaline-1,4-dioxide, 2-methylquinoxaline-1,4-oxide and 2,3-diphenylquinoxaline-1,4-dioxide. The results are correlated with structure. Some relationships exist between reduction potential and reported antimicrobial activity. A possible mechanism of drug action is addressed.  相似文献   

20.
H2O2 reacts with cytochrome c peroxidase in a variety of ways. The initial reaction produces cytochrome c peroxidase Compound I. If more than a 10-fold excess of H2O2 is added to the enzyme, a portion of the H2O2 will react with Compound I to produce molecular oxygen. The remainder oxidizes the heme group and various amino acid residues in the protein. If less than a 10-fold excess of H2O2 is added to the enzyme, essentially all the H2O2 is utilized by oxidation of amino acid residues in the protein. The oxidation of the amino acid residues by H2O2 substantially modifies the reactivity of cytochrome c peroxidase. The modification of reactivity could be the direct result of amino acid oxidation or an indirect result caused by a perturbation of the protein structure at the active site. The products oxidized at pH 8 lose their ability to react with H2O2. The products oxidized at pH4 react with H2O2 but their reactivity toward Fe(CN)4-6 is substantially reduced.  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