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.
Barry Halliwell 《Planta》1978,140(1):81-88
The enzyme horseradish peroxidase (EC 1.11.1.7) catalyses oxidation of NADH. NADH oxidation is prevented by addition of the enzyme superoxide dismutase (EC 1.15.1.1) to the reaction mixture before adding peroxidase but addition of dismutase after peroxidase has little inhibitory effect. Catalase (EC 1.11.1.6) inhibits peroxidase-catalysed NADH oxidation when added at any time during the reaction. Apparently the peroxidase uses hydrogen peroxide (H2O2) generated by non-enzymic breakdown of NADH to catalyse oxidation of NADH to a free-radical, NAD., which reduces oxygen to the superoxide free-radical ion, O2 .-. Some of the O2 .- reacts with peroxidase to give peroxidase compound III, which is catalytically inactive in NADH oxidation. The remaining O2 .- undergoes dismutation to O2 and H2O2. O2 .- does not react with NADH at significant rates. Mn2+ or lactate dehydrogenase stimulate NADH oxidation by peroxidase because they mediate a reaction between O2 .- and NADH. 2,4-Dichlorophenol, p-cresol and 4-hydroxycinnamic acid stimulate NADH oxidation by peroxidase, probably by breaking down compound III and so increasing the amount of active peroxidase in the reaction mixture. Oxidation in the presence of these phenols is greatly increased by adding H2O2. The rate of NADH oxidation by peroxidase is greatest in the presence of both Mn2+ and those phenols which interact with compound III. Both O2 .- and H2O2 are involved in this oxidation, which plays an important role in lignin synthesis.  相似文献   

2.
The two peroxidase isoenzyme groups (GI and GIII) localized in the cell walls of tobacco (Nicotiana tabacum L.) tissues were compared with respect to their capacity for NADH-dependent H2O2 formation. Peroxidases of the GIII group are slightly more active than those of the GI group when both are assayed under optimal conditions. This difference is probably not of major regulatory importance. NADH-dependent formation of H2O2 required the presence of Mn2+ and a phenol as cofactors. The addition of H2O2 to the reaction mixture accelerated subsequent NADH-dependent H2O2 formation. In the presence of both cofactors or Mn2+ alone, catalase oxidized NADH. However, if the cofactors were absent or if only dichlorophenol was present, catalase inhibited NADH oxidation. No H2O2 accumulation occurred in the presence of catalase. Superoxide dismutase inhibited NADH oxidation quite significantly indicating the involvement of the superoxide radical in the peroxidase reaction. These results are interpreted to mean that the reactions whereby tobacco cell wall peroxidases catalyze NADH-dependent H2O2 formation are similar to those proposed for horseradish peroxidase (Halliwell 1978 Planta 140: 81-88).  相似文献   

3.
A lignin-degrading basidiomycete, Phanerochaete chrysosporium, produces an extracellular peroxidase which in turn produces H2O2 by catalyzing the oxidation of NADH and NADPH. The high enzyme activity was observed in the culture grown under nutrient nitrogen limitation (low-N) and high oxygen tension (high-O2). The enzyme activity was absent in non-ligninolytic agitated culture and in the cultures of non-ligninolytic mutant strains of this organism. The culture method using polyurethane foam cubes as a support for the growing mycelia showed the beneficial effect of producing a large amount of the enzyme. The enzyme is capable of catalyzing the oxidation of NADH and NADPH in the absence of added H2O2, and its activity was inhibited strongly by catalase and superoxide dismutase. It is suggested that this peroxidase participates in the ligninolytic system of Phanerochaete chrysosporium as a donor of H2O2, which is required for the lignin-peroxidase reaction, by oxidizing extracellular NADH and NADPH.  相似文献   

4.
We report here that the Leishmania major ascorbate peroxidase (LmAPX), having similarity with plant ascorbate peroxidase, catalyzes the oxidation of suboptimal concentration of ascorbate to monodehydroascorbate (MDA) at physiological pH in the presence of added H2O2 with concurrent evolution of O2. This pseudocatalatic degradation of H2O2 to O2 is solely dependent on ascorbate and is blocked by a spin trap, α-phenyl-n-tert-butyl nitrone (PBN), indicating the involvement of free radical species in the reaction process. LmAPX thus appears to catalyze ascorbate oxidation by its peroxidase activity, first generating MDA and H2O with subsequent regeneration of ascorbate by the reduction of MDA with H2O2 evolving O2 through the intermediate formation of O2. Interestingly, both peroxidase and ascorbate-dependent pseudocatalatic activity of LmAPX are reversibly inhibited by SCN in a concentration dependent manner. Spectral studies indicate that ascorbate cannot reduce LmAPX compound II to the native enzyme in presence of SCN. Further kinetic studies indicate that SCN itself is not oxidized by LmAPX but inhibits both ascorbate and guaiacol oxidation, which suggests that SCN blocks initial peroxidase activity with ascorbate rather than subsequent nonenzymatic pseudocatalatic degradation of H2O2 to O2. Binding studies by optical difference spectroscopy indicate that SCN binds LmAPX (Kd = 100 ± 10 mM) near the heme edge. Thus, unlike mammalian peroxidases, SCN acts as an inhibitor for Leishmania peroxidase to block ascorbate oxidation and subsequent pseudocatalase activity.  相似文献   

5.
The rate of ascorbate and nicotinamide adenine dinucleotide plus hydrogen (NADH) cooxidation (i.e., their nonenzymic oxidation by peroxidase/H2O2-generated phenoxyl radicals of three hydroxycinnamates: caffeate, ferulate and p-coumarate) was studied in vitro. The reactions initiated by different sources of peroxidase (EC 1.11.1.7) [isolates from soybean (Glycine max L.) seed coat, maize (Zea mays L.) root-cell wall, and commercial horseradish peroxidase] were monitored. Native electrophoresis of samples and specific staining for peroxidase activity revealed various isoforms in each of the three enzyme sources. The peroxidase sources differed both in the rate of H2O2-dependent hydroxycinnamate oxidation and in the order of affinity for the phenolic substrates. The three hydroxycinnamates did not differ in their ability to cooxidize ascorbate, whereas NADH cooxidation was affected by substitution of the phenolic ring. Thus, p-coumarate was more efficient than caffeate in NADH cooxidation, with ferulate not being effective at all. Metal ions (Zn2+ and Al3+) inhibited the reaction of peroxidase with p-coumarate and affected the cooxidation rate of ascorbate and the peroxidase reaction in the same manner with all substrates used. However, inhibition of p-coumarate oxidation by metal ions did not affect NADH cooxidation rate. We propose that both the ascorbate and NADH cooxidation systems can function as mechanisms to scavenge H2O2 and regenerate phenolics in different cellular compartments, thus contributing to protection from oxidative damage. Electronic supplementary material The online version of this article (doi:) contains supplementary material, which is available to authorized users.  相似文献   

6.
Vanadate in the polymeric form of decavanadate, but not other forms, stimulated oxidation of NADH to NAD+ NADPH was also oxidized with comparable rates. This oxidation of NADH was accompanied by uptake of oxygen and generated hydrogen peroxide with the following stoichiometry: NADH + H+ + O2 → NAD+ + H2O2. The reaction followed second-order kinetics. The rate was dependent on the concentration of both NADH and vanadate and increased with decreasing pH. The reaction had an obligatory requirement for phosphate ions. Esr studies in the presence of the spin trap dimethyl pyrroline N oxide indicated the involvement of Superoxide anion as an intermediate. The reaction was sensitive to Superoxide dismutase and other scavengers of superoxide anions.  相似文献   

7.
Ferric leghemoglobin reductase (FLbR) from soybean (Glycine max [L.] Merr) nodules catalyzed oxidation of NADH, reduction of ferric leghemoglobin (Lb+3), and reduction of dichloroindophenol (diaphorase activity). None of these reactions was detectable when O2 was removed from the reaction system, but all were restored upon readdition of O2. In the absence of exogenous electron carriers and in the presence of O2 and excess NADH, FLbR catalyzed NADH oxidation with the generation of H2O2 functioning as an NADH oxidase. The possible involvement of peroxide-like intermediates in the FLbR-catalyzed reactions was analyzed by measuring the effects of peroxidase and catalase on FLbR activities; both enzymes at low concentrations (about 2 μg/mL) stimulated the FLbR-catalyzed NADH oxidation and Lb+3 reduction. The formation of H2O2 during the FLbR-catalyzed NADH oxidation was confirmed using a sensitive assay based on the fluorescence emitted by dichlorofluorescin upon reaction with H2O2. The stoichiometry ratios between the FLbR-catalyzed NADH oxidation and Lb+3 reduction were not constant but changed with time and with concentrations of NADH and O2 in the reaction solution, indicating that the reactions were not directly coupled and electrons from NADH oxidation were transferred to Lb+3 by reaction intermediates. A study of the affinity of FLbR for O2 showed that the enzyme required at least micromolar levels of dissolved O2 for optimal activities. A mechanism for the FLbR-catalyzed reactions is proposed by analogy with related oxidoreductase systems.  相似文献   

8.
Scopoletin was found to be a substrate for a single anodic isoperoxidase isolated from tobacco callus tissue W-38. Isolation of this peroxidase was accomplished using DEAE-cellulose chromatography. This isoperoxidase catalysed the destruction of scopoletin in the presence of H2O2 only. An enzyme assay for the scopoletin reaction was developed. The pH optimum of the enzyme was 5·5 and the apparent Kms for scopoletin and H2O2 were 0·6 and 0·9 rnM respectively.  相似文献   

9.
In cells, mitochondria, endoplasmic reticulum, and peroxisomes are the major sources of reactive oxygen species (ROS) under physiological and pathophysiological conditions. Cytochrome c (cyt c) is known to participate in mitochondrial electron transport and has antioxidant and peroxidase activities. Under oxidative or nitrative stress, the peroxidase activity of Fe3+cyt c is increased. The level of NADH is also increased under pathophysiological conditions such as ischemia and diabetes and a concurrent increase in hydrogen peroxide (H2O2) production occurs. Studies were performed to understand the related mechanisms of radical generation and NADH oxidation by Fe3+cyt c in the presence of H2O2. Electron paramagnetic resonance (EPR) spin trapping studies using 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) were performed with NADH, Fe3+cyt c, and H2O2 in the presence of methyl-β-cyclodextrin. An EPR spectrum corresponding to the superoxide radical adduct of DMPO encapsulated in methyl-β-cyclodextrin was obtained. This EPR signal was quenched by the addition of the superoxide scavenging enzyme Cu,Zn-superoxide dismutase (SOD1). The amount of superoxide radical adduct formed from the oxidation of NADH by the peroxidase activity of Fe3+cyt c increased with NADH and H2O2 concentration. From these results, we propose a mechanism in which the peroxidase activity of Fe3+cyt c oxidizes NADH to NAD, which in turn donates an electron to O2, resulting in superoxide radical formation. A UV-visible spectroscopic study shows that Fe3+cyt c is reduced in the presence of both NADH and H2O2. Our results suggest that Fe3+cyt c could have a novel role in the deleterious effects of ischemia/reperfusion and diabetes due to increased production of superoxide radical. In addition, Fe3+cyt c may play a key role in the mitochondrial “ROS-induced ROS-release” signaling and in mitochondrial and cellular injury/death. The increased oxidation of NADH and generation of superoxide radical by this mechanism may have implications for the regulation of apoptotic cell death, endothelial dysfunction, and neurological diseases. We also propose an alternative electron transfer pathway, which may protect mitochondria and mitochondrial proteins from oxidative damage.  相似文献   

10.
In the presence of NADH, and the reductase and rubredoxin components of the ω-hydroxylation system of Pseudomonas oleovorans, epinephrine is oxidized to adrenochrome at pH 7.8, and the reaction is strongly inhibited by the addition of superoxide dismutase (SDM). Boiled SDM has no effect on the reaction rate. The oxidation reaction is oxygen-dependent, and approximately 1 mole of H2O2 is produced per mole of O2 consumed. The stoichiometry between NADH oxidation and adrenochrome formation is approximately 2:1. Epoxidation and epinephrine oxidation are mutually competitive reactions, despite the fact that the epoxidation reaction is not stimulated by a superoxide generating system nor inhibited by SDM.  相似文献   

11.
In potato (Solatium tuberosum L. cv. Bintje and Doré) callus a very active hydrox-amate-stimulated NADH-dependent O2-uptake develops during the growth of the callus, which is caused by a peroxidase. More than 95% of the peroxidase activity is found in the 40000 g supernatant. The total activity may be as high as 1000 times the respiratory acitivity of the callus tissue. At least two fractions, obtained by Sephadex gel filtration, can be distinguished showing this peroxidase activity, one of about 15 kDa and one > 50 kDa. The main properties of both fractions are: a) Hydroxamate at 0.2–0.5 mM gives half-maximal stimulation. Maximal stimulation is observed with 1–3 mM benzhydroxamate (BHAM) and 1–15 mM salicylhydroxamate (SHAM). Higher concentrations, especially of BHAM, give less or no stimulation. b) Hydroxamates are not consumed during the reaction. c) Both NADH and NADPH can serve as the electron donor for the reaction. The affinity for NAD(P)H is very low (Km near 10 mM). In the absence of hydroxamates NAD(P)H is only slowly oxidized, with an even lower affinity. d) The peroxidase can carry out two reactions: an O2-consuming and a H2O2-consuming reaction. In both reactions one NAD(P)H is consumed. In the first reaction H2O2 is formed which can be consumed in the second reaction, resulting in an overall stoichiometry of 2 NADH consumed for each O2 molecule and in the production of H2O. e) The reaction is completely blocked by cyanide, superoxide dismutase (EC 1.15.1.1) and (excess) catalase (EC 1.11.1.6), but not by antimycin A or azide. This peroxidase-mediated O2-uptake might interfere with respiratory measurements. In experiments with isolated mitochondria this interference can be prevented by the addition of catalase to the reaction mixture. The use of high concentrations of hydroxamate is not allowed because of inhibitory effects on the cytochrome pathway. In intact callus tissue hydroxamates only stimulate O2-uptake in the presence of exogenous NADH. In vivo the peroxidase does not appear to function in O2-uptake, probably because of its localization (at least partly in the cell wall) and/or its low affinity for NADH. The use of hydroxamates in the determination of cytochrome and alternative pathway activity is discussed.  相似文献   

12.
H2O2 production by coupled mitochondrial fractions from the protozoan, Crithidia fasciculata, has been measured spectrophotometrically by the formation of the stable enzyme-substrate complex with yeast cytochrome c peroxidase. H2O2 formation was observed with succinate, l-α-glycerophosphate, l-proline, α-ketoglutarate, and with endogenous substrate. The maximum rate of H2O2 generation obtained with each substrate in the presence of antimycin A was about 10% of the state 4 rate of O2 respiration, and only 1–2% of the carbonylcyanide m-fluorophenylhydrazone-uncoupled respiratory rate. Therefore, excess O2 uptake due to the formation of H2O2 cannot satisfactorily account for the low ADP:O ratios previously reported.Cytochrome c peroxidase activity was measured in mitochondrial preparations by recording the decrease in absorbance at 550 nm during the oxidation of horse heart ferrocytochrome c which was observed after addition of H2O2. The distribution of activity after sonic disruption of mitochondrial preparations was that expected for a soluble enzyme. The activity was proportional to the amount of enzyme protein added, and was abolished by heating at 100 °C for 3 min. Total cytochrome c peroxidase activity in mitochondrial fractions isolated from C. fasciculata was calculated to be 0.3% that of isolated yeast mitochondria, but it is suggested that the in vivo activity may be considerably higher than this estimate.  相似文献   

13.
Sites of H2O2 generation in lightly prefixed, intact thyroid follicles were studied by two cytochemical reactions: peroxidase-dependent DAB oxidation and cerium precipitation. In both cases reaction product accumulated on the apical surface of the follicle cell at the membrane-colloid interface. The former reaction was inhibited by the peroxidase inhibitor, aminotriazole; both reactions were blocked by the presence of catalase. NADH in the medium slightly increased the amount of cerium precipitation. The ferricyanide technique for oxidoreductase activity was also applied; reaction product again was associated with the apical surface. These results strongly imply that the follicle cells have a NADH oxidizing system generating H2O2 at the apical plasma membrane.  相似文献   

14.
Brück TB  Brück DW 《Biochimie》2011,93(2):217-226
Mitoxantrone (MH2X), an anthraquinone-type anti-cancer agent used clinically in the treatment of human malignancies, is oxidatively activated by the peroxidase/H2O2 enzyme system. In contrast to the enzymatic mechanisms of drug oxidation, the chemical transformations of MH2X are not well described. In this study, MH2X metabolites, produced by the horseradish, lacto- or lignin peroxidase (respectively HRP, LPO and LIP)/H2O2 system, were investigated by steady-state spectrokinetic and HPLC-MS methods. At an equimolar mitoxantrone/H2O2 ratio, the efficacy of the enzyme-catalyzed oxidation of mitoxantrone decreased in the following order: LPO > HRP > LIP, which accorded with the decreasing size of the substrate access channel in the enzyme panel examined. In all cases, the central drug oxidation product was the redox-active cyclic metabolite, hexahydronaphtho-[2,3-f]-quinoxaline-7,12-dione (MH2), previously identified in the urine of mitoxantrone-treated patients. As the reaction progressed, data gathered in this study suggests that further oxidation of the MH2 side-chains occurred, yielding the mono- and dicarboxylic acid derivatives respectively. Based on the available data a further MH2 derivative is proposed, in which the amino-alkyl side-chain(s) are cyclised. With increasing H2O2 concentrations, these novel MH2 derivatives were oxidised to additional metabolites, whose spectral properties and MS data indicated a stepwise destruction of the MH2 chromophore due to an oxidative cleavage of the 9,10-anthracenedione moiety. The novel metabolites extend the known sequence of peroxidase-induced mitoxantrone metabolism, and may contribute to the cytotoxic effects of the drug in vivo. Based on the structural features of the proposed MH2 oxidation products we elaborate on various biochemical mechanisms, which extend the understanding of mitoxantrone’s pharmaceutical action and its clinical effectiveness with a particular focus on peroxidase-expressing solid tumors, such as breast carcinoma.  相似文献   

15.
Both ellagic and gallic acids non competitively inhibited guaiacol oxidation by turnip peroxidase. The Ki values were 3 and 26 μm for ellagic and gallic acid respectively. Enzymatic oxidation of gallic acid by the isolated major turnip peroxidase was characterized with respect to spectral behaviour, affinity constant and pH effect. The Km for H2O2 and gallic acid are 2.5 and 8.0 mM for turnip peroxidase. The pH optimum for gallic acid oxidation is about 6.5 and the rate constant k4 decreased with the increase of pH in presence of both guaiacol and Gallic acid. When the gallic acid oxidation products were subjected to chromatographic analysis, it was found to be converted mainly to ellagic and an unknown quinone.  相似文献   

16.
During the germination of fenugreek (Trigonella foenum graecum L.) sulfhydryl groups rapidly declined in cotyledon and seedling axis, while peroxidase activity increased. Studies on purified isoenzymes showed that GSH was oxidized by the isoenzymes and was accomplished in presence of cofactors, Mn2+ and DCP along with H2O2 (0.01 mM). This reaction was found to be peroxidatic in nature. The oxidation was inhibited by catechol but was enhanced by malic acid.  相似文献   

17.
Two pure peroxidase isoenzymes B1 and D4 were isolated from the upper parts of 10-day-old wheat seedlings by means of gel and ion-exchange chromatography. Their MWs were 85000 and 24000 respectively. B1 was unstable and under various conditions it was converted to another isoenzyme, electrophoretically identical with D4. B1 contains about 40% of neutral sugars: 17.2% arabinose, 15.3% galactose, 5% glucose and traces of mannose. D4 is free of neutral sugars. None of the isoenzymes contained amino sugars. B1 oxidizes ferulic and p-coumaric acids. This oxidation has two pH optima of 4.4 and 5.4–5.6 and is inhibited by high concentrations of substrates, cyanide and azide. B1 oxidizes IAA in the presence of phenolic cofactor and Mn2+ ions. IAA oxidation has two pH optima of 4.5 and 5.6 and is inhibited by high substrate concentration, cyanide and azide, and by a number of indole derivatives. The main products of IAA oxidation are 3-methyleneoxindole and indole-3-methanol. o- and p- diphenols induce a lag period prior to IAA oxidation. Ferulic acid is oxidized during this lag period, probably to a dimer. B1 is able to produce H2O2 from oxygen. Mn2+ ions, a phenolic cofactor and an electron donor (IAA or NADH) are needed. B1 oxidizes α-keto-γ- methylmercaptobutyric acid to ethylene. D4 has a low peroxidatic activity and is inactive as an IAA oxidase. Thus B1 is probably an active cell wall-bound peroxidase isoenzyme, whereas D4 is its decomposition product.  相似文献   

18.
All aerobic biological systems, including N2-fixing root nodules, are subject to O2 toxicity that results from the formation of reactive intermediates such as H2O2 and free radicals of O2. H2O2 may be removed from root nodules in a series of enzymic reactions involving ascorbate peroxidase, dehydroascorbate reductase, and glutathione reductase. We confirm here the presence of these enzymes in root nodules from nine species of legumes and from Alnus rubra. Ascorbate peroxidase from soybean nodules was purified to near homogeneity. This enzyme was found to be a hemeprotein with a molecular weight of 30,000 as determined by sodium dodecyl sulfate gel electrophoresis. KCN, NaN3, CO, and C2H2 were potent inhibitors of activity. Nonphysiological reductants such as guaiacol, o-dianisidine, and pyrogallol functioned as substrates for the enzyme. No activity was detected with NAD(P)H, reduced glutathione, or urate. Ascorbate peroxidation did not follow Michaelis-Menten kinetics. The substrate concentration which resulted in a reaction rate of ½ Vmax was 70 micromolar for ascorbate and 3 micromolar for H2O2. The high affinity of ascorbate peroxidase for H2O2 indicates that this enzyme, rather than catalase, is responsible for most H2O2 removal outside of peroxisomes in root nodules.  相似文献   

19.
《Phytochemistry》1987,26(12):3133-3136
Cell wall-bound peroxidase (EC 1.11.1.7) from lupin (Lupinus albus) shows a transition from oxidase to peroxidase activity when it oxidizes NADH. The oxidase phase represents a lag period in the time course of the reaction. This phase is phenol-dependent and responsible for hydrogen peroxide formation. Guaiacol, an assay substrate, and p-coumaric, ferulic and sinapic acids, precursors of the cinnamyl alcohols used in the lignification process affect both the length of lag period and the rate of the peroxidase phase of NADH oxidation. The effect of different phenols on the time course of the reaction is related to the efficacy (Vmax/Km ratio) of the enzyme when it is acting on them as a peroxidese.  相似文献   

20.
BackgroundHorseradish peroxidase (HRP) catalyzes H2O2 dismutation while undergoing heme inactivation. The mechanism underlying this process has not been fully elucidated. The effects of nitroxides, which protect metmyoglobin and methemoglobin against H2O2-induced inactivation, have been investigated.MethodsHRP reaction with H2O2 was studied by following H2O2 depletion, O2 evolution and heme spectral changes. Nitroxide concentration was followed by EPR spectroscopy, and its reactions with the oxidized heme species were studied using stopped-flow.ResultsNitroxide protects HRP against H2O2-induced inactivation. The rate of H2O2 dismutation in the presence of nitroxide obeys zero-order kinetics and increases as [nitroxide] increases. Nitroxide acts catalytically since its oxidized form is readily reduced to the nitroxide mainly by H2O2. The nitroxide efficacy follows the order 2,2,6,6-tetramethyl-piperidine-N-oxyl (TPO) > 4-OH-TPO > 3-carbamoyl proxyl > 4-oxo-TPO, which correlates with the order of the rate constants of nitroxide reactions with compounds I, II, and III.ConclusionsNitroxide catalytically protects HRP against inactivation induced by H2O2 while modulating its catalase-like activity. The protective role of nitroxide at μM concentrations is attributed to its efficient oxidation by P940, which is the precursor of the inactivated form P670. Modeling the dismutation kinetics in the presence of nitroxide adequately fits the experimental data. In the absence of nitroxide the simulation fits the observed kinetics only if it does not include the formation of a Michaelis-Menten complex.General SignificanceNitroxides catalytically protect heme proteins against inactivation induced by H2O2 revealing an additional role played by nitroxide antioxidants in vivo.  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