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.
The autoxidation and monoamine oxidase (MAO)-mediated metabolism of dopamine (3-hydroxytyramine; DA) cause a continuous production of hydroxyl radical (*OH), which is further enhanced by the presence of iron (ferrous iron, Fe(2+) and ferric ion, Fe(3+)). The accumulation of hydrogen peroxide (H2O2) in the presence of Fe(2+) appears to discard the involvement of the Fenton reaction in this process. It has been found that the presence of DA significantly reduces the formation of thiobarbituric acid reagent substances (TBARS), which under physiological conditions takes place in mitochondrial preparations. The presence of DA is also able to reduce TBARS formation in mitochondrial preparations even in the presence of iron (Fe(2+) and Fe(3+)). However, DA boosted the carbonyl content of mitochondrial proteins, which was further increased in the presence of iron (Fe(2+) and Fe(3+)). This latter effect is also accompanied by a significant reduction in thiol content of mitochondrial proteins. It has also been observed how the pre-incubation of mitochondria with pargyline, an acetylenic MAO inhibitor, reduces the production of *OH and increases the formation of TBARS. Although, the MAO-mediated metabolism of DA increases MAO-B activity, the presence of iron inhibits both MAO-A and MAO-B activities. Consequently, DA has been shown to be a double-edged sword, because it displays antioxidant properties in relation to both the Fenton reaction and lipid peroxidation and exhibits pro-oxidant properties by causing both generation *OH and oxidation of mitochondrial proteins. Evidently, these pro-oxidant properties of DA help explain the long-term side effects derived from l-DOPA treatment of Parkinson's disease and its exacerbation by the concomitant use of DA metabolism inhibitors.  相似文献   

2.
Formation of H2O2 during the oxidation of three lignin-derived hydroquinones by the ligninolytic versatile peroxidase (VP), produced by the white-rot fungus Pleurotus eryngii, was investigated. VP can oxidize a wide variety of phenols, including hydroquinones, either directly in a manner similar to horseradish peroxidase (HRP), or indirectly through Mn3+ formed from Mn2+ oxidation, in a manner similar to manganese peroxidase (MnP). From several possible buffers (all pH 5), tartrate buffer was selected to study the oxidation of hydroquinones as it did not support the Mn2+-mediated activity of VP in the absence of exogenous H2O2 (unlike glyoxylate and oxalate buffers). In the absence of Mn2+, efficient hydroquinone oxidation by VP was dependent on exogenous H2O2. Under these conditions, semiquinone radicals produced by VP autoxidized to a certain extent producing superoxide anion radical (O2*-) that spontaneously dismutated to H2O2 and O2. The use of this peroxide by VP produced quinone in an amount greater than equimolar to the initial H2O2 (a quinone/H2O2 molar ratio of 1 was only observed under anaerobic conditions). In the presence of Mn2+, exogenous H2O2 was not required for complete oxidation of hydroquinone by VP. Reaction blanks lacking VP revealed H2O2 production due to a slow conversion of hydroquinone into semiquinone radicals (probably via autooxidation catalysed by trace amounts of free metal ions), followed by O2*- production through semiquinone autooxidation and O2*- reduction by Mn2+. This peroxide was used by VP to oxidize hydroquinone that was mainly carried out through Mn2+ oxidation. By comparing the activity of VP to that of MnP and HRP, it was found that the ability of VP and MnP to oxidize Mn2+ greatly increased hydroquinone oxidation efficiency.  相似文献   

3.
Oxygen activation during oxidation of the lignin-derived hydroquinones 2-methoxy-1,4-benzohydroquinone (MBQH(2)) and 2, 6-dimethoxy-1,4-benzohydroquinone (DBQH(2)) by laccase from Pleurotus eryngii was examined. Laccase oxidized DBQH(2) more efficiently than it oxidized MBQH(2); both the affinity and maximal velocity of oxidation were higher for DBQH(2) than for MBQH(2). Autoxidation of the semiquinones produced by laccase led to the activation of oxygen, producing superoxide anion radicals (Q(*-) + O(2) <--> Q + O(2)(*-)). As this reaction is reversible, its existence was first noted in studies of the effect of systems consuming and producing O(2)(*-) on quinone formation rates. Then, the production of H(2)O(2) in laccase reactions, as a consequence of O(2)(*-) dismutation, confirmed that semiquinones autoxidized. The highest H(2)O(2) levels were obtained with DBQH(2), indicating that DBQ(*-) autoxidized to a greater extent than did MBQ(*-). Besides undergoing autoxidation, semiquinones were found to be transformed into quinones via dismutation and laccase oxidation. Two ways of favoring semiquinone autoxidation over dismutation and laccase oxidation were increasing the rate of O(2)(*-) consumption with superoxide dismutase (SOD) and recycling of quinones with diaphorase (a reductase catalyzing the divalent reduction of quinones). These two strategies made the laccase reaction conditions more natural, since O(2)(*-), besides undergoing dismutation, reacts with Mn(2+), Fe(3+), and aromatic radicals. In addition, quinones are continuously reduced by the mycelium of white-rot fungi. The presence of SOD in laccase reactions increased the extent of autoxidation of 100 microM concentrations of MBQ(*-) and DBQ(*-) from 4.5 to 30.6% and from 19.6 to 40.0%, respectively. With diaphorase, the extent of MBQ(*-) autoxidation rose to 13.8% and that of DBQ(*-) increased to 39.9%.  相似文献   

4.
H2O2 is catalytically metabolized by ferric lactoperoxidase (LPO)----compound (cpd) I----cpd II----ferric LPO cycles. An excess of the substrate, however, is degraded by a ferric LPO----cpd I----cpd II----cpd III----ferrous LPO----ferric LPO cycle. This latter pathway leads to the partial or total irreversible inactivation of the enzyme depending on the excess of H2O2 (H. Jenzer, W. Jones, and H. Kohler (1986) J. Biol. Chem. 261, 15550-15556). Spin-trapping/ESR data indicate that in the course of the reaction superoxide (HO2./O2-) and hydroxyl radicals (OH.) are formed. Since many substances known to scavenge radicals, such as a spin trap (e.g., 5,5-dimethyl-1-pyrroline-N-oxide) desferrioxamine, albumin, or mannitol, do not prevent enzyme inactivation, we conclude that OH. generation is a site-specific reaction at or near the active center of LPO where bulky scavenger molecules may not be able to penetrate. We suggest the formation of OH. by a Fenton-like reaction between H2O2 and the intermediate ferrous state of the enzyme, which substitutes for Fe2+ in the Fenton reaction. OH. is a powerful oxidant which in turn may attack rapidly the nearest partner available, either H2O2 to produce HO2. and H2O, or the prosthetic group to give rise to oxidative cleavage of the porphyrin ring structure of the heme moiety of LPO and thus to the liberation of iron.  相似文献   

5.
Iron and copper toxicity has been presumed to involve the formation of hydroxyl radical (*OH) from H2O2 in the Fenton reaction. The aim of this study was to verify that Fe2+-O2 and Cu+-O2 chemistry is capable of generating *OH in the quasi physiological environment of Krebs-Henseleit buffer (KH), and to compare the ability of the Fe2+-O2 system and of the Fenton system (Fe2+ + H2O2) to produce *OH. The addition of Fe2+ and Cu+ (0-20 microM) to KH resulted in a concentration-dependent increase in *OH formation, as measured by the salicylate method. While Fe3+ and Cu2+ (0-20 microM) did not result in *OH formation, these ions mediated significant *OH production in the presence of a number of reducing agents. The *OH yield from the reaction mediated by Fe2+ was increased by exogenous Fe3+ and Cu2+ and was prevented by the deoxygenation of the buffer and reduced by superoxide dismutase, catalase, and desferrioxamine. Addition of 1 microM, 5 microM or 10 microM Fe2+ to a range of H2O2 concentrations (the Fenton system) resulted in a H2O2-concentration-dependent rise in *OH formation. For each Fe2+ concentration tested, the *OH yield doubled when the ratio [H2O2]:[Fe2+] was raised from zero to one. In conclusion: (i) Fe2+-O2 and Cu+-O2 chemistry is capable of promoting *OH generation in the environment of oxygenated KH, in the absence of pre-existing superoxide and/or H2O2, and possibly through a mechanism initiated by the metal autoxidation; (ii) The process is enhanced by contaminating Fe3+ and Cu2+; (iii) In the presence of reducing agents also Fe3+ and Cu2+ promote the *OH formation; (iv) Depending on the actual [H2O2]:[Fe2+] ratio, the efficiency of the Fe2+-O2 chemistry to generate *OH is greater than or, at best, equal to that of the Fe2+-driven Fenton reaction.  相似文献   

6.
The brown rot fungus Gloeophyllum trabeum uses an extracellular hydroquinone-quinone redox cycle to reduce Fe(3+) and produce H(2)O(2). These reactions generate extracellular Fenton reagent, which enables G. trabeum to degrade a wide variety of organic compounds. We found that G. trabeum secreted two quinones, 2,5-dimethoxy-1,4-benzoquinone (2,5-DMBQ) and 4,5-dimethoxy-1,2-benzoquinone (4,5-DMBQ), that underwent iron-dependent redox cycling. Experiments that monitored the iron- and quinone-dependent cleavage of polyethylene glycol by G. trabeum showed that 2,5-DMBQ was more effective than 4,5-DMBQ in supporting extracellular Fenton chemistry. Two factors contributed to this result. First, G. trabeum reduced 2,5-DMBQ to 2,5-dimethoxyhydroquinone (2,5-DMHQ) much more rapidly than it reduced 4,5-DMBQ to 4,5-dimethoxycatechol (4,5-DMC). Second, although both hydroquinones reduced ferric oxalate complexes, the predominant form of Fe(3+) in G. trabeum cultures, the 2,5-DMHQ-dependent reaction reduced O(2) more rapidly than the 4,5-DMC-dependent reaction. Nevertheless, both hydroquinones probably contribute to the extracellular Fenton chemistry of G. trabeum, because 2,5-DMHQ by itself is an efficient reductant of 4,5-DMBQ.  相似文献   

7.
We have examined the action of nitric oxide (NO) on the ability of Fenton's reagent (ferrous iron and hydrogen peroxide), to oxidize a number of organic optical probes. We found that NO is able to arrest the oxidation of organic compounds at concentrations of NO found in brain, in vivo. We present evidence that Fenton's reagent proceeds via a ferryl intermediate ([Fe[double bond]O]2+), before the generation of hydroxyl radical *OH. NO reacts rapidly with this ferryl, blocking the production of *OH. We propose that NO has an important role in protecting biological tissues, and the brain in particular, from Fenton chemistry.  相似文献   

8.
Cultured hepatocytes pretreated with the ferric iron chelator deferoxamine were resistant to the toxicity of H2O2 generated by either glucose oxidase or by the metabolism of menadione (2-methyl-1,4-naphthoquinone). Ferric, ferrous, or cupric ions restored the sensitivity of the cells to H2O2. Deferoxamine added to hepatocytes previously treated with this chelator prevented the restoration of cell killing by only ferric iron. The free radical scavengers mannitol, thiourea, benzoate, and 4-methylmercapto-2-oxobutyrate protected either native cells exposed to H2O2 or pretreated hepatocytes exposed to H2O2 and given ferric or ferrous iron. Superoxide dismutase prevented the killing of native hepatocytes by either glucose oxidase or menadione. With deferoxamine-pretreated hepatocytes, superoxide dismutase prevented the cell killing dependent upon the addition of ferric but not ferrous iron. Catalase prevented the killing by menadione of deferoxamine-pretreated hepatocytes given either ferric or ferrous iron. Deferoxamine pretreatment did not prevent the toxicity of t-butyl hydroperoxide but did, however, prevent that of cumene hydroperoxide. It is concluded that both ferric iron and superoxide ions are required for the killing of cultured hepatocytes by H2O2. The toxicity of H2O2 is also dependent upon its reaction with ferrous iron to form hydroxyl radicals by the Fenton reaction. The ferrous iron needed for this reaction is formed by the reduction of cellular ferric iron by superoxide ions. Such a sequence corresponds to the so-called iron-catalyzed Haber-Weiss reaction, and the present report documents its participation in the killing of intact hepatocytes by H2O2. Cumene hydroperoxide but not t-butyl hydroperoxide closely models the toxicity of hydrogen peroxide.  相似文献   

9.
Evidence that hydroxyl radicals mediate auxin-induced extension growth   总被引:17,自引:0,他引:17  
Schopfer P  Liszkay A  Bechtold M  Frahry G  Wagner A 《Planta》2002,214(6):821-828
Reactive oxygen intermediates, i.e. the superoxide radical (O*-)(2), hydrogen peroxide (H2O2) and the hydroxyl radical (*OH), are generally regarded as harmful products of oxygenic metabolism causing cell damage in plants, animals and microorganisms. However, oxygen radical chemistry may also play a useful role in polymer breakdown leading to wall loosening during extension growth of plant cells controlled by the phytohormone auxin. Backbone cleavage of cell wall polysaccharides can be accomplished in vitro by (*OH) produced from H2O2 in a Fenton reaction or in a reaction catalyzed by peroxidase supplied with O2 and NADH. Here, we show that coleoptile growth of maize seedlings is accompanied by the release of reactive oxygen intermediates in the cell wall. Auxin promotes release of (O*-)(2) and subsequent generation of (*OH)when inducing elongation growth. Experimental generation of (*OH) in the wall causes an increase in wall extensibility in vitro and replaces auxin in inducing growth. Auxin-induced growth can be inhibited by scavengers of (O*-)(2), H2O2 or (*OH), or inhibitors interfering with the formation of these molecules in the cell wall. These results provide the experimental background for a novel hypothesis on the mechanism of plant cell growth in which (*OH), produced from (O*-)(2) and H2O2 by cell wall peroxidase, acts as a wall-loosening agent.  相似文献   

10.
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   

11.
A simple chemical system consisting of FeSO4 and H2O2 (Fenton's reagent) was shown to emit light (chemiluminescence). The addition of tryptophan to the reaction markedly enhanced light production. Very little chemiluminescence was observed when H2O2 was omitted from the reaction and when ferric, instead of ferrous, ions were used. Hydroxyl radical (OH.) and singlet oxygen (1 deltagO2) quenchers suppressed chemiluminescence of the FeSO4 + tryptophan + H2O2 system; and, deuterium oxide (2H2O) enhanced chemiluminescence of both FeSO4 reactions. These observations suggest that a radical chain reaction involving both OH. and 1 deltag O2 is responsible for the chemiluminescent reactions. Six iron-containing proteins, some of which are located within granulocytes, all emitted light in the presence of H2O2. Since iron and H2O2 are present in metabolically stimulated granulocytes, it is likely that chemiluminescent reactions similar to the ones demonstrated in this study account for part of the chemiluminescence of activated granulocytes.  相似文献   

12.
Two laccase isoenzymes produced by Pleurotus eryngii were purified to electrophoretic homogeneity (42- and 43-fold) with an overall yield of 56.3%. Laccases I and II from this fungus are monomeric glycoproteins with 7 and 1% carbohydrate content, molecular masses (by sodium dodecyl sulfate-polyacrylamide gel electrophoresis) of 65 and 61 kDa, and pIs of 4.1 and 4.2, respectively. The highest rate of 2,2'-azino-bis(3-ethylbenzothiazoline-6-sulfonate) oxidation for laccase I was reached at 65 degrees C and pH 4, and that for laccase II was reached at 55 degrees C and pH 3.5. Both isoenzymes are stable at high pH, retaining 60 to 70% activity after 24 h from pH 8 to 12. Their amino acid compositions and N-terminal sequences were determined, the latter strongly differing from those of laccases of other basidiomycetes. Antibodies against laccase I reacted with laccase II, as well as with laccases from Pleurotus ostreatus, Pleurotus pulmonarius, and Pleurotus floridanus. Different hydroxy- and methoxy-substituted phenols and aromatic amines were oxidized by the two laccase isoenzymes from P. eryngii, and the influence of the nature, number, and disposition of aromatic-ring substituents on kinetic constants is discussed. Although both isoenzymes presented similar substrate affinities, the maximum rates of reactions catalyzed by laccase I were higher than those of laccase II. In reactions with hydroquinones, semiquinones produced by laccase isoenzymes were in part converted into quinones via autoxidation. The superoxide anion radical produced in the latter reaction dismutated, producing hydrogen peroxide. In the presence of manganous ion, the superoxide union was reduced to hydrogen peroxide with the concomitant production of manganic ion. These results confirmed that laccase in the presence of hydroquinones can participate in the production of both reduced oxygen species and manganic ions.  相似文献   

13.
The metal-independent production of hydroxyl radicals (*OH) from H(2)O(2) and tetrachloro-1,4-benzoquinone (TCBQ), a carcinogenic metabolite of the widely used wood-preservative pentachlorophenol, was studied by electron spin resonance methods. When incubated with the spin trapping agent 5,5-dimethyl-1-pyrroline N-oxide (DMPO), TCBQ and H(2)O(2) produced the DMPO/*OH adduct. The formation of DMPO/*OH was markedly inhibited by the *OH scavenging agents dimethyl sulfoxide (DMSO), ethanol, formate, and azide, with the concomitant formation of the characteristic DMPO spin trapping adducts with *CH(3), *CH(CH(3))OH, *COO(-), and *N(3), respectively. The formation of DMPO/*OH and DMPO/*CH(3) from TCBQ and H(2)O(2) in the absence and presence, respectively, of DMSO was inhibited by the trihydroxamate compound desferrioxamine, accompanied by the formation of the desferrioxamine-nitroxide radical. In contrast, DMPO/*OH and DMPO/*CH(3) formation from TCBQ and H(2)O(2) was not affected by the nonhydroxamate iron chelators bathophenanthroline disulfonate, ferrozine, and ferene, as well as the copper-specific chelator bathocuproine disulfonate. A comparative study with ferrous iron and H(2)O(2), the classic Fenton system, strongly supports our conclusion that *OH is produced by TCBQ and H(2)O(2) through a metal-independent mechanism. Metal-independent production of *OH from H(2)O(2) was also observed with several other halogenated quinones.  相似文献   

14.
L-Histidine (L-His) enhances the clastogenic effects of hydrogen peroxide (H(2)O(2)). We previously suggested the involvement of active transport in the efficient influx of an L-His--H(2)O(2) adduct into cells (Oya-Ohta et al. [1]). In this study, we detected intracellular H(2)O(2) by monitoring formation of 2',7'-dichlorofluorescein (DCF) from its precursor. More fluoroproduct accumulated dose-dependently in cells treated with a mixture of L-His and H(2)O(2) (mixture) than with H(2)O(2) alone. This observation supports our hypothesis that active transport is involved in the enhanced incorporation of H(2)O(2) into cells. Moreover, both mixture and the L-His--H(2)O(2) adduct were less active in the generation of hydroxyl radicals (*OH) upon addition of FeCl(2) than was H(2)O(2) alone in a cell-free system. This result suggests that the Fenton reaction might occur more effectively around the nucleus in cells. An immunohistochemical assay using 8-oxodG-specific monoclonal antibodies did not reveal whether the accumulation of H(2)O(2) generates 8-oxodeoxyguanosine (8-oxodG). No 8-oxodG was evident in cells treated with mixture or with H(2)O(2) alone, or even in cells treated with H(2)O(2) at high doses up to 20 mM and, in some cases, pre-treated with catalase inhibitors. It appears, therefore, that *OH and, specifically, *OH derived from intracellular Fenton reactions, might not play a role in the formation of 8-oxodG. However, exposure to UV-C of cells treated with H(2)O(2) yielded more 8-oxodG in the presence of L-His than in the absence of L-His. Thus, the previously observed enhancing effects of L-His were also noted during the induction of formation of 8-oxodG by UV-C plus H(2)O(2). The formation of 8-oxodG in response to UV-C alone was very limited and, hence, H(2)O(2) seemed to be an effective source of *OH only in the presence of UV-C. It is suggested that the *OH that induces formation of 8-oxodG is not *OH formed via intracellular Fenton reactions but is *OH formed via the dissociation of H(2)O(2) under UV-C.  相似文献   

15.
Zhao G  Arosio P  Chasteen ND 《Biochemistry》2006,45(10):3429-3436
Overexpression of human H-chain ferritin (HuHF) is known to impart a degree of protection to cells against oxidative stress and the associated damage to DNA and other cellular components. However, whether this protective activity resides in the protein's ability to inhibit Fenton chemistry as found for Dps proteins has never been established. Such inhibition does not occur with the related mitochondrial ferritin which displays much of the same iron chemistry as HuHF, including an Fe(II)/H(2)O(2) oxidation stoichiometry of approximately 2:1. In the present study, the ability of HuHF to attenuate hydroxyl radical production by the Fenton reaction (Fe(2+) + H(2)O(2) --> Fe(3+) + OH(-) + *OH) was examined by electron paramagnetic resonance (EPR) spin-trapping methods. The data demonstrate that the presence of wild-type HuHF during Fe(2+) oxidation by H(2)O(2) greatly decreases the amount of .OH radical produced from Fenton chemistry whereas the ferroxidase site mutant 222 (H62K + H65G) and human L-chain ferritin (HuLF) lack this activity. HuHF catalyzes the pairwise oxidation of Fe(2+) by the detoxification reaction [2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(O)OH(core) + 4H(+)] that occurs at the ferroxidase site of the protein, thereby preventing the production of hydroxyl radical. The small amount of *OH radical that is produced in the presence of ferritin (相似文献   

16.
The redox cycle of 2,5-dimethoxybenzoquinone (2,5-DMBQ) is proposed as a source of reducing equivalent for the regeneration of Fe2+ and H2O2 in brown rot fungal decay of wood. Oxalate has also been proposed to be the physiological iron reductant. We characterized the effect of pH and oxalate on the 2,5-DMBQ-driven Fenton chemistry and on Fe3+ reduction and oxidation. Hydroxyl radical formation was assessed by lipid peroxidation. We found that hydroquinone (2,5-DMHQ) is very stable in the absence of iron at pH 2 to 4, the pH of degraded wood. 2,5-DMHQ readily reduces Fe3+ at a rate constant of 4.5 x 10(3) M(-1)s(-1) at pH 4.0. Fe2+ is also very stable at a low pH. H2O2 generation results from the autoxidation of the semiquinone radical and was observed only when 2,5-DMHQ was incubated with Fe3+. Consistent with this conclusion, lipid peroxidation occurred only in incubation mixtures containing both 2,5-DMHQ and Fe3+. Catalase and hydroxyl radical scavengers were effective inhibitors of lipid peroxidation, whereas superoxide dismutase caused no inhibition. At a low concentration of oxalate (50 micro M), ferric ion reduction and lipid peroxidation are enhanced. Thus, the enhancement of both ferric ion reduction and lipid peroxidation may be due to oxalate increasing the solubility of the ferric ion. Increasing the oxalate concentration such that the oxalate/ferric ion ratio favored formation of the 2:1 and 3:1 complexes resulted in inhibition of iron reduction and lipid peroxidation. Our results confirm that hydroxyl radical formation occurs via the 2,5-DMBQ redox cycle.  相似文献   

17.
Mono-substituted hypocrellin B (MHB) and di-substituted hypocrellin B (DHB) by mercaptoacetic acid are new photosensitizers synthesized to improve the red absorption and water solubility of the parent hypocrellin B (HB). The photochemistries (Type I and/or Type II) of MHB and DHB have been studied in homogeneous solutions using electron paramagnetic resonance (EPR) and spectrophotometric methods. In anaerobic homogeneous DMSO solution, DHB*- (or MHB*-) was predominantly photoproduced via self-electron transfer between the excited- and ground-state species. The presence of an electron donor significantly promotes the formation of the reduced form of DHB (or MHB). As compared with hypocrellin B, the efficiencies of DHB*- and MHB*- generation was enhanced obviously. When oxygen-saturated solutions of DHB (or MHB) were illuminated with 532 nm light, singlet oxygen (1O2), superoxide anion radical (O2*-), hydroxyl radical (*OH) and hydrogen peroxide (H2O2) were formed. DHB and MHB generate 1O2 with quantum yields of 0.18 and 0.22, respectively, which are much lower than that of HB (0.76) in chloroform. The superoxide anion radical was significantly enhanced by the presence of electron donors. The rate of O2*- production was also dependent on the concentration of DHB or MHB. Moreover, O2*- upon disproportionation can generate H2O2 and ultimately the highly reactive *OH via the Fenton reaction and other pathway with the involvement of DHB*- (or MHB*-). As in the case of DHB*- (or MHB*-), the efficiencies of O2*- and *OH generation by DHB and MHB were also enhanced obviously, consistent with the fact that DHB*- (or MHB*-) acts as the precursor of O2* and thus *OH. These findings suggest that the photodynamic actions of DHB and MHB may proceed via enhanced Type I mechanism and reduced Type II mechanism as compared with that of HB.  相似文献   

18.
应用脱氧核糖降解法研究了离体条件下Cu,Zn-SOD与H2O2反应产生·OH,并对其机理进行了探讨。H2O2可使Cu,Zn-SOD失活,在失活过程中有·OH产生,甲酸钠和苯甲酸钠均能不同程度地保护Cu,Zn-SOD和降低H2O2与Cu,Zn-SOD反应中·OH的产额;热失活SOD也可和H2O2反应生成·OH,且效能高于活性Cu,Zn-SOD;用螫合剂脱去Cu,Zn-SOD的金属辅基后,脱辅基的SOD蛋白不能和H2O2反应产生·OH;Cu2+和H2O2反应产生·OH的效率很高,而Zn2+产生·OH的效率很低。实验结果提示Cu,Zn-SOD与H2O2反应产生的·OH可能是SOD活性中心的Cu2+与H2O2发生Fenton反应的结果.  相似文献   

19.
The interaction of NADPH with ferric complexes to catalyze microsomal generation of reactive oxygen intermediates has been well studied. Experiments were carried out to characterize the ability of NADH to interact with various ferric chelates to promote microsomal lipid peroxidation and generation of .OH-like species. In the presence of NADH and iron, microsomes produced .OH as assessed by the oxidation of a variety of .OH scavenging agents. Rates of NADH-dependent .OH production were 50 to 80% those of the NADPH-catalyzed reaction. The oxidation of dimethyl sulfoxide or t-butyl alcohol was inhibited by catalase and competitive .OH scavengers but not by superoxide dismutase or carbon monoxide. NADH-dependent .OH production was effectively catalyzed by ferric-EDTA and ferric-diethylenetriaminepentaacetic acid (DTPA), whereas ferric-ATP and ferric-citrate were poor catalysts. All these ferric chelates were reduced by microsomes in the presence of NADH (and NADPH). H2O2 was produced in the presence of NADH in a reaction stimulated by the addition of ferric-EDTA, consistent with the increase in .OH production. The latter appeared to be limited by the rate of H2O2 generation rather than the rate of reduction of the ferric chelate. NADH-dependent lipid peroxidation was much lower than the NADPH-catalyzed reaction and showed an opposite response to catalysis by ferric complexes compared to .OH generation as production of thiobarbituric acid-reactive material was increased with ferric-ATP and -citrate, but not with ferric-EDTA or- DTPA, and was not affected by catalase, SOD, or .OH scavengers. These results indicate that NADH can support microsomal reduction of ferric chelates, with the subsequent production of .OH-like species and peroxidation of lipids. The pattern of response of the NADH-dependent reactions with respect to catalytic effectiveness of ferric chelates and sensitivity to radical scavengers is similar to that found with NADPH. Many of the metabolic actions of ethanol have been ascribed to production of NADH as a consequence of oxidation by alcohol dehydrogenase. Since the cytosol normally maintains a highly oxidized NAD+/NADH redox ratio, it is interesting to speculate that increased availability of NADH from the oxidation of ethanol may support microsomal reduction of iron complexes, with the subsequent generation of reactive oxygen intermediates.  相似文献   

20.
Menadione (Md)-resistant variants of V79 Chinese hamster cells were derived by culturing cells in progressively higher concentrations of this drug. Along with Md resistance these cells acquired cross-resistance to H2O2, which was retained when cells were cultured back in the absence of Md for 18 cell passages. This indicates that some stable alteration is responsible for the modification and may suggest that the toxicity of Md is mediated by oxygen activation. In agreement with the latter a 2.8-fold increase in catalase activity and a 1.5-fold increase in glutathione content were observed in Md-resistant cells as compared to parental cells, whereas superoxide dismutase and glutathione peroxidase remained unaltered. The use of inhibitors of Fenton reaction, inhibitors of enzymatic and nonenzymatic lipid peroxidation and OH radical scavengers, indicated that both DNA damage and cytotoxic effects of H2O2 and Md are mediated by OH radical, without intervention of lipid peroxides. However whereas the catalysis by iron was required for toxicity of extracellular H2O2, it was not involved in Md-induced toxicity. Possible explanations for this difference have been considered, one of them assuming that an excess of the semiquinone form of Md in the cell might replace iron II as a H2O2 reductant, producing OH radical by an organic Fenton reaction.  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