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 kinetics of iron binding by deferrioxamine B mesylate and the ramifications of this process upon iron-catalyzed lipid peroxidation were assessed. The relative rates of Fe(III) binding by deferrioxamine varied for the chelators tested as follows: ADP greater than AMP greater than citrate greater than histidine greater than EDTA. The addition of a fivefold molar excess of deferrioxamine to that of Fe(III) did not result in complete binding (within 10 min) for any of the Fe(III) chelates tested except ADP:Fe(III). The rates of Fe(III) binding by deferrioxamine were greater at lower pH and when the competing chelator concentration was high in relationship to iron. The relatively slow binding of Fe(III) by deferrioxamine also affected lipid peroxidation, an iron-dependent process. The addition of deferrioxamine to an ascorbate- and ADP:Fe(III)-dependent lipid peroxidation system resulted in a time-dependent inhibition or stimulation of malondialdehyde formation (i.e., lipid peroxidation), depending on the ratio of deferrioxamine to iron. Converse to Fe(III), the rates of Fe(II) binding by deferrioxamine from the chelators tested above were rapid and complete (within 1 min), and resulted in the oxidation of Fe(II) to Fe(III). Lipid peroxidation dependent on Fe(II) autoxidation was stimulated by the addition of deferrioxamine. Malondialdehyde formation in this system was inhibited by the addition of catalase, and a similar extent of lipid peroxidation was achieved by substituting hydrogen peroxide for deferrioxamine. Collectively, these results suggest that the kinetics of Fe(III) binding by deferrioxamine is a slow, variable process, whereas Fe(II) binding is considerably faster. The binding of either valence of iron by deferrioxamine may result in variable effects on iron-catalyzed processes, such as lipid peroxidation, either via slow binding of Fe(III) or the rapid binding of Fe(II) with concomitant Fe(II) oxidation.  相似文献   

2.
Hemoglobin: A mechanism for the generation of hydroxyl radicals   总被引:4,自引:0,他引:4  
Oxyhemoglobin (HbO2) reduces Fe(III) NTA aerobically to become methemoglobin (metHb) and Fe(II)NTA. These conditions are favorable for the generation via Fenton chemistry of the hydroxyl radical that was measured by HPLC using salicylate as a probe. The levels of hydroxyl radicals generated are a function of both the percent metHb formed and the chemical nature of the buffer. The rates of formation of both metHb and hydroxyl radicals were dependent upon the concentration of Fe(III)NTA. Of the buffers tested, HEPES was the most effective scavenger of hydroxyl radicals while the other buffers scavenged in the order: HEPES > Tris > MOPS > NaCl ≈ unbuffered. The addition of catalase to remove H202 or bathophenanthroline to chelate Fe(II) inhibited virtually all hydroxyl radical formation. Carbonyl formation from free radical oxidation of amino acids was found to be 0.1 mol/mol of hemoglobin. These experiments demonstrate the ability of hemoglobin to participate directly in the generation of hydroxyl radicals mediated by redox metals, and provide insight into potential oxidative damage from metals released into the blood during some pathologic disorders including iron overload.  相似文献   

3.
The reaction of ferrous bleomycin with dioxygen is reexamined to clarify whether radical species derived from molecular oxygen are generated. Detection of low levels of spin-trapped oxyradicals confirm the production of OH during this reaction when bleomycin is present in excess, but not when iron and drug concentrations are equal. In phosphate buffer, hydroxyl radicals continue to be spin trapped for at least 15 min after Fe(II)bleomycin has been oxidized to Fe(III)bleomycin. In HEPES buffer, detection of a HEPES radical in the absence of spin trap over the same period independently supports the conclusion that reactive radicals are present after the initial oxidation of Fe(II)bleomycin is complete. When glutathione is included in the aerobic reaction mixture, thiyl radical species are spin trapped. The reaction of Fe(III)bleomycin with cysteine produces thiyl radical without spin-trapped hydroxyl radical.  相似文献   

4.
Cytochrome c(3+)-catalyzed peroxidation of phosphatidylcholine liposomes by hydrogen peroxide (H2O2) was indicated by the production of thiobarbituric acid reactive substances, oxygen consumption, and emission of spontaneous chemiluminescence. The iron chelator diethylenetriaminepentaacetic acid (DTPA) only partially inhibited peroxidation when H2O2 concentrations were 200 microM or greater. In contrast, iron compounds such as ferric chloride, potassium ferricyanide, and hemin induced H2O2-dependent lipid peroxidation which was totally inhibitable by DTPA. Cyanide and urate, which react at or near the cytochrome-heme, completely prevented lipid peroxidation, while hydroxyl radical scavengers and superoxide dismutase had very little or no inhibitory effect. Changes in liposome surface charge did not influence cytochrome c3+ plus H2O2-dependent peroxidation, but a net negative charge was critical in favoring cytochrome c(3+)-dependent, H2O2-independent lipid auto-oxidative processes. These results show that reaction of cytochrome c with H2O2 promotes membrane oxidation by more than one chemical mechanism, including formation of high oxidation states of iron at the cytochrome-heme and also by heme iron release at higher H2O2 concentrations. Cytochrome c3+ could react with mitochondrial H2O2 to yield "site-specific" mitochondrial membrane lipid peroxidation during tissue oxidant stress.  相似文献   

5.
J M Gutteridge 《FEBS letters》1984,172(2):245-249
Iron salts stimulate lipid peroxidation by decomposing lipid peroxides to produce alkoxyl and peroxyl radicals which initiate further oxidation. In aqueous solution ferrous salts produce OH. radicals, a reactive species able to abstract hydrogen atoms from unsaturated fatty acids, and so can initiate lipid peroxidation. When iron salts are added to lipids, containing variable amounts of lipid peroxide, the former reaction is favoured and OH. radicals contribute little to the observed rate of peroxidation. When iron is complexed with EDTA, however, lipid peroxide decomposition is prevented, but the complex reacts with hydrogen peroxide to form OH. radicals which are seen to initiate lipid peroxidation. Superoxide radicals appear to play an important part in reducing the iron complex.  相似文献   

6.
The antioxidative effect of α-tocopherol incorporated into lecithin liposomes was studied. Lipid peroxidation of liposome membranes, assayed as malondialdehyde production, was catalyzed by ascorbic acid and Fe2+. The peroxidation reaction, which did not involve the formation of singlet oxygen, superoxide, hydrogen peroxide, or a hydroxyl radical, was inhibited by α-tocopherol and a model compound of α-tocopherol, 2,2,5,7,8-pentamethyl-6-hydroxy-chroman (TMC), but not by phytol, α-tocopherylquinone, or α-tocopheryl acetate. One mole of α-tocopherol completely prevented peroxidation of about 100 moles of polyunsaturated fatty acid. Decrease in membrane fluidity by lipid peroxidation, estimated as increase of fluorescence polarization of 1,6-diphenyl-1,3,5-hexatriene (DPH) embedded in the membrane, was also inhibited by α-tocopherol and TMC, reflecting their antioxidant functions. Cholesterol did not act as an antioxidant, even when incorporated in large amount into the liposome membranes, but it increased the antioxidative efficiency of α-tocopherol. When a mixture of liposomes with and without α-tocopherol was incubated with Fe2+ and ascorbic acid, α-tocopherol did not protect the liposomes not containing α-tocopherol from peroxidation. However, preincubation of the mixture, or addition of Triton X-100 allowed the α-tocopherol to prevent peroxidation of the liposomes not containing α-tocopherol. In contrast, in similar experiments, liposomes containing TMC prevented peroxidation of those without TMC without preincubation. Tocopherol in an amount so small as to exhibit only a slight antioxidative effect was oxidized when incorporated in egg lecithin liposomes, but it mostly remained unoxidized when incorporated in dipalmitoyllecithin liposomes, indicating that oxygen activated by ascorbic acid-Fe2+ does not oxidize α-tocopherol directly. Thus, decomposition of α-tocopherol may be caused by its interaction with peroxy and/or alkoxyl radicals generated in the process of lipid peroxidation catalyzed by Fe2+ and ascorbic acid.  相似文献   

7.
Aluminium salts do not themselves stimulate peroxidation of ox-brain phospholipid liposomes, but they greatly accelerate the peroxidation induced by iron(II) salts at acidic pH values. This effect of Al(III) is not seen at pH 7.4, perhaps because Al(III) salts form insoluble complexes at this pH in aqueous solution. Peroxidation of liposomes in the presence of Al(III) and Fe(II) salts is inhibited by the chelating agent desferrioxamine, and by EDTA and diethylenetriaminepentaacetic acid at concentrations greater than those of Fe(II) salt. Aluminium salts slightly stimulate the peroxidation of peroxide-depleted linolenic acid micelles, but they do not accelerate the peroxidation induced by addition of iron(II) salts to the micelles at acidic pH. Aluminium salts accelerate the peroxidation observed when human erythrocytes are treated with hydrogen peroxide at pH 7.4. Desferrioxamine decreases the peroxidation. We suggest that Al(III) ions produce an alteration in membrane structure that facilitates lipid peroxidation, and that the increased formation of fluorescent age pigments in the nervous system of patients exposed to toxic amounts of Al(III) may be related to this phenomenon. The ability of desferal to bind both iron (III) and aluminium(III) salts and to inhibit lipid peroxidation makes it an especially useful chelating agent in the treatment of 'aluminium overload'.  相似文献   

8.
Hematin- and peroxide-catalyzed peroxidation of phospholipid liposomes   总被引:3,自引:0,他引:3  
The effect of hydroperoxides on hematin-catalyzed initiation and propagation of lipid peroxidation was examined utilizing soybean phosphatidylcholine liposomes as model membranes. Polarographic and spectrophotometric methods revealed a bimodal pseudocatalytic activity for hematin. A slow initiation phase of peroxidation was observed in the presence of low peroxide concentrations, whereas a fast propagative phase was observed at higher peroxide levels. Peroxide levels were manipulated enzymatically by the combination of phospholipase A2 and lipoxidase or by the direct addition of linoleic acid hydroperoxide, cumene hydroperoxide, or hydrogen peroxide. In addition, the effect of two different techniques for liposome preparation, i.e., sonication and extrusion, were compared on the basis of peroxidation kinetics. High pressure liquid chromatography analysis showed that sonicated liposomes contained higher levels of endogenous peroxides than the extruded ones. These sonicated liposomes also exhibited more rapid peroxidation following hematin addition. Extruded liposomes were more resistant to hematin-catalyzed peroxidation but became better substrates when exogenous hydroperoxides were added. All three peroxides reacted with hematin during which decomposition of peroxide and irreversible oxidation of hematin took place. Spectral analysis of hematin indicated that a higher oxidation state of hematin iron may be transiently formed during reaction with hydroperoxides and accounts for the propagation of lipid peroxidation when reactions proceed in the presence of soybean phosphatidylcholine liposomes. Of the three peroxides studied, linoleic acid hydroperoxide was most efficient in supporting hematin-catalyzed lipid peroxidation. The relevance of our findings is discussed in terms of the concentration dependence for lipid peroxides in determining the rate and extent of radical propagation chain reactions catalyzed by heme-iron catalysts such as hematin. Variation of hematin and linoleic hydroperoxide concentrations may provide an efficient and reproducible method for inducing and manipulating the rates and extent of lipid peroxidation through facilitation of the propagative phase of lipid peroxidation. In addition, we address a problem inherent to in vitro studies of heme-catalyzed lipid peroxidation where preparations of peroxide-free membranes should be of concern.  相似文献   

9.
Iron storage proteins, ferritin and haemosiderin, release iron to a range of chelators and reducing agents, including citrate, acetate and ascorbate. Released iron promotes both hydroxyl radical formation in the presence of hydrogen peroxide and lipid peroxidation in liposomes. Ferritin protein is modified in such reactions, both by free radical cleavage and addition reactions with aldehyde products of lipid peroxidation.  相似文献   

10.
Miura T  Muraoka S  Fujimoto Y 《Life sciences》2002,70(22):2611-2621
Lipid peroxidation was investigated to evaluate the deleterious effect on tissues by phenylbutazone (PB). PB induced lipid peroxidation of microsomes in the presence of horseradish peroxidase and hydrogen peroxide (HRP-H2O2). The lipid peroxidation was completely inhibited by catalase but not by superoxide dismutase. Mannitol and dimethylsulfoxide had no effect. These results indicated no paticipation of superoxide and hydroxyl radical in the lipid peroxidation. Reduced glutathione (GSH) efficiently inhibited the lipid peroxidation. PB radicals emitted electron spin resonance (ESR) signals during the reaction of PB with HRP-H2O2. Microsomes and arachidonic acid strongly diminished the ESR signals, indicating that PB radicals directly react with unsaturated lipids of microsomes to cause thiobarbituric acid reactive substances. GSH sharply diminished the ESR signals of PB radicals, suggesting that GSH scavenges PB radicals to inhibit lipid peroxidation. Also, 2-methyl-2-nitrosopropan strongly inhibited lipid peroxidation. R-Phycoerythrin, a peroxyl radical detector substance, was decomposed by PB with HRP-H2O2. These results suggest that lipid peroxidation of microsomes is induced by PB radicals or peroxyl radicals, or both.  相似文献   

11.
Quinolinate (pyridine-2,3-dicarboxylic acid, Quin) is a neurotoxic tryptophan metabolite produced mainly by immune-activated macrophages. It is implicated in the pathogenesis of several brain disorders including HIV-associated dementia. Previous evidence suggests that Quin may exert its neurotoxic effects not only as an agonist on the NMDA subtype of glutamate receptor, but also by a receptor-independent mechanism. In this study we address ability of ferrous quinolinate chelates to generate reactive oxygen species. Autoxidation of Quin-Fe(II) complexes, followed in Hepes buffer at pH 7.4 using ferrozine as the Fe(II) detector, was found to be markedly slower in comparison with iron unchelated or complexed to citrate or ADP. The rate of Quin-Fe(II) autoxidation depends on pH (squared hydroxide anion concentration), is catalyzed by inorganic phosphate, and in both Hepes and phosphate buffers inversely depends on Quin concentration. These observations can be explained in terms of anion catalysis of hexaaquairon(II) autoxidation, acting mainly on the unchelated or partially chelated pool of iron. In order to follow hydroxyl radical generation in the Fenton chemistry, electron paramagnetic resonance (EPR) spin trapping with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) was employed. In the mixture consisting of 100 mM DMPO, 0.1 mM Fe(II), and 8.8 mM hydrogen peroxide in phosphate buffer pH 7.4, 0.5 mM Quin approximately doubled the yield of DMPO-OH adduct, and higher Quin concentration increased the spin adduct signal even more. When DMPO-OH was pre-formed using Ti3+/hydrogen peroxide followed by peroxide removal with catalase, only addition of Quin-Fe(II), but not Fe(II), Fe(III), or Quin-Fe(III), significantly promoted decomposition of pre-formed DMPO-OH. Furthermore, reaction of Quin-Fe(II) with hydrogen peroxide leads to initial iron oxidation followed by appearance of iron redox cycling, detected as slow accumulation of ferrous ferrozine complex. This phenomenon cannot be abolished by subsequent addition of catalase. Thus, we propose that redox cycling of iron by a Quin derivative, formed by initial attack of hydroxyl radicals on Quin, rather than effects of iron complexes on DMPO-OH stability or redox cycling by hydrogen peroxide, is responsible for enhanced DMPO-OH signal in the presence of Quin. The present observations suggest that Quin-Fe(II) complexes display significant pro-oxidant characteristics that could have implications for Quin neurotoxicity.  相似文献   

12.
Ergothioneine is a product of plant origin that accumulates in animal tissues. Its suggested ability to act as an antioxidant has been evaluated. Ergothioneine is a powerful scavenger of hydroxyl radicals (.OH) and an inhibitor of iron or copper ion-dependent generation of .OH from hydrogen peroxide (H2O2). It is also an inhibitor of copper ion-dependent oxidation of oxyhaemoglobin, and of arachidonic acid peroxidation promoted by mixtures of myoglobin (or haemoglobin) and H2O2. Ergothioneine is a powerful scavenger of hypochlorous acid, being able to protect alpha 1-antiproteinase against inactivation by this molecule. By contrast, it does not react rapidly with superoxide (O2-) or hydrogen peroxide (H2O2) and it does not inhibit microsomal lipid peroxidation in the presence of iron ions. Overall, our results show that ergothioneine at the concentrations present in vivo could act as an antioxidant.  相似文献   

13.
Quinolinate (pyridine-2,3-dicarboxylic acid, Quin) is a neurotoxic tryptophan metabolite produced mainly by immune-activated macrophages. It is implicated in the pathogenesis of several brain disorders including HIV-associated dementia. Previous evidence suggests that Quin may exert its neurotoxic effects not only as an agonist on the NMDA subtype of glutamate receptor, but also by a receptor-independent mechanism. In this study we address ability of ferrous quinolinate chelates to generate reactive oxygen species. Autoxidation of Quin-Fe(II) complexes, followed in Hepes buffer at pH 7.4 using ferrozine as the Fe(II) detector, was found to be markedly slower in comparison with iron unchelated or complexed to citrate or ADP. The rate of Quin-Fe(II) autoxidation depends on pH (squared hydroxide anion concentration), is catalyzed by inorganic phosphate, and in both Hepes and phosphate buffers inversely depends on Quin concentration. These observations can be explained in terms of anion catalysis of hexaaquairon(II) autoxidation, acting mainly on the unchelated or partially chelated pool of iron. In order to follow hydroxyl radical generation in the Fenton chemistry, electron paramagnetic resonance (EPR) spin trapping with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) was employed. In the mixture consisting of 100 mM DMPO, 0.1 mM Fe(II), and 8.8 mM hydrogen peroxide in phosphate buffer pH 7.4, 0.5 mM Quin approximately doubled the yield of DMPO-OH adduct, and higher Quin concentration increased the spin adduct signal even more. When DMPO-OH was pre-formed using Ti3+/hydrogen peroxide followed by peroxide removal with catalase, only addition of Quin-Fe(II), but not Fe(II), Fe(III), or Quin-Fe(III), significantly promoted decomposition of pre-formed DMPO-OH. Furthermore, reaction of Quin-Fe(II) with hydrogen peroxide leads to initial iron oxidation followed by appearance of iron redox cycling, detected as slow accumulation of ferrous ferrozine complex. This phenomenon cannot be abolished by subsequent addition of catalase. Thus, we propose that redox cycling of iron by a Quin derivative, formed by initial attack of hydroxyl radicals on Quin, rather than effects of iron complexes on DMPO-OH stability or redox cycling by hydrogen peroxide, is responsible for enhanced DMPO-OH signal in the presence of Quin. The present observations suggest that Quin-Fe(II) complexes display significant pro-oxidant characteristics that could have implications for Quin neurotoxicity.  相似文献   

14.
Evidence for mitochondrial DNA damage by lipid peroxidation   总被引:8,自引:0,他引:8  
When mitochondria of rat liver were incubated in an in vitro system containing NADPH and ferrous chloride, marked lipid peroxidation occurred, as evidenced by the evolution of malonic dialdehyde. DNA isolated from these peroxidized mitochondrial preparations had completely different electrophoretic mobility than DNA isolated from mitochondria protected from peroxidation. Scavengers of superoxide anion, hydrogen peroxide and hydroxyl radicals offered no protection against either lipid peroxidation or DNA damage. However, alpha-tocopherol protected against both lipid peroxidation and damage to the mitochondrial genome. These results support the hypothesis that lipid peroxidation can mediate DNA damage.  相似文献   

15.
《Free radical research》2013,47(3-6):143-148
Iron plays a central role in oxidative injury, reportedly because it catalyzes superoxide- and hydrogen peroxide-dependent reactions yielding a powerful oxidant such as the hydroxyl radical. Iron is also thought to mediate the cardiotoxic and antitumour effects of adriamycin and related compounds. NADPH-supplemented microsomes reduce adriamycin to a semiquinone radical, which in turn re-oxidizes in the presence of oxygen to form superoxide and hence hydrogen peroxide. During this redox cycling membrane-bound nonheme iron undergoes superoxide dismutase- and catalase-insensitive reductive release. Membrane iron mobilization triggers lipid peroxidation, which is markedly enhanced by simultaneous addition of superoxide dismutase and catalase. The results indicate that : i) lipid peroxidation is mediated by the release of iron, yet the two reactions are governed by different mechanisms; and ii) oxygen radicals are not involved in or may actually inhibit adriamycin-induced lipid peroxidation. Microsomal iron delocalization and lipid peroxidation might represent oxyradical-independent mechanisms of adriamycin toxicity.  相似文献   

16.
Lipid peroxidation in kidney of rats fed with vitamin B-6 deficient diet for a period of 12 weeks was studied with pair-fed controls. The basal lipid peroxide level as well as the degree of susceptibility to lipid peroxidation in presence of promotors such as NADPH, ascorbate, t-butyl hydroperoxide, Fe2+, Cu2+ and oxalate, were increased in vitamin B-6 deficient kidney. The observed increased lipid peroxidation in vitamin B-6 deficient kidney was correlated with high levels of lipids, copper, iron, calcium and oxalate, low levels of antioxidants and antioxidant enzymes and increased levels of hydroperoxides and hydroxyl radicals.  相似文献   

17.
Probucol, a clinically used cholesterol lowering and antioxidant drug, was investigated for possible protection against lipid peroxidation and DNA damage induced by iron nitrilotriacetate (Fe-NTA) plus hydrogen peroxide (H2O2). Fe-NTA is a potent nephrotoxic agent and induces acute and subacute renal proximal tubular necrosis by catalyzing the decomposition of H2O2-derived production of hydroxyl radicals, which are known to cause lipid peroxidation and DNA damage. Fe-NTA is associated with a high incidence of renal adenocarcinoma in rodents. Lipid peroxidation and DNA damage are the principal manifestation of Fe-NTA induced toxicity, which could be mitigated by probucol. Incubation of renal microsomal membrane and/or calf thymus DNA with H2O2 (40 mM) in the presence of Fe-NTA (0.1 mM) induces renal microsomal lipid peroxidation and DNA damage to about 2.4-fold and 5.9-fold, respectively, as compared to control (P < 0.05). Induction of renal microsomal lipid peroxidation and DNA damage was inhibited by probucol in a concentration-dependent manner. In lipid peroxidation protection studies, probucol treatment showed a concentration-dependent inhibition (10-34% inhibition; P < 0.05) of Fe-NTA plus H2O2-induced lipid peroxidation as measured by thiobarbituric acid reacting species' (TBARS) formation in renal microsomes. Similarly, in DNA damage protection studies, probucol treatment also showed a concentration-dependent strong inhibition (36-71% inhibition; P < 0.05) of DNA damage. From these studies, it was concluded that probucol inhibits peroxidation of microsomal membrane lipids and DNA damage induced by Fe-NTA plus H2O2. However, because the lipid peroxidation and DNA damage studied here are regarded as early markers of carcinogenesis, we suggest that probucol may be developed as a cancer chemopreventive agent against renal carcinogenesis and other adverse effects of Fe-NTA exposure in experimental animals, in addition to being a cholesterol-lowering drug, useful for the control of hypercholestrolemia.  相似文献   

18.
Oxidative DNA damage was investigated by free radicals generated from HEPES (2-[4-(2-hydroxyethyl)-1-piperazinyl]ethanesulfonic acid) buffer, which is widely used in biochemical or biological studies, in the presence of Au(III). The effect of free radicals on the DNA damage was ascertained by gel electrophoresis, electron spin resonance (ESR) spectroscopy and circular dichroism (CD) spectroscopy. ESR results indicated the generation of nitrogen-centered cationic free radicals from the HEPES in the presence of Au(III) which cause the DNA damage. No ESR spectra were observed for phosphate, tris(hydroxymethyl)aminomethane (Tris-HCl) and acetate buffers in the presence of Au(III) or for HEPES buffer in the presence of other metal ions such as Mn(II), Fe(III), Co(II), Ni(II), Cu(II), Zn(II) and Pd(II) or [Au(III)(TMPyP)](5+) and [Pd(II)(TMPyP)](4+), where [H(2)(TMPyP)](4+) denotes tetrakis(1-methylpyridium-4-yl)porphyrin. Consequently, no DNA damage was observed for these buffer agents (e.g., phosphate, Tris-HCl or acetate) in the presence of Au(III) or for HEPES in the presence of other metal ions or the metalloporphyrins mentioned above. No detectable inhibitory effect on the DNA damage was observed by using the typical scavengers of reactive oxygen species (ROS) ()OH, O(2)(-) and H(2)O(2). This non-inhibitory effect indicated that no reactive oxygen species were generated during the incubation of DNA with HEPES and Au(III). The drastic change in CD spectra from positive ellipticity to negative ellipticity approximately at 270 nm with increasing concentration of Au(III) also indicated the significant damage of DNA. Only HEPES or Au(III) itself did not damage DNA. A mechanism for the damaging of DNA is proposed.  相似文献   

19.
A model lipid peroxidation system dependent upon the hydroxyl radical, generated by Fenton's reagent, was compared to another model system dependent upon the enzymatic generation of superoxide by xanthine oxidase. Peroxidation was studied in detergent-dispersed linoleic acid and in phospholipid liposomes. Hydroxyl radical generation by Fenton's reagent (FeCl2 + H2O2) in the presence of phospholipid liposomes resulted in lipid peroxidation as evidenced by malondialdehyde and lipid hydroperoxide formation. Catalase, mannitol, and Tris-Cl were capable of inhibiting activity. The addition of EDTA resulted in complete inhibition of activity when the concentration of EDTA exceeded the concentration of Fe2+. The addition of ADP resulted in slight inhibition of activity, however, the activity was less sensitive to inhibition by mannitol. At an ADP to Fe2+ molar ratio of 10 to 1, 10 mm mannitol caused 25% inhibition of activity. Lipid peroxidation dependent on the enzymatic generation of superoxide by xanthine oxidase was studied in liposomes and in detergent-dispersed linoleate. No activity was observed in the absence of added iron. Activity and the apparent mechanism of initiation was dependent upon iron chelation. The addition of EDTA-chelated iron to the detergent-dispersed linoleate system resulted in lipid peroxidation as evidenced by diene conjugation. This activity was inhibited by catalase and hydroxyl radical trapping agents. In contrast, no activity was observed with phospholipid liposomes when iron was chelated with EDTA. The peroxidation of liposomes required ADP-chelated iron and activity was stimulated upon the addition of EDTA-chelated iron. The peroxidation of detergent-dispersed linoleate was also enhanced by ADP-chelated iron. Again, this peroxidation in the presence of ADP-chelated iron was not sensitive to catalase or hydroxyl radical trapping agents. It is proposed that initiation of superoxide-dependent lipid peroxidation in the presence of EDTA-chelated iron occurs via the hydroxyl radical. However, in the presence of ADP-chelated iron, the participation of the free hydroxyl radical is minimal.  相似文献   

20.
J M Gutteridge 《FEBS letters》1986,201(2):291-295
Hydrogen peroxide and organic hydroperoxides react with haemoglobin to release iron which can be complexed to apotransferrin, bleomycin and desferrioxamine. This released iron promotes deoxyribose degradation by a Fenton reaction, DNA degradation in the presence of bleomycin and stimulates lipid peroxidation. It is likely that iron released from haemoglobin is the true generator of hydroxyl radicals in the 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