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 46 毫秒
1.
The involvement of "free" iron in damage caused by oxidative stress is well recognized. Superoxide generated in a short burst and at a relatively high flux by the xanthine/xanthine oxidase couple is known to release iron from ferritin in the presence of phenanthroline derivatives as iron chelators. However, superoxide generation via xanthine oxidase is accompanied by the simultaneous direct generation of hydrogen peroxide and, in the presence of ferritin, there is also a superoxide-independent release of iron. In this study it was found that the iron chelator employed attenuates superoxide formation from the xanthine/xanthine oxidase couple. The reaction of ferritin and transferrin with a clean chemical source of superoxide, di(4-carboxybenzyl)hyponitrite (SOTS-1) was therefore investigated. The efficiency of superoxide-induced iron release from ferritin increases dramatically as the superoxide flux is decreased, reaching as high as 0.5 Fe per O2*-. Treatment of ferritin for 16 h with SOTS-1 yielded as many as 130 Fe atoms/ferritin molecule, which greatly exceeds the amount of possible "contaminating" iron absorbed on the protein shell.  相似文献   

2.
Human ferritins sequester and store iron as a stable FeOOH((s)) mineral core within a protein shell assembled from 24 subunits of two types, H and L. Core mineralization in recombinant H- and L-subunit homopolymer and heteropolymer ferritins and several site-directed H-subunit variants was investigated to determine the iron oxidation/hydrolysis chemistry as a function of iron flux into the protein. Stopped-flow absorption spectrometry, UV spectrometry, and electrode oximetry revealed that the mineral core forms by at least three pathways, not two as previously thought. They correspond to the ferroxidase, mineral surface, and the Fe(II) + H2O2 detoxification reactions, respectively: [see reactions]. The H-subunit catalyzed ferroxidase reaction 1 occurs at all levels of iron loading of the protein but decreases with increasing iron added (48-800 Fe(II)/protein). Reaction 2 is the dominant reaction at 800 Fe(II)/protein, whereas reaction 3 occurs largely at intermediate iron loadings of 100-500 Fe(II)/protein. Some of the H2O2 produced in reaction 1 is consumed in the detoxification reaction 3; the 2/1 Fe(II)/H2O2 stoichiometry of reaction 3 minimizes hydroxyl radical production during mineralization. Human L-chain ferritin and H-chain variants lacking functional nucleation and/or ferroxidase sites deposit their iron largely through the mineral surface reaction 2. H2O2 is shown to be an intermediate product of dioxygen reduction in L-chain as well as in H-chain and H-chain variant ferritins.  相似文献   

3.
Mobilization of iron from ferritin by xanthine oxidase was studied under aerobic and anaerobic conditions. Aerobic iron release amounted to approx. 3.7 nmol/ml in 10 min. This amount was decreased by approx. 30% under anaerobic conditions. Aerobic iron mobilization involved two mechanisms. About 70% was released by O2.- generated by xanthine oxidase. The rest was released by O2(.-)-independent mechanisms, which also accounted for the total iron release when O2 was absent. A possible transfer of reducing equivalents directly from xanthine oxidase to ferritin is discussed. The results imply that, in pathological conditions with increased formation of O2.-, iron may be released from ferritin. Furthermore, in hypoxic tissues xanthine oxidase can release iron from ferritin by an O2(.-)-independent process. Free iron is liable to catalyse the formation of the extremely reactive and damaging OH. radical.  相似文献   

4.
Xanthine oxidase is able to mobilize iron from ferritin. This mobilization can be blocked by 70% by superoxide dismutase, indicating that part of its action is mediated by superoxide (O2-). Uric acid induced the release of ferritin iron at concentrations normally found in serum. The O2(-)-independent mobilization of ferritin iron by xanthine oxidase cannot be attributed to uric acid, because uricase did not influence the O2(-)-independent part and acetaldehyde, a substrate for xanthine oxidase, also revealed an O2(-)-independent part, although no uric acid was produced. Presumably the amount of uric acid produced by xanthine oxidase and xanthine is insufficient to release a measurable amount of iron from ferritin. The liberation of iron from ferritin by xanthine oxidase has important consequences in ischaemia and inflammation. In these circumstances xanthine oxidase, formed from xanthine dehydrogenase, will stimulate the formation of a non-protein-bound iron pool, and the O2(-)-produced by xanthine oxidase, or granulocytes, will be converted by 'free' iron into much more highly toxic oxygen species such as hydroxyl radicals (OH.), exacerbating the tissue damage.  相似文献   

5.
The origin of previously observed variations in stoichiometry of iron oxidation during the oxidative deposition of iron in ferritin has been poorly understood. Knowledge of the stoichiometry of Fe(II) oxidation by O2 is essential to establishing the mechanism of iron core formation. In the present work, the amount of Fe(II) oxidized was measured by M?ssbauer spectrometry and the O2 consumed by mass spectrometry. The number of protons produced in the reaction was measured by "pH stat" titration and hydrogen peroxide production by the effect of the enzyme catalase on the measured stoichiometry. For protein samples containing low levels of iron (24 Fe(II)/protein) the stoichiometry was found to be 1.95 +/- 0.18 Fe(II)/O2 with H2O2 being a product, viz. Equation 1. 2Fe2+ + O2 + 4H2O----2FeOOH + H2O2 + 4H+ (1) EPR spin trapping experiments showed no evidence of superoxide radical formation. The stoichiometry markedly increased with additional iron (240-960 Fe/protein), to a value of 4 Fe(II)/O2 as in Equation 2. 4Fe2+ + O2 + 6H2O----4FeOOH + 8H+ (2) As the iron core is progressively laid down, the mechanism of iron oxidation changes from a protein dominated process with H2O2 being the primary product of O2 reduction to a mineral surface dominated process where H2O is the primary product. These results emphasize the importance of the apoferritin shell in facilitating iron oxidation in the early stage of iron deposition prior to significant development of the polynuclear iron core.  相似文献   

6.
The reaction of horse spleen ferritin (HoSF) with Fe2+ at pH 6.5 and 7.5 using O2, H2O2 and 1:1 a mixture of both showed that the iron deposition reaction using H2O2 is approximately 20- to 50-fold faster than the reaction with O2 alone. When H2O2 was added during the iron deposition reaction initiated with O2 as oxidant, Fe2+ was preferentially oxidized by H2O2, consistent with the above kinetic measurements. Both the O2 and H2O2 reactions were well defined from 15 to 40 degrees C from which activation parameters were determined. The iron deposition reaction was also studied using O2 as oxidant in the presence and absence of catalase using both stopped-flow and pumped-flow measurements. The presence of catalase decreased the rate of iron deposition by approximately 1.5-fold, and gave slightly smaller absorbance changes than in its absence. From the rate constants for the O2 (0.044 s(-1)) and H2O2 (0.67 s(-1)) iron-deposition reactions at pH 7.5, simulations of steady-state H2O2 concentrations were computed to be 0.45 microM. This low value and reported Fe2+/O2 values of 2.0-2.5 are consistent with H2O2 rapidly reacting by an alternate but unidentified pathway involving a system component such as the protein shell or the mineral core as previously postulated [Biochemistry 22 (1983) 876; Biochemistry 40 (2001) 10832].  相似文献   

7.
Protein ferroxidase site(s), which catalyze the reaction between ferrous ion and dioxygen, have long been thought to play a role in core formation in ferritin; however, the mechanism of the reaction has never been studied in detail. In the present work, the enzymatic activity of ferritin was examined using oximetry, the net Fe2+ oxidation reaction being as follows. [formula: see text] The reaction exhibits saturation kinetics with respect to both Fe2+ and O2 (apparent Michaelis constants: Km,Fe = 0.35 +/- 0.01 mM and Km,O2 = 0.14 +/- 0.03 mM). The enzyme has a turnover number kcat = 80 +/- 3 min-1 at 20 degrees C with maximal activity at pH 7. The kinetics are discussed in terms of two mechanisms, one involving monomeric and the other dimeric iron protein complexes. In both instances Fe(II) oxidation occurs in 1-electron steps. Zinc(II) is a competitive inhibitor of iron(II) oxidation at Zn2+/apoprotein ratios > or = 6 (inhibitor constant KI,Zn = 0.067 +/- 0.011 mM) but appears to be a noncompetitive inhibitor at lower ratios (< or = 2), indicating the presence of more than one type of zinc binding site on the protein. At increments of 50 Fe2+/protein or less, all of the iron is oxidized via the protein ferroxidase site(s), independent of the amount of core already present. However, when larger increments are employed, some iron oxidation appears to occur on the surface of the mineral core. The results of these studies emphasize the role of the protein shell in all phases of core growth and confirm the presence of a functionally important catalytic site in ferritin in addition to other binding sites on the protein for iron.  相似文献   

8.
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   

9.
Despite previous detection of hydroxyl radical formation during iron deposition into ferritin, no reports exist in the literature concerning how it might affect ferritin function. In the present study, hydroxyl radical formation during Fe(II) oxidation by apoferritin was found to be contingent on the "ferroxidase" activity (i.e., H subunit composition) exhibited by apoferritin. Hydroxyl radical formation was found to affect both the stoichiometry and kinetics of Fe(II) oxidation by apoferritin. The stoichiometry of Fe(II) oxidation by apoferritin in an unbuffered solution of 50 mM NaCl, pH 7.0, was approximately 3.1 Fe(II)/O(2) at all iron-to-protein ratios tested. The addition of HEPES as an alternate reactant for the hydroxyl radical resulted in a stoichiometry of about 2 Fe(II)/O(2) at all iron-to-protein ratios. HEPES functioned to protect apoferritin from oxidative modification, for its omission from reaction mixtures containing Fe(II) and apoferritin resulted in alterations to the ferritin consistent with oxidative damage. The kinetic parameters for the reaction of recombinant human H apoferritin with Fe(II) in HEPES buffer (100 mM) were: K(m) = 60 microM, k(cat) = 10 s(-1), and k(cat)/K(m) = 1.7 x 10(5) M(-1) x (-1). Collectively, these results contradict the "crystal growth model" for iron deposition into ferritin and, while our data would seem to imply that the ferroxidase activity of ferritin is adequate in facilitating Fe(II) oxidation at all stages of iron deposition into ferritin, it is important to note that these data were obtained in vitro using nonphysiologic conditions. The possibility that these findings may have physiological significance is discussed.  相似文献   

10.
铁核结构对马脾铁蛋白释放铁动力学的影响   总被引:1,自引:0,他引:1  
建立H^% 参与马脾铁蛋白释放铁的动力方程,H^ 以1/2级反应方式参与铁蛋白释放铁核表层的铁。在酸性介质(PH6.5)中,铁蛋白释放铁的总平均速率(332Fe^3 /HSF.min)比在碱性介质(P8H8.0)中放铁的总平均速率(73Fe^3 /HSF.min)高4.6倍,铁蛋白的铁核结构和外加的磷酸盐均能影响该蛋白释放的速率,但并不改变其反应级数。  相似文献   

11.
Aminoacetone (AA) is a threonine and glycine metabolite overproduced and recently implicated as a contributing source of methylglyoxal (MG) in conditions of ketosis. Oxidation of AA to MG, NH4+, and H
2
O
2
has been reported to be catalyzed by a copper-dependent semicarbazide sensitive amine oxidase (SSAO) as well as by copper- and iron ion-catalyzed reactions with oxygen. We previously demonstrated that AA-generated O2•al (AA
) induce dose-dependent Fe(II) release from horse spleen ferritin (HoSF); no reaction occurs under nitrogen. In the present study we further explored the mechanism of iron release and the effect of AA on the ferritin apoprotein. Iron chelators such as EDTA, ATP and citrate, and phosphate accelerated AA-promoted iron release from HoSF, which was faster in horse spleen isoferritins containing larger amounts of phosphate in the core. Incubation of apoferritin with AA (2.5-50 mM, after 6 h) changes the apoprotein electrophoretic behavior, suggesting a structural modification of the apoprotein by AA-generated ROS. Superoxide dismutase (SOD) was able to partially protect apoferritin from structural modification whereas catalase, ethanol, and mannitol were ineffective in protection. Incubation of apoferritin with AA (1-10 mM) produced a dose-dependent decrease in tryptophan fluorescence (13-30%, after 5 h), and a partial depletion of protein thiols (29% after 24 h). The AA promoted damage to apoferritin produced a 40% decrease in apoprotein ferroxidase activity and an 80% decrease in its iron uptake ability. The current findings of changes in ferritin and apoferritin may contribute to intracellular iron-induced oxidative stress during AA formation in ketosis and diabetes mellitus.  相似文献   

12.
Formation of OH radicals in the stomach is possible by Fenton-type reactions, as gastric juice contains ascorbic acid (AA), iron ions and H2O2. An objective of the present study is to elucidate the effects of salivary SCN- and NO2- on the hydroxylation of salicylic acid which was induced by H2O2/Fe(II) and AA/H2O2/Fe(II) systems. Thiocyanate ion inhibited the hydroxylation of salicylic acid by the above systems in acidic buffer solutions and in acidified saliva. The inhibition by SCN- was deduced to be due to SCN- -dependent scavenging of OH radicals. Nitrite ion could enhance the SCN- -dependent inhibition of the hydroxylation induced by AA/H2O2/Fe(II) systems. The enhancement was suggested to be due to scavenging of OH radicals by NO which was formed by the reactions among AA, HNO2 and SCN- contained in the reaction mixture. The concentrations of SCN- and NO2-, which were effective for the inhibition, were in ranges of their normal salivary concentrations. These results suggest that salivary SCN- can cooperate with NO2- to protect stomach from OH radicals formed by AA/H2O2/Fe(II) systems under acidic conditions.  相似文献   

13.
The ability of superoxide anion (O2-) from stimulated human neutrophils (PMNs) to release ferrous iron (Fe2+) from transferrin was assessed. At pH 7.4, unstimulated PMNs released minimal amounts of O2- and failed to facilitate the release of Fe2+ from holosaturated transferrin. In contrast, incubation of phorbol myristate acetate (PMA)-stimulated PMNs with holosaturated transferrin at pH 7.4 enhanced the release of Fe2+ from transferrin eightfold in association with marked generation of O2-. The release of Fe2+ was inhibited by addition of superoxide dismutase (SOD), indicating that the release of Fe2+ was dependent on PMN-derived extracellular O2-. In contrast, at physiologic pH (7.4), incubation of transferrin at physiological levels of iron saturation (e.g. 32%) with unstimulated or PMA stimulated PMNs failed to facilitate the release of Fe2+. The effect of decreasing the pH on the release of Fe2+ from transferrin by PMN-derived O2- was determined. Decreasing the pH greatly facilitated the release of Fe2+ from both holosaturated transferrin and from transferrin at physiological levels of iron saturation by PMN-derived O2-. Release of Fe2+ occurred despite a decrease in the amount of extracellular O2- generated by PMNs in an acidic environment. These results suggest that transferrin at physiologic levels of iron saturation may serve as a source of Fe2+ for biological reactions in disease states where activated phagocytes are present and there is a decrease in tissue pH. The unbound iron could participate in biological reactions including promoting propagation of lipid peroxidation reactions or hydroxyl radical formation following reaction with phagocytic cell-derived hydrogen peroxide.  相似文献   

14.

Background

Most models for ferritin iron release are based on reduction and chelation of iron. However, newer models showing direct Fe(III) chelation from ferritin have been proposed. Fe(III) chelation reactions are facilitated by gated pores that regulate the opening and closing of the channels.

Scope of review

Results suggest that iron core reduction releases hydroxide and phosphate ions that exit the ferritin interior to compensate for the negative charge of the incoming electrons. Additionally, chloride ions are pumped into ferritin during the reduction process as part of a charge balance reaction. The mechanism of anion import or export is not known but is a natural process because phosphate is a native component of the iron mineral core and non-native anions have been incorporated into ferritin in vitro. Anion transfer across the ferritin protein shell conflicts with spin probe studies showing that anions are not easily incorporated into ferritin. To accommodate both of these observations, ferritin must possess a mechanism that selects specific anions for transport into or out of ferritin. Recently, a gated pore mechanism to open the 3-fold channels was proposed and might explain how anions and chelators can penetrate the protein shell for binding or for direct chelation of iron.

Conclusions and general significance

These proposed mechanisms are used to evaluate three in vivo iron release models based on (1) equilibrium between ferritin iron and cytosolic iron, (2) iron release by degradation of ferritin in the lysosome, and (3) metallo-chaperone mediated iron release from ferritin.  相似文献   

15.
The mechanism of ascorbate-promoted ferritin iron reduction under aerobic conditions was studied. The initial rate of ferritin iron release was determined by spectrophotometric measurement of the Fe(ferrozine)3(2+) complex which absorbs at 562 nm. Variation of the initial ferrozine concentration had no influence on the rate of iron release suggesting that ferrozine does not participate in the rate-determining step. Experimental measurements of the initial rate of iron release as a function of ascorbate concentration resulted in saturation kinetics with Vmax = 2.0 X 10(-7) M.min-1 and KM = 1.3 X 10(-3) M. The effect of pH was quite pronounced with a maximal rate of iron release at pH 7.0. Stoichiometric measurements on the reaction mixture, with added catalase, resulted in a ratio of 2 Fe(II) released per ascorbate. Ascorbate-mediated iron release was inhibited 85% by superoxide dismutase, but 0% inhibition was noted with aposuperoxide dismutase. It is proposed that superoxide ion, generated during the iron-promoted oxidation of ascorbate, acts as a reductant of ferritin iron. A mechanism of ferritin iron release consistent with these experimental observations is discussed.  相似文献   

16.
This report describes studies yielding additional evidence that superoxide anion (O2) production by some biological oxidoreductase systems is a potential source of hydroxyl radical production. The phenomenon appears to be an intrinsic property of certain enzyme systems which produce superoxide and H2O2, and can result in extensive oxidative degradation of membrane lipids. Earlier studies had suggested that iron (chelated to maintain solubility) augmented production of the hydroxyl radical in such systems according to the following reaction sequence: O2 + Fe3+ leads to O2 + Fe2+ Fe2+ + H2O2 leads to Fe3+ + HO-+OH-. The data reported below provide additional support for the occurrence of these reactions, especially the reduction of Fe3+ by superoxide. Because the conditions for such reactions appear to exist in animal tissues, the results indicate a mechanism for the initiation and promotion of peroxidative attacks on membrane lipids and also suggest that the role of antioxidants in intracellular metabolism may be to inhibit initiation of degradative reactions by the highly reactive radicals formed extraneously during metabolic activity. This report presents the following new information: (1) Fe3+ is reduced to Fe2+ during xanthine oxidase activity and a significant part of the reduction was oxygen dependent. (2) Mn2+ appears to function as an efficient superoxide anion scavenger, and this function can be inhibited by EDTA. (3) The O2-dependent reduction of Fe3+ to Fe2+ by xanthine oxidase activity is inhibited by Mn2+, which, in view of statement 2 above, is a further indication that the reduction of the iron involves superoxide anion. (4) Free radical scavengers prevent or reverse the Fe3+ inhibiton of cytochrome c3+ reduction by xanthine oxidase. (5) The inhibition of xanthine oxidase-catalyzed reduction of cyt c3+ by Fe3+ does not affect uric acid production by the xanthine oxidase system. (6) The reoxidation of reduced cyt c in the xanthine oxidase system is markedly enhanced by Fe3+ and is apparently due to enhanced HO-RADICAL formation since the Fe3+-stimulated reoxidation is inhibited by free radical scavengers, including those with specificity for the hydroxyl radical.  相似文献   

17.
Bacterioferritin (BFR) from Escherichia coli is a member of the ferritin family of iron storage proteins and has the capacity to store very large amounts of iron as an Fe(3+) mineral inside its central cavity. The ability of organisms to tap into their cellular stores in times of iron deprivation requires that iron must be released from ferritin mineral stores. Currently, relatively little is known about the mechanisms by which this occurs, particularly in prokaryotic ferritins. Here we show that the bis-Met-coordinated heme groups of E. coli BFR, which are not found in other members of the ferritin family, play an important role in iron release from the BFR iron biomineral: kinetic iron release experiments revealed that the transfer of electrons into the internal cavity is the rate-limiting step of the release reaction and that the rate and extent of iron release were significantly increased in the presence of heme. Despite previous reports that a high affinity Fe(2+) chelator is required for iron release, we show that a large proportion of BFR core iron is released in the absence of such a chelator and further that chelators are not passive participants in iron release reactions. Finally, we show that the catalytic ferroxidase center, which is central to the mechanism of mineralization, is not involved in iron release; thus, core mineralization and release processes utilize distinct pathways.  相似文献   

18.
Phosvitin, a phosphoprotein known as an iron-carrier in egg yolk, binds almost all the yolk iron. In this study, we investigated the effect of phosvitin on Fe(II)-catalyzed hydroxyl radical ((.-)OH) formation from H(2)O(2) in the Fenton reaction system. Using electron spin resonance (ESR) with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO) and deoxyribose degradation assays, we observed by both assays that phosvitin more effectively inhibited (.-)OH formation than iron-binding proteins such as ferritin and transferrin. The effectiveness of phosvitin was related to the iron concentration, indicating that phosvitin acts as an antioxidant by chelating iron ions. Phosvitin accelerates Fe(II) autoxidation and thus decreases the availability of Fe(II) for participation in the (.-)OH-generating Fenton reaction. Furthermore, using the plasmid DNA strand breakage assay, phosvitin protected DNA against oxidative damage induced by Fe(II) and H(2)O(2). These results provide insight into the mechanism of protection of the developing embryo against iron-dependent oxidative damage in ovo.  相似文献   

19.
Iron (2+ and 3+) is believed to transfer through the three-fold channels in the ferritin shell during iron deposition and release in animal ferritins. However, the rate of iron transit in and out through these channels has not been reported. The recent synthesis of [Fe(CN)6]3-, Prussian Blue (PB) and desferrioxamine (DES) all trapped within the horse spleen ferritin (HoSF) interior makes these measurements feasible. We report the rate of Fe2+ penetrating into the ferritin interior by adding external Fe2+ to [Fe(CN)6]3- encapsulated in the HoSF interior and measuring the rate of formation of the resulting encapsulated PB. The rate at which Fe2+ reacts with [Fe(CN)6]3- in the HoSF interior is much slower than the formation of free PB in solution and is proceeded by a lag period. We assume this lag period and the difference in rate represent the transfer of Fe2+ through the HoSF protein shell. The calculated diffusion coefficient, D approximately 5.8x10(-20) m2/s corresponds to the measured lag time of 10-20 s before PB forms within the HoSF interior. The activation energy for Fe2+ transfer from the outside solution through the protein shell was determined to be 52.9 kJ/mol by conducting the reactions at 10 approximately 40 degrees C. The reaction of Fe3+ with encapsulated [Fe(CN)6]4- also readily forms PB in the HoSF interior, but the rate is faster than the corresponding Fe2+ reaction. The rate for Fe3+ transfer through the ferritin shell was confirmed by measuring the rate of the formation of Fe-DES inside HoSF and an activation energy of 58.4 kJ/mol was determined. An attempt was made to determine the rate of iron (2+ and 3+) transit out from the ferritin interior by adding excess bipyridine or DES to PB trapped within the HoSF interior. However, the reactions are slow and occur at almost identical rates for free and HoSF-encapsulated PB, indicating that the transfer of iron from the interior through the protein shell is faster than the rate-limiting step of PB dissociation. The method described in this work presents a novel way of determining the rate of transfer of iron and possibly other small molecules through the ferritin shell.  相似文献   

20.
《Free radical research》2013,47(1):153-159
Ceruloplasmin (CP) effectively inhibited superoxide and ferritin-dependent peroxidation of phospholipid liposomes, using xanthine oxidase or gamma irradiation of water as sources of superoxide. In addition, CP inhibited superoxide-dependent mobilization of iron from ferritin. suggesting that CP inhibited lipid peroxidation by decreasing the availability of iron from ferritin. CP also exhibited some superoxide scavenging activity as evidenced by its inhibition of superoxide-dependent cytochrome c reduction. However, superoxide scavenging by CP did not quantitatively account for its inhibitory effects on iron release. The effects of CP on iron-catalyzed lipid peroxidation in systems containing exogenously added ferrous iron was also investigated. CP exhibited prooxidant and antioxidant effects; CP stimulated at lower concentrations, reached a maximum. and inhibited at higher concentrations. However. the addition of apoferritin inhibited CP and Fe(II)-catalyzed lipid peroxidation at all concentrations of CP. In addition, CP catalyzed the incorporation of Fe(II) into apoferritin. Collectively these data suggest that CP inhibits superoxide and ferritin-dependent lipid peroxidation via its ability to incorporate reductively-mobilized iron into ferritin.  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