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 15 毫秒
1.
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   

2.
The effects of various maneuvers on the handling of 59Fe-labeled heat-damaged red cells (59Fe HDRC) by the reticuloendothelial system were studied in rats. Raising the saturation of transferrin with oral carbonyl iron had little effect on splenic release of 59Fe but markedly inhibited hepatic release. Splenic 59Fe release was, however, inhibited by the prior administration of unlabeled HDRC or by the combination of carbonyl iron and unlabeled HDRC. When carbonyl iron was administered with unlabeled free hemoglobin, the pattern of 59Fe distribution was the same as that observed when carbonyl iron was given alone. 59Fe ferritin was identified in the serum after the administration of 59Fe HDRC but the size of the fraction was not affected by raising the saturation of transferrin. Sizing column analyses of tissue extracts from the spleen at various times after the administration of 59Fe HDRC revealed a progressive shift from hemoglobin to ferritin, with only small amounts present in a small molecular weight fraction. The small molecular weight fraction was greater in hepatic extracts, with the difference being marked in animals that had received prior carbonyl iron. The increased hepatic retention of 59Fe associated with a raised saturation of transferrin was reduced by a hydrophobic ferrous chelator (2,2'-bipyridine), a hydrophilic ferric chelator (desferrioxamine), and an extracellular hydrophilic ferric chelator (diethylene-triaminepentacetic acid). Transmembrane iron transport did not seem to be a rate-limiting factor in iron release, since no differences in 59Fe membrane fractions were noted in the different experimental settings. These findings are consistent with a model in which RE cells release iron from catabolized red cells at a relatively constant rate. When the saturation of transferrin is raised, a significant proportion of the iron is transported from the spleen to the liver either in small molecular weight complexes or in ferritin. Although a saturated transferrin had no effect on the release of iron from reticuloendothelial cells, prior loading with HDRC conditions them to release less iron.  相似文献   

3.
Membrane transport of non-transferrin-bound iron by reticulocytes   总被引:8,自引:0,他引:8  
The transport of non-transferrin-bound iron into rabbit reticulocytes was investigated by incubating the cells in 0.27 M sucrose with iron labelled with 59Fe. In most experiments the iron was maintained in the reduced state, Fe(II), with mercaptoethanol. The iron was taken up by cytosolic, haem and stromal fractions of the cells in greater amounts than transferrin-iron. The uptake was saturable, with a Km value of approx. 0.2 microM and was competitively inhibited by Co2+, Mn2+, Ni2+ and Zn2+. It ceased when the reticulocytes matured into erythrocytes. The uptake was pH and temperature sensitive, the pH optimum being 6.5 and the activation energy for iron transport into the cytosol being approx. 80 kJ/mol. Ferric iron and Fe(II) prepared in the absence of reducing agents could also be transported into the cytosol. Sodium chloride inhibited Fe(II) uptake in a non-competitive manner. Similar degrees of inhibition was found with other salts, suggesting that this effect was due to the ionic strength of the solution. Iron chelators inhibited Fe(II) uptake by the reticulocytes, but varied in their ability to release 59Fe from the cells after it had been taken up. Several lines of evidence showed that the uptake of Fe(II) was not being mediated by transferrin. It is concluded that the reticulocyte can transport non-transferrin-bound iron into the cytosol by a carrier-mediated process and the question is raised whether the same carrier is utilized by transferrin-iron after its release from the protein.  相似文献   

4.
Rat liver mitochondria and rat liver mitoplasts mobilize iron from ferritin by a mechanism which depends on a respiratory substrate (preferentially succinate), a small molecular weight electron mediator (FMN, phenazine methosulphate or methylene blue) and (near) anaerobic conditions.The release process under optimized conditions (approx. 50 μmol/l FMN, 1 mmol/l succinate, 0.35 mmol/l Fe(III) (as ferritin iron), 37°C and pH 7.40) amounts to 0.9–1.2 nmol iron/mg protein per min.The results suggest that ferritin might function as an intermediate in the cytosolic transport of iron to the mitochondria.  相似文献   

5.
The removal of ferritin-bound iron by the physiologic dithiol DL-dihydrolipoate was studied over the pH range 5.5-9.0. A novel method was devised for the determination of iron removal, making it possible to study the actual release of iron from ferritin, regardless of the oxidation state or complexation form. The overall iron-removal process appears to depend upon a balance between the deprotonation of the dithiol and the protolytic dissolution of the iron core inside the ferritin molecule. The amount of iron removed at equilibrium increases with the pH, at any of the dihydrolipoate/ferritin iron ratios tested. The formation of the binuclear iron-dithiol complex [Fe2(dihydrolipoate)3]-3 is not strictly required for iron mobilization, but it seems to affect the efficiency of the dithiol in iron mobilization by providing a stable complexation form for the released iron outside the ferritin protein shell. Comparison of the release of ferritin-bound iron by free and immobilized dihydrolipoate indicates that mobility of the dithiol is mandatory for the removal process to take place.  相似文献   

6.
Apo-lactoferrin and apo-transferrin protect against iron-ion-dependent hydroxyl-radical (.OH) generation from H2O2 in the presence of superoxide radicals or ascorbic acid at pH 7.4, whether the necessary iron is added as ionic iron or as ferritin. Iron-loaded transferrin and lactoferrin [2 mol of Fe(III)/mol] show no protective ability, but do not themselves accelerate .OH production unless chelating agents are present in the reaction mixture, especially if the proteins are incorrectly loaded with iron. At acidic pH values, the protective ability of the apoproteins is diminished, and the fully iron-loaded proteins can release some iron in a form able to accelerate .OH generation. The physiological significance of these observations is discussed.  相似文献   

7.
The utilization of ferritin as a source of iron for the ferrochelatase reaction has been studied in isolated rat liver mitochondria. 1. It was found that isolated rat liver mitochondria utilized ferritin as a source of iron for the ferrochelatase reaction in the presence of succinate plus FMN (or FAD). 2. Under optimal experimental conditions, i.e., approx. 50 micromol/1 FMN, 37 degrees C, pH 7.4 and 0.5 mmol/l Fe(III) (as ferritin iron), the release process, as shown by the formation of deuteroheme, amounted to approx. 0.5 nmol iron/min per mg protein. 3. The release process could not be elicited by ultrasonically treated mitochondria, lysosomes, microsomes or cytosol, i.e., the release of iron from ferritin was due to mitochondria and was a function of the in situ orientation of the mitochondrial inner membrane. 4. The release of iron from ferritin by the mitochrondria might be of relevance not only for the in situ synthesis of heme in the hepatocyte, but also with respect to the mechanism(s) by means of which iron is mobilized for transport to the erythroid tissue.  相似文献   

8.
We investigated the iron release from ferritin by irradiation from a white fluorescent light in the absence or presence of ADP. Irradiation of a ferritin solution at 17,000 lx in the absence of ADP slightly induces iron release from ferritin but only at acidic pH conditions (pH 5.0 or pH 6.0). Irradiation in the presence of ADP markedly enhances iron release from ferritin under the same conditions. In the absence of irradiation, the iron release from ferritin was low even in the presence of ADP. The induction of the iron release by irradiation in the presence of ADP was also affected by various factors such as irradiation dose and acidity, but not temperature (4-47°C), oxygen concentration, or free radical generations during the irradiation. The iron release during the irradiation ceased to increase by turning off the light and was found to increase again after additional irradiation. These results suggest that visible light directly induces iron release from ferritin via the photoreduction of iron stored inside ferritin.  相似文献   

9.
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   

10.
On the limited ability of superoxide to release iron from ferritin   总被引:3,自引:0,他引:3  
Reductive release of iron from ferritin may catalyze cytotoxic radical reactions like the Haber-Weiss reaction. The ability of .O2- to mobilize Fe(II) from ferritin was studied by using the xanthine/xanthine oxidase reaction, with and without superoxide dismutase, and with bathophenanthroline sulphonate as the chelator. Not more than one or two Fe(II)/ferritin molecules could be released by an .O2(-)-dependent mechanism, even after repeated exposures of ferritin to bursts of .O2-. The amount of releaseable iron depended on the size and the age of the iron core, but not on the iron content of the protein shell of ferritin which was manipulated by chelators and addition of FeCl3. The kinetic characteristics of the .O2(-)-mediated iron release indicated the presence of a small pool of readily available iron at the surface of the core. The very limited .O2(-)-dependent release of iron from ferritin is compatible with a protective role of ferritin against toxic iron-catalyzed reactions.  相似文献   

11.
The carbohydrate composition of horse spleen ferritin was studied. 1 mol of the apoferritin, the protein moiety of ferritin, contains 25 mol of hexose, 3 mol of hexosamine and 10 mol of fucose. Same carbohydrate composition was detected in the apoferritin from iron rich ferritins. These results indicate that horse spleen ferritin is composed of non-identical subunits as regards its carbohydrate composition.  相似文献   

12.
To understand the function of the Fe2+-complexing compound nicotianamine (NA) in the iron metabolism of plants we have localized iron and other elements in the NA-containing tomato wild type (Lycopersicon esculentum) and its NA-free mutant chloronerva by quantitative x-ray microanalysis. Comparison of element composition of the rhizodermal cell walls indicated that the wild type accumulated considerable amounts of iron and phosphorus in the cell wall, whereas in the mutant iron and phosphorus were detected in the cytoplasm and vacuoles of the rhizodermis. In mutant leaves containing high iron concentrations in the symplast, electron-dense inclusions were detected in chloroplasts and phloem. Such particles, consisting mainly of iron and phosphorus, were never found in the wild type and were very rarely detected in young chlorotic mutant leaves or after treatment of the mutant with NA. For further characterization the electron-dense inclusions in mutant leaves were isolated and compared by sodium dodecyl sulfate-gel electrophoresis and immunoblotting to ferritin from iron-loaded Phaseolus vulgaris leaves. Antibodies raised against purified Phaseolus leaf ferritin were used. Neither in mutant nor in wild type (iron loaded and control) was ferritin protein detected. These results suggest that the electron-dense inclusions in mutant leaves are not identical with ferritin. It is concluded that NA is necessary to complex ferrous iron in a soluble and available form within the cells. In the absence of NA the precipitation of excessive iron in the form of insoluble ferric phosphate compounds could protect the cells from iron overload.  相似文献   

13.
The ability of transferrin to potentiate oxygen free radical-mediated endothelial cell injury was assessed. 51Cr-labeled endothelial cells derived from rat pulmonary arteries (RPAECs) were incubated with hydrogen peroxide (H2O2) in the presence and absence of holosaturated human transferrin, and the effect of transferrin on H2O2-mediated endothelial cell toxicity was determined. Addition of holosaturated transferrin potentiated H2O2-mediated RPAEC cytotoxicity at concentrations of H2O2 greater than 10 microM, suggesting that transferrin may provide a source of iron for free radical-mediated endothelial cell injury. Free radical-mediated injury is dependent on non-protein-bound iron. The ability of RPAECs to facilitate the release of iron from transferrin was assessed. We determined that RPAECs facilitate the release of transferrin-derived iron by reduction of transferrin-bound ferric iron (Fe3+) to ferrous iron (Fe2+). The reduction and release of transferrin-derived Fe2+ were inhibited by apotransferrin and chloroquine, indicating a dependence on receptor-specific binding of transferrin to the RPAEC cell surface, with subsequent endocytosis, acidification, and reduction of transferrin-bound Fe3+ to Fe2+. The release of transferrin-derived Fe2+ was potentiated by diethyldithiocarbamate, an inhibitor of intracellular superoxide dismutase (SOD). In contrast, exogenous SOD did not alter iron release, suggesting that intracellular superoxide anion (O2-) may play an important role in mediating the reduction and release of transferrin-derived iron. Results of this study suggest that transferrin may provide a source of iron for oxygen free radical-mediated endothelial cell injury and identify a novel mechanism by which endothelial cells may mediate the reduction and release of transferrin-derived iron.  相似文献   

14.
Intracellular ferritin in newt (Triturus cristatus) erythroblasts was accessible to the chelating effects of EDTA and pyridoxal phosphate. EDTA (0.5-1 mM) promoted release of radioactive iron from ferritin of pulse-labelled erythroblasts during chase incubation, but its continuous presence was not necessary for ferritin iron mobilization. Brief exposure to EDTA was sufficient to release 60-70% of ferritin 59Fe content during ensuing chase in EDTA-free medium. EDTA also suppressed cellular iron uptake and utilization for heme synthesis, but these activities were restored upon its removal. Pyridoxal-5'-phosphate (0.5-5 mM) also stimulated loss of radioactive iron from ferritin; however, ferritin iron release by pyridoxal phosphate required its continued presence. Unlike EDTA, pyridoxal phosphate did not interfere with iron uptake or its utilization for heme synthesis. Chelator-mobilized ferritin iron accumulated initially in the hemolysate as a low-molecular-weight component and appeared to be eventually released into the medium. No radioactive ferritin was found in the medium of chelator-treated cells, indicating that secretion or loss of ferritin was not responsible for decreasing cellular ferritin 59Fe content. Moreover, there was no transfer of radioactive iron between the low-molecular-weight component released into the medium and plasma transferrin. These results indicate that chelator-released ferritin iron is not available for cellular utilization in heme synthesis and that ferritin iron released by this process is not an alternative or complementary iron source for heme synthesis. Correlation of these data with effects of succinylacetone inhibition of heme synthesis and with previous studies indicates that the main role of erythroid cell ferritin is absorption and storage of excess iron not used for heme synthesis.  相似文献   

15.
Ferritin iron loading was studied in the presence of physiological serum phosphate concentrations (1 mM), elevated serum concentrations (2–5 mM), and intracellular phosphate concentrations (10 mM). Experiments compared iron loading into homopolymers of H and L ferritin with horse spleen ferritin. Prior to studying the reactions with ferritin, a series of control reactions were performed to study the solution chemistry of Fe2+ and phosphate. In the absence of ferritin, phosphate catalyzed Fe2+ oxidation and formed soluble polymeric Fe(III)-phosphate complexes. The Fe(III)-phosphate complexes were characterized by electron microscopy and atomic force microscopy, which revealed spherical nanoparticles with diameters of 10–20 nm. The soluble Fe(III)-phosphate complexes also formed as competing reactions during iron loading into ferritin. Elemental analysis on ferritin samples separated from the Fe(III)-phosphate complexes showed that as the phosphate concentration increased, the iron loading into horse ferritin decreased. The composition of the mineral that does form inside horse ferritin has a higher iron/phosphate ratio (~1:1) than ferritin purified from tissue (~10:1). Phosphate significantly inhibited iron loading into L ferritin, due to the lack of the ferroxidase center in this homopolymer. Spectrophotometric assays of iron loading into H ferritin showed identical iron loading curves in the presence of phosphate, indicating that the ferroxidase center of H ferritin efficiently competes with phosphate for the binding and oxidation of Fe2+. Additional studies demonstrated that H ferritin ferroxidase activity could be used to oxidize Fe2+ and facilitate the transfer of the Fe3+ into apo transferrin in the presence of phosphate.  相似文献   

16.
Ferric minerals in ferritins are protected from cytoplasmic reductants and Fe2+ release by the protein nanocage until iron need is signaled. Deletion of ferritin genes is lethal; two critical ferritin functions are concentrating iron and oxidant protection (consuming cytoplasmic iron and oxygen in the mineral). In solution, opening/closing (gating) of eight ferritin protein pores controls reactions between external reductant and the ferritin mineral; pore gating is altered by mutation, low heat, and physiological urea (1 mm) and monitored by CD spectroscopy, protein crystallography, and Fe2+ release rates. To study the effects of a ferritin pore gating mutation in living cells, we cloned/expressed human ferritin H and H L138P, homologous to the frog open pore model that was unexpressable in human cells. Human ferritin H L138P behaved like the open pore ferritin model in vitro as follows: (i) normal protein cage assembly and mineralization, (ii) increased iron release (t1/2) decreased 17-fold), and (iii) decreased alpha-helix (8%). Overexpression (> 4-fold), in HeLa cells, showed for ferritin H L138P equal protein expression and total cell 59Fe but increased chelatable iron, 16%, p < 0.01 (59Fe in the deferoxamine-containing medium), and decreased 59Fe in ferritin, 28%, p < 0.01, compared with wild type. The coincidence of decreased 59Fe in open pore ferritin with increased chelatable 59Fe in cells expressing the ferritin open pore mutation suggests that ferritin pore gating influences to the amount of iron (59Fe) in ferritin in vivo.  相似文献   

17.
In a previous study (Minotti, G., 1989, Arch. Biochem. Biophys. 268, 398-403) NADPH-supplemented microsomes were found to reduce adriamycin (ADR) to semiquinone free radical (ADR-.), which in turn autoxidized at the expense of oxygen to regenerate ADR and form O2-. Redox cycling of ADR was paralleled by reductive release of membrane-bound nonheme iron, as evidenced by mobilization of bathophenanthroline-chelatable Fe2+. In the present study, iron release was found to increase with concentration of ADR in a superoxide dismutase- and catalase-insensitive manner. This suggested that membrane-bound iron was reduced by ADR-. with negligible contribution by O2-. or interference by its dismutation product H2O2. Following release from microsomes, Fe2+ was reconverted to Fe3+ via two distinct mechanisms: (i) catalase-inhibitable oxidation by H2O2 and (ii) catalase-insensitive autoxidation at the expense of oxygen, which occurred upon chelation by ADR and increased with the ADR:Fe2+ molar ratio. Malondialdehyde formation, indicative of membrane lipid peroxidation, was observed when approximately 50% of Fe2+ was converted to Fe3+. This occurred in presence of catalase and low concentrations of ADR, which prevented Fe2+ oxidation and favored only partial Fe2+ autoxidation, respectively. Lipid peroxidation was inhibited by superoxide dismutase via increased formation of H2O2 from O2-. and excessive Fe2+ oxidation. Lipid peroxidation was also inhibited by high concentrations of ADR, which favored maximum Fe2+ release but also caused excessive Fe2+ autoxidation via formation of very high ADR:Fe2+ molar ratios. These results highlighted multiple and diverging effects of ADR, O2-., and H2O2 on iron release, iron (auto-)oxidation and lipid peroxidation. Stimulation of malondialdehyde formation by catalase suggested that lipid peroxidation was not promoted by reaction of Fe2+ with H2O2 and formation of hydroxyl radical. The requirement for both Fe2+ and Fe3+ was indicative of initiation by some type of Fe2+/Fe3+ complex.  相似文献   

18.
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   

19.
Liver depot iron can be divided into two fractions: ferritin iron and non-ferritin depot iron. Three methods intended to measure the non-ferritin depot iron in the rat liver were compared using livers of normal rats and livers of rats loaded with iron by transfusion of erythrocytes. Liver depot iron varied between 75 and 850 μg Fe/g liver. Non-ferritin depot iron, measured as the iron fraction sedimentable at 10 000 × g, was in the range 4–22 μg Fe/g liver. This fraction did contain ferritin. When measured as the difference between total liver depot iron and heat-stable iron (ferritin iron), the range was 10–270 μg Fe/g liver but this fraction also includes some ferritin iron.The values derived with both methods were linearly proportional to the total liver depot iron values.Non-ferritin depot iron, when measured as the difference between total liver depot iron and total ferritin iron, ranged from 0 to 190 μg Fe/g liver. In this last method no ferritin iron is included. This method provides the best estimate of the non-ferritin depot iron fraction. The concentrations obtained with this method were not always linearly proportional to the total liver depot iron concentration. Intravenous injection of rat liver ferritin resulted in a rapid accumulation of ferritin iron in the liver, together with an increase of the non-ferritin depot iron fraction from 18 μg Fe/g liver to 55 μg Ge/g liver. This confirms a relationship between ferritin catabolism and the non-ferritin depot iron fraction.  相似文献   

20.
Lindsay S  Brosnahan D  Watt GD 《Biochemistry》2001,40(11):3340-3347
The reaction of Fe2+ with O2 in the presence of horse spleen ferritin (HoSF) results in deposition of FeOH3 into the hollow interior of HoSF. This reaction was examined at low Fe2+/HoSF ratios (5-100) under saturating air at pH 6.5-8.0 to determine if H2O2 is a product of the iron deposition reaction. Three methods specific for H2O2 detection were used to assess H2O2 formation: (1) a fluorometric method with emission at 590 nm, (2) an optical absorbance method based on the reaction H2O2 + 3I- + 2H+ = I3- + 2H2O monitored at 340 nm for I3- formation, and (3) a differential pulsed electrochemical method that measures O2 and H2O2 concentrations simultaneously. Detection limits of 0.25, 2.5, and 5.0 microM H2O2 were determined for the three methods, respectively. Under constant air-saturation conditions (20% O2) and for a 5-100 Fe2+/HoSF ratio, Fe2+ was oxidized and the resulting Fe3+ was deposited within HoSF but no H2O2 was detected as predicted by the reaction 2Fe2+ + O2 + 6H2O = 2Fe(OH)3 + H2O2 + 4H+. Two other sets of conditions were also examined: one with excess but nonsaturating O2 and another with limiting O2. No H2O2 was detected in either case. The absence of H2O2 formation under these same conditions was confirmed by microcoulometric measurements. Taken together, the results show that under low iron loading conditions (5-100 Fe2+/HoSF ratio), H2O2 is not produced during iron deposition into HoSF using O2 as an oxidant. This conclusion is inconsistent with previous, carefully conducted stoichiometric and kinetic measurements [Xu, B., and Chasteen, N. D. (1991) J. Biol. Chem. 266, 19965], predicting that H2O2 is a quantitative product of the iron deposition reaction with O2 as an oxidant, even though it was not directly detected. Possible explanations for these conflicting results are considered.  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