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.
5-Aminolevulinic acid (ALA), a heme precursor overproduced in various porphyric disorders, has been implicated in iron-mediated oxidative damage to biomolecules and cell structures. From previous observations of ferritin iron release by ALA, we investigated the ability of ALA to cause oxidative damage to ferritin apoprotein. Incubation of horse spleen ferritin (HoSF) with ALA caused alterations in the ferritin circular dichroism spectrum (loss of a alpha-helix content) and altered electrophoretic behavior. Incubation of human liver, spleen, and heart ferritins with ALA substantially decreased antibody recognition (51, 60, and 28% for liver, spleen, and heart, respectively). Incubation of apoferritin with 1-10mM ALA produced dose-dependent decreases in tryptophan fluorescence (11-35% after 5h), and a partial depletion of protein thiols (18% after 24h) despite substantial removal of catalytic iron. The loss of tryptophan fluorescence was inhibited 35% by 50mM mannitol, suggesting participation of hydroxyl radicals. The damage to apoferritin had no effect on ferroxidase activity, but produced a 61% decrease in iron uptake ability. The results suggest a local autocatalytic interaction among ALA, ferritin, and oxygen, catalyzed by endogenous iron and phosphate, that causes site-specific damage to the ferritin protein and impaired iron sequestration. These data together with previous findings that ALA overload causes iron mobilization in brain and liver of rats may help explain organ-specific toxicities and carcinogenicity of ALA in experimental animals and patients with porphyria.  相似文献   

2.
5-Aminolevulinic acid (ALA), a heme precursor that accumulates in acute intermittent porphyria (AIP) and lead poisoning, undergoes enolization and subsequent iron-catalyzed oxidation at neutral pH. Iron is released from horse spleen ferritin (HoSF) by both ALA-generated O2•− and enoyl radical (ALA√), which amplifies the chain of ALA oxidation (autocatalysis). Iron chelators such as EDTA, ATP, but not citrate, and phosphate accelerate this process and ALA-promoted iron release from HoSF is faster in horse spleen isoferritins containing larger amounts of phosphate in the core. ALA (+0.377 V versus standard hydrogen electrode) is less effective in releasing iron from ferritin than are thioglycollic acid, 6-hydroxydopamine, and N,N,N′,N′-tetramethyl-p-phenylenediamine. During electrochemical one electron oxidation of ALA in a nitrogen atmosphere, spin trapping experiments with 3,5-dibromo-4-nitrosobenzenesulfonic acid demonstrated the formation of a spin adduct characterized by a six line signal, indicating a secondary carbon-centered radical and attributed to a resonant ALA√ radical. Iron is also released in such anaerobic electrochemical oxidations of ALA in the presence of ferritin, suggesting that, in addition to O2•−, ALA√ can promote iron mobilization from ferritin. Hence, ALA√ may amplify the metal-catalyzed oxidation of ALA, damaging ALA-accumulating cells and possibly contributing to the symptoms of porphyria.  相似文献   

3.
Subunit dimers in sheep spleen apoferritin. The effect on iron storage   总被引:6,自引:0,他引:6  
Ferritin with high and low iron content, 2000 and 790 iron atoms/molecule, was isolated from the spleens of copper-poisoned and control lambs, respectively. Differences in the iron content in vivo were reflected in the properties of the apoferritin protein shells, since the apoprotein from the low iron ferritin took up iron relatively more slowly (0.52 +/- 0.09) and released it more rapidly (1.68 +/- 0.06) in vitro. Although the two types of apoferritin were indistinguishable in terms of surface charge (pI range 4.98-5.43) and in consisting of both heavy and light subunits, the subunit interactions differed markedly; 40-50% of the subunits of low iron ferritin were in dimers stable to reduction and carboxylmethylation, 4% mercaptoethanol, 8% sodium dodecyl sulfate, and 100 degrees C for 30 min, 70% formic acid, and 30% methanol. Subunit dimers were also observed in liver ferritin from mouse and neonatal pig and were enriched in a low iron fraction of horse spleen ferritin. Based on cyanogen bromide fragmentation and NH2-terminal analysis, the natural and chemically cross-linked subunit dimers had two peptides in common; natural subunit dimers also appeared to have a second region cross-linked, suggesting the possibility of both intra- and intersubunit links in the natural dimers. In sheep spleen ferritin, both heavy and light subunits appeared to participate in subunit dimerization. Natural subunit dimers were enriched in low iron ferritin fractions of all ferritin preparations tested (linear correlation = 0.94) and can explain, at least in part, the previously observed effects of iron core size on the apoferritin shell. Whether the subunit cross-links represent part of the subunit assembly process subsequently cleaved by iron (or copper) or whether the cross-links form after iron core formation in vivo has yet to determined. In either case, it is clear that such post-translational variations can affect iron uptake and release and emphasize the importance of the protein shell in determining the iron storage properties of ferritin.  相似文献   

4.
The ferritin superfamily of iron storage proteins includes ferritin proper and Dps (DNA binding protein from starved cells) along with bacterioferritin. We examined the release of Fe from the Dps of Trichodesmium erythraeum (Dps(tery)) and compared it to the release of Fe from horse spleen ferritin (HoSF) under various conditions. Both desferrioxamine B (DFB), a Fe(III) chelator, and ascorbic acid were able to mobilize Fe from Dps(tery) at rates comparable to those observed for HoSF. The initial Fe release rate from both proteins increased linearly with the concentration of DFB, suggesting that the chelator binds to Fe in the protein. A small but significant rate obtained by extrapolation to zero concentration of DFB implies that Dps(tery) and HoSF might release Fe(III) spontaneously. A similar result was observed for HoSF in the presence of sulfoxine. In a different experiment, Fe(III) was transferred from holoferritin to apotransferrin across a dialysis membrane in the absence of chelator or reducing agent. The apparent spontaneous release of Fe from HoSF and Dps(tery) brings forth the hypothesis that the Fe core in Fe storage proteins might be continuously dissolving and re-precipitating in vivo, thus maintaining it in a highly reactive and bioavailable form.  相似文献   

5.
Uptake of iron by apoferritin from a ferric dihydrolipoate complex   总被引:1,自引:0,他引:1  
A study was made on the uptake of iron by horse spleen apoferritin, by using as an iron source the same ferric dihydrolipoate complex which represents the major product in the anaerobic removal of ferritin-bound iron by dihydrolipoate at neutral pH. The ferric dihydrolipoate complex was chemically synthesized and used as an iron donor to apoferritin. Iron uptake was studied, at slightly alkaline pH and in anaerobic conditions, as a function of the concentration of both the iron donor and apoferritin. Isolation of ferritin from mixtures of ferric dihydrolipoate and apoferritin, and subsequent identification of the oxidation state of ferritin-bound iron, showed that the first metal atoms were taken up in the ferrous form and that this early step was accompanied by accumulation of ferric iron. Total iron uptake increased with the molar ratio of complex to apoprotein and ranged over 25-40% of the iron being supplied. The amount of ferrous iron found inside the protein did not exceed 50-60 mol iron/mol ferritin after a 48-h incubation. At this time, ferric iron represented a significant fraction of the iron found in the isolated ferritin. Analytical and spectroscopic data indicated that fractional rates and equilibria for disassembly of the ferric complex in the presence of apoferritin were independent of the concentration of the protein and of the complex itself.  相似文献   

6.
Apolipoprotein B (apoB) is known to be a ferritin-binding protein. Here we show that apoB binds to ferritin through hemin-mediated binding. Human apoB bound to bovine spleen, horse spleen, and canine liver ferritins, but did not bind to bovine apoferritin, even after incorporation of iron into it. Incubation of apoferritin with hemin resulted in apoB binding with apoferritin at the same level as with holoferritin. In contrast, hemin inhibited binding of apoB to ferritin. Bovine spleen apoferritin bound biotinylated hemin, and hemin inhibited the binding between the apoferritin and biotinylated hemin, suggesting that ferritin binds hemin directly. ApoB and LDL containing apoB bound biotinylated hemin, and their bindings were also inhibited by hemin, but not protoporphyrin IX. These data demonstrate that binding of apoB to ferritin is mediated through ferritin’s binding to hemin, and also that apoB binds hemin directly.  相似文献   

7.
Variations in the turnover of storage iron have been attributed to differences in apoferritin and in the cytoplasm but rarely to differences in the structure of the iron core (except size). To explore the idea that the iron environment in soluble iron complexes could vary, we compared horse spleen ferritin to pharmaceutically important model complexes of hydrous ferric oxide formed from FeCl3 and dextran (Imferon) or chondroitin sulfate (Blutal), using x-ray absorption (EXAFS) and M?ssbauer spectroscopy. The results show that the iron in the chondroitin sulfate complex was more ordered than in either horse spleen ferritin or the dextran complex (EXAFS), with two magnetic environments (M?ssbauer), one (80%-85%) like Fe2O3 X nH2O (ferritinlike) and one (15%-20%) like Fe2O3 (hematite); since sulfate promotes the formation of inorganic hematite, the sulfate in the chondroitin sulfate most likely nucleated Fe2O3 and hydroxyl/carboxyls, which are ligands common to chondroitin sulfate, ferritin and dextran most likely nucleated Fe2O3 X nH2O. Differences in the structure of the iron complexed with chondroitin sulfate or dextran coincide with altered rates of iron release in vivo and in vitro and provide the first example relating function to local iron structure. Differences might also occur among ferritins in vivo, depending on the apoferritin (variations in anion-binding sites) or the cytoplasm (anion concentration).  相似文献   

8.
Ferritins are a class of iron storage protein spheres found mainly in the liver and spleen, which have attracted many research interests due to their unique structural features and biological properties. Recently, ferritin and apoferritin (ferritin devoid of the iron core), have been employed as chemically addressable nanoscale building blocks for functional materials development. However, the reactive residues of apoferritin or ferritin have never been specified and it is still unclear about the chemoselectivity of apoferritin towards different kinds of bioconjugation reagents. In this work, matrix-assisted laser desorption/ionization time-of-flight (MALDI-TOF) mass spectrometry combined with enzymatic digestion analysis was used to identify the reactive lysine residues of horse spleen apoferritin when conjugated with N-hydroxysuccinimide reagents. The result demonstrated that among all the lysine residues, K97, K83, K104, K67 and K143 are the reactive ones that can be addressed.  相似文献   

9.
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   

10.
Apo horse spleen ferritin (apo HoSF) was reconstituted to various core sizes (100-3500 Fe3+/HoSF) by depositing Fe(OH)3 within the hollow HoSF interior by air oxidation of Fe2+. Fe2+ and phosphate (Pi) were then added anaerobically at a 1:4 ratio, and both Fe2+ and Pi were incorporated into the HoSF cores. The resulting Pi layer consisted of Fe2+ and Pi at about a 1:3 ratio which is strongly attached to the reconstituted ferritin mineral core surface and is stable even after air oxidation of the bound Fe2+. The total amount of Fe2+ and Pi bound to the iron core surface increases as the core volume increases up to a maximum near 2500 iron atoms, above which the size of the Pi layer decreases with increasing core size. M?ssbauer spectroscopic measurements of the Pi-reconstituted HoSF cores using 57Fe2+ show that 57Fe3+ is the major species present under anaerobic conditions. This result suggests that the incoming 57Fe2+ undergoes an internal redox reaction to form 57Fe3+ during the formation of the Pi layer. Addition of bipyridine removes the 57Fe3+ bound in the Pi layer as [57Fe(bipy)3]2+, showing that the bound 57Fe2+ has not undergone irreversible oxidation. This result is related to previous studies showing that 57Fe2+ bound to native core is reversibly oxidized under anaerobic conditions in native holo bacterial and HoSF ferritins. Attempts to bury the Pi layer of native or reconstituted HoSF by adding 1000 additional iron atoms were not successful, suggesting that after its formation, the Pi layer "floats" on the developing iron mineral core.  相似文献   

11.
Ferritin utilizes ferroxidase activity to incorporate iron. Iron uptake kinetics of bovine spleen apoferritin (H: L = 1 : 1.1) were compared with those of recombinant H chain ferritin and L chain ferritin homopolymers. H chain ferritin homopolymer showed an iron uptake rate identical to bovine spleen apoferritin (0.19 and 0.21 mmol/min/micromol of protein, respectively), and both showed iron concentration-dependent uptake. In contrast, the L chain homopolymer, which lacks ferroxidase, did not incorporate iron and showed the same level of iron autoxidation in the absence of ferritin. Bovine spleen apoferritin was shown to have two iron concentration-dependent uptake pathways over a range of 0.02-0.25 mM ferrous ammonium sulfate (FAS) by an Eadie-Scatchard plot (v/[FAS] versus v), whereas the H chain ferritin homopolymer was found to have only one pathway. Of the two Km values found in bovine spleen apoferritin, the lower mean Km value was 9.0 microM, while that of the H chain homopolymer was 11.0 microM. H chain ferritin homopolymer reached a saturating iron uptake rate at 0.1 mM FAS, while bovine spleen apoferritin incorporated more iron even at 0.25 mM FAS. These results suggest that the intrinsic ferroxidase of ferritin plays a significant role in iron uptake, and the L chain cooperates with the H chain to increase iron uptake.  相似文献   

12.
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   

13.

Background

Horse fibrinogen has been identified as a plasma specific ferritin-binding protein. There are two ways in the binding of ferritin-binding protein with ferritin: one is direct binding and the other is indirect binding which is heme-mediated. The aim of this study was to analyze the binding between horse fibrinogen and ferritin.

Findings

Although fibrinogen in horse plasma did not show the binding to ferritin coated on the plate wells, after following heat-treatment (60°C, 30 min) of horse plasma, plasma fibrinogen as well as purified horse fibrinogen bound to plates coated with horse spleen ferritin, but not with its apoferritin which lost heme as well as iron after the treatment of reducing reagent. Binding of purified or plasma fibrinogen to ferritin was inhibited by hemin and Sn-protoporphyrin IX (Sn-PPIX), but not by PPIX or Zn-PPIX.

Conclusions

Heat-treatment of horse plasma enabled plasma fibrinogen to bind to plate well coated with holo-ferritin. From the binding analysis of fibrinogen and ferritin, it is suggested that horse fibrinogen recognized iron or tin in complexed with the heme- or the hemin-ring, and also suggest that some fibrinogens circulate in the form of a complex with ferritin and/or heat-labile factors which inhibit the binding of fibrinogen with ferritin.
  相似文献   

14.
The in vitro effects of four different species of arsenic (arsenate, arsenite, monomethylarsonic acid, and dimethylarsinic acid) in mobilizing iron from horse spleen ferritin under aerobic and anaerobic conditions were investigated. Dimethylarsinic acid (DMA(V)) and dimethylarsinous acid (DMA(III)) significantly released iron from horse spleen ferritin either with or without the presence of ascorbic acid, a strong synergistic agent. Ascorbic acid-mediated iron release was time-dependent as well as both DMA(III) and ferritin concentration-dependent. Iron release from ferritin by DMA(III)) alone or with ascorbic acid was not significantly inhibited by superoxide dismutase (150 or 300 units/ml). However, the iron release was greater under anaerobic conditions (nitrogen gas), which indicates direct chemical reduction of iron from ferritin by DMA(III), with or without ascorbic acid. Both DMA(V) and DMA(III)) released iron from both horse spleen and human liver ferritin. Further, the release of ferritin iron by DMA(III)) with ascorbic acid catalyzed bleomycin-dependent degradation of calf thymus DNA. These results indicate that exogenous methylated arsenic species and endogenous ascorbic acid can cause (a) the release of iron from ferritin, (b) the iron-dependent formation of reactive oxygen species, and (c) DNA damage. This reactive oxygen species pathway could be a mechanism of action of arsenic carcinogenesis in man.  相似文献   

15.
Large-molecule oxidants oxidize Fe(II) to form Fe(III) cores in the interior of ferritins at rates comparable to or faster than the iron deposition reaction using O(2) as oxidant. Iron deposition into horse spleen ferritin (HoSF) occurs using ferricyanide ion, 2,6-dichlorophenol-indophenol, and several redox proteins: cytochrome c, stellacyanin, and ceruloplasmin. Cytochrome c also loads iron into recombinant human H-chain (rHF), human L-chain (rLF), and A. vinelandii bacterioferritin (AvBF). The enzymatic activities of ferritins were monitored anaerobically using stopped-flow kinetic spectrophotometry. The reactions exhibit saturation kinetics with respect to the large oxidant concentrations, giving apparent Michaelis constants for cytochrome c as oxidant: K(m)=39.6 microM for HoSF and 6.9 microM for AvBF. Comparison of the kinetic parameters with that of iron deposition by O(2) shows that large oxidants load iron into HoSF and AvBF more effectively than O(2) and may use a mechanism different than the ferroxidase center. Large oxidants did not deposit iron as efficiently with rHF and rLF. The results suggest that the heme groups in AvBF and the protein redox centers present in heteropolymers may assist in anaerobic iron deposition by large oxidants. The physiological relevance of iron deposition by large molecules, including protein oxidants is discussed.  相似文献   

16.
Conformational changes were induced in human spleen ferritin by partial or complete removal of iron, and the immunoreactivity of the ferritin samples with variable iron content was analyzed. We established that a decrease in iron content resulted in bimodal changes in immunoreactivity of the epitopes recognized by the monoclonal antibodies G10 and F11. Immunoreactivity demonstrated a 3-6-fold decrease on lowering iron content from 800 to 40 atoms per protein molecule, followed by a sharp (4-14-fold) increase that was observed when low-iron ferritin was converted to iron-free apoferritin. These bimodal changes suggest the presence of more than two conformational states of ferritin with local alterations of the epitopes recognized by the monoclonal antibodies. The global conformation of ferritin, however, remained essentially unaltered, as demonstrated by ferritin interaction with polyclonal antibodies. Together, the results indicate that local conformational changes in the ferritin protein shell occur on progressive iron removal that results in low-iron and iron-free forms of ferritin. These changes are most clearly seen in apoferritin when compared to low-iron ferritin.  相似文献   

17.
Zhao G  Bou-Abdallah F  Yang X  Arosio P  Chasteen ND 《Biochemistry》2001,40(36):10832-10838
The ferritins are a class of iron storage and detoxification proteins that play a central role in the biological management of iron. These proteins have a catalytic site, "the ferroxidase site", located on the H-type subunit that facilitates the oxidation of Fe(II) to Fe(III) by O(2). Measurements during the past 10 years on a number of vertebrate ferritins have provided evidence that H(2)O(2) is produced at this diiron ferroxidase site. Recently reported experiments using three different analytical methods with horse spleen ferritin (HoSF) have failed to detect H(2)O(2) production in this protein [Lindsay, S., Brosnahan, D., and Watt, G. D. (2001) Biochemistry 40, 3340-3347]. These findings contrast with earlier results reporting H(2)O(2) production in HoSF [Xu, B., and Chasteen, N. D. (1991) J. Biol. Chem. 266, 19965-19970]. Here a sensitive fluorescence assay and an assay based on O(2) evolution in the presence of catalase were used to demonstrate that H(2)O(2) is produced in HoSF as previously reported. However, because of the relatively few H-chain ferroxidase sites in HoSF and the reaction of H(2)O(2) with the protein, H(2)O(2) is more difficult to detect in this ferritin than in recombinant human H-chain ferritin (HuHF). The proper sequence of addition of reagents is important for measurement of the total amount of H(2)O(2) produced during the ferroxidation reaction.  相似文献   

18.
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   

19.
猪脾和马脾铁蛋白理化特性的比较   总被引:10,自引:1,他引:9  
黄河清  张凤章 《动物学报》1997,43(2):170-177
H^+,OH^-均能参与猪脾和马脾铁蛋白铁核组成,迫使它们分别释放铁核中对酸碱不稳定的铁组份。在可见光谱中,猪脾和马脾铁蛋释放铁的动力学过程可分为一级快速反应和零级慢速反应,但猪脾铁蛋白释放铁一级反应速度明显大于马脾铁蛋白释放铁的一级反应的速率,推测这些现象均与各自蛋白的蛋白壳自身调节能力有着密切联系。  相似文献   

20.
Methylglyoxal (MG) is an endogenous metabolite which is present in increased concentrations in diabetics and reacts with amino acids to form advanced glycation end products. In this study, we investigated whether ferritin enhances DNA cleavage by the reaction of MG with lysine. When plasmid DNA was incubated with MG and lysine in the presence of ferritin, DNA strand breakage was increased in a dose-dependent manner. The ferritin/MG/lysine system-mediated DNA cleavage was significantly inhibited by reactive oxygen species (ROS) scavengers. These results indicated that ROS might participate in the ferritin/MG/lysine system-mediated DNA cleavage. Incubation of ferritin with MG and lysine resulted in a time-dependent release of iron ions from the protein molecules. Our data suggest that DNA cleavage caused by the ferritin/MG/lysine system via the generation of ROS by the Fenton-like reaction of free iron ions released from oxidatively damaged ferritin. [BMB Reports 2013; 46(4): 225-229]  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