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 93 毫秒
1.
Ferritin was isolated from bovine spleen and used to prepare apoferritin and reconstituted ferritin. The mol. wt of bovine ferritin was 464,000 with monomer subunits about 18,000-19,500. Gel electrophoresis showed three bands each for ferritin, apoferritin and reconstituted ferritin; all stained for protein and carbohydrate. Only apoferritin failed to stain for iron. Bovine ferritin had higher concentrations of proline, threonine, and valine than equine or human ferritin. The iron:protein ratio of bovine ferritin was 0.161 and of equine ferritin was 0.192. After iron uptake by the apoferritins the iron:protein ratios were 0.186 and 0.278 for the bovine and equine ferritins, respectively.  相似文献   

2.
Iron oxidation and incorporation into apoferritins of different subunit composition, namely the recombinant H and L homopolymers and the natural horse spleen heteropolymer (10-15% H), have been followed by steady-state and time-resolved fluorescence. After aerobic addition of 100 Fe(II) atoms/polymer, markedly different kinetic profiles are observed. In the rL-homopolymer a slow monotonic fluorescence quenching is observed which reflects binding, slow oxidation at the threefold apoferritin channels, and diffusion into the protein cavity. In the rH-homopolymer a fast fluorescence quenching is followed by a partial, slow recovery. The two processes have been attributed to Fe(II) binding and oxidation at the ferroxidase centers and to Fe(III) released into the cavity, respectively. The fluorescence kinetics of horse spleen apoferritin is dominated by the H chain contribution and resembles that of the H homopolymer. It brings out clearly that the rate of the overall process is limited by the rate at which Fe(III) leaves the ferroxidase centers of the H chains where binding of incoming Fe(II) and its oxidation take place. The data obtained upon stepwise addition of iron and the results of optical absorption measurements confirm this picture. The correspondence between steady-state and time-resolved data is remarkably good; this is manifest when the latter are used to calculate the change in fluorescence intensity as apparent in the steady-state measurements.  相似文献   

3.
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   

4.
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   

5.
6.
Iron incorporation by bovine spleen apoferritin either with ferrous ammonium sulfate in different buffers or with ferrous ammonium sulfate and phosphate was studied. Iron uptake and iron autoxidation were recorded spectrophotomerically. The buffers [4-(2-hydroxyethyl)-1-piperazinyl]ethanesulphonic acid (Hepes) and tris(hydroxymethyl)aminoethane (Tris) exhibited pH-dependent iron autoxidation, with Tris showing less iron autoxidation than Hepes. An Eadie-Scatchard plot (v/[s] versus v) of the iron uptake rate in Hepes was a curved rather than a straight line, suggesting that there are two iron uptake pathways. On the other hand, the Eadie-Scatchard plots of Tris and of Hepes after the addition of phosphate showed a straight line. Phosphate accelerated the iron uptake rate. The iron loading kinetics of apoferritin in Hepes was dependent on apoferritin concentration. The Km value obtained from iron uptake kinetics was 4.5 M, corresponding to the physiological iron concentration. These results demonstrate that iron loading of apoferritin was accomplished at physiological iron concentrations, which is essential for iron uptake, via two uptake pathways of dependent on iron concentration.  相似文献   

7.
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   

8.
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   

9.
Ferritin molecules contain 24 polypeptide chains folded as four-helix bundles and arranged as a hollow shell capable of storing up to 4500 Fe(III) atoms. H chains contain ferroxidase centres which lie within the bundle, about 12?Å (1.2?nm) from the outside surface and 8?Å from the inner surface of the protein shell. Catalysis of Fe(II) oxidation precedes storage of Fe(III) as ferrihydrite, with the formation of μ-oxo-bridged Fe(III) dimers as intermediates. Factors influencing the movement of μ-oxo-bridged Fe(III) from the ferroxidase centre to the ferritin cavity are uncertain. Assistance by small chelators is one possibility. The aim of this investigation was to determine whether iron at the dinuclear centres of three ferritins (human H chain homopolymer, HuHF, the non-haem ferritin of Escherichia coli, EcFTN, and horse spleen ferritin, HoSF) is accessible to chelators. Forty-eight Fe(II) atoms/molecule were added to the apoferritins followed, 2?min later, by the addition of chelator (1,10-phenanthroline, 2,2-bipyridine, desferrioxamine or 3,4-dihydroxybenzaldehyde). Iron species were analysed by Mössbauer spectroscopy or visible absorbance. Competition between chelators and apoferritin for Fe(II) was also investigated. The main conclusions of the study are that: (1) dinuclear iron and iron in small iron-cores in HuHF and EcFTN is mobilisable by all four chelators; (2) the chelators penetrate the shell; (3) 3,4-dihydroxybenzaldehyde is the most efficient in mobilising Fe(III) but the least successful in competing for Fe(II); (4) Fe(III) is more readily released from EcFTN than from HuHF; (5) 2,2′-bipyridine aids the movement of Fe(III) from ferroxidase centre to core.  相似文献   

10.
The human ferritin L-chain cDNA was cloned into a vector for overproduction in Escherichia coli, under the regulation of a lambda promoter. The plasmid obtained contains the full L-chain coding region modified at the first two codons. It is able to direct the synthesis of the L-chain which can constitute up to 15% of the total soluble protein of bacterial extract. The L-chains assemble to form a ferritin homopolymer with electrophoretic mobility, molecular weight, thermal stability, spectroscopic, and immunological properties analogous to natural ferritin from human liver (95% L-chain). This recombinant L-ferritin is able to incorporate and retain iron in solution at physiological pH values. At variance with the H-ferritin, the L form does not uptake iron at acidic pH values and does not show detectable ferroxidase activity. It is concluded that ferritin L-chain lacks the ferroxidase site present in the H-chain and that the two chains may have specialized functions in intracellular iron metabolism.  相似文献   

11.
Mineralization in Ferritin: An Efficient Means of Iron Storage   总被引:22,自引:0,他引:22  
Ferritins are a class of iron storage and mineralization proteins found throughout the animal, plant, and microbial kingdoms. Iron is stored within the protein shell of ferritin as a hydrous ferric oxide nanoparticle with a structure similar to that of the mineral "ferrihydrite." The eight hydrophilic channels that traverse the protein shell are thought to be the primary avenues by which iron gains entry to the interior of eukaryotic ferritins. Twenty-four subunits constitute the protein shell and, in mammalian ferritins, are of two types, H and L, which have complementary functions in iron uptake. The H chain contains a dinuclear ferroxidase site that is located within the four-helix bundle of the subunit; it catalyzes the oxidation of ferrous iron by O(2), producing H(2)O(2). The L subunit lacks this site but contains additional glutamate residues on the interior surface of the protein shell which produce a microenvironment that facilitates mineralization and the turnover of iron(III) at the H subunit ferroxidase site. Recent spectroscopic studies have shown that a di-Fe(III) peroxo intermediate is produced at the ferroxidase site followed by formation of a mu-oxobridged dimer, which then fragments and migrates to the nucleation sites to form incipient mineral core species. Once sufficient core has developed, iron oxidation and mineralization occur primarily on the surface of the growing crystallite, thus minimizing the production of potentially harmful H(2)O(2).  相似文献   

12.
Ferritin from the spleen of the Antarctic teleost Trematomus bernacchii is composed of a single subunit that contains both the ferroxidase center residues, typical of mammalian H chains, and the carboxylate residues forming the micelle nucleation site, typical of mammalian L chains. Comparison of the amino-acid sequence with those available from lower vertebrates indicates that T. bernacchii ferritin can be classified as an M-type homopolymer. Interestingly, the T. bernacchii ferritin chain shows 85.7% identity with a cold-inducible ferritin chain of the rainbow trout Salmo gairdneri. The structural and functional properties indicate that cold acclimation and functional adaptation to low temperatures are achieved without significant modification of the protein stability. In fact, the stability of T. bernacchii ferritin to denaturation induced by acid or temperature closely resembles that of mesophilic mammalian ferritins. Moreover iron is taken up efficiently and the activation energy of the reaction is 74.9 kJ.mol(-1), a value slightly lower than that measured for the human recombinant H ferritin (80.8 kJ.mol(-1)).  相似文献   

13.
Ferritins are iron storage proteins made of 24 subunits forming a hollow spherical shell. Vertebrate ferritins contain varying ratios of heavy (H) and light (L) chains; however, known ferritin structures include only one type of chain and have octahedral symmetry. Here, we report the 1.9A structure of a secreted insect ferritin from Trichoplusia ni, which reveals equal numbers of H and L chains arranged with tetrahedral symmetry. The H/L-chain interface includes complementary features responsible for ordered assembly of the subunits. The H chain contains a ferroxidase active site resembling that of vertebrate H chains with an endogenous, bound iron atom. The L chain lacks the residues that form a putative iron core nucleation site in vertebrate L chains. Instead, a possible nucleation site is observed at the L chain 3-fold pore. The structure also reveals inter- and intrasubunit disulfide bonds, mostly in the extended N-terminal regions unique to insect ferritins. The symmetrical arrangement of H and L chains and the disulfide crosslinks reflect adaptations of insect ferritin to its role as a secreted protein.  相似文献   

14.
To elucidate the physiological role of canine serum ferritin, we measured clearance rates of biotinylated ferritins in beagle. Biotinylated canine tissue ferritins were cleared rapidly from circulation. The clearance time (T1/2) of liver ferritin (H/L subunit ratio=0.43) was 6.8 to 11.8 min, and that of heart ferritin (H/L=3.69) was 9.3 to 25.0 min. T1/2 of biotinylated canine liver ferritin was independent of iron content, whereas canine heart apoferritin (T1/2=31.2 and 32.7 min) was more slowly removed from circulation than the holoferritin. On the other hand, biotinylated recombinant bovine H-chain ferritin homopolymer show a much slower rate of removal (T1/2=153.8 and 155.0 min) compared with the L-chain ferritin homopolymer (T1/2=26.4 and 31.3 min). The rapid clearance of canine tissue ferritin suggests that serum ferritin is an iron transporter in canines.  相似文献   

15.
Early embryonic lethality of H ferritin gene deletion in mice   总被引:17,自引:0,他引:17  
Ferritin molecules play an important role in the control of intracellular iron distribution and in the constitution of long term iron stores. In vitro studies on recombinant ferritin subunits have shown that the ferroxidase activity associated with the H subunit is necessary for iron uptake by the ferritin molecule, whereas the L subunit facilitates iron core formation inside the protein shell. However, plant and bacterial ferritins have only a single type of subunit which probably fulfills both functions. To assess the biological significance of the ferroxidase activity associated with the H subunit, we disrupted the H ferritin gene (Fth) in mice by homologous recombination. Fth(+/-) mice are healthy, fertile, and do not differ significantly from their control littermates. However, Fth(-/-) embryos die between 3.5 and 9.5 days of development, suggesting that there is no functional redundancy between the two ferritin subunits and that, in the absence of H subunits, L ferritin homopolymers are not able to maintain iron in a bioavailable and nontoxic form. The pattern of expression of the wild type Fth gene in 9.5-day embryos is suggestive of an important function of the H ferritin gene in the heart.  相似文献   

16.
Ferritin-binding proteins circulating in mammalian blood are thought to be involved in the clearance of ferritin. The present study characterizes canine serum autoantibodies (IgM and IgA) that react with ferritin. Canine IgM and IgA bound to bovine spleen ferritin as well as canine liver ferritin. To examine the specificity of canine IgM and IgA to ferritin H and L subunits, we used canine heart ferritin and canine liver ferritin with H/L subunit ratios of 3.69 and 0.43, respectively. Canine IgM and IgA recognized both of the H- and L-subunit-rich isoferritins, showing that their binding activities to ferritin depend on the H-subunit content. Recombinant bovine H-chain ferritin homopolymer expressed in a baculovirus expression system bound more with IgM and IgA than the recombinant L-chain homopolymer expressed under the same conditions. These results suggest that canine IgM and IgA recognize H-subunit-rich isoferritins, and that H-subunit-rich isoferritins are cleared from the circulation more rapidly than L-subunit-rich isoferritins.  相似文献   

17.
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   

18.
Ferritin plays an important role in iron metabolism and our aim is to understand the mechanisms by which iron is sequestered within its protein shell as the mineral ferrihydrite. We present M?ssbauer spectroscopic data on recombinant human and horse spleen ferritin from which we draw the following conclusions: (1) that apoferritin catalyses Fe(II) oxidation as a first step in ferrihydrite deposition, (2) that the catalysis of Fe(II) oxidation is associated with residues situated within H chains, at the postulated 'ferroxidase centre' and not in the 3-fold inter-subunit channels previously suggested as the initial Fe(II) binding and oxidation site; (3) that both isolated Fe(III) and Fe(III) mu-oxo-bridged dimers found previously by M?ssbauer spectroscopy to be intermediates in iron-core formation in horse spleen ferritin, are located on H chains; and (4) that these dimers form at ferroxidase centres. The importance of the ferroxidase centre is suggested by the conservation of its ligands in many ferritins from vertebrates, invertebrates and plants. Nevertheless iron-core formation does occur in those ferritins that lack ferroxidase centres even though the initial Fe(II) oxidation is relatively slow. We compare the early stages of core formation in such variants and in horse spleen ferritin in which only 10-15% of its chains are of the H type. We discuss our findings in relation to the physiological role of isoferritins in iron storage processes.  相似文献   

19.
Previous kinetics studies with homopolymer ferritins (bullfrog M-chain, human H-chain and Escherichia coli bacterial ferritins) have established that a mu-1,2-peroxo diferric intermediate is formed during Fe(II) oxidation by O2 at the ferroxidase site of the protein. The present study was undertaken to determine whether such an intermediate is formed also during iron oxidation in horse spleen ferritin (HoSF), a naturally occurring heteropolymer ferritin of H and L-subunits (approximately 3.3 H-chains/HoSF), and to assess its role in the formation of the mineral core. Multi-wavelength stopped-flow spectrophotometry of the oxidative deposition of iron in HoSF demonstrated that a transient peroxo complex (lambda(max) approximately 650 nm) is produced in this protein as for other ferritins. The peroxo complex in HoSF is formed about fourfold slower than in human H-chain (HuHF) and decays more slowly (approximately threefold) as well, at an iron level of two Fe(II)/H-chain. However, as found for HuHF, a second intermediate is formed in HoSF as a decay product of the peroxo complex. Only one-third of the expected peroxo complex forms at the ferroxidase centers of HoSF when two Fe(II)/H-subunits are added to the protein, dropping to only approximately 14% when 20 Fe(II)/H-chain are added, indicating a declining role of the peroxo complex in iron deposition. In contrast to HuHF, HoSF does not enzymatically regenerate the observable peroxo complex. The kinetics of mineralization in HoSF are modeled satisfactorily by a mechanism in which the ferroxidase site rapidly produces an incipient core from a single turnover of iron, upon which subsequent Fe(II) is oxidized autocatalytically to build the Fe(O)OH(s) mineral core. This model supports a role for the L-chain in iron mineralization and helps to explain the widespread occurrence of heteropolymer ferritins in tissues of vertebrates.  相似文献   

20.
Ceruloplasmin, a copper ferroxidase, promotes the incorporation of Fe(III) into the iron storage protein, apoferritin. The product formed is identical to ferritin as judged by polyacrylamide electrophoresis and iron/protein measurements. Of several proteins examined, only apoferritin accumulates the Fe(III) produced by ceruloplasmin. When ceruloplasmin was replaced by tyrosinase, which we have shown to have ferroxidase activity, no iron incorporation into apoferritin was observed. It is proposed that Fe(III) is transferred directly and specifically to apoferritin. These data support a more specific role for ceruloplasmin in iron metabolism than has previously been proposed.  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