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 250 毫秒
1.
Assembly of intra- and interspecies hybrid apoferritins   总被引:1,自引:0,他引:1  
An intraspecies hybrid apoferritin was assembled by mixing subunits of horse heart ferritin, which consists mainly of H-type subunits, and horse spleen ferritin, in which L-type subunits predominate. Interspecies hybrid apoferritins were reconstituted from subunits of human liver-horse spleen ferritins and from rat liver-horse spleen ferritins. All the hybrid ferritins migrated as single zones with electrophoretic mobilities intermediate between those of the parent ferritins. Isoelectric focusing data and immunological patterns were consistent with the view that the reassembled apoferritins were composite molecules that contained subunits from each of the interacting forms. Reconstitution occurred in a random manner, as there was no apparent preference for assembly of homologous subunits. These results suggest that intersubunit interaction domains and recognition mechanisms that dictate formation of the highly specific quaternary structure assumed by this protein are common for different species of ferritins.  相似文献   

2.
Structural similarities between ferritins and bacterioferritins have been extensively demonstrated. However, there is an essential difference between these two types of ferritins: whereas bacterioferritins bind haem, in-vivo, as Fe(II)-protoporphyrin IX (this haem is located in a hydrophobic pocket along the 2-fold symmetry axes and is liganded by two axial Met 52 residues), eukaryotic ferritins are non-haem iron proteins. However, in in-vivo studies, a cofactor has been isolated from horse spleen apoferritin similar to protoporphyrin IX; in in-vitro experiments, it has been shown that horse spleen apoferritin is able to interact with haemin (Fe(III)-protoporphyrin IX). Studies of haemin incorporation into horse spleen apoferritin have been carried out, which show that the metal free porphyrin is found in a pocket similar to that which binds haem in bacterioferritins (Précigoux et al. 1994 Acta Cryst D50, 739–743). A mechanism of demetallation of haemin by L-chain apoferritins was subsequently proposed (Crichton et al. 1997 Biochem 36, 15049–15054) which involved four Glu residues (E 53,56,57,60) situated at the entrance of the hydrophobic pocket and appeared to be favoured by acidic conditions. To verify this mechanism, these four Glu have been mutated to Gln in recombinant horse L-chain apoferritin. We report here the EPR spectra of recombinant horse L-chain apoferritin and its mutant with haemin in basic and acidic conditions. These studies confirm the ability of recombinant L-chain apoferritin and its mutant to incorporate and demetallate the haemin in acidic and basic conditions.  相似文献   

3.
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   

4.
Mouse liver ferritin is composed almost exclusively of polypeptide chains similar in molecular mass (22 kDa) to that characteristic of the major chain (H) found in heart ferritin isolated from human, horse or rat. In these species the predominant polypeptide of liver (L) is smaller (about 20 kDa). Here we show that mouse liver and horse spleen ferritins and apoferritins exhibit extensive structural homology as judged by the similarity in the diffraction patterns of their crystals grown from cadmium sulphate solutions. Implications of this finding are discussed.  相似文献   

5.
When either horse spleen apoferritin (containing more than 90% of L chains) or recombinant horse L apoferritin are modified with glycineamide or taurine in the presence of a water-soluble carbodiimide, a total of 11 to 12 carboxyl groups per subunit are modified, and iron incorporation is effectively abolished. In contrast, when horse spleen ferritin (containing on average 2500 atoms per molecule) is modified under similar conditions, seven to eight carboxyl groups are modified. When apoferritin is prepared from this modified ferritin, it retains full iron incorporation activity. Apoferritin in which seven to eight carboxyls per subunit have been modified by glycineamide can subsequently be modified by taurine; a total of three to four carboxyl groups are modified accompanied by total loss of iron incorporation. Additional studies confirm that three carboxyl groups per subunit are protected from modification by glycineamide by Cr(III) inhibition of iron incorporation. Using tandem mass spectroscopy we have looked for taurine-labelled peptides in tryptic digests of succinylated apoferritins after taurine modification. In the sample where the residues involved in iron uptake have been modified with taurine, we have identified the peptide: This corresponds to residues 53–59 of the L subunit, where it is part of a region of the B-helix which is directed towards the inside of the apoferritin protein shell. The same peptide was identified using classical protein sequencing techniques after (1,2-3H)-taurine modification. We conclude that in L-chain apoferritins the Glu residues at positions 53, 56 and 57 are involved in the mechanism of iron incorporation. Glu 53 and 56 are conserved in L but not in H ferritins, and are located in close proximity to each other within the three-dimensional structure. There is ample room for rotation of Glu 57 to join with the other two to form an iron-binding site. This may represent a site of iron incorporation (most probably involving nucleation) unique to L-chain ferritins, and may explain the predominant L-chain involvement in conditions of iron overload.  相似文献   

6.
Horse ferritins from different organs show heterogeneity on electrofocusing in Ampholine gradients. Both ferritin and apoferritin from liver and spleen could be fractionated with respect to surface charge by serial precipitation with (NH4)2SO4. In the ferritin fractions, increasing iron content parallels increasing isoelectric point. After removal of their iron, those fractions which originally contained most iron accumulated added iron at the fastest rates. When unfractionated ferritins from different organs were compared the average isoelectric point increased in order spleen less than liver less than kidney less than heart. The order of initial rates of iron uptake by the apoferritins was spleen greater than kidney greater than heart and initial average iron contents also followed this order. The relatively low rates of iron accumulation by iron-poor molecules may have been due to structural alteration, to degradation, to activation of the iron-rich molecules or to other factors.  相似文献   

7.
Ferritin was isolated from the seeds of pea (Pisum sativum) and lentil (Lens esculenta). The homogeneity of the phytoferritins was established by polyacrylamide-gel electrophoresis. The subunit molecular weights were respectively 20 300 and 21 400 for hte pea and lentil proteins. A neutron low-angle scattering study established the molecular weight of the oligomer as 480 000 for pea apoferritin and 510 000 for lentil apoferritin. Although the quaternary structure of 24 polypeptide chains is preserved, the phytoferritins have a larger cavity in the interior than mammalian ferritins and can thus potentially store 1.2-1.4 times as much iron. The amino acid composition of the phytoferritins show some similarities to those of mammalian apoferritins; tryptic 'fingerprinting' reveals that there are many differences in the amino acid sequence of plant and mammalian apoferritins.  相似文献   

8.
The X-ray structure of recombinant horse L-chain (rL) apoferritin, solved at 2.0?Å resolution with a final R factor of 17.9%, gives evidence that the residue at position 93 in the sequence is a proline and not a leucine, as found in earlier sequencing studies. The structure is isomorphous with other apoferritin structures, and we thus draw particular attention to those structural features which can be related to the stability and function of the protein. Analysis of hydrogen bonding and salt bridge interactions shows that dimers and tetramers are the most stable molecular entities within the protein shell: a result confirming earlier biophysical experiments. The stability of horse rL apoferritin to both dissociation into subunits at acidic pH values and to complete unfolding in guanidine chloride solutions is compared with that of other apoferritins. This emphasizes the role played by the salt bridge in the stability of this protein family. The horse rL apoferritin is significantly more resistant to denaturation than horse spleen ferritin, which in turn is more resistant than any human rH apoferritins, even those for which a salt bridge is restored. Finally, this structure determination not only establishes that a preformed pocket exists in L-chain apoferritin, at a site known to be able to bind porphyrin, but also underlines the particular function of a cluster of glutamic acids (E53, E56, E57 and E60) located at the entrance of this porphyrin-binding pocket.  相似文献   

9.
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   

10.
Ferritin from clover (Trifolium subterraneum) seeds was isolated and characterized. It was shown to contain polypeptide units of 28–26.5 kDa. The apparent molecular mass of the native protein was 560 kDa. The average iron cores were 4 nm in diameter and contained 1300 iron atoms, with Fe:P = 4:1. Purified clover apoferritin was shown to be functional by means of iron uptake experiments. Plots of initial velocities of iron uptake (at pH 6.7) into clover and horse spleen apoferritins were found to have similar profiles.  相似文献   

11.
We have reinvestigated the association and dissociation of ferritin and apoferritin in phosphate buffer (pH 7.2, I = 0.05). When oligomer-enriched solutions of horse spleen ferritin were mixed with more concentrated, but unenriched solutions of horse spleen apoferritin, there was dissociation of the ferritin oligomers, as determined by polyacrylamide gel electrophoresis and from iron/protein ratios. Some evidence was also obtained for association of monomers in the mixture of ferritin and apoferritin after pelleting and redissolution of pellets in minimal volumes of the phosphate buffer. Monomer-enriched, biosynthetically labeled rat liver ferritin was pelleted, redissolved in minimal volumes of phosphate buffer, and separated by polyacrylamide gel electrophoresis; the fractions were isolated and counted. The results revealed that an association of monomers of the rat liver ferritin had taken place which doubled the concentration of dimers. However, our results also indicate that association by concentration was limited to a fraction of monomers.  相似文献   

12.
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   

13.
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   

14.
An essential difference between eukaryotic ferritins and bacterioferritins is that the latter contain naturally, in vivo haem as Fe-protoporphyrin IX. This haem is located in a hydrophobic pocket along the 2-fold symmetry axes and is liganded by two Met 52. However, in in vivo studies, a cofactor has been isolated in horse spleen apoferritin similar to protoporphyrin IX; in in vitro experiments, it has been shown that horse spleen apoferritin is able to interact with haem. Studies of haemin (Fe(III)-PPIX) incorporation into horse spleen apoferritin have been carried out, which show that the metal free porphyrin is found in a corresponding pocket to haem in bacterioferritins [Précigoux, G., Yariv, J., Gallois, B., Dautant, A., Courseille, C. and Langlois, d'Estaintot B. (1994) A crystallographic study of haem binding to ferritin. Acta Cryst. D 50, 739-743]. A mechanism of demetallation of haemin by L-chain apoferritin was proposed [Crichton, R.R., Soruco, J.A., Roland, F., Michaux, M.A., Gallois, B., Précigoux, G., Mahy, J.P. and Mansuy. (1997) Remarkable ability of horse spleen apoferritin to demetallate hemin and to metallate protoporphyrin IX as a function of pH. J. P. Biochem. 36, 49, 15049-15054]: this involved four Glu residues (53,56,57,60) situated at the entrance of the hydrophobic pocket and appeared to be favoured by acidic conditions. To verify this mechanism, we have mutated these four Glu to Gln and examined demetallation in both acidic and basic conditions. In this paper, we report the mass spectrometry studies of L-chain apoferritin and its mutant incubated with haemin and analysed after different times of incubation: 15 days, 2 months, 6 months, 9 months and 12 months. These studies show that the recombinant L-chain apoferritin and its mutant are able to demetallate haemin to give a hydroxyethyl protoporphyrin IX derivative in a dimeric form [Macieira, S., Martins, B. M. and Huber, R. (2003) Oxygen-dependent coproporphyrinogen IX oxidase from Escherichia coli: one-step purification and biochemical characterization. FEMS. Microbiology Letters 226, 31-37].  相似文献   

15.
Ferritin purified from horse heart and applied to nondenaturing polyacrylamide gel electrophoresis migrated as a single band that stained for both iron and protein. This ferritin contained almost equal amounts of fast- and slow-sedimenting components of 58 S and 3-7 S, which could be separated on sucrose density gradients. Iron removal reduced the sedimentation coefficient of the fast-sedimenting ferritin to 18 S, and sedimentation equilibrium gave a molecular weight 650,000, with some preparations containing ferritin of 500,000 molecular weight as well. Sedimentation rates of the 3 S and 7 S ferritins were not affected by iron removal, and sedimentation equilibrium data were consistent with Mr's 40,000 and 180,000, respectively. Preparations of ferritin extracted from horse spleen contained only 67 S (holo) or 16 S (apo) ferritin and no slow-sedimenting species. When examined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis, all of the ferritins contained the usual H and L subunits (23 and 20 kDa, respectively), but the slow-sedimenting (3 S and 7 S) heart apoferritins also contained appreciable quantities (ca 25%) of three larger subunits of 42, 55, and 65 kDa. All the subunits reacted positively in Western blots to polyclonal antibodies made against specially purified large heart or spleen ferritins containing only 20- and 23-kDa subunits. Similar results were obtained for ferritins from rat heart. The results indicate that mammalian heart tissue is peculiar not just in having an abnormally large iron-rich ferritin but also in having iron-poor ferritins of much lower molecular weight, partly composed of larger subunits.  相似文献   

16.
In situ functional assay of each ferritin molecule in single-layer 2D arrays for horse spleen apoferritin and recombinant horse L- and human H-apoferritins was conducted by observing the iron-cores formed in the arrays by TEM. The study of the time-course, pH-dependence, and temperature-dependence of the function confirmed the iron-core formation to be due to the native function of apoferritins in array. Dark-field TEM imaging revealed that there was crystallinity in the cores in the array of recombinant human H-apoferritin. This iron-core formation was perfectly preserved in the array even after 3 months of storage at room temperature and low humidity. Moreover, about 50% of the function was found to remain in the array after it was exposed to 150 ° C in vacuum for 1 hr.  相似文献   

17.
Ferritin was isolated from the fungus Phycomyces blakesleeanus and compared biochemically and immunologically with horse spleen ferritin. Phycomyces and horse spleen ferritins were shown to exhibit similar electrophoretic patterns on polyacrylamide gels. Both preparations yielded an identical single band on sodium dodecyl sulfate-containing polyacrylamide gels. Tryptic digests of Phycomyces ferritin yielded 17 ninhydrin-positive spots as compared to 26 for horse spleen ferritin tryptic digests. Phycomyces ferritin was immunologically unrelated to horse spleen ferritin.  相似文献   

18.
Sodium dodecyl sulfate-capillary gel electrophoresis (SDS-CGE) was used to characterize the H- and L-subunit ratios of several mammalian ferritins and one bacterioferritin. Traditionally, SDS-PAGE has been used to characterize the H- and L-subunit ratios in ferritin; however, this technique is relatively slow and requires staining, destaining, and scanning before the data can be processed. In addition, the H- and L-subunits of ferritin are fairly close in molecular weight (approximately 21,000 and approximately 20,000, respectively) and are often difficult to resolve in SDS-PAGE slab gels. In contrast, SDS-CGE requires no staining or destaining procedures and the peak quantitation is superior to SDS-PAGE. SDS-CGE is effective in quickly resolving the H- and L-subunits of ferritins from horse spleen, human liver, recombinant human H and L homopolymers, and mixtures of the two- and the single-subunit of a bacterioferritin from Escherichia coli. The technique has also proven useful in assaying the quality of the protein sample from both commercial and recombinant sources. Significant amounts of low-molecular-weight degradation products were detected in all commercial sources of horse spleen ferritin. Most commercial horse spleen ferritins lacked intact H-subunits under denaturing conditions.  相似文献   

19.
Apoferritins prepared from horse spleen and heart and rat heart and liver were dissociated by treatment with acetic acid (pH 1.3-3.0). Sedimentation velocity studies showed that apoferritins of spleen and liver (16-17 S) and heart (18-19 S) dissociated into material sedimenting near 3.2 S. Sedimentation equilibrium measurements determined that most of the material had a molecular weight of 38,000-43,000, corresponding to subunit dimers. Failure to dissociate into subunit monomers was confirmed by gel chromatography on Sephadex G-75 and G-150. With the exception of boiling in sodium dodecyl sulfate, further treatments with 0.1-0.4 M KCl, NaCl, 4-9 M urea, 0.01-0.5 M KSCN, 0.1-0.5% Triton X-100, 5-52% dimethylsulfoxide, 10% ethylene glycol, or 0.1% trifluoroacetic acid all failed to cause dissociation into individual subunits, as did exposure to 6 M guanidine-HCl or formic acid, or prior succinylation and/or nitration of the protein. Reassociation occurred between pH 4 and 7 but was not aided by the addition of Fe(II) or reducing agents. It is concluded that ferritins readily dissociate to subunit dimer units and that further dissociation does not occur without full denaturation of the protein.  相似文献   

20.
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   

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