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.
Ferritin catalyzes the oxidation of Fe2+ by O2 to form a reconstituted Fe3+ oxy-hydroxide mineral core, but extensive studies have shown that the Fe2+ to O2 stoichiometry changes with experimental conditions. At Fe2+ to horse spleen ferritin (HoSF) ratios greater than 200, an upper limit of Fe2+ to O2 of 4 is typically measured, indicating O2 is reduced to 2H2O. In contrast, a lower limit of Fe2+ to O2 of approximately 2 is measured at low Fe2+ to HoSF ratios, implicating H2O2 as a product of Fe2+ deposition. Stoichiometric amounts of H2O2 have not been measured, and H2O2 is proposed to react with an unknown system component. Evidence is presented that identifies this component as amine buffers, including 3-N-morpholinopropanesulfonic acid (MOPS), which is widely used in ferritin studies. In the presence of non-amine buffers, the Fe2+ to O2 stoichiometry was approximately 4.0, but at high concentrations of amine buffers (0.10 M) the Fe2+ to O2 stoichiometry is approximately 2.5 for iron loadings of eight to 30 Fe2+ per HoSF. Decreasing the concentration of amine buffer to zero resulted in an Fe2+ to O2 stoichiometry of approximately 4. Direct evidence for amine buffer modification during Fe2+ deposition was obtained by comparing authentic and modified buffers using mass spectrometry, NMR, and thin layer chromatography. Tris(hydroxymethyl)aminomethane, MOPS, and N-methylmorpholine (a MOPS analog) were all rapidly chemically modified during Fe2+ deposition to form N-oxides. Under identical conditions no modification was detected when amine buffer, H2O2, and O2 were combined with Fe2+ or ferritin separately. Thus, a short-lived ferritin intermediate is required for buffer modification by H2O2. Variation of the Fe2+ to O2 stoichiometry versus the Fe2+ to HoSF ratio and the amine buffer concentration are consistent with buffer modification.  相似文献   

2.
An optical flow cell provided a means to conveniently measure the rate of successive Fe(2+) oxidation reactions catalyzed by horse spleen ferritin (HoSF) to determine if both ferroxidase and mineral core Fe(2+) oxidation reactions occur. The oxygen concentration and pH were held constant and multiple additions of Fe(2+)/HoSF ratios of 1, 10, 100, 150, 250 and 400 were conducted, creating core sizes ranging from 12 to 2800. During these oxidations, the absence of nonspecific Fe(OH)(3) formation and the presence (>95%) of Fe(OH)(3) deposited within the core of HoSF demonstrated the validity of monitoring iron deposition into HoSF by this procedure. Initial rates for oxidation of 5-50 Fe(2+)/HoSF established that the reaction is overall first order in Fe(2+) concentration. However, when full progress curves were analyzed at a variety of Fe(2+)/HoSF ratios, two first-order reactions (k(1) approximately 0.035 s(-1) and k(2) approximately 0.007 s(-1)) were found to contribute to the overall Fe(2+) oxidation reaction. The proportion of the fast reaction increased with increasing Fe(2+)/HoSF ratio until at approximately 400, it was the dominant reaction. For the Fe(2+)/HoSF ratios examined, the overall rate of iron deposition is independent of the size of the mineral core, a result suggesting that an increasing mineral core size does not enhance the rate of Fe(2+) oxidation. Comparison of successive additions of 1.0 Fe(2+)/HoSF showed that oxidation of the first 8-10 Fe(2+) produced a Fe(III) species with a lower molar absorptivity per Fe(III) than that of the bulk core. Measurement of the H(+)/Fe(2+) ratio confirmed this difference in behavior by giving an H(+)/Fe(2+) ratio of approximately 1.0 below and 2.0 for ratios >30 Fe(2+)/HoSF. The faster reaction was attributed to ferroxidase catalysis and the slow reaction to nonspecific ferroxidase activity of the HoSF protein shell.  相似文献   

3.
Fe2+ binding to both apo- and holo- bacterial ferritin from Azotobacter vinelandii (AVBF) was measured as a function of pH under carefully controlled anaerobic conditions. Fe2+ binding to apo-AVBF is strongly pH dependent with 25 Fe2+ ions/apo-AVBF binding tightly at pH 5.5 and over 150 Fe2+/apo-AVBF at pH 9.0. Holo-AVBF gave a similar pH-dependent binding profile with over 400 Fe2+/AVBF binding at pH of 9.0. Proton release per Fe2+ bound to either AVBF protein increases with increasing pH until a total of about two protons are released at pH 9.0. These binding results are both qualitatively and quantitatively different from corresponding measurements (Jacobs et al., 1989) on apo- and holo- mammalian ferritin (MF) where less Fe2+ binds in both cases. The high level of Fe2+ binding to holo-AVBF relative to that of mammalian ferritin is a consequence of the higher phosphate content in the core of AVBF. Reduction of AVBF by either dithionite or methyl viologen in the absence of chelating agents demonstrated that phosphate, but not Fe2+, is released from the AVBF core in amounts commensurate with the degree of iron reduction, although even at 100% reduction considerable phosphate remains associated with the reduced mineral core. Fe2+ binding to holo-AVBF made deficient in phosphate was lower than that of native AVBF, while the addition of phosphate to native holo-AVBF increased the Fe2+ binding capacity. These results clearly support the role of phosphate as the site of interaction of Fe2+ with the AVBF mineral core.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

4.
Role of phosphate in initial iron deposition in apoferritin   总被引:1,自引:0,他引:1  
Y G Cheng  N D Chasteen 《Biochemistry》1991,30(11):2947-2953
Ferritins from microorganisms to man are known to contain varying amounts of phosphate which has a pronounced effect on the structural and magnetic properties of their iron mineral cores. The present study was undertaken to gain insight into the role of phosphate in the early stages of iron accumulation by ferritin. The influence of phosphate on the initial deposition of iron in apoferritin (12 Fe/protein) was investigated by EPR, 57Fe M?ssbauer spectroscopy, and equilibrium dialysis. The results indicate that phosphate has a significant influence on iron deposition. The presence of 1 mM phosphate during reconstitution of ferritin from apoferritin, Fe(II), and O2 accelerates the rate of oxidation of the iron 2-fold at pH 7.5. In the presence or absence of phosphate, the rate of oxidation at 0 degrees C follows simple first-order kinetics with respect to Fe(II) with half-lives of 1.5 +/- 0.3 or 2.8 +/- 0.2 min, respectively, consistent with a single pathway for iron oxidation when low levels of iron are added to the apoprotein. This pathway may involve a protein ferroxidase site where phosphate may bind iron(II), shifting its redox potential to a more negative value and thus facilitating its oxidation. Following oxidation, an intermediate mononuclear Fe(III)-protein complex is formed which exhibits a transient EPR signal at g' = 4.3. Phosphate accelerates the rate of decay of the signal by a factor of 3-4, producing EPR-silent oligonuclear or polynuclear Fe(III) clusters. In 0.5 mM Pi, the signal decays according to a single phase first-order process with a half-life near 1 min.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

5.
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   

6.
The binding of Fe2+ to both apo and holo mammalian ferritin has been investigated under anaerobic conditions as a function of pH. In the pH range 6.0-7.5, 8.0 +/- 0.5 Fe2+ ions bind to each apoferritin molecule, but above pH 7.5, a pH-dependent Fe2+ binding profile is observed with up to 80 Fe2+ ions binding at pH 10.0. This Fe2+ binding is reversible and is accompanied by up to two H+ being released per Fe2+ bound at pH 10.0. The Fe2+ binding to apoferritin probably occurs in the 3-fold channels. A much larger and more complex pH-dependent Fe2+ binding stoichiometry was observed for holoferritin with up to 300 Fe2+ ions binding at pH 10.0. This pH-dependent Fe2+ binding was interpreted as Fe2+ interaction at the FeOOH mineral surface with displacement of H+ from -OH or phosphate surface groups by the incoming Fe2+ ions. Mossbauer spectroscopic measurements using 57Fe-labeled Fe2+ under anaerobic conditions showed that 57Fe2+ binding to holoferritin was accompanied by electron transfer to the core, yielding 57Fe3+, presumably bound to the mineral surface. Removal of added iron by Fe2+-specific chelating agents yielded 57Fe2+, demonstrating the reversibility of this electron-transfer process. The Fe2+ bound to apo- and holoferritin is readily converted to Fe3+ by exposure to O2 and strongly retained by the respective ferritin species.  相似文献   

7.
Ferritin cores isolated from human spleen, limpet (Patella vulgata) hemolymph and bacterial (Pseudomonas aeruginosa) cells have been investigated by high resolution transmission electron microscopy, electron diffraction and chemical analysis. Hemosiderin particles isolated from thalassemic spleens also have been studied. The results show that there is a marked difference in structure and composition of the biomineral phases. Human ferritin and hemosiderin particles are single domain crystals of hydrated iron (III) oxide (ferrihydrite). Lattice fringes were low in contrast and often discontinuous within the central regions of the core. Heat treatment of human ferritins results in a 5 A shrinkage in particle size and an increase in the single crystalline nature of the core. In contrast, lattice images and electron diffraction of limpet and bacterial cores show no evidence of long-range crystallographic order. Chemical analysis indicates a high inorganic phosphate (Pi) (Fe/Pi = 1.71) content in bacterial ferritin compared with human ferritin (thalassemic) (Fe/Pi = 21.0). The high Pi content of bacterial ferritin suggests a hydrated amorphous iron (III) phosphate mineral core. Structural disorder within the limpet and bacterial cores may be associated with increased Pi content and increased oxidation in Fe(II), resulting in rapid mineral deposition. Growth of the iron (III) oxide cores in human ferritin is discussed on the basis of high resolution electron microscopy results.  相似文献   

8.
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   

9.
M J Yablonski  E C Theil 《Biochemistry》1992,31(40):9680-9684
Ferritin is a large protein, highly conserved among higher eukaryotes, which reversibly stores iron as a mineral of hydrated ferric oxide. Twenty-four polypeptides assemble to form a hollow coat with the mineral inside. Multiple steps occur in iron core formation. First, Fe2+ enters the protein. Then, several alternate paths may be followed which include oxidation at site(s) on the protein, oxidation on the core surface, and mineralization. Sequence variations occur among ferritin subunits which are classified as H or L; Fe2+ oxidation at sites on the protein appears to be H-subunit-specific or protein-specific. Other steps of ferritin core formation are likely to involve conserved sites in ferritins. Since incorporation of Fe2+ into the protein must precede any of the other steps in core formation, it may involve sites conserved among the various ferritin proteins. In this study, accessibility of Fe2+ to 1,10-phenanthroline, previously shown to be inaccessible to Fe2+ inside ferritin, was used to measure Fe2+ incorporation in two different ferritins under various conditions. Horse spleen ferritin (L/H = 10-20:1) and sheep spleen ferritin (L/H = 1:1.6) were compared. The results showed that iron incorporation measured as inaccessibility of Fe2+ to 1,10-phenanthroline increased with pH. The effect was the same for both proteins, indicating that a step in iron core formation common among ferritins was being measured. Conserved sites previously proposed for different steps in ferritin core formation are at the interfaces of pairs and trios of subunits. Dinitrophenol cross-links, which modify pairs of subunits and affect iron oxidation, had no effect on Fe2+ incorporation.(ABSTRACT TRUNCATED AT 250 WORDS)  相似文献   

10.
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   

11.
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   

12.
Zhang B  Harb JN  Davis RC  Choi S  Kim JW  Miller T  Chu SH  Watt GD 《Biochemistry》2006,45(18):5766-5774
Azotobacter vinelandii bacterioferritin (AvBF) containing 800-1500 Co or Mn atoms as Co(III) and Mn(III) oxyhydroxide cores (Co-AvBF, Mn-AvBF) was synthesized by the same procedure used previously for horse spleen ferritin (HoSF). The kinetics of reduction of Co-AvBF and Mn-AvBF by ascorbic acid are first-order in each reactant. The rate constant for the reduction of Mn-AvBF (8.52 M(-1) min(-1)) is approximately 12 times larger than that for Co-AvBF (0.72 M(-1) min(-1)), which is consistent with a previous observation that Mn-HoSF is reduced approximately 10-fold faster than Co-HoSF [Zhang, B. et al. (2005) Inorg. Chem. 44, 3738-3745]. The rates of reduction of M-AvBF (M = Co and Mn) are more than twice that for the reduction of the corresponding M-HoSF. HoSF containing reduced Fe(II) cores (Fe(II)-HoSF), prepared by methyl viologen and CO, also reduces M-HoSF and M-AvBF species, with both cores remaining within ferritin, suggesting that electrons transfer through the ferritin shell. Electron transfer from Fe(II)-HoSF to Co-AvBF occurs at a rate approximately 3 times faster than that to Co-HoSF, indicating that the Co cores in AvBF are more accessible to reduction than the Co cores in HoSF. The presence of nonconductive (SiO2) or conductive (gold) surfaces known to bind ferritins enhances the rate of electron transfer. A more than approximately 4-fold increase in the apparent reaction rate is observed in the presence of gold. Although both surfaces (SiO2 and gold) enhance reaction by providing binding sites for molecular interaction, results show that ferritins with different mineral cores bound to a gold surface transfer electrons through the gold substrate so that direct contact of the reacting molecules is not required.  相似文献   

13.
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   

14.
Stabilization of iron in a bioavailable form is the function of ferritin, a protein of 24 subunits forming a coat around a core of less than or equal to 4500 hydrated iron atoms. The core of ferritin isolated from tissues contains Fe3+, but Fe2+ is required for experimental core formation in protein coats; reduction of Fe3+ to Fe2+ facilitates iron removal from protein coats. Using the differences in x-ray absorption spectra (x-ray absorption near edge structure) between Fe2+ and Fe3+ to monitor reconstitution of ferritin from Fe2+ and protein coats, we observed stabilization of Fe2+, apparently inside the coat. Mixtures of Fe2+ and Fe3+ persisted for greater than or equal to 16 h in air indicating that, in vivo, some iron in ferritin could be stored as Fe2+ and with Fe3+ could yield magnetite.  相似文献   

15.
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   

16.
Reconstituted and native iron-cores of bacterioferritin and ferritin   总被引:9,自引:0,他引:9  
The structural and magnetic properties of the iron-cores of reconstituted horse spleen ferritin and Azotobacter vinelandii bacterioferritin have been investigated by high-resolution transmission electron microscopy, electron diffraction and Mossbauer spectroscopy. The structural properties of native horse spleen ferritin, native Az. vinelandii, and native and reconstituted Pseudomonas aeruginosa bacterioferritins have also been determined. Reconstitution in the absence of inorganic phosphate at pH 7.0 showed sigmoidal behaviour in each protein but was approximately 30% faster in initial rate for the Az. vinelandii protein when compared with horse spleen apoferritin. The presence of Zn2+ reduced the initial rate of Fe(II) oxidation in Az. vinelandii to 22% of the control rate. The iron-cores of the reconstituted bacterioferritins adopt defect ferrihydrite structures and are more highly ordered than their native counterparts, which are both amorphous. However, the blocking temperature for reconstituted Az. vinelandii (22.2 K) is almost identical to that for the native protein (20 K). Particle size measurements indicate that the reconstituted Az. vinelandii cores are smaller in median diameter than the native cores and this reduction in particle volume (V) offsets the increased magnetocrystalline contribution to the magnetic anisotropy constant (K) in such a way that the magnetic anisotropy barrier (KV), and hence the blocking temperature, is similar for both proteins. Reconstituted horse spleen ferritin exhibits a similar blocking temperature (38 K) to that determined for the native protein, although it is structurally more disordered. The possibility of introducing structural and compositional modifications in both horse ferritin and bacterioferritins by in-vitro reconstitution suggests that these proteins do not function primarily as a crystallochemical-specific interface for core development in vivo.  相似文献   

17.
Ferritin is a complex of a hollow, spherical protein and a hydrous, ferric oxide core of less than or equal to 4500 iron atoms inside the apoprotein coat; the apoprotein has multiple (ca. 12) binding sites for monoatomic metal ions, e.g., Fe(II), V(IV), Tb(III), that may be important in the initiation of iron core formation. In an earlier study we observed that the oxidation of Fe(II) vacated some, but not all, of the metal-binding sites, suggesting migration of some Fe during oxidation, possibly to form nucleation clusters; some Fe(III) remained bound to the protein. Preliminary extended X-ray absorbance fine structure (EXAFS) analysis of the same Fe(III)-apoferritin complex showed an environment distinct from ferritin cores, but the data did not allow a test of the Fe cluster hypothesis. In this paper, with improved EXAFS data and with M?ssbauer data on the same complex formed with 57Fe, we clearly show that the Fe(III) in the distinctive environment is polynuclear (Fe atoms with Fe-Fe = 3.5 A and TB = 7 K). Moreover, the arrangement of atoms is such that Fe(III) atoms appear to have both carboxylate-like ligands, presumably from apoferritin, and oxo bridges to the other iron atoms. Thus the protein provides sites not only for initiation but also for nucleation of the iron core. Sites commodious enough and with sufficient conserved carboxylate ligands to accommodate such a nucleus occur inside the protein coat at the subunit dimer interfaces. Such Fe(III)-apoferritin nucleation complexes can be used to study the properties of the several members of the apoferritin family.  相似文献   

18.
A comparison of an undecairon(III) complex with the ferritin iron core   总被引:1,自引:0,他引:1  
The iron core of ferritin is comprised of up to 4,500 Fe(III) atoms as Fe2O3.nH2O, which is maintained in solution by a surrounding, spherical coat of protein. Organisms as diverse as bacteria and man use the ferritin iron-protein complex as a reservoir of stored iron for other essential proteins. To extend studies of the steps in polynuclear iron core formation, a recently characterized undecairon(III) oxo-hydroxo aggregate [Fe11 complex] (Gorun et al., J. Am. Chem. Soc. 109, 3337 [1987]) was examined by x-ray absorption spectroscopy as a model for an intermediate. The results, which are comparable to the previous x-ray diffraction studies, show near neighbors (Fe-O) at 1.90 A that are distinct from those in ferritin and a longer distance of 2.02 A. However, contributions from neighbors (Fe-C) known to exist at ca. 2.7 A were obscured by a highly ordered Fe-Fe interaction and were not detectable in the Fe11 complex in contrast to a previously characterized Fe(III) cluster bound to the protein coat. Of the two Fe-Fe interactions detectable in the Fe11 complex, the shortest, at 3.0 A is particularly interesting, occurring at the same distance as a full shell (CN = 6) in ferritin, but having fewer Fe neighbors (CN = 2-3) characteristic of an intermediate in core formation. The incomplete Fe-Fe shell is much more ordered than in ferritin, suggesting that the disorder in ferritin cores may be associated with the later steps of the core growth. Differences between the Fe11 complex and the full core of ferritin indicate the possibility of intermediates in ferritin iron formation that might be like Fe11.  相似文献   

19.
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   

20.
Ferritin stores iron within a hollow protein shell as a polynuclear Fe(III) hydrous oxide core. Although iron uptake into ferritin has been studied previously, the early stages in the creation of the core need to be clarified. These are dealt with in this paper by using M?ssbauer spectroscopy, a technique that enables several types of Fe(II) and Fe(III) to be distinguished. Systematic M?ssbauer studies were performed on samples prepared by adding 57Fe(II) atoms to apoferritin as a function of pH (5.6-7.0), n [the number of Fe/molecule (4-480)], and tf (the time the samples were held at room temperature before freezing). The measurements made at 4.1 and 90 K showed that for samples with n less than or equal to 40 at pH greater than or equal to 6.25 all iron was trivalent at tf = 3 min. Four different Fe(III) species were identified: solitary Fe(III) atoms giving relaxation spectra, which can be identified with the species observed before by EPR and UV difference spectroscopy; oxo-bridged dimers giving doublet spectra with large splitting, observed for the first time in ferritin; small Fe(III) clusters giving doublets of smaller splitting and larger antiferromagnetically coupled Fe(III) clusters, similar to those found previously in larger ferritin iron cores, which, for samples with n greater than or equal to 40, gave magnetically split spectra at 4.1 K. Both solitary Fe(III) and dimers diminished with time, suggesting that they are intermediates in the formation of the iron core. Two kinds of divalent iron were distinguished for n = 480, which may correspond to bound and free Fe(II).  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