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 343 毫秒
1.
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   

2.
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   

3.
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   

4.
Protein ferroxidase site(s), which catalyze the reaction between ferrous ion and dioxygen, have long been thought to play a role in core formation in ferritin; however, the mechanism of the reaction has never been studied in detail. In the present work, the enzymatic activity of ferritin was examined using oximetry, the net Fe2+ oxidation reaction being as follows. [formula: see text] The reaction exhibits saturation kinetics with respect to both Fe2+ and O2 (apparent Michaelis constants: Km,Fe = 0.35 +/- 0.01 mM and Km,O2 = 0.14 +/- 0.03 mM). The enzyme has a turnover number kcat = 80 +/- 3 min-1 at 20 degrees C with maximal activity at pH 7. The kinetics are discussed in terms of two mechanisms, one involving monomeric and the other dimeric iron protein complexes. In both instances Fe(II) oxidation occurs in 1-electron steps. Zinc(II) is a competitive inhibitor of iron(II) oxidation at Zn2+/apoprotein ratios > or = 6 (inhibitor constant KI,Zn = 0.067 +/- 0.011 mM) but appears to be a noncompetitive inhibitor at lower ratios (< or = 2), indicating the presence of more than one type of zinc binding site on the protein. At increments of 50 Fe2+/protein or less, all of the iron is oxidized via the protein ferroxidase site(s), independent of the amount of core already present. However, when larger increments are employed, some iron oxidation appears to occur on the surface of the mineral core. The results of these studies emphasize the role of the protein shell in all phases of core growth and confirm the presence of a functionally important catalytic site in ferritin in addition to other binding sites on the protein for iron.  相似文献   

5.
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   

6.
The early redox events involved in iron reduction and mobilization in mammalian ferritin have been investigated by several techniques. Sedimentation velocity measurements of ferritin samples with altered core sizes, prepared by partial reduction and Fe2+ chelation, suggest two different events occur during iron loss from the ferritin core. Reductive optical titrations confirm this biphasic behavior by showing that the first 20-30% of core reduction has different optical properties than the latter 70-80%. Proton uptake during initial core reduction is near zero, but as the percent core reduction increases, the proton uptake (H+/e) values increase to 2 H+/e (2 H+/Fe3+ reduced) as core reduction approaches 1 e/Fe3+. Coulometric reduction of ferritin by mediators of different redox potential and different cross-sectional areas show a two-phase sigmoidal reaction pattern in which initial core reduction occurs at a slower rate than later core reduction. The above experiments were all conducted in the absence of iron chelators so that the observed results were all attributed to core reduction rather than the combined effects of core reduction accompanied by Fe2+ chelation. The coulometric reduction of ferritin by various mediators shows a correlation more with reduction potential than with molecular cross-sectional area. The role of the ferritin channels in core reduction is considered in terms of the reported results.  相似文献   

7.
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   

8.
R K Watt  R B Frankel  G D Watt 《Biochemistry》1992,31(40):9673-9679
Apo horse spleen ferritin undergoes a 6.3 +/- 0.5 electron redox reaction at -310 mV at pH 6.0-8.5 and 25 degrees C to form reduced apoferritin (apoMFred). Reconstituted ferritin containing up to 50 ferric ions undergoes reduction at the same potential, taking up one electron per ferric ion and six additional electrons by the protein. We propose that apo mammalian ferritin (apoMF) contains six redox centers that can be fully oxidized forming oxidized apoferritin (apoMFox) or fully reduced forming apoMFred. ApoMFred can be prepared conveniently by dithionite or methyl viologen reduction. ApoMFred is slowly oxidized by molecular oxygen but more rapidly by Fe(CN)6(3-) to apoMFox. Fe(III)-cytochrome c readily oxidizes apoMFred to apoMFox with a stoichiometry of 6 Fe(III)-cytochrome c per apoMFred, demonstrating a rapid interprotein electron-transfer reaction. Both redox states of apoMF react with added Fe3+ and Fe2+. Addition of eight Fe2+ to apoMFox under anaerobic conditions produced apoMFred and Fe3+, as evidenced by the presence of a strong g = 4.3 EPR signal. Subsequent addition of bipyridyl produced at least six Fe(bipyd)3(2+) per MF, establishing the reversibility of this internal electron-transfer process between the redox centers of apoMF and bound iron. Incubation of apoMFred with the Fe(3+)-ATP complex under anaerobic conditions resulted in the formation and binding of two Fe2+ and four Fe3+ by the protein. The various redox states formed by the binding of Fe2+ and Fe3+ to apoMFox and apoMFred are proposed and discussed. The yellow color of apoMF appears to be an integral characteristic of the apoMF and is possibly associated with its redox activity.  相似文献   

9.
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   

10.
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   

11.
Iron deposition within the iron storage protein ferritin involves a complex series of events consisting of Fe(2+) binding, transport, and oxidation at ferroxidase sites and mineralization of a hydrous ferric oxide core, the storage form of iron. In the present study, we have examined the thermodynamic properties of Fe(2+) binding to recombinant human H-chain apoferritin (HuHF) by isothermal titration calorimetry (ITC) in order to determine the location of the primary ferrous ion binding sites on the protein and the principal pathways by which the Fe(2+) travels to the dinuclear ferroxidase center prior to its oxidation to Fe(3+). Calorimetric titrations show that the ferroxidase center is the principal locus for Fe(2+) binding with weaker binding sites elsewhere on the protein and that one site of the ferroxidase center, likely the His65 containing A-site, preferentially binds Fe(2+). That only one site of the ferroxidase center is occupied by Fe(2+) implies that Fe(2+) oxidation to form diFe(III) species might occur in a stepwise fashion. In dilute anaerobic protein solution (3-5 microM), only 12 Fe(2+)/protein bind at pH 6.51 increasing to 24 Fe(2+)/protein at pH 7.04 and 7.5. Mutation of ferroxidase center residues (E62K+H65G) eliminates the binding of Fe(2+) to the center, a result confirming the importance of one or both Glu62 and His65 residues in Fe(2+) binding. The total Fe(2+) binding capacity of the protein is reduced in the 3-fold hydrophilic channel variant S14 (D131I+E134F), indicating that the primary avenue by which Fe(2+) gains access to the interior of ferritin is through these eight channels. The binding stoichiometry of the channel variant is one-third that of the recombinant wild-type H-chain ferritin whereas the enthalpy and association constant for Fe(2+) binding are similar for the two with an average values (DeltaH degrees = 7.82 kJ/mol, binding constant K = 1.48 x 10(5) M(-)(1) at pH 7.04). Since channel mutations do not completely prevent Fe(2+) binding to the ferroxidase center, iron gains access to the center in approximately one-third of the channel variant molecules by other pathways.  相似文献   

12.
Formation of the ferritin iron mineral occurs in plastids.   总被引:1,自引:0,他引:1       下载免费PDF全文
Ferritin in plants is a nuclear-encoded, multisubunit protein found in plastids; an N-terminal transit peptide targets the protein to the plastid, but the site for formation of the ferritin Fe mineral is unknown. In biology, ferritin is required to concentrate Fe to levels needed by cells (approximately 10(-7) M), far above the solubility of the free ion (10(-18) M); the protein directs the reversible phase transition of the hydrated metal ion in solution to hydrated Fe-oxo mineral. Low phosphate characterizes the solid-phase Fe mineral in the center of ferritin of the cytosolic animal ferritin, but high phosphate is the hallmark of Fe mineral in prokaryotic ferritin and plant (pea [Pisum sativum L.] seed) ferritin. Earlier studies using x-ray absorption spectroscopy showed that high concentrations of phosphate present during ferritin mineralization in vivo altered the local structure of Fe in the ferritin mineral so that it mimicked the prokaryotic type, whether the protein was from animals or bacteria. The use of x-ray absorption spectroscopy to analyze the Fe environment in pea-seed ferritin now shows that the natural ferritin mineral in plants has an Fe-P interaction at 3.26A, similar to that of bacterial ferritin; phosphate also prevented formation of the longer Fe-Fe interactions at 3.5A found in animal ferritins or in pea-seed ferritin reconstituted without phosphate. Such results indicate that ferritin mineralization occurs in the plastid, where the phosphate content is higher; a corollary is the existence of a plastid Fe uptake system to allow the concentration of Fe in the ferritin mineral.  相似文献   

13.

Background

Most models for ferritin iron release are based on reduction and chelation of iron. However, newer models showing direct Fe(III) chelation from ferritin have been proposed. Fe(III) chelation reactions are facilitated by gated pores that regulate the opening and closing of the channels.

Scope of review

Results suggest that iron core reduction releases hydroxide and phosphate ions that exit the ferritin interior to compensate for the negative charge of the incoming electrons. Additionally, chloride ions are pumped into ferritin during the reduction process as part of a charge balance reaction. The mechanism of anion import or export is not known but is a natural process because phosphate is a native component of the iron mineral core and non-native anions have been incorporated into ferritin in vitro. Anion transfer across the ferritin protein shell conflicts with spin probe studies showing that anions are not easily incorporated into ferritin. To accommodate both of these observations, ferritin must possess a mechanism that selects specific anions for transport into or out of ferritin. Recently, a gated pore mechanism to open the 3-fold channels was proposed and might explain how anions and chelators can penetrate the protein shell for binding or for direct chelation of iron.

Conclusions and general significance

These proposed mechanisms are used to evaluate three in vivo iron release models based on (1) equilibrium between ferritin iron and cytosolic iron, (2) iron release by degradation of ferritin in the lysosome, and (3) metallo-chaperone mediated iron release from ferritin.  相似文献   

14.
The removal of ferritin-bound iron by the physiologic dithiol DL-dihydrolipoate was studied over the pH range 5.5-9.0. A novel method was devised for the determination of iron removal, making it possible to study the actual release of iron from ferritin, regardless of the oxidation state or complexation form. The overall iron-removal process appears to depend upon a balance between the deprotonation of the dithiol and the protolytic dissolution of the iron core inside the ferritin molecule. The amount of iron removed at equilibrium increases with the pH, at any of the dihydrolipoate/ferritin iron ratios tested. The formation of the binuclear iron-dithiol complex [Fe2(dihydrolipoate)3]-3 is not strictly required for iron mobilization, but it seems to affect the efficiency of the dithiol in iron mobilization by providing a stable complexation form for the released iron outside the ferritin protein shell. Comparison of the release of ferritin-bound iron by free and immobilized dihydrolipoate indicates that mobility of the dithiol is mandatory for the removal process to take place.  相似文献   

15.
To provide a framework for understanding the hyperthermostability of some rubredoxins, a comprehensive analysis of the thermally induced denaturation of rubredoxin (Rd) from the mesophile, Clostridium pasteurianum was undertaken. Rds with three different metals in its M(SCys)4 site (M = Fe3+/2+, Zn2+, or Cd2+) were examined. Kinetics of metal ion release were monitored anaerobically at several fixed temperatures between 40 and 100 degrees C, and during progressive heating of the iron-containing protein. Both methods gave a thermal stability of metal binding in the order Fe2+ < Fe3+ < Zn2+ < Cd2+. The temperature at which half of the iron was released from the protein in temperature ramp experiments was 69 degrees C for Fe2+ Rd and 83 degrees C for Fe3+ Rd. Temperature-dependent changes in the protein structure were monitored by differential scanning calorimetry, tryptophan fluorescence, binding of a fluorescent hydrophobic probe, and 1H NMR. Major but reversible structural changes, consisting of swelling of the hydrophobic core and opening of a loop region, were found to occur at temperatures (50-70 degrees C) much lower than those required for loss of the metal ion. For the three divalent metal ions, the results suggest that the onset of the reversible, lower-temperature structural changes is dependent on the size of the MS4 site, whereas the final, irreversible loss of metal ion is dependent on the inherent M-SCys bond strength. In the case of Fe3+ Rd, stoichiometric Fe3+/cysteine-ligand redox chemistry also occurs during metal ion loss. The results indicate that thermally induced unfolding of the native Cp Rd must surmount a significant kinetic barrier caused by stabilizing interactions both within the protein and within the M(SCys)4 site.  相似文献   

16.
Detailed circular dichroism (CD), steady-state and time-resolved tryptophan fluorescence studies on the holo- and apo- forms of high potential iron protein (HiPIP) from Chromatium vinosum and its mutant protein have been carried out to investigate conformational properties of the protein. CD studies showed that the protein does not have any significant secondary structure elements in the holo- or apo- HiPIP, indicating that the metal cluster does not have any effect on formation of secondary structure in the protein. Steady-state fluorescence quenching studies however, suggested that removal of the iron-sulfur ([Fe(4)S(4)](3+)) cluster from the protein leads to an increase in the solvent accessibility of tryptophans, indicating change in the tertiary structure of the protein. CD studies on the holo- and apo- HiPIP also showed that removal of the metal prosthetic group drastically affects the tertiary structure of the protein. Time-resolved fluorescence decay of the wild type protein was fitted to a four-exponentials model and that of the W80N mutant was fitted to a three-exponentials model. The time-resolved fluorescence decay was also analyzed by maximum entropy method (MEM). The results of the MEM analysis agreed with those obtained from discrete exponentials model analysis. Studies on the wild type and mutants helped to assign the fast picosecond lifetime component to the W80 residue, which exhibits fast fluorescence energy transfer to the [Fe(4)S(4)](3+) cluster of the protein. Decay-associated fluorescence spectra of each tryptophan residues were calculated from the time-resolved fluorescence results at different emission wavelengths. The results suggested that W80 is in the hydrophobic core of the protein, but W60 and W76 are partially or completely exposed to the solvent.  相似文献   

17.
The origin of previously observed variations in stoichiometry of iron oxidation during the oxidative deposition of iron in ferritin has been poorly understood. Knowledge of the stoichiometry of Fe(II) oxidation by O2 is essential to establishing the mechanism of iron core formation. In the present work, the amount of Fe(II) oxidized was measured by M?ssbauer spectrometry and the O2 consumed by mass spectrometry. The number of protons produced in the reaction was measured by "pH stat" titration and hydrogen peroxide production by the effect of the enzyme catalase on the measured stoichiometry. For protein samples containing low levels of iron (24 Fe(II)/protein) the stoichiometry was found to be 1.95 +/- 0.18 Fe(II)/O2 with H2O2 being a product, viz. Equation 1. 2Fe2+ + O2 + 4H2O----2FeOOH + H2O2 + 4H+ (1) EPR spin trapping experiments showed no evidence of superoxide radical formation. The stoichiometry markedly increased with additional iron (240-960 Fe/protein), to a value of 4 Fe(II)/O2 as in Equation 2. 4Fe2+ + O2 + 6H2O----4FeOOH + 8H+ (2) As the iron core is progressively laid down, the mechanism of iron oxidation changes from a protein dominated process with H2O2 being the primary product of O2 reduction to a mineral surface dominated process where H2O is the primary product. These results emphasize the importance of the apoferritin shell in facilitating iron oxidation in the early stage of iron deposition prior to significant development of the polynuclear iron core.  相似文献   

18.
The ferritin consists of a protein shell constructed of 24 subunits and an iron core. The liver ferritin of Sphyrna zygaena (SZLF) purified by column chromatography is a protein composed of eight ferritins containing varying iron numbers ranging from 400+/-20 Fe3+/SZLF to 1890+/-20 Fe3+/SZLF within the protein shell. Nature SZLF (SZLFN) consisting of holoSZLF and SZLF with unsaturated iron (SZLFUI) to have been purified with polyacrylamide gel electrophoresis (PAGE) exhibited five ferritin bands with different pI values ranging from 4.0 to 7.0 in the gel slab of isoelectric focusing (IEF). HoloSZLF purified by PAGE (SZLFE) not only had 1890+/-20 Fe3+/SZLFE but also showed an identical size of iron core observed by transmission electron microscopy (TEM). Molecular weight of approximately 21 kDa for SZLFE subunit was determined by sodium dodecyl sulfate-polyacrylamide gel electrophoresis (SDS-PAGE). Four peaks of molecular ions at mass/charge (m/z) ratios of 10611.07, 21066.52, 41993.16, and 63555.64 that come from the SZLFE were determined by matrix-assisted laser desorption ionization/time-of-flight mass spectrometry (MALDI-TOF MS), which were identified as molecular ions of the ferritin subunit (M+) and its polymers, namely, [M]2+, [M]+, [2M]+, and [3M]+, respectively. Both SZLFE and a crude extract from shark liver of S. zygaena showed similar kinetic characteristics of complete iron release with biphasic behavior. In addition, a combined technique of visible spectrometry and column chromatography was used for studying ratio of phosphate to Fe3+ within the SZLFE core. Interestingly, this ratio maintained invariable even after the iron release, which differed from that of other mammal ferritins.  相似文献   

19.
In the marine teleost intestine the secretion of bicarbonate increases pH of the lumen (pH 8.4 -9.0) and importantly reduces Ca2+ and Mg2+ concentrations by the formation of insoluble divalent ion carbonates. The alkaline intestinal environment could potentially also cause essential metal carbonate formation reducing bioavailability. Iron accumulation was assessed in the Gulf toadfish (Opsanus beta) gut by mounting intestine segments in modified Ussing chambers fitted to a pH-stat titration system. This system titrates to maintain lumen pH constant and in the process prevents bicarbonate accumulation. The luminal saline pH was clamped to pH 5.5 or 7.0 to investigate the effect of proton concentrations on iron uptake. In addition, redox state was altered (gassing with N2, addition of dithiothreitol (DTT) and ascorbate) to evaluate Fe3+ versus Fe2+ uptake, enabling us to compare a marine teleost intestine model for iron uptake to the mammalian system for non-haem bound iron uptake that occurs via a ferrous/proton (Fe2+/H+) symporter called Divalent Metal Transporter 1 (DMT1). None of the redox altering strategies affected iron (Fe3+ or Fe2+) binding to mucus, but the addition of ascorbate resulted in a 4.6-fold increase in epithelium iron accumulation. This indicates that mucus iron binding is irrespective of valency and suggests that ferrous iron is preferentially transported across the apical surface. Altering luminal saline pH from 7.0 to 5.5 did not affect ferric or ferrous iron uptake, suggesting that if iron is entering via DMT1 in marine fish intestine this transporter works efficiently under circumneutral conditions.  相似文献   

20.
The Fe(3+) binding site of recombinant nFbp, a ferric-binding protein found in the periplasmic space of pathogenic Neisseria, has been characterized by physicochemical techniques. An effective Fe(3+) binding constant in the presence of 350 microm phosphate at pH 6.5 and 25 degrees C was determined as 2.4 x 10(18) m(-1). EPR spectra for the recombinant Fe(3+)nFbp gave g' = 4.3 and 9 signals characteristic of high spin Fe(3+) in a strong ligand field of low (orthorhombic) symmetry. (31)P NMR experiments demonstrated the presence of bound phosphate in the holo form of nFbp and showed that phosphate can be dialyzed away in the absence of Fe(3+) in apo-nFbp. Finally, an uncorrected Fe(3+/2+) redox potential for Fe-nFbp was determined to be -290 mV (NHE) at pH 6.5, 20 degrees C. Whereas our findings show that nFbp and mammalian transferrin have similar Fe(3+) binding constants and EPR spectra, they differ greatly in their redox potentials. This has implications for the mechanism of Fe transport across the periplasmic space of Gram-negative bacteria.  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