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 62 毫秒
1.
Release of iron from ferritin requires reduction of ferric to ferrous iron. The iron can participate in the diabetogenic action of alloxan. We investigated the ability of ascorbate to catalyze the release of iron from ferritin in the presence of alloxan. Incubation of ferritin with ascorbate alone elicited iron release (33 nmol/10 min) and the generation of ascorbate free radical, suggesting a direct role for ascorbate in iron reduction. Iron release by ascorbate significantly increased in the presence of alloxan, but alloxan alone was unable to release measurable amounts of iron from ferritin. Superoxide dismutase significantly inhibited ascorbate-mediated iron release in the presence of alloxan, whereas catalase did not. The amount of alloxan radical (A·) generated in reaction systems containing both ascorbate and alloxan decreased significantly upon addition of ferritin, suggesting that A· is directly involved in iron reduction. Although release of iron from ferritin and generation of A· were also observed in reactions containing GSH and alloxan, the amount of iron released in these reactions was not totally dependent on the amount of A· present, suggesting that other reductants in addition to A· (such as dialuric acid) may be involved in iron release mediated by GSH and alloxan. These results suggest that A· is the main reductant involved in ascorbate-mediated iron release from ferritin in the presence of alloxan and that both dialuric acid and A· contribute to GSH/alloxan-mediated iron release.  相似文献   

2.
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   

3.
Reductive release of ferritin iron: a kinetic assay   总被引:1,自引:0,他引:1  
Ferritin iron release, a process of considerable interest in biology and medicine, occurs most readily in the presence of reducing agents. Here is described a kinetic assay for measuring the rate of ferritin iron removal promoted by various reductants. The new procedure uses ferrozine as a chromophoric, high-affinity chelator for the product, Fe(II). The initial rate of iron release is quantified by continuous spectrophotometric measurement of the Fe(ferrozine)2/3+ complex which absorbs maximally at 562 nm. The initial rate of iron mobilization is dependent on reductant concentration, but not on the concentration of the chelating agent, ferrozine. Saturation kinetics are observed for all reductants, including dihydroxyfumarate, cysteine, caffeic acid, ascorbate, and glutathione. Superoxide dismutase greatly inhibits ferritin iron release by ascorbate, but has little or no effect on the reducing action of dihydroxyfumarate, cysteine, caffeic acid, or glutathione. Ferritin iron removal by dihydroxyfumarate was inhibited by various metal ions. This new assay may be used for rapid screening of test compounds for treatment of iron overload and for investigation of the mechanistic aspects of ferritin iron reduction.  相似文献   

4.
The mechanism of ascorbate-promoted ferritin iron reduction under aerobic conditions was studied. The initial rate of ferritin iron release was determined by spectrophotometric measurement of the Fe(ferrozine)3(2+) complex which absorbs at 562 nm. Variation of the initial ferrozine concentration had no influence on the rate of iron release suggesting that ferrozine does not participate in the rate-determining step. Experimental measurements of the initial rate of iron release as a function of ascorbate concentration resulted in saturation kinetics with Vmax = 2.0 X 10(-7) M.min-1 and KM = 1.3 X 10(-3) M. The effect of pH was quite pronounced with a maximal rate of iron release at pH 7.0. Stoichiometric measurements on the reaction mixture, with added catalase, resulted in a ratio of 2 Fe(II) released per ascorbate. Ascorbate-mediated iron release was inhibited 85% by superoxide dismutase, but 0% inhibition was noted with aposuperoxide dismutase. It is proposed that superoxide ion, generated during the iron-promoted oxidation of ascorbate, acts as a reductant of ferritin iron. A mechanism of ferritin iron release consistent with these experimental observations is discussed.  相似文献   

5.
Ascorbate is catalytically oxidized by a coupled iron-ceruloplasmin system, the iron ions functioning as a red/ox cycling intermediate between ceruloplasmin and ascorbate. Serum albumin, an iron binding compound, was found to stimulate the ascorbate oxidation rate. It is proposed that ferrous ions react more rapidly with ceruloplasmin when they are bound to albumin. A K m value of 39 m was estimated for Fe2+-albumin. Citrate and urate inhibit the iron-ceruloplasmin-dependent ascorbate oxidation by chelating ferric ions. In the presence of albumin only citrate reduced the oxidation rate, the observation suggesting the following order of iron binding ability: citrate > albumin > urate. Physiological aspects of the results have been discussed.  相似文献   

6.
Formation of 7,8-dihydro-8-oxo-2′-deoxyguanosine (8-oxo-dG) in solutions of free 2′-deoxyguanosine (dG) and calf thymus DNA (DNA) was compared for the diffusion-dependent and localised production of oxygen radicals from phosphate-mediated oxidation of ferrous iron (Fe2+) to ferric iron (Fe3+). The oxidation of Fe2+ to Fe3+ was followed at 304 nm at pH 7.2 under aerobic conditions. Given that the concentration of Fe2+ ≥phosphate concentration, the rate of Fe2+ oxidation was significantly higher in DNA-phosphate as compared for the same concentration of inorganic phosphate. Phosphate catalysed oxidation of ferrous ions in solutions of dG or DNA led through the production of reactive oxygen species to the formation of 8-oxo-dG. The yield of 8-oxo-dG in solutions of dG or DNA correlated positively with the inorganic-/DNA-phosphate concentrations as well as with the concentrations of ferrous ions added. The yield of 8-oxo-dG per unit oxidised Fe2+ were similar for dG and DNA; thus, it differed markedly from radiation-induced 8-oxo-dG, where the yield in DNA was several fold higher.For DNA in solution, the localisation of the phosphate ferrous iron complex relative to the target is an important factor for the yield of 8-oxo-dG. This was supported from the observation that the yield of 8-oxo-dG in solutions of dG was significantly increased over that in DNA only when Fe2+ was oxidised in a high excess of inorganic phosphate (50 mM) and from the lower protection of DNA damage by the radical scavenger (hydroxymethyl)aminomethane (Tris)–HCl.  相似文献   

7.

Background

To satisfy their requirement for iron while at the same time countering the toxicity of this highly reactive metal ion, prokaryotes have evolved proteins belonging to two distinct sub-families of the ferritin family: the bacterioferritins (BFRs) and the bacterial ferritins (Ftns). Recently, Ftn homologues have also been identified and characterised in archaeon species. All of these prokaryotic ferritins function by solubilising and storing large amounts of iron in the form of a safe but bio-available mineral.

Scope of review

The mechanism(s) by which the iron mineral is formed by these proteins is the subject of much current interest. Here we review the available information on these proteins, with particular emphasis on significant advances resulting from recent structural, spectroscopic and kinetic studies.

Major conclusions

Current understanding indicates that at least two distinct mechanisms are in operation in prokaryotic ferritins. In one, the ferroxidase centre acts as a true catalytic centre in driving Fe2+ oxidation in the cavity; in the other, the centre acts as a gated iron pore by oxidising Fe2+ and transferring the resulting Fe3+ into the central cavity.

General significance

The prokaryotic ferritins exhibit a wide variation in mechanisms of iron core mineralisation. The basis of these differences lies, at least in part, in structural differences at and around the catalytic centre. However, it appears that more subtle differences must also be important in controlling the iron chemistry of these remarkable proteins.  相似文献   

8.
Free radical damage caused by ferrous iron is involved in the pathogenesis of secondary brain injury after intracerebral hemorrhage (ICH). NF‐E2‐related factor 2 (Nrf2), a major phase II gene regulator that binds to antioxidant response element, represents an important cellular cytoprotective mechanism against oxidative damage. We hypothesized that Nrf2 might protect astrocytes from damage by Fe2+. Therefore, we examined cytotoxicity in primary astrocytes induced by iron overload and evaluated the effects of Fe2+ on Nrf2 expression. The results demonstrated that 24‐h Fe2+ exposure exerted time‐ and concentration‐dependent cytotoxicity in astrocytes. Furthermore, Fe2+ exposure in astrocytes resulted in time‐ and concentration‐dependent increases in Nrf2 expression, which preceded Fe2+ toxicity. Nrf2‐specific siRNA further knocked down Nrf2 levels, resulting in greater Fe2+‐induced astrocyte cytotoxicity. These data indicate that induction of Nrf2 expression could serve as an adaptive self‐defense mechanism, although it is insufficient to completely protect primary astrocytes from Fe2+‐induced neurotoxicity.  相似文献   

9.
l-Lactate-driven ferric and nitrate reduction was studied in Escherichia coli E4. Ferric iron reduction activity in E. coli E4 was found to be constitutive. Contrary to nitrate, ferric iron could not be used as electron acceptor for growth. Ferric iron reductase activity of 9 nmol Fe2+ mg-1 protein min-1 could not be inhibited by inhibitors for the respiratory chain, like Rotenone, quinacrine, Actinomycin A, or potassium cyanide. Active cells and l-lactate-driven nitrate respiration in E. coli E4 leading to the production of nitrite, was reduced to about 20% of its maximum activity with 5 mM ferric iron, or to about 50% in presence of 5 mM ferrous iron. The inhibition was caused by nitric oxide formed by a purely chemical reduction of nitrite by ferrous iron. Nitric oxide was further chemically reduced by ferrous iron to nitrous oxide. With electron paramagnetic resonance spectroscopy, the presence of a free [Fe2+-NO] complex was shown. In presence of ferrous or ferric iron and l-lactate, nitrate was anaerobically converted to nitric oxide and nitrous oxide by the combined action of E. coli E4 and chemical reduction reactions (chemodenitrification).  相似文献   

10.
The ferrozine assay is a widely used colorimetric method for determining soluble iron concentrations. We provide evidence for a heretofore unrecognized interference of ferric ions (Fe3 +) on ferrous iron (Fe2 +) measurements performed in the dark. Fe3 + concentrations affected the absorbance measurements, which linearly increased with incubation time.  相似文献   

11.
Ferritin has a high capacity as an iron store, incorporating some 4500 iron atoms as a microcrystalline ferric oxide hydrate. Starting from apoferritin, or ferritin of low iron content, Fe2+ and an oxidizing agent, the uptake of iron can be recorded spectrophotometrically. Progress curves were obtained and the reconstituted ferritin was shown by several physical methods to be similar to natural ferritin. The progress curves of iron uptake by apoferritin are sigmoidal; those for ferritins of low iron content are hyperbolic. The rate of iron uptake is dependent on the amount of iron already present in the molecule. The distribution of iron contents among reconstituted ferritin molecules is inhomogeneous. These findings are interpreted in terms of a crystal growth model. The surface area of the crystallites forming inside the protein increases until the molecule is half full, and then declines. This surface controls the rate at which new material is deposited. The experimental results can best be accounted for by a two-stage mechanism, an initial slow `nucleation' stage, which is apparently zero order with respect to [Fe2+], followed by a more rapid `growth' stage. The rate of Fe2+ oxidation is increased in the presence of apoferritin as compared with controls. Ferritin can therefore be regarded as an enzyme to which the product remains firmly attached. The protein appears to increase the rate of `nucleation'. The apparent zero order of this stage suggests the presence of binding sites on the protein, which are saturated with respect to Fe2+. These sites are presumed also to be oxidation sites. The oxidation and subsequent formation of the ferric oxide hydrate may proceed according to one of three alternative models.  相似文献   

12.
[3H]Acetylcholine (ACh) release, malonaldehyde formation and45calcium-uptake were measured in rat cerebral cortical nerve terminal that were exposed to various concentrations of ferrous and ascorbate ions. At a constant molar ratio of 25:1, ferrous:ascorbate, these ions increased malonaldehyde (MA) synthesis in a concentration-dependent manner. Treatment with these ions in the same ratio also induced a dose-related inhibition of the K+-depolarization-induced release of newly synthesized [3H]ACh. Combined exposure to Fe2+/ascorbate also reduced calcium ionophore A23187-induced [3H]ACh release. Neither ferrous nor ascorbate ions alone altered depolarization-or ionophore-induced [3H]ACh release over this concentration range. Depolarization- and A23187-induced45calcium uptake were not affected by peroxidation, suggesting that membrane peroxidation influenced some process in the release-process subsequent to calcium influx in a manner similar to what is observed during aging.  相似文献   

13.
The effects of ascorbate and a-tocopherol as antioxidants and as co-operative factors against NADPH-dependent lipid peroxidation in human placental mitochondria have been studied. The addition of ascorbate at low concentration (up to 50 M) to the NADPH-generating system resulted in increasing lipid peroxidation and Fe3+ to Fe2+ reduction. High concentration of ascorbate (150 M), which produced maximal rate of ascorbate-dependent lipid peroxidation, was found to inhibit almost completely NADPH-dependent lipid peroxidation by maintaining too much iron in its reduced form. Either stimulatory or inhibitory effect of ascorbate on NADPH-dependent lipid peroxidation depends on the appropriate Fe3+/Fe2+ ratio. -Tocopherol caused a decrease of NADPH-dependent lipid peroxidation, inhibiting completely this process at 150 M concentration. The inhibitory effect of -tocopherol increased rapidly with the increasing ascorbate concentration, almost complete inhibition of NADPH-dependent lipid peroxidation being obtained at 25 M -tocopherol and 50 M ascorbate. This strong inhibitory combined effect of -tocopherol and ascorbate was independent of the Fe3+/Fe2+ ratio, as a-tocopherol is not able to reduce Fe3+ to Fe2+ under the conditions employed. These findings suggest that antioxidant effects of ascorbate in placental mitochondria are mediated by recycling of a-tocopherol rather than by strong reduction of Fe3+ to Fe2+. On the basis of the results obtained, we assume that adequate concentrations of a-tocopherol and ascorbate in placental tissue may prevent the release of lipid peroxide from placental mitochondria and therefore could be protective against the development of preeclampsia.  相似文献   

14.
Iron stored in phytoferritin plays an important role in the germination and early growth of seedlings. The protein is located in the amyloplast where it stores large amounts of iron as a hydrated ferric oxide mineral core within its shell-like structure. The present work was undertaken to study alternate mechanisms of core formation in pea seed ferritin (PSF). The data reveal a new mechanism for mineral core formation in PSF involving the binding and oxidation of iron at the extension peptide (EP) located on the outer surface of the protein shell. This binding induces aggregation of the protein into large assemblies of ∼400 monomers. The bound iron is gradually translocated to the mineral core during which time the protein dissociates back into its monomeric state. Either the oxidative addition of Fe2+ to the apoprotein to form Fe3+ or the direct addition of Fe3+ to apoPSF causes protein aggregation once the binding capacity of the 24 ferroxidase centers (48 Fe3+/shell) is exceeded. When the EP is enzymatically deleted from PSF, aggregation is not observed, and the rate of iron oxidation is significantly reduced, demonstrating that the EP is a critical structural component for iron binding, oxidation, and protein aggregation. These data point to a functional role for the extension peptide as an iron binding and ferroxidase center that contributes to mineralization of the iron core. As the iron core grows larger, the new pathway becomes less important, and Fe2+ oxidation and deposition occurs directly on the surface of the iron core.The chemistry of iron and oxygen in a number of non-heme di-iron proteins has been a subject of intense interest because of their varied roles in oxygen activation and catalysis, substrate hydrolysis, oxygen transport, redox reactions, H2O2, and iron detoxification and iron storage. Di-iron centers have similar structural motifs consisting of a combination of carboxylate and histidine ligands that either bind or bridge the two metal ions of the di-nuclear active site; di-iron proteins containing these centers include methane monooxygenase, ribonucleotide reductase, rubrerythrin, stearoyl desaturase, purple acid phosphatase, hemerythrin, the Dps proteins, and ferritins (16). Despite their similar di-nuclear centers, each of these proteins fulfills a distinct biological role that seems to be mediated by the nature of the first and second coordination sphere of the di-iron center.Ferritins are a class of intracellular iron storage and detoxification proteins that facilitate the oxidation of iron by molecular oxygen or hydrogen peroxide to form a hydrous ferric oxide mineral core within their interiors (13). Rapid Fe2+ oxidation occurs at the di-iron ferroxidase center located on the H-subunit of the mammalian protein. Unlike the H-chain, the more acidic L-subunit lacks the ferroxidase center but contains a putative nucleation site responsible for slower iron oxidation and mineralization (7). The shapes of both the H- and L-subunit are nearly cylindrical and composed of a four-α-helix bundle containing two antiparallel helix pairs (A, B and C, D) connected by a long non-helical stretch, the BC-loop, between B and C helices. A fifth short helix (E helix) lies at one end of the bundle at 60° to its axis (1, 2).From an evolutionary point of view, plant and animal ferritins arose from a common ancestor, but plant ferritins exhibit various specific features as compared with animal ferritins. Plant ferritins are observed in plastids (chloroplasts in leaves, amyloplasts in tubers and seeds, etc.) where iron is incorporated into the ferritin shell to form the mineral core, whereas animal ferritins are largely found in the cytoplasm of cell (1, 8). Ferritins from dried soybean and pea seed consist of two subunits of 26.5 and 28.0 kDa, which are designated H-1 and H-2, respectively, share ∼80% amino acid sequence identity (9, 10), and contain ferroxidase centers. The two subunits are synthesized from a 32-kDa precursor with a unique two-domain N-terminal sequence containing a transit peptide (TP)2 and an extension peptide (EP). The TP is responsible for precursor targeting to plastids (11). Upon transport to plastids, the TP is cleaved from the subunit precursor, resulting in the formation of the mature subunit which assembles into a 24-mer apoferritin within the plastids (11, 12). The EP domain is kept at the N-terminal extremity of the mature plant ferritins, but its function remains unknown. It is absent in animal ferritins. For soybean ferritin, further processing of the 26.5-kDa subunit occurs through excising a short C-terminal amino acid sequence (16 amino acid residues) that corresponds to the E helix of the mammalian ferritin subunit (9). Compared with their own precursors, H-1 is devoid of both the N-terminal TP domain and the C-terminal E helix, whereas H-2 lacks the N-terminal TP domain only (13). However, the amino acids constituting the ferroxidase center are strictly conserved in all the plant ferritins except for pea seed ferritin where His-62 replaces Glu-62 (14, 15).Iron oxidation/mineralization in human ferritin occurs by at least three reaction pathways (16, 17). After Fe2+ binding at the ferroxidase site, the protein-catalyzed oxidation of Fe2+ occurs, H2O2 is the product of O2 reduction, and a mineral core of Fe3+ is produced, written for simplicity as Fe(O)OH(core) (Equation 1) (18, 19). Some of the H2O2 generated in Equation 1 reacts with additional Fe2+ in the Fe2+ + H2O2 detoxification reaction (Equation 2) to produce H2O (16, 20). Once a mineral core of sufficient size has developed, Fe2+ autoxidation becomes significant, and iron oxidation and hydrolysis occurs primarily on the growing surface of the mineral through an autocatalytic process where O2 is reduced completely to H2O (Equation 3) (1, 7, 17). Based on the above reactions and on identification of intermediates by resonance Raman spectroscopy, Mössbauer spectroscopy, and EXAFS (extended x-ray absorption fine structure), the mechanism of mineral core formation in mammalian ferritins has been reasonably well established (18, 2123). Initially two Fe2+ are oxidized by O2 at the ferroxidase center to form a transient μ-1,2-peroxodi-Fe3+ intermediate, which rapidly decays to form a μ-oxo di-Fe3+ complex(es), releasing H2O2 to the solution. The μ-oxo(hydroxo) bridged di-Fe3+ dimer(s) then translocates from the ferroxidase center to the inner cavity to form an incipient core, which ultimately leads to the formation of the mineral core itself. However, this mechanism of oxidative deposition of iron is only applicable to ferritins such as horse spleen ferritin and human H-chain ferritin, which regenerate activity at their ferroxidase centers. In contrast, the ferroxidase centers of human mitochondrial ferritin (24), Escherichia coli bacterioferritin (25), E. coli bacterial ferritin A (26), and pea seed ferritin (PSF) (27) lack significant regeneration activity, and Fe3+ produced at their centers migrates with difficulty from the center to the cavity to form the initial core. These observations raise the question as to whether alternate pathways exist for core formation in these proteins, in particular in phytoferritins, the focus of this work.In the present study we have looked for alternate mechanisms of iron deposition in PSF and demonstrate that ferritin-ferritin aggregation occurs during the oxidative deposition of iron in PSF when more than 48 Fe2+/shell are added to the apoprotein (apoPSF), an amount exceeding the 24 ferroxidase center binding capacity for Fe2+/3+. Such aggregation was also induced by the addition of Fe3+ directly to PSF. The data indicate that binding occurs on the outer surface of the ferritin shell at sites involving the extension peptide. Upon standing, the iron induced aggregate reversibly dissociates to its original undissociated form, whereas the iron migrates to the mineral core. This finding represents an alternate pathway for iron deposition in phytoferritin.  相似文献   

15.
Iron and copper ions, in their unbound form, may lead to the generation of reactive oxygen species via Haber–Weiss and/or Fenton reactions. In addition, it has been shown that copper ions can irreversibly and non-specifically bind to thiol groups in proteins. This non-specific binding property has not been fully addressed for iron ions. Thus, the present study compares both the pro-oxidant and the non-specific binding properties of Fe3+ and Cu2+, using rat liver cytosol and microsomes as biological systems. Our data show that, in the absence of proteins, Cu2+/ascorbate elicited more oxygen consumption than Fe3+/ascorbate under identical conditions. Presence of cytosolic and microsomal protein, however, differentially altered oxygen consumption patterns. In addition, Cu2+/ascorbate increased microsomal lipid peroxidation and decreased cytosolic and microsomal content of thiol groups more efficiently than Fe3+/ascorbate. Finally, Fe3+/ascorbate and Cu2+/ascorbate inhibited in different ways cytosolic and microsomal glutathione S-transferase (GST) activities, which are differentially sensitive to oxidants. Moreover, in the absence of ascorbate, only Cu2+ decreased the content of cytosolic and microsomal thiol groups and inhibited cytosolic and microsomal GST activities. Catechin partially prevented the damage to thiol groups elicited by Fe3+/ascorbate and Cu2+/ascorbate but not by Cu2+ alone. N-Acetylcysteine completely prevented the damage elicited by Cu2+/ascorbate, Fe3+/ascorbate and Cu2+ alone. N-Acetylcysteine also completely reversed the damage to thiol groups elicited by Fe3+/ascorbate, partially reversed that of Cu2+/ascorbate but failed to reverse the damage promoted by Cu2+ alone. Our data are discussed in terms to the potential damage that the accumulation of iron and copper ions can promote in biological systems.  相似文献   

16.
Summary The release of iron from ferritin is important in the formation of iron proteins and for the management of diseases in both animals and plants associated with abnormal accumulations of ferritin iron. Much more iron can be released experimentally by reduction of the ferric hydrous oxide core than by chelation of Fe3+ which has led to the notion that reduction is also the major aspect of iron release in vivo. Variations in the kinetics of reduction of the mineral core of ferritin have been attributed to the redox potential of the reductant, redox properties of the iron core, the structure of the protein coat, the analytical method used to detect Fe2+ and reactions at the surface of the mineral. Direct measurements of the oxidation state of the iron during reduction has never been used to analyze the kinetics of reduction, although Mössbauer spectroscopy has been used to confirm the extent of reduction after electrochemical reduction using dispersive X-ray absorption spectroscopy (DXAS). We show that the near edge of X-ray absorption spectra (XANES) can be used to quantify the relative amounts of Fe2+ and Fe3+ in mixtures of the hydrated ions. Since the nearest neighbors of iron in the ferritin iron core do not change during reduction, XANES can be used to monitor directly the reduction of the ferritin iron core. Previous studies of iron core reduction which measured by Fe2+ · bipyridyl formation, or coulometric reduction with different mediators, suggested that rates depended mainly on the redox potential of the electron donor. When DXAS was used to measure the rate of reduction directly, the initial rate was faster than previously measured. Thus, previously measured differences in reduction rates appear to be influenced by the accessibility of Fe2+ to the complexing reagent or by the electrochemical mediator. In the later stages of ferritin iron core dissolution, reduction rates drop dramatically whether measured by DXAS or formation of Fe2+ complexes. Such results emphasize the heterogeneity of ferritin core structure.  相似文献   

17.
Reduction and release of ferritin iron by plant phenolics   总被引:1,自引:0,他引:1  
The reductive release of ferritin iron by several naturally occurring o-diphenols was studied. The initial rate of iron release was quantified by spectrophotometric measurement of the Fe(ferrozine)3(2+) complex, which absorbs maximally at 562 nm. The initial rate of iron release was dependent upon o-diphenol concentration, but not on the concentration of the chromophoric chelating agent, ferrozine, Stoichiometric measurements resulted in a ratio of 2Fe(II) released per molecule of o-diphenol. The series of o-diphenols studied included, caffeic acid, chlorogenic acid, dihydrocaffeic acid, 3,4-dihydroxybenzoic acid, and several analogs. These reductants represent an oxidation reduction potential range of 0.38 volts. A direct correlation between reducing power of the o-diphenols and rate of ferritin iron release was observed. Superoxide dismutase, catalase, mannitol, or general radical traps had no effect on the rate of iron removal; however, EDTA and oxalate inhibited iron release. A mechanism for ferritin iron reduction and release by o-diphenols consistent with the experimental observations is discussed.  相似文献   

18.
It has been previously reported that iron release from ferritin could be promoted by nitric oxide (NO) generated from sodium nitroprusside. It was thus proposed that some of the toxic effects of NO could be related to its ability to increase intracellular free iron concentrations and generate an oxidative stress. On the contrary, the iron exchange experiments reported here show that NO from S-nitrosothiols is unable to promote iron release from ferritin. The discrepancy may be explained by the disregarded ability of ferrozine, the ferrous trap used in the previous report, to mobilize iron both from ferritin and from sodium nitroprusside spontaneously.  相似文献   

19.
The release of iron from horse spleen ferritin by the chelating agents desferrioxamine B, rhodotorulic acid, 2,3-dihydroxybenzoate, 2,2′-bipyridyl and pyridine-2-aldehyde-2-pyridyl hydrazone (Paphy) has been studied in vitro. Ferritin prepared by classical procedures involving thermal denaturation releases its iron less effectively than ferritin isolated by a modified procedure that avoids this step. Desferrioxamine B and rhodotorulic acid are the most effective in releasing iron from both preparations of ferritin. When FMN is added, iron release by desferrioxamine B, rhodotorulic acid, and 2,3-dihydroxybenzoate was effectively blocked, whereas both bipyridyl and Paphy showed a marked simulation. A substantial increase in iron release was also observed for bipyridyl and Paphy with ascorbate; a less important increase was noted for rhodotorulic acid. EDTA exerted a marked inhibition of iron release from ferritin with rhodotorulic acid, 2,3-dihydroxybenzoate, bipyridyl, and Paphy. The effects of citrate and oxalate on iron release by the chelators was small. The effect of the concentration of flavin on iron release from ferritin by bipyridyl and desferrioxamine B have been studied. Desferrioxamine is unable to mobilize FeII from ferritin following reduction by reduced FMN, whereas bipyridyl can rapidly complex the ferrous iron. The results are discussed in the context of our current concepts of storage iron mobilization in the treatment of iron overload.  相似文献   

20.

Iron is implicated in ocular diseases such as in age-related macular degeneration. Light is also considered as a pathological factor in this disease. Earlier, two studies reported the influence of constant light environment on the pattern of expressions of iron-handling proteins. Here, we aimed to see the influence of light in 12-h light–12-h dark (12L:12D) cycles on the expression of iron-handling proteins in chick retina. Chicks were exposed to 400 lx (control) and 5000 lx (experimental) light at 12L:12D cycles and sacrificed at variable timepoints. Retinal ferrous ion (Fe2+) level, ultrastructural changes, lipid peroxidation level, immunolocalization and expression patterns of iron-handling proteins were analysed after light exposure. Both total Fe2+ level (p?=?0.0004) and lipid peroxidation (p?=?0.002) significantly increased at 12-, 48- and 168-h timepoint (for Fe2+) and 48- and 168-h timepoint (for lipid peroxidation), and there were degenerative retinal changes after 168 h of light exposure. Intense light exposure led to an increase in the levels of transferrin and transferrin receptor-1 (at 168-h) and ferroportin-1, whereas the levels of ferritins, hephaestin, (at 24-, 48- and 168-h timepoint) and ceruloplasmin (at 168-h timepoint) were decreased. These changes in iron-handling proteins after light exposure are likely due to a disturbance in the iron storage pool evident from decreased ferritin levels, which would result in increased intracellular Fe2+ levels. To counteract this, Fe2+ is released into the extracellular space, an observation supported by increased expression of ferroportin-1. Ceruloplasmin was able to convert Fe2+ into Fe3+ until 48 h of light exposure, but its decreased expression with time (at 168-h timepoint) resulted in increased extracellular Fe2+ that might have caused oxidative stress and retinal cell damage.

  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