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 46 毫秒
1.
The discrepancy between predicted and measured H(2)O(2) formation during iron deposition with recombinant heavy human liver ferritin (rHF) was attributed to reaction with the iron protein complex [Biochemistry 40 (2001) 10832-10838]. This proposal was examined by stopped-flow kinetic studies and analysis for H(2)O(2) production using (1) rHF, and Azotobacter vinelandii bacterial ferritin (AvBF), each containing 24 identical subunits with ferroxidase centers; (2) site-altered rHF mutants with functional and dysfunctional ferroxidase centers; and (3) recombinant human liver light ferritin (rLF), containing no ferroxidase center. For rHF, nearly identical pseudo-first-order rate constants of 0.18 s(-1) at pH 7.5 were measured for Fe(2+) oxidation by both O(2) and H(2)O(2), but for rLF, the rate with O(2) was 200-fold slower than that for H(2)O(2) (k = 0.22 s(-1)). A Fe(2+)/O(2) stoichiometry near 2.4 was measured for rHF and its site altered forms, suggesting formation of H(2)O(2). Direct measurements revealed no H(2)O(2) free in solution 0.5-10 min after all Fe(2+) was oxidized at pH 6.5 or 7.5. These results are consistent with initial H(2)O(2) formation, which rapidly reacts in a secondary reaction with unidentified solution components. Using measured rate constants for rHF, simulations showed that steady-state H(2)O(2) concentrations peaked at 14 muM at approximately 600 ms and decreased to zero at 10-30 s. rLF did not produce measurable H(2)O(2) but apparently conducted the secondary reaction with H(2)O(2). Fe(2+)/O(2) values of 4.0 were measured for AvBF. Stopped-flow measurements with AvBF showed that both H(2)O(2) and O(2) react at the same rate (k = 0.34 s(-1)), that is faster than the reactions with rHF. Simulations suggest that AvBF reduces O(2) directly to H(2)O without intermediate H(2)O(2) formation.  相似文献   

2.
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   

3.
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   

4.
The reaction of horse spleen ferritin (HoSF) with Fe2+ at pH 6.5 and 7.5 using O2, H2O2 and 1:1 a mixture of both showed that the iron deposition reaction using H2O2 is approximately 20- to 50-fold faster than the reaction with O2 alone. When H2O2 was added during the iron deposition reaction initiated with O2 as oxidant, Fe2+ was preferentially oxidized by H2O2, consistent with the above kinetic measurements. Both the O2 and H2O2 reactions were well defined from 15 to 40 degrees C from which activation parameters were determined. The iron deposition reaction was also studied using O2 as oxidant in the presence and absence of catalase using both stopped-flow and pumped-flow measurements. The presence of catalase decreased the rate of iron deposition by approximately 1.5-fold, and gave slightly smaller absorbance changes than in its absence. From the rate constants for the O2 (0.044 s(-1)) and H2O2 (0.67 s(-1)) iron-deposition reactions at pH 7.5, simulations of steady-state H2O2 concentrations were computed to be 0.45 microM. This low value and reported Fe2+/O2 values of 2.0-2.5 are consistent with H2O2 rapidly reacting by an alternate but unidentified pathway involving a system component such as the protein shell or the mineral core as previously postulated [Biochemistry 22 (1983) 876; Biochemistry 40 (2001) 10832].  相似文献   

5.
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   

6.
Zhao G  Bou-Abdallah F  Yang X  Arosio P  Chasteen ND 《Biochemistry》2001,40(36):10832-10838
The ferritins are a class of iron storage and detoxification proteins that play a central role in the biological management of iron. These proteins have a catalytic site, "the ferroxidase site", located on the H-type subunit that facilitates the oxidation of Fe(II) to Fe(III) by O(2). Measurements during the past 10 years on a number of vertebrate ferritins have provided evidence that H(2)O(2) is produced at this diiron ferroxidase site. Recently reported experiments using three different analytical methods with horse spleen ferritin (HoSF) have failed to detect H(2)O(2) production in this protein [Lindsay, S., Brosnahan, D., and Watt, G. D. (2001) Biochemistry 40, 3340-3347]. These findings contrast with earlier results reporting H(2)O(2) production in HoSF [Xu, B., and Chasteen, N. D. (1991) J. Biol. Chem. 266, 19965-19970]. Here a sensitive fluorescence assay and an assay based on O(2) evolution in the presence of catalase were used to demonstrate that H(2)O(2) is produced in HoSF as previously reported. However, because of the relatively few H-chain ferroxidase sites in HoSF and the reaction of H(2)O(2) with the protein, H(2)O(2) is more difficult to detect in this ferritin than in recombinant human H-chain ferritin (HuHF). The proper sequence of addition of reagents is important for measurement of the total amount of H(2)O(2) produced during the ferroxidation reaction.  相似文献   

7.
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   

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.
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   

10.
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   

11.
Listeria innocua Dps (DNA binding protein from starved cells) affords protection to DNA against oxidative damage and can accumulate about 500 iron atoms within its central cavity through a process facilitated by a ferroxidase center. The chemistry of iron binding and oxidation in Listeria Dps (LiDps, formerly described as a ferritin) using H(2)O(2) as oxidant was studied to further define the mechanism of iron deposition inside the protein and the role of LiDps in protecting DNA from oxidative damage. The relatively strong binding of 12 Fe(2+) to the apoprotein (K(D) approximately 0.023 microM) was demonstrated by isothermal titration calorimetry, fluorescence quenching, and pH stat experiments. Hydrogen peroxide was found to be a more efficient oxidant for the protein-bound Fe(2+) than O(2). Iron(II) oxidation by H(2)O(2) occurs with a stoichiometry of 2 Fe(2+)/H(2)O(2) in both the protein-based ferroxidation and subsequent mineralization reactions, indicating complete reduction of H(2)O(2) to H(2)O. Electron paramagnetic resonance (EPR) spin-trapping experiments demonstrated that LiDps attenuates the production of hydroxyl radical by Fenton chemistry. DNA cleavage assays showed that the protein, while not binding to DNA itself, protects it against the deleterious combination of Fe(2+) and H(2)O(2). The overall process of iron deposition and detoxification by LiDps is described by the following equations. For ferroxidation, Fe(2+) + Dps(Z)--> [(Fe(2+))-Dps](Z+1) + H(+) (Fe(2+) binding) and [(Fe(2+))-Dps](Z+1) + Fe(2+) + H(2)O(2) --> [(Fe(3+))(2)(O)(2)-Dps](Z+1) + 2H(+) (Fe(2+) oxidation/hydrolysis). For mineralization, 2Fe(2+) + H(2)O(2) + 2H(2)O --> 2Fe(O)OH((core)) + 4H(+) (Fe(2+) oxidation/hydrolysis). These reactions occur in place of undesirable odd-electron redox processes that produce hydroxyl radical.  相似文献   

12.
Ferrous ion (Fe(2+)) is long thought to be the most likely active species, producing oxidants through interaction of Fe(2+) with oxygen (O(2)). Because current iron overload therapy uses only Fe(3+) chelators, such as desferrioxamine (DFO), we have tested a hypothesis that addition of a Fe(2+) chelator, 2,2'-dipyridyl (DP), may be more efficient and effective in preventing iron-induced oxidative damage in human liver HepG2 cells than DFO alone. Using ferrozine as an assay for iron measurement, levels of cellular iron in HepG2 cells treated with iron compounds correlated well with the extent of lipid peroxidation (r = 0.99 after log transformation). DP or DFO alone decreased levels of iron and lipid peroxidation in cells treated with iron. DFO + DP together had the most significant effect in preventing cells from lipid peroxidation but not as effective in decreasing overall iron levels in the cells. Using ESR spin trapping technique, we further tested factors that can affect oxidant-producing activity of Fe(2+) with dissolved O(2) in a cell-free system. Oxidant formation enhanced with increasing Fe(2+) concentrations and reached a maximum at 5 mM of Fe(2+). When the concentration of Fe(2+) was increased to 50 mM, the oxidant-producing activity of Fe(2+) sharply decreased to zero. The initial ratio of Fe(3+):Fe(2+) did not affect the oxidant producing activity of Fe(2+). However, an acidic pH (< 3.5) significantly slowed down the rate of the reaction. Our results suggest that reaction of Fe(2+) with O(2) is an important one for oxidant formation in biological system, and therefore, drugs capable of inhibiting redox activity of Fe(2+) should be considered in combination with a Fe(3+) chelator for iron overload chelation therapy.  相似文献   

13.
The redox reactivities of air-oxidized apo horse spleen ferritin (HoSF) and apo rat liver ferritin (RaF) were examined by microcoulometry and reductive optical titrations. Microcoulometry on several independent lots of commercial HoSF revealed two distinct types of redox activity: one requiring 3-4 electrons and one requiring 6-7 electrons for full reduction of the protein shell. ApoRaF required 8-9 electrons to fully reduce the oxidized form. Reductive optical titrations confirmed the microcoulometric reduction stoichiometry and, in addition, showed that the spectra of both oxidized and reduced apoHoSF were distinct and possessed absorbances tailing into the visible region. The redox reactivity of both apoRaF and apoHoSF correlated with their H-subunit composition. Identical microcoulometric and optical experiments were conducted with recombinant apo human liver heavy (rHuHF) and light (rHuLF) ferritins, but neither was redox-active. These results suggest that the redox reactivity of native ferritins is due to their heteropolymeric nature. This was confirmed by mixing various proportions of rHuHF and rHuLF, dissociating the 24-mers into individual subunits with guanidine hydrochloride at pH 3.5, and renaturing to form heteropolymeric 24-mers. Microcoulometric measurements of these apoheteropolymers reassembled in vitro showed that they were redox-active like their native apoheteropolymer counterparts. The redox activity of these apoheteropolymers increased with H-subunit composition, reached a maximum near 12 H- and 12 L-subunits, and then declined to zero with increasing L-subunit composition. The decline in redox reactivity at high L-subunit concentrations indicates that both H- and L-subunits are involved in forming the observed redox centers. Apoheteropolymers formed from rHuLF and W93F (an H-chain mutant) were redox-inactive, suggesting that the conserved tryptophan is necessary for redox center formation.  相似文献   

14.
Schwartz JK  Liu XS  Tosha T  Diebold A  Theil EC  Solomon EI 《Biochemistry》2010,49(49):10516-10525
DNA protection during starvation (Dps) proteins are miniferritins found in bacteria and archaea that provide protection from uncontrolled Fe(II)/O radical chemistry; thus the catalytic sites are targets for antibiotics against pathogens, such as anthrax. Ferritin protein cages synthesize ferric oxymineral from Fe(II) and O(2)/H(2)O(2), which accumulates in the large central cavity; for Dps, H(2)O(2) is the more common Fe(II) oxidant contrasting with eukaryotic maxiferritins that often prefer dioxygen. To better understand the differences in the catalytic sites of maxi- versus miniferritins, we used a combination of NIR circular dichroism (CD), magnetic circular dichroism (MCD), and variable-temperature, variable-field MCD (VTVH MCD) to study Fe(II) binding to the catalytic sites of the two Bacillus anthracis miniferritins: one in which two Fe(II) react with O(2) exclusively (Dps1) and a second in which both O(2) or H(2)O(2) can react with two Fe(II) (Dps2). Both result in the formation of iron oxybiomineral. The data show a single 5- or 6-coordinate Fe(II) in the absence of oxidant; Fe(II) binding to Dps2 is 30× more stable than Dps1; and the lower limit of K(D) for binding a second Fe(II), in the absence of oxidant, is 2-3 orders of magnitude weaker than for the binding of the single Fe(II). The data fit an equilibrium model where binding of oxidant facilitates formation of the catalytic site, in sharp contrast to eukaryotic M-ferritins where the binuclear Fe(II) centers are preformed before binding of O(2). The two different binding sequences illustrate the mechanistic range possible for catalytic sites of the family of ferritins.  相似文献   

15.
The hollow sphere-shaped 24-meric ferritin can store large amounts of iron as a ferrihydrite-like mineral core. In all subunits of homomeric ferritins and in catalytically active subunits of heteromeric ferritins a diiron binding site is found that is commonly addressed as the ferroxidase center (FC). The FC is involved in the catalytic Fe(II) oxidation by the protein; however, structural differences among different ferritins may be linked to different mechanisms of iron oxidation. Non-heme ferritins are generally believed to operate by the so-called substrate FC model in which the FC cycles by filling with Fe(II), oxidizing the iron, and donating labile Fe(III)–O–Fe(III) units to the cavity. In contrast, the heme-containing bacterial ferritin from Escherichia coli has been proposed to carry a stable FC that indirectly catalyzes Fe(II) oxidation by electron transfer from a core that oxidizes Fe(II). Here, we put forth yet another mechanism for the non-heme archaeal 24-meric ferritin from Pyrococcus furiosus in which a stable iron-containing FC acts as a catalytic center for the oxidation of Fe(II), which is subsequently transferred to a core that is not involved in Fe(II)-oxidation catalysis. The proposal is based on optical spectroscopy and steady-state kinetic measurements of iron oxidation and dioxygen consumption by apoferritin and by ferritin preloaded with different amounts of iron. Oxidation of the first 48 Fe(II) added to apoferritin is spectrally and kinetically different from subsequent iron oxidation and this is interpreted to reflect FC building followed by FC-catalyzed core formation.  相似文献   

16.
Ferritins are ubiquitous iron mineralizing and storage proteins that play an important role in iron homeostasis. Although excess iron is stored in the cytoplasm, most of the metabolically active iron is processed in the mitochondria of the cell. Little is known about how these organelles regulate iron homeostasis and toxicity. The recently discovered human mitochondrial ferritin (MtF), unlike other mammalian ferritins, is a homopolymer of 24 subunits that has a high degree of sequence homology with human H-chain ferritin (HuHF). Parallel experiments with MtF and HuHF reported here reveal striking differences in their iron oxidation and hydrolysis chemistry despite their similar diFe ferroxidase centers. In contrast to HuHF, MtF does not regenerate its ferroxidase activity after oxidation of its initial complement of Fe(II) and generally has considerably slower ferroxidation and mineralization activities as well. MtF exhibits sigmoidal kinetics of mineralization more characteristic of an L-chain than an H-chain ferritin. Site-directed mutagenesis reveals that serine 144, a residue situated near the ferroxidase center in MtF but absent from HuHF, is one player in this impairment of activity. Additionally only one-half of the 24 ferroxidase centers of MtF are functional, further contributing to its lower activity. Stopped-flow absorption spectrometry of Fe(II) oxidation by O(2) in MtF shows the formation of a transient diiron(III) mu-peroxo species (lambda(max) = 650 nm) as observed in HuHF. Also, as for HuHF, minimal hydroxyl radical is produced during the oxidative deposition of iron in MtF using O(2) as the oxidant. However, the 2Fe(II) + H(2)O(2) detoxification reaction found in HuHF does not occur in MtF. The structural differences and the physiological implications of the unique iron oxidation properties of MtF are discussed in light of these results.  相似文献   

17.
Based on their nanocage architectures, ferritins show their potential applications in medical imaging and therapeutic delivery systems. However, the recombinant human H-chain ferritin (rHF) is prone to form inclusion bodies in Escherichia coli. In our study, the cDNA of rHF was cloned into plasmid pET28a under the control of a T7 promoter. Molecular chaperones, including GroES, GroEL, and trigger factor, were coexpressed with rHF to facilitate its correct folding. The results showed that the solubility of rHF was increased more than threefold with the help of molecular chaperones. Taking advantages of its N-terminal His-tag, rHF was then purified with Ni-affinity chromatography. With a yield of 15?mg/L from bacterial culture, the purified rHF was analyzed by circular dichroism spectrometry for its secondary structure. Moreover, the rHF nanocages were characterized by transmission electron microscopy and dynamic light scattering. Our results indicate that rHF is able to self-assemble into nanocages with a narrow size distribution.  相似文献   

18.
Redox buffering by melanin and Fe(II) in Cryptococcus neoformans.   总被引:3,自引:0,他引:3       下载免费PDF全文
Melanin is a fungal extracellular redox buffer which, in principle, can neutralize antimicrobial oxidants generated by immunologic effector cells, but its source of reducing equivalents is not known. We wondered whether Fe(II) generated by the external ferric reductase of fungi might have the physiologic function of reducing fungal melanin and thereby promoting pathogenesis. We observed that exposure of a melanin film electrode to reductants decreased the open-circuit potential (OCP) and reduced the area of a cyclic voltammetric reduction wave whereas exposure to oxidants produced the opposite effects. Exposure to 10, 100, 1,000 or 10,000 microM Fe(II) decreased the OCP of melanin by 0.015, 0.038, 0.100, and 0.120 V, respectively, relative to a silver-silver chloride standard, and decreased the area of the cyclic voltammetric reduction wave by 27, 35, 50, and 83%, respectively. Moreover, exposure to Fe(II) increased the buffering capacity by 44%, while exposure to millimolar dithionite did not increase the buffering capacity. The ratio of the amount of bound iron to the amount of the incremental increase in the following oxidation wave was approximately 1.0, suggesting that bound iron participates in buffering. Light absorption by melanin suspensions was decreased 14% by treatment with Fe(II), consistent with reduction of melanin. Light absorption by suspensions of melanized Cryptococcus neoformans was decreased 1.3% by treatment with Fe(II) (P < 0.05). Cultures of C. neoformans generated between 2 and 160 microM Fe(II) in culture supernatant, depending upon the strain and the conditions [the higher values were achieved by a constitutive ferric reductase mutant in high concentrations of Fe(III)]. We infer that Fe(II) can reduce melanin under physiologic conditions; moreover, it binds to melanin and cooperatively increases redox buffering. The data support a model for physiologic redox cycling of fungal melanin, whereby electrons exported by the yeast to form extracellular Fe(II) maintain the reducing capacity of the extracellular redox buffer.  相似文献   

19.
Despite previous detection of hydroxyl radical formation during iron deposition into ferritin, no reports exist in the literature concerning how it might affect ferritin function. In the present study, hydroxyl radical formation during Fe(II) oxidation by apoferritin was found to be contingent on the "ferroxidase" activity (i.e., H subunit composition) exhibited by apoferritin. Hydroxyl radical formation was found to affect both the stoichiometry and kinetics of Fe(II) oxidation by apoferritin. The stoichiometry of Fe(II) oxidation by apoferritin in an unbuffered solution of 50 mM NaCl, pH 7.0, was approximately 3.1 Fe(II)/O(2) at all iron-to-protein ratios tested. The addition of HEPES as an alternate reactant for the hydroxyl radical resulted in a stoichiometry of about 2 Fe(II)/O(2) at all iron-to-protein ratios. HEPES functioned to protect apoferritin from oxidative modification, for its omission from reaction mixtures containing Fe(II) and apoferritin resulted in alterations to the ferritin consistent with oxidative damage. The kinetic parameters for the reaction of recombinant human H apoferritin with Fe(II) in HEPES buffer (100 mM) were: K(m) = 60 microM, k(cat) = 10 s(-1), and k(cat)/K(m) = 1.7 x 10(5) M(-1) x (-1). Collectively, these results contradict the "crystal growth model" for iron deposition into ferritin and, while our data would seem to imply that the ferroxidase activity of ferritin is adequate in facilitating Fe(II) oxidation at all stages of iron deposition into ferritin, it is important to note that these data were obtained in vitro using nonphysiologic conditions. The possibility that these findings may have physiological significance is discussed.  相似文献   

20.
Iron is the most abundant transition metal in the earth's crust. It cycles easily between ferric (oxidized; Fe(III)) and ferrous (reduced; Fe(II)) and readily forms complexes with oxygen, making this metal a central player in respiration and related redox processes. However, 'loose' iron, not within heme or iron-sulfur cluster proteins, can be destructively redox-active, causing damage to almost all cellular components, killing both cells and organisms. This may explain why iron is so carefully handled by aerobic organisms. Iron uptake from the environment is carefully limited and carried out by specialized iron transport mechanisms. One reason that iron uptake is tightly controlled is that most organisms and cells cannot efficiently excrete excess iron. When even small amounts of intracellular free iron occur, most of it is safely stored in a non-redox-active form in ferritins. Within nucleated cells, iron is constantly being recycled from aged iron-rich organelles such as mitochondria and used for construction of new organelles. Much of this recycling occurs within the lysosome, an acidic digestive organelle. Because of this, most lysosomes contain relatively large amounts of redox-active iron and are therefore unusually susceptible to oxidant-mediated destabilization or rupture. In many cell types, iron transit through the lysosomal compartment can be remarkably brisk. However, conditions adversely affecting lysosomal iron handling (or oxidant stress) can contribute to a variety of acute and chronic diseases. These considerations make normal and abnormal lysosomal handling of iron central to the understanding and, perhaps, therapy of a wide range of diseases.  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