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 218 毫秒
1.
The availability of Fe(III)‐bearing minerals for dissimilatory Fe(III) reduction was evaluated in sediments from a petroleum‐contaminated sandy aquifer near Bemidji, Minnesota (USA). First, the sediments from a contaminated area of the aquifer, in which Fe(III) reduction was the predominant terminal electron accepting process, were compared with sediments from a nearby, uncontaminated site. Data from 0.5 m HCl extraction of different size fractions of the sediments revealed that the clay size fraction contributed a significant portion of the ‘bio‐available’ Fe(III) in the background sediment and was the most depleted in ‘bio‐available’ Fe(III) in the iron‐reducing sediment. Analytical transmission electron microscopy (TEM) revealed the disappearance of thermodynamically unstable Fe(III) and Mn(IV) hydroxides (ferrihydrite and Fe vernadite), as well as a decrease in the abundance of goethite and lepidocrocite in the clay size fraction from the contaminated sediment. TEM observations and X‐ray diffraction examination did not provide strong evidence of Fe(III)‐reduction‐related changes within another potential source of ‘bio‐available’ Fe(III) in the clay size fraction – ferruginous phyllosilicates. However, further testing in the laboratory with sediments from the methanogenic portion of the aquifer that were depleted in microbially reducible Fe(III) revealed the potential for microbial reduction of Fe(III) associated with phyllosilicates. Addition of a clay size fraction from the uncontaminated sediment, as well as Fe(III)‐coated kaolin and ferruginous nontronite SWa‐1, as sources of poorly crystalline Fe(III) hydroxides and structural iron of phyllosilicates respectively, lowered steady‐state hydrogen concentrations consistent with a stimulation of Fe(III) reduction in laboratory incubations of methanogenic sediments. There was no change in hydrogen concentration when non‐ferruginous clays or no minerals were added. This demonstrated that Fe(III)‐bearing clay size minerals were essential for microbial Fe(III) reduction and suggested that both potential sources of ‘bio‐available’ Fe(III) in the clay size fraction, poorly crystalline Fe(III) hydroxides and structural Fe(III) of phyllosilicates, were important sources of electron acceptor for indigenous iron‐reducing microorganisms in this aquifer.  相似文献   

2.
Rieske dioxygenases catalyze the reductive activation of O2 for the formation of cis-dihydrodiols from unactivated aromatic compounds. It is known that O2 is activated at a mononuclear non-heme iron site utilizing electrons supplied by a nearby Rieske iron sulfur cluster. However, it is controversial whether the reactive species is an Fe(III)-(hydro)peroxo or an Fe(II)-(hydro)peroxo (or electronically equivalent species formed by breaking the O-O bond). Here it is shown that benzoate 1,2 dioxygenase oxygenase component (BZDO) prepared in a form with the Rieske cluster oxidized and the mononuclear iron in the Fe(III) state can utilize H2O2 as a source of reduced oxygen to form the correct cis-dihydrodiol product from benzoate. The reaction approaches stoichiometric yield relative to the mononuclear Fe(III) concentration, being limited to a single turnover by inefficient product release from the Fe(III)-product complex. EPR and M?ssbauer studies show that the iron remains ferric throughout this single turnover "peroxide shunt" reaction. These results strongly support Fe(III)-(hydro)peroxo (or Fe(V)-oxo-hydroxo) as the reactive species because there is no source of additional reducing equivalents to form the Fe(II)-(hydro)peroxo state. This conclusion could be further tested in the case of BZDO because the peroxide shunt occurs very slowly compared with normal turnover, allowing the reactive intermediate to be trapped for spectroscopic analysis. We attribute the slow reaction rate to a forced change in the normally strict order of the substrate binding and enzyme reduction steps that regulate the catalytic cycle. The reactive intermediate is a high-spin ferric species exhibiting an unusual negative zero field splitting and other EPR and M?ssbauer spectroscopic properties reminiscent of previously characterized side-on-bound peroxide adducts of Fe(III) model complexes. If the species in BZDO is a similar adduct, its isomer shift is most consistent with an Fe(III)-hydroperoxo reactive state.  相似文献   

3.
Ferritin molecules contain 24 polypeptide chains folded as four-helix bundles and arranged as a hollow shell capable of storing up to 4500 Fe(III) atoms. H chains contain ferroxidase centres which lie within the bundle, about 12?Å (1.2?nm) from the outside surface and 8?Å from the inner surface of the protein shell. Catalysis of Fe(II) oxidation precedes storage of Fe(III) as ferrihydrite, with the formation of μ-oxo-bridged Fe(III) dimers as intermediates. Factors influencing the movement of μ-oxo-bridged Fe(III) from the ferroxidase centre to the ferritin cavity are uncertain. Assistance by small chelators is one possibility. The aim of this investigation was to determine whether iron at the dinuclear centres of three ferritins (human H chain homopolymer, HuHF, the non-haem ferritin of Escherichia coli, EcFTN, and horse spleen ferritin, HoSF) is accessible to chelators. Forty-eight Fe(II) atoms/molecule were added to the apoferritins followed, 2?min later, by the addition of chelator (1,10-phenanthroline, 2,2-bipyridine, desferrioxamine or 3,4-dihydroxybenzaldehyde). Iron species were analysed by Mössbauer spectroscopy or visible absorbance. Competition between chelators and apoferritin for Fe(II) was also investigated. The main conclusions of the study are that: (1) dinuclear iron and iron in small iron-cores in HuHF and EcFTN is mobilisable by all four chelators; (2) the chelators penetrate the shell; (3) 3,4-dihydroxybenzaldehyde is the most efficient in mobilising Fe(III) but the least successful in competing for Fe(II); (4) Fe(III) is more readily released from EcFTN than from HuHF; (5) 2,2′-bipyridine aids the movement of Fe(III) from ferroxidase centre to core.  相似文献   

4.
The enzymatic system mainly responsible for the reduction of labile iron ions in mammalian cells is still unknown. Using isolated organelles of the rat liver, i.e. mitochondria, microsomes, nuclei and the cytosol, we here demonstrate that Fe(III), added as Fe(III)-ATP complex, is predominantly reduced by an NADH-dependent enzyme system associated with mitochondria (65% of the overall enzymatic Fe(III) reduction capacity within liver cells). Microsomes showed a significantly smaller Fe(III) reduction capacity, whereas the cytosol and nuclei hardly reduced Fe(III). Studying the mitochondrial iron reduction, this NADH-dependent process was not mediated by superoxide, ascorbic acid, or NADH itself, excluding low-molecular-weight reductants. No evidence was found for the involvement of complex I and III of the respiratory chain. Submitochondrial preparations revealed the highest specific activity reducing Fe(III) in the outer membrane fraction. In conclusion, an NADH-dependent mitochondrial enzyme system, most likely the NADH-cytochrome c reductase system, located at the outer membrane, should decisively contribute to the enzymatic reduction of labile iron within liver cells, especially under pathological conditions.  相似文献   

5.
Kwok EY  Severance S  Kosman DJ 《Biochemistry》2006,45(20):6317-6327
In high-affinity iron uptake in the yeast Saccharomyces cerevisiae, Fe(II) is oxidized to Fe(III) by the multicopper oxidase, Fet3p, and the Fe(III) produced is transported into the cell via the iron permease, Ftr1p. These two proteins are likely part of a heterodimeric or higher order complex in the yeast plasma membrane. We provide kinetic evidence that the Fet3p-produced Fe(III) is trafficked to Ftr1p for permeation by a classic metabolite channeling mechanism. We examine the (59)Fe uptake kinetics for a number of complexes containing mutant forms of both Fet3p and Ftr1p and demonstrate that a residue in one protein interacts with one in the other protein along the iron trafficking pathway as would be expected in a channeling process. We show that, as a result of some of these mutations, iron trafficking becomes sensitive to an added Fe(III) chelator that inhibits uptake in a strictly competitive manner. This inhibition is not strongly dependent on the chelator strength, however, suggesting that Fe(III) dissociation from the iron uptake complex, if it occurs, is kinetically slow relative to iron permeation. Metabolite channeling is a common feature of multifunctional enzymes. We constructed the analogous ferroxidase, permease chimera and demonstrate that it supports iron uptake with a kinetic pattern consistent with a channeling mechanism. By analogy to the Fe(III) trafficking that leads to the mineralization of the ferritin core, we propose that ferric iron channeling is a conserved feature of iron homeostasis in aerobic organisms.  相似文献   

6.
Reduction of Fe(III) to Fe(II) by Fe(III) chelate reductase is thought to be an obligatory step in iron uptake as well as the primary factor in making iron available for absorption by all plants except grasses. Fe(III) chelate reductase has also been suggested to play a more general role in the regulation of cation absorption. In order to experimentally address the importance of Fe(III) chelate reductase activity in the mineral nutrition of plants, three Arabidopsis thaliana mutants (frd1-1, frd1-2 and frd1-3), that do not show induction of Fe(III) chelate reductase activity under iron-deficient growth conditions, have been isolated and characterized. These mutants are still capable of acidifying the rhizosphere under iron-deficiency and accumulate more Zn and Mn in their shoots relative to wild-type plants regardless of iron status. frd1 mutants do not translocate radiolabeled iron to the shoots when roots are presented with a tightly chelated form of Fe(III). These results: (1) confirm that iron must be reduced before it can be transported, (2) show that Fe(III) reduction can be uncoupled from proton release, the other major iron-deficiency response, and (3) demonstrate that Fe(III) chelate reductase activity per se is not necessarily responsible for accumulation of cations previously observed in pea and tomato mutants with constitutively high levels of Fe(III) chelate reductase activity.  相似文献   

7.
Iron homeostasis is essential for maintaining the physiological requirement for iron while preventing iron overload. Cell toxicity is caused by the generation of hydroxyl-free radicals that result from redox reactions involving Fe(II). Multicopper ferroxidases regulate the oxidation of Fe(II) to Fe(III), circumventing the generation of these harmful by-products. Ceruloplasmin (Cp) is the major multicopper ferroxidase in blood; however, hephaestin (Hp), a membrane-bound Cp homolog, was recently discovered and has been implicated in the export of iron from duodenal enterocytes into blood. In the intracellular milieu, it is likely that iron exists as reduced Fe(II), yet transferrin (Tf), the plasma iron transporter, is only capable of binding oxidized Fe(III). Due to the insoluble and reactive nature of free Fe(III), the oxidation of Fe(II) upon exiting the duodenal enterocyte may require an interaction between a ferroxidase and the iron transporter. As such, it has been suggested that as a means of preventing the release of unbound Fe(III), a direct protein-protein interaction may occur between Tf and Hp during intestinal iron export. In the present study, the putative interaction between Tf and both Cp and a soluble form of recombinant human Hp was investigated. Utilizing native polyacrylamide gel electrophoresis, covalent cross-linking and surface plasmon resonance (SPR), a stable interaction between the two proteins was not detected. We conclude that a stable complex between these ferroxidases and Tf does not occur under the experimental conditions used. We suggest alternative models for loading Tf with Fe(III) during intestinal iron export.  相似文献   

8.
Yeast biomass enriched with iron could represent a new and safer solution for prevention from anaemia development. Such an iron source is less toxic and has better absorbability in organisms. The purpose of our research was the determination of the most suitable iron source in the cultivation medium for the yeast Saccharomyces cerevisiae, regarding good growth and iron accumulation in cells. Iron(III) citrate, iron(III) chloride, iron(III) nitrate and Fe-EDTA complex were used. The uptake of the chosen iron compound, Fe(III) citrate, by the yeasts Candida intermedia and Kluyveromyces marxianus was also investigated. Different growth behaviour of the three yeast strains in the presence of Fe(III) citrate was observed. The highest amounts of accumulated iron in S. cerevisiae, C. intermedia and K. marxianus biomass were about 13, 20 and 34mgFeg(-1)dry wt., respectively. To optimise the accumulation of iron in K. marxianus and to characterise iron enriched yeast biomass, further experiments are needed.  相似文献   

9.
Exchange of iron by gallium in siderophores   总被引:2,自引:0,他引:2  
T Emery 《Biochemistry》1986,25(16):4629-4633
Siderophores are iron transport compounds produced by numerous microorganisms and which strongly chelate Fe(III), but not Fe(II). Other trivalent metals, such as Al(III), Cr(III), or Ga(III), are not capable of significantly displacing iron from siderophores. However, I demonstrate here that Ga(III) can effectively displace iron under reducing conditions. With ascorbate as reductant and ferrozine as Fe(II) trapping agent, the kinetics of reductive displacement of iron by Ga(III) were followed spectroscopically by the increase of absorbance at 562 nm due to formation of the Fe(II)-ferrozine complex. No significant reduction of siderophore occurred in the absence of Ga(III). With excess Ga(III), the displacement was quantitative and very rapid. The rate of metal exchange was pseudo first order with respect to Ga(III) concentration and highly pH dependent, suggesting that siderophore ligands are displaced from the iron in a concerted mechanism by Ga(III) and protonation to expose the Fe(III) to reduction by ascorbate. Reaction rates were dependent upon the structure of the siderophore, being greatest for ferric rhodotorulic acid and slowest for ferrichrome A at pH 5.4. The pH profile for ferric rhodotorulic acid was unusual in that it showed a maximum at pH 6.5, while all other siderophores examined showed an increase in rate as pH was lowered from 7.0. The physiological significance of this reaction to the clinical use of gallium is discussed.  相似文献   

10.
Iron is essential for most living organisms. In addition, its biogeochemical cycling influences important processes in the geosphere (e.g., the mobilization or immobilization of trace elements and contaminants). The reduction of Fe(III) to Fe(II) can be catalysed microbially, particularly by metal-respiring bacteria utilizing Fe(III) as a terminal electron acceptor. Furthermore, Gram-positive fermentative iron reducers are known to reduce Fe(III) by using it as a sink for excess reducing equivalents, as a form of enhanced fermentation. Here, we use the Gram-positive fermentative bacterium Clostridium acetobutylicum as a model system due to its ability to reduce heavy metals. We investigated the reduction of soluble and solid iron during fermentation. We found that exogenous (resazurin, resorufin, anthraquinone-2,6-disulfonate) as well as endogenous (riboflavin) electron mediators enhance solid iron reduction. In addition, iron reduction buffers the pH, and elicits a shift in the carbon and electron flow to less reduced products relative to fermentation. This study underscores the role fermentative bacteria can play in iron cycling and provides insights into the metabolic profile of coupled fermentation and iron reduction with laboratory experiments and metabolic network modelling.  相似文献   

11.
The kinetics of iron binding by deferrioxamine B mesylate and the ramifications of this process upon iron-catalyzed lipid peroxidation were assessed. The relative rates of Fe(III) binding by deferrioxamine varied for the chelators tested as follows: ADP greater than AMP greater than citrate greater than histidine greater than EDTA. The addition of a fivefold molar excess of deferrioxamine to that of Fe(III) did not result in complete binding (within 10 min) for any of the Fe(III) chelates tested except ADP:Fe(III). The rates of Fe(III) binding by deferrioxamine were greater at lower pH and when the competing chelator concentration was high in relationship to iron. The relatively slow binding of Fe(III) by deferrioxamine also affected lipid peroxidation, an iron-dependent process. The addition of deferrioxamine to an ascorbate- and ADP:Fe(III)-dependent lipid peroxidation system resulted in a time-dependent inhibition or stimulation of malondialdehyde formation (i.e., lipid peroxidation), depending on the ratio of deferrioxamine to iron. Converse to Fe(III), the rates of Fe(II) binding by deferrioxamine from the chelators tested above were rapid and complete (within 1 min), and resulted in the oxidation of Fe(II) to Fe(III). Lipid peroxidation dependent on Fe(II) autoxidation was stimulated by the addition of deferrioxamine. Malondialdehyde formation in this system was inhibited by the addition of catalase, and a similar extent of lipid peroxidation was achieved by substituting hydrogen peroxide for deferrioxamine. Collectively, these results suggest that the kinetics of Fe(III) binding by deferrioxamine is a slow, variable process, whereas Fe(II) binding is considerably faster. The binding of either valence of iron by deferrioxamine may result in variable effects on iron-catalyzed processes, such as lipid peroxidation, either via slow binding of Fe(III) or the rapid binding of Fe(II) with concomitant Fe(II) oxidation.  相似文献   

12.
Cucumber plants grown in hydroponics containing 10 μM Cd(II), Ni(II) and Pb(II), and iron supplied as Fe(III) EDTA or Fe(III) citrate in identical concentrations, were investigated by total-reflection X-ray fluorescence spectrometry with special emphasis on the determination of iron accumulation and distribution within the different plant compartments (root, stem, cotyledon and leaves). The extent of Cd, Ni and Pb accumulation and distribution were also determined. Generally, iron and heavy-metal contaminant accumulation was higher when Fe(III) citrate was used. The accumulation of nickel and lead was higher by about 20% and 100%, respectively, if the iron supply was Fe(III) citrate. The accumulation of Cd was similar. In the case of Fe(III) citrate, the total amounts of Fe taken up were similar in the control and heavy-metal-treated plants (27-31 μmol/plant). Further, the amounts of iron transported from the root towards the shoot of the control, lead- and nickel-contaminated plants were independent of the iron(III) form. Although Fe mobility could be characterized as being low, its distribution within the shoot was not significantly affected by the heavy metals investigated.  相似文献   

13.
Polynuclear iron complexes of Fe(III) and phosphate occur in seawater and soils and in cells where the iron core of ferritin, the iron storage protein, contains up to 4500 Fe atoms in a complex with an average composition of (FeO.OH)8FeO.OPO3H2. Although phosphate influences the size of the ferritin core and thus the availability of stored iron, little is known about the nature of the Fe(III)-phosphate interaction. In the present study, Fe-phosphate interactions were analyzed in stable complexes of Fe(III).ATP which, in the polynuclear iron form, had phosphate at interior sites. Such Fe(III).ATP complexes are important not only as models but also because they may play a role in intracellular iron transport and in iron toxicity; the complexes were studied by extended x-ray absorption fine structure, EPR, NMR spectroscopy, and measurement of proton release. Mononuclear iron complexes exhibiting a g' = 4.3 EPR signal were formed at Fe:ATP ratios less than or equal to 1:3, and polynuclear iron complexes (Fe greater than or equal to 250, EPR silent at g' = 4.3) were formed at an Fe:ATP ratio of 4:1. No NMR signals due to ATP were observed when Fe was in excess (Fe:ATP = 4:1). Extended x-ray absorption fine structure analysis of the polynuclear Fe(III).ATP complex was able to distinguish an Fe-P distance at 3.27 A in addition to the octahedral O at 1.95 A and 4-5 Fe atoms at 3.36 A. The Fe-O and Fe-Fe distances are the same as in ferritin, and the Fe-P distance is analogous to that in another metal-ATP complex. An observable Fe-P environment in such a large polynuclear iron cluster as the Fe(III).ATP (4:1) complex indicates that the phosphate is distributed throughout rather than merely on the surface, in contrast to earlier models of chelate-stabilized iron clusters. Complexes of Fe(III) and ATP similar to those described here may form in vivo either as normal components of intracellular iron metabolism or during iron excess where the consequent alteration of free nucleotide triphosphate pools could contribute to the observed toxicity of iron.  相似文献   

14.
Iron acquisition by iron‐limited cyanobacteria is typically considered to be mediated mainly by siderophores, iron‐chelating molecules released by iron‐limited cyanobacteria into the environment. In this set of experiments, iron uptake by iron‐limited cells of the cyanobacterium Anabaena flos‐aquae (L.) Bory was investigated in cells resuspended in siderophore‐free medium. Removal of siderophores decreased iron‐uptake rates by ~60% compared to siderophore‐replete conditions; however, substantial rates of iron uptake remained. In the absence of siderophores, Fe(III) uptake was much more rapid from a weaker synthetic chelator [N‐(2‐hydroxyethyl)ethylenediamine‐N,N′,N′‐triacetic acid (HEDTA); log Kcond = 28.64 for Fe(III)HEDTA(OH)?] than from a very strong chelator [N,N′‐bis(2‐hydroxybenzyl)‐ethylenediamine‐N,N′‐diacetic acid (HBED); log Kcond = 31.40 for Fe(III)HBED?], and increasing chelator:Fe(III) ratios decreased the Fe(III)‐uptake rate; these results were evident in both short‐term (4 h; absence of siderophores) and long‐term (116 h; presence of siderophores) experiments. However, free (nonchelated) Fe(III) provided the most rapid iron uptake in siderophore‐free conditions. The results of the short‐term experiments are consistent with an Fe(III)‐binding/uptake mechanism associated with the cyanobacterial outer membrane that operates independently of extracellular siderophores. Iron uptake was inhibited by temperature‐shock treatments of the cells and by metabolically compromising the cells with diphenyleneiodonium; this finding indicates that the process is dependent on active metabolism to operate and is not simply a passive Fe(III)‐binding mechanism. Overall, these results point to an important, siderophore‐independent iron‐acquisition mechanism by iron‐limited cyanobacterial cells.  相似文献   

15.
Bovine heart microsomes have been found to contain a non-heme iron protein which serves as an electron acceptor for NADPH-cytochrome P-450 reductase and therefore stimulates NADPH oxidation. This protein, tentatively referred to as Microsomal Iron Protein (MIP), has been extracted with Triton N-101 and purified by ion exchange chromatography on CM- and DEAE-celluloses and gel filtration on Sepharose 6B. MIP is an Mr = 66,000 monomer with 17 atoms of Fe(III)/molecule. Incubation with dithionite removes iron from MIP and abolishes the stimulation of NADPH oxidation, but subsequent incubation with nitrilotriacetic-Fe(III) reincorporates iron and restores the stimulation of NADPH oxidation. Oxygen is the ultimate electron acceptor. In the presence of oxygen, the enzymatic reduction of MIP Fe(III) is followed by the reoxidation of Fe(II) at the expense of oxygen, generating superoxide anion and regenerating MIP Fe(III) for the continuous oxidation of NADPH. In the absence of oxygen, electron transfer from the reductase to MIP Fe(III) causes the release of Fe(II), which limits the ability of MIP to serve as an electron acceptor and stimulate NADPH oxidation. The--NH2-terminal of MIP has been sequenced, and no homology has been found with the sequence of other iron storage or transport proteins such as ferritin or transferrin.  相似文献   

16.
Mining-impacted sediments of Lake Coeur d'Alene, Idaho, contain more than 10% metals on a dry weight basis, approximately 80% of which is iron. Since iron (hydr)oxides adsorb toxic, ore-associated elements, such as arsenic, iron (hydr)oxide reduction may in part control the mobility and bioavailability of these elements. Geochemical and microbiological data were collected to examine the ecological role of dissimilatory Fe(III)-reducing bacteria in this habitat. The concentration of mild-acid-extractable Fe(II) increased with sediment depth up to 50 g kg(-1), suggesting that iron reduction has occurred recently. The maximum concentrations of dissolved Fe(II) in interstitial water (41 mg liter(-1)) occurred 10 to 15 cm beneath the sediment-water interface, suggesting that sulfidogenesis may not be the predominant terminal electron-accepting process in this environment and that dissolved Fe(II) arises from biological reductive dissolution of iron (hydr)oxides. The concentration of sedimentary magnetite (Fe(3)O(4)), a common product of bacterial Fe(III) hydroxide reduction, was as much as 15.5 g kg(-1). Most-probable-number enrichment cultures revealed that the mean density of Fe(III)-reducing bacteria was 8.3 x 10(5) cells g (dry weight) of sediment(-1). Two new strains of dissimilatory Fe(III)-reducing bacteria were isolated from surface sediments. Collectively, the results of this study support the hypothesis that dissimilatory reduction of iron has been and continues to be an important biogeochemical process in the environment examined.  相似文献   

17.
The distribution of Fe(III), its availability for microbial reduction, and factors controlling Fe(III) availability were investigated in sediments from a freshwater site in the Potomac River Estuary. Fe(III) reduction in sediments incubated under anaerobic conditions and depth profiles of oxalate-extractable Fe(III) indicated that Fe(III) reduction was limited to depths of 4 cm or less, with the most intense Fe(III) reduction in the top 1 cm. In incubations of the upper 4 cm of the sediments, Fe(III) reduction was as important as methane production as a pathway for anaerobic electron flow because of the high rates of Fe(III) reduction in the 0- to 0.5-cm interval. Most of the oxalate-extractable Fe(III) in the sediments was not reduced and persisted to a depth of at least 20 cm. The incomplete reduction was not the result of a lack of suitable electron donors. The oxalate-extractable Fe(III) that was preserved in the sediments was considered to be in a form other than amorphous Fe(III) oxyhydroxide, since synthetic amorphous Fe(III) oxyhydroxide, amorphous Fe(III) oxyhydroxide adsorbed onto clay, and amorphous Fe(III) oxyhydroxide saturated with adsorbed phosphate or fulvic acids were all readily reduced. Fe3O4 and the mixed Fe(III)-Fe(II) compound(s) that were produced during the reduction of amorphous Fe(III) oxyhydroxide in an enrichment culture were oxalate extractable but were not reduced, suggesting that mixed Fe(III)-Fe(II) compounds might account for the persistence of oxalate-extractable Fe(III) in the sediments. The availability of microbially reducible Fe(III) in surficial sediments demonstrates that microbial Fe(III) reduction can be important to organic matter decomposition and iron geochemistry. However, the overall extent of microbial Fe(III) reduction is governed by the inability of microorganisms to reduce most of the Fe(III) in the sediment.  相似文献   

18.
The catalytic activity of phosvitin in Fe(II) oxidation and the addition of iron to transferrin were studied under various conditions. It was concluded that the Fe(II) oxidized by phosvitin would bind to apotransferrin, although an appreciable fraction of Fe(III) remained bound to phosvitin. Fe(III) also migrated from phosvitin to apotransferrin. This reaction was first-order with respect to Fe(III)-phosvitin concentration with a half-time (t1/2) of 10 min, and a first-order rate constant, k=0.069min-1, in 700 muM-phosphate buffer, pH 7.2, at 30 degrees C. The catalysis of the oxidation of Fe(III) by phosvitin was proportional to O2 concentration, and is quite different from the relative O2 independence of Fe(II) oxidation as catalysed by ferroxidase. A scheme for the mobilization and transfer of iron in the chicken, including the role of ferroxidase, phosyitin and transferrin, is presented.  相似文献   

19.
The potential of Desulfitobacterium frappieri strain G2, which was isolated from subsurface smectite bedding, to participate in iron redox reactions was investigated. Strain G2 can use poorly crystalline Fe(III) oxide, soluble forms of Fe(III) and Fe(III) in the structure of phyllosilicate minerals as electron acceptors. It can also oxidize Fe(II)-NTA or Fe(II) in the structure of phyllosilicate minerals with nitrate as the electron acceptor. These results suggest for the first time that strains of Desulfitobacterium frappieri may play an important role in iron cycling in sedimentary environments.  相似文献   

20.
FeCl3 or Fe(III) that attached to chelating ligands such as citrate or nitrilotriacetic acid (NTA) at a molar ratio of 1:1 had a toxic effect on PMN. Uptake of radiolabeled Staphylococcus aureus by PMN, preincubated for 2 hr at 37 degrees C in a medium containing Fe(III)-citrate or Fe(III)-NTA, was significantly lower than that of control PMN preincubated without excess iron (p less than 0.002). However, at a 1:2 molar ratio of Fe(III) to citrate or NTA, the iron was not toxic. In contrast, the iron-liganding molecules transferrin and deferoxamine protected the PMN against the noxious effect of iron at concentrations just high enough to sequester all the iron. Fe(III) increased the generation of luminol chemiluminescence by stimulated PMN, whereas the oxygen consumption of the cells was not altered in the presence of Fe(III); this suggests a catalytic effect of iron on the production by PMN of oxygen metabolites at some step beyond the formation of superoxide. No effect of iron was observed when the incubation was performed at 4 degrees C, nor when an oxygen-radical scavenger such as thiourea, mannitol, or catalase was present in the incubation medium. Also, Fe(III) had much less effect on the phagocytic function of PMN of a patient with chronic granulomatous disease. The results indicate that the Fe(III)-induced defect in the phagocytic capacity of PMN depends on the nature and the concentration of the ligand attached to the iron ion, and also suggest that the noxious effect of iron on the PMN function is a result of its ability to catalyze the generation of toxic oxygen species by these cells.  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