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.
Ferrous iron [Fe(II)] reductively transforms heavy metals in contaminated groundwater, and the bacterial reduction of indigenous ferric iron [Fe(III)] to Fe(II) has been proposed as a means of establishing redox reactive barriers in the subsurface. The reduction of Fe(III) to Fe(II) can be accomplished by stimulation of indigenous dissimilatory metal-reducing bacteria (DMRB) or injection of DMRB into the subsurface. The microbially produced Fe(II) can chemically react with contaminants such as Cr(VI) to form insoluble Cr(III) precipitates. The DMRB Shewanella algae BrY reduced surface-associated Fe(III) to Fe(II), which in batch and column experiments chemically reduced highly soluble Cr(VI) to insoluble Cr(III). Once the chemical Cr(VI) reduction capacity of the Fe(II)/Fe(III) couple in the experimental systems was exhausted, the addition of S. algae BrY allowed for the repeated reduction of Fe(III) to Fe(II), which again reduced Cr(VI) to Cr(III). The research presented herein indicates that a biological process using DMRB allows the establishment of a biogeochemical cycle that facilitates chromium precipitation. Such a system could provide a means for establishing and maintaining remedial redox reactive zones in Fe(III)-bearing subsurface environments.  相似文献   

2.
Lakes formed because of coal mining are characterized by low pH and high concentrations of Fe(II) and sulfate. The anoxic sediment is often separated into an upper acidic zone (pH 3; zone I) with large amounts of reactive iron and a deeper slightly acidic zone (pH 5.5; zone III) with smaller amounts of iron. In this study, the impact of pH on the Fe(III)-reducing activities in both of these sediment zones was investigated, and molecular analyses that elucidated the sediment microbial diversity were performed. Fe(II) was formed in zone I and III sediment microcosms at rates that were approximately 710 and 895 nmol cm−3 day−1, respectively. A shift to pH 5.3 conditions increased Fe(II) formation in zone I by a factor of 2. A shift to pH 3 conditions inhibited Fe(II) formation in zone III. Clone libraries revealed that the majority of the clones from both zones (approximately 44%) belonged to the Acidobacteria phylum. Since moderately acidophilic Acidobacteria species have the ability to oxidize Fe(II) and since Acidobacterium capsulatum reduced Fe oxides at pHs ranging from 2 to 5, this group appeared to be involved in the cycling of iron. PCR products specific for species related to Acidiphilium revealed that there were higher numbers of phylotypes related to cultured Acidiphilium or Acidisphaera species in zone III than in zone I. From the PCR products obtained for bioleaching-associated bacteria, only one phylotype with a level of similarity to Acidithiobacillus ferrooxidans of 99% was obtained. Using primer sets specific for Geobacteraceae, PCR products were obtained in higher DNA dilutions from zone III than from zone I. Phylogenetic analysis of clone libraries obtained from Fe(III)-reducing enrichment cultures grown at pH 5.5 revealed that the majority of clones were closely related to members of the Betaproteobacteria, primarily species of Thiomonas. Our results demonstrated that the upper acidic sediment was inhabited by acidophiles or moderate acidophiles which can also reduce Fe(III) under slightly acidic conditions. The majority of Fe(III) reducers inhabiting the slightly acidic sediment had only minor capacities to be active under acidic conditions.  相似文献   

3.
《Inorganica chimica acta》1986,123(3):127-131
The modified method of preparation of water soluble metalloporphyrins is presented. As a ligand 5,10,15,20-tetra-p(N-ethyl-N,N-dimethyl)anilinporphyrinium disulphate was used. The structure of the obtained metalloporphyrins for the following metal cations: Mg(II), Zn(II), Cd(II), Ag(II), Ru(Il), Rh(II), Ni(II), Fe(III), Mn(III), Co(III) and Sn(IV), was confirmed by electron, IR spectra and elemental analyses.  相似文献   

4.
Ferrous iron [Fe(II)] reductively transforms heavy metals in contaminated groundwater, and the bacterial reduction of indigenous ferric iron [Fe(III)] to Fe(II) has been proposed as a means of establishing redox reactive barriers in the subsurface. The reduction of Fe(III) to Fe(II) can be accomplished by stimulation of indigenous dissimilatory metal-reducing bacteria (DMRB) or injection of DMRB into the subsurface. The microbially produced Fe(II) can chemically react with contaminants such as Cr(VI) to form insoluble Cr(III) precipitates. The DMRB Shewanella algae BrY reduced surface-associated Fe(III) to Fe(II), which in batch and column experiments chemically reduced highly soluble Cr(VI) to insoluble Cr(III). Once the chemical Cr(VI) reduction capacity of the Fe(II)/Fe(III) couple in the experimental systems was exhausted, the addition of S. algae BrY allowed for the repeated reduction of Fe(III) to Fe(II), which again reduced Cr(VI) to Cr(III). The research presented herein indicates that a biological process using DMRB allows the establishment of a biogeochemical cycle that facilitates chromium precipitation. Such a system could provide a means for establishing and maintaining remedial redox reactive zones in Fe(III)-bearing subsurface environments.  相似文献   

5.
The high-sulfur coal desulfurization process completed by A. ferrooxidans LY01 cells domesticated with either ferrous iron [Fe(II)] or pyrite (FeS2) was investigated in this article. The desulfurization rate for 13 d was as high as 67.8% for the LY01 cells domesticated with pyrite but was only 45.6% for the LY01 cells domesticated with Fe(II). Bacterial adsorption experiments indicated that the bacterial adsorption quantity onto the pyrite particles was similar to the desulfurization efficiency. FTIR analysis showed that chemical composition of the two cell types was similar, but the LY01 cells domesticated with pyrite had higher levels of hydrophobic aromatic R-O groups than cells domesticated with Fe(II). The amount of extracellular polymeric substances (EPS) from the pyrite-domesticated LY01 cells was 1820 μg C/1010 cells, which was five times more than the amount of EPS in the Fe(II)-domesticated cells; the EPS readily bound Fe(III) with a maximum binding capacity of 0.21 mg Fe(III) per mg C EPS. Strains of pyrite-domesticated LY01 with a high amount of Fe(III) in their EPS possess greater oxidation activity than Fe(II)-domesticated strains with fewer Fe(III). These experiments showed the importance of the substrate-specific differences in the oxidative activity of A. ferrooxidans LY01. In addition, this study provides theoretical guidance for the future optimization of the biodesulfurization process.  相似文献   

6.
The observed biological differences in safety and efficacy of intravenous (IV) iron formulations are attributable to physicochemical differences. In addition to differences in carbohydrate shell, polarographic signatures due to ferric iron [Fe(III)] and ferrous iron [Fe(II)] differ among IV iron formulations. Intravenous iron contains Fe(II) and releases labile iron in the circulation. Fe(II) generates toxic free radicals and reactive oxygen species and binds to bacterial siderophores and other in vivo sequestering agents. To evaluate whether differences in Fe(II) content may account for some observed biological differences between IV iron formulations, samples from multiple lots of various IV iron formulations were dissolved in 12 M concentrated HCl to dissociate and release all iron and then diluted with water to achieve 0.1 M HCl concentration. Fe(II) was then directly measured using ferrozine reagent and ultraviolet spectroscopy at 562 nm. Total iron content was measured by adding an excess of ascorbic acid to reduce Fe(III) to Fe(II), and Fe(II) was then measured by ferrozine assay. The Fe(II) concentration as a proportion of total iron content [Fe(III) + Fe(II)] in different lots of IV iron formulations was as follows: iron gluconate, 1.4 and 1.8 %; ferumoxytol, 0.26 %; ferric carboxymaltose, 1.4 %; iron dextran, 0.8 %; and iron sucrose, 10.2, 15.5, and 11.0 % (average, 12.2 %). The average Fe(II) content in iron sucrose was, therefore, ≥7.5-fold higher than in the other IV iron formulations. Further studies are needed to investigate the relationship between Fe(II) content and increased risk of oxidative stress and infections with iron sucrose.  相似文献   

7.
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   

8.
The involvement of iron in the biogenesis of the cyanide-insensitive respiration in the yeast Saccharomycopsis lipolytica has been established on the following basis: (1) endogenous metal chelation by either benzyl- or salicylhydroxamic acid, EDTA or nitrilotriacetate prevented the biogenesis of the cyanide-insensitive respiratory pathway in S. lipolytica. (2) Addition of Fe(III) during the biogenesis increased both the rate of the appearance of the alternative respiratory pathway and its extent. Neither Fe(II), nor Co(II), Cu(II), Al(III), La(III), Mn(II) or Mg(II) could substitute for Fe(III). (3) The biogenesis of the alternative respiratory pathway could be dissociated into two steps: (a) a first one, slow, cycloheximide-sensitive, temperature-dependent, iron-independent, leading to cells still fully cyanide-sensitive, presumably involving the de novo biosynthesis of an inactive protein moiety and (b) a second step, fast, iron-dependent, temperature-independent, cycloheximide-insensitive, leading to cells with a cyanide-insensitive respiration, presumably the activation by iron of the inactive precursor.  相似文献   

9.
Iron transport across polarized intestinal epithelium was studied by using Caco-2 cells grown in bicameral chambers. When cells were grown under conditions of low, normal, or high iron concentration not only was the iron content of the cells markedly altered but the low iron cells exhibited a nearly 2-fold increase in transepithelial electrical resistance (TEER). 59Fe uptake from the apical surface into cells and transport into the basal chamber was affected both by the valency of the iron and the iron status of the cells. Uptake from 59Fe(II)-ascorbate was about 600 pmol 59Fe/h per mg protein, increased about 2-fold in low iron cells, and was about 13-200-fold greater than uptakes from 59Fe(III) chelated to nitrilotriacetic acid, BSA, or citrate. Transport into the basal chamber from 59Fe(II)-ascorbate was 3.7 +/- 1.7 pmol/h per cm2 for Fe-deficient cells vs. 0.72 +/- 0.1 pmol/h per cm2 for normal-Fe cells and from 59Fe(III)-BSA 1.1 +/- 0.2 pmol/h per cm2 vs. 0.3 +/- 0.03 pmol/h per cm2 for deficient vs. normal iron cells, respectively. The greater transport of iron both from Fe(II) and in iron deficient cells supports the use of the Caco-2 cells as a model for iron transport.  相似文献   

10.
Iron Uptake by Symbiosomes from Soybean Root Nodules   总被引:26,自引:2,他引:24       下载免费PDF全文
To identify possible iron sources for bacteroids in planta, soybean (Glycine max L. Merr.) symbiosomes (consisting of the bacteroid-containing peribacteroid space enclosed by the peribacteroid membrane [PBM]) and bacteroids were assayed for the ability to transport iron supplied as various ferric [Fe(III)]-chelates. Iron presented as a number of Fe(III)-chelates was transported at much higher rates across the PBM than across the bacteroid membranes, suggesting the presence of an iron storage pool in the peribacteroid space. Pulse-chase experiments confirmed the presence of such an iron storage pool. Because the PBM is derived from the plant plasma membrane, we reasoned that it may possess a ferric-chelate reductase activity similar to that present in plant plasma membrane. We detected ferric-chelate reductase activity associated with the PBM and suggest that reduction of Fe(III) to ferrous [Fe(II)] plays a role in the movement of iron into soybean symbiosomes.  相似文献   

11.
A new series of Schiff base complexes [Fe(III), VO(II), Pd(II), Cu(II), and Ni(II)] has been developed. The ligand possesses bulky t-pentyl groups at the 3- and 5-positions. The iron (III) complex is obtained in monomeric form with a square-pyramidal configuration while the copper complex is with square-planar configuration.  相似文献   

12.
Light-induced interaction of Fe(II) cations with the donor side of Mn-depleted photosystem II (PS II(–Mn)) results in the binding of iron cations and blocking of the high-affinity (HAZ) Mn-binding site. The pH dependence of the blocking was measured using the diphenylcarbazide/2,6-dichlorophenolindophenol test. The curve of the pH dependence is bell-shaped with pK 1 = 5.8 and pK 2 = 8.0. The pH dependence of the O2-evolution mediated by PS II membranes is also bellshaped (pK 2 = 7.6). The pH dependence of the process of electron donation from exogenous donors in PS II(–Mn) was studied to determine the location of the alkaline pH sensitive site of the electron transport chain. The data of the study showed that the decrease in the iron cation binding efficiency at pH > 7.0 during blocking was determined by the donor side of the PS II(–Mn). Mössbauer spectroscopy revealed that incubation of PS II(–Mn) membranes in a buffer solution containing 57Fe(II) + 57Fe(III) was accompanied by binding only Fe(III) cations. The pH dependence of the nonspecific Fe(III) cation binding is also described by the same bell-shaped curve with pK 2 = 8.1. The treatment of the PS II(–Mn) membranes with the histidine modifier diethylpyrocarbonate resulted in an increase in the iron binding strength at alkaline pH. It is suggested that blocking efficiency at alkaline pH is determined by competition between OH and histidine ligand for Fe(III). Because the high-affinity Mn-binding site contains no histidine residue, this fact can be regarded as evidence that histidine is located at another (other than high-affinity) Fe(III) binding site. In other words, this means that the blockage of the high-affinity Mn-binding site is determined by at least two iron cations. We assume that inactivation of oxygen-evolving complex and inhibition of photoactivation in the alkaline pH region are also determined by competition between OH and a histidine residue involved in coordination of manganese cation outside the high-affinity site.  相似文献   

13.
《Inorganica chimica acta》1987,138(3):205-214
Electronic and 57Fe Mössbauer spectra are reported for two new water-soluble porphyrinato iron(III) complexes. Equilibrium constants for μ-oxo bishaem formation are calculated assuming two protons are released.Comparisons are made of the data with other porphyrinato iron(III) systems and it is shown that, in the absence of well-defined fifth ligands, the mononuclear species in acidic solution probably contain two axial water ligands. The μ-oxo bishaems do not contain water or hydroxide coordinated to iron but may hold water by hydrogen-bonding to the oxygen bridge or possibly by aquation of the porphyrin ligands.μ-Oxo bridge formation is controlled by the acid strength of the water coordinated to the iron in the mononuclear species, low pKa values assisting oxo-bridge formation. Such low pKa values are assisted by electron-attracting substituents on the porphyrin periphery. It is noted that this same property assists the stabilisation of iron(II) complexes. Steric inhibition of oxo-bridge formation requires large substituents, unsubstituted phenyl groups being apparently not large enough.  相似文献   

14.
Adsorption of heavy metals and radionuclides (HMR) onto iron and manganese oxides has long been recognized as an important reaction for the immobilization of these compounds. However, in environments containing elevated concentrations of these HMR the adsorptive capacity of the iron and manganese oxides may well be exceeded, and the HMR can migrate as soluble compounds in aqueous systems. Here we demonstrate the potential of a bioremediative strategy for HMR stabilization in reducing environments based on the recently described anaerobic nitrate-dependent Fe(II) oxidation by Dechlorosoma species. Bio-oxidation of 10 mM Fe(II) and precipitation of Fe(III) oxides by these organisms resulted in rapid adsorption and removal of 55 μM uranium and 81 μM cobalt from solution. The adsorptive capacity of the biogenic Fe(III) oxides was lower than that of abiotically produced Fe(III) oxides (100 μM for both metals), which may have been a result of steric hindrance by the microbial cells on the iron oxide surfaces. The binding capacity of the biogenic oxides for different heavy metals was indirectly correlated to the atomic radius of the bound element. X-ray absorption spectroscopy indicated that the uranium was bound to the biogenically produced Fe(III) oxides as U(VI) and that the U(VI) formed bidentate and tridentate inner-sphere complexes with the Fe(III) oxide surfaces. Dechlorosoma suillum oxidation was specific for Fe(II), and the organism did not enzymatically oxidize U(IV) or Co(II). Small amounts (less than 2.5 μM) of Cr(III) were reoxidized by D. suillum; however, this appeared to be inversely dependent on the initial concentration of the Cr(III). The results of this study demonstrate the potential of this novel approach for stabilization and immobilization of HMR in the environment.  相似文献   

15.
《BBA》1987,893(2):138-148
Four representative inhibitors of Photosystem II (PS II) QA to QB electron transfer were shown to bind, at high concentrations, to PS II reaction centers having the acceptor-side non-heme iron in the Fe(III) state. Three of the inhibitors studied, DCMU, o-phenanthroline and dinoseb, modified the EPR spectrum of the Fe(III) relative to that obtained by ferricyanide oxidation in the absence of inhibitor. o-Phenanthroline gave particularly axial symmetry, while DCMU and dinoseb gave more rhombic configurations. The herbicide inhibitor, atrazine and its analogue, terbutryn, had no effect. The dissociation constants for inhibitor binding to reaction centers in the Fe(III) state were measured directly and also estimated from shifts in the midpoint potential of the Fe(III)/Fe(II) couple and were shown to increase by factors of approx. 100, approx. 10 and 10–15 for DCMU (pH 7.5), atrazine (pH 7.0) and o-phenanthroline (pH 7.0), respectively, upon oxidation of the iron. Atrazine and o-phenanthroline, which induce the smallest changes in the midpoint potential of the Fe(III)/Fe(II) couple, were shown to inhibit light-induced oxidation of the Fe(II) by phenyl-p-BQ, described in the preceding paper (Petrouleas, V. and Diner, B.A. (1987) Biochim. Biophys. Acta 893, 126–137). The extent of inhibition was much greater than would be predicted from a simple shift in the midpoint potential for Fe(III)/Fe(II) and we conclude that phenyl-p-BQ and the other quinones, which show light-induced oxidation, act through the QB binding site. It is also argued that reduction and oxidation of the iron by ferro- and ferricyanide, respectively, occur through this site. The effects of these inhibitors and of various quinones on the Fe(III) environment are discussed with reference to the known contact points between the protein and o-phenanthroline and terbutryn in the QB binding pocket of Rhodopseudomonas viridis reaction centers (Michel, H., Epp, O. and Deisenhofer, J. (1986) EMBO J. 5, 2445–2451). The Fe(III) EPR spectrum is thus a new and sensitive probe of the contact points at which molecules bind to the QB binding site.  相似文献   

16.
Desulfovibrio gigas desulforedoxin (Dx) consists of two identical peptides, each containing one [Fe-4S] center per monomer. Variants with different iron and zinc metal compositions arise when desulforedoxin is produced recombinantly from Escherichia coli. The three forms of the protein, the two homodimers [Fe(III)/Fe(III)]Dx and [Zn(II)/Zn(II)]Dx, and the heterodimer [Fe(III)/Zn(II)]Dx, can be separated by ion exchange chromatography on the basis of their charge differences. Once separated, the desulforedoxins containing iron can be reduced with added dithionite. For NMR studies, different protein samples were prepared labeled with (15)N or (15)N + (13)C. Spectral assignments were determined for [Fe(II)/Fe(II)]Dx and [Fe(II)/Zn(II)]Dx from 3D (15)N TOCSY-HSQC and NOESY-HSQC data, and compared with those reported previously for [Zn(II)/Zn(II)]Dx. Assignments for the (13)C(alpha) shifts were obtained from an HNCA experiment. Comparison of (1)H-(15)N HSQC spectra of [Zn(II)/Zn(II)]Dx, [Fe(II)/Fe(II)]Dx and [Fe(II)/Zn(II)]Dx revealed that the pseudocontact shifts in [Fe(II)/Zn(II)]Dx can be decomposed into inter- and intramonomer components, which, when summed, accurately predict the observed pseudocontact shifts observed for [Fe(II)/Fe(II)]Dx. The degree of linearity observed in the pseudocontact shifts for residues >/=8.5 A from the metal center indicates that the replacement of Fe(II) by Zn(II) produces little or no change in the structure of Dx. The results suggest a general strategy for the analysis of NMR spectra of homo-oligomeric proteins in which a paramagnetic center introduced into a single subunit is used to break the magnetic symmetry and make it possible to obtain distance constraints (both pseudocontact and NOE) between subunits.  相似文献   

17.
Shewanella oneidensis MR-1 is a facultatively anaerobic bacterium capable of using soluble and insoluble forms of manganese [Mn(III/IV)] and iron [Fe(III)] as terminal electron acceptors during anaerobic respiration. To assess the structural association of two outer membrane-associated c-type decaheme cytochromes (i.e., OmcA [SO1779] and MtrC [SO1778]) and their ability to reduce soluble Fe(III)-nitrilotriacetic acid (NTA), we expressed these proteins with a C-terminal tag in wild-type S. oneidensis and a mutant deficient in these genes (i.e., Delta omcA mtrC). Endogenous MtrC copurified with tagged OmcA in wild-type Shewanella, suggesting a direct association. To further evaluate their possible interaction, both proteins were purified to near homogeneity following the independent expression of OmcA and MtrC in the Delta omcA mtrC mutant. Each purified cytochrome was confirmed to contain 10 hemes and exhibited Fe(III)-NTA reductase activity. To measure binding, MtrC was labeled with the multiuse affinity probe 4',5'-bis(1,3,2-dithioarsolan-2-yl)fluorescein (1,2-ethanedithiol)2, which specifically associates with a tetracysteine motif engineered at the C terminus of MtrC. Upon titration with OmcA, there was a marked increase in fluorescence polarization indicating the formation of a high-affinity protein complex (Kd < 500 nM) between MtrC and OmcA whose binding was sensitive to changes in ionic strength. Following association, the OmcA-MtrC complex was observed to have enhanced Fe(III)-NTA reductase specific activity relative to either protein alone, demonstrating that OmcA and MtrC can interact directly with each other to form a stable complex that is consistent with their role in the electron transport pathway of S. oneidensis MR-1.  相似文献   

18.
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   

19.
R C Blake  K J White  E A Shute 《Biochemistry》1991,30(39):9443-9449
Rusticyanin is a soluble blue copper protein found in abundance in the periplasmic space of Thiobacillus ferrooxidans, an acidophilic bacterium capable of growing chemolithotrophically on soluble ferrous sulfate. The one-electron-transfer reactions between soluble iron and purified rusticyanin were studied by stopped-flow spectrophotometry in acidic solutions containing each of 14 different anions. The second-order rate constants for both the Fe(II)-dependent reduction and the Fe(III)-dependent oxidation of the rusticyanin varied as a function of the identity of the principal anion in solution. Analogous electron-transfer reactions between soluble iron and bis(dipicolinato)cobaltate(III) or bis(dipicolinato)ferrate(II) were studied by stopped-flow spectrophotometry under solution conditions identical with those of the rusticyanin experiments. Similar anion-dependent reactivity patterns were obtained with soluble iron whether the other reaction partner was rusticyanin or either of the two organometallic complexes. The Marcus theory of outer-sphere electron transfer reactions was applied to this set of kinetic data to demonstrate that the rusticyanin may possess at least two electron-transfer pathways for liganded iron, one where the pattern of electron-transfer reactivity is controlled largely by protein-independent activation parameters and one where the protein exhibits an anion-dependent kinetic specificity. The exact role of rusticyanin in the iron-dependent respiratory electron transport chain of T. ferrooxidans remains unclear.  相似文献   

20.
Novel thermophilic crenarchaea have been observed in Fe(III) oxide microbial mats of Yellowstone National Park (YNP); however, no definitive work has identified specific microorganisms responsible for the oxidation of Fe(II). The objectives of the current study were to isolate and characterize an Fe(II)-oxidizing member of the Sulfolobales observed in previous 16S rRNA gene surveys and to determine the abundance and distribution of close relatives of this organism in acidic geothermal springs containing high concentrations of dissolved Fe(II). Here we report the isolation and characterization of the novel, Fe(II)-oxidizing, thermophilic, acidophilic organism Metallosphaera sp. strain MK1 obtained from a well-characterized acid-sulfate-chloride geothermal spring in Norris Geyser Basin, YNP. Full-length 16S rRNA gene sequence analysis revealed that strain MK1 exhibits only 94.9 to 96.1% sequence similarity to other known Metallosphaera spp. and less than 89.1% similarity to known Sulfolobus spp. Strain MK1 is a facultative chemolithoautotroph with an optimum pH range of 2.0 to 3.0 and an optimum temperature range of 65 to 75°C. Strain MK1 grows optimally on pyrite or Fe(II) sorbed onto ferrihydrite, exhibiting doubling times between 10 and 11 h under aerobic conditions (65°C). The distribution and relative abundance of MK1-like 16S rRNA gene sequences in 14 acidic geothermal springs containing Fe(III) oxide microbial mats were evaluated. Highly related MK1-like 16S rRNA gene sequences (>99% sequence similarity) were consistently observed in Fe(III) oxide mats at temperatures ranging from 55 to 80°C. Quantitative PCR using Metallosphaera-specific primers confirmed that organisms highly similar to strain MK1 comprised up to 40% of the total archaeal community at selected sites. The broad distribution of highly related MK1-like 16S rRNA gene sequences in acidic Fe(III) oxide microbial mats is consistent with the observed characteristics and growth optima of Metallosphaera-like strain MK1 and emphasizes the importance of this newly described taxon in Fe(II) chemolithotrophy in acidic high-temperature environments of YNP.  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