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 15 毫秒
1.
Alcoholic solutions of FeCl(2) and Fe(II)(Hmmi)(2)Cl(2) (Hmmi=1-methyl-2-mercaptoimidazole) induce calf thymus DNA condensation from aqueous solutions buffered at pH 7.4. A 1:1 Fe(II)-(DNA monomer) stoichiometry is assumed. The (57)Fe M?ssbauer hyperfine parameters suggest an octahedral coordination environment, severely distorted, in both Fe(II)-(DNA monomer) and [Fe(II)(Hmmi)(2)]-(DNA monomer) condensates. The dynamic properties of iron nuclei in freeze-dried samples were investigated by means of variable temperature (57)Fe M?ssbauer spectroscopy. Mean square displacements, (T), were calculated, such as the effective vibrating mass and the M?ssbauer lattice temperature of the solids. increases linearly with the temperature in the whole temperature range explored; the absolute values are typical for lattice or solid-state vibrations. Very similar values for the effective vibrating masses were extracted, suggesting comparable covalency of the bonding interaction between the metal atom and its ligands, while the M?ssbauer lattice temperatures show a softening of the lattice for [Fe(II)(Hmmi)(2)]-(DNA monomer) with respect to Fe(II)-(DNA monomer) condensate.  相似文献   

2.
The involvement of the Fe cations in autoxidation in cells and tissues is well documented. DNA is a major target in such reaction, and can chelate Fe cation in many ways. The present study was designed to examine the interaction of calf-thymus DNA with Fe(II) and Fe(III), in aqueous solution at pH 6.5 with cation/DNA (P) (P = phosphate) molar ratios (r) of 1:160 to 1:2. Capillary electrophoresis and Fourier transform infrared (FTIR) difference spectroscopic methods were used to determine the cation binding site, the binding constant, helix stability and DNA conformation in Fe-DNA complexes. Structural analysis showed that at low cation concentration (r = 1/80 and 1/40), Fe(II) binds DNA through guanine N-7 and the backbone PO(2) group with specific binding constants of K(G) = 5.40 x 10(4) M(1) and K(P) = 2.40 x 10(4) M(1). At higher cation content, Fe(II) bindings to adenine N-7 and thymine O-2 are included. The Fe(III) cation shows stronger interaction with DNA bases and the backbone phosphate group. At low cation concentration (r = 1:80), Fe(III) binds mainly to the backbone phosphate group, while at higher metal ion content, cation binding to both guanine N-7 atom and the backbone phosphate group is prevailing with specific binding constants of K(G) = 1.36 x 10(5) M(-1) and K(P) = 5.50 x 10(4) M(-1). At r = 1:10, Fe(II) binding causes a minor helix destabilization, whereas Fe(III) induces DNA condensation. No major DNA conformational changes occurred upon iron complexation and DNA remains in the B-family structure.  相似文献   

3.
The kinetics and mechanism of Fe(III) reduction to Fe(II) were studied in pure batch cultures of Pseudomonas sp. 200. The rate of iron reduction has been mechanistically related to aqueous phase iron speciation. In the absence of microbial activity the iron reduction rate was negligible. Initial rates of microbial iron reduction were accelerated more than 20-fold by the addition of equimolar quantities of nitrilotriacetic acid (NTA) to media initially containing 1.86 x 10(-3)M total Fe(III). Numerical techniques were utilized to quantify relationships between the observed rate of Fe(II) production and the calculated (equilibrium) aqueous phase speciation. These results indicate that soluble ferric iron species are not equivalent in terms of their susceptibility to bacterial (dissimilative) iron reduction. The concentration of Fe(NTA)(OH)(2) (2-) correlated strongly with observed iron reduction rates. Ferrous iron species appeared to inhibit the reduction process.  相似文献   

4.
Jin S  Kurtz DM  Liu ZJ  Rose J  Wang BC 《Biochemistry》2004,43(11):3204-3213
The X-ray crystal structure of recombinant Desulfovibrio vulgaris rubrerythrin (Rbr) that was subjected to metal constitution first with zinc and then iron, yielding ZnS(4)Rbr, is reported. A [Zn(SCys)(4)] site with no iron and a diiron site with no appreciable zinc in ZnS(4)Rbr were confirmed by analysis of the anomalous scattering data. Partial reduction of the diiron site occurred during the synchrotron X-ray irradiation at 95 K, resulting in two different diiron site structures in the ZnS(4)Rbr crystal. These two structures can be classified as containing mixed-valent Fe1(III)(mu-OH(-))(mu-GluCO(2)(-))(2)Fe2(II) and Fe1(II)(mu-GluCO(2)(-))(2)Fe2(III)-OH(-) cores. The data do not show any evidence for alternative positions of the protein or solvent ligands. The iron and ligand positions of the solvent-bridged site are close to those of the diferric site in all-iron Rbr. The diiron site with only the two carboxylato bridges differs by an approximately 2 A shift in the position of Fe1, which changes from six- to four-coordination. The Fe1- - -Fe2 distance (3.6 A) in this latter site is significantly longer than that of the site with the additional solvent bridge (3.4 A) but significantly shorter than that previously reported for the diferrous site (4.0 A) in all-iron Rbr. The apparent redox-induced movement of Fe1 at 95 K in the ZnS(4)Rbr crystal implies an extremely low activation barrier, which is consistent with the rapid (approximately 30 s(-1)) room temperature turnover of the all-iron Rbr during its catalysis of two-electron reduction of hydrogen peroxide. ZnS(4)Rbr does not show peroxidase activity, presumably because the [Zn(SCys)(4)] site, unlike the [Fe(SCys)(4)] site, cannot mediate electron transfer to the diiron site. One or both of the diiron site structures in the cryoreduced ZnS(4)Rbr crystal are likely to represent that (those) of transient mixed-valent diiron site(s) that must occur upon return of the diferric to the diferrous oxidation level during peroxidase turnover.  相似文献   

5.
The reactions of RO(2)* radicals with Fe(H(2)O)(6)(2+) were studied, R[double bond]H; CH(3); CH(2)COOH; CH(2)CN; CH(2)C(CH(3))(2)OH; CH(2)OH; CHCl(2)/CCl(3). All these processes involve the following reactions: Fe(H(2)O)(6)(2+)+RO(2)*<==>(H(2)O)(5)Fe(III)[bond]OOR(2+) K(1) approximately 250 M(-1); (H(2)O)(5)Fe(III)[bond]OOR(2+)+H(3)O(+)/H(2)O-->Fe(H(2)O)(6)(3+)+ROOH+H(2)O/OH(-); (H(2)O)(5)Fe(III)[bond]OOR(2+)+2Fe(H(2)O)(6)(2+)-->3Fe(H(2)O)(6)(3+)+ROH; 2 RO(2)*-->Products; RO(2)*+(H(2)O)(5)Fe(III)[bond]OOR(2+)-->Fe(H(2)O)(6)(2+)+products. The values of k(1) and k(3) [reaction is clearly not an elementary reaction] approach the ligand exchange rate of Fe(H(2)O)(6)(2+), i.e. these reactions follow an inner sphere mechanism and the rate determining step is the ligand exchange step. The rate of reaction is several orders of magnitude faster than that of the Fenton reaction. Surprisingly enough the K(1) values are nearly independent of the redox potential of the radical and are considerably higher than calculated from the relevant redox potentials. These results indicate that the ROO(-) ligands considerably stabilise the Fe(III) complex, this stabilisation is smaller for radicals with electron withdrawing groups which raise the redox potential of the radical but decrease the basicity of the ROO(-) ligands, two effects which seem to nearly cancel each other. Finally, the results clearly indicate that reaction (5) is relatively fast and affects the nature of the final products. The contribution of these reactions to oxidation processes involving 'Fenton-like' processes is discussed.  相似文献   

6.
The electron transfer protein rubredoxin from Clostridium pasteurianum contains an Fe(S-Cys)(4) active site. Mutant proteins C9G, C9A, C42G and C42A, in which cysteine ligands are replaced by non-ligating Gly or Ala residues, have been expressed in Escherichia coli. The C42A protein expresses with a Fe(III)(2)S(2) cluster in place. In contrast, the other proteins are isolated in colourless forms, although a Fe(III)(2)S(2) cluster may be assembled in the C42G protein via incubation with Fe(III)and sulfide. The four mutant proteins were isolated as stable mononuclear Hg(II)forms which were converted to unstable mononuclear Fe(III)preparations that contain both holo and apo protein. The Fe(III)systems were characterized by metal analysis and mass spectrometry and by electronic, electron paramagnetic resonance, X-ray absorption and resonance Raman spectroscopies. The dominant Fe(III) form in the C9A preparation is a Fe(S-Cys)(3)(OH) centre, similar to that observed previously in the C6S mutant protein. Related centres are present in the proteins NifU and IscU responsible for assembly and repair of iron-sulfur clusters in both prokaryotic and eukaryotic cells. In addition to Fe(S-Cys)(3)(OH) centres, the C9G, C42G and C42A preparations contain a second four-coordinate Fe(III)form in which a ligand appears to be supplied by the protein chain.  相似文献   

7.
The DNA binding of iron(II) mixed-ligand complexes containing 1,10-phenanthroline(phen) and 4,7-diphenyl-1,10-phenanthroline(dip), [Fe(phen)(3)](2+), [Fe(phen)(2)(dip)](2+) and [Fe(phen)(dip)(2)](2+) has been characterized by spectrophotometric titration and melting temperature measurements. The salt concentration dependence of the binding constant has allowed us to dissect the DNA-binding constant and free energy change of each iron(II) complex into the nonelectrostatic and polyelectrolyte contributions. A comparison of the nonelectrostatic components in the binding free energy changes among iron(II) complexes has made it possible to rigorously evaluate the contribution of the ligand substituents to the DNA-binding event. The peripheral substitution of phen by two phenyl groups increases the nonelectrostatic binding constant of the iron(II) complex more than 20 times, which is equivalent to approximately 7.5 kJ mol(-1) of more favorable contribution to the DNA binding. In general, the iron(II) complexes studied have higher affinity towards the more facile A-T sequence than the G-C sequence. This preferential binding may be attributed to the steric effect induced by the ancillary part of the ligands in the course of DNA binding. The binding of disubstituted iron(II) complex to DNA is quite strong as reflected in the modest increase in the denaturation temperature (T(m)) of double helical DNA upon the interaction with the iron(II) complex.  相似文献   

8.
15N T(1), T(2) and (1)H-(15)N NOE were measured for the thermophilic Fe(7)S(8) protein from Bacillus schlegelii and for the Fe(4)S(4) HiPIP protein from Chromatium vinosum, which is a mesophilic protein. The investigation was performed at 276, 300, and 330 K at 11.7 T for the former, whereas only the 298 K data at 14.1 T for the latter were acquired. The data were analyzed with the model-free protocol after correcting the measured parameters for the effect of paramagnetism, because both proteins are paramagnetic. Both thermophilic and mesophilic proteins are quite rigid, with an average value of the generalized order parameter S2at room temperature of 0.92 and 0.94 for Fe(7)S(8) and Fe(4)S(4) proteins, respectively. The analyzed nitrogens for the Fe(7)S(8) protein showed a significant decrease in S2with increasing temperature, and at the highest temperature >70% of the residues had an internal correlation time. This research shows that subnanosecond rigidity is not related to thermostability and provides an estimate of the effect of increasing temperature on this time scale.  相似文献   

9.
The reaction of FeCl(2) or FeCl(3) with sodium ethanethiolate (SEt) in N-methylformamide (NMF) has been reevaluated to rectify a previous Fe(II) oxidation artifact. On titrating Fe(II) with EtS(-) concentrations up to 12 mol Eq, new features in the UV/vis spectrum (epsilon(344)=(3.1+/-0.2)x10(3) M(-1) cm(-1); epsilon(486)=(4.5+/-0.1)x10(2) M(-1) cm(-1)) indicated that the first observable step was the formation of a single complex different from the known tetrahedral tetrathiolate, [Fe(SEt)(4)](2-) . As the EtS(-) concentration increased past 12.5 mol Eq the UV/vis spectrum gradually transformed to that of [Fe(SEt)(4)](2-) (lambda(max)=314 nm). A Hill-formalism fit to the titration data of the initially formed complex indicated cooperative ligation by three ethanethiolate ions, with K(coop)=(1.7+/-0.1)x10(3) M(-3) and Hill "n"=2.4+/-0.1 (r=0.997). The 3:1 EtS(-)-Fe(II) complex is proposed to be [Fe(2)(SEt)(6)](2-). Titration of Fe(III) with EtS(-) showed direct cooperative formation of [Fe(SEt)(4)](-) [epsilon(340)=(3.4+/-0.5)x10(3) M(-1) cm(-1)] with a Hill-formalism K(coop)=(4.3+/-0.1)x10(2) M(-4) and a Hill coefficient "n"=3.7+/-0.2 (r=0.996). Further ligation past [Fe(SEt)(4)](-) was observed at EtS(-) concentrations above 35 mol Eq. The Fe(III) Hill constants are at variance with our previous report. However, the UV/vis spectrum of Fe(III) in NMF solution was found to change systematically over time, consistent with a slow progressive deprotonation of [Fe(nmf)](3+). The observed time-to-time differences in the equilibrium chemistry of Fe(III) with ethanethiolate in NMF thus reflect variation in the microscopic solution composition of FeCl(3) in alkaline NMF solvent. These results are related to the chemistry of nitrogenase FeMo cofactor in alkaline NMF solution.  相似文献   

10.
This is the first time to described the dissimilatory Fe(III) reducing characteristics of Shewanella cinica D14T. The effects of O2, light, temperature and pH on dissimilatory Fe(III) reduction were examined. The results suggested that the rate of Fe(III) reduction decreased with increasing Fe(III) concentration. Fe(III) reduction was partially inhibited by the presence of either O2 or light. The optimum temperature for Fe(III) reduction is 37 degrees C. At pH 6.0-10.0, strain D14T can reduce Fe(III). The soluble Fe(III) is more easy to be reduced than the insoluble one. Results of protein denaturants SDS and OGP suggest that the Fe(III) reduction activity of S. cinica is mostly localized to the soluble outer membrane fraction. The azo dye decolorization and Fe(III) reduction in strain D14T were enhanced in the presence of Fe(III) and dye.  相似文献   

11.
In a previous study (Minotti, G., and Ikeda-Saito, M. (1991) J. Biol. Chem. 266, 20011-20017) we demonstrated the existence of a M(r) 66,000 microsomal iron protein (MIP) which stimulates NADPH oxidation by shunting electrons from NADPH-cytochrome P-450 reducase to its bound Fe(III). In the present study, purified MIP was depleted of iron and the apoMIP was examined for its ability to incorporate Fe(III) upon an incubation with Fe(II). It was found that apoMIP had an oxygen-dependent ferroxidase activity coupled with the incorporation of Fe(III). The reconstituted MIP exhibited a Fe(III) content and an NADPH oxidation activity similar to those of native MIP. However, the reconstitution of MIP from apoMIP and Fe(II) had to be performed in the presence of detergents to prevent the formation of protein aggregates and the oxidative incorporation of an iron which could not react with NADPH-cytochrome P-450 reductase. This redox inactive iron was probably bound nonspecifically to artifactual sites formed by the protein aggregates.  相似文献   

12.
Biological systems usually contain cysteine, glutathione or other sulfur-containing biomolecules. These S-nucleophiles were found to affect drastically the [Fe(4)(mu(3)-S)(3)(NO)(7)](-) photolysis pathway generating products completely different from that of the neat cluster, which produces Fe(II) and NO and S(2-). The effect is interpreted in terms of formation of a pseudo-cubane adduct, [Fe(4)(mu(3)-S)(3)(mu(3)-SR)(NO)(7)](2-), whose existence in equilibrium with the parent complex has no detectable influence on the spectral properties, whereas shifts the redox potential and induces photoconversion leading to the Fe(III) species and N(2)O. Characteristic bond lengths, bond angles and atomic Mulliken charges were calculated using semi-empirical quantum chemical methods for the RBS anion and a series of pseudo-cubane complexes with S-donor or N-donor ligands. The results justify the hypothesis of the adduct formation and show that only in case of S-ligands the higher contribution of the Fe(III)-NO(-) components in adduct than in RBS is observed, which on excitation can undergo heterolytic cleavage yielding Fe(III) and NO(-), converted rapidly into N(2)O. These results are crucial in understanding the physiological activity of RBS. Fe(III) formation can be detected only when the S-ligand enables formation of a stable Fe(III) compound; the effect was recorded in the presence of sulfide, thioglycolate, 2-mercaptopropionate, mercaptosuccinate, penicillamine, 2,3-dimercaptosuccinate, 2,3-dimercaptopropanol, and thiocyanate. For all these S-ligands the Fe(III) photoproducts were identified and characterised. In the case of other thiolates, their excess results in fast reduction of Fe(III) to Fe(II), whereas N(2)O can be still detected. Quantum yields of Fe(III) formation in the presence of the S-ligands are considerably higher than that of the Fe(II) photoproduction from neat [Fe(4)(mu(3)-S)(3)(NO)(7)](-).  相似文献   

13.
Complexes of the type [Al(HL)(OH)Cl(2)], [M(HL)(OH)(2)Cl] and [M'(HL)(L')(OH)Cl], where HL = 5-iodouracil; HL' = histidine; M = Cr(III), Fe(III) and M' = Al(III), Cr(III), Fe(III), were synthesized and characterized. The complexes are polymeric showing high decomposition points and are insoluble in water and common organic solvents. The mu(eff) values, electronic spectral bands and ESR spectra suggest a polymeric 6-coordinate spin-free octahedral stereochemistry for the Cr(III) and Fe(III) complexes. 5-Iodouracil acts as a monodentate ligand coordinating to the metal ion through the O atom of C((4)) = O while histidine through the O atom of -COO(- ) and the N atom of -NH(2) group. In vivo antitumour effect of 5-iodouracil and its complexes was examined on C(3)H /He mice against P815 murine mastocytoma. As evident from their T/C values, Cr(III) and Fe(III) complexes display significant and higher antitumour activity compared to the 5-iodouracil ligand. The in vitro results of the complexes on the same cells indicate that Cr(III) and Fe(III) complexes show higher inhibition on (3)H-thymidine and (3)H-uridine incorporation in DNA and RNA replication, respectively, at a dose of 5 microg/mL.  相似文献   

14.
In this report we study the effect of Fe(III) on lipid peroxidation induced by Fe(II)citrate in mitochondrial membranes, as assessed by the production of thiobarbituric acid-reactive substances and antimycin A-insensitive oxygen uptake. The presence of Fe(III) stimulates initiation of lipid peroxidation when low citrate:Fe(II) ratios are used ( 4:1). For a citrate:total iron ratio of 1:1 the maximal stimulation of lipid peroxidation by Fe(III) was observed when the Fe(II):Fe(III) ratio was in the range of 1:1 to 1:2. The lag phase that accompanies oxygen uptake was greatly diminished by increasing concentrations of Fe(III) when the citrate:total iron ratio was 1:1, but not when this ratio was higher. It is concluded that the increase of lipid peroxidation by Fe(III) is observed only when low citrate:Fe(II) ratios were used. Similar results were obtained using ATP as a ligand of iron. Monitoring the rate of spontaneous Fe(II) oxidation by measuring oxygen uptake in buffered medium, in the absence of mitochondria, Fe(III)-stimulated oxygen consumption was observed only when a low citrate:Fe(II) ratio was used. This result suggests that Fe(III) may facilitate the initiation and/or propagation of lipid peroxidation by increasing the rate of Fe(II)citrate-generated reactive oxygen species.  相似文献   

15.
Li W  Liu N  Cai LL  Jiang JL  Chen JM 《Bioresource technology》2011,102(3):3049-3054
Biological reduction of Fe(III) to Fe(II) is a key step in nitrogen oxide (NO(x)) removal by the integrated chemical absorption-biological reduction process. NO(x) removal efficiency strongly depends on the concentration of Fe(II) in the scrubbing liquid. In this study, a newly isolated strain, Enterococcus sp. FR-3, was used to reduce Fe(III) chelated with citrate to Fe(II). Strain FR-3 reduced citrate-chelated Fe(III) with an efficiency of up to 86.9% and an average reduction rate of 0.21 mM h(-1). SO(4)(2-) was not inhibitory whereas NO(2)(-) and SO(3)(2-) inhibited cell growth and thus affected Fe(III) reduction. Models based on the Logistic equation were used to describe the relationship between growth and Fe(III) reduction. These findings provide some useful data for Fe(III) reduction, scrubber solution regeneration and NO(x) removal process design.  相似文献   

16.
A severe inactivation is obtained upon the addition of H(2)O(2) to bovine heart F(1)ATPase samples containing Fe(III) in the nucleotide-independent site, and Fe(II) in the ATP-dependent site. EPR spectra at 4.9 K of these samples indicate that H(2)O(2) produces the complete oxidation of Fe(II) to Fe(III) and the concomitant appearance of two protein-derived radical species. The two signals (g = 2.036 and g = 2.007) display a different temperature dependence and saturation behavior. The relaxation properties of the radical at g = 2.036 suggest magnetic interaction with one of the two iron centers. Such events are not observed when H(2)O(2) is added either to native F(1)ATPase containing a high amount of Fe(II) and low amount of Fe(III) or to F(1)ATPase deprived of endogenous Fe and subsequently loaded with only Fe(III) in both sites. It is hypothesized that in F(1)ATPase samples containing both Fe(III) and Fe(II), intramolecular long-range electron transfer may occur from Fe(II) to a high oxidation state species of Fe formed in the nucleotide-independent site upon oxidation of Fe(III) by H(2)O(2).  相似文献   

17.
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   

18.
The banded iron formations (BIF) of Brazil are composed of silica and Fe(III) oxide lamina, and are largely covered by a rock cap of BIF fragments in a goethite matrix (canga). Despite both BIF and canga being highly resistant to erosion and poorly soluble, >3,000 iron ore caves (IOCs) have formed at their interface. Fe(III) reducing microorganisms (FeRM) can reduce the Fe(III) oxides present in the BIF and canga, which could account for the observed speleogenesis. Here, we show that IOCs contain a variety of microbial taxa with member species capable of dissimilatory Fe(III) reduction, including the Chloroflexi, Acidobacteria and the Alpha- Beta- and Gammaproteobacteria; however, Fe(III) reducing enrichment cultures from IOCs indicate the predominance of Firmicutes and Enterobacteriaceae, despite varying the carbon/electron donor, Fe(III) type, and pH. We used model-based inference to evaluate multiple candidate hypotheses that accounted for the variation in medium chemistry and culture composition. Model selection indicated that none of the tested variables account for the dominance of the Firmicutes in these cultures. The addition of H2 to the headspace of the enrichment cultures enhanced Fe(III) reduction, while addition of N2 resulted in diminished Fe(III) reduction, indicating that these Enterobacteriaceae and Firmicutes were reducing Fe(III) during fermentative growth. These results suggest that fermentative reduction of Fe(III) may play a larger role in iron-rich environments than expected. Our findings also demonstrate that FeRM are present within the IOCs, and that their reductive dissolution of Fe(III) oxides, combined with mass transport of solubilized Fe(II) by groundwater, could contribute to IOC formation.  相似文献   

19.
The equilibrium and solution structural properties of the iron(III) and copper(II) complexes of an asymmetric salen-like ligand (N,N'-bis(2-hydroxybenzyl)-2,3-diamino-propionic acid, H(3)bhbdpa) bearing a pendant carboxylate group were characterized in aqueous solution by potentiometric, pH-dependent electron paramagnetic resonance (EPR) and UV-Vis (UV-Visible) measurements. In the equimolar systems the pentadentate ligand forms very stable, differently protonated mononuclear complexes with both metal ions. In the presence of iron(III) {NH, PhO(-), COO(-)}, {2NH, 2PhO(-), COO(-)} and {2NH, 2PhO(-), COO(-), OH(-)} coordinated complexes are dominant. The EPR titrations reflected the presence of microscopic complex formation pathways, leading to the formation of binding isomers in case of Cu(H(2)bhbdpa)(+), Cu(Hbhbdpa) and Cu(bhbdpa)(-). The {2NH, 2PhO(-)+COO(-)/H(2)O} coordinated Cu(bhbdpa) is the only species between pH 6-11. At twofold excess of metal ion dinuclear complexes were detected with both iron(III) and copper(II). In presence of iron(III) a mu-carboxylato-mu-hydroxo-bridged dinuclear complex (Fe(2)(bhbdpa)(OH)(3)) is formed from Fe(H(2)bhbdpa)(2+) through overlapping proton release processes, providing one of the rare examples for the stabilization of an endogenous carboxylate bridged diiron core in aqueous solution. The complex Cu(2)(bhbdpa)(+) detected in the presence of copper(II) is a paramagnetic (S=1) species with relatively weakly coupled metal ions.  相似文献   

20.
A neutrophilic Fe(II)-oxidizing bacterium was isolated from the redox zone of a low-salinity spring in Krasnodar krai (Russia), at the FeS-Fe(OH)(3) interface deposited at the sediment surface. The cells of strain Sp-1 were short, thin motile vibrioids with one polar flagellum dividing by binary fission. The optimal values and ranges for pH and temperature were pH 6.2 (5.5-8) and 35?°C (5-45?°C), respectively. The organism was a facultative anaerobe. Strain Sp-1 was capable of organotrophic, lithoheterotrophic and mixotrophic growth with Fe(II) as an electron donor. The denitrification chain was 'disrupted'. Oxidation of Fe(II) was coupled to reduction of NO3 - to NO2 - or of N(2) O to N(2) , as well as under microaerobic conditions, with O(2) as an electron acceptor. The DNA G+C content was 64.2?mol%. According to the results of phylogenetic analysis, the strain was 10.6-12% remote from the closest relatives, members of the genera Sneathiella, Inquilinus, Oceanibaculum and Phaeospirillum within the Alphaproteobacteria. Based on its morphological, physiological and taxonomic characteristics, together with the results of phylogenetic analysis, strain Sp-1 is described as a member of a new genus Ferrovibrio gen. nov., with the type species Ferrovibrio denitrificans sp. nov. and the type strain Sp-1(T) (=?LMG 25817(T) =?VKM B-2673(T) ).  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