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 40 毫秒
1.
An alternative hypothesis for the origin of the banded iron formations and the synthesis of prebiotic molecules is presented here. I show the importance of considering water near its supercritical point and at alkaline pH. It is based on the chemical equation for the anoxic oxidation of ferrous iron into ferric iron at high-subcritical conditions of water and high pH, that I extract from E-pH diagrams drawn for corrosion purposes (Geophysical Research Abstracts Vol 15, EGU2013–22 Bassez 2013, Orig Life Evol Biosph 45(1):5-13, Bassez 2015, Procedia Earth Planet Sci 17, 492-495, Bassez 2017a, Orig Life Evol Biosph 47:453-480, Bassez 2017b). The sudden change in solubility of silica, SiO2, at the critical point of water is also considered. It is shown that under these temperatures and pressures, ferric oxides and ferric silicates can form in anoxic terrains. No FeII oxidation by UV light, neither by oxygen is needed to explain the minerals of the Banded Iron Formations. The intervention of any kind of microorganisms, either sulfate-reducing, or FeII-oxidizing or O2-producing, is not required. The chemical equation for the anoxic oxidation of ferrous iron is applied to the hydrolyses of fayalite, Fe2SiO4 and ferrosilite, FeSiO3. It is shown that the BIF minerals of the Hamersley Group, Western Australia, and the Transvaal Supergroup, South Africa, are those of fayalite and ferrosilite hydrolyses and carbonations. The dissolution of crustal fayalite and ferrosilite during water-rock interaction needs to occur at T&P just below the critical point of water and in a rising water which is undersaturated in SiO2. Minerals of BIFs which can then be ejected at the surface from venting arcs are ferric oxide hydroxides, hematite, FeIII-greenalite, siderite. The greenalite dehydrated product minnesotaite forms when rising water becomes supersaturated in SiO2, as also riebeckite and stilpnomelane. Long lengths of siderite without ferric oxides neither ferric silicates can occur since the exothermic siderite formation is not so much dependent in T&P. It is also shown that the H2 which is released during hydrolysis/oxidation of fayalite/ferrosilite can lead to components of life, such as macromolecules of amino acids which are synthesized from mixtures of (CO, N2, H2O) in Sabatier-Senderens/Fischer-Tropsch & Haber-Bosch reactions or microwave or gamma-ray excitation reactions. I propose that such geobiotropic synthesis may occur inside fluid inclusions of BIFs, in the silica chert, hematite, FeIII-greenalite or siderite. Therefore, the combination of high-subcritical conditions of water, high solubility of SiO2 at these T&P values, formation of CO also at these T&P, high pH and anoxic water, leads to the formation of ferric minerals and prebiotic molecules in the process of geobiotropy.  相似文献   

2.
In the presence of methanoate as electron donor, Shewanella putrefaciens, a Gram‐negative, facultative anaerobe, is able to transform lepidocrocite (γ‐FeOOH) to secondary Fe (II–III) minerals such as carbonated green rust (GR1) and magnetite. When bacterial cells were added to a γ‐FeOOH suspension, aggregates were produced consisting of both bacteria and γ‐FeOOH particles. Recently, we showed that the production of secondary minerals (GR1 vs. magnetite) was dependent on bacterial cell density and not only on iron reduction rates. Thus, γ‐FeOOH and S. putrefaciens aggregation pattern was suggested as the main mechanism driving mineralization. In this study, lepidocrocite bioreduction experiments, in the presence of anthraquinone disulfonate, were conducted by varying the [cell]/[lepidocrocite] ratio in order to determine whether different types of aggregate are formed, which may facilitate precipitation of GR1 as opposed to magnetite. Confocal laser scanning microscopy was used to analyze the relative cell surface area and lepidocrocite concentration within the aggregates and captured images were characterized by statistical methods for spatial data (i.e. variograms). These results suggest that the [cell]/[lepidocrocite] ratio influenced both the aggregate structure and the nature of the secondary iron mineral formed. Subsequently, a [cell]/[lepidocrocite] ratio above 1 × 107 cells mmol?1 leads to densely packed aggregates and to the formation of GR1. Below this ratio, looser aggregates are formed and magnetite was systematically produced. The data presented in this study bring us closer to a more comprehensive understanding of the parameters governing the formation of minerals in dense bacterial suspensions and suggest that screening mineral–bacteria aggregate structure is critical to understanding (bio)mineralization pathways.  相似文献   

3.
The stability of (all-E)-β-carotene toward dietary iron was studied in a mildly acidic (pH 4) micellar solution as a simple model of the postprandial gastric conditions. The oxidation was initiated by free iron (FeII, FeIII) or by heme iron (metmyoglobin, MbFeIII). FeII and metmyoglobin were much more efficient than FeIII at initiating β-carotene oxidation. Whatever the initiator, hydrogen peroxide did not accumulate. Moreover, β-carotene markedly inhibited the conversion of FeII into FeIII. β-Carotene oxidation induced by FeII or MbFeIII was maximal with 5–10 eq FeII or 0.05–0.1 eq MbFeIII and was inhibited at higher iron concentrations, especially with FeII. UPLC/DAD/MS and GC/MS analyses revealed a complex distribution of β-carotene-derived products including Z-isomers, epoxides, and cleavage products of various chain lengths. Finally, the mechanism of iron-induced β-carotene oxidation is discussed. Altogether, our results suggest that dietary iron, especially free (loosely bound) FeII and heme iron, may efficiently induce β-carotene autoxidation within the upper digestive tract, thereby limiting its supply to tissues (bioavailability) and consequently its biological activity.  相似文献   

4.
An oxalate-bridged binuclear iron(III) complex, [(acac)2Fe(μ-ox)Fe(acac)2], (acac=acetylacetonate anion and ox2−=oxalate anion) was prepared. The complex crystallized as two types of crystals under different conditions: one had 1,2-dichloroethane as a solvent molecule of crystallization 2, the other did not 1. Both compounds have been characterized by X-ray crystallography, infrared spectroscopy, and thermogravimetric analysis. Compound 1 has also been characterized by UV-Vis and 1H NMR spectroscopies, mass spectrometry, and electrochemistry. In both crystals, each iron(III) is coordinated in an octahedral arrangement by the oxygen atoms of an oxalate-bridging ligand and four oxygen atoms belonging to peripheral acac ligands in an octahedral arrangement. The intermetallic distance of Fe?Fe is 5.4368(9) Å in 1 and 5.438(2) Å in 2. Two iron(III) ions in each crystal are bridged by the oxalate and both lie in the oxalate-plane. The results of thermal analyses imply that the thermal stability of 2 is lower than that of 1. Cyclic voltammograms of 1 in acetonitrile and dichloromethane at low temperature showed two consecutive, quasi-Nernstian, one-electron reduction steps corresponding to the reduction of FeIII-FeIII to FeIII-FeII followed by the reduction of FeIII-FeII to FeII-FeII. The electrochemical comproportionation constants (Kc) of the equilibrium (FeIII-FeIII) + (FeII-FeII) ? 2(FeIII-FeII) are 108.9 in acetonitrile medium and 108.5 in dichloromethane, respectively. The considerably large Kc values indicate that the main factor contributing to the stabilization of the FeIII-FeII mixed-valence state is electronic delocalization through the oxalate-bridge.  相似文献   

5.
Several anaerobic bacteria isolated from the sediments of Contrary Creek, an iron-rich environment, produced magnetite when cultured in combinations but not when cultured alone in synthetic iron oxyhydroxide medium. When glucose was added as a carbon source, the pH of the medium decreased (to 5.5) and no magnetite was formed. When the same growth medium without glucose was used, the pH increased (to 8.5) and magnetite was formed. In both cases, Fe2+ was released into the growth medium. Geochemical equilibrium equations with Eh and pH as master variables were solved for the concentrations of iron and inorganic carbon that were observed in the system. Magnetite was predicted to be the dominant iron oxide formed at high pHs, while free Fe2+ or siderite were the dominant forms of iron expected at low pHs. Thus, magnetite formation occurs because of microbial alteration of the local Eh and pH conditions, along with concurrent reduction of ferric iron (direct biological reduction or abiological oxidation-reduction reactions).  相似文献   

6.
Pyridine-2,6-bis(monothiocarboxylic acid), also known as pyridine-2,6-dithiocarboxylic acid (pdtc), is a unique and powerful metal chelator produced by Pseudomonas stutzeri and Pseudomonas putida. The actual physiological roles of pdtc in these pseudomonads are not known with certainty, though it is likely that the compound acts as a siderophore, an antibiotic, or both. The stability constant of FeIII(pdtc)2 2- was determined in previous work to be 1033.36. Here we determined that the stability constant of FeII(pdtc)2 2- is 1012. We determined this stability constant through potentiometric and spectrophotometric measurements of a ligand-ligand competition study using 2,6-pyridine dicarboxylic acid as the competitor for iron. Comparing the stability constant for FeII(pdtc)2 2- to the constant for FeIII(pdtc)2 2- shows that the stability constant of FeII(pdtc)2 2- is approximately 21 orders of magnitude smaller. This represents a very significant decrease in the binding strength of pdtc toward iron. Thus, if the host cell produces pdtc as a siderophore for sequestering Fe(III), it is likely that a second metabolite or a membrane protein of the host cell is used for reduction of the chelated iron at or near the cell membrane in order to facilitate its release from pdtc for cellular use.  相似文献   

7.
Nicotianamine (NA) occurs in all plants and chelates metal cations, including FeII, but reportedly not FeIII. However, a comparison of the FeII and ZnII affinity constants of NA and various FeIII-chelating aminocarboxylates suggested that NA should chelate FeIII. High-voltage electrophoresis of the FeNA complex formed in the presence of FeIII showed that the complex had a net charge of 0, consistent with the hexadentate chelation of FeIII. Measurement of the affinity constant for FeIII yielded a value of 1020.6, which is greater than that for the association of NA with FeII (1012.8). However, capillary electrophoresis showed that in the presence of FeII and FeIII, NA preferentially chelates FeII, indicating that the FeIINA complex is kinetically stable under aerobic conditions. Furthermore, Fe complexes of NA are relatively poor Fenton reagents, as measured by their ability to mediate H2O2-dependent oxidation of deoxyribose. This suggests that NA will have an important role in scavenging Fe and protecting the cell from oxidative damage. The pH dependence of metal ion chelation by NA and a typical phytosiderophore, 2′-deoxymugineic acid, indicated that although both have the ability to chelate Fe, when both are present, 2′-deoxymugineic acid dominates the chelation process at acidic pH values, whereas NA dominates at alkaline pH values. The consequences for the role of NA in the long-distance transport of metals in the xylem and phloem are discussed.  相似文献   

8.
Kovács K  Kuzmann E  Tatár E  Vértes A  Fodor F 《Planta》2009,229(2):271-278
Distinct chemical species of iron were investigated by Mössbauer spectroscopy during iron uptake into cucumber roots grown in unbuffered nutrient solution with or without 57Fe-citrate. Mössbauer spectra of iron deficient roots supplied with 10–500 μM 57Fe-citrate for 30–180 min and 24 h and iron-sufficient ones, were recorded. The roots were analysed for Fe concentration and Fe reductase activity. The Mössbauer parameters in the case of iron-sufficient roots revealed high-spin iron(III) components suggesting the presence of FeIII-carboxylate complexes, hydrous ferric oxides and sulfate–hydroxide containing species. No FeII was detected in these roots. However, iron-deficient roots supplied with 0.5 mM 57FeIII-citrate for 30 min contained significant amount of FeII in a hexaaqua complex form. This is a direct evidence for the Strategy I iron uptake mechanism. Correlation was found between the decrease in Fe reductase activity and the ratio of FeII–FeIII components as the time of iron supply was increased. The data may refer to a higher iron reduction rate as compared to its uptake/reoxidation in the cytoplasm in accordance with the increased reduction rate in iron deficient Strategy I plants.  相似文献   

9.
The reactions of hydroxylamine (HA) with several water-soluble iron(III) porphyrinate compounds, namely iron(III) meso-tetrakis-(N-ethylpyridinium-2yl)-porphyrinate ([FeIII(TEPyP)]5+), iron(III) meso-tetrakis-(4-sulphonatophenyl)-porphyrinate ([FeIII(TPPS)]3−), and microperoxidase 11 ([FeIII(MP11)]) were studied for different [FeIII(Porph)]/[HA] ratios, under anaerobic conditions at neutral pH. Efficient catalytic processes leading to the disproportionation of HA by these iron(III) porphyrinates were evidenced for the first time. As a common feature, only N2 and N2O were found as gaseous, nitrogen-containing oxidation products, while NH3 was the unique reduced species detected. Different N2/N2O ratios obtained with these three porphyrinates strongly suggest distinctive mechanistic scenarios: while [FeIII(TEPyP)]5+ and [FeIII(MP11)] formed unknown steady-state porphyrinic intermediates in the presence of HA, [FeIII(TPPS)]3− led to the well characterized soluble intermediate, [FeII(TPPS)NO]4−. Free-radical formation was only evidenced for [FeIII(TEPyP)]5+, as a consequence of a metal centered reduction. We discuss the catalytic pathways of HA disproportionation on the basis of the distribution of gaseous products, free radicals formation, the nature of porphyrinic intermediates, the FeII/FeIII redox potential, the coordinating capabilities of each complex, and the kinetic analysis. The absence of revealed either that no HAO-like activity was operative under our reaction conditions, or that , if formed, was consumed in the reaction milieu.  相似文献   

10.
We have examined the influence of carbon source on both the rate of iron reduction and the mineralogy of the reduction products with Shewanella putrefaciens strain W3-18-1. When pyruvate was the carbon source, the secondary products were spherules composed of siderite. When uridine was used as the carbon source, the products were hexagonal plate-like structures identified as iron carbonate hydroxide hydrate, also known as carbonate green rust, a precursor to fougerite. When lactate was used as the carbon source, products were a mixture of iron carbonate hydroxide and magnetite. In terms of reaction stoichiometry, there were differences in the amount of acetate produced depending on the starting organic carbon source. Incubation with pyruvate produced a relatively large amount of acetate compared to incubation with uridine and lactate. There were also differences in the final pH of the cultures. While the pH for incubations with lactate started at 8.6 and ended between 8.0–8.3, the pH of cultures incubated with uridine was found to be almost a full unit lower at the conclusion of the experiment (~7.4). Solubility diagrams based on the chemistry found in our experiments predict that the production of Fe2+ (aq) should always lead to the formation of magnetite. However, strain W3-18-1 produced different minerals depending on the carbon source utilized as the electron acceptor.  相似文献   

11.
Amyloid beta (Aβ) aggregation and oxidative stress are two of the central events in Alzheimer's Disease (AD). Both these phenomena can be caused by the interaction of Aβ with metal ions. In the last years the interaction between ZnII, CuII, and Aβ was much studied, but between iron and Aβ it is still little known. In this work we determine how three Aβ peptides, present in AD, interact with FeIII‐citrate. The three Aβ peptides are: full length Aβ1‐42, an isoform truncated at Glutamic acid in position three, Aβ3‐42, and its pyroglutamated form AβpE3‐42. Conformation and morphology of the three peptides, aggregated with and without FeIII‐citrate were studied. Besides, we have determined the strength of the interactions Aβ/FeIII‐citrate studying the effect of ethylenediaminetetraacetic acid as chelator. Results reported here demonstrate that FeIII‐citrate promotes the aggregation in all the three peptides. Moreover, Aspartic acid 1, Glutamic acid 3, and Tyrosine 10 have an important role in the coordination with iron, generating a more stable complex for Aβ1‐42 compared to that for the truncated peptides.  相似文献   

12.
A novel heterobinuclear mixed valence complex [FeIIICuII(BPBPMP)(OAc)2]ClO4, 1, with the unsymmetrical N5O2 donor ligand 2-bis[{(2-pyridylmethyl)aminomethyl}-6-{(2-hydroxybenzyl)(2-pyridylmethyl)}aminomethyl]-4-methylphenol (H2BPBPMP) has been synthesized and characterized. A combination of data from mass spectrometry, potentiometric titrations, X-ray absorption and electron paramagnetic resonance spectroscopy, as well as kinetics measurements indicates that in ethanol/water solutions an [FeIII–()OH–CuIIOH2]+ species is generated which is the likely catalyst for 2,4-bis(dinitrophenyl)phosphate and DNA hydrolysis. Insofar as the data are consistent with the presence of an FeIII-bound hydroxide acting as a nucleophile during catalysis, 1 presents a suitable mimic for the hydrolytic enzyme purple acid phosphatase. Notably, 1 is significantly more reactive than its isostructural homologues with different metal composition (FeIIIMII, where MII is ZnII, MnII, NiII, or FeII). Of particular interest is the observation that cleavage of double-stranded plasmid DNA occurs even at very low concentrations of 1 (2.5 M), under physiological conditions (optimum pH of 7.0), with a rate enhancement of 2.7×107 over the uncatalyzed reaction. Thus, 1 is one of the most effective model complexes to date, mimicking the function of nucleases.Electronic Supplementary Material Supplementary material is available for this article at .  相似文献   

13.
Three new chiral ligands bearing an O,O′,N donor set (OmethoxyOhydroxyNpyridine) were synthesised and coordinated to FeIII, FeII, NiII, CuII and ZnII to yield complexes with the general formula [M(OON)Clx]y. While the pyridine N and the hydroxy O atoms coordinate strongly to all applied metal ions, the methoxy donor seems not to be involved in coordination, although some evidence for a weak interaction between OMe and the ZnII were found in NMR spectra. In the bidentate O′,N coordination mode the new ligands exhibit several coordination geometries as analysed in the solid compounds by XRD, EXAFS and EPR and in solution by UV-Vis absorption, cyclic voltammetry, EXAFS, EPR or NMR spectroscopy.  相似文献   

14.
Minerals that contain ferric iron, such as amorphous Fe(III) oxides (A), can inhibit methanogenesis by competitively accepting electrons. In contrast, ferric iron reduced products, such as magnetite (M), can function as electrical conductors to stimulate methanogenesis, however, the processes and effects of magnetite production and transformation in the methanogenic consortia are not yet known. Here we compare the effects on methanogenesis of amorphous Fe (III) oxides (A) and magnetite (M) with ethanol as the electron donor. RNA-based terminal restriction fragment length polymorphism with a clone library was used to analyse both bacterial and archaeal communities. Iron (III)-reducing bacteria including Geobacteraceae and methanogens such as Methanosarcina were enriched in iron oxide-supplemented enrichment cultures for two generations with ethanol as the electron donor. The enrichment cultures with A and non-Fe (N) dominated by the active bacteria belong to Veillonellaceae, and archaea belong to Methanoregulaceae and Methanobacteriaceae, Methanosarcinaceae (Methanosarcina mazei), respectively. While the enrichment cultures with M, dominated by the archaea belong to Methanosarcinaceae (Methanosarcina barkeri). The results also showed that methanogenesis was accelerated in the transferred cultures with ethanol as the electron donor during magnetite production from A reduction. Powder X-ray diffraction analysis indicated that magnetite was generated from microbial reduction of A and M was transformed into siderite and vivianite with ethanol as the electron donor. Our data showed the processes and effects of magnetite production and transformation in the methanogenic consortia, suggesting that significantly different effects of iron minerals on microbial methanogenesis in the iron-rich coastal riverine environment were present.  相似文献   

15.
Roots of grasses in response to iron deficiency markedly increase the release of chelating substances (`phytosiderophores') which are highly effective in solubilization of sparingly soluble inorganic FeIII compounds by formation of FeIIIphytosiderophores. In barley (Hordeum vulgare L.), the rate of iron uptake from FeIIIphytosiderophores is 100 to 1000 times faster than the rate from synthetic Fe chelates (e.g. Fe ethylenediaminetetraacetate) or microbial Fe siderophores (e.g. ferrichrome). Reduction of FeIII is not involved in the preferential iron uptake from FeIIIphytosiderophores by barley. This is indicated by experiments with varied pH, addition of bicarbonate or of a strong chelator for FeII (e.g. batho-phenanthrolinedisulfonate). The results indicate the existence of a specific uptake system for FeIIIphytosiderophores in roots of barley and all other graminaceous species. In contrast to grasses, cucumber plants (Cucumis sativus L.) take up iron from FeIIIphytosiderophores at rates similar to those from synthetic Fe chelates. Furthermore, under Fe deficiency in cucumber, increased rates of uptake of FeIIIphytosiderophores are based on the same mechanism as for synthetic Fe chelates, namely enhanced FeIII reduction and chelate splitting. Two strategies are evident from the experiments for the acquisition of iron by plants under iron deficiency. Strategy I (in most nongraminaceous species) is characterized by an inducible plasma membrane-bound reductase and enhancement of H+ release. Strategy II (in grasses) is characterized by enhanced release of phytosiderophores and by a highly specific uptake system for FeIIIphytosiderophores. Strategy II seems to have several ecological advantages over Strategy I such as solubilization of sparingly soluble inorganic FeIII compounds in the rhizosphere, and less inhibition by high pH. The principal differences in the two strategies have to be taken into account in screening methods for resistance to `lime chlorosis'.  相似文献   

16.
The rates of electron exchange between ferricytochrome c (CIII)3 and ferrocytochrome c (CII) were observed as a function of the concentrations of ferrihexacyanide (FeIII) and ferrohexacyanide (FeII) by monitoring the line widths of several proton resonances of the protein. Addition of FeII to CIII homogeneously increased the line widths of the two downfield paramagnetically shifted heme methyl proton resonances to a maximal value. This was interpreted as indicating the formation of a stoichiometric complex, CIII·FeII, in the over-all reaction:
CIII+FeII?k?1k1CIII·FeII?k?2k2CII·FeIII?k?3k3CIII+FeII
Values for k1k?1 = 0.4 × 103m?1and k2 = 208 s?1, respectively, were calculated from the maximal change in line width observed at pH 7.0 and 25 °C. Changes in the line width of CIII in the presence of FeII and either KCl or FeIII suggest that complexation is principally ionic, that FeIII and FeII compete for a common site. Addition of saturating concentrations of FeIII to CIII produced only minor changes in the nuclear magnetic resonance spectrum of CIII suggesting that complexation occurs on the protein surface.Addition of FeIII to CII in the presence of excess FeII (to retain most of the protein as CII) increased the line width of the methyl protons of ligated methionine 80. A value for k?2 ≈ 2.08 × 104 s?1 was calculated from the dependence of linewidth on the concentration of FeII at 24 °C. These rates are shown to be consistent with the over-all rates of reduction and oxidation previously determined by stopped flow measurements, indicating that k2 and k?2 were rate limiting. From the temperature dependence the enthalpies of activation are 7.9 and 15.2 kcal/mol for k2 and k?2, respectively.  相似文献   

17.
In this work we report on the synthesis, crystal structure, and physicochemical characterization of the novel dinuclear [FeIIICdII(L)(μ-OAc)2]ClO4·0.5H2O (1) complex containing the unsymmetrical ligand H2L = 2-bis[{(2-pyridyl-methyl)-aminomethyl}-6-{(2-hydroxy-benzyl)-(2-pyridyl-methyl)}-aminomethyl]-4-methylphenol. Also, with this ligand, the tetranuclear [Fe2IIIHg2II(L)2(OH)2](ClO4)2·2CH3OH (2) and [FeIIIHgII(L)(μ-CO3)FeIIIHgII(L)](ClO4)2·H2O (3) complexes were synthesized and fully characterized. It is demonstrated that the precursor [FeIII2HgII2(L)2(OH)2](ClO4)2·2CH3OH (2) can be converted to (3) by the fixation of atmospheric CO2 since the crystal structure of the tetranuclear organometallic complex [FeIIIHgII(L)(μ-CO3)FeIIIHgII(L)](ClO4)2·H2O (3) with an unprecedented {FeIII(μ-Ophenoxo)2(μ-CO3)FeIII} core was obtained through X-ray crystallography. In the reaction 2 → 3 a nucleophilic attack of a FeIII-bound hydroxo group on the CO2 molecule is proposed. In addition, it is also demonstrated that complex (3) can regenerate complex (2) in aqueous/MeOH/NaOH solution. Magnetochemical studies reveal that the FeIII centers in 3 are antiferromagnetically coupled (J = − 7.2 cm− 1) and that the FeIII-OR-FeIII angle has no noticeable influence in the exchange coupling. Phosphatase-like activity studies in the hydrolysis of the model substrate bis(2,4-dinitrophenyl) phosphate (2,4-bdnpp) by 1 and 2 show Michaelis-Menten behavior with 1 being ~ 2.5 times more active than 2. In combination with kH/kD isotope effects, the kinetic studies suggest a mechanism in which a terminal FeIII-bound hydroxide is the hydrolysis-initiating nucleophilic catalyst for 1 and 2. Based on the crystal structures of 1 and 3, it is assumed that the relatively long FeIII…HgII distance could be responsible for the lower catalytic effectiveness of 2.  相似文献   

18.
We document the discovery of the first granular iron formation (GIF) of Archaean age and present textural and geochemical results that suggest these formed through microbial iron oxidation. The GIF occurs in the Nconga Formation of the ca. 3.0–2.8 Ga Pongola Supergroup in South Africa and Swaziland. It is interbedded with oxide and silicate facies micritic iron formation (MIF). There is a strong textural control on iron mineralization in the GIF not observed in the associated MIF. The GIF is marked by oncoids with chert cores surrounded by magnetite and calcite rims. These rims show laminated domal textures, similar in appearance to microstromatolites. The GIF is enriched in silica and depleted in Fe relative to the interbedded MIF. Very low Al and trace element contents in the GIF indicate that chemically precipitated chert was reworked above wave base into granules in an environment devoid of siliciclastic input. Microbially mediated iron precipitation resulted in the formation of irregular, domal rims around the chert granules. During storm surges, oncoids were transported and deposited in deeper water environments. Textural features, along with positive δ56Fe values in magnetite, suggest that iron precipitation occurred through incomplete oxidation of hydrothermal Fe2+ by iron‐oxidizing bacteria. The initial Fe3+‐oxyhydroxide precipitates were then post‐depositionally transformed to magnetite. Comparison of the Fe isotope compositions of the oncoidal GIF with those reported for the interbedded deeper water iron formation (IF) illustrates that the Fe2+ pathways and sources for these units were distinct. It is suggested that the deeper water IF was deposited from the evolved margin of a buoyant Fe2+aq‐rich hydrothermal plume distal to its source. In contrast, oncolitic magnetite rims of chert granules were sourced from ambient Fe2+aq‐depleted shallow ocean water beyond the plume.  相似文献   

19.
 Previous studies have demonstrated that 2-hydroxy-1-naphthaldehyde isonicotinoyl hydrazone (NIH) and several other aroylhydrazone chelators possess anti-neoplastic activity due to their ability to bind intracellular iron. In this study we have examined the structure and properties of NIH and its FeIII complex in order to obtain further insight into its anti-tumour activity. Two tridentate NIH ligands deprotonate upon coordination to FeIII in a meridional fashion to form a distorted octahedral, high-spin complex. Solution electrochemistry of [Fe(NIH–H)2]+ shows that the trivalent oxidation state is dominant over a wide potential range and that the FeII analogue is not a stable form of this complex. The fact that [Fe(NIH–H)2]+ cannot cycle between the FeII and FeIII states suggests that the production of toxic free-radical species, e.g. OH . or O2 . , is not part of this ligand's cytotoxic action. This suggestion is supported by cell culture experiments demonstrating that the addition of FeIII to NIH prevents its anti-proliferative effect. The chemistry of this chelator and its FeIII complex are discussed in the context of understanding its anti-tumour activity. Received: 12 November 1998 / Accepted: 9 February 1999  相似文献   

20.
Hydrogen bonding networks proximal to metal centers are emerging as a viable means for controlling secondary coordination spheres. This has led to the regulation of reactivity and isolation of complexes with new structural motifs. We have used the tridenate ligand bis[(N′-tert-butylureido)-N-ethyl]-N-methylaminato ([H21]2−) that contains two hydrogen bond donors to examine the oxidation of the FeII-acetate complex, [FeIIH212-OAc)] with dioxygen, amine N-oxides, and xylyl azide. A complex with FeIII-O-FeIII core results from the oxidation with dioxygen and amine N-oxides, in which the oxo ligand is involved in hydrogen bonding to the [H21]2− ligand. A distinctly different hydrogen bonding network was found in FeIII dimer isolated from the reaction with the xylyl azide: a rare FeIII-N(R)-FeIII core was observed that does not have hydrogen bonds to the bridging nitrogen atom. The intramolecular H-bond networks within these dimers appear to adjust to the presence of the bridging species and rearrange to its size and electron density.  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