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 31 毫秒
1.
 In this study, the effect of ferric ion and cell concentrations on the oxidation of ferrous ion by T. ferrooxidans was investigated. Ferric ions competitively inhibited ferrous ion oxidation by the bacteria. The inhibitory effect of ferric ion was, however, reduced by increasing cell concentration. The apparent ferric ion inhibition constant did not change with increasing cell concentration. The ferrous ion oxidation kinetics in the absence and presence of ferric ion changes from the standard Michaelis-Menten type at low cell concentrations to pseudo-first-order kinetics at high cell concentration. Received: 8 August 1995/Received revision: 31 October 1995/Accepted: 10 November 1995  相似文献   

2.
A chemo-biochemical process using Thiobacillus ferrooxidans for desulphurization of gaseous fuels and emissions containing hydrogen sulphide (H2S) has been developed. In the first stage, H2S present in fuel gas and emissions is selectively oxidized to elemental sulphur using ferric sulphate. The ferrous sulphate produced in the first stage of the process is oxidized to ferric sulphate using Thiobacillus ferrooxidans for recycle and reuse in the process. The effects of process variables, temperature, pH, total dissolved solids (TDS), elemental sulphur, ferric and magnesium ions on bio-oxidation of ferrous ions to ferric ions were investigated using flask culture experiments. The bio-oxidation of ferrous ions to ferric ions could be achieved efficiently in the temperature range of 20(+/-1)-44(+/-1) degrees C. A pH range of 1.8(+/-0.02)-2.2(+/-0.02) was optimum for the growth of culture and effective bio-oxidation of ferrous ions to ferric ions. The effect of TDS on bio-oxidation of ferrous ions indicated that a preacclimatized culture in a growth medium containing high dissolved solid was required to achieve effective bio-oxidation of ferrous ions. Elemental sulphur ranging from 1000 to 100,000 mg/l did not have any effect on efficiency of ferrous ion oxidation. The efficiency of bio-oxidation of ferrous ions to ferric ions was not affected in the presence of ferric ions up to a concentration of 500 mg/l while 3 mg/l of magnesium ion was optimal for achieving effective bio-oxidation.  相似文献   

3.
Phosphatidylethanolamine (PE) is generally more oxidizable than phosphatidylcholine (PC). To determine the difference in reactivities to oxidation between PE and PC, it is necessary for their fatty acid moieties to be uniform. Experimental results of the ferrous ion-catalyzed oxidation of dilinoleoylphosphatidylcholine, dilinoleoylphosphatidylethanolamine, and dilinoleoyldiglyceride revealed that the rate of oxidation depends on the type of base. Ferrous ion possessed a high catalytic activity in hydroperoxide formation at pH 5.8. Iron ions might initiate the oxidation of phospholipids by forming free radicals. Phosphoethanolamine was capable of trapping ferrous ion and preventing it from being autoxidized to ferric ion. Trapping of ferrous ion might be responsible for the significant oxidizability of PE at pH 5.8 ~ 7.0. In the ferrous ion-ascorbic acid (AsA) catalyzed oxidation system, PC oxidation was remarkably enhanced at pH 7.0. In this case, no reduction of ferric ion occurred, but AsA had a prooxidant effect of accelerating the formation of free radicals.  相似文献   

4.
The effect of dilution rate and feed solids concentration on the bacterial leaching of a pyrite/arsenopyrite ore concentrate was studied. A mathematical model was developed for the process based on the steady-state data collected over the range of dilution rates (20 to 110 h) and feed solids concentrations (6 to 18% w/v) studied. A modified Monod model with inhibition by arsenic was used to model bacterial ferrous ion oxidation rates. The model assumes that (i) pyrite and arsenopyrite leaching occurs solely by the action of ferric iron produced from the bacterial oxidation of ferrous iron and (ii) bacterial growth rates are proportional to ferrous ion oxidation rate. The equilibrium among the various ionic species present in the leach solution that are likely to have a significant effect on the bioleach process were included in the model. (c) 1994 John Wiley & Sons, Inc.  相似文献   

5.
V Favaudon  J M Lhoste 《Biochemistry》1975,14(21):4739-4744
The oxidation-reduction reactions of tetraacetylriboflavine in the presence of various metal ions in dimethylformamide have been investigated using the stopped-flow technique under anaerobic conditions. Dismutation kinetics in the presence of redox-inactive dissociated divalent metal ions such as Cd2+, Zn2+, and Fe2+ are typically triphasic. Metal ions act primarily upon an intermediate flavine dimer formed by fast association of flavoquinone and flavohydroquinone, resulting in a parallel formation and neutral and chelated radicals. A competition between metal ions and proton donors, e.g. the neutral flavohydroquinone (FredH3), is observed at the level of this intermediate complex. Small spectral changes occur secondarily as an ill-defined intermediate phase which could correspond to the reorganization of the solvation of radical chelate. The neutral radical is finally chelated at a much slower rate, the yield of total radical formation remaining almost unchanged during this kinetic phase. The oxidation of flavohydroquinone by ferric ions, either dissociated or strongly coordinated within a porphyrin, is complete and proceeds through biphasic kinetics. The first phase (Fred leads to F) is much faster than the second one (F leads to Fox). Dismutation resulting from the transient accumulation of neutral flavosemiquinone competes with the direct oxidation with ferric ions for the completion of the second oxidation step. The relative rate of dismutation is essentially limited by acidic-basic reactions in the absence of an excess of ferrous ion. The kinetic analysis of the direct oxidation reactions favors an outer-sphere mechanism for the electron transfer to the ferric ion, either free or strongly coordinated. The formation of a ferrous radical chelate can result from the dismutation reactions only when the amount of ferric ion initially present is not sufficient for complete oxidation.  相似文献   

6.
Stannous and cuprous ion oxidation by Thiobacillus ferrooxidans.   总被引:7,自引:0,他引:7  
Oxidation of stannous chloride by Thiobacillus ferrooxidans was studied manometrically. At low stannous ion concentrations, initial oxidation rate was proportional to concentration. Optimum pH for oxidation was 2.3 optimum temperature was 37-40 degrees C. Spectrophotometry showed reduction of cytochromes in suspensions of whole cells on addition of ferrous, stannous, or cuprous salts. Cytochrome c reductase activity in cell-free extracts was assayed with ferrous, stannous, or cuprous ions as electron donors. It appears unlikely that an essential non-biological reaction, the reduction of ferric ions by stannous or cuprous ions, is involved. Growth of T. ferrooxidans was not obtained with either stannous chloride or stannous sulphate as sole energy source.  相似文献   

7.
The concentrations of ferrous and ferric ions change dramatically during the course of the batch experiments usually performed to study the kinetics of the bacterial oxidation of ferrous ions and sulfide minerals. This change in concentration of the iron species during the course of the experiment often makes it difficult to interpret the results of these experiments, as is evidenced by the lack of consensus concerning the mechanism of bacterial leaching. If the concentrations of ferrous and ferric ions were constant throughout the course of the batch experiment, then the role of the bacteria could be easily established, because the rate of the chemical leaching should be the same at a given redox potential in the presence and in the absence of bacteria. In this paper we report an experiment designed to obtain kinetic data under these conditions. The redox potential is used as a measure of the concentrations of ferrous and ferric ions, and the redox potential of the leaching solution is controlled throughout the experiment by electrolysis. The effects of ferrous, ferric, and arsenite ions on the rate of growth of Thiobacillus ferrooxidans on ferrous ions in this redox-controlled reactor are presented. In addition, the growth of this bacterium on ferrous ions in batch culture was also determined, and it is shown that the parameters obtained from the batch culture and the redox-controlled batch culture are the same. An analysis of the results from the batch culture indicates that the initial number of bacteria that are adapted to the solution depends on the concentrations of ferrous and arsenite ions.  相似文献   

8.
Kinetic constants for the oxidation of uranous and ferrous ions byThiobacillus ferrooxidans were estimated. The kinetics indicate a direct biological mechanism for uranium oxidation. The complex interrelations of ferric, uranyl and uranous ion inhibition are considered.  相似文献   

9.
The FRAP reagent contains 2,4,6-tris(2-pyridyl)-s-triazine, which forms a blue–violet complex ion in the presence of ferrous ions. Although the FRAP (ferric reducing/antioxidant power) assay is popular and has been in use for many years, the correct molar extinction coefficient of this complex ion under FRAP assay conditions has never been published, casting doubt on the validity of previous calibrations. A previously reported value of 19,800 is an underestimate. We determined that the molar extinction coefficient was 21,140. The value of the molar extinction coefficient was also shown to depend on the type of assay and was found to be 22,230 under iron assay conditions, in good agreement with published data. Redox titration indicated that the ferrous sulfate heptahydrate calibrator recommended by Benzie and Strain, the FRAP assay inventors, is prone to efflorescence and, therefore, is unreliable. Ferrous ammonium sulfate hexahydrate in dilute sulfuric acid was a more stable alternative. Few authors publish their calibration data, and this makes comparative analyses impossible. A critical examination of the limited number of examples of calibration data in the published literature reveals only that Benzie and Strain obtained a satisfactory calibration using their method.  相似文献   

10.
The effects of pH, ferrous and ferric ion concentrations on iron oxidation by Thiobacillus ferrooxidans were examined. The initial temperature and bacterial concentration were maintained at 37°C and 2±1×104cells/ml, respectively. The iron oxidation rate increased with increased initial ferrous iron concentration to 4g/l and thereafter decreased. The presence of iron(III) showed a negative effect on the bacterial iron oxidation rate. The increase of pH also showed an increase in the oxidation rate up to pH 1.75. The oxidation rate followed first order kinetics for the parameters studied. A rate equation has been developed.  相似文献   

11.
The effects of pH, ferrous and ferric ion concentrations on iron oxidation by Thiobacillus ferrooxidans were examined. The initial temperature and bacterial concentration were maintained at 37°C and 2±1×104cells/ml, respectively. The iron oxidation rate increased with increased initial ferrous iron concentration to 4g/l and thereafter decreased. The presence of iron(III) showed a negative effect on the bacterial iron oxidation rate. The increase of pH also showed an increase in the oxidation rate up to pH 1.75. The oxidation rate followed first order kinetics for the parameters studied. A rate equation has been developed.  相似文献   

12.
Uptake of iron by apoferritin from a ferric dihydrolipoate complex   总被引:1,自引:0,他引:1  
A study was made on the uptake of iron by horse spleen apoferritin, by using as an iron source the same ferric dihydrolipoate complex which represents the major product in the anaerobic removal of ferritin-bound iron by dihydrolipoate at neutral pH. The ferric dihydrolipoate complex was chemically synthesized and used as an iron donor to apoferritin. Iron uptake was studied, at slightly alkaline pH and in anaerobic conditions, as a function of the concentration of both the iron donor and apoferritin. Isolation of ferritin from mixtures of ferric dihydrolipoate and apoferritin, and subsequent identification of the oxidation state of ferritin-bound iron, showed that the first metal atoms were taken up in the ferrous form and that this early step was accompanied by accumulation of ferric iron. Total iron uptake increased with the molar ratio of complex to apoprotein and ranged over 25-40% of the iron being supplied. The amount of ferrous iron found inside the protein did not exceed 50-60 mol iron/mol ferritin after a 48-h incubation. At this time, ferric iron represented a significant fraction of the iron found in the isolated ferritin. Analytical and spectroscopic data indicated that fractional rates and equilibria for disassembly of the ferric complex in the presence of apoferritin were independent of the concentration of the protein and of the complex itself.  相似文献   

13.
Fago A  Mathews AJ  Brittain T 《IUBMB life》2008,60(6):398-401
We propose a new hypothesis for the molecular mechanism by which neuroglobin exerts its protective effect in hypoxia-induced cell death. Our recent observation of a very rapid electron-transfer reaction between ferrous neuroglobin and ferric cytochrome c is central to this hypothesis. In contrast to previously suggested roles for neuroglobin, related to its putative but unlikely oxygen storage/transport properties or its ability to react with nitrogen oxides, we suggest that ferrous neuroglobin exerts its protective effect via modulation of the early events in the intrinsic apoptotic pathway. We suggest this is achieved by the rapid reduction of cytosolic ferric cytochrome c by neuroglobin. The maintenance of cytochrome c in the nonapoptotic ferrous oxidation state and the concomitant generation of ferric neuroglobin in this reaction fit well with known feedback processes in the early events of the intrinsic apoptotic pathway. Our hypothesis also fits well with a number of previously uncorrelated findings, including the localization of neuroglobin in close proximity to mitochondria, the high concentration of neuroglobin in cells whose basal rates of aerobic metabolism are extremely high, and the cell types which are subject to large calcium ion fluxes in their normal physiology.  相似文献   

14.
The influence of temperature, pH, and substrate and product concentrations on the oxidation rate of ferrous iron by biofilm of Thiobacillus ferrooxidans was determined. The experiments were performed in an inverse fluidized-bed biofilm reactor in which the biofilm thickness was kept constant at 80 mum. Oxygen concentration and diffusion through the biofilm did not limit the oxidation rate. The oxidation rate was almost unaffected by temperature between 13 and 38 degrees C, pH between 1.3 and 2.2, ferric iron concentration up to 14 g/L, or ferrous iron concentration from 4 to 13 g/L. The kinetics of the process was described by the Monod equation with respect to the mass of the biofilm and with ferrous ions as the limiting substrate.  相似文献   

15.
Previously, we showed the presence in radish (Raphanus sativus L.) plasmalemma vesicles of an NAD(P)H oxidase, active at pH 4.5-5.0, which elicits the formation of anion superoxide (Vianello and Macrí (1989) Biochim. Biophys. Acta 980, 202-208). In this work, we studied the role of hydrogen peroxide and iron ions upon this oxidase activity. NADH oxidation was stimulated by ferrous ions and, to a lesser extent, by ferric ions. Salicylate and benzoate, two known hydroxyl radical scavengers, inhibited both basal and iron-stimulated NADH oxidase activity. The iron chelators EDTA (ethylenediaminetetraacetic acid) and DFA (deferoxamine melysate) at high concentrations (2 mM) inhibited the NADH oxidation, whereas they were ineffective at lower concentrations (80 microM); the subsequent addition of ferrous ions caused a rapid and limited increase of oxygen consumption which later ceased. Hydrogen peroxide was not detected during NADH oxidation but, in the presence of salicylate, its formation was found in significant amounts. NADH oxidase activity was also associated to a Fe2+ oxidation which was only partially inhibited by salicylate. Ferrous ion oxidation was partially inhibited by catalase and prevented by superoxide dismutase, while ferric ion reduction was abolished by catalase and unaffected by superoxide dismutase. These results show that during NADH oxidation iron ions undergo oxidoreduction and that hydrogen peroxide is produced and rapidly consumed. As previously suggested, this oxidation appears linked to the univalent oxidoreduction of iron ions by a reduced flavoprotein of radish plasmalemma which is then converted to a radical form. The latter, reacting with oxygen generates the superoxide anion which dismutases to H2O2. Hydrogen peroxide, through a Fenton's reaction, may react with Fe2+ to produce hydroxyl radicals, or with Fe3+ to generate the superoxide anion.  相似文献   

16.
The results of recent research have shown that the bioleaching of sulfide minerals occurs via a two‐step mechanism. In this mechanism, the sulfide mineral is chemically oxidized by the ferric‐iron in the bioleaching liquor. The ferrous‐iron produced is subsequently oxidized to ferric‐iron by the microorganism. Further research has shown that the rates of both the ferric leaching and ferrous‐iron oxidation are governed by the ferric/ferrous‐iron ratio (i.e., the redox potential). During the steady‐state operation of a bioleach reactor, the rate of iron turnover between the chemical ferric leaching of the mineral and the bacterial oxidation of the ferrous‐iron will define the rate and the redox potential at which the system will operate. The balance between the two rates will in turn depend on the species used, the microbial concentration, the residence time employed, the nature of the sulfide mineral being leached, and its active surface area. The model described proposes that the residence time and microbial species present determine the microbial growth rate, which in turn determines the redox potential in the bioleach liquor. The redox potential of the solution, in turn, determines the degree of leaching of the mineral; that is, conversion in the bioleach reactor. © 1999 John Wiley & Sons, Inc. Biotechnol Bioeng 64: 671–677, 1999.  相似文献   

17.
Human porphyria cutanea tarda is an unusual consequence of common hepatic disorders such as alcoholic liver disease. Hepatic iron plays a key role in the expression of the metabolic lesions, i.e., defective hepatic decarboxylation of porphyrinogens, catalyzed by uroporphyrinogen decarboxylase. This prompted the present study to determine the in vitro effects of iron on the uroporphyrinogen substrate in the absence and presence of atmospheric oxygen. We observed that (i) unless oxygen is the limiting reactant, autoxidation of ferrous iron and iron-catalyzed oxidation of uroporphyrinogen occurred soon after initiating the reaction at pH 7.4 and 30 degrees C in buffers which are non- or poor chelators of iron; (ii) the rates of uroporphyrinogen oxidation were proportional to the initial concentration of ferrous ion; (iii) about 70% of the oxidations of uroporphyrinogen were accountable due to a free-radical chain reaction pathway involving superoxide radical and hence inhibitable by superoxide dismutase; (iv) uroporphyrinogen could be further oxidized to completion by the hydroxyl radical since the reaction was partially inhibited by both mannitol and catalase which prevent hydroxyl radical production; (v) the oxidizing effects of ferric ion on uroporphyrinogen were none or negligible as compared to those of ferrous ion. Ferric was reduced to ferrous ion in the presence of dithiothreitol. When the ferrous ion thus formed was reoxidized in the presence of atmospheric oxygen, minor but definite oxidations of both uroporphyrinogen and dithiothreitol were observed. The oxidations of Fe2+ and uroporphyrinogen could be blocked by 1,10-phenanthroline, a ferrous iron chelator. The data suggest that ferrous is the reactive form of iron that may contribute to pathogenic development of the disease by irreversibly oxidizing the porphyrinogen substrates to nonmetabolizable porphyrins, which accumulate in porphyric liver.  相似文献   

18.
The stoichiometry and kinetics of the spontaneous, chemical reaction between pyrite and ferric iron was studied at 30, 45, and 70 degrees C in shake flasks at pH 1.5 by monitoring the ferrous iron, total iron, elemental sulfur, and sulfate concentration profiles in time. It was found that the sulfur moiety of pyrite was oxidized completely to sulfate. Elemental sulfur was not produced in detectable amounts. The iron moiety of pyrite was released as ferrous iron. All observed initial reaction rates could be fitted into an empirical equation. This equation includes the concentrations of ferric iron and pyrite, and a constant which is dependent on the temperature and the nature of the main anion present. It was observed that ferrous iron formed during the reaction slowed down the oxidation of pyrite by ferric iron. The extent of this effect decreased with increasing temperature. With the aid of the empirical equation, the contribution of the chemical oxidation of pyrite by ferric iron to the overall oxidation in a hypothetical plug-flow reactor, in which biologically mediated oxdidation of pyrite and ferrous iron by oxygen also takes place, can be assessed. At 30, 45, and 70 degrees C, respectively, 2, 8-17, and 43% of the pyrite was oxidized chemically by ferric iron. Therefore, it is expected that only in reactors operating at high temperatures with extremely thermophilic bacteria, will chemical oxidation cause a significant deviation from the apparent first order overall kinetics of biological pyrite oxidation.  相似文献   

19.
Mechanism of Bacterial Pyrite Oxidation   总被引:14,自引:1,他引:13       下载免费PDF全文
The oxidation by Ferrobacillus ferrooxidans of untreated pyrite (FeS(2)) as well as HCl-pretreated pyrite (from which most of the acid-soluble iron species were removed) was studied manometrically. Oxygen uptake was linear during bacterial oxidation of untreated pyrite, whereas with HCl-pretreated pyrite both a decrease in oxygen uptake at 2 hr and nonlinear oxygen consumption were observed. Ferric sulfate added to HCl-pretreated pyrite restored approximately two-thirds of the decrease in total bacterial oxygen uptake and caused oxygen uptake to revert to nearly linear kinetics. Ferric sulfate also oxidized pyrite in the absence of bacteria and O(2); recovery of ferric and ferrous ions was in excellent agreement with the reaction Fe(2)(SO(4))(3) + FeS(2) = 3FeSO(4) + 2S, but the elemental sulfur produced was negligible. Neither H(2)S nor S(2)O(3) (2-) was a product of the reaction. It is probable that two mechanisms of bacterial pyrite oxidation operate concurrently: the direct contact mechanism which requires physical contact between bacteria and pyrite particles for biological pyrite oxidation, and the indirect contact mechanism according to which the bacteria oxidize ferrous ions to the ferric state, thereby regenerating the ferric ions required for chemical oxidation of pyrite.  相似文献   

20.
The dual functions of alpha-tocopherol in the oxidation of lipids in aqueous dispersions in the presence of iron were studied, aiming specifically at elucidating the effect of interaction between alpha-tocopherol and iron. Ferrous ion decomposed hydroperoxide rapidly and induced the free radical chain oxidation of soybean phosphatidylcholine liposomes. alpha-Tocopherol acted primarily as a radical scavenger in the oxidation induced by ferrous ion and acted as an antioxidant. Ferric ion decomposed hydroperoxide much more slowly than ferrous ion, but it also induced the oxidation of liposomal membranes. alpha-Tocopherol incorporated into artificial liposomal membranes reduced ferric ion rapidly to give more reactive ferrous ion, and alpha-tocopherol acted either as an antioxidant or as a prooxidant depending on the experimental conditions. When alpha-tocopherol was depleted by the interaction with ferric ion, it acted solely as a prooxidant, whereas if some alpha-tocopherol remained, it acted as an antioxidant. On the other hand, alpha-tocopherol residing in the intact erythrocyte membranes did not reduce ferric ion in the aqueous region.  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