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.
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   

2.
The influence of lithotrophic Fe(II)-oxidizing bacteria on patterns of ferric oxide deposition in opposing gradients of Fe(II) and O(2) was examined at submillimeter resolution by use of an O(2) microelectrode and diffusion microprobes for iron. In cultures inoculated with lithotrophic Fe(II)-oxidizing bacteria, the majority of Fe(III) deposition occurred below the depth of O(2) penetration. In contrast, Fe(III) deposition in abiotic control cultures occurred entirely within the aerobic zone. The diffusion microprobes revealed the formation of soluble or colloidal Fe(III) compounds during biological Fe(II) oxidation. The presence of mobile Fe(III) in diffusion probes from live cultures was verified by washing the probes in anoxic water, which removed ca. 70% of the Fe(III) content of probes from live cultures but did not alter the Fe(III) content of probes from abiotic controls. Measurements of the amount of Fe(III) oxide deposited in the medium versus the probes indicated that ca. 90% of the Fe(III) deposited in live cultures was formed biologically. Our findings show that bacterial Fe(II) oxidation is likely to generate reactive Fe(III) compounds that can be immediately available for use as electron acceptors for anaerobic respiration and that biological Fe(II) oxidation may thereby promote rapid microscale Fe redox cycling at aerobic-anaerobic interfaces.  相似文献   

3.
Batch cultures of a lithotrophic Fe(II)-oxidizing bacterium, strain BrT, isolated from the rhizosphere of a wetland plant, were grown in bioreactors and used to determine the significance of microbial Fe(II) oxidation at circumneutral pH and to identify abiotic variables that affect the partitioning between microbial oxidation and chemical oxidation. Strain BrT grew only in the presence of an Fe(II) source, with an average doubling time of 25 h. In one set of experiments, Fe(II) oxidation rates were measured before and after the cells were poisoned with sodium azide. These experiments indicated that strain BrT accounted for 18 to 53% of the total iron oxidation, and the average cellular growth yield was 0.70 g of CH2O per mol of Fe(II) oxidized. In a second set of experiments, Fe(II) was constantly added to bioreactors inoculated with live cells, killed cells, or no cells. A statistical model fitted to the experimental data demonstrated that metabolic Fe(II) oxidation accounted for up to 62% of the total oxidation. The total Fe(II) oxidation rates in these experiments were strongly limited by the rate of Fe(II) delivery to the system and were also influenced by O2 and total iron concentrations. Additionally, the model suggested that the microbes inhibited rates of abiotic Fe(II) oxidation, perhaps by binding Fe(II) to bacterial exopolymers. The net effect of strain BrT was to accelerate total oxidation rates by up to 18% compared to rates obtained with cell-free treatments. The results suggest that neutrophilic Fe(II)-oxidizing bacteria may compete for limited O2 in the rhizosphere and therefore influence other wetland biogeochemical cycles.  相似文献   

4.
The leaching of iron pyrite by Thiobacillus ferrooxidans was studied in a continuous stirred tank reactor at a variety of dilution rates (0.012-0.22 h(-1)), pyrite surface areas (18-194 m(2)/L), and inlet soluble substrate (Fe(2+)) concentrations (0-3000 ppm). The bacterial leaching rate was found to increase with increasing pyrite surface area, dilution rate, and inlet Fe(2+) concentration. The concentration of bacteria in solution was related to the concentration of bacteria attached to the pyrite surface by a Langmuir-type adsorption-desorption relation. Fitting the experimental data to this relation yielded a value for the area occupied per bacterium of 86 mum(2). This result is consistent with the concept of preferential bacterial attachment of certain sites on the solid. A bacterial growth model was developed that included both bacterial growth in solution and growth of bacteria attached to the pyrite surface. The specific growth rate of the attached bacteria was calculated from this model and was found to increase with increasing solid dilution rate and to decrease with increasing pyrite surface area and soluble substance concentration. An explanation of these results based on an active-inactive site mechanisms was also developed.  相似文献   

5.
Various species of soluble iron in pyrite‐grown cultures of Thiobacillus ferrooxidans were determined by colorimetry, atomic absorption spectrometry, and ultraviolet spectroscopy. All the cultures were incubated for six weeks before iron analysis. The effects of the following factors were investigated: particle size, initial pH, shaking (aeration), concentration of pyrite, and concentration of yeast extract. Shaking, but not initial pH nor particle size, influenced the relative proportion of different iron species. Polynomial regressions could be used to describe the functional relationship between the different iron species and concentration of pyrite; fewer relationships were evident with respect to concentration of yeast extract. The variance‐covariance matrices indicated a linear dependence among the different iron species. Canonical correlations indicated perfect correlations between group variables of iron, copper, and zinc, with the exception of an absence of significant correlation with the hydroxy complex of iron (FeOH2+).

The dissolved ferrous iron (dissociated and weakly chelated) always remained less than 7% of the total iron in solution. The total ferrous iron, which included complexed species, amounted to 7–34% of the total iron in solution. The concentrations of dissociated ferrous and ferric iron and their weak chelates (the dissolved iron) remained mostly constant, irrespective of the concentration of the total iron in solution. Most of the total iron was complexed as ferric species and the amount correlated with culture conditions. The hydroxy complex (FeOH2+), which was indicative of the relative amount of hydrolyzable ferric iron upon dilution in CO2‐free water, usually ranged between 60 and 80% of the total iron. The amount of the total iron in uninoculated controls was less than 12% of that solu‐bilized in the presence of iron‐oxidizing thiobacilli.

T. ferrooxidans was enumerated by a most‐probable‐number technique after three and six weeks of growth on pyrite. The counts after three weeks indicated an increase in the number of free and loosely attached bacteria, followed by a decline of about one order of magnitude in bacterial numbers after six weeks. The technique for bacterial enumeration was deemed unsatisfactory because it could not account for cells attached on pyrite.  相似文献   

6.
7.
The influence of lithotrophic Fe(II)-oxidizing bacteria on patterns of ferric oxide deposition in opposing gradients of Fe(II) and O2 was examined at submillimeter resolution by use of an O2 microelectrode and diffusion microprobes for iron. In cultures inoculated with lithotrophic Fe(II)-oxidizing bacteria, the majority of Fe(III) deposition occurred below the depth of O2 penetration. In contrast, Fe(III) deposition in abiotic control cultures occurred entirely within the aerobic zone. The diffusion microprobes revealed the formation of soluble or colloidal Fe(III) compounds during biological Fe(II) oxidation. The presence of mobile Fe(III) in diffusion probes from live cultures was verified by washing the probes in anoxic water, which removed ca. 70% of the Fe(III) content of probes from live cultures but did not alter the Fe(III) content of probes from abiotic controls. Measurements of the amount of Fe(III) oxide deposited in the medium versus the probes indicated that ca. 90% of the Fe(III) deposited in live cultures was formed biologically. Our findings show that bacterial Fe(II) oxidation is likely to generate reactive Fe(III) compounds that can be immediately available for use as electron acceptors for anaerobic respiration and that biological Fe(II) oxidation may thereby promote rapid microscale Fe redox cycling at aerobic-anaerobic interfaces.  相似文献   

8.
Batch cultures of a lithotrophic Fe(II)-oxidizing bacterium, strain BrT, isolated from the rhizosphere of a wetland plant, were grown in bioreactors and used to determine the significance of microbial Fe(II) oxidation at circumneutral pH and to identify abiotic variables that affect the partitioning between microbial oxidation and chemical oxidation. Strain BrT grew only in the presence of an Fe(II) source, with an average doubling time of 25 h. In one set of experiments, Fe(II) oxidation rates were measured before and after the cells were poisoned with sodium azide. These experiments indicated that strain BrT accounted for 18 to 53% of the total iron oxidation, and the average cellular growth yield was 0.70 g of CH2O per mol of Fe(II) oxidized. In a second set of experiments, Fe(II) was constantly added to bioreactors inoculated with live cells, killed cells, or no cells. A statistical model fitted to the experimental data demonstrated that metabolic Fe(II) oxidation accounted for up to 62% of the total oxidation. The total Fe(II) oxidation rates in these experiments were strongly limited by the rate of Fe(II) delivery to the system and were also influenced by O2 and total iron concentrations. Additionally, the model suggested that the microbes inhibited rates of abiotic Fe(II) oxidation, perhaps by binding Fe(II) to bacterial exopolymers. The net effect of strain BrT was to accelerate total oxidation rates by up to 18% compared to rates obtained with cell-free treatments. The results suggest that neutrophilic Fe(II)-oxidizing bacteria may compete for limited O2 in the rhizosphere and therefore influence other wetland biogeochemical cycles.  相似文献   

9.
Accumulation of elemental sulfur during pyrite oxidation lowers the efficiency of coal desulfurization and bioleaching. In the case of pyrite bioleaching by Leptospirillum ferrooxidans, an iron(II)-ion-oxidizing organism without sulfur-oxidizing capacity, from the pyritic sulfur moiety about 10% elemental sulfur, 2% pentathionate, and 1% tetrathionate accumulated by a recently described cyclic pyrite oxidation mechanism. In the case of pure cultures of Thiobacillus ferrooxidans and mixed cultures of L. ferrooxidans and T. thiooxidans, pyrite was nearly completely oxidized to sulfate because of the capacity of these cultures to oxidize both iron(II) ions and sulfur compounds. Pyrite oxidation in acidic solutions, mediated chemically by iron(III) ion, resulted in an accumulation of similar amounts of sulfur compounds as obtained with L. ferrooxidans. Changes of pH to values below 2 or in the iron ion concentration are not decisive for diverting the flux of sulfur compounds. The literature on pyrite bioleaching is in agreement with the findings indicating that the chemistry of direct and indirect pyrite leaching is identical. Received: 20 April 1998 / Received revision: 27 August 1998 / Accepted: 3 September 1998  相似文献   

10.
Microbial biofilms preferentially colonized pyrite surfaces of black shale incubated in groundwater in the Newark Basin (northeastern United States) for 1 month. SEM observation revealed the co-occurrence of bacteria-shaped pits and secondary iron minerals on pyrite, which indicate biological involvement in pyrite weathering and secondary solid formation. Of the 24 16S rDNA sequences obtained from bacterial communities on pyrite, arsenopyrite and quartz sand, 22 belonged to the phylum proteobacteria, including 5 identified as β or ?-proteobacteria capable of oxidizing iron or sulfur, 16 identified as members of the Fe(III)-reducing Geobacteraceae in the δ-proteobacteria and one identified as the Fe(III)-reducing Ferribacterium. Results indicate that microbes play an essential role in the oxidation of iron sulfides (via direct contact and indirect pathways) and the reduction of iron oxides in pyrite-bearing substrata of a slightly acidic black shale aquifer.  相似文献   

11.
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   

12.
Anaerobic microbial oxidation of Fe(II) was only recently discovered and very little is known about this metabolism. We recently demonstrated that several dissimilatory perchlorate-reducing bacteria could utilize Fe(II) as an electron donor under anaerobic conditions. Here we report on a more in-depth analysis of Fe(II) oxidation by one of these organisms, Dechlorosoma suillum. Similarly to most known nitrate-dependent Fe(II) oxidizers, D. suillum did not grow heterotrophically or lithoautotrophically by anaerobic Fe(II) oxidation. In the absence of a suitable organic carbon source, cells rapidly lysed even though nitrate-dependent Fe(II) oxidation was still occurring. The coupling of Fe(II) oxidation to a particular electron acceptor was dependent on the growth conditions of cells of D. suillum. As such, anaerobically grown cultures of D. suillum did not mediate Fe(II) oxidation with oxygen as the electron acceptor, while conversely, aerobically grown cultures did not mediate Fe(II) oxidation with nitrate as the electron acceptor. Anaerobic washed cell suspensions of D. suillum rapidly produced an orange/brown precipitate which X-ray diffraction analysis identified as amorphous ferric oxyhydroxide or ferrihydrite. This is similar to all other identified nitrate-dependent Fe(II) oxidizers but is in contrast to what is observed for growth cultures of D. suillum, which produced a mixed-valence Fe(II)-Fe(III) precipitate known as green rust. D. suillum rapidly oxidized the Fe(II) content of natural sediments. Although the form of ferrous iron in these sediments is unknown, it is probably a component of an insoluble mineral, as previous studies indicated that soluble Fe(II) is a relatively minor form of the total Fe(II) content of anoxic environments. The results of this study further enhance our knowledge of a poorly understood form of microbial metabolism and indicate that anaerobic Fe(II) oxidation by D. suillum is significantly different from previously described forms of nitrate-dependent microbial Fe(II) oxidation.  相似文献   

13.
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   

14.
Microbial desulfurization of coal by pyrite oxidizing bacterial enrichment cultures has been studied in air-agitated slurry reactors of 4- and 20-L volumes. Batch experiments showed that inoculation with an active bacterial culture is essential to minimize the lag phase, although a considerable number of pyrite oxidizing bacteria was found on the coal prior to desulfurization. For detailed investigations of kinetics, energy requirements, and technical applicability, a bioreactor equipment consisting of a cascade of eight stages was developed and operated continuously. Microbial desulfurization of coal-monitored by measuring the axial profile of dissolved iron concentration, real and maximum oxygen consumption rates, and cell concentration-at pulp densities to 30% was performed over a period of 200 days without any disturbances concerning the aeration system, fluidization, transport of solids and microbial growth. At a pulp density of 20%, a pyrite conversion of 68% was achieved after the third reactor stage at a total residence time of five days in the first three stages. The kinetics of pyrite degradation were found to be well described by a rate equation of first order in pyrite surface area concentration if the pyrite is directly accessible for microbial attack. Rate constants were determined to 0.48 mg pyrite/(cm(2) day) in the first and to 0.24 mg pyrite/(cm(2) day) in the following reactor stages. Kinetic models taking into account adsorption/desorption as well as growth kinetics failed to describe the observed reaction rates. However, a model treating pyrite degradation and microbial growth kinetics formalistically seems to be applicable when backmixing between the reactor stages can be avoided. The advantage of a multistage reactor in comparison to single-stage equipment was shown by calculation. To obtain a pyrite conversion of 68%, a three-stage reactor would require only 58% of the volume of single-stage equipment.Measurement of oxygen consumption rates proved to provide quickly and easily measurable parameters to observe microbial coal desulfurization in technical scale: the real oxygen consumption rate is correlated to the pyrite oxidation rate and the maximum oxygen consumption rate is correlated to the concentration of viable cells. The Y(o/s) coefficient for the amount of oxygen consumed per mass unit of pyrite oxygen was determined to approximately 0.33 in comparison to 1.0 which can be calculated from stoichiornetry. This could yet not be explained. Chemical leaching experiments as well as sulfur analyses of desulfurized coal samples showed that the microorganisms play the main role in degradation of pyrite from coal and that pyrite oxidation by ferric iron can be neglected.  相似文献   

15.
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   

16.
Shewanella oneidensis strain MR-1 utilizes soluble and insoluble ferric ions as terminal electron acceptors during anaerobic respiration. The components of respiratory metabolism are localized in the membrane fractions which include the outer membrane and cytoplasmic membrane. Many of the biological components that interact with the various iron forms are proposed to be localized in these membrane fractions. To identify the iron-binding proteins acting either as an iron transporter or as a terminal iron reductase, we used metal-catalyzed oxidation reactions. This system catalyzed the oxidation of amino acids in close proximity to the iron binding site. The carbonyl groups formed from this oxidation can then be labeled with fluoresceinamine (FLNH(2)). The peptide harboring the FLNH(2) can then be proteolytically digested, purified by HPLC and then identified by MALDI-TOF tandem MS. A predominant peptide was identified to be part of SO2907 that encodes a putative TonB-dependent receptor. Compared with wild type (wt), the so2907 gene deletion (ΔSO2907) mutant has impaired ability to reduce soluble Fe(III), but retains the same ability to respire oxygen or fumarate as the wt. The ΔSO2907 mutant was also impacted in reduction of insoluble iron. Iron binding assays using isothermal titration calorimetry and fluorescence tryptophan quenching demonstrated that a truncated form of heterologous-expressed SO2907 that contains the Fe(III) binding site, is capable of binding soluble Fe(III) forms with K(d) of approximate 50 μm. To the best of our knowledge, this is the first report of the physiological role of SO2907 in Fe(III) reduction by MR-1.  相似文献   

17.
Pyrite (FeS2) is the most abundant sulfide mineral on Earth and represents a significant reservoir of reduced iron and sulfur both today and in the geologic past. In modern environments, oxidative transformations of pyrite and other metal sulfides play a key role in terrestrial element partitioning with broad impacts to contaminant mobility and the formation of acid mine drainage systems. Although the role of aerobic micro‐organisms in pyrite oxidation under acidic‐pH conditions is well known, to date there is very little known about the capacity for aerobic micro‐organisms to oxidize pyrite at circumneutral pH. Here, we describe two enrichment cultures, obtained from pyrite‐bearing subsurface sediments, that were capable of sustained cell growth linked to pyrite oxidation and sulfate generation at neutral pH. The cultures were dominated by two Rhizobiales species (Bradyrhizobium sp. and Mesorhizobium sp.) and a Ralstonia species. Shotgun metagenomic sequencing and genome reconstruction indicated the presence of Fe and S oxidation pathways in these organisms, and the presence of a complete Calvin–Benson–Bassham CO2 fixation system in the Bradyrhizobium sp. Oxidation of pyrite resulted in thin (30–50 nm) coatings of amorphous Fe(III) oxide on the pyrite surface, with no other secondary Fe or S phases detected by electron microscopy or X‐ray absorption spectroscopy. Rates of microbial pyrite oxidation were approximately one order of magnitude higher than abiotic rates. These results demonstrate the ability of aerobic microbial activity to accelerate pyrite oxidation and expand the potential contribution of micro‐organisms to continental sulfide mineral weathering around the time of the Great Oxidation Event to include neutral‐pH environments. In addition, our findings have direct implications for the geochemistry of modern sedimentary environments, including stimulation of the early stages of acid mine drainage formation and mobilization of pyrite‐associated metals.  相似文献   

18.
Abundance and structure of the communities of neutrophilic lithotrophic iron-oxidizing bacteria (FeOB) inhabiting four low-mineralized ferruginous springs of the Marcial Waters Resort (South Karelia, Russia) and the brackish chalybeate spring of the Staraya Russa Resort (Novgorod region, Russia), were investigated, as well as the physicochemical conditions of these environments. In fresh iron-containing precipitates collected near the spring outlets and within the spring-discharge areas, as well as along the spring watercourses, the numbers of unicellular FeOB enumerated on nutrient media ranged from 105 to 107 cells per 1 mL of sediments irrespective of the initial Fe(II) concentration (11–126 mg L−1). In the spring waters and along the spring watercourses inhabited by iron-oxidizing bacteria, the concentration of dissolved oxygen did not exceed 0.05–0.1 mg L−1. Unicellular FeOB were predominant in three springs, while in the springs with relatively low Fe(II) concentrations (11–22 mg L−1), various morphological forms of Gallionella and uncultured forms of the iron-oxidizing bacterium Toxothrix trichogenes prevailed. In the model experiments with the water samples collected in the ferruginous springs and bogs under controlled conditions, the fractionation of stable iron isotopes and the rate of iron oxidation were found to depend on the oxygen regime and, to a lesser extent, on the initial Fe(II) concentration. The maximum enrichment of the iron oxides formed during the simulation experiments with the light 54Fe isotope was observed during bacterial oxidation under microaerobic conditions at O2 concentrations of 0.1–0.3 mg L−1 and in the cultures of iron-oxidizing bacteria. During the abiogenic oxidation of Fe(II), the extent of stable isotope fractionation was 1.5–2 times lower. Enrichment of Fe(III) oxides with the light 54Fe isotope (3- to 5-fold) was considerably lower at high rates of both the biogenic and abiogenic processes of iron oxidation under aerobic conditions; however, it was more intense during the bacterial processes. Comparison of the rates of enrichment of Fe(III) oxides with the light isotope during the model experiments with pure and enrichment cultures of iron-oxidizing bacteria from the sediments of ferruginous springs and bogs revealed that the biogenic factor plays a key role in the oxidation processes of the iron cycle, as well as in the differentiation of the composition of stable iron isotopes in the studied ecosystems.  相似文献   

19.
Particulate materials (flowers of sulfur, fluorapatite, glass beads, and pyrite) inhibited the growth of Thiobacillus ferrooxidans on ferrous iron at pH 1.5. The degree of inhabitation varied with the type of the particulate and culture conditions. The inhibition caused by pyrite involved a soluble toxic component. The inhibition caused by sulfur particles and glass beads was alleviated by incubation under static culture conditions (i.e., no shaking) but oxidation rates declined because of lack of aeration an increase in the pH above 2 also relieved the inhibition. The nature of inhibition. The nature of inhibition by different particles varied from declined iron oxidation rates to prolonged lag periods depending on the concentration of the test material. Thus, the results indicate that the bacterial activity at the liquid-solid interface is different from that in the bulk liquid phase and is influenced by the proton concentration and by the physiochemical properties of the solids.  相似文献   

20.
Summary Pyrite was microbiologically removed by Thiobacillus ferrooxidans in pure and mixed cultures from German bituminous coal at 10% pulp density with maximum pyrite oxidation rate of 350 mg pyritic S/l per day. However, at pulp densities above 20% bacterial growth and consequently pyrite oxidation were completely prevented both in a conventional airlift reactor and in a stirred-tank reactor. Modifying the airlift reactor by adapting a conical bottom part, bacterial growth and pyrite oxidation could be achieved even at 30% pulp density, resulting in a pyrite removal of more than 90% at a pyrite oxidation rate of 230 mg pyritic S/l per day.Dedicated to Prof. Dr. H. Jüntgen on the occasion of his 60th birthday  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