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.
Available cultures of Thiobacillus ferrooxidans were found to be contaminated with bacteria very similar to Thiobacillus acidophilus. The experiments described were performed with a homogeneous culture of Thiobacillus ferrooxidans.Pyrite (FeS2) was oxidized by Thiobacillus ferrooxidans grown on iron (Fe2+), elemental sulphur (So) or FeS2.Evidence for the direct utilization of the sulphur moiety of pyrite by Thiobacillus ferrooxidans was derived from the following observations: a. Known inhibitors of Fe2+ and So oxidation, NaN3 and NEM, respectively, partially abolished FeS2 oxidation. b. A b-type cytochrome was detectable in FeS2-and So-grown cells but not in Fe2+-grown cells. c. FeS2 and So reduced b-type cytochromes in whole cells grown on So. d. CO2 fixation at pH 4.0 per mole of oxygen consumed was the highest with So, lowest with Fe2+ and medium with FeS2 as substrate. e. Bacterial Fe2+ oxidation was found to be negligible at pH 5.0 whereas both FeS2 and So oxidation was still appreciable above this pH. f. Separation of pyrite and bacteria by means of a dialysis bag caused a pronounced drop of the oxidation rate which was similar to the reduction of pyrite oxidation by NEM; indirect oxidation of the sulphur moiety by Fe3+ was not affected by separation of pyrite and bacteria.Bacterial oxidation and utilization of the sulphur moiety of pyrite were relatively more important with increasing pH.  相似文献   

2.
Iron has a central role in bioleaching and biooxidation processes. Fe2+ produced in the dissolution of sulfidic minerals is re-oxidized to Fe3+ mostly by biological action in acid bioleaching processes. To control the concentration of iron in solution, it is important to precipitate the excess as part of the process circuit. In this study, a bioprocess was developed based on a fluidized-bed reactor (FBR) for Fe2+ oxidation coupled with a gravity settler for precipitative removal of ferric iron. Biological iron oxidation and partial removal of iron by precipitation from a barren heap leaching solution was optimized in relation to the performance and retention time (τFBR) of the FBR. The biofilm in the FBR was dominated by Leptospirillum ferriphilum and “Ferromicrobium acidiphilum.” The FBR was operated at pH 2.0 ± 0.2 and at 37 °C. The feed was a barren leach solution following metal recovery, with all iron in the ferrous form. 98–99% of the Fe2+ in the barren heap leaching solution was oxidized in the FBR at loading rates below 10 g Fe2+/L h (τFBR of 1 h). The optimal performance with the oxidation rate of 8.2 g Fe2+/L h was achieved at τFBR of 1 h. Below the τFBR of 1 h the oxygen mass transfer from air to liquid limited the iron oxidation rate. The precipitation of ferric iron ranged from 5% to 40%. The concurrent Fe2+ oxidation and partial precipitative iron removal was maximized at τFBR of 1.5 h, with Fe2+ oxidation rate of 5.1 g Fe2+/L h and Fe3+ precipitation rate of 25 mg Fe3+/L h, which corresponded to 37% iron removal. The precipitates had good settling properties as indicated by the sludge volume indices of 3–15 mL/g but this step needs additional characterization of the properties of the solids and optimization to maximize the precipitation and to manage sludge disposal.  相似文献   

3.
As an alternative to commercial Ni‐ and Co‐based intercalation‐type cathode materials, conversion‐type metal fluoride (MFx) cathodes are attracting more interest due to their promises to increase cell‐level energy density when coupled with lithium (Li) or silicon (Si)‐based anodes. Among metal fluorides, iron fluorides (FeF2 and FeF3) are regarded as some of the most promising candidates due to their high capacity, moderately high potential and the very low cost of Fe. In this study, the impacts of electrolyte composition on the performance and stability of nanostructured FeF2 cathodes are systematically investigated. Dramatic impacts of Li salt composition, Li salt concentration, solvent composition, and cycling potential range on the cathode's most critical performance parameters—stability, capacity, rate, and voltage hysteresis are discovered. In contrast to previous beliefs, it is observed that even if the Fe2+ cation dissolution could be avoided, the dissolution of F? anions may still negatively affect cathode performance. Formation of the more favorable cathode solid electrolyte interface (CEI) is found to minimize both processes.  相似文献   

4.
5.
Bioleaching of sulfide minerals at neutral pH has been rarely reported. In this study, a bacterium, Bacillus megaterium QM B1551, was isolated from Jinchuan sulfide tailings and used to leach a complex sulfide flotation concentrate for the extraction of Co2+, Ni2+ and Cu2+ at near neutral pH. A total of 38.2% Co, 44.7% Ni and 3.6% Cu were extracted from the sulfide concentrate in 5 days with an initial pH of 6. An enhanced Co2+, Ni2+ and Cu2+ extraction extent was achieved by first bioleaching the concentrate with Bacillus megaterium QM B1551 at 35°C and then followed by chemical leaching with 4 M sulfuric acid at 90°C. As a result, a total of 60.7% Co2+, 76.3% Ni2+ and 39.8% Cu2+ were extracted. On an industrial scale, the profits from the metal recovery by such a combined leaching procedure are optimum if considering the cost-benefit ratio.  相似文献   

6.
A number of methods for denitrification were studied including Electro-Fenton method, hydrogenotrophic as well as innovative Fe2+ mediated denitrification and their technical feasibility in terms of changes in TOC and nitrate concentrations, effect of different Fenton’s reagent dosage, current and the effect of the pH was investigated. This study was carried out using tailor made electrodialytic reactor. It was found that the highest TOC removal was achieved at pH 2.2 and 2.4 (77.1% and 97.8%, respectively) at the anode and the lowest accumulation of 33% at pH of 6.2 at the cathode. The highest TOC removal in terms of using different H2O2 concentrations was achieved at 40 mM reaching as high as 97.3%. Regardless experimental strategy, initially nitrates migrated towards the cathode due to the strong hydraulic gradient under the applied electric current. During the course of experiments, nitrates were transported towards the anode where their concentration decreased. The highest nitrate removal was achieved at 0.12 mA cm?2 electric current density (94.8%) at the anode and a complete removal at the cathode. Hydrogenotrophic denitrification was the highest reaching 92.5%, however, when Fe2+ ions as electron donor was used for the destruction of nitrates, only 66.6% removal was achieved. Denitrification using only Fe2+ ions was a factor 1.4 less than using electrically generated hydrogen or a Fenton’s reagent.  相似文献   

7.
An extracellular acid phytase was purified to homogeneity from the culture supernatant of the Saccharomyces cerevisiae CY strain by ultrafiltration, DEAE-Sepharose column chromatography, and Sephacryl S-300 gel filtration. The molecular weight of the purified enzyme was estimated to be 630 kDa by gel filtration. Removing the sugar chain by endoglycosidase H digestion revealed that the molecular mass of the protein decreased to 446 kDa by gel filtration and gave a band of 55 kDa by SDS-PAGE. The purified enzyme was most active at pH 3.6 and 40 °C and was fairly stable from pH 2.5 to 5.0. The phytase displayed broad substrate specificity and had a Km value of 0.66 mM (sodium phytate, pH 3.6, 40 °C). The phytase activity was completely inhibited by Fe3+ and Hg2+, and strongly inhibited (maximum of 91%) by Ba2+, Co2+, Cu+, Cu2+, Fe2+, Mg2+, and Sn2+ at 5 mM concentrations.  相似文献   

8.
Li‐rich electrode materials of the family x Li2MnO3·(1?x )LiNia Cob Mnc O2 (a + b + c = 1) suffer a voltage fade upon cycling that limits their utilization in commercial batteries despite their extremely high discharge capacity, ≈250 mA h g?1. Li‐rich, 0.35Li2MnO3·0.65LiNi0.35Mn0.45Co0.20O2, is exposed to NH3 at 400 °C, producing materials with improved characteristics: enhanced electrode capacity and a limited average voltage fade during 100 cycles in half cells versus Li. Three main changes caused by NH3 treatment are established. First, a general bulk reduction of Co and Mn is observed via X‐ray photoelectron spectroscopy and X‐ray absorption near edge structure. Next, a structural rearrangement lowers the coordination number of Co? O and Mn? O bonds, as well as formation of a surface spinel‐like structure. Additionally, Li+ removal from the bulk causes the formation of surface LiOH, Li2CO3, and Li2O. These structural and surface changes can enhance the voltage and capacity stability of the Li‐rich material electrodes after moderate NH3 treatment times of 1–2 h.  相似文献   

9.
Lithium metal batteries (LMBs) combining a Li metal anode with a transition metal (TM) cathode can achieve higher practical energy densities (Wh L?1) than Li/S or Li/O2 cells. Research for improving the electrochemical behavior of the Li metal anode by, for example, modifying the liquid electrolyte is often conducted in symmetrical Li/Li or Li/Cu cells. This study now demonstrates the influence of the TM cathode on the Li metal anode, thus full cell behavior is analyzed in a way not considered so far in research with LMBs. Therefore, the deposition/dissolution behavior of Li metal and the resulting morphology is investigated with three different cathode materials (LiNi0.5Mn1.5O4, LiNi0.6Mn0.2Co0.2O2, and LiFePO4) by post mortem analysis with a scanning electron microscope. The observed large differences of the Li metal morphology are ascribed to the dissolution and crossover of TMs found deposited on Li metal and in the electrolyte by X‐ray photoelectron spectroscopy, energy‐dispersive X‐ray spectroscopy, and total reflection X‐ray fluorescence analysis. To support this correlation, the TM dissolution is simulated by adding Mn salt to the electrolyte. This study offers new insights into the cross talk between the Li metal anodes and TM cathodes, which is essential, when investigating Li metal electrodes for LMB full cells.  相似文献   

10.
Massive pyrite was shown to produce soluble iron, hydrogen, and sulfate ions on exposure to air and water. The rate of this process was directly proportional to the surface area of the mineral; it was unaffected by a drop in the pH and the presence of the ferrous and sulfate ions formed. Cupic ion had no effect but ferric ion accelerated pyrite degradation until all the ferric ion was consumed, in accordance with FeS2 + 2Fe3+ —>‐3Fe2+ + 2S°. Thiobacillus ferrooxidans increased pyrite degradation considerably; the presence of Thiobacillus thiooxidans had no influence on pyrite degradation.  相似文献   

11.
Rate equations and kinetic parameters were obtained for various reactions involved in the bacterial oxidation of pyrite. The rate constants were 3.5 μM Fe2+ per min per FeS2 percent pulp density for the spontaneous pyrite dissolution, 10 μM Fe2+ per min per mM Fe3+ for the indirect leaching with Fe3+, 90 μM O2 per min per mg of wet cells per ml for the Thiobacillus ferrooxidans oxidation of washed pyrite, and 250 μM O2 per min per mg of wet cells per ml for the T. ferrooxidans oxidation of unwashed pyrite. The Km values for pyrite concentration were similar and were 1.9, 2.5, and 2.75% pulp density for indirect leaching, washed pyrite oxidation by T. ferrooxidans, and unwashed pyrite oxidation by T. ferrooxidans, respectively. The last reaction was competitively inhibited by increasing concentrations of cells, with a Ki value of 0.13 mg of wet cells per ml. T. ferrooxidans cells also increased the rate of Fe2+ production from Fe3+ plus pyrite.  相似文献   

12.
Fe3+-EDTA chelates react with the superoxide radical at physiological pH values (k = 1.3 × 106M?1 s?1 at pH 7 but is lower at more alkaline pH values) but do not appear to catalyze O2? dismutation at a significant rate. Complexes of Fe3+ with desferrioxamine, bathophenanthroline, or diethylenetriaminepentaacetic acid react much more slowly, if at all. Fe2+ complexes of EDTA, ATP, and diethylenetriaminepentaacetic acid also react with O2? at alkaline pH values. The significance of these reactions in the mechanism of the “iron-catalyzed Haber-Weiss reaction” is discussed.  相似文献   

13.
Lithium–sulfur (Li–S) batteries have great promise to support the next‐generation energy storage if their sluggish redox kinetics and polysulfide shuttling can be addressed. The rational design of sulfur electrodes plays key roles in tacking these problems and achieving high‐efficiency sulfur electrochemistry. Herein, a synergetic defect and architecture engineering strategy to design highly disordered spinel Ni–Co oxide double‐shelled microspheres (NCO‐HS), which consist of defective spinel NiCo2O4–x (x = 0.9 if all nickel is Ni2+ and cobalt is Co2.13+), as the multifunctional sulfur host material is reported. The in situ constructed cation and anion defects endow the NCO‐HS with significantly enhanced electronic conductivity and superior polysulfide adsorbability. Meanwhile, the delicate nanoconstruction offers abundant active interfaces and reduced ion diffusion pathways for efficient Li–S chemistry. Attributed to these synergistic features, the sulfur composite electrode achieves excellent rate performance up to 5 C, remarkable cycling stability over 800 cycles and good areal capacity of 6.3 mAh cm?2 under high sulfur loading. This proposed strategy based on synergy engineering could also inform material engineering in related energy storage and conversion fields.  相似文献   

14.
Using degenerate polymerase chain reaction (PCR) and thermal asymmetric interlaced PCR, a 1,347-bp full-length complementary DNA fragment encompassing the gene man5A, which encodes a 429-amino acid β-mannanase with a calculated mass of 46.8 kDa, was cloned from acidophilic Bispora sp. MEY-1. The deduced amino acid sequence (catalytic domain) displayed highest identity (54.1%) with the Emericella nidulans endo-β-1,4-d-mannanase, a member of the glycoside hydrolase family 5. Recombinant MAN5A was overexpressed in Pichia pastoris, and its activity in the culture medium reached 500 U ml−1. The enzyme was acidophilic, with highest activity at pH 1.0–1.5, lower than any known mannanases, and optimal temperature for activity was 65°C. MAN5A had good pH adaptability, excellent thermal and pH stability, and high resistance to both pepsin and trypsin. The specific activity, K m, and V max for locust bean gum substrate was 3,373 U mg−1, 1.56 mg ml−1, and 6,587.6 μmol min−1 mg−1, respectively. The enzymatic activity was not significantly affected by ions such as Ca2+, Cr3+, Co2+, Zn2+, Na+, K+, and Mg2+ and enhanced by Ni2+, Fe3+, Mn2+ and Ag+. These favorable properties make MAN5A a potential candidate for use in various industrial applications.  相似文献   

15.
《Process Biochemistry》2014,49(7):1196-1204
Laccase from a tree legume, Leucaena leucocephala, was purified to homogeneity using a quick two-step procedure: alginate bead entrapment and celite adsorption chromatography. Laccase was purified 110.6-fold with an overall recovery of 51.0% and a specific activity of 58.5 units/mg. The purified laccase was found to be a heterodimer (∼220 kDa), containing two subunits of 100 and 120 kDa. The affinity of laccase was found to be highest for catechol and lowest for hydroquinone, however, highest Kcat and Kcat/Km were obtained for hydroquinone. Purified laccase exhibited pH and temperature optima of 7.0 and 80 °C, respectively. Mn2+, Cd2+, Fe2+, Cu2+ and Na+ activated laccase while Ca2+ treatment increased laccase activity up to 3 mM, beyond which it inhibited laccase. Co2+, Hg2+, DTT, SDS and EDTA showed an inhibition of laccase activity. The Leucaena laccase was found to be fairly tolerant to organic solvents; upon exposure for 1 h individually to 50% (v/v) each of ethanol, DMF, DMSO and benzene, more than 50% of the activity was retained, while in the presence of 50% (v/v) each of methanol, isopropanol and chloroform, a 40% residual activity was observed. The purified laccase efficiently decolorized synthetic dyes such as indigocarmine and congo red in the absence of any redox mediator.  相似文献   

16.
Oxidative reactions can result in the formation of electronically excited species that undergo radiative decay depending on electronic transition from the excited state to the ground state with subsequent ultra‐weak photon emission (UPE). We investigated the UPE from the Fe2+–EDTA (ethylenediaminetetraacetic acid)–AA (ascorbic acid)–H2O2 (hydrogen peroxide) system with a multitube luminometer (Peltier‐cooled photon counter, spectral range 380–630 nm). The UPE, of 92.6 μmol/L Fe2+, 185.2 μmol/L EDTA, 472 μmol/L AA, 2.6 mmol/L H2O2, reached 1217 ± 118 relative light units during 2 min measurement and was about two times higher (P < 0.001) than the UPE of incomplete systems (Fe2+–AA–H2O2, Fe2+–EDTA–H2O2, AA–H2O2) and medium alone. Substitution of Fe2+ with Cr2+, Co2+, Mn2+ or Cu2+ as well as of EDTA with EGTA (ethylene glycol‐bis(β‐aminoethyl ether)‐N,N,N′,N′‐tetraacetic acid) or citrate powerfully inhibited UPE. Experiments with scavengers of reactive oxygen species (dimethyl sulfoxide, mannitol, sodium azide, superoxide dismutase) revealed the dependence of UPE only on hydroxyl radicals. Dimethyl sulfoxide at the concentration of 0.74 mmol/L inhibited UPE by 79 ± 4%. Plant phenolics (ferulic, chlorogenic and caffec acids) at the concentration of 870 μmol/L strongly enhanced UPE by 5‐, 13.9‐ and 46.8‐times (P < 0.001), respectively. It is suggested that augmentation of UPE from Fe2+–EDTA–AA–H2O2 system can be applied for detection of these phytochemicals.  相似文献   

17.
Anaerobic microbial dissolution of several crystalline, water-insoluble forms of metal oxides commonly associated with the waste from energy production was investigated. An anaerobic N-fixing Clostridium sp. with an acetic, butyric, and lactic acid fermentation pattern, isolated from coal-cleaning waste, solubilized Fe2O3 and MnO2 by direct enzymatic reduction; CdO, CuO, PbO, and ZnO were solubilized by indirect action due to the production of metabolites and the lowering of the pH of the growth medium. Extracellular heat-labile components of the cell-free spent medium obtained from cultures without oxide solubilized a significant amount of Fe2O3 (1.7 μmol); however, direct contact with the bacterial cells resulted in the complete dissolution (4.8 μmol) of the oxide. Under identical conditions, the cell-free spent medium solubilized only a small amount of MnO2 (0.07 μmol), whereas 2.3 μmol of the oxide was solubilized by direct bacterial contact. Reduction of Fe2O3 and MnO2 by Clostridium sp. proceeds at different rates and, possibly, by different enzymatic systems. Fe(III) and Mn(IV) oxides appear to be used as sinks for excess electrons generated from glucose fermentation, since there is no apparent increase in growth of the bacterium concomitant with the reduction of the oxides. Dialysis bag experiments with Co2O3 indicate that there is a slight dissolution of Co (0.16 μmol) followed by precipitation or biosorption. Although Mn2O3, Ni2O3, and PbO2 may undergo reductive dissolution from a higher to a lower oxidation state, dissolution by direct or indirect action was not observed. Also, Cr2O3 and NiO were not solubilized by direct or indirect action. Significant amounts of solubilized Cd, Cu, and Pb were immobilized by the bacterial biomass, and the addition of Cu2+ inhibited the growth of the bacterium.  相似文献   

18.
S.G. Lu  C. Tang  Z. Rengel 《Plant and Soil》2004,264(1-2):231-245
The combination effects of waterlogging and salinity on redox potential (Eh), pH, electric conductivity (EC), water-soluble cations (NH4 +, K+, Na+, Ca2+, Mg2+, Fe2+, and Mn2+) and water-dispersible clay (WDC) were studied in six soils collected near salt lakes in western Australia. The soils with various salinity levels were incubated under a waterlogged condition at 30 °C for 12 weeks. The Eh, pH, EC, and cations of soil solutions were monitored over the waterlogged period. The Eh values generally dropped to the lowest point within 12 days of waterlogging, then increased slightly, and reached equilibrium after 4 weeks of waterlogging. Increasing salinity levels increased soil Eh. While waterlogging increased soil pH in the first 3–4 weeks, increasing salinity level decreased soil pH during the entire waterlogging period. Waterlogging increased the EC values in the first 2 weeks, partly due to dissolution of insoluble salts. The concentrations of water-soluble NH4 + were significantly increased with salinity level and waterlogging, and reached maximum values at week 2, and then declined to the initial level. Waterlogging and salinity increased the concentrations of water-soluble K+, Ca2+, Mg2+, Fe2+, and Mn2+ ions, but the magnitudes of changes were greatly affected by soil properties. Increases in water-soluble K+, Ca2+ and Mg2+ were attributed to increased solubility of insoluble salts, and increased competition for the adsorption sites of the soil exchange complex due to elevated concentrations of Na+, Fe2+ and Mn2+. Increases in water-soluble Fe2+ and Mn2+ induced by waterlogging were attributed to the dissolution of Fe and Mn oxides under reduced conditions. Waterlogging increased, but salinity decreased, the amounts of water-dispersible clay in the soils of low EC value. The higher salinity level can counteract the adverse effect of waterlogging on clay flocculation.  相似文献   

19.
The objective of the present communication is to describe the role played by combinations between diethydithiocarbamate (DDC) and divalent metals in hemolysis of human RBC. RBC which had been treated with DDC (10–50 μM) were moderately hemolyzed (about 50%) upon the addition of subtoxic amounts of Cu2+ (50 μM). However, a much stronger and a faster hemolysis occurred either if mixtures of RBC-DDC were immediately treated either by Co2+ (50 μM) or by a premixture of Cu2+ and Co2+ (Cu:Co) (50 μM).

While Fe2+ and Ni2+, at 50 μM, initiated 30–50% hemolysis when combined with DDC (50 μM), on a molar basis, Cd2+ was at least 50 fold more efficient than any of the other metals in the initiation of hemolysis by DDC. On the other hand, neither Mn2+ nor Zn2+, had any hemolysis-initiating effects. Co2+ was the only metal which totally blocked hemolysis if added to DDC prior to the addition of the other metals.

Hemolysis by mixtures of DDC + (Cu:Co) was strongly inhibited by anaerobiosis (flushing with nitrogen gas), by the reducing agents glutathione, N-acetyl cysteine, mercaptosuccinate, ascorbate, TEMPO, and α-tocopherol, by the PLA2 inhibitor bromophenacylbromide (BrPACBr), by tetracycline as well as by phosphatidyl choline, cholesterol and by trypan blue. However, TEMPO, BrPACBr and PC were the only agents which inhibited hemolysis induced by DDC:Cd2+ complexes.

On the other hand, none of the classical scavengers of reactive oxygen species (ROS) employed e.g dimethylthiourea, catalase, histidine, mannitol, sodium benzoate, nor the metal chelators desferal and phenanthroline, had any appreciable inhibitory effects on hemolysis induced by DDC + (Cu:Co).

DDC oxidized by H2O2 lost its capacity to act in concert either with Cu2+ or with Cd2+ to hemolyze RBC.

While either heating RBC to temperatures greater than 37°C or exposure of the cells to glucose-oxidase-generated peroxide diminished their susceptibility to hemolysis, exposure to the peroxyl radical from AAPH, enhanced hemolysis by DDC + (Cu:Co).

The cyclovoltammetry patterns of DDC were drastically changed either by Cu2+, Co2+ or by Cd2+ suggesting a strong interaction of the metals with DDC. Also, while the absorbance spectrum of DDC at 280 nm was decreased by 50% either by Co2+, Cd2+ or by H2O2, a 90% reduction in absorbance occurred if DDC + H2O2 mixtures were treated either by Cu2+ or by Co2+, but not by Cd2+.

Taken together, it is suggested that DDC-metal chelates can induce hemolysis by affecting the stability and the integrity of the RBC membrane, and possibly also of the cytoskeleton and the role played by reducing agents as inhibitors might be related to their ability to deplete oxygen which is also supported by the inhibitory effects of anaeobiosis.  相似文献   

20.
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