首页 | 官方网站   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.
The reaction products formed during the leaching of bornite in either ferric chloride or ferric sulfate media depend on the leaching conditions as well as the particle size of the bornite. The extent of dissolution is always more vigorous in the ferric chloride system and increases with increasing temperature in either system. The reaction initially involves the rapid outward diffusion of copper to form slightly nonstoichiometric bornite (Cu5-xFeS4), chalcopyrite, and covellite. The non-stoichiometric bornite is progressively converted to a Cu3FeS4 phase, which varies considerably in its composition, and to covellite. Although the reaction at low temperature terminates at the Cu3FeS4 phase, leaching at higher temperatures results in further dissolution to elemental sulfur and soluble Cu2+ and Fe2+. The leaching ofmassive bornite illustrates the complexities of the leaching reaction more clearly than is observed for the finelypaniculate bornite. In leached massive bornite, a distinct covellite zone appears in the Cu3FeS4 phase; as well, chalcopyrite exsolution lamellae rimmed by a copper sulfide (possibly digenite) appear in the covellite zone, in the Cu3FeS4 phase, and in the nonstoichiometric bornite. The experimental leaching results, especially those involving massive bornite, are generally consistent with the mineralogical trends produced by supergene alteration of bornite ores, but a significant difference is that the Cu3FeS4 phase does not correspond closely to the mineral idaite.  相似文献   

2.
Leaching of natural bornite in a sulfuric acid solution with oxygen as oxidant was investigated using the parameters: temperature, particle size, initial concentration of ferrous, ferric and cupric ions, and using microscopic, X-ray and electronprobe microanalysis to characterize the reaction products. Additionally, stirring rate, pH and PO2 were varied. Dissolution curves for percent copper extracted as a function of time were sigmoidal in shape with three distinct periods of reaction: induction, autocatalytic and post-autocatalytic which levelled off at 28% dissolution of copper. The length of the induction period was not reproducible, causing the dissolution curves to be shifted with respect to time. The dissolution curves in the autocatalytic and post-autocatalytic regions were reproducible, and this property was utilized to treat much of the kinetic data. The iron dissolution curves had four dissolution regions. An initial small but rapid release of iron to solution preceded the three periods just given for copper dissolution. Aside from this initial iron release, the iron and copper dissolution curves were almost identical.Stirring rate had no effect on dissolution of copper above 400 min?1 nor did oxygen flow rate in the range 20–40 cm3/min. Dissolution rate was slightly dependent on oxygen partial pressure for PO2 < 0.67. Hydrogen ion concentration had no effect except that sufficient acid was required to prevent hydrolysis and precipitation of iron salts.The dissolution rate was directly dependent on the reciprocal of particle diameter indicating possible surface chemical reaction control, but the activation energy of 35.9 kJ/mol (8.58 kcal/mol) for the autocatalytic region of copper dissolution is slightly too small for that, though not unreasonable. Initial addition of Fe2+ had a rather complex effect and markedly enhanced dissolution of copper, as also did initial addition of Fe3+. Microscopic analysis showed nuclei of two new phases, covellite and Cu3FeS4, in the induction region. The new phases grow rapidly in the autocatalytic stage, which is controlled by nuclei formation and chemical reaction. The post-autocatalytic region is characterized by complete transformation of bornite into covellite on the particle surfaces and Cu3FeS4 as an internal product with an X-ray spectrum very similar to that of chalcopyrite. The post-autocatalytic region is controlled by autocatalytic growth of newly formed phases. Further reaction beyond the autocatalytic region (percent copper dissolution > 28%) occurs so slowly with oxygen as oxidant that it was not studied.The rate of copper dissolution appears to be controlled by the rate of iron dissolution. Using that and the other experimental evidence a mechanism for reaction is proposed in which iron-deficient bornite, Cu5Fe?S4, is formed on the surface by initial preferential iron dissolution. Labile Cu+ diffuses into this from Cu5Fe?SO4 and unreacted bornite to produce CuS on the surface. Depletion of labile Cu+ ions from Cu5FeS4 produces Cu3FeS4 in the interior of the mineral particles.  相似文献   

3.
Oxidation of stoichiometric iron sulfide was investigated. Rectangular plates of dense FeS were oxidized in an Ar-O2 gas mixture at 1023 to 1123 K. Oxygen partial pressure was varied between 1.01 × 103 and 2.03 × 104 Pa. During the initial five minutes of oxidation, a magnetite layer of about 10 μm in thickness was formed on the surface without the evolution of SO2 gas. Diffusion of iron from the interior of the sulfide to the sulfide/magnetite interface controlled the oxidation rate. Mass transfer through the gaseous boundary layer at the sample surface also affects the oxidation rate at lower oxygen partial pressures. Following this rapid formation of magnetite, the magnetite layer continued to grow for several hours in accordance with the parabolic rate law. Diffusion of iron through the magnetite layer controlled the oxidation rate during this stage. A thin layer of hematite was also formed on the outer surface of magnetite. When the composition of the inner sulfide core reached Fe0.9S, the oxidation proceeded irregularly into the interior of the remaining sulfide. Porous oxide was formed and SO2 gas was evolved. Former Graduate Student  相似文献   

4.
An amount of 80 mg of molten copper matte of a pseudo-ternary Cu2S-FeS-Fe system contained in a slender alumina sample tube was oxidized at 1503 and 1533 K in a mixed O2-Ar gas stream and the oxidation path was followed, comparing the overall rate of oxidation with the gaseous diffusion in the sample tube. The following successive reactions were found to be controlled by gas diffusion. Initially, Fe was oxidized to form FeO. After the melt composition reached a pseudo-ternary Cu2S-FeS-FeO system, FeS was oxidized to form FeO. As the amount of FeO increased, Fe3O4 was also formed and subsequently Cu was produced by the oxidation of Cu2S. In the latter stage, the Cu was oxidized, and the final product under the condition of gas diffusion control was composed of Cu2O, Fe3O4, and CuFeO2. On the other hand, the rate of formation of Fe2O3, CuO, and CuFe2O4 was much slower and they were not formed during the oxidation duration where the overall rate of oxidation was controlled by gas diffusion.  相似文献   

5.
In an acidified ferric chloride solution, bornite leaches in two stages of reaction with the first being relatively much more rapid than the second; the first terminates at 28 pct copper dissolution. The first-stage dissolution reaction is electrochemical and is mixed kinetics-controlled; ferric-ion transfer through the solution boundary layer and reduction on the surface to release Cu2+ into solution are both important in controlling the rate. The concentration of labile Cu+ in the bornite lattice governs the potential of the surface reaction, and, once Cu+ is depleted from the original bornite, stage-I reaction ceases. The solid reaction intermediate formed is Cu3FeS4. Minute subcrystallites formed at the latter part of stage I leach topochemically in stage II. This reaction which commences at 28 pct Cu dissolution is characterized by a change in mechanism at about 40 pct copper dissolution, though the overall chemical equation for reaction is unchanged in stage II; cupric and ferrous ions and sulfur as a solid residue are products of reaction. The region 28 to about 40 pct Cu dissolution is designated as a transition period to stage-II reaction. Reaction rate in this period is interpreted as being controlled by reduction of Fe3+ on active product sulfur surface sites, and hence the reaction rate is controlled by the rate of nucleation and growth of sulfur on the Cu3FeS4 intermediate surfaces. Strain in the Cu3FeS4 crystal lattice is released during this period by diffusion from the lattice of Cu+ remaining from the labile copper initially present in the bornite. After about 40 pct Cu dissolution the rate of reaction is controlled by diffusion through the fully formed sulfur layer in an equiaxial geometrically controlled reaction.  相似文献   

6.
Dense plate of cobalt sulfide was oxidized in a mixed O2-N2 gas stream at 873 to 1123 K. Atomic fraction of sulfur of the sulfide was between 0.505 and 0.525, and the partial pressure of oxygen in the gas stream was varied between 5.05 x 103 and 2.02 x 104 Pa. At lower temperature, cobalt diffused from the interior of sulfide to the surface due to the lower sulfur activity, and a dense oxide layer was formed without the evolution of SO2 gas. The oxidation rate was controlled by the diffusion of cobalt in the sulfide in the initial few minutes, and it was controlled by the diffusion of cobalt through the oxide layer in the subsequent oxidation. At higher temperature, the oxidation of cobalt sulfide proceeded accompanying the evolution of SO2 gas due to the higher sulfur activity, and a porous oxide was formed. The oxidation rate was determined by the mass transfer of oxygen through the gas boundary film and the oxide layer.  相似文献   

7.
《Acta Metallurgica》1988,36(1):167-180
A combination of thermogravimetry, optical microscopy, scanning and transmission electron microscopy, electron probe microanalysis and differential scanning calorimetry has been used to investigate the oxidation behaviour of amorphous Fe40Ni40P14B6. The oxide layer formed on amorphous Fe40Ni40uB6 has a whisker-like α-Fe2O3 structure, which grows very rapidly to build up a thick layer of oxide. Kinetic data indicate that the oxidation of amorphous Fe40Ni40P14B6 obeys a parabolic rate law as long as the alloy remains amorphous and the rate controlling process is diffusion of iron in the amorphous alloy matrix. However, the oxidation rate drops sharply if crystallization of amorphous Fe40Ni40P14B6 takes place during oxidation annealing. Crystalline Fe40Ni40P14B6 also obeys a parabolic rate law but with as much smaller rate constant than the amorphous alloy. The rate controlling process for oxidation of crystalline Fe40Ni40P14B6 is diffusion of iron and nickel in the multiphase oxide layer, which consists of a fine scale mixture of NiO, Fe3O4, Fe2O3 and NiFe2O4, crystals. The difference in oxidation behaviour between amorphous and crystallized Fe40Ni40P14B6 is caused by the different alloy microstructures.  相似文献   

8.
The oxide layer on the surface of the particles of a nitrogen atomized 304L stainless steel powder was quantitatively characterized using X-ray photoelectron spectroscopy (XPS), secondary-ion microprobe (IMP) analysis, selective reduction of surface iron oxide by hydrogen, and chemical analysis for total oxygen, all as a function of particle size. The composition and thickness of the oxide layer do not depend significantly on particle size. A 5.8-nm-thick outer layer made of MnO and Fe2O3 islands covers a 2.4-nm-thick inner layer of Cr2O3. The reaction of the powder with O2 impurity in either H2 or N2 exhibits different kinetics and mechanism in both cases. In H2, a selective oxidation of the alloying elements Mn, Cr, and Si takes place above 400 °C with a parabolic isothermal rate law. In N2, iron also is oxidized, isolated Fe2O3-rich microcrystals form over a Cr2O3-rich uniform underlayer, and the isothermal kinetics are accelerated.  相似文献   

9.
A fluid-dynamics computer model of the flash-converting furnace shaft, which is based on basic principles, is presented. The model is fully three-dimensional and incorporates the transport of momentum, heat, and mass and the reaction kinetics between the gas and particles in a particle-laden turbulent gas jet. The k-ɛ model was used to describe gas-phase turbulence in an Eulerian framework. The particle-cloud model was used to track the particle phase in a Lagrangian framework. The coupling of gas and particle equations was performed through the source terms in the Eulerian gas-phase governing equations. Copper matte particles were represented as Cu2S · yFeS x . Based on experimental observation, the oxidation products were assumed to be Cu2O, CuO, Fe3O4, and SO2. A reaction mechanism involving the external mass transfer of oxygen from the gas to the particle surface and diffusion of the oxygen through the successive layers of Cu2O-Fe3O4 and CuO-Fe3O4 was proposed. The predictions of the computer model were compared with the experimental data collected in a large laboratory furnace. Reasonable agreement between the model predictions and the measurements was obtained in terms of the fractional completion of the oxidation reactions and the sulfur remaining in the reacted particles. The relevance of the computational model for further analysis and optimization of an industrial flash-converting operation is discussed.  相似文献   

10.
Thermal decomposition of the mineral sulfide concentrates derived from froth flotation was studied in the temperature range between 1073 K and 1473 K. The experiments were carried out by heating the mineral sulfide concentrates under inert (argon) atmosphere and, in the presence and absence of carbon. The thermal decomposition of the mineral sulfide concentrates were analyzed by plotting the weight loss of the sample against the reaction time which were obtained from the thermogravimeric analysis (TGA) and characterizing the reacted samples by X-ray diffraction (XRD) and microscopic techniques. The extent and rate of thermal decomposition of the samples were temperature dependent. The high sulfur minerals (CuFeS2 and FeS2) were decomposed to low sulfur and stable phases (Cu1.1Fe1.1S2, Cu5FeS4, and FeS). However, the thermal decomposition of the mineral sulfides were complex because of liquid phase formation at high temperature such that the low sulfur phases appear to have precipitated from the liquid phase during cooling of the sample. Metallic copper was produced in the mineral concentrates containing hydrated and carbonated compounds. It was concluded that the formation of copper cannot occur in the absence of internal oxidation in the sample, when the mineral concentrates are heated under inert atmosphere. The phases were compared with the thermodynamic predictions.  相似文献   

11.
《Acta Metallurgica》1988,36(8):2293-2305
A combination of thermogravimetry, scanning and transmission electron microscopy, electron probe microanalysis and differential scanning calorimetry has been used to investigate the oxidation kinetics, and oxide morphology, structure and composition in amorphous and crystalline Fe78Si9B13 alloys. Kinetic data indicate that the oxidation reactions of both amorphous and crystalline Fe78Si9B13 obey a parabolic rate law over the temperature range 300 to 450°C with activation energies of 120 and 86 kJ/mol respectively, indicating that grain boundary diffusion is probably the rate controlling process. The parabolic rate constant for oxidation of crystalline Fe78Si9B13 is consistently higher than for amorphous Fe78Si9B13 over the temperature range 300–450°C, so that the amorphous alloy always shows a better oxidation resistance. Electron microscopy and electron probe microanalysis show that the oxide scales formed on both amorphous and crystalline Fe78Si9B13 consist of SiO2, Fe3O4 and Fe2O3, but the detailed microstructure and compositions are different. The oxide scale formed on amorphous Fe78Si9B13 contains more SiO2 and has a small particle size, while the oxide scale formed on crystalline Fe78Si9B13 contains more Fe3O4 and consists of larger particles. The difference in oxide growth between amorphous and crystalline Fe78Si9B13 is caused by the difference in alloy microstructure.  相似文献   

12.
Austenitic Fe–18 wt% Mn–0.6 wt% C steels were oxidized at 1273, 1373, and 1473 K for up to 2 h in either atmospheric air or an 85%N2–10%CO2–5%O2 gas mixture. The alloys oxidized faster in air than in the mixed gas, but the morphology and composition of the oxide scale formed were similar in both atmospheres. The scales that consisted primarily of FeO, Fe2O3, and MnFe2O4 were highly susceptible to cracking and spallation due to the severe oxidation condition. Since Mn was consumed to form MnFe2O4, the original γ‐matrix changed to an α‐matrix in the subscale area, in which Mn‐rich internal oxide precipitates formed locally.  相似文献   

13.
The formation of zinc ferrite (ZnFe2O4) during the roasting of iron-bearing zinc concentrates requires substantial additional processing to recover the zinc from this compound by leaching and to eliminate the iron from the leachate. The phase changes that occur in the particles of a typical industrial zinc sulfide concentrate during roasting in a fluidized bed at 1223 K were investigated by the use of light microscopy, electron microprobe analysis, and SEM with EDS. The processes which the iron undergoes during its eventual transformation into ferrite have been clarified by examination of the phases and the morphology of partially roasted marmatitic sphalerite particles (Zn, Fe)S, and by reference to the known phase equilibria involved in the Zn-Fe-S-0 system. The oxidation of ironbearing sphalerite occurs in three stages. The first involves the selective diffusion of most of the iron to the particle surface resulting in the formation of an iron oxide shell enclosing a largely unreacted zinc sulfide kernel. In the second stage, this kernel is oxidized to form a solid solution of zinc oxide and iron oxide. The iron is initially present in the ferrous state but, with the disappearance of the sulfide kernel, is oxidized to ferric iron. In the final stage, this dissolved iron oxide and the iron oxide shell react with the surrounding zinc oxide to form the refractory spinel zinc ferrite.  相似文献   

14.
Chalcopyrite is reduced by solutions of copper(I) sulfate and copper(I) chloride to chalcocite (Cu2S) and bornite (Cu5FeS4) whilst the iron reports to the solution. Factors which affect the rate and efficiency of reduction are examined. The reaction is rapid on fresh surfaces of chalcopyrite but slows markedly as a film of chalcocite or bornite forms. The reduction in the presence of copper metal goes to completion and gives a material which is more readily leached by oxidising agents than is chalcopyrite. Thus 99% of the copper in the reduced chalcopyrite is leached when copper(II) sulfate in aqueous acetonitrile is the oxidising agent, whereas only 30% of the copper is leached from pure chalcopyrite under similar conditions. Concentrated solutions of copper(I) salts are less effective in reducing CuFeS2 in a heterogeneous solid-liquid reaction than is copper metal in a “galvanic” solid-solid reaction. Solutions of copper(II) sulfate plus concentrated copper(I) sulfate in dilute acetonitrile (4 M) containing copper sheets are an effective reductant for chalcopyrite.  相似文献   

15.
The oxidation of FeS powders in flowing dry air was investigated over the temperature range of 648 to 923 K. Thermodynamic calculations and experimental observations showed that the initial stages of oxidation are characterized by the formation of FeS2 and Fe3O4 or Fe2O3. Sub-sequently, the oxidation process goes through a formation and eventual oxidation of Fe2(SO4)3 to Fe2O3. The kinetics of oxidation of FeS are believed to be controlled by this last step, whose activation energy, 192 kJ-mol-1, agrees reasonably well with a corresponding value measured for the oxidation of Fe2(SO4)3, 219 kJ.mol-1. Thermodynamic calculations of phase stability diagrams in the Fe-O-S system are presented in support of the proposed oxidation mechanism.  相似文献   

16.
Using the methods of X-ray photoelectron spectroscopy (XPES), X-ray phase analysis, gamma-resonance (Mössbauer) spectroscopy, thermogravimetry of solid samples and atomic absorption, and atomic emission spectroscopy for solutions, the forms of occurrence of nonferrous metals in oxyhydroxide sediments (nitration hydroxides, NH) from nitration of refining of rhodium, iridium, and ruthenium are investigated. According to the XPES data, nitration hydroxides involve oxidized and unoxidized (selenium) transition metals, namely, Fe(III), Sn(II), Te(IV), Cu(I), Cu(II), Ni(II), Pb(II), As(III), Se(0), and Se(IV), as well as nonmetals O, Cl, and C. All nitration hydroxides are formed mainly by an X-ray amorphous phase (>90%). During heating of nitration hydroxides up to t = 900°C in an inert atmosphere, dehydration of hydroxides and crystallization of certain phases take place, but the X-ray amorphous phase amounts to more than 50%. Purely oxide phases or mixed oxide phases Fe3O4, SnO2, CuIFeIIIO2, and Pb2O3 (NH-1); Ni0.4Fe2.6O4, CuIFeIIIO2, and CuNiSnO4 (NH-2); and Fe3O4, Fe2O3, and SnO2 (NH-3) are formed in this case.  相似文献   

17.
A twin balance apparatus capable of continuously weighing both the solid and liquid phase was constructed to measure the magnetite-matte reaction. The rate of magnetite consumption during the reaction with iron sulfide increased from 2.5 to 10.8 g Fe3O4/min with increasing temperature from 1473 to 1623 K. In the magnetite-(Fe-S-O) melt reactions, the dissolution rate decreased from 5.4 to 1.5 g Fe3O4/min with increasing oxygen content in the bulk liquid phase. The rate approached zero when the bulk phase contained approximately 32 at. pct O which was close to the magnetite saturation limit. The rate of reaction between magnetite and liquids of the FeS-Cu2S binary system decreased from 5.4 to 0.2 g Fe3O4/min with increasing Cu2S content from 0 to 60 mol pct Cu2S. Examination of the reaction rate data of the Fe-S-O system in conjunction with a magnetite pellet profile study and oxygen analyses in the matte showed that the dissolution mechanism was one of diffusion enhanced by natural convection. The results of reacting magnetite with FeS-Cu2S melts suggested that the same mechanism operated. The industrial significance of this investigation is discussed briefly in relation, to the problem of the presence of solid magnetite in copper smelting processes. Formerly with the Department of Metallurgy and Materials Science, University of Toronto  相似文献   

18.
Compound layers composed of γ′-Fe4N1-x on α-Fe substrates were oxidized in oxygen-containing gas atmospheres at 603 K. The microstructural changes were analyzed applying X-ray diffraction, dilatometry, and light, scanning electron, and transmission electron microscopy. Compositional changes were traced, in particular, with scanning Auger electron spectroscopy. Dual-phase, Fe3O4 (magnetite) and α-Fe2O3 (hematite), oxide layers formed at the compound layer surface. At the interface between the oxide and nitride layers, ε-Fe2N1=x nucleated because of a local nitrogen enrichment caused by the conversion of γ′ nitride into oxide. The volume of each of the three phases formed (hematite, magnetite, and ε nitride) increased with the square root of oxidation time, indicating solid-state diffusion-controlled layer growth. No evidence was obtained for the existence of γ′ or e oxynitrides. Within the γ′ layer, ferrite precipitated during oxidation as a consequence of an iron supersaturation of γ′ nitride due to a production temperature higher than the oxidation temperature. During this iron precipitation, stress relaxation occurred, as was concluded from X-ray diffractometric and dilatometric analyses. The stress relaxation was rate-controlled by nitrogen diffusion in γ′ nitride. The present findings were also used to explain microstructural changes during (commercial) oxidation of ε-Fe2(N,C)1-x compound layers. Formerly Graduate Student at the Laboratory of Metallurgy, Delft University of Technology  相似文献   

19.
《Hydrometallurgy》2001,59(2-3):177-185
The dissolution of metal sulfides is controlled by their solubility product and thus, the [H+] concentration of the solution, and further enhanced by several chemical mechanisms which lead to a disruption of sulfide chemical bonds. They include extraction of electrons and bond breaking by [Fe3+], extraction of sulfur by polysulfide and iron complexes forming reactants [Y+] and electrochemical dissolution by polarization of the sulfide [high Fe3+ concentration]. All these mechanisms have been exploited by sulfide and iron-oxidizing bacteria. Basically, the bacterial action is a catalytic one during which [H+], [Fe3+] and [Y+] are breaking chemical bonds and are recycled by the bacterial metabolism. While the cyclic bacterial oxidative action via [H+] and [Fe3+] can be called indirect, bacteria had difficulties harvesting chemical energy from an abundant sulfide such as FeS2, the electron exchange properties of which are governed by coordination chemical mechanisms (extraction of electrons does not lead to a disruption of chemical bonds but to an increase of the oxidation state of interfacial iron). Here, bacteria have evolved alternative strategies which require an extracellular polymeric layer for appropriately conditioned contact with the sulfide. Thiobacillus ferrooxidans cycles [Y+] across such a layer to disrupt FeS2 and Leptospirillum ferrooxidans accumulates [Fe3+] in it to depolarize FeS2 to a potential where electrochemical oxidation to sulfate occurs. Corrosion pits and high resolution electron microscopy leave no doubt that these mechanisms are strictly localized and depend on specific conditions which bacteria create. Nevertheless, they cannot be called ‘direct’ because the definition would require an enzymatic interaction between the bacterial membrane and the cell. Therefore, the term ‘contact’ leaching is proposed for this situation. In practice, multiple patterns of bacterial leaching coexist, including indirect leaching, contact leaching and a recently discovered cooperative (symbiotic) leaching where ‘contact’ leaching bacteria are feeding so wastefully that soluble and particulate sulfide species are supplied to bacteria in the surrounding electrolyte.  相似文献   

20.
Carbonyl iron ultrafine powder with sizes ranging from 5 to 30 nm was characterized by conventional transmission electron microscopy (CTEM), electron energy loss spectroscopy (EELS), high-resolution electron microscopy (HREM), and microdiffraction. It was found that the carbonyl iron ultrafine powder is composed of particles of a-Fe enclosed in about 5-nm-thick skins of polycrystalline Fe3O4. Some of the particles are completely oxidized into polycrystalline γ-Fe2O3 with grains 5 to 10 nm in size, which form about a 5-nm-thick shell with a hollow core in the center. It was observed that the oxidation of Fe3O4 can be enhanced by dispersion.  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-23

京公网安备 11010802026262号