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 46 毫秒
1.
The precipitation of hematite from ferric chloride media at temperatures <100 °C and at ambient pressure was studied as part of a program to recover a marketable iron product from metallurgical processing streams or effluents. Hematite (Fe2O3) can be formed in preference to ferric oxyhydroxides (e.g., β-FeO·OH) at temperatures as low as 60 °C by controlling the precipitation conditions, especially seeding. The hematite product typically contains >66 pct Fe and <1 pct Cl, and its composition does not change appreciably on repeated recycling. The amount of product formed increases significantly with increasing FeCl3 concentrations to ∼0.2 M FeCl3, but nearly constant product yields are obtained thereafter; the precipitates consist only of hematite, provided that an adequate amount of seed is present. The contamination with Zn, Ca, and Na is <0.1 pct, even for high concentrations of dissolved ZnCl2, CaCl2, or NaCl. The extent of the precipitation reaction depends principally on the temperature and the free-acid concentration; accordingly, the controlled addition of a base allows the nearly complete elimination of the iron from metallurgical processing streams or effluents, as readily filterable Fe2O3.  相似文献   

2.
The aging behavior of iron-nitrogen martensitic alloys (0.8 to 7.0 at. pct N) between -190 °C and 450 °C was investigated by quantitative analysis of the corresponding changes in volume and enthalpy. Martensitic specimens were prepared by gaseous nitriding of pure iron in a mixture of NH3 and H2 and subsequent quenching in brine and liquid nitrogen. Both X-ray diffraction analysis and metallography (light microscopical analysis and microhardness measurement) were used for interpretation of the structural changes. Analysis of the transformation kinetics was achieved by employing a range of heating rates. At least five different stages of structural change could be distinguished, which were quantitatively analyzed in terms of their effects on volume and enthalpy: (1) transformation of retained austenite into martensite (between-160 °C and -40 °C); (2) segregation and “ordering” of nitrogen atoms (below 100 °C); (3) precipitation of incoherent α′ nitride (between 100 °C and 220 °C); (4) conversion of α′ nitride into γ′ nitride (between 220 °C and 290 °C); and (5) decomposition of retained austenite (between 240 °C and 350 °C). Differences with the tempering behavior of analogous iron-carbon martensites were discussed.  相似文献   

3.
Experimental study of phase equilibria in the Al-Fe-Zn-O system in air   总被引:1,自引:0,他引:1  
The phase equilibria in the Al-Fe-Zn-O system in the range 1250 °C to 1695 °C in air have been experimentally studied using equilibration and quenching techniques followed by electron probe X-ray microanalysis. The phase diagram of the binary Al2O3-ZnO system and isothermal sections of the Al2O3-“Fe2O3”-ZnO system at 1250 °C, 1400 °C, and 1550 °C have been constructed and reported for the first time. The extents of solid solutions in the corundum (Al,Fe)2O3, hematite (Fe,Al)2O3, Al2O3*Fe2O3 phase (Al,Fe)2O3, spinel (Al,Fe,Zn)O4, and zincite (Al,Zn,Fe)O primary phase fields have been measured. Corundum, hematite, and Al2O3*Fe2O3 phases dissolve less than 1 mol pct zinc oxide. The limiting compositions of Al2O3*Fe2O3 phase measured in this study at 1400 °C are slightly nonstoichiometric, containing more Al2O3 then previously reported. Spinel forms an extensive solid solution in the Al2O3-“Fe2O3”-ZnO system in air with increasing temperature. Zincite was found to dissolve up to 7 mole pct of aluminum in the presence of iron at 1550 °C in air. A meta-stable Al2O3-rich phase of the approximate composition Al8FeZnO14+x was observed at all of the conditions investigated. Aluminum dissolved in the zincite in the presence of iron appears to suppress the transformation from a round to platelike morphology.  相似文献   

4.
Neodymium naphthenate-loaded organic phase stripping using sodium oxalate solution was studied to explore the feasibility of synchronous rare earth-loaded organic phase stripping, rare earth precipitation, and blank organic phase saponification. Experimental results show that loaded organic phase stripping, rare earth precipitation, and blank organic phase saponification can be realized simultaneously. When using 20% excess of sodium oxalate over the stoichiometry with the volume ratio of organic phase to aqueous phase of 1:1 at 25 °C for 40 min, the single stage stripping rate and saponification value are about 40% and 0.29 mol/L, respectively. After 16 stages of countercurrent continuous stripping, the stripping rate of neodymium can reach 99%, the saponification value is 0.42 mol/L, the Nd3+ concentration in saponified organic phase is less than 0.0020 mol/L, and the main phase in precipitation is Nd2(C2O4)3·10H2O. Afterwards, this saponified organic phase can be used in the extraction of NdCl3 solution, and then the loaded organic phases (neodymium naphthenate) with 0.16 mol/L Nd3+ can be retrieved. The morphology, particle size distribution, and composition of the Nd2(C2O4)3·10H2O products are similar to those of the current direct precipitation products. The neodymium oxide prepared by continuous calcination of neodymium oxalate meets the national standard of China (GB/T 5240?2015). These results prove the feasibility of stripping neodymium naphthenate-loaded organic phase by using sodium oxalate solution. Sodium oxalate can serve as a stripping agent, a saponifier, and a precipitator, thereby simplifying rare earth extraction and separation. This study provides theoretical and technical support for the development of a novel method for rare earth extraction and separation.  相似文献   

5.
The conversion of hematite to magnetite was investigated in a mixture of H2, H2O, and N2. In the temperature range of 500° to 700°C, the magnetic roast reaction gives a sigmoidal kinetic curve with a finite induction time. The induction time decreases with rise in temperature and increases in the presence of alkali and alkaline earth oxides. The magnetic roast reaction was also studied in the presence of low concentrations of silica (quartz structure) and alumina (5.0 and 4.3 wt pct, respectively, which correspond to 8.4 × 10-4 g atom of Si or Al per gram mixture). The addition of SiO2 to hematite decreases the induction time. At temperatures below 550°C, alumina increases the induction time; at higher temperatures it has about the same effect as silica. For pellets containing SiO2, a maximum in the relative decrease in induction time, A, was observed at 578°C; for pellets containing A12O3, there was a steady increase in A with increasing temperature. Because the α β quartz transition occurs at 575°C, the enhanced surface activity at 578°C in the presence of quartz is attributed to the Hedvall effect of solid-state chemistry. The induction period of the magnetic roast reaction was exceptionally prolonged in the presence of lithia. Mixing of hematite with silica, alumina, and lithia (8.4 × 10-4 g atom, respectively, per gram mixture) was found to eliminate the beneficial effect of quartz by inhibiting its α β transformation.  相似文献   

6.
The phase equilibria in the FeO-Fe2O3-ZnO system have been experimentally investigated at oxygen partial pressures between metallic iron saturation and air using a specially developed quenching technique, followed by electron probe X-ray microanalysis (EPMA) and then wet chemistry for determination of ferrous and ferric iron concentrations. Gas mixtures of H2, N2, and CO2 or CO and CO2 controlled the atmosphere in the furnace. The determined metal cation ratios in phases at equilibrium were used for the construction of the 1200 °C isothermal section of the Fe-Zn-O system. The univariant equilibria between the gas phase, spinel, wustite, and zincite was found to be close to pO2=1 · 10−8 atm at 1200 °C. The ferric and ferrous iron concentrations in zincite and spinel at equilibrium were also determined at temperatures from 1200 °C to 1400 °C at pO2 = 1·10−6 atm and at 1200 °C at pO2 values ranging from 1 · 10−4 to 1 · 10−8 atm. Implications of the phase equilibria in the Fe-Zn-O system for the formation of the platelike zincite, especially important for the Imperial Smelting Process (ISP), are discussed.  相似文献   

7.
Abstract

The hydrothermal conversion of K jarosite, Pb jarosite, Na jarosite, Na–Ag jarosite, AsO4 containing Na jarosite and in situ formed K jarosite and Na jarosite to hematite was investigated. Potassium jarosite is the most stable jarosite species. Its conversion to hematite in the absence of Fe2O3 seed occurred only partially after 5 h reaction at >240°C. In the presence of Fe2O3 seed, the conversion to hematite was nearly complete within 2 h at 225°C and was complete at 240°C. The rate of K jarosite precipitation, in situ at 225°C in the presence of 50 g L?1 Fe2O3 seed, is faster than its rate of hydrothermal conversion to hematite. In contrast, complete conversion of either Pb jarosite or Na–Pb jarosite to hematite and insoluble PbSO4 occurs within 0·75 h at 225°C in the presence of 20 g L?1 Fe2O3 seed. Dissolved Fe(SO4)1·5 either inhibits the conversion of Pb jarosite or forms Pb jarosite from any PbSO4 generated. The hydrothermal conversion of Na–Ag jarosite to hematite was complete within 0·75 h at 225°C in the presence of 20 g L?1 Fe2O3 seed. The Ag dissolved during hydrothermal conversion and reported to the final solution. However, the presence of sulphur or sulphide minerals caused the reprecipitation of the dissolved Ag. The conversion of AsO4 containing Na jarosite at 225°C in the presence of 20 g L?1 Fe2O3 seed was complete within 2 h, for H2SO4 concentrations <0·4M. Increasing AsO4 contents in the Na jarosite resulted in a linear increase in the AsO4 content of the hematite, and ~95% of the AsO4 remained in the conversion product. Increasing temperatures and Fe2O3 seed additions significantly promote the hydrothermal conversion of in situ formed Na jarosite at 200–240°C. However, the conversion of previously synthesised Na jarosite seems to proceed to a greater degree than that of in situ formed Na jarosite.

On a examiné la conversion hydrothermale en hématite de la jarosite de K, de la jarosite de Pb, de la jarosite de Na, de la jarosite de Na-Ag, de la jarosite de Na contenant de l’AsO4, et de la jarosite de K et de la jarosite de Na qui sont formées in situ. La jarosite de potassium est la plus stable des espèces de jarosite. Sa conversion en hématite ne se produisait que partiellement après 5 h de réaction à >240°C en l’absence d’amorce de Fe2O3. En présence d’amorce de Fe2O3, la conversion en hématite était presque complète à moins de 2 h à 225°C et était complète à 240°C. La vitesse de précipitation de la jarosite de K, in situ à 225°C en présence de 50 g L?1 d’amorce de Fe2O3, est plus rapide que sa vitesse de conversion hydrothermale en hématite. Par contraste, la conversion complète soit de la jarosite de Pb ou de la jarosite de Na-Pb en hématite et en PbSO4 insoluble se produit à moins de 0·75 h à 225°C en présence de 20 g L?1 d’amorce de Fe2O3. Le Fe(SO4)1·5 dissous soit inhibe la conversion de la jarosite de Pb ou forme de la jarosite de Pb à partir de tout PbSO4 produit. La conversion hydrothermale de la jarosite de Na-Ag en hématite était complète à moins de 0·75 h à 225°C en présence de 20 g L?1 d’amorce de Fe2O3. L’Ag se dissolvait lors de la conversion hydrothermale et se rapportait dans la solution finale. Cependant, la présence de soufre ou de minéraux sulfurés avait pour résultat la re-précipitation de l’Ag dissous. La conversion de la jarosite de Na contenant de l’AsO4 à 225°C en présence de 20 g L?1 d’amorce de Fe2O3 était complète à moins de 2 h, avec des concentrations d’H2SO4 <0·4 M. L’augmentation de la teneur en AsO4 de la jarosite de Na avait pour résultat une augmentation linéaire de la teneur en AsO4 de l’hématite et ~95% de l’AsO4 demeurait dans le produit de conversion. L’augmentation de la température et d’additions d’amorce de Fe2O3 favorisait significativement la conversion hydrothermale de la jarosite de Na qui est formée in situ à 220–240°C. Cependant, la conversion de la jarosite de Na synthétisée antérieurement semblait se produire à un plus grand degré que celle de la jarosite de Na qui est formée in situ.  相似文献   

8.
Aqueous oxidation of chalcopyrite in hydrochloric acid   总被引:1,自引:0,他引:1  
The aqueous oxidation of chalcopyrite flotation concentrate is faster in HC1 than in H2SO4. Under certain conditions, FeCl2 formed during reaction undergoes oxidation and hydrolysis (or vice versa) yielding α-Fe2O3 or β-FeOOH depending on the initial acidity, O2 pressure, and temperature. A small fraction of sulfur oxidizes to soluble SO 4 2- and a part precipitates as Fe(OH)SO4. The precipitation of FeOOH is more enhanced in HC1 than in H2SO4 solution. Optimum leaching conditions are 110°C, 2010 kPaO2, 2 N HC1, molar ratio CuFeS2/HCl 0.8 to 0.9, and time of reaction 45 min. Under these conditions practically all the iron precipitates and the reaction can be described by the equations: CuFeS2 + 2 HCl + 5/4 O2 → CuCl2 + FeOOH + 1/2H2O + 2S and CuFeS2 + HC1 + O2 → CuCl + FeOOH + 2S. Copper ferrite, CuFe2O4, precipitates if the final pH of the leach solution is ≥2.2, while CuCl precipitates if the initial acidity is increased to 3 N HC1.  相似文献   

9.
The sulfidation of wustite in H2S−H2O−H2−Ar atmospheres has been studied at temperatures of 700, 800, and 900°C with thermogravimetric techniques. Polycrystalline wustite wafers were equilibrated in a flowing H2O−H2−Ar gas stream and then sulfidizedin situ. During an initial transient stage a protective layer of FeS formed on the sample, and an intermediate layer of Fe3O4 formed between the FeO and FeS layers. Subsequently, the reaction followed a parabolic rate law. The parabolic rate constant varied from 0.22×10−2 mg2 cm−4 min−1 at 700°C to 6.5×10−2 mg2 cm−4 min−1 at 900°C. The reaction rate was limited by the diffusion of iron through the intermediate Fe3O4 layer which grew concurrently with the FeS layer and at the expense of the FeO core. After the FeO core was completely converted to Fe3O4, the process entered a passive stage during which no further mass changes could be detected. SCOTT McCORMICK, formerly Graduate Student, Purdue University is currently Assistant Professor, Department of Metallurgical and Materials Engineering, Illinois Institute of Technology, Chicago, Illinois 60616.  相似文献   

10.
llmenite concentrates were heated in argon and oxygen in the temperature range 700 °C to 1000 °C to study the behavior of the pseudorutile phase and other changes which occur. Pseudorutile does not persist in argon or oxygen in the temperature range studied. In argon at 700 °C, pseudorutile decomposes into hematite and rutile, while at 1000 °C, it combines with ilmenite to form ferrous-ferritic pseudobrookite solid solution. A new phase “Fe2O3-2TiO2” was identified as an intermediate product during the heating of ilmenite or pseudorutile in oxygen. This compound decomposes into hematite and rutile below 800 °C and to pseudobrookite and rutile above 800 °C. The sequence of reactions during the heating of ilmenite and pseudorutile in oxygen is proposed. Formerly with the Department of Metallurgy, Imperial College of Science, Technology and Medicine, London. Formerly with the Department of Metallurgy, Imperial College of Science, Technology and Medicine, London.  相似文献   

11.
Potential transformation of oolitic hematite into magnetite by mixing iron powder using the mechanochemical method has been achieved and discussed in this paper. The phase transition of pure hematite in the preliminary test was identified by X-ray diffractometer (XRD) and Fourier Transform Infrared Spectroscopy (FTIR) techniques. The experimental results have shown that the crystallographic planes of magnetite, (220), (311), (400), and (511) were observed clearly in the Fe/α-Fe2O3 mixture after milling for 15 h, indicating that α-Fe2O3 had been effectively transformed into Fe3O4. The diffraction peaks of magnetite were also observed at d = 0.29605 nm (2θ = 30.163°), 0.25226 nm (2θ = 35.559°), 0.24156 nm (2θ = 37.190°), and 0.20898 nm (2θ = 43.458°) after 13 h milling-time. It suggests that the oolitic hematite is transformed into magnetite successfully by mechanochemical processing. The processing might be applied potentially for the magnetic separation of oolitic hematite.  相似文献   

12.
13.
Phase equilibria and thermodynamics in the FeO-TiO2-Ti2O3 ternary system were studied at 1500 °C and 1600 °C. In particular, the liquid slag-phase region and its saturation boundary with respect to metallic iron, titania, and lower titanium oxides was investigated. The liquid slag-phase region extends substantially toward an anosovite (Ti3O5) composition, and considerable concentrations of divalent iron coexist with trivalent titanium in the liquid-slag phase. This seems to be a consequence of the complete solid solution between ferrous pseudobrookite (FeTi2O5) and anosovite (Ti3O5), which exists at subsolidus temperatures. The liquid-slag field is significantly enlarged toward the anosovite composition upon increasing the temperature from 1500 °C to 1600 °C. Activities of the components “FeO” and TiO2 in the liquid-slag region were determined by Gibbs-Duhem integration of the measured oxygen partial pressures at 1500 °C. The FeO shows moderate negative deviation, while titania shows a slight negative deviation in FeO-rich slags and a positive deviation in high-titania slags. The experimentally measured activity values were modeled using regular and biregular solution models, and good agreement was obtained with the biregular solution model.  相似文献   

14.
Complex silico-ferrites of calcium and aluminium (low-Fe form, denoted as SFCA; and high-Fe, low-Si form, denoted as SFCA-I) constitute up to 50 vol pct of the mineral composition of fluxed iron ore sinter. The reaction sequences involved in the formation of these two phases have been determined using an in-situ X-ray diffraction (XRD) technique. Experiments were carried out under partial vacuum over the temperature range of T=22 °C to 1215 °C (alumina-free compositions) and T=22 °C to 1260 °C (compositions containing 1 and 5 wt pct Al2O3) using synthetic mixtures of hematite (Fe2O3), calcite (CaCO3), quartz (SiO2), and gibbsite (Al(OH)3). The formation of SFCA and SFCA-I is dominated by solid-state reactions, mainly in the system CaO-Fe2O3. Initially, hematite reacts with lime (CaO) at low temperatures (T ∼ 750 °C to 780 °C) to form the calcium ferrite phase 2CaO·Fe2O3 (C2F). The C2F phase then reacts with hematite to produce CaO·Fe2O3 (CF). The breakdown temperature of C2F to produce the higher-Fe2O3 CF ferrite increases proportionately with the amount of alumina in the bulk sample. Quartz does not react with CaO and hematite, remaining essentially inert until SFCA and SFCA-I began to form at around T=1050 °C. In contrast to previous studies of SFCA formation, the current results show that both SFCA types form initially via a low-temperature solid-state reaction mechanism. The presence of alumina increases the stability range of both SFCA phase types, lowering the temperature at which they begin to form. Crystallization proceeds more rapidly after the calcium ferrites have melted at temperatures close to T=1200 °C and is also faster in the higher-alumina-containing systems.  相似文献   

15.
The kinetics of hydrogen reduction of thin, dense strips of hematite were investigated in the range 245 °C to 482 °C. Pure hydrogen gas at 1 atm was used as the reducing agent. Because of the relative thinness (only 136 /μm thick) of the specimens used, the pore-diffusion of gases offered no significant resistance to the reduction process. The interfacial-reaction-rate constantk s * , which has been corrected for film-mass-transfer effects, is found to be given by logk s * = −1.032 (±0.138) -[7860 (±200)]/2.303r where k s * is in g · atom O · cm−2 · s−1 · atm−1. The activation energy for the reduction process is found to be 65,325 (±1650) J · mol−1; the rate-controlling step appears to be the Fe3O4 → Fe conversion step.  相似文献   

16.
《Hydrometallurgy》2007,85(2-4):183-192
Iron(III) can be used as an oxidant in the leaching of uranium ore in an acid medium. The oxidation of iron(II) to iron(III) using an SO2/O2 gas mixture was investigated in order to provide an iron(III) stream for uranium extraction. The effects of pH, temperature and SO2/O2 volumetric ratios were considered. Oxidation of iron(II) by SO2/O2 was controlled by diffusion of SO2 or O2 at pH 2 and 40 °C. However as the pH decreased below pH 1, the reaction was controlled by a slow chemical step and the reaction rate decreased. Increasing the temperature increased the oxidation rate at pH 0.8, and at 70 °C the rate again became dependent on SO2 or O2 diffusion. The oxygen efficiency for a fixed reactor set-up was dependent on the SO2/O2 ratio and total flow rate of the gas. In leach tests, the uranium extraction achieved with iron(III) solution prepared by SO2/O2 oxidation was the same as that for a standard uranium leach with conventional oxidant.  相似文献   

17.
The phase transformations occurring during magnetizing roasting of leucoxene concentrate in the temperature range 600–1300°C are studied. It is demonstrated, that in the temperature range 600–800°C, only iron oxides are reduced to a metallic state; at temperatures above 800°C, combined reduction of iron and titanium oxides takes place. At 1050°C, reduced specimens are represented by the Ti5O9 and Ti6O13 Magnéli phases. The formation of iron metatitanate (FeTiO3), under reduction conditions and the existence of ferrous iron ions in the Magnéli phases slightly degrade the magnetic properties of the products of magnetizing roasting. In high temperature region (1200–1300°C), a similar effect is exerted by the formation of iron dititanate or anosovite in the system. The possibilities of eliminating the undesired factors decreasing the magnetic properties of the products of magnetizing roasting are determined.  相似文献   

18.
Kinetics of silver leaching from a manganese-silver associated ore in sulfuric acid solution in the presence of H2O2 has been investigated in this article. It is found that sulfuric acid and hydrogen peroxide have significant effects on the leaching rate of silver. The reaction orders of H2SO4 and H2O2 were determined as 0.80 and 0.68, respectively. It is found that the effects of temperature on the leaching rate are not marked, the apparent activation energy is attained to be 8.05 kJ/mol within the temperature range of 30 °C to 60 °C in the presence of H2O2. Silver leaching is found to be diffusion-controlled and follows the kinetic model: 1−2x/3−(1−x)2/3=Kt. It is also found that particle size presents a clear effect on silver leaching rate, and the rate constant (k) is proportional to d −2 0 .  相似文献   

19.
The kinetics of the chlorination of gallium oxide in chlorine atmosphere was studied between 650 °C and 800 °C. The calculations of the Gibbs standard free energy variation with temperature for the reaction Ga2O3(S)+3Cl2 (g)→2GaCl3(g)+1.5O2 (g) show that direct chlorination is favorable above 850 °C. Thermogravimetric experiments were performed under isothermal and nonisothermal conditions. The effect of temperature, gas flow rate, and Cl2 partial pressure were studied. The solids were characterized by X-ray diffraction (XRD) and scanning electronic microscopy (SEM). The nonisothermal results showed that chlorination of Ga2O3 starts at approximately 650 °C, with a mass loss of 50 pct at 850 °C. The isothermal results between 650 °C and 800 °C indicated that the reaction rate increased with temperature. The correlation of the experimental data with different solid-gas reaction models showed that the results are adequately represented by the model proposed by Shieh and Lee: X=1−{1−b 22[b 21 t+e −b 21 t−1]}1/(1−γ). From this model, it was found that the rate of reaction for the chlorination of Ga2O3 is of the order 0.68 with respect to Cl2 and the activation energy is 113.23 kJ/mol. On the other hand, the order of the activation rate of the interface surface is 0.111 with respect to Cl2 and its activation energy is 23.81 kJ/mol.  相似文献   

20.
It is possible, in some cases, for ground coal particles to react with gasifier gas during combustion, allowing the ash material in the coal to form phases besides the expected slag phase. One of these phases is metallic iron, because some gasifiers are designed to operate under a reducing atmosphere (pO2{p_{\rm {O}_{2}}} of approximately 10−4 atm). Metallic iron can become entrained in the gas stream and deposit on, and foul, downstream equipment. To improve the understanding of the reaction between different metallic iron particles and gas, which eventually oxidizes them, and the slag that the resulting oxide dissolves in, the kinetics of iron reaction on slag were predicted using gas-phase mass-transfer limitations for the reaction and were compared with diffusion in the slag; the reaction itself was observed under confocal scanning laser microscopy. The expected rates for iron droplet removal are provided based on the size and effective partial pressure of oxygen, and it is found that decarburization occurs before iron reaction, leading to an extra 30- to 100-second delay for carbon-saturated particles vs pure iron particles. A pure metallic iron particle of 0.5 mg should be removed in about 220 seconds at 1400 °C and in 160 seconds at 1600 °C.  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