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.
Physical modeling of gas/liquid mass transfer in a gas stirred ladle   总被引:3,自引:0,他引:3  
The absorption of gas through the plume eye and of an injected gas in a steelmaking ladle process was investigated, using a physical model of CO2 absorption into a NaOH solution. The results show that the inert gas escaping through the plume eye is ineffective in protecting the bath from the atmosphere, and placing an oil layer (simulated slag) decreases the absorption rate significantly. Increasing the flow rate of the inert gas not only exposes more of the liquid surface to the CO2 atmosphere, but also increases the mass transfer coefficient at the surface. The overall mass transfer between an injected CO2 gas and NaOH solution includes the mass transfer through the surface of the bath as well as the mass transfer in the bubble dispersion zone. The difference between the mass transfer in the bubble dispersion zone and the overall mass transfer was found to be significant for relatively low gas flow rates. The mass transfer coefficient of CO2 in the bubble dispersion zone was estimated using available information regarding the bubble size and velocity. Mass transfer coefficient estimated for the constant bubble frequency regime shows a dependence on gas flow rate. However, if a constant characteristic size of bubbles is assumed as an alternative approach, the mass transfer coefficient is independent of the gas flow rate.  相似文献   

2.
An antimony microelectrode was prepared by quenching a molten Sb-Sb2O3 mixture (2 pct Sb2O3). The local pH in the vicinity of a nickel-plated copper cathode was directly measured using the microelectrode during the chloride electrowinning of nickel for a MCLE (matte chlorine leach electrowinning) process, where nickel metal is electrodeposited with a high current efficiency, 94 to 97 pct, from low-pH baths. The local pH at 328 K was increased by proton consumption during the electrolysis of aqueous electrolytes containing NiCl2 (1.20 mol dm−3) and NaCl (0.43 mol dm−3) with the same concentrations as employed for the MCLE process. The difference in pH between the cathode surface and bulk solution increased with increasing cathodic current density. Nickel deposits with a metallic luster were obtained when the difference was not more than 1.2 pH units. The current efficiency was a maximum for electrolysis with a current density of 265 A m−2 and bulk pH of 1.0 to 1.5; these optimal conditions coincided with those reported for the MCLE process: temperature 328 to 333 K, bulk pH 1.1 to 1.5, and current density 230 to 260 A m−2. Electrolytes with lower NiCl2 and NaCl concentrations resulted in a drop in current efficiency.  相似文献   

3.
Densities, electrical conductivities and viscosities of CuSO4-H2SO4-H2O solutions are reported for electrorefining electrolytes over the ranges: Temperature 50 to 70°C Copper concentration 40 to 55 g · dm-3 H2SO4 concentration 160 to 220 g · dm-3 and for electrowinning electrolytes over the ranges: Temperature 20 to 70°C Copper concentration 10 to 60 g · dm-3 H2SO4 concentration 10 to 170 g · dm-3 Empirical and semiempirical equations describing the measured properties are presented.  相似文献   

4.
The oxidation of Fe(II) with dissolved molecular oxygen was studied in sulfuric acid solutions containing 0.2 mol · dm−3 FeSO4 at temperatures ranging from 343 to 363 K. In solutions of sulfuric acid above 0.4 mol · dm−3, the oxidation of Fe (II) was found to proceed through two parallel paths. In one path the reaction rate was proportional to both [Fe−2+]2 andp o 2 exhibiting an activation energy of 51.6 · kJ mol−1. In another path the reaction rate was proportional to [Fe2+]2, [SO 4 2 ], andp o 2 with an activation energy of 144.6 kJ · mol−1. A reaction mechanism in which the SO 4 2 ions play an important role was proposed for the oxidation of Fe(II). In dilute solutions of sulfuric acid below 0.4 mol · dm−3, the rate of the oxidation reaction was found to be proportional to both [Fe(II)]2 andp o 2, and was also affected by [H+] and [SO 4 2 ]. The decrease in [H+] resulted in the increase of reaction rate. The discussion was further extended to the effect of Fe (III) on the oxidation reaction of Fe (II).  相似文献   

5.
Extraction of copper(II) with capric acid in chloroform was investigated at 25°C and aqueous ionic strength of 0.33 mol dm–3 (Na2SO4). The composition of the complex extracted was determined using the slope analysis method. Copper(II) complex exists as a monomeric species CuL2.2HL in the lower concentration range of capric acid and a dimeric species (CuL2.HL)2 in the higher concentration range, while both monomer and dimer were extracted at a total capric acid concentration of 0.1–0.2 mol dm–3. Extraction and dimerization constants of the above species were also determined. UV-Visible and FT-IR spectroscopy have also been applied for the study of the ligand and their copper(II) species. Electronic spectrum of the organometallic complexes suggests the octahedral geometry.  相似文献   

6.
The alkaline decomposition of Argentian rubidium jarosite in NaOH media is characterized by an induction period and a progressive conversion period in which the sulfate and rubidium ions pass to the solution, leaving an amorphous iron hydroxide residue. The process is chemically controlled and the order of reaction with respect to hydroxide concentration in the range of 1.75 and 20.4?mol OH? m?3 is 0.94, while activation energy in the range of temperatures of 298?K to 328?K (25?°C to 55?°C) is 91.3?kJ mol?1. Cyaniding of Argentian rubidium jarosite in NaOH media presents a reaction order of 0 with respect to NaCN concentration (in the range of 5 to 41?mol m?3) and an order of reaction of 0.62 with respect to hydroxide concentration, in the range of 1.1 and 30?mol [OH?] m?3. In this case, the cyaniding process can be described, as in other jarosites, as the following two-step process: (1) a step (slow) of alkaline decomposition that controls the overall process followed by (2) a fast step of silver complexation. The activation energy during cyaniding in the range of temperatures of 298?K to 333?K (25?°C to 60?°C) is 43.5?kJ mol?1, which is characteristic of a process controlled by chemical reaction. These results are quite similar to that observed for several synthetic jarosites and that precipitated in a zinc hydrometallurgical plant (Industrial Minera México, San Luis Potosi).  相似文献   

7.
Herein,we report the experimental methodology to optimize the operational parameters of the photocatalytic degradation of acriflavine dye using TiO_2 and Gd~(3+) as dopant.A series of Gd~(3+) doped TiO_2 nanoparticles were synthesized via hydrothermal route and characterized using various techniques like FT-IR,UV,XRD,FESEM and EDS.It is observed that synthesized particles are in the range of 25-30 nm with spherical shape in nature.TiO_2 has rutile phase and the average particle size was estimated from Scherrer's equation.Energy bandgap was estimated using Tauc's plot.The photodegradation was carried out under UV light in pseudofirst order condition.To understand the kinetics,four experimental parameters were chosen as independent variables like percentage of dopant,initial concentration of dye,dosage of catalyst and pH of reaction medium.The degradation efficiency of 92% was observed for 0.5%Gd doped TiO_2 at catalyst dosage of 0.3 g/dm~3,pH 10 and dye concentration of 3×10~(-6) mol/dm~3.It is observed that,the photocatalytic activity of TiO_2 can be increased by using gadolinium as dopant only in optimum concentration.Further,this photocatalyst can be employed to degrade other o rganic pollutants.  相似文献   

8.
A chalcopyrite concentrate containing 17 pct pyrite was oxidized in 1 mol/dm3 sulfuric acid solution at 90 °C (363.2 K). The suspension potential (Ptvs SCE, in the presence of Fe3+/Fe2+) was maintained constant in the range 0.30 to 0.65 V by controlled additions of KMnO4 solution. The oxidation appeared to be under surface reaction control. The rate constant was nearly independent of total Fe concentration (0.01 to 0.5 mol/dm3), but increased rapidly with a rise in suspension potential until it reached a maximum at 0.40 to 0.43 V, after which there was marked decrease at around 0.45 V. Chalcopyrite in the concentrate was oxidized to form elemental sulfur over the whole of the suspension potential range, whereas the oxidation of pyrite took place only above 0.45 V and yielded sulfate ion. At 0.40 V the apparent activation energy was 47 kJ/mol. An analogy between the potential dependence of the rate and the Tafel correlation for an electrode process is discussed.  相似文献   

9.
Effect of the bubble size and chemical reactions on slag foaming   总被引:5,自引:0,他引:5  
Slag foams have been investigated with smaller bubbles than those used in the previous studies.[5,6,7] The bubbles were generated by argon gas injection with the nozzle of multiple small orifices and by the slag/metal interfacial reaction of FeO in the slag with carbon in the liquid iron. The foam stability in terms of the foam index for a bath-smelting type of slag (CaO-SiO2-Al2O3-FeO) was determined for different bubble sizes. The average diameter of bubbles in the foam was measured by an X-ray video technique. When the foam was generated by the slag/metal interfacial reaction at 1450 °C, it was found that the average bubble diameter varied from less than 1 to more than 5 mm as a function of the sulfur activity in the carbon-saturated liquid iron. The foam index was found to be inversely proportional to the average bubble diameter. A general correlation is obtained by dimensional analysis in order to predict the foam index from the physical properties of the liquid slag and the average size of the gas bubbles in the foam. Formerly Graduate Student and Research Associate, Department of Materials Science and Engineering, Carnegie Mellon University.  相似文献   

10.
The leaching of chalcopyrite with ferric sulfate   总被引:1,自引:0,他引:1  
The leaching kinetics of natural chalcopyrite crystals with ferric sulfate was studied. The morphology of the leached chalcopyrite and the electrochemical properties of chalcopyrite electrodes also were investigated. The leaching of chalcopyrite showed parabolic-like kinetics initially and then showed linear kinetics. In the initial stage, a dense sulfur layer formed on the chalcopyrite surface. The growth of the layer caused it to peel from the surface, leaving a rough surface. In the linear stage, no thick sulfur layer was observed. In this investigation, chalcopyrite leaching in the linear stage was principally studied. The apparent activation energy for chalcopyrite leaching was found to range from 76.8 to 87.7 kJ mol−1, and this suggests that the leaching of chalcopyrite is chemically controlled. The leaching rate of chalcopyrite increases with an increase in Fe(SO4)1.5 concentration up to 0.1 mol dm−3, but a further increase of the Fe(SO4)1.5 concentration has little effect on the leaching rate. The dependency of the mixed potential upon Fe(SO4)1.5 concentration was found to be 79 mV decade−1 from 0.01 mol dm−3 to 1 mol dm−3 Fe(SO4)1.5. Both the leaching rate and the mixed potential decreased with an increased FeSO4 concentration. The anodic current of Fe(II) oxidation on the chalcopyrite surface in a sulfate medium was larger than that in a chloride medium.  相似文献   

11.
The gasification of graphite by carbon dioxide was studied under atmospheric pressure in a fixed bed reactor in the temperature range of 1173–1773?K, CO2 partial pressures 2–10?kPa and gas flow rate 0·5–2·0?L?min?1. Iron presented in a small amount in graphite ash had a catalytic effect on the gasification reaction at 1373?K; this effect was weaker at 1473?K due to the melting of iron saturated with carbon. The gasification rate increased with increasing CO2 partial pressure and total gas flow rate.  相似文献   

12.
The kinetics of decarburization of liquid nickel in CO2-CO mixtures have been studied at 1400 and 1500°C, using the experimental arrangement of the impinging jet. At carbon concentrations above about 1 wt pct, pressures of CO2 ⪯ 0.1 atm, and for total gas flow-rates above about 40 l/min (STP) impinging on a metal surface of 2.08 cm2, it is concluded that the interfacial reaction step controls the rate. Comparison with isotope exchange studies indicates that dissociative chemisorption of CO2 is the rate determining step. Rate constants, based on the nominal surface area, are 1.2 ×10−3 and 1.4 × 10−3 mol/cm2 · s · atm at 1400 and 1500°C, respectively. on leave of absence from the Homer Research Laboratories of the Bethlehem Steel Co.  相似文献   

13.
《Hydrometallurgy》2008,93(3-4):107-114
Crystalline FeOOH and Zn(II) doped FeOOH samples were prepared by a co-precipitation method using ferrous sulphate heptahydrate, zinc sulphate heptahydrate, ammonia, 3% hydrogen peroxide at 80 °C and pH value 3.0. These samples were characterized by XRD and were also chemically analyzed. The XRD patterns of pure FeOOH matched with that of Zn(II) doped FeOOH sample and showed no shift in their d-values, which indicated that Zn2+ ions were adsorbed on the surface of FeOOH or on the surface of the crystalline lattice, and were not captured into the ferrite lattice. Chemical analysis showed that the molar ratio of Zn(II)/Fe in Zn(II) doped FeOOH sample was only 0.038 when this ratio was 0.5 in an initial reaction solution. These samples were treated in ammoniacal or sodium hydroxide medium in the presence of ferrous sulphate or cobalt sulphate under hydrothermal conditions. The converted products were analyzed both for ferrous, total iron, cobalt and zinc. Characterization was carried out using instrumental techniques such as SEM, FT-IR, XRD and magnetization characterization. Though the XRD patterns of the converted products showed presence of only magnetite in the NH3–Fe2+, Co2+–H2O converted medium, the FT-IR spectra and chemical analysis showed an un-reacted FeOOH in the converted product. Even though, the saturation magnetization of this magnetite was higher to 97.489 emu g 1. Furthermore, higher temperature increased the concentration of Co2+ captured into spinel structure. The conversion were restrained in NH3 concentration of less than 12 g/L and more than 25 g/L. However, increasing of Fe2+/Co2+ ratio in the NH3–Fe2+, Co2+–H2O medium made no distinct difference in the introduction of Co2+ into spinel structure when concentration of Fe2+ was enough. Compared with Zn-free FeOOH, it was difficult for Zn-doped FeOOH to convert to magnetite in the NH3 (NaOH)–Fe2+–H2O medium. Furthermore, alkali species were critical in the above conversion. The content of Zn2+ which was out of Zn-doped FeOOH solid state and entered into the outer solution was more in the NH3 medium than that of the NaOH medium, but conversion from Zn–FeOOH into magnetite was easy in the NaOH medium. The explanations of above facts were given in the text.  相似文献   

14.
《钢铁冶炼》2013,40(3):195-202
Abstract

The present study has been conducted to investigate the effects of operating conditions, which include gas flowrate, tuyere size, tuyere number, and height of iron phase, on the extent of mixing between molten iron and molten slag in the direct iron ore smelting reduction process. A transparent acrylic water model, 30% of the size of the actual smelter, was constructed to study the mass transfer phenomena. In the water model, water and spindle oil were used to simulate molten iron and molten slag, respectively, while air was used to replace the bottom blown nitrogen gas. In addition, thymol (C10H14O6) was used as a tracer material in the water model, added to the water at the beginning of the experiment. As mixing between water and spindle oil proceeded owing to stirring by the bottom blown gas, the concentration of thymol in the water decreased and that in the spindle oil increased. Water samples were taken from the bottom and 12 cm above the bottom of the water model at various operating times. Concentrations of thymol were then measured using a diode array ultraviolet visible spectrophotometer. By analysing the concentration data, the mass transfer rate kwA, which is a direct index for evaluating the mixing efficiency, could be derived. The process conditions under investigation included 40-500 L min-1 gas flowrate, 0·3-1 cm tuyere size, four or five tuyeres, and 20-30 cm height of the water phase. The test results indicate that when the gas flowrate increases, the value of kwA increases, which indicates better mixing between oil and water phases. However, as the gas flowrate approaches 40 L min-1, the improvement becomes less obvious. The smaller tuyere gives better mixing, and the design of five tuyeres results in better mixing compared with four tuyeres when they are blown with the same total gas flowrate. However, mixing efficiency decreases with increased height of the water phase. Also, as the gas flowrate of bottom blowing approaches 40 L min-1, gas blowing from the top has little effect on the mixing behaviour in the liquid bath. For a four tuyere system, the process conditions of height of oil phase 5 cm, height of water phase 25 cm, diameter of tuyere 0·75 cm, and gas flowrate for each tuyere 40 L min-1, appear to be the optimal design.  相似文献   

15.
A novel hydrometallurgical process for recovering Zn from ZnS in alkaline solution via chemical conversion with PbCO3 was developed in this work. The S originally present in ZnS can be converted into PbS, while the Zn can be converted into Na2Zn(OH)4 in the alkaline solution in the presence of PbCO3. And then, the Pb in PbS deposited in the leach residues can be converted into PbCO3 again in the Na2CO3 solution. It was found that over 90% of Zn can be extracted from ZnS when the leaching process is operated in 6 mol/L NaOH solution at 90 °C with PbCO3 as additive, and over 95% of PbS in the leach residues can be converted into PbCO3 by stirring the leach residues in Na2CO3 solution with air bubble at a temperature of 80 °C. The leaching solution can be used to produce metallic zinc powder by electrowinning after chemical separation of impurities.  相似文献   

16.
Shaotao Cao  Yifei Zhang  Yi Zhang   《Hydrometallurgy》2009,98(3-4):298-303
A new process to produce monosodium aluminate hydrates (MAH) by fast crystallization from the leach liquor of a diasporic bauxite in concentrated NaOH solution is presented. The crystallization of MAH was carried out easily compared to the precipitation of gibbsite and the effect of agitation, initial concentration of sodium aluminate, seed amount and the presence of red mud were systematically studied in a batch crystallizer. The apparent kinetics of crystallization followed a second order rate law with an apparent activation energy for MAH crystallization of 38.0 kJ/mol which implies a surface-diffusion controlled mechanism. X-Ray diffraction and scanning electron microscopy identified the structure of MAH as Na2O·Al2O3·2.5H2O with a flake crystal morphology. The molar ratio α of Na2O to Al2O3 in the MAH products was < 1.2 after a simple wash by dilute sodium aluminate.  相似文献   

17.
Measurements were made on the transport and equilibrium properties of dissolved chlorine in aqueous HC1, HCl-ZnCl2, HCl-MgCl2, and water. These measurements included solubility, absorption rates during bubbling, stripping rates during nitrogen bubbling, and cathodic reduction rates. The solubility of chlorine was found to be affected by speciation into aqueous Cl2, HC1O, and C13 . With increasing HCl concentration, the solubility of chlorine decreased to a minimum at 0.2 mol dm−3, followed by a slow and linear increase. Metal chloride salts depressed the chlorine solubility approximately in proportion to concentration. Mass transfer of aqueous chlorine was found to support a current of about 100 A m−2 from a chlorine-saturated ZnCl2-HCl solution under typical zinc electrowinning conditions. Comparisons with published zinc electrowinning papers indicate that air sparging would eliminate dissolved chlorine sufficiently to remove the need for diaphragm cell separation of dissolved chlorine, insofar as current inefficiencies due to cathodic chlorine reduction are concerned.  相似文献   

18.
A thermodynamic equilibrium model is developed for the distribution of copper between the phases for the system copper-sulfuric acid-xylene-/3-alkenyl 8-hydroxyquinoline. Chemical equilibrium data were obtained for the range of experimental parameters of [Cu] aq, i T (0.005 to 0.02 mol/dm3), [S]T (0.05 to 0.2 mol/dm3), extractant (0.02 to 0.06 mol/dm3), and temperature (20 to 50 ?C). The nonlinear regression analysis best fit model is ln(Dαo/γCusu++) = 3.83 + 642(1/T - 1/308) +2 ln([HR]/[H+H +) whereD is the distribution coefficient, αo is the reciprocal of the degree of formation of copper ion, and γ refers to activity coefficients of ionic species. The analysis of the data considers the aqueous phase ionic equilibria and a technique to estimate the impurities present in the reagent. The results indicate that the enthalpy of the reaction and the entropy change are -5338 J/mol and 14.5 J/mol · K, respectively. Also, the stoichiometric coefficient of two, determined for the extractant with cupric ion, is in agreement with other investigators.  相似文献   

19.
The diffusion coefficients of NiCl2, ZnCl2, and CdCl2 in the aqueous solution systems of MC12 and MC12-HC1 were measured at 298 K using a diaphragm-cell method. The data are listed as a function of molar concentrations of MC12 at the HC1 concentrations of 0, 0.1, 0.5, 1.0, and 2.0 mol dm-3. It was found that the concentration dependencies of the diffusion coefficients for these metal chlorides in single-electrolyte solutions differed from each other. This could be explained in terms of changes in the mean activity coefficients of chloride and in the viscosities of those solutions. The diffusion coefficient of metal chloride in MC12-HC1 solution was greatly affected by the HC1 concentration; however, the behavior of the diffusion coefficients varied, depending on the kind of chloride involved. In NiCl2-HCl solutions, an increase in HC1 con-centration caused a decrease in the diffusion coefficient value, while in ZnCl2-HCl solutions, the addition of 2 mol dm-3 HC1 caused an increase in the diffusion coefficient of ZnCl2. These phenomena are quite different from those of the sulfate systems reported in our previous work. It was also demonstrated that the concentration dependency of the diffusion coefficients of MC12 in aqueous MC12-HC1 solutions could be attributed to the diffusion potential as well as the changes in the mean activity coefficient and viscosity. Masami Aoki, formerly Graduate Student, Department of Met-allurgy, Kyoto University.  相似文献   

20.
A complete characterization was carried out on a jarositic residue from the zinc industry. This residue consists of ammonium jarosite, with some contents of H3O+, Ag+, Pb2+, Na+ and K+ in the alkaline “sites” and, Cu2+ and Zn2+ as a partial substitution of iron. The formula is: [Ag0.001Na0.07K0.02Pb0.007(NH4)0.59(H3O)0.31]Fe3(SO4)2(OH)6. Some contents of franklinite (ZnO·Fe2O3), gunninguite (ZnSO4·H2O) and quartz were also detected. The jarosite is interconnected rhombohedral crystals of 1–2 μm, with a size distribution of particles of 2–100 μm, which could be described by the Rosin–Rammler model.The alkaline decomposition curves exhibit an induction period followed by a progressive conversion period; the experimental data are consistent with the spherical particle with shrinking core model for chemical control. The alkaline decomposition of the ammonium jarosite can be shown by the following stoichiometric formula:NH4Fe3(SO4)2(OH)6(s)+3OH(aq)→(NH)4(aq)++3Fe(OH)3(s)+2SO4(aq)2−.The decomposition (NaOH) presents an order of reaction of 1.1 with respect to the [OH] and an activation energy of 77 kJ mol−1. In NaOH/CN media, the process is of 0.8 order with respect to the OH and 0.15 with respect to the CN. The activation energy was 46 kJ mol−1. Products obtained are amorphous. Franklinite was not affected during the decomposition process. The presence of this phase is indicative that the franklinite acted like a nucleus during the ammonium jarosite precipitation.  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