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.
The possibility of direct reduction of nickel from its sulfides (heazlewoodite, pentlandite, and disulfide) by intrinsic sulfide sulfur in a sodium hydroxide melt is confirmed experimentally. This process can be performed in a temperature range of 550–700°C reaching a 95–98% degree of metallization. The chemistry of utilization reactions of sulfur varies as a function of the duration of the phase contact (8–12 minutes). At the initial stage, sodium polysulfides are accumulated in alkali fusion cake and are decomposed in time, participating in the disproportionation reaction with the formation of monosulfide sulfur. The process of reduction of nickel from sulfides is limited by the intradiffusion kinetics.  相似文献   

2.
The leaching of covellite (CuS), chalcocite (Cu2S), bornite (Cu5FeS4), and chalcopyrite (CuFeS2) was carried out in a small, shaking autoclave in perchloric acid solutions using moderate pressures of oxygen. The temperature range of investigation was 105° to 140°C. It was found that covellite, chalcocite, and bornite leach at approximately similar rates, with chalcopyrite being an order of magnitude slower. It was found that chalcocite leaching can be divided into two stages; first, the rapid transformation to covellite with an activation energy of 1.8 kcal/mole, followed by a slower oxidation stage identified as covelite dissolution with an activation energy of 11.4 kcal/mole. These two stages of leaching were also observed in bornite with chalcocite (or digenite) and covellite appearing as an intermediate step. No such transformations were observed in covellite or chalcopyrite. Two separate reactions were recognized as occurring simultaneously for all four minerals during the oxidation process; an electrochemical reaction yielding elemental sulfur and probably accounting for pits produced on the mineral surface, and a chemical reaction producing sulfate. The first reaction dominates in strongly acidic conditions, being responsible for about 85 pct of the sulfur released from the mineral, but the ratio of sulfate to elemental sulfur formed increases with decreasing acidity. Above 120°C the general oxidation process appears to be inhibited by molten sulfur coating the mineral particles; the sulfate producing reaction, however, is not noticeably affected above this temperature. For chalcopyrite, activation energies were determined separately for the oxygen consumption reaction and for the production of sulfate, with values of 11.3 and 16.0 kcal/mole respectively. This paper is based upon a thesis submitted by F. LOEWEN in partial fulfillment of the requirements of the degree of M.A. Sc. in Metallurgical Engineering at The University of British Columbia.  相似文献   

3.
Experiments on the extraction of sulfur from the sulfides and oxy-sulfides such as FeS, MgS, CaS, BaS, RE2S3 and RE2O2S (RE - rare earth metals) by hydrogen reduction were carried out at 600 – 900°C by means of hydrogen reduction-polarographic analysis method. The reduction degree of various sulfides by H2 at 800°C is in decreasing order as follows: FeS > MgS ≥ RE2S3 > RE2O2S ≥ CaS ? BaS. The equilibrium constants for reaction MgS(s) = [Mg] + [S] and MgO(s) = [Mg] + [O] obtained by using vapour-liquid equilibrium method are: KMgS = 2.0·10?4 at 1873 K, log KMgO = 1.27 - 13670/T, eOMg = 644.5 - 1406000/T.  相似文献   

4.
Energy storage and recovery using the redox reaction of tungsten/tungsten-oxide is proposed. The system will store energy as tungsten metal by reducing the tungsten oxide with hydrogen. Thereafter, steam will be used to reoxidize the metal and recover the hydrogen. The volumetric energy density of W for storing hydrogen by this process is 21 kWh/L based on the lower heating value (LHV) of hydrogen. The main objective of this investigation was to study the kinetics of the reduction process of tungsten oxide (WO3) and determine the optimum parameters for rapid and complete reduction. Theoretical treatment of isothermal kinetics has been extended in the current work to the reduction of tungsten oxide in powder beds. Experiments were carried out using a thermogravimetric technique under isothermal conditions at different temperatures. The reaction at 1073 K (800 °C) was found to take place in the following sequence: WO3 → WO2.9 → WO2.72 → WO2 → W. Expressions for the last three reaction rate constants and activation energies have been calculated based on the fact that the intermediate reactions proceed as a front moving at a certain velocity while the first reaction occurs in the entire bulk of the oxide. The gas–solid reaction kinetics were modeled mathematically in terms of the process parameters. This model of the reduction has been found to be accurate for bed heights above 1.5 mm and hydrogen partial pressures greater than 3 pct, which is ideal for implementing the energy storage concept.  相似文献   

5.
The kinetics of the smelting reduction of iron ore by a graphite crucible and carbon-saturated molten iron was investigated between 1400 °C and 1550 °C, and its reaction phenomena were continuously observed in situ by X-ray fluoroscopy. In the smelting reduction by graphite, it was shown from the observation results that the smelting reduction reaction proceeded by the following two stages: an initial quiet reduction without foaming (stage I) and a following highly active reduction with severe foaming (stage II). At 1500 °C, by the graphite crucible, the reduction rate of iron ore was found to be 8.88×10−5 mol/cm2 · s, and by the molten iron, 8.25×10−5 mol/cm2·s. The activation energies for the reduction by the graphite crucible and the molten iron were 24.1 and 22.9 kcal/mol, respectively. Based on the results of kinetic research and X-ray fluoroscopic observations, it can be concluded that these two types of smelting reduction reactions of iron ore by the graphite crucible and by the molten iron are essentially the same.  相似文献   

6.
《Hydrometallurgy》2007,85(2-4):95-102
Pressure oxidation kinetics of a massive orpiment (As2S3) sample in sulfuric acid solution were systematically investigated. The effects of temperature (170 to 230 °C), mean particle size (48 to 125 μm diameter) and oxygen partial pressure (345 to 1035 kPa) were evaluated. Oxidation rates were found to be affected significantly by changes in temperature and particle size, but were relatively insensitive to changes in oxygen partial pressure. Kinetics appear to be controlled by product layer diffusion over the temperature range of 170 to 210 °C, due to the formation of elemental sulfur on particle surfaces. However, Arrhenius activation energies over this range are large (> 40 kJ/mol) and increase with decreasing temperature, perhaps reflecting the kinetics of sulfur oxidation rather than diffusion. Above 210 °C the rate-controlling step is a surface reaction with an activation energy of approximately 33.8 kJ/mol. The reaction order at 210 °C is approximately 0.2 with respect to oxygen partial pressure.  相似文献   

7.
A novel ironmaking technology is under development at the University of Utah. The purpose of this research was to determine comprehensive kinetics of the flash reduction reaction of magnetite concentrate particles by hydrogen. Experiments were carried out in the temperature range of 1423 K to 1673 K (1150 °C to 1400 °C) with the other experimental variables being hydrogen partial pressure and particle size. The nucleation and growth kinetics expression was found to describe the reduction rate of fine concentrate particles and the reduction kinetics had a 1/2-order dependence on hydrogen partial pressure and an activation energy of 463 kJ/mol. Unexpectedly, large concentrate particles reacted faster at 1423 K and 1473 K (1150 °C and 1200 °C), but the effect of particle size was negligible when the reduction temperature was above 1573 K (1300 °C). A complete reaction rate expression incorporating all these factors was formulated.  相似文献   

8.
The effect of temperature on the dissolution rate of carbon from coke in molten iron was investigated using a sampling technique in the temperature range of 1723?K to 1923?K (1450?°C to 1650?°C). The dissolution rate of carbon from coke in molten iron increased as the temperature increased. At 1923?K (1650?°C), the rate-determining step was the mass transfer of carbon in the boundary layer adjacent to the metal-carbon interface. At 1723?K (1450?°C), the rate-determining step changed from the mass transfer to the interfacial chemical reaction as the reaction proceeded. At 1823?K (1550?°C), both reaction steps affected the apparent reaction rates. Sulfur dissolution did not affect the carbon dissolution rates in molten iron, so it was considered that the sulfur adsorption at the metal/coke interface was not so significant. The apparent activation energy of the carbon dissolution of coke in molten iron was estimated to be 442?kJ/mol.  相似文献   

9.
Reactions of Cu2S with Cu2O, CuS with Cu2O and CuS with CuO in the molten state were examined in the presence of one atmosphere of argon at 1200°C. A rate law of the form,r SO2 =kNSNO was applicable for each reaction system studied. Comparison of the rate constants for the systems, under conditions of similar initial mole fraction of sulfur to mole fraction of oxygen ratios, showed that Cu2O was much more reactive than CuO in its reaction with copper sulfides. These results are incorporated in a mechanism in which Cu2O reacts with the sulfide in the rate determining step. Experiments carried out in the presence of oxygen indicated the importance of a CuO-Cu2O equilibrium in the overall reaction mechanism.  相似文献   

10.
Differnetial thermal analysis (DTA) and thermal gravimetric analysis (TGA), at a heating rate of 10 °C/min, revealed a complete reduction of NiCl2 by hydrogen in a temperature interval of 375 °C to 450 °C. However, addition of 0.1 mass pct of Pd, Cu, or Ni to the sample caused the reduction to occur at considerably lower temperatures, in the rather narrow range of 315 °C to 370 °C. The activation energy of NiCl2 reduction by hydrogen (between 300 °C and 550 °C) without additives is 54 kJ/mol, and with Pd and Cu or Ni added, under isothermal conditions (from 260 °C to 380 °C), is 33 and 50 kJ/mol, respectively. These values confirm a positive effect of additives on the reduction kinetics. The positive effect of Pd is a consequence of the dissociation and spillover of hydrogen, whereas in the case of Cu and Ni(HCOO)2, it is manifested in a decrease in bonds energy in the nickel lattice because of good Cu solubility, and in the formation of artificial nickel nuclei that intensify the reduction, respectively. Scanning electron microscopy (SEM) analysis of nickel powders obtained under isothermal conditions shows relatively rounded spherical particles (0.321 to 0.780 μm in size) of powder samples with additives, and particles of irregular shape (2.085 μm mean size) of the sample without additives. This illustrates the positive effect of Pd, Cu, or Ni added in the reduction process, in decreasing the size of nickel particles and in the production of a more uniform particle shape.  相似文献   

11.
Coordinated fracture mechanics and surface chemistry experiments were carried out to develop further understanding of environment enhanced subcritical crack growth in high strength steels. The kinetics of crack growth were determined for an AISI 4340 steel (tempered at 204°C) in hydrogen and in water, and the kinetics for the reactions of water with the same steel were also determined. A regime of rate limited (Stage II) crack growth was observed in each of the environments. Stage II crack growth was found to be thermally activated, with an apparent activation energy of 14.7 ±2.9 kJ/mole for crack growth in hydrogen, and 33.5 ± 5.0 kJ/mole in water. Fractographic evidence indicated that the fracture path through the microstructure was the same for these environments, and suggested hydrogen to be the embrittling species for environment enhanced crack growth in hydrogen and in water/water vapor. A slow step in the surface reaction of water vapor with steel was identified, and exhibited an activation energy of 36 ± 14 kJ/ mole. This reaction step was identified to be that for the nucleation and growth of oxide. The hydrogen responsible for embrittlement was presumed to be produced during this reaction. On the basis of a comparison of the activation energies, in conjunction with other supporting data, this slow step in the water/metal surface reaction was unambiguously identified as the rate controlling process for crack growth in water/water vapor. The inhibiting effect of oxygen and the influence of water vapor pressure on environment enhanced subcritical crack growth were considered. The influence of segregation of alloying and residual impurity elements on crack growth was also considered.  相似文献   

12.
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   

13.
The reduction of calcium sulfate to produce calcium sulfide is a part of the cyclic process for converting sulfur dioxide to elemental sulfur that is described in Part I. The kinetics of the hydrogen reduction of nickel-catalyzed calcium-sulfate pellets were investigated using a thermogravimetric analysis (TGA) technique at reaction temperatures between 1023 and 1088 K and hydrogen partial pressures between 12.9 and 86.1 kPa. The reactivity of nickel-catalyzed calcium-sulfate pellets was demonstrated by the conversion of 70 pct fresh nickel-catalyzed calcium sulfate to calcium sulfide in 20 minutes at 1073 K under a hydrogen partial pressure of 86.1 kPa. Furthermore, the reactivity remained relatively intact after ten cycles of reactions and regenerations. This observed characteristic of the pellets is important because the solids must be reusable for repeated cycles to avoid generating secondary pollutants. The nucleation and growth rate expression was found to be useful in describing the kinetics of the reaction, which had an activation energy of about 167 kJ/mol (∼40 kcal/mol) in all reaction cycles except for the first regenerated samples that were lower at 146 kJ/mol (35 kcal/mol). The reaction order with respect to hydrogen partial pressure was 0.22 in all cycles with the exception of the first regenerated sample for which it was 0.37.  相似文献   

14.
The oxidation kinetics of molten Cu2S baths, during top lancing with oxygen/nitrogen (argon) mixtures, have been investigated as a function of oxygen partial pressure (0.2 to 0.78), bath temperature (1200 °C to 1300 °C), gas flow rate (1 to 4 L/min), and bath mixing. Surface-tension-driven flows (the Marangoni effect) were observed both visually and photographically. Thus, the oxidation of molten Cu2S was found to progress in two distinct stages, the kinetics of which are limited by the mass transfer of oxygen in the gas phase to the melt surface. During the primary stage, the melt is partially desulfurized while oxygen dissolves in the liquid sulfide. Upon saturation of the melt with oxygen, the secondary stage commences in which surface and bath reactions proceed to generate copper and SO2 electrochemically. A mathematical model of the reaction kinetics has been formulated and tested against the measurements. The results of this study shed light on the process kinetics of the copper blow in a Peirce-Smith converter or Mitsubishi reactor.  相似文献   

15.
The corrosion (dissolution) kinetics of solid molybdenum, niobium, chromium, and yttrium in molten aluminum were investigated at temperatures between 700 °C and 915 °C under hydrodynamic conditions using the rotating disc method. Dissolution was governed by diffusion under laminar flow conditions (in the angular velocity range of 10 to 32 rad/s) regardless of the number of intermetallic compound layers formed at the solid-liquid interface. The solubility limit (C S ), dissolutíon rate constant(K), and diffusion coefficient(D) were determined. It was found that temperature dependencies of the solubility, dissolution rate, and diffusion coefficient for each system obeyed Arrhenius-type relationships; from these, the activation energies were calculated. The single or multiphase intermetallic layer growth occurring at the solid-liquid interface during dissolution in an unsaturated melt between 700 °C and 915 °C was characterized.  相似文献   

16.
The kinetics of decarburization of liquid iron have been studied between 1160 and 1600°C under conditions where mass transport of reactants is not rate determining. Studies with continuously carbon-saturated iron and of iron with varying carbon concentration have been used to show that the slow step at high concentrations of carbon is independent of carbon concentration and is first order with respect to the pressure of CO2. For high purity iron, the forward rate constant, in mole cm2 s-1 atm-1, is given by the equation ln kf = -11,700/T-0.48. It is concluded that the data are consistent with the chemisorption process as the rate limiting step. A marked sensitivity of the rate to trace amounts of sulfur has been found and it is shown that this is consistent with ideal adsorption of sulfur and is in fair accord with the existing measurements of the depression of the surface tension of iron-carbon alloys by sulfur. D. R. Sain was formerly a Graduate Student.  相似文献   

17.
The reaction between Ni3S2 (liquid) and NiO (solid) resulting in the formation of Ni and SO2 was investigated in the temperature range 800° to 1200°C under a reduced pressure of <0.1 mm Hg. From the kinetic studies in the temperature range 950° to 1150°C, the reaction was found to proceed in three stages: i) Up to about 25 pct reduction, the rate of reaction was high and followed approximately a cubic rate law. During this stage, the reaction is thought to be under mixed control. Activation energy for the first 10 pct reduction was found to be approximately 45 kcal per mole. ii) From about 25 to 90 pct reduction, it obeyed the parabolic rate law, with an activation energy of 86±6 kcal per mole. This value is in agreement with the activation energy reported in the literature for the diffusion of sulfur in nickel. iii) Beyond 90 pct reduction, the reaction was very sluggish owing to the poor availability of the reactants. Optimum conditions for preparing nickel sponge by the above reaction and its processing into thin strips have been standardized. Some of the properties of the metal thus produced have also been incorporated.  相似文献   

18.
A novel flash ironmaking process based on hydrogen-containing reduction gases is under development at the University of Utah. The goal of this work was to study the possibility of the re-oxidation of iron particles in a H2-H2O gas mixture in the lower part of the flash reactor from the kinetic point of view. The last stage of hydrogen reduction of iron oxide, i.e., the reduction of wustite, is limited by equilibrium. As the reaction mixture cools down, the re-oxidation of iron could take place because of the decreasing equilibrium constant and the high reactivity of the freshly reduced fine iron particles. The effects of temperature and H2O partial pressure on the re-oxidation rate were examined in the temperature range of 823 K to 973 K (550 °C to 700 °C) and H2O contents of 40 to 100 pct. The nucleation and growth kinetics model was shown to best describe the re-oxidation kinetics. The partial pressure dependence with respect to water vapor was determined to be of first order, and the activation energy of re-oxidation reaction was 146 kJ/mol. A complete rate equation that adequately represents the experimental data was developed.  相似文献   

19.
The kinetics of reduction of nickel chloride with hydrogen were investigated in the temperature range of 533 (260 °) to 788 K (515 °). Most experiments were done with porous NiCl2 granules having ?8+10 mesh (Tyler) size. The effects of temperature, sample size, granule size and hydrogen partial pressure on the reduction kinetics were investigated. In the temperature range 533 (260 °) to 651 K (378 °) the reduction is dominated by chemical kinetic factors. At higher temperatures diffusional effects appear to become quite significant. The temperature-dependence of the chemical kinetic rate constantK is given by $$\log K = 6.744 - \frac{{22,240}}{{2.303RT}};K\,\,is\,\,in\,\,\min ^{ - 1} $$ The activation energy for the reduction was found to be 22,240 cal /mol (or 93,050 J/mol) in the chemical kinetic regime.  相似文献   

20.
Gaseous hydrogen embrittlement of quenched and tempered 4130 steel was studied as a function of temperature from −42° to 164°C in a partially dissociated hydrogen environment at low molecular hydrogen pressures (≈8 × 10−3 torr). Atomic hydrogen was created by dissociation of molecular hydrogen on a hot tungsten filament located near a crack opening. The presence of atomic hydrogen was found to increase the rate of hydrogen-induced, slow crack growth by several orders of magnitude and to significantly alter the temperature dependence of embrittlement from what is observed in the presence of molecular hydrogen alone. Based on a previous study, these observations are interpreted in terms of a difference between the hydrogen-transport reaction step controlling hydrogen-induced, slow crack growth in the molecular hydrogen and the atomic-molecular hydrogen environments. Finally, a comparison is made between the kinetics of hydrogen-induced, slow crack growth observed in the presence of atomic-molecular hydrogen and the kinetics of known, possible hydrogen-transport reactions in an effort to identify the reaction step controlling hydrogen embrittlement in the presence of atomic hydrogen.  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