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 15 毫秒
1.
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   

2.
The dissolution rate of heazlewoodite in nitric acid solution has been determined. The effects of nitric acid concentration, temperature, particle size, stirring intensity and addition of Cu2+ ions have been investigated. Solid residues after leaching were examined by SEM, X-ray diffraction and chemical analysis. In the solutions containing less than 2.0 M HNO3, dissolution was observed to be completely inhibited after 30 min leaching time, and the rate of hydrogen sulphide production was faster than its oxidation to S0 and HSO4?. In 3 M HNO3, an abrupt increase in dissolution rate of Ni3S2 was found. Two different regions of the dissolution of heazlewoodite were observed below and above 50°C. At temperatures below 50°C, the dissolution rate was very slow, even in 3.0 M HNO3 solution, and H2S gas was evolved. Above 50°C, the dissolution rate rapidly increased. Over the temperature interval 60–90°C in 3.0 M HNO3 dissolution followed a linear rate law, and the activation energy was found to be 42.1 kJ mol?1. Most of the oxidized sulphide ion was found in the solution as sulphate. The leaching rate was independent of stirring speed. The rate-controlling step of the Ni3S2 dissolution is the oxidation of hydrogen sulphide to elemental sulphur or sulphate ions on the Ni3S2 surface. Addition of small amounts of Cu2+ ions to the nitric acid acted as catalyst for the dissolution of Ni3S2. Bubbling air through the leach suspension increased the dissolution rate of Ni3S2 in solutions containing less than 2.0 M HNO3.  相似文献   

3.
《Hydrometallurgy》1987,17(2):201-214
The kinetics of dissolution of synthetic Ni3S2 in nitric acid solution containing ⩽ 2.0 M HNO3 in the presence of cupric and ferric ions was investigated. The effects of stirring, particle size, temperature and concentration of cupric and ferric ions were examined. Solid residues at various levels of nickel extraction were examined by SEM, X-ray diffraction, electron microprobe and chemical analysis. The constant dissolution rate of Ni3S2 was attributed to the electrochemical reaction occurring on the surface of flat-plate type Ni3S2 particles. The reaction kinetics were found to be independent of both cupric and ferric ion concentrations up to 1 mM. Cupric ion did not act as a catalyst at temperatures below 60°C. At 80°C cupric and ferric ions show the same catalytic activity. The activation energy in the presence of ferric ions was 103.6 ± 4.2 kJ/mol. A mechanism for Ni3S2 dissolution in the presence of cupric and ferric ions is proposed.  相似文献   

4.
The reduction kinetics from Ni(OH)2 to fine Ni metal powder under hydrothermal and H2 pressure conditions using anthraquinon as an activator were investigated. The reduction ratio increased with temperature up to 225 °C. The H2 molecule was activated by anthraquinone. This activated hydrogen acts as a reduction reagent from Ni2+ to Ni0. The process may proceed by heterogeneous reaction of Ni2+ ion and solid anthraquinon as follows: (1) anthraquinon is hydrated, (2) Ni2+ ion is absorbed on anthraquinon hydride, and (3) this complex decomposes to Ni0 powder, anthraquinon, and H2O. The particle size of nickel powder obtained increased with increasing temperature and with decreasing pH. The average particle size was about 350 nm. The reduction kinetics were in good agreement with a core model equation with surface reaction, that is, 1 − (1 −x)1/3 =kt, wherex is a reaction ratio andt is reaction time. Arrhenius plots showed two slopes with activation energies 65.9 kJ/mol at higher temperature and 377.2 kJ/mol at lower one. This result shows that the hydration of anthraquinon and adsorption of Ni2+ ion onto anthraquinon hydride and decomposition to Ni metal and anthraquinone proceed by consecutive reactions and rate determining steps change in each temperature range.  相似文献   

5.
The simultaneous oxidation-sulfidation rate of nickel has been measured as a function of SO2 pressure (0.04 to 1 atm) in Ar-SO2 gas mixtures at 603°C. The observed corrosion rates are about 107 times faster than the oxidation rate of nickel in oxygen at 1 atm. The product scale consists of an inner Ni3S2 layer and an outer two-phase layer of NiO and Ni3S2. A linear rate law is observed during an initial time period, and the most probable rate-controlling step is dissociation of SO2. An increase in the scale-gas interfacial area increases the corrosion rate during intermediate time periods. With increasing time, parabolic corrosion rates are measured for SO2 pressures of 0.25 and 1 atm. Values of the nickel diffusivity in Ni3S2 calculated from our measured parabolic-rate constants are in good agreement with recently reported values. This agreement indicates that an interconnected Ni3S2 phase in the outer two-phase layer provides rapid transport paths for nickel diffusion through the scale. Formerly a Graduate Student in the Department of Metallurgy and Materials Science at the University of Pennsylvania, Philadelphia, PA  相似文献   

6.
The kinetics and mechanism of the reduction of M3S2 by hydrogen have been investigated between 1133° and 1300°C. When high flow rates of hydrogen and argon or helium bubbling through the melt are maintained the rate-determining step is a chemical process which can be expressed by a rate law of the form $$\begin{gathered} r_{H_2 S} = k_{expt} (N_S - \alpha )^2 p_{H_2 }^{1/2} \hfill \\ p_{H_2 } \geqslant 0.88atm \hfill \\ \end{gathered} $$ where kexpt = 85.1 atm-1/2 min-1, α = 0.17 at 1250°C. The experimental activation energy for this process is 20.1 ±3.0 kcal per mole. These results are discussed in terms of possible catalysis by nickel.  相似文献   

7.
The diffusion couple method and electron probe microanalysis have been used to determine the ternary isotherms of Ni?Cr?S at 600°, 700°, 800°, and 850°C for sulfur compositions lying below 60 at. pct. It has been found that the solid and liquid phases Ni3+x S2 have a very shallow penetration into the diagram whereas the Ni1?x S, Cr3S4, and Cr2S3 phases penetrate very deeply into the diagram. The phase compositions NiCr2S4 and NiCrS3 have been detected near the limit of the latter two phase fields. The binary phases Cr2S3, Cr3S4, Cr5S6, Cr7S8, and CrS have all been identified and their existence ranges and free energies of formation have been established at 700°C. The phase diagram observed and proposed is consistent with the sequence of free energies of formation of nickel and chromium sulfides.  相似文献   

8.
Recovery of pure nickel from nickel sulfide (Ni3S2) was studied by following to completion the hydrogen reduction reaction in the presence of calcium oxide. The effects of reaction temperature, molar ratio of calcium oxide to nickel sulfide, bed depth, and particle size of the nickel sulfide powder on the reaction were experimentally investigated. A simple empirical integrated rate equation describing the relationship among these variables over the temperature range 773 to 973 K was derived. The activation energy for the scavenged reaction was found to be 101.9 kJ from the experimental data. Over the range of experimental conditions, both globular and fibrous forms of metallic nickel were observed.  相似文献   

9.
An investigation has been carried out to study the feasibility of the removal of nickel from ammoniacal solutions using the reducing agent sodium dithionite. The results obtained indicate that dithionite can precipitate over 95% of nickel values in solution in the form of Ni3S2 and metallic nickel. A temperature of 45°C has been found to be optimal for the precipitation process.  相似文献   

10.
The vapor pressure of sulfur over Ni-S melts of various compositions was calculated from the equilibrium weight of the melt in gas streams of known H2S-H2 composition. The Gibbs-Duhem equation was used to calculate the activity of nickel and other thermodynamic properties. For the reaction: 3Ni(S) + S2(g) ⇌ Ni3S2(l) the suggested free energy rslationship is: ΔG° = -57,910 + 15.89T (800° to 1100°C). The Calculations were extrapolated to predict that for the reaction: Ni(s) + 1/2S2(g) ⇌ NiS(l), ΔG° = -26.730 + 10.5T (1000° to 1100°C)  相似文献   

11.
The formation conditions for the recovery of nickel from sulfate solutions as Ni3S2 have been investigated; this sulfide is more reactive than NiS in subsequent leaching operations. Hydrogen sulfide gas at atmospheric pressure is introduced into a NiSO4-Na2SO4-MgSO4-Al2(SO4)3 solution in the presence of reduced iron powder. Although the formation of Ni3S2 is compctitive with that of NiS, the nickel precipitation efficiency and the ratio of nickel as Ni3S2 to the total nickel precipitated reached 99.5 to 99.9 and 90 to 95 pct, respectively, under the following conditions: 363 K, Ph2s 31 kPa, Ni2+ 4.0 g · dm-3, 3[Feo]/[Ni2+] 1.25 to 1.5, H2S flow rate 70 to 100 cm3 · min-1, and 45 to 60 minutes retention time. Selective formation of Ni3S2 is achieved within 10 minutes, and a reaction on the surface of the iron is rate-determining during the early stages of precipitation. Since the iron is almost totally consumed after 1 to 2 hours of reaction, the precipitated Ni3S2 is gradually converted to NiS. Calculations considering the buffer action of sulfate ion and sulfate complex formation with polyvalent metal cations as well as with nickel ions confirmed that significant nickel precipitation as Ni3S2 should occur under the test conditions.  相似文献   

12.
Combustion synthesis (SHS) of Ni3Ti-TiB2 metal matrix composites (MMCs) was selected to investigate the effect of gravity in a reaction system that produced a light, solid ceramic particle (TiB2) synthesized in situ in a large volume (>50 pct) of the liquid metallic matrix (Ni3Ti). The effects of composition, green density of pellets, and nickel particle size on the combustion characteristics are presented. Combustion reaction temperature, wave velocity, and combustion behavior changed drastically with change in reaction parameters. Two types of density effects were observed when different nickel particle sizes were used. The structures of the combustion zones were characterized using temperature profile analysis. The combustion zone can be divided into preflame, reaction, and after-burning zones. The combustion mechanism was studied by quenching the combustion front. It was found that the combustion reactions proceeded in the following sequence: formation of liquid Ni-Ti eutectic at 940 °C → Ni3Ti+NiTi phases → reduction of NiTi with B→TiB2+Ni3Ti.  相似文献   

13.
Reaction synthesis of Ni-Al-based particle composite coatings   总被引:1,自引:0,他引:1  
Electrodeposited metal matrix/metal particle composite (EMMC) coatings were produced with a nickel matrix and aluminum particles. By optimizing the process parameters, coatings were deposited with 20 vol pct aluminum particles. Coating morphology and composition were characterized using light optical microscopy (LOM), scanning electron microscopy (SEM), and electron probe microanalysis (EPMA). Differential thermal analysis (DTA) was employed to study reactive phase formation. The effect of heat treatment on coating phase formation was studied in the temperature range 415 °C to 1000 °C. Long-time exposure at low temperature results in the formation of several intermetallic phases at the Ni matrix/Al particle interfaces and concentrically around the original Al particles. Upon heating to the 500 °C to 600 °C range, the aluminum particles react with the nickel matrix to form NiAl islands within the Ni matrix. When exposed to higher temperatures (600 °C to 1000 °C), diffusional reaction between NiAl and nickel produces (γ′)Ni3Al. The final equilibrium microstructure consists of blocks of (γ′)Ni3Al in a γ(Ni) solid solution matrix, with small pores also present. Pore formation is explained based on local density changes during intermetallic phase formation, and microstructural development is discussed with reference to reaction synthesis of bulk nickel aluminides.  相似文献   

14.
The kinetics and mechanism of reactions of molten Fe, Co, Cu, and Pb sulfides with hydrogen have been investigated between 1080° and 1340°C. While maintaining high flow rates of hydrogen and argon to ensure chemical control, the rate of reduction of these sulfides was found to be second order in sulfur concentration and half-order in hydrogen pressure. These dependences are similar to those reported earlier for molten nickel sulfide reduction by hydrogen. The kinetics and mechanistic proposal suggest a metal catalyzed step. If the activation energies for the rate determining step of S2 reacting with hydrogen in the melt for the various sulfides are compared with the values for the hydrogen-liquid sulfur or hydrogen-gaseous sulfur reactions, the order of catalytic activity seems to follow generally the order of the pure metals themselves as found for other processes involving activation of molecular hydrogen.  相似文献   

15.
The formation of the Ni3Al layer in NiAl (55 at. pct Ni)-pure Ni diffusion couples at temperatures above 1000°C has been found to be controlled almost completely by volume diffusion. At 1000°C and below, the relatively small grain size of the Ni3Al compound in the layers caused such a large contribution from grain boundary diffusion, that the layer growth rates at 1000°C exceeded those at 1100°C and even those at 1200°C. In Ni3Al (75at. pct Ni)-pure Ni diffusion couples the Ni3Al compound rapidly converted into the solid solution of aluminum in nickel. Volume-diffusion coefficients calculated by the Boltzmann-Matano method yielded heats of activation of 55, 64, and 65 kcal·mol?1 for NiAl, Ni3Al and the solid solution of aluminum in nickel, respectively. In addition, eleven different types of diffusion couples were prepared from various Ni?Al alloys and annealed at 1000°C. Marker interface displacements and observations of porosity in these couples yielded a more detailed picture of the Kirkendall-effect than earlier work had done. The ratio of the intrinsic diffusion coefficients at the marker interface,D NI/D Al, is greater than one in the nickel-rich NiAl phase. For the Ni3Al phase no statement can be made on the basis of this work. When the marker interface is located in the nickel solid solution,D Ni/D Al is smaller than one. The phase boundary concentrations in these couples did not show the expected deviation from the equilibrium concentrations in two-phase alloys; this finding is discussed with regard to the free-energycomposition diagram.  相似文献   

16.
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   

17.
The activity of sulfur in the Fe-S and Ni-S binary and the Fe-Ni-S ternary systems was investigated at 1250°C by equilibrating molten mattes with gaseous mixtures of H2S and H2. From the sulfur activity data the activities of iron and nickel in the Fe-S and Ni-S binary systems as well as the activities of FeS and Ni3S2 in the ternary system were calculated. Isoactivity diagrams are presented for S, FeS, and Ni3S2 in the Fe-Ni-S ternary system. Some comments are offered on the industrial problems of nickel slag loss and ferronickel separation.  相似文献   

18.
Published measurements of sulfur vapor pressure were used to determine thermodynamic properties of the solid Ni-S system above 800 K. They were Gibbs-Duhem integrated to estimate the formation properties of stoichiometric Ni3S2, NiS, and NiS2. Statistical thermodynamics was applied to find partition functions, interaction energies, and free energies to find possible atom arrangements in Nix S2 and Ni1−x S. Theoretical calculations indicated that nickel vacancies may exist in quasichemical order in the former phase and randomly in the latter phase. A strong indication was that Nix S2 should have a sulfur bcc structure with nickel atoms distributed in octahedral sites having a Ni5S4 configuration. The possible existence of sulfur vacancies was theoretically investigated.  相似文献   

19.
The selective chloridization of nickel in a lateritic iron ore by gaseous HCl is based on the principle of relative thermal stability of iron and nickel chlorides. This aspect has been discussed with differential thermal analysis (DTA) and thermogravimetric (TG) data of the hydrated chlorides of iron and nickel. The kinetics of chloridization of nickel in a lateritic nickel ore from Orissa, India, have been studied by using both pure HCl (g) and the HCl (g) + N2 mixture. The sharp decrease in the rate of chloridization of nickel at temperatures above 250 °C is attributed to the rapid decomposition of molten ferric chloride hydrate (FeCl3 · 3H2O), which blocks the pores of the reactant solid. Therefore, kinetics of chloridization follow both the pore-blocking model (logarithmic rate law) and diffusion-controlled mechanisms. Very low values of apparent activation energy and effective diffusivity derived from the rate constants of the diffusion-controlled process suggest that diffusion of HCl (g) takes place either in a dissolved state in the molten ferric chloride (at 100 °C to 150 °C) or through cracks and fissures formed on the surface due to rapid decomposition of ferric chloride at 200 °C to 250 °C. Because of the complexity of the reaction system, the rate of chloridization of nickel is almost independent of grain size.  相似文献   

20.
The macrokinetic regularities of the reactivity of synthesized Ni–Re (20 and 60 wt %) alloys in a sulfuric acid solution (100 g/L, 25–40°C) during direct current polarization are studied using physicochemical methods. The phase composition of the synthesized alloys is determined by the formation of solid solutions as a function of the initial Ni/Re weight ratio. These are two types of nickel solid solutions (Ni16Re0.2 and Ni14Re0.9) and one rhenium solution (Ni1.1Re). These solid solutions are anodically oxidized in the sequence of their structural rearrangement Ni16Re0.2 → Ni14Re0.9 → Ni1.1Re with a combined transition of the metals into an electrolyte solution. These solid solutions provide the reduction of Ni3+ to Ni2+ due to the depolarization ability of rhenium, being their component.  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