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 640 毫秒
1.
Sulfation studies were conducted on CoO samples coated with a thin film of Na2SO4 and exposed at 600 to 900 °C in O2-SO2-SO3 environments containing 0.02 to 2 pct (SO2+SO3). Within this range of conditions, Co3O4 (s), CoSO4 (s), and liquid Na2SO4-CoSO4 solutions were observed as reaction products. At lower concentrations of SO3 where CoSO4 (s) was unstable, the reaction product consisted of a thick Co3O4 layer close to the CoO, a layer of Na2SO4-CoSO4 liquid, and large Co3O4 particles at the gas/salt interface. At higher concentrations of SO3, the reaction procuct was a two phase mixture of solid CoSO4 and an Na2SO4-CoSO4 liquid. In gas mixtures containing 0.15 pct and 1 pct (SO2+SO3), the highest reaction rates were observed at about 750 to 800 °C. Some of these results are quite similar to those observed during low temperature hot corrosion of cobalt-base alloys. The overall reaction mechanism has been described in terms of two processes: oxidation of CoO to Co3O4 and sulfation of cobalt oxides.  相似文献   

2.
Accelerated oxidation tests have been conducted on a number of Co-Cr, Co-Al, and Co-Cr-Al alloys coated with a thin film of Na2SO4 and exposed at 600 to 750 °C in O2-SO2-SO3 environments containing 0.0095-5 pct (SO2 + SO3). Generally, Co-Cr and Co-Cr-Al alloys reacted nonuniformly, usually in the form of pits, and Co-Al alloys suffered broad frontal attack. The morphology of the reaction product was observed to be dependent on temperature andP SO3 Under all conditions, a thin sulfur-rich band containing sulfides was observed at the alloy/scale interface, and cobalt dissolved near this interface and formed Co3O4 and/or CoSO4(s) in different regions of the reaction product.  相似文献   

3.
The kinetics of dissolution of cobalt oxides Co2O3 and Co3O4 in aqueous solutions of acids (H2SO4, EDTA) is experimentally studied. Dissolution rate W increases with the temperature or the EDTA concentration. The reaction orders of dissolution for hydrogen ions in sulfuric acid and EDTA (dlogW/dpH = 0.5 ± 0.1) and for anions (dlogW/dlog[An ] = 0.5 ± 0.1) are determined. A specific feature of the dissolution kinetics in EDTA is a maximum in the dissolution rate of the cobalt oxides at pH −1. The activation energy of the process E a is 70 kJ/mol in H2SO4 and 60 kJ/mol in EDTA. The modeling of the process shows that the CoOH+ ion is a surface particle controlling the dissolution rate in mineral acids and the CoHY ion, in the complexone.  相似文献   

4.
The modeling of dissolution of cobalt, nickel, and iron oxides in sulfuric acid shows that the rate-determining step of the dissolution is the passage of oxide complexes formed on the surface of oxide particles to a solution. A system analysis of the dissolution curves (??-??) of the oxides is developed to calculate the kinetic parameters ( $n_{H^ + } $ ,E a). Cobalt oxide Co3O4 dissolves in H2SO4 more rapidly than nickel and iron oxides (Ni2O3 and Fe2O3).  相似文献   

5.
《Hydrometallurgy》2005,76(3-4):225-232
The recycling of silver from metallic scraps can be performed through O3 leaching at an ambient temperature and low (∼0.1 M) H2SO4 concentration. The main by-product is O2, which can be recycled to the O3 generation or used as leaching agent in a pretreatment step. The stoichiometry and the effects of the stirring speed, ozone and acid concentration and temperature on the leaching of silver were investigated. Silver dissolved as Ag2+(aq) in the range 10−3–1 M H2SO4, but for pH ≥4, insoluble Ag2O2 was the main reaction product. Kinetics appeared to be controlled by mass transfer of O3(aq) to the solid–liquid interface, showing first order dependency with respect to [O3]aq and PO3. Specific rates were only slightly dependent on the temperature in the interval 10–50 °C, but decreased at 60 °C due to the fall in O3 solubility. The mass transfer coefficients showed an average activation energy of 17 kJ/mol. No significant effect of [H2SO4] on mass transfer coefficients was observed for 10−2–1 M. Leaching rate gradually diminished for pH >2, as a consequence of the influence of the [H+] in the transport control.  相似文献   

6.
Leaching of natural bornite in a sulfuric acid solution with oxygen as oxidant was investigated using the parameters: temperature, particle size, initial concentration of ferrous, ferric and cupric ions, and using microscopic, X-ray and electronprobe microanalysis to characterize the reaction products. Additionally, stirring rate, pH and PO2 were varied. Dissolution curves for percent copper extracted as a function of time were sigmoidal in shape with three distinct periods of reaction: induction, autocatalytic and post-autocatalytic which levelled off at 28% dissolution of copper. The length of the induction period was not reproducible, causing the dissolution curves to be shifted with respect to time. The dissolution curves in the autocatalytic and post-autocatalytic regions were reproducible, and this property was utilized to treat much of the kinetic data. The iron dissolution curves had four dissolution regions. An initial small but rapid release of iron to solution preceded the three periods just given for copper dissolution. Aside from this initial iron release, the iron and copper dissolution curves were almost identical.Stirring rate had no effect on dissolution of copper above 400 min?1 nor did oxygen flow rate in the range 20–40 cm3/min. Dissolution rate was slightly dependent on oxygen partial pressure for PO2 < 0.67. Hydrogen ion concentration had no effect except that sufficient acid was required to prevent hydrolysis and precipitation of iron salts.The dissolution rate was directly dependent on the reciprocal of particle diameter indicating possible surface chemical reaction control, but the activation energy of 35.9 kJ/mol (8.58 kcal/mol) for the autocatalytic region of copper dissolution is slightly too small for that, though not unreasonable. Initial addition of Fe2+ had a rather complex effect and markedly enhanced dissolution of copper, as also did initial addition of Fe3+. Microscopic analysis showed nuclei of two new phases, covellite and Cu3FeS4, in the induction region. The new phases grow rapidly in the autocatalytic stage, which is controlled by nuclei formation and chemical reaction. The post-autocatalytic region is characterized by complete transformation of bornite into covellite on the particle surfaces and Cu3FeS4 as an internal product with an X-ray spectrum very similar to that of chalcopyrite. The post-autocatalytic region is controlled by autocatalytic growth of newly formed phases. Further reaction beyond the autocatalytic region (percent copper dissolution > 28%) occurs so slowly with oxygen as oxidant that it was not studied.The rate of copper dissolution appears to be controlled by the rate of iron dissolution. Using that and the other experimental evidence a mechanism for reaction is proposed in which iron-deficient bornite, Cu5Fe?S4, is formed on the surface by initial preferential iron dissolution. Labile Cu+ diffuses into this from Cu5Fe?SO4 and unreacted bornite to produce CuS on the surface. Depletion of labile Cu+ ions from Cu5FeS4 produces Cu3FeS4 in the interior of the mineral particles.  相似文献   

7.
This paper discusses the leaching of enargite (Cu3AsS4) in sulfuric acid–sodium chloride solution using oxygen as oxidant at atmospheric pressure. The dissolution of arsenic from enargite in this medium proceeds with elemental sulfur as a reaction product according to:2Cu3AsS4 + 6H2SO4 + 5.5O2 → 6CuSO4 + 2H3AsO4 + 8S° + 3H2OThe dissolution rate of arsenic was found to be very slow. About 6% of arsenic was dissolved in 7 h of leaching at 100 °C in a solution containing 0.25 M H2SO4 and 1.5 M NaCl under a flow of oxygen of 0.3 l/min.The kinetics of arsenic dissolution is well represented by a shrinking core model for spherical particles controlled by surface reaction. An apparent energy of activation of 65 kJ/mol was obtained for the temperature range 80 to 100 °C.  相似文献   

8.
According to the “oxide ion” theory, sulfidation attack does not occur until oxide ions present in the fused Na2SO4 melt react with the normally protective oxide scale. It has already been shown that chromia reacts with and decreases the oxide ion content of sodium sulfate and inhibits sulfidation attack. Based upon the results reported herein, the reduction of the oxide ion content of sodium sulfate is a necessary but not sufficient condition for sulfidation inhibition. It is shown that the oxides of molybdenum as well as vanadium react with and decrease the oxide ion content of Na2SO4. It is shown that the addition of either Mo or V to nickel imparts sulfidation resistance. However, it is also shown that whereas the addition of Cr2O3 to Na2SO4-coated nickel-base superalloys prevents or inhibits sulfidation attack, no beneficial effects are noted when either MoO3 or V2O5 are codeposited with Na2SO4 onto nickel-base superalloy substrates. The reactions between V2O5 with metal oxides were also studied. V2O5 readily fluxes Al2O3 and slowly reacts with Na2SO4. The relationship between accelerated oxidation, oxide ion content of a fused melt and the fluxing of the normally protective oxide scale by liquid metal oxides is discussed.  相似文献   

9.
The reaction of cobalt with 1 pct Cl2 in 1,10, and 50 pct O2/Ar atmospheres has been studied at 650 °C with thermogravimetry and mass spectrometry. The principal vapor species appear to be CoCl2 and CoCl3. In all cases, CoCl2 (s) forms at the oxide/metal interface and equilibration of the volatile chlorides with Co3O4 does not occur in the early stages of the reaction. In the 1 pct C12-1 pct O2-Ar case, continuous volatilization occurs. In the 1 pct C12-10 pct O2-Ar and 1 pct C12-50 pct O2-Ar cases, volatilization occurs only in the first few minutes of reaction. Afterwards, the reaction is predominantly oxidation.  相似文献   

10.
In an acidified ferric chloride solution, bornite leaches in two stages of reaction with the first being relatively much more rapid than the second; the first terminates at 28 pct copper dissolution. The first-stage dissolution reaction is electrochemical and is mixed kinetics-controlled; ferric-ion transfer through the solution boundary layer and reduction on the surface to release Cu2+ into solution are both important in controlling the rate. The concentration of labile Cu+ in the bornite lattice governs the potential of the surface reaction, and, once Cu+ is depleted from the original bornite, stage-I reaction ceases. The solid reaction intermediate formed is Cu3FeS4. Minute subcrystallites formed at the latter part of stage I leach topochemically in stage II. This reaction which commences at 28 pct Cu dissolution is characterized by a change in mechanism at about 40 pct copper dissolution, though the overall chemical equation for reaction is unchanged in stage II; cupric and ferrous ions and sulfur as a solid residue are products of reaction. The region 28 to about 40 pct Cu dissolution is designated as a transition period to stage-II reaction. Reaction rate in this period is interpreted as being controlled by reduction of Fe3+ on active product sulfur surface sites, and hence the reaction rate is controlled by the rate of nucleation and growth of sulfur on the Cu3FeS4 intermediate surfaces. Strain in the Cu3FeS4 crystal lattice is released during this period by diffusion from the lattice of Cu+ remaining from the labile copper initially present in the bornite. After about 40 pct Cu dissolution the rate of reaction is controlled by diffusion through the fully formed sulfur layer in an equiaxial geometrically controlled reaction.  相似文献   

11.
《Hydrometallurgy》2007,89(1-4):154-169
The effect of the change in phase constitution, particle size distribution, surface area, crystallite size, strain and lattice parameters introduced by mechanical activation of the altered beach sand ilmenite from Manavalakurichi region, India on the dissolution kinetics of HCl and H2SO4 was investigated. The altered ilmenite showed different physico-chemical characteristics and was found to be more resistant to acid leaching than the less altered ilmenite from the Chatrapur beach sands, India investigated earlier. The dissolution behavior was also different in H2SO4 and HCl. For sulfuric acid leaching, the dissolution of Fe and Ti increased monotonically with time of milling and showed a continuous increase with time of leaching, whereas hydrolysis of titanium occurs in HCl medium, especially for the activated samples at lower acid concentration, lower solid to liquid ratio and higher temperature leading to lower solution recoveries. The dissolution kinetics in both H2SO4 and HCl prior to hydrolysis conforms initially to the reaction rate control model and for higher leaching times to the shrinking core model where diffusion through the product layer is rate controlling. It is postulated that the anatase formed by hydrolysis in milled samples impedes the further progress of leaching. The activation energies for the dissolution of Fe and Ti decreased with time of milling and were marginally lower in HCl than in H2SO4. An attempt has also been made to correlate the decrease in activation energy to the increase in the energy input to the material through high-energy milling. The relative contribution of the increase in surface area and structural disorder on the enhancement of the dissolution rates has been evaluated.  相似文献   

12.
High-temperature experiments and Refractory-Slag-Metal-Inclusion (ReSMI) multiphase reaction simulations were carried out to determine the effect of the ladle slag composition on the formation behavior of non-metallic inclusions in molten steel. Immediately after the slag-metal reaction, magnesium migrated to the molten steel and a MgAl2O4 spinel inclusion was formed due to a reaction between magnesium and alumina inclusions. However, the spinel inclusion changed entirely into a liquid oxide inclusion via the transfer of calcium from slag to metal in the final stage of the reaction. Calcium transfer from slag to metal was more enhanced for lower SiO2 content in the slag. Consequently, the spinel inclusion was modified to form a liquid CaO-Al2O3-MgO-SiO2 inclusion, which is harmless under steelmaking conditions. The modification reaction was more efficient as the SiO2 content in the slag decreases.  相似文献   

13.
Abstract

The reactions that occur when Co3O4 is sulphated. By sulphur trioxide in the presence of sodium sulphate and/or sodium chloride have been studied by differential thermal analysis. Sodium chloride forms a eutectic with sodium sulphate, lowering its melting point to a temperature within the roasting range. This facilitates the dissolution of SO3 in the melt and thus promotes the formation of sodium pyrosulphate the sulphation agent. In the absence of added sodium sulphate, sodium chloride reacts with sulphur trioxide to produce Na2SO4, which is, in turn, converted to sodium-pyrosulphate.

Résumé

Les réactions se déroulant lors de sulfatation du Co3O4 par l'anhydride sulfureux en présence de sulfate de sodium ou de chlorure de sodium ont été étudiees par analyse thermique différentielle. Le chlorure de sodium forme un eutectique avec le sulfate de sodium, abaissant son point de fusion jusqu'à une température située dans l'intervalle de température du grillage. Ceci facilite la dissolution du SO3 dans ce liquide et ainsi favorise la formation de pyrosulfate de sodium, l'agent sulfatant. Sans sulfate de sodium, le chlorure de sodium réagit avec le SO3 pour former du Na2SO4, qui est, à son tour, transformé en pyrosulfate de sodium.  相似文献   

14.
One of the most frequent causes of refractoriness in precious metals leaching is their occlusion or fine dissemination into a pyritic matrix. This study experimentally explores the acid leaching of pyrite with ozone, suggests the stoichiometry of the reaction, estimates its activation energy and defines the effect of the main variables on the leaching kinetics. The results of stoichiometry tests allow establishing that one mole of pyrite requires 7.7 moles of ozone to produce one mole of ferric ion and 2 moles of HSO4? ions. A decrease in the particle size, solution pH and solids’ concentration of the leaching system increases pyrite dissolution. The type of acid (nitric, sulphuric and hydrochloric) does not affect pyrite dissolution rate. Up to 60% of pyrite is dissolved when the optimal experimental conditions are employed (1?g pyrite (?25?µm), 800?mL of 0.18?M of H2SO4, 800?rev?min?1, 1.2?L?min?1 gas stream O2/O3 with 0.079?g O3?L?1 and 25°C). The apparent activation energy of the pyrite-ozone reaction is 14.92?kJ?mol?1, and the absence of a passive layer on the pyrite surface and the linearity of the dissolution profiles suggest that the dissolution kinetics is controlled by the chemical reaction.  相似文献   

15.
The refractory–slag–metal–inclusion multiphase reaction model was developed by integrating the refractory–slag, slag–metal, and metal–inclusion elementary reactions in order to predict the evolution of inclusions during the secondary refining processes. The mass transfer coefficient in the metal and slag phase, and the mass transfer coefficient of MgO in the slag were employed in the present multiphase reactions modeling. The “Effective Equilibrium Reaction Zone (EERZ) Model” was basically employed. In this model, the reaction zone volume per unit step for metal and slag phase, which is dependent on the ‘effective reaction zone depth’ in each phase, should be defined. Thus, we evaluated the effective reaction zone depth from the mass transfer coefficient in metal and slag phase at 1873 K (1600 °C) for the desulfurization reaction which was measured in the present study. Because the dissolution rate of MgO from the refractory to slag phase is one of the key factors affecting the slag composition, the mass transfer coefficient of MgO in the ladle slag was also experimentally determined. The calculated results for the variation of the composition of slag and molten steel as a function of reaction time were in good agreement with the experimental results. The MgAl2O4 spinel inclusion was observed at the early to middle stage of the reaction, whereas the liquid oxide inclusion was mainly observed at the final stage of the refining reaction. The content of CaO sharply increased, and the SiO2 content increased mildly with the increasing reaction time, while the content of Al2O3 in the inclusion drastically decreased. Even though there is slight difference between the calculated and measured results, the refractory–slag–metal multiphase reaction model constructed in the present study exhibited a good predictability of the inclusion evolution during ladle refining process.  相似文献   

16.
The kinetics of the evolution of SO2 gas from a liquid synthetic blast furnace -type slag (CaO-Al2O3-SiO2) in an atmosphere of O2 + Ar gas at a total pressure of 1 atm for 0.003 ≤ Po2 ≤ 1.00 atm has been studied in the range 1360 to 1460°C. The process has been followed by collecting and analyzing the SO2 as it forms, and also by observing the change in weight of the slag sample with time. The effect of slag composition has also been studied. For partial pressures of oxygen less than about 0.1 atm, the rate is very rapid and is controlled by transport in the gas phase. At greater values of Po2, the rate is much slower and is controlled by a chemical process. In the high Po2 region, the process is half-order with respect to the concentration of sulfur in the slag. This half-order dependence on sulfur concentration in the slag may be explained by an initial fast irreversible reaction to form two intermediate species which then decompose at equal rates to give the final products. Additions to the slag of iron or manganese oxides greatly accelerate the rate of evolution of SO2 at Po2 = 1.00 atm. This is interpreted to mean that a charge transfer process, possibly involving S2- and O2- ions, is rate-controlling at Po2 = 1 atm. It is also apparent that Fe2+ and Fe3+ (or Mn2+ and Mn3+) ions can act as charge carriers. Some measurements with actual industrial blast furnace slags are also reported. Formerly Visiting Scientist, Massachusetts Institute of Technology  相似文献   

17.
The dissolution behaviour of alumina, MgO and MgAl2O4 particles in a 36%CaO‐21%Al2O3‐42%SiO2 (mass contents in %) slag has been studied using a confocal scanning laser microscope in the temperature range of 1470 ‐ 1550 °C. The double hot thermocouple technique and metallographic examination was also used to characterize the reaction product between the particles and the slag. The total dissolution times for MgO and MgAl2O4 particles decreased with increased temperature and were almost identical (approximately 200 s for a 150 μm diameter particle at 1500 °C.). This phenomenon can be explained by the formation of MgAl2O4 layer on the surface of the MgO particle during dissolution. Alumina dissolution times were slightly longer (approximately 250 s for a 150 μm diameter particle at 1500 °C) than that of MgO and MgAl2O4 particles and no reaction layer was observed on the surface of the particles.  相似文献   

18.
When coal is burned in the presence of limestone in an atmospheric fluidized-bed combustor (AFBC), the sulfur emission can be kept below acceptable EPA levels. Calcining of the limestone produces CaO, which then forms solid CaSO4 by a reaction with the SO2 produced during coal combustion. The internal components (e.g., heat exchanger tubes) of the bed, however, become coated with a compact layer of CaSO4, CaO, and ash during combustion. It has been suggested that the presence of the sulfate on these hot metal surfaces is the cause of observed instances of accelerated oxidation-sulfidation. This paper presents results which support the above suggestion. The reactions between Cr, Ni, Co, Fe, alloy 800, 2.25 Cr-1 Mo, 9 Cr-1 Mo steels, or 304 stainless steel with CaSO4 were studied using differential thermal and thermogravimetric analyses. The reaction products were analyzed using X-ray diffraction, optical microscopy, and in some instances, X-ray energy dispersive analyses. The chromium-calcium sulfate reaction is the only case studied in which a sulfide is not formed. In that case, CaCr2O4 is the reaction product. In all other cases, the reactions are oxidation-sulfidation processes.  相似文献   

19.
The formation of a liquid phase during the early stages of the roasting reaction is a common problem in the sodium chromate manufacturing process. The molten salt phase, which is primarily constituted of a binary mixture of Na2CrO4 and Na2CO3, creates major operational problems such as the granulation and blocking of the kilns. In addition to the operational problems, it was observed that the molten salt also affects the transport of oxygen toward the reaction interface. The mechanism of the soda-ash roasting reaction has been analyzed for improving the yield of sodium chromate. It was observed that the conversion efficiency of the roasting process changed dramatically, depending on the origin and the type of the chromite ores used. Thermal and scanning electron microscopic analyses of the products of roasting were carried out to establish the reaction mechanism. It was observed that the presence of silicates in the chromite ores interferes with the formation of sodium chromate involving the binary Na2CO3-Na2CrO4 liquid. The roasting reaction proceeds in a certain crystallographic direction in the chromite spinel in the presence of a nonsilicate molten salt, whereas a complete dissolution of chromite appears to take place in the binary liquid containing silicate phases present in the ore. This article is based on a presentation given in the Mills Symposium entitled “Metals, Slags, Glasses: High Temperature Properties & Phenomena,” which took place at The Institute of Materials in London, England, on August 22–23, 2002.  相似文献   

20.
Chalcopyrite reacts readily with SO3 at about 100°C to form water-soluble sulfates; the reaction is approximately: 3CuFeS2+26SO3→3CuSO4+FeSO4+Fe2(SO4)3+25SO2 The presence of about 4 pct O2 in the gas phase greatly accelerates the reaction presumably due to the complete transformation of ferrous into ferric sulfate in an extremely porous form: 2CuFeS2+17SO3+1/2O2→2CuSO4+Fe2(SO4)3+16SO2 A stoichiometric mixture of SO2+1/2O2 behaves towards chalcopyrite in nearly the same way as SO3 although only in the temperature range 350° to 700°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