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.
Experimental studies on phase equilibria in the multicomponent system PbO-ZnO-CaO-SiO2-FeO-Fe2O3 in air have been conducted to characterize the phase relations of a complex slag system used in commercial lead oxidation smelting. The liquidus in the pseudo-ternary section ZnO-“Fe2O3”-(PbO + CaO + SiO2) with the CaO/SiO2 weight ratio of 0.35 and the PbO/(CaO + SiO2) weight ratio of 5.0 has been constructed using results of over 100 high-temperature equilibration and quenching experiments followed by electron probe X-ray microanalysis. The liquidus in this pseudoternary section contains primary phase fields of spinel (zinc ferrite) Zn x Fe3−x O4+y , zincite Zn u Fe1−u O, melilite Pb v Ca2−v Zn w Fe1−w Si2O7, hematite Fe2O3, magneto-plumbite PbFe10O16, and dicalcium silicate Ca2−t Pb t SiO4. The laboratory results are compared with the slags obtained from an industrial reactor.  相似文献   

2.
Modified coulometric titrations on the galvanic cell: O in liquid Bi, Sn or Ge/ZrO2( + CaO)/Air, Pt, were performed to determine the oxygen activities in liquid bismuth and tin at 973, 1073 and 1173 and in liquid germanium at 1233 and 1373 K. The standard Gibbs energy of solution of oxygen in liquid bismuth, tin and germanium for 1/2 O2 (1 atm) →O (1 at. pct) were determined respectively to be ΔG° (in Bi) = −24450 + 3.42T (±200), cal· g-atom−1 = − 102310 + 14.29T (±900), J·g-atom−1, ΔG° (in Sn) = −42140 + 4.90T (±350), cal· g-aton−1 = −176300 + 20.52T (± 1500), J-g-atom−1, ΔG° (inGe) = −42310 + 5.31 7 (±300), cal·g-atom−1 = −177020 + 22.21T(± 1300), J· g-atom−1, where the reference state for dissolved oxygen was an infinitely dilute solution. It was reconfirmed that the modified coulometric titration method proposed previously by two of the present authors produced far more reliable results than those reported by other investigators. TOYOKAZU SANO, formerly a Graduate Student, Osaka University  相似文献   

3.
The depressions of the freezing temperatures of MgF2, CaF2, and BaF2 by adding 2MO · SiO2, 3MO · 2SiO2, MO · SiO2, 2MO · 3SiO2, and MO · 2SiO2 where M = Mg, Ca and Ba, have been measured, as have the depressions of the freezing temperature of PbF2 resulting from additions of 2Pb · SiO2, 3PbO · 2SiO2 and PbO · SiO2. The variations of the activities of the fluorides with liquidus composition have been calculated. These are shown to be in good agreement with a proposed theoretical model of the constitutions of these melts. In the alkaline earth systems with MO/SiO2 > 1.5 linear chain silicate ions and free F ions are postulated and in melts of MO/SiO2 < 1.5 reaction between F and silicate ions to form polyfluorosilicate anions is postulated. These conclusions are in agreement with those drawn from infrared absorption studies of CaF2-CaO-SiO2 glasses. The activity behavior in the reciprocal systems MF2-M′O · SiO2 and M′F2-MO · SiO2 is explained in terms of polymerization and preferred ionic association effects within the melts. In the lead fluoride-silicate systems fluorination of the silicate ions occurs at PbO/SiO2 = 2 and, in contrast with the alkali and alkaline earth systems, it appears that polyfluorosilicate anions and free O2− anions can coexist in lead fluorosilicate melts.  相似文献   

4.
The standard enthalpy of formation of Sc5Si3 has been determined by solute-solvent drop calorimetry at (1473±2) K. The following value is reported: ΔH f o (mean)=−(719.1±34.0) kJ mol−1. This result is compared with corresponding published values for the enthalpies of formation of Me5Si3, with Me=Mn, Cr, V, Ti. This comparison shows regularly increasing negative enthalpies of formation from Mn5Si3 to Sc5Si3. LETTTIA TOPOR, formerly Senior Research Associate, The James Franck Institute, The University of Chicago,  相似文献   

5.
The thermodynamic properties of Mg48Zn52 were investigated by calorimetry. The standard entropy of formation at 298 K, Δf S 298 o , was determined from measuring the heat capacity, C p , from near absolute zero (2 K) to 300 K by the relaxation method. The standard enthalpy of formation at 298 K, Δf H 298 o , was determined by solution calorimetry in hydrochloric acid solution. The standard Gibbs energy of formation at 298 K, Δf G 298 o , was determined from these data. The obtained results were as follows: Δf H 298 o (Mg48Zn52)=(−1214±(300) kJ · mol−1fS 298 o (Mg48Zn52)=(−123±0.36) J · K−1 · mol−1; and Δf G 298 o (Mg48Zn52)=(−1177±(300) kJ · mol−1. The electronic contribution to the heat capacity of Mg48Zn52 was found to be approximately equal to pure magnesium, indicating that the density of states in the vicinity of the Fermi level follows the free electron parabolic law.  相似文献   

6.
Interaction between molten salts of the type LiCl-KCl-MeCl (Me = Na, Rb, Cs, x MeCl = 0 to 0.5, x KCl/x LiCl = 0.69) and zeolite 4A have been studied at 823 K. The main interactions between these salts and zeolite are molten salt occlusion to form salt-loaded zeolite and ion exchange between the molten salt and salt-loaded zeolite. No chemical reaction has been observed. The extent of occlusion is a function of the concentration of MeCl in the zeolite and is equal to 11±1 Cl per zeolite unit cell, (AlSiO4)12, at infinite MeCl dilution. The ion-exchange mole fraction equilibrium constants (separation factors) with respect to Li are decreasing functions of concentration of MeCl in the zeolite. At infinite MeCl dilution, they are equal to 0.84, 0.87, and 2.31 for NaCl, RbCl, and CsCl, respectively, and increase in the order Na<Rb<Cs at identical MeCl concentrations. The standard ion-exchange chemical potentials are equal to −(0.0±0.5) kJ·mol−1, −(0.4±0.3) kJ·mol−1, and −(6.5±0.5) kJ·mol−1 for Na, Rb−1, and Cs+, respectively.  相似文献   

7.
The standard enthalpies of formation of TiSi2 and VSi2 have been measured by a new calorimetric method. The following results are reported: ΔH f ° (TiSi2) = −(170.9 ± 8.3) kJ mol−1 and ΔH {f °} (VSi2) = −(112.4 ± 6.0) kJ mol−1. These results are compared with experimental, assessed, and predicted values reported in the literature and with our own data for the corresponding borides. Estimates are given for the enthalpies of formation of the silicides of scandium and chromium.  相似文献   

8.
Abstract

The thermodynamic properties of PbO-SiO2 liquid slags have been investigated by measuring the emfs of the following cell:

Pt, O2(1 atm)/0.90 ZrO2 + 0.10 CaO/(PbO-SiO2)liq.sol'n Pb(l)

From the results, the activities and the other partial molar properties of the system have been calculated for compositions between 40 and 85 mole % PbO in the temperature range 720 to 110°C.

Résumé

Les propriétés thermodynamiques des scories liquids PbO-SiO2 ont été recherchees en mesurant la f.e.m. de la pile suivante:

Pt,O2(1 atm.)/0.90 ZrO2 + 0.10 CaO/(PbO-SiO2)liq.sol'nPb(l)

A partir des résultats, les activités et les autres propriétés molales partielles de ce système ont été calculées pour des compositions molales variant de 40 à 85% en PbO, sur un intelvalle de température s'etendant de 720 a 1100°C.  相似文献   

9.
The formation of gaseous silicon monosulfide according to the reaction 1/2 SiS2(s + 1/2 Si(s) = SiS(g) has been studied by the Knudsen effusion-gravimetric technique over the temperature range 863 to 1161 K. The equilibrium SiS pressure may be expressed by the equation logP SiS(atm) = 7.205 -12,600T −1+ 0.4677 logT -4.508 × 10−4 T + 29.88T −1/2 The experimental free energy of the reaction may be represented by the two-term expression ΔF 863-1161 K 0 = 54,270 -38.34T Combination of these results and recent experimental data provide the standard free energy of formation of gaseous silicon monosulfide as expressed by ΔF 863-1686 K 0 = 12,740 -19.18T Si(s ) + 1/2 S2(g) = SiS(g) ΔF 1686-2953 K 0 = 640 -12.02T Si(l) + 1/2 S2(g) = SiS(g)  相似文献   

10.
Measurements have been made on the thermal capacity of γ-Gd2Se3 at 58.88–298.34 K. Values have been obtained for the thermal capacity, entropy, reduced Gibbs energy, and enthalpy under standard conditions: C°p = 125.87 ± 0.5 J· mole−1 · K−1; S°(298.15 K) = 196.5 · 1.6 J · mole−1 · K−1; Φ°(298.15 K) = 103.6 ± 1.6 J · mole−1 · K−1; H°(298.15 K)-H°(0) = 27681 ± 138 J · mole−1. The enthalpy of Gd2Se3 has been measured and the major thermodynamic functions have been calculated for the solid and liquid states over the temperature range 450–2300 K. The temperature dependence of the enthalpy in the ranges 300–1800 K and 2000–2300 K are represented: H°(T)-H°(298.15 K) = = 1.1949 · 10−2 · T2 + 122.38 · T + 347402 · T−1 − 38716 and H°(T)-H°(298.15 K) = 262.81 · T-− 196047, respectively. The calculated temperature, enthalpy, and entropy of melting for Gd2Se3 are: Tm = 1925 ± 40 K, ΔmH° (Gd2Se3) = 68.5 kJ · mole-1, ΔmS°(Gd2Se3) = 35.6 J · mole−1 · K−1. __________ Translated from Poroshkovaya Metallurgiya, Nos. 3–4(448), pp. 56–61, March–April, 2006.  相似文献   

11.
12.
Enthalpies of formation of (Pd + In) alloys have been obtained by direct reaction calorimetry using a very high temperature calorimeter between 1425 and 1679 K in the concentration range 0 <x Pd < 0.66. They are very negative with a minimum Δmix H o m, = -59.6 /2.5 kJ · mol-1 atx Pd = 0.59 and independent of temperature within the experimental error. The integral molar enthalpy of mixing is given by ΔmixΔH m o /· mol-1 =x(1 -x)·- (-126.94 - 92.653x-83.231.x 2 - 734.49.x 3 + 949.07x 4), wherex = x Pd. The limiting partial molar enthalpy of palladium in indium was calculated as Δh m(Pd liquid in ∞ liquid In) = -127 ± 5 kJ·mol-1. The results are discussed and compared with the enthalpies of formation of solid alloys. The anomalous behavior of the partial enthalpy of Pd is assumed to be due to the charge transfer of, at most, two electrons of In to Pd. Formerly Ph.D. Student, Université de Provence.  相似文献   

13.
Adiabatic oxygen combustion calorimetry has been used to determine the enthalpies of combustion of the chromium carbides Cr23C6, Cr7C3 and Cr3C2 to be—15,057.6±12.4 kJ ·mole−1,—4985.3±3.8 kJ ·mole−1 and—2400.5±0.9 kJ ·mole−1 respectively. The products of combustion in all cases were Cr2O3 and CO2. Using standard data for Cr2O3 and CO2, the enthalpies of formation of the carbides have been calculated to be:fΔH 298 o Cr23C6=−290.0±27.6 kJ·mole−1 fΔH 298 o Cr7C3=−149.2±8.5 kJ·mole−1 fΔH 298 o Cr3C2=−81.1±2.9 kJ·mole−1  相似文献   

14.
Experimental studies on phase equilibria in the multicomponent system PbO-ZnO-CaO-SiO2-FeO-Fe2O3 in air have been conducted to characterize the phase relations of a complex slag system used in lead and zinc smelting. The liquidus in the pseudoternary section ZnO-“Fe2O3”-(PbO + CaO + SiO2) with a CaO/SiO2 weight ratio of 0.35 and a PbO/(CaO + SiO2) weight ratio of 3.2 has been constructed to describe liquidus temperatures as a function of composition in the range of commercial operating conditions employed by the Lead Isasmelt smelting process. The section contains the primary phase fields of spinel (zinc ferrite, Zn x Fe3−y O4+z ), zincite (Zn u Fe1−u O), melilite (Pb v Ca2−v Zn w Fe1−w -Si2O7), hematite (Fe2O3), magnetoplumbite (PbFe10O16), and wollastonite (CaSiO3).  相似文献   

15.
16.
The standard enthalpies of formation of eight samarium alloys with late transition metals have been determined by direct synthesis calorimetry at 1273±2 K. The following values of ΔH f 0, in kJ·(mole atom), are reported: SmNi5, −27.4±0.5; Sm5Rh4, −66.5±1.0; SmRh2, −65.5±1.2; SmPd, −82.4±2.0; Sm3Pd4, −87.2±2.5; SmPd3, −82.9±2.5; SmPt, −108.7±3.5; and SmPt2, −100.2±2.6. The results are compared with predicted values from the Miedema model, with available literature data for SmNi5, SmPd, and SmPt, and with earlier values for similar compounds formed by other lanthanide metals reported by this laboratory. The observed relationships between the enthalpies of formation and the number of f-electrons in the considered binary alloys RE n Me m (RE=lanthanide elements; and Me=Group VIII elements) are discussed.  相似文献   

17.
The possibility of using quantitative differential thermal analysis to investigate phase transformations is examined. The temperature, enthalpy, and entropy of polymorphic transformations in LaGe1.8 and SmSi2 are determined: Ttr = 724 K, ΔtrH = 1635 ± 79 J · mole−1, ΔtrS = 2.3 ± 0.1 J · mole−1 · K−1 (LaGe1.8); Ttr = 658 K, ΔtrH = 1384 ± 69 J · mole−1, ΔtrS = 2.1 ± 0.1 J · mole−1 · K−1 (SmSi2). __________ Translated from Poroshkovaya Metallurgiya, Vol. 46, No. 3–4 (454), pp. 72–78, 2007.  相似文献   

18.
The enthalpies of formation of liquid (Cu + Mn) alloys were measured in the isoperibolic heat-flux calorimeter at 1573 K in the entire range of compositions. The integral molar enthalpy of mixing was found to be negative in the range of molar fractions 0 < x Mn < 0.31, with ΔH(min) = −0.69 ± 0.27 kJ mol−1 at x Mn = 0.12, and positive in the range 0.31 < x Mn < 1, with ΔH(max) = 3.67 ± 0.36 kJ mol−1 at x Mn = 0.75. Limiting partial molar enthalpies of manganese and copper were calculated as = −18.0 ± 6.6 kJ mol−1 and = 29.1 ± 4.9 kJ mol−1, respectively. The results are discussed in comparison with the thermodynamic data available in the literature and the equilibrium phase diagram.  相似文献   

19.
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   

20.
Speciation models for aqueous solutions of UO2SO4−U(SO4)2−H2SO4−HF and UO2Cl2−UCl4−HCl−HF were proposed based on chemical reaction equilibria, mass balances, charge balance, and stoichiometry of UF4(s). The equilibrium concentrations of uranium and fluoride species in these solutions were calculated at 298 K, and are of relevance to the electrolytic reduction of U(VI), followed by the precipitation of UF4(s). In these calculations, the reduction ratios of U(VI) were set at 25, 50, 75, and 100 pct. In the sulfate system the stable domains of U4+, U(SO4) n 4−2n , UF n 4−n , and UF4(s) as U(IV) species and UO 2 2+ , UO2(SO4) n 2−2n , and UO2F n 2−n , as U(VI) species are strongly dependent on theC T(F)/C T(U(IV)) value. On the other hand, the stable domains of U4+ UCl3+, UF n /4−n , and UF4(s) as U(IV) species and UO 2 2+ , UO2Cl+, and UO2F n 2−n as U(VI) species are also strongly affected by theC T(F)/C T(U(IV)) ratio in the chloride system. The initiation and precipitation of UF4(s) in both the sulfate and chloride systems are a function of the reduction ratio of U(VI). The higher the reduction ratio, the lower theC T(F)/C T(U(IV)) values required. Compared to the chloride system, UF4(s) precipitation in the sulfate system starts at a lower value ofC T(F)/C T(U(IV)). The addition of an excess amount of HF does not cause the dissolution of UF4(s) precipitates because HF is a weak acid. KOJI SATO, formerly Graduate Student, Department of Metallurgy, Kyoto University, Kyoto, Japan  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