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 tracer diffusion coefficient, D1O, of oxide ions in LaFeO3 single crystal was determined over the temperature range of 900–1100°C by the gas-solid isotopic exchange technique using 18O as a tracer. For the determination of D1O, the depth profile of 18O was measured by means of a secondary ion mass spectrometer (SIMS). The surface exchange reaction was found to be slow and the surface exchange rate constant, k, was determined together with D1O. It was found that D1O at 950°C is proportional to P?0.58O2, where PO2 is an oxygen pressure. The vacancy mechanism was determined for the diffusion of oxide ions from the PO2 dependence. The vacancy diffusion coefficient, DV, for LaFeO3 was nearly the same as that for LaCoO3 at the same temperature. The activation energy for migration of oxide ion vacancies was 74 kJ · mole?1 for both oxides.  相似文献   

2.
Yttrium self-diffusion in monocrystalline yttrium oxide (Y2O3) is studied by means of the classical radio tracer technique. The few reliable diffusion data obtained in the temperature range 1600–1700°C lead to the following diffusion coefficient
D=3.5×10?9exp?72RT(kcal/mole) m2sec?1
.Experimental errors on the above numerical values are large and give, for the preexponential and energy terms, respectively:
2.10?7<D0<3.10?10m2sec?
62<Q<82 kcal/mole
.Nevertheless these results seem in good agreement with those deduced from high-temperature and low-stress creep experiments. The theoretical aspect of self-diffusion of yttrium in Y2O3 is studied in terms of point defects and lattice disorder due to the equilibrium between the oxide and its environment. This last part is confined to the restricted range of high oxygen partial pressure in which oxygen interstitials are supposed to be majority defects. Intrinsic and extrinsic diffusion behavior are both considered on the basis of a vacancy diffusion mechanism.  相似文献   

3.
In order to elucidate the defect structure of the perovskite-type oxide solid solution La1?xSrxFeO3?δ (x = 0.0, 0.1, 0.25, 0.4, and 0.6), the nonstoichiometry, δ, was measured as a function of oxygen partial pressure, PO2, at temperatures up to 1200°C by means of the thermogravimetric method. Below 200°C and in an atmosphere of PO2 ≥ 0.13 atm, δ in La1?xSrxFeO3?δ was found to be close to 0. With decreasing log PO2, δ increased and asymptotically reached x2. The log(PO2atm) value corresponding to δ = x2 was about ?10 at 1000°C. With further decrease in log PO2, δ slightly increased. For LaFeO3?δ, the observed δ values were as small as <0.015. It was found that the relation between δ and log PO2 is interpreted on the basis of the defect equilibrium among Sr′La (or V?La for the case of LaFeO3?δ), V··O, Fe′Fe, and Fe·Fe. Calculations were made for the equilibrium constants Kox of the reaction
12O2(g) + V··o + 2FexFe = Oxo + 2Fe·Fe
and Ki for the reaction
2FexFe = FeFe + Fe·Fe·
Using these constants, the defect concentrations were calculated as functions of PO2, temperature, and composition x. The present results are discussed with respect to previously reported results of conductivity measurements.  相似文献   

4.
The electrical conductivity and departure from the stoichiometry of Nd2O3 have been measured over the temperature range of 900° to 1100°C and oxygen partial pressure of 1 to 10?16 atm. The hole conductivity of Nd2O3 is found to be proportional to P1nO2, where n are 4.6, 4.9, and 5.1 at 900°, 1000°, and 1100°C, respectively. From the oxygen partial pressure dependence of the hole conductivity, it is shown that the predominant point defects in nonstoichiometric NdO1·+x are fully ionized and partially doubly ionized metal vacancies. From the thermogravimetric measurements, the departure from stoichiometry, x in NdO1·5+x, is 2.0 × 10?3 at 1000°C and 1 atm. By combining the electrical conductivity and weight change data, it is shown that the hole mobility is 6.3 × 10?4 (cm2/V·sec) at 1000°C and 1 atm.  相似文献   

5.
CsSbF6(II) under ambient conditions is trigonal, space group D3d5-R3m. At 187.8°C it undergoes a phase transition with an enthalpy change of 5.267 ± 0.316 kJ mole?1, to phase CsSbF6(I). CsSbF6 decomposes with loss of fluorine at atmospheric pressure at high temperatures, but under pressure the decomposition is prevented and a melting point of 310°C at atmospheric pressure can be inferred. The III phase boundary and melting curve were studied as functions of pressure. The infrared and Raman spectra of CsSbF6(II) were studied in the temperature range of ?256 to 20°C, at ambient pressure. The crystal chemistry of the CsSbF6 and its relationship with other related compounds is discussed.  相似文献   

6.
The electrical conductivity of polycrystalline SrTiO3 was determined for the oxygen partial pressure range of 10° to 10?22 atm and temperature range of 800 to 1050°C. The data were found to be proportional to the ?16th power of the oxygen partial pressure for the oxygen pressure range 10?15–10?22 atm, proportional to P?14O2 for the oxygen pressure range 10?8–10?15 atm, and proportional to P+14O2 for the oxygen pressure range 100–10?3 atm. These data are consistent with the presence of very small amounts of acceptor impurities in SrTiO3.  相似文献   

7.
The effect of substitution extent on formation of superstructure-ordered vacancies in zinc-substituted lacunar spinels of type (Zn2+xFe3+1?x)A(Fe3+(5+x)3(1?x)3)BO2?4 was investigated using ir spectrometry. Only those lacunar phases whose substitution extent x is less than about 0.3 show a vacancy ordering on octahedral sites. In addition, referring to the disappearance of the 635-cm?1 absorption band, which is characteristic of these lacunar spinels, we show that the transformation temperature of the γ phases into αFe2O3 increases with zinc substitution extent. For the α phase obtained at 700°C we have found a linear variation between the intensity difference of the 390- and 450-cm?1 absorption bands and the percentage of αFe2O3.  相似文献   

8.
The kinetics, mechanism, and activation energy of the isothermal decomposition of CuCrO4 was studied using an isothermal TG method and an X-ray high-temperature diffraction technique in either air or a flowing atmosphere of N2. The enthalpy change ΔH of the decomposition reaction
2CuCrO4CuO+CuO+CuCr2O4+32O2
was determined by DSC analysis. The mechanism of the thermal decomposition of CuCrO4 is well represented by the standard Avrami-Erofeev kinetic equation [?ln(1 ? α)]12 = kt. According to this mechanism, the reaction rate is controlled by the formation and growth of nuclei on the surface of the reactant. The activation energy EA of the process in air is EA = (248 ± 8) kJ mole?1, in flowing atmosphere of nitrogen EA = (229 ± 8) kJ mole?1. ΔH in air is 110 kJ mole?1, in flowing nitrogen 67 kJ mole?1. The lower values of ΔH and EA in the flowing atmosphere of nitrogen are due to the fast elimination of O2 from the reaction interface. However, the decay of the crystalline portion of CuCrO4 during its thermal decomposition, studied by the X-ray diffraction, is controlled by a different reaction mechanism (first-order kinetics). The reaction mechanism is discussed in the relation to the crystal structure of the reactants.  相似文献   

9.
The theta temperature for the system poly(o-chlorostyrene)-methyl ethyl ketone has been determined as 24·5°. The samples used in the determination were prepared by radical polymerization. The dependence of intrinsic viscosity on molecular weight has been measured in methyl ethyl ketone at 24·5° and found to be ηθ = 4·68 × 10?4MwM12. The ratio 〈s=2〉/M was found, by light scattering, to be 5·60 × 10?18 cm2. Analysis of the solution properties indicates that the Kurata-Yamakawa theory is valid in the vicinity of the Flory temperature (UCST).  相似文献   

10.
The electrical conductivity of polycrystalline strontium titanate with (SrTi = 0.996, 0.99, and 0.98 was determined for the oxygen partial pressure range of 100 to 10?22 atm and the temperature range of 850–1050°C. These data were found to be similar to that obtained for the sample with ideal cationic ratio. The observed data were proportional to the ?16 power of oxygen partial pressure for PO2 < 10?15atm, proportional to P?14O2 for the pressure range 10?8–10?15 atm, and proportional to P+14O2 for PO2 > 10?4atm. The deviation from the ideal Sr-to-Ti ratio was found to be accommodated by neutral vacancy pairs, (V″Sr V″0. The results indicate that the single-phase field of strontium titanate extends beyond 50.505 mole% TiO2 at elevated temperatures.  相似文献   

11.
The electrical conductivity of polycrystalline CaTiO3 was measured over the temperature range 800–1100°C while in thermodynamic equilibrium with oxygen partial pressures from 10?22 to 100 atm. The data were found to be proportional to the ?16th power of the oxygen partial pressure for the oxygen pressure range 10?16 – 10?22 atm, proportional to P?14O2 for the oxygen pressure range 10?8 – 10?15 atm, and proportional to P+14O2 for the oxygen pressure range greater than 10?4 atm. The region of linearity where the electrical conductivity varies as ?14th power of PO2 increased as the temperature was decreased. The observed data are consistent with the presence of small amounts of acceptor impurities in CaTiO3. The band-gap energy (extrapolated to zero temperature) was estimated to be 3.46 eV.  相似文献   

12.
Calorimetric measurements of the enthalpy of solution of cesium chromate gave ΔHsoln = (7622 ± 24) calth mol?1 for a dilution of Cs2CrO4·21128H2O. This result, along with the enthalpy of dilution gave the standard enthalpy of solution, ΔHsolno = (7512 ± 31) calth mol?1, whence the standard enthalpy of formation, ΔHf0(Cs2CrO4, c, 298.15 K), was calculated to be ?(341.78 ± 0.46) kcalth mol?1. Recomputed thermodynamic data for the formation of the other alkali metal chromates have been tabulated. From their solubilities and enthalpies of solution, the standard entropies, S0(298 K), of BaCrO4 and PbCrO4 were estimated to be (38.9 ± 0.9) and (43.7 ± 1.2) calth K?1 mol?1, respectively. There is evidence that ΔHf0(SrCrO4, c, 298.15 K) may be in error. Thermochemical, solubility, and equilibrium data, have been combined to update the thermodynamic properties of the aqueous chromate (CrO42?), bichromate (HCrO4?), and dichromate (Cr2O72?) ions. The new values at 298.15 K are as follows:
  相似文献   

13.
The heat capacity of the solid solution Mn3.2Ga0.8N was measured between 5 to 330 K by adiabatic calorimetry. A sharp anomaly with first-order character was detected at TA = (160.5±0.5) K, corresponding to a magnetic rearrangement and a lattice expansion. No sharp anomaly was observed at Tc ≈ 260 K where the magnetic ordering takes place; instead, a smooth shoulder was detected. The thermodynamic functions at 298.15 K are Cp,mR = 15.16, SmoR = 18.57, {Hmo(T)?Hmo(0)}R = 2896 K, ?{Gmo(T)?Hmo(0)}RT = 8.85. At low temperatures the coefficient for the linear electronic contribution to the heat capacity was derived: γ = (0.031±0.003) J·K?2·mol?1. Moreover, the different contributions to the heat capacity were obtained and the electronic origin of the phase transitions was established.  相似文献   

14.
The compounds Ba4Fe2S6[S23(S2)13] and Ba3.6Al0.4Fe2S6[S0.6(S2)0.4], designated I and II, were prepared by reacting BaS, Fe, and S powders and Al foils in graphite containers sealed in evacuated quartz ampoules at approximately 1100°C. The crystal structure of I was determined using 1682 independent, nonzero X-ray reflections, while 3589 were used for II. They are triclinic, Al:
a=9.002(2)A?,b=6.7086(8)A?,c=24.658(4)A?α91.49(2)°,
β=105.10(2)°y=90.74(2)°,ψcalc=4.15g/cm3,for I:
a=8.993(6)A?,b=6.708(7)A?,c=24.70(1)A?α91.11(6)°,
β=105.04(6)°y=90.90(9)°,ψcalc=3.90g/cm3,for II:
BaS6 trigonal prisms share edges to form distorted hexagonal rings which form one-dimensional chains leaving two free lateral edges. The chains link in a stairstep manner with the rings offset along the [301] direction. These stairsteps join in a complicated manner to form a three-dimensional network. Fe ions are in two sites forming isolated FeS4 tetrahedra and isolated Fe2S6 dimers by edge-sharing tetrahedra. The Al substitution occurs in the trigonal prisms which have free edges with Al replacing Ba. Room-temperature Mössbauer isomer shifts are 0.20 mm/sec. for I and 0.30 mm/sec for II. These data indicate that upon Al substitution charge compensation occurs by reducing Fe3+. Valence calculations indicate that Fe in edge-sharing tetrahedra are reduced while the Fe in the isolated tetrahedron remains unchanged. The effective charge distribution in the Al substituted compound is approximately Fe3+, Fe2.5+ with electron delocalization across the shared edge. Room temperature electrical resistivity is 105 ohm/cm. The compositions of the crystals are best represented by the formulas [Ba4Fe2S7]23·[Ba4Fe2S6(S2)]13 and [Ba3AlFe2S7]0.4·[Ba4Fe2S7]0.2·[Ba4Fe2S6(S2)]0.4. The replacement of a sulfide by a disulfide ion is thought to be strongly dependent on the sulfur activity during the preparation.  相似文献   

15.
The standard enthalpy of formation of γ-UO3 has been critically assessed; the value ?(292.5 ± 0.2) kcalth mol?1 is suggested.The enthalpies of solution of β-UO3 and γ-UO3 in 3 M H2SO4 have been measured and used to derive:
ΔHf°(β?UO3, 298.15 K) = ?(291.6 ± 0.2) kcalth mol?
  相似文献   

16.
The electrical conductivity of sintered specimens of nonstoichiometric CeO2?x was measured as a function of temperature (750–1500°C) and oxygen pressure (1–10?22 atm). The isothermal compositional dependence of the electrical conductivity of CeO2?x was determined by combining recently obtained thermodynamic data, x = x(PO2, T), with the conductivity data. The compositional and temperature dependence of the electrical conductivity may be represented by the expression
σ=410[x]e?(0.158+x)kT(ohm cm)?1
over the temperature range 750–1500°C and from x = 0.001 to x = 0.1.This expression was rationalized in terms of the following simple relations for (a) the electron carrier concentration
ncece=8xa03
where nCe′Ce is the number of Ce′Ce per cm3 and a0 is the lattice parameter and (b) the electron mobility
μ=5.2(10?2)e?(0.158+x)kT(cm2/V sec)
.  相似文献   

17.
Paramagnetic resonance and magnetic measurements were performed on powdered samples of GdGa2. The magnetic data indicated ferrimagnetic behavior with Tc ? 181° K. Above 250° K the susceptibility obeys the Curie-Weiss law χg = 2.662 × 10?2(T ? 27.6)emu/g-Oe which corresponds to an effective moment of 7.95 Bohr magnetons. Over the range from 190 to 300°K the data obey a Néel type law, χg?1 = 35.95 (T ? 12.5) ? 2.20 × 104(T ? 177), which is indicative of ferrimagnetic order. The resonance measurements were performed at 9.013 gHz at 247, 296, and 349°K. The spectra were analyzed with a computerized curve-fitting technique that involves a linear combination of Lorentzian absorption and dispersion susceptibility components. Following demagnetization corrections, the g-factor was found to be 1.9832 while the half-power, half-linewidth was 592.7 Oersteds.  相似文献   

18.
Proton NMR relaxation times (T2T1, and T1?) and absorption spectra are reported for the compounds H1.71MoO3 (red monoclinic) and H0.36MoO3 (blue orthorhombic) in the temperature range 77 K < T < 450 K. Rigid lattice dipolar spectra show that both compounds contain proton pairs, as OH2 groups coordinated to Mo atoms in H1.71MoO3 and as pairs of OH groups in H0.36MoO3. The room temperature lineshape for H1.71MoO3 shows that the average chemical shielding tensor has a total anisotropy of 20.1 ppm. The relaxation measurements confirm that hydrogen diffusion occurs and give EA = 22 kJ mole?1 and τ0C ? 10?13sec for H1.71MoO3 and EA = 11 kJ mole?1 and τ0C ? 3 × 10?8sec for H0.36MoO3.  相似文献   

19.
The extraction of In(III) from 1M (Na,H)(Cl,ClO4) media with 4-acylpyrazol-5-ones (HL) in toluene at 25°C is described by equilibria In 3+ + 3 HL ? InL3 + 3 H+ (log K = 1.48, 1.03, 0.87 with acyl = benzoyl, lauroyl, 2-thenoyl), InCl 2+ + 2 HL ? InClL2 + 2 H+ (log K = 0.26, ?0.45, ?0.35 respectively) and In3+ + m Cl? ? InClm(3-m)+ (log βm available from literature). The extraction from 1M (Na,H)(Cl,NO3) medium is enhanced by addition of aliquat (TOMA+,Cl?) and the following synergic equilibrium takes place : InCl2 + (TOMA+,Cl?) ? (TOMA+, InCl2L2? (log K = 5.49, 5.25, 5.21 respectively). Cl? of (TOMA+,Cl?) is exchanged by NO3? with the equilibrium constant log K = 1.50. If (TOMA+,Cl?) is replaced by tri-n-octylammonium chloride, the synergic effect is largely reduced (log K = 4.17 with acyl = benzoyl). The extraction from chloride solutions containing ClO4? remains unchanged by addition of ammonium salts.  相似文献   

20.
4-Vinylpyridine was polymerized by cumyl barium in THF at 0°C. Detailed conductance studies at various concentrations of the living oligomer solution gave similar experimental results as for 2-vinylpyridine, i.e. triple ions exist in thermodynamic equilibrium with free ions and ion pairs. The dissociation constant of ion pairs increases from 1.96 × 10?10 M at 15°C to 4.35 × 10?10 M at ?70°C with enthapy and entropy of dissociation of ?3.9 kJ/mol and ?200 J mol?1 K?1 respectively. The extent of dissociation of ion pairs of living oligo-4-vinylpyridine is comparable to that of living polystyrene (K1 ? 10?10) but higher than that of living poly-2-vinylpyridine (K1 ? 10?11). This result is interpreted in terms of intramolecular complexation which is no longer favourable when the nitrogen atom is situated at the 4-position of the pyridine nucleus. Studies of propagation kinetics revealed that 4-vinylpyridine polymerizes by the same reaction mechanism as the 2-isomer, the active sites being essentially ion pairs. At 0°C, its propagation constant was found to be 211.1 M?1 sec?1 compared with 151.04 M?1 sec?1 for living poly-2-vinylpyridine. As for free radical polymerization, the polymers contain only about 26% isotactic triads.  相似文献   

S0/calth K?1 mol?1ΔHf0/kcalth mol?1ΔGf0/kcalth mol?1
CrO42?(aq)(13.8 ± 0.5)?(210.93 ± 0.45)?(174.8 ± 0.5)
HCrO4?(aq)(46.6 ± 1.8)?(210.0 ± 0.7)?(183.7 ± 0.5)
Cr2O72?(aq)(67.4 ± 3.9)?(356.5 ± 1.5)?(312.8 ± 1.0)
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