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 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   

2.
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   

3.
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   

4.
The tracer diffusion coefficient, D1O, of oxide ions in LaCoO3 single crystal was determined over the temperature range of 700–1000°C by a gas-solid isotopic exchange technique using 18O tracer. For the determination, two methods, the gas phase analysis and the depth profile measurement, were employed. Under an oxygen pressure of 34 Torr, the temperature dependence of D1O in LaCoO3 was expressed by
D1O(cm2·sec?1) = 3.63 × 104exp? (74 ± 5)kcal · mole?1RT
D1O at 950°C was found to be proportional to P?0.35O2. The diffusion of oxide ions occurs through a vacancy mechanism. The activation energy for the migration of oxide ion vacancies was estimated as 18 kcal · mole?1.  相似文献   

5.
The mutual solubilities of {xCH3CH2CH2CH2OH+(1-x)H2O} have been determined over the temperature range 302.95 to 397.75 K at pressures up to 2450 atm. An increase in temperature and pressure results in a contraction of the immiscibility region. The results obtained for the critical solution properties are: To(U.C.S.T.) = 397.85 K and xo = 0.110 at 1 atm; (dTodp) = ?(12.0±0.5)×10?3K atm?1 at p < 400 atm and (dTodp) = ?(7.0±0.7)×10?3K atm?1 at 800 atm < p < 2500 atm; (dxodT) = ?(4.0±0.5)×10?4K?1.  相似文献   

6.
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   

7.
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   

8.
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   

9.
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   

10.
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   

11.
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   

12.
Benzophenone (BP) in low concentrations (<0.001 mol 1?1) produces a rate enhancing effect in the H2O2-induced bulk photopolymerization of MMA. Rp is proportional to [H2O2]0.4 and [BP]0.4, and kp2k1 at 30° is 1.00 × 10?2 1.mol?1 sec?1. In diluted systems, different solvents produce different kinetic effects, reaction order with respect to monomer being negative for IPA and THF as solvent, positive but <1.0 for benzene and chloroform, 1.2 for acetonitrile, CCl4 and t-butanol and 1.8 for DMA. The variable solvent effect is attributed to modification of the initiation process by the various solvents to different extents. Kinetic analysis of data for bulk photopolymerization gives evidence for primary radical termination and degradative initiator transfer.  相似文献   

13.
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   

14.
Geometric constraints present in A2BO4 compounds with the tetragonal-T structure of K2NiF4 impose a strong pressure on the BOIIB bonds and a stretching of the AOIA bonds in the basal planes if the tolerance factor is t ? RAO√2 RBO < 1, where RAO and RBO are the sums of the AO and BO ionic radii. The tetragonal-T phase of La2NiO4 becomes monoclinic for Pr2NiO4, orthorhombic for La2CuO4, and tetragonal-T′ for Pr2CuO4. The atomic displacements in these distorted phases are discussed and rationalized in terms of the chemistry of the various compounds. The strong pressure on the BOIIB bonds produces itinerant σ1x2?y2 bands and a relative stabilization of localized dz2 orbitals. Magnetic susceptibility and transport data reveal an intersection of the Fermi energy with the d2z2 levels for half the copper ions in La2CuO4; this intersection is responsible for an intrinsic localized moment associated with a configuration fluctuation; below 200 K the localized moment smoothly vanishes with decreasing temperature as the d2z2 level becomes filled. In La2NiO4, the localized moments for half-filled dz2 orbitals induce strong correlations among the σ1x2?y2 electrons above Td ? 200 K; at lower temperatures the σ1x2?y2 electrons appear to contribute nothing to the magnetic susceptibility, which obeys a Curie-Weiss law giving a μeff corresponding to S = 12, but shows no magnetic order to lowest temperatures. These surprising results are verified by comparison with the mixed systems La2Ni1?xCuxO4 and La2?2xSr2xNi1?xTixO4. The onset of a charge-density wave below 200 K is proposed for both La2CuO4 and La2NiO4, but the atomic displacements would be short-range cooperative in mixed systems. The semiconductor-metallic transitions observed in several systems are found in many cases to obey the relation Ea ? kTmin, where ? = ?0exp(?EakT) and Tmin is the temperature of minimum resistivity ?. This relation is interpreted in terms of a diffusive charge-carrier mobility with Ea ? ΔHm ? kT at T = Tmin.  相似文献   

15.
The kinetics of the interaction of hexaaquochromium(III) ion with potassium octacyanomolybdate(IV) have been studied using conductance and spectrophotometric data. The mechanism of the reaction is discussed and the effect of H+ ion and the ionic strength on the rate of the reaction determined. The reaction is found to be pseudo-first order with respect to potassium octacyanomolybdate(IV) and inverse first order with [H3O+]. The rate of the reaction increases with increase in ionic strength and temperature. Activation parameters have been calculated using the Arrhenius equation and have the values ΔE* = 1.3 × 102 kJ mol?1, ΔH* = 129 kJ mol?1, ΔS* = ?315 e.u., ΔF* = 2.3 × 102 kJ and A = 1.5 × 10?3. The mechanism proposed is based on ion-pair formation and the rate equation obtained is given by: kobs=[kKE[H3O+]+k′K′kEkh][Mo(CN)84?][H3O+]+kh+[KE[H3O+]+K′Ekh][Mo(CN)84?]  相似文献   

16.
The polymerization of acrylamide, initiated with permanganate/oxalic acid, has been studied at 35 ± 0.2 in an aqueous medium under nitrogen. Samples of polyacrylamide were fractionated by the triangular fractionation method using methanol as non-solvent. Molecular weights of the fractions have been determined by viscometry and osmometry. Integral and differential distribution curves were plotted using the fractionation data. The narrow molecular weight distribution for high conversion polymers has been discussed. For fractionated samples of polyacrylamide in water at 30°, the equation [η]ml/g = 6.5 × 10?3Mn0.82 is applicable for the molecular weight range 4 × 104 to 127 × 104. This equation is very similar to the equation [η]ml/g = 6.31 × 10?3M/0.80 of Scholtan. Other parameters (osmotic second virial coefficient and unperturbed dimension) have also been evaluated.  相似文献   

17.
18.
The extraction of Co(II) with mixtures of 1-phenyl-3-methyl-4-benzoyl-pyrazol-5-one ((H)PMBP) and tri-n-octylamine (TOA) is investigated in order to explore the influence of diluents and inorganic anions with synergistic acidic extractant + liquid anion exchanger systems. Although it is proved that the same species [HTOA]+ [Co(PMBP)3]? is extracted from various inorganic media, with toluene as the diluent, the presence of ClO4? SO42? or Cl? anion modifies the distribution of the anions which are associated to (HTOA)+ in the organic phase, leading to different synergistic equilibria; with Cl? or SO42?: CO(PMBP)2 + (HTOA+,PMBP?) ?(HTOA+,Co(PMBP)3? (log K = 6.10) and with ClO4? : Co(PMBP)2 + HPMBP + (HTOA+,ClO4? ? (HTOA+,Co(PMBP)3? + H+ + ClO4? (log K = 2.34) The same synergistic equilibrium is observed for the extraction of Ni(II) from ClO4? medium, with a comparable value of the constant (log K = 2.45). The synergistic effect is cancelled in n-octanol.  相似文献   

19.
Electron spin resonance spectra attributed to four Fe3+ centers designated Oa, Ob, Ta, Tb have been observed in crystals of muscovite and phlogopite. The results are discussed using the spin Hamiltonian
Hs=geμBH·S+D(Sz2?13S?(S+1)+E(S?x2?E(S?y2)
with ge ~ 2.002. The angular variation of the resonance lines is used to determine the ESR axes of the four different sites. Two species are octahedrally coordinated (Oa and Ob) and are assigned to two different surroundings of Fe3+ in the octahedral sheet. The remaining two species (Ta and Tb) may be assigned to the tetrahedral FeO4. The Ta sites have a symmetry axis lying along one of the FeO bonds. The symmetry axis is created by an excess of negative charge on the oxygen bound to the neighboring tetrahedral substitution. Rhombic symmetry of the Tb sites is due to the presence of fluorine anions substituting some hydroxyl ions. One of the ESR axes is directed toward the fluorine ion.  相似文献   

20.
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   

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