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.
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   

2.
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   

3.
A model theory of viscosity η for moderately concentrated polymer solutions is based on the assumption of a “local viscosity” effect and intermolecular hydrodynamic and thermodynamic interactions. It is shown that η is given by
η = ηo{1 + γc[η]}12·expHoRT1 ? aø
where γ is 0–0.4 and depends on the quality of the solvent, a varies between 0,4 and 0.8 and depends on the fraction of the “free volume” of the systems, H0 is the activation energy of the solvent and π is the polymer volume concentration. The dependence of η and “activation energy” of π and T for various molecular weights and qualities of solvents is described quantitatively. Anomalous dependences of [η] and of η on M for low polymer are obtained. An expression for η is proposed:
ηηo1 ? 2K= {1 + (1 ? 2K)c[η]}F(π)
where K is the Huggins-Martin coefficient and F(π) = 1 for most solutions when T is > Tg. For poor solvents the H vs c curve (where H is the activation energy of η of solution) has a minimum value at moderate concentrations. For good solvents, H depends slightly on the molecular weight according to an empirical equation:
H = Ho + 660α31nηηo
Expressions are given from the viscosities of solutions of miscible and also solutions of immiscible polymers.  相似文献   

4.
Photon-correlation spectroscopy has been used to measure the diffusion coefficient D and first-mode intramolecular relaxation time τ1 of polystyrene in a theta solvent, cyclohexane at 34.5°C. Measurements were made on five narrow fractions (Mw = 2.88 × 106 to 9.35 × 106) as a function of concentration c, in the dilute regime. D varied linearly with c, D = Do (I + kDc), with Do = (1.4 ± 0.2) × 10?4Mw?(0.508±0.007) cm2 s?1. Although the values obtained for τ1 were reproducible to within 5%, no systematic variation with c was detected. The results are fitted by the relation τ1 = (7.7 ? 0.3) × 10?8Mw(1.42±0.05) μs, which agrees well with the theoretical expression of Zimm for the non-draining bead-and-spring model, modified for the light-scattering case.  相似文献   

5.
The η-hexamethylbenzenehydridoruthenium(II) complexes RuHCl(η-C6Me6)L (L = PPh3 (11), AsPh3 (12), P(C6H4-p-F)3 (14), P(C6H4-p-Me)3 (15), P(C6H4-p-OMe)3 (16), P-t-BuPh2 (17), P-i-PrPh2 (18), P-i-Pr3 (19), PCy3 (20) and P-t-BuMe2 (21)) have been made by heating [RuCl2(η-C6Me6)]2, the ligand and sodium carbonate in propan-2-ol. The triarylphosphine complexes 11, 14 and 15 react with methyllithium to give aryl ortho-metallated hydridoruthenium(II) complexes such as RuH(o-C6H4PPh2)(η-C6Me6) (22) and 19 similarly gives the isopropyl cyclometallated complex RuH(CH2CHMeP-i-Pr2(η-C6Me6) (29) as a mixture of diastereomers. Reaction of 17 with methyllithium gives initially the t-butyl cyclometallated complex RuH(CH2CMe2PPh2)(η-C6Me6) (25) which isomerizes by a first order process (k0?.2 h?1 in C6D6 or THF-d8 at 50°C) to the aryl ortho-metallated complex RuH(o-C6H4P-t-BuPh)(η-C6Me6) (26). The similarly generated isopropyl cyclometallated complex RuH(CH2CHMePPh2)(η-C6Me6) (27) has not been isolated in a pure state owing to rapid isomerization to RuH(o-C6H4P-i-PrPh)(η-C6Me6) (28); both 27 and 28 exist as a pair of diastereomers. The formation of the cyclometallated complexes and the isomerizations are thought to involve intermediate 16-electron ruthenium(O) complexes Ru(η-C6Me6)L.  相似文献   

6.
The Sr2+1?yLa3+yFeO3 system with 0.1 ≦ y ≦ 0.6 has been studied mainly by the Mössbauer effect. The results are discussed referring to the Ca1?xSrxFeO3 system. The following four kinds of electronic phases have been observed: the paramagnetic and the antiferromagnetic average valence phases and the corresponding mixed valence phases. Two kinds of Fe ions coexist, in general, in the mixed valence phases. In the antiferromagnetic mixed valence phase, typically at 4 K, the magnetic hyperfine field and the center shift each takes a wide range of value depending on the composition, while a beautiful correlation is kept between them. The extreme values are close to those expected for Fe3+ and Fe5+. The appropriate chemical formulas are, therefore, Ca1?xSrxFe(3+Δ)+0.5Fe(5?Δ)+0.5O3 and Sr1?yLayFe(3+δ)+(1+y)2Fe(5?δ)+(1?y)2O3.  相似文献   

7.
Treatment of [{Ir(COD)(μ-Cl)}2] with excess of the electron-rich olefin [CN(Ar)(CH2)2NAr]2 (abbreviated as (LAr)2, Ar = C6H4Me-p or C6H4OMe-p) affords the ortho-metallated tricycle [Ir(LAr)3], which for Ar = C6H4Me-p (Ia) with HCL yields [Ir(LAr)2(LAr)]Cl (IV); X-ray data show that in IV there is an unexpectedly close Ir?C(o-aryl) contact (2;52(1) Å) involving the “free” LAr which compares with an IrC(o-aryl) distance of 2.09(3) Å in Ia or 2.07(3) Å in the ortho-metallated LAr ligand of complex IV.  相似文献   

8.
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   

9.
From the study of the CuCuBrCu2?εSC solid cell in the range of cubic digenite and “high temperature” hexagonal chalcocite we have deduced the laws of variation of the deviation from stoichiometry and the holes concentration with the equilibrium partial pressure of sulfur (δ and p are found to be proportional to p14S2). The electronic model corresponding to the formation of associations (V×CuVCu) in the presence of neutral vacancies V×Cu allows one to explain these laws. In the low temperature range (range of “low temperature” hexagonal chalcocite and orthorhombic chalcocite) the study of thermal variations of Hall coefficient permits us to propose the following models: (a) the deviation from stoichiometry of the “low temperature” hexagonal chalcocite is due to the simple ionized vacancies VCu; (b) the deviation from stoichiometry of the orthorhombic chalcocite is due to the vacancies VCu and to the associations (V×CuV×Cu).  相似文献   

10.
11.
Oxygen defect K2NiF4-type oxides La2?xSrxCuO4?x2 have been synthesized for a wide composition range: 0 ≤ x ≤ 1.34. From the X-ray and electron diffraction study three domains have been characterized: orthorhombic compounds with La2CuO4 structure for 0 ≤ x < 0.10, tetragonal oxides similar to LaSrCuO4 for 0.10 ≤ x < 1 and several superstructures derived from the tetragonal cell (a ? n.aLaSrCuO4 with n = 3, 4, 4.5, 5, 6) for 1 ≤ x ≤ 1.34. The compounds corresponding to 0 < x < 1 differ from the other oxides in that they are characterized by the presence of copper with two oxidation states: + 2 and + 3. A model structure for La0.8Sr1.2CuλO3.4, in which copper has only the + 2 oxidation state, and for which the actual cell is tegragonal—a = 18.804 Å and c = 12.94 Å—has been established. The particular structural evolution of these compounds is discussed in terms of a competition between the capability of Cu(II) to be oxidized to Cu(III) and the ordering of oxygen vacancies.  相似文献   

12.
Magnetic study and Mössbauer resonance measurements of the tellurites Fe2Te3O9 and Fe2Te4O11 characterize antiferromagnetic ordering. The transition temperatures determined by Mössbauer resonance, are 34 and 27 K, respectively. At 295 K the values of chemical shifts, 0.35 and 0.39 mm/sec, are typical of high-spin Fe(III) in octahedral coordination. Neutron powder diffraction was used to determine the magnetic structure of Fe2Te4O11 at 4.2 K. It shows antiferromagnetic interactions between Fe3+ ions belonging to [Fe2O10] groups. The magnetic space group is P2a21c.  相似文献   

13.
Osmotic pressure measurements on polystyrene (Mn = 396. 000) in trans-decalin for concentrations up to 140 kg m?3 and from 20 to 40 are reported. The θ-temperature is 20.8 . The ratio
?(πc)?cc=0?(πc)?cc=c+
where c+ is the concentration at which a homogeneous segment distribution is assumed to prevail, increases with temperature up to the plateau value of 0.7. From the temperature dependence of the osmotic pressure, the partial molar enthalpy. Δh1, and entropy, Δs1. of mixing are found to be positive. The solvent-solute interaction parameter increases linearly with concentration at all temperatures.  相似文献   

14.
The study of the K2NiF4 structure by the “method of invariants” leads to the relationship
0.99615 V13BA212
with V = a2c (βB and ψA are invariant values associated with cations B and A) in compounds with K2NiF4 structures. Some values of ψA and examples are proposed.  相似文献   

15.
Variable temperature 1H NMR spectroscopy has been used in the study of 1,3-intramolecular shifts of the M(CO)5 moiety in complexes of the general formula [M(CO)5L], (M = Cr or w), L = SCH2SCH2SCH2, SCH2SCH2CH2CH2 and SCH(Me)SCH2CH2CH2. For the 1,3,5-trithian complexes precise energy barriers for the process have been obtained by detailed computer simulation of the static and dynamic spectra. Our results suggest that the magnitude of ΔG (298.15 K) for the 1,3-shift is largely dependent upon the skeletal flexibility of the ligand system. In this context we have investigated the X-ray crystal structure of the highly substituted trithian complex [W(CO)5{β-SCH(Me)SCH(Me)SCH(Me)}].  相似文献   

16.
The study of K2NiF4 and perovskite structure type by the “method of invariants” leads to the relationship: (A-X)9 212 ? (A-X)12 = constant, where (A-X)9 and (A-X)12 are the invariant values associated with cation A in coordination number 9 and 12. In the case where A = K+ and X = F?, we propose the relationship:
(K+?F)R = 2.832 R111.4
where R is the coordination number.  相似文献   

17.
A neutron diffraction study has been carried out on Sr0.5La1.5Li0.5Fe0.5O4 of K2NiF4-type derived structure and it has shown that iron in the tetravalent state has a high spin configuration (t32ge1g) and that the material has some stacking defects. At room temperature this compound shows an ordering between iron and lithium atoms leading to a nuclear cell a0√2, a0√2, c0 (a0 and c0 are the parameters of the K2NiF4-type cell). At low temperature (T < TN2) the magnetic structure can be described as antiferromagnetic, corresponding likely to a colinear pattern with a propagation vector of 0.5 (a0) along the [110] axis. At higher temperature (TN2 < T < TN) helimagnetic structure is consistent with a propagation vector of 0.47 (a0) [110].  相似文献   

18.
We present the heat capacities measured by adiabatic calorimetry from 6 to 350 K, and by differential scanning calorimetry from 300 to 500 K, of CsCrCl3 and RbCrCl3. A first-order transition at Tc = (171.1±0.1) K was detected for CsCrCl3. The RbCrCl3 showed at Tc = (193.3±0.1) K a transition with thermal hysteresis at temperatures just below the maximum. At T1 = (440±10) K a continuous transition was also detected. Furthermore, at TN ≈ 16 K, and for both compounds, a small bump due to magnetic long-range ordering was observed. The thermodynamic functions at 298.15 K are
  相似文献   

19.
The lanthanum β-alumina phase doped with europium was investigated by X-ray diffraction and fluorescence. This nonstoichiometric phase exists over the composition range: 11Al2O31La2O3 to 14Al2O31La2O3. The unit cell is hexagonal hexagonal with a = 5.560 ± 0.003 Å, c = 22.001 ± 0.003 Å and belongs to the P63mmc space group. X-ray diffraction patterns do not vary between both boundary compositions, but fluorescence spectra show that the structure of the mirror plane in which the lanthanide ions are located is deeply modified. The atomic structure of the mirror plane is of “β-type” (like β(Na) or β(Ag)) for the lower alumina contents; it gradually changes to a “magnetoplumbite type” for higher alumina contents.  相似文献   

20.
The Gibbs energy of formation of V2O3-saturated spinel CoV2O4 has been measured in the temperature range 900–1700 K using a solid state galvanic cell, which can be represented as Pt, Co + CoV2O4 + V2O3(CaO)ZrO2Co + CoO, Pt. The standard free energy of formation of cobalt vanadite from component oxides can be represented as CoO (rs) + V2O3 (cor) → CoV2O4 (sp), ΔG° = ?30,125 ? 5.06T (± 150) J mole?1. Cation mixing on crystallographically nonequivalent sites of the spinel is responsible for the decrease in free energy with increasing temperature. A correlation between “second law” entropies of formation of cubic 2–3 spinels from component oxides with rock salt and corundum structures and cation distribution is presented. Based on the information obtained in this study and trends in the stability of aluminate and chromite spinels, it can be deduced that copper vanadite is unstable.  相似文献   

Cp,mRSmoR{Hmo(T)?Hmo(0)}RK?{Gmo(T)?Hmo(0}RT
CsCrCl315.3826.493503.214.735
RbCrCl315.7625.993556.814.384
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