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.
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   

2.
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   

3.
The solid-, cholesteric- and liquid-state polymerizations of cholesteryl-vinyl-succinate (CVS) are studied. Only one of the three polymorphic modifications of CVS oligomerizes in the solid state into oligomers of degree P = 2 to 6 in a homogeneous topochemical reaction. The rate of polymerization in the cholesteric state is lower than that in the liquid state at the same temperature. Kinetic constants were measured at 85° using benzoyl peroxide as initiator and the Banfield radical, giving Eoverall = 15.4, 36.4 kcal/mole?1; f = 0.52, 0.19; kp/kt12 = 0.0167, 0.0192, (1/mole sec)12, (Ep ? 12Et) = 0.40, 21.4 kcal mole?1. The values refer to the liquid- and the cholesteric-state reactions, respectively. The average degree of polymerization is low in both cases (P = 20 and 22). It was concluded that the molecular weights are controlled by chain transfer and that the initiation reaction is mostly dependent on the phase where the reaction takes place.  相似文献   

4.
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   

5.
If in a potentiometric titration each addition of reagent equals Δv and if the observed potential steps are, in decreasing order, ΔEm, ΔE1 and ΔE2, the ratio ρα = ΔE2E1is formed; ραΔv is the difference between the equivalence point and the known volume of reagent next to it. Both an extraordinary increase in the precision and a valuable simplification of the analysis are obtained. For Q = ΔEmΔE1 < 2.5 or ρα < 0.25 this value is approximated; but in this case a new series with doubled intervals may be formed from the potential steps observed in which ρα is very near to 0.5, found hence absolutely accurate.  相似文献   

6.
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   

7.
The isothermal decomposition of any ternary oxide AxByOz on liberation of n moles of oxygen at a constant pressure is found to be driven by the mixing entropy ΔSm = ?nRln PO2 of the total entropy change ΔS = ΔS° + ΔSm. The stability of AxByOz towards isothermal decomposition into a biphasic solid mixture is derived from the equilibrium condition ΔG1 = 0 as functions of standard changes ΔH° and ΔS°. Assuming ΔS° = 44n and calculating ΔH° in terms of lattice energies U(ABO3) and U(A2O3), the stability of perovskites St(ABO3) = ?log P1O2 (A = La, Sm, Dy; B = Mn, Fe) is given as a function of the ionic radius of the A3+ ion. The calculated stability agrees well with that observed. The effect of electronic entropy change ΔSe on ΔS° is demonstrated for AFeO3 (A = La, Sm, Dy).  相似文献   

8.
9.
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   

10.
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   

11.
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   

12.
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   

13.
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   

14.
It is shown that the eigenvalues Ei of a Hermitian matrix H with matrix elements Hij = ΣkAkijak, where Akij are known numbers and ak a set of parameters, can be exactly expanded as Ei = Σk(?Ei?ak)ak. This property is applied to the analysis of the optical spectra of transition metal ions in crystals proposed by L. Pueyo, M. Bermejo, and J. W. Richardson (J. Solid State Chem.31, 217, 1980), and it is shown that this method represents the best fit of the Hamiltonian eigenvalues to the observed (or calculated) spectrum. Further advantages of using this property, in connection with the spectral analysis, are the minimization of the errors associated with the numerical approximations and a reduction in computer time. In the molecular orbital calculation of the optical or uv spectra of these systems, this linear expansion of the eigenvalues give a detailed interpretation of the improvements produced by refined calculations, such as those including configuration interaction. In particular, the changes in one-electron energy and in open-shell repulsion interactions associated with the refinement can be clearly and easily formulated. As examples, the computed spectra of CrF4?6 and CrF3?6 are discussed.  相似文献   

15.
Very high resolution measurements of hyperfine structure on the P(13) and R(15), 43?0, 3gp0+u1Σg+ transitions in iodine 127 were made using laser molecular beam spectroscopy. The observed linewidth was 300 kHz (fwhm) giving a resolution of 5 × 10?10 The observed spectrum was fitted to obtain a quadrupole coupling strength difference of ΔeQq = 1906 ± 2 MHz and a spin rotation interaction strength difference of ΔCI = 181 ± 7 kHz between the upper and lower levels of the P(13) transition. For the R (15) transition ΔeQq = 1905 ± 2 MHz and ΔCI = 167 ± 5 kHz.  相似文献   

16.
Single crystals of PdPSe were shown to be n-type semiconductors. Weak Pauli paramagnetic behavior was observed, which is consistent with the presence of delocalized electrons. Electrical measurements showed a room-temperature resistivity ? = 70 ohm-cm, activation energy of resistivity Ea = 0.32 eV, and Hall mobility μ = 34 cm2 V?1 sec?1. Photoelectronic measurements in aqueous solutions of I?I?3 indicate that PdPSe has high quantum efficiencies below 800 nm. The indirect optical band gap is 1.28(2) eV.  相似文献   

17.
The thermodynamic properties of the spinel Mg2SnO4 have been determined by emf measurements on the solid oxide galvanic cell,
Pt,W,MgO+Mg2SnO4+SnY2O3?ThO2Sn+SnO2,W,Pt
in the temperature range 600 to 1000°C. The Gibbs' free energy of formation of Mg2SnO4 from the component oxides can be expressed as
2MgO (r.s)+SnO2 (rut)→Mg2SnO4 (sp)
,
ΔG°=1420?2.96T (±100)cal mole?1
These values are in good agreement with the information obtained by Jackson et al. [Earth Planet. Sci. Lett.24, 203 (1974)] on the high pressure decomposition of magnesium stannate into component oxides at different temperatures. The thermodynamic data suggest that the spinel phase is entropy stabilized, and would be unstable below 207 (±25)°C at atmospheric pressure. Based on the information obtained in this study and trends in the stability of aluminate and chromite spinels, it can be deduced that the stannates of nickel and copper(II) are unstable.  相似文献   

18.
Enthalpies of reaction, ΔHr, of the monouranates of lithium, potassium, and rubidium with 1 mol dm?3 HCl have been measured calorimetrically. From these measurements, and auxiliary determinations of the enthalpies of solution in acid of the chlorides of lithium, potassium, and rubidium and of uranyl chloride, the standard enthalpies of formation of the uranates, ΔHfo, have been derived. The results obtained are as follows:
  相似文献   

19.
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:
ΔHr/kcalth mol?1ΔHfo(c, 298.15 K)/kcalth mol?1
α-Li2UO4?(41.77±0.02)?(463.31±0.84)
K2UO4?(42.07±0.05)?(451.39±0.83)
Rb2UO4?(41.30±0.05)?(452.00±0.85)
  相似文献   

20.
The paper refers to the synthesis and properties of some bifunctional initiators, viz. 4.4′-azo-bis(4-cyanovaleryl)bisbenzoyl diperoxide (I) and 4,4′-azobis(4-cyanovaleryl)bisacetyl diperoxide (II), obtained from the acid chloride of cyanovaleric acid and condensed with perbenzoic acid or peracetic acid. The structures of the products were established by i.r. and NMR spectroscopy, as well as by elemental analysis. Kinetic studies on the thermolyses of the two initiators led to the following results: for I. t12azo, 360°K = 129 min, t12peroxy, 360°K = 245 min, Eazo = 174.2 kJ/mol, Eperoxy = 206.5 kJ/mol; for II, t12azo, 357°K = 152 min, t12peroxy, 357°K = 208 min, Eazo = 155.4 kJ/mol, Eperoxy = 188.5 kJ/mol.  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