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 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   

2.
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   

3.
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   

4.
Translational diffusion, velocity sedimentation and viscometry of polyamidobenzimidazole (PABI) solutions in the range of M = (1–61) · 103 have been investigated in N,N-dimethylacetamide (DMA) and 98% H2SO4. The dependences of D0, S0 and [η] on M were obtained. Tsvetkov-Klenin's hydrodynamic invariant was found to be A0 = 3.55 · 10?10erg deg?1mol?13. The equilibrium rigidity of PABI molecules was characterized by the length of the Kuhn segment A = 250 ± 100 A?. The chain diameter was 7 ± 4 A?. The values of A in 98% H2SO4 and in an aprotic solvent, DMA, were virtually identical, implying that the rigid-chain conformation of PABI molecules in 98% H2SO4 is due to their geometrical structure rather than to the protonization of amide bonds. The significance of the latter evidently increases in PABI solutions in 100% H2SO4 in which A is 1.5 times as high. The decrease in rigidity of PABI as compared to that of poly-p-phenylene terephthalamide (A = 400 ± 100 A?) is in reasonable agreement with the presence of imidazole rings in PABI molecules. The presence of these rings results in kinks in the PABI chain with angles of about 30° and hence, in the depature from parallelism of rotating bonds.  相似文献   

5.
6.
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   

7.
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   

8.
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   

9.
The radiation induced solid-state polymerization and post-polymerization of crystalline acetaldehyde were studied in a diathermic calorimeter by measuring the heat evolution during polymerization. The heat of melting of crystalline acetaldehyde was found to be 1,4 ± 0,07 kcal mol?1 and the heat of polymerization 2,5 ± 0,5 kcal mol?1 at 80–150°K. Under isothermal conditions the rate of the solid state polymerization of acetaldehyde increased with irradiation time up to a maximum and thereafter it decreased. This phenomenon is connected with an increase of the concentration of active centres during irradiation. The propagation rate constant is kp ? 5 × 10?4exp(?11,000/RT) cm3 sec?1 at 130–140°K and the average time of addition of one monomer unit is 10?1–10?2 sec.  相似文献   

10.
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   

11.
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   

12.
Very narrow fractions of polyvinylpyrrolidone (PVP) for a range of low molecular weight from 20·103 to &{;103 were prepared by gel filtration using sephadex gel. Interactions between fractions of different Mn and small molecules (iodine and 1-anilinonaphthaline-8-sulphonate) and polymers (polymethacrylic and polyacrylic acids) were studied in aqueous solution. The ability of PVP to form complexes with low molecular weight compounds depends on chainlength. The greatest loss of the ability was observed for PVP with Mn on passing from 5·103 to &{;103. The chainlength effects for PVP are accounted for an unlike dehydration and unlike “local” link concentration near links for these macromolecules in water. For PAA and PMAA. the threshold values of M?n, below which there is the beginning of weakening of complexation for PVP, are 6·103 and 2.5·103 respectively. The difference of the complex formation for these polyacids appears to be related to hydrophobic interactions between the χ-methyl groups of PMAA and nonpolar regions of PVP.  相似文献   

13.
Compounds MIIMeIVF6 requently undergo phase transitions from the cubic ordered ReO3 to the trigonal LiSbF6 structure when lowering the temperature. In case of a strongly Jahn-Teller unstable cation in the MII position additional phases may occur. Results of powder neutron-diffraction studies on CaSnF6, FeZrF6, and CrZrF6 at different temperatures are reported. The high-temperature phases have the space group Fm3m; the F? ligands are either statistically displaced from the MIIMeIV directions or undergo a strong thermal motion perpendicular to these directions (?MIIFMeIV: 165–180°). The thermal ellipsoids of the CrF bonds are strongly indicative of a dynamical Jahn-Teller effect in addition. In the low-temperature phases of CaSnF6 and FeZrF6 (space group R3) the ?MIIFMeIV is more distinctly bent (?155–160°). CrZrF6 undergoes two reversible phase transitions, which are determined to occur at 415 ± 5 K (cubic → tetragonal, dynamic to static Jahn-Teller distortion of CrF6 octahedra and 150 ± 10 K (tetragonal → (pseudo)monoclinic).  相似文献   

14.
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   

15.
The relaxation of Pb2+-vacancy and Cd2+-vacancy dipoles in purified KCl crystals was studied using a double crystal dc polarization method which self-corrects relaxation effects due to spurious causes. In the radius range from Cd2+ to Ba2+ these results and most others support Dreyfus' model. The reciprocal relaxation times for each impurity are given by
τ?1Cd=2.06x1012exp(?s0.64±0.01eV)kT
;
τ?1Pb=1.19x1013exp(?0.69±0.01eV)kT
.It is also shown that the presence of H2O or its products greatly perturb the relaxation times observed.  相似文献   

16.
The reaction between nitric oxide and vibrationally excited ozone was studied in a fast flow reactor by monitoring the visible emission from electronically excited NO21. The antisymmetric mode (ν3) of O3 was excited with a Q-switched 9.6 μm CO2 laser, and a laser-induced signal was detected, with a rise rate constant of (4.0 ± 0.5) × 1011 cm3/mole sec and a decay rate constant of (1.1 ± 0.1) × 1011 cm3/mole sec for an NO-rich mixture. The latter was unaffected by addition of large amounts of He or Ar, indicating that the signal was not a thermal effect. Most of the measurements were made at 350°K; however, the He and Ar dilution results suggest that the enhanced reaction rate is not very sensitive to temperature. In order to explain the observed rise times, it was necessary to postulate an intermediate step prior to the chemical reaction. A model which is consistent with our data has energy transferred from ν3 to ν2 (the bending mode) at a rate of (2.9 ± 0.5) × 1011 cm3/mole sec for NO and a rate of (1.1 ± 0.2) × 1011 cm3/mole sec for He. According to this model, the rate constant for the reaction of NO with O3 (ν2= 1) producing vibrationally excited ground state NO22,
NO + O32 (010) 3 NO22 + O2
is (1.5 ± 0.2) × 1011 cm3/mole sec, and the relative rate for the reaction of O3 (ν2 = 1) and O32 = 0) with NO was estimated to be k3(1)k3(0) ≈ 22.  相似文献   

17.
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   

18.
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   

19.
The energies of combustion of 3,4- and 3,5-dimethylbenzoic acids have been revised by combustion calorimetry. Vapour pressures of very pure samples of the six dimethylbenzoic acids have been determined over a range of temperatures near 298 K by the Knudsen-effusion technique. From the experimental results and our previously published thermochemical quantities the following results for the six C6H3(CH3)2CO2H isomers at 298.15 K have been derived.
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)
  相似文献   

20.
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   

IsomerΔfHmo(cr)ΔsubHmoΔfHmo(g)
kJ·mol?1kJ·mol?1kJ·mol?1
2,6-?440.7 ± 1.799.1 ± 0.2?341.6 ± 1.7
2,3-?450.4 ± 1.7104.6 ± 0.4?345.8 ± 1.7
2,5-?456.1 ± 1.6105.0 ± 0.6?351.1 ± 1.7
2,4-?458.5 ± 1.7103.5 ± 0.3?355.0 ± 1.7
3,4-?468.8 ± 1.9106.4 ± 0.3?362.4 ± 1.9
3,5-?466.8 ± 1.7102.3 ± 0.3?364.5 ± 1.7
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