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.
Specific heat capacities (Cp) of polycrystalline samples of BaCeO3 and BaZrO3 have been measured from about 1.6 K up to room temperature by means of adiabatic calorimetry. We provide corrected experimental data for the heat capacity of BaCeO3 in the range T < 10 K and, for the first time, contribute experimental data below 53 K for BaZrO3. Applying Debye's T3-law for T → 0 K, thermodynamic functions as molar entropy and enthalpy are derived by integration. We obtain Cp = 114.8 (±1.0) J mol−1 K−1, S° = 145.8 (±0.7) J mol−1 K−1 for BaCeO3 and Cp = 107.0 (±1.0) J mol−1 K−1, S° = 125.5 (±0.6) J mol−1 K−1 for BaZrO3 at 298.15 K. These results are in overall agreement with previously reported studies but slightly deviating, in both cases. Evaluations of Cp(T) yield Debye temperatures and identify deviations from the simple Debye-theory due to extra vibrational modes as well as anharmonicity. The anharmonicity turns out to be more pronounced at elevated temperatures for BaCeO3. The characteristic Debye temperatures determined at T = 0 K are Θ0 = 365 (±6) K for BaCeO3 and Θ0 = 402 (±9) K for BaZrO3.  相似文献   

2.
The (p,ρ,T) and (ps,ρs,Ts) properties of {(1−x)CH3OH + xLiBr} over a wide range of state parameters are reported for the first time. The experiments were carried out in a constant volume piezometer over a temperature range from 298.15 K to 398.15 K, at 0.08421, 0.13617, 0.19692, 0.23133 and 0.26891 mole fractions and from atmospheric pressure up to 60 MPa. The experimental uncertainties are ΔT=±3 mK for temperature, Δp=±5·10−2 MPa for high pressure and Δp=±5·10−4 MPa for atmospheric pressure, Δρ=±3·10−2 kg · m−3 for density. An equation of state was derived for correlation of the experimental data of the solutions.  相似文献   

3.
Heat capacity and enthalpy increments of ternary bismuth tantalum oxides Bi4Ta2O11, Bi7Ta3O18 and Bi3TaO7 were measured by the relaxation time method (2-280 K), DSC (265-353 K) and drop calorimetry (622-1322 K). Temperature dependencies of the molar heat capacity in the form Cpm=445.8+0.005451T−7.489×106/T2 J K−1 mol−1, Cpm=699.0+0.05276T−9.956×106/T2 J K−1 mol−1 and Cpm=251.6+0.06705T−3.237×106/T2 J K−1 mol−1 for Bi3TaO7, Bi4Ta2O11 and for Bi7Ta3O18, respectively, were derived by the least-squares method from the experimental data. The molar entropies at 298.15 K, S°m(298.15 K)=449.6±2.3 J K−1 mol−1 for Bi4Ta2O11, S°m(298.15 K)=743.0±3.8 J K−1 mol−1 for Bi7Ta3O18 and S°m(298.15 K)=304.3±1.6 J K−1 mol−1 for Bi3TaO7, were evaluated from the low-temperature heat capacity measurements.  相似文献   

4.
The (p, ρ, T) properties of pure methanol, the (p, ρ, T) properties and apparent molar volumes V? of ZnBr2 in methanol at T = (298.15 to 398.15) K and pressures up to p = 40 MPa are reported, and apparent molar volumes have been evaluated. The experimental (p, ρ, T, m) values were described by an equation of state. For the solutions the experiments were carried out at molalities m = (0.05772, 0.37852, 0.71585 and 1.95061) mol · kg−1 of zinc bromide.  相似文献   

5.
In vitro degradation of poly(ethyl glyoxylate) (PEtG), a functionalised polyacetal, was investigated. First, the thermodynamic polymerization parameters and the ceiling temperature (Tc) were determined (ΔHp = 28 ± 3 kJ mol−1, ΔSp = 98 ± 7 J mol−1 K−1, Tc = 310 ± 4 K). Secondly, PEtG hydrolysis was investigated using potentiometry, weight loss measurements, SEC and 1H NMR. The results show that PEtG is stable for at least 7 days in aqueous media. Then degradation occurs and releases ethanol and glyoxylic acid hydrate as final products. A scheme for the degradation mechanism involving chain scission and ester hydrolysis is proposed.  相似文献   

6.
The heat capacity and the heat content of bismuth niobate BiNbO4 and bismuth tantalate BiTaO4 were measured by the relaxation method and Calvet-type heat flux calorimetry. The temperature dependencies of the heat capacities in the form Cpm=128.628+0.03340 T−1991055/T2+136273131/T3 (J K-1 mol-1) and 133.594+0.02539 T−2734386/T2+235597393/T3 (J K-1 mol-1) were derived for BiNbO4 and BiTaO4, respectively, by the least-squares method from the experimental data. Furthermore, the standard molar entropies at 298.15 K Sm(BiNbO4)=147.86 J K-1 mol-1 and Sm(BiTaO4)=149.11 J K-1 mol-1 were assessed from the low temperature heat capacity measurements. To complete a set of thermodynamic data of these mixed oxides an attempt was made to estimate the values of the heat of formation from the constituent binary oxides.  相似文献   

7.
Sipos P  Hefter G  May PM 《Talanta》2006,70(4):761-765
27Al NMR and Raman spectra of alkaline aluminate solutions with 0.005 M ≤ [Al(III)]T ≤ 3 M in various M′OH solutions (M′+ = Na+, K+ and Li+) were recorded and analysed. Caustic concentrations up to 20 M were used to explore whether higher aluminium hydroxo complexes are formed at extremely high concentrations of hydroxide. A single peak was observed on the 27Al NMR spectrum of each solution. The chemical shift of this peak shifts significantly upfield with increasing [M′OH]T in solutions with [Al(III)]T < 0.8 M. This variation shows a strong dependence on the cation of the solution and practically disappears in systems with [Al(III)]T ≥ 0.8 M. For Raman spectra of solutions with [Al(III)]T = 0.8 M and [NaOH]T ≥ 10 M, the peak maximum of the symmetric ν1-AlO4 stretching of Al(OH)4 shifted progressively from ∼620 to ∼625 cm−1 and decreased in intensity with increasing [NaOH]T. In parallel, modes centred at ∼720 and ∼555 cm−1 (cf. ∼705 and ∼535 cm−1 at lower [NaOH]T, ascribed to a dimeric aluminate species appeared, and their intensities increased with increasing [NaOH]T. These variations in the 27Al NMR and Raman spectra can be interpreted in terms of contact ion-pairs formed between the cation of the medium and the well-established Al(OH)4 or the dimeric aluminate species. Assumption of higher aluminium hydroxo complex species (e.g., Al(OH)63−) is not necessary to explain the spectroscopic effects observed.  相似文献   

8.
The magnetic structure of the Fe2P-type R6CoTe2 phases (R=Gd-Er, space group P6¯2m) has been investigated through magnetization measurement and neutron powder diffraction. All phases demonstrate high-temperature ferromagnetic and low-temperature transitions: TC=220 K and TCN=180 K for Gd6CoTe2, TC=174 K and TCN=52 K for Tb6CoTe2, TC=125 K and TCN=26 K for Dy6CoTe2, TCN=60 K and TN=22 K for Ho6CoTe2 and TCN∼30 K and TN∼14 K for Er6CoTe2.Between 174 and 52 K Tb6CoTe2 has a collinear magnetic structure with K0=[0, 0, 0] and with magnetic moments along the c-axis, whereas below 52 K it adopts a non-collinear ferromagnetic one.Below 60 K the magnetic structure of Ho6CoTe2 is that of a non-collinear ferromagnet. The holmium magnetic components with a K0=[0, 0, 0] wave vector are aligned ferromagneticaly along the c-axis, whereas the magnetic component with a K1=[1/2, 1/2, 0] wave vector are arranged in the ab plane. The low-temperature magnetic transition at ∼22 K coincides with the reorientation of the Ho magnetic component with the K0 vector from the collinear to the non-collinear state.Below 30 K Er6CoTe2 shows an amplitude-modulate magnetic structure with a collinear arrangement of magnetic components with K0=[0, 0, 0] and K1=[1/2, 1/2, 0]. The low-temperature magnetic transition at ∼14 K corresponds to the variation in the magnitudes of the MErK0 and MErK1 magnetic components.In these phases, no local moment was detected on the cobalt site.The magnetic entropy of Gd6CoTe2 increases from ΔSmag=−4.5 J/kg K at 220 K up to ΔSmag=−6.5 J/kg K at 180 K for the field change Δμ0H=0-5 T.  相似文献   

9.
(pρT) Measurements and visual observations of the meniscus for isobutane were carried out carefully in the critical region over the range of temperatures: −15 mK ? (T − Tc) ? 35 mK, and of densities: −7.5 kg · m−3 ? (ρ − ρc) ? 7.5 kg · m−3 by a metal-bellows volumometer with an optical cell. Vapor pressures were also measured at T = (310, 405, 406, 407, and 407.5) K. The critical point of Tc and ρc was determined by the image analysis of the critical opalescence which is proposed in this study. The critical pressure pc was determined to be the pressure measurement at the critical point. Comparisons of the critical parameters with values given in the literature are presented.  相似文献   

10.
We determined apparent molar volumes V? from densities measured with a vibrating-tube densimeter at 278.15 ? (T/K) ? 368.15 and apparent molar heat capacities Cp,? with a twin fixed-cell, differential, temperature-scanning calorimeter at 278.15 ? (T/K) ? 363.15 for aqueous solutions of N-acetyl-d-glucosamine at m from (0.01 to 1.0) mol · kg−1 and at p = 0.35 MPa. We also determined V? at 278.15 ? (T/K) ? 368.15 and Cp,? at 278.15 ? (T/K) ? 393.15 for aqueous solutions of N-methylacetamide at m from (0.015 to 1.0) mol · kg−1 and at p = 0.35 MPa. Empirical functions of m and T for each compound were fitted to our results, which are then compared to those for N,N-dimethylacetamide. Estimated values of ΔrVm(mT) and ΔrCp,m(mT) for formation of aqueous N-acetyl-d-glucosamine from aqueous d-glucose and aqueous acetamide are calculated and discussed.  相似文献   

11.
The thermal conductivity and heat capacity of high-purity single crystals of yttrium titanate, Y2Ti2O7, have been determined over the temperature range 2 K?T?300 K. The experimental heat capacity is in very good agreement with an analysis based on three acoustic modes per unit cell (with the Debye characteristic temperature, θD, of ca. 970 K) and an assignment of the remaining 63 optic modes, as well as a correction for CpCv. From the integrated heat capacity data, the enthalpy and entropy relative to absolute zero, are, respectively, H(T=298.15 K)−H0=34.69 kJ mol−1 and S(T=298.15 K)−S0=211.2 J K−1 mol−1. The thermal conductivity shows a peak at ca. θD/50, characteristic of a highly purified crystal in which the phonon mean free path is about 10 μm in the defect/boundary low-temperature limit. The room-temperature thermal conductivity of Y2Ti2O7 is 2.8 W m−1 K−1, close to the calculated theoretical thermal conductivity, κmin, for fully coupled phonons at high temperatures.  相似文献   

12.
Thermodynamics of chromium acetylacetonate sublimation   总被引:1,自引:0,他引:1  
The equilibrium sublimation pressure Cr(acac)3(s) = Cr(acac)3(g) was measured in the range 320 ≤ T (K) ≤ 476 by two procedures. One of them is Knudsen's effusion procedure with mass spectrometric analysis of the composition of the gas phase, which proved to be good in measuring low pressure. The second is mass spectrometric procedure “calibrated volume method” (CVM), which helped us to expand the possibilities of the effusion method toward high pressure range. Experimental data are in good agreement with each other.For this process were obtained ln(P (Pa)) = 39.197 − 15 308.5/T, enthalpy ΔsubH° (T) = 127.28 ± 0.22 kJ mol−1 and entropy ΔsubS° (T) = 230.1 ± 0.5 J mol−1 K−1.  相似文献   

13.
Nuclear magnetic resonance (1H NMR and 19F NMR) measurements performed at 90-295 K, inelastic incoherent neutron scattering (IINS) spectra and neutron powder diffraction (NPD) patterns registered at 22-190 K, and X-ray powder diffraction (XRPD) measurements performed at 86-293 K, provided evidence that the crystal of [Zn(NH3)4](BF4)2 has four solid phases. The phase transitions occurring at: TC3=101 K, TC2=117 K and TC1=178 K, as were detected earlier by differential scanning calorimetry (DSC), were connected on one hand only with an insignificant change in the crystal structure and on the other hand with a drastic change in the speed of the anisotropic, uniaxial reorientational motions of the NH3 ligands and BF4 anions (at TC3 and at TC2) and with the dynamical orientational order-disorder process (“tumbling”) of tetrahedral [Zn(NH3)4]2+ and BF4 ions (at TC1). The crystal structure of [Zn(NH3)4](BF4)2 at room temperature was determined by XRPD as orthorhombic, space group Pnma (No. 62), a=10.523 Å, b=7.892 Å, c=13.354 Å and Z=4. Unfortunately, it was not possible to determine the structure of the intermediate and the low-temperature phase. However, we registered the change of the lattice parameters and unit cell volume as a function of temperature and we can observe only a small deviation from near linear dependence of these parameters upon temperature in the vicinity of the TC1 phase transition.  相似文献   

14.
Magnetism for the LixMn2O4 samples with 0.07≤x≤1, which are prepared by a chemical reaction in HNO3 solution, is investigated by direct current susceptibility (χ) and muon-spin rotation/relaxation (μSR) measurements. The effective magnetic moment (μeff) of Mn ions decreases monotonically with decreasing x, indicating that Mn3+ ions with S=2 () are oxidized to Mn4+ ions with S=3/2 () with decreasing x. On the other hand, as x decreases from 1 to 0.6, the Curie-Weiss temperature (Θp) increases monotonically from ∼260 to 100 K, and then levels off to 100 K with further decreasing x. This indicates that the antiferromagnetic interaction is dominant in the whole x range. For the x=0.48 sample, the temperature dependence of χ in field-cooling mode clearly deviates from that in zero-field-cooling mode below ∼63 K (=Tm). Furthermore, the hysteresis loop is observed in the magnetization vs. field curve at 5 K. Since the zero-field μSR spectrum is well fitted by a strongly damped oscillation function, the Mn moments for the x=0.48 sample are in a highly disordered fashion down to the lowest temperature measured.  相似文献   

15.
[Ni(ND3)6](ClO4)2 has three solid phases between 100 and 300 K. The phase transitions temperatures at heating (TC1h=164.1 K and TC2h=145.1 K) are shifted, as compared to the non-deuterated compound, towards the lower temperature of ca. 8 and 5 K, respectively. The ClO4 anions perform fast, picosecond, isotropic reorientation with the activation energy of 6.6 kJ mol−1, which abruptly slow down at TC1c phase transition, during sample cooling. The ND3 ligands perform fast uniaxial reorientation around the Ni-N bond in all three detected phases, with the effective activation energy of 2.9 kJ mol−1. The reorientational motion of ND3 is only slightly distorted at the TC1 phase transition due to the dynamical orientational order-disorder process of anions. The low value of the activation energy for the ND3 reorientation suggests that this reorientation undergoes the translation-rotation coupling, which makes the barrier to the rotation of the ammonia ligands not constant but fluctuating. The phase polymorphism and the dynamics of the molecular reorientations of the title compound are similar but not quite identical with these of the [Ni(NH3)6](ClO4)2.  相似文献   

16.
The crystal structure of an Li-bearing double-ring silicate mineral, sogdianite ((Zr1.18Fe3+0.55Ti0.24Al0.03)(?1.64,Na0.36)K0.85[Li3Si12O30], P6/mcc, a≈10.06 Å, c≈14.30 Å, Z=2), was investigated by neutron powder diffraction from 300 up to 1273 K. Rietveld refinements of displacement parameters revealed high anisotropic Li motions perpendicular to the crystallographic c-axis, indicating an exchange process between tetrahedral T2 and octahedral A sites. AC impedance spectra of a sogdianite single crystal (0.04×0.09×0.25 cm3) show that the material is an ionic conductor with conductivity values of σ=4.1×10−5 S cm−1 at 923 K and 1.2×10−3 S cm−1 at 1219 K perpendicular to the c-axis, involving two relaxation processes with activation energies of 1.26(3) and 1.08(3) eV, respectively.  相似文献   

17.
Phase stability of the type-I clathrate compound Ba8AlxSi46−x and the thermoelectric property dependence on chemical composition are presented. Polycrystalline samples were prepared by argon arc melting and annealing. Results of powder X-ray diffraction and electron microprobe analysis show that the type-I structure is formed without framework deficiency for 8≤x≤15. Lattice constant a increases linearly with the increase of x. Thermoelectric properties were measured for x=12, 14 and 15. The Seebeck coefficients are negative, with the absolute values increasing with x. The highest figure of merit zT=0.24 was observed for x=15 at T=1000 K, with carrier electron density n=3×1021 cm−3. A theoretical calculation based on the single parabolic band model reveals the optimum carrier concentration to be n∼4×1020 cm−3, where zT∼0.7 at T=1000 K is predicted.  相似文献   

18.
Copper(II) compounds {CuCA(phz)(H2O)2}n (H2CA = chloranilic acid, phz = phenazine) having a layer structure of -CuCA(H2O)2- polymer chains and phenazine were studied by 35Cl nuclear quadrupole resonance (NQR). The single NQR line observed at 35.635 MHz at 261.5 K increased to 35.918 MHz at 4.2 K. The degree of reduction of electric field gradient due to lattice vibrations was similar to that of chloranilic acid crystal. Temperature dependence of spin-lattice relaxation time, T1, of the 35Cl NQR signal below 20 K, between 20 and 210 K, and above 210 K, was explained by (1) a decrease of effective electron-spin density caused by antiferromagnetic interaction, (2) a magnetic interaction between Cl nuclear-spin and electron-spins on paramagnetic Cu(II) ions, and (3) an increasing contribution from reorientation of ligand molecules, respectively. The electron spin-exchange parameter ∣J∣ between the neighboring Cu(II) electrons was estimated to be 0.33 cm−1 from the T1 value of the range 20−210 K. Comparing this value with that of J = −1.84 cm−1 estimated from the magnetic susceptibility, it is suggested that the magnetic dipolar coupling with the electron spins on Cu(II) ions must be the principal mechanism for the 35Cl NQR spin-lattice relaxation of {CuCA(phz)(H2O)2}n but a delocalization of electron spin over the chloranilate ligand has to be taken into account.  相似文献   

19.
The potential energy surface for the reaction of CH3S with CO was calculated at the G3MP2//B3LYP/6-311++G(d,p) level. The rate constants for feasible channels leading to several products were calculated by TST and multichannel-RRKM theory. The results show that addition–elimination mechanism is dominant, while hydrogen abstraction mechanism is uncompetitive. The major channel is the addition of CO to CH3S leading to an intermediate CH3SCO which then decomposes to CH3 + OCS. In the temperature range of 200–3000 K, the overall rate constants are positive temperature dependence and pressure independence, and it can be described by the expression as k = 1.10 × 10−16T1.57exp(−3359/T) cm3 molecule−1 s−1. At temperature between 208 and 295 K, the calculated rate constants are in good agreement with the experimental upper limit data. At T = 1000 and 2000 K, the major product is CH3 + OCS at lower pressure; while at higher pressure, the stabilization of IM1 is dominant channel.  相似文献   

20.
Lithium monosilicide (LiSi) was formed at high pressures and high temperatures (1.0-2.5 GPa and 500-700°C) in a piston-cylinder apparatus. This compound was previously shown to have an unusual structure based on 3-fold coordinated silicon atoms arranged into interpenetrating sheets. In the present investigation, lowered synthesis pressures permitted recovery of large (150-200 mg) quantities of sample for structural studies via NMR spectroscopy (29Si and 7Li), Raman spectroscopy and electrical conductivity measurements. The 29Si chemical shift occurs at −106.5 ppm, intermediate between SiH4 and Si(Si(CH3)3)4, but lies off the trend established by the other alkali monosilicides (NaSi, KSi, RbSi, CsSi), that contain isolated Si44− anions. Raman spectra show a strong peak at 508 cm−1 due to symmetric Si-Si stretching vibrations, at lower frequency than for tetrahedrally coordinated Si frameworks, due to the longer Si-Si bonds in the 3-coordinated silicide. Higher frequency vibrations occur due to asymmetric stretching. Electrical conductivity measurements indicate LiSi is a narrow-gap semiconductor (Eb∼0.057 eV). There is a rapid increase in conductivity above T=450 K, that might be due to the onset of Li+ mobility.  相似文献   

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