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 15 毫秒
1.
A small sample adiabatic calorimeter for measuring heat capacities in the temperature range 60–350 K using the Nernst method has been constructed. The sample cell of the calorimeter is 6 cm3 in the internal volume, equipped with a miniature platinum thermometer and surrounded by two adiabatic shields. Two sets of 6-junction chromel-copel thermocouples were mounted between the cell and the shields to indicate the temperature differences between them. The adiabatic conditions of the cell were automatically controlled by two sets of temperature controller. A mechanical pump was used to pump out the vapour of liquid nitrogen in the cryostat to solidify N2 (1), and 60 K or even lower temperature was obtained. The performance of this apparatus was evaluated by heat capacity measurements on α-alumina. The deviations of experimental results from a smoothed curve lie within ±0.2%, while the inaccuracy is within ±0.5% compared with the recommended reference data in the wole temperature range.  相似文献   

2.
Low-temperature heat capacity of natural zinnwaldite was measured at temperatures from 6 to 303 K in a vacuum adiabatic calorimeter. An anomalous behavior of heat capacity function C p(T) has been revealed at very low temperatures, where this function does not tend to zero. Thermodynamic functions of zinnwaldite have been calculated from the experimental data. At 298.15 K, heat capacity C p(T) = 339.8 J K−1mol−1, calorimetric entropy S o(Т) – S o(6.08) = 329.1 J K−1 mol−1, and enthalpy Н o(Т) − Н o(6.08) = 54,000 J mol−1. Heat capacity and thermodynamic functions at 298.15 K for zinnwaldite having theoretical composition were estimated using additive method of calculation.  相似文献   

3.
4.
The heat capacity of calcium monouranate CaUO4 and strontium monouranate SrUO4 have been measured over the temperature range (2 to 300) K. The results are significantly different from earlier measurements, confirming the conclusions from our previous study on BaUO4. The standard entropy at T = 298.15 K of orthorhombic α-SrUO4 is consistent with the values for the isostructural BaUO4 as well as the alkali uranates Na2UO4 and Cs2UO4. The standard entropy of the rhombohedral CaUO4 is appreciably different, which is attributed to the different structures of the uranium sublattices.  相似文献   

5.
The results of heat capacity measurements for several crystalline uranoborates over the temperature range 0–340 K were discussed and analyzed. Low-temperature heat capacities (T < 50 K) were considered using the Debye theory of the heat capacity of solids and its multifractal generalization. The fractal dimensions of compounds were calculated and the heterodynamic characteristics of their structures determined.  相似文献   

6.
The low-temperature heat capacity of synthetic huttonite ThSiO4 has been measured from T = (2 to 300) K. The sample was synthesised successfully from SiO2 and ThO2 by solid-state reaction at T = 1873 K at atmospheric pressure. From the calorimetric results, the value for the standard entropy Sm° (ThSiO4, huttonite, 298.15 K) = (104.3 ± 2.0) J · K?1 · mol?1 has been obtained. This value indicates that the entropy of reaction from SiO2 and ThO2 is negative, giving a positive entropy term (?T · ΔrS) of the Gibbs free energy of reaction. The implications of this finding are discussed extensively.  相似文献   

7.
The dynamic heat capacity of a simple polymeric, model glassformer was computed using molecular dynamics simulations by sinusoidally driving the temperature and recording the resultant energy. The underlying potential energy landscape of the system was probed by taking a time series of particle positions and quenching them. The resulting dynamic heat capacity demonstrates that the long time relaxation is the direct result of dynamics resulting from the potential energy landscape. Moreover, the equilibrium (low frequency) portion of the potential energy landscape contribution to the heat capacity is found to increase rapidly at low temperatures and at high packing fractions. This increase in the heat capacity is explained by a statistical mechanical model based on the distribution of minima in the potential energy landscape.  相似文献   

8.
Evident discrepancy (by six times) between low-temperature heat capacities of two different samples of magnesioferrite (MgFe2O4) with a different thermal history is analogous to that (by four times) for franklinite (ZnFe2O4) found 50 years ago. The reason is supposed to be in the different magnetic ordering caused by different cation ordering after quenching–annealing. Magnetic nature of the discrepancy in low-temperature thermodynamic functions of minerals with variable cation ordering is out of quantitative consideration so far. This effect can also be found in all oxide phases of magnetic cations with the inversion parameter.  相似文献   

9.
The heat capacities of isobutyl tert-butyl ether in crystalline, liquid, supercooled liquid, and glassy states were measured by vacuum adiabatic calorimetry over the temperature range from (7.68 to 353.42) K. The purity of the substance, the glass-transition temperature, the triple point and fusion temperatures, and the enthalpy and entropy of fusion were determined. Based on the experimental data, the thermodynamic functions (absolute entropy and changes of the enthalpy and Gibbs free energy) were calculated for the solid and liquid states over the temperature range studied and for the ideal gas state at T = 298.15 K. The ideal gas heat capacity and other thermodynamic functions in wide temperature range were calculated by statistical thermodynamics method using molecular parameters determined from density-functional theory. Empirical correction for coupling of rotating groups was used to calculate the internal rotational contributions to thermodynamic functions. This correction was found by fitting to the calorimetric entropy values.  相似文献   

10.
聚吡咯的电子能带结构   总被引:1,自引:1,他引:0  
严继民  张普文 《化学学报》1986,44(10):1081-1086
本文用EHMO方法计算了类苯型,类等键长型及类醌型等聚吡咯的电子能带.分析了原子电荷分布及价带,导带与能隙的宽度,并由此探讨了聚吡咯可能的电导机理.  相似文献   

11.
The heat capacities of starch and starch—water have been measured with adiabatic calorimetry and standard differential scanning calorimetry and are reported from 8 to 490 K. The amorphous starch containing 11–26 wt % (53–76 mol %) water shows a partial glass transition decreasing from 372 to 270 K, respectively. Even the dry amorphous starch gradually increases in heat capacity above 270 K beyond that set by the vibrational density of states. This gradual increase in the heat capacity is identified as part of the glass transition of dry starch that is, however, not completed at the decomposition temperature. The heat capacities of the glassy, dry starch are linked to an approximate group vibrational spectrum with 44 degrees of freedom. The Tarasov equation is used to estimate the heat capacity contribution due to skeletal vibrations with the parameters Θ1 = 795.5 K, Θ2 = 159 K, and Θ3 = 58 K for 19 degrees of freedom. The calculated and experimental heat capacities agree better than ±3% between 8 and 250 K. Similarly, the vibrational heat capacity has been estimated for glassy water by being linked to an approximate group vibrational spectrum and the Tarasov equation (Θ1 = 1105.5 K and Θ3 = 72.4 K, with 6 degrees of freedom). Below the glass transition, the heat capacity of the solid starch—water system has been estimated from the appropriate sum of its components and also from a direct fitting to skeletal vibrations. Above the glass transition, the differences are interpreted as contributions of different conformational heat capacities from chains of the carbohydrates interacting with water. The conformational parts are estimated from the experimental heat capacities of dry starch and starch—water, decreased by the vibrational and external contributions to the heat capacity. © 2001 John Wiley & Sons, Inc. J Polym Sci Part B: Polym Phys 39: 3038–3054, 2001  相似文献   

12.
The heat capacity, thermodynamic properties of fusion, and purity of the ethyl ester of butanoic acid were determined by adiabatic calorimetry in the temperature range from 8 to 372 K. The pT-parameters of the ester for the equilibrium liquid-vapor were measured by comparative ebulliometry in the “atmospheric” range of pressure from 10.8 to 101.7 kPa. The obtained data were used to derive the normal boiling temperature (T n.b), the enthalpies of vaporization at T = 298.15 K and T n.b, and the main thermodynamic functions (changes of S, H, G) in the crystal and liquid states of the temperature interval studied and in the ideal gas state at T = 298.15 K. The experimental vapor pressures of the narrow temperature interval, ΔT = 62 K were extended to the entire range of the liquid, T crT tp0 = 394.3 K, from the triple, T tp0, to the critical, T cr, temperatures.  相似文献   

13.
14.
15.
Lithium-ion batteries(LIBs) have evolved into the mainstream power source of ene rgy sto rage equipment by reason of their advantages such as high energy density,high power,long cycle life and less pollution.With the expansion of their applications in deep-sea exploration,aerospace and military equipment,special working conditions have placed higher demands on the low-temperature performance of LIBs.However,at low temperatures,the severe polarization and inferior electrochemical activity of electrode materials cause the acute capacity fading upon cycling,which greatly hindered the further development of LIBs.In this review,we summarize the recent important progress of LIBs in low-temperature operations and introduce the key methods and the related action mechanisms for enhancing the capacity of the various cathode and anode materials.It aims to promote the development of high-performance electrode materials and broaden the application range of LIBs.  相似文献   

16.
In treating the experimental data on the heat capacity of solids, the essence of any model application is in the searching for the scaling factors (k i or 1/Θi) which transform a set of independent functions C P,i(T) for every substance into a function C P(T·k i) universal for the particular set of substances. DSC heat capacities of I–III–VI2 compounds at elevated temperatures exceed the upper limit of 12R (3R per mole of atoms) and make impossible application of any model. Nevertheless, the temperature scaling of heat capacity can be solved as a pure mathematical problem without any physical model (theory). The benefits of the model-free scaling are illustrated with the case of four isostructural chalcogenides (LiInS2, LiInSe2, LiGaS2, and LiGaSe2) measured recently with DSC in a temperature range from 180 to 460 K. The upper limit of C P(T·k i) functions was expanded up to 635 K. Low-temperature heat capacity of LiInSe2 published in 1995 made it possible to derive the thermodynamic functions (enthalpy and entropy) for LiInS2 (0–590 K), LiGaS2 (0–640 K), and LiGaSe2 (0–490 K) and expand those data for LiInSe2 from 300 to 460 K.  相似文献   

17.
Automatic computer programs are developed to calculate one- two-, and three-dimensional Debye functions. Prior tables of these functions are critically reviewed. Also, strategies are derived to calculate Debye temperatures from heat capacities. Both, simple three-dimensional Debye analyses and Tarasov analyses were carried out on 35 linear macromolecules. The experimental heat capacities for these analyses were collected in the ATHAS data bank. It is shown that the skeletal heat capacity of linear macromolecules is often best represented by only two vibrations per chain atom. For most of the all-carbon chain macromolecules the intramolecular skeletal heat capacity can be given by Cvs=D1[520 (28/MW)1/2] whereMW is the molecular mass andD 1 represents the one-dimensional Debye function. Polyoxides show a higher intramolecular theta temperature, but a lower intermolecular theta temperature. Double bonds and phenylene groups in the chain increase the intramolecular theta temperature.Dedicated to Prof. Dr. F. H. Müller.On leave from the Lumumba Peoples' Friendship University, Moscow, USSR.  相似文献   

18.
Using the semiempirical quantum-chemical AM1 method, the dependence of geometry and electronic structure on the torsion angle δ between neighbouring pyrrole rings of quaterpyrrole in various charge and spin states is investigated. The magnitudes of changes in the electronic structure of polypyrrole caused by geometrical changes of the chain may be comparable with those caused by charge and spin fluctuations. A limited amount of negative charge may be spontaneously stabilized on the polypyrrole chain due to chain deformations caused by solid state influences without any charged dopants. Charge fluctuations are most probably concentrated at nitrogen atoms. N—C bonds are less dependent on δ variations as well as on charge and spin perturbations than C—C bonds.  相似文献   

19.
Quantitative techniques for evaluating magnetic and calorimetric parameters of narrow-band and heavy fermion metals are discussed. The behavior of the electronic heat capacity at temperatures near a transition to magnetic order is emphasized. The discussion illustrates the methods used in establishing a previously published correlation between the electronic heat capacity and magnetic susceptibility of heavy Fermi liquids and their low temperature ground states.  相似文献   

20.
A simple and accurate group contribution method to estimate the heat capacity of ionic liquids is presented. The method considers groups previously defined for a successful method used to estimate critical properties of ionic liquids. Additionally a structural parameter known as mass connectivity index recently defined by the authors has been incorporated to define the model equation. To better define the values of the groups, heat capacity data at 298 K for 126 organic substances were used with the 469 heat capacity data for 32 ionic liquids. The results were compared with experimental data and with values reported by other available estimation methods. Results show that the new group contribution method gives low deviations and can be used with confidence in thermodynamic and engineering calculations.  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