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.
《Thermochimica Acta》1986,109(1):105-109
The heat capacity of gold has been measured by laserflash calorimetry in the temperature range 80–1000 K. The results are compared with available low- and high-temperature heat capacities, and revised thermodynamic values of gold are given.  相似文献   

2.
Heat capacities Cp of a polydiacetylene-bis(toluene sulfonate) single crystal and its monomer have been measured in the temperature range from 3 to 300 K. The temperature dependence of Cp for both monomer and polymer crystals differs from that for monoatomic solids. By applying a chain lattice model for a polymer crystal, the temperature dependence of the heat capacity can be described assuming a phonon density of states given by bending and stretching modes of the polymer backbone. With a combination of one-dimensional and three-dimensional elastic continuum approximations, the heat capacity has been calculated and a good fit to the data has been obtained. A small peak in Cp was detected at 161 K for the monomer and at 198 K for the polymer. This may be ascribed to a lower-temperature phase transition in the polydiacetylene crystals evidenced by previous x-ray and spectroscopic measurements.  相似文献   

3.
The heat capacity of orthorhombic (marcasite-type structure) cobalt ditelluride has been measured from 5 to 1 030 K by adiabatic-shield calorimetry with alternate energy inputs and equilibrations. Above 900 K a marked increase in heat capacity occurs which probably signals a change in the composition of the CoTe2-phase towards higher tellurium content. Values at 298.15 and 1 000 K in J K–1 mol–1 of the heat capacity (C p,m), entropy [S m ° (T)S m ° (0)], andGibbs energy function – [G m ° (T)H m ° (0)]T –1 are 75.23, 114.5, 49.93, and 132.4, 216.2, 139.17, respectively. Consistent with the metallic behavior of CoTe2, deviation of the heat capacity from theDebye T 3-law was found at low temperatures. Comparison with the heat capacity of FeTe2 shows aSchottky-like deviation with a maximum of 7.3 J K–1 mol–1 at 80 K and evidences the influence of the additional 3 d-electron in cobalt compared to iron. Heat capacity measurements were made on CoTe2.33 to ascertain the existence range of the CoTe2+x -phase and the entropy of the associated structural disorder.The portion of this research done at Ann Arbor was supported in part by the Structural Chemistry and Chemical Thermodynamics Program of the Division of Chemistry of the National Science Foundation under Grant No. CHE-7710049.  相似文献   

4.
Thermodynamic properties of sodium borosilicate glasses {56.7 SiO2, (43.7   x)B2O3,xNa2O} wherex =  14.4, 22.9, and 32.5, have been studied. The heat capacity was measured using an adiabatic calorimeter at temperatures between 13 K and 300 K. The thermodynamic functions were calculated from the smoothed values ofCp, m . The results differ from an additive model with pure glassy SiO2, B2O3, and crystalline Na2O as components. A model based on the assumption that the contribution of structural units of glasses to the heat capacity is equal to those of glasses with the same molecular formula is proposed.  相似文献   

5.
Thermodynamic properties of β-alanine in the temperature range 6.3–301 K were studied. No phase transitions were observed for the sample specially prepared to contain no solvent inclusions. At 298.15 K the calorimetric entropy and the difference in the enthalpy values are equal, respectively, to 126.6 JK−1 mol−1 and 19.220 Jmol−1. The C p (T) in the temperature range 6–16 K can be well described by Debye equation C p  = AT 3. A comparison of the data on the entropies of glycine polymorphs and of β-alanine was used to show, that the empirical Parks–Huffman rule holds in the case of these compounds.  相似文献   

6.
Heat capacities of three cubic lithium tungsten bronze samples (LixWO3) with x values of 0.363, 0.437, and 0.478 were measured from 200 to 800°K. Heat capacities per gram-atom at the same temperature of Li0.363WO3 and Li0.437WO3 were equal within experimental error and also equal to those of Na0.485WO3, Na0.698WO3, and Na0.794WO3, regardless of the difference of the composition. λ-type heat capacity anomalies were observed around 330, 460, and 590°K in Li0.363WO3 and around 330 and 460°K in Li0.437WO3 and Li0.478WO3, showing the existence of second-order phase transitions. The entropy increments of the transitions were obtained as 1.36, 0.45, and 1.68 J mole?1 K?1 for Li0.363WO3, 1.09 and 0.59 J mole?1 K?1 for Li0.437WO3, and 1.42 and 0.50 J mole?1 K?1 for Li0.478WO3.  相似文献   

7.
Measurements of the heat capacity by quasi-adiabatic, intermittent energy increments from 5 to 350 K show small high heat-capacity anomalies near 7 and 10 K which are attributed to superconducting transitions seen by magnetic measurements on the same carefully synthesized and well-characterized sample of (Hf0.934Zr0.057)(N0.97). Although no previous heat capacity measurements over the cryogenic region are known, the estimated 298.15 K standard entropy values (S/R) vary in the literature from about 200 per cent higher to 5 per cent lower than our measured value of (5.28±0.01)R–1 when the formula is represented as above. A simple scheme to represent and predict values based on both molar volumes and atomic masses for related materials is presented which seems more reliable on a limited sample than do others despite the intrusion of lanthanide contraction.This revised version was published online in November 2005 with corrections to the Cover Date.  相似文献   

8.
The heat capacity at constant pressure C p (T) of terbium diboride synthesized from elements via an intermediate hydride phase was studied experimentally within 5–300 K. A ferromagnetic phase transition manifests itself in the C p (T) dependence as a sharp maximum at 142.4 ± 0.1 K. The C p (T) dependence was used to calculate the tempreature dependences of the enthalpy, entropy, and the Gibbs energy and to determine the parameters of the electronic, lattice, and magnetic contributions to the heat capacity of TbB2.  相似文献   

9.
Heat capacity at constant pressure C p (T) of a dysprosium boride DyB62 single crystal obtained by zone melting was studied experimentally in the temperature range of 2 to 300 K. Abnormally high values of dysprosium boride heat capacity were revealed in the range of 2–20 K, due to the magnetic contribution and the effect of disorder in the boride lattice. Temperature changes in DyB62 enthalpy, entropy, Gibbs energy, and standard values of these thermodynamic functions were calculated.  相似文献   

10.
11.
The heat capacity of polyhexene-1 was measured between 20 and 300°K. The apparatus, an adiabatic calorimeter giving results with a random error of 0.2–0.4%, is briefly described. The characterization of the sample by x-ray diffraction patterns established that it was amorphous at all temperatures. Gold foil was incorporated with the sample to increase the apparent thermal diffusivity and so to decrease the time needed for the measurements. The glass transition temperature was found to be 215.5 ± 1°K. On the Cp curve, no subglass anomaly was detected, unlike the results of experiments described elsewhere. The calculation of Cv is discussed, and an explanation is given for the choice of the number of intramolecular vibrational modes per monomer which are assumed to contribute to Cv. A linear continuum model with characteristic temperature θ1 = 736°K allows us to fit the experimental curve over a temperature range of 140°K.  相似文献   

12.
Heat capacity of α-glycylglycine was measured using adiabatic calorimetry (6 to 304 K) and DSC (264 to 443 K), and then thermodynamic functions were calculated. Heat capacity has no anomalies. The molecular crystal melts at 493 K (enthalpy of melting is about 62 kJ mol–1). The melting is accompanied by decomposition. C P(T) function for glycylglycine is very similar to those of three glycine polymorphs. The ‘universal’ curve consists of two parts: low-temperature Debye-like function (from 0 to about 120 K) and a straight line (up to the melting point). At very low temperatures rigid molecules oscillate as a whole, and the Debye temperature is proportional to the molecular mass to the power of 3/2.  相似文献   

13.
The radicals formed in poly(methyl methacrylate) (PMMA) under vacuum by UV irradiation at room temperature were carefully examined from 77 K to 300 K by electron spin resonance (ESR). The conventional nine-line spectrum was observed with significant overall intensity changes in contrast to previous reports. The intensity decreases greatly as the temperature increases from 77 K to 100 K. The intensity of the ESR spectrum increases as the temperature increases gradually from 100 K to 260 K. The spectral changes were reversible at all temperatures. Three different models are considered to interpret the temperature dependence of the intensity of the ESR spectrum. The results indicate that the ESR spectrum depends on (1) the steady-state concentration of the propagating radical in the polymer, (2) the conformational distributions of the radicals, and (3) the environmental structures of the polymer matrix.  相似文献   

14.
Heat capacities of U1–yLayO2 were measured by means of direct heating pulse calorimetry in the temperature range from 300 to 1500 K. An anomalous increase in the heat capacity curve of each sample was observed similarly to the case of U1–yGdyO2, found recently in our laboratory. As the lanthanum content of U1–yLayO2 increased, the onset temperature of an anomalous increase in the heat capacity decreased and the excess heat capacity increased. The enthalpy of activation (Hf) and the entropy of activation (Sf) of the thermally excited process, which cause the excess heat capacity were obtained to be 2.14, 1.63 and 1.50 eV and 39.4, 34.2 and 31.8 J·K–1·mol–1 for U0.956La0.044O2, U0.910La0.090O2 and U0.858La0.142O2, respectively. The values at zero La content extrapolated by using the data of Hf and Sf for U1–yLayO2 were in good agreement with the experimental values of undoped UO2 so far reported, similarly to the case of Gddoped UO2. The electrical conductivities of U1–yLayO2 (y=0.044 and 0.142) were also measured as a function temperature. No anomaly was seen in the electrical conductivity curve. It may be concluded that the excess heat capacity originates from the predominant contribution of the formation of oxygen clusters and from the small contribution of the formation of electron-hole pairs.  相似文献   

15.
The heat capacity of indium phosphide was measured over the temperature range 360–760 K using a DSM-2M differential scanning calorimeter. The thermodynamic functions of InP were calculated. Original Russian Text ? A.S. Pashinkin, A.S. Malkova, M.S. Mikhailova, 2009, published in Zhurnal Fizicheskoi Khimii, 2009, Vol. 83, No. 6, pp. 1191–1192.  相似文献   

16.
The heat capacities of NbCl5 and Nb3Cl8 samples with less than 1·10–3 mass-% of impurities were determined over the range 6–320 K by an adiabatic calorimeter. An anomaly was found in Nb3Cl8 within the 7–13 K range. Comparisons ofC p values of Nb3Cl8 are made with the theoretical works of Tarassov. The qualitative fit is quite good.
Zusammenfassung Die Wärmekapazitäten von NbCl5- und Nb3Cl8-Proben mit Verunreinigungsgehalten 10–3 Masse-% wurden mit einem adiabatischen Kalorimeter bei 6 bis 320 K bestimmt. Am Nb3Cl8 wurde eine Anomalie bei 7 bis 13 K nachgewiesen. Die Cp-Werte von Nb3Cl8 werden mit theorischen Ergebnissen von Tarassov verglichen, die qualitative Übereinstimmung ist gut.

6–320 NbCl5 Nb3Cl8 , 1· 10–3 . Nb3Cl3 7–13 . p Nb3Cl8 , .
  相似文献   

17.
18.
The heat capacities of cesium and rubidium molybdates, Cs2MoO4 and Rb2MoO4, have been measured by differential scanning calorimetry (DSC) in the temperature range 300–800 K. These values have been combined with published low-temperature heat capacity data for Cs2MoO4 to obtain thermodynamic functions to 800 K. For Rb2MoO4, however, these functions could not be calculated because low-temperature heat capacities are unavailable. Instead, only heat capacity data are reported.  相似文献   

19.
The heat capacity of MnAs0.88P0.12 has been measured by adiabatic shield calorimetry from 10 to 500 K. It is shown that very small energy changes are connected with two magnetic order-order transitions, indicating that these can be regarded as mainly “noncoupled” magnetic transitions. At higher temperatures contributions to the excess heat capacity arises from a magnetic order-disorder transition, a conversion from low- to high-spin state for manganese, and a MnP- to NiAs-type structural transition. The observed heat capacity is resolved into contributions from the different physical phenomena, and the character of the transitions is discussed. In particular it is substantiated that the dilational contribution, which includes magnetoelastic and magnetovolume terms as well as normal anharmonicity terms, plays a major role in MnAs0.88P0.12. The entropy of the magnetic order-disorder transition is smaller than should be expected from a complete randomization of the spins, assuming a purely magnetic transition. Thermodynamic functions have been evaluated and the respective values of Cp, {SOm(T) - SOm(0)}, and -{GOm(T) - HOm(0)}/T at 298.15 K are 68.74, 72.09, and 32.30 J K−1 mole−1, and at 500 K 56.05, 108.12, and 56.64 J K−1 mole−1.  相似文献   

20.
The heat capacity of polycrystalline germanium disulfide α-GeS2 has been measured by relaxation calorimetry, adiabatic calorimetry, DSC and heat flux calorimetry from T = (2 to 1240) K. Values of the molar heat capacity, standard molar entropy and standard molar enthalpy are 66.191 J · K?1 · mol?1, 87.935 J · K?1 · mol?1 and 12.642 kJ · mol?1. The temperature of fusion and its enthalpy change are 1116 K and 23 kJ · mol?1, respectively. The thermodynamic functions of α-GeS2 were calculated over the range (0 ? T/K ? 1250).  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