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.
Up to 40 kbar and 1100°C, CaSi2 is dimorphic. Trigonal/rhombohedral CaSi2I (CaSi2-type structure) with corrugated layers of three-connected Si atoms can be transformed by a high pressure-high temperature treatment into tetragonal CaSi2II (α-ThSi2-type structure) with a three-dimensional net of three-connected Si atoms. The silicon net of CaSi2II is slightly distorted from the topologically simplest tetragonal three-dimensional three-connected net derived on a geometrical basis. In order to correlate crystal chemical with thermochemical data the transformation between both polymorphs of CaSi2 has been studied at equilibrium and nonequilibrium conditions. The pressure-temperature phase diagram of CaSi2 has been investigated by X-ray technique in quenched samples. From the slope of the equilibrium line and the change in molar volume the approximate values of the entropy and energy of transformation CaSi2(I-II) have been determined ΔS = 3.2 e.u., ΔU = 4.9 kcal/mole. Under nonequilibrium conditions the transformation CaSi2(II-I) yielded ΔH = ?4.2 kcal/mole at 500°C and ambient pressure in a DTA apparatus. Complete transformation of metastable CaSi2II can be achieved within 5 min at a heating rate of 20°C/min. Due to the relatively high speed of transformation simple structural relations between both polymorphs of CaSi2 are discussed.  相似文献   

2.
The glass–ceramic electrolytes of (100?x)(0.8Li2S·0.2P2S5xLiI (in mole percent; x?=?0, 2, 5, 10, 15, 20, and 30) were prepared by mechanical milling and subsequent heat treatment. Crystalline phases analogous to the thio-LISICON region II or III in the Li2S–GeS2–P2S5 system were precipitated. The thio-LISICON III analog phase was mainly precipitated at the composition x?=?0, and the thio-LISICON II analog phase was precipitated in the composition range from x?=?2 to 15. The X-ray diffraction peaks of the thio-LISICON II analog phase shifted to the lower diffraction angle side with increasing the LiI content. High conductivities above 2?×?10?3?S?cm?1 at room temperature were observed in the glass–ceramics at the wide composition range from x?=?2 to 15. The glass–ceramic electrolyte at x?=?5 with the highest conductivity of 2.7?×?10?3?S?cm?1 showed a wide electrochemical window of about 10 V. The addition of LiI to the 80Li2S·20P2S5 (in mole percent) glass was effective in crystallizing the thio-LISICON II analog phase with high conductivity from the glass.  相似文献   

3.
A GC-MS analysis of the azobisisobutyronitrile thermal decomposition products of in solutions at 80°C showed that the ratio of recombination and disproportionation rates of the cyanoisopropyl radical does not depend on the medium viscosity, but increases when the internal pressure of the solvent increases according to the log(k dispr/k rec) = ?1.25 + 0.096 P int 0.5 law. This means that the activation volume corresponding to recombination is larger than that corresponding to disproportionation. It follows from the relationship log(k dispr/k rec) = (ΔV rec ? Δv dispr P/RT that, for the decomposition of the substrate in benzene under a pressure of 0.5–4.0 kbar, the difference between the activation volumes is ΔV rec ? ΔV dispr = 8 cm3/mol.  相似文献   

4.
Phase relations in the system BaOGeO2 were investigated in the pressure range 20–70 kbar in the temperature range 750–1200°C. Several new phases were identified in this system: an atmospheric phase of BaGe2O5 (monoclinic BaGe2O5 I), two high-pressure phases of BaGe2O5 (monoclinic BaGe2O5 II and tetragonal BaGe2O5 III), and a high-pressure phase of Ba2Ge5O12. The phase boundary curve between BaGe2O5 II and BaGe2O5 III was preliminarily determined as P(kbar) = 7.7 + 0.047T (°C). The high-pressure phases of BaGeO3, which were previously reported by Y. Shimizu, Y. Syono, and S. Akimoto (High Temp.-High Pressures2, 113 (1970)) in the pressure range 15–95 kbar, were interpreted to be not single-phase materials but complicated mixtures of more than two phases in the system BaOGeO2. X-Ray powder diffraction data for the new compounds synthesized in this study are given.  相似文献   

5.
The anionic polymerization of three monomers, 2-isopropenyl-4,5-dimethyloxazole(I), 2-isopropenylthiazole(II), and 2-isopropenylpyridine(III), was studied in THF. These monomers produced red-colored living polymers on addition of sodium naphthalene or living α-methylstyrene tetramer as an initiator. It was observed that a considerable amount of monomer remained in the respective living polymer–monomer system, indicating that an equilibrium between the polymer and the monomer existed as in the case of α-methylstyrene. At lower temperatures, the conversion of the monomer to the polymer increased. The equilibrium monomer concentrations [Me] were determined at different temperatures, and the heats (ΔH) and the entropies (ΔS°) of polymerization were obtained by plotting In(1/[Me]) against 1/T as ΔH = ?9.4, ?6.8, and ?6.2 kcal/mole, ΔS°S = ?22.9, ?16.5, and ?16.6, eu for I, II, and III, respectively.  相似文献   

6.
Coulometric titrations using solid zirconia ionic conductors have been employed to determine the phase diagram of the ternary system CuGeO in the temperature range from 750 to 950°C. CuGeO3 was found to be the only existing ternary compound in the system. It is in equilibrium with Cu2O, CuO, GeO2, and oxygen of atmospheric pressure. Cu and Cu2O may coexist with GeO2. The standard Gibbs energy of formation of CuGeO3 was found to be ΔG°f (CuGeO3) = ?424.5 kJ/mole at 900°C. The standard enthalpy and entropy of formation are ΔH0f = ?756.8 kJ/mole and ΔS°f = ?283 J/mole·K, respectively.  相似文献   

7.
The phase diagram of Li2WO4, previously studied by Yamaoka et al. (J. Solid State Chem.6, 280 (1973)) has been revised. Li2WO4 II is stable at atmospheric pressure below ~310°C. This phase appears to be a modified spinel, and is tetragonal, a, c = 11.941, 8.409Å, Z = 16, space group I41amd. The melting curve of phenacite-type Li2WO4 I rises with pressure with a slope of 0.9°C/kbar to the III/I/liquid triple point at 3.1 kbar, 743°C, beyond which the melting curve of orthorhombic Li2WO4 III rises steeply with pressure (initial slope 31°C/kbar). The Li2WO4IIII transition line at 3 kbar is almost independent of temperature, i.e., the IIII transition entropy is zero. Li2WO4 II is 21.3% denser than Li2WO4 I at ambient conditions.  相似文献   

8.
Equilibrium positions between intramolecular OH ? N hydrogen bonded and free OH forms of some 3-piperidinols, decahydroisoquinolinols, a decahydroquinolinol, lupinine and N-methyl-3-piperidinemethanol have been determined from dilute solution IR spectral data at 33°. Conformational free energies of the H-bonds (ΔG°OH?N, attractive) have been calculated. The results suggest a linear relationship between the apparent value of ΔG°OH?N, as defined by the method of calculation, and the strength of the OH ? N bond expressed as Δν, within the limits of 0·5 ± 0·2kcal/mole per 100 cm?1, from Δν 90 to 350 cm?1. For cis-decahydroisoquinoline (N-Me or N-H) systems, a 0·4 kcal/mole difference has been calculated between the two possible ring-fused conformations, in favor of the so-called steroid form. For the corresponding cis-decahydroqumoline equilibrium, a 0·8 kcal/mole difference has been calculated, in favor of the nonsteroid form.  相似文献   

9.
The melting curve of NH4HF2 I rises from 125.2°C at atmospheric pressure to a triple point II/I/liquid at 9.3 kbar, 220°C. The I/II phase boundary is terminated at a triple point III/I/II at ∼45 kbar, 295°C. The melting curve of the new phase NH4HF2 II passes through a broad maximum at ∼39 kbar, 306°C, and is terminated at a triple point III/II/liquid at 46.3 kbar, 301°C. The melting curve of NH4HF2 III rises with pressure. The NH4HF2 III may be a dense hydrogen-bonded phase. Liquid NH4HF2 appears to be anomalous in several respects, and has a high compressibility relative to the solid phases.  相似文献   

10.
The polymerization of styrene in bulk at pressures up to 273 MPa and temperatures between 3 and 49°C with the use of γ-radiation as the initiator has been studied. The polymerization rate and the molecular weight of the polymer increased with increasing pressure; the molecular weight increased at a slightly faster rate. The difference in the rate is a theoretical expectation which has not previously been observed because chain-transfer reactions obscure the effect in chemically initiated systems. A small but significant retardation of the initiation reaction occurs as the pressure is increased. The results of previous workers are critically reviewed. Chain transfer at 25°C for pressures below 220 MPa is negligible when γ-radiation is the initiator. The activation energy for bulk polymerization decreased with increasing pressure from 28.1 kJ/mole at 0.1013 MPa to 22.3 kJ/mole at 203 MPa. Volumes of activation at 25°C for 0.1013 < p < 273 MPa were calculated to be Initiation, +4.0 < ΔV < +4.4 cm3/mole; polymerization; ?Δ = ?20.9 cm3/mole; degree of polymerization; ΔV = ?25.3 cm3/mole; propagation/termination; ?ΔV = ?22.7 cm3/mole.  相似文献   

11.
Pressure effects on the two‐site jumping of sodium and potassium cations in a 2,5‐di‐tert‐butyl‐1,4‐benzoquinone ion pair have been studied using a high‐pressure EPR technique. The rate constants of the intramolecular and intermolecular migrations for Na+ and K+ were determined from an EPR spectral simulation. The migration rates were found to be accelerated by increasing the external pressure. Using the pressure dependence of the migration rates, we estimated the activation volumes of the intramolecular (ΔV1?) and intermolecular (ΔV2?) processes for the Na+ and K+ migrations: ΔV1? = ?5.3 cm3 mol?1 and ΔV2? = ?29 cm3 mol?1 for Na+, and ΔV1? = ?8.3 cm3 mol?1 and ΔV2? = ?0.85 cm3 mol?1 for K+. Based on the results, the mechanisms for the two‐site jumping of Na+ and K+ are discussed in terms of volume. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 397–401, 2001  相似文献   

12.
Infrared spectra of 1,4-dibromobutyne-2 have been recorded over the 4000-200 cm?1 region in the vapour, liquid, amorphous and crystalline states Raman spectra were extended to ca. 20 cm?1 in the same states of aggregation, except for the non-recorded vapour phase spectrum. The temperature was varied between ?190 and 160 °C, and the pressure up to 10 kbar.A high proportion of the molecules exhibited free, internal rotation in the vapour and liquid phases, but to a smaller extent in the amorphous state at ?190 °C. For those molecules not being excited beyond the potential barrier, an unsymmetric conformation was preferred, whereas in the crystalline state the molecules possessed the anti conformation (C2h) both at low temperature and at high pressure at ambient temperature.A vibrational analysis based upon force field calculations was carried out and the mean amplitudes of vibration computed. The data have been related to preliminary results from dipole moment and electron diffraction investigations.  相似文献   

13.
From measurements of the degree of association and of the optical rotation of solutions of ethylmagnesium bromide in (+)(S)-1-ethoxy-2-methylbutane [(+)-(S)-L] and benzene it is concluded that these solutions contain the species EtMgBr·2L (I), (EtMgBr)2·3L (III) and (EtMgBr·L)2 (II) which have molecular rotations [M]25D of 3.79°, 5.92° and 4.26° respectively. Equilibrium constants of 103 mole/l and 1.41 mole/l at 25° were calculated for the equilibria between these species.  相似文献   

14.
Radiation-induced polymerization of methyl methacrylate (MMA) was studied up to 7500 kg/cm2 at 20°C. The rate of polymerization increased to 3000 kg/cm2 with overall activation volume ΔVpol? of -23.6 cm3/mole, and then the pressure dependence of the rate was very small in the pressure range between 3000 and 3700 kg/cm2. The rate of polymerization increased again above 3700 kg/cm2 up to the crystallization pressure of MMA (5500 kg/cm2) with ΔVpol? of -13.7 cm3/ mole with increasing pressure. The volume contraction by polymerization decreased with increasing pressure up to 3000 kg/cm2 but hardly decreased with increasing pressure above 3000 kg/cm2. The stereoregulzarity (triad probability) of PMMA changed slightly at 3000 kg/cm; above 3000 kg/cm2, syndiotactic addition decreased and heterotactic addition increased. Marked change in P-V isotherms of MMA, however, was not observed about 3000 kg/cm2. We concluded from these facts that an alignment of monomer molecules, which does not cause large volume change, was realized about 3000 kg/cm2. Polymerization proceeded above the crystallization pressure by long time irradiation, and isotactic addition increased clearly in the solid-state polymerization.  相似文献   

15.
Radicals giving the usual triplet ESR spectrum have been generated in polystyrene by γ-irradiation at room temperature. The decay of the radicals has been investigated in the temperature interval between 90° and 200° and pressures ranging from 1 to 8000 atm. The effect of pressure on the mechanism of the free-radical decay is discussed. There are two regions of free-radical decay showing different activation volumes: VI = 11.5 cm3/mole and VII = 66 cm3/mole. The correlation between molecular motion in the α-relaxation region and radical decay is pointed out.  相似文献   

16.
Heat capacities have been measured for single crystals of V2O3, either pure or doped with 1 and 1.4 mole% Cr2O3 and Al2O3 over the temperature range 100–700°K. V2O3 undergoes a fairly sharp transition at low temperatures (~170°K) but fails to exhibit any thermal anomaly above 300°K. The thermal behavior of (MxV1?x)2O3, M = Cr, Al, is manifested by two transitions: one at low temperatures, 170–180°K for x = 0.01 and 180–190°K for x = 0.014, and the other at high temperatures. For x = 0.01, the high-temperature (HT) anomaly extended over the range 325–345°K (Cr-doped V2O3) and 345–365°K (Al-doped V2O3), respectively. The corresponding ranges for x = 0.014 were found to be 260–280°K and 270–290°K, respectively. Further, the HT anomaly was characterized by a large hysteresis (~50°K). The values of lattice heat capacity of pure and doped V2O3 were, however, found to be almost the same and could be empirically represented by the Debye (D)?Einstein (E) function D(580T) + 4E(θT) with θ values 430°K (T = 100–230°K) and 465°K (T > 230°K), respectively. Further, the enthalpy change ΔH associated with the HT anomaly in doped V2O3 (80 ≤ ΔH ≤ 510 J/mole) was 5–10 times smaller than the ΔH corresponding to the lower-temperature transition. The results cited here appear incompatible with the Mott transition model that has been invoked to explain the HT anomaly.  相似文献   

17.
Dimorphic SrSi2 is the first compound for which the two simplest three-dimensional three-connected nets are found in its polymorphs. The cubic net of three-connected silicon atoms (SrSi2 type of structure) can be transformed into the tetragonal one (α-ThSi2 type of structure) by a high-pressure-high-temperature treatment. The tetragonal phase is quenchable. Heating of this phase to 600–700°C at ambient pressure results in transformation into the cubic one. At a heating rate of 20°C/min complete transformation can be achieved within 5 min in a DTA apparatus. The energy of transformation has been obtained from the peak areas of the DTA curves to ?1.6 ± 0.3 kcal/mole. Although the transformation between the three-dimensional three-connected sets in SrSi2 must be formally classified as a reconstructive one, a relatively small entropy change (ΔS = 1 ·1 cal/deg · mole) has been calculated from the change in molar volume and p-T equilibrium conditions. Therefore, structural relations between the cubic and the tetragonal nets are discussed.  相似文献   

18.
The values of ΔG(O2), ΔH(O2), and ΔS(O2) have been determined from electrochemical cell measurements, within the whole homogeneity range of WO3?x, between 700 and 900°C. The samples have been previously prepared by equilibration of WO3 pellets with COCO2 mixtures and their composition has been determined by thermogravimetry. A single phase has been found between WO3 and WO2.9760. The results may be understood by considering a structure involving point defects, singly ionized oxygen vacancies V·O between WO3 and WO2.9880. For larger departure from stoichiometry, the variations of ΔH(O2) and ΔS(O2) suggest the formation of more complex defects. The enthalpy of formation of V·O has been calculated: 78 kcal · mole?1.  相似文献   

19.
The possibility of ?-caprolactam (CPL) to coordinate to manganese(II), cobalt(II), and nickel(II) rhodanides has been investigated. New complexes trans-[M(CPL)4(NCS)2], where M = Mn (I), Co (II), and Ni (III), have been synthesized. The complexes have been studied by chemical analysis and IR spectroscopy. According to X-ray diffraction, complexes are isostructural to each other and crystallize in monoclinic space group P21/c, Z = 2. For I: a = 6.9457(2) ?, b = 17.7751(6) 0A, c = 12.8999(4) 0A, ?? = 104.2670(10)°, V = 1543.51(8) ?3, ??calc = 1.342 g/cm3, R 1 = 0.0426. For II: a = 6.8925(2) ?, b = 17.8189(8) ?, c = 12.7278(6) ?, ?? = 104.421(2)°, V = 1513.93(11) ?3, ??calc = 1.377 g/cm3, R 1 = 0.0280. For III: a = 6.7804(2) ?, b = 18.4631(4) ?, c = 12.4841(3) ?, ?? = 105.2950(10)°, V = 1507.49(7) ?3, ??calc = 1.382 g/cm3, R 1 = 0.0273. Structures I?CIII are molecular; the metal atom in each of them coordinates four CPL molecules and two NCS groups via oxygen and nitrogen atoms, respectively.  相似文献   

20.
We correlate here the specific heat Cp with the frequency shifts (1/V) (?V/?T)p and the thermal expansivity αp with the (1/ν) (?V/?P)T close to the I–II transition in NH4Br. This correlation is performed for the Raman mode of νs (140 cm?1) using the molar volume data for NH4Br. It is shown here that spectroscopic modifications of the Pippard relations are applied satisfactorily to the I–II phase transition by using a lattice mode of NH4Br.  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