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.
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   

2.
After hydrothermal and thermovaporous treatment of chemically pure amorphous aqueous silicic acid in solutions of NaOH and NH4OH and in water vapour it is possible, using complex thermal analysis, to detect the weight loss and heat effects corresponding to evaporation of various forms of combined water, and to estimate the heats of evaporation of these forms. From the obtained data, the following water forms have been identified: (1) at 200–300° capillary-condensed water formations of the cluster type evaporate;ΔH deh is about 8 kcal/mole H2O; (2) at 250–400°, molecules of water linked by hydrogen bonds with hydroxyl groups on the surface and in the volume of the particles;ΔH deh. is about 5 kcal/mole H2O; (3) at 350 600°, molecules of water coordinated to silicon atoms in the volume of the particles;ΔH deh is approximately 1 kcal/mole H2O. The total evaporation heat changes from 10 kcal/mole H2O when water of form 1 predominates, to 5 kcal/mole H2O when forms 2 and 3 predominate.  相似文献   

3.
Solid-solid transitions in trimorphic BaSi2 have been investigated up to 40 kbar and 1000°C by X-ray powder technique in quenched samples. All transformations between orthorhombic BaSi2I with isolated Si-tetrahedra, trigonal BaSi2II with corrugated Si-layers, and cubic BaSi2III with a three-dimensional three-connected Si-net can be performed in approximately 5 min at high-pressure-high-temperature conditions. At ambient conditions the difference in molar volume between BaSI2I and BaSi2III is relatively large (ΔVI–III = ?6.79 cm3/mole) and that between BaSi2III and BaSi2II very small (ΔVIII-II = ?0.05 cm3/mole). Consequently in the pressure-temperature phase diagram the boundary (I–III) shows a strong pressure dependence contrary to that of (III-II) which is less dependent on variation of pressure. The triple point between the three solid phases is near 11 kbar and 925°C. Substitution of divalent metal and quadrivalent metalloid can easily influence the phase relations in BaSi2.  相似文献   

4.
Heterogeneous decompositions of MgSO4 · 7H2O (Epsomite) monocrystals were studied with thermal (DTA, DSC, TG) and thermo-optical methods. The polythermal reaction is controlled by nucleation of the reactant. This process has been considered by the Avrami-Erofe'ev equation: $$kt = [ - \ln (1 - \alpha )]^{{\raise0.7ex\hbox{$1$} \!\mathord{\left/ {\vphantom {1 3}}\right.\kern-\nulldelimiterspace}\!\lower0.7ex\hbox{$3$}}} $$ The plots and the slope which give the activation energyE+=23.5 kcal/mole (760 Torr N2, 50–100°), are obtained from the Freeman-Carroll equation. The DSC technique was used to determine the heat of decomposition (ΔH=42.3 kcal/mole, 760 Torr N2, 50–100°). The heat of transformation for the reaction 39–47° $$MgSO_4 \cdot 7H_2 O\xrightarrow{{39 - 47^ \circ }}MgSO_4 \cdot 6H_2 O + H_2 O$$ wasΔH=2.8 kcal/mole. The isothermal reaction (20°, 10?6 Torr) is controlled by first-order kinetic.  相似文献   

5.
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   

6.
The dissociation energies of the gaseous molecules CuGe, AgGe, AuGe, Ge2, and Cu2 have been determined by mass spectrometric investigations of the vapour phases above the liquid alloys Ge?Cu, Ge?Ag, and Ge?Au. The evaluation according to the third-law method leads to the following values for the dissociation energies:D 0°(CuGe)=49,0±5 kcal/mole;D 0°(AgGe)=40,8±5 kcal/mole;D 0°(AuGe)=65,3±3,5 kcal/Mol;D 0°(Ge2)=64,5±5 kcal/Mol;D 0°(Cu2)=64,5±5 kcal/Mol.  相似文献   

7.
The crystal structure of [Si(CH3)(t-C4H9)]4 has been determined by single crystal X-ray diffraction. The crystals are tetragonal, P42/n; a = b = 13.069(4), c = 7.880(2) Å, Z = 2. The structure was determined using 745 independent data and refined with anisotropic least-squares to a final unweighted R-value of 3.5%. Each tetrameric molecule was found to be arranged about a 4 axis, with the independent crystallographic unit comprising one silicon atom, one methyl and one tert-butyl group. The four-membered ring of silicon atoms is nonplanar with an unusually large dihedral angle of 36.8°. The principal mean bond lengths are SiSi 2.377(1), SiC(methyl) 1.893(4), SiC(tert-butyl) 1.918(3) Å, and the SiSiSi bond angle is 86.99°. The SiSi bond length is somewhat longer than in other polysilanes.  相似文献   

8.
The complex formation of Ytterbium (III) with Bromopyrogallol red has been studied. spectrophotometrically in an attempt to establish composition, stability, thermodynamic parameters and optimum conditions for determining small amounts of ytterbium. The violet complex of ytterbium has λmax at 620nm against a reagent blank. The composition determined by different methods is 1:1 at pH 6.2±0.1. The mean value of log K, free energy (ΔG), the heat content (ΔH) and entropy (ΔS) changes, of the complex are found to be 6.0, —8.1.kcal/mole, —3.5 Kcal/mole and 15.53 e.U. respectively at 30°C. The net molar absorptivity and Sandell's sensitivity is 19850 and 0.0087 μg of ytterbium /cm2. The effect of diverse ions was examined with thirteen cations and ten anions in the determination of ytterbium.  相似文献   

9.
The superstructure of cubic SiP2O7 has a volume 27 times as large as the substructure of this compound. Because conventional methods failed in solving the superstructure, computer simulation of the structure was applied. Computer simulation depends on the accurate prediction of individual interatomic distances in a structure and on an appropriate weighting scheme. The predicted distances are treated as observations in a distance least-squares refinement, in which the positional coordinates of the atoms are varied until the calculated distances conform to the predicted values. Cubic SiP2O7 (a = 22.418Å, space group Pa3, Z = 108, Dx = 3.22) has 50 atoms in the asymmetric unit. The initial R-value for the simulated structure was 0.18, which dropped to 0.061 after three cycles of least squares refinement using Fobsd = 1382. In the substructure, all POP angles in the diphosphate groups are straight because of symmetry requirements, and the angles POSi are 164°. In the superstructure, the average angle POP is 150°; the average POSi angle is 149°. The tendency to decrease these angles may be responsible for the formation of the superstructure. However, two diphosphate groups retain, even in the superstructure, the 180° configuration. Such a feature is usually only observed in high temperature polymorphs and is explainable as a positional disorder of bent P2O7 groups, or by the assumption of a highly anisotropic motion of the bridging oxygen atom. The silicon atoms in cubic SiP2O7 are octahedrally coordinated, as they are in the two monoclinic polymorphs of SiP2O7. However, the three modifications are topologically distinct from each other as can be proved by considering the three-dimensional nets on which the three structures are based.  相似文献   

10.
Near Hartree-Fock level ab initio molecular orbital calculations on H3O+ and a minimum energy structure with θ(HOH) = 112.5° and r(OH) = 0.963 Å and an inversion barrier of 1.9 kcal/mole. By comparing these results to calculations on NH3 and H2O, where precise experimental geometries are known, we estimate the “true” geometry of isolated H3O+ to have a structure with θ(HOH) = 110-112°, r(OH) = 0.97–0.98 Å and an inversion barrier of 2–3 kcal/mole. Our prediction for the proton affinity of water is ≈ 170 kcal/mole, which is somewhat smaller than the currently accepted value.  相似文献   

11.
The rate of the gas phase reaction has been measured spectrophotometrically over the range 480°–550°K. The rate constant fits the equation where θ = 2.303RT in kcal/mole. This result, together with the assumption that the activation energy for the back reaction is 0 ± 1 kcal/mole, allows calculation of DH (Δ? CH2? H) = 97.4 ± 1.6 kcal/mole and ΔH (Δ? CH2·) = 51.1 ± 1.6 kcal/mole. These values correspond to a stabilization energy of 0.4 ± 1.6 kcal/mole in the cyclopropylcarbinyl radical.  相似文献   

12.
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   

13.
A study of the dissociation pressure of crystalline K2CoCl4·2H2O. The reactions can be summed up as K2CoCl4·nH2O(c) = K2CoCl4·mH2O(c)+(nm)H2O (v). Below 50°C, n = 2 and 1, m = 1 and 0, above 50°C, n = 2 and m = 0. Below 50°C, the dihydrate is octahedral, the monohydrate and anhydrous compounds are tetrahedral. ΔH° and ΔS° are respectively 7.0 kcal and 14.2 e.u. for the loss of the first mole of water and 12.7 kcal and 32.0 e.u. for the loss of the last mole of water. Above 50°C, ΔH° and ΔS° are respectively 29.8 kcal and 77.6 e.u. for the loss of both waters. The changes in structure are discussed using the spectral and magnetic properties as indications for structural changes.  相似文献   

14.
Equilibria among the cyclic compounds (Me2Si)n where n = 5, 6 and 7 have been studied between 30–58°C. Thermodynamic values for the redistribution reactions between pairs of compounds are, for n = 5 → 6, ΔH = ?18 kcal/mole, ΔS = ?20 cal/deg. mole; for n = 7 → 6, ΔH ?3, ΔS +33; for n = 7 → 5, ΔH +18, ΔS + 51. The enthalpies indicate that the stabilities of the rings increase in the order (Me2Si)5 < (Me2Si)7 < (Me2Si)6. The differences are smaller than corresponding differences among the cycloalkanes, probably because the silicon compounds are less affected by steric repulsions and angle strain.  相似文献   

15.
M. Branik  H. Kessler 《Tetrahedron》1974,30(6):781-786
Z, E-Isomerism of the urethane bond of t-BOC-glycine was observed by 1H- and 13C-NMR spectroscopy at various temperatures in several solvents. The special stabilization of the Z isomer at low temperatures in CDCl3 has been explained by intra- and intermolecular H-bond forming a 7-membered ring. Thermodynamic data have been determined for the ground state (AH°= ?7 kcal/mol, Δ° = ?25 Clausius) as well asfor the barrier of interconversion (ΔG° = 15·4 kcal/mole for the deuterated title compound) in CDCl3. The equilibrium between the Z and the E conformation is shifted towards the E conformation in more polar solvents (acetonitril-d3, acetone-d6, DMSO-d6), in which cyclization of the Z conformation is not important.  相似文献   

16.
Copper(II) hypophosphite has been shown to exist as several polymorphs. The crystal structures of monoclinic α‐, ortho­rhombic β‐ and ortho­rhombic γ‐Cu(H2PO2)2 have been determined at different temperatures. The geometry of the hypophosphite anion in all three polymorphs is very close to the idealized one, with point symmetry mm2. Despite having different space groups, the structures of the α‐ and β‐polymorphs are very similar. The polymeric layers formed by the Cu atoms and the hypophosphite ions, which are identical in the α‐ and β‐polymorphs, stack in the third dimension in different ways. Each hypophosphite anion is coordinated to three Cu atoms. On cooling, a minimum amount of contraction was observed in the direction normal to the layers. The structure of the polymeric layers in the γ‐­polymorph is quite different. There are two symmetry‐independent hypophosphite anions; the first is coordinated to two Cu atoms, while the second is coordinated to four Cu atoms. In all three polymorphs, the Cu atoms are coordinated by six O atoms of six hypophosphite anions, forming tetragonal bipyramids; in the α‐ and β‐polymorphs, there are four short and two long Cu—O distances, while in the γ‐polymorph, there are four long and two short Cu—O distances.  相似文献   

17.
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   

18.
The saturation vapour pressures of WOBr4 and WO2Br2 and their reaction equilibria have been determined by means of a membrane zero manometer and ampoule quenching experiments, respectively. From the pressuretemperature dependence the following sublimation data were estimated: Δ H° (subl., WOBr4, 298) = 29.4 (± 1.0) kcal/mole; Δ H° (subl., WO2Br2, 298) = 36.6 (±1.5) kcal/mole; Δ S° (subl., WOBr4, 298) = 50.1 (± 1) cl; Δ S° (subl. WO2Br2, 298) = 53.0 (±1.5) cl. For the decomposition reaction of solid WO2Br2 were obtained: Δ H° (s, 690) 37.5 (± 0.7) kcal/mole, Δ S° (s, 690) = 49.0 (± 0.5) cl; and for the decomposition of gaseous WO2Br2: Δ H° (g, 690) = ?29.6 (± 2.0) kcal/mole, Δ S°. (g, 690) = ?44.5 (± 1.5) cl.  相似文献   

19.
The I2-catalyzed isomerization of allyl chloride to cis- and trans- l-chloro-l-propene was measured in a static system in the temperature range 225–329°C. Propylene was found as a side product, mainly at the lower temperatures. The rate constant for an abstraction of a hydrogen atom from allyl chloride by an iodine atom was found to obey the equation log [k,/M?1 sec?1] = (10.5 ± 0.2) ?; (18.3 ± 10.4)/θ, where θ is 2.303RT in kcal/mole. Using this activation energy together with 1 ± 1 kcal/mole for the activation energy for the reaction of HI with alkyl radicals gives DH0 (CH2CHCHCl? H) = 88.6 ± 1.1 kcal/mole, and 7.4 ± 1.5 kcal/mole as the stabilization energy (SE) of the chloroallyl radical. Using the results of Abell and Adolf on allyl fluoride and allyl bromide, we conclude DH0 (CH2CHCHF? H) = 88.6 ± 1.1 and DH0 (CH2CHCHBr? H) = 89.4 ± 1.1 kcal/ mole; the SE of the corresponding radicals are 7.4 ± 2.2 and 7.8 ± 1.5 kcal/mole. The bond dissociation energies of the C? H bonds in the allyl halides are similar to that of propene, while the SE values are about 2 kcal/mole less than in the allyl radical, resulting perhaps more from the stabilization of alkyl radicals by α-halogen atoms than from differences in the unsaturated systems.  相似文献   

20.
The first and second bond dissociation energies for H2O have been calculated in anab initio manner using a multistructure valence-bond scheme. The basis set consisted of a minimal number of non-orthogonal atomic orbitals expressed in terms of gaussian-lobe functions. The valence-bond structures considered properly described the change in the molecular system as the hydrogen atoms were individually removed to infinity. The calculated equilibrium geometry for the H2O molecule has an O-H bond length of 1.83 Bohrs and an HOH bond angle of 106.5°. With 49 valence-bond structures the energy of H2O at this geometry was ?76.0202 Hartrees. The calculated equilibrium bond length for the OH radical was 1.86 Bohrs and the energy, using the same basis set, was ?75.3875 Hartrees. After correction for zero point energies the calculated bond dissociation energies are: H2O → OH + H, D1=75.38 kcal/mole and OH → O+H, D2=54.79 kcal/mole.  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