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 187 毫秒
1.
Depolarization ratios ρ of the Raman bands due to CH3 stretching at 2907 cm?1 and the Si? O skeletal mode at 491 cm?1 have been measured in polydimethylsiloxane gum as a function of temperature from 100°C to ?45°C. Below 0°C the changes in p have been interpreted in terms of the formation of helical regions in the gum. The enthalpy of helix formation ΔH has been determined as 3200 ± 600 cal/mole. An upper limit on the entropy change, ΔS, of 16 ± 3 e.u./mole and minimum values of helix content at different temperatures have been found. The Raman spectrum of crystalline polydimethylsiloxane is presented.  相似文献   

2.
Energy-deformation characteristics for the primary T, S, and U conformational units of tie molecules were obtained from the analysis of data generated from a constrained minimization algorithm. Energy-deformation profiles (covering the range from compact equilibrium defect structures to the fully extended chain) are reported for the S0 and S1 members of the Sλ family and for the U00 member of the Umn family. Estimates of the energy content V0 and the elastic modulus E were obtained from the computed energy-deformation data in the vicinity of the equilibrium Structure—S0 → {60°, 180°, ?60°}, V = 1.7 kcal/mole, E = 60 kcal/cm3 [250 × 1010 dyn/cm2];S1 → {60°, 180°, 180°, 180°, ?60°}: V = 1.7 kcal/mole, E = 25 kcal/cm3 [100 × 1010 dyn/cm2]; and U00 → {60°, 180°, 60°, 180°, 60°}: V = 2.7 kcal/mole, E = 80 kcal/cm3 [340 × 1010 dyn/cm2]. Although the elastic modulus of the U00 unit is comparable to the elastic modulus of the fully extended chain, the highenergy content of this unit (V0 = 2.7 Kcal/mole) prohibits a significant population and thereby mitigates an appreciable reinforcing effect from this rigid unit. A model for a surrogate force constant is introduced to generalize the results from this study to any member of the Sλ or Umn family as well as any combination of Sλ and Umn units. This generalization provides a basis for estimating the deformation characteristics of tie molecules comprised of various populations of these primary conformational building blocks.  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.
Pyrolytic decay of carbon diselenide was monitored by ultraviolet absorption spectroscopy in reflected shock waves in the temperature range of 1600–2600°K. The temperature dependence of the absorption coefficient of CSe2 at 2308 Å was determined and was used to provide kinetic information along with a deconvolution procedure which accounted for and removed systematic distortions of the fast time-resolved absorbance profile. For temperatures of 1600–2600°K and argon densities of 1.5–7.0 × 10?5 mol/cm3 dilute (1.0–9.0 × 10?9 mol/cm3) CSe2 pyrolyzed with measured first-order decay rates in the range of log10 k1 (sec?1) = 3.0?5.7; at midrange (2100°K and 4.3 × 10?5 mol/cm3 in Ar) k1 ≈ 3 × 104 sec?1. The decay probably occurs via a unimolecular low-pressure process, first order in both CSe2 and Ar, for which k2 ± 109 cm3/mol·sec at 2100°K. The deconvoluted data yield Arrhenius activation energies of 53.2 kcal/mol under second-order treatment, but the activation energy is less reliable than the general magnitude of the rate constant. A comparison of CSe2 with other molecules which are isoelectronic in their valence shells (CO2, CS2, OCS, and N2O) is made.  相似文献   

5.
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   

6.
Rate parameters for the reaction of ground-state atomic sulfur, S(3P), with the olefins cis-2-butene and tetramethylethylene have been determined over a temperature range of ∽280°K. A major finding of this study was that the rate constants for both reactions showed negative temperature dependencies. When k is expressed in the form of an Arrhenius equation, this necessarily leads to negative activation energies: k1 = (4.68 ± 0.70) × 10?12 exp (+0.23 ± 0.09 kcal/mole)/RT (219°-500°K) k2 = (4.68 ± 1.70) × 10?12 exp (+1.29 ± 0.23 kcal/mole)/RT (252°-500°K) Units are cm3 molec?1s?1. When a threshold energy of 0.0 kcal/mole is assumed for reaction (2), the temperature dependence of the preexponential term has a value of T?2. Making the usual simplifying assumptions, neither collision theory nor transition state theory leads to a preexponential factor with a strong enough negative temperature dependence. A comparison of these results with those derived from studies of the reactions of atomic oxygen, O(3P), with the same olefins shows that in both studies simple bimolecular processes were being examined. Also discussed are the possible experimental and theoretical ramifications of these new results.  相似文献   

7.
Abstract

It has been shown that host compound 1,1,6,6-tetraphenylhexa-2,4-diyne-1,6-diol is able to include polar guests and now we report on its ability to form clathrate compounds with apolar guests. The structures of this host with cyclohexane (1) and the ortho (2), meta (3) and para (4) xylenes have been determined and are discussed. Crystal data: (1) 2C30H22O2C6H12, M r = 913.20 g mol?1, mono-clinic, C2/c, a = 22.851(6), b = 14.010(2), c = 17.076(6) Å, β = 108.71(3)°, V = 5178(2) Å3, Z = 4, D c = 1.17g cm?3, N = 3326, R = 0.092. (2) 2C30H22O21 ½C8H10, M r = 1976.5 g mol?1, triclinic, P 1, a = 13.185(3), b = 15.466(3), c = 16.573(2) Å, α = 96.39(13)°, β = 106.96(15)°, γ = 114.94(18)°, V = 2822(2) Å3, Z = 2, D c = 1.16 g cm?3, N = 6152, R = 0.075. (3) 2C30H22O21 ½C8H10, M r = 1976.5 g mol?1, triclinic, P 1, a = 13.267(5), b = 15.453(3), c = 16.654(5) Å, α = 97.12(2)°, β = 107.09(3)°, γ = 114.68(3)°, V = 2843(2) Å3, Z = 2, D c = 1.15 g cm?3, N = 6505, R = 0.083. (4) 2C30H22O21 ½C8H10, M r = 1976.5 g mol?1, triclinic, P 1, α = 13.070(2), b = 15.348(3), c = 16.776(3) Å, α = 67.88<2)°, β = 74.27(1)°, γ = 65.29(1)°, V = 2817(1) Å3, Z = 2, D c = 1.15 g cm?3, N = 6711, R = 0.050. Thermal analysis studies were also performed in order to examine their stability and the strength with which the guest species are held in the crystal lattice.  相似文献   

8.
The kinetics of the reactions of O(3P) and D atoms with cyclohexane have been investigated using fast-flow techniques. The rates of reaction were computed by monitoring changes in both atom and cyclohexane concentrations using electron spin resonance and mass spectrometric methods, respectively. The O(3P) + C6H12 reaction was studied over a temperature range of 344 to 513°K and we obtain a specific rate constant of (3.2±0.6) × 1014 exp (?4400±400/RT) cm3/mole˙sec for this reaction. The only reaction product detectable mass spectrometrically under flow conditions of excess oxygen atoms is formaldehyde. The D + C6H12 reaction was studied over a temperature range of 297 to 596°K. A specific rate constant of (4.1±1.0) × 1013 exp (?4000±300/RT) cm3/mole˙sec was obtained for this reaction. On the basis of the results obtained in these studies, the important primary process in both the O(3P) and D atom reactions is concluded to be abstraction of a hydrogen atom from the cyclohexane molecule.  相似文献   

9.
X-Ray diffraction, density, and electrical conductivity measurements were performed on the perovskite-like mixed oxide La0.84Sr0.16MnO3. A rhombohedral crystalline structure with lattice parameters a = 3.893 Å and α = 90°29′16″ was assigned to the powder prepared by standard ceramic technique. Its theoretical density is therefore 6.576 g/cm3, while the experimental density was determined as 6.48 g/cm3. The conductivity measured at 1000°C is 133 Ω?1 cm?1. The temperature dependence of the conductivity indicates that the charge carriers are small polarons. The activation energy of the mobility is 9.6 kJ/mole.  相似文献   

10.
Yang  Wen‐Bin  Lu  Can‐Zhong  Zhuang  Hong‐Hui 《中国化学》2003,21(8):1066-1072
Since two interesting inorganic “host‐guest” polyoxomolybdates 1 and 2 have been reported previously, we have now succeeded in selectively isolating three new acetated “host‐guest” polyoxomolybdates 3–5, which considerably extend the range of structures in the cyclic polyoxomolybdate catalogue. 3 crystallizes in the triclinic space group P‐1 with a = 1.22235(1) nm, b = 1.52977(2) nm, c = 1.54022(1) nm, a = 113.746(1)°, β = 96.742(1)°, γ = 101.564(1)°, V = 2.51892(4) nm3, Z =1, Dc = 2.568 g. cm?3. 4 and 5 crystallize in the monoclinic system: P2(1)/n, a = 1.08298(2) nm, b = 1.54029(1) nm, c = 2.78893(5) nm, β =94.2730(10)°, V = 4.63929(12) nm3, Z = 2 and Dc = 2.671 g. cm?3 for 4, and C2/c, a =2.59907(8) nm, b = 1.65992(3) nm, c = 2.28473(7) nm, β‐93.4370(10)°, V = 9.8392(5) nm3, Z = 4 and Dc = 2.556 g. cm?3 for 5. The structures of 3, 4 and 5 consist of 18‐membered “host‐guest” polyoxoanions [ Na (X)2| ∈ |(μ3‐OH)4Moy8MoVI1052(μ2‐CH3COO)2]?(R+9 (X = CH3COO?for 3, DMF for 4 and H2O for 5), which are connected via Na* ions or hydrogen bonds into infinite extended frameworks.  相似文献   

11.
Hu  Mao-Lin  Huang  Zhen-Yan  Cheng  Ya-Qian  Wang  Shm  Lin  Juan-Juan  Hu  Yi  Xu  Duan-Jun  Xu  Yuan-Zhi 《中国化学》1999,17(6):637-643
The title complex Eu(III)(TTA)3(phen) (where TTA = thenoyltrifluoroacetone monoanion; phen = 1,10-phenanthroline) has been synthesized in mixed solvents of acetone and ethanol (1:1 volume ratio) and its crystal structure has been determined by X-ray diffraction. The complex crystals are triclinic, space group P 1 (# 2) with cell dimensions of a = 1.32.41 (2) nm, b = 1.5278(4) nm, c = 0.9755(3) nm, α = 92.49 (2)°, β = 102.57(2)°, γ = 91.62(2)°, V = 1.9268(8) nm3, Z = 2, μ (Mo Ka)= 18.77 cm?1, Dx=1.720 g/cm3. The coordination geometry of Eu atom is a distorted square antiprism, and the encapsulated structure that can meet the structural requirement of the typical europium luminescent sensor. The fluorescence spectrum suggests that the complex is a strong photoluminescent material.  相似文献   

12.
The ignition of COS + O2 mixtures diluted in argon was studied behind reflected shocks in a single-pulse shock tube over the temperature range of 1100–1700°K. Ignition delay times and the distribution of reaction products before and after ignition were determined experimentally. From a total of 63 tests run at varying initial conditions, the following correlation for the induction times was derived: where β1 = +0.30, β2 = 1.12, and E = 16.9 kcal/mole. Using a reaction scheme of 14 steps, the following values were obtained by a computer modeling of the induction times: β1 = +0.22, β2 = 1.55, and E = 17.3 kcal/mole. The calculations showed that the reaction COS + S → CO + S2 caused the inhibiting effect of the COS. The reaction COS → O ± CO2 + S has a very strong accelerating effect, whereas the parallel channel COS + O → CO + SO shows the opposite effect. It was also shown that the reaction O + S2 → SO + O is very slow and does not contribute to the overall oxidation reaction. It is suggested that the rate constant given to the four-center reaction COS + SO → CO2 + S2, that is, 1011 cm3/mole · sec at 300°K is incorrect. This constant is not much higher than 108 cm3/mole · sec at 1300°K.  相似文献   

13.
Second harmonic generation (SHG) was used to measure the temperature dependence of the reorientation activation volume of 4-(diethylamino)-4′-nitrotolane (DEANT) in poly(methyl methacrylate) (PMMA). The decay of the SHG signal from films of DEANT/PMMA was recorded at hydrostatic pressures up to 3060 atm and at different temperatures between 25°C below the glass transition temperature to 35°C above it. The activation volume, ΔV*αβ associated with the long range α-type motion of the polymer remained constant at 213 ± 10 Å3 between Tg − 25°C and Tg + 10°C. At higher temperatures, ΔV*αβ decreased linearly with increasing temperature. The activation volume, ΔV*αβ, associated with short range secondary relaxations was constant over the entire temperature range with a value of 77 ± 10 Å3. The data suggest that above Tg chromophore reorientation is coupled to both the long range and local motions of the polymer; whereas, well below Tg chromophore reorientation is closely coupled to the local relaxations of the polymer. © 1998 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 36: 901–911, 1998  相似文献   

14.
The title compound crystallizes in the monoclinic space group P21/n with a = 8.706(1), b = 9.192(1), c = 15.261(2) Å, β = 94.740(3)°, V = 1217.0(3) Å3, Z = 4, Dcalc = 2.278 g cm–3. The tris(imidazolyl)borate ligand bridges between three thallium atoms. The structure consists of one‐dimensional twisted ladder‐like strands.  相似文献   

15.
The transport properties of oxygen in poly(dimethyl siloxane) have been measured using the method of quenching of fluorescence. This paper discusses the uniqueness of this method and its use in measuring the diffusion coefficient of oxygen in unfilled PDMS. The results show (1) a large value for the diffusion coefficient of oxygen in pure PDMS at 25°C, D = 3.55 × 10?5 cm2/s, (2) a low value of the acitivation energy, ED = 4.77 kcal/mole, which was not temperature dependent in the ranges evaluated, and (3) a large value of the preexponential term, D0 = 0.115 cm2/s. The diffusion coefficient was found to be independent of both the oxygen concentration and fluorophor concentration in the pressure and temperature ranges used in these experiments. The import of these experiments lies in their application to a unique biomedical oxygen sensor which is fast, sensitive, and does not consume oxygen.  相似文献   

16.
In recent years the volume of activation Δ V* has become a powerful tool in chemical kinetics. High resolution NMR. spectroscopy is now one of the most common techniques used in the study of the kinetics of labile chemical systems. In order to measure Δ V* by this technique, we have built a 1H-probe-head, for a Fourier transform spectrometer, working up to 4 kbar and with a resolution of 0.6 Hz. The temperature is regulated and measured with an accuracy better than 0.2°. The high pressure probe-head has been tested on a chemical system showing a dissociative-associative crossover for the ligand substitution mechanism. It had been shown previously that the ligand exchange TaBr5 · L + *L ? TaBr5 · *L + L proceeds via a D mechanism when L=Me2O, and via an Ia mechanism when L=Me2S. As expected, ΔV is positive (+30.5 ± 2.0 cm3 mol?1) for the dissociative process and negative (?12.6 ± 1.0 cm3 mol?1) for the associative one.  相似文献   

17.
The temperature dependence of the NMR spectrum of 5,5-dimethyl-3,7-dithia-1,2-benzocyclo-heptene-( 1 ) is described and discussed. This compound occurs in two conformers with different topographies of the seven membered ring. From the chemical shift of the 2 geminal methyl groups (obtained by low-temperature spectra) it can be shown that the ring occurs in one case in the chair form and in the other case in the twist form. The free conformational energy of the twist form is only about 20 cal/mole. Two conformational changes can be distinguished: the conversion between chair and twist forms and the pseudorotation of the twist forms. The free activation enthalpies of the conversion (ΔGV) and the pseudorotation (ΔGP) are 12·6 ± 0·1 Kcal/mole (at ?12°C) and 8·3 ± 0·3 Kcal/mole (at ?95°C) respectively.  相似文献   

18.
Thermodynamic properties (ΔH°f(298), S°(298) and Cp(T) from 300 to 1500 K) for reactants, adducts, transition states, and products in reactions of CH3 and C2H5 with Cl2 are calculated using CBSQ//MP2/6‐311G(d,p). Molecular structures and vibration frequencies are determined at the MP2/6‐311G(d,p), with single‐point calculations for energy at QCISD(T)/6‐311 + G(d,p), MP4(SDQ)/CbsB4, and MP2/CBSB3 levels of calculation with scaled vibration frequencies. Contributions of rotational frequencies for S°(298) and Cp(T)'s are calculated based on rotational barrier heights and moments of inertia using the method of Pitzer and Gwinn [1]. Thermodynamic parameters, ΔH°f(298), S°(298), and CP(T), are evaluated for C1 and C2 chlorocarbon molecules and radicals. These thermodynamic properties are used in evaluation and comparison of Cl2 + R· → Cl· + RCl (defined forward direction) reaction rate constants from the kinetics literature for comparison with the calculations. Data from some 20 reactions in the literature show linearity on a plot of Eafwd vs. ΔHrxn,fwd, yielding a slope of (0.38 ± 0.04) and intercept of (10.12 ± 0.81) kcal/mole. A correlation of average Arrhenius preexponential factor for Cl· + RCl → Cl2 + R· (reverse rxn) of (4.44 ± 1.58) × 1013 cm3/mol‐sec on a per‐chlorine basis is obtained with EaRev = (0.64 ± 0.04) × ΔHrxn,Rev + (9.72 ± 0.83) kcal/mole, where EaRev is 0.0 if ΔHrxn,Rev is more than 15.2 kcal/mole exothermic. Kinetic evaluations of literature data are also performed for classes of reactions. Eafwd = (0.39 ± 0.11) × ΔHrxn,fwd + (10.49 ± 2.21) kcal/mole and average Afwd = (5.89 ± 2.48) × 1012 cm3/mole‐sec for hydrocarbons: Eafwd = (0.40 ± 0.07) × ΔHrxn,fwd + (10.32 ± 1.31) kcal/mole and average Afwd = (6.89 ± 2.15) × 1011 cm3/mole‐sec for C1 chlorocarbons: Eafwd = (0.33 ± 0.08) × ΔHrxn,fwd + (9.46 ± 1.35) kcal/mole and average Afwd = (4.64 ± 2.10) × 1011 cm3/mole‐sec for C2 chlorocarbons. Calculation results on the methyl and ethyl reactions with Cl2 show agreement with the experimental data after an adjustment of +2.3 kcal/mole is made in the calculated negative Ea's. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 548–565, 2000  相似文献   

19.
The title compound crystallizes in the monoclinic space group P21 with a = 7.928(6), b = 9.306(4), c = 17.16(2) Å, β = 92.06(8)°, V = 1265(2) Å3, Z = 2, Dcalc = 2.191 g cm–3. From two independent molecular units, metal-ligand strands are formed based on electrostatic interactions between the thallium centers and pyrazolyl π manifolds from neighboring molecules.  相似文献   

20.
Ammonium hydrogen hexaaquacobaltate(III) isopolyvanadate [(NH4)2][Co(H2O)6] · H[V10O28] · 8H2O was synthesized and studied by X-ray diffraction, thermogravimetry, and IR and NMR spectroscopy. The crystals are triclinic, space group P [`1]\bar 1, a = 8.216(2), b = 9.965(2), c = 11.796(2) ?, α = 77.71(3)°, β = 71.11(3)°, γ = 86.11(3)°, V = 867.0(3) ?3, ρ(calcd.)= 2.501 g/cm3, Z = 1.  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