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 265 毫秒
1.
Binary solid solutions of n-paraffins (n-C50H102/n-C60H122) epitaxially prepared on potassium hydrogen phthalate substrate from the vapor phase have been studied by electron diffraction to characterize their phase transitions and structure. The continuity of solid solution in the n-C50H102/n-C60H122 system is demonstrated once lamellar ordering of the crystal packing is achieved. However, such ordering is achieved only by annealing and proceeds through a series of intermediate chain packings. At first, the electron diffraction patterns from all samples prepared at room temperature resemble those from polyethylene, in which no spots corresponding to the interlayer spacings appear. Longitudinal chain translations are induced by annealing to cause the lamellar reflections to appear, while the “polyethylene” subcell reflections remain unsplit until the crystal structure with well-defined methyl end planes is reached.  相似文献   

2.
《Fluid Phase Equilibria》1999,163(2):165-173
Monte Carlo simulation has been applied to calculate the pVT relationship of CO2+butane (n-C4H10 and i-C4H10) systems at 310.93 K and up to 9.5 MPa. CO2 is treated as a single-site molecule and butanes are treated as four-site molecules. The Lennard–Jones (12–6) potential is used as the site–site potentials and the combining rules proposed by Jorgensen et al. [W.L. Jorgensen, J.D. Madura, C.J. Swenson, J. Am. Chem. Soc. 106 (1984) 6638.] are adopted for unlike site pairs. The calculated results of the pVT relationship show good agreement with the experimental data [T. Tsuji, S. Honda, T. Hiaki, M. Hongo, J. Supercrit. Fluids 13 (1998) 15.] by introducing an intersite interaction parameter between unlike molecules. Furthermore, the radial distribution functions and the number of CO2 and butanes around butanes are calculated as a fundamental information on the microscopic structures. It is found that the radial distribution functions and the number of CO2 and n-C4H10 around n-C4H10 are different from those of CO2 and i-C4H10 around i-C4H10.  相似文献   

3.
Penicillin potassium salt (penicillin-K) is found to show hydrotrope action, which can increase the solubility of cationic surfactant CTAB in water. Penicillin-K also shows hydrotrope-solubilization action, which makes the W/O and O/W microemulsion more stable and increases the solubilized amount of n-C5H11OH in O/W microemulsion and that of water in W/O microemulsion for CTAB/n-C5H11OH/H2O system. However, in this system, the presence of penicillin-K can decrease the stability of the lamellar liquid crystal phase due to its structure change to bicontinuous, which are proved by the mechanism of its hydrotrope-solubilization action.  相似文献   

4.
The Raman spectrum of n-C36H74 has been measured in the region between 0 and 150 cm?1. Eleven newly observed bands have been assigned to the intramolecular skeletal vibrations and rotatory lattice vibrations on the basis of the dispersion curves calculated previously.  相似文献   

5.
Controlled release of cephanone from hexadecyltrimethylammonium bromide (CTAB) micelles and CTAB/n-C5H11OH/H2O microemulsions was studied. The results showed that the release rate of cephanone was reduced in CTAB micelles and CTAB/n-C5H11OH/H2O microemulsions, because of the solubilization of cephanone in micelles and microemulsions. The release of cephanone from CTAB micelles and CTAB/n-C5H11OH/H2O microemulsions was characterized by Fickian diffusion and non-Fickian diffusion.  相似文献   

6.
(η-C5H5)(CO)2W[(η3-C5H5)(C5H5)2], I, containing two tilted five-membered rings, is converted into the bridged ferrocene derivative (η-C5H5)(CO)2W{(η3-C5H5)}[(η-C5 H4)2Fe]} II by successive reaction with Na and FeCl2.  相似文献   

7.
The low-frequency Raman spectra of triclinic n-paraffins, n-C8H18 through n-C24H50, were observed. The normal-coordinate treatments of crystal vibrations of n-C8H18 through n-C18H38 were carried out. Six characteristic series of the observed Raman lines were assigned to rotatory lattice vibrations and intramolecular skeletal vibrations.  相似文献   

8.
The perovskite-type layered compounds decylammonium tetrachlorocobaltate (n-C10H21NH3)2CoCl4 and dodecylammonium tetrachlorocobaltate (n-C12H25NH3)2CoCl4 exist the solid–solid phase transition in the temperatures range 330–380 K. These laminar materials contain bilayers sandwiched between metal halide layers. The experimental subsolidus binary phase diagram of (n-C10H21NH3)2CoCl4–(n-C12H25NH3)2CoCl4 has been established over the whole composition range by differential thermal analysis and X-ray diffraction. In the phase diagram, one intermediate compound (n-C10H21NH3) (n-C12H25NH3)CoCl4 at $ W_{{{\text{C}}_{ 1 0} {\text{Co}}}} \% = 50. 9 2 $ and two eutectoid invariants points at $ W_{{{\text{C}}_{ 10} {\text{Co}}}} \% = 2 9. 8 3 $ and $ W_{{{\text{C}}_{ 10} {\text{Co}}}} \% = 7 7. 9 8 $ were observed, two eutectoids temperatures are about 343 ± 1 and 340 ± 1 K. There are three noticeable solid solution ranges (α, β, γ) at the left and right boundary and middle of the phase diagram.  相似文献   

9.
The non-empirical generalized Kirkwood, Unsöld, and the single-Δ Unsöld methods (with double-zeta quality SCF wave-functions) are used to calculate isotropic dispersion (and induction) energy coefficients C2n, with n ? 5, for interactions involving ground state CH4, C2H6, C3H8, n-C4H10 and cyclo-C3H6. Results are also given for the related multipole polarizabilities αl, multipole sums S1/(0) and S1(?1) which are evaluated using sum rules, and the permanent multipole moments. for l = 1 (dipole) to l = 3 (octupole). Estimates of the reliability of the non-empirical methods, for the type of molecules considered, are obtained by a comparison with accurate literature values of α1S1(?1) and C6. This, and the asymptotic properties of the multipolar expansion of the dispersion energy, the use to discuss recommended representation for the isotropic long range interaction energies through R?10 where R is the intermolecular separation.  相似文献   

10.
Temperature study of Raman spectra of alkylgermanes (Alk = Bun, n-C6H13) in liquid, polycrystalline, and glassy states was carried out. The spectra of liquid samples are complicated, because compounds exist as equilibrium mixtures of rotational isomers due to hindered rotation about C—C and Ge—C bonds. In crystals, only the most stable conformer persists, which results in simplification of the spectrum. Unlike n-hexylgermanes, n-butylgermanes crystallize on cooling with difficulties. Quantum-chemical calculations of the geometry and normal coordinate analysis of the possible conformers of the BunGeH3, BunGeCl3, and Bun 2GeCl2 molecules were performed. The unusually large (65 cm–1) difference between the experimental (Ge—C) stretching frequencies of the trans and gauche conformers about the C—C bond nearest to the Ge atom is attributed to the difference in the electronic structure of these conformers, which is manifested in the differences in the Ge—C bond lengths, molecular geometry, and their force fields and, consequently, affects the mode frequencies and eigenvectors.  相似文献   

11.
The rate coefficients for the gas-phase reactions of C2H5O2 and n-C3H7O2 radicals with NO have been measured over the temperature range of (201–403) K using chemical ionization mass spectrometric detection of the peroxy radical. The alkyl peroxy radicals were generated by reacting alkyl radicals with O2, where the alkyl radicals were produced through the pyrolysis of a larger alkyl nitrite. In some cases C2H5 radicals were generated through the dissociation of iodoethane in a low-power radio frequency discharge. The discharge source was also tested for the i-C3H7O2 + NO reaction, yielding k298 K = (9.1 ± 1.5) × 10−12 cm3 molecule−1 s−1, in excellent agreement with our previous determination. The temperature dependent rate coefficients were found to be k(T) = (2.6 ± 0.4) × 10−12 exp{(380 ± 70)/T} cm3 molecule−1 s−1 and k(T) = (2.9 ± 0.5) × 10−12 exp{(350 ± 60)/T} cm3 molecule−1 s−1 for the reactions of C2H5O2 and n-C3H7O2 radicals with NO, respectively. The rate coefficients at 298 K derived from these Arrhenius expressions are k = (9.3 ± 1.6) × 10−12 cm3 molecule−1 s−1 for C2H5O2 radicals and k = (9.4 ± 1.6) × 10−12 cm3 molecule−1 s−1 for n-C3H7O2 radicals. © 1996 John Wiley & Sons, Inc.  相似文献   

12.
The reaction betweeen (η5-C5H5Mo(CO)3I and RNC is catalysed by [η5 -C5H5Mo(CO)3]2 and readily yields η5-C5H5Mo(CO)3?n(RNC)nI (n = 1–3). A free radical mechanism is consistent with experimental data.  相似文献   

13.
Infrared and Raman spectra were obtained for n-C3D7I and n-C4D9I that showed the existence of rotational isomers. The conformer with a co-planar chain of carbons and iodine is the one stable in the crystals of both compounds, analogous to the protonated species. Vibrational assignments were made with the aid of normal coordinate calculations. Force constants were determined by simultaneously fitting the calculated to the observed frequencies for 155 values of the eight molecules, trans and gauche n-C3H7I, trans and gauche n-C 3D7I, TT and TG n-C4H9I, and TT and TG n-C4D9I. The average error for the 155 frequencies is 6.3 cm?1.  相似文献   

14.
Compounds of the type π-C5H5NiPBu3SC(S)X (X = R, OR and NRH) are obtained from reactions between [π-C5H5Ni(PBu3)2]+Cl? and SC(S)X? in aqueous solution. Compounds such as π-C5H5NiPBu3SC(S)NRH are also obtained by reactions of π-C5H5NiPBu3SH with RNCS.Reactions of C6H5NCS with π-C5H5NiPBu3SEt or [π-C5H5NiPBu3S(CH2)n]2 (n = 1, 2 and 3) give π-C5H5NiPBu3[SC(NC6H5)SC2H5] or [π-C5H5NiPBu3 {SC(NC6H5)S(CH2)n}]2 respectively.Similar reactions of π-C5H5NiPBu3SH and RNCO given π-C5H5NiPBu3SC(O)NRH.Treatment of π-C5H5NiPBu3SC(S)R with HCl gives π-C5H5NiSC(S)R.  相似文献   

15.
Reactions of n-C4H9O radicals have been investigated in the temperature range 343–503 K in mixtures of O2/N2 at atmospheric pressure. Flow and static experiments have been performed in quartz and Pyrex vessels of different diameters, walls passivated or not towards reactions of radicals, and products were analyzed by GC/MS. The main products formed are butyraldehyde, hydroperoxide C4H8O3 of MW 104, 1-butanol, butyrolactone, and n-propyl hydroperoxide. It is shown that transformation of these RO radicals occurs through two reaction pathways, H shift isomerization (forming C4H8OH radicals) and decomposition. A difference of activation energies ΔE = (7.7 ± 0.1 (σ)) kcal/mol between these reactions and in favor of the H-shift is found, leading to an isomerization rate constant kisom (n-C4H9O) = 1.3 × 1012 exp(− 9,700/RT). Oxidation, producing butyraldehyde, is proposed to occur after isomerization, in parallel with an association reaction of C4H8OH radicals with O2 producing OOC4H8OH radicals which, after further isomerization lead to an hydroperoxide of molecular weight 104 as a main product. Butyraldehyde is mainly formed from the isomerized radical HOCCCC˙ + O2 ··· → O (DOUBLE BOND) CCCC + HO2, since (i) the ratio butyraldehyde/(butyraldehyde + isomerization products) = 0.290 ± 0.035 (σ) is independent of oxygen concentration from 448 to 496 K, and (ii) the addition of small quantities of NO has no influence on butyraldehyde formation, but decreases concentration of the hydroperoxides (that of MW 104 and n-propyl hydroperoxide). By measuring the decay of [MW 104] in function of [NO] added (0–22.5 ppm) at 487 K, an estimation of the isomerization rate constant OOC4H8OH → HOOC4H7OH, κ5 ≅ 1011exp(−17,600/RT) is made. Implications of these results for atmospheric chemistry and combustion are discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
The Fourier transform infrared and Raman spectra of the cations [η5-C9H7Fe(CO)n dppa]+ (n = 1, 2; dppa=bisdiphenylphosphinoalkane, where alkane=methane, ethane, butane, hexane and octane) and [{η5-C9H7Fe(CO)2}2-μ-dppa]2+ indicate that the alkyl chain lengths have effects on the structures of the bidentate cations resulting in increased back-donation to carbonyl groups as the chain length increases. In contrast the alkyl chain lengths have no similar effects in the unidentate mononuclear and bridged cations.  相似文献   

17.
The NMR second moment of a uniaxially oriented mat of single crystals of n-C32H66 (in the orthorhombic form) was measured at temperatures from ?170°C to 70°C and at various alignment angles γ between the orientation axis (preferential direction of the molecular chains) and the NMR magnetic field. Accurate expressions are given for the NMR second moment of an orthorhombic normal paraffin CnH2n+2 of arbitrary molecular chain length n for n ≥ 10, in the following states of molecular motion: no motion (a rigid lattice), rotation of CH3 groups, and rotation of the chains around their axes with superimposed rotation of CH3 groups. In addition to these well-known motions, n-C32H66 is found to exhibit an α process. The corresponding decrease of the NMR second moment shows the dependence on γ predicted for “flip-flop” motion, i.e., rotational jumps of the chain molecules around their axes through 180° and a simultaneous translation along these axes by one CH2 group. The overall decrease in second moment occuring at the transition to the hexagonal rotator phase in n-C32H66 can be quantitatively accounted for. The dependence of this decrease on the alignment angle γ, however, is in disagreement with calculations based on a simple rotation of the chains around their axes. Considerable torsion of the chains superimposed on the rotation would improve agreement between theory and experiment.  相似文献   

18.
The reductive high-pressure carbonylation of tetrachloro(η5-cyclopentadienyl)niobium using sodium as reduction agent and Cu/Al mixture as halogen acceptor system cleanly yields tetracarbonyl (η5-cyclopentadienyl)niobium which can be synthesized by means of this new procedure on a 35 g-scale, with yields ranging between 89 and 94%. Under photolysis conditions, (η5-C5H5) Nb(CO)4 is converted into the solvent complex (η5-C5H5) Nb(CO)3THF which, in turn, incorporates 13Co or 13C18O under mild conditions to give the labelled derivatives (η5-C5H5) Nb(CO)4 ? n(13CO)n and (η5-C5H5) Nb(CO)4 ? n(13C18O)n, respectively.  相似文献   

19.
Irradiation of solutions of n5-C5H5W(CO)3R (R  CH3n1-CH2C6H5) in cyclohexane at ca. 310490 nm leads to the formation of [n5-C5H5W(CO)3]2 and methane and of n5-C5H5W5(CO)2(n3-CH2C6H5) and some [n5-C5H5W(CO)3]2, respectively. When the irradiation is carried out in the presence of excess P(C6H5)3, the photoproducts are n5-C5H5W(CO)2[P(C6H5)3]CH3 (R  CH3) and n5-C5H5W(CO)2(n3-CH2C6H5) and trace [n5-C5H5W(CO)3]2 (R  n1-CH2C6H5). Photolysis of the n5-C5H5W(CO)3R in the presence of benzyl chloride affords n5-C5H5W(CO)3Cl (R  CH3) and both n5-C5H5W(CO)2(n3-CH2C2H5) and n5-C5H5W(CO)3Cl (R  n1-CH2C6H5), the relative amounts of the latter products depending on the quantity of added C6H5CH2Cl. Irradiation of n5-C5H5W(CO)3-CH3 in the presence of both P(C6h5)3 and C6H5CH2Cl affords n5-C5H5W(CO)2-[P(C6H5)3]CH3, but no n5-C5H5W(CO)3Cl. It is proposed that the primary photo-reaction in these transformations is dissociation of a CO group from n5-C5H5W-(CO)3R to generate n5-C5H5W(CO)2R, which can either combine with L to form a stable 18 electron complex, n5-C5H5W(CO)2(L)R (L  CO, P(C5H5)3; LR  n3-CH2C6H5), or lose the group R in a competing, apparently slower step. This proposal receives support from the observation that, light intensifies being equal, n5-C5H5W(CO)3CH3 undergoes a considerably faster photoconversion to [n5-C5H5W(CO)3]2 under argon than under carbon monoxide.  相似文献   

20.
The energy band dispersion E = E(k) is determined for the valence bands in a long-chain alkane n-C36 H74 by angle-resolved photoemission from an oriented polycrystalline sample using synchrotron radiation. Comparison with theoretical calculations for a single chain shows good qualitative agreement for the position, width, and dispersion of the valence band. This is the first observation of energy band dispersion in an organic molecular solid.  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