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 10 毫秒
1.
The rate coefficients for the reactions of hydrogen atoms with n-C3H7Br, s-C3H7Br, n-C4H9Br, and s-C4H9Br were determined in a discharge flow-reactor at 298 K and a pressure of 4 mbar. Molecular-beam sampling and subsequent mass-spectrometric detection with electron-impact ionisation was used for the measurement of the bromo-hydrocarbon concentration. The rate coefficients obtained are (in 1010 cm3 mol−1 s−1): 2.3±1.2 for n-C3H7Br, 2.3±1.2 for s-C3H7Br, 2.4±1.2 for n-C4H9Br, and 2.8±1.4 for s-C4H9Br. The results are compared with predictions from bond-energy bond-order (BEBO) calculations, where a reasonable agreement is found. Furthermore, also by BEBO calculations, the relative importance of bromine abstraction as compared to hydrogen abstraction is estimated. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 721–727, 1998  相似文献   

2.
A study was carried out on the rate constant ratio (k RH/k EtH) in reactions of alkanes C3H8, n-C4H10, n-C5H12, n-C6H14, i-C4H10, c-C5H10, and c-C6H12 with OH radicals in water at 5-55°C and the relative activation parameters A RH/A EtH and E EtHE RH. The values of E EtHE RH in water and the gas phase have opposite signs. The values of k RH/k EtH decrease with increasing temperature in the gas phase but increase in water. The behavior of these reactions in water may be attributed to a solvent cage effect.  相似文献   

3.
The 13C n.m.r. spectra of forty alkoxysilanes of the general type XnSi(OR)4–n (X = CH3, C6H5, H; R = CH3, C2H5, n-C3H7, i-C3H7, n-C4H9, i-C4H9, s-C4H9, n-C5H11, CH(CH3)(C6H5), C6H5) have been recorded and assigned. The chemical shifts of the α-carbon resonances of the alkoxy groups are shown to depend on both the nature of the alkoxy group and the number and type of substituents on the silicon. Regression analyses of the data give empirical substituent chemical shift (SCS) parameters for the silyl substituents. The β-carbon resonances are shown to be dependent on the presence of the silyl group, but not the specific silyl substituents.  相似文献   

4.
Trejo Rodríguez, A. and Patterson, D., 1984. Prediction of activity coefficients and Henry's constants at infinite dilution for mixtures of n-alkanes. Fluid Phase Equilibria, 17: 265–279.The corresponding-states principle (CSP) for chain molecules has been used to predict activity coefficients γi at infinite dilution and Henry's constants Hi,j for mixtures of n-alkanes. The mixtures for which predictions of γi are made include n-C4 through n-C10 as the solute in several pure higher n-alkanes, at several temperatures from 30 to 90°C. Predictions of Hi,j are made for mixtures where the solute is C2 through n-C8 and the solvent any n-alkane from n-C4 through n-C22, also at several temperatures, from –15 to 177°C. The predicted values are compared with experimental data from the literature, and in all cases the agreement is remarkably good. The temperature dependence of Hi,j for some mixtures is used to derive heats of solution ΔHs, and comparison is again carried out with available experimental values.  相似文献   

5.
n-C26H54, n-C27H56, n-C28H58, n-C29H60, n-C30H62, n-C31H64, n-C32H66, n-C33H68, stigmasterol, β-sitosterol and two unidentified compounds: a saturated ester (mp 79-81°C); an unsaturated alcohol (mp 87-88°C) were isolated from the whole herb of Moghania strotilifera.  相似文献   

6.
Methods are discussed for the production and detection of the hydroperoxyl radical for use in gas phase kinetic studies. Rate constants for gas phase reactions of the hydroperoxyl radical with itself, H2, H2O, CO, NO, SO2, O3, C2H6, C3H8, i-and n-C4H10, C2H4, i-C4H8, HCHO, C2H5CHO, n-C3H7CHO, Br, O, OH, and H are critically evaluated. Recommended or estimated rate constant expressions with associated error limits are given applicable over specified temperature ranges (normally 300–1000°K). The reactivity of HO2 compared with OH, O, H, F, Cl, Br, CH3, and CH3O is presented in tabular form and the implications for atmospheric chemistry are discussed.  相似文献   

7.
The Flory–Huggins interaction parameters χ for 23 gases (He, Ne, Ar, Kr, Xe, H2, N2, O2, N2O, CO2, CH4, C2H4, C2H6, C3H6, C3H8, 1,3-C4H6, four C4H8's, n-C4H10, iso-C4H10, and n-C5H12) in five rubbery polymers (1,2-polybutadiene (PB), poly(ethylene-co-vinyl acetate)) (EVAc), polyethylene (PE), polypropylene (PP), and poly(dimethyl siloxane) (PDMS) were determined from either literature data on Henry's law coefficient and partial molar volume or those on sorptive dilation for each polymer/gas system. Values of χ for the gases increased in the order of PDMS < PP ≡ PB < EVAc ≡ PE. Among the gases except He and H2 whose χ values are not reliable, Ne and Xe have respectively the highest and the lowest values of χ for the polyolefins. The χ values of the hydrocarbons were compared together with previously reported χ values of n-alkanes C3-C10. The dependencies of χ upon concentration and temperature were discussed on the basis of the literature data. © 1997 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 35 : 1049–1053, 1997  相似文献   

8.
A method is described for extrapolating existing experimental data on the reactions of OH radicals with alkanes to higher temperatures using conventional transition-state theory. Expressions are developed for the estimation of the structural properties of the activated complex necessary for calculating ΔS± and ΔH±. The vibrational frequencies and internal rotations of the activated complex are given by those of the reacting alkane or the analogous alcohol and a set of additional internal modes that is the same for all OH + alkane reactions considered. Differences between primary, secondary, and tertiary hydrogen attack are discussed, and the validity of representing the activated complexes of all OH + alkane reactions by a fixed set of vibrational frequencies and other internal modes is evaluated. Calculations are presented for the reaction of OH with CH4, C2H6, C3H8, n-C4H10, i-C4H10, c-C4H8, c-C5H10, c-C6H12, (CH3)2CHCH(CH3)2, (CH3)3CCH(CH3)2, (CH3)4C, and (CH3)3CC(CH3)3, and the results are compared with experiments.  相似文献   

9.
The kninetics of acid-catalyzed acetalization and ketalization of poly(vinyl alcohol) (PVA) were systematically studied in completely homogeneous media with carefully selected solvents. Thus the acetalization reaction was run in water with six aldehydes [R1CHO (R1 = H, CH3, C2H5, n-C3H7, i-C3H7, ClCH2)], whereas the ketalization in dimethylslfoxide with 11 ketones [R2CH3CO (R2 = CH3, C2H5, n-C3H7, i-C3H7, n-C4H9, i-C4H9, tert-C4H9, C6H5CH2, C6H5CH2CH2), cyclopentanone, and cyclohexanone]. The latter was difficult to proceed in aqueous media. Both reactions were reversible and bimolecular and, despite the use of different solvents, gave similar heats of reaction (7.5 kcal/mol) and activation energies (ca. 15 kcal/mol) except for the case of formaldehyde and chloroacetaldehyde; however the equilibrium constants at 25°C showed that the acetalization is thermodynamically much more favored than the ketalization (ca. 5000 vs. 0.01–0.9), probably because of steric hindrance of the ketone substrate. The rate constants of hydrolysis (reverse reactions) for the poly(vinyl acetal) and poly(vinyl ketal) followed the Hammett-Taft equation to give a single p* (=3.60) that is very close to that for the hydrolysis of diethyl acetal and ketal. From these and other data, it was concluded that the polymer hydrolysis, as well as PVA acetalization and ketalization, are all electrophilic reaction where the formation of hemiacetal or hemiketal is the rate-determining step. © 1996 John Wiley & Sons, Inc.  相似文献   

10.
Methyl radical reactions with matrix molecules in glasses C2H5OH, (CH2OH)2, n- and i-C3H7OH, n- and i-C4H9OH, n- and i-C5H11OH, C2D5OH, and i-C3D7OD, and the reactions of ?2H5, ?3H7, ?4H9, ?5H11 with methanol glasses have been studied. Alkyl radicals were produced by photolysis of diphenylamine–alkylhalide–alcohol mixtures using ultraviolet light. In all cases the alkyl radical decay follows the law c = c0 exp(-kt). The √t law should not be associated with alkyl radical diffusion in a matrix. A method of processing the kinetics of those reactions in which one paramagnetic species changes into another with the total concentration being constant and the electron spin resonance spectra of both species overlapping, is described.  相似文献   

11.
Poly(2,3-dialkylbutanediol-1,4 terephthalates) with the alkyl substituents CH3, C2H5, n-C3H7, iso-C3H7, n-C4H9, and n-C10H21, andn-C16H33 were synthesized from the corresponding 2,3-dialkylbutanediols-1,4 and dimethyl terephthalate or terephthaloyl chloride. The substituents of the butanediol-1,4 portion of the polyterephthalates influence the 13C NMR chemical shifts of the carbon atoms near the branching site, the glass transition (Tg), and the crystallizability. Small alkyl substituents do not change the Tg of the polymers, whereas bulky substituents such as the isopropyl group increase the Tg and long normal alkyl groups as substituents decrease the Tg of the polymers. Crystallinity in these polyterephthalates was found only with CH3 and C16H33 as the 2,3-dialkyl substituents in the butanediol-1,4 portion of the polyester. This crystallinity of polyterephthalate of 2,3-di-C16H33 substituted butanediol-1,4 could be assigned to side-chain crystallization of the paraffinic groups.  相似文献   

12.
The kinetics of thermal decomposition of solid In(S2CNR2)3 complexes, (R=CH3, C2H5, n-C3H7,i-C3H7, n-C4H9 and i-C4H9), has been studied using isothermal and non-isothermal thermogravimetry. Superimposed TG/DTG/DSC curves show that thermal decomposition reactions occur in the liquid phase, except for the In(S2CNMe2)3 and In(S2CNPri 2)3 compounds. This revised version was published online in July 2006 with corrections to the Cover Date.  相似文献   

13.
The reactions of ?-C3H3+ (propargylium cation) with acetylene and diacetylene have been modeled kinetically. Data were obtained from Fourier Transform Ion Cyclotron Resonance (FTICR) experiments on these systems, which are themselves models for soot particle initiation. Acetylene forms an encounter complex with ?-C3H3+, but, in the absence of a third body collision, the complex decomposes to acetylene and c-C3H3+ (cyclopropenylium cation) at about 1/3 the rate it decomposes to acetylene and ?-C3H3+, in spite of the fact that c-C3H3+ is ca. 115 kJ/mol more stable than ?-C3H3+. The encounter complex is long enough lived, and energetic enough, to scramble deuterium in reactions between ?-C3H3+ and C2D2. These reactions have been successfully modeled, yielding a nearly statistical distribution of deuterium, and a rather large kinetic isotope effect. The more complex reactions of ?-C3H3+ with diacetylene have also been modeled.  相似文献   

14.
The removal of *UF6 (A state) molecules by selected alkanes has been investigated at 25°C. The following rate constants (units of 1011 l/mol·sec) were evaluated: iso-C4F10, 0.0432 ± 0.0115; n-C4F10, 0.0764 ± 0.020; C2F6, 0.0192 ± 0.0052; CH4, 0.0612 ± 0.0061; C2H6, 3.78 ± 0.60; C3H8, 5.08 ± 0.60; n-C4H10, 5.05 ± 0.78; iso-C4H10, 4.17 ± 1.15; neo-C5H12, 6.59 ± 0.93; CF3? CH3, 0.0385 ± 0.0056; CF2H? CF2H, 0.0729 ± 0.0074; and CF2H? CFH2, 0.149 ± 0.015. The perfluoro-alkane quenching of *UF6 proceeds via a physical mechanism. The other alkane quenching reactions are consistent with a chemical mechanism also contributing in varying degrees which may involve removal of two hydrogens from the alkane.  相似文献   

15.
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   

16.
The thermodynamic and kinetic parameters of the thermal decomposition of Zn(S2CNR2)2 complexes (R=CH3, C2H5 and n-C3H7) were determined with the dynamic thermogravimetric method. Superimposed TG/DTG/DSC curves show that thermal decomposition reactions for chelates with R=C2H5 and n-C3H7 occur in the liquid phase, at temperatures far away from their melting points, whereas for the complex with R=CH3 the thermal decomposition begins at a temperature closer to its melting point, suggesting a rather complex decomposition mechanism. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

17.
Polymerization of 1-methylthio-1-alkynes (MeSC?CR; R = Et, n-Bu, n-C6H13, and n-C8H17) was studied by use of transition metal catalysts. A 1 : 2 mixture of MoCl5 and Ph3SiH provided polymers having M?w over 1 × 105 in 30–50% yields from these monomers. The length of the alkyl group hardly affected the polymerization. The monomer, MeSC?C-n-C6H13, showed low reactivity in homopolymerization, but higher reactivity than that of MeC?C-n-C5H11 in copolymerization. Poly(1-methylthio-1-alkyne)s were colorless solids, and those with long alkyl pendants (R = n-C6H13, n-C8H17) were soluble in various organic solvents. The present polymers were thermally more stable than poly(2-alkyne)s, the corresponding hydrocarbon polymers.  相似文献   

18.
The molecular and electronic structures and thermodynamic parameters of 1,2-, 1,7-, and 1,12-dicarba-closo-dodecaborane(12) molecules in the singlet ground state were calculated by the Hartree-Fock, DFT, and MP2 (including B3LYP/6-311G(2d,2p) and MP2/6-311+G(d,2p) methods). The energies and character of spatial localization of the frontier MOs in o-, m-, p-carboranes(12) correlate with the changes in the configuration stabilities and reactivities in the reactions of carboranes with electrophilic and nucleophilic agents and bases. The electrostatic potential distributions in the molecules and the atomic charge distribution of hydrogen atoms correlate with the known chemical properties of carboranes(12). The thermodynamic parameters of two isomerization reactions, o-C2B10H12m-C2B10H12 and m-C2B10H12p-C2B10H12, calculated for the temperature range 298–1000 K agree with experimental data within the limits of measuring error. The values of the electron density and the Laplacian of the electron density at the (3, −1) critical points of the B-H and C-H bonds correlate with the reactivities of the title compounds. Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 12, pp. 2074–2080, December, 2006.  相似文献   

19.
Solid state reactions of acids RCOOH (R = n-C7H15, BuC(Et)H, n-C9H19, PhCH2, PhCH2CH2, H2C=CH(CH2)8, or MeOOC(CH2)3) with Pb(OAc)4 combined with KCl, NaCl, CdCl2, or NH4Cl in the absence of a solvent and without mechanical activation afford chlorohydrocarbons RCl. The corresponding reactions of acids HOOC(CH2)nCOOH (n = 3–6) give dichloroalkanes Cl(CH2)nCl and γ-butyrolactone (n = 3).__________Published in Russian in Izvestiya Akademii Nauk. Seriya Khimicheskaya, No. 10, pp. 2105–2109, October, 2004.  相似文献   

20.
Copolymers with the equimolar composition and presumably alternating microstructure are synthesized by the radical copolymerization of methylene alkanes, the products of dimerization of α-olefins with the general formula RCH2CH2C(=CH2)R (R = n-C4H7, n-C6H13, n-C8H17, n-C10H21, n-C12H25), with maleic anhydride. The products of reactions between the obtained copolymers and 1-octadecanamine or 1-octadecanol are isolated and spectrally characterized. The depressor efficiency of the copolymers with respect to the solutions of paraffins in n-alkanes is studied. Qualitative differences between the copolymers of maleic anhydride with methylene alkanes and reference copolymers based on 1-octadecene are estimated using vibrational viscometry combined with analysis of the size and morphology of paraffin crystals. It is shown that the copolymers with methylene alkanes more effectively decrease the cold filter plugging point (CFPP) of paraffin solutions in n-decane.  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