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.
Hydrogen, cycloalkene, and bicycloalkyl were found to be the principal products which account for ≈?97% of all products formed in the gas-phase radiolysis of water vapor containing low concentrations of cycloalkanes. From the ratios of cycloalkene-to-bicycloalkyl yields extrapolated to the zero dose, the self- and cross-disproportionation-to-recombination rate constant ratios Δ = kd/kc were determined for the following 12 reactions: Δ(c-C5H9, c-C5H9) = 0.73; Δ(c-C5D9, c-C5D9) = 0.58; Δ(c-C6H11, cC6H11) = 0.59; Δ(c-C6D11, c-C6D11) = 0.46; Δ(c-C5H9, c-C6H11) = 0.28; Δ(c-C5D9, c-C6H11) = 0.28; Δ(c-C5H9, c-C6D11) = 0.24; Δ(c-C5D9, c-C6D11) = 0.24; Δ(c-C6H11, c-C5H9) = 0.33; Δ(c-C6H11, c-C5D9) = 0.25; Δ(c-C6D11, c-C5H9) = 0.35; and Δ(c-C6D11, c-C5D9) = 0.28, where in the case of the cross-disproportionation the symbol Δ(R1,R2) is used to represent kd/kc for the disproportionation in which radical R1 captures a hydrogen (deuterium) atom from radial R2. The geometrical mean rule holds in the cross-combination reactions of cyclopentyl and cyclohexyl radicals. The kinetic isotope effect in the disproportionation reaction was determined as 1.24 ± 0.06.  相似文献   

2.
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   

3.
The reactions of the cyclic molecules C6H6 (benzene), c-C3H6 (cyclopropane) and c-C6H12 (cyclohexane) with ArH+ (ArD+), H3+, N2H+, CH5+, HCO+, OCSH+, C2H3+, CS2H+ and H3O+ have been studied at 300 K using a SIFT apparatus. All the reactions except those of C2H3+ proceed via proton transfer and all are fast except the H3O+ and CS2H+ reactions with c-C6H12 which are endothermic and which establish that the proton affinity of c-C6H12 is 160 ± 1 kcal mol−1, which is considerably lower than the published value. In the c-C3H6 and the c-C6H12 reactions multiple products are observed and hence “breakdown curves” for the protonated molecules are constructed and the appearance energies of the various ion products are consistent with available thermochemical data. The reactions of C2H3+ with these cyclic molecules are atypical within this series of reactions in that they appear to proceed largely via hydride ion transfer. The implications of the results of this study to interstellar chemistry are alluded to.  相似文献   

4.
Relative vibrational energy transfer efficiencies are determined for cyclobutane-t chemically activated to an average energy of 5 eV by recoil tritium replacement reaction. The pressure and composition dependence of the stabilization-decomposition ratio indicates relative efficiencies of 1.00, 1.05, and 0.32 for c-C4H8, CF4, and Ne bath gases.  相似文献   

5.
Vapour-phase radiolysis of c-C6D12 and n-C7D16 leads to a relatively large HD yield, up to ~ 20% of the total hydrogen, which cannot be accounted for by incomplete deuteration of c-C6D12 and n-C7D16. In order to elucidate the mechanism of the HD formation, we have examined the effect of additives and of physical conditions on the HD yield. By coating the vessel with a layer of Aquadag the HD yield is greatly decreased without appreciable variation of the total hydrogen yield. The HD yield from nontreated vessels also decreases remarkably with increasing total dose; it is increased upon the addition of traces of water; and is less affected by small amounts of c-C6H12. The addition of SF6 selectively reduces the HD yield. These experimental results lead us to conclude that inert hydrocarbon ions react by proton transfer with water desorbed from the vessel wall to give hydronium ions which yield HD on subsequent neutralization.  相似文献   

6.
The decomposition of the [C6H5CO]+ ions produced from eight alkyl benzoates by electron impact has been studied. By calculating the heat of formation of [C6H5CO]+ ions from the appearance potential value, it is shown that the ions from C6H5COOR when R?H, CH3, C2H5 have some excess energy, and those where R = n-C3H7, iso-C3H7, n-C4H9, iso-C4H9, iso-C5H11 are produced with more excess energy. It is also shown that by taking this excess energy into account, there is a linear relationship between the heat of formation of the activated complex produced in the reaction [C6H5CO]+→[C6H5]+ + CO and the vibrational degree of freedom of the neutral fragment ? OR.  相似文献   

7.
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   

8.
Infrared laser-induced fluorescence measurements of vibrational relaxation in cyclopropane are presented. Following laser excitation of the CH-stretch vibrations υ6 and υ8 time-dependent fluorescence signals from υ10 and υ711 were recorded. Activation and deactivation rate constants for C3H6C3H6 collisions were found for υ10 and υ711. A simplified model for the vibrational relaxation of cyclopropane is discussed.  相似文献   

9.
The collision-induced dissociation (CID) spectra of five alkylmethyleneimmonium ions (H2C-N+R1R2, (a) R1 = R2 = C2H5, (b) R1 = n-C3H7, R2 = H, (c) R1 = n-C3H7, R2 = CH3, (d) R1 = n-C3H7, R2 = C2H5, (e) R1 = R2 = n-C3H7) are reported and discussed in terms of the mechanism of alkane loss. The most abundant alkane losses result from 2-azaallylic bond cleavages within R1 and R2 leading to daughter ions of m/z 84. Ion d (R1 = n-C3H7, R2 = C2H5) was chosen for a deuterium-labelling study because it exhibited methane loss nearly free from interferences with other fragmentations. The methane lost consists to a great extent (95%) of the methyl moiety of R2. Whereas the methyl moiety obviously stays intact during the fragmentation process, the hydrogen additionally needed originates from all positions of R1 and the double-bonded methylene in an approximately random distribution, suggesting extensive hydrogen migrations preceding the transfer step.  相似文献   

10.
Six examples of newly synthesized α,α’-bis (aryl)-2,3:5,6-bis (pentame thylene)pyridyliron complexes [2,3:5,6-{C4H8C(NAr)}2C5HN]FeCl2 (Ar = 2-(c-C5H9)-6-MeC6H3 Fe1 , 2-(c-C6H11)-6-MeC6H3 Fe2 , 2-(c-C8H15)-6-MeC6H3 Fe3 , 2-(c-C5H9)-4,6-Me2C6H2 Fe4 , 2-(c-C6H11)-4,6-Me2C6H2 Fe5 , 2-(c-C8H15)-4,6-Me2C6H2 Fe6 ; c refers as cyclic), on activation with methylalumoxane (MAO) or modified MAO (MMAO), exhibit high activities towards ethylene polymerization, producing strictly linear polyethylenes with terminal vinyl groups. The catalytic performances are systematically investigated along with various polymerization parameters as well as the microstructures of resultant polyethylenes. The steric hindrances of ortho-cycloalkyl substituents of Nimino-aryl groups significantly affect the activities of the corresponding iron precatalysts as well as the microstructures of resultant polyethylenes: higher steric hindrance the ortho-cycloalkyl substituents, higher activity the iron precatalyst, lower molecular weight the resultant polyethylenes. Experimental observations are additionally supported by the computational study. The resultant polyethylenes exhibited excellent hydrophobicity.  相似文献   

11.
Isothermal three-phase equilibria of gas, aqueous, and hydrate phases for the {xenon (Xe) + cyclopropane (c-C3H6)} mixed-gas hydrate system were measured at two different temperatures (279.15 and 289.15) K. The structural phase transitions from structure-I to structure-II and back to structure-I, depending on the mole fraction of guest mixtures, occur in the (Xe + c-C3H6) mixed-gas hydrate system. The isothermal pressure–composition relations have two local pressure minima. The most important characteristic in the (Xe + c-C3H6) mixed-gas hydrate system is that the equilibrium pressure–composition relations exhibit the complex phase behavior involving two structural phase transitions and two homogeneous negative azeotropes. One of two structural phase transitions exhibits the heterogeneous azeotropic-like behavior.  相似文献   

12.
We have determined the differences in the parameters log A and E of the Arrhenius equations for the kinetic isotope effect (KIE) (c-C6H12/c-C6D12) and the 5/6 effect (c-C5H10/c-C6H12) in reactions of the C—H bonds of cycloalkanes with adamantyl (Ad+) carbocations (1-adamantanol in 92.8% H2SO4, 40-97 °C). We have established the compensation relations between log A and E for the kinetic isotope effect and the 5/6 effect for anthracene (AH+), hydroxymethyl (CH2OH+), Ad+ carbocations and the hypothetical "infinitely strong reagent," supporting a hydride transfer mechanism in such reactions.  相似文献   

13.
Abstract

The calcium vanadyl tartrate complex [Ca(VO)(d,l-C4H2O6)(H2O)4] has been synthesized and characterized by spectroscopic methods. Its crystal structure was solved by X-ray methods. The compound is monoclinic, space group P21/c, with a = 8.0282(5), b = 17.1568(8), c = 7.6113(3)Å, β = 94.269(4)° and Z = 4. The structure consists of centrosymmetric vanadyl tartrate dimers, [(VO)(d,l-C4H2O6)]2 4-, and calcium cations placed between them. As a result, dimers form chains in the [101] direction. Neighbouring chains are linked by the coordination of the calcium ion to the oxygen atom of a vanadyl group of a different chain, thus forming a two-dimensional structure. Different layers are linked by hydrogen bonds. Spectroscopic studies show the existence of intra-dimeric interactions between vanadium atoms.  相似文献   

14.
Kinetic modelling is used in conjunction with measurements of product yields to develop a mechanism for the pyrolysis of ethylene at 896K and ethylene pressures ranging from approximately 3 to 78 kPa. An induction period was observed for all products except H2, and was followed by a steady rate, which was of second-order for all products except 1,3-C4H6, the most abundant product. The mechanism quantitatively accounts for the yields of H2, CH4, C2H6, C3H6, 1-C4H8 and 1,3-C4H6. The reaction is initiated by disproportionation of C2H4 and the product 1,3-C4H6 results from decomposition of the C4H7 radical, formed by addition of C2H3 to C2H4. The other organic products that were measured are formed as a result of reactions involving the C2H5 radical. The hydrogen is produced by abstraction from C2H4 by atomic hydrogen and its rate is controlled by the reaction C2H5 → C2H4 + H which is nearly equilibrated. The main termination reaction is recombination of C2H5. The auto-acceleration which is evident particularly in the yields of H2, CH4, C2 H6, and C3H6 is accounted for by the decomposition of 1-C4H8. © 1996 John Wiley & Sons Inc.  相似文献   

15.
A gas-chromatographic procedure was developed for determining impurities (CH4, C2H6, C3H8, C4H10, iso-C4H10, C5H12, iso-C5H12, neo-C5H12, CH3Cl, C2H5Cl, CH2Cl2, CHCl3, CO, and CO2) in hydrogen chloride using two columns and a column switching technique in an isothermal mode with a flame ionization detector; the detection limits were 0.01–0.1 ppm. The matrix was separated in a precolumn packed with urea. CO and CO2 were determined by reaction gas chromatography with their conversion into methane.  相似文献   

16.
The deactivation of I(2P½) by R-OH compounds (R = H, CnH2n+1) was studied using time-resolved atomic absorption at 206.2 nm. The second-order quenching rate constants determined for H2O, CH3OH, C2H5OH, n-C3H7OH, i-C3H7OH, n-C4H9OH, i-C4H9OH, s-C4H9OH, t-C4H9OH, are respectively, 2.4 ± 0.3 × 10−12, 5.5 ± 0.8 × 10−12, 8 ± 1 × 10−12, 10 ± 1 × 10−12, 10 ± 1 × 10−12, 11.1 ± 0.9 × 10−12, 9.8 ± 0.9 × 10−12, 7.1 ± 0.7 × 10−12, and 4.1 ± 0.4× 10−12 cm3 molec−1 s−1 at room temperature. It is believed that a quasi-resonant electronic to vibrational energy transfer mechanism accounts for most of the features of the quenching process. The influence of the alkyl group and its role in the total quenching rate is also discussed. © 1997 John Wiley & Sons, Inc.  相似文献   

17.
The hydrogen bonding complexes formed between the H2O and OH radical have been completely investigated for the first time in this study using density functional theory (DFT). A larger basis set 6‐311++G(2d,2p) has been employed in conjunction with a hybrid density functional method, namely, UB3LYP/6‐311++G(2d,2p). The two degenerate components of the OH radical 2Π ground electronic state give rise to independent states upon interaction with the water molecule, with hydrogen bonding occurring between the oxygen atom of H2O and the hydrogen atom of the OH radical. Another hydrogen bond occurs between one of the H atoms of H2O and the O atom of the OH radical. The extensive calculation reveals that there is still more hydrogen bonding form found first in this investigation, in which two or three hydrogen bonds occur at the same time. The optimized geometry parameter and interaction energy for various isomers at the present level of theory was estimated. The infrared (IR) spectrum frequencies, IR intensities, and vibrational frequency shifts are reported. The estimates of the H2O · OH complex's vibrational modes and predicted IR spectra for these structures are also made. It should be noted that a total of 10 stationary points have been confirmed to be genuine minima and transition states on the potential energy hypersurface of the H2O · HO system. Among them, four genuine minima were located. © 2002 Wiley Periodicals, Inc. Int J Quantum Chem, 2002  相似文献   

18.
《Chemical physics letters》1987,136(2):187-191
The reactions of F atoms with C2H5I, C2F5I, and n-C3H5I were studied by the crossed beam laser-induced fluorescence techniques within the 570–620 nm wavelength region. The vibrational and rotational excitation spectra of the reaction product IF were measured. The relative vibrational population densities of v = 3,4, and 5 vibrational levels, and some of the relative detailed vibrational rate constants, the rotational temperatures, and the mean fractions of rotational energy in individual vibrational states of the reaction product IF were obtained. The reaction mechanism was discussed.  相似文献   

19.
Permeability, solubility, and diffusion coefficients have been determined for cyclopropane (c-C3H6) in silicone rubber at temperatures between ?8 and 70°C at relative pressures from 0.04 to 0.30. The permeability coefficients, , are of the order of 10?6 cm3 (STP) · cm/(s · cm2 · cmHg). increases slightly with increasing penetrant pressure and decreases with increasing temperature, the energy of activation for permeation being ?1.27 kcal/gmol at zero pressure. The solubility of cyclopropane in silicone rubber can be represented over the experimental concentration range by the Flory-Huggins equation. The solubility decreases with increasing temperature and the partial molar heat of solution is ?4.95 kcal/gmol. The solubility coefficient in the Henry's law limit, S(0), for cyclopropane and many other gases and vapors can be correlated with (Tc/T)2, where T and Tc are the experimental and critical temperatures, respectively. The mutual diffusion coefficients, D, increase with increasing concentration and temperature, the energy of activation for diffusion being 3.68 kcal/gmol. The pressure dependence of &\[\bar P\] is described satisfactorily by a free-volume model proposed by Fujita and extended by Stern, Frisch, and coworkers. The permeability, diffusion, and solubility behavior of cyclopropane in silicone rubber is similar to that of propane (C3H8).  相似文献   

20.
Reactions of Fe+ and FeL+ [L=O, C4H6, c-C5H6, C5H5, C6H6, C5H4(=CH2)] with thiophene, furan, and pyrrole in the gas phase by using Fourier transform mass spectrometry are described. Fe+, Fe(C5H5)+, and FeC6H 6 + yield exclusive rapid adduct formation with thiophene, furan, and pyrrole. In addition, the iron-diene complexes [FeC4H 6 + and Fe(c-C5H6)+], as well as FeC5H4(=CH2)+ and FeO+, are quite reactive. The most intriguing reaction is the predominant direct extrusion of CO from furan by FeC4H6 +, Fe(c-C5H6)+, and FeC5H4(=CH2)+. In addition, FeC4H 6 + and Fe(c-C5H6)+ cause minor amounts of HCN extrusion from pyrrole. Mechanisms are presented for these CO and HCN extrusion reactions. The absence of CS elimination from thiophene may be due to the higher energy requirements than those for CO extrusion from furan or HCN extrusion from pyrrole. The dominant reaction channel for reaction of Fe(c-C5H6)+ with pyrrole and thiophene is hydrogen-atom displacement, which implies DO(Fa(N5H5)+-C4H4X)>DO(Fe(C5H5)+-H)=46±5 kcal mol?1. DO(Fe+-C4H4S) and DO(Fe+-C4H5N)=DO(Fe+-C4H6)=48±5 kcal mol?1. Finally, 55±5 kcal mol?1=DO(Fe+-C6H6)>DO(Fe+-C4H4O)>DO(Fe+-C2H4)=39.9±1.4 kcal mol?1. FeO+ reacts rapidly with thiophene, furan, and pyrrole to yield initial loss of CO followed by additional neutral losses. DO(Fe+-CS)>DO(Fe+-C4H4S)≈48±5 kcal mol?1 and DO(Fe+-C4H5N)≈48±5 kcal mol?1>DO(Fe+-HCN)>DO(Fe+-C2H4)=39.9±1.4 kcal mil?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