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 0 毫秒
1.
The Cl-transfer reaction between CFCl3 and c-C6H11 radicals (R) was studied in liquid cyclohexane (RH). The Arrhenius parameters for Cl abstraction were determined in the RH-CFCl3 system versus the termination reaction between cyclohexyl radicals and competitively versus addition to C2Cl4 in the RH-CFCl3-C2Cl4 system. The two sets of results are in very good agreement and give the following Arrhenius expression for the reaction R + CFCL3 → RCl + CFCl2 (2): where θ = 2.303RT in kcal/mol. Comparison with Cl-transfer data of other chloromethanes and chloroethanes shows that an increase in the C? Cl bond dissociation energy is the main cause of the reduced reactivity of CFCl3. Based on a previously developed correlation, D(CFCl2 ? Cl) is estimated to be equal to 74.4 kcal/mol.  相似文献   

2.
The kinetics of the thermal elimination of HF from 1,2-difluoroethane have been studied in a static system over the temperature range 734–820°K. The reaction was shown to be first order and homogeneous, with a rate constant of where θ = 2.303RT in kcal/mole. The A-factor falls within the normal range for such reactions and is in line with transition state theory; the activation energy is similarly consistent with an estimate based on data for the analogous reactions of ethyl fluoride and other alkyl halides. The above activation energy has been compared with values of the critical energy calculated from data on the decomposition of chemically activated 1,2-difluoroethane by the RRKM theory and the bond dissociation energy, D(CH2F? CH2F) = 88 ± 2 kcal/mole, derived. It follows from thermochemistry that ΔHf0(CH2F) = -7.8 and D(CH2F? H) = 101 ± 2 kcal/mole. Bond dissociation energies in fluoromethanes and fluoroethanes are discussed.  相似文献   

3.
Density functional theory studies of the series of isomeric d(6) (pentacarbonyl)metal complexes (CO)(5)M(eta(1)-SO(2))(nq), (CO)(5)M(eta(1)-OSO)(nq)(), and (CO)(5)M(eta(2)-SO(2))(nq) (M = Ti-Hf, nq = 2-; M = V-Ta, nq = 1-; M = Cr -W, nq = 0; M = Mn-Re, nq = 1+; M = Fe-Os, nq = 2+) provide accurate structural modeling and quantitative prediction of the relative stabilities of the isomers. The eta(1)-S-bound complexes display planar SO(2) moieties that adopt staggered orientations with respect to the carbonyl ligands, in keeping with experimental observations. The OSO chain in the eta(1)-O-bound complexes generally adopts the u-shape with a staggered orientation. The dianions (CO)(5)(Ti-Hf)(eta(1)-OSO)(2-) differ in that the OSO chain adopts the eclipsed z-shape orientation. The eta(2)-SO(2) complexes exhibit a facial interaction and are stable only for anionic and neutral complexes, supporting the view that this motif involves substantial M --> SO(2) pi-back-bonding. The relative stabilities of the isomers generally follow u-shaped trends both across a row and down a family. This fits with qualitative ideas that the bond dissociation energies (BDEs) for the (CO)(5)M(SO(2))(nq) complexes track competition between relative hardness/softness of the metal fragment and its capacity for pi-back-bonding. Quantitatively, examination of BDEs by bond energy decomposition approaches suggests that electrostatic considerations dominate bonding for the eta(1)-SO(2) complexes and covalent effects dominate for the eta(2)-SO(2) species, while both are important for eta(1)-OSO complexes.  相似文献   

4.
5.
The bond dissociation energy (D(0)) of the water dimer is determined by using state-to-state vibrational predissociation measurements following excitation of the bound OH stretch fundamental of the donor unit of the dimer. Velocity map imaging and resonance-enhanced multiphoton ionization (REMPI) are used to determine pair-correlated product velocity and translational energy distributions. H(2)O fragments are detected in the ground vibrational (000) and the first excited bending (010) states by 2 + 1 REMPI via the C? (1)B(1) (000) ← X? (1)A(1) (000 and 010) transitions. The fragments' velocity and center-of-mass translational energy distributions are determined from images of selected rovibrational levels of H(2)O. An accurate value for D(0) is obtained by fitting both the structure in the images and the maximum velocity of the fragments. This value, D(0) = 1105 ± 10 cm(-1) (13.2 ± 0.12 kJ/mol), is in excellent agreement with the recent theoretical value of D(0) = 1103 ± 4 cm(-1) (13.2 ± 0.05 kJ∕mol) suggested as a benchmark by Shank et al. [J. Chem. Phys. 130, 144314 (2009)].  相似文献   

6.
The acidities of the two different sites in naphthalene (1alpha and 1beta) and the electron affinities of the alpha- and beta-naphthyl radicals were measured using a Fourier transform mass spectrometer. Both carbon-hydrogen bond dissociation energies for naphthalene also were obtained, in this case via the application of a thermodynamic cycle. The final results are DeltaH(o)acid (1alpha) = 394.2+/-1.2 kcal mol(-1), DeltaH(o)acid (1beta) = 395.5+/-1.3 kcal mol(-1), EA(alpha) = 31.6+/-0.5 kcal mol(-1), EA(beta) = 31.6+/-0.5 kcal mol(-1), BDE(1alpha) = 112.2+/-1.3 kcal mol(-1) and BDE(1alpha) = 111.9+/-1.4 kcal mol(-1), and they are compared to benzene and phenyl radical as well as ab initio and density functional theory (B3LYP) calculations.  相似文献   

7.
Crystal Structures of the Silylated Phosphaneimines Me3SiNP(c-C6H11)3 and (Me3SiNPPh2)2CH2 The crystal structures of Me3SiNP(c-C6H11)3 ( 1 ) and (Me3SiNPPh2)2CH2 ( 2 ) are determined by X-ray diffraction at single crystals. In both compounds the PN distances correspond to double bonds, the SiN distances to single bonds. With 149.8° the SiNP bond angle in 1 is noticeably large, while it is only 138.5° in 2 , which shows C2 symmetry. 1 : Space group P21/n, Z = 4, lattice dimensions at –70 °C: a = 1143.0(1), b = 1743.0(2), c = 1152.5(1) pm, β = 90.42(1)°; R = 0.0677. 2 : Space group I41/a, Z = 8, lattice dimensions at –60 °C: a = b = 1959.7(1), c = 1695.8(1) pm, R = 0.0433.  相似文献   

8.
Using a recent, full-dimensional, ab initio potential energy surface [Y. Wang, X. Huang, B. C. Shepler, B. J. Braams, and J. M. Bowman, J. Chem. Phys. 134, 094509 (2011)] together with rigorous diffusion Monte Carlo calculations of the zero-point energy of the water trimer, we report dissociation energies, D(0), to form one monomer plus the water dimer and three monomers. The calculations make use of essentially exact zero-point energies for the water trimer, dimer, and monomer, and benchmark values of the electronic dissociation energies, D(e), of the water trimer [J. A. Anderson, K. Crager, L. Fedoroff, and G. S. Tschumper, J. Chem. Phys. 121, 11023 (2004)]. The D(0) results are 3855 and 2726 cm(-1) for the 3H(2)O and H(2)O + (H(2)O)(2) dissociation channels, respectively, and 4206 and 2947 cm(-1) for 3D(2)O and D(2)O + (D(2)O)(2) dissociation channels, respectively. The results have estimated uncertainties of 20 and 30 cm(-1) for the monomer plus dimer and three monomer of dissociation channels, respectively.  相似文献   

9.
The bond dissociation energies for losing one water from Cd(2+)(H(2)O)(n) complexes, n = 3-11, are measured using threshold collision-induced dissociation in a guided ion beam tandem mass spectrometer coupled with a thermal electrospray ionization source. Kinetic energy dependent cross sections are obtained for n = 4-11 complexes and analyzed to yield 0 K threshold measurements for loss of one, two, and three water ligands after accounting for multiple collisions, kinetic shifts, and energy distributions. The threshold measurements are converted from 0 to 298 K values to give the hydration enthalpies and free energies for sequentially losing one water from each complex. Theoretical geometry optimizations and single point energy calculations are performed on reactant and product complexes using several levels of theory and basis sets to obtain thermochemistry for comparison to experiment. The charge separation process, Cd(2+)(H(2)O)(n) → CdOH(+)(H(2)O)(m) + H(+)(H(2)O)(n-m-1), is also observed for n = 4 and 5 and the competition between this process and water loss is analyzed. Rate-limiting transition states for the charge separation process at n = 3-6 are calculated and compared to experimental threshold measurements resulting in the conclusion that the critical size for this dissociation pathway of hydrated cadmium is n(crit) = 4.  相似文献   

10.
Quantum mechanical density functional theory (DFT) and coupled DFT/molecular mechanics (QMMM) studies of the compounds (H(3)P)(3)M(eta(1)-SO(2)) and (Me(n)Ph(3-n)P)(3)M(eta(1)-SO(2)) (M = Ni, Pd, Pt; n = 0-3) model the experimental data well, particularly the planar/pyramidal geometries at sulfur. Bond dissociation energy (BDE) calculations confirm that Pd/Pt systems with pyramidal SO(2) ligands exhibit M-S BDEs smaller by 30-50% than Ni systems with planar SO(2). However, scans of the potential energy surfaces show that flexing the planar/pyramidal torsion angle within ranges of 20-30 degrees requires little energy. Bond energy decomposition calculations indicate that the electrostatic Delta E(elstat) term determines the BDE for Pd/Pt molecules where the sulfur is pyramidal, whereas all three terms matter when the sulfur is planar, as for Ni compounds. However, this accounts only for a fraction of the BDE differences; orbital energy matching accounts for the balance.  相似文献   

11.
The title reaction has been investigated in the temperature range of 490-573 K. Initial reactant pressures were varied in the range of 0.2-5.2 torr (I2) and 2-20 torr (C6H5SiH3). The rate of iodine consumption, monitored spectrophotometrically, was found to obey both by initial rate and integrated equation fitting procedures. The effect of added initial HI conformed to this expression. The data are consistent with a conventional I-atom propagated chain reaction, and for the step the rate constant is given by From this is derived the bond dissociation energy value C6H5SiH2? H = 374 kJ/mol(88 kcal/mol). A comparison with other Si? H dissociation energy values indicates that the “silabenzyl” stabilization energy is small, ≈7 kJ/mol.  相似文献   

12.
A laser-induced fluorescence spectrum was observed in the 500-560 nm region when a mixture of 1,4-cyclohexadiene and oxalyl chloride was photolyzed at 193 nm. The observed excitation spectrum was assigned to the A (2)A(2)<--X (2)B(1) transition of the cyclohexadienyl radical c-C6H7, produced by abstraction of a hydrogen atom from 1,4-cyclohexadiene by Cl atoms. The origin of the A<--X transition of c-C(6)H(7) was at 18 207 cm(-1). From measurements of the dispersed fluorescence spectra and ab initio calculations, the frequencies of several vibrational modes in both the ground and excited states of c-C(6)H(7) were determined: nu(5)(C-H in-plane bend)=1571, nu(8)(C-H in-plane bend)=1174, nu(10)(C-C-C in-plane bend)=981, nu(12)(C-C-C in-plane bend)=559, nu(16)(C-C-C out-of-plane bend)=375, and nu(33)(C-C-C in-plane bend)=600 cm(-1) for the ground state and nu(8)=1118, nu(10)=967, nu(12)=502, nu(16)=172, and nu(33)=536 cm(-1) for the excited states.  相似文献   

13.
The reactions of CF3 radicals with the C3 to C7 cyclanes and spiropentane were studied and the following Arrhenius parameters were obtained for the reaction CF3 + c-RH → CF3H + c-R:
c-RH log A (cm3mole?1sec?1) E (kcal/mole) D(c-R—H) (kcal/mole)
Cyclopropane 11.54 8.73 100.7
Cyclobutane 11.66 6.48 95.7
Cyclopentane 12.30 6.18 94.3
Cyclohexane 12.12 6.26 94.9
Cycloheptane 12.43 5.89 94.0
Spiropentane 11.91 8.12 98.8
The CF3 radicals were generated by photolysis of hexafluoroacetone or CF3I and a comparison is made of the utility of the two compounds as radical sources. The Arrhenius parameters are compared with those for corresponding reactions of CH3 radicals with cyclanes, and the general reactivity of cyclic compounds toward free radicals is discussed. An Evans-Polanyi treatment is used to derive C? H bond dissociation energies in cyclanes; and these results, based on the reactions of CF3 with cyclanes, agree well with those previously obtained using CH3 plus cyclanes. The final mean values are shown above.  相似文献   

14.
15.
The reactions (R = CF3, C2F5, and i,-C3F7) have been studied competitively in the gas phase over the range of 27–231°C. The following Arrhenius parameters were obtained:
log A,(cm3/mol · s) E,(kJ/mol)
R = CF3 13.99 ± 0.21 17.1 ± 2.0
R = C2F5 13.97 ± 0.20 11.5 ± 2.0
R = i,-C3F7 14.18 ± 0.20 10.2 ± 2.0
The above data lead to bond dissociation energies D(R-I) which are compared with previous published results. The following values are recommended: D,(CF3-I) = 224, D,(C2F5-I) = 219, and D,(i,-C3F7-I) = 215 kJ/mol.  相似文献   

16.
Pincer complexes of the type [2,6-(R(2)PO)(2)C(6)H(3)]NiSC(6)H(4)Z (R = Ph and i-Pr; Z = p-OCH(3), p-CH(3), H, p-Cl, and p-CF(3)) have been synthesized from [2,6-(R(2)PO)(2)C(6)H(3)]NiCl and sodium arylthiolate. X-ray structure determinations of these thiolate complexes have shown a somewhat constant Ni-S bond length (approx. 2.20 ?) but an almost unpredictable orientation of the thiolate ligand. Equilibrium constants for various thiolate exchange (between a nickel thiolate complex and a free thiol, or between two different nickel thiolate complexes) reactions have been measured. Evidently, the thiolate ligand with an electron-withdrawing substituent prefers to bond with "[2,6-(Ph(2)PO)(2)C(6)H(3)]Ni" rather than "[2,6-(i-Pr(2)PO)(2)C(6)H(3)]Ni", and bonds least favourably with hydrogen. The reactions of the thiolate complexes with halogenated compounds such as PhCH(2)Br, CH(3)I, CCl(4), and Ph(3)CCl have been examined and several mechanistic pathways have been explored.  相似文献   

17.
A model for taking into account surface temperature effects in molecule-surface reactions is reported and applied to the dissociation of H(2) and D(2) on Cu(111). In contrast to many models developed before, the model constructed here takes into account the effects of static corrugation of the potential energy surface rather than energy exchange between the impinging hydrogen molecule and the surface. Such an approximation is a vibrational sudden approximation. The quality of the model is assessed by comparison to a recent density functional theory study. It is shown that the model gives a reasonable agreement with recently performed ab initio molecular dynamics calculations, in which the surface atoms were allowed to move. The observed broadening of the reaction probability curve with increasing surface temperature is attributed to the displacement of surface atoms, whereas the effect of thermal expansion is found to be primarily a shift of the curve to lower energies. It is also found that the rotational quadrupole alignment parameter is generally lowered at low energies, whereas it remains approximately constant at high energies. Finally, it is shown that the approximation of an ideal static surface works well for low surface temperatures, in particular for the molecular beams for this system (T(s) = 120 K). Nonetheless, for the state-resolved reaction probability at this surface temperature, some broadening is found.  相似文献   

18.
The reaction of copper(I) iodide with 1, 3-imidazolidine-2-thione (SC3H6N2) in a 1:2 molar ratio (M/L) has formed unusual 1D polymers, {Cu6(mu3-SC3H6N2)4(mu-SC3H6N2)2(mu-I)2I4}n (1) and {Cu6(mu3-SC3H6N2)2(mu-SC3H6N2)4(mu-I)4I2}n (1a). A similar reaction with copper(I) bromide has formed a polymer {Cu6(mu3-SC3H6N2)2(mu-SC3H6N2)4(mu-Br)4Br2}n (3a), similar to 1a, along with a dimer, {Cu2(mu-SC3H6N2)2(eta1-SC3H6N2)2Br2} (3). Copper(I) chloride behaved differently, and only an unsymmetrical dimer, {Cu2(mu-SC3H6N2)(eta1-SC3H6N2)3Cl2} (4), was formed. Finally, reactions of copper(I) thiocyanate in 1:1 or 1:2 molar ratios yielded a 3D polymer, {Cu2(mu-SC3H6N2)2(mu-SCN)2}n (2). Crystal data: 1, C9H18Cu3I3N6S3, triclinic, P, a = 9.6646(11) A, b = 10.5520(13) A, c = 12.6177(15) A, alpha = 107.239(2) degrees , beta = 99.844(2) degrees , gamma = 113.682(2) degrees , V = 1061.8(2) A(3), Z = 2, R = 0.0333; 2, C(4)H(6)CuN(3)S(2), monoclinic, P2(1)/c, a = 7.864(3) A, b = 14.328(6) A, c = 6.737(2) A, beta = 100.07(3) degrees , V = 747.4(5), Z = 4, R = 0.0363; 3, C12H24Br2Cu2N8S4, monoclinic, C2/c, a = 19.420(7) A, b = 7.686(3) A, c = 16.706(6) A, beta = 115.844(6) degrees , V = 2244.1(14) A(3), Z = 4, R = 0.0228; 4, C12H24Cl2Cu2N8S4, monoclinic, P2(1)/c, a = 7.4500(6) A, b = 18.4965(15) A, c = 16.2131(14) A, beta = 95.036(2) degrees , V = 2225.5(3) A(3), Z = 4, R = 0.0392. The 3D polymer 2 exhibits 20-membered metallacyclic rings in its structure, while synthesis of linear polymers, 1 and 1a, represents an unusual example of I (1a)-S (1) bond isomerism.  相似文献   

19.
We use protonated benzene (C(6)H(7)(+)) and cyclohexadienyl radical (c-C(6)H(7)) to demonstrate a new method that has some advantages over other methods currently used. C(6)H(7)(+) and c-C(6)H(7) were produced on electron bombardment of a mixture of benzene (C(6)H(6)) and para-hydrogen during deposition onto a target at 3.2 K. Infrared (IR) absorption lines of C(6)H(7)(+) decreased in intensity when the matrix was irradiated at 365 nm or maintained in the dark for an extended period, whereas those of c-C(6)H(7) increased in intensity. Observed vibrational wavenumbers, relative IR intensities, and deuterium isotopic shifts agree with those predicted theoretically. This method, providing a wide spectral coverage with narrow lines and accurate relative IR intensities, can be applied to larger protonated polyaromatic hydrocarbons and their neutral species which are difficult to study with other methods.  相似文献   

20.
Contributions to the Chemistry of Phosphorus. 99. A Heterocyclophosphane with a P4N Ring Skeleton: Synthesis and Properties of (PC6H5)4N(c-C6H11) The first cycloazaphosphane, (PC6H5)4N(c-C6H11) ( 1 ), is obtained by the reaction of N, N-dichlorocyclohexylamine with an excess of dipotassium phenylphosphide K—(PC6H5)n—K (n = 3, 4). Besides, considerable amounts of (PC6H5)5 are formed and occasionally some (PC6H5)4 and (PC6H5)6 can be found. 1 could be purely isolated in 12% yield. The compound is relatively stable against heating but light sensitive. It was characterized by elemental analysis, mass-, NMR, and vibrational spectra. The phenyl substituents at adjacent phosphorus atoms are arranged in trans position. After the reaction of K(R)P—P(R)K with N, N-dihalogenamines no clear indications for the formation of azadiphosphiranes with a P2N three-membered ring skeleton could be found.  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