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 15 毫秒
1.
By employing time‐resolved Fourier transform infrared emission spectroscopy, the fragments HCl (v=1–3), HBr (v=1), and CO (v=1‐3) are detected in one‐photon dissociation of 2‐bromopropionyl chloride (CH3CHBrCOCl) at 248 nm. Ar gas is added to induce internal conversion and to enhance the fragment yields. The time‐resolved high‐resolution spectra of HCl and CO were analyzed to determine the rovibrational energy deposition of 10.0±0.2 and 7.4±0.6 kcal mol?1, respectively, while the rotational energy in HBr is evaluated to be 0.9±0.1 kcal mol?1. The branching ratio of HCl(v>0)/HBr(v>0) is estimated to be 1:0.53. The bond selectivity of halide formation in the photolysis follows the same trend as the halogen atom elimination. The probability of HCl contribution from a hot Cl reaction with the precursor is negligible according to the measurements of HCl amount by adding an active reagent, Br2, in the system. The HCl elimination channel under Ar addition is verified to be slower by two orders of magnitude than the Cl elimination channel. With the aid of ab initio calculations, the observed fragments are dissociated from the hot ground state CH3CHBrCOCl. A two‐body dissociation channel is favored leading to either HCl+CH3CBrCO or HBr+CH2CHCOCl, in which the CH3CBrCO moiety may further undergo secondary dissociation to release CO.  相似文献   

2.
The thermal decomposition of formaldehyde was investigated behind shock waves at temperatures between 1675 and 2080 K. Quantitative concentration time profiles of formaldehyde and formyl radicals were measured by means of sensitive 174 nm VUV absorption (CH2O) and 614 nm FM spectroscopy (HCO), respectively. The rate constant of the radical forming channel (1a), CH2O + M → HCO + H + M, of the unimolecular decomposition of formaldehyde in argon was measured at temperatures from 1675 to 2080 K at an average total pressure of 1.2 bar, k1a = 5.0 × 1015 exp(‐308 kJ mol?1/RT) cm3 mol?1 s?1. The pressure dependence, the rate of the competing molecular channel (1b), CH2O + M → H2 + CO + M, and the branching fraction β = k1a/(kA1a + k1b) was characterized by a two‐channel RRKM/master equation analysis. With channel (1b) being the main channel at low pressures, the branching fraction was found to switch from channel (1b) to channel (1a) at moderate pressures of 1–50 bar. Taking advantage of the results of two preceding publications, a decomposition mechanism with six reactions is recommended, which was validated by measured formyl radical profiles and numerous literature experimental observations. The mechanism is capable of a reliable prediction of almost all formaldehyde pyrolysis literature data, including CH2O, CO, and H atom measurements at temperatures of 1200–3200 K, with mixtures of 7 ppm to 5% formaldehyde, and pressures up to 15 bar. Some evidence was found for a self‐reaction of two CH2O molecules. At high initial CH2O mole fractions the reverse of reaction (6), CH2OH + HCO ? CH2O + CH2O becomes noticeable. The rate of the forward reaction was roughly measured to be k6 = 1.5 × 1013 cm3 mol?1 s?1. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 157–169 2004  相似文献   

3.
Laser flash photolysis coupled with resonance fluorescence detection of Br atoms was employed to investigate the temperature dependence of the reaction Br + neo‐C5H12 (1) between 688 and 775 K. The following Arrhenius preexponential factor and activation energy were determined (±1 σ): A1 = (6.89 ± 2.27) 1014 cm3 mol−1 s−1 and EA,1 = 57.61 ± 2.05 kJ mol1 The only other kinetic parameters reported for the reaction of Br atoms with neo‐C5H12 were obtained from competitive kinetic experiments relative to Br + C2H6. Comparison with our direct results is hampered by uncertainties in the kinetic data for the reference reaction that may need reinvestigation. The standard enthalpy of formation for the neo‐C5H11 radical was estimated to be 34.7 and 41.6 kJ mol−1, depending on the value of the activation energy assumed for the reverse reaction neo‐C5H11 + HBr (−1). © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 33: 49–55, 2001  相似文献   

4.
Upon irradiation with ultraviolet wavelengths, Fe2(S2C3H6)(CO)6, a simple model of the [FeFe]‐hydrogenase active site, undergoes CO dissociation to form the unsaturated Fe2(S2C3H6)(CO)5 species and successively a solvent adduct at the vacant coordination site. In the present work, the CO‐photolysis of Fe2(S2C3H6)(CO)6 was investigated by density functional theory (DFT) and time‐dependent DFT (TDDFT). Trans Fe2(S2C3H6)(CO)5 form and the corresponding trans heptane or acetonitrile solvent adducts are the lowest energy ground state forms. CO dissociation barriers computed for the lowest triplet state are roughly halved with respect to those for the ground state suggesting that some low‐lying excited potential energy surface (PES) could be loosely bound with respect to Fe? C bond cleavage. The TDDFT excited state PESs and geometry optimizations for the excited states likely involved in the CO‐photolysis suggest that the Fe? S bond elongation and the partial isomerization toward the rotated form could take place simultaneously, favoring the trans CO photodissociation. © 2014 Wiley Periodicals, Inc.  相似文献   

5.
UV photolysis (λ=248 or 255 nm) of cyclic S2N2 isolated in solid argon matrices yields two open‐shell S2N2 isomers, trans SNSN (3A′′) and cis SNSN (3A′′), as well as a closed‐shell C2v dimer (SN)2 (1A1). These novel isomers have been characterized by their IR spectra and mutual photo‐interconversion reactions. Quantum chemical calculations support the experimental results and also provide insight into the complex potential energy surface of S2N2.  相似文献   

6.
The kinetics of C6H5 reactions with n‐CnH2n+2 (n = 3, 4, 6, 8) have been studied by the pulsed laser photolysis/mass spectrometric method using C6H5COCH3 as the phenyl precursor at temperatures between 494 and 1051 K. The rate constants were determined by kinetic modeling of the absolute yields of C6H6 at each temperature. Another major product C6H5CH3 formed by the recombination of C6H5 and CH3 could also be quantitatively modeled using the known rate constant for the reaction. A weighted least‐squares analysis of the four sets of data gave k (C3H8) = (1.96 ± 0.15) × 1011 exp[?(1938 ± 56)/T], and k (n‐C4H10) = (2.65 ± 0.23) × 1011 exp[?(1950 ± 55)/T] k (n‐C6H14) = (4.56 ± 0.21) × 1011 exp[?(1735 ± 55)/T], and k (n?C8H18) = (4.31 ± 0.39) × 1011 exp[?(1415 ± 65)T] cm3 mol?1 s?1 for the temperature range studied. For the butane and hexane reactions, we have also applied the CRDS technique to extend our temperature range down to 297 K; the results obtained by the decay of C6H5 with CRDS agree fully with those determined by absolute product yield measurements with PLP/MS. Weighted least‐squares analyses of these two sets of data gave rise to k (n?C4H10) = (2.70 ± 0.15) × 1011 exp[?(1880 ± 127)/T] and k (n?C6H14) = (4.81 ± 0.30) × 1011 exp[?(1780 ± 133)/T] cm3 mol?1 s?1 for the temperature range 297‐‐1046 K. From the absolute rate constants for the two larger molecular reactions (C6H5 + n‐C6H14 and n‐C8H18), we derived the rate constant for H‐abstraction from a secondary C? H bond, ks?CH = (4.19 ± 0.24) × 1010 exp[?(1770 ± 48)/T] cm3 mol?1 s?1. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 36: 49–56, 2004  相似文献   

7.
The relative energies of 11 [C3H3O]+ ions are calculated by different molecular orbital methods (MINDO/3, MNDO, ab initio with 3-21G and 4-31G* basis set and configuration interaction). The four most stable structures are: a ([CH2?CH? CO]+), b c ([CH?C? CHOH]+) and d ([CH2?C?COH]+); their relative energies at the CI/4-31G*//3-21G level are 0, 117, 171 and 218 kJ mol?1, respectively. The isomerizations c→[CH?CH? CHO]+→[CH2?C? CHO]+a and dissociations into [C2H3]++CO and [HCO]++C2H2 are explored. The calculated potential energy profile reveals that the energy-determining step is the 1,3-H migration c→[CH?CH? CHO]+. This explains the value of unity of the branching ratio and the spread of kinetic energy released for the two dissociation channels.  相似文献   

8.
The gas‐phase reaction mechanism between methane and rhodium monoxide for the formation of methanol, syngas, formaldehyde, water, and methyl radical have been studied in detail on the doublet and quartet state potential energy surfaces at the CCSD(T)/6‐311+G(2d, 2p), SDD//B3LYP/6‐311+G(2d, 2p), SDD level. Over the 300–1100 K temperature range, the branching ratio for the Rh(4F) + CH3OH channel is 97.5–100%, whereas the branching ratio for the D‐CH2ORh + H2 channel is 0.0–2.5%, and the branching ratio for the D‐CH2ORh + H2 channel is so small to be ruled out. The minimum energy reaction pathway for the main product methanol formation involving two spin inversions prefers to both start and terminate on the ground quartet state, where the ground doublet intermediate CH3RhOH is energetically preferred, and its formation rate constant over the 300–1100 K temperature range is fitted by kCH3RhOH = 7.03 × 106 exp(?69.484/RT) dm3 mol?1 s?1. On the other hand, the main products shall be Rh + CH3OH in the reactions of RhO + CH4, CH2ORh + H2, Rh + CO +2H2, and RhCH2 + H2O, whereas the main products shall be CH2ORh + H2 in the reaction of Rh + CH3OH. Meanwhile, the doublet intermediates H2RhOCH2 and CH3RhOH are predicted to be energetically favored in the reactions of Rh + CH3OH and CH2ORh + H2 and in the reaction of RhCH2 + H2O, respectively. © 2009 Wiley Periodicals, Inc. J Comput Chem 2010  相似文献   

9.
The kinetics of C6H5 reactions with C2H6 (1) and neo‐C5H12 (2) have been studied by the pulsed laser photolysis/mass spectrometric method using C6H5COCH3 as the phenyl precursor at temperatures between 565 and 1000 K. The rate constants were determined by kinetic modeling of the absolute yields of C6H6 at each temperature. Another major product, C6H5CH3, formed by the recombination of C6H5 and CH3, could also be quantitatively modeled using the known rate constant for the reaction. A weighted least‐squares analysis of the two sets of data gave k1 = 1011.32±0.05 exp[−(2236 ± 91)/T] cm3 mol−1 s−1 and k2 = 1011.37±0.03 exp[−(1925 ± 48)/T] cm3 mol−1 s−1 for the temperature range studied. The result of our sensitivity analysis clearly supports that the yields of C6H6 and C6H5CH3 depend primarily on the abstraction reactions and C6H5 + CH3, respectively. From the absolute rate constants for the two reactions we obtained the value for the H‐abstraction from a primary C‐H bond, k‐CH = 1010.40±0.06 exp(−1790 ± 102/T) cm3 mol−1 s−1. This result is utilized for analysis of other kinetic data measured for C6H5 reactions with alkanes in solution as well as in the gas phase. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 33: 64–69, 2001  相似文献   

10.
Using density functional calculations, we demonstrate that the planarity of the nonclassical planar tetracoordinate carbon (ptC) arrangement can be utilized to construct new families of flat, tubular, and cage molecules which are geometrically akin to graphenes, carbon nanotubes, and fullerenes but have fundamentally different chemical bonds. These molecules are assembled with a single type of hexagonal blocks called starbenzene (D6h C6Be6H6) through hydrogen‐bridge bonds that have an average bonding energy of 25.4–33.1 kcal mol?1. Starbenzene is an aromatic molecule with six π electrons, but its carbon atoms prefer ptC arrangements rather than the planar trigonal sp2 arrangements like those in benzene. Various stability assessments indicate their excellent stabilities for experimental realization. For example, one starbenzene unit in an infinite two‐dimensional molecular sheet lies on average 154.1 kcal mol?1 below three isolated linear C2Be2H2 (global minimum) monomers. This value is close to the energy lowering of 157.4 kcal mol?1 of benzene relative to three acetylene molecules. The ptC bonding in starbenzene can be extended to give new series of starlike monocyclic aromatic molecules (D4h C4Be4H42?, D5h C5Be5H5?, D6h C6Be6H6, D7h C7Be7H7+, D8h C8Be8H82?, and D9h C9Be9H9?), known as starenes. The starene isomers with classical trigonal carbon sp2 bonding are all less stable than the corresponding starlike starenes. Similarly, lithiated C5Be5H5 can be assembled into a C60‐like molecule. The chemical bonding involved in the title molecules includes aromaticity, ptC arrangements, hydrogen‐bridge bonds, ionic bonds, and covalent bonds, which, along with their unique geometric features, may result in new applications.  相似文献   

11.
The absolute bimolecular rate constants for the reactions of C6H5 with 2‐methylpropane, 2,3‐dimethylbutane and 2,3,4‐trimethylpentane have been measured by cavity ringdown spectrometry at temperatures between 290 and 500 K. For 2‐methylpropane, additional measurements were performed with the pulsed laser photolysis/mass spectrometry, extending the temperature range to 972 K. The reactions were found to be dominated by the abstraction of a tertiary C H bond from the molecular reactant, resulting in the production of a tertiary alkyl radical: C6H5 + CH(CH3)3 → C6H6 + t‐C4H9 (1) (1) C6H5 + (CH3)2CHCH(CH3)2 → C6H6 + t‐C6H13 (2) (2) C6H5 + (CH3)2CHCH(CH3)CH(CH3)2 → C6H6 + t‐C8H17 (3) (3) with the following rate constants given in units of cm3 mol−1 s−1: k1 = 10(11.45 ± 0.18) e−(1512 ± 44)/T k2 = 10(11.72 ± 0.15) e−(1007 ± 124)/T k3 = 10(11.83 ± 0.13) e−(428 ± 108)/T © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 645–653, 1999  相似文献   

12.
Five examples of unsymmetrical 1,2‐bis (arylimino) acenaphthene ( L1 – L5 ), each containing one N‐2,4‐bis (dibenzocycloheptyl)‐6‐methylphenyl group and one sterically and electronically variable N‐aryl group, have been used to prepare the N,N′‐nickel (II) halide complexes, [1‐[2,4‐{(C15H13}2–6‐MeC6H2N]‐2‐(ArN)C2C10H6]NiX2 (X = Br: Ar = 2,6‐Me2C6H3 Ni1 , 2,6‐Et2C6H3 Ni2 , 2,6‐i‐Pr2C6H3 Ni3 , 2,4,6‐Me3C6H2 Ni4 , 2,6‐Et2–4‐MeC6H2 Ni5 ) and (X = Cl: Ar = 2,6‐Me2C6H3 Ni6 , 2,6‐Et2C6H3 Ni7 , 2,6‐i‐Pr2C6H3 Ni8 , 2,4,6‐Me3C6H2 Ni9 , 2,6‐Et2–4‐MeC6H2 Ni10 ), in high yield. The molecular structures Ni3 and Ni7 highlight the extensive steric protection imparted by the ortho‐dibenzocycloheptyl group and the distorted tetrahedral geometry conferred to the nickel center. On activation with either Et2AlCl or MAO, Ni1 – Ni10 exhibited very high activities for ethylene polymerization with the least bulky Ni1 the most active (up to 1.06  ×  107 g PE mol?1(Ni) h?1 with MAO). Notably, these sterically bulky catalysts have a propensity towards generating very high molecular weight polyethylene with moderate levels of branching and narrow dispersities with the most hindered Ni3 and Ni8 affording ultra‐high molecular weight material (up to 1.5  ×  106 g mol?1). Indeed, both the activity and molecular weights of the resulting polyethylene are among the highest to be reported for this class of unsymmetrical 1,2‐bis (imino)acenaphthene‐nickel catalyst.  相似文献   

13.
Third-order Møller–Plesset perturbation theory (MP 3) with a 6-31G** basis set was applied to study the relative stabilities of H+(X)2 conformations (X ? CO and N2) and their clustering energies. The effect of both basis set extensions and electron correlation is not negligible on the relative stabilities of the H+(CO)2 clusters. The most stable conformation of H+(CO)2 is found to be a Cv structure in which a carbon atom of CO bonds to the proton of H+(CO), whereas that of H+(N2)2 is a symmetry Dh structure. The second lowest energy conformations of H+(CO)2 and H+(N2)2 lie within 2 kcal/mol above the energies of the most stable structures. Clustering energies computed using MP 3 method with the 6-31G** basis set are in good agreement with the experimental findings of Hiraoka, Saluja, and Kebarle. The low-lying singlet conformations of H+(X)3 (X ? CO and N2) have been studied by the use of the Hartree–Fock MO method with the 6-31G** basis set and second-order Møller–Plesset perturbation theory with a 4-31G basis set. The most stable structure is a T-shaped structure in which a carbon atom of CO (or a nitrogen atom of N2) attacks the proton of the most stable conformation of H+(X)2 clusters.  相似文献   

14.
Butadiene cation radicals are produced symmetrically from the ring and side-chain of the vinylcyclohexene cation radical near the onset of the fragmentation. The appearance energies of C4H6+? and C4H2D4+? from (3,3,6,6-D4)vinylcyclohex ene were measured as 11.07 ± 0.05 and 11.06 ± 0.06 eV, respectively. This sets the barrier to retro-Diels-Alder decomposition at 1140 kJ mol?1 above the energy of 1 and 44 kJ mol?1 above the thermochemical threshold corresponding to C4H6+? + C4H6. Topological molecular orbital calculations indicate that this lowest-energy path involves a sequential rupture of the C3C4 and C5C6 bonds, with a calculated barrier of 211 kJ mol?1. The second, two-step reaction channel proceeds by subsequent fission of the C5C6 and C3C4 bonds with a barrier of 299 kJ mol?1. This channel is found experimentally as a break on the ionization efficiency curve at 12.1 eV. Both the supra-supra and the supra-antara pericyclic reactions go through energy maxima and are therefore forbidden. The supra-supra process is the most favorable route for decomposition from the first excited state, the activation energy being 333 kJ mol?1. The preference for the two-step mechanism is due to hyperconjugative stabilization of intermediate molecular configurations.  相似文献   

15.
The pyrolysis of anisole (C6H5OCH3) was studied behind reflected shock waves via highly sensitive absorption measurements of CO concentration using a rotational transition in the fundamental vibrational band near 4.7 µm. Time‐resolved CO mole fractions were monitored in shock‐heated C6H5OCH3/Ar mixtures between 1000 and 1270 K at 1.3–1.6 bar. The decomposition of C6H5OCH3 proceeds exclusively via homolytic dissociation, with reaction rate k 1, forming methyl (CH3) and phenoxy (C6H5O) radicals. The subsequent decomposition of C6H5O by ring rearrangement and bond dissociation yields CO. To determine the rate constant k 2 of C6H5O decomposition avoiding secondary reactions, allyl phenyl ether (C6H5OC3H5) was used as an alternative source for C6H5O. Its decomposition was studied between 970 and 1170 K at ∼1.4 bar. The potential‐energy surface of C6H5O dissociation has been reevaluated at the G4 level of theory. Rate constants determined from unimolecular rate theory are in good agreement with the present experiments. However, the obtained rates k 2 = 9.1 × 1013 exp(−220.3 kJ mol−1/RT )s−1 are significantly higher than those reported before (factor 6, 2, and 1.5 faster than those data reported by Lin and Lin, J. Phys. Chem . 1986, 90, 425–431; Frank et al., 1994; Carstensen and Dean, 2012, respectively). Good agreement was found between the measured CO concentration profiles and simulations based on the mechanism of Nowakowska et al. after substituting k 2 by the value obtained from experiments on C6H5OC3H5 in this work. The bimolecular reaction of C6H5O and CH3 toward cresol was identified as the most important reaction influencing the CO concentration at longer reaction time.  相似文献   

16.
The vulcanization of rubber by sulfur is a large‐scale industrial process that is only poorly understood, especially the role of zinc oxide, which is added as an activator. We used the highly symmetrical cluster Zn4O4 (Td) as a model species to study the thermodynamics of the initial interaction of various vulcanization‐related molecules with ZnO by DFT methods, mostly at the B3LYP/6‐31+G* level. The interaction energy of Lewis bases with Zn4O4 increases in the following order: CO62H43H62S2<1,4‐C5H82O2S3N?CH3COO?. The corresponding binding energies range from ?57 to ?262 kJ mol?1. However, Brønsted acids react with the Zn4O4 cluster with proton transfer from the ligand molecule to one of the oxygen atoms of Zn4O4, and these reactions are all strongly exothermic [binding energies [kJ mol?1] in parentheses: H2O (?183), MeOH (?171), H2S (?245), MeSH (?230), C3H6 (?121), and CH3COOH (?255)]. The important vulcanization accelerator mercaptobenzothiazole (C7H5NS2, MBT) containing several donor sites reacts with the Zn4O4 cluster with proton transfer from the NH group to one of the oxygen atoms of ZnO, and in addition the exocyclic thiono sulfur atom and the nitrogen atom coordinate to one and the same zinc atom, resulting in a binding energy of ?247 kJ mol?1. A second isomer of [(MBT)Zn4O4] with a strong O? H???N hydrogen bond rather than a Zn? N bond is only slightly less stable (binding energy ?243 kJ mol?1). The NH form of free MBT is 36 kJ mol?1 more stable than the tautomeric SH form, while the sulfurized MBT derivative benzothiazolyl hydrodisulfide C7H5NS3 (BtSSH) is most stable with the connectivity >CSSH.  相似文献   

17.
Palladacyclic compounds [Pd(C6H4(C6H5C?O)C?N? R)(N? N)] [X] (R = Et, iPr, 2,6‐iPr2C6H3; N? N = bpy = 2,2′‐bipyridine, or 1,4‐(o,o′‐dialkylaryl)‐1,4‐diazabuta‐1,3‐dienes; [X]? = [BF4]? or [PF6]?) were synthesized from the dimers [{Pd(C6H4(C6H5C?O)C?N? R)(μ‐Cl)}2] and N? N ligands. Their interionic structure in CD2Cl2 was determined by means of 19F,1H‐HOESY experiments and compared with that in the solid state derived from X‐ray single‐crystal studies. [Pd(C6H4(C6H5C?O)C?N? R)(N? N)] [X] complexes were found to copolymerize CO and p‐methylstyrene affording syndiotactic or isotactic copolymers when bpy or 1,4‐(o,o′‐dimethylaryl)‐1,4‐diazabuta‐1,3‐dienes were used, respectively. The reactions with CO and p‐methylstyrene of the bpy derivatives were investigated. Two intermediates derived from a single and a double insertion of CO into the Pd? C bonds were isolated and completely characterized in solution.  相似文献   

18.
A new series of palladium complexes ( Pd1–Pd5 ) ligated by symmetrical 2,3‐diiminobutane derivatives, 2,3‐bis[2,6‐bis{bis(4‐FC6H4)2CH}2‐4‐(alkyl)C6H2N]C4H6 (alkyl = Me L1 , Et L2 , i Pr L3 , t Bu L4 ) and 2,3‐bis[2,6‐bis{bis(C6H5)2CH}2‐4‐{(CH3)3C}C6H2N]C4H6 L5 , have been prepared and well characterized, and their catalytic scope toward ethylene polymerization have been investigated. Upon activation with MAO, all palladium complexes ( Pd1–Pd5) exhibited good activities (up to 1.44 × 106 g (PE) mol?1(Pd) h?1) and produced higher molecular weight polyethylene in the range of 105 g mol?1 with precise molecular weight distribution (M w/M n = 1.37–1.77). One of the long‐standing limiting features of the Brookhart type α‐diimine Pd(II) catalysts is that they produce highly branched (ca. 100/1000 C atoms) and totally amorphous polymer. Conversely, herein Pd5 produced polymers having dramatically lower branching number (28/1000) as well as improved melting temperature up to 73.1 °C showing well‐controlled linear architecture, and very similar to polyethylene materials generated by early‐transition‐metal based catalysts. © 2017 Wiley Periodicals, Inc. J. Polym. Sci., Part A: Polym. Chem. 2017 , 55 , 3214–3222  相似文献   

19.
The 2‐aminobenzothiazole sulfonation intermediate 2,3‐dihydro‐1,3‐benzothiazol‐2‐iminium monohydrogen sulfate, C7H7N2S+·HSO4, (I), and the final product 2‐iminio‐2,3‐dihydro‐1,3‐benzothiazole‐6‐sulfonate, C7H6N2O3S2, (II), both have the endocyclic N atom protonated; compound (I) exists as an ion pair and (II) forms a zwitterion. Intermolecular N—H...O and O—H...O hydrogen bonds are seen in both structures, with bonding energy (calculated on the basis of density functional theory) ranging from 1.06 to 14.15 kcal mol−1. Hydrogen bonding in (I) and (II) creates DDDD and C(8)C(9)C(9) first‐level graph sets, respectively. Face‐to‐face stacking interactions are observed in both (I) and (II), but they are extremely weak.  相似文献   

20.
A new type of reaction pathway which involves a nontotally symmetric trifurcation was found and investigated for a typical SN2‐type reaction, NC + CH3X → NC? CH3 + X (X = F, Cl). A nontotally symmetric valley‐ridge inflection (VRI) point was located along the C3v reaction path. For X = F, the minimum energy path (MEP) starting from the transition state (TS) leads to a second‐order saddle point with C3v symmetry, which connects three product minima of Cs symmetry. For X = Cl, four product minima have been observed, of which three belong to Cs symmetry and one to C3v symmetry. The branching path from the VRI point to the lower symmetry minima was determined by a linear interpolation technique. The branching mechanism is discussed based on the reaction path curvature and net atomic charges, and the possibility of a nonotally symmetric n‐furcation is discussed. © 2015 Wiley Periodicals, Inc.  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