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.
[2+2] Cycloaddition reactions of P2 with alkenes were predicted to have concerted paths, that is, pseudoexcitation, distorted 2πs+2πs, and 2πs+2πa processes without any interventions of intermediates. The pseudoexcitation and/or distorted 2πs+2πs paths with retention of configuration of alkenes are kinetically preferred to the 2πs+2πa path with inversion of configuration. The reactions were predicted from the appreciable difference in the calculated enthalpies of activation to be stereospecific.  相似文献   

2.
Klaus Wittel 《Tetrahedron》1977,33(20):2687-2689
A theoretical analysis is presented for the [2s + 2s] cyclodimerisation of molecules in which the lowest unoccupied MO is of Rydberg nature. The concerted mechanism is forbidden by the Woodward-Hoffmann rules for the thermal and the photochemical path. However, the frontier MO approach predicts an enhancement of these reactions. Experiments related to this analysis are discussed.  相似文献   

3.
(2+2)- and (4+2)-cycloadducts are formed by thermal reaction of xanthenethione with allenes. Activation parameters for ths “concerted two step” (π2s+π2s+π2s) reaction are presented.  相似文献   

4.
Thermolysis of the anthracene photooxide 1a, in solution, can afford successively three dimers (I→II→III) the structures of which have been established by NMR and X-ray analysis for the first one. Their origin is the unstable o-quinodimethane diether 3a, coming from 1a by isomenzation, which at 80° dimerizes to 8a (dimer I) by an unusual [π8s+π6s] concerted cycloaddition. Above 110°, 8a isomerizes into 9a (dimer II) by a concerted [1,5] suprafacial sigmatropic migration. Finally, at higher temperature (~200°), both dimers are partly converted into the symetrical dimer 10a (dimer III), probably by a radical pathway. Thermolysis runned in presence of N-methylmaleimide show the non-reversibility of these reactions, giving two adducts, 15a and 16a, derived from 9a, the structures of which are deduced from NMR data. An explanation of the nature and high selectivity of the proceeding processes is given by considering orbital factors.  相似文献   

5.
The water exchange reactions in aquated Li+ and Be2+ ions were investigated with density functional theory calculations performed using the [Li(H2O)4]+·14H2O and [Be(H2O)4]2+·8H2O systems and a cluster‐continuum approach. A range of commonly used functionals predict water exchange rates several orders of magnitude lower than the experimental ones. This effect is attributed to the overstabilization of coordination number four by these functionals with respect to the five‐coordinated transition states responsible for the associative ( A ) or associative interchange ( Ia ) water exchange mechanisms. However, the M06 and M062X functionals provide results in good agreement with the experimental data: M062X/TZVP calculations yield a concerted Ia mechanism for the water exchange in [Be(H2O)4]2+·8H2O that gives an average residence time of water molecules in the first coordination sphere of 260 μs. For [Li(H2O)4]+·14H2O the water exchange reaction is predicted to follow an A mechanism with a residence time of inner‐sphere water molecules of 25 ps.  相似文献   

6.
Three possible reaction mechanisms of methanoyl fluoride with 2H2O include a concerted and a stepwise hydrolysis of HFCO into HCOOH + HF, and a pure catalytic decomposition of HFCO into HF + CO. Among these, the two H2O molecules acting as catalyst to decompose HFCO has the lowest calculated barrier, 25.1 kcal/mol with respect to the reactant‐adduct complex, whereas the barriers for the concerted and stepwise hydrolytic reactions in which one H2O acts as a reactant and the other H2O as catalyst are similar, 30.8 kcal/mol for concerted and 29.9 kcal/mol for stepwise. The formation of transoid HCOOH in the hydrolysis of HFCO is more favorable than cisoid HCOOH.  相似文献   

7.
The reaction between Ar2+ and C2H2 has been studied, at centre-of-mass collision energies ranging from 3 to 7 eV, using a position-sensitive coincidence technique to detect the monocation pairs, which are formed. Sixteen different reaction channels generating pairs of monocations have been observed, these channels arise from double-electron-transfer, single-electron-transfer and chemical reactions forming ArC+. Examination of the scattering diagrams and energetic information extracted from the coincidence data indicate that double-electron-transfer is a direct process, which does not involve a collision complex, and the derived energetics point towards a concerted, not stepwise, mechanism for the two-electron-transfer. As is commonly observed, single-electron-transfer from C2H2 to Ar2+ takes place via a direct mechanism, again not involving complexation. Most of the C2H2+ products that are formed in the single-electron-transfer reactions possess significant (12–15 eV) internal energy and fragment rapidly within the electric field of the partner Ar+ ion. The chemical reactions appear to proceed via a direct mechanism involving the initial formation of ArCH+, which subsequently fragments to form ArC+.  相似文献   

8.
The reactions of diethyl 4‐nitrophenyl phosphate ( 1 ) with a series of nucleophiles: phenoxides, secondary alicyclic (SA) amines, and pyridines are subjected to a kinetic study. Under excess of nucleophile, all the reactions obey pseudo‐first‐order kinetics and are first order in the nucleophile. The nucleophilic rate constants (kN) obtained are pH independent for all the reactions studied. The Brønsted‐type plot (log kN vs. pKa nucleophile) obtained for the phenolysis is linear with slope β=0.21; no break was found at pKa 7.5, consistent with a concerted mechanism. The Brønsted‐type plots for the SA aminolysis and pyridinolysis are linear with slopes β=0.39 and 0.43, respectively, also suggesting concerted processes. The concerted mechanisms for the latter reactions are proposed on the basis of the lack of break in the Brønsted‐type plots and the instability of the hypothetical pentacoordinate intermediates formed in these reactions. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 708–714, 2011  相似文献   

9.
Here we employ density functional theory calculations to systematically investigate the underlying mechanism of Cu(OTf)2-catalyzed [3+2] cycloaddition reactions in the synthesis of CF3-substituted pyrazolidines. About eight possible initial configurations of the [3+2] reaction is considered, and all relevant reactants, transition states, and products are optimized. Based on these structures, internal reaction coordinate paths, and wavefunction analysis results, we conclude that the Cu(OTf)2-catalyzed [3+2] cycloaddition follows a concerted asynchronous mechanism. The C N bond forms immediately after the formation of the C C bond. Among the eight reaction paths, the energy barrier for the [3+2] reaction that leads to the CF3-substituted syn-pyrazolidine is the lowest, ∼5.4 kcal/mol, which might result in the diastereoselectivity that is observed in the experiment. This work not only gives the detailed mechanism of the Cu(OTf)2-catalyzed [3+2] cycloaddition but can also be helpful for the future designation of Cu(OTf)2-based cycloaddition processes.  相似文献   

10.
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   

11.
The temperature dependence of the rate constant for the reactions of HO2 with OH, H, Fe2+ and Cu2+ has been determined using pulse radiolysis technique. The following rate constants, k (dm3 mol−1 s−1) at 20°C and activation energies, Ea (kJ mol−1) have been found. The reaction with OH was studied in the temperature range 20–296°C (k=7.0×109, Ea=7.4) and the reaction with H in the temperature range 5–149°C (k=8.5×109, Ea=17.5). The reaction with Fe2+ was studied in the temperature range 16–118°C (k=7.9×105, Ea=36.8) and the reaction with Cu2+ in the temperature range 17–211°C (k=1.1×108, Ea=14.9).  相似文献   

12.
Mn1 + 2sCr2 ? 3sSbsO4, a new series of spinels, have been prepared and studied using X-ray powder data. For s going from 0.05 to 0.30, a gradually increases from 8.441(1) to 8.472(1) Å, and u slightly decreases, from 0.262 to 0.258. Interatomic distances are given. The Mn1 + 2sCr2 ? 3sSbsO4 (0.05 < s < 0.30) series may be conceived as the result of partial substitution of Cr3 + by 2/3Mn2 + + 1/3Sb5 + in the normal spinel, MnCr2O 4.  相似文献   

13.
On the Chemical Transport of Cr2O3 with Cl2 and with HgCl2 — Experiments and Model Calculations The migration of Cr2O3 in a temperature gradient (1 000°C → 900°C) in the presence of low concentrations of chlorine and water (from the wall of silica ampoules) is a result from the endothermic reactions (1) Cr2O3,s + H2Og + 3 Cl2,g = 2 CrO2Cl2,g + 2 HClg (2) Cr2O3,s + 1/2 O2,g + 2 Cl2,g = 2 CrO2Cl2,g With higher concentrations of chlorine, the transport reaction is (3) Cr2O3,s + 5/2 Cl2,g = 3/2 CrO2Cl2,g + 1/2 CrCl4,g The gas phase of the transport system Cr2O3/Cl2 can be reduced step by step by adding small amounts of chromium, so that CrCl3 and finally also CrCl2 become more important. Further, at a lower ratio n°(Cl)/n°(Cr) three transport reactions have to be taken into consideration; with the participation of CrOCl2,g (5). (4) Cr2O3,s + 9/2 CrCl4,g = 3/2 CrO2Cl2,g + 5 CrCl3,g (5) Cr2O3,s + 3 CrCl4,g = 3 CrOCl2,g + 2 CrCl3,g (6) Cr2O3,s + H2,g + 4 HClg = 2 CrCl2,g + 3 H2Og The reactions (1), (2) and (6) become possible through the cooperation of two transport agents at a time. The migration of Cr2O3 with HgCl2 can also be described with reactions (1) – (3). The decomposition of HgCl2 Produces the small chlorine pressure for the transport reaction. The oxidation potential of the transport agent HgCl2 is too low for the oxidation of CrIII to CrVI.  相似文献   

14.
A number of model Diels‐Alder (D‐A) cycloaddition reactions (H2C?CH2 + cyclopentadiene and H2C?CHX + 1,3‐butadiene, with X = H, F, CH3, OH, CN, NH2, and NO) were studied by static (transition state ‐ TS and IRC) and dynamics (quasiclassical trajectories) approaches to establish the (a)synchronous character of the concerted mechanism. The use of static criteria, such as the asymmetry of the TS geometry, for classifying and quantifying the (a)synchronicity of the concerted D‐A reaction mechanism is shown to be severely limited and to provide contradictory results and conclusions when compared to the dynamics approach. The time elapsed between the events is shown to be a more reliable and unbiased criterion and all the studied D‐A reactions, except for the case of H2C?CHNO, are classified as synchronous, despite the gradual and quite distinct degrees of (a)symmetry of the TS structures. © 2015 Wiley Periodicals, Inc.  相似文献   

15.
The chemoselectivities in the (π4s + π2s) cycloaddition reactions of tetraphenylcyclopentadienone (TPCD) with ethyl- and 1,1-dimethylallene have been determined and are compared with those observed for reaction with maleic anhydride.  相似文献   

16.
The [2s + 2a] cycloaddition of ethylene and acetylene has been studied. A transition structure of C2 symmetry was located on the potential surface. Activation energies for the process are also reported.  相似文献   

17.
An experimental study of the kinetics of the B2H6/O (3P) system at room temperature, is presented. Modeling was based on a multiple-parameter fitting process to a complex kinetic mechanism. The aim of the study was to propose and evaluate a preferred set of elementary reactions which might be important for the oxidation of boron-hydrides. In this study, relative concentration-vs.-time profiles of the radicals OH and BO2 were measured in a low pressure flow reactor, by the technique of laser-induced-fluorescence. A range of almost two orders of magnitude in the initial fuel to oxygen ratio was covered, while the residence times of the gases in the reactor were up to 1 s. A comprehensive kinetic mechanism was constructed from the available measured and estimated data in literature. After the fitting process and dropping all of the reactions with negligible contribution, a 46 elementary reactions mechanism was obtained. In this mechanism, 27 reactions are not measured and their rate coefficients were used as the adjustable (within reasonable limits deduced from kinetic and thermodynamic considerations) parameters of the fitting process. Good agreement was obtained between all of the measured and calculated profiles. From sensitivity analysis it was found that only a limited group of these reactions highly contributes to the calculated concentrations of OH and BO2. With only the highly contributing reactions and all the oxygenhydrogen reactions which are measured and relatively well known, a 30 elementary reactions mechanism was obtained. In this mechanism only 13 reactions are not measured and the agreement between the measured and the calculated profiles is still reasonable. To demonstrate the possible usefulness of the proposed mechanism, the rate coefficient of the reaction B2H6 + OH → B2H5 + H2O which was not measured before, was directly measured in our experimental set-up. The rate coefficient that was obtained in the direct measurement is (3.3 ± 1.1) × 1011 cm3mol?1s?1, in excellent agreement with the predicted one by the multiple-parameter-fitting process, which was 2 × 1011 cm3mol?1s?1. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
The thermally and radiolytically induced chain decomposition of methanesulfonyl chloride (MeSO2Cl) in liquid cyclohexane (RH) was studied at 150°C. The main products, chlorocyclohexane, sulfur dioxide, and methane, are formed in almost equal yields, and a relatively small amount of methyl chloride is also observed. The formation and addition of SO2 strongly inhibit the chain decomposition reaction. By kinetic analysis it is shown that the formation of the main products can be explained only in terms of a mechanism that postulates the decomposition of MeSO2, and that the alternative mechanism of methane and SO2 formation via the methanesulfinic acid is inconsistent with the kinetic behavior of the system. For the reactions Me + MeSO2Cl → MeCl + MeSO2(2b), Me + RH → MeH + R (4), and Me + CCl4 → MeCCl4 → MeCl + CCl3 (15b) the following rate constant ratios are determined; k2b/k4=2.17±0.20 and k2b/k15b=2.63±0.52. For the reactions R + MeSO2Cl → RCl + MeSO2(2a) and R + CCl4 → RCl + CCl3 (15a), k2a/k15a is equal to 1.55±0.05. In addition the equilibrium constant K7 for the reaction R + SO2 ? RSO2 (7) is estimated as being equal to (9.4±3) × 103 mole/l.  相似文献   

19.
A low‐pressure discharge‐flow system equipped with laser‐induced fluorescence (LIF) detection of NO2 and resonance‐fluorescence detection of OH has been employed to study the self reactions CH2ClO2 + CH2ClO2 → products (1) and CHCl2O2 + CHCl2O2 → products (2), at T = 298 K and P = 1–3 Torr. Possible secondary reactions involving alkoxy radicals are identified. We report the phenomenological rate constants (kobs) k1obs = (4.1 ± 0.2) × 10−12 cm3 molecule−1 s−1 k2obs = (8.6 ± 0.2) × 10−12 cm3 molecule−1 s−1 and the rate constants derived from modelling the decay profiles for both peroxy radical systems, which takes into account the proposed secondary chemistry involving alkoxy radicals k1 = (3.3 ± 0.7) × 10−12 cm3 molecule−1 s−1 k2 = (7.0 ± 1.8) × 10−12 cm3 molecule−1 s−1 A possible mechanism for these self reactions is proposed and QRRK calculations are performed for reactions (1), (2) and the self‐reaction of CH3O2, CH3O2 + CH3O2 → products (3). These calculations, although only semiquantitative, go some way to explaining why both k1 and k2 are a factor of ten larger than k3 and why, as suggested by the products of reaction (1) and (2), it seems that the favored reaction pathway is different from that followed by reaction (3). The atmospheric fate of the chlorinated peroxy species, and hence the impact of their precursors (CH3Cl and CH2Cl2), in the troposphere are briefly discussed. HC(O)Cl is identified as a potentially important reservoir species produced from the photooxidation of these precursors. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 433–444, 1999  相似文献   

20.
The kinetics and mechanism of reduction of aqueous toluidine blue (TB+) by phenyl hydrazine (Pz), which exhibits nonlinear behavior, is studied spectrophotometrically at 630 nm. Typical kinetic curves exhibited autocatalytic characteristics. The role of H+ as an autocatalyst is established. Rate constants for the uncatalyzed and acid catalyzed reactions are determined. The forward rate constants for the uncatalyzed and acid catalyzed reactions were 1.4 × 10−2 M−1 s−1 and 60 M−1 s−1. Reaction products are toluidine white, phenol, and an azo dye. From the stoichiometric ratios, the major reaction is Pz + 2 TB+ + H2O = PhOH + 2 TBH + 2 H+ + N2. The rate expression and a detailed 12‐step reaction mechanism supported by simulations are proposed. ©1999 John Wiley & Sons, Inc. Int J Chem Kinet: 31: 83–88, 1999  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