共查询到20条相似文献,搜索用时 31 毫秒
1.
Rates of thermal gas‐phase elimination of eleven 2‐aryloxyacetic acid have been measured over a 45°C temperature range for each compound. Hammett correlation of the present kinetic data with the literature σ0 values of the given substituents gave a reaction ρ constant of 0.69 at 600 K; this is more than that for the gas‐phase elimination parameter of 2‐aryloxypropanoic acid (ρ = 0.26) and consistent with a transition state with some charge separation, suggesting a partial formation of carbocation. The implications of this observation for the thermal gas‐phase elimination of α‐aryloxycarboxylic acids are considered. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 612–616, 2001 相似文献
2.
The gas phase elimination kinetics of 2‐bromopropene was studied over the temperature range of 571–654 K and pressure range of 12–46 Torr using the seasoned static reaction system. Propyne was the only olefinic product formed and accounted for >98% of the reaction. This product was formed by homogeneous, unimolecular pathways with high‐pressure first‐order rate constant k∞ given by the equation k∞ = 1013.47 ± 0.6 exp?208.2 ± 6.7 (kJ mol?1)/RT. The error limits are 95% certainty limits. The observed Arrhenius parameters are consistent with the four centered activated complex. The presence of methyl group on α‐carbon lowers the activation energy by 41 kJ mol?1. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 39: 1–5, 2007 相似文献
3.
Elisey Mazilov Evgeniya Ogurtsova Alexander Shamov Grigorii Khrapkovskii 《International journal of quantum chemistry》2010,110(4):939-945
Using nonempirical methods and DFT‐methods the geometrical parameters formation enthalpies of molecules and radicals, energies dissociation of N? NO2 bonds of primary and secondary N‐nitramines have been investigated. The basic tendencies in the changes of the geometrical and electronic structures, formation enthalpies, and dissociation energies have been analyzed in basic homologous series of nitramines. Various alternative mechanisms of the gas‐phase monomolecular thermal decomposition have been studied by of the example of N‐methylnitramine. The process of the aci‐form formation and its further multistage destruction is the most advantageous way of decomposition of the primary N‐nitramines. © 2009 Wiley Periodicals, Inc. Int J Quantum Chem, 2010 相似文献
4.
R. Fournet F. Battin‐Leclerc P. A. Glaude B. Judenherc V. Warth G. M. Cme G. Scacchi A. Ristori G. Pengloan P. Dagaut M. Cathonnet 《国际化学动力学杂志》2001,33(10):574-586
Since n‐hexadecane or cetane is a reference fuel for the estimation of cetane numbers in diesel engines, a detailed chemical model of its gas‐phase oxidation and combustion will help to enhance diesel performance and reduce the emission of pollutants at their outlet. However, until recently the gas‐phase reactions of n‐hexadecane had not been experimentally studied, prohibiting a validation of oxidation models which could be written. This paper presents a modeling study of the oxidation of n‐hexadecane based on experiments performed in a jet‐stirred reactor, at temperatures ranging from 1000 to 1250 K, 1‐atm pressure, a constant mean residence time of 0.07 s, and high degree of nitrogen dilution (0.03 mol% of fuel) for equivalence ratios equal to 0.5, 1, and 1.5. A detailed kinetic mechanism was automatically generated by using the computer package (EXGAS) developed in Nancy. The long linear chain of this alkane necessitates the use of a detailed secondary mechanism for the consumption of the alkenes formed as a result of primary parent fuel decomposition. This high‐temperature mechanism includes 1787 reactions and 265 species, featuring satisfactory agreement for both the consumption of reactants and the formation of products. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 574–586, 2001 相似文献
5.
Libin Du 《Journal of fluorine chemistry》2008,129(10):1044-1046
Trifluoromethyl fluoroformate is prepared through a radical reaction between CF3OF and CO initiated by elemental fluorine. The reaction could be integrated into a continuous synthesis process. This method represents a convenient synthesis of the title compound under mild condition. 相似文献
6.
Edgar Marquez Tania Cordova Gabriel Chuchani 《International journal of quantum chemistry》2012,112(3):724-730
The mechanism of the gas‐phase elimination kinetics of 2‐ethoxypyridine has been studied through the electronic structure calculations using density functional methods: B3LYP/6‐31G(d,p), B3LYP/6‐31++G(d,p), B3PW91/6‐31G(d,p), B3PW91/6‐31++G(d,p), MPW1PW91/6‐31G(d,p), MPW1PW91/6‐31++G(d,p), PBEPBE/6‐31G(d,p), PBEPBE/6‐31++G(d,p), PBE1PBE1/6‐31G(d,p), and PBE1PBE1/6‐31++G(d,p). The elimination reaction of 2‐ethoxypyridine occurs through a six‐centered transition state geometry involving the pyridine nitrogen, the substituted carbon of the aromatic ring, the ethoxy oxygen, two carbons of the ethoxy group, and a hydrogen atom, which migrates from the ethoxy group to the nitrogen to give 2‐pyridone and ethylene. The reaction mechanism appears to occur with the participation of π‐electrons, similar to alkyl vinyl ether elimination reaction, with simultaneous ethylene formation and hydrogen migration to the pyridine nitrogen producing 2‐pyridone. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011 相似文献
7.
Annia Galano J. Raúl Alvarez‐Idaboy Armando Cruz‐Torres Ma. Esther Ruiz‐Santoyo 《国际化学动力学杂志》2003,35(5):212-221
Unrestricted density functional theory (BHandHLYP) calculations have been performed, using the 6‐311G(d,p) basis sets, to study the gas‐phase OH hydrogen abstraction reaction from methionine. The structures of the different stationary points are discussed. Ring‐like structures are found for all the transition states. Reaction profiles are modeled including the formation of prereactive complexes, and negative net activation energy is obtained for the gamma H‐abstraction channel. A complex mechanism is proposed, and the rate coefficients are calculated using transition state theory over the temperature range 250–350 K. The rate coefficients are proposed for the first time and it was found that in gas phase the hydrogen abstraction occurs almost exclusively from the gamma site. The large overall rate coefficient for the methionine + OH reaction compared to other free amino acids could explain the significant role of methionine in the oxidative processes. The following expressions in [L/(mol s)] are obtained for the alpha, beta, and gamma H‐abstraction channels, and for the overall temperature‐dependent rate constants, respectively: kα = (3.42 ± 0.11) × 108 exp[(?1118 ± 9)/T], kβ = (1.13 ± 0.03) × 108 exp[(?1070 ± 8)/T], kγ = (2.11 ± 0.26) × 107 exp[(2049 ± 34)/T], and ktot = (2.12 ± 0.26) × 107 exp[(2047 ± 34)/T]. © 2003 Wiley Periodicals, Inc. Int J Chem Kinet 35: 212–221, 2003 相似文献
8.
The study of the kinetics and mechanism of dehydrochlorination reaction of 2‐methyl benzyl chloride in the gas phase was carried out by means of electronic structure calculations using ab initio Móller‐Plesset MP2/6‐31G(d,p), and Density Functional Theory (DFT) methods: B3LYP/6‐31G(d,p), B3LYP/6‐31++G(d,p), MPW1PW91/6‐31G(d,p), MPW1PW91/6‐31++G(d,p)], PBE/6‐31G(d,p), PBE/6‐31++G(d,p). Investigated reaction pathways comprise: Mechanism I, a concerted reaction through a six‐centered cyclic transition state (TS) geometry; Mechanism II, a 1,3‐chlorine shift followed by beta‐elimination and Mechanism III, a single‐step elimination with simultaneous HCl and benzocyclobutene formation through a bicyclic type of TS. Calculated parameters ruled out Mechanism III and suggest the elimination reaction may occur by either unimolecular Mechanism I or Mechanism II. However, the TS of the former is 20 kJ/mole more stable than the TS of the latter. Consequently, the Mechanism I seem to be more probable to occur. The rate‐determining process is the breaking of C‐Cl bond. The involvement of π‐electrons of the aromatic system was demonstrated by NBO charges and bond order calculations. The reaction is moderately polar in nature. © 2011 Wiley Periodicals, Inc. Int J Chem Kinet 43: 537–546, 2011 相似文献
9.
Rates and products of reaction and Arrhenius activation parameters were determined for the gas‐phase thermolysis of 14 substrates of the title compounds using sealed pyrex reactor tubes and HPLC/UV‐VIS to monitor substrate pyrolysis. The 14 compounds under study are N‐phenyl‐3‐oxo‐ ( 1 ), N‐(p‐chlorophenyl)‐3‐oxo‐ ( 2 ), N‐(p‐methylphenyl)‐3‐oxo‐ ( 3 ), and N‐(p‐methoxyphenyl)‐3‐oxobutanamide ( 4 ), in addition to (i) four substrates ( 5–8 ) obtained by the replacement of the pairs of methylene hydrogens at the 2‐position of compounds ( 1–4 ), each pair by a phenylhydrazono group; (ii) three arylhydrazono derivatives ( 9–11 ) in which Cl, CH3, or OCH3 groups are substituted at the para position of the phenylhydrazono moiety of compound 5 ; (iii) 3‐oxobutanamide (acetoacetamide, 12 ), N‐phenyl‐3‐oxo‐3‐phenylpropanamide ( 13 ), and N,N′‐diphenylpropanediamide ( 14 ). The reactions were conducted over 374–546 K temperature range, and the values of the Arrhenius log A(s?1) and Ea(kJ mol?1) of these reactions were, respectively, 12.0 ± 2.0 and 119.2 ± 17.0 for the ketoanilides ( 1–4, 12–14 ), and 13.0 ± 0.7 and 157.5 ± 8.6 for the arylhyrazono compounds ( 5–11 ). Kinetically, the arylhydrazono derivatives were found to be ca. 1.4 × 103 to 5.7 × 103 times less reactive than the parent ketoanilides. A mechanism is proposed to account for reaction products and to rationalize molecular reactivities. © 2006 Wiley Periodicals, Inc. Int J Chem Kinet 39: 82–91, 2007 相似文献
10.
José R. Mora Edgar Márquez Jesus Lezama Tania Córdova Gabriel Chuchani 《International journal of quantum chemistry》2011,111(15):4011-4019
The kinetics and mechanisms of the gas‐phase elimination reactions of neopentyl chloride and neopentyl bromide have been studied by means of electronic structure calculations using density functional methods: B3LYP/6‐31G(d,p), B3LYP/ 6‐31++G(d,p), MPW1PW91/6‐31G(d,p), MPW1PW91/6‐31++G(d,p), PBEPBE/6‐31G(d,p), PBEPBE /6‐31++G(d,p). The reaction channels that account in products formation have a common first step involving a Wagner‐Meerwein rearrangement. The migration of the halide from the terminal carbon to the more substituted carbon is followed by beta‐elimination of HCl or HBr to give two olefins: the Sayzeff and Hoffmann products. Theoretical calculations demonstrated that these eliminations proceed through concerted asynchronous process. The transition state (TS) located for the rate‐determining step shows the halide detached and bridging between the terminal carbon and the quaternary carbon, while the methyl group is also migrating in a concerted fashion. The TS is described as an intimate ion‐pair with a large negative charge at the halide atom. The concerted migration of methyl group provides stabilization of the TS by delocalizing the electron density between the terminal carbon and the quaternary carbon. The B3LYP/6‐31++G(d,p) allows to obtain reasonable energies and enthalpies of activation. The nature of these reactions is examined in terms of geometrical parameters, electron distribution, and bond order analysis. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011 相似文献
11.
Xiaoli Wang Yongcheng Wang Shuang Li Yuwei Zhang 《International journal of quantum chemistry》2017,117(18)
To explore the details of the reaction mechanism of Ti atom with HCN, the reactive site and reactivity have been predicted first, the potential energy surfaces have been systematically studied at different theoretical levels. Four different reaction pathways and product distribution are discussed by means of the activation strain model and Curtin–Hammett principle. In addition, the structures, bonding properties and the frontier molecular orbital interaction diagrams of main stationary points were analyzed by atoms in molecules and natural bond orbital. The results show that for this system, there are four reaction pathways, in which path b (HCN+Ti→IM1→TS1→IM2→T2b→IM4) is the most favorable pathway. 相似文献
12.
The kinetics and mechanism of the oxidation of carbon by NO2 in absence and presence of water vapor were studied in a fixed bed reactor. The rate of carbon oxidation by NO2 is enhanced in the presence of water vapor in the range of temperature 300–400°C. The benefit effect of water is attributed to the intermediate formation of traces of nitric and nitrous acids, which enhance the rate of the carbon oxidation without modifying the global mechanism reaction. Therefore, water acts as a catalyst for the carbon oxidation by NO2. A kinetic mechanism derived from this parametric study shows a decrease in the activation energy of carbon oxidation by NO2 in the presence of water vapor. This result is in agreement with the experimental observation. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 236–244, 2009 相似文献
13.
CF3CF2CH2OH is a new chlorofluorocarbon (CFC) alternative. However, there are few data about its atmospheric fate. The kinetics of its atmospheric oxidation, the OH radical reaction of CF3CF2CH2OH, has been investigated in a 2‐liter Pyrex reactor in the temperature range of 298 ∼ 356 K using gas chromatography (GC)–mass spectrometry (MS) for analysis in this study. The rate coefficient of k1 = (2.27) × 10−12 exp[−(900 ± 70)/T] cm3 molecule−1 s−1 was determined using the relative rate method. The results are in good agreement with the literature values and the prediction of Atkinson's structure–activity relationship (SAR) model. From these results, the atmospheric lifetime of CF3CF2CH2OH in the troposphere was deduced to be 0.34 year, which is 250 and 6 times shorter than those of CFC‐113 and hydrochlorofluorocarbons (HCFC‐225ca), respectively. Therefore CF3CF2CH2OH has significant potential for the replacement of CFC‐113 and HCFC‐225ca. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 73–78, 2000 相似文献
14.
Relative rate techniques were used to measure k(Cl + CF2ClCFClH) = (1.4) × 10−11 exp[−(2360 ± 400)/T] cm3 molecule−1 s−1; k(Cl + CF2ClCFClH) = 5.1 × 10−15 cm3 molecule−1 s−1 at 298 K. This result is discussed with respect to the available data. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 785–788, 1999 相似文献
15.
Manolis N. Romanias Vassileios G. Stefanopoulos Dimitrios K. Papanastasiou Vassileios C. Papadimitriou Panos Papagiannakopoulos 《国际化学动力学杂志》2010,42(12):724-734
The rate coefficient for the gas‐phase reaction of chlorine atoms with acetone was determined as a function of temperature (273–363 K) and pressure (0.002–700 Torr) using complementary absolute and relative rate methods. Absolute rate measurements were performed at the low‐pressure regime (~2 mTorr), employing the very low pressure reactor coupled with quadrupole mass spectrometry (VLPR/QMS) technique. The absolute rate coefficient was given by the Arrhenius expression k(T) = (1.68 ± 0.27) × 10?11 exp[?(608 ± 16)/T] cm3 molecule?1 s?1 and k(298 K) = (2.17 ± 0.19) × 10?12 cm3 molecule?1 s?1. The quoted uncertainties are the 2σ (95% level of confidence), including estimated systematic uncertainties. The hydrogen abstraction pathway leading to HCl was the predominant pathway, whereas the reaction channel of acetyl chloride formation (CH3C(O)Cl) was determined to be less than 0.1%. In addition, relative rate measurements were performed by employing a static thermostated photochemical reactor coupled with FTIR spectroscopy (TPCR/FTIR) technique. The reactions of Cl atoms with CHF2CH2OH (3) and ClCH2CH2Cl (4) were used as reference reactions with k3(T) = (2.61 ± 0.49) × 10?11 exp[?(662 ± 60)/T] and k4(T) = (4.93 ± 0.96) × 10?11 exp[?(1087 ± 68)/T] cm3 molecule?1 s?1, respectively. The relative rate coefficients were independent of pressure over the range 30–700 Torr, and the temperature dependence was given by the expression k(T) = (3.43 ± 0.75) × 10?11 exp[?(830 ± 68)/T] cm3 molecule?1 s?1 and k(298 K) = (2.18 ± 0.03) × 10?12 cm3 molecule?1 s?1. The quoted errors limits (2σ) are at the 95% level of confidence and do not include systematic uncertainties. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 724–734, 2010 相似文献
16.
The effects of ammonium sulfate aerosols on the kinetics of the hydroxyl radical reactions with C1–C6 aliphatic alcohols have been investigated using the relative rate technique. P‐xylene was used as a reference compound for the C2–C6 aliphatic alcohols study, and ethanol was used as a reference compound for the methanol study. Two different aerosol concentrations that are typical of polluted urban conditions were tested. The total surface areas of aerosols were 1400 μm2 cm?3 (condition I) and 3400 μm2 cm?3 (condition II). Results indicate that ammonium sulfate aerosols promote the ethanol/OH radical and 1‐propanol/OH radical reactions as compared to the p‐xylene/OH radical reaction. The relative rate of the ethanol/·OH reaction versus the p‐xylene/·OH reaction increased from 0.19 ± 0.01 in the absence of aerosols to 0.24 ± 0.01 and 0.26 ± 0.02 under aerosol conditions I and II, respectively. The relative rate of the 1‐propanol/·OH reaction versus the p‐xylene/·OH reaction increased from 0.45 ± 0.03 in the absence aerosols to 0.56 ± 0.02 and 0.55 ± 0.03 under aerosol conditions I and II, respectively. However, significant changes in the relative rates of the 1‐butanol/·OH, 1‐pentanol/·OH, and 1‐hexanol/·OH reactions versus the p‐xylene/·OH reaction were not observed for either aerosol concentration. The relative rates of the methanol/·OH reaction versus the ethanol/·OH reaction were identical in the absence and presence of aerosols. These results indicate that ammonium sulfate aerosols promote the methanol/·OH reaction as much as the ethanol/·OH reaction (as compared to the p‐xylene/·OH reaction). © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 422–430, 2001 相似文献
17.
H. Moustafa S. El‐Taher M. F. Shibl Rifaat Hilal 《International journal of quantum chemistry》2002,87(6):378-388
The ground‐state geometries of uracil, 6‐hydroxy‐uracil, and 6‐hydroxy‐, 6‐amino‐, 6‐methyl‐, 6‐trifluoro‐, and 6‐phenyl‐2‐thiouracil were optimized at the Hartree–Fock level. The molecular structures were fully optimized using the 6‐31G and 6‐31G* basis sets. The effect of substituents on the geometry and electronic structural features of 2‐thiouracils were examined. The perturbation effects of the OH and NH2 groups are by far more pronounced on the geometric features and the dipole moment magnitude and direction of 2‐thiouracil. The potential energy per atom criteria was used to compare the relative tightness of binding in the studied series. Proton affinity and deprotonation enthalpy on each of the possible sites in 2‐thiouracil and its derivatives have been calculated at the 6‐31G/MP2 level of theory. The obtained results show that thiouracils behave as bases where they possess a high tendency to abstract protons. Substituents in the 6‐position have the general effect of enhancing the basicity strength of the thiocabonyl site in the order Ph < CH3 ≈ NH2 < OH. The CF3 group has the effect of reducing considerably the basicity strength and enhances the acidity strength at both N1 and N3. © 2002 Wiley Periodicals, Inc. Int J Quantum Chem, 2002 相似文献
18.
The gas phase elimination kinetics of the title compound was studied over the temperature range of 260.1–315.0°C and pressure range of 20–70 Torr. This elimination, in seasoned static reaction system and in the presence of at least fourfold of the free radical inhibitor toluene, is homogeneous, unimolecular and follows a first‐order rate law. The reaction yielded mainly benzaldehyde, CO, and HBr, and small amounts of benzylbromide and CO2. The observed rate coefficients are expressed by the following Arrhenius equations: For benzaldehyde formation: log k1 (s−1) = (12.23 ± 0.26) − (164.9 ± 2.7) kJ mol−1 (2.303 RT)−1 For benzylbromide formation: log k1 (s−1) = (13.82 ± 0.50) − (192.8 ± 5.5) kJ mol−1 (2.303 RT)−1 The mechanisms are believed to proceed through a semi‐polar five‐membered cyclic transition state for the benzaldehyde formation, while a four‐centered cyclic transition state for benzylbromide formation. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 725–728, 1999 相似文献
19.
New and improved routes to trifluoromethyl fluoroformate were developed. Sterically hindered halogenated olefins initiated the reaction of CF3OF and CO under mild conditions, giving yields of up to 80%. The thermal reaction of CF3OF and CO in a flow system was highly dependent on temperature and the type of tubular reactor material. A PTFE reactor gave moderate conversion and high selectivity for the formate at 120 °C. 相似文献
20.
The gas‐phase thermal chlorination of CHCl3 has been studied up to high conversions by photometry and gas chromatography in a conditioned static quartz reaction vessel between 573 and 635 K. The initial pressures of both CHCl3 and Cl2 ranged from about 10–100 Torr, and the initial total pressure was varied between about 30–190 Torr. The reaction is rather complex because the produced CCl4 is not stable. The rate of consumption of Cl2 therefore increases in the course of time. This acceleration is explained quantitatively in terms of a radical mechanism and its kinetic and thermodynamic parameters. This reaction model is based on a known model for the pyrolysis of CCl4 to which only one reaction couple involving CHCl3 has been added. Analyses of the rates of the homogeneous elementary steps show that the primary source of Cl atoms is the second‐order dissociation of Cl2, which is rapidly superseded by a secondary source, the first‐order dissociation of the CCl4 primary product. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 466–472, 2000 相似文献