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.
Hexafluoro acetone CF3COCF3 has been shown to react rapidly with the CH2OO, CH3CHOO, and (CH3)2COO intermediates that are formed in the ozonolysis of C2H4, trans‐2‐C4H8, and 2,3‐dimethyl‐2‐butene, respectively, and to form products tentatively assigned to the corresponding secondary ozonides. Relative rate method applied to the C2H4 ozonolysis has indicated that CF3COCF3 reacts ∼13 times faster than CH3CHO. © 1999 John Wiley & Sons, Inc., Int J Chem Kinet 31: 261–269, 1999  相似文献   

2.
The reactions of Criegee intermediates in the gas phase are reviewed. These intermediates are formed by the reaction of olefins with ozone. In the gas phase Criegee intermediates have a biradical character. Initially they are formed as vibrationally hot species. After deactivation by collision with a third body, they can participate in bimolecular reactions with aldehydes, NOx, SO2, water, and so on. Reaction mechanisms are discussed.  相似文献   

3.
4.
Using a relative rate method, rate constants for the gas phase reactions of O3 with 1‐ and 3‐methylcyclopentene, 1‐, 3‐, and 4‐methylcyclohexene, 1‐methylcycloheptene, cis‐cyclooctene, 1‐ and 3‐methylcyclooctene, 1,3‐ and 1,5‐cyclooctadiene, and 1,3,5,7‐cyclooctatetraene have been measured at 296 ± 2 K and atmospheric pressure of air. The rate constants obtained (in units of 10?18 cm3 molecule?1 s?1) are 1‐methylcyclopentene, 832 ± 24; 3‐methylcyclopentene, 334 ± 12; 1‐methylcyclohexene, 146 ± 10; 3‐methylcyclohexene, 55.3 ± 2.6; 4‐methylcyclohexene, 73.1 ± 3.6; 1‐methylcycloheptene, 930 ± 24; cis‐cyclooctene, 386 ± 23; 1‐methylcyclooctene, 1420 ± 100; 3‐methylcyclooctene, 139 ± 9; cis,cis‐1,3‐cyclooctadiene, 20.0 ± 1.4; 1,5‐cyclooctadiene, 152 ± 10; and 1,3,5,7‐cyclooctatetraene, 2.60 ± 0.19, where the indicated errors are two least‐squares standard deviations and do not include the uncertainties in the rate constants for the reference alkenes (propene, 1‐butene, cis‐2‐butene, trans‐2‐butene, 2‐methyl‐2‐butene, and terpinolene). These rate data are compared with the few available literature data, and the effects of methyl substitution discussed. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 37: 183–190, 2005  相似文献   

5.
The electrophilic additions of hydroperoxyl (HO 2 ), alkylperoxyl (RO 2 ), and halogenated alkylperoxyl radicals to ethylene were studied using the AM1 and PM3 semiempirical MO methods at the SCF/UHF level. Reactantlike transition states were predicted for the title additions. The AM1 activation enthalpies (ΔH f * ) were found to be increased in the order HO 2 <CH3O 2 <C2H5O 2 <i‐C3H7O 2 . The reactivity of an alkylperoxyl radical toward ethylene was found to be increased as the degree of halogen substitution on the alkyl group increased. A good correlation was established between ΔH f * and the Taft polar substituent constants, σ*. The Evans–Polanyi correlation between ΔH f * and ΔH r ° was justified and the validity of the Hammond postulate was indicated. The calculated results were compared with the available experimental findings. ©1999 John Wiley & Sons, Inc. Int J Quant Chem 71: 273–283, 1999  相似文献   

6.
An experimental investigation of the hydroxyl radical initiated gas‐phase photooxidation of 1‐propanol in the presence of NO was carried out in a reaction chamber using gas chromatography mass spectrometry. The products identified in the OH radical reactions of 1‐propanol were propionaldehyde and acetaldehyde, with corresponding formation yields of 0.719 ± 0.058 and 0.184 ± 0.030, respectively. Errors represent ±2σ. The experimental product yields were compared to predictions made using chemical mechanisms. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 810–818, 1999  相似文献   

7.
The bimolecular rate constant of k (9.4 ± 2.4 × 10?12 cm3 molecule?1 s?1 was measured using the relative rate technique for the reaction of the nitrate radical (NO3?) with 4‐(2,6,6‐trimethyl‐1‐cyclohexen‐1‐yl)‐3‐buten‐2‐one (β‐ionone) at (297 ± 3) K and 1 atmosphere total pressure. In addition, the products of β‐ionone + NO3? reaction were also investigated. The identified reaction products were glyoxal (HC(?O)C(?O)H), and methylglyoxal (CH3C(?O)C(?O)H). Derivatizing agents O‐(2,3,4,5,6‐pentafluorobenzyl)hydroxylamine and N,O‐bis(trimethylsilyl)trifluoroacetamide were used to propose the other major reaction products: 3‐oxobutane‐1,2‐diyl nitrate, 2,6,6‐trimethylcyclohex‐1‐ene‐carbaldehyde, 2‐oxo‐1‐(2,6,6‐trimethylcyclohex‐1‐en‐1‐yl)ethyl nitrate, pentane‐2,4‐dione, 3‐oxo‐1‐(2,6,6‐trimethylcyclohex‐1‐en‐1‐yl)butane‐1,2‐diyl dinitrate, 3,3‐dimethylcyclohexane‐1,2‐dione, and 4‐oxopent‐2‐enal. The elucidation of these products was facilitated by mass spectrometry of the derivatized reaction products coupled with plausible β‐ionone + NO3? reaction mechanisms based on previously published volatile organic compound + NO3? gas‐phase mechanisms. The additional gas‐phase products 5‐acetyl‐2‐ethylidene‐3‐methylcyclopentyl nitrate, 1‐(1‐hydroxy‐7,7‐dimethyl‐2,3,4,5,6,7‐hexahydro‐1 H‐inden‐2‐yl)ethanone, 1‐(1‐hydroxy‐3a,7‐dimethyl‐2,3,3a,4,5,6,‐hexahydro‐1 H‐inden‐2‐yl)ethanone, and 5‐acetyl‐2‐ethylidene‐3‐methylcyclopentanone are proposed to be the result of cyclization through a reaction intermediate. © 2009 Wiley Periodicals, Inc. *
  • 1 This article is a U.S. Government work and, as such, is in the public domain of the United States of America.
  • Int J Chem Kinet 41: 629–641, 2009  相似文献   

    8.
    Rate coefficients have been determined for the gas‐phase reaction of the hydroxyl (OH) radical with the aromatic dihydroxy compounds 1,2‐dihydroxybenzene, 1,2‐dihydroxy‐3‐methylbenzene and 1,2‐dihydroxy‐4‐methylbenzene as well as the two benzoquinone derivatives 1,4‐benzoquinone and methyl‐1,4‐benzoquinone. The measurements were performed in a large‐volume photoreactor at (300 ± 5) K in 760 Torr of synthetic air using the relative kinetic technique. The rate coefficients obtained using isoprene, 1,3‐butadiene, and E‐2‐butene as reference hydrocarbons are kOH(1,2‐dihydroxybenzene) = (1.04 ± 0.21) × 10−10 cm3 s−1, kOH(1,2‐dihydroxy‐3‐methylbenzene) = (2.05 ± 0.43) × 10−10 cm3 s−1, kOH(1,2‐dihydroxy‐4‐methylbenzene) = (1.56 ± 0.33) × 10−10 cm3 s−1, kOH(1,4‐benzoquinone) = (4.6 ± 0.9) × 10−12 cm3 s−1, kOH(methyl‐1,4‐benzoquinone) = (2.35 ± 0.47) × 10−11 cm3 s−1. This study represents the first determination of OH radical reaction‐rate coefficients for these compounds. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 696–702, 2000  相似文献   

    9.
    The laser photolysis–resonance fluorescence technique has been used to determine the absolute rate coefficient for the Cl atom reaction with a series of ethers, at room temperature (298 ± 2) K and in the pressure range 15–60 Torr. The rate coefficients obtained (in units of cm3 molecule−1 s−1) are dimethyl ether (1.3 ± 0.2) × 10−10, diethyl ether (2.5 ± 0.3) × 10−10, di‐n‐propyl ether (3.6 ± 0.4) × 10−10, di‐n‐butyl ether (4.5 ± 0.5) × 10−10, di‐isopropyl ether (1.6 ± 0.2) × 10−10, methyl tert‐butyl ether (1.4 ± 0.2) × 10−10, and ethyl tert‐butyl ether (1.5 ± 0.2) × 10−10. The results are discussed in terms of structure–reactivity relationship. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 105–110, 2000  相似文献   

    10.
    The pulsed laser photolysis‐resonance fluorescence technique has been used to determine the absolute rate coefficient for the Cl atom reaction with a series of ketones, at room temperature (298 ± 2) K and in the pressure range 15–60 Torr. The rate coefficients obtained (in units of cm3 molecule−1 s−1) are: acetone (3.06 ± 0.38) × 10−12, 2‐butanone (3.24 ± 0.38) × 10−11, 3‐methyl‐2‐butanone (7.02 ± 0.89) × 10−11, 4‐methyl‐2‐pentanone (9.72 ± 1.2) × 10−11, 5‐methyl‐2‐hexanone (1.06 ± 0.14) × 10−10, chloroacetone (3.50 ± 0.45) × 10−12, 1,1‐dichloroacetone (4.16 ± 0.57) × 10−13, and 1,1,3‐trichloroacetone (<2.4 × 10−12). © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 62–66, 2000  相似文献   

    11.
    Gas‐phase reactions of ozone with two butenes (1‐butene and isobutene) and two methyl‐substituted butenes (2‐methyl‐1‐butene and 3‐methyl‐1‐butene) have been studied in an indoor chamber at 295–351 K. The O3 concentrations were monitored by Model 49C‐Ozone analyzer. The butene concentrations were measured by gas chromatography–flame ionization detector. The Arrhenius expressions of k=3.50×10?15e(?1756±84)/T cm3 molecule?1 s?1, k=3.39×10?15e(?1697±52)/T cm3 molecule?1 s?1, k=6.18×10?15e?(1822±80)/T cm3 molecule?1 s?1, and k=7.24×10?14e?(2741±139)/T cm3 molecule?1 s?1 were obtained for the ozonolysis reactions of 1‐butene, isobutene, 2‐methyl‐1‐butene, and 3‐methyl‐1‐butene, respectively. Both the reaction rate constant and activation energy obtained in this work are in good agreement with those reported by using different techniques in the literature. © 2011 Wiley Peiodicals, Inc. Int J Chem Kinet 43: 238–246, 2011  相似文献   

    12.
    The relative‐rate method has been used to determine the rate coefficients for the reactions of OH radicals with three C5 biogenic alcohols, 2‐methyl‐3‐buten‐2‐ol (k1), 3‐methyl‐3‐buten‐1‐ol (k2), and 3‐methyl‐2‐buten‐1‐ol (k3), in the gas phase. OH radicals were produced by the photolysis of CH3ONO in the presence of NO. Di‐n‐butyl ether and propene were used as the reference compounds. The absolute rate coefficients obtained with the two reference compounds agreed well with each other. The O3 and O‐atom reactions with the target alcohols were confirmed to have a negligible contribution to their total losses by using two kinds of light sources with different relative rates of CH3ONO and NO2 photolysis. The absolute rate coefficients were obtained as the weighted mean values for the two reference compound systems and were k1 = (6.6 ± 0.5) × 10?11, k2 = (9.7 ± 0.7) × 10?11, and k3 = (1.5 ± 0.1) × 10?10 cm3 molecule?1 s?1 at 298 ± 2 K and 760 torr of air. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 379–385 2004  相似文献   

    13.
    The chemistry and energetics of the Criegee intermediate, a primary product of the ozonolysis of alkenes, are discussed in terms of recent ab-initio calculations and laboratory studies. The experimental observations in O3–alkene systems can be rationalized on the basis of a general mechanism: where the ? represents the range of internal excitation energies available to the planar dioxymethylenes (i.e., the Criegee intermediates) initially formed via exothermic O3-alkene reactions. Estimates are given for the rate constants of these reactions, and a critique is provided of the possible role of the Criegee intermediate and its isomers in the formation of alkanoic acid anhydrides in O3–alkene systems and in the formation of H2SO4 aerosols in O3–alkene–SO2 systems.  相似文献   

    14.
    A laser flash photolysis–resonance fluorescence technique has been employed to study the kinetics of the reactions of atomic chlorine with acetone (CH3C(O)CH3; k1), 2‐butanone (C2H5C(O)CH3; k2), and 3‐pentanone (C2H5C(O)C2H5; k3) as a function of temperature (210–440 K) and pressure (30–300 Torr N2). No significant pressure dependence is observed for any of the reactions studied. Arrhenius expressions (units are 10?11 cm3 molecule?1 s?1) obtained from the data are k1(T) = (1.53 ± 0.19) exp[(?594 ± 33)/T], k2(T) = (2.77 ± 0.33) exp[(+76 ± 33)/T], and k3(T) = (5.66 ± 0.41) exp[(+87 ± 22)/T], where uncertainties are 2σ and represent precision only. The accuracy of reported rate coefficients is estimated to be ±15% over the entire range of pressure and temperature investigated. The room temperature rate coefficients reported in this study are in good agreement with a majority of literature values. However, the activation energies reported in this study are in poor agreement with the literature values, particularly for 2‐butanone and 3‐pentanone. Possible explanations for discrepancies in published kinetic parameters are proposed, and the potential role of Cl + ketone reactions in atmospheric chemistry is discussed. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 40: 259–267, 2008  相似文献   

    15.
    Using a relative rate method, rate constants for the gas‐phase reactions of OH radicals with allyl alcohol, 3‐buten‐1‐ol, 3‐buten‐2‐ol, and 2‐methyl‐3‐buten‐2‐ol have been measured at 296 ± 2 K and atmospheric pressure of air. Using 1,3,5‐trimethylbenzene as the reference compound, the rate constants (in units of 10−11 cm3 molecule−1 s−1) were: allyl alcohol, 5.46 ± 0.35; 3‐buten‐1‐ol, 5.50 ± 0.20; 3‐buten‐2‐ol, 5.93 ± 0.23; and 2‐methyl‐3‐buten‐2‐ol, 5.67 ± 0.13; where the indicated errors are two least‐squares standard deviations and do not include the uncertainty in the rate constant for 1,3,5‐trimethylbenzene. The H‐atom abstraction products acrolein and methyl vinyl ketone were observed from the allyl alcohol and 3‐buten‐2‐ol reactions, respectively, with respective yields of 5.5 ± 0.7 and 4.9 ± 1.4%. No evidence for formation of acrolein from 3‐buten‐1‐ol or 3‐buten‐2‐ol was obtained, with upper limits to the acrolein yields of ≤1.2 and ≤0.5%, respectively, being determined. Reaction mechanisms are discussed. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 142–147, 2001  相似文献   

    16.
    The ozonolysis of olefinic species is an important tropospheric process impacting on climate and human health. However, few studies have investigated these reactions as a function of temperature and even less information is available upon the effects of alkene heteroatomic substitution on the Arrhenius parameters. The electron‐withdrawing capacity of substituents about the olefinic bond strongly influences the rate of alkene ozonolysis. To understand better the effect of these substitutions, the temperature‐dependence of a series of ozone–chloroalkene reactions is investigated. Experiments were conducted in the EXTreme RAnge (EXTRA) chamber, over the range of 292–409 K and 760 Torr. The experimentally determined rate coefficients were fitted using an Arrhenius‐type analysis to yield the following activation energies: 30.80 ± 0.79, 23.18 ± 0.59, 65.2 ± 2.8, 116.9 ± 5.6, 29.5 ± 1.8, and 18.67 ± 0.96 kJ mol?1 and preexponential A‐factors 1.22+0.39?0.29×10?15, 9.3+6.7?5.4×10?16, 1.6+2.5?1.0×10?10, 6+22?3.9×10?4, 1.7+1.6?0.8×10?14, and 4.2+1.9?1.3×10?15 cm3 molecule?1 s?1 for cis‐1,2‐dichloroethene, trans‐1,2‐dichloroethene, trichloroethene, tetrachloroethene, 2‐chloropropene, and 3‐chloro‐1‐butene, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 43: 120–129, 2011  相似文献   

    17.
    Highly reactive acyl nitroso intermediates were formed in situ by transition metals-catalyzed hydrogen peroxide oxidation of hydroxamic acids 1a-b and these transient species trapped with alkenes 2a-c to afford the corresponding ene products 3a-d and 4b up to 91% yield, and halocyclization of 3d gave substituted oxazolidinone 5a in 77% yield.  相似文献   

    18.
    Motivated by the need for chemical strategies designed to tune peptide fragmentation to selective cleavage reactions, benzyl ring substituent influence on the relative formation of carbocation elimination (CCE) products from peptides with benzylamine‐derivatized lysyl residues has been examined using collision‐induced dissociation (CID) tandem mass spectrometry. Unsubstituted benzylamine‐derivatized peptides yield a mixture of products derived from amide backbone cleavage and CCE. The latter involves side‐chain cleavage of the derivatized lysyl residue to form a benzylic carbocation [C7H7]+ and an intact peptide product ion [(MHn)n+ – (C7H7)+](n‐1)+. The CCE pathway is contingent upon protonation of the secondary ε‐amino group (Nε) of the derivatized lysyl residue. Using the Hammett methodology to evaluate the electronic contributions of benzyl ring substituents on chemical reactivity, a direct correlation was observed between changes in the CCE product ion intensity ratios (relative to backbone fragmentation) and the Hammett substituent constants, σ, of the corresponding substituents. There was no correlation between the substituent‐influenced gas‐phase proton affinity of Nε and the relative ratios of CCE product ions. However, a strong correlation was observed between the π orbital interaction energies (ΔEint) of the eliminated benzylic carbocation and the logarithm of the relative ratios, indicating the predominant factor in the CCE pathway is the substituent effect on the level of hyperconjugation and resonance stability of the eliminated benzylic carbocation. This work effectively demonstrates the applicability of σ (and ΔEint) as substituent selection parameters for the design of benzyl‐based peptide‐reactive reagents which tune CCE product formation as desired for specific applications. Copyright © 2012 John Wiley & Sons, Ltd.  相似文献   

    19.
    The reaction of Cl atoms with a series of C2–C5 unsaturated hydrocarbons has been investigated at atmospheric pressure of 760 Torr over the temperature range 283–323 K in air and N2 diluents. The decay of the hydrocarbons was followed using a gas chromatograph with a flame ionization detector (GC‐FID), and the kinetic constants were determined using a relative rate technique with n‐hexane as a reference compound. The Cl atoms were generated by UV photolysis (λ ≥ 300 nm) of Cl2 molecules. The following absolute rate constants (in units of 10−11 cm3 molecule−1 s−1, with errors representing ±2σ) for the reaction at 295 ± 2 K have been derived from the relative rate constants combined to the value 34.5 × 10−11 cm3 molecule−1 s−1 for the Cl + n‐hexane reaction: ethene (9.3 ± 0.6), propyne (22.1 ± 0.3), propene (27.6 ± 0.6), 1‐butene (35.2 ± 0.7), and 1‐pentene (48.3 ± 0.8). The temperature dependence of the reactions can be expressed as simple Arrhenius expressions (in units of 10−11 cm3 molecule−1 s−1): kethene = (0.39 ± 0.22) × 10−11 exp{(226 ± 42)/T}, kpropyne = (4.1 ± 2.5) × 10−11 exp{(118 ± 45)/T}, kpropene = (1.6 ± 1.8) × 10−11 exp{(203 ± 79)/T}, k1‐butene = (1.1 ± 1.3) × 10−11 exp{(245 ± 90)/T}, and k1‐pentene = (4.0 ± 2.2) × 10−11 exp{(423 ± 68)/T}. The applicability of our results to tropospheric chemistry is discussed. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 478–484, 2000  相似文献   

    20.
    Terpenes and terpene alcohols are prevalent compounds found in a wide variety of consumer products including soaps, flavorings, perfumes, and air fresheners used in the indoor environment. Knowing the reaction rate of these chemicals with the nitrate radical is an important factor in determining their fate indoors. In this study, the bimolecular rate constants of k (16.6 ± 4.2) × 10?12, k (12.1 ± 3) × 10?12, and k (2.3 ± 0.6) × 10?14 cm3 molecule?1 s?1 were measured using the relative rate technique for the reaction of the nitrate radical (NO3?) with 2,6‐dimethyl‐2,6‐octadien‐8‐ol (geraniol), 3,7‐dimethyl‐6‐octen‐1‐ol (citronellol), and 2,6‐dimethyl‐7‐octen‐2‐ol (dihydromyrcenol) at (297 ± 3) K and 1 atmosphere total pressure. Using the geraniol, citronellol, or dihydromyrcenol + NO3? rate constants reported here, pseudo‐first‐order rate lifetimes (k′) of 1.5, 1.1, and 0.002 h?1 were determined, respectively. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 669–675, 2010  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