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 31 毫秒
1.
The rate coefficients for the reactions of Cl atoms with CH3Br, (k1) and CH2Br2, (k2) were measured as functions of temperature by generating Cl atoms via 308 nm laser photolysis of Cl2 and measuring their temporal profiles via resonance fluorescence detection. The measured rate coefficients were: k1 = (1.55 ± 0.18) × 10?11 exp{(?1070 ± 50)/T} and k2 = (6.37 ± 0.55) × 10?12 exp{(?810 ± 50)/T} cm3 molecule?1 s?1. The possible interference of the reaction of CH2Br product with Cl2 in the measurement of k1 was assessed from the temporal profiles of Cl at high concentrations of Cl2 at 298 K. The rate coefficient at 298 K for the CH2Br + Cl2 reaction was derived to be (5.36 ± 0.56) × 10?13 cm3 molecule?1 s?1. Based on the values of k1 and k2, it is deduced that global atmospheric lifetimes for CH3Br and CH2Br2 are unlikely to be affected by loss via reaction with Cl atoms. In the marine boundary layer, the loss via reaction (1) may be significant if the Cl concentrations are high. If found to be true, the contribution from oceans to the overall CH3Br budget may be less than what is currently assumed. © 1994 John Wiley & Sons, Inc.  相似文献   

2.
The homogeneous gas-phase decomposition kinetics of methylsilane and methylsilane-d3 have been investigated by the comparative-rate-single-pulse shock-tube technique at total pressures of 4700 torr in the 1125–1250 K temperature range. Three primary processes occur: CH3SiH3 → CH3SiH + H2 (1), CH3SiH3 → CH4 + SiH2 (2), and CH3SiH3 → CH2 = SiH2 + H2 (3). The high-pressure rate constants for the primary processes in CH3SiH3 obtained by RRKM calculations are log (k1 + k3) (s?1) = 15.2 - 64,780 Cal/θ and log k2 (s?) = 14.50 - 67,600 → 2800 Cal/θ. For CH3SiD3 these same rate constants are log k1 (s?) = 14.99 - 64,700 cal/θ log k2 (s?) = 14.68 – 66,700 → 2000 cal/θ, and log k3 (s?) = 14.3 ? 64,700 cal/θ.  相似文献   

3.
The 220 MHz 1H NMR spectrum of an ether solution of CH3Li and LiBr in 10–1 ratio has been examined as a function of temperature. At low temperature distinct resonances, assignable to Li4(CH3)4 and Li4(CH3)3Br, are seen. Methyl group exchange between the two tetramers is observed in the NMR spectra in the temperature interval ?32 to 0°. The exchange is shown to be much slower than the dissociation of Li4(CH3)4 tetramer, measured in other work. It is proposed that the rate-determining step is dissociation of Li4(CH3)3Br to form Li2(CH3)2 and Li2(CH3)Br. The rate constant for dissociation, k2, obeys the equation ln k2 = 36.0?83303/T.  相似文献   

4.
The rate coefficient of the reaction CH2 + O2OH → HO2 + CH2O, has been measured at 300 K by the LMR flow-tube method, and found to have the unexpectedly large value k = (2?1+2) × 10?12 cm3 molecule?1 s?1. This reaction, preceded by isomerization, may be an important route for the oxidation of CH3O in the upper atmosphere.  相似文献   

5.
The reactions of ethyldiphenylphosphine with a number of cis-dioxomolybdenum(VI) Schiff base coordination complexes are described. These molybdenum complexes incorporate tridentate Schiff base ligands obtained from the condensation of 5-X-salicylaldehyde (X = Cl, Br, H, CH3O) with o-aminobenzenethiol. Oxomolybdenum(IV) Schiff base complexes were observed as products of the reaction of these Mo(VI) complexes with PEtPh2. The kinetics for these reactions were followed spectrophotometrically and the applicable rate law is ? d[MoO2L]/dt = k1[MoO2L][PEtPh2]. The k1's were shown to vary systematically as the X-substituent on the ligand was changed. For MoO2(5-X-SSP), the specific rate constants at 30°C span the range from 19.6 × 10?4 M?1 sec?1 (X = Br) to 8.4 × 10?4 M?1 sec?1 (X = CH3O). It was also observed that a correlation exists between the cathodic reduction potentials (Epc) and the k1's within the series. The rate of reaction of MoO2(5-X-SSP) with PEtPh2 was altered and systematically controlled through ligand design.  相似文献   

6.
The ratio of the Br + Br2 halogen atom exchange rate, k2, to the Br + HI reaction rate k1, has been determined experimentally by using an isotopically selective, laser-initiated chemical reaction. At 294 K, k2/k1 = 4, and thus, k2 = 4 × 10?11 cm3 molecule?1 s?1.  相似文献   

7.
Deactivation rate constants of spin-orbital excited Br atoms in the reactions Br(2P12) + O2 → Br(2P32) + O2 (k1), and Br(2P12) + NO → Br(2P32) + NO (k4) have been measured with a photodissociative IBr laser on the electronic transition 2P12?2P32 in the Br atom (λ = 2.7 μm). The values obtained are (6.4 ± 1.8) × 10?14 cm3 s?1 and (1.9 ± 0.6) × 10?12 cm3 s?1, respectively. Comparison with published data leads to the conclusion that, contrary to a widely accepted point of view, the high rate constants for the quenching of excited halogen atoms are due to resonant energy transfer processes and not to the paramagnetic nature of the quencher.  相似文献   

8.
The fast flow method with laser induced fluorescence detection of CH3C(O)CH2 was employed to obtain the rate constant of k1 (298 K) = (1.83 ± 0.12 (1σ)) × 1010 cm3 mol?1 s?1 for the reaction CH3C(O)CH2 + HBr ? CH3C(O)CH3 + Br (1, ?1). The observed reduced reactivity compared with n‐alkyl or alkoxyl radicals can be attributed to the partial resonance stabilization of the acetonyl radical. An application of k1 in a third law estimation provides ΔfH(CH3C(O)CH2) values of ?24 kJ mol?1 and ?28 kJ mol?1 depending on the rate constants available for reaction ( ‐1 ) from the literature. © 2005 Wiley Periodicals, Inc. Int J Chem Kinet 38: 32–37, 2006  相似文献   

9.
The absolute rate constants of the reactions F + H2CO → HF + HCO (1) and Br + H2CO → HBr + HCO (2) have been measured using the discharge flow reactor-EPR method. Under pseudo-first-order conditions (¦H2CO¦?¦F¦or¦Br¦), the following values were obtained at 298 K: k1 = (6.6 ± 1.1) × 10?11 and k2 = (1.6± 0.3) × 10?12, Units are cm3 molecule?1s?1. The stratospheric implication of these data is discussed and the value obtained for k makes reaction (2) a possible sink for Br atoms in the stratosphere.  相似文献   

10.
The reactions of CH3O2 with SO2 and NO have been studied by steady state photolysis of azomethane in the presence of O2SO2→NO mixtures at 296 K and 1 atm total pressure. The quantum yield of NO oxidation by CH3O2 radicals is increased substantially when SO2 is added to the system indicating an SO2 induced chain oxidation of NO. The rate law gives k1/k2 = (2.5 ± 0.5) × 10?3 for CH3O2 + SO2 → CH3O2SO2 (1), CH3O2 + NO → CH3O + NO2 (2). Combining this ratio with the absolute value of k1 = 8.2 × 10?15 cm3 s?1 gives k2 = 10?11.5 ± 02 cm3 s?1.  相似文献   

11.
Gas-phase reactions typical of the Earth’s atmosphere have been studied for a number of partially fluorinated alcohols (PFAs). The rate constants of the reactions of CF3CH2OH, CH2FCH2OH, and CHF2CH2OH with fluorine atoms have been determined by the relative measurement method. The rate constant for CF3CH2OH has been measured in the temperature range 258–358 K (k = (3.4 ± 2.0) × 1013exp(?E/RT) cm3 mol?1 s?1, where E = ?(1.5 ± 1.3) kJ/mol). The rate constants for CH2FCH2OH and CHF2CH2OH have been determined at room temperature to be (8.3 ± 2.9) × 1013 (T = 295 K) and (6.4 ± 0.6) × 1013 (T = 296 K) cm3 mol?1 s?1, respectively. The rate constants of the reactions between dioxygen and primary radicals resulting from PFA + F reactions have been determined by the relative measurement method. The reaction between O2 and the radicals of the general formula C2H2F3O (CF3CH2? and CF3?HOH) have been investigated in the temperature range 258–358 K to obtain k = (3.8 ± 2.0) × 108exp(?E/RT) cm3 mol?1 s?1, where E = ?(10.2 ± 1.5) kJ/mol. For the reaction between O2 and the radicals of the general formula C2H4FO (? HFCH2O, CH2F?HOH, and CH2FCH2?) at T = 258–358 K, k = (1.3 ± 0.6) × 1011exp(?E/RT) cm3 mol?1 s?1, where E = ?(5.3 ± 1.4) kJ/mol. The rate constant of the reaction between O2 and the radicals with the general formula C2H3F2O (?F2CH2O, CHF2?HOH, and CHF2CH2?) at T = 300 K is k = 1.32 × 1011 cm3 mol?1 s?1. For the reaction between NO and the primary radicals with the general formula C2H2F3O (CF3CH2? and CF3?HOH), which result from the reaction CF3CH2OH + F, the rate constant at 298 K is k = 9.7 × 109 cm3 mol?1 s?1. The experiments were carried out in a flow reactor, and the reaction mixture was analyzed mass-spectrometrically. A mechanism based on the results of our studies and on the literature data has been suggested for the atmospheric degradation of PFAs.  相似文献   

12.
The relative-rate method with gas-chromatographic product analysis was applied to study the kinetics of the reactions Br + CH3Br → CH2Br + HBr (1) and Br + CH2ClBr → CHClBr + HBr (2) The rate coefficient ratio of k 1/ k 2 = (1.6 ± 0.2) exp[(-15.2 ± 0.3) kJ mol-1/ RT] was determined in the temperature range of 353 - 410 K. This revised version was published online in August 2006 with corrections to the Cover Date.  相似文献   

13.
By allowing dimethyl peroxide (10?4M) to decompose in the presence of nitric oxide (4.5 × 10?5M), nitrogen dioxide (6.5 × 10?5M) and carbon tetrafluoride (500 Torr), it has been shown that the ratio k2/k2′ = 2.03 ± 0.47: CH3O + NO → CH3ONO (reaction 2) and CH3O + NO2 → CH3ONO2 (reaction 2′). Deviations from this value in this and previous work is ascribed to the pressure dependence of both these reactions and heterogeneity in reaction (2). In contrast no heterogeneous effects were found for reaction (2′) making it an ideal reference reaction for studying other reactions of the methoxy radical. We conclude that the ratio k2/k2′ is independent of temperature and from k1 = 1010.2±0.4M?1 sec?1 we calculate that k2′ = 109.9±0.4M?1 sec?1. Both k2 and k2′ are pressure dependent but have reached their limiting high-pressure values in the presence of 500 Torr of carbon tetrafluoride. Preliminary results show that k4 = 10.9.0±0.6 10?4.5±1.1M?1 sec?1 (Θ = 2.303RT kcal mole?1) and by k4 = 108.6±0.6 10?2.4±1.1M?1 sec?1: CH3O + O2 → CH2O + HO2 (reaction 4) and CH3O + t-BuH → CH3OH + (t-Bu) (reaction 4′).  相似文献   

14.
A jet-stream kinetic technique and the resonance fluorescence method applied to detection of iodine atoms were used to measure the rate constants of the reactions of the IO· radical with the halohydrocarbons CHFCl-CF2Cl (k = (3.2 ± 0.9) × 10?16 cm3 molecule s?1) and CH2ClF (k = (9.4 ± 1.3) × 10?16 cm3 molecule s?1), the hydrogen-containing haloethers CF3-O-CH3 (k = (6.4 ± 0.9) × 10?16 cm3 molecule s?1) and CF3CH2-O-CHF2 (k = (1.2 ± 0.6) × 10?15 cm3 molecule s?1), and hydrogen iodide (k = (1.3 ± 0.9) × 10?12 cm3 molecule s?1) at 323 K.  相似文献   

15.
The multiple-channel reactions Cl + Si(CH3)4 and Br + Si(CH3)4 are investigated by direct dynamics method. The minimum energy path is calculated at the MP2/6-31+G(d,p) level, and energetic information is further refined by the MC-QCISD (single-point) method. The rate constants for individual reaction channel are calculated by the improved canonical variational transition state theory with small-curvature tunneling correction over the temperature range 200–3,000 K. The theoretical three-parameter expression k 1(T) = 9.97 × 10?13 T 0.54exp(613.22/T) and k 2(T) = 1.16 × 10?17 T 2.30exp(?3525.88/T) (in unit of cm3 molecule?1 s?1) are given. Our calculations indicate that hydrogen abstraction channel is the major channel due to the smaller barrier height among feasible channels considered.  相似文献   

16.
Relative rate techniques were used to study the title reactions in 930–1200 mbar of N2 diluent. The reaction rate coefficients measured in the present work are summarized by the expressions k(Cl + CH2F2) = 1.19 × 10?17 T2 exp(?1023/T) cm3 molecule?1 s?1 (253–553 K), k(Cl + CH3CCl3) = 2.41 × 10?12 exp(?1630/T) cm3 molecule?1 s?1 (253–313 K), and k(Cl + CF3CFH2) = 1.27 × 10?12 exp(?2019/T) cm3 molecule?1 s?1 (253–313 K). Results are discussed with respect to the literature data. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 401–406, 2009  相似文献   

17.
Double fluorescence of p-dimethylacetophenone (DMAPh) in CH3CN and m-methyl-p-cyanodimethylaniline (MCDMA) in CH2Cl2 has been observed and analyzed in terms of reversible excited state isomerisation of the primary excited form b* to the strongly polar rotamer a*. Using the oxygen quenching technique, the kinetics of the reactions have been solved and all rate constants separated. The “formal” lifetimes of the species b*, τb ≡ (kbf + kbd + kba)?1, are 1 ps and 2.2 ps for DMAPh and MCDMA, respectively. The first value fits well to the reorientation relaxation time of acetonitrile.  相似文献   

18.
The kinetics of 1,1-dimethylpropyl peroxy radicals recombination in polar solvents—water, methanol, and their mixtures—was studied by EPR spectroscopy in combination with the stopped-flow method, and the rate constants of this reaction were determined. Peroxyl radicals were generated by mixing solutions of Ce4+ sulfate and 1,1-dimethylpropyl hydroperoxide. The observed EPR signal of the peroxyl radical is a singlet with a g-factor of 2.015 ± 0.001, and a line width of ΔH = (1.36 ± 0.02) × 10?3 T for methanol and ΔH = (9.7 ± 0.2) × 10?4 T for water. The measured rate constants of (CH3)2C(O2·)CH2CH3 radical recombination at 298 K are 2kt = (3.9 ± 0.4) × 104 L mol?1 s?1 for water and 2kt = (5.2 ± 0.5) × 103 L mol?1 s?1 for methanol. A linear relationship between ln(2kt) and the Kirkwood function (ε?1)/(2ε + 1), where e is the dielectric constant of the medium, has been established, indicating an important role of nonspecific solvation in the recombination of tertiary peroxyl radicals.  相似文献   

19.
The rate coefficients for the reaction OH + CH3CH2CH2OH → products (k1) and OH + CH3CH(OH)CH3 → products (k2) were measured by the pulsed‐laser photolysis–laser‐induced fluorescence technique between 237 and 376 K. Arrhenius expressions for k1 and k2 are as follows: k1 = (6.2 ± 0.8) × 10?12 exp[?(10 ± 30)/T] cm3 molecule?1 s?1, with k1(298 K) = (5.90 ± 0.56) × 10?12 cm3 molecule?1 s?1, and k2 = (3.2 ± 0.3) × 10?12 exp[(150 ± 20)/T] cm3 molecule?1 s?1, with k2(298) = (5.22 ± 0.46) × 10?12 cm3 molecule?1 s?1. The quoted uncertainties are at the 95% confidence level and include estimated systematic errors. The results are compared with those from previous measurements and rate coefficient expressions for atmospheric modeling are recommended. The absorption cross sections for n‐propanol and iso‐propanol at 184.9 nm were measured to be (8.89 ± 0.44) × 10?19 and (1.90 ± 0.10) × 10?18 cm2 molecule?1, respectively. The atmospheric implications of the degradation of n‐propanol and iso‐propanol are discussed. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 42: 10–24, 2010  相似文献   

20.
The reaction dynamics for C–Br dissociation within BrH2C–C≡CH(ads) adsorbed on an Ag(111) surface has been investigated by combining density functional theory-based molecular dynamics simulations with short-time Fourier transform (STFT) analysis of the dipole moment autocorrelation function. Two possible reaction pathways for C–Br scission within BrH2C–C≡CH(ads) have been proposed on the basis of different initial structural models. Firstly, the initial perpendicular orientation of adsorbed BrH2C–C≡CH(ads) with a stronger C–Br bond will undergo dynamic rotation leading to the final parallel orientation of BrH2C–C≡CH(ads) to cause the C–Br scission, namely, an indirect dissociation pathway. Secondly, the initial parallel orientation of adsorbed BrH2C–C≡C(ads) with a weaker C–Br bond will directly cause the C–Br scission within BrH2C–C≡CH(ads), namely, a direct dissociation pathway. To further investigate the evolution of different vibrational modes of BrH2C–C≡CH(ads) along these two reaction pathways, the STFT analysis is performed to illustrate that the infrared (IR) active peaks of BrH2C–C≡CH(ads) such as vCH2 [2956 cm?1(s) and 3020 cm?1(as)], v≡CH (3320 cm?1) and vC≡C (2150 cm?1) gradually vanish as the rupture of C–Br bond occurs and then the resulting IR active peaks such as C=C=C (1812 cm?1), ω-CH2 (780 cm?1) and δ-CH (894 cm?1) appear due to the formation of H2C=C=CH(ads) which are in a good agreement with experimental reflection adsorption infrared spectrum (RAIRS) at temperatures of 110 and 200 K, respectively. Finally, the total energy profiles indicate that the reaction barriers for the scission of C–Br within BrH2C–C≡CH(ads) along both direct and indirect dissociation pathways are very close due to a similar rupture of C–Br bond leading to a similar transition state.  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