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 46 毫秒
1.
The rate coefficient of the reaction CH+O2 → products was determined by measuring CH-radical concentration profiles in shock-heated 100–150 ppm ethane/1000 ppm O2 mixtures in Ar using cw, narrow-linewidth laser absorption at 431.131 nm. Comparing the measured CH concentration profiles to ones calculated using a detailed kinetics model, yielded the following average value for the rate coefficient independent of temperature over the range 2200–2600 K: The experimental conditions were chosen such that the calculated profiles were sensitive mainly to the reactions CH+O2 → products and CH3+M → CH+H2+M. For the methyl decomposition reaction channel, the following rate-coefficient expression provided the best fit of the measured CH profiles: Additionally, the rate coefficient of the reaction CH2+H→CH+ H2 was determined indirectly in the same system: © 1997 John Wiley & Sons, Inc.  相似文献   

2.
The formation and consumption of CH radicals during shock-induced pyrolysis of a few ppm ethane diluted in argon was measured by a ring-dye laser spectrometer. Absorption-over-time profiles, measured at a resonance line in the Q-branch of the A2Δ − X2Π band of CH at λ = 431.1311 nm, were recorded and transformed into CH concentrations by known absorption coefficients. By adding some hundred ppm of CO2 or O2 to the initial mixtures, the CH concentration profiles were significantly perturbed. Both the perturbed and unperturbed CH concentration profiles have been compared with calculations based on a reaction kinetic model. A sensitivity analysis revealed that the perturbation process was dominated by direct reactions of CH with the added molecules. By fitting calculated to observed CH profiles the following rate coefficients were obtained The experiments were performed in the temperature range between 2500 K and 3500 K. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
Iodinated hydrocarbons are often used as precursors for hydrocarbon radicals in shock-tube experiments. The radicals are produced by C─I bond fission reaction, and their formation can be followed through time-resolved monitoring of the complementary I-atom concentrations, for example, by I-atom resonance absorption spectroscopy (I-ARAS). This very sensitive technique requires, however, an independent calibration. As a very clean source of I atoms, CH3I is particularly well suited as calibration system for I-ARAS presumed the yield of I atoms and the rate coefficient of I-atom formation from CH3I are known with sufficient accuracy. But if the formation of I atoms from CH3I by I-ARAS is to be characterized, an independent calibration system is required. In this study, we propose a cross-calibration approach for I-ARAS based on the simultaneous time-resolved monitoring of I and H atoms by ARAS in C2H5I pyrolysis experiments. For this reaction system, it can be shown that at sufficiently short reaction times very similar amounts of I and H atoms are formed (difference <1%). As calibration of H-ARAS, with mixtures of N2O and H2, is a well-established technique, we calibrated I-atom absorption–time profiles with respect to simultaneously recorded H-atom concentration–time profiles. Using this approach, we investigated the thermal decomposition of CH3I in the temperature range 950–2050 K behind reflected shock waves at two different nominal pressures (p ∼ 0.4 and 1.6 bar, bath gas: Ar). From the obtained absolute I-atom concentration–time profiles at temperatures T < 1250 K, we inferred a second-order rate coefficient k(T) = (1.7 ± 0.7) × 1015 exp(–20020 K/T) cm3 mol–1 s–1 for the reaction CH3I + Ar → CH3 + I + Ar. A small mechanism to describe the pyrolysis of CH3I under shock-tube conditions is presented and discussed.  相似文献   

4.
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   

5.
Rate coefficients, k, for the gas‐phase reaction CH3CO + Cl2 → products (2) were measured between 253 and 384 K at 55–200 Torr (He). Rate coefficients were measured under pseudo‐first‐order conditions in CH3CO with CH3CO produced by the 248‐nm pulsed‐laser photolysis of acetone, CH3C(O)CH3, or 2,3‐butadione, CH3C(O)C(O)CH3. The loss of CH3CO was monitored by cavity ring‐down spectroscopy (CRDS) at 532 nm. Rate coefficients were determined by first‐order kinetic analysis of the CH3CO temporal profiles for [Cl2] < 1 × 1014 molecule cm?3 and the analysis of the CRDS profiles by the simultaneous kinetics and ring‐down method for experiments performed with [Cl2] > 1 × 1014 molecule cm?3. k2(T) was found to be independent of pressure, with k2(296 K) = (3.0 ± 0.5) × 10?11 cm3 molecule?1 s?1. k2(T) showed a weak negative temperature dependence that is well reproduced by the Arrhenius expression k2(T) = (2.2 ± 0.8) × 10?11 exp[(85 ± 120)/T] cm3 molecule?1 s?1. The quoted uncertainties in k2(T) are at the 2σ level (95% confidence interval) and include estimated systematic errors. A comparison of the present work with previously reported rate coefficients for the CH3CO + Cl2 reaction is presented. © 2009 Wiley Periodicals, Inc. Int J Chem Kinet 41: 543–553, 2009  相似文献   

6.
The UV absorption spectrum and kinetics of CH2I and CH2IO2 radicals have been studied in the gasphase at 295 K using a pulse radiolysis UV absorption spectroscopic technique. UV absorption spectra of CH2I and CH2IO2 radicals were quantified in the range 220–400 nm. The spectrum of CH2I has absorption maxima at 280 nm and 337.5 nm. The absorption cross-section for the CH2I radicals at 337.5 nm was (4.1 ± 0.9) × 10?18 cm2 molecule?1. The UV spectrum of CH2IO2 radicals is broad. The absorption cross-section at 370 nm was (2.1 ± 0.5) × 10?18 cm2 molecule?1. The rate constant for the self reaction of CH2I radicals, k = 4 × 10?11 cm3 molecule?1 s?1 at 1000 mbar total pressure of SF6, was derived by kinetic modelling of experimental absorbance transients. The observed self-reaction rate constant for CH2IO2 radicals was estimated also by modelling to k = 9 × 10?11 cm3 molecule?1 s?1. As part of this work a rate constant of (2.0 ± 0.3) × 10?10 cm3 molecule?1 s?1 was measured for the reaction of F atoms with CH3I. The branching ratios of this reaction for abstraction of an I atom and a H atom were determined to (64 ± 6)% and (36 ± 6)%, respectively. © 1994 John Wiley & Sons, Inc.  相似文献   

7.
Rate constants for the reaction OH + CO have been measured as functions of temperature (340–1220 K) and water concentration in the presence of 1 atm of argon. Results at zero water concentration yield the expression, log k? (cm3 molecule?1 S?1) = ?12.96 + 4.7 × 10?4 T, for the reaction rate constant as a function of temperature. These results are in very good agreement with previous direct measurements and in reasonable agreement with flame and shock tube measurements. Explanations are offered for the involvement of the water molecule in the present experiments and earlier measurements from this laboratory throughout the entire temperature range. Results are consistent with previous results showing little, if any, pressure effect of Ar on the reaction up to 1 atm of Ar.  相似文献   

8.
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   

9.
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   

10.
The rate coefficients for the reactions were determined using mixtures of HNO3/CO/Ar and HNO3/HNCO/Ar in incident shock wave experiments. Simultaneous OH and CO2 absorption time-histories were obtained via cw uv narrow-linewidth absorption at 32606.56 cm−1 (λ = 306.687 nm) and cw infrared narrow-linewidth absorption at 2380.72 cm−1 (λ = 4.2004 μm), respectively. The measurements of k1 determined from measured CO2 time-histories are in good agreement with those determined from previous measurements of OH time-histories at this laboratory. The rate coefficient for the overall reaction of HNCO + OH → Products was determined from analysis of OH data traces. The uncertainty in k2 was found to be +22% −16%. By incorporating data from a previous low-temperature study, the following empirical expression was determined for the bimolecular reaction: over the temperature range 620–1860 K. From analysis of CO2 data traces, an upper limit on the branching fraction (α = k2a/k2) for reaction (2a) of 10% was found, independent of temperature over the range 1250–1860 K. © 1996 John Wiley & Sons, Inc.  相似文献   

11.
The reaction chemistry of C2N2? Ar and C2N2? NO? Ar mixtures has been investigated behind incident shock waves. Progress of the reaction was monitored by observing the cyano radical (CN) in absorption at 388.3 nm. A quantitative spectroscopic model was used to determine concentration histories of CN. From initial slopes of CN concentration during cyanogen pyrolysis, the rate constant for C2N2 + M → 2CN + M (1) was determined to be k1 = (4.11 ± 1.8) × 1016 exp(?47,070 ± 1400/T) cm3/mol · s. A reaction sequence for the C2N2? NO system was developed, and CN profiles were computed. By comparison with experimental CN profiles the rate constant for the reaction CN + NO → NCO + N (3) was determined to be k3 = 10(14.0 ± 0.3) exp(?21,190 ± 1500/T) cm3/mol · s. In addition, the rate of the four-centered reaction CN + NO → N2 + CO (2) was estimated to be approximately three orders of magnitude below collision frequency.  相似文献   

12.
The rate constant of the title reaction is determined during thermal decomposition of di-n-pentyl peroxide C5H11O( )OC5H11 in oxygen over the temperature range 463–523 K. The pyrolysis of di-n-pentyl peroxide in O2/N2 mixtures is studied at atmospheric pressure in passivated quartz vessels. The reaction products are sampled through a micro-probe, collected on a liquid-nitrogen trap and solubilized in liquid acetonitrile. Analysis of the main compound, peroxide C5H10O3, was carried out by GC/MS, GC/MS/MS [electron impact EI and NH3 chemical ionization CI conditions]. After micro-preparative GC separation of this peroxide, the structure of two cyclic isomers (3S*,6S*)3α-hydroxy-6-methyl-1,2-dioxane and (3R*,6S*)3α-hydroxy-6-methyl-1,2-dioxane was determined from 1H NMR spectra. The hydroperoxy-pentanal OHC( )(CH2)2( )CH(OOH)( )CH3 is formed in the gas phase and is in equilibrium with these two cyclic epimers, which are predominant in the liquid phase at room temperature. This peroxide is produced by successive reactions of the n-pentoxy radical: a first one generates the CH3C·H(CH2)3OH radical which reacts with O2 to form CH3CH(OO·)(CH2)3OH; this hydroxyperoxy radical isomerizes and forms the hydroperoxy HOC·H(CH2)2CH(OOH)CH3 radical. This last species leads to the pentanal-hydroperoxide (also called oxo-hydroperoxide, or carbonyl-hydroperoxide, or hydroperoxypentanal), by the reaction HOC·H(CH2)2CH(OOH)CH3+O2→O()CH(CH2)2CH(OOH)CH3+HO2. The isomerization rate constant HOCH2CH2CH2CH(OO·)CH3→HOC·HCH2CH2CH(OOH)CH3 (k3) has been determined by comparison to the competing well-known reaction RO2+NO→RO+NO2 (k7). By adding small amounts of NO (0–1.6×1015 molecules cm−3) to the di-n-pentyl peroxide/O2/N2 mixtures, the pentanal-hydroperoxide concentration was decreased, due to the consumption of RO2 radicals by reaction (7). The pentanal-hydroperoxide concentration was measured vs. NO concentration at ten temperatures (463–523 K). The isomerization rate constant involving the H atoms of the CH2( )OH group was deduced: or per H atom: The comparison of this rate constant to thermokinetics estimations leads to the conclusion that the strain energy barrier of a seven-member ring transition state is low and near that of a six-member ring. Intramolecular hydroperoxy isomerization reactions produce carbonyl-hydroperoxides which (through atmospheric decomposition) increase concentration of radicals and consequently increase atmospheric pollution, especially tropospheric ozone, during summer anticyclonic periods. Therefore, hydrocarbons used in summer should contain only short chains (<C4) hydrocarbons or totally branched hydrocarbons, for which isomerization reactions are unlikely. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 875–887, 1998  相似文献   

13.
CH and C-atom concentration-time histories were measured during pyrolysis of highly dilute mixtures (6 to 100 ppm) of ethane or methane in argon behind reflected shock waves over the temperature range 2500 to 3800 K and pressure range 0.5 to 1.3 atm. CH was detected using narrow-linewidth laser absorption at 431 nm. C-atom concentrations were measured using atomic resonance absorption spectroscopy (ARAS) at 156.1 nm. These data allow improved understanding of dilute hydrocarbon pyrolysis. A pyrolysis reaction mechanism was developed which fits essential characteristics of the CH and C-atom profiles (time to peak, peak concentration, and 50% decay time) within ±25%. Critical reactions for which rate coefficient data were not previously available are: Best-fit rate coefficients, valid over the range 2500 to 3800 K, are: k4 = 5.0 × 1015 exp(?42800 K/T), k5 = 4.0 × 1015 exp(?41800 K/T), k6 = 1.3 × 1014 exp(?29700 K/T), and k7 = 1.9 × 1014 exp(?33700 K/T) cm3 mol?1 s?1  相似文献   

14.
Using a pulse-radiolysis transient UV–VIS absorption system, rate constants for the reactions of F atoms with CH3CHO (1) and CH3CO radicals with O2 (2) and NO (3) at 295 K and 1000 mbar total pressure of SF6 was determined to be k1=(1.4±0.2)×10−10, k2=(4.4±0.7)×10−12, and k3=(2.4±0.7)×10−11 cm3 molecule−1 s−1. By monitoring the formation of CH3C(O)O2 radicals (λ>250nm) and NO2 (λ=400.5nm) following radiolysis of SF6/CH3CHO/O2 and SF6/CH3CHO/O2/NO mixtures, respectively, it was deduced that reaction of F atoms with CH3CHO gives (65±9)% CH3CO and (35±9)% HC(O)CH2 radicals. Finally, the data obtained here suggest that decomposition of HC(O)CH2O radicals via C C bond scission occurs at a rate of <4.7×105 s−1. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 913–921, 1998  相似文献   

15.
The kinetics and mechanism of the gas-phase reaction of Cl atoms with CH2CO have been studied with a FTIR spectrometer/smog chamber apparatus. Using relative rate methods the rate of reaction of Cl atoms with ketene was found to be independent of total pressure over the range 1–700 torr of air diluent with a rate constant of (2.7 ± 0.5) × 10−10 cm3 molecule−1 s−1 at 295 K. The reaction proceeds via an addition mechanism to give a chloroacetyl radical (CH2ClCO) which has a high degree of internal excitation and undergoes rapid unimolecular decomposition to give a CH2Cl radical and CO. Chloroacetyl radicals were also produced by the reaction of Cl atoms with CH2ClCHO; no decomposition was observed in this case. The rates of addition reactions are usually pressure dependent with the rate increasing with pressure reflecting increased collisional stabilization of the adduct. The absence of such behavior in the reaction of Cl atoms with CH2CO combined with the fact that the reaction rate is close to the gas kinetic limit is attributed to preferential decomposition of excited CH2ClCO radicals to CH2Cl radicals and CO as products as opposed to decomposition to reform the reactants. As part of this work ab initio quantum mechanical calculations (MP2/6-31G(d,p)) were used to derive ΔfH298(CH2ClCO) = −(5.4 ± 4.0) kcal mol−1. © 1996 John Wiley & Sons, Inc.  相似文献   

16.
Data for the title reaction have been fitted using an RRKM/master equation approach. Energy transfer was modeled using an exponential decay with downward step sizes, ΔEd, as a fitting parameter. The low temperature (200 < T (K) < 300) combination of CH3 with Cl atoms in He can be accommodated with ΔEd (cm?1) = 400. Higher temperature (1600 < T (K) < 2100) decomposition in Ar required ΔEd(T) (cm?1) = 694(T/300)0.46. Previous analysis of the analogous system CH4 = CH3 + H required ΔEd(T) (cm?1) = 100(T/300) for He and ΔEd(T) (cm?1) = 150(T/300) for Ar. Understanding of the magnitudes and temperature dependence of ΔEd remains the greatest detriment to quantitative calculation, extrapolation, and prediction of unimolecular rate constants. © 2008 Wiley Periodicals, Inc. Int J Chem Kinet 41: 245–254, 2009  相似文献   

17.
The thermal decomposition of CH3NO2 highly diluted in Ar has been studied in shock waves at 900 < T < 1500 K and 1.5 · 10?5 < [Ar] < 3.5 · 10?4 mol/cm3. Concentration profiles of CH3NO2 and NO2 were recorded. The unimolecular reaction was found to be in its fall-off range. Limiting low pressure rate constants of k0 = [Ar] · 1017.1 exp(?42(kcal/mol)/RT) cm3/ mol sec in the range 900 < T < 1400 K and limiting high pressure rate constants of k = 1016.25 exp (?(58.5 ± 0.5 kcal/mol)/RT) sec?1 have been derived. A rate constant of 1.3 · 1013 cm3/mol sec was found for the first subsequent reaction CH3+NO2 → CH3O+NO.  相似文献   

18.
The combination of sensitive detection of formaldehyde by 174 nm absorption and use of ethyl iodide as a hydrogen atom source allowed direct measurements of the reaction H + CH2O → H2 + HCO behind shock waves. The rate constant was determined for temperatures from 1510 to 1960 K to be k2 = 6.6 × 1014 exp(?40.6 kJ mol?1/RT) cm3 mol?1 s?1 (Δ log k2 = ± 0.22) Considering the low uncertainty in k2, which accounts both for experimental and mechanism‐induced contributions, this result supports the upper range of previously reported, largely scattered high temperature rate constants. Vis–UV light of 174 nm was generated by a microwave N2 discharge lamp. At typical reflected shock wave conditions of 1750 K and 1.3 atm, as low as 33 ppm formaldehyde could be detected. High temperature absorption cross sections of CH2O and other selected species have been determined. © 2002 Wiley Periodicals, Inc. Int J Chem Kinet 34: 374–386, 2002  相似文献   

19.
The reaction of atomic chlorine with CH3CH2OD has been examined using a discharge fast flow system coupled to a mass spectrometer combined with the relative rate method (RR/DF/MS). At 298 ± 2 K, the rate constant for the Cl + CH3CH2OD reaction was determined using cyclohexane as a reference and found to be k3 = (1.13 ± 0.21) × 10?10 cm3 molecule?1 s?1. Mass spectral studies of the reaction products resulted in yields greater than 97% for the combined hydrogen abstraction at the α and β sites (3a + 3b) and less than 3% at the hydroxyl site (3c). As a calibration of the apparatus and the RR/DF/MS technique, the rate constant of the Cl + CH3CH2OH reaction was also determined using cyclohexane as the reference, and a value of k2 = (1.05 ± 0.07) × 10?10 cm3 molecule?1 s?1 was obtained at 298 ± 2 K, which was in excellent agreement with the value given in current literature. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 36: 584–590, 2004  相似文献   

20.
Rate coefficients for the reactions of CH3 + Br2 (k2), CH3CO + Br2 (k3), and Cl + Br2 (k5) were measured using the laser‐pulsed photolysis method combined with detection of the product Br atoms using resonance fluorescence. For the reactions involving organic radicals, the rate coefficients were observed to increase with decreasing temperature and within the temperature range explored, were adequately described by Arrhenius‐like expressions: k2 (224–358 K) = 1.83 × 10?11 exp(252/T) and k3 (228–298 K) = 2.92 × 10?11 exp(361/T) cm3 molecule?1 s?1. The total, temperature‐independent uncertainty for each reaction (including possible systematic errors in Br2 concentration measurement) was estimated as ~7% for k2 and 10% for k3. Accurate data on k5 was obtained at 298 K, with a value of 1.88 × 10?10 cm3 molecule?1 s?1 obtained (with an associated error of 6%). A limited data set at 228 K suggests that k5 is, within experimental uncertainty, independent of temperature. © 2010 Wiley Periodicals, Inc. Int J Chem Kinet 42: 575–585, 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