共查询到20条相似文献,搜索用时 15 毫秒
1.
《Chemical physics letters》1985,116(4):345-352
Absolute rate coefficients for the reactions of OH with C3H6 and C3D6 were measured at temperatures from 293 to 896 K and at pressures from 25 to 600 Torr helium. Mechanistic information of importance to atmospheric and combustion modeling was obtained. 相似文献
2.
The reaction of OH with acetylene was studied in a discharge flow system at room temperature. OH was generated by the reaction of atomic hydrogen with NO2 and was monitored throughout the reaction using ESR spectroscopy. Mass-spectrometric analysis of the reaction products yielded the following results: (1) less than 3 molecules of OH were consumed, and less than 2 molecules of H2O were formed for every molecule of acetylene that reacted; (2) CO was identified as the major carbon-containing product; (3) NO, formed in the generation of OH, reacted with a reaction intermediate to give among other products N2O. These observations placed severe limitations on the choice of a reaction mechanism. A mechanism containing the reaction OH + C2H2 → HC2O + H2 better accounted for the experimental results than one involving the abstraction reaction OH + C2H2 → C2H + H2O. The rate constant for the initial reaction was measured as 1.9 ± 0.6 × 10?13 cm3 molecule?1 sec?1. 相似文献
3.
Jasper AW Klippenstein SJ Harding LB Ruscic B 《The journal of physical chemistry. A》2007,111(19):3932-3950
The CH3 + OH bimolecular reaction and the dissociation of methanol are studied theoretically at conditions relevant to combustion chemistry. Kinetics for the CH3 + OH barrierless association reaction and for the H + CH2OH and H + CH3O product channels are determined in the high-pressure limit using variable reaction coordinate transition state theory and multireference electronic structure calculations to evaluate the fragment interaction energies. The CH3 + OH --> 3CH2 + H2O abstraction reaction and the H2 + HCOH and H2 + H2CO product channels feature localized dynamical bottlenecks and are treated using variational transition state theory and QCISD(T) energies extrapolated to the complete basis set limit. The 1CH2 + H2O product channel has two dynamical regimes, featuring both an inner saddle point and an outer barrierless region, and it is shown that a microcanonical two-state model is necessary to properly describe the association rate for this reaction over a broad temperature range. Experimental channel energies for the methanol system are reevaluated using the Active Thermochemical Tables (ATcT) approach. Pressure dependent, phenomenological rate coefficients for the CH3 + OH bimolecular reaction and for methanol decomposition are determined via master equation simulations. The predicted results agree well with experimental results, including those from a companion high-temperature shock tube determination for the decomposition of methanol. 相似文献
4.
Hemibonding between hydroxyl radical and water 总被引:1,自引:0,他引:1
Chipman DM 《The journal of physical chemistry. A》2011,115(7):1161-1171
The ultraviolet absorption peak commonly used to identify OH radical in liquid water is mainly due to a charge-transfer-from-solvent transition that is prominent when OH is hemibonded, rather than more stable hydrogen bonded, to H(2)O. This work computationally characterizes the hemibonding interaction and the extent of the geometrical region over which it is significant. Hemibonding is found to be associated with an enlarged energy separation between the two lowest-lying electronic states, which are otherwise always quite close to one another. The lower state, wherein the hemibonding occurs, retains an attractive interaction energy between OH and H(2)O that can be as much as one-half as strong as the optimum hydrogen-bonding interaction, while the enlarged separation between the states is mainly due to the upper state becoming repulsive over most of the hemibonding region. Hemibonding also leads to a considerable drop in the energy and a considerable increase in the oscillator strength of the characteristic charge-transfer transition. The region of significant hemibonding is found to lie within a moderate range of O-O azimuthal angles and over quite wide ranges of O-O separation distances and hydroxyl OH tilt angles. In particular, significant hemibonding interactions can extend down to surprisingly short O-O distances, where the oscillator strength for the charge-transfer-from-solvent transition becomes very large. 相似文献
5.
The kinetics of the decay of hydroxyl radicals in the presence of excess acetylene were studied at pressures in the vicinity of 1 torr and at ambient temperature in a tubular discharge-flow reactor. Hydroxyl radicals were produced by the reaction of atomic hydrogen with nitrogen dioxide, H + NO2 → OH + NO. The concentration of hydroxyl was followed by line absorption photometry at 308.939 nm and 308.328 nm. Second-order rate coeffcients were determined in two sets of experiments. The initial concentration ratio [C2H2]0/[OH]0 was in the range of 2.3 to 13.2 in the first set, and 14 to 125 (owing to greater hydroxyl detection sensitivity) in the second set. Values of the second-order rate coefficient obtained were nk5 = (2.9 ± 0.3) × 10?13 cm3/molec-sec in the first set, and nk5 = (2.1 ± 0.6) × 10?13 cm3/molec-sec in the second set, where n is the stoichiometric coefficient of OH. A value of the bimolecular rate constant k5 = (2.0 ± 0.6) × 10?13 cm/molec-sec is consistent with both sets of data, as well as an earlier determination. 相似文献
6.
Absolute rate coefficients for the reactions of the hydroxyl radical with ethane (k1, 297–300 K) and propane (k2, 297–690 K) were measured using the flash photolysis–resonance fluorescence technique. The rate coefficient data were fit by the following temperature-dependent expressions, in units of cm3/molecule·s: k1(T) = 1.43 × 10?14T1.05 exp (?911/T) and k2(T) = 1.59 × 10?15T1.40 exp (-428/T). Semiquantitative separation of OH-propane reactivity into primary and secondary H-atom abstraction channels was obtained. 相似文献
7.
The gas phase reaction of the hydroxyl radical (OH) with cyclohexanol (COL) has been studied. The rate coefficient was determined to be (19.0 ± 4.8) × 10−12 cm3 molecule−1 s−1 (at 297 ± 3 K and 1 atmosphere total pressure) using the relative rate technique with pentanal, decane, and tridecane as the reference compounds. Assuming an average OH concentration of 1 × 106 molecules cm−3, an atmospheric lifetime of 15 h is calculated for cyclohexanol. Products of the OH + COL reaction were determined to more clearly define COL's atmospheric degradation mechanism. The observed products and their formation yields were: cyclohexanone (0.55 ± 0.06), hexanedial (0.32 ± 0.15), 3‐hydroxycyclohexanone (0.31 ± 0.14), and 4‐hydroxycyclohexanone (0.08 ± 0.04). Consideration of the potential reaction pathways suggests that each of these products is formed via hydrogen abstraction at a different site on the COL ring. The products and their relative amounts are discussed in light of the predicted yields for each reaction channel. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 108–117, 2001 相似文献
8.
The relative rate technique has been used to measure the hydroxyl radical (OH) reaction rate constant of hexamethyldisiloxane (MM, (CH3)3Si-O-Si(CH3)3), octamethyltrisiloxane (MDM, (CH3)3Si-O-Si(CH3)2-O-Si(CH3)3), and decamethyltetrasiloxane (MD2M, (CH3)3Si-O-Si(CH3)2-O-Si(CH3)2-O-Si(Ch3)3). Hexamethyldisiloxane, octamethyltrisiloxane, and decamethyltetrasiloxane react with OH with bimolecular rate constants of 1.32 ± 0.05 × 10−12 cm3molecule−1s−1, 1.83 ± 0.09 × 10−12 cm3 molecule−1s−1, and 2.66 ± 0.13 × 10−12 cm3molecule−1s−1, respectively. Investigation of the OH + siloxane reaction products yielded trimethylsilanol, pentamethyldisiloxanol, heptamethyltetrasiloxanol, hexamethylcyclotrisiloxane, octamethylcyclotetrasiloxane, and other compounds. Several of these products have not been reported before because these siloxanes and the proposed reaction mechanisms yielding these products are complicated. Some unusual cyclic siloxane products were observed and their formation pathways are discussed in light of current understanding of siloxane atmospheric chemistry. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet 29: 445–451, 1997. 相似文献
9.
The mechanisms and the kinetics of the OH (OD) radicals with methyl acetate CH3C(O)OCH3 are investigated theoretically. The dual-level direct dynamics method is employed in the calculation of the rate constants. The optimized geometries and frequencies and the gradients of the stationary points are calculated at the MP2/6-311G(d,p) level. The energetic information of potential energy surfaces is further refined by the multicoefficient correlation method based on QCISD (MC-QCISD) using the MP2/6-311G(d,p) geometries. Four channels are found for the title reaction. The calculated results reveal that there exists an attractive well (reactant complex) in each entrance H-abstraction channel, that is, the H-abstraction reaction makes a stepwise mechanism. The rate constants are calculated by the canonical variational transition-state theory (CVT) with the interpolated single-point energies (ISPE) approach in the temperature range of 200-1200 K. The small-curvature tunneling effect (SCT) approximation is used to evaluate the transmission coefficient. The calculated rate constants are in good agreement with the experimental ones in the measured temperature range. It is shown that the "out-of-plane hydrogen abstraction" from the methoxy end is the dominant channel at the lower temperatures, and the other two H-abstraction channels should be taken into account with the temperatures increasing. The kinetic isotope effects (KIEs) for the three H-abstraction channels and the total reaction are "inverse", and these theoretically calculated KIEs as a function of temperature are expected to be useful for the future laboratory investigation. 相似文献
10.
11.
The relative rate technique has been used to measure the rate coefficient for the reaction of the hydroxyl radical (OH) with methyl isobutyrate (MIB, (CH3)2 CHC(O) O CH3) to be (1.7 ± 0.4) × 10−12cm3molecule−1s−1 at 297 ± 3 K and 1 atmosphere total pressure. To more clearly define MIB's atmospheric degradation mechanism, the products of the OH + MIB reaction were also determined. The observed products and their yields were: acetone (97 ± 1%, (CH3)2C(O)) and methyl pyruvate (MP, 3.3 ± 0.3%, CH3C(O)C(O) O CH3). The products' formation pathways are discussed in light of current understanding of the atmospheric chemistry of oxygenated organic compounds. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 551–557, 1999 相似文献
12.
A kinetic study of the reaction of the 4-methylphenyl radical (4-C6H4CH3) with the oxygen molecule was conducted using experimental and theoretical approaches. The absorption spectrum for the λ = 266 nm photolysis of the 4-C6H4CH3X (X = Cl, Br)/N2/O2 mixture was measured in the wavelength range of λ = 503-512 nm using N2 as the buffer gas at a total pressure of 40 Torr using a cavity ring-down spectroscopy apparatus coupled with a pulsed laser photolysis system. Based on the absorbance of the product of the 4-C6H4CH3 + O2 reaction at λ = 504 nm, the reaction rate coefficient for the 4-C6H4CH3 + O2 reaction was determined to be k = (1.21 ± 0.10) × 10−11 cm3 molecule−1 s−1 and k = (1.18 ± 0.21) × 10−11 cm3 molecule−1 s−1 using 4-C6H4CH3Cl and 4-C6H4CH3Br, respectively, as the radical precursor. And there was no pressure dependence in the total pressure range of 10-90 Torr varying partial pressure of N2 buffer gas at T = 296 ± 5 K. The geometries, vibration frequencies, and potential energy surfaces of the reactants, major products, and transition states in the 4-C6H4CH3 + O2 reaction were determined using the CBS-QB3 method. The k value at the high-pressure limit was calculated to be 1.26 × 10−11 cm3 molecule−1 s−1 using the variational transition-state theory. The calculated value of k was consistent with the experimental value, which indicated that the 4-C6H4CH3 + O2 reaction reaches the high-pressure limit at 10 Torr. Therefore, the oxidation of the 4-C6H4CH3 radical is almost 10 times faster than that of the benzyl radical, which has the same chemical formula, at the high-pressure limit. 相似文献
13.
J. R. Wells Floyd L. Wiseman Dale C. Williams J. Steven Baxley D. F. Smith 《国际化学动力学杂志》1996,28(7):475-480
The gas-phase reaction products of the OH radical with 2-ethoxyethyl acetate (EEA, CH3C(O)OCH2CH2OCH2CH3) have been investigated. 1,2-Ethanediol acetate formate (EAF, CH3C(O)OCH2CH2OC(O)H) and ethyl formate (EF, HC(O)OCH2CH3) were identified as the two main products. A third product, ethylene glycol diacetate (EGD, CH3C(O)OCH2CH2OC(O)CH3), was also observed. EAF, EF, and EGD formation yields were determined to be 0.37 ± 0.03 and 0.328 ± 0.018 and 0.040 ± 0.005, respectively. Proposed reaction mechanisms are discussed and compared with these data. © 1996 John Wiley & Sons, Inc. 相似文献
14.
The rate coefficient for the reaction of the hydroxyl radical, OH, with propane has been measured at 1220 K in shock tube experiments, and a value of (1.58 ± 0.24) × 1013 cm3/mol s was obtained. This measured value is compared with previous experimental results and a transition-state theory calculation. 相似文献
15.
Ab initio molecular orbital theory and density functional theory calculations have been carried out on the reactions of the trifluoromethyl radical with the hydroxyl and the hydrogen radicals. These reactions are key reactions that underlie a new fire extinguishing mechanism of non-bromine-containing halon replacements. The activation energies calculated by the MP2 and QCISD methods are in good agreement with the experimental values. The B3LYP, as well as MP2 and QCISD, give good results for the calculations of the heats of reactions. The GAUSSIAN-1 and GAUSSIAN-2 theory calculations present the most acxcurate results on both the activation energies and the heats of reactions. The effects of the scaling factors on the activation energies and the heats of reactions are also evaluated. © 1998 John Wiley & Sons, Inc. J Comput Chem 19: 277–289, 1998 相似文献
16.
Anglada JM 《Journal of the American Chemical Society》2004,126(31):9809-9820
The gas phase reaction between formic acid and hydroxyl radical has been investigated with high level quantum mechanical calculations using DFT-B3LYP, MP2, CASSCF, QCISD, and CCSD(T) theoretical approaches in connection with the 6-311+G(2df,2p) and aug-cc-pVTZ basis sets. The reaction has a very complex mechanism involving several elementary processes, which begin with the formation of a reactant complex before the hydrogen abstraction by hydroxyl radical. The results obtained in this investigation explain the unexpected experimental fact that hydroxyl radical extracts predominantly the acidic hydrogen of formic acid. This is due to a mechanism involving a proton coupled electron-transfer process. The calculations show also that the abstraction of formyl hydrogen has an increased contribution at higher temperatures, which is due to a conventional hydrogen abstraction radical type mechanism. The overall rate constant computed at 298 K is 6.24 x 10(-13) cm3 molecules(-1) s(-1), and compares quite well with the range from 3.2 +/- 1 to 4.9 +/- 1.2 x 10(-13) cm3 molecules(-1) s(-1), reported experimentally. 相似文献
17.
James Lee Kacee Caster Trey Maddaleno Zachery Donnellan Talitha M. Selby Fabien Goulay 《国际化学动力学杂志》2020,52(11):838-851
The gas phase reaction of the ground state cyano-radical (CN (X2∑+)) with 2-methylfuran (2-MF) is investigated in a quasi-static reaction cell at pressures ranging from 2.2 to 7.6 Torr and temperatures ranging from 304 to 440 K. The CN radicals are generated in their ground electronic state by pulsed laser photolysis of gaseous cyanogen iodide (ICN) at 266 nm. Their concentration is monitored as a function of reaction time using laser-induced fluorescence at 387.3 nm on the B2∑+ (ν′ = 0) ← X2∑+ (ν″ = 0) vibronic band. The reaction rate coefficient is found to be rapid and independent of pressure and temperature. Over the investigated temperature and pressure ranges, the rate coefficient is measured to be 2.83 (± 0.18) × 10−10 cm3 molecules s−1. The enthalpies of the stationary points and transition states on the CN + 2-MF potential energy surface are calculated using the CBS-QB3 computational method. The kinetic results suggest the lack of a prereactive complex on the reaction entrance channel with either a very small or nonexistent entrance energy barrier. In addition, the potential energy surface calculations reveal only submerged barriers along the minimum energy path. Based on comparisons between previous CN reactions with unsaturated hydrocarbons, the most likely reaction pathway is CN addition onto one of the unsaturated carbons followed by either H or methyl elimination. The implications for the transformation of biomass-derived fuels in nitrogen-rich flames is discussed. 相似文献
18.
The relative rate technique has been used to measure the hydroxyl radical (OH) reaction rate constant of ethyl 3-ethoxypropionate (EEP, CH3CH2(SINGLE BOND)O(SINGLE BOND)CH2CH2C(O)O(SINGLE BOND)CH2CH3). EEP reacts with OH with a bimolecular rate constant of (22.9±7.4)×10−12 cm3 molecule−1s−1 at 297±3 K and 1 atmosphere total pressure. In order to more clearly define EEP's atmospheric reaction mechanism, an investigation into the OH+EEP reaction products was also conducted. The OH+EEP reaction products and yields observed were: ethyl glyoxate (EG, 25±1% HC((DOUBLE BOND)O)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl (2-formyl) acetate (EFA, 4.86±0.2%, HC((DOUBLE BOND)O)(SINGLE BOND)CH2(SINGLE BOND)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl (3-formyloxy) propionate (EFP, 30±1%, HC((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH2(SINGLE BOND)C((DOUBLE BOND)O)(SINGLE BOND)O(SINGLE BOND)CH2CH3), ethyl formate (EF, 37±1%, HC((DOUBLE BOND)O)O(SINGLE BOND)CH2CH3), and acetaldehyde (4.9±0.2%, HC((DOUBLE BOND)O)CH3). Neither the EEP's OH rate constant nor the OH/EEP reaction products have been previously reported. The products' formation pathways are discussed in light of current understanding of oxygenated hydrocarbon atmospheric chemistry. © 1997 John Wiley & Sons, Inc. 相似文献
19.
《International Journal of Radiation Applications and Instrumentation. Part C. Radiation Physics and Chemistry》1991,37(5):467-472
The reaction of the potent carcinogen N-nitrosodimethylamine (NDMA) with hydroxyl radical generated via radiolysis was studied using EPR techniques. Attempts to spin trap NDMA radical intermediates with 3,5-dibromo-4-nitrosobenzene sulfonate (DBNBS) produced only unusual DBNBS radicals. One of these radicals was shown to be generated by both reaction of DBNBS with nitric oxide, and direct oxidation of DBNBS with an inorganic oxidant (.Br-2). Another DBNBS radical was identified as a sulfite spin adduct resulting from the degradation of DBNBS by a NDMA reactive intermediate. In the absence of DBNBS, hydroxyl radical reaction with NDMA gave the dimethylnitroxide radical. Unexpectedly, addition of DBNBS to a solution containing dimethylnitroxide produced an EPR spectrum nearly identical to that of NDMA solutions with DBNBS added before radiolysis. A proposed mechanism accounting for these observations is presented. 相似文献