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 546 毫秒
1.
The molecular modulation spectroscopic technique was employed to study the kinetics of NO3 radicals produced in the 253.7 nm photolysis of flowing gas mixtures of HNO3/CH4/O2 at room temperature. By computer fitting of the NO3 temporal behavior, a rate coefficient of (2.3 ± 0.7) × 10?12 cm3 molecule?1 s?1 was obtained for the reaction between NO3 and CH3O2 at 298 K.  相似文献   

2.
Rate coefficients for the gas‐phase reaction of isoprene with nitrate radicals and with nitrogen dioxide were determined. A Teflon collapsible chamber with solid phase micro extraction (SPME) for sampling and gas chromatography with flame ionization detection (GC/FID) and a glass reactor with long‐path FTIR spectroscopy were used to study the NO3 radical reaction using the relative rate technique with trans‐2‐butene and 2‐buten‐1‐ol (crotyl alcohol) as reference compounds. The rate coefficients obtained are k(isoprene + NO3) = (5.3 ± 0.2) × 10?13 and k(isoprene + NO3) = (7.3 ± 0.9) × 10?13 for the reference compounds trans‐2‐butene and 2‐buten‐1‐ol, respectively. The NO2 reaction was studied using the glass reactor and FTIR spectroscopy under pseudo‐first‐order reaction conditions with both isoprene and NO2 in excess over the other reactant. The obtained rate coefficient was k(isoprene + NO2) = (1.15 ± 0.08) × 10?19. The apparent rate coefficient for the isoprene and NO2 reaction in air when NO2 decay was followed was (1.5 ± 0.2) × 10?19. The discrepancy is explained by the fast formation of peroxy nitrates. Nitro‐ and nitrito‐substituted isoprene and isoprene‐peroxynitrate were tentatively identified products from this reaction. All experiments were conducted at room temperature and at atmospheric pressure in nitrogen or synthetic air. All rate coefficients are in units of cm3 molecule?1 s?1, and the errors are three standard deviations from a linear least square analyses of the experimental data. © 2004 Wiley Periodicals, Inc. Int J Chem Kinet 37: 57–65, 2005  相似文献   

3.
The gas‐phase kinetics of CHBr2 + NO2 and CH3CHBr + NO2 reactions have been studied in direct time resolved measurements using a tubular flow reactor coupled to a photoionization mass spectrometer. The radicals were generated by pulsed laser photolysis of bromoform and 1,1‐dibromoethane at 248 nm. The subsequent decays of the radical concentrations were monitored as a function of [NO2] under pseudo–first‐order conditions. The rate coefficients of both reactions are independent of bath gas (He) pressure and display negative temperature dependence under the conditions of 2–6 Torr pressure (He) and 250–480 K. The obtained bimolecular rate coefficients are k(CHBr2 + NO2) = (9.8 ± 0.4) × 10?12 (T/300 K)?1.65 ± 0.18 cm3 s?1 (288–483 K) and k(CH3CHBr + NO2) = (2.27 ± 0.06) × 10?11 (T/300 K)?1.28 ± 0.11 cm3 s?1 (250–483 K), with the uncertainties given as one standard error. Estimated overall uncertainties in the measured bimolecular reaction rate coefficients are ±25%. The reaction products identified were CBr2O for the CHBr2 + NO2 reaction and CHBrO and CH3CHO with minor amounts of CH3 for the CH3CHBr + NO2 reaction, respectively. © 2012 Wiley Periodicals, Inc. Int J Chem Kinet 44: 767–777, 2012  相似文献   

4.
The rate constants for the reaction of NO3· with sulfur compounds in acetonitrile have been determined by the flash photolysis method. The rate constant for dimethyl sulfone (2.7 × 104 M?1s?1 at ?10°C) is larger than that of the deuterium derivative, indicating that NO3· abstracts the hydrogen atom from dimethyl sulfone. In the case of dimethyl sulfide, the rate constant was evaluated to be 1.5 × 109 M?1 s?1 at ?10°C; the transient absorption band attributable to the cation radical was observed after the decay of NO3·, suggesting the electron transfer reaction from the sulfide to NO3·. For diphenyl sulfide and dimethyl disulfide, the electron transfer reactions were also confirmed. For dimethyl sulfoxide, the reaction rate constant of 1.2 × 109 M?1 s?1 (at ?10°C) was not practically affected by the deuterium substitution, suggesting that NO3· adds to sulfur atom forming (CH3)2?(O)-ONO2. On the other hand, for diphenyl sulfoxide, the electron transfer reaction occurs. By the comparison of these rate constants in acetonitrile solution with the reported rate constants in the gas phase, the change of the reaction paths was revealed.  相似文献   

5.
No reliable rate constant is available for the self-reaction of tert-;butoxy radicals. We have set up a competition between hydrogen abstraction and self-reaction of tert-butoxy radicals in a flash photolysis electron spin resonance study to extract this information. Experimental values of hydrogen abstraction product radical concentrations under various hydrogen donor concentrations were then compared with theoretically calculated values with different values of 2k4 to obtain the best fit. Hydrogen donors such as cyclopentane, anisole, methyl tert-butyl ether, and methanol were chosen for the study. A value of (1.3 ± 0.5) × 109M?1 sec?1 for the rate constant of the self-reaction of tert-butoxy radicals has been determined at 293°K.  相似文献   

6.
A pulse radiolysis system was used to study the kinetics of the reaction of FC(O)O2 radicals with NO2. By monitoring the rate of the decay of NO2 using its absorption at 400 nm the reaction rate constant was determined to be (5.5 ± 0.6) × 10?12 cm3 molecule?1 s?1 at 296 K and 500–1000 mbar pressure of SF6 diluent. A long path length Fourier transform infrared spectrometer was used to investigate the thermal stability of the product FC(O)O2NO2. The rate of thermal decomposition of FC(O)O2NO2 was independent of the total pressure of N2 diluent over the range 100–700 torr and was fit by the expression k?3 = 6.0 × 1016 exp(?14150/T) s?1. The results are discussed in the context of the atmospheric chemistry of FCOx radicals. © 1995 John Wiley & Sons, Inc.  相似文献   

7.
The dark reaction of NOx and H2O vapor in 1 atm of air was studied for the purpose of elucidating the recently discussed unknown radical source in smog chambers. Nitrous acid and nitric oxide were found to be formed by the reaction of NO2 and H2O in an evacuable and bakable smog chamber. No nitric acid was observed in the gas phase. The reaction is not stoichiometric and is thought to be a heterogeneous wall reaction. The reaction rate is first order with respect to NO2 and H2O, and the concentrations of HONO and NO initially increase linearly with time. The same reaction proceeds with a different rate constant in a quartz cell, and the reaction of NO2 and H218O gave H18ONO exclusively. Taking into consideration the heterogeneous reaction of NO2 and H2O, the upper limit of the rate constant of the third-order reaction NO + NO2 + H2O → 2HONO was deduced to be (3.0 ± 1.4) × 10?10 ppm?2-min?1, which is one order of magnitude smaller than the previously reported value. Nitrous acid formed by the heterogeneous dark reaction of NO2 and H2O should contribute significantly to both an initially present HONO and a continuous supply of OH radicals by photolysis in smog chamber experiments.  相似文献   

8.
Nitrate radicals, NO3, were produced for the first time by 193 nm laser flash photolysis of N2O5 and HNO3. Detection was achieved due to NO3's strong absorption at 622.7 nm confirmed by measurements of the absorption spectrum in the range of 617–625 nm using both NO3 precursors. Time‐resolved kinetic studies allowed the determination of room temperature rate coefficients for the reactions of NO3 with 2‐methylbut‐2‐ene and NO2 of (1.28 ± 0.11) × 10?11 and (8.4 ± 1.2) × 10?13 cm3 molecule?1 s?1, respectively. The rate coefficients compare well to previous measurements with alternative techniques, suggesting that the reported method is valid and may be applied in follow‐up studies. The rate coefficient for 2‐methylbut‐2‐ene is compared to previous measurements and predictions for the alkene as well as the related alkenol. The new data are consistent with a previously suggested deactivation of the reactive site of the double bond if adjacent to an OH group. A calculated atmospheric lifetime for 2‐methylbut‐2‐ene with respect to NO3‐initiated oxidation of less than 3 min suggests predominant removal by NO3 in the atmosphere.  相似文献   

9.
Flash photolysis of CH3CHO and H2CO in the presence of NO has been investigated by the intracavity laser spectroscopy technique. The decay of HNO formed by the reaction HCO + NO → HNO + CO was studied at NO pressures of 6.8–380 torr. At low NO pressure HNO was found to decay by the reaction HNO + HNO → N2O + H2O. The rate constant of this reaction was determined to be k1 = (1.5 ± 0.8) × 10?15 cm3/s. At high NO pressure the reaction HNO + NO → products was more important, and its rate constant was measured to be k2 = (5 ± 1.5) × 10?19 cm3/s. NO2 was detected as one of the products of this reaction. Alternative mechanisms for this reaction are discussed.  相似文献   

10.
Reactions of HCCCO and NCCO radicals with O2 have been studied by a combination of pulsed laser photolysis and photoionization mass spectrometry. HCCCO was produced by 193‐nm photolysis of methylpropiolate or 3‐butyn‐2‐one, and NCCO was formed by 193‐nm photolysis of acetylcyanide. The rate constants obtained at 298 ± 3 K were (6.5 ± 0.7) × 10?12 cm3 molecule?1 s?1 for the HCCCO + O2 reaction, and no pressure dependence was observed between 1.5 and 16 Torr of N2 as a bath gas. Because HCO and HCCO radicals were observed as reaction products, it was confirmed that the reaction proceeds by a two‐body reaction. On the other hand, the rate constants of NCCO with O2 depended on the total pressure and were (5.4–8.8) × 10?13 cm3 molecule?1 s?1 for total pressures 2.0–15.5 Torr of N2, confirming that the reaction proceeds by a three‐body process. © 2001 John Wiley & Sons, Inc. Int J Chem Kinet 33: 440–448, 2001  相似文献   

11.
The recombination reaction O + O2 → O3 was studied by laser flash photolysis of pure O2 in the pressure range 3–20 atm, and of N2O? O2 mixtures in the bath gases Ar, N2, (CO2, and SF6) in the pressure range 3–200 atm. Fall-off curves of the reaction have been derived. Low-pressure rate coefficients were found to agree well with literature data. A high-pressure rate coefficient of k = (2.8 ± 1.0) × 10?12 cm3 molecule?1 s?1 was obtained by extrapolation.  相似文献   

12.
The rate constant of the reaction of OH with DMS has been measured relative to OH + ethene in a 420 l reaction chamber at 760 torr total pressure and 298 ± 3 K in N2 + O2 buffer gas using the 254 nm photolysis of H2O2 as the OH source. In agreement with a recent absolute rate determination of the reaction the measured effective rate constant was found to increase with increasing partial pressure of O2 in the system, for 760 torr air a rate constant of (8.0 ± 0.5) × 10?12 cm3 s?1 was obtained. Product studies have been performed on the reaction in air using FTIR absorption spectrometry for detection of reactants and products. On a molar basis, SO2 was formed with a yield of 70% and dimethyl sulfone (CH3SO2CH3) with a yield of approximately 20%. These results are considerably different to those obtained in other product studies which were carried out in the presence of NOx. These differences are compared and their relevance for the atmospheric oxidation mechanisms of DMS is discussed.  相似文献   

13.
An upper limit for the reaction rate of CO with the nitrate radical NO3 has been determined equal to 4 × 10?19 cm+3 molec?1 s?1 at 295 ± 2 K. In the experiment the isotopic species C13O16 and C13O18 mixed at 1–2 ppmv level in synthetic air have been reacted with NO3 and the reaction followed using long path infrared absorption FT spectroscopy. The result is of interest in the studies on the role played by NO3 in nighttime tropospheric chemistry.  相似文献   

14.
The thermal decomposition of CCl3O2NO2 has been reevaluated based on new rate data for the reaction of Cl + NO2. The revised rate coefficient for CCl3O2NO2 thermal decay at about 1 atm total pressure (mainly N2) is 1.42 × 1016 × exp(?11500/T) s?1 from 268–298 K.  相似文献   

15.
The rate coefficient for the reaction OH + HO2 =H2O + O2 has been determined from measurements of the steady-state absorption of HO2 at 210 nm, in low-frequency square-wave modulated photolysis of O3 + H2O mixtures. The value obtained was (9.9 ± 2.5) × 10?11 cm3 molecule?1 s?1 at 308 K and 1 atm pressure.  相似文献   

16.
Rate constants for the gas-phase reactions of the biogenically emitted monoterpene β-phellandrene with OH and NO3 radicals and O3 have been measured at 297 ± 2 K and atmospheric pressure of air using relative rate methods. The rate constants obtained were (in cm3 molecule?1 s?1 units): for reaction with the OH radical, (1.68 ± 0.41) × 10?10; for reaction with the NO3 radical, (7.96 ± 2.82) × 10?12; and for reaction with O3, (4.77 ± 1.23) × 10?17, where the error limits include the estimated uncertainties in the reference reaction rate constants. Using these rate constants, the lifetime of β-phellandrene in the lower troposphere due to reaction with these species is calculated to be in the range of ca. 1–8 h, with the OH radical reaction being expected to dominate over the O3 reaction as a loss process for β-phellandrene during daylight hours.  相似文献   

17.
The rate coefficients for the reaction of O(3P) with the biogenic hydrocarbons Δ3-carene, α-pinene, and isoprene have been measured using a direct method for the first time. O(3P) was generated from the pulsed photolysis of NO2 or O3 at 308 nm, and measured by resonance fluorescence at 131 nm. Rate coefficients at room temperature for the biogenics are similar: (3.4 ± 0.6) × 10?11, (3.7 ± 0.6) × 10?11, and (3.5 ± 0.6) × 10?11 cm3 molec?1 s?1, for Δ3-carene, α-pinene, and isoprene, respectively. The rate coefficients for the reaction of O(3P) with NO2 and ethene were also measured with the same method, and these values are within 4% and 10% of the currently recommended values, respectively. The correlation between OH and O(3P)-alkene reaction rate coefficients is updated and discussed. © 1995 John Wiley & Sons, Inc.  相似文献   

18.
Rate constants for the gas-phase reactions of the four oxygenated biogenic organic compounds cis-3-hexen-1-ol, cis-3-hexenylacetate, trans-2-hexenal, and linalool with OH radicals, NO3 radicals, and O3 have been determined at 296 ± 2 K and atmospheric pressure of air using relative rate methods. The rate constants obtained were (in cm3 molecule?1 s?1 units): cis-3-hexen-1-ol: (1.08 ± 0.22) × 10?10 for reaction with the OH radical; (2.72 ± 0.83) × 10?13 for reaction with the NO3 radical; and (6.4 ± 1.7) × 10?17 for reaction with O3; cis-3-hexenylacetate: (7.84 ± 1.64) × 10?11 for reaction with the OH radical; (2.46 ± 0.75) × 10?13 for reaction with the NO3 radical; and (5.4 ± 1.4) × 10?17 for reaction with O3; trans-2-hexenal: (4.41 ± 0.94) × 10?11 for reaction with the OH radical; (1.21 ± 0.44) × 10?14 for reaction with the NO3 radical; and (2.0 ± 1.0) × 10?18 for reaction with O3; and linalool: (1.59 ± 0.40) × 10?10 for reaction with the OH radical; (1.12 ± 0.40) × 10?11 for reaction with the NO3 radical; and (4.3 ± 1.6) × 10?16 for reaction with O3. Combining these rate constants with estimated ambient tropospheric concentrations of OH radicals, NO3 radicals, and O3 results in calculated tropospheric lifetimes of these oxygenated organic compounds of a few hours. © 1995 John Wiley & Sons, Inc.  相似文献   

19.
Absolute rate coefficients for the reaction between the important environmental free radical oxidant NO3. and a series of N‐ and C‐protected amino acids, di‐ and tripeptides were determined using 355 nm laser flash photolysis of cerium(IV) ammonium nitrate in the presence of the respective substrates in acetonitrile at 298±1 K. Through combination with computational studies it was revealed that the reaction with acyclic aliphatic amino acids proceeds through hydrogen abstraction from the α‐carbon, which is associated with a rate coefficient of about 1.8×106 m ?1 s?1 per abstractable hydrogen atom. The considerably faster reaction with phenylalanine [k=(1.1±0.1)×107 m ?1 s?1] is indicative for a mechanism involving electron transfer. An unprecedented amplification of the rate coefficient by a factor of 7–20 was found with di‐ and tripeptides that contain more than one phenylalanine residue. This suggests a synergistic effect between two aromatic rings in close vicinity, which makes such peptide sequences highly vulnerable to oxidative damage by this major environmental pollutant.  相似文献   

20.
A steady-state system involving the photolysis of NO2 in an excess of I2 as a source of IO radicals has been used to study the reaction IO + DMS in 760 Torr N2 at 296 K. IO radicals were found to react rapidly with DMS, one molecule of DMSO being produced for each molecule of DMS consumed. Numerical analysis of the experimental results yielded a rate constant of (3.0 ± 1.5) × 10?11 cm3 s?1 for the reaction IO + DMS → DMSO + I.  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