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.
Using a relative rate method, rate constants for the gas-phase reactions of 2-methyl-3-buten-2-ol (MBO) with OH radicals, ozone, NO3 radicals, and Cl atoms have been investigated using FTIR. The measured values for MBO at 298±2 K and 740±5 torr total pressure are: kOH=(3.9±1.2)×10−11 cm3 molecule−1 s−1, kO3=(8.6±2.9)×10−18 cm3 molecule−1 s−1, k=(8.6±2.9)×10−15 cm3 molecule−1 s−1, and kCl=(4.7±1.0)×10−10 cm3 molecule−1 s−1. Atmospheric lifetimes have been estimated with respect to the reactions with OH, O3, NO3, and Cl. The atmospheric relevance of this compound as a precursor for acetone is, also, briefly discussed. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet: 30: 589–594, 1998  相似文献   

2.
Using relative rate methods, rate constants have been measured for the gas-phase reactions of 3-methylfuran with NO3 radicals and O3 at 296 ± 2 K and atmospheric pressure of air. The rate constants determined were (1.31 ± 0.461) × 10−11 cm3 molecule−1 s−1 for the NO3 radical reaction and (2.05 ± 0.52) × 10−17 cm3 molecule−1 s−1 for the O3 reaction, where the indicated errors include the estimated overall uncertainties in the rate constants for the reference reactions. Based on the cyclohexanone plus cyclohexanol yield in the presence of sufficient cyclohexane to scavenge > 95% of OH radicals formed, it is estimated that the O3 reaction leads to the formation of OH radicals with a yield of 0.59, uncertain to a factor of ca. 1.5. In the troposphere, 3-methylfuran will react dominantly with the OH radical during daylight hours, and with the NO3 radical during nighttime hours for nighttime NO3 radical concentrations > 107 molecule cm −3. © 1996 John Wiley & Sons, Inc.  相似文献   

3.
The atmospheric role of photochemical processes involving NO2 beyond its dissociation limit (398 nm) is controversial. Recent experiments have confirmed that excited NO2* beyond 420 nm reacts with water according to NO2*+H2O→HONO+OH. However, the estimated kinetic constant for this process in the gas phase is quite small (k≈10−15–3.4×10−14 cm3 molecule−1 s−1) suggesting minor atmospheric implications of the formed radicals. In this work, ab initio molecular dynamics simulations of NO2 adsorbed at the air–water interface reveal that the OH production rate increases by about 2 orders of magnitude with respect to gas phase, attaining ozone reference values for NO2 concentrations corresponding to slightly polluted rural areas. This finding substantiates the argument that chemistry on clouds can be an additional source of OH radicals in the troposphere and suggests directions for future laboratory experimental studies.  相似文献   

4.
Rate constants have been measured at 296 ± 2 K for the gas‐phase reactions of camphor with OH radicals, NO3 radicals, and O3. Using relative rate methods, the rate constants for the OH radical and NO3 radical reactions were (4.6 ± 1.2) × 10−12 cm3 molecule−1 s−1 and <3 × 10−16 cm3 molecule−1 s−1, respectively, where the indicated error in the OH radical reaction rate constant includes the estimated overall uncertainty in the rate constant for the reference compound. An upper limit to the rate constant for the O3 reaction of <7 × 10−20 cm3 molecule−1 s−1 was also determined. The dominant tropospheric loss process for camphor is calculated to be by reaction with the OH radical. Acetone was identified and quantified as a product of the OH radical reaction by gas chromatography, with a formation yield of 0.29 ± 0.04. In situ atmospheric pressure ionization tandem mass spectrometry (API‐MS) analyses indicated the formation of additional products of molecular weight 166 (dicarbonyl), 182 (hydroxydicarbonyl), 186, 187, 213 (carbonyl‐nitrate), 229 (hydroxycarbonyl‐nitrate), and 243. A reaction mechanism leading to the formation of acetone is presented, as are pathways for the formation of several of the additional products observed by API‐MS. © 2000 John Wiley and Sons, Inc. Int J Chem Kinet 33: 56–63, 2001  相似文献   

5.
The rate constants for the gas-phase reactions of di-tert-butyl ether (DTBE) with chlorine atoms, hydroxyl radicals, and nitrate radicals have been determined in relative rate experiments using FTIR spectroscopy. Values of k(DTBE+CI) = (1.4 ± 0.2) × 10−10,k(DTBE+OH) = (3.7 ± 0.7) × 10−12, and k(DTBE+N03) = (2.8 ± 0.9) × 10−16 cm3 molecule−1 s−1 were obtained. Tert-butyl acetate was identified as the major product of both Cl atom and OH radical initiated oxidation of DTBE in air in the presence of NOx. The molar tert-butyl acetate yield was 0.85 ± 0.11 in the Cl atom experiments and 0.84 ± 0.11 in OH radical experiments. As part of this work the rate constant for reaction of Cl atoms with tert-butyl acetate at 295 K was determined to be (1.6 ± 0.3) × 10−11 cm3 molecule−1 s−1. The stated errors are two standard deviations (2σ). © 1996 John Wiley & Sons, Inc.  相似文献   

6.
Rate constants for the reactions of OH, NO3, and O3 with pinonaldehyde and the structurally related compounds 3-methylbutanal, 3-methylbutan-2-one, cyclobutyl-methylketone, and 2,2,3-trimethyl-cyclobutyl-1-ethanone have been measured at 300±5 K using on-line Fourier transform infrared spectroscopy. The rate constants obtained for the reactions with pinonaldehyde were: kOH=(9.1±1.8)×10−11 cm3 molecule−1 s−1, kNO3=(5.4±1.8)×10−14 cm3 molecule−1 s−1, and kO3=(8.9±1.4)×10−20 cm3 molecule−1 s−1. The results obtained indicate a chemical lifetime of pinonaldehyde in the troposphere of about two hours under typical daytime conditions, [OH]=1.6×106 molecule cm−3. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet 29: 527–533, 1997.  相似文献   

7.
The kinetics of the gas-phase reactions of OH radicals, NO3 radicals, and O3 with indan, indene, fluorene, and 9,10-dihydroanthracene have been studied at 297 ± 2 K and atmospheric pressure of air. The rate constants, or upper limits thereof, for the O3 reactions were (in cm3 molecule−1 s−1 units): indan, < 3 × 10−19; indene, (1.7 ± 0.5) × 10−16, fluorene, < 2 × 10−19; and 9,10-dihydroanthracene, (9.0 ± 2.0) × 10−19. Using a relative rate method, the rate constants for the OH radical and NO3 radical reactions, respectively, were (in cm3 molecule−1 s−1 units): indan, (1.9 ± 0.5) × 10−11 and (6.6 ± 2.0) × 10−15; indene, (7.8 ± 2.0) × 10−11 and (4.1 ± 1.5) × 10−12; fluorene, (1.6 ± 0.5) × 10−11 and (3.5 ± 1.2) × 10−14; and 9,10-dihydroanthracene, (2.3 ± 0.6) × 10−11 and (1.2 ± 0.4) × 10−12. These kinetic data were used to assess the relative contributions of the various reaction pathways. © 1997 John Wiley & Sons, Inc. Int J Chem Kinet 29: 299–309, 1997.  相似文献   

8.
By using relative rate methods, rate constants for the gas‐phase reactions of OH and NO3 radicals with propanal, butanal, pentanal, and hexanal have been measured at 296 ± 2 K and atmospheric pressure of air. By using methyl vinyl ketone as the reference compound, the rate constants obtained for the OH radical reactions (in units of 10−12 cm3 molecule−1 s−1) were propanal, 20.2 ± 1.4; butanal, 24.7 ± 1.5; pentanal, 29.9 ± 1.9; and hexanal, 31.7 ± 1.5. By using methacrolein and 1‐butene as the reference compounds, the rate constants obtained for the NO3 radical reactions (in units of 10−15 cm3 molecule−1 s−1) were propanal, 7.1 ± 0.4; butanal, 11.2 ± 1.5; pentanal, 14.1 ± 1.6; and hexanal, 14.9 ± 1.3. The dominant tropospheric loss process for the aldehydes studied here is calculated to be by reaction with the OH radical, with calculated lifetimes of a few hours during daytime. © 2000 John Wiley & Sons, Inc. Int J Chem Kinet 32: 79–84, 2000  相似文献   

9.
Using a relative rate method, rate constants have been measured for the gas-phase reactions of the OH radical with 1-hexanol, 1-methoxy-2-propanol, 2-butoxyethanol, 1,2-ethanediol, and 1,2-propanediol at 296±2 K, of (in units of 10−12 cm3 molecule−1 s−1): 15.8±3.5; 20.9±3.1; 29.4±4.3; 14.7±2.6; and 21.5±4.0, respectively, where the error limits include the estimated overall uncertainties in the rate constants for the reference compounds. These OH radical reaction rate constants are higher than certain of the literature values, by up to a factor of 2. Rate constants were also measured for the reactions of 1-methoxy-2-propanol and 2-butoxyethanol with NO3 radicals and O3, with respective NO3 radical and O3 reaction rate constants (in cm3 molecule−1 s−1 units) of: 1-methoxy-2-propanol, (1.7±0.7)×10−15, and <1.1×10−19; and 2-butoxyethanol, (3.0±1.2)×10−15, and <1.1×10−19. The dominant tropospheric loss process for the alcohols, glycols, and glycol ethers studied here is calculated to be by reaction with the OH radical, with lifetimes of 0.4–0.8 day for a 24 h average OH radical concentration of 1.0×106 molecule cm−3. © 1998 John Wiley & Sons, Inc. Int J Chem Kinet 30: 533–540, 1998  相似文献   

10.
Kinetic studies on reactions of ozone with trans-1,2-dichloroethene (DCE) and vinyl chloride (VC) were performed in air. In the presence of scavengers of radicals, such as CH3CHO, the rates for both reactions are second order (first order in each reactant). Observed rate constants are (1.80 ± 0.29) X 10?19 cm3/molecule·s for DCE and (2.45 ± 0.45) × 10?19 cm3/molecule·s for VC. In the presence of CH3CHO, propene ozonide \documentclass{article}\pagestyle{empty}\begin{document}$ ({\rm CH}_3 \overline {{\rm CHOOCH}_2 {\rm O}}) $\end{document} was observed as a product in the case of VC. Peroxyformic acid (HC)O(OOH) was detected in both reactions. The Criegee mechanism was proposed to play a major role in the reaction of ozone with chloroolefins. The branching ratio of O3 + CH2=CHCl → CH2OO + HCOCl (6a), CHClOO + HCHO (6b) was obtained as 76:24, and the fraction of the stabilized CH2OO was estimated to be 0.25 of that produced in reaction (6a).  相似文献   

11.
The kinetics and nitroarene product yields of the gas-phase reactions of naphthalene-d8, fluoranthene-d10, and pyrene with OH radicals in the presence of NOx and in N2O5? NO3? NO2? air mixtures have been investigated at 296 ± 2 K and atmospheric pressure of air. Using a relative rate method, naphthalene-d8 was shown to react in N2O5? NO3? NO2? air mixtures a factor of 1.22 ± 0.10 times faster than did naphthalene, with the 1- and 2-nitronaphthalene-d7 product yields being similar to those of 1- and 2-nitronaphthalene from naphthalene. From the measured PAH concentrations and the nitroarene product yields, formation yields of 2-, 7-, and 8-nitrofluoranthene-d9 and 2- and 4-nitropyrene of 0.03, 0.01, 0.003, 0.005, and 0.0006, respectively, were determined from the OH radical-initiated reactions. Effective rate constants for the reactions of fluoranthene-d10 and pyrene with N2O5 in N2O5? ;NO3? NO2? air mixtures of ca. 1.8 × 10?17 cm3 molecule?1 s?1 and ca. 5.6 × 10?17 cm3 molecule?1 s?1, respectively, were derived. Formation yields of 2-nitrofluoranthene-d9 and 4-nitropyrene of ca. 0.24 and ca. 0.0006, respectively, were estimated for these reaction systems. 2-Nitropyrene was also observed to be formed in these N2O5? NO3? NO2 reactions, but was found to be a function of the NO2 concentration and, therefore, would be a negligible product under ambient NO2 concentrations. These product and kinetic data are consistent with ambient air measurements of the nitroarene concentrations.  相似文献   

12.
Using a relative rate method, rate constants for the gas-phase reactions of the NO3 radical with methacrolein and methyl vinyl ketone were determined to be (4.4 ± 1.7) × 10−15 cm3 molecule−1 s−1 and <6 × 10−16 cm3 molecule−1 s−1, respectively, at 296 ± 2 K. The molar formation yields of methacrolein and methyl vinyl ketone from the gas-phase reaction of the NO3 radical with isoprene at 296 ± 2 K and atmospheric pressure of air were measured to be 0.035 ± 0.014 each. The tropospheric implications of these kinetic and product data are discussed, and it is concluded that the nighttime NO3 radical reactions with methacrolein and methyl vinyl ketone are not important. However, during nighttime the formation of methacrolein and methyl vinyl ketone from the reaction of isoprene with the NO3 radical may dominate over their formation from the O3 reaction with isoprene. Atmospheric pressure ionization tandem mass spectrometry (API-MS/MS) was used to investigate the products of the reactions of the NO3 radical with isoprene and isoprene-d8, and C5-nitrooxycarbonyl(s) (e.g., O2NOCH2C(CH3) (DOUBLEBOND) CHCHO), C5-hydroxynitrate(s) (e.g., O2NOCH2C(CH3)(DOUBLEBOND) CHCH2OH), C5-nitrooxyhydroperoxide(s) (e.g., O2NOCH2C(CH3)(DOUBLEBOND) CHCH2OOH), and C5-hydroxycarbonyl(s) (e.g., HOCH2CH(DOUBLEBOND) C(CH3)CHO) and their deuterated analogs were observed from these reactions. © 1996 John Wiley & Sons, Inc.  相似文献   

13.
The gas-phase reaction of ozone with eight alkenes including six 1,1-disubstituted alkenes has been investigated at ambient T (285–298 K) and p = 1 atm. of air. The reaction rate constants are, in units of 10−18 cm3 molecule−1 s−1, 9.50 ± 1.23 for 3-methyl-1-butane, 13.1. ± 1.8 for 2-methyl-1-pentene, 11.3 ± 3.2 for 2-methyl-1,3-butadiene (isoprene), 7.75 ± 1.08 for 2,3,3-trimethyl-1-butene, 3.02 ± 0.52 for 3-methyl-2-isopropyl-1-butene, 3.98 ± 0.43 for 3,4-diethyl-2-hexene, 1.39 ± 17 for 2,4,4-trimethyl-2-pentene, and >370 for (cis + trans)-3,4-dimethyl-3-hexene. For isoprene, results from this study and earlier literature data are consistent with: k (cm3 molecule−1 s−1) = 5.59 (+ 3.51, &minus 2.16) × 10−15 e(−3606±279/RT), n = 28, and R = 0.930. The reactivity of the other alkenes, six of which have not been studied before, is discussed in terms of alkyl substituent inductive and steric effects. For alkenes (except 1,1-disubstituted alkenes) that bear H, CH3, and C2H5 substituents, reactivity towards ozone is related to the alkene ionization potential: In k<(10−18 cm3 molecule−1 s−1) = (32.89 ± 1.84) − (3.09 ± 0.20) IP (eV), n = 12, and R = 0.979. This relationship overpredicts the reactivity of C≥3 1-alkenes, of 1,1-disubstituted alkenes, and of alkenes with bulky substituents, for which reactivity towards ozone is lower due to substituent steric effects. The atmospheric persistence of the alkenes studied is briefly discussed. © 1996 John Wiley & Sons, Inc.  相似文献   

14.
The atmospheric oxidation of amines proceeds via initial radical attack at C–H or N–H bonds to form carbon- and nitrogen-centered radicals, respectively. It is conventionally assumed that nitrogen-centered aminyl radicals react slowly with oxygen in the troposphere and associate predominantly with the radicals NO and NO2 to form toxic nitrosamines and nitramines. We have used theoretical kinetic modeling techniques to study the prototypical CH3NH + O2 reaction and have shown that it proceeds to CH2NH + HO2 under tropospheric conditions with a rate coefficient of 3.6 × 10−17 cm3 molecule−1 s−1. Although this value is low compared to the competing NOx reactions (∼10−11 cm3 molecule−1 s−1), the much higher concentration of O2 versus NOx in air makes it the dominant process in the atmospheric oxidation of methylamine for NOx concentrations below 100 ppb. The mechanism identified here is available to amines with primary, secondary, and tertiary α carbons and suggests that they may be less likely to form nitramines and nitrosamines than is currently thought.  相似文献   

15.
Kinetics for reactions of phenoxy radical, C6H5O, with itself and with O3 were examined at 298 K and low pressure (1 Torr) using discharge flow coupled with mass spectrometry (DF/MS). The rate constant for the phenoxy radical self‐reaction was determined to be k1 = (1.49 ± 0.53) × 10−11 cm3 molecule−1 s−1 defined by d[C6H5O]/dt=−2 k1[C6H5O]2. The rate constant for the C6H5O reaction with O3 was measured to be k2 = (2.86 ± 0.35) × 10−13 cm3 molecule−1 s−1, which may be a lower limit value. Because of much higher atmospheric abundance of ozone than that of both NO and phenoxy, the reaction of C6H5O with ozone may represent the principal fate of the phenoxy radical in the atmosphere. Products from reaction of C6H5O + C6H5O, NO, and NO2 were also investigated, and (C6H5O)2 (m/e = 186), C6H5O(NO) (m/e = 123), and C6H5O(NO2) (m/e = 139) adducts were observed as products for the reactions of C6H5O with itself, NO, and NO2, respectively. © 1999 John Wiley & Sons, Inc. Int J Chem Kinet 31: 65–72, 1999  相似文献   

16.
The atmospheric chemistry of methyl ethyl ether, CH3CH2OCH3, was examined using FT‐IR/relative‐rate methods. Hydroxyl radical and chlorine atom rate coefficients of k (CH3CH2OCH3+OH) = (7.53 ± 2.86) × 10−12 cm3 molecule−1 s−1 and k (CH3CH2OCH3+Cl) = (2.35 ± 0.43) × 10−10 cm3 molecule−1 s−1 were determined (297 ± 2 K). The Cl rate coefficient determined here is 30% lower than the previous literature value. The atmospheric lifetime for CH3CH2OCH3 is approximately 2 days. The chlorine atom–initiated oxidation of CH3CH2OCH3 gives CH3C(O)H (9 ± 2%), CH3CH2OC(O)H (29 ± 7%), CH3OC(O)H (19 ± 7%), and CH3C(O)OCH3 (17 ± 7%). The IR absorption cross section for CH3CH2OCH3 is (7.97 ± 0.40) × 10−17 cm molecule−1 (1000–3100 cm−1). CH3CH2OCH3 has a negligible impact on the radiative forcing of climate.  相似文献   

17.
The reactions between OH radicals and hydrogen halides (HCl, HBr, HI) have been studied between 298 and 460 K by using a discharge flow-electron paramagnetic resonance technique. The rate constants were found to be kHCl(298 K) = (7.9 ± 1.3) × 10−13 cm3 molecule−1 s−1 with a weak positive temperature dependence, kHBr (298-460 K) = (1.04 ± 0.2) × 10−11 cm3 molecule−1 s−1, and kHI(298 K) = (3.0 ± 0.3) × 10−11 cm3 molecule−1 s−1, respectively. The homogeneous nature of these reactions has been experimentally tested.  相似文献   

18.
The gas phase reaction of OH radicals with hydrogen iodide (HI) has been studied using a Laser Photolysis-Resonance Fluorescence (LP-RF) apparatus, recently developed in our group. The measured rate constant at 298 K was (2.7 ± 0.2) × 10−11 cm3 molecule−1 s−1. This rate constant is compared with the ones of the reactions OH + HCl and OH + HBr. The role of the reaction OH + HI in marine tropospheric chemistry is discussed. In addition, the LP-RF apparatus was tested and validated by measuring the following rate constants (in cm3 molecule−1 s−1 units): 𝓀(OH + HNO3) = (1.31 ± 0.06) × 10−13 at p = 27 and 50 Torr of argon and 𝓀(OH + C3H8) = (1.22 ± 0.08) × 10−12. These rate constants are in very good agreement with the literature data.  相似文献   

19.
The homogeneous gas-phase reaction of N2H4 with O3 in air atmospheric pressure has been used to generate OH radicals in the dark, allowing the determination of relative OH radical rate constants for compounds which photolyze rapidly. This technique was first validated by determining the OH radical rate constant ratios for n-butane/cyclohexane and methanol/dimethyl ether, both of which are in excellent agreement with the literature values. The rate constant for the reaction of OH radicals with methyl nitrite at 300 ± 3 K was then determined relative to those for the reaction of OH radicals with n-hexane and dimethyl ether. The resulting rate constant of 1.8 × 10?13 cm3/molecule·s is about seven times lower than those of previous measurements which employed a different nonphotolytic relative rate method.  相似文献   

20.
Rate coefficients for the reactions of OH with n, s, and iso-butanol have been measured over the temperature range 298 to ∼650 K. The rate coefficients display significant curvature over this temperature range and bridge the gap between previous low-temperature measurements with a negative temperature dependence and higher temperature shock tube measurements that have a positive temperature dependence. In combination with literature data, the following parameterizations are recommended: k1,OH + n-butanol(T) = (3.8 ± 10.4) × 10−19T2.48 ± 0.37exp ((840 ± 161)/T) cm3 molecule−1 s−1 k2,OH + s-butanol(T) = (3.5 ± 3.0) × 10−20T2.76 ± 0.12exp ((1085 ± 55)/T) cm3 molecule−1 s−1 k3,OH + i-butanol(T) = (5.1 ± 5.3) × 10−20T2.72 ± 0.14exp ((1059 ± 66)/T) cm3 molecule−1 s−1 k4,OH + t-butanol(T) = (8.8 ± 10.4) × 10−22T3.24 ± 0.15exp ((711 ± 83)/T) cm3 molecule−1 s−1 Comparison of the current data with the higher shock tube measurements suggests that at temperatures of ∼1000 K, the OH yields, primarily from decomposition of β-hydroxyperoxy radicals, are ∼0.3 (n-butanol), ∼0.3 (s-butanol) and ∼0.2 (iso-butanol) with β-hydroxyperoxy decompositions generating OH, and a butene as the main products. The data suggest that decomposition of β-hydroxyperoxy radicals predominantly occurs via OH elimination.  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