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 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   

2.
The kinetics of the atmospherically important gas-phase reactions of acenaphthene and acenaphthylene with OH and NO3 radicals, O3 and N2O5 have been investigated at 296 ± 2 K. In addition, rate constants have been determined for the reactions of OH and NO3 radicals with tetralin and styrene, and for the reactions of NO3 radicals and/or N2O5 with naphthalene, 1- and 2-methylnaphthalene, 2,3-dimethylnaphthalene, toluene, toluene-α,α,α-d3 and toluene-d8. The rate constants obtained (in cm3 molecule?1 s?1 units) at 296 ± 2 K were: for the reactions of O3; acenaphthene, <5 × 10?19 and acenaphthylene, ca. 5.5 × 10?16; for the OH radical reactions (determined using a relative rate method); acenaphthene, (1.03 ± 0.13) × 10?10; acenaphthylene, (1.10 ± 0.11) × 10?10; tetralin, (3.43 ± 0.06) × 10?11 and styrene, (5.87 ± 0.15) × 10?11; for the reactions of NO3 (also determined using a relative rate method); acenaphthene, (4.6 ± 2.6) × 10?13; acenaphthylene, (5.4 ± 0.8) × 10?12; tetralin, (8.6 ± 1.3) × 10?15; styrene, (1.51 ± 0.20) × 10?13; toluene, (7.8 ± 1.5) × 10?17; toluene-α,α,α-d3, (3.8 ± 0.9) × 10?17 and toluene-d8, (3.4 ± 1.9) × 10?17. The aromatic compounds which were observed to react with N2O5 and the rate constants derived were (in cm3 molecule?1 s?1 units): acenaphthene, 5.5 × 10?17; naphthalene, 1.1 × 10?17; 1-methylnaphthalene, 2.3 × 10?17; 2-methylnaphthalene, 3.6 × 10?17 and 2,3-dimethylnaphthalene, 5.3 × 10?17. These data for naphthylene and the alkylnaphthalenes are in good agreement with our previous absolute and relative N2O5 reaction rate constants, and show that the NO3 radical reactions with aromatic compounds proceed by overall H-atom abstraction from substituent-XH bonds (where X = C or O), or by NO3 radical addition to unsaturated substituent groups while the N2O5 reactions only occur for aromatic compounds containing two or more fused six-membered aromatic rings.  相似文献   

3.
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   

4.
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   

5.
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   

6.
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   

7.
Unlike the chemistry underlying the self‐coupling of phenoxy (C6H5O) radicals, there are very limited kinetics data at elevated temperatures for the reaction of the phenoxy radical with other species. In this study, we investigate the addition reactions of O2, OH, and NO2 to the phenoxy radical. The formation of a phenoxy‐peroxy is found to be very slow with a rate constant fitted to k = 1.31 × 10?20T2.49 exp (?9300/T) cm3/mol/s in the temperature range of (298–2,000 K) where the addition occurs predominantly at the ortho site. Our rate constant is in line with the consensus of opinions in the literature pointing to the observation of no discernible reaction between the oxygen molecule and the resonance‐stabilized phenoxy radical. Addition of OH at the ortho and para sites of the phenoxy radical is found to afford adducts with sizable well depths of 59.8 and 56.0 kcal/mol, respectively. The phenoxy‐NO2 bonds are found to be among the weakest known phenoxy‐radical bonds (1.7–8.7 kcal/mol). OH‐ and O2‐initiated mechanisms for the degradation of atmospheric phenoxy appear to be negligible and the fate of atmospheric phenoxy is found to be controlled by its reaction with NO2. © 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2011  相似文献   

8.
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   

9.
A differential optical absorption spectrometer (DOAS) system was operated at Long Beach, CA during the 1987 SCAQS Fall episodes to measure atmospheric concentrations of nitrous acid (HONO), as well as ambient levels of nitrogen dioxide (NO2) and formaldehyde (HCHO). The rapid scanning (-3000 spectra per min) spectrometer was interfaced to a 25 m basepath open, multiple reflection system operated routinely at a total optical path of 800 m. During several of the Fall episodes at Long Beach, levels of gaseous HONO were the highest (>15 ppb) reported to date by the DOAS technique. Although approximately half, to all, of the measured nighttime HONO concentrations could be accounted for by proposed heterogeneous formation pathways involving NO2, HONO concentrations correlated well with primary pollutants such as CO and NO, suggesting that elevated nighttime HONO concentrations in the western end of the Los Angeles basin may be influenced by emissions of HONO from combustion sources. This has significant implications for models which assume HONO arises only from secondary formation, rather than a combination of direct emissions and atmospheric reactions. Estimates of hydroxyl (OH) radical production show that photolysis of HONO shortly after sunrise on these episode days produces a large pulse of OH radicals at a time of the day when OH production from photolysis of O3 and HCHO is low. In terms of integrated OH radical production, HONO is of comparable importance to HCHO and much more important than O3 during these Fall periods.  相似文献   

10.
The kinetics of the gas-phase reactions of allyl chloride and benzyl chloride with the OH radical and O3 were investigated at 298 ± 2 K and atmospheric pressure. Direct measurements of the rate constants for reactions with ozone yielded values of ??(O3 + allyl chloride) = (1.60 ± 0.18) × 10?18 cm3 molecule?1 s?1 and ??(O3 + benzyl chloride) < 6 × 10?20 cm3 molecule?1 s?1. With the use of a relative rate technique and ethane as a scavenger of chlorine atoms produced in the OH radical reactions, rate constants of ??(OH + allyl chloride) = (1.69 ± 0.07) × 10?11 cm3 molecule?1 s?1 and ??(OH + benzyl chloride) = (2.80 ± 0.19) × 10?12 cm3 molecule?1 s?1 were measured. A study of the OH radical reaction with allyl chloride by long pathlength FT-IR absorption spectroscopy indicated that the co-products ClCH2CHO and HCHO account for ca. 44% of the reaction, and along with the other products HOCH2CHO, (ClCH2)2CO, and CH2 ? CHCHO account for 84 ± 16% of the allyl chloride reacting. The data indicate that in one atmosphere of air in the presence of NO the chloroalkoxy radical formed following OH radical addition to the terminal carbon atom of the double bond decomposes to yield HOCH2CHO and the CH2Cl radical, which becomes a significant source of the Cl atoms involved in secondary reactions. A product study of the OH radical reaction with benzyl chloride identified only benzaldehyde and peroxybenzoyl nitrate in low yields (ca. 8% and ?4%, respectively), with the remainder of the products being unidentified.  相似文献   

11.
The production of dimethyl sulfoxide (DMSO) and dimethyl sulfone (DMSO2) in the dimethyl sulfide (DMS) degradation scheme initiated by the hydroxyl (OH) radical has been shown to be very sensitive to nitrogen oxides (NOx) levels. In the present work we have explored the potential energy surfaces corresponding to several reaction pathways which yield DMSO2 from the CH3S(O)(OH)CH3 adduct [including the formation of CH3S(O)(OH)CH3 from the reaction of DMSO with OH] and the reaction channels that yield DMSO or/and DMSO2 from the CH3S(O2)(OH)CH3 adduct are also studied. The formation of the CH3S(O2)(OH)CH3 adduct from CH3S(OH)CH3 (DMS‐OH) and O2 was analyzed in our previous work. All these pathways due to the presence of NOx (NO and NO2) and also due to the reactions with O2, OH and HO2 are compared with the objective of inferring their kinetic relevance in the laboratory experiments that measure DMSO2 (and DMSO) formation yields. In particular, our theoretical results clearly show the existence of NOx‐dependent pathways leading to the formation of DMSO2, which could explain some of these experimental results in comparison with experimental measurements carried out in NOx‐free conditions. Our results indicate that the relative importance of the addition channel in the DMS oxidation process can be dependent on the NOx content of chamber experiments and of atmospheric conditions. © 2008 Wiley Periodicals, Inc. J Comput Chem, 2009  相似文献   

12.
Azulene, which is isomeric with naphthalene, was studied to determine whether it behaves like a polycyclic aromatic hydrocarbon or an alkene in its gas-phase reactions with OH and NO3 radicals and O3. Using relative rate methods, rate constants for the reactions of azulene with OH and NO3 radicals and O3 of (2.73 ± 0.56) × 10?10 cm3 molecule?1 s?1, (3.9) × 10?10 cm3 molecule?1 s?1, and <7 × 10?17 cm3 molecule?1 s?1, respectively, were obtained at 298 ± 2 K. The observation that the NO3 radical reaction did not involve NO2 in the rate determining step indicates that azulene behaves more like an alkene than a polycyclic aromatic hydrocarbon in this reaction. No conclusive evidence for the formation of nitroazulene(s) from either the OH or NO3 radical-initiated reaction of azulene (in the presence of NOx) was obtained.  相似文献   

13.
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   

14.
The gas-phase reaction of methyl vinyl ketone with the OH radical, in the presence of NOx, was investigated at 298 ± 2 K and atmospheric pressure of air. Glycolaldehyde and methylglyoxal were observed to be the major products, with a combined yield of 0.89 ± 0.16. The sum of the yields of the two other main products, formaldehyde and peroxyacetyl nitrate, were found to be essentially unity. The product yield data for glycolaldehyde and methylglyoxal indicate that OH radical addition to the terminal carbon atom of the >C?C< bond accounts for 72 ± 21% of the overall reaction, with the remaining 28 ± 9% proceeding via addition to the inner carbon atom of the double bond.  相似文献   

15.
The kinetics of the gas-phase reactions of naphthalene, 2-methylnaphthalene, and 2,3-dimethylnaphthalene with O3 and with OH radicals have been studied at 295 ± 1 K in one atmosphere of air. Upper limit rate constants for the O3 reactions of <3 × 10?19, <4 × 10?19, and <4 × 10?19 cm3 molecule?1 s?1 were obtained for naphthalene, 2-methylnaphthalene, and 2,3-dimethylnaphthalene, respectively. For the OH radical reactions, rate constants of (in units of 10?11 cm3 molecule?1 s?1) 2.59 ± 0.24, 5.23 ± 0.42, and 7.68 ± 0.48 were determined for naphthalene, 2±methylnaphthalene, and 2,3-dimethylnaphthalene, respectively. These data show that under atmospheric conditions these naphthalenes will react mainly with the OH radical, with life-times due to this reaction ranging from ca. 11 h for naphthalene to ca. 4 h for 2,3-dimethylnaphthalene.  相似文献   

16.
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   

17.
The aromatic ring-retaining products formed from the gas–phase reactions of the OH radical with benzene and toluene, in the presence of NOx, have been identified and their formation yields determined. These products, and their formation yields, are as follows: from benzene – phenol, 0.236 ± 0.044; nitrobenzene, {(0.0336 ± 0.0078) + (3.07 ± 0.92) × 10?16[NO2]}; from toluene – benzaldehyde, 0.0645 ± 0.0080; benzyl nitrate, 0.0084 ± 0.0017; o?cresol, 0.204 ± 0.027; m? + p?cresol, 0.048 ± 0.009; m-nitrotoluene, {(0.0135 ± 0.0029) + (1.90 ± 0.25) × 10?16[NO2]}, where the NO2 concentration is in molecule cm?3 units. The formation yields of o- and p-nitrotoluene from toluene were ca. 0.07 and 0.35 that of m-nitrotoluene, respectively. The observations that the nitro-aromatic yields do not extrapolate to zero as the NO2 concentration approaches zero are not consistent with current chemical mechanisms for these OH radical-initiated reactions, and suggest that under the experimental conditions employed in this study the hydroxycyclohexadienyl radicals formed from OH radical addition to the aromatic ring react with NO2 rather than with O2. However, these data concerning the nitroaromatic yields are consistent with our previous conclusions that many of the nitrated polycyclic aromatic hydrocarbons present in ambient air are formed, at least in part, in the atmosphere from OH radical reactions.  相似文献   

18.
A previous technique for the calculation of rate constants for the gas-phase reactions of the OH radical with organic compounds has been updated and extended to include sulfur- and nitrogen-containing compounds. The overall OH radical reaction rate constants are separated into individual processes involving (a) H-atom abstraction from C? H and O? H bonds in saturated organics, (b) OH radical addition to >C?C< and ? C?C? unsaturated bonds, (c) OH radical addition to aromatic rings, and (d) OH radical interaction with ? NH2, >NH, >N? , ? SH, and ? S? groups. During its development, this estimation technique has been tested against the available database, and only for 18 out of a total of ca. 300 organic compounds do the calculated and experimental room temperature rate constants disagree by more than a factor of 2. This suggests that this technique has utility in estimating OH radical reaction rate constants at room temperature and atmospheric pressure of air, and hence atmospheric lifetimes due to OH radical reaction, for organic compounds for which experimental data are not available. In addition, OH radical reaction rate constants can be estimated over the temperature range ca. 250–1000 K for those organic compounds which react via H-atom abstraction from C? H and O? H bonds, and over the temperature range ca. 250–500 K for compounds containing >C?C< bond systems.  相似文献   

19.
The application of non-thermal plasma generated by electron beam (EB) was investigated in laboratory scale to study decomposition of polycyclic aromatic hydrocarbons like naphthalene and acenaphthene in flue gas. PAH compounds were treated by EB with the dose up to 8 kGy in dry and humid base gas mixtures. Experimentally established G-values gained 1.66 and 3.72 mol/100 eV for NL and AC at the dose of 1 kGy. NL and AC removal was observed in dry base gas mixtures showing that the reaction with OH radical is not exclusive pathway to initialize PAH decomposition, however in the presence of water remarkably higher decomposition efficiency was observed. As by-products of NL decomposition were identified compounds containing one aromatic ring and oxygen atoms besides CO and CO2. It led to the conclusion that PAH decomposition process in humid flue gas can be regarded as multi-step oxidative de-aromatization analogical to its atmospheric chemistry.  相似文献   

20.
It has been four decades since the phenomenon of photochemical air pollution was first characterized and, in the same year, a tragic London smog episode caused 4,000 excess deaths. Since then, there has been a substantial increase in our understanding of the chemistry involved in both types of air pollution, and a recognition that there is a very close chemical interrelationship between them. In this overview, we provide a brief historical perspective on the atmospheric chemistry of photochemical smog and illustrate how fundamental studies on the gas-phase chemistry of uv-irradiated mixtures of volatile organic compounds (VOC) and oxides of nitrogen (NOx in polluted laboratory and ambient air masses have contributed to our understanding of three environmental problems: the atmospheric formation of ozone, nitric acid and airborne mutagens. In particular, we demonstrate the central role played by nitrogen dioxide and the hydroxyl radical in each case. We also show how certain reactive toxic and acidic species, e.g., formaldehyde and nitrous and formic acids, have been characterized in smog chambers and ambient smog by long pathlength spectroscopic techniques. It is shown that by using the same methods they now have been identified unequivocally, along with NO2, in certain common types of polluted indoor atmospheres ... and at much higher concentrations than outdoors. This has significant health implications for indoor HCHO and quite possibly the acids. We then trace the history of the direct mutagenicity of respirable particles in polluted ambient air and show how, through use of the Ames test in biologically-directed assays of products coupled with fundamental studies of gas-phase reactions of polycyclic aromatic hydrocarbons (PAH) and NOx in irradiated air, much of this activity can be accounted for in terms of the formation of nitro-PAH and oxygenated derivatives. Finally, we discuss the application of basic kinetic, mechanistic and analytical, experimental techniques and theoretical concepts to the development of a new set of “reactivity-based” regulatory controls on motor vehicle emissions of VOC’s. This novel regulatory approach applied by California’s Air Resources Board, which takes effect in 1994, illustrates the continuing need for fundamental research in the area of atmospheric chemistry and how it may be applied to “real world” environmental problems.  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