Night-Time Atmospheric Fate of Acrolein and ... - ACS Publications

Feb 20, 2008 - Night-Time Atmospheric Fate of. Acrolein and Crotonaldehyde. M. S. SALGADO,* E. MONEDERO,. F. VILLANUEVA, P. MARTÍN, A. TAPIA, ...
0 downloads 0 Views 490KB Size
Environ. Sci. Technol. 2008, 42, 2394–2400

Night-Time Atmospheric Fate of Acrolein and Crotonaldehyde M. S. SALGADO,* E. MONEDERO, F. VILLANUEVA, P. MARTÍN, A. TAPIA, AND B. CABAÑAS Departamento de Química Física, Facultad de Ciencias Químicas, Universidad de Castilla La Mancha. Avda Camilo José Cela 10, 13071 Ciudad Real, Spain.

Received October 5, 2007. Revised manuscript received December 19, 2007. Accepted January 8, 2008.

The absolute rate coefficients for the gas-phase reaction of the NO3 radical with acrolein and crotonaldehyde have been measuredoverthetemperaturerange249–330K,usingadischarge flow system and monitoring the NO3 radical by laser induced fluorescence (LIF). The obtained rate coefficients at 298 K for NO3 reactions with acrolein and crotonaldehyde were (3.30 ( 0.39) × 10-15 cm3 molecule-1 s-1 for acrolein and (1.35 ( 0.04) × 10-14 cm3 molecule-1 s-1 for crotonaldehyde, and the proposed Arrhenius expressions are k(T) ) (1.72 ( 0.5) × 10-13 exp[(-1190 ( 43)/T] and k(T) ) (5.02 ( 0.7) × 10-13 exp[(-1076 ( 47)/T], respectively, in units of cm3 molecule-1 s-1. In addition, the products and mechanisms were investigated using an environmental chamber/FTIR absorption system. Formaldehyde, CO, and acryloylperoxy nitrate were identified as the main products for the acrolein reaction with molar yields of 31.6 ( 2.0, 20.9 ( 1.9, and 47 ( 3, respectively. In the crotonaldehyde reaction the main products detected were crotonylperoxy nitrate and CO with yields of 93.6 ( 4.3 and 8.3 ( 1.1, respectively. On the basis of the rate constant measured, the activation energy calculated, and the identified products, abstraction of the aldehydic H seems to be the main degradation pathway at room temperature for the reaction of acrolein with NO3. For crotonaldehyde, the mechanism is unclear on the basis of the experimental results. The atmospheric implications of the reactions in question are also discussed.

Introduction Unsaturated carbonyl compounds play an important role in the chemistry of the polluted troposphere. The principal unsaturated compounds of atmospheric interest are R,βunsaturated carbonyls such as acrolein (prop-2-en-1-al) and crotonaldehyde (but-2-en-1-al). Both of these materials are released into the atmosphere through different sources such as combustion processes and chemical industries (1). For example, they are emitted in the combustion of wood, polymers, tobacco, and gasoline. Acrolein is also a highvolume industrial commodity because it is an intermediate in the manufacture of acrylic acid and other organic compounds (e.g., glycerine, plastics, polyurethane, polyester resins, perfumes, and various pharmaceuticals). Moreover, acrolein is formed in the atmospheric degradation of 1,3dienes (2). Average acrolein concentrations measured in the United States at various monitoring stations range from 0.5 to 3.2 ppbv (3). As far as crotonaldehyde is concerned, this * Corresponding author e-mail: [email protected] 2394

9

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 42, NO. 7, 2008

compound can be found in at least two industrial sectors: cellulosic man-made fibers and industrial organic chemicals (4). Interest in the chemistry of acrolein and crotonaldehyde has been stimulated by the recent observation of acryloylperoxy nitrate (CH2dCHCOO2NO2, APAN, or vinyl PAN references) in ambient air (5, 6). For example, Tanimoto and Akimoto (5) observed this species in both an urban site (suburban Tokyo) and a more remote site (Rishiri Island, northern Japan). Roberts et al. (6) identified APAN in ambient air near Houston. Elevated APAN mixing ratios in this region were attributed in part to large emissions of 1,3-butadiene and/or acrolein from the petrochemical industry. It is well-established that the main degradation processes of carbonyl compounds in the gas phase are controlled by photolysis or reaction with OH or NO3 radicals and eventually with O3 in the case of unsaturated compounds. In coastal and marine areas, the reactions with Cl atoms during daylight hours could also be important (7). Because these processes can lead to the formation of additional radicals, they could be significant with regard to the atmospheric oxidation capacity and local and regional formation of ozone and other photooxidants. Photolysis and OH-reactions of acrolein and crotonaldehyde have been the subjects of numerous studies (8–14), but at present, very little is known about the kinetics and mechanisms of the reactions of NO3 with these compounds. The available information corresponds to kinetic studies, most of which were carried out using the relativerate technique (15–18). Additionally, the activation energies for the reactions in question have previously been obtained in our laboratory in the temperature range 298–430 K in the only absolute study reported to date (16). Nonetheless, complete accounting for the product yields and, hence, an accurate assessment of the branching ratio for NO3 addition and abstraction channels in the indicated reactions has yet to be achieved. Moreover, information on the reactivity at low temperatures is not currently available, and these studies are of crucial interest, bearing in mind that the temperature of the troposphere can reach values much lower than 220 K (19). The work reported here concerned a complete study of the reaction of the NO3 radical with acrolein and crotonaldehyde. First, a low-temperature analysis of the reactions was developed using a flow-tube discharge system with laser induced fluorescence (LIF) detection of the NO3 radical. This study enabled us to confirm the activation energy previously determined by considering new data for temperatures below room temperature and also to determine the possible influence of temperature on the reaction mechanism. Additionally, an environmental chamber/FTIR absorption system was used to determine the end products of the NO3initiated oxidation of both acrolein and crotonaldehyde in order to establish the night-time degradation mechanism of these aldehydes in the atmosphere and the possible atmospheric implications.

Experimental Section Kinetic Study. The kinetic measurements of the absolute rate constants were carried out using a discharge-flow tube reactor with LIF detection for the nitrate radical excited at λ ) 662 nm. The experimental setup has been described elsewhere (20, 21), and so only the relevant details are given here. Helium was used as the carrier gas. Nitrate radicals, generated in a sidearm tube by the fast reaction F + HNO3 f NO3 + HF, were admitted to the flow tube through a fixed port. Fluorine atoms were obtained by passing F2-He mixtures 10.1021/es702533u CCC: $40.75

 2008 American Chemical Society

Published on Web 02/20/2008

FIGURE 1. First-order plot at different reactant concentrations for the reactions of NO3 with acrolein at 260 K. The error limits given for the rate coefficients k′ include only the precision for the fit to our experimental data (twice the standard deviation ( 2σ).

FIGURE 2. Plots of the pseudofirst-order rate coefficients k′ versus concentration of acrolein at different temperatures. The second-order rate coefficients (k′ ( 2σ) were extracted from the linear-least-squares analysis. through a microwave discharge (2450 Mz) to give initial NO3 concentrations in the range of (2.29–32.8) × 1012 molecule cm-3. In all the experiments, the HNO3 concentration was in a sufficiently large excess over the F atom concentration to prevent secondary chemistry. Absolute concentrations of NO3 were determined before and after each kinetic run by chemical titration with known amounts of tetramethylethene (TME) (22). Under the experimental conditions used, the loss of nitrate radicals on the wall of the flow tube was not observed. Organic reactants were added to the flow through a sliding injector. The time evolution of the reaction was monitored by changing the distance between the injector and the detection cell. Contact times between the NO3 radical and the reactant ranged from 4 to 80 ms. The concentrations were in the range (1.51-11.61) × 1014 molecule cm-3 in the case of acrolein and (1.06-4.18) × 1014 molecule cm-3 for crotonaldehyde, and these concentrations were measured by the drop in pressure with time from their calibrated volume storage bulbs.

The experiments were conducted at a total pressure of (1.1–1.5) ( 0.1 Torr and at temperatures in the range 249-330 K ((1 K). The flow tube was cooled and heated at the desired temperature by circulating ethanol through an external jacket connected by Viton tubing to a thermostatic bath, which controlled the temperature. The desired bimolecular rate coefficient for the reactions between these aldehydes and NO3 radicals, k, is defined by eq 1: d|NO3| ) -k[NO3][aldehyde] dt

(1)

Kinetic experiments for acrolein were performed under pseudofirst-order conditions, with organic reactant present in a large excess over NO3. The corresponding integrated rate expression in this case is shown in eq 2. ln

[NO3]0 ) (k[aldehyde])t ) k ′ t [NO3]t

(2)

where VOL. 42, NO. 7, 2008 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

9

2395

k′ ) k[aldehyde]

(3)

The vapor pressure of crotonaldehyde does not allow us to perform the experiments with a large excess of reactant over NO3, so in this case the data were analyzed using the second-order integrated rate expression (23), assuming a 1:1 stoichiometry, ln

(M - Xa) [M(1 - Xa)]

) (Bo - A0)kt

(4)

where M ) [Crotonaldehyde]o/[NO3]o Bo ) [Crotonaldehyde]o Ao ) [NO3]o Xa )

[NO3]o - [NO3]t [NO3]o

(5)

M ) [Crotonaldehyde]o/[NO3]o The second-order rate constant may be obtained in this case as the slope of plots of 1/(B0-A0) ln[(M-Xa)/M(1-Xa)] versus time. Chemicals used in this study and their sources and purities were as follows: Helium (Carburos Metálicos, C50) was passed through an oxygen-removing column (Oxisorb, Messer Griesheim) and through a trap containing molecular sieves. Molecular fluorine (5% in He) was supplied by Praxair. Anhydrous gaseous HNO3 in a He carrier was prepared by bubbling He through a mixture of H2SO4/HNO3 (P.A. Panreac). Acrolein and crotonaldehyde (Purity 99%) were supplied by the Aldrich Chemical Co. and were purified by successive trap-to-trap distillations. Products Study. All of the experiments were carried out in a 405 L Pyrex cylindrical glass reactor (1.5 m length and 60 cm inner diameter) with Teflon coated metal end flanges. This reactor has previously been described in detail (24). In brief, a White mirror system (base path length 1.4 m) mounted inside the reactor is coupled, by an external mirror system, to a Fourier transform-spectrometer (Nicolet Magna 550), and this arrangement enables the in situ monitoring of both reactants and products by long-path infrared absorption using a total path length of 50.4 m and a resolution of 1 cm-1. The spectrometer was directly controlled by OMNIC software provided by Nicolet and running on a personal computer, which was also used to store raw data. Nitrate radicals were generated by thermal decomposition of N2O5. The N2O5 was synthesized in a separate setup by the reaction of excess O3 with NO2 and trapped and stored at 195 K. Products of the NO3-oxidation of acrolein and crotonaldehyde were measured at 760 ( 10 Torr total pressure (synthetic air) and at 298 ( 3 K. Quantification of reactant and products was carried out by comparison with quantitative reference spectra of the organic compounds, N2O5, NO, NO2, HNO3, HCHO, and CO, which were taken from a calibrated infrared spectra data bank archived by the laboratory. In the case of peroxyacyl nitrates, APAN and crotonylperoxy nitrate (CPAN) spectra were obtained from the final product spectrum of the reaction after the absorption bands of the identified products and reactants were subtracted and were identified by comparison with a reference spectrum previously published (12). This spectrum shows several intense absorptions at around 796, 1037, 1300, 1741, and 1834 cm-1, which are characteristic of the NO2 scissors, NO2 symmetric stretch, NO2 asymmetric stretch, and CO stretching modes in peroxyacyl nitrates, respectively. 2396

9

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 42, NO. 7, 2008

FIGURE 3. Plots of 1/B0 - A0 ln [M - XA/M (1 - XA)] versus time obtained for the reaction of NO3 with crotonaldehyde at different temperatures. The second-order rate coefficients (k ( 2σ) were extracted from the linear-least squares analysis. The experimental procedure was as follows. First, to obtain the wall losses, the organic compound was introduced alone into the reaction chamber under reduced pressure, and the wall loss rate (k3) for the compound was derived from the slope of plots of ln([C0]/[Ct]) against time, where C0 is the initial concentration of the compound at t ) 0 and Ct its concentration at time t. For the studied compounds, wall deposition losses were estimated to be negligible (10-5-10-6) s-1, although this was corrected in the experiments. After this determination, N2O5 was flushed into the chamber through a Teflon line by evaporating solid N2O5. In total, the measurement time period for one experiment was about 30 min, with 128 scans recorded per spectrum over a period of 2 min and 15 such spectra collected. Once the N2O5 has been consumed, possible wall deposition of the products formed was studied as described above; once again, the losses were negligible. Analysis of the complex spectra was carried out by successively subtracting the characteristic absorption of identified compounds using calibrated spectra. The known concentration of the reference spectra and the subtraction factors allowed the concentration of each identified compound to be determined. The initial organic reactant concentration range was (2.2–3.1) × 1014 molecule cm-3 and the initial N2O5 concentration range was (1.4–2.1) × 1014 molecule cm-3. A minimum of six experiments was performed for each organic compound. Between experiments the chamber was evacuated to approximately 10-3 mbar using the pumping system.

Results and Discussion A typical first-order plot for the reaction of NO3 with acrolein at 260 K is shown in Figure 1. The slight positive intercepts observed are attributed to the initial consumption of NO3 radical by a rapid reaction with impurities, and so long as the impurities react rapidly, the plots will be linear and their slopes will correctly represent the second-order rate coefficients. The resulting first-order rate constants k′ at each temperature were plotted against the reactant concentration, and second-order rate constants were obtained as the slope of least-squares fits of these data (Figure 2). For crotonaldehyde, plots of 1/B0-A0 ln [M - XA/M (1 - XA)] versus time, according to eq 4, directly yield the bimolecular rate constants (Figure 3). The positive intercepts obtained in the secondorder plots are related with a small heterogeneous effect on the walls of the flow tube because the sticky nature of aldehydes. In this case, the second-order plot will appear to be straight, and it is the true homogeneous gas-phase rate constant, although will have an intercept.

TABLE 1. Experimental Kinetic Data for the Reaction of NO3 with Acrolein and Crotonaldehydea compound

acrolein (CH2)CHCHO)

kNO3 (10-15 cm3molec-1 s-1) (technique) (ref)

peroxyacyl nitrate identified as product

2.5 ( 0.4 (FTD-LIF) (16) 1.1 ( 0.4 (R) (17) 8.9 ( 2.8 (FTD-LIF) (17) 1.1 ( 0.2 (R) (18) 3.3 ( 0.4 (A) this work 16.1 ( 0.2 (DFT-LIF) (16) 5.1 ( 0.2 (R) (18) 13.5 ( 0.6 (A) this work

crotonaldehyde (CH3CH)CHCHO)

APAN

CPAN

band (cm-1)

absorption cross section (molecule-1 cm2)

793 950 1193 1300 1740 1810 793 1070 1205 1300 1748 1818

1.1 × 10-18 1.6 × 10-18 2.9 × 10-19 1.1 × 10-18 2.4 × 10-18 7.8 × 10-19 1.3 × 10-18 6.9 × 10-19 2.1 × 10-19 1.2 × 10-18 2.8 × 10-18 6.8 × 10-19

a

R, relative technique; PLP-LIF, pulsed laser photolisis-laser induced fluorescence; FTD-LIF, flow tube discharge-laser induced fluorescence.

FIGURE 4. Arrhenius plots for the reactions of NO3 with acrolein and crotonaldehyde compared with previous studies (16). The four regression lines represented correspond to 2σ statistical analysis error of any of the four series of experimental data plotted. Values of the rate coefficients at room temperature obtained here are compared with available literature data in Table 1. For the acrolein reaction, our value is consistent with that reported by Cabañas et al. (16) and is higher than the values of Canosa-Mas et al. (17) and Atkinson et al. (18), these latter values were determined using a relative technique. Canosa-Mas et al. also determined an absolute rate constant for this reaction, but the value exceeds the expected one due to secondary chemistry, as explained by the authors. For crotonaldehyde, absolute rate constants are again higher than those obtained in relative experiments. The absence of secondary chemistry in our experiments was tested with plots of the second-order intercepts versus reactant concentration and versus temperature. No dependence was observed in both cases. Although no pressure dependence has previously been described for aldehydes reactions with NO3 radical, it possibly will be necessary to obtain more kinetic data for acrolein and crotonaldehyde reactions at varying pressure conditions to uncover the reason for the discrepancies observed using different experimental conditions. According to the Arrhenius expression, activation energies of the considered reactions can be obtained by plotting ln k(T) versus 1/T (Figure 4). The following expressions were found: k(T) ) (1.72 ( 0.5) × 10-13 exp[(-1190 ( 43)/T] cm3

molecule-1 s-1 (giving an activation energy of 9.89 ( 0.72 kJ mol-1) for the reaction of acrolein with NO3 and k(T) ) (5.02 ( 0.7) × 10-13 exp[(-1076 ( 47)/T] (activation energy 8.95 ( 1.60 kJ mol-1) for crotonaldehyde. Activation energies for these reactions have previously been calculated (16) in the temperature range 298–433 K, with values of 26.9 and 20.1 kJ mol-1 obtained for acrolein and crotonaldehyde reactions, respectively. These values are more than twice those obtained in the study reported here at low temperatures. Some experiments were carried out at high temperatures in our experimental system (data also plotted in Figure 4), obtaining rate constants consistent with those previously reported. Additionally, data reported here at low and high temperatures did not adjust to any mathematical function. We think that these facts support the idea that there are different activation energies depending on the temperature range used for the experiments and, possibly, a change in reaction mechanism. In fact, temperature dependence for abstraction channels has previously been observed in the reaction of Cl with acrolein (7), where the detected products were different depending on the temperature used in the experiments. The kinetic study was completed by determining the products of these reactions using an FTIR chamber as explained above. Thus, FTIR spectroscopy and comparison VOL. 42, NO. 7, 2008 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

9

2397

FIGURE 5. Spectrum A shows a typical product obtained after substraction of the spectral features belonging to water, reactants (acrolein, N2O5, NO, and HNO3), spectra B, C, and D show reference spectra of APAN (obtained in this work), HCHO, and CO, respectively. Spectrum E corresponds to the reference spectrum of APAN (12).

FIGURE 6. Concentration-time profiles of acrolein and products identified in its NO3-initiated oxidation. with the characteristic IR spectra showed that formaldehyde, CO, and APAN were the main reaction products in the NO3 radical-initiated oxidation of acrolein. An example of the FTIR data for one typical experiment on acrolein-N2O5 performed in the 405 L Pyrex glass reactor is shown in Figure 5. The concentration-time profile of reactants and identified products in a typical NO3-initiated oxidation of acrolein is shown in Figure 6. To determine the concentration of APAN, we estimated that all acrolein consumed was used to form HCHO, CO, and APAN (because other products are not observed). The concentration of CO and HCHO could be calculated using calibrated spectra for these compounds, 2398

9

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 42, NO. 7, 2008

and therefore, from the difference with acrolein consumed we can ascertain the concentration of APAN. To confirm the reliability of the calculation method, we used the absorption cross section previously calculated by Orlando (12) for knowing APAN concentrations, and similar results have been obtained. The corrected product concentrations were plotted versus the amount of acrolein consumed, and the corrected yields derived from the least-squares analysis of such plots were as follows: 47 ( 3, 20.9 ( 1.9, and 31.6 ( 2.0 for APAN, HCHO, and CO, respectively. The same procedure in the case of the reactions of crotonaldehyde with NO3 gave CPAN and CO as products with yields of 93.6 ( 4.3 and 8.3 ( 1.1, respectively.

SCHEME 1. Proposed mechanism for the reaction of acrolein with NO3

As explained above, CPAN was also identified by comparison with a reference spectrum (12). The errors quoted are a combination of the 2σ statistical errors from the regression analysis together with the errors from the spectral subtraction procedure. For APAN and CPAN, the yields are upper-limit yield due to the simulation method used for knowing its concentrations. Absorption cross sections of APAN and CPAN at each wavelength were obtained by monitoring the infrared absorption as a function of the product of ([Peroxyacyl nitrate] × l) and applying Beer’s law. Values are listed in Table 1. Our data are in good agreement (deviation always lower than 15%) with data previously determined by Orlando et al. (12), confirming the reliability of the calculation method that no other secondary reactions have significant influence in the areas of the bands obtained for products and, moreover, confirming products identified as dominant in these experimental studies. Finally, taking into account the order of the kinetic constant value, the positive activation energy, and the products identified for the reaction of NO3 with acrolein, a mechanism at room temperature can be proposed and is shown in Scheme 1. Night-time atmospheric degradation of acrolein occurs by reaction with NO3 radicals through a mechanism involving abstraction of the aldehydic H atom, and HCHO and APAN are the main detected products. The resulting peroxyacyl nitrate is very stable at room temperature and at atmospheric pressure because a significant decrease in the absorption bands is not observed — even more than 30 min after reaction time. Additionally, when NO was added to the reaction chamber, APAN did not react in a significant proportion, again confirming the stability of the formed compound.

The system used for product analysis was not designed for temperature variations, so we were unable to assess whether the products (or yields) at lower temperatures are different, which would have supported the change of mechanism suggested above. In the case of crotonaldehyde, the results cannot confirm the nature of the dominant pathway for this reaction at room temperature, because CO obtained as a product could be generated either in an H-abstraction initiation channel or by NO3 initial addition to the double bond. Several qualitative experiments were made in our laboratory for this reaction using GC-MS (Shimadzu, GC-MS-QP5000) as the detection system and obtaining an unidentifiable methyl cetonic compound and acetic acid as products, but not in a significant way. These products could indicate an addition channel for this reaction. Thermal instability of CPAN or other compounds in CG-MS did not allow us to quantify the branching ratios to the competing pathways.

Atmospheric Implications The atmospheric lifetimes for acrolein and crotonaldehyde with respect to the NO3 radical reaction can be calculated according to eq 6, τ)

1 [R] k2

(6)

where [R] is the estimated ambient tropospheric concentration of the radical. The rate constants at room temperature calculated in this work for NO3 reactions can be used to give the following lifetime values: 168 and 41 h for acrolein and crotonaldehyde, respectively (using [NO3] ) 5 × 108 molecules cm-3 (25)). Calculated tropospheric lifetimes of these aldeVOL. 42, NO. 7, 2008 / ENVIRONMENTAL SCIENCE & TECHNOLOGY

9

2399

hydes with respect to daily OH- and Cl-reaction (8–13, 26, 27), which are on the order of a few hours, are shorter than the values estimated here with respect to the NO3 reaction. However, at night the very low concentration of the OH radical (due to its photolytic generation) indicates that the NO3 reaction is the relevant process for acrolein and crotonaldehyde degradation; therefore, the products obtained in these reactions must also be considered. HNO3 is one of the products generated in H-abstraction reactions and also by the decomposition of PANs, and it is one of main contributors to precipitation acidity. In addition, PAN compounds obtained in these reactions are important atmospheric trace species through their role as a reservoir for NOx, and they are closely related to the photochemical ozone formation and to the oxidizing potential of the atmosphere (28, 29). Moreover, the long lifetime of PANs in the upper troposphere allows them to transport NO2 over wide areas (30). Furthermore, peroxyacyl nitrates have also received attention as eye irritants, mutagens, possible skin cancer agents, and phytotoxins (31–33). Finally, another important product detected in this study was HCHO, formaldehyde, the presence of which in the air at levels at or above 0.1 ppm causes serious health effects (34).

Acknowledgments E. M. and F. V. thank the Ministerio de Educación y Cultura and Junta de Comunidades de Castilla La Mancha, respectively, for personal grants. This work was supported by project No. BQU-2001-1574, granted by the CICYT (Comisión Interministerial de Ciencia y Tecnología). The authors also acknowledge the cooperation of the Wuppertal University, especially that of Ian Barnes and the Physikalische Chemie for the analysis of products.

Literature Cited (1) Grosjean, D. Atmospheric chemistry of toxic contaminants. 3. Unsaturated aliphatics: Acrolein, acrylonitrile, maleic anhydride. Air Waste Manage. Assoc. 1990, 40, 1664–1668. (2) Tuazon, E. C.; Alvarado, A.; Aschmann, S. M.; Atkinson, R.; Arey, J. Products of the Gas-Phase Reactions of 1,3-Butadiene with OH and NO3 Radicals. Environ. Sci. Technol. 1999, 33, 3586– 3595. (3) Agency for Toxic Substances and Disease Registry (ATSDR), Toxicological Profile for Acrolein; U. S. Public Health Service, U.S. Department Health and Human Services: Atlanta, Georgia, 2007. (4) U.S. Environmental Protection Agency. http://www.epa.gov. (5) Tanimoto, H.; Akimoto, H. A new peroxycarboxylic nitric anhydride identified in the atmosphere: CH2)CHC(O)OONO2 (APAN). Geophys. Res. Lett. 2001, 28 (14), 2831–2834. (6) Roberts, J. M.; Flocke, F.; Weinheimer, A.; Tanimoto, H.; Jobson, B. T.; Riemer, D.; Apel, E.; Atlas, E.; Donnelly, S.; Stroud, V.; Johnson, K.; Weaver, R.; Fehsenfeld, F. C. Observations of APAN during TexAQS 2000. J. Geophys. Res. 2002, 107, 4554. (7) Aranda, A.; Díaz de Mera, Y.; Rodríguez, A.; Rodríguez, D.; Martínez, E. A kinetic and mechanistic study of the reaction of Cl atoms with acrolein: Temperature dependence for abstraction channel. J. Phys. Chem. A. 2003, 107, 5717–5721. (8) Maldotti, A.; Chiorboli, C.; Bignozzi, C. A.; Bartocci, C.; Carassiti, V. Photooxidation of 1,3-butadiene containing systems: Rate constant determination for the reaction of acrolein with OH radicals. Int. J. Chem. Kinet. 1980, 12, 905–913. (9) Kerr, J. A.; Sheppard, D. W. Kinetics of the reactions of hydroxyl radicals with aldehydes studied under atmospheric conditions. Environ. Sci. Technol. 1981, 8, 960–964. (10) Atkinson, R.; Aschmann, S. M.; Pitts, J. N. Jr. Kinetics of the gas-phase reactions of hydroxyl radical with a series of R,βunsaturated carbonyls at 299 ( 2 K. Int. J. Chem. Kinet. 1983, 15, 75–81. (11) Grosjean, E.; Williams, E. L. II.; Grosjean, D. Atmospheric chemistry of acrolein. Sci. Total Environ. 1994, 153, 195–202.

2400

9

ENVIRONMENTAL SCIENCE & TECHNOLOGY / VOL. 42, NO. 7, 2008

(12) Orlando, J. J.; Tyndall, G. S. Mechanisms for the reactions of OH with two unsaturated aldehydes: Crotonaldehyde and acrolein. J. Phys. Chem. A. 2002, 106, 12252–12259. (13) Orlando, J. J.; Tyndall, G. S.; Burkholder, J. B.; Bertman, S. B.; Chen, W. The atmospheric chemistry of the unsaturated aldehydes methacrolein, acrolein, and crotonaldehyde. Abstracts of Papers, 223rd ACS National Meeting, Orlando, United States, 2002. American Chemical Society: Washington DC, 2002. (14) Magneron, I.; Thevenet, R.; Mellouki, A.; Le Bras, G.; Moortgat, G. K.; Wirtz, K. A study of the photolysis and OH-initiated oxidation of acrolein and trans-crotonaldehyde. J. Phys. Chem. A. 2002, 106, 2526–2537. (15) Carlier, R.; Hannachi, H.; Mouvier, G. The chemistry of carbonyl compounds in the atmosphere: A review. Atmos. Environ. 1986, 20, 2079. (16) Cabañas, B.; Salgado, S.; Martín, P.; Baeza, M. T.; Martínez, E. Night-time atmospheric loss process for unsaturated aldehydes: reaction with NO3 radicals. J. Phys. Chem. A. 2001, 105, 4440– 4445. (17) Canosa-Mas, C. E.; Carr, S.; King, M. D.; Shallcross, D. E.; Thompson, K. C.; Wayne, R. P. A kinetic study of the reactions of NO3 with methyl vinyl ketone, methacrolein, acrolein, methyl acrylate and methyl methacrylate. Phys. Chem. Chem. Phys. 1999, 1, 4195–4202. (18) Atkinson, R.; Aschmann, S. M.; Goodman, M. A. Kinetics of the gas-phase reactions of nitrate radicals with a series of alkynes, haloalkenes, and R,β-unsaturated aldehydes. Int. J. Chem. Kinet. 1987, 19, 299–307. (19) Toon, O. B.; Hamill, P.; Turzo, R. P.; Pinto, J. Condensation of nitric acid and hydrochloric acid in the winter polar stratospheres. Geophys. Res. Lett. 1986, 13, 1284–1287. (20) Martínez, E.; Cabañas, B.; Aranda, A.; Wayne, R. P. Temperature dependence of the reactions of the nitrate radical with dichloroalkenes followed by LIF detection. J. Chem. Soc., Faraday Trans. 1996, 92, 53–58. (21) Martínez, E.; Cabañas, B.; Aranda, A.; Martín, P.; Notario, A. Salgado. Study on the NO3 Radical Reactivity Reactions with Cyclic Alkenes S. J. Phys. Chem. A. 1999, 103, 5321–5327. (22) Sander, S. P. Temperature dependence of the NO3 absorption spectrum. J. Phys. Chem, 1986, 90, 1435. (23) Laidler, J. K. Chemical Kinetics, 3rd ed.; Harper Collins Publisher, Inc, 1987. (24) Bierbach, A. Produktuntersuchungen und Kinetik der OHinitiierten Gasphasenasenoxidation aromatischer Kohlenwasserstoffe sowie ausgewähter carbonyllischer Folgenprodukte; Ph.D. dissertation, BUGH Wuppertal, Wuppertal, 1994. (25) Shu, Y.; Atkinson, R. Atmospheric lifetimes and fates of a series of sesquiterpenes. J. Geophys. Res 1995, 100, 7275. (26) Thevenet, R.; Mellouki, A.; Le Bras, G. Kinetics of OH and Cl reactions with a series of aldehydes. Int. J. Chem. Kinet. 2000, 32, 676–685. (27) Cuevas, C. A.; Notario, A.; Martinez, E.; Albadalejo, J. Rate coefficients in function of temperature for the gas-phase reaction of Cl atoms with selected aldehydes. Atmos. Environ. 2006, 40, 3845–3854. (28) Atkinson, R. Atmospheric chemistry of VOCs and NOx. Atmos. Environ. 2000, 34, 2063–2102. (29) Roberts, J. M. Composition Chemistry and Climate of the Atmosphere; Singh, H. B. , Ed.; VNR-Verlag, New York,1995. (30) Olszyna, K. J.; Bailey, E. M.; Simonaitis, R.; Meagher, J. F. O3 and NOy relationships at a rural site. J. Geo. Res A. 1994, 99, 14557– 14563. (31) Taylor, O. C. Importance of peroxyacetyl nitrate (PAN) as a phytotoxic air pollutant. J. Air Pollut. Control Assoc. 1969, 19347, 19351. (32) Peak, M. J.; Belser, W. L. Some effects of the air pollutant, peroxyacetyl nitrate, upon deoxyribonucleic acid and upon nucleic acid bases. Atmos. Environ. 1969, 3, 385. (33) Kleindienst, T. E.; Shepson, P. B.; Smith, D. F.; Hudgens, E. E.; Nero, C. M.; Cupitt, L. T.; Bufalini, J. J.; Claxton, L. D. Comparison of Mutagenic Activities of Several Peroxyacyl Nitrates. Environ. Mol. Mutagen. 1990, 16, 70–80. (34) International Agency for Research on Cancer. Monographs on the evaluation of carcinogenic risks to humans; World Health Organization: 2002. http://monographs.iarc.fr.

ES702533U