Acessibilidade / Reportar erro

Rate Coefficient for the Reaction of Cl Atoms with cis-3-Hexene at 296 ± 2 K

Abstract

The rate coefficient of the cis-3-hexene + Cl atoms reaction at 296 ± 2 K and 750 ± 10 Torr was determined using the relative rate technique. The reaction was investigated using an 80 L Teflon reaction bag and a gas chromatograph coupled with flame-ionization detection. Chlorine atoms were produced by the photolysis of trichloroacetyl chloride. No previous experimental data was available in the literature, to the best of our knowledge. The mean second-order rate coefficient value found was (4.13 ± 0.51) × 10-10 cm3 molecule-1 s-1. The experimental value agrees with the rate coefficient estimated by structure-reactivity analysis, 4.27 × 10-10 cm3 molecule-1 s-1. Moreover, both addition and hydrogen abstraction channels contribute to the global kinetics, with branching ratios 70:30. Effective lifetime with respect to Cl atoms is predicted as 67.2 hours; however, the cis-3-hexene + Cl channel is suggested to be non-negligible at atmospheric conditions. Other atmospheric implications are discussed.

Keywords:
cis-3-hexene + Cl atoms; rate coefficient; experimental relative rate method


Introduction

Volatile organic compounds (VOCs) are emitted to the troposphere from different sources of biogenic and anthropogenic origin, playing a fundamental role in atmospheric chemistry. Among the anthropogenic origin compounds released, cis-3-hexene is of a particular interest, emitted in gasoline vapor.11 Heath, J. S.; Koblis, K.; Sager, S. L.; J. Soil Contam. 1993, 2, 1.,22 Ramnäs, O.; Östermark, U.; Petersson, G.; J. Chromatogr. 1993, 638, 65. In the troposphere, cis-3-hexene reacts with hydroxyl radicals, ozone and nitrate radicals. Barbosa et al.33 Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714. have studied the kinetic for the reaction of cis-3-hexene with hydroxyl radicals using the relative rate method and the experimental mean second-order rate coefficient value was determined as (6.27 ± 0.66) × 10-11cm3molecule-1 s-1. The kinetics of the reaction with ozone and nitrate radicals have been investigated at room temperature using the relative method and the rate coefficients values (in cm3molecule-1s-1) have been determined as (1.44 ± 0.17) × 10-16 and (4.37 ± 0.49) × 10-13, respectively.44 Calvert, J. G.; Atkinson, R.; Kerr, J. A.; Madronich, S.; Moortgat, G. K.; Wallington, T. J.; Yarwood, G.; The Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press, Inc.: New York. 2000.,55 Pfrang, C.; Martin, R. S.; Nalty, A.; Waring, R.; Canosa-Mas, C. E.; Wayne, R. P.; Phys. Chem. Chem. Phys. 2005, 7, 2506.

Chlorine atoms are also important atmospheric oxidants and have been observed in the marine boundary layer. In the early morning, the concentration of Cl atoms reaches the highest value and reactions of VOCs with Cl could be even more important than the reactions with OH radicals, the major daytime oxidant.66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. To the best of our knowledge, the reaction of Cl atoms with cis-3-hexene, despite the importance of such reaction to Atmospheric Chemistry, has not been studied yet.

The main goal of this work is the experimental study of the kinetics of the cis-3-hexene + Cl reaction. The rate coefficient at 298 K and atmospheric pressure is reported for the first time based on the use of the relative rate method. Aspects of the reaction mechanism and atmospheric implications are also discussed.

Experimental

The experimental study was performed at the Instituto de Investigaciones en Fisico Química de Córdoba, Argentina. An 80 L collapsible Teflon bag was used and the rate coefficients were determined by the relative rate method. This methodology has been carried out at this laboratory for the study of other reactions.33 Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714.,77 Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Teruel, M. A.; Lane, S. I.; Atmos. Environ. 2006, 40, 7298.

8 Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2010, 44, 1749.
-99 Abrate, J. P. A.; Pisso, I.; Peirone, S. A.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2013, 67, 85.

Briefly, the reactant and the reference compound were introduced into the chamber using nitrogen or ultrapure air. Cl atoms were generated by the trichloroacetyl chloride photolysis using three germicide lamps (Philips 30 W), with a λ maximum around 254 nm, and the time of photolysis varied from 20 seconds to 1 minute. A gas syringe (Hamilton gas tight 5 mL) was periodically used to collect samples and the gas sample was analyzed using a gas chromatograph (Clarus 500, PerkinElmer) equipped with an Elite-1 column (PerkinElmer, length: 30 m, inner diameter 0.32 mm, film thickness: 0.25 µm) and a flame ionization detector (FID). The temperature of the column was 33 °C for 20 min. Helium was used as the carrier gas with flow rate of 0.8 mL min-1.

cis-3-Hexene and reference compounds with trichloroacetyl chloride were introduced into the chamber and left in the dark for 2 hours. Under such conditions, no evidence for a reaction between cis-3-hexene and the reference compounds has been found. No reaction was observed for the organic specie and trichloroacetyl chloride in the absence of UV light.

The reactant and the reference compounds decay through the following reactions:

(1) Cl + cis 3 hexene p r o d u c t s , k hex

(2) Cl + reference p r o d u c t s , k ref

where, khex and kref are the rate coefficients for reactions of the cis-3-hexene and the reference compound with Cl atoms, respectively.

Assuming that the cis-3-hexene and the reference compounds are lost entirely due to the reactions 1 and 2, the following relation can be obtained:

(3) ln cis 3 hexene 0 cis 3 hexene t = k hex k ref × ln reference 0 reference t

In equation 3, the subscripts 0 and t correspond to the time instants 0 and t, respectively.

Rate coefficients for the cis-3-hexene with Cl atoms were obtained, in each experiment, at 298 ± 2 K and atmospheric pressure 750 ± 10 Torr, relative to the rate coefficients of the Cl atoms reactions with n-heptane and cyclopentane used as reference compounds.

The initial concentrations used during the experiments were in the range of 4.67-9.59 × 101414 Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Troe, J.; J. Phys. Chem. Ref. Data 1989, 18, 881. molecule cm-3 for the cyclopentane, 7.38 × 101414 Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Troe, J.; J. Phys. Chem. Ref. Data 1989, 18, 881. molecule cm-3 for the cis-3-hexene, 6.15 × 101414 Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Troe, J.; J. Phys. Chem. Ref. Data 1989, 18, 881. molecule cm-3 for n-heptane and 1.0-3.4 × 101515 Atkinson, R.; Chem. Rev. 1986, 86, 69. molecule cm-3 for the trichloroacetyl chloride.

The chemicals used were trichloroacetyl chloride (99%, CAS: 76-02-8) supplied by Sigma-Aldrich, n-heptane (99.9%, CAS: 142-82-5) and cyclopentane (99.9%, CAS: 287-92-3) supplied by Mallinckrodt and cis-3-hexene (97%, CAS: 7642-09-3) supplied by Alfa Aesar. Synthetic air (purity 99.999%) and chromatographic gases (N2, H2 and chromatographic air) were obtained from Linde. The organic reagents were degassed by repeated freeze-pump-thaw cycling and purified by vacuum distillation before use.

The infrared (IR) spectrum of cis-3-hexene was recorded with a Nicolet FTIR (Fourier transform infrared) spectrometer, with 1.0 cm-1 resolution. The absorption cell used was a Pyrex cell sealed with NaCl windows and with an optical path-length equal to 23.0 ± 0.1 cm. Gas sample pressures were measured with a capacitance manometer (MKS Baratron, range 10 Torr). Background spectra were measured with the sample cell under vacuum. The infrared spectrum, recorded in the 500-1500 cm-1 region at 298 K, was used to calculate radiative efficiencies (RE)1010 Pinnock, S.; Hurley, M. D.; Shine, K. P.; Wallington, T. J.; Smyth, T. J.; J. Geophys. Res. 1995, 100, 23227. and the global warming potential.

The model adopted for calculating the radiative efficiencies considers a uniform distribution of the compound over the troposphere. The RE values for short-lived compounds calculated from this model can be significantly lower, since the concentration should strongly decrease with altitude. Taking this assumption into account, the calculated RE values, as estimated in this work, should be better considered as an upper limit.

The global warming potential (GWP) is calculated relative to CO2 over a specified time horizon from a model which also takes into account the RE values and tropospheric lifetimes. In some cases, CFCl3 (CFC-11) is used as the standard, and this GWP is called the halocarbon global warming potential (HGWP). The HGWP was calculated relative to CFC-11 using the following expression:1111 Fisher, D. A.; Hales, C. H.; Wang, W. C.; Ko, M. K. W.; Sze, N. D.; Nature 1990, 344, 513.

(4) HGWP = τ cis 3 hexene τ CFCl 3 M CFCl M cis 3 hexene RE cis 3 hexene RE CFCl 1 e 1 / τ cis 3 hexene 1 e 1 / τ CFCl 3

where τcis-3-hexene and τCFC-11CFCl3 ) are the corresponding tropospheric lifetimes; Mcis-3-hexene and MCFC-11 (MCFCl3 ) are the corresponding molar masses; REcis-3-hexene and RECFC-11 (RECFCl3) are the radiative efficiencies of the hexene and CFCl3, respectively, and t is the time horizon over which the RE is integrated.

The GWPs of the cis-3-hexene, relative to CO2, were calculated by multiplying the HGWP values by the scaling factors of 6730 and 4750 on a time horizon of 20 and 100 years, respectively.1212 Forster, P.; Ramaswamy, V.; Artaxo, P.; Berntsen, T.; Betts, R.; Fahey, D. W.; Haywood, J.; Lean, J.; Lowe, D. C.; Myhre, G.; Nganga, J.; Prinn, R.; Raga, G.; Schulz, M.; van Dorland, R. In Climate Change 2007: The Physical Science Basis; Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K. B.; Tignor, M.; Miller, H. L., eds.; Cambridge University Press: Cambridge, United Kingdom. 2007, ch. 2. These scaling factors are the GWP values of the CFC-11.

Results and Discussion

Rate coefficient for cis-3-hexene + Cl → products

The rate coefficient for the cis-3-hexene + Cl atoms reaction is the sum of the coefficients for the addition and abstraction channels and was measured at 296 ± 2 K and atmospheric pressure. The reference reactions are:

(5) Cl + n heptane p r o d u c t s , k 1

(6) Cl + cyclopentane p r o d u c t s , k 2

where k1 and k2 are the rate coefficients (in cm3molecule-1s-1): k1= (3.97 ± 0.27) × 10-10 and k2= (3.26 ± 1.0) × 10-10, as reported by Ezell et al.66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. and Wallington et al.,1313 Wallington, T. S.; Skewes, L. M.; Siegl, W. O.; J. Phys. Chem. 1989, 93, 3649. respectively. Four experiments with each reference compound were performed.

Different initial concentrations of the reference compounds were used in each experiment.

Figures 1 and 2 show the plot of ln([cis-3-hexene]0/[cis-3-hexene]t) vs. ln([reference]0/[reference]t), using n-heptane and cyclopentane as reference compounds, respectively.

Figure 1
Plot of the kinetic data for the reaction of cis-3-hexene with Cl atom using n-heptane as reference compound. [S]0 and [S]t are the concentrations of the cis-3-hexene at times 0 and t, respectively; [R]0 and [R]t are the concentrations of the reference compound at times 0 and t, respectively.

Figure 2
Plot of the kinetic data for the reaction of cis-3-hexene with Cl atom using cyclopentane as reference compound. [S]0 and [S]t are the concentrations of the cis-3-hexene at times 0 and t, respectively; [R]0 and [R]t are the concentrations of the reference compound at times 0 and t, respectively.

The linearity of the data points is observed in all experiments, with correlation coefficients greater than 0.99. Moreover, the intercepts are close to zero, suggesting that the contribution of secondary reactions with the products of the reactions studied can be neglected.

The initial concentrations of cis-3-hexene and of the reference compounds are presented in Table 1, as well as the number of experiments, the khex/kref ratio and the khex rate coefficient.

Table 1
Initial concentration of the reactants (hex: cis-3-hexene and ref: reference compound), rate constant ratios (khex/kref) and the relative rate constant (khex) for the reaction of OH radicals with cis-3-hexene at 298 K and atmospheric pressure

Error propagation was considered to estimate the uncertainty on rate coefficients. As previously discussed,33 Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714. these uncertainties have been calculated by assuming both the standard error of the slopes of the logarithm concentration curves and the reported errors on the reference rate coefficients.66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813.,1313 Wallington, T. S.; Skewes, L. M.; Siegl, W. O.; J. Phys. Chem. 1989, 93, 3649. Errors due to sample handling and chromatographic method were introduced in the standard error of the slope.

At least four experiments using two reference compounds were performed and the khex rate coefficient was determined. Consequently, the final rate coefficient is a mean value achieved from all experiments and the uncertainty is equal to twice the standard deviation.

Reactivity and reaction mechanism

A comparison of the room temperature rate coefficients determined for the reactions of cis-3-hexene with Cl atoms and OH radicals33 Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714. shows that the former is 6.6 times higher. In order to (i) evaluate the rate coefficient for the reaction of Cl atoms with cis-3-hexene determined in this work, (ii) explain the higher reactivity towards Cl atoms and (iii) to infer about the reaction mechanism, the reactivity of a series of alkenes was compared and the contributions of hydrogen abstraction and electrophilic addition channels were evaluated. These issues can be assessed by comparing the rate coefficients of the reactions of a series of alkenes with Cl and OH radicals.

Concerning the differences between the rate coefficients for the reaction of the alkene with OH radicals and Cl atoms, let us first note that the electrophilic character of chlorine atom and OH radicals are different and can be investigated from the experimental electron affinity values, which are 3.61 and 1.827 eV for Cl and OH, respectively. Since both reactions are initiated by the electrophilic attack of the oxidation agent, the higher electrophilic character of the Cl atoms explains the higher rate coefficient for the reaction of hydrocarbons with this specie.

Moreover, the rate coefficients for the reactions of OH radicals with alkenes are always lower than the rate coefficients for the corresponding reactions with Cl. Rate coefficients, in cm3 molecule-1 s-1, at 296 ± 2 K and 1 atm, for the reactions of OH radicals with alkenes are: kOH,propene= 3.0 × 10-11,1414 Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Troe, J.; J. Phys. Chem. Ref. Data 1989, 18, 881. kOH,ethylene= 8.5 × 10-12, kOH,1-butene= 3.1 × 10-11, kOH,cis-2-butene= 5.6 × 10-11, kOH,1-pentene= 3.1 × 10-11, kOH,2-methyl-2-butene= 8.7 × 10-11, kOH,2,3-dimethyl-2-butene= 1.1 × 10-10,1515 Atkinson, R.; Chem. Rev. 1986, 86, 69. kOH,1-hexene= 3.7 × 10-11,1616 Atkinson, R.; Aschmann, S. M.; Int. J. Chem. Kinet. 1985, 17, 33. kOH,1-heptene= 3.9 × 10-11, kOH,1-octene= 4.1 × 10-11 and kOH,1-nonene= 4.3 × 10-11.1717 Aschmann, S. M.; Atkinson, R.; Phys. Chem. Chem. Phys. 2008, 10, 4159. Room temperature rate coefficients (cm3 molecule-1 s-1), at 1 atm, for the reactions of chlorine atoms with the same alkenes are: kCl,ethylene= 3.0 × 10-10,1818 Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Rossi, M. J.; Troe, J.; J. Phys. Chem. Ref. Data 1997, 26, 521. kCl,propene= 2.6 × 10-10,66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. kCl,1-butene= 3.0 × 10-10,1919 Orlando, J. J.; Tyndall, G. S.; Apel, E. C.; Riemer, D. D.; Paulson, S. E.; Int. J. Chem. Kinet. 2003, 35, 334. kCl,1-pentene= 3.9 × 10-10, kCl,2-methyl-2-butene= 4.0 × 10-10,66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. kCl,cis-2-butene= 3.1 × 10-10,2020 Kaiser, E. W.; Donahue, C. J.; Pala, I. R.; Wallington, T. J.; Hurley, M. D.; J. Phys. Chem. A 2007, 111, 1286. kCl,1-hexene= 4.0 × 10-10, kCl,1-heptene= 4.4 × 10-10, kCl,1-octene= 5.5 × 10-10 and kCl,1-nonene= 5.9 × 10-10.2121 Walavalkar, M.; Sharma, A.; Alwe, H. D.; Pushpa, K. K.; Dhanya, S.; Naik, P. D.; Bajaj, P. N.; Atmos. Environ. 2013, 67, 93.

The Cl and OH reactions with alkenes and the corresponding rate coefficients are collected in two different groups. In the first group, hydrogen atoms connected to carbon atoms at the double bound are replaced by methyl groups (ethylene, propene, cis-2-butene, 2-methyl-2-butene and 2,3-dimethyl-2-butene) and an increase on the rate coefficients, from kOH,ethylene to kOH,2,3-dimethyl-2-butene (or from kCl,propene to kCl,2-methyl-2-butene), is observed, as evidenced in Figure 3.

In Figure 3, rate coefficients for the OH reactions with alkenes are highly correlated with the number of hydrogen atoms replaced by methyl groups in the molecule (triangles), whereas similar correlation is not observed for the rate coefficients for Cl reactions (circles). However, neglecting the ethylene from this group, a much better correlation is found between the rate coefficients for the Cl reactions with alkenes and the number of replaced hydrogen atoms (black line), showing that the contribution of the replacement of a hydrogen atom by a methyl group to the reactivity towards chlorine atoms is greater than for the kinetics of OH reactions with alkenes. In fact, the rate coefficients were expected to increase, since the replacement of the hydrogen atom by an alkyl group causes the electronic density on the π orbitals to increase, favoring the electrophilic attack. Therefore, the higher slope observed for the kCl rate coefficients can also be attributed to the higher electrophilic character of the Cl atoms.

Figure 3
Replacement of hydrogen atoms by methyl groups in alkenes and comparison of the reactivity of their reactions with chlorine atoms and with OH radicals.

The second group comprises the OH and Cl reactions (and the corresponding rate coefficients) with 1-alkenes (Figure 4). The values of rate coefficients for OH radicals with 1-alkenes suggest that the increase of the side chain along a homolog series has a small effect in the rate coefficient when the double bond is at the terminal carbon atom.

Figure 4
Homologous series of alkenes and comparison of the reactivity between their reactions with chlorine atoms and with OH radicals.

Different from the trend observed for the kOH values, a significant increase is observed for the kCl rate coefficients, as evidenced in Figure 4. Note that the kOH rate coefficients (squares) are found in the range from 3.0 × 10-11 to 4.3 × 10-11cm3 molecule-1 s-1, whereas the kCl rate coefficients (circles) are found in the range from 2.5 × 10-10 to 6.0 × 10-10 cm3 molecule-1 s-1. Since the increase of the side chain along a homolog series produces a minor effect over the electronic density on the π orbitals, thus representing a minor contribution to the electrophilic addition, the different slopes observed for kOH and kCl in this group can only be attributed to the hydrogen abstraction channel. As the side chain increases in the homolog series, the number of hydrogen atoms also increases and therefore the contribution of the hydrogen abstraction channel to the global kinetics can be increased.

The question now is how to predict how much the contribution of each possible channel to the global kinetics is.

The rate coefficients for alkane reactions with OH radicals and chlorine atoms, where only the hydrogen abstraction channel is predominant, were also compared. The contribution of the carbon chain length increase in the alkanes reactivity towards OH radicals and Cl atoms is shown in Figure 5.

In Figure 5, the rate coefficients for the reaction of the OH radicals with propane (1.1 × 10-12cm3molecule-1s-1), butane (2.4 × 10-12cm3molecule-1 s-1), hexane (5.2 × 10-12cm3molecule-1s-1), heptane (6.8 × 10-12cm3molecule-1s-1), octane (8.1 × 10-12cm3molecule-1 s-1) and nonane (9.7 × 10-12cm3molecule-1 s-1) were determined by Atkinson,2222 Atkinson, R.; Atmos. Chem. Phys. 2003, 3, 2233. whereas the rate coefficient for the pentane + OH reaction (3.9 × 10-12 cm3 molecule-1 s-1) was determined by Sivaramakrishnan and Michael.2323 Sivaramakrishnan, R.; Michael, J. V.; J. Phys. Chem. A 2009, 113, 5047. The following rate coefficients (cm3 molecule-1 s-1) for the reaction of the Cl atoms with alkanes have been used: 1.4 × 10-10(propane),1919 Orlando, J. J.; Tyndall, G. S.; Apel, E. C.; Riemer, D. D.; Paulson, S. E.; Int. J. Chem. Kinet. 2003, 35, 334. 2.1 × 10-10 (butane),2424 Tyndall, G. S.; Orlando, J. J.; Wallington, T. J.; Dill, M.; Kaiser, E. W.; Int. J. Chem. Kinet. 1997, 29, 43. 2.5 × 10-10(pentane), 3.1 × 10-10(hexane), 3.6 × 10-10 (heptane), 4.1 × 10-10 (octane),2525 Hooshiyar, P. A.; Niki, H.; Int. J. Chem. Kinet. 1995, 27, 1197. and 4.3 × 10-10 (nonane).2626 Aschmann, S. M.; Atkinson, R.; Int. J. Chem. Kinet. 1995, 27, 613.

Figure 5
Homologous series of alkanes and comparison of the reactivity between their reactions with chlorine atoms and with OH radicals.

The comparison of the slopes for kCl rate coefficients (Figures 4 and 5) suggests that the contribution of the increase in carbon length to the reactivity is similar for alkanes and alkenes, suggesting that the contribution from the hydrogen abstraction channel to the kinetics of the alkenes + Cl reaction is similar to that for the corresponding alkane + Cl reactions. Same comparison for kOH shows that the rate coefficients for the reactions of alkenes increase with the carbon length faster: the slope for the kOH rate coefficients of the reactions of alkenes is almost twice the slope for the reactions of alkanes, suggesting that the hydrogen abstraction channel is of minor contribution to the alkene + OH reactions. Therefore, the hydrogen abstraction channel contributes more to the overall kinetics of the reactions of Cl with alkenes, than for the reactions of OH with alkenes.

A possible estimate for the addition and hydrogen abstraction branching ratios can be done on the basis of the structure-reactivity scheme suggested by Ezellet al.,66 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. which considers the rate coefficient (k) as a sum of terms corresponding to the direct abstraction of non-allylic hydrogen atoms, addition to the double bond and abstraction of the allylic hydrogen atoms:

(7) k = k alkyl + k add + k allyl

The first term in this sum is obtained from the following expression:

(8) k alkyl = k 1 o F x + k 2 o F X F Y + k 3 o F X F Y F Z

where k1o, k2o and k3o are the contributions of hydrogen abstractions from primary, secondary and tertiary groups, respectively, to the overall rate coefficient and F(X), F(Y) and F(Z) are factors necessary to take into account the neighboring group effects. Values for individual parameters were given by Atkinson.2727 Atkinson, R.; J. Phys. Chem. Ref. Data 1997, 26, 215.

The second term in equation 7, kadd, represents the contribution of the addition rate coefficient for the overall rate coefficient and takes into account the possible formation of primary and secondary radicals, two equivalent secondary radicals, tertiary and primary radicals or tertiary and secondary radicals.

The kallyl term in equation 7 is the contribution of the allylic hydrogen abstraction to the overall rate coefficient and assumes three possible values for the hydrogen bounded to primary, secondary or tertiary carbon atoms.

Taking this scheme into account, the overall rate coefficient for cis-3-hexene + Cl reaction can be predicted as 4.27 × 10-10 cm3 molecule-1 s-1, in agreement with our experimental result ((4.13 ± 0.51) × 10-10 cm3 molecule-1 s-1). Moreover, the branching ratios for addition and hydrogen abstraction channels are 70 and 30%, supporting the conclusion that the contribution of the hydrogen abstraction channel is not negligible.

Atmospheric implications

In the atmosphere, cis-3-hexene can also be removed by reactions with O3, NO3 and OH radicals. The effective lifetimes of cis-3-hexene with respect to reaction with each of the oxidants were calculated using the relationship expressed as:

(9) τ X = 1 k X X

with X = Cl atoms, OH and NO3 radicals and O3 molecules, using the estimated 12 h average day-time global concentration of OH radicals (1 × 1066 Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813. radicals cm-3),2828 Prinn, R. G.; Huang, J.; Weiss, R. F.; Cunnold, D. M.; Fraser, P. J.; Simmonds, P. G.; McCulloch, A.; Harth, C.; Salameh, P.; O'Doherty, S.; Wang, R. H. J.; Porter, L.; Miller, B. R.; Science 2001, 292, 1882. the 12 h average night-time concentration of NO3 radicals (5 × 1088 Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2010, 44, 1749. molecule cm-3)2929 Shu, Y.; Atkinson, R.; J. Geophys. Res. 1995, 100, 7275. and 24 h average O3 concentration (7 × 1011 molecule cm-3)3030 Logan, J. A.; J. Geophys. Res. 1985, 90, 10463. and considering average global concentrations of 1 × 104 atoms cm-3 of chlorine.3131 Wingenter, O. W.; Kubo, M. K.; Blake, N. J.; Smith, T. W.; Blake, D. R.; Rowland, F. S.; J. Geophys. Res. 1996, 101, 4331. From the rate coefficients available in the literature33 Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714.

4 Calvert, J. G.; Atkinson, R.; Kerr, J. A.; Madronich, S.; Moortgat, G. K.; Wallington, T. J.; Yarwood, G.; The Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press, Inc.: New York. 2000.
-55 Pfrang, C.; Martin, R. S.; Nalty, A.; Waring, R.; Canosa-Mas, C. E.; Wayne, R. P.; Phys. Chem. Chem. Phys. 2005, 7, 2506. and the experimental rate coefficient obtained in this work, the lifetimes were calculated and ranged from 1.27 h to 2.8 days. These calculations do not take into account local atmospheric conditions and seasonal variations which are capable of changing these oxidant concentrations.

The tropospheric lifetimes shown in Table 2 indicate that cis-3-hexene is rapidly removed by OH radicals and O3 during day-time and by NO3 radicals at night. The contribution of Cl atoms is small, but maybe significant in those areas with higher Cl concentrations. For instance, with an OH concentration of 5 × 105 cm-3, observed in the early morning hours,3232 Brauers, T.; Aschmutat, U.; Brandenburger, U.; Dorn, H.-P.; Hausmann, M.; Heβling, M.; Hofzumahaus, A.; Holland, F.; Plass-Dülmer, C.; Ehhalt, D. H.; Geophys. Res. Lett. 1996, 23, 2545. atomic chlorine concentrations of only 1% of that OH could contribute significantly to the chemical removal of volatile organic compounds in the marine boundary layer.3333 Finlayson-Pitts, B. J.; Keoshian, C. J.; Buehler, B.; Ezell, A. A.; Int. J. Chem. Kinet. 1999, 31, 491. Moreover, maximum Cl concentrations as high as 1 × 105 atom cm-3 have been reported in the marine boundary layer at mid-latitudes at dawn emphasizing the locally significant effect of Cl atoms on the concentration and lifetimes of some atmospheric organic compounds in both remote marine boundary layer and coastal urban regions.3434 Spicer, C. W.; Chapman, E. G.; Finlayson-Pitts, B. J.; Plastridge, R. A.; Hubbe, J. M.; Fast, J. D.; Berkowitz, C. M.; Nature 1998, 394, 353. Significant Cl concentrations may also be found in mid-continental polluted areas from photolysis of ClNO2, a particular pollutant formed at night by reactions of soluble chloride specie (emitted by anthropogenic sources) with nitrogen oxides, as reported by Thornton et al.3535 Thornton, J. A.; Kercher, J. P.; Riedel, T. P.; Wagner, N. L.; Cozic, J.; Holloway, J. S.; Dube, W. P.; Wolfe, G. M.; Quinn, P. K.; Middlebrook, A. M.; Alexander, B.; Brown, S. S.; Nature 2010, 464, 271. Therefore, the importance of chlorine reactions, with respect to the OH reactions, with hydrocarbons may also be inferred beyond the marine boundary layer.

Table 2
Calculated atmospheric lifetimes, τX (hours), of cis-3-hexene with OH and NO3 radicals, O3 molecules and Cl atoms

Another atmospheric concern of VOCs is their contribution to the greenhouse warming, as expected by the global warming potential (GWP) which is calculated relative to CO2 over a specified time horizon. In fact, GWPs for short-lived compounds have been reported recently.3636 Nilsson, E. J. K.; Nielsen, O. J.; Johnson, M. S.; Hurley, M. D.; Wallington, T. J.; Chem. Phys. Lett. 2009, 473, 233.

37 Baasandorj, M.; Knight, G.; Papadimitriou, V. C.; Talukdar, R. K.; Ravishankara, A. R.; Burkholder, J. B.; J. Phys. Chem. A 2010, 114, 4619.

38 Jiménez, E.; Antiñolo, M.; Ballesteros, B.; Martínez, E.; Albaladejo, J.; ChemPhysChem 2011, 11, 4079.
-3939 Peirone, S. A.; Barrera, J. A.; Taccone, R. A.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2014, 85, 92.

The plot of the cross-sections (cm2 molecule-1 cm-1) as a function of wavenumber (cm-1) of the cis-3-hexene is shown in Figure 6.

Figure 6
Infrared spectra of cis-3-hexene at 298 K.

The integrated IR absorption cross-section (500 and 1500 cm-1) value for the cis-3-hexene is 1.35 × 10-17 cm2mol-1cm-1. Uncertainties in the cross-section measurement arise from the following sources: the sample concentration (1%), sample purity (3%), path length (1%), spectrum noise and residual baseline offset after subtraction of background (1.5%). Considering these individual uncertainties, we quote a conservative uncertainty of ± 6%. Unfortunately, there are no literature data for the absorption cross-section of the studied hexene to compare with.

Table 3 shows the calculated value of the RE for the cis-3-hexene with the RE of CFC-1112 (in units of W m-2) and the calculated values of HGWP and GWP on a time horizon of 20 and 100 years.

Table 3
Global lifetime (τglobal), estimated RE, HGWP and GWP for the cis-3-hexene, over a specified time horizon of 20 and 100 years

Summarizing, the lifetime of the studied compound indicates that they will be removed from the troposphere in few hours. In addition, it is clear from the GWP values that these compounds will not have a significant contribution to the radiative efficiencies of climate change.

Conclusions

In this work the rate coefficient for the reactions of Cl atoms with cis-3-hexene has been determined using the relative rate method. The rate coefficients, obtained from the experiments with the two reference compounds (n-heptane and cyclopentane), showed negligible dispersion among the determinations. Our final recommended value is (4.13 ± 0.51) × 10-10 cm3 molecule-1 s-1.

This value is in agreement with the estimate from structure-reactivity relations, 4.27 × 10-10 cm3 molecule-1 s-1. Such analysis also suggests that 30% of the rate coefficient for the reaction of cis-3-hexene with chlorine atoms is due to abstraction of hydrogen atoms.

The tropospheric lifetimes indicate that cis-3-hexene is rapidly removed by the typical oxidants in troposphere and that the contribution of Cl atoms to the global chemical removal can be significant in areas with high concentrations of this species.

Acknowledgments

The authors thank the Brazilian National Council for Scientific and Technological Development, CNPq (PROSUL, Proc. 490252/2011-7), and CONICET, ANPCyT-FONCyT, MinCyT and SECyT-UNC of Argentina for financial support of this research.

References

  • 1
    Heath, J. S.; Koblis, K.; Sager, S. L.; J. Soil Contam. 1993, 2, 1.
  • 2
    Ramnäs, O.; Östermark, U.; Petersson, G.; J. Chromatogr. 1993, 638, 65.
  • 3
    Barbosa, T. S.; Peirone, S.; Barrera, J. A.; Abrate, J. P. A.; Lane, S. I.; Arbillla, G.; Bauerfeldt, G. F.; Phys. Chem. Chem. Phys. 2015, 17, 8714.
  • 4
    Calvert, J. G.; Atkinson, R.; Kerr, J. A.; Madronich, S.; Moortgat, G. K.; Wallington, T. J.; Yarwood, G.; The Mechanisms of Atmospheric Oxidation of the Alkenes; Oxford University Press, Inc.: New York. 2000
  • 5
    Pfrang, C.; Martin, R. S.; Nalty, A.; Waring, R.; Canosa-Mas, C. E.; Wayne, R. P.; Phys. Chem. Chem. Phys. 2005, 7, 2506.
  • 6
    Ezell, M. J.; Wang, W.; Ezell, A. A.; Soskin, G.; Finlayson-Pitts, B. J.; Phys. Chem. Chem. Phys. 2002, 4, 5813.
  • 7
    Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Teruel, M. A.; Lane, S. I.; Atmos. Environ. 2006, 40, 7298.
  • 8
    Dalmasso, P. R.; Taccone, R. A.; Nieto, J. D.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2010, 44, 1749.
  • 9
    Abrate, J. P. A.; Pisso, I.; Peirone, S. A.; Cometto, P. M.; Lane, S. I.; Atmos. Environ. 2013, 67, 85.
  • 10
    Pinnock, S.; Hurley, M. D.; Shine, K. P.; Wallington, T. J.; Smyth, T. J.; J. Geophys. Res. 1995, 100, 23227.
  • 11
    Fisher, D. A.; Hales, C. H.; Wang, W. C.; Ko, M. K. W.; Sze, N. D.; Nature 1990, 344, 513.
  • 12
    Forster, P.; Ramaswamy, V.; Artaxo, P.; Berntsen, T.; Betts, R.; Fahey, D. W.; Haywood, J.; Lean, J.; Lowe, D. C.; Myhre, G.; Nganga, J.; Prinn, R.; Raga, G.; Schulz, M.; van Dorland, R. In Climate Change 2007: The Physical Science Basis; Solomon, S.; Qin, D.; Manning, M.; Chen, Z.; Marquis, M.; Averyt, K. B.; Tignor, M.; Miller, H. L., eds.; Cambridge University Press: Cambridge, United Kingdom. 2007, ch. 2.
  • 13
    Wallington, T. S.; Skewes, L. M.; Siegl, W. O.; J. Phys. Chem. 1989, 93, 3649.
  • 14
    Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Troe, J.; J. Phys. Chem. Ref. Data 1989, 18, 881.
  • 15
    Atkinson, R.; Chem. Rev. 1986, 86, 69.
  • 16
    Atkinson, R.; Aschmann, S. M.; Int. J. Chem. Kinet. 1985, 17, 33.
  • 17
    Aschmann, S. M.; Atkinson, R.; Phys. Chem. Chem. Phys. 2008, 10, 4159.
  • 18
    Atkinson, R.; Baulch, D. L.; Cox, R. A.; Hampson Jr., R. F.; Kerr, J. A.; Rossi, M. J.; Troe, J.; J. Phys. Chem. Ref. Data 1997, 26, 521.
  • 19
    Orlando, J. J.; Tyndall, G. S.; Apel, E. C.; Riemer, D. D.; Paulson, S. E.; Int. J. Chem. Kinet. 2003, 35, 334.
  • 20
    Kaiser, E. W.; Donahue, C. J.; Pala, I. R.; Wallington, T. J.; Hurley, M. D.; J. Phys. Chem. A 2007, 111, 1286.
  • 21
    Walavalkar, M.; Sharma, A.; Alwe, H. D.; Pushpa, K. K.; Dhanya, S.; Naik, P. D.; Bajaj, P. N.; Atmos. Environ. 2013, 67, 93.
  • 22
    Atkinson, R.; Atmos. Chem. Phys. 2003, 3, 2233.
  • 23
    Sivaramakrishnan, R.; Michael, J. V.; J. Phys. Chem. A 2009, 113, 5047.
  • 24
    Tyndall, G. S.; Orlando, J. J.; Wallington, T. J.; Dill, M.; Kaiser, E. W.; Int. J. Chem. Kinet. 1997, 29, 43.
  • 25
    Hooshiyar, P. A.; Niki, H.; Int. J. Chem. Kinet. 1995, 27, 1197.
  • 26
    Aschmann, S. M.; Atkinson, R.; Int. J. Chem. Kinet. 1995, 27, 613.
  • 27
    Atkinson, R.; J. Phys. Chem. Ref. Data 1997, 26, 215.
  • 28
    Prinn, R. G.; Huang, J.; Weiss, R. F.; Cunnold, D. M.; Fraser, P. J.; Simmonds, P. G.; McCulloch, A.; Harth, C.; Salameh, P.; O'Doherty, S.; Wang, R. H. J.; Porter, L.; Miller, B. R.; Science 2001, 292, 1882.
  • 29
    Shu, Y.; Atkinson, R.; J. Geophys. Res. 1995, 100, 7275.
  • 30
    Logan, J. A.; J. Geophys. Res. 1985, 90, 10463.
  • 31
    Wingenter, O. W.; Kubo, M. K.; Blake, N. J.; Smith, T. W.; Blake, D. R.; Rowland, F. S.; J. Geophys. Res. 1996, 101, 4331.
  • 32
    Brauers, T.; Aschmutat, U.; Brandenburger, U.; Dorn, H.-P.; Hausmann, M.; Heβling, M.; Hofzumahaus, A.; Holland, F.; Plass-Dülmer, C.; Ehhalt, D. H.; Geophys. Res. Lett. 1996, 23, 2545.
  • 33
    Finlayson-Pitts, B. J.; Keoshian, C. J.; Buehler, B.; Ezell, A. A.; Int. J. Chem. Kinet. 1999, 31, 491.
  • 34
    Spicer, C. W.; Chapman, E. G.; Finlayson-Pitts, B. J.; Plastridge, R. A.; Hubbe, J. M.; Fast, J. D.; Berkowitz, C. M.; Nature 1998, 394, 353.
  • 35
    Thornton, J. A.; Kercher, J. P.; Riedel, T. P.; Wagner, N. L.; Cozic, J.; Holloway, J. S.; Dube, W. P.; Wolfe, G. M.; Quinn, P. K.; Middlebrook, A. M.; Alexander, B.; Brown, S. S.; Nature 2010, 464, 271.
  • 36
    Nilsson, E. J. K.; Nielsen, O. J.; Johnson, M. S.; Hurley, M. D.; Wallington, T. J.; Chem. Phys. Lett. 2009, 473, 233.
  • 37
    Baasandorj, M.; Knight, G.; Papadimitriou, V. C.; Talukdar, R. K.; Ravishankara, A. R.; Burkholder, J. B.; J. Phys. Chem. A 2010, 114, 4619.
  • 38
    Jiménez, E.; Antiñolo, M.; Ballesteros, B.; Martínez, E.; Albaladejo, J.; ChemPhysChem 2011, 11, 4079.
  • 39
    Peirone, S. A.; Barrera, J. A.; Taccone, R. A.; Cometto, P. M.; Lane, S. I.; Atmos. Environ 2014, 85, 92.

Publication Dates

  • Publication in this collection
    Nov 2017

History

  • Received
    20 Jan 2017
  • Accepted
    03 May 2017
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br