Вы находитесь на странице: 1из 7

Atmospheric Environment 43 (2009) 5996–6002

Contents lists available at ScienceDirect

Atmospheric Environment
journal homepage: www.elsevier.com/locate/atmosenv

Temperature-dependent rate coefficients for the reactions of Cl atoms


with methyl methacrylate, methyl acrylate and butyl methacrylate
at atmospheric pressure
Marı́a B. Blanco a, Iustinian Bejan b,1, Ian Barnes b, Peter Wiesen b, Mariano A. Teruel a, *
a
Instituto de Investigaciones en Fisicoquı́mica, I.N.F.I.Q.C., Departamento de Fisicoquı́mica, Facultad de Ciencias Quı́micas, Universidad Nacional de Córdoba,
Ciudad Universitaria, 5000 Córdoba, Argentina
b
Physikalische Chemie/FBC, Bergische Universitaet Wuppertal, Wuppertal, Germany

a r t i c l e i n f o a b s t r a c t

Article history: Rate coefficients for the gas-phase reactions of Cl atoms with a series of unsaturated esters
Received 5 June 2009 CH2]C(CH3)C(O)OCH3 (MMA), CH2]CHC(O)OCH3 (MAC) and CH2]C(CH3)C(O)O(CH2)3CH3 (BMA) have
Received in revised form been measured as a function of temperature by the relative technique in an environmental chamber with
24 August 2009
in situ FTIR detection of reactants. The rate coefficients obtained at 298 K in one atmosphere of nitrogen or
Accepted 24 August 2009
synthetic air using propene, isobutene and 1,3-butadiene as reference hydrocarbons were (in units of
1010 cm3 molecule1 s1) as follows: k(ClþMMA) ¼ 2.82  0.93, k(ClþMAC) ¼ 2.04  0.54 and k(ClþBMA) ¼
Keywords:
3.60  0.87. The kinetic data obtained over the temperature range 287–313 K were used to derive the
Acrylate-esters
Chlorine atoms following Arrhenius expressions (in units of cm3 molecule1 s1): k(ClþMMA) ¼ (13.9  7.8)  1015
Arrhenius parameters exp[(2904  420)/T], k(ClþMAC) ¼ (0.4  0.2)  1015 exp[(3884  879)/T], k(ClþBMA) ¼ (0.98  0.42)  1015
Troposphere exp[(3779  850)/T]. All the rate coefficients display a slight negative temperature dependence which
Marine environment points to the importance of the reversibility of the addition mechanism for these reactions. This work
constitutes the first kinetic and temperature dependence study of the reactions cited above.
An analysis of the available rates of addition of Cl atoms and OH radicals to the double bond of alkenes
and unsaturated and oxygenated volatile organic compounds (VOCs) at 298 K has shown that they can be
related by the expression: log kOH ¼ 1.09 log kCl  0.10. In addition, a correlation between the reactivity of
unsaturated VOCs toward OH radicals and Cl atoms and the HOMO of the unsaturated VOC is presented.
Tropospheric implications of the results are also discussed.
Ó 2009 Elsevier Ltd. All rights reserved.

1. Introduction Unsaturated VOCs are also extensively used in different


industries. Acrylates and methacrylates for example are employed
Isoprene is the major biogenic species released to the atmo- in the extraction of chemicals, as intermediates in the manufacture
sphere (Graedel et al., 1986; Koppmann, 2007). The atmospheric of polymers and in the manufacture of circuit-boards (Graedel
degradation of this unsaturated volatile organic compound (VOC) et al., 1986). This family of unsaturated esters is toxic and volatile
leads to the formation of the unsaturated oxidation products and is a potentially major source of environmental concern in local
methacrolein and methyl vinyl ketone (Calvert et al., 2000; Atkin- industrial areas. Substantial amounts of oxygenated unsaturated
son, 1986, 1997, 2000; Atkinson and Arey, 2003). On the other hand, VOCs are continually introduced into the atmosphere and the
1,3-butadiene is the major anthropogenic unsaturated VOC emitted kinetics and degradation pathways under atmospheric conditions
to the troposphere (Graedel et al., 1986; Koppmann, 2007). Its of many of these compounds are still largely unknown. The main
atmospheric degradation produces mainly the a,b-unsaturated loss processes in the troposphere are reaction with OH and NO3
compounds acrolein, 3-butenal and methyl vinyl ketone (Calvert radicals, O3 molecules and at high altitudes photolysis. The fast
et al., 2000; Atkinson, 1986, 1997, 2000; Atkinson and Arey, 2003). degradation of VOCs in the presence of NOx (NO þ NO2) promote
local ozone formation and the formation of other photooxidants,
which are ubiquitous in polluted urban areas. The oxidation
* Corresponding author. Fax: þ54 351 4334188.
E-mail address: mteruel@fcq.unc.edu.ar (M.A. Teruel).
products include semi-volatile substances that may contribute to
1
Present address: Department of Chemistry, University College Cork, College the formation of atmospheric particulate matter (Noda and
Road, Cork, Ireland. Ljungstrom, 2002).

1352-2310/$ – see front matter Ó 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.atmosenv.2009.08.032
M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002 5997

Chlorine atoms, present at an average global concentration of coupled Teflon mixing fans are mounted inside each chamber to
about 1  104 atoms cm3 in the atmosphere (Wingenter et al., ensure homogeneous mixing of the reactants. The photolysis
1996), can also play an important role in the atmospheric chemistry systems consist of 18 (405 L reactor) or 32 (1080 L reactor)
since atomic Cl generally reacts with organic compounds (VOCs) superactinic fluorescent lamps (Philips TL05 40 W), which are
much faster than OH radicals. Moreover, in coastal urban areas for evenly spaced around the reaction vessels. These lamps emit light
a brief period at dawn in the marine boundary layer it is estimated in the wavelength region from 320 nm to 480 nm, with
that the concentration of Cl atoms can be as high as 105 atoms cm3 a maximum intensity at 360 nm. The lamps are wired in parallel
or more (Spicer et al., 1998; Ezell et al., 2002). Thus, reactions with and can be switched individually, which allows a variation of the
chlorine atoms in coastal areas may represent a significant sink for light intensity, and thus also the photolysis frequency/atom
VOCs. While there are many studies of Cl atom reactions with production rate, within the chamber. Each chamber is equipped
a variety of oxygenated volatile organic compounds (OVOCs), only with a White type multiple-reflection mirror system for sensitive
limited information is available for unsaturated OVOCs (Orlando in situ long path absorption monitoring of reactants and products
et al., 2003; Ullerstam et al., 2001; Rodrı́guez et al., 2005, 2007). The in the IR spectral range 4000–700 cm1. The White systems were
database for the reactions of unsaturated OVOCs with Cl atoms as operated at total optical path lengths of 50.4  0.1 m (405 L
a function of temperature is very limited (Rodrı́guez et al., 2007), reactor) and 484.7  0.8 m (1080 L reactor). IR spectra were
and there are virtually no temperature-dependent studies for the recorded with a spectral resolution of 1 cm1 using a Nicolet
reactions of Cl atoms with unsaturated esters. Magna 550 spectrometer (405 L reactor) and a Nicolet Nexus FT-
As part of a systematic study on the kinetics and oxidation IR spectrometer (1080 L reactor), each spectrometer is equipped
products of the atmospheric reactions of unsaturated oxygenated with a liquid nitrogen cooled mercury-cadmium-tellurium (MCT)
VOCs, we report in this work, relative determinations of rate detector.
coefficients for the reactions of Cl atoms with methyl methacrylate Chlorine atoms were generated by photolysis of molecular Cl2:
(MMA), methyl acrylate (MAC) and butyl methacrylate (BMA) over
the temperature range 287–313 K: Cl2 D hn / 2Cl (4)

In the presence of Cl atoms, the esters and the references decay


Cl D CH2 ]CðCH3 ÞCðOÞOCH3 / products (1)
through the following reactions:

Cl D CH2 ]CHCðOÞOCH3 / products (2) Cl D ester / products; kester (5)

Cl D CH2 ]CðCH3 ÞCðOÞOðCH2 Þ3 CH3 / products (3) Cl D reference / products; kref (6)
The studies have been performed in a large volume photoreactor Provided that the reference compound and the reactant are lost
using FTIR to monitor the reactants. To the best of our knowledge, only by reactions (5) and (6), then it can be shown that:
the rate coefficients and temperature dependencies of the above
reactions have not been previously reported. This work, therefore,    
½Ester0 kester ½Reference0
represents the first kinetic and temperature dependence study of ln ¼ ln (I)
½Estert kref ½Referencet
the cited reactions.
The aim of this study was to extend our earlier work on the where, [Ester]0, [Reference]0, [Ester]t and [Reference]t are the
reactivity of unsaturated esters toward OH radicals (Teruel et al., concentrations of the esters and reference compound at times t ¼ 0
2006; Blanco et al., 2006; Blanco and Teruel, 2008) and Cl atoms as and t, respectively and kester and kref are the rate coefficients of
a function of temperature. Additionally, the reactivity of the reactions (5) and (6), respectively.
unsaturated esters toward Cl atoms has been correlated with the The relative rate technique relies on the assumption that both
reactivity of the same unsaturated esters toward OH radicals and the esters and reference organics are removed solely by reaction
with the HOMO of the esters of the Frontier Molecular Orbital with Cl atoms, with irradiation times of 15 min maximum. To verify
Theory (FMOT). this assumption, mixtures of chlorine, air or nitrogen and both
Kinetic data for these Cl reactions are needed to facilitate organics (ester and reference hydrocarbon) were prepared and
a better understanding of the oxidation mechanisms of esters so allowed to stand in the dark for 30 min. In all cases, reaction of the
that their role in tropospheric atmospheric chemistry can be better organic species with the Cl-atom precursor Cl2, in the absence of UV
assessed. The measurements will also help to develop a more light, was of negligible importance over the typical experimental
reliable structure–reactivity relationship for these unsaturated time periods used in this work. Additionally, to test for possible
compounds. Residence times in the atmosphere for the esters photolysis of the reactants mixtures of the esters and reference
studied have been calculated and their atmospheric implications compounds in air or nitrogen in the absence of molecular chlorine
assessed. were irradiated using the output from all the fluorescent lamps
surroundings the chambers for 30 min. Photolysis was not observed
2. Experimental for any of the reactants.
To test for possible OH radical formation via secondary reactions
All the experiments were performed in synthetic air in with O2 that could interfere in the kinetic results (Kaiser and
a 1080 L quartz–glass reaction chamber at a total pressure of Wallington, 1996) separate experiments were performed in the
760 Torr (760 Torr ¼ 101.325 kPa). Additionally, for the reaction of presence and in the absence of O2 for the reaction of MMA with Cl
Cl with MMA experiments were also performed in a 405 L in the 1080 L reactor.
borosilicate glass reactor using synthetic air or nitrogen as bath The following typical initial concentrations in ppmV
gas. Detailed descriptions of the reactors can be found elsewhere (1 ppmV ¼ 2.46  1013 molecule cm3 at 298 K and 760 Torr of total
and only a brief general overall description is given here (Barnes pressure) were used for the esters and reference hydrocarbons in
et al., 1982, 1983, 1994). The chambers can be evacuated by the experiments: (0.5–0.9) ppmV for MMA; (0.3–0.6) ppmV for
pumping systems consisting of a turbo-molecular pump backed MAC, (0.6–0.9) ppmV for BMA; (1.4–1.8) ppmV for propene;
by a double stage rotary fore pump to 103 Torr. Magnetically (1.4–1.9) ppmV for isobutene; and (0.9–1) ppmV for 1,3-butadiene.
5998 M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002

The initial concentration of Cl2 was typically between 4.6 and 0.60
12.4 ppmV.
The chemicals N2 (Air Liquide, 99.999%), synthetic air (Air
Liquide, 99.999%) MMA (Aldrich, 99%), MAC (Aldrich, 99%), BMA
0.45
(Aldrich, 99%), isobutene (Messer Griesheim 99%), 1,3-butadiene
(Aldrich 99þ%), propene (Messer Griesheim 99.95%) and Cl2

ln[MAC]0/[MAC]t
(Messer Griesheim, 2.8) were used without further purification.
0.30

3. Results

3.1. Room-temperature rate coefficients


0.15

Figs. 1–3 show plots of the kinetic data according to equation (I)
for reactions (1–3) at 298 K using propene, isobutene and 1,3-
butadiene as reference compounds. For each organic reactant 0.00
studied, several runs were performed for each rate coefficient 0.0 0.3 0.6 0.9 1.2
determination. In all cases the plots showed good linearity with ln[Reference]0/[Reference]t
zero or near zero intercepts. Rate coefficient ratios derived from
linear least-squares analyses of each experimental data set are Fig. 2. Plots of the kinetic data for the reaction of Cl atoms with MAC in the 1080 L
listed in Table 1. The ratios in Table 1 were placed on an absolute reactor in air at 298 K using isobutene (,), 1,3-butadiene (B) and propene (6) as
reference hydrocarbons.
basis using the following 298 K rate coefficients for the reactions
of Cl with the reference hydrocarbons: k(Cl þ propene) ¼ (2.37 
0.47)  1010 cm3 molecule1 s1 (Ceacero Vega et al., in prepara-
Therefore, in all cases we prefer to quote rate coefficients for
tion); k(Cl þ isobutene) ¼ (3.40  0.28)  1010 cm3 molecule1 s1
the reactions of Cl with the esters, which are averages of all the
(Ezell et al., 2002); k(Cl þ 1,3-butadiene) ¼ (4.20  0.4) 
determinations under the different conditions, i.e. reference
1010 cm3 molecule1 s1 (Ragains and Finlayson-Pitts, 1997).
hydrocarbon and bath gas. Averaging the values of the rate
For all the esters, there is good agreement between the values of
coefficients and taking errors which encompass the extremes of
kester determined using two or three different reference
the determinations for each reaction results in the following
compounds. Additionally, in the case of MMA there is excellent
values for the reaction rate coefficients at 298 K (listed as kfinal
agreement with the results obtained in two different environ-
in Table 1):
mental chambers and bath gases, i.e. the average rate coefficient
obtained in the 405 L reactor k(ClþMMA) ¼ (2.82  0.40) 
kðClDMMAÞ [ ð2:82 ± 0:93Þ 3 10L10 cm3 moleculeL1 sL1 (II)
1010 cm3 molecule1 s1 is in good agreement with the average
values of (2.80  0.70)  1010 and (2.85  0.47) 
1010 cm3 molecule1 s1 obtained in the 1080 L reactor in air and kðClDMACÞ [ ð2:04 ± 0:54Þ 3 10L10 cm3 moleculeL1 sL1 (III)
N2, respectively. The good agreement between the results for Cl
with MMA in both air and N2 shows the kinetic rate determinations
kðClDBMAÞ [ ð3:60 ± 0:87Þ 3 10L10 cm3 moleculeL1 sL1 (IV)
were independent of the bath gas used and also that possible
inference by any OH radicals formed when using air as diluent gas The uncertainties are a combination of the 2s statistical errors
was negligible. from the linear regression analysis, to which an estimated error of

1.00

1.2

0.75

0.9
ln[MMA]0/[MMA]t

ln[BMA]0/[BMA]t

0.50

0.6

0.25

0.3

0.00
0.0 0.3 0.6 0.9 1.2
ln[Reference]0/[Reference]t 0.0
0.0 0.3 0.6 0.9 1.2

Fig. 1. Plots of the kinetic data for the reaction of Cl atoms with MMA at 298 K using ln[Reference]0/[Reference]t
propene in the 1080 L reactor in air (6), isobutene in the 1080 L reactor in air (,) and
in N2 (-), isobutene in the 405 L reactor in air ( ) and 1,3-butadiene in the 1080 L Fig. 3. Plots of the kinetic data for the reaction of Cl atoms with BMA in the 1080 L
reactor in air (B), 1,3-butadiene in the 1080 L reactor in N2 (C) and 1,3-butadiene in reactor in air at 298 K using isobutene (,), 1,3-butadiene (B) and propene (6), as
the 405 L reactor in air ( ) as reference hydrocarbons. reference hydrocarbons.
M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002 5999

Table 1
Rate coefficient ratios and rate coefficients for the reactions of Cl atoms with methyl methacrylate (MMA), methyl acrylate (MAC) and butyl methacrylate (BMA) obtained at
room temperature (298 K) and atmospheric pressure using either air or nitrogen as bath gas.

Ester Bath gas Reference kester/kref kester/1010 cm3 molecule1 s1 kaverage/1010 cm3 molecule1 s1 kfinal/1010 cm3 molecule1 s1
a
MMA Air Isobutene 0.79  0.01 2.69  0.24 2.82  0.40 2.82  0.93
1,3-Butadiene 0.70  0.01 2.94  0.32

MMAb Air Isobutene 0.85  0.02 2.90  0.31 2.80  0.70


1,3-Butadiene 0.73  0.02 3.07  0.38
Propene 1.03  0.01 2.44  0.50
N2 Isobutene 0.82  0.02 2.78  0.30 2.85  0.47
1,3-Butadiene 0.69  0.02 2.92  0.36

MACb Air Isobutene 0.62  0.02 2.11  0.24 2.04  0.54 2.04  0.54
1,3-Butadiene 0.51  0.01 2.16  0.25
Propene 0.78  0.02 1.85  0.41

BMAb Air Isobutene 1.11  0.01 3.77  0.34 3.60  0.87 3.60  0.87
1,3-Butadiene 0.90  0.01 3.78  0.40
Propene 1.37  0.02 3.25  0.69
a
405 L reactor.
b
1080 L reactor.

20% has been added to cover errors in the rate coefficients of the reaction of Cl atoms with the unsaturated esters studied to be peer-
reference compounds. reviewed. Rate coefficients for the reactions Cl atoms with MAC
and MMA of 2.0  1010 and 2.5  1010 cm3 molecule1 s1,
3.2. Temperature dependencies of the rate coefficients respectively, determined at room temperature have been previ-
ously reported in a doctoral thesis (Thévenet, 2000) but to the
The temperature dependences of the reactions 1–3 were studied authors’ knowledge have not been subjected to peer review.
between 287 and 313 K using propene as a reference compound. The values are in good agreement with the values determined in
The rate coefficient ratios derived from linear least-squares the present work.
analyses of each experimental data set at each temperature are
listed in Table 2 together with the second-order absolute rate 4.1. Reactivity trends
coefficients derived from the ratios using the rate coefficients for Cl
with propene calculated for the respective temperatures from the Since the reactions of Cl and OH proceed via similar addition
following Arrhenius expression between 287 and 313 K (Ceacero mechanisms (Orlando et al., 2003) it is interesting to compare the
Vega et al., in preparation): reactivity of the unsaturated esters studied in this work toward Cl
atoms with that toward OH radicals. Such a comparison is made in
kðCl D propeneÞ [ ð1:59 ± 0:22Þ 3 10L14 expðð2864 ± 400Þ=TÞ Table 3 where rate coefficients for the reactions of Cl and OH with
(V) other unsaturated VOCs are also included. It is evident from Table 3
For all three compounds, the reaction rate coefficients were that the olefinic units of the esters are more reactive toward Cl than
found to decrease with increasing temperature in the temperature OH and that the effect of substituents on the reactivity of the
range 287–313 K, as shown by plots of ln k versus 1/T in Fig. 4. A olefinic unit toward Cl atoms is much less important than that for
linear least-squares analysis of the plots yields the activation OH reactions. The rate coefficients for the reactions of Cl atoms with
energy and the pre-exponential factor for each reaction as given in unsaturated compounds are close to the gas kinetic limited value
Table 2. The following Arrhenius expressions, derived from the data from gas collision theory, thus, the reaction probability of Cl atoms
in Table 2, can be used to represent the reactions of Cl with the
esters in the range 287–313 K: Table 2
Summary of the rate coefficient ratios, individual values of the absolute rate coef-
kðClDMMAÞ [ ð13:9 ± 7:8Þ 3 10L15 exp½ð2904 ± 420Þ=T (VI) ficients and Arrhenius parameters for the reactions of Cl atoms with the unsaturated
esters obtained in this work over the temperature range 287–313 K.

T kester/ k  1010 (cm3 Reactant Ea/R (K) A  1015 (cm3


kðClDMACÞ [ ð0:4 ± 0:2Þ 3 10L15 exp½ð3884 ± 879Þ=T (VII)
(K) kpropene molecule1 s1) molecule1 s1)
287 0.98  0.01 3.35  0.76 Methyl 2904  420 13.9  7.8
kðClDBMAÞ [ ð0:98 ± 0:42Þ 3 10L15 exp½ð3779 ± 850Þ=T (VIII) 293 0.99  0.01 2.76  0.75 methacrylate
298 1.03  0.01 2.44  0.67
The errors in the activation energies and the pre-exponential 303 1.02  0.02 2.06  0.63
313 0.97  0.01 1.45  0.58
factors are the 2s random errors from the fits to the data presented
in Table 2 and plotted in Fig. 4. 287 0.88  0.02 3.01  0.93 Methyl 3884  879 0.4  0.2
293 0.81  0.01 2.26  0.84 acrylate
298 0.78  0.02 1.85  0.77
4. Discussion 303 0.69  0.01 1.39  0.59
313 0.68  0.01 1.01  0.51
In this work, rate coefficients for the reactions of Cl with 287 1.45  0.01 4.96  1.25 Butyl 3779  850 0.98  0.42
CH2]C(CH3)C(O)OCH3 (MMA), CH2]CHC(O)OCH3 (MAC) and 293 1.42  0.01 3.96  1.01 methacrylate
CH2]C(CH3)C(O)O(CH2)3CH3 (BMA), at temperatures between 287 298 1.37  0.02 3.25  0.97
and 313 K have been obtained. To the best of our knowledge these 303 1.18  0.02 2.38  0.91
313 1.17  0.01 1.74  0.79
are the first reported room-temperature rate coefficients for the
6000 M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002

-21.0 Cl with the unsaturated ester can be attributed to Cl addition to


the double bond with the sum of the contributions from
abstractable hydrogens in the molecule playing a lesser role. This
-21.6 is evident if one compares the rate coefficient of MAC with Cl
(2.04  1010 cm3 molecule1 s1) with that of butyl acrylate Cl
ln k (cm molecule s )

(2.50  1010 cm3 molecule1 s1) where an increase in the rate


-1
-1

-22.2 coefficient by a factor of only 1.23 is observed upon introduction 6


abstractable hydrogens. Similarly, for the reactions of Cl with
MMA (2.82  1010 cm3 molecule1 s1) and Cl with BMA
(3.60  1010 cm3 molecule1 s1) an increase by a factor of 1.28
3

-22.8
is observed upon introduction 6 abstractable hydrogens. The
observed increases are in line with other studies on the reactions
of Cl with different alkenes which predict a contribution of
-23.4
around 20% or less for the H-atom abstraction channel (Ezell
et al., 2002; Orlando et al., 2003).
It is also apparent from the comparison in Table 3 that the
3.1 3.2 3.3 3.4 3.5 3.6
-1
reactivity of the double bond toward Cl is only marginally influ-
1000 / T (K ) enced by its position in the ester, i.e. adjacent to the carbonyl group
Fig. 4. Arrhenius plots for the reactions of Cl atoms with MMA (,); MAC(B) and BMA
as in MAC (2.04  1010 cm3 molecule1 s1) or adjacent to the
(6) between 287 and 313 K. alkoxy group as in vinyl acetate (2.68  1010 cm3 molecule1 s1).

4.2. Temperature dependencies


with unsaturated VOCs is high and the chemical structure conse-
quently plays only a limited role in determining the magnitude of There are no prior experimental determinations of the Arrhe-
the rate coefficient in comparison to the reactions of OH with the nius parameters for the reactions of Cl with MMA, MAC and BMA.
same unsaturated VOCs where the rates are an order of magnitude Hence, this is the first temperature dependence study for these
or more lower. reactions. This is also the first rate coefficient temperature depen-
The rate coefficients for Cl with the unsaturated esters are not dence study of the reactions of Cl atoms with unsaturated esters in
substantially different to those reported in the literature for their general. However, the values of the rate coefficients and activation
corresponding alkene (Ezell et al., 2002; Ullerstam et al., 2001), energies obtained in the present work appear entirely consistent
i.e. this would be propene in the case of MAC and isobutene in the with the available literature data for the reactions of Cl atoms with
case of MMA. Therefore, the ester –C(O)OR functionality is not unsaturated alcohols where Ea/R vales from 2670 and 8000 K
significantly deactivating the double bond toward Cl attack and have been reported (Rodrı́guez et al., 2007, 2008).
the overwhelming proportion of the reactivity in the reactions of The negative temperature dependence of these reactions can
be rationalized on a qualitative basis assuming that the lifetime
of the excited bimolecular complex, formed between the Cl atom
Table 3 and the ester, with respect to decomposition back to the reac-
Rate coefficients of the reactions of Cl atoms and OH radicals with different unsat-
urated VOCs and HOMO energies for the corresponding VOCs.
tants decreases as the temperature increases. Therefore, the
probability of the excited adduct being stabilized by collision with
VOC EHOMO kCl  1010 (cm3 kOH  1011 (cm3 a third body falls with increasing temperature. Alternatively one
(eV) molecule1 s1) molecule1 s1)
can treat the termolecular reactions as the sum of bimolecular
1 CH2]CHCN 10.886 1.11a 1.11a
reactions which result in an overall effective activation energy
2 CH2] CHCH3 10.105 2.45b 3.01b
3 CH2]CHOCH3 9.6205 – 4.50c for the global reaction, i.e. the overall activation energy is the
4 CH2]CHOCH2CH3 9.5707 3.30b 7.30b sum of the energies for the individual steps. If the activation
5 CH2]CHO(CH2)2CH3 9.5969 – 10.00b energy for the energized complex back to reactants is greater than
6 CH2]CHO(CH2)3CH3 9.5985 – 11.00c the sum of the reactions to form products, then the effective
7 CH2]CHOCH2CH(CH3)CH3 9.5274 – 10.80c
8 CH2]CHCH2OH 10.3632 1.72d 4.55b
activation energy for the reaction becomes negative and the rate
9 CH2]CHC(O)H 10.6936 2.10b 1.96b coefficient decreases as the temperature increases (Finlayson-Pitts
10 CH2]CHC(O)CH3 10.7266 2.20b 1.85b and Pitts, 2000).
11 CH2]CHC(O)OH 11.1457 3.99a 1.75a
12 CH2]CHC(O)OCH3 11.0667 2.04e 1.30f
4.3. Free energy relationships
13 CH2]CHC(O)OCH2CH3 11.0401 – 1.70f
14 CH2]CHC(O)O(CH2)3CH3 11.0441 2.50g 1.80h
15 CH2]C(CH3)C(O)OCH3 10.561 2.82e 4.15i Several authors have proposed a linear correlation between the
16 CH2]C(CH3)C(O)OCH2CH3 10.5258 – 4.58i rate coefficients of OH addition reactions and those of other
17 CH2]C(CH3)C(O)O(CH2)3CH3 10.5328 3.60e 7.08i tropospheric oxidants like NO3 radicals or O3 molecules which have
18 CH2]CHCH2OC(O)CH3 10.4073 1.30g 3.06b
19 CH2]CHOC(O)CH3 10.64 2.68j 2.48g
the same primary reaction mechanism (Atkinson, 1997; Wayne
a
et al., 1991; Baumgartner et al., 2002).
Teruel et al. (2007).
b In this work we extend the correlation with the corresponding
NIST-Chemical Kinetics Data Base on the Web.
c
Thiault and Mellouki (2006). for the Cl reactions with the same unsaturated VOCs.
d
Rodrı́guez et al. (2007). In this work we present for the first time a correlation between
e
This work. kOH and kCl for a wide range of different alkenes and unsaturated
f
Teruel et al. (2006). VOCs which contain halogen, alcohol, carbonyl, ether and ester
g
Blanco et al. (2009).
h
Blanco and Teruel (2008).
functionalities. Table 3 shows the rate coefficients for the reactions
i
Blanco et al. (2006). of different unsaturated VOCs with Cl atoms and OH radicals at
j
Blanco et al. (2009). 298 K.
M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002 6001

-19.5
CH2=C(CH3)C(O)O(CH2)3CH3
1E-10 CH3CH2CH2CH=CHC(O)H
CH2=CHC(CH3)2OH

CH2=CHCH3
-21.0
CH CH2CH=CHC(O)H CH2=C(CH3)C(O)OCH3 17
3 4 11

ln k (cm molecule s )
2 19

-1
14
k(OH)(cm molecule s )

CH2=CHCH2OC(O)CH3 10
15
-1

CH2=CHCH2CH3

-1
8 12
-22.5 9
-1

CH2=CHOC(O)CH3 6 18
CH2=CHCH2OH 5 1
CH2=CHCN 7
CH2=CHC(O)O(CH2)3CH3 4 17
CH2=CHC(O)H 8 16
1E-11 CH3CH2CH2CH2CH=CHC(O)H
-24.0 3 15

3
CH2=CHC(O)OCH3 2 18 14
3

CH2=CHC(O)CH3 9 11
CH2=CH2 10 13
CH2=CHCl 12
-25.5 1

-27.0
1E-12

1E-10 2E-10 3E-10 4E-10


9.1 9.8 10.5 11.2 11.9
3 -1 -1
k(Cl) (cm molecule s ) EHOMO (eV)

Fig. 5. Linear free energy plot of log kOH against log kCl at room temperature for
Fig. 6. Correlation plot of ln (k/cm3 molecule1 s1) versus calculated EHOMO for the
a series of unsaturated volatile organic compounds (VOCs) including the room-
reactions of OH radicals (B) and Cl atoms (,) with unsaturated VOCs. Filled squares
temperature rate coefficients for the reactions of Cl atoms with CH2]C(CH3)C(O)OCH3,
and circles are unsaturated esters from this work (red) and literature (black). The
CH2]CHC(O)OCH3 and CH2]C(CH3)C(O)O(CH2)3CH3 obtained in this work (red
numbers correspond to the numbered compounds listed in Table 3. (For interpretation
squares). (For interpretation of the references to colour in this figure legend, the reader
of the references to colour in this figure legend, the reader is referred to the web
is referred to the web version of this article.)
version of this article.)

The correlation obtained between the rate coefficients for the theory (MP4-SCF) using an ab initio Hamiltonian with a 6–311þþ
reactions of Cl atoms with a given alkene or unsaturated VOC and G(d,p) bases set. The correlations obtained in this work for this
those for the corresponding reactions with OH radicals is shown in group of reactions are as follows:
Fig. 5. A reasonable correlation is obtained and a least-squares
treatment of the data points in Fig. 5 yields the following expres- ln kOH ðcm3 moleculeL1 sL1 Þ [ Lð1:1 ± 0:2ÞEHOMO L ð12:6 ± 1:7Þ
sion (with the rate coefficients in units of cm3 molecule1 s1): (XI)

log kOH [ 1:09 log kCl L 0:10 (IX)


ln kCl ðcm3 moleculeL1 sL1 Þ [ Lð0:1 ± 0:1ÞEHOMO Lð21:2 ±2:9Þ
This correlation gives evidence that the mechanism of the (XII)
reactions of Cl atoms with the unsaturated VOCs is similar to that
The good quality of these correlations is such that estimations
observed for the reactions of OH with unsaturated VOCs, i.e. addi-
can be made of the rate coefficients for reactions which have not yet
tion to the double bond of the unsaturated VOC occurs in a primary
been investigated.
reversible step forming an adduct which can react to form products
in subsequent fast reactions (Calvert et al., 2000; Atkinson, 1986,
4.4. Tropospheric implications
1997, 2000; Atkinson and Arey, 2003).
The reactivity of different kinds of electrophiles such as O
With respect to the atmospheric implications of the reactions
atoms, OH and NO3 radicals and O3 molecules toward alkenes,
studied, the rate coefficients summarized in Table 1 can be used to
methyl-substituted alkenes and more recently to unsaturated
calculate the atmospheric lifetimes of the unsaturated esters with
oxygenated VOCs has been found to correlate with the ionization
respect to reaction with Cl atoms, which can be compared to the
potentials (IP) or the energy of the highest occupied molecular
corresponding lifetimes for their reactions with the other major
orbital (HOMO) of the unsaturated VOC (Atkinson, 1997; Wayne
atmospheric oxidants OH, NO3 and O3. Lifetimes have been
et al., 1991; Baumgartner et al., 2002). The EHOMO for the reactions
calculated using the expression: sx ¼ 1/kx[X] with X ¼ OH, Cl, NO3
of the saturated and unsaturated VOCs are presented in the
or O3, where kx is the rate coefficient for the reaction of the
literature in the form:
oxidant X with the ester and [X] is the typical atmospheric
ln kðcm3 moleculeL1 sL1 Þ [ aEHOMO D b (X) concentration of the oxidant. For the calculations the following
rate coefficients values were used: k1 ¼ 2.82  1010, k2 ¼ 2.04 
Thus, the use of the rate coefficients determined in this work 1010 and k3 ¼ 3.60  1010 cm3 molecule1 s1 all from this
together with other values for different kinds of unsaturated and work, k(OH þ MMA) ¼ 4.15  1011 (Blanco et al., 2006), k(OH þ
unsaturated oxygenated VOCs from the literature enables a rela- MAC) ¼ 1.33  1011 (Teruel et al., 2006), k(OH þ BMA) ¼
tionship with EHOMO (Table 3) to be proposed for OH/unsaturated 7.08  1011 (Blanco et al., 2006), k(NO3 þ MMA) ¼ 3.6  1015
VOCs (alkenes and oxygenated VOCs) and Cl/unsaturated VOCs. (Canosa-Mas et al., 1999), k(NO3 þ MAC) ¼ 1.0  1016 (Canosa-
These correlations are presented in Fig. 6. Mas et al., 1999), k(O3 þ MMA) ¼ 7.51  1018 (Grosjean et al.,
The HOMO energies for the alkenes and unsaturated oxygenated 1993), k(O3 þ MAC) ¼ 1.05  1018 (Grosjean and Grosjean, 1998).
VOCs (ethers, esters, aldehydes, ketones and alcohols) listed in Typical oxidant concentrations have been used: a 12 h average
Table 3 were calculated using the Gaussian 03 package. The OH radical concentration of [OH] ¼ 2  106 radicals cm3 (day-
geometry optimizations and initial values of energies were time) (Hein et al., 1997); [Cl] ¼ 1  104 atoms cm3 (Wingenter
obtained at the Hartree–Fock (HF) level, and ab initio Hamiltonian et al., 1996); [NO3] ¼ 5  108 radicals cm3 (nigth-time) (Shu and
with a 6–31þþ G(d,p) bases set was used. The self-consistent field Atkinson, 1995) and a 24 h average ozone concentration
energies were then calculated by Moller–Plesset perturbation [O3] ¼ 7  1011 molecules cm3 (Logan, 1985).
6002 M.B. Blanco et al. / Atmospheric Environment 43 (2009) 5996–6002

Table 4 Blanco, M.B., Taccone, R.A., Lane, S.I., Teruel, M.A., 2006. On the OH-initiated
Calculated atmospheric lifetimes of MAC, MMA and BMA with different tropospheric degradation of methacrylates in the troposphere: gas-phase kinetics and
oxidants. formation of pyruvates. Chem. Phys. Lett. 429 (4–6), 389–394.
Blanco, M.B., Teruel, M.A., 2008. Photodegradation of butyl acrylate in the tropo-
Concentration smethyl methacrylate smethyl acrylate sbutyl methacrylate sphere by OH radicals: kinetics and fate of 1,2-hydroxyalcoxy radicals. J. Phys.
(molecule cm3) Org. Chem. 21, 397–401.
Calvert, J.G., Atkinson, R., Kerr, J.A., Madronich, S., Moortgat, G.K., Wallington, T.J.,
Cl 1  104 4 days 6 days 3 days
Yarwood, G., 2000. Mechanisms of Atmospheric Oxidation of the Alkanes.
OH 2  106 3 h 10 h 2h
Oxford University Press.
NO3 5  108 6 days 232 days –
Canosa-Mas, C.E., Carr, S., King, M.D., Shallcross, D.E., Thompson, K.C., Wayne, R.P.,
O3 7  1011 2 days 16 days – 1999. A kinetic study of the reactions of NO3 with methyl vinyl ketone, meth-
acrolein, acrolein, methyl acrylate and methyl methacrylate. Phys. Chem. Chem.
Phys. 1, 4195–4202.
Ceacero Vega, A.A., Albaladejo, J., Ballesteros, B., Barnes, I., Bejan, I. Chem. Phys. Lett.,
The estimated tropospheric lifetimes at room temperature of in preparation.
the esters with the tropospheric oxidants are listed in Table 4. Ezell, M.J., Wang, W., Ezell, A.A., Soskin, G., Finlayson-Pitts, B.J., 2002. Kinetics of
Photolytic loss of the esters will be negligible since they are reactions of chlorine atoms with a series of alkenes at 1 atm and 298 K:
structure and reactivity. Phys. Chem. Chem. Phys. 4, 5813–5820.
photolytically stable in the actinic region of the electromagnetic Finlayson-Pitts, B.J., Pitts Jr., J.N., 2000. Chemistry of the Upper and Lower Atmo-
spectrum as pointed out previously (Teruel et al., 2006; Blanco sphere. Academic Press, N. Y.
et al., 2006). Graedel, T.E., Hawkins, D.T., Claxton, L.D., 1986. Atmospheric Compounds: Sources,
Occurrence, and Bioassay. Academic Press, Orlando, Florida.
Unfortunately, no data are available on the reactions of NO3 Grosjean, E., Grosjean, D., 1998. Rate constants for the gas-phase reaction of ozone
radicals or O3 molecules with BMA, however, on the basis of with unsaturated oxygenates. Int. J. Chem. Kinet. 30, 21–29.
structural similarities it is probable that it will show a similar Grosjean, D., Grosjean, E., Williams II, E.L., 1993. Rate constants for the gas-phase
reactions of ozone with unsaturated alcohols, esters, and carbonyls. Int. J. Chem.
reactivity toward NO3 radicals and O3 as MAC and MMA and thus
Kinet. 25, 783–794.
will have a similar lifetime with respect to reaction with these Hein, R., Crutzen, P.J., Heimann, M., 1997. An inverse modeling approach to
oxidants. Therefore the lifetimes indicate that the esters are likely investigate the global atmospheric methane cycle. Global Biogeochem. Cycles
11, 43–76.
to be removed rapidly in the gas phase, the reaction with OH being
Kaiser, E.W., Wallington, T.J., 1996. Kinetics of the reactions of chlorine atoms with
the major loss process for all the esters studied. The short lifetimes C2H4 (k1) and C2H2 (k2): a determination of DHf,298 for C2H3. J. Phys. Chem. A
for the unsaturated esters, in the range of 2–10 h, implies that they 100, 4111–4119.
are likely to be removed rapidly in the gas phase close to their Koppmann, R. (Ed.), 2007. Volatile Organic Compounds in the Atmosphere. Wiley-
Blackwell, New York.
source of emission. Nevertheless, in coastal areas and in the marine Logan, J.A., 1985. Tropospheric ozone-seasonal behaviour, trends, and anthropo-
boundary layer, where peak concentrations of Cl atoms as high as genic influence. J. Geophys. Res. 90, 463–482.
1  105 atom cm3 can occur during sunrise (Spicer et al., 1998; NIST-Chemical Kinetics Database on the Web – Standard Reference Database 17,
Version (Web Version), http://physics.nist.gov/PhysRefData/Contents.html.
Ezell et al., 2002), Cl-atom initiated degradation of methyl meth- Noda, J., Ljungstrom, E., 2002. Aerosol formation in connection with NO3 oxidation
acrylate, methyl acrylate and butyl methacrylate could contribute of unsaturated alcohols. Atmos. Environ. 36, 521–525.
to their atmospheric chemical processing. Orlando, J.J., Tyndall, G.S., Apel, E.C., Riemer, D.D., Paulson, S.E., 2003. Rate coeffi-
cients and mechanisms of the reaction of Cl-atoms with a series of unsaturated
hydrocarbons under atmospheric conditions. Int. J. Chem. Kinet. 35, 334–353.
Acknowledgement Ragains, M.L., Finlayson-Pitts, B.J., 1997. Kinetics and mechanism of the reaction of
Cl atoms with 2-methyl-1,3-butadiene (isoprene) at 298 K. J. Phys. Chem. A 101,
1509–1517.
The authors wish to acknowledge DAAD-PROALAR (Germany), Rodrı́guez, D., Rodrı́guez, A., Notario, A., Aranda, A., Dı́az-de-Mera, Y., Martı́nez, E.,
the Deutsche Forschungsgemeinschaft (DFG), the EU project 2005. Kinetic study of the gas-phase reaction of atomic chlorine with a series of
aldehydes. Atmos. Chem. Phys. 5, 5167–5182.
EUROCHAMP, SECYT (Argentina), CONICET (Argentina), ANPCyT- Rodrı́guez, D., Rodrı́guez, A., Soto, A., Aranda, A., Dı́az-de-Mera, Y., Notario, A., 2008.
FONCYT (Argentina), SECyT-UNC (Córdoba, Argentina), Fundación Kinetics of the reactions of chlorine atoms with 2-buten-1-ol, 2methyl-2-
Antorchas (Argentina), TWAS (Italy) and RSC (UK) for financial propen-1-ol, and 3-methyl-2-buten-1-ol as a function of temperature. J. Atmos.
Chem. 59, 187–197.
support of this research. Rodrı́guez, A., Rodrı́guez, D., Soto, A., Notario, A., Aranda, A., Dı́az-de-Mera, Y., Bravo, I.,
2007. Relative rate measurements of reactions of unsaturated alcohols with
atomic chlorine as a function of temperature. Atmos. Environ. 41, 4693–4702.
References Shu, Y., Atkinson, R., 1995. Atmospheric lifetimes and fates of a series of sesqui-
terpenes. J. Geophys. Res. 100, 7275–7282.
Atkinson, R., Arey, J., 2003. Atmospheric degradation of volatile organic compounds. Spicer, C.W., Chapman, E.G., Finlayson-Pitts, B.J., Plastridge, R.A., Hubbe, J.M.,
Chem. Rev. 103, 4605–4638. Fast, J.D., Berkowitz, C.M., 1998. Unexpectedly high concentrations of molecular
Atkinson, R., 2000. Atmospheric chemistry of VOCs and NOx. Atmos. Environ. 34, chlorine in coastal air. Nature 394, 353–356.
2063–2101. Teruel, M.A., Blanco, M.B., Luque, G.R., 2007. Atmospheric fate of acrylic acid and
Atkinson, R., 1997. Gas-phase tropospheric chemistry of volatile organic acrylonitrile: rate constants with Cl atoms and OH radicals in the gas phase.
compounds: 1. Alkanes and alkenes. J. Phys. Chem. Ref. Data 26, 215–290. Atmos. Environ. 41, 5769–5777.
Atkinson, R., 1986. Kinetics and mechanism of the gas-phase reactions of the Teruel, M.A., Lane, S.I., Mellouki, A., Solignac, G., Le Bras, G., 2006. OH reaction rate
hydroxyl radical with organic compounds under atmospheric conditions. Chem. constants and UV absorption cross-sections of unsaturated esters. Atmos.
Rev. 86, 69–201. Environ. 40 (20), 3764–3772.
Barnes, I., Bastian, V., Becker, K.H., Fink, E.H., Zabel, F., 1982. Reactivity studies of Thévenet, R., 2000. Ph.D. thesis: Etudes des cinétiques et mécanismes de dégra-
organic substances towards hydroxyl radicals under atmospheric conditions. dation atmosphérique de composés organiques volatils oxygénés: Aldéhydes,
Atmos. Environ. 16, 545–550. Cétones et Esters (Emissions automobiles et solvant). Orléans, France.
Barnes, I., Becker, K.H., Fink, E.H., Reimer, A., Zabel, F., Niki, H., 1983. Rate constant Thiault, G., Mellouki, A., 2006. Rate constants for the reaction of OH radicals with
and products of the reaction CS2 þ OH in the presence of O2. Int. J. Chem. Kinet. n-propyl, n-butyl, iso-butyl and tert-butyl vinyl ethers. Atmos. Environ. 40,
15, 631–645. 5566–5573.
Barnes, I., Becker, K.H., Mihalopoulos, N., 1994. An FTIR product study of the Ullerstam, M., Ljungstrom, E., Langer, S., 2001. Reactions of acrolein, crotonaldehyde
photooxidation of dimethyl disulfide. J. Atmos. Chem. 18, 267–289. and pivalaldehyde with Cl atoms: structure–activity relationship and compar-
Baumgartner, M.T., Teruel, M.A., Taccone, R.A., Lane, S.I., 2002. Theoretical study of ison with OH and NO3 reactions. Phys. Chem. Chem. Phys. 3, 986–992.
the relative reactivity of chloroethenes with atmospheric oxidants OH, NO3, Wayne, R.P., Barnes, I., Biggs, P., Burrows, J.P., Canosa - Mas, C.E., Hjorth, J., Le
O(3P), Cl(2P) and Br(2P. Phys. Chem. Chem. Phys. 4, 1028–1032. Bras, G., Moortgat, G.K., Pernon, D., Poulet, G., Restelli, G., Sidebottom, H., 1991.
Blanco, M.B., Bejan, I., Barnes, I., Wiesen, P., Teruel, M.A., 2009. OH-initiated The nitrate radical: physics, chemistry, and the atmosphere. Atmos. Environ.
degradation of unsaturated esters in the atmosphere: kinetics in the temper- Part A 25, 1–203.
ature range of 287–313 K. J. Phys. Chem. A 113, 5958–5965. Wingenter, O.W., Kubo, M.K., Blake, N.J., Smith, T.W., Blake, D.R., Rowland, F.S., 1996.
Blanco, M.B., Bejan, I., Barnes, I., Wiesen, P., Teruel, M.A., 2009. The Hydrocarbon and halocarbon measurements as photochemical and dynamical
Cl-initiated degradation of CH2]CHOC(O)CH3 , CH2]CHCH2OC(O)CH2CH3 and indicators of atmospheric hydroxyl, atomic chlorine, and vertical mixing
CH2]CHC(O)O(CH2)3CH3 in the troposphere. Environ. Sci. Pollut. Res. 16, 641–648. obtained during Lagrangian flights. J. Geophys. Res. 101, 4331–4340.

Вам также может понравиться