1
Kinetic measurements for the reactions of ozone with crotonaldehyde and its methyl derivatives and calculations of transition-state theory

2
The rate coefficients for the reactions of O3 with six unsaturated carbonyls have been measured with the relative-rate method in the presence of a sufficient radical scavenger.

3
The experiments were conducted using a 6-m3 reaction chamber combined with a long-path FTIR system.

4
The rate coefficients (measured in 10−18 cm3 molecule−1 s−1) were 1.58 ± 0.23 for crotonaldehyde, 1.59 ± 0.22 for trans-2-pentenal, 1.82 ± 0.26 for 3-methyl-2-butenal, 5.34 ± 0.73 for trans-2-methyl-2-butenal, 29.5 ± 4.1 for 3-pentene-2-one and 8.3 ± 1.1 for mesityl oxide.

5
Conventional transition-state theory (CTST) calculations based on ab initio molecular orbital and density functional methods were performed to evaluate the rate constants for nine unsaturated carbonyls including six compounds examined in the present experiment as well as acrolein, methacrolein, and methyl vinyl ketone.

6
A log–log plot of the rate coefficients measured in the present and previous works vs. the calculated results of the rate constants, showed a linear relationship.

Introduction

7
Unsaturated carbonyls are produced through the atmospheric oxidation of conjugated dienes emitted into the lower troposphere from vegetations1 and human activities such as road vehicles.2

8
The unsaturated carbonyls are then oxidized with OH radicals and/or O3.

9
Thus, kinetic data for this class of molecules are often necessary in the atmospheric model calculations.

10
For example, some unsaturated carbonyls with complex structures have recently been paid attention because these are possible intermediates of the formations of the secondary organic aerosols.3

11
However, for complex molecules, there are only preliminarily kinetic data4–6 because of the high reactivity and difficulties in the preparation of these molecules.

12
To predict unknown but necessary kinetic data, structure–activity relationships have often been applied to the electrophillic reactions.7,8

13
In this method, the logarithms of the rate coefficients are plotted as a function of the ionization potentials (IP) or the highest-occupied molecular orbital (HOMO) energies.

14
This plot shows an almost linear correlation for the reactions of OH radical with unsaturated hydrocarbons.

15
On the other hand, a correlation of the plot for the reactions of O3 with unsaturated hydrocarbons is much poorer than that for the reactions of OH radical.

16
As an example of the poor correlation, Grosjean and Grosjean pointed out that the rate coefficient for 1,2-distributed alkene, e.g., trans-2-butene, is ca. 20 times larger than that for 1,1-distributed alkene, e.g., isobutene, despite these ionization potentials being close to each other.9

17
The poor correlation is tentatively attributed to the difference in the steric factor between 1,1- and 1,2-distributed alkenes.

18
If this is the case, the method such as conventional transition-state theory (CTST), in which the pre-exponential factor is taken into account, must be applicable to predictions of unknown rate constants for the reactions with O3.

19
In this study, the rate coefficients have been measured for the reactions of O3 with crotonaldehyde (2-butenal) and five its methyl derivatives including both 1,1- and 1,2-methyl substituted unsaturated carbonyls.

20
The rate coefficients measured are compared with results of CTST calculations based on ab initio molecular orbital (MO) and density functional theory (DFT) calculations.

21
The purpose of this work is to determine these rate coefficients experimentally and to check whether CTST is applicable to the evaluations of these rate constants.

Experimental

22
Experiments were conducted with a bakeable and evacuable 6-m3 photochemical chamber whose inner surface was coated with perfluoroethylene–perfluoroalkyl vinyl ether copolymer.10,11

23
The rate coefficients were determined with a relative-rate method in the presence of sufficient diethyl ether (Et2O) as a scavenger of OH radicals produced through the O3 + alkene reactions.

24
Propylene and isobutene were used as reference alkenes.

25
Prior to each experiment, the chamber was filled with purified air under 101 kPa.

26
The desired pressures of the sample alkene, the reference alkene and Et2O were collected into calibrated bulbs.

27
These gases were flushed into the chamber with N2 carrier gas.

28
A diluent of O3/O2 was prepared as needed with an ozone generator (Nippon ozone, 2058) and then injected into the chamber.

29
Rapid mixing was ensured using two stirring fans.

30
The initial mixing ratios of the sample alkene, the reference alkene, Et2O and O3 were 0.5, 0.5, 19–84 and 1.1–2.4 ppmv, respectively.

31
Under these conditions, the concentrations for the sample and reference alkenes are presented in eqn. (1), assuming that both alkenes are only consumed through the reactions with O3where [X]0 and [X]t are the concentrations of X at times 0 and t, respectively; k1 and k2 are the rate constants for the reactions of O3 with the sample alkene and reference alkene, respectively.

32
The experimental data were analyzed according to eqn. (1).

33
The concentrations of the sample and reference alkenes were obtained from infrared absorption spectra measured by an FTIR spectrometer (Nicolet, Nexus 670) with a multi-reflection mirror system whose optical path length was 221.5 m.

34
The chemicals used and their stated purities were crotonaldehyde (CA: Wako, 99%, mainly trans), 3-methyl-2-butenal (MBA32: Aldrich, 97%), trans-2-methyl-2-butenal (MBA22: Aldrich, 96%), mesityl oxide (MSO: Aldrich, 90%), 3-pentene-2-one (PO32: Aldrich, 65%, mainly trans), trans-2-pentenal (PA2: Aldrich, 95%), Et2O (Kanto, 99.5%), propylene (Takachiho, 99%) and isobutene (Takachiho, 95%).

35
The impurity of PO32 was identified to be MSO with the FTIR analysis.

36
The concentration of MSO was shown to be ca.

37
15% of that of the used reagent.

38
All experiments were carried out at 298 ± 2 K.

Method of calculations

39
The rate constants were calculated with CTST for the reactions of O3 with nine unsaturated carbonyls, i.e., six molecules examined in the present experiment as well as acrolein (AC), methacrolein (MAC) and methyl vinyl ketone (MVK).

40
These reactions are believed to proceed on the singlet ground-state potential-energy surface via the addition of O3 to the CC double bond.

41
In these reactions, O3 approaches the double bond from the directions vertical to the alkene plane, leading to the transition-state (TS) involving a five-membered ring.12

42
There are two possible TS conformers referred as syn- and anti-TS;13where Ri and X indicate the alkyl and carbonyl groups, respectively.

43
Since the energies and all the bond lengths are close to each other between these TS conformers, it was assumed that the rate constants are the same between two reaction pathways via these TS conformers.

44
The rate constants were calculated with CTST for reaction pathways viasyn-TS molecules and were then doubled to determine the overall rate constants (k1CTST);14,15where E0, L and QX represent the barrier-height energy, the statistical factor and the partition function of the internal motions for molecule X, respectively.

45
The statistical factor was set to 4 because there are two configurations of ozone, i.e., O1–O2–O3 and O3–O2–O1, for each of two attack sites located in both sides of the alkene plane.

46
The partition functions for the intra-molecular rotations of methyl groups in unsaturated carbonyls were calculated with a method proposed by Troe.16

47
All rate constants were calculated for temperature of 298 K.

48
The vibrational frequencies, the rotational constants and the barrier-height energies necessary in the calculations of eqn. (2) were determined by ab initio MO and DFT calculations.17,18

49
Prior to the calculations, a conformer with the lowest energy was found with an RHF/6-31G(d) method for each reactant.

50
For the conformers found, the geometries for the reactant and the TS were optimized by the RHF/6-31G(d) method and the hybrid density functional consisting of Becke's three parameter nonlocal hybrid exchange potential with the nonlocal correlation functional of Lee, Yang, and Parr (B3LYP) method19 with the 6-31G(d,p) basis set.

51
Harmonic vibrational frequencies were calculated at the optimized geometries by the respective methods.

52
The results of vibrational frequencies were scaled with factors of 0.89 and 0.96 for the results of the RHF and B3LYP methods, respectively.

53
The single-point energies were calculated by the fourth order Møller–Plesset perturbation theory (MP4) with the 6-31G(d,p) basis set and the coupled-cluster singles and doubles including a perturbational estimate of triple excitations (CCSD(T)) method20 with the 6-311G(d,p) basis set for the geometries optimized by the RHF/6-31G(d) and B3LYP/6-31G(d,p) methods, respectively.

54
In order to examine validities of the basis sets used in the above calculations, geometry optimizations were also carried out at by the B3LYP/6-311G(d,p) method for the reactants and TS for the O3 + ACR reaction, and the single-point energies were also calculated at the CCSD(T) level with 6-311G(d,p), 6-311++G(d,p), 6-311G(2d,2p) and 6-311++G(2d,2p) basis sets.

55
The HOMO energies of unsaturated carbonyls were calculated by the B3LYP/6-31+G(d,p) method for the geometries optimized at the RHF/6-31G(d) level of theory.

Results and discussion

Experimental results of rate coefficient k1

56
The concentrations of the sample and the reference alkenes were determined from the infrared spectra measured as follows: A contribution of Et2O to the infrared spectrum measured was subtracted prior to each analysis.

57
The concentration of the reference alkene was then determined with an absorption line peaked at 912 cm−1 for propylene or 890 cm−1 for isobutene.

58
The concentration of the sample alkene was determined with an absorption line attributed to the CC stretching mode at around 1640 cm−1.

59
For PO32, the CC absorption line contained a contribution of the MSO impurity of ≤25% in the absorbance.

60
This component was subtracted with an absorption line at 968 cm−1 before the evaluation of the concentration of PO32.

61
All the concentrations could be determined without interference from any product absorption.

62
The concentrations of the sample and reference alkenes are plotted in Fig. 1 in accordance with eqn. (1) for the measurements with propylene as the reference alkene.

63
Since the plots for all measurements examined were almost linear, the relative rate coefficients (k1/k2) were obtained from the slopes with the linear least-squares analysis.

64
The results of k1/k2 are listed in Table 1 along with initial experimental conditions.

65
The initial conditions were varied to check validities of k1/k2 obtained as follows: First, k1/k2 of each compound were measured at both [Et2O]0 ≈ 80 and ≈ 20 ppmv to confirm that OH radicals were entirely scavenged.

66
For example, for CA, k1/k2 obtained at 18.9 ppmv (0.152 ± 0.016) shows good agreement with that obtained at 82.6 ppmv (0.152 ± 0.008).

67
Similar results were obtained for all the other compounds examined as listed in Table 1.

68
These results indicate that OH radicals are entirely scavenged even at [Et2O]0 ≈ 18 ppmv.

69
Next, if the decay rate of the sample alkene due to the reaction with O3 (k1[O3]t) was comparably as low as that due to the wall loss, k1/k2 obtained would be overestimated.

70
To check that there is no such overestimation, a dependence of k1/k2 upon [O3]0 was measured for MBA32.

71
As shown in Table 1, k1/k2 is almost independent of [O3]0 between 1.21 and 2.38 ppmv.

72
This overestimation is negligible even at [O3]0 = 1.21 ppmv.

73
Finally, for MSO, both propylene and isobutene were employed as the reference alkene. k1/k2 obtained for isobutene was 0.767 ± 0.011.

74
Multiplying this value to a literature value of k2 of isobutene,21 the rate coefficient k1 is determined to be (8.82 ± 0.35) × 10−18 cm3 molecule−1 s−1.

75
The result of k1 agrees with that obtained from propylene data, (8.3 ± 1.1) × 10−18 cm3 molecule−1 s−1, which is described later.

76
This suggests that a systematic error due to the use of propylene is negligible.

77
Since the possible systematic errors could be ruled out for all the results of k1/k2, the results of k1/k2 for each compound were averaged.

78
The results of averaged values are listed in Table 1.

79
The rate coefficients k1 were calculated with the averaged values of k1/k2 and a literature value of k2 for propylene,22 (1.04 ± 0.14) × 10−17 cm3 molecule−1 s−1.

80
The rate coefficients k1 determined are listed in Table 2 along with previous data.

81
To our knowledge, the previous kinetic data are available for CA and PO32.

82
For CA, two literature values have been reported.23,24

83
Both data were measured by monitoring the O3 decay curves with UV absorption methods in the presence of excess CA.

84
A literature value,21 (1.74 ± 0.20) × 10−18 cm3 molecule−1 s−1, is consistent with the present result, (1.58 ± 0.23) × 10−18 cm3 molecule−1 s−1, within the experimental uncertainties.

85
However, another literature value,24 (0.90 ± 0.18) × 10−18 cm3 molecule−1 s−1, is a factor of 0.57 lower than the present result.

86
For PO32, two literature values are available.24,25

87
A literature value,25 (35.0 ± 8.9) × 10−18 cm3 molecule−1 s−1, was measured with a relative-rate method.

88
This data agrees with the present result, (29.5 ± 4.1) × 10−18 cm3 molecule−1 s−1.

89
Another literature value, (21.3 ± 3.9) × 10−18 cm3 molecule−1 s−1, which is taken from ref. 24 is again slightly lower than the present data.

90
The rate coefficients measured in the present and previous works for unsaturated carbonyls were compared with those for unsaturated hydrocarbons to study the substituent effects on the rate coefficients.

91
Fig. 2 shows the available data of the rate coefficients plotted as a function of a series of five molecular structures, i.e., CH2CHX, CH3CHCHX, (CH3)2CCHX, C2H5CHCHX and CH3CHC(CH3)X, and compares the kinetic data for molecules with XH (kHC), CHO (kALD) and COCH3 (kKET).

92
The previous kinetic data plotted in the figure for ACR, MAC and MVK are listed in Table 2,12,23,24,26,27 and these for hydrocarbons, in units of 10−18 cm3 molecule−1 s−1, are 1.43 ± 0.19 for ethylene,22 10.4 ± 1.4 for propylene,22 11.5 ± 4.6 for isobutene,21 9.80 ± 0.29 for 1-butene21 and 129 ± 9 for cis-2-butene.21

93
Fig. 2 shows that the ratio of kALD to kHC was kept almost constant among all five structures.

94
An averaged value of kALD/kHC with an error of one standard deviation was 0.14 ± 0.06.

95
This means that the rate coefficient for unsaturated hydrocarbon is suppressed with a substitution of the H atom in the unsaturated hydrocarbon with an electron-withdrawing CHO group.

96
An averaged value of kKET/kALD was also calculated to be 14 ± 9.

97
In contrast to the result of kALD/kHC, the rate coefficient for unsaturated aldehyde is enhanced with a substitution of the H atom in the CHO group with an electron-releasing methyl group.

98
The results of the substituent effects suggest that the reactions of O3 with alkenes are electrophillic reactions.

Conventional structure–activity relationships

99
Assuming that the reactions of O3 with alkenes are electrophillic, a relationship of the present and previous experimental results of the rate coefficients with –EHOMO was studied.

100
Prior to this analysis, validities of the present results of –EHOMO were checked with previous experimental data of IP.

101
The results of –EHOMO calculated for nine compounds, i.e. six molecules examined in the present experiment as well as ACR, MAC and MVK are listed in Table 2 together with the available data of IP.28

102
Fig. 3 shows the data of IP plotted as a function of the results of –EHOMO.

103
With the linear least-square analysis, the data plotted can be approximated with a straight line, IP = (1.43 ± 0.08) × (–EHOMO) − 0.50 ± 0.59.

104
The errors quoted for the slope and the intercept are two standard deviations.

105
The correlation coefficient (R2) was 0.994 and was nearly the same as unity.

106
This suggests that the results of –EHOMO can be used instead of the experimental data of IP.

107
Fig. 3b shows the experimental rate coefficients for the reactions of O3 with nine unsaturated carbonyls plotted as a function of −EHOMO.

108
The data plotted in the figure are taken from Table 2.

109
All the data plotted in Fig. 3b were fitted with the linear least-squares analysis.

110
The line fitted was represented as log10k1 = −(2.83 ± 0.80) × (−EHOMO) + 2.50 ± 5.72.

111
The errors for the slope and the intercept are two standard deviations.

112
The correlation coefficient was 0.723 and was smaller than unity.

113
This indicates that the correlation of log10k1 with −EHOMO was so poor that the plot cannot be used for the predictions.

114
The data for seven compounds, i.e., PO32, MBA22, MVK, MAC, PA2, CA and ACR, can be approximated with a single straight line as shown in Fig. 3b with a dotted line.

115
The rate coefficients for 1,1-dimethyl alkenes, i.e., MBA32 and MSO, are lower than those predicted with the dotted line.

116
This result would be consistently attributed to the steric hinderance of 1,1-distributed substituents, as described by the previous workers.9

117
However, the data for 1-methyl-1-carbonyl alkenes, i.e., MBA22 and MAC, are fitted with a single line together with the data for 1,2-distributed alkenes, PO32, PA2 and CA.

118
This suggests that the present results of the scattered plot cannot simply be interpreted with the steric hinderance of the 1,1-distributed substituents.

Results of ab initio and DFT calculations

119
In order to study the reaction dynamics of O3 with unsaturated carbonyls and check whether accurate rate constants can be predicted, the calculations of the rate constants were performed with CTST based on the ab initio MO and DFT calculations.

120
In the previous theoretical studies on the reactions of O3 with alkenes, the barrier heights have been successfully predicted by employing the MP4/6-31G(d,p)//MP2/6-31G(d,p) method29 and the CCSD(T)/6-31G(d)//B3LYP/6-31G(d,p) method.30

121
In this work, we have employed the MP4(SDQ)//RHF/6-31G(d) and CCSD(T)/6-311G(d,p)//B3LYP/6-31G(d,p) methods for the reactions of O3 with nine unsaturated carbonyls.

122
The results of the energies and the zero-point energies (ZPE) calculated at the RHF/6-31G(d) level are listed in Table 3.

123
Table 4 shows the barrier-height energies (E0) for the respective reactions, calculated from the single-point energies and the ZPEs, as well as the available experimental data for MAC and MVK.27

124
For MAC, the barrier-height energies were evaluated as 23.9, 21.5, and 10.8 kJ mol−1 for the MP4(SDQ), CCSD, and CCSD(T) methods, respectively.

125
These values are close to the experimental activation energy of 17.5 kJ mol−1 and distribute within a range of ±8.4 kJ mol−1 (= ±2 kcal mol−1).

126
It is noted that an estimation of the barrier height with an error smaller than ±4.2–8.4 kJ mol−1 may be difficult even at a higher level of theory.

127
The higher the level of theory, the lower the result of the barrier height.

128
The difference in E0 between the CCSD and CCSD(T) methods are larger than that between the MP4(SDQ) and CCSD methods.

129
The similar tendency can be seen for the other reactions, as listed in Table 4.

130
This suggests that contributions of highly electronic excited configurations are important in the present reaction systems.

131
To check the dependence of the results of E0 upon the basis sets, CCSD(T) calculations were carried out for the O3 + ACR reaction, with 6-311G(d,p), 6-311++G(d,p), 6-311G(2d,2p), and 6-311++G(2d,2p) basis sets for geometries optimized by the B3LYP/6-311G(d,p) method.

132
The results are listed in Table 4.

133
The barrier-height energies determined by CCSD(T)/6-311G(d,p)//B3LYP/6-31G(d,p) and CCSD(T)/6-311G(d,p)//B3LYP/6-311G(d,p) are 20.1 and 22.9 kJ mol−1, respectively, and their difference is just 2.8 kJ mol−1.

134
For the same B3LYP/6-311G(d,p) geometries, the CCSD(T) barrier-height energies were evaluated as 22.9, 22.5, 15.8 and 20.4 kJ mol−1 with the 6-311311G(d,p), 6-311++G(d,p), 6-311G(2d,2p), and 6-311++G(2d,2p) basis sets, respectively.

135
As shown in these results, no systematic trend is seen as to the basis sets.

136
The width of the distribution of the results of E0 is ca. ±4.2 kJ mol−1.

137
This value is used as an uncertainty of the results of E0 in the present CTST calculations described in the next section.

Rate constants calculated with CTST method

138
The CTST calculations were performed using the results of E0 obtained at the MP4(SDQ), CCSD and CCSD(T) levels.

139
The pre-exponential factors, ACTST and k1CTST, were calculated following eqn. (2) for nine unsaturated carbonyls.

140
These results are listed in Table 4 together with results of the ab initio calculations of E0.

141
Fig. 4a shows the experimental results of the rate coefficients plotted as a function of k1CTST calculated with the results of E0 for the MP4(SDQ) method.

142
In Fig. 4a, the errors of the experimental data represent the experimental uncertainties, whereas the errors of k1CTST represent those evaluated taking into account the uncertainties of the results of E0, as described previously.

143
With the linear least-squares analysis, the data plotted in Fig. 4a were fitted with a straight line:log10k1 = a log10k1CTST + b,where a and b are fitting parameters.

144
Fit results of the parameters with errors of two standard deviations were a = 0.585 ± 0.107 and b = –7.04 ± 1.93.

145
The correlation coefficient was 0.863.

146
Figs. 4b and 4c show similar plots with E0 obtained with the CCSD and CCSD(T) methods, respectively.

147
These plots were also fitted with eqn. (3).

148
The parameters fitted were a = 0.352 ± 0.107 and b = −11.6 ± 1.8 for the CCSD method, and these were a = 0.395 ± 0.097 and b = −11.6 ± 1.5 for the CCSD(T) method.

149
The correlation coefficients were 0.692 and 0.778 for the CCSD and CCSD(T) methods, respectively.

150
The slopes of the fitted lines for Figs. 4a–c are smaller than unity.

151
This is because the results of E0 still contain errors even for the CCSD(T)/6-311G(d,p)//B3LYP/6-31G(d,p) method.

152
However, all the plots for the MP4(SDQ), CCSD and CCSD(T) methods show linear relationships within the errors of k1CTST.

153
Especially the correlation coefficients for the MP4(SDQ) and CCSD(T) methods are larger than that for the plot of log10k1vs. of −EHOMO (0.723).

154
CTST calculations combined with an ab initio MO and DFT methods can predict the relative values of the rate constants with a higher accuracy than those predicted with the conventional structure–activity relationships.

155
Finally, the difference in the rate coefficient between 1,2-methyl substituted alkene (MBA22) and 1,1-methyl substituted alkene (MBA32) is discussed using the present results of ACTST and E0.

156
The rate coefficients for the reaction of MBA22, (5.34 ± 0.73) × 10−18 cm3 molecule−1 s−1, is 2.9 times larger than that for the reaction of MBA32, (1.82 ± 0.26) × 10−18 cm3 molecule−1 s−1, as listed in Table 2.

157
If this is caused by the steric hindrance of 1,1-methyl substituents, the potential energy curves for the TS for the reaction of MBA32 should be tighter than those for the reaction of MBA22, and thus the result of ACTST for MBA32 should be smaller than that for MBA22.

158
However, the results of ACTST obtained at the B3LYP level for MBA32 (9.63 × 10−15 cm3 molecule−1 s−1) is very close to that for MBA22 (1.02 × 10−14 cm3 molecule−1 s−1) as listed in Table 4.

159
In addition, ACTST for MBA32 is also close to those for the other two isomers, PO32 (1.04 × 10−14 cm3 molecule−1 s−1) and PA2 (2.77 × 10−14 cm3 molecule−1 s−1).

160
These results indicate that the pre-exponential factor is almost independent of the molecular structure.

161
On the other hand, the result of E0 obtained at the CCSD(T) level for MBA32 (11.7 kJ mol−1) is higher than that for MBA22 (2.6 kJ mol−1), i.e., E0 depends on the molecular structure.

162
This suggests that the rate constant is determined by E0 rather than the pre-exponential factor.

163
Fig. 5 shows the geometries optimized with the B3LYP/6-31G(d,p) method for MBA32 and MBA32-TS molecules.

164
The reactant alkene has a structure with Cs symmetry.

165
In contrast, the methyl and carbonyl substituents in the TS move out of the original Cs plane into the opposite side of the O3 group.

166
The methyl groups move further from the original geometry via the intramolecular rotations.

167
These results suggest that motion of the TS along the intrinsic reaction coordinate consists of the relative translation between the O3 and alkene groups as well as the symmetric out-of-plane vibration of the alkene group and the intramolecular rotations of the methyl groups.

168
Since the geometry change of the alkene group is necessary for the formation of the TS, there is a substantial barrier for the reactions of O3 with alkenes.

169
The normal mode frequency assigned to the methyl rotation for MBA32 (127 cm−1) is larger than that for MBA22 (90 cm−1).

170
Since the tightness of the potential energy curves for the internal motions of alkene depends on the combination of the substituted sites, the barrier-height energy depends on the molecular structure.

Conclusion

171
In summary, the rate coefficients of the reactions of O3 with six unsaturated carbonyls were measured with a relative-rate method in the presence of sufficient diethyl ether.

172
Using the present and previous experimental results of the rate coefficients, the correlation of the rate coefficients with the HOMO energies was investigated, and it was found that the linearity of the plot is poor.

173
To evaluate the rate constants taking into account the reaction dynamics, conventional transition-state theory calculations based on ab initio molecular orbital and DFT methods were performed.

174
A log–log plot of the experimental rate coefficients vs. the results of the calculated rate constants show linear relationships for the results with the MP4(SDQ), CCSD and CCSD(T) barrier-height energies.

175
Among four isomers of MBA32, the result of the pre-exponential factor is almost independent of the reactant, while the result of the barrier-height energy depends on the reactant.

176
These results suggest that the rate constant is determined by the barrier-height energy rather than the pre-exponential factor.

177
In future works, both theoretical and experimental studies of the rate coefficients are necessary for the reactions of O3 with unsaturated carbonyls, which are considered to be possible intermediates of the secondary organic aerosols, such as cyclic carbonyl alkenes, dicarbonyl alkenes and carbonyl dienes.