the Creative Commons Attribution 4.0 License.
the Creative Commons Attribution 4.0 License.
Oligomer formation from the gas-phase reactions of Criegee intermediates with hydroperoxide esters: mechanism and kinetics
Long Chen
Yu Huang
Yonggang Xue
Zhihui Jia
Wenliang Wang
Download
- Final revised paper (published on 16 Nov 2022)
- Supplement to the final revised paper
- Preprint (discussion started on 08 Jun 2022)
- Supplement to the preprint
Interactive discussion
Status: closed
-
RC1: 'Comment on acp-2022-376', Anonymous Referee #1, 02 Jul 2022
General comments: The authors use density functional theory (DFT) and transition state theory (TST) to study two categories of Criegee intermediate (CI) reactions for the parent, anti-methyl, syn-methyl, and dimethyl CI. The first category is the reaction of each CI with HCOOH and the second category is the reaction of each CI with the hydroperoxy ester formed in the HCOOH reaction. The novelty is considering the hydroperoxy ester + CI reactions as a possible oligomerization mechanism. The theoretical methods are qualitatively reasonable and there are a few chemical insights. However, the atmospheric relevance of the work is either misrepresented or under-discussed.
Specific comments:
- The authors should explain their variational TST calculations for barrierless reactions (p. 7) in more detail, particularly since they consistently predict higher CI + HCOOH rate constants than experiment (pp. 10-11).
- The trend in exothermicity with substitution pattern (pp. 8-9) should be explained.
- The analysis of possible bimolecular CI reactions (p. 21) should be extended to the three substituted CIs.
- Since the CI is clearly the limiting reactant in the CI + HCOOH reaction, the atmospheric concentration of HPMF (and the other hydroperoxy esters) is much better estimated to be the CI concentration. (This, of course, will greatly lower the predicted pseudo-first-order rate constants for the CI + HPMF reaction.)
- Since a big motivation for the computations is the potential for CI + hydroperoxy ester reactions to lead to SOA, there should be some specific discussion, perhaps buttressed by rough calculations, of how many cycles of CI addition are required before a given adduct is expected to have low volatility. The approach of Chhantyal-Pun et al. (ACS Earth Space Chem. 2018, 2, 8, 833–842) is an example of the approach the authors should take.
Technical corrections:
- On p. 6, line 145: "saddle point" should be "minimum"
- On p. 6, line 162: "precision" should be "accuracy"
- On p. 7, line 182: "decomposes" should be "rearranges"
- On p. 14, lines 341-322, use a non-breaking hyphen
- On p. 15, line 372, "intermolecular" should be "intramolecular"
- On p. 17, lines 413-414, use a non-breaking hyphen
Citation: https://doi.org/10.5194/acp-2022-376-RC1 -
AC1: 'Reply on RC1', Y. Huang, 25 Aug 2022
Prof. Yu Huang
State Key Lab of Loess and Quaternary Geology
Institute of Earth Environment, Chinese Academy of Sciences, Xi’an, 710061, China
Tel./Fax: (86) 29-62336261
E-mail: huangyu@ieecas.cn
Aug. 25, 2022
Dear Prof. Kourtchev,
Revision for Manuscript ACP-2022-376
We thank you very much for giving us the opportunity to revise our manuscript. We highly appreciate the reviewers for their comments and suggestions on the manuscript entitled “Oligomer formation from the gas-phase reactions of Criegee intermediates with hydroperoxide esters: mechanism and kinetics”. We have made revisions of our manuscript carefully according to the comments and suggestions of reviewers. The revised contents are marked in blue color. The response letter to reviewers is attached at the end of this cover letter.
We hope that the revised manuscript can meet the requirement of Atmospheric Chemistry & Physics. Any further modifications or revisions, please do not hesitate to contact us.
Look forward to hearing from you as soon as possible.
Best regards,
Yu Huang
Comments of reviewer #1
- The authors should explain their variational TST calculations for barrierless reactions (p.7) in more detail, particularly since they consistently predict higher CI + HCOOH rate constants than experiment (p.10-11).
Response: In the original manuscript, the rate coefficients for the barrierless reactions are calculated by employing the variational transition state theory (VTST), and the rate coefficients for the bimolecular reactions with the tight transition states are computed by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction. For the initiation reactions of distinct stabilized Criegee intermediates (SCIs) with HCOOH, there are four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4), in which Entry 1 is barrierless and Entry 2-4 have the tight transition states. The total rate coefficient for the reaction of SCIs with HCOOH is equal to the sum of the rate coefficient of each pathway. For the barrierless 1,4 O-H insertion reaction, the VTST is approximated with a Morse potential function, V(R) = De{1-exp[-β(R-Re)]}2, along with an anisotropy potential function to stand for the minimum energy path, which is used to calculate the rate coefficients (Raghunath et al., 2017). Here, De is the bond energy excluding the zero-point energy, R is the reaction coordinate, and Re is the equilibrium value of R. It is assumed that the stretching potential in an anisotropy potential is used in conjunction with a potential form of Vanisotropy = V0[1-cos2(θ1–θ1e) × cos2(θ2–θ2e)] (Raghunath et al., 2017). Here, V0 is the stretching potential, which stands for by a Morse potential, θ1 and θ1e represent the rotational angle between fragment 1 and the reference axis and the equilibrium bond angle of fragment 1, θ2 and θ2e stand for the rotational angle between fragment 2 and the reference axis and the equilibrium bond angle of fragment 2. The association curve for the reaction of 1,4 O-H insertion of SCIs into HCOOH is computed at the M06-2X/6-311+G(2df,2p) level of theory to cover a range from 0.97 to 1.97 Å at step size 0.1 Å for O-H bond and from 1.44 to 2.44 Å at step size 0.1 Å for C-O bond, while other structural parameters are fully optimized. The computed potential energies are fitted to the Morse potential function. However, the calculated rate coefficients for the reactions of SCIs with HCOOH are higher than the prior experimental measurements. The reason is ascribed to the fact that the approximation of VTST using a Morse potential function in conjunction with an anisotropy potential function is unsuitable to predict the rate coefficients for the barrierless 1,4 O-H insertion reaction.
In the revised manuscript, the rate coefficients for the barrierless reactions are computed by employing the inverse Laplace transformation (ILT) method, and the rate coefficients for the bimolecular reactions with the tight transition states are calculated by utilizing CTST in conjunction with Eckart tunneling correction. The ILT and CTST/Eckart calculations are performed by using the MESMER 6.0 and KiSThelP 2019 programs, respectively (Glowacki et al., 2012; Canneaux et al., 2013). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken, 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OO is estimated to be 1.4 × 10-10 cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of HCOOH with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO (Table S4-S6). At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
Table S3 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of CH2OO with HCOOH computed at different temperatures
T/K
k (TSent1)
k (TSent2)
k (TSent3)
k (TSent4)
ktot-CH2OO
273
1.7 × 10-10
3.6 × 10-12
1.0 × 10-22
3.6 × 10-12
1.8 × 10-10
280
1.6 × 10-10
2.9 × 10-12
1.2 × 10-22
3.1 × 10-12
1.7 × 10-10
298
1.4 × 10-10
1.9 × 10-12
2.2 × 10-22
2.3 × 10-12
1.4 × 10-10
300
1.4 × 10-10
1.8 × 10-12
2.4 × 10-22
2.2 × 10-12
1.4 × 10-10
320
1.3 × 10-10
1.2 × 10-12
4.9 × 10-22
1.6 × 10-12
1.3 × 10-10
340
1.3 × 10-10
8.2 × 10-13
1.0 × 10-21
1.3 × 10-12
1.3 × 10-10
360
1.2 × 10-10
5.9 × 10-13
2.2 × 10-21
1.0 × 10-12
1.2 × 10-10
380
1.2 × 10-10
4.5 × 10-13
4.5 × 10-21
8.2 × 10-13
1.2 × 10-10
400
1.2 × 10-10
3.5 × 10-13
9.0 × 10-21
6.9 × 10-13
1.2 × 10-10
Table S4 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of anti-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-anti)
k (TSent2-anti)
k (TSent3-anti)
k (TSent4-anti)
ktot-anti
273
5.9 × 10-10
4.2 × 10-11
5.5 × 10-22
6.1 × 10-11
6.9 × 10-10
280
5.7 × 10-10
3.8 × 10-11
6.7 × 10-22
4.9 × 10-11
6.6 × 10-10
298
5.4 × 10-10
2.3 × 10-11
1.2 × 10-21
3.0 × 10-11
5.9 × 10-10
300
5.3 × 10-10
2.0 × 10-11
1.3 × 10-21
2.8 × 10-11
5.8 × 10-10
320
5.0 × 10-10
1.5 × 10-11
2.6 × 10-21
1.7 × 10-11
5.3 × 10-10
340
4.7 × 10-10
9.4 × 10-12
5.4 × 10-21
1.1 × 10-11
4.9 × 10-10
360
4.5 × 10-10
7.0 × 10-12
1.1 × 10-20
7.8 × 10-12
4.7 × 10-10
380
4.4 × 10-10
3.6 × 10-12
2.1 × 10-20
5.6 × 10-12
4.5 × 10-10
400
4.3 × 10-10
2.0 × 10-12
4.0 × 10-20
4.2 × 10-12
4.4 × 10-10
Table S5 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of syn-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-syn)
k (TSent2-syn)
k (TSent3-syn)
k (TSent4-syn)
ktot-syn
273
3.1 × 10-10
9.5 × 10-13
4.6 × 10-27
7.5 × 10-16
3.1× 10-10
280
2.8 × 10-10
8.0 × 10-13
7.1 × 10-27
6.4 × 10-16
2.8× 10-10
298
2.7 × 10-10
5.4 × 10-13
8.9 × 10-26
5.5 × 10-16
2.7× 10-10
300
2.7 × 10-10
5.2 × 10-13
9.9 × 10-26
4.6 × 10-16
2.7× 10-10
320
2.5 × 10-10
3.6 × 10-13
3.0 × 10-25
3.8 × 10-16
2.5× 10-10
340
2.5 × 10-10
2.6 × 10-13
9.1 × 10-25
3.1 × 10-16
2.5× 10-10
360
2.3 × 10-10
2.0 × 10-13
2.6 × 10-24
3.0 × 10-16
2.3× 10-10
380
2.2 × 10-10
1.5 × 10-13
7.2 × 10-24
2.4 × 10-16
2.2× 10-10
400
2.2 × 10-10
1.2 × 10-13
1.8 × 10-23
2.2 × 10-16
2.2× 10-10
Table S6 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of (CH3)2OO with HCOOH computed at different temperatures
T/K
k (TSent1-dim)
k (TSent2-dim)
k (TSent3-dim)
k (TSent4-dim)
ktot-dim
273
5.3 × 10-10
6.8 × 10-12
1.4 × 10-26
4.4 × 10-15
5.4 × 10-10
280
5.1 × 10-10
5.2 × 10-12
2.2 × 10-26
4.2 × 10-15
5.2 × 10-10
298
4.8 × 10-10
2.8 × 10-12
8.0 × 10-26
4.0 × 10-15
4.8 × 10-10
300
4.7 × 10-10
2.6 × 10-12
9.2 × 10-26
3.9 × 10-15
4.7 × 10-10
320
4.5 × 10-10
1.4 × 10-12
3.6 × 10-25
3.7 × 10-15
4.5 × 10-10
340
4.2 × 10-10
8.6 × 10-13
1.3 × 10-24
3.6 × 10-15
4.2 × 10-10
360
3.9 × 10-10
5.5 × 10-13
4.5 × 10-24
3.5 × 10-15
3.9 × 10-10
380
3.7 × 10-10
3.7 × 10-13
1.4 × 10-23
3.4 × 10-15
3.7 × 10-10
400
3.7 × 10-10
2.6 × 10-13
3.9 × 10-23
3.4 × 10-15
3.7 × 10-10
Corresponding descriptions have been added in the page 7 line 173-190, page 11 line 303-315, page 12 line 330-338 and page 13 line 346-351 of the revised manuscript:
The rate coefficients for the barrierless reactions are determined by employing the inverse Laplace transformation (ILT) method. The ILT calculations are performed with the MESMER 6.0 program (Glowacki et al., 2012). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken et al., 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients for the bimolecular reactions with the tight transition states are calculated by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction (Truhlar et al., 1996; Eckart, 1930). The CTST/Eckart calculations are performed with the KiSThelP 2019 program (Canneaux et al., 2013).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OOis estimated to be 1.4 × 10-10cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant.
Equivalent to the case of CH2OO reaction with HCOOH, the rate coefficient of each elementary pathway involved in the anti-CH3CHOO + HCOOH reaction also decreases with the temperature increasing (Table S4). This table shows that Entry 1 is kinetically favored over Entry 2, 3 and 4, and Entry 2 is competitive with Entry 4 in the range 273-400 K. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of syn-CH3CHOO and (CH3)2COO with HCOOH that Entry 1 is the dominant pathway (Table S5-S6). It deserves mentioning that the competition of Entry 2 is significantly greater than that of Entry 4 in the syn-CH3CHOO + HCOOH and (CH3)2COO + HCOOH systems. At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
- The trend in exothermicity with substitution pattern (p.8-9) should be explained.
Response: Based on the Reviewer’s suggestion, the relevance explanations on the trend in exothermicity have been added in the revised manuscript. The exothermicity of 1,4 O-H insertion reactions of distinct SCIs with HCOOH is assessed by the reaction enthalpy (), which is defined as the difference between the enthalpies of formation () of the products and reactants (). To the best of our knowledge, there are no literature values available on the enthalpies of formation of carbonyl oxides and hydroperoxide esters except the simplest carbonyl oxide CH2OO. Therefore, the isodesmic reaction method is adopted to obtain the enthalpies of formation, and the results are listed in Table S2. An isodesmic reaction is a hypothetical reaction, in which the type of chemical bonds in the reactants is the similar as that of chemical bonds in the products. The following isodesmic reaction is constructed because the experimental values of H2, CH4 and H2O are available ((H2) = 0.00 kcal·mol-1; (CH4) = -17.82 kcal·mol-1; (H2O) = -57.79 kcal·mol-1).
(4)
As seen in Table S2, the enthalpy of formation of CH2OO is calculated to be 23.23 kcal·mol-1, which is in good agreement with the available literature values (Karton et al., 2013; Chen et al., 2016). This result implies that the theoretical method employed herein is reasonable to predict the thermochemical parameters. The enthalpies of formation of carbonyl oxides and hydroperoxide esters significantly decrease with increasing the number of methyl groups. Notably, the decreased values in the enthalpies of formation of carbonyl oxides are greater than those of hydroperoxide esters under the condition of the same number of methyl groups. For example, the enthalpy of formation of anti-CH3CHOO decreases by 12.95 kcal·mol-1 compared to the enthalpy of formation of CH2OO, and the enthalpy of formation of Pent1b decreases by 12.12 kcal·mol-1 compared to the enthalpy of formation of Pent1a. The reaction enthalpies decrease in the order of -44.69 (CH2OO + HCOOH → Pent1a) < -43.86 (anti-CH3CHOO + HCOOH → Pent1b) < -38.13 (syn-CH3CHOO + HCOOH → Pent1c) < -37.12 kcal·mol-1 ((CH3)2COO + HCOOH → Pent1d), indicating that the reaction enthalpies are highly dependent on the number and location of methyl groups. The trend in reaction enthalpies is consistent with the trend in the enthalpies of formation of carbonyl oxides. The reason might be attributed to the decreased values in the enthalpies of formation of carbonyl oxides greater than those of hydroperoxide esters under the condition of the same number of methyl groups.
Table S2 Enthalpies of formation () for the various carbonyl oxides and hydroperoxide esters computed at the CCSD(T)//M06-2X/6-311+G(2df,2p) level of theory
Species
Cal (kcal·mol-1)
Refs. (kcal·mol-1)
CH2OO
23.23
22.92a
24.59b
anti-CH3OO
10.28
syn-CH3CHOO
6.73
(CH3)2COO
-6.77
HCOOH
-90.62 (exp)
HC(O)OCH2OOH (Pent1a)
-112.08
HC(O)OCH(CH3)OOH (Pent1b)
-124.20
HC(O)OCH(CH3)OOH (Pent1c)
-122.02
HC(O)OC(CH3)2OOH (Pent1d)
-134.51
Exp is taken from NIST Chemistry Webbook
a the value is obtained at the G4 level of theory (Chen et al., 2016)
b the value is obtained at the W3-F12 level of theory (Karton et al., 2013)
Corresponding descriptions have been added in the page 9 line 240-247 and page 10 line 248-271 of the revised manuscript:
The exothermicity of 1,4 O-H insertion reactions of distinct SCIs with HCOOH is assessed by the reaction enthalpy (), which is defined as the difference between the enthalpies of formation () of the products and reactants (). To the best of our knowledge, there are no literature values available on the enthalpies of formation of carbonyl oxides and hydroperoxide esters except the simplest carbonyl oxide CH2OO. Therefore, the isodesmic reaction method is adopted to obtain the enthalpies of formation, and the results are listed in Table S2. An isodesmic reaction is a hypothetical reaction, in which the type of chemical bonds in the reactants is the similar as that of chemical bonds in the products. The following isodesmic reaction is constructed because the experimental values of H2, CH4 and H2O are available ((H2) = 0.00 kcal·mol-1; (CH4) = -17.82 kcal·mol-1; (H2O) = -57.79 kcal·mol-1).
(4)
As seen in Table S2, the enthalpy of formation of CH2OO is calculated to be 23.23 kcal·mol-1, which is in good agreement with the available literature values (Chen et al., 2016; Karton et al., 2013). This result implies that the theoretical method employed herein is reasonable to predict the thermochemical parameters. The enthalpies of formation of carbonyl oxides and hydroperoxide esters significantly decrease with increasing the number of methyl groups. Notably, the decreased values in the enthalpies of formation of carbonyl oxides are greater than those of hydroperoxide esters under the condition of the same number of methyl groups. For example, the enthalpy of formation of anti-CH3CHOO decreases by 12.95 kcal·mol-1 compared to the enthalpy of formation of CH2OO, and the enthalpy of formation of Pent1b decreases by 12.12 kcal·mol-1 compared to the enthalpy of formation of Pent1a. The reaction enthalpies decrease in the order of -44.69 (CH2OO + HCOOH → Pent1a) < -43.86 (anti-CH3CHOO + HCOOH → Pent1b) < -38.13 (syn-CH3CHOO + HCOOH → Pent1c) < -37.12 kcal·mol-1 ((CH3)2COO + HCOOH → Pent1d), indicating that the reaction enthalpies are highly dependent on the number and location of methyl groups. The trend in reaction enthalpies is consistent with the trend in the enthalpies of formation of carbonyl oxides. The reason might be attributed to the decreased values in the enthalpies of formation of carbonyl oxides greater than those of hydroperoxide esters under the condition of the same number of methyl groups.
- The analysis of possible bimolecular CI reactions (p.21) should be extended to the three substituted CIs.
Response: Kalinowski et al. has comfirmed that the central CO bond of carbonyl oxides is a double bond, while the terminal OO bond is a single bond (Kalinowski et al., 2014). It is therefore that the maximum degree of substitution of carbonyl oxides is two. To further evaluate the relative importance of the complex SCIs reactions with coreactant, the bimolecular reactions of methyl vinyl ketone oxide (MVK-OO) with H2O, HCOOH, SO2 and HPMF have been considered in the revised manuscript. MVK-OO, formed with 21 to 23% yield from the ozonolysis of isoprene, is a four carbon, asymmetric, resonance-stabilized Criegee intermediate (Barber et al., 2018). MVK-OO has four conformers, syn-trans-, syn-cis-, anti-trans-, and anti-cis- as shown in Fig. S10. Herein, syn and anti refer to the orientation of the -CH3 group relative to the terminal oxygen of MVK-OO, whereas cis and trans refer to the orientation of the C8=C9 bond relative to the C1=O2 bond. According to the results shown in the Fig. S10, the lowest-energy conformer is syn-trans-MVK-OO, which is lower than syn-cis-, anti-trans-, and anti-cis-MVK-OO by 1.42, 2.43 and 2.69 kcal·mol-1, respectively. Therefore, the lowest-energy conformer syn-trans-MVK-OO is selected as the model compound to study its bimolecular reactions. As shown in Table 2, the rate coefficient of H2O reaction with syn-trans-MVK-OO is lower than with other SCIs by 2 to 3 orders of magnitude. The reason is likely to be that the existence of methyl and vinyl groups hinders the occurrence of bimolecular reaction with water vapour. Consequently, a fraction of syn-trans-MVK-OO may survive in the presence of water vapour and react with other species. keff(MVK-OO+H2O) is nearly identical to keff(MVK-OO+HCOOH), which is greater than keff(MVK-OO+SO2) and keff(MVK-OO+HPMF) when the concentration of HPMF is the same as that of HCOOH. keff(MVK-OO+H2O) and keff(MVK-OO+HCOOH) are greater than keff(MVK-OO+SO2), which, in turn, are greater than keff(MVK-OO+HPMF) when the concentration of HPMF is equal to that of SCIs. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
trans
cis
syn
0.00
1.42
anti
2.43
2.69
Figure S10. The optimized geometries and relative energies (kcal·mol-1) computed for the four conformers of MVK-oxide. Geometries are optimized at the M06-2X/6-311+g(2df,2p) level of theory. Single point energies are calculated at the CCSD(T)/6-311+g(2df,2p) level of theory.
Corresponding descriptions have been added in the page 24 line 611-619 and page 25 line 620-636 of the revised manuscript:
To further evaluate the relative importance of the complex SCIs reactions with coreactant, the bimolecular reactions of methyl vinyl ketone oxide (MVK-OO) with H2O, HCOOH, SO2, and HPMF are considered. MVK-OO, formed with 21 to 23% yield from the ozonolysis of isoprene, is a four carbon, asymmetric, resonance-stabilized Criegee intermediate (Barber et al., 2018). MVK-OO has four conformers, syn-trans-, syn-cis-, anti-trans-, and anti-cis- as shown in Fig. S10. Herein, syn and anti refer to the orientation of the –CH3 group relative to the terminal oxygen of MVK-OO, whereas cis and trans refer to the orientation of the C8=C9 bond relative to the C1=O2 bond. According to the results shown in the Fig. S10, the lowest-energy conformer is syn-trans-MVK-OO, which is lower than syn-cis-, anti-trans-, and anti-cis-MVK-OO by 1.42, 2.43 and 2.69 kcal·mol-1, respectively. Therefore, the lowest-energy conformer syn-trans-MVK-OO is selected as the model compound to study its bimolecular reactions. As shown in Table 2, the rate coefficient of H2O reaction with syn-trans-MVK-OO is lower than with other SCIs by 2 to 3 orders of magnitude. The reason is likely to be that the existence of methyl and vinyl groups hinders the occurrence of bimolecular reaction with water vapour. Consequently, a fraction of syn-trans-MVK-OO may survive in the presence of water vapour and react with other species. keff(MVK-OO+H2O) is nearly identical to keff(MVK-OO+HCOOH), which is greater than keff(MVK-OO+SO2) and keff(MVK-OO+HPMF) when the concentration of HPMF is the same as that of HCOOH. keff(MVK-OO+H2O) and keff(MVK-OO+HCOOH) are greater than keff(MVK-OO+SO2), which, in turn, are greater than keff(MVK-OO+HPMF) when the concentration of HPMF is equal to that of SCIs. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
- Since the CI is clearly the limiting reactant in the CI + HCOOH reaction, the atmospheric concentration of HPMF (and the other hydroperoxy esters) is much better estimated to be the CI concentration. (This, of course, will greatly lower the predicted pseudo-first-order rate constants for the CI + HPMF reaction.)
Response: Based on the Reviewer’s suggestion, the relevance explanations on the predicted pseudo-first-order rate constants have been added in the revised manuscript. It is of interest to assess whether the reactions of distinct SCIs with HPMF can compete well with the losses to reactions with trace species (e.g., H2O, HCOOH and SO2), because it is well known that the reactions with trace species are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2 and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmospheric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken, 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al., 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF) is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105 molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF) is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters.
Table 2 The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with HPMF, H2O, HCOOH and SO2 at the tropical forest environments
SCIs
Coreactant
[Coreactant]
(molecules cm-3)
k
(cm3 molecule-1 s-1)
keff
(s-1)
Reference
CH2OO
H2O
3.9-6.1 × 1017
< 1.5 × 10-15
5.9-9.2 × 102
Chao et al., (2015)
HCOOH
5.0-10.0 × 1010
[1.1 ± 0.1] × 10-10
5.5-11
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[3.9 ± 0.7] × 10-11
0.7-3.5
Welz et al., (2012)
HPMF
-
2.7 × 10-11
-
This work
anti-CH3CHOO
H2O
3.9-6.1 × 1017
[1.0 ± 0.4] × 10-14
3.9-6.1 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[5 ± 3] × 10-10
25.0-50.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[6.7 ± 1.0] × 10-11
1.1-6.0
Taatjes et al., (2013)
HPMF
-
3.3 × 10-10
-
This work
syn-CH3CHOO
H2O
3.9-6.1 × 1017
< 4.0 × 10-15
1.6-2.4 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[2.5 ± 0.3] × 10-10
12.5-25.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[2.4 ± 0.3] × 10-11
0.4-2.2
Taatjes et al., (2013)
HPMF
-
1.7 × 10-13
-
This work
(CH3)2COO
H2O
3.9-6.1 × 1017
< 1.5 × 10-16
58.5-91.5
Huang et al., (2015)
HCOOH
5.0-10.0 × 1010
4.5 × 10-10
22.5-45.0
Sipilä et al., (2014)
SO2
1.7-9.0 × 1010
1.3 × 10-10
2.2-11.7
Huang et al., (2015)
HPMF
-
2.2 × 10-11
-
This work
syn-trans-MVK-OO
H2O
3.9-6.1 × 1017
< 4.0 × 10-17
15.6-24.4
Caravan et al., (2020)
HCOOH
5.0-10.0 × 1010
[3.0 ± 0.1] × 10-10
15.0-30.0
Caravan et al., (2020)
SO2
1.7-9.0 × 1010
[4.2 ± 0.6] × 10-11
0.7-3.8
Caravan et al., (2020)
HPMF
-
3.0 × 10-11
-
This work
Corresponding descriptions have been added in the page 23 line 573-590 and page 24 line 591-610 of the revised manuscript:
It is of interest to assess whether the reactions of distinct SCIs with HPMF can compete well with the losses to reactions with trace species (e.g., H2O, HCOOH and SO2), because it is well known that the reactions with trace species are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2, and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmospheric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken, 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al., 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF) is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105 molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF) is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters.
- Since a big motivation for the computations is the potential for CI + hydroperoxy ester reactions to lead to SOA, there should be some specific discussion, perhaps buttressed by rough calculations, of how many cycles of CI addition are required before a given adduct is expected to have low volatility. The approach of Chhantyal-Pun et al. (ACS Earth Space Chem. 2018, 2, 8, 833-842) is an example of the approach the authors should take.
Response: Based on the Reviewer’s suggestion, the vapour pressure and volatility of adduct products formed from the successive reactions of SCIs with hydroperoxide esters have been added in the revised manuscript. The assessment of Barley and McFiggans (2010) and O'Meara et al. (2014) found that the combination of boiling point estimation from Nannoolal et al. (2004) and vapour pressure estimation from Nannoolal et al. (2008) gives the lowest mean bias error of vapour pressure for atmospherically relevant compounds. Therefore, the saturated vapour pressure (P0) of adduct products at room temperature is estimated by using the Nannoolal-Nannoolal method, and the results are listed in Table S10.
From Table S10, it can be seen that the P0 of adduct products involved in the successive reactions of CH2OO with HCOOH increases first and then decreases with increasing the number of CH2OO. The P0 of the adduct product HC(O)O(CH2OO)3H is maximum when the number of CH2OO is equal to three. The P0 of adduct products included in the successive reactions of anti-CH3CHOO with HCOOH decreases significantly as the number of anti-CH3CHOO is increased. Similar phenomenon is also observed from the successive reactions of syn-CH3CHOO and (CH3)2COO with HCOOH. Notably, the P0 of adduct products decreases obviously when the size of SCIs increases. For example, the P0 of the adduct product HC(O)O(CH2OO)3H in the nCH2OO + HCOOH reaction is estimated to be 4.43 × 10-3 atm, which is greater than those of the corresponding adduct products in the nanti-CH3CHOO + HCOOH (7.12 × 10-4), nsyn-CH3CHOO + HCOOH (7.12 × 10-4), and n(CH3)2COO + HCOOH (1.27 × 10-4) reactions by 6.22, 6.22 and 34.88 times, respectively.
A classify scheme of various organic compounds is based on their volatility, as presented by Donahue et al. (2012) The volatility of organic compounds is described by their effective saturation concentration. The saturated concentrations (c0) of adduct products formed from the successive reactions of SCIs with HCOOH are predicted by using the SIMPOL.1 method proposed by Pankow and Asher (2008), and the results are listed in Table S10. As shown in Table S10, the c0 of adduct products involved in the nCH2OO + HCOOH reaction decreases with increasing the number of CH2OO. According to the Volatility Basis Set (VBS) of organic compounds (Donahue et al., 2012), these adduct products belong to volatile organic compounds (VOC, c0 > 3 × 106 ug/m3). Similarly, the c0 of adduct products included in the nanti-CH3CHOO + HCOOH, nsyn-CH3CHOO + HCOOH, and n(CH3)2COO + HCOOH reactions decreases when the number of SCIs increases. It deserves mentioning that the adduct products in the nanti-CH3CHOO + HCOOH and nsyn-CH3CHOO + HCOOH reactions belong to intermediate volatility organic compounds (IVOC, 300 < c0 < 3 × 106 ug/m3) when the number of SCIs is equal to five. However, the adduct products in the n(CH3)2COO + HCOOH reaction become IVOC when the number of (CH3)2COO is greater than or equal to two. Based on the above discussions, it can be concluded that the volatility of adduct products is significantly affected by the number and size of SCIs in the successive reaction of SCIs with HCOOH.
Table S10 Predicted saturated vapour pressure (P0) and saturated concentrations (c0) for the adduct products of the successive reactions of SCIs with HCOOH
formula
P0 (atm)
c0 (ug/m3)
n CH2OO + HCOOH
n = 1
HC(O)OCH2OOH
2.12 × 10-3
7.86 × 107
n = 2
HC(O)O(CH2OO)2H
3.80 × 10-3
3.99 × 107
n = 3
HC(O)O(CH2OO)3H
4.43 × 10-3
3.91 × 107
n = 4
HC(O)O(CH2OO)4H
4.21 × 10-3
3.29 × 107
n = 5
HC(O)O(CH2OO)5H
3.59 × 10-3
2.12 × 107
n anti-CH3CHOO + HCOOH
n = 1
HC(O)OCH(CH3)OOH
1.25 × 10-3
8.32 × 106
n = 2
HC(O)O(CH(CH3)OO)2H
1.13 × 10-3
7.57 × 106
n = 3
HC(O)O(CH(CH3)OO)3H
7.12 × 10-4
6.49 × 106
n = 4
HC(O)O(CH(CH3)OO)4H
3.90 × 10-4
4.50 × 106
n = 5
HC(O)O(CH(CH3)OO)5H
2.01 × 10-4
2.81 × 106
n syn-CH3CHOO + HCOOH
n = 1
HC(O)OCH(CH3)OOH
1.25 × 10-3
8.32 × 106
n = 2
HC(O)O(CH(CH3)OO)2H
1.13 × 10-3
7.57 × 106
n = 3
HC(O)O(CH(CH3)OO)3H
7.12 × 10-4
6.49 × 106
n = 4
HC(O)O(CH(CH3)OO)4H
3.90 × 10-4
4.50 × 106
n = 5
HC(O)O(CH(CH3)OO)5H
2.01 × 10-4
2.81 × 106
n (CH3)2COO + HCOOH
n = 1
HC(O)OC(CH3)2OOH
7.23 × 10-4
3.50 × 106
n = 2
HC(O)O(C(CH3)2OO)2H
3.50 × 10-4
2.74 × 106
n = 3
HC(O)O(C(CH3)2OO)3H
1.27 × 10-4
1.38 × 106
n = 4
HC(O)O(C(CH3)2OO)4H
4.27 × 10-5
5.90 × 105
n = 5
HC(O)O(C(CH3)2OO)5H
1.40 × 10-5
2.36 × 105
Corresponding descriptions have been added in the page 27 line 644-671 and page 28 line 672-682 of the revised manuscript:
The assessment of Barley and McFiggans (2010) and O'Meara et al. (2014) found that the combination of boiling point estimation from Nannoolal et al. (2004) and vapour pressure estimation from Nannoolal et al. (2008) gives the lowest mean bias error of vapour pressure for atmospherically relevant compounds. Therefore, the saturated vapour pressure (P0) of adduct products at room temperature is estimated by using the Nannoolal-Nannoolal method, and the results are listed in Table S10. From Table S10, it can be seen that the P0 of adduct products involved in the successive reactions of CH2OO with HCOOH increases first and then decreases with increasing the number of CH2OO. The P0 of the adduct product HC(O)O(CH2OO)3H is maximum when the number of CH2OO is equal to three. The P0 of adduct products included in the successive reactions of anti-CH3CHOO with HCOOH decreases significantly as the number of anti-CH3CHOO is increased. Similar phenomenon is also observed from the successive reactions of syn-CH3CHOO and (CH3)2COO with HCOOH. Notably, the P0 of adduct products decreases obviously when the size of SCIs increases. For example, the P0 of the adduct product HC(O)O(CH2OO)3H in the nCH2OO + HCOOH reaction is estimated to be 4.43 × 10-3 atm, which is greater than those of the corresponding adduct products in the nanti-CH3CHOO + HCOOH (7.12 × 10-4), nsyn-CH3CHOO + HCOOH (7.12 × 10-4), and n(CH3)2COO + HCOOH (1.27 × 10-4) reactions by 6.22, 6.22 and 34.88 times, respectively.
A classify scheme of various organic compounds is based on their volatility, as presented by Donahue et al. (2012) The volatility of organic compounds is described by their effective saturation concentration. The saturated concentrations (c0) of adduct products formed from the successive reactions of SCIs with HCOOH are predicted by using the SIMPOL.1 method proposed by Pankow and Asher (2008), and the results are listed in Table S10. As shown in Table S10, the c0 of adduct products involved in the nCH2OO + HCOOH reaction decreases with increasing the number of CH2OO. According to the Volatility Basis Set (VBS) of organic compounds (Donahue et al., 2012), these adduct products belong to VOC (c0 > 3 × 106 ug/m3). Similarly, the c0 of adduct products included in the nanti-CH3CHOO + HCOOH, nsyn-CH3CHOO + HCOOH, and n(CH3)2COO + HCOOH reactions decreases when the number of SCIs increases. It deserves mentioning that the adduct products in the nanti-CH3CHOO + HCOOH and nsyn-CH3CHOO + HCOOH reactions belong to intermediate volatility organic compounds (IVOC, 300 < c0 < 3 × 106 ug/m3) when the number of SCIs is equal to five. However, the adduct products in the n(CH3)2COO + HCOOH reaction become IVOC when the number of (CH3)2COO is greater than or equal to two. Based on the above discussions, it can be concluded that the volatility of adduct products is significantly affected by the number and size of SCIs in the successive reaction of SCIs with HCOOH.
- On p. 6, line 145: "saddle point" should be "minimum".
Response: The word “saddle point” has been replaced by “minimum” in the revised manuscript.
- On p. 6, line 162: "precision" should be "accuracy".
Response: The word “precision” has been replaced by “accuracy” in the revised manuscript.
- On p. 7, line 182: "decomposes" should be "rearranges".
Response: The word “decomposes” has been replaced by “rearranges” in the revised manuscript.
- On p. 14, lines 341-342, use a non-breaking hyphen.
Response: A non-breaking hyphen has been used in the revised manuscript.
- On p. 15, line 372, "intermolecular" should be "intramolecular".
Response: The word “intermolecular” has been replaced by “intramolecular” in the revised manuscript.
- On p. 17, lines 413-414, use a non-breaking hyphen.
Response: A non-breaking hyphen has been used in the revised manuscript.
References
Barber, V. P., Pandit, S., Green, A. M., Trongsiriwat, N., Walsh, P. J., Klippenstein, S. J., and Lester, M. I.: Four-carbon Criegee intermediate from isoprene ozonolysis: methyl vinyl ketone oxide synthesis, infrared spectrum, and OH production, J. Am. Chem. Soc., 140, 10866-10880, https://doi.org/10.1021/jacs.8b06010, 2018.
Barley, M. H., and McFiggans, G.: The critical assessment of vapour pressure estimation methods for use in modelling the formation of atmospheric organic aerosol, Atmos. Chem. Phys., 10, 749-767, https://doi.org/10.5194/acp-10-749-2010, 2010.
Canneaux, S., Bohr, F., and Henon, E.: KiSThelP: a program to predict thermodynamic properties and rate constants from quantum chemistry results, J. Comput. Chem., 35, 82-93, https://doi.org/10.1002/jcc.23470, 2013.
Caravan, R. L., Vansco, M. F., Au, K., Khan, M. A. H., Li, Y. L., Winiberg, F. A. F., Zuraski, K., Lin, Y. H., Chao, W., Trongsiriwat, N., Walsh, P. J., Osborn, D. L., Percival, C. J., Lin, J. J. M., Shallcross, D. E., Sheps, L., Klippenstein, S. J., Taatjes, C. A., and Lester, M. I.: Direct kinetic measurements and theoretical predictions of an isoprene-derived Criegee intermediate, Proc. Natl. Acad. Sci. U.S.A., 117, 9733-9740, https://doi.org/10.1073/pnas.1916711117, 2020.
Chao, W., Hsieh, J. T., Chang, C. H., and Lin, J. J. M.: Direct kinetic measurement of the reaction of the simplest Criegee intermediate with water vapor, Science, 347, 751-754, https://doi.org/10.1126/science.1261549, 2015.
Chen, L., Wang, W., Wang, W., Liu, Y., Liu, F., Liu, N., and Wang, B: Water‑catalyzed decomposition of the simplest Criegee intermediate CH2OO, Theor. Chem. Acc., 135, 131-143, https://doi.org/10.1007/s00214-016-1894-9, 2016.
Chung, C. A., Su, J. W., and Lee, Y. P.: Detailed mechanism and kinetics of the reaction of Criegee intermediate CH2OO with HCOOH investigated via infrared identification of conformers of hydroperoxymethyl formate and formic acid anhydride, Phys. Chem. Chem. Phys., 21, 21445-21455, https://doi.org/10.1039/c9cp04168k, 2019.
Donahue, N. M., Kroll, J. H., Pandis, S. N., and Robinson, A. L.: A two-dimensional volatility basis set – Part 2: Diagnostics of organic-aerosol evolution, Atmos. Chem. Phys., 12, 615-634, https://doi.org/10.5194/acp-12-615-2012, 2012.
Gilbert, R. G., and Smith, S. C.: Theory of unimolecular and recombination reactions; Blackwell Scientific: Carlton, Australia, 1990.
Glowacki, D. R., Liang, C. H., Morley, C., Pilling, M. J., and Robertson, S. H.: MESMER: an open-source master equation solver for multi-energy well reactions, J. Phys. Chem. A, 116, 9545-9560, https://doi.org/10.1021/jp3051033, 2012.
Huang, H. L., Chao, W., and Lin, J. J. M.: Kinetics of a Criegee intermediate that would survive high humidity and may oxidize atmospheric SO2, Proc. Natl. Acad. Sci. U.S.A., 112, 10857-10862, https://doi.org/ 10.1073/pnas.1513149112, 2015.
Kalinowski, J., Räsänen, M., Heinonen, P., Kilpeläinen, I., and Gerber, R. B.: Isomerization and decomposition of a Criegee intermediate in the ozonolysis of alkenes: dynamics using a multireference potential, Angew. Chem., 126, 269-272, https://doi.org/10.1002/ange.201307286, 2014.
Karton, A., Kettner, M., and Wild, D. A.: Sneaking up on the Criegee intermediate from below: Predicted photoelectron spectrum of the CH2OO- anion and W3-F12 electron affinity of CH2OO, Chem. Phys. Lett., 585, 15-20, http://doi.org/10.1016/j.cplett.2013.08.075, 2013.
Khan, M. A. H., Percival, C. J., Caravan, R. L., Taatjes, C. A., and Shallcross, D. E.: Criegee intermediates and their impacts on the troposphere, Environ. Sci.: Processes Impacts, 20, 437-453, https://doi.org/10.1039/C7EM00585G, 2018.
Long, B., Bao, J. L., and Truhlar, D. G.: Atmospheric chemistry of Criegee intermediates: unimolecular reactions and reactions with water, J. Am. Chem. Soc., 138, 14409-14422, https://doi.org/10.1021/jacs.6b08655, 2016.
Nannoolal, Y., Rarey, J., and Ramjugernatha, D.: Estimation of pure component properties Part 3. Estimation of the vapor pressure of non-electrolyte organic compounds via group contributions and group interactions, Fluid Phase Equilibria, 269, 117-133, https://doi.org/10.1016/j.fluid.2008.04.020, 2008.
Nannoolal, Y., Rarey, J., Ramjugernatha, D., and Cordesb, W.: Estimation of pure component properties Part 1. Estimation of the normal boiling point of non-electrolyte organic compounds via group contributions and group interactions, Fluid Phase Equilibria, 226, 45-63, https://doi.org/10.1016/j.fluid.2004.09.001, 2004.
Novelli, A., Hens, K., Ernest, C. T., Martinez, M., Nölscher, A. C., Sinha, V., Paasonen, P., Petäjä, T., Sipilä, M., Elste, T., Plass-Dülmer, C., Phillips, G. J., Kubistin, D., Williams, J., Vereecken, L., Lelieveld, J., and Harder, H.: Estimating the atmospheric concentration of Criegee intermediates and their possible interference in a FAGE-LIF instrument, Atmos. Chem. Phys., 17, 7807-7826, https://doi.org/10.5194/acp-17-7807-2017, 2017.
O'Meara, S., Booth, A. M., Barley, M. H., Topping, D., and McFiggans, G.: An assessment of vapour pressure estimation methods, Phys. Chem. Chem. Phys., 16, 19453-19469, https://doi.org/10.1039/C4CP00857J, 2014.
Pankow, J. F., and Asher, W. E.: SIMPOL.1: a simple group contribution method for predicting vapor pressures and enthalpies of vaporization of multifunctional organic compounds, Atmos. Chem. Phys., 8, 2773-2796, https://doi.org/10.5194/acp-8-2773-2008, 2008.
Peltola, J., Seal, P., Inkilä, A., and Eskola, A.: Time-resolved, broadband UV-absorption spectrometry measurements of Criegee intermediate kinetics using a new photolytic precursor: unimolecular decomposition of CH2OO and its reaction with formic acid, Phys. Chem. Chem. Phys., 22, 11797-11808, https://doi.org/10.1039/d0cp00302f, 2020.
Raghunath, P., Lee, Y. P., and Lin, M. C.: Computational chemical kinetics for the reaction of Criegee intermediate CH2OO with HNO3 and its catalytic conversion to OH and HCO, J. Phys. Chem. A, 121, 3871-3878, https://doi.org/10.1021/acs.jpca.7b02196, 2017.
Sipilä, M., Jokinen, T., Berndt, T., Richters, S., Makkonen, R., Donahue, N. M., Mauldin Iii, R. L., Kurtén, T., Paasonen, P., Sarnela, N., Ehn, M., Junninen, H., Rissanen, M. P., Thornton, J., Stratmann, F., Herrmann, H., Worsnop, D. R., Kulmala, M., Kerminen, V. M., and Petäjä, T.: Reactivity of stabilized Criegee intermediates (sCIs) from isoprene and monoterpene ozonolysis toward SO2 and organic acids, Atmos. Chem. Phys., 14, 12143-12153, https://doi.org/10.5194/acp-14-12143-2014, 2014.
Taatjes, C. A., Welz, O., Eskola, A. J., Savee, J. D., Scheer, A. M., Shallcross, D. E., Rotavera, B., Lee, E. P. F., Dyke, J. M., Mok, D. K. W., Osborn, D. L., and Percival, C. J.: Direct measurements of conformer-dependent reactivity of the Criegee intermediate CH3CHOO, Science, 340, 177-180, https://doi.org/10.1126/science.1234689, 2013.
Vereecken, L., Harder, H., and Novelli, A.: The reaction of Criegee intermediates with NO, RO2, and SO2, and their fate in the atmosphere, Phys. Chem. Chem. Phys., 14, 14682-14695, https://doi.org/10.1039/c2cp42300f, 2012.
Welz, O., Eskola, A. J., Sheps, L., Rotavera, B., Savee, J. D., Scheer, A. M., Osborn, D. L., Lowe, D., Booth, A. M., Xiao, P., Khan, M. A. H., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Rate coefficients of C(1) and C(2) Criegee intermediate reactions with formic and acetic Acid near the collision limit: direct kinetics measurements and atmospheric implications, Angew. Chem. Int. Ed., 53, 4547-4550, https://doi.org/10.1002/anie.201400964, 2014.
Welz, O., Savee, J. D., Osborn, D. L., Vasu, S. S., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Direct kinetic measurements of Criegee intermediate (CH2OO) formed by reaction of CH2I with O2, Science, 335, 204-207, https://doi.org/10.1126/science.1213229, 2012.
-
RC2: 'Comment on acp-2022-376', Anonymous Referee #2, 04 Jul 2022
This paper has investigated the oligomerization mechanisms and kinetics of distinct stabilized Criegee intermediates with HCOOH and their products using quantum chemical and kinetics modeling methods. Also, the effect of methyl groups on the oligomerization were discussed.
However, a deeper discussion is required for the data in this paper. For example, in lines 263-266 “At room temperature, ktot is estimated to be 3.6 × 10-10 cm3 molecule-1 s-1, which is greater by a factor of ~3 than that reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10 cm3 molecule-1 s-1), Chung et al. (2019) ([1.4 ± 0.3] × 10-10 cm3 molecule-1 s-1), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10 cm3 molecule-1 s-1)”. What is the reason for the difference of the k value about three times?
Furthermore, this paper should also exhibit some extended discussions about atmospheric implications of these reactions and their products. For example, what is the role of the formed oligomers on the atmosphere? It follows in the requirements of ACP journal “The journal scope is focused on studies with important implications for our understanding of the state and behavior of the atmosphere. Articles with a local focus must clearly explain how the results extend and compare with current knowledge”.
Hence, as a quick assessment, some deeper and extended discussions should be required and strengthened, such as the nature of the reactions, the detailed atmospheric implications, if this paper is published in the ACP journal.
Citation: https://doi.org/10.5194/acp-2022-376-RC2 -
AC2: 'Reply on RC2', Y. Huang, 25 Aug 2022
Prof. Yu Huang
State Key Lab of Loess and Quaternary Geology
Institute of Earth Environment, Chinese Academy of Sciences, Xi’an, 710061, China
Tel./Fax: (86) 29-62336261
E-mail: huangyu@ieecas.cn
Aug. 25, 2022
Dear Prof. Kourtchev,
Revision for Manuscript ACP-2022-376
We thank you very much for giving us the opportunity to revise our manuscript. We highly appreciate the reviewers for their comments and suggestions on the manuscript entitled “Oligomer formation from the gas-phase reactions of Criegee intermediates with hydroperoxide esters: mechanism and kinetics”. We have made revisions of our manuscript carefully according to the comments and suggestions of reviewers. The revised contents are marked in blue color. The response letter to reviewers is attached at the end of this cover letter.
We hope that the revised manuscript can meet the requirement of Atmospheric Chemistry & Physics. Any further modifications or revisions, please do not hesitate to contact us.
Look forward to hearing from you as soon as possible.
Best regards,
Yu Huang
Comments of reviewer #2
- However, a deeper discussion is required for the data in this paper. For example, in lines 263-266 “At room temperature, ktot is estimated to be 3.6 × 10-10 cm3 molecule-1 s-1, which is greater by a factor of ~3 than that reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10 cm3 molecule-1 s-1), Chung et al. (2019) ([1.4 ± 0.3] × 10-10 cm3 molecule-1 s-1), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10 cm3 molecule-1 s-1)”. What is the reason for the difference of the k value about three times?
Response: In the original manuscript, the rate coefficients for the barrierless reactions are calculated by employing the variational transition state theory (VTST), and the rate coefficients for the bimolecular reactions with the tight transition states are computed by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction. For the initiation reactions of distinct stabilized Criegee intermediates (SCIs) with HCOOH, there are four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4), in which Entry 1 is barrierless and Entry 2-4 have the tight transition states. The total rate coefficient for the reaction of SCIs with HCOOH is equal to the sum of the rate coefficient of each pathway. For the barrierless 1,4 O-H insertion reaction, the VTST is approximated with a Morse potential function, V(R) = De{1-exp[-β(R-Re)]}2, along with an anisotropy potential function to stand for the minimum energy path, which is used to calculate the rate coefficients (Raghunath et al., 2017). Here, De is the bond energy excluding the zero-point energy, R is the reaction coordinate, and Re is the equilibrium value of R. It is assumed that the stretching potential in an anisotropy potential is used in conjunction with a potential form of Vanisotropy = V0[1-cos2(θ1–θ1e) × cos2(θ2–θ2e)] (Raghunath et al., 2017). Here, V0 is the stretching potential, which stands for by a Morse potential, θ1 and θ1e represent the rotational angle between fragment 1 and the reference axis and the equilibrium bond angle of fragment 1, θ2 and θ2e stand for the rotational angle between fragment 2 and the reference axis and the equilibrium bond angle of fragment 2. The association curve for the reaction of 1,4 O-H insertion of SCIs into HCOOH is computed at the M06-2X/6-311+G(2df,2p) level of theory to cover a range from 0.97 to 1.97 Å at step size 0.1 Å for O-H bond and from 1.44 to 2.44 Å at step size 0.1 Å for C-O bond, while other structural parameters are fully optimized. The computed potential energies are fitted to the Morse potential function. However, the calculated rate coefficients for the reactions of SCIs with HCOOH are higher than the prior experimental measurements. The reason is ascribed to the fact that the approximation of VTST using a Morse potential function in conjunction with an anisotropy potential function is unsuitable to predict the rate coefficients for the barrierless 1,4 O-H insertion reaction.
In the revised manuscript, the rate coefficients for the barrierless reactions are computed by employing the inverse Laplace transformation (ILT) method, and the rate coefficients for the bimolecular reactions with the tight transition states are calculated by utilizing CTST in conjunction with Eckart tunneling correction. The ILT and CTST/Eckart calculations are performed by using the MESMER 6.0 and KiSThelP 2019 programs, respectively (Glowacki et al., 2012; Canneaux et al., 2013). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken, 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OO is estimated to be 1.4 × 10-10 cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of HCOOH with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO (Table S4-S6). At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
Table S3 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of CH2OO with HCOOH computed at different temperatures
T/K
k (TSent1)
k (TSent2)
k (TSent3)
k (TSent4)
ktot-CH2OO
273
1.7 × 10-10
3.6 × 10-12
1.0 × 10-22
3.6 × 10-12
1.8 × 10-10
280
1.6 × 10-10
2.9 × 10-12
1.2 × 10-22
3.1 × 10-12
1.7 × 10-10
298
1.4 × 10-10
1.9 × 10-12
2.2 × 10-22
2.3 × 10-12
1.4 × 10-10
300
1.4 × 10-10
1.8 × 10-12
2.4 × 10-22
2.2 × 10-12
1.4 × 10-10
320
1.3 × 10-10
1.2 × 10-12
4.9 × 10-22
1.6 × 10-12
1.3 × 10-10
340
1.3 × 10-10
8.2 × 10-13
1.0 × 10-21
1.3 × 10-12
1.3 × 10-10
360
1.2 × 10-10
5.9 × 10-13
2.2 × 10-21
1.0 × 10-12
1.2 × 10-10
380
1.2 × 10-10
4.5 × 10-13
4.5 × 10-21
8.2 × 10-13
1.2 × 10-10
400
1.2 × 10-10
3.5 × 10-13
9.0 × 10-21
6.9 × 10-13
1.2 × 10-10
Table S4 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of anti-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-anti)
k (TSent2-anti)
k (TSent3-anti)
k (TSent4-anti)
ktot-anti
273
5.9 × 10-10
4.2 × 10-11
5.5 × 10-22
6.1 × 10-11
6.9 × 10-10
280
5.7 × 10-10
3.8 × 10-11
6.7 × 10-22
4.9 × 10-11
6.6 × 10-10
298
5.4 × 10-10
2.3 × 10-11
1.2 × 10-21
3.0 × 10-11
5.9 × 10-10
300
5.3 × 10-10
2.0 × 10-11
1.3 × 10-21
2.8 × 10-11
5.8 × 10-10
320
5.0 × 10-10
1.5 × 10-11
2.6 × 10-21
1.7 × 10-11
5.3 × 10-10
340
4.7 × 10-10
9.4 × 10-12
5.4 × 10-21
1.1 × 10-11
4.9 × 10-10
360
4.5 × 10-10
7.0 × 10-12
1.1 × 10-20
7.8 × 10-12
4.7 × 10-10
380
4.4 × 10-10
3.6 × 10-12
2.1 × 10-20
5.6 × 10-12
4.5 × 10-10
400
4.3 × 10-10
2.0 × 10-12
4.0 × 10-20
4.2 × 10-12
4.4 × 10-10
Table S5 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of syn-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-syn)
k (TSent2-syn)
k (TSent3-syn)
k (TSent4-syn)
ktot-syn
273
3.1 × 10-10
9.5 × 10-13
4.6 × 10-27
7.5 × 10-16
3.1× 10-10
280
2.8 × 10-10
8.0 × 10-13
7.1 × 10-27
6.4 × 10-16
2.8× 10-10
298
2.7 × 10-10
5.4 × 10-13
8.9 × 10-26
5.5 × 10-16
2.7× 10-10
300
2.7 × 10-10
5.2 × 10-13
9.9 × 10-26
4.6 × 10-16
2.7× 10-10
320
2.5 × 10-10
3.6 × 10-13
3.0 × 10-25
3.8 × 10-16
2.5× 10-10
340
2.5 × 10-10
2.6 × 10-13
9.1 × 10-25
3.1 × 10-16
2.5× 10-10
360
2.3 × 10-10
2.0 × 10-13
2.6 × 10-24
3.0 × 10-16
2.3× 10-10
380
2.2 × 10-10
1.5 × 10-13
7.2 × 10-24
2.4 × 10-16
2.2× 10-10
400
2.2 × 10-10
1.2 × 10-13
1.8 × 10-23
2.2 × 10-16
2.2× 10-10
Table S6 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of (CH3)2OO with HCOOH computed at different temperatures
T/K
k (TSent1-dim)
k (TSent2-dim)
k (TSent3-dim)
k (TSent4-dim)
ktot-dim
273
5.3 × 10-10
6.8 × 10-12
1.4 × 10-26
4.4 × 10-15
5.4 × 10-10
280
5.1 × 10-10
5.2 × 10-12
2.2 × 10-26
4.2 × 10-15
5.2 × 10-10
298
4.8 × 10-10
2.8 × 10-12
8.0 × 10-26
4.0 × 10-15
4.8 × 10-10
300
4.7 × 10-10
2.6 × 10-12
9.2 × 10-26
3.9 × 10-15
4.7 × 10-10
320
4.5 × 10-10
1.4 × 10-12
3.6 × 10-25
3.7 × 10-15
4.5 × 10-10
340
4.2 × 10-10
8.6 × 10-13
1.3 × 10-24
3.6 × 10-15
4.2 × 10-10
360
3.9 × 10-10
5.5 × 10-13
4.5 × 10-24
3.5 × 10-15
3.9 × 10-10
380
3.7 × 10-10
3.7 × 10-13
1.4 × 10-23
3.4 × 10-15
3.7 × 10-10
400
3.7 × 10-10
2.6 × 10-13
3.9 × 10-23
3.4 × 10-15
3.7 × 10-10
Corresponding descriptions have been added in the page 7 line 173-190, page 11 line 303-315, page 12 line 330-338 and page 13 line 346-351 of the revised manuscript:
The rate coefficients for the barrierless reactions are determined by employing the inverse Laplace transformation (ILT) method. The ILT calculations are performed with the MESMER 6.0 program (Glowacki et al., 2012). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken et al., 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients for the bimolecular reactions with the tight transition states are calculated by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction (Truhlar et al., 1996; Eckart, 1930). The CTST/Eckart calculations are performed with the KiSThelP 2019 program (Canneaux et al., 2013).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OOis estimated to be 1.4 × 10-10cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant.
Equivalent to the case of CH2OO reaction with HCOOH, the rate coefficient of each elementary pathway involved in the anti-CH3CHOO + HCOOH reaction also decreases with the temperature increasing (Table S4). This table shows that Entry 1 is kinetically favored over Entry 2, 3 and 4, and Entry 2 is competitive with Entry 4 in the range 273-400 K. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of syn-CH3CHOO and (CH3)2COO with HCOOH that Entry 1 is the dominant pathway (Table S5-S6). It deserves mentioning that the competition of Entry 2 is significantly greater than that of Entry 4 in the syn-CH3CHOO + HCOOH and (CH3)2COO + HCOOH systems. At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
- Furthermore, this paper should also exhibit some extended discussions about atmospheric implications of these reactions and their products. For example, what is the role of the formed oligomers on the atmosphere? It follows in the requirements of ACP journal “The journal scope is focused on studies with important implications for our understanding of the state and behavior of the atmosphere. Articles with a local focus must clearly explain how the results extend and compare with current knowledge”.
Response: Based on the Reviewer’s suggestion, the atmospheric implication of the reactions of SCIs with hydroperoxide esters and the role of the formed oligomers have been added in the revised manuscript. It is well known that the reactions with trace species (e.g., H2O, HCOOH and SO2) are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). The relative importance of distinct SCIs reactions with hydroperoxide esters and trace species is taken into account. In the present study, the hydroperoxymethyl formate (HPMF) is selected as the model compound since it is the simplest hydroperoxide ester formed from the barrierless reaction of 1,4 O-H insertion of CH2OO into HCOOH. The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2, and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmospheric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken, 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al., 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF) is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105 molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF) is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
The vapour pressure and volatility of the formed oligomers are estimated in the revised manuscript. The assessment of Barley and McFiggans (2010) and O'Meara et al. (2014) found that the combination of boiling point estimation from Nannoolal et al. (2004) and vapour pressure estimation from Nannoolal et al. (2008) gives the lowest mean bias error of vapour pressure for atmospherically relevant compounds. Therefore, the saturated vapour pressure (P0) of adduct products at room temperature is estimated by using the Nannoolal-Nannoolal method, and the results are listed in Table S10. From Table S10, it can be seen that the P0 of adduct products involved in the successive reactions of CH2OO with HCOOH increases first and then decreases with increasing the number of CH2OO. The P0 of the adduct product HC(O)O(CH2OO)3H is maximum when the number of CH2OO is equal to three. The P0 of adduct products included in the successive reactions of anti-CH3CHOO with HCOOH decreases significantly as the number of anti-CH3CHOO is increased. Similar phenomenon is also observed from the successive reactions of syn-CH3CHOO and (CH3)2COO with HCOOH. Notably, the P0 of adduct products decreases obviously when the size of SCIs increases. For example, the P0 of the adduct product HC(O)O(CH2OO)3H in the nCH2OO + HCOOH reaction is estimated to be 4.43 × 10-3 atm, which is greater than those of the corresponding adduct products in the nanti-CH3CHOO + HCOOH (7.12 × 10-4), nsyn-CH3CHOO + HCOOH (7.12 × 10-4), and n(CH3)2COO + HCOOH (1.27 × 10-4) reactions by 6.22, 6.22 and 34.88 times, respectively.
A classify scheme of various organic compounds is based on their volatility, as presented by Donahue et al. (2012) The volatility of organic compounds is described by their effective saturation concentration. The saturated concentrations (c0) of adduct products formed from the successive reactions of SCIs with HCOOH are predicted with the SIMPOL.1 method proposed by Pankow and Asher (2008), and the results are listed in Table S10. As shown in Table S10, the c0 of adduct products involved in the nCH2OO + HCOOH reaction decreases with increasing the number of CH2OO. According to the Volatility Basis Set (VBS) of organic compounds (Donahue et al., 2012), these adduct products belong to volatile organic compound (VOC) (c0 > 3 × 106 ug/m3). Similarly, the c0 of adduct products included in the nanti-CH3CHOO + HCOOH, nsyn-CH3CHOO + HCOOH, and n(CH3)2COO + HCOOH reactions decreases when the number of SCIs increases. It deserves mentioning that the adduct products in the nanti-CH3CHOO + HCOOH and nsyn-CH3CHOO + HCOOH reactions belong to intermediate volatility organic compounds (IVOC, 300 < c0 < 3 × 106 ug/m3) when the number of SCIs is equal to five. However, the adduct products in the n(CH3)2COO + HCOOH reaction become IVOC when the number of (CH3)2COO is greater than or equal to two. Based on the above discussions, it can be concluded that the volatility of adduct products is significantly affected by the number and size of SCIs in the successive reaction of SCIs with HCOOH.
Table 2 The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with HPMF, H2O, HCOOH and SO2 at the tropical forest environments
SCIs
Coreactant
[Coreactant]
(molecules cm-3)
k
(cm3 molecule-1 s-1)
keff
(s-1)
Reference
CH2OO
H2O
3.9-6.1 × 1017
< 1.5 × 10-15
5.9-9.2 × 102
Chao et al., (2015)
HCOOH
5.0-10.0 × 1010
[1.1 ± 0.1] × 10-10
5.5-11
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[3.9 ± 0.7] × 10-11
0.7-3.5
Welz et al., (2012)
HPMF
-
2.7 × 10-11
-
This work
anti-CH3CHOO
H2O
3.9-6.1 × 1017
[1.0 ± 0.4] × 10-14
3.9-6.1 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[5 ± 3] × 10-10
25.0-50.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[6.7 ± 1.0] × 10-11
1.1-6.0
Taatjes et al., (2013)
HPMF
-
3.3 × 10-10
-
This work
syn-CH3CHOO
H2O
3.9-6.1 × 1017
< 4.0 × 10-15
1.6-2.4 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[2.5 ± 0.3] × 10-10
12.5-25.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[2.4 ± 0.3] × 10-11
0.4-2.2
Taatjes et al., (2013)
HPMF
-
1.7 × 10-13
-
This work
(CH3)2COO
H2O
3.9-6.1 × 1017
< 1.5 × 10-16
58.5-91.5
Huang et al., (2015)
HCOOH
5.0-10.0 × 1010
4.5 × 10-10
22.5-45.0
Sipilä et al., (2014)
SO2
1.7-9.0 × 1010
1.3 × 10-10
2.2-11.7
Huang et al., (2015)
HPMF
-
2.2 × 10-11
-
This work
Table S10 Predicted saturated vapour pressure (P0) and saturated concentrations (c0) for the adduct products of the successive reactions of SCIs with HCOOH
formula
P0 (atm)
c0 (ug/m3)
n CH2OO + HCOOH
n = 1
HC(O)OCH2OOH
2.12 × 10-3
7.86 × 107
n = 2
HC(O)O(CH2OO)2H
3.80 × 10-3
3.99 × 107
n = 3
HC(O)O(CH2OO)3H
4.43 × 10-3
3.91 × 107
n = 4
HC(O)O(CH2OO)4H
4.21 × 10-3
3.29 × 107
n = 5
HC(O)O(CH2OO)5H
3.59 × 10-3
2.12 × 107
n anti-CH3CHOO + HCOOH
n = 1
HC(O)OCH(CH3)OOH
1.25 × 10-3
8.32 × 106
n = 2
HC(O)O(CH(CH3)OO)2H
1.13 × 10-3
7.57 × 106
n = 3
HC(O)O(CH(CH3)OO)3H
7.12 × 10-4
6.49 × 106
n = 4
HC(O)O(CH(CH3)OO)4H
3.90 × 10-4
4.50 × 106
n = 5
HC(O)O(CH(CH3)OO)5H
2.01 × 10-4
2.81 × 106
n syn-CH3CHOO + HCOOH
n = 1
HC(O)OCH(CH3)OOH
1.25 × 10-3
8.32 × 106
n = 2
HC(O)O(CH(CH3)OO)2H
1.13 × 10-3
7.57 × 106
n = 3
HC(O)O(CH(CH3)OO)3H
7.12 × 10-4
6.49 × 106
n = 4
HC(O)O(CH(CH3)OO)4H
3.90 × 10-4
4.50 × 106
n = 5
HC(O)O(CH(CH3)OO)5H
2.01 × 10-4
2.81 × 106
n (CH3)2COO + HCOOH
n = 1
HC(O)OC(CH3)2OOH
7.23 × 10-4
3.50 × 106
n = 2
HC(O)O(C(CH3)2OO)2H
3.50 × 10-4
2.74 × 106
n = 3
HC(O)O(C(CH3)2OO)3H
1.27 × 10-4
1.38 × 106
n = 4
HC(O)O(C(CH3)2OO)4H
4.27 × 10-5
5.90 × 105
n = 5
HC(O)O(C(CH3)2OO)5H
1.40 × 10-5
2.36 × 105
Corresponding descriptions have been added in the page 23 line 573-590, page 24 line 591-610, page 27 line 645-671 and page 28 line 672-682 of the revised manuscript:
It is of interest to assess whether the reactions of distinct SCIs with HPMF can compete well with the losses to reactions with trace species (e.g., H2O, HCOOH and SO2), because it is well known that the reactions with trace species are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2, and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmospheric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken, 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al., 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF)is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF)is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
The assessment of Barley and McFiggans (2010) and O'Meara et al. (2014) found that the combination of boiling point estimation from Nannoolal et al. (2004) and vapour pressure estimation from Nannoolal et al. (2008) gives the lowest mean bias error of vapour pressure for atmospherically relevant compounds. Therefore, the saturated vapour pressure (P0) of adduct products at room temperature is estimated by using the Nannoolal-Nannoolal method, and the results are listed in Table S10. From Table S10, it can be seen that the P0 of adduct products involved in the successive reactions of CH2OO with HCOOH increases first and then decreases with increasing the number of CH2OO. The P0 of the adduct product HC(O)O(CH2OO)3H is maximum when the number of CH2OO is equal to three. The P0 of adduct products included in the successive reactions of anti-CH3CHOO with HCOOH decreases significantly as the number of anti-CH3CHOO is increased. Similar phenomenon is also observed from the successive reactions of syn-CH3CHOO and (CH3)2COO with HCOOH. Notably, the P0 of adduct products decreases obviously when the size of SCIs increases. For example, the P0 of the adduct product HC(O)O(CH2OO)3H in the nCH2OO + HCOOH reaction is estimated to be 4.43 × 10-3 atm, which is greater than those of the corresponding adduct products in the nanti-CH3CHOO + HCOOH (7.12 × 10-4), nsyn-CH3CHOO + HCOOH (7.12 × 10-4), and n(CH3)2COO + HCOOH (1.27 × 10-4) reactions by 6.22, 6.22 and 34.88 times, respectively.
A classify scheme of various organic compounds is based on their volatility, as presented by Donahue et al. (2012) The volatility of organic compounds is described by their effective saturation concentration. The saturated concentrations (c0) of adduct products formed from the successive reactions of SCIs with HCOOH are predicted by using the SIMPOL.1 method proposed by Pankow and Asher (2008), and the results are listed in Table S10. As shown in Table S10, the c0 of adduct products involved in the nCH2OO + HCOOH reaction decreases with increasing the number of CH2OO. According to the Volatility Basis Set (VBS) of organic compounds (Donahue et al., 2012), these adduct products belong to VOC (c0 > 3 × 106 ug/m3). Similarly, the c0 of adduct products included in the nanti-CH3CHOO + HCOOH, nsyn-CH3CHOO + HCOOH, and n(CH3)2COO + HCOOH reactions decreases when the number of SCIs increases. It deserves mentioning that the adduct products in the nanti-CH3CHOO + HCOOH and nsyn-CH3CHOO + HCOOH reactions belong to intermediate volatility organic compounds (IVOC, 300 < c0 < 3 × 106 ug/m3) when the number of SCIs is equal to five. However, the adduct products in the n(CH3)2COO + HCOOH reaction become IVOC when the number of (CH3)2COO is greater than or equal to two. Based on the above discussions, it can be concluded that the volatility of adduct products is significantly affected by the number and size of SCIs in the successive reaction of SCIs with HCOOH.
- Hence, as a quick assessment, some deeper and extended discussions should be required and strengthened, such as the nature of the reactions, the detailed atmospheric implications, if this paper is published in the ACP journal.
Response: Based on the Reviewer’s suggestion, the deeper discussions on the nature of the reactions of SCIs with hydroperoxide esters have been added in the revised manuscript. A schematic potential energy surface (PES) for the addition reaction 2CH2OO + Pent1a is drawn in Fig. 2. As seen in Fig. 2, the successive insertion of CH2OO into Pent1a eventually leads to the formation of oligomers P2a and P2b composed of CH2OO as the repeat unit. These oligomerization reactions are strongly exothermic and spontaneous (> 83 kcal·mol-1), implying that they are feasible thermodynamically. The addition reaction 2CH2OO + Pent1a initially proceeds through two possible pathways, namely (1) –OOH insertion reaction R1a, and (2) –CH insertion reaction R1b. For the –OOH insertion reaction R1a, the pre-reactive intermediate IM1a with a seven-membered ring structure is formed in the entrance channel, which is stabilized by the hydrogen bond interactions between the H4 atom of Pent1a and the O6 atom of CH2OO (D(O6-H4) = 1.706 Å), and between the H6 atom of CH2OO and the O3 atom of Pent1a (D(O3-H6) = 2.115 Å). Then IM1a converts into P1a (C3H6O6, HC(O)O–(CH2OO)2–H) via a concerted process of O4-H4 bond breaking in the Pent1a and O4-C3 and H4-O6 bonds forming with a barrier of 8.1 kcal·mol-1. For the –CH insertion reaction R1b, the pre-reactive intermediate IM1b with a seven-membered ring structure is formed in the entrance channel, which is stabilized by the van der Waals (vdW) interactions between the O3 atom of Pent1a and the C3 atom of CH2OO (D(O3-C3) = 2.602 Å), and between the O6 atom of CH2OO and the C1 atom of Pent1a (D(O6-C1) = 2.608 Å). Due to the absence of hydrogen bond in IM1b, the energy of IM1b is lower than that of IM1a by 3.0 kcal·mol-1. IM1b transforms into P1b (C3H6O6, HO2CH2OC(O)CH2OOH) via a concerted process of C1-H1 bond breaking in the Pent1a and C1-C3 and H1-O6 bonds forming with a barrier of 21.5 kcal·mol-1. By comparing the barriers of R1a and R1b, it can be concluded that the –OOH insertion reaction is favored over the –CH insertion reaction. The high reaction barrier of R1b is attributed to the large bond dissociation energy (BDE) of C-H bond in the Pent1a. To further insight into the reaction mechanism of R1a, the natural bond orbital (NBO) analysis of the donor-accepter orbitals involved in the TS1a is performed using the M06-2X wave function. The possible donor-accepter interactions are estimated by using the second order perturbation theory. As illustrated in Fig. S4, the strong interactions are identified as the interaction of the lone pair orbital of O6 atom and the antibonding orbital of O4-H4 bond, and the interaction of the lone pair orbital of O4 atom and the antibonding orbital of C3-O5 bond.
Similarly, the addition reaction CH2OO + P1a proceeds through the formation of the pre-reactive intermediates IM2a and IM2b in the entrance channel, which are stabilized by a hydrogen bond between the terminal oxygen atom of CH2OO and the reacting hydrogen atom of P1a, and a van der Waals (vdW) interaction between the central carbon atom of CH2OO and the carbonyl oxygen atom of P1a. The relative energies of IM2a and IM2b with respect to the separate reactants P1a and CH2OO are -1.2 and 3.2 kcal·mol-1, respectively, below the energies of the initial reactants 2CH2OO and Pent1a are 41.6 and 37.2 kcal·mol-1, respectively. Then they immediately transform into the respective products P2a and P2b through the –OOH and –CH insertion transition states TS2a and TS2b with the barriers of 10.1 and 21.6 kcal·mol-1. This result again shows that the –OOH insertion reaction is favored kinetically. It deserves mentioning that the barrier of –OOH insertion reaction increases as the number of CH2OO is increased. From the viewpoint of the geometrical parameters of TS2a and TS2b, the breaking O-H and C-H bonds are elongated by 14.8% and 20.6%, respectively, with respect to the equilibrium structures of IM2a and IM2b, while the forming C-O and C-C bond length are 2.013 and 2.264 Å, respectively. The result reveals that TS2a and TS2b are structurally reactant-like, which are consistent with the Hammond’s hypothesis that the earlier transition states are generally exothermic (Hammond, 1955).
Figure 2. PES (ΔG and ΔE, in italics) for the 2CH2OO+ Pent1a reaction at the M06-2X/ma-TZVP//M06-2X/6-311+G(2df,2p) level of theory
Figure S4. Natural bond orbital (NBO) analysis of the donor-acceptor orbitals involved in the TS1a
To further elucidate the effect of the number and location of methyl substituents on the reactivity of carbonyl oxides toward hydroperoxide esters, Pent1a (also called as HPMF) is selected as the model compound since it is the simplest hydroperoxide ester formed from the barrierless reaction of 1,4 O-H insertion of CH2OO into HCOOH. As mentioned above, –OOH insertion reaction in the oligomerization reactions is the most favorable pathway. Therefore, this type of reaction is merely considered in the reactions of distinct SCIs with Pent1a. The corresponding PES is displayed in Fig. 6. As shown in Fig. 6, each pathway starts with the formation of a pre-reactive intermediate, and then it overcomes a modest barrier to reaction. The barrier of the reaction of CH2OO with Pent1a is calculated to be 8.1 kcal·mol-1, which is higher than that of the anti-CH3CHOO + Pent1a reaction by 2.5 kcal·mol-1. The reason of low barrier can be explained by the NPA atomic charges, as presented in Fig. S9. As seen in Fig. S9, the charges of the central carbon atom C1 and the terminal oxygen atom O1 of CH2OO are 0.186e and -0.459e, respectively, indicating that CH2OO is indeed a zwitterion. The C1 atom charge becomes more positive (0.393e), while the O1 atom charge becomes more negative (-0.497e) when a methyl substituent occurs at the anti-position. This result suggests that the anti-methyl substituent enhances the characteristic of carbonyl oxides zwitterion and reduces the reaction barriers. Compared with the barrier of the CH2OO + Pent1a reaction, the barriers increase by about 3.0 kcal·mol-1 when a methyl group is introduced at the syn-position and dimethyl substituent. Although syn-methyl and dimethyl substituent promote the raise of carbonyl oxides zwitterion, the steric hindrance effect and intramolecular hydrogen bond are obviously dominant for syn-CH3CHOO and (CH3)2COO, that are not thus conducive to the nucleophilic attack of hydroperoxide esters. It is worth noting that the exothermicity of distinct SCIs reactions with Pent1a obviously decreases as the number of methyl group is increased, and the exothermicity of anti-methyl substituent is higher than that of syn-methyl substituent.
Figure 6. PES (ΔG and ΔE, in italics) for the distinct SCIs + Pent1a reactions at the M06-2X/ma-TZVP//M06-2X/6-311+G(2df,2p) level of theory
Figure S9 The NPA charges of different atoms in the distinct SCIs computed at the M06-2X/6-311+g(2df,2p) level of theory
Corresponding descriptions have been added in the page 15 line 388-398, page 16 line 399-438 and page 21 line 530-556 of the revised manuscript:
A schematic PES for the addition reaction 2CH2OO + Pent1a is drawn in Fig. 2. As seen in Fig. 2, the successive insertion of CH2OO into Pent1a eventually leads to the formation of oligomers P2a and P2b composed of CH2OO as the repeat unit. These oligomerization reactions are strongly exothermic and spontaneous (> 83 kcal·mol-1), implying that they are feasible thermodynamically. The addition reaction 2CH2OO + Pent1a initially proceeds through two possible pathways, namely (1) –OOH insertion reaction R1a, and (2) –CH insertion reaction R1b. For the –OOH insertion reaction R1a, the pre-reactive intermediate IM1a with a seven-membered ring structure is formed in the entrance channel, which is stabilized by the hydrogen bond interactions between the H4 atom of Pent1a and the O6 atom of CH2OO (D(O6-H4) = 1.706 Å), and between the H6 atom of CH2OO and the O3 atom of Pent1a (D(O3-H6) = 2.115 Å). Then IM1a converts into P1a (C3H6O6, HC(O)O–(CH2OO)2–H) via a concerted process of O4-H4 bond breaking in the Pent1a and O4-C3 and H4-O6 bonds forming with a barrier of 8.1 kcal·mol-1. For the –CH insertion reaction R1b, the pre-reactive intermediate IM1b with a seven-membered ring structure is formed in the entrance channel, which is stabilized by the van der Waals (vdW) interactions between the O3 atom of Pent1a and the C3 atom of CH2OO (D(O3-C3) = 2.602 Å), and between the O6 atom of CH2OO and the C1 atom of Pent1a (D(O6-C1) = 2.608 Å). Due to the absence of hydrogen bond in IM1b, the energy of IM1b is lower than that of IM1a by 3.0 kcal·mol-1. IM1b transforms into P1b (C3H6O6, HO2CH2OC(O)CH2OOH) via a concerted process of C1-H1 bond breaking in the Pent1a and C1-C3 and H1-O6 bonds forming with a barrier of 21.5 kcal·mol-1. By comparing the barriers of R1a and R1b, it can be concluded that the –OOH insertion reaction is favored over the –CH insertion reaction. The high reaction barrier of R1b is attributed to the large bond dissociation energy (BDE) of C-H bond in the Pent1a. To further insight into the reaction mechanism of R1a, the natural bond orbital (NBO) analysis of the donor-accepter orbitals involved in the TS1a is performed using the M06-2X wave function. The possible donor-accepter interactions are estimated by using the second order perturbation theory. As illustrated in Fig. S4, the strong interactions are identified as the interaction of the lone pair orbital of O6 atom and the antibonding orbital of O4-H4 bond, and the interaction of the lone pair orbital of O4 atom and the antibonding orbital of C3-O5 bond.
Similarly, the addition reaction CH2OO + P1a proceeds through the formation of the pre-reactive intermediates IM2a and IM2b in the entrance channel, which are stabilized by a hydrogen bond between the terminal oxygen atom of CH2OO and the reacting hydrogen atom of P1a, and a van der Waals (vdW) interaction between the central carbon atom of CH2OO and the carbonyl oxygen atom of P1a. The relative energies of IM2a and IM2b with respect to the separate reactants P1a and CH2OO are -1.2 and 3.2 kcal·mol-1, respectively, below the energies of the initial reactants 2CH2OO and Pent1a are 41.6 and 37.2 kcal·mol-1, respectively. Then they immediately transform into the respective products P2a and P2b through the –OOH and –CH insertion transition states TS2a and TS2b with the barriers of 10.1 and 21.6 kcal·mol-1. This result again shows that the –OOH insertion reaction is favored kinetically. It deserves mentioning that the barrier of –OOH insertion reaction increases as the number of CH2OO is increased. From the viewpoint of the geometrical parameters of TS2a and TS2b, the breaking O-H and C-H bonds are elongated by 14.8% and 20.6%, respectively, with respect to the equilibrium structures of IM2a and IM2b, while the forming C-O and C-C bond length are 2.013 and 2.264 Å, respectively. The result reveals that TS2a and TS2b are structurally reactant-like, which are consistent with the Hammond’s hypothesis that the earlier transition states are generally exothermic (Hammond, 1955).
To further elucidate the effect of the number and location of methyl substituents on the reactivity of carbonyl oxides toward hydroperoxide esters, Pent1a (also called as HPMF) is selected as the model compound since it is the simplest hydroperoxide ester formed from the barrierless reaction of 1,4 O-H insertion of CH2OO into HCOOH. As mentioned above, –OOH insertion reaction in the oligomerization reactions is the most favorable pathway. Therefore, this type of reaction is merely considered in the reactions of distinct SCIs with Pent1a. The corresponding PES and the optimized geometries of all stationary points are displayed in Figs. 6 and S8, respectively. As seen in Fig. 6, each pathway starts with the formation of a pre-reactive intermediate, and then it overcomes a modest barrier to reaction. The barrier of the reaction of CH2OO with Pent1a is calculated to be 8.1 kcal·mol-1, which is higher than that of the anti-CH3CHOO + Pent1a reaction by 2.5 kcal·mol-1. The reason of low barrier can be explained by the NPA atomic charges, as presented in Fig. S9. As seen in Fig. S9, the charges of the central carbon atom C1 and the terminal oxygen atom O1 of CH2OO are 0.186e and -0.459e, respectively, indicating that CH2OO is indeed a zwitterion. The C1 atom charge becomes more positive (0.393e), while the O1 atom charge becomes more negative (-0.497e) when a methyl substituent occurs at the anti-position. This result suggests that the anti-methyl substituent enhances the characteristic of carbonyl oxides zwitterion and reduces the reaction barriers. Compared with the barrier of the CH2OO + Pent1a reaction, the barriers increase by about 3.0 kcal·mol-1 when a methyl group is introduced at the syn-position and dimethyl substituent. Although syn-methyl and dimethyl substituent promote the raise of carbonyl oxides zwitterion, the steric hindrance effect and intramolecular hydrogen bond are obviously dominant for syn-CH3CHOO and (CH3)2COO, that are not thus conducive to the nucleophilic attack of hydroperoxide esters. It is worth noting that the exothermicity of distinct SCIs reactions with Pent1a obviously decreases as the number of methyl group is increased, and the exothermicity of anti-methyl substituent is higher than that of syn-methyl substituent.
Reference
Barley, M. H., and McFiggans, G.: The critical assessment of vapour pressure estimation methods for use in modelling the formation of atmospheric organic aerosol, Atmos. Chem. Phys., 10, 749-767, https://doi.org/10.5194/acp-10-749-2010, 2010.
Canneaux, S., Bohr, F., and Henon, E.: KiSThelP: a program to predict thermodynamic properties and rate constants from quantum chemistry results, J. Comput. Chem., 35, 82-93, https://doi.org/10.1002/jcc.23470, 2013.
Chao, W., Hsieh, J. T., Chang, C. H., and Lin, J. J. M.: Direct kinetic measurement of the reaction of the simplest Criegee intermediate with water vapor, Science, 347, 751-754, https://doi.org/10.1126/science.1261549, 2015.
Chung, C. A., Su, J. W., and Lee, Y. P.: Detailed mechanism and kinetics of the reaction of Criegee intermediate CH2OO with HCOOH investigated via infrared identification of conformers of hydroperoxymethyl formate and formic acid anhydride, Phys. Chem. Chem. Phys., 21, 21445-21455, https://doi.org/10.1039/c9cp04168k, 2019.
Donahue, N. M., Kroll, J. H., Pandis, S. N., and Robinson, A. L.: A two-dimensional volatility basis set – Part 2: Diagnostics of organic-aerosol evolution, Atmos. Chem. Phys., 12, 615-634, https://doi.org/10.5194/acp-12-615-2012, 2012.
Gilbert, R. G., and Smith, S. C.: Theory of unimolecular and recombination reactions; Blackwell Scientific: Carlton, Australia, 1990.
Glowacki, D. R., Liang, C. H., Morley, C., Pilling, M. J., and Robertson, S. H.: MESMER: an open-source master equation solver for multi-energy well reactions, J. Phys. Chem. A, 116, 9545-9560, https://doi.org/10.1021/jp3051033, 2012.
Hammond, G. S.: A correlation of reaction rates, J. Am. Chem. Soc., 77, 334-338, https://doi.org/10.1021/ja01607a027, 1955.
Huang, H. L., Chao, W., and Lin, J. J. M.: Kinetics of a Criegee intermediate that would survive high humidity and may oxidize atmospheric SO2, Proc. Natl. Acad. Sci. U.S.A., 112, 10857-10862, https://doi.org/ 10.1073/pnas.1513149112, 2015.
Khan, M. A. H., Percival, C. J., Caravan, R. L., Taatjes, C. A., and Shallcross, D. E.: Criegee intermediates and their impacts on the troposphere, Environ. Sci.: Processes Impacts, 20, 437-453, https://doi.org/10.1039/C7EM00585G, 2018.
Long, B., Bao, J. L., and Truhlar, D. G.: Atmospheric chemistry of Criegee intermediates: unimolecular reactions and reactions with water, J. Am. Chem. Soc., 138, 14409-14422, https://doi.org/10.1021/jacs.6b08655, 2016.
Nannoolal, Y., Rarey, J., and Ramjugernatha, D.: Estimation of pure component properties Part 3. Estimation of the vapor pressure of non-electrolyte organic compounds via group contributions and group interactions, Fluid Phase Equilibria, 269, 117-133, https://doi.org/10.1016/j.fluid.2008.04.020, 2008.
Nannoolal, Y., Rarey, J., Ramjugernatha, D., and Cordesb, W.: Estimation of pure component properties Part 1. Estimation of the normal boiling point of non-electrolyte organic compounds via group contributions and group interactions, Fluid Phase Equilibria, 226, 45-63, https://doi.org/10.1016/j.fluid.2004.09.001, 2004.
Novelli, A., Hens, K., Ernest, C. T., Martinez, M., Nölscher, A. C., Sinha, V., Paasonen, P., Petäjä, T., Sipilä, M., Elste, T., Plass-Dülmer, C., Phillips, G. J., Kubistin, D., Williams, J., Vereecken, L., Lelieveld, J., and Harder, H.: Estimating the atmospheric concentration of Criegee intermediates and their possible interference in a FAGE-LIF instrument, Atmos. Chem. Phys., 17, 7807-7826, https://doi.org/10.5194/acp-17-7807-2017, 2017.
O'Meara, S., Booth, A. M., Barley, M. H., Topping, D., and McFiggans, G.: An assessment of vapour pressure estimation methods, Phys. Chem. Chem. Phys., 16, 19453-19469, https://doi.org/10.1039/C4CP00857J, 2014.
Pankow, J. F., and Asher, W. E.: SIMPOL.1: a simple group contribution method for predicting vapor pressures and enthalpies of vaporization of multifunctional organic compounds, Atmos. Chem. Phys., 8, 2773-2796, https://doi.org/10.5194/acp-8-2773-2008, 2008.
Peltola, J., Seal, P., Inkilä, A., and Eskola, A.: Time-resolved, broadband UV-absorption spectrometry measurements of Criegee intermediate kinetics using a new photolytic precursor: unimolecular decomposition of CH2OO and its reaction with formic acid, Phys. Chem. Chem. Phys., 22, 11797-11808, https://doi.org/10.1039/d0cp00302f, 2020.
Raghunath, P., Lee, Y. P., and Lin, M. C.: Computational chemical kinetics for the reaction of Criegee intermediate CH2OO with HNO3 and its catalytic conversion to OH and HCO, J. Phys. Chem. A, 121, 3871-3878, https://doi.org/10.1021/acs.jpca.7b02196, 2017.
Sipilä, M., Jokinen, T., Berndt, T., Richters, S., Makkonen, R., Donahue, N. M., Mauldin Iii, R. L., Kurtén, T., Paasonen, P., Sarnela, N., Ehn, M., Junninen, H., Rissanen, M. P., Thornton, J., Stratmann, F., Herrmann, H., Worsnop, D. R., Kulmala, M., Kerminen, V. M., and Petäjä, T.: Reactivity of stabilized Criegee intermediates (sCIs) from isoprene and monoterpene ozonolysis toward SO2 and organic acids, Atmos. Chem. Phys., 14, 12143-12153, https://doi.org/10.5194/acp-14-12143-2014, 2014.
Taatjes, C. A., Welz, O., Eskola, A. J., Savee, J. D., Scheer, A. M., Shallcross, D. E., Rotavera, B., Lee, E. P. F., Dyke, J. M., Mok, D. K. W., Osborn, D. L., and Percival, C. J.: Direct measurements of conformer-dependent reactivity of the Criegee intermediate CH3CHOO, Science, 340, 177-180, https://doi.org/10.1126/science.1234689, 2013.
Vereecken, L., Harder, H., and Novelli, A.: The reaction of Criegee intermediates with NO, RO2, and SO2, and their fate in the atmosphere, Phys. Chem. Chem. Phys., 14, 14682-14695, https://doi.org/10.1039/c2cp42300f, 2012.
Welz, O., Eskola, A. J., Sheps, L., Rotavera, B., Savee, J. D., Scheer, A. M., Osborn, D. L., Lowe, D., Booth, A. M., Xiao, P., Khan, M. A. H., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Rate coefficients of C(1) and C(2) Criegee intermediate reactions with formic and acetic Acid near the collision limit: direct kinetics measurements and atmospheric implications, Angew. Chem. Int. Ed., 53, 4547-4550, https://doi.org/10.1002/anie.201400964, 2014.
Welz, O., Savee, J. D., Osborn, D. L., Vasu, S. S., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Direct kinetic measurements of Criegee intermediate (CH2OO) formed by reaction of CH2I with O2, Science, 335, 204-207, https://doi.org/10.1126/science.1213229, 2012.
-
AC2: 'Reply on RC2', Y. Huang, 25 Aug 2022
-
RC3: 'Comment on acp-2022-376', Anonymous Referee #3, 08 Jul 2022
General comments:
Chen et al. studied the the oligomerization reaction mechanisms and kinetics of several stabilized Criegee intermediates (SCIs) in the presence of HCOOH using quantum chemical and kinetics modeling methods. The major conclusion is that the oligomerization is mainly initiated by barrierless reactions of SCIs with HCOOH and proceeds through highly exothermic insertion of SCIs into hydroperoxide ester. Meanwhile, the influence of methyl substitution on SCIs were examined. Overall, this paper provides a good amount of data on SCIs chemistry. However, some of the conclusions need to be carefully validated and further discussions are suggested in terms of the mechanism and atmospheric implication.
Specific comments:
For Entry 1 of the initiation reaction, how is it validated that 1,4 O-H insertion is barrierless? Is there a multi-point potential energy surface showing that no barrier is found along the reaction coordinate?
The calculated ktot in this study is greater by a factor of ~3 than several previous studies. Since this is related to one of the major conclusions of the paper, the authors should carefully validate this result. For example, what could be the reason they underestimate the value? Which value can have a better interpretation of the experimental or atmospheric data?
Why are k(TSent2) and k(TSent4) decrease with increasing temperature as they both have positive energy barrier? (Table S2)
The oligomerization reactions are highly dependent on the concentration of the monomers. Here the monomer are highly reactive SCIs and usually has very low concentration in the atmosphere. It seems that the high exothermicity of the oligomerization reaction results from the “stabilization” of SCIs in oligomerization. Also, the calculated free energies represent standard condition. Could the authors correct the Gibbs free energies by incorporating the atmospheric concentrations of SCIs (i.e., RTln(P/Pref)) to check whether this oligomerization is favored in the atmospheric conditions?
Additionally, it would be helpful if there is some estimation about how much the oligomerization process could contribute to the regional or global SOA.
Technical corrections:
Line 39, “with increasing the number of SCIs” is a bit confusing, it would be better to say “with increasing the number of SCIs added to the oligomer”.
Line 491, “netative” should be “negative”.
Line 499, “neartly” should be “nearly”.
Citation: https://doi.org/10.5194/acp-2022-376-RC3 -
AC3: 'Reply on RC3', Y. Huang, 25 Aug 2022
Prof. Yu Huang
State Key Lab of Loess and Quaternary Geology
Institute of Earth Environment, Chinese Academy of Sciences, Xi’an, 710061, China
Tel./Fax: (86) 29-62336261
E-mail: huangyu@ieecas.cn
Aug. 25, 2022
Dear Prof. Kourtchev,
Revision for Manuscript ACP-2022-376
We thank you very much for giving us the opportunity to revise our manuscript. We highly appreciate the reviewers for their comments and suggestions on the manuscript entitled “Oligomer formation from the gas-phase reactions of Criegee intermediates with hydroperoxide esters: mechanism and kinetics”. We have made revisions of our manuscript carefully according to the comments and suggestions of reviewers. The revised contents are marked in blue color. The response letter to reviewers is attached at the end of this cover letter.
We hope that the revised manuscript can meet the requirement of Atmospheric Chemistry & Physics. Any further modifications or revisions, please do not hesitate to contact us.
Look forward to hearing from you as soon as possible.
Best regards,
Yu Huang
Comments of reviewer #3
- For Entry 1 of the initiation reaction, how is it validated that 1,4 O-H insertion is barrierless? Is there a multi-point potential energy surface showing that no barrier is found along the reaction coordinate?
Response: Based on the Reviewer’s suggestion, the relevance descriptions on the barrierless 1,4 O-H insertion reactions have been added in the revised manuscript. The potential energy surface (PES) of the initiation reactions of distinct stabilized Criegee intermediates (SCIs) (CH2OO, syn-, anti-CH3CHOO and (CH3)2COO) with HCOOH is drawn in Fig. 1. As shown in Fig. 1, the bimolecular reaction of distinct SCIs with HCOOH proceeds via four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4). For Entry 1, the addition reaction of CH2OO with HCOOH proceeds through the 1,4 O-H insertion of CH2OO into HCOOH to form a hydroperoxide ester HC(O)O-CH2OO-H with a exoergicity of 37.6 kcal·mol-1. The formation of HC(O)O-CH2OO-H is obtained through a concerted process of O2-H2 bond breaking in the HCOOH and O4-H2 and C2-O1 bonds forming. Despite an attempt by various methods, the corresponding transition state is still not located in the effort of optimization. To further validate the barrierless process of 1,4 O-H insertion reaction, a relaxed scan over the O4-H2 and C2-O1 bonds is performed at the M06-2X/6-311+G(2df,2p) level of theory. The scans start from the optimized structure of the adduct product HC(O)O-CH2OO-H, and the O4-H2 and C2-O1 bond length are then increased in steps of 0.10 Å. The relaxed scan energy profiles are presented in Fig. S2. As seen in Fig. S2a, the relative energy of the minimum energy path from reactant to product decreases monotonically when the bond length of O4-H2 and C2-O1 bonds decreases, suggesting that the transition state is not exist in the 1,4 O-H insertion reaction of CH2OO with HCOOH. Similar conclusion is also obtained from the relaxed scan energy profiles for the HCOOH + anti-CH3CHOO, HCOOH + syn-CH3CHOO and HCOOH + (CH3)2COO (Fig. S2b-d) reactions that 1,4 O-H insertion reactions are barrierless. This conclusion is further supported by the analogous reaction systems that 1,4 O-H insertion reactions of carbonyl oxides with carboxylic acids are a barrierless process including concerted hydrogen atom transfer and new bond formation (Long et al., 2009; Vereecken, 2017; Cabezas and Endo, 2019; Lin et al., 2019; Chhantyal-Pun et al., 2017).
Figure 1. Schematic PES for the possible entrance pathways of the initiation reactions of HCOOH with various SCIs (black, pink, blue, and red lines represent 1,4 O-H insertion, 1,2 O-H insertion, C-H insertion, and C=O cycloaddition reactions, respectively)
Figure S2. Relaxed scan energy profiles calculated using the M06-2X/6-311+G(2df,2p) method for varying the C-O and O-H bonds in the 1,4-insertion reactions CH2OO + HCOOH (a), anti-CH3CHOO + HCOOH (b), syn-CH3CHOO + HCOOH (c) and (CH3)2COO + HCOOH (d) (the black solid line represents the minimum energy path)
Corresponding descriptions have been added in the page 8 line 215-239 of the revised manuscript:
The potential energy surface (PES) of distinct SCIs (CH2OO, syn-, anti-CH3CHOO and (CH3)2COO) reactions with HCOOH is drawn in Fig. 1. As shown in Fig. 1, the bimolecular reaction of distinct SCIs with HCOOH proceeds via four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4). For Entry 1, the addition reaction of CH2OO with HCOOH proceeds through the 1,4 O-H insertion of CH2OO into HCOOH to form a hydroperoxide ester HC(O)O-CH2OO-H with a exoergicity of 37.6 kcal·mol-1. The formation of HC(O)O-CH2OO-H is obtained through a concerted process of O2-H2 bond breaking in the HCOOH and O4-H2 and C2-O1 bonds forming. Despite an attempt by various methods, the corresponding transition state is still not located in the effort of optimization. To further validate the barrierless process of 1,4 O-H insertion reaction, a relaxed scan over the O4-H2 and C2-O1 bonds is performed at the M06-2X/6-311+G(2df,2p) level of theory. The scans start from the optimized structure of the adduct product HC(O)O-CH2OO-H, and the O4-H2 and C2-O1 bond length are then increased in steps of 0.10 Å. The relaxed scan energy profiles are presented in Fig. S2. As seen in Fig. S2a, the relative energy of the minimum energy path from reactant to product decreases monotonically when the bond length of O4-H2 and C2-O1 bonds decreases, suggesting that the transition state is not exist in the 1,4 O-H insertion reaction of CH2OO with HCOOH. Similar conclusion is also obtained from the relaxed scan energy profiles for the HCOOH + anti-CH3CHOO, HCOOH + syn-CH3CHOO and HCOOH + (CH3)2COO (Fig. S2b-d) reactions that 1,4 O-H insertion reactions are barrierless. This conclusion is further supported by the analogous reaction systems that 1,4 O-H insertion reactions of carbonyl oxides with carboxylic acids are a barrierless process including concerted hydrogen atom transfer and new C-O bond formation (Chhantyal-Pun et al., 2017; Long et al., 2009; Vereecken, 2017; Cabezas and Endo, 2019; Lin et al., 2019).
- The calculated ktot in this study is greater by a factor of ~3 than several previous studies. Since this is related to one of the major conclusions of the paper, the authors should carefully validate this result. For example, what could be the reason they underestimate the value? Which value can have a better interpretation of the experimental or atmospheric data?
Response: In the original manuscript, the rate coefficients for the barrierless reactions are calculated by employing the variational transition state theory (VTST), and the rate coefficients for the bimolecular reactions with the tight transition states are computed by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction. For the initiation reactions of distinct SCIs with HCOOH, there are four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4), in which Entry 1 is barrierless and Entry 2-4 have the tight transition states. The total rate coefficient for the reaction of SCIs with HCOOH is equal to the sum of the rate coefficient of each pathway. For the barrierless 1,4 O-H insertion reaction, the VTST is approximated with a Morse potential function, V(R) = De{1-exp[-β(R-Re)]}2, along with an anisotropy potential function to stand for the minimum energy path, which is used to calculate the rate coefficients (Raghunath et al., 2017). Here, De is the bond energy excluding the zero-point energy, R is the reaction coordinate, and Re is the equilibrium value of R. It is assumed that the stretching potential in an anisotropy potential is used in conjunction with a potential form of Vanisotropy = V0[1-cos2(θ1–θ1e) × cos2(θ2–θ2e)] (Raghunath et al., 2017). Here, V0 is the stretching potential, which stands for by a Morse potential, θ1 and θ1e represent the rotational angle between fragment 1 and the reference axis and the equilibrium bond angle of fragment 1, θ2 and θ2e stand for the rotational angle between fragment 2 and the reference axis and the equilibrium bond angle of fragment 2. The association curve for the reaction of 1,4 O-H insertion of SCIs into HCOOH is computed at the M06-2X/6-311+G(2df,2p) level of theory to cover a range from 0.97 to 1.97 Å at step size 0.1 Å for O-H bond and from 1.44 to 2.44 Å at step size 0.1 Å for C-O bond, while other structural parameters are fully optimized. The computed potential energies are fitted to the Morse potential function. However, the calculated rate coefficients for the reactions of SCIs with HCOOH are higher than the prior experimental measurements. The reason is ascribed to the fact that the approximation of VTST using a Morse potential function in conjunction with an anisotropy potential function is unsuitable to predict the rate coefficients for the barrierless 1,4 O-H insertion reaction.
In the revised manuscript, the rate coefficients for the barrierless reactions are computed by employing the inverse Laplace transformation (ILT) method, and the rate coefficients for the bimolecular reactions with the tight transition states are calculated by utilizing CTST in conjunction with Eckart tunneling correction. The ILT and CTST/Eckart calculations are performed by using the MESMER 6.0 and KiSThelP 2019 programs, respectively (Glowacki et al., 2012; Canneaux et al., 2013). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken et al., 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OO is estimated to be 1.4 × 10-10 cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of HCOOH with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO (Table S4-S6). At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
Table S3 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of CH2OO with HCOOH computed at different temperatures
T/K
k (TSent1)
k (TSent2)
k (TSent3)
k (TSent4)
ktot-CH2OO
273
1.7 × 10-10
3.6 × 10-12
1.0 × 10-22
3.6 × 10-12
1.8 × 10-10
280
1.6 × 10-10
2.9 × 10-12
1.2 × 10-22
3.1 × 10-12
1.7 × 10-10
298
1.4 × 10-10
1.9 × 10-12
2.2 × 10-22
2.3 × 10-12
1.4 × 10-10
300
1.4 × 10-10
1.8 × 10-12
2.4 × 10-22
2.2 × 10-12
1.4 × 10-10
320
1.3 × 10-10
1.2 × 10-12
4.9 × 10-22
1.6 × 10-12
1.3 × 10-10
340
1.3 × 10-10
8.2 × 10-13
1.0 × 10-21
1.3 × 10-12
1.3 × 10-10
360
1.2 × 10-10
5.9 × 10-13
2.2 × 10-21
1.0 × 10-12
1.2 × 10-10
380
1.2 × 10-10
4.5 × 10-13
4.5 × 10-21
8.2 × 10-13
1.2 × 10-10
400
1.2 × 10-10
3.5 × 10-13
9.0 × 10-21
6.9 × 10-13
1.2 × 10-10
Table S4 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of anti-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-anti)
k (TSent2-anti)
k (TSent3-anti)
k (TSent4-anti)
ktot-anti
273
5.9 × 10-10
4.2 × 10-11
5.5 × 10-22
6.1 × 10-11
6.9 × 10-10
280
5.7 × 10-10
3.8 × 10-11
6.7 × 10-22
4.9 × 10-11
6.6 × 10-10
298
5.4 × 10-10
2.3 × 10-11
1.2 × 10-21
3.0 × 10-11
5.9 × 10-10
300
5.3 × 10-10
2.0 × 10-11
1.3 × 10-21
2.8 × 10-11
5.8 × 10-10
320
5.0 × 10-10
1.5 × 10-11
2.6 × 10-21
1.7 × 10-11
5.3 × 10-10
340
4.7 × 10-10
9.4 × 10-12
5.4 × 10-21
1.1 × 10-11
4.9 × 10-10
360
4.5 × 10-10
7.0 × 10-12
1.1 × 10-20
7.8 × 10-12
4.7 × 10-10
380
4.4 × 10-10
3.6 × 10-12
2.1 × 10-20
5.6 × 10-12
4.5 × 10-10
400
4.3 × 10-10
2.0 × 10-12
4.0 × 10-20
4.2 × 10-12
4.4 × 10-10
Table S5 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of syn-CH3CHOO with HCOOH computed at different temperatures
T/K
k (TSent1-syn)
k (TSent2-syn)
k (TSent3-syn)
k (TSent4-syn)
ktot-syn
273
3.1 × 10-10
9.5 × 10-13
4.6 × 10-27
7.5 × 10-16
3.1× 10-10
280
2.8 × 10-10
8.0 × 10-13
7.1 × 10-27
6.4 × 10-16
2.8× 10-10
298
2.7 × 10-10
5.4 × 10-13
8.9 × 10-26
5.5 × 10-16
2.7× 10-10
300
2.7 × 10-10
5.2 × 10-13
9.9 × 10-26
4.6 × 10-16
2.7× 10-10
320
2.5 × 10-10
3.6 × 10-13
3.0 × 10-25
3.8 × 10-16
2.5× 10-10
340
2.5 × 10-10
2.6 × 10-13
9.1 × 10-25
3.1 × 10-16
2.5× 10-10
360
2.3 × 10-10
2.0 × 10-13
2.6 × 10-24
3.0 × 10-16
2.3× 10-10
380
2.2 × 10-10
1.5 × 10-13
7.2 × 10-24
2.4 × 10-16
2.2× 10-10
400
2.2 × 10-10
1.2 × 10-13
1.8 × 10-23
2.2 × 10-16
2.2× 10-10
Table S6 Rate coefficients (cm3 molecule-1 s-1) of each elementary pathway involved in the initiation reaction of (CH3)2OO with HCOOH computed at different temperatures
T/K
k (TSent1-dim)
k (TSent2-dim)
k (TSent3-dim)
k (TSent4-dim)
ktot-dim
273
5.3 × 10-10
6.8 × 10-12
1.4 × 10-26
4.4 × 10-15
5.4 × 10-10
280
5.1 × 10-10
5.2 × 10-12
2.2 × 10-26
4.2 × 10-15
5.2 × 10-10
298
4.8 × 10-10
2.8 × 10-12
8.0 × 10-26
4.0 × 10-15
4.8 × 10-10
300
4.7 × 10-10
2.6 × 10-12
9.2 × 10-26
3.9 × 10-15
4.7 × 10-10
320
4.5 × 10-10
1.4 × 10-12
3.6 × 10-25
3.7 × 10-15
4.5 × 10-10
340
4.2 × 10-10
8.6 × 10-13
1.3 × 10-24
3.6 × 10-15
4.2 × 10-10
360
3.9 × 10-10
5.5 × 10-13
4.5 × 10-24
3.5 × 10-15
3.9 × 10-10
380
3.7 × 10-10
3.7 × 10-13
1.4 × 10-23
3.4 × 10-15
3.7 × 10-10
400
3.7 × 10-10
2.6 × 10-13
3.9 × 10-23
3.4 × 10-15
3.7 × 10-10
Corresponding descriptions have been added in the page 7 line 173-190, page 11 line 303-315, page 12 line 330-338 and page 13 line 346-351 of the revised manuscript:
The rate coefficients for the barrierless reactions are determined by employing the inverse Laplace transformation (ILT) method. The ILT calculations are performed with the MESMER 6.0 program (Glowacki et al., 2012). In the ILT treatment, the rotational constants, vibrational frequencies, molecular weights, energies and other input parameters are obtained from the M06-2X/6-311+G(2df,2p) or M06-2X/ma-TZVP methods. For the barrierless reaction of 1,4 O-H insertion of SCIs into HCOOH, SCIs and HCOOH are assigned as the deficient and excess reactants, respectively. The concentration of HCOOH is given a value of 5.0 × 1010 molecules cm-3 in the simulation, which is taken from the typical concentration of HCOOH in the tropical forest environments (Vereecken et al., 2012). N2 is applied as the buffer gas. A single exponential down model is employed to simulate the collision transfer (<ΔE>down = 200 cm-1). The collisional Lennard-Jones parameters are estimated with the empirical formula described by Gilbert and Smith (1990).
The rate coefficients for the bimolecular reactions with the tight transition states are calculated by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction (Truhlar et al., 1996; Eckart, 1930). The CTST/Eckart calculations are performed with the KiSThelP 2019 program (Canneaux et al., 2013).
The rate coefficients of each elementary pathway included in the initiation reactions of distinct SCIs with HCOOH are calculated in the temperature range of 273-400 K, as listed in Table S3-S6. As shown in Table S3, the total rate coefficients ktot-CH2OO of CH2OO reaction with HCOOH are in excess of 1.0 × 10-10 cm3 molecule-1 s-1, and they exhibit a slightly negative temperature dependence in the temperature range studied. ktot-CH2OOis estimated to be 1.4 × 10-10cm3 molecule-1 s-1 at 298 K, which is in good agreement with the experimental values reported by Welz et al. (2014) ([1.1 ± 0.1] × 10-10), Chung et al. (2019) ([1.4 ± 0.3] × 10-10), and Peltola et al. (2020) ([1.0 ± 0.03] × 10-10). k(TSent1) is approximately equal to ktot-CH2OO in the whole temperature range, and it decreases in the range of 1.7 × 10-10 (273 K) to 1.2 × 10-10 (400 K) cm3 molecule-1 s-1 with increasing temperature. k(TSent1) is several orders of magnitude greater than k(TSent2), k(TSent3) and k(TSent4) over the temperature range from 273 to 400 K. The result again shows that the barrierless 1,4 O-H insertion reaction is predominant.
Equivalent to the case of CH2OO reaction with HCOOH, the rate coefficient of each elementary pathway involved in the anti-CH3CHOO + HCOOH reaction also decreases with the temperature increasing (Table S4). This table shows that Entry 1 is kinetically favored over Entry 2, 3 and 4, and Entry 2 is competitive with Entry 4 in the range 273-400 K. Similar conclusion is also obtained from the results of the rate coefficients for the reactions of syn-CH3CHOO and (CH3)2COO with HCOOH that Entry 1 is the dominant pathway (Table S5-S6). It deserves mentioning that the competition of Entry 2 is significantly greater than that of Entry 4 in the syn-CH3CHOO + HCOOH and (CH3)2COO + HCOOH systems. At ambient temperature, the total rate coefficients of HCOOH reactions with anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO are estimated to be 5.9, 2.7 and 4.8 × 10-10 cm3 molecule-1 s-1, respectively, which are consistent with the prior experimental measurements of 5 ± 3, 2.5 ± 0.3 and 4.5 × 10-10 cm3 molecule-1 s-1 (Welz et al., 2014; Chung et al., 2019; Sipilä et al., 2014).
- Why are k(TSent2) and k(TSent4) decrease with increasing temperature as they both have positive energy barrier? (Table S3)
Response: Based on the Reviewer’s suggestion, the relevance descriptions on the negative temperature dependence of k(TSent2) and k(TSent4) in Table S3 have been added in the revised manuscript. The rate coefficients for the bimolecular reactions with the tight transition states are calculated by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction. The initiation reaction of CH2OO with HCOOH proceeds through four possible pathways, namely (1) 1,4 O-H insertion (Entry 1), (2) 1,2 O-H insertion (Entry 2), (3) C-H insertion (Entry 3), and (4) C=O cycloaddition (Entry 4). A schematic PES for the possible entrance pathways is drawn in Fig. 1. As shown in Fig. 1, the entrance pathway Entry2 consists of two elementary steps: (i) an intermediate IMent2 is formed via a barrierless process; (ii) then, it rearranges to the product Pent2 through a tight transition state TSent2. The whole reaction process can be described as Eq. (1):
(1)
Assuming the rapid equilibrium is established between the IMent2 and reactants. According to the steady-state approximation (SSA), the total rate coefficient is approximately expressed as Eq. (2):
(2)
The equilibrium constant Keq is written as Eq. (3):
(3)
where σ refers to the reaction symmetry number, QIM(T), QR1(T) and QR2(T) denote the partition functions of intermediate, reactants R1 and R2, which are equal to the multiplication of translational, rotational, vibrational and electronic partition functions (Q = QrotQvibQtransQelec). T is the temperature in Kelvin, R is the ideal gas constant, GR and GIM are the total Gibbs free energies of reactant and intermediate, respectively. Similar methodology is adopted to calculate the rate coefficient of each elementary pathway in Entry 4.
The calculated Keq-ent2, k2-ent2, and k(TSent2) (k(TSent2) = Keq-ent2 × k2-ent2) in Entry 2 are listed in Table S7. This table shows that Keq-ent2 significantly decreases with increasing temperature, and k2-ent2 increases as the temperature is increased. However, the decreased value in Keq-ent2 is greater than the increased value in k2-ent2 under the same temperature range. For example, Keq-ent2 deceases by a factor of 6.3 and k2-ent2 increases by a factor of 2.9 at 298 K compared with the values of Keq-ent2 and k2-ent2 at 273 K. It is therefore that k(TSent2) decreases with the temperature increasing. Similar conclusion is also obtained from the results of the rate coefficients in Entry 4 that k(TSent4) exhibits a negative temperature dependence in the temperature range studied (Table S8). The aforementioned results imply that k(TSent2) and k(TSent4) are mediated by the pre-reactive complexes IMent2 and IMent4 in the Entry 2 and 4 of the CH2OO + HCOOH reaction.
Table S7 Keq-ent2 (cm3 molecule-1), k2-ent2 (s-1) and k(TSent2) (cm3 molecule-1 s-1) in Entry 2 computed at different temperatures
T/K
Keq-ent2
k2-ent2
k(TSent2)
273
8.2 × 10-17
4.4 × 104
3.6 × 10-12
280
4.7 × 10-17
6.3 × 104
2.9 × 10-12
298
1.3 × 10-17
1.5 × 105
1.9 × 10-12
300
1.1 × 10-17
1.6 × 105
1.8 × 10-12
320
3.2 × 10-18
3.7 × 105
1.2 × 10-12
340
1.1 × 10-18
7.6 × 105
8.2 × 10-13
360
4.1 × 10-19
1.5 × 106
5.9 × 10-13
380
1.7 × 10-19
2.6 × 106
4.5 × 10-13
400
8.0 × 10-20
4.4 × 106
3.5 × 10-13
Table S8 Keq-ent4 (cm3 molecule-1), k2-ent4 (s-1) and k(TSent4) (cm3 molecule-1 s-1) in Entry 4 computed at different temperatures
T/K
Keq-ent4
k2-ent4
k(TSent4)
273
6.3 × 10-20
5.7 × 107
3.6 × 10-12
280
4.5 × 10-20
7.0 × 107
3.1 × 10-12
298
2.0 × 10-20
1.1 × 108
2.3 × 10-12
300
1.8 × 10-20
1.2 × 108
2.2 × 10-12
320
8.4 × 10-21
1.9 × 108
1.6 × 10-12
340
4.3 × 10-21
2.9 × 108
1.3 × 10-12
360
2.4 × 10-21
4.2 × 108
1.0 × 10-12
380
1.4 × 10-21
5.9 × 108
8.2 × 10-13
400
8.8 × 10-22
7.9 × 108
6.9 × 10-13
Corresponding descriptions have been added in the page 7 line 186-206 and page 12 line 315-326 of the revised manuscript:
The rate coefficients for the bimolecular reactions with the tight transition states are calculated by using the canonical transition state theory (CTST) along with one-dimensional asymmetric Eckart tunneling correction (Truhlar et al., 1996; Eckart, 1930). The CTST/Eckart calculations are performed with the KiSThelP 2019 program (Canneaux et al., 2013). As shown in Fig. 1, the entrance pathway Entry2 of R1R2COO reaction with HCOOH consists of two steps: (i) an intermediate IMent2 is formed via a barrierless process; (ii) then, it rearranges to the product Pent2 through a tight transition state TSent2. The whole reaction process can be described as Eq. (1):
(1)
Assuming the rapid equilibrium is established between the IMent2 and reactants. According to the steady-state approximation (SSA), the total rate coefficient is approximately expressed as Eq. (2) (Zhang et al., 2012):
(2)
The equilibrium constant Keq is written as Eq. (3):
(3)
where σ refers to reaction symmetry number, QIM(T), QR1(T) and QR2(T) denote the partition functions of intermediate, reactants R1 and R2, which are equal to the multiplication of translational, rotational, vibrational and electronic partition functions (Q = QrotQvibQtransQelec) (Mendes et al., 2014), T is the temperature in Kelvin, R is the ideal gas constant, GR and GIM are the total Gibbs free energies of reactant and intermediate, respectively.
The calculated Keq-ent2, k2-ent2, and k(TSent2) (k(TSent2) = Keq-ent2 × k2-ent2) in Entry 2 are listed in Table S7. This table shows that Keq-ent2 significantly decreases with increasing temperature, and k2-ent2 increases as the temperature is increased. However, the decreased value in Keq-ent2 is greater than the increased value in k2-ent2 under the same temperature range. For example, Keq-ent2 deceases by a factor of 6.3 and k2-ent2 increases by a factor of 2.9 at 298 K compared with the values of Keq-ent2 and k2-ent2 at 273 K. It is therefore that k(TSent2) decreases with the temperature increasing. Similar conclusion is also obtained from the results of the rate coefficients in Entry 4 that k(TSent4) exhibits a negative temperature dependence in the temperature range studied (Table S8). The aforementioned results imply that k(TSent2) and k(TSent4) are mediated by the pre-reactive complexes IMent2 and IMent4 in the Entry 2 and 4.
- The oligomerization reactions are highly dependent on the concentration of the monomers. Here the monomer are highly reactive SCIs and usually has very low concentration in the atmosphere. It seems that the high exothermicity of the oligomerization reaction results from the “stabilization” of SCIs in oligomerization. Also, the calculated free energies represent standard condition. Could the authors correct the Gibbs free energies by incorporating the atmospheric concentrations of SCIs (i.e., RTln(P/Pref)) to check whether this oligomerization is favored in the atmospheric conditions?
Response: Based on the Reviewer’s suggestion, the relative importance of distinct SCIs reactions with hydroperoxide esters and trace species (e.g., H2O, HCOOH and SO2) has been added in the revised manuscript. It is well known that the reactions with trace species are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). In the present study, the hydroperoxymethyl formate (HPMF) is selected as the model compound since it is the simplest hydroperoxide ester formed from the barrierless reaction of 1,4 O-H insertion of CH2OO into HCOOH. The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2, and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmosphereric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken et al., 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al, 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF) is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105 molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF) is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
Table 2 The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with HPMF, H2O, HCOOH and SO2 at the tropical forest environments
SCIs
Coreactant
[Coreactant]
(molecules cm-3)
k
(cm3 molecule-1 s-1)
keff
(s-1)
Reference
CH2OO
H2O
3.9-6.1 × 1017
< 1.5 × 10-15
5.9-9.2 × 102
Chao et al., (2015)
HCOOH
5.0-10.0 × 1010
[1.1 ± 0.1] × 10-10
5.5-11
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[3.9 ± 0.7] × 10-11
0.7-3.5
Welz et al., (2012)
HPMF
-
2.7 × 10-11
-
This work
anti-CH3CHOO
H2O
3.9-6.1 × 1017
[1.0 ± 0.4] × 10-14
3.9-6.1 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[5 ± 3] × 10-10
25.0-50.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[6.7 ± 1.0] × 10-11
1.1-6.0
Taatjes et al., (2013)
HPMF
-
3.3 × 10-10
-
This work
syn-CH3CHOO
H2O
3.9-6.1 × 1017
< 4.0 × 10-15
1.6-2.4 × 103
Taatjes et al., (2013)
HCOOH
5.0-10.0 × 1010
[2.5 ± 0.3] × 10-10
12.5-25.0
Welz et al., (2014)
SO2
1.7-9.0 × 1010
[2.4 ± 0.3] × 10-11
0.4-2.2
Taatjes et al., (2013)
HPMF
-
1.7 × 10-13
-
This work
(CH3)2COO
H2O
3.9-6.1 × 1017
< 1.5 × 10-16
58.5-91.5
Huang et al., (2015)
HCOOH
5.0-10.0 × 1010
4.5 × 10-10
22.5-45.0
Sipilä et al., (2014)
SO2
1.7-9.0 × 1010
1.3 × 10-10
2.2-11.7
Huang et al., (2015)
HPMF
-
2.2 × 10-11
-
This work
Corresponding descriptions have been added in the page 23 line 573-590 and page 24 line 591-610 of the revised manuscript:
It is of interest to assess whether the reactions of distinct SCIs with HPMF can compete well with the losses to reactions with trace species (e.g., H2O, HCOOH and SO2), because it is well known that the reactions with trace species are expected to be the dominant chemical sinks for SCIs in the atmosphere (Taatjes et al., 2013; Long et al., 2016). The reported concentrations of coreactant, the rate coefficients k, and the effective pseudo-first-order rate constants (keff = k[coreactant]) for distinct SCI reactions with H2O, HCOOH, SO2, and HPMF are summarized in Table 2. As seen in Table 2, the rate coefficient of a particular SCI reaction with trace species is strongly dependent on its structure. The methyl group substitution may alter the rate coefficient by several to tens of times. The atmospheric concentrations of H2O, HCOOH and SO2 in the tropical forest environments are measured to be 3.9-6.1 × 1017, 5.0-10 × 1010, and 1.7-9.0 × 1010 molecules cm-3, respectively (Vereecken, 2012). For the reactions of CH2OO with H2O, HCOOH, and SO2, the experimental rate coefficients are determined to be < 1.5 × 10-15, [1.1 ± 0.1] × 10-10, and [3.9 ± 0.7] × 10-11 cm3 molecule-1 s-1, respectively (Welz et al., 2012 and 2014; Chao et al., 2015), which translate into keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2) of 5.9-9.2 × 102, 5.5-11, and 0.7-3.5 s-1, respectively. The result reveals that the reaction of CH2OO with H2O is the most important bimolecular reaction. keff(CH2OO+HCOOH) is greater by a factor of 3-8 than keff(CH2OO+SO2), indicating that the reaction of CH2OO with HCOOH is favored over reaction with SO2. Similar conclusion is also obtained from the results of keff for the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with H2O, HCOOH and SO2 that SCIs reactions with H2O are faster than with HCOOH, which, in turn, are faster than with SO2.
According to the results shown in the Table 2, the room temperature rate coefficient for the reaction of CH2OO with HPMF is calculated to be 2.7 × 10-11 cm3 molecule-1 s-1. However, to the best of our knowledge, the atmospheric concentration of HPMF has not been reported up to now. If we assume that the concentration of HPMF is the same as that of HCOOH, keff(CH2OO+HPMF) is estimated to be 1.4-2.7 s-1, which is significantly lower than keff(CH2OO+H2O) and keff(CH2OO+HCOOH). keff(CH2OO+HPMF) is nearly identical to keff(CH2OO+SO2), indicating that the CH2OO + HPMF reaction is competitive with the CH2OO + SO2 system. Previous model-measurement studies have estimated the surface-level SCIs concentrations in the range of 1.0 × 104 to 1.0 × 105 molecules cm-3 (Khan et al., 2018; Novelli et al., 2017). If we assume that the concentration of HPMF is equal to that of SCIs, keff(CH2OO+HPMF) is calculated to be 2.7-27 × 10-7 s-1, which is several orders of magnitude lower than keff(CH2OO+H2O), keff(CH2OO+HCOOH) and keff(CH2OO+SO2). This result indicates that the reaction of CH2OO with HPMF is of less importance. Similar conclusion is also obtained from the reactions of anti-CH3CHOO, syn-CH3CHOO and (CH3)2COO with HPMF. Based on the above discussions, it can be concluded that the relative importance of carbonyl oxides reactions with hydroperoxide esters is significantly dependent on the concentrations of hydroperoxide esters. These reactions may play a certain role in the formation of organic new particle in some regions where low concentration of water vapour and high concentration of hydroperoxide esters occur.
- Additionally, it would be helpful if there is some estimation about how much the oligomerization process could contribute to the regional or global SOA.
Response: Sakamoto et al. (2013) investigated the ozonolysis of ethylene in a Teflon bag reactor, and found that CH2OO plays a critical role in the formations of oligomers and secondary organic aerosol (SOA) in the gas phase and particle phase. They proposed a possible formation mechanism for the oligomeric hydroperoxides, which includes the successive addition of CH2OO to hydroperoxides. Sadezky et al. (2008) studied the gas-phase ozonolysis of small enol ethers in a 570 l spherical glass reactor at atmospheric conditions in the absence of seed aerosol. They found that the oligomers composed of Criegee intermediate as the repeated chain unit are the main constituents of SOA. Zhao et al. (2015) studied the ozonolysis of trans-3-hexene in both the static Teflon chamber and glass flow reactor under different relative humidity conditions. It was found that the oligomers having Criegee intermediate as the chain unit are the dominant components of SOA. These findings may help in understanding the potential pathway for the formation of SOA in the atmosphere. However, to the best of our knowledge, the contribution of the oligomerization reaction composed of Criegee intermediate as the chain unit to SOA remains unknown. In the future work, we will adopt the combination of quantum chemistry and numerical simulation to estimate the contribution of oligomerization reaction to the regional and global SOA.
- Line 39, “with increasing the number of SCIs” is a bit confusing, it would be better to say “with increasing the number of SCIs added to the oligomer”.
Response: The sentence “with increasing the number of SCIs” has been replaced by “with increasing the number of SCIs added to the oligomer” in the revised manuscript.
- Line 491, “netative” should be “negative”.
Response: The word “netative” has been replaced by “negative” in the revised manuscript.
- Line 499, “neartly” should be “nearly”.
Response: The word “neartly” has been replaced by “nearly” in the revised manuscript.
References
Cabezas, C., and Endo, Y.: The Criegee intermediate-formic acid reaction explored by rotational spectroscopy, Phys. Chem. Chem. Phys., 21, 18059-18064, https://doi.org/10.1039/c9cp03001h, 2019.
Canneaux, S., Bohr, F., and Henon, E.: KiSThelP: a program to predict thermodynamic properties and rate constants from quantum chemistry results, J. Comput. Chem., 35, 82-93, https://doi.org/10.1002/jcc.23470, 2013.
Chao, W., Hsieh, J. T., Chang, C. H., and Lin, J. J. M.: Direct kinetic measurement of the reaction of the simplest Criegee intermediate with water vapor, Science, 347, 751-754, https://doi.org/10.1126/science.1261549, 2015.
Chhantyal-Pun, R., McGillen, M. R., Beames, J. M., Khan, M. A. H., Percival, C. J., Shallcross, D. E., and Orr-Ewing, A. J.: Temperature Dependence of the Rates of Reaction of Trifluoracetic Acid with Criegee Intermediates, Angew. Chem. Int. Ed., 129, 9172-9175, https://doi.org/10.1002/anie.201703700, 2017.
Chung, C. A., Su, J. W., and Lee, Y. P.: Detailed mechanism and kinetics of the reaction of Criegee intermediate CH2OO with HCOOH investigated via infrared identification of conformers of hydroperoxymethyl formate and formic acid anhydride, Phys. Chem. Chem. Phys., 21, 21445-21455, https://doi.org/10.1039/c9cp04168k, 2019.
Gilbert, R. G., and Smith, S. C.: Theory of unimolecular and recombination reactions; Blackwell Scientific: Carlton, Australia, 1990.
Glowacki, D. R., Liang, C. H., Morley, C., Pilling, M. J., and Robertson, S. H.: MESMER: an open-source master equation solver for multi-energy well reactions, J. Phys. Chem. A, 116, 9545-9560, https://doi.org/10.1021/jp3051033, 2012.
Huang, H. L., Chao, W., and Lin, J. J. M.: Kinetics of a Criegee intermediate that would survive high humidity and may oxidize atmospheric SO2, Proc. Natl. Acad. Sci. U.S.A., 112, 10857-10862, https://doi.org/ 10.1073/pnas.1513149112, 2015.
Khan, M. A. H., Percival, C. J., Caravan, R. L., Taatjes, C. A., and Shallcross, D. E.: Criegee intermediates and their impacts on the troposphere, Environ. Sci.: Processes Impacts, 20, 437-453, https://doi.org/10.1039/C7EM00585G, 2018.
Lin, X., Meng, Q., Feng, B., Zhai, Y., Li, Y., Yu, Y., Li, Z., Shan, X., Liu, F., Zhang, L., and Sheng, L.: Theoretical study on Criegee intermediate’s role in ozonolysis of acrylic acid, J. Phys. Chem. A, 123, 1929-1936, https://doi.org/10.1021/acs.jpca.8b11671, 2019.
Long, B., Bao, J. L., and Truhlar, D. G.: Atmospheric chemistry of Criegee intermediates: unimolecular reactions and reactions with water, J. Am. Chem. Soc., 138, 14409-14422, https://doi.org/10.1021/jacs.6b08655, 2016.
Long, B., Cheng, J. R., Tan, X. F., and Zhang, W. J.: Theoretical study on the detailed reaction mechanisms of carbonyl oxide with formic acid, J. Mol. Struc.: Theochem, 916, 159-167, https://doi.org/10.1016/j.theochem.2009.09.028, 2009.
Novelli, A., Hens, K., Ernest, C. T., Martinez, M., Nölscher, A. C., Sinha, V., Paasonen, P., Petäjä, T., Sipilä, M., Elste, T., Plass-Dülmer, C., Phillips, G. J., Kubistin, D., Williams, J., Vereecken, L., Lelieveld, J., and Harder, H.: Estimating the atmospheric concentration of Criegee intermediates and their possible interference in a FAGE-LIF instrument, Atmos. Chem. Phys., 17, 7807-7826, https://doi.org/10.5194/acp-17-7807-2017, 2017.
Peltola, J., Seal, P., Inkilä, A., and Eskola, A.: Time-resolved, broadband UV-absorption spectrometry measurements of Criegee intermediate kinetics using a new photolytic precursor: unimolecular decomposition of CH2OO and its reaction with formic acid, Phys. Chem. Chem. Phys., 22, 11797-11808, https://doi.org/10.1039/d0cp00302f, 2020.
Raghunath, P., Lee, Y. P., and Lin, M. C.: Computational chemical kinetics for the reaction of Criegee intermediate CH2OO with HNO3 and its catalytic conversion to OH and HCO, J. Phys. Chem. A, 121, 3871-3878, https://doi.org/10.1021/acs.jpca.7b02196, 2017.
Sadezky, A., Winterhalter, R., Kanawati, B., Rompp, A., Spengler, B., Mellouki, A., Bras, G. L., Chaimbault, P., and Moortgat, G. K.: Oligomer formation during gas-phase ozonolysis of small alkenes and enol ethers: new evidence for the central role of the Criegee Intermediate as oligomer chain unit, Atmos. Chem. Phys., 8, 2667-2699, https://doi.org/10.5194/acp-8-2667-2008, 2008.
Sakamoto, Y., Inomata, S., and Hirokawa, J.: Oligomerization reaction of the Criegee intermediate leads to secondary organic aerosol formation in ethylene ozonolysis, J. Phys. Chem. A, 117, 12912-12921, https://doi.org/10.1021/jp408672m, 2013.
Sipilä, M., Jokinen, T., Berndt, T., Richters, S., Makkonen, R., Donahue, N. M., Mauldin Iii, R. L., Kurtén, T., Paasonen, P., Sarnela, N., Ehn, M., Junninen, H., Rissanen, M. P., Thornton, J., Stratmann, F., Herrmann, H., Worsnop, D. R., Kulmala, M., Kerminen, V. M., and Petäjä, T.: Reactivity of stabilized Criegee intermediates (sCIs) from isoprene and monoterpene ozonolysis toward SO2 and organic acids, Atmos. Chem. Phys., 14, 12143-12153, https://doi.org/10.5194/acp-14-12143-2014, 2014.
Taatjes, C. A., Welz, O., Eskola, A. J., Savee, J. D., Scheer, A. M., Shallcross, D. E., Rotavera, B., Lee, E. P. F., Dyke, J. M., Mok, D. K. W., Osborn, D. L., and Percival, C. J.: Direct measurements of conformer-dependent reactivity of the Criegee intermediate CH3CHOO, Science, 340, 177-180, https://doi.org/10.1126/science.1234689, 2013.
Vereecken, L., Harder, H., and Novelli, A.: The reaction of Criegee intermediates with NO, RO2, and SO2, and their fate in the atmosphere, Phys. Chem. Chem. Phys., 14, 14682-14695, https://doi.org/10.1039/c2cp42300f, 2012.
Vereecken, L.: The reaction of Criegee intermediates with acids and enols, Phys. Chem. Chem. Phys., 19, 28630-28640, https://doi.org/10.1039/c7cp05132h, 2017.
Welz, O., Eskola, A. J., Sheps, L., Rotavera, B., Savee, J. D., Scheer, A. M., Osborn, D. L., Lowe, D., Booth, A. M., Xiao, P., Khan, M. A. H., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Rate coefficients of C(1) and C(2) Criegee intermediate reactions with formic and acetic Acid near the collision limit: direct kinetics measurements and atmospheric implications, Angew. Chem. Int. Ed., 53, 4547-4550, https://doi.org/10.1002/anie.201400964, 2014.
Welz, O., Savee, J. D., Osborn, D. L., Vasu, S. S., Percival, C. J., Shallcross, D. E., and Taatjes, C. A.: Direct kinetic measurements of Criegee intermediate (CH2OO) formed by reaction of CH2I with O2, Science, 335, 204-207, https://doi.org/10.1126/science.1213229, 2012.
Zhao, Y., Wingen, L. M., Perraud, V., Greaves, J., and Finlayson-Pitts, B. J.: Role of the reaction of stabilized Criegee intermediates with peroxy radicals in particle formation and growth in air, Phys. Chem. Chem. Phys., 17, 12500-12514, https://doi.org/10.1039/c5cp01171j, 2015.
-
AC3: 'Reply on RC3', Y. Huang, 25 Aug 2022