Reactivity of Criegee Intermediates toward Carbon Dioxide - The

Dec 18, 2017 - Recent theoretical work by Kumar and Francisco suggested that the high reactivity of Criegee intermediates (CIs) could be utilized for ...
1 downloads 8 Views 842KB Size
Subscriber access provided by UNIV OF DURHAM

Letter

The Reactivity of Criegee Intermediates Towards Carbon Dioxide Yen-Hsiu Lin, Kaito Takahashi, and Jim Jr-Min Lin J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.7b03154 • Publication Date (Web): 18 Dec 2017 Downloaded from http://pubs.acs.org on December 20, 2017

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

The Journal of Physical Chemistry Letters is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

The Reactivity of Criegee Intermediates towards Carbon Dioxide Yen-Hsiu Lin,1,2 Kaito Takahashi,1 and Jim Jr-Min Lin1,2* 1

Institute of Atomic and Molecular Sciences, Academia Sinica, Taipei 10617, Taiwan 2

Department of Chemistry, National Taiwan University, Taipei 10617, Taiwan *To whom correspondence should be addressed: [email protected]

Abstract A recent theoretical work by Kumar and Francisco suggested that the high reactivity of Criegee intermediates (CIs) could be utilized for designing efficient carbon capture technologies. Because the anti-CH3CHOO+CO2 reaction has the lowest barrier in their study, we chose to investigate it experimentally. We probed anti-CH3CHOO with its strong UV absorption at 365 nm and measured the rate coefficient to be ≤ 2×10−17 cm3 molecule−1 s−1 at 298 K, which is consistent with our theoretical value of 2.1×10−17 cm3 molecule−1 s−1 at QCISD(T)/CBS//B3LYP/6-311+G(2d,2p) level, but inconsistent with their results obtained at M06-2X/aug-cc-pVTZ level, which tends to underestimate the barrier heights. The experimental result indicates that the reaction of a Criegee intermediate with atmospheric CO2 (400 ppmv) would be inefficient (keff < 0.2 s−1), and cannot compete with other decay processes of Criegee intermediates like reactions with water vapor (~103 s−1) or thermal decomposition (~102 s−1).

TOC graphic

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The growing level of greenhouse gases (primarily CO2) in our atmosphere is undoubtedly a very important issue in this century. Very recently, Kumar and Francisco calculated the reactions of CO2 with various Criegee intermediates and proposed that the high reactivity of Criegee intermediates could be utilized for designing carbon capture technologies.1 While fast reactions of Criegee intermediates with a number of atmospheric species, like SO2, NO2, water and water dimer, organic and inorganic acids, etc., have been reported,2–7 the reaction of CO2 with a Criegee intermediate might not be expected to be efficient, due to the inertness of CO2. The motivation of this work is to examine the above counterintuitive idea. Kumar and Francisco1 found that the barrier height of a Criegee intermediate reaction with CO2 depends strongly on the substitution groups (with H, F, Cl, CH3, and/or CF3 substitutions). While the syn-substituents tend to raise the reaction barrier (we assume this is due to steric hindrance), the anti-substituents lower the reaction barrier. For anti-Criegee intermediates, an inverse correlation between the barrier height and the charge separation of the Criegee intermediate is reported; this finding offers a handle for designing an efficient carbon fixation method.1 Among their studied Criegee intermediates, the anti-CHFOO and anti-CH3CHOO have the largest charge separation (the most positive Criegee carbons and the most negative Criegee terminal oxygens), making them the most reactive towards CO2.1 To examine the reactivity of Criegee intermediates towards CO2, we chose to investigate the kinetics of anti-CH3CHOO reaction with CO2. A mixture of syn-CH3CHOO and anti-CH3CHOO was prepared using a well-established reaction of CH3CHI + O2 → CH3CHOO + I.8 The concentrations of syn-CH3CHOO and anti-CH3CHOO were monitored in real time by their strong UV absorption at 365 nm.9–11 To separate the signals of syn-CH3CHOO and anti-CH3CHOO, we added CH3OH to scavenge anti-CH3CHOO. We also performed ab initio calculations on related reaction paths at various levels of theory to compare with the experimental results.

2

ACS Paragon Plus Environment

Page 2 of 16

Page 3 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Figure 1. Time traces of the absorbance change (with respect to that before the photolysis pulse) of the CH3CHI2/O2 photolysis system at various [CO2]. The time zero is set as the photolysis laser pulse. The data in this Figure are Exp. 2d; see Supporting Information for details.

In Figure 1, the time traces of the absorbance change show the formation and decay of CH3CHOO. The absorption cross section of CH3CHOO is on the order of 10−17 cm2 at 365 nm,9–11 much larger than those of other involved species like CH3CHI2 ( 0 have been shifted down vertically by 1, 2, or 3 units. Note that the [CO2]=0 trace is not shifted; the vertical scale of −8 would correspond to a value of e−8 = 3×10−4 for the peak absorbance. The lines are linear fit to the data from 0.3 to 9 ms. The data in this Figure are Exp. 2d; see Supporting Information for details.

The slope of the plot in Figure 2 would correspond to the effective 1st-order rate coefficient kobs, which includes the decay rate independent of CO2 (k0) and the decay rate due to CO2 (kCO2[CO2]) as indicated in Equations (1) and (2).



d [CI] = k0 [CI] + kCO 2 [CO 2 ][CI] = (k0 + kCO 2 [CO 2 ])[CI] = kobs [CI] dt

− 1 d [CI] d ln[CI] =− = kobs [CI] dt dt

(1)

(2)

Again, the slope does not show significant variation for different CO2 concentrations (Figure 2). A linear fit to the data gives estimated values of kobs. When CO2 is absent, kobs = k0 ≅ 115 s−1. This is mostly due to the reactions of Criegee intermediates with radical species in the reactor, including iodine atoms, Criegee intermediates (in this case, both syn and anti forms), OH radicals (from the decomposition of syn-CH3CHOO), etc.14 Even under a high concentration of CO2, the increase in kobs is only about 20 to 30 s−1. 6

ACS Paragon Plus Environment

Page 7 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Figure 3. Effective decay rate of anti-CH3CHOO as a function of [CO2]. The error bar indicates one standard deviation of the data points. The solid line is the linear fit to all the data points. The upper penal plots the data of Exp. 1 and the lower for Exp. 2. The shapes of the symbols indicate the data acquisition sequence ([CO2] was varied within each sequence), in the order of square, circle, up triangle, and down triangle.

Figure 3 shows a summary of the resulted kobs plotted as a function of CO2 concentration. The best linear fit to the data gives slightly positive slopes. The values with 1σ error bar are (0.48±0.32)×10−17 and (1.25±0.34)×10−17 cm3 molecule−1 s−1 for Exp. 1 and Exp. 2, respectively. However, a horizontal line, signifying zero reactivity, is still acceptable. Thus, we set an upper limit of 2×10−17 cm3 molecule−1 s−1 for the rate coefficient of anti-CH3CHOO reaction with CO2. This result is consistent with the theoretical value of 2.1×10−17 cm3 molecule−1 s−1 obtained at QCISD(T)/CBS//B3LYP/6-311+G(2d,2p) level (Table 1).

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

It is not trivial to measure the slow rate of the anti-CH3CHOO reaction with CO2. This experiment requires a low concentration of CH3CHOO (~ 2×1011 cm−3). At higher concentrations, anti-CH3CHOO would react quickly with itself, syn-CH3CHOO, and other radical species like I atoms and OH radicals, resulting in a shorter lifetime14,16 (then it would be even harder to observe the small rate increase due to CO2). As shown in Figure 3, the variation of k0 is less than 20 s−1 in our system. While this is already a small value, the effect of adding CO2 is only about the same magnitude. We may estimate the effect of atmospheric CO2 (~400 ppmv, 1×1016 cm−3) toward a Criegee intermediate. With kCO2 ≤ 2×10−17 cm3 molecule−1 s−1 for anti-CH3CHOO, the effective 1st-order decay rate coefficient is kCO2[CO2] ≤ 0.2 s−1. Since other Criegee intermediates would be less reactive than anti-CH3CHOO, their effective rates would be even smaller. The main decay pathway of a Criegee intermediate in the atmosphere depends on its structure.3 The anti-Criegee intermediate and CH2OO would mainly be consumed by the reaction with water vapor (monomer and dimer) with effective rates faster than 103 s−1,17–19 the syn and double alkyl substituted Criegee intermediates would decay mainly by thermal decomposition with an effective rate faster than 102 s−1.16,20,21 It is clear that for anti-Criegee intermediates and CH2OO, the reaction with CO2 cannot compete with the water reactions. Considering that the syn-substituents tend to increase the barrier height for the CO2 reaction by a few kcal mol−1, the reactions of syn and dialkyl Criegee intermediates with CO2 would be slowed down by several orders of magnitude and thus, they would not compete with the fast thermal decomposition reactions. Therefore, we may conclude that the effective CO2 reaction rate for a Criegee intermediate is much slower than its normal decay pathways in the troposphere. Another perspective is the following. The concentrations of Criegee intermediates in the atmosphere are expected to be very low due to their fast decay processes; the impact of Criegee intermediate reaction toward CO2 would be negligible due to small kCO2[CI]. Novelli et al. have estimated an average concentration of Criegee intermediate to be 5×104 cm−3 (with an order of magnitude uncertainty).22 Even if we assume [CI] = 5×105 cm−3 (the upper limit given by Novelli et al.), the value of kCO2[CI] would be 3×103 years. For anti-CH3CHOO reaction with CO2, the energy of the transition state is 0.1 kcal mol−1 lower than that of the separated reactants at M06-2X/aug-cc-pVTZ level of theory.1 However, calculation at a 8

ACS Paragon Plus Environment

Page 8 of 16

Page 9 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

higher level (see Figure 4 and Table 1), raises the transition state energy by ~3 kcal mol−1, indicating a much lower reactivity. A similar trend is observed for Criegee intermediate reactions with CH3OH, suggesting that M06-2X/aug-cc-pVTZ tends to underestimate the barrier heights. This has also been observed for several similar reactions as those with H2O, H2S, enols, and alkenes.23–25 In fact, for the CH2OO+CO2 reaction, Kumar and Francisco reported that CCSD(T)/aug-cc-pVTZ gives a barrier which is 2.7 kcal mol−1 higher than that by M06-2X.1 Detailed discussion on the systematic deviations of the barrier heights by using various theoretical methods will be published elsewhere.

Figure 4. Schematic potential energy diagram for CO2 reactions with selected Criegee intermediates. The energy zero is the energy of the separated reactants. The calculation level is QCISD(T)/CBS//B3LYP/6-311+G(2d,2p) with zero-point energy correction at B3LYP/6-311+G(2d,2p).

Although Criegee intermediates are found to be very reactive towards a number of atmospheric gases, including SO2, NO2, organic and inorganic acids, water vapor, etc., the reactivity towards CO2 is measured to be quite limited, indicating no practical impact toward carbon fixation. Evaluation of different theoretical methods suggests that one should be cautious about the systematic uncertainty of different quantum chemistry methods in estimating the barrier heights.

Experimental Method Most of the experimental setup has been described in our earlier works.10,16-19 Thus, only the most 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 16

relevant parts are given here. Methyl substituted Criegee intermediate (acetaldehyde oxide, CH3CHOO) was synthesized by using a well-established method of CH3CHI + O2 → CH3CHOO + I in a flow reactor.8–10 The CH3CHI radical was prepared by pulsed photolysis of CH3CHI2 precursor (Aldrich, ≥98%) at 248 nm (KrF excimer laser, Coherent, CompExPro 205). The concentration of CH3CHOO was monitored by its intense UV absorption at 365 nm using a continuous light source (Energetiq, EQ-99), a bandpass filter (Edmund Optics, 65130, 10nm OD4 band pass filter at 365 nm) and a balanced photodiode detector (Thorlabs, PDB450A). The output of the detector was recorded by a digital oscilloscope (LeCroy HDO4034, 12-bit vertical resolution). The experimental temperature was controlled to be 298±1 K; the total pressure was 250 Torr. See Table S1. When high pressure of CO2 (PCO2 ≤ 60 Torr) was used, we observed that the yield of CH3CHOO was slightly decreased (~ 5%). This is due to the fact that CO2 has a better efficiency in collisional cooling compared to N2, such that more CH3CHIOO adduct would form at high PCO2. To compensate this effect, we added He, which is less efficient in collisional cooling, as part of the buffer gas. We found the yield of CH3CHOO could maintain a constant at various PCO2 under the condition of PHe = 3PCO2 (N2 balance). See Figure S6.

Computational Method We calculated the reaction paths for the relevant reactions at B3LYP26,27/6-311+G(2d,2p)28,29 and M06-2X30/aug-cc-pVTZ31,32 levels of theory. Furthermore, we used the vibrational second order perturbation

theory

(VPT2)33,34

to

obtain

the

anharmonic

vibrational

frequencies

at

B3LYP/6-311+G(2d,2p). Using these B3LYP geometries, we performed complete basis set (CBS) extrapolation35 using Dunning’s aug-cc-pVXZ (X = D, T, Q) basis sets with QCISD(T). All density functional theory calculations were done with the Gaussian09 program36 while all QCISD(T) calculations were performed using the MOLPRO program37. The XYZ geometries are given in Tables S6 and S7. More details concerning the accuracy of the calculation are given in the SI. The bimolecular reaction rate coefficients were obtained using conventional transition state theory, assuming thermal equilibrium between the reactants and the pre-reactive complex as well as the steady-state approximation for the activated complex using the Multiwell program38. For the reaction with CO2 we obtained only one path, but for the reaction with CH3OH, similar to previous studies on H2O and H2S reactions, we obtained two pathways. We note that Anglada and Sole have tested the effect of the variational transition state for Criegee intermediate reaction with water and reported effects of a factor of 2 at most.39 Furthermore, previous studies have also mentioned that multireference effects may cause slight errors on the barrier energies obtained by single reference 10

ACS Paragon Plus Environment

Page 11 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

methods.19,40 We estimated the theoretical rate coefficient to have an uncertainty within a factor of 6, after considering the errors in the quantum chemistry energies, partition function calculation, and tunneling corrections. See Supporting Information for more details.

Supporting Information Available: Experimental and computational details; Figures S1-S7; Tables S1-S7. This material is available free of charge via the Internet at http://pubs.acs.org.

Acknowledgements: This work was supported by Academia Sinica and Ministry of Science and Technology, Taiwan (MOST 106-2113-M-001-026-MY3; MOST 106-2113-M-001-007-MY3).

References: (1)

Kumar, M.; Francisco, J. S. Reactions of Criegee Intermediates with Non-Water Greenhouse Gases: Implications for Metal Free Chemical Fixation of Carbon Dioxide. J. Phys. Chem. Lett.

2017, 8, 4206–4213. (2)

Welz, O.; Savee, J. D.; Osborn, D. L.; Vasu, S. S.; Percival, C. J.; Shallcross, D. E.; Taatjes, C. A. Direct Kinetic Measurements of Criegee Intermediate (CH2OO) Formed by Reaction of CH2I with O2. Science 2012, 335, 204–207.

(3)

Lin, J. J.-M.; Chao, W. Structure-Dependent Reactivity of Criegee Intermediates Studied with Spectroscopic Methods. Chem. Soc. Rev. 2017, 46, 7483-7497.

(4)

Osborn, D. L.; Taatjes, C. A. The Physical Chemistry of Criegee Intermediates in the Gas Phase. Int. Rev. Phys. Chem. 2015, 34, 309–360.

(5)

Lee, Y.-P. Perspective: Spectroscopy and Kinetics of Small Gaseous Criegee Intermediates. J. Chem. Phys. 2015, 143, 20901.

(6)

Welz, O.; Eskola, A. J.; Sheps, L.; Rotavera, B.; Savee, J. D.; Scheer, A. M.; Osborn, D. L.; Lowe, D.; Murray Booth, A.; Xiao, P.; et al. Rate Coefficients of C1 and C2 Criegee 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Intermediate Reactions with Formic and Acetic Acid near the Collision Limit: Direct Kinetics Measurements and Atmospheric Implications. Angew. Chemie - Int. Ed. 2014, 53, 4547–4550. (7)

Foreman, E. S.; Kapnas, K. M.; Murray, C. Reactions between Criegee Intermediates and the Inorganic Acids HCl and HNO3: Kinetics and Atmospheric Implications. Angew. Chemie - Int. Ed. 2016, 55, 10419–10422.

(8)

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.; et al. Direct Measurements of Conformer-Dependent Reactivity of the Criegee Intermediate CH3CHOO. Science 2013, 340, 177–180.

(9)

Sheps, L.; Scully, A. M.; Au, K. UV Absorption Probing of the Conformer-Dependent Reactivity of a Criegee Intermediate CH3CHOO. Phys. Chem. Chem. Phys. 2014, 16, 26701– 26706.

(10)

Lin, L.-C.; Chao, W.; Chang, C.-H.; Takahashi, K.; Lin, J. J. Temperature Dependence of the Anti-CH3CHOO Reaction with Water Vapor. Phys. Chem. Chem. Phys. 2016, 18, 28189– 28197.

(11)

Smith, M. C.; Ting, W.-L.; Chun-Hung, C.; Takahashi, K.; Boering, K. A.; Lin, J. J.-M. The UV Absorption Spectrum of the C2 Criegee Intermediate CH3CHOO. J. Chem. Phys. 2014, 141, 74302.

(12)

Sander, S. P.; Friedl, R. R.; Barker, J. R.; Golden, D. M.; Kurylo, M. J.; Wine, P. H.; Abbat, J. P. D.; Kolb, C. . E.; Moortgat, G. K.; Huie, R. E.; et al. Chemical Kinetics and Photochemical Data for Use in Atmospheric Studies Evaluation Number 17. JPL Publ. 10-6 2011, 1–153.

(13)

Ting, W.-L.; Chen, Y.-H.; Chao, W.; Smith, M. C.; Lin, J. J.-M. The UV Absorption Spectrum of the Simplest Criegee Intermediate CH2OO. Phys. Chem. Chem. Phys. 2014, 16, 10438– 12

ACS Paragon Plus Environment

Page 12 of 16

Page 13 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

10443. (14)

Ting, W. L.; Chang, C. H.; Lee, Y. F.; Matsui, H.; Lee, Y. P.; Lin, J. J. M. Detailed Mechanism of the CH2I + O2 Reaction: Yield and Self-Reaction of the Simplest Criegee Intermediate CH2OO. J. Chem. Phys. 2014, 141, 104308.

(15)

McGillen, M. R.; Curchod, B. F. E.; Chhantyal-Pun, R.; Beames, J. M.; Watson, N.; Khan, M. A. H.; McMahon, L.; Shallcross, D. E.; Orr-Ewing, A. J. Criegee Intermediate–Alcohol Reactions, A Potential Source of Functionalized Hydroperoxides in the Atmosphere. ACS Earth Sci. Chem. 2017, DOI: 10.1021/ acsearthspacechem.7b00108.

(16)

Smith, M. C.; Chao, W.; Takahashi, K.; Boering, K. A.; Lin, J. J.-M. Unimolecular Decomposition Rate of the Criegee Intermediate (CH3)2COO Measured Directly with UV Absorption Spectroscopy. J. Phys. Chem. A 2016, 120, 4789–4798.

(17)

Chao, W.; Hsieh, J.; Chang, C.; Lin, J. J. Direct Kinetic Measurement of the Reaction of the Simplest Criegee Intermediate with Water Vapor. Science 2015, 347, 751–754.

(18)

Smith, M. C.; Chang, C.-H.; Chao, W.; Lin, L.-C.; Takahashi, K.; Boering, K. A.; Lin, J. J.-M. Strong Negative Temperature Dependence of the Simplest Criegee Intermediate CH2OO Reaction with Water Dimer. J. Phys. Chem. Lett. 2015, 6, 2708–2713.

(19)

Lin, L.-C.; Chang, C.-H.; Chao, W.; Smith, M. C.; Chang, C.-H.; Lin, J. J.; Takahashi, K. Competition between H2O and (H2O)2 Reactions with CH2OO/CH3CHOO. Phys. Chem. Chem. Phys. 2016, 18, 4557–4568.

(20)

Fang, Y.; Liu, F.; Barber, V. P.; Klippenstein, S. J.; Mccoy, A. B.; Lester, M. I. Deep Tunneling in the Unimolecular Decay of CH3CHOO Criegee Intermediates to OH Radical Products. J. Chem. Phys. 2016, 145, 234308.

(21)

Fang, Y.; Barber, V. P.; Klippenstein, S. J.; McCoy, A. B.; Lester, M. I. Tunneling Effects in 13

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

the Unimolecular Decay of (CH3)2COO Criegee Intermediates to OH Radical Products. J. Chem. Phys. 2017, 146, 134307. (22)

Novelli, A.; Hens, K.; Ernest, C. T.; Martinez, M.; Nölscher, A. C.; Sinha, V.; Paasonen, P.; Petäjä, T.; Sipilä, M.; Elste, T.; et al. Estimating the Atmospheric Concentration of Criegee Intermediates and Their Possible Interference in a FAGE-LIF Instrument. Atmos. Chem. Phys.

2017, 17, 7807–7826. (23)

Vereecken, L.; Harder, H.; Novelli, A. The Reactions of Criegee Intermediates with Alkenes, Ozone, and Carbonyl Oxides. Phys. Chem. Chem. Phys. 2014, 16, 4039–4049.

(24)

Smith, M. C.; Chao, W.; Kumar, M.; Francisco, J. S.; Takahashi, K.; Lin, J. J.-M. Temperature-Dependent Rate Coefficients for the Reaction of CH2OO with Hydrogen Sulfide. J. Phys. Chem. A 2017, 121, 938–945.

(25)

Vereecken, L. The Reaction of Criegee Intermediates with Acids and Enols. Phys. Chem. Chem. Phys. 2017, 19, 28630–28640.

(26)

Becke, A. D. Density-Functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652.

(27)

Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789.

(28)

Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A. Self-Consistent Molecular Orbital Methods. XX. A Basis Set for Correlated Wave Functions. J. Chem. Phys. 1980, 72, 650–654.

(29)

Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-Consistent Molecular Orbital Methods 25. Supplementary Functions for Gaussian Basis Sets. J. Chem. Phys. 1984, 80, 3265–3269.

(30)

Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and 14

ACS Paragon Plus Environment

Page 14 of 16

Page 15 of 16 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry Letters

Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215–241. (31)

Dunning, T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023.

(32)

Kendall, R. A.; Dunning, T. H.; Harrison, R. J. Electron Affinities of the First-Row Atoms Revisited. Systematic Basis Sets and Wave Functions. J Chem Phys 1992, 96, 6796–6806.

(33)

Barone, V. Anharmonic Vibrational Properties by a Fully Automated Second-Order Perturbative Approach. J. Chem. Phys. 2005, 122, 1–10.

(34)

Barone, V.; Biczysko, M.; Bloino, J. Fully Anharmonic IR and Raman Spectra of Medium-Size Molecular Systems: Accuracy and Interpretation. Phys. Chem. Chem. Phys.

2013, 16, 1759–1787. (35)

Peterson, K. A.; Woon, D. E.; Dunning, T. H. Benchmark Calculations with Correlated Molecular Wave Functions. IV. The Classical Barrier Height of the H+H2→H2+H Reaction. J. Chem. Phys. 1994, 100, 7410–7415.

(36)

Gaussian09; Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; et al. Gaussian09 A.02. Gaussian, Inc.: Wallingford CT 2009.

(37)

MOLPRO; Werner, H.-J.; Knowles, P. J.; Manby, F. R.; Schütz, M.; Celani, P.; Knizia, G.; Korona, T.; Lindh, R.; Mitrushenkov, A.; et al. MOLPRO Ver 2010.1. 2011.

(38)

Barker, J. R. Reaction Systems . I . MultiWell Computer Program Suite. Int. J. Chem. Kinet.

2000, 232. (39)

Anglada, J. M.; Sole, A. Impact of Water Dimer on the Atmospheric Reactivity of Carbonyl Oxides. Phys. Chem. Chem. Phys. 2016, 18, 17698–17712. 15

ACS Paragon Plus Environment

The Journal of Physical Chemistry Letters 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(40)

Kjaergaard, H. G.; Kurtén, T.; Nielsen, L. B.; Jørgensen, S.; Wennberg, P. O. Criegee Intermediates React with Ozone. J. Phys. Chem. Lett. 2013, 4, 2525–2529.

16

ACS Paragon Plus Environment

Page 16 of 16