Reactions of Criegee Intermediates with Non-Water Greenhouse

Aug 21, 2017 - Considering that there is a strong nexus between carbon dioxide levels in the atmosphere and global warming, the high reactivity of Cri...
0 downloads 14 Views 1MB Size
Subscriber access provided by UNIVERSITY OF ADELAIDE LIBRARIES

Letter

Reactions of Criegee Intermediates with Non-Water Greenhouse Gases: Implications For Metal Free Chemical Fixation of Carbon Dioxide Manoj Kumar, and Joseph S. Francisco J. Phys. Chem. Lett., Just Accepted Manuscript • DOI: 10.1021/acs.jpclett.7b01762 • Publication Date (Web): 21 Aug 2017 Downloaded from http://pubs.acs.org on August 22, 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 27

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

Reactions of Criegee Intermediates with Non-Water Greenhouse Gases: Implications For Metal Free Chemical Fixation of Carbon Dioxide Manoj Kumar, and Joseph S. Francisco*

Department of Chemistry, University of Nebraska-Lincoln, Lincoln, Nebraska, 68588, United States.

AUTHOR INFORMATION Corresponding Author *Email: [email protected]

ACS Paragon Plus Environment

1

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

ABSTRACT. High level theoretical calculations suggest that Criegee intermediate preferably interacts with carbon dioxide compared to other two greenhouse gases, nitrous oxide and methane. The results also suggest that the interaction between Criegee intermediates and carbon dioxide involves a cycloaddition reaction, which results in the formation of a cyclic carbonatetype adduct with a barrier of 6.0-14.0 kcal/mol. These results are in contrast to previous assumption that the reaction occurs barrierlessly. The subsequent decomposition of the cyclic adduct into formic acid and carbon dioxide follows both concerted and stepwise mechanisms. The latter mechanism has been overlooked in a previous study. Under formic acid catalysis, the concerted decomposition of the cyclic carbonate may be favored under tropospheric conditions. Considering that there is a strong nexus between carbon dioxide levels in atmosphere and global warming, the high reactivity of Criegee intermediate could be utilized for designing efficient carbon capture technologies.

TOC GRAPHICS

Criegee reactions play an important role in influencing the tropospheric budgets of hydroxyl radicals, organic acids, hydroperoxides, nitrates, sulfates and particulate material.1-8 Criegee

ACS Paragon Plus Environment

2

Page 2 of 27

Page 3 of 27

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

intermediates are zwitter ionic/biradical species that mediate the ozone-alkene reactions.9 Though the involvement of Criegee intermediates in ozonolysis reactions was proposed in 1949 by Rudolf Criegee,10 their direct detection in gas-phase was eluded until Taatjes and co-workers produced one and two carbon Criegee intermediates from the reactions of α-iodoalkyl radicals with O2 in a flow reactor and unambiguously detected them by tunable synchrotron photoionization mass spectrometery.6,7 This has opened up new avenues for directly studying the reaction kinetics of Criegee intermediates with several important tropospheric species including SO2 and NO2. Besides from its broad tropospheric profile, the Criegee intermediate in the condensed-phase has been used to prepare carbonyl compounds via a synthetic route that avoids the hazards associated with the formation and isolation of ozonides or other peroxide products.11,12 Because of its wide importance, the unimolecular and bimolecular reactions of Criegee intermediate as well as factors influencing these reactions have been extensively investigated by means of experimental and theoretical means.13 Surprisingly, the possible interaction of Criegee intermediate with non-water greenhouse gases, and its implications for the global environment have never been discussed in literature. Greenhouse gases cause global warming and consequently climate change.14 Apart from water vapor, carbon dioxide (CO2), nitrous oxide (N2O) and methane (CH4) are the most prominent greenhouse gases in the Earth’s atmosphere. In 21st century, atmospheric concentrations of these gases are at levels that are unparalleled in nearly 800,000 years.15 Concentrations of CO2, CH4 and N2O have all shown significant increases since 1750 (40%, 150% and 20%, respectively). CO2 levels are rising at the fastest observed decadal rate of change (2.0 ± 0.1 ppm/yr) for 2002–2011. According to the UN World Meteorological organization,16 the current global levels of CO2 are above 400 parts per million

ACS Paragon Plus Environment

3

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

(ppm) for the first time in 3 million years, and are 144 percent of pre-industrial levels of 278 ppm. After being stable for almost one decade since the late 1990s, CH4 levels have started increasing since 2007. N2O levels have gradually increased at a rate of 0.73 ± 0.03 ppb/yr over the last three decades. Clearly, there is global urgency to better understand chemistries involving these greenhouse gases and find efficient solutions to minimize their impact on global climate. Herein we examined the interaction of the simplest Criegee intermediate, CH2OO with three greenhouse gases, CO2, N2O, and CH4. The hybrid M06-2X17 DFT functional along with the aug-cc-pVTZ18 basis set was employed for studying the gas-phase chemistries involving Criegee intermediate and greenhouse gases. The energies were further improved by performing single point calculations at the coupled cluster single and double substitution method with a perturbative treatment of triple excitations (CCSD(T))19/aug-cc-pVTZ level. All calculations were performed with Gaussian0920 quantum chemistry package. The structural coordinates of key species are provided in the Supplemental Information. See Methods section for additional computational details. Criegee intermediate forms the strongest and the weakest complexes with CO2 (∆E = -5.2) and CH4 (∆E = -1.2), respectively. N2O can interact with the Criegee carbon via either its N end or O end (Figure S1). The calculations show that the O end approach is more favorable as it accounts for 1.1 kcal/mol additional stabilization of the CH2OO⋅⋅N2O complex (∆E = -4.3 kcal/mol). The calculated equilibrium constants (Keq) for these Criegee complexes in the atmospherically relevant temperature range of 200-300 K are given in Figure 1 and Table S1. The equilibrium constants have been evaluated from the relative energy calculated at the CCSD(T)/aug-cc-pVTZ level of theory and partition functions computed at M06-2X/aug-cc-

ACS Paragon Plus Environment

4

Page 4 of 27

Page 5 of 27

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. The CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ calculated equilibrium constants (molecule cm-3) for the complexes of the simplest Criegee intermediate, CH2OO with three greenhouse gases, carbon dioxide (green), nitrous oxide (blue), and methane (red), respectively from 200 to 300 K. Note that nitrous oxide can bind to Criegee carbon via either its N end or O end (see Figure S1). However, the O end binding of N2O is more favorable and is, thus, considered for the equilibrium constant calculations.

pVTZ level of theory. The CH2OO⋅⋅CO2 complex has the larger Keq values than CH2OO⋅⋅N2O or CH2OO⋅⋅CH4 complex at all temperatures considered. This is in line with the trends predicted for their binding energies. Temperature has an interesting effect on the Keq values of the

ACS Paragon Plus Environment

5

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

CH2OO⋅⋅CH4 complex; below 260 K, the CH2OO⋅⋅CH4 complex has smaller Keq values than CH2OO⋅⋅N2O, but at higher temperatures, it has relatively larger Keq values. Considering that CO2 is the second most abundant greenhouse gas in the Earth’s atmosphere, and interacts better with Criegee intermediate than N2O or CH4, we next examined the possible gas-phase reaction between a Criegee intermediate and CO2 at the M06-2X/aug-cc-pVTZ level of theory. Surprisingly, the reaction between Criegee intermediate and CO2 has been relatively less explored, presumably due to the inert nature of CO2 that may make this reaction less likely. Though preliminary calculations on the Criegee-CO2 reaction are reported in literature,21 our results suggest that several important mechanistic features of this reaction including new reaction paths and transition states, have been overlooked previously. Our calculations indicate that the cycloaddition reaction between CH2OO and CO2 involves a barrier of 9.4 kcal/mol relative to the prereaction complex (Int1) and results in the formation of a cyclic carbonate- or ozonide-type adduct (Figure 2a). The cyclic adduct is 34.9 kcal/mol more stable than separated reactants. Interestingly, Aplincourt and Ruiz-Lopez previously studied the CH2OO-CO2 reaction at the CCSD(T)/6-311G(d,p)//B3LYP/6-31G(d,p) level of theory.21 They observed the barrierless formation of cyclic carbonate adduct with a reaction energy of -28.7 kcal/mol. However, our CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ calculations suggest that though the cyclic adduct is formed with a similar reaction energy of -29.0 kcal/mol, its formation is preceded by a transition state that lies 6.9 kcal/mol above the separated reactants. The effective forward barrier for the adduct formation is 11.0 kcal/mol relative to Int1.

ACS Paragon Plus Environment

6

Page 6 of 27

Page 7 of 27

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 2. (a) M06-2X/aug-cc-pVTZ calculated reaction profile for the gas-phase reaction of the various Criegee intermediates, (R1)(R2)COO with carbon dioxide (298.15 K and 1 atm). The values in parentheses are calculated at the CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ level of theory. The total energy of separated reactants, (R1)(R2)COO and carbon dioxide is set as the zero of the energy scale. (b) The correlation plots between the barrier heights for the reactions of various anti substituted Criegee intermediates with carbon dioxide and Mulliken charges on the central Criegee carbon (purple color) and terminal Criegee oxygen atoms (orange color), respectively. The barrier heights are measured relative to the prereaction complex, in which the Criegee intermediate is non-covalently bonded to carbon dioxide. The energies in both panels are zero-point-corrected electronic energies and are given in kcal/mol units.

ACS Paragon Plus Environment

7

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

We next examined the effect of various substituents on the CH2OO-CO2 reaction. The results suggest that the reaction barrier could be further lowered by making judicious substitutions at the Criegee carbon. Though the syn substituents tends to raise the reaction barrier by upto 4.0 kcal/mol irrespective of their nature (Figure S2), the anti substituents lower the reaction barrier by ~4.0 kcal/mol (Figure 2b and Table S2). For example, the reaction between CH3CHOO and CO2 has an effective barrier of 6.0 and 11.4 kcal/mol for the anti and syn orientations of methyl substituent relative to terminal Criegee oxygen, respectively. The nature and size of the anti substituents plays an important role in lowering the reaction barrier. Among the F, Cl, and CF3 electron withdrawing substituents considered, only F and Cl, which are relatively smaller in size, lower the reaction barrier. The anti F substitution lowers the barrier to 6.7 kcal/mol where the anti Cl substitution lowers the barrier to 8.4 kcal/mol. The effect of anti substituents on the reaction barrier could be explained in terms of the electrophilicity of Criegee carbon and nucleophilicity of terminal Criegee oxygen i.e., there exists an inverse correlation between the reaction barrier and the positive charge on the Criegee carbon, or the negative charge on the terminal Criegee oxygen in the CO2 reactions of anti substituted Criegee intermediate (Figure 2b). The anti-CHFOO and anti-CH3CHOO have the most positive Criegee carbons and the most negative Criegee oxygens that makes their CO2 reactions the lowest barrier reactions among the considered ones. Clearly, the electronic and steric nature of anti substituents in Criegee intermediates would play an important role in guiding the design of Criegee-based frameworks for the CO2 removal from air. Alternatively, the Int2, which is formed from the CH2OO-CO2 reaction with an excess energy of 29.0 kcal/mol, could decompose further. The CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ calculations suggest that the Int2 decomposition could occur either concertedly or in a stepwise

ACS Paragon Plus Environment

8

Page 8 of 27

Page 9 of 27

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

manner (Figure 3). Interestingly, both pathways could lead to the formation of formic acid (HCOOH) with CO2 being retrieved at the end of the reaction. In this decomposition reaction, CO2 participates in an oxygen-exchange reaction with Criegee intermediate, CH2OO, i.e., CO2 accepts the terminal Criegee oxygen of CH2OO and donates one of its oxygens to Criegee carbon. This leads to the Criegee isomerization into HCOOH in a catalytic manner. As far as we know, this is the first example showing the catalytic role of CO2 in an atmospherically relevant chemical reaction.

Figure 3. uCCSD(T)/aug-cc-pVTZ//uM06-2X/aug-cc-pVTZ calculated concerted and stepwise pathways for the decomposition of cyclic carbonate-type adduct, Int2 (298.15 K, 1 atm). The total energy of separated reactants, CH2OO and carbon dioxide is set as the zero of the energy scale. The energies values here are zero-point-corrected electronic energies and are given in kcal/mol units.

ACS Paragon Plus Environment

9

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 27

The decomposition of Int2 into HCOOH and CO2 has an extremely favorable reaction energy of -115.7 kcal/mol. The comparative analysis of the concerted and stepwise pathways implies that the latter pathway, which is mediated by the homolytic cleavage of O-O bond and the subsequent formation of bismethoxy species, is favored. The transition state for the ratelimiting formation of bismethoxy biradical in the stepwise pathway (TS5) is calculated to be 10.0 kcal/mol lower in energy than that for the concerted formation of HCOOH and CO2 (TS2). It is important to mention here that the stepwise decomposition of Int2 via TS3 is mediated by many biradical stationary points. Those stationary points are carefully characterized using the unrestricted formalism. Specifically, we have used the guess=(read,mix) option, as implemented in the Gaussian0920 software, to generate the unrestricted broken-symmetry wavefunctions for the singlet biradical stationary points. Our calculated energy of bismethoxy radical (∆E = -14.0 kcal/mol) relative to CH2OO is in a very good agreement with a recent CCSDT-F12a/AVTZ value of -13.0 kcal/mol from the full-dimensional potential energy study of CH2OO.22 This supports the authenticity of our chosen theoretical approach for describing the singlet biradical stationary points. Bismethoxy species can either decompose into CO2 and H2 or isomerize into HCOOH. The latter channel involves 0.7 kcal/mol lower barrier than the former one. Interestingly, the more favorable stepwise mechanism for the formation of HCOOH and CO2 as well as the CO2 and H2 forming pathway have been overlooked in previous calculations of Aplincourt and Ruiz-Lopez.21 The concerted decomposition of Int2 involves the breakage of O-O and O-C bonds as well as the hydrogen atom transfer from carbon to oxygen atom. The barrier for the concerted decomposition is quite high, i.e., the transition state (TS2) lies 21.0 kcal/mol above the zero of the reaction. However, carboxylic acids that are present in appreciable amounts in the

ACS Paragon Plus Environment

10

Page 11 of 27

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

troposphere,23-27 and are also formed in the Criegee-CO2 reaction, are capable of catalyzing hydrogen atom transfer chemistries.28 We, thus, next estimated the catalytic effect of HCOOH on the concerted decomposition of Int2, which also involves hydrogen atom transfer. The calculations suggest that the conversion of Int2 into HCOOH and CO2 under HCOOH catalysis is now mediated by prereaction (Int1FA) and postreaction complexes (Int2FA, Figure 4). Int1FA has a binding energy of 6.4 kcal/mol relative to Int2 and HCOOH catalyst whereas Int2FA is 17.9 kcal/mol more stable than separated products (HCOOH + CO2) and HCOOH catalyst due to extensive hydrogen bonding involving CO2 and two HCOOH molecules. More importantly, the transition state for the HCOOH-catalyzed concerted decomposition of Int2 (TSFA) lies at the zero

Figure 4. CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ calculated concerted pathway for the decomposition of the cyclic carbonate-type adduct formed from the reaction between the simplest Criegee intermediate, CH2OO and carbon dioxide with and without formic acid (298.15 K and 1 atm). The total energy of separated reactants, CH2OO and carbon dioxide is set as the zero of the energy scale. The energies values here are zero-point-corrected electronic energies and are given in kcal/mol units.

ACS Paragon Plus Environment

11

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 12 of 27

of the reaction, which is 11.0 kcal/mol lower than that for the formation of bismethoxy species in the stepwise pathway. These results suggests that under tropospheric conditions, acid catalysis may alter the mechanistic nature of Int2 decomposition from being stepwise to concerted and result in the exclusive isomerization of Criegee intermediate into HCOOH. Since HCOOH is formed in the Criegee-CO2 reaction, the decomposition of Int2 may be autocatalytic. The role of acid catalysis in effeciently decomposing cyclic peroxides, which is a newly discovered isomer of γ-ketohydroperoxide, has recently been demonstrated under low-temperature conditions.29 The predicted rate coefficients for this decomposition were found to be in excellent agreement with experiment when due consideration was made for the acid-catalyzed isomerization.

Table 1. Calculated kinetic data for the formation of cyclic adduct from the reaction of the simplest Criegee intermediate, CH2OO with carbon dioxide. Temperature (K)

Keq(molecule cm-3)

k2 (s-1)

kTS (cm3 molecule-1 s-1)

keff (s-1)

200 210 220 230 240 250 260 270 280 290 298.15 300

3.7 x 10-20 2.3 x 10-20 1.5 x 10-20 1.1 x 10-20 7.5 x 10-21 5.5 x 10-21 4.1 x 10-21 3.2 x 10-21 2.5 x 10-21 2.0 x 10-21 1.7 x 10-22 1.7 x 10-22

3.8 x 10-1 1.3 x 100 4.1 x 100 1.1 x 101 2.9 x 101 7.0 x 101 1.6 x 102 3.3 x 102 6.5 x 102 1.2 x 103 2.0 x 103 2.2 x 103

1.4 x 10-21 3.1 x 10-21 6.3 x 10-21 1.2 x 10-20 2.2 x 10-20 3.8 x 10-20 6.5 x 10-20 1.0 x 10-19 1.6 x 10-19 2.5 x 10-19 3.5 x 10-19 3.7 x 10-19

1.4 x 10-5 3.1 x 10-5 6.3 x 10-5 1.2 x 10-4 2.2 x 10-4 3.8 x 10-4 6.5 x 10-4 1.0 x 10-3 1.6 x 10-3 2.5 x 10-3 3.5 x 10-3 3.7 x 10-3

ACS Paragon Plus Environment

12

Page 13 of 27

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

To gain deeper insights into the CH2OO-CO2 reaction, we next calculated the rate of formation of the cyclic adduct, Int2 from the gas-phase reaction between CH2OO and CO2 within the framework of variational transition state theory. According to the reaction profiles shown in Figure 2, the cyclic adduct formation follows a two-step mechanism as described by equation 1, where the prereaction complex Int1 is in equilibrium with CH2OO and CO2, and the reaction proceeds through the unimolecular decomposition of this complex.

(1) In the steady state approximation, the forward rate constant can be approximated by 

k    k   K k  

(2)



K   e

 

   

where Keq is the equilibrium constant for the formation of the Int1

complex, the various Q denote the partition functions of the reactants and, the prereaction complex, and k2 is the unimolecular rate constant for its isomerization into the cyclic carbonate adduct as given in equation 1. In these equations, the equilibrium constant has been evaluated from the relative energy obtained at CCSD(T)/aug-cc-pVTZ level of theory and partition functions computed at M06-2X/aug-ccpVTZ level of theory, whereas for k2, we have carried out variational transition state theory calculations employing energies obtained at CCSD(T)/aug-cc-pVTZ level of theory and partition functions computed at M06-2X/aug-cc-pVTZ level of theory. The partition functions were calculated using a rigid rotator-harmonic oscillator for the vibrational degrees-of-freedom. See Computational Methods for additional details.

ACS Paragon Plus Environment

13

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 14 of 27

The calculated rate coefficient for the CH2OO-CO2 reaction shows a positive temperature dependence, i.e., the rate coefficient increases by nearly two orders of magnitude in going from 200 to 300 K (Table 1). At 300 K, the calculated rate coefficient is 3.5 x 10-19 cm3 molecule-1 s-1 whereas at 200 K, the rate coefficient is 1.4 x 10-21 cm3 molecule-1 s-1. The calculated rate coefficient for the CO2 reaction at 298.15 K is 3.5 x 10-19 cm3 molecule-1 s-1, which is much smaller than those estimated for H2O (10-16),30 (H2O)2 (10-11),31 H2S(10-15),32 SO2 (10-11),6 NO2 (10-12),6 NH3 (10-13)33, HO2 (10-10),34 HO (10-12),35 RO2 (10-12),35 H2SO4 (10-12),35 HNO3 (1010 36,37

)

and carboxylic acids (10-10),38 but is nearly 5 orders of magnitude larger than that for the

CO reaction (10-24),39 and is only an order of magnitude smaller than that for the NO reaction (10-18, Table 2).35 The direct comparison between the rate coefficients of various bimolecular Criegee reactions is deceptive because it does not take into account the atmospheric concentrations of coreactants. Thus, we next calculated the effective rates for the various bimolecular CH2OO reactions according to the following equation: keff = kTS[coreactant] (3) where [coreactant] is the tropospheric concentration of a trace gas reacting with CH2OO. We have used previously calculated kTS values for the non-CO2 reactions at 298 K (Table 2). For the coreactant concentration, we have used [H2O] ~ 1017 molecules/cm3, [(H2O)2] ~ 1014 molecules/cm3, [RCOOH] ~ 1010 molecules/cm3, [HO2] ~ 108 molecules/cm3, [HO] ~ 105 molecules/cm3, [RO2] ~ 108 molecules/cm3, [SO2] ~ 108 molecules/cm3, [H2SO4] ~ 106 molecules/cm3, [CO] ~ 1012 molecules/cm3, [NO] ~ 108 molecules/cm3, and [NO2] ~ 109 molecules/cm3. These values were adopted from a recent study by Vereecken et al.,35 in which they have summarized concentrations of various species in different environments based on

ACS Paragon Plus Environment

14

Page 15 of 27

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

Table 2. Calculated effective rates for the reaction of the simplest Criegee intermediate, CH2OO with various coreactants at 298 K and 1 atm.

coreactant H2O (H2O)2 H2S SO2 NO2 NH3 HO2 OH RO2 H2SO4 HNO3 RCOOH CO NO CO2

kTS (cm3 molecule-1 s-1) 10-16 10-11 10-15 10-11 10-12 10-13 10-10 10-12 10-12 10-12 10-10 10-10 10-24 10-18 10-19

molecule cm-3 1017 1014 1011 1010 109 1011 108 105 108 106 1011 1010 1012 108 1016

keff (s-1) 101 103 10-4 10-1 10-3 10-2 10-2 10-7 10-4 10-6 101 100 10-12 10-10 10-3

reference 30 31 32 6 6 33 34 35 35 35 36,37 38 39 35 Present work

previous literature estimates.40,41 The H2S concentration of 1011 molecules/cm3 is used.42,43 For NH3 concentration, we have used ~ 1011 molecules/cm3.44 The HNO3 concentration of 1011 molecules/cm3 is used.45 The current [CO2] in troposphere is 406 ppmV,46 which translates into 1.0 x 1016 molecules cm-3. The keff for the CH2OO-CO2 reaction at 298.15 K is 10-3 s-1, which is larger than those estimated for the CH2OO-H2SO4, CH2OO-NO, CH2OO-OH, CH2OO-CO, CH2OO-OH, CH2OO-H2S, and CH2OO-RO2, and are comparable to those estimated for the CH2OO-NO2 reaction. Only the CH2OO-H2O, CH2OO-(H2O)2, CH2OO-NH3, CH2OO-HO2, CH2OO-SO2, CH2OO-HCOOH and CH2OO-HNO3 reactions have larger keff values than the CH2OO-CO2 reaction. The calculated barrier for the anti-CH3CHOO-CO2 reaction at the M062X/aug-cc-pVTZ level is 3.4 kcal/mol lower than that for the CH2OO-CO2 reaction, implying

ACS Paragon Plus Environment

15

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 16 of 27

that the CO2 reaction of anti-CH3CHOO may even compete with the NH3 and SO2 reactions. For larger Criegee intermediates (C3), the water reaction is less significant,47 suggesting that the CO2 reaction will only face competition from the SO2, NH3, and acid reactions.

Figure 5. ONIOM(M06-2X/aug-cc-pVTZ: UFF)-EE calculated reaction profile for the CH2OOCO2 reaction on a water droplet of 191 water molecules at 298.15 K and 1 atm. For the sake of clarity, only the relevant part of the entire system is shown here. The CH2OO and CO2 are shown in ball-stick representation whereas the droplet water molecules are shown in wire representation. The energies values here are the zero-point-corrected electronic energies and are given in kcal/mol units.

Considering that the air-water interfaces are ubiquitous in troposphere and are characteristics of oceans surfaces, lakes, fog waters and water microdroplets,48 we next examined the CH2OO-CO2 reaction on a water droplet of 191 water molecules using the

ACS Paragon Plus Environment

16

Page 17 of 27

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

ONIOM49-52 (QM:MM) method, an integrated quantum mechanics:molecular mechanics (QM:MM) approach, as implemented in the Gaussian0920 program. The ONIOM calculations were performed with a two-layer ONIOM electronic embedding (EE) scheme, in which the highlevel region (CH2OO and CO2) was treated with the M06-2X density functional method in conjunction with the aug-cc-pVTZ basis set, and the low-level region was treated with universal force field (UFF),53-55 i.e., ONIOM(M06-2X/aug-cc-pVTZ:UFF)-EE. See Computational Methods for more details. The results suggest that the interfacial reaction has a barrier of 8.3 kcal/mol, which is 1.1 kca/mol lower than that in the gas-phase (Figure 5), suggesting that the air-water interface may promote the CH2OO-CO2 reaction. The cyclic adduct formed is 30.8 kcal/mol more stable than the preaction complex, and is 4.1 kcal/mol less stable than in the gas-phase. The Criegee-CO2 reaction may have important implications for carbon capture. CO2 as the primary greenhouse gas has raised a great deal of concern owing to its effect on global warming and climate change.14 This has led to the global urgency for developing efficient CO2 capture and storage technologies.56 Converting CO2 and epoxides into cyclic carbonates is one of the most prominent green approaches used for recycling CO2 into valuable chemicals by chemical fixation.57,58 The Criegee-CO2 reaction could represent an important metal free chemical fixation approach that leads to cyclic carbonate-type adduct under moderate temperatures and may help in reducing the carbon footprint. Since formation of the adduct, Int2 via CH2OO-CO2 reaction involves a moderate barrier and has a reaction energy of -29.0 kcal/mol, the judicious substitution at the Criegee carbon as well as the tuning of reaction conditions could be exploited to dominantly favor the formation of this cyclic carbonate-type adduct. These cyclic carbonates may be utilized as substrates for producing polycarbonates and

ACS Paragon Plus Environment

17

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 18 of 27

other value-added chemicals. To date, Criegee chemistry has been mainly studied because of its tropospheric significance.13 However, our results suggest that the high reactivity of Criegee intermediate could be harnessed to design new chemical fixation technologies for CO2 capture. Considering that the gas-phase laboratory synthesis of Criegee intermediate is now possible,6,7,59 our calculations provide crucial leads for extending the usefulness of Criegee chemistry beyond troposphere. Though the role of homogenous and heterogeneous catalysis as well as that of metal organic frameworks for CO2 capture is well-documented in literature,57-59 this is the first time that the possible role of Criegee reactivity in removing harmful CO2 from atmosphere is being reported. In summary, high level theoretical calculations have been used to study bimolecular chemistries between Criegee intermediate and greenhouse gases. The results suggest that the interaction between Criegee intermediate and carbon dioxide is the most favourable and leads to the exothermic formation of cyclic carbonate-type adduct with a barrier of 6.0-14.0 kcal/mol. This is in contrast to previous calculations suggesting the barrierless formation of the cyclic adduct. The rate calculations suggest that the reaction of Criegee with CO2 could be an important sink in addition to the water, water dimer, NH3, SO2, and acid reactions, and could be utilized in designing efficient carbon capture technologies. The subsequent decomposition of cyclic adduct, which results in formic acid and carbon dixoide, follows both concerted and stepwise mechanisms. The latter mechanism has been overlooked in a previous theoretical study. Under formic acid catalysis, the concerted decomposition of cyclic carbonate into formic acid and carbon dioxide is preferred suggesting that carbon dioxide and formic acid may catalyse the isomerization of Criegee intermediate into formic acid. These results reveal that the Criegee intermediate may serve as a new source of carboxylic acids in the troposphere.

ACS Paragon Plus Environment

18

Page 19 of 27

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

COMPUTATIONAL METHODS All quantum chemical calculations reported in this work were performed using Gaussian0920 software. In the first step, the possible interactions of the simplest Criegee intermediate, CH2OO with three greenhouse gases, CO2, CH4 and N2O were examined. The equilibrium constants for the complexes of CH2OO with CO2, CH4 and N2O were calculated using the relative energy obtained at the coupled cluster single and double substitution method with a perturbative treatment of triple excitations (CCSD(T))19/aug-cc-pVTZ18 level of theory and partition functions computed at M06-2X17/aug-cc-pVTZ level of theory. In specific, the zero-pointcorrected CCSD(T)/aug-cc-pVTZ electronic energies at the M06-2X/aug-cc-pVTZ-optimized geometries were used. In the second step, the gas-phase cycloaddition reactions of nine different Criegee intermediates with CO2 (Figure 2), resulting in the formation of a cyclic carbonate-type adduct, were investigated at the M06-2X/aug-cc-pVTZ level of theory. The reaction of the simplest Criegee intermediate, CH2OO with CO2 was explored in greater detail, i.e., the subsequent decomposition of the cyclic adduct was also examined. The energetics of the CH2OO-CO2 reaction was further improved by performing single point calculations at the CCSD(T)/aug-ccpVTZ//M06-2X/aug-cc-pVTZ level of theory. The decomposition of the cyclic carbonate-type adduct with and without formic acid catalyst was also examined. For accurately describing the various singlet biradical stationary points, geometry optimizations were performed using the unrestricted formalism. Specifically, we used the guess = (read, mix) keyword to generate the unrestricted broken-symmetry singlet wavefunctions for the open shell biradical stationary points. The wavefunction stability for such cases was confirmed using the stable=opt keyword. Carboxylic acids that are present in appreciable amounts in the troposphere,23-27 and are also

ACS Paragon Plus Environment

19

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 20 of 27

formed in the Criegee-CO2 reaction, are capable of catalyzing hydrogen atom transfer chemistries.28 We, thus, estimated the catalytic effect of HCOOH on the concerted decomposition of cyclic adduct, which also involves hydrogen atom transfer. For all the reactions, the M06-2X/aug-cc-pVTZ calculated harmonic vibrational frequencies were used to estimate the zero-point correction for the reactants, products, transition states, and intermediates. Intrinsic reaction coordinate calculations were carried out to ensure that a given transition state connects with the desired reactant and product. The rate constant calculations were also performed to gain deeper insights into the formation of cyclic carbonate adduct from the reaction of CH2OO with CO2. All rate constant calculations were performed with POLYRATE,60 version 2015. The rate constants were calculated using variational transition state theory. The energies obtained at CCSD(T)/aug-cc-pVTZ//M06-2X/aug-cc-pVTZ level of theory, and partition functions computed at M06-2X/aug-cc-pVTZ level of theory were used for estimating the rate constants for the bimolecular CH2OO-CO2 reaction. Finally, we examined the CH2OO-CO2 reaction on a water droplet of 191 water molecules

using

the

ONIOM49-52

(QM:MM)

method,

an

integrated

quantum

mechanics:molecular mechanics (QM:MM) approach as implemented in the Gaussian0920 program. The initial geometry of the water droplet system for the ONIOM calculations was extracted from the liquid bulk water (40 Å × 40 Å × 40 Å) that was first equilibrated up to 3 nanoseconds by performing classical MD simulations using TIP3P force field for water molecules in Material Studio software developed by Accelrys Incorporation (Cambridge, UK). Subsequently, the water droplet of 191 molecules was extracted from the bulk water and was equilibrated upto 5 picoseconds by performing the Born Oppenheimer molecular dynamics simulations. These simulations were performed using the Cp2K61 code. Using the equilibrated

ACS Paragon Plus Environment

20

Page 21 of 27

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

geometry of the water droplet system, the ONIOM calculations were performed with a two-layer electronic embedding (EE) scheme, in which the high-level region (CH2OO and CO2) was treated with the M06-2X density functional method in conjunction with the aug-cc-pVTZ basis set, and the low-level region was treated with universal force field (UFF),53-55 i.e., ONIOM(M062X/aug-cc-pVTZ:UFF)-EE. UFF is a molecular mechanics force field designed to model the entire periodic table, and has been successfully applied to organic molecules, metallic complexes, and main group compounds.57-59

ASSOCIATED CONTENT Supporting Information. Equilibrium geometries of two CH2OO⋅⋅N2O conformers, figure showing the barrier heights of syn-substituted Criegee intermediates, tables containing the calculated equilibrium constants for the complexes of CH2OO with three greenhouse gases, barrier heights for the reactions of various syn and anti substituted Criegee intermediates with carbon dioxide, and Cartesian coordinates of key species involved in the reactions of various Criegee intermediate with carbon dioxide.

AUTHOR INFORMATION Corresponding Author *Email: [email protected] Notes

ACS Paragon Plus Environment

21

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 22 of 27

The authors declare no competing financial interests. ACKNOWLEDGMENT This work is supported by the University of Nebraska Holland Computing Center. REFERENCES (1) Leather, K. E.; McGillen, M. R.; Cooke, M. C.; Utembe, S. R.; Archibald, A. T.; Jenkin, M. E.; Derwent, R. G.; Shallcross, D. E.; Percival, C. J. Acid-Yield Measurements of the Gas-Phase Ozonolysis of Ethene as a Function of Humidity Using Chemical Ionisation Mass Spectrometry (CIMS). Atmos. Chem. Phys. 2012, 12, 469-479. (2) Mauldin, R. L.; Berndt, T.; Sipil., M.; Paasonen, P.; Petäjä, T.; Kim, S.; Kurtén, T.; Stratmann, F.; Kerminen, V.-M.; Kulmala, M. A New Atmospherically Relevant Oxidant of Sulfur Dioxide. Nature 2012, 488, 193-196. (3) Heaton, K. J.; Sleighter, R. L.; Hatcher, P. G.; Hall IV, W. A.; Johnston, M. V. Composition Domains in Monoterpene Secondary Organic Aerosol. Environ. Sci. Technol. 2009, 43, 7797-7802. (4) Ma, Y.; Porter, R. A.; Chappell, D.; Russell, A. T.; Marston, G. Mechanisms for the Formation of Organic Acids in the Gas-Phase Ozonolysis of 3-Carene. Phys. Chem. Chem. Phys. 2009, 11, 4184-4197. (5) Taatjes, C. A.; Welz, O.; Eskola, A. J.; Savee, J. D.; Osborn, D. L.; Lee, E. P. F.; Dyke, J. M.; Mok, D.W. K.; Shallcross, D. E.; Percival, C. J. Direct Measurement of Criegee Intermediate (CH2OO) Reactions with Acetone, Acetaldehyde, and Hexafluoroacetone. Phys. Chem. Chem. Phys. 2012, 14, 10391-10400. (6) 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. (7) 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. (8) Vereecken, L. Lifting the Veil On An Old Mystery. Science 2013, 340, 154-155.

ACS Paragon Plus Environment

22

Page 23 of 27

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

(9) Johnson, D.; Marston, G. The Gas-Phase Ozonolysis of Unsaturated Volatile Organic Compounds in the Troposphere. Chem. Soc. Rev. 2008, 37, 699-716. (10)

Criegee, R. Mechanism of Ozonolysis. Angew. Chem. Int. Ed. 1975, 14, 745-752.

(11)

Ragan, J. A.; am Ende, D. J.; Brenek, S. J.; Eisenbreis, S. A.; Singer, R. A.;

Tickner, D. L.; Teixeira, J. J.; Vanderplas, B. C.; Weston, N. Safe Execution of a LargeScale Ozonolysis:  Preparation of the Bisulfite Adduct of 2-Hydroxyindan-2carboxaldehyde and Its Utility in a Reductive Amination. Org. Process Res. Dev. 2003, 7, 155-160. (12)

Van Ornum, S. G.; Champeau, R. M.; Pariza, R. Ozonolysis Applications in Drug

Synthesis. Chem. Rev. 2006, 106, 2990-3001. (13)

Vereecken, L.; Glowacki, D. R.; Pilling, M. J. Theoretical Chemical Kinetics in

Tropospheric Chemistry: Methodologies and Applications. Chem. Rev. 2015, 115, 40634114. (14)

Allen, M. R.; Frame, D. J.; Huntingford, C.; Jones, C. D.; Lowe, J. A.;

Meinshausen, M.; Meinshausen, N. Warming Caused by Cumulative Carbon Emissions Towards the Trillionth Tonne. Nature 2009, 458, 1163-1166. (15)

IPCC, 2014: Climate Change 2014: Synthesis Report. Contribution of Working

Groups I, II and III to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change (Core Writing Team, R.K. Pachauri and L.A. Meyer, eds.). IPCC, Geneva, 151 pp. (16)

WMO Greenhouse Gas Bulletin: The State of Greenhouse Gases in the

Atmosphere Based on Global Observations through 2015, No. 12, 24 October 2016. (17)

Zhao, Y.; Truhlar, D. G. The M06 Suite of Density Functionals for Main Group

Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215-241. (18)

Kendall, R. A.; Dunning, Jr., 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. (19)

Noga, J.; Bartlett, R. J. The Full CCSD(T) Model for Molecular Electronic

Structure. J. Chem. Phys. 1987, 86, 7041-7050.

ACS Paragon Plus Environment

23

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

(20)

Page 24 of 27

Gaussian 09, Revision D.01, Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria,

G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A. et al. Gaussian, Inc., Wallingford CT, 2009. (21)

Aplincourt, P.; Ruiz-Lopez, M. F. Theoretical Investigation of Reaction

Mechanisms for Carboxylic Acid Formation in the Atmosphere. J. Am. Chem. Soc. 2000, 122, 8990-8997. (22)

Li, J.; Guo, H. Full-Dimensional Potential Energy Surface and Ro-vibrational

Levels of Dioxirane. J. Phys. Chem. A 2016, 120, 2991-2998. (23)

Chebbi, A.; Carlier, P. Carboxylic Acids in the Troposphere, Occurrence,

Sources, and Sinks: A Review. Atmos. Environ. 1996, 30, 4233-4249. (24)

Granby, K.; Christensen, C. S.; Lohse, C. Urban and Semi-Rural Observations of

Carboxylic Acids and Carbonyls. Atmos. Environ. 1997, 31, 1403-1415. (25)

Shephard, M. W.; Goldman, A.; Clough, S. A.; Mlawer, E. J. Spectroscopic

Improvements Providing Evidence of Formic Acid in AERI-LBLRTM Validation Spectra. J. Quant. Spectrosc. Radiat. Transfer 2003, 82, 383-390. (26)

Grutter, M.; Glatthor, N.; Stiller, G. P.; Fischer, H.; Grabowski, U.; Höpfner, M.;

Kellmann, S.; Linden, A.; von Clarmann, T. Global Distribution and Variability of Formic Acid as Observed by MIPAS-ENVISAT. J. Geophys. Res. 2010, 115, D10303. (27)

Rinsland, C. P.; Mahieu, E.; Zander, R.; Goldman, A.; Wood, S.; Chiou, L. Free

Tropospheric Measurements of Formic Acid (HCOOH) from Infrared Ground-Based Solar Absorption Spectra: Retrieval Approach, Evidence for a Seasonal Cycle, and Comparison with Model Calculations. J. Geophys. Res. 2004, 109, D18308. (28)

Kumar, M.; Sinha, A.; Francisco, J. S. Role of Double Hydrogen Atom Transfer

Reactions in Atmospheric Chemistry. Acc. Chem. Res. 2016, 49, 877-883. (29)

Jalan, A.; Alecu, I. M.; Meana-Paneda, R.; Aguilera-Iparraguirre, J.; Yang, K. R.;

Merchant, S. S.; Truhlar, D. G.; Green, W. H. New Pathways for Formation of Acids and Carbonyl Products in Low-Temperature Oxidation: The Korcek Decomposition of γKetohydroperoxides. J. Am. Chem. Soc. 2013, 135, 11100-11114. (30)

Long, B.; Bao, J. L.; Truhlar, D. G. Atmospheric Chemistry of Criegee

Intermediates. Unimolecular Reactions and Reactions with Water. J. Am. Chem. Soc. 2016, 138, 14409-14422.

ACS Paragon Plus Environment

24

Page 25 of 27

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

(31)

Anglada, J. M.; Sole, A. Impact of the Water Dimer on the Atmospheric

Reactivity of Carbonyl Oxides. Phys. Chem. Chem. Phys. 2017, 18, 17698-17712. (32)

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. (33)

Jorgensen, S.; Gross, A. Theoretical Investigation of the Reaction between

Carbonyl Oxides and Ammonia. J. Phys. Chem. A 2009, 113, 10284-10290. (34)

Long, B.; Tan, X.; Long, Z.; Wang, Y.; Ren, D.; Zhang. W. Theoretical Studies

on Reactions of the Stabilized H2COO with HO2 and the HO2···H2O Complex. J. Phys. Chem. A 2011, 115, 6559-6567. (35)

Vereecken, L.; Harder, H.; Novelli, A. The Reaction of Criegee Intermediates

with NO, RO2, and SO2, and Their Fate in the Atmosphere. Phys. Chem. Chem. Phys. 2012, 14, 14682-14695. (36)

Foreman, E. S.; Kapnas, K. M.; Murray, C. Reactions between Criegee

Intermediate and the Inorganic Acids HCl and HNO3: Kinetics and Atmospheric Implications. Angew. Chem. Int. Ed. 2016, 55, 10419-10422. (37)

Raghunath, P.; Lee, Y.-P.; 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 2017, 121, 3871-3878. (38)

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. et al. Rate Coefficients of C1 and C2 Criegee Intermediate Reactions with Formic and Acetic Acid Near the Collision Limit: Direct Kinetics Measurements and Atmospheric Implications. Angew. Chem. Int. Ed. 2014, 53, 1-5. (39)

Kumar, M.; Busch, D. H.; Subramaniam, B.; Thompson, W. H. Criegee

Intermediate Reaction with CO: Mechanism, Barriers, Conformer-Dependence, and Implications for Ozonolysis Chemistry. J. Phys. Chem. A 2014, 118, 1887-1894. (40)

Junninen, H.; Lauri, A.; Keronen, P.; Aalto, P.; Hiltunen, V.; Hari, P.; Kulmala,

M. Smart-SMEAR: On-line Data Exploration and Visualization Tool for SMEAR Stations. Boreal Environ. Res. 2009, 14, 447-457.

ACS Paragon Plus Environment

25

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

(41)

Page 26 of 27

Petaja, T.; Mauldin, III, R. L.; Kosciuch, E.; McGrath, J.; Nieminen, T.;

Paasonen, P.; Boy, M.; Adamov, A.; Kotiaho, T.; Kulmala, M. Sulfuric Acid and OH Concentrations in a Boreal Forest Site. Atmos. Chem. Phys. 2009, 9, 7435-7448. (42)

Air Quality Guidelines for Europe, 2nd ed.; World Health Organization, Ed.;

WHO Regional Publications; World Health Organization, Regional Office for Europe; Copenhagen, 2000. (43)

Brimblecombe, P. Air Composition and Chemistry; Cambridge University Press:

Cambridge, U. K., 1996. (44)

Chen, H.; Varner, M. E.; Gerber, R. B.; Finlayson-Pitts, B. J. Reactions of

Methanesulfonic Acid with Amines and Ammonia as a Source of New Particles in Air. J. Phys. Chem. B 2016, 120, 1526-1536. (45)

Crisp, T. A.; Lerner, B. M.; Williams, E. J.; Quinn, P. K.; Bates, T. S.; Bertram,

T. H. Observations of Gas Phase Hydrochloric Acid in the Polluted Marine Boundary Layer. J. Geophys. Res. 2014, 119, 6897-6915. (46)

Trends in Atmospheric Carbon Dioxide; Earth System Research Laboratory,

National Oceanic & Atmospheric Administration, 2017. (47)

Huang, H.-L.; Chao, W.; 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. 2015, 112, 10857-10862. (48)

Ravishankara, A. R. Heterogeneous and Multiphase Chemistry in the

Troposphere. Science 1997, 276, 1058-1065. (49)

Maseras, F.; Morokuma, K. IMOMM: A New Integrated ab Initio + Molecular

Mechanics Geometry Optimization Scheme of Equilibrium Structures and Transition States. J. Comput. Chem. 1995, 16, 1170-1179. (50)

Dapprich, S.; Komaromi, I.; Byun, K. S.; Morokuma, K.; Frisch, M. J. A New

ONIOM Implementation in Gaussian98. Part I. The Calculation of Energies, Gradients, Vibrational Frequencies and Electric Field Derivatives. J. Mol. Struct. (THEOCHEM) 1999, 461, 1-21. (51)

Tschumper, G. S.; Morokuma, K. Gauging the Applicability of ONIOM

(MO/MO) Methods to Weak Chemical Interactions in Large Systems: Hydrogen Bonding in Alcohol Dimers. J. Mol. Struct. (THEOCHEM) 2002, 592, 137-147.

ACS Paragon Plus Environment

26

Page 27 of 27

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

(52)

Chung, L. W.; Hirao, H.; Li, X.; Morokuma, K. The ONIOM Method: Its

Foundation and Applications to Metalloenzymes and Photobiology. WIREs Comput. Mol. Sci. 2012, 2, 327-350. (53)

Rappe, A. K.; Casewit, C. J.; Colwell, K. S.; Goddard III, W. A.; Skiff, W. M.

UFF, a Full Periodic Table Force Field for Molecular Mechanics and Molecular Dynamics Simulations. J. Am. Chem. Soc. 1992, 114, 10024-10035. (54)

Casewit, C. J.; Colwell, K. S.; Rappe, A. K. Application of a Universal Force

Field to Organic Molecules. J. Am. Chem. Soc. 1992, 114, 10035-10046. (55)

Casewit, C. J.; Colwell, K. S.; Rappe, A. K. Application of a Universal Force

Field to Main Group Compounds. J. Am. Chem. Soc. 1992, 114, 10046-10053. (56)

Sumida, K.; Rogow, D. L.; Mason, J. A.; McDonald, T. M.; Bloch, E. D.; Herm,

Z. R.; Bae, T.-H.; Long, J. R. Carbon Dioxide Capture in Metal–Organic Frameworks. Chem. Rev. 2012, 112, 724-781. (57)

Ema, T.; Miyazaka, Y.; Shimonishi, J.; Maeda, C.; Hasegawa, J.-Y. Bifunctional

Porphyrin Catalysts for the Synthesis of Cyclic Carbonates from Epoxides and CO2: Structural Optimization and Mechanistic Study. J. Am. Chem. Soc. 2014, 136, 1527015279. (58)

Nakano, K.; Kobayashi, K.; Nozaki, K. Tetravalent Metal Complexes as a New

Family of Catalysts for Copolymerization of Epoxides with Carbon Dioxide. J. Am. Chem. Soc. 2011, 133, 10720-10723. (59)

Beames, J. M.; Liu, F.; Lu, L.; Lester, M. I. Ultraviolet Spectrum and

Photochemistry of the Simplest Criegee Intermediate CH2OO. J. Am. Chem. Soc. 2012, 134, 20045-20048. (60)

Zheng, J.; Zhang, S.; Lynch, B. J.; Corchado, J. C.; Chuang, Y.-Y.; Fast, P. L.;

Hu, W.-P.; Liu, Y.-P.; Lynch, G. C.; Nguyen, K. A. et al. POLYRATE-version 2015. University of Minnesota, Minneapolis, 2015. (61)

VandeVondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.;

Hutter, J. Quickstep: Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach. Comput. Phys. Commun. 2005, 167, 103-128.

ACS Paragon Plus Environment

27