Next Article in Journal
Immunocytochemical Analysis of Bifid Trichomes in Aldrovanda vesiculosa L. Traps
Previous Article in Journal
Melatonin-Assisted Cisplatin Suppresses Urinary Bladder Cancer Cell Proliferation and Growth through Inhibiting PrPC-Regulated Cell Stress and Cell Proliferation Signaling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Theoretical Study on the Gas Phase and Gas–Liquid Interface Reaction Mechanism of Criegee Intermediates with Glycolic Acid Sulfate

Environment Research Institute, Shandong University, Qingdao 266237, China
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(4), 3355; https://doi.org/10.3390/ijms24043355
Submission received: 20 December 2022 / Revised: 9 January 2023 / Accepted: 11 January 2023 / Published: 8 February 2023
(This article belongs to the Topic Theoretical, Quantum and Computational Chemistry)

Abstract

:
Criegee intermediates (CIs) are important zwitterionic oxidants in the atmosphere, which affect the budget of OH radicals, amines, alcohols, organic/inorganic acids, etc. In this study, quantum chemical calculation and Born–Oppenheimer molecular dynamic (BOMD) simulation were performed to show the reaction mechanisms of C2 CIs with glycolic acid sulfate (GAS) at the gas-phase and gas–liquid interface, respectively. The results indicate that CIs can react with COOH and OSO3H groups of GAS and generate hydroperoxide products. Intramolecular proton transfer reactions occurred in the simulations. Moreover, GAS acts as a proton donor and participates in the hydration of CIs, during which the intramolecular proton transfer also occurs. As GAS widely exists in atmospheric particulate matter, the reaction with GAS is one of the sink pathways of CIs in areas polluted by particulate matter.

1. Introduction

The reactions of ozone with unsaturated hydrocarbons are important sources of free radicals and particulate matter in the atmosphere [1]. Criegee intermediates (CIs) are tropospheric biradical/zwitterionic species that are derived from the ozonolysis of alkenes [2]. Ozone undergoes 1,3-cycloaddition with a double bond of alkene to form a primary ozonide that subsequently decomposes into a carbonyl oxide (also called Criegee intermediate) and a carbonyl compound [3]. The energized Criegee intermediate, which is produced by exothermic decomposition and contains vibrational excitation, undergoes unimolecular decay or forms a stable Criegee intermediate via collisional quenching [4]. In addition to the above reaction, the contribution that formed through other reactions to the total CIs is inappreciable. The reaction of CH3O2 with OH radicals, affecting the concentrations of HO2 and O3 in the oceanic boundary layer, has been determined to produce trace amounts of CH2OO (yield of CH2OO < 5%) [5]. The oxidation of dimethyl sulfoxide forms CH2OO at low temperatures, but the yield is too low to be a pivotal source of CIs [6].
The unimolecular decomposition of CIs, which occurs via the vinyl hydroperoxide pathway or the ester pathway, plays a significant role in the production of atmospheric OH radicals [7,8]. Elshorbany et al. [9] reported that the decomposition of CIs has a 24% contribution to the daytime OH radical formation. In addition, CI decomposition is a steady and dominant source of OH radicals at nighttime [10]. The PUMA campaign (a project that measured pollution of the urban midlands atmosphere in the UK) showed that ozonolysis of alkenes dominates the OH production in winter and accounts for more than 50% of the production in summer [11].
Even if most CIs decay through unimolecular pathways, a fraction of CIs survives for long enough to react with other substances. The reaction with water vapor is the most efficient scavenging route for CIs in the troposphere [12]. Part of CIs react with water molecules and form low vapor pressure substances that act as cloud condensation nuclei (CCN) [13]. CH2OO reacts with NO2 to enhance NO3 production in the urban atmosphere [14]. Dimethyl-substituted Criegee intermediate (CH3)2COO survives at high humidity to react with atmospheric SO2, and the reaction rate (1.3 × 10−10 cm3s−1) is close to the gas kinetic limit [15]. The reactions of CIs with atmospheric organic matter play a significant role in the formation of secondary organic aerosol (SOA). The product of the CH2OO + acrolein reaction is the secondary ozonide that has the potential to form SOA because of its large molecular weight [16]. Experimental and theoretical studies show that the reaction of CIs with alcohols represents a considerable source of α-alkoxyalkyl hydroperoxides (AAAHs) that form SOA with a production rate of 24 Gg year−1 in tropical forests [17]. Despite the fact that the reactions of CIs with organics have been internationally recognized, the understanding of a Criegee–organosulfate reaction in the atmosphere is still developing.
Organosulfates (OSs), also known as sulfate esters or sulfate derivatives, are important organosulfur compounds that contain -OSO3H group in the atmosphere [18,19]. Due to hydrophobic (the hydrocarbon group) and hydrophilic (the sulfate group) parts, OSs lower the surface tension of atmospheric particles, eventually changing the ability of particles to uptake water and to form CCN [20,21]. As OSs absorb sunlight, they significantly change the optical properties of organic aerosols and affect the energy balance of the atmosphere [22]. OSs react with oxidants such as OH, O3, and NOX in the atmospheric lifetime, which promotes the formation of SOA [23]. Accordingly, OSs have the potential to react with CIs. In the atmosphere, glycolic acid sulfate (GAS) is a common organosulfate with a -OSO3H group and a -COOH group [24]. Some studies have reported the formation of aqueous-phase and heterogeneous OSs [25,26]. Most of the GAS is present in the ambient aerosols in the atmosphere [27]. Detection of OSs in summer in Beijing revealed that GAS is the most abundant among all quantified species, the concentrations of GAS in the thirteen OSs measured ambient aerosols ranged from 3.9 to 58.2 ng/m3 and the average concentration was 19.5 ng/m3 [28]. Olson et al. measured concentrations of OSs in particulate matter in the United States, Pakistan, and Mexico, where GAS concentrations ranged from 1.9–11.3 ng/m3 [29]. Therefore, it is necessary to investigate the reaction mechanism of CIs with GAS in the atmosphere.
This work explored the reaction mechanism of C2 CIs (anti- and syn-CH3CHOO) with GAS using density functional theory (DFT). Due to the hydrophobic methyl group, C2 CIs have a longer lifetime at the gas–liquid interface compared to C1 CIs [30]. Therefore, the Born–Oppenheimer molecular dynamics (BOMD) was carried out to simulate the heterogeneous reactions of CH3CHOO with GAS on the droplet surface. For the gas–liquid interface reactions, we consider the collision of GAS molecules in the atmosphere with CIs on the droplet rather than the collision reaction of CIs with GAS that is dissolved in the droplet.

2. Results and Discussion

The most stable configuration of GAS is shown in Figure 1. A seven-membered ring structure is formed in the molecule. The hydrogen bond interaction was observed between the terminal hydrogen of the OSO3H group and the carbonyl oxygen of the COOH group.

2.1. Gas-Phase Reactions

The values of relative Gibbs free energy (CCSD/6-311++G(2d,2p)) of the reactions between CH3CHOO and GAS are displayed in Figure 2. The configurations of reaction complexes, transition states, and products are depicted in Figure S1. The reaction energy barriers of syn-CH3CHOO with OSO3H and COOH groups are 1.00 and 2.87 kcal/mol, respectively. The reactions of anti-CH3CHOO with OSO3H and COOH groups need to cross energy barriers of 3.30 and 4.09 kcal/mol, respectively. The low energy barriers suggest that these reactions are feasible in the atmosphere. Previous studies have shown that most reactions of CIs with carboxylic acids follow a barrierless pathway [31,32]. However, we observed that the reactions of the COOH group with CH3CHOO require overcoming the energy barrier, which may be due to the formation of a six-membered ring that increases the stability of GAS.
Water molecules are one of the most abundant species in the atmosphere and have a significant impact on atmospheric chemical processes [33]. Therefore, the relative Gibbs free energy (CCSD/6-311++G(2d,2p)) of the water-mediated reactions between CH3CHOO and GAS was calculated (Figure 3). The two lowest energy barriers are 0.78 and 2.55 kcal/mol, which are determined for the water-mediated reaction of syn-CH3CHOO with the OSO3H group and the reaction of syn-CH3CHOO with the COOH group, respectively. Compared with the direct reactions of syn-CH3CHOO, the water molecule lowers the energy barrier of the reaction with the OSO3H group but increases the energy barrier of the reaction with the COOH group. In contrast, the water-mediated reaction barrier of anti-CH3CHOO with OSO3H group is increased to 6.11 kcal/mol, and that of anti-CH3CHOO with COOH group is decreased to 3.43 kcal/mol. The participation of water molecules leads to the change of the reaction energy barrier to be less than 3 kcal/mol, indicating that the effect of water molecules on the reaction of CH3CHOO with GAS is weak. The configurations of the water-mediated reaction stages are presented in Figure S2.

2.2. Gas–Liquid Interface Reactions

The gas–liquid interface plays a pivotal and ubiquitous role in atmospheric chemistry, including absorbing various pollutants and changing chemical reaction mechanisms. For example, Shang et al. [34] reported the interfacial reaction of SO2 with oleic acid, which is a new pathway to form organosulfur in the atmosphere. In this study, the gas–liquid interface reaction mechanism of anti-CH3CHOO with GAS was explored at the molecular level. In order to eliminate the effect of reaction location, thirty simulations were performed at different locations of the droplet. The reactions of anti-CH3CHOO with GAS occurred in 16 simulations. The direct and water-mediated reactions of anti-CH3CHOO with the COOH group occurred six and two times, respectively. The direct and water-mediated reactions of anti-CH3CHOO with the OSO3H group both occurred four times. The hydration reactions of anti-CH3CHOO occurred 11 times.

2.2.1. Reaction of Anti-CH3CHOO with the COOH Group of GAS

Figure 4a shows the structure and bond length variations of the direct reaction between anti-CH3CHOO and the COOH group of GAS. Based on the properties of hydrophobicity and hydrophilicity, the methyl group of anti-CH3CHOO is placed on the side away from the droplet. At 0 ps, the distances of H-O3, C-O1, and H-O2 are 0.97, 2.66, and 2.52 Å, respectively. Subsequently, the H atom of the COOH group gradually approaches the terminal oxygen of anti-CH3CHOO, and the carbonyl oxygen of the COOH group approaches the α-carbon atom of anti-CH3CHOO. At 0.14 ps, the distances of H-O3, C-O1, and H-O2 are 1.53, 2.13, and 1.00 Å, respectively, where the transition-state-like structure is formed. At 0.20 ps, the H atom on the COOH group binds to the terminal oxygen of anti-CH3CHOO, and the C-O1 and H-O2 bonds are formed and remain stable, indicating the formation of the reaction product.
The mechanism of the water-mediated reaction between the anti-CH3CHOO and COOH group of GAS (Figure 4b) is different from that of the direct reaction. At 0 ps, the initial distances of H2-O3, H1-O3, H2-O2, H1-O4, and C-O1 are 0.98, 2.40, 2.05, 0.97, and 3.22 Å, respectively. At 0.12 ps, the structure of the reactants is similar to the transition state; the distances of H2-O3, H1-O3, H2-O2, H1-O4, and C-O1 are 1.05, 1.22, 1.48, 1.24, and 2.66 Å, respectively. At 0.24 ps, H2-O2 and H1-O4 bonds are formed. However, the C-O1 bond (1.56 Å) is formed at 0.57 ps. In water-mediated reactions, the C-O1 bond is formed later than other bonds, and the water molecule is the bridge of proton transfer.

2.2.2. Reaction of Anti-CH3CHOO with the OSO3H Group of GAS

The direct and water-mediated reactions between the anti-CH3CHOO and OSO3H groups of GAS were observed in our simulations (Figure 5). For the direct reaction, the distances of H-O2, H-O3, and C-O1 are 0.98, 2.48, and 4.04 Å, respectively, at 0 ps. From 0.38 to 0.63 ps, the H atom of the OSO3H group vibrates between O3 and O2 atoms. The length of H-O3 fluctuates around 0.98 Å from 0.63 ps, but the distance of C-O1 is still decreasing. At 0.71ps, the O1 atom of GAS binds to the C atom of anti-CH3CHOO, indicating the formation of a hydroperoxide product.
For the water-mediated reaction (Figure 5b), the initial distances of H2-O2, H1-O3, and C-O1 are 2.30, 2.03, and 3.10 Å, respectively. The transition-state-like structure is observed at 0.26 ps, where the lengths of H2-O2, H1-O3, H1-O1, H2-O3, and C-O1 are 1.41, 1.09, 1.39, 1.10, and 2.43 Å, respectively. The H2-O2 and H1-O3 are formed at 0.30 and 0.27 ps, respectively. The O1 atom binds to the C atom at 0.52 ps, which occurs later than the formation of the H2-O2 bond. This phenomenon is also observed in the water-mediated reaction between the COOH group of anti-CH3CHOO and GAS, suggesting that water-mediated proton transfer initiates the reactions and promotes the binding of C and O atoms.

2.2.3. Intramolecular Proton Transfer Reaction of Anti-CH3CHOO with the OSO3H Group of GAS

Most of the previous studies have focused on the reaction of CIs with monofunctional species [35,36,37]. Even the study of CIs and multifunctional species is the independent reaction of a single functional group [38]. However, the reaction involving both functional groups of GAS is observed in this study (Figure 6). At 0 ps, the two interface water molecules are far away from anti-CH3CHOO and GAS, and the distances of H2-O4 and H3-O2 are 3.68 and 3.92 Å, respectively. The transition-state-like complex of anti-CH3CHOO, GAS, and water molecules is formed at 2.09 ps. The proton of the OSO3H group is transferred to the carbonyl oxygen of the COOH group at this time; the distances of H1-O1 and H1-O6 are 1.15 and 1.44 Å, respectively. The proton of the COOH group moves to the water molecule, and the distances of H2-O5 and H2-O4 are 1.26 and 1.28 Å, respectively. In this way, proton transfer occurs between the two functional groups of GAS, and the COOH group acts as the shuttle of proton transfer in this reaction. At 2.37 ps, the H3-O2 bond (the length is 0.99 Å) is formed, indicating proton transfer between the four reactant molecules is complete. The C-O1 bond is formed at 2.79 ps, resulting in the formation of new products. During the whole reaction process, both the COOH group of GAS and water molecules act as bridges of proton transfer.

2.2.4. GAS-Mediated Hydration of Anti-CH3CHOO

Kumar et al. [39] reported BOMD simulations of anti-CH3CHOO reacting with HNO3 at the gas–liquid interface, suggesting that HNO3-mediated anti-CH3CHOO hydration is the most dominant reaction. In this study, the GAS-mediated hydration of anti-CH3CHOO occurs and generates C2H3O6S- and H3O+ ions, and the reaction follows an obvious stepwise mechanism. As shown in Figure 7, anti-CH3CHOO and GAS are placed on the droplet surface, which is far from the water molecules at 0 ps. For the first step, the proton on the COOH group is transferred to the terminal oxygen of anti-CH3CHOO, and the intramolecular proton transfer occurs in GAS. At 1.86 ps, the distances of H1-O1 and H5-O7 are 1.18 and 1.24 Å, respectively. At 2.10 ps, the distances of H1-O1 and H5-O7 are 0.98 and 1.01 Å, respectively, indicating the protonated anti-CH3CHOO and C2H3O6Sȡ ion are formed. In the second step, the O atom of the water molecule combines with α-C of the anti-CH3CHOO, and the proton is transferred to another water molecule. At 2.52 ps, the distances of C-O2 and H2-O3 are 1.41 and 0.98 Å, respectively. In the last step, the water molecules transfer protons to each other, and the H3O+ ion is formed at 4.85 ps. In this pathway, GAS acts as a water molecule to provide protons for the hydration of anti-CH3CHOO.

3. Atmospheric Implications

These results deepen the understanding of CI fate in the atmosphere. GAS is an abundant nucleation precursor and has an important contribution to the formation of atmospheric particulate pollution [40,41,42]. The reaction with GAS is one of the sink pathways for CH3CHOO, especially the direct reaction with the COOH group is a nearly barrierless process (1.00 kcal/mol), which may be the dominant reaction in a dry atmosphere. The water-mediated reaction between CH3CHOO and OSO3H groups is also a nearly barrierless process (0.78 kcal/mol), which may be the dominant reaction in polluted areas with high humidity.
Furthermore, the BOMD simulations reveal the reaction mechanism of anti-CH3CHOO with GAS on droplets. Both direct and water-mediated reactions are observed, with the formation of ring structures during the reaction. Among them, the water-mediated reaction follows the proton transfer mechanism, and the water molecule acts as a bridge for proton transfer. Although it has been reported that CH3CHOO can exist stably on the droplet surface [30], all the reactions of anti-CH3CHOO with GAS on the droplet occur on the ps timescale. The terminal O atom of anti-CH3CHOO binding to the H atom of GAS in the reaction occurs slightly later than the α-C atom of anti-CH3CHOO binding to the O atom of GAS, suggesting the reactions are initiated by the binding of O and H atoms. Both quantum chemical calculations and molecular simulations indicate that water molecules can participate in the reaction of CH3CHOO with GAS. In addition, though the reaction of CIs with various species on droplets is discussed in the literature [43,44,45,46], intramolecular proton transfer was observed in our simulations. The hydrogen bond and six-membered ring structure are formed inside the GAS molecule, resulting in the intramolecular proton transfer. It is believed that similar reactions can occur in some substances with longer carbon chains and multiple functional groups in the atmosphere. GAS-promoted hydration of anti-CH3CHOO as a proton donor was also observed in the simulations. Previous studies have pointed out that nitric acid and methanesulfonic acid promote CIs hydration [39,47]; GAS also tends to perform intramolecular proton transfer while donating a proton, which has appeared many times in the simulations.
The product of CH3CHOO reacting with GAS is hydroperoxides. Hydroperoxides participate in the formation of new particles, and SOA in the atmosphere [48,49,50], which have toxic effects on human health and plants [51,52,53], and some macromolecular hydroperoxides can act as CCN [39]. The reactions of CH3CHOO with GAS increase the carbon chain length and the ability to form hydrogen bonds, which has great potential in particle formation.

4. Materials and Methods

The 72 configurations of GAS were optimized, and the most stable configuration was selected as the reactant. The configuration optimization and transition state searching were conducted with the M06-2X method [54] in conjunction with the 6-311++G(2d,2p) basis set. The analysis of intrinsic reaction coordinates (IRC) [55,56] was used to verify that all transition states connect to the expected reactants and products. Single point energies of reactants, complexes, transition states, and products were calculated at the CCSD/6-311++G(2d,2p) level. All calculations of electronic structures were performed with Gaussian 09 program [57].
Thirty gas–liquid interface simulations were carried out by BOMD. All simulations were performed by CP2K [58] software based on the DFT method. The droplet consisted of 30 water molecules, and reactants were placed in a cubic box with the side length set as 35 Å. The Becke–Lee–Yang–Parr (BLYP) [59,60] functional, double-ζ Gaussian basis set (DZVP) [61] and Goedecker–Teter–Hutter (GTH) norm-conserved pseudopotentials [62,63] were adopted to handle electronic exchange interaction, valence and core electrons, respectively. All BOMD simulations were performed in the constant volume and temperature (NVT) ensemble with the integration step of 1 fs. The system temperature (300 K) was controlled using the Nose-Hoover chain method.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24043355/s1.

Author Contributions

Conceptualization, L.L., Y.W. and Q.W.; Data curation, L.L., Y.W. and Q.W.; Formal analysis, L.L.; Funding acquisition, Q.Z.; Methodology, L.L.; Supervision, W.W.; Writing—original draft, L.L. and Y.W.; Writing—review and editing, Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by NSFC (National Natural Science Foundation of China, project Nos. 22236004, 21976107, 42075106, 42175122).

Data Availability Statement

The configurations of the direct and water-mediated reaction stages (reaction complex, transition state and product).

Conflicts of Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

References

  1. Johnson, D.; Marston, G. The gas-phase ozonolysis of unsaturated volatile organic compounds in the troposphere. Chem. Soc. Rev. 2008, 37, 699–716. [Google Scholar] [CrossRef] [PubMed]
  2. Criegee, R. Mechanism of Ozonolysis. Angew. Chem. Int. Ed. 1975, 14, 745–752. [Google Scholar] [CrossRef]
  3. Taatjes, C.A.; Shallcross, D.E.; Percival, C.J. Research frontiers in the chemistry of Criegee intermediates and tropospheric ozonolysis. Phys. Chem. Chem. Phys. 2014, 16, 1704–1718. [Google Scholar] [CrossRef]
  4. Horie, O.; Moortgat, G.K. Gas-Phase Ozonolysis of Alkenes. Recent Advances in Mechanistic Investigations. Acc. Chem. Res. 1998, 31, 387–396. [Google Scholar] [CrossRef]
  5. Assaf, E.; Sheps, L.; Whalley, L.; Heard, D.; Tomas, A.; Schoemaecker, C.; Fittschen, C. The Reaction between CH3O2 and OH Radicals: Product Yields and Atmospheric Implications. Environ. Sci. Technol. 2017, 51, 2170–2177. [Google Scholar] [CrossRef]
  6. Asatryan, R.; Bozzelli, J.W. Formation of a Criegee intermediate in the low-temperature oxidation of dimethyl sulfoxide. Phys. Chem. Chem. Phys. 2008, 10, 1769–1780. [Google Scholar] [CrossRef]
  7. Lester, M.I.; Klippenstein, S.J. Unimolecular Decay of Criegee Intermediates to OH Radical Products: Prompt and Thermal Decay Processes. Acc. Chem. Res. 2018, 51, 978–985. [Google Scholar] [CrossRef]
  8. Novelli, A.; Vereecken, L.; Lelieveld, J.; Harder, H. Direct observation of OH formation from stabilised Criegee intermediates. Phys. Chem. Chem. Phys. 2014, 16, 19941–19951. [Google Scholar] [CrossRef]
  9. Elshorbany, Y.; Barnes, I.; Becker, K.H.; Kleffmann, J.; Wiesen, P. Sources and Cycling of Tropospheric Hydroxyl Radicals—An Overview. Z. Phys. Chem. 2010, 224, 967–987. [Google Scholar] [CrossRef]
  10. Stone, D.; Whalley, L.K.; Heard, D.E. Tropospheric OH and HO2 radicals: Field measurements and model comparisons. Chem. Soc. Rev. 2012, 41, 6348–6404. [Google Scholar] [CrossRef] [Green Version]
  11. Harrison, R.; Yin, J.; Tilling, R.; Cai, X.; Seakins, P.; Hopkins, J.; Lansley, D.; Lewis, A.; Hunter, M.; Heard, D.; et al. Measurement and modelling of air pollution and atmospheric chemistry in the U.K. West Midlands conurbation: Overview of the PUMA Consortium project. Sci. Total Environ. 2006, 360, 5–25. [Google Scholar] [CrossRef]
  12. Khan, M.A.H.; Percival, C.J.; Caravan, R.L.; Taatjes, C.A.; Shallcross, D.E. Criegee intermediates and their impacts on the troposphere. Environ. Sci. Process. Impacts 2018, 20, 437–453. [Google Scholar] [CrossRef] [PubMed]
  13. Docherty, K.S.; Ziemann, P.J. Effects of Stabilized Criegee Intermediate and OH Radical Scavengers on Aerosol Formation from Reactions of β-Pinene with O3. Aerosol Sci. Technol. 2003, 37, 877–891. [Google Scholar] [CrossRef]
  14. Ouyang, B.; McLeod, M.W.; Jones, R.L.; Bloss, W.J. NO3 radical production from the reaction between the Criegee intermediate CH2OO and NO2. Phys. Chem. Chem. Phys. 2013, 15, 17070–17075. [Google Scholar] [CrossRef]
  15. 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. USA 2015, 112, 10857–10862. [Google Scholar] [CrossRef]
  16. Zhou, X.; Chen, Y.; Liu, Y.; Li, X.; Dong, W.; Yang, X. Kinetics of CH2OO and Syn-CH3CHOO Reaction with Acrolein. Phys. Chem. Chem. Phys. 2021, 23, 13276–13283. [Google Scholar] [CrossRef]
  17. McGillen, M.R.; Curchod, B.F.; Chhantyal-Pun, R.; Beames, J.M.; Watson, N.P.; 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 Space Chem. 2017, 1, 664–672. [Google Scholar] [CrossRef]
  18. Reemtsma, T.; These, A.; Venkatachari, P.; Xia, X.; Hopke, P.K.; Springer, A.; Linscheid, M. Identification of Fulvic Acids and Sulfated and Nitrated Analogues in Atmospheric Aerosol by Electrospray Ionization Fourier Transform Ion Cyclotron Resonance Mass Spectrometry. Anal. Chem. 2006, 78, 8299–8304. [Google Scholar] [CrossRef]
  19. Surratt, J.D.; Kroll, J.H.; Kleindienst, T.E.; Edney, E.O.; Claeys, M.; Sorooshian, A.; Ng, N.L.; Offenberg, J.H.; Lewandowski, M.; Jaoui, M.; et al. Evidence for Organosulfates in Secondary Organic Aerosol. Environ. Sci. Technol. 2007, 41, 517–527. [Google Scholar] [CrossRef] [PubMed]
  20. Hansen, A.M.K.; Hong, J.; Raatikainen, T.; Kristensen, K.; Ylisirniö, A.; Virtanen, A.; Petäjä, T.; Glasius, M.; Prisle, N.L. Hygroscopic properties and cloud condensation nuclei activation of limonene-derived organosulfates and their mixtures with ammonium sulfate. Atmos. Meas. Tech. 2015, 15, 14071–14089. [Google Scholar] [CrossRef] [Green Version]
  21. Nozière, B.; Ekström, S.; Alsberg, T.; Holmström, S. Radical-initiated Formation of Organosulfates and Surfactants in Atmospheric Aerosols. Geophys. Res. Lett. 2010, 37, L05806. [Google Scholar] [CrossRef]
  22. Fleming, L.T.; Ali, N.N.; Blair, S.L.; Roveretto, M.; George, C.; Nizkorodov, S.A. Formation of Light-Absorbing Organosulfates during Evaporation of Secondary Organic Material Extracts in the Presence of Sulfuric Acid. ACS Earth Space Chem. 2019, 3, 947–957. [Google Scholar] [CrossRef]
  23. Romero, F.; Oehme, M. Organosulfates—A New Component of Humic-Like Substances in Atmospheric Aerosols? J. Atmos. Chem. 2005, 52, 283–294. [Google Scholar] [CrossRef]
  24. Surratt, J.D.; Gómez-González, Y.; Chan, A.W.H.; Vermeylen, R.; Shahgholi, M.; Kleindienst, T.E.; Edney, E.O.; Offenberg, J.H.; Lewandowski, M.; Jaoui, M.; et al. Organosulfate Formation in Biogenic Secondary Organic Aerosol. J. Phys. Chem. A 2008, 112, 8345–8378. [Google Scholar] [CrossRef]
  25. McNeill, V.F. Aqueous Organic Chemistry in the Atmosphere: Sources and Chemical Processing of Organic Aerosols. Environ. Sci. Technol. 2015, 49, 1237–1244. [Google Scholar] [CrossRef]
  26. McNeill, V.F.; Woo, J.L.; Kim, D.D.; Schwier, A.N.; Wannell, N.J.; Sumner, A.J.; Barakat, J.M. Aqueous-Phase Secondary Organic Aerosol and Organosulfate Formation in Atmospheric Aerosols: A Modeling Study. Environ. Sci. Technol. 2012, 46, 8075–8081. [Google Scholar] [CrossRef] [PubMed]
  27. Brüggemann, M.; Xu, R.; Tilgner, A.; Kwong, K.C.; Mutzel, A.; Poon, H.Y.; Otto, T.; Schaefer, T.; Poulain, L.; Nin Chan, M.; et al. Organosulfates in Ambient Aerosol: State of Knowledge and Future Research Directions on Formation, Abundance, Fate, and Importance. Environ. Sci. Technol. 2020, 54, 3767–3782. [Google Scholar] [CrossRef]
  28. Wang, Y.; Hu, M.; Guo, S.; Wang, Y.; Zheng, J.; Yang, Y.; Zhu, W.; Tang, R.; Li, X.; Liu, Y.; et al. The secondary formation of organosulfates under interactions between biogenic emissions and anthropogenic pollutants in summer in Beijing. Atmos. Meas. Tech. 2018, 18, 10693–10713. [Google Scholar] [CrossRef]
  29. Olson, C.N.; Galloway, M.M.; Yu, G.; Hedman, C.J.; Lockett, M.R.; Yoon, T.; Stone, E.A.; Smith, L.M.; Keutsch, F.N. Hydroxycarboxylic Acid-Derived Organosulfates: Synthesis, Stability, and Quantification in Ambient Aerosol. Environ. Sci. Technol. 2011, 45, 6468–6474. [Google Scholar] [CrossRef] [PubMed]
  30. Zhong, J.; Kumar, M.; Zhu, C.Q.; Francisco, J.S.; Zeng, X.C. Surprising Stability of Larger Criegee Intermediates on Aqueous Interfaces. Angew. Chem. Int. Ed. Engl. 2017, 56, 7740–7744. [Google Scholar] [CrossRef] [PubMed]
  31. Chhantyal-Pun, R.; McGillen, M.R.; Beames, J.M.; Khan, M.A.H.; Percival, C.J.; Shallcross, D.E.; Orr-Ewing, A.J. Temperature-Dependence of the Rates of Reaction of Trifluoroacetic Acid with Criegee Intermediates. Angew. Chem. Int. Ed. 2017, 56, 9044–9047. [Google Scholar] [CrossRef] [PubMed]
  32. Vereecken, L. The reaction of Criegee intermediates with acids and enols. Phys. Chem. Chem. Phys. 2017, 19, 28630–28640. [Google Scholar] [CrossRef] [PubMed]
  33. Buszek, R.J.; Francisco, J.S.; Anglada, J.M. Water effects on atmospheric reactions. Int. Rev. Phys. Chem. 2011, 30, 335–369. [Google Scholar] [CrossRef]
  34. Shang, J.; Passananti, M.; Dupart, Y.; Ciuraru, R.; Tinel, L.; Rossignol, S.; Perrier, S.; Zhu, T.; George, C. SO2 Uptake on Oleic Acid: A New Formation Pathway of Organosulfur Compounds in the Atmosphere. Environ. Sci. Technol. Lett. 2016, 3, 67–72. [Google Scholar] [CrossRef]
  35. Chhantyal-Pun, R.; Rotavera, B.; McGillen, M.R.; Khan, M.A.H.; Eskola, A.J.; Caravan, R.L.; Blacker, L.; Tew, D.P.; Osborn, D.L.; Percival, C.J.; et al. Criegee Intermediate Reactions with Carboxylic Acids: A Potential Source of Secondary Organic Aerosol in the Atmosphere. ACS Earth Space Chem. 2018, 2, 833–842. [Google Scholar] [CrossRef]
  36. Cornwell, Z.A.; Harrison, A.W.; Murray, C. Kinetics of the Reactions of CH2OO with Acetone, alpha-Diketones, and beta-Diketones. J. Phys. Chem. A 2021, 125, 8557–8571. [Google Scholar] [CrossRef] [PubMed]
  37. Elsamra, R.M.I.; Jalan, A.; Buras, Z.J.; Middaugh, J.E.; Green, W.H. Temperature and Pressure-Dependent Kinetics of CH2OO + CH3COCH3 and CH2OO + CH3CHO: Direct Measurements and Theoretical Analysis. Int. J. Chem. Kinet. 2016, 48, 474–488. [Google Scholar] [CrossRef]
  38. Ma, X.; Zhao, X.; Wei, Y.; Wang, W.; Xu, F.; Zhang, Q.; Wang, W. Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications. Sci. Total Environ. 2020, 715, 136812. [Google Scholar] [CrossRef]
  39. Kumar, M.; Zhong, J.; Zeng, X.C.; Francisco, J.S. Reaction of Criegee Intermediate with Nitric Acid at the Air–Water Interface. J. Am. Chem. Soc. 2018, 140, 4913–4921. [Google Scholar] [CrossRef]
  40. Hettiyadura, A.P.S.; Al-Naiema, I.M.; Hughes, D.D.; Fang, T.; Stone, E.A. Organosulfates in Atlanta, Georgia: Anthropogenic influences on biogenic secondary organic aerosol formation. Atmos. Meas. Tech. 2019, 19, 3191–3206. [Google Scholar] [CrossRef] [Green Version]
  41. Le Breton, M.; Wang, Y.; Hallquist, Å.M.; Pathak, R.K.; Zheng, J.; Yang, Y.; Shang, D.; Glasius, M.; Bannan, T.J.; Liu, Q.; et al. Online Gas- and Particle-phase Measurements of Organosulfates, Organosulfonates and Nitrooxy Organosulfates in Beijing Utilizing a FIGAERO ToF-CIMS. Atmos. Chem. Phys. 2018, 18, 10355–10371. [Google Scholar] [CrossRef]
  42. Wang, Y.; Zhao, Y.; Wang, Y.; Yu, J.-Z.; Shao, J.; Liu, P.; Zhu, W.; Cheng, Z.; Li, Z.; Yan, N.; et al. Organosulfates in atmospheric aerosols in Shanghai, China: Seasonal and interannual variability, origin, and formation mechanisms. Atmos. Meas. Tech. 2021, 21, 2959–2980. [Google Scholar] [CrossRef]
  43. Liu, J.; Liu, Y.; Yang, J.; Zeng, X.C.; He, X. Directional Proton Transfer in the Reaction of the Simplest Criegee Intermediate with Water Involving the Formation of Transient H3O+. J. Phys. Chem. Lett. 2021, 12, 3379–3386. [Google Scholar] [CrossRef]
  44. Qiu, J.; Ishizuka, S.; Tonokura, K.; Colussi, A.J.; Enami, S. Reactivity of Monoterpene Criegee Intermediates at Gas–Liquid Interfaces. J. Phys. Chem. A 2018, 122, 7910–7917. [Google Scholar] [CrossRef] [PubMed]
  45. Zhang, T.; Wen, M.; Ding, C.; Zhang, Y.; Ma, X.; Wang, Z.; Lily, M.; Liu, J.; Wang, R. Multiple evaluations of atmospheric behavior between Criegee intermediates and HCHO: Gas-phase and air-water interface reaction. J. Environ. Sci. 2023, 127, 308–319. [Google Scholar] [CrossRef]
  46. Zhu, C.; Kumar, M.; Zhong, J.; Li, L.; Francisco, J.S.; Zeng, X.C. New Mechanistic Pathways for Criegee–Water Chemistry at the Air/Water Interface. J. Am. Chem. Soc. 2016, 138, 11164–11169. [Google Scholar] [CrossRef] [PubMed]
  47. Ma, X.; Zhao, X.; Huang, Z.; Wang, J.; Lv, G.; Xu, F.; Zhang, Q.; Wang, W. Determination of reactions between Criegee intermediates and methanesulfonic acid at the air-water interface. Sci. Total Environ. 2020, 707, 135804. [Google Scholar] [CrossRef]
  48. Docherty, K.S.; Wu, W.; Bin Lim, Y.; Ziemann, P.J. Contributions of Organic Peroxides to Secondary Aerosol Formed from Reactions of Monoterpenes with O3. Environ. Sci. Technol. 2005, 39, 4049–4059. [Google Scholar] [CrossRef]
  49. Ehn, M.; Thornton, J.A.; Kleist, E.; Sipilä, M.; Junninen, H.; Pullinen, I.; Springer, M.; Rubach, F.; Tillmann, R.; Lee, B.; et al. A large source of low-volatility secondary organic aerosol. Nature 2014, 506, 476–479. [Google Scholar] [CrossRef]
  50. Li, H.; Chen, Z.; Huang, L.; Huang, D. Organic peroxides’ gas-particle partitioning and rapid heterogeneous decomposition on secondary organic aerosol. Atmos. Meas. Tech. 2016, 16, 1837–1848. [Google Scholar] [CrossRef] [Green Version]
  51. Marklund, S. Actions of Hydroxymethylhydroperoxide and Bis(Hydroxymethyl)Peroxide on Fumarate Hydratase, Lactate-Dehydrogenase, Aspartate Aminotransferase, Glucose Oxidase, and Acid-Phosphatase. Biochim. Biophys. Acta Biomembr. 1972, 258, 9–16. [Google Scholar] [CrossRef] [PubMed]
  52. Marklund, S. Mechanisms of the irreversible inactivation of horseradish peroxidase caused by hydroxymethylhydroperoxide. Arch. Biochem. Biophys. 1973, 154, 614–622. [Google Scholar] [CrossRef]
  53. Wang, Z.; Herbinet, O.; Hansen, N.; Battin-Leclerc, F. Exploring hydroperoxides in combustion: History, recent advances and perspectives. Prog. Energy Combust. Sci. 2019, 73, 132–181. [Google Scholar] [CrossRef]
  54. 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 M06-class functionals and 12 other functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  55. Fukui, K. The path of chemical reactions—The IRC approach. Acc. Chem. Res. 1981, 14, 363–368. [Google Scholar] [CrossRef]
  56. Hratchian, H.P.; Schlegel, H.B. Accurate reaction paths using a Hessian based predictor–corrector integrator. J. Chem. Phys. 2004, 120, 9918–9924. [Google Scholar] [CrossRef]
  57. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision B.01; Gaussian Inc.: Wallingford, CT, USA, 2010. [Google Scholar]
  58. 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. [Google Scholar] [CrossRef]
  59. Becke, A.D. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  60. 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. [Google Scholar] [CrossRef]
  61. VandeVondele, J.; Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007, 127, 114105. [Google Scholar] [CrossRef] [Green Version]
  62. Goedecker, S.; Teter, M.; Hutter, J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996, 54, 1703–1710. [Google Scholar] [CrossRef] [PubMed]
  63. Hartwigsen, C.; Goedecker, S.; Hutter, J. Relativistic separable dual-space Gaussian pseudopotentials from H to Rn. Phys. Rev. B 1998, 58, 3641–3662. [Google Scholar] [CrossRef] [Green Version]
Figure 1. The most stable configuration of GAS.
Figure 1. The most stable configuration of GAS.
Ijms 24 03355 g001
Figure 2. The minimum energy pathway of the direct reactions between CH3CHOO and GAS.
Figure 2. The minimum energy pathway of the direct reactions between CH3CHOO and GAS.
Ijms 24 03355 g002
Figure 3. The minimum energy pathway of the water-mediated reactions between CH3CHOO and GAS.
Figure 3. The minimum energy pathway of the water-mediated reactions between CH3CHOO and GAS.
Ijms 24 03355 g003
Figure 4. The direct and water-mediated reactions between anti-CH3CHOO and COOH group of GAS at gas–liquid interface ((a): direct reaction; (b): water-mediated reaction).
Figure 4. The direct and water-mediated reactions between anti-CH3CHOO and COOH group of GAS at gas–liquid interface ((a): direct reaction; (b): water-mediated reaction).
Ijms 24 03355 g004
Figure 5. The direct and water-mediated reactions between anti-CH3CHOO and OSO3H group of GAS at gas–liquid interface ((a): direct reaction; (b): water-mediated reaction).
Figure 5. The direct and water-mediated reactions between anti-CH3CHOO and OSO3H group of GAS at gas–liquid interface ((a): direct reaction; (b): water-mediated reaction).
Ijms 24 03355 g005
Figure 6. The intramolecular proton transfer reaction between anti-CH3CHOO and OSO3H group of GAS at gas–liquid interface.
Figure 6. The intramolecular proton transfer reaction between anti-CH3CHOO and OSO3H group of GAS at gas–liquid interface.
Ijms 24 03355 g006
Figure 7. The GAS-mediated hydration of anti-CH3CHOO at gas–liquid interface.
Figure 7. The GAS-mediated hydration of anti-CH3CHOO at gas–liquid interface.
Ijms 24 03355 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Li, L.; Zhang, Q.; Wei, Y.; Wang, Q.; Wang, W. Theoretical Study on the Gas Phase and Gas–Liquid Interface Reaction Mechanism of Criegee Intermediates with Glycolic Acid Sulfate. Int. J. Mol. Sci. 2023, 24, 3355. https://doi.org/10.3390/ijms24043355

AMA Style

Li L, Zhang Q, Wei Y, Wang Q, Wang W. Theoretical Study on the Gas Phase and Gas–Liquid Interface Reaction Mechanism of Criegee Intermediates with Glycolic Acid Sulfate. International Journal of Molecular Sciences. 2023; 24(4):3355. https://doi.org/10.3390/ijms24043355

Chicago/Turabian Style

Li, Lei, Qingzhu Zhang, Yuanyuan Wei, Qiao Wang, and Wenxing Wang. 2023. "Theoretical Study on the Gas Phase and Gas–Liquid Interface Reaction Mechanism of Criegee Intermediates with Glycolic Acid Sulfate" International Journal of Molecular Sciences 24, no. 4: 3355. https://doi.org/10.3390/ijms24043355

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop