Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications

Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications

Journal Pre-proof Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications Xiaohui Ma,...

1MB Sizes 0 Downloads 10 Views

Journal Pre-proof Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications

Xiaohui Ma, Xianwei Zhao, Yuanyuan Wei, Wei Wang, Fei Xu, Qingzhu Zhang, Wenxing Wang PII:

S0048-9697(20)30322-3

DOI:

https://doi.org/10.1016/j.scitotenv.2020.136812

Reference:

STOTEN 136812

To appear in:

Science of the Total Environment

Received date:

29 November 2019

Revised date:

30 December 2019

Accepted date:

18 January 2020

Please cite this article as: X. Ma, X. Zhao, Y. Wei, et al., Effect of multifunctional compound monoethanolamine on Criegee intermediates reactions and its atmospheric implications, Science of the Total Environment (2018), https://doi.org/10.1016/ j.scitotenv.2020.136812

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2018 Published by Elsevier.

Journal Pre-proof

Effect of Multifunctional Compound Monoethanolamine on Criegee Intermediates Reactions and Its Atmospheric Implications Xiaohui Ma, Xianwei Zhao, Yuanyuan Wei, Wei Wang, Fei Xu, Qingzhu Zhang*, Wenxing Wang

of

Environment Research Institute, Shandong University,

Jo ur

na

lP

re

-p

ro

Qingdao 266237, P. R. China

Keywords: Criegee Intermediates; Multifunctional Compound; Reaction Rates; BOMD ___________________________________________________________ *

Corresponding authors. E-mail: [email protected]

Fax: 86-531-8836 1990 1

Journal Pre-proof

Abstract The reactions of Criegee intermediates with trace gases (such as alcohols, amines, and acids) are primarily dependent on the trace gases’ functional group activity. In this study, we used density functional theory calculations and ab initio dynamics simulation methods to explore the synergistic effect of NH2 and OH groups, in the

of

multifunctional compound monoethanolamine (MEA), on the Criegee reaction. The results showed that among the four evaluated MEA configurations, two functional

ro

groups in the g′Gg′ and tGg′ configurations, -NH2 and -OH, have the synergistic effect

-p

on the C2 stabilized Criegee intermediates (sCIs). At the gas-liquid interface, sCIs

re

react with NH2 groups of MEA molecules directly or are mediated by water molecules,

lP

resulting in additional product formation. The rate calculation indicated that the

na

reaction of sCIs with NH2 groups of MEA molecules is prior to that with OH groups. In addition, OH groups promote the reactions between sCIs and NH2 groups of MEA,

Jo ur

while the presence of NH2 groups weakens the reactions of sCIs and OH groups of MEA to some extent. At 298 K, the total rate constant of anti-CH3CHOO with NH2 group of MEA is 4.26×10-11 cm3 molecule-1 s-1, which is four orders of magnitude higher than that of anti-CH3CHOO hydration. Under low humidity conditions, the reactions between sCIs and MEA could contribute to the removal of sCIs.

2

Journal Pre-proof

1 Introduction Criegee chemistry has always attracted significant attention from the research community due to its contribution to hydroxyl radical cycling, SO2 oxidation, and aerosol formation. Criegee intermediates in the atmosphere are derived from

of

ozonolysis of alkenes, during which ozone molecules attack the alkenes’ double bond

ro

to form primary epoxides. The epoxides then rapidly decompose, which leads to

-p

carbonyl compound and carbonyl oxide formation (i.e., Criegee intermediates). The

re

Criegee intermediate formation is accompanied by a large amount of vibration energy, and 37 - 50% intermediates undergo rapid single molecule cleavage. This results in

lP

the formation of CO, CO2, organic acids, and OH radicals, while the remaining

na

intermediates lose energy via collisions to form stabilized Criegee intermediates (sCIs)

Jo ur

(Anglada et al., 1999; Gutbrod et al., 1997; Gutbrod et al., 1996; Horie and Moortgat, 1998; Johnson and Marston, 2008; Neeb et al., 1998). The latter is removed by reactions with gaseous species in the atmosphere. Among the Criegee’s bimolecular reactions, hydration is considered the main removal channel due to the abundance of tropospheric water molecules. The simplest sCI (CH2OO) reacts with H2O monomers at rate constants of 10-16 to 10-17 cm3 molecule-1 s-1 (Berndt et al., 2015; Stone et al., 2014). The addition of a second H2O molecule results in significantly decrease for hydration barrier and a faster rate constant of 6.5×10-12 cm3 molecule-1 s-1 (Chao et al., 2015). Recent studies have shown that two or more methyl-substituted sCIs can stay at the water interface for a long time, which provides an environment for sCIs to react 3

Journal Pre-proof

with other trace gases (Zhong et al., 2017). The sCIs bimolecular reaction scavenges for numerous trace gases in the atmosphere. In addition to the hydroxyl radical, sCIs are also powerful oxidants that are closely related to sulfate and nitrate formation from SO2 and NO2 , respectively (Berndt et al., 2012; Berndt et al., 2014; Boy et al., 2013; Li et al., 2013; Mauldin et

of

al., 2012; Welz et al., 2012). Moreover, the reactions between sCIs and acids are close to the collision limit, and thus serve as the main removal channel of acids in the

ro

atmosphere (Chhantyal-Pun et al., 2017; Chhantyal-Pun et al., 2018; Kurten et al.,

-p

2007; Welz et al., 2014). Due to their zwitterionic property, sCIs can also interact with

re

alkaline gases. In such reactions, the NH2 group’s N atom in the alkaline gas

lP

combines with the sCI’s C atom. The NH2 group transfers a H atom to the sCI’s

na

terminal oxygen to form a functionalized hydroperoxide adduct (Chao et al., 2019; Chhantyal-Pun et al., 2019; Jørgensen and Gross, 2009; Kumar and Francisco, 2019).

Jo ur

Furthermore, hydroxyl species also exhibit high activity for the sCIs (McGillen et al., 2017; Tadayon et al., 2018; Watson et al., 2019). Studies have shown that in the presence of H2O monomers or at a gas-liquid interface, sCIs react with alcohols in competition with H2O molecules (Enami and Colussi, 2017; Lin et al., 2018). For the reactions with C13 sCIs, relative rate constant measurements indicate that heptanoic acid and formic acid exhibit the highest reactivity, followed by formaldehyde, 2-propanol, methanol, and finally water (Tobias and Ziemann, 2001). Adducts containing sCI chain units have a demonstrated ability to form aerosols (Inomata et al., 2014; Sadezky et al., 2008; Sakamoto et al., 2013). Recent research has focused on 4

Journal Pre-proof

the reactions of sCIs with a specific class of functional group substances, but the mechanism research on multifunctional group substances is still lacking. Monoethanolamine (MEA) is a potential Criegee reaction receptor because of its basic amino group and alcoholic properties; as NH2 and OH functional groups can react with the sCIs. MEA is an excellent adsorption solvent for post-combustion

of

CO2 capture, and MEA volatilization is the main MEA atmospheric supply source. Research has shown that for each tonne of captured CO2, 0.1-0.8 kg MEA will be

ro

released into the atmosphere (Karl et al., 2011; Veltman et al., 2010). The alkalinity is

-p

a key factor in determining the reactivity of the sCIs with the NH2 group. The stronger

re

the alkalinity, the lower the reaction barrier (Kumar and Francisco, 2019). From the

lP

standpoint of alkalinity, MEA is stronger than NH3 and weaker than methylamine

na

(MA). Therefore, the reactivity of sCIs with MEA is between those of sCIs with NH3 and MA, i.e., MA > MEA > NH3. With respect to structure, the OH group, on one

Jo ur

hand, weakens the basicity of the amine, and on the other hand, acts as a sCI hydrogen bond donor, increasing the reaction complexity. For example, when the sCIs react with an OH group, the NH2 group of MEA can form intermolecular interaction with the sCI’s O atom, thereby impacting the reaction process. In addition, reactions with sCIs host competition between the NH2 group and OH groups in MEA. In this study, we explored the reaction mechanisms between sCIs and difunctional compound MEA in the gas-phase and at the gas-liquid interface using density functional theory (DFT) calculations and ab initio dynamics simulations. Four MEA configurations were selected based on the relative positions of the NH2 and OH 5

Journal Pre-proof

groups within it. The reaction mechanisms between NH2 groups of MEA and sCIs were compared with those between OH groups of MEA and sCIs. Also, the reaction rate constants of sCIs reacting with MEA were compared with those of sCIs reacting with H2O, CH3CH2NH2 and CH3CH2OH to evaluate the potential atmospheric significance. Finally, the interactions between sCIs and MEA at the gas-liquid

of

interface were simulated. This study further expands the understanding of sCIs’ reactions with complex atmospheric gases and demonstrating Criegee's atmospheric

-p

ro

behavior.

re

2 Methods

lP

MEA has 13 non-equivalent configurations (Vorobyov et al., 2002; Wang et

na

al., 2009) that are distinguished by the Radom nomenclature (Radom et al., 1973).

Jo ur

Four representative configurations (g′Gg′, tGt, tGg′, and tTg) were selected to investigate the reactions between sCIs and MEA. Among the selected configurations, the g′Gg′ configuration is the most stable in the gas phase. We selected the C2 sCIs as the reactants, and both the syn- and anti- structures were considered. In this study, the structure optimization and the transition state search were performed using the M06-2X method (Zhao and Truhlar, 2008) coupled with the 6-311++G(3df,3pd) basis set. The local minimum points and the transition states were confirmed by the frequency calculation. No imaginary frequency was observed for the minimum points and one was observed for the transition states. The zero-point energy (ZPE) correction values were obtained at the same theoretical level. The intrinsic 6

Journal Pre-proof

reaction coordinates (IRC) (Fukui, 1981; Hratchian and H Bernhard, 2004) calculations demonstrated that the minimum energy pathway connects to the reaction complex and product. The structure energy was further determined by the ab initio CCSD(T) method combined with the cc-pVTZ basis set. All electronic structures were calculated using the Gaussian 09 program (Ortiz et al., 2009). The reactions between

of

sCIs and MEA begin with the formation of a reaction complex and then overcome a reaction barrier to form a substituted hydroperoxide. The reaction pathway can be

ro

expressed as Keq

-p

Kuni  A  B  A+B   C

re

where Keq is the equilibrium constant between the reactant molecules and the reaction

lP

complex, and Kuni is the unimolecular reaction rate constant between the complex and

na

the product. The total reaction rate can be described as

k  KeqKuni

Jo ur

The canonical variational transition state theory (CVT) with small-curvature tunneling (SCT) correction was adopted to calculate the rate constant, which was carried out by the Polyrate 9.7 program (Corchado et al., 2007). The reaction of sCIs with MEA at the gas-liquid interface was simulated by the Born-Oppenheimer molecular dynamics (BOMD) method. Based on DFT and the plane-wave method, BOMD simulations have shown outstanding performance in exploring the bond cleavage and formation in the reaction process and have been extensively applied in the gas-liquid interface reactions (Kumar et al., 2017; Zhong et al., 2018; Zhong et al., 2019; Zhong et al., 2015). In these dynamics simulations, we 7

Journal Pre-proof

employed Becke-Lee-Yang-Par (BLYP) functionals (Becke, 1988; Lee et al., 1988) to address the electronic exchange interactions. The Grimme’s dispersion correction (D3 correction) (Grimme et al., 2010) was added to better reveal weak intermolecular interactions. Valence electrons and core electrons were treated with double-ζ Gaussian basis set (DZVP) (VandeVondele and Hutter, 2007) and Goedecker-Teter-Hutter

of

(GTH) norm-conserved pseudopotentials (Goedecker et al., 1996; Hartwigsen et al., 1998), respectively. A cutoff energy of 280 Ry was programmed for the plane-wave

ro

base set and 40 Ry was selected for the Gaussian basis set. In this study, the constant

-p

temperature and constant volume NVT ensemble was used and the temperature was

re

maintained at 300 K by a Nose-Hoover thermostat. The droplet system, which

lP

consisted of 191 H2O molecules, was equilibrated by BOMD simulation for 6 ps. A

na

three-dimensional unit cell (35×35×35 Å3) was chosen to avoid the interaction between adjacent repeat units, as this method has proven successful in previous

Jo ur

studies (Kumar et al., 2018; Li et al., 2016). The simulation time step was set to 1 fs. All BOMD simulations were carried out using the CP2K program (VandeVondele et al., 2005).

3 Results and Discussion The four MEA configurations selected for this study are shown in Figure 1. The g′Gg′ configuration is characterized by an O-H···O intramolecular hydrogen bond, while an N-H···O hydrogen bond is formed in the tGt configuration. Based on the g′Gg′ configuration, the energies of tGt, tGg′, and tTg are 1.71, 2.51 and 2.82 8

Journal Pre-proof kcal/mol higher than that of g′Gg′, respectively.

3.1 Reaction of sCI with NH2 Group Figure 2 depicts the C2 sCIs’ potential energy profile while the NH2 of MEA acts as a reactive group. The complex formation from g′Gg′ and syn-CH3CHOO emits 4.14 kcal/mol of energy, which is the least exothermic of the examined

of

complexes. Using the reaction complexes as the benchmark, the lowest reaction

ro

energy barriers were observed in the reactions of syn-CH3CHOO with g′Gg′ (4.41

-p

kcal/mol) and tGt (4.11 kcal/mol), followed by the reactions with tGg′ (5.85 kcal/mol)

re

and tTg (7.57 kcal/mol). From the configuration perspective, in addition to NH2 group

lP

interaction with the syn-CH3CHOO, an O-H···O intermolecular interaction was also observed in the g′Gg′ and syn-CH3CHOO complex, as shown in Figure S1. However,

na

the distance between the H atom of NH2 group and the terminal O atom of

Jo ur

syn-CH3CHOO can reach 3.51 Å, which also explains the complex’s lower exothermic value. In the hydroperoxide product molecules, O-H···N intramolecular hydrogen bonds were formed and an additional O-H···O hydrogen bond was observed in the product comprised of g′Gg′ and CH3CHOO. When the OH group of MEA was replaced by a H atom, the reaction barrier associated with the syn-CH3CHOO and CH3CH2NH2 reaction was 4.43 kcal/mol. This value is close to the barrier values of the g′Gg′ and tGt configurations and slightly lower than that of the tGg′ and tTg configurations, which indicates that the OH group has little effect on the reaction energy barrier of syn-CH3CHOO with the NH2 group of MEA.

9

Journal Pre-proof

With respect to the reactions of anti-CH3CHOO with NH2 groups of tGt or tTg molecules, the OH groups do not directly interact with the CH3CHOO, thus the heat release from complex formation and the transition state barrier are similar. When g′Gg′ and tGg′ participated in the reactions, an additional intermolecular hydrogen bond between the OH group of MEA and the terminal O atom of anti-CH3CHOO was

of

formed. In this case, the heat released from complex formation and the transition state barrier were thermodynamically more advantageous than that of the reactions

ro

involving tGt and tTg. It is worth noting that during complex formation, the tGg′ NH2

-p

group’s orientation was reversed due to the attraction between tGg′ and

re

anti-CH3CHOO, resulting in maximum heat release. When the OH group of MEA

lP

was replaced by a H atom, its reaction barrier with anti-CH3CHOO was 3.56 kcal/mol,

na

which is higher than the barrier associated with anti-CH3CHOO reacting with all four MEA configurations. These results indicate that the OH group of MEA promotes the

Jo ur

addition of NH2 to anti-CH3CHOO, which is most evident in the reactions involving the g′Gg′ and tGg′ configurations. Furthermore, the reaction between anti-CH3CHOO and the NH2 group is thermodynamically more favorable than that of syn-CH3CHOO. For example, the heat released from complex formation of anti-CH3CHOO and g′Gg′ (-10.19 kcal/mol) was greater than that of the corresponding complex from syn-CH3CHOO and g′Gg′ (-4.14 kcal/mol); while the reaction energy barrier of the anti-CH3CHOO and g′Gg′ (0.57 kcal/mol) was lower than that of syn-CH3CHOO and g′Gg′ (4.41 kcal/mol). This phenomenon occurs mainly because the anti-CH3CHOO configuration is more stretched than the syn-CH3CHOO, and the steric hindrance of 10

Journal Pre-proof

NH2 addition to the CH3CHOO C atom is much smaller.

3.2 Reaction of sCI with OH Group Figure 3 depicts the potential energy profile while the OH group of MEA reacts with the sCIs. The reaction complex structures, transition state, and products are shown in Figure S3. When the OH group faces toward the syn-CH3CHOO active

of

center, the NH2 groups of g′Gg′ and tGg′ form additional hydrogen bonds with the

ro

terminal O atom of CH3CHOO. In the transition state, the two sets of N-H···O

-p

interactions make the OH group of tGg′ configuration combined with the

re

syn-CH3COO the most favorable reaction with the lowest barrier (9.31 kcal/mol). In

lP

the same way, the NH2 group of MEA was replaced with a H atom. The reaction between syn-CH3COO and CH3CH2OH was induced by the reaction complex. After

na

releasing 9.96 kcal/mol of energy, it overcame the 8.95 kcal/mol energy barrier and

Jo ur

formed corresponding hydroperoxides. Thus, the CH3CH2OH barrier is more thermodynamically favorable than those of the four MEA configurations, demonstrating that the NH2 group has little effect on the reactions between the syn-CH3CHOO and OH groups of MEA molecules. Compared with the reactions between the syn-CH3CHOO and OH groups of MEA, the reactions with anti-CH3CHOO were more accessible. Among the formed reaction complexes, those in which anti-CH3CHOO was involved were more exothermic than complexes with the corresponding syn-CH3CHOO. For example, formation of the anti-CH3CHOO and g′Gg′ complex released 10.30 kcal/mol of

11

Journal Pre-proof energy, while formation of the syn-CH3CHOO and g′Gg′ complex released 9.05 kcal/mol of energy. Structurally, the C-O distance in the anti-CH3CHOO and MEA (ex: tGg′) complex (2.75 Å) was significantly smaller than that in the syn-CH3CHOO and MEA (ex: tGg′) complex (3.09 Å). Since the steric hindrance of the OH group addition to the anti-CH3CHOO is smaller than addition to the syn-CH3CHOO, the

of

reactive energy barriers for the reactions between anti-CH3CHOO and MEA were also

ro

lower than those for the reactions between syn-CH3CHOO and MEA.

-p

3.3 Reactions at the Gas-liquid Interface

re

The gas-liquid interface is an important reaction medium in atmospheric

lP

chemistry, as it facilitates formation of HONO induced by N2O4 and H2O clusters (de

na

Jesus Medeiros and Pimentel, 2011; Miller et al., 2009), reactions between NOx and HCl at the droplet interface (Hammerich et al., 2012; Njegic et al., 2010), and

Jo ur

photochemical reactions at the water and ice interfaces (Anglada et al., 2014). In this work, the high activity anti-CH3CHOO and the most stable g′Gg′ configuration were selected to investigate the reaction mechanism between sCIs and MEA at the gas-liquid interface. Ten simulations were carried out to avoid effects of the initial conformations at the interface. The reaction trajectory and key bond length evolution are shown in Figure 4. At the beginning of the reaction, the C-N and O-H distances in the anti-CH3CHOO and g′Gg′ were 4.23 and 3.79 Å, respectively, indicating there was no direct interaction between the two reaction molecules (Figure 4a). At 1.18 ps, a transition-like structure appeared, in which the N-H, O-H, and C-N distances were 12

Journal Pre-proof

1.28, 1.30, and 1.47 Å, respectively. Subsequently, the O-H distance was maintained at around 1.00 Å, indicating the formation of a new OH bond. In addition, water molecules at the interface could participate in the anti-CH3CHOO and g′Gg′ reaction (Figure 4b). After 2.85 ps, the H1-O1 and H2-O2 distances fluctuated at around 2.50 and 1.60 Å, respectively, indicating the formation of a g′Gg′-H2O-CH3CHOO complex. A transition-like structure occurred at 3.65 ps, in which the H1 atom from

of

the NH2 in the g′Gg′ configuration was transferred to a nearby H2O molecule, while

ro

the H2O molecule released its own H2 atom to the terminal O2 atom of

-p

anti-CH3CHOO. Simultaneously, the active C atom of anti-CH3CHOO combined

re

with the N atom of g′Gg′, and formed a new addition product. In the reaction, water

lP

molecule acted as a proton transfer bridge. In both cases, the proton transfer and the

na

C-N bond formation were simultaneously completed, and no step mechanism was observed. Due to the reactivity of OH groups is weaker than that of NH2 groups, no

interface.

Jo ur

reaction was observed between the anti-CH3CHOO and the OH groups at the droplet

4 Reaction Rates

In this study, the reaction rate constants between sCIs and two MEA functional groups (-NH2 and -OH) were assessed using the CVT/SCT correction method.

The

kinetic

parameters

were

obtained

at

the

CCSD(T)/cc-pVTZ//M06-2X/6-311++G(3df,3pd) theoretical level. The equilibrium constants, unimolecular rate constants, and total rate constants for the CH3CHOO 13

Journal Pre-proof

with MEA reactions are listed in Table S3-S5. As shown in Table S3, the rate constants of syn-CH3CHOO with NH2 groups of the four MEA molecules change from 10-15 to 10-17 cm3 molecule-1 s-1. At 298 K, the rate constant of most stable g′Gg′ configuration is 4.91×10-17 cm3 molecule-1 s-1, which is slightly higher than the value of syn- sCIs and H2O monomers (10-18-10-19 cm3 molecule-1 s-1) (Anglada et al., 2011; Long et al., 2016; Long et al., 2018). For the reaction of anti-CH3CHOO with the NH2

of

group of most stable g′Gg′ configuration, the rate constant is as high as 5.75×10-11

ro

cm3 molecule-1 s-1, while the reaction with tGg′ is close to the bimolecular collision

-p

limit. At 298K, the rate constant of anti-CH3CHOO and tGg′ NH2 group obtained

re

from the collision theory is 1.74×10-10 cm3 molecule-1 s-1.

lP

In order to determine the total rate constant of the four MEA configurations

na

with C2 sCIs, we derived the proportions of all 13 configurations according to the

Jo ur

Boltzmann formula, which is mathematically expressed as follows: wi 

e

 ( Gi  Gmin )/ RT

e

 ( Gi Gmin )/ RT

i

where Gi is the free energy of each configuration, Gmin is the free energy of the most stable g′Gg′ configuration, R is the gas constant, and T is the absolute temperature. To obtain more accurate conformational proportions of MEA, the free energy values were refined at the CCSDT/aug-cc-pVTZ//M06-2X/6-311++G(3df,3pd) level. As shown in Table S6, the most stable g′Gg′ configuration accounts for 61% of the total MEA concentration at 298 K, which is consistent with the results from Xie's study (Xie et al., 2014). The four MEA configuration concentrations in this study account for 72% of the total MEA concentration at 300 K. The total rate constants of 14

Journal Pre-proof

CH3CHOO with four MEA configurations at the different temperatures are listed in Table 1. The reaction of sCIs with NH2 groups is 2-3 orders of magnitude faster than with OH groups. For example, at 300K, the rate constant of syn-CH3CHOO with NH2 groups of MEA is 6.19×10-16 cm3 molecule-1 s-1, and the rate constant with the OH group is 9.29×10-19 cm3 molecule-1 s-1. This indicates that C2 sCIs react more

of

preferentially with NH2 groups in MEA configurations. Moreover, compared with syn configuration sCIs, the reactivity of anti configuration sCIs is significantly higher. At

ro

298 K, the reaction rate constant of anti-CH3CHOO and NH2 group of MEA is

-p

4.26×10-11 cm3 molecule-1 s-1, which is four orders of magnitude higher than that of

re

anti-CH3CHOO hydration (5.21×10-15 cm3 molecule-1 s-1) (Long et al., 2016). Under

na

removal of sCIs.

lP

low humidity conditions, the reactions between sCIs and MEA could contribute to the

To elucidate the effect of additional functional groups on the Criegee

Jo ur

reactions, the reaction rate constants of C2 sCIs with CH3CH2NH2 and CH3CH2OH were evaluated, as shown in Table S7. When the OH groups of MEA are replaced by H atoms, the reaction rate constants of C2 sCIs and CH3CH2NH2 are slightly lower than those of sCIs and NH2 groups of MEA. For example, at 298K, the rate constant of anti-CH3CHOO + CH3CH2NH2 reaction is estimated to be 1.60×10-12 cm3 molecule-1 s-1, while the reaction rate constant of anti-CH3CHOO with NH2 groups of MEA is 4.26×10-11 cm3 molecule-1 s-1, which indicates that OH groups promote the reactions of sCIs and NH2 groups to some extent. The reactions of sCIs with OH groups of MEA are just the opposite. When the NH2 groups of MEA are replaced by 15

Journal Pre-proof

H atoms, the reaction rate constants of sCIs and CH3CH2OH are one~three order of magnitude higher than those of corresponding sCIs with OH groups of MEA, indicating that the presence of NH2 groups weakens the reactions between sCIs and OH groups of MEA to some extent.

of

5 Conclusions

ro

In this study, we explored the reaction mechanisms of C2 sCIs with NH2

-p

and OH groups of MEA in the gas phase. In addition, the interaction between

re

anti-CH3CHOO and the g′Gg′ configuration at the gas-liquid interface was investigated using ab initio dynamics simulation methods. According to the reaction

lP

type and rate comparison, we can draw the following conclusions:

na

(1) When C2 sCIs react with the NH2 group in the four MEA configurations, only the

Jo ur

OH groups of the g′Gg′ and tGg′ configurations participate in the interactions with sCIs. Similarly, when C2 sCIs react with the OH group in MEA configurations, only the NH2 group of the g′Gg′ and tGg′ configurations participate in the interactions with sCIs.

(2) The reaction of sCIs with NH2 groups of MEA is prior to that with OH groups. In addition, for the reaction with NH2 groups, the activity of anti-CH3CHOO is much higher than that of syn-CH3CHOO. (3) It can be inferred from the comparison of rate constants that OH groups promote the reactions between sCIs and NH2 groups of MEA, while NH2 groups could inhibite the activity of sCIs and OH groups of MEA to some extent. 16

Journal Pre-proof

(3) At 298 K, the total rate constant of anti-CH3CHOO with four MEA configurations is 4.26×10-11 cm3 molecule-1 s-1, which is four orders of magnitude higher than that of anti-CH3CHOO hydration. In a dry environment, the reactions between sCIs and MEA could contribute to the removal of sCIs.

of

Acknowledgment

ro

This work was supported by NSFC (National Natural Science Foundation of China, Project No. 21677089) and Taishan Scholars (No. ts201712003)

-p

References

re

Anglada JM, Crehuet R, Bofill JM. The Ozonolysis of ethylene: a theoretical study of

lP

the gas-phase reaction mechanism. Chem.- Eur. J. 1999; 5: 1809-1822.

na

Anglada JM, Gonzalez J, Torrent-Sucarrat M. Effects of the substituents on the reactivity of carbonyl oxides. a theoretical study on the reaction of substituted

Jo ur

carbonyl oxides with water. Phys. Chem. Chem. Phys. 2011; 13: 13034-13045. Anglada JM, Martins-Costa M, Ruiz-López MF, Francisco JS. Spectroscopic signatures of ozone at the air-water interface and photochemistry implications. Proc. Natl. Acad. Sci. U. S. A. 2014; 111: 11618-11623. Becke AD. Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988; 38: 3098. Berndt T, Jokinen T, Mauldin RL, Petäjä T, Herrmann H, Junninen H, et al. Gas-phase ozonolysis of selected olefins: the yield of stabilized Criegee intermediate and the reactivity toward SO2. J. Phys. Chem. Lett. 2012; 3: 2892-2896. 17

Journal Pre-proof

Berndt T, Jokinen T, Sipilä M, Mauldin RL, Herrmann H, Stratmann F, et al. H2SO4 formation from the gas-phase reaction of stabilized Criegee intermediates with SO2: influence of water vapour content and temperature. Atmos. Environ. 2014; 89: 603-612. Berndt T, Kaethner R, Voigtländer J, Stratmann F, Pfeifle M, Reichle P, et al. Kinetics

of

of the unimolecular reaction of CH2OO and the bimolecular reactions with the water monomer, acetaldehyde and acetone under atmospheric conditions. Phys.

ro

Chem. Chem. Phys. 2015; 17: 19862-19873.

-p

Boy M, Mogensen D, Smolander S, Zhou L, Nieminen T, Paasonen P, et al. Oxidation

re

of SO2 by stabilized Criegee intermediate (sCI) radicals as a crucial source for

na

3865-3879.

lP

atmospheric sulfuric acid concentrations. Atmos. Chem. Phys. 2013; 13:

Chao W, Hsieh J-T, Chang C-H. Direct kinetic measurement of the reaction of the

Jo ur

simplest Criegee intermediate with water vapor. Science 2015; 347: 751-754. Chao W, Yin C, Takahashi K, Lin JJ. Effects of water vapor on the reaction of CH2OO with NH3. Phys. Chem. Chem. Phys. 2019; 21: 22589-22597. Chhantyal-Pun R, McGillen MR, Beames JM, Khan MAH, Percival CJ, Shallcross DE, et al. Temperature-dependence of the rates of reaction of trifluoroacetic acid with Criegee intermediates. Angew. Chem. Int. Ed. 2017; 56: 9044-9047. Chhantyal-Pun R, Rotavera B, McGillen MR, Khan MAH, Eskola AJ, Caravan RL, et al. Criegee intermediate reactions with carboxylic acids: a potential source of secondary organic aerosol in the atmosphere. ACS Earth Space Chem. 2018; 2: 18

Journal Pre-proof

833-842. Chhantyal-Pun R, Shannon RJ, Tew DP, Caravan RL, Duchi M, Wong C, et al. Experimental and computational studies of Criegee intermediate reactions with NH3 and CH3NH2. Phys. Chem. Chem. Phys. 2019; 21: 14042-14052. Corchado J, Chuang Y-Y, Fast P, Hu W-P, Liu YP, Lynch G, et al. POLYRATE,

of

version 9.7. University of Minnesota, Minneapolis 2007. de Jesus Medeiros D, Pimentel AS. New insights in the atmospheric HONO formation:

ro

new pathways for N2O4 isomerization and NO2 dimerization in the presence of

-p

water. J. Phys. Chem. A 2011; 115: 6357-65.

re

Enami S, Colussi AJ. Reactions of Criegee intermediates with alcohols at air-aqueous

lP

interfaces. J. Phys. Chem. A 2017; 121: 5175-5182.

14: 363-368.

na

Fukui K. The path of chemical reactions—the IRC approach. Acc. Chem. Res. 1981;

Jo ur

Goedecker S, Teter M, Hutter J. Separable dual-space Gaussian pseudopotentials. Phys. Rev. B 1996; 54: 1703. Grimme S, Antony J, Ehrlich S, Krieg H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010; 132: 154104. Gutbrod R, Kraka E, Schindler RN, Cremer D. Kinetic and theoretical investigation of the gas-phase ozonolysis of isoprene: carbonyl oxides as an important source for OH radicals in the atmosphere. J. Am. Chem. Soc. 1997; 119: 7330-7342. Gutbrod R, Schindler RN, Kraka E, Cremer D. Formation of OH radicals in the gas 19

Journal Pre-proof

phase ozonolysis of alkenes: the unexpected role of carbonyl oxides. Chem. Phys. Lett. 1996; 252: 221-229. Hammerich AD, Finlayson-Pitts BJ, Gerber RB. NOx reactions on aqueous surfaces with gaseous HCl: formation of a potential precursor to atmospheric Cl atoms. J. Phys. Chem. Lett. 2012; 3: 3405-10.

of

Hartwigsen C, Gœdecker S, Hutter J. Relativistic separable dual-space Gaussian pseudopotentials from H to Rn. Phys. Rev. B 1998; 58: 3641.

ro

Horie O, Moortgat GK. Gas-phase ozonolysis of alkenes. recent advances in

-p

mechanistic investigations. Acc. Chem. Res. 1998; 31: 387-396.

re

Hratchian HP, H Bernhard S. Accurate reaction paths using a Hessian based

lP

predictor-corrector integrator. J. Chem. Phys. 2004; 120: 9918-9924.

na

Inomata S, Sato K, Hirokawa J, Sakamoto Y, Tanimoto H, Okumura M, et al. Analysis of secondary organic aerosols from ozonolysis of isoprene by proton transfer

Jo ur

reaction mass spectrometry. Atmos. Environ. 2014; 97: 397-405. Jørgensen S, Gross A. Theoretical investigation of the reaction between carbonyl oxides and ammonia. J. Phys. Chem. A 2009; 113: 13041-13041. Johnson D, Marston G. The gas-phase ozonolysis of unsaturated volatile organic compounds in the troposphere. Chem. Soc. Rev. 2008; 37: 699-716. Karl M, Wright RF, Berglen TF, Denby B. Worst case scenario study to assess the environmental impact of amine emissions from a CO2 capture plant. Int. J. Greenhouse Gas Control 2011; 5: 439-447. Kumar M, Francisco JS. Elucidating the molecular mechanisms of Criegee-amine 20

Journal Pre-proof

chemistry in the gas phase and aqueous surface environments. Chem. Sci. 2019; 10: 743-751. Kumar M, Li H, Zhang X, Zeng XC, Francisco JS. Nitric acid-amine chemistry in the gas phase and at the air-water interface. J. Am. Chem. Soc. 2018; 140: 6456-6466.

of

Kumar M, Zhong J, Francisco JS, Zeng XC. Criegee intermediate-hydrogen sulfide chemistry at the air/water interface. Chem. Sci. 2017; 8: 5385-5391.

ro

Kurten T, Bonn B, Vehkamäki H, Kulmala M. Computational study of the reaction

re

Chem. A 2007; 111: 3394-3401.

-p

between biogenic stabilized Criegee intermediates and sulfuric acid. J. Phys.

lP

Lee C, Yang W, Parr RG. Development of the Colle-Salvetti correlation-energy

na

formula into a functional of the electron density. Phys. Rev. B 1988; 37: 785. Li J, Ying Q, Yi B, Yang P. Role of stabilized Criegee intermediates in the formation

Jo ur

of atmospheric sulfate in eastern United States. Atmos. Environ. 2013; 79: 442-447.

Li L, Kumar M, Zhu C, Zhong J, Francisco JS, Zeng XC. Near-barrierless ammonium bisulfate formation via a loop-structure promoted proton-transfer mechanism on the surface of water. J. Am. Chem. Soc. 2016; 138: 1816-1819. Lin YH, Yin C, Lin WH, Li YL, Takahashi K, Lin JJ. Criegee intermediate reaction with alcohol is enhanced by a single water molecule. J. Phys. Chem. Lett. 2018. Long B, Bao JL, Truhlar DG. Atmospheric chemistry of Criegee intermediates: 21

Journal Pre-proof

unimolecular reactions and reactions with water. J. Am. Chem. Soc. 2016; 138: 14409-14422. Long B, Bao JL, Truhlar DG. Unimolecular reaction of acetone oxide and its reaction with water in the atmosphere. Proc. Natl. Acad. Sci. U. S. A. 2018; 115: 6135-6140.

of

Mauldin RL, 3rd, Berndt T, Sipila M, Paasonen P, Petaja T, Kim S, et al. A new atmospherically relevant oxidant of sulphur dioxide. Nature 2012; 488:

ro

193-196.

-p

McGillen MR, Curchod BF, Chhantyal-Pun R, Beames JM, Watson N, Khan MAH, et

re

al. Criegee intermediate-alcohol reactions, a potential source of functionalized

lP

hydroperoxides in the atmosphere. ACS Earth Space Chem. 2017; 1: 664-672.

na

Miller Y, Finlayson-Pitts BJ, Gerber RB. Ionization of N2O4 in contact with water: mechanism, time scales and atmospheric implications. J. Am. Chem. Soc.

Jo ur

2009; 131: 12180-12185.

Neeb P, Horie O, Moortgat GK. The ethene-ozone reaction in the gas phase. J. Phys. Chem. A 1998; 102: 6778-6785. Njegic B, Raff JD, Finlayson-Pitts BJ, Gordon MS, Gerber RB. Catalytic role for water in the atmospheric production of ClNO. J. Phys. Chem. A 2010; 114: 4609-4618. Ortiz J, Cioslowski J, Fox D. Gaussian 09, revision B. 01. Wallingford CT 2009. Radom L, Lathan WA, Hehre WJ, Pople JA. Molecular orbital theory of the electronic structure of organic compounds. XVII. internal rotation in 1, 2-disubstituted 22

Journal Pre-proof

ethanes. J. Am. Chem. Soc. 1973; 95: 693-698. Sadezky A, Winterhalter R, Kanawati B, Römpp A, Spengler B, Mellouki A, et al. Oligomer formation during gas-phase ozonolysis of small alkenes and enol ethers: new evidence for the central role of the Criegee intermediate as oligomer chain unit. Atmos. Chem. Phys. 2008; 8: 2667-2699.

of

Sakamoto Y, Inomata S, Hirokawa J. Oligomerization reaction of the Criegee intermediate leads to secondary organic aerosol formation in ethylene

ro

ozonolysis. J. Phys. Chem. A 2013; 117: 12912-12921.

-p

Stone D, Blitz M, Daubney L, Howes NU, Seakins P. Kinetics of CH2OO reactions

re

with SO2, NO2, NO, H2O and CH3CHO as a function of pressure. Phys. Chem.

lP

Chem. Phys. 2014; 16: 1139-1149.

na

Tadayon SV, Foreman ES, Murray C. Kinetics of the reactions between the Criegee intermediate CH2OO and alcohols. J. Phys. Chem. A 2018; 122: 258-268.

Jo ur

Tobias HJ, Ziemann PJ. Kinetics of the gas-phase reactions of alcohols, aldehydes, carboxylic acids, and water with the C13 stabilized Criegee intermediate formed from ozonolysis of 1-tetradecene. J. Phys. Chem. A 2001; 105: 6129-6135. Torrent-Sucarrat M, Francisco JS, Anglada JM. Sulfuric acid as autocatalyst in the formation of sulfuric acid. J. Am. Chem. Soc. 2012; 134: 20632-44. VandeVondele J, Hutter J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. J. Chem. Phys. 2007; 127: 114105. VandeVondele J, Krack M, Mohamed F, Parrinello M, Chassaing T, Hutter J. 23

Journal Pre-proof

Quickstep: fast and accurate density functional calculations using a mixed Gaussian and plane waves approach. Comput. Phys. Commun. 2005; 167: 103-128. Veltman K, Singh B, Hertwich EG. Human and environmental impact assessment of postcombustion CO2 capture focusing on emissions from amine-based

of

scrubbing solvents to air. Environ. Sci. Technol. 2010; 44: 1496-1502. Vorobyov I, Yappert MC, DuPré DB. Hydrogen bonding in monomers and dimers of

ro

2-aminoethanol. J. Phys. Chem. A 2002; 106: 668-679.

-p

Wang K, Shan X, Chen X. Electron propagator theory study of 2-aminoethanol

re

conformers. J. Mol. Struct.: THEOCHEM 2009; 909: 91-95.

lP

Watson NAI, Black JA, Stonelake TM, Knowles PJ, Beames JM. An extended

na

computational study of Criegee intermediate-alcohol reactions. J. Phys. Chem. A 2019; 123: 218-229.

Jo ur

Welz O, Eskola AJ, Sheps L, Rotavera B, Savee JD, Scheer AM, 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: 4547-4550. Welz O, Savee JD, Osborn DL, Vasu SS, Percival CJ, Shallcross DE, et al. Direct kinetic measurements of Criegee intermediate (CH2OO) formed by reaction of CH2I with O2. Science 2012; 335: 204-207. Xie H-B, Li C, He N, Wang C, Zhang S, Chen J. Atmospheric chemical reactions of monoethanolamine initiated by OH radical: mechanistic and kinetic study. 24

Journal Pre-proof

Environ. Sci. Technol. 2014; 48: 1700-1706. Zhao Y, Truhlar DG. 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;

of

120: 215-241. Zhong J, Kumar M, Francisco JS, Zeng XC. Insight into chemistry on cloud/aerosol

ro

water surfaces. Acc. Chem. Res. 2018; 51: 1229-1237.

-p

Zhong J, Kumar M, Zhu CQ, Francisco JS, Zeng XC. Surprising stability of larger

re

Criegee intermediates on aqueous interfaces. Angew. Chem. Int. Ed. 2017; 56:

lP

7740-7744.

na

Zhong J, Li H, Kumar M, Liu J, Liu L, Zhang X, et al. Mechanistic insight into the reaction of organic acids with SO3 at the air-water interface. Angew. Chem. Int.

Jo ur

Ed. 2019.

Zhong J, Zhao Y, Li L, Li H, Francisco JS, Zeng XC. Interaction of the NH 2 radical with the surface of a water droplet. J. Am. Chem. Soc. 2015; 137: 12070-12078.

25

Journal Pre-proof

Jo ur

na

lP

re

-p

ro

of

Figure 1 The four configurations of monoethanolamine (MEA).

26

Journal Pre-proof

f o

l a

o r p

e

r P

n r u

o J

Figure 2 Reaction potential energy surface of C2 Criegee intermediates with the NH2 group of MEA. (in kcal/mol) RC: reaction complex; TS: transition state; P: product

27

Journal Pre-proof

f o

l a

o r p

e

r P

n r u

o J

Figure 3 Reaction potential energy surface of C2 Criegee intermediates with the OH group of MEA. (in kcal/mol) RC: reaction complex; TS: transition state; P: product

28

Journal Pre-proof

f o

l a

o r p

e

r P

n r u

o J

Figure 4 Reaction evolution of anti-CH3CHOO with the NH2 group of g′Gg′ configuration at the gas-liquid interface (a: direct reaction; b: water-mediated reaction).

29

Journal Pre-proof Table 1 Total rate constants (k, cm3 molecule-1 s-1) for the addition reaction of CH3CHOO with four MEA configurations at the different temperatures. 260

270

280

290

298

300

310

syn-CH3CHOO+MEA-NH2

1.23×10-16

7.13×10-16

1.39×10-16

6.48×10-16

6.32×10-16

6.19×10-16

1.38×10-16

anti-CH3CHOO+MEA-NH2

4.51×10-10

2.26×10-10

1.16×10-10

6.53×10-11

4.26×10-11

3.86×10-11

2.20×10-11

syn-CH3CHOO+MEA-OH

2.17×10-19

6.72×10-19

3.22×10-19

8.41×10-19

9.13×10-19

9.29×10-19

4.38×10-19

anti-CH3CHOO+MEA-OH

1.43×10-13

3.12×10-13

8.14×10-14

1.71×10-13

1.38×10-13

1.31×10-13

3.51×10-14

Jo ur

na

lP

re

-p

ro

of

T (K)

30

Journal Pre-proof Declaration of interests ☒ 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.

Jo ur

na

lP

re

-p

ro

of

☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests:

31

Journal Pre-proof

na

Highlights

lP

re

-p

ro

of

Graphical abstract

The reactions of sCIs with NH2 and OH groups of MEA were compared.



The NH2 and OH groups in the g′Gg′ and tGg′ configurations have a

Jo ur



synergistic effect on reactions with sCIs. 

C2 sCIs react more preferentially with NH2 groups of MEA configurations.



At 298 K, the total rate constant of anti-CH3CHOO with NH2 groups of MEA is 4.26×10-11 cm3 molecule-1 s-1.

32