Atmospheric Initial Nucleation Containing Carboxylic Acids - The

Apr 11, 2019 - ... enhances by 44 times with respect to the ones at the underground level. ... The Chamber Wall Index for Gas–Wall Interactions in A...
0 downloads 0 Views 2MB Size
Subscriber access provided by SAM HOUSTON STATE UNIV

A: Environmental, Combustion, and Atmospheric Chemistry; Aerosol Processes, Geochemistry, and Astrochemistry

Atmospheric Initial Nucleation Containing Carboxylic Acids Xia Sheng, Benjin Wang, Xue Song, Cleopatra Ashley Ngwenya, Yuyu Wang, and Hailiang Zhao J. Phys. Chem. A, Just Accepted Manuscript • DOI: 10.1021/acs.jpca.9b01104 • Publication Date (Web): 11 Apr 2019 Downloaded from http://pubs.acs.org on April 11, 2019

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 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 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.

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 37 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

Atmospheric Initial Nucleation Containing Carboxylic Acids Xia Sheng,1 Benjin Wang,1 Xue Song,1 Cleopatra Ashley Ngwenya,1 Yuyu Wang,2 and Hailiang Zhao1,* 1

College of Chemistry, Chemical and Environmental Engineering, Henan University of

Technology, Lianhua Street 100, 450001 Zhengzhou, China 2

College of Mathematical Science, Tianjin Normal University, Binshui West Road 393,

300387 Tianjin, China e-mail: [email protected]

1 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 The possible involvement of chemical components in atmospheric new particle formation has received increased attention in recent years. However, the deep understanding of the clusters formed between atmospheric gas-phase organic acids is incomplete. In this work, the chemical and physical properties of the cluster formed between three organic acids (glyoxylic acid (GA), oxalic acid (OA), and pyruvic acid (PA)) with common atmospheric nucleation precursors (methyl hydrogen sulfate (MHS), methanesulfonic acid (MSA), and hydroxymethanesulfonic acid (HMSA)) have been investigated with density functional theory (DFT) and ab initio coupled-cluster singles and doubles with perturbative triples (CCSD(T)) theory. Six- to nine-membered cyclic ring structures are mainly arranged via two classes of inter-molecular hydrogen bonds: SO−H∙∙∙O and CO−H∙∙∙O. The GA/OA/PA−MSA/MHS/HMSA complexes with the nineand eight-membered cyclic ring structures are thermodynamically more stable than the others. Large red shifts of the OH-stretching vibrational frequencies of both SO−H∙∙∙O (354-794 cm-1) and CO−H∙∙∙O (320-481 cm-1) are obtained with regard to the isolated gas monomers. Atoms in molecules (AIM) topological analysis reveals that the Laplacian of the charge density of the bimolecular interactions in the GA/OA/PA−MHS/MSA/HMSA complexes are higher than the upper value of the hydrogen bond criteria. The thermodynamic data, dipole moments, and atmospheric mixing ratios indicate that the MHS- and MSA-containing complexes possibly take part in atmospheric new particle formation. Additionally, the environmental factors, such as temperatures, pressures, are also important in atmospheric particle nucleation, and the gas-mixing ratios of the clusters at 12 km are much enhanced by 18-44 times with respect to the ones at the ground level. This study suggests that small cluster calculations may be helpful in simulating atmospheric new particle formation.

2 ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37 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

1. Introduction Aerosols and particulate matters in the atmosphere directly or indirectly impact solar radiation which in turn effects unprecedented changes in the global climate. 1,2 Aerosols directly affect radiative balance by absorbing and scattering solar radiation in the atmosphere, and indirectly by altering cloud properties.3 The indirect effects take place as a result of aerosol particles acting as cloud condensation nuclei (CCN). 4,5 To be considered of paramount importance in the formation of condensation nuclei are organic compounds that contribute to new particle formation (NPF), 6 which is the formation of thermodynamically consistent compounds from condensable vapors like ammonia, sulfuric acid and other gaseous precursors, accompanied by the growth of these clusters to notable size.3 The phrase “new particle formation” refers to the restructuring of particles into a new, different mode resulting in the growth of these, as measurable by continuous size distribution measurements.7 In China’s major and highly polluted cities like the capital Beijing, NPF has been investigated and observed.5,8 A previous study has critically inspected growth of clusters to possible CCN conversion basing on a case study of the regional aerosol measured in South Beijing.5 The authors further reported that the CCN concentration increased by a factor of 2-4 as a result of NPF. Up to 80% of CCN concentration was contributed by incessant growth from the nucleation mode.8 Meanwhile, the presence of atmospheric aerosols poses a potential threat to human health hence the attention given to them in recent years.9 Water soluble organic acids are a crucial component of aerosols. This is because of their ability in having strong molecular interactions with the nucleation precursors.10 Critical studies of water soluble organic acids show that they constitute a critical fraction in atmospheric particles. However, the NPF process is still the least understood process. It is because that the species involved in NPF is still highly unknown. H2SO4 has been placed under scrutiny over the years and accepted as one of the main precursors in NPF.11,12 According to Sihto et al., there is actually a direct correlation to the power 1-2 between the concentration of atmospheric H2SO4 and NPF, thus proving that H2SO4 plays a weighty role in NPF.13 In its gaseous form, the atmospheric concentrations range between 106-107 molecules cm-3, and it is an essential circumstance for NPF.14 H2SO4 in its gaseous form is necessary in the process of NPF, or at least a 3 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

compound containing H2SO4. Moreover, some recent studies at molecular levels show that the atmospheric processes of relevant pollutants can generate organic acids (such as, pyruvic acid, oxalic acid, etc.).15-23 The heterogeneous oxidation of catechol and substituted catechols can produce carboxylic acids.19-21 Moreover, the O3-initiated degradation of the substituted catechols, such as 3-methylcatechol and 4-methylcatechol, also can produce carboxylic acids, for example, 2-oxopropanoic acid, 2-oxoacetic acid. 22 The photooxidation of phenolic compounds are also a source of carboxylic acids. 23 On the other hand, organic acids are expected to be involved in NPF as well. 24,25 The studies by Zhang et al. are revealed that NPF involving H2SO4 is enhanced in the presence of organic acid, where 5-dicarboxylic acids can make a great contribution to nucleate by interacting with H2SO4 and NH3.24 Zhao et al. later claimed that ratios of cis-pinonic acid:H2SO4 were found to be 1:3 to 1:5 in the critical nucleus.25 Meanwhile, the photochemistry of organic acids, such as pyruvic acid and glyoxylic acid, can generate ketyl and acetyl radicals to reduce the uncertainty associated with SOA sources. 15-18 With the green revolution and the adoption of green technologies in the modern world, there has been a massive reduction of the H2SO4 levels in the atmosphere. From a base analysis of the gaseous phase H2SO4, the massive drop is due to the fact that most of the H2SO4 in the atmosphere is a result of industrialization and industrial emissions. In recent years, the development of strategies to curb global climate change and global warming has seen many renowned production companies reducing emissions into the atmosphere. Consequently, the volume concentrations of the gas phase H 2SO4 have dropped drastically. Methanesulfonic acid (MSA) is one of the significant sulfur containing organic acid found in the troposphere.14 It is a hygroscopic acid, a result of the oxidation reaction between dimethyl sulfide (DMS) and hydroxyl (OH) radical. Though it is not a major constituent of NPF, MSA is pivotal in particle growth, which is the backbone of atmospheric aerosols. The atmospheric volume concentration of MSA ranges from 10 5107 molecules cm-3.14 MSA on the other hand is one major constituent of marine air, with atmospheric concentrations in marine aerosols ranging from 0.039 to 0.322 ppt in Pacific and Indian oceans and Miami, Florida according to a study by Saltzman et al.26 Both hydroxymethanesulfonic acid (HMSA) and methyl hydrogen sulfate (MHS) are the 4 ACS Paragon Plus Environment

Page 4 of 37

Page 5 of 37 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

organic acids similar to MSA. HMSA is however very stable and not readily oxidized, thus it accumulates in the atmosphere in high concentrations. 27,28 It has been known to be produced by the reaction between SO2 and formaldehyde in acidic atmospheric droplets.29 It is about 6.29 ppt in PM10 of nine sites in Germany.28 Meanwhile, the concentrations are about 0.88-4.11 ppt in urban atmospheric particles by a 1H NMR study.29 The MHS is formed through the reaction of H2SO4 and methanol (MeOH) was recorded in their condensed phase as well as at the vapor and liquid interface. 30-33 In field study, the atmospheric dimethyl sulfate and MHS have been observed in Los Angeles, and the corresponding total gas phase concentrations of methyl sulfates were estimated at 0.716 to 3.987 ppt during the smog period.34 Consequently, many researches have focused on the nucleation of organic acids with organic species.35 In our previous study, we have noticed that the organic acids, such as pyruvic acid (PA), oxalic acid (OA), and glyoxylic acid (GA) are common in the atmosphere. These organic acids, especially OA, are representatives of familiar organic acids in the atmosphere, with significant concentrations and relatively low vapor pressures, thus they are expected to participate in nucleation.10 Organic acids are play an important role in NPF The recent studies show that MSA is bound to play an active role in the process of NPF to an extent of undoubtedly replacing H2SO4.14,26 Thus, to assess the impact of MHS, MSA, and HMSA on atmospheric NPF are absolutely necessary. In this research, we investigate and report the viability of MHS, MSA, HMSA as a basis of NPF instead of H2SO4 in atmospheric clusters formation with common organic acid nucleation precursors (PA, OA, and GA). The electronic structures and thermochemical parameters in this study could shed light on the mechanism for the research on atmospheric NPF involving MHS, MSA, and HMSA.

2. Computational details All electronic structure simulations were carried out by making use of the Becke three-parameter and Lee-Yang-Parr hybrid functional with the D3 version of Grimme’s dispersion (B3LYP-D3) and employing the Dunning’s augmented triple-ζ correlationconsistent basis set aug-cc-pVTZ. However, the aug-cc-pV(T+Q)Z basis set was used for sulfur. B3LYP-D3 has good performance on atmospheric clusters containing common 5 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

nucleation precursors, and its predictions for geometries, vibrational transition frequencies, interaction energies agreed best with experiments as compared to other density functionals.10,14,36-40 During the optimization, a small cutoff (Root Mean Square=1×10-6 a.u.) on force constants of the “verytight” convergence criterion (“optimization=verytight”) was utilized instead of the default cutoff (Root Mean Square=3×10-4 a.u.). Meanwhile, the “integral=ultrafine” option uses a more accurate numerical integration grid. For the hydrogen bonded atmospheric clusters, these two additional options (“optimization=verytight” and “integral=ultrafine”) have demonstrated good vibrational transition frequencies as well as thermochemical corrections to the molecular interaction energies.41,42 For each stationary point, IR frequency simulations were confirmed that no negative frequencies existed. The zero-point vibrational energy (ZPVE) was employed to correct binding energies (BEs). Furthermore, the basis set superposition error (BSSE) was corrected for BEs with the typical counterpoise (CP) method.43 In addition, BEs in the acetonitrile−HCl clusters obtained with MP2 and B3LYP-D3 were in good agreement with the experimental data as well as with a computational benchmarking study, while the ab initio coupled-cluster singles and doubles with perturbative triples (CCSD(T)) theory results show high-quality in predicting binding energies.44 Thus, the single point calculations of interaction energies were also carried out at the CCSD(T)/aug-cc-pVTZ (aug-cc-pV(T+Q)Z for sulfur) level on the B3LYPD3/aug-cc-pVTZ (aug-cc-pV(T+Q)Z for sulfur) optimized geometries. The interaction energy gained by CCSD(T) was often recommended as a theoretical benchmark. The thermal corrections achieved at the B3LYP-D3 level were employed to correct the CCSD(T) interaction energies. All the calculations were performed using the Gaussian 09 (revision E.01) program package.45 The topological analysis was carried out by utilizing atoms in molecules (AIM) theory as implemented in AIM2000 program package.46 The wavefunction files used for input to AIM analyses were generated by Gaussian 09 program package (revision E.01). The AIM analyses used wave functions obtained at B3LYP-D3 to examine the chemical and physical properties of hydrogen bonded complexes. Furthermore, the topological parameters (e.g., charge density ρ(r) and the Laplacian density ∇2ρ(r)) at bond critical 6 ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37 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

points (BCPs) were utilized to assess the SO−H∙∙∙O and CO−H∙∙∙O hydrogen bond strength in atmospheric complexes.47

3. Results and discussion 3.1. Molecular interaction energies The bimolecular complexes and the monomers were minimized at B3LYP-D3. Our previous studies have already focused on the monomers of MSA, MHS, GA, OA, and PA.10,14,36 For HMSA, it also has five different conformers and all the conformers and their relative energies are displayed in Figure S1 (Supporting Information). The conformers shown in Figure 1 are the most stable monomers. The same as in our previous studies, only the most stable monomers were considered in the complexation.10,14,36 During complexation, similar to our previous study cyclic structures were formed via two types of hydrogen bonds: (i) the OH group of carboxylic acids (GA, OA, PA) approaching the O atom of the S=O groups of MHS/MSA/HMSA to arrange a CO−H∙∙∙O hydrogen bond; (ii) the SOH group of MHS/MSA/HMSA interacting with the O atom of the C=O group of the carboxylic acids to generate a SO−H∙∙∙O hydrogen bond. For HMSA, the SOH group is not only the hydrogen bond donor, the alcohol group can also act as a hydrogen bond donor. Thus, there are several possible conformers when the alcohol group is the hydrogen bond donor, but these cyclic structures are high in energy, and they are exhibited in Figure S2 (Supporting Information). The corresponding interaction energies are listed in Table S1 (Supporting Information). This study only mainly concentrated on the lowest-energy structures of the bimolecular clusters. For each bimolecular complex, there are several low-lying stable structures and only the most stable structures (nine- to six-membered cyclic rings) were discussed. These structures are similar to the ones in our previous study on the carboxylic acid−H 2SO4 complexes.10 The corresponding structures of the GA/OA/PA−MHS, GA/OA/PA−MSA, and GA/OA/PA−HMSA complexes are shown in Figures 2-4. The seven-membered cyclic structures of the PA−MHS and PA−MSA complexes cannot be acquired at DFT/B3LYP-D3. They were converged into other stable structures at B3LYP-D3. Then, the corresponding optimized structures of the PA−MHS and PA−MSA complexes at

7 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 8 of 37

B3LYP were used in all analysis in order to compare with other seven-membered cyclic structures. The binding energies (BEs), zero-point vibrational energies (ZPVEs), basis set superposition errors (BSSEs), enthalpies of formation (Δ𝐻 formation (Δ𝐺

) and Gibbs energies of

) for the GA/OA/PA−MHS/MSA/HMSA complexes (298.15 K and 1

atm) at B3LYP-D3 are listed in Tables 1-3. ZPVEs vary from 2.8 to 5.2 kJ mol-1 for the studied systems, and BSSEs range from 1.0 to 1.9 kJ mol-1. Consequently, BEs of the GA/OA/PA−MHS/MSA/HMSA complexes were corrected with ZPVE and BSSE, and those of the nine- to eight-membered cyclic structures are larger than -47.1 kJ mol -1. While, the seven- to six-membered cyclic structures are less stable, and their BEs range between -37.1 and -24.0 kJ mol-1. For the six-membered cyclic structures, one OH group in carboxylic acid plays as the hydrogen bond donor as well as the acceptor at the same time. For the seven-membered cyclic structures, the SOH group in MHS/MSA/HMSA acts as both the donor and the acceptor. Nevertheless, in the nine- to eight-membered cyclic structures, one of the hydrogen bonds is produced with one OH in GA/OA/PA bonding to an O atom in MHS/MSA/HMSA, and another hydrogen bond is formed with bond formations between the OH in MHS/MSA/HMSA and an O atom in the carboxylic acid. In several previous studies, the eight-membered cyclic structures are the most stable conformers in some H2SO4-containing complexes: formic acid/acetic acid−H2SO4,48 cispinonic acid−H2SO4,25 and pinic acid−H2SO4.49 So an eight-membered cyclic structures is the characteristic feature of the atmospheric clusters between H 2SO4 and carboxylic acids. Thus, it is in line with our study where the more stable structures are eight- and nine-membered rings. The interactions of GA/OA/PA with H2SO4 have been studied at B3LYP-D3 in our previous study.10 The BEs of the most stable GA/OA/PA−H2SO4 complexes were obtained at -68.0 to -61.8 kJ mol-1, that is in line with this study, -67.1 to -57.3 kJ mol -1. Since they are acid-acid interactions, it is not favored as the atmospheric acid-base interactions, such as amine−H2SO4 (-134.5 to -95.5 kJ mol-1, RI-CC2/aug-ccpV(T+d)Z).50 In comparison base-base interactions are much weaker, such as DMA−DMA (-18.3 kJ mol-1, QCISD/aug-cc-pVTZ) and DMA−TMA (-22.0 kJ mol-1, CCSD(T)-F12a/VDZ-F12).51,52 8 ACS Paragon Plus Environment

for DMA−DMA and

Page 9 of 37 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

3.2. Structural analysis The most thermodynamically stable conformers of the GA, OA, and PA monomers are planar, and they have weak intra-molecular hydrogen bonds. 10 Figures 2-4 displays the most stable structures of the bimolecular complexes calculated at B3LYP-D3. The corresponding geometric details of the bimolecular clusters are listed in Table 4. The changes of the MHS/MSA/HMSA OH bond lengths upon complexation are in the range of 0.019 to 0.038 Å, while the changes of the GA/OA/PA OH bond lengths after the complex formation are much slightly smaller (< 0.026 Å) than the changes of the MHS/MSA/HMSA OH bond upon complexation. The changes of the OH bond lengths in the OA−MHS/MSA/HMSA complexes are relatively small as compared with the GA/PA−MHS/MSA/HMSA complexes. This is because that the OA intra-molecular hydrogen bond weakens the inter-molecular hydrogen bond upon complexation. Meanwhile, the angles of the inter-molecular hydrogen bonds in the most stable conformers deviate by less than 20º from the linear orientation. Two oxygen atoms inter-connecting through a π-conjugated double bonds (−O−H∙∙∙O=) form a homonuclear bond which is classified as resonance-assisted hydrogen bond.53-56 The contact distance d(O---O) is the length between the two O atoms in the −O−H∙∙∙O= hydrogen bond, and it has been used to determine the hydrogen bond strength.53 Basing to the contact distances, the strength of a hydrogen bond can be sorted as very strong if d(O---O) is less 2.50 Å, strong if d(O---O) is greater than 2.50 Å and less than 2.65 Å, medium if d(O---O) is greater than 2.65 Å and less than 2.80 Å, and weak if d(O---O) is greater than 2.80 Å. The calculated contact distances are listed in Table 4. For SO−H∙∙∙O in the GA/PA−MHS/MSA/HMSA complexes, the values of d(O---O) are in the range of 2.613-2.635 Å. Therefore, these hydrogen bonds are strong hydrogen bonds according to the contact distances. The remaining hydrogen bonds vary from 2.645 to 2.683 Å, so the strength of the hydrogen bonds is catalogued as medium. 3.3. OH-stretching vibrational frequencies A red shift (Δν) happens when a hydrogen bond is formed, and it is the wavenumber changes of OH-stretching vibrational transitions between a free molecule and its corresponding hydrogen bonded one (∆ν = ν

 ν

9 ACS Paragon Plus Environment

). The red shift is very

The Journal of Physical Chemistry 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 37

helpful to classify the relative strength of hydrogen bond interactions. The simulated OHstretching vibrational transition frequencies and the corresponding red shifts upon complex formation of the most stable cyclic conformers at B3LYP-D3 are listed in Table 5. Red shifts in OH-stretching vibrational transitions are very large for both CO−H∙∙∙O and SO−H∙∙∙O. In this study, the changes in the OH bond length upon complexation has recorded to have a linear correlation (R2 = 0.998) with the IR spectral shifts (Figure 5). The line implies that the larger Δr(OH), the higher red shifts of the OH-stretching vibrational frequencies. The values of Δ ν reveal that the strength of the SO−H∙∙∙O hydrogen bond is substantially stronger than that of CO−H∙∙∙O. The OH-stretching transitions of SO−H∙∙∙O can be sorted as follows: PA−MHS/MSA/HMSA > GA−MHS/MSA/HMSA > OA−MHS/MSA/HMSA. In contrast, the red shifts of the OHstretching vibrational frequencies of CO−H∙∙∙O have another order as follows: GA−MHS/MSA/HMSA

>

PA−MHS/MSA/HMSA

Furthermore, the relative shifts (%, ∆ν /ν

>

OA−MHS/MSA/HMSA.

) are 10.4-26.8% and 9.6-15.1% for

SO−H∙∙∙O and CO−H∙∙∙O, respectively. This means that the strength of CO−H∙∙∙O and SO−H∙∙∙O in the GA/OA/PA−MSA complexes may be classified as weak to intermediate ( weak < 12%, intermediate 12-22%), while the SO−H∙∙∙O hydrogen bonds in the GA/PA/OA−MHS/HMSA complexes belong to strong hydrogen bond (strong 25-80%) basing on the criteria proposed by Novak.57 In the similar studies, the OH-stretching vibrational frequencies of SO−H∙∙∙O were red shifted by 651 cm-1 (B3LYP/6-311++G(2d,2p)) for the phthalic acid−H2SO4 complex,58 and above 1000 cm-1 (B3LYP/6-31G(d,p)) for the cis-pinonic acid−H2SO4 and benzoic acid−H2SO4 complexes.25 Their corresponding red shifts of the OHstretching vibrational frequencies of CO−H∙∙∙O were also small, about 400 cm -1 for the organic acid−H2SO4 complexes, where organic acid is phthalic acid, cis-pinonic acid and benzoic acid.25,58 In argon matrices, the red shifts of the OH-stretching vibrational frequencies of CO−H∙∙∙O in the GA−H2O complex was about 330-447 cm-1.59 3.4. Topological analysis Besides the geometrical analysis, topological analysis using AIM of the molecular structure is illustrated by the electron density stationary points as well as the electron 10 ACS Paragon Plus Environment

Page 11 of 37 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

density gradient paths which originate and terminate at these points. It can be used to prove the existences of the hydrogen bond, and it supplies an alternative technique to assess the strength of a hydrogen bond. Therefore, AIM analysis was launched to deepen the nature of the bimolecular hydrogen bond in the atmospheric clusters of carboxylic acids and MHS/MSA/HMSA. The wavefunctions were calculated at the B3LYP-D3 level. The AIM plots of the GA/OA/PA−MHS/MSA/HMSA clusters with bond critical points (BCPs), ring critical points (RCPs) and electron density paths are displayed in Figure 6. The AIM topological parameters, such as electron density ρ(BCP), Laplacian density ∇2ρ(BCP) at the BCPs, and change in atomic charge Δq(H) at the H atom with the B3LYP-D3 method are listed in Table 6. In accordance with the criteria defined by Popelier, ρ(BCP) at the BCPs needs to be in the range of 0.002 to 0.040 a.u. for a hydrogen bond.60,61 Moreover, ρ(BCP) is a good indicator for the strength of hydrogen bond, where a larger the ρ(BCP) value would mean a stronger the hydrogen bond is.62 ρ(BCP) vary from 0.0274 to 0.0577 and from 0.0281 to 0.0494 a.u. for SO−H∙∙∙O and CO−H∙∙∙O, respectively. Meanwhile, ∇2ρ(BCP) is in the ranges of 0.1000-0.2662 and 0.1046-0.2396 a.u. for SO−H∙∙∙O and CO−H∙∙∙O, respectively. The values of ∇2ρ(BCP) reveal the chemical and physical properties of the molecular interaction. This means that a negative value of ∇2ρ(BCP) represents that there is a covalent interaction (e.g. shared interaction), and a positive value of ∇2ρ(BCP) implies closed-shell system interactions (such as van der Waals forces, and hydrogen bonding).62 These calculated values are much larger than the generally accepted data for a hydrogen bond, where ∇2ρ(BCP) is 0.014-0.139 a.u.60,61 This is due to strong acid-acid interactions. The trend is in line with previous study on the organic acid−H 2SO4 (organic acid = benzoic acid, cis-pinonic acid) complexes where ∇2ρ(BCP) overshot the upper value of the hydrogen bonding criteria.25 The length between a BCP and an RCP has also been employed as a criteria to determine the molecular stability of a hydrogen bond.63 In

these studied

PA/OA/GA−MHS/MSA/HMSA clusters, the length between the BCP (SO−H∙∙∙O) and the RCP are greater than their corresponding lengths between the BCP (CO−H∙∙∙O) hydrogen bond and the RCP. It proves that the relative strength of SO−H∙∙∙O are stronger than the corresponding CO−H∙∙∙O. 11 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 37

During hydrogen bond formation a charge transfer (CT) from the hydrogen bond acceptor to the donor is commonly recorded, resulting in decreasing in the atomic charge on the hydrogen atom.64 The changes of AIM atomic charge on the H atoms (Δq(H)) during hydrogen bond formation is listed in Table 6. Moreover, CT also plays an important part in the stabilization energy of an inter-molecular interacting system, and reveals electron delocalization interaction between an occupied molecular orbital of one molecule and an unoccupied molecular orbital of another molecule.65 The changes of Δq(H) for CO−H∙∙∙O are marginally shifted by 0.024-0.090 a.u., but the substantial increasements of 0.053-0.129 a.u. are noticed for SO−H∙∙∙O. The values of Δq(H) demonstrate that the strength of SO−H∙∙∙O is substantially larger than that of CO−H∙∙∙O. 3.5. Atmospheric implications Atmospheric aerosols often comprise a significant fraction of organic matter. 66 Organic acids in the atmosphere have been noticed to strengthen nucleation and growth of nanoparticles involving H2SO4.67 Meanwhile, thermodynamic analyses can offer great insight into the probability and possibility of atmospheric cluster formation in detail. Gibbs energy changes are used for evaluating the strength of the inter-molecular interaction and the spontaneity in the process of the cluster formation. Undoubtedly, thermodynamics favors the formation of the GA/OA/PA−MHS/MSA/HMSA at room temperature. In the view of the geometric parameters, topological analysis, and red shifts of the OH-stretching vibrational transitions discussed above, it has been found the GA/OA/PA−MHS/MSA/HMSA clusters are very stable. However, the interaction energies do not indicate that these complexes are important for the atmospheric nucleation. Determining the concentrations of the GA/OA/PA−MHS/MSA/HMSA clusters under a realistic atmospheric circumstance is therefore of great interest. The atmospheric concentrations can offer a possible reference for the corresponding clusters in the atmosphere. The chemical reaction, such as, between GA and MHS, is carried out as a case and it is delivered as follows: GA + MHS ⇄ GA−MHS

(1)

Then, the mass-balance relations result in the GA−MHS cluster concentrations under 12 ACS Paragon Plus Environment

Page 13 of 37 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

equilibrium circumstances. [GA−MHS] = [GA] × [MHS] 𝑒



(2)

The atmospheric concentration of GA−MHS rely on the Gibbs energies and the atmospheric concentrations of the participating gas molecules (GA and MHS). Where R is the molar gas constant, T is the atmospheric temperature in K, ΔG is the Gibbs energy for formation of the GA−MHS cluster. The net generation of GA in the atmosphere was measured at 2.95 (±0.89) × 10 10 mole year–1,68 with about 85% in the particle phase.69 GA is the most existed oxocarboxylic acids in the atmosphere with the concentration in the range of 1.0 × 10 71.0 × 109 molecules cm-3.70 PA is abundant species in the atmosphere, and it is a keto-acid intermediate produced by isoprene oxidation with a concentration of 100 ppt. 71 The total concentration of OA was chosen to be 5 ppb, and this value corresponds to typical and reasonable concentration of this pollutant species. 72 HMSA is about 0.158 ppb in PM10 of nine sites in Germany, which contribute about 0.21% to the total mass. 28 Apparently, the real condition in the atmosphere is much more complicated, and these simulations display a simplistic and limited approximation. Nevertheless, these simulations offer a general assessment of the essentiality of various clusters. With the purpose of assessing the environmental impact of MHS, MSA, HMSA to atmospheric NPF, the maximum limits of the atmospheric concentrations for each gas molecule (25 ppt for MSA,14 0.158 ppt for HMSA, 3.987 ppt for MHS, 100 ppt for PA, 5000 ppt for OA, and 40.8 ppt for GA) were used in the calculations. In quantum chemistry the CCSD(T) method has been known as the “gold-standard”, and it has been the first option for over twenty years to get accurate interaction energies and molecular properties. The Gibbs energies (298 K and 1 atm) of the most stable structures were calculated at the CCSD(T)//B3LYP-D3 level. The CCSD(T) interaction energies were corrected with the thermal correction delivered at B3LYP-D3, and presented in Table 7. Using Eq. (2) as well as the CCSD(T) Gibbs energies of formation, we have estimated the atmospheric concentrations of the most thermodynamically stable MHS/MSA/HMSAcontaining clusters. The calculated cluster concentrations of the MSA-containing clusters are 31-1202 molecules cm-3. The atmospheric concentrations of the MHS-containing clusters are one magnitude less than those of the MSA-containing complexes. In contrast, 13 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 HMSA-containing complexes are difficult to be found in the atmosphere. This is due to the initial HMSA gas molecule is much less than MSA and MHS. In our previous studies, we have demonstrated that the mixing ratios of MSA with the common atmospheric aerosol nucleation precursors (such as CH 3OH, HCOOH, CH2OCH3, etc.) are almost the same as H2SO4 in nucleation.14 Meanwhile, the interactions between MSA and amines (ethylamine, dimethylamine and trimethylamine) have been found to be important to NPF in coastal and agricultural areas. NPF from MSA is comparable with that from H2SO4, where the concentrations of the atmospheric gas phase MSA are about ~10-100% of that of H2SO4.73,74 In addition, MHS reveals a very larger atmospheric nucleation capability than H2SO4 and MSA, which is based on the simulated thermodynamic properties and the atmospheric concentrations of the participated gases.36 As compared to other relevant atmospheric clusters, the concentration of DMA−H2SO4 are 3×108-3.8×109 molecules cm-3 and for H2O−H2SO4, it is 4.5×102-6.8×107 molecules cm-3.14 As one can see, the atmospheric concentrations of the GA/OA/PA−MHS/MSA/HMSA clusters are negligible as compared to the two former mentioned common H2SO4-containing species. This means GA/OA/PA could play a role in NPF, but the contribution to NPF would be significantly lower than the common nucleation precursors, such as DMA and H 2O. However, the Gibbs energies of formation of the GA/OA/PA−MHS/MSA/HMSA clusters (-22.1 to -9.1 kJ mol -1) show that those clusters are more favored in thermal reactions than H2O−H2SO4 (-9.5 kJ mol-1, PW91PW91/6-311++G(3df,3pd)), although less favored than those of DMA−H 2SO4 (69.1 kJ mol-1, RI-CC2/aug-cc-pV(T+d)Z//RI-MP2/aug-cc-pV(D+d)Z).48,50 This means that GA/OA/PA may also make contributions to NPF in some highly GA/OA/PA polluted areas. In the troposphere (0-12 km), the temperature and the atmospheric pressure decrease as the altitude increases.76-78 The temperature in troposphere reduces by 6.49 K for every 1 km increase in altitude, and it is a constant temperate (216.69 K) from 11 km. 76,77 The atmospheric pressure reduces from 1 atm. (0 km) to 0.19 atm. (12 km). 78 To evaluate the specific Gibbs energies at various atmospheric heights, we carried out a series of simulations at different atmospheric altitude (Table S2, Supporting Information). As an example, the gas-phase mixing ratios of the most stable GA/OA/PA−MSA conformers is 14 ACS Paragon Plus Environment

Page 14 of 37

Page 15 of 37 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

constructed with the atmospheric altitude (Figure 7). It can be noticed that the concentrations of these complexes are dramatically increased by about 18-44 times at 12 km than the ones under the ground level. This is because the Gibbs energy reduces by about 9.4 kJ mol-1 as the height rises from 0 to 12 km. Although the atmospheric height effects on the thermo-dynamical parameters of complexes have been illustrated to be essential for explaining the atmospheric nucleation mechanism, 10 the concentrations of the participated gas molecules might decrease with altitude. However, there is a deficiency of concentrations as a function of atmospheric height in field studies, thus it is not easy to assess the corresponding atmospheric relevance as a function of atmospheric altitude. In this study, the method for calculating the concentration of clusters is only a limited to simplistic approximation, and the real atmospheric circumstance is actually much more complicated. This is only a qualitative assessment of the importance of the clusters in the atmosphere. The IIN (ion-induced nucleation) of single particle was considered by Nadykto et al., and they claimed that the dipole-charge attraction is very important in contributing the size of the ion clusters when polar gas molecules participate in the nucleation process. 79 As compared to the dipole moments for the cis-pinonic acid−H2SO4, H2O−H2SO4, H2SO−H2SO4, NH3−H2SO4 clusters are 3.8-5.2 Debyes,25 the dipole moments of the GA/OA/PA−MHS/MSA/HMSA clusters range from 2.67 to 4.28 Debyes (Table 7). The dipole moments of the monomers were listed in Table S3 (Supporting Information). The larger the dipole moment, the stronger the dipole-charge attraction and the bigger the size of the clusters, which is why the GA/OA/PA−MHS/MSA/HMSA clusters have high a dipole moment, and the dipole-ion interaction effect increases the importance of ions in mediating the new particle formation in the atmosphere.

4. Conclusions Understanding the properties of atmospheric particles is a very difficult issue. In the present study, quantum chemical simulations were performed by using the DFT and CCSD(T) methods to elucidate the possibility of the molecular clusters of atmospheric relevance that incorporate MHS/MSA/HMSA and three carboxylic acids (GA, OA, and PA)

in

atmospheric

particle

formation.

Structural

15 ACS Paragon Plus Environment

analyses

reveal

that

the

The Journal of Physical Chemistry 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 37

GA/OA/PA−MHS/HMSA clusters have two types of hydrogen bonds forming nine- to six-membered cyclic ring structures. The GA/OA/PA−MSA/MHS complexes were calculated to be existed in atmospherically appreciable quantities. Therefore, they likely participate the initial steps of atmospheric nucleation. This research simulates the possible forms of carboxylic acids with MHS/MSA/HMSA when taking part in initial nucleation, and more theoretical and experimental researches are still desired to depict the nucleation mechanism.

Conflicts of interest There are no conflicts to declare.

Supporting Information The

supporting

information

includes

the

optimized

conformers

of

the

hydroxymethanesulfonic acid (HMSA) monomer at the B3LYP-D3/aug-cc-pVTZ (augcc-pV(T+d)Z for sulfur) level (Figure S1); the less stable optimized conformers of the GA/OA/PA−HMSA complexes (Figure S2); calculated interaction energies of the conformers in Figure S1 (Table S1); calculated the Gibbs energies and the corresponding atmospheric concentrations at different temperatures, atmospheric pressures and heights in the Earth atmosphere (Table S2); calculated dipole moments of the MHS-, MSA- and HMSA-containing complexes (Table S3).

Acknowledgements This work is supported by the National Natural Science Foundation of China (Grant No.: 21607037), the Henan University of Technology (2017BS022), and the Fundamental Research Funds for the Henan Provincial Colleges and Universities in Henan University of Technology (Grant Nos.: 2017QNJH27, 2016QNJH05). We also thank the High Performance Computing Centre of Shandong University for providing Gaussian 09 software package and high-performance computation.

16 ACS Paragon Plus Environment

Page 17 of 37 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

References 1. Sarnela, N.; Jokinen, T.; Nieminen, T.; Lehtipalo, K.; Junninen, H.; Kangasluoma, J.; Hakala, J.; Taipale, R.; Schobesberger, S.; Sipilä, M.; Larnimaa, K.; Westerholm, H.; Heijari, J.; Kerminen, V.-M.; Petäjä, T.; Kulmala, M., Sulphuric Acid and Aerosol Particle Production in the Vicinity of an Oil Refinery. Atmos. Environ. 2015, 119, 156166. 2. Zhang, X.; Yin, Y.; Lin, Z.; Han, Y.; Hao, J.; Yuan, L.; Chen, K.; Chen, J.; Kong, S.; Shan, Y.; Xiao, H.; Tan, W., Observation of Aerosol Number Size Distribution and New Particle Formation at a Mountainous Site in Southeast China. Sci. Total Environ. 2017, 575, 309-320. 3. Kulmala, M.; Kerminen, V.-M.; Anttila, T.; Laaksonen, A.; O'Dowd, C. D., Organic Aerosol Formation via Sulphate Cluster Activation. J. Geophys. Res.-Atmos. 2004, 109 (D4), D04205. 4. Betha, R.; Spracklen, D. V.; Balasubramanian, R., Observations of New Aerosol Particle Formation in a Tropical Urban Atmosphere. Atmos. Environ. 2013, 71, 340-351. 5. Yue, D. L.; Hu, M.; Zhang, R. Y.; Wu, Z. J.; Su, H.; Wang, Z. B.; Peng, J. F.; He, L. Y.; Huang, X. F.; Gong, Y. G.; Wiedensohler, A., Potential Contribution of New Particle Formation to Cloud Condensation Nuclei in Beijing. Atmos. Environ. 2011, 45 (33), 6070-6077. 6. Cai, C.; Zhang, X.; Wang, K.; Zhang, Y.; Wang, L.; Zhang, Q.; Duan, F.; He, K.; Yu, S.-C., Incorporation of New Particle Formation and Early Growth Treatments into WRF/Chem: Model Improvement, Evaluation, and Impacts of Anthropogenic Aerosols over East Asia. Atmos. Environ. 2016, 124, 262-284. 7. Dameto de España, C.; Wonaschütz, A.; Steiner, G.; Rosati, B.; Demattio, A.; Schuh, H.; Hitzenberger, R., Long-term Quantitative Field Study of New Particle Formation (NPF) Events as a Source of Cloud Condensation Nuclei (CCN) in the Urban Background of Vienna. Atmos. Environ. 2017, 164, 289-298. 8. Wang, Z. B.; Hu, M.; Pei, X. Y.; Zhang, R. Y.; Paasonen, P.; Zheng, J.; Yue, D. L.; Wu, Z. J.; Boy, M.; Wiedensohler, A., Connection of Organics to Atmospheric New particle Formation and Growth at an Urban Site of Beijing. Atmos. Environ. 2015, 103, 717. 9. Huang, X.; Wang, C.; Peng, J.; He, L.; Cao, L.; Zhu, Q.; Cui, J.; Wu, Z.; Hu, M., Characterization of Particle Number Size Distribution and New Particle Formation in Southern China. J. Environ. Sci. (China) 2017, 51, 342-351. 10. Zhao, H.; Zhang, Q.; Du, L., Hydrogen Bonding in Cyclic Complexes of Carboxylic Acid-Sulfuric Acid and Their Atmospheric Implications. RSC Adv. 2016, 6 (75), 71733-71743. 11. Sipilä, M.; Berndt, T.; Petäjä, T.; Brus, D.; Vanhanen, J.; Stratmann, F.; Patokoski, J.; Mauldin, R. L.; Hyvärinen, A.-P.; Lihavainen, H.; Kulmala, M., The Role of Sulfuric Acid in Atmospheric Nucleation. Science 2010, 327 (5970), 1243-1246. 12. Kirkby, J.; Curtius, J.; Almeida, J.; Dunne, E.; Duplissy, J.; Ehrhart, S.; Franchin, A.; Gagne, S.; Ickes, L.; Kurten, A.; Kupc, A.; Metzger, A.; Riccobono, F.; Rondo, L.; Schobesberger, S.; Tsagkogeorgas, G.; Wimmer, D.; Amorim, A.; Bianchi, F.; Breitenlechner, M.; David, A.; Dommen, J.; Downard, A.; Ehn, M.; Flagan, R. C.; Haider, S.; Hansel, A.; Hauser, D.; Jud, W.; Junninen, H.; Kreissl, F.; Kvashin, A.; Laaksonen, A.; Lehtipalo, K.; Lima, J.; Lovejoy, E. R.; Makhmutov, V.; Mathot, S.; Mikkila, J.; 17 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Minginette, P.; Mogo, S.; Nieminen, T.; Onnela, A.; Pereira, P.; Petaja, T.; Schnitzhofer, R.; Seinfeld, J. H.; Sipila, M.; Stozhkov, Y.; Stratmann, F.; Tome, A.; Vanhanen, J.; Viisanen, Y.; Vrtala, A.; Wagner, P. E.; Walther, H.; Weingartner, E.; Wex, H.; Winkler, P. M.; Carslaw, K. S.; Worsnop, D. R.; Baltensperger, U.; Kulmala, M., Role of Sulphuric Acid, Ammonia and Galactic Cosmic Rays in Atmospheric Aerosol Nucleation. Nature 2011, 476 (7361), 429-433. 13. Vuollekoski, H.; Nieminen, T.; Paasonen, P.; Sihto, S.-L.; Boy, M.; Manninen, H.; Lehtinen, K.; Kerminen, V.-M.; Kulmala, M., Atmospheric Nucleation and Initial Steps of Particle Growth: Numerical Comparison of Different Theories and Hypotheses. Atmos. Res. 2010, 98 (2), 229-236. 14. Zhao, H.; Jiang, X.; Du, L., Contribution of Methane Sulfonic Acid to New Particle Formation in the Atmosphere. Chemosphere 2017, 174, 689-699. 15. Eugene, A. J.; Guzman, M. I., Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid. J. Phys. Chem. A 2017, 121 (15), 2924-2935. 16. Xia, S.-S.; Eugene, A. J.; Guzman, M. I., Cross Photoreaction of Glyoxylic and Pyruvic Acids in Model Aqueous Aerosol. J. Phys. Chem. A 2018, 122 (31), 6457-6466. 17. Eugene, A. J.; Xia, S.-S.; Guzman, M. I., Aqueous Photochemistry of Glyoxylic Acid. J. Phys. Chem. A 2016, 120 (21), 3817-3826. 18. Eugene, A. J.; Guzman, M. I., Reply to “Comment on ‘Reactivity of Ketyl and Acetyl Radicals from Direct Solar Actinic Photolysis of Aqueous Pyruvic Acid’”. J. Phys. Chem. A 2017, 121 (45), 8741-8744. 19. Pillar, E. A.; Guzman, M. I., Oxidation of Substituted Catechols at the Air–Water Interface: Production of Carboxylic Acids, Quinones, and Polyphenols. Environ. Sci. Technol. 2017, 51 (9), 4951-4959. 20. Pillar, E. A.; Zhou, R.; Guzman, M. I., Heterogeneous Oxidation of Catechol. J. Phys. Chem. A 2015, 119 (41), 10349-10359. 21. Pillar-Little, E. A.; Guzman, M. I., An Overview of Dynamic Heterogeneous Oxidations in the Troposphere. Environments 2018, 5 (9), 104. 22. Sun, J.; Wei, B.; Mei, Q.; An, Z.; Wang, X.; He, M., Ozonation of 3Methylcatechol and 4-Methylcatechol in the Atmosphere and Aqueous Particles: Mechanism, Kinetics and Ecotoxicity Assessment. Chem. Eng. J. 2019, 358, 456-466. 23. Huang, D. D.; Zhang, Q.; Cheung, H. H. Y.; Yu, L.; Zhou, S.; Anastasio, C.; Smith, J. D.; Chan, C. K., Formation and Evolution of aqSOA from Aqueous-Phase Reactions of Phenolic Carbonyls: Comparison between Ammonium Sulfate and Ammonium Nitrate Solutions. Environ. Sci. Technol. 2018, 52 (16), 9215-9224. 24. Zhang, R.; Suh, I.; Zhao, J.; Zhang, D.; Fortner, E. C.; Tie, X.; Molina, L. T.; Molina, M. J., Atmospheric New Particle Formation Enhanced by Organic Acids. Science 2004, 304 (5676), 1487-1490. 25. Zhao, J.; Khalizov, A.; Zhang, R.; McGraw, R., Hydrogen-Bonding Interaction in Molecular Complexes and Clusters of Aerosol Nucleation Precursors. J. Phys. Chem. A 2009, 113 (4), 680-689. 26. Saltzman, E. S.; Savoie, D. L.; Zika, R. G.; Prospero, J. M., Methane Sulfonic Acid in the Marine Atmosphere. J. Geophys. Res.-Oceans 1983, 88 (C15), 10897-10902.

18 ACS Paragon Plus Environment

Page 18 of 37

Page 19 of 37 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

27. Yasuyuki, K.; Natsuji, S.; Yukiko, A.; Hiroshi, O.; Manabu, I., Enhanced Dissolution of SO2 into Dewwater by Forming Hydroxyalkanesulfonate. Chem. Lett. 1996, 25 (3), 197-198. 28. Scheinhardt, S.; van Pinxteren, D.; Müller, K.; Spindler, G.; Herrmann, H., Hydroxymethanesulfonic Acid in Size-segregated Aerosol Particles at Nine Sites in Germany. Atmos. Chem. Phys. 2014, 14 (9), 4531-4538. 29. Suzuki, Y.; Kawakami, M.; Akasaka, K., 1H NMR Application for Characterizing Water-soluble Organic Compounds in Urban Atmospheric Particles. Environ. Sci. Technol. 2001, 35 (13), 2656-64. 30. Van Loon, L. L.; Allen, H. C., Uptake and Surface Reaction of Methanol by Sulfuric Acid Solutions Investigated by Vibrational Sum Frequency Generation and Raman Spectroscopies. J. Phys. Chem. A 2008, 112 (34), 7873-7880. 31. Van Loon, L. L.; Allen, H. C.; Wyslouzil, B. E., Effective Diffusion Coefficients for Methanol in Sulfuric Acid Solutions Measured by Raman Spectroscopy. J. Phys. Chem. A 2008, 112 (43), 10758-10763. 32. Van Loon, L. L.; Allen, H. C., Methanol Reaction with Sulfuric Acid:  A Vibrational Spectroscopic Study. J. Phys. Chem. B 2004, 108 (45), 17666-17674. 33. Schobesberger, S.; Junninen, H.; Bianchi, F.; Lönn, G.; Ehn, M.; Lehtipalo, K.; Dommen, J.; Ehrhart, S.; Ortega, I. K.; Franchin, A.; Nieminen, T.; Riccobono, F.; Hutterli, M.; Duplissy, J.; Almeida, J.; Amorim, A.; Breitenlechner, M.; Downard, A. J.; Dunne, E. M.; Flagan, R. C.; Kajos, M.; Keskinen, H.; Kirkby, J.; Kupc, A.; Kürten, A.; Kurtén, T.; Laaksonen, A.; Mathot, S.; Onnela, A.; Praplan, A. P.; Rondo, L.; Santos, F. D.; Schallhart, S.; Schnitzhofer, R.; Sipilä, M.; Tomé, A.; Tsagkogeorgas, G.; Vehkamäki, H.; Wimmer, D.; Baltensperger, U.; Carslaw, K. S.; Curtius, J.; Hansel, A.; Petäjä, T.; Kulmala, M.; Donahue, N. M.; Worsnop, D. R., Molecular Understanding of Atmospheric Particle Formation from Sulfuric Acid and Large Oxidized Organic Molecules. P. Natl. Acad. Sci. U.S.A. 2013, 110 (43), 17223-17228. 34. Eatough, D. J.; White, V. F.; Hansen, L. D.; Eatough, N. L.; Cheney, J. L., Identification of Gas-Phase Dimethyl Sulfate and Monoethyl Sulfate in the Los Angeles Atmosphere. Environ. Sci. Technol. 1986, 20 (9), 867-872. 35. Hong, Y.; Liu, Y.-R.; Wen, H.; Miao, S.-K.; Huang, T.; Peng, X.-Q.; Jiang, S.; Feng, Y.-J.; Huang, W., Interaction of Oxalic Acid with Methylamine and its Atmospheric Implications. RSC Adv. 2018, 8 (13), 7225-7234. 36. Sheng, X.; Zhao, H.; Du, L., Molecular Understanding of the Interaction of Methyl Hydrogen Sulfate with Ammonia/Dimethylamine/Water. Chemosphere 2017, 186, 331-340. 37. Cheng, S.; Tang, S.; Tsona, N. T.; Du, L., The Influence of the Position of the Double Bond and Ring Size on the Stability of Hydrogen Bonded Complexes. Sci. Rep. 2017, 7 (1), 11310. 38. Zhang, Q.; Du, L., Hydrogen Bonding in the Carboxylic Acid–Aldehyde Complexes. Comput. Theor. Chem. 2016, 1078, 123-128. 39. Li, S.; Kjaergaard, H. G.; Du, L., Infrared Spectroscopic Probing of Dimethylamine Clusters in an Ar Matrix. J. Environ. Sci.-China 2016, 40, 51-59. 40. Du, L.; Tang, S.; Hansen, A. S.; Frandsen, B. N.; Maroun, Z.; Kjaergaard, H. G., Subtle Differences in the Hydrogen Bonding of Alcohol to Divalent Oxygen and Sulfur. Chem. Phys. Lett. 2017, 667, 146-153. 19 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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. Hansen, A. S.; Du, L.; Kjaergaard, H. G., The Effect of Fluorine Substitution in Alcohol-Amine Complexes. Phys. Chem. Chem. Phys. 2014, 16 (41), 22882-22891. 42. Hansen, A. S.; Du, L.; Kjaergaard, H. G., Positively Charged Phosphorus as a Hydrogen Bond Acceptor. J. Phys. Chem. Lett. 2014, 5 (23), 4225-4231. 43. Boys, S. F.; Bernardi, F., The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19 (4), 553-566. 44. Bork, N.; Du, L.; Reiman, H.; Kurtén, T.; Kjaergaard, H. G., Benchmarking Ab Initio Binding Energies of Hydrogen-Bonded Molecular Clusters Based on FTIR Spectroscopy. J. Phys. Chem. A 2014, 118 (28), 5316-5322. 45. 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.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J., Gaussian 09, Revision E.01. Gaussian, Inc.: Wallingford, CT, USA, 2013. 46. Lane, J. R.; Contreras-Garcia, J.; Piquemal, J.-P.; Miller, B. J.; Kjaergaard, H. G., Are Bond Critical Points Really Critical for Hydrogen Bonding? J. Chem. Theory Comput. 2013, 9 (8), 3263-3266. 47. Bader, R. F., A Quantum Theory of Molecular Structure and Its Applications. Chem. Rev. 1991, 91 (5), 893-928. 48. Nadykto, A. B.; Yu, F., Strong Hydrogen Bonding between Atmospheric Nucleation Precursors and Common Organics. Chem. Phys. Lett. 2007, 435 (1-3), 14-18. 49. Elm, J.; Kurten, T.; Bilde, M.; Mikkelsen, K. V., Molecular Interaction of Pinic Acid with Sulfuric Acid: Exploring the Thermodynamic Landscape of Cluster Growth. J. Phys. Chem. A 2014, 118 (36), 7892-7900. 50. Kurtén, T.; Loukonen, V.; Vehkamaki, H.; Kulmala, M., Amines are Likely to Enhance Neutral and Ion-Induced Sulfuric Acid-Water Nucleation in the Atmosphere More Effectively than Ammonia. Atmos. Chem. Phys. 2008, 8 (14), 4095-4103. 51. Du, L.; Kjaergaard, H. G., Fourier Transform Infrared Spectroscopy and Theoretical Study of Dimethylamine Dimer in the Gas Phase. J. Phys. Chem. A 2011, 115 (44), 12097-12104. 52. Du, L.; Lane, J. R.; Kjaergaard, H. G., Identification of the DimethylamineTrimethylamine Complex in the Gas Phase. J. Chem. Phys. 2012, 136 (18), 184305. 53. Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G., Evidence for Resonance-Assisted Hydrogen Bonding. 4. Covalent Nature of the Strong Homonuclear Hydrogen Bond. Study of the O-H···O System by Crystal Structure Correlation Methods. J. Am. Chem. Soc. 1994, 116 (3), 909-915.

20 ACS Paragon Plus Environment

Page 20 of 37

Page 21 of 37 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

54. Alabugin, I. V.; Manoharan, M.; Peabody, S.; Weinhold, F., Electronic Basis of Improper Hydrogen Bonding:  A Subtle Balance of Hyperconjugation and Rehybridization. J. Am. Chem. Soc. 2003, 125 (19), 5973-5987. 55. Gilli, P.; Bertolasi, V.; Pretto, L.; Gilli, G., Outline of a Transition-State Hydrogen-Bond Theory. J. Mol. Struct. 2006, 790 (1–3), 40-49. 56. Pakiari, A. H.; Eskandari, K., The Chemical Nature of very Strong Hydrogen Bonds in Some Categories of Compounds. J. Mol. Struct.: THEOCHEM 2006, 759 (1–3), 51-60. 57. Novak, A., Hydrogen Bonding in Solids Correlation of Spectroscopic and Crystallographic Data. Struct. Bond. 1974, 18, 177-216. 58. Xu, W.; Zhang, R., Theoretical Investigation of Interaction of Dicarboxylic Acids with Common Aerosol Nucleation Precursors. J. Phys. Chem. A 2012, 116 (18), 45394550. 59. Lundell, J.; Olbert-Majkut, A., Isolated Glyoxylic Acid-Water 1:1 Complexes in Low Temperature Argon Matrices. Spectrochim. Acta A 2015, 136, 113-121. 60. Koch, U.; Popelier, P., Characterization of CHO Hydrogen Bonds on the Basis of the Charge Density. J. Phys. Chem. 1995, 99 (24), 9747-9754. 61. Grabowski, S. J., Hydrogen Bonding Strength-Measures Based on Geometric and Topological Parameters. J. Phys. Org. Chem. 2004, 17 (1), 18-31. 62. Bader, R. F., Atoms in Molecules. A Quantum Theory. Oxford University Press: New York, 1990. 63. Popelier, P., Characterization of a Dihydrogen Bond on the Basis of the Electron Density. J. Phys. Chem. A 1998, 102 (10), 1873-1878. 64. Bushmarinov, I. S.; Lyssenko, K. A.; Antipin, M. Y., Atomic Energy in the 'Atoms in Molecules' Theory and its Use for Solving Chemical Problems. Russ. Chem. Rev. 2009, 78 (4), 283-302. 65. Umeyama, H.; Morokuma, K., Origin of Alkyl Substituent Effect in the Proton Affinity of Amines, Alcohols, and Ethers. J. Am. Chem. Soc. 1976, 98 (15), 4400-4404. 66. Shi, X.; Zhang, R.; Sun, Y.; Xu, F.; Zhang, Q.; Wang, W., A Density Functional Theory Study of Aldehydes and Their Atmospheric Products Participating in Nucleation. Phys. Chem. Chem. Phys. 2018, 20 (2), 1005-1011. 67. Khare, P.; Kumar, N.; Kumari, K. M.; Srivastava, S. S., Atmospheric Formic and Acetic Acids: An Overview. Rev. Geophys. 1999, 37 (2), 227-248. 68. Lin, G.; Sillman, S.; Penner, J. E.; Ito, A., Global Modeling of SOA: the Use of Different Mechanisms for Aqueous-Phase Formation. Atmos. Chem. Phys. 2014, 14 (11), 5451-5475. 69. Eugene, A. J.; Xia, S.-S.; Guzman, M. I., Aqueous Photochemistry of Glyoxylic Acid. J. Phys. Chem. A 2016, 120 (21), 3817-3826. 70. Liu, L.; Zhang, X.; Li, Z.; Zhang, Y.; Ge, M., Gas-phase Hydration of Glyoxylic Acid: Kinetics and Atmospheric Implications. Chemosphere 2017, 186, 430-437. 71. Reed Harris, A. E.; Cazaunau, M.; Gratien, A.; Pangui, E.; Doussin, J.-F.; Vaida, V., Atmospheric Simulation Chamber Studies of the Gas-Phase Photolysis of Pyruvic Acid. J. Phys. Chem. A 2017, 121 (44), 8348-8358. 72. Peng, X.-Q.; Liu, Y.-R.; Huang, T.; Jiang, S.; Huang, W., Interaction of Gas Phase Oxalic Acid with Ammonia and its Atmospheric Implications. Phys. Chem. Chem. Phys. 2015, 17 (14), 9552-9563. 21 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

73. Chen, H.; Finlayson-Pitts, B. J., New Particle Formation from Methanesulfonic Acid and Amines/Ammonia as a Function of Temperature. Environ. Sci. Technol. 2017, 51 (1), 243-252. 74. Dawson, M. L.; Varner, M. E.; Perraud, V.; Ezell, M. J.; Wilson, J.; Zelenyuk, A.; Gerber, R. B.; Finlayson-Pitts, B. J., Amine–Amine Exchange in Aminium– Methanesulfonate Aerosols. J. Phys. Chem. C 2014, 118 (50), 29431-29440. 75. Zhao, H.; Du, L., Atmospheric Implication of the Hydrogen Bonding Interaction in Hydrated Clusters of HONO and Dimethylamine in the Nighttime. Environ. Sci. Proc. Impacts 2017, 19 (1), 65-77. 76. Gonzalez, J.; Anglada, J. M.; Buszek, R. J.; Francisco, J. S., Impact of Water on the OH plus HOCl Reaction. J. Am. Chem. Soc. 2011, 133 (10), 3345-3353. 77. Ji, Y. M.; Wang, H. H.; Gao, Y. P.; Li, G. Y.; An, T. C., A Theoretical Model on the Formation Mechanism and Kinetics of Highly Toxic Air Pollutants from Halogenated Formaldehydes Reacted with Halogen Atoms. Atmos. Chem. Phys. 2013, 13 (22), 1127711286. 78. Seinfeld, J. H.; Pandis, S. N., Atmospheric Chemistry and Physics: from Air Pollution to Climate Change. second ed.; John Wiley & Sons, Inc.: New Jersey, 2006. 79. Nadykto, A. B.; Mäkelä, J. M.; Yu, F.; Kulmala, M.; Laaksonen, A., Comparison of the Experimental Mobility Equivalent Diameter for Small Cluster Ions with Theoretical Particle Diameter Corrected by Effect of Vapour Polarity. Chem. Phys. Lett. 2003, 382 (1–2), 6-11.

22 ACS Paragon Plus Environment

Page 22 of 37

Page 23 of 37 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

Table 1 B3LYP-D3 calculated binding energy (BE), zero-point vibrational energy (ZPVE), basis set superposition error (BSSE), enthalpy of formation (Δ𝐻 energy of formation (Δ𝐺 Type Nine-membered ring Eight-membered ring Seven-membered ring Six-membered ring a

) and Gibbs

) for the MHS-containing complexes (298.15 K and 1 atm) a

Conformer

BE b

ZPVE

BSSE

Δ𝐻

Δ𝐺

GA−MHS

-47.1

4.1

1.6

-47.3

-6.6

OA−MHS

-58.2

4.1

1.8

-59.1

-16.1

PA−MHS

-47.6

3.8

1.7

-48.0

-7.6

GA−MHS

-66.7

4.4

1.6

-67.8

-23.4

OA−MHS

-58.1

4.2

1.6

-59.1

-14.2

PA−MHS

-61.0

4.2

1.8

-62.3

-18.7

GA−MHS

-25.2

3.8

1.3

-24.8

13.6

OA−MHS

-30.5

3.8

1.6

-30.8

10.0

PA−MHS c

-37.9

3.8

1.8

-38.0

4.8

GA−MHS

-35.1

3.2

1.1

-34.2

2.6

OA−MHS

-32.0

2.8

1.1

-31.0

5.9

PA−MHS

-35.3

3.1

1.0

-34.2

2.4

All interaction energies are in kJ mol-1. b BEs corrected with ZPVE and BSSE. c Acquired

from single-point B3LYP-D3 calculation on the B3LYP optimized structure.

23 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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 24 of 37

Table 2 B3LYP-D3 calculated binding energy (BE), zero-point vibrational energy (ZPVE), basis set superposition error (BSSE), enthalpy of formation (Δ𝐻 energy of formation (Δ𝐺 Type Nine-membered ring Eight-membered ring Seven-membered ring Six-membered ring a

Conformer

) and Gibbs

) for the MSA-containing complexes (298.15 K and 1 atm) a BE b

ZPVE

BSSE

Δ𝐻

Δ𝐺

GA−MSA

-46.8

4.8

1.5

-47.8

-4.9

OA−MSA

-57.6

4.6

1.8

-59.0

-15.7

PA−MSA

-47.8

4.6

1.7

-49.1

-5.1

GA−MSA

-65.4

4.8

1.5

-66.8

-22.8

OA−MSA

-57.4

4.5

1.5

-58.6

-14.3

PA−MSA

-59.9

4.7

1.5

-61.3

-18.2

GA−MSA

-24.0

3.6

1.1

-23.4

14.0

OA−MSA

-29.3

4.0

1.5

-29.7

10.0

PA−MSA c

-36.2

4.4

1.6

-36.7

9.3

GA−MSA

-34.6

3.5

1.0

-34.0

3.7

OA−MSA

-31.8

3.0

1.1

-31.1

6.3

PA−MSA

-34.2

3.4

1.0

-33.5

3.9

All interaction energies are in kJ mol-1. b BEs corrected with ZPVE and BSSE. c Acquired

from single-point B3LYP-D3 calculation on the B3LYP optimized structure.

24 ACS Paragon Plus Environment

Page 25 of 37 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

Table 3 B3LYP-D3 calculated binding energy (BE), zero-point vibrational energy ) and Gibbs

(ZPVE), basis set superposition error (BSSE), enthalpy of formation (Δ𝐻 energy of formation (Δ𝐺

) for the HMSA-containing complexes (298.15 K and 1 atm)

a

Type Nine-membered ring Eight-membered ring Seven-membered ring Six-membered ring a

Conformer

BE b

ZPVE

BSSE

Δ𝐻

Δ𝐺

GA−HMSA

-48.8

4.9

1.7

-50.1

-2.9

OA−HMSA

-56.1

4.6

1.9

-57.8

-9.8

PA−HMSA

-50.1

4.7

1.9

-51.5

-4.5

GA−HMSA

-62.0

4.4

1.6

-63.3

-17.9

OA−HMSA

-55.4

4.0

1.7

-56.7

-11.6

PA−HMSA

-58.4

4.2

1.7

-59.8

-16.1

GA−HMSA

-29.9

4.7

1.4

-30.4

15.4

OA−HMSA

-37.1

5.1

1.7

-38.5

10.0

PA−HMSA c

-28.6

3.0

1.2

-30.1

14.4

GA−HMSA

-35.0

5.1

1.5

-35.6

10.8

OA−HMSA d

--

--

--

--

--

PA−HMSA

-36.7

5.2

1.7

-37.6

8.9

All interaction energies are in kJ mol-1. b BEs corrected with ZPVE and BSSE. c Acquired

from single-point B3LYP-D3 calculation on the B3LYP optimized structure. possible to obtain either by B3LYP-D3 or B3LYP.

25 ACS Paragon Plus Environment

d

Not

The Journal of Physical Chemistry 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 26 of 37

Table 4 The most stable geometric properties in the SO−H···O and CO−H···O hydrogen bonded complexes at the B3LYP-D3 level (angles in degrees; lengths/distances in Å) Conformer

SO−H···O (H2SO4 donor) Δr(OH) a

r(HB) b

θ(HB) c

CO−H···O (GA/OA/PA donor) dd

Δr(OH) a

r(HB) b

θ(HB) c

dd

GA−MHS

0.038

1.621

178.0

2.627

0.023

1.671

176.4

2.666

OA−MHS

0.019

1.697

171.2

2.677

0.015

1.718

160.8

2.672

PA−MHS

0.041

1.605

178.9

2.613

0.020

1.685

176.9

2.678

GA−MSA

0.036

1.632

178.2

2.635

0.026

1.648

175.4

2.644

OA−MSA

0.021

1.689

175.7

2.676

0.018

1.685

162.4

2.647

PA−MSA

0.038

1.618

178.5

2.623

0.022

1.660

176.2

2.655

GA−HMSA

0.038

1.620

174.7

2.623

0.026

1.652

176.6

2.649

OA−HMSA

0.020

1.697

175.7

2.683

0.019

1.699

157.7

2.645

PA−HMSA

0.040

1.605

179.3

2.614

0.021

1.666

175.1

2.660

a

Δr(OH) = rdimer – rmonomer, is the change in the OH bond length upon complexation. b Inter-

molecular hydrogen bond distance.

c

Inter-molecular hydrogen bond angle.

distance between the two oxygen atoms in the hydrogen bond (−O−H∙∙∙O=).

26 ACS Paragon Plus Environment

d

Contact

Page 27 of 37 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

Table 5 B3LYP-D3 calculated OH-stretching vibrational wavenumbers and red shifts (in cm-1) of the GA/OA/PA−MHS/MSA/HMSA complexes Conformer

a

SO−H···O (MHS/MSA/HMSA) Δν a

ν

CO−H···O (GA/OA/PA)

fD/fM b

ν

Δν a

fD/fM b

GA−MHS

3014

743

9

3222

441

24

OA−MHS

3403

354

16

3327

320

3

PA−MHS

2970

787

13

3250

380

19

GA−MSA

3045

721

5

3190

473

29

OA−MSA

3369

397

16

3271

376

4

PA−MSA

3010

756

9

3214

416

23

GA−HMSA

3004

755

7

3182

481

25

OA−HMSA

3377

382

14

3254

393

3

PA−HMSA

2965

794

12

3213

417

20

∆ν = ν



ν

.

b

fD/fM represents the enhancement of intensity upon

complexation.

27 ACS Paragon Plus Environment

The Journal of Physical Chemistry 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

Page 28 of 37

Table 6 AIM topological properties for the GA/OA/PA−MHS/MSA/HMSA complexes obtained at B3LYP-D3 (distances in Å; all other values in a.u.) Conformer

a

SO−H···O (H2SO4 donor) Δq(H) a

ρ(BCP) b

GA−MHS

0.121

0.0329

0.2541

OA−MHS

0.092

0.0274

PA−MHS

0.129

GA−MSA

CO−H···O (GA/OA/PA donor) Δq(H) a

ρ(BCP) b

2.144

0.083

0.0314

0.2238

2.174

0.2092

2.063

0.066

0.0281

0.1978

2.545

0.0337

0.2662

2.145

0.062

0.0309

0.2174

2.151

0.119

0.0323

0.2447

2.099

0.0322

0.2396

2.170

OA−MSA

0.088

0.0280

0.2103

1.997

0.090 0.071

0.0294

0.2158

2.525

PA−MSA GA−HMSA

0.118 0.053

0.0330 0.0559

0.2558 0.1006

2.105 2.183

0.074 0.034

0.0318 0.0494

0.2340 0.1050

2.148 2.167

OA−HMSA

0.057

0.0438

0.1063

2.342

0.024

0.0414

0.1098

2.170

PA−HMSA

0.058

0.0577

0.1000

2.180

0.027

0.0481

0.1046

2.145

∇2ρ(BCP) c

rd

∇2ρ(BCP) c

rd

Change in atomic charge Δq(H) at the H atom. b Electron density at bond critical point. c Laplacian density at bond critical point. d

The distance between a BCP and an RCP.

28 ACS Paragon Plus Environment

Page 29 of 37 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

Table 7 CCSD(T)//B3LYP-D3 calculated binding energy (BE), enthalpy of formation (Δ𝐻

), Gibbs energy of formation (Δ𝐺

), concentrations and dipole moments of

the MHS-, MSA- and HMSA-containing complexes at ambient conditions (298.15 K and 1 atm) a Conformer

BE

Δ𝐻

Δ𝐺

Concentrations Dipole moment

GA−MHS

-67.1

-66.7

-22.2

77

3.90

OA−MHS

-59.6

-58.7

-15.7

277

3.79

PA−MHS

-62.2

-61.7

-18.1

6

4.56

GA−MSA

-65.8

-65.7

-21.7

384

3.92

OA−MSA

-58.5

-58.2

-14.8

1202

4.00

PA−MSA

-60.9

-60.7

-17.6

31

4.28

GA−HMSA

-62.2

-61.9

-16.6