Journal Pre-proof Formation mechanism of methanesulfonic acid and ammonia clusters: A kinetics simulation study Dongping Chen, Danfeng Li, Changwei Wang, Fengyi Liu, Wenliang Wang PII:
S1352-2310(19)30800-3
DOI:
https://doi.org/10.1016/j.atmosenv.2019.117161
Reference:
AEA 117161
To appear in:
Atmospheric Environment
Received Date: 26 July 2019 Revised Date:
14 November 2019
Accepted Date: 16 November 2019
Please cite this article as: Chen, D., Li, D., Wang, C., Liu, F., Wang, W., Formation mechanism of methanesulfonic acid and ammonia clusters: A kinetics simulation study, Atmospheric Environment (2019), doi: https://doi.org/10.1016/j.atmosenv.2019.117161. 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. © 2019 Published by Elsevier Ltd.
1
Formation mechanism of methanesulfonic acid and ammonia clusters:
2
a kinetics simulation study
3
Dongping Chen, Danfeng Li, Changwei Wang, Fengyi Liu, Wenliang Wang ∗
4
Key Laboratory for Macromolecular Science of Shaanxi Province, School of Chemistry and Chemical
5
engineering, Shaanxi Normal University, Xi’an 710119, Shaanxi, China.
6
Abstract: The formation mechanism of methanesulfonic acid (MSA) and ammonia
7
(NH3) clusters is investigated using density functional theory (DFT) and Atmospheric
8
Cluster Dynamic Code (ACDC) in different conditions. The results suggest that
9
hydrogen bonding and electrostatic interactions induced by proton transfer could
10
provide the primary driving force that forms these clusters. NH3 can effectively
11
promote the formation of MSA-based clusters at pptv levels, but high concentrations
12
of precursors ([MSA] ≥ 2 × 107 molecules cm−3 and [NH3] ≥ 1 ppbv) is necessary for
13
effective formation. The formation of the initial (MSA)2 dimer is a rate-determining
14
step of cluster growth. The formation rate is proportional to the monomer
15
concentration and inversely proportional to the temperature in the troposphere.
16
Hydration has a great influence on the evaporation rate, formation rate, and nucleation
17
mechanism of the MSA-NH3 system. The relative evaporation rate of all clusters is
18
significantly affected by humidity, especially for the (MSA)(NH3) dimer. The
19
evaporation rate of the (MSA)(NH3) dimer can be reduced by approximately a factor
20
of 10−5 at a relative humidity (RH) ≥ 40%. The formation rate increases significantly
21
with RH and reached up to a factor of 105 at RH = 100%. The formation of the initial
22
(MSA)(NH3) dimer is the rate-determining step, which indicates that the nucleation
23
mechanism is fundamentally different from anhydrous cases. In addition, the results
24
showed that the formation of MSA-NH3 molecular clusters is relatively weak under
25
typical atmospheric conditions. In addition to the high concentration of precursors and
26
atmospheric humidity, the formation of MSA-based ternary clusters through the
27
participation of other species such as SA that may be more important in promoting
28
effective nucleation of MSA-NH3 system in the coastal atmosphere. ∗
Corresponding authors: Tel: +86-029-81530815; E-mail:
[email protected] (W. L.Wang). 1
29
Keywords: Methanesulfonic acid, Evaporation rate, Formation rate, Growth pathway,
30
Hydration
31
1. Introduction
32
New particle formation (NPF) from low-volatile gas precursors, as an important
33
source of atmospheric aerosol particles (Pope and Dockery, 2012; Seinfeld and Pandis,
34
2006), remains one of the least understood micro-processes in the atmospheric
35
troposphere. Atmospheric aerosol particles significantly contribute to cloud
36
condensation nuclei (CCN) (Dameto de España et al., 2017), atmospheric
37
transmittance, air quality, morbidity and premature mortality of human beings and
38
remain one of the leading uncertainties in global climate modeling and prediction
39
(Stocker et al., 2013). Despite recent advances in experimental and theoretical
40
methods, the mechanism governing the formation of seeds for thermally stable aerosol
41
particles involving multiple components is still unclear (Kumar et al., 2018),
42
especially in Northern China. It is well-known that sulfuric acid (SA) is closely
43
related to atmospheric NPF events (Weber et al., 2001) and is therefore of decisive
44
importance for the formation of atmospheric aerosol particles. It has also become
45
clear within the last decades that more stabilizing species, such as organic acids,
46
nitrogenous compounds, and highly oxygenated molecules (HOMs) (Ahlberg et al.,
47
2017) likely participate in the nucleation of atmospheric NPF (Wang et al., 2019;
48
Zhang et al., 2012). However, the nucleation mechanism involving organic acids and
49
nitrogenous bases remains unclear.
50
Methanesulfonic (MSA), one of the simplest organic organosulfur acids in the
51
atmosphere, is a prominent oxidation product from organosulfur compounds that
52
originate from biological processes, biomass combustion, industrial emissions, and
53
agriculture (Barnes et al., 2006), which appreciably contribute to atmospheric NPF
54
events in certain conditions (Chen and Finlayson-Pitts, 2017; Chen et al., 2016;
55
Dawson et al., 2012). MSA has been measured in atmospheric aerosol particles nearly
56
all geographic regions, ranging from coastal areas (Sorooshian et al., 2009) to the
57
hinterland (Gaston et al., 2010), and is present in concentrations of ~ 10-50% of 2
58
gaseous SA concentration (Berresheim, 2002). In a recent study (Dall'Osto et al.,
59
2012), the gaseous concentration of MSA was found to decrease during marine
60
particle formation events, suggesting that MSA contributed to the formation of initial
61
clusters. Station observations also reported a strong correlation between the
62
concentration of MSA and NPF events (Willis et al., 2016). In addition, results from
63
later studies by Dawson (Dawson et al., 2012) and Bzdek (Bzdek et al., 2011) found
64
that ammonia/amines distinctly increased the formation rate of MSA-based aerosol
65
particles using a flow tube experiment with simple ab initio calculations. Ammonia,
66
an important component of increasing atmospheric NPF (Jiang and Xia, 2017), is
67
ubiquitous in the atmosphere due to various emission sources ranging from livestock
68
farming to marine organism biodegradation (Ge et al., 2011). The concentration of
69
ammonia ranges from tens to tens of thousands of pptv in the troposphere (Ge et al.,
70
2011), which easily participates in forming particles (Glasoe et al., 2015; Kurtén et al.,
71
2008; Zollner et al., 2012). An increased MSA concentration was detected in small
72
particles when NH3 participated in NPF in several field observations (Kerminen et al.,
73
1997), which ultimately substantiated the role of NH3. Therefore, understanding the
74
mechanism governing atmospheric NPF and its relationship to MSA and NH3 is
75
important (Karl et al., 2007). The studies on the binary systems of MSA and ammonia
76
(Chen et al., 2016; Dawson et al., 2012) provide meaningful results regarding cluster
77
formation as well as useful information on the electronic structure. Numerous studies
78
involving MSA-NH3-H2O (Chen and Finlayson-Pitts, 2017; Chen et al., 2016; Li et
79
al., 2007; Wen et al., 2019) experimentally revealed that humidity has an important
80
influence on the formation of MSA-NH3 system. The research results (Bork et al.,
81
2014) showed that MSA-enhanced clustering involves clusters containing one MSA
82
molecule that contributes significantly to the growth of the SA-DMA system at low
83
temperatures. However, several fundamental questions involving the formation and
84
growth of clusters at the initial stage including the detailed dynamic growth
85
mechanism, the rate-determining step of clusters formation, and the effect of humidity
86
on the formation mechanism at the molecular level, remain unresolved.
87
To our knowledge, theoretical studies regarding the formation mechanism and 3
88
the influence of humidity on the formation mechanism of MSA-NH3 clusters at the
89
molecular level have not been previously reported. In addition, DFT and ACDC
90
methods have been successfully used to study the mechanism governing the formation
91
of SA-based systems in recent years (Kurtén et al., 2006; Li et al., 2017; Li et al.,
92
2018; Loukonen et al., 2010; Ortega et al., 2012). Herein, adopting the same method
93
based on DFT and ACDC, the thermodynamic properties, evaporation rate, formation
94
rate, growth pathway, and hydration level were investigated in order to gain a deeper
95
understanding into the mechanism governing the formation of (MSA)x(NH3)y (x, y ≤
96
4) clusters under different monomer concentrations (MSA and NH3), temperature, and
97
relative humidity (RH) conditions.
98
2. Computational details
99
2.1 Theoretical calculation and analysis methods
100
The Gibbs free energy (G) of clusters is the most critical parameter for
101
identifying the formation mechanism of clusters. The calculated reliability of G was
102
determined by the structure of the global minimum (here it refers to energy minimum)
103
of (MSA)x(NH3)y (x, y ≤ 4) clusters. The global minimum sampling technique is an
104
efficient and convenient tool for searching for atmospheric clusters (Elm et al., 2013b;
105
Elm et al., 2016a; Elm et al., 2015). In this study, this method was used to locate
106
global minima for (MSA)x(NH3)y (x, y ≤ 4) clusters; a detailed flow chart is shown in
107
Fig. S1. Theoretical calculations and kinetics simulations were performed using the
108
ABCluster (Zhang and Dolg, 2015; Zhang and Dolg, 2016), MOPAC (Stewart, 1990)
109
and GAUSSIAN 09 programs (Frisch et al., 2009). The molecular structure of the
110
precursor was described using the CHARMM36 force field (Yin and MacKerell, 1998)
111
in the ABCluster program. The M06-2X functional and 6-31++G(d,p) basis set was
112
used to describe the noncovalent interaction and equilibrium structure of the
113
atmospheric clusters (Elm et al., 2013a; Elm and Kristensen, 2017). Single point
114
energy calculations at the DLPNO-CCSD(T) (domain-based local pair natural orbital
115
coupled cluster)/aug-cc-pVTZ level (Riplinger and Neese, 2013; Riplinger et al., 2013)
116
were performed in ORCA version 4.0.0 (Neese, 2012). The DLPNO-CCSD(T)/ 4
117
aug-cc-pVTZ level of theory was utilized to yield a mean absolute error of 1.3
118
kJ·mol−1 compared to CCSD(T) complete basis set limit (CBS) estimates, based on a
119
test set of 10 small atmospheric cluster reactions (Myllys et al., 2016). Herein, the
120
value of G for monomers and clusters was estimated by combining the single-point
121
energy with the Gibbs free energy correction term at the DLPNO-CCSD(T)/
122
aug-cc-pVTZ and M06-2X/6-31++G(d,p) level at 298.15K. The 6-31++G(d,p) basis
123
set is widely used for the optimization of atmospheric clusters because of its fast
124
convergence rate and reliable results compared with other larger basis sets (Elm et al.,
125
2013a; Elm and Kristensen, 2017). Additionally, the Gibbs free energy of the
126
formation (∆Gref) of the clusters was calculated at different temperatures under the
127
approximation that the enthalpy of formation (∆Href) and entropy of formation (∆Sref)
128
remain unchanged in the troposphere. Considering the concentration of precursors in
129
the atmosphere, the actual Gibbs free energy of the formation (∆Gact) of the clusters
130
was obtained by converting the ∆Gref from 1 atm to the actual vapor pressure of a
131
particular species. ∆Gact is different from ∆Gref, which can be used as a
132
thermodynamic criterion to evaluate the spontaneity of cluster formation (a detailed
133
description of ∆Gref and ∆Gact are shown in the SI). Analyses of Atoms in Molecules
134
(AIM) theory (Bader, 1990) and the noncovalent interaction (NCI) index (Johnson et
135
al., 2010) for the dimers were performed using Multiwfn 4.3 software (Lu and Chen,
136
2012) in this study. (MSA)x(NH3)yWz (x, y ≤ 2, where “W” stands for water, and z is
137
the number of water molecules) clusters were investigated while considering the
138
effect of hydration on the MSA-NH3 system. This was similar to procedures used for
139
anhydrous clusters in order to search for the global minimum of the hydrated clusters.
140
Moreover, for the purposes of comparing the formation rate of SA-NH3 and the
141
MSA-NH3 system directly, the corresponding ∆G values of (SA)x(NH3)y (x, y ≤ 4)
142
clusters were re-calculated at the same theoretical level (DLPNO-CCSD(T)/
143
aug-cc-pVTZ//M06-2X/6-31++G(d,p)) based on those cluster structures reported
144
(Olenius et al., 2017).
145
2.2 Kinetics simulation
146
The kinetics simulation of the atmospheric clusters is important for 5
147
understanding their nucleation mechanism in a real atmospheric environment. One
148
note is that the key input file for a kinetics simulation is the ∆G values of relevant
149
clusters based on the DLPNO-CCSD(T)/aug-cc-pVTZ//M06-2X/6-31++G(d,p) level.
150
ACDC is widely used to determine certain key parameters of cluster growth, such as
151
the evaporation rate, formation rate, and growth pathway. Moreover, ACDC is useful
152
for understanding the dynamic mechanism of cluster growth at the molecular level.
153
We refer readers to the study from McGrath et al (McGrath et al., 2012) for a detailed
154
theoretical description of the ACDC program. In brief, equations for the time
155
derivative of the concentration of any cluster are generated in the code using the
156
Matlab ode15s routine (Shampine and Reichelt, 1997), which are then used to solve
157
differential equations and simulate the time-dependent concentration of various
158
clusters. These differential equations, called birth-death equations, mainly comprise
159
source terms for small molecules (clusters) that collide to form large clusters, as well
160
as large clusters that evaporate into small molecules (clusters) and so on. In addition,
161
the formation rate of clusters in the ACDC program is defined as the flux outside the
162
system. The flux outside of the system for a cluster is determined by the boundary
163
condition. In an ACDC simulation, hydration is considered by regarding H2O as a
164
nucleation species that influence the evaporation rate, formation rate, and nucleation
165
mechanism of the clusters (Henschel et al., 2016). A “4 × 4 box” system was
166
simulated for the anhydrous system, where 4 is the maximum of MSA or NH3
167
molecules in the studied system. This is required for (MSA)5(NH3)5 to grow outside
168
the MSA-NH3 system and all other clusters crossing the box edge are brought back to
169
the simulation box by monomer evaporations (detailed discussion of the boundary
170
condition can be found in the SI). Low temperature is very helpful for the NPF
171
because of the lower entropy penalty from the second term in the free energy (∆G =
172
∆H − T∆S). In this work, in order to comprehensively evaluate the influence of
173
temperature on the formation rate of MSA-NH3 clusters, a temperature range from
174
220 to 290 K was chosen in our ACDC simulations. The primary ACDC simulation
175
was conducted at 278.15 K, while additional simulations were performed to study the
176
effect of temperature at 220, 240, 260, 280, and 290 K. Regarding the SA-containing 6
177
system, 2.6 × 10−3 s−1, was used as the constant coagulation sink coefficient, which
178
corresponds to a typical value observed in a boreal forest (Olenius et al., 2013). In
179
addition, it was reported that the constant coagulation sink coefficient has a slight
180
effect on the formation rate in the SA-containing system (Olenius et al., 2017).
181
Therefore, a value of 2.6 × 10−3 s−1 was used as the constant coagulation sink
182
coefficient in the MSA-containing system. In addition, the concentration of MSA was
183
defined in the range from 104 to 108 molecules·cm−3, and the concentration of NH3
184
ranged from 10 to 1000 pptv; these ranges are also relevant for atmospheric NPF
185
(Almeida et al., 2013). It is worth mentioning that the acid concentration is considered
186
to be the total concentration of any neutral cluster that includes one acid molecule
187
(Almeida et al., 2013). Therefore, the simulated system becomes a “2 × 2 box” when
188
hydration is considered. For (MSA)x(NH3)y (0 ≤ x, y ≤ 2) clusters, the average
189
collision and evaporation coefficients based on the hydrate distribution are used in the
190
birth-death equations for [MSA] = 106 molecules·cm−3, [NH3] = 100 pptv, and T =
191
278.15 K. For each cluster, we calculated the equilibrium hydrate distribution with the
192
equilibrium constant in terms of the value of ∆Gref for the hydrate (Henschel et al.,
193
2016).
194
3. Results and discussion
195
3.1 Structure, topology, and average partial charge analysis
196
(MSA)x(NH3)y indicates a cluster containing x (x ≤ 4) MSA and y (y ≤ 4) NH3
197
molecules, and therefore the proton transfer status does not need to be clearly
198
specified. The structures of pure (NH3)y (y = 2 to 4) clusters were also discussed in
199
previous studies (Ling et al., 2017). Herein, (MSA)x(NH3)y (1 ≤ x ≤ 4 and y ≤ 4)
200
clusters will be the primary focus. All structures are shown in Fig. S2. In general,
201
proton transfer does not occur in (MSA)x (x = 2 to 4) and (NH3)y (y = 2 to 4)
202
homomolecular clusters, and they are typically stabilized by hydrogen bonding.
203
Regarding pure (MSA)x clusters, more complicated network structures were formed
204
due to the presence of more O−H···O hydrogen bonds from x = 2 to 4 (see Fig. S2).
205
Similar to pure MSA clusters, a typical parallelogram ring, six-membered ring, and 7
206
eight-membered ring form by two, three, and four N−H hydrogen bonds in the (NH3)2,
207
(NH3)3 and (NH3)4 clusters, respectively (detailed structural information is shown in
208
Fig. S2). In pure clusters, the length of a hydrogen bond (N−H···N) in pure NH3
209
clusters is longer than that in (O−H···O) of a pure MSA cluster. (MSA)x (2 ≤ x ≤ 4)
210
and (NH3)y (2 ≤ y ≤ 4) clusters tend to form spherical and planar configurations due to
211
the structural differences in MSA and NH3 molecules as x and y increase, respectively.
212
This also indicates that hydrogen bonding in pure MSA clusters is stronger than that
213
in pure NH3 clusters.
214
For all heteromolecular clusters, proton transfer occurred from an MSA molecule
215
to an NH3 molecule when the number of NH3 molecules is equal or larger than that of
216
MSA molecules, except for (MSA)(NH3). All stabilized via hydrogen bonding and
217
electrostatic interactions between positive and negative species. Several studies (Elm,
218
2017; Elm et al., 2016b; Olenius et al. 2013; Olenius et al. 2017) show that clusters
219
with an equal number of acid and base molecules are more stable and play an
220
important role in subsequent cluster growth. Herein, the structures of the
221
(MSA)x(NH3)y (x ≠ y ≠ 0) clusters are not described in detail; more information is
222
presented in Fig. S2. In an (MSA)(NH3) dimer, no proton is transferred, but one
223
six-membered ring is formed by one strong hydrogen bond (O−H···N) and one weak
224
hydrogen bond (N−H···O) (see Fig. S2). This is consistent with the conclusion from
225
Born-Oppenheimer molecular dynamics (BOMD) simulations, i.e., proton transfer
226
between one MSA and one NH3 molecule in the absence of water at 298.15 K and
227
1atm is impossible (Kumar and Francisco, 2017). In (MSA)2(NH3)2, four strong and
228
four weak hydrogen bonds form alternately between MSA and NH3 molecules. It
229
should be mentioned here is that (MSA)2(NH3)2 is the most stable cluster due to its
230
high structural symmetry and saturated hydrogen bonds. In (MSA)3(NH3)3, six strong
231
hydrogen bonds form by anchoring two NH3 molecules to form a discus shape
232
composed of three MSA molecules on the upper and lower sides. One should note that
233
the unique net structure forms through proton transfer in (MSA)4(NH3)4. Here, four
234
NH+4 and four CH3 SO3 ions alternately occupy 8 vertices in this structure. For
235
(MSA)x(NH3)y (x = y ≠ 1) clusters, proton transfer helps form a more stable 8
236
three-dimensional spherical structure. In addition, the patterns of proton transfer in the
237
MSA-NH3 system (see Fig. S2) are different from those in the SA-NH3 system
238
(Olenius et al., 2017; Olenius et al., 2013) due to the difference in the number of −OH
239
and −CH3 groups of acid molecules. The steric hindrance of the –CH3 in clusters is
240
unfavorable for cluster formation. Moreover, the contribution of the stronger acidity
241
of SA than MSA cannot be ignored in the different acid-base clusters (Elm et al.,
242
2016b; Elm et al., 2015). These factors together cause the formation capability of
243
MSA to be far lower than that of SA in the formation of atmospheric clusters.
244
Results from previous studies (Olenius et al., 2017) showed that the formation of
245
initial dimers is critical for the overall formation process. Herein, only three initial
246
dimers, i.e. (NH3)2, (MSA)2, and (MSA)(NH3), were chosen for AIM and NCI
247
analyses to obtain useful information about the formation of the initial clusters.
248
Topological analysis of AIM indicates that the electron density (ρ) at bond critical
249
points (BCPs) in the N−H…N, O−H…O, and O−H…N hydrogen bonds in (NH3)2,
250
(MSA)2 and (MSA)(NH3) clusters are 0.013 (0.013), 0.040 (0.039), and 0.079 a.u.,
251
respectively (see SI Text and Table S1). The value of ρ represents the hydrogen bond
252
strength. As an extension of AIM theory, the NCI analysis also shows that there are
253
low-reduced gradient spikes (blue or green color) at low density of the sign(λ2)ρ value
254
of −0.013, −0.040 and −0.079 a.u., which suggests that two weak, two strong, and one
255
strong hydrogen bonds are present in (NH3)2, (MSA)2, and (MSA)(NH3), respectively
256
(see the lower panel in Fig. 1). In addition, the color-coding of these bonding
257
isosurfaces in the dimers switches from green to blue in the upper panel in Fig. 1,
258
indicating that the strength of a hydrogen bond ranges from weak to strong. From the
259
three analyses presented above, hydrogen bonding in (MSA)2 and (MSA)(NH3) is far
260
stronger than that in (NH3)2, especially (MSA)2 (see SI Text). Average partial charge
261
(δA) analysis also shows that δA in NH3 ranges from 0.8280 to 0.8845, except for one
262
(MSA)(NH3) cluster (δA = 0.1285) (see Table S2). Larger values (δA > 0.8) indicate
263
that electrostatic force from ion pairs formed by proton transfer from MSA to NH3
264
becomes stronger, which favors the growth of large clusters (see SI Text and Table
265
S2). 9
266
3.2 Thermodynamic properties and evaporation rate
267
∆Gref for the (MSA)x(NH3)y (x, y ≤ 4) clusters from their component molecules
268
at 298.15 K is shown in Fig. 2a, whereas the corresponding thermodynamic properties
269
(∆Href and ∆Sref) are shown in Table S3. For the (MSA)x(NH3)y (x, y ≤ 4) clusters, the
270
values of ∆Gref and ∆Href decreased as x and y increased, respectively. Although Href is
271
related to the stability (Curtiss and Blander, 1988), the Gibbs free energy is essentially
272
the determining factor for the stability of a cluster. The entropy contribution to the
273
formation process of clusters is very important. This indicates that the stability of the
274
cluster was gradually enhanced as the total number of cluster molecules increased.
275
Herein, ∆Gref for the three initial dimers ((NH3)2, (MSA)2, and (MSA)(NH3)) is
276
highlighted because its value is important for the initial formation and subsequent
277
growth of clusters. ∆Gref = −8.92, −5.37, and 3.95 kcal·mol−1 for (MSA)2,
278
(MSA)(NH3), and (NH3)2, respectively. From a thermodynamics perspective, the
279
formation of (MSA)2 is the most favorable. In addition, for the MSA-NH3 and
280
SA-NH3 systems, a ∆Gref for (MSA)x(NH3)y (x, y ≤ 4) clusters is generally higher than
281
that for the corresponding (SA)x(NH3)y (x, y ≤ 4) clusters (see Fig. S3). For example,
282
the noncovalent interaction in (MSA)x(NH3)y clusters is weaker than that of the
283
(SA)x(NH3)y clusters. The main reason for this is the difference in the ability to form
284
hydrogen bonds and steric hindrance between −OH and −CH3 functional groups. Two
285
−OH groups in SA are more favorable for the formation of hydrogen bonds than one
286
−OH and one −CH3 in MSA. It is tempting to conclude that the −CH3 functional
287
group in an MSA molecule hinders the formation of (MSA)x(NH3)y clusters.
288
Considering the growth of clusters at a specific concentration, the formation rate
289
(details are present in section 3.3) can be exported by comparing the collision and
290
evaporation rates of the clusters, which is important for better understanding their
291
formation mechanism. Although the collision and evaporation rates of clusters can be
292
easily determined using ACDC kinetics simulations, the difference in collision rates
293
for all clusters is small; thus, the evaporation rate is used to evaluate the stability of
294
clusters in this section. Those with an evaporation rate in the range of 10−4 to 10−3 s−1
295
are considered stable when the concentration of the precursors (MSA and NH3) is 10
296
approximately equal to or above pptv levels (Kulmala, 2013; Kulmala et al., 2000).
297
(MSA)3(NH3)3 can be considered stable because its evaporation rate is lowest and is
298
far lower than that of all others in the system studied at 278.15 K (see Fig. 2b). All the
299
evaporation pathways were also checked (see Table S4), and it was found that
300
evaporation of the NH3 monomer was the main decomposition channel for all relevant
301
clusters when x = y. When x < y, evaporation of the NH3 monomer is always preferred
302
in the studied clusters. For example, an (MSA)2(NH3)3 cluster is easily converted to
303
(MSA)2(NH3)2 by NH3 monomer evaporation due to its high evaporation rate (4.92 ×
304
106 s−1). Evaporation of an (MSA)2 dimer or MSA monomer is the dominant
305
dissociation channels when x > y. In other words, when an (MSA)3(NH3)2 cluster
306
changes to (MSA)2(NH3)2 by evaporation of an MSA monomer, its evaporation rate
307
reaches 104 s−1. Moreover, when the total number of molecules in different clusters is
308
equal, the evaporation rate of abundant NH3 clusters is higher than the corresponding
309
abundant MSA clusters, and the evaporation of an NH3 monomer is faster than that of
310
MSA because MSA combines with clusters more easily than NH3. Overall, the
311
evaporation rate of clusters on or near the diagonal is far lower than that of clusters
312
away from the diagonal of the MSA-NH3 system. This also indicates that the channel
313
with a low evaporation rate is more favorable for cluster growth. It is also true that the
314
ability for an acid to bond is stronger than that of a base, and evaporation of a base
315
monomer is the primary decay route for larger clusters in the SA-NH3 (McGrath et al.,
316
2012), SA-MA (Olenius et al., 2017), SA-DMA (McGrath et al., 2012), and SA-MEA
317
systems (Xie et al., 2017). It is also interesting to compare the total evaporation rate of
318
the MSA-NH3 and SA-NH3 systems in identical simulation conditions. In general, the
319
evaporation rate of the MSA-NH3 system is higher than that of the SA-NH3 system,
320
showing that the SA-NH3 system is more stable (see Fig. S4). The essential
321
differences are that SA is much more acidic than MSA, and the ability to form the
322
non-covalent interaction for −OH is far stronger than that for −CH3. The former is to
323
form hydrogen bonding whereas the latter is Van der Waals interaction.
324
3.3 (MSA)2 dimer concentration and formation rate of the MSA-NH3 system
325
Similar to the discussion regarding SA-base systems (Olenius et al., 2017; Xie et 11
326
al., 2017), the concentration of the precursors also affects cluster formation, especially
327
the initial dimers. As described in section 3, for three initial dimers, the evaporation
328
rate of (MSA)2 (102 s−1) is lower than that of (MSA)(NH3) (5 × 105 s−1) and (NH3)2 (3
329
× 1012 s−1) by 3 and 10 orders of magnitude, respectively. This confirms that (MSA)2
330
forms more easily than (MSA)(NH3) and (NH3)2 during cluster growth. Herein, the
331
concentration of the (MSA)2 dimer (hereafter called [(MSA)2]) will be the primary
332
focus. As shown in Fig. 3a, [(MSA)2] increases in proportion to [MSA] and [NH3],
333
and the low [NH3] (≤ 100 pptv) has little effect on [(MSA)2]. In contrast, the high
334
[NH3] (> 100 pptv) has an apparent influence on [(MSA)2] in the aforementioned
335
condition, where T = 278.15 K and P = 1 atm. Overall, [(MSA)2] ranges from 10−3 to
336
107 molecules·cm−3, which is higher than that of [(MSA)(NH3)] and [(NH3)2]. Thus, it
337
is clear that [(MSA)2] is critical for determining the formation rate and the overall
338
formation process. Regarding the SA-NH3 system, [(SA)2] and [SA] or [NH3] exhibit
339
linear relationships at low and medium [SA], which is slightly different compared to
340
the MSA-NH3 system (see Fig. S5a).
341
The formation rate, an important quantification index for cluster formation, is
342
used to evaluate the capability of NH3 to promote MSA-based cluster formation in
343
this section; this is similar to the case with SA-based clusters (Olenius et al., 2017).
344
Fig. 3b shows that the formation rate is a function of the monomer concentration
345
([MSA]: 104, 105, 106, 107, and 108 molecules·cm−3, [NH3]: 10, 100, and 1000 pptv)
346
in the MSA-NH3 system at 278.15 K. Generally, the formation rate is proportional to
347
[MSA] and [NH3], and [NH3] is inversely proportional to [MSA] when the formation
348
rate is constant. This figure also shows that the MSA-NH3 system progressively
349
inclines toward saturation at high [MSA], thus cluster growth is dominated by MSA.
350
The simulation results also show that NH3 could reinforce MSA-based NPF when the
351
atmospheric concentration of MSA and NH3 reach or exceed ppbv levels, which is
352
consistent with the experimental results (Chen et al., 2016). Regarding the MSA-NH3
353
system, the formation rate between MSA and NH3 exhibits a linear relationship with
354
[MSA] when 104 ≤ [MSA] ≤ 107 molecules·cm−3. However, when [MSA] > 107
355
molecules·cm−3, the formation rate increased moderately due to the significant 12
356
deposition of pure MSA clusters. In addition, the effect of high [MA] on the formation
357
rate was less than that of high [MSA] for the MSA-NH3 system (see Fig. 3b). This
358
difference in the SA-NH3 system arises because the −OH functional group acts as a
359
strong donor in a SA molecule, which contains more −OH than an MSA molecule. In
360
addition, the formation rate is inversely proportional to the temperature within the 220
361
to 290 K range. The reduced formation rate at high temperatures is more obvious than
362
that at low temperatures (see Fig. S6). The reverse effect of temperature dependence
363
on the concentration of precursors is not clear. In this work, considering the
364
concentration of precursors and temperature in the lower troposphere, subsequent
365
kinetics simulations were performed with the NH3 concentration set to 100 pptv, MSA
366
concentration set to 106 molecules·cm−3 and temperature set to 278.15K. Furthermore,
367
the formation rates of the MSA-NH3 (4 × 4) system are far lower than those of the
368
SA-NH3 (4 × 4) system under typical atmospheric conditions (see Fig. S5b). This
369
further implies that, in addition to the high concentration of precursors, the
370
participation of other species is also very important to promote the effective
371
nucleation of the MSA-NH3 system by forming stable MSA-based ternary clusters in
372
the lower troposphere over coastal areas.
373
3.4 Actual Gibbs free energy and growth pathway
374
The actual Gibbs free energy (∆Gact) and growth pathway in the MSA-NH3
375
system are shown in Figs. 4 and 5a, where [MSA] = 106 molecules·cm−3, [NH3] = 100
376
pptv, and T = 278.15 K. Considering the concentration of precursors in the
377
atmosphere, ∆Gact is obtained by converting the free energy change from 1 atm to the
378
actual vapor pressure of a particular species. ∆Gact for the clusters along or near the
379
diagonal is much lower than that far from the diagonal. For the MSA-NH3 system, the
380
growth pathway along or near the diagonal is the most advantageous, but free energy
381
barriers in all directions in this system signified that these small clusters are difficult
382
to grow in a typical atmosphere. Herein, we note that ∆Gact is different from ∆Gref.
383
The ∆Gact is greater than zero for all clusters, which cannot be used as a
384
thermodynamic criterion to assess the spontaneity of the process (see SI Eqs. 7 and 8).
385
Combined with the discussion in the previous section, (MSA)2 is a greater 13
386
determinant in the first step of cluster formation than (MSA)(NH3) and (NH3)2 from
387
the perspective of topology analyses, thermodynamics, and evaporation rate. The
388
primary flux out of the simulated box is the (MSA)4(NH3)5 cluster, and some features
389
can be observed through the growth pathway and the Gibbs free energy of the
390
formation of clusters. First, the cluster growth pathways do not fully follow the lowest
391
free energy pathway; they deviate from the diagonal in the acid-base grid. The first
392
step of cluster formation is primarily controlled by collisions between MSA molecules
393
due to strong hydrogen bonding and the low evaporation rate of the (MSA)2 dimer
394
(see section 3.1 and 3.2). The second step is primarily controlled by the collision
395
between (MSA)2 and an NH3 monomer because the concentration of (MSA)2 is larger
396
than the other dimers. Thus, (MSA)2 plays a key role in subsequent cluster growth,
397
which is also an important reason why growth deviates from the diagonal in the grid
398
at the initial stage. Thirdly, the primary flux out, i.e., (MSA)4(NH3)5 cluster, has two
399
pathways. The initial stage before formation of (MSA)2(NH3)2 is the same, and
400
subsequent growth is divided along two pathways as follows: (1) (MSA)2(NH3)2→
401
(MSA)4(NH3)4 → (MSA)4(NH3)5 and (2) (MSA)2(NH3)2 → (MSA)4(NH3)2 →
402
(MSA)4(NH3)3→(MSA)4(NH3)4→(MSA)4(NH3)5. Regarding the flux out, pathway (1)
403
is more important because its proportional contribution is as high as 89%. From the
404
discussion in section 3.1, although (MSA)3(NH3)3 is the most stable cluster in the
405
whole system, however, the (MSA)3(NH3)3 cluster is difficult to form, as both the
406
(MSA)2(NH3)3 and (MSA)3(NH3)2 clusters have high evaporation rates. Therefore, no
407
(MSA)3(NH3)3 cluster is present in the favorable growth pathways on the acid-base
408
grid. Finally, one can conclude that the formation of an (MSA)2 dimer is the
409
rate-determining step for cluster growth when the growth pathway and evaporation
410
rate are considered. One should note that temperature has no effect on the growth
411
pathways in the initial stage; however, it has obvious effects in the subsequent stage
412
after the formation of an (MSA)2(NH3)2 cluster (see Fig. S7).
413
The growth pathways in the MSA-NH3 and SA-NH3 systems are presented in
414
Figs. 5a and S8 under the same simulation conditions, respectively. One common
415
characteristic is that homogeneous dimers initially form from two MSA or two SA 14
416
molecules in the initial stage at 278.15 K, which is critical for initiating cluster growth.
417
Nevertheless, there are some differences regarding all growth pathways for the
418
MSA-NH3 and SA-NH3 systems at different temperatures (see Figs. 5a, S7, and S8).
419
Regarding the SA-NH3 system, the formation of the initial (SA)(NH3) and (SA)2
420
dimers are more important for determining subsequent growth channels at low (220 K)
421
and high (278.15 K) temperatures, respectively (see Fig. S8). Based on the results of a
422
previous study (McGrath et al., 2012), one conclusion can be drawn that a collision
423
comprising (SA)(NH3) is dominant for the SA-NH3 cluster growth at low
424
temperatures, where growth proceeds along the diagonal in the acid-base grid.
425
Regarding the MSA-NH3 system, the effect of temperature on the growth pathway is
426
different from that in the SA-NH3 system. The temperature has no effect on the initial
427
formation stage for (MSA)x(NH3)y cluster and vice versa. For example, the cluster
428
growth pathway is complicated at low temperature, as shown in Figs, S7 and S8a.
429
3.5 Effect of hydration on the nucleation mechanism of MSA-NH3 system
430
Water is more abundant than acids and bases in the troposphere by several to 10
431
orders of magnitude. Therefore, the effects of hydration on the evaporation rate,
432
formation rate, and growth pathway of clusters were investigated in this work
433
(DePalma et al., 2014). As found in previous studies (DePalma et al., 2014; Henschel
434
et al., 2016), clusters containing SA, ammonia, or simple organic amines are primarily
435
hydrated by a few water molecules. Herein, small anhydrous clusters and a few water
436
molecules are considered together to study the effect of hydration on the formation
437
mechanism in the studied system. In addition, for the purposes of saving computing
438
resources, small (MSA)x(NH3)y (x, y ≤ 2) clusters were selected as the testing system
439
to study hydration. The equalizing hydrate distribution was converted from the ∆Gref
440
of clusters hydrated at different RH values, a temperature of 278.15 K and 1 atm. One
441
can conclude that pure MSA and heterogeneous clusters are hydrated from Fig. S9. A
442
detailed discussion regarding ∆Gref, structure, and average number of water molecules
443
in individual clusters was presented in SI Text, Figs. S10, and S11, respectively. In
444
this study, the effect of hydration on the kinetics simulation for cluster formation was
445
the primary focus. 15
446
In principle, the formation rate of clusters can be affected by hydration due to
447
changes in the collision rate and evaporation rate. However, hydration has little or no
448
effect on the collision rate because the collision diameter, which is a significant factor
449
for the collision rate in the kinetic collision theory adopted in ACDC, has less
450
dependence on hydration (Henschel et al., 2016). Herein, the effect of hydration on
451
the evaporation rate, formation rate, and growth pathway are discussed. Fig. 6 show
452
that the evaporation rate and formation rate serve as a function of RH at 278.15 K in
453
comparison to anhydrous conditions. Overall, the presence of water produces a
454
variety of effects on the evaporation rates for specific clusters.
455
Fig. 6a shows that the RH value can slightly affect the evaporation rate of
456
(MSA)2(NH3), whereas the RH value has almost no effect on that of (NH3)2 and
457
(MSA)(NH3)2, but it effectively increases the evaporation rate of (MSA)2 by
458
approximately a factor of 103. In addition, the evaporation rates of (MSA)(NH3) and
459
(MSA)2(NH3)2 are significantly reduced by factors of 10−5 and 10−3 compared to the
460
anhydrous case, respectively. From Table S5, the extra stabilization energy from water
461
dominated the change of the evaporation rate of clusters. The results of the
462
Finlayson-Pitts group (Chen et al., 2016) showed that the effect of hydration (RH =
463
42%) on the formation was up to a factor of 106, which is generally consistent with
464
our simulated results. Therefore, it is clear that hydration has the most obvious effect
465
on the evaporation rate of the initial (MSA)(NH3) cluster, which is critical for the
466
overall formation process. Similar to the discussion presented in section 3.3, the
467
primary flux out of the simulated box (2 × 2) consists of (MSA)3(NH3)2 and
468
(MSA)3(NH3)3. The growth pathways of the (MSA)3(NH3)3 (77%) and (MSA)3(NH3)2
469
(23%) clusters are as follows: MSA→(MSA)(NH3)→(MSA)2(NH3)2→(MSA)3(NH3)3
470
and MSA→(MSA)(NH3)→(MSA)2(NH3)2→(MSA)3(NH3)2, respectively (see Fig. 5b).
471
When Figs. 5a and 5b are compared, hydration clearly changes the initial cluster
472
formation compared with anhydrous cases. Apparently, (MSA)(NH3) plays a more
473
important role than (MSA)2 in cluster formation at the initial stage, which is
474
completely different from anhydrous conditions. As shown in Fig. 6b, the relative
475
formation rate can be increased by approximately 105 times when RH = 100%, which 16
476
further confirms that humidity has a major influence on the studied system. This also
477
explains why humidity can affect the formation rate and mechanism. Based on the
478
discussion presented above, SA is more important than MSA in the NPF process in
479
anhydrous cases, but the formation rate of the MSA-NH3 system increases
480
significantly in the presence of water (up to a factor of 106), which is even higher than
481
that in the water-free SA-NH3 system. This indicates that the effect of hydration on
482
the formation rate in the MSA-NH3 system cannot be ignored in a heavily polluted
483
atmosphere ([NH3] ≥ 1ppbv) with high humidity, especially in coastal areas where the
484
concentration of precursors are also relatively high ([MSA] ≥ 2 × 107 molecules cm−3).
485
In addition, one should note that the influence of humidity on the formation rate is
486
closely related to the structure of nucleation precursors. Due to the steric hindrance of
487
−CH3 in MSA, hydrogen bonds are less likely to form in the MSA-NH3 system, and
488
the formation rate of the MSA-NH3 system is lower than that of the SA-NH3 system
489
in anhydrous cases. Although a box (2 × 2) can introduce bias into the absolute
490
formation rate compared with the actual formation rate, attention should be paid to the
491
fact that the relative formation rate stated is favorable to reduce the bias generated
492
through the small box used in this work. Results from kinetics simulations involving
493
small hydrated clusters allow one to conclude that hydration has a significant effect
494
on the formation mechanism and greatly increases the formation rate of the MSA-NH3
495
system. Although the use of larger clusters and more water molecules in
496
contemporary studies should not lead to different qualitative conclusions, such a case
497
is still worth studying in the future, with the goal of reaching a conclusion regarding
498
the effect of humidity on the kinetics governing MSA-NH3 cluster formation.
499
4. Conclusions
500
In this work, the MSA-NH3 system was investigated using quantum mechanics
501
and kinetics simulations under different conditions. The findings suggest that
502
hydrogen bonding and electrostatic interactions provide the primary force that drives
503
cluster formation. The different concentrations of MSA and NH3 have a significant
504
influence on the formation rate of MSA-NH3 clusters under the water-free condition, 17
505
and the formation of (MSA)2 dimer is the rate-determining step. Compared to
506
anhydrous cases, hydration causes the formation of (MSA)(NH3) dimer to be the
507
rate-determining step. The formation rate increased significantly by 105 times. This
508
suggests that the effect of humidity on the formation of the MSA-NH3 clusters is
509
significant in the coastal atmosphere. Generally, the effective nucleation for the
510
MSA-NH3 system is difficult to occur due to its weak stability under typical
511
atmospheric conditions. The high concentration of precursor and atmospheric
512
humidity is necessary for the effective nucleation of the MSA-NH3 system in an
513
atmospheric environment. Other species such as SA are involved in forming the
514
MSA-SA-NH3 ternary clusters that may be more effective to promote the occurrence
515
of effective nucleation. The results presented in this study are essential for better
516
predicting NPF. Therefore, this study provides theoretical guidance for a better
517
understanding of atmospheric pollution.
518
Acknowledgments
519
This work was supported by the National Natural Science Foundation of China
520
(No: 21473108, 21873060, 21636006) and Fundamental Research Funds for the
521
Central Universities (Grant No. GK201901007).
522
We gratefully acknowledge the valuable help of Hanna Vehkamäki (University
523
of Helsinki), Tinja Olenius (Stockholm University), Zhang Xiuhui (Beijing Institute
524
of Technology), LUO Yi (Dalian University of Technology), Liu Yirong (University
525
of Science and Technology of China), ZHANG Jun (University of Illinois), ZHAO
526
Xianwei (Shandong University) and LIN Jinfei (Shenzhen Transsion Holdings
527
Limited).
528
We thank Dr. Anand Parkash and LetPub (www.letpub.com) for its linguistic
529
assistance during the preparation of this manuscript.
530
Reference:
531 532 533 534 535
Ahlberg, E., Falk, J., Eriksson, A., Holst, T., Brune, W.H., Kristensson, A., Roldin, P., Svenningsson, B., 2017. Secondary organic aerosol from VOC mixtures in an oxidation flow reactor. Atmos. Environ. 161, 210-220. Almeida, J., Schobesberger, S., Kurten, A., Ortega, I.K., Kupiainen-Maatta, O., Praplan, A.P., Adamov, A., Amorim, A., Bianchi, F., Breitenlechner, M., David, A., Dommen, J., Donahue, N.M., Downard, 18
536 537 538 539 540 541 542 543 544 545 546 547 548 549 550 551 552 553 554 555 556 557 558 559 560 561 562 563 564 565 566 567 568 569 570 571 572 573 574 575 576 577 578 579
A., Dunne, E., Duplissy, J., Ehrhart, S., Flagan, R.C., Franchin, A., Guida, R., Hakala, J., Hansel, A., Heinritzi, M., Henschel, H., Jokinen, T., Junninen, H., Kajos, M., Kangasluoma, J., Keskinen, H., Kupc, A., Kurten, T., Kvashin, A.N., Laaksonen, A., Lehtipalo, K., Leiminger, M., Leppa, J., Loukonen, V., Makhmutov, V., Mathot, S., McGrath, M.J., Nieminen, T., Olenius, T., Onnela, A., Petaja, T., Riccobono, F., Riipinen, I., Rissanen, M., Rondo, L., Ruuskanen, T., Santos, F.D., Sarnela, N., Schallhart, S., Schnitzhofer, R., Seinfeld, J.H., Simon, M., Sipila, M., Stozhkov, Y., Stratmann, F., Tome, A., Trostl, J., Tsagkogeorgas, G., Vaattovaara, P., Viisanen, Y., Virtanen, A., Vrtala, A., Wagner, P.E., Weingartner, E., Wex, H., Williamson, C., Wimmer, D., Ye, P., Yli-Juuti, T., Carslaw, K.S., Kulmala, M., Curtius, J., Baltensperger, U., Worsnop, D.R., Vehkamaki, H., Kirkby, J., 2013. Molecular understanding of sulphuric acid-amine particle nucleation in the atmosphere. Nature 502, 359-363. Bader, R.F.W., 1990. Atoms in Molecules: A Quantum Theory. Oxford University Press, Oxford, UK. Barnes, I., Hjorth, J., Mihalopoulos, N., 2006. Dimethyl Sulfide and Dimethyl Sulfoxide and Their Oxidation in the Atmosphere. Chem. Rev. 106, 940-975. Berresheim, H., 2002. Gas-aerosol relationships of H2SO4, MSA, and OH: Observations in the coastal marine boundary layer at Mace Head, Ireland. J. Geophys. Res. 107, 7101-7116. Bork, N., Elm, J., Olenius, T., Vehkamäki, H., 2014. Methane sulfonic acid-enhanced formation of molecular clusters of sulfuric acid and dimethyl amine. Atmos. Chem. Phys. 14, 12023-12030. Bzdek, B.R., Ridge, D.P., Johnston, M.V., 2011. Reactivity of methanesulfonic acid salt clusters relevant to marine air. J. Geophys. Res. 116, D03301. Chen, H., Finlayson-Pitts, B.J., 2017. New Particle Formation from Methanesulfonic Acid and Amines/Ammonia as a Function of Temperature. Environ. Sci. Technol. 51, 243-252. Chen, H., Varner, M.E., Gerber, R.B., Finlayson-Pitts, B.J., 2016. Reactions of Methanesulfonic Acid with Amines and Ammonia as a Source of New Particles in Air. J. Phys. Chem. B 120, 1526-1536. Curtiss, L.A., Blander, M., 1988. Thermodynamic properties of gas-phase hydrogen-bonded complexes. Chem. Rev. 88, 827-841. Dall'Osto, M., Ceburnis, D., Monahan, C., Worsnop, D.R., Bialek, J., Kulmala, M., Kurtén, T., Ehn, M., Wenger, J., Sodeau, J., Healy, R., O'Dowd, C., 2012. Nitrogenated and aliphatic organic vapors as possible drivers for marine secondary organic aerosol growth. J. Geophys. Res-Atmos. 117, D12311. Dameto de España, C., Wonaschütz, A., Steiner, G., Rosati, B., Demattio, A., Schuh, H., Hitzenberger, R., 2017. 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. 164, 289-298. Dawson, M.L., Varner, M.E., Perraud, V., Ezell, M.J., Gerber, R.B., Finlayson-Pitts, B.J., 2012. Simplified mechanism for new particle formation from methanesulfonic acid, amines, and water via experiments and ab initio calculations. Proc. Natl. Acad. Sci. U. S. A. 109, 18719-18724. DePalma, J.W., Doren, D.J., Johnston, M.V., 2014. Formation and growth of molecular clusters containing sulfuric acid, water, ammonia, and dimethylamine. J. Phys. Chem. A 118, 5464-5473. Elm, J., 2017. Elucidating the Limiting Steps in Sulfuric Acid–Base New Particle Formation. The J. Phys. Chem. A 121, 8288-8295. Elm, J., Bilde, M., Mikkelsen, K.V., 2013a. Assessment of binding energies of atmospherically relevant clusters. Phys. Chem. Chem. Phys. 15, 16442-16445. Elm, J., Fard, M., Bilde, M., Mikkelsen, K.V., 2013b. Interaction of Glycine with Common 19
580 581 582 583 584 585 586 587 588 589 590 591 592 593 594 595 596 597 598 599 600 601 602 603 604 605 606 607 608 609 610 611 612 613 614 615 616 617 618 619 620 621 622 623
Atmospheric Nucleation Precursors. J. Phys. Chem. A 117, 12990-12997. Elm, J., Jen, C.N., Kurten, T., Vehkamaki, H., 2016a. Strong Hydrogen Bonded Molecular Interactions between Atmospheric Diamines and Sulfuric Acid. J. Phys. Chem. A 120, 3693-3700. Elm, J., Kristensen, K., 2017. Basis set convergence of the binding energies of strongly hydrogen-bonded atmospheric clusters. Phys. Chem. Chem. Phys. 19, 1122-1133. Elm, J., Myllys, N., Hyttinen, N., Kurtén, T., 2015. Computational Study of the Clustering of a Cyclohexene Autoxidation Product C6H8O7 with Itself and Sulfuric Acid. J. Phys. Chem. A 119, 8414-8421. Elm, J., Myllys, N., Kurtén, T., 2016b. Phosphoric acid – a potentially elusive participant in atmospheric new particle formation. Mol. Phys. 115, 2168-2179. Frisch, M.J., Trucks, G.W., Schlegel, H. B., Scuseria, G. E., Robb,M.A., Cheeseman,, J. R., S., G., Barone, V., Mennucci, B., Petersson, G. A., Nakatsuji, H., Caricato, M., Li, X., Hratchian, H., P., I., 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., O., F., Bearpark, M., Heyd, J. J., Brothers, E., Kudin, K. N., Staroverov, V. N., Kobayashi, R., Normand,, J., R., K., Rendell, A., Burant, J. C. Iyengar, S. S. Tomasi, J. Cossi, M. Rega, Millam, N. J., Klene, M., Knox, J.E., Cross, J. B., Bakken, V., Adamo, C., Jaramillo, J., Gomperts, R. E. Stratmann, O. Yazyev,A. J.Austin,, R. Cammi, C.P., J.W. Ochterski, R. Martin, R. L., Morokuma, K., Zakrzewski, V. G., Voth, G. A., Salvador,, P., D., J. J., Dapprich, S., Daniels,A. D., Farkas, O., Foresman, J. B., Ortiz, J. V., Cioslowski, J., and Fox,, J., D., 2009. Gaussian 09, Revision C.1. Gaussian, Inc., Wallingford CT. Gaston, C.J., Pratt, K.A., Qin, X., Prather, K.A., 2010. Real-Time Detection and Mixing State of Methanesulfonate in Single Particles at an Inland Urban Location during a Phytoplankton Bloom. Environ. Sci. Technol. 44, 1566-1572. Ge, X., Wexler, A.S., Clegg, S.L., 2011. Atmospheric amines – Part I. A review. Atmos. Environ. 45, 524-546. Glasoe, W.A., Volz, K., Panta, B., Freshour, N., Bachman, R., Hanson, D.R., McMurry, P.H., Jen, C., 2015. Sulfuric acid nucleation: An experimental study of the effect of seven bases. J. Geophys. Res-Atmos. 120, 1933-1950. Henschel, H., Kurten, T., Vehkamaki, H., 2016. Computational Study on the Effect of Hydration on New Particle Formation in the Sulfuric Acid/Ammonia and Sulfuric Acid/Dimethylamine Systems. J. Phys. Chem. A 120, 1886-1896. Jiang, B., Xia, D., 2017. Role identification of NH 3 in atmospheric secondary new particle formation in haze occurrence of China. Atmos. Environ. 163, 107-117. Johnson, E.R., Keinan, S., Mori-Sánchez, P., Contreras-García, J., Cohen, A.J., Yang, W., 2010. Revealing Noncovalent Interactions. J. Am. Chem. Soc. 132, 6498-6506. Karl, M., Gross, A., Leck, C., Pirjola, L., 2007. Intercomparison of dimethylsulfide oxidation mechanisms for the marine boundary layer: Gaseous and particulate sulfur constituents. J. Geophys. Res-Atmos. 112, D15304. Kerminen, V.-M., Aurela, M., Hillamo, R.E., Virkkula, A., 1997. Formation of particulate MSA: deductions from size distribution measurements in the Finnish Arctic. Tellus B 49, 159-171. Kulmala, M., Kontkanen, J., Junninen, H., Lehtipalo, K., Manninen, H.E., Nieminen, T., Petaja, T., Sipila, M., Schobesberger, S., Rantala, P., Franchin, A., Jokinen, T., Jarvinen, E., Aijala, M., Kangasluoma, J., Hakala, J., Aalto, P.P., Paasonen, P., Mikkila, J., Vanhanen, J., Aalto, J., Hakola, 20
624 625 626 627 628 629 630 631 632 633 634 635 636 637 638 639 640 641 642 643 644 645 646 647 648 649 650 651 652 653 654 655 656 657 658 659 660 661 662 663 664 665 666 667
H., Makkonen, U., Ruuskanen, T., Mauldin, R.L., 3rd, Duplissy, J., Vehkamaki, H., Back, J., Kortelainen, A., Riipinen, I., Kurten, T., Johnston, M.V., Smith, J.N., Ehn, M., Mentel, T.F., Lehtinen, K.E., Laaksonen, A., Kerminen, V.M., Worsnop, D.R., 2013. Direct observations of atmospheric aerosol nucleation. Science 339, 943-946. Kulmala, M., Pirjola, L., Makela, J.M., 2000. Stable sulphate clusters as a source of new atmospheric particles. Nature 404, 66-69. Kumar, M., Francisco, J.S., 2017. Ion pair particles at the air-water interface. Proc. Natl. Acad. Sci. U. S. A. 114, 12401-12406. Kumar, M., Li, H., Zhang, X., Zeng, X.C., Francisco, J.S., 2018. Nitric Acid-Amine Chemistry in the Gas Phase and at the Air-Water Interface. J. Ame. Chem. Soc. 140, 6456-6466. Kurtén, T., Loukonen, V., Vehkamäki, H., Kulmala, M., 2008. Amines are likely to enhance neutral and ion-induced sulfuric acid-water nucleation in the atmosphere more effectively than ammonia. Atmos. Chem. Phys. 8, 4095-4103. Kurtén, T., Sundberg, M.R., Vehkamäki, H., Noppel, M., Blomqvist, J., Kulmala, M., 2006. Ab Initio and Density Functional Theory Reinvestigation of Gas-Phase Sulfuric Acid Monohydrate and Ammonium Hydrogen Sulfate. J. Phys. Chem. A 110, 7178-7188. Li, H., Kupiainen-Määttä, O., Zhang, H., Zhang, X., Ge, M., 2017. A molecular-scale study on the role of lactic acid in new particle formation: Influence of relative humidity and temperature. Atmos. Environ. 166, 479-487. Li, H., Zhang, X., Zhong, J., Liu, L., Zhang, H., Chen, F., Li, Z., Li, Q., Ge, M., 2018. The role of hydroxymethanesulfonic acid in the initial stage of new particle formation. Atmos. Environ. 189, 244-251. Li, S., Zhang, L., Qin, W., Tao, F.-M., 2007. Intermolecular structure and properties of the methanesulfonic acid–ammonia system in small water clusters. Chem. Phys. Lett. 447, 33-38. Ling, J., Ding, X., Li, Z., Yang, J., 2017. First-Principles Study of Molecular Clusters Formed by Nitric Acid and Ammonia. J. Phys. Chem. A 121, 661-668. Loukonen, V., Kurtén, T., Ortega, I.K., Vehkamäki, H., Pádua, A.A.H., Sellegri, K., Kulmala, M., 2010. Enhancing effect of dimethylamine in sulfuric acid nucleation in the presence of water - a computational study. Atmos. Chem. Phys. 10, 4961-4974. Lu, T., Chen, F., 2012. Multiwfn: A Multifunctional Wavefunction Analyzer. J. Comput. Chem. 33, 580-592. McGrath, M.J., Olenius, T., Ortega, I.K., Loukonen, V., Paasonen, P., Kurtén, T., Kulmala, M., Vehkamäki, H., 2012. Atmospheric Cluster Dynamics Code: a flexible method for solution of the birth-death equations. Atmos. Chem. Phys. 12, 2345-2355. Myllys, N., Elm, J., Halonen, R., Kurten, T., Vehkamaki, H., 2016. Coupled Cluster Evaluation of the Stability of Atmospheric Acid-Base Clusters with up to 10 Molecules. J. Phys. Chem. A 120, 621-630. Neese, F., 2012. The ORCA program system. Wires. Comput. Mol. Sci. 2, 73-78. Olenius, T., Halonen, R., Kurtén, T., Henschel, H., Kupiainen-Määttä, O., Ortega, I.K., Jen, C.N., Vehkamäki, H., Riipinen, I., 2017. New particle formation from sulfuric acid and amines: Comparison of monomethylamine, dimethylamine, and trimethylamine. J. Geophys. Res-Atmos. 122, 7103-7118. Olenius, T., Kupiainen-Maatta, O., Ortega, I.K., Kurten, T., Vehkamaki, H., 2013. Free energy barrier in the growth of sulfuric acid-ammonia and sulfuric acid-dimethylamine clusters. J. Chem. Phys. 21
668 669 670 671 672 673 674 675 676 677 678 679 680 681 682 683 684 685 686 687 688 689 690 691 692 693 694 695 696 697 698 699 700 701 702 703 704 705 706 707 708 709 710 711
139, 084312-084324. Ortega, I.K., Kupiainen, O., Kurten, T., Olenius, T., Wilkman, O., McGrath, M.J., Loukonen, V., Vehkamaki, H., 2012. From quantum chemical formation free energies to evaporation rates. Atmos. Chem. Phys. 12, 225-235. Pope, C.A., Dockery, D.W., 2012. Health Effects of Fine Particulate Air Pollution: Lines that Connect. J. Air Waste Manage. 56, 709-742. Riplinger, C., Neese, F., 2013. An efficient and near linear scaling pair natural orbital based local coupled cluster method. J. Chem. Phys. 138, 034106. Riplinger, C., Sandhoefer, B., Hansen, A., Neese, F., 2013. Natural triple excitations in local coupled cluster calculations with pair natural orbitals. J. Chem. Phys. 139, 134101. Seinfeld, J.H., Pandis, S.N., 2006. Atmospheric Chemistry and Physics From Air Pollution to Climate Change. Wiley Interscience, New York. Shampine, L.F., Reichelt, M.W., 1997. The MATLAB ODE Suite. SIAM J. Sci.Comput. 18, 1-22. Sorooshian, A., Padró, L.T., Nenes, A., Feingold, G., McComiskey, A., Hersey, S.P., Gates, H., Jonsson, H.H., Miller, S.D., Stephens, G.L., Flagan, R.C., Seinfeld, J.H., 2009. On the link between ocean biota emissions, aerosol, and maritime clouds: Airborne, ground, and satellite measurements off the coast of California. Global Biogeochem.Cy. 23, GB4007(15). Stewart, J.J.P., 1990. MOPAC: A semiempirical molecular orbital program. J. Comput. Aid. Mol. Des. 4, 1-105. Stocker, T.F., Qin, D., Plattner, G.-K., Tignor, M., Allen, S.K., Boschung, J., Nauels, A., Xia, Y., Bex, V., Midgley, P.M., Eds., 2013. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change, IPCC, 2013. Climate Change 2013: The Physical Science Basis;. Cambridge University Press, Cambridge. Wang, C.-Y., Jiang, S., Wang, Z.-Q., Liu, Y.-R., Wen, H., Huang, T., Han, Y.-J., Huang, W., 2019. Can formaldehyde contribute to atmospheric new particle formation from sulfuric acid and water? Atmos. Environ. 201, 323-333. Weber, R.J., Chen, G., Davis, D.D., Mauldin, R.L., Tanner, D.J., Eisele, F.L., Clarke, A.D., Thornton, D.C., Bandy, A.R., 2001. Measurements of enhanced H2SO4 and 3–4 nm particles near a frontal cloud during the First Aerosol Characterization Experiment (ACE 1). J. Geophys. Res-Atmos. 106, 24107-24117. Wen, H., Wang, C.-Y., Wang, Z.-Q., Hou, X.-F., Han, Y.-J., Liu, Y.-R., Jiang, S., Huang, T., Huang, W., 2019. Formation of atmospheric molecular clusters consisting of methanesulfonic acid and sulfuric acid: Insights from flow tube experiments and cluster dynamics simulations. Atmos. Environ. 199, 380-390. Willis, M.D., Burkart, J., Thomas, J.L., Kollner, F., Schneider, J., Bozem, H., Hoor, P.M., Aliabadi, A.A., Schulz, H., Herber, A.B., Leaitch, W.R., Abbatt, J.P.D., 2016. Growth of nucleation mode particles in the summertime Arctic: a case study. Atmos. Chem. Phys. 16, 7663-7679. Xie, H.B., Elm, J., Halonen, R., Myllys, N., Kurten, T., Kulmala, M., Vehkamaki, H., 2017. Atmospheric Fate of Monoethanolamine: Enhancing New Particle Formation of Sulfuric Acid as an Important Removal Process. Environ. Sci. Technol. 51, 8422-8431. Yin, D., MacKerell, A.D., 1998. Combined ab initio/empirical approach for optimization of Lennard– Jones parameters. J. Comput. Chem. 19, 334-348. Zhang, J., Dolg, M., 2015. ABCluster: the artificial bee colony algorithm for cluster global optimization. Phys. Chem. Chem. Phys. 17, 24173-24181. 22
712 713 714 715 716 717 718
Zhang, J., Dolg, M., 2016. Global Optimization of Clusters of Rigid Molecules Using the Artificial Bee Colony Algorithm. Phys. Chem. Chem. Phys. 18, 3003-3010. Zhang, R., Khalizov, A., Wang, L., Hu, M., Xu, W., 2012. Nucleation and growth of nanoparticles in the atmosphere. Chem. Rev. 112, 1957-2011. Zollner, J.H., Glasoe, W.A., Panta, B., Carlson, K.K., McMurry, P.H., Hanson, D.R., 2012. Sulfuric acid nucleation: power dependencies, variation with relative humidity, and effect of bases. Atmos. Chem. Phys. 12, 4399-4411.
23
a
b
c
Fig. 1 NCI (lower) and RDG (upper) analyses among global minima for (a) (NH3)2, (b) (MSA)2 and (c) (MSA)(NH3) clusters. Red, yellow, blue, gray and light gray spheres represent oxygen, sulfur, nitrogen, carbon and hydrogen atoms, respectively.
a
b Fig. 2 ∆Gref (a) and total evaporation rate (b) for (MSA)x(NH3)y clusters (x, y ≤ 4) by the DLPNO-CCSD(T)/aug-cc-pVTZ//M06-2X/6-31++G(d,p) level on the MSA-NH3 grid at 298.15 K and 278.15 K, respectively. 7
10
[NH3] = 10 pptv
-3 [(MSA)2] (molecules⋅ cm )
[NH3] = 100 pptv [NH3] = 1000 pptv
4
10
-4
2.4x10
-4
1
2.2x10
10
-4
2x10
-4
1.8x10 -2
10
-4
1.6x10
4
10 4
10
5
10
6
4
4
7
10 10 -3 [MSA] (molecules⋅ cm ) a
4
4
4
10 .08x10 .12x10 .16x10 1.2x10 1.04x 1 1 1 8
10
9
10
2
10
-3
10
[NH3] = 10 pptv [NH3] = 100 pptv
-8
[NH3] = 1000 pptv
-3
-1
J (cm ⋅ s )
10
-13
10
-18
10
-23
10
-28
10
4
10
5
10
6
7
10 10 -3 [MSA] (molecules⋅ cm )
8
10
b Fig. 3 Concentration of (MSA)2 dimer (molecules·cm−3) (a) and the formation rate J (cm−3·s−1) of the studied system (b) as a function of MSA monomer concentration at 278.15 K.
Fig. 4 ∆Gact for (MSA)x(NH3)y (x, y ≤ 4) clusters at 278.15 K. ([MSA] = 106 molecules·cm−3 and [NH3] = 100 pptv).
a
b Fig. 5 Main clustering pathways and flux out for (a) (MSA)x(NH3)y (x, y ≤ 4) at RH = 0% and (b) (MSA)x(NH3)y(W)z (x, y ≤ 2, z ≤ 4) at RH = 80% where [MSA] = 106 molecules·cm−3, [NH3] = 102 pptv, and T = 278.15 K. Pathways contributing less than 5% to the flux of the cluster are not shown for clarity.
2
Relative Evaporation Rate
10
1
10
0
10
-1
10
-2
10
(MSA)2
(MSA)(NH3)2
(NH3)2
(MSA)2(NH3)
(MSA)(NH3)
(MSA)2(NH3)2
-3
10
-4
10
-5
10
0
20
40
60
80
100
RH %
a 4
8x10
Jrel
4
Relative Formation Rate (Jrel)
7x10
4
6x10
4
5x10
4
4x10
4
3x10
4
2x10
4
1x10
0 0
20
40 RH %
60
80
100
b Fig. 6 Relative evaporation rate (a) and relative formation rate (b) of (MSA)x(NH3)y (x, y ≤ 2) clusters as a function of RH where [MSA] = 106 molecules·cm−3, [NH3] = 100 pptv and T = 278.15 K.
Highlight: 1.
Hydrogen bonding is the primary driving force that forms the MSA-NH3 clusters.
2.
NH3 effectively promotes the formation of MSA-based clusters at ppt levels.
3.
Formation of (MSA)2 is a rate-determining step under anhydrous condition.
4.
Formation of (MSA)(NH3) is a rate-determining step under hydrous condition.
5.
The formation rate increases with RH, reaching up to a factor of 105 at RH = 100%.
Dongping Chen, Fengyi Liu and Wenliang Wang analyzed the results and wrote the manuscript. Dongping Chen prepared Figs. 2-6 and Tables S1–S6. Changwei Wang prepared Figs. 1, S1, S2, S9, and S10. Danfeng Li prepared Figs. S3-S8 and S11. All authors contributed to the manuscript.
The authors declared that they have no conflicts of interest to this work.