Abstract

Aerosols play an important role in climate and air quality; however, the mechanisms behind aerosol particle formation in the atmosphere are poorly understood. Studies have identified sulfuric acid, water, oxidized organics, and ammonia/amines as key precursors for forming aerosol particles in the atmosphere. Theoretical and experimental investigations have indicated that other species, such as organic acids, may be involved in atmospheric nucleation and growth of freshly formed aerosol particles. Organic acids, such as dicarboxylic acids, which are abundant in the atmosphere, have been measured in ultrafine aerosol particles. These observations suggest that organic acids may contribute to new particle formation in the atmosphere but their role remains ambiguous. This study examines how malonic acid interacts with sulfuric acid and dimethylamine to form new particles at warm boundary layer conditions using experimental observations from a laminar flow reactor and quantum chemical calculations coupled with cluster dynamics simulations. Observations reveal that malonic acid does not contribute to the initial steps (formation of <1 nm diameter particle) of nucleation with sulfuric acid-dimethylamine. In addition, malonic acid was found to not participate in the subsequent growth of the freshly nucleated 1 nm particles from sulfuric acid-dimethylamine reactions to diameters of 2 nm.
Introduction
Atmospheric new particle formation (NPF), which comprises nucleation and subsequent growth of particles, contributes significantly to the formation of cloud condensation nuclei (CCN), thus affecting the Earth’s radiative budget.1,2 Studies have shown that nucleation, the process where trace amounts of precursor vapors react or cluster to form stable particles, produces approximately 50% of the global CCN.3−7 Understanding nucleation and initial particle growth is crucial to better predict the impact of aerosol particles on clouds and the climate. Sulfuric acid nucleation with ammonia and amines in the atmosphere has been widely established by several studies as important pathways for nucleation in the polluted boundary layer.8−13 However, the sulfuric acid concentrations in the atmosphere are generally too low to enable the growth of clusters to CCN sizes, leaving atmospherically abundant organic acids as plausible agents for the growth of nucleated clusters.
Organic dicarboxylic acids are highly oxidized molecules that have been measured in ultrafine aerosol particles (diameters < 100 nm) around the world.14−22 As a result, these molecules may contribute to the observed nucleation and growth rates in the atmosphere. Dicarboxylic acids in the atmosphere come from various biogenic and anthropogenic sources.23,24 The primary sources of dicarboxylic acid vapors in the atmosphere include animal waste, chemical plants, wildfires, tobacco smoke, and others.25−28 Atmospheric dicarboxylic acids are also generated via secondary pathways such as photo-oxidation of volatile organic precursors.29,30 Though dicarboxylic acids have been measured in ultrafine particles, their role in NPF from gas-phase reactions in the atmosphere is still not well known.
Computational chemistry studies predict that dicarboxylic acids can offer some stabilization to sulfuric acid clusters.31−34 Dicarboxylic acids are very polar and contain sites for hydrogen bonding with nucleation precursors such as sulfuric acid and amines. In addition, small dicarboxylic acids are intermediate volatility compounds, which make them potential candidates for NPF in the atmosphere.35,36 On a molecular level, the mechanisms through which small dicarboxylic acids, such as malonic acid, participate in the initial steps of nucleation and growth of clusters from sulfuric acid-amine reactions have not been investigated extensively. Sulfuric acid and ammonia/amine reactions are responsible for nucleation events observed in various parts of the world.8,11−13,37−48 Computational studies of cluster formation involving sulfuric acid, ammonia, and malonic acid have previously been performed by Zhang et al.34 Extremely cold temperatures (218 K) were required for malonic acid to enhance NPF in any appreciable amount. The computational study of sulfuric acid, dimethylamine, and malonic acid clusters by Wang et al. found that malonic acid could potentially participate in sulfuric acid-dimethylamine nucleation.33 The nucleation rate of the sulfuric acid-dimethylamine-malonic acid system was found to be in between that of the sulfuric acid-dimethylamine-water and sulfuric acid-ammonia-water systems at 278 K. However, only 40 low-energy structures based on the PM7 semi-empirical level of theory were considered, which makes it possible that the global minimum structure was not observed. In addition, Wang et al. applied the RI-MP2/cc-pVTZ level of theory to calculate the binding energies of their clusters. It is well known that MP2 overestimates the binding energies of atmospheric molecular clusters49 and thereby their calculated cluster stabilities are expected to be overestimated.
Experimental evidence of the role of small dicarboxylic acids is still lacking. Specifically, malonic acid is abundant in the atmosphere and has a low vapor pressure (1.2 × 10–3 Pa at 298 K).50 Previous measurements from rural Lamont, OK show malonic acid at 107–109 cm–3, higher than sulfuric acid concentrations of 105–107 cm–3.51 Fang et al. measured 107–109 cm–3 of malonic acid at Pingyuan rural site and observed a correlation of nucleation rates with concentrations of diacids.52 The authors also observed clusters of small diacids, including malonic acid, and concluded that malonic acid has the potential to enhance the initial steps of sulfuric acid-dimethylamine particle formation and growth.
This study presents experiments performed in a laminar flow reactor to investigate the role of malonic acid in the initial steps of nucleation and growth of freshly formed clusters in the presence and absence of sulfuric acid and stabilizing bases such as dimethylamine. A chemical ionization mass spectrometer (CIMS) and a versatile water condensation particle counter (vWCPC) were used to examine molecular clusters and particle number concentrations, respectively. Complimentary quantum chemical calculations were also conducted on clusters consisting of sulfuric acid, dimethylamine, and malonic acid. Based on the calculated thermochemistry, the cluster population dynamics simulations are performed and compared to the experimental results.
Experimental Section
Malonic acid (MaA) and dimethylamine (DMA) were reacted with gaseous sulfuric acid to produce freshly formed clusters in a clean, laminar flow reactor. This glass flow reactor has been previously described in Fomete et al., with key details repeated here.53 Malonic acid vapor, humified nitrogen, and nitrogen carrier gas are injected into the top of the reactor (Figure 1). 2.0 sLpm of nitrogen carrier gas is supplied to the reactor from an ultra-high purity liquid nitrogen tank (99.999% purity, Matheson). Malonic acid at 1.0 sLpm injected into the reactor is produced by flowing nitrogen gas over a reservoir containing 15 wt % malonic acid in HPLC-grade water. 1.5 sLpm of humidified nitrogen is also injected into the reactor by flowing nitrogen over a reservoir containing HPLC-grade water. The flow reactor was operated at 1 atm, 20% RH, 300–303 K, and a total flow rate of 4.5–4.6 sLpm.
Figure 1.
Schematic of the malonic acid flow tube reactor. The reactor is set up in a vertical position during experiments.
Sulfuric acid was injected into the mixing region of the reactor, shown in Figure 1. Sulfuric acid concentrations in this study were maintained at a factor of ∼10 lower than malonic acid to mimic what has previously been observed in the field.51,52 Dimethylamine (DMA) was injected into the reactor’s nucleation region, resulting in a nucleation reaction time of ∼10 s. This time was determined using the center line flow velocity (7.3 m/s) calculated via computational fluid dynamics.54 Gas-phase DMA was produced by passing nitrogen over a temperature-controlled (300 K) permeation tube containing DMA (40 wt % solution in water, Fischer Scientific). The gas-phase concentration of DMA was varied (0–33 ptv) by the use of a double-stage dilution system.53 The range of DMA concentrations studied here also follows those previously observed in the atmosphere.55−57 DMA was injected into the centerline of the flow in the reactor to minimize vapor wall losses.53
The flow reactor was continuously purged with nitrogen, water, and malonic acid vapors when experiments were not being conducted to minimize contamination that may interfere with the targeted nucleation reactions.53 Baseline conditions were conducted each day to ensure clean and repeatable reaction conditions in the flow reactor before an experiment.9 These baseline measurements, shown in Figure S1a,b (SI-1) in the supplementary information (SI), involve measuring the malonic acid dimer concentrations ([MaA2]) at a given malonic acid monomer concentration ([MaA1] at 103 amu) and the sulfuric acid dimer concentrations ([SA2] at 195 amu) at a given sulfuric acid monomer concentration ([SA1] at 97 amu) without the addition of DMA. In addition, baseline 1-nm particle concentrations at various [MaA1] were taken and are shown in Figure S2 in SI-2. SAi and MaAi represent clusters containing i sulfuric acid or malonic acid molecules and were detected without any other base or water ligands. The heterodimer of malonic acid and sulfuric acid, which was detected without any attached ligands, is represented as MaA1·SA1. Other ligand molecules were likely attached to these clusters but evaporated upon ionization and/or measurement with the mass spectrometer.44
A previously described custom-built transverse chemical ionization inlet, coupled to an atmospheric pressure, long time-of-flight mass spectrometer (Tofwerk AG), called the Pittsburgh Cluster CIMS (PCC), was attached in-line with the malonic acid flow reactor to measure concentrations of gases and freshly formed clusters.58 Malonic and sulfuric acid concentrations, as well as the neutral clusters formed in the flow reactor, were measured in negative ion mode with the PCC using acetate as the chemical ionization reagent ion. Acetate ion, compared to nitrate, was chosen to ensure efficient ionization of malonic acid. The dominant acetate reagent ions are H2O · CH3CO2– (∼6 × 103 Hz), CH3CO2H · CH3CO2 (∼1 × 105 Hz), and CH3CO2– (∼6 × 104 Hz). The concentration of DMA in the sample flow was measured in positive ion mode using hydronium ions, (H2O)1 – 2 · H3O+. See SI-3 for an explanation on how the acid and amine concentrations were calculated. The important flows and voltage parameters used in the custom-built inlet of the PCC were chosen in order to result in a 25 ms chemical ionization reaction time.53 The systematic uncertainty of the PCC is estimated to be a factor of 2.58 In addition, a versatile water condensation particle counter (vWCPC, TSI 3789) was connected through one of the side ports in the nucleation region of the reactor to detect the total particle number concentration of freshly nucleated particles.59 The vWCPC operated at 1.5 sLpm total flow to minimize diffusional losses within the sampling line. The operating temperatures of the vWCPC were set to 1.5, 90, and 22 °C for the conditioner, initiator, and moderator stages, respectively, to measure particles down to ∼1 nm (50% cut-off size, d50). For the d50 of 2 nm, the temperatures were 10, 90, and 22 °C for the conditioner, initiator, and moderator stages respectively.
Computational Details
A subset of the relevant SA0–3DMA0–3MaA0–1 clusters was chosen as a test case to simulate cluster formation of the SA–DMA–MaA system (see SI-4 for the list of clusters). Atmospheric Cluster Dynamics Code (ACDC),60 representing the cluster birth-death equations, was used to determine concentrations of various formed clusters. ACDC incorporates all possible collision and evaporation patterns for all monomers and clusters. At each given initial monomer concentration, a 10 s simulation was performed at standard condition (298.15 K and 1 atm) to mimic the flow tube experiment. Wall losses for flow tube experiments with a tube radius of 2.5 cm were also included (see ACDC for more details). The clusters that were allowed to grow out of the simulation system were set to be SAi≥3DMAj≥3MaAk≥0. These clusters are assumed to be stable enough to grow further into particles. The formation rate of outgrowing clusters thus defines the particle formation rate. Clusters that are not within our simulation scheme and within the outgrowing clusters are assumed to be unstable and fragment back into the simulation system. The simulation scheme is quite small. However, the calculated particle formation rate will have only a small offset but the studied trends should be correct. In other words, the simulated system is large enough to support the experimental data and reveal the stability of clusters containing malonic acid, as well as its effect on the atmospheric NPF of SA–DMA system.
The collision probabilities of each cluster type are calculated from kinetic gas theory.60 The evaporation probabilities are calculated from the balance equation and cluster binding free energies obtained from quantum chemical calculations.60 The structures and energies of the SA–DMA clusters were taken from Kubečka et al.61 The remaining clusters containing one MaA were constructed through the same configurational sampling protocol, and the binding free energies were evaluated at the same level of theory (i.e., DLPNO–CCSD(T)62−66/aug-cc-pVTZ//ωB97X-D67/6-31++G(d,p)) and using the same computational software (i.e., ABCluster,68,69 XTB,70,71 Gaussian,72 and ORCA.73,74 The DLPNO–CCSD(T)/aug-cc-pVTZ//ωB97X-D/6-31++G(d,p) level of theory has been thoroughly benchmarked both with regards to the structures75,76 and binding energies.49,77 This methodology is generally recommended for atmospheric cluster formation studies by our group78 and others.79 More technical details are given in Kubečka et al.61 The identified structure coordinates and binding free energies of all clusters are presented in SI-4.
Results and Discussion
Figure 2 presents the measured dimer concentrations of malonic acid ([MaA2]), the heterodimer of malonic acid and sulfuric acid ([MaA1·SA1]), and the sulfuric acid dimer concentration ([SA2]) as a function of malonic acid concentration ([MaA1]). Each color represents a different [DMA1] ranging from 0 to 33 pptv, and [SA1] was held constant at 3 × 108 cm–3. For Figure 2a, most malonic acid dimers were detected at 207 m/z (C3H4O4·C3H3O4–), with small amounts (<5%) detected with acetate ligands at 267 and 327 m/z. It can be seen in Figure 2a that increasing the [DMA1] from 0 to 33 pptv did not affect the measured [MaA2]. [MaA2] displays a squared dependency on [MaA1], suggesting that its formation depends primarily on the collision of two MaA1 with no involvement of DMA. This implies that amines, such as DMA do not enhance the formation of MaA2 and ternary nucleation of MaA-DMA-H2O in the absence of sulfuric acid is very unlikely at 300 K. This is corroborated by quantum chemical calculations as the binding free energy of malonic acid interacting with DMA (−4.0 kcal/mol) is relatively high compared to SA1·DMA1 (−11.4 kcal/mol). This implies that the [MaA1·DMA1] cluster will have a very low gas-phase concentration.
Figure 2.

(a) Measured malonic acid dimer concentration ([MaA2]) vs malonic acid monomer concentration ([MaA1]), (b) malonic acid and sulfuric acid heterodimer ([MaA1·SA1]) vs [MaA1], and (c) sulfuric acid dimer concentration ([SA2]) vs [MaA1]. Each color represents a different [DMA1] between 0 and 33 pptv. The sulfuric acid concentration ([SA1]) was constant at 3 × 108 cm–3.
Figure 2b shows the concentration of the heterodimer of malonic acid and sulfuric acid [MaA1·SA1] vs [MaA1] for [DMA1] = 0–33 pptv and [SA1] = 3 × 108 cm–3. The [MaA1·SA1] increases with increasing [MaA1] when [DMA1] is kept constant. In addition, [DMA1] from 0 to 2 pptv, the [MaA1·SA1] increases due to increased [MaA1] as evidenced by a continuation of the [DMA1] = 0 pptv curve. The computed free energies and the results of Wang et al.33 and our theoretical calculations show that hydrogen bonding is possible between sulfuric and malonic acid in this acid heterodimer which explains why MaA1·SA1 was experimentally measured (see Figure S3 in SI-4 for the heterodimer structure). As [MaA1] is increased in the reactor, more malonic acid is available to collide and cluster with sulfuric acid to form MaA1·SA1. However, when [DMA1] is increased in the reactor, the formation of MaA1·SA1 is reduced. For example, at [MaA1] = 3 × 109 cm–3, [MaA1·SA1] decreases from 6.2 × 107 to 1.8 × 107 cm–3 when [DMA1] is increased from 2 and 33 pptv. This suggests that the presence of DMA hinders the formation of MaA1·SA1 in the reactor. Although the theoretical simulations resulted in lower [MaA1·SA1] (see Figure S4a), the simulations reproduce similar DMA hindering effects. This hindrance can be understood from the free energies of these formed clusters. The reaction free energy for forming MaA1·SA1 is −5.6 kcal/mol, whereas the formation of SA1·DMA1 is −11.4 kcal/mol. This implies that the [SA1·DMA1] will build up over time, with DMA scavenging the free sulfuric acid available to form MaA1·SA1.9,80 Cluster growth of MaA1·SA1 to larger sizes with and without DMA is not favorable as no larger malonic acid clusters (MaAi≥2) containing either sulfuric acid and/or DMA were measured.
Figure 2c shows [SA2] vs [MaA1], where [DMA1] = 0–33 pptv and [SA1] = 3 × 108 cm–3. As previously shown, DMA enhances sulfuric acid dimer formation, which is seen by increased sulfuric acid dimer concentrations with increasing [DMA1].8,9,55,80−83 Notice that the concentration of sulfuric acid dimer is slightly lower at [DMA1] = 33 pptv than at 16 pptv. This is likely due to the loss of more sulfuric acid dimers to sulfuric acid dimer–dimer coagulation when [DMA1] is very high.9 Generally, increasing malonic acid monomer concentrations in Figure 2c leads to a slight increase in [SA2]. Due to the narrow range of [MaA1] explored in Figure 2, more measurements are needed at higher and lower [MaA1] to better examine its effect on the aforementioned cluster concentrations. Figure S4b shows that the simulated [SA2] does not change with [MaA1]. In addition, [SA2] decreases with [DMA1] because SA1DMA1 clusters are more stable than SA2 and DMA also hinders the SA2 formation. Figure S4c shows the concentration of all simulated clusters containing two SA, i.e., [N2] = [SA2DMA0–3MaA0–1] at different [DMA1]. Simulated [N2] reproduces well the trends observed in Figure 2c which shows the [SA2]. Note, measured [SA2] includes various ligands (e.g., SA2·(H2O)x, SA2·MaA1, and SA2·(DMA)x) which then fragment within the PCC due to the ionization; leaving the pure SA cluster ion to be detected.
Compared to our experimental results, Wang et al.33 reported the formation of a stable MaA1·SA1·DMA1 cluster from computational chemistry calculations, which predicts that malonic acid can form hydrogen bonds with sulfuric acid and a stable cyclic ring structure when DMA is added. However, MaA1·SA1·DMA1 was not observed experimentally, maybe due to the short lifetime of its ion form. Also, large MaA–SA clusters or MaA–DMA clusters were not measured. Therefore, it is not likely that MaA1·SA1·DMA1 contributes to the initial stages of nucleation as it, and even the large clusters containing more MaA, SA, and/or DMA, were not measured.
Based on our quantum chemical calculations and in agreement with the reported evaporation rates by Wang et al., the SA1·DMA1 cluster is more stable (ΔG = −11.4 kcal/mol) than MaA1·DMA1 cluster (ΔG = −4.0 kcal/mol).33 In MaA1·DMA1, no proton transfer occurs between the malonic acid and DMA, and thereby the cluster is held together purely by hydrogen-bonded interactions, as opposed to electrostatic in the SA1·DMA1 cluster. This explains the shorter lifetime of MaA1·DMA1 compared to SA1·DMA1.33,84 For the experiments carried out in this study, neither MaA1·DMA1 nor SA1·DMA1 were measured by the PCC due to evaporation of the base upon ionization.33,84 However, larger clusters containing only sulfuric acid and DMA were measured. This suggests that SA1·DMA1 grows rapidly to larger clusters faster than it evaporates. In contrast, clusters do not grow via the addition of MaA1·DMA1, as this cluster quickly evaporates. Although it is also possible that MaA1 could be lost from larger clusters upon ionization, overall, the initial stages of cluster stabilization and growth appear to be driven primarily by SA–DMA interactions.
Figure 3 shows the measured and simulated concentrations of SAiDMA0–iMaA0–1 (Ni) clusters as a function of [MaA1] at various [DMA1]. [SA1] was held constant at 3 × 108 cm–3 with up to 5% increase throughout the experiment due to the passivation of the sulfuric acid on the injection setup. Note, Ni represents clusters containing i sulfuric acid molecules with equal or fewer base molecules (clusters are typically unstable with more base molecules and one base molecule likely evaporates upon ionization),83,84 up to one MaA (assuming that more MaA destabilizes the clusters), and any number of water molecules. For a given [DMA1], an increase in [MaA1] does not lead to a noticeable increase in various Ni clusters, as shown in Figure 3a,b. This trend is confirmed by the cluster dynamics simulations shown in Figure 3c. Note, the slight increase in [Ni] observed in Figure 3 is likely due to 5% variation in [SA1]. However, increasing [DMA1] leads to a significant increase in [Ni] (see Figure 3a,b). A relatively similar trend is again reproduced in the dynamics simulation (Figure 3c), and it has been previously observed too.83 This is due to the fact that sulfuric acid is in excess compared to DMA during the measurements, and DMA stabilizes sulfuric acid dimers with barrierless growth to ∼1 nm particle sizes.9,80,85 Note, when [DMA1] = 33 pptv, the measured sulfuric acid tetramer concentration is greater than [N3] at high [DMA1] due to the formation of N4 from N2. At high [DMA1], the concentration of available N2 is high, which would result in a similar number of N2–N1 and N2–N2 collisions to form comparable [N3] and [N4]. Regardless, the observations of Figure 3 strongly suggest that the malonic acid likely does not contribute to the initial steps of SA–DMA atmospheric NPF.
Figure 3.
Measured concentrations of clusters with i sulfuric acid molecules ([Ni]) at varying malonic acid concentration ([MaA1]) where (a) is low [DMA1] = 2 ppt and (b) high [DMA1] = 33 ppt. [SA1] held constant at 3 × 108 cm–3. (c) Simulated SA–DMA cluster concentrations ([Ni]) at varying malonic acid concentration ([MaA1]) and DMA concentration.
The key uncertainty in the PCC measurements is the ionizability of large clusters containing sulfuric acid, malonic acid, and DMA by acetate. For SA–DMA, acetate has been shown to detect more types of large clusters compared to nitrate reagent ion.83 However, to our knowledge, no previous studies have measured SA–MaA–DMA clusters using acetate ions. Malonic acid is still a much stronger acid than acetic acid and should therefore be ionized by acetate ions. Therefore, since no malonic acid clusters with any number of DMA and/or sulfuric acids, except for the heterodimer, were measured, this implies that these clusters were not formed in the reactor.
To explore if malonic acid instead plays a role in particle growth, experiments were carried out in which the concentrations of particles >1 nm were measured by the vWCPC. Figure 4a shows the ratio of measured 2 to 1 nm particle number concentrations as a function of [MaA1] for different sulfuric acid concentrations injected into the reactor with [DMA1] kept constant at 7 pptv. For a given sulfuric acid concentration, the fraction of >2 to >1 nm particle number concentration remains constant with varying [MaA1]. However, increasing the concentration of sulfuric acid from 2 × 107 to 3 × 108 cm–3 leads to an increase in the fraction of >2 to >1 nm particle number concentrations from ∼0 to 65%. The observed increase in the fraction of >2 nm particles with increasing [SA1] is due to the condensation of sulfuric acid. In Figure 4b, the particle number concentration for >1 nm at a given sulfuric acid concentration does not change with an increase in malonic acid concentration. This suggests that the malonic acid does not help form 1 nm particles or grow them to larger sizes. The ACDC simulations also show that the particle formation rate is independent of the malonic acid concentrations over a wide range of concentrations (see Figure S5), which further corroborates the experimental findings. Error bars are not shown in Figure 4a because only one set of measurements for 2 nm particles at [DMA1] = 7 pptv was obtained. Uncertainty in the absolute particle number concentrations due to uncertainty in activation efficiency will not affect the conclusions of the vWCPC results as the trends of particle number concentrations are consistent for different sulfuric acid and DMA concentrations. Overall, these results suggest two conclusions for systems where malonic acid concentrations are much greater than sulfuric acid concentrations: (1) malonic acid does not help nucleate 1 nm particles or grow 1 nm particles to 2 nm, and (2) for sub-2 nm particle sizes, nucleation and subsequent particle growth is driven mainly by sulfuric acid.
Figure 4.
(a) Ratio of >2 to >1 nm particle number concentrations as a function of the malonic acid concentration ([MaA1]). (b) Number concentration of >1 nm particles vs [MaA1] for sulfuric acid concentrations of 2 × 107 cm–3 (green), 8 × 107 cm–3 (blue), and 3 × 108 cm–3 (red). [DMA1] was kept constant at 7 pptv.
Conclusions
The effect of malonic acid on sulfuric acid-dimethylamine nucleation and initial particle growth has been examined experimentally and using modeling supported by quantum chemical calculations. Malonic acid is a highly oxidized, acidic organic molecule that was previously predicted to enhance sulfuric acid new particle formation rates. However, this study shows that malonic acid does not enhance the formation of stable sulfuric acid-dimethylamine clusters or the growth of clusters from 1 to 2 nm. For sub-2 nm particles, sulfuric acid condensation is found to be the dominant pathway for nucleation and initial particle growth even in the presence of high malonic acid concentrations compared to sulfuric acid concentrations. Small organic acids, such as malonic acid likely contribute to the growth of larger particles >2 nm, as these organic acids have been measured in larger ultrafine particles globally. Further experiments should be conducted to examine the role of malonic acid on the growth of particles >2 nm. The results from this study are useful in narrowing the range of potential compounds that enhance sulfuric acid new particle formation in the atmosphere.
Acknowledgments
S.K.W.F. and C.N.J. acknowledge funding from NSF AGS-1913504. S.K.W.F. also acknowledges support from CMU Presidential Fellowship. J.K. and J.E. thank the Independent Research Fund Denmark grant number 9064-00001B and the Centre for Scientific Computing, Aarhus http://phys.au.dk/forskning/cscaa/, for computational resources.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsomega.3c01643.
Additional details for daily baseline measurements and methods for calculating cluster concentrations (PDF)
The authors declare no competing financial interest.
Supplementary Material
References
- Finlayson-Pitts B. J.; Pitts J. N.CHAPTER 8 - Acid Deposition: Formation and Fates of Inorganic and Organic Acids in the Troposphere. In Chemistry of the Upper and Lower Atmosphere; Finlayson-Pitts B. J., Pitts J. N., Eds.; Academic Press: San Diego, 2000; pp 294–348, 10.1016/B978-012257060-5/50010-1. [DOI] [Google Scholar]
- Seinfeld J. H.; Pandis S. N.. Atmospheric Chemistry and Physics: From Air Pollution to Climate Change, 3rd ed.; Wiley, 2016. [Google Scholar]
- Merikanto J.; Spracklen D. V.; Mann G. W.; Pickering S. J.; Carslaw K. S. Impact of Nucleation on Global CCN. Atmos. Chem. Phys. 2009, 9, 8601–8616. 10.5194/acp-9-8601-2009. [DOI] [Google Scholar]
- Gordon H.; Kirkby J.; Baltensperger U.; Bianchi F.; Breitenlechner M.; Curtius J.; Dias A.; Dommen J.; Donahue N. M.; Dunne E. M.; Duplissy J.; Ehrhart S.; Flagan R. C.; Frege C.; Fuchs C.; Hansel A.; Hoyle C. R.; Kulmala M.; Kürten A.; Lehtipalo K.; Makhmutov V.; Molteni U.; Rissanen M. P.; Stozkhov Y.; Tröstl J.; Tsagkogeorgas G.; Wagner R.; Williamson C.; Wimmer D.; Winkler P. M.; Yan C.; Carslaw K. S. Causes and Importance of New Particle Formation in the Present-Day and Preindustrial Atmospheres. J. Geophys. Res.: Atmos. 2017, 122, 8739–8760. 10.1002/2017JD026844. [DOI] [Google Scholar]
- Spracklen D. V.; Carslaw K. S.; Kulmala M.; Kerminen V.-M.; Sihto S.-L.; Riipinen I.; Merikanto J.; Mann G. W.; Chipperfield M. P.; Wiedensohler A.; Birmili W.; Lihavainen H. Contribution of Particle Formation to Global Cloud Condensation Nuclei Concentrations. Geophys. Res. Lett. 2008, 35, L06808. 10.1029/2007GL033038. [DOI] [Google Scholar]
- Wang M.; Penner J. E. Aerosol Indirect Forcing in a Global Model with Particle Nucleation. Atmos. Chem. Phys. 2009, 9, 239–260. 10.5194/acp-9-239-2009. [DOI] [Google Scholar]
- Yu F.; Luo G. Simulation of Particle Size Distribution with a Global Aerosol Model: Contribution of Nucleation to Aerosol and CCN Number Concentrations. Atmos. Chem. Phys. 2009, 9, 7691–7710. 10.5194/acp-9-7691-2009. [DOI] [Google Scholar]
- Chen M.; Titcombe M.; Jiang J.; Jen C.; Kuang C.; Fischer M. L.; Eisele F. L.; Siepmann J. I.; Hanson D. R.; Zhao J.; McMurry P. H. Acid–Base Chemical Reaction Model for Nucleation Rates in the Polluted Atmospheric Boundary Layer. Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 18713–18718. 10.1073/pnas.1210285109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jen C. N.; McMurry P. H.; Hanson D. R. Stabilization of Sulfuric Acid Dimers by Ammonia, Methylamine, Dimethylamine, and Trimethylamine. J. Geophys. Res.: Atmos. 2014, 119, 7502. 10.1002/2014JD021592. [DOI] [Google Scholar]
- Loukonen V.; Kurtén T.; Ortega I. K.; Vehkamäki H.; Pádua A. A. H.; Sellegri K.; Kulmala M. Enhancing Effect of Dimethylamine in Sulfuric Acid Nucleation in the Presence of Water – a Computational Study. Atmos. Chem. Phys. 2010, 10, 4961–4974. 10.5194/acp-10-4961-2010. [DOI] [Google Scholar]
- Zhao J.; Eisele F. L.; Titcombe M.; Kuang C.; McMurry P. H. Chemical Ionization Mass Spectrometric Measurements of Atmospheric Neutral Clusters Using the Cluster-CIMS. J. Geophys. Res. 2010, 115, D08205. 10.1029/2009jd012606. [DOI] [Google Scholar]
- Erupe M. E.; Viggiano A. A.; Lee S. H. The Effect of Trimethylamine on Atmospheric Nucleation Involving H2SO4. Atmos. Chem. Phys. 2011, 11, 4767–4775. 10.5194/acp-11-4767-2011. [DOI] [Google Scholar]
- Zollner J. H.; Glasoe W. A.; Panta B.; Carlson K. K.; McMurry P. H.; Hanson D. R. Sulfuric Acid Nucleation: Power Dependencies, Variation with Relative Humidity, and Effect of Bases. Atmos. Chem. Phys. 2012, 12, 4399–4411. 10.5194/acp-12-4399-2012. [DOI] [Google Scholar]
- Riipinen I.; Pierce J. R.; Yli-Juuti T.; Nieminen T.; Häkkinen S.; Ehn M.; Junninen H.; Lehtipalo K.; Petäjä T.; Slowik J.; Chang R.; Shantz N. C.; Abbatt J.; Leaitch W. R.; Kerminen V.-M.; Worsnop D. R.; Pandis S. N.; Donahue N. M.; Kulmala M. Organic Condensation: A Vital Link Connecting Aerosol Formation to Cloud Condensation Nuclei (CCN) Concentrations. Atmos. Chem. Phys. 2011, 11, 3865–3878. 10.5194/acp-11-3865-2011. [DOI] [Google Scholar]
- Paasonen P.; Nieminen T.; Asmi E.; Manninen H. E.; Petäjä T.; Plass-Dülmer C.; Flentje H.; Birmili W.; Wiedensohler A.; Hõrrak U.; Metzger A.; Hamed A.; Laaksonen A.; Facchini M. C.; Kerminen V.-M.; Kulmala M. On the Roles of Sulphuric Acid and Low-Volatility Organic Vapours in the Initial Steps of Atmospheric New Particle Formation. Atmos. Chem. Phys. 2010, 10, 11223–11242. 10.5194/acp-10-11223-2010. [DOI] [Google Scholar]
- Metzger A.; Verheggen B.; Dommen J.; Duplissy J.; Prevot A. S. H.; Weingartner E.; Riipinen I.; Kulmala M.; Spracklen D. V.; Carslaw K. S.; Baltensperger U. Evidence for the Role of Organics in Aerosol Particle Formation under Atmospheric Conditions. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 6646–6651. 10.1073/pnas.0911330107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Glicker H. S.; Lawler M. J.; Ortega J.; de Sá S. S.; Martin S. T.; Artaxo P.; Vega Bustillos O.; de Souza R.; Tota J.; Carlton A.; Smith J. N. Chemical Composition of Ultrafine Aerosol Particles in Central Amazonia during the Wet Season. Atmos. Chem. Phys. 2019, 19, 13053–13066. 10.5194/acp-19-13053-2019. [DOI] [Google Scholar]
- Jimenez J. L.; Canagaratna M. R.; Donahue N. M.; Prevot A. S. H.; Zhang Q.; Kroll J. H.; DeCarlo P. F.; Allan J. D.; Coe H.; Ng N. L.; Aiken A. C.; Docherty K. S.; Ulbrich I. M.; Grieshop A. P.; Robinson A. L.; Duplissy J.; Smith J. D.; Wilson K. R.; Lanz V. A.; Hueglin C.; Sun Y. L.; Tian J.; Laaksonen A.; Raatikainen T.; Rautiainen J.; Vaattovaara P.; Ehn M.; Kulmala M.; Tomlinson J. M.; Collins D. R.; Cubison M. J.; Dunlea E. J.; Huffman J. A.; Onasch T. B.; Alfarra M. R.; Williams P. I.; Bower K.; Kondo Y.; Schneider J.; Drewnick F.; Borrmann S.; Weimer S.; Demerjian K.; Salcedo D.; Cottrell L.; Griffin R.; Takami A.; Miyoshi T.; Hatakeyama S.; Shimono A.; Sun J. Y.; Zhang Y. M.; Dzepina K.; Kimmel J. R.; Sueper D.; Jayne J. T.; Herndon S. C.; Trimborn A. M.; Williams L. R.; Wood E. C.; Middlebrook A. M.; Kolb C. E.; Baltensperger U.; Worsnop D. R. Evolution of Organic Aerosols in the Atmosphere. Science 2009, 326, 1525–1529. 10.1126/science.1180353. [DOI] [PubMed] [Google Scholar]
- Yang L.; Yu L. E. Measurements of Oxalic Acid, Oxalates, Malonic Acid, and Malonates in Atmospheric Particulates. Environ. Sci. Technol. 2008, 42, 9268–9275. 10.1021/es801820z. [DOI] [PubMed] [Google Scholar]
- Arquero K. D.; Gerber R. B.; Finlayson-Pitts B. J. The Role of Oxalic Acid in New Particle Formation from Methanesulfonic Acid, Methylamine, and Water. Environ. Sci. Technol. 2017, 51, 2124–2130. 10.1021/acs.est.6b05056. [DOI] [PubMed] [Google Scholar]
- Pope F. D.; Dennis-Smither B. J.; Griffiths P. T.; Clegg S. L.; Cox R. A. Studies of Single Aerosol Particles Containing Malonic Acid, Glutaric Acid, and Their Mixtures with Sodium Chloride. I. Hygroscopic Growth. J. Phys. Chem. A 2010, 114, 5335–5341. 10.1021/jp100059k. [DOI] [PubMed] [Google Scholar]
- Elm J.; Myllys N.; Kurtén T. What Is Required for Highly Oxidized Molecules To Form Clusters with Sulfuric Acid?. J. Phys. Chem. A 2017, 121, 4578–4587. 10.1021/acs.jpca.7b03759. [DOI] [PubMed] [Google Scholar]
- Kulmala M.; Kontkanen J.; Junninen H.; Lehtipalo K.; Manninen H. E.; Nieminen T.; Petäjä T.; Sipilä M.; Schobesberger S.; Rantala P.; Franchin A.; Jokinen T.; Järvinen E.; Äijälä M.; Kangasluoma J.; Hakala J.; Aalto P. P.; Paasonen P.; Mikkilä J.; Vanhanen J.; Aalto J.; Hakola H.; Makkonen U.; Ruuskanen T.; Mauldin R. L.; Duplissy J.; Vehkamäki H.; Bäck J.; Kortelainen A.; Riipinen I.; Kurtén T.; Johnston M. V.; Smith J. N.; Ehn M.; Mentel T. F.; Lehtinen K. E. J.; Laaksonen A.; Kerminen V.-M.; Worsnop D. R. Direct Observations of Atmospheric Aerosol Nucleation. Science 2013, 339, 943–946. 10.1126/science.1227385. [DOI] [PubMed] [Google Scholar]
- Zhang R.; Wang L.; Khalizov A. F.; Zhao J.; Zheng J.; McGraw R. L.; Molina L. T. Formation of Nanoparticles of Blue Haze Enhanced by Anthropogenic Pollution. Proc. Natl. Acad. Sci. U. S. A. 2009, 106, 17650–17654. 10.1073/pnas.0910125106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawamura K.; Kaplan I. R. Motor Exhaust Emissions as a Primary Source for Dicarboxylic Acids in Los Angeles Ambient Air. Environ. Sci. Technol. 1987, 21, 105–110. 10.1021/es00155a014. [DOI] [Google Scholar]
- Rozaini M. Z. H.The Chemistry of Dicarboxylic Acids in the Atmospheric Aerosols. In Atmospheric Aerosols - Regional Characteristics - Chemistry and Physics 2012, 10.5772/50127. [DOI] [Google Scholar]
- Liang Y.; Jen C. N.; Weber R. J.; Misztal P. K.; Goldstein A. H. Chemical Composition of PM2.5 in October 2017 Northern California Wildfire Plumes. Atmos. Chem. Phys. 2021, 21, 5719–5737. 10.5194/acp-21-5719-2021. [DOI] [Google Scholar]
- Narukawa M.; Kawamura K.; Takeuchi N.; Nakajima T. Distribution of Dicarboxylic Acids and Carbon Isotopic Compositions in Aerosols from 1997 Indonesian Forest Fires. Geophys. Res. Lett. 1999, 26, 3101–3104. 10.1029/1999GL010810. [DOI] [Google Scholar]
- Martinelango P. K.; Dasgupta P. K.; Al-Horr R. S. Atmospheric Production of Oxalic Acid/Oxalate and Nitric Acid/Nitrate in the Tampa Bay Airshed: Parallel Pathways. Atmos. Environ. 2007, 41, 4258–4269. 10.1016/j.atmosenv.2006.05.085. [DOI] [Google Scholar]
- Madronich S.; Calvert J. G. Permutation Reactions of Organic Peroxy Radicals in the Troposphere. J. Geophys. Res.: Atmos. 1990, 95, 5697–5715. 10.1029/JD095iD05p05697. [DOI] [Google Scholar]
- Chen J.; Jiang S.; Liu Y.-R.; Huang T.; Wang C.-Y.; Miao S.-K.; Wang Z.-Q.; Zhang Y.; Huang W. Interaction of Oxalic Acid with Dimethylamine and Its Atmospheric Implications. RSC Adv. 2017, 7, 6374–6388. 10.1039/C6RA27945G. [DOI] [Google Scholar]
- Arquero K. D.; Xu J.; Gerber R. B.; Finlayson-Pitts B. J. Particle Formation and Growth from Oxalic Acid, Methanesulfonic Acid, Trimethylamine and Water: A Combined Experimental and Theoretical Study. Phys. Chem. Chem. Phys. 2017, 19, 28286–28301. 10.1039/C7CP04468B. [DOI] [PubMed] [Google Scholar]
- Wang Z.-Q.; Liu Y.-R.; Wang C.-Y.; Jiang S.; Feng Y.-J.; Huang T.; Huang W. Multicomponent Nucleation of Malonic Acid Involved in the Sulfuric Acid - Dimethylamine System and Its Atmospheric Implications. Atmos. Environ. 2021, 267, 118558 10.1016/j.atmosenv.2021.118558. [DOI] [Google Scholar]
- Zhang H.; Li H.; Liu L.; Zhang Y.; Zhang X.; Li Z. The Potential Role of Malonic Acid in the Atmospheric Sulfuric Acid - Ammonia Clusters Formation. Chemosphere 2018, 203, 26–33. 10.1016/j.chemosphere.2018.03.154. [DOI] [PubMed] [Google Scholar]
- Donahue N. M.; Chuang W.; Epstein S. A.; Kroll J. H.; Worsnop D. R.; Robinson A. L.; Adams P. J.; Pandis S. N. Why Do Organic Aerosols Exist? Understanding Aerosol Lifetimes Using the Two-Dimensional Volatility Basis Set. Environ. Chem. 2013, 10, 151. 10.1071/EN13022. [DOI] [Google Scholar]
- Donahue N. M.; Epstein S. A.; Pandis S. N.; Robinson A. L. A Two-Dimensional Volatility Basis Set: 1. Organic-Aerosol Mixing Thermodynamics. Atmos. Chem. Phys. 2011, 11, 3303–3318. 10.5194/acp-11-3303-2011. [DOI] [Google Scholar]
- 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.; 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, 429–433. 10.1038/nature10343. [DOI] [PubMed] [Google Scholar]; http://www.nature.com/nature/journal/v476/n7361/abs/nature10343.html#supplementary-information
- Kerminen V. M.; Petäjä T.; Manninen H. E.; Paasonen P.; Nieminen T.; Sipilä M.; Junninen H.; Ehn M.; Gagné S.; Laakso L.; Riipinen I.; Vehkamäki H.; Kurten T.; Ortega I. K.; Dal Maso M.; Brus D.; Hyvärinen A.; Lihavainen H.; Leppä J.; Lehtinen K. E. J.; Mirme A.; Mirme S.; Hõrrak U.; Berndt T.; Stratmann F.; Birmili W.; Wiedensohler A.; Metzger A.; Dommen J.; Baltensperger U.; Kiendler-Scharr A.; Mentel T. F.; Wildt J.; Winkler P. M.; Wagner P. E.; Petzold A.; Minikin A.; Plass-Dülmer C.; Pöschl U.; Laaksonen A.; Kulmala M. Atmospheric Nucleation: Highlights of the EUCAARI Project and Future Directions. Atmos. Chem. Phys. 2010, 10, 10829–10848. 10.5194/acp-10-10829-2010. [DOI] [Google Scholar]
- Kulmala M.; Kerminen V.-M. On the Formation and Growth of Atmospheric Nanoparticles. Atmos. Res. 2008, 90, 132–150. 10.1016/j.atmosres.2008.01.005. [DOI] [Google Scholar]
- 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, 1243–1246. 10.1126/science.1180315. [DOI] [PubMed] [Google Scholar]
- Weber R. J.; Marti J. J.; McMurry P. H.; Eisele F. L.; Tanner D. J.; Jefferson A. Measurements of New Particle Formation and Ultrafine Particle Growth Rates at a Clean Continental Site. J. Geophys. Res.: Atmos. 1997, 102, 4375–4385. 10.1029/96JD03656. [DOI] [Google Scholar]
- Zhang R.; Khalizov A.; Wang L.; Hu M.; Xu W. Nucleation and Growth of Nanoparticles in the Atmosphere. Chem. Rev. 2012, 112, 1957–2011. 10.1021/cr2001756. [DOI] [PubMed] [Google Scholar]
- Nadykto A. B.; Herb J.; Yu F.; Xu Y. Enhancement in the Production of Nucleating Clusters Due to Dimethylamine and Large Uncertainties in the Thermochemistry of Amine-Enhanced Nucleation. Chem. Phys. Lett. 2014, 609, 42–49. 10.1016/j.cplett.2014.03.036. [DOI] [Google Scholar]
- Ortega I. K.; Olenius T.; Kupiainen-Mättä O.; Loukonen V.; Kurtén T.; Vehkamäki H. Electrical Charging Changes the Composition of Sulfuric Acid–Ammonia/Dimethylamine Clusters. Atmos. Chem. Phys. 2014, 14, 7995–8007. 10.5194/acp-14-7995-2014. [DOI] [Google Scholar]
- Olenius T.; Halonen R.; Kurtén T.; Henschel H.; Kupiainen-Mättä O.; Ortega I. K.; Jen C. N.; Vehkamäki H.; Riipinen I. New Particle Formation from Sulfuric Acid and Amines: Comparison of Monomethylamine, Dimethylamine, and Trimethylamine. J. Geophys. Res.: Atmos. 2017, 122, 7103–7118. 10.1002/2017JD026501. [DOI] [Google Scholar]
- Sihto S. L.; Kulmala M.; Kerminen V. M.; Dal Maso M.; Petäjä T.; Riipinen I.; Korhonen H.; Arnold F.; Janson R.; Boy M.; Laaksonen A.; Lehtinen K. E. J. Atmospheric Sulphuric Acid and Aerosol Formation: Implications from Atmospheric Measurements for Nucleation and Early Growth Mechanisms. Atmos. Chem. Phys. 2006, 6, 4079–4091. 10.5194/acp-6-4079-2006. [DOI] [Google Scholar]
- Elm J.; Passananti M.; Kurtén T.; Vehkamäki H. Diamines Can Initiate New Particle Formation in the Atmosphere. J. Phys. Chem. A 2017, 121, 6155–6164. 10.1021/acs.jpca.7b05658. [DOI] [PubMed] [Google Scholar]
- Elm J.; Jen C. N.; Kurtén T.; Vehkamäki H. Strong Hydrogen Bonded Molecular Interactions between Atmospheric Diamines and Sulfuric Acid. J. Phys. Chem. A 2016, 120, 3693–3700. 10.1021/acs.jpca.6b03192. [DOI] [PubMed] [Google Scholar]
- Schmitz G.; Elm J. Assessment of the DLPNO Binding Energies of Strongly Noncovalent Bonded Atmospheric Molecular Clusters. ACS Omega 2020, 5, 7601–7612. 10.1021/acsomega.0c00436. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bilde M.; Barsanti K.; Booth M.; Cappa C. D.; Donahue N. M.; Emanuelsson E. U.; McFiggans G.; Krieger U. K.; Marcolli C.; Topping D.; Ziemann P.; Barley M.; Clegg S.; Dennis-Smither B.; Hallquist M.; Hallquist Å. M.; Khlystov A.; Kulmala M.; Mogensen D.; Percival C. J.; Pope F.; Reid J. P.; Ribeiro da Silva M. A. V.; Rosenoern T.; Salo K.; Soonsin V. P.; Yli-Juuti T.; Prisle N. L.; Pagels J.; Rarey J.; Zardini A. A.; Riipinen I. Saturation Vapor Pressures and Transition Enthalpies of Low-Volatility Organic Molecules of Atmospheric Relevance: From Dicarboxylic Acids to Complex Mixtures. Chem. Rev. 2015, 115, 4115–4156. 10.1021/cr5005502. [DOI] [PubMed] [Google Scholar]
- Hodshire A. L.; Lawler M. J.; Zhao J.; Ortega J.; Jen C.; Yli-Juuti T.; Brewer J. F.; Kodros J. K.; Barsanti K. C.; Hanson D. R.; McMurry P. H.; Smith J. N.; Pierce J. R. Multiple New-Particle Growth Pathways Observed at the US DOE Southern Great Plains Field Site. Atmos. Chem. Phys. 2016, 16, 9321–9348. 10.5194/acp-16-9321-2016. [DOI] [Google Scholar]
- Fang X.; Hu M.; Shang D.; Tang R.; Shi L.; Olenius T.; Wang Y.; Wang H.; Zhang Z.; Chen S.; Yu X.; Zhu W.; Lou S.; Ma Y.; Li X.; Zeng L.; Wu Z.; Zheng J.; Guo S. Observational Evidence for the Involvement of Dicarboxylic Acids in Particle Nucleation. Environ. Sci. Technol. Lett. 2020, 7, 388–394. 10.1021/acs.estlett.0c00270. [DOI] [Google Scholar]
- Fomete S. K. W.; Johnson J. S.; Casalnuovo D.; Jen C. N. A Tutorial Guide on New Particle Formation Experiments Using a Laminar Flow Reactor. J. Aerosol Sci. 2021, 157, 105808 10.1016/j.jaerosci.2021.105808. [DOI] [Google Scholar]
- Panta B.; Glasoe W. A.; Zollner J. H.; Carlson K. K.; Hanson D. R. Computational Fluid Dynamics of a Cylindrical Nucleation Flow Reactor with Detailed Cluster Thermodynamics. J. Phys. Chem. A 2012, 116, 10122–10134. 10.1021/jp302444y. [DOI] [PubMed] [Google Scholar]
- Jen C. N.; Bachman R.; Zhao J.; McMurry P. H.; Hanson D. R. Diamine-Sulfuric Acid Reactions Are a Potent Source of New Particle Formation. Geophys. Res. Lett. 2016, 43, 867. 10.1002/2015GL066958. [DOI] [Google Scholar]
- Kürten A.; Bergen A.; Heinritzi M.; Leiminger M.; Lorenz V.; Piel F.; Simon M.; Sitals R.; Wagner A. C.; Curtius J. Observation of New Particle Formation and Measurement of Sulfuric Acid, Ammonia, Amines and Highly Oxidized Organic Molecules at a Rural Site in Central Germany. Atmos. Chem. Phys. 2016, 16, 12793–12813. 10.5194/acp-16-12793-2016. [DOI] [Google Scholar]
- Cai R.; Yan C.; Yang D.; Yin R.; Lu Y.; Deng C.; Fu Y.; Ruan J.; Li X.; Kontkanen J.; Zhang Q.; Kangasluoma J.; Ma Y.; Hao J.; Worsnop D. R.; Bianchi F.; Paasonen P.; Kerminen V.-M.; Liu Y.; Wang L.; Zheng J.; Kulmala M.; Jiang J. Sulfuric Acid–Amine Nucleation in Urban Beijing. Atmos. Chem. Phys. 2021, 21, 2457–2468. 10.5194/acp-21-2457-2021. [DOI] [Google Scholar]
- Fomete S. K. W.; Johnson J. S.; Myllys N.; Jen C. N. Experimental and Theoretical Study on the Enhancement of Alkanolamines on Sulfuric Acid Nucleation. J. Phys. Chem. A 2022, 126, 4057–4067. 10.1021/acs.jpca.2c01672. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hering S. V.; Lewis G. S.; Spielman S. R.; Eiguren-Fernandez A.; Kreisberg N. M.; Kuang C.; Attoui M. Detection near 1-Nm with a Laminar-Flow, Water-Based Condensation Particle Counter. Aerosol Sci. Technol. 2017, 51, 354–362. 10.1080/02786826.2016.1262531. [DOI] [Google Scholar]
- McGrath M. J.; Olenius T.; Ortega I. K.; Loukonen V.; Paasonen P.; Kurtén T.; Kulmala M.; Vehkamäki H. Atmospheric Cluster Dynamics Code: A Flexible Method for Solution of the Birth-Death Equations. Atmos. Chem. Phys. 2012, 12, 2345–2355. 10.5194/acp-12-2345-2012. [DOI] [Google Scholar]
- Kubečka J.; Neefjes I.; Besel V.; Qiao F.; Xie H.-B.; Elm J. Atmospheric Sulfuric Acid–Multi-Base New Particle Formation Revealed through Quantum Chemistry Enhanced by Machine Learning. J. Phys. Chem. A 2023, 127, 2091–2103. 10.1021/acs.jpca.3c00068. [DOI] [PubMed] [Google Scholar]
- Riplinger C.; Neese F. An Efficient and near Linear Scaling Pair Natural Orbital Based Local Coupled Cluster Method. J. Chem. Phys. 2013, 138, 034106 10.1063/1.4773581. [DOI] [PubMed] [Google Scholar]
- Riplinger C.; Sandhoefer B.; Hansen A.; Neese F. Natural Triple Excitations in Local Coupled Cluster Calculations with Pair Natural Orbitals. J. Chem. Phys. 2013, 139, 134101. 10.1063/1.4821834. [DOI] [PubMed] [Google Scholar]
- Riplinger C.; Pinski P.; Becker U.; Valeev E. F.; Neese F. Sparse Maps—A Systematic Infrastructure for Reduced-Scaling Electronic Structure Methods. II. Linear Scaling Domain Based Pair Natural Orbital Coupled Cluster Theory. J. Chem. Phys. 2016, 144, 024109 10.1063/1.4939030. [DOI] [PubMed] [Google Scholar]
- Dunning T. H. Gaussian Basis Sets for Use in Correlated Molecular Calculations. I. The Atoms Boron through Neon and Hydrogen. J. Chem. Phys. 1989, 90, 1007–1023. 10.1063/1.456153. [DOI] [Google Scholar]
- Kendall R. A.; Dunning T. H.; Harrison R. J. Electron Affinities of the First-row Atoms Revisited. Systematic Basis Sets and Wave Functions. J. Chem. Phys. 1992, 96, 6796–6806. 10.1063/1.462569. [DOI] [Google Scholar]
- Chai J.-D.; Head-Gordon M. Long-Range Corrected Hybrid Density Functionals with Damped Atom–Atom Dispersion Corrections. Phys. Chem. Chem. Phys. 2008, 10, 6615–6620. 10.1039/B810189B. [DOI] [PubMed] [Google Scholar]
- Zhang J.; Dolg M. ABCluster: The Artificial Bee Colony Algorithm for Cluster Global Optimization. Phys. Chem. Chem. Phys. 2015, 17, 24173–24181. 10.1039/C5CP04060D. [DOI] [PubMed] [Google Scholar]
- Zhang J.; Dolg M. Global Optimization of Clusters of Rigid Molecules Using the Artificial Bee Colony Algorithm. Phys. Chem. Chem. Phys. 2016, 18, 3003–3010. 10.1039/C5CP06313B. [DOI] [PubMed] [Google Scholar]
- Grimme S.; Bannwarth C.; Shushkov P. A Robust and Accurate Tight-Binding Quantum Chemical Method for Structures, Vibrational Frequencies, and Noncovalent Interactions of Large Molecular Systems Parametrized for All Spd-Block Elements (Z = 1–86). J. Chem. Theory Comput. 2017, 13, 1989–2009. 10.1021/acs.jctc.7b00118. [DOI] [PubMed] [Google Scholar]
- Pracht P.; Caldeweyher E.; Ehlert S.; Grimme S. A Robust Non-Self-Consistent Tight-Binding Quantum Chemistry Method for Large Molecules. ChemRxiv 2019, 10.26434/chemrxiv.8326202.v1. [DOI] [Google Scholar]
- Frisch M. J.; Trucks G. W.; Schlegel H. B.; Scuseria G. E.; Robb M. A.; Cheeseman J. R.; Scalmani G.; Barone V.; Petersson G. A.; Nakatsuji H.; Li X.; Caricato M.; Marenich A. V.; Bloino J.; Janesko B. G.; Gomperts R.; Mennucci B.; Hratchian H. P.; Ortiz J. V.; Izmaylov A. F.; Sonnenberg J. L.; Williams; Ding F.; Lipparini F.; Egidi F.; Goings J.; Peng B.; Petrone A.; Henderson T.; Ranasinghe D.; Zakrzewski V. G.; Gao J.; Rega N.; Zheng G.; Liang W.; Hada M.; Ehara M.; Toyota K.; Fukuda R.; Hasegawa J.; Ishida M.; Nakajima T.; Honda Y.; Kitao O.; Nakai H.; Vreven T.; Throssell K.; Montgomery J. A. Jr.; Peralta J. E.; Ogliaro F.; Bearpark M. J.; Heyd J. J.; Brothers E. N.; Kudin K. N.; Staroverov V. N.; Keith T. A.; Kobayashi R.; Normand J.; Raghavachari K.; Rendell A. P.; Burant J. C.; Iyengar S. S.; Tomasi J.; Cossi M.; Millam J. M.; Klene M.; Adamo C.; Cammi R.; Ochterski J. W.; Martin R. L.; Morokuma K.; Farkas O.; Foresman J. B.; Fox D. J.. Gaussian 16 Rev. C.01, 2016.
- Neese F. The ORCA Program System. WIREs Comput. Mol. Sci. 2012, 2, 73–78. 10.1002/wcms.81. [DOI] [Google Scholar]
- Neese F. Software Update: The ORCA Program System, Version 4.0. WIREs Comput. Mol. Sci. 2018, 8, e1327 10.1002/wcms.1327. [DOI] [Google Scholar]
- Elm J.; Mikkelsen K. V. Computational Approaches for Efficiently Modelling of Small Atmospheric Clusters. Chem. Phys. Lett. 2014, 615, 26–29. 10.1016/j.cplett.2014.09.060. [DOI] [Google Scholar]
- Myllys N.; Elm J.; Kurten T. Density Functional Theory Basis Set Convergence of Sulfuric Acid-Containing Molecular Clusters. Comput. Theor. Chem. 2016, 1098, 1–12. 10.1016/j.comptc.2016.10.015. [DOI] [Google Scholar]
- Jensen A. B.; Kubečka J.; Schmitz G.; Christiansen O.; Elm J. Massive Assessment of the Binding Energies of Atmospheric Molecular Clusters. J. Chem. Theory Comput. 2022, 18, 7373–7383. 10.1021/acs.jctc.2c00825. [DOI] [PubMed] [Google Scholar]
- Elm J.; Kubecka J.; Besel V.; Jäskeläinen M. J.; Halonen R.; Kurten T.; Vehkamäki H. Modeling the Formation and Growth of Atmospheric Molecular Clusters: A Review. J. Aerosol Sci. 2020, 149, 105621 10.1016/j.jaerosci.2020.105621. [DOI] [Google Scholar]
- Smith J. N.; Draper D. C.; Chee S.; Dam M.; Glicker H.; Myers D.; Thomas A. E.; Lawler M. J.; Myllys N. Atmospheric Clusters to Nanoparticles: Recent Progress and Challenges in Closing the Gap in Chemical Composition. J. Aerosol Sci. 2021, 153, 105733 10.1016/j.jaerosci.2020.105733. [DOI] [Google Scholar]
- Kürten A.; Jokinen T.; Simon M.; Sipilä M.; Sarnela N.; Junninen H.; Adamov A.; Almeida J.; Amorim A.; Bianchi F.; Breitenlechner M.; Dommen J.; Donahue N. M.; Duplissy J.; Ehrhart S.; Flagan R. C.; Franchin A.; Hakala J.; Hansel A.; Heinritzi M.; Hutterli M.; Kangasluoma J.; Kirkby J.; Laaksonen A.; Lehtipalo K.; Leiminger M.; Makhmutov V.; Mathot S.; Onnela A.; Petäjä T.; Praplan A. P.; Riccobono F.; Rissanen M. P.; Rondo L.; Schobesberger S.; Seinfeld J. H.; Steiner G.; Tomé A.; Tröstl J.; Winkler P. M.; Williamson C.; Wimmer D.; Ye P.; Baltensperger U.; Carslaw K. S.; Kulmala M.; Worsnop D. R.; Curtius J. Neutral Molecular Cluster Formation of Sulfuric Acid–Dimethylamine Observed in Real Time under Atmospheric Conditions. Proc. Natl. Acad. Sci. U. S. A. 2014, 15019. 10.1073/pnas.1404853111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 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. Molecular Understanding of Sulphuric Acid-Amine Particle Nucleation in the Atmosphere. Nature 2013, 502, 359–363. 10.1038/nature12663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kürten A.; Li C.; Bianchi F.; Curtius J.; Dias A.; Donahue N. M.; Duplissy J.; Flagan R. C.; Hakala J.; Jokinen T.; Kirkby J.; Kulmala M.; Laaksonen A.; Lehtipalo K.; Makhmutov V.; Onnela A.; Rissanen M. P.; Simon M.; Sipilä M.; Stozhkov Y.; Tröstl J.; Ye P.; McMurry P. H. New Particle Formation in the Sulfuric Acid-Dimethy Lamine-Water System: Reevaluation of CLOUD Chamber Measurements and Comparison to an Aerosol Nucleation and Growth Model. Atmos. Chem. Phys. Discuss. 2017, 2017, 1–31. 10.5194/acp-2017-636. [DOI] [Google Scholar]
- Jen C. N.; Zhao J.; McMurry P. H.; Hanson D. R. Chemical Ionization of Clusters Formed from Sulfuric Acid and Dimethylamine or Diamines. Atmos. Chem. Phys. 2016, 16, 12513–12529. 10.5194/acp-16-12513-2016. [DOI] [Google Scholar]
- Ortega I. K.; Kupiainen O.; Kurtén T.; Olenius T.; Wilkman O.; McGrath M. J.; Loukonen V.; Vehkamäki H. From Quantum Chemical Formation Free Energies to Evaporation Rates. Atmos. Chem. Phys. 2012, 12, 225–235. 10.5194/acp-12-225-2012. [DOI] [Google Scholar]
- Olenius T.; Kupiainen-Mättä O.; Ortega I. K.; Kurtén T.; Vehkamäki H. Free Energy Barrier in the Growth of Sulfuric Acid–Ammonia and Sulfuric Acid–Dimethylamine Clusters. J. Chem. Phys. 2013, 139, 084312 10.1063/1.4819024. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.



