Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Significance of halogen bonding in the synergistic nucleation of iodine oxoacids and iodine oxides

Rongjie Zhang a, Yueyang Liua, Rujing Yina, Fangfang Maa, Deming Xiaab, Jingwen Chena, Hong-Bin Xie*a and Joseph S. Francisco*b
aKey Laboratory of Industrial Ecology and Environmental Engineering (Ministry of Education), School of Environmental Science and Technology, Dalian University of Technology, Dalian 116024, China. E-mail: hbxie@dlut.edu.cn
bDepartment of Earth and Environmental Science, University of Pennsylvania, Philadelphia, PA 19104-6316, USA. E-mail: frjoseph@sas.upenn.edu

Received 3rd April 2025 , Accepted 19th July 2025

First published on 21st July 2025


Abstract

Congeneric iodine oxoacids and iodine oxides, key nucleating vapours in the marine atmosphere, have been reported to nucleate individually. However, whether they can nucleate together remains unknown. Here, we provide molecular-level evidence that I2O4, the iodine oxide with the highest nucleation potential towards iodine oxoacids, can synergistically nucleate with HIO3–HIO2. The nucleation rate of HIO3–HIO2–I2O4 is 1.5 to 6.8 times higher than that of the known most efficient iodine-associated two-component (HIO3–HIO2) nucleation at 278.15 K, enhancing the role of iodine-containing species in marine atmospheric particle formation. Microscopic analysis of the three-component cluster configurations revealed that an unexpected acid–base reaction between I2O4 and HIO2/HIO3 is a key driver of this efficient synergistic nucleation, besides hydrogen bonds and halogen bonds. We identified halogen bond-induced basicity enhancement as the chemical nature of I2O4 behaving as a base in the nucleation process with HIO2/HIO3. Such a basicity enhancement effect can be extended to other iodine-containing species, e.g., HIO2 and the more acidic HIO3, suggesting a common feature in interactions between iodine-containing species. Our findings clarify the synergistic nucleation of iodine oxoacids and iodine oxides and highlight the necessity of considering the effect of halogen bond-induced basicity enhancement on the formation of iodine-containing particles.


Introduction

New particle formation (NPF) initiated by the nucleation of condensable vapours produces more than half of all atmospheric aerosols.1–3 These nascent particles act as cloud condensation nuclei (CCN), significantly influencing the earth's atmosphere radiative balance and modulating global climate patterns.4,5 Marine NPF plays a pivotal role in the global climate model, as marine clouds have high albedo and are susceptible to changes in CCN availability.6–10 However, the influence of marine NPF on climate remains uncertain, mainly due to the limited understanding of the nucleation process in the marine atmosphere, thus resulting in extensive uncertainty within global models.

Conventional theory has been based on the assumption that sulfuric acid (SA)–ammonia (NH3) nucleation is the dominant mechanism in NPF in marine environments.11,12 However, recent research has highlighted the importance of iodine oxides (IxOy) and iodine oxoacids (HIOz) as crucial nucleation precursors.10,13–28 These species are formed mainly by the transformation of iodine (I2) and methyl iodide (CH3I), which are emitted from the ocean surface.29–33 Although iodine-initiated NPF events have been frequently observed,13,34–37 the specific nucleation mechanism is still not fully understood. One study reported that in flow tube experiments, the nucleation of iodine oxides, particularly I2O2, I2O3, and I2O4 at high concentrations, is rapid.14 However, Cosmics Leaving Outdoor Droplets (CLOUD) chamber experiments have shown that nucleation involving iodic acid (HIO3) and iodous acid (HIO2) can occur rapidly under marine boundary layer conditions.38 In these CLOUD chamber experiments, the concentration of formed iodine oxides is relatively low when the precursor concentration is low, which could be one of the reasons why iodine oxide nucleation cannot compete with HIO3–HIO2 nucleation. Iodine oxides and iodine oxoacids are known to coexist on the basis of their homology and can form strong halogen bonds (typical halogen bond energies ΔE ranging from 1 to 45 kcal mol−1 calculated by Oliveira et al.39), which are non-covalent interactions formed between an electrophilic halogen atom and a nucleophilic heteroatom (i.e., with lone-pair electrons).39,40 Therefore, they may be able to nucleate together, so investigating the joint nucleation mechanism for iodine oxides and iodine oxoacids is important.

In this study, the multicomponent nucleation mechanism of iodine oxides and iodine oxoacids was investigated via quantum chemical calculations and atmospheric cluster dynamic coding. Here, multicomponent nucleation was considered to occur among HIO3, HIO2 and specific iodine oxides since HIO3 and HIO2 have been found to be able to nucleate at realistic atmospheric concentration ranges.38,41 Since dimer can present the foundational intermolecular interaction between iodine oxides and iodine oxoacids, and the computational cost of dimer is low, the dimer formation free energy (ΔG) between iodine oxides (IO, I2O2, I2O3, I2O4, and I2O5) and iodine oxoacids (HIO3 and HIO2) was used to screen the propensity of specific iodine oxides towards nucleation with iodine oxoacids. I2O4 was found to have the greatest nucleation potential among the selected iodine oxides and can efficiently nucleate with HIO3–HIO2 as a base. Halogen bond-induced basicity enhancement of I2O4 was found to be the chemical nature for I2O4 behaving as a base during the nucleation with HIO3–HIO2. This study reveals that iodine oxides and iodine oxoacids can synergistically nucleate, breaking through the previous findings of their independent nucleation14,38 and providing a new chemical mechanism for iodine-containing particles.

Methods

A multistep sampling scheme was employed to identify the global minimum for all of the studied clusters, including IxOy–HIO2–3 dimer clusters and HIO3–HIO2–I2O4 system clusters. This procedure has been widely used in many previous studies on atmospheric cluster formation.42–44 First, 5000–8000 initial configurations were generated for each cluster using the ABCluster program45 and were optimized via the semiempirical PM7 method. The single point energy was subsequently calculated at the M06-2X/def2-TZVP level for all the converged geometries. Configurations up to 15 kcal mol−1 higher than the lowest energy configuration were further fully optimized at the M06-2X/6-31++G(d,p) (for H O atoms) + aug-cc-pVTZ-PP (a basis set with relativistic pseudopotential for I atoms46) level. Since there are tens of configurations we have to optimize for each cluster in this step, the adopted larger basis set such as aug-cc-pVTZ(-PP) (aug-cc-pVTZ for H O atoms and aug-cc-pVTZ-PP for I atoms) would lead to the high-cost for the computational time. Hence, we employed the smaller basis set here to reduce computational cost. Our previous study found that the adopted basis set here for optimization presents similar ΔG for HIO3–HIO2 clusters to that of the aug-cc-pVTZ(-PP) basis set (Table S1).41 Moreover, we also tested several clusters including (HIO2)1(I2O4)1, (HIO3)2(I2O4)2 and (HIO3)1(HIO2)1(I2O4)1 with the other basis set def2-TZVP. It was found that the low-energy configurations (within 1–2 kcal mol−1 of the lowest-energy configuration) from def2-TZVP basis set are the same as those from 6-31++G(d,p) + aug-cc-pVTZ-PP (Fig. S1). Therefore, the adopted basis set is acceptable for the first-round geometry optimization for HIO3–HIO2–I2O4 system. Finally, for the lowest-energy configurations within 1–2 kcal mol−1, reoptimization was performed at the M06-2X/aug-cc-pVTZ(-PP) level, and the single point energy was refined at the DLPNO-CCSD(T)/aug-cc-pVTZ(-PP) level. If the calculation results failed during optimization or ended up with imaginary frequencies, the initial configurations were adjusted and reoptimized until a successful optimization without imaginary frequencies was finished. The configuration with the lowest Gibbs free energy (G) at 298.15 K and 1 atm was chosen as the global minimum. The G values at other temperatures and pressures were obtained via a GoodVibes Python script.47 The ΔG for the global minimum was calculated by subtracting the G of the constituent molecules from that of the cluster. We also tested the influence of the dispersion correction by reoptimizing the global minima for selected three clusters ((HIO2)1(I2O4)1, (HIO3)2(I2O4)2 and (HIO3)1(HIO2)1(I2O4)1) using M06-2X/aug-cc-pVTZ(-PP), with Grimme's dispersion correction with ZERO Damping GD3.48–50 The root-mean-square deviation (RMSD) for optimized structure and difference in ΔG from method with and without dispersion correction are 0.001–0.003 Å and −0.004–0.012 kcal mol−1, indicating that the dispersion correction causes slight influence on the results of this study (Table S2). All the PM7 and M06-2X calculations were carried out with the Gaussian 16 program51 and DLPNO-CCSD(T) calculations were performed with the software ORCA 5.0.3.52,53 In addition, natural bond orbital (NBO) analysis was carried out by Gaussian 16 program to give a detailed insight into intermolecular interactions.

The cluster formation rate and growth mechanism were analysed via Atmospheric Cluster Dynamics Code (ACDC) software, which simulates the cluster kinetics by means of explicit solution of the birth–death equations.54 Herein, the simulated system was treated as a 3 × 3 box containing (HIO3)0–3(HIO2)0–3, (HIO3)0–3(I2O4)0–3, (HIO2)0–3(I2O4)0–3 and (HIO3)x(HIO2)y(I2O4)z (x = 1–3, y + z = 2–3) clusters. The boundary cluster settings can be found in the ESI. The simulation mainly ran at 278.15 K and 1 atm with a coagulation sink rate coefficient (kcoag) of 2 × 10−3 s−1, a typical value in coastal regions where iodine species are nucleating.55 Additional simulations were performed at 298.15, 253, 223.15 K to probe the effect of temperature, and further simulations were conducted at 0.5 and 0.1 atm to test the effect of pressure. The concentration of HIO3 ([HIO3]) was set to 105 to 108 cm−3 based on the common atmospheric concentration range.13,36–38 The concentrations of HIO2 and I2O4 were set to approximately 1/30 [HIO3] and 1/100 [HIO3], respectively, corresponding to the measured steady-state concentrations in the CLOUD chamber experiment.38 The detailed discussion on the selection of [I2O4] was presented in the ESI. We accounted for the contribution of long-range interactions via an enhancement factor of 2.4 for the collision rate coefficient.41,56 In addition, the cluster formation rate of the SA–HIO3–HIO2 system ((SA)x(HIO3)y(HIO2)z (0 ≤ x + y ≤ 3, 1 ≤ z ≤ 3) clusters) was simulated for comparison, with the concentration of SA ranging from 106 to 107 cm−3.37,57

The similar ACDC simulations based on thermodynamics data from the same theoretical methods as ones used here have been employed in the previous studies on SA–HIO3–HIO2 and HIO3–HIO2 nucleation, producing consistent cluster formation rates with those from CLOUD experiments.26,41 Therefore, it is reasonably believed that adopted computational level of theories combined with ACDC simulations can be reliable to describe the interactions among the iodine species and predict the cluster formation mechanism for the HIO3–HIO2–I2O4 system.

To investigate the correlation between the formation of halogen bonds (XBs) and the change in basicity of I2O4, a series of I2O4-containing dimer clusters that contain only one XB were manually constructed and optimized at the M06-2X/aug-cc-pVTZ(-PP) level. After optimization, the single point energies (E) of these dimer clusters were calculated at the DLPNO-CCSD(T)/aug-cc-pVTZ(-PP) level. The bond energy (ΔE) was calculated to represent the strength of the formed XB:

 
ΔE = |E(precursor–I2O4 dimer) − E(precursor) − E(I2O4)| (1)

Previous study has shown that the change in electrostatic potential (ΔESP) at specific basic site can be employed to present the change in basicity of the compounds.58 Herein, we employed ΔESP for the basic site of I2O4 to represent the change in basicity of I2O4 when it forms XBs with various precursors. The computational details for ΔESP was presented in the ESI. Furthermore, we also directly calculated the gas basicity (GB) of I2O4 before and after the formation of XBs, which is defined as the Gibbs free energy change (ΔG) for reactions (2) and (3), respectively:

 
I2O4H+ (g) → I2O4 (g) + H+ (g) (2)
 
precursor–I2O4H+ (g) → precursor–I2O4 (g) + H+ (g) (3)

It is noted that the terminal O atom of I2O4 with the most negative average ESP after forming XB was taken as the basic site for ΔESP and GB calculation. The geometries of positive precursor–I2O4H+ dimers and I2O4H+ were manually constructed by adding H+ to the basic site of I2O4.

Results and discussion

Of all iodine oxides, I2O4 is the precursor with the highest potential towards iodine oxoacids

The nucleation potential of iodine oxides towards iodine oxoacids was checked by their dimerization with HIO2 or HIO3. Fig. 1 presents the ΔG values and global minimum configurations of the HIO2-containing and HIO3-containing dimer clusters. The ΔG values for all dimers except IO–HIO2 and IO–HIO3 are negative, ranging from −9.7 to −21.6 kcal mol−1. Of all the HIO2-containing and HIO3-containing dimers, HIO2–I2O4 and HIO3–I2O4 exhibit the lowest ΔG values, followed by HIO2–I2O2 and HIO3–I2O2, respectively. Therefore, I2O4 is expected to have the greatest intrinsic nucleation potential towards iodine oxoacids, followed by I2O2. In addition, the ΔG values of HIO2–I2O4 and HIO3–I2O4 are even lower than that of HIO3–HIO2 (−16.79 kcal mol−1), indicating that I2O4 has a greater ability to bind with HIO2 or HIO3 than HIO3 and HIO2 do with each other. It is noteworthy that Engsvang et al. also calculated ΔG values for dimers of iodine oxides (I2O4 and I2O5) with iodine oxoacids in a recent study.59 They found I2O4–HIO2–3 dimers have lower ΔG values than I2O5–HIO2–3 dimers, consistent with results predicted here, although the specific ΔG values in this study are lower.
image file: d5sc02517f-f1.tif
Fig. 1 Formation free energy (ΔG) (A) and global minimum configurations (B) of IxOy–HIO2–3 dimers at the DLPNO-CCSD(T)/aug-cc-pVTZ(-PP)//M06-2X/aug-cc-pVTZ(-PP) level of theory at 298.15 K. The red balls represent oxygen atoms, purple ones are for iodine atoms, and white ones are for hydrogen atoms. The red dashed lines in the configurations represent hydrogen bonds (HBs), and the blue dashed lines represent XBs. The red and blue numbers indicate the bond lengths of HBs and XBs, respectively. The bond lengths are given in Å.

Observation of the configurations and NBO analysis of the dimers found that both I2O4 and I2O2 accept a proton from HIO3 in HIO3–I2O4 and HIO3–I2O2 clusters in addition to forming XBs and hydrogen bonds (HBs). Accordingly, both I2O4 and I2O2 behave as bases towards HIO3. However, for dimers of other iodine oxides with HIO3, there is no proton transfer even though XBs and HBs are formed. Therefore, the additional electrostatic interaction caused by the proton transfer reaction is expected to be one reason that I2O4 and I2O2 have lower dimer ΔG values with HIO3 than other iodine oxides do. It is noted that the structure of (HIO3)1(I2O4)1 resembles two IO3 radicals clustered with a hypoiodous acid HOI. However, (HIO3)1(I2O4)1 could not dissociate into HOI + 2·IO3 rather than HIO3 + I2O4, since the ΔG value for (HIO3)1(I2O4)1 dissociating into HOI + 2·IO3 (76.14 kcal mol−1) is much higher than HIO3 + I2O4 (18.01 kcal mol−1). For the dimers of HIO2 with I2O4 and I2O2, two XBs are formed. Two XBs are known to form for the dimers of HIO2 with other iodine oxides except for IO. The lower dimer ΔG values of I2O4 and I2O2 with HIO2 result from the stronger strength of XBs formed between them, as evidenced by the lower average energy gap between antibonding orbital δ*(O–I) and lone-pair orbital LP(O), which are two critical molecular orbitals for the formation of XBs with HIO2 (Table S3). As a representative example, Fig. S2 illustrates the δ*(O–I) and LP(O) orbitals involved in the (I2O4)1(HIO2)1 dimer. Therefore, although I2O5 has more electronically deficient iodine centers than I2O4/I2O2, the additional electrostatic interaction caused by the proton transfer reaction between I2O4/I2O2 and HIO3, and the stronger average strength of two XBs between I2O4/I2O2 and HIO2, make I2O4/I2O2 form a more stable complex with iodine oxoacids compared to I2O5.

As discussed above, I2O4 is expected to have the highest intrinsic nucleation potential towards iodine oxoacids. In addition, in photochemical reactions of I2 or CH3I with ozone (O3) under laboratory conditions, the gaseous concentration of I2O4 produced was higher than that of all other iodine oxides except IO.15,60 Therefore, I2O4 has the greatest nucleation potential towards iodine oxoacids. In the following section, we describe our detailed investigation of the joint nucleation of I2O4 and iodine oxoacids.

Configurations of HIO3–HIO2–I2O4 clusters featuring halogen-induced acid–base reaction

We present the cluster configurations for the HIO3–HIO2–I2O4 system in Fig. 2 and S3. Since HIO3–HIO2 configurations have already been investigated in our previous study,41 we mainly focus on HIO3–I2O4 and HIO2–I2O4 two-component clusters (Fig. S3) and HIO3–HIO2–I2O4 three-component clusters (Fig. 2) here. Observation and analysis of the cluster configurations shows that I2O4 accepts a proton from HIO3 and HIO2 in almost all HIO3–I2O4 and HIO2–I2O4 two-component clusters. Therefore, I2O4 behaves as a Brønsted–Lowry base when it clusters with HIO3 or HIO2. To our knowledge, this is the first time that I2O4 has been revealed to behave as a base in clustering with HIO3 or HIO2. In HIO3–HIO2–I2O4 three-component clusters, only HIO3, with the greater acidity, can donate a proton, and I2O4 or HIO2 can accept a proton. In addition, I2O4 has a greater ability to accept a proton from HIO3 than HIO2 does. Therefore, the behaviour of I2O4 as a base is a common feature of HIO3–I2O4 and HIO2–I2O4 two-component clusters and HIO3–HIO2–I2O4 three-component clusters, in addition to the formation of typical XBs and HBs.
image file: d5sc02517f-f2.tif
Fig. 2 Global minimum configurations of the (HIO3)x(HIO2)y(I2O4)z (x = 1–3, y + z = 2–3) clusters calculated at the DLPNO-CCSD(T)/aug-cc-pVTZ(-PP)//M06-2X/aug-cc-pVTZ(-PP) level of theory. The red balls represent oxygen atoms, purple ones are for iodine atoms, and white ones are for hydrogen atoms. The red dashed lines represent HBs, and the blue dashed lines represent XBs. The red and blue numbers indicate the bond lengths of HBs and XBs, respectively. The bond lengths are given in Å.

Why I2O4 can behave as a base towards HIO2–3 is an interesting topic for discussion. The calculated gas basicity (GB) of I2O4 is 199 kcal mol−1, close to that of NH3 (196 kcal mol−1), which has been found to be unable to accept a proton from HIO3 in the (HIO3)1(NH3)1 dimer.17,18 Therefore, the GB of I2O4 cannot explain its acid–base reaction with HIO2 and HIO3. According to observation of the configurations of HIO3–HIO2–I2O4 clusters, the occurrence of acid–base reactions with I2O4 as a base is always accompanied by the formation of XBs. Therefore, it is reasonable to speculate that the formation of XBs can induce the basicity enhancement of I2O4, further driving the acid–base reaction.

To verify this speculation, ΔESP for the basic site and GB of I2O4 when it forms XBs with various precursors, were determined. Here, the selected precursors included amines, carbonyl compounds, sulfides, alcohols, benzothiazole, peroxides and hydroperoxymethyl, which can form XBs with I2O4 at various bond strengths. Since I2O4 can form two types of XBs via its two lowest energy antibonding orbitals δ*(O–I), the effects of these two types of XBs on ΔESP and GB were considered. Fig. 3A shows that the ΔESP values for the O atom as a basic site are negative when I2O4 forms a XB with precursors (identity of selected precursors was presented in Table S4), which leads to an increased basicity of I2O4. Moreover, a higher XB energy generally results in a greater (more negative) ΔESP value. As shown in Fig. 3B, the GB of I2O4–precursor dimer clusters is significantly greater than that of the I2O4 monomer regardless of the type of XBs. In addition, a roughly positive correlation between the formed XB energy and the GB values of I2O4–precursor dimer clusters is observed. More importantly, even when a relatively weak XB (approximately 13 kcal mol−1) is formed, the GB of I2O4 can be increased to 215 kcal mol−1, which is close to that of dimethylamine (DMA) (215 kcal mol−1), a typical base for enhancing SA-driven nucleation.57 Therefore, the formation of XBs indeed increases the basicity of I2O4, increasing it from a level similar to that of NH3 to a level that is comparable to or even greater than that of DMA.


image file: d5sc02517f-f3.tif
Fig. 3 Change in electrostatic potential (ΔESP) values (kcal mol−1) for the O atom as a basic site of I2O4 (A) and the gas basicity (GB) (kcal mol−1) of precursor–I2O4 dimers (B) as a function of the halogen bond energy (kcal mol−1). The purple and red spheres represent two different types of XBs. O-atom 1 is the basic site of I2O4 when it forms XB type 1 with precursors. O-atom 2 is the basic site of I2O4 when it forms XB type 2 with precursors.

Notably, the present study and an earlier study also revealed that HIO2 behaves as a base towards HIO3.41 Similarly, the ΔESP for the basic site of HIO2 when it forms XBs with various precursors was determined. An effect similar to that of the formed XB on the ΔESP for the terminal O atom as a basic site was also found for the HIO2 system (Fig. S4). Therefore, the formation of XBs can greatly increase the basicity of I2O4 and HIO2. To our knowledge, this is the first study to reveal that XBs induce acid–base reactions in HIO3–HIO2 and HIO3–HIO2–I2O4 nucleation systems.

Since XBs are weak non-covalent interactions, it is interesting to check whether other weak non-covalent interactions like HBs, commonly formed in the acid–base nucleation system, can also induce the basicity enhancement of compounds. Here, possible basicity enhancement effect of the HBs was checked by taking NH3 as the test case. It was found that HBs indeed can induce the basicity enhancement of NH3 (see ΔESP values for the N atom as the basic site of NH3 when it forms HBs with precursors in Table S5). Still, when XBs and HBs are considered as non-covalent interactions, the correlation of non-covalent interaction energy with ΔESP is similar to that of XB energy with ΔESP, i.e. a higher non-covalent interaction energy generally results in a greater (more negative) ΔESP value (Fig. S5). Since the HB energy is weaker than XB energy considered here, the basicity enhancement effect of HBs is much lower than that of the XBs (Fig. S5). If the stronger or more HBs are formed in the SA–base systems, the basicity enhancement effect of HBs could be improved. In view of potential role of HBs in inducing basicity enhancement, future investigation on effect of HB-induced basicity enhancement in different nucleation systems is warranted.

Low formation free energy of I2O4-containing clusters

The cluster formation free energy of the HIO3–HIO2–I2O4 system at 278.15 K and 1 atm is presented in Fig. S6. As shown in Fig. S6, the ΔG values of all HIO3–I2O4 and HIO2–I2O4 two-component clusters are lower than those of the corresponding HIO3–HIO2 two-component clusters. The ΔG values of all I2O4-rich clusters are lower than those of the corresponding HIO3 or HIO2-rich clusters in these two-component clusters. In addition, the ΔG values of the I2O4-containing three-component clusters are lower than those of the corresponding HIO3–HIO2 two-component clusters with the same number of acids (HIO3) and bases (HIO2 and I2O4). All of the aforementioned results indicate that I2O4 has a greater ability to bind with HIO2 or HIO3 than HIO3 and HIO2 do with each other, which is consistent with the conclusion from the ΔG analysis of the dimers of I2O4 with HIO2 and HIO3. The higher binding ability of I2O4 leads to the high stability of most I2O4-containing clusters, as evidenced by their low evaporation rates (Fig. S7). Therefore, the participation of I2O4 in HIO3–HIO2 nucleation may be favorable. It is noteworthy that (HIO2)3(I2O4)3 and (HIO2)3(I2O4)2 clusters own high evaporation rates despite they have low ΔG values. It is understandable since the evaporation rate is not only dependent on the ΔG value of the individual cluster, but also on the stability of its all possible daughter clusters.54 The high evaporation rate of (HIO2)3(I2O4)3 and (HIO2)3(I2O4)2 clusters result from the higher stability of their daughter clusters (e.g. (HIO2)2(I2O4)2). Tables S6 and S7 also display the ΔG values of the HIO3–HIO2–I2O4 system clusters at other temperatures and pressures. The ΔG values of most clusters become more negative with decreasing temperature, while they become higher with decreasing pressure.

Previous experimental and theoretical studies have shown that HBs are favorable for new particle formation, particularly for those involving organic acids.61,62 Herein, we also compared the effects of HBs and XBs on cluster formation by examining the formation free energy of two types of specific dimers: organic acids (benzoic acid, cis-pinonic acid, formic acid)–NH3 dimers formed by HBs,62 and HIO2–I2O4 and HIO2–HIO2 dimers formed by XBs.41 It is found that the formation free energies of HIO2–I2O4 and HIO2–HIO2 dimers are lower than those of organic acids (benzoic acid, cis-pinonic acid, formic acid)–NH3 dimers (Table S8). This implies that XBs here are stronger than HBs.

Cluster formation rates of the HIO3–HIO2–I2O4 system

The variation in the cluster formation rates (J) of the three-component HIO3–HIO2–I2O4 system with [HIO3] (105–108 cm−3) ([HIO2] = 1/30 [HIO3], I2O4 = 1/100 [HIO3]) at 278.15 K, 1 atm and kcoag = 0.002 s−1, along with that of the two-component HIO3–HIO2 system, is presented in Fig. 4A. The detailed computational methods for J are presented in ESI. Note that pure I2O4 nucleation is not compared with three-component nucleation since its rate is much lower than that of HIO3–HIO2 nucleation when [I2O4] = 1/100 [HIO3] (Fig. S8). As shown in Fig. 4A, the J values of the HIO3–HIO2–I2O4 system increase with [HIO3]. The J values of the HIO3–HIO2–I2O4 system are higher than those of the two-component HIO3–HIO2 system under the studied conditions, especially at low [HIO3]. Therefore, I2O4 can enhance HIO3–HIO2 nucleation. To clearly demonstrate the enhancement effect of I2O4, the enhancement coefficient (RI2O4), which is the ratio of the J value of the HIO3–HIO2–I2O4 system relative to that of the HIO3–HIO2 system, was calculated. The calculated RI2O4 ranges from 1.5 to 6.8 at 278.15 K, depending on [HIO3]. The lower [HIO3] is, the higher RI2O4 is. The RI2O4 of 6.8 applies when the absolute J is very low and the maximum RI2O4 for non-negligible absolute J (around 7.5 cm−3 s−1) is about 2–3. The higher enhancement effect of I2O4 at low [HIO3] decreases the lower limit concentration that HIO3-driven nucleation has a pronounced nucleation rate. The required [HIO3] values are 3.45 × 106 cm−3 and 4.35 × 106 cm−3 for J values of approximately 1 cm−3 s−1 in the systems with and without I2O4, respectively. This provides further support for the idea that iodine oxides play a key role in iodine oxoacid nucleation. We also noted that a very recent study by Engsvang and Elm found “oxoacid-assisted oxide” nucleation efficient.63 In their study, it was found that the predicted rate of HIO3–HIO2 nucleation is much lower than the experimental rates based on their selected theoretical method.63 In our early studies, the predicted HIO3–HIO2 and SA–HIO3–HIO2 nucleation rates using the same theoretical methods as in this study are comparable to that from CLOUD data.26,41 Besides, the concentration of iodine oxide is much lower than that of iodine oxoacid in the ambient atmosphere.38 Hence, we treat HIO3–HIO2–I2O4 synergistic nucleation as “oxide-assisted oxoacid” here. In fact, both studies found that iodine oxoacid and iodine oxide can nucleate together.
image file: d5sc02517f-f4.tif
Fig. 4 Cluster formation rates (J) (cm−3 s−1) of the HIO3–HIO2 and HIO3–HIO2–I2O4 systems and enhancement coefficient (RI2O4) (the ratio of the J value of the HIO3–HIO2–I2O4 system relative to that of the HIO3–HIO2 system) (A) and comparison coefficient (r) (the ratio of the J value of the HIO3–HIO2–I2O4 system relative to that of the SA–HIO3–HIO2 system) (B) as a function of precursor concentration at 278.15 K, 1 atm and kcoag = 0.002 s−1.

The J of HIO3–HIO2–I2O4 system and RI2O4 as a function of temperature at 1 atm are presented in Fig. S9. J values rise as temperature declines, while further decreasing temperature exhibits a diminished effect on J below 253 K. The temperature dependence of J is more pronounced at lower [HIO3]. As shown in Fig. S9B, enhancement of I2O4 on HIO3–HIO2 nucleation becomes higher at higher temperature (298.15 K). Fig. S10 illustrates the J and RI2O4 as a function of pressure at 278.15 K. J of HIO3–HIO2–I2O4 system and RI2O4 increase slightly as pressure decreases from 1 to 0.5 atm but remain nearly constant when pressure declines from 0.5 to 0.1 atm.

A recent study revealed that SA can also enhance HIO3–HIO2 nucleation.26 Therefore, comparing the enhancement potential of I2O4 with that of SA would be interesting. The variation in the comparison coefficient (r), which is the ratio of the J value of the HIO3–HIO2–I2O4 system relative to that of the SA–HIO3–HIO2 system, with [HIO3] and [SA], is presented in Fig. 4B. At low [HIO3] and therefore lower [I2O4], r is much less than 1, indicating that the enhancement potential of I2O4 is lower than that of SA. When [HIO3] ≥ 5 × 106 cm−3 and the corresponding [I2O4] ≥ 5 × 104 cm−3, r becomes approximately 1, i.e., 1.12–1.50 for [SA] = 1 × 106 cm−3; 0.20–1.38 for [SA] = 5 × 106 cm−3, indicating that the enhancement potential of I2O4 is comparable to or even greater than that of SA for HIO3–HIO2 nucleation. In most cases for which r is approximately 1, [I2O4] is much lower than [SA]. Therefore, I2O4, as a base, has a greater enhancement efficiency than SA for HIO3–HIO2 nucleation.

Synergistic nucleation mechanism of the HIO3–HIO2–I2O4 system

The cluster growth pathway for the HIO3–HIO2–I2O4 system is shown in Fig. 5. Overall, the cluster growth pathway involves three channels: the (I) HIO3–HIO2 pathway, (II) HIO3–I2O4 pathway and (III) HIO3–HIO2–I2O4 pathway. The contributions of the two-component pathways (HIO3–HIO2 and HIO3–I2O4) are 30% and 12%, respectively, and the contribution of the HIO3–HIO2–I2O4 pathway is 38%. At the very early cluster formation stage, two-component clusters of HIO3–HIO2 and HIO3–I2O4 evolved into three-component HIO3–HIO2–I2O4 clusters. Note that almost all clusters growing out of the box had O/I ratios ranging from 2.3 to 2.5, which is consistent with the O/I ratio of approximately 2.5 reported in a previous laboratory study,15 supporting the occurrence of the proposed synergistic nucleation mechanism of iodine oxoacids and iodine oxides in the atmosphere.
image file: d5sc02517f-f5.tif
Fig. 5 Cluster formation pathways for the HIO3–HIO2–I2O4 system at 278.15 K, [HIO3] = 1 × 107 cm−3, [HIO2] = 3.33 × 105 cm−3, [I2O4] = 1 × 105 cm−3 and kcoag = 0.002 s−1. The pathways contributing less than 15% to the flux of cluster formation are not show.

Implications

This study reveals that I2O4, HIO3 and HIO2 can nucleate together, breaking through the previous findings that iodine oxoacids and iodine oxides independently nucleate.14,38 More importantly, I2O4, HIO3 and HIO2 nucleate in a synergistic way. This synergistic nucleation increases the contribution of iodine-containing species to NPF beyond what was previously thought. Fig. 6 shows the significance of HIO3–HIO2–I2O4 nucleation on a global scale. It is noted that HIO3–HIO2 system was selected as the base case, since HIO3–HIO2 is the only two-component system that was verified to have high nucleation rate at the realistic atmospheric concentration ranges of HIO2 and HIO3 (ref. 38) and has the highest nucleation rate among three two-component systems in most conditions. As shown in Fig. 6, the nucleation rate increased by at least 1.5 times in areas where HIO3 was detected once I2O4 was involved in nucleation. Under conditions of relatively high temperature and 107 cm−3 HIO3 in Helsinki, the nucleation rate can be increased by up to 23.3 times. In addition, I2O4 enhances HIO3–HIO2 nucleation as a base. Atmospheric acids such as SA and methanesulfonic acid (MSA) have been proposed to enhance HIO3–HIO2 nucleation.25,26,64,65 Therefore, multicomponent nucleation involving HIO2, HIO3, I2O4, SA, MSA and other acids can further increase the contribution of iodine-containing species to NPF, which deserves future ambient field and laboratory study.
image file: d5sc02517f-f6.tif
Fig. 6 Cluster formation rate (J) (cm−3 s−1) of the HIO3–HIO2–I2O4 and HIO3–HIO2 systems and the enhancement coefficient (RI2O4) in different areas with different temperatures (T) and [HIO3]. The T and [HIO3] in different areas were obtained from He et al.38 The kcoag in Helsinki and Mace Head was set to 2 × 10−3 s−1,55 and kcoag in Aboa, Greenland, and Ny-Ålesund was set to 1 × 10−4 s−1.38

This study revealed that XBs can induce the basicity enhancement of I2O4 and HIO2, further driving their acid–base reactions and therefore facilitating nucleation. Since the formation of XBs is common in iodine and oxygen-containing clusters, the mechanism by which XBs induce basicity enhancement could also be extended to other nucleation systems containing iodine oxoacids and iodine oxides. We found that XBs even can induce the basicity enhancement of more acidic HIO3 in our preliminary study. Therefore, the mechanism by which XBs induce basicity enhancement is expected to be a common mechanism in nucleation systems containing iodine oxoacids and iodine oxides, necessitating further study. In addition, XB-induced basicity enhancement is also expected to occur during the growth of iodine oxoacids and iodine oxides-containing particles. However, more species and more complicated interactions should be involved in particle growth. The possible role of such a mechanism in particle growth is worthy of further investigation.

Conclusions

In this study, quantum chemical methods and ACDC were employed to investigate the nucleation mechanism and kinetics of iodine oxides and iodine oxoacids. Of all iodine oxides, I2O4 was found to have the strongest nucleation potential towards HIO3–HIO2. I2O4 can synergistically nucleate with HIO3–HIO2, breaking through the previous findings that iodine oxoacids and iodine oxides independently nucleate.14,38 The synergistic nucleation rate of HIO3–HIO2–I2O4 is 1.5 to 6.8 times higher than that of the known most efficient iodine-associated two-component (HIO3–HIO2) nucleation at 278.15 K. The high synergistic nucleation rate enhances the role of iodine-containing species in marine atmospheric particle formation. Microscopic analysis of the three-component cluster configurations revealed that an unexpected acid–base reaction between I2O4 and HIO2/HIO3 is a key driver of this efficient synergistic nucleation, in addition to traditional HBs and XBs. The halogen bond-induced basicity enhancement was further identified as the chemical nature of I2O4 behaving as a base in the nucleation with HIO2 or HIO3. Such a basicity enhancement effect can be extended to other iodine-containing species, e.g., HIO2 and even more acidic HIO3, suggesting that this is a common feature in interactions between iodine-containing species. Our findings clarify the synergistic nucleation of iodine oxoacids and iodine oxides and highlight the importance of halogen bond-induced basicity enhancement in the formation of iodine-containing particles.

Data availability

The data supporting this article have been included as part of the ESI.

Author contributions

H. B. X. and J. S. F. designed the study. R. J. Z. and Y. Y. L. performed the quantum chemical calculation and analyzed data. R. J. Z., Y. Y. L., R. J. Y., F. F. M., H. B. X. and J. S. F. wrote the manuscript. D. M. X., J. W. C., H. B. X. and J. S. F. commented on and revised the manuscript. All coauthors participated in relevant scientific discussion of the manuscript.

Conflicts of interest

The authors declare no competing financial interest.

Acknowledgements

This work was supported by the National Key Research and Development Program of China (2022YFC3701000, Task 1) and the National Natural Science Foundation of China (22236004, 22176022 and 22406017).

Notes and references

  1. J. Merikanto, D. V. Spracklen, G. W. Mann, S. J. Pickering and K. S. Carslaw, Impact of nucleation on global CCN, Atmos. Chem. Phys., 2009, 9, 8601–8616 CrossRef CAS.
  2. S.-H. Lee, H. Gordon, H. Yu, K. Lehtipalo, R. Haley, Y. Li and R. Zhang, New particle formation in the atmosphere: from molecular clusters to global climate, J. Geophys. Res.:Atmos., 2019, 124, 7098–7146 CrossRef CAS.
  3. F. Yu and G. Luo, 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 CrossRef CAS.
  4. H. Gordon, J. Kirkby, U. Baltensperger, F. Bianchi, M. Breitenlechner, J. Curtius, A. Dias, J. Dommen, N. M. Donahue, E. M. Dunne, J. Duplissy, S. Ehrhart, R. C. Flagan, C. Frege, C. Fuchs, A. Hansel, C. R. Hoyle, M. Kulmala, A. Kürten, K. Lehtipalo, V. Makhmutov, U. Molteni, M. P. Rissanen, Y. Stozkhov, J. Tröstl, G. Tsagkogeorgas, R. Wagner, C. Williamson, D. Wimmer, P. M. Winkler, C. Yan and K. S. Carslaw, Causes and importance of new particle formation in the present-day and preindustrial atmospheres, J. Geophys. Res.:Atmos., 2017, 122, 8739–8760 CrossRef.
  5. J. Haywood and O. Boucher, Estimates of the direct and indirect radiative forcing due to tropospheric aerosols: A review, Rev. Geophys., 2000, 38, 513–543 CrossRef CAS.
  6. E. Gryspeerdt, J. Quaas, S. Ferrachat, A. Gettelman, S. Ghan, U. Lohmann, H. Morrison, D. Neubauer, D. G. Partridge, P. Stier, T. Takemura, H. Wang, M. Wang and K. Zhang, Constraining the instantaneous aerosol influence on cloud albedo, Proc. Natl. Acad. Sci. U. S. A., 2017, 114, 4899–4904 CrossRef CAS PubMed.
  7. R. J. Charlson, J. E. Lovelock, M. O. Andreae and S. G. Warren, Oceanic Phytoplankton, Atmospheric Sulphur, Cloud Albedo and Climate, Nature, 1987, 326, 655–661 CrossRef CAS.
  8. C. D. O'Dowd and G. de Leeuw, Marine aerosol production: a review of the current knowledge, Philos. Trans. R. Soc., A, 2007, 365, 1753–1774 CrossRef.
  9. N. Takegawa, T. Seto, N. Moteki, M. Koike, N. Oshima, K. Adachi, K. Kita, A. Takami and Y. Kondo, Enhanced New Particle Formation above the Marine Boundary Layer over the Yellow Sea: Potential Impacts on Cloud Condensation Nuclei, J. Geophys. Res.:Atmos., 2020, 125, e2019JD031448 CrossRef.
  10. G. Zheng, Y. Wang, R. Wood, M. P. Jensen, C. Kuang, I. L. McCoy, A. Matthews, F. Mei, J. M. Tomlinson, J. E. Shilling, M. A. Zawadowicz, E. Crosbie, R. Moore, L. Ziemba, M. O. Andreae and J. Wang, New particle formation in the remote marine boundary layer, Nat. Commun., 2021, 12, 527 CrossRef CAS PubMed.
  11. E. M. Dunne, H. Gordon, A. Kürten, J. Almeida, J. Duplissy, C. Williamson, I. K. Ortega, K. J. Pringle, A. Adamov, U. Baltensperger, P. Barmet, F. Benduhn, F. Bianchi, M. Breitenlechner, A. Clarke, J. Curtius, J. Dommen, N. M. Donahue, S. Ehrhart, R. C. Flagan, A. Franchin, R. Guida, J. Hakala, A. Hansel, M. Heinritzi, T. Jokinen, J. Kangasluoma, J. Kirkby, M. Kulmala, A. Kupc, M. J. Lawler, K. Lehtipalo, V. Makhmutov, G. Mann, S. Mathot, J. Merikanto, P. Miettinen, A. Nenes, A. Onnela, A. Rap, C. L. S. Reddington, F. Riccobono, N. A. D. Richards, M. P. Rissanen, L. Rondo, N. Sarnela, S. Schobesberger, K. Sengupta, M. Simon, M. Sipilä, J. N. Smith, Y. Stozkhov, A. Tomé, J. Tröstl, P. E. Wagner, D. Wimmer, P. M. Winkler, D. R. Worsnop and K. S. Carslaw, Global atmospheric particle formation from CERN CLOUD measurements, Science, 2016, 354, 1119–1124 CrossRef CAS PubMed.
  12. T. Jokinen, M. Sipilä, J. Kontkanen, V. Vakkari, P. Tisler, E.-M. Duplissy, H. Junninen, J. Kangasluoma, H. E. Manninen, T. Petäjä, M. Kulmala, D. R. Worsnop, J. Kirkby, A. Virkkula and V.-M. Kerminen, Ion-induced sulfuric acid–ammonia nucleation drives particle formation in coastal Antarctica, Sci. Adv., 2018, 4, eaat9744 CrossRef CAS PubMed.
  13. M. Sipilä, N. Sarnela, T. Jokinen, H. Henschel, H. Junninen, J. Kontkanen, S. Richters, J. Kangasluoma, A. Franchin, O. Peräkylä, M. P. Rissanen, M. Ehn, H. Vehkamäki, T. Kurten, T. Berndt, T. Petäjä, D. Worsnop, D. Ceburnis, V.-M. Kerminen, M. Kulmala and C. O'Dowd, Molecular-scale evidence of aerosol particle formation via sequential addition of HIO3, Nature, 2016, 537, 532–534 CrossRef.
  14. J. C. Gómez Martín, T. R. Lewis, M. A. Blitz, J. M. C. Plane, M. Kumar, J. S. Francisco and A. Saiz-Lopez, A gas-to-particle conversion mechanism helps to explain atmospheric particle formation through clustering of iodine oxides, Nat. Commun., 2020, 11, 4521 CrossRef.
  15. J. C. Gómez Martín, O. Gálvez, M. T. Baeza-Romero, T. Ingham, J. M. C. Plane and M. A. Blitz, On the mechanism of iodine oxide particle formation, Phys. Chem. Chem. Phys., 2013, 15, 15612–15622 RSC.
  16. J. C. Gómez Martín, T. R. Lewis, A. D. James, A. Saiz-Lopez and J. M. C. Plane, Insights into the Chemistry of Iodine New Particle Formation: The Role of Iodine Oxides and the Source of Iodic Acid, J. Am. Chem. Soc., 2022, 144, 9240–9253 CrossRef.
  17. H. Rong, J. Liu, Y. Zhang, L. Du, X. Zhang and Z. Li, Nucleation mechanisms of iodic acid in clean and polluted coastal regions, Chemosphere, 2020, 253, 126743 CrossRef CAS PubMed.
  18. D. Xia, J. Chen, H. Yu, H.-b. Xie, Y. Wang, Z. Wang, T. Xu and D. T. Allen, Formation Mechanisms of Iodine–Ammonia Clusters in Polluted Coastal Areas Unveiled by Thermodynamics and Kinetic Simulations, Environ. Sci. Technol., 2020, 54, 9235–9242 CrossRef CAS PubMed.
  19. A. Ning, L. Liu, L. Ji and X. Zhang, Molecular-level nucleation mechanism of iodic acid and methanesulfonic acid, Atmos. Chem. Phys., 2022, 22, 6103–6114 CrossRef CAS.
  20. S. Zhang, S. Li, A. Ning, L. Liu and X. Zhang, Iodous Acid – A More Efficient Nucleation Precursor than Iodic Acid, Phys. Chem. Chem. Phys., 2022, 24, 13651–13660 RSC.
  21. A. Ning, L. Liu, S. Zhang, F. Yu, L. Du, M. Ge and X. Zhang, The critical role of dimethylamine in the rapid formation of iodic acid particles in marine areas, npj Clim. Atmos. Sci., 2022, 5, 92 CrossRef CAS.
  22. N. Wu, A. Ning, L. Liu, H. Zu, D. Liang and X. Zhang, Methanesulfonic Acid and Iodous Acid Nucleation: A Novel Mechanism of Marine Aerosols, Phys. Chem. Chem. Phys., 2023, 25, 16745–16752 RSC.
  23. J. B. Burkholder, J. Curtius, A. R. Ravishankara and E. R. Lovejoy, Laboratory studies of the homogeneous nucleation of iodine oxides, Atmos. Chem. Phys., 2004, 4, 19–34 CrossRef CAS.
  24. F. Ma, H.-B. Xie, R. Zhang, L. Su, Q. Jiang, W. Tang, J. Chen, M. Engsvang, J. Elm and X.-C. He, Enhancement of Atmospheric Nucleation Precursors on Iodic Acid-Induced Nucleation: Predictive Model and Mechanism, Environ. Sci. Technol., 2023, 57, 6944–6954 CrossRef CAS PubMed.
  25. R. Zhang, F. Ma, Y. Zhang, J. Chen, J. Elm, X.-C. He and H.-B. Xie, HIO3–HIO2-Driven Three-Component Nucleation: Screening Model and Cluster Formation Mechanism, Environ. Sci. Technol., 2024, 58, 649–659 CrossRef CAS PubMed.
  26. X.-C. He, M. Simon, S. Iyer, H.-B. Xie, B. Rörup, J. Shen, H. Finkenzeller, D. Stolzenburg, R. Zhang, A. Baccarini, Y. J. Tham, M. Wang, S. Amanatidis, A. A. Piedehierro, A. Amorim, R. Baalbaki, Z. Brasseur, L. Caudillo, B. Chu, L. Dada, J. Duplissy, I. El Haddad, R. C. Flagan, M. Granzin, A. Hansel, M. Heinritzi, V. Hofbauer, T. Jokinen, D. Kemppainen, W. Kong, J. Krechmer, A. Kürten, H. Lamkaddam, B. Lopez, F. Ma, N. G. A. Mahfouz, V. Makhmutov, H. E. Manninen, G. Marie, R. Marten, D. Massabò, R. L. Mauldin, B. Mentler, A. Onnela, T. Petäjä, J. Pfeifer, M. Philippov, A. Ranjithkumar, M. P. Rissanen, S. Schobesberger, W. Scholz, B. Schulze, M. Surdu, R. C. Thakur, A. Tomé, A. C. Wagner, D. Wang, Y. Wang, S. K. Weber, A. Welti, P. M. Winkler, M. Zauner-Wieczorek, U. Baltensperger, J. Curtius, T. Kurtén, D. R. Worsnop, R. Volkamer, K. Lehtipalo, J. Kirkby, N. M. Donahue, M. Sipilä and M. Kulmala, Iodine oxoacids enhance nucleation of sulfuric acid particles in the atmosphere, Science, 2023, 382, 1308–1314 CrossRef CAS PubMed.
  27. R.-J. Huang, T. Hoffmann, J. Ovadnevaite, A. Laaksonen, H. Kokkola, W. Xu, W. Xu, D. Ceburnis, R. Zhang, J. H. Seinfeld and C. O'Dowd, Heterogeneous iodine-organic chemistry fast-tracks marine new particle formation, Proc. Natl. Acad. Sci. U. S. A., 2022, 119, e2201729119 CrossRef CAS PubMed.
  28. R. W. Saunders and J. M. C. Plane, Formation Pathways and Composition of Iodine Oxide Ultra-Fine Particles, Environ. Chem., 2005, 2, 299–303 CrossRef CAS.
  29. A. Saiz-Lopez, J. M. C. Plane, A. R. Baker, L. J. Carpenter, R. von Glasow, J. C. Gómez Martín, G. McFiggans and R. W. Saunders, Atmospheric Chemistry of Iodine, Chem. Rev., 2012, 112, 1773–1804 CrossRef CAS PubMed.
  30. R. W. Saunders, R. Kumar, J. C. G. Martín, A. S. Mahajan, B. J. Murray and J. M. C. Plane, Studies of the Formation and Growth of Aerosol from Molecular Iodine Precursor, Z. Phys. Chem., 2010, 224, 1095–1117 CrossRef CAS.
  31. C. D. O'Dowd, J. L. Jimenez, R. Bahreini, R. C. Flagan, J. H. Seinfeld, K. Hämeri, L. Pirjola, M. Kulmala, S. G. Jennings and T. Hoffmann, Marine Aerosol Formation from Biogenic Iodine Emissions, Nature, 2002, 417, 632–636 CrossRef PubMed.
  32. L. J. Carpenter, S. M. MacDonald, M. D. Shaw, R. Kumar, R. W. Saunders, R. Parthipan, J. Wilson and J. M. C. Plane, Atmospheric Iodine Levels Influenced by Sea Surface Emissions of Inorganic Iodine, Nat. Geosci., 2013, 6, 108–111 CrossRef CAS.
  33. L. J. Carpenter, Iodine in the Marine Boundary Layer, Chem. Rev., 2003, 103, 4953–4962 CrossRef CAS PubMed.
  34. H. Yu, L. Ren, X. Huang, M. Xie, J. He and H. Xiao, Iodine speciation and size distribution in ambient aerosols at a coastal new particle formation hotspot in China, Atmos. Chem. Phys., 2019, 19, 4025–4039 CrossRef CAS.
  35. R. C. Thakur, L. Dada, L. J. Beck, L. L. J. Quéléver, T. Chan, M. Marbouti, X. C. He, C. Xavier, J. Sulo, J. Lampilahti, M. Lampimäki, Y. J. Tham, N. Sarnela, K. Lehtipalo, A. Norkko, M. Kulmala, M. Sipilä and T. Jokinen, An evaluation of new particle formation events in Helsinki during a Baltic Sea cyanobacterial summer bloom, Atmos. Chem. Phys., 2022, 22, 6365–6391 CrossRef CAS.
  36. A. Baccarini, L. Karlsson, J. Dommen, P. Duplessis, J. Vüllers, I. M. Brooks, A. Saiz-Lopez, M. Salter, M. Tjernström, U. Baltensperger, P. Zieger and J. Schmale, Frequent New Particle Formation over the High Arctic Pack Ice by Enhanced Iodine Emissions, Nat. Commun., 2020, 11, 4924 CrossRef CAS PubMed.
  37. L. J. Beck, N. Sarnela, H. Junninen, C. J. M. Hoppe, O. Garmash, F. Bianchi, M. Riva, C. Rose, O. Peräkylä, D. Wimmer, O. Kausiala, T. Jokinen, L. Ahonen, J. Mikkilä, J. Hakala, X.-C. He, J. Kontkanen, K. K. E. Wolf, D. Cappelletti, M. Mazzola, R. Traversi, C. Petroselli, A. P. Viola, V. Vitale, R. Lange, A. Massling, J. K. Nøjgaard, R. Krejci, L. Karlsson, P. Zieger, S. Jang, K. Lee, V. Vakkari, J. Lampilahti, R. C. Thakur, K. Leino, J. Kangasluoma, E.-M. Duplissy, E. Siivola, M. Marbouti, Y. J. Tham, A. Saiz-Lopez, T. Petäjä, M. Ehn, D. R. Worsnop, H. Skov, M. Kulmala, V.-M. Kerminen and M. Sipilä, Differing Mechanisms of New Particle Formation at Two Arctic Sites, Geophys. Res. Lett., 2021, 48, e2020GL091334 CrossRef CAS.
  38. X.-C. He, Y. J. Tham, L. Dada, M. Wang, H. Finkenzeller, D. Stolzenburg, S. Iyer, M. Simon, A. Kürten, J. Shen, B. Rörup, M. Rissanen, S. Schobesberger, R. Baalbaki, D. S. Wang, T. K. Koenig, T. Jokinen, N. Sarnela, L. J. Beck, J. Almeida, S. Amanatidis, A. Amorim, F. Ataei, A. Baccarini, B. Bertozzi, F. Bianchi, S. Brilke, L. Caudillo, D. Chen, R. Chiu, B. Chu, A. Dias, A. Ding, J. Dommen, J. Duplissy, I. El Haddad, L. Gonzalez Carracedo, M. Granzin, A. Hansel, M. Heinritzi, V. Hofbauer, H. Junninen, J. Kangasluoma, D. Kemppainen, C. Kim, W. Kong, J. E. Krechmer, A. Kvashin, T. Laitinen, H. Lamkaddam, C. P. Lee, K. Lehtipalo, M. Leiminger, Z. Li, V. Makhmutov, H. E. Manninen, G. Marie, R. Marten, S. Mathot, R. L. Mauldin, B. Mentler, O. Möhler, T. Müller, W. Nie, A. Onnela, T. Petäjä, J. Pfeifer, M. Philippov, A. Ranjithkumar, A. Saiz-Lopez, I. Salma, W. Scholz, S. Schuchmann, B. Schulze, G. Steiner, Y. Stozhkov, C. Tauber, A. Tomé, R. C. Thakur, O. Väisänen, M. Vazquez-Pufleau, A. C. Wagner, Y. Wang, S. K. Weber, P. M. Winkler, Y. Wu, M. Xiao, C. Yan, Q. Ye, A. Ylisirniö, M. Zauner-Wieczorek, Q. Zha, P. Zhou, R. C. Flagan, J. Curtius, U. Baltensperger, M. Kulmala, V.-M. Kerminen, T. Kurtén, N. M. Donahue, R. Volkamer, J. Kirkby, D. R. Worsnop and M. Sipilä, Role of iodine oxoacids in atmospheric aerosol nucleation, Science, 2021, 371, 589–595 CrossRef CAS PubMed.
  39. V. Oliveira, E. Kraka and D. Cremer, The intrinsic strength of the halogen bond: electrostatic and covalent contributions described by coupled cluster theory, Phys. Chem. Chem. Phys., 2016, 18, 33031–33046 RSC.
  40. M. Kumar, A. Saiz-Lopez and J. S. Francisco, Single-Molecule Catalysis Revealed: Elucidating the Mechanistic Framework for the Formation and Growth of Atmospheric Iodine Oxide Aerosols in Gas-Phase and Aqueous Surface Environments, J. Am. Chem. Soc., 2018, 140, 14704–14716 CrossRef CAS PubMed.
  41. R. Zhang, H.-B. Xie, F. Ma, J. Chen, S. Iyer, M. Simon, M. Heinritzi, J. Shen, Y. J. Tham, T. Kurtén, D. R. Worsnop, J. Kirkby, J. Curtius, M. Sipilä, M. Kulmala and X.-C. He, Critical Role of Iodous Acid in Neutral Iodine Oxoacid Nucleation, Environ. Sci. Technol., 2022, 56, 14166–14177 CrossRef CAS PubMed.
  42. J. Elm, Clusteromics II: Methanesulfonic Acid–Base Cluster Formation, ACS Omega, 2021, 6, 17035–17044 CrossRef CAS PubMed.
  43. B. Temelso, E. F. Morrison, D. L. Speer, B. C. Cao, N. Appiah-Padi, G. Kim and G. C. Shields, Effect of Mixing Ammonia and Alkylamines on Sulfate Aerosol Formation, J. Phys. Chem. A, 2018, 122, 1612–1622 CrossRef CAS PubMed.
  44. J. Kubečka, V. Besel, T. Kurtén, N. Myllys and H. Vehkamäki, Configurational Sampling of Noncovalent (Atmospheric) Molecular Clusters: Sulfuric Acid and Guanidine, J. Phys. Chem. A, 2019, 123, 6022–6033 CrossRef.
  45. J. Zhang and M. Dolg, ABCluster: the artificial bee colony algorithm for cluster global optimization, Phys. Chem. Chem. Phys., 2015, 17, 24173–24181 RSC.
  46. K. A. Peterson, D. Figgen, E. Goll, H. Stoll and M. Dolg, Systematically convergent basis sets with relativistic pseudopotentials. II. Small-core pseudopotentials and correlation consistent basis sets for the post-d group 16–18 elements, J. Chem. Phys., 2003, 119, 11113–11123 CrossRef CAS.
  47. g. Funes-Ardoiz; and R. S. Paton, GoodVibes, Version 1.0.1, Zenodo, 2016,  DOI:10.5281/zenodo.60811.
  48. S. Grimme, J. Antony, S. Ehrlich and H. Krieg, A consistent and accurateab initioparametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu, J. Chem. Phys., 2010, 132, 154104 CrossRef PubMed.
  49. S. Grimme, S. Ehrlich and L. Goerigk, Effect of the damping function in dispersion corrected density functional theory, J. Comput. Chem., 2011, 32, 1456–1465 CrossRef CAS.
  50. L. Goerigk and S. Grimme, A thorough benchmark of density functional methods for general main group thermochemistry, kinetics, and noncovalent interactions, Phys. Chem. Chem. Phys., 2011, 13, 6670–6688 RSC.
  51. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16 Rev. A.01, Gaussian, Inc., Wallingford, CT, 2016 Search PubMed.
  52. F. Neese, The ORCA Program System, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2012, 2, 73–78 CAS.
  53. F. Neese, Software update: The ORCA program system—Version 5.0, Wiley Interdiscip. Rev.:Comput. Mol. Sci., 2022, 12, e1606 Search PubMed.
  54. M. J. McGrath, T. Olenius, I. K. Ortega, V. Loukonen, P. Paasonen, T. Kurtén, M. Kulmala and H. Vehkamäki, Atmospheric Cluster Dynamics Code: A Flexible Method for Solution of the Birth-Death Equations, Atmos. Chem. Phys., 2012, 12, 2345–2355 CrossRef CAS.
  55. M. Dal Maso, M. Kulmala, K. E. J. Lehtinen, J. M. Mäkelä, P. Aalto and C. D. O'Dowd, Condensation and Coagulation Sinks and Formation of Nucleation Mode Particles in Coastal and Boreal Forest Boundary Layers, J. Geophys. Res.:Atmos., 2002, 107, 1–10 CrossRef.
  56. Y. Georgievskii and S. J. Klippenstein, Long-range transition state theory, J. Chem. Phys., 2005, 122, 194103 CrossRef PubMed.
  57. L. Yao, O. Garmash, F. Bianchi, J. Zheng, C. Yan, J. Kontkanen, H. Junninen, S. B. Mazon, M. Ehn, P. Paasonen, M. Sipila, M. Wang, X. Wang, S. Xiao, H. Chen, Y. Lu, B. Zhang, D. Wang, Q. Fu, F. Geng, L. Li, H. Wang, L. Qiao, X. Yang, J. Chen, V.-M. Kerminen, T. Petaja, D. R. Worsnop, M. Kulmala and L. Wang, Atmospheric new particle formation from sulfuric acid and amines in a Chinese megacity, Science, 2018, 361, 278–281 CrossRef CAS PubMed.
  58. A. Shiroudi, M. Śmiechowski, J. Czub and M. A. Abdel-Rahman, Computational analysis of substituent effects on proton affinity and gas-phase basicity of TEMPO derivatives and their hydrogen bonding interactions with water molecules, Sci. Rep., 2024, 14, 8434 CrossRef CAS PubMed.
  59. M. Engsvang, H. Wu and J. Elm, Iodine Clusters in the Atmosphere I: Computational Benchmark and Dimer Formation of Oxyacids and Oxides, ACS Omega, 2024, 9, 31521–31532 CrossRef CAS PubMed.
  60. A. Ning, J. Zhong, L. Li, H. Li, J. Liu, L. Liu, Y. Liang, J. Li, X. Zhang, J. S. Francisco and H. He, Chemical Implications of Rapid Reactive Absorption of I2O4 at the Air-Water Interface, J. Am. Chem. Soc., 2023, 145, 10817–10825 CrossRef CAS PubMed.
  61. R. Zhang, I. Suh, J. Zhao, D. Zhang, E. C. Fortner, X. Tie, L. T. Molina and M. J. Molina, Atmospheric New Particle Formation Enhanced by Organic Acids, Science, 2004, 304, 1487–1490 CrossRef CAS PubMed.
  62. J. Zhao, A. Khalizov, R. Zhang and R. McGraw, Hydrogen-Bonding Interaction in Molecular Complexes and Clusters of Aerosol Nucleation Precursors, J. Phys. Chem. A, 2009, 113, 680–689 CrossRef CAS PubMed.
  63. M. Engsvang and J. Elm, Iodine Clusters in the Atmosphere II: Cluster Formation Potential of Iodine Oxyacids and Iodine Oxides, ACS Omega, 2025, 10, 24887–24896 CrossRef PubMed.
  64. H. Zu, S. Zhang, L. Liu and X. Zhang, The vital role of sulfuric acid in iodine oxoacids nucleation: impacts of urban pollutants on marine atmosphere, Environ. Res. Lett., 2024, 19, 014076 CrossRef CAS.
  65. J. Li, N. Wu, B. Chu, A. Ning and X. Zhang, Molecular-level study on the role of methanesulfonic acid in iodine oxoacid nucleation, Atmos. Chem. Phys., 2024, 24, 3989–4000 CrossRef CAS.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5sc02517f
R. J. Z. and Y. Y. L. contributed equally.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.