Next Article in Journal
Reponses of Land Surface Albedo to Global Vegetation Greening: An Analysis Using GLASS Data
Next Article in Special Issue
Relative Humidity Impact on Organic New Particle Formation from Ozonolysis of α- and β-Pinene at Atmospherically Relevant Mixing Ratios
Previous Article in Journal
Deterministic–Probabilistic Prediction of Forest Fires from Lightning Activity Taking into Account Aerosol Emissions
Previous Article in Special Issue
Theoretical Perspectives on the Gas-Phase Oxidation Mechanism and Kinetics of Carbazole Initiated by OH Radical in the Atmosphere
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

HONO Formation from the Oxidation Reactions of ClO, NO, and Water in the Gas-Phase and at the Air-Water Interface

1
Environment Research Institute, Shandong University, Qingdao 266237, China
2
Shenzhen Research Institute, Shandong University, Shenzhen 518057, China
3
College of Environment and Safety Engineering, Qingdao University of Science & Technology, Qingdao 266042, China
*
Author to whom correspondence should be addressed.
Atmosphere 2023, 14(1), 30; https://doi.org/10.3390/atmos14010030
Submission received: 19 November 2022 / Revised: 13 December 2022 / Accepted: 20 December 2022 / Published: 24 December 2022
(This article belongs to the Special Issue New Insights into Secondary Organic Aerosol Formation)

Abstract

:
Nitrous acid (HONO) plays a key role in atmospheric chemistry. Nevertheless, the HONO formation mechanism in the atmosphere, especially in the marine boundary layer, remains to be fully understood. Here, Born–Oppenheimer molecular dynamic and metadynamics simulations were performed to study the formation mechanism of HONO from the oxidation reactions of ClO radical and NO with the addition of (H2O)1–2, considering a monohydrated system ((ClO)(NO)(H2O)1) and dihydrated system ((ClO)(NO)(H2O)2), as well as at the air-water interface. This study shows that HONO formation follows a single-water mechanism in gas-phase and air-water interface systems. The free-energy barrier of the (ClO)(NO)(H2O)1 system was 9.66 kJ mol−1, whereas the (ClO)(NO)(H2O)2 system was a barrierless reaction. HONO formation at the air-water interface was faster than that in monohydrated and dihydrated systems. Although the concentration of ClO radical in the marine boundary layer is two orders higher than that of Cl radical, the production rates of HONO from the (ClO)(NO)(H2O)1 system are six orders lower than that from the (Cl)(NO)(H2O)1 system, which means that Cl radical dominates HONO formation rather than ClO radical in the marine boundary layer. These results can deepen our understanding of the HONO formation mechanism and be used to reduce HONO emissions and establish HONO-control strategies.

1. Introduction

In the atmosphere, the hydroxyl radical (OH) is generally considered a “detergent” due to its contribution to atmospheric oxidizing capacity, which can clean up primary pollutants as well as initiate the photochemical process to further produce ozone and secondary pollutants [1,2,3]. Nitrous acid (HONO), an active nitrogen-containing component, is a significant reservoir of OH radicals. Field and modeling studies have evaluated that photolysis of HONO can contribute to 25–60% of the daytime OH radical on average (even about 90% during winter noon) [4,5,6,7]. Hence, HONO can indirectly affect atmospheric oxidizing capacity and influence regional air quality and global climate.
HONO has been detected in various environments, including ocean, rural, urban, forest, and polar [8,9,10,11,12]. For example, the average concentration of HONO was 0.44 ppb during a shipboard-based observation in the East China Sea [9] and 0.74 ppb at an urban site in Guangzhou [13]. The main sources of HONO can be categorized into direct emissions and secondary emissions. Direct emission sources contain vehicle exhaust [14,15], soil microbiological activities [16,17], biomass burning [18], and indoor combustion processes [19], as well as ship emissions in the marine [20]. The secondary emissions of HONO are mainly from homogeneous and heterogeneous chemical reactions of NOx (i.e., NO and NO2). During the daytime, HONO can be mainly produced through the reaction of NO and OH radicals [21,22] and the reaction of NO and Cl radicals in the introduction of water molecules [23]. During nighttime, heterogeneous conversions of NO2 on various wet surfaces, like aerosols, ground, vegetation, and sea, have always been considered the main nocturnal HONO formation route [10,24,25,26,27]. For example, HONO can be produced on indoor material surfaces by heterogeneous hydrolysis of NO2 [28].
Despite the number of advances in exploring the HONO sources, the concentrations of HONO predicted by models and coefficient calculations based on current known HONO formation mechanisms are much lower than that of observed HONO. For example, the HONO concentration predicted by the Community Multiscale Air Quality (CMAQ) model only accounted for 60% of the observed HONO concentration [29]. The predicted HONO concentration calculated from the rate coefficient method was two and four times lower than the measured concentrations of HONO in severe haze and clean periods [30], respectively. Therefore, the sources and formation mechanisms of HONO are still unclearly and should be further investigated, especially in the marine boundary layer (MBL). For instance, the calculated production/loss rates of HONO in the East China Sea showed that the unknown HONO production rates at two monitoring sites are 1.52 ppb h−1 and 1.14 ppb h−1, respectively [9]. Additionally, larger marine-derived HONO production rates in MBL than that in the land were detected in field observations [27,31]. Yang et al. revealed that the NO2-to-HONO conversion rate in the “sea case” is four times faster than that in the “land case”, where the sea surface microlayer is likely to be a medium for the formation of HONO [32]. Therefore, further investigations of the HONO formation mechanism in the MBL are much appreciated.
The air-water interfaces are ubiquitous and can influence most tropospheric chemical reactions in aerosols, fogs, and clouds [33]. The air-water interfaces can provide an effective adsorption and reaction site, which improves atmospheric reaction rates or creates novel mechanisms [34,35,36,37]. Previous studies have indicated that OH radical shows a higher potential to react with other species at the air-water interface than that in the gas phase [38,39,40]. Xia et al. unveiled a significant mechanism of HONO formation from NO2 dimer catalyzed by CO2 at the air-water interface by combining the simulation and flow system experiment [41]. Li et al. revealed that water droplets reduce the free-energy barrier to 0.5 kcal mol−1 in HONO formation from NH3-promoted hydrolysis of NO2 dimer, which can be ignored at room temperature [42]. Since oceans have more abundant water and air-water interfaces than inland areas, air-water interfaces in MBL likely play a more important role in the formation of HONO than that in inland areas.
Field observations found halogen sources in polar, open ocean, coastal areas, and salt lake regions [43]. Chlorine monoxide radical (ClO) is one of the reactive halogen species which can play an important role in stratospheric ozone depletion through a catalytic cycle [44,45]. In the coastal MBL, the average concentration for ClO radical is 1.00 × 107 molecules cm−3, while the value is 6.00 × 105 molecules cm−3 for OH radical [46]. Interestingly, the rate constants of the oxidation reaction from OH and ClO radicals with caffeine showed a similar or close order of magnitude [47], indicating that OH and ClO radicals could have comparable reactivity and the importance of ClO in MBL should not be ignored. Meanwhile, the average concentration of Cl radical is 3.00 × 105 molecules cm−3 in the coastal MBL [46], which is two orders of magnitude lower than that of ClO radical. Our previous study proposed that the reaction of Cl radical, NO, and one water molecule can explain 40.3–53.8% of the unknown HONO production rate [23]. Considering the much higher concentration of ClO radical than that of Cl radical in MBL, as the following investigation, the HONO formation from ClO radical, NO, and water molecules need to be further explored.
Hence, in this study, the HONO formations from the reactions of ClO radical and NO with the addition of one to two water molecules and at the air-water interface in the MBL were simulated by using the Born–Oppenheimer molecular dynamics (BOMD) simulation and metadynamics (MTD) sampling. Density functional theory (DFT) calculations were further performed to verify the simulation results. The rate constants of ClO radical and NO molecule in the presence of one water molecule were calculated by using the canonical variational transition-state theory (CVT) with the small curvature tunneling contribution (SCT) method. Furthermore, the contributions to HONO formation initiated by Cl and ClO radicals were compared to evaluate their atmospheric significance [23].

2. Materials and Methods

2.1. Born–Oppenheimer Molecular Dynamics (BOMD) Simulation

BOMD simulation for all reactions in gas-phase and at air-water interface was carried out using the Quickstep module of the CP2K program package [48]. The simulation systems in the gas-phase were (ClO)(NO)(H2O)n (n = 1–2), and the droplet system was consisted of 191 water molecules in a spherical model. A cubic box with 20 × 20 × 20 Å3 was set for (ClO)(NO)(H2O)1–2 systems, while 35 × 35 × 35 Å3 was chosen as cubic box for the droplet system to minimize the interactions between adjacent periodic units (Figure S1). By using the Becke–Lee–Yang–Parr (BLYP) correction functional [49,50], we treated the exchange and correction interactions. Grimme’s dispersion correction (D3) method [51] was added to depict the weak dispersion interactions. To model the valence and core electrons, we chose a triple-ζ plus polarization level Gaussian basis set [52] and the Goedecker–Teter–Hutter (GTH) norm-conserved pseudopotentials [53], respectively. A 280 Ry energy cutoff was set for the plane-wave basis set, and a 40 Ry cutoff was selected for the Gaussian basis set. All BOMD simulations were employed under the canonical (NVT) ensemble with a temperature of 298 K, and the Nose–Hoover chain method [54] was adopted to control the temperature. The simulation step was set as 0.5 fs and 1 fs for gas-phase and air-water interface systems, respectively.

2.2. Metadynamics (MTD) Simulation

MTD simulation was employed using the PLUMED 2.4 plugin [55], combined with CP2K package. The distance between the O atom of ClO radical and H atom of water (dO-H) was chosen as the collective variable (CV) for the (ClO)(NO)(H2O)1–2 systems. To reduce the unnecessary sampling, an upper wall was selected if the dO-H was greater than 4.0 Å with force constant of 150.00 kcal mol−1. For the Gaussian functions, we set their height and width to 0.05 kcal mol−1 and 0.1 Å, and every 100 steps will add a new Gaussian function.

2.3. Quantum Chemistry Calculations

To verify the results from dynamics simulations, a high-accuracy molecular orbital method was used to perform all density functional theory (DFT) calculations of reactants (R), pre-reaction compound (RC), transition state (TS), and product compound (PC) by using the Gaussian 16 software package [56], along with the hybrid meta function M06-2X method [57]. The geometric optimizations and vibrational frequency calculations were carried out at the 6-311++G(3df,3pd) level. The frequency calculations can ensure that there is only one imaginary frequency for TS. The intrinsic reaction coordinate (IRC) calculations [58] were also carried out at the same level to confirm that the TS is connected with desired RC and PC. For higher accuracy, the CCSD(T) method [59,60] with the aug-cc-pVTZ basis set was adopted to calculate single-point energies of the optimized geometries.

2.4. Rate Constant Calculations

The canonical variational transition-state (CVT) theory, along with the small curvature tunneling (SCT) contribution [61,62,63,64], was adopted to obtain the rate constant of the unimolecular reaction from RC to PC with the Polyrate 9.7 program [65]. Moreover, the KiSThelP program [66] was used to calculate the equilibrium constants between R and RC (details in the Supplementary Materials). It should be noted that all constants used in the study were based on a temperature range of 220–310 K.

3. Results and Discussion

HONO formation from oxidation reactions of ClO radical, NO, and water involves gas-phase reactions and air-water interface reactions. Two ab initio molecular dynamics simulations, BOMD simulation, and MTD simulation, as well as DFT quantum chemistry calculations, were used to present the dynamics and thermal analyses. For the gas-phase reactions, HONO formation from ClO radical, NO, and one water molecule (i.e., monohydrated system or (ClO)(NO)(H2O)1 system) and HONO formation from the ClO radical, NO, and two water molecules (i.e., dihydrated system or (ClO)(NO)(H2O)2 system) were considered by using both ab initio molecular dynamics simulation and DFT calculation methods. For the air-water interface reactions, HONO formation was mainly analyzed by BOMD simulation.

3.1. HONO Formation from the Monohydrated System in the Gas-Phase

Figure 1 shows the variations of structures and key bond lengths in the HONO formation from the ClO radical, NO, and one water molecule by using BOMD simulation within 2.5 ps. Initially, the distance from the Cl atom of the ClO radical to the O atom of H2O (Cl-O2) and the distance from the N atom of NO to the O atom of H2O (N-O2) was kept at 3.00 Å. The distance from the O atom of ClO to the H atom of H2O (O1-H1) and the distance from the O atom of H2O to the H atom of H2O (O2-H1) fluctuated around 4.00 Å and 1.00 Å, respectively. Up to 1.20 ps, the O2-H1 bond length was 1.08 Å, while at approximately 1.23 ps, the H1 atom transferred to the O1 atom, forming a stable O1-H1 bond with a strength of 1.00 Å. After the H1 transferring, the O2-H2 group of H2O moved to the N atom of NO immediately at 1.24 ps, forming O2-N with a bond length of around 1.50 Å. These results can be interpreted as describing the dynamic process of HONO formation in the monohydrated system, which occurs at about 1.24 ps.
In order to calculate the free-energy variation of the dynamic process, a 28 ps MTD simulation was conducted based on the results of the BOMD simulation. As presented in Figure 2, there is a low free-energy barrier of 9.66 kJ mol−1 for the monohydrated system considering the O1-H1 bond length as the Collective Variable (CV). It also can be seen from Figure S2a that the O1-H1 bond decreases from 4.00 Å to 1.00 Å in a dynamic equilibrium process after 0.80 ps.
DFT calculations were carried out to confirm the results from the BOMD and MTD simulations. Figure 3 shows the free energy profile of the HONO formation from the monohydrated system at the CCSD(T)/aug–cc–pVTZ//M06–2X/6–311++G(3df,3pd) level. Firstly, the ClO radical and NO molecule collide with the water molecule to form ClO-H2O and NO-H2O compounds, with a binding energy of 14.08 and 14.82 kJ mol−1, respectively. As pointed out in Figure 3, the subsequent formation of the ClO-NO-H2O pre-reaction compound (RC) is an endoergic reaction with a free energy of 23.71 kJ mol−1, followed by the formation of the product compound (PC) derived from the RC with a free energy of −51.57 kJ mol−1. This step can occur with a free-energy barrier of 18.75 kJ mol−1, resulting in the formation of HONO and HOCl, agreeing well with the MTD free-energy barrier of 9.66 kJ mol−1 for the (ClO)(NO)(H2O)1 system.
In our previous study, we simulated the HONO formation from the reaction of the Cl radical and the NO molecule with the addition of one water molecule [23]. It is interesting to compare the effect of Cl and ClO radicals on the HONO formation. For example, the MTD free-energy barrier of the HONO formation from the (Cl)(NO)(H2O)1 system is 3.97 kJ mol−1 [23], which is much lower than the value of 9.66 kJ mol−1 for the (ClO)(NO)(H2O)1 system in this study. Moreover, the DFT calculated free-energy barrier of the HONO formation in the (Cl)(NO)(H2O)1 system is 16.26 kJ mol−1, which is also much lower than that in the (ClO)(NO)(H2O)1 system. In addition, the formation of the RC from the ClO radical, NO, and one water molecule is endoergic by 23.71 kJ mol−1, while the formation of RC from the Cl radical, NO, and one water molecule is exoergic by −0.79 kJ mol−1. Thus, the HONO formation from the Cl radical, NO, and one water molecule is energetically more favorable than from the ClO radical, NO, and one water molecule, which matches well with the reported phenomenon that the Cl radical is more reactive than ClO radical [67,68].

3.2. HONO Formation from the Dihydrated System in the Gas-Phase

In a similar manner, the HONO formation from the ClO radical, NO, and two water molecules was also investigated by 2.5 ps BOMD and 28 ps MTD simulations, along with DFT calculations at the CCSD(T)/aug–cc–pVTZ//M06–2X/6–311++G(3df,3pd) level. In the previous study of HONO forming through the NO2 dimers hydrolysis reaction catalyzed by NH3, Li and coworkers had proposed a dual-water mechanism for their dihydrated system and believed that every water molecule was involved in the hydrogen transfer process [42]. However, in this study, Figure 4a clearly shows that the dihydrated system follows a single-water mechanism, which means only one water molecule is involved in the reaction of the ClO radical and the NO molecule, while the second water molecule acts as the “solvent”. In Figure 4b, the O2-H1 bond in H2O remains at 1.00 Å at the beginning, and then the H1 transfers to the O atom of ClO to form a new hydrogen bond (O1-H1) at around 0.43 ps. In the next step, the remnant OH radical is connected quickly to the N atom of NO to form the HONO molecule.
To verify the effect of the “solvent” molecule, MTD simulation was employed to obtain the free energy variation of the dihydrated system. It can be obviously shown in Figure 5 that the HONO formation in the dihydrated system is a barrierless reaction, with the free-energy barrier being much lower than that in the monohydrated system (9.66 kJ mol−1). This demonstrates that the HONO formation from the dihydrated system is energetically more favorable than that from the monohydrated system. Although the second water molecule does not directly participate in the HONO formation and does not provide the hydrogen bond for the reaction, it can decrease the free-energy barrier effectively and play a positive catalytic effect on HONO formation.
The DFT calculation was also carried out to present the indirect evidence for the fact that the second water molecule serves as the “solvent”. In the HONO formation mechanism, the NO2 dimer hydrolyzes, and NH3 acts as a catalyst proposed by Li and coworkers [42]; the free-energy barrier of the dihydrated system was calculated based on a loop structure formed by two water molecules participating in hydrogen transfer. However, as presented in Figure 6, the formation of the product (PC_2) from the pre-reaction compound (RC_2) with two water molecules directly participating in the reaction via a loop structure exhibits a free-energy barrier of 121.69 kJ mol−1, which is much higher than the value of 18.75 kJ mol−1 in the monohydrated system. Therefore, the HONO formation from the dihydrated system tends to follow the single-water mechanism, which means that one water molecule participates in the hydrogen transfer process, and the second one acts as the “solvent” and has no interaction with other molecules. Overall, the HONO formation in both monohydrated and dihydrated systems follows the single-water mechanism rather than the dual-water mechanism [42].

3.3. HONO Formation at the Air-Water Interface

Considering the fact that the air-water interface is ubiquitous in MBL, the HONO formation from the ClO radical and the NO molecule at the air-water interface was studied by BOMD simulation. Through the 5 ps BOMD simulation presented in Figure 7, it is obvious that only one water molecule participates in the hydrogen transfer process, resulting in HONO and HOCl molecules. In other words, similar to the gas-phase reactions, the HONO formation from the ClO radical and the NO molecule at the air-water interface also exhibits a single-water mechanism. In the beginning, the ClO radical and the NO molecule were placed in the surface region without any interaction with neighboring water molecules. The distance from the O atom of the ClO radical to the H1 atom of H2O (O1-H1) decreased rapidly from 3.84 Å at 0.00 ps to 1.00 Å at 0.18 ps, resulting in the formation of the O1-H1 bond and the HOCl molecule. Next, the remnant OH group of the water molecule moved towards NO and yielded the HONO molecule, in which the distance between the O atom of H2O and the N atom of NO (O2-N) decreased from 2.56 Å to 1.50 Å until 0.19 ps. Subsequently, the bond lengths of O1-H1 and O2-N remained stable, indicating the reaction had stopped. It is necessary to compare the HONO formation in the gas-phase systems with the HONO formation in the air-water interface system. HONO formation from the ClO radical and NO at the air-water interface only needs 0.19 ps, which is much faster than the formation in the monohydrated system (1.24 ps) and dihydrated system (0.45 ps), respectively, implying that air-water interface has a positive catalytic influence on the formation of HONO. This agrees well with the previous study that the time scale for the CH2OO reacting with water at the air-water interface was two to three orders of magnitude shorter than in the gas phase [69].
As displayed in Figure S5, a 5 ps BOMD simulation was performed for the ClO radical and NO, initially placed in the interior of the droplet. During the whole process, there is no water molecule reacting with the ClO radical and the NO molecule, implying the weak reactivity of water molecules inside the droplet. This phenomenon is similar to previously published research [44], in which the ClO radical preferred to be adsorbed on the air-water interface rather than to be dissolved in bulk.

3.4. Atmospheric Implication

The unimolecular reaction rate constants for the reaction from RC to PC and total rate constants were calculated to determine the contribution of the monohydrated system in the marine atmosphere. The equilibrium constants (Keq) were calculated with the KisThelP program [66], and the unimolecular rate constants (kuni) were obtained by CVT/SCT theory by using the Polyrate program [65]. All equilibrium constants, unimolecular rate constants, and total rate constants with temperature ranges from 220 K to 310 K are listed in Table S1. From Table S1, the total rate constants of the (ClO)(NO)(H2O)1 system vary from 4.65 × 10−40 cm6 molecule−2 s−1 (310 K) to 2.11 × 10−39 cm6 molecule−2 s−1 (220 K), which are 8–11 orders of magnitude lower than the (Cl)(NO)(H2O)1 system from 6.10 × 10−32 cm6 molecule−2 s−1 (310 K) to 2.07 × 10−28 cm6 molecule−2 s−1 (220 K) [23].
In addition to the rate constants, the reactant concentrations are also a critical element in evaluating the significance of the monohydrated system. According to field measurements, the average concentration of NO is 0.5 ppbv over the eastern Bohai Sea [70] and can reach approximately 350 ppbv in the East China Sea [9]. The average concentration of the ClO radical is 1.00 × 107 molecules cm−3 [46], with a maximum concentration of 15 pptv (3.67 × 108 molecules cm−3) in the coastal MBL [71]. The water molecule has a concentration of 5.18 × 1017 molecules cm−3 [72]. Four cases were considered based on the concentrations of ClO, NO, and water molecule, namely cases 1–4. All the reaction rates at 300 K for the monohydrated system are depicted in Table S2. The highest production rate of HONO is 1.25 × 10−7 ppb h−1 when the [NO] is 350 ppbv and [ClO] is 3.67 × 108 molecules cm−3, which is six orders of magnitude lower than the value of 6.13 × 10−1 ppb h−1 for the (Cl)(NO)(H2O)1 system in the case of [NO] being 350 ppbv and [Cl] being 8.00 × 106 molecule cm−3 [23]. Although the concentration of the ClO radical in coastal MBL is two orders of magnitude higher than that of the Cl radical, the Cl radical dominates the HONO formation rather than the ClO radical in the MBL.

4. Conclusions

In this study, the HONO formation mechanisms from the ClO radical and the NO molecule in the presence of one to two water molecules and at the air-water interface were investigated by BOMD simulations and MTD samplings as well as DFT calculations. The monohydrated system exhibits a free-energy barrier of 9.66 kJ mol−1, whereas the dihydrated system is barrierless. All systems follow the single-water mechanism, which has one water molecule participating in the proton transfer of the reactions directly; other water molecules in the dihydrated system and in a water droplet just serve as the “solvent”. The “solvent” water molecules can decrease the free-energy barrier in the dihydrated system or accelerate the simulation time at the air-water interface, which plays a positive catalytic effect on HONO formation. Although the concentration of the ClO radical in the MBL is two orders higher than that of the Cl radical, the production rates of HONO from the (ClO)(NO)(H2O)1 system are six orders lower than that from the (Cl)(NO)(H2O)1 system, which means that the Cl radical dominates the HONO formation rather than the ClO radical in the MBL. These new insights for the HONO formation can be input into HONO control and prediction models as detailed parameters and then narrow the gap between the predicted and measured values of the HONO concentration.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/atmos14010030/s1, Figure S1: The BOMD simulation systems for (a) the monohydrated system, (b) the dihydrated system and (c) the air-water interface system with 191 water molecules; Figure S2: Length of O1-H1 versus time in the 28 ps metadynamics simulation for (a) the monohydrated system and (b) the dihydrated system; Figure S3: Snapshots of the system with ClO radical initially placed in the interior of the water droplet; Figure S4: Snapshots of the system with NO radical initially placed in the interior of the water droplet; Figure S5: Snapshots of the system with NO and the ClO radical initially placed in the interior of the water droplet; Table S1: The equilibrium constants (Keq, cm3 molecule−1), unimolecular rate constants (kuni, s−1) and total rate constants (ktotal, cm6 molecule−2 s−1) for the monohydrated system at 220–310 K; Table S2: The reaction rates (ppb h−1) for the monohydrated system under different concentrations of the ClO radical and the NO molecule (molecule cm−3) at 300 K.

Author Contributions

Conceptualization, F.X.; methodology, Q.Z. and X.Z.; software, X.Z. and M.H.H.; formal analysis, Q.Z.; investigation, X.W. and X.B.; data curation, Q.Z.; writing—original draft, Q.Z.; writing—review and editing, F.X. and M.H.H.; visualization, Y.S.; supervision, F.X.; funding acquisition, F.X. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by NSFC (National Natural Science Foundation of China, Project Nos. 22236004, 22076103, 42075106, 21976107).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data that support the findings of this study are available from the corresponding author upon reasonable request.

Acknowledgments

The authors thank Donald G. Truhlar for providing the POLYRATE 9.7 program.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Cheng, Y.; Zheng, G.; Wei, C.; Mu, Q.; Zheng, B.; Wang, Z.; Gao, M.; Zhang, Q.; He, K.; Carmichael, G.; et al. Reactive Nitrogen Chemistry in Aerosol Water as a Source of Sulfate during Haze Events in China. Sci. Adv. 2016, 2, e1601530. [Google Scholar] [CrossRef] [Green Version]
  2. Wang, G.; Zhang, R.; Gomez, M.E.; Yang, L.; Zamora, M.L.; Hu, M.; Lin, Y.; Peng, J.; Guo, S.; Meng, J.; et al. Persistent Sulfate Formation from London Fog to Chinese Haze. Proc. Natl. Acad. Sci. USA 2016, 113, 13630–13635. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Fu, X.; Wang, T.; Zhang, L.; Li, Q.; Wang, Z.; Xia, M.; Yun, H.; Wang, W.; Yu, C.; Yue, D.; et al. The Significant Contribution of HONO to Secondary Pollutants during a Severe Winter Pollution Event in Southern China. Atmos. Chem. Phys. 2019, 19, 1–14. [Google Scholar] [CrossRef] [Green Version]
  4. Alicke, B.; Platt, U.; Stutz, J. Impact of Nitrous Acid Photolysis on the Total Hydroxyl Radical Budget during the Limitation of Oxidant Production/Pianura Padana Produzione Di Ozono Study in Milan. J. Geophys. Res. Atmos. 2002, 107, 8196. [Google Scholar] [CrossRef]
  5. Kleffmann, J. Daytime Sources of Nitrous Acid (HONO) in the Atmospheric Boundary Layer. ChemPhysChem 2007, 8, 1137–1144. [Google Scholar] [CrossRef]
  6. Yang, Q.; Su, H.; Li, X.; Cheng, Y.; Lu, K.; Cheng, P.; Gu, J.; Guo, S.; Hu, M.; Zeng, L.; et al. Daytime HONO Formation in the Suburban Area of the Megacity Beijing, China. Sci. China Chem. 2014, 57, 1032–1042. [Google Scholar] [CrossRef]
  7. Hendrick, F.; Clémer, K.; Wang, P.; De Mazière, M.; Fayt, C.; Gielen, C.; Hermans, C.; Ma, J.Z.; Pinardi, G.; Stavrakou, T.; et al. Four Years of Ground-Based MAX-DOAS Observations of HONO and NO2 in the Beijing Area. Atmos. Chem. Phys. 2014, 14, 765–781. [Google Scholar] [CrossRef] [Green Version]
  8. Beine, H.J.; Dominé, F.; Simpson, W.; Honrath, R.E.; Sparapani, R.; Zhou, X.; King, M. Snow-Pile and Chamber Experiments during the Polar Sunrise Experiment ‘Alert 2000’: Exploration of Nitrogen Chemistry. Atmos. Environ. 2002, 36, 2707–2719. [Google Scholar] [CrossRef]
  9. Cui, L.; Li, R.; Fu, H.; Li, Q.; Zhang, L.; George, C.; Chen, J. Formation Features of Nitrous Acid in the Offshore Area of the East China Sea. Sci. Total Environ. 2019, 682, 138–150. [Google Scholar] [CrossRef]
  10. Li, X.; Brauers, T.; Häseler, R.; Bohn, B.; Fuchs, H.; Hofzumahaus, A.; Holland, F.; Lou, S.; Lu, K.D.; Rohrer, F.; et al. Exploring the Atmospheric Chemistry of Nitrous Acid (HONO) at a Rural Site in Southern China. Atmos. Chem. Phys. 2012, 12, 1497–1513. [Google Scholar] [CrossRef]
  11. Liu, J.; Liu, Z.; Ma, Z.; Yang, S.; Yao, D.; Zhao, S.; Hu, B.; Tang, G.; Sun, J.; Cheng, M.; et al. Detailed Budget Analysis of HONO in Beijing, China: Implication on Atmosphere Oxidation Capacity in Polluted Megacity. Atmos. Environ. 2021, 244, 117957. [Google Scholar] [CrossRef]
  12. Zhang, N.; Zhou, X.; Bertman, S.; Tang, D.; Alaghmand, M.; Shepson, P.B.; Carroll, M.A. Measurements of Ambient HONO Concentrations and Vertical HONO Flux above a Northern Michigan Forest Canopy. Atmos. Chem. Phys. 2012, 12, 8285–8296. [Google Scholar] [CrossRef] [Green Version]
  13. Yu, Y.; Cheng, P.; Li, H.; Yang, W.; Han, B.; Song, W.; Hu, W.; Wang, X.; Yuan, B.; Shao, M.; et al. Budget of Nitrous Acid (HONO) at an Urban Site in the Fall Season of Guangzhou, China. Atmos. Chem. Phys. 2022, 22, 8951–8971. [Google Scholar] [CrossRef]
  14. Li, S.; Song, W.; Zhan, H.; Zhang, Y.; Zhang, X.; Li, W.; Tong, S.; Pei, C.; Wang, Y.; Chen, Y.; et al. Contribution of Vehicle Emission and NO2 Surface Conversion to Nitrous Acid (HONO) in Urban Environments: Implications from Tests in a Tunnel. Environ. Sci. Technol. 2021, 55, 15616–15624. [Google Scholar] [CrossRef]
  15. Xu, Z.; Wang, T.; Wu, J.; Xue, L.; Chan, J.; Zha, Q.; Zhou, S.; Louie, P.K.K.; Luk, C.W.Y. Nitrous Acid (HONO) in a Polluted Subtropical Atmosphere: Seasonal Variability, Direct Vehicle Emissions and Heterogeneous Production at Ground Surface. Atmos. Environ. 2015, 106, 100–109. [Google Scholar] [CrossRef]
  16. Mushinski, R.M.; Phillips, R.P.; Payne, Z.C.; Abney, R.B.; Jo, I.; Fei, S.; Pusede, S.E.; White, J.R.; Rusch, D.B.; Raff, J.D. Microbial Mechanisms and Ecosystem Flux Estimation for Aerobic NOy Emissions from Deciduous Forest Soils. Proc. Natl. Acad. Sci. USA 2019, 116, 2138–2145. [Google Scholar] [CrossRef] [Green Version]
  17. Wu, D.; Horn, M.A.; Behrendt, T.; Müller, S.; Li, J.; Cole, J.A.; Xie, B.; Ju, X.; Li, G.; Ermel, M.; et al. Soil HONO Emissions at High Moisture Content Are Driven by Microbial Nitrate Reduction to Nitrite: Tackling the HONO Puzzle. ISME J. 2019, 13, 1688–1699. [Google Scholar] [CrossRef]
  18. Cui, L.; Li, R.; Fu, H.; Meng, Y.; Zhao, Y.; Li, Q.; Chen, J. Nitrous Acid Emission from Open Burning of Major Crop Residues in Mainland China. Atmos. Environ. 2021, 244, 117950. [Google Scholar] [CrossRef]
  19. Zhang, J.; An, J.; Qu, Y.; Liu, X.; Chen, Y. Impacts of Potential HONO Sources on the Concentrations of Oxidants and Secondary Organic Aerosols in the Beijing-Tianjin-Hebei Region of China. Sci. Total Environ. 2019, 647, 836–852. [Google Scholar] [CrossRef]
  20. Sun, L.; Chen, T.; Jiang, Y.; Zhou, Y.; Sheng, L.; Lin, J.; Li, J.; Dong, C.; Wang, C.; Wang, X.; et al. Ship Emission of Nitrous Acid (HONO) and Its Impacts on the Marine Atmospheric Oxidation Chemistry. Sci. Total Environ. 2020, 735, 139355. [Google Scholar] [CrossRef]
  21. Li, S.; Matthews, J.; Sinha, A. Atmospheric Hydroxyl Radical Production from Electronically Excited NO2 and H2O. Science 2008, 319, 1657–1660. [Google Scholar] [CrossRef]
  22. Zhang, L.; Wang, T.; Zhang, Q.; Zheng, J.; Xu, Z.; Lv, M. Potential Sources of Nitrous Acid (HONO) and Their Impacts on Ozone: A WRF-Chem Study in a Polluted Subtropical Region. J. Geophys. Res. Atmos. 2016, 121, 3645–3662. [Google Scholar] [CrossRef] [Green Version]
  23. Zhao, X.; Shi, X.; Ma, X.; Wang, J.; Xu, F.; Zhang, Q.; Li, Y.; Teng, Z.; Han, Y.; Wang, Q.; et al. Simulation Verification of Barrierless HONO Formation from the Oxidation Reaction System of NO, Cl, and Water in the Atmosphere. Environ. Sci. Technol. 2021, 55, 7850–7857. [Google Scholar] [CrossRef] [PubMed]
  24. Huang, R.J.; Yang, L.; Cao, J.; Wang, Q.; Tie, X.; Ho, K.F.; Shen, Z.; Zhang, R.; Li, G.; Zhu, C.; et al. Concentration and Sources of Atmospheric Nitrous Acid (HONO) at an Urban Site in Western China. Sci. Total Environ. 2017, 593–594, 165–172. [Google Scholar] [CrossRef] [PubMed]
  25. Qin, M.; Xie, P.; Su, H.; Gu, J.; Peng, F.; Li, S.; Zeng, L.; Liu, J.; Liu, W.; Zhang, Y. An Observational Study of the HONO-NO2 Coupling at an Urban Site in Guangzhou City, South China. Atmos. Environ. 2009, 43, 5731–5742. [Google Scholar] [CrossRef]
  26. Su, H.; Cheng, Y.F.; Cheng, P.; Zhang, Y.H.; Dong, S.; Zeng, L.M.; Wang, X.; Slanina, J.; Shao, M.; Wiedensohler, A. Observation of Nighttime Nitrous Acid (HONO) Formation at a Non-Urban Site during PRIDE-PRD2004 in China. Atmos. Environ. 2008, 42, 6219–6232. [Google Scholar] [CrossRef]
  27. Zha, Q.; Xue, L.; Wang, T.; Xu, Z.; Yeung, C.; Louie, P.K.K.; Luk, C.W.Y. Large Conversion Rates of NO2 to HNO2 Observed in Air Masses from the South China Sea: Evidence of Strong Production at Sea Surface? Geophys. Res. Lett. 2014, 41, 7710–7715. [Google Scholar] [CrossRef]
  28. Pandit, S.; Grassian, V.H. Gas-Phase Nitrous Acid (HONO) Is Controlled by Surface Interactions of Adsorbed Nitrite (NO2-) on Common Indoor Material Surfaces. Environ. Sci. Technol. 2022, 56, 12045–12054. [Google Scholar] [CrossRef]
  29. Sarwar, G.; Roselle, S.J.; Mathur, R.; Appel, W.; Dennis, R.L.; Vogel, B. A Comparison of CMAQ HONO Predictions with Observations from the Northeast Oxidant and Particle Study. Atmos. Environ. 2008, 42, 5760–5770. [Google Scholar] [CrossRef]
  30. Hou, S.; Tong, S.; Ge, M.; An, J. Comparison of Atmospheric Nitrous Acid during Severe Haze and Clean Periods in Beijing, China. Atmos. Environ. 2016, 124, 199–206. [Google Scholar] [CrossRef]
  31. Ye, C.; Zhou, X.; Pu, D.; Stutz, J.; Festa, J.; Spolaor, M.; Tsai, C.; Cantrell, C.; Mauldin, R.L.; Campos, T.; et al. Rapid Cycling of Reactive Nitrogen in the Marine Boundary Layer. Nature 2016, 532, 489–491. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, J.; Shen, H.; Guo, M.Z.; Zhao, M.; Jiang, Y.; Chen, T.; Liu, Y.; Li, H.; Zhu, Y.; Meng, H.; et al. Strong Marine-Derived Nitrous Acid (HONO) Production Observed in the Coastal Atmosphere of Northern China. Atmos. Environ. 2021, 244, 117948. [Google Scholar] [CrossRef]
  33. Zhong, J.; Kumar, M.; Francisco, J.S.; Zeng, X.C. Insight into Chemistry on Cloud/Aerosol Water Surfaces. Acc. Chem. Res. 2018, 51, 1229–1237. [Google Scholar] [CrossRef] [PubMed]
  34. Kumar, M.; Francisco, J.S. Elucidating the Molecular Mechanisms of Criegee-Amine Chemistry in the Gas Phase and Aqueous Surface Environments. Chem. Sci. 2019, 10, 743–751. [Google Scholar] [CrossRef] [Green Version]
  35. Zhong, J.; Li, H.; Kumar, M.; Liu, J.; Liu, L.; Zhang, X.; Zeng, X.C.; Francisco, J.S. Mechanistic Insight into the Reaction of Organic Acids with SO3 at the Air-Water Interface. Angew. Chem. 2019, 131, 8439–8443. [Google Scholar] [CrossRef]
  36. Martins-Costa, M.T.C.; Anglada, J.M.; Francisco, J.S.; Ruiz-López, M.F. Impacts of Cloud Water Droplets on the OH Production Rate from Peroxide Photolysis. Phys. Chem. Chem. Phys. 2017, 19, 31621–31627. [Google Scholar] [CrossRef]
  37. Martins-Costa, M.T.C.; Anglada, J.M.; Francisco, J.S.; Ruiz-López, M.F. Theoretical Investigation of the Photoexcited NO2 + H2O Reaction at the Air-Water Interface and Its Atmospheric Implications. Chem. Eur. J. 2019, 25, 13899–13904. [Google Scholar] [CrossRef]
  38. Hadizadeh, M.H.; Pan, Z.; Azamat, J. Investigation of OH Radical in the Water Nanodroplet during Vapor Freezing Process: An Ab Initio Molecular Dynamics Study. J. Mol. Liq. 2021, 343, 117597. [Google Scholar] [CrossRef]
  39. Hadizadeh, M.H.; Pan, Z.; Azamat, J. A New Insight into the Interaction of Hydroxyl Radical with Supercooled Nanodroplet in the Atmosphere. J. Mol. Liq. 2022, 359, 119261. [Google Scholar] [CrossRef]
  40. Hadizadeh, M.H.; Yang, L.; Fang, G.; Qiu, Z.; Li, Z. The Mobility and Solvation Structure of a Hydroxyl Radical in a Water Nanodroplet: A Born-Oppenheimer Molecular Dynamics Study. Phys. Chem. Chem. Phys. 2021, 23, 14628–14635. [Google Scholar] [CrossRef]
  41. Xia, D.; Zhang, X.; Chen, J.; Tong, S.; Xie, H.B.; Wang, Z.; Xu, T.; Ge, M.; Allen, D.T. Heterogeneous Formation of HONO Catalyzed by CO2. Environ. Sci. Technol. 2021, 55, 12215–12222. [Google Scholar] [CrossRef] [PubMed]
  42. Li, L.; Duan, Z.; Li, H.; Zhu, C.; Henkelman, G.; Francisco, J.S.; Zeng, X.C. Formation of HONO from the NH3-Promoted Hydrolysis of NO2 Dimers in the Atmosphere. Proc. Natl. Acad. Sci. USA 2018, 115, 7236–7241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Simpson, W.R.; Brown, S.S.; Saiz-Lopez, A.; Thornton, J.A.; Von Glasow, R. Tropospheric Halogen Chemistry: Sources, Cycling, and Impacts. Chem. Rev. 2015, 115, 4035–4062. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Shiyu, D.; Francisco, J.S.; Schenter, G.K.; Garrett, B.C. Interaction of ClO Radical with Liquid Water. J. Am. Chem. Soc. 2009, 131, 14778–14785. [Google Scholar]
  45. Bedjanian, Y.; Poulet, G. Kinetics of Halogen Oxide Radicals in the Stratosphere. Chem. Rev. 2003, 103, 4639–4655. [Google Scholar] [CrossRef]
  46. Chang, C.T.; Liu, T.H.; Jeng, F.T. Atmospheric Concentrations of the Cl Atom, CIO Radical, and OH Radical in the Coastal Marine Boundary Layer. Environ. Res. 2004, 94, 67–74. [Google Scholar] [CrossRef]
  47. Li, M.; Mei, Q.; Han, D.; Wei, B.; An, Z.; Cao, H.; Xie, J.; He, M. The Roles of HO•, ClO• and BrO• Radicals in Caffeine Degradation: A Theoretical Study. Sci. Total Environ. 2021, 768, 144733. [Google Scholar] [CrossRef]
  48. Vandevondele, J.; Krack, M.; Mohamed, F.; Parrinello, M.; Chassaing, T.; Hutter, J. Quickstep: Fast and Accurate Density Functional Calculations Using a Mixed Gaussian and Plane Waves Approach. Comput. Phys. Commun. 2005, 167, 103–128. [Google Scholar] [CrossRef] [Green Version]
  49. Becke, A.D. Density-Functional Exchange-Energy Approximation with Correct Asymptotic Behavior. Phys. Rev. A 1988, 38, 3098–3100. [Google Scholar] [CrossRef]
  50. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti Correlation-Energy Formula into a Functional of the Electron Density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  51. Grimme, S.; Antony, J.; Ehrlich, S. A Consistent and Accurate Ab Initio Parametrization of Density Functional Dispersion Correction (DFT-D) for the 94 Elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [PubMed]
  52. Vandevondele, J.; Hutter, J. Gaussian Basis Sets for Accurate Calculations on Molecular Systems in Gas and Condensed Phases. J. Chem. Phys. 2007, 127, 114105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Goedecker, S.; Teter, M.; Hutter, J. Separable Dual-Space Gaussian Pseudopotentials. Phys. Rev. B 1996, 54, 1703–1710. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Nosé, S. A Unified Formulation of the Constant Temperature Molecular Dynamics Methods. J. Chem. Phys. 1984, 81, 511–519. [Google Scholar] [CrossRef] [Green Version]
  55. Tribello, G.A.; Bonomi, M.; Branduardi, D.; Camilloni, C.; Bussi, G. Plumed 2: New Feathers for an Old Bird. Comput. Phys. Commun. 2014, 185, 604–613. [Google Scholar] [CrossRef] [Green Version]
  56. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Fox, D.J. Gaussian 16, Revision C. 01; Gaussian, Inc.: Wallingford, CT, USA, 2019. [Google Scholar]
  57. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Function. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef] [Green Version]
  58. Hratchian, H.P.; Schlegel, H.B. Using Hessian Updating to Increase the Efficiency of a Hessian Based Predictor-Corrector Reaction Path Following Method. J. Chem. Theory Comput. 2005, 1, 61–69. [Google Scholar] [CrossRef]
  59. Purvis, G.D.; Bartlett, R.J. A Full Coupled-Cluster Singles and Doubles Model: The Inclusion of Disconnected Triples. J. Chem. Phys. 1982, 76, 1910–1918. [Google Scholar] [CrossRef]
  60. Pople, J.A.; Head-Gordon, M.; Raghavachari, K. Quadratic Configuration Interaction.A General Technique for Determining Electron Correlation Energies. J. Chem. Phys. 1987, 87, 5968–5975. [Google Scholar] [CrossRef]
  61. Baldridge, K.K.; Gordon, M.S.; Steckler, R.; Truhlar, D.G. Ab Initio Reaction Paths and Direct Dynamics Calculations. J. Phys. Chem. 1989, 93, 5107–5119. [Google Scholar] [CrossRef]
  62. Gonzalez-Lafont, A.; Truong, T.N.; Truhlar, D.G. Interpolated Variational Transition-State Theory: Practical Methods for Estimating Variational Transition-State Properties and Tunneling Contributions to Chemical Reaction Rates from Electronic Structure Calculations. J. Chem. Phys. 1991, 95, 8875–8894. [Google Scholar] [CrossRef]
  63. Garrett, B.C.; Truhlar, D.G. Classical Mechanical Theory and Applications to Collinear Reactions of Hydrogen Molecule. J. Phys. Chem. 1979, 83, 1052–1079. [Google Scholar] [CrossRef]
  64. Fernandez-Ramos, A.; Ellingson, B.A.; Garrett, B.C.; Truhlar, D.G. Variational Transition State Theory with Multidimensional Tunneling. Rev. Comput. Chem. 2007, 23, 125–232. [Google Scholar]
  65. Corchado, J.C.; Chuang, Y.Y.; Fast, P.L.; Hu, W.P.; Liu, Y.P.; Lynch, G.C.; Nguyen, K.A.; Jackels, C.F.; Fernandez-Ramos, A.; Ellingson, B.A. Polyrate, Version 9.7; University of Minnesota: Minneapolis, MN, USA, 2007. [Google Scholar]
  66. Ebastien Canneaux, S.; Ed Eric Bohr, F.; Henon, E. KiSThelP: A Program to Predict Thermodynamic Properties and Rate Constants from Quantum Chemistry Results. J. Comput. Chem. 2014, 35, 82–93. [Google Scholar] [CrossRef] [PubMed]
  67. Anglada, J.M.; Solé, A. The Atmospheric Oxidation of HONO by OH, Cl, and ClO Radicals. J. Phys. Chem. A 2017, 121, 9698–9707. [Google Scholar] [CrossRef] [PubMed]
  68. Zhou, S.; Zhang, W.; Sun, J.; Zhu, S.; Li, K.; Meng, X.; Luo, J.; Shi, Z.; Zhou, D.; Crittenden, J.C. Oxidation Mechanisms of the UV/Free Chlorine Process: Kinetic Modeling and Quantitative Structure Activity Relationships. Environ. Sci. Technol. 2019, 53, 4335–4345. [Google Scholar] [CrossRef]
  69. Zhu, C.; Kumar, M.; Zhong, J.; Li, L.; Francisco, J.S.; Zeng, X.C. New Mechanistic Pathways for Criegee-Water Chemistry at the Air/Water Interface. J. Am. Chem. Soc. 2016, 138, 11164–11169. [Google Scholar] [CrossRef]
  70. Wen, L.; Chen, T.; Zheng, P.; Wu, L.; Wang, X.; Mellouki, A.; Xue, L.; Wang, W. Nitrous Acid in Marine Boundary Layer over Eastern Bohai Sea, China: Characteristics, Sources, and Implications. Sci. Total Environ. 2019, 670, 282–291. [Google Scholar] [CrossRef]
  71. Stutz, J.; Ackermann, R.; Fast, J.D.; Barrie, L. Atmospheric Reactive Chlorine and Bromine at the Great Salt Lake, Utah. Geophys. Res. Lett. 2002, 29, 1380. [Google Scholar] [CrossRef] [Green Version]
  72. Torrent-Sucarrat, M.; Francisco, J.S.; Anglada, J.M. Sulfuric Acid as Autocatalyst in the Formation of Sulfuric Acid. J. Am. Chem. Soc. 2012, 134, 20632–20644. [Google Scholar] [CrossRef]
Figure 1. (a) BOMD simulation results of the monohydrated system. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 1. (a) BOMD simulation results of the monohydrated system. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g001
Figure 2. The free energy variation versus the corresponding collective variable of the O1−H1 bond length in the monohydrated system. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively. The arrow is used to show the free energy value.
Figure 2. The free energy variation versus the corresponding collective variable of the O1−H1 bond length in the monohydrated system. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively. The arrow is used to show the free energy value.
Atmosphere 14 00030 g002
Figure 3. The free energy profile and corresponding configurations of the monohydrated system at the CCSD(T)/cc−pVTZ//M06−2X/6−311++G(3df,3pd) level. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 3. The free energy profile and corresponding configurations of the monohydrated system at the CCSD(T)/cc−pVTZ//M06−2X/6−311++G(3df,3pd) level. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g003
Figure 4. (a) BOMD simulation results of the dihydrated system. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 4. (a) BOMD simulation results of the dihydrated system. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g004
Figure 5. The free energy variation versus the corresponding collective variable of the O1–H1 bond length in the dihydrated system. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 5. The free energy variation versus the corresponding collective variable of the O1–H1 bond length in the dihydrated system. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g005
Figure 6. The free energy profile and corresponding configurations of the dihydrated system at the CCSD(T)/cc–pVTZ//M06–2X/6–311++G(3df,3pd) level. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 6. The free energy profile and corresponding configurations of the dihydrated system at the CCSD(T)/cc–pVTZ//M06–2X/6–311++G(3df,3pd) level. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g006
Figure 7. (a) BOMD simulation results of the reaction of the ClO radical and the NO molecule on the surface of a water droplet. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Figure 7. (a) BOMD simulation results of the reaction of the ClO radical and the NO molecule on the surface of a water droplet. (b) Time evolution of the O1-H1, O2-H1, and O2-N bond lengths during the BOMD simulation. The white, blue, red, and green spheres represent H, N, O, and Cl atoms, respectively.
Atmosphere 14 00030 g007
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhang, Q.; Hadizadeh, M.H.; Wang, X.; Zhao, X.; Bai, X.; Xu, F.; Sun, Y. HONO Formation from the Oxidation Reactions of ClO, NO, and Water in the Gas-Phase and at the Air-Water Interface. Atmosphere 2023, 14, 30. https://doi.org/10.3390/atmos14010030

AMA Style

Zhang Q, Hadizadeh MH, Wang X, Zhao X, Bai X, Xu F, Sun Y. HONO Formation from the Oxidation Reactions of ClO, NO, and Water in the Gas-Phase and at the Air-Water Interface. Atmosphere. 2023; 14(1):30. https://doi.org/10.3390/atmos14010030

Chicago/Turabian Style

Zhang, Qi, Mohammad Hassan Hadizadeh, Xiaotong Wang, Xianwei Zhao, Xurong Bai, Fei Xu, and Yanhui Sun. 2023. "HONO Formation from the Oxidation Reactions of ClO, NO, and Water in the Gas-Phase and at the Air-Water Interface" Atmosphere 14, no. 1: 30. https://doi.org/10.3390/atmos14010030

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop