Adsorption and heterogeneous reactions of ClONO2 and N2O5 on/with NaCl aerosol

Yanhui Sunab, Qingzhu Zhang*a and Wenxing Wanga
aEnvironment Research Institute, Shandong University, Jinan 250100, P. R. China. E-mail: zqz@sdu.edu.cn; Fax: +86-531-88361990
bCollege of Environment and Safety Engineering, Qingdao University of Science and Technology, Qingdao 266042, Shandong, P. R. China

Received 13th February 2016 , Accepted 5th May 2016

First published on 6th May 2016


Abstract

The adsorption and heterogeneous reactions of ClONO2 and N2O5 on the NaCl (100) surface have been investigated by performing density functional theory (DFT) calculations. Possible adsorption structures, energies and electronic properties of ClONO2 and N2O5 were studied by considering multiple possible adsorption sites without symmetry restriction. The most stable adsorption configurations are a vertical adsorption mode for ClONO2 and a horizontal adsorption mode for N2O5, with the adsorption energies of 14.81 and 13.67 kcal mol−1, respectively. Two possible reactions were proposed on the NaCl (100) surface. Adsorbed ClONO2 and N2O5 could react with HCl or hydrolyze on the NaCl surface to form HNO3 and other gaseous species (Cl2, HOCl and ClNO2). More importantly, ClONO2 and N2O5 can actively participate in the heterogeneous reaction with NaCl to generate Cl2 and ClNO2, which could be photolyzed to produce reactive chlorine atoms. The adsorbed H2O molecule plays an important role in the key elementary step and promotes the formation of the products. The present results rationalize previous experimental studies well and enrich our understanding of the source of halogen atoms in the marine boundary layer.


1. Introduction

Approximately 70% of the earth's surface is covered by oceans, making the exchange of marine aerosols and trace gases between the marine boundary layer (MBL) and the troposphere potentially important for atmospheric processes.1 Sea salt aerosols, which are generated by wind-induced wave action and bubble bursting of seawater, contribute to most of the marine aerosol present in the atmosphere, with uncertain annual emission estimates ranging from 0.3 Pg to 30 Pg.2,3 During the last few decades, growing interest has been given to understanding of the fundamental chemical and physical processes in the marine boundary layer, especially in the heterogeneous reactions of nitrogen oxides such as NO2, N2O5 and ClONO2 with sea salt aerosols, with the goal to reveal the role they play in atmospheric chemistry.4–7 It has been proposed that the heterogeneous reactions taking place on or in sea salt aerosols may provide the main possible sources for gaseous chlorine-containing species (including HCl, Cl2, ClNO2, and HOCl etc.). Upon photolysis, the inert halogen reservoir species can convert into active halogen atoms, which may affect the ozone balance and oxidative potential of the local atmosphere, specifically in the marine boundary layer.

Of particular interest is the heterogeneous loss of ClONO2 and N2O5 on the surface of solid NaCl (major component of sea salts aerosols). ClONO2 is one of the most important temporary reservoir species for reactive chlorine. For example, in the winter, ClONO2 accounts for approximately 50% of the total inorganic chlorine (Cly = HCl + ClO + HOCl + ClONO2) at mid-northern latitudes stratosphere.8,9 N2O5 represents a significant reactive intermediate in the atmospheric chemistry of nitrogen oxides and nitrate aerosol, especially during nighttime.10 Potential fates of ClONO2 and N2O5 in the troposphere include deposition as well as reacting with HCl or their hydrolysis on surfaces. Furthermore, the gaseous ClONO2 and N2O5 could react directly with NaCl according to:

 
ClONO2 + NaCl → Cl2 + NaNO3 (1)
 
N2O5 + NaCl → ClNO2 + NaNO3 (2)

In the atmosphere, the gaseous products Cl2 and ClNO2 will photolyze rapidly (λ > 290 nm) and generate highly reactive chlorine atoms.11 Chlorine atoms are known to have a significant impact on the formation and fate of tropospheric ozone. It can react with O3 or react readily with organics in a manner similar to OH. In the presence of NOx, the Cl-organic reaction may result in the formation of O3, rather than its destruction.

Several experimental studies have been devoted to investigating the heterogeneous interactions of nitrogen oxides with NaCl. In 1994, Keyser and co-workers performed the quantitative study on the reaction of ClONO2 with NaCl over a temperature range 220–300 K. They found Cl2 was the sole product, and the determined uptake coefficient at 296 K was γ = (4.6 ± 3.0) × 10−3 on dry NaCl substrates. With added H2O, a trace of HOCl was observed in addition to the Cl2 product.12 Subsequently, in the experiments of Gebel et al. and Zelenov et al., the reactions have been further confirmed. Two possible reaction paths were taken into consideration when water was present, ClONO2 could react with NaCl to form Cl2 and NaNO3, or hydrolyze on the surface to generate HOCl and HNO3.5,13 The heterogeneous reaction of N2O5 with NaCl has been investigated by different techniques.10,11,14,15 Results indicated that the uptake coefficients (γ) were in a wide range of 10−4 to 10−2, and increased with increase in relative humidity (RH). The importance of water vapor in the real atmosphere has been proposed and demonstrated. Like ClONO2, two different processes were also founded during the reaction of N2O5 with deliquesced NaCl or solid NaCl that holds surface-adsorbed water (SAW), both ClNO2 and HNO3 were observed as gaseous products from the N2O5–salt interactions.4,10

In order to assess the source and role of halogen atoms, it is important to elucidate the kinetics, products, and mechanisms for reactions of nitrogen oxides with sea salt, especially for its major component, NaCl. However, while the overall kinetics and products are well-known by experiments,4,5,10–15 there have been no studies conducted to the interaction mechanisms between ClONO2 or N2O5 and NaCl. The experimental analysis helped to establish that the (100) surface of NaCl is dominant. Herein, we chose the (100) surface to study the adsorption and heterogeneous reactions of ClONO2 and N2O5 using the density functional theory (DFT) slab calculations. This theoretical method has been successfully applied in the environmental investigations and can provide accurate predictions for the reaction mechanism through energy calculations.16–18 Indices of crystal face is the inverse ratio of intercept coefficient for crystal face in three crystallographic axis (a, b and c). NaCl (100) surface is the crystal face, which is parallel to the two crystal axis (b, c), and the intercept in axis a is C. That is equivalent to the plane X = C when in the XYZ space. Stable adsorption structures as well as transition states for both clean NaCl (100) surface and water adsorbed NaCl have been determined. Particular attention was paid to explore the role of water in the formation of chlorine-containing species. We anticipate that results from the study of ClONO2 and N2O5 adsorption and reaction on the ideal NaCl (100) surface will provide some insights into chemistry of the aerosol surfaces.

2. Computational methods

All calculations were performed in the framework of DFT by using the program package DMol3 in Materials Studio (version 6.0) of Accelrys Inc.19 The exchange–correlation energy was calculated within the GGA–PW91 approximation.20,21 The double-numeric quality basis set with polarization functions (DNP) was used in the calculations. The fine quality mesh was used for numerical integration, in which the tolerances of energy, gradient, displacement, and self-consistent field (SCF) convergence were 1 × 10−5 hartree, 2 × 10−3 hartree Å−1, 5 × 10−3 Å, and 1 × 10−6, respectively. The real space cutoff radius was maintained as 5.2 Å. On the basis of the convergence test, the surface Brillouin zone was sampled with a 2 × 2 × 1 Monkhorst–Pack k-point grid.22

The (100) surface of NaCl was modeled by a p (3 × 3) supercell with a dimension of 12.10 Å × 12.10 Å × 23.55 Å. Thirty-six NaCl molecular units in the slab were distributed in four layers. The vacuum region separating the slabs along the z direction was set to 15 Å. In all calculations, the bottom two layers were kept fixed in their bulk lattice positions, while the top two layers together with the adsorbates were allowed to relax. Fig. 1 shows the super cell model used in this work. The flow of charges was estimated by Mulliken population analysis, which has been shown to be a useful tool.23,24 Spin polarization was applied to all calculations.


image file: c6ra03961h-f1.tif
Fig. 1 Structures of the clean NaCl (100) surface, ClONO2 and N2O5. O atoms are red, N atoms are blue, Cl atoms are green and Na atoms are purple. Angles are in degree.

Following usual convention, the adsorption energy Ead was defined as:

 
Ead = Etotal − (Eadsorbate + Esurf) (3)
where Etotal, Eadsorbate, and Esurf represent the energies of the adsorbed system, free adsorbate, and the clean surface, respectively. With this definition, a negative Ead value denotes an exothermic adsorption process, and more negative Ead shows stronger adsorption between free adsorbate and the clean surface.

To determine the activation energy for a specific reaction path, a transition state (TS) connecting two structures through a minimum energy path was identified by complete synchronous transit (LST) and quadratic synchronous transit (QST) search methods. Each TS was confirmed by vibrational frequency analysis with one and only one imaginary frequency. The energy barrier (ΔEb) and reaction heat (ΔEr) of each elementary step were calculated by:

 
ΔEb = ETSER (4)
 
ΔEr = EPER (5)
where ETS, ER, and EP are the energies of the transition state, reactant, and product, respectively.

3. Results and discussion

To test the reliability of the theoretical calculations, we compared several calculated results with available experimental values. (i) The optimized unit cell parameter of the bulk NaCl is 5.597 Å, while the experimental value is 5.640 Å.25 (ii) The geometrical parameters of H2O, ClONO2 and N2O5 were calculated at the GGA–PW91 level. As shown in Fig. S1 of ESI, the values match well with the available experimental data, and the maximum relative error is less than 2.0%.26–28 (iii) The calculated adsorption energy of H2O on the NaCl (100) surface is −13.99 kcal mol−1, while previous experimental findings are in a range of −13.87 to −11.58 kcal mol−1.29,30 All the calculated data show good agreement with the corresponding reference values, indicating the acceptable accuracy and reliability of the method employed.

3.1 Adsorption of ClONO2 and N2O5 on the NaCl (100) surface

ClONO2 is a planar, near prolate molecule with a free electron pair residing on the N atom.31 In order to get an insight into the nature of ClONO2–substrate interaction, we have carried out an extensive study for ClONO2 molecule on NaCl (100) by considering different adsorption modes (flat and vertical) and various adsorption sites (top, bridge, etc.). The optimized stable adsorption configurations and key structural parameters are shown in Fig. S2 of ESI. Table 1 lists the corresponding adsorption energies (Ead) calculated by DMol3. As displayed in Fig. S2, configurations I-1 to I-16 represent the vertical manner with single-atom adsorption, where the ClONO2 molecule could either adsorb at the Cl top site (Ead, −7.35 to −5.08 kcal mol−1) or adsorb at the Na+ top site (Ead, −9.45 to −7.32 kcal mol−1). According to the definitions of Eads, the calculated adsorption energies on Na+ site are more negative than that on Cl site, implying that the ClONO2–Na+ interaction is much stronger than the ClONO2–Cl interaction. Configurations II-1 to II-16 are the vertical manner with double-atom adsorption. Two atoms of ClONO2 are bonded with the surface in a bridge configuration, including the Cl–Cl, Na–Na and Na–Cl sites. Table 1 reveals that the ClONO2 adsorption at Na–Na bridge site (II-2, II-6, and II-10) is more energetically favorable than at the Cl–Cl bridge site (II-1, II-5 and II-9) and Na–Cl bridge site (II-3, II-4, II-7, II-8, II-11 and II-12). Configurations III-1 to III-7 stand for the flat manner with multi-atom adsorption. ClONO2 can be horizontally adsorbed on NaCl (100) surface via three or more atoms as illustrated in Fig. S2.
Table 1 Adsorption energy (Ead) and Mulliken charges (e) for various adsorption configurations of ClONO2 on the NaCl (100) surface
Species Ead (kcal mol−1) Charges Species Ead (kcal mol−1) Charges
I-1 −6.76 −0.012 II-5 −7.15 −0.015
I-2 −7.35 −0.014 II-6 −9.25 −0.002
I-3 −9.45 −0.001 II-7 −8.16 −0.002
I-4 −8.22 −0.002 II-8 −8.60 −0.008
I-5 −5.08 −0.004 II-9 −7.00 −0.017
I-6 −5.11 −0.005 II-10 −9.11 −0.002
I-7 −7.32 0.011 II-11 −7.54 −0.011
I-8 −7.33 0.009 II-12 −8.36 −0.011
I-9 −5.70 −0.005 II-13 −12.67 −0.172
I-10 −6.26 −0.001 II-14 −8.00 −0.001
I-11 −8.75 0.010 II-15 −14.81 −0.227
I-12 −8.46 0.008 II-16 −8.57 −0.005
I-13 −6.93 −0.174 III-1 −9.95 −0.004
I-14 −6.66 −0.173 III-2 −10.11 −0.001
I-15 −8.46 −0.015 III-3 −9.42 0.001
I-16 −8.21 −0.014 III-4 −10.19 −0.002
II-1 −5.52 −0.009 III-5 −8.95 −0.002
II-2 −9.67 0.004 III-6 −10.87 0.001
II-3 −8.49 0.009 III-7 −9.03 0.000
II-4 −8.05 0.010 III-8 −10.30 −0.008


In all the optimized ClONO2-adsorbed structures, the bridgelike configuration II-15 is found to be most stable with the adsorption energy of −14.81 kcal mol−1. In II-15, ClONO2 adsorbs on NaCl (100) in an upright fashion, where the O atom of ClONO2 binds to the surface Na+ ion, and the Cl atom of ClONO2 attaches to an adjacent Cl ion. The distances of the newly formed O–Na and Cl–Cl bonds are 2.695 and 2.548 Å, respectively. The Mulliken population analysis32 was performed to understand the charge redistribution upon ClONO2 adsorption on NaCl. The results are summarized in Table 1. It's clearly show that, in configurations II-15, the adsorbed ClONO2 molecule is partially negatively charged (−0.227|e|), indicating that ClONO2 acts as a Lewis acid and accepts the electrons from the NaCl surface.

N2O5 is a flat-shaped molecule, with C2 symmetry.28 Studies show that N2O5 can either lie flat or stand vertically on the surface. Twenty-one different adsorption configurations of N2O5 molecule on the NaCl (100) surface were explored. The optimized structures and calculated adsorption energies (Ead) are summarized in Fig. S3 and Table 2. Configurations I-1′ to I-14′ represent the vertical adsorption manners with single or multiple atom adsorptions. In these configurations, N2O5 molecule is initially placed perpendicular to the NaCl surface. According to the calculated adsorption energies in Table 2, N2O5 preferentially adsorbs on the Na+ site over the Cl site, which is similar to the property of ClONO2 adsorption on the NaCl (110) surface. Configurations II-1′ to II-7′ are the parallel adsorption manners with multi-atom adsorption. The plane of the N2O5 molecule is nearly parallel to the surface, with adsorption energies ranging from −13.67 to −11.21 kcal mol−1. The parallel adsorption has stronger interaction with NaCl surface than the vertical adsorption.

Table 2 Adsorption energy (Ead) and Mulliken populations (e) for various adsorption configurations of N2O5 on the NaCl (100) surface
Species Ead (kcal mol−1) Charges Species Ead (kcal mol−1) Charges
I-1′ −9.00 −0.011 I-12′ −9.21 −0.015
I-2′ −9.71 −0.008 I-13′ −12.08 0.004
I-3′ −11.86 −0.008 I-14′ −12.24 −0.004
I-4′ −9.64 −0.013 II-1′ −12.96 −0.003
I-5′ −9.07 −0.013 II-2′ −11.77 −0.015
I-6′ −9.30 −0.014 II-3′ −11.21 −0.021
I-7′ −10.30 −0.011 II-4′ −13.42 −0.013
I-8′ −9.76 −0.008 II-5′ −12.98 −0.003
I-9′ −8.29 −0.013 II-6′ −11.24 −0.019
I-10′ −12.56 0.003 II-7′ −13.67 −0.002
I-11′ −11.50 0.007      


Comparing above different adsorption models of N2O5 on the NaCl (100) surface, structures (II-4′) and (II-7′) are identified as the favorable adsorption configurations. The calculated adsorption energies for the two structures are −13.42 and −13.67 kcal mol−1, respectively. Mulliken population analysis shows that the net electronic charges transferred from the surface to N2O5 in (II-4′) and (II-7′) are −0.013|e| and −0.002|e|, respectively.

3.1.1 Reactions of ClONO2 and N2O5 with HCl or H2O on the NaCl (100) surface. The chemistry of ClONO2 and N2O5 with HCl or their hydrolysis has been recognized as a major contributor to ozone depletion, especially in the Antarctic stratosphere. Both experimental and theoretical investigations suggested that these reactions cannot occur in the gas phase easily and thus must be heterogeneous. Water clusters, ice, sulfate aerosols, as well as the sea salt aerosols, have been demonstrated to promote these processes.4,33–35 In order to understand the catalytic effect of sea salt aerosols, we explored the relevant reaction mechanisms of ClONO2 and N2O5 on the NaCl (100) surface.

Previous studies indicate that the interaction between ClONO2 and HCl or H2O molecule are prone to present planar bimolecular reactions, thus we chose the flat configuration III-6 in the following researches.35,36 For N2O5, considering the effect of steric hindrance on the reactions, the stable configuration II-4′ is selected for the following discussions. The optimized geometries of intermediates and transition states involved in these reactions on the NaCl (100) surface are shown in Fig. 2. Fig. 3 is the corresponding profile of the potential energy surface (PES), where the sum of the energies of the isolated molecule (ClONO2 or N2O5) and the clean NaCl surface is taken as zero energy. In order to show the zero-point vibrational energy contributions to the barriers and reaction energies, the reaction schemes labeled with energy barriers (ΔEb) and reaction heats (ΔEr) are supplied in Fig. S5 of ESI.


image file: c6ra03961h-f2.tif
Fig. 2 Optimized geometries for the intermediates and the transition states involved in the reactions on the NaCl (100) surface. Distances are in angstroms.

image file: c6ra03961h-f3.tif
Fig. 3 Calculated potential energy surface profiles for the reactions on the NaCl (100) surface.

As show in Fig. 2, HCl molecule could adsorb on the NaCl surface firstly, combine with ClONO2 and form a ClONO2–HCl complex (IM1). The distance of H–Cl bond is 2.078 Å. Then via a six-membered cyclic transition state TS1, the products coadsorbed HNO3 and Cl2 can be obtained. In TS1, the O–Cl bond in ClONO2 and H–Cl bond in HCl stretch synchronously, and the adjacent O–H and Cl–Cl bond lengths decreased to 1.462 Å and 2.683 Å, respectively. The energy barrier is calculated to be 5.42 kcal mol−1, which is notably lower than that occurs in pure gas phase (40.90–64.00 kcal mol−1).35 If give the energy of 9.26 kcal mol−1 to P1, desorption reaction can be occur to get “free products” P1-p. The optimized geometries of the “free products” are displayed in Fig. S4 of ESI.

H2O molecule could also adsorb on the NaCl surface, combine with ClONO2 and form a complex IM2. This process is strongly exothermic and releases energy of −18.22 kcal mol−1. The O–Cl⋯O in IM2 has an almost linear configuration. Upon complexation, the hydrolysis reaction on the surface may happen subsequently. TS2 is the transition state corresponding to this process, in which the H atom of H2O transfers to the O atom of ClONO2, meanwhile the Cl atom of ClONO2 migrates to the O atom of H2O. The optimized structure and its key parameters are depicted in Fig. 2. The products HNO3 and HOCl are still absorbed on the NaCl surface. As can be seen from the PES profile (Fig. 3), the reaction could readily occur on NaCl surface. This result is consistent with the experimental studies of Timonen et al. and Zelenov et al.5,12

The two different reactions with regard to N2O5 on the NaCl (100) surface are exhibited in Fig. 3. Unlike the reaction of ClONO2, HCl molecule doesn't need to adsorb on the NaCl surface, the reaction of N2O5 with HCl proceeds via the formation of a complex (named IM3). The bond lengths of O–H and Cl–N in IM3 are 2.040 and 3.480 Å, respectively. Once formed, IM3 can perform H–Cl and O–N bonds cleavage through transition state TS3 lead to the formation of ClONO2 and adsorbed HNO3. As shown in the structure of TS3, the distances of O–H and Cl–N bond have been significantly shortened to 1.462 and 2.699 Å, respectively. Energy barrier for this transformation is found to be only 4.80 kcal mol−1, indicating facile occurrence of such a process. The overall reaction is calculated to be exothermic by 15.67 kcal mol−1.

The hydrolysis of N2O5 on particle surfaces including NaCl is considered as a major source of nitric acid aerosols in the atmosphere. As shown in Fig. 2, IM4 is the reactant complex of N2O5 with H2O, which is calculated to be more stable by 9.42 kcal mol−1 than the separated reactants. Along this channel, IM4 is converted to two HNO3 via transition state TS4, with a lower barrier of 5.75 kcal mol−1. In TS4, the breaking N–O bond is elongated to 2.344 Å and the forming H–O and O–N bonds of HNO3 decreased to 1.428 Å and 1.990 Å, respectively. The adsorbed HNO3 has been observed experimentally from the heterogeneous interactions of N2O5 with NaCl aerosol.4,10 The homogeneous gas-phase reaction of N2O5 with H2O is slow and considered to be relatively unimportant, while the heterogeneous reaction on NaCl surface is more easier and can occur rapidly. In the subsequent reaction, the product HNO3 can react further with NaCl to form gaseous HCl.37

3.1.2 H2O adsorption and reconstruction on the NaCl (100) surface. Water is a ubiquitous atmospheric constituent and plays a key role in many atmospheric heterogeneous chemical reactions. In particular, its adsorption on salt aerosols is an important process that can affect their surface structure and chemical reactivity.38 Recent studies have demonstrated the catalytic effect of water during the interaction between NaCl and gaseous pollutants, such as ClONO2, N2O5, and NO2.10,13 However, the role of adsorbed water in the reaction mechanisms at the molecular level is not well understood. Thus, we proposed a possible H2O adsorption and reconstruction mechanism on the NaCl surface. As shown in Fig. 4, water is preferably adsorbed on the NaCl (100) surface, producing a partially hydrated NaCl surface (named IM5), in which the water molecule locates near the top site with its O atom adjacent to Na+ and its H atoms attracted by the nearest Cl. The distance of the newly formed O–Na bond is 2.399 Å. The calculated adsorption energy of H2O on the NaCl (100) surface is −13.99 kcal mol−1 (∼0.61 eV), which is in reasonable agreement with the previous experimental results of 0.50–0.60 eV.29,30 Subsequently, the adsorbed H2O molecule can enter into NaCl lattice via transition state TS5 with a barrier of 14.03 kcal mol−1 and form intermediate IM6. In TS5, the O–Na bond has been slightly stretched to 2.470 Å and the adjacent Cl ion emerges from the plane. IM6 is the reconstructed NaCl configuration, where the water molecule embedded into the bulk NaCl, and the Cl ion move upward about 1.610 Å from the surface. Apparently, the raised chloride ion could act as the reactive component of the NaCl and positively participate in further reactions.
image file: c6ra03961h-f4.tif
Fig. 4 Potential energy surface (PES) profile with the optimized geometries of intermediate, transition state, and product for H2O adsorption and reconstruction on the NaCl (100) surface. Distances are in angstroms.
3.1.3 Heterogeneous reactions of ClONO2 and N2O5 with the reconstructed NaCl (100) surface. There are four different Na+ ions around the adsorbed H2O molecule and the raised Cl ion, which are labeled as Na1+, Na2+, Na3+ and Na4+. Calculations show that, ClONO2 and N2O5 molecule preferentially adsorb on the unrestricted Na1+ or Na2+ ions. Starting from these stable adsorption structures, we have searched reasonable mechanisms for ClONO2 and N2O5 with the reconstructed NaCl (100) surface. Fig. 5 shows the optimized geometries of intermediates and transition states involved in the heterogeneous reactions, and Fig. 6 is the calculated potential energy surface (PES) profile. The reaction schemes labeled with energy barriers (ΔEb) and reaction heats (ΔEr) are supplied in Fig. S7 of ESI. For convenience, the oxygen atoms in ClONO2 and N2O5 have been numbered. As seen in Fig. 5, ClONO2 molecule adsorbs on the NaCl surface in a slant fashion (IM7), where the O1 atom binds to the Na1+ site. The newly generated O1–Na1 bond is 2.442 Å, and the calculated adsorption energy is −20.37 kcal mol−1. After adsorption, the Na1+ ion in IM7 shifts 0.10 Å upward with respect to the original position on the surface. Then, the ClONO2 moiety tilts toward the surface through its O3 atom binding to adjacent Na2+ ion, forming a bridgelike configuration IM8. In IM8, ClONO2 was adsorbed at the Na1–Na2 site through O1–Na1 and O3–Na3 bonds with the distances of 2.368 Å and 3.787 Å, respectively. From the calculated PES profile, IM8 is calculated to 0.24 kcal mol−1 higher than IM7. Next, the bidentate configuration IM8 can be further converted into products NaNO3 and Cl2 through transition state TS6, where the O–Cl bond is elongated to 2.616 Å and Cl–Cl bond is shortened to 2.122 Å. This process is exothermic by 5.35 kcal mol−1 and needs to overcome a barrier of 5.36 kcal mol−1. The generated product Cl2 adsorbs on the NaCl (100) surface in a tilted mode with two adjacent Na+ ions. ClONO2 molecule can also adsorb on the Na2+ site, resulting in a unidentate configuration IM9, in which the O2–Na2 distance is 2.448 Å. After absorbing the energy of 0.45 kcal mol−1, ClONO2 is found to bind the surface with two O–Na bonds. The further conversion of the adsorbed ClONO2 in IM10 is shown in Fig. 5. In this process, the Cl atom is transferred from the adsorbed ClONO2 to the raised Cl ion through a transition state TS7, producing NaNO3 and adsorbed Cl2 (P6). This process is exothermic by 5.35 kcal mol−1 with a low energy barrier of 4.38 kcal mol−1. As shown in Fig. 6, adsorbed Cl2 could deviate from NaCl easily after absorb little energy. The product Cl2 is a photochemical labile species, given its absorption cross section in the actinic region, it will photolyze rapidly and generate highly active chlorine atom, which may affect the fate of tropospheric ozone and the oxidative balance of the atmosphere.13
image file: c6ra03961h-f5.tif
Fig. 5 Optimized geometries for the intermediates and the transition states involved in the heterogeneous reactions on the reconstructed NaCl (100) surface. Distances are in angstroms.

image file: c6ra03961h-f6.tif
Fig. 6 Calculated potential energy surface profiles for the heterogeneous reactions on the reconstructed NaCl (100) surface.

Similar to ClONO2, there are two different Na+ ions (Na1+ and Na2+) participate in the following reactions of N2O5 with the reconstructed NaCl surface. As displayed in Fig. 5, N2O5 molecule initially binds to the Na1+ ion through its O3 atom, forming configuration IM11. PES profile shows that this adsorption process is exothermic by 14.35 kcal mol−1. The distance of formed O3–Na1 bond is 2.972 Å. The adsorption of N2O5 causes the Na1+ ion to relax downward by 0.07 Å along the z direction. Next, O1 atom of N2O5 could weakly bind to adjacent Na2+ ion and transform into configuration IM12, where O1 and O3 coadsorbed onto the NaCl surface. It is a barrierless process and slightly endothermic by 0.99 kcal mol−1. The structure of the adsorbed N2O5 in IM12 is distorted significantly with respect to that of the free N2O5 molecule. Subsequently, IM12 can react with raised Cl ion and perform the rupture of N–O3 bond and the formation of N–Cl bond, finally resulting in NaNO3 and absorbed ClNO2. As shown in Fig. 5, the distance between O3 and N increased to 2.147 Å from 1.704 Å and the N–Cl bond length significantly decreased to 2.511 Å from 3.250 Å. This process has a potential barrier of 4.14 kcal mol−1 and is exothermic by 10.63 kcal mol−1. The low energy barrier and the strong exothermicity of the reaction indicate that the formation of ClNO2 on the NaCl (100) surface is both kinetically and thermodynamically favorable. IM13 represents the stable configuration for N2O5 adsorption on the Na2+ site with an adsorption energy of 14.29 kcal mol−1, and it could further rearrange to a bidentate configuration IM14. Once formed, IM14 would convert into NaNO3 and ClNO2 through breaking the N–O3 bond and forming the new N–Cl bond. The barrier associated with the reaction is found to be 3.78 kcal mol−1. And the overall reaction is calculated to be exothermic by 9.91 kcal mol−1. The nitrate species NaNO3 is a common product of sea salt aerosols with nitrogen oxides. Gaseous molecule ClNO2 can subsequently be photo-dissociated by sunlight to give chlorine atoms and NO2.

As discussed above, the energy barriers for the formation of Cl2 and ClNO2 are low (3.78–5.36 kcal mol−1), and the exothermicity are large (5.35–10.63 kcal mol−1), which demonstrates that with the aid of water the heterogeneous reactions of ClONO2 and N2O5 with the reconstructed NaCl (100) surface are facile. According to previous experiments, the heterogeneous reactions could occur in a few minutes.4,5,13 The results are in accordance with the experimental discoveries, and further complement the experiment researches.10,13 We conjecture that the suggested mechanism involved with water could be applied to other similar heterogeneous reaction of sea salt aerosols with other nitrogen oxides.

4. Conclusion

In this paper, the detailed adsorption modes and heterogeneous reaction mechanism of ClONO2 and N2O5 on the NaCl (100) surface have been systematically analyzed by the density functional theory (DFT) method (periodic DMol3). Several valuable findings are obtained:

(1) Various adsorption configurations for ClONO2 and N2O5 on the clean NaCl (100) surface have been explored. For ClONO2, the vertical adsorption mode II-15 is the most stable with the adsorption energy of −14.81 kcal mol−1, and for N2O5, the horizontal adsorption mode II-7′ is determinately the most favorable with the adsorption energy of −13.67 kcal mol−1.

(2) On the NaCl (100) surface, adsorbed ClONO2 and N2O5 could react with HCl or via hydrolyze to form chlorine-containing species (Cl2, HOCl and ClNO2) and nitric acid. Acting as a catalyst, NaCl can promote these reactions.

(3) Possible heterogeneous reaction mechanism of ClONO2 and N2O5 with the reconstructed NaCl (100) surface in the presence of adsorbed water was proposed. Water molecule plays an important role for the formation of active chloride ion and insures the consequent reactions. Cl2, ClNO2 as well as NaNO3 are the relevant products. The present results may help to improve our understanding of the interaction between nitrogen oxides and sea salt aerosols.

Acknowledgements

This work was supported by NSFC (National Natural Science Foundation of China, Project No. 21337001, 21177076).

References

  1. N. K. Richards and B. J. Finlayson-Pitts, Environ. Sci. Technol., 2012, 46, 10447–10454 CrossRef CAS PubMed.
  2. E. R. Lewis and S. E. Schwartz, Sea salt aerosol production: mechanisms, methods, measurements, and models, American Geophysical Union, Washington, DC, 2004 Search PubMed.
  3. M. Spada, O. Jorba, C. P. García-Pando, Z. Janjic and J. M. Baldasano, Atmos. Environ., 2015, 101, 41–48 CrossRef CAS.
  4. R. C. Hoffman, M. E. Gebel, B. S. Fox and B. J. Finlayson-Pitts, Phys. Chem. Chem. Phys., 2003, 5, 1780–1789 RSC.
  5. V. V. Zelenov, E. V. Aparina, S. A. Kashtanov, D. V. Shestakov and Y. M. Gershenzon, J. Phys. Chem. A, 2006, 110, 6771–6780 CrossRef CAS PubMed.
  6. N. K. Richards and B. J. Finlayson-Pitts, Environ. Sci. Technol., 2012, 46, 10447–10454 CrossRef CAS PubMed.
  7. C. Zhang, X. Zhang, L. Kang, N. Wang, M. Wang, X. Sun and W. Wang, Sci. Total Environ., 2015, 524, 195–200 CrossRef PubMed.
  8. J. A. Thornton, J. P. Kercher, T. P. Riedel, N. L. Wagner, J. Cozic, J. S. Holloway, W. P. Dubé, G. M. Wolfe, P. K. Quinn and A. M. Middlebrook, Nature, 2010, 464, 271–274 CrossRef CAS PubMed.
  9. J. Orphal, M. Morillon-Chapey, A. Diallo and G. Guelachvili, J. Phys. Chem. A, 1997, 101, 1062–1067 CrossRef CAS.
  10. W. L. Chang, P. V. Bhave, S. S. Brown, N. Riemer, J. Stutz and D. Dabdub, Aerosol Sci. Technol., 2011, 45, 665–695 CrossRef.
  11. B. J. Finlayson-Pitts and J. C. Hemminger, J. Phys. Chem. A, 2000, 104, 11463–11477 CrossRef CAS.
  12. R. S. Timonen, L. T. Chu, M. Leu and L. F. Keyser, J. Phys. Chem., 1994, 98, 9509–9517 CrossRef CAS.
  13. M. E. Gebel and B. J. Finlayson-Pitts, J. Phys. Chem. A, 2001, 105, 5178–5187 CrossRef CAS.
  14. F. F. Fenter, F. Caloz and M. J. Rossi, J. Phys. Chem., 1996, 100, 1008–1019 CrossRef CAS.
  15. D. J. Stewart, P. T. Griffiths and R. A. Cox, Atmos. Chem. Phys., 2004, 4, 1381–1388 CrossRef CAS.
  16. W. Ji, Z. Shen, Q. Tang, B. Yang and M. Fan, Chem. Eng. J., 2016, 289, 349–355 CrossRef CAS.
  17. C. Zhang, X. Zhang, L. Kang, N. Wang, M. Wang, X. Sun and W. Wang, Sci. Total Environ., 2015, 524–525, 195–200 CrossRef CAS PubMed.
  18. H. Xu, W. Chu, X. Huang, W. Sun, C. Jiang and Z. Liu, Appl. Surf. Sci., 2016 DOI:10.1016/j.apsusc.2016.01.236.
  19. Accelrys Software Inc., Material Studio Modeling Environment, Release 6.0, Accelrys Software Inc., San Diego, 2011, p. 36 Search PubMed.
  20. J. P. Perdew and W. Yue, Phys. Rev. B: Condens. Matter Mater. Phys., 1986, 33, 8800–8802 CrossRef.
  21. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244–13249 CrossRef.
  22. H. J. Monkhorst and J. D. Pack, Phys. Rev. B: Condens. Matter Mater. Phys., 1976, 13, 5188–5192 CrossRef.
  23. Z. Liu, J. Li, S. I. Woo and H. Hu, Catal. Lett., 2013, 143, 912–918 CrossRef CAS.
  24. Y. Li, W. Guo, H. Zhu, L. Zhao, M. Li, S. Li, D. Fu, X. Lu and H. Shan, Langmuir, 2012, 28, 3129–3137 CrossRef CAS PubMed.
  25. J. E. Nickels, M. A. Fineman and W. E. Wallace, J. Phys. Chem., 1949, 53, 625–628 CrossRef CAS PubMed.
  26. A. R. Hoy and P. R. Bunker, J. Mol. Spectrosc., 1979, 74, 1–8 CrossRef CAS.
  27. B. Casper, P. Lambotte, R. Minkwitz and H. Oberhammer, J. Phys. Chem., 1993, 97, 9992–9995 CrossRef CAS.
  28. B. W. McClelland, L. Hedberg, K. Hedberg and K. Hagen, J. Am. Chem. Soc., 1983, 105, 3789–3793 CrossRef CAS.
  29. L. W. Bruch, A. Glebov, J. P. Toennies and H. Weiss, J. Chem. Phys., 1995, 103, 5109–5120 CrossRef CAS.
  30. B. Li, A. Michaelides and M. Scheffler, Surf. Sci., 2008, 602, L135–L138 CrossRef CAS.
  31. R. A. Butler, D. T. Petkie, P. Helminger, F. C. De Lucia and Z. Kisiel, J. Mol. Spectrosc., 2007, 243, 1–9 CrossRef CAS.
  32. R. S. Mulliken, J. Chem. Phys., 1955, 23, 1833–1840 CrossRef CAS.
  33. Q. Shi, J. T. Jayne, C. E. Kolb, D. R. Worsnop and P. Davidovits, J. Geophys. Res., 2001, 106, 24259–24274 CrossRef CAS.
  34. R. L. Apodaca, D. M. Huff and W. R. Simpson, Atmos. Chem. Phys., 2008, 8, 7451–7463 CrossRef CAS.
  35. K. Nam and Y. Kim, J. Chem. Phys., 2009, 130, 144310–144320 CrossRef PubMed.
  36. Y. Kim, C. Park and K. H. Kim, Bull. Korean Chem. Soc., 2005, 26, 1953–1961 CrossRef CAS.
  37. B. N. Fong, K. V. Newhouse and H. Ali, React. Kinet., Mech. Catal., 2015, 116, 273–283 CrossRef CAS.
  38. Q. Dai, J. Hu and M. Salmeron, J. Phys. Chem. B, 1997, 101, 1994–1998 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Supplementary material contains Fig. S1–S7. See DOI: 10.1039/c6ra03961h

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.