Guangliang Chena,
Mingqiang Huang*a,
Guangzhen Gaob,
Xingqiang Liuc,
Changjin Hud,
Weixiong Zhaod,
Xuejun Gud and
Weijun Zhangd
aFujian Provincial Key Laboratory of Modern Analytical Science and Separation Technology, College of Chemistry & Chemical Engineering and Environment, Minnan Normal University, Zhangzhou 363000, China. E-mail: huangmingqiang@mnnu.edu.cn
bCollege of Physics and Electronic Engineering, Jiangsu Normal University, Xuzhou 221116, China
cSchool of Environment Science & Engineering, Tan Kah Kee College, Xiamen University, Zhangzhou 363105, China
dLaboratory of Atmospheric Physico-Chemistry, Anhui Institute of Optics and Fine Mechanics, Chinese Academy of Sciences, Hefei 230031, China
First published on 3rd September 2025
Nitrous acid (HONO) is a vital pollutant gas and the nitrogen-containing organic compounds (NOCs) produced by its reaction are the main components of aerosols. The reaction mechanisms and kinetics of HONO and the simplest aromatic Criegee intermediate (PhCHOO) are investigated by density functional theory and transition state theory in this study. The results demonstrate that cycloaddition of HONO and PhCHOO to form heteroozonide with the highest activation energy and smallest rate constant does not easily occur. Pathways of oxygen atom transfer and cycloaddition can form in situ HNO3 and benzoic acid. Meanwhile, the hydrogen atom transfer pathway results in the generation of phenyl hydroperoxide methyl nitrite (Ph-HPMN), which has the lowest activation energy, dominating the reaction between HONO and PhCHOO with a rate constant (5.68 × 10−13 cm3 per molecule per s) close to that with OH radicals (4.83 × 10−12 cm3 per molecule per s). These results provide a theoretical reference for clarifying the mechanism of generation of NOCs formed from ozonolysis of styrene and other olefin compounds in the presence of HONO.
Styrene, which is emitted into the atmosphere by human activities such as solvent use and automobile exhaust, is the second most highly reactive aromatic volatile organic compound (AVOC) in the field of environmental chemistry, second only to toluene.13,14 Styrene itself is toxic and has potential mutagenic and carcinogenic properties. Exposure to styrene may cause damage to the central nervous system and reproductive system of the human body.15 In addition, when emitted into the atmosphere, styrene mainly undergoes oxidation reactions with ozone. Ozone is first added to the vinyl CC bond by 1,3-cycloaddition, generating a five-membered ring ozonation intermediate, which breaks C–C and O–O bonds to produce aldehydes, CH2OO and other Criegee intermediates.14,16,17 Researches have shown that except for isomerization and decomposition reactions, Criegee intermediates can stabilize through collision with other atmospheric species, and can undergo bimolecular reactions with trace gases such as H2O, SO2, NO, NO2, HNO3, etc.18–21 The rate constant of bimolecular reaction for CH2OO can be as high as 10−10 cm3 per molecule per s. For example, for the reaction of CH2OO + HNO3, its rate constant is 5.1 × 10−10 cm3 per molecule per s at 295 K.21 The reaction between CH2OO and HNO3 produces excited intermediate nitrooxymethyl hydroperoxide (NO3CH2OOH*), which undergoes cleavage to generate OH radical and CH2(O)NO3 as a major reaction pathway, with a branch ratio as high as 0.79.21
CH2(O)NO3 and other NOCs are important components of secondary organic aerosol (SOA) particles.22 These NOCs have strong stimulating effects and can cause brain and lung edema, induce DNA mutations, and pose a threat to health when absorbed by the human body.22–24 Also, NOCs containing the NO chromophore are a type of brown carbon with strong light absorption ability. They can absorb and scatter sunlight, resulting in reduced atmospheric visibility, and are one of the main culprits of haze pollution,22,25 which has attracted widespread attention. Although various aromatic NOCs have been detected in field measurement experiments, the formation mechanism of NOCs still needs to be further studied. Styrene is the second most highly reactive AVOC, and SOA generated by its reaction with ozone is an important part of atmospheric anthropogenic SOA.14,17 C6H5CHOO (PhCHOO) produced by styrene ozonation is the simplest aromatic Criegee intermediate. However, theoretical studies on the subsequent reaction of PhCHOO to produce NOCs have not been reported. Therefore, the generation process for PhCHOO produced by the ozonolysis of styrene is investigated in this study, and the formation mechanism of NOCs formed from the reaction of PhCHOO and HONO is further discussed in depth. These results provide a theoretical reference for the generation mechanism of NOCs produced by ozonolysis of styrene and other olefin compounds in the presence of HONO.
The traditional transition state theory (TST) combined with zero curvature tunneling is exploited to calculate rate constant (k(T)) of each reaction path:33,34
![]() | ||
Fig. 1 Optimized configurations at M06-2X/6-311++G(d,p) level and potential energy surface at CCSD(T)/6-311++G(3df,3pd) level for ozonolysis of styrene (bond lengths: Å). |
The generated C1 intermediate produces formaldehyde, benzaldehyde and Criegee intermediates (C6H5CHOO (PhCHOO) and CH2OO) by breaking the bonds of C–C and O–O through TS2A and TS2B subsequently. As illustrated in Fig. S1(b) and (c) in the SI, TS2A and TS2B are connected with corresponding reactants and products. The distance between two C atoms of vinyl in these two transition states is 1.882 Å and 1.900 Å, respectively, while the lengths of O⋯O are 1.976 Å and 1.981 Å, presenting the breaking trend of C–C and O–O bond (see Fig. 1). The imaginary frequencies of TS2A and TS2B are 546i and 557i cm−1, respectively, which correspond to breaking vibrations of C–C and O–O bonds. Compared to C1 intermediate, the energy barrier for producing formaldehyde and PhCHOO is 18.79 kcal mol−1, with reaction energy of −2.17 kcal mol−1. The energy barrier and reaction energy for generation of benzaldehyde and CH2OO are 22.62 and −0.50 kcal mol−1, respectively (see Fig. 1 and Table S4). Due to the low energy barrier and high reaction heat for producing formaldehyde and PhCHOO, it is the main reaction channel for ozonolysis of styrene. Therefore, the mechanism of the reaction of PhCHOO and HONO is investigated in the following sections.
HONO exists in two forms: trans- and cis-HONO. Calculated results in this work show that at the theoretical level for CCSD(T)/6-311++G(3df,3pd)//M06-2X/6-311++G(d,p), trans-HONO is 0.48 kcal mol−1 more stable than cis-HONO, which is relatively consistent with the research results reported by Guo et al.,37 indicating that the calculation method adopted in this work can provide accurate and reliable energy values. PhCHOO formed by the cleavage of C1 intermediate also has two forms: cis- and trans-PhCHOO. According to the theoretical calculation research results of Anglada et al.38 and Taatjes et al.,39 compared with syn-CH3CHOO, anti-CH3CHOO has a smaller steric hindrance and is more likely to form corresponding adducts with H2O, SO2, etc. The configurations of anti-PhCHOO and syn-PhCHOO optimized at the M06-2X/6-311++G(d,p) level are shown in Fig. 2. This is similar to the result of Du et al.,40 where anti-PhCHOO is 0.55 kcal mol−1 more stable than syn-PhCHOO at the level of CCSD(T)/6-311++G(3df,3pd)//M06-2X/6-311++G(d,p). It can be seen from Fig. 2 that the terminal O atom of syn-PhCHOO is close to the benzene ring, and the distance between it and a hydrogen atom in the benzene ring is 2.121 Å, forming an O⋯C–H hydrogen bond. Thus, syn-PhCHOO has a large steric hindrance and is not conducive to the reaction with HONO. On the contrary, the terminal O atom of anti-PhCHOO is far away from the benzene ring and has a smaller steric hindrance, making it easier to react with HONO. For the reactions of syn-PhCHOO with anti-HONO or cis-HONO, Gaussian 09 software30 is used to search for and optimize adducts and transition states with M06-2X/6-311++G(d,p) multiple times, but it failed to afford the corresponding adducts and transition states. Therefore, the trans-PhCHOO conformation is selected as the initial reactant to study the mechanism of its reaction with cis- and trans-HONO.
![]() | ||
Fig. 2 Optimized configurations at M06-2X/6-311++G(d,p) level for trans-PhCHOO and cis-PhCHOO (bond length: Å). |
![]() | ||
Fig. 3 Optimized configurations at M06-2X/6-311++G(d,p) level for oxygen atom transfer reaction of trans-PhCHOO and HONO (bond lengths: Å). |
![]() | ||
Fig. 4 Potential energy surface at CCSD(T)/6-311++G(3df,3pd) level for the reaction of trans-PhCHOO and trans-HONO. |
![]() | ||
Fig. 5 Potential energy surface at CCSD(T)/6-311++G(3df,3pd) level for the reaction of trans-PhCHOO and cis-HONO. |
The rate constant corrected by the zero curvature tunneling effect for each reaction pathway calculated using Polyrate 1735 in the temperature range of 238–338 K is listed in Table 1. The rate constant (kot) of the oxygen atom transfer path for trans-PhCHOO and trans-HONO changes from 4.67 × 10−23 to 1.82 × 10−20 cm3 per molecule per s. While the kot value corresponding to trans-PhCHOO and cis-HONO varies from 2.48 × 10−28 to 6.62 × 10−24 cm3 per molecule per s in the range 238–338 K. Especially at 298 K, kot for the trans-PhCHOO + trans-HONO reaction (2.57 × 10−21 cm3 per molecule per s) is higher than that for the trans-PhCHOO + cis-HONO reaction (2.40 × 10−25 cm3 per molecule per s), about 4 orders of magnitude higher. The Arrhenius diagram for each reaction channel at 238–338 K is described in Fig. 6 and 7, respectively. Based on Arrhenius diagrams, the activation energy (Ea) of the oxygen atom transfer path for trans-PhCHOO and trans-HONO is calculated to be 4.90 kcal mol−1, which is lower than the Ea for trans-PhCHOO + cis-HONO reaction (8.28 kcal mol−1). This indicates that the oxygen atom transfer path for trans-PhCHOO and trans-HONO is the main channel for producing benzaldehyde and HNO3.
T (K) | Trans-PhCHOO + trans-HONO | Trans-PhCHOO + cis-HONO | |||||
---|---|---|---|---|---|---|---|
kot | koz | kca | kha | kot | koz | kca | |
238 | 4.67 × 10−23 | 5.58 × 10−35 | 4.69 × 10−25 | 6.85 × 10−12 | 2.48 × 10−28 | 3.82 × 10−32 | 5.62 × 10−29 |
248 | 1.03 × 10−22 | 3.41 × 10−34 | 7.08 × 10−25 | 5.19 × 10−12 | 9.65 × 10−28 | 1.95 × 10−31 | 1.24 × 10−28 |
258 | 2.14 × 10−22 | 1.82 × 10−33 | 1.04 × 10−24 | 3.02 × 10−12 | 3.41 × 10−27 | 8.76 × 10−31 | 2.57 × 10−28 |
268 | 4.23 × 10−22 | 8.57 × 10−33 | 1.48 × 10−24 | 1.66 × 10−12 | 1.10 × 10−26 | 3.53 × 10−30 | 5.08 × 10−28 |
278 | 8.02 × 10−22 | 3.62 × 10−32 | 2.07 × 10−24 | 9.47 × 10−13 | 3.29 × 10−26 | 1.29 × 10−29 | 9.54 × 10−28 |
288 | 1.46 × 10−21 | 1.39 × 10−31 | 2.82 × 10−24 | 7.15 × 10−13 | 9.18 × 10−26 | 4.33 × 10−29 | 1.72 × 10−27 |
298 | 2.57 × 10−21 | 4.88 × 10−31 | 3.78 × 10−24 | 5.68 × 10−13 | 2.40 × 10−25 | 1.34 × 10−28 | 2.99 × 10−27 |
308 | 4.37 × 10−21 | 1.58 × 10−30 | 4.98 × 10−24 | 3.52 × 10−13 | 5.93 × 10−25 | 3.86 × 10−28 | 5.03 × 10−27 |
318 | 7.21 × 10−21 | 4.77 × 10−30 | 6.45 × 10−24 | 2.26 × 10−13 | 1.39 × 10−24 | 1.04 × 10−27 | 8.19 × 10−27 |
328 | 1.16 × 10−20 | 1.35 × 10−29 | 8.24 × 10−24 | 1.02 × 10−13 | 3.10 × 10−24 | 2.65 × 10−27 | 1.29 × 10−26 |
338 | 1.82 × 10−20 | 3.59 × 10−29 | 1.03 × 10−23 | 8.93 × 10−14 | 6.62 × 10−24 | 6.39 × 10−27 | 2.01 × 10−26 |
![]() | ||
Fig. 6 Arrhenius diagram of each channel for the reaction of trans-PhCHOO and trans-HONO at 238–338 K. |
![]() | ||
Fig. 7 Arrhenius diagram of each channel for the reaction of trans-PhCHOO and cis-HONO at 238–338 K. |
The transition states of trans-TSoz and cis-TSoz for heteroozonides generated by terminal O and C atoms of trans-PhCHOO combine with O and N atoms of HONO in the first type of cycloaddition reaction, as displayed in Fig. 8. The IRC profiles displayed in Fig. S1(f) and (g) of the SI clearly demonstrate that trans-TSoz and cis-TSoz connect the corresponding reactants and products. The distance of the terminal O atom of PhCHOO and O atom of HONO in the two transition states is 1.917 Å and 1.894 Å, while the distance between the C atom in CHOO and N atom of HONO is 1.746 Å and 1.651 Å, showing the trend towards the formation of a five-membered ring adduct. The imaginary frequencies of trans-TSoz and cis-TSoz are 366i and 248i cm−1, respectively, which correspond to vibrations of terminal O and C atoms in trans-PhCHOO attacking O and N atoms of HONO. As displayed in Fig. 4 and 5, the energy barrier for heteroozonide generated by the reaction of trans-PhCHOO with trans-HONO and cis-HONO is 19.52 and 15.31 kcal mol−1, respectively. It is worth noting that these two reactions are endothermic, with reaction energies of 9.13 and 10.16 kcal mol−1, which do not easily occur under atmospheric conditions.
![]() | ||
Fig. 8 Optimized configurations at M06-2X/6-311++G(d,p) level for the first type of cycloaddition reaction of trans-PhCHOO and HONO (bond lengths: Å). |
In the second type of cycloaddition reaction, two steps lead to the formation of benzoic acid. In the first step, a five-membered ring adduct (RCca) is formed between trans-PhCHOO and HONO via transition state of trans-TS1ca or cis-TS1ca. Compared with trans-TSoz and cis-TSoz shown in Fig. 8, in the above two transition states, the terminal O and C atoms of trans-PhCHOO combine with N and O atoms of HONO to form RCca. The distance of the terminal O atom of PhCHOO and N atom of HONO in the two transition states is 1.945 Å and 1.915 Å, while the distance between the C atom of CHOO and O atom of HONO is 1.879 Å and 1.924 Å (see Fig. 9 and 10), showing the trend towards the formation of a five-membered ring adduct. The imaginary frequencies of trans-TS1ca and cis-TS1ca are 363i and 368i cm−1, respectively, which correspond to vibrations of terminal O and C atoms in trans-PhCHOO attacking N and O atoms of HONO. IRC curves of trans-TS1ca and cis-TS1ca show that they are associated with corresponding reactants and products (see Fig. S1(h) and (i) in SI). Relative to the reactants trans-PhCHOO and trans-HONO, the energy barrier for formation of RCca is 2.60 kcal mol−1, with reaction energy of −17.56 kcal mol−1, while the energy barrier for trans-PhCHOO and cis-HONO to produce RCca is 7.22 kcal mol−1, with reaction energy of −17.99 kcal mol−1 (see Fig. 4 and 5).
In the subsequent second reaction step, the five membered ring adduct RCca is converted to benzoic acid and HONO through transition state trans-TS2ca or cis-TS2ca. The imaginary frequencies of trans-TS2ca and cis-TS2ca are 1099i and 1206i cm−1, respectively, which correspond to stretching vibrations for N–O and O–H bonds. As illustrated in Fig. 9 and 10, in these two transition states, the O⋯N distance is 1.682 Å and 1.590 Å, C⋯H distance is 1.206 Å and 1.209 Å, and O⋯H distance is 1.543 Å and 1.524 Å, respectively, showing the trend of N–O and C–H breaking and forming benzoic acid and HONO. Trans-TS2ca and cis-TS2ca are confirmed to be connected with corresponding reactants and products by IRC curves as displayed in Fig. S1(j) and (k) of the SI. Compared to trans-PhCHOO and HONO, the energy barrier for generation of benzoic acid is 4.55 kcal mol−1 (for trans-TS2ca) and 7.96 kcal mol−1 (for cis-TS2ca), with reaction energy of −113.02 kcal mol−1 (see Fig. 4 and 5).
![]() | ||
Fig. 9 Optimized configurations at M06-2X/6-311++G(d,p) level for the second type of cycloaddition reaction of trans-PhCHOO and trans-HONO (bond lengths: Å). |
![]() | ||
Fig. 10 Optimized configurations at M06-2X/6-311++G(d,p) level for the second type of cycloaddition reaction of trans-PhCHOO and cis-HONO (bond lengths: Å). |
It should be noted that with RCca as the initial stationary point, the energy barrier for generation of benzoic acid is 22.11 kcal mol−1 (for trans-TS2ca) and 25.95 kcal mol−1 (for cis-TS2ca), with reaction energy of −95.46 kcal mol−1 (for trans-RCca) and −95.03 kcal mol−1 (for cis-RCca) (see Fig. 4 and 5). The second step actually has a higher energy barrier, which is the rate-controlled step for cycloaddition reaction leading to benzoic acid formation. Rate constants for generation pathways of heteroozonide (koz) and benzoic acid (kca) in the temperature range of 238–338 K using Polyrate 1735 are shown in Table 1. The koz of heteroozonide produced by reaction of trans-PhCHOO and trans-HONO ranges from 5.58 × 10−35 to 3.59 × 10−29 cm3 per molecule per s, while koz values of trans-PhCHOO and cis-HONO to generate heteroozonide change from 3.82 × 10−32 to 6.39 × 10−27 cm3 per molecule per s in the range 238–338 K. koz for reaction of trans-PhCHOO and cis-HONO is about 3 orders of magnitude higher than that corresponding to trans-HONO.
On the contrary, kca for benzoic acid produced by reaction of trans-PhCHOO and trans-HONO ranges from 4.69 × 10−25 to 1.03 × 10−23 cm3 per molecule per s, which is about 4 orders higher than the corresponding kca of cis-HONO (5.62 × 10−29 to 2.01 × 10−26 cm3 per molecule per s) in the range 238–338 K. It is worth noting that when RCca is taken as the initial stationary point of the reaction, the energy barrier for generation benzoic acid is higher than that for formation of heteroozonide, but its reaction energy is −95.46 kcal mol−1 (for trans-RCca) or −95.03 kcal mol−1 (for cis-RCca), which is an exothermic reaction (while the production of heteroozonide is an endothermic reaction). Thus, the rate constant of the reaction of trans-PhCHOO with HONO to produce benzoic acid is always higher than rate constant of the heteroozonide formation path. According to the Arrhenius diagrams in Fig. 6 and 7, Ea of trans-PhCHOO and HONO to produce heteroozonide is 10.92 kcal mol−1 (for trans-HONO) and 9.87 kcal mol−1 (for cis-HONO). While, compared to trans-PhCHOO and HONO, Ea for benzoic acid formation pathway is 4.63 kcal mol−1 (for trans-HONO) and 7.35 kcal mol−1 (for cis-HONO). It should be pointed out that for the reaction of trans-PhCHOO and HONO to produce heteroozonide with higher Ea, its koz is smaller than that of the benzoic acid formation path. This reaction pathway is almost impossible to occur under atmospheric conditions. The cycloaddition reaction between trans-PhCHOO and HONO mainly generates benzoic acid.
![]() | ||
Fig. 11 Optimized configurations at M06-2X/6-311++G(d,p) level for hydrogen atom transfer reaction of trans-PhCHOO and trans-HONO (bond lengths: Å). |
It should be pointed out that it may be more accurate to consider the stable complex of trans-RCha as the initial stationary point for oxygen atom transfer, cycloaddition, and HAT reaction profiles, since dissociation of the complex must occur for these reactions to proceed. However, regrettably, for oxygen transfer and cycloaddition reactions, taking trans-RCha as the initial stationary point of the reaction, Gaussian 09 software30 is utilized to search and optimize repeatedly, but it failed to afford the corresponding transition states. Furthermore, regarding the reactions of anti-PhCHOO and cis-HONO, multiple searches and optimizations are also conducted, but do not lead to the complex cis-RCha. Taking anti-PhCHOO and anti-HONO (or cis-HONO) as the initial stationary point of the reaction, transition states of trans-TSot, cis-TSot, trans-TSoz, cis-Tsoz, trans-TS1ca and cis-TS1ca are obtained. The IRC curves shown in Fig. S1(d)–(i) of the SI indicate that these transition states connect the corresponding reactants and products. Therefore, anti-PhCHOO and anti-HONO (or cis-HONO) are used as the initial stationary point to better compare and analyze the pathways of oxygen transfer, cycloaddition and HAT reactions.
It should be pointed out that with trans-RCha as the initial stationary point for HAT, the reaction energy barrier is 9.35 kcal mol−1 (see Fig. 4). Trans-HONO is close to trans-PhCHOO and forms pre-reaction complex trans-RCha, whose equilibrium constant is Keq. Trans-RCha then generates Ph-HPMN through trans-TSha, with rate constant of k2:
The rate constant of hydrogen transfer reaction (kha) is equal to the product of Keq and k2 (kha = Keq × k2).33,34 Keq and rate constants (k2 and kha) corrected by the zero curvature tunneling effect for the hydrogen transfer reaction pathway calculated using Polyrate 1735 in the range 238–338 K are provided in Table S5 of the SI. Values of kha are also listed in Table 1. In the temperature range of 238–338 K, kha changes from 8.93 × 10−14 to 6.85 × 10−12 cm3 per molecule per s. Especially at 298 K, kha for the HAT reaction is 5.68 × 10−13 cm3 per molecule per s, which is about 8 orders of magnitude higher than kot of the oxygen atom transfer reaction (2.57 × 10−21 cm3 per molecule per s), and is about 11 orders of magnitude higher than kca for the benzoic acid formation pathway (3.78 × 10−24 cm3 per molecule per s). In addition, the Arrhenius diagram as shown in Fig. 6 indicates that the HAT path dominates compared to other reaction paths in the range 238–338 K. As displayed in Fig. 6, the difference between rate constants of kha and those of other paths decreases as temperature increases. This is mainly because the activation energy of the HAT reaction is negative (Ea = −2.49 kcal mol−1, compared to trans-PhCHOO and trans-HONO), exhibiting anti-Arrhenius behavior, while the activation energy of other reaction paths is positive, presenting Arrhenius behavior. It is worth noting that multiple optimizations in trans-PhCHOO and cis-HONO reaction did not afford the corresponding complex and transition state of Ph-HPMN in this work. The HAT reaction is only feasible in the case of trans-HONO, but does not occur for cis-HONO (see Fig. 4 and 11). Both the computational result in this work and experimental value reported in the literature37 show that trans-HONO is more stable than cis-HONO, and HONO mainly exists in the form of trans-HONO. Compared with other reaction paths, the HAT reaction between trans-PhCHOO and trans-HONO has a higher rate constant, which dominates the reaction between HONO and PhCHOO.
It is well known that Cl, OH and ClO radicals are the main oxidants in the troposphere and can undergo oxidation reactions with HONO.6 Since the rate constant of the HAT reaction is about 8–18 orders of magnitude higher than that of other pathways, for evaluating the relative influence of PhCHOO on HONO, only the HAT reaction path is selected for comparison with HONO reactions with Cl, OH and ClO radicals. The rate constant of the reaction of PhCHOO with HONO at 298 K (5.68 × 10−13 cm3 per molecule per s) is about 4 orders of magnitude higher than that of the reaction of ClO radical with HONO (1.62 × 10−17 cm3 per molecule per s) obtained by Anglada et al.6 and about 2 orders of magnitude lower than that of the reaction of Cl radical with HONO (7.38 × 10−11 cm3 per molecule per s). It is worth noting that rate constant of the reaction between PhCHOO and HONO obtained in this work is slightly smaller than that between OH radical and HONO (4.83 × 10−12 cm3 per molecule per s).6 This indicates that the reaction between PhCHOO and HONO can occur under atmospheric conditions, which can compete with reaction between HONO and OH radicals.
However, it should be pointed out that in evening or early morning, the concentration of OH is low,42 and the reaction between PhCHOO and HONO may be an important chemical reaction process for the consumption of HONO at night, affecting the budget of atmospheric HONO. In addition, Ph-HPMN formed by the reaction of PhCHOO and HONO is a common nitrogen-containing organic component with strong light absorption ability in atmospheric aerosol particles. It can be clearly seen from Fig. 11 that Ph-HPMN contains a peroxy bond, which can be dissociated in the atmosphere to generate OH radicals and serve as the reservoir for OH radicals at night.43 The results of this work indicate that the HAT pathway forms Ph-HPMN with the minimum activation energy and the maximum reaction rate constant, which dominates the reaction between HONO and PhCHOO. These results provide a theoretical basis for clarifying the formation mechanism of NOCs produced by ozonolysis of styrene and other olefin compounds in the presence of HONO.
Supplementary information is available. Cartesian coordinates and frequencies for all optimized structures at M06-2X level are provided in Table S1 and S2. Electronic energy at M06-2X and CCSD(T) level, T1 diagnostic value for various species at CCSD(T) level are listed in Table S3. While, activation energy and reaction energy of each reaction at CCSD(T)level are offered in Table S4. Keq, rate constant (k2 and kha) corrected by tunneling effect for hydrogen atom transfer reaction path in 238–338 K are supplied in Table S5. Also, IRC profile for the reactions at M06-2X level are illustrated in Fig. S1. See DOI: https://doi.org/10.1039/d5ra03441h.
This journal is © The Royal Society of Chemistry 2025 |