Mechanism Studies on the Degradation of Hexachlorocyclopentadiene in the Atmosphere

Jiao-Xue YANG Guo-Chun LV Yan WANG Ze-Hua WANG Xiao-Min SUN Zhi-Qiang LI

Citation:  Jiao-Xue YANG, Guo-Chun LV, Yan WANG, Ze-Hua WANG, Xiao-Min SUN, Zhi-Qiang LI. Mechanism Studies on the Degradation of Hexachlorocyclopentadiene in the Atmosphere[J]. Chinese Journal of Structural Chemistry, 2020, 39(11): 1925-1932. doi: 10.14102/j.cnki.0254–5861.2011–2833 shu

Mechanism Studies on the Degradation of Hexachlorocyclopentadiene in the Atmosphere

    通讯作者: 孙孝敏,
    李志强,

English

  • Since it was firstly identified as a chlorocarbon in 1930, hexachlorocyclopentadiene (HCCP) has received much attention due to its chlorinated property, high reactivity and volatility[1]. And HCCP is also widely used in the building block so as to obtain the biocidal and fire-retarding features, and in the synthesis of pesticides such as Chlordane, Heptachlor, Aldrin and Dieldrin through its Diels-Alder (D-A) condensation with the substrate[2-5].

    Some studies have shown that HCCP is toxic for organism[6-8]. In the range-finding study with rats the threshold of toxicity for HCCP was 0.11~0.5 ppm[4]. Considering its use in the processes of pesticide synthesis, HCCP was qualitatively identified as a contaminant in the discharge of pesticide production plants[9]. Although the annual United States production of HCCP is not a matter of public record, the yield of HCCP may be annually more than about 50 million pounds by estimating annual production of the cyclodiene insecticides[5, 10]. In addition to a large amount of environmental content, the conversion products of HCCP will also bring potential harm to the environment. Some studies have shown that a chlorine atom can be easily removed from HCCP when the temperature increased. Under oxidative conditions, these chlorinated aromatics can become precursors of PCDD/F, which are some of the most important pollutants due to their carcinogenic, teratogenic, and mutagenic effects[11-13].

    There have been numerous studies on the chemical reactivity of HCCP. In previous study, an experimental and theoretical investigation of the kinetics of HCCP into the pentachlorocyclopentadienyl radical (PCCP•) was researched[2]. The results show that the rate constant for Cl fission from HCCP was found to be k = 1.45 × 1015exp(–222 ± 9 kJ⋅mol-1⋅RT-1)⋅s-1[10]. Moreover, in aqueous phase the HCCP was photo-oxidized very efficiently[14, 15]. The majority of photoproducts in different sensitizer substrates are respectively hexachlorocyclopentenone (C5Cl6O), tetrachlorobutadiene (C4H2C14) or tetrachlorobutyne (C4Cl4)[16]. The diene reaction of HCCP with maleic acid was also researched, and their products can be utilized as a major constituent in the synthesis of polyester resins that are unusually flame resistant[3]. However, little is known about the degradation mechanism of HCCP and the products in the atmosphere. Thus, in this study we focus on the reactions between HCCP with different oxidants (•OH, •NO3, •HO2 radicals and O3) to gain insight into the understanding of molecular mechanisms of HCCP, which is significant for ecological safety and human health.

    To access the atmospheric behavior of HCCP, it is critical to understand their atmospheric reaction. Some previous studies have shown that quantum chemical calculation has been successfully applied to the study on the chemistry degradation mechanism of pesticides[17-19], such as DDT and folpet. This study is targeted at investigating the reaction mechanism and rate constants for the reaction of HCCP with the typical atmospheric oxidants by using quantum chemical method. And the kinetic model is constructed based on the rate constants of element reactions within the temperature range of 200~320 K.

    All the quantum chemical calculations were carried out using the Gaussian 09 program[20]. In this study, the geometries of the reactants, transition states, intermediates and products, as well as its vibrational frequencies were optimized using the M06-2X method[21] with a standard 6-311+G(d, p) basis set[22]. M06-2X has been proven to be a reliable calculation for transformation mechanisms and kinetics of organic pollutants[22, 23]. Meanwhile, in order to verify the transition states connected with the corresponding reactants and products, the intrinsic reaction coordinate (IRC)[24-26] in both directions from a saddle point was computed. The single point energy of all species was optimized at the M062X/6-311++G(3df, 3pd) level in order to obtain more accurate energy.

    The transition state theory (TST)[27, 28] is a method to calculate the dynamic parameters of the elementary reaction by using molecular structure and thermodynamic properties. It has been widely applied to calculate the rate processes[29-31]. In this paper, the KiSThelP program[32] is used to calculate the rate constants based on TST with Wigner tunneling correction.

    The initial reaction of HCCP with the oxidants can be described as follows:

    ${\text{HCCP + X}}\underset{{{k_{ - 1}}}}{\overset{{{k_1}}}{\rightleftarrows}}{\text{HCCP}} \cdot \cdot \cdot {\text{X}}$

    $ {\text{HCCP}} \cdot \cdot \cdot {\text{X}}\xrightarrow{{k_2}}{\text{post reactive complex}} $

    In which X is the typical atmospheric oxidants, such as •NO3, •HO2, •OH and O3.

    The total rate constant is calculated using the following equation:

    $ k_{total} = K_{eq}{\text{ }}k_2 $

    Where Keq is the equilibrium constant, Keq = K1/K-1 and K2 represents the unimolecular reaction rate constants.

    The initial reactions of •HO2, •OH, and •NO3 with HCCP are similar that the oxidants are added to the carbon site of the double bonds of HCCP. Meanwhile, the unsaturated double bonds of HCCP can react with O3 and form an oxide by the cycloaddition reaction[33], which is different from its reactions with •NO3, •HO2 and •OH. The primary ozonide is unstable, which can further dissociate to produce the Criegee intermediates[34]. Although there are two double bonds of HCCP, only one of them is taken into consideration owning to the symmetric structure of HCCP. All possible addition pathways are depicted in Fig. S1.

    The potential energy surface (PES) of the initiated transformation pathways of HCCP is shown in Fig. 1. When •HO2 approaches to HCCP, complexes C1A and C2A can be formed with the energy release of 5.11 and 5.22 kcal⋅mol-1, respectively. For path 1A, C1A can continue to transform into the intermediate IM1A via a transition state TS1A with the energy barrier of 15.18 kcal⋅mol-1. The IM2A in path 2A can be produced from C1A, but needs to pass through the transition state TS2A and overcome the energy barrier of 15.34 kcal⋅mol-1. It can be found that the two pathways have similar features. It is reasonable because only reaction sites are different for the two pathways.

    Figure 1

    Figure 1.  Schematic diagram of potential energy for the addition reaction of HCCP with HO2, OH and NO3 at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    OH radical is a key oxidant in the initiation of tropospheric degradation of gas-phase organic compounds, typically during daylight hours[35]. For the •OH-initiated reaction of HCCP, after the •OH is added to the C1 or C2 atom, the post-reactive complexes (C1B and C2B) can be generated through releasing energy of 4.20 kcal⋅mol-1. The reaction proceeds via the transition state TS1B (or TS2B) with the energy barrier of 3.96 kcal⋅mol-1 (or 4.42 kcal⋅mol-1) to produce the final intermediate IM1B (or IM2B).

    The reactions, NO + O3 → NO2 + O2 and NO2 + O3 → NO3 + O2, lead to the formation of the NO3 radical[36]. NO3 radical concentrations remain low in the daytime but can rise to measurable levels at night[37]. The reactions of •NO3 with HCCP are similar with the reactions of •OH and •HO2. As shown in Fig. 1, it is clear that the O atom of NO3 can attack the double bonds of HCCP. In this process, a complex (C1C or C2C) compound was formed by releasing energy of 5.76 kcal⋅mol-1. Furthermore, the energy barrier of the pathways from C1C (or C2C) to IM1C (or IM2C) was determined as 1.22 and 0.84 kcal⋅mol-1, and the reaction energies for the two pathways are –33.24 and –25.04 kcal⋅mol-1, respectively. Considering all the pathways of the addition reaction, it is clear that the reaction energies are negative (–13.51~–45.0 kcal⋅mol-1). The results imply that these processes are exothermic and potentially contribute to the transformation of HCCP from standpoint thermodynamics.

    Ozone is a powerful oxidizing agent, which can oxidize many organic compounds by addition to C=C double bonds. In the troposphere, O3 plays an important role in the oxidation of olefinic compounds. In recent years, field measurements have indicated very high concentrations of O3 in or near megacities of China[38]. The pathways for the O3-initiated reactions of HCCP are depicted in Fig. 2. The formation processes of IM1D include two steps, which correspond to the HCCP + O3 → the complex C1D and C1D → IMID. The transformation of HCCP into C1D is exothermic with the reaction energy of 4.46 kcal⋅mol-1. For C1D → IMID, this process is highly exothermic with reaction energy of 64.44 kcal⋅mol-1 and needs to overcome the barrier of 10.58 kcal⋅mol-1. Then a five-membered heterocyclic compound is formed. And the primary ozonide product is unstable. It could open the ring to further the subsequent reactions. As shown in Fig. 2, there are two possible pathways in the ring opening reactions. And the energy barriers in the processes are 24.01 and 25.66 kcal⋅mol-1, respectively.

    Figure 2

    Figure 2.  Schematic diagram of potential energy for the ozonization reaction of HCCP with O3 at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    Comparing these initiated reactions of HCCP, the energy barrier of NO3 pathways was lower than those of •OH, •HO2 and O3 pathways. These results mean that the NO3 pathways could be more kinetically favorable than the other pathways.

    In order to further assess the dominant subsequent reactions of HCCP in atmosphere, it is necessary to quantitatively the contribution of all initial reactions. The rate constants are calculated with the TST method at the temperature ranging from 200 to 320 K. Because the HCCP structure is symmetric, the rate constants initiated by •NO3, •OH, and •HO2 were calculated according to the following formula $ k_{total}^{\text{X}} $= 2(Radd1 + Radd2). As for the rate constants initiated by O3, it could be calculated with the formula $ k_{total}^{\text{X}} $ = 2R. The rate constants ($ k_{total}^{\text{X}} $) of •NO3, •OH, •HO2, and O3 initiated reactions are 2.49 × 10-12, 2.46 × 10-13, 2.44 × 10-22, and 1.33 × 10-20 cm3⋅molecule-1⋅s-1 at 298 K, respectively, which are listed in Table 1. The initiated reaction rate constants of HCCP can be described as •NO3 > •OH > O3 > •HO2 at 298 K.

    Table 1

    Table 1.  Calculated Rate Constants (cm3⋅molecule-1⋅s-1) between 200 and 320 K in the Reaction of HCCP and Oxidants
    DownLoad: CSV
    T(K) 200 K 230 K 260 K 298 K 320 K
    NO3 Radd1 8.88×10-12 2.29×10-12 8.30×10-13 3.18×10-13 2.06×10-13
    Radd2 3.35×10-11 7.73×10-12 2.60×10-12 9.30×10-13 5.86×10-13
    $ k_{{\text{total}}}^{{\text{N}}{{\text{O}}_3}} $ 8.48×10-11 2.00×10-11 6.86×10-12 2.49×10-12 1.58×10-12
    OH Radd1 1.02×10-13 8.98×10-14 8.33×10-14 7.99×10-14 7.94×10-14
    Radd2 4.02×10-14 3.99×10-14 4.08×10-14 4.30×10-14 4.46×10-14
    $ k_{{\text{total}}}^{{\text{OH}}} $ 2.84×10-13 2.59×10-13 2.48×10-13 2.46×10-13 2.48×10-13
    HO2 Radd1 1.34×10-26 3.41×10-25 4.25×10-24 5.23×10-23 9.43×10-25
    Radd2 1.74×10-26 4.47×10-25 5.61×10-24 6.94×10-23 2.32×10-22
    $k_{{\text{total}}}^{{\text{H}}{{\text{O}}_{\text{2}}}}$ 6.16×10-26 1.58×10-24 1.97×10-23 2.44×10-22 4.65×10-22
    O3 R 3.75×10-23 2.80×10-22 1.36×10-21 6.67×10-21 1.44×10-20
    $ k_{{\text{total}}}^{{{\text{O}}_{\text{3}}}} $ 7.49×10-23 5.59×10-22 2.72×10-21 1.33×10-20 2.87×10-20

    Considering the concentrations of •NO3, •OH, O3 and •HO2 in the atmosphere, the degradation rate can be expressed as the formula r = $ k_{total}^{\text{X}} $ [X] [HCCP], where [X] is the concentration of •NO3, •OH, •HO2 and O3. Although the rate constants of $ k_{total}^{{\rm{N}}{{\rm{O}}_{\rm{3}}}}$ and $ k_{total}^{{\text{OH}}} $ are much higher than $ k_{total}^{{{\rm{O}}_{\rm{3}}}}$, the role of O3 cannot be ignored due to its high concentration (the value of [O3] is about 7.0 × 1011 molecule⋅cm-3[39], and [•OH] is about 9.7 × 105 molecule⋅cm-3[40], [•HO2] = 1.0 × 108 molecule⋅cm-3[41], [•NO3] = 1.2 × 107 molecule⋅cm-3[42]), and O3 can participate in atmospheric reactions all day. O3 is a highly reactive gas and a strong oxidant, which can undergo rapid heterogeneous chemical reactions with organic matters. Thus, the O3 is indispensable to assess the subsequent reactions of HCCP.

    3.3.1   Subsequent reactions of •NO3/•OH

    Because O2 and NO are abundant in the atmosphere and O2 is the common oxidant, the intermediates can undergo further reactions in the presence of O2 and NO in the atmosphere after the addition reactions initiated by •NO3, •HO2, •OH with HCCP. Owing to the rate constants of •NO3 and •OH with HCCP to be the fastest, we choose the •NO3/•OH-initiated reactions as examples to calculate the subsequent degradation reactions.

    The intermediates formed in the •NO3-initiated reaction could be further oxidized by ubiquitous O2 in the atmosphere. The subsequent reactions of NO3 radical are shown in Fig. 3. Obviously, the reaction of IM1C with O2 is strongly exothermic (43.95 kcal⋅mol-1). In the troposphere, the reaction of RO2 + NO → RONO2 is a radical chain terminating step that can suppress ozone production. Thus, the IM3C peroxy radical can be regarded as a good depressor of photochemical smog. For IM4C → IM5C + NO2, the NO2 elimination reaction has a barrier of 28.08 kcal⋅mol-1, and this process is endothermic with the reaction energy of 9.47 kcal⋅mol-1. Furthermore, IM5C will open the C–C bond and eliminate a molecule of NO2 to generate the product P1 after overcoming the energy barrier of 4.59 kcal⋅mol-1. And this is a high exothermic reaction with 54.09 kcal⋅mol-1.

    OH radicals are the most important oxidant in the atmosphere. It controls the oxidative removal of most trace gases in the atmosphere, and plays an important role in the generation of photochemical smog. Although the rate constant of •OH is slower than the •NO3-initiated reaction, we also calculate the subsequent reaction of OH. As shown in Fig. 3, the subsequent reactions of •OH are similar with •NO3. The reaction of IM1B with O2 can release energy of 50.07 kcal⋅mol-1. Then, the IM3B peroxy radical can also react with ubiquitous NO in the troposphere. The pathway can release energy of 24.21 kcal⋅mol-1. Moreover, the NO2 elimination reaction occurs, which is exothermic with the reaction energy of 9.59 kcal⋅mol-1 and needs to overcome the energy barrier of 24.77 kcal⋅mol-1. This process is the rate determining step in the reaction pathway owning to the high energy barrier. In the end, P2 can be produced through the transition state TS4B and overcomes the energy barrier of 1.27 kcal⋅mol-1.

    Figure 3

    Figure 3.  Schematic diagram of potential energy for the subsequent pathways of IM1C and IM1B at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    The subsequent reactions of IM2C and IM2B are shown in Fig. S2 and Fig. S3. The subsequent pathway of IM2C also contains four elementary steps: O2 barrierless reaction, NO addition, NO2 elimination and C–C bond cleavage. What's more, the C–C cleavage reactions of IM8C are a little different from IM5C maybe owing to the influence of Cl atom. The reaction of IM8C is divided to two steps, in which the C–C bond is broken firstly, and then the NO2 is removed.

    3.3.2   Subsequent reactions of Criegee intermediate

    The reaction of O3 with unsaturated hydrocarbons produces short-lived molecules termed Criegee intermediates[43]. The fate of the Criegee intermediates determines the end products of the ozonolysis reaction and can have a substantial impact on the atmosphere[44]. It can react with radicals and common gases in atmosphere like O2, SO2 and NO2 because of the active feature[45, 46]. In this study, the subsequent degradation reactions of Criegee intermediate were calculated.

    The intermediate IM2D formed by ozonization reactions is taken as an example to access the subsequent reaction of HCCP. The energy profile for the subsequent reaction of IM2D with SO2, O2 and NO2 and the optimized geometries are shown in Fig. 4. For the O2 reactions, a five-membered heterocyclic compound IM4D is formed by a barrierless reaction of IM2D with O2. Then, the transformation of IM4D into P1 releases the energy of 8.28 kcal⋅mol-1 and needs to overcome the barrier of 27.33 kcal⋅mol-1.

    Figure 4

    Figure 4.  Schematic diagram of the subsequent pathways of Criegee intermediate IM2D initiated by NO2, O2 and SO2

    In the stratosphere, NO2 is involved in the conversion of active halogen oxides into less active species. In the troposphere, it is one of the most important O3 precursors[47]. In this work, two pathways for the reactions of NO2 with IM2D are calculated. For the first pathway, the formation processes of P1 include two steps, corresponding to the reactions IM2D + NO2 → IM5D and IM5D → IMID + NO3. The transformation of IM2D into IM5D is exothermic with the reaction energy of 10.44 kcal⋅mol-1 and needs to overcome the energy barrier of 5.22 kcal⋅mol-1 via a transition state TS4D. Then the O–O bond breaks to form P1 and NO3, with releasing the energy of 38.52 kcal⋅mol-1. The second reaction pathway of NO2 is the N atom added to the terminal C atom of IM2D. In this process, the energy barrier of transition state TS5D is 5.55 kcal⋅mol-1 and this is an exothermic process with reaction energy 17.67 kcal⋅mol-1. Then a cyclic intermediate IM7D is formed by the intramolecular reaction of O–O bond, which requires to overcome a high energy barrier of 42.74 kcal⋅mol-1 and absorb the reaction energy of 41.21 kcal⋅mol-1. Next, the pentacyclic compound IM7D can react with H2O in atmosphere, which is an endothermic reaction with the energy barrier of 29.43 kcal⋅mol-1.

    For the reaction of SO2 with IM2D, a complex compound C3D was formed by releasing energy of 9.89 kcal⋅mol-1. The reaction proceeds via the transition state TS8D with the energy barrier of 2.89 kcal⋅mol-1 to produce the cyclic intermediate IM9D. The pentacyclic intermediate is not stable, and can open the O–O bond simultaneously to form the product P1 and SO3. In this process, the energy barrier of the pathways from IM9D to P1 and SO3 is 23.31 kcal⋅mol-1 with the reaction energy being 51.49 kcal⋅mol-1.

    In this paper, the detailed description of degradation mechanism of HCCP initiated by •OH, •NO3, •HO2 and O3 have been researched with the computational methods. Considering the degradation reactions of HCCP with typical atmospheric oxidants (•NO3, •OH, •HO2, and O3), all the pathways are exothermic, and the maximum energy barrier is 15.34 kcal⋅mol-1. Moreover, the energy barriers of •NO3 and •OH initiated reaction are smaller than the reactions with •HO2 and O3.

    The kinetic analysis of HCCP is performed with the rate constants. The rate constants of •NO3, •HO2, •OH, and O3 initiated reactions are 2.49 × 10-12, 2.44 × 10-22, 2.46 × 10-13, 1.33 × 10-20 cm3⋅molecule-1⋅s-1 at 298 K, respectively. Obviously, the rate constants of •NO3, •OH are faster than those of •HO2 and O3. But the role of O3 cannot be ignored in atmosphere owning to the average concentration is high. Once the intermediates are produced in the initial reactions, they can react with O2 subsequently to generate the peroxy radical isomers, which can further undergo reaction via combination with tropospheric NO. The subsequent reactions can form the open-loop chlorinated products. This research provides an efficient way to investigate the gas-phase atmospheric chemistry transformation of HCCP.


    1. [1]

      McBee, E. T.; Hatton, R. E. Production of hexachlorobutadiene. Ind. Eng. Chem. 1949, 41, 809−812. doi: 10.1021/ie50472a031

    2. [2]

      Dharmarathne, N. K.; Mackie, J. C.; Kennedy, E. M.; Stockenhuber, M. Mechanism and rate of thermal decomposition of hexachlorocyclopentadiene and its importance in PCDD/F formation from the combustion of cyclodiene pesticides. J. Phys. Chem. A 2017, 121, 5871−5883. doi: 10.1021/acs.jpca.7b05209

    3. [3]

      Robitschek, P.; Bean, C. T. Flame-resistant polyesters from hexachlorocyclopentadiene. Ind. Eng. Chem. 1954, 46, 1628−1632. doi: 10.1021/ie50536a034

    4. [4]

      Zhang, C.; Wang, D.; Song, J.; Li, C.; Mo, Y. Bonding of Diels-Alder reactions of substituted 2-borabicyclo(1.1. 0)but-1(3)-enes: a theoretical study. Theor. Chem. Acc. 2019, 138, 106−5. doi: 10.1007/s00214-019-2491-5

    5. [5]

      Lu, P. Y.; Metcalf, R. L.; Hirwe, A. S.; Williams, J. W. Evaluation of environmental distribution and fate of hexachlorocyclopentadiene, chlordene, heptachlor, and heptachlor epoxide in a laboratory model ecosystem. J. Agric. Food Chem. 1975, 23, 967−973. doi: 10.1021/jf60201a016

    6. [6]

      Podowski, A. A.; Sclove, S. L.; Pilipowicz, A.; Khan, M. A. Q. Biotransformation and disposition of hexachlorocyclopentadiene in fish. Arch. Environ. Contam. Toxicol. 1991, 20, 488−496. doi: 10.1007/BF01065837

    7. [7]

      Rand, G. M.; Nees, P. O.; Calo, C. J.; Alexander, D. J.; Clark, G. C. Effects of inhalation exposure to hexachlorocyclopentadiene on rats and monkeys. J. Toxicol. Env. Health Part A 1982, 9, 743−760. doi: 10.1080/15287398209530201

    8. [8]

      Abdo, K. M.; Montgomery, C. A.; Kluwe, W. M.; Farnell, D. R.; Prejean, J. D. Toxicity of hexachlorocyclopentadiene: subchronic (13-week) administration by gavage to F344 rats and B6C3F1, mice. J. Appl. Toxicol. 1984, 4, 75−81. doi: 10.1002/jat.2550040204

    9. [9]

      Spehar, R. L.; Veith, G. D.; DeFoe, D. L.; Bergstedt, B. V. Toxicity and bioaccumulation of hexachlorocyclopentadiene, hexachloronorbornadiene and heptachloronorbornene in larval and early juvenile fathead minnows, Pimephales promelas. Bull. Environ. Contam. Toxicol. 1979, 21, 576−583. doi: 10.1007/BF01685472

    10. [10]

      Dharmarathne, N. K.; Mackie, J. C.; Lucas, J.; Kennedy, E. M.; Stockenhuber, M. Mechanisms of thermal decomposition of cyclodiene pesticides, identification and possible mitigation of their toxic products. Proc. Combust. Inst. 2019, 37, 1143−1150. doi: 10.1016/j.proci.2018.06.121

    11. [11]

      Sun, X.; Zhang, C.; Zhao, Y.; Bai, J.; Zhang, Q.; Wang, W. Atmospheric chemical reactions of 2, 3, 7, 8-tetrachlorinated dibenzofuran initiated by an OH radical: mechanism and kinetics study. Environ. Sci. Technol. 2012, 46, 8148−8155. doi: 10.1021/es301413v

    12. [12]

      Lin, L. F.; Lee, W. J.; Li, H. W.; Wang, M. S.; Chang-Chien, G. P. Characterization and inventory of PCDD/F emissions from coal-fired power plants and other sources in Taiwan. Chemosphere 2007, 68, 1642−1649. doi: 10.1016/j.chemosphere.2007.04.002

    13. [13]

      Yan, M.; Qi, Z.; Yang, J.; Li, X.; Ren, J.; Xu, Z. Effect of ammonium sulfate and urea on PCDD/F formation from active carbon and possible mechanism of inhibition. J. Environ, Sci. 2014, 26, 2277−2282. doi: 10.1016/j.jes.2014.09.012

    14. [14]

      Nubbe, M. E.; Adams, V. D.; Moore, W. M. The direct and sensitized photo-oxidation of hexachlorocyclopentadiene. Water Res. 1995, 29, 1287−1293. doi: 10.1016/0043-1354(94)00276-D

    15. [15]

      Wolfe, N. L.; Zepp, R. G.; Schlotzhauer, P.; Sink, M. Transformation pathways of hexachlorocyclopentadiene in the aquatic environment. Chemosphere 1982, 11, 91−101. doi: 10.1016/0045-6535(82)90160-6

    16. [16]

      Stemmler, E. A.; Hites, R. A. Methane enhanced negative ion mass spectra of hexachlorocyclopentadiene derivatives. Anal Chem. 1985, 57, 684−692. doi: 10.1021/ac00280a025

    17. [17]

      Shi, J.; Liu, H.; Sun, L.; Hou, H.; Xu, Y.; Wang, Z. Theoretical study on hydrophilicity and thermodynamic properties of polyfluorinated dibenzofurans. Chemosphere 2011, 84, 296−304. doi: 10.1016/j.chemosphere.2011.04.016

    18. [18]

      Zhou, Q.; Sun, X.; Gao, R.; Hu, J. Mechanism and kinetic properties for OH-initiated atmospheric degradation of the organophosphorus pesticide diazinon. Atmos. Environ. 2011, 45, 3141−3148. doi: 10.1016/j.atmosenv.2011.02.064

    19. [19]

      Gao, R.; Sun, X.; Yu, W.; Zhang, Q.; Wang, W. Mechanism and rate constants for complete series reactions of 19 fluorophenols with atomic H. J. Environ. Sci. 2014, 26, 154−159. doi: 10.1016/S1001-0742(13)60392-7

    20. [20]

      Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery Jr., J. A.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision B. 01, Gaussian, Inc., Wallingford CT 2009.

    21. [21]

      Wang, Y.; Verma, P.; Jin, X.; Truhlar, D. G.; He, X. Revised M06 density functional for main-group and transition-metal chemistry. Proc. Natl. Acad. Sci. U. S. A 2018, 115, 10257−10262. doi: 10.1073/pnas.1810421115

    22. [22]

      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 functionals. Theor. Chem. Acc. 2008, 120, 215−241. doi: 10.1007/s00214-007-0310-x

    23. [23]

      Li, C.; Xie, H. B.; Chen, J.; Yang, X.; Zhang, Y.; Qiao, X. Predicting gaseous reaction rates of short chain chlorinated paraffins with ·OH: overcoming the difficulty in experimental determination. Environ. Sci. Technol. 2014, 48, 13808−13816. doi: 10.1021/es504339r

    24. [24]

      Zheng, J.; Frisch, M. J. Efficient geometry minimization and transition structure optimization using interpolated potential energy surfaces and iteratively updated hessians. J. Chem. Theory Comput. 2017, 13, 6424−6432. doi: 10.1021/acs.jctc.7b00719

    25. [25]

      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. doi: 10.1021/ct0499783

    26. [26]

      Maeda, S.; Harabuchi, Y.; Ono, Y.; Taketsugu, T.; Morokuma, K. Intrinsic reaction coordinate: calculation, bifurcation, and automated search. Int. J. Quantum Chem. 2015, 115, 258−269. doi: 10.1002/qua.24757

    27. [27]

      Truhlar, D. G.; Garrett, B. C.; Klippenstein, S. J. Current status of transition-state theory. J. Phys. Chem. 1996, 100, 12771−12800. doi: 10.1021/jp953748q

    28. [28]

      Pechukas, P. Transition state theory. Annu. Rev. Phys. Chem. 1981, 32, 159−177. doi: 10.1146/annurev.pc.32.100181.001111

    29. [29]

      Berndt, T.; Böge, O. Atmospheric reaction of OH radicals with 1, 3-butadiene and 4-hydroxy-2-butenal. J. Phys. Chem. A 2007, 111, 12099−12105. doi: 10.1021/jp075349o

    30. [30]

      Cheng, G. J.; Zhang, X.; Chung, L. W.; Xu, L.; Wu, Y. D. Computational organic chemistry: bridging theory and experiment in establishing the mechanisms of chemical reactions. J. Am. Chem. Soc. 2015, 137, 1706−1725. doi: 10.1021/ja5112749

    31. [31]

      Truong, T. N. Reaction class transition state theory: hydrogen abstraction reactions by hydrogen atoms as test cases. J. Chem. Phys. 2000, 113, 4957−4964. doi: 10.1063/1.1287839

    32. [32]

      Canneaux, S.; 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. doi: 10.1002/jcc.23470

    33. [33]

      Lan, Y.; Wheeler, S. E.; Houk, K. N. Extraordinary difference in reactivity of ozone (OOO) and sulfur dioxide (OSO): a theoretical study. J. Chem. Theory Comput. 2011, 7, 2104−2111. doi: 10.1021/ct200293w

    34. [34]

      Wadt, W. R.; Goddard, W. A. Electronic structure of the Criegee intermediate. Ramifications for the mechanism of ozonolysis. J. Am. Chem. Soc. 1975, 97, 3004−3021. doi: 10.1021/ja00844a016

    35. [35]

      Gligorovski, S.; Strekowski, R.; Barbati, S.; Vione, D. Environmental implications of hydroxyl radicals (•OH). Chem. Rev. 2015, 115, 13051−13092. doi: 10.1021/cr500310b

    36. [36]

      Tsuji, M.; Kawahara, T.; Uto, K.; Hayashi, J.; Tsuji, T. Photochemical removal of NO2 in air at atmospheric pressure using side-on type 172-nm Xe2 excimer lamp. Int. J. Environ. Sci. Technol. 2019, 16, 5685−5694. doi: 10.1007/s13762-018-2131-y

    37. [37]

      Zelenov, V. V.; Aparina, E. V.; Kozlovskiy, V. I.; Sulimenkov, I. V.; Nosyrev, A. E. Kinetics of NO3 uptake on pyrene as a representative organic aerosols. Russ. J. Phys. Chem. B 2018, 12, 343−351. doi: 10.1134/S1990793118020136

    38. [38]

      Li, W.; Chen, Y.; Tong, S.; Guo, Y.; Zhang, Y.; Ge, M. Kinetic study of the gas-phase reaction of O3 with three unsaturated alcohols. J. Environ. Sci. 2018, 71, 292−299. doi: 10.1016/j.jes.2018.04.009

    39. [39]

      Logan, J. A. Tropospheric ozone: seasonal behavior, trends, and anthropogenic influence. Journal of Geophysical Research: Atmos 1985, 90, 10463−10482. doi: 10.1029/JD090iD06p10463

    40. [40]

      Lawrence, M. G.; Jöckel, P.; von Kuhlmann, R. What does the global mean OH concentration tell us? Atmos. Chem. Phys. 2001, 1, 37−49. doi: 10.5194/acp-1-37-2001

    41. [41]

      Holland, F.; Hofzumahaus, A.; Schäfer, J.; Kraus, A.; Pätz, H. W. Measurements of OH and HO2 radical concentrations and photolysis frequencies during BERLIOZ. J. Geophys. Res. Atmos. 2003, 108, 2−23.

    42. [42]

      Vrekoussis, M.; Kanakidou, M.; Mihalopoulos, N.; Crutzen, P. J.; Lelieveld, J.; Perner, D.; Berresheim, H.; Baboukas, E. Role of the NO3 radicals in oxidation processes in the eastern Mediterranean troposphere during the MINOS campaign. Atmos. Chem. Phys. 2004, 4, 169−182. doi: 10.5194/acp-4-169-2004

    43. [43]

      Horie, O.; Moortgat, G. K. Decomposition pathways of the excited Criegee intermediates in the ozonolysis of simple alkenes. Atmos. Environ. 1991, 25, 1881−1896. doi: 10.1016/0960-1686(91)90271-8

    44. [44]

      Taatjes, C. A.; Meloni, G.; Selby, T. M.; Trevitt, A. J.; Osborn, D. L.; Percival, C. J.; Shallcross, D. E. Direct observation of the gas-phase Criegee intermediate (CH2OO). J. Am. Chem. Soc. 2008, 130, 11883−11885. doi: 10.1021/ja804165q

    45. [45]

      Jiang, L.; Xu, Y. S.; Ding, A. Z. Reaction of stabilized Criegee intermediates from ozonolysis of limonene with sulfur dioxide: ab initio and DFT study. J. Phys. Chem. A 2010, 114, 12452−12461. doi: 10.1021/jp107783z

    46. [46]

      Long, B.; Bao, J. L.; Truhlar, D. G. Atmospheric chemistry of Criegee intermediates: unimolecular reactions and reactions with water. J. Am. Chem. Soc. 2016, 138, 14409−14422. doi: 10.1021/jacs.6b08655

    47. [47]

      Richter, A.; Burrows, J. P. Tropospheric NO2 from GOME measurements. Adv. Space Res. 2002, 29, 1673−1683. doi: 10.1016/S0273-1177(02)00100-X

  • Figure 1  Schematic diagram of potential energy for the addition reaction of HCCP with HO2, OH and NO3 at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    Figure 2  Schematic diagram of potential energy for the ozonization reaction of HCCP with O3 at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    Figure 3  Schematic diagram of potential energy for the subsequent pathways of IM1C and IM1B at the M06-2X/6-311+G(d, p)//M06-2X/6-311++G(3df, 3pd) level

    Figure 4  Schematic diagram of the subsequent pathways of Criegee intermediate IM2D initiated by NO2, O2 and SO2

    Table 1.  Calculated Rate Constants (cm3⋅molecule-1⋅s-1) between 200 and 320 K in the Reaction of HCCP and Oxidants

    T(K) 200 K 230 K 260 K 298 K 320 K
    NO3 Radd1 8.88×10-12 2.29×10-12 8.30×10-13 3.18×10-13 2.06×10-13
    Radd2 3.35×10-11 7.73×10-12 2.60×10-12 9.30×10-13 5.86×10-13
    $ k_{{\text{total}}}^{{\text{N}}{{\text{O}}_3}} $ 8.48×10-11 2.00×10-11 6.86×10-12 2.49×10-12 1.58×10-12
    OH Radd1 1.02×10-13 8.98×10-14 8.33×10-14 7.99×10-14 7.94×10-14
    Radd2 4.02×10-14 3.99×10-14 4.08×10-14 4.30×10-14 4.46×10-14
    $ k_{{\text{total}}}^{{\text{OH}}} $ 2.84×10-13 2.59×10-13 2.48×10-13 2.46×10-13 2.48×10-13
    HO2 Radd1 1.34×10-26 3.41×10-25 4.25×10-24 5.23×10-23 9.43×10-25
    Radd2 1.74×10-26 4.47×10-25 5.61×10-24 6.94×10-23 2.32×10-22
    $k_{{\text{total}}}^{{\text{H}}{{\text{O}}_{\text{2}}}}$ 6.16×10-26 1.58×10-24 1.97×10-23 2.44×10-22 4.65×10-22
    O3 R 3.75×10-23 2.80×10-22 1.36×10-21 6.67×10-21 1.44×10-20
    $ k_{{\text{total}}}^{{{\text{O}}_{\text{3}}}} $ 7.49×10-23 5.59×10-22 2.72×10-21 1.33×10-20 2.87×10-20
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  2
  • 文章访问数:  622
  • HTML全文浏览量:  17
文章相关
  • 发布日期:  2020-11-01
  • 收稿日期:  2020-04-02
  • 接受日期:  2020-05-27
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章