A theoretical study of CH3ONO + H reaction
De-Quan Wang,Yan Li,Xuri Huáng,Hui-Ling Liu,Chia-Chung Sun
DOI: https://doi.org/10.1016/j.theochem.2009.02.026
2009-01-01
Abstract:A detailed theoretical study is performed at the UMP2/6-311++G(d,p) and CCSD(T)/aug-cc-pVTZ (single-point) levels in order to explore the mechanism of the reaction between cis -CH 3 ONO/ trans -CH 3 ONO and H. For cis -CH 3 ONO + H reaction, six products are obtained. P 1 (CH 3 OH + NO), and P 6 (CH 3 O + HNO) are the most feasible products. P 3 (CH 3 + trans -HONO), P 4 (CH 3 + cis -HONO), and P 5 ( cis -CH 2 ONO + H 2 ) are the second feasible products, followed by the least feasible product P 2 (CH 4 + ONO). For trans -CH 3 ONO + H reaction, five dissociation products are obtained. P 1 (CH 3 OH + NO), P 6 (CH 3 O + HNO), and P 7 ( trans -CH 2 ONO + H 2 ) are the most feasible products, the lesser followed competitive product is P 4 (CH 3 + cis -HONO), while P 2 (CH 4 + ONO) is even much less feasible. Our theoretical results are in consistent with the available experiments. Because the rate-determining transition states involved in the feasible pathways lie above the reactant R, the title reactions may be important at higher temperatures. The present paper may be helpful for modeling of methyl nitrite–hydrogen combustion chemistry. Keywords Theoretical calculations Reaction mechanism Potential energy surface (PES) CH 3 ONO H 1 Introduction Methyl nitrite (CH 3 ONO) is one of the very few prototypes of medium-sized polyatomic molecules. Its infrared spectrum has been investigated by several groups [1–8] . It is believed that methyl nitrite CH 3 ONO, has two conformers with comparable amounts. Moreover, the microwave analysis of methyl nitrite, CH 3 ONO has shown that both conformers have a heavy atom skeleton with the C–O bond either cis or trans to the N–O bond [9] . Over the last 20 years, photodissociation dynamics of the methyl nitrite has been extensively studied both experimentally [10–23] and theoretically [24–31] . The CH 3 ONO reactions with H, F, OH and Cl have also been studied [32–37] . Among these reactions, the reaction with H attracts our interest. However to our best knowledge, only one experimental study has been performed on the title reaction. The measured rate constant of the CH 3 ONO + H reaction can be expressed by k 1 = (4.3 ± 0.9) × 10 −13 exp[−(1900 ± 110) cal/mol RT] cm 3 mol −1 s −1 in the temperature range of 223–398 K. A number of products can be listed for the CH 3 ONO + H reaction which are summarized as follows: CH 3 ONO + H → CH 3 OH + NO −62.6 kcal/mol →CH 2 ONO + H 2 −11 kcal/mol →CH 3 O + HNO −8.1 kcal/mol →CH 2 O + H 2 + NO −42.2 kcal/mol →CH 4 + NO 2 −46.1 kcal/mol →CH 3 + HONO −21.2 kcal/mol In Moortgat et al.’s studies, the primary products were identified as CH 3 OH + NO (47 ± 5%), CH 2 ONO + H 2 and CH 3 O + HNO (53 ± 5%). However, Moortgat et al. [37] did not provide further unambiguous information on the distribution of CH 2 ONO + H 2 and CH 3 O + HNO. Furthermore, products CH 2 O + H 2 + NO, CH 4 + NO 2 , and CH 3 + HONO are not suggested. Yet, in our opinion, these three products may be formed through the CH 3 ONO + H reaction at least thermodynamically feasible. In view of their potential importance and rather limited information, a detailed theoretical study on the doublet potential energy surfaces (DPESs) of the CH 3 ONO + H reaction is carried out in the present paper. 2 Computational methods All the calculations are carried out using the GAUSSIAN03 [38] and MOLPRO [39] sets of programs. Geometries of the reactants, products, intermediates, and transition states (TSs) are optimized using the unrestricted Møller–Plesset secondorder perturbation (UMP2) [40] method in conjunction with 6-311++G(d,p) basis set. Vibrational frequencies are obtained at the same level to characterize the different stationary points as local minima or transition states and to evaluate zero-point energy (ZPE). The ZPE and vibrational frequencies were scaled by a factor of 0.95 for anharmonicity correction [41] . To confirm that the transition states connect designated intermediates, intrinsic reaction coordinate (IRC) [42] calculations are performed at the UMP2/6-311++G(d,p) level. To obtain more reliable energetic data, higher level single-point energy calculations are performed at the coupled-cluster CCSD(T) [43] method with Dunning’s correlation-consistent augmented aug-cc-pVTZ basis sets [44] by using the UMP2/6-311++G(d,p) optimized geometries. For doublet systems, the expectation values of < S 2 > after annihilation range from 0.7502 to 0.7698 (the exact value for a pure doublet is 0.7500). This suggests that the wave function is not severely contaminated by states of higher multiplicity [45,46] . The thermodynamic functions (Δ H and Δ G ) are estimated within the ideal gas, rigid-rotor, and harmonic oscillator approximations. A temperature of 298.15 K and a pressure of 1 atm were assumed. 3 Results and discussion For the present system, one complex, three intermediate isomers, seven products and 16 transition states are located. The structures of the reactants, complex, products and intermediates are shown in Fig. 1 , while the structures of transition states are shown in Fig. 2 . Table 1 lists the energies of reactants, complex, intermediates, transition states and products. By means of the interrelation among the reactant, isomers, transition states and products as well as their corresponding relative energies, the schematic profiles of the potential energy surfaces for cis -CH 3 ONO + H, and trans -CH 3 ONO + H reactions are depicted in Figs. 3 a and b, respectively. Noted that CH 3 ONO have two isomers, i.e., cis -CH 3 ONO and trans -CH 3 ONO, the energy of cis -CH 3 ONO + H is set at zero as a reference for other species. The symbol TSm-n is used to denote the transition state connecting isomers m and n. Unless specified otherwise, the relative energies at CCSD(T)/aug-cc-pVTZ//UMP2/6-311++G(d,p)+ZPE level are used throughout. 3.1 cis -CH 3 ONO + H reaction For the reaction of cis -CH 3 ONO + H, six products are obtained, which are P 1 (CH 3 OH + NO), P 2 (CH 4 + ONO), P 3 (CH 3 + trans -HONO) P 4 (CH 3 + cis -HONO), P 5 ( cis -CH2ONO + H 2 ) and P 6 (CH 3 O + HNO). In the following parts, we focus our attention on the formation pathways of these six products. 3.1.1 P 1 (CH 3 OH + NO) From Fig. 3 a, we find that three pathways are possible to form P 1 . They can be written as: Path1 RP 1 (1): R → TSR-1 → 1 → TS1-C1 → C1 → P 1 Path2 RP 1 (2): R → TSR-1 → 1 → TS1-2 → 2 → TS2-C1 → C1 → P 1 Path2 RP 1 (3): R → TSR-C1 → C1 → P 1 Hydrogen atom attacks the N-atom of cis -CH 3 ONO lead to isomer 1 CH 3 ONHO (−29.93) via TSR-1(7.97). The values in parentheses are CCSD(T)/aug-cc-pVTZ//UMP2/6-311++G(d,p)+ZPE relative energies in kcal/mol with reference to reactant R(0.0). Subsequently, 1 can undergo either concerted 2,3-H-shift accompanied by internal N–O bond rupture to give P 1 as Path P 1 (1), or continuously 1,2 H-shift and concerted 1,3-H-shift along with N–O bond rupture to generate 2 cis -CH 3 ONOH (−33.46) then to the weakly bound complex C1 CH 3 OH…NO (−59.49). Finally, C1 dissociate to P 1 as in Path P 1 (2). In addition, the weakly bound complex C1 can be obtained via TSR-C1(1.91) as in Path P 1 (3). 3.1.2 P 2 (CH 4 + ONO) Hydrogen atom attacks the C-atom of cis -CH 3 ONO accompanied by C–O bond rupture will give rise to P 2 viaTSR-P 2 (26.92). Such a simple process can be depicted as: Path P 2 : R → TSR-P2 → P 2 3.1.3 P 3 (CH 3 + trans -HONO) For product P 3 , Only one pathway is found, which can be written as: Path P 3 : R → TSR-3 → 3 → TS3-P 3 → P 3 Hydrogen atom can attack the terminal O-atom of cis -CH 3 ONO to form 3 trans -CH 3 ONOH (−30.01) via TSR-3(13.43), followed by C–O bond fission to give P 3 via TS3-P 3 (−4.20). 3.1.4 P 4 (CH 3 + cis -HONO) Only one pathway is associated with the formation of P 4 , which can be depicted as: Path P 4 : R → TSR-1 → 1 → TS1-2→→2 → TS2-P 4 → P 4 The formation of 2 cis -CH 3 ONOH (−33.46) is the same as that of Path P 1 (2). Subsequently, 2 takes C–O bond rupture to form P 4 via TS2-P 4 (−3.06). 3.1.5 P 5 cis -CH2ONO + H 2 Direct H-abstraction from cis -CH 3 ONO to hydrogen atom to give P 5 can be achieved through TSR-P 5 (11.08). The formation pathway of P 5 can be written as: Path P 5 : R → TSR-P 5 → P 5 3.1.6 P 6 (CH 3 O + HNO) The formation of 1 CH 3 ONHO (−29.93) has been discussed in Section 3.1.1 . 1 undergoes internal N–O bond cleavage to give P 6 via TS1-P 6 (−0.58). Such a simple process can be depicted as: Path1 RP 6 : R → TSR-1 → 1 → TS1-P 6 → P 6 3.1.7 Reaction mechanism As shown in preceding sections, we have obtained 8 reaction pathways for the cis -CH 3 ONO + H reaction and three of them are associated with the formation of P 1 . Among the three formation pathways of P 1 , Path P 1 (1) should be the most competitive one because the relative energy of the rate-determining transition state TSR-1(7.97) in Path1 P 1 (1) lies much lower than that of TS1-2(12.16) in Path RP 1 (2) and TSR-C1(22.04) in Path RP 1 (3). Now, let us compare the feasibility of Path RP 1 (1), Path P 2 , Path P 3 , Path P 4 , Path P 5 , and Path P 6 . By comparison, we find that the relative energies of rate-determining transition state increases as follows: 7.97 kcal/mol (TSR-1 in Path P 1 (1)) = 7.97 kcal/mol (TSR-1 in Path P 6 ) → 11.08 kcal/mol (TSR-P 5 in Path P 5 ) → 12.16 kcal/mol (TS1-2 in Path2 RP 1 (2)) → 13.43 kcal/mol (TS3-P 3 in Path P3) → 26.92 kcal/mol (TSR-P 2 in Path P 2 ). Thus, we expect that Path P 2 should be the least feasible channel. Path P 3 , Path P 4 , and Path P 5 should be the second feasible pathways because the relative energies of the rate-determining transition states are very close within 3 kcal/mol. Path P 1 (1) and Path P 6 should be the most competitive pathways. Reflected in the final product distributions, we predict that P 1 and P 6 may be the most favorable products with comparable yield, P 3 , P 4 , and P 5 should be the second feasible products, while P 2 is the least competitive product. 3.2 trans -CH 3 ONO + H reaction The attack of hydrogen atom on trans -CH 3 ONO radical may have four kinds of entrance pathways (i) H-abstraction to form P 2 via TSR′-P 2 (28.68), or P 7 via TSR′-P 7 (10.86), (ii) internal O-atom attack accompanied by N–O bond rupture to give the weakly bond complex C1 CH 3 OH…NO (−59.49) before the final product P 1 . The relative energies of TSR′-C1 is 20.45 kcal/mol. (iii) N-attack to form 1 CH 3 ONHO (−29.93) via TSR′-1(9.06) (iv) terminal O-attack to form 2 cis -CH 3 ONOH (−33.46) via TSR′-2(12.89). Noted that further changes of isomers 1 and 2 have been discussed in Section 3.2 . For simplicity, we decide not to discuss it again. The pathways involved in Fig. 3 b are listed as: Path P 1 (1) R′ → TSR′-1 → 1 → TS1-C1 → C1 → P 1 Path P 1 (2) R′ → TSR′-2 → 2 → TS2-C1 → C1 → P 1 Path P 1 (3) R′ → TSR′-1 → 1 → TS1-2 → 2 → TS2-C1 → C1 → P 1 Path P 1 (4) R′ → TSR′-C1 → C1 → P 1 Path P 2 R′ → TSR′-P 2 → P 2 Path P 4 (1):R′ → TSR′-2 → 2 → TS2-P 4 → P 4 Path P 4 (2):R′ → TSR′-1 → 1 → TS1-2 → 2 → TS2-P 4 → P 4 Path RP 6 : R′ → TS R′-1 → 1 → TS1-P 6 → P 6 Path P 7 : R′ → TSR′-P 7 → P 7 3.2.1 Reaction mechanism Similar to the discussion of the cis -CH 3 ONO+H reaction, in the first place, we also compare which is the most probable formation channel of P 1 . Considering that the rate-determining transition state TSR′-1(9.06) in Path P 1 (1) lies lower than that of TSR′-2(12.89) in Path P 1 (2), TS1-2 in Path P 1 (3), and TSR′-C1(20.45) in Path P 1 (4). Thus the optimal channel to generate P 1 is Path P 1 (1). There are two pathways to form P4, that is Path R′P 4 (1) and Path R′P 4 (2). We expect that R′P 4 (1) should be competitive than R′P 4 (2) because Path P 4 (1) is simpler than Path P 4 (2). Now let us compare the feasibility of Path P 1 (1), Path P 2 , Path P 4 (1), Path P 6 , and Path P 7 . The relative energies of rate-determining transition states increased in sequence as 9.06 kcal/mol (TSR′-1 in Path P 1 (1) and Path P 6 ), 10.86 kcal/mol (TSR′-P 7 in Path P 7 ), 12.89 kcal/mol (TSR′-2 in Path P 4 ), and 28.68 kcal/mol (TSR′-P 2 in Path P 2 ). Thus we expect that the least competitive pathway should be Path P 2 . Furthermore, the relative energy of TSR′-2 in Path P 4 is higher than those of TSR′-1 in Path P 1 (1) and Path P 6 , and TSR′-P 7 in Path P 7 . Thus, we expect that Path P 4 cannot compete with Path P 1 (1), Path P 6 , and Path P 7 . On the other hand, because the relative energies of the rate-determining transition state in Path P 1 (1), Path P 6 , and Path P 7 are very close, these three channels may have comparable contribution to title reaction. As a result, reflected in the final product distributions, we predict that P 1 , P 6 and P 7 may be the most favorable products with comparable yield, followed by P 4 being the second feasible product, while P 2 is the least competitive product. In summary, for CH 3 ONO reactions with H, P 1 (CH 3 OH + NO), and P 6 (CH 3 O + HNO) may be the major products followed by the minor products P 3 (CH 3 + trans -HONO) P 4 (CH 3 + cis -HONO), P 5 ( cis -CH2ONO + H 2 ), and P 7 ( trans -CH2ONO + H 2 ), the contribution of P 2 (CH 4 + ONO) to the final product may be negligible. 3.3 Comparison with experiments In 1977, Moortgat et al. [37] performed experimental studies on the reaction of CH 3 ONO + H in a fast-flow system using photoionization mass spectrometry and excess atomic hydrogen in the temperature range of 223–398 K. The experimental results show the products and distributions are as: CH 3 OH + NO (47 ± 5%), CH 2 ONO + H 2 and CH 3 O + HNO (53 ± 5%). Among these products, CH 3 OH + NO corresponds to P 1 in our result, CH 2 ONO + H 2 corresponds to two products, which are P 5 and P 7 , CH 3 O + HNO corresponds to P 6 . On the whole, this result is in agreement with our theoretical calculations. However, the experimental studies do not give the distribution of products CH 2 ONO + H 2 and CH 3 O + HNO, while based on our theoretical studies, CH 3 O + HNO should have a larger distribution than CH 2 ONO + H 2 . Moreover, based on our theoretical results, P 5 and P 7 , which were ignored by Moortgat et al. may also contribute to the final products. Furthermore, the rate-determining steps in the most feasible pathways for the cis -CH 3 ONO + H and trans -CH 3 ONO + H reactions are a barrier-consumed process, the title reaction may be important at higher temperatures. Therefore, further reinvestigation of the title reaction at higher temperatures is still desirable. 4 Conclusion A detailed theoretical study was performed on the reaction of CH 3 ONO + H, the main calculated results can be summarized as follows: for cis -CH 3 ONO + H system, the most favorable products should be P 1 (CH 3 OH + NO) and P 6 (CH 3 O + HNO), P 3 (CH 3 + trans -HONO) P 4 (CH 3 + cis -HONO), and P 5 ( cis -CH2ONO + H 2 ) should be the second feasible product followed by the least feasible product P 2 (CH 4 + ONO). For trans -CH 3 ONO + H system, P 1 (CH 3 OH + NO), P 6 (CH 3 O + HNO), and P 7 ( trans -CH 2 ONO + H 2 ) should be the most favorable products, followed by P 4 (CH 3 + cis -HONO) and P 2 (CH 4 + ONO) being the lesser and least feasible product. Some conclusions are in agreement with the experimental investigations, and we hope the results may provide useful information for understanding the atom-molecule reaction in atmosphere. Acknowledgement This work is supported by the National Natural Science Foundation of China (No. 20773048). References [1] P. Tarte J. Chem. Phys. 20 1952 1570 [2] H.W. Brown G.C. Pimentel J. Chem. Phys. 29 1958 883 [3] P. Klaboe D. Jones E.R. Lippincott Spectrochim. Acta 23A 1967 2957 [4] J.F. Ogilvie J. Chem. Soc. Chem. Commun. 1973 450 [5] A. Hartford Jr. Chem. Phys. Lett. 53 1978 503 [6] F.L. Rook M.E. Jacox J. Mol. Spectrosc. 93 1982 101 [7] M.E. Jacox F.L. Rook J. Phys. Chem. 86 1982 2899 [8] R.S. Ruoff T.J. Kulp J.D. McDonald J. Chem. Phys 81 1984 4414 [9] P.H. Turner M.J. Corkill A.P. Cox J. Phys. Chem. 83 1979 1473 [10] F. Lahmani C. Lardeux D. Solgadi Chem. Phys. Lett. 102 1983 523 [11] O. Benoist D’azy F. Lahmani C. Lardeux D. Solgadi Chem. Phys. 94 1985 247 [12] F. Lahmani C. Lardeux D. Solgadi Chem. Phys. Lett. 129 1986 24 [13] M. Hippler F.A.H. Al-Janabi J. Pfab Chem. Phys. Lett. 192 1992 173 [14] M. Hippler M.R.S. McCoustra J. Pfab Chem. Phys. Lett. 198 1992 68 [15] M.R.S. McCoustra M. Hippler J. Pfab Chem. Phys. Lett. 200 1992 451 [16] U. Brühlmann M. Dubs J.R. Huber J. Chem. Phys. 86 1987 1249 [17] U. Brühlmann J.R. Huber Chem. Phys. Lett. 143 1988 199 [18] E. Kades M. Rösslein J.R. Huber Chem. Phys. Lett. 209 1993 275 [19] B.A. Keller P. Felder J.R. Huber Chem. Phys. Lett. 124 1986 135 [20] P. Farmanara V. Stert W. Radloff Chem. Phys. Lett. 303 1999 521 [21] J.W. Winniczek R.L. Dubs J.R. Appling V. McKoy M.G. White J. Chem. Phys. 90 1989 949 [22] H.-M. Yin J.-L. Sun Y.-M. Li K.-L. Han G.-Z. He J. Chem. Phys. 118 2003 8248 [23] G. Inoue M. Kawasaki H. Sato T. Kikuchi S. Kobayashi T. Arikawa J. Chem. Phys. 87 1987 5722 [24] M. Nonella J.R. Huber A. Untch R. Schinke J. Chem. Phys. 91 1989 194 [25] A. Untch K. Weide R. Schinke Chem. Phys. Lett. 180 1991 265 [26] A. Untch R. Schinke R. Cotting J.R. Huber J. Chem. Phys. 99 1993 9553 [27] M. Nonella J.R. Huber Chem. Phys. Lett. 131 1986 376 [28] S. Hennig V. Engel R. Schinke M. Nonella J.R. Huber J. Chem. Phys. 87 1987 3522 [29] V. Engel R. Schinke S. Hennig H. Metiu J. Chem. Phys. 92 1990 1 [30] V. Engel H. Metiu J. Chem. Phys. 92 1990 2317 [31] X.F. Yue J.L. Sun Z.F. Liu Q. Wei K.L. Han Chem. Phys. Lett. 426 2006 57 [32] O. Sokolov M.D. Hurley J.C. Ball T.J. Wallington W. Nelsen I. Barnes K.H. Becker Int. J. Chem. Kinet. 31 1999 357 [33] O.J. Nielsen H.W. Sidebottom M. Donlon J. Treacy J. Chem. Kinet. 23 1991 1095 [34] P. Biggs C.E. Canosa-Mas J.-M. Fracheboud D.E. Shallcross R.P. Wayne J. Chem. Soc. Faraday Trans. 93 1997 2481 [35] A.S. Manocha D.W. Setzer M.A. Wickramaaratchi Chem. Phys. 76 1983 129 [36] S. Dobe T. Berces F. Temps H.Gg. Wagner H. Ziemer Symp. Int. Combust. Proc. 25 1994 75 [37] G.K. Moortgat F. Glemr P. Warneck Int. J. Chem. Kinet. 4 1977 249 [38] M.J. Frisch et al., GAUSSIAN 03, Revision C02, Gaussian, Inc., Wallingford, CT, 2004, p. 32. [39] MOLPRO is a package of ab initio programs written by H.J. Werner, P.J. Knowles, with contributions from J. Almolf, R.D. Amos, M.J. O Deegan, S.T. Elbert, C. Hampel, W. Meyer, K. Peterson, R. Pitzer, A.J. Stone, R. Lindh, 2006. [40] H.B. Schlegel J. Chem. Phys. 84 1986 4530 [41] M.T. Nguyen S. Creve L.G. Vanquickenborne J. Phys. Chem. 100 1996 18422 [42] C. Gonzalez H.B. Schlegel J. Phys. Chem. 94 1990 5523 [43] G.D. Purvis R.J. Bartlett J. Chem. Phys. 76 1982 1910 [44] T.H. Dunning Jr. J. Chem. Phys. 90 1989 1007 [45] L. Famell J.A. Pople L. Radom J. Phys. Chem. 87 1983 79 [46] J.J.W. McDouall H.B. Schlegel J. Chem. Phys. 90 1989 2363