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
What problem does this paper attempt to address?
-
Synthesis, Dynamic Behavior, and Catalytic Activity of the Ether-Phosphine Complex Rh(P.intrsec.O)(P.apprx.O)Cl and Its Reactivity Toward Hydrogen, Oxygen, Methyl Iodide, and Acetyl Chloride
E LINDNER,QY WANG,HA MAYER,A BADER,H KUHBAUCH,P WEGNER
DOI: https://doi.org/10.1021/om00032a059
1993-01-01
Organometallics
Abstract:The complex ClRh(P approximately O)(P O) (1; P O = eta2(O,P)-chelated Cy2PCH2CH2OCH3 ligand; P approximately O = eta1(P) coordinated) exhibits fluxional behavior on the P-31 NMR time scale. Line-shape analysis of variable-temperature P-31{H-1) NMR spectra of 1 yields values of DELTAH(double-dagger) = 43.0 +/- 3.4 kJ mol-1 and DELTAS(double dagger) = -42 +/- 11.8 J mol-I K-1, indicating an associative mechanism of the dynamic process involving a five-coordinate species. Hydrogenation of 1-hexene to n-hexane over complex 1 as precatalyst proceeds with 100% selectivity and excellent activity (turnover number 7920 h-1) under mild conditions (300 K, 40 bar of H-2). Oxidative addition of H-2 and 02 results in the formation of the unstable dihydridorhodium(III) complex ClRhH2(P approximately O)2 (2) and the relatively stable (dioxygen)rhodium(III) complex ClRhO2(P approximately O)(P O) (3), respectively. The reaction of methyl iodide and acetyl chloride with complex 1 affords the products of the oxidative addition in trans fashion (CH3(Cl)(I)Rh(P approximately 0)(P O) (5a), CH3C(O)Cl2Rh(P approximately 0)(P O) (6a)) and cis fashion (CH3(Cl)(I)Rh(P approximately O)2 (5b), CH3(I)(CI)Rh(P approximately O)2 (5c), CH3C(O)Cl2Rh(P approximately O)2 (6b)). The use of labeled (CH3I)-C-13 and (CH3C(O)Cl)C-13 supports the structural characterization of all isomers by P-31{H-1} and C-13{H-1} NMR spectroscopy. When it is warmed in CDC13, 5a' is irreversibly converted into 5b' and 5c' (5a', 5b', and 5c' = (CH3)-C-13-labeled isomers of complex 5). First-order kinetics in 5a' with a rate constant of 7.67 X 10(-5) s-I at 287 K is established. The carbonylation of 5b (5b') gives a mixture of IRhCO(P approximately 0)2 (8) and CH3C(0)Rh(I) (Cl) (P approximately O)2 (9; 9' = (CH3C)-C-13-(O)-labeled analogue of 9) via the unstable intermediate 7. The progress of the reaction is monitored, and the compounds 8,9 (9'), and (CH3Cl)-C-13 are identified by P-31{H-1} and C-13{H-1} NMR spectroscopy.
-
Radical Reaction Hcno + (Nh)-N-3: A Mechanistic Study
Yan Li,Hui-Ling Liu,Yan-Bo Sun,Zhuo Li,Xu-Ri Huang,Chia-Chung Sun
DOI: https://doi.org/10.1007/s00214-009-0591-3
2009-01-01
Theoretical Chemistry Accounts
Abstract:The complex triplet potential energy surface for the reaction of HCNO with NH is investigated at the G3B3 level using the B3LYP/6-311++G(d,p), and QCISD/6-311++G(d,p) geometries. Various possible isomerization and dissociation pathways are probed. The initial association between HCNO and NH is found to be carbon to nitrogen attack leading to HNCHNO 2a, which can convert to 2b, 2c, and 2d. Subsequently, 1,4-H-shift of 2a to form NCHNOH 3a followed by dissociation to P (2) ((HCN)-H-1 + (HON)-H-3) is the most feasible pathway. Much less competitively, 2d undergoes successive 1,3-H-shift and C-N cleavage to form HNCNOH 8b, and then to product P (3) ((HNC)-H-1 + (HON)-H-3), the second feasible pathway. 8b can alternatively isomerize to 8c followed by N-O bond rupture to generate P (6) ((OH)-O-2 + (HNCN)-H-2), the lesser followed feasible pathway. In addition, 2b takes continuously 1,3- and 1,2-H-shift to form NC(H)NHO 6a, then to ONHCNH 7a which can convert to 7b. Eventually, 7b may take C-N bond fission to produce P (5) ((HNC)-H-1 + (HNO)-H-3), the least feasible pathway. The present paper may be helpful for future experimental identification of the product distributions for the title reaction, and may be helpful to deeply understand the mechanism of the title reaction.
-
Reaction Mechanism and Product Branching Ratios of OH+C2H3F Reaction: A Theoretical Study?
Chih-Hao Chin,Tong Zhu,John Zeng-Hui Zhang
DOI: https://doi.org/10.1063/1674-0068/cjcp2001016
IF: 1.09
2020-01-01
Chinese Journal of Chemical Physics
Abstract:Ab initio CCSD(T)/CBS//B3LYP/6-311G(d,p) calculations of the potential energy surface for possible dissociation channels of HOC2H3F, as well as Rice-Ramsperger-Kassel-Marcus (RRKM) calculations of rate constants, were carried out, in order to predict statistical product branching ratios in dissociation of HOC2H3F at various internal energies. The most favorable reaction pathway leading to the major CH2CHO+HF products is as the following: OH+C2H3F→i2→TS14→i6→TS9→i3→TS3→CH2CHO+HF, where the ratedetermining step is HF elimination from the CO bridging position via TS11, lying above the reactants by 3.8 kcal/mol. The CH2O+CH2F products can be formed by F atom migration from Cβ to Cα position via TS14, then H migration from O to Cα position via TS16, and C–C breaking to form the products via TS5, which is 1.8 kcal/mol lower in energy than the reactants, and 4.0 kcal/mol lower than TS11.
-
Radical reaction HCNO + 3 NH: a mechanistic study
Yan Li,Hui-ling Liu,Yan-bo Sun,Zhuo Li,Xu-ri Huang,Chia-chung Sun
DOI: https://doi.org/10.1007/s00214-009-0591-3
2009-01-01
Theoretical Chemistry Accounts
Abstract:The complex triplet potential energy surface for the reaction of HCNO with NH is investigated at the G3B3 level using the B3LYP/6-311++G(d,p), and QCISD/6-311++G(d,p) geometries. Various possible isomerization and dissociation pathways are probed. The initial association between HCNO and NH is found to be carbon to nitrogen attack leading to HNCHNO 2a , which can convert to 2b , 2c , and 2d . Subsequently, 1,4-H-shift of 2a to form NCHNOH 3a followed by dissociation to P 2 ( 1 HCN + 3 HON) is the most feasible pathway. Much less competitively, 2d undergoes successive 1,3-H-shift and C-N cleavage to form HNCNOH 8b , and then to product P 3 ( 1 HNC + 3 HON), the second feasible pathway. 8b can alternatively isomerize to 8c followed by N–O bond rupture to generate P 6 ( 2 OH + 2 HNCN), the lesser followed feasible pathway. In addition, 2b takes continuously 1,3- and 1,2-H-shift to form NC(H)NHO 6a , then to ONHCNH 7a which can convert to 7b . Eventually, 7b may take C-N bond fission to produce P 5 ( 1 HNC + 3 HNO), the least feasible pathway. The present paper may be helpful for future experimental identification of the product distributions for the title reaction, and may be helpful to deeply understand the mechanism of the title reaction.
-
Study on Conformation Interconversion of 3-Alkyl-4-acetyl-3;4-dihydro-2h-1;4-benzoxazines from Dynamic NMR Experiments and Ab Initio Density Functional Calculations
G Yang,XW Han,WP Zhang,XM Liu,PY Yang,YG Zhou,XH Bao
DOI: https://doi.org/10.1021/jp050874d
IF: 3.466
2005-01-01
The Journal of Physical Chemistry B
Abstract:Variable-temperature NMR experiments and ab initio density functional calculations were carried out to investigate the conformation interconversion of novel chiral 3-alkyl-3,4-dihydro-2H-benzo[1,4]oxazine derivatives. With CDCl3 as the solvent, the coalescence temperatures of H-2, H-3, H-11 and H-19 of product 1 are about 289, 304, 292, and 316 K, with the corresponding activation free energies at 58.0 +/- 6.7, 60.9 +/- 7.1, 58.3 +/- 6.8, and 59.6 +/- 6.9 kJ(.)mol(-1), respectively. When dimethyl sulfoxide (DMSO-d(6)) was used as the solvent, H-1 and C-13 NMR signals were completely assigned at 375 K. The effects of solvent and temperature were investigated through a polarizable continuum model. At each theoretical level (MP2 or B3LYP), the changing tendencies of the calculated activation free energies and interconversion rates agree well with those of the NMR results. In addition, the interconversion rate at each specified temperature was calculated to be about 1.5 times faster in DMSO-d6 than in CDCl3. Accordingly, we failed to observe the coalescence phenomena of H-3 and H-19 in DMSO-d(6) by NMR measurements from 296 to 375 K. The substitution effect at the R-1-R-5 positions was considered using density functional calculations, with the activation barriers decreasing as follows: product 6 > 3 > 1 > 7 > 2. This sequence is consistent with that of the reaction heats, except for product 7, implying that the interconversion processes may be thermodynamically controlled. Surprisingly, the substituted groups near the acetyl group in product 2 and 7 do not elevate the activation barrier but, instead, lower it somewhat, with the possible reasons for this provided in the paper.
-
Reaction mechanism of CH + C3H6: A theoretical study
Yan Li,Huiling Liu,Zhongjun Zhou,Xuri Huáng,Chiachung Sun
DOI: https://doi.org/10.1021/jp102029w
2010-01-01
Abstract:A detailed theoretical study is performed at the B3LYP/6-311G(d,p) and G3B3 (single-point) levels as an attempt to explore the reaction mechanism of CH with C3H6. It is shown that the barrierless association of CH with C3H6 forms two energy-rich isomers CH3-cCHCHCH(2) (1), and CH2CH2CHCH2 (4). Isomers 1 and 4 are predicted to undergo subsequent isomerization and dissociation steps leading to ten dissociation products P-1 (CH3-cCHCHCH + H), P-2 (CH3-eCCHCH(2) + H), P-3 (cCHCHCH(2) + CH3), P-4 (CH3CHCCH2 + H), P-5 (cis-CH2CHCHCH2 + H), P-6 (trans-CH2CHCHCH2 + H), P-7 (C2H4 C2H3), P-8 (CH3CCH + CH3), P-9 (CH3CCCH3 + H) and P-12 (CH2CCH2 + CH3), which are thermodynamically and kinetically possible. Among these products, P5, P6, and Pi may be the most favorable products with comparable yields; P-1, P-2, and P-3 may be the much less competitive products, followed by the almost negligible P-4, P-8, P-9, and P-12. Since the isomers and transition states involved in the CH + C3H6 reaction all lie lower than the reactant, the title reaction is expected to be fast, which is consistent with the measured large rate constant in experiment. The present study may lead us to a deep understanding of the CH radical chemistry.
-
Theoretical study on the mechanism of C 2 Cl 3 + NO 2 reaction
Yan Li,Hui-ling Liu,Xu-ri Huang,De-quan Wang,Chia-chung Sun,Au-chin Tang
DOI: https://doi.org/10.1007/s00214-009-0549-5
2009-01-01
Theoretical Chemistry Accounts
Abstract:The radical-molecule reaction of C 2 Cl 3 with NO 2 is explored at the B3LYP/6-311G(d,p) and CCSD(T)/6-311+G(d,p) (single-point) levels. On the singlet potential energy surface (PES), the association between C 2 Cl 3 and NO 2 is found to be carbon-to-nitrogen attack forming the adduct C 2 Cl 3 NO 2 ( 1 ) without any encounter barrier, followed by isomerization to C 2 Cl 3 ONO ( 2 ). Starting from 2 , the most feasible pathway is the N–O1 bond cleavage which lead to P 1 (C 2 Cl 3 O + NO). Much less competitively, 2 transforms to the three-membered ring isomer c-OCCl 2 C–ClNO ( 4 a ) which can easily interconvert to c-OCCl 2 C–ClNO 4 b . Then 4 ( 4 a , 4 b ) takes direct C1–C2 and C2–O1 bonds cleavage to give P 2 (COCl 2 + ClCNO). The lesser competitive channel is the 4 a isomerizes to the four-membered ring intermediate O-c-CNClOCCl 2 ( 5 ) followed by dissociation to P 3 (CO + ClNOCCl 2 ). The concerted 1,2-Cl shift along with C1–O1 bond rupture of 4 b to form ONC(O)CCl 3 ( 6 ) followed by dissociation to P 4 (ClNO + OCCCl 2 ) is even much less feasible. Moreover, some of P 3 and P 4 can further dissociate to P 5 (ClNO + CO + CCl 2 ). Compared with the singlet pathways, the triplet pathways may have less contribution to the title reaction. Our results are in marked difference from previous theoretical studies which showed that two initial adducts C 2 Cl 3 –NO 2 and C 2 Cl 3 –ONO are obtained. Moreover, in the present paper we focus our main attentions on the cyclic isomers in view of only the chain-like isomers are considered by previous studies. The present study may be helpful for understanding the halogenated vinyl chemistry.
-
Theoretical Study on Mechanisms of the High-Temperature Reactions C2H3 + H2O and C2H4 + OHElectronic Supplementary Information (ESI) Available: Figures of Optimized Geometries of the Fragments of C2H5O Radicals, C2H5O Isomers and C2H5O Transition States and Table of Harmonic Vibration Frequencies of C2H3, H2O, TSR/P1, C2H4 and OH. See Http://Www.rsc.org/suppdata/cp/b1/b109758j/
Guixia Liu,Ding Ye,Zesheng Li,Qiang Fu,Xuri Huang,Chenglin Sun,An‐Na Tang
DOI: https://doi.org/10.1039/b109758j
2002-01-01
Abstract:The potential energy surface of the radical-molecule reaction C2H3 + H2O in the gas phase is explored at the 6-31G(d,p) and 6-311G(d,p) B3LYP and single-point QCISD(T)/6-311G(2df,p) levels. The most favorable channel is the direct H-abstraction from H2O to C2H3 leading to product P1 C2H4 + OH, whereas the other channels leading to the products P2 CH3 + CH2O, P3 CH3CHO + H, P4cis-CH2CHOH + H and P4′ trans-CH2CHOH + H are kinetically much less competitive. For the direct H-abstraction channel, high-level energetic calculations at the QCISD(T)/6-311G(2df,p), QCISD(T)/6-311+G(2df,2p) and G2 levels using the B3LYP/6-31G(d,p) and QCISD/6-31G(d,p) optimized geometries are further performed to estimate the thermal rate constants over a wide temperature range 200–5000 K for comparison with future laboratory measurements. The calculated barrier heights at the QCISD(T)/6-311+G(2df,2p) and G2 levels based on the QCISD/6-31G(d,p) geometries with zero-point vibrational energy (ZPVE) correction are 12.6 and 13.0 kcal mol−1, respectively. The results indicate that the C2H3 + H2O reaction might play an important role at high temperatures (T > 1800 K) in the presence of gaseous water and should be incorporated in the C2H3-modeling of hydrocarbon-fuel combustion processes. Discussions are also made in comparison with the analogous reactions C2H3 + H2 and C2H + H2O. While the addition-elimination mechanism of another important radical-molecule reaction C2H4 + OH has been the subject of extensive theoretical and experimental studies, its H-abstraction process leading to C2H3 + H2O has received little attention. For the C2H4 + OH → C2H3 + H2O channel, our calculations predict ZPVE-corrected barriers, 5.6 and 5.4 kcal mol−1, respectively, at the QCISD(T)/6-311+G(2df,2p)//QCISD/6-31G(d,p) and G2//QCISD/6-31G(d,p) levels, and reveal its importance at high temperatures (T > 560 K). In the range 720–1173 K, the calculated high-level rate constants are quantitatively in good agreement with the measured values. However, our calculated activation energy, 9.5 and 9.3 kcal mol−1 at the QCISD(T)/6-311+G(2df,2p)//QCISD/6-31G(d,p) and G2//QCISD/6-31G(d,p) levels with ZPVE correction, respectively, suggests that the experimentally determined value 4–5 kcal mol−1 may be underestimated and future rate constant measurements over a wide temperature range including T > 1200 K may be desirable.
-
DFT study on the mechanism of the CF3O + NO reaction
Guohua Xu,Chengyin Shen,HaiYan Han,Jianquan Li,Höngmei Wang,Yannan Chu
DOI: https://doi.org/10.1016/j.theochem.2010.02.019
2010-01-01
Abstract:The singlet potential energy surface of the CF3O+NO reaction has been studied at the B3LYP/6-311+G(3df) level of theory. The relative energies were calculated by using the CCSD(T)/aug-cc-pVDZ and the G3B3 methods at the B3LYP/6-311+G(3df) optimized geometries. The study shows that the reaction starts via an exothermic barrierless addition of NO to the CF3O radical to produce cis-CF3ONO, which will isomerize to trans-CF3ONO, followed by trans-CF3ONO dissociating to the products CF2O+FNO. trans-CF3ONO can also rearrange to trans-CF3OON and further isomerize to cis-CF3OON. Once cis-CF3OON is formed, it will finally dissociate to CF2O+FNO. This is another energetically facile reaction route to produce CF2O+FNO. The transition states involved in above reaction pathways all lie below the reactants in energy, thus the present calculations suggest the overall rate coefficient of the title reaction may exhibit a negative temperature dependence, in agreement with most experimental results. Additionally, the enthalpies of formation of CF3NO2, trans-CF3ONO and cis-CF3ONO were computed to be ΔfH298.15∘(CF3NO2)=−160.99kcal/mol, ΔfH298.15∘(trans-CF3ONO)=−176.76kcal/mol and ΔfH298.15∘(cis-CF3ONO)=−173.24kcal/mol, respectively.
-
Theoretical Study of the C(3P) + Trans-C4h8 Reaction.
Yan Li,Hui-ling Liu,Xu-ri Huang,De-Quan Wang,Chia-chung Sun
DOI: https://doi.org/10.1021/jp810312h
2009-01-01
Abstract:The complex triplet potential energy surface for the reaction of ground-state carbon atom C((3)P) with trans-C(4)H(8) is theoretically investigated at the B3LYP/6-311G(d,p) and G3B3(single-point) levels. Various possible isomerization and dissociation pathways are probed. The initial association between C((3)P) and trans-C(4)H(8) is found to be the C((3)P) addition to the C=C bond of trans-C(4)H(8) to barrierlessly generate the three-membered cyclic isomer 1 CH(3)-cCHCCH-CH(3). Subsequently, 1 undergoes a ring-opening process to form the chainlike isomer 3a cis-trans-CH(3)CHCCHCH(3), which can either lead to P(6)((2)CH(3)CHCCCH(3) + (2)H) via the C-H bond cleavage or to P(7)((2)CH(3)CHCCH + (2)CH(3)) via C-C bond rupture. These two paths are the most favorable channels of the title reaction. Other channels leading to products P(1)((2)CH(3)-cCHCCH + (2)CH(3)), P(2)((2)CH(3)-cCHCC-CH(3) + (2)H), P(3)(trans-(2)CH(3)CHCH + (2)C(2)H(3)), P(4)(cis-(2)CH(3)CHCH + (2)C(2)H(3)), P(5)((3)CH(3)CH + (1)CH(3)CCH), P(8)(cis-(2)CH(3)CHCHCCH(2) + (2)H), P(9)(trans-(2)CH(3)CHCHCCH(2) + (2)H), P(10)((2)CH(3)CCCH(2) + (2)CH(3)), and P(11)((2)CH(3)CHCCHCH(2) + (2)H), however, are much less competitive due to either kinetic or thermodynamic factors. Because the intermediates and transition states involved in the C((3)P) + trans-C(4)H(8) reaction all lie below the reactant, the title reaction is expected to be rapid, as is consistent with the measured large rate constant. Our results may be helpful for future experimental investigation of the title reaction.
-
Formation of C3H2, C5H2, C7H2, and C9H2 from reactions of CH, C3H, C5H, and C7H radicals with C2H2
Yi-Lun Sun,Wen-Jian Huang,Shih-Huang Lee
DOI: https://doi.org/10.1039/c5cp06072a
2016-01-21
Abstract:The Cm+2H2 family can be classified into two categories - C2n+1H2 and C2n+2H2. Cm+2H2 are important intermediates in the syntheses of large carbonaceous molecules. An understanding of the formation mechanisms of both odd and even carbon-numbered Cm+2H2 is beneficial to atmospheric, astronomical, and combustion chemistry. HC2n+2H (polyynes) are believed to be producible from C2nH + C2H2 and C2H + C2nH2 reactions but C2n+1H2 (n≥ 2) attract less attention to their formation mechanisms. In the present study, we make up for the lack of knowledge on C2n+1H2 formation mechanisms by investigating the reactions C2n-1H + C2H2→ C2n+1H2 + H with n = 1-4. The dynamics of reactions of C2n-1H radicals with C2H2 are explored in crossed molecular beams using products C2n+1H2. The translational-energies and angular distributions of the hydrogen-loss channels of products are unraveled by measuring time-of-flight spectra and photoionization-efficiency spectra of C2n+1H2 with tunable synchrotron vacuum-ultraviolet ionization. The C2n+1H2 product includes two isomers, c-(1)HC2n-1(C)CH and (3)HC2n+1H, which are identified by the maximal translational-energy release and the photoionization threshold. Furthermore, quantum-chemical calculations indicate that the title reactions incur a small or negligible entrance barrier and are nearly isoergic except for the barrierless exothermic reaction CH + C2H2→ C3H2 + H. We demonstrate for the first time that C5H2, C7H2, and C9H2 are producible from the title reactions. In conjunction with studies on the C2nH + C2H2 reactions, a brief picture for the CmH (m = 1-8) + C2H2→ Cm+2H2 + H reactions can be outlined.
-
Theoretical Studies on the Initial Reaction Kinetics and Mechanisms of P-, M- and O-Nitrotoluene
Meng Yang,Caiyue Liao,Chenglong Tang,Peng Zhang,Zuohua Huang,Jianling Li
DOI: https://doi.org/10.1039/d0cp05935h
2021-01-01
Abstract:The potential energy surfaces (PESs) of three nitrotoluene isomers, such as p-nitrotoluene, m-nitrotoluene, and o-nitrotoluene, have been theoretically built at the CCSD(T)/CBS level. The geometries of reactants, transition states (TSs) and products are optimized at the B3LYP/6-311++G(d,p) level. Results show that reactions of -NO2 isomerizing to ONO, and C-NO2 bond dissociation play important roles among all of the initial channels for p-nitrotoluene and m-nitrotoluene, and that the H atom migration and C-NO2 bond dissociation are dominant reactions for o-nitrotoluene. In addition, there exist pathways for three isomer conversions, but with high energy barriers. Rate constant calculations and branching ratio analyses further demonstrate that the isomerization reactions of O transfer are prominent at low to intermediate temperatures, whereas the direct C-NO2 bond dissociation reactions prevail at high temperatures for p-nitrotoluene and m-nitrotoluene, and that H atom migration is a predominant reaction for o-nitrotoluene, while C-NO2 bond dissociation becomes important by increasing the temperature.
-
Riddles of the structure and vibrational dynamics of HO3 resolved near the ab initio limit
Marcus A Bartlett,Arianna H Kazez,Henry F Schaefer,Wesley D Allen
DOI: https://doi.org/10.1063/1.5110291
2019-09-07
Abstract:The hydridotrioxygen (HO3) radical has been investigated in many previous theoretical and experimental studies over several decades, originally because of its possible relevance to the tropospheric HOx cycle but more recently because of its fascinating chemical bonding, geometric structure, and vibrational dynamics. We have executed new, comprehensive research on this vexing molecule via focal point analyses (FPA) to approach the ab initio limit of optimized geometric structures, relative energies, complete quartic force fields, and the entire reaction path for cis-trans isomerization. High-order coupled cluster theory was applied through the CCSDT(Q) and even CCSDTQ(P) levels, and CBS extrapolations were performed using cc-pVXZ (X = 2-6) basis sets. The cis isomer proves to be higher than trans by 0.52 kcal mol-1, but this energetic ordering is achieved only after the CCSDT(Q) milestone is reached; the barrier for cis → trans isomerization is a minute 0.27 kcal mol-1. The FPA central re(O-O) bond length of trans-HO3 is astonishingly long (1.670 Å), consistent with the semiexperimental re distance we extracted from microwave rotational constants of 10 isotopologues using FPA vibration-rotation interaction constants (αi). The D0(HO-O2) dissociation energy converges to a mere 2.80 ± 0.25 kcal mol-1. Contrary to expectation for such a weakly bound system, vibrational perturbation theory performs remarkably well with the FPA anharmonic force fields, even for the torsional fundamental near 130 cm-1. Exact numerical procedures are applied to the potential energy function for the torsional reaction path to obtain energy levels, tunneling rates, and radiative lifetimes. The cis → trans isomerization occurs via tunneling with an inherent half-life of 1.4 × 10-11 s and 8.6 × 10-10 s for HO3 and DO3, respectively, thus resolving the mystery of why the cis species has not been observed in previous experiments executed in dissipative environments that allow collisional cooling of the trans-HO3 product. In contrast, the pure ground eigenstate of the cis species in a vacuum is predicted to have a spontaneous radiative lifetime of about 1 h and 5 days for HO3 and DO3, respectively.
-
The Atmospherically Important Reaction Of Hydroxyl Radicals With Methyl Nitrate: A Theoretical Study Involving The Calculation Of Reaction Mechanisms, Enthalpies, Activation Energies, And Rate Coefficients
Maggie Ng,Daniel K W Mok,Edmond P F Lee,John M Dyke
DOI: https://doi.org/10.1021/acs.jpca.7b05035
2017-01-01
Abstract:A theoretical study, involving the calculation of reaction enthalpies, activation energies, mechanisms, and rate coefficients, was made of the reaction of hydroxyl radicals with methyl nitrate, an important process for methyl nitrate removal in the earth's atmosphere. Four reaction channels were considered: formation of H2O + CH2ONO2, CH3OOH + NO2, CH3OH + NO3, and CH3O + HNO3. For all channels, geometry optimization and frequency calculations were performed at the M06-2X/6-31+G** level, while relative energies were improved at the UCCSD(T*)-F12/CBS level. The major channel is found to be the H abstraction channel, to give the products H2O + CH2ONO2. The reaction enthalpy (Delta H-298 KRX) of this channel is computed as -17.90 kcal mol(-1). Although the other reaction channels are also exothermic, their reaction barriers are high (>24 kcal mol(-1)), and therefore these reactions do not contribute to the overall rate coefficient in the temperature range considered (200-400 K). Pathways via three transition states were identified for the H abstraction channel. Rate coefficients were calculated for these pathways at various levels of variational transition state theory including tunneling. The results obtained are used to distinguish between two sets of experimental rate coefficients, measured in the temperature range of 200-400 K, one of which is approximately an order of magnitude greater than the other. This comparison, as well as the temperature dependence of the computed rate coefficients, shows that the lower experimental values are favored. The implications of the results to atmospheric chemistry are discussed.
-
Theoretical Study on the Reaction Mechanism of CH3CH2+O(3P)
Yang Yong,Zhang Weijun,Gao Xiaoming,Pei Shixin,Shao Jie,Huang Wei,Qu Jun
DOI: https://doi.org/10.3969/j.issn.1674-0068.2005.04.010
IF: 1.09
2005-01-01
Chinese Journal of Chemical Physics
Abstract:The reaction for CH 3CH 2 + O( 3P) was studied by ab initio method. The geometries of the reactants, intermediates, transition states and products were optimized at MP2/6-311 + G(d,p) level. The corresponding vibration frequencies were calculated at the same level. The single - point calculations for all the stationary points were carried out at the QCISD(T)/6-311 +G(d,p) level using the MP2/6-311 + G(d,p) optimized geometries. The results of the theoretical study indicate that the major products are the CH 2O + CH 3, CH 3CHO + H and CH 2CH 2 + OH in the reaction. For the products CH 2O + CH 3 and CH 3CHO + H, the major production channels are Al: (R)→ IM1→ TS3 → (A) and B1: (R) → IM1→ TS4 → (B), respectively. The majority of the products CH 2CH 2 + OH are formed via the direct abstraction channels C1 and C2: (R) → TS1 (TS2) →(C). In addition, the results suggest that the barrier heights to form the CO reaction channels are very high, so the CO is not a major product in the reaction.
-
Stationary points on the energy hypersurface of the reaction O3 + H•→ [•O3H]* ⇆ O2 + •OH and thermodynamic functions of •O3H at G3MP2B3, CCSD(T)-CBS (W1U) and MR-ACPF-CBS levels of theory
W. M. F. Fabian,J. Kalcher,R. Janoschek
DOI: https://doi.org/10.1007/s00214-005-0659-7
2005-09-01
Theoretical Chemistry Accounts
Abstract:The relative enthalpies, ΔHo (0) and ΔHo (298.15), of stationary points (four minimum and three transition structures) on the •O3H potential energy surface were calculated with the aid of the G3MP2B3 as well as the CCSD(T)–CBS (W1U) procedures from which we earlier found mean absolute deviations (MAD) of 3.9 kJ mol−1 and 2.3 kJ mol−1, respectively, between experimental and calculated standard enthalpies of the formation of a set of 32 free radicals. For CCSD(T)-CBS (W1U) the well depth from O3 + H• to trans-•O3H, ΔHowell(298.15) = −339.1 kJ mol−1, as well as the reaction enthalpy of the overall reaction O3 + H•→O2 + •OH, ΔrHo(298.15) = −333.7 kJ mol−1, and the barrier of bond dissociation of trans-•O3H → O2 + •OH, ΔHo(298.15) = 22.3 kJ mol−1, affirm the stable short-lived intermediate •O3H. In addition, for radicals cis-•O3H and trans-•O3H, the thermodynamic functions heat capacity Cop(T), entropy So (T), and thermal energy content Ho(T) − Ho(0) are tabulated in the range of 100 − 3000 K. The much debated calculated standard enthalpy of the formation of the trans-•O3H resulted to be ΔfHo(298.15) = 31.1 kJ mol −1 and 32.9 kJ mol −1, at the G3MP2B3 and CCSD(T)-CBS (W1U) levels of theory, respectively. In addition, MR-ACPF-CBS calculations were applied to consider possible multiconfiguration effects and yield ΔfHo(298.15) = 21.2 kJ mol −1. The discrepancy between calculated values and the experimental value of −4.2 ± 21 kJ mol−1 is still unresolved.
chemistry, physical
-
CH3NHNH2 + OH Reaction: Mechanism and Dynamics Studies
Hong-Xia Liu,Ying Wang,Lei Yang,Jing-Yao Liu,Hong Gao,Ze-Sheng Li,Chia-Chung Sun
DOI: https://doi.org/10.1002/jcc.21228
2009-01-01
Journal of Computational Chemistry
Abstract:A direct dynamics study was carried out for the multichannel reaction of CH(3)NHNH(2) with OH radical. Two stable Conformers (I, II) of CH(3)NHNH(2) are identified by the rotation of the -CH(3) group. For each conformer, five hydrogen-abstraction channels are found. The reaction mechanisms of product radicals (CH(3)NNH(2) and CH(3)NHNH) with OH radical are also investigated theoretically. The electronic structure information on the potential energy surface is obtained at the B3LYP/6-311G(d,p) level and the energetics along the reaction path is refined by the BMC-CCSD method. Hydrogen-bonded complexes are presented at both the reactant and product sides of the five channels, indicating that the reaction may proceed via an indirect mechanism. The influence of the basis set superposition error (BSSE) on the energies of all the complexes is discussed by means of the CBS-QB3 method. The rate constants of CH(3)NHNH(2) + OH are calculated using canonical variational transition-state theory with the small-curvature tunneling correction (CVT/SCT) in the temperature range of 200-1000 K. Slightly negative temperature dependence of rate constant is found in the temperature range from 200 to 345 K. The agreement between the theoretical and experimental results is good. It is shown that for Conformer I, hydrogen-abstraction from -NH- position is the primary pathway at low temperature; the hydrogen-abstraction from -NH(2) is a competitive pathway as the temperature increases. A similar case can be concluded for Conformer II. The overall rate constant is evaluated by considering the weight factors of each conformer from the Boltzmann distribution function, and the three-term Arrhenius expressions are fitted to be k(T) = 1.6 x 10(-24)T(4.03)exp (1411.5/T) cm(3) molecule(-1) s(-1) between 200-1000 K.
-
Role of Conformational Structures and Torsional Anharmonicity in Controlling Chemical Reaction Rates and Relative Yields: Butanal + HO2reactions
Jingjing Zheng,Prasenjit Seal,Donald G. Truhlar
DOI: https://doi.org/10.1039/c2sc21090h
IF: 8.4
2013-01-01
Chemical Science
Abstract:Aldehyde-radical reactions are important in atmospheric and combustion chemistry, and the reactions studied here also serve more generally to illustrate a fundamental aspect of chemical kinetics that has been relatively unexplored from a quantitative point of view, in particular the roles of multiple structures and torsional anharmonicity in determining the rate constants and branching ratios (product yields). We consider hydrogen abstraction from four carbon sites of butanal (carbonyl-C, alpha-C, beta-C and gamma-C) by hydroperoxyl radical. We employed multi-structural variational transition state theory for studying the first three channels; this uses a multi-faceted dividing surface and allows us to include the contributions of multiple structures of both reacting species and transition states. Multi-configurational Shepard interpolation (MCSI) was used to obtain the geometries and energies of the potential energy surface along the minimum-energy paths, with gradients and Hessians calculated by the M08-HX/maugcc-pVTZ method. We find the numbers of structures obtained for the transition states are 46, 60, 72 and 76respectively for the H abstraction at the carbonyl C, the alpha position, the beta position and the gamma position. Our results show that neglecting the factors arising from multiple structures and torsional anharmonicity would lead to errors at 300, 1000 and 2400 K of factors of 8, 11 and 10 for abstraction at the carbonyl-O, 2, 11 and 25 at the alpha-C position, 2, 23 and 47 at the beta-C position, and 0.6, 8 and 18 at the gamma-C position. The errors would be even larger at high temperature for the reverse of the H abstraction at the beta-C. Relative yields are changed as much as a factor of 7.0 at 200 K, a factor of 5.0 at 298 K, and a factor of 3.7 in the other direction at 2400 K. The strong dependence of the product ratios on the multi-structural anharmonicity factors shows that such factors play an important role in controlling branching ratios in reaction mechanism networks.
-
Theoretical study of the radical–radical reactions between HOCH2OO and OH
Xiang, Tiancheng
DOI: https://doi.org/10.1007/s00214-022-02900-x
2022-07-28
Theoretical Chemistry Accounts
Abstract:The reaction mechanism of HOCH 2 OO with OH was elucidated through density functional theory calculations, which indicated that the reaction could occur on two potential energy surfaces. On the singlet PES, the major products are HCOOH + H 2 O + O( 1 D), which are produced from the adduct 1 im2 by surmounting an energy barrier of 109.0 kJ mol −1 . On the triplet PES, the calculational results indicated a minimal energy path that begins from a barrierless addition of HOCH 2 OO to OH, generating the adduct 3 im2, followed by a H transfer from 3 im2, which results in the formation of H 2 CO + H 2 O + O 2 ( 3 P). The rate constants of four channels were calculated in the temperature range of 200–1000 K. The results show that the title reaction mainly occurs on the singlet PES and the dominant products are HCOOH + H 2 O + O( 1 D) in the calculating temperature range. The H 2 CO + H 2 O + O 2 ( 3 P) should be the minor products at high temperature.
chemistry, physical
-
Theoretical Study On The Ion-Molecule Reaction Of Nh+ With Ch2o
JianChao Song,Huiling Liu,Zhongjun Zhou,Xuri Huáng
DOI: https://doi.org/10.1002/qua.23044
2012-01-01
International Journal of Quantum Chemistry
Abstract:An in-depth theoretical study is carried out at the B3LYP/6-311G(d,p), M062X/aug-cc-pVDZ and CCSD(T)/6-311++G(3df,2dp) (single-point) levels as an attempt to explore the mechanism of the little-understood ionmolecule reaction between NH+ and CH2O. Various possible reaction pathways are taken into account. It is shown that six dissociation products, including P1 (2N + CH2OH+), P2 (4N + CH2OH+), P3 (3NH + CH2O+), P4 (NH2 + HCO+), P5 (NH?3+ + CO), and P9 (H + CONH?2+) are all accessible both kinetically and thermodynamically. Among these products, P4 is the most competitive product with predominant abundance, and the second most feasible product is P3, followed by P2 and P1. The remaining products, P5 and P9, may have negligible yield under room temperature condition. As the intermediates and transition states involved in the NH+ + CH2O reaction all stay below the reactant, the title reaction is expected to be rapid, which is consistent with the measured large rate constant in experiment. The present study will enrich our knowledge of the chemistry of NH+. Furthermore, our calculated result is compared with the previous experimental research, and, meanwhile, it provides a useful guide for understanding analogous reaction, NH+ with CH2NH. (c) 2011 Wiley Periodicals, Inc. Int J Quantum Chem, 2012