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.
-
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.
-
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.
-
ATM: A mediator of multiple responses to genotoxic stress
G. Rotman,Y. Shiloh
DOI: https://doi.org/10.1038/sj.onc.1203124
IF: 8.756
1999-11-01
Oncogene
Abstract:
-
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.
-
An ab initio/RRKM study of the reaction mechanism and product branching ratios of CH3OH+ and CH3OH++ dissociation
Cuiyu Li,Chih-Hao Chin,Tong Zhu,John Zeng Hui Zhang
DOI: https://doi.org/10.1016/j.molstruc.2020.128410
IF: 3.841
2020-01-01
Journal of Molecular Structure
Abstract:Regarding CH3OH+ and CH3OH++, theoretical calculations have used a variety of methods to describe geometric structures and potential energy surfaces. Rice–Ramsperger–Kassel–Marcus (RRKM) theory has been applied to compute the rate constants and product branching ratios of various channels on potential energy surfaces. The dissociate ions produced from CH3OH+ include CH2+, HCOH+ and CH2OH+, whereas H+, H2+, H3+, COH+/HCO+, and CH3O+ fragments are generated from CH3OH++. By roaming, we imply that a neutral hydrogen molecule fragment explores relatively flat regions of the intrinsic reaction coordinate calculations from the minimum energy path.
-
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
-
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 of Stereodynamics for Reaction O(~3P)+HCl
Zhu Tong,Hu Guo-Dong,Chen Jian-Zhong,Liu Xin-Guo,Zhang Qing-Gang
DOI: https://doi.org/10.1088/1674-1056/19/8/083402
2010-01-01
Chinese Physics B
Abstract:The vector correlation between products and reagents for reaction O(P-3)+HCl -> OH+Cl is studied using a quasi-classical trajectory (QCT) method on the benchmark potential energy surface of the ground (3)A" state [Ramachandran and Peterson, J. Chem. Phys. 119 (2003) 9550]. The generalised differential cross section (2 pi/sigma)(d sigma 00/d omega t) is presented in the centre of mass frame. The distribution of dihedral angles. P(phi(T)),), and the distribution of angles between k and j', P(theta(T)), are calculated. The influence of the collision energy and the influence of the reagent rotation and vibration on the product polarization are studied in the present work. The calculated results indicate that the rotational polarization of product molecule is almost independent of collision energy but sensitive to the reagent rotation and vibration.
-
Quasi-classical Trajectory Study of the Reaction O(3P) + HCl → OH + Cl and O(3P) + DCl → OD + Cl: Vector and Scalar Properties
Tong Zhu,Guodong Hu,Qinggang Zhang
DOI: https://doi.org/10.1016/j.theochem.2010.02.015
2010-01-01
Journal of Molecular Structure THEOCHEM
Abstract:Quasi-classical trajectory calculations are carried out for the O(3P)+HCl and its isotopic reactions on the benchmark potential energy surfaces for both 3A′′ and 3A′ electronic states. The results indicate that the reaction probability calculated on the 3A′ PES is much smaller than that on the 3A′′ PES. The product rotational polarization calculated on the 3A′ PES is stronger than that on the 3A′′ PES, implies that the effect of the van der Waals minima has significant influence on the product rotational polarization. There are notable variations in product rotational polarization on the 3A′′ PES when the H atom is substituted by the D atom, and the discrepancy can be attributed to the indirect reactive mechanism and the mass factor of these two reactions.
-
Quasi-Classical Trajectory Study Of Reaction O (P-3) Plus Hcl (V=2; J=1, 6, 9) -> Oh Plus Cl
Zhu Tong,Hu Guo-Dong,Zhang Qing-Gang
DOI: https://doi.org/10.1088/0256-307X/27/3/033102
2010-01-01
Chinese Physics Letters
Abstract:The reaction O (P-3)+HCl (v = 2; j = 1, 6, 9) -> OH+Cl is theoretically studied with a quasi-classical trajectory method (QCT) on the benchmark potential energy surface of the ground 3A '' state [J. Chem. Phys. 119(2003)9550]. The QCT-calculated state-resolved rotational distributions are in good agreement with the experimental results. The rotational polarization of the product OH molecule becomes weaker as the initial HCl rotation is excited. The calculated results can be explained from the large mass factor cos(2) beta of the title reaction, the van der Waals well in the potential energy surface and the secondary encounters in the exit channel.
-
Formation of alkoxymethyl hydroperoxides and alkyl formates from simplest Criegee intermediate (CH2OO) + ROH (R=CH3, CH3CH2, and (CH3)2CH) reaction systems
Manas Ranjan Dash,Balaganesh Muthiah,Subhashree Subhadarsini Mishra
DOI: https://doi.org/10.1007/s00214-024-03104-1
2024-03-23
Theoretical Chemistry Accounts
Abstract:Gas-phase reactions involving simplest Criegee intermediate (CH 2 OO) have been the current hot topic due to its vital role in atmospheric chemistry. In this study, high-level ab initio calculations are used to investigate the energetics and kinetics for the reaction of CH 2 OO + ROH → ROCHO + H 2 O (R=CH 3 , CH 3 CH 2 and (CH 3 ) 2 CH). Energies of the stationary points are computed at the CCSD(T)/M06-2X/6-311++G(3d,3pd)//M06-2X/6-311++G(3d,3pd) level of theory. Reaction is going through a 1,2-addition and water elimination step leading to the formation of alkoxymethyl hydroperoxides and alkyl formates, respectively. The barrier heights for the 1,2-addition step with methanol, ethanol, and isopropanol were found to be − 3.1, − 3.7, and − 4.8 kcal mol −1 , and water elimination steps were found to be 2.2, 1.5, and 1.6 kcal mol −1 , respectively, relative to the energies of the starting reactants. The rate constants for addition and elimination channels were calculated using canonical variational transition state theory in conjugation with small-curvature tunneling and the interpolated single point energy method between the temperature range of 200 and 500 K. In addition, the thermochemistry analysis indicates that addition and elimination channels are thermodynamically feasible and the formation of alkyl formates is entropically more favored when compared to the formation of alkoxymethyl hydroperoxide along the reaction path in the potential energy surface. The pressure-dependent microcanonical rate constants for both addition and elimination channels were also estimated using the Rice–Ramsperger–Kassel–Marcus theory and discussed in this study.
chemistry, physical
-
Gas-phase Kinetics Study of Reaction of OH Radical with CH3NHNH2 by Second-Order Multireference Perturbation Theory.
Hongyan Sun,Peng Zhang,Chung K. Law
DOI: https://doi.org/10.1021/jp3021529
2012-01-01
The Journal of Physical Chemistry A
Abstract:The gas-phase kinetics of H-abstraction reactions of monomethylhydrazine (MMH) by OH radical was investigated by second-order multireference perturbation theory and two-transition-state kinetic model. It was found that the abstractions of the central and terminal amine H atoms by the OH radical proceed through the formation of two hydrogen bonded preactivated complexes with energies of 6.16 and 5.90 kcal mol(-1) lower than that of the reactants, whereas the abstraction of methyl H atom is direct. Due to the multireference characters of the transition states, the geometries and ro-vibrational frequencies of the reactant, transition states, reactant complexes, and product complexes were optimized by the multireference CASPT2/aug-cc-pVTZ method, and the energies of the stationary points of the potential energy surface were refined at the QCISD(T)/CBS level via extrapolation of the QCISD(T)/cc-pVTZ and QCISD(T)/cc-pVQZ energies. It was found that the abstraction reactions of the central and two terminal amine H atoms of MMH have the submerged energy barriers with energies of 2.95, 2.12, and 1.24 kcal mol(-1) lower than that that of the reactants respectively, and the abstraction of methyl H atom has a real energy barrier of 3.09 kcal mol(-1). Furthermore, four MMH radical-H2O complexes were found to connect with product channels and the corresponding transition states. Consequently, the rate coefficients of MMH + OH for the H-abstraction of the amine H atoms were determined on the basis of a two-transition-state model, with the total energy E and angular momentum J conserved between the two transition-state regions. In units of cm(3) molecule(-1) s(-1), the rate coefficient was found to be k(1) = 3.37 x 10(-16)T(1.295) exp(1126.17/T) for the abstraction of the central amine H to form the (CH3NNH2)-N-center dot radical, k(2) = 2.34 X 10(-17)T(1.907) exp(1052.26/T) for the abstraction of the terminal amine H to form the trans-(CH3NHNH)-H-center dot radical, k(3) = 7.41 X 10(-20)T(2.428) exp(1343.20/T) for the abstraction of the terminal amine H to form the cis-(CH3NHNH)-H-center dot radical, and k(4) = 9.13 x 10(-21)T(2.964) exp(-114.09/T) for the abstraction of the methyl H atom to form the (CH2NHNH2)-H-center dot radical, respectively. Assuming that the rate coefficients are additive, the total rate coefficient of these theoretical predictions quantitatively agrees with the measured rate constant at temperatures of 200-650 K, with no adjustable parameters.
-
Intermediates Involved in the Oxidation of Nitrogen Monoxide: Photochemistry of the Cis-N(2)o(2)o(2) Complex and of Sym-N(2)o(4) in Solid Ne Matrices.
Helmut Beckers,Xiaoqing Zeng,Helge Willner
DOI: https://doi.org/10.1002/chem.200902406
2009-01-01
Chemistry - A European Journal
Abstract:Pure sym-N(2)O(4) isolated in solid Ne was obtained by passing cold neon gas over solid N(2)O(4) at -115 degrees C and quenching the resulting gaseous mixture at 6.3 K. Filtered UV irradiation (260-400 nm) converts sym-N(2)O(4) into trans-ONONO(2), a weakly interacting (NO(2))(2) radical pair, and traces of the cis-N(2)O(2)O(2) complex. Besides the weakly bound ONO(2) complex, cis-N(2)O(2)O(2) was also obtained by co-deposition of NO and O(2) in solid Ne at 6.3 K, and both complexes were characterised by their matrix IR spectra. Concomitantly formed cis-N(2)O(2) dissociated on exposure to filtered IR irradiation (400-8000 cm(-1)), and the cis-N(2)O(2)O(2) complex rearranged to sym-N(2)O(4) and trans-ONONO(2). Experiments using (18)O(2) in place of (16)O(2) revealed a non-concerted conversion of cis-N(2)O(2)O(2) into these species, and gave access to four selectively di-(18)O-substituted trans-ONONO(2) isotopomers. No isotopic scrambling occurred. The IR spectra of sym-N(2)O(4) and of trans-ONONO(2) in solid Ne were recorded. IR fundamentals of trans-ONONO(2) were assigned based on experimental (16/18)O isotopic shifts and guided by DFT calculations. Previously reported contradictory measurements on cis- and trans-ONONO(2) are discussed. Dinitroso peroxide, ONOONO, a proposed intermediate in the IR photoinduced rearrangement of cis-N(2)O(2)O(2) to the various N(2)O(4) species, was not detected. Its absence in the photolysis products indicates a low barrier (<or=10 kJ mol(-1)) for its exothermic O-O bond homolysis into a (NO(2))(2) radical pair.
-
A Kinetic Study of the Gas-Phase O( 1 D) + CH3OH and O( 1 D) + CH3CN Reactions. Low Temperature Rate Constants and Atomic Hydrogen Product Yields
Kevin K.M. Hickson,Jean-Christophe Loison
DOI: https://doi.org/10.1021/acs.jpca.2c01946
2022-07-07
Abstract:Atomic oxygen in its first excited singlet state, O(1 D), is an important species in the photochemistry of several planetary atmospheres and has been predicted to be a potentially important reactive species on interstellar ices. Here, we report the results of a kinetic study of the reactions of O(1 D) with methanol, CH3OH, and acetonitrile, CH3CN, over the 50-296 K temperature range. A continuous supersonic flow reactor was used to attain these low temperatures coupled with pulsed laser photolysis and pulsed laser induced fluorescence to generate and monitor O(1 D) atoms respectively. Secondary experiments examining the atomic hydrogen product channels of these reactions were also performed, through laser induced fluorescence measurements of H(2 S) atom formation. On the kinetics side, the rate constants for these reactions were seen to be large (> 2 x 10-10 cm 3 s-1) and consistent with barrierless reactions, although they display contrasting dependences as a function of temperature. On the product formation side, both reactions are seen to yield non-negligible quantities of atomic hydrogen. For the O(1 D) + CH3OH reaction, the derived yields are in good agreement with the conclusions of previous experimental and theoretical work. For the O(1 D) + CH3CN reaction, whose H-atom formation channels had not previously been investigated, electronic structure calculations of several new product formation channels were performed to explain the observed H-atom yields. These calculations demonstrate the barrierless and exothermic nature of the relevant exit channels, confirming that atomic hydrogen is also an important product of the O(1 D) + CH3CN reaction.
Chemical Physics
-
Dissection of the Multichannel Reaction O(3P) + C2H2: Differential Cross-Sections and Product Energy Distributions
Shuwen Zhang,Qixin Chen,Junxiang Zuo,Xixi Hu,Daiqian Xie
DOI: https://doi.org/10.3390/molecules27030754
IF: 4.6
2022-01-01
Molecules
Abstract:The O(3P) + C2H2 reaction plays an important role in hydrocarbon combustion. It has two primary competing channels: H + HCCO (ketenyl) and CO + CH2 (triplet methylene). To further understand the microscopic dynamic mechanism of this reaction, we report here a detailed quasi-classical trajectory study of the O(3P) + C2H2 reaction on the recently developed full-dimensional potential energy surface (PES). The entrance barrier TS1 is the rate-limiting barrier in the reaction. The translation of reactants can greatly promote reactivity, due to strong coupling with the reaction coordinate at TS1. The O(3P) + C2H2 reaction progress through a complex-forming mechanism, in which the intermediate HCCHO lives at least through the duration of a rotational period. The energy redistribution takes place during the creation of the long-lived high vibrationally (and rotationally) excited HCCHO in the reaction. The product energy partitioning of the two channels and CO vibrational distributions agree with experimental data, and the vibrational state distributions of all modes of products present a Boltzmann-like distribution.
-
Formations of C6H from reactions C3 + C3H2 and C3H + C3H and of C8H from reactions C4 + C4H2 and C4H + C4H
Yi-Lun Sun,Wen-Jian Huang,Shih-Huang Lee
DOI: https://doi.org/10.1063/5.0184683
2024-01-28
Abstract:We interrogated C6H and C8H produced separately from the reactions C3 + C3H2/C3H + C3H/C3H2 + C3 → C6H + H and C4 + C4H2/C4H + C4H/C4H2 + C4 → C8H + H using product translational and photoionization spectroscopy. Individual contributions of the three reactions to the product C6H or C8H were evaluated with reactant concentrations. Translational-energy distributions, angular distributions, and photoionization efficiency curves of products C6H and C8H were unraveled. The product C6H (C8H) was recognized as the most stable linear isomer by comparing its photoionization efficiency curve with that of l-C6H (l-C8H), produced exclusively from the reaction C2 + C4H2 → l-C6H + H (C2 + C6H2 → l-C8H + H). The ionization threshold after deconvolution was determined to be 9.3 ± 0.1 eV for l-C6H and 8.9 ± 0.1 eV for l-C8H, which is in good agreement with theoretical values. Quantum-chemical calculations indicate that the reactions of C3 + C3H2 and C3H + C3H (C4 + C4H2 and C4H + C4H) incur no energy barriers that lie above the corresponding reactant and the most stable product l-C6H (l-C8H) with H on the lower-lying potential-energy surfaces. The theoretical calculation is in accord with the experimental observation. This work implies that the reactions of C3 + C3H2/C3H + C3H and C4 + C4H2/C4H + C4H need to be taken into account for the formation of interstellar C6H and C8H, respectively.
-
Ab Initio Kinetics for Thermal Decomposition of CH3N•NH2, Cis-Ch3nhn•h, Trans-Ch3nhn•h, and C•H2NNH2 Radicals.
Hongyan Sun,Peng Zhang,Chung K. Law
DOI: https://doi.org/10.1021/jp3045675
2012-01-01
The Journal of Physical Chemistry A
Abstract:The thermal decomposition of the CH3N center dot NH2, cis-(CH3NHNH)-H-center dot, trans-(CH3NHNH)-H-center dot, and (CH2NNH2)-H-center dot radicals, which are the four radical products from the H-abstraction reactions of monomethylhydrazine, were theoretically studied by using ab initio Rice-Ramsperger-Kassel-Marcus (RRKM) transition-state theory and master equation analysis. Various decomposition pathways were identified by using either the QCISD(T)/cc-pV infinity Z//CASPT2/aug-cc-pVTZ or the QCISD(T)/cc-pV infinity Z//B3LYP/6-311++G(d,p) quantum chemistry methods. The results reveal that the beta-scission of NH2 to form methyleneimine is the predominant channel for the decomposition of the (CH2NNH2)-H-center dot radical due to its small energy barrier of 13.8 kcal mol(-1). The high pressure limit rate coefficient for the reaction is fitted by 3.88 x 10(19)T(-1.672) exp(-9665.13/T) s(-1). In addition, the pressure dependent rate coefficients exhibit slight temperature dependence at temperatures of 1000-2500 K. The cis-(CH3NHNH)-H-center dot and trans-(CH3NHNH)-H-center dot radicals are the two distinct spatial isomers with an energy barrier of 26 kcal mol(-1) for their isomerization. The beta-scission of CH3 from the cis-(CH3NHNH)-H-center dot radical to form trans-diazene has an energy barrier of 35.2 kcal mol(-1), and the beta-scission of CH3 from the trans-(CH3NHNH)-H-center dot radical to form cis-diazene has an energy barrier of 39.8 kcal mol(-1). The (CH3NNH2)-N-center dot radical undergoes the beta-scission of methyl hydrogen and amine hydrogen to form CH2=NNH2, trans-CH3N=NH, and cis-CH3N=NH products, with the energy barriers of 42.8, 46.0, and 50.2 kcal mol(-1), respectively. The dissociation and isomerization rate coefficients for the reactions were calculated via the E/J resolved RRKM theory and multiple well master equation analysis at temperatures of 300-2500 K and pressures of 0.01-100 atm. The calculated rate coefficients associated with updated thermochemical property data are essential components in the development of kinetic mechanisms for the pyrolysis and oxidation of MMH and its derivatives.
-
Anab Initiobased Full-Dimensional Potential Energy Surface for OH + O2⇄ HO3and Low-Lying Vibrational Levels of HO3
Xixi Hu,Junxiang Zuo,Changjian Xie,Richard Dawes,Hua Guo,Daiqian Xie
DOI: https://doi.org/10.1039/c9cp02206f
2019-01-01
Abstract:To provide an in-depth understanding of the HO3 radical and its dissociation to OH + O2, a six-dimensional potential energy surface (PES) has been constructed by fitting 2087 energy points for the electronic ground state of HO3 (X2A'') using the permutation invariant polynomial-neural network (PIP-NN) approach. The energy points were calculated using an explicitly-correlated and Davidson-corrected multi-reference configuration interaction method with the correlation-consistent polarized valence double zeta basis (MRCI(Q)-F12/VDZ-F12). On the PES, the trans-HO3 isomer is found to be the global minimum, 33.0 cm-1 below the cis-HO3 conformer, which is consistent with previous high-level theoretical investigations. The dissociation to the OH + O2 asymptote from both conformers is shown to be barrierless. As a benchmark from a recently developed high-accuracy thermochemistry protocol, D0 for trans-HO3 is calculated to be 2.29 ± 0.36 kcal mol-1, only slightly deeper than the value of 2.08 kcal mol-1 obtained using the PES, and in reasonable agreement with the experimentally estimated value of 2.93 ± 0.07 kcal mol-1. Using this PES, low-lying vibrational energy levels of HO3 are determined using an exact quantum Hamiltonian and compared with available experimental results.