全部 标题 作者
关键词 摘要

OALib Journal期刊
ISSN: 2333-9721
费用:99美元

查看量下载量

相关文章

更多...

Exploring Multiple Potential Energy Surfaces: Photochemistry of Small Carbonyl Compounds

DOI: 10.1155/2012/268124

Full-Text   Cite this paper   Add to My Lib

Abstract:

In theoretical studies of chemical reactions involving multiple potential energy surfaces (PESs) such as photochemical reactions, seams of intersection among the PESs often complicate the analysis. In this paper, we review our recipe for exploring multiple PESs by using an automated reaction path search method which has previously been applied to single PESs. Although any such methods for single PESs can be employed in the recipe, the global reaction route mapping (GRRM) method was employed in this study. By combining GRRM with the proposed recipe, all critical regions, that is, transition states, conical intersections, intersection seams, and local minima, associated with multiple PESs, can be explored automatically. As illustrative examples, applications to photochemistry of formaldehyde and acetone are described. In these examples as well as in recent applications to other systems, the present approach led to discovery of many unexpected nonadiabatic pathways, by which some complicated experimental data have been explained very clearly. 1. Introduction In order to theoretically unravel the entire photochemical reaction processes of a given system, one has to explore and characterize systematically several excited as well as the ground state potential energy surfaces (PESs). In the Franck-Condon (FC) approximation, a reaction starts from the FC point on an excited state PES. Then, the system undergoes either adiabatic or nonadiabatic pathways depending on a given excess energy and topographies on the PESs. In adiabatic paths, a bond reorganization occurs directly on the excited state PES through transition states (TSs). In nonadiabatic paths, on the other hand, the system makes nonadiabatic transition to a lower PES and then reacts on the lower PES. The system may cascade through several PESs via nonadiabatic transitions. Seams of intersection between two PESs often play a key role in various reactions [1–7]; near an intersection seam, nonadiabatic transitions can take place efficiently. When the two states have the same spin and space symmetry, the intersection hyperspace consists of f-2 dimensions and is called “conical intersection (CI),” where f is the number of vibrational degrees of freedom. If the spin multiplicity and/or the space symmetry are different, the PESs cross simply in f-1 dimensional hyperspace. In this paper, both of these two types of intersections, that is, conical intersection and simple intersection seam, are denoted as “seam-of-crossing.” In many kinetic studies, minimum energy structures on seam-of-crossing hyperspaces (MSXs)

References

[1]  F. Bernardi, M. Olivucci, and M. A. Robb, “Potential energy surface crossings in organic photochemistry,” Chemical Society Reviews, vol. 25, no. 5, pp. 321–328, 1996.
[2]  D. R. Yarkony, “Conical intersections: diabolical and often misunderstood,” Accounts of Chemical Research, vol. 31, no. 8, pp. 511–518, 1998.
[3]  D. Schr?der, S. Shaik, and H. Schwarz, “Two-state reactivity as a new concept in organometallic chemistry,” Accounts of Chemical Research, vol. 33, no. 3, pp. 139–145, 2000.
[4]  A. L. Sobolewski, W. Domcke, C. Dedonder-Lardeux, and C. Jouvet, “Excited-state hydrogen detachment and hydrogen transfer driven by repulsive 1πσ* states: a new paradigm for nonradiative decay in aromatic biomolecules,” Physical Chemistry Chemical Physics, vol. 4, no. 7, pp. 1093–1100, 2002.
[5]  R. Poli and J. N. Harvey, “Spin forbidden chemical reactions of transition metal compounds. New ideas and new computational challenges,” Chemical Society Reviews, vol. 32, no. 1, pp. 1–8, 2003.
[6]  B. G. Levine and T. J. Martínez, “Isomerization through conical intersections,” Annual Review of Physical Chemistry, vol. 58, pp. 613–634, 2007.
[7]  S. Nanbu, T. Ishida, and H. Nakamura, “Future perspectives of nonadiabatic chemical dynamics,” Chemical Science, vol. 1, pp. 663–674, 2010.
[8]  N. Koga and K. Morokuma, “Determination of the lowest energy point on the crossing seam between two potential surfaces using the energy gradient,” Chemical Physics Letters, vol. 119, no. 5, pp. 371–374, 1985.
[9]  M. R. Manaa and D. R. Yarkony, “On the intersection of two potential energy surfaces of the same symmetry. Systematic characterization using a Lagrange multiplier constrained procedure,” The Journal of Chemical Physics, vol. 99, no. 7, pp. 5251–5256, 1993.
[10]  J. M. Anglada and J. M. Bofill, “A reduced-restricted-quasi-Newton-Raphson method for locating and optimizing energy crossing points between two potential energy surfaces,” Journal of Computational Chemistry, vol. 18, no. 8, pp. 992–1003, 1997.
[11]  M. J. Bearpark, M. A. Robb, and H. Bernhard Schlegel, “A direct method for the location of the lowest energy point on a potential surface crossing,” Chemical Physics Letters, vol. 223, no. 3, pp. 269–274, 1994.
[12]  F. Sicilia, L. Blancafort, M. J. Bearpark, and M. A. Robb, “New algorithms for optimizing and linking conical intersection points,” Journal of Chemical Theory and Computation, vol. 4, no. 2, pp. 257–266, 2008.
[13]  C. Ciminelli, G. Granucci, and M. Persico, “The photoisomerization mechanism of azobenzene: a semiclassical simulation of nonadiabatic dynamics,” Chemistry, vol. 10, no. 9, pp. 2327–2341, 2004.
[14]  B. G. Levine, J. D. Coe, and T. J. Martínez, “Optimizing conical intersections without derivative coupling vectors: application to multistate multireference second-order perturbation theory (MS-CASPT2),” Journal of Physical Chemistry B, vol. 112, no. 2, pp. 405–413, 2008.
[15]  T. W. Keal, A. Koslowski, and W. Thiel, “Comparison of algorithms for conical intersection optimisation using semiempirical methods,” Theoretical Chemistry Accounts, vol. 118, no. 5-6, pp. 837–844, 2007.
[16]  S. Maeda, K. Ohno, and K. Morokuma, “Updated branching plane for finding conical intersections without coupling derivative vectors,” Journal of Chemical Theory and Computation, vol. 6, no. 5, pp. 1538–1545, 2010.
[17]  J. W. McIver Jr. and A. Komornicki, “Structure of transition states in organic reactions. General theory and an application to the cyclobutene-butadiene isomerization using a semiempirical molecular orbital method,” Journal of the American Chemical Society, vol. 94, no. 8, pp. 2625–2633, 1972.
[18]  A. Komornicki, K. Ishida, K. Morokuma, R. Ditchfield, and M. Conrad, “Efficient determination and characterization of transition states using ab-initio methods,” Chemical Physics Letters, vol. 45, no. 3, pp. 595–602, 1977.
[19]  H. B. Schlegel, “Optimization of equilibrium geometries and transition structures,” Journal of Computational Chemistry, vol. 3, no. 2, pp. 214–218, 1982.
[20]  ? Farkas and H. B. Schlegel, “Methods for optimizing large molecules. II. Quadratic search,” Journal of Chemical Physics, vol. 111, no. 24, pp. 10806–10814, 1999.
[21]  C. J. Cerjan and W. H. Miller, “On finding transition states,” The Journal of Chemical Physics, vol. 75, no. 6, pp. 2800–2801, 1981.
[22]  P. Császár and P. Pulay, “Geometry optimization by direct inversion in the iterative subspace,” Journal of Molecular Structure, vol. 114, no. 1-2, pp. 31–34, 1984.
[23]  A. Banerjee, N. Adams, J. Simons, and R. Shepard, “Search for stationary points on surfaces,” Journal of Physical Chemistry, vol. 89, no. 1, pp. 52–57, 1985.
[24]  H. B. Schlegel, “Exploring potential energy surfaces for chemical reactions: an overview of some practical methods,” Journal of Computational Chemistry, vol. 24, no. 12, pp. 1514–1527, 2003.
[25]  F. Jensen, Introduction to Computational Chemistry, John Wiley & Sons, Chichester, UK, 2nd edition, 2007.
[26]  H. B. Schlegel, “Geometry optimization,” Wiley Interdisciplinary Reviews: Computational Molecular Science, vol. 1, no. 5, pp. 790–809, 2011.
[27]  T. A. Halgren and W. N. Lipscomb, “The synchronous-transit method for determining reaction pathways and locating molecular transition states,” Chemical Physics Letters, vol. 49, no. 2, pp. 225–232, 1977.
[28]  M. J. S. Dewar, E. F. Healy, and J. J. P. Stewart, “Location of transition states in reaction mechanisms,” Journal of the Chemical Society, Faraday Transactions 2: Molecular and Chemical Physics, vol. 80, no. 3, pp. 227–233, 1984.
[29]  R. Elber and M. Karplus, “A method for determining reaction paths in large molecules: application to myoglobin,” Chemical Physics Letters, vol. 139, no. 5, pp. 375–380, 1987.
[30]  G. Henkelman, B. P. Uberuaga, and H. Jónsson, “Climbing image nudged elastic band method for finding saddle points and minimum energy paths,” Journal of Chemical Physics, vol. 113, no. 22, pp. 9901–9904, 2000.
[31]  W. Ren and E. Vanden-Eijnden, “String method for the study of rare events,” Physical Review B, vol. 66, no. 5, Article ID 052301, 2002.
[32]  B. Peters, A. Heyden, A. T. Bell, and A. Chakraborty, “A growing string method for determining transition states: comparison to the nudged elastic band and string methods,” Journal of Chemical Physics, vol. 120, no. 17, pp. 7877–7886, 2004.
[33]  S. Maeda and K. Ohno, “A new approach for finding a transition state connecting a reactant and a product without initial guess: applications of the scaled hypersphere search method to isomerization reactions of HCN, (H2O)2, and alanine dipeptide,” Chemical Physics Letters, vol. 404, no. 1–3, pp. 95–99, 2005.
[34]  S. Maeda and K. Ohno, “Conversion pathways between a fullerene and a ring among C20 clusters by a sphere contracting walk method: remarkable difference in local potential energy landscapes around the fullerene and the ring,” Journal of Chemical Physics, vol. 124, no. 17, Article ID 174306, 2006.
[35]  M. ?ernohorsky, S. Kettou, and J. Ko?a, “VADER : new software for exploring interconversions on potential energy surfaces,” Journal of Chemical Information and Computer Sciences, vol. 39, no. 4, pp. 705–712, 1999.
[36]  A. Laio and M. Parrinello, “Escaping free-energy minima,” Proceedings of the National Academy of Sciences of the United States of America, vol. 99, no. 20, pp. 12562–12566, 2002.
[37]  S. K. Burger and P. W. Ayers, “Dual grid methods for finding the reaction path on reduced potential energy surfaces,” Journal of Chemical Theory and Computation, vol. 6, no. 5, pp. 1490–1497, 2010.
[38]  J. Q. Sun and K. Ruedenberg, “Gradient extremals and steepest descent lines on potential energy surfaces,” The Journal of Chemical Physics, vol. 98, no. 12, pp. 9707–9714, 1993.
[39]  K. Bondensg?rd and F. Jensen, “Gradient extremal bifurcation and turning points: an application to the H2CO potential energy surface,” Journal of Chemical Physics, vol. 104, no. 20, pp. 8025–8031, 1996.
[40]  W. Quapp, M. Hirsch, O. Imig, and D. Heidrich, “Searching for saddle points of potential energy surfaces by following a reduced gradient,” Journal of Computational Chemistry, vol. 19, no. 9, pp. 1087–1100, 1998.
[41]  K. K. Irikura and R. D. Johnson, “Predicting unexpected chemical reactions by isopotential searching,” Journal of Physical Chemistry A, vol. 104, no. 11, pp. 2191–2194, 2000.
[42]  E. M. Müller, A. de Meijere, and H. Grubmüller, “Predicting unimolecular chemical reactions: chemical flooding,” Journal of Chemical Physics, vol. 116, no. 3, pp. 897–905, 2002.
[43]  K. Ohno and S. Maeda, “A scaled hypersphere search method for the topography of reaction pathways on the potential energy surface,” Chemical Physics Letters, vol. 384, no. 4–6, pp. 277–282, 2004.
[44]  S. Maeda and K. Ohno, “Global mapping of equilibrium and transition structures on potential energy surfaces by the scaled hypersphere search method: applications to ab initio surfaces of formaldehyde and propyne molecules,” Journal of Physical Chemistry A, vol. 109, no. 25, pp. 5742–5753, 2005.
[45]  K. Ohno and S. Maeda, “Global reaction route mapping on potential energy surfaces of formaldehyde, formic acid, and their metal-substituted analogues,” Journal of Physical Chemistry A, vol. 110, no. 28, pp. 8933–8941, 2006.
[46]  S. Maeda and K. Morokuma, “Communications: a systematic method for locating transition structures of A+B→X type reactions,” Journal of Chemical Physics, vol. 132, no. 24, Article ID 241102, 2010.
[47]  S. Maeda, S. Komagawa, M. Uchiyama, and K. Morokuma, “Finding reaction pathways for multicomponent reactions: the passerini reaction is a four-component reaction,” Angewandte Chemie, vol. 50, no. 3, pp. 644–649, 2011.
[48]  S. Maeda and K. Morokuma, “"Finding reaction pathways of type A + B → X: toward systematic prediction of reaction mechanisms,” Journal of Chemical Theory and Computation, vol. 7, no. 8, pp. 2335–2345, 2011.
[49]  S. Maeda, K. Ohno, and K. Morokuma, “Automated global mapping of minimal energy points on seams of crossing by the anharmonic downward distortion following method: a case study of H2CO,” Journal of Physical Chemistry A, vol. 113, no. 9, pp. 1704–1710, 2009.
[50]  S. Maeda, R. Saito, and K. Morokuma, “Finding minimum structures on the seam of crossing in reactions of type A + B → X: exploration of nonadiabatic ignition pathways of unsaturated hydrocarbons,” Journal of Physical Chemistry Letters, vol. 2, no. 8, pp. 852–857, 2011.
[51]  S. Maeda, K. Ohno, and K. Morokuma, “A theoretical study on the photodissociation of acetone: insight into the slow intersystem crossing and exploration of nonadiabatic pathways to the ground state,” Journal of Physical Chemistry Letters, vol. 1, no. 12, pp. 1841–1845, 2010.
[52]  K. Fukui, “A formulation of the reaction coordinate,” Journal of Physical Chemistry, vol. 74, no. 23, p. 4161, 1970.
[53]  K. Ishida, K. Morokuma, and A. Komornicki, “The intrinsic reaction coordinate. An ab initio calculation for HNC→HCN and H? + CH4→CH4 + H?,” The Journal of Chemical Physics, vol. 66, no. 5, pp. 2153–2156, 1976.
[54]  K. Müller and L. D. Brown, “Location of saddle points and minimum energy paths by a constrained simplex optimization procedure,” Theoretica Chimica Acta, vol. 53, no. 1, pp. 75–93, 1979.
[55]  M. Page and J. W. McIver Jr., “On evaluating the reaction path Hamiltonian,” The Journal of Chemical Physics, vol. 88, no. 2, pp. 922–935, 1988.
[56]  C. Gonzalez and H. Bernhard Schlegel, “An improved algorithm for reaction path following,” The Journal of Chemical Physics, vol. 90, no. 4, pp. 2154–2161, 1989.
[57]  H. P. Hratchian and H. B. Schlegel, “Using Hessian updating to increase the efficiency of a Hessian based predictor-corrector reaction path following method,” Journal of Chemical Theory and Computation, vol. 1, no. 1, pp. 61–69, 2005.
[58]  S. Maeda and K. Ohno, “Structures of water octamers (H2O)8: exploration on ab initio potential energy surfaces by the scaled hypersphere search method,” Journal of Physical Chemistry A, vol. 111, no. 20, pp. 4527–4534, 2007.
[59]  H.-J. Werner, P. J. Knowles, F. R. Manby, et al., MOLPRO, version 2006.1, a package of ab initio programs, http://www.molpro.net.
[60]  S. Maeda, Y. Osada, K. Morokuma, and K. Ohno, GRRM11, http://kmweb.fukui.kyoto-u.ac.jp/~smaeda/GRRM_PR000.html.
[61]  W. M. Gelbart, M. L. Elert, and D. F. Heller, “Photodissociation of the formaldehyde molecule: does it or doesn't it?” Chemical Reviews, vol. 80, no. 5, pp. 403–416, 1980.
[62]  C. B. Moore and J. C. Weisshaar, “Formaldehyde photochemistry,” Annual Review of Physical Chemistry, vol. 34, pp. 525–555, 1983.
[63]  D. J. Clouthier and D. A. Ramsay, “The spectroscopy of formaldehyde and thioformaldehyde,” Annual Review of Physical Chemistry, vol. 34, pp. 31–58, 1983.
[64]  W. H. Green Jr., C. B. Moore, and W. F. Polik, “Transition states and rate constants for unimolecular reactions,” Annual Review of Physical Chemistry, vol. 43, no. 1, pp. 591–626, 1992.
[65]  C. B. Moore, “A spectroscopist's view of energy states, energy transfers, and chemical reactions,” Annual Review of Physical Chemistry, vol. 58, pp. 1–33, 2007.
[66]  D. M. Hayes and K. Morokuma, “Theoretical studies of carbonyl photochemistry. I. ab initio potential energy surfaces for the photodissociation H2CO*→H + HCO,” Chemical Physics Letters, vol. 12, no. 4, pp. 539–543, 1972.
[67]  R. L. Jaffe, D. M. Hayes, and K. Morokuma, “Photodissociation of formaldehyde-4: potential energy surfaces for H2CO* → H2 + CO,” The Journal of Chemical Physics, vol. 60, no. 12, pp. 5108–5109, 1974.
[68]  J. A. Hernandez and R. Carbo, “Theoretical-study of formaldehyde photodissociation,” Journal de Chimie Physique et de Physico-Chimie Biologique, vol. 72, no. 9, pp. 959–960, 1975.
[69]  R. L. Jaffe and K. Morokuma, “MCSCF potential energy surface for photodissociation of formaldehyde,” The Journal of Chemical Physics, vol. 64, no. 12, pp. 4881–4886, 1976.
[70]  J. D. Goddard and H. F. Schaefer, “The photodissociation of formaldehyde: potential energy surface features,” The Journal of Chemical Physics, vol. 70, no. 11, pp. 5117–5134, 1979.
[71]  N. C. Handy and S. Carter, “An improved potential surface for formaldehyde,” Chemical Physics Letters, vol. 79, no. 1, pp. 118–124, 1981.
[72]  R. Schinke, “Rotational state distributions of H2 and CO following the photofragmentation of formaldehyde,” The Journal of Chemical Physics, vol. 84, no. 3, pp. 1487–1491, 1986.
[73]  G. E. Scuseria and H. F. Schaefer, “The photodissociation of formaldehyde: a coupled cluster study including connected triple excitations of the transition state barrier height for H2CO→H2 + CO,” The Journal of Chemical Physics, vol. 90, no. 7, pp. 3629–3636, 1989.
[74]  Y. T. Chang, C. Minichino, and W. H. Miller, “Classical trajectory studies of the molecular dissociation dynamics of formaldehyde: H2CO→H2+CO,” The Journal of Chemical Physics, vol. 96, no. 6, pp. 4341–4355, 1992.
[75]  L. Deng, T. Ziegler, and L. Fan, “A combined density functional and intrinsic reaction coordinate study on the ground state energy surface of H2CO,” The Journal of Chemical Physics, vol. 99, no. 5, pp. 3823–3835, 1993.
[76]  W. Chen, W. L. Hase, and H. B. Schlegel, “Ab initio classical trajectory study of H2CO→H2+CO dissociation,” Chemical Physics Letters, vol. 228, no. 4-5, pp. 436–442, 1994.
[77]  G. H. Peslherbe and W. L. Hase, “Semiempirical MNDO, AM1, and PM3 direct dynamics trajectory studies of formaldehyde unimolecular dissociation,” Journal of Chemical Physics, vol. 104, no. 20, pp. 7882–7894, 1996.
[78]  H. Nakano, K. Nakayama, K. Hirao, and M. Dupuis, “Transition state barrier height for the reaction H2CO→H2+CO studied by multireference M?ller-Plesset perturbation theory,” Journal of Chemical Physics, vol. 106, no. 12, pp. 4912–4917, 1997.
[79]  F. Jensen, “Stationary points on the H2CO potential energy surface: dependence on theoretical level,” Theoretical Chemistry Accounts, vol. 99, no. 5, pp. 295–300, 1998.
[80]  D. Feller, M. Dupuis, and B. C. Garrett, “Barrier for the H2CO→H2+CO reaction: a discrepancy between high-level electronic structure calculations and experiment,” Journal of Chemical Physics, vol. 113, no. 1, pp. 218–226, 2000.
[81]  X. Li, J. M. Millam, and H. B. Schlegel, “Ab initio molecular dynamics studies of the photodissociation of formaldehyde, H2CO→H2 + CO: direct classical trajectory calculations by MP2 and density functional theory,” Journal of Chemical Physics, vol. 113, no. 22, pp. 10062–10067, 2000.
[82]  H. B. Schlegel, S. S. Iyengar, X. Li et al., “Ab initio molecular dynamics: propagating the density matrix with Gaussian orbitals. III. Comparison with Born-Oppenheimer dynamics,” Journal of Chemical Physics, vol. 117, no. 19, pp. 8694–8704, 2002.
[83]  T. Yonehara and S. Kato, “Role of isomerization channel in unimolecular dissociation reaction H2CO→H2+CO: Ab initio global potential energy surface and classical trajectory analysis,” Journal of Chemical Physics, vol. 117, no. 24, pp. 11131–11138, 2003.
[84]  X. Zhang, S. Zou, L. B. Harding, and J. M. Bowman, “A global ab initio potential energy surface for formaldehyde,” Journal of Physical Chemistry A, vol. 108, no. 41, pp. 8980–8986, 2004.
[85]  X. Zhang, J. L. Rheinecker, and J. M. Bowman, “Quasiclassical trajectory study of formaldehyde unimolecular dissociation: H2 CO→ H2 + CO, H + HCO,” Journal of Chemical Physics, vol. 122, no. 11, Article ID 114313, 8 pages, 2005.
[86]  T. Yonehara and S. Kato, “Quantum dynamics study on multichannel dissociation and isomerization reactions of formaldehyde,” Journal of Chemical Physics, vol. 125, no. 8, Article ID 084307, 2006.
[87]  J. M. Bowman and X. Zhang, “New insights on reaction dynamics from formaldehyde photodissociation,” Physical Chemistry Chemical Physics, vol. 8, no. 3, pp. 321–332, 2006.
[88]  L. B. Harding, S. J. Klippenstein, and A. W. Jasper, “Ab initio methods for reactive potential surfaces,” Physical Chemistry Chemical Physics, vol. 9, no. 31, pp. 4055–4070, 2007.
[89]  S. Maeda and K. Ohno, “A new global reaction route map on the potential energy surface of H2CO with unrestricted level,” Chemical Physics Letters, vol. 460, no. 1–3, pp. 55–58, 2008.
[90]  Y. Yamaguchi, S. S. Wesolowski, T. J. Van Huis, and H. F. Schaefer, “The unimolecular dissociation of H2CO on the lowest triplet potential-energy surface,” Journal of Chemical Physics, vol. 108, no. 13, pp. 5281–5288, 1998.
[91]  H. M. Yin, S. H. Kable, X. Zhang, and J. M. Bowman, “Signatures of H2CO photodissociation from two electronic states,” Science, vol. 311, no. 5766, pp. 1443–1446, 2006.
[92]  K. K. Baeck, “The analytic gradient for the equation-of-motion coupled-cluster energy with a reduced molecular orbital space: an application for the first excited state of formaldehyde,” Journal of Chemical Physics, vol. 112, no. 1, pp. 1–4, 2000.
[93]  J. B. Simonsen, N. Rusteika, M. S. Johnson, and T. I. S?lling, “Atmospheric photochemical loss of H and H2 from formaldehyde: the relevance of ultrafast processes,” Physical Chemistry Chemical Physics, vol. 10, no. 5, pp. 674–680, 2008.
[94]  R. D. Van Zee, M. F. Foltz, and C. B. Moore, “Evidence for a second molecular channel in the fragmentation of formaldehyde,” Journal of Chemical Physics, vol. 99, no. 3, pp. 1664–1673, 1993.
[95]  D. Townsend, S. A. Lahankar, S. K. Lee et al., “The roaming atom: straying from the reaction path in formaldehyde decomposition,” Science, vol. 306, no. 5699, pp. 1158–1161, 2004.
[96]  A. G. Suits, “Roaming atoms and radicals: a new mechanism in molecular dissociation,” Accounts of Chemical Research, vol. 41, no. 7, pp. 873–881, 2008.
[97]  N. Herath and A. G. Suits, “Roaming radical reactions,” Journal of Physical Chemistry Letters, vol. 2, no. 6, pp. 642–647, 2011.
[98]  J. M. Bowman and B. C. Shepler, “Roaming radicals,” Annual Review of Physical Chemistry, vol. 62, pp. 531–553, 2011.
[99]  P. L. Houston and S. H. Kable, “Photodissociation of acetaldehyde as a second example of the roaming mechanism,” Proceedings of the National Academy of Sciences of the United States of America, vol. 103, no. 44, pp. 16079–16082, 2006.
[100]  B. C. Shepler, B. J. Braams, and J. M. Bowman, “Quasiclassical trajectory calculations of acetaldehyde dissociation on a global potential energy surface indicate significant non-transition state dynamics,” Journal of Physical Chemistry A, vol. 111, no. 34, pp. 8282–8285, 2007.
[101]  B. R. Heazlewood, M. J. T. Jordan, S. H. Kable et al., “Roaming is the dominant mechanism for molecular products in acetaldehyde photodissociation,” Proceedings of the National Academy of Sciences of the United States of America, vol. 105, no. 35, pp. 12719–12724, 2008.
[102]  V. Goncharov, N. Herath, and A. G. Suits, “Roaming dynamics in acetone dissociation,” Journal of Physical Chemistry A, vol. 112, no. 39, pp. 9423–9428, 2008.
[103]  C. Chen, B. Braams, D. Y. Lee, J. M. Bowman, P. L. Houston, and D. Stranges, “Evidence for vinylidene production in the photodissociation of the allyl radical,” Journal of Physical Chemistry Letters, vol. 1, no. 12, pp. 1875–1880, 2010.
[104]  E. Kamarchik, L. Koziol, H. Reisler, J. M. Bowman, and A. I. Krylov, “Roaming pathway leading to unexpected water + vinyl products in C2H4OH dissociation,” Journal of Physical Chemistry Letters, vol. 1, no. 20, pp. 3058–3065, 2010.
[105]  M. P. Grubb, M. L. Warter, A. G. Suits, and S. W. North, “Evidence of roaming dynamics and multiple channels for molecular elimination in NO3 photolysis,” The Journal of Physical Chemistry Letters, vol. 1, pp. 2455–2458, 2010.
[106]  M. P. Grubb, M. L. Warter, K. M. Johnson, and S. W. North, “Ion imaging study of NO3 radical photodissociation dynamics: characterization of multiple reaction pathways,” Journal of Physical Chemistry, vol. 115, no. 15, pp. 3218–3226, 2011.
[107]  H. Xiao, S. Maeda, and K. Morokuma, “Excited-state roaming dynamics in photolysis of a nitrate radical,” Journal of Physical Chemistry Letters, vol. 2, no. 9, pp. 934–938, 2011.
[108]  S. W. North, “Roaming in the dark,” Nature Chemistry, vol. 3, no. 7, pp. 504–505, 2011.
[109]  A. Fernández-Ramos, J. A. Miller, S. J. Klippenstein, and D. G. Truhlar, “Modeling the kinetics of bimolecular reactions,” Chemical Reviews, vol. 106, no. 11, pp. 4518–4584, 2006.
[110]  L. B. Harding and S. J. Klippenstein, “Roaming radical pathways for the decomposition of alkanes,” Journal of Physical Chemistry Letters, vol. 1, no. 20, pp. 3016–3020, 2010.
[111]  P. Zhang, Applications of electronic structure theory in the study of molecular processes, Ph.D. thesis, Emory University, 2005.
[112]  B. C. Shepler, E. Epifanovsky, P. Zhang, J. M. Bowman, A. I. Krylov, and K. Morokuma, “Photodissociation dynamics of formaldehyde initiated at the T1/S0 minimum energy crossing configurations,” Journal of Physical Chemistry A, vol. 112, no. 51, pp. 13267–13270, 2008.
[113]  M. Araujo, B. Lasorne, M. J. Bearpark, and M. A. Robb, “The photochemistry of formaldehyde: internal conversion and molecular dissociation in a single step?” Journal of Physical Chemistry A, vol. 112, no. 33, pp. 7489–7491, 2008.
[114]  P. Zhang, S. Maeda, K. Morokuma, and B. J. Braams, “Photochemical reactions of the low-lying excited states of formaldehyde: T1/S0 intersystem crossings, characteristics of the S1 and T1 potential energy surfaces, and a global T1 potential energy surface,” Journal of Chemical Physics, vol. 130, no. 11, Article ID 114304, 2009.
[115]  M. Araújo, B. Lasorne, A. L. Magalh?es, G. A. Worth, M. J. Bearpark, and M. A. Robb, “The molecular dissociation of formaldehyde at medium photoexcitation energies: a quantum chemistry and direct quantum dynamics study,” Journal of Chemical Physics, vol. 131, no. 14, Article ID 144301, 2009.
[116]  M. Arauéjo, B. Lasorne, A. L. Magalhaes, M. J. Bearpark, and M. A. Robb, “Controlling product selection in the photodissociation of formaldehyde: direct quantum dynamics from the S1 barrier,” Journal of Physical Chemistry A, vol. 114, no. 45, pp. 12016–12020, 2010.
[117]  B. Fu, B. C. Shepler, and J. M. Bowman, “hree-state trajectory surface hopping studies of the photodissociation dynamics of formaldehyde on ab initio potential energy surfaces,” Journal of the American Chemical Society, vol. 133, no. 20, pp. 7957–7968, 2011.
[118]  Y. Haas, “Photochemical α-cleavage of ketones: revisiting acetone,” Photochemical and Photobiological Sciences, vol. 3, no. 1, pp. 6–16, 2004.
[119]  R. A. Copeland and D. R. Crosley, “Radiative, collisional and dissociative processes in triplet acetone,” Chemical Physics Letters, vol. 115, no. 4-5, pp. 362–368, 1985.
[120]  H. Zuckermann, B. Schmitz, and Y. Haas, “Dissociation energy of an isolated triplet acetone molecule,” Journal of Physical Chemistry, vol. 92, no. 17, pp. 4835–4837, 1988.
[121]  S. W. North, D. A. Blank, J. Daniel Gezelter, C. A. Longfellow, and Y. T. Lee, “Evidence for stepwise dissociation dynamics in acetone at 248 and 193?nm,” The Journal of Chemical Physics, vol. 102, no. 11, pp. 4447–4460, 1995.
[122]  D. Liu, W. H. Fang, and X. Y. Fu, “An ab initio study on photodissociation of acetone,” Chemical Physics Letters, vol. 325, no. 1–3, pp. 86–92, 2000.
[123]  E. W. G. Diau, G. K?tting, and A. H. Zewail, “Femtochemistry of norrish type-1 reactions. I. Experimental and theoretical studies of acetone and related ketones on the s1 surface,” ChemPhysChem, vol. 2, no. 5, pp. 273–293, 2001.
[124]  E. W. G. Diau, C. K?tting, T. I. S?lling, and A. H. Zewail, “Femtochemistry of Norrish type-I reactions. III. Highly excited ketones—theoretical,” ChemPhysChem, vol. 3, no. 1, pp. 57–78, 2002.
[125]  T. I. S?lling, E. W. G. Diau, C. K?tting, S. de Feyter, and A. H. Zewail, “Femtochemistry of Norrish type-I reactions. IV. Highly excited ketones—experimental,” ChemPhysChem, vol. 3, no. 1, pp. 79–97, 2002.
[126]  W. K. Chen, J. W. Ho, and P. Y. Cheng, “Ultrafast photodissociation dynamics of the acetone 3s Rydberg state at 195?nm: a new mechanism,” Chemical Physics Letters, vol. 380, no. 3-4, pp. 411–418, 2003.
[127]  W. K. Chen, J. W. Ho, and P. Y. Cheng, “Ultrafast photodissociation dynamics of acetone at 195?nm. I. Initial-state, intermediate, and product temporal evolutions by femtosecond mass-selected multiphoton ionization spectroscopy,” Journal of Physical Chemistry A, vol. 109, no. 31, pp. 6805–6817, 2005.
[128]  W. K. Chen and P. Y. Cheng, “Ultrafast photodissociation dynamics of acetone at 195?nm: II. Unraveling complex three-body dissociation dynamics by femtosecond time-resolved photofragment translational spectroscopy,” Journal of Physical Chemistry A, vol. 109, no. 31, pp. 6818–6829, 2005.
[129]  H. Somnitz, M. Fida, T. Ufer, and R. Zellner, “Pressure dependence for the CO quantum yield in the photolysis of acetone at 248?nm: a combined experimental and theoretical study,” Physical Chemistry Chemical Physics, vol. 7, no. 18, pp. 3342–3352, 2005.
[130]  V. Khamaganov, R. Karunanandan, A. Rodriguez, and J. N. Crowley, “Photolysis of CH3C(O)CH3 (248?nm, 266?nm), CH3C(O)C2H5 (248?nm) and CH3C(O)Br (248?nm): pressure dependent quantum yields of CH3 formation,” Physical Chemistry Chemical Physics, vol. 9, no. 31, pp. 4098–4113, 2007.
[131]  R. Nádasdi, G. Kovács, I. Szilágyi et al., “Exciplex laser photolysis study of acetone with relevance to tropospheric chemistry,” Chemical Physics Letters, vol. 440, no. 1–3, pp. 31–35, 2007.
[132]  I. Antol, M. Eckert-Maksi?, M. On?ák, P. Slaví?ek, and H. Lischka, “Photodissociation pathways of acetone upon excitation into the 3s Rydberg state: adiabatic versus diabatic mechanism,” Collection of Czechoslovak Chemical Communications, vol. 73, no. 11, pp. 1475–1494, 2008.
[133]  H. Somnitz, T. Ufer, and R. Zellner, “Acetone photolysis at 248?nm revisited: pressure dependence of the CO and CO2 quantum yields,” Physical Chemistry Chemical Physics, vol. 11, no. 38, pp. 8522–8531, 2009.
[134]  T. Shibata and T. Suzuki, “Photofragment ion imaging with femtosecond laser pulses,” Chemical Physics Letters, vol. 262, no. 1-2, pp. 115–119, 1996.
[135]  N. Rusteika, K. B. M?ller, and T. I. S?lling, “New insights on the photodynamics of acetone excited with 253–288?nm femtosecond pulses,” Chemical Physics Letters, vol. 461, no. 4–6, pp. 193–197, 2008.
[136]  R. Y. Brogaard, T. I. S?lling, and K. B. M?ller, “Initial dynamics of the Norrish Type I reaction in acetone: probing wave packet motion,” Journal of Physical Chemistry A, vol. 115, no. 15, pp. 556–561, 2011.
[137]  M. J. Frisch, G. W. Trucks, H. B. Schlegel, et al., “Gaussian 09 Revision A.2,” Wallingford CT, 2009.
[138]  R. Nádasdi, G. L. Zügner, M. Farkas, S. D?bé, S. Maeda, and K. Morokuma, “Photochemistry of methyl ethyl ketone: quantum yields and S1/S0-diradical mechanism of photodissociation,” ChemPhysChem, vol. 11, no. 18, pp. 3883–3895, 2010.
[139]  B. Lasorne, G. A. Worth, and M. A. Robb, “Excited-state dynamics,” Wiley Interdisciplinary Reviews: Computational Molecular Science, vol. 1, no. 3, pp. 460–475, 2011.
[140]  J. K?stner, “Umbrella sampling,” Wiley Interdisciplinary Reviews: Computational Molecular Science, vol. 1, no. 6, pp. 932–942, 2011.
[141]  H. Hu and W. Yang, “Free energies of chemical reactions in solution and in enzymes with ab initio quantum mechanics/molecular mechanics methods,” Annual Review of Physical Chemistry, vol. 59, pp. 573–601, 2008.
[142]  A. Barducci, M. Bonomi, and M. Parrinello, “Metadynamics,” Wiley Interdisciplinary Reviews: Computational Molecular Science, vol. 1, no. 5, pp. 826–843, 2011.
[143]  G. C. Schatz, “The analytical representation of electronic potential-energy surfaces,” Reviews of Modern Physics, vol. 61, no. 3, pp. 669–688, 1989.
[144]  M. A. Collins, “Molecular potential-energy surfaces for chemical reaction dynamics,” Theoretical Chemistry Accounts, vol. 108, no. 6, pp. 313–324, 2002.
[145]  J. M. Bowman, B. J. Braams, S. Carter et al., “Ab-initio-based potential energy surfaces for complex molecules and molecular complexes,” Journal of Physical Chemistry Letters, vol. 1, no. 12, pp. 1866–1874, 2010.
[146]  J. M. Bowman, G. Czakó, and B. Fu, “High-dimensional ab initio potential energy surfaces for reaction dynamics calculations,” Physical Chemistry Chemical Physics, vol. 13, no. 18, pp. 8094–8111, 2011.
[147]  C. R. Evenhuis, X. Lin, D. H. Zhang, D. Yarkony, and M. A. Collins, “Interpolation of diabatic potential-energy surfaces: quantum dynamics on ab initio surfaces,” Journal of Chemical Physics, vol. 123, no. 13, Article ID 134110, 12 pages, 2005.
[148]  Z. H. Li, R. Valero, and D. G. Truhlar, “Improved direct diabatization and coupled potential energy surfaces for the photodissociation of ammonia,” Theoretical Chemistry Accounts, vol. 118, no. 1, pp. 9–24, 2007.
[149]  Z. Wang, I. S. K. Kerkines, K. Morokuma, and P. Zhang, “Analytical potential energy surfaces for N3 low-lying doublet states,” Journal of Chemical Physics, vol. 130, no. 4, Article ID 044313, 2009.
[150]  X. Zhu and D. R. Yarkony, “On the representation of coupled adiabatic potential energy surfaces using quasi-diabatic Hamiltonians: description of accidental seams of conical intersection,” Molecular Physics, vol. 108, no. 19-20, pp. 2611–2619, 2010.
[151]  A. Warshel and M. Levitt, “Theoretical studies of enzymatic reactions: dielectric, electrostatic and steric stabilization of the carbonium ion in the reaction of lysozyme,” Journal of Molecular Biology, vol. 103, no. 2, pp. 227–249, 1976.
[152]  U. C. Singh and P. A. Kollman, “A combined ab initio quantum mechanical and molecular mechanical method for carrying out simulations on complex molecular systems: applications to the CH3Cl + Cl? exchange reaction and gas phase protonation of polyethers,” Journal of Computational Chemistry, vol. 7, no. 6, pp. 718–730, 1986.
[153]  M. J. Field, P. A. Bash, and M. Karplus, “A combined quantum mechanical and molecular mechanical potential for molecular dynamics simulations,” Journal of Computational Chemistry, vol. 11, no. 6, pp. 700–733, 1990.
[154]  F. Maseras and K. Morokuma, “IMOMM—a new integrated ab-initio plus molecular mechanics geometry optimization scheme of equilibrium structures and transition-states,” Journal of Computational Chemistry, vol. 16, no. 9, pp. 1170–1179, 1995.
[155]  M. Svensson, S. Humbel, R. D. J. Froese, T. Matsubara, S. Sieber, and K. Morokuma, “ONIOM: a multilayered integrated MO + MM method for geometry optimizations and single point energy predictions. A test for Diels-Alder reactions and Pt(P(t-Bu)3)2 + H2 oxidative addition,” Journal of Physical Chemistry, vol. 100, no. 50, pp. 19357–19363, 1996.
[156]  T. Vreven, K. Morokuma, ? Farkas, H. B. Schlegel, and M. J. Frisch, “Geometry optimization with QM/MM, ONIOM, and other combined methods. I. Microiterations and constraints,” Journal of Computational Chemistry, vol. 24, no. 6, pp. 760–769, 2003.
[157]  T. Vreven, M. J. Frisch, K. N. Kudin, H. B. Schlegel, and K. Morokuma, “Geometry optimization with QM/MM methods II: explicit quadratic coupling,” Molecular Physics, vol. 104, pp. 701–714, 2006.
[158]  S. Maeda, K. Ohno, and K. Morokuma, “An automated and systematic transition structure explorer in large flexible molecular systems based on combined global reaction route mapping and microiteration methods,” Journal of Chemical Theory and Computation, vol. 5, no. 10, pp. 2734–2743, 2009.

Full-Text

comments powered by Disqus

Contact Us

service@oalib.com

QQ:3279437679

WhatsApp +8615387084133