Skip to main content
Log in

Second-order Saddle Dynamics in Isomerization Reaction

  • Published:
Regular and Chaotic Dynamics Aims and scope Submit manuscript

Abstract

The role of second-order saddle in the isomerization dynamics was investigated by considering the \(E-Z\) isomerization of guanidine. The potential energy profile for the reaction was mapped using the ab initio wavefunction method. The isomerization path involved a torsional motion about the imine (C-N) bond in a clockwise or an anticlockwise fashion resulting in two degenerate transition states corresponding to a barrier of 23.67 kcal/mol. An alternative energetically favorable path (\(\sim\)1 kcal/mol higher than the transition states) by an in-plane motion of the imine (N-H) bond via a second-order saddle point on the potential energy surface was also obtained. The dynamics of the isomerization was investigated by ab initio classical trajectory simulations. The trajectories reveal that isomerization happens via the transition states as well as the second-order saddle. The dynamics was found to be nonstatistical with trajectories exhibiting recrossing and the higher energy second-order saddle pathway preferred over the traditional transition state pathway. Wavelet based time-frequency analysis of internal coordinates indicate regular dynamics and existence of long-lived quasi-periodic trajectories in the phase space.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

References

  1. Steinfeld, J. I., Francisco, J. S., and Hase, W. L., Chemical Kinetics and Dynamics, Upper Saddle River, N.J.: Prentice Hall, 1999.

    Google Scholar 

  2. Levine, R. D., Molecular Reaction Dynamics, Cambridge: Cambridge Univ. Press, 2005.

    Book  Google Scholar 

  3. Fukui, K., The Path of Chemical Reactions: The IRC Approach, Acc. Chem. Res., 1981, vol. 14, no. 12, pp. 363–368.

    Article  Google Scholar 

  4. Eyring, H., The Activated Complex in Chemical Reactions, J. Chem. Phys., 1935, vol. 3, no. 2, pp. 107–115.

    Article  Google Scholar 

  5. Evans, M. G. and Polanyi, M., Some Applications of the Transition State Method to the Calculation of Reaction Velocities, Especially in Solution, Trans. Faraday Soc., 1935, vol. 31, pp. 875–894.

    Article  Google Scholar 

  6. Pechukas, P., Transition State Theory, Annu. Rev. Phys. Chem., 1981, vol. 32, no. 1, pp. 159–177.

    Article  Google Scholar 

  7. Heidrich, D. and Quapp, W., Saddle Points of Index \(2\) on Potential Energy Surfaces and Their Role in Theoretical Reactivity Investigations, Theor. Chem. Acc., 1986, vol. 70, no. 2, pp. 89–98.

    Article  Google Scholar 

  8. Breulet, J. and Schaefer, H. F. III, Conrotatory and Disrotatory Stationary Points for the Electrocyclic Isomerization of Cyclobutene to cis-Butadiene, J. Am. Chem. Soc., 1984, vol. 106, no. 5, pp. 1221–1226.

    Article  Google Scholar 

  9. Chen, J. S., Houk, K., and Foote, C. S., Theoretical Study of the Concerted and Stepwise Mechanisms of Triazolinedione Diels – Alder Reactions, J. Am. Chem. Soc., 1998, vol. 120, no. 47, pp. 12303–12309.

    Article  Google Scholar 

  10. Minyaev, R. M., Getmanskii, I. V., and Quapp, W., A Second-Order Saddle Point in the Reaction Coordinate for the Isomerization of the NH\({}_{5}\) Complex: Ab initio Calculations, Russ. J. Phys. Chem., 2004, vol. 78, no. 9, pp. 1494–1498; see also: Zh. Fiz. Khim., 2004, vol. 78, no. 9, pp. 1700-1705.

    Google Scholar 

  11. Quapp, W. and Bofill, J. M., Embedding of the Saddle Point of Index Two on the PES of the Ring Opening of Cyclobutene, Int. J. Quantum Chem., 2015, vol. 115, no. 23, pp. 1635–1649.

    Article  Google Scholar 

  12. Harabuchi, Y., Ono, Y., Maeda, S., Taketsugu, T., Keipert, K., and Gordon, M. S., Nontotally Symmetric Trifurcation of an SN\({}_{2}\) Reaction Pathway, J. Comput. Chem., 2016, vol. 37, no. 5, pp. 487–493.

    Article  Google Scholar 

  13. Ezra, G. S. and Wiggins, S., Phase-Space Geometry and Reaction Dynamics near Index \(2\) Saddles, J. Phys. A, 2009, vol. 42, no. 20, 205101, 25 pp.

    Article  MathSciNet  Google Scholar 

  14. Collins, P., Ezra, G. S., and Wiggins, S., Index \(k\) Saddles and Dividing Surfaces in Phase Space with Applications to Isomerization Dynamics, J. Chem. Phys., 2011, vol. 134, no. 24, 244105, 19 pp.

    Article  Google Scholar 

  15. Mauguière, F., Collins, P., Ezra, G., and Wiggins, S., Bond Breaking in a Morse Chain under Tension: Fragmentation Patterns, Higher Index Saddles, and Bond Healing, J. Chem. Phys., 2013, vol. 138, no. 13, 134118, 17 pp.

    Article  Google Scholar 

  16. Litman, Y., Richardson, J. O., Kumagai, T., and Rossi, M., Elucidating the Nuclear Quantum Dynamics of Intramolecular Double Hydrogen Transfer in Porphycene, J. Am. Chem. Soc., 2019, vol. 141, no. 6, pp. 2526–2534.

    Article  Google Scholar 

  17. Yoshikawa, T., Sugawara, S., Takayanagi, T., Shiga, M., and Tachikawa, M., Quantum Tautomerization in Porphycene and Its Isotopomers: Path-Integral Molecular Dynamics Simulations, Chem. Phys., 2019, vol. 394, no. 1, pp. 46–51.

    Article  Google Scholar 

  18. Accardi, A., Barth, I., Kühn, O., and Manz, J., From Synchronous to Sequential Double Proton Transfer: Quantum Dynamics Simulations for the Model Porphine, J. Phys. Chem. A, 2010, vol. 114, no. 42, pp. 11252–11262.

    Article  Google Scholar 

  19. García-Garrido, V. J., Agaoglou, M., and Wiggins, S., Exploring Isomerization Dynamics on a Potential Energy Surface with an Index-2 Saddle Using Lagrangian Descriptors, Commun. Nonlinear Sci. Numer. Simul., 2020, vol. 89, 105331, 29 pp.

    Article  MathSciNet  Google Scholar 

  20. Haller, G., Uzer, T., Palacián, J., Yanguas, P., and Jaffé, Ch., Transition State Geometry near Higher-Rank Saddles in Phase Space, Nonlinearity, 2011, vol. 24, no. 2, pp. 527–561.

    Article  MathSciNet  Google Scholar 

  21. Nagahata, Y., Teramoto, H., Li, Ch.-B., Kawai, Sh., and Komatsuzaki, T., Reactivity Boundaries for Chemical Reactions Associated with Higher-Index and Multiple Saddles, Phys. Rev. E, 2013, vol. 88, no. 4, 042923, 11 pp.

    Article  Google Scholar 

  22. Harding, L. B., Klippenstein, S. J., and Jasper, A. W., Separability of Tight and Roaming Pathways to Molecular Decomposition, J. Phys. Chem. A, 2012, vol. 116, no. 26, pp. 6967–6982.

    Article  Google Scholar 

  23. Pradhan, R. and Lourderaj, U., Can Reactions Follow Non-Traditional Second-Order Saddle Pathways Avoiding Transition States?, Phys. Chem. Chem. Phys., 2019, vol. 21, no. 24, pp. 12837–12842.

    Article  Google Scholar 

  24. Coyle, S. and Glaser, R., Asymmetric Imine N-Inversion in \(3\)-Methyl-\(4\)-Pyrimidinimine. Molecular Dipole Analysis of Solvation Effects, J. Org. Chem., 2011, vol. 76, no. 10, pp. 3987–3996.

    Article  Google Scholar 

  25. Frisch, M. J., Trucks, G. W., Schlegel, H. B., Scuseria, G. E., Robb, M. A., Cheeseman, J. R., Scalmani, G., Barone, V., Petersson, G. A., Nakatsuji, H., Li, X., Caricato, M., Marenich, A. V., Bloino, J., Janesko, B. G., Gomperts, R., Mennucci, B., Hratchian, H. P., Ortiz, J. V., Izmaylov, A. F., Sonnenberg, J. L., Williams-Young, D., Ding, F., Lipparini, F., Egidi, F., Goings, J., Peng, B., Petrone, A., Henderson, T., Ranasinghe, D., Zakrzewski, V. G., Gao, J., Rega, N., Zheng, G., Liang, W., Hada, M., Ehara, M., Toyota, K., Fukuda, R., Hasegawa, J., Ishida, M., Nakajima, T., Honda, Y., Kitao, O., Nakai, H., Vreven, T., Throssell, K., Montgomery, J. A., Jr., Peralta, J. E., Ogliaro, F., Bearpark, M. J., Heyd, J. J., Brothers, E. N., Kudin, K. N., Staroverov, V. N., Keith, T. A., Kobayashi, R., Normand, J., Raghavachari, K., Rendell, A. P., Burant, J. C., Iyengar, S. S., Tomasi, J., Cossi, M., Millam, J. M., Klene, M., Adamo, C., Cammi, R., Ochterski, J. W., Martin, R. L., Morokuma, K., Farkas, O., Foresman, J. B., and Fox, D. J., Gaussian 16 Revision C.01., 0 ().

  26. Valiev, M., Bylaska, E. J., Govind, N., Kowalski, K., Straatsma, T. P., Van Dam, H. J. J., Wang, D., Nieplocha, J., Apra, E., Windus, T. L., and de Jong, W. A., NWChem: A Comprehensive and Scalable Open-Source Solution for Large Scale Molecular Simulations, Comput. Phys. Commun., 2010, vol. 181, no. 9, pp. 1477–1489.

    Article  Google Scholar 

  27. Lourderaj, U., Sun, R., Kohale, S. C., Barnes, G. L., de Jong, W. A., Windus, T. L., and Hase, W. L., The VENUS/NWChem Software Package: Tight Coupling between Chemical Dynamics Simulations and Electronic Structure Theory, Comput. Phys. Commun., 2014, vol. 185, no. 3, pp. 1074–1080.

    Article  Google Scholar 

  28. Sun, L. and Hase, W. L., Born – Oppenheimer Direct Dynamics Classical Trajectory Simulations, in Reviews in Computational Chemistry, K. B. Lipkowitz, R. Larter, Th. R. Cundari (Eds.), New York: Wiley, 2003, pp. 79–146..

    Chapter  Google Scholar 

  29. Hare, S. R., Bratholm, L. A., Glowacki, D. R., and Carpenter, B. K., Low Dimensional Representations along Intrinsic Reaction Coordinates and Molecular Dynamics Trajectories Using Interatomic Distance Matrices, Chem. Sci., 2019, vol. 10, no. 43, pp. 9954–9968.

    Article  Google Scholar 

  30. Tsutsumi, T., Ono, Y., Arai, Z., and Taketsugu, T., Visualization of Dynamics Effect: Projection of On-the-Fly Trajectories to the Subspace Spanned by the Static Reaction Path Network, J. Chem. Theory Comput., 2020, vol. 16, no. 7, pp. 4029–4037.

    Article  Google Scholar 

  31. Zhu, L. and Hase, W. L., A General RRKM Program, Program No. QCPE 644 (1993).

  32. Ezra, G. S., Waalkens, H., and Wiggins, S., Microcanonical Rates, Gap Times, and Phase Space Dividing Surfaces, J. Chem. Phys., 2009, vol. 130, no. 16, 164118, 15 pp.

    Article  Google Scholar 

  33. Manikandan, P. and Keshavamurthy, S., Dynamical Traps Lead to the Slowing Down of Intramolecular Vibrational Energy Flow, Proc. Natl. Acad. Sci. USA,, 2014, vol. 111, no. 40, pp. 14354–14359.

    Article  Google Scholar 

  34. Vela-Arevalo, L. B. and Wiggins, S., Time-frequency Analysis of Classical Trajectories of Polyatomic Molecules, Internat. J. Bifur. Chaos Appl. Sci. Engrg., 2001, vol. 11, no. 5, pp. 1359–1380.

    Article  Google Scholar 

  35. Chandre, C., Wiggins, S., and Uzer, T., Time-Frequency Analysis of Chaotic Systems, Phys. D, 2003, vol. 181, no. 3–4, pp. 171–196.

    MathSciNet  Google Scholar 

  36. Guillery, N. and Meiss, J. D., Diffusion and Drift in Volume-Preserving Maps, Regul. Chaotic Dyn., 2017, vol. 22, no. 6, pp. 700–720.

    Article  MathSciNet  Google Scholar 

  37. Arnol’d, V. I., Small Denominators and Problems of Stability of Motion in Classical and Celestial Mechanics, Russian Math. Surveys, 1963, vol. 18, no. 6, pp. 85–191; see also: Uspekhi Mat. Nauk, 1963, vol. 18, no. 6(114), pp. 91-192.

    Article  MathSciNet  Google Scholar 

  38. Karmakar, S., Yadav, P. K., and Keshavamurthy, S., Stable Chaos and Delayed Onset of Statisticality in Unimolecular Dissociation Reactions, Commun. Chem., 2020, vol. 3, no. 4, pp. 11.

    Google Scholar 

  39. Lourderaj, U. and Hase, W. L., Theoretical and Computational Studies of Non-RRKM Unimolecular Dynamics, J. Phys. Chem. A, 2009, vol. 113, no. 11, pp. 2236–2253.

    Article  Google Scholar 

  40. Jayee, B. and Hase, W. L., Nonstatistical Reaction Dynamics, Annu. Rev. Phys. Chem., 2020, vol. 71, pp. 289–313.

    Article  Google Scholar 

  41. Wiggins, S., The Role of Normally Hyperbolic Invariant Manifolds (NHIMs) in the Context of the Phase Space Setting for Chemical Reaction Dynamics, Regul. Chaotic Dyn., 2016, vol. 21, no. 6, pp. 621–638.

    Article  MathSciNet  Google Scholar 

  42. Manikandan, P., Semparithi, A., and Keshavamurthy, S., Decoding the Dynamical Information Embedded in Highly Excited Vibrational Eigenstates: State Space and Phase Space Viewpoints, J. Phys. Chem. A, 2009, vol. 113, no. 9, pp. 1717–1730.

    Article  Google Scholar 

Download references

ACKNOWLEDGMENTS

The authors thank NISER Bhubaneswar for the computational facility and Mr. Rupayan Biswas for the help with the figures.

Funding

This work was supported by the grant of the Science and Engineering Board (SERB), India (No. EMR/2017/004843).

Author information

Authors and Affiliations

Authors

Corresponding authors

Correspondence to Richa Rashmi, Komal Yadav, Upakarasamy Lourderaj or Manikandan Paranjothy.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

MSC2010

92E99

SUPPLEMENTARY INFORMATION

APPENDIX

APPENDIX

The sampling procedure for trajectories initiated at the reactant and transition state regions are discussed elsewhere [28]. Here, the sampling procedure for trajectories initiated at the SOS region from a microcanonical ensemble is presented. The procedure involved the following steps:

  1. 1)

    The classical harmonic vibrational frequencies, \(\omega_{i}=2\pi\nu_{i}\) (\(i=1,\ldots,3N-6\)) and their corresponding eigenvectors (normal modes, \(Q_{i}\)) are first obtained by diagonalizing the mass-weighted force constant (\(\mathbf{F}\)) matrix given by \(\mathbf{\Lambda}=\mathbf{L^{-1}FL}\). At the SOS there are two imaginary frequencies corresponding to two normal mode vectors \(Q_{1}\) and \(Q_{2}\).

  2. 2)

    The energy \(E\) available to the SOS is distributed randomly among the \(m=3N-8\) normal modes with positive frequencies such that the \(i^{\text{th}}\) mode has an energy

    $$E_{i}=\left(E-\sum_{j=1}^{i-1}\right)(1-R_{i}^{1/(m-i)}),$$
    (A.1)

    where \(R_{i}\) is a random number between 0 and 1.

  3. 3)

    The remaining energy \(E_{2}=E-\sum_{j=1}^{m}E_{j}\) is then calculated and part of it is assigned randomly to the mode \(Q_{1}\) and the rest of the energy is then assigned to the mode \(Q_{2}\).

  4. 4)

    Once the energy for each normal mode is assigned, the normal mode coordinates and their conjugate momenta are chosen randomly using a random phase for each mode given by

    $$Q_{i}=\left(\frac{\sqrt{2E_{i}}}{\omega_{i}}\right)\cos(2\pi R_{i})\text{ and }P_{i}=-(\sqrt{2E_{i}})\sin(2\pi R_{i}),$$
    (A.2)

    where \(R_{i}\) is a new random number between 0 and 1 different from the one given in Eq. (A.1).

  5. 5)

    The normal mode coordinates and momenta \(\mathbf{Q}\) and \(\mathbf{P}\) are then transformed to Cartesian coordinates \(\mathbf{q}\) and momenta \(\mathbf{p}\) according to

    $$\mathbf{q}=\mathbf{q_{0}}+\mathbf{m}^{-1/2}\mathbf{L}\mathbf{Q}\text{ and }\mathbf{p}=\mathbf{m}^{-1/2}\mathbf{L}\mathbf{P},$$
    (A.3)

    where \(\mathbf{m}\) is a diagonal matrix with atomic masses as its elements and the vector \(\mathbf{q_{0}}\) corresponds to the equilibrium (SOS) coordinates of the molecule.

  6. 6)

    Any spurious angular momentum that arises from the linear transformation is subtracted. Then the energy \(E_{\text{chosen}}\) for the chosen Cartesian coordinates and momenta is calculated. If \(E_{\text{chosen}}\) and the desired energy \(E\) do not agree, the Cartesian coordinates and momenta are rescaled by a factor of \((E/E_{\text{chosen}})\) and any spurious center-of-mass translation energy arising from the rescaling is then subtracted. This step is repeated until the agreement between \(E_{\text{chosen}}\) and \(E\) is within 0.1 %.

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Rashmi, R., Yadav, K., Lourderaj, U. et al. Second-order Saddle Dynamics in Isomerization Reaction. Regul. Chaot. Dyn. 26, 119–130 (2021). https://doi.org/10.1134/S1560354721020027

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1560354721020027

Keywords

Navigation