Skip to main content
Log in

The Role of Depth and Flatness of a Potential Energy Surface in Chemical Reaction Dynamics

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

This article has been updated

Abstract

In this study, we analyze how changes in the geometry of a potential energy surface in terms of depth and flatness can affect the reaction dynamics. We formulate depth and flatness in the context of one- and two-degree-of-freedom (DOF) Hamiltonian normal form for the saddle-node bifurcation and quantify their influence on chemical reaction dynamics [1, 2]. In a recent work, García-Garrido et al. [2] illustrated how changing the well-depth of a potential energy surface (PES) can lead to a saddle-node bifurcation. They have shown how the geometry of cylindrical manifolds associated with the rank-1 saddle changes en route to the saddle-node bifurcation. Using the formulation presented here, we show how changes in the parameters of the potential energy control the depth and flatness and show their role in the quantitative measures of a chemical reaction. We quantify this role of the depth and flatness by calculating the ratio of the bottleneck width and well width, reaction probability (also known as transition fraction or population fraction), gap time (or first passage time) distribution, and directional flux through the dividing surface (DS) for small to high values of total energy. The results obtained for these quantitative measures are in agreement with the qualitative understanding of the reaction dynamics.

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
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13
Fig. 14
Fig. 15
Fig. 16

Similar content being viewed by others

Change history

  • 22 January 2021

    Changes in the attributes of the ImageObject (img) tag in the HTML file

References

  1. Borondo, F., Zembekov, A. A., and Benito, R. M., Saddle-Node Bifurcations in the LiNC/LiCN Molecular System: Classical Aspects and Quantum Manifestations, J. Chem. Phys., 1996, vol. 105, no. 12, pp. 5068–5081.

    Article  Google Scholar 

  2. García-Garrido, V. J., Naik, Sh. and Wiggins, S., Tilting and Squeezing: Phase Space Geometry of Hamiltonian Saddle-Node Bifurcation and Its Influence on Chemical Reaction Dynamics, Internat. J. Bifur. Chaos Appl. Sci. Engrg., 2020, vol. 30, no. 4, 2030008, 35 pp.

    Article  MathSciNet  Google Scholar 

  3. Wales, D., Energy Landscapes: Applications to Clusters, Biomolecules and Glasses, Cambridge: Cambridge Univ. Press, 2004.

    Book  Google Scholar 

  4. Steinfeld, J. I., Francisco, J. S., and Hase, W. L., Chemical Kinetics and Dynamics, Englewood Cliffs, N.J.: Prentice Hall, 1989.

    Google Scholar 

  5. Levine, R. D., Molecular Reaction Dynamics, Cambridge: Cambridge Univ. Press, 2009.

    Google Scholar 

  6. Agaoglou, M., Aguilar-Sanjuan, B., García-Garrido, V. J., García-Meseguer, R., González-Montoya, F., Katsanikas, M., Krajňák, V., Naik, Sh., and Wiggins, S., Chemical Reactions: A Journey into Phase Space, https://www.chemicalreactions.io (2019).

  7. Preuss, R., Buenker, R. J., and Peyerimhoff, S. D., Theoretical Study of the Electronically Excited States of the HNSi Molecule, Chem. Phys. Lett., 1979, vol. 62, no. 1, pp. 21–25.

    Article  Google Scholar 

  8. Bittererová, M. and Biskupič, S., Ab initio Calculation of Stationary Points on the \(\mathrm{HF}_{2}\) Potential Energy Surface, Chem. Phys. Lett., 1999, vol. 299, no. 2, pp. 145–150.

    Article  Google Scholar 

  9. Koseki, Sh. and Gordon, M. S., Intrinsic Reaction Coordinate Calculations for Very Flat Potential Energy Surfaces: Application to Singlet Disilenylidene Isomerization, J. Phys. Chem., 1989, vol. 93, no. 1, pp. 118–125.

    Article  Google Scholar 

  10. Nummela, J. A. and Carpenter, B. K., Nonstatistical Dynamics in Deep Potential Wells: A Quasiclassical Trajectory Study of Methyl Loss from the Acetone Radical Cation, J. Am. Chem. Soc., 2002, vol. 124, no. 29, pp. 8512–8513.

    Article  Google Scholar 

  11. Bowman, J. M. and Suits, A. G., Roaming Reactions: The Third Way, Phys. Today, 2011, vol. 64, no. 11, pp. 33–37.

    Article  Google Scholar 

  12. Shepler, B. C., Han, Y., and Bowman, J. M., Are Roaming and Conventional Saddle Points for \(\rm{H_{2}CO}\) and \(\rm{CH_{3}CHO}\) Dissociation to Molecular Products Isolated from Each Other?, J. Phys. Chem. Lett., 2011, vol. 2, no. 7, pp. 834–838.

    Article  Google Scholar 

  13. Mauguiére, F. A. L., Collins, P., Kramer, Z. C., Carpenter, B. K., Ezra, G. S., Farantos, S. C., and Wiggins, S., Roaming: A Phase Space Perspective, Annu. Rev. Phys. Chem., 2017, vol. 68, no. 1, pp. 499–524.

    Article  Google Scholar 

  14. Hare, S. R. and Tantillo, D. J., Post-Transition State Bifurcations Gain Momentum — Current State of the Field, Pure Appl. Chem., 2017, vol. 89, no. 6, pp. 679–698.

    Article  Google Scholar 

  15. Wiggins, S., Ordinary Differential Equations, https://open.umn.edu/opentextbooks/textbooks/ordinarydifferential-equations (2017).

  16. Uzer, T., Jaffé, Ch., Palacián, J., Yanguas, P., and Wiggins, S., The Geometry of Reaction Dynamics, Nonlinearity, 2002, vol. 15, no. 4, pp. 957–992.

    Article  MathSciNet  Google Scholar 

  17. 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 

  18. Naik, Sh. and Wiggins, S., Finding Normally Hyperbolic Invariant Manifolds in Two and Three Degrees of Freedom with Hénon – Heiles-Type Potential, Phys. Rev. E, 2019, vol. 100, no. 2, 022204, 14 pp.

    Article  MathSciNet  Google Scholar 

  19. de Souza, R. T., Huizenga, J. R., and Schröder, W. U., Effect of a Steep Gradient in the Potential Energy Surface on Nucleon Exchange, Phys. Rev. C, 1988, vol. 37, no. 5, pp. 1901–1919.

    Article  Google Scholar 

  20. Doye, J. P. K. and Wales, D. J., On Potential Energy Surfaces and Relaxation to the Global Minimum, J. Chem. Phys., 1996, vol. 105, no. 18, pp. 8428–8445.

    Article  Google Scholar 

  21. Suits, A. G., Roaming Reactions and Dynamics in the van der Waals Region, Annu. Rev. Phys. Chem., 2020, vol. 71, no. 1, pp. 77–100.

    Article  Google Scholar 

  22. De Leon, N. and Berne, B. J., Intramolecular Rate Process: Isomerization Dynamics and the Transition to Chaos, J. Chem. Phys., 1981, vol. 75, no. 7, pp. 3495–3510.

    Article  MathSciNet  Google Scholar 

  23. Waalkens, H. and Wiggins, S., Direct Construction of a Dividing Surface of Minimal Flux for Multi-Degree-of-Freedom Systems That Cannot Be Recrossed, J. Phys. A, 2004, vol. 37, no. 35, pp. L435–L445.

    Article  MathSciNet  Google Scholar 

  24. Katsanikas, M., García-Garrido, V. J., and Wiggins, S., The Dynamical Matching Mechanism in Phase Space for Caldera-Type Potential Energy Surfaces, Chem. Phys. Lett., 2020, vol. 743, 137199, pp.

    Article  Google Scholar 

  25. Crawford, J. D., Introduction to Bifurcation Theory, Rev. Mod. Phys., 1991, vol. 63, no. 4, pp. 991–1037.

    Article  MathSciNet  Google Scholar 

  26. Lyu, W., Naik, Sh., and Wiggins, S., UPOsHam: A Python Package for Computing Unstable Periodic Orbits in Two-Degree-of-Freedom Hamiltonian Systems, J. Open Source Softw., 2020, vol. 5, no. 45, pp. 1684–1689.

    Article  Google Scholar 

  27. 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, 44 pp.

    Article  Google Scholar 

  28. Ezra, G. S. and Wiggins, S., Sampling Phase Space Dividing Surfaces Constructed from Normally Hyperbolic Invariant Manifolds (NHIMs), J. Phys. Chem. A, 2018, vol. 122, no. 42, pp. 8354–8362.

    Article  Google Scholar 

  29. Naik, Sh. and Ross, Sh. D., Geometry of Escaping Dynamics in Nonlinear Ship Motion, Commun. Nonlinear Sci. Numer. Simul., 2017, vol. 47, pp. 48–70.

    Article  MathSciNet  Google Scholar 

  30. Ross, Sh. D., BozorgMagham, A. E., Naik, Sh., and Virgin, L. N., Experimental Validation of Phase Space Conduits of Transition between Potential Wells, Phys. Rev. E, 2018, vol. 98, no. 5, 052214, 6 pp.

    Article  Google Scholar 

  31. Marston, C. C. and De Leon, N., Reactive Islands As Essential Mediators of Unimolecular Conformational Isomerization: A Dynamical Study of 3-Phospholene, J. Chem. Phys., 1989, vol. 91, no. 6, pp. 3392–3404.

    Article  Google Scholar 

  32. De Leon, N., Mehta, M. A., and Topper, R. Q., Cylindrical Manifolds in Phase Space As Mediators of Chemical Reaction Dynamics and Kinetics: 1. Theory, J. Chem. Phys., 1991, vol. 94, no. 12, pp. 8310–8328.

    Article  MathSciNet  Google Scholar 

  33. De Leon, N., Mehta, M. A., and Topper, R. Q., Cylindrical Manifolds in Phase Space As Mediators of Chemical Reaction Dynamics and Kinetics: 2. Numerical Considerations and Applications to Models with Two Degrees of Freedom, J. Chem. Phys., 1991, vol. 94, no. 12, pp. 8329–8341.

    Article  MathSciNet  Google Scholar 

  34. Naik, Sh. and Wiggins, S., Finding Normally Hyperbolic Invariant Manifolds in Two and Three Degrees of Freedom with Hénon – Heiles-Type Potential, Phys. Rev. E, 2019, vol. 100, no. 2, 022204, 14 pp.

    Article  MathSciNet  Google Scholar 

  35. Waalkens, H., Burbanks, A., and Wiggins, S., A Formula to Compute the Microcanonical Volume of Reactive Initial Conditions in Transition State Theory, J. Phys. A, 2005, vol. 38, no. 45, pp. L759–L768.

    Article  MathSciNet  Google Scholar 

Download references

ACKNOWLEDGMENTS

The authors would like to acknowledge the London Mathematical Society (Summer Research Bursary 2019) and the School of Mathematics at the University of Bristol for supporting WL. We would like to thank the anonymous reviewer for providing critical feedback and pointing towards using steepness as an alternative to flatness.

Funding

We acknowledge the support of EPSRC Grant No. EP/P021123/1 and Office of Naval Research Grant No. N00014-01-1-0769.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Shibabrat Naik.

Ethics declarations

The authors declare that they have no conflicts of interest.

Additional information

MSC2010

37J05,37J15,37J20,34C23,70H05,37G05

APPENDIX

Derivation of the Normal Form Hamiltonian for the Saddle-node Bifurcation

The following is concerned with the classical Hamiltonian saddle-node bifurcation and using it to model some phenomena of interest relevant to chemical reactions —bifurcations of NHIMs, depth and flatness of the potential energy surface.

The normal form for the one DOF Hamiltonian saddle node bifurcation is given by

$$H(q,p)=\frac{p^{2}}{2}-\mu q+\frac{q^{3}}{3},$$
(A.1)
with corresponding Hamilton’s equations
$$\displaystyle\dot{q}=p,$$
$$\displaystyle\dot{p}=-q^{2}+\mu,$$
(A.2)
and \(\mu\) is the bifurcation parameter. The equilibria are given by \(q^{2}=\mu,p=0\).

The Jacobian of the vector field is

$$\left(\begin{array}[]{cc}0&1\\ -2q&0\end{array}\right).$$
(A.3)
We evaluate the Jacobian at each equilibrium point to determine stability:
$$(-\sqrt{\mu},0):\qquad\left(\begin{array}[]{cc}0&1\\ 2\sqrt{\mu}&0\end{array}\right),\qquad\mbox{saddle},$$
(A.4)
$$(\sqrt{\mu},0):\qquad\left(\begin{array}[]{cc}0&1\\ -2\sqrt{\mu}&0\end{array}\right),\qquad\mbox{center}.$$
(A.5)

In the Hamiltonian saddle node described above the equilibria move as \(\mu\) is varied. It will be useful to fix the saddle point at the origin. To do this we introduce the following coordinate transformation: \(q=x-\sqrt{\mu},p=y\). Substituting this into (A.2) gives

$$\displaystyle\dot{q}=\dot{x}=p=y,$$
$$\displaystyle\dot{p}=\dot{y}=-(x-\sqrt{\mu})^{2}+\mu,$$
$$\displaystyle=-x^{2}+2x\sqrt{\mu}$$
(A.6)
or
$$\displaystyle\dot{x}=y,$$
$$\displaystyle\dot{y}=2\sqrt{\mu}x-x^{2},$$
(A.7)
with the corresponding Hamiltonian
$$H(x,y)=\frac{y^{2}}{2}-\sqrt{\mu}x^{2}+\frac{x^{3}}{3}.$$
(A.8)

The equilibria are given by

$$(x,y)=(0,0),(2\sqrt{\mu},0).$$
(A.9)
The Jacobian of the vector field is
$$\left(\begin{array}[]{cc}0&1\\ -2x+2\sqrt{\mu}&0\end{array}\right).$$
(A.10)

We evaluate the Jacobian at each equilibrium point to determine stability:

$$\displaystyle(0,0):\qquad\left(\begin{array}[]{cc}0&1\\ 2\sqrt{\mu}&0\end{array}\right),\qquad\mbox{saddle},$$
(A.11)
$$\displaystyle(2\sqrt{\mu},0):\qquad\left(\begin{array}[]{cc}0&1\\ -2\sqrt{\mu}&0\end{array}\right),\qquad\mbox{center}.$$
(A.12)

The depth of the potential well is controlled by the cubic term in the potential energy surface. Therefore, we introduce a parameter that allows us to vary the amplitude of this term:

$$H(x,y)=\frac{y^{2}}{2}-\sqrt{\mu}x^{2}+\frac{\alpha x^{3}}{3},\qquad\alpha>0,$$
(A.13)
$$\displaystyle\dot{x}=y,$$
$$\displaystyle\dot{y}=2\sqrt{\mu}x-\alpha x^{2}.$$
(A.14)
The equilibria are given by
$$(x,y)=(0,0),\left(\frac{2}{\alpha}\sqrt{\mu},0\right).$$
(A.15)
The Jacobian of the vector field is given by
$$\left(\begin{array}[]{cc}0&1\\ 2\sqrt{\mu}-2\alpha x&0\end{array}\right).$$
(A.16)
We evaluate the Jacobian at the equilibria to determine their stability:
$$\displaystyle(0,0):\qquad\left(\begin{array}[]{cc}0&1\\ 2\sqrt{\mu}&0\end{array}\right),\qquad\mbox{saddle}.$$
(A.17)
$$\displaystyle\left(\frac{2}{\alpha}\sqrt{\mu},0\right):\qquad\left(\begin{array}[]{cc}0&1\\ -3\sqrt{\mu}&0\end{array}\right),\qquad\mbox{center}.$$
(A.18)

The depth of the potential energy surface is determined by the difference between the potential evaluated at the saddle minus the potential evaluated at the center (minumum of the well). The potential energy function is given by

$$V(x)=-\sqrt{\mu}x^{2}+\frac{\alpha x^{3}}{3},\qquad\alpha>0,$$
(A.19)
and this difference is given by
$$\displaystyle V(0)-V\left(\frac{2}{\alpha}\sqrt{\mu}\right)=\frac{4}{\alpha^{2}}\mu\sqrt{\mu}-\frac{\alpha}{3}\frac{4}{\alpha^{2}}\mu\frac{2}{\alpha}\sqrt{\mu},$$
$$\displaystyle=\left(\frac{4}{\alpha^{2}}-\frac{8\alpha}{3\alpha^{3}}\right)\mu\sqrt{\mu},$$
$$\displaystyle=\left(\frac{12}{3\alpha^{2}}-\frac{8}{3\alpha^{2}}\right)\mu\sqrt{\mu},$$
$$\displaystyle=\frac{4}{3\alpha^{2}}\mu\sqrt{\mu}.$$
(A.20)

Hence, for a fixed \(\mu\) (that is, the distance apart from the two equilibria) the potential is made less deep by taking large \(\alpha\).

Derivation of the Expression for the Ratio of the Bottleneck Width and the Well Width for the Coupled two DOF System

The following expression is defined as \(A\):

$$A=e-\left(\dfrac{\alpha}{3}x^{3}-\sqrt{\mu}x^{2}+\dfrac{\varepsilon}{2}x^{2}\right)+\frac{2x^{2}\varepsilon^{2}}{4(\omega^{2}+\varepsilon)}.$$
(A.21)

Then the ratio of the bottleneck width and the well width can be written as

$$\displaystyle R_{bw}=\dfrac{w_{b}}{w_{w}}=\dfrac{\text{width of the bottleneck}}{\text{width of the well}}$$
$$\displaystyle=\dfrac{\sqrt{\dfrac{2e}{\omega^{2}+\varepsilon}}}{\sqrt{A|_{x=x^{e},V=e}}\sqrt{\dfrac{2}{\omega^{2}+\varepsilon}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e-(\dfrac{\alpha}{3}(x^{e})^{3}-\sqrt{\mu}(x^{e})^{2}+\dfrac{\varepsilon}{2}(x^{e})^{2})+\dfrac{2(x^{e})^{2}\varepsilon^{2}}{4(\omega^{2}+\varepsilon)}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e-(\dfrac{\alpha}{3}x^{e}-\sqrt{\mu}+\dfrac{\varepsilon}{2}-\dfrac{2\varepsilon^{2}}{4(\omega^{2}+\varepsilon)})((x^{e})^{2})}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e-(\dfrac{\alpha}{3}\dfrac{1}{\alpha}(2\sqrt{\mu}-\dfrac{\omega^{2}\varepsilon}{\omega^{2}+\varepsilon})-\sqrt{\mu}+\dfrac{\varepsilon}{2}-\dfrac{2\varepsilon^{2}}{4(\omega^{2}+\varepsilon)})\dfrac{1}{\alpha^{2}}(2\sqrt{\mu}-\dfrac{\omega^{2}\varepsilon}{\omega^{2}+\varepsilon})^{2}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e-\dfrac{1}{\alpha^{2}}(\dfrac{2\sqrt{\mu}}{3}-\dfrac{\omega^{2}\varepsilon}{3(\omega^{2}+\varepsilon)}-\sqrt{\mu}+\dfrac{\varepsilon}{2}-\dfrac{2\varepsilon^{2}}{4(\omega^{2}+\varepsilon)})(2\sqrt{\mu}-\dfrac{\omega^{2}\varepsilon}{\omega^{2}+\varepsilon})^{2}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e-\dfrac{1}{\alpha^{2}}(-\dfrac{\sqrt{\mu}}{3}+\dfrac{\omega^{2}\varepsilon}{6(\omega^{2}+\varepsilon)})(2\sqrt{\mu}-\dfrac{\omega^{2}\varepsilon}{\omega^{2}+\varepsilon})^{2}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e+\dfrac{1}{6\alpha^{2}}(2\sqrt{\mu}-\dfrac{\omega^{2}\varepsilon}{\omega^{2}+\varepsilon})^{3}}}$$
$$\displaystyle=\sqrt{\dfrac{e}{e+\mathcal{D_{\varepsilon}}}.}$$
(A.22)

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Lyu, W., Naik, S. & Wiggins, S. The Role of Depth and Flatness of a Potential Energy Surface in Chemical Reaction Dynamics. Regul. Chaot. Dyn. 25, 453–475 (2020). https://doi.org/10.1134/S1560354720050044

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1134/S1560354720050044

Keywords

Navigation