Get 20M+ Full-Text Papers For Less Than $1.50/day. Start a 14-Day Trial for You or Your Team.

Learn More →

Tilting and Squeezing: Phase space geometry of Hamiltonian saddle-node bifurcation and its influence on chemical reaction dynamics

Tilting and Squeezing: Phase space geometry of Hamiltonian saddle-node bifurcation and its... In this article we present the in uence of a Hamiltonian saddle-node bifurcation on the high- dimensional phase space structures that mediate reaction dynamics. To achieve this goal, we identify the phase space invariant manifolds using Lagrangian descriptors, which is a trajectory- based diagnostic suitable for the construction of a complete \phase space tomography" by means of analyzing dynamics on low-dimensional slices. First, we build a Hamiltonian system with one degree-of-freedom (DoF) that models reaction, and study the e ect of adding a parameter to the potential energy function that controls the depth of the well. Then, we extend this framework to a saddle-node bifurcation for a two DoF Hamiltonian, constructed by coupling a harmonic oscillator, i.e. a bath mode, to the other reactive DoF in the system. For this problem, we describe the phase space structures associated with the rank-1 saddle equilibrium point in the bottleneck region, which is a Normally Hyperbolic Invariant Manifold (NHIM) and its stable and unstable manifolds. Finally, we address the qualitative changes in the reaction dynamics of the Hamiltonian system due to changes in the well depth of the potential energy surface that gives rise to the saddle-node bifurcation. arXiv:1907.03322v1 [nlin.CD] 7 Jul 2019 Keywords: Saddle-Node bifurcation, Hamiltonian systems, Phase space structure, La- grangian descriptors, Chemical reaction dynamics. Contents I. Introduction 2 II. Hamiltonian Saddle-Node Bifurcation Models 7 A. One Degree-of-Freedom Hamiltonian 7 B. Two degrees-of-freedom Hamiltonian 15 III. Results and Discussion: In uence of well-depth on reaction dynamics 21 A. One Degree-of-Freedom Hamiltonian 21 B. Two degree-of-freedom Hamiltonian 24 IV. Conclusions 34 Acknowledgments 36 References 36 A. Lagrangian Descriptors 39 B. Visualization of phase space structure for the quadratic normal form Hamiltonian 45 C. Computation of NHIM and its invariant manifolds for the 2 DoF system 48 I. INTRODUCTION Background. Molecular reaction dynamics is concerned with the breaking and forming of bonds between the atoms that make up a molecule. The description of such mechanisms requires analyzing how the bonds change in time, that is, it is a dynamical mechanism re- quiring a phase space description. The phase space will have coordinates that correspond to the bond con gurations with a corresponding canonical momentum for each bond con gura- tion coordinate. Thus, the dynamics of chemical bonds can be formulated in the language of 2 Hamiltonian dynamics and its resulting phase space geometry. In recent years there has been signi cant developments in the mathematical description and analysis of chemical reaction 1{3 dynamics in phase space . Our objective is to describe the change in a bond that signals the occurrence of a chemical reaction in terms of a unique characteristic of a trajectory of Hamilton's equations. This characteristic requires an understanding of the geometry of the phase space of the molecule in a way that enables us to divide an appropriate volume of phase space into a region corresponding to \reactants" and a region corresponding to \products". The passage from reactants to products, that is \reaction", occurs when a trajectory crosses the \dividing surface" (DS) between the reactants and products. In this framework for understanding chemical reactions, the ux through such a dividing surface would be related to the reaction rate, and therefore the construction of DS between such regions is of interest in reaction dynamics. The description of the regions of reactants and products can often be inferred from the nature of the development of the coordinates used in the mathematical model of the chemical reaction. Once the model is developed, the DS must be constructed in the context of this model. In relating the ux through the DS to a reaction rate it is desirable that trajectories crossing the DS from the reactant side proceed to the product side before possible recrossing the DS back to the reactant region. Thus, we require the 4,5 DS to have the \no-recrossing" property . Hamiltonian dynamics conserves energy and therefore the trajectories evolve on a xed energy surface of one less dimension (referred to as codimension one) than the phase space. The DS is required to be codimension one in the energy surface. Hence, we require a DS between reactants and products to be codimension one in the energy surface and to have the no-recrossing property. We note that Wigner had 6,7 already described these properties for a DS in phase space much earlier , and a review of classical and quantum versions of DS constructed in phase space can be found in . Traditionally, the construction of DS was initially focused on critical points of the po- tential energy surface (PES), that is, in the con guration space describing the molecular system. Critical points on the PES do have signi cance in phase space; they are the equilib- rium points for zero momentum. But they continue to have in uence for nonzero momentum for a range of energies above the energy of the equilibrium point. The precise manner of this 1,5,9 dynamical in uence has only been understood recently and we will describe this shortly . The construction of a DS separating the phase space into two parts, reactants and products, 3 has been a focus from the dynamical systems point of view in recent years. However, the lack of a rm theoretical basis for the construction of such surfaces for molecular systems with three and more degree-of-freedom (DoF) has until recently been a major obstacle in the development of the theory. In phase space, that is for nonzero momentum, the role of the saddle point is played by an invariant manifold of saddle stability type, the nor- 10{12 mally hyperbolic invariant manifold (NHIM) (see x: II B) . In order to fully appreciate the NHIM and its role in reaction rate theory, it is useful to begin with a precursor con- cept { the periodic orbit dividing surface or PODS. For systems with two DoF described by a natural Hamiltonian, kinetic plus potential energy, the problem of constructing the 13{16 DS in phase space was solved during the 1970s by McLa erty, Pechukas and Pollak . They demonstrated that the DS at a speci c energy is related to an invariant phase space structure, an unstable periodic orbit (UPO). The UPO de nes (it is the boundary of ) the bottleneck in phase space through which the reaction occurs and the DS which intersects trajectories evolving from reactants to products can be shown to have the geometry of a 5,17 hemisphere in phase space whose boundary is the unstable PO . The same construction can be carried out for a DS intersecting trajectories crossing from products to reactants and these two hemispheres form a sphere for which the UPO is the equator. Generalisation of this construction of DS to high dimensional systems has been a central question in reaction 2,17 dynamics and has only received a satisfactory answer in recent years . The key diculty concerns the high dimensional analogue of the unstable PO used in the two DoF system for the construction of the DS. This diculty is resolved by considering the NHIM, which has the appropriate dimensionality for anchoring the dividing surface in phase space. Results from dynamical systems theory show that transport in phase space is controlled by high dimensional manifolds, NHIMs, which are the natural generalisation of the UPO of the two DoF case . Normal hyperbolicity of these invariant manifolds means that their stability, in a precise sense, is of saddle type in the transverse direction, which implies that they possess stable and unstable invariant manifolds that are impenetrable barriers and mediate transport in phase space. These invariant manifolds of the NHIM are structurally stable, that is, stable under perturbation . For two DoF systems, the NHIM is an unstable PO, and for an n > 2 DoF system at a xed energy, the NHIM has the topology of a (2n 3)-dimensional sphere and is the equator of a (2n 2)-dimensional sphere which constitutes the DS. The DS can be used to divide the (2n 1)-dimensional energy surface 4 1,3,9,18,19 into two parts, reactants and products . An elementary description of the role of the NHIM in reaction dynamics is given in . Fundamental theorems assure the existence of the phase space structures |NHIM and its invariant manifolds |for a range of energies above that of the saddle . However, the precise extent of this range, as well as the nature and consequences of any bifurcations of the phase space structures that might occur as energy 21{26 is increased, is not known and is a topic of continuing research . Motivation. In this article, we investigate the changes in geometry of the phase space structures in a normal form Hamiltonian that undergoes saddle-node bifurcation. In general, a Hamiltonian system with a saddle equilibrium point whose eigenvalues depend on the potential well depth parameter will exhibit bifurcation when the potential energy barrier height decreases. If the potential energy surface has a single well, then the result is a collision of the stable and saddle equilibrium points as we illustrate in Figs. 3 and 5. This leads to drastic changes in the geometry of phase space structures | NHIM and its invariant manifolds | which can then be used to account for corrections to Kramers' reaction rate as barrier height decreases . Furthermore, quantifying rates of crossing low or vanishing barrier is signi cant for experimental study of single bond dynamics of molecules and control 27{30 of micro and nano-electromechanical devices . We would also like to point out that the questions addressed in this article are further motivated by the work of Borondo and co- 31{33 authors who noted the signi cance of a saddle-node bifurcation in the isomerization of LiCN/LiNC molecule. Conservative dynamics on an open potential well has received considerable attention because the geometry of phase space structures explains the intricate fractal structure of 34{36 ionization rates . Furthermore, the discrepancies in observed and predicted ionization rates in atomic systems has also been explained by accounting for the topology of these phase space structures. These have been connected with the breakdown of the ergodic assumption that is the basis for using ionization and dissociation rate formulae . This rich literature on chaotic escape of electrons from atoms sets a precedent for delineating the changes in phase space structures due to saddle-node bifurcation that can be expected in open potential wells where the eigenvalues of the saddle equilibrium point depend on the 36,38{41 well depth parameter . Approach in this article. We rst develop a Hamiltonian model that exhibits saddle- node bifurcation when the well depth is decreased and obtain a relationship between the 5 parameters of the model for the bifurcation in the 2 DoF system. Then, the NHIM (an unstable PO in the 2 DoF system) is computed using di erential correction and numeri- cal continuation, and its invariant manifolds are computed using globalization. The role of these invariant manifolds in reaction dynamics and the implications of the change in their geometry by varying well depth is discussed. We compare the results with another numerical method that can be used for identifying high dimensional phase space (4 or more dimensions) structures called Lagrangian descriptors (LDs). Lagrangian descriptors is a trajectory di- agnostic method for revealing invariant manifolds that mediate phase space transport. The method was originally developed in the context of Lagrangian transport studies in uid dy- namics , but the wide applicability of the method has recently been recognized as a tool to 33,43{48 construct the dividing surface via identi cation of NHIM . The method is straightfor- ward to implement computationally, and it provides a high resolution method for locating high dimensional invariant manifolds using low dimensional slices of the high dimensional 49{51 phase space . Therefore, this technique provides us with a rst step towards realizing a complete \phase space tomography" of high dimensional invariant manifolds. It also ap- plies to both Hamiltonian and non-Hamiltonian systems as well as to stochastic dynamical systems . Moreover, Lagrangian descriptors can be applied directly to data sets obtained 54{56 from the numerical solution of geophysical models and satelite observations . We provide more details on our use of Lagrangian descriptors in Appendix A. This paper is outlined as follows. Section II is devoted to the description of the Hamilto- nian models with one and two DoF that we will use in our study of saddle-node bifurcation phenomena in phase space. In Section III we explore the implications of these model prob- lems for reaction dynamics, and discuss our results by analyzing the phase space geometrical structures responsible for the escaping dynamics from the PES and their relevance for chem- ical reactions. Finally, in Section IV we summarize the conclusions of this work. The reader can nd all the relevant information related to the method of Lagrangian Descriptors and its implementation details in Appendix A. Furthermore, Appendix C describes the numerical techniques used (di erential correction and manifold globalization) for the computation of the NHIM and its stable and unstable manifolds associated to the rank-1 saddle point in the 2 DoF Hamiltonian system that characterizes the saddle-node phase space bifurcation. 6 II. HAMILTONIAN SADDLE-NODE BIFURCATION MODELS In this section we introduce the normal form Hamiltonian that exhibit saddle-node bi- furcation and their implications for reaction dynamics. First, we describe the normal form Hamiltonian system with one DoF, and parametrize this model by adding a parameter that controls the potential well depth. Next, we extend this model to a two DoF Hamiltonian system by coupling a bath mode (a harmonic oscillator) to the reaction DoF. A. One Degree-of-Freedom Hamiltonian The normal form for the one DoF Hamiltonian that undergoes a saddle-node bifurcation in phase space can be written as 1 1 2 3 H (u; v) = v  u + u ; (1) 2 3 where  2 R is the bifurcation parameter. Hamilton's equations are given by @H > u _ = = v @v ; (2) > @H : 2 v _ = =  u @u and the equilibria of this system are located at (u; v) = ( ; 0). We observe that the dynamical system has two di erent equilibrium points for  > 0, which approach each other as the bifurcation parameter goes to zero. When  = 0, both equilibria \collide" into one equilibrium point at (0; 0). For  < 0, the resulting dynamical system has no equilibria. In order to study the stability of the equilibria, we linearize the vector eld and calculate the Jacobian 0 1 2 2 0 1 @ H @ H B 2 C 0 1 @u@v @v B C @ A J(u; v) = = : (3) @ 2 2 A @ H @ H 2u 0 @u @v@u Evaluating the Jacobian at the equilibrium point ( ; 0), we get 0 1 0 1 @ A J( ; 0) = ; (4) 2  0 7 p which has eigenvalues  4, and hence the equilibrium point ( ; 0) is a saddle. The Jacobian at ( ; 0) is 0 1 0 1 @ A J( ; 0) = ; ; (5) 2  0 which has eigenvalues  4jj i,yielding thst the equilibrium point ( ; 0) is a center. For = 0, zero is the only eigenvalue of the Jacobian matrix, and therefore the linearization does not provide enough information, and one needs to include higher order terms to determine the stability of the equilibrium point (0; 0). We compute LDs with the goal of detecting the invariant manifolds of the dynamical system given in Eq. (2). As shown by the phase portraits in Fig. 1, trajectories of (2) can escape o to in nity in nite time depending on the total energy of the system and the location of the initial conditions. Therefore, this issue needs to be accounted for in the computation of LDs as explained in Appendix A, and the approach we will follow to 44,45 resolve this problem has also been adopted in the chemical reactions literature . In particular we will use the p-norm in the LD de nition (A4) with p = 1=2. Therefore, we evolve initial conditions forward and backward in time for  = 8 and whenever a trajectory leaves the domain de ned by the circle of radius 15 about the origin, we stop the numerical integration of that trajectory. In Fig. 1 we show the LD contour maps for three di erent values of the bifurcation parameter  = 0:25; 0; 0:25 and compare the results obtained by means of LDs with the corresponding phase portraits. We can clearly see in Fig. 1 how the method succesfully recovers all the relevant phase space structures. The stable and unstable manifolds are highlighted by the singularities present in the LD contour map. Furthermore, another interesting aspect to highlight from Fig. 1D is that, for  = 0, the LD values in the neighborhood of the origin seem to indicate that a bifurcation is going to take place in the phase space structure, as one can observe that a `ghost' center structure is about to be created close to the cusp at the origin. a. Fixing the saddle point at the origin. In the Hamiltonian saddle-node model (1), the equilibria move as the bifurcation parameter  is varied. In order to simplify the bifurcation analysis necessary to address the implications for chemical reaction dynamics, we x the saddle point at the origin. Moreover, this form of non-moving saddle will become useful when discussing the 2 DoF model of the saddle-node bifurcation. This transformation will also facilitate the construction of the dividing surface from the NHIM associated with the 8 A) B) C) D) E) F) FIG. 1: Comparison between the phase space portrait of the dynamical system represented by Eq. (2) (left column) and the output of variable time LDs calculated using the p-norm de nition with p = 1=2 and  = 8 (right column). A) and B) correspond to the value of the bifurcation parameter = 0:25; C) and D) for  = 0; D) and E) for  = 0:25. We have marked di erent trajectories in the phase portraits in blue, invariant manifolds in red, and equilibrium points as yellow dots. 9 `non-moving' rank-1 saddle. Let us consider the linear change of coordinates u = q ;  2 R [f0g ; (6) v = p and substitute this transformation into Eq. (2) to yield the following dynamical system in the new coordinates q_ = p ; (7) : 2 p _ = 2  q q with corresponding Hamiltonian 1 p 1 2 2 3 H (q; p) = p  q + q : (8) 2 3 The equilibrium points of Eq. (7) are (0; 0) and (2 ; 0). Their stability is characterized by the eigenvalues of the Jacobian matrix 0 1 2 2 @ H @ H 0 1 B C 0 1 @q@p @p B C @ A J(q; p) = = (9) 2 2 p @ A @ H @ H 2  2q 0 @q @p@q evaluated at each equilibrium point 0 1 0 1 0 1 0 1 @ A @ A J(0; 0) = ; J(2 ; 0) = (10) p p 2  0 2  0 with eigenvalues  4 and  4jj i, respectively. Hence, the origin is a saddle and (2 ; 0) is a center equilibrium point. This is also supported by the result that eigenvalues do not change under translational change of coordinates. In Fig. 2, we show the variable integration time LD contour maps for Eq. (7) at two di erent values of the bifurcation parameter,  = 0 and  = 0:25. These plots illustrate how the stable and unstable manifolds of the saddle equilibrium point are identi ed by the singular features, points where LDs are non-di erentiable, in the LD contour maps. These singular features can be visualized by taking one dimensional slices of the LD contour map, where jump discontinuities mark the initial conditions on an invariant manifold. 10 A) B) C) D) FIG. 2: Comparison between the output of variable time LDs using the -norm de nition with = 1=2 and  = 8 (left column) and the M function value along lines parallel to the q-axis (right column). A) and B) correspond to  = 0 and the line of initial conditions is p = 0:2; C) and D) use  = 0:25 and the lines of initial conditions are p = 0:25 and p = 0:75. b. Introducing a parameter to control \depth" of the PES. The depth of the well on the PES is controlled by the cubic term that appears in the Hamiltonian (8). Therefore, we introduce a parameter that allows us to vary the amplitude of this term, and consequently the strength of the nonlinearity in the vector eld (7). Thus, the Hamiltonian becomes 1 p 2 2 3 H (q; p) = p  q + q  T (p) + V (q) ; (11) 2 3 11 + + where  2 R [f0g and 2 R are the two parameters, T is the kinetic energy of the DoF and V its potential energy. Hamilton's equations are given by @H q_ = = p @p : (12) > @H p _ = = 2  q q @q The equilibria are (0; 0) and 2 = ; 0 , and the Jacobian of the vector eld is given by 0 1 0 1 @ A J (q; p) = : (13) 2  2 q 0 We determine the stability of the equilibria by evaluating the Jacobian which gives 0 1 0 1 0 1 0 1 @ A @ A J (0; 0) = ; J (2 = ; 0) = : (14) p p 2  0 2  0 The eigenvalues of the Jacobian at the equilibrium point (0; 0) are  4 with correspond- ing eigenvectors (1; 4). The eigenvalues for the Jacobian at the equilibrium point p p 4 4 2 = ; 0 are  4 i with eigenvectors (1; 4 i). Clearly, the eigenvalues of the Hamil- tonian (8) are perserved, and (0; 0) is a saddle, while 2 = ; 0 is a center. We note here that the eigenvalues and eigenvectors of the Jacobian evaluated at the equilibrium points only depend on . \Depth" of the well on the PES, referred to as well-depth from here on, is determined by the di erence between the potential energy of the saddle and the potential energy of the center (minimum of the well) equilibrium points. The potential energy function is given by 2 3 V (q) =  q + q ; (15) and this di erence is given by F  V (0) V = : (16) Hence, for a xed , the well-depth is increased by decreasing . In fact, if we denote D = 2 = as the distance between the saddle and the center equilibrium point, we get F = D ; (17) 12 therefore, as increases, the well-depth approaches zero faster than the distance between the saddle and the center equlibrium points. Furthermore, the rate at which the well-depth changes as we change the distance between equilibria is given by dF 1 = D =  D : (18) dD 3 3 which is proportional to the product of the square of the eigefrequency associated to the saddle equilibrium and the distance between equilibria. In Fig. 3(a) we illustrate the well-depth together with the distance between the saddle and the center. In Fig. 3(b) shows how the potential energy function changes, for a xed value  = 1, as the well-depth parameter (dstrength of the nonlinearity) is increased. a) b) FIG. 3: A) Potential energy function described by Eq. (15) illustrating the \depth" parameter together with the horizontal distance from the saddle to the center equilibrium. B) Evolution of he potential energy function for a xed value of  = 1, as is varied. Recall that, as we discussed in the introduction, in the context of chemical reaction dynamics the escape from a potential well problem can be identi ed for instance with dis- sociation or fragmentation reactions, where a chemical transformation takes place if a bond of a molecule A breaks up, giving rise to two products B and C. In this situation, the equilibrium conformation of the given molecule A is represented by a potential well in a PES, and dissociation into B and C takes place if the system has sucient energy to croos the potential barrier that separates bounded (vibration) from unbounded (bond breakup) motion. This setting can be modeled for instance by the potential energy function in Eq. (15), and is illustrated in Fig. 4(a), where a locally cubic potential energy function about 13 the potential barrier describes this phenomenon, which could be the result of a saddle-node bifurcation that has occurred in the phase space of the Hamiltonian system. In Fig. 4(b) we show the phase portrait for the dynamical system (12) which shows the equilibrium points and the homoclinic orbit. Fig. 4(c-d) display the changes in the phase space structure using the variable integration time LDs for di erent values of the well-depth parameter . (a) (b) (c) (d) FIG. 4: (a) Cubic potential energy function as a basic model for dissociation chemical reactions. The shape of this potential is the building block (normal form) for saddle-node bifurcations in phase space. (b) Phase portrait showing the location of the saddle and center equilibria in terms of the model parameters and . (c-d) Variable time LDs calculated using the p-norm de nition with p = 1=2 and  = 5 for the dynamical system described in Eq. (12) using (c)  = 1 and = 2 (d)  = 1 and = 4. 14 B. Two degrees-of-freedom Hamiltonian In this section we introduce the normal form for the two DoF Hamiltonian that undergoes a saddle-node bifurcation by extending the one DoF Hamiltonian discussed above. To do so, we add another degree of freedom in the form of a harmonic oscillator with mass m = 1 and frequency !, known as a bath mode in the terminology of chemical reaction dynamics, and parametrize the coupling of the reaction and bath modes. Thus, the Hamiltonian becomes 1 p ! " 2 2 2 3 2 H (q; x; p; p ) = p + p  q + q + x + (x q) ; (19) 2 3 2 2 where > 0 is the well-depth parameter and measures the strength of the nonlinearity, ! > 0 is the frequency of the harmonic oscillator or the bath mode,   0 is the bifurcation parameter, and " > 0 is the strength of the coupling between the reaction and the bath mode. Identifying this Hamiltonian's kinetic energy, T (p; p ), potential energy, V (q; x), and assuming H (q; x; p; p ) = T (p; p ) + V (q; x), we get that x x 1 ! " 2 2 2 3 2 T (p; p ) = p + p ; V (q; x) =  q + q + x + (x q) (20) 2 3 2 2 The corresponding Hamilton's equations are given by: @H > q_ = = p @p @H x _ = = p < x @p (21) @H > p _ = = 2  q q + "(x q) @q > @H p _ = = ! x + "(q x) @x For this Hamiltonian system, the phase space is four dimensional, and since energy is con- served, the trajectories evolve on a three dimensional energy surface. The equilibria for this e e system are located at x = (0; 0; 0; 0) and x = (q ; x ; 0; 0) where e e 1 2 2  ! " " q = ; x = q (22) e e e 2 2 (! + ") ! + " The energy of the system at the equilibrium points is: 4 2 2 2 p ! " 2! " q e e e H (x ) = 0 ; H (x ) = 2  + (23) 1 2 (! + ") 15 One can easily show that there exists a critical value of the coupling strength (also interpreted as a perturbation to the one DoF model) given by 2  ! " = ; (24) ! 2 for which the Hamiltonian (21) has only one equilibrium point at the origin. We note here that this critical value is independent of the nonlinearity strength parameter and requires ! > 2  to be satis ed. Moreover, it is characterized by a functional relationship between the squares of the eigenfrequencies of the reactive and bath modes of the uncoupled system (" = 0). We will analyze the in uence of the perturbation strength " on the geometry of the phase space structures up until this critical condition in Section III. We illustrate in Fig. 5 how the geometry of the PES (20) and the equipotentials in con guration space change as we vary the coupling strength (perturbation parameter) ". For this visualization, we have used the following values of the model parameters  = 0:25, = 2, and ! = 1:25. The reason for doing so is that they satisfy the condition ! > 2  and consequently a critical value of the perturbation parameter exists, given by " = 25=9, at which the two equilibrium points `collide' at the origin resulting in a saddle- node bifurcation. The dynamics after this collision is beyond the scope of this study and therefore we will only focus on the description of the system dynamics for di erent values of the coupling strength when " < " . We observe in Fig. 5(a,b) that the location of the well when the DoF are uncoupled (" = 0) lies on the q-axis. As " is increased, Fig. 5(c,d) show that its e ect is to tilt the PES with respect to the con guration space plane, and the position of the center equilibrium point moves o the q-axis towards the origin. Finally, the situation for which the coupling strength reaches the critical value " is shown in Fig. 5(e,f ) when the collision has happened and there is only one equilibrium point at the origin. a. Linear stability analysis of the equilibrium point at the origin. We show here that e e the equilibrium point x is a rank-1 saddle on the PES and that x is a center corresponding 1 2 to the bottom of the potential well. This linear stability analysis will be useful later for computing the rank-1 saddle's NHIM, which carries its in uence to higher energies, and also to determine the NHIM's invariant stable and unstable manifolds that act as codimension one impenetrable barriers on the constant energy surface. The global geometry of the invariant manifolds is paramount for quantifying reaction rates such as escape rate from potential well. 16 (a) (b) (c) (d) (e) (f ) FIG. 5: Geometry of the PES (left-column) and equipotentials in con guration space (right column) for the model parameters  = 0:25, = 2 and ! = 1:25. (a) and (b) correspond to " = 0; (c) and (d) represent the case " = 0:25; (e) and (f ) are for the critical coupling strength " = 25=9. 17 We determine the stability of the equilibria by linearizing Eq. (21), which gives the Jacobian matrix 0 1 2 2 2 2 @ H @ H @ H @ H B C @q@p @x@p @p @p @p x 0 1 B C B C 0 0 1 0 2 2 2 2 @ H @ H @ H @ H B C B C B C B C B C @q@p @x@p @p@p @p 0 0 0 1 x x x x B C B C J(q; x; p; p ) = = B C B C p 2 2 2 2 B C @ H @ H @ H @ H B C 2  2 q " " 0 0 @ A B C B @q @x@q @p@q @p @qC x 2 B C " ! " 0 0 @ 2 2 2 2 A @ H @ H @ H @ H @q@x @x @p@x @p @x (25) The stability of x = (0; 0; 0; 0) is given by the eigenvalues of the Jacobian 0 1 0 0 1 0 B C B C 0 0 0 1 B C J(x ) = B C (26) B C 2  " " 0 0 @ A " ! " 0 0 which has the characteristic equation e 2 2 2 2 det (J(x ) I) = 2  + " + ! + " " = 0 (27) where are the eigenvalues and I the identity matrix. When the reaction and bath modes are decoupled, that is " = 0 in (27), the eigenvalues are given by  and ! i, where 0 0 = 4 ; ! = ! (28) 0 0 Therefore, the origin is a rank-1 saddle equilibrium point, since the linearized system has ex- actly one pair of real eigenvalues,  and  and the saddle plane is spanned by their eigen- 0 0 12,58 vectors. We know from the Moser's generalization of the Lyapunov Subcenter Theorem that when the energy of the system is above that of the rank-1 saddle, there is a two dimen- sional plane spanned by the eigenvectors of ! i, known as the center invariant manifold. This invariant manifold with normal hyperbolicity is referred to as the normally hyperbolic 2N3 invariant manifold (NHIM) and has the topology of a (2N 3)-sphere, that is S . In the two DoF setting, a NHIM is simply an unstable periodic orbit whose geometry is topo- logically equivalent to a circle, i.e. S . 18 As the energy of the system is increased above the energy of the rank-1 saddle equilibrium point, a bottleneck opens in the phase space connecting dynamically di erent phase space regions and allowing trajectories to move between them. This phenomenon results in a phase space transport mechanism. This framework of understanding chemical reactions is realized by computing the stable and unstable manifolds associated with the unstable periodic orbit. These invariant manifolds have cylindrical geometry, that is Rx , and their global behavior is referred to as tube dynamics. The cylindrical manifolds on the constant energy surface are impenetrable barriers (since they are two dimensional on the three dimensional energy surface) separating the reactive and non-reactive trajectories in the phase space. Thus, they determine the initial conditions that will pass through the bottleneck in some future time (or had passed through it in the past) during their evolution . The unstable periodic orbit provides us with the sca olding to construct the dividing surface that separates the trapped motion in the well region of the PES and the escape to in nity of particle trajectories through the phase space bottleneck. The local (linearized) dynamics mediated by these phase space structures is shown in Appendix B. We will resume the discussion on computing the NHIM and its invariant manifolds in x:III B. When " 6= 0, the reaction and bath modes are coupled and the Hamilton's equations (21) are non-integrable. This means we can expect chaotic trajectories to appear that might lead to reaction by escaping the potential well. However, the geometry of the NHIM and its invariant manifolds still governs this reacting and non-reacting behavior. To simplify the eigenvalue problem arising in the stability analysis, we observe that Eq. (27) can be rewritten in terms of the eigenvalues of the uncoupled system as 2 2 2 2 2 4 2 2 2 2 2 2 2 + " + ! + " " = + !  + 2" + !  "  ! = 0 (29) 0 0 0 0 0 0 0 0 Introducing  = the solutions are 2 2 2 2 !  + ! 0 0 0 0 = " + " : (30) 2 2 We note here that the two possible values of  have opposite signs for all values " > 0 of 2 2 2 2 the coupling strength when !   , and for 0 < " < " whenever the condition ! >  is 0 0 0 0 satis ed, where the critical coupling strength " is given by Eq. (24). If we denote  and c 1;" as the positive and negative roots respectively in Eq. (30), then the eigenvalues of the 2;" 19 Jacobian matrix are  and ! i where " " =  ; ! = j j (31) " 1;" " 2;" This shows that in this case the equilibrium point at the origin is a rank-1 saddle. In particular, when the coupling between the reaction and the bath mode is weak, that is "  1, we have q q 2 2 " ; !  ! + " (32) " " 0 0 which con rms that the rank-1 saddle structure of the equilibrium point at the origin persists under small perturbations. We nish the linear stability analysis of the equilibrium point at the origin with the computation of the eigenvectors. Let us denote the eigenvector of the matrix J (x ) by v = [v ; v ; v ; v ] , then the eigenvalue satis es 1 2 3 4 > v = v 1 3 v = v 2 4 J (x )v = v , (33) v = ( ") v + "v > 3 1 2 v = "v (! + ") v 4 1 2 The eigenvectors associated with the real eigenvalues =  correspond to the tangent directions of the unstable and stable manifolds, respectively, of the NHIM at the rank-1 saddle, and are given by "  " u = 1; ; ; : (34) 2 2 2 2 " + ( + ! ) " + ( + ! ) " 0 " 0 The eigenvectors associated with the complex eigenvalues = ! i span the center subspace of the NHIM at the rank-1 saddle and are given by " ! " w = ; 1; i;! i : (35) 2 2 2 2 " ( + ! ) " ( + ! ) 0 " 0 " where ! is the magnitude of the complex eigenvalue in (31). Therefore, we can write the solution to the linearized system at the rank-1 saddle as t  t i! t " " " x(t) = C e u + C e u + 2Re e w (36) 1 + 2 + where C ; C 2 R and  =  +  i 2 C are constants to be determined from an initial 1 2 1 2 condition. We will use this general solution to the linear system to select an initial guess to search for the NHIM using the di erential correction and numerical continuation method . 20 For the sake of completeness, we show next the linear stability analysis for the equilibrium point x . The Jacobian is given by 0 1 0 0 1 0 B C B C 0 0 0 1 2! " B C J(x ) = ; where = (37) B C B C ! + " 2  " " 0 0 @ A " ! " 0 0 The characteristic equation det (J (x ) I) = 0 becomes 2 2 2 2 2 +  + " + ! + " " = 0 0 0 (38) 4 2 2 2 2 2 2 2 + ! +  + 2" + ! +  " + !  = 0 0 0 0 0 0 0 Introducing  = the solutions are 2 2 2 2 ( + ! ) + ! 0 0 0 0 = " + " (39) 2 2 It can be easily shown that both solutions are negative, and consequently J(x ) has two pairs of complex eigenvalues. This implies that the equilibrium point x is a center. III. RESULTS AND DISCUSSION: INFLUENCE OF WELL-DEPTH ON REAC- TION DYNAMICS In this section we discuss the implications of changing the well-depth in terms of the geometry of the invariant manifolds that mediate reacting trajectories (or escape from the potential well). This is done by identifying the invariant manifolds using Lagrangian de- scriptors and a numerical continuation with globalization method (see Appendix C). The geometry of these invariant manifolds and character of changes a ected by varying the po- tential well-depth are involved in the computation of reaction fraction (or escape rates). A. One Degree-of-Freedom Hamiltonian Here we return to the 1 DoF Hamiltonian (11) where the reaction is de ned as the change in the sign of the con guration coordinate q. For the rank-1 saddle equilibrium point at (0; 0), the energy is H (0; 0) = H = 0 which we call the critical energy, and 21 p that of the center equilibrium point at (2 = ; 0), at the bottom of the potential well, is H (2 = ; 0) = H = 4  =3 . In this model system, the phase space is two- dimensional and since the total energy is conserved, the trajectories evolve on the one- dimensional constant energy lines or isoenergetic contours of the Hamiltonian in the 2D phase space. Therefore, the system is completely integrable, and for a xed energy H the trajectories evolve on the one-dimensional curve 1 p 2 2 3 H = p  q + q (40) 2 3 The value of total energy H with respect to H and H characterizes the nature of trajec- 0 c w tories as follows We discuss the nature of the trajectories depending on the energy H of the system: Case H < H : The initial conditions that satisfy q < = , which lie to the left 0 w of the potential energy barrier at the origin. Initial conditions in this con guration space initially climb the potential and at some point their velocity reverses direction and they roll down the potential going to in nity. In the phase space, they come from in nity and y-by the rank-1 saddle on their way towards in nity. They can not cross the barrier and thus do not lead to reaction shown by the blue curve in Fig. 6. Case H  H < 0: Now two types of trajectories are possible. For the initial w 0 conditions that satisfy =  q < 0, trajectories will show same y-by behavior as in the previous case. If initial conditions satisfy 0 < q < 3 = , then they are con ned in the potential well on the right of the potential barrier. Since they are bounded by the homoclinic orbit, trajectories will never escape the well, and thus do not lead to reaction and shown as green curves in Fig. 6. Case H = H = 0: When the total energy of the initial conditions is equal to the 0 c potential energy of the barrier, resulting trajectories approach the barrier asymptoti- cally in forward and backward time. The initial conditions that start on the left of the potential barrier and asymptotically approach it in forward and backward time form pieces of the stable and unstable manifolds of the saddle equilibrium point. The initial conditions that start on the right of the potential barrier also approach the barrier as time goes to in nity. These trajectories combine to form the homoclinic orbit and shown as a black curves in Fig. 6. 22 8 Reactive Reactive Non-reactive Non-reactive 6 NHIM NHIM DS DS -2 -2 8 8 8 6 6 6 6 6 4 4 4 -4 4 4 -4 4 2 2 2 2 0 0 0 0 0 -2 -2 -2 -2 -2 -2 -6 -4 -4 -4 -6 -4 -4 -4 -6 -6 -6 -6 -6 -6 -8 -8 -8 -5 0 5 -8 -8 -5 0 5 -5 0 5 -8 -5 0 5 -5 0 5 -5 0 5 -8 -8 -5 0 5 -5 0 5 (a) (b) FIG. 6: Trajectory behaviors dictated by the potential energy (15). (a)  = 1 and = 1 with the insets  = 2; 4; 6 for = 1 (b)  = 1 and = 4 with the insets  = 2; 4; 6 for = 4. Here we are showing the reactive and nonreactive trajectories (red and blue, respectively) partitioned by the homoclinic orbit (shown in black) formed by the unstable and stable manifolds of the NHIM (shown as a magenta cross), and the reactive trajectory (red) is at energy H = 2, while the two non-reactive trajectories are at H = 1;5. The dividing surface that the reactive trajectory must cross in order for reaction to occur is marked by green asterisks. Case H > H = 0: The energy of the system is above that of the barrier so trajecto- 0 c ries can escape from the well, that is the con guration space coordinate q can change sign and lead to reaction. The dividing surface that trajectories must cross once in a given direction (locally no recrossing), that is when escaping from the potential well, is given by n o D = (q; p) 2 R j q = 0 ; = H > 0 = 0; 2H : (41) 0 0 We note that in this case the dividing surface for a xed energy and given direction of crossing is a point, or has a geometry of S , thus it can partition the one-dimensional (isoenergetic) curves into \reactant" and \`product" regions. To further characterize the reaction dynamics, we resort to the linear stability analysis near the saddle equi- librium point at the origin. The eigenvalues of the saddle equilibrium point are  4 and the corresponding eigenvectors 1= 4; 1 . Therefore, the origin is a NHIM 23 61 since the dynamics is hyperbolic in directions normal to it . Furthermore, we remark that the stability of the NHIM only depends on the  parameter. The eigenvectors tell us how the unstable and stable manifolds of the NHIM are oriented and can be used to numerically globalize the linear approximation. However, we can compute the stable and unstable manifolds of the saddle equilibrium point analytically by noting that they lie on the zero level curve of the Hamiltonian (total energy). Thus, the stable and unstable manifolds are given by s s u u W (0; 0) = [W (0; 0) ; W (0; 0) = [W (0; 0) (42) l l where is the homoclinic orbit = (q; p) 2 R j q > 0 ; H (q; p) = 0 (43) and s 2 W (0; 0) = (q; p) 2 R j q < 0 ; p > 0 ; H (q; p) = 0 (44) u 2 W (0; 0) = (q; p) 2 R j q < 0 ; p < 0 ; H (q; p) = 0 We note here that the geometry of the invariant manifolds is R  S , that is codimension-1 in R , and hence can form the impenetrable barrier between the re- active and non-reactive trajectories as shown in Fig. 6. B. Two degree-of-freedom Hamiltonian We return to the 2 DoF Hamiltonian (19) which introduces the bath degree-of-freedom as a harmonic oscillator into the reaction dynamics. We show the changes in the geometry of the invariant manifolds due to the changes in the well-depth which leads to a saddle-node bifurcation. This is done using Lagrangian descriptors along with numerical continuation and globalization for computing the NHIM and its invariant manifolds. Let us consider a xed energy H and since the model has 2 DoF we know that the dynamics is on a three-dimensional energy surface given by 1 p ! " 4 2 2 2 3 2 S (H ) = (q; x; p; p ) 2 R p + p  q + q + x + (x q) = H (45) 0 x 0 2 3 2 2 24 The projection of the energy surface onto the (q; x) con guration space is the region of energetically possible motion for a xed energy H , and is given by C(H ) = (q; x) 2 R V (q; x) 6 H 0 0 2 (46) p ! " 2 2 3 2 = (q; x) 2 R  q + q + x + (x q) 6 H 3 2 2 This projection denotes con gurations with positive kinetic energy and has been known in classical mechanics as the Hill's region. The boundary of M (H ) is de ned as the locus of points in the (q; x) plane where the kinetic energy is zero, that is (H V (x; y)) = 0, and is called the zero velocity curve. The trajectories are only able to move on the side of the curve where the kinetic energy is positive, shown as white regions in Fig. 7(b-d). To identify the invariant manifolds, we take two dimensional slices of the energy surface and determine the intersection of the invariant manifolds with these low-dimensional slices. In particular, we calculate Lagrangian desriptor and compare with qualitative understand- ings from Poincar e surface-of-section. The isoenergetic surfaces-of-section are + 4 U = (q; x; p; p ) 2 R x = 0 ; p (q; x; p; H )  0 (47) x x 0 qp + 4 U = (q; x; p; p ) 2 R q = q ; p(q; x; p ; H )  0 (48) x e x 0 xp where q is the con guration space coordinate of the equilibrium point at the bottom of the well on the PES (22). We return to the Hamiltonian (19) where the \reaction" and \bath" DoF are uncoupled, that is " = 0. Therefore, the system is integrable and the trajectories are regular, thus the separable Hamiltonian is 1 p 1 ! 2 2 3 2 2 H (q; x; p; p ) = p q + q + p + x (49) 2 3 2 2 | {z } | {z } H (q;p) H (x;p ) r b x where H is the Hamiltonian for the reaction and H is the Hamiltonian for the bath DoF. r b Given a xed total energy of the system H , the necessary condition for the reaction to take place is when the total energy is above that of the barrier of the PES located at the origin, which is zero. For H  0 the energy surface divides phase space into two disconnected regions as illustrated in Fig. 7 so that we have bounded motion in the potential well region. In Fig. 8, we compare the bounded trajectories (regular dynamics) of the system for H = 0 by computing LDs and Poincar e section on the surface-of-section, U (47). qp 25 (a) (b) (c) (d) FIG. 7: (a) Phase space energy surface, inside which motion takes place, for three di erent values of the energy; (b) Con guration space projection for an energy H = 0:025 below the barrier energy; (c) for the rank-1 saddle energy H = 0; (d) for the energy H = 0:1 which is above the barrier energy. The parameter values chosen are  = 0:25, = 2, ! = 1:25 and " = 0. We have marked the energy accessible regions in white and the forbidden regions in dark grey. The equipotential in black marks the zero velocity curve for which the kinetic energy of the system is zero. Both methods clearly recover, the trajectories on tori, as known for integrable Hamiltonian systems , that lls the energy surface. We observe that the high values (white regions in 8) in the LD contour map recovers the quasiperiodic trajectories, which is a consequence of the relationship between the convergence of time averages of LDs with the Ergodic Partition Theorem as explained in Appendix A. 26 A) B) FIG. 8: Quasiperiodic trajectories describing regular motion in the potential well region of the PES for the uncoupled (" = 0) Hamiltonian system with energy H = 0. The model parameters chosen for this calculation are  = 0:25, = 2 and ! = 1:25; A) Poincar e map; B) LDs obtained for an integration time  = 20. The curve in magenta depicts the energy boundary. When the energy of the system is above the barrier, that is H > 0, the topology of the energy surface changes and a phase space bottleneck opens up in the barrier region allowing the escape from the well as shown in Fig. 7(d). Thus, reaction can take place by crossing the bottleneck and passing from the q > 0 to q < 0 or vice versa. Therefore, q = 0 is a natural choice for de ning a dividing surface (DS) separating reactants (bounded motion in the potential well region) from products (escape to in nity) in phase space or vice versa. Hence, the isoenergetic DS is given by 4 2 2 2 2 D(H ) = (q; x; p; p ) 2 R j q = 0 ; 2H = p + p + ! x (50) 0 x 0 2 3 which has the geometry of a 2-sphere, that is S in R . To be precise, it is an ellipsoid with p p semi-major axis 2H in the p, p axis, and 2H =! in the x axis. This ellipsoid has two 0 x 0 hemispheres, known as the forward and backward DS with the form n o 2 2 2 D (H ) = (q; x; p; p ) 2 R j q = 0 ; p = 2H p ! x f 0 x 0 n o (51) 2 2 2 D (H ) = (q; x; p; p ) 2 R j q = 0 ; p = + 2H p ! x b 0 x 0 Forward \reaction" occurs when trajectories cross D and back \reaction" when trajecto- ries cross D . The forward and backward DS is joined at the equator along the normally hyperbolic invariant manifold given by 4 2 2 2 N (H ) = (q; x; p; p ) 2 R j q = p = 0 ; 2H = p + ! x (52) 0 x 0 27 p which has the topology of S . To be precise, it is an ellipse with semiaxis 2H in the p 0 x direction and 2H =! in the x direction. As discussed earlier, for a 2 DoF Hamiltonian the NHIM is an unstable periodic orbit which extends the in uence of the rank-1 saddle equilibrium point in the bottleneck of the PES (a con guration space concept) into phase space. We note that for H > 0 the NHIM is the correct phase space structure that anchors the barriers to the reaction and carry the e ect of the saddle equilibrium point to a range of energies as given by Moser's generalization of Lyapunov Subcenter Manifold Theorem . The stable and unstable manifolds of the unstable periodic orbit are s s u u W = [W ; W = [W (53) l l where is the homoclinic orbit = (q; x; p; p ) 2 R j q > 0 ; H (q; p) = 0 ; H (x; p ) = H (54) x r b x 0 and the left branches are s 4 W = (q; x; p; p ) 2 R j q < 0 ; p > 0 ; H (q; p) = 0 ; H (x; p ) = H x r b x 0 (55) u 4 W = (q; x; p; p ) 2 R j q < 0 ; p < 0 ; H (q; p) = 0 ; H (x; p ) = H x r b x 0 We note that the stable and unstable invariant manifolds have the structure of a cartesian product of a curve in the (q; p) saddle space and an ellipse in the (x; p ) center space, and thus become cylindrical (or tube ) manifolds. We show the energy surface in Fig. 7(a) at 3 energy values for  = 0 where the bottleneck only opens for H > 0. The NHIM computed using di erential correction and continuation, and its invariant manifolds computed using globalization are shown in Fig. 9(a) for H = 0:05 and  = 0. When " 6= 0, the \reaction" and \bath" modes are coupled and for " small, we can think about the resulting dynamics as a perturbation of the uncoupled case . Given an energy of the system below that of the barrier, that is H  0, the phase space bottleneck is closed and trajectories are trapped in the potential well region. However, due to the perturbation, the unstable regular motion on the tori in the uncoupled system is destroyed as given by the KAM theorem. Therefore, chaotic motion arises in some regions of the phase space due to a non-zero coupling of the \reaction" and \bath" modes. We illustrate the system's behavior for the energy H = 0 (energy of the rank-1 saddle) in Fig. 10 using LDs and Poincar e + + maps on the surfaces of section U (47) and U (48). We observe that there is a distinct qp xp 28 (a) (b) (c) (d) FIG. 9: Con guration space and phase space view of the energy surface and the cylindrical mani- folds of the unstable periodic orbit associated with the rank-1 saddle in the bottleneck. Blue and red denote the stable and unstable manifolds respectively, and the green surface is the isosurface of total energy H = 0:05. The magenta curve represents the unstable periodic orbit. (a-b)  = 0 and (c-d)  = 0:25. correlation between the qualitative dynamics revealed by the Lagrangian descriptor (LD) contour maps and Poincar e sections. That is, chaotic regions of phase space which appear as a sea of points in the Poincar e section and hide the underlying structures of stable and unstable manifolds are completely resolved by LD contour maps where the tangled geometry of the manifolds is revealed by the points where the function attains a local minimum as has also been shown in . The capability of LDs to identify the phase space structures relevant 29 43,49,50 in chemical reaction dynamics can also be found in recent literature . In addition, the generation of Poincar e section relies on tracking the crossing of a 2D surface which can not be guaranteed in high dimensional phase space, while LDs just accumulate a positive scalar quantity along the trajectory and thus have potential to reveal high dimensional phase space structures. A) B) C) D) FIG. 10: Phase space structures of the coupled Hamiltonian with model parameters  = 0:25, = 2, ! = 1:25 and " = 0:25. The total energy of the system is H = 0 (barrier energy). A) Poincar e map on the surface of section U ; B) LDs calculated for  = 75 on the surface of section qp + + U ; C) Poincar e map on the surface of section U ; D) LDs calculated for  = 75 on the surface qp xp of section U . We have marked the energy boundary with a magenta curve. xp Now let us consider the dynamics when the total energy is above that of the rank-1 30 saddle, that is H > 0, and in particular we will set H = 0:05 in our analysis. In this 0 0 situation, a phase space bottleneck opens up in the barrier region, as shown in Fig. 7. In order to describe the structures that mediate reaction dynamics, that is the NHIM (or unstable periodic orbit in this case) and its stable and unstable manifolds, we use di erential correction and continuation along with globalization as described in Appendix C. In Fig. 9(c- d), we have shown the cylindrical manifolds along with the energy surface and the unstable periodic orbit for a xed energy. In order to recover the homoclinic tangle geometry of + + the invariant manifolds, we calculate LDs on the surfaces of section U and U , and qp xp compare with the direct numerical construction of these invariant manifolds. This LD based diagnostic is similar to performing a \phase space tomography" of the high dimensional phase space structures using a low dimensional slice. In Fig. 11 we show the computation of variable time LD for an integration time  = 10 on the slice U . We observe that LD qp clearly identify, by means of points which attain a minimum, the location of the stable and unstable manifolds and the NHIM at the intersection of the invariant manifolds. Since we are using a small integration time of  = 10 to compute LDs in Fig. 11, the complete geometry of the homoclinic tangle is not fully revealed. Therefore, in order to recover a FIG. 11: Variable time LDs calculated on the surface of section U for  = 10. The NHIM and its qp stable and unstable manifolds are revealed as points where the LD eld in non-di erentiable and attains a local minimum. We have marked the energy boundary with a magenta curve. more complete and intricate dynamical picture of the homoclinic tangle, the integration 31 time to compute LDs has to be increased. This is shown in Fig. 12(b), where  = 30 and we observe that the regular motion obtained in the middle of the Poincar e section displayed in Fig. 12(a) corresponds to trajectories that remain trapped in the potential well region and never escape. The trapped dynamics on the tori are also captured by the LD contour map shown in 12(b) which also reveals the homoclinic tangle of the stable and unstable manifolds and the resulting lobe dynamics . We note that in all these computations we are using the variable time de nition of LDs, since the open potential surface causes trajectories to escape to in nity through the bottleneck in nite time and resulting in NaN values in the LD contour map. This issue is discussed in Appendix A and illustrated in Fig. 14, and would hide the important underlying phase space structures making the interpretation of results dicult. In addition, when we analyze the dynamics using Poincar e sections, we can not ensure that trajectories return to the surface of section when escaping to in nity is possible. This will result in blank regions in the Poincar e sections as shown in Figs. 12(a) and (c). However, we note that trapped trajectories in the potential well corresponding to regular (motion on the tori) and chaotic motion are highlighted as expected in the Poincar e sections. These trajectories are non-reactive and will remain so until they satisfy the sucient condition for reaction which is entering the cylindrical manifolds. Identifying the regions inside the cylindrical (tube) stable and unstable manifolds can be done using globalization or using the LD contour map on appropriate isoenergetic surfaces. We compute LD contour maps to recover the intersections of these tube manifolds with the surface of section (48). The intersection of the invariant manifolds with the isoenergetic surface of section yields topological ellipses, known in the chemical reactions literature as 48,64{66 reactive islands and is involved in calculating reaction rates/fraction. In order to illustrate the capability of LDs to recover the reactive island structure, we have compared the LD contour map and manifold intersection with the surface of section U obtained xp using globalization in Fig. 13. We have shown the rst intersection of the stable and unstable manifold with the surface of section as a blue/red curve superimposed on the LD contour map. We observe that the minima in the LD contour maps and the manifolds intersections are in agreement, thus verifying the LD based identi cation of reaction islands. We also observe that successive fold and resulting intersection of the tube manifolds with the surface of section are also revealed by the points with minima in the LD values. We note here that the rst intersection of the manifolds encloses a large phase space volume in the potential 32 (a) (b) (c) (d) FIG. 12: Phase space structures of the coupled Hamiltonian with model parameters  = 0:25, = 2, ! = 1:25 and " = 0:25. The total energy of the system is H = 0:05, which is above the barrier energy. (a) Poincar e map on the surface of section U (b) LDs calculated for  = 30 on qp + + the surface of section U (c) Poincar e map on the surface of section U (d) LDs calculated for qp xp = 30 on the surface of section U . We have marked with a magenta curve the energy boundary. xp well which indicates that a large portion of the potential well escapes to in nity through the bottleneck. This is despite the fact that we have chosen a very small value for the energy of the system, H = 0:05, compared to the energy of the barrier. This is a consequence 33 of using a high coupling strength, " = 0:25, which makes the vase-like shaped potential well region small and narrow as the center equilibrium point at the bottom of the potential well approaches the rank-1 saddle equilibrium point located at the origin as " is increased. Movies illustrating the change in both the shape of the energy surface and the equipotentials in con gurations space as we vary the coupling strength can be found at here and here, respectively. We can see that the e ect of increasing the coupling strength from zero is to tilt and squeeze the vase-like shape of the energy surface that qualitatively increases the number of reactive trajectories. This action of tilting and squeezing of the vase-like container (boundary de ned by the total energy surface) is to pour out its contents (the reactive trajectories) and hence the increase in reaction fraction. A quantitative investigation of this phenomenon is current work in progress and beyond the scope of this article. IV. CONCLUSIONS In this article we have presented new results, by means of using a normal form Hamil- tonian that models a saddle-node bifurcation, to investigate the in uence on phase space dynamics of decreasing the potential well depth of a PES. The decrease in potential well depth causes the saddle and center equilibrium points to collide |an e ect that takes place in the con guration space. The resulting saddle-node bifurcation, when the well depth is decreased, manifests in the phase space as tilting and squeezing of the energy surface. This change in geometry of the energy surface due to decreasing well depth leads to more reacting (or escaping) trajectories for a given energy. This is supported by the increase in the area of the rst intersection of the invariant manifolds with an appropriate surface of section. The trajectory diagnostic method of Lagrangian descriptors also identi es these changes in the geometry of the phase space structures. This method also provides an approach for revealing the in uence of potential well depth on high dimensional phase space structures. Detecting the qualitative changes in the geometry of the invariant manifolds using Lagrangian de- scriptors gets us closer to the goal of achieving a complete high-dimensional \phase space tomography" for realistic molecular systems. 34 A) B) FIG. 13: Phase space structures revealed by LDs using  = 30, for the coupled Hamiltonian with parameters  = 0:25, = 2, ! = 1:25 and total energy H = 0:05. A) Coupling strength " = 0:25. On the top/bottom left we show the forward/backward LD and, superimposed, the rst intersection of the stable/unstable manifold with U as a blue/red curve. On the right, the total xp LD is depicted (addition of forward and backward LD) together with the rst intersection of the stable and unstable manifolds, known in chemistry as reactive islands since trajectories inside these regions will escape the potential well through the bottleneck in forward/backward time respectively. The magenta curve represents the energy boundary. B) same analysis using " = 0:125. 35 Acknowledgments We acknowledge the support of EPSRC Grant No. EP/P021123/1 and Oce of Naval Research Grant No. N00014-01-1-0769. The authors would like to acknowledge the EPSRC sponsored Chemistry and Mathematics in Phase Space (CHAMPS) workshop held March 19, 2019 at the University of Bristol where they had the opportunity to discuss with Florentino Borondo about the role of the saddle-node bifurcation in the isomerization of lithium cyanide. Electronic address: vjose.garcia@uah.es Electronic address: s.naik@bristol.ac.uk Electronic address: s.wiggins@bristol.ac.uk T. Komatsuzaki and R. S. Berry, J. Mol. Struct. THEOCHEM 506, 55 (2000). T. Uzer, C. Ja  e, J. Palaci an, P. Yanguas, and S. Wiggins, Nonlinearity 15, 957 (2002). T. Komatsuzaki and R. S. Berry, J. Phys. Chem. A 106, 10945 (2002). R. S. MacKay, Phys. Lett. A 145, 425 (1990). H. Waalkens and S. Wiggins, J. Phys. A: Math. Gen. 37, L435 (2004). E. P. Wigner, Trans. Faraday Soc. 34, 29 (1938). E. P. Wigner, J. Chem. Phys. 7, 646 (1939). H. Waalkens, R. Schubert, and S. Wiggins, Nonlinearity 21, R1 (2008). T. Komatsuzaki and M. Nagoaka, Chem. Phys. Lett. 265, 91 (1997). S. Wiggins, Global bifurcations and chaos: Analytical methods (Springer-Verlag, 1988). S. Wiggins, Physica D 44, 471 (1990). S. Wiggins, Normally hyperbolic invariant manifolds in dynamical systems, vol. 105 (Springer Science & Business Media, 2013). P. Pechukas and F. J. McLa erty, J. Chem. Phys. 58, 1622 (1973), URL http://scitation. aip.org/content/aip/journal/jcp/58/4/10.1063/1.1679404. P. Pechukas and E. Pollak, J. Chem. Phys. 67, 5976 (1977), URL http://scitation.aip. org/content/aip/journal/jcp/67/12/10.1063/1.434777. E. Pollak and P. Pechukas, J. Chem. Phys. 69, 1218 (1978). P. Pechukas and E. Pollak, J. Chem. Phys. 71, 2062 (1979), URL http://scitation.aip. 36 org/content/aip/journal/jcp/71/5/10.1063/1.438575. S. Wiggins, L. Wiesenfeld, C. Ja  e, and T. Uzer, Physical Review Letters 86, 5478 (2001). R. E. Gillilan and G. S. Ezra, J. Chem. Phys. 94, 2648 (1991). T. Komatsuzaki and M. Nagaoka, J. Chem. Phys. 105, 10838 (1996). S. Wiggins, Regular and Chaotic Dynamics 21, 621 (2016). C. B. Li, M. Toda, and T. Komatsuzaki, J. Chem. Phys. 130, 124116 (2009). M. Inarrea, J. F. Palacian, A. I. Pascual, and J. P. Salas, J. Chem. Phys. 135, 014110 (2011). A. Allahem and T. Bartsch, J. Chem. Phys. 137, 214310 (2012), submitted to Arxiv.org. F. A. L. Maugui ere, P. Collins, G. S. Ezra, and S. Wiggins, Int. J. Bifurcat. Chaos 23 (2013). R. S. MacKay and D. C. Strub, Nonlinearity 27, 859 (2014). R. S. MacKay and D. C. Strub, Nonlinearity 28, 4303 (2015), 1501.00266. D. Hathcock and J. P. Sethna, arXiv preprint arXiv:1902.07382 (2019). J. Husson, M. Dogterom, and F. Pincet, The Journal of Chemical Physics 130, 051103 (2009). N. J. Miller and S. W. Shaw, Physical Review E 85, 046202 (2012). C. Herbert and F. Bouchet, Physical Review E 96, 030201 (2017). F. Borondo, A. A. Zembekov, and R. M. Benito, Chemical Physics Letters 246, 421 (1995). F. Borondo, A. Zembekov, and R. Benito, The Journal of chemical physics 105, 5068 (1996). F. Revuelta, R. Benito, and F. Borondo, Physical Review E 99, 032221 (2019). K. A. Mitchell, J. P. Handley, B. Tighe, J. B. Delos, and S. K. Knudson, Chaos: An Interdisci- plinary Journal of Nonlinear Science 13, 880 (2003). K. A. Mitchell, J. P. Handley, J. B. Delos, and S. K. Knudson, Chaos: An Interdisciplinary Journal of Nonlinear Science 13, 892 (2003). K. A. Mitchell, J. P. Handley, B. Tighe, A. Flower, and J. B. Delos, Physical Review Letters 92 (2004). N. De Leon and B. J. Berne, The Journal of Chemical Physics 75, 3495 (1981). K. A. Mitchell, J. P. Handley, B. Tighe, A. Flower, and J. B. Delos, Physical Review A 70 (2004). K. A. Mitchell and B. Ilan, Physical Review A 80 (2009). K. A. Mitchell and J. B. Delos, Physica D: Nonlinear Phenomena 229, 9 (2007). L. Wang, H. F. Yang, X. J. Liu, H. P. Liu, M. S. Zhan, and J. B. Delos, Physical Review A 82 (2010). 37 42 J. A. J. Madrid and A. M. Mancho, Chaos 19, 013111 (2009). G. T. Craven and R. Hernandez, Physical Chemistry Chemical Physics 18, 4008 (2016). G. T. Craven and R. Hernandez, Physical review letters 115, 148301 (2015). G. T. Craven, A. Junginger, and R. Hernandez, Physical Review E 96, 022222 (2017). A. Junginger, L. Duvenbeck, M. Feldmaier, J. Main, G. Wunner, and R. Hernandez, The Journal of chemical physics 147, 064101 (2017). M. Feldmaier, A. Junginger, J. Main, G. Wunner, and R. Hernandez, Chemical Physics Letters 687, 194 (2017). S. Patra and S. Keshavamurthy, Physical Chemistry Chemical Physics 20, 4970 (2018). A. S. Demian and S. Wiggins, International Journal of Bifurcation and Chaos 27, 1750225 (2017). S. Naik, V. J. Garc a-Garrido, and S. Wiggins, arXiv preprint:1903.10264 (Under Review) (2019). S. Naik and S. Wiggins, arXiv preprint:1904.10018 (Under Review) (2019). C. Lopesino, F. Balibrea-Iniesta, V. J. Garc a-Garrido, S. Wiggins, and A. M. Mancho, Inter- national Journal of Bifurcation and Chaos 27, 1730001 (2017). F. Balibrea-Iniesta, C. Lopesino, S. Wiggins, and A. M. Mancho, International Journal of Bifurcation and Chaos 26, 1630036 (2016). V. J. Garcia-Garrido, A. Ramos, A. M. Mancho, J. Coca, and S. Wiggins, Marine Pollution Bulletin. pp. 1{10 (2016). A. G. Ramos, V. J. Garc a-Garrido, A. M. Mancho, S. Wiggins, J. Coca, S. Glenn, O. Scho eld, J. Kohut, D. Aragon, J. Kerfoot, et al., Scient c Reports 4, 4575 (2018). F. Balibrea-Iniesta, J. Xie, V. J. Garc a-Garrido, L. Bertino, A. M. Mancho, and S. Wiggins, Quarterly Journal of the Royal Meteorological Society 145, 76 (2019). S. Wiggins (2017), URL https://figshare.com/articles/Ordinary_Differential_ Equations/5311612. S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, vol. 2 (Springer Science & Business Media, 2003). S. Wiggins, L. Wiesenfeld, C. Ja , and T. Uzer, Physical Review Letters 86, 5478 (2001). W. S. Koon, M. W. Lo, J. E. Marsden, and S. D. Ross, Dynamical systems, the three-body problem and space mission design (Marsden books, 2011), ISBN 978-0-615-24095-4. 38 61 S. Wiggins, Regular and Chaotic Dynamics 21, 621 (2016). K. R. Meyer, G. R. Hall, and D. On, Introduction to Hamiltonian Dynamical Systems and the N-Body Problem (Springer, 2009). D. Beigie and S. Wiggins, Physical Review A 45, 4803 (1992), ISSN 1050-2947, 1094-1622. N. De Leon, M. A. Mehta, and R. Q. Topper, The Journal of Chemical Physics 94, 8310 (1991). N. De Leon, M. A. Mehta, and R. Q. Topper, The Journal of Chemical Physics 94, 8329 (1991). N. De Leon, The Journal of Chemical Physics 96, 285 (1992). A. M. Mancho, S. Wiggins, J. Curbelo, and C. Mendoza, Communications in Nonlinear Science and Numerical Simulation 18, 3530 (2013). I. Mezic and S. Wiggins, Chaos 9, 213 (1999). A. Junginger, G. T. Craven, T. Bartsch, F. Revuelta, F. Borondo, R. Benito, and R. Hernandez, Physical Chemistry Chemical Physics 18, 30270 (2016). V. J. Garc a-Garrido, J. Curbelo, A. M. Mancho, S. Wiggins, and C. R. Mechoso, Regular and Chaotic Dynamics 23, 551 (2018). S. Wiggins, Normally hyperbolic invariant manifolds in dynamical systems. (Springer-Verlag, New York, 2014), ISBN 978-1-4612-8734-6. S. Naik and S. D. Ross, Communications in Nonlinear Science and Numerical Simulation 47, 48 (2017). S. D. Ross, A. E. BozorgMagham, S. Naik, and L. N. Virgin, Physical Review E 98, 052214 (2018). T. S. Parker and L. O. Chua, Practical Numerical Algorithms for Chaotic Systems (Springer- Verlag New York, Inc., New York, NY, USA, 1989), ISBN 0-387-96689-7. J. E. Marsden and S. D. Ross, Bulletin of the American Mathematical Society 43, 43 (2006). Appendix A: Lagrangian Descriptors The computational tool that we use in this work to explore the template of geometrical structures governing phase space transport is the method of Lagrangian descriptors (LDs). This mathematical technique is a trajectory-based diagnostic that was originally developed 42,67 in the context of Lagrangian transport studies in geophysical uid dynamics . The funda- mental idea behind this methodology is to integrate a positive scalar function along particle 39 trajectories of a dynamical system of general time-dependence in the form: dx = v(x; t) ; x 2 R ; t 2 R ; (A1) dt where the vector eld v(x; t) 2 C (r  1) in x and continuous in time. 42,67 Lagrangian Descriptors were rst introduced in by means of a scalar function, referred to as the function M , to identify distinguished hyperbolic trajectories, i.e. moving saddles, of a dynamical system with general time dependence. In this original approach, the function M was based on the computation of the arclength of a trajectory starting at an initial condition x(t ) = x as it evolves forward and backward for a speci ed time period  > 0. In this 0 0 context, LDs were de ned as: t + M (x ; t ;  ) = kv(x(t; x ); t)k dt ; (A2) 0 0 0 where kk stands for the Euclidean distance and v(x; t) is the vector eld of the dynamical system de ned in Eq. (A1). The connection between the function M and invariant manifolds in phase space has been primarily demonstrated through numerical simulation experiments. Hyperbolic trajectories and their stable and unstable manifolds are revealed by the function M through sharp changes, which we call singular features, in the values of the M eld where the gradient becomes very large and changes abruptly. Moreover, it is important to highlight that if the function M is broken into forward and backward integration: (f) (b) M (x ; t ;  ) = M (x ; t ;  ) + M (x ; t ;  ) 0 0 0 0 0 0 where we have that: Z Z t + t 0 0 (f) (b) M = kv(x(t; x ); t)k dt ; M = kv(x(t; x ); t)k dt 0 0 t t 0 0 (f) (b) then M detects the stable manifolds in phase space and M does the same for unstable manifolds. Other structures such as tori-like invariant manifolds are highlighted due to the relation of the time averages of the function M with the Ergodic Partition Theory . This can be done by de ning the time average: t + M (x ; t ;  ) = kv(x(t; x ); t)k dt ; (A3) 0 0 0 and analyzing its convergence as  ! 1. A detailed description of the detection and visu- 52,67 alization of phase space structures with the function M can be found in . It is important 40 to note here that in the context of chemical reaction dynamics, this de nition of LDs would measure phase space arclength of trajectories. In particular, in the Transition State Theory 44{46,69 literature , con guration space arclength has been used for the computation of LDs. Recently, a rigorous mathematical foundation for Lagrangian descriptors has been estab- lished in by means of introducing an alternative de nition based on the p-norm of the components of the vector eld that de ne the dynamical system in Eq. (A1). Consider the scalar function: t + M (x ; t ;  ) = jv (x(t; x ); t)j dt ; p 2 (0; 1] (A4) p 0 0 i 0 i=1 For this LD one can mathematically prove in certain model problems that hyperbolic points and their stable and unstable manifolds are detected as singularities of the M eld, that is, points in which the function is non-di erentiable. Therefore, this alternative de nition provides a characterization for the concept of singular features. Moreover, tori-like invariant structures are also related to time averages of M by means of the Ergodic Partition Theory. An important aspect to highlight from the alternative de nition given in Eq. (A4) is that it allows to decompose the phase space analysis by separating the integral into the di erent DoFs of the system under study, making it possible to isolate and assess their elliptic and hyperbolic dynamical contributions separately. This property has been shown to be rele- vant for the detection of unstable periodic orbits in Hamiltonian system with two DoF for the classical H enon-Heiles system . Furthermore, this property has been used recently to address the Barbanis system with three DoF , and also to mathematically prove that for the normal form of a Hamiltonian system with three DoFs with an rank-1 saddle, the NHIM and its stable and unstable manifolds are located at the singularities, which are also local minima, of the LD eld. Hence, it is shown that if we decompose Eq. (A4) into forward and backward integration as we also did for Eq. (A2), for a suciently large integration time we have: u (b) s (f) W (x ; t ) = argmin M (x ; t ;  ) ; W (x ; t ) = argmin M (x ; t ;  ) (A5) 0 0 0 0 0 0 0 0 p p u s where W and W are, respectively, the unstable and stable manifolds calculated at time t and argmin denotes the phase space coordinates x that minimize the function M . In 0 0 p addition, the NHIM at time t can be calculated as the intersection of the stable and unstable manifolds: u s N (x ; t ) = W (x ; t )\W (x ; t ) = argmin M (x ; t ;  ) (A6) 0 0 0 0 0 0 p 0 0 41 At this stage, we would like to point out other relevant properties reated to the method of Lagrangian descriptors. First, the integration parameter  that appears in the de nition of both M and M plays a critical role in the identi cation of the underlying dynamical structures in phase space. For applications, the value of  is chosen so that the dynamical history of particle trajectories covers all signi cant timescales for the problem under study. Therefore, there is no general rule on how to choose the appropriate value for  in order to achieve the successful visualization of phase space structures. Consequently, all dynamical systems have to be investigated on a case by case basis by means of trial and error simula- tions. Observe that small values of  would yield blurry phase space structures, which are barely recognizable, since this regime resembles the Eulerian (instantaneous) description of the ow. On the other hand, large  values result in a richer and more complex geometrical description of phase space structures, because LDs resolve them in great detail, making the task of interpreting transport from the obained picture dicult. So there is always a compromise between these two situations. Another interesting feature of LDs is that its computational implementation is straightforward, even for high dimensions, and that the method can be parallelized easily to run on high performance computers, because it deals with initial conditions separately. But probably the most important capability of LDs in all its variants is that it provides us with a high-resolution methodology to explore and visualize high-dimensional phase space dynamics. The successful results obtained by applying this tool to uncover the phase space geometry are in part vindicated from its naive approach of emphasizing initial conditions of particle trajectories rather than focusing on their precise location and long-term evolu- tion. As a result of this crucial point, it o ers tremendous advantages for the analysis of high-dimensional phase space, where the evolution of ensembles of initial conditions may yield trajectories that become disperse and \lost" with respect to each other, making the interpretation of phase space structures problematic and challenging. The goal of identifying phase space structures for high-dimensional systems with LDs is achieved by analyzing the 49{51,70 behavior of initial conditions over low-dimensional phase space slices . In fact, any low-dimensional surface can be selected as a probe, which can be sampled with arbitrary high resolution by de ning an adequate grid of initial conditions. Therefore, no resolution will be lost in the analysis, as the trajectories corresponding to these initial conditions evolve in time, since phase space structure is encoded in the initial conditions of the trajectories 42 themselves. 46,51 In recent studies it has been highlighted that computing xed-time LDs, that is, integrating all initial conditions chosen on a phase space surface for the same integration time , could give rise to some issues. First, it can obscure the detection of the NHIM, which is a crucial step, for instance in Transition State Theory, in order to determine chemical reaction rates by analyzing ux across the dividing surface constructed from the NHIM. This diculty appears as a result of bounded trajectories recrossing the barrier region that surrounds the NHIM, due to the Poincar e recurrence theorem. Consequently, multiple minima and singularities occur in the LD plots, which makes the location of the true NHIM trajectories a challenging task. Another issue, which takes place in the saddle-node bifurcation problem that we study in this paper, is that some of the trajectories that escape the potential well of the PES can go to in nity in nite time. The trajectories that show this behavior will give NaN values in the LD scalar eld, hiding some regions of the phase space, and therefore obscuring the detection of invariant manifolds. In order to illustrate this problem, we have calculated the xed-time M function given in Eq. (A4) with p = 1=2 for the 1 DoF Hamiltonian described by Eq. (11) using the model parameters  = 0:25 and = 2. The results are shown in Fig. 14 for di erent values of the integration time. In Fig. 14A, which corresponds to the value  = 3, we can clearly see how LDs detect the hyperbolic xed point at the origin and the elliptic point at (1=2; 0). Moreover, the stable and unstable manifolds that originate from the saddle point are barely visible, as well as the homoclinic orbit. This is a consequence of the integration time being small. However, at regions on the left part of the phase space, corresponding to initial conditions that have escaped to in nity in nite time, start to appear as a result of NaN values in the LD. This e ect is emphasized further in Fig. 14B, where we have calculated xed-time LDs using  = 5. As we increase the integration time more trajectories escape to in nity and, consequently, a larger region of phase space disappears in the LD picture. In order to circumvent these issues, we will apply in this work the approach that has 46,51 been recently adopted in the literature known as variable time Lagrangian Descriptors. In this methodology, LDs are calculated for a given initial condition until the trajectory leaves a certain barrier (or saddle) region R de ned in the phase space. Therefore, the total integration time in this strategy depends on the initial condition itself, that is  (x ). In this 43 A) B) FIG. 14: Fixed-time M function calculated with p = 1=2 for the 1 DoF Hamiltonian described by Eq. (11) using the model parameters  = 0:25 and = 2. Panel A corresponds to  = 3 and B is calculated for  = 5. formulation, the p-norm de nition of LDs has the form: Z + t + x X M (x ; t ;  ) = jv (x(t; x ); t)j dt ; p 2 (0; 1] : (A7) p 0 0 i 0 0 x 0 i=1 and, for a xed integration time  , the total integration time is de ned as: n o = min  ; t (A8) 0 x(t ;x )2R = where t and t are the times for which the trajecory leaves the barrier region R in forward and backward time, respectively. Since the origin is an rank-1 sddle for the Hamiltonian models that we are using to analyze saddle-node bifurcations, we choose for the barrier region: R = x = (q; x; p; p ) 2 R jqj < 15 ; jpj < 15 (A9) To conclude, it is important to point out here that if the selected barrier region is large enough, the variable time LD de nition given above in Eq. (A7) will approach the xed- time LD de nition in Eq. (A7). Normally hyperbolic invariant manifolds and their stable and unstable manifolds will be captured by the phase space points for which the LD is non-di erentiable and tori-like structures can be determined from time averages of LDs. Moreover, if the barrier region is very small, the NHIM and its invariant stable and unstable manifolds will appear as local maxima in the LD eld, while for large barrier regions the local minimum behavior given in Eqs. (A5) and (A6) is recovered. Consequently, the variable 44 integration time LD provides us with a suitable methodology to study the phase space geometrical structures that characterize the type of saddle-node problems that we discuss in this work, since it avoids the issue of trahectories escaping from the potential well of the PES to in nity in nite time. A detailed analysis on the theoretical background on the variable integration time Lagrangian Descriptors technique will be carried out in future work. Appendix B: Visualization of phase space structure for the quadratic normal form Hamiltonian Consider the normal form for a quadratic Hamiltonian system with two DoF given by 2 2 2 2 H (q ; q ; p ; p ) = (p q ) + (q + p ) ;  ; ! > 0 (B1) 1 1 1 2 1 1 2 2 2 2 | {z } | {z } H H r b where Hamilton's equations are @H >q_ = =  p 1 1 > @p > @H p _ = =  q < 1 1 @q (B2) @H q_ = = ! p > 2 2 > @p @H :p _ = = ! q 2 2 @q The equilibirum point is located at (0; 0; 0; 0) and has zero energy. It is trivial to check that the eigenvalues of the linearized system about the equilibrium point are  and ! i, and hence the equilibrium point is of saddlecenter stability type, which is known as a rank-1 saddle. In this form, the Hamiltonian (B1) is decoupled into the \reactive" mode given by H and the \bath" mode given by H . For this reason, it is known as a separable quadratic r b Hamiltonian (SQH). This representation allows us to address the phase space structures and discuss the distribution of the total energy of the system between the two modes in uncoupled coordinates. In this form, a chemical reaction is said to have occurred when the q coordinate of a trajectory changes sign and thus, an isoenergetic, H = H dividing 1 0 surface (DS) can be de ned by the q = 0 hypersurface. The constant energy de nes a three-dimensional energy surface in the four dimensional phase space given by 2 2 2 2 p q + p + q = H + H = H > 0 ; H > 0 ; H  0 (B3) r b 0 r b 1 1 2 2 2 2 45 The dividing surface, q = 0, for a constant energy is 2 2 2 p + p + q = H + H = H > 0 ; H > 0 ; H  0 (B4) r b 0 r b 1 2 2 2 2 which is a two dimensional surface, and has the geometry S , that is, a 2-sphere on the three dimensional energy surface. Thus, it is codimension-1 and partitions the energy surface into reactant p q > 0 and product p q < 0 regions by the forward and backward \reaction" 1 1 1 1 dividing surfaces as shown in Ref. , which are given by 2 ! 2 2 (B5) p =  H (p + q ) ; forward/backward DS 1 0 2 2 The forward and backward DS join at p = 0 giving n o 4 2 2 N (H ) = (q ; q ; p ; p ) 2 R j q = p = 0 ; p + q = H  0 ; NHIM (B6) 0 1 2 1 2 1 1 0 2 2 which is of geometry S , that is a circle centered at the origin with radius 2H =! in the (q ; p ) plane. This is a normally hyperbolic invariant manifold (NHIM) associated with the 2 2 rank-1 saddle and parametrized by total energy H (q ; q ; p ; p ) = H . Invariance follows 1 2 1 2 0 from the vector eld (B2), since when we have that q = p = 0, this gives q_ = p _ = 0. 1 1 1 1 Thus q and p always remain zero, and trajectories with these initial conditions remain on 1 1 the NHIM, that is, q = p = 0 is invariant. It is normally hyperbolic since the directions 1 1 normal to the NHIM, that is, the (q ; p ) plane, have linear saddle-like dynamics. For a 1 1 two DoF system, this NHIM is more commonly referred to in the literature as an unstable periodic orbit and is shown in Fig. 15. In order to understand the relationship between the NHIM and the rank-1 saddle point, we note that for H = H = 0 the NHIM reduces to the point (q ; q ; p ; p ) = (0; 0; 0; 0), r b 1 2 1 2 which is the rank-1 saddle point on the energy surface H = 0. Moreover, recall that this rank-1 saddle equilibrium point is a con guration space concept and is located on the potential energy surface. Therefore, as the total energy of the system is increased from 0, with the bath mode energy H increasing from zero, the NHIM \grows" from the index- 1 saddle point on the zero energy surface into an invariant 1-sphere. This shows how the \in uence" of the rank-1 saddle point is carried to higher energy sufaces on which the saddle point does not exist. The stable and unstable manifolds of the NHIM (B6) are given by u 4 2 2 W = (q ; q ; p ; p ) 2 R j q = p ; p + q = H > 0 (B7) 1 2 1 2 1 1 b 2 2 s 4 2 2 W = (q ; q ; p ; p ) 2 R j q = p ; p + q = H > 0 (B8) 1 2 1 2 1 1 b 2 2 46 1 which are two-dimensional surfaces and have the geometry of RS for a xed energy. Thus, the codimension-1 geometry of the manifolds partition the phase space into \reactive" and \non-reactive" trajectories as shown in Fig. 15. A) B) C) FIG. 15: Phase space structures in the neighborhood of the bottleneck for the uncoupled (" = 0) Hamiltonian system (19). A) Linearized dynamics about the rank-1 saddle in the saddle space corresponding to the reactive DoF; B) Dynamics in the center space associated to the harmonic oscillator DoF; C) Description of the phase space bottleneck region obtained for an energy of the system above that of the rank-1 saddle (the barrier of the PES). The stable and unstable manifolds of the NHIM act as conduits connecting reactants (well region) and products (escape to in nity) and characterize reaction dynamics. 47 Appendix C: Computation of NHIM and its invariant manifolds for the 2 DoF system In this aapendix we describe the steps followed in order to calculate the NHIM and its stable and unstable invariant manifolds associated with the rank-1 saddle located at the origin for the Hamiltonian system with 2 DoF given by Eq. (19), which models the saddle- node bifurcation phenomena in phase space. Step 1: Select an excess energy above the critical value. The Lyapunov subcenter theorem tells us that when the energy of the Hamiltonian system is raised above that of the equilibrium point at the origin of saddlecenter (rank-1 saddle) stability, which is known 2N3 as critical energy, a family of NHIM with geometry S bifurcates, where N indicates the number of DoF. In our problem, N = 2 so the topology of the NHIM is S , that is, the NHIM is an unstable periodic orbit (UPO). So we pick a value for the total energy H greater than the critical value H = 0, which gives an excess energy H = H H = H . c 0 c 0 This total energy is also the energy of trajectories on the invariant manifolds which partition the phase space into nonreactive and reactive trajectories at the same energy. The excess energy can be an arbitrary value, up to the value at which the NHIM bifurcates. Step 2: Obtain the NHIM at the selected excess energy. As we have discussed, for the 2 DoF problem the NHIM is an UPO which touches the equipotential contour corresponding to the selected energy at the bottleneck region. However, due to the saddle- like dynamics in the transverse directions, any numerical error in the computation of the periodic orbit will get exponentially ampli ed with each time step, which will eventually destroy the periodic nature of the trajectory, since it will fail to come back to its starting point. In order to generate the desired unstable periodic orbit corresponding to a selected excess energy, we consider a procedure that starts with a small initial condition (\seed") obtained from the linearized equations of motion near the rank-1 saddle, and uses di erential cor- 60,72,73 rection and numerical continuation on that initial guess. The result is an unstable periodic orbit at target energy H of period T which will be close to 2=! , where ! i 0 " " is the pair of imaginary eigenvalues of the linearization about the rank-1 saddle point. In order to choose an initial guess for the search of the UPO we can use the linearized equations of motion about the rank-1 saddle equilibrium point x = 0. The Jacobian matrix at the 48 origin is given by Eq. (26) and the general solution to the linearized dynamical system can be written as: e  t  t i! t " " " x(t) = x + C e v + C e v + 2Re e w (C1) 1 + 2 + where C ; C 2 R and  =  +  i 2 C. The real eigenvalues  with corresponding 1 2 1 2 " eigenvectors v ; v are described in Eqs. (31) and (34) and the pair of complex eigenvalues ! i with corresponding eigenvectors w ; w are described in Eqs. (31) and (35). The idea " + is to use the complex eigenvalues and their corresponding eigenvectors to obtain a starting guess as an initial condition to initialize the method that searches for the UPO. To do so, we choose a small amplitude, A  2  10 , periodic orbit in the center manifold of the linearized system by selecting in Eq. (C1) the values  = A =2 (this eliminates the factor 2 in the formula), C = C = 0 and t = 0. Thus, the initial guess is: 1 2 A " x  = (q ; x ; 0; 0) = x + 2Re(w ) = ;A ; 0; 0 (C2) 0;g 0;g 0;g + x " ( + ! ) 0 " which has a period of T = 2=! . 0;g " Step 3: Di erential correction of the initial guess. The di erential correction procedure that we apply in this step to the initial guess will only yield a good approximation to the true UPO of the nonlinear Hamiltonian system whenever A  1, which corresponds to a much smaller excess energy H than the one we originally selected in step 1. The reason for this is that the construction of the initial guess is based on the linear approximation near the rank-1 saddle equilibrium point. The convergence criterion that we use is based on the basic property that a periodic orbit returns to the starting point after a given period T , which is in fact the period of the PO. If x  = x  (0) is a true initial condition on the PO 0 po x of period T , the convergence is checked using the condition: po kx  (T ) x  (0)k <  (C3) po po for some tolerance   1. In this approach, we hold the con guration coordinate q constant, while applying correction to the x con guration coordinate of the initial guess. While the p momentum coordinate is used as a stopping criterion for the di erential correction procedure, the p momentum coordinate is used as a terminating event (crossing the p = 0 plane in phase space) for the integration. It is to be noted that this combination of coordinates is suitable for the structure of the initial guess at hand, so for other problems or other forms of the 49 initial guess, would require some permutation of the phase space coordinates to achieve a stable implementation. Let us denote the ow map of a di erential equation x _ = f (x) with initial condition x(t ) = x by (t; x ). Thus, the displacement of a reference trajectory x (t) after a time t 0 0 0 becomes: x (t + t) = (t + t; x  + x  ) (t; x  ) : (C4) 0 0 0 Thus, a Taylor expansion of the displacement after t + t gives: 1 1 @(t ; x  ) @(t ; x  ) 1 0 1 0 x (t + t ) = x  + t + h:o:t (C5) 1 1 0 1 @x @t 0 1 where the rst term on the right hand side is the state transition matrix, (t ; t ), evaluated 1 0 along the reference trajectory initialized at t . The state transition matrix along a trajec- tory is obtained from the numerical solution of the variational equations along with the Hamilton's equation of motion . Suppose that we want to land at a point x (this would be the starting initial condition for a periodic orbit), after an integration time interval t , and starting from the initial guess x  , then we have: 0;g x (t ) = (t ; x  ) = x  = x x  (C6) 1 1 0;g 1 d 1 where the error x  is the applied rst order correction obtained from the state transition matrix evaluated along the trajectory with initial condition x  and integrated for t time 0;g 1 units. For the rank-1 saddle equilibrium point under consideration, we initialize the guess as obtained in Eq. (C2) with x  . Using numerical integration, the initial condition is 0;g integrated until a p = 0 event occurs with a high speci ed tolerance (typically 10 ). This results in x(t ), which for the guess initial condition denotes the half-period location t = T =2. Then, we evaluate the state transition matrix (t ; 0) for the trajectory obtained 1 0;g 1 from the guess intial condition. Now, this is used to correct the initial value of x while 0;g keeping q constant and iterating until p = 0. Since the x con guration coordinate is kept 0;g x constant, the rst order correction is given by: p =  x + p _ t + h:o:t 1 32 0 1 1 p =  x + p _ t + h:o:t x 42 0 x 1 1 1 50 th where  is the (i; j) entry of (t ; 0) and the acceleration terms come from the equations ij 1 of motion evaluated at the crossing t = t when p = p = 0. Thus, we obtain the rst 1 1 1 order correction x as: p _ x    p ; x ! x x (C7) 0 42 32 x 0 0 0 p _ which is iterated until jp j = jp j <  for some tolerance , since we want the nal x x 1 1 point of the periodic orbit to be of the form x  = (q ; x ; 0; 0). We remark that in all t 1 1 this process, di erential correction assumes that the guess periodic orbit has a small error (for example in this system, of the order of 10 ) and can be corrected using rst order form of the correction terms. If, however, large a corrective step is used, the half-orbit overshoots between successive steps leading to failure in converging to a closed orbit. Once the appropriate conditions are chosen, di erential correction generates a family of periodic orbits and takes 2-3 iterations per unstable periodic orbit. Step 4: Numerical continuation to the UPO at the selected excess energy. The procedure described above yields an initial condition for an unstable periodic orbit from an initial guess. Since our initial guess came from the linearization near the rank-1 saddle equilibrium pointgiven by (C1), we can use this procedure for small amplitudes of order 2 10 . We remark that this procedure is based on computations presented for this problem and will vary for a di erent nonlinear system. This small amplitude corresponds to small excess energy, typically of the order 10 , and to obtain the unstable periodic orbit of arbitrarily large amplitude, we resort to numerical continuation for generating a family of periodic orbits that reach the selected excess energy. This procedure starts with the initial conditions for two nearby unstable periodic orbits of small amplitude to obtain an initial guess for the next periodic orbit. The initial guess obtained by a simple extrapolation can then be corrected using di erential correction. To this end, we proceed as follows. Suppose we nd two nearby small amplitude unstable (1) (2) periodic orbits with initial conditions x and x , correct to within the tolerance d tol 0 0 obtained using the di erential correction procedure described above. We can then generate a family of periodic orbits with increasing amplitudes around x  as: eq (2) (1) = x  x  = (q ; x ; 0; 0) (C8) 0 0 0 0 51 (3) A linear extrapolation to an initial guess of slightly larger amplitude, x is given by: (3) (2) (2) (2) (3) (3) x  = x  +  = (q + q ; x + x ; 0; 0) = (q ; x ; 0; 0) (C9) 0 0 0;g 0 0 0 0 0 Thus, we can use di erential correction on this guess initial condition to compute an accurate (3) (3) solution x  from the initial guess x  and repeat as an iterative step to generate a family 0 0;g of unstable periodic orbits. Next, to compute an unstable periodic orbit at the selected excess energy, we track (k) the energy of each unstable periodic orbit in the family until we have two solutions, x (k+1) and x  , whose energy brackets the selected excess energy H . Then, we can resort to combining a bisection type method with di erential correction on the two periodic orbits until we converge to the desired periodic orbit to within a speci ed tolerance. Thus, the result is an unstable periodic orbit at the selected energy H and is speci ed by the initial condition x  (0) and time period T . po po Step 5: Computation of invariant manifolds of the NHIM. We nd the global ap- proximation to the unstable and stable manifolds of the periodic orbit from the eigenvectors of the monodromy matrix. The local linear approximation of the unstable (or stable) man- ifolds (initial conditions displaced along the saddle space eigenvectors) is integrated using the full nonlinear equations of motion to produce the global approximation of the unstable (or stable) manifolds. This procedure is known as globalization of the manifolds and we proceed as follows. First, the state transition matrix (t) along the periodic orbit with initial condition X can be obtained by numerical integration of the variational equations along with the equations of motion from t = 0 to t = T . This is gives the monodromy matrix M = (T ) po po and its eigenvalues are obtained. For Hamiltonian systems , it is known that the eigenvalues of M are of the form: > 1;  = ;  =  = 1 (C10) 1 2 3 4 The eigenvector associated with eigenvalue  is in the unstable direction, the eigenvector associated with eigenvalue  is in the stable direction. Let e (x  (0)) denote the normalized 2 po stable eigenvector, and e (x  (0)) denote the normalized unstable eigenvector. We can po compute the invariant manifolds by initializing along these eigenvectors as: s=u s=u X (x (0)) = x (0) + e (x (0)) (C11) po po po 52 for the stable/unstable manifold of the desired periodic orbit. Here the small displacement from X is denoted by  > 0 and its magnitude is taken to be small enough for the validity of the linearization, yet not so small that the time of ight becomes too large due to asymptotic nature of the stable and unstable manifolds. It has been suggested that typical values of around 10 correspond to nondimensional position displacements of magnitude around 10 . By numerically integrating the unstable vector forwards in time, using both  and , for the forward and backward branches respectively, we generate trajectories shadowing the two u u branches, W and W . Similarly, by integrating the stable vector backwards in time, using both  and , for forward and backward branches respectively, we generate trajectories shadowing the stable manifold, W . For the manifold at X(t), one can simply use the +; state transition matrix to transport the eigenvectors from X to X(t) u=s u=s X (X(t)) = (t; 0)X (X ) (C12) It is to be noted that since the state transition matrix does not preserve the norm, and hence the resulting vector must be normalized. The globalized invariant manifolds associated with rank-1 saddles are known as Conley-McGehee tubes . These tubes form the impenetrable phase space conduits (codimension-1 barriers) for the trajectories to react by crossing the q = 0 dividing surface. In summary, the computation of invariant manifolds of the unstable periodic orbit associ- ated with the rank-1 saddle begins with the linearized equations of motion. This is obtained after a coordinate transformation to the rank-1 saddle equilibrium point and a Taylor expan- sion of the equations of motion. Keeping the rst order terms in this expansion, we obtain the eigenvalues and eigenvectors of the linearized system. The eigenvectors corresponding to the center subspace provide the starting guess for computing the unstable periodic orbits of small excess energy, H << 1. This iterative procedure performs a small correction to the starting guess for initial condition based on the terminal condition of the periodic orbit until a desired tolerance is satis ed. This procedure is known as di erential correction and generates initial condition for an unstable periodic orbit at small excess energy. Next, a numerical continuation procedure is adopted to follow the small energy (small amplitude) periodic orbit out to high excess energy. Once the unstable periodic orbit is obtained, the globalization of its invariant manifolds is done using the initial conditions along the eigen- 53 vectors of the state transition matrix computed along the unstable periodic orbit. The result of the steps described above is shown in Fig. 9. http://www.deepdyve.com/assets/images/DeepDyve-Logo-lg.png Nonlinear Sciences arXiv (Cornell University)

Tilting and Squeezing: Phase space geometry of Hamiltonian saddle-node bifurcation and its influence on chemical reaction dynamics

Loading next page...
 
/lp/arxiv-cornell-university/tilting-and-squeezing-phase-space-geometry-of-hamiltonian-saddle-node-cc0SrVlxgI
ISSN
0218-1274
eISSN
ARCH-3338
DOI
10.1142/S0218127420300086
Publisher site
See Article on Publisher Site

Abstract

In this article we present the in uence of a Hamiltonian saddle-node bifurcation on the high- dimensional phase space structures that mediate reaction dynamics. To achieve this goal, we identify the phase space invariant manifolds using Lagrangian descriptors, which is a trajectory- based diagnostic suitable for the construction of a complete \phase space tomography" by means of analyzing dynamics on low-dimensional slices. First, we build a Hamiltonian system with one degree-of-freedom (DoF) that models reaction, and study the e ect of adding a parameter to the potential energy function that controls the depth of the well. Then, we extend this framework to a saddle-node bifurcation for a two DoF Hamiltonian, constructed by coupling a harmonic oscillator, i.e. a bath mode, to the other reactive DoF in the system. For this problem, we describe the phase space structures associated with the rank-1 saddle equilibrium point in the bottleneck region, which is a Normally Hyperbolic Invariant Manifold (NHIM) and its stable and unstable manifolds. Finally, we address the qualitative changes in the reaction dynamics of the Hamiltonian system due to changes in the well depth of the potential energy surface that gives rise to the saddle-node bifurcation. arXiv:1907.03322v1 [nlin.CD] 7 Jul 2019 Keywords: Saddle-Node bifurcation, Hamiltonian systems, Phase space structure, La- grangian descriptors, Chemical reaction dynamics. Contents I. Introduction 2 II. Hamiltonian Saddle-Node Bifurcation Models 7 A. One Degree-of-Freedom Hamiltonian 7 B. Two degrees-of-freedom Hamiltonian 15 III. Results and Discussion: In uence of well-depth on reaction dynamics 21 A. One Degree-of-Freedom Hamiltonian 21 B. Two degree-of-freedom Hamiltonian 24 IV. Conclusions 34 Acknowledgments 36 References 36 A. Lagrangian Descriptors 39 B. Visualization of phase space structure for the quadratic normal form Hamiltonian 45 C. Computation of NHIM and its invariant manifolds for the 2 DoF system 48 I. INTRODUCTION Background. Molecular reaction dynamics is concerned with the breaking and forming of bonds between the atoms that make up a molecule. The description of such mechanisms requires analyzing how the bonds change in time, that is, it is a dynamical mechanism re- quiring a phase space description. The phase space will have coordinates that correspond to the bond con gurations with a corresponding canonical momentum for each bond con gura- tion coordinate. Thus, the dynamics of chemical bonds can be formulated in the language of 2 Hamiltonian dynamics and its resulting phase space geometry. In recent years there has been signi cant developments in the mathematical description and analysis of chemical reaction 1{3 dynamics in phase space . Our objective is to describe the change in a bond that signals the occurrence of a chemical reaction in terms of a unique characteristic of a trajectory of Hamilton's equations. This characteristic requires an understanding of the geometry of the phase space of the molecule in a way that enables us to divide an appropriate volume of phase space into a region corresponding to \reactants" and a region corresponding to \products". The passage from reactants to products, that is \reaction", occurs when a trajectory crosses the \dividing surface" (DS) between the reactants and products. In this framework for understanding chemical reactions, the ux through such a dividing surface would be related to the reaction rate, and therefore the construction of DS between such regions is of interest in reaction dynamics. The description of the regions of reactants and products can often be inferred from the nature of the development of the coordinates used in the mathematical model of the chemical reaction. Once the model is developed, the DS must be constructed in the context of this model. In relating the ux through the DS to a reaction rate it is desirable that trajectories crossing the DS from the reactant side proceed to the product side before possible recrossing the DS back to the reactant region. Thus, we require the 4,5 DS to have the \no-recrossing" property . Hamiltonian dynamics conserves energy and therefore the trajectories evolve on a xed energy surface of one less dimension (referred to as codimension one) than the phase space. The DS is required to be codimension one in the energy surface. Hence, we require a DS between reactants and products to be codimension one in the energy surface and to have the no-recrossing property. We note that Wigner had 6,7 already described these properties for a DS in phase space much earlier , and a review of classical and quantum versions of DS constructed in phase space can be found in . Traditionally, the construction of DS was initially focused on critical points of the po- tential energy surface (PES), that is, in the con guration space describing the molecular system. Critical points on the PES do have signi cance in phase space; they are the equilib- rium points for zero momentum. But they continue to have in uence for nonzero momentum for a range of energies above the energy of the equilibrium point. The precise manner of this 1,5,9 dynamical in uence has only been understood recently and we will describe this shortly . The construction of a DS separating the phase space into two parts, reactants and products, 3 has been a focus from the dynamical systems point of view in recent years. However, the lack of a rm theoretical basis for the construction of such surfaces for molecular systems with three and more degree-of-freedom (DoF) has until recently been a major obstacle in the development of the theory. In phase space, that is for nonzero momentum, the role of the saddle point is played by an invariant manifold of saddle stability type, the nor- 10{12 mally hyperbolic invariant manifold (NHIM) (see x: II B) . In order to fully appreciate the NHIM and its role in reaction rate theory, it is useful to begin with a precursor con- cept { the periodic orbit dividing surface or PODS. For systems with two DoF described by a natural Hamiltonian, kinetic plus potential energy, the problem of constructing the 13{16 DS in phase space was solved during the 1970s by McLa erty, Pechukas and Pollak . They demonstrated that the DS at a speci c energy is related to an invariant phase space structure, an unstable periodic orbit (UPO). The UPO de nes (it is the boundary of ) the bottleneck in phase space through which the reaction occurs and the DS which intersects trajectories evolving from reactants to products can be shown to have the geometry of a 5,17 hemisphere in phase space whose boundary is the unstable PO . The same construction can be carried out for a DS intersecting trajectories crossing from products to reactants and these two hemispheres form a sphere for which the UPO is the equator. Generalisation of this construction of DS to high dimensional systems has been a central question in reaction 2,17 dynamics and has only received a satisfactory answer in recent years . The key diculty concerns the high dimensional analogue of the unstable PO used in the two DoF system for the construction of the DS. This diculty is resolved by considering the NHIM, which has the appropriate dimensionality for anchoring the dividing surface in phase space. Results from dynamical systems theory show that transport in phase space is controlled by high dimensional manifolds, NHIMs, which are the natural generalisation of the UPO of the two DoF case . Normal hyperbolicity of these invariant manifolds means that their stability, in a precise sense, is of saddle type in the transverse direction, which implies that they possess stable and unstable invariant manifolds that are impenetrable barriers and mediate transport in phase space. These invariant manifolds of the NHIM are structurally stable, that is, stable under perturbation . For two DoF systems, the NHIM is an unstable PO, and for an n > 2 DoF system at a xed energy, the NHIM has the topology of a (2n 3)-dimensional sphere and is the equator of a (2n 2)-dimensional sphere which constitutes the DS. The DS can be used to divide the (2n 1)-dimensional energy surface 4 1,3,9,18,19 into two parts, reactants and products . An elementary description of the role of the NHIM in reaction dynamics is given in . Fundamental theorems assure the existence of the phase space structures |NHIM and its invariant manifolds |for a range of energies above that of the saddle . However, the precise extent of this range, as well as the nature and consequences of any bifurcations of the phase space structures that might occur as energy 21{26 is increased, is not known and is a topic of continuing research . Motivation. In this article, we investigate the changes in geometry of the phase space structures in a normal form Hamiltonian that undergoes saddle-node bifurcation. In general, a Hamiltonian system with a saddle equilibrium point whose eigenvalues depend on the potential well depth parameter will exhibit bifurcation when the potential energy barrier height decreases. If the potential energy surface has a single well, then the result is a collision of the stable and saddle equilibrium points as we illustrate in Figs. 3 and 5. This leads to drastic changes in the geometry of phase space structures | NHIM and its invariant manifolds | which can then be used to account for corrections to Kramers' reaction rate as barrier height decreases . Furthermore, quantifying rates of crossing low or vanishing barrier is signi cant for experimental study of single bond dynamics of molecules and control 27{30 of micro and nano-electromechanical devices . We would also like to point out that the questions addressed in this article are further motivated by the work of Borondo and co- 31{33 authors who noted the signi cance of a saddle-node bifurcation in the isomerization of LiCN/LiNC molecule. Conservative dynamics on an open potential well has received considerable attention because the geometry of phase space structures explains the intricate fractal structure of 34{36 ionization rates . Furthermore, the discrepancies in observed and predicted ionization rates in atomic systems has also been explained by accounting for the topology of these phase space structures. These have been connected with the breakdown of the ergodic assumption that is the basis for using ionization and dissociation rate formulae . This rich literature on chaotic escape of electrons from atoms sets a precedent for delineating the changes in phase space structures due to saddle-node bifurcation that can be expected in open potential wells where the eigenvalues of the saddle equilibrium point depend on the 36,38{41 well depth parameter . Approach in this article. We rst develop a Hamiltonian model that exhibits saddle- node bifurcation when the well depth is decreased and obtain a relationship between the 5 parameters of the model for the bifurcation in the 2 DoF system. Then, the NHIM (an unstable PO in the 2 DoF system) is computed using di erential correction and numeri- cal continuation, and its invariant manifolds are computed using globalization. The role of these invariant manifolds in reaction dynamics and the implications of the change in their geometry by varying well depth is discussed. We compare the results with another numerical method that can be used for identifying high dimensional phase space (4 or more dimensions) structures called Lagrangian descriptors (LDs). Lagrangian descriptors is a trajectory di- agnostic method for revealing invariant manifolds that mediate phase space transport. The method was originally developed in the context of Lagrangian transport studies in uid dy- namics , but the wide applicability of the method has recently been recognized as a tool to 33,43{48 construct the dividing surface via identi cation of NHIM . The method is straightfor- ward to implement computationally, and it provides a high resolution method for locating high dimensional invariant manifolds using low dimensional slices of the high dimensional 49{51 phase space . Therefore, this technique provides us with a rst step towards realizing a complete \phase space tomography" of high dimensional invariant manifolds. It also ap- plies to both Hamiltonian and non-Hamiltonian systems as well as to stochastic dynamical systems . Moreover, Lagrangian descriptors can be applied directly to data sets obtained 54{56 from the numerical solution of geophysical models and satelite observations . We provide more details on our use of Lagrangian descriptors in Appendix A. This paper is outlined as follows. Section II is devoted to the description of the Hamilto- nian models with one and two DoF that we will use in our study of saddle-node bifurcation phenomena in phase space. In Section III we explore the implications of these model prob- lems for reaction dynamics, and discuss our results by analyzing the phase space geometrical structures responsible for the escaping dynamics from the PES and their relevance for chem- ical reactions. Finally, in Section IV we summarize the conclusions of this work. The reader can nd all the relevant information related to the method of Lagrangian Descriptors and its implementation details in Appendix A. Furthermore, Appendix C describes the numerical techniques used (di erential correction and manifold globalization) for the computation of the NHIM and its stable and unstable manifolds associated to the rank-1 saddle point in the 2 DoF Hamiltonian system that characterizes the saddle-node phase space bifurcation. 6 II. HAMILTONIAN SADDLE-NODE BIFURCATION MODELS In this section we introduce the normal form Hamiltonian that exhibit saddle-node bi- furcation and their implications for reaction dynamics. First, we describe the normal form Hamiltonian system with one DoF, and parametrize this model by adding a parameter that controls the potential well depth. Next, we extend this model to a two DoF Hamiltonian system by coupling a bath mode (a harmonic oscillator) to the reaction DoF. A. One Degree-of-Freedom Hamiltonian The normal form for the one DoF Hamiltonian that undergoes a saddle-node bifurcation in phase space can be written as 1 1 2 3 H (u; v) = v  u + u ; (1) 2 3 where  2 R is the bifurcation parameter. Hamilton's equations are given by @H > u _ = = v @v ; (2) > @H : 2 v _ = =  u @u and the equilibria of this system are located at (u; v) = ( ; 0). We observe that the dynamical system has two di erent equilibrium points for  > 0, which approach each other as the bifurcation parameter goes to zero. When  = 0, both equilibria \collide" into one equilibrium point at (0; 0). For  < 0, the resulting dynamical system has no equilibria. In order to study the stability of the equilibria, we linearize the vector eld and calculate the Jacobian 0 1 2 2 0 1 @ H @ H B 2 C 0 1 @u@v @v B C @ A J(u; v) = = : (3) @ 2 2 A @ H @ H 2u 0 @u @v@u Evaluating the Jacobian at the equilibrium point ( ; 0), we get 0 1 0 1 @ A J( ; 0) = ; (4) 2  0 7 p which has eigenvalues  4, and hence the equilibrium point ( ; 0) is a saddle. The Jacobian at ( ; 0) is 0 1 0 1 @ A J( ; 0) = ; ; (5) 2  0 which has eigenvalues  4jj i,yielding thst the equilibrium point ( ; 0) is a center. For = 0, zero is the only eigenvalue of the Jacobian matrix, and therefore the linearization does not provide enough information, and one needs to include higher order terms to determine the stability of the equilibrium point (0; 0). We compute LDs with the goal of detecting the invariant manifolds of the dynamical system given in Eq. (2). As shown by the phase portraits in Fig. 1, trajectories of (2) can escape o to in nity in nite time depending on the total energy of the system and the location of the initial conditions. Therefore, this issue needs to be accounted for in the computation of LDs as explained in Appendix A, and the approach we will follow to 44,45 resolve this problem has also been adopted in the chemical reactions literature . In particular we will use the p-norm in the LD de nition (A4) with p = 1=2. Therefore, we evolve initial conditions forward and backward in time for  = 8 and whenever a trajectory leaves the domain de ned by the circle of radius 15 about the origin, we stop the numerical integration of that trajectory. In Fig. 1 we show the LD contour maps for three di erent values of the bifurcation parameter  = 0:25; 0; 0:25 and compare the results obtained by means of LDs with the corresponding phase portraits. We can clearly see in Fig. 1 how the method succesfully recovers all the relevant phase space structures. The stable and unstable manifolds are highlighted by the singularities present in the LD contour map. Furthermore, another interesting aspect to highlight from Fig. 1D is that, for  = 0, the LD values in the neighborhood of the origin seem to indicate that a bifurcation is going to take place in the phase space structure, as one can observe that a `ghost' center structure is about to be created close to the cusp at the origin. a. Fixing the saddle point at the origin. In the Hamiltonian saddle-node model (1), the equilibria move as the bifurcation parameter  is varied. In order to simplify the bifurcation analysis necessary to address the implications for chemical reaction dynamics, we x the saddle point at the origin. Moreover, this form of non-moving saddle will become useful when discussing the 2 DoF model of the saddle-node bifurcation. This transformation will also facilitate the construction of the dividing surface from the NHIM associated with the 8 A) B) C) D) E) F) FIG. 1: Comparison between the phase space portrait of the dynamical system represented by Eq. (2) (left column) and the output of variable time LDs calculated using the p-norm de nition with p = 1=2 and  = 8 (right column). A) and B) correspond to the value of the bifurcation parameter = 0:25; C) and D) for  = 0; D) and E) for  = 0:25. We have marked di erent trajectories in the phase portraits in blue, invariant manifolds in red, and equilibrium points as yellow dots. 9 `non-moving' rank-1 saddle. Let us consider the linear change of coordinates u = q ;  2 R [f0g ; (6) v = p and substitute this transformation into Eq. (2) to yield the following dynamical system in the new coordinates q_ = p ; (7) : 2 p _ = 2  q q with corresponding Hamiltonian 1 p 1 2 2 3 H (q; p) = p  q + q : (8) 2 3 The equilibrium points of Eq. (7) are (0; 0) and (2 ; 0). Their stability is characterized by the eigenvalues of the Jacobian matrix 0 1 2 2 @ H @ H 0 1 B C 0 1 @q@p @p B C @ A J(q; p) = = (9) 2 2 p @ A @ H @ H 2  2q 0 @q @p@q evaluated at each equilibrium point 0 1 0 1 0 1 0 1 @ A @ A J(0; 0) = ; J(2 ; 0) = (10) p p 2  0 2  0 with eigenvalues  4 and  4jj i, respectively. Hence, the origin is a saddle and (2 ; 0) is a center equilibrium point. This is also supported by the result that eigenvalues do not change under translational change of coordinates. In Fig. 2, we show the variable integration time LD contour maps for Eq. (7) at two di erent values of the bifurcation parameter,  = 0 and  = 0:25. These plots illustrate how the stable and unstable manifolds of the saddle equilibrium point are identi ed by the singular features, points where LDs are non-di erentiable, in the LD contour maps. These singular features can be visualized by taking one dimensional slices of the LD contour map, where jump discontinuities mark the initial conditions on an invariant manifold. 10 A) B) C) D) FIG. 2: Comparison between the output of variable time LDs using the -norm de nition with = 1=2 and  = 8 (left column) and the M function value along lines parallel to the q-axis (right column). A) and B) correspond to  = 0 and the line of initial conditions is p = 0:2; C) and D) use  = 0:25 and the lines of initial conditions are p = 0:25 and p = 0:75. b. Introducing a parameter to control \depth" of the PES. The depth of the well on the PES is controlled by the cubic term that appears in the Hamiltonian (8). Therefore, we introduce a parameter that allows us to vary the amplitude of this term, and consequently the strength of the nonlinearity in the vector eld (7). Thus, the Hamiltonian becomes 1 p 2 2 3 H (q; p) = p  q + q  T (p) + V (q) ; (11) 2 3 11 + + where  2 R [f0g and 2 R are the two parameters, T is the kinetic energy of the DoF and V its potential energy. Hamilton's equations are given by @H q_ = = p @p : (12) > @H p _ = = 2  q q @q The equilibria are (0; 0) and 2 = ; 0 , and the Jacobian of the vector eld is given by 0 1 0 1 @ A J (q; p) = : (13) 2  2 q 0 We determine the stability of the equilibria by evaluating the Jacobian which gives 0 1 0 1 0 1 0 1 @ A @ A J (0; 0) = ; J (2 = ; 0) = : (14) p p 2  0 2  0 The eigenvalues of the Jacobian at the equilibrium point (0; 0) are  4 with correspond- ing eigenvectors (1; 4). The eigenvalues for the Jacobian at the equilibrium point p p 4 4 2 = ; 0 are  4 i with eigenvectors (1; 4 i). Clearly, the eigenvalues of the Hamil- tonian (8) are perserved, and (0; 0) is a saddle, while 2 = ; 0 is a center. We note here that the eigenvalues and eigenvectors of the Jacobian evaluated at the equilibrium points only depend on . \Depth" of the well on the PES, referred to as well-depth from here on, is determined by the di erence between the potential energy of the saddle and the potential energy of the center (minimum of the well) equilibrium points. The potential energy function is given by 2 3 V (q) =  q + q ; (15) and this di erence is given by F  V (0) V = : (16) Hence, for a xed , the well-depth is increased by decreasing . In fact, if we denote D = 2 = as the distance between the saddle and the center equilibrium point, we get F = D ; (17) 12 therefore, as increases, the well-depth approaches zero faster than the distance between the saddle and the center equlibrium points. Furthermore, the rate at which the well-depth changes as we change the distance between equilibria is given by dF 1 = D =  D : (18) dD 3 3 which is proportional to the product of the square of the eigefrequency associated to the saddle equilibrium and the distance between equilibria. In Fig. 3(a) we illustrate the well-depth together with the distance between the saddle and the center. In Fig. 3(b) shows how the potential energy function changes, for a xed value  = 1, as the well-depth parameter (dstrength of the nonlinearity) is increased. a) b) FIG. 3: A) Potential energy function described by Eq. (15) illustrating the \depth" parameter together with the horizontal distance from the saddle to the center equilibrium. B) Evolution of he potential energy function for a xed value of  = 1, as is varied. Recall that, as we discussed in the introduction, in the context of chemical reaction dynamics the escape from a potential well problem can be identi ed for instance with dis- sociation or fragmentation reactions, where a chemical transformation takes place if a bond of a molecule A breaks up, giving rise to two products B and C. In this situation, the equilibrium conformation of the given molecule A is represented by a potential well in a PES, and dissociation into B and C takes place if the system has sucient energy to croos the potential barrier that separates bounded (vibration) from unbounded (bond breakup) motion. This setting can be modeled for instance by the potential energy function in Eq. (15), and is illustrated in Fig. 4(a), where a locally cubic potential energy function about 13 the potential barrier describes this phenomenon, which could be the result of a saddle-node bifurcation that has occurred in the phase space of the Hamiltonian system. In Fig. 4(b) we show the phase portrait for the dynamical system (12) which shows the equilibrium points and the homoclinic orbit. Fig. 4(c-d) display the changes in the phase space structure using the variable integration time LDs for di erent values of the well-depth parameter . (a) (b) (c) (d) FIG. 4: (a) Cubic potential energy function as a basic model for dissociation chemical reactions. The shape of this potential is the building block (normal form) for saddle-node bifurcations in phase space. (b) Phase portrait showing the location of the saddle and center equilibria in terms of the model parameters and . (c-d) Variable time LDs calculated using the p-norm de nition with p = 1=2 and  = 5 for the dynamical system described in Eq. (12) using (c)  = 1 and = 2 (d)  = 1 and = 4. 14 B. Two degrees-of-freedom Hamiltonian In this section we introduce the normal form for the two DoF Hamiltonian that undergoes a saddle-node bifurcation by extending the one DoF Hamiltonian discussed above. To do so, we add another degree of freedom in the form of a harmonic oscillator with mass m = 1 and frequency !, known as a bath mode in the terminology of chemical reaction dynamics, and parametrize the coupling of the reaction and bath modes. Thus, the Hamiltonian becomes 1 p ! " 2 2 2 3 2 H (q; x; p; p ) = p + p  q + q + x + (x q) ; (19) 2 3 2 2 where > 0 is the well-depth parameter and measures the strength of the nonlinearity, ! > 0 is the frequency of the harmonic oscillator or the bath mode,   0 is the bifurcation parameter, and " > 0 is the strength of the coupling between the reaction and the bath mode. Identifying this Hamiltonian's kinetic energy, T (p; p ), potential energy, V (q; x), and assuming H (q; x; p; p ) = T (p; p ) + V (q; x), we get that x x 1 ! " 2 2 2 3 2 T (p; p ) = p + p ; V (q; x) =  q + q + x + (x q) (20) 2 3 2 2 The corresponding Hamilton's equations are given by: @H > q_ = = p @p @H x _ = = p < x @p (21) @H > p _ = = 2  q q + "(x q) @q > @H p _ = = ! x + "(q x) @x For this Hamiltonian system, the phase space is four dimensional, and since energy is con- served, the trajectories evolve on a three dimensional energy surface. The equilibria for this e e system are located at x = (0; 0; 0; 0) and x = (q ; x ; 0; 0) where e e 1 2 2  ! " " q = ; x = q (22) e e e 2 2 (! + ") ! + " The energy of the system at the equilibrium points is: 4 2 2 2 p ! " 2! " q e e e H (x ) = 0 ; H (x ) = 2  + (23) 1 2 (! + ") 15 One can easily show that there exists a critical value of the coupling strength (also interpreted as a perturbation to the one DoF model) given by 2  ! " = ; (24) ! 2 for which the Hamiltonian (21) has only one equilibrium point at the origin. We note here that this critical value is independent of the nonlinearity strength parameter and requires ! > 2  to be satis ed. Moreover, it is characterized by a functional relationship between the squares of the eigenfrequencies of the reactive and bath modes of the uncoupled system (" = 0). We will analyze the in uence of the perturbation strength " on the geometry of the phase space structures up until this critical condition in Section III. We illustrate in Fig. 5 how the geometry of the PES (20) and the equipotentials in con guration space change as we vary the coupling strength (perturbation parameter) ". For this visualization, we have used the following values of the model parameters  = 0:25, = 2, and ! = 1:25. The reason for doing so is that they satisfy the condition ! > 2  and consequently a critical value of the perturbation parameter exists, given by " = 25=9, at which the two equilibrium points `collide' at the origin resulting in a saddle- node bifurcation. The dynamics after this collision is beyond the scope of this study and therefore we will only focus on the description of the system dynamics for di erent values of the coupling strength when " < " . We observe in Fig. 5(a,b) that the location of the well when the DoF are uncoupled (" = 0) lies on the q-axis. As " is increased, Fig. 5(c,d) show that its e ect is to tilt the PES with respect to the con guration space plane, and the position of the center equilibrium point moves o the q-axis towards the origin. Finally, the situation for which the coupling strength reaches the critical value " is shown in Fig. 5(e,f ) when the collision has happened and there is only one equilibrium point at the origin. a. Linear stability analysis of the equilibrium point at the origin. We show here that e e the equilibrium point x is a rank-1 saddle on the PES and that x is a center corresponding 1 2 to the bottom of the potential well. This linear stability analysis will be useful later for computing the rank-1 saddle's NHIM, which carries its in uence to higher energies, and also to determine the NHIM's invariant stable and unstable manifolds that act as codimension one impenetrable barriers on the constant energy surface. The global geometry of the invariant manifolds is paramount for quantifying reaction rates such as escape rate from potential well. 16 (a) (b) (c) (d) (e) (f ) FIG. 5: Geometry of the PES (left-column) and equipotentials in con guration space (right column) for the model parameters  = 0:25, = 2 and ! = 1:25. (a) and (b) correspond to " = 0; (c) and (d) represent the case " = 0:25; (e) and (f ) are for the critical coupling strength " = 25=9. 17 We determine the stability of the equilibria by linearizing Eq. (21), which gives the Jacobian matrix 0 1 2 2 2 2 @ H @ H @ H @ H B C @q@p @x@p @p @p @p x 0 1 B C B C 0 0 1 0 2 2 2 2 @ H @ H @ H @ H B C B C B C B C B C @q@p @x@p @p@p @p 0 0 0 1 x x x x B C B C J(q; x; p; p ) = = B C B C p 2 2 2 2 B C @ H @ H @ H @ H B C 2  2 q " " 0 0 @ A B C B @q @x@q @p@q @p @qC x 2 B C " ! " 0 0 @ 2 2 2 2 A @ H @ H @ H @ H @q@x @x @p@x @p @x (25) The stability of x = (0; 0; 0; 0) is given by the eigenvalues of the Jacobian 0 1 0 0 1 0 B C B C 0 0 0 1 B C J(x ) = B C (26) B C 2  " " 0 0 @ A " ! " 0 0 which has the characteristic equation e 2 2 2 2 det (J(x ) I) = 2  + " + ! + " " = 0 (27) where are the eigenvalues and I the identity matrix. When the reaction and bath modes are decoupled, that is " = 0 in (27), the eigenvalues are given by  and ! i, where 0 0 = 4 ; ! = ! (28) 0 0 Therefore, the origin is a rank-1 saddle equilibrium point, since the linearized system has ex- actly one pair of real eigenvalues,  and  and the saddle plane is spanned by their eigen- 0 0 12,58 vectors. We know from the Moser's generalization of the Lyapunov Subcenter Theorem that when the energy of the system is above that of the rank-1 saddle, there is a two dimen- sional plane spanned by the eigenvectors of ! i, known as the center invariant manifold. This invariant manifold with normal hyperbolicity is referred to as the normally hyperbolic 2N3 invariant manifold (NHIM) and has the topology of a (2N 3)-sphere, that is S . In the two DoF setting, a NHIM is simply an unstable periodic orbit whose geometry is topo- logically equivalent to a circle, i.e. S . 18 As the energy of the system is increased above the energy of the rank-1 saddle equilibrium point, a bottleneck opens in the phase space connecting dynamically di erent phase space regions and allowing trajectories to move between them. This phenomenon results in a phase space transport mechanism. This framework of understanding chemical reactions is realized by computing the stable and unstable manifolds associated with the unstable periodic orbit. These invariant manifolds have cylindrical geometry, that is Rx , and their global behavior is referred to as tube dynamics. The cylindrical manifolds on the constant energy surface are impenetrable barriers (since they are two dimensional on the three dimensional energy surface) separating the reactive and non-reactive trajectories in the phase space. Thus, they determine the initial conditions that will pass through the bottleneck in some future time (or had passed through it in the past) during their evolution . The unstable periodic orbit provides us with the sca olding to construct the dividing surface that separates the trapped motion in the well region of the PES and the escape to in nity of particle trajectories through the phase space bottleneck. The local (linearized) dynamics mediated by these phase space structures is shown in Appendix B. We will resume the discussion on computing the NHIM and its invariant manifolds in x:III B. When " 6= 0, the reaction and bath modes are coupled and the Hamilton's equations (21) are non-integrable. This means we can expect chaotic trajectories to appear that might lead to reaction by escaping the potential well. However, the geometry of the NHIM and its invariant manifolds still governs this reacting and non-reacting behavior. To simplify the eigenvalue problem arising in the stability analysis, we observe that Eq. (27) can be rewritten in terms of the eigenvalues of the uncoupled system as 2 2 2 2 2 4 2 2 2 2 2 2 2 + " + ! + " " = + !  + 2" + !  "  ! = 0 (29) 0 0 0 0 0 0 0 0 Introducing  = the solutions are 2 2 2 2 !  + ! 0 0 0 0 = " + " : (30) 2 2 We note here that the two possible values of  have opposite signs for all values " > 0 of 2 2 2 2 the coupling strength when !   , and for 0 < " < " whenever the condition ! >  is 0 0 0 0 satis ed, where the critical coupling strength " is given by Eq. (24). If we denote  and c 1;" as the positive and negative roots respectively in Eq. (30), then the eigenvalues of the 2;" 19 Jacobian matrix are  and ! i where " " =  ; ! = j j (31) " 1;" " 2;" This shows that in this case the equilibrium point at the origin is a rank-1 saddle. In particular, when the coupling between the reaction and the bath mode is weak, that is "  1, we have q q 2 2 " ; !  ! + " (32) " " 0 0 which con rms that the rank-1 saddle structure of the equilibrium point at the origin persists under small perturbations. We nish the linear stability analysis of the equilibrium point at the origin with the computation of the eigenvectors. Let us denote the eigenvector of the matrix J (x ) by v = [v ; v ; v ; v ] , then the eigenvalue satis es 1 2 3 4 > v = v 1 3 v = v 2 4 J (x )v = v , (33) v = ( ") v + "v > 3 1 2 v = "v (! + ") v 4 1 2 The eigenvectors associated with the real eigenvalues =  correspond to the tangent directions of the unstable and stable manifolds, respectively, of the NHIM at the rank-1 saddle, and are given by "  " u = 1; ; ; : (34) 2 2 2 2 " + ( + ! ) " + ( + ! ) " 0 " 0 The eigenvectors associated with the complex eigenvalues = ! i span the center subspace of the NHIM at the rank-1 saddle and are given by " ! " w = ; 1; i;! i : (35) 2 2 2 2 " ( + ! ) " ( + ! ) 0 " 0 " where ! is the magnitude of the complex eigenvalue in (31). Therefore, we can write the solution to the linearized system at the rank-1 saddle as t  t i! t " " " x(t) = C e u + C e u + 2Re e w (36) 1 + 2 + where C ; C 2 R and  =  +  i 2 C are constants to be determined from an initial 1 2 1 2 condition. We will use this general solution to the linear system to select an initial guess to search for the NHIM using the di erential correction and numerical continuation method . 20 For the sake of completeness, we show next the linear stability analysis for the equilibrium point x . The Jacobian is given by 0 1 0 0 1 0 B C B C 0 0 0 1 2! " B C J(x ) = ; where = (37) B C B C ! + " 2  " " 0 0 @ A " ! " 0 0 The characteristic equation det (J (x ) I) = 0 becomes 2 2 2 2 2 +  + " + ! + " " = 0 0 0 (38) 4 2 2 2 2 2 2 2 + ! +  + 2" + ! +  " + !  = 0 0 0 0 0 0 0 Introducing  = the solutions are 2 2 2 2 ( + ! ) + ! 0 0 0 0 = " + " (39) 2 2 It can be easily shown that both solutions are negative, and consequently J(x ) has two pairs of complex eigenvalues. This implies that the equilibrium point x is a center. III. RESULTS AND DISCUSSION: INFLUENCE OF WELL-DEPTH ON REAC- TION DYNAMICS In this section we discuss the implications of changing the well-depth in terms of the geometry of the invariant manifolds that mediate reacting trajectories (or escape from the potential well). This is done by identifying the invariant manifolds using Lagrangian de- scriptors and a numerical continuation with globalization method (see Appendix C). The geometry of these invariant manifolds and character of changes a ected by varying the po- tential well-depth are involved in the computation of reaction fraction (or escape rates). A. One Degree-of-Freedom Hamiltonian Here we return to the 1 DoF Hamiltonian (11) where the reaction is de ned as the change in the sign of the con guration coordinate q. For the rank-1 saddle equilibrium point at (0; 0), the energy is H (0; 0) = H = 0 which we call the critical energy, and 21 p that of the center equilibrium point at (2 = ; 0), at the bottom of the potential well, is H (2 = ; 0) = H = 4  =3 . In this model system, the phase space is two- dimensional and since the total energy is conserved, the trajectories evolve on the one- dimensional constant energy lines or isoenergetic contours of the Hamiltonian in the 2D phase space. Therefore, the system is completely integrable, and for a xed energy H the trajectories evolve on the one-dimensional curve 1 p 2 2 3 H = p  q + q (40) 2 3 The value of total energy H with respect to H and H characterizes the nature of trajec- 0 c w tories as follows We discuss the nature of the trajectories depending on the energy H of the system: Case H < H : The initial conditions that satisfy q < = , which lie to the left 0 w of the potential energy barrier at the origin. Initial conditions in this con guration space initially climb the potential and at some point their velocity reverses direction and they roll down the potential going to in nity. In the phase space, they come from in nity and y-by the rank-1 saddle on their way towards in nity. They can not cross the barrier and thus do not lead to reaction shown by the blue curve in Fig. 6. Case H  H < 0: Now two types of trajectories are possible. For the initial w 0 conditions that satisfy =  q < 0, trajectories will show same y-by behavior as in the previous case. If initial conditions satisfy 0 < q < 3 = , then they are con ned in the potential well on the right of the potential barrier. Since they are bounded by the homoclinic orbit, trajectories will never escape the well, and thus do not lead to reaction and shown as green curves in Fig. 6. Case H = H = 0: When the total energy of the initial conditions is equal to the 0 c potential energy of the barrier, resulting trajectories approach the barrier asymptoti- cally in forward and backward time. The initial conditions that start on the left of the potential barrier and asymptotically approach it in forward and backward time form pieces of the stable and unstable manifolds of the saddle equilibrium point. The initial conditions that start on the right of the potential barrier also approach the barrier as time goes to in nity. These trajectories combine to form the homoclinic orbit and shown as a black curves in Fig. 6. 22 8 Reactive Reactive Non-reactive Non-reactive 6 NHIM NHIM DS DS -2 -2 8 8 8 6 6 6 6 6 4 4 4 -4 4 4 -4 4 2 2 2 2 0 0 0 0 0 -2 -2 -2 -2 -2 -2 -6 -4 -4 -4 -6 -4 -4 -4 -6 -6 -6 -6 -6 -6 -8 -8 -8 -5 0 5 -8 -8 -5 0 5 -5 0 5 -8 -5 0 5 -5 0 5 -5 0 5 -8 -8 -5 0 5 -5 0 5 (a) (b) FIG. 6: Trajectory behaviors dictated by the potential energy (15). (a)  = 1 and = 1 with the insets  = 2; 4; 6 for = 1 (b)  = 1 and = 4 with the insets  = 2; 4; 6 for = 4. Here we are showing the reactive and nonreactive trajectories (red and blue, respectively) partitioned by the homoclinic orbit (shown in black) formed by the unstable and stable manifolds of the NHIM (shown as a magenta cross), and the reactive trajectory (red) is at energy H = 2, while the two non-reactive trajectories are at H = 1;5. The dividing surface that the reactive trajectory must cross in order for reaction to occur is marked by green asterisks. Case H > H = 0: The energy of the system is above that of the barrier so trajecto- 0 c ries can escape from the well, that is the con guration space coordinate q can change sign and lead to reaction. The dividing surface that trajectories must cross once in a given direction (locally no recrossing), that is when escaping from the potential well, is given by n o D = (q; p) 2 R j q = 0 ; = H > 0 = 0; 2H : (41) 0 0 We note that in this case the dividing surface for a xed energy and given direction of crossing is a point, or has a geometry of S , thus it can partition the one-dimensional (isoenergetic) curves into \reactant" and \`product" regions. To further characterize the reaction dynamics, we resort to the linear stability analysis near the saddle equi- librium point at the origin. The eigenvalues of the saddle equilibrium point are  4 and the corresponding eigenvectors 1= 4; 1 . Therefore, the origin is a NHIM 23 61 since the dynamics is hyperbolic in directions normal to it . Furthermore, we remark that the stability of the NHIM only depends on the  parameter. The eigenvectors tell us how the unstable and stable manifolds of the NHIM are oriented and can be used to numerically globalize the linear approximation. However, we can compute the stable and unstable manifolds of the saddle equilibrium point analytically by noting that they lie on the zero level curve of the Hamiltonian (total energy). Thus, the stable and unstable manifolds are given by s s u u W (0; 0) = [W (0; 0) ; W (0; 0) = [W (0; 0) (42) l l where is the homoclinic orbit = (q; p) 2 R j q > 0 ; H (q; p) = 0 (43) and s 2 W (0; 0) = (q; p) 2 R j q < 0 ; p > 0 ; H (q; p) = 0 (44) u 2 W (0; 0) = (q; p) 2 R j q < 0 ; p < 0 ; H (q; p) = 0 We note here that the geometry of the invariant manifolds is R  S , that is codimension-1 in R , and hence can form the impenetrable barrier between the re- active and non-reactive trajectories as shown in Fig. 6. B. Two degree-of-freedom Hamiltonian We return to the 2 DoF Hamiltonian (19) which introduces the bath degree-of-freedom as a harmonic oscillator into the reaction dynamics. We show the changes in the geometry of the invariant manifolds due to the changes in the well-depth which leads to a saddle-node bifurcation. This is done using Lagrangian descriptors along with numerical continuation and globalization for computing the NHIM and its invariant manifolds. Let us consider a xed energy H and since the model has 2 DoF we know that the dynamics is on a three-dimensional energy surface given by 1 p ! " 4 2 2 2 3 2 S (H ) = (q; x; p; p ) 2 R p + p  q + q + x + (x q) = H (45) 0 x 0 2 3 2 2 24 The projection of the energy surface onto the (q; x) con guration space is the region of energetically possible motion for a xed energy H , and is given by C(H ) = (q; x) 2 R V (q; x) 6 H 0 0 2 (46) p ! " 2 2 3 2 = (q; x) 2 R  q + q + x + (x q) 6 H 3 2 2 This projection denotes con gurations with positive kinetic energy and has been known in classical mechanics as the Hill's region. The boundary of M (H ) is de ned as the locus of points in the (q; x) plane where the kinetic energy is zero, that is (H V (x; y)) = 0, and is called the zero velocity curve. The trajectories are only able to move on the side of the curve where the kinetic energy is positive, shown as white regions in Fig. 7(b-d). To identify the invariant manifolds, we take two dimensional slices of the energy surface and determine the intersection of the invariant manifolds with these low-dimensional slices. In particular, we calculate Lagrangian desriptor and compare with qualitative understand- ings from Poincar e surface-of-section. The isoenergetic surfaces-of-section are + 4 U = (q; x; p; p ) 2 R x = 0 ; p (q; x; p; H )  0 (47) x x 0 qp + 4 U = (q; x; p; p ) 2 R q = q ; p(q; x; p ; H )  0 (48) x e x 0 xp where q is the con guration space coordinate of the equilibrium point at the bottom of the well on the PES (22). We return to the Hamiltonian (19) where the \reaction" and \bath" DoF are uncoupled, that is " = 0. Therefore, the system is integrable and the trajectories are regular, thus the separable Hamiltonian is 1 p 1 ! 2 2 3 2 2 H (q; x; p; p ) = p q + q + p + x (49) 2 3 2 2 | {z } | {z } H (q;p) H (x;p ) r b x where H is the Hamiltonian for the reaction and H is the Hamiltonian for the bath DoF. r b Given a xed total energy of the system H , the necessary condition for the reaction to take place is when the total energy is above that of the barrier of the PES located at the origin, which is zero. For H  0 the energy surface divides phase space into two disconnected regions as illustrated in Fig. 7 so that we have bounded motion in the potential well region. In Fig. 8, we compare the bounded trajectories (regular dynamics) of the system for H = 0 by computing LDs and Poincar e section on the surface-of-section, U (47). qp 25 (a) (b) (c) (d) FIG. 7: (a) Phase space energy surface, inside which motion takes place, for three di erent values of the energy; (b) Con guration space projection for an energy H = 0:025 below the barrier energy; (c) for the rank-1 saddle energy H = 0; (d) for the energy H = 0:1 which is above the barrier energy. The parameter values chosen are  = 0:25, = 2, ! = 1:25 and " = 0. We have marked the energy accessible regions in white and the forbidden regions in dark grey. The equipotential in black marks the zero velocity curve for which the kinetic energy of the system is zero. Both methods clearly recover, the trajectories on tori, as known for integrable Hamiltonian systems , that lls the energy surface. We observe that the high values (white regions in 8) in the LD contour map recovers the quasiperiodic trajectories, which is a consequence of the relationship between the convergence of time averages of LDs with the Ergodic Partition Theorem as explained in Appendix A. 26 A) B) FIG. 8: Quasiperiodic trajectories describing regular motion in the potential well region of the PES for the uncoupled (" = 0) Hamiltonian system with energy H = 0. The model parameters chosen for this calculation are  = 0:25, = 2 and ! = 1:25; A) Poincar e map; B) LDs obtained for an integration time  = 20. The curve in magenta depicts the energy boundary. When the energy of the system is above the barrier, that is H > 0, the topology of the energy surface changes and a phase space bottleneck opens up in the barrier region allowing the escape from the well as shown in Fig. 7(d). Thus, reaction can take place by crossing the bottleneck and passing from the q > 0 to q < 0 or vice versa. Therefore, q = 0 is a natural choice for de ning a dividing surface (DS) separating reactants (bounded motion in the potential well region) from products (escape to in nity) in phase space or vice versa. Hence, the isoenergetic DS is given by 4 2 2 2 2 D(H ) = (q; x; p; p ) 2 R j q = 0 ; 2H = p + p + ! x (50) 0 x 0 2 3 which has the geometry of a 2-sphere, that is S in R . To be precise, it is an ellipsoid with p p semi-major axis 2H in the p, p axis, and 2H =! in the x axis. This ellipsoid has two 0 x 0 hemispheres, known as the forward and backward DS with the form n o 2 2 2 D (H ) = (q; x; p; p ) 2 R j q = 0 ; p = 2H p ! x f 0 x 0 n o (51) 2 2 2 D (H ) = (q; x; p; p ) 2 R j q = 0 ; p = + 2H p ! x b 0 x 0 Forward \reaction" occurs when trajectories cross D and back \reaction" when trajecto- ries cross D . The forward and backward DS is joined at the equator along the normally hyperbolic invariant manifold given by 4 2 2 2 N (H ) = (q; x; p; p ) 2 R j q = p = 0 ; 2H = p + ! x (52) 0 x 0 27 p which has the topology of S . To be precise, it is an ellipse with semiaxis 2H in the p 0 x direction and 2H =! in the x direction. As discussed earlier, for a 2 DoF Hamiltonian the NHIM is an unstable periodic orbit which extends the in uence of the rank-1 saddle equilibrium point in the bottleneck of the PES (a con guration space concept) into phase space. We note that for H > 0 the NHIM is the correct phase space structure that anchors the barriers to the reaction and carry the e ect of the saddle equilibrium point to a range of energies as given by Moser's generalization of Lyapunov Subcenter Manifold Theorem . The stable and unstable manifolds of the unstable periodic orbit are s s u u W = [W ; W = [W (53) l l where is the homoclinic orbit = (q; x; p; p ) 2 R j q > 0 ; H (q; p) = 0 ; H (x; p ) = H (54) x r b x 0 and the left branches are s 4 W = (q; x; p; p ) 2 R j q < 0 ; p > 0 ; H (q; p) = 0 ; H (x; p ) = H x r b x 0 (55) u 4 W = (q; x; p; p ) 2 R j q < 0 ; p < 0 ; H (q; p) = 0 ; H (x; p ) = H x r b x 0 We note that the stable and unstable invariant manifolds have the structure of a cartesian product of a curve in the (q; p) saddle space and an ellipse in the (x; p ) center space, and thus become cylindrical (or tube ) manifolds. We show the energy surface in Fig. 7(a) at 3 energy values for  = 0 where the bottleneck only opens for H > 0. The NHIM computed using di erential correction and continuation, and its invariant manifolds computed using globalization are shown in Fig. 9(a) for H = 0:05 and  = 0. When " 6= 0, the \reaction" and \bath" modes are coupled and for " small, we can think about the resulting dynamics as a perturbation of the uncoupled case . Given an energy of the system below that of the barrier, that is H  0, the phase space bottleneck is closed and trajectories are trapped in the potential well region. However, due to the perturbation, the unstable regular motion on the tori in the uncoupled system is destroyed as given by the KAM theorem. Therefore, chaotic motion arises in some regions of the phase space due to a non-zero coupling of the \reaction" and \bath" modes. We illustrate the system's behavior for the energy H = 0 (energy of the rank-1 saddle) in Fig. 10 using LDs and Poincar e + + maps on the surfaces of section U (47) and U (48). We observe that there is a distinct qp xp 28 (a) (b) (c) (d) FIG. 9: Con guration space and phase space view of the energy surface and the cylindrical mani- folds of the unstable periodic orbit associated with the rank-1 saddle in the bottleneck. Blue and red denote the stable and unstable manifolds respectively, and the green surface is the isosurface of total energy H = 0:05. The magenta curve represents the unstable periodic orbit. (a-b)  = 0 and (c-d)  = 0:25. correlation between the qualitative dynamics revealed by the Lagrangian descriptor (LD) contour maps and Poincar e sections. That is, chaotic regions of phase space which appear as a sea of points in the Poincar e section and hide the underlying structures of stable and unstable manifolds are completely resolved by LD contour maps where the tangled geometry of the manifolds is revealed by the points where the function attains a local minimum as has also been shown in . The capability of LDs to identify the phase space structures relevant 29 43,49,50 in chemical reaction dynamics can also be found in recent literature . In addition, the generation of Poincar e section relies on tracking the crossing of a 2D surface which can not be guaranteed in high dimensional phase space, while LDs just accumulate a positive scalar quantity along the trajectory and thus have potential to reveal high dimensional phase space structures. A) B) C) D) FIG. 10: Phase space structures of the coupled Hamiltonian with model parameters  = 0:25, = 2, ! = 1:25 and " = 0:25. The total energy of the system is H = 0 (barrier energy). A) Poincar e map on the surface of section U ; B) LDs calculated for  = 75 on the surface of section qp + + U ; C) Poincar e map on the surface of section U ; D) LDs calculated for  = 75 on the surface qp xp of section U . We have marked the energy boundary with a magenta curve. xp Now let us consider the dynamics when the total energy is above that of the rank-1 30 saddle, that is H > 0, and in particular we will set H = 0:05 in our analysis. In this 0 0 situation, a phase space bottleneck opens up in the barrier region, as shown in Fig. 7. In order to describe the structures that mediate reaction dynamics, that is the NHIM (or unstable periodic orbit in this case) and its stable and unstable manifolds, we use di erential correction and continuation along with globalization as described in Appendix C. In Fig. 9(c- d), we have shown the cylindrical manifolds along with the energy surface and the unstable periodic orbit for a xed energy. In order to recover the homoclinic tangle geometry of + + the invariant manifolds, we calculate LDs on the surfaces of section U and U , and qp xp compare with the direct numerical construction of these invariant manifolds. This LD based diagnostic is similar to performing a \phase space tomography" of the high dimensional phase space structures using a low dimensional slice. In Fig. 11 we show the computation of variable time LD for an integration time  = 10 on the slice U . We observe that LD qp clearly identify, by means of points which attain a minimum, the location of the stable and unstable manifolds and the NHIM at the intersection of the invariant manifolds. Since we are using a small integration time of  = 10 to compute LDs in Fig. 11, the complete geometry of the homoclinic tangle is not fully revealed. Therefore, in order to recover a FIG. 11: Variable time LDs calculated on the surface of section U for  = 10. The NHIM and its qp stable and unstable manifolds are revealed as points where the LD eld in non-di erentiable and attains a local minimum. We have marked the energy boundary with a magenta curve. more complete and intricate dynamical picture of the homoclinic tangle, the integration 31 time to compute LDs has to be increased. This is shown in Fig. 12(b), where  = 30 and we observe that the regular motion obtained in the middle of the Poincar e section displayed in Fig. 12(a) corresponds to trajectories that remain trapped in the potential well region and never escape. The trapped dynamics on the tori are also captured by the LD contour map shown in 12(b) which also reveals the homoclinic tangle of the stable and unstable manifolds and the resulting lobe dynamics . We note that in all these computations we are using the variable time de nition of LDs, since the open potential surface causes trajectories to escape to in nity through the bottleneck in nite time and resulting in NaN values in the LD contour map. This issue is discussed in Appendix A and illustrated in Fig. 14, and would hide the important underlying phase space structures making the interpretation of results dicult. In addition, when we analyze the dynamics using Poincar e sections, we can not ensure that trajectories return to the surface of section when escaping to in nity is possible. This will result in blank regions in the Poincar e sections as shown in Figs. 12(a) and (c). However, we note that trapped trajectories in the potential well corresponding to regular (motion on the tori) and chaotic motion are highlighted as expected in the Poincar e sections. These trajectories are non-reactive and will remain so until they satisfy the sucient condition for reaction which is entering the cylindrical manifolds. Identifying the regions inside the cylindrical (tube) stable and unstable manifolds can be done using globalization or using the LD contour map on appropriate isoenergetic surfaces. We compute LD contour maps to recover the intersections of these tube manifolds with the surface of section (48). The intersection of the invariant manifolds with the isoenergetic surface of section yields topological ellipses, known in the chemical reactions literature as 48,64{66 reactive islands and is involved in calculating reaction rates/fraction. In order to illustrate the capability of LDs to recover the reactive island structure, we have compared the LD contour map and manifold intersection with the surface of section U obtained xp using globalization in Fig. 13. We have shown the rst intersection of the stable and unstable manifold with the surface of section as a blue/red curve superimposed on the LD contour map. We observe that the minima in the LD contour maps and the manifolds intersections are in agreement, thus verifying the LD based identi cation of reaction islands. We also observe that successive fold and resulting intersection of the tube manifolds with the surface of section are also revealed by the points with minima in the LD values. We note here that the rst intersection of the manifolds encloses a large phase space volume in the potential 32 (a) (b) (c) (d) FIG. 12: Phase space structures of the coupled Hamiltonian with model parameters  = 0:25, = 2, ! = 1:25 and " = 0:25. The total energy of the system is H = 0:05, which is above the barrier energy. (a) Poincar e map on the surface of section U (b) LDs calculated for  = 30 on qp + + the surface of section U (c) Poincar e map on the surface of section U (d) LDs calculated for qp xp = 30 on the surface of section U . We have marked with a magenta curve the energy boundary. xp well which indicates that a large portion of the potential well escapes to in nity through the bottleneck. This is despite the fact that we have chosen a very small value for the energy of the system, H = 0:05, compared to the energy of the barrier. This is a consequence 33 of using a high coupling strength, " = 0:25, which makes the vase-like shaped potential well region small and narrow as the center equilibrium point at the bottom of the potential well approaches the rank-1 saddle equilibrium point located at the origin as " is increased. Movies illustrating the change in both the shape of the energy surface and the equipotentials in con gurations space as we vary the coupling strength can be found at here and here, respectively. We can see that the e ect of increasing the coupling strength from zero is to tilt and squeeze the vase-like shape of the energy surface that qualitatively increases the number of reactive trajectories. This action of tilting and squeezing of the vase-like container (boundary de ned by the total energy surface) is to pour out its contents (the reactive trajectories) and hence the increase in reaction fraction. A quantitative investigation of this phenomenon is current work in progress and beyond the scope of this article. IV. CONCLUSIONS In this article we have presented new results, by means of using a normal form Hamil- tonian that models a saddle-node bifurcation, to investigate the in uence on phase space dynamics of decreasing the potential well depth of a PES. The decrease in potential well depth causes the saddle and center equilibrium points to collide |an e ect that takes place in the con guration space. The resulting saddle-node bifurcation, when the well depth is decreased, manifests in the phase space as tilting and squeezing of the energy surface. This change in geometry of the energy surface due to decreasing well depth leads to more reacting (or escaping) trajectories for a given energy. This is supported by the increase in the area of the rst intersection of the invariant manifolds with an appropriate surface of section. The trajectory diagnostic method of Lagrangian descriptors also identi es these changes in the geometry of the phase space structures. This method also provides an approach for revealing the in uence of potential well depth on high dimensional phase space structures. Detecting the qualitative changes in the geometry of the invariant manifolds using Lagrangian de- scriptors gets us closer to the goal of achieving a complete high-dimensional \phase space tomography" for realistic molecular systems. 34 A) B) FIG. 13: Phase space structures revealed by LDs using  = 30, for the coupled Hamiltonian with parameters  = 0:25, = 2, ! = 1:25 and total energy H = 0:05. A) Coupling strength " = 0:25. On the top/bottom left we show the forward/backward LD and, superimposed, the rst intersection of the stable/unstable manifold with U as a blue/red curve. On the right, the total xp LD is depicted (addition of forward and backward LD) together with the rst intersection of the stable and unstable manifolds, known in chemistry as reactive islands since trajectories inside these regions will escape the potential well through the bottleneck in forward/backward time respectively. The magenta curve represents the energy boundary. B) same analysis using " = 0:125. 35 Acknowledgments We acknowledge the support of EPSRC Grant No. EP/P021123/1 and Oce of Naval Research Grant No. N00014-01-1-0769. The authors would like to acknowledge the EPSRC sponsored Chemistry and Mathematics in Phase Space (CHAMPS) workshop held March 19, 2019 at the University of Bristol where they had the opportunity to discuss with Florentino Borondo about the role of the saddle-node bifurcation in the isomerization of lithium cyanide. Electronic address: vjose.garcia@uah.es Electronic address: s.naik@bristol.ac.uk Electronic address: s.wiggins@bristol.ac.uk T. Komatsuzaki and R. S. Berry, J. Mol. Struct. THEOCHEM 506, 55 (2000). T. Uzer, C. Ja  e, J. Palaci an, P. Yanguas, and S. Wiggins, Nonlinearity 15, 957 (2002). T. Komatsuzaki and R. S. Berry, J. Phys. Chem. A 106, 10945 (2002). R. S. MacKay, Phys. Lett. A 145, 425 (1990). H. Waalkens and S. Wiggins, J. Phys. A: Math. Gen. 37, L435 (2004). E. P. Wigner, Trans. Faraday Soc. 34, 29 (1938). E. P. Wigner, J. Chem. Phys. 7, 646 (1939). H. Waalkens, R. Schubert, and S. Wiggins, Nonlinearity 21, R1 (2008). T. Komatsuzaki and M. Nagoaka, Chem. Phys. Lett. 265, 91 (1997). S. Wiggins, Global bifurcations and chaos: Analytical methods (Springer-Verlag, 1988). S. Wiggins, Physica D 44, 471 (1990). S. Wiggins, Normally hyperbolic invariant manifolds in dynamical systems, vol. 105 (Springer Science & Business Media, 2013). P. Pechukas and F. J. McLa erty, J. Chem. Phys. 58, 1622 (1973), URL http://scitation. aip.org/content/aip/journal/jcp/58/4/10.1063/1.1679404. P. Pechukas and E. Pollak, J. Chem. Phys. 67, 5976 (1977), URL http://scitation.aip. org/content/aip/journal/jcp/67/12/10.1063/1.434777. E. Pollak and P. Pechukas, J. Chem. Phys. 69, 1218 (1978). P. Pechukas and E. Pollak, J. Chem. Phys. 71, 2062 (1979), URL http://scitation.aip. 36 org/content/aip/journal/jcp/71/5/10.1063/1.438575. S. Wiggins, L. Wiesenfeld, C. Ja  e, and T. Uzer, Physical Review Letters 86, 5478 (2001). R. E. Gillilan and G. S. Ezra, J. Chem. Phys. 94, 2648 (1991). T. Komatsuzaki and M. Nagaoka, J. Chem. Phys. 105, 10838 (1996). S. Wiggins, Regular and Chaotic Dynamics 21, 621 (2016). C. B. Li, M. Toda, and T. Komatsuzaki, J. Chem. Phys. 130, 124116 (2009). M. Inarrea, J. F. Palacian, A. I. Pascual, and J. P. Salas, J. Chem. Phys. 135, 014110 (2011). A. Allahem and T. Bartsch, J. Chem. Phys. 137, 214310 (2012), submitted to Arxiv.org. F. A. L. Maugui ere, P. Collins, G. S. Ezra, and S. Wiggins, Int. J. Bifurcat. Chaos 23 (2013). R. S. MacKay and D. C. Strub, Nonlinearity 27, 859 (2014). R. S. MacKay and D. C. Strub, Nonlinearity 28, 4303 (2015), 1501.00266. D. Hathcock and J. P. Sethna, arXiv preprint arXiv:1902.07382 (2019). J. Husson, M. Dogterom, and F. Pincet, The Journal of Chemical Physics 130, 051103 (2009). N. J. Miller and S. W. Shaw, Physical Review E 85, 046202 (2012). C. Herbert and F. Bouchet, Physical Review E 96, 030201 (2017). F. Borondo, A. A. Zembekov, and R. M. Benito, Chemical Physics Letters 246, 421 (1995). F. Borondo, A. Zembekov, and R. Benito, The Journal of chemical physics 105, 5068 (1996). F. Revuelta, R. Benito, and F. Borondo, Physical Review E 99, 032221 (2019). K. A. Mitchell, J. P. Handley, B. Tighe, J. B. Delos, and S. K. Knudson, Chaos: An Interdisci- plinary Journal of Nonlinear Science 13, 880 (2003). K. A. Mitchell, J. P. Handley, J. B. Delos, and S. K. Knudson, Chaos: An Interdisciplinary Journal of Nonlinear Science 13, 892 (2003). K. A. Mitchell, J. P. Handley, B. Tighe, A. Flower, and J. B. Delos, Physical Review Letters 92 (2004). N. De Leon and B. J. Berne, The Journal of Chemical Physics 75, 3495 (1981). K. A. Mitchell, J. P. Handley, B. Tighe, A. Flower, and J. B. Delos, Physical Review A 70 (2004). K. A. Mitchell and B. Ilan, Physical Review A 80 (2009). K. A. Mitchell and J. B. Delos, Physica D: Nonlinear Phenomena 229, 9 (2007). L. Wang, H. F. Yang, X. J. Liu, H. P. Liu, M. S. Zhan, and J. B. Delos, Physical Review A 82 (2010). 37 42 J. A. J. Madrid and A. M. Mancho, Chaos 19, 013111 (2009). G. T. Craven and R. Hernandez, Physical Chemistry Chemical Physics 18, 4008 (2016). G. T. Craven and R. Hernandez, Physical review letters 115, 148301 (2015). G. T. Craven, A. Junginger, and R. Hernandez, Physical Review E 96, 022222 (2017). A. Junginger, L. Duvenbeck, M. Feldmaier, J. Main, G. Wunner, and R. Hernandez, The Journal of chemical physics 147, 064101 (2017). M. Feldmaier, A. Junginger, J. Main, G. Wunner, and R. Hernandez, Chemical Physics Letters 687, 194 (2017). S. Patra and S. Keshavamurthy, Physical Chemistry Chemical Physics 20, 4970 (2018). A. S. Demian and S. Wiggins, International Journal of Bifurcation and Chaos 27, 1750225 (2017). S. Naik, V. J. Garc a-Garrido, and S. Wiggins, arXiv preprint:1903.10264 (Under Review) (2019). S. Naik and S. Wiggins, arXiv preprint:1904.10018 (Under Review) (2019). C. Lopesino, F. Balibrea-Iniesta, V. J. Garc a-Garrido, S. Wiggins, and A. M. Mancho, Inter- national Journal of Bifurcation and Chaos 27, 1730001 (2017). F. Balibrea-Iniesta, C. Lopesino, S. Wiggins, and A. M. Mancho, International Journal of Bifurcation and Chaos 26, 1630036 (2016). V. J. Garcia-Garrido, A. Ramos, A. M. Mancho, J. Coca, and S. Wiggins, Marine Pollution Bulletin. pp. 1{10 (2016). A. G. Ramos, V. J. Garc a-Garrido, A. M. Mancho, S. Wiggins, J. Coca, S. Glenn, O. Scho eld, J. Kohut, D. Aragon, J. Kerfoot, et al., Scient c Reports 4, 4575 (2018). F. Balibrea-Iniesta, J. Xie, V. J. Garc a-Garrido, L. Bertino, A. M. Mancho, and S. Wiggins, Quarterly Journal of the Royal Meteorological Society 145, 76 (2019). S. Wiggins (2017), URL https://figshare.com/articles/Ordinary_Differential_ Equations/5311612. S. Wiggins, Introduction to Applied Nonlinear Dynamical Systems and Chaos, vol. 2 (Springer Science & Business Media, 2003). S. Wiggins, L. Wiesenfeld, C. Ja , and T. Uzer, Physical Review Letters 86, 5478 (2001). W. S. Koon, M. W. Lo, J. E. Marsden, and S. D. Ross, Dynamical systems, the three-body problem and space mission design (Marsden books, 2011), ISBN 978-0-615-24095-4. 38 61 S. Wiggins, Regular and Chaotic Dynamics 21, 621 (2016). K. R. Meyer, G. R. Hall, and D. On, Introduction to Hamiltonian Dynamical Systems and the N-Body Problem (Springer, 2009). D. Beigie and S. Wiggins, Physical Review A 45, 4803 (1992), ISSN 1050-2947, 1094-1622. N. De Leon, M. A. Mehta, and R. Q. Topper, The Journal of Chemical Physics 94, 8310 (1991). N. De Leon, M. A. Mehta, and R. Q. Topper, The Journal of Chemical Physics 94, 8329 (1991). N. De Leon, The Journal of Chemical Physics 96, 285 (1992). A. M. Mancho, S. Wiggins, J. Curbelo, and C. Mendoza, Communications in Nonlinear Science and Numerical Simulation 18, 3530 (2013). I. Mezic and S. Wiggins, Chaos 9, 213 (1999). A. Junginger, G. T. Craven, T. Bartsch, F. Revuelta, F. Borondo, R. Benito, and R. Hernandez, Physical Chemistry Chemical Physics 18, 30270 (2016). V. J. Garc a-Garrido, J. Curbelo, A. M. Mancho, S. Wiggins, and C. R. Mechoso, Regular and Chaotic Dynamics 23, 551 (2018). S. Wiggins, Normally hyperbolic invariant manifolds in dynamical systems. (Springer-Verlag, New York, 2014), ISBN 978-1-4612-8734-6. S. Naik and S. D. Ross, Communications in Nonlinear Science and Numerical Simulation 47, 48 (2017). S. D. Ross, A. E. BozorgMagham, S. Naik, and L. N. Virgin, Physical Review E 98, 052214 (2018). T. S. Parker and L. O. Chua, Practical Numerical Algorithms for Chaotic Systems (Springer- Verlag New York, Inc., New York, NY, USA, 1989), ISBN 0-387-96689-7. J. E. Marsden and S. D. Ross, Bulletin of the American Mathematical Society 43, 43 (2006). Appendix A: Lagrangian Descriptors The computational tool that we use in this work to explore the template of geometrical structures governing phase space transport is the method of Lagrangian descriptors (LDs). This mathematical technique is a trajectory-based diagnostic that was originally developed 42,67 in the context of Lagrangian transport studies in geophysical uid dynamics . The funda- mental idea behind this methodology is to integrate a positive scalar function along particle 39 trajectories of a dynamical system of general time-dependence in the form: dx = v(x; t) ; x 2 R ; t 2 R ; (A1) dt where the vector eld v(x; t) 2 C (r  1) in x and continuous in time. 42,67 Lagrangian Descriptors were rst introduced in by means of a scalar function, referred to as the function M , to identify distinguished hyperbolic trajectories, i.e. moving saddles, of a dynamical system with general time dependence. In this original approach, the function M was based on the computation of the arclength of a trajectory starting at an initial condition x(t ) = x as it evolves forward and backward for a speci ed time period  > 0. In this 0 0 context, LDs were de ned as: t + M (x ; t ;  ) = kv(x(t; x ); t)k dt ; (A2) 0 0 0 where kk stands for the Euclidean distance and v(x; t) is the vector eld of the dynamical system de ned in Eq. (A1). The connection between the function M and invariant manifolds in phase space has been primarily demonstrated through numerical simulation experiments. Hyperbolic trajectories and their stable and unstable manifolds are revealed by the function M through sharp changes, which we call singular features, in the values of the M eld where the gradient becomes very large and changes abruptly. Moreover, it is important to highlight that if the function M is broken into forward and backward integration: (f) (b) M (x ; t ;  ) = M (x ; t ;  ) + M (x ; t ;  ) 0 0 0 0 0 0 where we have that: Z Z t + t 0 0 (f) (b) M = kv(x(t; x ); t)k dt ; M = kv(x(t; x ); t)k dt 0 0 t t 0 0 (f) (b) then M detects the stable manifolds in phase space and M does the same for unstable manifolds. Other structures such as tori-like invariant manifolds are highlighted due to the relation of the time averages of the function M with the Ergodic Partition Theory . This can be done by de ning the time average: t + M (x ; t ;  ) = kv(x(t; x ); t)k dt ; (A3) 0 0 0 and analyzing its convergence as  ! 1. A detailed description of the detection and visu- 52,67 alization of phase space structures with the function M can be found in . It is important 40 to note here that in the context of chemical reaction dynamics, this de nition of LDs would measure phase space arclength of trajectories. In particular, in the Transition State Theory 44{46,69 literature , con guration space arclength has been used for the computation of LDs. Recently, a rigorous mathematical foundation for Lagrangian descriptors has been estab- lished in by means of introducing an alternative de nition based on the p-norm of the components of the vector eld that de ne the dynamical system in Eq. (A1). Consider the scalar function: t + M (x ; t ;  ) = jv (x(t; x ); t)j dt ; p 2 (0; 1] (A4) p 0 0 i 0 i=1 For this LD one can mathematically prove in certain model problems that hyperbolic points and their stable and unstable manifolds are detected as singularities of the M eld, that is, points in which the function is non-di erentiable. Therefore, this alternative de nition provides a characterization for the concept of singular features. Moreover, tori-like invariant structures are also related to time averages of M by means of the Ergodic Partition Theory. An important aspect to highlight from the alternative de nition given in Eq. (A4) is that it allows to decompose the phase space analysis by separating the integral into the di erent DoFs of the system under study, making it possible to isolate and assess their elliptic and hyperbolic dynamical contributions separately. This property has been shown to be rele- vant for the detection of unstable periodic orbits in Hamiltonian system with two DoF for the classical H enon-Heiles system . Furthermore, this property has been used recently to address the Barbanis system with three DoF , and also to mathematically prove that for the normal form of a Hamiltonian system with three DoFs with an rank-1 saddle, the NHIM and its stable and unstable manifolds are located at the singularities, which are also local minima, of the LD eld. Hence, it is shown that if we decompose Eq. (A4) into forward and backward integration as we also did for Eq. (A2), for a suciently large integration time we have: u (b) s (f) W (x ; t ) = argmin M (x ; t ;  ) ; W (x ; t ) = argmin M (x ; t ;  ) (A5) 0 0 0 0 0 0 0 0 p p u s where W and W are, respectively, the unstable and stable manifolds calculated at time t and argmin denotes the phase space coordinates x that minimize the function M . In 0 0 p addition, the NHIM at time t can be calculated as the intersection of the stable and unstable manifolds: u s N (x ; t ) = W (x ; t )\W (x ; t ) = argmin M (x ; t ;  ) (A6) 0 0 0 0 0 0 p 0 0 41 At this stage, we would like to point out other relevant properties reated to the method of Lagrangian descriptors. First, the integration parameter  that appears in the de nition of both M and M plays a critical role in the identi cation of the underlying dynamical structures in phase space. For applications, the value of  is chosen so that the dynamical history of particle trajectories covers all signi cant timescales for the problem under study. Therefore, there is no general rule on how to choose the appropriate value for  in order to achieve the successful visualization of phase space structures. Consequently, all dynamical systems have to be investigated on a case by case basis by means of trial and error simula- tions. Observe that small values of  would yield blurry phase space structures, which are barely recognizable, since this regime resembles the Eulerian (instantaneous) description of the ow. On the other hand, large  values result in a richer and more complex geometrical description of phase space structures, because LDs resolve them in great detail, making the task of interpreting transport from the obained picture dicult. So there is always a compromise between these two situations. Another interesting feature of LDs is that its computational implementation is straightforward, even for high dimensions, and that the method can be parallelized easily to run on high performance computers, because it deals with initial conditions separately. But probably the most important capability of LDs in all its variants is that it provides us with a high-resolution methodology to explore and visualize high-dimensional phase space dynamics. The successful results obtained by applying this tool to uncover the phase space geometry are in part vindicated from its naive approach of emphasizing initial conditions of particle trajectories rather than focusing on their precise location and long-term evolu- tion. As a result of this crucial point, it o ers tremendous advantages for the analysis of high-dimensional phase space, where the evolution of ensembles of initial conditions may yield trajectories that become disperse and \lost" with respect to each other, making the interpretation of phase space structures problematic and challenging. The goal of identifying phase space structures for high-dimensional systems with LDs is achieved by analyzing the 49{51,70 behavior of initial conditions over low-dimensional phase space slices . In fact, any low-dimensional surface can be selected as a probe, which can be sampled with arbitrary high resolution by de ning an adequate grid of initial conditions. Therefore, no resolution will be lost in the analysis, as the trajectories corresponding to these initial conditions evolve in time, since phase space structure is encoded in the initial conditions of the trajectories 42 themselves. 46,51 In recent studies it has been highlighted that computing xed-time LDs, that is, integrating all initial conditions chosen on a phase space surface for the same integration time , could give rise to some issues. First, it can obscure the detection of the NHIM, which is a crucial step, for instance in Transition State Theory, in order to determine chemical reaction rates by analyzing ux across the dividing surface constructed from the NHIM. This diculty appears as a result of bounded trajectories recrossing the barrier region that surrounds the NHIM, due to the Poincar e recurrence theorem. Consequently, multiple minima and singularities occur in the LD plots, which makes the location of the true NHIM trajectories a challenging task. Another issue, which takes place in the saddle-node bifurcation problem that we study in this paper, is that some of the trajectories that escape the potential well of the PES can go to in nity in nite time. The trajectories that show this behavior will give NaN values in the LD scalar eld, hiding some regions of the phase space, and therefore obscuring the detection of invariant manifolds. In order to illustrate this problem, we have calculated the xed-time M function given in Eq. (A4) with p = 1=2 for the 1 DoF Hamiltonian described by Eq. (11) using the model parameters  = 0:25 and = 2. The results are shown in Fig. 14 for di erent values of the integration time. In Fig. 14A, which corresponds to the value  = 3, we can clearly see how LDs detect the hyperbolic xed point at the origin and the elliptic point at (1=2; 0). Moreover, the stable and unstable manifolds that originate from the saddle point are barely visible, as well as the homoclinic orbit. This is a consequence of the integration time being small. However, at regions on the left part of the phase space, corresponding to initial conditions that have escaped to in nity in nite time, start to appear as a result of NaN values in the LD. This e ect is emphasized further in Fig. 14B, where we have calculated xed-time LDs using  = 5. As we increase the integration time more trajectories escape to in nity and, consequently, a larger region of phase space disappears in the LD picture. In order to circumvent these issues, we will apply in this work the approach that has 46,51 been recently adopted in the literature known as variable time Lagrangian Descriptors. In this methodology, LDs are calculated for a given initial condition until the trajectory leaves a certain barrier (or saddle) region R de ned in the phase space. Therefore, the total integration time in this strategy depends on the initial condition itself, that is  (x ). In this 43 A) B) FIG. 14: Fixed-time M function calculated with p = 1=2 for the 1 DoF Hamiltonian described by Eq. (11) using the model parameters  = 0:25 and = 2. Panel A corresponds to  = 3 and B is calculated for  = 5. formulation, the p-norm de nition of LDs has the form: Z + t + x X M (x ; t ;  ) = jv (x(t; x ); t)j dt ; p 2 (0; 1] : (A7) p 0 0 i 0 0 x 0 i=1 and, for a xed integration time  , the total integration time is de ned as: n o = min  ; t (A8) 0 x(t ;x )2R = where t and t are the times for which the trajecory leaves the barrier region R in forward and backward time, respectively. Since the origin is an rank-1 sddle for the Hamiltonian models that we are using to analyze saddle-node bifurcations, we choose for the barrier region: R = x = (q; x; p; p ) 2 R jqj < 15 ; jpj < 15 (A9) To conclude, it is important to point out here that if the selected barrier region is large enough, the variable time LD de nition given above in Eq. (A7) will approach the xed- time LD de nition in Eq. (A7). Normally hyperbolic invariant manifolds and their stable and unstable manifolds will be captured by the phase space points for which the LD is non-di erentiable and tori-like structures can be determined from time averages of LDs. Moreover, if the barrier region is very small, the NHIM and its invariant stable and unstable manifolds will appear as local maxima in the LD eld, while for large barrier regions the local minimum behavior given in Eqs. (A5) and (A6) is recovered. Consequently, the variable 44 integration time LD provides us with a suitable methodology to study the phase space geometrical structures that characterize the type of saddle-node problems that we discuss in this work, since it avoids the issue of trahectories escaping from the potential well of the PES to in nity in nite time. A detailed analysis on the theoretical background on the variable integration time Lagrangian Descriptors technique will be carried out in future work. Appendix B: Visualization of phase space structure for the quadratic normal form Hamiltonian Consider the normal form for a quadratic Hamiltonian system with two DoF given by 2 2 2 2 H (q ; q ; p ; p ) = (p q ) + (q + p ) ;  ; ! > 0 (B1) 1 1 1 2 1 1 2 2 2 2 | {z } | {z } H H r b where Hamilton's equations are @H >q_ = =  p 1 1 > @p > @H p _ = =  q < 1 1 @q (B2) @H q_ = = ! p > 2 2 > @p @H :p _ = = ! q 2 2 @q The equilibirum point is located at (0; 0; 0; 0) and has zero energy. It is trivial to check that the eigenvalues of the linearized system about the equilibrium point are  and ! i, and hence the equilibrium point is of saddlecenter stability type, which is known as a rank-1 saddle. In this form, the Hamiltonian (B1) is decoupled into the \reactive" mode given by H and the \bath" mode given by H . For this reason, it is known as a separable quadratic r b Hamiltonian (SQH). This representation allows us to address the phase space structures and discuss the distribution of the total energy of the system between the two modes in uncoupled coordinates. In this form, a chemical reaction is said to have occurred when the q coordinate of a trajectory changes sign and thus, an isoenergetic, H = H dividing 1 0 surface (DS) can be de ned by the q = 0 hypersurface. The constant energy de nes a three-dimensional energy surface in the four dimensional phase space given by 2 2 2 2 p q + p + q = H + H = H > 0 ; H > 0 ; H  0 (B3) r b 0 r b 1 1 2 2 2 2 45 The dividing surface, q = 0, for a constant energy is 2 2 2 p + p + q = H + H = H > 0 ; H > 0 ; H  0 (B4) r b 0 r b 1 2 2 2 2 which is a two dimensional surface, and has the geometry S , that is, a 2-sphere on the three dimensional energy surface. Thus, it is codimension-1 and partitions the energy surface into reactant p q > 0 and product p q < 0 regions by the forward and backward \reaction" 1 1 1 1 dividing surfaces as shown in Ref. , which are given by 2 ! 2 2 (B5) p =  H (p + q ) ; forward/backward DS 1 0 2 2 The forward and backward DS join at p = 0 giving n o 4 2 2 N (H ) = (q ; q ; p ; p ) 2 R j q = p = 0 ; p + q = H  0 ; NHIM (B6) 0 1 2 1 2 1 1 0 2 2 which is of geometry S , that is a circle centered at the origin with radius 2H =! in the (q ; p ) plane. This is a normally hyperbolic invariant manifold (NHIM) associated with the 2 2 rank-1 saddle and parametrized by total energy H (q ; q ; p ; p ) = H . Invariance follows 1 2 1 2 0 from the vector eld (B2), since when we have that q = p = 0, this gives q_ = p _ = 0. 1 1 1 1 Thus q and p always remain zero, and trajectories with these initial conditions remain on 1 1 the NHIM, that is, q = p = 0 is invariant. It is normally hyperbolic since the directions 1 1 normal to the NHIM, that is, the (q ; p ) plane, have linear saddle-like dynamics. For a 1 1 two DoF system, this NHIM is more commonly referred to in the literature as an unstable periodic orbit and is shown in Fig. 15. In order to understand the relationship between the NHIM and the rank-1 saddle point, we note that for H = H = 0 the NHIM reduces to the point (q ; q ; p ; p ) = (0; 0; 0; 0), r b 1 2 1 2 which is the rank-1 saddle point on the energy surface H = 0. Moreover, recall that this rank-1 saddle equilibrium point is a con guration space concept and is located on the potential energy surface. Therefore, as the total energy of the system is increased from 0, with the bath mode energy H increasing from zero, the NHIM \grows" from the index- 1 saddle point on the zero energy surface into an invariant 1-sphere. This shows how the \in uence" of the rank-1 saddle point is carried to higher energy sufaces on which the saddle point does not exist. The stable and unstable manifolds of the NHIM (B6) are given by u 4 2 2 W = (q ; q ; p ; p ) 2 R j q = p ; p + q = H > 0 (B7) 1 2 1 2 1 1 b 2 2 s 4 2 2 W = (q ; q ; p ; p ) 2 R j q = p ; p + q = H > 0 (B8) 1 2 1 2 1 1 b 2 2 46 1 which are two-dimensional surfaces and have the geometry of RS for a xed energy. Thus, the codimension-1 geometry of the manifolds partition the phase space into \reactive" and \non-reactive" trajectories as shown in Fig. 15. A) B) C) FIG. 15: Phase space structures in the neighborhood of the bottleneck for the uncoupled (" = 0) Hamiltonian system (19). A) Linearized dynamics about the rank-1 saddle in the saddle space corresponding to the reactive DoF; B) Dynamics in the center space associated to the harmonic oscillator DoF; C) Description of the phase space bottleneck region obtained for an energy of the system above that of the rank-1 saddle (the barrier of the PES). The stable and unstable manifolds of the NHIM act as conduits connecting reactants (well region) and products (escape to in nity) and characterize reaction dynamics. 47 Appendix C: Computation of NHIM and its invariant manifolds for the 2 DoF system In this aapendix we describe the steps followed in order to calculate the NHIM and its stable and unstable invariant manifolds associated with the rank-1 saddle located at the origin for the Hamiltonian system with 2 DoF given by Eq. (19), which models the saddle- node bifurcation phenomena in phase space. Step 1: Select an excess energy above the critical value. The Lyapunov subcenter theorem tells us that when the energy of the Hamiltonian system is raised above that of the equilibrium point at the origin of saddlecenter (rank-1 saddle) stability, which is known 2N3 as critical energy, a family of NHIM with geometry S bifurcates, where N indicates the number of DoF. In our problem, N = 2 so the topology of the NHIM is S , that is, the NHIM is an unstable periodic orbit (UPO). So we pick a value for the total energy H greater than the critical value H = 0, which gives an excess energy H = H H = H . c 0 c 0 This total energy is also the energy of trajectories on the invariant manifolds which partition the phase space into nonreactive and reactive trajectories at the same energy. The excess energy can be an arbitrary value, up to the value at which the NHIM bifurcates. Step 2: Obtain the NHIM at the selected excess energy. As we have discussed, for the 2 DoF problem the NHIM is an UPO which touches the equipotential contour corresponding to the selected energy at the bottleneck region. However, due to the saddle- like dynamics in the transverse directions, any numerical error in the computation of the periodic orbit will get exponentially ampli ed with each time step, which will eventually destroy the periodic nature of the trajectory, since it will fail to come back to its starting point. In order to generate the desired unstable periodic orbit corresponding to a selected excess energy, we consider a procedure that starts with a small initial condition (\seed") obtained from the linearized equations of motion near the rank-1 saddle, and uses di erential cor- 60,72,73 rection and numerical continuation on that initial guess. The result is an unstable periodic orbit at target energy H of period T which will be close to 2=! , where ! i 0 " " is the pair of imaginary eigenvalues of the linearization about the rank-1 saddle point. In order to choose an initial guess for the search of the UPO we can use the linearized equations of motion about the rank-1 saddle equilibrium point x = 0. The Jacobian matrix at the 48 origin is given by Eq. (26) and the general solution to the linearized dynamical system can be written as: e  t  t i! t " " " x(t) = x + C e v + C e v + 2Re e w (C1) 1 + 2 + where C ; C 2 R and  =  +  i 2 C. The real eigenvalues  with corresponding 1 2 1 2 " eigenvectors v ; v are described in Eqs. (31) and (34) and the pair of complex eigenvalues ! i with corresponding eigenvectors w ; w are described in Eqs. (31) and (35). The idea " + is to use the complex eigenvalues and their corresponding eigenvectors to obtain a starting guess as an initial condition to initialize the method that searches for the UPO. To do so, we choose a small amplitude, A  2  10 , periodic orbit in the center manifold of the linearized system by selecting in Eq. (C1) the values  = A =2 (this eliminates the factor 2 in the formula), C = C = 0 and t = 0. Thus, the initial guess is: 1 2 A " x  = (q ; x ; 0; 0) = x + 2Re(w ) = ;A ; 0; 0 (C2) 0;g 0;g 0;g + x " ( + ! ) 0 " which has a period of T = 2=! . 0;g " Step 3: Di erential correction of the initial guess. The di erential correction procedure that we apply in this step to the initial guess will only yield a good approximation to the true UPO of the nonlinear Hamiltonian system whenever A  1, which corresponds to a much smaller excess energy H than the one we originally selected in step 1. The reason for this is that the construction of the initial guess is based on the linear approximation near the rank-1 saddle equilibrium point. The convergence criterion that we use is based on the basic property that a periodic orbit returns to the starting point after a given period T , which is in fact the period of the PO. If x  = x  (0) is a true initial condition on the PO 0 po x of period T , the convergence is checked using the condition: po kx  (T ) x  (0)k <  (C3) po po for some tolerance   1. In this approach, we hold the con guration coordinate q constant, while applying correction to the x con guration coordinate of the initial guess. While the p momentum coordinate is used as a stopping criterion for the di erential correction procedure, the p momentum coordinate is used as a terminating event (crossing the p = 0 plane in phase space) for the integration. It is to be noted that this combination of coordinates is suitable for the structure of the initial guess at hand, so for other problems or other forms of the 49 initial guess, would require some permutation of the phase space coordinates to achieve a stable implementation. Let us denote the ow map of a di erential equation x _ = f (x) with initial condition x(t ) = x by (t; x ). Thus, the displacement of a reference trajectory x (t) after a time t 0 0 0 becomes: x (t + t) = (t + t; x  + x  ) (t; x  ) : (C4) 0 0 0 Thus, a Taylor expansion of the displacement after t + t gives: 1 1 @(t ; x  ) @(t ; x  ) 1 0 1 0 x (t + t ) = x  + t + h:o:t (C5) 1 1 0 1 @x @t 0 1 where the rst term on the right hand side is the state transition matrix, (t ; t ), evaluated 1 0 along the reference trajectory initialized at t . The state transition matrix along a trajec- tory is obtained from the numerical solution of the variational equations along with the Hamilton's equation of motion . Suppose that we want to land at a point x (this would be the starting initial condition for a periodic orbit), after an integration time interval t , and starting from the initial guess x  , then we have: 0;g x (t ) = (t ; x  ) = x  = x x  (C6) 1 1 0;g 1 d 1 where the error x  is the applied rst order correction obtained from the state transition matrix evaluated along the trajectory with initial condition x  and integrated for t time 0;g 1 units. For the rank-1 saddle equilibrium point under consideration, we initialize the guess as obtained in Eq. (C2) with x  . Using numerical integration, the initial condition is 0;g integrated until a p = 0 event occurs with a high speci ed tolerance (typically 10 ). This results in x(t ), which for the guess initial condition denotes the half-period location t = T =2. Then, we evaluate the state transition matrix (t ; 0) for the trajectory obtained 1 0;g 1 from the guess intial condition. Now, this is used to correct the initial value of x while 0;g keeping q constant and iterating until p = 0. Since the x con guration coordinate is kept 0;g x constant, the rst order correction is given by: p =  x + p _ t + h:o:t 1 32 0 1 1 p =  x + p _ t + h:o:t x 42 0 x 1 1 1 50 th where  is the (i; j) entry of (t ; 0) and the acceleration terms come from the equations ij 1 of motion evaluated at the crossing t = t when p = p = 0. Thus, we obtain the rst 1 1 1 order correction x as: p _ x    p ; x ! x x (C7) 0 42 32 x 0 0 0 p _ which is iterated until jp j = jp j <  for some tolerance , since we want the nal x x 1 1 point of the periodic orbit to be of the form x  = (q ; x ; 0; 0). We remark that in all t 1 1 this process, di erential correction assumes that the guess periodic orbit has a small error (for example in this system, of the order of 10 ) and can be corrected using rst order form of the correction terms. If, however, large a corrective step is used, the half-orbit overshoots between successive steps leading to failure in converging to a closed orbit. Once the appropriate conditions are chosen, di erential correction generates a family of periodic orbits and takes 2-3 iterations per unstable periodic orbit. Step 4: Numerical continuation to the UPO at the selected excess energy. The procedure described above yields an initial condition for an unstable periodic orbit from an initial guess. Since our initial guess came from the linearization near the rank-1 saddle equilibrium pointgiven by (C1), we can use this procedure for small amplitudes of order 2 10 . We remark that this procedure is based on computations presented for this problem and will vary for a di erent nonlinear system. This small amplitude corresponds to small excess energy, typically of the order 10 , and to obtain the unstable periodic orbit of arbitrarily large amplitude, we resort to numerical continuation for generating a family of periodic orbits that reach the selected excess energy. This procedure starts with the initial conditions for two nearby unstable periodic orbits of small amplitude to obtain an initial guess for the next periodic orbit. The initial guess obtained by a simple extrapolation can then be corrected using di erential correction. To this end, we proceed as follows. Suppose we nd two nearby small amplitude unstable (1) (2) periodic orbits with initial conditions x and x , correct to within the tolerance d tol 0 0 obtained using the di erential correction procedure described above. We can then generate a family of periodic orbits with increasing amplitudes around x  as: eq (2) (1) = x  x  = (q ; x ; 0; 0) (C8) 0 0 0 0 51 (3) A linear extrapolation to an initial guess of slightly larger amplitude, x is given by: (3) (2) (2) (2) (3) (3) x  = x  +  = (q + q ; x + x ; 0; 0) = (q ; x ; 0; 0) (C9) 0 0 0;g 0 0 0 0 0 Thus, we can use di erential correction on this guess initial condition to compute an accurate (3) (3) solution x  from the initial guess x  and repeat as an iterative step to generate a family 0 0;g of unstable periodic orbits. Next, to compute an unstable periodic orbit at the selected excess energy, we track (k) the energy of each unstable periodic orbit in the family until we have two solutions, x (k+1) and x  , whose energy brackets the selected excess energy H . Then, we can resort to combining a bisection type method with di erential correction on the two periodic orbits until we converge to the desired periodic orbit to within a speci ed tolerance. Thus, the result is an unstable periodic orbit at the selected energy H and is speci ed by the initial condition x  (0) and time period T . po po Step 5: Computation of invariant manifolds of the NHIM. We nd the global ap- proximation to the unstable and stable manifolds of the periodic orbit from the eigenvectors of the monodromy matrix. The local linear approximation of the unstable (or stable) man- ifolds (initial conditions displaced along the saddle space eigenvectors) is integrated using the full nonlinear equations of motion to produce the global approximation of the unstable (or stable) manifolds. This procedure is known as globalization of the manifolds and we proceed as follows. First, the state transition matrix (t) along the periodic orbit with initial condition X can be obtained by numerical integration of the variational equations along with the equations of motion from t = 0 to t = T . This is gives the monodromy matrix M = (T ) po po and its eigenvalues are obtained. For Hamiltonian systems , it is known that the eigenvalues of M are of the form: > 1;  = ;  =  = 1 (C10) 1 2 3 4 The eigenvector associated with eigenvalue  is in the unstable direction, the eigenvector associated with eigenvalue  is in the stable direction. Let e (x  (0)) denote the normalized 2 po stable eigenvector, and e (x  (0)) denote the normalized unstable eigenvector. We can po compute the invariant manifolds by initializing along these eigenvectors as: s=u s=u X (x (0)) = x (0) + e (x (0)) (C11) po po po 52 for the stable/unstable manifold of the desired periodic orbit. Here the small displacement from X is denoted by  > 0 and its magnitude is taken to be small enough for the validity of the linearization, yet not so small that the time of ight becomes too large due to asymptotic nature of the stable and unstable manifolds. It has been suggested that typical values of around 10 correspond to nondimensional position displacements of magnitude around 10 . By numerically integrating the unstable vector forwards in time, using both  and , for the forward and backward branches respectively, we generate trajectories shadowing the two u u branches, W and W . Similarly, by integrating the stable vector backwards in time, using both  and , for forward and backward branches respectively, we generate trajectories shadowing the stable manifold, W . For the manifold at X(t), one can simply use the +; state transition matrix to transport the eigenvectors from X to X(t) u=s u=s X (X(t)) = (t; 0)X (X ) (C12) It is to be noted that since the state transition matrix does not preserve the norm, and hence the resulting vector must be normalized. The globalized invariant manifolds associated with rank-1 saddles are known as Conley-McGehee tubes . These tubes form the impenetrable phase space conduits (codimension-1 barriers) for the trajectories to react by crossing the q = 0 dividing surface. In summary, the computation of invariant manifolds of the unstable periodic orbit associ- ated with the rank-1 saddle begins with the linearized equations of motion. This is obtained after a coordinate transformation to the rank-1 saddle equilibrium point and a Taylor expan- sion of the equations of motion. Keeping the rst order terms in this expansion, we obtain the eigenvalues and eigenvectors of the linearized system. The eigenvectors corresponding to the center subspace provide the starting guess for computing the unstable periodic orbits of small excess energy, H << 1. This iterative procedure performs a small correction to the starting guess for initial condition based on the terminal condition of the periodic orbit until a desired tolerance is satis ed. This procedure is known as di erential correction and generates initial condition for an unstable periodic orbit at small excess energy. Next, a numerical continuation procedure is adopted to follow the small energy (small amplitude) periodic orbit out to high excess energy. Once the unstable periodic orbit is obtained, the globalization of its invariant manifolds is done using the initial conditions along the eigen- 53 vectors of the state transition matrix computed along the unstable periodic orbit. The result of the steps described above is shown in Fig. 9.

Journal

Nonlinear SciencesarXiv (Cornell University)

Published: Jul 7, 2019

There are no references for this article.