This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the pitchfork bifurcation of the folded node and other unbounded time-reversible connection problems in 3\mathbb{R}^{3}

K. Uldall Kristiansen
(Date: February 10, 2025)
Abstract.

In this paper, we revisit the folded node and the bifurcations of secondary canards at resonances μ\mu\in\mathbb{N}. In particular, we prove for the first time that pitchfork bifurcations occur at all even values of μ\mu. Our approach relies on a time-reversible version of the Melnikov approach in [27], used in [28] to prove the transcritical bifurcations for all odd values of μ\mu. It is known that the secondary canards produced by the transcritical and the pitchfork bifurcations only reach the Fenichel slow manifolds on one side of each transcritical bifurcation for all 0<ϵ10<\epsilon\ll 1. In this paper, we provide a new geometric explanation for this fact, relying on the symmetry of the normal form and a separate blowup of the fold lines. We also show that our approach for evaluating the Melnikov integrals of the folded node – based upon local characterization of the invariant manifolds by higher order variational equations and reducing these to an inhomogeneous Weber equation – applies to general, quadratic, time-reversible, unbounded connection problems in 3\mathbb{R}^{3}. We conclude the paper by using our approach to present a new proof of the bifurcation of periodic orbits from infinity in the Falkner-Skan equation and the Nosé equations.

Department of Applied Mathematics and Computer Science,
Technical University of Denmark,
2800 Kgs. Lyngby,
DK

1. Introduction

In slow-fast systems with one fast variable and two slow ones, the folded node pp is a singularity of the slow flow on the fold line of a critical manifold CC. See an illustration in Fig. 1. Upon desingularization pp corresponds to a stable node with eigenvalues λs<λw<0\lambda_{s}<\lambda_{w}<0, and its strong stable manifold υ\upsilon, tangent to the eigenvector associated with λs\lambda_{s}, produces a funnel region on the critical manifold, where orbits approach the singularity tangent to the weak eigendirection (associated with the eigenvalue λw\lambda_{w}). Due to the contraction within the funnel region, the folded node – upon composition with a global return mapping – provides a mechanism for producing attracting limit cycles Γϵ\Gamma_{\epsilon}, see [2]. In fact, a blowup of the folded node reveals one orbit γ\gamma, along which extended versions WcuW^{cu} and WcsW^{cs} of the attracting and repelling critical manifolds, respectively, twist or rotate; the number of rotations being described by μ:=λs/λw>1\mu:=\lambda_{s}/\lambda_{w}>1. The twisting is such that these manifolds intersect transversally whenever μ\mu\notin\mathbb{N}. As a consequence, for these values of μ\mu, there exists a ‘weak canard’ connecting extended versions of the Fenichel slow manifolds. This orbit acts, due to the twisting of WcuW^{cu} along γ\gamma, as a ‘center of rotation’ and trajectories on either side will therefore experience small oscillations before they leave a neighborhood of pp by following its unstable set; in Fig. 1 this unstable set coincides with the positive zz-axis. Consequently, the limit cycles Γϵ\Gamma_{\epsilon} will be of mixed-mode type where small oscillations are followed by larger ones. Such oscillations appear in many applications, perhaps most notably in chemical reaction dynamics, and the folded node has therefore gained glory as a (relatively) simple mathematical model of this phenomenon, see e.g. the review article [4].

Bifurcations of the weak canard occurs whenever μ\mu\in\mathbb{N}; in this case, for ϵ=0\epsilon=0, the twisting of WcuW^{cu} and WcuW^{cu} is such that these manifolds intersect tangentially along γ\gamma. These bifurcations were described for ϵ=0\epsilon=0 by the reference [28], working on the ‘normal form’

x˙=12μy(μ+1)z,y˙=1,z˙=x+z2,\displaystyle\begin{split}\dot{x}&=\frac{1}{2}\mu y-(\mu+1)z,\\ \dot{y}&=1,\\ \dot{z}&=x+z^{2},\end{split} (1.1)

and using the Melnikov approach developed in [27], following [26]. The system (1.1) is related to the blowup of the folded node pp for ϵ=0\epsilon=0. In particular for (1.1), γ\gamma takes the following form

γ:(x,y,z)=(14t2+12,t,12t),\displaystyle\gamma:\,(x,y,z)=\left(-\frac{1}{4}t^{2}+\frac{1}{2},t,\frac{1}{2}t\right), (1.2)

For each odd nn, it was shown that there is a transcritical bifurcation of ‘canards’ connecting Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu). For all 0<ϵ10<\epsilon\ll 1, this bifurcation produces additional transversal intersections of the slow manifold. The resulting new canards – the ‘secondary canards’ – produce bands on the attracting slow manifold where different number of small oscillations occur, see [2, 4, 28]. For any even nn, it was conjectured that a pitchfork bifurcation occurs. This was supported by numerical computations. Furthermore, in [18, App. A] a way was found to compute a ‘third order’ Melnikov integral using Mathematica for all even nn and explicit computations demonstrated that the integral was nonzero for even values of nn up to 2020. Following the work of [28] this also shows that a pitchfork bifurcation occurs, at least for these values.

In this paper, we prove the pitchfork bifurcation for every even nn by evaluating the third order Melnikov integral analytically. Our approach is based upon a time-reversible version of the Melnikov theory of [27]. However, the most important insight of this paper is to characterize the manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) locally by solutions to higher order variational equations; this is in contrast to [28] which uses an integral representation of these manifolds. We show that this approach, relying on reducing these variational equations to an inhomogeneous Weber equation, extends to a very general class of time-reversible, quadratic, unbounded connection problems in 3\mathbb{R}^{3}. Regardless, for the folded node the ‘time-reversible approach’ also allows us to provide a more detailed blowup picture of the folded node, including a rigorous description of the additional transverse intersections of WcsW^{cs} and WcuW^{cu} that arise due to the pitchfork bifurcation.

The bifurcations of canards for (1.1), is closely related to bifurcations of periodic orbits from heteroclinic cycles at infinity. The Falkner-Skan equation

x′′′+x′′x+μ(1x2)=0,\displaystyle x^{\prime\prime\prime}+x^{\prime\prime}x+\mu(1-x^{\prime 2})=0, (1.3)

and the Nosé equation:

x˙\displaystyle\dot{x} =yxz,\displaystyle=-y-xz, (1.4)
y˙\displaystyle\dot{y} =x,\displaystyle=x,
z˙\displaystyle\dot{z} =μ(1x2),\displaystyle=\mu(1-x^{2}),

are well-known examples of systems (without equilibria) possessing such bifurcations, see e.g. [23], and [17] for other examples. The Falkner-Skan equation (1.3) initially appeared in the study of boundary layers in fluid dynamics, see [6]. In this context, the physical relevant parameter regime is μ(0,2)\mu\in(0,2). However, the equation has subsequently been studied by other authors [22, 21, 23] for all μ>0\mu>0 on the basis of the rich dynamics it possesses (including chaotic dynamics and a novel bifurcation of periodic orbits from infinity). On the other hand, the Nosé equations (1.4) model the interaction of a particle with a heat bath [19]. The system also has interesting dynamics without any equilibrium and possesses many similar properties to (1.1) and (1.3). Nevertheless, the description of the bifurcating periodic orbits in both of these systems, is – as noted by [23] – long and cumbersome, and to a large extend, independent of standard methods of dynamical systems theory. Therefore, although the folded node will be our primary focus, a subsequent aim of the paper, is to apply the Melnikov theory, and our classification of WcsW^{cs} and WcuW^{cu} through solutions of an inhomogeneous Weber equation, to these bifurcations and present a simpler description of the emergence of periodic orbits, based on normal form theory and invariant manifolds and therefore more in tune with dynamical systems theory.

1.1. The folded node: Further background

Following [28, Proposition 2.1], any folded node can be brought into the ‘normal form’:

x˙=ϵ(12μy(μ+1)z+𝒪(x,ϵ,(y+z)2)),y˙=ϵ,z˙=x+z2+𝒪(xz2,z3,xyz)+ϵ𝒪(x,y,z,ϵ),\displaystyle\begin{split}\dot{x}&=\epsilon\left(\frac{1}{2}\mu y-(\mu+1)z+\mathcal{O}(x,\epsilon,(y+z)^{2})\right),\\ \dot{y}&=\epsilon,\\ \dot{z}&=x+z^{2}+\mathcal{O}(xz^{2},z^{3},xyz)+\epsilon\mathcal{O}(x,y,z,\epsilon),\end{split} (1.5)

by only using scalings, translations and a regular time transformation. Here μ:=λs/λw>1\mu:=\lambda_{s}/\lambda_{w}>1, and the critical manifold CC is approximately given by the parabolic cylinder x=z2x=-z^{2}, z<0z<0 (CaC_{a}) being stable and z>0z>0 (CrC_{r}) being unstable. See Fig. 1. Here CaC_{a} is in blue, CrC_{r} is in red, whereas the degenerate line F:x=z=0F:\,x=z=0, being the fold line, is in green. For (1.5), the folded node pp (pink), on FF, is at the origin. Furthermore, if we for simplicity ignore the 𝒪\mathcal{O}-terms in (1.5), then the reduced problem on CC is given by

y\displaystyle y^{\prime} =1,\displaystyle=1,
2zz\displaystyle 2zz^{\prime} =12μy+(μ+1)z.\displaystyle=-\frac{1}{2}\mu y+(\mu+1)z.

Consider CaC_{a} where z<0z<0. Then multiplication of the right hand side by 2z-2z gives the topologically equivalent system

y=2z,z=12μy(μ+1)z,\displaystyle\begin{split}y^{\prime}&=-2z,\\ z^{\prime}&=\frac{1}{2}\mu y-(\mu+1)z,\end{split} (1.6)

on CaC_{a}, see [28]. The point (y,z)=(0,0)(y,z)=(0,0) is then a stable node of these equations with eigenvalues 1-1 and μ-\mu and associated eigenvectors:

(2,1)T,\displaystyle(2,1)^{T}, (1.7)

and (2,μ)T(2,\mu)^{T}, respectively. See illustration of the reduced flow in Fig. 2. Notice that the orbits on CrC_{r}, where z>0z>0, are also orbits of (1.6), but their directions have to be reversed.

Refer to caption
Figure 1. The folded node singularity pp. Upon desingularization of the reduced problem, the folded node singularity becomes a stable node. The strong eigenvector associated with the node, gives rise to a strong stable manifold υ\upsilon (orange) that forms a boundary of a funnel region (shaded), bounded on the other side by FF, where trajectories approach the folded node pp (in finite time before desingularization), tangent to a weak eigendirection. For the system (1.5) without the 𝒪\mathcal{O}-terms, the weak eigenvector also produces an invariant space and an orbit γ\gamma, which we show in purple.
Refer to caption
Figure 2. The reduced flow on CC, recall Fig. 1, projected onto the (y,z)(y,z)-plane. The strong canard υ\upsilon is shown in purple, whereas the weak canard γ\gamma, obtained from (1.5) upon ignoring the 𝒪\mathcal{O}-terms, is shown in orange.

Blowup analysis

Compact submanifolds (with boundaries) SaS_{a} and SrS_{r} of CaC_{a} and CrC_{r}, respectively, bounded away from the fold line, perturb by Fenichel’s theory to attracting and repelling slow manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon} for all 0<ϵ10<\epsilon\ll 1, see [7, 8, 9, 12]. We will refer to these manifolds as ‘Fenichel’s (slow) manifolds’. They are nonunique but 𝒪(ec/ϵ)\mathcal{O}(e^{-c/\epsilon})-close. Extended versions of these invariant manifolds up close to the folded node pp is obtained in [24] by blowing up the point (x,y,z)=0(x,y,z)=0 for ϵ=0\epsilon=0. In further details, the authors apply the following blowup transformation :[0,r0)×S34\mathcal{B}:[0,r_{0})\times S^{3}\rightarrow\mathbb{R}^{4}, given by

r[0,r0),(x¯,y¯,z¯,ϵ¯)S3{x=r2x¯,y=ry¯,z=rz¯,ϵ=r2ϵ¯,\displaystyle r\in[0,r_{0}),\,(\bar{x},\bar{y},\bar{z},\bar{\epsilon})\in S^{3}\mapsto\begin{cases}x&=r^{2}\bar{x},\\ y&=r\bar{y},\\ z&=r\bar{z},\\ \epsilon&=r^{2}\bar{\epsilon},\end{cases} (1.8)

to the extended system ((1.5),ϵ˙=0)(\mbox{(\ref{eq:fnnfeps})},\dot{\epsilon}=0). For this extended system, x=y=z=ϵ=0x=y=z=\epsilon=0 is fully nonhyperbolic – its linearization having only zero eigenvalues – but upon blowup (1.8), we gain hyperbolicity of r=0r=0 after desingularization through division of the resulting right hand sides by rr. In particular, setting x¯=1\bar{x}=-1 in (1.8) produces the following local form of (1.8)

(r1,y1,z1,ϵ1){x=r12,y=r1y1,z=r1z1,ϵ=r12ϵ1.\displaystyle(r_{1},y_{1},z_{1},\epsilon_{1})\mapsto\begin{cases}x&=-r_{1}^{2},\\ y&=r_{1}y_{1},\\ z&=r_{1}z_{1},\\ \epsilon&=r_{1}^{2}\epsilon_{1}.\end{cases} (1.9)

The local coordinates (r1,y1,z1,ϵ1)(r_{1},y_{1},z_{1},\epsilon_{1}) provide a coordinate chart ‘x¯=1\bar{x}=-1’, covering [0,r0)×Sx¯<03[0,r_{0})\times S_{\bar{x}<0}^{3} where Sx¯<03:=S3{x¯<0}S_{\bar{x}<0}^{3}:=S^{3}\cap\{\bar{x}<0\}. Here r1=0r_{1}=0 corresponds to r=0r=0. In this chart, one gains hyperbolicity of CaC_{a} and CrC_{r} for r=0r=0 upon division of the right hand side by r1r_{1}. By center manifold theory, this then enables an extension of the Fenichel slow manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon} as the \mathcal{B}-image of ϵ=\epsilon=const. sections of three-dimensional invariant manifolds MaM_{a} and MrM_{r}, respectively, for all 0<ϵ10<\epsilon\ll 1. Following [16], we shall abbreviate these extended manifolds in the (x,y,z)(x,y,z)-space by Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}}, respectively; see [24] for further details.

Remark 1.1.

It is important to highlight that, due to the contraction towards the weak canard, the forward (backward) flow of the Fenichel manifold Sa,ϵS_{a,\epsilon} (Sr,ϵS_{r,\epsilon}, respectively) is only a subset of Sa,ϵS_{a,\sqrt{\epsilon}} (Sr,ϵS_{r,\sqrt{\epsilon}}). Therefore when we intersect Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}}, extended by the forward and backward flow, it does not follow directly that the Fenichel manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon} also intersect.

Notice that the blowup approach, following (1.9) and the conservation of ϵ\epsilon, provide control of Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}} up to 𝒪(ϵ)\mathcal{O}(\sqrt{\epsilon})-close to the folded node pp, justifying the use of the subscripts. To describe these manifolds beyond this, we have to look at the (scaling) chart obtained by setting ϵ¯=1\bar{\epsilon}=1. This produces the following local blowup transformation

(r2,x2,y2,z2){x=r22x2,y=r2y2,z=r2z2,ϵ=r22.\displaystyle(r_{2},x_{2},y_{2},z_{2})\mapsto\begin{cases}x&=r_{2}^{2}x_{2},\\ y&=r_{2}y_{2},\\ z&=r_{2}z_{2},\\ \epsilon&=r_{2}^{2}.\end{cases} (1.10)

using the chart-specified coordinates (x2,y2,z2,r2)(x_{2},y_{2},z_{2},r_{2}). The corresponding coordinate chart ‘ϵ¯=1\bar{\epsilon}=1’ covers [0,r0)×Sϵ¯>03[0,r_{0})\times S_{\bar{\epsilon}>0}^{3} where Sϵ¯>03:=S3{ϵ¯>0}S_{\bar{\epsilon}>0}^{3}:=S^{3}\cap\{\bar{\epsilon}>0\}. By inserting (1.10) into (1.5), dividing the right hand side by r2r_{2} and subsequently setting r2=ϵ=0r_{2}=\sqrt{\epsilon}=0, we obtain (1.1), repeated here for convenience:

x˙=12μy(μ+1)z,y˙=1,z˙=x+z2.\displaystyle\begin{split}\dot{x}&=\frac{1}{2}\mu y-(\mu+1)z,\\ \dot{y}&=1,\\ \dot{z}&=x+z^{2}.\end{split} (1.11)

In (1.11), we have also dropped the subscripts on (x2,y2,z2)(x_{2},y_{2},z_{2}). Two explicit algebraic solutions are known for this unperturbed system, one:

υ:(x,y,z)=(μ24t2+μ2,t,μ2t)\displaystyle\upsilon:\,\,(x,y,z)=\left(-\frac{\mu^{2}}{4}t^{2}+\frac{\mu}{2},t,\frac{\mu}{2}t\right)

corresponding to the ‘strong canard’, while γ\gamma in (1.2), repeated here for convenience:

γ:(x,y,z)=(14t2+12,t,12t),\displaystyle\gamma:\,(x,y,z)=\left(-\frac{1}{4}t^{2}+\frac{1}{2},t,\frac{1}{2}t\right), (1.12)

corresponds to the ‘weak canard’, which we will focus on in this paper.

Remark 1.2.

Notice that the projection of (1.12) onto the (y,z)(y,z)-plane coincides with the span of the weak eigenvector (1.7), explaining the use of ‘weak’ in ‘weak canard’. Also, the orbit (1.12) is unique as an orbit on the blowup sphere with these properties. This is obviously in contrast with reduced flow on CaC_{a} where all trajectories within the funnel is assumption to the weak canard.

On a related issue, notice we abuse notation slightly: Most often, γ\gamma will refer to (1.12) as an orbit of (1.11). But by the coordinate chart ‘ϵ¯=1\bar{\epsilon}=1’, this orbit also becomes a heteroclinic connection on r=0,Sϵ¯03r=0,\,S^{3}_{\bar{\epsilon}\geq 0}, connecting partially hyperbolic points on the equator ϵ¯=0\bar{\epsilon}=0 for the blowup system. We will use the same symbol for this orbit. At the same time, in Fig. 2 we also use the symbol γ\gamma to highlight the weak eigendirection of the folded node as an attracting node of the desingularized reduced problem on CaC_{a}. A similar misuse of notation occurs for υ\upsilon.

Restricting the center manifolds MaM_{a} and MrM_{r}, obtained in the chart x¯=1\bar{x}=-1, to r=0r=0 we obtain, when writing the result in the chart ϵ¯=1\bar{\epsilon}=1, center-stable Wcs(μ)W^{cs}(\mu) and center-unstable manifolds Wcu(μ)W^{cu}(\mu) of (1.11) and z±z\rightarrow\pm\infty, respectively, consisting of solutions that grow algebraically as t±t\rightarrow\pm\infty, respectively. Following [24], a simple calculation shows that Wcs(μ)W^{cs}(\mu) takes the local form:

Wloccs(μ):x=z2+12(μ+1)14μyz1+z2m(yz1,z2),\displaystyle W_{loc}^{cs}(\mu):x=-z^{2}+\frac{1}{2}(\mu+1)-\frac{1}{4}\mu yz^{-1}+z^{-2}m(yz^{-1},z^{-2}), (1.13)

for all zz sufficiently large and some smooth m:I×[0,δ]m:I\times[0,\delta]\rightarrow\mathbb{R} for an appropriate interval II\subset\mathbb{R} and δ>0\delta>0 sufficiently small. Due to the invariance of υ\upsilon and γ\gamma, mm also satisfies m(2,z2)=m(2/μ,z2)=0m(2,z^{-2})=m(2/\mu,z^{-2})=0. By using the time-reversible symmetry (x,y,z,t)(x,y,z,t)(x,y,z,t)\mapsto(x,-y,-z,-t) of (1.11), a simple calculation shows that the manifold Wcu(μ)W^{cu}(\mu) takes an identical form, with the expression in (1.13) now valid for all zz sufficiently negative. We illustrate the results of the blowup analysis in Fig. 3. See figure caption for further description. Guiding these manifolds along υ\upsilon and γ\gamma one obtains the global manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu). In particular, by considering the variational equations of (1.11) along γ\gamma the following was shown in [24].

Lemma 1.3.

Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) intersect transversally along γ\gamma if and only if μ\mu\notin\mathbb{N}.

By regular perturbation theory, the extension of the slow manifolds Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}} by the flow are therefore smoothly 𝒪(r2=ϵ)\mathcal{O}(r_{2}=\sqrt{\epsilon})-close to Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu), respectively, in compact subsets of the chart ϵ¯=1\bar{\epsilon}=1. Hence, as a consequence of Lemma 1.3, for every μ\mu\notin\mathbb{N} there exist a transverse intersection of Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}} which is 𝒪(ϵ)\mathcal{O}(\sqrt{\epsilon})-close to γ\gamma in fixed compact subsets of the scaling chart. In general, recall Remark 1.1, it seems that there do not exist any results on how far this perturbed ‘weak canard’ extends and whether ‘it’ (being nonunique) actually reaches the true Fenichel slow manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon}. The situation is different for υ\upsilon. First and foremost, Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) always intersect transversally along this orbit. Consequently, υ\upsilon always perturbs as a ‘strong maximal canard’ for all 0<ϵ10<\epsilon\ll 1, and this perturbed version always reaches the Fenichel manifolds. This latter property is a consequence of the repelling nature of υ\upsilon, ‘forcing’ Sa,ϵS_{a,\sqrt{\epsilon}} (Sr,ϵS_{r,\sqrt{\epsilon}}) and the forward (backward) flow Sa,ϵS_{a,\epsilon} (Sr,ϵS_{r,\epsilon}, respectively) to coincide near this object.

Refer to caption
Figure 3. Illustration of the blowup of pp for ϵ=0\epsilon=0 to a hemisphere Sϵ¯03:=S3{ϵ¯0}S^{3}_{\bar{\epsilon}\geq 0}:=S^{3}\cap\{\bar{\epsilon}\geq 0\}. In this figure, we represent r=0r=0, Sϵ¯0rS^{r}_{\bar{\epsilon}\geq 0} – by projection – as a solid ‘ball’ in the (x¯,y¯,z¯)(\bar{x},\bar{y},\bar{z})-space, emphasizing those objects that are inside by using dotted lines. The S2S^{2} sphere, being the boundary of the ball, corresponds to r=ϵ¯=0r=\bar{\epsilon}=0, whereas everything inside of the ball corresponds to r=0,ϵ¯>0r=0,\bar{\epsilon}>0. Outside of the ball, we represent r>0,ϵ¯=0r>0,\,\bar{\epsilon}=0, highlighting, in particular, the critical manifolds CaC_{a} and CrC_{r} and their reduced flow. Through the blowup we gain hyperbolicity of CaC_{a} and CrC_{r} for r=0r=0 (indicated by triple-headed arrows) along the lines (in blue and red, respectively) of partially hyperbolic equilibria. By center manifold theory, these lines produce two three-dimensional manifolds, MaM_{a} and MrM_{r} (not shown), having submanifolds within r=0r=0, denoted by WcuW^{cu} and WcsW^{cs}. These local two-dimensional manifolds are shown in lighter blue and red, respectively, since they extend inside the sphere. Also, within the sphere r=0,ϵ¯>0r=0,\,\bar{\epsilon}>0 we illustrate the orbits υ\upsilon (orange) and γ\gamma (purple), the ‘singular canards’, connecting partially hyperbolic points within r=ϵ¯=0r=\bar{\epsilon}=0 on WcuW^{cu} and WcsW^{cs}, respectively. The transversality of WcuW^{cu} and WcsW^{cs} along υ\upsilon, and along γ\gamma for any μ\mu\notin\mathbb{N}, produce, transverse intersections of Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}}, since these objects, obtained as ϵ=const.\epsilon=\text{const}. sections of MaM_{a} and MrM_{r}, respectively, are smoothly 𝒪(r2=ϵ)\mathcal{O}(r_{2}=\sqrt{\epsilon})-close on ϵ¯>0\bar{\epsilon}>0 to WcuW^{cu} and WcsW^{cs}, respectively.

1.2. Main result

Using a Melnikov approach, it was shown in [28, Theorem 3.1] that a transcritical bifurcation of the intersection of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) occurs along γ\gamma for any odd μ=2k1\mu=2k-1, kk\in\mathbb{N}. As a result, additional (secondary) ‘canards’, connecting Sa,ϵS_{a,\sqrt{\epsilon}} with Sr,ϵS_{r,\sqrt{\epsilon}}, exist near μ=2k1\mu=2k-1, for all 0<ϵ10<\epsilon\ll 1 by regular perturbation theory. In this paper, we prove the existence of a pitchfork bifurcation for μ=2k\mu=2k. We then have the following complete result regarding the bifurcations of ‘canards’ for (1.11):

Theorem 1.4.

Consider any nn\in\mathbb{N} and let kk\in\mathbb{N} be so that

n={2k1n=odd2kn=even.\displaystyle n=\begin{cases}2k-1&n=\textnormal{odd}\\ 2k&n=\textnormal{even}\end{cases}.

Set μ=n+α\mu=n+\alpha and let

D(v,α)=0,\displaystyle D(v,\alpha)=0, (1.14)

be the bifurcation equation (to be defined formally below in (2.17) locally near (v,α)=(0,0)(v,\alpha)=(0,0)) where each solution (v,α)(v,\alpha) corresponds to an intersection of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu). In particular, D(0,α)=0D(0,\alpha)=0 for all α\alpha due to the existence of the connection γ\gamma. Then

  1. (1)

    For n=oddn=\textnormal{odd}, (1.14) is locally equivalent with the transcritical bifurcation:

    v~(α~+(1)kv~)=0.\displaystyle\tilde{v}(\tilde{\alpha}+(-1)^{k}\tilde{v})=0. (1.15)
  2. (2)

    For n=evenn=\textnormal{even}, (1.14) is locally equivalent with the pitchfork bifurcation:

    v~(α~+v~2)=0.\displaystyle\tilde{v}(\tilde{\alpha}+\tilde{v}^{2})=0.

In each case, the local conjugacy ϕ:(v,α)(v~,α~)\phi:(v,\alpha)\mapsto(\tilde{v},\tilde{\alpha}) satisfies ϕ(0,0)=(0,0)\phi(0,0)=(0,0) and

Dϕ(0,0)=diag(d1(n),d2(n))with di(n)>0 for every k.\displaystyle D\phi(0,0)=\textnormal{diag}\,(d_{1}(n),d_{2}(n))\quad\mbox{with $d_{i}(n)>0$ for every $k$.} (1.16)

Theorem 1.4 item (1) is covered by [28]. In particular, it is shown (see [28, Propositions 3.2 & 3.3]) that

sign2Dv2(0,0)\displaystyle\textnormal{sign}\,\frac{\partial^{2}D}{\partial v^{2}}(0,0) =sign(1)k,\displaystyle=\textnormal{sign}\,(-1)^{k},
2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) >0,\displaystyle>0,

which produces (1.16) by singularity theory [10]. We will therefore only prove Theorem 1.4 item (2) in the following. Notice, however, that in [28], the Melnikov function is defined for all r2=ϵr_{2}=\sqrt{\epsilon} sufficiently small, measuring the intersection of Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}} directly (rather than measuring the ϵ=0\epsilon=0 objects Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu)). Nevertheless, seeing that the bifurcations in Theorem 1.4 are for the r2=ϵ=0r_{2}=\sqrt{\epsilon}=0 system, we will in this paper just focus on ϵ=0\epsilon=0 (and will only describe the perturbation of transverse intersection points into 0<ϵ10<\epsilon\ll 1, see e.g. Remark 4.3 and Remark 4.5).

1.3. Overview

The remainder of the paper is organized as follows: In the next Section 2, we review the Melnikov theory in [28] in further details and extend this approach to time-reversible systems, see also [13]. The result is collected in Theorem 2.8. This is relevant for (1.11), since this equation is time-reversible with respect the following symmetry

σ=diag(1,1,1):If (x,y,z)(t) is a solution of (1.11) then so is σ(x,y,z)(t).\displaystyle\sigma=\textnormal{diag}\,(1,-1,-1):\quad\mbox{If $(x,y,z)(t)$ is a solution of (\ref{eq:fnnfeps20}) then so is $\sigma(x,y,z)(-t)$.} (1.17)

This reduces the proof of our main result, Theorem 1.4 item (2) on the pitchfork bifurcation, to evaluating two integrals; one of which is already covered by [28], while the other one is the ‘third order’ Melnikov integral mentioned above. We evaluate these integrals by characterizing the manifolds Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu) locally through solutions of ‘higher order variational equations’ rather than, how it is done in [18], their (implicit) formulation through integral equations. We describe our approach further in Section 2.2 and how these variational equations can be solved upon reduction to an inhomogeneous Weber equation. In this section, we also present a general class of systems, that include the folded node normal form, the Falkner-Skan equation and the Nosé equations, for which we can show, see Theorem 2.9, that our method produces closed form expressions of the appropriate Melnikov integrals. These results rely on properties of Hermite polynomials HnH_{n}, n0n\in\mathbb{N}_{0}. All information on these orthogonal polynomials that is relevant to the present manuscript is available in Appendix A.

The proof of Theorem 1.4 is presented in Section 3 below, see Lemma 3.4 proven towards the end of the section. We show that our expressions agree with the computations in [18, App. A] in Appendix B. Next, in Section 4 we will use the time-reversible setting and the blowup approach to show that the additional intersections produced by the pitchfork bifurcation do not reach the actual Fenichel slow manifolds for any 0<ϵ10<\epsilon\ll 1, see Proposition 4.4 and Remark 4.5. They do therefore not produce ‘true’ canards. This is in contrast to the case μ=2k1\mu=2k-1 where it is known that true ‘secondary’ canards are produced for every μ>2k1\mu>2k-1 and 0<ϵ10<\epsilon\ll 1. We also provide a new geometric explanation for this property in Section 4, see Proposition 4.2. Although these statements about canards are probably known to most experts in the field, we believe that we present the first rigorous proofs of these facts. In our final Section 5, we consider some other equations: the two-fold, the Falkner-Skan equation (1.3), and the Nosé equations (1.4), for which our time-reversible Melnikov approach in Section 2 is also applicable. In particular, for the Falkner-Skan and for the Nosé equations we provide a new geometric proof of the emergence of symmetric periodic orbits from infinity in these systems. For the Nosé equations, we show that for μ>1\mu>1 periodic orbits only emerge for μ\mu\in\mathbb{N}, a result that escaped [23]. We conclude the paper in Section 6.

2. A Melnikov theory for time-reversible systems

The reference [27] describes a Melnikov theory for connection problems of nonhyperbolic points at infinity. In this section, we will review this approach in the context of time-reversible systems. For simplicity, we restrict to 3\mathbb{R}^{3} and consider a general smooth ODE

x˙\displaystyle\dot{x} =f(x,α),\displaystyle=f(x,\alpha), (2.1)

for x=x(t)3x=x(t)\in\mathbb{R}^{3}, depending on a parameter α\alpha\in\mathbb{R}, and assume the following:

  • (H1)

    There exists a time-reversible symmetry

    (x,t)(σx,t),\displaystyle(x,t)\mapsto(\sigma x,-t), (2.2)

    with σ3×3\sigma\in\mathbb{R}^{3\times 3} being an involution: σ2=id\sigma^{2}=\textnormal{id}, id=diag(1,1,1)\textnormal{id}=\text{diag}(1,1,1) being the identity matrix in 3×3\mathbb{R}^{3\times 3}, such that

    f(σx,α)=σf(x,α),f(\sigma x,\alpha)=-\sigma f(x,\alpha),

    for all xx and all α\alpha.

Therefore:

If x(t)x(t) is solution of (2.1) then so is σx(t)\sigma x(-t).

As is standard, we say that an orbit xx with parametrization x(t)x(t), which is a fix-point of the symmetry: x(t)=σx(t)x(t)=\sigma x(-t) for all tt, is ‘symmetric’. On the other hand, in general two orbits x1x2x_{1}\neq x_{2}, for which x2(t)=σx1(t)x_{2}(t)=\sigma x_{1}(-t), is said to be ‘symmetrically related’.

Furthermore, we say that a solution x(t)x(t) of (2.1) has algebraic growth for tt\rightarrow\infty if there exists a ν>0\nu>0 such that supt0|x(t)|(|t|+1)ν<\sup_{t\geq 0}|x(t)|(|t|+1)^{-\nu}<\infty for ν>0\nu>0 large enough. Specifically, we define the Banach space

Cb,+(ν):={xC([0,),3)|supt0{|x(t)|(|t|+1)ν<},\displaystyle C_{b,+}(\nu):=\{x\in C([0,\infty),\mathbb{R}^{3})|\sup_{t\geq 0}\{|x(t)|(|t|+1)^{-\nu}<\infty\},

for ν>0\nu>0 fixed, see [27]. Similarly, a solution x(t)x(t) of (2.1) has algebraic growth for tt\rightarrow-\infty if supt0|x(t)|(|t|+1)ν<\sup_{t\leq 0}|x(t)|(|t|+1)^{-\nu}<\infty for ν>0\nu>0 large enough. Accordingly, we define

Cb,(ν):={xC((,0],3)|supt0{|x(t)|(|t|+1)ν<}.\displaystyle C_{b,-}(\nu):=\{x\in C((-\infty,0],\mathbb{R}^{3})|\sup_{t\leq 0}\{|x(t)|(|t|+1)^{-\nu}<\infty\}.

We will suppress ν\nu in Cb,+(ν)C_{b,+}(\nu) and Cb,(ν)C_{b,-}(\nu) whenever it is convenient to do so.

Next, we assume

  • (H2)

    For α=0\alpha=0 there exists a symmetric solution γ\gamma with parametrization γ(t)\gamma(t), tt\in\mathbb{R}, of at most algebraic growth for t±t\rightarrow\pm\infty, i.e. γ(t)Cb,+(ν)\gamma(t)\in C_{b,+}(\nu) with ν>0\nu>0 large enough. Without loss of generality we suppose that

    γ(0)=0.\displaystyle\gamma(0)=0.
  • (H3)

    There exists a three-dimensional smooth invariant manifold WcsW^{cs} in the extended system

    x˙=f(x,α),α˙=0.\displaystyle\begin{split}\dot{x}&=f(x,\alpha),\\ \dot{\alpha}&=0.\end{split} (2.3)

    Here WcsW^{cs} denotes the center-stable manifold consisting of all solution curves (x(t),α)(x(t),\alpha) of (2.1) near (and including) (x,α)=(γ(t),0)(x,\alpha)=(\gamma(t),0) (in a sense specified below) for which x(t)Cb,+(ν)x(t)\in C_{b,+}(\nu), for ν>0\nu>0 large enough.

WcsW^{cs} is foliated by two-dimensional invariant manifolds Wcs(α)W^{cs}(\alpha) of (2.1) for fixed values of α\alpha, sufficiently small.

By (H1) and (H3), there exists a center-unstable manifold

Wcu(α):=σWcs(α):={σq|qWcs(α)},\displaystyle W^{cu}(\alpha):=\sigma W^{cs}(\alpha):=\{\sigma q|q\in W^{cs}(\alpha)\}, (2.4)

consisting of all solution curves (z(t),α)(z(t),\alpha) of (2.1) near (and including) (x,α)=(γ(t),0)(x,\alpha)=(\gamma(t),0) for which x(t)Cb,(ν)x(t)\in C_{b,-}(\nu).

  • (H4)

    Let U:=span(γ˙(0))U:=\textnormal{span}(\dot{\gamma}(0)). Then for α=0\alpha=0 there exists a one-dimensional linear space VV such that

    Tγ(0)Wcs(0)Tγ(0)Wcu(0)=UV,T_{\gamma(0)}W^{cs}(0)\cap T_{\gamma(0)}W^{cu}(0)=U\oplus V,

    is a two-dimensional subspace.

(H4) implies that the manifolds Wcs(0)W^{cs}(0) and Wcu(0)W^{cu}(0) intersect tangentially along γ\gamma for α=0\alpha=0. In fact, seeing that Wcu=σWcsW^{cu}=\sigma W^{cs} we have

Lemma 2.1.

The following statements are equivalent:

  1. (1)

    (H4) holds.

  2. (2)

    Tγ(0)Wcs(0)=Tγ(0)Wcu(0)T_{\gamma(0)}W^{cs}(0)=T_{\gamma(0)}W^{cu}(0) and the intersection of Wcs(0)W^{cs}(0) and WcuW^{cu} along γ\gamma is tangential.

  3. (3)

    Tγ(0)Wcs(0)T_{\gamma(0)}W^{cs}(0) is an invariant subspace for σ\sigma: xTγ(0)Wcs(0)x\in T_{\gamma(0)}W^{cs}(0) \Longrightarrow σxTγ(0)Wcs(0)\sigma x\in T_{\gamma(0)}W^{cs}(0).

  4. (4)

    The variational equation along γ\gamma for α=0\alpha=0:

    z˙\displaystyle\dot{z} =A(t)z,\displaystyle=A(t)z, (2.5)

    where A(t)=Dxf(γ(t),0)A(t)=D_{x}f(\gamma(t),0), has two linearly independent solutions z1(t)=γ˙(t)z_{1}(t)=\dot{\gamma}(t) and z2(t)z_{2}(t) for which ziCb,+Cb,z_{i}\in C_{b,+}\cap C_{b,-}.

Proof.

(1) \Leftrightarrow (2) is trivial, seeing that Wcs(0)W^{cs}(0) and Wcu(0)W^{cu}(0) are two-dimensional manifolds. (3) \Leftrightarrow (2) follows from the following computation: Tγ(0)Wcu(0)=Tγ(0)σWcs(0)=Tγ(0)Wcs(0)T_{\gamma(0)}W^{cu}(0)=T_{\gamma(0)}\sigma W^{cs}(0)=T_{\gamma(0)}W^{cs}(0) by (H1), recall (2.4). Finally, (1) \Leftrightarrow (4) is standard, see [24, Proposition 4.4]. Indeed, variations along the two-dimensional space Tγ(0)Wcs(0)Tγ(0)Wcu(0)T_{\gamma(0)}W^{cs}(0)\cap T_{\gamma(0)}W^{cu}(0) correspond to algebraic growth as t±t\rightarrow\pm\infty. ∎

Next, following [28] let

W=Tγ(0)Wcs(0).\displaystyle W=T_{\gamma(0)}W^{cs}(0)^{\perp}. (2.6)

Then 3=UVW\mathbb{R}^{3}=U\oplus V\oplus W. Let eue_{u}, eve_{v} and ewe_{w} be unit vectors spanning UU, VV and WW, respectively, and denote the coordinates of any x3x\in\mathbb{R}^{3} with respect to this basis {eu,ev,ew}\{e_{u},e_{v},e_{w}\} by (u,v,w)(u,v,w). Fix r>0r>0 small and let BrB_{r} be the ball of radius rr centered at γ(0)\gamma(0). We then define a local section Σ\Sigma transverse to γ\gamma at γ(0)=0\gamma(0)=0 by

Σ={VW}Br.\displaystyle\Sigma=\{V\oplus W\}\cap B_{r}.

Notice that Σ\Sigma – in the (u,v,w)(u,v,w)-coordinates – is contained within the (v,w)(v,w)-plane. Next, we write x=z+γ(t)x=z+\gamma(t) following [27] such that

z˙\displaystyle\dot{z} =A(t)z+g(t,z,α),\displaystyle=A(t)z+g(t,z,\alpha), (2.7)

where g(t,z,α)=f(γ(t)+z,α)f(γ(t),0)A(t)zg(t,z,\alpha)=f(\gamma(t)+z,\alpha)-f(\gamma(t),0)-A(t)z. Also g(t,0,0)=0,Dzg(t,0,0)=0g(t,0,0)=0,D_{z}g(t,0,0)=0 and notice that (2.5) is the variational equation along γ(t)\gamma(t). Furthermore:

Lemma 2.2.
σA(t)\displaystyle\sigma A(-t) =A(t)σ,\displaystyle=-A(t)\sigma,
σg(t,z,α)\displaystyle\sigma g(-t,z,\alpha) =g(t,σz,α),\displaystyle=-g(t,\sigma z,\alpha),

for all t,z,αt,z,\alpha.

Proof.

Follows directly from the time-reversible symmetry of ff, recall (H1). ∎

Let Φ(t,s)\Phi(t,s) be the state-transition matrix of (2.5). Then by (H3) and (H4) there exists a continuous projection P:[0,)3P:[0,\infty)\rightarrow\mathbb{R}^{3} such that

RangeP(0)=UV,kerP(0)\displaystyle\textnormal{Range}\,P(0)=U\oplus V,\quad\textnormal{ker}\,P(0) =W,\displaystyle=W,

and

P(t)Φ(t,s)=Φ(t,s)P(s),\displaystyle P(t)\Phi(t,s)=\Phi(t,s)P(s),

for all t,s0t,s\geq 0. Furthermore, if Q(s)=IP(s)Q(s)=I-P(s) then

kerQ(0)=UV,RangeQ(0)=W.\displaystyle\textnormal{ker}\,Q(0)=U\oplus V,\quad\textnormal{Range}\,Q(0)=W. (2.8)

and it follows that

Φ(t,s)P(s)K(ts+1)θ,Φ(s,t)Q(t)Keη(ts),\displaystyle\begin{split}\|\Phi(t,s)P(s)\|&\leq K(t-s+1)^{\theta},\\ \|\Phi(s,t)Q(t)\|&\leq Ke^{-\eta(t-s)},\end{split} (2.9)

for some K1K\geq 1, θ,η0\theta,\eta\geq 0 and all 0st0\leq s\leq t, see e.g. [27, 26]. By assumption (H1), Lemma 2.2 and (2.9) we also have:

Lemma 2.3.

Φ\Phi is symmetric in the following sense:

σΦ(t,s)=Φ(t,s)σ.\displaystyle\sigma\Phi(t,s)=\Phi(-t,-s)\sigma.

Also, tσP(t)σ1t\mapsto\sigma P(-t)\sigma^{-1} and tσQ(t)σ1t\mapsto\sigma Q(-t)\sigma^{-1} are continuous projection operators such that

Φ(t,s)σP(s)σ1\displaystyle\|\Phi(t,s)\sigma P(-s)\sigma^{-1}\| =σΦ(t,s)P(s)σ1K(st+1)θ,\displaystyle=\|\sigma\Phi(-t,-s)P(-s)\sigma^{-1}\|\leq K(s-t+1)^{\theta},
Φ(s,t)σQ(t)σ1\displaystyle\|\Phi(s,t)\sigma Q(-t)\sigma^{-1}\| =σΦ(s,t)Q(t)σ1Keη(st).\displaystyle=\|\sigma\Phi(-s,-t)Q(-t)\sigma^{-1}\|\leq Ke^{-\eta(s-t)}.

for all ts0t\leq s\leq 0.

Proof.

Straightforward calculation. ∎

Consider the adjoint equation of (2.5):

ψ˙+A(t)Tψ=0,\displaystyle\dot{\psi}+A(t)^{T}\psi=0, (2.10)

and notice that

(ψ,t)(σTψ,t),\displaystyle(\psi,t)\mapsto(\sigma^{T}\psi,-t), (2.11)

is a time-reversible symmetry for (2.10) by (H1). Then

Lemma 2.4.

Let ψ(t)\psi_{*}(t) be a solution of (2.10). Then ψ(t)\psi_{*}(t) decays exponentially for t±t\rightarrow\pm\infty if and only if ϕ(0)W\phi_{*}(0)\in W.

Proof.

Standard, see [27]. ∎

In the following, we fix a specific ψ(t)\psi_{*}(t) by setting ψ(0)=ew\psi_{*}(0)=e_{w}. Since ΦT(s,t)=ΦT(t,s)\Phi^{T}(s,t)=\Phi^{-T}(t,s) is a state-transition matrix of (2.10), we can then write ψ(t)\psi_{*}(t) as

ψ(t)=ΦT(0,t)ew.\displaystyle\psi_{*}(t)=\Phi^{T}(0,t)e_{w}. (2.12)

We now have the following important result.

Lemma 2.5.

VV and WW are one-dimensional invariant subspaces of σ\sigma and σT\sigma^{T}, respectively. Hence; there exists σi{±1}\sigma_{i}\in\{\pm 1\}, i=v,wi=v,w, such that

σ|V=σvid,σT|W=σwid,\sigma|_{V}=\sigma_{v}\textnormal{id},\quad\sigma^{T}|_{W}=\sigma_{w}\textnormal{id},

where σi=±1\sigma_{i}=\pm 1 for i=v,wi=v,w.

Proof.

First, regarding the σT\sigma^{T}-invariance of WW: The solution ψ(t)\psi_{*}(t) of (2.11) is exponentially decaying for t±t\rightarrow\pm\infty. Clearly, the symmetrically related solution σTψ(t)\sigma^{T}\psi_{*}(-t) satisfies the same properties, and hence σTψ(0)W\sigma^{T}\psi_{*}(0)\in W by Lemma 2.4; the σT\sigma^{T}-invariance of WW therefore follows. Next, since γ(t)\gamma(t) is symmetric it follows by differentiation with respect to t=0t=0 that σ|U=id\sigma|_{U}=-\textnormal{id}. But then since UVU\oplus V is invariant with respect to σ\sigma, recall Lemma 2.1 item (3), and σ2=id\sigma^{2}=\textnormal{id}, the statement about σ|V\sigma|_{V} also follows from a straightforward calculation. ∎

In fact, in the (u,v,w)(u,v,w)-coordinates

σ=diag(1,σv,σw).\displaystyle\sigma=\text{diag}\,(-1,\sigma_{v},\sigma_{w}). (2.13)

Next, for (2.7), it can by variation of constants – following [27] – be shown that z(t)Cb,+z(t)\in C_{b,+}, with z(0)Σz(0)\in\Sigma, if and only if there exists a vVv\in V such that

z(t)=Φ(t,0)v+0tP(t)Φ(t,s)g(s,z(s),α)𝑑s+tQ(t)Φ(t,s)g(s,z(s),α)𝑑s.\displaystyle z(t)=\Phi(t,0)v+\int_{0}^{t}P(t)\Phi(t,s)g(s,z(s),\alpha)ds+\int_{\infty}^{t}Q(t)\Phi(t,s)g(s,z(s),\alpha)ds. (2.14)

This enables an analytic characterization of the (nonunique) invariant manifold Wcs(μ)W^{cs}(\mu), which is essential for the Melnikov approach, as follows. Let the mapping zT(z)z\mapsto T(z) be defined on Cb,+C_{b,+} so that T(z)(t)T(z)(t) is the right hand side of (2.14) and consider a sufficiently small neighborhood NN of (v,α)=(0,0)(v,\alpha)=(0,0). Then, upon possible modification (or cut-off) of ff (and therefore of gg in (2.7)), as in center manifold theory [3], we obtain for each (v,α)N(v,\alpha)\in N, a unique fix-point z(v,α)z_{*}(v,\alpha) of TT: T(z)=zT(z_{*})=z_{*}, see [27]. Henceforth we will assume that such a modification of ff (and therefore of gg) has been made. It is also standard, see also [27], to show that zz_{*} is smooth, with each partial derivative belonging to Cb,+(ν)C_{b,+}(\nu) for ν\nu large enough. In this way,

Wcs(α)={z(v,α)(t)|(v,α)N,t}.\displaystyle W^{cs}(\alpha)=\{z_{*}(v,\alpha)(t)|(v,\alpha)\in N,\,t\in\mathbb{R}\}.
Remark 2.6.

In our examples, including (1.11), the invariant manifolds Wcs(μ)W^{cs}(\mu) will be obtained, not as fix-points of (2.14), but as center manifolds upon appropriate Poincaré compactification. For the analysis of the implications of the bifurcations of γ\gamma, it will be important to study the reduced problem on such center manifolds. For the Falkner-Skan equation and the Nosé equations the ‘selection’ of these nonunique manifolds will be crucial to our analysis.

2.1. The Melnikov function

For α\alpha sufficiently small, we write W0cs(α)=Wcs(α)ΣW_{0}^{cs}(\alpha)=W^{cs}(\alpha)\cap\Sigma and W0cu(α)=Wcu(α)ΣW_{0}^{cu}(\alpha)=W^{cu}(\alpha)\cap\Sigma locally within the (v,w)(v,w)-plane as smooth graphs

w=hcs(v,α),\displaystyle w=h_{cs}(v,\alpha), (2.15)

and

w=hcu(v,α),\displaystyle w=h_{cu}(v,\alpha), (2.16)

respectively, over vv. Following [27], we then define the Melnikov function as

D(v,α)\displaystyle D(v,\alpha) =hcu(v,α)hcs(v,α).\displaystyle=h_{cu}(v,\alpha)-h_{cs}(v,\alpha). (2.17)

Clearly, a root of DD corresponds to an intersection of W0cs(α)W_{0}^{cs}(\alpha) and W0cu(α)W_{0}^{cu}(\alpha) and, hence, to an intersection of Wcs(α)W^{cs}(\alpha) and Wcu(α)W^{cu}(\alpha). Furthermore, the intersection is transverse if and only if the root is simple. We now prove the following.

Lemma 2.7.
D(v,α)\displaystyle D(v,\alpha) =σwhcs(σvv,α)hcs(v,α).\displaystyle=\sigma_{w}h_{cs}(\sigma_{v}v,\alpha)-h_{cs}(v,\alpha). (2.18)
Proof.

Following (2.17), we simply have to show that

hcu(v,α)=σwhcs(σvv,α),\displaystyle h_{cu}(v,\alpha)=\sigma_{w}h_{cs}(\sigma_{v}v,\alpha), (2.19)

for all vv and α\alpha. We will show this using the integral representation (2.14) as follows. Let z(v,α)Cb,+z_{*}(v,\alpha)\in C_{b,+} be the fix-point of the mapping TT, with T(z)(t)T(z)(t) being defined as the right hand side of (2.14), so that z(v,α)(t)Wcs(α)z_{*}(v,\alpha)(t)\in W^{cs}(\alpha) for all tt and z(v,α)(0)=(v,hcs(v,α))z_{*}(v,\alpha)(0)=(v,h_{cs}(v,\alpha)) within Σ\Sigma in the (v,w)(v,w)-coordinates. Then σz(v,α)()Cb\sigma z_{*}(v,\alpha)(-\cdot)\in C_{b-} so that σz(v,α)(t)Wcu\sigma z_{*}(v,\alpha)(-t)\in W^{cu} with

σz(v,α)(0)=(σvv,σwhcs(v,α)),\sigma z_{*}(v,\alpha)(0)=(\sigma_{v}v,\sigma_{w}h_{cs}(v,\alpha)),

writing the right hand side in the (v,w)(v,w)-coordinates, recall (2.13). Since Wcu=σWcsW^{cu}=\sigma W^{cs}, we conclude from (2.16)v=σvv{}_{v=\sigma_{v}v} that

hcu(σvv,α)=σwhcs(v,α).\displaystyle h_{cu}(\sigma_{v}v,\alpha)=\sigma_{w}h_{cs}(v,\alpha).

This shows (2.19), seeing that σv2=1\sigma_{v}^{2}=1. ∎

Now, we are ready to present the following, final result on the Melnikov integral, which is a translation of [27, Theorem 1] to the time-reversible setting in 3\mathbb{R}^{3}.

Theorem 2.8.

Let tz(v,α)(t)Cb,+t\mapsto z_{*}(v,\alpha)(t)\in C_{b,+} be the solution of (2.7) with initial conditions

z(v,α)(0)=(v,hcs(v,α)),\displaystyle z_{*}(v,\alpha)(0)=(v,h_{cs}(v,\alpha)), (2.20)

with respect to the (v,w)(v,w)-coordinates, on W0cs(α)W^{cs}_{0}(\alpha) within Σ{u=0}\Sigma\subset\{u=0\}. Then

D(v,α)=0ψ(s),g(s,z(v,α)(s),α)σwg(s,z(σvv,α)(s),α)𝑑s.\displaystyle D(v,\alpha)=\int_{0}^{\infty}\langle\psi_{*}(s),g(s,z_{*}(v,\alpha)(s),\alpha)-\sigma_{w}g(s,z_{*}(\sigma_{v}v,\alpha)(s),\alpha)\rangle ds. (2.21)
Proof.

The result follows from [27, Theorem 1], upon setting hcsew=h+h_{cs}e_{w}=h_{+} and σThcsew=h\sigma^{T}h_{cs}e_{w}=h_{-}. In further details, we simply set t=0t=0 in (2.14)z=z(v,α){}_{z=z_{*}(v,\alpha)}:

z(v,α)(0)=vQ(0)0Φ(t,s)g(s,z(v,α)(s),α)𝑑s.\displaystyle z_{*}(v,\alpha)(0)=v-Q(0)\int_{0}^{\infty}\Phi(t,s)g(s,z_{*}(v,\alpha)(s),\alpha)ds. (2.22)

The last term on the right hand side – by (2.8) – belongs to WW for all |v|,|α|δ|v|,|\alpha|\leq\delta; whence,

hcs(v,α)=ew,Q(0)0Φ(0,s)g(s,z(v,α)(s),α)𝑑s.\displaystyle h_{cs}(v,\alpha)=\langle e_{w},-Q(0)\int_{0}^{\infty}\Phi(0,s)g(s,z_{*}(v,\alpha)(s),\alpha)ds\rangle. (2.23)

Then upon using (2.12), we obtain the desired form

hcs(v,α)\displaystyle h_{cs}(v,\alpha) =0ψ(s),g(s,z(v,α)(s),α)𝑑s.\displaystyle=-\int_{0}^{\infty}\langle\psi_{*}(s),g(s,z_{*}(v,\alpha)(s),\alpha)\rangle ds.

The result then follows from (2.18). ∎

2.2. A recipe for computing appropriate Melnikov integrals

We can use (2.21) to describe the bifurcations of heteroclinic connections of Wcs(α)W^{cs}(\alpha) and Wcs(α)W^{cs}(\alpha), provided that we can determine the partial derivatives of DD. We now describe a set of assumptions, covering all of the cases we study below, where we can describe a specific procedure for doing this. We start with the following.

  • (H5)

    Suppose that γ=(0,0,t)\gamma=(0,0,t) for all tt\in\mathbb{R} and all α\alpha.

Upon a linear change of coordinates, we may also suppose without loss of generality that σ\sigma is diagonal, recall (2.13).

  • (H6)

    Suppose that σ=diag(1,1,1)\sigma=\text{diag}(1,-1,-1).

Notice that γ\gamma is symmetric with respect to this σ\sigma.

  • (H7)

    Suppose that f(,α)f(\cdot,\alpha) is quadratic.

We can then show that all relevant Melnikov integrals can be evaluated in closed form.

Theorem 2.9.

Suppose (H3) and (H5)-(H7).

  1. (1)

    Then ff, upon scaling x3x_{3} and tt if necessary, takes the following form

    f(x,α)=(x2a2+x1x2a12+x1x3x1b1+x12b11+x22b221+x1c1+x12c11+x22c22+x2x3c23).\displaystyle f(x,\alpha)=\begin{pmatrix}x_{2}a_{2}+x_{1}x_{2}a_{12}+x_{1}x_{3}\\ x_{1}b_{1}+x_{1}^{2}b_{11}+x_{2}^{2}b_{22}\\ 1+x_{1}c_{1}+x_{1}^{2}c_{11}+x_{2}^{2}c_{22}+x_{2}x_{3}c_{23}\end{pmatrix}. (2.24)

    with each coefficient a2,a12,b1,b11,b22,c1,c11,c22,c23a_{2},a_{12},b_{1},b_{11},b_{22},c_{1},c_{11},c_{22},c_{23} depending smoothly upon the parameter α\alpha.

  2. (2)

    Furthermore, let

    β:=(a2b1+1).\displaystyle\beta:=-(a_{2}b_{1}+1).

    Then (H4) is satisfied if and only if β0\beta\in\mathbb{N}_{0}.

  3. (3)

    Next, let α=0\alpha=0 be so that β0\beta\in\mathbb{N}_{0} and consider D(v,α)D(v,\alpha) as in (2.21). Then D(0,α)=0D(0,\alpha)=0 for all α\alpha and we have the following:

    1. (a)

      Suppose β=odd\beta=\text{odd}. Then σv=1\sigma_{v}=-1, σw=1\sigma_{w}=1 and vD(v,α)v\mapsto D(v,\alpha), see (2.21), is an even function, so that 2jDv2j(0,0)=0\frac{\partial^{2j}D}{\partial v^{2j}}(0,0)=0 for all j0j\in\mathbb{N}_{0}. Furthermore, the following partial derivatives of DD can be evaluated in closed form:

      2Dvα(0,0),3Dv3(0,0).\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0),\quad\frac{\partial^{3}D}{\partial v^{3}}(0,0).

      If these quantities are nonzero, then the bifurcation equation D(v,α)=0D(v,\alpha)=0 is locally equivalent with the pitchfork normal form.

    2. (b)

      Suppose β=even\beta=\text{even}. Then σv=1\sigma_{v}=1, σw=1\sigma_{w}=-1. In this case, the following partial derivatives can be evaluated in closed form

      2Dvα(0,0),2Dv2(0,0).\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0),\quad\frac{\partial^{2}D}{\partial v^{2}}(0,0).

      If these quantities are nonzero, then the bifurcation equation D(v,α)=0D(v,\alpha)=0 is locally equivalent with the transcritical normal form.

Proof.

Regarding (1): The general form in (2.24) is a simple consequence of f(0,0,t,α)=(0,0,1)Tf(0,0,t,\alpha)=(0,0,1)^{T} by (H5), σf(x,α)+f(σx,α)=0\sigma f(x,\alpha)+f(\sigma x,\alpha)=0 for all xx and all α\alpha by (H6). Furthermore, we conclude, using a Poincaré compactification and center manifold theory, that (H3) implies that the term x1x3a13x_{1}x_{3}a_{13} in the first component of ff satisfies a13>0a_{13}>0 and subsequently that there is no term x2x3b23x_{2}x_{3}b_{23} in the second component of ff. Seeing that a13>0a_{13}>0, we finally obtain (2.24) by scaling x3x_{3} and tt as follows: x~3=a13x3\tilde{x}_{3}=\sqrt{a_{13}}x_{3}, t~=a13t\tilde{t}=\sqrt{a_{13}}t.

Next regarding (2): The general form (2.24) produces

A(t)=(ta20b100c1tc230).\displaystyle A(t)=\begin{pmatrix}t&a_{2}&0\\ b_{1}&0&0\\ c_{1}&tc_{23}&0\end{pmatrix}. (2.25)

But then, upon differentiating the first equation for z1z_{1} in the variational equation (2.5) one more time with respect to tt, we can write the equation for z1z_{1} as a Weber equation:

Lβz1=0,\displaystyle L_{\beta}z_{1}=0, (2.26)

where the second order operator LβL_{\beta} is defined by the general expression:

Lβq:=q¨tq˙+βq,\displaystyle L_{\beta}q:=\ddot{q}-t\dot{q}+\beta q, (2.27)

for any qC2q\in C^{2}. From z1z_{1}, z2z_{2} and z3z_{3} can be determined by successive integration of

z˙2\displaystyle\dot{z}_{2} =b1z1,\displaystyle=b_{1}z_{1}, (2.28)
z˙3\displaystyle\dot{z}_{3} =c1z1+tc23z2.\displaystyle=c_{1}z_{1}+tc_{23}z_{2}.

We have.

Lemma 2.10.

Let HnH_{n}, n0n\in\mathbb{N}_{0}, denote the nnth degree Hermite polynomial. Then for all non-negative integers mm and ll, the following holds

LmHl(t/2)=(ml)Hl(t/2).\displaystyle L_{m}H_{l}(t/\sqrt{2})=(m-l)H_{l}(t/\sqrt{2}). (2.29)

In particular,

Hm(/2)kerLm.\displaystyle H_{m}(\cdot/\sqrt{2})\in\textnormal{ker}\,L_{m}. (2.30)
Proof.

Follows from (A.1) and (A.2) in Appendix A. ∎

Therefore for β0\beta\in\mathbb{N}_{0} there exists by (2.30) an algebraic solution x(t)=Hβ(t/2)x(t)=H_{\beta}(t/\sqrt{2}) of (2.26). Inserting this solution into (2.28) produces, upon using (A.1) and (A.2) in Appendix A, an algebraic solution of (2.5):

z(t)=(Hβ(t/2)b12(β+1)Hβ+1(t/2)c12(β+1)Hβ+1(t/2)+b1c232(β+1)(12(β+3)Hβ+3(t/2)+Hβ+1(t/2))).\displaystyle z(t)=\begin{pmatrix}H_{\beta}(t/\sqrt{2})\\ \frac{b_{1}}{2(\beta+1)}H_{\beta+1}(t/\sqrt{2})\\ \frac{c_{1}}{2(\beta+1)}H_{\beta+1}(t/\sqrt{2})+\frac{b_{1}c_{23}}{2(\beta+1)}\left(\frac{1}{2(\beta+3)}H_{\beta+3}(t/\sqrt{2})+H_{\beta+1}(t/\sqrt{2})\right)\end{pmatrix}. (2.31)

β0(H4)\beta\in\mathbb{N}_{0}\Rightarrow(H4) is then a consequence of Lemma 2.1, see item 4. On the other hand, if β0\beta\notin\mathbb{N}_{0} then there are no algebraic solutions of (2.26), see e.g. [1], and therefore (H4) does not hold, see again Lemma 2.1 item 4. This completes the proof of (2).

To finish the proof, we just need to verify the claims about σv\sigma_{v}, σw\sigma_{w} and the partial derivatives of DD at v=α=0v=\alpha=0 in item (3). For this we first determine ψ\psi_{*}. Suppose that β0\beta\in\mathbb{N}_{0}. Then a simple computation shows that the adjoint equation can be written as a second order equation for z1z_{1}:

z¨1\displaystyle\ddot{z}_{1} =tz1(β+2)z1.\displaystyle=-tz_{1}-(\beta+2)z_{1}.

Substituting z1=et/2z~1z_{1}=e^{-t/2}\tilde{z}_{1} gives

Lβ+1z1=0,\displaystyle L_{\beta+1}z_{1}=0,

recall (2.27), upon dropping the tilde. Using (A.1), we obtain the following expression for ψ\psi_{*}:

ψ(t)=et2/2c(Hβ+1(t/2)a22Hβ(t/2)0),\displaystyle\psi_{*}(t)=e^{-t^{2}/2}c\begin{pmatrix}H_{\beta+1}(t/\sqrt{2})\\ -\frac{a_{2}}{\sqrt{2}}H_{\beta}(t/\sqrt{2})\\ 0\end{pmatrix}, (2.32)

for some constant cc, ensuring that ψ(0)=ew\psi_{*}(0)=e_{w} has length 11. The statements regarding σv\sigma_{v} and σw\sigma_{w} are then simple consequences of (2.31) and (2.32), recall Lemma 2.5.

Regarding the partial derivatives of DD, we focus on β=odd\beta=\text{odd} and the closed form expression for 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0) in (3a). Both 2Dvα(0,0)\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) and β=even\beta=\text{even} in (3b) are similar, but simpler, and therefore left out. Notice also that the statements about the local equivalence with the pitchfork and the transcritical normal form follows from singularity theory, see e.g. [10].

Let

z=(z1,z2,z3)T:=zv(0,0),z′′=(z1′′,z2′′,z3′′)=2zv2(0,0).\displaystyle z^{\prime}=(z^{\prime}_{1},z^{\prime}_{2},z^{\prime}_{3})^{T}:=\frac{\partial z_{*}}{\partial v}(0,0),\quad z^{\prime\prime}=(z^{\prime\prime}_{1},z^{\prime\prime}_{2},z^{\prime\prime}_{3})=\frac{\partial^{2}z_{*}}{\partial v^{2}}(0,0). (2.33)

Notice, by linearization of (2.14)z=z(v,α){}_{z=z_{*}(v,\alpha)} using z(0,0)=0z_{*}(0,0)=0, it follows that zz^{\prime} is determined by an appropriate normalization of (2.31). Then by (2.32) we conclude that 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0) is a linear combination of terms of the following form

0et2/2Hβ(t/2)zm(t)zn′′(t)𝑑t,\displaystyle\int_{0}^{\infty}e^{-t^{2}/2}H_{\beta}(t/\sqrt{2})z_{m}^{\prime}(t)z_{n}^{\prime\prime}(t)dt, (2.34)

for n,m=1,2,3n,m=1,2,3. Obviously, this linear combination can be stated explicitly in terms of (2.24). However, the details are not important here. Next, suppose that z′′z^{\prime\prime} is a finite sum of Hermite polynomials:

z′′=iIviHi(t/2),\displaystyle z^{\prime\prime}=\sum_{i\in I}v_{i}H_{i}(t/\sqrt{2}), (2.35)

with I0I\subset\mathbb{N}_{0} being a finite index set and vi3v_{i}\in\mathbb{R}^{3}, for any iIi\in I. Then upon inserting (2.35) into (2.34) we obtain a linear combination of terms of the following form

0et2/2Hβ(t/2)Hi(t/2)Hj(t/2)𝑑t,\displaystyle\int_{0}^{\infty}e^{-t^{2}/2}H_{\beta}(t/\sqrt{2})H_{i}(t/\sqrt{2})H_{j}(t/\sqrt{2})dt, (2.36)

with i,j0i,j\in\mathbb{N}_{0} and β0\beta\in\mathbb{N}_{0}. Again, the details are not important. However, each term of the form (2.36) can be determined in closed form using (A.7) and the statement of the theorem therefore follows once we have shown (2.35). For this we insert z=z(v,α)z=z_{*}(v,\alpha) into (2.7) and differentiate the resulting equation twice with respect to vv. We then evaluate the resulting expression at (v,α)=(0,0)(v,\alpha)=(0,0) and again use that z(0,0)=0z_{*}(0,0)=0. This gives a linear inhomogeneous equation for z′′z^{\prime\prime} of the following form

z˙′′=A(t)z′′+(2(a12z1z2+a13z1z3)2(b11(z1)2+b22(z2)2)2(c11(z1)2+c22(z2)2+c23z2z3)),\displaystyle\dot{z}^{\prime\prime}=A(t)z^{\prime\prime}+\begin{pmatrix}2(a_{12}z_{1}^{\prime}z_{2}^{\prime}+a_{13}z_{1}^{\prime}z_{3}^{\prime})\\ 2(b_{11}(z_{1}^{\prime})^{2}+b_{22}(z_{2}^{\prime})^{2})\\ 2(c_{11}(z_{1}^{\prime})^{2}+c_{22}(z_{2}^{\prime})^{2}+c_{23}z_{2}^{\prime}z_{3}^{\prime})\end{pmatrix},

By (2.25) we can therefore write the equation for z1′′z_{1}^{\prime\prime} as a second order equation and obtain an inhomogeneous Weber equation:

Lβz1′′=2ddt(a12z1z2+a13z1z3)+2a2(b11(z1)2+b22(z2)2).\displaystyle L_{\beta}z_{1}^{\prime\prime}=2\frac{d}{dt}\left(a_{12}z_{1}^{\prime}z_{2}^{\prime}+a_{13}z_{1}^{\prime}z_{3}^{\prime}\right)+2a_{2}\left(b_{11}(z_{1}^{\prime})^{2}+b_{22}(z_{2}^{\prime})^{2}\right). (2.37)

Notice that by (2.31) the right hand side of (2.37) is a sum of products of Hermite polynomials. The product rule in (A.7) in Appendix A allows us to write this sum of products as a sum of Hermite polynomials only. A simple calculation, using (A.2), then shows that this sum only consists of Hermite polynomials of even degree. We can then solve (2.37) using Lemma 2.10. In particular, by linearity and the fact that m=β=oddm=\beta=\text{odd} and l=evenl=\text{even}, it follows from (2.29) that there exists an algebraic solution of (2.37) of the form

z1′′(t)=jJdjH2j(t/2),\displaystyle z_{1}^{\prime\prime}(t)=\sum_{j\in J}d_{j}H_{2j}(t/\sqrt{2}),

for a finite index set J0J\subset\mathbb{N}_{0} and djd_{j}\in\mathbb{R}, jJj\in J. This is z1′′z_{1}^{\prime\prime}, since (a) it has the desired algebraic growth and (b) z1′′(0)=0z_{1}^{\prime\prime}(0)=0. The latter property (b) is a consequence of H2i(0)=0H_{2i}(0)=0 for each i0i\in\mathbb{N}_{0}. Upon integrating the equations for z2′′z_{2}^{\prime\prime} and z3′′z_{3}^{\prime\prime} we obtain a solution of the form (2.35). This completes the proof of Theorem 2.9.

Remark 2.11.

The general procedure for evaluating 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0), described in the proof above, can essentially be summarized as follows:

  • Step (a). Insert z(v,α)z_{*}(v,\alpha) into (2.7) and differentiate the resulting equation twice with respect to vv. This characterizes z′′z^{\prime\prime} as a solution of a higher order variational equation.

  • Step (b). This equation can by (H5)-(H7) be reduced to an inhomogeneous Weber equation, see (2.37), with right hand side as a finite sum of products of Hermite polynomials.

  • Step (c). We can then use (A.7) in Appendix A to reduce these products to sums of Hermite polynomials and solve the resulting equation for z′′z^{\prime\prime} using Lemma 2.10.

  • Step (d). Finally, we insert the resulting expression for z′′z^{\prime\prime} in step (c) into 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0) producing a sum of integrals of the form (2.36). Finally, these integrals can be evaluated using (A.8).

In principle, this procedure can be extended to cases where ff is a polynomial of higher degree. It is still possible to reduce the equation for z′′z^{\prime\prime} to an inhomogeneous Weber equation with a right hand side consisting of a sum of Hermite polynomials. However, I have not managed to find an appropriate general setting for this, where one can show that this right hand side does not involve Hβ(t/2)H_{\beta}(t/\sqrt{2}). These terms belong to the kernel of LβL_{\beta}, recall (2.30), and we can therefore not apply Lemma 2.10, as described in step (c), to inhomogeneous terms of this form. There are similar issues related to formalising the procedure iteratively to obtain closed form expressions for higher order derivatives of zz_{*} and DD.

Remark 2.12.

In [18], in the case of the folded node normal form, a more implicit, integral representation of z′′z^{\prime\prime} is obtained by differentiating the fix-point equation z=T(z)z_{*}=T(z_{*}), with TT defined by the right hand side of (2.14). Inserting this expression into the expression for 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0) gives a double integral, which (presumable) can be evaluated in Mathematica. The reference [18] computes all values up to n=20n=20 for the folded node. ( We compare these values with our closed form expressions in Appendix B.) However, it is unclear if it is possible to evaluate such double integral directly.

3. Application of Theorem 2.8 to the folded node: Proof of Theorem 1.4

In this section, we now prove Theorem 1.4. For this, we follow the recipe in Section 2.2, summarized in Remark 2.11. But first we follow [28] and rectify γ\gamma, recall (1.12), to the x3x_{3}-axis by introducing

x1:=x+z212,x2:=y2z,x3:=2z,\displaystyle\begin{split}x_{1}&:=x+z^{2}-\frac{1}{2},\\ x_{2}&:=y-2z,\\ x_{3}&:=2z,\end{split} (3.1)

so that

γ:x(t)=(0,0,t),t,\displaystyle\gamma:\,x(t)=\left(0,0,t\right),\,t\in\mathbb{R}, (3.2)

using – for simplicity – the same symbol for the same object in the new variables. Notice that [18] rectifies γ\gamma in a slightly different way, see Appendix B. Inserting (3.1) into (1.11) produces

x˙1=12μx2+x1x3,x˙2=2x1,x˙3=2x1+1,\displaystyle\begin{split}\dot{x}_{1}&=\frac{1}{2}\mu x_{2}+x_{1}x_{3},\\ \dot{x}_{2}&=-2x_{1},\\ \dot{x}_{3}&=2x_{1}+1,\end{split} (3.3)

which we, see also [28, Eq. (2.18)], will study in the following. The system (3.3) is time-reversible with respect to the same symmetry as in (1.17). It is easy to see that (3.3) satisfies the assumptions (H5)-(H7) and Theorem 2.9 applies. In particular, we have

β=μ1.\displaystyle\beta=\mu-1.

We will therefore apply the procedure used in the proof of this result, specifically see Theorem 2.9 item (3a) and Remark 2.11, to (3.3). This will enable a proof of Theorem 1.4.

In the following, let

μ=n+α.\mu=n+\alpha.

Then by Lemma 1.3 and the analysis in [28] the assumptions (H1)-(H4) are satisfied. At this stage, we keep nn\in\mathbb{N} general. For n=oddn=\textnormal{odd} the results in the following are therefore covered by [28]; the only exception is that we exploit the symmetry to simplify some of the expressions.

Writing x=γ(t)+zx=\gamma(t)+z gives

z˙1=tz1+n2z2+g(z,α),z˙2=2z1,z˙3=2z1,\displaystyle\begin{split}\dot{z}_{1}&=tz_{1}+\frac{n}{2}z_{2}+g(z,\alpha),\\ \dot{z}_{2}&=-2z_{1},\\ \dot{z}_{3}&=2z_{1},\end{split} (3.4)

where

g(z,α)=12αz2+z1z3.\displaystyle g(z,\alpha)=\frac{1}{2}\alpha z_{2}+z_{1}z_{3}. (3.5)

In comparison with (2.7), ‘gg’ for (3.4) is really (g(z,α),0,0)(g(z,\alpha),0,0), but it is useful to allow for a slight abuse of notation and let gg here refer to the first nontrivial coordinate function only. Setting g=0g=0 (ignoring the nonlinear terms) in (3.4) produces the variational equations about γ\gamma

z˙=A(t)zwithA(t)=(tn20200200).\displaystyle\dot{z}=A(t)z\quad\mbox{with}\quad A(t)=\begin{pmatrix}t&\frac{n}{2}&0\\ -2&0&0\\ 2&0&0\end{pmatrix}. (3.6)

By differentiating the first equation for z1z_{1} in (3.6) with respect to tt , we obtain a Weber equation for z1z_{1}:

Ln1z1=0,\displaystyle L_{n-1}z_{1}=0,

recall (2.27). For nn\in\mathbb{N}, it has an algebraic solution:

z1=Hn1(t/2).\displaystyle z_{1}=H_{n-1}(t/\sqrt{2}). (3.7)

Inserting (3.7) into the remaining equations for z2z_{2} and z3z_{3}, we obtain the following state-transition matrix Φ(t,s)\Phi(t,s) of (3.6):

Φ(t,0)\displaystyle\Phi(t,0) =(1Hn1(0)Hn1(t/2)02nHn1(0)Hn(t/2)02nHn1(0)Hn(t/2)1),n=odd,\displaystyle=\begin{pmatrix}\frac{1}{H_{n-1}(0)}H_{n-1}(t/\sqrt{2})&*&0\\ -\frac{\sqrt{2}}{nH_{n-1}(0)}H_{n}(t/\sqrt{2})&*&0\\ \frac{\sqrt{2}}{nH_{n-1}(0)}H_{n}(t/\sqrt{2})&*&1\end{pmatrix},\quad n=\textnormal{odd}, (3.8)
Φ(t,0)\displaystyle\Phi(t,0) =(n2Hn(0)Hn1(t/2)01Hn(0)Hn(t/2)011Hn(0)Hn(t/2)1),n=even,\displaystyle=\begin{pmatrix}*&-\frac{n}{\sqrt{2}H_{n}(0)}H_{n-1}(t/\sqrt{2})&0\\ *&\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})&0\\ *&1-\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})&1\end{pmatrix},\quad n=\textnormal{even}, (3.9)

see also [28, Eqns. (3.14)-(3.15)]. Following the notation in [28], the asterisks denote a separate linearly independent solution that we do not specify and which will play no role in the following. Setting t=0t=0 in the expressions for Φ\Phi above, it follows that the z1z2z_{1}z_{2}-plane has the following decomposition:

VW,\displaystyle V\oplus W,

where

V\displaystyle V =spanev,{ev=(1,0,0)Tn=oddev=(0,1,0)Tn=even,\displaystyle=\textnormal{span}\,e_{v},\quad\left\{\begin{array}[]{cc}e_{v}=(1,0,0)^{T}&n=\textnormal{odd}\\ e_{v}=(0,1,0)^{T}&n=\textnormal{even}\end{array}\right., (3.12)
W\displaystyle W =spanew,{ew=(0,1,0)Tn=oddew=(1,0,0)Tn=even,\displaystyle=\textnormal{span}\,e_{w},\quad\left\{\begin{array}[]{cc}e_{w}=(0,1,0)^{T}&n=\textnormal{odd}\\ e_{w}=(1,0,0)^{T}&n=\textnormal{even}\end{array}\right., (3.15)

recall (H4) and (2.6). Also U=span(0,0,1)TU=\textnormal{span}(0,0,1)^{T} for all nn\in\mathbb{N}. Therefore by (2.18):

Proposition 3.1.

For (3.3),

σv\displaystyle\sigma_{v} ={1n=odd1n=even,\displaystyle=\begin{cases}1&n=\textnormal{odd}\\ -1&n=\textnormal{even}\end{cases},
σw\displaystyle\sigma_{w} ={1n=odd1n=even,\displaystyle=\begin{cases}-1&n=\textnormal{odd}\\ 1&n=\textnormal{even}\end{cases},

and

D(v,α)={2hcs(v,α)n=oddhcs(v,α)hcs(v,α)n=even.\displaystyle D(v,\alpha)=\left\{\begin{array}[]{cc }-2h_{cs}(v,\alpha)&n=\textnormal{odd}\\ h_{cs}(-v,\alpha)-h_{cs}(v,\alpha)&n=\textnormal{even}\end{array}\right.. (3.18)

In particular,

  1. (1)

    D(0,α)=0D(0,\alpha)=0 for all α\alpha and any nn.

  2. (2)

    For n=evenn=\textnormal{even}, vD(v,α)v\mapsto D(v,\alpha) is an odd function for every α\alpha.

Proof.

Follows from the definition of σi\sigma_{i}, i=v,wi=v,w in Lemma 2.5 and from (2.18), see also Theorem 2.9. ∎

As a corollary, we have the following.

Corollary 3.2.

Consider n=oddn=\textnormal{odd}. Then solutions of D(v,α)=0D(v,\alpha)=0 bifurcating from v=α=0v=\alpha=0 correspond to symmetric solutions of (3.4), i.e. they are fix-points of the time-reversible symmetry σ\sigma.

Proof.

Follows from (3.18)n=odd{}_{n=\textnormal{odd}} and the fact that any solution of D(v,α)D(v,\alpha) in this case lies within w=0w=0, corresponding to x2=x3=0x_{2}=x_{3}=0, being the fix-point set of the symmetry σ\sigma. ∎

In contrary, when n=evenn=\textnormal{even} bifurcating solutions come in pairs (as a pitchfork bifurcation) that are related by the symmetry.

Now, finally by Theorem 2.8 we have.

Lemma 3.3.

Let z(v,α)()z_{*}(v,\alpha)(\cdot) be the solution with z(v,α)(0)=(v,hcs(v,α))W0cs(α)Σz_{*}(v,\alpha)(0)=(v,h_{cs}(v,\alpha))\in W_{0}^{cs}(\alpha)\subset\Sigma in the (v,w)(v,w)-coordinates. Then z(v,α)Cb,+z_{*}(v,\alpha)\in C_{b,+} and

D(v,α)\displaystyle D(v,\alpha) =0et2/2×\displaystyle=\int_{0}^{\infty}e^{-t^{2}/2}\times
{22nHn1(0)Hn(t/2)g(z(v,α)(t),α)n=odd1Hn(0)Hn(t/2)(g(z(v,α)(t),α)g(z(v,α)(t),α)n=evendt.\displaystyle\left\{\begin{array}[]{cc}\frac{2\sqrt{2}}{nH_{n-1}(0)}H_{n}(t/\sqrt{2})g(z_{*}(v,\alpha)(t),\alpha)&n=\textnormal{odd}\\ \frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})(g(z_{*}(v,\alpha)(t),\alpha)-g(z_{*}(-v,\alpha)(t),\alpha)&n=\textnormal{even}\end{array}\right.dt. (3.21)
Proof.

We have in these expressions used that

ψ(t)\displaystyle\psi_{*}(t) =et2/2(2nHn1(0)Hn(t/2)1Hn1(0)Hn1(t/2)0),for n odd,\displaystyle=e^{-t^{2}/2}\begin{pmatrix}\frac{\sqrt{2}}{nH_{n-1}(0)}H_{n}(t/\sqrt{2})\\ \frac{1}{H_{n-1}(0)}H_{n-1}(t/\sqrt{2})\\ 0\end{pmatrix},\quad\text{for $n$ odd},
ψ(t)\displaystyle\psi_{*}(t) =et2/2(1Hn(0)Hn(t/2)n2Hn(0)Hn1(t/2)0),for n even,\displaystyle=e^{-t^{2}/2}\begin{pmatrix}\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})\\ \frac{n}{\sqrt{2}H_{n}(0)}H_{n-1}(t/\sqrt{2})\\ 0\end{pmatrix},\quad\text{for $n$ even},

see [28, Eq. (3.12)], is the solution of the adjoint equation (2.10) with ψ(0)=ew\psi_{*}(0)=e_{w} which decays exponentially for t±t\rightarrow\pm\infty, recall Lemma 2.4. ∎

We are now ready to prove Theorem 1.4 item (2).

Proof of Theorem 1.4 item (2).

We now write n=2kn=2k, kk\in\mathbb{N}. By Proposition 3.1, items (1) and (2), it follows that

D(0,α)=2iDv2i(0,α)=0,\displaystyle D(0,\alpha)=\frac{\partial^{2i}D}{\partial v^{2i}}(0,\alpha)=0, (3.22)

for all α\alpha and all ii\in\mathbb{N}. Next, we have the following lemma.

Lemma 3.4.

For n=2kn=2k with kk\in\mathbb{N} the following expressions hold:

2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =π(2k)!!2(2k1)!!,\displaystyle=\frac{\sqrt{\pi}(2k)!!}{\sqrt{2}(2k-1)!!}, (3.23)
3Dv3(0,0)\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0) =32π(2k+1)(2k)!!4\displaystyle={3\sqrt{2\pi}(2k+1)(2k)!!^{4}}
×j=02k1(4k12j)!(2k12j)j!(j+1)!(2k1j)!2(2kj)!2.\displaystyle\times\sum_{j=0}^{2k-1}\frac{(4k-1-2j)!}{(2k-1-2j)j!(j+1)!(2k-1-j)!^{2}(2k-j)!^{2}}. (3.24)

Let cjkc_{jk} be the elements of the sum in (3.24):

ckj:=(4k12j)!(2k12j)j!(j+1)!(2k1j)!2(2kj)!2,\displaystyle c_{kj}:=\frac{(4k-1-2j)!}{(2k-1-2j)j!(j+1)!(2k-1-j)!^{2}(2k-j)!^{2}},

for j=0,,2k1j=0,\ldots,2k-1. Then for every kk\in\mathbb{N}

{ckj>0forj=0,,k1,ckj<0forj=k,,2k1,\displaystyle\left\{\begin{array}[]{ccc}c_{kj}>0&\textnormal{for}&j=0,\ldots,k-1,\\ c_{kj}<0&\textnormal{for}&j=k,\ldots,2k-1,\end{array}\right.

and

|ck(kl)ck(k+l1)|>22l12,\displaystyle\left|\frac{c_{k(k-l)}}{c_{k(k+l-1)}}\right|>2^{2l-1}\geq 2, (3.25)

for all l=1,,kl=1,\ldots,k.

We turn to the proof of Lemma 3.4 once we have shown that Lemma 3.4 implies Theorem 1.4. For this, we first estimate the negative terms (where j=k,,2k1j=k,\ldots,2k-1) of the sum in (3.24) using (3.25) to obtain the following positive lower bound,

3Dv3(0,0)>32π(2k+1)(2k)!!4j=0k112ckj,\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0)>{3\sqrt{2\pi}(2k+1)(2k)!!^{4}}\sum_{j=0}^{k-1}\frac{1}{2}c_{kj}, (3.26)

of 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0), with the right hand side being the sum of only positive terms. Consequently, the expressions (3.22), (3.23), (3.24) – together with singularity theory [10] – proves our main result Theorem 1.4 item (2) on the pitchfork bifurcation.

Proof of Lemma 3.4.

Let z(v,α)z_{*}(v,\alpha) be as described. Recall, that it has algebraic growth as tt\rightarrow\infty, and that z(0,α)=0z_{*}(0,\alpha)=0 for all α\alpha since γ\gamma is a solution for all α\alpha. Furthermore, by differentiating (3.4) with respect to vv and setting v=α=0v=\alpha=0, we obtain the following equation

z˙=A(t)z,\displaystyle\dot{z}^{\prime}=A(t)z^{\prime},

with z=zv(0,0)z^{\prime}=\frac{\partial z_{*}}{\partial v}(0,0), recall (2.33). Here A(t)A(t) is given in (3.6) with n=2kn=2k. Consequently, by (3.9) we have

z(t)=(2kH2k(0)H2k1(t/2)1H2k(0)H2k(t/2)11H2k(0)H2k(t/2)),\displaystyle z^{\prime}(t)=\begin{pmatrix}-\frac{\sqrt{2}k}{H_{2k}(0)}H_{2k-1}(t/\sqrt{2})\\ \frac{1}{H_{2k}(0)}H_{2k}(t/\sqrt{2})\\ 1-\frac{1}{H_{2k}(0)}H_{2k}(t/\sqrt{2})\end{pmatrix}, (3.27)

see also [28]. Let z′′(t)=2zv2(0,0)z^{\prime\prime}(t)=\frac{\partial^{2}z_{*}}{\partial v^{2}}(0,0), recall (2.33), denote the second partial derivative of zz_{*}. We now follow the steps in Remark 2.11.

Step (a). By differentiating (3.4) once more with respect to vv and setting v=α=0v=\alpha=0 we obtain a ‘higher order variational equation’

z˙′′=A(t)z′′+(z1z300).\displaystyle\dot{z}^{\prime\prime}=A(t)z^{\prime\prime}+\begin{pmatrix}z_{1}^{\prime}z_{3}^{\prime}\\ 0\\ 0\end{pmatrix}. (3.28)

We have the following.

Lemma 3.5.
2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =1H2k(0)20et2/2H2k(t/2)2𝑑t\displaystyle=\frac{1}{H_{2k}(0)^{2}}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})^{2}dt (3.29)
3Dv3(0,0)\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0) =32H2k(0)0et2/2H2k+1(t/2)z3(t)z3′′(t)𝑑t,\displaystyle=\frac{3}{\sqrt{2}H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k+1}(t/\sqrt{2})z_{3}^{\prime}(t)z_{3}^{\prime\prime}(t)dt, (3.30)

where z3z_{3}^{\prime} and z3′′z_{3}^{\prime\prime} in (3.30) are defined by (2.33), respectively.

Proof.

We use (3.21) with n=2kn=2k:

D(v,α)=1H2k(0)0et2/2H2k(t/2)(g(z(v,α)(t),α)g(z(v,α)(t),α)dt,\displaystyle D(v,\alpha)=\frac{1}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})(g(z_{*}(v,\alpha)(t),\alpha)-g(z_{*}(-v,\alpha)(t),\alpha)dt,

recall (3.5). To obtain (3.29) we differentiate this expression partially with respect to vv and α\alpha. This gives

2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =1H2k(0)0et2/2H2k(t/2)z2(t)𝑑t\displaystyle=\frac{1}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}{H_{2k}(t/\sqrt{2})}z_{2}^{\prime}(t)dt
=1H2k(0)20et2/2H2k(t/2)2𝑑t,\displaystyle=\frac{1}{H_{2k}(0)^{2}}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})^{2}dt,

by (3.27) upon setting v=α=0v=\alpha=0.

For (3.30), we also perform a direct calculation to obtain

3Dv3(0,0)=6H2k(0)0et2/2H2k(t/2)(z1(t)z3′′(t)+z1′′(t)z3(t))𝑑t.\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0)=\frac{6}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})\left(z_{1}^{\prime}(t)z_{3}^{\prime\prime}(t)+z_{1}^{\prime\prime}(t)z_{3}^{\prime}(t)\right)dt.

Following (3.4),

z1(i)=12z˙3(i),\displaystyle z_{1}^{(i)}=\frac{1}{2}\dot{z}_{3}^{(i)},

for i=1,2i=1,2, and hence

3Dv3(0,0)=3H2k(0)0et2/2H2k(t/2)ddt(z3(t)z3′′(t))𝑑t.\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0)=\frac{3}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})\frac{d}{dt}\left(z_{3}^{\prime}(t)z_{3}^{\prime\prime}(t)\right)dt.

By integration by parts, using z3(0)=0z_{3}^{\prime}(0)=0 and (A.1) in Appendix A, we then obtain the result. ∎

Using that H2kH_{2k} is an even function, the formula in (A.6) in Appendix A then produces the desired expression (3.23) for 2Dvα(0,0)\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) in Lemma 3.4.

To prove the remaining expression (3.24) in Lemma 3.4 for 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0), we determine z3′′z_{3}^{\prime\prime}, which is the only remaining unknown in the expression (3.30).

Step (b). We do so by first writing (3.28) as an inhomogeneous Weber equation for z1′′z_{1}^{\prime\prime}:

L2k1z1′′=2ddt(z1z3),\displaystyle L_{2k-1}z_{1}^{\prime\prime}=2\frac{d}{dt}\left(z_{1}^{\prime}z_{3}^{\prime}\right), (3.31)

recall the definition of second order linear differential operator L2k1L_{2k-1} in (2.27).

Step (c). We then use Lemma 2.10 to solve the linear, inhomogeneous equation (3.31) for the algebraic solution z1′′z_{1}^{\prime\prime} with z1′′(0)=0z_{1}^{\prime\prime}(0)=0, once we have written the right hand side of (3.31) as a finite sum of Hermite polynomials. For this we use (A.7):

Lemma 3.6.

The following holds true for any kk\in\mathbb{N}:

z1z3=2kH2k(0)2j=02k1(2k1j)(2kj)2jj!H4k12j(t/2)2kH2k(0)H2k1(t/2).\displaystyle z_{1}^{\prime}z_{3}^{\prime}=\frac{\sqrt{2}k}{H_{2k}(0)^{2}}\sum_{j=0}^{2k-1}\begin{pmatrix}2k-1\\ j\end{pmatrix}\begin{pmatrix}2k\\ j\end{pmatrix}2^{j}j!H_{4k-1-2j}(t/\sqrt{2})-\frac{\sqrt{2}k}{H_{2k}(0)}H_{2k-1}(t/\sqrt{2}).
Proof.

Calculation. ∎

Consequently, we have

Lemma 3.7.

The following holds true for any kk\in\mathbb{N}:

z1′′(t)\displaystyle z_{1}^{\prime\prime}(t) =2ddt(2kH2k(0)2j=02k112k12j(2k1j)(2kj)2jj!H4k12j(t/2)\displaystyle=-2\frac{d}{dt}\bigg{(}\frac{\sqrt{2}k}{H_{2k}(0)^{2}}\sum_{j=0}^{2k-1}\frac{1}{2k-1-2j}\begin{pmatrix}2k-1\\ j\end{pmatrix}\begin{pmatrix}2k\\ j\end{pmatrix}2^{j}j!H_{4k-1-2j}(t/\sqrt{2})
+2kH2k(0)H2k1(t/2)).\displaystyle\quad\quad\quad+\frac{\sqrt{2}k}{H_{2k}(0)}H_{2k-1}(t/\sqrt{2})\bigg{)}. (3.32)
z3′′(t)\displaystyle z_{3}^{\prime\prime}(t) =4(2kH2k(0)2j=02k112k12j(2k1j)(2kj)2jj!H4k12j(t/2)\displaystyle=-4\bigg{(}\frac{\sqrt{2}k}{H_{2k}(0)^{2}}\sum_{j=0}^{2k-1}\frac{1}{2k-1-2j}\begin{pmatrix}2k-1\\ j\end{pmatrix}\begin{pmatrix}2k\\ j\end{pmatrix}2^{j}j!H_{4k-1-2j}(t/\sqrt{2})
+2kH2k(0)H2k1(t/2)).\displaystyle\quad\quad\quad+\frac{\sqrt{2}k}{H_{2k}(0)}H_{2k-1}(t/\sqrt{2})\bigg{)}. (3.33)
Proof.

The expression in (3.32) follows from a simple calculation using Lemma 2.10, Lemma 3.6 and (A.2). (3.33) is then obtained by integrating z˙3′′=2z1′′\dot{z}_{3}^{\prime\prime}=2z_{1}^{\prime\prime}, recall (3.28) and using z3′′(0)=0z_{3}^{\prime\prime}(0)=0. ∎

Step (d). We then have.

Lemma 3.8.

The following holds for any kk\in\mathbb{N}:

3Dv3(0,0)\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0) =6kH2k(0)4j=02k112k12j(2k1j)(2kj)2jj!\displaystyle=\frac{6k}{H_{2k}(0)^{4}}\sum_{j=0}^{2k-1}\frac{1}{2k-1-2j}\begin{pmatrix}2k-1\\ j\end{pmatrix}\begin{pmatrix}2k\\ j\end{pmatrix}2^{j}j!
×et2/2H2k+1(t/2)H2k(t/2)H4k12j(t/2)dt.\displaystyle\times\int_{-\infty}^{\infty}e^{-t^{2}/2}H_{2k+1}(t/\sqrt{2})H_{2k}(t/\sqrt{2})H_{4k-1-2j}(t/\sqrt{2})dt.
Proof.

We simply insert the expressions for z3z_{3}^{\prime} and z3′′z_{3}^{\prime\prime} in (3.27) and (3.33), respectively, into (3.30). We then use (A.6) and the fact that the integrand is an even function of tt to simplify the expression. ∎

The expression (3.24) for 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0) in Lemma 3.4 then follows from (A.8) and (A.5).

To show (3.25) we simply expand the binomial coefficients in the expression for ckjc_{kj} and obtain

ck(kl)|ck(k+l1)|=(2k+2l)(2k+2l1)(2k+42l)(2k+32l)(k+l)2(k+l1)2(k+3l)2(k+2l)2,\displaystyle\frac{c_{k(k-l)}}{|c_{k(k+l-1)}|}=\frac{(2k+2l)(2k+2l-1)\cdots(2k+4-2l)(2k+3-2l)}{(k+l)^{2}(k+l-1)^{2}\cdots(k+3-l)^{2}(k+2-l)^{2}},

where the numerator and denominator both consist of 2(2l1)2(2l-1) factors. We simplify half of these factors by dividing up

ck(kl)|ck(k+l1)|=22l1(2k+2l1)(2k+2l3)(2k+52l)(2k+32l)(k+l)(k+l1)(k+3l)(k+2l).\displaystyle\frac{c_{k(k-l)}}{|c_{k(k+l-1)}|}=2^{2l-1}\frac{(2k+2l-1)(2k+2l-3)\cdots(2k+5-2l)(2k+3-2l)}{(k+l)(k+l-1)\cdots(k+3-l)(k+2-l)}.

We can write the last fraction as a product

(21/(k+l))(23/(k+l1))(21/(k+2l)),\displaystyle\left(2-1/(k+l)\right)\left(2-3/(k+l-1)\right)\cdots\left(2-1/(k+2-l)\right),

where each factor is >1>1 for every l=1,,kl=1,\ldots,k. This shows (3.25) and we have therefore completed the proof of Lemma 3.4. ∎

4. Secondary canards: a complete picture

In Fig. 4 we present a sketch of the compactified version of (1.11) using the Poincaré compactification induced by (1.8). The diagram is therefore identical to Fig. 3, but with r=0r=0 (and therefore ϵ=0\epsilon=0). Recall also that the three-dimensional hemisphere Sϵ¯03+={(x¯,y¯,z¯,ϵ¯)S3|ϵ¯0}S^{3}_{\bar{\epsilon}\geq 0}+=\{(\bar{x},\bar{y},\bar{z},\bar{\epsilon})\in S^{3}|\bar{\epsilon}\geq 0\}, is “flattened out” by projection onto the (x¯,y¯,z¯)(\bar{x},\bar{y},\bar{z})-space, so that the sketched two-dimensional-sphere (x¯,y¯,z¯)S2(\bar{x},\bar{y},\bar{z})\in S^{2} corresponds to the “equator” ϵ¯=0\bar{\epsilon}=0 of Sϵ¯03S_{\bar{\epsilon}\geq 0}^{3}. On the other hand, everything inside is ϵ¯>0\bar{\epsilon}>0. In the following, we let σ\sigma act on Sϵ¯03S^{3}_{\bar{\epsilon}\geq 0} as follows σ:(x¯,y¯,z¯,ϵ¯)(x¯,y¯,z¯,ϵ¯)\sigma:\,(\bar{x},\bar{y},\bar{z},\bar{\epsilon})\mapsto(\bar{x},-\bar{y},-\bar{z},\bar{\epsilon}). This action is consistent with (1.17). The red and blue curves on the equator sphere ϵ¯=0\bar{\epsilon}=0 correspond to the intersection with the critical manifold: x¯=z¯2\bar{x}=\bar{z}^{2}, which away from x¯=z¯=0\bar{x}=\bar{z}=0 has gained hyperbolicity, recall Fig. 3. Applying center manifold theory to these points gives rise to the local center manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) also illustrated as shaded surfaces extending into ϵ¯>0\bar{\epsilon}>0. (Recall, that these manifolds are (a) the ones obtained by restricting the 3D3D manifolds MrM_{r} and MaM_{a} to the sphere r=0r=0 and therefore (b) the ‘extensions’ of the critical manifolds CrC_{r} and CaC_{a}, respectively, onto the blowup sphere, recall Fig. 3. ) The manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) contain the strong and weak canards (orange and purple dotted lines, respectively), being heteroclinic orbits, within this framework, connecting partially hyperbolic points σps\sigma p_{s} and σpw\sigma p_{w}, given by

(x¯,y¯,z¯,ϵ¯)=(1,2,1,0),(x¯,y¯,z¯,ϵ¯)=(1,2/μ,1,0),\displaystyle(\bar{x},\bar{y},\bar{z},\bar{\epsilon})=(-1,-2,-1,0),\quad(\bar{x},\bar{y},\bar{z},\bar{\epsilon})=(-1,-2/\mu,-1,0),

with psp_{s} and pwp_{w}, respectively, on the equator sphere with ϵ¯=0\bar{\epsilon}=0. Another simple calculation in the ‘z¯=1\bar{z}=1’ chart shows that the points qoutq_{\textnormal{out}} and qin=σqoutq_{\textnormal{in}}=\sigma q_{\textnormal{out}} are hyperbolic attracting and repelling nodes, respectively. They correspond to the intersection of the nonhyperbolic critical fiber of the folded node pp (in Fig. 1 this fiber coincides with the zz-axis) with the blowup sphere. On the other hand, by working in the chart ‘y¯=1\bar{y}=1’ it follows that the points q±:x¯=z¯=ϵ¯=0,y¯=±1q_{\pm}:\,\bar{x}=\bar{z}=\bar{\epsilon}=0,\bar{y}=\pm 1 are fully nonhyperbolic. Notice also q+=σqq_{+}=\sigma q_{-}.

In the following, we write aba\prec b to mean that a<ba<b while bab-a is ‘sufficiently small’. We define aba\succ b similarly to mean that bab\prec a. Finally, aba\sim b will mean that |ba||b-a| is ‘sufficiently small’.

Recall that for (1.11), γ\gamma in (1.12) is the ‘weak canard’ written in the ϵ¯=1\bar{\epsilon}=1 chart. This special orbit divides the center manifolds into unique and nonunique subsets. To see this, notice that the local center manifold for z¯>0\bar{z}>0 and ϵ¯0\bar{\epsilon}\sim 0 is unique around the strong canard all up to the weak canard since these points coincide with the stable set of psp_{s}, see Fig. 4 and [28, Fig. 9]. We collect this result – using the xx-variables, recall (3.1) – as follows:

Lemma 4.1.

The local center manifold Wloccs(μ)W_{loc}^{cs}(\mu) is unique on the side x20x_{2}\leq 0 as the stable set of psp_{s} but nonunique for x2>0x_{2}>0. Indeed, every point on the nonunique side of Wloccs(μ)W_{loc}^{cs}(\mu) with ϵ¯>0\bar{\epsilon}>0 is forward asymptotic to the hyperbolic and attracting node qoutq_{\textnormal{out}}.

Proof.

Regarding the unique side of Wloccs(μ)W_{loc}^{cs}(\mu), we proceed as follows. In terms of the coordinates (x1,y1,ϵ1)(x_{1},y_{1},\epsilon_{1}) specified by the chart ‘z¯=1\bar{z}=1’, the point σps\sigma p_{s} is (x1,y1,ϵ1)=(0,2μ1,0)(x_{1},y_{1},\epsilon_{1})=(0,2\mu^{-1},0) whereas σpw\sigma p_{w} is (x1,y1,ϵ1)=(0,2,0)(x_{1},y_{1},\epsilon_{1})=(0,2,0). The center manifold WloccsW^{cs}_{loc} in (1.13) is therefore only unique for y12y_{1}\leq 2 which upon coordinate transformation becomes y22z2y_{2}\leq 2z_{2}, seeing that z21z_{2}\gg 1. The result then follows from the definition of x2x_{2} in (3.1).

On the other hand, to verify the statement about x2>0x_{2}>0, we blowup each q±q_{\pm} to a sphere by setting

x¯=ρ2x¯¯,z¯=ρz¯¯,ϵ¯=ρ3ϵ¯¯,\displaystyle\bar{x}=\rho^{2}\bar{\bar{x}},\,\bar{z}=\rho\bar{\bar{z}},\,\bar{\epsilon}=\rho^{3}\bar{\bar{\epsilon}}, (4.1)

leaving y¯\bar{y} untouched, where ρ0\rho\geq 0, (x¯¯,z¯¯,ϵ¯¯)S2(\bar{\bar{x}},\bar{\bar{z}},\bar{\bar{\epsilon}})\in S^{2}. Only Sϵ¯¯02:=S2{ϵ¯¯0}S^{2}_{\bar{\bar{\epsilon}}\geq 0}:=S^{2}\cap\{\bar{\bar{\epsilon}}\geq 0\} is relevant. Notice that these weights are the same as those used for blowing up the fold in 3\mathbb{R}^{3}, see [25]. The calculations are also essentially identical to those in [25], so we skip the details and just present the resulting diagrams, see Fig. 5 and Fig. 6 for the blowup of qq_{-} and q+q_{+}, respectively. In these figures, the spheres Sϵ¯¯02S^{2}_{\bar{\bar{\epsilon}}\geq 0}, obtained from the blowup (4.1), are shown in green. The consequence of these blowups are then that each point on Wloccs(μ)W^{cs}_{loc}(\mu) with x2>0x_{2}>0, ϵ¯>0\bar{\epsilon}>0, is forward asymptotic to qoutq_{out}. Seeing that qoutq_{out} is a hyperbolic and attracting node, this means that Wloccs(μ)W^{cs}_{loc}(\mu) is nonunique on this side of γ\gamma. ∎

Using the symmetry, we obtain a similar result for WcuW^{cu}. In particular, every point on the nonunique side of Wloccu(μ)W_{loc}^{cu}(\mu) with ϵ¯>0\bar{\epsilon}>0 is backwards asymptotic to qinq_{\textnormal{in}}.

4.1. The transcritical bifurcation

Now, consider the transcritical bifurcation near any odd integer n=2k1n=2k-1. Then by Theorem 1.4 item (1) and Corollary 3.2, we have a symmetric secondary canard γsc(μ)\gamma^{sc}(\mu) for any μ2k1\mu\sim 2k-1. For μ=2k1\mu=2k-1, γsc(2k1)=γ\gamma^{sc}(2k-1)=\gamma. Furthermore

Proposition 4.2.

The following holds for any kk\in\mathbb{N}:

  1. (1)

    For any μ2k1\mu\prec 2k-1, γsc(μ)\gamma^{sc}(\mu) is backwards asymptotic to qin=σqoutq_{\textnormal{in}}=\sigma q_{\textnormal{out}} and forward asymptotic to qoutq_{\textnormal{out}}. In this case, γsc(μ)\gamma^{sc}(\mu) is nonunique.

  2. (2)

    For any μ2k1\mu\succ 2k-1, γsc(μ)\gamma^{sc}(\mu) is backwards asymptotic to σps\sigma p_{s} and forward asymptotic to psp_{s}. In this case, γsc(μ)\gamma^{sc}(\mu) is unique as a heteroclinic connection.

Proof.

Firstly, the fact that γsc(μ)\gamma^{sc}(\mu) is either (1): backwards asymptotic to qinq_{\textnormal{in}} and forward asymptotic to qout=σqinq_{\textnormal{out}}=\sigma q_{\textnormal{in}} or (2): backwards asymptotic to σps\sigma p_{s} and forward asymptotic to psp_{s}, is a consequence of γsc(μ)\gamma^{sc}(\mu) being symmetric, recall Corollary 3.2. Similarly, the uniqueness of γsc(μ)\gamma^{sc}(\mu) is a consequence of the uniqueness of the center manifolds on one side of γ\gamma only, see discussion above and Lemma 4.1. To complete the proof, suppose that μ2k1\mu\succ 2k-1. (μ2k1\mu\prec 2k-1 is similar and therefore left out.) Therefore α0\alpha\succ 0 and by working with the normal form (1.15), recall also (1.16), we realise that γscWcsWcu\gamma^{sc}\subset W^{cs}\cap W^{cu} intersects Σ\Sigma along x2=0x_{2}=0. Let (x1(μ),0,0)(x_{1}(\mu),0,0) denote the intersection point. Then by (1.15)

signx1=sign(1)k+1.\displaystyle\textnormal{sign}\,x_{1}=\textnormal{sign}\,(-1)^{k+1}. (4.2)

We will now show that the x2x_{2}-component of γsc(μ)\gamma^{sc}(\mu) is negative for all tt sufficiently large. For this purpose, consider the first column of the state-transition matrix Φ\Phi in (3.8)n=2k-1 and multiply this column by the nonzero number H2k2(0)H_{2k-2}(0). This gives the following solution

(H2k2(t/2)12k1H2k1(t/2)22k1H2k1(t/2)),\displaystyle\begin{pmatrix}H_{2k-2}(t/\sqrt{2})\\ -\frac{1}{2k-1}H_{2k-1}(t/\sqrt{2})\\ \frac{\sqrt{2}}{2k-1}H_{2k-1}(t/\sqrt{2})\end{pmatrix}, (4.3)

of the variational equations (3.4) with an initial condition

(H2k2(0),0,0)T,\displaystyle(H_{2k-2}(0),0,0)^{T}, (4.4)

along VV; recall that UVU\oplus V is Tγ(0)Wcs(μ)T_{\gamma(0)}W^{cs}(\mu) for μ=2k1\mu=2k-1. Using (A.5) we realise that the first component of (4.4) has the same sign as (4.2). Fix therefore T>0T>0 large enough so that H2k1(t)1H_{2k-1}(t)\geq 1, say, for all tTt\geq T. Such TT exists since H2k1(t)H_{2k-1}(t) is polynomial with positive coefficient of the leading order term t2k1t^{2k-1}. Then specifically, the x2x_{2}-component of (4.3) is negative for all tTt\geq T, and consequently for μ2k1\mu\succ 2k-1, by regular perturbation theory, the x2x_{2}-component of the time tTt\geq T forward flow of γscΣ\gamma^{sc}\cap\Sigma is also negative. This completes the proof since by Lemma 4.1, γsc(μ)\gamma^{sc}(\mu) then belongs to the unique side of Wloccs(μ)W_{loc}^{cs}(\mu) with x2<0x_{2}<0 being forward asymptotic to psp_{s}. See also Fig. 4. ∎

Remark 4.3.

Here we recall some basic facts about canards from [2, 24, 28]. Whereas the strong canard always persists as a true (‘maximal’) canard for any 0<ϵ10<\epsilon\ll 1, connecting the Fenichel slow manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon}, the perturbation of the weak canard for 0<ϵ10<\epsilon\ll 1 to a true (‘maximal’) canard is clearly more involved. In particular, there is no candidate weak canard on the critical manifold, but rather a funnel of trajectories tangent at pp to the weak eigenvector at the folded node. However, seeing that γsc(μ)\gamma^{sc}(\mu) on the blowup sphere is asymptotic to psp_{s} and σps\sigma p_{s} for fixed μ2k1\mu\succ 2k-1, see Proposition 4.2 (2), this secondary canard has the same asymptotic properties as υ\upsilon and it therefore also perturbs into a true (‘maximal’) canard connecting the Fenichel slow manifolds Sa,ϵS_{a,\epsilon} and Sr,ϵS_{r,\epsilon} for 0<ϵ10<\epsilon\ll 1, see also [28].

In fact, the secondary canards appearing for μ2k1\mu\succ 2k-1 do not undergo additional bifurcations for μ>2k1\mu>2k-1. Therefore if μ\mu satisfies 2k1<μ<2k+12k-1<\mu<2k+1 for some kk, then there exists kk secondary canards for all 0<ϵ10<\epsilon\ll 1, see [28, Proposition 4.1]. These canards divide the Fenichel slow manifold into bands o(1)o(1)-close (with respect to ϵ0\epsilon\rightarrow 0) to the strong canard with different rotational properties [2]. These bands provide an explanation for mixed-mode oscillations, see also [4].

4.2. The pitchfork bifurcation

Next, we consider n=2kn=2k and the pitchfork bifurcation. Then by Theorem 1.4 item (2) there exists two secondary canards γsc(μ)\gamma^{sc}(\mu) and σγsc(μ)\sigma\gamma^{sc}(\mu) for any μ2k\mu\prec 2k (or α0\alpha\prec 0). For μ=2k\mu=2k, γsc(2k)=γ\gamma^{sc}(2k)=\gamma.

Proposition 4.4.

The secondary canards γsc(μ)\gamma^{sc}(\mu) and σγsc(μ)\sigma\gamma^{sc}(\mu) for μ2k\mu\prec 2k are nonunique heteroclinic connections. One connects σps\sigma p_{s} with qoutq_{\textnormal{out}} while the other one connects qin=σqoutq_{\textnormal{in}}=\sigma q_{\textnormal{out}} with psp_{s}.

Proof.

Straightforward working from the diagrams in Fig. 4, Fig. 5 and Fig. 6. These canards are nonunique since they intersect the nonunique parts of the local center manifolds; recall Lemma 4.1 and that γsc(μ)\gamma^{sc}(\mu) is not symmetric in this case. ∎

Together Proposition 4.2 and Proposition 4.4 provide a rigorous and geometric explanation of [28, Fig. 17]. In this figure, ‘ρ=v\rho=v’ and the ‘TPB’s are points beyond which γsc(μ)\gamma^{sc}(\mu) does not reach the fixed local version of Wcs(μ)W^{cs}(\mu), see (1.13).

Remark 4.5.

As in Remark 4.3, we will now describe the implications of Proposition 4.4 for 0<ϵ10<\epsilon\ll 1. Fix μ2k\mu\prec 2k and suppose without loss of generality that γsc(μ)\gamma^{sc}(\mu) is the connection from σps\sigma p_{s} to qoutq_{\textnormal{out}}. For all 0<ϵ10<\epsilon\ll 1, seeing that Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) are transverse along γsc(μ)\gamma^{sc}(\mu), this secondary canard produces, as for the transcritical bifurcation above, a connection between the extended manifolds Sa,ϵS_{a,\sqrt{\epsilon}} and Sr,ϵS_{r,\sqrt{\epsilon}}. But since γsc(μ)\gamma^{sc}(\mu) for ϵ=0\epsilon=0 is asymptotic to qoutq_{\textnormal{out}} in forward time, the perturbed ‘canard’ never reaches the Fenichel slow manifold Sr,ϵS_{r,\epsilon}. Instead it follows, upon blowing down, the nonhyperbolic critical fiber as ϵ0\epsilon\rightarrow 0. However, since γsc(μ)\gamma^{sc}(\mu) is close to the strong canard for all tt sufficiently negative, we can flow the perturbed version backwards and conclude that it does in fact originate from the Fenichel slow manifold Sa,ϵS_{a,\epsilon}. Here it also divides the subset of Sa,ϵS_{a,\epsilon} between the secondary canard due to the bifurcation at 2k12k-1 and the rest of the funnel into regions of separate rotational properties through the folded node, see also [28, Proposition 2.5].

Refer to caption
Figure 4. The global dynamics on the sphere S3S^{3}, representing Sϵ¯03:=S3{ϵ¯0}S^{3}_{\bar{\epsilon}\geq 0}:=S^{3}\cap\{\bar{\epsilon}\geq 0\} – by projection – as a solid ‘ball’ in (x¯,y¯,z¯)(\bar{x},\bar{y},\bar{z})-space. Here the unit sphere, being the boundary of the ball, corresponds to ϵ¯=0\bar{\epsilon}=0, whereas everything inside of the ball corresponds to ϵ¯>0\bar{\epsilon}>0. Within this framework, the strong and weak canard, υ\upsilon and γ\gamma, respectively, are symmetric heteroclinic connections of points on the sphere. These orbits belong to ϵ¯>0\bar{\epsilon}>0, i.e. inside the sphere, and are therefore indicated in orange and purple, recall also Fig. 1, using dotted lines. Indicated are also the invariant manifolds Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu) (suppressing the μ\mu-dependency in the figure), which are locally center manifolds of normally hyperbolic lines of equilibria (blue and red half-circles, respectively). These lines end in nonhyperbolic points, qq_{-} and q+q_{+} in green which correspond to the intersection of the fold line FF, see Fig. 1, with the sphere obtained by blowing up the folded node pp. The manifolds Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu) intersect along γ\gamma, doing so tangentially for any μ\mu\in\mathbb{N}. This ‘bifurcation’ produces secondary canards through transcritical and pitchfork bifurcations, see Theorem 1.4.
Refer to caption
Figure 5. Illustration of the blowup of qq_{-}, using the same viewpoint as in Fig. 4. This blowup allows us to conclude that every point on the local manifold WcuW^{cu}, close to the line of equilibria (in blue) and between pwp_{w} and qq_{-}, will be backwards asymptotic to qinq_{\textnormal{in}}.
Refer to caption
Figure 6. Illustration of the blowup of q+q_{+}, using the same viewpoint as in Fig. 4, except now the positive yy-axis is coming out of the diagram. This blowup allows us to conclude that every point on the local manifold Wcs(μ)W^{cs}(\mu) close to the line of equilibria (in red) and between pwp_{w} and q+q_{+}, will be forward asymptotic to qoutq_{\textnormal{out}}.

5. Other examples of time-reversible systems

In this section, we present a few other examples: The two-fold, the Falkner-Skan equation and finally the Nosé equations, of nonhyperbolic connection problems where our time-reversible version of the Melnikov theory in [27] can be applied to study bifurcations. Whereas our analysis of the two-fold is brief – postponing all of the details to future work – we do include complete, self-contained descriptions of the bifurcations of periodic orbits in the Falkner-Skan equation and the Nosé equations.

5.1. The two-fold: Bifurcations of ‘canards’

Singularly perturbed systems in 3\mathbb{R}^{3} that limit to the piecewise smooth two-fold singularity, see [11], also possess orbits that are reminiscent of weak and strong canards in the singular limit ϵ0\epsilon\rightarrow 0. In particular, upon blowup, the two-fold pp corresponds to a ‘0/00/0’-singularity of a reduced problem on a critical manifold, having attracting and repelling parts on either side of pp. Upon desingularization, in much the same way as in (1.6), pp becomes a stable node with eigenvalues λs<λw<0\lambda_{s}<\lambda_{w}<0 for a subset of parameters. The essential geometry is shown in Fig. 7, see further description in the figure caption. Notice the geometry is essentially different from the folded node, insofar that the attracting and repelling manifolds for the two-fold only ‘meet up’ at the point pp, whereas for the folded node they align along the fold line. Nevertheless, upon further blowup, [15] showed, see also [14], by working on a ‘normal form’, that center manifold extensions of slow-like manifolds for the two-fold also intersect along a ‘strong canard’ for all 0<ϵ10<\epsilon\ll 1, as well as along a ‘weak’ one provided that μ=λs/λw\mu=\lambda_{s}/\lambda_{w}\notin\mathbb{N}. The equations in the associated scaling chart for ϵ=0\epsilon=0 take the following form

x˙\displaystyle\dot{x} =β1c(1+ϕ(y))(1ϕ(y)),\displaystyle=\beta^{-1}c(1+\phi(y))-(1-\phi(y)), (5.1)
y˙\displaystyle\dot{y} =bz(1+ϕ(y))βx(1ϕ(y)),\displaystyle=b{z}(1+\phi(y))-\beta x(1-\phi(y)),
z˙\displaystyle\dot{z} =1+ϕ(y)+b1γ~(1ϕ(y)).\displaystyle=1+\phi(y)+b^{-1}\tilde{\gamma}(1-\phi(y)).

see [14, Eq. (93)]. Here ϕ\phi is any ‘regularization function’ satisfying ϕ>1\phi^{\prime}>1 and ϕ(y)±1\phi(y)\rightarrow\pm 1 as y±y\rightarrow\pm\infty. Associated with the strong and weak eigenvectors of pp, the system (5.1) has, under certain assumptions on the parameters of the system (b,c,γ~,βb,c,\tilde{\gamma},\beta and ϕ\phi), two algebraic solutions υ\upsilon and γ\gamma, respectively. These solutions each lie within sets of the form {y=const.}\{y=\text{const}.\} and their projections onto the (x,z)(x,z)-plane coincides with the strong and weak eigenspace, see further details in [15, 14]. Moreover, υ\upsilon and γ\gamma are fixed by the time-reversible symmetry of (5.1) given by σ=diag(1,1,1)\sigma=\textnormal{diag}\,(-1,1,-1) and correspond to ‘unbounded heteroclinic connections’ upon the compactification provided by the blowup for ϵ=0\epsilon=0. Moreover, even though (5.1) does not fit our general framework in Section 2.2, the variational equations along γ\gamma can also be reduced to the Weber equation, see [14, Eq. (101)]. In particular, for each μ\mu\in\mathbb{N}, this equation has an algebraic solution resulting in a bifurcation scenario similar to folded node, where ‘secondary canards’ (may) emerge. I expect that the details are very similar to the folded node above, see also the numerical exploration in [15, Section 8]. However, I also expect it to be slightly more involved due to the many parameters of the system. (In fact, it might be more advantageous to work with a different, further simplified, normal form, e.g. one derived from the ‘normal form’ in [11] based on the associated piecewise smooth Filippov system). I have therefore decided not pursue this problem further in the present manuscript.

Refer to caption
Figure 7. Local geometry of the visible-invisible two-fold at p=(0,0,0)p=(0,0,0) (pink). As a piecewise smooth system, the two-fold is the intersection of two fold lines on either side of a discontinuity set. In this (generic) ‘normal form’ picture, the discontinuity set is y=0y=0, while the xx-axis is a visible fold line for the system defined for y>0y>0 whereas the yy-axis is an invisible fold-line for the system below y<0y<0. The fold lines divide a neighborhood of pp on the discontinuity set into four quadrants: crossing downwards Σcr\Sigma_{cr}^{-}, crossing upwards Σcr+\Sigma_{cr}^{+}, stable sliding Σsl\Sigma_{sl}^{-} and unstable sliding Σsl+\Sigma_{sl}^{+}. See [11, 5] for further background on these piecewise smooth concepts. On the other hand, as a singular perturbed system, the system has, upon blowup of y=ϵ=0y=\epsilon=0, Σsl\Sigma_{sl}^{-} as an attracting critical manifold CaC_{a} (blue) and Σsl+\Sigma_{sl}^{+} as a repelling one CrC_{r} (red). The point pp is a degenerate point (fully nonhyperbolic half-circle for the blowup system). However, the reduced problem on C=CaCrC=C_{a}\cup C_{r} has a ‘0/00/0’-type of singularity where orbits, like canards, can pass from the attracting sheet to the repelling one. In fact, as for the folded node, one can apply desingularization so that pp becomes a stable node with eigenvalues λs<λw<0\lambda_{s}<\lambda_{w}<0 for the reduced problem on CaC_{a}. The two orbits shown υ\upsilon (orange) and γ\gamma (purple) are ‘strong’ and ‘weak’ canards. Furthermore, bifurcations of γ\gamma occur whenever μ=λs/λw\mu=\lambda_{s}/\lambda_{w}\in\mathbb{N}.

5.2. The Falkner-Skan equation: Bifurcations of unbounded periodic orbits

In [22] it was shown for the Falkner-Skan equation (1.3) that periodic orbits bifurcate from each integer value of μ\mu\in\mathbb{N}. As noted in [23], the proof is long, complicated and – to a large extend – not based upon dynamical systems theory. The aim of the following section, is therefore to give a simple proof using the Melnikov approach, in particular Theorem 2.8, and the recipe in Section 2.2, which is based upon – as is more standard in dynamical systems – invariant manifolds. See also [17], for a similar approach in this context. In this reference, however, periodic orbits are constructed through an analysis of a return mapping.

First we write the equation (1.3) as a first order system

x˙\displaystyle\dot{x} =y,\displaystyle=y, (5.2)
y˙\displaystyle\dot{y} =z,\displaystyle=z,
z˙\displaystyle\dot{z} =xzμ(1y2),\displaystyle=-xz-\mu(1-y^{2}),

which possesses two special solutions:

γ:(x,y,z)\displaystyle\gamma:\,(x,y,z) =(t,1,0),\displaystyle=(-t,-1,0),
υ:(x,y,z)\displaystyle\upsilon:\,(x,y,z) =(t,1,0)\displaystyle=(t,1,0)

and a time-reversible symmetry given by

σ=diag(1,1,1).\displaystyle\sigma=\textnormal{diag}\,(-1,1,-1).

Both γ\gamma and υ\upsilon are symmetric orbits. It is easy to see, upon rectifying γ\gamma to the x3x_{3}-axis, setting (x,y,z)=(x3,1+x1,x2)(x,y,z)=(-x_{3},1+x_{1},x_{2}), that (5.2) satisfies the assumptions in Section 2.2, recall (H5)-(H7), respectively. In particular,

β=2μ1.\displaystyle\beta=2\mu-1.

Theorem 2.9 therefore applies and we can evaluate the relevant integrals at bifurcations in closed form following the recipe outlined in Remark 2.11. To describe the global dynamics relevant for the bifurcations of periodic orbits, and obtain the invariant manifolds WcsW^{cs} and WcuW^{cu}, we will compactify the system. For our purposes I find it useful to just compactify (x,z)(x,z), leaving yy untouched, by setting

x=x¯w¯,z=z¯w¯,\displaystyle\begin{split}x&=\frac{\bar{x}}{\bar{w}},\\ z&=\frac{\bar{z}}{\bar{w}},\end{split} (5.3)

for (x¯,z¯,w¯)S2(\bar{x},\bar{z},\bar{w})\in S^{2}. To describe the dynamics near the equator defined by w¯=0\bar{w}=0, we consider the directional chart ‘x¯=1\bar{x}=-1’ defined by

z1:=z¯x¯,w1:=w¯x¯.\displaystyle\begin{split}z_{1}&:=-\frac{\bar{z}}{\bar{x}},\\ w_{1}&:=-\frac{\bar{w}}{\bar{x}}.\end{split} (5.4)

The smooth change of coordinates between the ‘w¯=1\bar{w}=1’ chart, defined in (5.3), and the ‘x¯=1\bar{x}=-1 chart, given by (5.4), is determined by the following equations

w1=x1,z1=zx1,\displaystyle\begin{split}w_{1}&=-x^{-1},\\ z_{1}&=-zx^{-1},\end{split} (5.5)

for x<0x<0. Using (5.5) we obtain the following equations in the ‘x¯=1\bar{x}=-1’ chart:

y˙=z1,z˙1=z1+w12(yz1μ(1y2)),w˙1=w13y,\displaystyle\begin{split}\dot{y}&=z_{1},\\ \dot{z}_{1}&=z_{1}+w_{1}^{2}(yz_{1}-\mu(1-y^{2})),\\ \dot{w}_{1}&=w_{1}^{3}y,\end{split} (5.6)

upon multiplication of the right hand sides by w1w_{1}, to ensure that w1=0w_{1}=0 – corresponding to w¯=0\bar{w}=0 under the coordinate map – is invariant. For this system, we notice that any point (y,0,0)(y,0,0) is a partially hyperbolic equilibrium of (5.6), the linearization having λ=1\lambda=1 as a single nonzero eigenvalue. Therefore, by center manifold theory there exists a local repelling center manifold Wloccs(μ)W_{loc}^{cs}(\mu). A simple calculation, using the invariance of γ\gamma and υ\upsilon, shows that it takes the following form

Wloccs(μ):z1=(1y2)w12(μ+w1m1(y,w1)),yI,w1[0,δ],\displaystyle W_{loc}^{cs}(\mu):\,z_{1}=(1-y^{2})w_{1}^{2}\left(\mu+w_{1}m_{1}(y,w_{1})\right),\quad y\in I,\,w_{1}\in[0,\delta], (5.7)

with II a fixed sufficiently large interval and where δ>0\delta>0 is sufficiently small. Also, m1m_{1} is a smooth function, also depending on μ\mu. In terms of (x,y,z)(x,y,z), Wloccs(μ)W_{loc}^{cs}(\mu) takes the following form

Wloccs(μ):z=(1y2)x1(μx1m1(y,x1))),yI,\displaystyle W_{loc}^{cs}(\mu):\,z=-(1-y^{2})x^{-1}\left(\mu-x^{-1}m_{1}(y,-x^{-1}))\right),\quad y\in I,\,

valid for all xx sufficiently negative.

Inserting (5.7) into (5.6) gives the reduced problem on WloccsW^{cs}_{loc}:

y˙=(1y2)(μ+w1m1(y,w1)),w˙1=w1y,\displaystyle\begin{split}\dot{y}&=(1-y^{2})\left(\mu+w_{1}m_{1}(y,w_{1})\right),\\ \dot{w}_{1}&=w_{1}y,\end{split} (5.8)

upon desingularization through division of the right hand side by w12w_{1}^{2}. Notice that y˙>0\dot{y}>0 for y(1,1)y\in(-1,1) and all w10w_{1}\geq 0 sufficiently small. In particular, (y,w1)=(1,0)(y,w_{1})=(-1,0) and (y,w1)=(1,0)(y,w_{1})=(1,0) are saddles, with the orbit y(1,1)y\in(-1,1), w1=0w_{1}=0 being a heteroclinic connection under the flow of (5.8). For later reference, let LL be the invariant set defined by

z1=w1=0,y[1,1].\displaystyle z_{1}=w_{1}=0,\quad y\in[-1,1]. (5.9)

It becomes (x¯,z¯,w¯)=(1,0,0)(\bar{x},\bar{z},\bar{w})=(-1,0,0), yIy\in I on the cylinder.

By applying the symmetry, we obtain a local manifold WloccuW^{cu}_{loc} for all xx sufficiently large. Combining this information we obtain the diagram in Fig. 8, see [21, Fig. 1] for a related figure.

The global manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) intersect along γ\gamma and υ\upsilon. In particular, along γ\gamma we have the following

Lemma 5.1.

The manifold Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) intersect transversally along γ\gamma if and only if 2μ2\mu\notin\mathbb{N}.

Proof.

We use Lemma 2.1 item (4). Consider therefore the variational equations about γ\gamma:

z˙1=z2,z˙2=z3,z˙3=tz32μz2,\displaystyle\begin{split}\dot{z}_{1}&=z_{2},\\ \dot{z}_{2}&=z_{3},\\ \dot{z}_{3}&=tz_{3}-2\mu z_{2},\end{split} (5.10)

which upon eliminating z1z_{1} and z2z_{2}, can be written as a Weber equation

L2μ1z3=0,\displaystyle L_{2\mu-1}z_{3}=0, (5.11)

recall (2.27). The result then follows from Lemma 2.10, see also proof of Theorem 2.9. In particular, for n=2μn=2\mu, we obtain the following algebraic solution of (5.10):

z=(12n(n+1)Hn+1(t/2)12nHn(t/2)Hn1(t/2)).\displaystyle z=\begin{pmatrix}\frac{1}{2n(n+1)}H_{n+1}(t/\sqrt{2})\\ \frac{1}{\sqrt{2}n}H_{n}(t/\sqrt{2})\\ H_{n-1}(t/\sqrt{2})\end{pmatrix}. (5.12)

Next, fix any nn\in\mathbb{N} and define α\alpha by

μ=n/2+α.\mu=n/2+\alpha.

Then we can define a local Melnikov function D(v,α)D(v,\alpha), the roots of which correspond to intersections of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) near γ\gamma. Using Theorem 2.8, and proceeding as in the proof of Theorem 1.4, we obtain the following.

Proposition 5.2.

Let kk\in\mathbb{N} be so that

n={2k1n=odd2kn=even.\displaystyle n=\begin{cases}2k-1&n=\textnormal{odd}\\ 2k&n=\textnormal{even}\end{cases}.

Then

  1. (1)

    For n=oddn=\textnormal{odd}, D(v,α)=0D(v,\alpha)=0 is locally equivalent with a pitchfork bifurcation:

    v~(α~+v~2)=0.\displaystyle\tilde{v}(\tilde{\alpha}+\tilde{v}^{2})=0. (5.13)
  2. (2)

    For n=evenn=\textnormal{even}, D(v,α)=0D(v,\alpha)=0 is locally equivalent with the transcritical bifurcation:

    v~(α~+(1)k+1v~)=0.\displaystyle\tilde{v}(\tilde{\alpha}+(-1)^{k+1}\tilde{v})=0. (5.14)

In each case, the local conjugacy ϕ:(v,α)(v~,α~)\phi:(v,\alpha)\mapsto(\tilde{v},\tilde{\alpha}) satisfies ϕ(0,0)=(0,0)\phi(0,0)=(0,0) and

Dϕ(0,0)=diag(d1(n),d2(n))with di(n)>0 for every n.\displaystyle D\phi(0,0)=\textnormal{diag}\,(d_{1}(n),d_{2}(n))\quad\mbox{with $d_{i}(n)>0$ for every $n$.}
Proof.

See Appendix C. ∎

The bifurcations of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu), described in the previous result, produce new transverse intersection for μn/2\mu\sim n/2 and every nn\in\mathbb{N}. Notice, however, that they will not always produce periodic orbits. Instead they may simply diverge (by following the invariant green curves in Fig. 8 defined by (x¯,z¯)=0(\bar{x},\bar{z})=0). To give rise to periodic orbits, the intersections have to be symmetric (which rules out the pitchfork bifurcation, whose symmetrically related solutions diverge either as t±t\rightarrow\pm\infty along the green curves in Fig. 8) and they have to be on the ‘right’ side so that they follow LL and υ\upsilon. This is qualitatively very similar to the bifurcation of canards, recall Proposition 4.2 and Proposition 4.4. However, whereas for canards, the ‘interesting’ orbits, the true canards, appeared on unique sides of the invariant manifolds, we will see that for the Falkner-Skan equation and the Nosé equations, described below, the ‘interesting’ periodic orbits now appear on the nonunique sides of the manifolds. To obtain closed orbits we will have to fix copies of the center manifolds. We do so by using the fix-point sets of the symmetries.

In fact, we obtain a new proof of the following result on the bifurcation of periodic orbits from infinity, see [22], now based upon bifurcation theory and invariant manifolds.

Theorem 5.3.

Let Γ\Gamma be the (singular) heteroclinic cycle obtained from concatenating (a) γ\gamma, (b) the ‘segment’ L:(x¯,z¯,w¯)=(1,0,0)L:\,(\bar{x},\bar{z},\bar{w})=(-1,0,0), yIy\in I, recall (5.9) in the ‘x¯=1\bar{x}=-1’ chart, (c) υ\upsilon, and finally (d) σL\sigma L, i.e. the symmetrically related version of the segment LL defined in (b). Then symmetric periodic orbits bifurcate from Γ\Gamma for each μ\mu\in\mathbb{N}.

In further, details let μ=k\mu=k (so that n=2kn=2k). Then symmetric periodic orbits only exist (‘locally’ to Γ\Gamma) for μk\mu\succ k.

Proof.

The manifolds WcsW^{cs} and Wcu=σWcsW^{cu}=\sigma W^{cs} are (again) nonunique. We select unique copies as follows: Consider the strip II defined by (0,y,0)(0,y,0) with y1y\sim 1. Notice that σI=I\sigma I=I. We then select a unique copy of Wcs(μ)W^{cs}(\mu) on the y1y\geq-1 side of γ\gamma by flowing this strip backwards (where Wcs(μ)W^{cs}(\mu) becomes attracting). Obviously, we let Wcu(μ)W^{cu}(\mu) be the symmetrically related version of this fixed manifold.

Now, the transcritical bifurcation (5.14) produce a secondary intersection γsc(μ)\gamma^{sc}(\mu) of Wcs(μ)W^{cs}(\mu) and Wcu(μ)=σWcs(μ)W^{cu}(\mu)=\sigma W^{cs}(\mu) for all μk\mu\sim k so that γsc(k)=γ\gamma^{sc}(k)=\gamma. In particular, we first note – following (C.4) in Appendix C – that γsc(μ)\gamma^{sc}(\mu) intersects Σ\Sigma along the yy-axis. Let (0,y0(μ),0)(0,y_{0}(\mu),0) denote this intersection point where y0(μ)1y_{0}(\mu)\sim-1. Consider μk\mu\succ k so that α0\alpha\succ 0. Then by (5.14) we have

sign(y0(μ)+1)=sign(1)k.\displaystyle\text{sign}(y_{0}(\mu)+1)=\text{sign}(-1)^{k}. (5.15)

Consider now the solution (5.12)n=2k of the variational equations (5.10), repeated here for convenience

z=(14k(2k+1)H2k+1(t/2)122kH2k(t/2)H2k1(t/2)),\displaystyle z=\begin{pmatrix}\frac{1}{4k(2k+1)}H_{2k+1}(t/\sqrt{2})\\ \frac{1}{2\sqrt{2}k}H_{2k}(t/\sqrt{2})\\ H_{2k-1}(t/\sqrt{2})\end{pmatrix}, (5.16)

with initial condition

(0,122kH2k(0),0)T.\displaystyle\left(0,\frac{1}{2\sqrt{2}k}H_{2k}(0),0\right)^{T}. (5.17)

By (A.5) in Appendix A, we realise that the sign of the second component of (5.16) coincides with the sign of (5.15). But then, since the second component of (5.16) is positive for all sufficiently large tt, we conclude that γsc(μ)\gamma^{sc}(\mu) for μk\mu\succ k follows LL for tt large enough. Subsequently, by following υ\upsilon, see Fig. 8, we realise that γsc(μ)\gamma^{sc}(\mu) returns to x=0x=0. Since we have fixed the manifolds to intersect x=0x=0 in the strip II, this intersection is of the form (0,y1(μ),0)(0,y_{1}(\mu),0) with y1(μ)1y_{1}(\mu)\rightarrow 1 as μk+\mu\rightarrow k^{+}. But then upon applying the time-reversible symmetry, we obtain a closed orbit. The periodic orbit intersects the fix-point set of σ\sigma, defined by (0,y,0)(0,y,0) twice, once near y1y\sim-1 at (0,y0(μ),0)(0,y_{0}(\mu),0) and once near y1y\sim 1 at (0,y1(μ),0)(0,y_{1}(\mu),0).

In the remaining cases (μk\mu\prec k and the n=oddn=\textnormal{odd}), the ‘secondary intersection’ diverge as either t±t\rightarrow\pm\infty by following the green curves in Fig. 8 defined by (x¯,z¯)=(0,±1)(\bar{x},\bar{z})=(0,\pm 1). ∎

Remark 5.4.

Notice that the periodic orbits appearing for μk\mu\succ k will rotate (or twist) around γ\gamma, the number of rotations, as for the folded node, being determined by kk. See Fig. 9 for examples, and the figure caption for further description. As for the folded node, and the twisting of the secondary canards around the weak one, these rotations around γ\gamma can be explained by the solutions (C.14) of variational equations, see [28] for further details.

Refer to caption
Figure 8. Illustration of the compactification of (5.2). Our viewpoint is from the negative yy-axis, seeing the disc at y=1y=-1, containing γ\gamma (purple), from below. Notice that the circles at y=±1y=\pm 1 are not invariant; they are just emphasized for illustrative purposes. We find invariant manifolds Wcs(μ)W^{cs}(\mu) (red) and Wcu(μ)=σWcs(μ)W^{cu}(\mu)=\sigma W^{cs}(\mu) (blue) by application of center manifold theory to the partially hyperbolic lines LL and σL\sigma L. Together with γ\gamma and υ\upsilon (orange and dotted since it is on the disc at y=1y=1), these lines produce to a (singular) cycle. We obtain a new proof of the bifurcation of periodic orbits by using our time-reversible version of the Melnikov theory to perturb away from this cycle.
Refer to caption
Refer to caption
Figure 9. In (a): Three periodic orbits of the (compactified) Falkner-Skan equation, projected onto the (x¯,y)(\bar{x},y)-plane, for three different values of μ\mu (μ=1.1\mu=1.1 in blue, μ=2.3\mu=2.3 in red, μ=3.4\mu=3.4 in green). These orbits are determined by appropriate backward integration from the set II, described in the proof of Theorem 5.3. In (b): The periodic orbits in (a) are now projected onto the (y,z¯)(y,\bar{z})-plane with a zoom near γ\gamma, appearing as a point (1,0)(-1,0) in this projection. Notice that the periodic orbits twist (one twist defined as one a 360360^{\circ} complete rotation) around γ\gamma. For μ=1.1\mu=1.1 (blue) there is 1/21/2 a twist, for μ=2.3\mu=2.3 (red) there are 3/23/2 twists, and finally for μ=3.4\mu=3.4 (green) there are 5/25/2 twists.

5.3. The Nosé equation: Bifurcations of unbounded periodic orbits

The system (1.4) has two time-reversible symmetries given by

σx:=diag(1,1,1)\displaystyle\sigma^{x}:=\text{diag}\,(1,-1,-1)

and

σy:=diag(1,1,1),\displaystyle\sigma^{y}:=\text{diag}\,(-1,1,-1),

as well as one symmetry given by σz:(x,y,z)(1,1,1)\sigma^{z}:(x,y,z)\mapsto(-1,-1,1), the superscripts xx, yy and zz indicating the coordinates that are fixed by the transformation. Notice that σz=σxσy\sigma^{z}=\sigma^{x}\sigma^{y} and the group of symmetries of (1.4) is therefore the group, consisting of 44 elements, generated by σx\sigma^{x} and σz\sigma^{z}. It is isomorphic to 22\mathbb{Z}_{2}\rtimes\mathbb{Z}_{2}.

Furthermore, periodic solutions bifurcate from each integer value of μ1\mu^{-1}\in\mathbb{N}, see [23]. For μ<1\mu<1, it is known that periodic solutions only bifurcate at these values. In [22] they also show that for μ>1\mu>1, (different) periodic solutions bifurcate for every μ\mu\in\mathbb{N}, but they do not prove whether periodic solutions bifurcate from other values of μ\mu, see remark following [23, Theorem 2]. In the following, we will prove this using our time-reversible version of the Melnikov theory:

Theorem 5.5.

For μ>1\mu>1 periodic solutions only bifurcate from ‘infinity’ for μ\mu\in\mathbb{N}. In particular, periodic orbits only emerge for μn\mu\prec n for each integer nn.

We will prove this theorem in the remainder of this section. For this purpose, it will be convenient to scale (1.4) in the following way: Let

κ:=1μ1,\displaystyle\kappa:=\sqrt{1-\mu^{-1}},

and define x~\tilde{x} and y~\tilde{y} by

x\displaystyle x =κx~,\displaystyle=\kappa\tilde{x},
y\displaystyle y =κy~.\displaystyle=\kappa\tilde{y}.

Then

x˙\displaystyle\dot{x} =yxz,\displaystyle=-y-xz, (5.18)
y˙\displaystyle\dot{y} =x,\displaystyle=x,
z˙\displaystyle\dot{z} =μ+(μ1)x2,\displaystyle=-\mu+(\mu-1)x^{2},

upon dropping the tildes again. For this system, there exists three special solutions of (1.4)

υ:(x,y,z)=(0,0,t),\displaystyle\upsilon:\,(x,y,z)=(0,0,-t),

as well as

γ:\displaystyle\gamma: (x,y,z)=(1,t,t),\displaystyle\,\,(x,y,z)=(1,t,-t),
σzγ:\displaystyle\sigma^{z}\gamma: (x,y,z)=(1,t,t).\displaystyle\,\,(x,y,z)=(-1,-t,-t).

We introduce the Poincaré sphere (x¯,y¯,z¯,w¯)S3(\bar{x},\bar{y},\bar{z},\bar{w})\in S^{3} by setting

x\displaystyle x =x¯w¯,\displaystyle=\frac{\bar{x}}{\bar{w}},
y\displaystyle y =y¯w¯,\displaystyle=\frac{\bar{y}}{\bar{w}},
z\displaystyle z =z¯w¯.\displaystyle=\frac{\bar{z}}{\bar{w}}.

By working in the charts defined by ‘z¯=1\bar{z}=-1’, y¯=±1\bar{y}=\pm 1, and applying the symmetries defined by σx\sigma^{x}, σy\sigma^{y} and σz\sigma^{z} we obtain the diagram in Fig. 10. Here we adopt the same visualization technique (by projection) used above: The outer sphere corresponds to the equator sphere defined by w¯=0\bar{w}=0, whereas everything inside is w¯>0\bar{w}>0. Notice, in particular, that the invariant manifolds Wloccs(μ)W_{loc}^{cs}(\mu) and Wloccu(μ)W_{loc}^{cu}(\mu) are obtained as local center manifolds in the charts ‘z¯=1\bar{z}=\mp 1’, respectively. The reduced flow on these local manifolds, can be desingularized along w¯=0\bar{w}=0 to produce the dynamics indicated in the figure. The associated global manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) contain γ\gamma, σzγ\sigma^{z}\gamma and υ\upsilon. These solution curves are shown in purple and orange. (Notice that these lines are actually not straight lines in the projection used in Fig. 10. The figure is therefore ‘artistic’.) We leave out the simple details. In particular, using thick lines, we visualize a (singular) heteroclinic cycle Γ\Gamma. It consists of (a) γ\gamma, (b) a segment LL (through the fully nonhyperbolic point q+q_{+} at (x¯,y¯,z¯,w¯)=(0,1,0,0)(\bar{x},\bar{y},\bar{z},\bar{w})=(0,-1,0,0)) connecting ‘the end of’ γ\gamma with ‘the beginning of’ σzγ\sigma^{z}\gamma, (c) σzγ\sigma^{z}\gamma and finally σzL\sigma^{z}L. The periodic orbits of Theorem 5.5 will appear as bifurcations from this cycle through bifurcations of intersections of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu). But notice the following:

Refer to caption
Figure 10. Poincaré compactification of the Nosé equations for μ>1\mu>1. There exists three special solutions ν\nu, γ\gamma and σzγ\sigma^{z}\gamma, connecting partially hyperbolic points at infinity. In the figure, we indicate the invariant manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu), obtained as center manifolds of these partially hyperbolic points. These points include the sets LL and σzL\sigma^{z}L, shown in orange, which, together with the special solutions γ\gamma and σzγ\sigma^{z}\gamma make up a closed (singular) cycle. Our main result shows that periodic orbits bifurcate from these cycles for μ=k\mu=k for every k2k\geq 2 integer only.

Firstly, ‘new’ intersection of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) may not produce periodic orbits. Similar to the bifurcation of secondary canards, these intersections may just converge to the points qinq_{\text{in}}: (x¯,y¯,z¯,w¯)=(0,0,1,0)(\bar{x},\bar{y},\bar{z},\bar{w})=(0,0,1,0), qoutq_{\text{out}}: (x¯,y¯,z¯,w¯)=(0,0,1,0)(\bar{x},\bar{y},\bar{z},\bar{w})=(0,0,-1,0) as tt\rightarrow\mp\infty. Indeed, qinq_{\text{in}} and qoutq_{\text{out}} are hyperbolic, being a source and a sink, for the desingularized slow flow on WloccsW^{cs}_{loc} and WloccuW^{cu}_{loc}, respectively. Notice that these manifolds are unique on the corresponding side of γ\gamma as stable and unstable sets of qoutq_{\text{out}} and qinq_{\text{in}}, respectively. Consequently, to produce periodic orbits, additional intersections of Wcu(μ)W^{cu}(\mu) and Wcs(μ)W^{cs}(\mu) have to be on the ‘nonunique side’ of γ\gamma.

Secondly, there are other (singular) heteroclinic cycles. For example: (a) γ\gamma, (b) a piece of LL before (c) jumping off on a ‘fast’ connecting (shown in black near the green point q+q_{+}), (d) follow a separate piece on LL, (e) σzγ\sigma^{z}\gamma, etc. However, these cycles do not produce periodic orbits since they are not symmetric. The cycle Γ\Gamma is the only symmetric cycle.

To prove Theorem 5.5, we first realise, upon rectifying γ\gamma to the x3x_{3}-axis, setting (x,y,z)=(1+x1,x3,x3+x2)(x,y,z)=(1+x_{1},x_{3},-x_{3}+x_{2}), that the system (5.18) satisfies the assumptions of Theorem 2.9. In particular,

β=2(μ1).\displaystyle\beta=2(\mu-1).

Consequently:

Lemma 5.6.

The manifolds Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) intersect transversally along γ\gamma if and only if 2(μ1)2(\mu-1)\notin\mathbb{N}.

Proof.

Follows from Theorem 2.9, but for completeness notice the following: The variational equations about γ\gamma takes the following form:

z˙1\displaystyle\dot{z}_{1} =tz1z2z3,\displaystyle=tz_{1}-z_{2}-z_{3}, (5.19)
z˙2\displaystyle\dot{z}_{2} =z1,\displaystyle=z_{1},
z˙3\displaystyle\dot{z}_{3} =2(μ1)z1,\displaystyle=2(\mu-1)z_{1},

which upon eliminating z2z_{2} and z3z_{3}, can be written as a Weber equation:

L2(μ1)z1=0,\displaystyle L_{2(\mu-1)}z_{1}=0, (5.20)

recall (2.27). The result therefore follows from Lemma 2.10 and Lemma 2.1. Notice, in particular, that for n:=2(μ1)n:=2(\mu-1)\in\mathbb{N} we obtain, using (A.2) in Appendix A, the following algebraic solution of (5.19)

z=(Hn(t/2)12(n+1)Hn+1(t/2)n2(n+1)Hn+1(t/2)).\displaystyle z=\begin{pmatrix}H_{n}(t/\sqrt{2})\\ \frac{1}{\sqrt{2}(n+1)}H_{n+1}(t/\sqrt{2})\\ \frac{n}{\sqrt{2}(n+1)}H_{n+1}(t/\sqrt{2})\end{pmatrix}. (5.21)

Completely analogously to Section 5.2 for the Falkner-Skan equation, we fix any nn\in\mathbb{N} and define α\alpha by

μ=n2+1+α,\displaystyle\mu=\frac{n}{2}+1+\alpha, (5.22)

and let D(v,α)D(v,\alpha) denote the resulting Melnikov function, the roots of which correspond to intersections of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) near γ\gamma. Following Theorem 2.9 we can therefore evaluate the appropriate Melnikov integrals in closed form by using the recipe in Remark 2.11. In this way, we obtain the following result.

Proposition 5.7.

Let kk\in\mathbb{N} be so that

n={2k1n=odd2kn=even.\displaystyle n=\begin{cases}2k-1&n=\textnormal{odd}\\ 2k&n=\textnormal{even}\end{cases}.

Then

  1. (1)

    For n=oddn=\textnormal{odd}, D(v,α)=0D(v,\alpha)=0 is locally equivalent with a pitchfork bifurcation:

    v~(α~+v~2)=0.\displaystyle\tilde{v}(\tilde{\alpha}+\tilde{v}^{2})=0. (5.23)
  2. (2)

    For n=evenn=\textnormal{even}, D(v,α)=0D(v,\alpha)=0 is locally equivalent with the transcritical bifurcation:

    v~(α~+(1)kv~)=0.\displaystyle\tilde{v}(\tilde{\alpha}+(-1)^{k}\tilde{v})=0. (5.24)

In each case, the local conjugacy ϕ:(v,α)(v~,α~)\phi:(v,\alpha)\mapsto(\tilde{v},\tilde{\alpha}) satisfies ϕ(0,0)=(0,0)\phi(0,0)=(0,0) and

Dϕ(0,0)=diag(d1(n),d2(n))with di(n)>0 for every n.\displaystyle D\phi(0,0)=\textnormal{diag}\,(d_{1}(n),d_{2}(n))\quad\mbox{with $d_{i}(n)>0$ for every $n$.}
Proof.

See Appendix D. ∎

To complete the proof of Theorem 5.5, we first realise that when n=2k1=oddn=2k-1=\textnormal{odd} – in which case μ\mu\notin\mathbb{N} – then the roots of DD produce additional symmetrically related solutions, coexisting for μn2+1=k+12\mu\prec\frac{n}{2}+1=k+\frac{1}{2}. Working from the diagram in Fig. 10, we conclude, see also Proposition 4.4, that one of these solutions is asymptotic to qoutq_{\text{out}} whereas the other one is backwards asymptotic to qinq_{\text{in}}. Hence no periodic bifurcate from n=oddn=\textnormal{odd}. On the other hand, for n=2k=evenn=2k=\textnormal{even}, then the transcritical bifurcation produce for μk+1\mu\sim k+1 a ‘secondary’ intersection γsc(μ)\gamma^{sc}(\mu), with γsc(k)=γ\gamma^{sc}(k)=\gamma, of Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu). Since each γsc(μ)\gamma^{sc}(\mu) is symmetric, it will be asymptotic to qoutq_{\text{out}} and qinq_{\text{in}} for t±t\rightarrow\pm\infty on one side of α=0\alpha=0. On the other side, however, it will follow LL, recall Fig. 10 for tt large enough. To distinguish the two cases, we proceed as in the proofs of Proposition 4.2 and Theorem 5.3. In particular, we first note – following (D.6) in Appendix D – that γsc(μ)\gamma^{sc}(\mu) intersects Σ\Sigma along the xx-axis. Denote the intersection point by (x(μ),0,0)(x(\mu),0,0) and suppose that μk+1\mu\prec k+1. Then α0\alpha\prec 0 and by (5.24),

sign(x(μ)1)=sign(1)k.\displaystyle\text{sign}\,(x(\mu)-1)=\text{sign}\,(-1)^{k}. (5.25)

Consider the solution (5.21) of (5.19), repeated here for convenience

z=(H2k(t/2)12(2k+1)H2k+1(t/2)n2(2k+1)H2k+1(t/2)),\displaystyle z=\begin{pmatrix}H_{2k}(t/\sqrt{2})\\ \frac{1}{\sqrt{2}(2k+1)}H_{2k+1}(t/\sqrt{2})\\ \frac{n}{\sqrt{2}(2k+1)}H_{2k+1}(t/\sqrt{2})\end{pmatrix}, (5.26)

with initial condition

(H2k(0),0,0)T.\displaystyle(H_{2k}(0),0,0)^{T}. (5.27)

By (A.5) in Appendix A, we realise that the sign of the first component in (5.27) coincides with the sign of (5.25). But then, since the second component of (5.26) is positive for all tt sufficiently large, we conclude that γsc(μ)\gamma^{sc}(\mu) for μk+1\mu\prec k+1 follows LL for tt large enough. To construct periodic orbits, we fix Wcs(μ)W^{cs}(\mu) by flowing the points near LL, of the form (0,y,0)(0,y,0) for yy large enough, backwards. This fixes a copy of Wcs(μ)W^{cs}(\mu). In this way, γsc(μ)\gamma^{sc}(\mu) intersects z=0z=0 for the first time in forward time in a point (0,y(μ),0)(0,y(\mu),0), where y(μ)y(\mu)\rightarrow\infty as μk+1\mu\rightarrow k+1^{-}. Since γsc(μ)\gamma^{sc}(\mu) is symmetric with respect to the time-reversible symmetry σx\sigma^{x}, the first intersection in backwards time is at the point (0,y(μ),0)(0,-y(\mu),0). But then upon applying the symmetry σz\sigma^{z}, we obtain a closed orbit that approaches the singular heteroclinic cycle Γ\Gamma as μk+1\mu\rightarrow k+1^{-}. The periodic orbit intersects z=0z=0 four times: Once near γΣ\gamma\cap\Sigma at (x,y,z)=(x(μ),0,0)(x,y,z)=(x(\mu),0,0), once at (0,y(μ),0)(0,y(\mu),0), then near σzγΣ\sigma^{z}\gamma\cap\Sigma at (x,y,z)=(x(μ),0,0)(x,y,z)=(-x(\mu),0,0), and then finally at (0,y(μ),0)(0,-y(\mu),0).

Periodic orbits therefore only bifurcate from infinity for μ>1\mu>1 when μ\mu\in\mathbb{N}, appearing for μn\mu\prec n for each integer n2n\geq 2.

Remark 5.8.

For μ<1\mu<1, κi\kappa\in i\mathbb{R} and therefore only υ\upsilon exists. In fact, υ\upsilon bifurcates in a pitchfork-like bifurcation at μ=1\mu=1 in such a way that qoutq_{\text{out}} becomes a saddle for the reduced flow on WloccsW^{cs}_{loc}. By following Theorem 2.9, and reducing the variational equations of (1.4) along υ\upsilon to a Weber equation, it is again straightforward to show that Wcs(μ)W^{cs}(\mu) and Wcu(μ)W^{cu}(\mu) intersect transversally along υ\upsilon if and only if μ1\mu^{-1}\notin\mathbb{N}.

6. Conclusion

In this paper, we have applied a time-reversible version of the Melnikov theory for nonhyperbolic unbounded connection problems in [27] to the bifurcations of canards in the folded node normal form. In particular, we proved – for the first time – the existence of a pitchfork bifurcation for μ=even\mu=\text{even}. Our time-reversible setting also allowed for a new description of the ‘secondary canards’ emerging from the bifurcations at μ\mu\in\mathbb{N}, see Section 4. The connection to the Weber equation as well as properties of the Hermite polynomials were essential to our proof of Theorem 1.4. But the results in Section 2.2, specifically see Theorem 2.9 and Remark 2.11, highlight that the Weber equation is ‘synonymous’ with quadratic, time-reversible systems satisfying (H4) and (H2) with the unbounded symmetric orbit γ\gamma linear in tt and independent of α\alpha. This is also expected, as noted by [23], since the Weber equation is the ‘simplest’ non-autonomous equation with a non-trivial time-reversible symmetry. I believe that it is possible to obtain closed-form expressions for the Melnikov integrals in [18] related to the bifurcations of faux canards for the folded saddle singularity. Although these problems do not fit our general setting in Section 2.2, the Weber equation also appears naturally for these problems.

In Section 5, we also applied our approach to the Falkner-Skan equation and the Nosé equations. In particular, we provided a new proof of the emergence of periodic orbits, bifurcating from heteroclinic cycles at infinity, in these systems using more standard methods of dynamical systems theory. In particular, we showed that for the Nosé equations periodic orbits only bifurcate from μ\mu\in\mathbb{N}, a result that had escaped [23]. In future work, it would be natural to use the geometric framework provided by this theory to study the emergence of chaos in these two systems.

Acknowledgement

I would like to thank Martin Wechselberger for his encouragement and for providing valuable feedback on an earlier version of this manuscript.

Appendix A Properties on the Hermite polynomials

The following properties of the “physicist” Hermite polynomials:

Hn(x)=(2xddx)n1,\displaystyle H_{n}(x)=\left(2x-\frac{d}{dx}\right)^{n}\cdot 1,

is standard, see e.g. [20].

Lemma A.1.

For every nn\in\mathbb{N}

Hn+1(t)\displaystyle H_{n+1}(t) =2sHn(s)Hn(s),\displaystyle=2sH_{n}(s)-H_{n}^{\prime}(s), (A.1)
Hn(s)\displaystyle H_{n}^{\prime}(s) =2nHn1(s),\displaystyle=2nH_{n-1}(s), (A.2)
Hn(0)\displaystyle H_{n}(0) ={0n= odd(2)n/2(n1)!!n= even\displaystyle=\left\{\begin{array}[]{cc}0&\text{$n=$ odd}\\ (-2)^{n/2}(n-1)!!&\text{$n=$ even}\end{array}\right. (A.5)

and

et2/2Hn(t/2)Hm(t/2)=2π2nn!δnm,\displaystyle\int_{-\infty}^{\infty}e^{-t^{2}/2}H_{n}(t/\sqrt{2})H_{m}(t/\sqrt{2})=\sqrt{2\pi}2^{n}n!\delta_{nm}, (A.6)

where δnm\delta_{nm} is the Kronecker delta.

Furthermore, for every nn, mm\in\mathbb{N}:

Hn(s)Hm(s)=j=0min(n,m)(mj)(nj)2jj!Hn+m2j(s),\displaystyle H_{n}(s)H_{m}(s)=\sum_{j=0}^{\text{min}(n,m)}\begin{pmatrix}m\\ j\end{pmatrix}\begin{pmatrix}n\\ j\end{pmatrix}2^{j}j!H_{n+m-2j}(s), (A.7)

Finally, for every (n,m,l)3(n,m,l)\in\mathbb{N}^{3} that satisfies the triangle property and for which s=(n+m+l)/2s=(n+m+l)/2\in\mathbb{N}:

et2/2Hn(t/2)Hm(t/2)Hl(t/2)𝑑t=2π2sn!m!l!(sn)!(sm)!(sl)!.\displaystyle\int_{-\infty}^{\infty}e^{-t^{2}/2}H_{n}(t/\sqrt{2})H_{m}(t/\sqrt{2})H_{l}(t/\sqrt{2})dt=\sqrt{2\pi}2^{s}\frac{n!m!l!}{(s-n)!(s-m)!(s-l)!}. (A.8)

If (n,m,l)(n,m,l) does not satisfy the triangle inequality or if ss\notin\mathbb{N}, then the integral in (A.8) is 0.

Appendix B Comparison with [18]

In [18, App. A, p. 595] the third order Melnikov integral (3.24) is evaluated numerical for k=1,,10k=1,\ldots,10. However, we cannot compare the results directly since this reference considers μ(0,1)\mu\in(0,1); in this case υ\upsilon is the weak canard while γ\gamma is the strong one. Nevertheless, these two cases are obviously equivalent and we can map the μ(0,1)\mu\in(0,1) system into the μ>1\mu>1 system, considered in the present paper, through the following transformation:

(x,y,z,t,μ){x~=1μx,y~=μy,z~=1μz,t~=μt,μ~=μ1,\displaystyle(x,y,z,t,\mu)\mapsto\begin{cases}\tilde{x}=\frac{1}{\mu}x,\\ \tilde{y}=\sqrt{\mu}y,\\ \tilde{z}=\frac{1}{\sqrt{\mu}}z,\\ \tilde{t}=\sqrt{\mu}t,\\ \tilde{\mu}=\mu^{-1},\end{cases}

upon dropping the tildes. Specifically, this transformation maps γ\gamma and υ\upsilon for μ(0,1)\mu\in(0,1) into υ\upsilon and γ\gamma for μ1>1\mu^{-1}>1, respectively. But [18] also rectifies γ\gamma (which is υ\upsilon for μ(0,1)\mu\in(0,1)) in a slightly different way. A simple computation shows that (x~,y~,z~)(\tilde{x},\tilde{y},\tilde{z}) in [18, Eq. (15)] is related to (x1,x2,x3)(x_{1},x_{2},x_{3}) in (3.1) as follows:

x~\displaystyle\tilde{x} =1μx1,\displaystyle=\frac{1}{\mu}x_{1}, (B.1)
y~\displaystyle\tilde{y} =μ(x2+x3),\displaystyle=\sqrt{\mu}(x_{2}+x_{3}),
z~\displaystyle\tilde{z} =12μx2,\displaystyle=-\frac{1}{2\sqrt{\mu}}x_{2},

upon also replacing μ\mu by μ1\mu^{-1}. Let DMW(v,μ1)D_{MW}(v,\mu^{-1}) be the Melnikov function in [18, Proposition 29] for ρ=v\rho=v and r=0r=0. Then from (B.1) and [18, Eq. (89)] it follows that:

D(v,α)=(n+α)DMW(12n+αv,1n+α),\displaystyle D(v,\alpha)=(n+\alpha)D_{MW}\left(-\frac{1}{2\sqrt{n+\alpha}}v,\frac{1}{n+\alpha}\right),

where DD is the Melnikov function used in the present paper. Hence,

3Dv3(0,0)\displaystyle\frac{\partial^{3}D}{\partial v^{3}}(0,0) =18n3DMWv3(0,0).\displaystyle=-\frac{1}{8\sqrt{n}}\frac{\partial^{3}D_{MW}}{\partial v^{3}}(0,0). (B.2)

In Table 1, we compare each side of this equation, using our analytical expression (3.24) on the left hand side, whereas on the right hand side we use the numerical values in the table on [18, p. 595]. See further explanation in the table caption. We conclude that the results are in agreement (and attribute the tiny differences, indicated in red, to round off errors).

n=2kn=2k 3DMWv3(0,0)\frac{\partial^{3}D_{MW}}{\partial v^{3}}(0,0) 18n3DMWv3(0,0)-\frac{1}{8\sqrt{n}}\frac{\partial^{3}D_{MW}}{\partial v^{3}}(0,0) 3Dv3(0,0)\frac{\partial^{3}D}{\partial v^{3}}(0,0)
22 4.0837336724863×103-4.0837336724863...\times 10^{3} 360.9544714360.9544714... 360.9544714360.9544714...
44 9.1263550336787×105-9.1263550336787...\times 10^{5} 57039.7189557039.71895... 57039.718957039.718966...
66 1.2403985652051×108-1.2403985652051...\times 10^{8} 6.329882421×1066.329882421...\times 10^{6} 6.329882426.329882420×106...\times 10^{6}
88 1.3867566218372×1010-1.3867566218372...\times 10^{10} 6.128656321×1086.128656321...\times 10^{8} 6.128656326.128656321818×108...\times 10^{8}
1010 1.3996176586682×1012-1.3996176586682...\times 10^{12} 5.532474570×10105.532474570...\times 10^{10} 5.53247455.53247456868×1010...\times 10^{10}
1212 1.3282386742790×1014-1.3282386742790...\times 10^{14} 4.792868474×10124.792868474...\times 10^{12} 4.792868474×10124.792868474...\times 10^{12}
1414 1.2108610331032×1016-1.2108610331032...\times 10^{16} 4.045202792×10144.045202792...\times 10^{14} 4.045202792×10144.045202792...\times 10^{14}
1616 1.0738223745005×1018-1.0738223745005...\times 10^{18} 3.355694922×10163.355694922...\times 10^{16} 3.355694923.355694920×1016...\times 10^{16}
1818 9.3381989535112×1019-9.3381989535112...\times 10^{19} 2.751293251×10182.751293251...\times 10^{18} 2.751293251×10182.751293251...\times 10^{18}
2020 8.0059501510523×1021-8.0059501510523...\times 10^{21} 2.237731095×10202.237731095...\times 10^{20} 2.237731095×10202.237731095...\times 10^{20}
Table 1. Comparison of our closed-form Melnikov integral (3.24) with the values in [18]. The first and second column show all of the even values considered in [18]. The third column shows the values of the right hand side of (B.2), using the values in the first two columns, whereas the final column uses the expression in (3.24). In red we indicate the slight deviations between the last two columns. We attribute these tiny differences to round off errors.

Appendix C The Falkner-Skan equation: Proof of Proposition 5.2

Let z~\tilde{z} be defined as (x,y,z)=γ(t)+z~(x,y,z)=\gamma(t)+\tilde{z}. Then we have

z˙1\displaystyle\dot{z}_{1} =z2,\displaystyle=z_{2},
z˙2\displaystyle\dot{z}_{2} =z3,\displaystyle=z_{3},
z˙3\displaystyle\dot{z}_{3} =tz3nz2+g(z,α),\displaystyle=tz_{3}-nz_{2}+g(z,\alpha),

where

g(z,α):=2αz2z1z3+(n2+α)z22,\displaystyle g(z,\alpha):=-2\alpha z_{2}-z_{1}z_{3}+\left(\frac{n}{2}+\alpha\right)z_{2}^{2},

upon dropping the tildes. Then by (5.12), we obtain the following regarding the state transition matrix:

Φ(t,0)\displaystyle\Phi(t,0) =(112n(n+1)Hn1(0)(Hn+1(t/2)Hn+1(0))012nHn1(0)Hn(t/2)01Hn1(0)Hn1(t/2)),n=odd,\displaystyle=\begin{pmatrix}1&*&\frac{1}{2n(n+1)H_{n-1}(0)}\left(H_{n+1}(t/\sqrt{2})-H_{n+1}(0)\right)\\ 0&*&\frac{1}{\sqrt{2}nH_{n-1}(0)}H_{n}(t/\sqrt{2})\\ 0&*&\frac{1}{H_{n-1}(0)}H_{n-1}(t/\sqrt{2})\end{pmatrix},\quad n=\textnormal{odd},
Φ(t,0)\displaystyle\Phi(t,0) =(112(n+1)Hn(0)Hn+1(t/2)01Hn(0)Hn(t/2)02nHn(0)Hn1(t/2)),n=even.\displaystyle=\begin{pmatrix}1&\frac{1}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})&*\\ 0&\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})&*\\ 0&\frac{\sqrt{2}n}{H_{n}(0)}H_{n-1}(t/\sqrt{2})&*\end{pmatrix},\quad n=\textnormal{even}. (C.1)

Consequently,

V\displaystyle V =spanev,{ev=(0,0,1)Tn=oddev=(0,1,0)Tn=even,\displaystyle=\textnormal{span}\,e_{v},\quad\left\{\begin{array}[]{cc}e_{v}=(0,0,1)^{T}&n=\textnormal{odd}\\ e_{v}=(0,1,0)^{T}&n=\textnormal{even}\end{array}\right., (C.4)
W\displaystyle W =spanew,{ew=(0,1,0)Tn=oddew=(0,0,1)Tn=even,\displaystyle=\textnormal{span}\,e_{w},\quad\left\{\begin{array}[]{cc}e_{w}=(0,1,0)^{T}&n=\textnormal{odd}\\ e_{w}=(0,0,1)^{T}&n=\textnormal{even}\end{array}\right., (C.7)

recall (H4) and (2.6). Also U=span(1,0,0)TU=\textnormal{span}(1,0,0)^{T} for all nn\in\mathbb{N}. Therefore by (2.18):

σv\displaystyle\sigma_{v} ={1n=odd1n=even,\displaystyle=\begin{cases}-1&n=\textnormal{odd}\\ 1&n=\textnormal{even}\end{cases},
σw\displaystyle\sigma_{w} ={1n=odd1n=even\displaystyle=\begin{cases}1&n=\textnormal{odd}\\ -1&n=\textnormal{even}\end{cases}

and hence

D(v,α)={hcs(v,α)hcs(v,α)n=odd2hcs(v,α)n=even.\displaystyle D(v,\alpha)=\left\{\begin{array}[]{cc }h_{cs}(-v,\alpha)-h_{cs}(v,\alpha)&n=\textnormal{odd}\\ -2h_{cs}(v,\alpha)&n=\textnormal{even}\end{array}\right.. (C.10)

Consequently, for n=oddn=\textnormal{odd}, vD(v,α)v\mapsto D(v,\alpha) is an odd function for every α\alpha. On the other hand, for n=evenn=\textnormal{even} roots of D(,α)D(\cdot,\alpha) correspond to symmetric solutions, being fixed with respect to the symmetry σ\sigma. Furthermore, using

ψ(t)\displaystyle\psi_{*}(t) =(01Hn1(0)et2/2Hn1(t/2)12Hn1(0))et2/2Hn(t/2)),n=odd\displaystyle=\begin{pmatrix}0\\ \frac{1}{H_{n-1}(0)}e^{-t^{2}/2}H_{n-1}(t/\sqrt{2})\\ -\frac{1}{\sqrt{2}H_{n-1}(0))}e^{-t^{2}/2}H_{n}(t/\sqrt{2})\end{pmatrix},n=\textnormal{odd}
ψ(t)\displaystyle\psi_{*}(t) =(02Hn(0)et2/2Hn1(t/2)1Hn(0)et2/2Hn(t/2),n=even.\displaystyle=\begin{pmatrix}0\\ -\frac{\sqrt{2}}{H_{n}(0)}e^{-t^{2}/2}H_{n-1}(t/\sqrt{2})\\ \frac{1}{H_{n}(0)}e^{-t^{2}/2}H_{n}(t/\sqrt{2}\end{pmatrix},n=\textnormal{even}.

which follows from a simple calculation, we obtain

D(v,α)\displaystyle D(v,\alpha) =20et2/2×\displaystyle=2\int_{0}^{\infty}e^{-t^{2}/2}\times
{122nHn1(0)Hn(t/2)(g(z(v,α)(t),α)g(z(v,α)(t),α))n=odd1Hn(0)Hn(t/2)g(z(v,α)(t),α)n=evendt.\displaystyle\left\{\begin{array}[]{cc}\frac{1}{2\sqrt{2}nH_{n-1}(0)}H_{n}(t/\sqrt{2})\left(g(z_{*}(-v,\alpha)(t),\alpha)-g(z_{*}(v,\alpha)(t),\alpha)\right)&n=\textnormal{odd}\\ \frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})g(z_{*}(v,\alpha)(t),\alpha)&n=\textnormal{even}\end{array}\right.dt. (C.13)

We now focus on n=evenn=\textnormal{even}, which is easier, and prove the transcritical case. The details of n=oddn=\textnormal{odd} and the pitchfork are lengthier, but similar to the details of the proof of Theorem 1.4 item (2), see also Remark 2.11, and therefore left out.

Let therefore n=2kn=2k, such that μ=k+α\mu=k+\alpha, and write z:=vz(0,0)z^{\prime}:=\frac{\partial}{\partial v}z_{*}(0,0). Following (C.1), we have

z=(12(n+1)Hn(0)Hn+1(t/2)1Hn(0)Hn(t/2)2nHn(0)Hn1(t/2)),\displaystyle z^{\prime}=\begin{pmatrix}\frac{1}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})\\ \frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})\\ \frac{\sqrt{2}n}{H_{n}(0)}H_{n-1}(t/\sqrt{2})\end{pmatrix}, (C.14)

Then upon differentiating (C.13)n=even{}_{n=\textnormal{even}} with respect to α\alpha and vv we have

2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =2H2k(0)0et2/2H2k(t/2)(2z2)𝑑t\displaystyle=\frac{2}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})(-2z_{2}^{\prime})dt
=2H2k(0)2et2/2H2k(t/2)2𝑑t\displaystyle=-\frac{2}{H_{2k}(0)^{2}}\int_{-\infty}^{\infty}e^{-t^{2}/2}{H_{2k}(t/\sqrt{2})}^{2}dt
=22π(2k)!(2k1)!!2=22π(2k)!!(2k1)!!,\displaystyle=-\frac{2\sqrt{2\pi}(2k)!}{(2k-1)!!^{2}}=-\frac{2\sqrt{2\pi}(2k)!!}{(2k-1)!!},

using (C.1) as well as (A.5) and (A.6) in Appendix A. Similarly, by differentiating (C.13)n=even{}_{n=\textnormal{even}} twice with respect to vv we have

2Dv2(0,0)\displaystyle\frac{\partial^{2}D}{\partial v^{2}}(0,0) =2H2k(0)0et2/2H2k(t/2)(2z1z3+2k(z2)2)𝑑t\displaystyle=\frac{2}{H_{2k}(0)}\int_{0}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})(-2z_{1}^{\prime}z_{3}^{\prime}+2k(z_{2}^{\prime})^{2})dt
:=I1+I2,\displaystyle:=I_{1}+I_{2},

where

I1\displaystyle I_{1} =4k(2k+1)H2k(0)3et2/2H2k(t/2)H2k+1(t/2)H2k1(t/2)𝑑t,\displaystyle=-\frac{4k}{(2k+1)H_{2k}(0)^{3}}\int_{-\infty}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})H_{2k+1}(t/\sqrt{2})H_{2k-1}(t/\sqrt{2})dt,
I2\displaystyle I_{2} =2kH2k(0)3et2/2H2k(t/2)3𝑑t,\displaystyle=\frac{2k}{H_{2k}(0)^{3}}\int_{-\infty}^{\infty}e^{-t^{2}/2}H_{2k}(t/\sqrt{2})^{3}dt,

using (C.14). To compute these integrals we use (A.8) in Appendix A and obtain the following expression

I2\displaystyle I_{2} =2kH2k(0)32π23k(2k)!3k!3=(1)k2k2π(2k)!3(2k1)!!3k!3=(1)k2k2π(2k)!!3k!3,\displaystyle=\frac{2k}{H_{2k}(0)^{3}}\sqrt{2\pi}2^{3k}\frac{(2k)!^{3}}{k!^{3}}=(-1)^{k}\frac{2k\sqrt{2\pi}(2k)!^{3}}{(2k-1)!!^{3}k!^{3}}=(-1)^{k}\frac{2k\sqrt{2\pi}(2k)!!^{3}}{k!^{3}},

using (A.5), and, after some simple calculations,

I1\displaystyle I_{1} =1k+1I2.\displaystyle=-\frac{1}{k+1}I_{2}.

Consequently,

2Dv2(0,0)=kk+1I2=(1)k2k22π(2k)!!3(k+1)k!3\displaystyle\frac{\partial^{2}D}{\partial v^{2}}(0,0)=\frac{k}{k+1}I_{2}=(-1)^{k}\frac{2k^{2}\sqrt{2\pi}(2k)!!^{3}}{(k+1)k!^{3}}

By singularity theory [10] this proves the transcritical bifurcation and the local equivalence (upon replacing DD by D-D) with the normal form (5.14).

Appendix D The Nosé equations: Proof of Proposition 5.7

Let (x,y,z)=γ(t)+z~(x,y,z)=\gamma(t)+\tilde{z}. Then

z˙1\displaystyle\dot{z}_{1} =tz1z2z3+g1(z,α),\displaystyle=tz_{1}-z_{2}-z_{3}+g_{1}(z,\alpha),
z˙2\displaystyle\dot{z}_{2} =z1,\displaystyle=z_{1},
z˙3\displaystyle\dot{z}_{3} =nz1+g3(z,α),\displaystyle=nz_{1}+g_{3}(z,\alpha),

where

g1(z,α)\displaystyle g_{1}(z,\alpha) :=z1z3,\displaystyle:=-z_{1}z_{3}, (D.1)
g3(z,α)\displaystyle g_{3}(z,\alpha) :=2αz1+(12n+α)z12,\displaystyle:=2\alpha z_{1}+\left(\frac{1}{2}n+\alpha\right)z_{1}^{2},

upon dropping the tildes. By (5.21), we can compute the following relevant quantities:

Φ(t,0)\displaystyle\Phi(t,0) =(2(n+1)1+n2Hn+1(0)Hn(t/2)011+n2Hn+1(0)Hn+1(t/2)12n1+n2Hn+1(0)Hn+1(t/2)12)Vn,n=odd,\displaystyle=\begin{pmatrix}\frac{\sqrt{2}(n+1)}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n}(t/\sqrt{2})&0&*\\ \frac{1}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n+1}(t/\sqrt{2})&\frac{1}{\sqrt{2}}&*\\ \frac{n}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n+1}(t/\sqrt{2})&-\frac{1}{\sqrt{2}}&*\end{pmatrix}V_{n},\quad n=\textnormal{odd}, (D.2)
Φ(t,0)\displaystyle\Phi(t,0) =(1Hn(0)Hn(t/2)012(n+1)Hn(0)Hn+1(t/2)12n2(n+1)Hn(0)Hn+1(t/2)12)Vn,n=even,\displaystyle=\begin{pmatrix}\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})&0&*\\ \frac{1}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})&\frac{1}{\sqrt{2}}&*\\ \frac{n}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})&-\frac{1}{\sqrt{2}}&*\end{pmatrix}V_{n},\quad n=\textnormal{even}, (D.3)

for some (unspecified) constant matrix VnV_{n} (to ensure that Φ(0,0)=id\Phi(0,0)=\textnormal{id}), which will not be important in the following, and consequently

V\displaystyle V =spanev,{ev=(0,11+n2,n1+n2)Tn=oddev=(1,0,0)Tn=even,\displaystyle=\textnormal{span}\,e_{v},\quad\left\{\begin{array}[]{cc}e_{v}=\left(0,\frac{1}{\sqrt{1+n^{2}}},\frac{n}{\sqrt{1+n^{2}}}\right)^{T}&n=\textnormal{odd}\\ e_{v}=(1,0,0)^{T}&n=\textnormal{even}\end{array}\right., (D.6)
W\displaystyle W =spanew,{ew=(1,0,0)Tn=oddew=(0,1,1)Tn=even,\displaystyle=\textnormal{span}\,e_{w},\quad\left\{\begin{array}[]{cc}e_{w}=(1,0,0)^{T}&n=\textnormal{odd}\\ e_{w}=(0,1,1)^{T}&n=\textnormal{even}\end{array}\right., (D.9)

recall (H4) and (2.6). Also U=span(0,1,1)TU=\textnormal{span}(0,1,-1)^{T} for all nn\in\mathbb{N}. Therefore by (2.18):

σv\displaystyle\sigma_{v} ={1n=odd1n=even,\displaystyle=\begin{cases}-1&n=\textnormal{odd}\\ 1&n=\textnormal{even}\end{cases},
σw\displaystyle\sigma_{w} ={1n=odd1n=even\displaystyle=\begin{cases}1&n=\textnormal{odd}\\ -1&n=\textnormal{even}\end{cases}

with respect to the symmetry σx\sigma^{x}, and hence

D(v,α)={hcs(v,α)hcs(v,α)n=odd2hcs(v,α)n=even.\displaystyle D(v,\alpha)=\left\{\begin{array}[]{cc }h_{cs}(-v,\alpha)-h_{cs}(v,\alpha)&n=\textnormal{odd}\\ -2h_{cs}(v,\alpha)&n=\textnormal{even}\end{array}\right.. (D.12)

Consequently, for n=oddn=\textnormal{odd}, vD(v,α)v\mapsto D(v,\alpha) is an odd function for every α\alpha. On the other hand, for n=evenn=\textnormal{even} roots of D(,α)D(\cdot,\alpha) correspond to symmetric solutions, being fixed with respect to the symmetry σx\sigma^{x}. Furthermore, using

ψ(t)\displaystyle\psi_{*}(t) =(1Hn+1(0)et2/2Hn+1(t/2)2Hn+1(0)et2/2Hn(t/2)2Hn+1(0)et2/2Hn(t/2)),n=odd\displaystyle=\begin{pmatrix}\frac{1}{H_{n+1}(0)}e^{-t^{2}/2}{H_{n+1}(t/\sqrt{2})}\\ -\frac{\sqrt{2}}{{H_{n+1}(0)}}e^{-t^{2}/2}{H_{n}(t/\sqrt{2})}\\ -\frac{\sqrt{2}}{H_{n+1}(0)}e^{-t^{2}/2}{H_{n}(t/\sqrt{2})}\end{pmatrix},n=\textnormal{odd}
ψ(t)\displaystyle\psi_{*}(t) =(12Hn(0)et2/2Hn+1(t/2)1Hn(0)et2/2Hn(t/2)1Hn(0)et2/2Hn(t/2)),n=even.\displaystyle=\begin{pmatrix}-\frac{1}{{\sqrt{2}H_{n}(0)}}e^{-t^{2}/2}{H_{n+1}(t/\sqrt{2})}\\ \frac{1}{{H_{n}(0)}}e^{-t^{2}/2}{H_{n}(t/\sqrt{2})}\\ \frac{1}{H_{n}(0)}e^{-t^{2}/2}{H_{n}(t/\sqrt{2})}\end{pmatrix},n=\textnormal{even}.

which follows from a simple calculation, we obtain D(v,α)=D1(v,α)+D3(v,α)D(v,\alpha)=D_{1}(v,\alpha)+D_{3}(v,\alpha) where

D1(v,α)\displaystyle D_{1}(v,\alpha) =0et2/2×\displaystyle=\int_{0}^{\infty}e^{-t^{2}/2}\times
{1Hn+1(0)Hn+1(t/2)(g1(z(v,α)(t),α)g1(z(v,α)(t),α))n=odd12Hn(0)Hn+1(t/2)g1(z(v,α)(t),α)n=evendt,\displaystyle\left\{\begin{array}[]{cc}\frac{1}{H_{n+1}(0)}H_{n+1}(t/\sqrt{2})\left(g_{1}(z_{*}(v,\alpha)(t),\alpha)-g_{1}(z_{*}(-v,\alpha)(t),\alpha)\right)&n=\textnormal{odd}\\ -\frac{1}{\sqrt{2}H_{n}(0)}H_{n+1}(t/\sqrt{2})g_{1}(z_{*}(v,\alpha)(t),\alpha)&n=\textnormal{even}\\ \end{array}\right.dt, (D.15)
D3(v,α)\displaystyle D_{3}(v,\alpha) =0et2/2×\displaystyle=\int_{0}^{\infty}e^{-t^{2}/2}\times
{2Hn+1(0)Hn(t/2)(g3(z(v,α)(t),α)g3(z(v,α)(t),α))n=odd1Hn(0)Hn(t/2)g3(z(v,α)(t),α)n=evendt.\displaystyle\left\{\begin{array}[]{cc}-\frac{\sqrt{2}}{H_{n+1}(0)}H_{n}(t/\sqrt{2})\left(g_{3}(z_{*}(v,\alpha)(t),\alpha)-g_{3}(z_{*}(-v,\alpha)(t),\alpha)\right)&n=\textnormal{odd}\\ \frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})g_{3}(z_{*}(v,\alpha)(t),\alpha)&n=\textnormal{even}\\ \end{array}\right.dt. (D.18)

As described in Theorem 2.9, we are able to evaluate these integrals by following the procedure in Remark 2.11:

Lemma D.1.

Let kk\in\mathbb{N} be so that

n={2k1n=odd2kn=even.\displaystyle n=\begin{cases}2k-1&n=\textnormal{odd}\\ 2k&n=\textnormal{even}\end{cases}.

Then

  1. (1)

    For n=oddn=\textnormal{odd}, the following holds

    2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =22π(2k1)(2k)!!1+(2k1)2.\displaystyle=-\frac{2\sqrt{2\pi}(2k-1)(2k)!!}{\sqrt{1+(2k-1)^{2}}}.

    Furthermore, let

    ckj\displaystyle c_{kj} :=(2(2k1)2j)!j!(j+1)!(2k1j)!42k12k12j×\displaystyle:=\frac{(2(2k-1)-2j)!}{j!(j+1)!(2k-1-j)!^{4}}\frac{2k-1}{2k-1-2j}\times
    (522kj)(2k(1+j+12kj)+1+j),\displaystyle\left(5-\frac{2}{2k-j}\right)\left(2k\left(1+\frac{j+1}{2k-j}\right)+1+j\right),
    dkj\displaystyle d_{kj} :=(2(2k1)2j)!j!(j+1)!(2k1j)!42k(j+1)2kj,\displaystyle:=\frac{(2(2k-1)-2j)!}{j!(j+1)!(2k-1-j)!^{4}}\frac{2k(j+1)}{2k-j},

    for j=0,,2k1j=0,\ldots,2k-1. Then

    Dvvv′′′(0,0)\displaystyle D^{\prime\prime\prime}_{vvv}(0,0) =3(2k)!!42π(2k1)2k(1+(2k1)2)3/2j=02k1(ckj+dkj)\displaystyle=-\frac{3(2k)!!^{4}\sqrt{2\pi}(2k-1)}{2k(1+(2k-1)^{2})^{3/2}}\sum_{j=0}^{2k-1}\left(c_{kj}+d_{kj}\right)
  2. (2)

    For n=evenn=\textnormal{even}, then

    2Dvα(0,0)\displaystyle\frac{\partial^{2}D}{\partial v\partial\alpha}(0,0) =2k2π(2k1)!!,\displaystyle=2k\sqrt{2\pi}(2k-1)!!,

    and

    Dvv′′(0,0)\displaystyle D^{\prime\prime}_{vv}(0,0) =8(1)kk42π(2k1)!!3(4k)!3(1+16k2(2k+1)).\displaystyle=\frac{8(-1)^{k}k^{4}\sqrt{2\pi}(2k-1)!!^{3}}{(4k)!^{3}}\left(1+16k^{2}(2k+1)\right).
Proof.

We simply differentiate the expressions (D.15) and (D.18) and use (D.1),

z\displaystyle z^{\prime} =(2(n+1)1+n2Hn+1(0)Hn(t/2)11+n2Hn+1(0)Hn+1(t/2)n1+n2Hn+1(0)Hn+1(t/2)),n=odd,\displaystyle=\begin{pmatrix}\frac{\sqrt{2}(n+1)}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n}(t/\sqrt{2})\\ \frac{1}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n+1}(t/\sqrt{2})\\ \frac{n}{\sqrt{1+n^{2}}H_{n+1}(0)}H_{n+1}(t/\sqrt{2})\end{pmatrix},\quad n=\textnormal{odd},
z\displaystyle z^{\prime} =(1Hn(0)Hn(t/2)12(n+1)Hn(0)Hn+1(t/2)n2(n+1)Hn(0)Hn+1(t/2)),n=even,\displaystyle=\begin{pmatrix}\frac{1}{H_{n}(0)}H_{n}(t/\sqrt{2})\\ \frac{1}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})\\ \frac{n}{\sqrt{2}(n+1)H_{n}(0)}H_{n+1}(t/\sqrt{2})\end{pmatrix},\quad n=\textnormal{even}, (D.19)

by (D.3) and (D.2), where z:=zv(0,0)z^{\prime}:=\frac{\partial z_{*}}{\partial v}(0,0). Following Remark 2.11, see step (a), we then characterize z=′′2zv2(0,0)z{{}^{\prime\prime}}=\frac{\partial^{2}z_{*}}{\partial v^{2}}(0,0) using the higher variational equations:

z˙1′′\displaystyle\dot{z}_{1}^{\prime\prime} =tz1′′z2′′z3′′2z1z3,\displaystyle=tz_{1}^{\prime\prime}-z_{2}^{\prime\prime}-z_{3}^{\prime\prime}-2z_{1}^{\prime}z_{3}^{\prime},
z˙2′′\displaystyle\dot{z}_{2}^{\prime\prime} =z1′′,\displaystyle=z_{1}^{\prime\prime},
z˙3′′\displaystyle\dot{z}_{3}^{\prime\prime} =nz1′′+n(z1)2.\displaystyle=nz_{1}^{\prime\prime}+n(z_{1}^{\prime})^{2}.

By the remaining steps b,c,d we obtain the results. The details are identical to Lemma 3.4 and therefore left out. ∎

For n=oddn=\textnormal{odd}, notice that whereas all dkj>0d_{kj}>0, the sign of ckjc_{kj} – due to the factor 2k12j2k-1-2j in the denominator – changes from j=k1j=k-1 to j=kj=k, in such a way that

ckj{<0for alljk>0for alljk1.\displaystyle c_{kj}\begin{cases}<0&\text{for all}\quad j\geq k\\ >0&\text{for all}\quad j\leq k-1\end{cases}.

However, a simple calculation shows that

|ck(k1l)ck(k+l)|=|5(k+l)+35(kl)2×k+l+1kl×(2k+2l)(2k2l1)(k+l+1)2(k+1l)2|>1×1×1,\displaystyle\left|\frac{c_{k(k-1-l)}}{c_{k(k+l)}}\right|=\left|\frac{5(k+l)+3}{5(k-l)-2}\times\frac{k+l+1}{k-l}\times\frac{(2k+2l)\cdots(2k-2l-1)}{(k+l+1)^{2}\cdots(k+1-l)^{2}}\right|>1\times 1\times 1,

for all l=0,,k1l=0,\ldots,k-1. But then

j=02k1ckj=l=0k|ck(k+l)|(|ck(k1l)ck(k+l)|1)>0,\sum_{j=0}^{2k-1}c_{kj}=\sum_{l=0}^{k}|c_{k(k+l)}|\left(\left|\frac{c_{k(k-1-l)}}{c_{k(k+l)}}\right|-1\right)>0,

and hence Dvvv′′′(0,0)<0D^{\prime\prime\prime}_{vvv}(0,0)<0 for all kk. By singularity theory [10], these expressions therefore complete the proof of Proposition 5.7.

References

  • [1] M. Abramowitz and A. Stegun. Handbook of Mathematical Functions, volume 55. National Bureau of Standards Applied Mathematics Series, Washington, DC, 1964.
  • [2] M. Brøns, M. Krupa, and M. Wechselberger. Mixed mode oscillations due to the generalized canard phenomenon. In W. Nagata and N. Sri Namachchivaya, editors, Bifurcation Theory and Spatio-Temporal Pattern Formation, volume 49 of Fields Institute Communications, pages 39–64. American Mathematical Society, 2006.
  • [3] J. Carr. Applications of centre manifold theory, volume 35. New York: Springer-Verlag, 1981.
  • [4] M. Desroches, J. Guckenheimer, B. Krauskopf, C. Kuehn, H. M. Osinga, and M. Wechselberger. Mixed-mode oscillations with multiple time scales. Siam Review, 54(2):211–288, 2012.
  • [5] M. di Bernardo, C. J. Budd, A. R. Champneys, and P. Kowalczyk. Piecewise-smooth Dynamical Systems: Theory and Applications. Springer Verlag, 2008.
  • [6] V. M. Falkner and S. W. Skan. Solutions of the boundary-layer equations in aerodynamics. Philosophical Magazine, 12:865–896, 865–896, 1931.
  • [7] N. Fenichel. Persistence and smoothness of invariant manifolds for flows. Indiana University Mathematics Journal, 21:193–226, 1971.
  • [8] N. Fenichel. Asymptotic stability with rate conditions. Indiana University Mathematics Journal, 23:1109–1137, 1974.
  • [9] N. Fenichel. Geometric singular perturbation theory for ordinary differential equations. J. Diff. Eq., 31:53–98, 1979.
  • [10] M. Golubitsky and D. G. Schaeffer. Singularities and Groups in Bifurcation Theory, volume 51. Springer, 1984.
  • [11] M. R. Jeffrey and A. Colombo. The two-fold singularity of discontinuous vector fields. SIAM Journal on Applied Dynamical Systems, 8(2):624–640, January 2009.
  • [12] C.K.R.T. Jones. Geometric Singular Perturbation Theory, Lecture Notes in Mathematics, Dynamical Systems (Montecatini Terme). Springer, Berlin, 1995.
  • [13] J. Knobloch. Bifurcation of degenerate homoclinic orbits in reversible and conservative systems. Journal of Dynamics and Differential Equations, 9(3):427–444, 1997.
  • [14] K. U. Kristiansen and S. J. Hogan. Resolution of the piecewise smooth visible-invisible two-fold singularity in R3 using regularization and blowup. Journal of Nonlinear Science, 29(2):723–787, 2018.
  • [15] K. Uldall Kristiansen and S. J. Hogan. On the use of blowup to study regularizations of singularities of piecewise smooth dynamical systems in R3. SIAM Journal on Applied Dynamical Systems, 14(1):382–422, 2015.
  • [16] M. Krupa and M. Wechselberger. Local analysis near a folded saddle-node singularity. Journal of Differential Equations, 248(12):2841–2888, 2010.
  • [17] J. Llibre and M. Messias. Large amplitude oscillations for a class of symmetric polynomial differential systems in R3. Anais Da Academia Brasileira De Ciencias, 79(4):563–575, 2007.
  • [18] J. Mitry and M. Wechselberger. Folded saddles and faux canards. Siam Journal on Applied Dynamical Systems, 16(1):546–596, 2017.
  • [19] S. Nosé. A unified formulation of the constant temperature molecular-dynamics methods. Journal of Chemical Physics, 81(1):511–519, 1984.
  • [20] F. W. J. Olver, D. W. Lozier, R. F. Boisvert, and C. W. Clark. Nist handbook of mathematical functions. Journal of Geometry and Symmetry in Physics, pages 99–104, 2011.
  • [21] C. Sparrow and H. P.F. Swinnerton-Dyer. The Falkner-Skan equation II: Dynamics and the bifurcations of P- and Q-orbits. Journal of Differential Equations, 183(1):1–55, 2002.
  • [22] H.P.F. Swinnerton-Dyer and C.T. Sparrow. The Falkner-Skan Equation I. The Creation of Strange Invariant Sets. Journal of Differential Equations, 1995.
  • [23] P. Swinnerton-Dyer and T. Wagenknecht. Some third-order ordinary differential equations. Bulletin of the London Mathematical Society, 40(5):725–748, 2008.
  • [24] P. Szmolyan and M. Wechselberger. Canards in 3\mathbb{R}^{3}. J. Diff. Eq., 177(2):419–453, December 2001.
  • [25] P. Szmolyan and M. Wechselberger. Relaxation oscillation in ℝ3. Journal of Differential Equations, 200(1):69–104, 2004.
  • [26] A. Vanderbauwhede. Bifurcation of degenerate homoclinics. Results in Mathematics, 21:211–223, 1992.
  • [27] M. Wechselberger. Extending Melnikov theory to invariant manifolds on non-compact domains. Dynamical Systems, 17(3):215–233, 2002.
  • [28] M. Wechselberger. Existence and bifurcation of canards in 3\mathbb{R}^{3} in the case of a folded node. SIAM Journal on Applied Dynamical Systems, 4(1):101–139, January 2005.