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

Oscillatory motions in the Restricted 3-Body Problem: A Functional Analytic Approach

Jaime Paradela Departament de Matemàtiques, Universitat Politècnica de Catalunya, Diagonal 647, 08028 Barcelona, Spain [email protected]  and  Susanna Terracini Dipartimento di Matematica Giuseppe Peano, Università di Torino, Via Carlo Alberto 10, 10123, Torino, Italy [email protected]
Abstract.

A fundamental question in Celestial Mechanics is to analyze the possible final motions of the Restricted 33-body Problem, that is, to provide the qualitative description of its complete (i.e. defined for all time) orbits as time goes to infinity. According to the classification given by Chazy back in 1922, a remarkable possible behaviour is that of oscillatory motions, where the motion qq of the massless body is unbounded but returns infinitely often inside some bounded region:

lim supt±|q(t)|=andlim inft±|q(t)|<.\limsup_{t\to\pm\infty}|q(t)|=\infty\qquad\qquad\text{and}\qquad\qquad\liminf_{t\to\pm\infty}|q(t)|<\infty.

In contrast with the other possible final motions in Chazy’s classification, oscillatory motions do not occur in the 22-body Problem, while they do for larger numbers of bodies. A further point of interest is their appearance in connection with the existence of chaotic dynamics.

In this paper we introduce new tools to study the existence of oscillatory motions and prove that oscillatory motions exist in a particular configuration known as the Restricted Isosceles 33-body Problem (RI3BP) for almost all values of the angular momentum. Our method, which is global and not limited to nearly integrable settings, extends the previous results [GPSV21] by blending variational and geometric techniques with tools from nonlinear analysis such as topological degree theory. To the best of our knowledge, the present work constitutes the first complete analytic proof of existence of oscillatory motions in a non perturbative regime.

Key words and phrases:
Restricted 33-body problem, parabolic motions, oscillatory motions, chaotic dynamics
2020 Mathematics Subject Classification:
34C28, 37B10, 70G75 70F07, 37C83, 70K44.
This work was partially done while the first author was visiting the Mathematics Department of the University of Turin and he thanks the department for their hospitality and pleasant working atmosphere. This project has received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No 757802).

1. Introduction

One of the oldest questions in Dynamical Systems is to understand the mechanisms driving the global dynamics of the 33-body problem, which models the motion of three bodies interacting through Newtonian gravitational force. The 33-body Problem is called restricted if one of the bodies has mass zero and the other two have strictly positive masses. In this limit problem, the massless body is affected by, but does not affect, the motion of the massive bodies. A fundamental question concerning the global dynamics of the Restricted 33-body Problem is the study of its possible final motions, that is, the qualitative description of its complete (defined for all time) orbits as time goes to infinity. In 1922 Chazy gave a complete classification of the possible final motions of the Restricted 33-body Problem [Cha22]. To describe them we denote by qq the position of the massless body in a Cartesian reference frame with origin at the center of mass of the primaries.

Theorem 1.1 ([Cha22]).

Every solution of the Restricted 33-body Problem defined for all (future) times belongs to one of the following classes

  • B (bounded): supt0|q(t)|<\sup_{t\geq 0}|q(t)|<\infty.

  • P (parabolic) |q(t)||q(t)|\to\infty and |q˙(t)|0|\dot{q}(t)|\to 0 as tt\to\infty.

  • H (hyperbolic): |q(t)||q(t)|\to\infty and |q˙(t)|c>0|\dot{q}(t)|\to c>0 as tt\to\infty.

  • O (oscillatory) lim supt|q(t)|=\limsup_{t\to\infty}|q(t)|=\infty and lim inft|q(t)|<\liminf_{t\to\infty}|q(t)|<\infty.

Notice that this classification also applies for tt\to-\infty. We distingish both cases adding a superindex ++ or - to each of the cases, e.g. H+H^{+} and HH^{-}.

Bounded, parabolic and hyperbolic motions also exist in the 22-body Problem, and examples of each of these classes of motion in the Restricted 33-body Problem were already known by Chazy. However, the existence of oscillatory motions in the Restricted 33-body Problem was an open question for a long time. Their existence was first established by Sitnikov in a particular configuration of the Restricted 33-body Problem nowadays known as the Sitnikov problem.

1.1. The Moser approach to the existence of oscillatory motions

After Sitnikov’s work, Moser gave a new proof of the existence of oscillatory motions in the Sitnikov problem [Mos01]. His approach makes use of tools from the geometric theory of dynamical systems, in particular, hyperbolic dynamics. More concretely, Moser considered an invariant periodic orbit “at infinity” (see Section 1.3) which is degenerate (the linearized vector field vanishes) but posseses stable and unstable invariant manifolds. Then, he proved that its stable and unstable manifolds intersect transversally. Close to this intersection, he built a section Σ\Sigma transverse to the flow and established the existence of a (non trivial) locally maximal hyperbolic set 𝒳\mathcal{X} for the Poincaré map ΦΣ\Phi_{\Sigma} induced on Σ\Sigma. The dynamics of ΦΣ\Phi_{\Sigma} restricted to 𝒳Σ\mathcal{X}\subset\Sigma is moreover conjugated to the shift

σ:(σω)k=ωk+1\sigma:\mathbb{N}^{\mathbb{Z}}\to\mathbb{N}^{\mathbb{Z}}\qquad\qquad(\sigma\omega)_{k}=\omega_{k+1}

acting on the space of infinite sequences. Namely, 𝒳\mathcal{X} is a horseshoe with ”infinitely many legs” for the map ΦΣ\Phi_{\Sigma}. By construction, sequences ω=(,ωn,ωn+1ω0,ωn1ωn)\omega=(\dots,\omega_{-n},\omega_{-n+1}\dots\omega_{0},\dots\omega_{n-1}\dots\omega_{n}\dots)\in\mathbb{N}^{\mathbb{Z}} for which lim supnωn\limsup_{n\to\infty}\omega_{n} (respectively lim supnωn\limsup_{n\to-\infty}\omega_{n}) correspond to complete motions of the Sitnikov problem which are oscillatory in the future (in the past).

Moser’s ideas have been very influential. In [LS80a] Simó and Llibre implemented Moser’s approach in the Restricted Circular 33-body Problem (RC3BP) in the region of the phase space with large Jacobi constant provided the values of the ratio between the masses of the massive bodies is small enough. Their result was later extended by Xia [Xia94] and closed by Guardia, Martín and Seara in [GMS16] where oscillatory motions for the RC3BP for all mass ratios are constructed in the region of the phase space with large Jacobi constant. The same result is obtained in [CGM+21] for low values of the Jacobi constant relying on a computer assisted proof. In [GSMS17] and [SZ20], the Moser approach is applied to the Restricted Elliptic 33-body Problem and the Restricted 4 Body Problem respectively. For the 33-body Problem, results in certain symmetric configurations (which reduce the dimension of the phase space) were obtained in [Ale69] and [LS80b]. Another interesting result, which however holds for non generic choices of the 3 masses, is obtained in [Moe07]. In the recent preprint [GMPS22], the first author together with Guardia, Martín and Seara, has proved the existence of oscillatory motions in the planar 33-body Problem (5 dimensional phase space after symplectic reductions) for all choices of the masses (except all equal) and large total angular momentum.

The first main ingredient in Moser’s strategy is the detection of a transversal intersection between the invariant manifolds of the periodic orbit at infinity. Yet, checking the occurrence of this phenomenon in a physical model is rather problematic, and in general little can be said except for perturbations of integrable systems with a hyperbolic fixed point whose stable and unstable manifolds coincide along a homoclinic manifold. As far as the authors know, all the previous works concerning the existence of oscillatory motions in the 33-body Problem (restricted or not) adopt a perturbative approach to prove the existence of transversal intersections between the stable and unstable manifolds of infinity. In some cases the perturbative regime is obtained by assuming that certain parameter related to the motion of the massive bodies (in general the ratio between the masses of the massive bodies or the eccentriticy of their orbit) is small, and others by working in a region of the phase space where the massless body is located far away from the primaries. The latter situation falls in what is usually called singular perturbation theory and (in general) needs a much more involved analysis than the former one, usually referred to as regular perturbation theory.

The second key ingredient is the construction of a horseshoe close to the transversal intersections of the invariant manifolds. For the Sitnikov and Isosceles Restricted 33-body Problem (which is introduced in Section 1.3) are non autonomous Hamiltonian systems with 1+1/21+1/2 degrees of freedom (33 dimensional phase space), its dynamics can be reduced to the study of a two dimensional area preserving map in which the periodic orbit at infinity becomes a fixed point which, despite being degenerate, behaves as a hyperbolic fixed point. The same happens in the RC3BP after reducing by rotational symmetry and in certain symmetric configurations of the 3BP. In all of these problems Moser’s ideas for constructing a horseshoe close to the transverse intersections between the invariant manifolds of the parabolic fixed point can be implemented directly. In the planar 33-body Problem, the dynamics can be reduced to a 4 dimensional symplectic map and the parabolic fixed point becomes a 2 dimensional (degenerate) normally hyperbolic invariant manifold. Due to the existence of central directions the construction of the horseshoe in [GMPS22] becomes much more involved. In [Moe07], the author analyzes orbits which pass close to triple collision. In this setting, the close encounters with triple collision, produce stretching also in the central directions.

An approach different in nature from Moser’s is developed by Galante and Kaloshin in [GK11]. By making use of Aubry-Mather theory and semi-infinite regions of instability, the authors prove the existence of oscillatory orbits for the RC3BP with a realistic value of the mass ratio.

1.2. The fundamental question in Celestial Mechanics

Besides the question of its existence, is the question of their abundance. In the conference in honor of the 70th anniversary of Alexeev, Arnol’d posed the following question (cfr [GK11]).

Question 1.2.

Is the Lebesgue measure of the set of oscillatory motions positive?

This question was considered by Arnol’d to be the fundamental issue of Celestial Mechanics. It has been conjectured by Alexeev that the Lebesgue measure is zero. Neverthless, this conjecture remains wide open. The only partial results in this direction are due to Gorodetski and Kaloshin [GK12]. They consider the RC3BP and the Sitnikov problem and prove that for both problems and a Baire generic subset of an open set of parameters (eccentricity in the Sitnikov problem and mass ratio in the RC3BP), the Hausdorff dimension of the set of oscillatory motions is maximal.

1.3. The Isosceles configuration of the Restricted 33-body Problem

In the present work we consider a particular configuration of the Restricted 33-body Problem known as the Restricted Isosceles 33-body Problem. In this configuration, the two primaries have equal masses m0=m1=1/2m_{0}=m_{1}=1/2 and move periodically on a degenerate ellipse of eccentricity one (a line), according to the Kepler laws for the motion of the 22-body Problem. The massless particle moves on the plane perpendicular to the line along which the primaries move (see Figure 1.1).

In the plane of motion of the massless body we fix a Cartesian reference frame with origin at the point where the line along which the primaries move intersects the plane. Then, in Cartesian coordinates (q,p,t)4×𝕋{q=0}(q,p,t)\in\mathbb{R}^{4}\times\mathbb{T}\setminus\{q=0\}, the motion of the massless body is given by the Hamiltonian system

H(q,p,t)=|p|22Vcart(q,t)Vcart(q,t)=1|q|2+ρ2(t).H(q,p,t)=\frac{|p|^{2}}{2}-V_{\mathrm{cart}}(q,t)\qquad\qquad V_{\mathrm{cart}}(q,t)=\frac{1}{\sqrt{|q|^{2}+\rho^{2}(t)}}.

where ρ(t):𝕋[0,1/2]\rho(t):\mathbb{T}\to[0,1/2] is a half of the distance between the primaries.

Remark 1.

One can obtain an explicit expression of the function ρ(t)\rho(t) after introducing the change of variables t=usinut=u-\sin u, commonly known as the Kepler equation. When expressed in terms of the new variable uu (which is the eccentric anomaly) we have ρ(t(u))=(1cosu)/2\rho(t(u))=(1-\cos u)/2. Yet, our analysis does not require to have an explicit expression of the function ρ(t)\rho(t), so we work directly with the original variable tt.

Refer to caption
Figure 1.1. Sketch of the motion in the Restricted Isosceles 33-body Problem.

It will be convenient for our analysis to introduce polar coordinates (r,α,t,y,G)+×𝕋2×2(r,\alpha,t,y,G)\in\mathbb{R}_{+}\times\mathbb{T}^{2}\times\mathbb{R}^{2} where q=(rcosα,rsinα)q=(r\cos\alpha,r\sin\alpha) and (y,G)(y,G) denote the conjugate momenta to (r,α)(r,\alpha). In polar coordinates, the Hamiltonian of the Restricted Isosceles 33-body Problem reads

(1.1) H(r,t,y,G)=y22+G22r2V(r,t)V(r,t)=1r2+ρ2(t).H(r,t,y,G)=\frac{y^{2}}{2}+\frac{G^{2}}{2r^{2}}-V(r,t)\qquad\qquad V(r,t)=\frac{1}{\sqrt{r^{2}+\rho^{2}(t)}}.

We inmediately notice that GG is a conserved quantity for the flow of (1.1). It is therefore natural to consider the one-parameter family of Hamiltonian systems

(1.2) HG(r,t,y)=H(r,t,y,G)(r,t,y)+×𝕋×.H_{G}(r,t,y)=H(r,t,y,G)\qquad\qquad(r,t,y)\in\mathbb{R}_{+}\times\mathbb{T}\times\mathbb{R}.

Since limrV(r,t)=0\lim_{r\to\infty}V(r,t)=0, for all GG\in\mathbb{R} the Hamiltonian (1.2) posses a periodic orbit at infinity

(1.3) γ={r=,y=0}+×𝕋×.\gamma_{\infty}=\{r=\infty,\ y=0\}\subset\mathbb{R}_{+}\times\mathbb{T}\times\mathbb{R}.

In [GPSV21], the first author together with M. Guardia, T. Seara and C.Vidal, proved the following result.

Theorem 1.3 ([GPSV21]).

Consider the Hamiltonian system HGH_{G} defined in (1.2). Denote by X+X^{+} (respectively YY^{-}) either H+,P+,B+H^{+},P^{+},B^{+} or OS+OS^{+} (respectively H,P,BH^{-},P^{-},B^{-} or OSOS^{-}) according to Chazy’s classification in Theorem 1.1. Then, there exists G1G_{*}\gg 1 such that for all GG\in\mathbb{R} such that |G|G|G|\geq G_{*}, the Hamiltonian system HGH_{G} satisfies

X+YX^{+}\cap Y^{-}\neq\emptyset

for all possible combinations of X+X^{+} and YY^{-}.

Theorem 1.3 is proved by exploiting the fact that for GG large enough, in a suitable region of the phase space, the Hamiltonian HGH_{G} can be studied as a perturbation of the (integrable) 22-body Problem. This allowed the authors to prove that the periodic orbit γ\gamma_{\infty} posses global stable and unstable invariant manifolds which intersect transversally (see Theorem 4.1). As a corollary of this result, a rather straightforward implementation of Moser’s ideas shows the truth of Theorem 1.3.

The following is the first main result of the present work.

Theorem 1.4.

Consider the Hamiltonian system HGH_{G} defined in (1.2). Denote by X+X^{+} (respectively YY^{-}) either H+,P+,B+H^{+},P^{+},B^{+} or OS+OS^{+} (respectively H,P,BH^{-},P^{-},B^{-} or OSOS^{-}) according to Chazy’s classification in Theorem 1.1. Then, for almost all GG\in\mathbb{R} the Hamiltonian system HGH_{G} satisfies

X+YX^{+}\cap Y^{-}\neq\emptyset

for all possible combinations of X+X^{+} and YY^{-}.

To the best of our knowledge, Theorem 1.4 is the first complete analytic proof of the existence of oscillatory motions relying upon a global analytical approach rather than on perturbative techniques. Some interesting related works, where the existence of oscillatory motions is obtained in a setting which is not close to integrable, are [Moe07] and [CGM+21]. While in [Moe07] the author shows the existence of oscillatory motions in the 33-body Problem close to triple collision (small values of the total angular momentum), in [CGM+21] the authors obtain a computer assisted proof of the existence of oscillatory motions in the Restricted Circular 33-body Problem for small values of the Jacobi constant.

Theorem 1.4 is indeed obtained as a consequence of the following result.

Theorem 1.5.

Let {lj}\{l_{j}\}\subset\mathbb{Z} be an increasing sequence and define the time intervals Ij=[(ljlj1)/2,(lj+1lj)/2]I_{j}=[(l_{j}-l_{j-1})/2,\ (l_{j+1}-l_{j})/2]. Then, for almost all GG\in\mathbb{R}, all ε>0\varepsilon>0 and all RR sufficiently large, there exists an orbit rh(s):+r_{\mathrm{h}}(s):\mathbb{R}\to\mathbb{R}_{+} of (1.2) homoclinic to γ\gamma_{\infty} and a constant L>0L>0 such that if the sequence {lj}\{l_{j}\}\subset\mathbb{Z} satisfies lj+1ljLl_{j+1}-l_{j}\geq L, then, for any sequence σ={σj}{0,1}\sigma=\{\sigma_{j}\}\subset\{0,1\}^{\mathbb{Z}} there exists an orbit rσ(s):+r_{\sigma}(s):\mathbb{R}\to\mathbb{R}_{+} of (1.1) such that , if σj=0\sigma_{j}=0

|rσ|C1(Ij)R|r_{\sigma}|_{C^{1}(I_{j})}\geq R

and if σj=1\sigma_{j}=1

|rσrh|C1(Ij)ε,|r_{\sigma}-r_{\mathrm{h}}|_{C^{1}(I_{j})}\leq\varepsilon,

Moreover, if σ\sigma has only a finite number of non zero entries, then rσr_{\sigma} is a homoclinic solution.

Theorem 1.5 can be read as follows. For almost all GG\in\mathbb{R} there exist an orbit rhr_{h} of (1.2) homoclinic to γ\gamma_{\infty} such that the following holds. Let z=(r,y,t)=(rh(0),r˙h(0),0)+××𝕋z_{*}=(r,y,t)=(r_{h}(0),\dot{r}_{h}(0),0)\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T}, let z=(r,y,t)=(,0,0)=γ{t=0}+××𝕋z_{\infty}=(r,y,t)=(\infty,0,0)=\gamma_{\infty}\cap\{t=0\}\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T} and denote by Φ\Phi the Poincaré map induced on the section {t=0}\{t=0\} by the flow to the Hamiltonian (1.2). Then, for any δ>0\delta>0 and any sequence {zk}k{z,z}\{z_{k}\}_{k\in\mathbb{Z}}\subset\{z_{\infty},z_{*}\}^{\mathbb{Z}} there exists a point zBδ(z0)z\in B_{\delta}(z_{0}) and a sequence {nk}k\{n_{k}\}_{k\in\mathbb{Z}}\in\mathbb{N}^{\mathbb{Z}} such that Φnk(z0)Bδ(zk)\Phi^{n_{k}}(z_{0})\in B_{\delta}(z_{k}) 111By Bδ(z)B_{\delta}(z_{\infty}) we mean the set {|y|δ,|r|1δ}\{|y|\leq\delta,\ |r|^{-1}\leq\delta\}.. The statement in Theorem 1.5 is indeed stronger since it also provides control on the orbit in all the intervals [(nknk1)/2,(nk+nk+1)/2][(n_{k}-n_{k-1})/2,(n_{k}+n_{k+1})/2].

The following corollary of Theorem 1.5 can obtained by nowadays well known arguments (see for example [MNT99] and [Koz83]).

Corollary 1.6.

For almost all GG\in\mathbb{R} the Restricted Isosceles 33-body Problem is not CωC^{\omega} integrable and has positive topological entropy.

1.4. Outline of the proof: new tools for the study of oscillatory motions

As in Moser’s approach, the first main step in our construction is to prove the existence of a homoclinic orbit to γ\gamma_{\infty}. Yet, in the setting of Theorem 1.5, geometric perturbation theory is not available since the Hamiltonian system HGH_{G} in (1.2) is not nearly integrable. Instead, we will adopt a global approach and deploy the powerful machinery of the theory of calculus of variations. In particular, we rephrase the problem of existence of homoclinic orbits to γ\gamma_{\infty} as that of the existence of critical points of a certain action functional 𝒜G\mathcal{A}_{G} (cfr 4.2) defined in a suitable Hilbert space D1,2D^{1,2} (cfr (4.1)). The existence of critical points of the action functional 𝒜G\mathcal{A}_{G} is obtained by a minmax argument tailored made for the present problem. The use of minmax techniques to study the existence and multiplicity results for homoclinic orbits in Hamiltonian systems has already been widely exploited in the literature (see for example [Sér92, CZES90, CZR91] and [MNT99]). In the variational approach to our problem, we face two main difficulties at this step: the phase space is not compact and the vector field presents singularities (corresponding to possible collision with the massive bodies). In order to overcome the first difficulty we make use of a renormalized action functional (see Remark 4) defined on a appropriately chosen functional space D1,2D^{1,2}. In order to avoid singularities and gain compactness we then perform a constrained deformation argument. With these techniques, together with a compactness property of the map d𝒜G:D1,2D1,2\mathrm{d}\mathcal{A}_{G}:D^{1,2}\to D^{1,2} (Struwe’s monotonicity trick), we are able to show that, for almost all values of the angular momentum GG 222See the discussion at the beginning of Section 4.2.1., there exists a Palais-Smale sequence in D1,2D^{1,2} which converges to a critical point of the action functional 𝒜G\mathcal{A}_{G}. This proves the existence of an orbit r~h\tilde{r}_{h} homoclinic to γ\gamma_{\infty}, which actually correspond to a doubly parabolic motion of our problem. It is worthwhile pointing out that half parabolic and hyperbolic motions for the nn-body problem have been obtained using variational methods in [MV09, MV20] with a different technique.

The homoclinc orbit r~h\tilde{r}_{h} obtained in this way is associated with an intersection between the stable and unstable manifolds of the periodic orbit γ\gamma_{\infty}. To proceed further, though we can not tell whether this intersection is transversal or not, we may rely on our minmax construction to deduce some topological transversality. This can be achieved by a topological degree argument based on a general result by Hofer ([Hof86]). More precisely, we exploit the mountain pass characterization of r~h\tilde{r}_{h} to show that for almost all values of the angular momentum GG (except possibly a finite set of values) there exists a (possibly different) critical point rhr_{h} of the action functional 𝒜G\mathcal{A}_{G} for which the Leray-Schauder index of the map 𝒜G:D1,2D1,2\nabla\mathcal{A}_{G}:D^{1,2}\to D^{1,2} at rhr_{h} is well defined and different from zero 333In Proposition 5.10 show that the topological degree being non zero implies that the intersection between the invariant manifolds of γ\gamma_{\infty} at rhr_{h} is topologically transverse.. This allows us to shadow finite segments of the homoclinic orbit rhr_{h}. The proof of Theorem 1.5 is then obtained by combining a suitable parabolic version of the Lambda lemma close to γ\gamma_{\infty} with the outer dynamics wich shadows finite segments of rhr_{h}.

1.5. Organization of the paper

In Section 2 we recall some well known facts about the 22-body Problem. Then, in Section 3 we analyze the dynamics around the periodic orbit γ\gamma_{\infty}. In particular, the existence of stable and unstable manifolds W±(γ;G)W^{\pm}(\gamma_{\infty};G) and a parabolic version of the lambda lemma close to γ\gamma_{\infty}. In Section 4 we introduce the variational formulation and prove the existence of a homoclinic orbit to γ\gamma_{\infty} by means of a minmax argument. Then, in Section 5 we obtain a (possibly different) homoclinic orbit associated with a topologically transverse intersection between W±(γ;G)W^{\pm}(\gamma_{\infty};G). Finally in Section 6 we combine the parabolic Lambda lemma of Section 3 together with the robustness of the topological degree under perturbations to construct “multibump” homoclinics and finish the proof of Theorem 1.5.

2. The 22-body Problem

In this section we recall some well known facts about the 22-body Problem (2BP) which will be used in the following. In polar coordinates, the Hamiltonian of the 2BP reads (compare (1.1))

(2.1) H2BP(r,α,y,G)=y22+G22r21r.H_{\mathrm{2BP}}(r,\alpha,y,G)=\frac{y^{2}}{2}+\frac{G^{2}}{2r^{2}}-\frac{1}{r}.

As for (1.1), the rotational symmetry implies that GG is a conserved quantity, so we look at (2.1) as a one-parameter family of Hamiltonian functions H2BP,G(r,y)H_{\mathrm{2BP},G}(r,y). For each GG\in\mathbb{R} the Hamiltonian H2BP,G(r,y)H_{\mathrm{2BP},G}(r,y) is integrable and the motion can be classified in terms of the value of the energy: negative values correspond to elliptic motions, positive energies correspond to hyperbolic motions and for zero energy the motion is parabolic.

It is also straightforward to check that for all GG\in\mathbb{R}

z={r=,y=0}+×.z_{\infty}=\{r=\infty,\ y=0\}\subset\mathbb{R}_{+}\times\mathbb{R}.

is a fixed point for the flow of (2.1)444To analyze this fixed point properly one should work in McGehee coordinates, which are introduced in Section 3.2.. Moreover, for all GG\in\mathbb{R} the fixed point zz_{\infty} posses stable and unstable manifolds which coincide along a one dimensional homoclinic manifold W2BPh(z,G)W^{h}_{2BP}(z_{\infty},G). The homoclinic orbit Wh(z,G)W^{h}(z_{\infty},G) is indeed the parabolic orbit of the 2BP with angular momentum GG.

Lemma 2.1.

There exist real analytic functions r0(u;G)r_{0}(u;G) and y0(u;G)y_{0}(u;G), defined for all uu\in\mathbb{R}, such that

(2.2) W2BPh(z;G)={r=r0(u;G),y=y0(u;G),u}.W^{h}_{\mathrm{2BP}}(z_{\infty};G)=\{r=r_{0}(u;G),\ y=y_{0}(u;G),\ u\in\mathbb{R}\}.

Moroever, r0(u;G)G2/2r_{0}(u;G)\geq G^{2}/2 for all uu\in\mathbb{R} and

r0(u;G)u2/3y0(u;G)u1/3asu±.r_{0}(u;G)\sim u^{2/3}\qquad\qquad y_{0}(u;G)\sim u^{-1/3}\qquad\qquad\text{as}\qquad u\to\pm\infty.

In addition, for any G,GG,G_{*}\in\mathbb{R} we have

|r0(u;G)r0(u;G)||G2G02|asu±.|r_{0}(u;G)-r_{0}(u;G_{*})|\lesssim|G^{2}-G_{0}^{2}|\qquad\qquad\text{as}\qquad u\to\pm\infty.
Remark 2.

In the last item of Lemma 2.1 we compare solutions associated with different values of the angular momentum GG. The fact that we need that kind of information in our argument is due to a technical step (Struwe’s monotonicity trick) in Section 4 (see Remark 4 and Lemma 4.7).

Proof.

A proof of the first two items can be found in [MP94], where the authors also show that

rh(u;G)=G2(τ2(u)+1)2foru=G32(τ(u)+τ3(u)3).r_{h}(u;G)=\frac{G^{2}(\tau^{2}(u)+1)}{2}\qquad\qquad\text{for}\qquad\qquad u=\frac{G^{3}}{2}\left(\tau(u)+\frac{\tau^{3}(u)}{3}\right).

One can check that for τ\tau\in\mathbb{R} the second equality admits the unique inverse

τ(u)=(3G3u+9G6u21)1/3(3G3u+9G6u21)1/3\tau(u)=\left(3G^{-3}u+\sqrt{9G^{-6}u^{2}-1}\right)^{1/3}-\left(3G^{-3}u+\sqrt{9G^{-6}u^{2}-1}\right)^{-1/3}

which for large uu yields that

τ(u)=(6G3u)1/3(1+𝒪(u1)).\tau(u)=(6G^{-3}u)^{1/3}\left(1+\mathcal{O}(u^{-1})\right).

Therefore, as u±u\to\pm\infty

rh(u;G)=G22+(6u)2/32(1+𝒪(u1))r_{h}(u;G)=\frac{G^{2}}{2}+\frac{(6u)^{2/3}}{2}\left(1+\mathcal{O}(u^{-1})\right)

and the conclusion follows. ∎

Define the local stable and unstable manifolds555One can prove that orbits starting at points in W2BP,loc+(z;G)W^{+}_{2BP,loc}(z_{\infty};G) (respectively W2BP,loc(z;G)W^{-}_{2BP,loc}(z_{\infty};G) ) are confined in the region {r>G2/2,y0}\{r>G^{2}/2,y\geq 0\} for all positive times (respectively in the region {r>G2/2,y0}\{r>G^{2}/2,y\leq 0\} for all negative times).

W2BP,loc+(z;G)=W2BPh(z;G){y>0}W2BP,loc(z;G)=W2BPh(z;G){y<0}.\begin{split}W^{+}_{2BP,loc}(z_{\infty};G)=&W^{h}_{2BP}(z_{\infty};G)\cap\{y>0\}\\ W^{-}_{2BP,loc}(z_{\infty};G)=&W^{h}_{2BP}(z_{\infty};G)\cap\{y<0\}.\end{split}

It is a standard fact that W2BP,loc±(z;G)W^{\pm}_{2BP,loc}(z_{\infty};G) are exact Lagrangian submanifolds so they can therefore be parametrized in terms of a generating function.

Lemma 2.2.

There exists S0(r;G):(G2/2,)+S_{0}(r;G):(G^{2}/2,\infty)\to\mathbb{R}_{+}, which satisfies

H2BP;G(r,rS0(r;G))=0H_{2BP;G}(r,\partial_{r}S_{0}(r;G))=0

and such that

W2BP,loc±(z;G)={(r,±rS0(r;G))+×:r>G2/2}.W^{\pm}_{2BP,loc}(z_{\infty};G)=\{(r,\pm\partial_{r}S_{0}(r;G))\in\mathbb{R}_{+}\times\mathbb{R}\colon r>G^{2}/2\}.

3. The dynamics close to γ\gamma_{\infty}

In this section we study the dynamics in a neighbourhood of the periodic orbit at infinity defined in (1.3). Despite being degenerate (the linearized vector field vanishes at γ\gamma_{\infty}) the flow close to the periodic orbit γ\gamma_{\infty} behaves in a similar way to the flow on a neighbourhood of a hyperbolic periodic orbit.

3.1. The local invariant manifolds

Let ϕGs\phi_{G}^{s} be the time ss flow associated with the Hamiltonian HGH_{G} defined in (1.2). It is a classical result by McGehee [McG73] (see also [BF04]) that γ\gamma_{\infty} posses local stable and unstable invariant manifolds (by πr,πy\pi_{r},\pi_{y} we denote the projection on the rr and yy coordinates of a point (r,y,t)+××𝕋(r,y,t)\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T})

(3.1) Wloc,R+(γ;G)={x+××𝕋:πrϕGs(x)R,πyϕs(x)1/R,s0}Wloc,R(γ;G)={x+××𝕋:πrϕGs(x)R,πyϕs(x)1/R,s0}\begin{split}W^{+}_{\mathrm{loc},R}(\gamma_{\infty};G)=&\{x\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T}\colon\pi_{r}\phi^{s}_{G}(x)\geq R,\ \pi_{y}\phi^{s}(x)\leq 1/R,\ \forall s\geq 0\}\\ W^{-}_{\mathrm{loc},R}(\gamma_{\infty};G)=&\{x\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T}\colon\pi_{r}\phi^{s}_{G}(x)\geq R,\ \pi_{y}\phi^{s}(x)\leq 1/R,\ \forall s\leq 0\}\\ \end{split}

It is also a standard fact that Wloc,R±(γ;G)W^{\pm}_{\mathrm{loc},R}(\gamma_{\infty};G) are exact Lagrangian submanifolds so they can be parametrized in terms of a generating function. The following result follows directly from the arguments in the proof of Theorem 4.4. in [GPSV21] (see Remark 3).

Proposition 3.1 ([GPSV21]).

Let HGH_{G} be the one parameter family of Hamiltonians defined in (1.2) and fix any G>0G_{*}>0. Then, there exist R>0R>0 such that for all G[G,G]G\in[-G_{*},G_{*}] there exist two functions S±(r,t;G):[R,)×𝕋S^{\pm}(r,t;G):[R,\infty)\times\mathbb{T}\to\mathbb{R}, real analytic on rr and GG, solutions to the Hamilton-Jacobi equation

HG(r,t,rS±(r,t;G))+tS±(r,t;G)=0H_{G}(r,t,\partial_{r}S^{\pm}(r,t;G))+\partial_{t}S^{\pm}(r,t;G)=0

and such that

Wloc,R±(γ;G)={(r,y,t)+××:r[R,),y=rS±(r,t;G)}.W^{\pm}_{\mathrm{loc},R}(\gamma_{\infty};G)=\{(r,y,t)\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{R}\colon r\in[R,\infty),\ y=\partial_{r}S^{\pm}(r,t;G)\}.

Moreover, if we let S0(r;G)S_{0}(r;G) be the function defined in Lemma 2.2, we have that

S±(r;G)S0(r;G)r3/2asr.S^{\pm}(r;G)-S_{0}(r;G)\sim r^{-3/2}\qquad\qquad\text{as}\qquad\qquad r\to\infty.
Remark 3.

In Theorem 4.4. in [GPSV21] the authors only show the existence of the generating functions S±(r,t;G)S^{\pm}(r,t;G) for large values of GG. The reason is that, under the hypothesis of large GG, they can extend the generating functions to a common domain where they can measure their diference. However, if we are only concerned with the existence and behaviors of the generating functions close to infinity, the problem is already perturbative, and the very same arguments apply to obtain the conlcusion in Proposition 3.1.

Define the global stable and unstable invariant manifolds

(3.2) W+(γ;G)=s0ϕGs(Wloc,R+(γ;G))W(γ;G)=s0ϕGs(Wloc,R(γ;G)).W^{+}(\gamma_{\infty};G)=\bigcup_{s\leq 0}\phi^{s}_{G}(W^{+}_{\mathrm{loc},R}(\gamma_{\infty};G))\qquad\qquad W^{-}(\gamma_{\infty};G)=\bigcup_{s\geq 0}\phi^{s}_{G}(W^{-}_{\mathrm{loc},R}(\gamma_{\infty};G)).

The analytic dependence of the functions S±(r,t;G)S^{\pm}(r,t;G) on rr and GG will be key to prove that transversal intersections (whenever they exist) between the global stable and unstable invariant manifolds (3.2) are topologically transverse except for (possibly) a finite subset of values of GG. This is key for the multibump construction. On the other hand, the estimate S±S0r3/2S^{\pm}-S^{0}\sim r^{-3/2} as rr\to\infty will be needed in the proof of certain technical steps in Lemma 5.5 (see Appendix A).

3.2. The parabolic Lambda Lemma

We now analyze the topology of the flow lines close to the periodic orbit γ\gamma_{\infty}. For that, it is convenient to introduce the McGehee transformation r=2/x2r=2/x^{2} in which the equations of motion associated with the Hamiltonian system HGH_{G} in (1.2) read

x˙=x34HGy=x3y4y˙=x34HGx=x441(1+x4ρ2(t)4)3/2+x6G28.\dot{x}=-\frac{x^{3}}{4}\frac{\partial H_{G}}{\partial y}=-\frac{x^{3}y}{4}\qquad\qquad\qquad\dot{y}=\frac{x^{3}}{4}\frac{\partial H_{G}}{\partial x}=-\frac{x^{4}}{4}\frac{1}{(1+\frac{x^{4}\rho^{2}(t)}{4})^{3/2}}+\frac{x^{6}G^{2}}{8}.
Refer to caption
Figure 3.1. Phase portrait of the 2BP in McGehee coordinates. The fixed point zz_{\infty} corresponds in McGehee coordinates to the origin in the (x,y)2(x,y)\in\mathbb{R}^{2} plane.

In this variables, the periodic orbit at infinity (1.3) now corresponds to the periodic orbit γ^={x=y=0,t𝕋}\hat{\gamma}_{\infty}=\{x=y=0,\ t\in\mathbb{T}\}. Following Moser [Mos01], we now straighten the stable and unstable directions associated with this periodic orbit. To that end, we introduce the change of variables

q~=xy2p~=x+y2.\tilde{q}=\frac{x-y}{2}\qquad\qquad\tilde{p}=\frac{x+y}{2}.

In these coordinates

(3.3) q~˙=14(q~+p~)3(q+𝒪3(q~,p~))p~˙=14(q~+p~)3(p~+𝒪3(q~,p~))\dot{\tilde{q}}=\frac{1}{4}(\tilde{q}+\tilde{p})^{3}(q+\mathcal{O}_{3}(\tilde{q},\tilde{p}))\qquad\qquad\qquad\dot{\tilde{p}}=-\frac{1}{4}(\tilde{q}+\tilde{p})^{3}(\tilde{p}+\mathcal{O}_{3}(\tilde{q},\tilde{p}))

so it is clear that the local stable and unstable invariant manifolds associated with the periodic orbit γ~={q~=p~=0,t𝕋}\tilde{\gamma}_{\infty}=\{\tilde{q}=\tilde{p}=0,t\in\mathbb{T}\}, which, by the work of McGehee [McG73] (see also Proposition 3.1) we already know exist, are close (for small |q~|,|p~||\tilde{q}|,|\tilde{p}|) to {q~=0}\{\tilde{q}=0\} and {p~=0}\{\tilde{p}=0\} respectively. Let now, for sufficiently small δ>0\delta>0, define the set

Qδ={(q~,p~,t)2×𝕋:|q~|δ,|p~|δ}.Q_{\delta}=\{(\tilde{q},\tilde{p},t)\in\mathbb{R}^{2}\times\mathbb{T}\colon|\tilde{q}|\leq\delta,\ |\tilde{p}|\leq\delta\}.

and, let (0,p~,t)Qδ(p~,γs(p~,t),t)Qδ(0,\tilde{p},t)\in Q_{\delta}\to(\tilde{p},\gamma^{s}(\tilde{p},t),t)\subset Q_{\delta} and (q~,0,t)Qδ(q~,γu(q~,t),t)Qδ(\tilde{q},0,t)\in Q_{\delta}\to(\tilde{q},\gamma^{u}(\tilde{q},t),t)\subset Q_{\delta} be graph parametrizations of these local invariant manifolds. Introduce new variables on QδQ_{\delta} given by

q=q~γs(p~,t)p=p~γu(q~,t).q=\tilde{q}-\gamma^{s}(\tilde{p},t)\qquad\qquad p=\tilde{p}-\gamma^{u}(\tilde{q},t).

From the invariance equation satisfied by γu,s\gamma^{u,s} one can deduce their Taylor expansion around q~=p~=0\tilde{q}=\tilde{p}=0. Then, an easy computation, shows that

(3.4) q˙=q4((q+p)3+𝒪4(q,p))p˙=p4((q+p)3+𝒪4(q,p))\dot{q}=-\frac{q}{4}\left((q+p)^{3}+\mathcal{O}_{4}(q,p)\right)\qquad\qquad\dot{p}=\frac{p}{4}\left((q+p)^{3}+\mathcal{O}_{4}(q,p)\right)

so in coordinates (q,p,t)Qδ(q,p,t)\subset Q_{\delta} the local stable and unstable manifolds are the sets {p=0}Qδ\{p=0\}\cap Q_{\delta} and {q=0}Qδ\{q=0\}\cap Q_{\delta} respectively. Define now, for a<δa<\delta the sections (see Figure 3.2)

Σa+={(q,p,t)Q2δ:p=δ, 0<qa}Σa={(q,p,t)Q2δ×𝕋:q=δ, 0<pa}\Sigma^{+}_{a}=\{(q,p,t)\in Q_{2\delta}\colon p=\delta,\ 0<q\leq a\}\qquad\qquad\Sigma^{-}_{a}=\{(q,p,t)\in Q_{2\delta}\times\mathbb{T}\colon q=\delta,\ 0<p\leq a\}

and the associated Poincaré map Φloc:Σa+Σa\Phi_{\mathrm{loc}}:\Sigma^{+}_{a}\to\Sigma^{-}_{a^{\prime}}, associated with the flow (3.3), whenever is well defined. Lemma 3.2 shows that a parabolic version of the Lambda Lemma holds for the degenerate periodic orbit {p=q=0}\{p=q=0\}. In order to build orbits whose final motions are hyperbolic, we also introduce the outer sections

Σa,hyp+={(q,p,t)Q2δ:p=δ,aq<0}Σa,hyp={(q,p,t)Q2δ:q=δ,ap<0}.\Sigma^{+}_{a,\mathrm{hyp}}=\{(q,p,t)\in Q_{2\delta}\colon p=\delta,\ -a\leq q<0\}\qquad\qquad\Sigma^{-}_{a,\mathrm{hyp}}=\{(q,p,t)\in Q_{2\delta}\colon q=\delta,\ -a\leq p<0\}.
Refer to caption
Figure 3.2. The sections Σa±\Sigma_{a}^{\pm}. The Poincaré map Φloc:Σa+Σa\Phi_{\mathrm{loc}}:\Sigma^{+}_{a}\to\Sigma^{-}_{a^{\prime}} sends the blue line in the section Σa+\Sigma_{a}^{+} into the blue line in the section Σa\Sigma_{a}^{-}, which accumulates to {p=0}\{p=0\}.

The proof of the following proposition follows plainly from the arguments in Chapter IV of [Mos01], where an analogous result is proved for the Sitnikov problem. See also Theorem 5.4. in [GMPS22].

Lemma 3.2.

Fix any G>0G_{*}>0. Then, there exists C>0C>0 sufficiently large and δ>0\delta>0 sufficiently small such that for any G[G,G]G\in[-G_{*},G_{*}] and any a(0,δ/2)a\in(0,\delta/2) the Poincaré map

Φloc:Σa+Σa1Cδ\Phi_{\mathrm{loc}}:\Sigma^{+}_{a}\longrightarrow\Sigma^{-}_{a^{1-C\delta}}

is well defined. Moreover, for any t1t_{1} sufficiently large there exist unique qq and p1p_{1}, which satisfy

q1+Cδp1q1Cδq3(1Cδ)/2t1q3(1+Cδ)/2,q^{1+C\delta}\leq p_{1}\leq q^{1-C\delta}\qquad\qquad\quad q^{-3(1-C\delta)/2}\lesssim t_{1}\lesssim q^{-3(1+C\delta)/2},

for which Φloc(q,a,0)=(a,p1,t1)\Phi_{\mathrm{loc}}(q,a,0)=(a,p_{1},t_{1}).

In addition, for any (q,a,0)Σa,hyp+(q,a,0)\in\Sigma^{+}_{a,\mathrm{hyp}} (respectively (a,p,0)Σa,hyp(a,p,0)\in\Sigma^{-}_{a,\mathrm{hyp}}), the orbit (qhyp(s),phyp(s),s)(q_{\mathrm{hyp}}(s),p_{\mathrm{hyp}}(s),s) of (3.4) with initial condition (q,a,0)(q,a,0) (respectively (a,p,0)(a,p,0)) is defined for all forward (respectively backward) times and satisfies

limsy(qhyp(s),(phyp(s))>0(respectivelylimsy(qhyp(s),(phyp(s))<0).\lim_{s\to\infty}y(q_{\mathrm{hyp}}(s),(p_{\mathrm{hyp}}(s))>0\qquad\qquad(\text{respectively}\ \lim_{s\to-\infty}y(q_{\mathrm{hyp}}(s),(p_{\mathrm{hyp}}(s))<0).

The first item in Lemma 3.2 shows that the iterates of curves which are transversal to the local stable manifold accumulate along the unstable manifold (see also Figure 3.2). The second item ensures that orbits with initial conditions on Σa,hyp+\Sigma^{+}_{a,\mathrm{hyp}} (respectively Σa,hyp\Sigma^{-}_{a,\mathrm{hyp}}) have forward (respectively backward) hyperbolic final motions. We now translate these results to the original coordinates. To that end we introduce the sections

ΛR,δ+={(r,y,t):r=R, 0<rS+(R,t;G)yδ,t𝕋}ΛR,δ={(r,y,t):r=R, 0<yrS(R,t;G)δ,t𝕋}\begin{split}\Lambda^{+}_{R,\delta}=&\{(r,y,t)\colon r=R,\ 0<\partial_{r}S^{+}(R,t;G)-y\leq\delta,\ t\in\mathbb{T}\}\\ \Lambda^{-}_{R,\delta}=&\{(r,y,t)\colon r=R,\ 0<y-\partial_{r}S^{-}(R,t;G)\leq\delta,\ t\in\mathbb{T}\}\end{split}

and the map Φloc,R1,R2:ΛR1,δ+ΛR2,δ\Phi_{\mathrm{loc},R_{1},R_{2}}:\Lambda^{+}_{R_{1},\delta}\to\Lambda^{-}_{R_{2},\delta^{\prime}} whenever is well defined. We also define the sections leading to hyperbolic final motions

ΛR,δ+={(r,y,t):r=R,δrS+(R,t;G)y<0,t𝕋}ΛR,δ={(r,y,t):r=R,δyrS(R,t;G)<0,t𝕋}.\begin{split}\Lambda^{+}_{R,\delta}=&\{(r,y,t)\colon r=R,\ -\delta\leq\partial_{r}S^{+}(R,t;G)-y<0,\ t\in\mathbb{T}\}\\ \Lambda^{-}_{R,\delta}=&\{(r,y,t)\colon r=R,\ -\delta\leq y-\partial_{r}S^{-}(R,t;G)<0,\ t\in\mathbb{T}\}.\end{split}
Lemma 3.3.

Fix any G>0G_{*}>0. Then, there exist R>0R>0 sufficiently large such that for any R1,R2RR_{1},R_{2}\geq R there exists δ0(R1,R2)\delta_{0}(R_{1},R_{2}) such that for all G[G,G]G\in[-G_{*},G_{*}] the Poincaré map

Φloc,R1,R2:ΛR1,δ+ΛR2,δ\Phi_{\mathrm{loc},R_{1},R_{2}}:\Lambda^{+}_{R_{1},\delta}\to\Lambda^{-}_{R_{2},\delta^{\prime}}

is well defined for δδ0\delta\leq\delta_{0} and some δ(R1,R2,δ)>0\delta^{\prime}(R_{1},R_{2},\delta)>0. There exists TT_{*} such that for any TTT\geq T_{*} there exist unique y0,y1y_{0},y_{1} such that Φloc,R1,R2(R1,y0,0)=(R2,y1,T)\Phi_{\mathrm{loc},R_{1},R_{2}}(R_{1},y_{0},0)=(R_{2},y_{1},T). Moreover, for any ε>0\varepsilon>0 there exists TT_{**} such that, if TTT\geq T_{**} and Φloc,R1,R2(R1,y0,0)=(R2,y1,T)\Phi_{\mathrm{loc},R_{1},R_{2}}(R_{1},y_{0},0)=(R_{2},y_{1},T), then

rS+(R1,0;G)y0εy1rS+(R2,T;G)ε.\partial_{r}S^{+}(R_{1},0;G)-y_{0}\leq\varepsilon\qquad\qquad y_{1}-\partial_{r}S^{+}(R_{2},T;G)\leq\varepsilon.

In addition, the orbit (rhyp(s),yhyp(s),s)(r_{\mathrm{hyp}}(s),y_{\mathrm{hyp}}(s),s) of (1.2) with initial condition (R1,y,0)ΛR1,δ,hyp+(R_{1},y,0)\in\Lambda^{+}_{R_{1},\delta,\mathrm{hyp}} (respectively (R2,y,0)ΛR2,δ,hyp(R_{2},y,0)\in\Lambda^{-}_{R_{2},\delta,\mathrm{hyp}}), is defined for all forward (respectively backward) times and satisfies

limsyhyp(s)>0(respectivelylimsyhyp(s)<0).\lim_{s\to\infty}y_{\mathrm{hyp}}(s)>0\qquad\qquad(\text{respectively}\ \lim_{s\to-\infty}y_{\mathrm{hyp}}(s)<0).

4. Existence of homoclinic orbits to γ\gamma_{\infty}

In this section we establish the existence of orbits of the Hamiltonian (1.1), which are homoclinic to γ\gamma_{\infty}. For |G|1|G|\gg 1, the Hamiltonian (1.1) can be considered as a perturbation of the integrable 2BP, in which there exists a homoclinic manifold to γ\gamma_{\infty} (see Lemma 2.1). Therefore, for |G|1|G|\gg 1, one can use geometric perturbation theory to prove that the global invariant manifolds W+(γ;G)W^{+}(\gamma_{\infty};G) and W(γ;G)W^{-}(\gamma_{\infty};G) defined in (3.2) intersect transverally. This was the approach used in [GPSV21] where the following result was proved.

Theorem 4.1 ([GPSV21]).

There exists G<G_{*}<\infty such that for all GG such that |G|G|G|\geq G_{*} the global stable and unstable manifolds W+(γ;G)W^{+}(\gamma_{\infty};G) and W(γ;G)W^{-}(\gamma_{\infty};G) defined in (3.2), intersect transversally.

Yet, for a fixed GG\in\mathbb{R}, the Hamiltonian (1.1) is not close to the 2BP. Therefore, geometric perturbation theory cannot help to study the existence of transversal intersections between W+(γ;G)W^{+}(\gamma_{\infty};G) and W(γ;G)W^{-}(\gamma_{\infty};G). We however exploit the variational formulation of the problem, in which the powerful techniques from nonlinear functional analysis are available.

More concretely, in Section 4.1 we introduce a suitable action functional, defined on a suitable Hilbert space, whose critical points are indeed orbits of (1.1) which are homoclinic to γ\gamma_{\infty}. Then, in Section 4.2 we establish the existence of a critical point of the aforementioned action functional using a minmax argument. The minmax characterization of the critical point obtained is crucial for the construction in Section 6.

4.1. The Variational Formulation

We introduce the vector space of real valued functions

(4.1) D1,2={φC():vφL2()such that φ(s)=φ(0)+0svφ(t)dts}.D^{1,2}=\{\varphi\in C(\mathbb{R})\colon\exists v_{\varphi}\in L^{2}(\mathbb{R})\ \text{such that }\varphi(s)=\varphi(0)+\int_{0}^{s}v_{\varphi}(t)\mathrm{d}t\ \ \forall s\in\mathbb{R}\}.

In the following, we will write φ˙=vφ\dot{\varphi}=v_{\varphi} (i.e. vφv_{\varphi} is the weak derivative of φ\varphi). It is easy to chek that

φ,ψD1,2=|φ(0)ψ(0)|+φ˙,ψ˙L2\langle\varphi,\psi\rangle_{D^{1,2}}=|\varphi(0)\psi(0)|+\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}}

defines an inner product on D1,2D^{1,2} for which the functional space D1,2D^{1,2} equiped with this inner product is a Hilbert space. We write

φD1,2=(φ,φD1,2)1/2.\lVert\varphi\rVert_{D^{1,2}}=\left(\langle\varphi,\varphi\rangle_{D^{1,2}}\right)^{1/2}.

for the induced norm. Notice that for all φD1,2\varphi\in D^{1,2} and all ss\in\mathbb{R}

|φ(s)||φ(0)|+φ˙L2|s|.|\varphi(s)|\leq|\varphi(0)|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|}.

After the introduction of the functional space D1,2D^{1,2} it is an easy computation to show that the existence of orbits of (1.2) homoclinic to the periodic orbit at infinity γ={r=,y=0,t𝕋}\gamma_{\infty}=\{r=\infty,y=0,t\in\mathbb{T}\} is equivalent to the existence of critical points of the action functional 𝒜G:D1/2\mathcal{A}_{G}:D^{1/2}\to\mathbb{R} given by

(4.2) 𝒜G(φ;G0)=ren(φ,φ˙,s;G,G0)ds,\mathcal{A}_{G}(\varphi;G_{0})=\int_{\mathbb{R}}\mathcal{L}_{\mathrm{ren}}(\varphi,\dot{\varphi},s;G,G_{0})\mathrm{d}s,

where

(4.3) ren(φ,φ˙,s;G,G0)=φ˙22+VG(r0+φ)V0(r0)r¨0φ,\mathcal{L}_{\mathrm{ren}}(\varphi,\dot{\varphi},s;G,G_{0})=\frac{\dot{\varphi}^{2}}{2}+V_{G}(r_{0}+\varphi)-V_{0}(r_{0})-\ddot{r}_{0}\varphi,

VGV_{G} stands for the effective potential

(4.4) VG(r,t)=G22r21r2+ρ2(t),V_{G}(r,t)=\frac{G^{2}}{2r^{2}}-\frac{1}{\sqrt{r^{2}+\rho^{2}(t)}},

and V0(r0)=G022r021r0V_{0}(r_{0})=\frac{G_{0}^{2}}{2r_{0}^{2}}-\frac{1}{r_{0}} with r0r_{0} being the parabolic orbit of the 2BP with angular momentum G0G_{0}\in\mathbb{R} (see Remark 5).

Remark 4.

ren(φ,φ˙,s;G)\mathcal{L}_{\mathrm{ren}}(\varphi,\dot{\varphi},s;G) is indeed a renormalized Lagrangian, that is, we have substracted the term V0(r0)V_{0}(r_{0}) in the integrand of what would be the “natural” action functional. The reason behind the definition of (4.2) is that the action of a parabolic orbit is infinite. Indeed, the Lagrangian of the 2BP reads

0(r0,r˙0)=r˙022V0(r0)\mathcal{L}_{0}(r_{0},\dot{r}_{0})=\frac{\dot{r}_{0}^{2}}{2}-V_{0}(r_{0})

and for a parabolic orbit r0(s)s2/3r_{0}(s)\sim s^{2/3} for s±s\to\pm\infty.

Remark 5.

It might seem surprising that when defining the renormalized Lagrangian ren\mathcal{L}_{\mathrm{ren}}, we let G0G_{0} be an independent parameter instead of taking G=G0G=G_{0}. The reason is that in this way, for a fixed G0G_{0}\in\mathbb{R} and fixed φD1,2\varphi\in D^{1,2} the function G𝒜G(φ)G\to\mathcal{A}_{G}(\varphi) is monotonely decreasing. This will allow us to use a monotonicity trick due to Struwe which is key to obtain uniform bounds for certain (Palais-Smale) sequences {φn}nD1,2\{\varphi_{n}\}_{n\in\mathbb{N}}\subset D^{1,2} for which d𝒜G(φn)0\mathrm{d}\mathcal{A}_{G}(\varphi_{n})\to 0 (see Section 4.2.1 and, in particular, 4.7). On the other hand, the asymptotic behavior of parabolic solutions as s±s\to\pm\infty becomes independent of the value of the angular momentum GG (see Lemma 2.1) so the definition of the renormalized Lagrangian ren\mathcal{L}_{\mathrm{ren}} makes sense for GG0G\neq G_{0}.

Remark 6.

Throughout the rest of the paper the value G0+G_{0}\in\mathbb{R}_{+} will be fixed. Thus, we omit the dependence of all quantities on G0G_{0}. Having fixed G0+G_{0}\in\mathbb{R}_{+}, we state results for G[G0,G0]G\in[-G_{0},G_{0}] (or full measure subsets of this set). This choice is completely arbitrary: the results proved below are certainly true if we replace [G0,G0][-G_{0},G_{0}] by any other bounded subset. However, since we have always the freedom to choose G0G_{0} as large as we want it is enough to state results for G[G0,G0]G\in[-G_{0},G_{0}].

The following observation will play an important role in our construction.

Lemma 4.2.

Let τ\tau\in\mathbb{Z} and define the translation operator

Tτ(φ)(s)=φ(s+τ)+r0(s+τ)r0(s).T_{\tau}(\varphi)(s)=\varphi(s+\tau)+r_{0}(s+\tau)-r_{0}(s).

Then, for all τ\tau\in\mathbb{Z}

𝒜G(Tτ(φ))=𝒜G(φ).\mathcal{A}_{G}(T_{\tau}(\varphi))=\mathcal{A}_{G}(\varphi).

We now state a technical lemma which will prove useful in later compactness arguments.

Lemma 4.3.

Let γ0\gamma\geq 0 and let Lγ2L^{2}_{\gamma} be the weighted L2L^{2} space with norm given by

φLγ2=(|φ|2r03+γ)1/2.\lVert\varphi\rVert_{L^{2}_{\gamma}}=\left(\int_{\mathbb{R}}\frac{|\varphi|^{2}}{r_{0}^{3+\gamma}}\right)^{1/2}.

Then, D1,2D^{1,2} is continuously embedded in Lγ2L^{2}_{\gamma} for γ0\gamma\geq 0 and compactly embedded in Lγ2L^{2}_{\gamma} for γ>0\gamma>0.

Proof.

The proof of the continuous embedding for γ0\gamma\geq 0 is obtained by the very same argument used in the proof of Proposition 3.2. in [BDFT21] taking into account that r0(s)s2/3r_{0}(s)\sim s^{2/3} for s±s\to\pm\infty and r0(s)G02/2sr_{0}(s)\geq G_{0}^{2}/2\ \forall s\in\mathbb{R}. We now prove that the embedding for γ>0\gamma>0 is compact. Take any bounded sequence {φn}nD1,2\{\varphi_{n}\}_{n\in\mathbb{N}}\subset D^{1,2} such that φn0\varphi_{n}\to 0 weakly in D1,2D^{1,2}. In particular φn(s)0\varphi_{n}(s)\to 0 pointwise for all ss\in\mathbb{R}. Since, for any φD1,2\varphi\in D^{1,2} and any ss\in\mathbb{R} we have

|φ(s)||φ(0)|+φ˙L2|s||\varphi(s)|\leq|\varphi(0)|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|}

we obtain that for all ss\in\mathbb{R}

|φn(s)|2r03+γ(s)φnD1,221+|s|1+γ.\frac{|\varphi_{n}(s)|^{2}}{r_{0}^{3+\gamma}(s)}\lesssim\frac{\lVert\varphi_{n}\rVert^{2}_{D^{1,2}}}{1+|s|^{1+\gamma}}.

Therefore, a direct application of the dominated convergence theorem shows that

limnφnLγ22=limn|φn|2r03+γ(s)=0.\lim_{n\to\infty}\lVert\varphi_{n}\rVert_{L_{\gamma}^{2}}^{2}=\lim_{n\to\infty}\int_{\mathbb{R}}\frac{|\varphi_{n}|^{2}}{r_{0}^{3+\gamma}(s)}=0.

We now show that 𝒜G\mathcal{A}_{G} is continuous and has a continuous differential on a suitable subset QD1,2Q\subset D^{1,2}.

Lemma 4.4.

Let K>0K>0 and m¯>0\underline{m}>0 be two fixed constants and define

Q={φD1,2:φD1,2K,minsr0(s)+φ(s)m¯}Q=\{\varphi\in D^{1,2}\colon\lVert\varphi\rVert_{D^{1,2}}\leq K,\ \min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\geq\underline{m}\}

Then, for any G[G0,G0]G\in[-G_{0},G_{0}] we have 𝒜GC1(int(Q),)\mathcal{A}_{G}\in C^{1}(\mathrm{int}(Q),\mathbb{R}).

Proof.

Let φ,ψQ\varphi,\psi\in Q and make use of the mean value theorem to write

(4.5) 𝒜G(φ)𝒜G(ψ)=12(φ˙+ψ˙)(φ˙ψ˙)+rVG(r0+ξ)(φψ)r¨0(φψ)\mathcal{A}_{G}(\varphi)-\mathcal{A}_{G}(\psi)=\int_{\mathbb{R}}\frac{1}{2}(\dot{\varphi}+\dot{\psi})(\dot{\varphi}-\dot{\psi})+\partial_{r}V_{G}(r_{0}+\xi)(\varphi-\psi)-\ddot{r}_{0}(\varphi-\psi)

where ξ=λφ+(1λ)ψ\xi=\lambda\varphi+(1-\lambda)\psi for some λ(s)[0,1]\lambda(s)\in[0,1]. On one hand,

|(φ˙+ψ˙)(φ˙ψ˙)|(|φ˙+ψ˙|2)1/2(|φ˙ψ˙|2)1/20\left|\int_{\mathbb{R}}(\dot{\varphi}+\dot{\psi})(\dot{\varphi}-\dot{\psi})\right|\leq\left(\int_{\mathbb{R}}|\dot{\varphi}+\dot{\psi}|^{2}\right)^{1/2}\left(\int_{\mathbb{R}}|\dot{\varphi}-\dot{\psi}|^{2}\right)^{1/2}\to 0

as φψD1,20\lVert\varphi-\psi\rVert_{D^{1,2}}\to 0. On the other hand, since for φ,ψD1,2\varphi,\psi\in D^{1,2}

minsr0(s)+ξ(s)=minsr0(s)+λφ(s)+(1λ)ψ(s)=minsλ(r0(s)+φ(s))+(1λ)(r0(s)+ψ(s))minsλm¯+(1λ)m¯=m¯>0\begin{split}\min_{s\in\mathbb{R}}\ r_{0}(s)+\xi(s)=&\min_{s\in\mathbb{R}}\ r_{0}(s)+\lambda\varphi(s)+(1-\lambda)\psi(s)=\min_{s\in\mathbb{R}}\ \lambda(r_{0}(s)+\varphi(s))+(1-\lambda)(r_{0}(s)+\psi(s))\\ \geq&\min_{s\in\mathbb{R}}\ \lambda\underline{m}+(1-\lambda)\underline{m}=\underline{m}>0\end{split}

and convergence in D1,2D^{1,2} implies uniform convergence in compact intervals, we have, taking into account the expression of VGV_{G} (4.4), that

(rVG(r0+ξ)r¨0)(φψ)0\left(\partial_{r}V_{G}(r_{0}+\xi)-\ddot{r}_{0}\right)(\varphi-\psi)\to 0

pointwise as φψD1,20\lVert\varphi-\psi\rVert_{D^{1,2}}\to 0. Moreover, for s±s\to\pm\infty

r0(s)+φ(s)r0(s)(|φ(0)|+φ˙L2|s|)s2/3r_{0}(s)+\varphi(s)\geq r_{0}(s)-\left(|\varphi(0)|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|}\right)\sim s^{2/3}

so, from the definition of VGV_{G} in (4.4), a straightforward computation shows that for r0r_{0}\to\infty

rVG(r0)r¨0r03.\partial_{r}V_{G}(r_{0})-\ddot{r}_{0}\sim r_{0}^{-3}.

Thus, using again that minsr0(s)+ξ(s)m¯>0\min_{s\in\mathbb{R}}\ r_{0}(s)+\xi(s)\geq\underline{m}>0, we obtain the existence of C>0C>0 depending only on KK and m¯\underline{m} such that for all ss\in\mathbb{R}

|rVG(r0(s)+ξ(s))r¨0(s)|Cr03(s)|\partial_{r}V_{G}(r_{0}(s)+\xi(s))-\ddot{r}_{0}(s)|\leq Cr_{0}^{-3}(s)

Therefore,

|(rVG(r0+ξ)r¨0)(φψ)|(|(rVG(r0+ξ)r¨0)|)1/2×(|(rVG(r0+ξ)r¨0)||φψ|2)1/2C|φψ|2r03=CφψL02,\begin{split}\left|\int_{\mathbb{R}}\left(\partial_{r}V_{G}(r_{0}+\xi)-\ddot{r}_{0}\right)(\varphi-\psi)\right|\leq&\left(\int_{\mathbb{R}}|\left(\partial_{r}V_{G}(r_{0}+\xi)-\ddot{r}_{0}\right)|\right)^{1/2}\\ &\times\left(\int_{\mathbb{R}}|\left(\partial_{r}V_{G}(r_{0}+\xi)-\ddot{r}_{0}\right)||\varphi-\psi|^{2}\right)^{1/2}\\ \leq&C\int_{\mathbb{R}}\frac{|\varphi-\psi|^{2}}{r_{0}^{3}}=C\lVert\varphi-\psi\rVert_{L^{2}_{0}},\end{split}

and the continuity of the map 𝒜G:QD1,2\mathcal{A}_{G}:Q\subset D^{1,2}\to\mathbb{R} is implied by Lemma 4.3. The proof that d𝒜G:QD1,2D1,2\mathrm{d}\mathcal{A}_{G}:Q\subset D^{1,2}\to D^{1,2} is a continuous map follows from similar arguments. ∎

Lemma 4.5.

Let K>0K>0 and m¯>0\underline{m}>0 be two fixed constants and let QD1,2Q\subset D^{1,2} be the subset defined in Lemma 4.4. Then, for any for any G[G0,G0]G\in[-G_{0},G_{0}], d𝒜G:int(Q)D1,2\mathrm{d}\mathcal{A}_{G}:\mathrm{int}(Q)\to D^{1,2} is a compact perturbation of the identity. In partiuclar, this implies that for any compact set FD1,2F\subset D^{1,2} the set Q(d𝒜G)1(F)Q\cap(\mathrm{d}\mathcal{A}_{G})^{-1}(F) is compact.

Proof.

We write

(4.6) d𝒜G(φ)[ψ]=φ˙,ψ˙L2(r0+φ((r0+φ)2+ρ2)3/21r02)ψ+(G2(r0+φ)3G0r03)ψ=φ˙,ψ˙L2+2φψr03(r0+φ((r0+φ)2+ρ2)3/21r02+2φr03)ψ+(G2(r0+φ)3G0r03)ψ=φ˙,ψ˙L2+2φψr03+P(φ)[ψ]\begin{split}\mathrm{d}\mathcal{A}_{G}(\varphi)[\psi]=&\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}}-\int_{\mathbb{R}}\left(\frac{r_{0}+\varphi}{((r_{0}+\varphi)^{2}+\rho^{2})^{3/2}}-\frac{1}{r_{0}^{2}}\right)\psi+\int_{\mathbb{R}}\left(\frac{G^{2}}{(r_{0}+\varphi)^{3}}-\frac{G_{0}}{r_{0}^{3}}\right)\psi\\ =&\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}}+2\int_{\mathbb{R}}\frac{\varphi\psi}{r_{0}^{3}}-\int_{\mathbb{R}}\left(\frac{r_{0}+\varphi}{((r_{0}+\varphi)^{2}+\rho^{2})^{3/2}}-\frac{1}{r_{0}^{2}}+\frac{2\varphi}{r_{0}^{3}}\right)\psi\\ &+\int_{\mathbb{R}}\left(\frac{G^{2}}{(r_{0}+\varphi)^{3}}-\frac{G_{0}}{r_{0}^{3}}\right)\psi\\ =&\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}}+2\int_{\mathbb{R}}\frac{\varphi\psi}{r_{0}^{3}}+P(\varphi)[\psi]\end{split}

where we have introduced the functional

P(φ)[ψ]=(G2(r0+φ)3G0r03)ψ(r0+φ((r0+φ)2+ρ2)3/21r02+2φr03)ψP(\varphi)[\psi]=\int_{\mathbb{R}}\left(\frac{G^{2}}{(r_{0}+\varphi)^{3}}-\frac{G_{0}}{r_{0}^{3}}\right)\psi-\int_{\mathbb{R}}\left(\frac{r_{0}+\varphi}{((r_{0}+\varphi)^{2}+\rho^{2})^{3/2}}-\frac{1}{r_{0}^{2}}+\frac{2\varphi}{r_{0}^{3}}\right)\psi

Thanks to Lemma 4.3 we can take

φ,ψD1,2=2φψr03+φ˙,ψ˙L2\langle\langle\varphi,\psi\rangle\rangle_{D^{1,2}}=2\int_{\mathbb{R}}\frac{\varphi\psi}{r_{0}^{3}}+\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}}

as an equivalent inner product in D1,2D^{1,2}. It follows from Lemma 4.4 that for all φQ\varphi\in Q, d𝒜G(φ):D1,2\mathrm{d}\mathcal{A}_{G}(\varphi):D^{1,2}\to\mathbb{R} and P(φ):D1,2P(\varphi):D^{1,2}\to\mathbb{R} are continuous linear functionals and thanks to Riesz representation theorem, for every φQD1,2\varphi\in Q\subset D^{1,2} there exist unique ηA(φ),ηP(φ)D1,2\eta_{A}(\varphi),\eta_{P}(\varphi)\in D^{1,2} such that

ηA(φ),ψD1,2=d𝒜G(φ)[ψ]ηP(φ),ψD1,2=P(φ)[ψ].\langle\langle\eta_{A}(\varphi),\psi\rangle\rangle_{D^{1,2}}=\mathrm{d}\mathcal{A}_{G}(\varphi)[\psi]\qquad\qquad\langle\langle\eta_{P}(\varphi),\psi\rangle\rangle_{D^{1,2}}=P(\varphi)[\psi].

and ηA=Id+ηP\eta_{A}=\mathrm{Id}+\eta_{P}. After writing

ηP(φ)ηP(φ),ηP(φ)ηP(φ)D1,2=P(φ)[ηP(φ)ηP(φ)]P(φ)[ηP(φ)ηPφ)],\langle\langle\eta_{P}(\varphi_{*})-\eta_{P}(\varphi),\eta_{P}(\varphi_{*})-\eta_{P}(\varphi)\rangle\rangle_{D^{1,2}}=P(\varphi_{*})[\eta_{P}(\varphi_{*})-\eta_{P}(\varphi)]-P(\varphi)[\eta_{P}(\varphi_{*})-\eta_{P}\varphi)],

a tedious but straightforward computation shows that for any φ,φQ\varphi_{*},\varphi\in Q

(4.7) ηP(φ)ηP(φ)D1,22φφL1/42ηP(φ)ηP(φ)L1/42φφL1/42ηP(φ)ηP(φ)D1,2\lVert\eta_{P}(\varphi_{*})-\eta_{P}(\varphi)\rVert_{D^{1,2}}^{2}\leq\lVert\varphi_{*}-\varphi\rVert_{L^{2}_{1/4}}\lVert\eta_{P}(\varphi^{*})-\eta_{P}(\varphi)\rVert_{L^{2}_{1/4}}\leq\lVert\varphi_{*}-\varphi\rVert_{L^{2}_{1/4}}\lVert\eta_{P}(\varphi^{*})-\eta_{P}(\varphi)\rVert_{D^{1,2}}

what implies that ηP:QD1,2\eta_{P}:Q\to D^{1,2} is a compact operator (recall that the embedding of D1,2D^{1,2} in L1/42L^{2}_{1/4} is compact). The second item in the lemma plainly follows after writing

ηA(φ)=φ+ηP(φ).\eta_{A}(\varphi)=\varphi+\eta_{P}(\varphi).

Indeed, for a sequence {φn}nQD1,2\{\varphi_{n}\}_{n\in\mathbb{N}}\subset Q\subset D^{1,2} whose image under d𝒜G\mathrm{d}\mathcal{A}_{G} is contained in a compact subset FD1,2F\subset D^{1,2} there exists a subsequence (which we do not relabel) for which {ηA(φn)}n\{\eta_{A}(\varphi_{n})\}_{n\in\mathbb{N}} is convergent in D1,2D^{1,2}. Then, the proof is finished since ηP\eta_{P} being a compact operator and implies that (up to passing to a further subsequence) {ηP(φn)}n\{\eta_{P}(\varphi_{n})\}_{n\in\mathbb{N}} is also convergent in D1,2D^{1,2}. ∎

From now on we will omit the subscript in the inner product and norm defined in D1,2D^{1,2}.

4.2. Existence of critical points of the action functional

In this section we prove the existence of critical points of the action functional 𝒜G\mathcal{A}_{G} defined in (4.2) using a minmax argument. In particular, we will employ a constrained version of the celebrated mountain pass theorem of Ambrosetti and Rabinowitz [AR73]. The first step is to verify that the level sets of 𝒜G\mathcal{A}_{G} have a mountain pass geometry. This is the content of the following proposition.

Proposition 4.6.

Take any constant M>0M>0. Then, for all G[G0,G0]{0}G\in[-G_{0},G_{0}]\setminus\{0\} there exist ψ1,ψ2D1,2\psi_{1},\psi_{2}\in D^{1,2} such that

𝒜G(ψi)Mi=1,2.\mathcal{A}_{G}(\psi_{i})\leq-M\qquad\qquad i=1,2.

Moreover, there exists M>0M^{*}>0 such that if we take MMM\geq M^{*}, then for any curve γC([0,1],D1,2)\gamma\in C([0,1],D^{1,2}) joining ψ1\psi_{1} and ψ2\psi_{2} there exist a point ψγ\psi_{\gamma} for which

𝒜G(ψγ)M/2.\mathcal{A}_{G}(\psi_{\gamma})\geq-M/2.
Proof.

Let μ>0\mu>0 so

𝒜G(μ)=VG(r0+μ)V0(r0)=1((r0+μ)2+ρ2)1/21r0G22(r0+μ)2+G022r02μ1r0+μ1r0.\begin{split}\mathcal{A}_{G}(\mu)=&\int_{\mathbb{R}}V_{G}(r_{0}+\mu)-V_{0}(r_{0})=\int_{\mathbb{R}}\frac{1}{((r_{0}+\mu)^{2}+\rho^{2})^{1/2}}-\frac{1}{r_{0}}-\frac{G^{2}}{2(r_{0}+\mu)^{2}}+\frac{G_{0}^{2}}{2r_{0}^{2}}\\ \leq&\int_{\mu}\frac{1}{r_{0}+\mu}-\frac{1}{r_{0}}.\end{split}

It follows from Fatou’s lemma that

lim supμ𝒜G(μ)=lim supμ1r0+μ1r01r0=.\limsup_{\mu\to\infty}\mathcal{A}_{G}(\mu)=\limsup_{\mu\to\infty}\int_{\mathbb{R}}\frac{1}{r_{0}+\mu}-\frac{1}{r_{0}}\leq-\int_{\mathbb{R}}\frac{1}{r_{0}}=-\infty.

On the other hand, take η(0,1/2)\eta\in(0,1/2). Then, for some finite (and uniform for η(0,1/2)\eta\in(0,1/2)) C>0C>0 we have

𝒜G(η)=VG(r0+η)V0(r0)=1((r0+η)2+ρ2)1/21r0+G022r02G22(r0+η)2C+011r0+ηG22(r0+η)2.\begin{split}\mathcal{A}_{G}(\eta)=&\int_{\mathbb{R}}V_{G}(r_{0}+\eta)-V_{0}(r_{0})=\int_{\mathbb{R}}\frac{1}{((r_{0}+\eta)^{2}+\rho^{2})^{1/2}}-\frac{1}{r_{0}}+\frac{G_{0}^{2}}{2r_{0}^{2}}-\frac{G^{2}}{2(r_{0}+\eta)^{2}}\\ \leq&C+\int_{0}^{1}\frac{1}{r_{0}+\eta}-\frac{G^{2}}{2(r_{0}+\eta)^{2}}.\\ \end{split}

Using that r0(s)=1/2+s2+𝒪(s3)r_{0}(s)=1/2+s^{2}+\mathcal{O}(s^{3}) for s0s\to 0 (this can be deduced from the proof of Lemma 2.1) one can easily check that

lim supη1/2𝒜G(η)=.\limsup_{\eta\to 1/2}\mathcal{A}_{G}(\eta)=-\infty.

The first part of the lemma is proven by taking ψ1=μ\psi_{1}=\mu with μ\mu large enough and ψ2=η\psi_{2}=\eta with η1/2\eta\to 1/2. In order to prove the second item of the lemma we let R>0R>0 be such that

rr2VG(r)0rR\partial^{2}_{rr}V_{G}(r)\geq 0\qquad\qquad\forall r\geq R

and denote by TT the value of ss for which r0(s)Rr_{0}(s)\geq R for all ss such that |s|T|s|\geq T. Notice that RR exists because of the convexity of VG(r)V_{G}(r) for large values of rr, which can be checked explicitely from the expression of VGV_{G} in (4.4). We now take φD1,2\varphi\in D^{1,2} such that minsr0(s)+φ(s)=R\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)=R. We claim that 𝒜G(φ)M/2\mathcal{A}_{G}(\varphi)\geq-M/2 so the lemma follows since, by continuity, for all γC([0,1],D1,2)\gamma\in C([0,1],D^{1,2}) joining ψ1\psi_{1} and ψ2\psi_{2} there exist a point φγ\varphi\in\gamma for which

minsr0(s)+φ(s)=R.\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)=R.

We now prove the claim. Lemma 4.2 implies that, withouth lost of generality, we can suppose that the minimum is attained at the interval s[0,1]s\in[0,1]. We express

𝒜G(φ)=φ˙L222+J(φ)+J(φ)+E(φ)\mathcal{A}_{G}(\varphi)=\frac{\lVert\dot{\varphi}\rVert_{L^{2}}^{2}}{2}+J_{\leq}(\varphi)+J_{\geq}(\varphi)+E(\varphi)

where

J(φ)=|s|T1((r0+φ)2+ρ2)1/21(r02+ρ2)1/2+r0φ(r02+ρ2)3/2G2|s|T12(r0+φ)212r02+φr03J(φ)=|s|T1((r0+φ)2+ρ2)1/21(r02+ρ2)1/2+r0φ(r02+ρ2)3/2G2|s|T12(r0+φ)212r02+φr03\begin{split}J_{\geq}(\varphi)=&\int_{|s|\geq T}\frac{1}{((r_{0}+\varphi)^{2}+\rho^{2})^{1/2}}-\frac{1}{(r_{0}^{2}+\rho^{2})^{1/2}}+\frac{r_{0}\varphi}{(r_{0}^{2}+\rho^{2})^{3/2}}\\ &-G^{2}\int_{|s|\leq T}\frac{1}{2(r_{0}+\varphi)^{2}}-\frac{1}{2r_{0}^{2}}+\frac{\varphi}{r_{0}^{3}}\\ J_{\leq}(\varphi)=&\int_{|s|\leq T}\frac{1}{((r_{0}+\varphi)^{2}+\rho^{2})^{1/2}}-\frac{1}{(r_{0}^{2}+\rho^{2})^{1/2}}+\frac{r_{0}\varphi}{(r_{0}^{2}+\rho^{2})^{3/2}}\\ &-G^{2}\int_{|s|\geq T}\frac{1}{2(r_{0}+\varphi)^{2}}-\frac{1}{2r_{0}^{2}}+\frac{\varphi}{r_{0}^{3}}\end{split}

and

E(φ)=s1(r02+ρ2)1/2r0φ(r02+ρ2)3/21r0+φr02+(G02G2)(12r02φr03)\begin{split}E(\varphi)=&\int_{s\in\mathbb{R}}\frac{1}{(r_{0}^{2}+\rho^{2})^{1/2}}-\frac{r_{0}\varphi}{(r_{0}^{2}+\rho^{2})^{3/2}}-\frac{1}{r_{0}}+\frac{\varphi}{r_{0}^{2}}+(G_{0}^{2}-G^{2})\left(\frac{1}{2r_{0}^{2}}-\frac{\varphi}{r_{0}^{3}}\right)\end{split}

For the first term, after applying the mean value theorem twice, we obtain that

J(φ)=|s|Trr2VG(r0+ξ)ηφJ_{\geq}(\varphi)=\int_{|s|\geq T}\partial^{2}_{rr}V_{G}(r_{0}+\xi)\eta\varphi

with η=σφ\eta=\sigma\varphi, 0σ10\leq\sigma\leq 1 and ξ=λη\xi=\lambda\eta, 0λ10\leq\lambda\leq 1. Since

min(r0+ξ)min(r0,r0+φ)R\min(r_{0}+\xi)\geq\min(r_{0},r_{0}+\varphi)\geq R

we have J(φ)0J_{\geq}(\varphi)\geq 0 by the definition of RR. For the second term we use that minsr0(s)+φ(s)R>0\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\geq R>0 and that for all ss\in\mathbb{R} we have |φ(s)||φ(0)|+φ˙L2|s||\varphi(s)|\leq|\varphi(0)|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|} so we obtain

J(φ)C+|s|Tr0φ(r02+ρ2)3/2G2|s|T12R212r02φr03C+|s|T(r0(r02+ρ2)3/2G02r03)φC(1+φ˙L2)\begin{split}J_{\leq}(\varphi)\geq&-C+\int_{|s|\leq T}\frac{r_{0}\varphi}{(r_{0}^{2}+\rho^{2})^{3/2}}-G^{2}\int_{|s|\leq T}\frac{1}{2R^{2}}-\frac{1}{2r_{0}^{2}}-\frac{\varphi}{r_{0}^{3}}\\ \geq&-C+\int_{|s|\leq T}\left(\frac{r_{0}}{(r_{0}^{2}+\rho^{2})^{3/2}}-\frac{G_{0}^{2}}{r_{0}^{3}}\right)\varphi\geq-C(1+\lVert\dot{\varphi}\rVert_{L^{2}})\end{split}

for some C>0C>0 which depends only on RR. An analogous computation shows that for the third term we have

E(φ)s1(r02+ρ2)1/21r0+(1r02(r0+ρ)(r02+ρ2)3/2+(G02G2)r03)φC(1+φ˙L2)E(\varphi)\geq\int_{s\in\mathbb{R}}\frac{1}{(r_{0}^{2}+\rho^{2})^{1/2}}-\frac{1}{r_{0}}+\left(\frac{1}{r_{0}^{2}}-\frac{(r_{0}+\rho)}{(r_{0}^{2}+\rho^{2})^{3/2}}+\frac{(G_{0}^{2}-G^{2})}{r_{0}^{3}}\right)\varphi\geq-C(1+\lVert\dot{\varphi}\rVert_{L^{2}})

for some C>0C>0 which depends only on RR. Therefore

𝒜G(φ)φ˙L222C(1+φ˙L2)\mathcal{A}_{G}(\varphi)\geq\frac{\lVert\dot{\varphi}\rVert^{2}_{L^{2}}}{2}-C(1+\lVert\dot{\varphi}\rVert_{L^{2}})

for CC depending only on RR and the result follows after enlarging MM (if necessary) while keeping RR fixed. ∎

We now have established the existence of the mountain pass geometry for the level sets of the functional 𝒜G\mathcal{A}_{G}. The next natural step would be to apply the classical deformation lemma to obtain a Palais-Smale (PS) sequence for the functional 𝒜G\mathcal{A}_{G}. There are however two difficulties. The first one is that, a priori, a suboptimal path, might contain points φD1,2\varphi\in D^{1,2} for which mins(r0+φ)(s)=0\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)=0, at which the functional φ𝒜G\varphi\mapsto\mathcal{A}_{G} is not continuous. The second difficulty is that, even if we can guarantee that mins(r0+φ)(s)>0\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)>0 for all φ\varphi in the region where we carry the deformation argument, without further constraints we are not able to show that the PS sequence obtained is precompact. For that reason, we take m¯>0\overline{m}>0 large enough and we carry the deformation argument in the region

(4.8) m¯={φD1,2:mins(r0+φ)(s)m¯}.\mathcal{F}_{\overline{m}}=\left\{\varphi\in D^{1,2}\colon\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)\leq\overline{m}\right\}.

In Lemma 4.11 we show that on a suitable subset m¯,δ,bm¯\mathcal{F}_{\overline{m},\delta,b}\subset\mathcal{F}_{\overline{m}}, the functional 𝒜G(φ)\mathcal{A}_{G}(\varphi) is bounded and coercive, from where we deduce a uniform bound for φ\lVert\varphi\rVert when φm¯,δ,b\varphi\in\mathcal{F}_{\overline{m},\delta,b}. This will be crucial to obtain uniformly bounded PS sequences.

4.2.1. The deformation argument

We now introduce the set of curves

(4.9) Γ={γC([0,1],D1,2):γ(0)=ψ1,γ(1)=ψ2}\varGamma=\left\{\gamma\in C([0,1],D^{1,2})\colon\gamma(0)=\psi_{1},\ \gamma(1)=\psi_{2}\right\}

and for m¯>0\overline{m}>0 large enough the candidate to critical value

(4.10) cG=infγΓmax{𝒜G(γ(t)):γ(t)m¯,t[0,1]}c_{G}=\inf_{\gamma\in\varGamma}\ \max\{\mathcal{A}_{G}(\gamma(t))\colon\gamma(t)\in\mathcal{F}_{\overline{m}},\ t\in[0,1]\}

The first step in the deformation argument is to prove that there exists a positive δ\delta such that for all bounded φ{φD1,2:|𝒜Gφ|δ}\varphi\in\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}-\varphi|\geq\delta\}, we have mins(r0+φ)(s)>0\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)>0. To that end we notice that

(4.11) 𝒜G(φ)=A(φ)G2B(φ)\mathcal{A}_{G}(\varphi)=A(\varphi)-G^{2}\ B(\varphi)

with

(4.12) A(φ)=φ˙22+1((r0+φ)2+ρ2)1/21r0+φr02+G02(12r02φr03)B(φ)=(r0+φ)2\begin{split}A(\varphi)=&\int_{\mathbb{R}}\frac{\dot{\varphi}^{2}}{2}+\frac{1}{((r_{0}+\varphi)^{2}+\rho^{2})^{1/2}}-\frac{1}{r_{0}}+\frac{\varphi}{r_{0}^{2}}+G_{0}^{2}\left(\frac{1}{2r_{0}^{2}}-\frac{\varphi}{r_{0}^{3}}\right)\\ B(\varphi)=&\int_{\mathbb{R}}(r_{0}+\varphi)^{-2}\end{split}

and apply a monotonicity trick due to Struwe (see [Str88] and [Str00]) to show that for almost every GG, the functional B(φ)B(\varphi) is bounded if |𝒜G(φ)cG||\mathcal{A}_{G}(\varphi)-c_{G}| is small enough (see Remark 7). The following version of the monotonicity trick was proved in [Jea99]. We provide the proof for the sake of self completeness.

Lemma 4.7.

There exists a full measure subset J[G0,G0]J\subset[-G_{0},G_{0}] such that for all GJG\in J there exists constants δ>0\delta>0 and C>0C>0 for which if |𝒜G(φ)cG|δ|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta then B(φ)CB(\varphi)\leq C.

Proof.

Since B(φ)0B(\varphi)\geq 0 it follows from expression (4.11) and the definition of cGc_{G} in (4.10) that GcGG\mapsto c_{G} is a monotone decreasing function. Therefore, it is differentiable on a subset JJ\subset\mathbb{R} whose complement has zero measure. Let GJG^{*}\in J, δ>0\delta>0 and take φ\varphi such that |IG(φ)cG|δ|I_{G^{*}}(\varphi)-c_{G^{*}}|\leq\delta. Take now G<GG<G^{*}, then, by decreasing (if necessary) the value of δ\delta we can assume that

𝒜G(φ)cG(GG)𝒜G(φ)cG+(GG)\mathcal{A}_{G}(\varphi)\geq c_{G^{*}}-(G^{*}-G)\qquad\qquad\mathcal{A}_{G^{*}}(\varphi)\leq c_{G^{*}}+(G^{*}-G)

Then

B(φ)=𝒜G(φ)𝒜G(φ)GGcG+(GG)cG+(GG)GGB(\varphi)=\frac{\mathcal{A}_{G^{*}}(\varphi)-\mathcal{A}_{G}(\varphi)}{G^{*}-G}\leq\frac{c_{G}+(G^{*}-G)-c_{G}^{*}+(G^{*}-G)}{G^{*}-G}

By the hypothesis on GG^{*} there exists an open neighbourhood around GG^{*} for which

cG1cGcGGGcG+1-c^{\prime}_{G^{*}}-1\leq\frac{c_{G}-c_{G^{*}}}{G^{*}-G}\leq-c^{\prime}_{G^{*}}+1

and the lemma is proven. ∎

Boundedness of the functional B(φ)B(\varphi) allows us to obtain an a priori estimate for mins(r0+φ)(s)\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s) if φD1,2\varphi\in D^{1,2} is bounded.

Lemma 4.8.

Let φD1,2\varphi\in D^{1,2} be such that B(φ)CB(\varphi)\leq C. Then, there exists a constant m¯>0\underline{m}>0, depending only on φ˙L2\lVert\dot{\varphi}\rVert_{L^{2}} such that

r0(s)+φ(s)m¯s.r_{0}(s)+\varphi(s)\geq\underline{m}\qquad\qquad\forall s\in\mathbb{R}.
Proof.

Suppose there exists ss_{*}\in\mathbb{R} such that limssr0(s)+φ(s)=0\lim_{s\to s_{*}}r_{0}(s)+\varphi(s)=0. Since φD1,2\varphi\in D^{1,2} it holds that |s|<|s_{*}|<\infty and we can assume without loss of generalitiy that r0(s)+φ(s)>0r_{0}(s)+\varphi(s)>0 for all s<ss<s_{*}. Take now s0=s1s_{0}=s_{*}-1 and write r(s)=r0(s)+φ(s)r(s)=r_{0}(s)+\varphi(s). Then, by the fundamental theorem of calculus, for any s[s0,s)s\in[s_{0},s_{*})

ln(r(s))ln(r(s0))=r(s0)r(s)r1dr=s0sr1(t)r˙(t)dt\ln(r(s))-\ln(r(s_{0}))=\int_{r(s_{0})}^{r(s)}r^{-1}\mathrm{d}r=\int_{s_{0}}^{s}r^{-1}(t)\dot{r}(t)\mathrm{d}t

and Hölder’s inequality implies

|ln(r(s))ln(r(s0))|B(φ)(s0s(r˙0+φ˙)2)1/2C(1+φ˙L2).|\ln(r(s))-\ln(r(s_{0}))|\leq B(\varphi)\left(\int_{s_{0}}^{s}(\dot{r}_{0}+\dot{\varphi})^{2}\right)^{1/2}\leq C\left(1+\lVert\dot{\varphi}\rVert_{L^{2}}\right).

Remark 7.

In Lemma 5.1, we show that for all G{0}G\in\mathbb{R}\setminus\{0\}, we have mins(r0+φ)(s)G2/2\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)\geq G^{2}/2 for orbits of (1.2) which are homoclinic to γ\gamma_{\infty}. However, that argument does not allow us to conclude that there exist δ>0\delta>0 such that for all bounded φ{φD1,2:|𝒜Gφ|δ}\varphi\in\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}-\varphi|\geq\delta\} we have mins(r0+φ)(s)>0\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)>0. Therefore, is not clear how to incorporate the a priori estimate in Lemma 5.1 to obtain a minmax critical point.

Assume now that φ\varphi is such that |𝒜G(φ)cG|δ|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta. Therefore, thanks to Lemmas 4.7 and 4.8 it is possible to obtain an inequality of the form

(4.13) 𝒜G(φ)φ˙L22Cφ\mathcal{A}_{G}(\varphi)\geq\frac{\lVert\dot{\varphi}\rVert_{L^{2}}}{2}-C\lVert\varphi\rVert

for some C>0C>0. Thus, if we moreover assume that φm¯\varphi\in\mathcal{F}_{\overline{m}} and that infs(r0+φ)(s)\inf_{s\in\mathbb{R}}(r_{0}+\varphi)(s) happens for s[0,1]s\in[0,1] we can obtain a uniform bound for the D1,2D^{1,2} norm of φ\varphi. In general, in problems in which the action functional is invariant under integer time translations, the latter assumption introduces no loss of generality and this argument can be employed to obtain uniformly bounded PS sequences.

However, in the present problem, the translation operator Tτ(φ)=φ(s+τ)+r0(s+τ)r0(s)T_{\tau}(\varphi)=\varphi(s+\tau)+r_{0}(s+\tau)-r_{0}(s), for which we have 𝒜G(Tτ(φ))=𝒜G(φ)\mathcal{A}_{G}(T_{\tau}(\varphi))=\mathcal{A}_{G}(\varphi), is not an isometry in D1,2D^{1,2}. This introduces certain technicalities in the deformation argument. In order to overcome this technical annoyance we introduce the following definition.

Definition 4.9.

Given φD1,2\varphi\in D^{1,2} we define its barycenter as the functional Bar:D1,2\mathrm{Bar}:D^{1,2}\to\mathbb{R} given by

Bar(φ)=((1+(r0+φ)2)2ds)1s(1+(r0+φ)2)2ds.\mathrm{Bar}(\varphi)=\left(\int_{\mathbb{R}}(1+(r_{0}+\varphi)^{2})^{-2}\mathrm{d}s\right)^{-1}\int_{\mathbb{R}}s(1+(r_{0}+\varphi)^{2})^{-2}\mathrm{d}s.

The following properties of the barycenter functional will be crucial for the deformation argument.

Lemma 4.10.

Let Bar(φ)\mathrm{Bar}(\varphi) be the functional introduced in Definition 4.9. The following statements hold:

  • Behaviour under translations: For any τ\tau\in\mathbb{Z}

    B(Tτφ)=B(φ)τ.B(T_{\tau}\varphi)=B(\varphi)-\tau.

    where the translation operator TτT_{\tau} was introduced in Lemma 4.2.

  • Local Lipschitzianity: For any K>0K>0 there exists LBar>0L_{\mathrm{Bar}}>0 such that

    supφK,φK|Bar(φ)Bar(φ)|φφLBar.\sup_{\lVert\varphi\rVert\leq K,\lVert\varphi^{\prime}\rVert\leq K}\frac{|\mathrm{Bar}(\varphi)-\mathrm{Bar}(\varphi^{\prime})|}{\lVert\varphi-\varphi^{\prime}\rVert}\leq L_{\mathrm{Bar}}.
Proof.

The proof of the first part is a trivial computation. For the second one we express

B(φ)=B2(φ)/B1(φ)B(\varphi)=B_{2}(\varphi)/B_{1}(\varphi)

with

B1(φ)=(1+(r0+φ)2)2dsB2(φ)=s(1+(r0+φ)2)2ds.B_{1}(\varphi)=\int_{\mathbb{R}}(1+(r_{0}+\varphi)^{2})^{-2}\mathrm{d}s\qquad\qquad B_{2}(\varphi)=\int_{\mathbb{R}}s(1+(r_{0}+\varphi)^{2})^{-2}\mathrm{d}s.

First we notice that there exists C>0C>0 such that for all φK\lVert\varphi\|\leq K we have B1(φ)C>0B_{1}(\varphi)\geq C>0 and |B2(φ)|C|B_{2}(\varphi)|\leq C. Indeed, for all ss\in\mathbb{R}

|φ(s)||φ(0)|+φ˙L2|s|(1+|s|)φC(1+|s|)|\varphi(s)|\leq|\varphi(0)|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|}\leq(1+\sqrt{|s|})\lVert\varphi\rVert\leq C(1+\sqrt{|s|})

so there exists T>0T>0, depending only on φ\lVert\varphi\rVert, such that

r0(s)+φ(s)r0(s)(1𝒪(s1/6))r_{0}(s)+\varphi(s)\geq r_{0}(s)(1-\mathcal{O}(s^{-1/6}))

for all |s|T|s|\geq T. Therefore, for CC depending only on TT,

|B2(φ)|C+|s|T|s|(1+(r0+φ)2)2dsC(1+|s|T|s|r04(s)ds)C.|B_{2}(\varphi)|\leq C+\int_{|s|\geq T}|s|(1+(r_{0}+\varphi)^{2})^{-2}\mathrm{d}s\leq C\left(1+\int_{|s|\geq T}|s|r_{0}^{-4}(s)\mathrm{d}s\right)\leq C.

The uniform bound from below for B1(φ)B_{1}(\varphi) follows since there exists CC and TT, depending only on φ\lVert\varphi\rVert, such that

r0(s)+φ(s)C(1+|s|)r_{0}(s)+\varphi(s)\leq C(1+\sqrt{|s|})

for all |s|T|s|\leq T. Take now φ,φD1,2\varphi,\varphi^{*}\in D^{1,2} and write

B(φ)B(φ)=(B2(φ)B2(φ))/B1(φ)+(B1(φ)B1(φ))B2(φ)/B1(φ)B1(φ)B(\varphi^{*})-B(\varphi)=(B_{2}(\varphi^{*})-B_{2}(\varphi))/B_{1}(\varphi)+(B_{1}(\varphi^{*})-B_{1}(\varphi))B_{2}(\varphi)/B_{1}(\varphi^{*})B_{1}(\varphi)

Let g(φ)=(1+(r0+φ)2)2g(\varphi)=(1+(r_{0}+\varphi)^{2})^{-2}. Then for φ,φD1,2\varphi,\varphi^{*}\in D^{1,2} we can write

B2(φ)B2(φ)=s(φφ)01g(λ(φφ))dλdsB_{2}(\varphi^{*})-B_{2}(\varphi)=\int_{\mathbb{R}}s(\varphi^{*}-\varphi)\int_{0}^{1}g^{\prime}(\lambda(\varphi^{*}-\varphi))\mathrm{d}\lambda\mathrm{d}s

On one hand for all ss\in\mathbb{R}

|φ(s)φ(s)||φ(0)φ(0)|+φ˙φ˙L2|s|(1+|s|)φφ|\varphi^{*}(s)-\varphi(s)|\leq|\varphi^{*}(0)-\varphi(0)|+\lVert\dot{\varphi}^{*}-\dot{\varphi}\rVert_{L^{2}}\sqrt{|s|}\leq(1+\sqrt{|s|})\lVert\varphi^{*}-\varphi\rVert

and it follows that

|B2(φ)B2(φ)|Cφφ(1+|s|3/2)01|r0+λ(φφ)|(1+|r0+λφφ|2)3dλdsCφφ(1+|s|3/2)r05Cφφ.\begin{split}|B_{2}(\varphi^{*})-B_{2}(\varphi)|\leq&C\lVert\varphi^{*}-\varphi\rVert\int_{\mathbb{R}}(1+|s|^{3/2})\int_{0}^{1}|r_{0}+\lambda(\varphi^{*}-\varphi)|(1+|r_{0}+\lambda\varphi^{*}-\varphi|^{2})^{-3}\mathrm{d}\lambda\mathrm{d}s\\ \leq&C\lVert\varphi^{*}-\varphi\rVert\int_{\mathbb{R}}(1+|s|^{3/2})r_{0}^{-5}\leq C\lVert\varphi^{*}-\varphi\rVert.\end{split}

The same computation shows that there exists CC such that

|B1(φ)B1(φ)|Cφφ|B_{1}(\varphi^{*})-B_{1}(\varphi)|\leq C\lVert\varphi^{*}-\varphi\rVert

and the lemma is proven. ∎

Together, Lemma 4.7 and Lemma 4.10 imply the following result, which is key in our constrained deformation argument.

Lemma 4.11.

Let J[G0,G0]J\subset[-G_{0},G_{0}]\subset\mathbb{R} be the subset obtained in Lemma 4.7. Let GJG\in J, let δ>0\delta>0 be the constant in Lemma 4.7, let b>0b>0 and define

(4.14) m¯,δ,b=m¯{φD1,2:|𝒜G(φ)cG|δ,|Bar(φ)|b}\mathcal{F}_{\overline{m},\delta,b}=\mathcal{F}_{\overline{m}}\cap\left\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta,\ \ |\mathrm{Bar}(\varphi)|\leq b\right\}

Then, there exists K>0K>0 such that

supφm¯,δ,bφK\sup_{\varphi\in\mathcal{F}_{\overline{m},\delta,b}}\lVert\varphi\rVert\leq K
Proof.

Let

𝒮={s¯:mins(r0+φ)(s)=(r0+φ)(s¯)}\mathcal{S}=\{\bar{s}\in\mathbb{R}\colon\min_{s\in\mathbb{R}}(r_{0}+\varphi)(s)=(r_{0}+\varphi)(\bar{s})\}

We claim that under the hypothesis of the lemma there exists C>0C>0 such that

|Bar(φ)s¯|Cs¯𝒮|\mathrm{Bar}(\varphi)-\bar{s}|\leq C\qquad\qquad\forall\bar{s}\in\mathcal{S}

Suppose not, then, by continuity, there exist s¯𝒮\bar{s}\in\mathcal{S} and sequences {φn}\{\varphi_{n}\}, {s¯n}\{\bar{s}_{n}\} such that φnφ\varphi_{n}\to\varphi in D1,2D^{1,2}, s¯ns¯\bar{s}_{n}\to\bar{s} and

φnm¯,|𝒜G(φn)cG|δ,and|Bar(φn)s¯n|.\varphi_{n}\in\mathcal{F}_{\overline{m}},\qquad\qquad|\mathcal{A}_{G}(\varphi_{n})-c_{G}|\leq\delta,\qquad\qquad\text{and}\qquad\qquad|\mathrm{Bar}(\varphi_{n})-\bar{s}_{n}|\to\infty.

By invariance of the action functional under translation Tτ(φ)(s)=φ(s+τ)+r0(s+τ)r0(s)T_{\tau}(\varphi)(s)=\varphi(s+\tau)+r_{0}(s+\tau)-r_{0}(s) (see Lemma 4.2) and the fact that

B(Tτ(φ))=B(φ)τB(T_{\tau}(\varphi))=B(\varphi)-\tau

the sequence φ~n=T[s¯n]φn\tilde{\varphi}_{n}=T_{[\bar{s}_{n}]}\varphi_{n}, satisfies

φ~nm¯,|𝒜G(φ~n)cG|δand|Bar(φ~n)|.\tilde{\varphi}_{n}\in\mathcal{F}_{\overline{m}},\qquad\qquad|\mathcal{A}_{G}(\tilde{\varphi}_{n})-c_{G}|\leq\delta\qquad\qquad\text{and}\qquad\qquad|\mathrm{Bar}(\tilde{\varphi}_{n})|\to\infty.

However, the first two properties imply the existence of KK such φ~nK\lVert\tilde{\varphi}_{n}\rVert\leq K for all nn\in\mathbb{N}. Indeed, by the construction of φ~n\tilde{\varphi}_{n}, for all φ~n\tilde{\varphi}_{n} we have

|φ(s)|m¯+φ~˙nL2|s+1||\varphi(s)|\leq\overline{m}+\lVert\dot{\tilde{\varphi}}_{n}\rVert_{L^{2}}\sqrt{|s+1|}

Moreover, since GJG\in J and |𝒜G(φ~n)cG|δ|\mathcal{A}_{G}(\tilde{\varphi}_{n})-c_{G}|\leq\delta , Lemma 4.7 implies that there exists C>0C>0 such that B(φ~n)CB(\tilde{\varphi}_{n})\leq C for all nn\in\mathbb{N}. Therefore, it is easy to check that there exists C>0C>0 such that

𝒜G(φ~n)φ~˙nL222C(1+φ~˙nL2)\mathcal{A}_{G}(\tilde{\varphi}_{n})\geq\frac{\lVert\dot{\tilde{\varphi}}_{n}\rVert^{2}_{L^{2}}}{2}-C(1+\lVert\dot{\tilde{\varphi}}_{n}\rVert_{L^{2}})

for all nn\in\mathbb{N}. Thus, since |𝒜G(φ~n)cG|δ|\mathcal{A}_{G}(\tilde{\varphi}_{n})-c_{G}|\leq\delta and |cG||c_{G}| is bounded, the sequence {φ~n}\{\tilde{\varphi}_{n}\} must be uniformly bounded. It is easy to check that the existence of KK such that φ~nK\lVert\tilde{\varphi}_{n}\rVert\leq K is in contradition with Bar(φ~n)\mathrm{Bar}(\tilde{\varphi}_{n}) being unbounded.

Once we know that the claim |Bar(φ)s¯|C|\mathrm{Bar}(\varphi)-\bar{s}|\leq C holds, we obtain that s¯C+b\bar{s}\leq C+b for all s¯𝒮\bar{s}\in\mathcal{S}. Therefore, since φm¯\varphi\in\mathcal{F}_{\overline{m}}, we have that

|φ(0)||φ(s¯)|+φ˙L2|s¯|C(1+φ˙L2)|\varphi(0)|\leq|\varphi(\bar{s})|+\lVert\dot{\varphi}\rVert_{L^{2}}\sqrt{|\bar{s}|}\leq C(1+\lVert\dot{\varphi}\rVert_{L^{2}})

for some C>0C>0 depending only on m¯,δ\overline{m},\delta and bb. The result follows since now we can show that for all φn\varphi_{n} we have

𝒜G(φn)φ˙nL222C(1+φ˙nL2)\mathcal{A}_{G}(\varphi_{n})\geq\frac{\lVert\dot{\varphi}_{n}\rVert^{2}_{L^{2}}}{2}-C(1+\lVert\dot{\varphi}_{n}\rVert_{L^{2}})

for some CC uniform on nn. ∎

In Proposition 4.13 we show the existence of a PS sequence contained in m¯,δ,b\mathcal{F}_{\overline{m},\delta,b} for large enough values of m¯\overline{m} and bb. Notice that in particular, thanks to Lemma 4.11 this sequence will be uniformly bounded.

We split the proof of Proposition 4.13 in two parts. First, we assume by contradiction that there exists no critical point of the action functional 𝒜G\mathcal{A}_{G} in m¯,δ,b\mathcal{F}_{\overline{m},\delta,b}. Under this assumption, we build a pseudo gradient vector field for 𝒜G\mathcal{A}_{G}, this is the content of Proposition 4.12. Then, in Proposition 4.13 we use this pseudo gradient vector field to build a localized deformation which yields points φm¯,δ,b\varphi\in\mathcal{F}_{\overline{m},\delta,b} for which 𝒜G(φ)<cG\mathcal{A}_{G}(\varphi)<c_{G}, a contradiction.

Before stating Proposition 4.12 some definitions are in order. Let b>0b>0 and 0<ε<δ/20<\varepsilon<\delta/2 where δ>0\delta>0 is the constant in Lemma 4.7. For we want the flow along the pseudogradient vector field to leave D1,2m¯D^{1,2}\setminus\mathcal{F}_{\overline{m}} positively invariant, we express it as the convex combination of two localized vector fields: a gradient-like vector field supported on

(4.15) P={φD1,2:|𝒜G(φ)cG|ε,minsr0(s)+φ(s)m¯,Bar(φ)2b}P=\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\varepsilon,\ \min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\leq\overline{m},\ \mathrm{Bar}(\varphi)\leq 2b\}

and a vector field supported on

(4.16) Y={φD1,2:|𝒜G(φ)cG|δ,|minsr0(s)+φ(s)m¯|ε,Bar(φ)2b}.Y=\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta,\ |\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)-\overline{m}|\leq\varepsilon,\ \mathrm{Bar}(\varphi)\leq 2b\}.

for which D1,2m¯D^{1,2}\setminus\mathcal{F}_{\overline{m}} is positively invariant. This construction is made explicit in the following proposition.

Proposition 4.12.

Let J[G0,G0]J\subset[G_{0},G_{0}]\mathbb{R} be the subset obtained in Lemma 4.7. Let δ>0\delta>0 be the constant in Lemma 4.7 and take m¯>0\overline{m}>0 large enough. Assume that for all b>0b>0 there exists α>0\alpha>0 for which

inf{𝒜G(φ):φm¯,δ,2b}α\inf\left\{\lVert\nabla\mathcal{A}_{G}(\varphi)\rVert\colon\varphi\in\mathcal{F}_{\overline{m},\delta,2b}\right\}\geq\alpha

Then, there exists b0b_{0} such that for all bb0b\geq b_{0} there exists a Lipschitz pseudogradient vector field W:D1,2D1,2W:D^{1,2}\to D^{1,2} such that

  • W1\lVert W\rVert\leq 1,

  • There exists a constant β>0\beta>0 such that

    d𝒜G(φ)[W(φ)]βφPY,\mathrm{d}\mathcal{A}_{G}(\varphi)\left[W(\varphi)\right]\leq-\beta\qquad\qquad\forall\varphi\in P\cup Y,
  • The region D1,2m¯D^{1,2}\setminus\mathcal{F}_{\overline{m}} is positively invariant under the flow of WW.

Proof.

Let ε<δ/2\varepsilon<\delta/2, define the sets

Y={φD1,2:|𝒜G(φ)cG|δ,|minsr0(s)+φ(s)m¯|ε,Bar(φ)2b}Z={φD1,2:|𝒜G(φ)cG|δ,|minsr0(s)+φ(s)m¯|2ε,Bar(φ)2b}\begin{split}Y=&\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta,\ |\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)-\overline{m}|\leq\varepsilon,\ \mathrm{Bar}(\varphi)\leq 2b\}\\ Z=&\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\delta,\ |\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)-\overline{m}|\geq 2\varepsilon,\ \mathrm{Bar}(\varphi)\leq 2b\}\end{split}

and the function

Ψ=distZdistY+distZ,\Psi=\frac{\mathrm{dist}Z}{\mathrm{dist}Y+\mathrm{dist}Z},

which satisfies Ψ=0\Psi=0 on ZZ and Ψ=1\Psi=1 on YY. We also introduce

P={φD1,2:|𝒜G(φ)cG|ε,minsr0(s)+φ(s)m¯,Bar(φ)2b}Q={φD1,2:|𝒜G(φ)cG|δ,minsr0(s)+φ(s)m¯,Bar(φ)2b}\begin{split}P=&\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\leq\varepsilon,\ \min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\leq\overline{m},\ \mathrm{Bar}(\varphi)\leq 2b\}\\ Q=&\{\varphi\in D^{1,2}\colon|\mathcal{A}_{G}(\varphi)-c_{G}|\geq\delta,\ \min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\leq\overline{m},\ \mathrm{Bar}(\varphi)\leq 2b\}\\ \end{split}

and define the function

Φ=distQdistP+distQ.\Phi=\frac{\mathrm{dist}Q}{\mathrm{dist}P+\mathrm{dist}Q}.

which satisfies Φ=0\Phi=0 on QQ and Φ=1\Phi=1 on PP. Take now a sufficiently small open neighbourhood UD1,2U\subset D^{1,2}, supp(Φ)D1,2\mathrm{supp}(\Phi)\subset D^{1,2}. Notice that by Lemma 4.11, there exists K>0K>0 such that

sup{φ:φU}K.\sup\{\lVert\varphi\rVert\colon\varphi\in U\}\leq K.

Then, since G𝒥G\in\mathcal{J}, Lemmas 4.7 and 4.8, impy that there exists m¯>0\underline{m}>0 such that

inf{minsr0(s)+φ(s):φU}m¯.\inf\{\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\colon\varphi\in U\}\geq\underline{m}.

Therefore, by Lemma 4.4 we have that d𝒜GC1(U,D1,2)\mathrm{d}\mathcal{A}_{G}\in C^{1}(U,D^{1,2}), what implies the existence of a constant C>0C>0 such that distP+distQ>C\mathrm{dist}P+\mathrm{dist}Q>C. We introduce now the pseudogradient vector field

(4.17) W=12((1Ψ)Φ𝒜G𝒜G+Ψv)W=\frac{1}{\sqrt{2}}(-(1-\Psi)\Phi\frac{\nabla\mathcal{A}_{G}}{\lVert\nabla\mathcal{A}_{G}\rVert}+\Psi v)

where vv is the constant vector field given by the constant v=1D1,2v=1\in D^{1,2}. Notice that for a large enough fixed m¯\overline{m}, and for all b>0b>0 there exists α~>0\tilde{\alpha}>0 such that

(4.18) sup{d𝒜G(φ)[v]:φsuppΨ}α~\sup\{\mathrm{d}\mathcal{A}_{G}(\varphi)[v]\colon\varphi\in\mathrm{supp}\Psi\}\leq-\tilde{\alpha}

Indeed

d𝒜G(φ)[v]=(G2(r0+φ)3r0+φ((r0+φ)2+ρ2)3/2)v\mathrm{d}\mathcal{A}_{G}(\varphi)[v]=\int_{\mathbb{R}}\left(\frac{G^{2}}{(r_{0}+\varphi)^{3}}-\frac{r_{0}+\varphi}{((r_{0}+\varphi)^{2}+\rho^{2})^{3/2}}\right)v

and the claim follows since for large enough m¯\overline{m} the integrand is non positive and moreover it is strictly negative on a positive measure subset of the real line since (thanks to Lemma 4.11) there exists K>0K>0 such that supφsuppΨφK\sup_{\varphi\in\mathrm{supp}\Psi}\lVert\varphi\rVert\leq K. It is straightforward to chek that the pseudogradient vector field WW introduced in (4.17) satisfies the properties listed in the statement in the lemma with β=min{α,α~}>0\beta=\min\{\alpha,\tilde{\alpha}\}>0.

Proposition 4.13.

Let J[G0,G0]J\subset[-G_{0},G_{0}]\subset\mathbb{R} be the subset obtained in Lemma 4.7. Let δ>0\delta>0 be the constant in Lemma 4.7 and take m¯>0\overline{m}>0 large enough. Then, for b>0b>0 large enough there exists a Palais-Smale sequence {φn}nm¯,δ,2b\left\{\varphi_{n}\right\}_{n\in\mathbb{N}}\subset\mathcal{F}_{\overline{m},\delta,2b} for 𝒜G\mathcal{A}_{G} at the level cGc_{G}.

Proof.

Let ΓC([0,1],D1,2)\varGamma\subset C([0,1],D^{1,2}) be the set defined in (4.9). Let ε<δ/2\varepsilon<\delta/2 and take a suboptimal path γεΓ\gamma_{\varepsilon}\subset\varGamma for which 𝒜G(γε(t))cG+ε\mathcal{A}_{G}(\gamma_{\varepsilon}(t))\leq c_{G}+\varepsilon for all t[0,1]t\in[0,1] such that minsr0(s)+(γε(t))(s)m¯\min_{s\in\mathbb{R}}r_{0}(s)+(\gamma_{\varepsilon}(t))(s)\leq\overline{m}. For all γΓ\gamma\in\varGamma we define the set

γ{Bar(γε(t)):|𝒜G(γ(t))cG|ε,minsr0(s)+(γ(t))(s)m¯,t[0,1]}.\mathcal{B}_{\gamma}\equiv\left\{\mathrm{Bar}(\gamma_{\varepsilon}(t))\colon|\mathcal{A}_{G}(\gamma(t))-c_{G}|\geq\varepsilon,\ \ \min_{s\in\mathbb{R}}r_{0}(s)+(\gamma(t))(s)\leq\overline{m},\ \ t\in[0,1]\right\}.

Notice that for each γΓ\gamma\in\varGamma, the set γ\mathcal{B}_{\gamma} is a compact subset of \mathbb{R}. Denote by

bmin=minγεbmax=maxγε\begin{split}b_{\mathrm{min}}=\min\mathcal{B}_{\gamma_{\varepsilon}}\qquad\qquad b_{\mathrm{max}}=\max\mathcal{B}_{\gamma_{\varepsilon}}\end{split}

and consider the translated path γε1=Tb0γε\gamma_{\varepsilon}^{1}=T_{b_{0}}\gamma_{\varepsilon} for b0=[bmin]b_{0}=[b_{\mathrm{min}}]. It satisfies that

γε1[0,bmaxb0+1]\mathcal{B}_{\gamma_{\varepsilon}^{1}}\subset[0,b_{\mathrm{max}}-b_{0}+1]

Let W:D1,2D1,2W:D^{1,2}\to D^{1,2} be the pseudogradient vector field built in Proposition 4.12 and denote by ητ\eta_{\tau} its time τ\tau flow. Notice that since WW is Lipschitz the flow ητ\eta_{\tau} is well defined at least for sufficiently small τ\tau. Let β>0\beta>0 be the constant in Proposition 4.12. We claim that the deformed path γε1ητ\gamma_{\varepsilon}^{1}\circ\eta_{\tau^{*}} with τ=2ε/β\tau_{*}=2\varepsilon/\beta satisfies

max{𝒜G(ητ(γε1)(t)):minsr0(s)+(ητ(γε1)(t))(s)m¯,|Bar(ητ(γε1)(t))|b,t[0,1]}cGε.\begin{split}\max\{&\mathcal{A}_{G}(\eta_{\tau_{*}}(\gamma_{\varepsilon}^{1})(t))\colon\min_{s\in\mathbb{R}}r_{0}(s)+\left(\eta_{\tau_{*}}(\gamma_{\varepsilon}^{1})(t)\right)(s)\leq\overline{m},\ |\mathrm{Bar}(\eta_{\tau_{*}}(\gamma_{\varepsilon}^{1})(t))|\leq b,\ t\in[0,1]\}\\ &\leq c_{G}-\varepsilon.\end{split}

To verify the claim we first notice that the maximal displacement is bounded by

ητ(φ)φτWτ=2ε/β.\lVert\eta_{\tau^{*}}(\varphi)-\varphi\rVert\leq\tau^{*}\lVert W\rVert\leq\tau^{*}=2\varepsilon/\beta.

Therefore, applying Lemma 4.10, we obtain that for any φγε\varphi\in\gamma_{\varepsilon} (taking bb sufficiently large)

|Bar(ητ(φ))Bar(φ)|LBar2ε/βb/4.|\mathrm{Bar}(\eta_{\tau^{*}}(\varphi))-\mathrm{Bar}(\varphi)|\leq L_{\mathrm{Bar}}2\varepsilon/\beta\leq b/4.

Thus, since the region {minsr0(s)+φ(s)m¯}\{\min_{s\in\mathbb{R}}r_{0}(s)+\varphi(s)\geq\overline{m}\} is forward invariant by the flow ητ\eta_{\tau} and

ddτ(𝒜Gη)0,\frac{\mathrm{d}}{\mathrm{d}\tau}(\mathcal{A}_{G}\circ\eta)\leq 0,

in order to verify the claim, it is enough to check that there does not exist

φ{minsr0(s)+(γε1(t))(s)m¯,Bar(γε1(t))5b/4,t[0,1]}\varphi\in\left\{\min_{s\in\mathbb{R}}r_{0}(s)+\left(\gamma_{\varepsilon}^{1}(t)\right)(s)\leq\overline{m},\ \ \mathrm{Bar}(\gamma_{\varepsilon}^{1}(t))\leq 5b/4,\ \ t\in[0,1]\right\}

for which ητ(γε1)Pτ[0,τ]\eta_{\tau}(\gamma_{\varepsilon}^{1})\in P\ \ \forall\tau\in[0,\tau^{*}] where PD1,2P\subset D^{1,2} is the set defined in (4.15). This is clearly not possible since for φP\varphi\in P we have

ddτ(𝒜Gη)β\frac{\mathrm{d}}{\mathrm{d}\tau}(\mathcal{A}_{G}\circ\eta)\leq-\beta

so

𝒜G(ητ(γε1))𝒜G(γε1)τβcGε\mathcal{A}_{G}(\eta_{\tau^{*}}(\gamma_{\varepsilon}^{1}))\leq\mathcal{A}_{G}(\gamma_{\varepsilon}^{1})-\tau^{*}\beta\leq c_{G}-\varepsilon

a contradiction. Now that the claim is verified consider the path

γε2=Tb0(γε1).\gamma_{\varepsilon}^{2}=T_{-b_{0}}(\gamma_{\varepsilon}^{1}).

It satisfies that

γε2[b0+b,bmax+1]\mathcal{B}_{\gamma_{\varepsilon}^{2}}\subset[b_{0}+b,b_{\mathrm{max}}+1]

and

max{𝒜G(γε2(t)):minsr0(s)+(γε2(t))(s)m¯,Bar(γε2(t))b0+b,t[0,1]}cGε\max\{\mathcal{A}_{G}(\gamma_{\varepsilon}^{2}(t))\colon\min_{s\in\mathbb{R}}r_{0}(s)+\left(\gamma_{\varepsilon}^{2}(t)\right)(s)\leq\overline{m},\ \ \mathrm{Bar}(\gamma_{\varepsilon}^{2}(t))\leq b_{0}+b,\ \ t\in[0,1]\}\leq c_{G}-\varepsilon

If bmaxb0+1bb_{\mathrm{max}}-b_{0}+1\leq b the proposition is proved. In the case bmaxb0+1bb_{\mathrm{max}}-b_{0}+1\geq b we repeat the argument above with the path γε2\gamma_{\varepsilon}^{2} to obtain a path γε3\gamma_{\varepsilon}^{3} satisfying

γε3[b0+2b,bmax+1]\mathcal{B}_{\gamma_{\varepsilon}^{3}}\subset[b_{0}+2b,b_{\mathrm{max}}+1]

and

max{𝒜G(γε3(t)):minsr0(s)+(γε3(t))(s)m¯,Bar(γε3(t))b0+2b,t[0,1]}cGε\max\{\mathcal{A}_{G}(\gamma_{\varepsilon}^{3}(t))\colon\min_{s\in\mathbb{R}}r_{0}(s)+\left(\gamma_{\varepsilon}^{3}(t)\right)(s)\leq\overline{m},\ \ \mathrm{Bar}(\gamma_{\varepsilon}^{3}(t))\leq b_{0}+2b,\ \ t\in[0,1]\}\leq c_{G}-\varepsilon

The result follows after repeating the construction no more than [(bmaxb0+1)/b][(b_{\mathrm{max}}-b_{0}+1)/b] steps. ∎

Finally, we obtain the existence of a critical point of the functional 𝒜G\mathcal{A}_{G} at a level cGc_{G} .

Theorem 4.14.

Let J[G0,G0]J\subset[-G_{0},G_{0}]\subset\mathbb{R} be the subset obtained in Lemma 4.7. Then, for all GJG\in J there exists a critical point of the action functional 𝒜G\mathcal{A}_{G} at the level cGc_{G}.

Proof.

By Proposition 4.13, for sufficiently large m¯\overline{m} and bb there exists a Palais-Smale sequence {φn}nm¯,δ,2b\left\{\varphi_{n}\right\}_{n\in\mathbb{N}}\subset\mathcal{F}_{\overline{m},\delta,2b} so it follows from Lemma 4.11 that it is bounded. The theorem is then proved since the Palais Smale sequence {φn}n\{\varphi_{n}\}_{n\in\mathbb{N}} satisfies the hypothesis for the set QQ of the compactness Lemma 4.5. ∎

5. Topological transversality between the stable and unstable manifolds

For the choice of G0>0G_{0}>0 was arbitrary, Theorem 4.14 implies that for any compact subset [G0,G0][-G_{0},G_{0}] of the real line, there exists a full measure subset J[G0,G0]J\subset[-G_{0},G_{0}] such that for all GJG\in J there exists an orbit of (1.1) which is homoclinic to γ\gamma_{\infty}. Another way of rephrasing Theorem 4.14 is that the invariant manifolds manifolds W±(γ,G)W^{\pm}(\gamma_{\infty},G) defined in (3.2) intersect for almost all values of GG in [G0,G0][-G_{0},G_{0}]. However, Theorem 4.14 contains no information about the geometry of the intersection, in particular wether it is transversal or not.

Theorem 4.1, proved in [GPSV21], shows that the intersection between W±(γ,G)W^{\pm}(\gamma_{\infty},G) is transverse for all GG sufficiently large. Moreover, the local stable manifolds Wloc±(γ;G)W^{\pm}_{\mathrm{loc}}(\gamma_{\infty};G) (see (3.1)) depend analytically on rr and GG (see Proposition 3.1). We want to exploit this facts to deduce that the intersection of the manifolds W±(γ,G)W^{\pm}(\gamma_{\infty},G) (which, by Theorem 4.14 we already know that exists for almost all values of G[G0,G0]G\in[-G_{0},G_{0}]) must be topologically transverse for almost all values of GG in [G0,G0][-G_{0},G_{0}].

Remark 8.

In the following we fix a sufficiently large value of G0G_{0} and work with GJ[G0,G0]G\in J\subset[-G_{0},G_{0}].

The first step is to obtain an a priori estimate from below for minsrh(s)\min_{s\in\mathbb{R}}r_{h}(s).

Lemma 5.1.

Let GG\in\mathbb{R} and let rh(s;G):r_{h}(s;G):\mathbb{R}\to\mathbb{R} be an orbit of of the Hamiltonian HGH_{G} in (1.1) which is homoclinic to γ\gamma_{\infty}. Then, for all ss\in\mathbb{R} we have

rh(s)G22.r_{h}(s)\geq\frac{G^{2}}{2}.
Proof.

Since rh(s;G):r_{h}(s;G):\mathbb{R}\to\mathbb{R} is an orbit of of the Hamiltonian HGH_{G} we have that

(5.1) r¨h=G2rh3rh(rh2+ρ2)3/2G2rh31rh2.\ddot{r}_{h}=\frac{G^{2}}{r_{h}^{3}}-\frac{r_{h}}{(r_{h}^{2}+\rho^{2})^{3/2}}\geq\frac{G^{2}}{r_{h}^{3}}-\frac{1}{r_{h}^{2}}.

Let now II\subset\mathbb{R} be an interval in which r˙h(s)0\dot{r}_{h}(s)\leq 0 for all sIs\in I. Then, multiplying (5.1) by r˙h\dot{r}_{h}, for all sIs\in I we obtain

dds(r˙h22+G22rh21rh)0,\frac{\mathrm{d}}{\mathrm{d}s}\left(\frac{\dot{r}_{h}^{2}}{2}+\frac{G^{2}}{2r_{h}^{2}}-\frac{1}{r_{h}}\right)\leq 0,

that is, the energy

E(s)=r˙h2(s)2+G22rh2(s)1rh(s)E(s)=\frac{\dot{r}_{h}^{2}(s)}{2}+\frac{G^{2}}{2r_{h}^{2}(s)}-\frac{1}{r_{h}(s)}

is non increasing on the interval II. Let II be a maximal interval in which rh(s)r_{h}(s) is decreasing: we distinguish between two alternatives, either

  • I=(,s1]I=(-\infty,s_{1}], and limsrh(s)=\lim_{s\to-\infty}r_{h}(s)=\infty, limsr˙h(s)=0\lim_{s\to-\infty}\dot{r}_{h}(s)=0 and r˙h(s1)=0\dot{r}_{h}(s_{1})=0, or

  • I=[s0,s1]I=[s_{0},s_{1}] and r˙h(s0)=0\dot{r}_{h}(s_{0})=0, r¨h(s0)0\ddot{r}_{h}(s_{0})\leq 0 and r˙h(s1)=0\dot{r}_{h}(s_{1})=0

for some <s0<s1<-\infty<s_{0}<s_{1}<\infty. In the first case we have

limsE(s)=0.\lim_{s\to-\infty}E(s)=0.

In the second case, using that r˙h(s0)=0\dot{r}_{h}(s_{0})=0, r¨h(s0)0\ddot{r}_{h}(s_{0})\leq 0 and the inequality (5.1), we obtain

E(s0)=r˙h2(s0)2+G22rh2(s0)1rh(s0)=G22rh2(s0)1rh(s0)G2rh2(s0)1rh(s0)0.E(s_{0})=\frac{\dot{r}_{h}^{2}(s_{0})}{2}+\frac{G^{2}}{2r_{h}^{2}(s_{0})}-\frac{1}{r_{h}(s_{0})}=\frac{G^{2}}{2r_{h}^{2}(s_{0})}-\frac{1}{r_{h}(s_{0})}\leq\frac{G^{2}}{r_{h}^{2}(s_{0})}-\frac{1}{r_{h}(s_{0})}\leq 0.

Therefore, in both cases, for all sIs\in I we have E(s)0E(s)\leq 0, which implies that

r(s)r(s1)G22.r(s)\geq r(s_{1})\geq\frac{G^{2}}{2}.

Lemma 5.1 implies that for G0G\neq 0, homoclinic orbits do not intersect the section {r=0}\{r=0\}. This fact allows us to exploit the analytic dependence of the Hamiltonian HGH_{G} in the parameter GG to prove the following result.

Lemma 5.2.

The set of values of G{0}G\in\mathbb{R}\setminus\{0\} for which W(γ,G)=W+(γ,G)W^{-}(\gamma_{\infty},G)=W^{+}(\gamma_{\infty},G), is finite.

Proof.

Fix any δ>0\delta>0 and let GG_{*} be the constant in Theorem 4.1 and let 1R1<R21\ll R_{1}<R_{2} be such that for all G[2G,2G]G\in[-2G_{*},2G_{*}] the generating function S+(r,t;G)S^{+}(r,t;G) associated with the local stable manifold (see 3.1) is well defined for all (r,t)[R1,R2]×𝕋(r,t)\in[R_{1},R_{2}]\times\mathbb{T}. Define the set

Q={(r,y,t)+××𝕋:r(R1,R2),y>0,t=0}.Q=\{(r,y,t)\in\mathbb{R}_{+}\times\mathbb{R}\times\mathbb{T}\colon r\in(R_{1},R_{2}),\ y>0,\ t=0\}.

Whenever it exists, denote by γGQ𝒲u(γ;G)\gamma_{G}^{-}\subset Q\cap\mathcal{W}^{u}(\gamma_{\infty};G) the connected component of Q𝒲u(γ;G)Q\cap\mathcal{W}^{u}(\gamma_{\infty};G) associated with the first backwards intersection of 𝒲u(γ;G)\mathcal{W}^{u}(\gamma_{\infty};G) with QQ (see Figure 5.1). Define now the set

𝒢~={G:δ|G|2G,γGandφGCω([R1,R2],)such thatγG=graph(φG)}.\mathcal{\widetilde{G}}=\{G\in\mathbb{R}\colon\delta\leq|G|\leq 2G_{*},\ \gamma_{G}^{-}\neq\emptyset\ \text{and}\ \exists\ \varphi^{-}_{G}\in C^{\omega}\left([R_{1},R_{2}],\mathbb{R}\right)\ \text{such that}\ \gamma_{G}^{-}=\mathrm{graph}(\varphi^{-}_{G})\}.
Refer to caption
Figure 5.1. The domain QQ and a sketch of the intersection of the stable manifolds W±(γ;G)W^{\pm}(\gamma_{\infty};G) with QQ for a value of G𝒢~G\in\mathcal{\widetilde{G}}.

Clearly, 𝒢𝒢~\mathcal{G}\subset\mathcal{\widetilde{G}} where

𝒢={G:δ|G|2G,W+(γ;G)=W(γ;G)}.\mathcal{G}=\{G\in\mathbb{R}\colon\delta\leq|G|\leq 2G_{*},\ W^{+}(\gamma_{\infty};G)=W^{-}(\gamma_{\infty};G)\}.

In view of Lemma 5.1, for all G𝒢G\in\mathcal{G}

(5.2) dist(W±(γ,G),{r=0})G2/2.\mathrm{dist}(W^{\pm}(\gamma_{\infty},G),\{r=0\})\geq G^{2}/2.

so 𝒢\mathcal{G} is a closed set. Moreover, since the Hamiltonian (1.2) depends analytically on GG, and, by (5.2), for all G𝒢G\in\mathcal{G} the vector field associated with (1.2) is analytic on a neighbourhood of W±(γ,G)W^{\pm}(\gamma_{\infty},G), there exists an open subset 𝒪𝒢~\mathcal{O}\subset\mathcal{\widetilde{G}} such that 𝒢𝒪\mathcal{G}\subset\mathcal{O} and in which φGCω([R1,R2]×𝒪)\varphi^{-}_{G}\in C^{\omega}([R_{1},R_{2}]\times\mathcal{O}). Define now the function Δ(r,G):[R1,R2]×𝒢~\Delta(r,G):[R_{1},R_{2}]\times\mathcal{\widetilde{G}}\to\mathbb{R} given by

Δ(r,G)=φG(r)rS+(r,0;G)\Delta(r,G)=\varphi^{-}_{G}(r)-\partial_{r}S^{+}(r,0;G)

which satisfies Δ(r,G)=0\varDelta(r,G)=0 for all G𝒢~G\in\mathcal{\widetilde{G}} and ΔCω([R1,R2]×𝒪)\varDelta\in C^{\omega}\left([R_{1},R_{2}]\times\mathcal{O}\right). Suppose now that 𝒢={δ|G|2G}\mathcal{G}=\{\delta\leq|G|\leq 2G_{*}\}. Then, Δ(r,G)=0\Delta(r,G)=0 for all (r,G)[R1,R2]×{δ|G|2G}(r,G)\in[R_{1},R_{2}]\times\{\delta\leq|G|\leq 2G_{*}\} and we obtain a contradiction with the fact that for |G|G|G|\geq G_{*} the manifolds Wu(γ,G)W^{u}(\gamma_{\infty},G), Ws(γ,G)W^{s}(\gamma_{\infty},G) intersect transversally (see Theorem 4.1). Therefore, 𝒢{δ|G|2G}\mathcal{G}\subsetneq\{\delta\leq|G|\leq 2G_{*}\}. We now show that, moreover, 𝒢\mathcal{G} cannot contain any accumulation point. To see this suppose that there exists Gmax𝒢G_{max}\in\mathcal{G} such that

Gmax=max{G𝒢:Gis an accumulation point of𝒢}.G_{max}=\max\{G\in\mathcal{G}\colon G\ \text{is an accumulation point of}\ \mathcal{G}\}.

Since Gmax𝒢G_{max}\in\mathcal{G} there exists an open interval 𝒱𝒪\mathcal{V}\subset\mathcal{O} such that G𝒱G\in\mathcal{V}. Then, the fact that 𝒱𝒪\mathcal{V}\subset\mathcal{O} implies thatΔ(r,G)Cω([R1,R2]×𝒱)\varDelta(r,G)\in C^{\omega}\left([R_{1},R_{2}]\times\mathcal{V}\right) and since GmaxG_{max} is an acuumulation point of 𝒢\mathcal{G} we conclude that Δ(r,G)=0\Delta(r,G)=0 on [R1,R2]×𝒱[R_{1},R_{2}]\times\mathcal{V}. Then 𝒱𝒢\mathcal{V}\subset\mathcal{G}, so there exists G~𝒱𝒢\tilde{G}\in\mathcal{V}\subset\mathcal{G} such that G~>Gmax\tilde{G}>G_{max}, a contradiction with the definition of GmaxG_{max}. ∎

Denote now by 𝒥\mathcal{J}\subset\mathbb{R} the set

(5.3) 𝒥={GJ:G0,W+(γ;G)W(γ;G)}\mathcal{J}=\{G\in J\colon G\neq 0,\ W^{+}(\gamma_{\infty};G)\neq W^{-}(\gamma_{\infty};G)\}

where JJ was defined in Lemma 4.7 (see also Theorem 4.14).

Lemma 5.3.

For all G𝒥G\in\mathcal{J} the set Crit(𝒜G)={φD1,2:d𝒜G(φ)=0}\mathrm{Crit}(\mathcal{A}_{G})=\{\varphi\in D^{1,2}\colon\mathrm{d}\mathcal{A}_{G}(\varphi)=0\} is isolated in D1,2D^{1,2}.

Proof.

Following [MNT99], we define the map TR:Crit(𝒜)D1,2T_{R}:\mathrm{Crit}(\mathcal{A})\subset D^{1,2}\to\mathbb{R} given by

TR=sup{s:r0(s)+φ(s)=R,φCrit(𝒜)D1,2}.T_{R}=\sup\{s\in\mathbb{R}\colon r_{0}(s)+\varphi(s)=R,\ \varphi\in\mathrm{Crit}(\mathcal{A})\subset D^{1,2}\}.

We now show that the set TR(Crit(𝒜))T_{R}(\mathrm{Crit}(\mathcal{A})) is isolated in \mathbb{R}. Suppose on the contrary that there exists an accumulation point TTR(Crit(𝒜))T_{*}\in T_{R}(\mathrm{Crit}(\mathcal{A})), then, there exist {(φn,tn)}nCrit(𝒜)×\{(\varphi_{n},t_{n})\}_{n\in\mathbb{N}}\subset\mathrm{Crit}(\mathcal{A})\times\mathbb{R} and R+R\in\mathbb{R}_{+} such that tnTRt_{n}\to T_{R}, (r0+φn)(tn)=R(r_{0}+\varphi_{n})(t_{n})=R and

((r0+φn)(tn),(r˙0+φ˙n)(tn))Wloc+(γ;G).((r_{0}+\varphi_{n})(t_{n}),(\dot{r}_{0}+\dot{\varphi}_{n})(t_{n}))\in W^{+}_{\mathrm{loc}}(\gamma_{\infty};G).

Thus, there exist infinitely many different homoclinic points contained in the piece of the local stable manifold γ+={y=rS+(r,t),t=T,r[R,R1]}\gamma_{+}=\{y=\partial_{r}S^{+}(r,t),\ t=T_{*},\ r\in[R,R_{1}]\} for any R1<RR_{1}<R. This would imply the existence of T<TT_{**}<T_{*}, R2<R3R_{2}<R_{3} such that γ+ϕTT(γ)\gamma_{+}\cap\phi^{T_{*}-T_{**}}(\gamma_{-}) intersect at infinitely many points, where γ={y=rS(R,t),t=T,r[R2,R3]}\gamma_{-}=\{y=\partial_{r}S^{-}(R,t),\ t=T_{**},\ r\in[R_{2},R_{3}]\}.However, γ+\gamma_{+} and γ\gamma_{-} are compact analytic curves, and since G𝒥G\in\mathcal{J} they cannot intersect at infinitely many points.

By Lemma 3.3. in [MNT99], the function TR:Crit(𝒜)D1,2T_{R}:\mathrm{Crit}(\mathcal{A})\subset D^{1,2}\to\mathbb{R} is continuous, so the lemma is proven, for if it were to be false there would exist an accumulation point TTR(Crit(𝒜))T_{*}\in T_{R}(\mathrm{Crit}(\mathcal{A})). ∎

The fact that the critical points are isolated implies the following non-degeneracy property at, at least, one of the critical points of 𝒜G\mathcal{A}_{G} at the level cGc_{G}. We say that φCrit(𝒜G)\varphi_{*}\in\mathrm{Crit}(\mathcal{A}_{G}) has a local mountain pass structure if, for all neighbourhood UD1,2U\subset D^{1,2} of φ\varphi_{*}, the set {φU:𝒜G(φ)<𝒜G(φ)}\{\varphi\in U\colon\mathcal{A}_{G}(\varphi)<\mathcal{A}_{G}(\varphi_{*})\} is not path connected. The following result is a direct consequence of Lemma 5.3 and Theorem 1 in [Hof86].

Proposition 5.4.

For all G𝒥G\in\mathcal{J} there exists φCrit(𝒜G)\varphi_{*}\in\mathrm{Crit}(\mathcal{A}_{G}) such that 𝒜G(φ)=cG\mathcal{A}_{G}(\varphi_{*})=c_{G} which has a local mountain pass structure.

Remark 9.

In all the forthcoming sections we fix G𝒥G\in\mathcal{J} where 𝒥\mathcal{J} is the set defined in (5.3) and omit the dependence on GG.

5.1. The reduced action functional

For n{0}n\in\mathbb{N}\setminus\{0\} we denote by H1([n,n])H^{1}([-n,n]) the usual Sobolev space consisting of functions defined on the interval [n,n][-n,n]\subset\mathbb{R} with one weak derivative in L2([n,n])L^{2}([-n,n]) and introduce the restriction operator

(5.4) j:D1,2\displaystyle j:D^{1,2} H1([n,n])\displaystyle\longrightarrow H^{1}([-n,n])
φ\displaystyle\varphi j(φ)=φ|[n,n]\displaystyle\longmapsto j(\varphi)=\varphi|_{[-n,n]}

Then, for a sufficiently small neighbourhood U~H1([n,n])\tilde{U}\subset H^{1}([-n,n]) of a point φ~=j(φ)\tilde{\varphi}_{*}=j(\varphi_{*}) where φD1,2\varphi_{*}\in D^{1,2} and a sufficiently large nn\in\mathbb{N} (depending on φ\varphi_{*}) we define the reduced action functional 𝒜~:U~H1([n,n])\widetilde{\mathcal{A}}:\tilde{U}\subset H^{1}([-n,n])\rightarrow\mathbb{R} given by

𝒜~(φ~)=nnren(φ~,φ~˙,s)dsS+((r0+φ~)(n))+S((r0+φ~)(n))+r˙0(n)(φ~(n)φ~(n)),\widetilde{\mathcal{A}}(\tilde{\varphi})=\int_{-n}^{n}\mathcal{L}_{\mathrm{ren}}(\tilde{\varphi},\dot{\tilde{\varphi}},s)\mathrm{d}s-S^{+}((r_{0}+\tilde{\varphi})(n))+S^{-}((r_{0}+\tilde{\varphi})(-n))+\dot{r}_{0}(n)(\tilde{\varphi}(n)-\tilde{\varphi}(-n)),

where the renormalized Lagrangian ren\mathcal{L}_{\mathrm{ren}} is defined in (4.3) and S±S^{\pm} are the generating functions of the local stable and unstable manifolds which were obtained in Proposition 3.1. Notice that for nn sufficiently large (depending on φ\varphi_{*}) and φ~\tilde{\varphi} sufficiently close to j(φ)j(\varphi_{*}) the values (r0+φ~)(±n)(r_{0}+\tilde{\varphi})(\pm n) are contained in Dom(S±)\mathrm{Dom}(S^{\pm}).

We now want to translate the results we have obtained for the functional 𝒜\mathcal{A}, in particular Proposition 5.4, in results for the functional 𝒜~\mathcal{\widetilde{A}}. To that end, given any constant cc\in\mathbb{R} and nn\in\mathbb{N} we introduce the functional spaces

D+1,2(c,n)={φC([n,)):vφL2([n,))such that φ(s)=c+nsvφ(t)dt,s[n,)}\begin{split}D^{1,2}_{+}(c,n)=\{\varphi\in C([n,\infty))\colon&\exists v_{\varphi}\in L^{2}([n,\infty))\ \text{such that }\\ &\varphi(s)=c+\int_{n}^{s}v_{\varphi}(t)\mathrm{d}t,\ \forall s\in[n,\infty)\}\end{split}

and

D1,2(c,n)={φC((,n]):vφL2((,n])such that φ(s)=csnvφ(t)dt,s(,n]}\begin{split}D^{1,2}_{-}(c,n)=\{\varphi\in C((-\infty,-n])\colon&\exists v_{\varphi}\in L^{2}((-\infty,-n])\ \text{such that }\\ &\varphi(s)=c-\int_{s}^{-n}v_{\varphi}(t)\mathrm{d}t,\ \forall s\in(-\infty,-n]\}\end{split}

Define also the weakly closed subsets

D~+1,2(c,n)={φD+1,2(c,n):r0(s)+φ(s)r0(n)+c,s[n,)}D~1,2(c,n)={φD1,2(c,n):r0(s)+φ(s)r0(n)+c,s(,n]}\begin{split}\tilde{D}^{1,2}_{+}(c,n)=&\{\varphi\in D^{1,2}_{+}(c,n)\colon r_{0}(s)+\varphi(s)\geq r_{0}(n)+c,\ \forall s\in[n,\infty)\}\\ \tilde{D}^{1,2}_{-}(c,n)=&\{\varphi\in D^{1,2}_{-}(c,n)\colon r_{0}(s)+\varphi(s)\geq r_{0}(-n)+c,\ \forall s\in(-\infty,-n]\}\\ \end{split}

Then, we define the asymptotic actions

(5.5) 𝒜±(φ)=±±n±ren(φ,φ˙,s)ds\mathcal{A}^{\pm}(\varphi)=\pm\int_{\pm n}^{\pm\infty}\mathcal{L}_{\mathrm{ren}}(\varphi,\dot{\varphi},s)\mathrm{d}s
Lemma 5.5.

For all cc\in\mathbb{R} there exists n0n_{0}\in\mathbb{N} such that for all nn0n\geq n_{0} there exists a unique φ±D~±1,2(c,n)\varphi_{\pm}\in\tilde{D}^{1,2}_{\pm}(c,n) such that

𝒜±(φ±)=min{𝒜±(ψ):ψD~±1,2(c,n))}.\mathcal{A}^{\pm}(\varphi_{\pm})=\min\{\mathcal{A}^{\pm}(\psi)\colon\psi\in\tilde{D}^{1,2}_{\pm}(c,n))\}.

Moreover,

𝒜±(φ±)=S±((r0(±n)+c)±S0((r0)(±n))r˙0(±n)c.\mathcal{A}^{\pm}(\varphi_{\pm})=\mp S^{\pm}((r_{0}(\pm n)+c)\pm S^{0}((r_{0})(\pm n))\mp\dot{r}_{0}(\pm n)c.
Proof.

A simple computation shows that

rr2V(r,t)=3G2r4+3r2(r2+ρ2(t))5/21(r2+ρ2(t))3/2\partial_{rr}^{2}V(r,t)=-\frac{3G^{2}}{r^{4}}+\frac{3r^{2}}{(r^{2}+\rho^{2}(t))^{5/2}}-\frac{1}{(r^{2}+\rho^{2}(t))^{3/2}}

from where we easily deduce that there exists R>0R>0 such that, if

r0(n)+cRr_{0}(n)+c\geq R

then, the functional φ𝒜+(φ)\varphi\mapsto\mathcal{A}^{+}(\varphi) in (5.5) is strictly convex on the strictly convex set D~+1,2(c,n)\tilde{D}^{1,2}_{+}(c,n). Therefore, there exists a unique minimizer φ+D~+1,2(c,n)\varphi_{+}\in\tilde{D}^{1,2}_{+}(c,n) for which

𝒜+(φ+)=min{𝒜+(ψ):ψD~+1,2(φ~(n))}.\mathcal{A}^{+}(\varphi_{+})=\min\{\mathcal{A}^{+}(\psi)\colon\psi\in\tilde{D}^{1,2}_{+}(\tilde{\varphi}(n))\}.

Moreover, is easy to check that φ+\varphi_{+} is a critical point of the functional 𝒜+(φ)\mathcal{A}^{+}(\varphi). Consequently, r(s)=r0(s)+φ+(s)r(s)=r_{0}(s)+\varphi_{+}(s) is an orbit of (1.1) asymptotic in the future to γ\gamma_{\infty}.

Let now S+(r,s)S^{+}(r,s) be the generating function of the local stable manifold introduced in Proposition 3.1. By uniqueness of the local stable manifold, the function φ+(s)\varphi_{+}(s) satisfies that

(r˙0+φ˙+)(s)=rS+(r0(s)+φ+(s),s)(\dot{r}_{0}+\dot{\varphi}_{+})(s)=\partial_{r}S^{+}(r_{0}(s)+\varphi_{+}(s),s)

for all s[n,)s\in[n,\infty). In particular, since moreover φ+(s)D+1,2\varphi_{+}(s)\in D^{1,2}_{+} Lemma A.2 in Appendix A implies that |φ˙+(s)|s1/6|\dot{\varphi}_{+}(s)|\lesssim s^{1/6} as ss\to\infty and since r˙0(s)s1/3\dot{r}_{0}(s)\sim s^{-1/3} as ss\to\infty , we can integrate by parts to obtain

nren(φ+,φ˙+,s)=nφ˙+22+V(r0+φ+)V0(r0)r¨0φ+=r˙0(n)c+nφ˙+22+r˙0φ˙+V(r0+φ+)V0(r0).\begin{split}\int_{n}^{\infty}\mathcal{L}_{\mathrm{ren}}(\varphi_{+},\dot{\varphi}_{+},s)=&\int_{n}^{\infty}\frac{\dot{\varphi}_{+}^{2}}{2}+V(r_{0}+\varphi_{+})-V_{0}(r_{0})-\ddot{r}_{0}\varphi_{+}\\ =&-\dot{r}_{0}(n)c+\int_{n}^{\infty}\frac{\dot{\varphi}_{+}^{2}}{2}+\dot{r}_{0}\dot{\varphi}+V(r_{0}+\varphi_{+})-V_{0}(r_{0}).\end{split}

On the other hand,

nφ˙+22+r˙0φ˙+V(r0+φ+)V0(r0)=n((r˙0+φ˙+)rS+(r0+φ+)H(r0+φ+,rS+(r0+φ+),s)r˙0rS0(r0)H0(r0,rS0(r0)))=nddsS+((r0+φ+)(s))ddsS0(r0(s))=S+((r0(n)+c)+S0(r0(n))\begin{split}\int_{n}^{\infty}\frac{\dot{\varphi}_{+}^{2}}{2}+\dot{r}_{0}\dot{\varphi}+V(r_{0}+\varphi_{+})-&V_{0}(r_{0})\\ =&\int_{n}^{\infty}\left((\dot{r}_{0}+\dot{\varphi}_{+})\partial_{r}S^{+}(r_{0}+\varphi_{+})-H(r_{0}+\varphi_{+},\partial_{r}S^{+}(r_{0}+\varphi_{+}),s)\right.\\ &\left.-\dot{r}_{0}\ \partial_{r}S^{0}(r_{0})-H_{0}(r_{0},\partial_{r}S^{0}(r_{0}))\right)\\ =&\int_{n}^{\infty}\frac{\mathrm{d}}{\mathrm{d}s}S^{+}((r_{0}+\varphi_{+})(s))-\frac{\mathrm{d}}{\mathrm{d}s}S^{0}(r_{0}(s))\\ =&-S^{+}((r_{0}(n)+c)+S^{0}(r_{0}(n))\end{split}

where we have used that H(r0+φ+,rS+(r0+φ+),s)+tS+(r0+φ+,s)=0H(r_{0}+\varphi_{+},\partial_{r}S^{+}(r_{0}+\varphi_{+}),s)+\partial_{t}S^{+}(r_{0}+\varphi_{+},s)=0 and the fact that

limsS+((r0+φ+)(s))S0(r0(s))=0,\lim_{s\to\infty}S^{+}((r_{0}+\varphi_{+})(s))-S^{0}(r_{0}(s))=0,

which is also proved in Lemma A.2. ∎

Introduce now the extension operator E:U~H1([n,n])D1,2E:\tilde{U}\subset H^{1}([-n,n])\to D^{1,2}

(5.6) E(φ~)={E(φ~)fors(,n)φ~fors[n,n]E+(φ~)fors(n,)E(\tilde{\varphi})=\left\{\begin{array}[]{ccr}E_{-}(\tilde{\varphi})&\text{for}&\quad s\in(-\infty,-n)\\ \tilde{\varphi}&\text{for}&\quad s\in[-n,n]\\ E_{+}(\tilde{\varphi})&\text{for}&\quad s\in(n,\infty)\\ \end{array}\right.

where

E±(φ~)={φD~±1,2(φ~(±n)):𝒜±(φ)𝒜±(ψ),ψD~±1,2(φ~(±n))}.\begin{split}E_{\pm}(\tilde{\varphi})=&\{\varphi\in\tilde{D}^{1,2}_{\pm}(\tilde{\varphi}(\pm n))\colon\mathcal{A}^{\pm}(\varphi)\leq\mathcal{A}^{\pm}(\psi),\ \forall\psi\in\tilde{D}^{1,2}_{\pm}(\tilde{\varphi}(\pm n))\}.\\ \end{split}

From the proof of Lemma 5.5 we deduce the following.

Lemma 5.6.

Let φD1,2\varphi\in D^{1,2}, let nn\in\mathbb{N} sufficiently large, let φ~=j(φ)\tilde{\varphi}=j(\varphi) and let U~H1([n,n])\widetilde{U}\subset H^{1}([-n,n]) a sufficiently small neighbourhood of φ~\tilde{\varphi}. Then, the extension operator (5.6) is well defined on U~\widetilde{U}.

The proof of the following Lemma is an straightforward consequence of the definition of the extension operator EE.

Lemma 5.7.

Let φD1,2\varphi_{*}\in D^{1,2}. Then, for nn\in\mathbb{N} sufficiently large and all φ\varphi contained in a sufficiently small neigubourhood UD1,2U\subset D^{1,2} of φ\varphi_{*}

𝒜~(j(φ))𝒜(φ).\mathcal{\widetilde{A}}(j(\varphi))\leq\mathcal{A}(\varphi).

Also, for all φ~\tilde{\varphi} in a sufficiently small neighbourhood U~H1([n,n])\tilde{U}\subset H^{1}([-n,n]) of j(φ)j(\varphi_{*})

𝒜~(φ~)=𝒜(E(φ~)).\mathcal{\widetilde{A}}(\tilde{\varphi})=\mathcal{A}(E(\tilde{\varphi})).

Moreover, for φD1,2\varphi_{*}\in D^{1,2} such that d𝒜(φ)=0\mathrm{d}\mathcal{A}(\varphi_{*})=0 we have d𝒜~(j(φ))=0\mathrm{d}\mathcal{\widetilde{A}}(j(\varphi_{*}))=0.

We can now translate the result for 𝒜\mathcal{A} stated in Proposition 5.3 in an analogous result for 𝒜~\mathcal{\tilde{A}}.

Proposition 5.8.

There exists nn\in\mathbb{N} and φ~H1([n,n])\tilde{\varphi}_{*}\subset H^{1}([-n,n]) which is a critical point of 𝒜~\widetilde{\mathcal{A}} and has a local mountain pass structure.

Proof.

The proof is a simple combination of the proof of Theorem 1 in [Hof86] together with the relationship between the functionals 𝒜\mathcal{A} and 𝒜~\tilde{\mathcal{A}} which was obtained in Lemma 5.7. We sketch here the details for the sake of completeness.

Denote by Crit(𝒜,cG)={φCrit(𝒜)D1,2:𝒜(φ)=cG}\mathrm{Crit}(\mathcal{A},c_{G})=\{\varphi\in\mathrm{Crit}(\mathcal{A})\subset D^{1,2}\colon\mathcal{A}(\varphi)=c_{G}\} where cGc_{G} is the critical value defined in (4.10). Lemma 5.3 implies, in particular, that Crit(𝒜,cG)\mathrm{Crit}(\mathcal{A},c_{G}) is an isolated subset in D1,2D^{1,2}. Moreover, fixing m¯\overline{m} sufficiently large Crit(𝒜,cG)m¯\mathrm{Crit}(\mathcal{A},c_{G})\subset\mathcal{F}_{\overline{m}} where m¯\mathcal{F}_{\overline{m}} was defined in (4.8). Let now ε>0\varepsilon>0 and γεΓD1,2\gamma_{\varepsilon}\subset\varGamma\subset D^{1,2} be a suboptimal path at level cGc_{G}. Then, γε\gamma_{\varepsilon} intersects a finite number of elements in Crit(𝒜,cG)\mathrm{Crit}(\mathcal{A},c_{G}), which we denote by {φ1,,φk}\{\varphi_{1},\dots,\varphi_{k}\} for some finite kk. Let now δ>0\delta>0 sufficiently small and denote by i,δD1,2\mathcal{B}_{i,\delta}\subset D^{1,2} the ball of radius δ\delta around φi\varphi_{i}. Without loss of generality we can assume that γε\gamma_{\varepsilon} intersects each i,δ\mathcal{B}_{i,\delta} only once so we can define (see Figure 5.2)

ti=inf{t[0,1]:γ(t)i,δ}ti+=sup{t[0,1]:γ(t)i,δ}.t_{i}^{-}=\inf\{t\in[0,1]\colon\gamma(t)\in\mathcal{B}_{i,\delta}\}\qquad\qquad t_{i}^{+}=\sup\{t\in[0,1]\colon\gamma(t)\in\mathcal{B}_{i,\delta}\}.
Refer to caption
Figure 5.2. Sketch of the suboptimal path γε\gamma_{\varepsilon}.

and ei=γ(ti)e^{-}_{i}=\gamma(t_{i}^{-}) and ei+=γ(ti+)e^{+}_{i}=\gamma(t_{i}^{+}). Let now nn\in\mathbb{N} large enough and δ\delta small enough so the restricition operator jj in(5.4) is well defined on 1iki,δ\cup_{1\leq i\leq k}\mathcal{B}_{i,\delta}. Let ~i,δ=j(i,δ)\tilde{\mathcal{B}}_{i,\delta}=j(\mathcal{B}_{i,\delta}). Again, without loss of generality, we can assume that for all i=1,,ki=1,\dots,k, ei±D1,2e_{i}^{\pm}\in D^{1,2} have minimizing tails, that is, ei±|[n,)D~1,2(ei±(n),n)e_{i}^{\pm}|_{[n,\infty)}\subset\tilde{D}^{1,2}(e_{i}^{\pm}(n),n) is the unique minimizer of 𝒜+\mathcal{A}^{+} on D~1,2(ei±(n),n)\tilde{D}^{1,2}(e_{i}^{\pm}(n),n) and ei±|(,n]D~1,2(ei±(n),n)e_{i}^{\pm}|_{(-\infty,-n]}\subset\tilde{D}^{1,2}(e_{i}^{\pm}(-n),n) is the unique minimizer of 𝒜\mathcal{A}^{-} on D~1,2(ei±(n),n)\tilde{D}^{1,2}(e_{i}^{\pm}(-n),n). Now, define the paths

γ~i=j(γ|[titi+])H1([n,n])\tilde{\gamma}_{i}=j(\gamma|_{[t_{i}^{-}t_{i}^{+}]})\subset H^{1}([-n,n])

for i=1,,ki=1,\dots,k, and the points φ~i=j(φi)H1([n,n])\tilde{\varphi}_{i}=j(\varphi_{i})\in H^{1}([-n,n]), which are indeed critical points of the reduced action functional 𝒜~\mathcal{\tilde{A}}. Suppose the point φ~i\tilde{\varphi}_{i} does not have a local mountain pass structure. Then, we can build (see Lemma 1 in [Hof86]) a continuous deformation η:[0,1]×~i,δH1([n,n])\eta:[0,1]\times\tilde{\mathcal{B}}_{i,\delta}\to H^{1}([-n,n]) such that

η({1}×({φ~~i,δ:𝒜~(φ)cG+ε}~i,δ/2)){φ~~i,δ:𝒜~(φ)cGε}\eta\left(\{1\}\times(\{\tilde{\varphi}\in\tilde{\mathcal{B}}_{i,\delta}\colon\mathcal{\tilde{A}}(\varphi)\leq c_{G}+\varepsilon\}\setminus\tilde{\mathcal{B}}_{i,\delta/2})\right)\subset\{\tilde{\varphi}\in\tilde{\mathcal{B}}_{i,\delta}\colon\mathcal{\tilde{A}}(\varphi)\leq c_{G}-\varepsilon\}
η([0,1]×Cl(~i,δ/2))~i,δ\eta\left([0,1]\times\mathrm{Cl}(\tilde{\mathcal{B}}_{i,\delta/2})\right)\subset\tilde{\mathcal{B}}_{i,\delta}
η(z,φ)=φ(z,φ)[0,1]×{φ~~i,δ:|𝒜~(φ)cG|ε}.\eta(z,\varphi)=\varphi\qquad\forall(z,\varphi)\in[0,1]\times\{\tilde{\varphi}\in\tilde{\mathcal{B}}_{i,\delta}\colon|\mathcal{\tilde{A}}(\varphi)-c_{G}|\geq\varepsilon\}\ .

where by Cl(~i,δ/2)\mathrm{Cl}(\tilde{\mathcal{B}}_{i,\delta/2}) we denote the closure of the ball ~i,δ/2\tilde{\mathcal{B}}_{i,\delta/2}. Write η(γ~i)=η({1}×γ~i)\eta(\tilde{\gamma}_{i})=\eta(\{1\}\times\tilde{\gamma}_{i}), which satisfies

𝒜~(η(γ~i))cGε\mathcal{\tilde{A}}(\eta(\tilde{\gamma}_{i}))\leq c_{G}-\varepsilon

and

η(γ~i)(ti)=j(ei)η(γ~i)(ti+)=j(ei+).\eta(\tilde{\gamma}_{i})(t_{i}^{-})=j(e_{i}^{-})\qquad\qquad\eta(\tilde{\gamma}_{i})(t_{i}^{+})=j(e_{i}^{+}).

Now, for the extension operator EE is well defined on ~i,δ\tilde{\mathcal{B}}_{i,\delta} (shrinking δ\delta if necessary) and η(γ~i)~i,δ\eta(\tilde{\gamma}_{i})\subset\tilde{\mathcal{B}}_{i,\delta}, we can define the path E(η(γ~i))D1,2E(\eta(\tilde{\gamma}_{i}))\subset D^{1,2} which, by construction, satisfies

𝒜(E(η(γ~i)))cGε\mathcal{A}(E(\eta(\tilde{\gamma}_{i})))\leq c_{G}-\varepsilon

and

η(γ~i)(ti)=eiη(γ~i)(ti+)=ei+.\eta(\tilde{\gamma}_{i})(t_{i}^{-})=e_{i}^{-}\qquad\qquad\eta(\tilde{\gamma}_{i})(t_{i}^{+})=e_{i}^{+}.

The proposition is therefore proved for, if none of the points φ~i\tilde{\varphi}_{i} posses a local mountain pass structure, the continuous path γD1,2\gamma\subset D^{1,2} defined by gluing (in the obvious way) the segments γε1ikγε|[ti,ti+]\gamma_{\varepsilon}\setminus\bigcup_{1\leq i\leq k}\gamma_{\varepsilon}|_{[t_{i}^{-},t_{i}^{+}]} with the segments E(ηi(γ~i))E(\eta_{i}(\tilde{\gamma}_{i})) satisfies 𝒜(γ)cε\mathcal{A}(\gamma)\leq c-\varepsilon, a contradiction. ∎

Proposition 5.8 entails a non degeneracy condition for the intersection of the invariant manifolds W+(γ)W^{+}(\gamma_{\infty}) and W(γ)W^{-}(\gamma_{\infty}) at the homoclinic orbit associated with φ~\tilde{\varphi}_{*}. We now make use of topological degree theory to exploit this non degeneracy condition. Let φ~H1([n,n])\tilde{\varphi}_{*}\in H^{1}([-n,n]) be the critical point obtained in Proposition 5.8 and consider a sufficiently small neighbourhood U~H1([n,n])\widetilde{U}\in H^{1}([-n,n]) such that φ~U~\tilde{\varphi}_{*}\in\widetilde{U}. By definition of the functional 𝒜~\widetilde{\mathcal{A}}, and the fact that

mins[n,n]r0(s)+φ~(s)>0\min_{s\in[-n,n]}r_{0}(s)+\tilde{\varphi}_{*}(s)>0

the differential d𝒜~(φ~):U~H1([n,n])\mathrm{d}\widetilde{\mathcal{A}}(\tilde{\varphi}):\widetilde{U}\to H^{1}([-n,n]) is a continuous linear functional and, for any φ~U~\tilde{\varphi}\in\widetilde{U} and any ψH1([n,n])\psi\in H^{1}([-n,n]), we can express

(5.7) d𝒜~(φ)[ψ]=φ˙,ψ˙L2([n,n])+2nnφψr03+P(φ)[ψ],\mathrm{d}\mathcal{\widetilde{A}}(\varphi)[\psi]=\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}([-n,n])}+2\int_{-n}^{n}\frac{\varphi\psi}{r_{0}^{3}}+P(\varphi)[\psi],

where we have introduced the functional (compare expression (4.6) in the proof of Lemma 4.5)

P~(φ)[ψ]=nn(G2(r0+φ)3G0r03)ψnn(r0+φ((r0+φ)2+ρ2)3/21r02+2φr03)ψ(rS+((r0+φ~)(n))r˙0(n))ψ(n)+(rS((r0+φ~)(n))r˙0(n))ψ(n).\begin{split}\widetilde{P}(\varphi)[\psi]=&\int_{-n}^{n}\left(\frac{G^{2}}{(r_{0}+\varphi)^{3}}-\frac{G_{0}}{r_{0}^{3}}\right)\psi-\int_{-n}^{n}\left(\frac{r_{0}+\varphi}{((r_{0}+\varphi)^{2}+\rho^{2})^{3/2}}-\frac{1}{r_{0}^{2}}+\frac{2\varphi}{r_{0}^{3}}\right)\psi\\ -&\left(\partial_{r}S^{+}((r_{0}+\tilde{\varphi})(n))-\dot{r}_{0}(n)\right)\psi(n)+\left(\partial_{r}S^{-}((r_{0}+\tilde{\varphi})(-n))-\dot{r}_{0}(-n)\right)\psi(-n).\end{split}

Since r0(s)>0r_{0}(s)>0 and the interval [n,n][-n,n] is compact, the expression

φ,ψ=φ˙,ψ˙L2([n,n])+2nnφψr03,\langle\langle\varphi,\psi\rangle\rangle=\langle\dot{\varphi},\dot{\psi}\rangle_{L^{2}([-n,n])}+2\int_{-n}^{n}\frac{\varphi\psi}{r_{0}^{3}},

defines an equivalent inner product on H1([n,n])H^{1}([-n,n]). For φ~U~\tilde{\varphi}\in\widetilde{U}, denote by 𝒜~(φ~)\nabla\widetilde{\mathcal{A}}(\tilde{\varphi}) the unique element of H1([n,n])H^{1}([-n,n]) such that for all ψH1([n,n])\psi\in H^{1}([-n,n])

(5.8) 𝒜~(φ~),ψ=d𝒜~(φ)[ψ].\langle\langle\nabla\widetilde{\mathcal{A}}(\tilde{\varphi}),\psi\rangle\rangle=\mathrm{d}\mathcal{\widetilde{A}}(\varphi)[\psi].

From (5.7) one easily deduces that the map 𝒜~:U~H1([n,n])\nabla\widetilde{\mathcal{A}}:\widetilde{U}\to H^{1}([-n,n]) is a compact perturbation of the identity. Therefore, for any subset V~U~H1([n,n])\widetilde{V}\subset\widetilde{U}\in H^{1}([-n,n]) and any point z~H1([n,n])\tilde{z}\in H^{1}([-n,n]) such that z~𝒜~(V~)\tilde{z}\notin\nabla\widetilde{\mathcal{A}}(\partial\widetilde{V}) the Leray-Schauder degree 666The Leray-Schauder degree is a generalization of the Brouwer degree to maps between infinite dimensional spaces wich are of the form identity+compact. Details about its definition and properties can be found in [Kie12]. associated with the triple (𝒜~,V~,z~)(\nabla\widetilde{\mathcal{A}},\widetilde{V},\tilde{z}), which we denote by

deg(𝒜~,V~,z~),\mathrm{deg}(\nabla\widetilde{\mathcal{A}},\widetilde{V},\tilde{z}),

is well defined. Proposition 5.8, together with Theorem 2 in [Hof86], imply the following result.

Proposition 5.9.

Let φ~H1([n,n])\tilde{\varphi}_{*}\in H^{1}([-n,n]) be the critical point of 𝒜~\widetilde{\mathcal{A}} which was obtained in Proposition 5.8 and, for ε>0\varepsilon>0, denote by Bε(φ~)H1([n,n])B_{\varepsilon}(\tilde{\varphi}_{*})\subset H^{1}([-n,n]) the ball of radius ε\varepsilon centered at φ~\tilde{\varphi}_{*}. Then, there exists ε0\varepsilon_{0} such that for all 0εε00\leq\varepsilon\leq\varepsilon_{0},

deg(A~,Bε(φ~),0)=1.\mathrm{deg}(\nabla{\widetilde{A}},B_{\varepsilon}(\tilde{\varphi}_{*}),0)=-1.

As a consequence of Proposition 5.9 we can prove that the manifolds W+(γ;G)W^{+}(\gamma_{\infty};G) and W(γ;G)W^{-}(\gamma_{\infty};G) intersect transversally for G𝒢G\in\mathcal{G}. First, we introduce some notation which will be useful in the proof of Proposition 5.10 and in Section 6. Let s[n,n]s\in[-n,n], then, we denote by evs:H1([n,n])\mathrm{ev}_{s}:H^{1}([-n,n])\to\mathbb{R} the evaluation operator given by

evsφ~=φ~(s).\mathrm{ev}_{s}\tilde{\varphi}=\tilde{\varphi}(s).

In addition we denote by 𝚎𝚟sH1([n,n])\mathtt{ev}_{s}\in H^{1}([-n,n]) the unique element such that for all ψH1([n,n])\psi\in H^{1}([-n,n])

𝚎𝚟s,ψ=evs(ψ).\langle\langle\mathtt{ev}_{s},\psi\rangle\rangle=\mathrm{ev}_{s}(\psi).
Proposition 5.10.

For all G𝒥G\in\mathcal{J} there exists a topologically transverse intersection between W+(γ;G)W^{+}(\gamma_{\infty};G) and W(γ;G)W^{-}(\gamma_{\infty};G).

Proof.

Let φ~H1([n,n])\tilde{\varphi}_{*}\subset H^{1}([-n,n]) be the critical point obtained in Proposition 5.8. Then, there exists ε0>0\varepsilon_{0}>0 such that for all 0εε00\leq\varepsilon\leq\varepsilon_{0}

𝒜~(φ~)=0and𝒜~(φ~)0φBε(φ){φ}.\nabla{\tilde{\mathcal{A}}}(\tilde{\varphi}_{*})=0\qquad\qquad\text{and}\qquad\qquad\nabla{\tilde{\mathcal{A}}}(\tilde{\varphi})\neq 0\qquad\forall\varphi\in B_{\varepsilon}(\varphi_{*})\setminus\{\varphi_{*}\}.

In particular, there exists δ0>0\delta_{0}>0 such that

supφBε(φ)𝒜~(φ~)δ0.\sup_{\varphi\in\partial B_{\varepsilon}(\varphi_{*})}\lVert\nabla{\tilde{\mathcal{A}}}(\tilde{\varphi})\rVert\geq\delta_{0}.

Define now, for δ\delta\in\mathbb{R}, the one parameter family of maps Fδ:H1([n,n])H1([n,n])F_{\delta}:H^{1}([-n,n])\to H^{1}([-n,n]) given by

Fδ(φ)=𝒜~(φ~)+δ𝚎𝚟n=(nnren(φ,φ˙,s))+rS((r0+φ)(n))𝚎𝚟n(rS+((r0+φ)(n))+δ)𝚎𝚟nF_{\delta}(\varphi)=\nabla{\tilde{\mathcal{A}}}(\tilde{\varphi}_{*})+\delta\mathtt{ev}_{n}=\nabla\left(\int_{-n}^{n}\mathcal{L}_{\mathrm{ren}}(\varphi,\dot{\varphi},s)\right)+\partial_{r}S^{-}((r_{0}+\varphi)(-n))\mathtt{ev}_{-n}-(\partial_{r}S^{+}((r_{0}+\varphi)(n))+\delta)\mathtt{ev}_{n}

Then, it is possible to find δ1>0\delta_{1}>0 such that Fδ(φ)F_{\delta}(\varphi) is an admissible homotopy for δ[δ1,δ1]\delta\in[-\delta_{1},\delta_{1}] so by invariance of the degree under admissible homotopies

deg(Fδ,Bε(φ),0)=1δ[δ1,δ1].\mathrm{deg}(F_{\delta},B_{\varepsilon}(\varphi_{*}),0)=-1\qquad\qquad\forall\delta\in[-\delta_{1},\delta_{1}].

We now show how this implies the desired conclusion. Let Q={φBε:Fδ(φ)=0,δ[δ1,δ1]}Q=\{\varphi\in B_{\varepsilon}\colon F_{\delta}(\varphi)=0,\ \delta\in[-\delta_{1},\delta_{1}]\}. Then, denoting by πr,πy\pi_{r},\pi_{y} the projections onto the r,yr,y coordinates of a point (r,y,t)2×𝕋(r,y,t)\in\mathbb{R}^{2}\times\mathbb{T}, and by ϕs\phi^{s} the flow at time ss associated to Hamiltonian (1.2) we have that

[δ1,δ1]{πyϕH2n(r,rS(r,n),n)rS+(πrϕH2n(r,rS(r,n),n):rRδ1}[-\delta_{1},\delta_{1}]\subset\{\pi_{y}\circ\phi^{2n}_{H}(r,\partial_{r}S^{-}(r,-n),-n)-\partial_{r}S^{+}(\pi_{r}\circ\phi^{2n}_{H}(r,\partial_{r}S^{-}(r,-n),-n)\colon r\in R_{\delta_{1}}\}

for Rδ1{r=(r0+φ)(n):φQ}R_{\delta_{1}}\{r=(r_{0}+\varphi)(-n)\colon\varphi\in Q\}. This completes the proof. ∎

6. Construction of multibump solutions

We now show how Proposition 5.9 together with the parabolic lambda Lemma 3.2 can be used to the deduce the existence of homoclinic orbits to γ\gamma_{\infty} which perform any arbitrary number of “bumps”. We start by stating the following lemma, which is nothing but a reformulation of the parabolic lambda Lemma 3.3.

Lemma 6.1.

There exists RR large enough such that for R0,R1RR_{0},R_{1}\geq R there exists TT_{*} such that for all TTT\geq T_{*} there exists a unique orbit r^(t;T,R0,R1)\hat{r}(t;T,R_{0},R_{1}) of (1.1) for which r^(0)=R0\hat{r}(0)=R_{0} and r^(T)=R1\hat{r}(T)=R_{1}. Moreover, for all ε>0\varepsilon>0 there exists T(ε)T_{**}(\varepsilon) such that for all TTT\geq T_{**} the unique solution r^(t;T,R0,R1)\hat{r}(t;T,R_{0},R_{1}) satisfies

rS+(R0)r^˙(0)εr^˙(T)rS(R1)ε.\partial_{r}S^{+}(R_{0})-\dot{\hat{r}}(0)\leq\varepsilon\qquad\qquad\dot{\hat{r}}(T)-\partial_{r}S^{-}(R_{1})\leq\varepsilon.

Given R0,R1RR_{0},R_{1}\geq R and TTT\geq T_{*} we denote by

v+(T,R0,R1)=r^˙(0;T,R0,R1)v(T,R0,R1)=r^˙(T;T,R0,R1).v^{+}(T,R_{0},R_{1})=\dot{\hat{r}}(0;T,R_{0},R_{1})\qquad\qquad v^{-}(T,R_{0},R_{1})=\dot{\hat{r}}(T;T,R_{0},R_{1}).

where r^(t;T,R0,R1)\hat{r}(t;T,R_{0},R_{1}) is the orbit segment found in Lemma 6.1.

6.1. Proof of Theorem 1.5

We are now ready to build the multibump solutions. By proposition 5.9 we know that there exists a critical point φ~H1([n,n])\tilde{\varphi}_{*}\in H^{1}([-n,n]) of 𝒜~\widetilde{\mathcal{A}} and ε00\varepsilon_{0}\geq 0 such that for all 0εε00\leq\varepsilon\leq\varepsilon_{0},

deg(A~,Bε(φ~),0)=1,\mathrm{deg}(\nabla{\widetilde{A}},B_{\varepsilon}(\tilde{\varphi}_{*}),0)=-1,

where Bε(φ~)H1([n,n])B_{\varepsilon}(\tilde{\varphi}_{*})\subset H^{1}([-n,n]) stands for the ball of radius ε\varepsilon centered at φ~\tilde{\varphi}_{*}. For any LL\in\mathbb{N}, introduce now the map

(6.1) F:(Bε(φ))L+1×({l:lT})L\displaystyle F:(B_{\varepsilon}(\varphi_{*}))^{L+1}\times(\{l\in\mathbb{N}\colon l\geq T_{**}\})^{L} (H1([n,n]))L+1\displaystyle\longrightarrow(H^{1}([-n,n]))^{L+1}
(φ1,,φL+1,l1,,lL)\displaystyle(\varphi_{1},\dots,\varphi_{L+1},l_{1},\dots,l_{L}) (F1,,FL+1)\displaystyle\longmapsto(F_{1},\dots,F_{L+1})

where the maps FjF_{j}, 1jL+11\leq j\leq L+1 are given by

F1=𝒜~G+(rS+(φ1(n),n)v+(l1,φ1(n),φ2(n)))𝚎𝚟nFL+1=𝒜~G+(v(lL,φL(n),φL+1(n))rS(φL+1(n),n))𝚎𝚟n\begin{split}F_{1}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(\partial_{r}S^{+}(\varphi_{1}(n),n)-v^{+}(l_{1},\varphi_{1}(n),\varphi_{2}(-n))\right)\mathtt{ev}_{n}\\ F_{L+1}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(v^{-}(l_{L},\varphi_{L}(n),\varphi_{L+1}(-n))-\partial_{r}S^{-}(\varphi_{L+1}(-n),-n)\right)\mathtt{ev}_{n}\\ \end{split}

and for 2jL2\leq j\leq L (this set is empty for L=1L=1)

Fj=𝒜~G+(rS+(φj(n),n)v+(lj,φj(n),φj+1(n)))𝚎𝚟n+(v(lj1,φj1(n),φj(n))rS((φj)(n),n))𝚎𝚟n.\begin{split}F_{j}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(\partial_{r}S^{+}(\varphi_{j}(n),n)-v^{+}(l_{j},\varphi_{j}(n),\varphi_{j+1}(-n))\right)\mathtt{ev}_{n}\\ &+\left(v^{-}(l_{j-1},\varphi_{j-1}(n),\varphi_{j}(-n))-\partial_{r}S^{-}((\varphi_{j})(-n),-n)\right)\mathtt{ev}_{-n}.\end{split}

The proof of the following result follows inmediately after from Proposition 5.9 and Lemma 6.1.

Theorem 6.2.

There exists φ~H1([n,n])\tilde{\varphi}_{*}\in H^{1}([-n,n]), T>0T>0 and ε>0\varepsilon>0 such that for all LL\in\mathbb{N}

deg(F,(Bε(φ))L+1×({l:lT}L,0)=(1)L\mathrm{deg}(F,(B_{\varepsilon}(\varphi_{*}))^{L+1}\times(\{l\in\mathbb{N}\colon l\geq T\}^{L},0)=(-1)^{L}

In particular, for any sequence 𝐥={lj}1jL({l:lT}L\mathbf{l}=\{l_{j}\}_{1\leq j\leq L}\subset(\{l\in\mathbb{N}\colon l\geq T\}^{L} there exists 𝛗(𝐥)={φj(𝐥)}1jL+1(H1([n,n]))L+1\boldsymbol{\varphi}(\mathbf{l})=\{\varphi_{j}(\mathbf{l})\}_{1\leq j\leq L+1}\subset(H^{1}([-n,n]))^{L+1} such that

F(𝝋(𝐥),𝐥)=0.F(\boldsymbol{\varphi}(\mathbf{l}),\mathbf{l})=0.

Theorem 6.2 shows the truth of the first item in Theorem 1.5 for sequences σ{0,1}\sigma\in\{0,1\}^{\mathbb{Z}} with finite, but arbitrarily large number of nonzero entries. For the time interval TT_{**} in the definition of (6.1) does not depend on LL, the existence of solutions rσr_{\sigma} such that σ\sigma has infinitely many non-zero entries follows by a standard diagonal argument in the Cloc1C^{1}_{\mathrm{loc}} topology. In order to deduce the second item, namely the existence of infinitely many periodic orbits, we define the functional

(6.2) Fper:(Bε(φ))L×({l:lT})L\displaystyle F_{\mathrm{per}}:(B_{\varepsilon}(\varphi_{*}))^{L}\times(\{l\in\mathbb{N}\colon l\geq T_{**}\})^{L} (H1([n,n]))L\displaystyle\longrightarrow(H^{1}([-n,n]))^{L}
(φ1,,φL+1,l1,,lL)\displaystyle(\varphi_{1},\dots,\varphi_{L+1},l_{1},\dots,l_{L}) (F1,,FL)\displaystyle\longmapsto(F_{1},\dots,F_{L})

with periodic boundary conditions

F1=𝒜~G+(rS+(φ1(n),n)v+(l1,φ1(n),φ2(n)))𝚎𝚟n+(v(lL,φL(n),φ1(n))rS((φ1)(n),n))𝚎𝚟nFL=𝒜~G+(v(lL,φL(n),φL+1(n))rS(φL+1(n),n))𝚎𝚟n+(v+(L,φL(n),φ1(n))rS+((φL)(n),n))𝚎𝚟n\begin{split}F_{1}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(\partial_{r}S^{+}(\varphi_{1}(n),n)-v^{+}(l_{1},\varphi_{1}(n),\varphi_{2}(-n))\right)\mathtt{ev}_{n}\\ &+\left(v^{-}(l_{L},\varphi_{L}(n),\varphi_{1}(-n))-\partial_{r}S^{-}((\varphi_{1})(-n),-n)\right)\mathtt{ev}_{-n}\\ F_{L}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(v^{-}(l_{L},\varphi_{L}(n),\varphi_{L+1}(-n))-\partial_{r}S^{-}(\varphi_{L+1}(-n),-n)\right)\mathtt{ev}_{n}\\ &+\left(v^{+}(L,\varphi_{L}(n),\varphi_{1}(-n))-\partial_{r}S^{+}((\varphi_{L})(-n),-n)\right)\mathtt{ev}_{-n}\end{split}

and such that for 2jL2\leq j\leq L (this set is empty for L=1L=1) FjF_{j} has the same expression as in in the non periodic case. The proof of Theorem 1.5 is complete.

6.2. Proof of Theorem 1.4

From the proof of Theorem 1.5 it follows that

X+YX^{+}\cap Y^{-}\neq\emptyset

for any possible combination of X+{P+,B+,OS+}X^{+}\in\{P^{+},B^{+},OS^{+}\} and Y{P,B,OS}Y^{-}\in\{P^{-},B^{-},OS^{-}\}. We now show that H+PH^{+}\cap P^{-}\neq\emptyset (the proof for the other combinations being similar). The following result is implied by the second part of Lemma 3.3.

Lemma 6.3.

There exists RR large enough and ε0\varepsilon_{0} such that for all RR0R\geq R_{0} and all 0εε00\leq\varepsilon\leq\varepsilon_{0} there exists a unique orbit r^(s;,R0,δ)\hat{r}(s;,R_{0},\delta) of (1.1) for which

r^(0)=R0,r^˙(0)=rS+(R0,0)+ε.\hat{r}(0)=R_{0},\qquad\qquad\qquad\dot{\hat{r}}(0)=\partial_{r}S^{+}(R_{0},0)+\varepsilon.

Moreover, r^(s;,R0,δ)\hat{r}(s;,R_{0},\delta) is defined for all s0s\geq 0 and satisfies

limsr^(s;R0,ε)=limsr^˙(s;R0,ε)>0\lim_{s\to\infty}\hat{r}(s;R_{0},\varepsilon)=\infty\qquad\qquad\qquad\lim_{s\to\infty}\dot{\hat{r}}(s;R_{0},\varepsilon)>0

Given R0RR_{0}\geq R and ε>0\varepsilon>0 we denote by

vhyp+(R0,ε)=r^˙(0;R0,ε)v^{+}_{\mathrm{hyp}}(R_{0},\varepsilon)=\dot{\hat{r}}(0;R_{0},\varepsilon)

where r^(s;R0,ε)\hat{r}(s;R_{0},\varepsilon) is the orbit segment found in Lemma 6.3. Fix 0εε00\leq\varepsilon\leq\varepsilon_{0}. The fact that H+PH^{+}\cap P^{-}\neq\emptyset follows from the fact that the functional

Fhyp:(Bε(φ))L+1×({l:lT})L\displaystyle F_{\mathrm{hyp}}:(B_{\varepsilon}(\varphi_{*}))^{L+1}\times(\{l\in\mathbb{N}\colon l\geq T_{**}\})^{L} (H1([n,n]))L+1\displaystyle\longrightarrow(H^{1}([-n,n]))^{L+1}
(φ1,,φL+1,l1,,lL)\displaystyle(\varphi_{1},\dots,\varphi_{L+1},l_{1},\dots,l_{L}) (F1,hyp,,FL+1,hyp)\displaystyle\longmapsto(F_{1,\mathrm{hyp}},\dots,F_{L+1,\mathrm{hyp}})

where the maps Fj,hypF_{j,\mathrm{hyp}}, 1jL+11\leq j\leq L+1 are given by

F1,hyp=𝒜~G+(rS+(φ1(n),n)v+(l1,φ1(n),φ2(n)))𝚎𝚟nFL+1,hyp=𝒜~G+(v(lL,φL(n),φL+1(n))rS(φL+1(n),n))𝚎𝚟n+(rS+(φL+1(n),n)vhyp+(φL+1(n),ε))𝚎𝚟n\begin{split}F_{1,\mathrm{hyp}}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(\partial_{r}S^{+}(\varphi_{1}(n),n)-v^{+}(l_{1},\varphi_{1}(n),\varphi_{2}(-n))\right)\mathtt{ev}_{n}\\ F_{L+1,\mathrm{hyp}}=&\nabla\mathcal{\widetilde{A}}_{G}+\left(v^{-}(l_{L},\varphi_{L}(n),\varphi_{L+1}(-n))-\partial_{r}S^{-}(\varphi_{L+1}(-n),-n)\right)\mathtt{ev}_{n}\\ &+\left(\partial_{r}S^{+}(\varphi_{L+1}(n),n)-v^{+}_{\mathrm{hyp}}(\varphi_{L+1}(n),\varepsilon)\right)\mathtt{ev}_{n}\end{split}

and for 2jL2\leq j\leq L (this set is empty for L=1L=1)

Fj,hyp=𝒜~Gi+(rS+(φj(n),n)v+(lj,φj(n),φj+1(n)))𝚎𝚟n+(v(lj1,φj1(n),φj(n))rS((φj)(n),n))𝚎𝚟n.\begin{split}F_{j,\mathrm{hyp}}=&\nabla\mathcal{\widetilde{A}}_{G}i+\left(\partial_{r}S^{+}(\varphi_{j}(n),n)-v^{+}(l_{j},\varphi_{j}(n),\varphi_{j+1}(-n))\right)\mathtt{ev}_{n}\\ &+\left(v^{-}(l_{j-1},\varphi_{j-1}(n),\varphi_{j}(-n))-\partial_{r}S^{-}((\varphi_{j})(-n),-n)\right)\mathtt{ev}_{-n}.\end{split}

In order to prove that P+OSP^{+}\cap OS^{-}\neq\emptyset we take LL\to\infty and argue as in the proof of Theorem 1.5. In order to show that H+BH^{+}\cap B^{-} we impose periodic boundary conditions. The proof of Theorem 1.4 is complete.

Appendix A Proof of of the technical claims in Lemma 5.5

We first prove the following result, which will be needed for the proof of Lemma A.2.

Lemma A.1.

Let S0(r;G)S_{0}(r;G) be the generating function defined in Lemma 2.2. Then, for any G,GG,G_{*}\in\mathbb{R} we have that

|S0(r;G)S0(r;G)||G2G2|r1/2.|S_{0}(r;G)-S_{0}(r;G_{*})|\lesssim\frac{|G^{2}-G_{*}^{2}|}{r^{1/2}}.
Proof.

Denote by u~(r)\tilde{u}(r) the unique function such that, for all y<0y<0, the yy component of the parametrization (2.2) is given by yh(u~(r))y_{h}(\tilde{u}(r)). Writing ΔS(r;G,G)=S0(r;G)S0(r;G)\Delta S(r;G,G_{*})=S_{0}(r;G)-S_{0}(r;G_{*}) we observe that ΔS(r;G,G)\Delta S(r;G,G_{*}) satisfies

yh(u~(r))rΔS+G2G22r2+(rΔS)22=0y_{h}(\tilde{u}(r))\partial_{r}\Delta S+\frac{G^{2}-G_{*}^{2}}{2r^{2}}+\frac{(\partial_{r}\Delta S)^{2}}{2}=0

Using now that yh(u~(r))r1/2y_{h}(\tilde{u}(r))\sim r^{-1/2} for large r1r\gg 1 we obtain that

rΔSG2G2r3/2+𝒪(r1/2(rΔS)2)\partial_{r}\Delta S\sim\frac{G^{2}-G_{*}^{2}}{r^{3/2}}+\mathcal{O}(r^{1/2}(\partial_{r}\Delta S)^{2})

so

|ΔS(r;G,G))||G2G2|r1/2|\Delta S(r;G,G_{*}))|\lesssim\frac{|G^{2}-G_{*}^{2}|}{r^{1/2}}

as was to be shown. ∎

The claims in Lemma 5.5 follow from the following result.

Lemma A.2.

Suppose that for G[G,G]G\in[-G_{*},G_{*}] there exists an orbit r(s;G):+r(s;G):\mathbb{R}\to\mathbb{R}_{+} of the Hamiltonian HGH_{G} in(1.2) which is homoclinic to γ\gamma_{\infty} and, for some G0[G,G]G_{0}\in[-G_{*},G_{*}] satisfies

|r(s;G)r0(s,G0)|s1/2ass±.|r(s;G)-r_{0}(s,G_{0})|\lesssim s^{1/2}\qquad\qquad\text{as}\qquad s\to\pm\infty.

Then,

|rS+(r(s;G),G)rS0(r0(s;G0),G0)|s5/6ass±,|\partial_{r}S^{+}(r(s;G),G)-\partial_{r}S_{0}(r_{0}(s;G_{0}),G_{0})|\lesssim s^{-5/6}\qquad\qquad\text{as}\qquad s\to\pm\infty,

and, in particular

lims±S±(r(s;G),s;G)S0(r0(s;G0),G0)=0.\lim_{s\to\pm\infty}S^{\pm}(r(s;G),s;G)-S^{0}(r_{0}(s;G_{0}),G_{0})=0.
Proof.

We write

rS+(r(s;G),G)rS0(r0(s;G0),G0)=(rS+(r(s;G),G)rS0(r(s;G),G))+(rS0(r(s;G),G)rS0(r0(s;G0),G))+(rS0(r0(s;G0),G)rS0(r0(s;G0),G0))=E1+E2+E3.\begin{split}\partial_{r}S^{+}(r(s;G),G)-\partial_{r}S_{0}(r_{0}(s;G_{0}),G_{0})=&\left(\partial_{r}S^{+}(r(s;G),G)-\partial_{r}S_{0}(r(s;G),G)\right)\\ &+\left(\partial_{r}S^{0}(r(s;G),G)-\partial_{r}S_{0}(r_{0}(s;G_{0}),G)\right)\\ &+\left(\partial_{r}S^{0}(r_{0}(s;G_{0}),G)-\partial_{r}S_{0}(r_{0}(s;G_{0}),G_{0})\right)\\ =&E_{1}+E_{2}+E_{3}.\end{split}

On one hand, it follows from the last item in Lemma 3.1 that as s±s\to\pm\infty

|E1|r5/2(s;G)r05/2(s;G)s5/3.|E_{1}|\lesssim r^{-5/2}(s;G)\lesssim r_{0}^{-5/2}(s;G)\lesssim s^{-5/3}.

On the other hand, it follows from the mean value theorem, the definition of S0(r;G)S^{0}(r;G) and the hypothesis in the statement of the lemma that as s±s\to\pm\infty

|E2|suprI|rr2S0(r;G)||r(s;G)r0(s,G0)|r02(s;G0)|r(s;G)r0(s,G0)|s5/6|E_{2}|\lesssim\sup_{r\in I}|\partial_{rr}^{2}S^{0}(r;G)||r(s;G)-r_{0}(s,G_{0})|\lesssim r_{0}^{-2}(s;G_{0})|r(s;G)-r_{0}(s,G_{0})|\lesssim s^{-5/6}

for I={r+:r=λr0(s;G0)+(1λ)r(s;G),λ[0,1]}I=\{r\in\mathbb{R}_{+}\colon r=\lambda r_{0}(s;G_{0})+(1-\lambda)r(s;G),\ \lambda\in[0,1]\}. Also, from Lemma 2.2 we deduce that

|E3|r3/2s1.|E_{3}|\lesssim r^{-3/2}\lesssim s^{-1}.

The proof of the first item follows combining the estimates for E1,E2,E3E_{1},E_{2},E_{3} and integrating. The second part follows from the obtained estimate and straightforward computations. ∎

References

  • [Ale69] V. M. Alekseev. Quasirandom dynamical systems. mat. Zametki, 6:489–498, 1969.
  • [AR73] Antonio Ambrosetti and Paul H. Rabinowitz. Dual variational methods in critical point theory and applications. J. Functional Analysis, 14:349–381, 1973.
  • [BDFT21] Alberto Boscaggin, Walter Dambrosio, Guglielmo Feltrin, and Susanna Terracini. Parabolic orbits in celestial mechanics: a functional-analytic approach. Proc. Lond. Math. Soc. (3), 123(2):203–230, 2021.
  • [BF04] I. Baldomá and E. Fontich. Stable manifolds associated to fixed points with linear part equal to identity. J. Differential Equations, 197(1):45–72, 2004.
  • [CGM+21] Maciej J Capiński, Marcel Guardia, Pau Martín, Tere Seara, and Piotr Zgliczyński. Oscillatory motions and parabolic manifolds at infinity in the planar circular restricted three body problem. arXiv preprint arXiv:2106.06254, 2021.
  • [Cha22] Jean Chazy. Sur l’allure du mouvement dans le problème des trois corps quand le temps croît indéfiniment. Ann. Sci. École Norm. Sup. (3), 39:29–130, 1922.
  • [CZES90] Vittorio Coti Zelati, Ivar Ekeland, and Éric Séré. A variational approach to homoclinic orbits in Hamiltonian systems. Math. Ann., 288(1):133–160, 1990.
  • [CZR91] Vittorio Coti Zelati and Paul H. Rabinowitz. Homoclinic orbits for second order Hamiltonian systems possessing superquadratic potentials. J. Amer. Math. Soc., 4(4):693–727, 1991.
  • [GK11] Joseph Galante and Vadim Kaloshin. Destruction of invariant curves in the restricted circular planar three-body problem by using comparison of action. Duke Math. J., 159(2):275–327, 2011.
  • [GK12] A. Gorodetski and V Kaloshin. Hausdorff dimension of oscillatory motions for restricted three body problems. Preprint, available at http://www.terpconnect.umd.edu/ vkaloshi, 2012.
  • [GMPS22] Marcel Guardia, Pau Martín, Jaime Paradela, and Tere M. Seara. Hyperbolic dynamics and oscillatory motions in the 3 body problem, 2022.
  • [GMS16] Marcel Guardia, Pau Martín, and Tere M. Seara. Oscillatory motions for the restricted planar circular three body problem. Invent. Math., 203(2):417–492, 2016.
  • [GPSV21] Marcel Guardia, Jaime Paradela, Tere M Seara, and Claudio Vidal. Symbolic dynamics in the restricted elliptic isosceles three body problem. Journal of Differential Equations, 294:143–177, 2021.
  • [GSMS17] Marcel Guardia, Tere M. Seara, Pau Martín, and Lara Sabbagh. Oscillatory orbits in the restricted elliptic planar three body problem. Discrete Contin. Dyn. Syst., 37(1):229–256, 2017.
  • [Hof86] Helmut Hofer. The topological degree at a critical point of mountain-pass type. In Nonlinear functional analysis and its applications, Part 1 (Berkeley, Calif., 1983), volume 45 of Proc. Sympos. Pure Math., pages 501–509. Amer. Math. Soc., Providence, RI, 1986.
  • [Jea99] Louis Jeanjean. On the existence of bounded Palais-Smale sequences and application to a Landesman-Lazer-type problem set on 𝐑N{\bf R}^{N}. Proc. Roy. Soc. Edinburgh Sect. A, 129(4):787–809, 1999.
  • [Kie12] Hansjörg Kielhöfer. Bifurcation theory, volume 156 of Applied Mathematical Sciences. Springer, New York, second edition, 2012. An introduction with applications to partial differential equations.
  • [Koz83] V. V. Kozlov. Integrability and nonintegrability in Hamiltonian mechanics. Uspekhi Mat. Nauk, 38(1(229)):3–67, 240, 1983.
  • [LS80a] Jaume Llibre and Carles Simó. Oscillatory solutions in the planar restricted three-body problem. Mathematische Annalen, 248(2):153–184, 1980.
  • [LS80b] Jaume Llibre and Carlos Simó. Some homoclinic phenomena in the three-body problem. J. Differential Equations, 37(3):444–465, 1980.
  • [McG73] Richard McGehee. A stable manifold theorem for degenerate fixed points with applications to celestial mechanics. J. Differential Equations, 14:70–88, 1973.
  • [MNT99] Piero Montecchiari, Margherita Nolasco, and Susanna Terracini. A global condition for periodic Duffing-like equations. Trans. Amer. Math. Soc., 351(9):3713–3724, 1999.
  • [Moe07] R. Moeckel. Symbolic dynamics in the planar three-body problem. Regul. Chaotic Dyn., 12(5):449–475, 2007.
  • [Mos01] Jürgen Moser. Stable and random motions in dynamical systems. Princeton Landmarks in Mathematics. Princeton University Press, Princeton, NJ, 2001. With special emphasis on celestial mechanics, Reprint of the 1973 original, With a foreword by Philip J. Holmes.
  • [MP94] Regina Martínez and Conxita Pinyol. Parabolic orbits in the elliptic restricted three body problem. J. Differential Equations, 111(2):299–339, 1994.
  • [MV09] E. Maderna and A. Venturelli. Globally minimizing parabolic motions in the Newtonian NN-body problem. Arch. Ration. Mech. Anal., 194(1):283–313, 2009.
  • [MV20] Ezequiel Maderna and Andrea Venturelli. Viscosity solutions and hyperbolic motions: a new PDE method for the NN-body problem. Ann. Math. (2), 192(2):499–550, 2020.
  • [Sér92] Eric Séré. Existence of infinitely many homoclinic orbits in Hamiltonian systems. Math. Z., 209(1):27–42, 1992.
  • [Str88] Michael Struwe. The existence of surfaces of constant mean curvature with free boundaries. Acta Math., 160(1-2):19–64, 1988.
  • [Str00] Michael Struwe. Variational methods, volume 34 of Ergebnisse der Mathematik und ihrer Grenzgebiete. 3. Folge. A Series of Modern Surveys in Mathematics [Results in Mathematics and Related Areas. 3rd Series. A Series of Modern Surveys in Mathematics]. Springer-Verlag, Berlin, third edition, 2000. Applications to nonlinear partial differential equations and Hamiltonian systems.
  • [SZ20] Tere M. Seara and Jianlu Zhang. Oscillatory orbits in the restricted planar four body problem. Nonlinearity, 33(12):6985–7015, 2020.
  • [Xia94] Zhihong Xia. Arnold diffusion and oscillatory solutions in the planar three-body problem. J. Differential Equations, 110(2):289–321, 1994.