1 Introduction

In this paper we consider the second order Hamiltonian system

$$\begin{aligned} \ddot{q} + V'(q)=0 \end{aligned}$$
(HS)

where \(q:{\mathbb {R}}\longrightarrow {\mathbb {R}}^2\) and \(V\in C^2({\mathbb {R}}^2{\setminus }\{e\},\, {\mathbb {R}})\) has a singularity at a point \(e\not =0\) such that

$$\begin{aligned} \displaystyle V(q) \sim -\frac{1}{|q-e|^\alpha } \quad \hbox {as} \, \, q\rightarrow e \, \hbox { with} \,\, \alpha \in ]0,2[. \end{aligned}$$
(1)

We will assume that V has a strict global maximum at \(q=0\). So 0 is a equilibrium point for (HS).

Our goal in the first part is studying the existence of nontrivial homoclinic solutions to 0 of (HS), i.e. solutions q of (HS) such that

$$\begin{aligned} q\ne 0 \quad \hbox {and }\lim _{t\rightarrow \pm \infty } q(t)\equiv q(\pm \infty )=0=\dot{q} (\pm \infty ). \end{aligned}$$

We note that the order \(\alpha \) in (1) plays an important role and we consider the existence of homoclinic solutions of (HS) under weak force case (\(\alpha \in ]0,2[\)). This case has been studied in several works which deal via variational methods with the periodic problem. See, e.g., [1, 2, 5, 9, 15, 17]. Let us now define the strong force condition:

(SF)    There exists a neighbourhood \(\Omega \) of e in \({\mathbb {R}}^2\) and \(U\in C^1(\Omega \setminus \{e\},{\mathbb {R}})\) such that

$$\begin{aligned}&|U(q)|\rightarrow \infty&\hbox {as }\, q\rightarrow e,\\&-V(q)\ge |U'(q)|^2&\hbox { for all } q\in \Omega \setminus \{e\}. \end{aligned}$$

Condition (SF) was introduced by Gordon [10]. For a potential \(\displaystyle V(q) \sim -\frac{1}{|q-e|^\alpha } \) as \(q\rightarrow e,\) (SF) is satisfied if and only if \(\alpha \ge 2\). In fact, for \(\alpha \ge 2\) we can take \(U(q)=-2^{-1}\ln |q-e|\). The major role of (SF) is the following property.

Lemma 1.1

Assume (SF) and \(V(q)\rightarrow -\infty \) as \(q\rightarrow e\). Let \( a<b \in R\) and \((q_m)\subset H^1([a,b],\Omega \setminus \{e\})\) which converges weakly in \(H^1([a,b], {\mathbb {R}}^2)\) to q such that \(q(t_0)=e\) for some \(t_0\in [a,b]\). Then

$$\begin{aligned} -\int _a^b V(q_m)dt \longrightarrow +\infty \end{aligned}$$

(and therefore \(\displaystyle \int _a^b\Big [ {1\over 2} |\dot{q}_m|^2 -V(q_m)\Big ]dt \longrightarrow +\infty \)).

The proof of this lemma can be found in ([11], Lemma 2.1) or in [13]. As a consequence, if (SF) holds then functions with bounded energy are uniformly away from the singularity e. Therefore, in such case, a standard variational arguments in [13] provided the existence of a a pair of homoclinic orbits that wind respectively around the singularity e in a positive and negative sense. These solutions were obtained by minimizing the energy functional

$$\begin{aligned} I(q)=\displaystyle \int _{\mathbb {R}}\Big [\frac{1}{2}|\dot{q}|^2- V(q)\Big ]dt \end{aligned}$$

on classes of sets with a fixed winding number around e (see also [6, 7] for multiplicity results). If this condition is dropped (weak force case), Rabinowitz [13] proved the existence of a “generalized” homoclinic solution of (HS) which may pass through the singularity.

In \({\mathbb {R}}^N\) with \(N\ge 3\), the existence of homoclinic solutions of (HS) was proved in [16] for strong force potentials (see also [8] in the case of time periodic potentials) and [3, 14] for weak force potentials like (1.1). In [3, 14], the authors introduced a strong force perturbed potential \(V_\varepsilon \) for \(\varepsilon \in ]0,1]\) such that \(V_\varepsilon (q)= V(q)- \displaystyle {\varepsilon }/{|q-e|^2}\) near \(q=e\) and proved through a min-max method from Bahri-Rabinowitz [4] the existence of non-collision solutions for approximated differential problems. Then they passed to the limit as \(\varepsilon \rightarrow 0\) with the aid of appropriate estimates to obtain a generalized homoclinic solution. In [3] we studied the Morse index of approximated functionals at critical points to estimate the number of collisions. In particular we established the existence of non-collision homoclinic solution for \(\alpha \in ]1,2[\) i.e. \(q(t)\ne e\) for all \(t\in {\mathbb {R}}\), while in [14] this result is obtained by assuming that V(q) is radially symmetric near \(q=e\).

The main purpose of Sect. 2 is to prove the existence of non-collision homoclinic orbits of (HS) in \({\mathbb {R}}^2\) for weak force potentials. By exploiting the topology of the plane and using a minimization method, we first show the existence of a generalized homoclinic solution of (HS) as a limit of solutions of perturbed problems with boundary conditions. Then and for the regularity of this solution, we will use a Tanaka’s rescaling argument to prove some additional properties of approximated solutions near collisions, and we will prove how the generalized homoclinic solution obtained is actually a non-collision orbit in the case \(\alpha \in ]1,2[\).

In Sect. 3, we assume that V has another global maximum at infinity i.e. \(\displaystyle \lim _{|x|\rightarrow +\infty } V(x)=V(0)\) and we study the existence of a heteroclinic orbit “at infinity" i.e. a solution q of (HS) satisfying

$$\begin{aligned} q(-\infty )=0,\,\, |q(+\infty )|=+\infty \,\,\, \hbox { and } \, \,\dot{q}(\pm \infty )=0. \end{aligned}$$

The problem in \({\mathbb {R}}^N\) was treated by Serra in [14] for regular potentials where \(V(q)\sim \displaystyle -{a}/{|q|^b } \; \hbox {as } \, |q|\rightarrow +\infty \, \hbox { with } \, a,\; b >0.\) He also treated the case of singular potentials which behaves like (1) when \(N\ge 3\) and established the existence of non-collision orbits using some results from [15] on the analysis of collisions solutions of minimization problems. In the present paper we deal with the case \(N=2\) and we will perturb V near e with a strong force term to get the existence of sequence \((q_n)\) of heteroclinic orbits at infinity for perturbed problems. We obtain uniform estimates to show that \((q_n)\) converges to a generalized solution. Some local properties of \(q_n\) near collisions and the fact that \(q_n\) is obtained via a minimization procedure permit us to obtain a non-collision heteroclinic solution at infinity.

2 Existence of Homoclinic Orbits

In this section, we consider the existence of a homoclinic solution of (HS) where the potential V satisfies the following assumptions:

  1. (V1)

    \( V\in C^2 ({\mathbb {R}}^2\setminus \{e\},{\mathbb {R}})\) for some \(e\ne 0 \);

  2. (V2)

    \( V(q)< V(0)=0\) for all \( q\in {\mathbb {R}}^2\setminus \{0,e\}\);

  3. (V3)

    V is of the form

    $$\begin{aligned} \displaystyle V(q)=- \frac{1}{|q-e|^\alpha } + W(q), \end{aligned}$$

with \(\alpha \in ]0,2[\) and W is such that

$$\begin{aligned} {|}q-e|^{\alpha -\nu } W(q), \;\; |q-e|^{\alpha -\nu +1} W'(q)\;\; \hbox {and }\; |q-e|^{\alpha -\nu +2} W''(q) \longrightarrow 0\, \text{ as } \, q\rightarrow e \end{aligned}$$

for some \(\nu \in ]0,\alpha [\);

  1. (V4)

    There are \(R>2|e|\) and a function \(W_{\infty } \in C^1 ({\mathbb {R}}^2, {\mathbb {R}})\) such that

\(|W_\infty (q)|\longrightarrow +\infty \quad \hbox {as}\quad |q|\rightarrow + \infty \) and \(-V(q)\ge |W_\infty '(q)|\) for \(|q|\ge R\).

Remark 2.1

  1. i)

    The condition (V3) remains valid when \(\nu =0\). In particular it involves that \(V\sim \displaystyle -1/|q-e|^\alpha \) near \(q=e\) with \( \alpha \in ]0,2[\).

  2. ii)

    The condition (V4) concerns the behavior of the potential at infinity. It will be satisfied if for example \( V(q)\sim -a|q|^\beta \) as \(|q|\rightarrow +\infty \) where \( a>0 \) and \(\beta \ge -2\).

Our main result of this section is

Theorem 2.2

Assume (V1)-(V4).

  1. 1)

    If \(\alpha \in ]1,2[\), then (HS) possesses at least one non-collision homoclinic solution.

  2. 2)

    If \(\alpha \in ]0,1]\), then (HS) possesses a non trivial generalized homoclinic solution q having at most one collision. Moreover, if \(q(t_0)=e\) then q(t) is a collision brake orbit, i.e. \(q(t+t_0)=q(t_0-t)\) for all \( t \in {\mathbb {R}}\).

Here, similarly as in [4, 17] for the periodic problem, we define a generalized homoclinic solution as a continuous function \(q:{\mathbb {R}}\longrightarrow {\mathbb {R}}^2\) such that

  1. (i)

    \( \dot{q} \in L^2({\mathbb {R}},\, {\mathbb {R}}^2)\) and \(I(q)<\infty \);

  2. (ii)

    \(D=\{t\in {\mathbb {R}}, \, q(t)=e\}\) is a set of measure 0;

  3. (iii)

    \(q\in C^2 ({\mathbb {R}}\setminus D,\, {\mathbb {R}}^2) \quad \hbox {and satisfies (HS) on}\, {\mathbb {R}}\setminus D\);

  4. (iv)

    \(\displaystyle \frac{1}{2}|\dot{q}(t)|^2 + V(q(t)) =0\) for \( t \in {\mathbb {R}}{\setminus } D\);

  5. (v)

    \(q(t)\longrightarrow 0\quad \hbox {and}\quad \dot{q}(t)\longrightarrow 0 \,\, \hbox {as}\;\, t\rightarrow \pm \infty \).

If \(D=\emptyset \), q is a classical (non-collision) homoclinic solution.

Remark 2.3

Since V is independent of t, \(q(-t)\) is a homoclinic solution of (HS) whenever q(t) is a homoclinic solution.

The proof of Theorem 2.2. is divided in various steps. We shall construct a homoclinic solution of (HS) as a limit of solutions of approximate value problems. We started by modifying the potential V near e. For \(\varepsilon \in ]0,1]\), we define \(V_\varepsilon \in C^2({\mathbb {R}}^2{\setminus }\{e\},\, {\mathbb {R}})\) such that \(V_1\le V_\varepsilon \le V\) and

$$\begin{aligned} V_\varepsilon (q)= \left\{ \begin{array}{ll} V(q)-\displaystyle \frac{\varepsilon }{|q-e|^2} &{} \quad ~\text {if}~~\,\, 0<|q-e|\le |e|/4,\\ 0 &{} \quad ~\text {if}~~\,\, \displaystyle |q-e|\ge |e|/2.\\ \end{array} \right. \end{aligned}$$

Remark that \(\displaystyle V_\varepsilon (q)\sim -\frac{\varepsilon }{|q-e|^2}\; \hbox {as}\; q\rightarrow e\). So \(V_\varepsilon \) satisfies the strong force condition.

Let \((\varepsilon _n)_{n \in {\mathbb {N}}^*}\subset ]0,1]\) be a non-increasing sequence converging to 0. We consider for each \(n\in {\mathbb {N}}^*\) the Dirichlet boundary value problem

figure a

The corresponding functional is

$$\begin{aligned} I_{0,n}(q)=\displaystyle \int _{0}^{n}\Big [\frac{1}{2}|\dot{q}|^2 -V_{\varepsilon _n}(q)\Big ]dt \in C^1 (\Lambda _n, {\mathbb {R}}) \end{aligned}$$

where

$$\begin{aligned} \Lambda _n= \{q\in H_0^1 ([0,n],\, {\mathbb {R}}^2); \quad q(t) \ne e,\, \forall \, t\in [0,n]\}. \end{aligned}$$

Let \(\hbox {ind}_{z_0}(\gamma )\) denote the winding number of a closed curve in \({\mathbb {C}}\) around a point \(z_0\). That is

$$\begin{aligned} \hbox {ind}_{z_0}(\gamma )={1\over {2 i \pi }}\int _\gamma {{dz}\over {z-z_0}} \end{aligned}$$

which is a integer representing the number of counterclockwise turns that \(\gamma \) makes around \(z_0\).

A critical point of \(I_{0,n}\) will be found as a minimizer of \(I_{0,n}\) over the set

$$\begin{aligned} \Gamma _n^{\pm 1} = \{ q\in \Lambda _n,\hbox { ind}_e(q)=\pm 1 \}. \end{aligned}$$

Clearly \(\Gamma _n^{\pm 1} \ne \emptyset \), so we can define

$$\begin{aligned} c_n^{\pm 1} = \displaystyle \inf _{q\in \Gamma _n^{\pm 1}}I_{0,n}(q). \end{aligned}$$
(2)

We remark that, since \(I_{0,n}(q)= I_{0,n}(q(n-.))\; \hbox {for all}\; q\in \Lambda _n\), then \(c_n^1 = c_n^{-1}\).

Proposition 2.4

  1. 1)

    There exist \(M_1, \, M_2 >0\) such that

    $$\begin{aligned} 0<M_1\le c_n^1\le M_2, \quad \forall n\in {\mathbb {N}}^*. \end{aligned}$$
    (3)
  2. 2)

    For every \(n\in {\mathbb {N}}^*\), there is \(q_n\in \Gamma _n^1\) such that \(I_{0,n}(q_n)=c_n^1\). Moreover \(q_n\) is a non trivial classical solution of \((D_n)\).

Proof

1) Let \(q\in \Gamma _n^1\). The fact that ind\(_e(q) =1 \) implies that \(||q||_{L^{\infty } ( [0,n],\, {\mathbb {R}}^2)}\ge |e|\). Since \(q(0)=q(n)=0\), there exist \(s_q<t_q\) such that

$$\begin{aligned} |q(s_q)|=\frac{|e|}{3},\,\, |q(t_q)|= \frac{2|e|}{3}\hbox { and } \frac{|e|}{3}\le |q(t)|\le \frac{2|e|}{3}\,\, \hbox {for all}\,\, t\in [s_q,t_q]. \end{aligned}$$

Using the Cauchy-Schwartz inequality, we have the general formula

$$\begin{aligned} \int _{t_1}^{t_2} \Big [\frac{1}{2}|\dot{u}|^2 -V(u)\Big ]dt&\ge \displaystyle {{|u(t_2)-u(t_1)|^2} \over {2(t_2-t_1)}} +(t_2-t_1)\min _{t\in [t_1,t_2]}-V(u(t)) \nonumber \\&\ge |u(t_2)-u(t_1)| \sqrt{2\min _{t\in [t_1,t_2]}-V(u(t))} \end{aligned}$$
(4)

where \(u\in H^1([t_1,t_2],{\mathbb {R}}^2)\).

We denote \(c=\displaystyle \min _{\frac{|e|}{3}\le |x|\le \frac{2|e|}{3}} -V(x)>0\). Then from (4), we get

$$\begin{aligned} I_{0,n}(q)&\ge \int _{s_q}^{t_q} \Big [\frac{1}{2}|\dot{q}|^2- V(q)\Big ]dt\\&\ge {{|e|}\over 3}\sqrt{2c}=M_1. \end{aligned}$$

Thus by the arbitrariness of q, we obtain \(c_n ^1 \ge M_1 >0\) for any \( n\in {\mathbb {N}}^*\).

In order to prove that \(c_n ^1\) is bounded from above, let \(q\in \Gamma _1^1\) and define

$$\begin{aligned} v_n(t)= \left\{ \begin{array}{ll} q(t) &{} \quad ~\text {if}~~ t\in [0,1],\\ 0 &{} \quad ~\text {if}~~ t\in ]1,n].\\ \end{array} \right. \end{aligned}$$

Clearly \(v_n \in \Gamma _n ^1 \) and then

$$\begin{aligned} c_n ^1&\le I_{0,n}(v_n)=\int _{0}^{1}\Big [\frac{1}{2}|\dot{q}|^2-V_{\varepsilon _n}(q)\Big ]dt\\&\le I_{0,1}(q). \end{aligned}$$

Therefore

$$\begin{aligned} c_n^1 \le \displaystyle \inf _{ q\in \Gamma _1 ^1} I_{0,1}(q)=M_2. \end{aligned}$$

2) Let \((u_m ) \) be a minimizing sequence for \(c_n^1\). We have from (3), \((u_m)\) is bounded in \(H^1_0([0,n],\, {\mathbb {R}}^2)\). It follows that along a subsequence \((u_m )\) converge weakly in \(H_0^1 ([0,n], \; {\mathbb {R}}^2)\) and uniformly in [0, n] to a function \(q_n\). Since \( \int _{0}^{n} -V_{\varepsilon _n}(u_m) dt \) is bounded independently of m and \(V_{\varepsilon _n}\) is a strong force, Lemma 1.1 shows that \(q_n\in \Lambda _n\). Moreover we know that the winding number is continuous with respect to uniform convergence of curves. Therefore \(\hbox {ind}_e(q_n)=\lim _{m\rightarrow +\infty } \hbox {ind}_e (u_m) = 1\) and so \(q_n\in \Gamma _n ^1\). Using the lower semi continuity of \(I_{0,n}\), we get \( I_{0,n}(q_n) \le \liminf _{m\rightarrow +\infty }I_{0,n}(u_m)= c_n^1\). That is \(I_{0,n}(q_n)=c_n^1\). Now in a standard way, we can see that \(q_n\) is a critical point of \(I_{0,n}\) and then a nontrivial classical solution of \((D_n)\). \(\square \)

As a consequence of Proposition 2.4, we get the following estimates:

Lemma 2.5

  1. (i)

    There is a constant \(C>0\) which is independent of n such that for any \(n\in {\mathbb {N}}^*\),

    $$\begin{aligned} ||\dot{q}_n||_{ L^2 ([0,n], \, {\mathbb {R}}^2)} \le C;\quad \displaystyle \int _{0}^{n}-V(q_n) dt \le C; \quad ||q_n||_{L^{\infty } ([0,n],\, {\mathbb {R}}^2)} \le C. \end{aligned}$$
  2. (ii)

    For every \(n\in {\mathbb {N}}^*\), there is a constant \(h_n>0\) such that

    $$\begin{aligned} \displaystyle \frac{1}{2}|\dot{q}_n(t)|^2 + V_{\varepsilon _n}(q_n(t))= h_n, \quad \forall \, t\in [0,n]. \end{aligned}$$

    Moreover, \(\displaystyle h_n={1\over 2} |\dot{q}_n(0)|^2= {1\over 2} |\dot{q}_n(n)|^2 \longrightarrow 0\).

Since \(q_n \in \Gamma _n^1\), we have \(\displaystyle \max _{t\in [0,n]} |q_n(t)| >{{|e|}/ 4}\). Otherwise we would have ind\(_e( q_n) =0\).

Then we can find numbers \(\tau _n^1,\, \tau _n^2 \in ]0,n[\) such that

$$\begin{aligned} \displaystyle |q_n(\tau _n^1)|=|q_n(\tau _n^2)|=|e|/4 \displaystyle \quad \hbox {and } |q_n(t)|< |e|/4\;\, \hbox { if }\;\, t\in [0,\tau _n^1[\cup ]\tau _n^2,n]. \end{aligned}$$

Note that in [3], it was also proved the existence of approximated solution \(q_n\) of \((D_n)\) in \({\mathbb {R}}^N\) (\(N\ge 3)\) such that

*:

\(\displaystyle \max _{t\in [0,n]} |q_n(t)| >{\rho }\) where \(\rho >0\) is a constant;

*:

\(|\dot{q}_n (0)| \rightarrow 0\) and \(|\dot{q}_n(n)|\rightarrow 0\).

Using the continuity theorem of solutions with respect to initials conditions, we can see in a similar way to Lemma 2.7 in [3],

$$\begin{aligned} \tau _n^1\rightarrow \infty \quad \hbox {and} \, \, n-\tau _n^2 \rightarrow \infty . \end{aligned}$$

Next we define

$$\begin{aligned} \tilde{q}_n(t)= \left\{ \begin{array}{ll} q_n(t+\tau _n ^1) &{} \quad ~\text {if}~~ t\in [-\tau _n ^1,n-\tau _n ^1] ,\\ 0 &{} \quad ~\text {if}~~ t\in {\mathbb {R}}\setminus [-\tau _n ^1,n-\tau _n ^1] .\\ \end{array} \right. \end{aligned}$$
(5)

Clearly \( |{\tilde{q}}_n (0)|=|e|/4\) and \({\tilde{q}}_n\) verifies

$$\begin{aligned}{} & {} \ddot{\tilde{q}}_n+V_{\varepsilon _n}'(\tilde{q}_n)=0\quad \hbox { in }] -\tau _n^1,\, n-\tau _n^1[,\\{} & {} \displaystyle \frac{1}{2}|\dot{\tilde{q}}_n|^2 +V_{\varepsilon _n}(\tilde{q}_n)= h_n \quad \hbox { in }] -\tau _n^1,\, n-\tau _n^1[. \end{aligned}$$

By (i) of Lemma 2.5, we can extract a subsequence -still denoted by \({\tilde{q}}_n\)- which converges in \( C_{loc}({\mathbb {R}},\, {\mathbb {R}}^2)\) to some function \(\tilde{q}\in C({\mathbb {R}},\, {\mathbb {R}}^2)\cap L^\infty ({\mathbb {R}},\,{\mathbb {R}}^2)\) with \(\dot{\tilde{q}}\in L^2({\mathbb {R}},\,{\mathbb {R}}^2)\). Since \(-\tau _n^1 \rightarrow -\infty \) and \(n-\tau _n^1 \rightarrow +\infty \), we can see \(\tilde{q}\) is a non trivial generalized homoclinic solution of (HS). The complete proofs to Lemma 2.5 and the last statements are ommited as they are similar to its analogues in [3].

In what follows, we focus our attention to study the regularity of \(\tilde{q}\). First we state some further properties of \(\tilde{q}_n\) and \(\tilde{q}\) near the singularity.

Let \(t \in {\mathbb {R}}\) such that \(|\tilde{q}_n(t)-e|< |e|/4\). From the definition of \(V_{\varepsilon _n},\, \, \tilde{q}_n(t)\) verifies

$$\begin{aligned}&\displaystyle \ddot{\tilde{q}}_n+ \alpha \frac{{\tilde{q}}_n-e}{|{\tilde{q}}_n -e|^{\alpha +2}}+ W'(\tilde{q}_n) +2\varepsilon _n \frac{\tilde{q}_n-e}{|\tilde{q}_n-e|^4}=0, \end{aligned}$$
(6)
$$\begin{aligned}&\displaystyle \frac{1}{2}|\dot{\tilde{q}}_n|^2- \frac{1}{|\tilde{q}_n-e|^\alpha }+ W(\tilde{q}_n)-\frac{\varepsilon _n}{|\tilde{q}_n-e|^2}=h_n. \end{aligned}$$
(7)

Then

$$\begin{aligned} {1\over 2} \displaystyle \frac{d^2}{dt^2}|{\tilde{q}}_{n}(t)-e|^2&= < \ddot{\tilde{q}}_n,\tilde{q}_n-e> +|\dot{\tilde{q}}_n|^2 \\&= \displaystyle \frac{2-\alpha }{|\tilde{q}_n-e|^\alpha }-W'(\tilde{q}_n)(\tilde{q_n}-e) - 2W(\tilde{q}_n)+2h_n \\&= \displaystyle \frac{1}{|\tilde{q}_n-e|^\alpha } [ 2-\alpha -|\tilde{q}_n-e|^\alpha W'(\tilde{q}_n)(\tilde{q_n}-e) -2|\tilde{q}_n-e|^\alpha W(\tilde{q}_n)\\&\quad +2h_n |\tilde{q}_n-e|^\alpha ]. \end{aligned}$$

By (V3) (see Remark 2.1 i)) and the fact that \(h_n \rightarrow 0\), we can find \(0<\delta < |e|/4\) such that for sufficiently large n,

$$\begin{aligned} \displaystyle \frac{1}{2}\frac{d^2}{dt^2}|\tilde{q}_n(t)-e|^2>0 \quad \hbox {if}\quad |\tilde{q}_n(t)-e|<\delta . \end{aligned}$$
(8)

Similarly, if \(\tilde{q}(t) \not = e\) then \(\tilde{q}(t)\) satisfies (HS) and of energy 0. From this, we obtain

$$\begin{aligned} {1\over 2} \frac{d^2}{dt^2}| {\tilde{q}}(t)-e|^2 = \displaystyle \frac{1}{|\tilde{q} -e|^\alpha } [ 2-\alpha -|\tilde{q}-e|^\alpha W'(\tilde{q})(\tilde{q}-e)-2|\tilde{q}-e|^\alpha W(\tilde{q})]. \end{aligned}$$

Thus the property (8) holds also for \(\tilde{q}\), i.e.

$$\begin{aligned} \displaystyle \frac{1}{2}\frac{d^2}{dt^2}|\tilde{q}(t)-e|^2 >0 \quad \hbox {if}\quad 0< |\tilde{q}(t)-e|<\delta . \end{aligned}$$
(9)

Taking into account the property (ii) of a generalized solution, (9) implies that the collisions times of \(\tilde{q}\) (if they exist) are isolated.

Now we suppose that \(\tilde{q}\) has a collision at \(t=\tilde{t}\) i.e. \(\tilde{q}(\tilde{t})=e\) for some \({ \tilde{t}} \in {\mathbb {R}}\). We will study the angle which describes \(\tilde{q}_n(t)\) around e when t is near \(\tilde{t}\). In particular we will show that \(\tilde{q}_n\) have one self intersection if \(\alpha \in ]1,2[\).

Since \(\tilde{q}(t)\longrightarrow 0\) as \(t\rightarrow \pm \infty \), there exist \(\tau _1< \tilde{t}<\tau _2\) such that

$$\begin{aligned} |\tilde{q}(\tau _1)-e|= |\tilde{q}(\tau _2)-e|= \displaystyle \frac{\delta }{2} \quad \hbox {and }\,\, 0< |\tilde{q}(t)-e|<\displaystyle \frac{\delta }{2}\quad \forall \, t\in ]\tau _1,\tau _2[\setminus \{\tilde{t}\}. \end{aligned}$$

Thus for sufficiently large n, we have

$$\begin{aligned} |\tilde{q}_n(\tau _i)-e|\ge \displaystyle \frac{\delta }{4}\quad \hbox {for}\quad i=1,2 \end{aligned}$$
(10)

and

$$\begin{aligned} |\tilde{q}_n(t)-e|<\delta \quad \forall \; t\in [\tau _1,\tau _2]. \end{aligned}$$
(11)

Let \(t_n \in [\tau _1,\tau _2]\) and \(\delta _n>0\) such that \( \delta _n = |\tilde{q}_n(t_n)-e|= \displaystyle \min _{t\in [\tau _1,\tau _2]} |\tilde{q}_n(t)-e|.\)

Clearly \(\delta _n \le |\tilde{q}_n({\tilde{t}})-e|\longrightarrow |\tilde{q}(\tilde{t})-e|=0\). So \(\delta _n\longrightarrow 0\). Moreover, up a subsequence, we have \(t_n\longrightarrow \tilde{t}\).

By (8), we have

$$\begin{aligned}&\displaystyle \frac{d}{dt}|\tilde{q}_n(t)-e| < 0 \quad \forall \, t\in [\tau _1,t_n [, \end{aligned}$$
(12)
$$\begin{aligned}&\displaystyle \frac{d}{dt}|\tilde{q}_n(t)-e| > 0 \quad \forall \, t\in ]t_n,\tau _2]. \end{aligned}$$
(13)

In the sequel we use a rescaling argument as in [17] and we introduce the function

$$\begin{aligned} x_n(s)= \delta _n ^{-1} \Big [ \tilde{q}_n \left( \delta _n^{\frac{\alpha +2}{2}} s +t_n \right) -e \Big ], \quad s\in {\mathbb {R}}. \end{aligned}$$

Remark that

$$\begin{aligned} |x_n(0)|=1 \quad \hbox {and} \quad (x_n(0),\, \dot{x} _n (0))=0. \end{aligned}$$
(14)

Let \(l>0\). For sufficiently large n, since \(\delta _n^{\frac{\alpha +2}{2}} s +t_n \in [\tau _1,\tau _2]\) for \(s\in [-l,l]\), we have from (11) and (6)-(7),

$$\begin{aligned}&\ddot{x}_n(s)+ \alpha \frac{x_n}{|x_n|^{\alpha +2}}+\delta _n^{\alpha +1} W'(\delta _nx_n+e)+ \frac{2\varepsilon _n}{\delta _n^{2-\alpha }}\frac{x_n}{|x_n|^4}=0 \quad \hbox {in}\; [-l,l], \end{aligned}$$
(15)
$$\begin{aligned}&\frac{1}{2}|\dot{x}_n|^2- \frac{1}{|x_n|^\alpha }+ \delta _n^\alpha W(\delta _nx_n+e)- \frac{\varepsilon _n}{\delta _n^{2-\alpha }} \frac{1}{|x_n|^2}= \delta _n^\alpha h_n \quad \hbox { in} \; [-l,l]. \end{aligned}$$
(16)

We extract a subsequence still indexed by n such that

$$\begin{aligned} d= \displaystyle \lim _{n\rightarrow +\infty }\frac{\varepsilon _n}{\delta _n^{2-\alpha }}\in [0, +\infty ] \end{aligned}$$
(17)

exists. For d we need to show that

Lemma 2.6

The quantity d defined in (17) is a finite one.

Proof

On the contrary, we assume that \(d= +\infty \). We will prove that \(\tilde{q}_n\) has a self intersection around e to find a contradiction. Let consider another rescaling of \(\tilde{q}_n\):

$$\begin{aligned} y_n(s)=\displaystyle \delta _n ^{-1} \Big [\tilde{q}_n \left( \varepsilon _n^{-\frac{1}{2}}\delta _n^2 s +t_n\right) -e \Big ], \quad s\in {\mathbb {R}}. \end{aligned}$$
(18)

Since \(\displaystyle \varepsilon _n ^{-\frac{1}{2}}\delta _n^2 = \Big ( \varepsilon _n ^{-1}\delta _n^{2-\alpha } \Big ) ^{\frac{1}{2}}\delta _n ^{1+ \frac{\alpha }{2}}\longrightarrow 0\), then for sufficiently large n, we have \(\varepsilon _n^{-\frac{1}{2}}\delta _n^2\,s +t_n \in [\tau _1,\tau _2], \,\forall \, s \in [-l,l]\). From (12)-(13), we get

$$\begin{aligned}{} & {} \displaystyle \frac{d}{ds}|y_n(s)|<0 \quad \forall \, \, s\in [-l,0[,\\{} & {} \displaystyle \frac{d}{ds}|y_n(s)|>0 \quad \forall \, \, s\in ]0,l]. \end{aligned}$$

As in [3], we can see that -up a subsequence-

$$\begin{aligned} y_n\longrightarrow \cos (\sqrt{2}s)e_1 + \sin (\sqrt{2}s)e_2 \quad \hbox { in } \, C^2_{\hbox {\tiny {loc}}} ({\mathbb {R}}, \, {\mathbb {R}}^2) \end{aligned}$$

where \((e_1, e_2)\) is an orthonormal basis of \({\mathbb {R}}^2\). Using polar coordinates, there exists a function \(\alpha _n \in C^2({\mathbb {R}},\, {\mathbb {R}})\) such that

$$\begin{aligned} y_n(s)= |y_n(s)|\Big ( \cos (\alpha _n(s))e_1 + \sin (\alpha _n(s))e_2\Big ). \end{aligned}$$

We take \(l> \sqrt{2}\pi \). Since \(\dot{\alpha }_n\longrightarrow \sqrt{2}\) uniformly on \([-l,l]\), then for sufficiently large n, there exist \(s_1<0<s_2\) such that

$$\begin{aligned} \alpha _n(0)-\alpha _n(s_1)=\alpha _n(s_2)-\alpha _n(0)=2\pi . \end{aligned}$$
(19)

We may assume that \(1< |y_n(s_1)|\le |y_n(s_2)|\). By continuity, there exists \(s_3\in ]0,s_2]\) such that \(|y_n(s_1)|=|y_n(s_3)|\). Since \({\dot{\alpha }}_n >0\), it follows from (19) that

$$\begin{aligned} \alpha _n(s_3)-\alpha _n(s_1)&=\alpha _n(s_3)-\alpha _n(0)+\alpha _n(0)-\alpha _n(s_1)\\&>2\pi . \end{aligned}$$

This implies the existence of \(s_1', s_2' \in [s_1, s_3]\) such that \(y_n(s_1')=y_n(s_2')\) and

ind\(_0 {y_n}|_{[s'_1,s'_2]}=1\). From (5) and (18), it follows the existence of \(t',\, t''\in ]0,n[\) such that \(q_n(t')=q_n(t'')\) and ind\(_e q_n|{_{[t',t'']}}=1\). But this contradicts the fact that \(q_n\) is a minimum of \(I_{0,n}\) over \(\Gamma ^1_n\). Indeed, let consider the function

$$\begin{aligned} \underline{q_n}(t)= \left\{ \begin{array}{ll} q_n(t) &{} \quad ~\text {if}~~ t\in [0,n]\setminus [t',t''],\\ q_n(t'+t''-t) &{} \quad ~\text {if}~~ t\in [t',t''].\\ \end{array} \right. \end{aligned}$$

Then \(\underline{q_n} \in \Gamma _n^{-1}\) and \(I_{0,n}(\underline{q_n})=I_{0,n}(q_n)=c_n^1=c_n^{-1}\). Therefore \(\underline{q_n}\) is a classical solution of \((D_n)\). By the uniqueness of the solution of ordinary differential equation, we deduce that \(\underline{q_n}=q_n\): clearly this is a contradiction since \(\hbox {ind}_e(\underline{q_n})=-1\) and ind\(_e(q_n)=1\). \(\square \)

Since \(d< +\infty \), by the continuity dependence of solutions on initial data and equation, we can see from (14)-(16) and (V3) the existence of an orthonormal basis \((e_1,e_2)\) of \( {\mathbb {R}}^2\) such that

$$\begin{aligned} x_n(s)\longrightarrow x_{\alpha ,d}(s)\quad \hbox {in } \, \, C^2_{\hbox {\tiny {loc}}} ({\mathbb {R}},\, {\mathbb {R}}^2) \end{aligned}$$

where \( x_{\alpha ,d}(s)\) is the solution of the initial value problem

$$\begin{aligned} \left\{ \begin{array}{ll} \displaystyle \ddot{x}+\frac{\alpha x}{|x|^{\alpha +2}} +2d \frac{x}{|x|^4}=0, \\ x(0)=e_1,\quad \dot{x}(0)=\sqrt{2(1+d)}e_2. \\ \end{array}\right. \end{aligned}$$

We use polar coordinates and write

$$\begin{aligned}{} & {} \tilde{q}_n(t)-e =|\tilde{q}_n(t)-e| \Big ( \cos (\tilde{\theta }_n(t))e_1 + \sin (\tilde{\theta }_n(t))e_2\Big ),\\{} & {} x_{\alpha ,d}(s)= |x_{\alpha ,d}(s)|\Big (\cos (\theta _{\alpha ,d}(s))e_1 + \sin (\theta _{\alpha ,d}(s))e_2\Big ), \end{aligned}$$

where \(\tilde{\theta }_n(s),\, \theta _{\alpha ,d}(s)\in {\mathbb {R}}\) with \(\theta _{\alpha ,d}(0)=0\). In [18] we observed the following properties for \(\theta _{\alpha ,d}\)

$$\begin{aligned}&\dot{\theta }_{\alpha ,d}(s)>0 \quad \forall \; s\in {\mathbb {R}}, \end{aligned}$$
(20)
$$\begin{aligned}&\Delta \theta _{\alpha ,d}=\displaystyle \lim _{s\rightarrow +\infty } (\theta _{\alpha ,d}(s) -\theta _{\alpha ,d}(-s))= \frac{2\pi \sqrt{1+d}}{2-\alpha }. \end{aligned}$$
(21)

We remark that \(\Delta \theta _{\alpha ,d}>\pi \, \, \forall \, \alpha \in ]0,1]\) and if \(\alpha \in ]1,2[\) then \(\Delta \theta _{\alpha ,d}>2\pi \).

Let \(B_r(e)\) denote the open ball of radius r about e. We will give a estimate of \(\tilde{\theta }_n(t)\) when \(\tilde{q}_n(t) \in B_\mu (e)\setminus B_{L\delta _n}(e)\) for sufficiently small \(\mu >0\) and large \(L>1\) and n.

We have for \(t <t'\),

$$\begin{aligned} |\tilde{\theta }_n(t')-\tilde{\theta }_n(t)|&\le \int _t^{t'}|\dot{\tilde{\theta }}_n(\tau )|d\tau \nonumber \\&=\int _t^{t'}\Big |{d\over {dt}} {{{\tilde{q}}_n(\tau ) -e}\over {|{{\tilde{q}}_n(\tau ) -e}|}}\Big |d\tau \end{aligned}$$
(22)

On the other hand, Tanaka [18] studied under the condition (V3) with \(e=0\) the behavior of generalized periodic solutions of singular Hamiltonian systems in \({\mathbb {R}}^N\). In a neighborhood of the singularity, the generalized solution is a limit of classical solutions of perturbed problems with potentials \(V_\varepsilon \) as in our case, so we can apply some locally property of approximated solutions near a collision. More precisely, modifying the argument in Proposition 1.5 slightly, we can see that for any \(\eta >0\) there exist constants \( \mu ,\, S>0\) and \(n_0\in {\mathbb {N}}^*\) such that for \(n\ge n_0\),

$$\begin{aligned}{} & {} \displaystyle \int _t^{t'}\Big |{d\over {dt}} {{{\tilde{q}}_n(\tau ) -e}\over {|{{\tilde{q}}_n(\tau ) -e}|}}\Big |d\tau \le \frac{\eta }{2} \quad \hbox {if } \,\, {\tilde{q}}_n(t),\,\, {\tilde{q}}_n(t')\in B_\mu (e) \ \ \hbox {and} \nonumber \\{} & {} \tau _1<t<t'< t_n-S\delta _n^{\frac{\alpha +2}{2}}\, \, \hbox {or }\,\, t_n+S \delta _n^{\frac{\alpha +2}{2}}<t<t'<\tau _2. \end{aligned}$$
(23)

Combining (22) and (23), we get

Lemma 2.7

For any \(\eta >0\), there are constants \( \mu \in ]0,\, \delta /4[,\, S>0\) such that for sufficiently large n, if \( \, {\tilde{q}}_n(t),\,\, {\tilde{q}}_n(t')\in B_\mu (e) \,\, \hbox {and}\)

$$\begin{aligned} \tau _1<t<t'< t_n-S\delta _n^{\frac{\alpha +2}{2}}\, \, \hbox {or } \,\, t_n+S \delta _n^{\frac{\alpha +2}{2}}<t<t'<\tau _2, \end{aligned}$$

then

$$\begin{aligned} |\tilde{\theta }_n(t')-\tilde{\theta }_n(t)|\le \displaystyle \frac{\eta }{2}. \end{aligned}$$

End of the proof of Theorem 2.2. 1) If \(\alpha \in ]1,2[\), there exists from (21) \(\eta >0\) such that \(\Delta \theta _{\alpha ,d}> 2\pi +\eta \). For this \(\eta \), we choose \(\mu \in ]0,\; \delta /4[,\; S>0\) and n sufficiently large as in Lemma 2.7.

From (21) again we can take \(S_1>S\) such that

$$\begin{aligned} \theta _{\alpha ,d}(S_1)- \theta _{\alpha ,d}(-S_1)> 2\pi +\eta . \end{aligned}$$

Then we obtain for sufficiently large n,

$$\begin{aligned} \displaystyle \tilde{\theta }_n\left( t_n+\delta _n^{\frac{\alpha +2}{2}}S_1\right) -\tilde{\theta }_n\left( t_n-\delta _n^{\frac{\alpha +2}{2}}S_1\right) >2\pi + \eta . \end{aligned}$$
(24)

On the other hand, since \(\displaystyle | \tilde{q}_n(t_n \pm S_1\delta _n^{\frac{\alpha +2}{2}})-e|\longrightarrow |\tilde{q}({\tilde{t}})-e|=0,\) we can assume that

$$\begin{aligned} \displaystyle |\tilde{q}_n\left( t_n \pm S_1\delta _n ^{\frac{\alpha +2}{2}}\right) -e|<\mu . \end{aligned}$$

We set \(t'_{1,n}= t_n - S_1\delta _n^{\frac{\alpha +2}{2}} \). Then we have from (10)

$$\begin{aligned} \displaystyle |\tilde{q}_n(t'_{1,n})-e|<\mu <{\delta \over 4} \le |\tilde{q}_n(\tau _{1})-e|. \end{aligned}$$

Therefore there exists \( t_{1,n}\in ]\tau _1,\, t'_{1,n}[\) such that

$$\begin{aligned} |\tilde{q}_n(t_{1,n})-e|= \mu . \end{aligned}$$

Similarly we set \(t_{2,n}= t_n +S_1\delta _n^{\frac{\alpha +2}{2}}\). Since \(\displaystyle |\tilde{q}_n(t_{2,n})-e|<\mu <{\delta \over 4} \le |\tilde{q}_n(\tau _{2})-e|\), there exists \( t'_{2,n}\in ] t_{2,n},\, \tau _2[\) such that

$$\begin{aligned} |\tilde{q}_n(t'_{2,n})-e|= \mu . \end{aligned}$$

Applying lemma 2.7 for \( t=t_{i,n}\) and \( t'=t'_{i,n}\,\, (i=1,2)\), we obtain

$$\begin{aligned} |\tilde{\theta }_n (t'_{i,n}) -\tilde{\theta }_n(t_{i,n})|\le \displaystyle \frac{\eta }{2} \quad \hbox { for}\quad i=1,\, 2. \end{aligned}$$
(25)

It follows from (24)-(25),

$$\begin{aligned} \tilde{\theta }_n(t'_{2,n})-\tilde{\theta }_n(t_{1,n})&= \tilde{\theta }_n(t'_{2,n}) -\tilde{\theta }_n(t_{2,n}) +\tilde{\theta }_n(t_{2,n}) -\tilde{\theta }_n(t'_{1,n})+\tilde{\theta }_n(t'_{1,n}) -\tilde{\theta }_n(t_{1,n}) \\&> \displaystyle -\frac{\eta }{2} +2\pi + \eta - \frac{\eta }{2}=2\pi . \end{aligned}$$

That is \(\tilde{q}_n\) describes an angle greater than \(2\pi \) in going from \(\partial B_\mu (e)\; \hbox {back to}\, \partial B_\mu (e)\) which implies the existence of \(t''_{1,n},\; t''_{2,n} \) with \(\tau _1<t''_{1,n}<t''_{2,n}<\tau _2\) such that

$$\begin{aligned} {\tilde{q}}_{n}(t''_{1,n})={\tilde{q}}_{n}(t''_{2,n}) \quad \hbox {and ind}_e {{\tilde{q}}_n|}_{{[t''_{1,n},t''_{2,n}]}}=1. \end{aligned}$$

Thus we deduce that \({q_n}\) has a self intersection around e for sufficiently large n. As in the proof of Lemma 2.6, we get a contradiction and then we conclude that \(\tilde{q}\) is a non collision homoclinic solution of (HS).

2) In the case \(\alpha \in ]0,1]\), the angle which describes \(\tilde{q}_n\) near e is greater than \(\pi \) and \(\tilde{q}_n\) cannot have a self intersection. The fact that the collisions times of \(\tilde{q}\) are isolated and since \(\tilde{q}(t)\longrightarrow 0 \) as \( t\rightarrow \pm \infty \), we get that the number of collisions of \(\tilde{q}\) is finite. Assume \(\tilde{q}(t)\) enters the singularity e and let

$$\begin{aligned} {t_0}=\displaystyle \min \{ t\in {\mathbb {R}}, \,\, \tilde{q}(t)=e\}. \end{aligned}$$

Since (HS) is time reversible, the function

$$\begin{aligned} {q}(t)= \left\{ \begin{array}{ll} \tilde{q}(t) &{} \quad ~\text {if}~~ t\le {t_0},\\ \tilde{q}(2 {t_0}-t) &{} \quad ~\text {if}~~ t\ge {t_0}, \end{array} \right. \end{aligned}$$

is a generalized homoclinic solution of (HS) and satisfies \({q}(t+ {t_0})= {q}({t_0}-t)\) for all t. Moreover q has one collision in \({\mathbb {R}}\). The proof of Theorem 2.2 is finally complete.

Remark 2.8

The assumption (V3) is far too restrictive in the case \(\alpha \in ]0,1]\) and the existence of a generalized homoclinic solution with finite number of collisions and then a solution as in Theorem 2.2 2) still holds if we replace (V3) by

  1. (V’3)
    1. (i)

      \(V(q)\rightarrow -\infty \) as \( q \rightarrow e\);

    2. (ii)

      There exists a constant \(\delta \in ]0,|e|/4[\) such that

      $$\begin{aligned} V(q)+ \displaystyle \frac{1}{2}V'(q)(q-e)<0\quad \hbox { for} \,\, 0<|q-e|\le \delta . \end{aligned}$$

We have kept (V3) in the case \(\alpha \in ]0,1]\), on the one hand to obtain a certain symmetry in the statements of Theorem 2.2, on the other hand the study of approximated solutions near collisions under (V3) will be useful in Sect. 3 to prove the existence of a non-collision heteroclinic orbit at infinity for every \(\alpha \in ]0,2[\) (see Theorem 3.1 below).

3 Existence of Heteroclinic Orbits

In this section, the existence of non-collision heteroclinic orbits at infinity for (HS) will be established. Consider the problem

$$\begin{aligned} \left\{ \begin{array}{ll} \ddot{q}+V'(q)=0, \\ q(-\infty )=0, \quad |q(+\infty )|= +\infty ,\\ \dot{q}(\pm \infty )=0,\\ \end{array}\right. \end{aligned}$$
(P)

where V behaves like (1) near e and satisfies the assumptions (V1)-(V3) of Theorem 2.2.

The natural condition for V at infinity for (P) is \(\displaystyle \lim _{|q|\rightarrow +\infty }V(q)=0\). More precisely, we assume

  1. (V’4)

    \(V(q)\sim -\displaystyle \frac{a}{|q|^b} \quad \hbox {as} \; |q|\rightarrow +\infty \,\hbox { for some }\, a>0,\, b>2\).

When \(\alpha \in ]0,1]\), we need a further property of V near e

  1. (V5)

    there exists \(\phi \in C^2 (]0,\, r[, \, {\mathbb {R}})\) for some \(r\in ]0,|e|/4[\) such that

    $$\begin{aligned} V(q)= \phi (|q-e|)\quad \forall \, \, q \in B_r(e). \end{aligned}$$

Theorem 3.1

Suppose (V1)-(V3), (V’4) and (V5)(only when \(\alpha \in ]0,1]\)).

Then there exists at least one non-collision orbit of (P).

We now pass to the proof of Theorem 3.1. Solutions of (P) can be found as critical points of the functional

$$\begin{aligned} \displaystyle I(q)= \int _{{\mathbb {R}}}\Big [ \frac{1}{2}|\dot{q}|^2 - V(q) \Big ]dt \end{aligned}$$

defined on the set

$$\begin{aligned} \displaystyle \Lambda _0^{\infty } =\{ q\in H;\; \, q(-\infty )=0, \, |q(+\infty )|=+\infty , \, q(t)\ne e\, \forall \,\, t\in {\mathbb {R}}\} \end{aligned}$$

where

$$\begin{aligned} H=\Big \{ q\in H_{loc} ^1 ({\mathbb {R}}, {\mathbb {R}}^N), \displaystyle \int _{{\mathbb {R}}} |\dot{q}|^2 dt < +\infty \Big \}. \end{aligned}$$

In [14] the case \(\alpha \ge 2\) (strong force case) has been studied and the existence of one classical solution of (P) has been found as a minimizer of I on \(\Lambda _0^{\infty }\). In our situation where \(0<\alpha <2\), we make a perturbation to the potential as in Theorem 2.2 and we consider for every n the problem

figure b

Since \(V_{\varepsilon _n}\) is a strong force, we can use Lemma 1.1, and a standard compactness argument provides the existence of a classical (non-collision) solution \(q_n\) of \((P_n)\) as a minimizer of the functional

$$\begin{aligned} \displaystyle I_n (q)=\int _{{\mathbb {R}} }\Big [ \frac{1}{2}|\dot{q}|^2 -V_{\varepsilon _n}(q) \Big ]dt \end{aligned}$$

on \(\Lambda _0^{\infty }\), i.e. \(q_n \in \Lambda _0 ^{\infty }\) such that

$$\begin{aligned} \displaystyle I_n(q_n)=\inf _{q\in \Lambda _0 ^{\infty }}I_n(q). \end{aligned}$$
(26)

By normalization, we can assume that

$$\begin{aligned} |q_n(0)|=\displaystyle \frac{|e|}{4}\quad \hbox {and}\quad |q_n(t)|< \frac{|e|}{4}\quad \forall \, t<0. \end{aligned}$$

Remark also that \(q_n\) has energy zero.

Now we observe that \(I_n(q_n ) \le \inf _{q\in \Lambda _0 ^{\infty }} I_1(q)=c_1< +\infty \). We deduce then the existence of a constant \(C>0\) independent of n such that \( ||q_n||_H \le C\) and \( \int _{{\mathbb {R}}}- V(q_n)dt\le C \). Thus there is a subsequence still denoted by \((q_n)\) and a function \(q\in H\) such that \( q_n\) converges weakly in H and uniformly in \( C_{loc}({\mathbb {R}},\, {\mathbb {R}}^2)\) to q. By Fatou’s lemma \(\int _{{\mathbb {R}}}-V(q) dt\le C\), so the set of collisions \(D=\{t\in {\mathbb {R}},\, q(t)=e \}\) is of measure 0. In a standard way, we can see that \(q\in C^2 ({\mathbb {R}}\setminus D, {\mathbb {R}}^2)\), satisfies (HS) and has energy zero in \({\mathbb {R}}\setminus D\), that is q is a generalized solution of (HS).

Lemma 3.2

\(q(t)\ne e \) for all \(\, t\in {\mathbb {R}}\).

Proof

We prove by contradiction assuming \(q({\tilde{t}})=e\) for some \(\tilde{t}\in {\mathbb {R}}\). From (V3) and the conservation of the energy, q satisfies the property (9) and then we can see that the collisions times of q are isolated. Moreover there is a sequence \((t_n)\) such that \(t_n\longrightarrow \tilde{t}\) and \(|q_n(t)-e|\) takes its local minimum at \(t=t_n\).

As in Theorem 2.2 we define \(\delta _n = |q_n(t_n)-e|\) and \(d=\displaystyle \lim _{n\rightarrow +\infty } \frac{\varepsilon _n}{\delta _n ^{2-\alpha }}\in [0, +\infty ]\) (we extract a subsequence if necessary).

If we suppose that \(d=+\infty \), we can see as in Lemma 2.6 that \(q_n\) has a self intersection i.e. there exist \(\sigma _1<\sigma _2\) such that \(q_n(\sigma _1)=q_n(\sigma _2)\) and ind\(_e q_n|_{[\sigma _1,\sigma _2]}=1\). Here we consider the function

$$\begin{aligned} {u_n}(t)= \left\{ \begin{array}{ll} q_n(t+\sigma _1-\sigma _2) &{} \quad ~\text {if}~~ t\le \sigma _2,\\ q_n(t) &{} \quad ~\text {if}~~ t \ge \sigma _2.\\ \end{array} \right. \end{aligned}$$

Then \(u_n \in \Lambda _0 ^\infty \) and it is easy to see that \(I_n(u_n)<I_n(q_n)\), which contradicts (26).

Therefore we get \(d<+\infty \). In that case, there is a function \(x_{\alpha ,d}\) such that after taking a subsequence still denoted by n,

$$\begin{aligned} \delta _n ^{-1} \Big [ q_n\left( \delta _n^{\frac{\alpha +2}{2}} s+ t_n\right) -e \Big ] \longrightarrow x_{\alpha ,d}(s)=|x_{\alpha ,d}(s)|\Big ( \cos (\theta _{\alpha ,d}(s)) e_1+\sin (\theta _{\alpha ,d}(s))e_2\Big ) \end{aligned}$$

in \( C_{loc}^2 ({\mathbb {R}}, {\mathbb {R}}^2)\) where \((e_1,e_2)\) is an orthonormal basis of \({\mathbb {R}}^2\) and \(\theta _{\alpha ,d}:{\mathbb {R}}\longrightarrow {\mathbb {R}}\) satisfies \(\theta _{\alpha ,d} (0)=0\) and the properties (20)-(21).

In polar coordinates, there exists \(\theta _n:{\mathbb {R}}\rightarrow {\mathbb {R}}\) such that

$$\begin{aligned} q_n (t) -e=|q_n(t)-e|\Big ( \cos (\theta _{n}(t)) e_1 +\sin (\theta _{n}(t))e_2\Big ). \end{aligned}$$

For \(\alpha \in ]1,2[\), we have from (21) \(\Delta \theta _{\alpha ,d}>2\pi \). Repeating the argument of Theorem 2.2, we get that \(q_n\) has a self intersection around e which is a contradiction as above.

For \(\alpha \in ]0,1]\), we will use (V5) to get a contradiction. Here \(\Delta \theta _{\alpha ,d}>\pi \) and \(q_n\) cannot have a self intersection. However there exists \(L>0\) such that \(\theta _{\alpha ,d}(L)-\theta _{\alpha ,d}(-L)> \pi \). Setting \(\sigma _{1,n}= t_n-\delta _n^{\frac{\alpha +2}{2}}L \) and \(\sigma _{2,n}=t_n+ \delta _n ^{\frac{\alpha +2}{2}}L\), for sufficiently large n we have

$$\begin{aligned}{} & {} |q_n(t)-e|\le r,\quad \forall \, \, t \in [\sigma _{1,n},\sigma _{2,n}], \nonumber \\{} & {} \theta _n(\sigma _{2,n})-\theta _n(\sigma _{1,n})>\pi ,\nonumber \\{} & {} \dot{ \theta }_n(t)>0 \quad \forall \, t\in [\sigma _{1,n},\sigma _{2,n}]. \end{aligned}$$
(27)

Let \(\sigma '_{1,n}, \sigma '_{2,n} \in [\sigma _{1,n}, \sigma _{2,n}]\) such that

$$\begin{aligned} \theta _n(\sigma '_{2,n})- \theta _n(\sigma '_{1,n})=\pi . \end{aligned}$$

We consider the function \(\hat{q}_n \) defined by

$$\begin{aligned}{} & {} \hat{q}_n(t)=q_n(t)\quad \hbox { if} \quad t\in {\mathbb {R}}\setminus [\sigma '_{1,n},\sigma '_{2,n}], \\{} & {} \hat{q}_n(t)-e = |q_n(t)-e|\Big ( \cos \big (-\theta _n(t)+2 \theta _n(\sigma '_{1,n})\big ) e_1 + \sin \big (-\theta _n(t)+2\theta _n(\sigma '_{1,n})\big )e_2\Big )\\{} & {} \quad \hbox {if}\quad t\in [\sigma '_{1,n},\sigma '_{2,n}]. \end{aligned}$$

That is \({\hat{q}_n|}_{[\sigma '_{1,n},\sigma '_{2,n}]}\) and \({{q}_n|}_{[\sigma '_{1,n},\sigma '_{2,n}]}\) are axially symmetric with respect to the axis joining the two points \(q_n(\sigma '_{1,n})\) and \(q_n(\sigma '_{2,n})\).

Clearly \(\hat{q}_n\in \Lambda _0 ^{\infty }\) and from (V5), since V is radially symmetric about e in \(B_r(e)\), we get that \(I_n(q_n)=I_n(\hat{q}_n)=\displaystyle \inf _{q\in \Lambda _0 ^{\infty }}I_n(q)\). It follows that \(\hat{q}_n\) is of class \(C^2\) and satisfies the equation \( \ddot{q}+V_{\varepsilon _n}'(q)=0 \). By the uniqueness of solution of ordinary differential equation, we deduce that \(q_n=\hat{q}_n\), which enters in contradiction with (27). Therefore we conclude that \(q(t)\ne e\) for all \(t\in {\mathbb {R}}\). \(\square \)

End of the proof of Theorem 3.1. To prove that q is a solution of (P), it remains to show that \(q(-\infty )=0,\, |q(+\infty )|= +\infty \) and \(\dot{q}( \pm \infty )=0\). Using the formula (4) and the fact that \(I(q)<+\infty \) one can see that \( |q(-\infty )| \) and \(|q(+\infty )| \) exist and they are 0 or \(+\infty \). Since \(|q(t)|=\lim |q_n(t)| \le |e|/4\, \, \forall \, t\le 0\), then \( q(-\infty )=0\).

To show that \(|q(+\infty )|=+\infty \), we suppose that \(|q(+\infty )|=0\). We will construct as in [12] a function \(Q_n\in \Lambda _0 ^{\infty }\) such that \(I_n(Q_n) < I_n(q_n)\). Indeed, let \(\varepsilon \in ]0,|e|/16[\) and \(T_\varepsilon >0\) such that \(q(T_\varepsilon ) \in B_\varepsilon \) the open ball of radius \(\varepsilon \) about 0. For sufficiently large n we have \(q_n(T_\varepsilon )\in B_{2\varepsilon }\). We consider the function \(Q_n\in \Lambda _0^{\infty }\) different from \(q_n\) for \(t<T_\varepsilon \) such that

$$\begin{aligned} Q_n(t)=\left\{ \begin{array}{lll} 0 &{}\hbox {if }\,\, t<T_\varepsilon -1,\\ (t-T_\varepsilon +1)q_n(T_\varepsilon ) &{}\hbox {if } \,\,t\in [T_\varepsilon -1,T_\varepsilon ],\\ q_n(t)&{}\hbox {if }\,\, t\ge T_\varepsilon . \end{array}\right. \end{aligned}$$

Since \(V_{\varepsilon _n}=V\) in \(B_{2\varepsilon }\) and \(V_{\varepsilon _n}\le V\), we have

$$\begin{aligned} I_n(Q_n)-I_n(q_n)\le 2 \varepsilon ^2 +\max _{x\in B_{2\varepsilon }}-V(x) -\int _{-\infty }^{T_\varepsilon } \Big [ \frac{1}{2}|\dot{q}_n|^2 - V(q_n) \Big ]dt. \end{aligned}$$
(28)

On the other hand, since \(|q_n(0)|=|e|/4\) and \(|q_n(T_\varepsilon )|\le 2\varepsilon <|e|/8\), there are \(t_1<t_2\) in \([0,T_\varepsilon ]\) such that

$$\begin{aligned} |q_n(t_1)|=\frac{|e|}{4},\,\, |q_n(t_2)|=\frac{|e|}{8}\hbox { and } \frac{|e|}{8}\le |q_n(t)| \le \frac{|e|}{4}\,\, \hbox {for all}\,\, t\in [t_1,t_2]. \end{aligned}$$

By the formula (4), it holds that

$$\begin{aligned} \int _{-\infty }^{T_\varepsilon } \Big [ \frac{1}{2}|\dot{q}_n|^2 - V(q_n) \Big ]dt&\ge \int _{t_1}^{t_2} \Big [ \frac{1}{2}|\dot{q}_n|^2 - V(q_n) \Big ]dt\nonumber \\&\ge {{|e|}\over 8} \sqrt{2m_0} \end{aligned}$$
(29)

where \(m_0=\displaystyle \min _{\frac{|e|}{8}\le |x|\le \frac{|e|}{4}} -V(x)>0\).

Then combining (28) and (29), we get \(I_n(Q_n)-I_n(q_n)<0\) for sufficiently small \(\varepsilon \), which contradicts (26). We conclude that \(|q(+\infty )|=+\infty \).

From the conservation of energy and the fact that \(V(q(t))\longrightarrow 0 \) as \(t \rightarrow \pm \infty \), it follows that \(\displaystyle \frac{1}{2}|\dot{q}(t)|^2 = -V(q(t))\longrightarrow 0\) as \(t\rightarrow \pm \infty \), that is \(\dot{q}( \pm \infty )=0\). The proof is complete.