1 Introduction

We study the existence and orbital stability of standing waves for the following nonlinear Schrödinger equation with inverse square potential on the half line

$$\begin{aligned} {\left\{ \begin{array}{ll}\displaystyle i u_t + u'' + c \frac{u}{x^2}+|u|^{p-1}u=0, \\ u(0)= u_0 \in H^1_0({\mathbb {R}}^+), \end{array}\right. } \end{aligned}$$
(1.1)

where \(u: {\mathbb {R}}\times {\mathbb {R}}^+ \rightarrow {\mathbb {C}}\), \(u_0: {\mathbb {R}}^+ \rightarrow {\mathbb {C}}\), \(1<p<\infty \), and \(0<c<1/4\).

There has been considerable interest recently in the study of the Schrödinger equation with inverse-square potential in three and higher dimensions. Classification of the so-called minimal mass blow-up solutions, global well-posedness, and stability of standing wave solutions were studied in [1, 6, 8, 22]. In the papers by Bensouilah et al. [1], and by Trachanas and Zographopoulos [22] the authors establish orbital stability of ground state solutions in the Hardy subcritical \((c<(N-2)^2/4)\) and Hardy critical \((c=(N-2)^2/4)\) case respectively for dimensions higher that three. In both cases, orbital stability is proved by showing the precompactness of minimizing sequences of the energy functional on an \(L^2\) constraint. Local well-posedness was established for the two-dimensional space by Suzuki in [21], and in three and higher dimensions by Okazawa et al. in [18]. The presence of the inverse square potential in one-dimensional space has also attracted attention. In [13] H. Kovarik and F. Truc established dispersive estimates for \(\partial _x^2+c/x^2\).

The dynamics of the equation is closely related to Hardy’s inequality (see [7])

$$\begin{aligned} c\int _0^\infty \frac{|u|^2}{x^2}dx\leqslant \int _0^\infty |u'|^2dx \text { for all } u \in C^\infty _0(0,\infty ), \end{aligned}$$
(1.2)

where \(c\leqslant 1/4\). We introduce the Hardy functional

$$\begin{aligned} H(u)=\int _0^\infty \left( |u'|^2-\frac{c}{x^2}|u|^2\right) dx, \end{aligned}$$

which is closely related to our problem. We will mainly focus on the case \(0<c<1/4\), when the natural energy space associated to (1.1) is \(H^1_0({\mathbb {R}}^+)\), and the semi-norm \(\left\| u'\right\| ^2_{L^2}\) is equivalent to H(u).

Let us consider the operator

$$\begin{aligned} H_c=-\frac{\partial ^2}{\partial x^2} - \frac{c}{x^2} \end{aligned}$$

acting on \(C^\infty _0({\mathbb {R}}^+)\). Owing to the Hardy inequality, if \(c<1/4\) the quadratic form \(\langle H_c\varphi ,\varphi \rangle \) is positive definite on \(C^\infty _0({\mathbb {R}}^+)\). It is natural to take the Friedrichs extension of \(H_c\), thereby defining a self-adjoint operator in \(L^2({\mathbb {R}}^+)\), which generates an isometry group in \(H^1_0({\mathbb {R}}^+)\).

Local well-posedness for parameters \(1<p<\infty \) and \(0<c<\frac{1}{4}\) follows by standard arguments (see e.g. in [3] Chapter 4). In particular, the following holds.

Theorem 1.1

Let \(1<p<\infty \) and \(c<1/4\). For any initial value \(u_0 \in H^1_0({\mathbb {R}}^+)\), there exist \(T_{\mathrm {min}},T_{\mathrm {max}} \in (0,\infty ]\) and a unique maximal solution \(u\in C((-T_{\mathrm {min}},T_{\mathrm {max}}),H^1_0({\mathbb {R}}^+))\) of (1.1), which satisfies for all \(t\in (-T_{\mathrm {min}},T_{\mathrm {max}})\) the conservation laws

$$\begin{aligned} \left\| u(t)\right\| _{L^2}=\left\| u_0\right\| _{L^2}, \quad E(u(t))=E(u_0), \end{aligned}$$
(1.3)

where the energy is defined as

$$\begin{aligned} E(u)=\frac{1}{2}\left\| u'\right\| ^2_{L^2}-\frac{c}{2}\left\| \frac{u}{x}\right\| ^2_{L^2}-\frac{1}{p+1}\left\| u\right\| ^{p+1}_{L^{p+1}}, \textit{ for } u\in H^1_0({\mathbb {R}}^+). \end{aligned}$$
(1.4)

Moreover, the so-called blow-up alternative holds: if \({T_\mathrm {max}}<\infty \) then \(\lim _{t\rightarrow T_{\mathrm {max}}}\left\| u'(t)\right\| _{L^2}=\infty \), (or \(T_{\mathrm {min}}<\infty \) then \(\lim _{t\rightarrow -T_{\mathrm {min}}}\left\| u'(t)\right\| _{L^2}=\infty \)).

In this work we address the existence of standing wave solutions and their orbital stability/instability. By introducing the ansatz \(u(t,x)=e^{i\omega t}\varphi (x)\), the standing wave equation to (1.1) reads as

$$\begin{aligned} \varphi ''+\frac{c}{x^2}\varphi -\omega \varphi +|\varphi |^{p-1}\varphi =0. \end{aligned}$$
(1.5)

First we will prove regularity of standing waves and the Pohozaev identities. To establish the existence of standing waves we carry out a minimization procedure on the Nehari manifold for the so-called action functional

$$\begin{aligned} S(v)=\frac{1}{2}\left\| v'\right\| ^2_{L^2}-\frac{c}{2}\left\| \frac{v}{x}\right\| ^2_{L^2}+\frac{\omega }{2}\left\| v\right\| ^2_{L^2}-\frac{1}{p+1}\left\| v\right\| ^{p+1}_{L^{p+1}}\quad v\in H^1_0({\mathbb {R}}^+). \end{aligned}$$

Owing to the non-compactness of the problem, we have to use a profile decomposition lemma, in the spirit of the article by Jeanjean and Tanaka [11]. To establish strong convergence of the minimizing sequence on the Nehari manifold we compare the minimization problem with the problem “at infinity”, i.e. when \(c=0\). Hence, we obtain that the set of bound states is not empty:

$$\begin{aligned} {\mathcal {A}}=\{ u\in H^1_0({\mathbb {R}}^+)\setminus \{0\} : u''+c u/x^2 -\omega u + |u|^{p-1}u=0\}\ne \varnothing . \end{aligned}$$

We are in particular interested in the orbital stability/instability of ground states, i.e., solutions which minimize the action functional. We denote the set of ground sate solutions by

$$\begin{aligned} {\mathcal {G}}=\{u\in {\mathcal {A}} : S(u)\leqslant S(v) \text { for all } v \in {\mathcal {A}} \}. \end{aligned}$$

We use Lions’ concentration-compactness principle to obtain a variational characterization of ground states on an \(L^2\)-constraint, thereby establishing the orbital stability of the set of ground states for nonlinearities with power \(1<p<5\). Finally, for \(p\geqslant 5\) we establish strong instability by a convexity argument.

2 Existence of bound states

We start by investigating the standing wave equation,

$$\begin{aligned} {\left\{ \begin{array}{ll} \varphi '' + \frac{c}{x^2}\varphi - \omega \varphi + |\varphi |^{p-1}\varphi =0,\\ \varphi \in H^1_0({\mathbb {R}}^+) \setminus \{0\}. \end{array}\right. } \end{aligned}$$
(2.1)

First, we prove the regularity of solutions to (2.1) by a bootstrap argument.

Proposition 2.1

Let \(\omega >0\) and \(c < 1/4\). Assume \(\varphi \in H^1_0({\mathbb {R}}^+)\) is a solution of (2.1) in \(H^{-1}({\mathbb {R}}^+)\). Then the following statements are true

  1. (1)

    \(\varphi \in W^{2,r}_0((\epsilon ,\infty ))\) for all \(r\in [2, +\infty )\) and \(\epsilon >0\), in particular \(\varphi \in C^1((\epsilon ,\infty ))\);

  2. (2)

    The solution is exponentially bounded, that is \(\mathrm {e}^{\sqrt{\omega } x}(|\varphi |+|\varphi '|)\in L^{\infty }({\mathbb {R}}^+)\);

Proof

(1) For \(\varphi \in H^1_0({\mathbb {R}}^+)\) we have \(\varphi \in L^q({\mathbb {R}}^+)\) for all \(q\in [2,\infty ]\). We get easily that \(|\varphi |^{p-1}\varphi \in L^q({\mathbb {R}}^+)\) for all \(q\in [2,\infty )\). By (2.1) we have for any \(\epsilon >0 \) that \(\varphi \in W^{2,q}_0((\epsilon ,\infty ))\) for all \(q\in [2,\infty )\). By Sobolev’s embedding we get \(\varphi \in C^{1,\delta }((\epsilon ,\infty ))\) for all \(\delta \in (0,1)\), hence \(|\varphi (x)|\rightarrow 0\), and \(|\varphi '(x)|\rightarrow 0\) as \(x\rightarrow \infty \).

(2) Let \(\omega >0\). Changing \(\varphi (x)\) to \(\varphi (x)=\omega ^{1/(p-1)}\varphi (\sqrt{\omega }x) \) we may assume that \(\omega =1\) in (2.1). Let \(\varepsilon >0\) and \(\theta _\varepsilon (x)=e^{\frac{x}{1+\varepsilon x}}\), for \(x\geqslant 0\). It is easy to see that \(\theta _\varepsilon \) is bounded, Lipschitz continuous, and \(|\theta '_\varepsilon (x)|\leqslant \theta _\varepsilon (x)\) for all \(x\in {\mathbb {R}}^+\). Additionally, \(\theta _\varepsilon (x)\rightarrow e^x\) uniformly on bounded sets of \({\mathbb {R}}^+\). Taking the scalar product of the equation (2.1) with \(\theta _\varepsilon \varphi \in H^1_0({\mathbb {R}}^+)\), we get

$$\begin{aligned} \mathop {{\mathrm{Re}}}\nolimits \int _{{\mathbb {R}}^+} \varphi '\cdot (\theta _\varepsilon {\bar{\varphi }})'dx - c\int _{{\mathbb {R}}^+} \theta _\varepsilon \frac{|\varphi |^2}{x^2} dx + \int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi |^2 dx = \int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi |^{p+1}dx. \end{aligned}$$

Using the inequality \(\mathop {{\mathrm{Re}}}\nolimits (\varphi '(\theta _\varepsilon {\bar{\varphi }})')\geqslant \theta _\varepsilon |\varphi '|^2-\theta _\varepsilon |\varphi ||\varphi '|\) and

$$\begin{aligned} \int _{{\mathbb {R}}^+} \theta _\varepsilon |\varphi ||\varphi '|dx \leqslant \frac{1}{2}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi |^2dx +\frac{1}{2}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi '|^2dx, \end{aligned}$$

we obtain

$$\begin{aligned} \frac{1}{2}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi '|^2dx + \frac{1}{2}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi |^2dx - c \int _{{\mathbb {R}}^+} \theta _\varepsilon \frac{|\varphi |^2}{x^2}dx \leqslant \int _{{\mathbb {R}}^+} \theta _\varepsilon |\varphi |^{p+1} dx. \end{aligned}$$

Let \(R>0\) such that if \(x>R\), then \(\frac{c}{x^2}\leqslant \frac{1}{8}\) and \(|\varphi (x)|^{p-1}\leqslant \frac{1}{8}\). Then we get

$$\begin{aligned}&c\int _{{\mathbb {R}}^+}\theta _\varepsilon \frac{|\varphi |^2}{x^2} dx + \int _{{\mathbb {R}}^+} \theta _\varepsilon |\varphi |^{p+1} \\&\quad \leqslant e^R \left( \int _0^R c\frac{|\varphi |^2}{x^2} dx +\int _0^R|\varphi |^{p+1}dx \right) + \frac{1}{4}\int _{{\mathbb {R}}^+} \theta _\varepsilon |\varphi |^2dx. \end{aligned}$$

From the last two inequalities it follows that

$$\begin{aligned} \frac{1}{2}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi '|^2dx + \frac{1}{4}\int _{{\mathbb {R}}^+}\theta _\varepsilon |\varphi |^2dx \leqslant e^R \left( \int _0^R c\frac{|\varphi |^2}{x^2} dx +\int _0^R|\varphi |^{p+1}dx \right) . \end{aligned}$$

By taking \(\varepsilon \downarrow 0\) we get

$$\begin{aligned} \frac{1}{2}\int _{{\mathbb {R}}^+} e^{x} |\varphi '|^2dx + \frac{1}{4}\int _{{\mathbb {R}}^+}e^{x}|\varphi |^2dx <\infty . \end{aligned}$$

Since both \(\varphi \) and \(\varphi '\) are Lipschitz continuous we deduce that \(|\varphi (x)|e^{x}\) and \(|\varphi '(x)|e^{x}\) are bounded. \(\square \)

We now prove that there exists a solution to (2.1). We define the action functional associated to (2.1) as follows

$$\begin{aligned} S(u)= \frac{1}{2} H(u) + \frac{\omega }{2}\left\| u\right\| ^2_{L^2} - \frac{1}{p+1}\left\| u\right\| ^{p+1}_{L^{p+1}}, \end{aligned}$$

for \(c<1/4\) and \(u\in H^1_0({\mathbb {R}}^+)\). Clearly, we have

$$\begin{aligned} S'(u)= - u''-\frac{c}{x^2}u + \omega u - |u|^{p-1}u. \end{aligned}$$

Therefore, to prove the existence of a solution to (2.1) amounts to show that S has a nontrivial critical point. A simple calculation yields the following identities.

Lemma 2.2

Assume \(p>1\), \(\omega >0\) and \(c< 1/4\). Let \(\varphi \in H^1_0({\mathbb {R}}^+)\) be a solution of (2.1) in \(H^{-1}({\mathbb {R}}^+)\). Then the following identities are true:

$$\begin{aligned} \left\| \varphi '\right\| ^2_{L^2} - c\left\| \frac{\varphi }{x}\right\| ^2_{L^2} + \omega \left\| \varphi \right\| ^2_{L^2}-\left\| \varphi \right\| ^{p+1}_{L^{p+1}}&=0, \end{aligned}$$
(2.2)
$$\begin{aligned} \left\| \varphi '\right\| ^2_{L^2} - c\left\| \frac{\varphi }{x}\right\| ^2_{L^2} - \frac{p-1}{2(p+1)}\left\| \varphi \right\| ^{p+1}_{L^{p+1}}&= 0. \end{aligned}$$
(2.3)

Proof

We obtain the first equality by multiplying (2.1) by \({\bar{\varphi }}\) and integrating over \({\mathbb {R}}^+\).

To prove the second equality, let us put \(\varphi _{\lambda }(x)=\lambda ^{1/2}\varphi (\lambda x)\) for \(\lambda >0\). We have that

$$\begin{aligned} S(\varphi _\lambda )=\frac{\lambda ^2}{2}\left\| \varphi '\right\| ^2_{L^2}-\frac{\lambda ^2 c}{2}\left\| \frac{\varphi }{x}\right\| ^2_{L^2}+\frac{\omega }{2}\left\| \varphi \right\| ^2_{L^2}-\frac{\lambda ^{(p-1)/2}}{p+1}\left\| \varphi \right\| ^{p+1}_{L^{p+1}}, \end{aligned}$$

from which we get

$$\begin{aligned} \frac{\partial }{\partial \lambda } S(\varphi _\lambda )\Big |_{\lambda =1}=\left\| \varphi '\right\| ^2_{L^2} - c \left\| \frac{\varphi }{x} \right\| _{L^2}^2-\frac{p-1}{2(p+1)}\left\| \varphi \right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

We also have that

$$\begin{aligned} \frac{\partial }{\partial \lambda }S(\varphi _\lambda )\Big |_{\lambda =1} = \left\langle S'(\varphi ), \frac{\partial \varphi _\lambda }{\partial \lambda }\Big |_{\lambda =1} \right\rangle . \end{aligned}$$

Now \(\frac{\partial \varphi _\lambda }{\partial \lambda }\Big |_{\lambda =1}=\frac{1}{2} \varphi + x \varphi '\) is in \(H^1({\mathbb {R}}^+)\), since \(\varphi \) and \(\varphi '\) are exponentially decaying at infinity by Proposition 2.1. We obtain that the right hand-side is well-defined. Since \(\varphi \) is a critical point of S, we obtain \(S'(\varphi )=0\), which concludes the proof. \(\square \)

Remark 2.3

Since (2.2) and (2.3) hold for solutions of (2.1), it follows for \(\omega \ne 0\) that

$$\begin{aligned} \omega \left\| \varphi \right\| ^2_{L^2}=\frac{p+3}{2(p+1)}\left\| \varphi \right\| ^{p+1}_{L^{p+1}}>0. \end{aligned}$$

Hence, non-trivial solution of (2.1) exists only if \(\omega >0\).

Let us define for all \(u\in H^1_0({\mathbb {R}}^+)\) the following functional:

$$\begin{aligned} J(u)= (S'(u),u)_{H^{-1},H^1_0} =H(u)+ \omega \left\| u\right\| ^2_{L^2}-\left\| u\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

It follows from Lemma 2.2, that \({\mathcal {N}}= \{ u\in H^1_0({\mathbb {R}}^+)\setminus \{0\} : J(u)=0 \}\) contains all nontrivial critical points of S. We aim to show that the infimum of the following minimization problem is attained

$$\begin{aligned} m=\inf \{S(u) : u \in {\mathcal {N}}\}=\frac{p-1}{2(p+1)}\inf \{\left\| u\right\| ^{p+1}_{L^{p+1}}:u \in {\mathcal {N}} \}. \end{aligned}$$
(2.4)

First we prove the following lemma.

Lemma 2.4

\({\mathcal {N}}\) is nonempty, and \(m>0\).

Proof

Let \(u\in H^1_0({\mathbb {R}}^+)\setminus \{0\}\). Take

$$\begin{aligned} t(u)=\left( \frac{H(u)+\omega \left\| u\right\| _{L^2}^2}{\left\| u\right\| ^{p+1}_{L^{p+1}}}\right) ^{1/(p-1)}. \end{aligned}$$

By simple calculation, we get that \(J(t(u)u)=0\), hence \(t(u)u\in {\mathcal {N}}\). We see that

$$\begin{aligned} m=\inf _{u\in {\mathcal {N}}} S(u)=\inf _{u\in {\mathcal {N}}}\left( S(u)-\frac{1}{p+1}J(u) \right) =\frac{p-1}{2(p+1)} \inf _{u\in {\mathcal {N}}} (H(u)+\omega \left\| u\right\| ^2_{L^2}). \end{aligned}$$

It follows from Sobolev’s and Hardy’s inequalities, that there exists \(C>0\) such that

$$\begin{aligned} H(u) + \omega \left\| u\right\| ^2_{L^2} =\left\| u\right\| ^{p+1}_{L^{p+1}} \leqslant C (H(u) + \omega \left\| u\right\| ^2_{L^2})^{(p+1)/2}, \end{aligned}$$

for all \(u \in {\mathcal {N}}\). Hence,

$$\begin{aligned} \left( \frac{1}{C}\right) ^{2/(p-1)}\leqslant H(u) + \omega \left\| u\right\| ^2_{L^2} \text { for all } u\in H^1_0({\mathbb {R}}^+), \end{aligned}$$

which implies that

$$\begin{aligned} m\geqslant \frac{p-1}{2(p+1)}\left( \frac{1}{C}\right) ^{2/(p-1)}>0. \end{aligned}$$

\(\square \)

Lemma 2.5

Let \(c<1/4\), and \(p>1\). Then if \(u\in H^1_0({\mathbb {R}}^+)\) is a minimizer of (2.4), then |u| is also a minimizer. In particular, we can search for the minimizers of (2.4) among the non-negative, real-valued functions of \(H^1_0({\mathbb {R}}^+)\).

Proof

Let \(u\in H^1_0({\mathbb {R}}^+)\) be a solution of the minimization problem (2.4). It is well-known that if \(u\in H^1_0({\mathbb {R}}^+)\) then \(|u|\in H^1_0({\mathbb {R}}^+)\) and \(\left\| |u|'\right\| _{L^2}\leqslant \left\| u'\right\| _{L^2}\). Moreover, \(\left\| |u|\right\| _{L^{p+1}}=\left\| u\right\| _{L^{p+1}}\). Therefore, \(J(|u|)\leqslant J(u)\). Hence there exists a \(\lambda \in (0,1]\) such that \(J(\lambda |u|)=J(u)=0\). Then

$$\begin{aligned} m\leqslant S(\lambda |u|)=\frac{p-1}{2(p+1)} \left\| \lambda u\right\| ^{p+1}_{L^{p+1}}\leqslant \frac{p-1}{2(p+1)} \left\| u\right\| ^{p+1}_{L^{p+1}}=m. \end{aligned}$$

Hence \(\lambda =1\), \(J(|u|)=0\), and \(S(|u|)=m\). \(\square \)

Let \(m\in {\mathbb {R}}\). We say that \(\{u_n\}_{n\in {\mathbb {N}}}\) is a Palais-Smale sequence for S at level m, if

$$\begin{aligned} S(u_n)\rightarrow m, \quad S'(u_n)\rightarrow 0 \textit{ in } H^{-1}({\mathbb {R}}^+), \end{aligned}$$

as \(n\rightarrow \infty \).

Lemma 2.6

Let \(c<1/4\), and \(p>1\). There exists a bounded Palais-Smale sequence \(\{u_n\}_{n\in {\mathbb {N}}}\subset {\mathcal {N}}\) for S at the level m. Namely, there is a sequence \(\{u_n\}_{n\in {\mathbb {N}}}\subset {\mathcal {N}}\) bounded in \(H^1({\mathbb {R}}^+)\) such that, as \(n\rightarrow \infty \),

$$\begin{aligned} S(u_n)\rightarrow m, \quad S'(u_n)\rightarrow 0 \textit{ in } H^{-1}({\mathbb {R}}^+). \end{aligned}$$

Proof

Since \({\mathcal {N}}\) is a closed manifold in \(H^1_0({\mathbb {R}}^+)\), it is a complete metric space. Hence, Ekeland’s variational principle (see pp. 51–53 in [20]) directly yields the existence of a Palais-Smale sequence at level m in \({\mathcal {N}}\).

We now show that if \(\{u_n\}_{n\in {\mathbb {N}}} \subset {\mathcal {N}}\) and \(\left\| u_n\right\| ^2_{H^1} \rightarrow \infty \), then \(S(u_n)\rightarrow \infty \). Indeed, since \(u_n\in {\mathcal {N}}\) from Hardy’s inequality we get that

$$\begin{aligned} S(u_n)= & {} \frac{p-1}{2(p+1)}(H(u_n)+\omega \left\| u_n\right\| ^2_{L^2}) \\\geqslant & {} \frac{p-1}{2(p+1)}(\min \{1,(1-4c)\}\left\| u'_n\right\| ^2_{L^2}+\omega \left\| u_n\right\| ^2_{L^2}). \end{aligned}$$

Therefore, any Palais-Smale sequence \(\{u_n\}_{n\in {\mathbb {N}}}\) is bounded in \(H^1_0({\mathbb {R}}^+)\). \(\square \)

Before proceeding to our next lemma, let us recall some classical results, see e.g. [3], concerning the case \(c=0\). It is well-known that the set of solutions of

$$\begin{aligned} q''-\omega q+|q|^{p-1}q=0, \quad \omega >0, \quad q \in H^1({\mathbb {R}}) \end{aligned}$$
(2.5)

is given by \(\{e^{i\theta }q(\cdot + y): y\in {\mathbb {R}}, \theta \in {\mathbb {R}}\}\), where q is a symmetric, positive solution of (2.5), explicitly given by

$$\begin{aligned} q(x)=\left( \frac{(p+1)\omega }{2}\text {sech}^2\left( \frac{(p-1)\sqrt{\omega }}{2}x \right) \right) ^{1/(p-1)}. \end{aligned}$$
(2.6)

Moreover, up to translation and phase invariance, it is the unique solution of the minimization problem

$$\begin{aligned} m^\infty&=\inf \{S^\infty (u): u\in H^1({\mathbb {R}})\setminus \{0\}, J^\infty (u)=0\}\\&=\frac{p-1}{2(p+1)}\inf \{\left\| u\right\| ^{p+1}_{L^{p+1}({\mathbb {R}})}: u\in H^1({\mathbb {R}})\setminus \{0\}, J^\infty (u)=0\}, \end{aligned}$$

where the functionals \(S^\infty \) and \(J^\infty \) are defined by

$$\begin{aligned} S^\infty (u)&=\frac{1}{2}\left\| u'\right\| ^2_{L^2({\mathbb {R}})}+\frac{\omega }{2}\left\| u\right\| ^2_{L^2({\mathbb {R}})}-\frac{1}{p+1}\left\| u\right\| ^{p+1}_{L^{p+1}({\mathbb {R}})},\\ J^\infty (u)&=\left\| u'\right\| ^2_{L^2({\mathbb {R}})}+\omega \left\| u\right\| ^2_{L^2({\mathbb {R}})}-\left\| u\right\| ^{p+1}_{L^{p+1}({\mathbb {R}})}. \end{aligned}$$

Lemma 2.7

Let \(0<c<1/4\), and \(p>1\). Then \(m<m^\infty \).

Proof

It is not hard to see that \(m\leqslant m^\infty \), we only need to prove that \(m\ne m^\infty \). Let us first note that if \(u\in H^1_0({\mathbb {R}}^+)\setminus \{0\}\) and \(J(u)<0\), then \(m<{\tilde{S}}(u)\), where

$$\begin{aligned} {\tilde{S}}(u)=\frac{p-1}{2(p+1)}\left( H(u)+\omega \left\| u\right\| ^2_{L^2} \right) . \end{aligned}$$

Indeed, if \(J(u)<0\), then let us define

$$\begin{aligned} t(u)=\left( \frac{H(u)+\omega \left\| u\right\| ^2_{L^2}}{\left\| u\right\| ^{p+1}_{L^{p+1}}}\right) ^{1/(p-1)}. \end{aligned}$$

Hence \(t(u) \in (0,1)\), \(t(u) u\in {\mathcal {N}}\), and

$$\begin{aligned} m\leqslant {\tilde{S}}(t(u) u)=t^2(u){\tilde{S}}(u)< {\tilde{S}}(u). \end{aligned}$$

Now let us define \(\psi _A(x)=q(x+A)-q(x-A)\) for \(x\geqslant 0\). For large enough A we obtain the following estimates (see Lemma 5.1 in the Appendix):

$$\begin{aligned} \int _0^\infty |\psi _A'|^2dx&= \int _{-\infty }^\infty |q'|^2dx+O\left( \left( 2A+\frac{1}{\sqrt{\omega }}\right) e^{-2\sqrt{\omega } A}\right) ,\\ \int _0^\infty |\psi _A|^2dx&= \int _{-\infty }^\infty |q|^2dx +O\left( \left( 2A+\frac{1}{\sqrt{\omega }}\right) e^{-2\sqrt{\omega } A}\right) ,\\ \int _0^\infty \frac{|\psi _A|^2}{x^2}dx&\leqslant \frac{4}{A^2}\int _{-\infty }^\infty |q|^2dx +O\left( \frac{1}{A^2}e^{-\sqrt{\omega } A}\right) ,\\ \int _0^\infty |\psi _A|^{p+1}dx&= \int _{-\infty }^\infty |q|^{p+1}dx + O\left( e^{-2\sqrt{\omega } A}\right) . \end{aligned}$$

Since \(0<c<1/4\), we obtain for \(A>0\) large enough

$$\begin{aligned} J(\psi _A)&\leqslant \left\| q'\right\| ^2_{L^2({\mathbb {R}})}+\omega \left\| q\right\| ^2_{L^2({\mathbb {R}})}-\left\| q\right\| ^{p+1}_{L^{p+1}({\mathbb {R}})}-\frac{4c}{A^2}\left\| q\right\| ^2_{L^2({\mathbb {R}})}+O\left( \frac{1}{A^2}e^{-\sqrt{\omega } A}\right) \\&=-\frac{4c}{A^2}\left\| q\right\| ^2_{L^2({\mathbb {R}})}+O\left( \frac{1}{A^2}e^{-\sqrt{\omega } A}\right) <0, \end{aligned}$$

and

$$\begin{aligned} {\tilde{S}}(\psi _A)&\leqslant \frac{p-1}{2(p+1)}\left( \left\| q'\right\| ^2_{L^2({\mathbb {R}})}+\omega \left\| q\right\| ^2_{L^2({\mathbb {R}})}-\frac{4c}{A^2}\left\| q\right\| ^2_{L^2({\mathbb {R}})}\right) +O\left( \frac{1}{A^2}e^{-\sqrt{\omega } A}\right) \\&=m^\infty -\frac{p-1}{2(p+1)}\frac{4c}{A^2}\left\| q\right\| ^2_{L^2({\mathbb {R}})}+O\left( \frac{1}{A^2}e^{-\sqrt{\omega } A}\right) < m^\infty . \end{aligned}$$

Since \(J(\psi _A)<0\), we get

$$\begin{aligned} m<{\tilde{S}}(\psi _A)< m^\infty , \end{aligned}$$

which concludes the proof. \(\square \)

We need the following lemma, which describes the behavior of bounded Palais-Smale sequences. We note that \(H^1_0({\mathbb {R}}^+)\) functions can be extended to functions in \(H^1({\mathbb {R}})\) by setting \(u\equiv 0\) on \({\mathbb {R}}^-\). The proof of the following statement is presented in the appendix.

Lemma 2.8

Let \(\{u_n\}_{n\in {\mathbb {N}}}\subset H^1_0({\mathbb {R}}^+)\) be a bounded Palais-Smale sequence for S at level m. Then there exists a subsequence still denoted by \(\{u_n\}_{n\in {\mathbb {N}}}\), a \(u_0 \in H^1_0({\mathbb {R}}^+)\) solution of

$$\begin{aligned} \varphi ''+\frac{c}{x^2}\varphi -\omega \varphi +|\varphi |^{p-1}\varphi =0, \end{aligned}$$

an integer \(k\geqslant 0\), \(\{x_n^i\}_{i=1}^k\subset {\mathbb {R}}^+\), and nontrivial solutions \(q_i\) of (2.5) satisfying

$$\begin{aligned}&u_n \rightharpoonup u_0 \quad \textit{weakly in} \quad H^1_0({\mathbb {R}}^+), \\&S(u_n)\rightarrow S(u_0)+\sum _{i=1}^kS^\infty (q_i), \\&u_n-(u_0+\sum _{i=1}^kq_i(x-x_n^i))\rightarrow 0 \quad \textit{strongly in}\quad H^1({\mathbb {R}}),\\&|x_n^i|\rightarrow \infty ,\quad |x_n^i-x_n^j|\rightarrow \infty \quad for \quad 1\leqslant i\ne j\leqslant k, \end{aligned}$$

where in case \(k=0\), the above holds without \(q_i\) and \(x_n^i\).

We only need to show that the critical point of S provided by Lemma 2.8 is non-trivial.

Theorem 2.9

Let \(0<c<1/4\). Then there exists \(u\in {\mathcal {N}}\setminus \{0\}\), \(u\geqslant 0\) a.e., such that \(S(u)=m\).

Proof

We only have to prove that the \(\{u_n\}_{n\in {\mathbb {N}}}\) bounded Palais-Smale sequence obtained in Lemma 2.6 admits a strongly convergent subsequence. Assume that it is not the case. Using Lemma 2.8 we see that \(k\geqslant 1\) and \(u_n\) is weakly convergent to \(u_0\) in \(H^1_0({\mathbb {R}}^+)\) up to a subsequence. Then

$$\begin{aligned} m=\lim _{n\rightarrow \infty } S(u_n)\geqslant S(u_0)+S^\infty (q)=S(u_0)+m^\infty . \end{aligned}$$

Now, \(S(u_0)\geqslant 0\) since \(J(u_0)=0\). Thus \(m\geqslant m^\infty \), which contradicts Lemma 2.7. Hence \(k=0\) and \(u_n\rightarrow u_0\) in \(H^1_0({\mathbb {R}}^+)\). \(\square \)

Lemma 2.10

Let \(p>1\) and \(\omega >0\). There exists a \(\mu >0\) such that

$$\begin{aligned} \int _0^\infty |u|^2dx= \mu , \text { for every } u \in {\mathcal {G}}. \end{aligned}$$

The mass of ground state solutions is \(\mu =\frac{m}{\omega }\frac{p+3}{p-1}\). Moreover, we have

$$\begin{aligned} \left\| u\right\| ^{p+1}_{L^{p+1}}=\frac{2(p+1)}{p-1}m, \text { and } H(u)=m \text { for every } u\in {\mathcal {G}}. \end{aligned}$$

Proof

Since \(u \in {\mathcal {G}}\) is a solution of (2.1), it satisfies (2.2) and (2.3). By subtracting the two identities we get

$$\begin{aligned} \omega \left\| u\right\| _{L^2}^2=\frac{p+3}{2(p+1)}\left\| u\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$
(2.7)

Additionally, since u is a ground state solution, it also solves the minimization problem (2.4). From (2.4) and (2.3) we get

$$\begin{aligned} \omega \left\| u\right\| ^2_{L^2}+\frac{p-5}{2(p+1)}\left\| u\right\| _{L^{p+1}}^{p+1}=2m. \end{aligned}$$
(2.8)

From (2.7) and (2.8) it follows

$$\begin{aligned} \left\| u\right\| ^2_{L^2}=\frac{m}{\omega }\frac{p+3}{p-1}>0. \end{aligned}$$

Thus, let \(\mu =\frac{m}{\omega }\frac{p+3}{p-1}\). Now it follows from (2.4) and (2.3) that

$$\begin{aligned} \left\| u\right\| ^{p+1}_{L^{p+1}}=\frac{2(p+1)}{p-1}m, \text { and } H(u)=m \text { for every } u\in {\mathcal {G}}. \end{aligned}$$

which concludes the proof. \(\square \)

3 Stability

In this section we consider nonlinearities with \(1<p<5\). Our aim is to prove orbital stability of the standing waves. To do so, we investigate the minimization problem:

$$\begin{aligned} I=\inf \{E(u) : u \in \Gamma \}, \end{aligned}$$
(3.1)

where

$$\begin{aligned} \Gamma = \{ u\in H^1_0({\mathbb {R}}^+) : \left\| u\right\| _{L^2}^2=\mu \}. \end{aligned}$$

and the energy E is defined by (1.4). We will rely on a of Lions’ concentration-compactness principle [15] and the arguments by Cazenave and Lions [4], see also in [3]. The main problem is to obtain compactness of minimizing sequences owing to the absence of translation invariance. We define the problem at infinity by

$$\begin{aligned} I^\infty =\inf \{ E^\infty (u) : u\in H^1({\mathbb {R}}) \text { and } \left\| u\right\| ^2_{L^2}=\mu \}, \end{aligned}$$
(3.2)

where

$$\begin{aligned} E^\infty (u)=\frac{1}{2}\int _{{\mathbb {R}}}|u'|^2dx-\frac{1}{p+1}\int _{\mathbb {R}}|u|^{p+1}dx. \end{aligned}$$

We recall some well-known facts about the minimization problem (3.2) (see [3, Chapter 8.]). For every \(\mu >0\), there exists a unique, positive, symmetric function \(q=q(\mu ) \in H^1({\mathbb {R}})\), such that

$$\begin{aligned} \left\| q\right\| _{L^2}=\mu , \quad E^\infty (q)=I^\infty , \end{aligned}$$

and q solves the nonlinear equation

$$\begin{aligned} q''-\lambda q+|q|^{p-1}q=0, \end{aligned}$$

where \(\lambda =\lambda (\mu )\). Moreover, there exists \(M>0\) such that

$$\begin{aligned} e^{\sqrt{\lambda } |x|}|q(x)|\leqslant M \textit{ and } e^{\sqrt{\lambda } |x|}|q'(x)|\leqslant M. \end{aligned}$$

We proceed by proving the following lemma:

Lemma 3.1

If \(0<c<1/4\), then the following inequality holds:

$$\begin{aligned} I<I^\infty . \end{aligned}$$

Proof

For \(A>0\), let C(A) be a normalizing factor specified later. Let us define

$$\begin{aligned} \Psi _A(x)=C(A)(q(x+A)-q(x-A)) \text { for } x\geqslant 0. \end{aligned}$$

Since q is even, we obtain \(\Psi _A \in H^1_0({\mathbb {R}}^+)\) and

$$\begin{aligned}&\int _0^\infty |\Psi _A(x)|^2dx= C^2(A)\left( \int _{-\infty }^\infty |q|^2dx -\int _{-\infty }^\infty q(x+A)q(x-A)dx\right) . \end{aligned}$$

We estimate the second integral by (see Lemma 5.1)

$$\begin{aligned} \int _{-\infty }^\infty q(x+A)q(x-A)dx = O\left( \left( 2A+\frac{1}{\sqrt{\lambda }}\right) e^{-2\sqrt{\lambda } A}\right) . \end{aligned}$$

We define

$$\begin{aligned} C(A)=\left( \frac{\mu }{\mu -\int _{-\infty }^\infty q(x+A)q(x-A)dx}\right) ^{1/2}. \end{aligned}$$

C(A) is a continuous function of A, \(C(A)\geqslant 1\), and \(C(A)\rightarrow 1\) exponentially fast as \(A\rightarrow \infty \). Thus, \(\left\| \Psi _A\right\| _{L^2}=\mu \) for all \(A>0\). By Lemma 5.1 in the Appendix, we obtain for \(A>0\) large enough that

$$\begin{aligned} \int _0^\infty |\Psi _A'|^2dx&= C^2(A)\int _{-\infty }^\infty |q'|dx+O\left( \left( 2A+\frac{1}{\sqrt{\lambda }}\right) e^{-2\sqrt{\lambda } A}\right) ,\\ \int _0^\infty \frac{|\Psi _A|^2}{x^2}dx&\leqslant \frac{4C^2(A)}{A^2}\int _0^\infty |\Psi _A|^2dx +O\left( \frac{1}{A^2}e^{-\sqrt{\lambda } A}\right) ,\\ \int _0^\infty |\Psi _A|^{p+1}dx&= C^{p+1}(A)\int _{-\infty }^\infty |q|^{p+1}dx + O(e^{-2\sqrt{\lambda } A}). \end{aligned}$$

Hence for A large enough we get

$$\begin{aligned} E(\Psi _A)&=\frac{1}{2}\int _0^\infty |\Psi _A'|^2dx -\frac{c}{2}\int _0^\infty \frac{|\Psi _A|^2}{x^2}dx- \frac{1}{p+1}\int _0^\infty |\Psi _A|^{p+1}dx\\&\leqslant C^2(A)\left( \frac{1}{2}\int _{-\infty }^\infty |q'|^2dx- \frac{C^{p-1}(A)}{p+1}\int _{-\infty }^\infty |q|^{p+1}dx \right) \\&\quad -\frac{c}{2}\frac{4C^2(A)}{A^2}\int _0^\infty |\Psi _A|^2dx +O\left( \frac{1}{A^2}e^{-\sqrt{\lambda } A}\right) . \end{aligned}$$

Owing to the exponential decay of the last term, for large A we get

$$\begin{aligned} E(\Psi _A)\leqslant E(q) -\frac{2c}{A^2}\mu =I^\infty -\frac{2c}{A^2}\mu . \end{aligned}$$

Since \(0<c<1/4\) we get that \(E(\Psi _A)<I^\infty \), which concludes the proof. \(\square \)

We need the following version of the concentration-compactness principle. The proof follows the same way as in the classical case (see [15]).

Lemma 3.2

Let \(0<c<1/4\), and \(\{ u_n\}_{n\in {\mathbb {N}}} \subset H^1_0({\mathbb {R}}^+)\) be a sequence satisfying

$$\begin{aligned} \lim _{n\rightarrow \infty } \left\| u_n\right\| ^2_{L^2} = M \text { and } \lim _{n\rightarrow \infty } H(u_n) < \infty . \end{aligned}$$

Then there exists a subsequence \(\{u_n\}_{n\in {\mathbb {N}}}\) such that it satisfies one of the following alternatives.

(Vanishing) \(\lim _{n\rightarrow \infty }\left\| u_n\right\| _{L^p}\rightarrow 0\) for all \(p\in (2,\infty )\).

(Dichotomy) There are sequences \(\{v_n\}_{n\in {\mathbb {N}}}, \{w_n\}_{n\in {\mathbb {N}}}\) in \(H^1_0({\mathbb {R}}^+)\) and a constant \(\alpha \in (0,1)\) such that:

  1. (1)

    \({{\,\mathrm{dist}\,}}({{\,\mathrm{supp}\,}}(v_n),{{\,\mathrm{supp}\,}}(w_n))\rightarrow \infty \);

  2. (2)

    \(|v_n|+|w_n|\leqslant |u_n|\);

  3. (3)

    \(\sup _{n\in {\mathbb {N}}}(\left\| v_n\right\| _{H^1}+\left\| w_n\right\| _{H^1})<\infty \);

  4. (4)

    \(\left\| v_n\right\| _{L^2}^2\rightarrow \alpha M\) and \(\left\| w_n\right\| _{L^2}^2\rightarrow (1-\alpha )M\) as \(n\rightarrow \infty \);

  5. (5)

    \(\lim _{n\rightarrow \infty }\left| \int _0^\infty |u_n|^qdx-\int _0^\infty |v_n|^qdx-\int _0^\infty |w_n|^qdx\right| =0\) for all \(q\in [2,\infty )\);

  6. (6)

    \(\liminf _{n\rightarrow \infty }\{H(u_n)-H(v_n)-H(w_n)\}\geqslant 0\).

(Compactness) There exists a sequence \(y_n \in {\mathbb {R}}^+\), such that for any \(\varepsilon >0\) there is an \(R>0\) with the property that

$$\begin{aligned} \int _{(y_n-R,y_n+R)\cap {\mathbb {R}}^+} |u_n|^2\geqslant M-\varepsilon . \end{aligned}$$

for all \(n\in {\mathbb {N}}\).

We are now in a position to prove the following lemma.

Lemma 3.3

Let \(1<p<5\), \(0<c<1/4\), and \(\omega >0\). Then the infimum in (3.1) is attained. Additionally, all minimizing sequences are relatively compact, that is if \(\{u_n\}_{n\in {\mathbb {N}}}\) satisfies \(\left\| u_n\right\| _{L^2}^2\rightarrow \mu \) and \(E(u_n)\rightarrow I\) then there exists a subsequence \(\{u_n\}_{n\in {\mathbb {N}}}\) which converges to a minimizer \(u\in H^1_0({\mathbb {R}}^+)\).

Proof

Step 1. We first show that \(0>I>-\infty \). Let \(u\in \Gamma \). For \(\lambda >0\), we define \(u_\lambda (x)=\lambda ^{1/2}u(\lambda x) \in \Gamma \). Clearly,

$$\begin{aligned} E(u_\lambda )=\frac{\lambda ^2}{2}\left\| u'\right\| ^2_{L^2}-\frac{c\lambda ^2}{2}\int _0^\infty \frac{|u|^2}{x^2}dx-\frac{\lambda ^{(p-1)/2}}{p+1}\left\| u\right\| ^{p+1}_{L^{p+1}} \end{aligned}$$

Since \(1<p<5\), we can choose a small \(\lambda >0\) such that \(E(u_\lambda )<0\). Hence \(I<0\).

Since \(c\in (0,1/4)\), we have \(H(u)\sim \left\| u'\right\| ^2_{L^2}\). We get from the Gagliardo-Nirenberg inequality that there exists \(C>0\) such that for all \(u\in H^1_0({\mathbb {R}}^+)\)

$$\begin{aligned} \int _0^\infty |u|^{p+1}dx\leqslant C H(u)^{\frac{p-1}{4}}\left( \int _0^\infty |u|^2dx\right) ^{1+\frac{p-1}{4}}. \end{aligned}$$

Since \(1<p<5\), this yields that there exists \(\delta >0\) and \(K>0\) such that

$$\begin{aligned} E(u)\geqslant \delta \left\| u\right\| ^2_{H^1}-K \text { for all } u \in \Gamma , \end{aligned}$$
(3.3)

from which follows that \(I>-\infty \).

Every minimizing sequence is bounded in \(H^1_0({\mathbb {R}}^+)\) and bounded from below in \(L^{p+1}({\mathbb {R}}^+)\). Indeed, let \(\{u_n\}_{n\in {\mathbb {N}}}\subset \Gamma \) be a minimizing sequence, then by (3.3) it is bounded in \(H^1_0({\mathbb {R}}^+)\). Furthermore, for n large enough we have \(E(u_n)<I/2\), thus

$$\begin{aligned} \left\| u_n\right\| _{L^{p+1}}^{p+1}>-\frac{p+1}{2}I. \end{aligned}$$
(3.4)

Now \(I<0\), hence the result follows.

Step 2. We now verify that all minimizing sequences have a subsequence which converges to a limit u in \(H^1_0({\mathbb {R}}^+)\). Let \(\{u_n\}_{n\in {\mathbb {N}}}\) satisfy \(\left\| u_n\right\| ^2_{L^2}\rightarrow \mu \) and \(E(u_n)\rightarrow I\). Since every minimizing sequence is bounded in \(H^1_0({\mathbb {R}}^+)\), \(\{u_n\}_{n\in {\mathbb {N}}}\) has a weak-limit \(u\in L^p({\mathbb {R}}^+)\) . We can apply the concentration-compactness principle (see Lemma 3.2) to the sequence \(\{u_n\}_{n\in {\mathbb {N}}}\). We note that since the sequence is bounded from below in \(L^{p+1}({\mathbb {R}}^+)\) vanishing cannot occur.

Now let us assume that dichotomy occurs. Let \(\alpha \in (0,1)\), \(\{v_n\}_{n\in {\mathbb {N}}}\) and \(\{w_n\}_{n\in {\mathbb {N}}}\) sequences as in Lemma 3.2. It follows from (5) and (6) of Lemma 3.2 that

$$\begin{aligned} \liminf _{n\rightarrow \infty }(E(u_n)-E(v_n)-E(w_n))\geqslant 0, \end{aligned}$$

hence

$$\begin{aligned} \limsup _{n\rightarrow \infty }(E(v_n)+E(w_n))\leqslant I. \end{aligned}$$
(3.5)

Observe that for \(u\in H^1_0({\mathbb {R}}^+)\), and \(a>0\), we have

$$\begin{aligned} E(u)=\frac{1}{a^2}E(au)+\frac{a^{p-1}-1}{p+1}\int _0^\infty |u|^{p+1}dx. \end{aligned}$$

Let \(a_n=\sqrt{\mu }/\left\| v_n\right\| _{L^2}\) and \(b_k^2=\sqrt{\mu }/\left\| w_n\right\| _{L^2}\). Hence, \(a_nv_n \in \Gamma \) and \(b_nw_n\in \Gamma \), which implies

$$\begin{aligned} E(v_n)&\geqslant \frac{I}{a^2_n}+\frac{a_n^{p-1}-1}{p+1}\int _0^\infty |v_n|^{p+1}dx, \\ E(w_n)&\geqslant \frac{I}{b^2_n}+\frac{b_n^{p-1}-1}{p+1}\int _0^\infty |w_n|^{p+1}dx. \end{aligned}$$

Therefore

$$\begin{aligned} E(v_n)+E(w_n)\geqslant I(a_n^{-2}+b_n^{-2})+\frac{a_n^{p-1}}{p+1}\int _0^\infty |v_n|^{p+1}+\frac{b_n^{p-1}}{p+1}\int _0^\infty |w_n|^{p+1}. \end{aligned}$$

Now we observe \(a_n^{-2}\rightarrow \alpha \) and \(b_n^{-2}\rightarrow (1-\alpha )\) by (4) of Lemma 3.2. Since \(\alpha \in (0,1)\), we get that \(\theta = \min \{\alpha ^{-(p-1)/2};(1-\alpha )^{-(p-1)/2}) \}>1\). Property (5) of Lemma 3.2 and (3.4) implies

$$\begin{aligned} \liminf _{n\rightarrow \infty }(E(v_n)+E(w_n))\geqslant I+ \frac{\theta -1}{p+1}\liminf _{n\rightarrow \infty }\int _0^\infty |u_n|^{p+1}dx, \geqslant I+\frac{\theta -1}{2}>I, \end{aligned}$$

which contradicts (3.5). Hence the following holds: there exists a sequence \(y_n\in {\mathbb {R}}^+\), such that for any \(\varepsilon >0\) there exists \(R>0\) with the property that

$$\begin{aligned} \int _{(y_n-R,y_n+R)\cap {\mathbb {R}}^+}|u_n|^2\geqslant \mu -\varepsilon . \end{aligned}$$
(3.6)

for all \(n\in {\mathbb {N}}\).

We now show that \(\{y_n\}_{n\in {\mathbb {N}}}\) is bounded in \({\mathbb {R}}^+\). First we show that if \(y_n\rightarrow \infty \), then

$$\begin{aligned} \lim _{n\rightarrow \infty }\int _0^\infty \frac{|u_n|^2}{x^2}dx=0. \end{aligned}$$
(3.7)

Let us assume by contradiction that

$$\begin{aligned} \int _0^\infty \frac{|u_n|^2}{x^2}dx\geqslant \delta >0, \end{aligned}$$
(3.8)

which implies together with Hardy’s inequality that

$$\begin{aligned} H(u_n)\geqslant (1/4-c)\delta . \end{aligned}$$
(3.9)

Let us take \(\xi \in C^\infty ({\mathbb {R}}^+)\), such that for \({\tilde{R}}>0\) and \(a>0\) we have that \(\xi (r)=1\) for \(0\leqslant r \leqslant {\tilde{R}}\), \(\xi (r)=0\) for \(r\geqslant {\tilde{R}}+a\), and \(\left\| \xi '\right\| _{L^\infty }\leqslant 2/a\). We introduce \(u_{n,1}=u_n\cdot \xi \) and \(u_{n,2}=u_n\cdot (1-\xi )\). Clearly, \(u_{n,1}\in H^1_0({\mathbb {R}}^+)\), \(u_{n,2}\in H^1_0({\mathbb {R}}^+)\) and \(u_n=u_{n,1}+u_{n,2}\). Moreover, the following inequalities hold

$$\begin{aligned} |u'_{n,1}|^2&\leqslant 2(4a^{-2}|u_n|^2+|u'_n|^2), \\ |u'_{n,2}|^2&\leqslant 2(4a^{-2}|u_n|^2+|u'_n|^2). \end{aligned}$$

We obtain by direct calculation that

$$\begin{aligned} E(u_n)=E(u_{n,1})+E(u_{n,2})+\rho _n \end{aligned}$$

where

$$\begin{aligned} \rho _n&=\frac{1}{2}\int _{{\tilde{R}}}^{{\tilde{R}}+a} \left[ (|u_n'|^2-|u'_{n,1}|^2-|u'_{n,2}|^2)-\frac{c}{x^2}(|u_n|^2-|u_{n,1}|^2-|u_{n,2}|^2) \right] dx \\ {}&\quad -\frac{1}{p+1}\int _{{\tilde{R}}}^{{\tilde{R}}+a}(|u_n|^{p+1}-|u_{n,1}|^{p+1}-|u_{n,2}|^{p+1})dx. \end{aligned}$$

We show that there exists \({\tilde{R}}>0\) and \(a>1\), such that for n large enough \(|\rho _n|\leqslant (1/4-c)\frac{\delta }{4}\). First we observe by the properties of the cut-off that

$$\begin{aligned} \left| \frac{1}{2}\int _{{\tilde{R}}}^{{\tilde{R}}+a} (|u_n'|^2-|u'_{n,1}|^2-|u'_{n,2}|^2)dx\right| \leqslant \frac{5}{2} \int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n'|^2dx+\frac{8}{a^2}\int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n|^2dx. \end{aligned}$$

We claim that there exist \({\tilde{R}}>0\) and \(a>1\) such that for a subsequence \(\{u_{n_k}\}\) we have

$$\begin{aligned} \int _{{\tilde{R}}}^{{\tilde{R}}+a}|u'_{n_k}|^2dx< \frac{1}{20}(1/4-c)\delta . \end{aligned}$$
(3.10)

Suppose that this claim does not hold, that is for all \(R>0\), \(a>1\) there exists \(k\in {\mathbb {N}}\) such that for all \(n\geqslant k\) the following holds

$$\begin{aligned} \int _R^{R+a}|u'_n|^2dx\geqslant \frac{1}{20}(1/4-c)\delta . \end{aligned}$$

Let \((R_1,R_1+a_1)\). There exists \(k_1\in {\mathbb {N}}\), such that for all \(n\geqslant k_1\) we have

$$\begin{aligned} \int _{R_1}^{R_1+a_1}|u'_n|^2dx\geqslant \frac{1}{20}(1/4-c)\delta . \end{aligned}$$

Now let \(R_2>R_1+a_1\) and \(a_2>1\). Then by our assumption there exists \(k_2\in {\mathbb {N}}\), such that for all \(n\geqslant k_2\) it holds that

$$\begin{aligned} \int _{R_2}^{R_2+a_2}|u'_n|^2dx\geqslant \frac{1}{20}(1/4-c)\delta . \end{aligned}$$

Hence, there exists a subsequence \(\{v_{n_k}\}_{k\in {\mathbb {N}}}\) such that for all \(j\in \{1,2\}\) it holds that

$$\begin{aligned} \int _{R_j}^{R_j+a_j}|u'_{n_k}|^2dx\geqslant \frac{1}{20}(1/4-c)\delta \end{aligned}$$

for all \(k\in {\mathbb {N}}\). Therefore, we can construct for all \(l\in {\mathbb {N}}\) a subsequence \(\{u_{n_k}\}_{k\in {\mathbb {N}}}\), such that for all \(1\leqslant j\leqslant l\) there are disjoint intervals \(A_j=(R_j,R_j+a_j)\), such that

$$\begin{aligned} \int _{A_j}|u'_{n_k}|^2dx\geqslant \frac{1}{20}(1/4-c)\delta . \end{aligned}$$

Hence for all \(l\in {\mathbb {N}}\) there exists a subsequence \(\{ u_{n_k}\}_{k\in {\mathbb {N}}}\), such that for all \(k\in {\mathbb {N}}\) we have

$$\begin{aligned} \int _0^\infty |u'_{n_k}|^2dx\geqslant \sum _{j=1}^{l}\int _{A_j}|u'_{n_k}|^2dx\geqslant \frac{l}{20}(1/4-c)\delta . \end{aligned}$$

This implies that \(\int _0^\infty |u'_{n_k}|^2dx\rightarrow \infty \), which is a contradiction since \(\{u_n\}_{n\in {\mathbb {N}}}\) is bounded in \(H^1_0({\mathbb {R}}^+)\). Hence the assertion (3.10) is true. Now we note that

$$\begin{aligned} \int _0^R |u_n|^{p+1}dx \leqslant \left\| u_n\right\| ^{p-1}_{L^\infty }\int _0^R|u_n|^2dx. \end{aligned}$$

Since \(\{u_n\}_{n\in {\mathbb {N}}}\) is bounded in \(L^\infty ({\mathbb {R}}^+)\), in view of (3.6) we obtain for \(R>0\) given in (3.6) that

$$\begin{aligned} \int _0^R|u_n|^2dx \rightarrow 0 \quad \text {implies} \quad \int _0^R |u_n|^{p+1}dx\rightarrow 0. \end{aligned}$$
(3.11)

For large n we have \({\tilde{R}}+a<y_n-R\), since \(y_n\rightarrow \infty \) by our assumption. Now (3.11) implies

$$\begin{aligned}&\left| \frac{8}{a^2}\int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n|^2dx\right| +\left| \int _{{\tilde{R}}}^{{\tilde{R}}+a}\frac{c}{x^2}(|u_n|^2-|u_{n,1}|^2-|u_{n,2}|^2) dx\right| \nonumber \\&\qquad +\left| \frac{1}{p+1}\int _{{\tilde{R}}}^{{\tilde{R}}+a}(|u_n|^{p+1}-|u_{n,1}|^{p+1}-|u_{n,2}|^{p+1})dx\right| \nonumber \\&\quad \leqslant \left| \frac{8}{a^2}\int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n|^2dx\right| +\frac{c}{{\tilde{R}}^2}\left| \int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n|^2(1-\xi ^2-(1-\xi )^2) dx\right| \nonumber \\&\qquad +\left| \frac{1}{p+1}\int _{{\tilde{R}}}^{{\tilde{R}}+a}|u_n|^{p+1}(1-\xi ^{p+1}-(1-\xi )^{p+1})dx\right| \nonumber \\&\quad \leqslant \frac{(1/4-c)\delta }{8}. \end{aligned}$$
(3.12)

for large n. Now (3.10) and (3.12) implies

$$\begin{aligned} |\rho _n|\leqslant \frac{(1/4-c)\delta }{4}. \end{aligned}$$
(3.13)

Let us observe that \(\left\| u_{n,1}\right\| _{L^{p+1}}\rightarrow 0\) by (3.11). Hence

$$\begin{aligned} E(u_{n,1})=\frac{1}{2}H(u_{n,1})+o(1). \end{aligned}$$

Now let us notice that \(\mathrm {supp} (u_{n,2}) \subset ({\tilde{R}}, \infty )\). Moreover, in view of (3.6),

$$\begin{aligned} \int _0^\infty |u_{n,2}|^2dx = \int _{y_n-R}^\infty |u_{n,2}|^2dx+o(1). \end{aligned}$$

Hence

$$\begin{aligned} \int _0^\infty \frac{|u_{n,2}|^2}{x^2}dx= \int _{y_n-R}^\infty \frac{|u_{n,2}|^2}{x^2}dx + o(1)\leqslant \frac{\mu }{|y_n-R|^2}. \end{aligned}$$

Now \(y_n\rightarrow \infty \) implies that

$$\begin{aligned} E(u_{n,2})= E^\infty (u_{n,2})+o(1). \end{aligned}$$

Thus,

$$\begin{aligned} E(u_n)=\frac{1}{2}H(u_{n,1})+E^\infty (u_{n,2})+\rho _n+o(1). \end{aligned}$$

From the properties of the cut-off and (3.6), we get

$$\begin{aligned} \left\| u_{n,2}\right\| _{L^2}^2=\left\| u_n\right\| ^2_{L^2}-\left\| u_{n,1}\right\| ^2_{L^2}-2\mathop {{\mathrm{Re}}}\nolimits \int _{R'}^{R'+a} u_{n,1}{\bar{u}}_{n,2}dx \rightarrow \mu . \end{aligned}$$

Since \(\frac{1}{2}H(u_{n,1})+\rho _n>0\) by (3.9) and (3.13), we obtain

$$\begin{aligned} I=\lim _{n\rightarrow \infty }E(u_n)\geqslant \lim _{n\rightarrow \infty }E^\infty (u_{n,2})\geqslant I^\infty . \end{aligned}$$

which is a contradiction, hence (3.7) follows.

Now, from (3.7) we obtain

$$\begin{aligned}&\lim _{n\rightarrow \infty }\left( \frac{1}{2}\int _0^\infty |u'_n|^2dx-\frac{c}{2}\int _0^\infty \frac{|u_n|^2}{x^2}dx-\frac{1}{p+1}\int _0^\infty |u_n|^{p+1}dx\right) = \\&\qquad =\lim _{n\rightarrow \infty } \left( \frac{1}{2}\int _0^\infty |u'_n|^2dx-\frac{1}{p+1}\int _0^\infty |u_n|^{p+1}dx\right) . \end{aligned}$$

Hence

$$\begin{aligned} I\geqslant I^\infty , \end{aligned}$$

which is again a contradiction. Thus \(\{y_n\}_{n\in {\mathbb {N}}}\) is bounded and has an accumulation point \(y^*\in {\mathbb {R}}^+\). Therefore, it follows that for any \(\varepsilon >0\) there is \(R>0\) such that

$$\begin{aligned} \int _0^R|u_n|^2\geqslant \mu -\varepsilon . \end{aligned}$$

for all \(n\in {\mathbb {N}}\). Hence \(u_n\rightarrow u\) strongly in \(L^{2}({\mathbb {R}}^+)\). Moreover, since \(\{u_n\}\) is bounded in \(H^1_0({\mathbb {R}}^+)\) it is also strongly convergent in \(L^{p+1}({\mathbb {R}}^+)\). By the weak-lower semicontinuity of H (see [17]), it follows that \(E(u)\leqslant \lim _{n\rightarrow \infty }E(u_n)=I\). Hence \(E(u)=I\), and \(E(u_n)\rightarrow E(u)\) implies that \(H(u_n)\rightarrow H(u)\), which concludes that proof. \(\square \)

Remark 3.4

If \(c<0\), the infimum is not attained on the \(L^2\) constraint. Indeed, let us assume that there exists \(v\in H^1_0({\mathbb {R}}^+)\), such that \(\left\| v\right\| ^2_{L^2}=\mu \) and \(E(v)=I\). Then taking translates of v, i.e. \(v(\cdot -y)\) for \(y>0\), we get \(E(v(\cdot -y))<I\), which is a contradiction.

Lemma 3.5

Let \(0<c<1/4\), \(\omega >0\) and \(1<p<5\). Let \(\mu \) be defined by Lemma 2.10. Then \(u\in H^1_0({\mathbb {R}}^+)\) is a ground state solution of (2.1) if and only if u solves the minimization problem

$$\begin{aligned} {\left\{ \begin{array}{ll} u \in \Gamma , \\ S(u)=\inf \{S(v): v\in \Gamma \}. \end{array}\right. } \end{aligned}$$
(3.14)

Proof

Step 1. Let us first define

$$\begin{aligned} m_{{\mathcal {A}}}=\inf \{S(u) : u \in {\mathcal {A}}\}, \end{aligned}$$

and

$$\begin{aligned} m_{\Gamma }=\inf \{S(u) : u \in \Gamma \}. \end{aligned}$$

If \(u\in {\mathcal {G}}\), then \(S(u)=m_\Gamma \). By Lemma 2.10 we know that \(u\in \Gamma \), hence \(m_{{\mathcal {A}}}\leqslant m_\Gamma \).

Step 2. We claim that every solution of (3.14) belongs to \({\mathcal {A}}\). Indeed, let us consider a solution u to (3.14). There exists a Lagrange multiplier \(\lambda _1\in {\mathbb {R}}\) such that \(S'(u)=\lambda _1 u\). Hence there exists \(\lambda \in {\mathbb {R}}\) such that

$$\begin{aligned} -u'' -\frac{c}{x^2}u+ \lambda \omega u= |u|^{p-1}u. \end{aligned}$$
(3.15)

Indeed, since u is a solution of (3.14), and for \(\lambda >0\) let

$$\begin{aligned} u_\lambda (x)=\lambda ^{1/2}u(\lambda x). \end{aligned}$$

We have \(u_\lambda \in \Gamma \). Since \(u_1\) is a solution of (3.14), we get from (3.15) and Lemma 2.2 that

$$\begin{aligned} \frac{\partial }{\partial \lambda }S(u_\lambda )|_{\lambda =1}=\left\| u'\right\| ^2_{L^2}-c\left\| \frac{u}{x}\right\| ^2_{L^2}-\frac{p-1}{2(p+1)} \left\| u\right\| ^{p+1}_{L^{p+1}}=0. \end{aligned}$$
(3.16)

We can deduce directly from (3.15) and (3.16) that

$$\begin{aligned} \lambda \omega \mu =\frac{p+3}{p-1} H(u), \end{aligned}$$

which implies that \(\lambda >0\). Let us define v by

$$\begin{aligned} u(x)=\lambda ^{1/(p-1)}v(\lambda ^{1/2}x). \end{aligned}$$

By (3.16), \(v\in {\mathcal {A}}\), hence

$$\begin{aligned} S(v)\geqslant m_{{\mathcal {A}}}. \end{aligned}$$

We obtain simple calculation that

$$\begin{aligned} m_{\Gamma }=S(u)=\lambda ^{2/(p-1)+1/2}S(v) + (1-\lambda )\frac{\omega \mu }{2}. \end{aligned}$$

Hence,

$$\begin{aligned} m_{{\mathcal {A}}}\geqslant \lambda ^{\frac{2}{p-1}+\frac{1}{2}}m_{{\mathcal {A}}}+(1-\lambda )\frac{\omega \mu }{2}. \end{aligned}$$

Since u is a solution of (3.15), we obtain from Lemma 2.2 that \(m_{{\mathcal {A}}}\geqslant 0\). By Lemma 2.2 and Lemma 2.10 we have that

$$\begin{aligned} \frac{\omega \mu }{2}=\left( \frac{2}{p-1}+\frac{1}{2} \right) m_{{\mathcal {A}}}, \end{aligned}$$

hence

$$\begin{aligned} 0\geqslant \lambda ^{\frac{2}{p-1}+\frac{1}{2}} - \lambda \left( \frac{2}{p-1}+\frac{1}{2} \right) + \left( \frac{2}{p-1}-\frac{3}{2} \right) . \end{aligned}$$

The right hand side is always strictly positive, except if \(\lambda =1\). Thus, \(\lambda =1\), which implies together with (3.16) that \(u\in {\mathcal {A}}\).

Step 3. It follows from Step 2, that \(m_\Gamma \leqslant m_{{\mathcal {A}}}\), hence \(m_\Gamma =m_{{\mathcal {A}}}\). In particular, it follows that if \(u\in {\mathcal {G}}\), then \(u \in \Gamma \) and \(S(u)=m_{{\mathcal {A}}}\), thus u satisfies (3.14). Conversely, let u be the solution of (3.14). Then by Step 2 \(u\in {\mathcal {A}}\), and \(S(u)=m_\Gamma =m_{{\mathcal {A}}}\), hence \(u \in {\mathcal {G}}\). \(\square \)

Theorem 3.6

Let \(0<c<1/4\), \(\omega >0\), and \(1<p<5\). If \(\varphi \) is a ground state solution of (2.1), then the standing wave \(u(t,x)=e^{i\omega t}\varphi (x)\) is an orbitally stable solution of (1.1), i.e. for all \(\varepsilon >0\) there is \(\delta >0\), such that if \(u(0)\in H^1_0({\mathbb {R}}^+)\) satisfies \(\left\| \varphi -u(0)\right\| _{H^1}<\delta \), then the corresponding maximal solution u of (1.1) satisfies

$$\begin{aligned} \sup _{t\in {\mathbb {R}}}\inf _{\theta \in {\mathbb {R}}}\left\| u(t)-e^{i\theta }\varphi \right\| _{H^1}<\varepsilon . \end{aligned}$$

Proof

Assume by contradiction that there exist a sequence \(\{\varphi _n\}_{n\in {\mathbb {N}}}\subset H^1_0({\mathbb {R}}^+)\), a sequence \(\{t_n\}_{n\in {\mathbb {N}}}\subset {\mathbb {R}}\), and \(\varepsilon >0\), such that

$$\begin{aligned} \lim _{n\rightarrow \infty } \left\| \varphi _n-\varphi \right\| _{H^1}=0, \end{aligned}$$

and the corresponding maximal solution \(u_n\) of (1.1) with initial value \(\varphi _n\) satisfies

$$\begin{aligned} \inf _{\theta \in {\mathbb {R}}}\left\| u_n(t_n)-e^{i\theta }\varphi \right\| _{H^1}\geqslant \varepsilon . \end{aligned}$$

Set \(v_n=u_n(t_n)\). Applying Lemma 3.5, we obtain

$$\begin{aligned} \lim _{n\rightarrow \infty }\inf _{\varphi \in {\mathcal {G}}}\left\| v_n-\varphi \right\| _{H^1}\geqslant \varepsilon . \end{aligned}$$
(3.17)

By the conservation of charge and energy, we obtain

$$\begin{aligned} \left\| v_n\right\| ^2_{L^2}\rightarrow \mu , \textit{ and } E(v_n)\rightarrow I. \end{aligned}$$

Hence \(\{v_n\}_{n\in {\mathbb {N}}}\) is a minimizing sequence of (3.1). It follows from Lemma 3.3, that there exists a solution u of the problem (3.1), such that \(\left\| v_n-u\right\| _{H^1}\rightarrow 0\). By Lemma 3.5 we obtain that \(u\in {\mathcal {G}}\), which contradicts (3.17). \(\square \)

4 Instability

In this section we assume that \(p \geqslant 5\). Let us define for \(v\in H^1_0({\mathbb {R}}^+)\) the functional

$$\begin{aligned} Q(v)= \left\| v'\right\| ^2_{L^2}-c\left\| \frac{v}{x}\right\| ^2_{L^2}-\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

In Lemma 2.2 we have shown that if v is a solution of (2.1), then \(Q(v)=0\). First, we prove the virial identities.

Proposition 4.1

Let \(u_0\in H^1_0({\mathbb {R}}^+)\) be such that \(x u_0 \in L^2({\mathbb {R}}^+)\) and u be the corresponding maximal solution to (1.1). Then \(x u(t) \in L^2 ({\mathbb {R}}^+)\) for any \(t\in (-T_{\mathrm {min}},T_{\mathrm {max}})\). Moreover, the following identities hold for all \(v\in H^1_0({\mathbb {R}}^+)\):

$$\begin{aligned} \frac{\partial }{\partial t} \left\| x u(t)\right\| ^2_{L^2}&= 4\mathop {{\mathrm{Im}}}\nolimits \int _0^\infty {\bar{u}}(t) x u'(t) dx, \\ \frac{\partial ^2}{\partial t^2} \left\| x u(t)\right\| ^2_{L^2}&= 8 Q(u(t)). \end{aligned}$$

Proof

The proof follows the same line as in [6]. \(\square \)

Proposition 4.2

Let \(p\geqslant 5\) and let \(u_0\in H^1_0({\mathbb {R}}^+)\) be such that

$$\begin{aligned} xu_0\in L^2({\mathbb {R}}^+) \text { and } E(u_0)<0. \end{aligned}$$

Then the maximal solution u to (1.1) with initial condition \(u_0\) blows up in finite time.

Proof

First, let us note that

$$\begin{aligned} Q(u(t))=2E(u(t))+\frac{5-p}{2(p+1)}\left\| u(t)\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

Since \(p\geqslant 5\), we get by the conservation of the energy that

$$\begin{aligned} Q(u(t))\leqslant 2E(u_0)<0 \text { for all } t\in (-T_{\mathrm {min}},T_{\mathrm {max}}). \end{aligned}$$

Hence, Proposition 4.1 implies that

$$\begin{aligned} \frac{\partial ^2}{\partial t^2} \left\| xu(t)\right\| ^2_{L^2}\leqslant 16 E(u_0) \text { for all } t\in (-T_{\mathrm {min}},T_{\mathrm {max}}). \end{aligned}$$

Integrating twice, we get

$$\begin{aligned} \left\| xu(t)\right\| ^2_{L^2}\leqslant 8E(u_0)t^2 +\left( 4\mathop {{\mathrm{Im}}}\nolimits \int _0^\infty {\bar{u}}_0 x u'_0dx \right) t +\left\| xu_0\right\| ^2_{L^2} \end{aligned}$$
(4.1)

The main coefficient of the second order polynomial on the right hand side is negative. Thus, it is negative for |t| large, what contradicts with \(\left\| xu(t)\right\| ^2_{L^2}\geqslant 0\) for all t. Therefore, \(-T_{\mathrm {min}}>-\infty \) and \(T_{\mathrm {max}}<+\infty \). \(\square \)

Theorem 4.3

Assume that \(\omega >0\) and \(p=5\). Then for any solution \(\varphi \in H^1_0({\mathbb {R}}^+)\) of (2.1) the standing wave \(e^{i\omega t} \varphi (x)\) is unstable by blow-up.

Proof

Since \(p=5\), we have for all \(v\in H^1_0({\mathbb {R}}^+)\), that \(2E(v)=Q(v)\). Hence from Lemma 2.2 we get that

$$\begin{aligned} E(\varphi )=0. \end{aligned}$$

Let us define \(\varphi _{n,0}=\left( 1+\frac{1}{n}\right) \varphi \). It is easy to see that \(E(\varphi _{n,0})<0\). By Lemma 2.1 we know that \(x \varphi _{n,0}\in L^2({\mathbb {R}}^+)\). The conclusion follows from Proposition 4.2. \(\square \)

Theorem 4.4

Let \(p>5\). Then for any ground state solution \(\varphi \) to (2.1), the corresponding standing wave \(e^{i\omega t}\varphi (x)\) is orbitally unstable.

We need to prove a series of Lemmas to establish Theorem 4.4.

Lemma 4.5

Let \(v\in H^1_0({\mathbb {R}}^+)\setminus \{0\}\) such that \(Q(v)\leqslant 0\), and set \(v_\lambda (x)=\lambda ^{1/2}v(\lambda x)\) for \(\lambda >0\). Then there exists \(\lambda ^* \in (0,1]\) such that the following assertions hold:

  1. (1)

    \(Q(v_{\lambda ^*})=0\).

  2. (2)

    \(\lambda ^*=1\) if and only if \(Q(v)=0\).

  3. (3)

    \(\frac{\partial }{\partial \lambda } S(v_\lambda ) =\frac{1}{\lambda } Q(v_\lambda )\).

  4. (4)

    \(\frac{\partial }{\partial \lambda } S(v_\lambda ) >0\) for all \(\lambda \in (0,\lambda ^*)\), and \(\frac{\partial }{\partial \lambda } S(v_\lambda ) <0 \) for all \(\lambda \in (\lambda ^*, +\infty )\).

  5. (5)

    The function \((\lambda ^*, +\infty )\ni \lambda \mapsto S(v_\lambda )\) is concave.

Proof

We get that by the scaling properties of \(\lambda \mapsto Q(v_\lambda )\) that

$$\begin{aligned} Q(v_\lambda )=\lambda ^2\left\| v'\right\| ^2_{L^2}-\lambda ^2 c \left\| \frac{v}{x}\right\| ^2_{L^2}-\lambda ^{\frac{p-1}{2}}\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

We get from the Hardy inequality that for \(c\in (0,1/4)\)

$$\begin{aligned}&(1-4c)\lambda ^2\left\| v'\right\| ^2_{L^2}-\lambda ^{\frac{p-1}{2}}\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}} \\&\quad \leqslant Q(v_\lambda )\leqslant \lambda ^2\left\| v'\right\| ^2_{L^2}-\lambda ^{\frac{p-1}{2}}\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

Since \(p>5\), there exists \(\lambda \in (0,1]\) small enough, such that \(Q(v_\lambda )>0\). Hence, there exists \(\lambda ^*\in (0,1]\), such that \(Q(v_{\lambda ^*})=0\). This proves (1). To prove (2), we first note that if \(\lambda ^*=1\), then clearly \(Q(v)=0\). Now assume that \(Q(v)=0\). Then

$$\begin{aligned} Q(v_\lambda )&=\lambda ^2Q(v)+(\lambda ^2-\lambda ^{\frac{p-1}{2}})\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}} \\&=(\lambda ^2-\lambda ^{\frac{p-1}{2}})\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}, \end{aligned}$$

which is positive for all \(\lambda \in (0,1)\), since \(p>5\). Hence, (2) follows. (3) follows form simple calculation:

$$\begin{aligned} \frac{\partial }{\partial \lambda } S(v_\lambda )&=\lambda \left\| v'\right\| ^2_{L^2}-\lambda c\left\| \frac{v}{x}\right\| ^2_{L^2}-\lambda ^{\frac{p-1}{2}-1}\frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}} \\&=\frac{1}{\lambda }Q(v_\lambda ). \end{aligned}$$

To show (4), we note that

$$\begin{aligned} Q(v_\lambda )=\frac{\lambda ^2}{(\lambda ^*)^2} Q(v_{\lambda ^*}) + \lambda ^2\left( (\lambda ^*)^{\frac{p-5}{2}} -\lambda ^{\frac{p-5}{2}} \right) \frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

Since \(p>5\) and \(Q(v_{\lambda ^*})=0\), we get that \(\lambda >\lambda ^*\) implies \(Q(v_\lambda )<0\), and \(\lambda <\lambda ^*\) implies \(Q(v_\lambda )>0\). This and (3), implies (4).

Finally, we get by simple calculation that

$$\begin{aligned} \frac{\partial ^2}{\partial \lambda ^2} S(v_\lambda )= \frac{1}{\lambda ^2} Q(v_\lambda )-\lambda ^{\frac{p-5}{2}}\left( \frac{p-1}{2}-2 \right) \frac{p-1}{2(p+1)}\left\| v\right\| ^{p+1}_{L^{p+1}}. \end{aligned}$$

Since \(p>5\), we obtain for \(\lambda >\lambda ^*\) that \(\frac{\partial ^2}{\partial \lambda ^2} S(v_\lambda )<0\) which concludes the proof of (5). \(\square \)

To prove orbital instability we prove a new variational characterization of the ground state. Let us define the following set

$$\begin{aligned} {\mathcal {M}}=\{v\in H^1_0({\mathbb {R}}^+)\setminus \{0\}: Q(v)=0, J(v)\leqslant 0\}, \end{aligned}$$

and the corresponding minimization problem

$$\begin{aligned} d=\inf _{W\in {\mathcal {M}}} S(W). \end{aligned}$$

Then we have the following.

Lemma 4.6

The following equality holds:

$$\begin{aligned} m=d, \end{aligned}$$

where m is defined by (2.4).

Proof

Let \(v\in {\mathcal {G}}\). Since v solves (2.1), by Lemma 2.2 we have that \(Q(v)=J(v)=0\), hence \({\mathcal {G}}\subset {\mathcal {M}}\), and

$$\begin{aligned} d \leqslant m. \end{aligned}$$

Let now \(v\in {\mathcal {M}}\). Assume first, that \(J(v)=0\). In this case \(v\in {\mathcal {N}}\), and \(m\leqslant S(v)\). Let us assume that \(J(v)<0\). Then for \(v_\lambda (x)=\lambda ^{1/2}v(\lambda x)\) we have

$$\begin{aligned} J(v_\lambda )=\lambda ^2\left\| v'\right\| ^2_{L^2}-\lambda ^2c\left\| \frac{v}{x}\right\| ^2_{L^2}+\omega \left\| v\right\| ^2_{L^2}-\lambda ^{(p-1)/2}\left\| v\right\| ^{p+1}_{L^{p+1}}, \end{aligned}$$

and \(\lim _{\lambda \downarrow 0} J(v_\lambda )>\omega \left\| v\right\| ^2_{L^2}\), thus there exists \(\lambda _1\in (0,1)\), such that \(J(v_{\lambda _1})=0\). By Proposition 2.9

$$\begin{aligned} m\leqslant S(v_{\lambda _1}). \end{aligned}$$

From \(Q(v)=0\) and Lemma 4.5 we have

$$\begin{aligned} S(v_{\lambda _1})\leqslant S(v), \end{aligned}$$

hence \(m\leqslant S(v)\) for all \(v\in {\mathcal {M}}\). Therefore \(m\leqslant d\), which concludes the proof. \(\square \)

We now define the manifold

$$\begin{aligned} {\mathcal {J}}=\{u\in H^1_0({\mathbb {R}}^+)\setminus \{0\}: J(u)<0, Q(u)<0, S(u)< d\}. \end{aligned}$$

We will prove the invariance of \({\mathcal {J}}\) under the flow of (1.1).

Lemma 4.7

Let \(u_0 \in {\mathcal {J}}\) and \(u\in C((-T_{\mathrm {min}},T_{\mathrm {max}}), H^1_0({\mathbb {R}}^+))\) the corresponding solution to (1.1). Then \(u(t)\in {\mathcal {J}}\) for all \(t\in (-T_{\mathrm {min}}, T_{\mathrm {max}})\).

Proof

Let \(u_0\in {\mathcal {J}}\) and \(u\in C((-T_{\mathrm {min}}, T_{\mathrm {max}}), H^1_0({\mathbb {R}}^+))\) the corresponding maximal solution. Since S is conserved under the flow of (1.1) we have for all \(t\in (-T_{\mathrm {min}}, T_{\mathrm {max}})\) that

$$\begin{aligned} S(u(t))=S(u_0)<d. \end{aligned}$$

We prove the assertion by contradiction. Suppose that there exists \(t\in (-T_{\mathrm {min}}, T_{\mathrm {max}})\) such that

$$\begin{aligned} J(u(t))\geqslant 0. \end{aligned}$$

Then, since J and u are continuous, there exists \(t_0\in (-T_{\mathrm {min}}, T_{\mathrm {max}})\) such that

$$\begin{aligned} J(u(t_0))=0, \end{aligned}$$

thus \(u(t_0) \in {\mathcal {N}}\). Then by Proposition 2.9 we have that

$$\begin{aligned} S(u(t_0))\geqslant d, \end{aligned}$$

which is a contradiction, thus \(J(u(t))<0\) for all \(t \in (-T_{\mathrm {min}},T_{\mathrm {max}})\). Let us suppose now that for some \(t\in (-T_{\mathrm {min}}, T_{\mathrm {max}})\) we have

$$\begin{aligned} Q(u(t))\geqslant 0. \end{aligned}$$

Again, by continuity, there exists \(t_1 \in (-T_{\mathrm {min}}, T_{\mathrm {max}})\) such that

$$\begin{aligned} Q(u(t_1))=0. \end{aligned}$$

Hence we that \(Q(u(t_1))=0\), and \(J(u(t_1))<0\). Therefore, by Lemma 4.6

$$\begin{aligned} S(u(t_1))\geqslant d, \end{aligned}$$

which is a contradiction. Hence,

$$\begin{aligned} Q(u(t))<0 \end{aligned}$$

for all \(t \in (-T_{\mathrm {min}},T_{\mathrm {max}})\), which concludes the proof. \(\square \)

Lemma 4.8

Let \(u_0 \in {\mathcal {J}}\) and \(u\in C((-T_{\mathrm {min}},T_{\mathrm {max}}),H^1_0({\mathbb {R}}^+))\). Then there exists \(\varepsilon >0\) such that \(Q(u(t))\leqslant -\varepsilon \) for all \(t\in (-T_{\mathrm {min}},T_{\mathrm {max}})\).

Proof

Let \(u_0 \in {\mathcal {J}}\) and let us define \(v:=u(t)\) and \(v_\lambda (x)=\lambda ^{1/2}v(\lambda x)\). By Lemma 4.5, there exists \(\lambda _0<1\) such that \(Q(v_{\lambda ^*})=0\). If \(J(v_{\lambda ^*})\leqslant 0\), then by Lemma 4.7 we get \(S(v_{\lambda ^*})\geqslant m.\) On the other hand, if \(J(v_{\lambda ^*})>0\), there exists \(\lambda _1\in (\lambda ^*,1)\), such that \(J(\lambda _1)=0\) and we replace \(\lambda ^*\) with \(\lambda _1\). In this case, by Lemma 4.6 we get \(S(v_{\lambda ^*})\geqslant m\). In conclusion, in both cases we obtain

$$\begin{aligned} S(v_{\lambda ^*})\geqslant d. \end{aligned}$$
(4.2)

By Lemma 4.5 we know that \(\lambda \mapsto S(v_\lambda )\) is concave on \((\lambda ^*, +\infty )\), thus

$$\begin{aligned} S(v)-S(v_{\lambda ^*})\geqslant (1-\lambda ^*)\frac{\partial }{\partial \lambda } S(v_\lambda )\Big |_{\lambda =1}. \end{aligned}$$
(4.3)

From Lemma 4.5 we have

$$\begin{aligned} \frac{\partial }{\partial \lambda }S(v_\lambda )\Big |_{\lambda =1}=Q(v). \end{aligned}$$
(4.4)

Moreover, since \(Q(v)<0\) and \(\lambda ^*\in (0,1)\), we have

$$\begin{aligned} (1-\lambda ^*)Q(v) > Q(v). \end{aligned}$$
(4.5)

Combining (4.2)–(4.5), we obtain

$$\begin{aligned} S(v)-d > Q(v). \end{aligned}$$

Define \(-\varepsilon = S(v)-d\). Then \(\varepsilon >0\), since \(v\in {\mathcal {J}}\). Owing to the conservation of the energy and mass, \(\varepsilon >0\) is independent from t, which concludes the proof. \(\square \)

Lemma 4.9

Let us take \(u_0\in {\mathcal {J}}\) such that \(x u_0 \in L^2({\mathbb {R}}^+)\). Then the maximal solution \(u\in C((-T_{\mathrm {min}},T_{\mathrm {max}}),H^1_0({\mathbb {R}}^+))\) corresponding to the initial value problem (1.1) blows up in finite time.

Proof

From Lemma 4.8 we know that there exists \(\varepsilon >0\) such that

$$\begin{aligned} Q(u(t))< -\varepsilon \text { for } t\in (-T_{\mathrm {min}},T_{\mathrm {max}}). \end{aligned}$$

From Proposition 4.1 we know that \(\frac{\partial ^2}{\partial t^2}\left\| xu(t)\right\| ^2_{L^2}=8Q(u(t))\), and by integration we get

$$\begin{aligned} \left\| xu(t)\right\| ^2_{L^2}\leqslant -4\varepsilon t^2 + C_1t+C_2. \end{aligned}$$
(4.6)

The right hand side of (4.6) is negative for large |t|, which contradicts with \(\left\| x u(t)\right\| ^2_{L^2}>0\) for all t. Therefore, \(T_{\mathrm {min}}>-\infty \) and \(T_{\mathrm {max}}<\infty \) and by local well-posedness it follows that

$$\begin{aligned} \lim _{t\downarrow -T_{\mathrm {min}}}\left\| u(t)\right\| _{H^1}=+\infty , \text { and } \lim _{t\uparrow T_{\mathrm {max}}} \left\| u(t)\right\| _{H^1}=+\infty . \end{aligned}$$

\(\square \)

Proof of Theorem 4.4

Let \(\varphi \in {\mathcal {G}}\). Owing to Lemma 4.9, it suffices to show that there exists a sequence \(\{\varphi _\lambda \}\subset {\mathcal {J}}\), which converges to \(\varphi \) in \(H^1_0({\mathbb {R}}^+)\). Let us put \(\varphi _\lambda (x)=\lambda ^{1/2}\varphi (\lambda x)\). By Lemma 4.5\(\{\varphi _\lambda \}\subset {\mathcal {J}}\) for all \(\lambda \in (0,1)\). Additionally, by Proposition 2.1, \(\varphi \) decays exponentially at infinity, and so does \(\varphi _\lambda \). Therefore, \(x\varphi _\lambda \in L^2({\mathbb {R}}^+)\). Clearly, \(\varphi _\lambda \rightarrow \varphi \) as \(\lambda \rightarrow 0\), and by Lemma 4.9 the maximal solution of (1.1) corresponding to \(\varphi _\lambda \), blows up in finite time for all \(\lambda \in (0,1)\). Hence, the conclusion follows. \(\square \)