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

On stability and instability of standing waves for 2d-nonlinear Schrödinger equations with point interaction

Noriyoshi Fukaya Department of Mathematics, Faculty of Science Division I, Tokyo University of Science, Tokyo, 162-8601, Japan [email protected] Vladimir Georgiev Dipartimento di Matematica Università di Pisa Largo B. Pontecorvo 5, 56100 Pisa, Italy
and
Faculty of Science and Engineering
Waseda University
3-4-1, Okubo, Shinjuku-ku, Tokyo 169-8555
Japan and IMI–BAS, Acad. Georgi Bonchev Str., Block 8, 1113 Sofia, Bulgaria
[email protected]
 and  Masahiro Ikeda Center for Advanced Intelligence Project, RIKEN, Japan/Department of Mathematics, Faculty of Science and Technology, Keio University, 3-14-1 Hiyoshi, Kohoku-ku, Yokohama 223-8522, Japan [email protected]/[email protected]
Abstract.

We study existence and stability properties of ground-state standing waves for two-dimensional nonlinear Schrödinger equation with a point interaction and a focusing power nonlinearity. The Schrödinger operator with a point interaction (Δα)α(-\Delta_{\alpha})_{\alpha\in\mathbb{R}} describes a one-parameter family of self-adjoint realizations of the Laplacian with delta-like perturbation. The operator Δα-\Delta_{\alpha} always has a unique simple negative eigenvalue. We prove that if the frequency of the standing wave is close to the negative eigenvalue, it is stable. Moreover, if the frequency is sufficiently large, we have the stability in the L2L^{2}-subcritical or critical case, while the instability in the L2L^{2}-supercritical case.

Key words and phrases:
nonlinear Schrödinger equation, point interaction, standing waves, stability, instability
2020 Mathematics Subject Classification:
35Q55; 35B35

1. Introduction

We consider the following nonlinear Schrödinger equation (NLS) with a focusing power nonlinearity in two spatial dimension:

(1.1) itu=Δαu|u|p1u,(t,x)×2,i\partial_{t}u=-\Delta_{\alpha}u-|u|^{p-1}u,\quad(t,x)\in\mathbb{R}\times\mathbb{R}^{2},

where p>1p>1 and Δα-\Delta_{\alpha} is the Laplacian with a point interaction with strength α\alpha\in\mathbb{R} at the origin (see (1.2)–(1.5) for the precise definition).

The study of the Laplace operator with point interactions in N\mathbb{R}^{N} (N=1N=1, 22, or 33) seems to become intensive research area in the last decades. The first rigorous attempt to define and study the spectral properties of these operators was done in [14] by Berezin and Faddeev in 1961. Their study was extended in the work [10] by Albeverio and Høegh-Krohn in 1981.

A simplest way to introduce a singular perturbation at a point is to consider the potential perturbation Δ+Vε-\Delta+V_{\varepsilon} of the Laplace operator with regular potentials VεV_{\varepsilon} spiking up and shrinking around a point in the limit ε0\varepsilon\to 0. This approach is well-known for dimension N=1N=1 [9], N=2N=2 [7], and N=3N=3 [6] (we also refer to [8] for a comprehensive overview). In the case of one singular point, this approximation gives a self-adjoint operator Δα-\Delta_{\alpha} depending on the parameter α\alpha\in\mathbb{R}. To be more precise, starting with the symmetric operator of Δ|Cc(N{0})-\Delta|_{C_{c}^{\infty}(\mathbb{R}^{N}\setminus\{0\})}, one can characterize all its nontrivial self-adjoint extension on L2(N)L^{2}(\mathbb{R}^{N}) by means of a parameter α\alpha\in\mathbb{R}. A detailed overview of the construction and the main properties of Δα-\Delta_{\alpha} can be found in [18]. Fractional powers of the Laplace operator Δα-\Delta_{\alpha} with singular perturbations are studied in [35, 49].

The dispersive properties of the Schrödinger group (eitΔα)t(e^{it\Delta_{\alpha}})_{t\in\mathbb{R}} have been studied intensively in the last years. In one-dimensional case, the dispersive and Strichartz estimates and boundedness wave operators have been studied, for example, in [4, 27, 39]. In higher dimensional case, such properties have been studied recently by [22, 59] in N=2N=2, [25] in N=3N=3. See also [24] for the weighted dispersive estimate in three dimension.

By using these dispersive properties, one can establish existence of nonlinear evolution flow. The local well-posedness for (1.1) (in two or three-dimensional cases) in the domain D(Δα)D(-\Delta_{\alpha}) has been studied in [18]. See, e.g., [4, 27] for the one-dimensional case.

Recently, the nonlinear dynamics around standing waves in one dimension are intensively studied in various contexts depending on whether the potential is attractive or repulsive. In the attractive regime, the orbital stability and instability of standing waves have been studied by [34, 36, 41], the asymptotic stability has been done in [23, 46, 47], and the strong instability has been done in [51, 29]. In the repulsive regime, the orbital stability and instability have been studied by [31, 45], and the global dynamics below ground states were studied in [40].

As related topics, we mention the NLS with the concentrated nonlinearity. See [3, 17, 42] for N=1N=1, [1, 2] for N=2N=2, and [5] for N=3N=3.

However, much less is known about the nonlinear dynamics and the existence and stability/instability properties of the ground states for (1.1). Up to our knowledge, there are no results treating such properties in the two-dimensional case. The main goal of the work is to cover this case and establish fundamental properties of ground states for (1.1) such as existence, uniqueness, nondegeneracy, and its orbital stability/instability.

To state our main results, we shortly give the definition of the operator Δα-\Delta_{\alpha} in two dimension (see [8, Chapter I.5] for more details). The class of self-adjoint extensions in L2(2)L^{2}(\mathbb{R}^{2}) of the positive and densely defined symmetric operator Δ|C0(2{0})-\Delta|_{C^{\infty}_{0}(\mathbb{R}^{2}\setminus\{0\})} is a one-parameter family of operators (Δα)α(,+](-\Delta_{\alpha})_{\alpha\in(-\infty,+\infty]}. The extension Δα=-\Delta_{\alpha=\infty} is the Friedrichs extension and is precisely the free Laplacian on L2(2)L^{2}(\mathbb{R}^{2}) with domain H2(2)H^{2}(\mathbb{R}^{2}). All other extensions for α\alpha\in\mathbb{R} represent nontrivial operators with a point interaction at the origin, and are characterized explicitly by

(1.2) D(Δα)={f+f(0)βα(λ)Gλ:fH2(2)},\displaystyle D(-\Delta_{\alpha})=\left\{f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\colon\,f\in H^{2}(\mathbb{R}^{2})\right\},
(1.3) (Δα+λ)g=(Δ+λ)ffor g=f+f(0)βα(λ)GλD(Δα).\displaystyle(-\Delta_{\alpha}+\lambda)g=(-\Delta+\lambda)f\quad\text{for }g=f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\in D(-\Delta_{\alpha}).

Here λ>0\lambda>0 is a fixed constant with λeα\lambda\neq-e_{\alpha} (see (1.11) for the definition of eαe_{\alpha}),

(1.4) βα(λ):=α+γ2π+12πlnλ2,\beta_{\alpha}(\lambda)\mathrel{\mathop{:}}=\alpha+\frac{\gamma}{2\pi}+\frac{1}{2\pi}\ln\frac{\sqrt{\lambda}}{2},

γ>0\gamma>0 denoting Euler–Mascheroni constant, and GλG_{\lambda} is the Green function of Δ+λ-\Delta+\lambda on 2\mathbb{R}^{2}, defined by the distributional relation (Δ+λ)Gλ=δ(-\Delta+\lambda)G_{\lambda}=\delta, that is

(1.5) Gλ(x):=12π1[1|ξ|2+λ](x)=1(2π)22eixξ|ξ|2+λdξ,G_{\lambda}(x)\mathrel{\mathop{:}}=\frac{1}{2\pi}\mathcal{F}^{-1}\biggl{[}\frac{1}{|\xi|^{2}+\lambda}\biggr{]}(x)=\frac{1}{(2\pi)^{2}}\int_{\mathbb{R}^{2}}\frac{e^{ix\cdot\xi}}{|\xi|^{2}+\lambda}\,d\xi,

where 1\mathcal{F}^{-1} is the inverse Fourier transform. Note that f(0)f(0) makes sense for fH2(2)f\in H^{2}(\mathbb{R}^{2}) from the embedding H2(2)C(2)H^{2}(\mathbb{R}^{2})\hookrightarrow C(\mathbb{R}^{2}).

From the the expression (1.5), we can easily check that

(1.6) GλH1ε(2),GλH1(2),\displaystyle G_{\lambda}\in H^{1-\varepsilon}(\mathbb{R}^{2}),\quad G_{\lambda}\notin H^{1}(\mathbb{R}^{2}),
(1.7) GλGμH3ε(2),GλGμH3(2)if λμ\displaystyle G_{\lambda}-G_{\mu}\in H^{3-\varepsilon}(\mathbb{R}^{2}),\quad G_{\lambda}-G_{\mu}\notin H^{3}(\mathbb{R}^{2})\quad\text{if $\lambda\neq\mu$}

for all ε>0\varepsilon>0 and λ,μ>0\lambda,\mu>0. The function GλG_{\lambda} is also represented as

(1.8) Gλ(x)=12πK0(λ|x|),G_{\lambda}(x)=\frac{1}{2\pi}K_{0}(\sqrt{\lambda}|x|),

where K0K_{0} is the modified Bessel function of the second kind (or the Macdonald function) of order zero. From the expression (1.8), GλG_{\lambda} is positive, radial, and strictly decreasing function with

(1.9) Gλ(r){ln(λr)as r0,r1/2eλras rG_{\lambda}(r)\sim\left\{\begin{array}[]{@{}c l@{}}-\ln(\sqrt{\lambda}r)&\text{as }r\to 0,\\[3.0pt] r^{-1/2}e^{-\sqrt{\lambda}r}&\text{as }r\to\infty\end{array}\right.

(see [26, Chapter 10] for more properties of KνK_{\nu}).

From the property (1.7) we can check that the definition of Δα-\Delta_{\alpha} is independent of λ\lambda. Indeed, if g=f+f(0)βα(λ)1Gλg=f+f(0)\beta_{\alpha}(\lambda)^{-1}G_{\lambda} for some fH2(2)f\in H^{2}(\mathbb{R}^{2}) and λeα\lambda\neq-e_{\alpha}, then for μeα\mu\neq-e_{\alpha}, we have the decomposition

g=f~+f~(0)βα(μ)Gμ,where f~:=f+f(0)βα(λ)(GλGμ)H2(2),g=\widetilde{f}+\frac{\widetilde{f}(0)}{\beta_{\alpha}(\mu)}G_{\mu},\quad\text{where }\widetilde{f}\mathrel{\mathop{:}}=f+\frac{f(0)}{\beta_{\alpha}(\lambda)}(G_{\lambda}-G_{\mu})\in H^{2}(\mathbb{R}^{2}),

and the relation

(Δα+μ)(f~+f~(0)βα(μ)Gμ)=(Δ+λ)f+(μλ)(f+f(0)βα(λ)Gλ).(-\Delta_{\alpha}+\mu)\biggl{(}\widetilde{f}+\frac{\widetilde{f}(0)}{\beta_{\alpha}(\mu)}G_{\mu}\biggr{)}=(-\Delta+\lambda)f+(\mu-\lambda)\biggl{(}f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\biggr{)}.

This means the independence.

The spectral properties of Δα-\Delta_{\alpha} are also known (see [8, Theorem 5.4]). The essential spectrum of Δα-\Delta_{\alpha} are given by

(1.10) σess(Δα)=σac(Δα)=[0,+),\displaystyle\sigma_{\mathrm{ess}}(-\Delta_{\alpha})=\sigma_{\mathrm{ac}}(-\Delta_{\alpha})=[0,+\infty), σsc(Δα)=,\displaystyle\sigma_{\mathrm{sc}}(-\Delta_{\alpha})=\emptyset,

where σac\sigma_{\mathrm{ac}} and σsc\sigma_{\mathrm{sc}} denote the set of the absolutely and singularly continuous spectrum, respectively. The operator Δα-\Delta_{\alpha} has a simple negative eigenvalue

(1.11) σp(Δα)={eα},\displaystyle\sigma_{\mathrm{p}}(-\Delta_{\alpha})=\{e_{\alpha}\}, eα:=4e4πα2γ.\displaystyle e_{\alpha}\mathrel{\mathop{:}}=-4e^{-4\pi\alpha-2\gamma}.

In this sense, Δα-\Delta_{\alpha} can be regarded as an Schrödinger operator with an attractive potential. The normalized eigenfunction corresponding to the eigenvalue eαe_{\alpha} is

χα:=GeαGeαL2.\chi_{\alpha}\mathrel{\mathop{:}}=\frac{G_{-e_{\alpha}}}{\|G_{-e_{\alpha}}\|_{L^{2}}}.

Note that βα(λ)\beta_{\alpha}(\lambda) defined in (1.4) is expressed as

βα(λ)=14πlnλeα,\beta_{\alpha}(\lambda)=\frac{1}{4\pi}\ln\frac{\lambda}{-e_{\alpha}},

so one has

βα(λ)>0λ>eα.\beta_{\alpha}(\lambda)>0\iff\lambda>-e_{\alpha}.

The energy space (or the form domain) Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) associated with Δα-\Delta_{\alpha} can be characterized by general results on the Kreĭn–Višik–Birman extension theory (see e.g. [48]), and it is given explicitly by

(1.12) Hα1(2)={{f+cGλ:fH1(2),c}if α,H1(2)if α=.H_{\alpha}^{1}(\mathbb{R}^{2})=\left\{\begin{array}[]{@{}c l@{}}\{f+cG_{\lambda}\colon\,f\in H^{1}(\mathbb{R}^{2}),\ c\in\mathbb{C}\}&\text{if }\alpha\in\mathbb{R},\\[3.0pt] H^{1}(\mathbb{R}^{2})&\text{if }\alpha=\infty.\end{array}\right.

Note that Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) is independent of the choice of α\alpha\in\mathbb{R} and λ>0\lambda>0 from (1.7). For λ>eα\lambda>-e_{\alpha}, one can define the maximal extension of the form

(Δα+λ)g,g=fL22+λfL22+|f(0)|2βα(λ)\langle(-\Delta_{\alpha}+\lambda)g,g\rangle=\|\nabla f\|_{L^{2}}^{2}+\lambda\|f\|_{L^{2}}^{2}+\frac{|f(0)|^{2}}{\beta_{\alpha}(\lambda)}

with g=f+f(0)βα(λ)1GλD(Δα)g=f+f(0)\beta_{\alpha}(\lambda)^{-1}G_{\lambda}\in D(-\Delta_{\alpha}). This extension defines a positive quadratic form well-defined on gHα1(2)g\in H_{\alpha}^{1}(\mathbb{R}^{2}) explicitly given by

(1.13) (Δα+λ)g,g=fL22+λfL22+βα(λ)|c|2\langle(-\Delta_{\alpha}+\lambda)g,g\rangle=\|\nabla f\|_{L^{2}}^{2}+\lambda\|f\|_{L^{2}}^{2}+\beta_{\alpha}(\lambda)|c|^{2}

for g=f+cGλHα1(2)g=f+cG_{\lambda}\in H_{\alpha}^{1}(\mathbb{R}^{2}). By using the relation

λgL22=λf+cGλL22=λfL22+2λRe[c(f,Gλ)L2]+|c|24π,\lambda\|g\|_{L^{2}}^{2}=\lambda\|f+cG_{\lambda}\|_{L^{2}}^{2}=\lambda\|f\|_{L^{2}}^{2}+2\lambda\operatorname{Re}[c(f,G_{\lambda})_{L^{2}}]+\frac{|c|^{2}}{4\pi},

we can rewrite (1.13) as

Δαg,g=fL222λRe[c(f,Gλ)L2]+(βα(λ)14π)|c|2.\langle-\Delta_{\alpha}g,g\rangle=\|\nabla f\|_{L^{2}}^{2}-2\lambda\operatorname{Re}[c(f,G_{\lambda})_{L^{2}}]+\left(\beta_{\alpha}(\lambda)-\frac{1}{4\pi}\right)|c|^{2}.

We denote the Hα1H_{\alpha}^{1}-norm depending on the parameter λ>eα\lambda>-e_{\alpha} by

(1.14) gHα,λ1:=(Δα+λ)g,g\|g\|_{H_{\alpha,\lambda}^{1}}\mathrel{\mathop{:}}=\sqrt{\langle(-\Delta_{\alpha}+\lambda)g,g\rangle}

for gHα1(2)g\in H_{\alpha}^{1}(\mathbb{R}^{2}). The following equivalence holds:

gHα,λ1gHα,μ1,λ,μ>eα.\|g\|_{H_{\alpha,\lambda}^{1}}\simeq\|g\|_{H_{\alpha,\mu}^{1}},\quad\lambda,\mu>-e_{\alpha}.

As a special case, we choose λ=1eα\lambda=1-e_{\alpha} and also use the notation

gHα1:=gHα,1eα1.\|g\|_{H_{\alpha}^{1}}\mathrel{\mathop{:}}=\|g\|_{H_{\alpha,1-e_{\alpha}}^{1}}.

To study the stability properties of the standing waves, we need the local well-posedness in the energy spaces Hα1(2)H^{1}_{\alpha}(\mathbb{R}^{2}). We have the following statement (see Appendix B for the proof).

Proposition 1.1.

Let α\alpha\in\mathbb{R} and p>1p>1. For each u0Hα1(2)u_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}), there exists the unique maximal solution

uC((Tmin,Tmax),Hα1(2))C1((Tmin,Tmax),Hα1(2))u\in C\bigl{(}(-T_{\min},T_{\max}),H_{\alpha}^{1}(\mathbb{R}^{2})\bigr{)}\cap C^{1}\bigl{(}(-T_{\min},T_{\max}),H_{\alpha}^{-1}(\mathbb{R}^{2})\bigr{)}

of (1.1) with the initial data u(0)=u0u(0)=u_{0}, where Hα1(2)H_{\alpha}^{-1}(\mathbb{R}^{2}) is the dual space of Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}). Moreover, uu satisfies the conservation laws of energy and the L2L^{2}-norm:

E(u(t))=E(u0),\displaystyle E(u(t))=E(u_{0}), u(t)L2=u0L2\displaystyle\|u(t)\|_{L^{2}}=\|u_{0}\|_{L^{2}}

for all t(Tmin,Tmax)t\in(-T_{\min},T_{\max}), where the energy is defined by

E(v):=12Δαv,v1p+1vLp+1p+1,vHα1(2).E(v)\mathrel{\mathop{:}}=\frac{1}{2}\langle-\Delta_{\alpha}v,v\rangle-\frac{1}{p+1}\|v\|_{L^{p+1}}^{p+1},\quad v\in H_{\alpha}^{1}(\mathbb{R}^{2}).

Now let us consider standing wave solutions with the form

(1.15) u(t,x)=eiωtϕ(x),u(t,x)=e^{i\omega t}\phi(x),

where ϕHα1(2)\phi\in H_{\alpha}^{1}(\mathbb{R}^{2}) is a nontrivial solution of the stationary equation

(1.16) Δαϕ+ωϕ|ϕ|p1ϕ=0,x2.-\Delta_{\alpha}\phi+\omega\phi-|\phi|^{p-1}\phi=0,\quad x\in\mathbb{R}^{2}.

Equation (1.16) can be rewritten as Sω(ϕ)=0S_{\omega}^{\prime}(\phi)=0, where SωS_{\omega} is the action functional defined by

(1.17) Sω(v)=Sα,ω(v):=12vHα,ω121p+1vLp+1p+1,vHα1(2).S_{\omega}(v)=S_{\alpha,\omega}(v)\mathrel{\mathop{:}}=\frac{1}{2}\|v\|_{H_{\alpha,\omega}^{1}}^{2}-\frac{1}{p+1}\|v\|_{L^{p+1}}^{p+1},\quad v\in H_{\alpha}^{1}(\mathbb{R}^{2}).

We denote the set of all nontrivial solution of (1.16) by

𝒜ω=𝒜α,ω\displaystyle\mathcal{A}_{\omega}=\mathcal{A}_{\alpha,\omega} :={ϕHα1(2):ϕ0,Sα,ω(ϕ)=0}\displaystyle\mathrel{\mathop{:}}=\{\phi\in H_{\alpha}^{1}(\mathbb{R}^{2})\colon\,\phi\neq 0,\ S_{\alpha,\omega}^{\prime}(\phi)=0\}

and the set of all ground states (minimal action solution) by

𝒢ω=𝒢α,ω\displaystyle\mathcal{G}_{\omega}=\mathcal{G}_{\alpha,\omega} :={ϕ𝒜α,ω:Sα,ω(ϕ)Sα,ω(ψ) for all ψ𝒜α,ω}.\displaystyle\mathrel{\mathop{:}}=\{\phi\in\mathcal{A}_{\alpha,\omega}\colon\,S_{\alpha,\omega}(\phi)\leq S_{\alpha,\omega}(\psi)\text{ for all }\psi\in\mathcal{A}_{\alpha,\omega}\}.

Now we state our main results. First, we state the results about the existence, symmetry, and uniqueness of ground states.

Theorem 1.2.

If ω>eα\omega>-e_{\alpha}, then the set 𝒢ω\mathcal{G}_{\omega} is not empty.

Theorem 1.3.

Let ω>eα\omega>-e_{\alpha} and ϕ𝒢ω\phi\in\mathcal{G}_{\omega}. If ϕ\phi is decomposed as ϕ=f+f(0)βα(ω)1Gω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega} with fH2(2)f\in H^{2}(\mathbb{R}^{2}), then there exists θ\theta\in\mathbb{R} such that eiθfe^{i\theta}f is positive, radial, and decreasing function. In particular, the function eiθϕe^{i\theta}\phi is also positive, radial, and decreasing.

Theorem 1.4.

There exists ω1>eα\omega_{1}>-e_{\alpha} such that if ω>ω1\omega>\omega_{1}, then there exists the unique positive radial ground state ϕω𝒢ω\phi_{\omega}\in\mathcal{G}_{\omega} such that the set of all ground states is characterized as

(1.18) 𝒢ω={eiθϕω:θ}.\mathcal{G}_{\omega}=\{e^{i\theta}\phi_{\omega}\colon\,\theta\in\mathbb{R}\}.

Next, we state the results about orbital stability and instability of standing waves. The definition of orbital stability is as follows.

Definition 1.5.

Let ω\omega\in\mathbb{R} and let ϕHα1(2)\phi\in H_{\alpha}^{1}(\mathbb{R}^{2}) be a nontrivial solution of (1.16). The standing wave eiωtϕe^{i\omega t}\phi is stable if for any ε>0\varepsilon>0 there exists δ>0\delta>0 such that for any u0Hα1(2)u_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}) satisfying u0ϕHα1<δ\|u_{0}-\phi\|_{H_{\alpha}^{1}}<\delta, the solution u(t)u(t) of (1.1) with u(0)=u0u(0)=u_{0} exists globally in time and satisfies

suptinfθu(t)eiθϕHα1<ε.\sup_{t\in\mathbb{R}}\inf_{\theta\in\mathbb{R}}\|u(t)-e^{i\theta}\phi\|_{H_{\alpha}^{1}}<\varepsilon.

Otherwise, the standing wave eiωtϕe^{i\omega t}\phi is unstable.

The following two statements are main results of this paper. The first one concerns the stability for ω\omega close to eα-e_{\alpha}.

Theorem 1.6.

For each α\alpha\in\mathbb{R} and p>1p>1, there exists ω>eα\omega_{*}>-e_{\alpha} such that if ω(eα,ω)\omega\in(-e_{\alpha},\omega_{*}) and ϕ𝒢ω\phi\in\mathcal{G}_{\omega}, the standing wave eiωtϕe^{i\omega t}\phi is stable.

The second one concerns the stability/instability for large frequency ω\omega. We denote the unique ground state given in Theorem 1.4 by ϕω\phi_{\omega}.

Theorem 1.7.

For each α\alpha\in\mathbb{R} and p>1p>1, there exists ω(ω1,)\omega^{*}\in(\omega_{1},\infty) such that the following is true.

  • If 1<p31<p\leq 3, then the standing wave eiωtϕωe^{i\omega t}\phi_{\omega} is stable for all ω>ω\omega>\omega^{*}.

  • If p>3p>3, then the standing wave eiωtϕe^{i\omega t}\phi is unstable for all ω>ω\omega>\omega^{*}.

One can observe the similarity between the results in [28, 30, 32, 33, 34] and Theorems 1.6 and 1.7. The paper [34] treats NLS with attractive δ\delta-potential in one dimension, and the papers [28, 30, 32, 33] concern NLS with general attractive potential V(x)V(x). Since Δα-\Delta_{\alpha} has a unique simple negative eigenvalue, we regard it as a Schrödinger operator with an attractive potential, so it is natural to choose to follow the approach in these papers.

Let us give a short outline of the proofs. The local well-posedness (Proposition 1.1) follows from the energy methods in [52] (see also [20, Chapter 3]) and the Strichartz estimates for the operator Δα-\Delta_{\alpha} obtained by [22]. The existence of ground states (Theorem 1.2) follows from a standard variational method by using the Nehari manifold. The positivity and symmetry of ground states (Theorem 1.3) follow from the maximal principle and the symmetric rearrangement. In particular, we use the result of Brothers and Ziemer [16] to obtain the radial symmetry and decrease of ground states.

To investigate the stability properties, we consider rescaled ground states and use a perturbation argument as in [30, 32, 33]. Let (ϕω)ω>eα(\phi_{\omega})_{\omega>-e_{\alpha}} be a family of positive ground states with ϕω=ϕα,ω𝒢α,ω\phi_{\omega}=\phi_{\alpha,\omega}\in\mathcal{G}_{\alpha,\omega}. For ω\omega close to eα-e_{\alpha}, we normalize the ground states as

ϕ^ω(x):=ϕωϕωL2.\widehat{\phi}_{\omega}(x)\mathrel{\mathop{:}}=\frac{\phi_{\omega}}{\lVert\phi_{\omega}\rVert_{L^{2}}}.

Then ϕ^ω\widehat{\phi}_{\omega} is a positive solution of

Δαϕ^+ωϕ^ϕωL2p1|ϕ^|p1ϕ^=0.-\Delta_{\alpha}\widehat{\phi}+\omega\widehat{\phi}-\lVert\phi_{\omega}\rVert_{L^{2}}^{p-1}|\widehat{\phi}|^{p-1}\widehat{\phi}=0.

Since we can verify ϕωL20\lVert\phi_{\omega}\rVert_{L^{2}}\to 0 as ωeα\omega\to-e_{\alpha}, ϕ^ω\widehat{\phi}_{\omega} can be regarded as a perturbation of the solution for the linear equation

(1.19) Δαϕ^=eαϕ^,\mathopen{}-\Delta_{\alpha}\widehat{\phi}=e_{\alpha}\widehat{\phi},

that is, ϕ^ω\widehat{\phi}_{\omega} is close to the eigenfunction χα\chi_{\alpha}. On the other hand, for large ω\omega, we rescale ϕω\phi_{\omega} as

ϕ~ω(x):=ω1/(p1)ϕω(x/ω).\widetilde{\phi}_{\omega}(x)\mathrel{\mathop{:}}=\omega^{-1/(p-1)}\phi_{\omega}(x/\sqrt{\omega}).

Then ϕ~ω\widetilde{\phi}_{\omega} is a positive solution of

Δα~ϕ~+ϕ~|ϕ~|p1ϕ~=0,-\Delta_{\widetilde{\alpha}}\widetilde{\phi}+\widetilde{\phi}-|\widetilde{\phi}|^{p-1}\widetilde{\phi}=0,

where

α~=α~(ω):=α+14πlnω.\widetilde{\alpha}=\widetilde{\alpha}(\omega)\mathrel{\mathop{:}}=\alpha+\frac{1}{4\pi}\ln\omega.

Since Δα=-\Delta_{\alpha=\infty} is the free Laplacian Δ-\Delta, ϕ~ω\widetilde{\phi}_{\omega} can be regarded as a perturbation of the solution for the equation

(1.20) Δϕ~+ϕ~|ϕ~|p1ϕ~=0.{-}\Delta\widetilde{\phi}+\widetilde{\phi}-\lvert\widetilde{\phi}\rvert^{p-1}\widetilde{\phi}=0.

Therefore, we can investigate the stability properties by using the limiting equation (1.19) for small ω\omega and (1.20) for large ω\omega.

The stability for small frequency (Theorem 1.6) follows from the argument of [33]. If ω\omega is sufficiently close to eα-e_{\alpha}, we can obtain the following coercivity property for the linearlized operator around the ground state.

Proposition 1.8.

For each α\alpha\in\mathbb{R} and p>1p>1, there exists ω>eα\omega_{*}>-e_{\alpha} such that if ω(eα,ω)\omega\in(-e_{\alpha},\omega_{*}) and ϕ𝒢ω\phi\in\mathcal{G}_{\omega}, the following holds: There exists a positive constant kk such that

Sω′′(ϕ)w,wkwHα12\langle S_{\omega}^{\prime\prime}(\phi)w,w\rangle\geq k\|w\|_{H_{\alpha}^{1}}^{2}

for any wHα1(2)w\in H_{\alpha}^{1}(\mathbb{R}^{2}) satisfying 2ϕw¯𝑑x=0\int_{\mathbb{R}^{2}}\phi\overline{w}\,dx=0.

It is known that this coercivity implies the stability (see, e.g., [37, 44]). Because we can prove Proposition 1.8 exactly in the same way as [33, Section 4], we omit the proof.

Remark 1.9.

In this paper, we do not discuss the uniqueness of ground states for small frequencies because we do not need it if we just prove the stability. However, we can obtain the uniqueness by the bifurcation theory. See, e.g., [53, 38, 47] for more details.

To investigate the properties of the ground states for large frequency ω\omega, we use the limiting equation (1.20). It is well known that (1.20) has the unique positive radial ground state ϕ~H1(2)\widetilde{\phi}_{\infty}\in H^{1}(\mathbb{R}^{2}) (see, e.g., [12, 43]), and it is nondegenerate in the radial space, that is, the kernel of the linearized operator L~:=Δ+1pϕ~p1\widetilde{L}_{\infty}\mathrel{\mathop{:}}=-\Delta+1-p\widetilde{\phi}_{\infty}^{p-1} is trivial: kerL~|Hrad1={0}\ker\widetilde{L}_{\infty}|_{H_{\mathrm{rad}}^{1}}=\{0\}. By using these properties, we establish the uniqueness (Theorem 1.18) and nondegeneracy (Lemma 6.1) for large ω\omega following the argument of [30, Proposition 2 (v)]. Moreover, we can obtain the regularity of the map ωϕω\omega\mapsto\phi_{\omega} (Corollary 8.2). To obtain the stability and instability, we use the following criteria.

Proposition 1.10 ([50, 54]).

For ω>ω1\omega>\omega_{1}, the standing wave eiωtϕωe^{i\omega t}\phi_{\omega} of (1.1) is stable if ddωϕωL22>0\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}>0 and unstable if ddωϕωL22<0\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}<0.

Remark 1.11.

Proposition 1.10 are well-known as the criteria of Grillakis, Shatah, and Strauss [37] (see also [55, 58]). To use their result, we need to investigate the spectral properties of the linearized operator Sω′′(ϕω)S_{\omega}^{\prime\prime}(\phi_{\omega}), but we do not discuss its spectra in this paper. Instead, we can apply the arguments of [54] for the stability and [50] for the instability because they only require the variational characterization on the Nehari manifold, the uniqueness, and the differentiability of the map ωϕω\omega\mapsto\phi_{\omega} with ϕω𝒢ω\phi_{\omega}\in\mathcal{G}_{\omega}. These properties are discussed in this paper.

From Proposition 1.10, the stability/instability problems can be reduced to the investigation of the sign of the derivative ddωϕωL22\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}. When α=\alpha=\infty, i.e., without interaction, one can show by the scaling invariance for the equation that the ground states ϕ,ω\phi_{\infty,\omega} of (1.16) satisfies ddωϕ,ωL22>0\frac{d}{d\omega}\|\phi_{\infty,\omega}\|_{L^{2}}^{2}>0 if 1<p<31<p<3 and ddωϕ,ωL22<0\frac{d}{d\omega}\|\phi_{\infty,\omega}\|_{L^{2}}^{2}<0 if p>3p>3 for all ω>0\omega>0. This means that when α=\alpha=\infty, the ground-state standing wave eiωtϕ,ωe^{i\omega t}\phi_{\infty,\omega} of (1.1) is stable if 1<p<31<p<3 [19] and unstable if p3p\geq 3 [13] (see [57] for p=3p=3).

To investigate the sign of ddωϕωL22\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2} for large ω\omega, we apply the argument of [30, 28]. Instead of ddωϕωL22\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}, we calculate the rescaled version ddωϕ~ωL22\frac{d}{d\omega}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2} and use the convergence ϕ~ωϕ~\widetilde{\phi}_{\omega}\to\widetilde{\phi}_{\infty}. To estimate some error terms, we establish and use a boundedness of the inverse linearized operator of ϕ~ω\widetilde{\phi}_{\omega}. After that, we can determine the sign of ddωϕωL22\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}, and combining Proposition 1.10 we obtain Theorem 1.7.

The difficulty of the proofs of our results mainly comes from the treatment of functions in the energy space Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) and the domain D(Δα)D(-\Delta_{\alpha}). Of special importance in one-dimensional case is the fact that the fundamental solution of (1Δ)(1-\Delta) is in H1()H^{1}(\mathbb{R}), so one can use H1()H^{1}(\mathbb{R}) as a natural space of the nonlinear flow associated with the corresponding NLS. The situation changes essentially in two dimension since we are forced to work with the perturbed Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) space, so there are nontrivial difficulties to apply of the variational technique from [34] and the cases of slowly decaying potentials [28, 30, 32, 33]. For a function in the spaces Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) or D(Δα)D(-\Delta_{\alpha}), we need to decompose it into the regular and singular parts and to treat these separately, and we have to avoid several difficult points requiring appropriate new treatments.

  • The local well-posedness for the standard 2d NLS with or without potential requires the use of Strichartz estimates in Sobolev spaces

    (1.21) Hs,p(2)={g=(1Δ)sψ:ψLp(2)},p(1,2),s(0,1],H^{s,p}(\mathbb{R}^{2})=\{g=(1-\Delta)^{-s}\psi\colon\,\psi\in L^{p}(\mathbb{R}^{2})\},\quad p\in(1,2),\ s\in(0,1],

    if a contraction argument is applied. On the other hand, the case of singular perturbed Laplacian Δα-\Delta_{\alpha} requires the replacement of the classical Sobolev space H1(2)H^{1}(\mathbb{R}^{2}) by the perturbed space Hα1(2)H^{1}_{\alpha}(\mathbb{R}^{2}), and we need to decompose functions gHα1(2)g\in H^{1}_{\alpha}(\mathbb{R}^{2}) as

    (1.22) g=f+cGλ.g=f+cG_{\lambda}.

    There is no flexible treatment (up to our knowledge) of appropriate generalization of generic spaces Hs,pH^{s,p} for the Laplacian of type Δα-\Delta_{\alpha}. For this we have chosen another approach based on compactness argument and the results in [52].

  • The existence of ground states seems to be closely connected with the inclusion

    H1(2)Hα1(2).H^{1}(\mathbb{R}^{2})\subset H^{1}_{\alpha}(\mathbb{R}^{2}).

    However, the ground states ϕ=f+f(0)βα(ω)1Gω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega} from Theorem 1.3 have nontrivial singular part, since eiθf(0)e^{i\theta}f(0) is positive. This fact shows that the ground states associated with α\alpha\in\mathbb{R} are different from ground states with α=\alpha=\infty. Moreover the ground state ϕ\phi from Theorem 1.3 is not in H1(2)H^{1}(\mathbb{R}^{2}).

  • The symmetry of the ground state for the classical NLS can be obtained by Schwartz symmetrization. Since we have the decomposition ϕ=f+f(0)βα(ω)1Gω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega} for any ϕD(Δα)\phi\in D(-\Delta_{\alpha}) into regular and singular parts, a formal symmetrization

    ϕ=(f+f(0)βα(λ)Gλ)\phi^{*}=\left(f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\right)^{*}

    cannot work. We have chosen the following symmetrization

    f+f(0)βα(λ)Gλf+|f(0)|βα(λ)Gλ.f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\to f^{*}+\frac{|f(0)|}{\beta_{\alpha}(\lambda)}G_{\lambda}.

    The technical difficulties associated with this symmetrization can be overcome by using the results in [11].

  • The uniqueness and nondegeneracy of ground states require careful use of the decomposition (1.22) and modify accordingly the approach in [32, 33, 30].

  • To determine the sign of ddωϕ~ωL22\frac{d}{d\omega}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}, we need to estimate the error term ωf~ω(0)\partial_{\omega}\widetilde{f}_{\omega}(0) coming from the interaction of Δα-\Delta_{\alpha}, where f~ω\widetilde{f}_{\omega} is the regular part of ϕ~ω\widetilde{\phi}_{\omega}. As in the previous work [28, 30], we use the boundedness of the inverse of the linearized operator L~ω\widetilde{L}_{\omega}. However, it is not trivial how to express and estimate the term ωf~ω(0)\partial_{\omega}\widetilde{f}_{\omega}(0) by using the operator L~ω1\widetilde{L}_{\omega}^{-1}. To overcome this difficulty we make a good use the expression (1.3) of the operator and the expression of the bilinear form as

    f(0)=(Δα+λ)Gλ,f+f(0)βα(λ)Gλf(0)=\left\langle(\Delta_{\alpha}+\lambda)G_{\lambda},f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\right\rangle

    for fH2(2;)f\in H^{2}(\mathbb{R}^{2};\mathbb{R}). For more details, see Lemma 9.4.

The rest of organization of this paper is as follows. In Section 2 we correct the properties of GλG_{\lambda} used in this paper. In Section 3 we prove Theorem 1.2 through the characterization with the Nehari functional. Section 4 is devoted to the proof of Theorem 1.3. In Section 5 we show that a family of rescaled ground states converges to the ground state of NLS without interaction (i.e. α=\alpha=\infty) as ω\omega\to\infty. In Section 6 we show lower boundedness and nondegeneracy of the linearized operator around the ground state for large ω\omega. This lower boundedness will be used in Section 9 as the boundedness of the inverse operator. In Section 7 we prove Theorem 1.18. In Section 8 we discuss the regularity of the map ωϕω\omega\mapsto\phi_{\omega} for large ω\omega. Finally we prove Theorem 1.7 in Section 9. In Appendix A we review the properties of wave operators and Strichartz estimates for the operator Δα-\Delta_{\alpha}. In Appendix B we prove Proposition 1.1.

2. Preliminaries

The aim of this section is to recall the main properties of the singular-perturbed Laplace operator Δα-\Delta_{\alpha} and the Green function GλG_{\lambda}.

Note that (1.6) and (1.9) imply

(2.1) GλLq(2)for all q[1,).G_{\lambda}\in L^{q}(\mathbb{R}^{2})\quad\text{for all }q\in[1,\infty).

This fact leads easily to the following Sobolev inequality

Lemma 2.1.

For any q[2,)q\in[2,\infty) and λ>eα\lambda>-e_{\alpha} there exists a constant C>0C>0 such that

(2.2) vLqCvHα,λ1\|v\|_{L^{q}}\leq C\|v\|_{H_{\alpha,\lambda}^{1}}

for all vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}).

Proof.

Any vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}) has the representation v=f+cGλv=f+cG_{\lambda} for some λ>eα\lambda>-e_{\alpha}, fH1(2)f\in H^{1}(\mathbb{R}^{2}), and cc\in\mathbb{C}. Then the property (2.1) implies that

vLqfLq+|c|fH1+|c|.\|v\|_{L^{q}}\lesssim\|f\|_{L^{q}}+|c|\lesssim\|f\|_{H^{1}}+|c|.

The relation (1.14) implies that

fH1+|c|vHα,λ1.\|f\|_{H^{1}}+|c|\sim\|v\|_{{H_{\alpha,\lambda}^{1}}}.

Hence, we have (2.2). ∎

Lemma 2.2.

For λ,μ>0\lambda,\mu>0, the inner product of GλG_{\lambda} and GμG_{\mu} is given by

(Gλ,Gμ)L2={lnλlnμ4π(λμ),μλ,14πλ,μ=λ.(G_{\lambda},G_{\mu})_{L^{2}}=\left\{\begin{array}[]{@{} c l @{}}\dfrac{\ln\lambda-\ln\mu}{4\pi(\lambda-\mu)},&\mu\neq\lambda,\\[9.0pt] \dfrac{1}{4\pi\lambda},&\mu=\lambda.\end{array}\right.
Proof.

The assertion follows from direct calculations. ∎

Lemma 2.3.

[xG1]=2(|ξ|2+1)1[G1]\mathcal{F}[x\cdot\nabla G_{1}]=-2(|\xi|^{2}+1)^{-1}\mathcal{F}[G_{1}].

Proof.

By a direct calculation, we have

[G1]\displaystyle\nabla\mathcal{F}[G_{1}] =ξπ(|ξ|2+1)2,\displaystyle=-\frac{\xi}{\pi(|\xi|^{2}+1)^{2}},
[xG1]\displaystyle\mathcal{F}[x\cdot\nabla G_{1}] =2[G1]ξ[G1]=1π(|ξ|2+1)+|ξ|2π(|ξ|2+1)2\displaystyle=-2\mathcal{F}[G_{1}]-\xi\cdot\nabla\mathcal{F}[G_{1}]=-\frac{1}{\pi(|\xi|^{2}+1)}+\frac{|\xi|^{2}}{\pi(|\xi|^{2}+1)^{2}}
=1π(|ξ|2+1)2=2(|ξ|2+1)1[G1].\displaystyle=-\frac{1}{\pi(|\xi|^{2}+1)^{2}}=-2(|\xi|^{2}+1)^{-1}\mathcal{F}[G_{1}].

Thus, we have the assertion. ∎

3. Existence of ground states

In this section, we prove existence of ground states for (1.16) by using a standard variational method and properties of the operator Δα-\Delta_{\alpha}. Throughout this section, we fix ω>eα\omega>-e_{\alpha}. We define the Nehari functional by

Kα,ω(v)\displaystyle K_{\alpha,\omega}(v) :=λSα,ω(λv)|λ=1\displaystyle\mathrel{\mathop{:}}=\partial_{\lambda}S_{\alpha,\omega}(\lambda v)|_{\lambda=1}
=Sα,ω(v),v=vHα,ω12vLp+1p+1\displaystyle=\langle S_{\alpha,\omega}^{\prime}(v),v\rangle=\|v\|_{H_{\alpha,\omega}^{1}}^{2}-\|v\|_{L^{p+1}}^{p+1}

for vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}). We denote

𝒦α,ω\displaystyle\mathcal{K}_{\alpha,\omega} :={vHα1(2):v0,Kα,ω(v)=0},\displaystyle\mathrel{\mathop{:}}=\{v\in H_{\alpha}^{1}(\mathbb{R}^{2})\colon\,v\neq 0,~{}K_{\alpha,\omega}(v)=0\},
dα(ω)\displaystyle d_{\alpha}(\omega) :=inf{Sα,ω(v):v𝒦α,ω},\displaystyle\mathrel{\mathop{:}}=\inf\{S_{\alpha,\omega}(v)\colon\,v\in\mathcal{K}_{\alpha,\omega}\},
α,ω\displaystyle\mathcal{M}_{\alpha,\omega} :={v𝒦α,ω:Sα,ω(v)=dα(ω)}.\displaystyle\mathrel{\mathop{:}}=\{v\in\mathcal{K}_{\alpha,\omega}\colon\,S_{\alpha,\omega}(v)=d_{\alpha}(\omega)\}.

For simplicity of notations, we shall often omit the subscript α\alpha like SωS_{\omega}, KωK_{\omega}, and so on. We note that 𝒢ω𝒜ω𝒦ω\mathcal{G}_{\omega}\subset\mathcal{A}_{\omega}\subset\mathcal{K}_{\omega}.

We will prove the following.

Proposition 3.1.

For any ω>eα\omega>-e_{\alpha},

𝒢ω=ω.\displaystyle\mathcal{G}_{\omega}=\mathcal{M}_{\omega}\neq\emptyset.

By using the functional KωK_{\omega}, we can rewrite the action as

(3.1) Sω(v)\displaystyle S_{\omega}(v) =p12(p+1)vHα,ω12+1p+1Kω(v)\displaystyle=\frac{p-1}{2(p+1)}\|v\|_{H_{\alpha,\omega}^{1}}^{2}+\frac{1}{p+1}K_{\omega}(v)
(3.2) =p12(p+1)vLp+1p+1+12Kω(v),\displaystyle=\frac{p-1}{2(p+1)}\|v\|_{L^{p+1}}^{p+1}+\frac{1}{2}K_{\omega}(v),

and dα(ω)d_{\alpha}(\omega) as

(3.3) dα(ω)\displaystyle d_{\alpha}(\omega) =inf{p12(p+1)vHα,ω12:v𝒦ω}\displaystyle=\inf\left\{\frac{p-1}{2(p+1)}\|v\|_{H_{\alpha,\omega}^{1}}^{2}\colon\,v\in\mathcal{K}_{\omega}\right\}
(3.4) =inf{p12(p+1)vLp+1p+1:v𝒦ω}.\displaystyle=\inf\left\{\frac{p-1}{2(p+1)}\|v\|_{L^{p+1}}^{p+1}\colon\,v\in\mathcal{K}_{\omega}\right\}.
Lemma 3.2.

ω𝒢ω\mathcal{M}_{\omega}\subset\mathcal{G}_{\omega}.

Proof.

Let ϕω\phi\in\mathcal{M}_{\omega}. By Kω(ϕ)=0K_{\omega}(\phi)=0, we have

(3.5) Kω(ϕ),ϕ=2ϕHα,ω12(p+1)ϕLp+1p+1=(p1)ϕLp+1p+1<0.\langle K_{\omega}^{\prime}(\phi),\phi\rangle=2\|\phi\|_{H_{\alpha,\omega}^{1}}^{2}-(p+1)\|\phi\|_{L^{p+1}}^{p+1}=-(p-1)\|\phi\|_{L^{p+1}}^{p+1}<0.

Therefore, there exists a Lagrange multiplier η\eta\in\mathbb{R} such that Sω(ϕ)=ηKω(ϕ)S_{\omega}^{\prime}(\phi)=\eta K_{\omega}^{\prime}(\phi). Moreover, since

ηKω(ϕ),ϕ=Sω(ϕ),ϕ=Kω(ϕ)=0,\eta\langle K_{\omega}^{\prime}(\phi),\phi\rangle=\langle S_{\omega}^{\prime}(\phi),\phi\rangle=K_{\omega}(\phi)=0,

it follows from (3.5) that η=0\eta=0, which implies Sω(ϕ)=0S_{\omega}^{\prime}(\phi)=0.

Furthermore, if ψ𝒜ω\psi\in\mathcal{A}_{\omega}, by ϕω\phi\in\mathcal{M}_{\omega} and ψ𝒦ω\psi\in\mathcal{K}_{\omega}, we have Sω(ϕ)Sω(ψ)S_{\omega}(\phi)\leq S_{\omega}(\psi). Thus, we obtain ϕ𝒢ω\phi\in\mathcal{G}_{\omega}. This completes the proof. ∎

Lemma 3.3.

If vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}) satisfies Kω(v)<0K_{\omega}(v)<0, then

p12(p+1)vLp+1p+1>dα(ω),\displaystyle\frac{p-1}{2(p+1)}\|v\|_{L^{p+1}}^{p+1}>d_{\alpha}(\omega), p12(p+1)vHα,ω12>dα(ω).\displaystyle\frac{p-1}{2(p+1)}\|v\|_{H_{\alpha,\omega}^{1}}^{2}>d_{\alpha}(\omega).
Proof.

Let vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}) satisfy Kω(v)<0K_{\omega}(v)<0. From the shape of the graph of the function λKω(λv)=λ2vHα,ω12λp+1vLp+1p+1\lambda\mapsto K_{\omega}(\lambda v)=\lambda^{2}\|v\|_{H_{\alpha,\omega}^{1}}^{2}-\lambda^{p+1}\|v\|_{L^{p+1}}^{p+1}, there exists λ0(0,1)\lambda_{0}\in(0,1) such that Kω(λ0v)=0K_{\omega}(\lambda_{0}v)=0. From (3.3), we obtain

dα(ω)p12(p+1)λ0vLp+1p+1=p12(p+1)λ0p+1vLp+1p+1<p12(p+1)vLp+1p+1.d_{\alpha}(\omega)\leq\frac{p-1}{2(p+1)}\|\lambda_{0}v\|_{L^{p+1}}^{p+1}=\frac{p-1}{2(p+1)}\lambda_{0}^{p+1}\|v\|_{L^{p+1}}^{p+1}<\frac{p-1}{2(p+1)}\|v\|_{L^{p+1}}^{p+1}.

Similarly, by using (3.4) we have dα(ω)<p12(p+1)vHα,ω12d_{\alpha}(\omega)<\frac{p-1}{2(p+1)}\|v\|_{H_{\alpha,\omega}^{1}}^{2}. ∎

We note that from Lemma 3.3, the expression (3.3) can be written as

(3.6) dα(ω)=inf{p12(p+1)vHα,ω12:vHα1(2),v0,Kω(v)0}.d_{\alpha}(\omega)=\inf\left\{\frac{p-1}{2(p+1)}\|v\|_{H_{\alpha,\omega}^{1}}^{2}\colon\,v\in H_{\alpha}^{1}(\mathbb{R}^{2}),\ v\neq 0,\ K_{\omega}(v)\leq 0\right\}.
Lemma 3.4.

dα(ω)>0d_{\alpha}(\omega)>0.

Proof.

Let v𝒦ωv\in\mathcal{K}_{\omega}. From Kω(v)=0K_{\omega}(v)=0 and the embedding Hα1(2)Lp+1(2)H_{\alpha}^{1}(\mathbb{R}^{2})\hookrightarrow L^{p+1}(\mathbb{R}^{2}) (Lemma 2.1), we have

vHα,ω12=vLp+1p+1vHα,ω1p+1.\displaystyle\|v\|_{H_{\alpha,\omega}^{1}}^{2}=\|v\|_{L^{p+1}}^{p+1}\lesssim\|v\|_{H_{\alpha,\omega}^{1}}^{p+1}.

Since v0v\neq 0 and p>1p>1, we have the uniform bound vHα,ω11\|v\|_{H_{\alpha,\omega}^{1}}\gtrsim 1. From the expression (3.3) we obtain the conclusion. ∎

Now we use the action and Nehari functional without potential defined by

S,ω(f)\displaystyle S_{\infty,\omega}(f) :=12fL22+ω2fL221p+1fLp+1p+1,\displaystyle\mathrel{\mathop{:}}=\frac{1}{2}\|\nabla f\|_{L^{2}}^{2}+\frac{\omega}{2}\|f\|_{L^{2}}^{2}-\frac{1}{p+1}\|f\|_{L^{p+1}}^{p+1},
K,ω(f)\displaystyle K_{\infty,\omega}(f) :=fL22+ωfL22fLp+1p+1,\displaystyle\mathrel{\mathop{:}}=\|\nabla f\|_{L^{2}}^{2}+\omega\|f\|_{L^{2}}^{2}-\|f\|_{L^{p+1}}^{p+1},

respectively. We denote the minimal action value without potential by

d(ω):=min{S,ω(v):vH1(2),v0,K,ω(v)=0}.\displaystyle d_{\infty}(\omega)\mathrel{\mathop{:}}=\min\{S_{\infty,\omega}(v)\colon\,v\in H^{1}(\mathbb{R}^{2}),\ v\neq 0,\ K_{\infty,\omega}(v)=0\}.

It is known that there exists the unique positive radial ground state ϕ,ωH1(2)\phi_{\infty,\omega}\in H^{1}(\mathbb{R}^{2}) of the equation

Δϕ+ωϕ|ϕ|p1ϕ=0,x2,-\Delta\phi+\omega\phi-|\phi|^{p-1}\phi=0,\quad x\in\mathbb{R}^{2},

and that ϕ,ω\phi_{\infty,\omega} satisfies S,ω(ϕ,ω)=d(ω)S_{\infty,\omega}(\phi_{\infty,\omega})=d_{\infty}(\omega).

We note that

Sα,ω(f)=S,ω(f),\displaystyle S_{\alpha,\omega}(f)=S_{\infty,\omega}(f), Kα,ω(f)=K,ω(f)\displaystyle K_{\alpha,\omega}(f)=K_{\infty,\omega}(f)

for all fH1(2)f\in H^{1}(\mathbb{R}^{2}), which implies dα(ω)d(ω)d_{\alpha}(\omega)\leq d_{\infty}(\omega). We prove the strict inequality.

Lemma 3.5.

dα(ω)<d(ω)d_{\alpha}(\omega)<d_{\infty}(\omega).

Proof.

Suppose that dα(ω)=d(ω)d_{\alpha}(\omega)=d_{\infty}(\omega). Then we have

Sα,ω(ϕ,ω)=S,ω(ϕ,ω)=d(ω)=dα(ω),\displaystyle S_{\alpha,\omega}(\phi_{\infty,\omega})=S_{\infty,\omega}(\phi_{\infty,\omega})=d_{\infty}(\omega)=d_{\alpha}(\omega),
Kα,ω(ϕ,ω)=K,ω(ϕ,ω)=0.\displaystyle K_{\alpha,\omega}(\phi_{\infty,\omega})=K_{\infty,\omega}(\phi_{\infty,\omega})=0.

This yields ϕ,ωα,ω\phi_{\infty,\omega}\in\mathcal{M}_{\alpha,\omega}. From Lemma 3.2, ϕ,ω\phi_{\infty,\omega} is also a solution of (1.16). This means that ϕ,ωD(Δα)H2(2)\phi_{\infty,\omega}\in D(-\Delta_{\alpha})\cap H^{2}(\mathbb{R}^{2}). From the definition of the domain D(Δα)D(-\Delta_{\alpha}) and the singularity of GωG_{\omega}, we see that ϕ,ω(0)=0\phi_{\infty,\omega}(0)=0, which contradicts the positivity of ϕ,ω\phi_{\infty,\omega}. ∎

Lemma 3.6.

Let (vn)n(v_{n})_{n\in\mathbb{N}} be a sequence in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) satisfy

Kω(vn)0,\displaystyle K_{\omega}(v_{n})\to 0, Sω(vn)dα(ω).\displaystyle S_{\omega}(v_{n})\to d_{\alpha}(\omega).

Then there exist v0Hα1(2){0}v_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2})\setminus\{0\} and a subsequence (vnj)j(v_{n_{j}})_{j\in\mathbb{N}} of (vn)n(v_{n})_{n\in\mathbb{N}} such that vnjv0v_{n_{j}}\to v_{0} in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) as jj\to\infty. In particular, v0ωv_{0}\in\mathcal{M}_{\omega}.

Remark 3.7.

If we just prove the existence of the minimizer v0ωv_{0}\in\mathcal{M}_{\omega}, it suffices to only consider a minimizing sequence (vn)n(v_{n})_{n\in\mathbb{N}} for dα(ω)d_{\alpha}(\omega), that is, Kω(vn)=0K_{\omega}(v_{n})=0 and Sω(vn)dα(ω)S_{\omega}(v_{n})\to d_{\alpha}(\omega). However, we show the stronger statement in Lemma 3.6 because it is used when we apply the argument of Shatah [54] for stability in Proposition 1.10.

Proof of Lemma 3.6.

We decompose vn=fn+cnGωv_{n}=f_{n}+c_{n}G_{\omega}. From the assumptions and the expressions (3.1) and (3.2), we have

(3.7) p12(p+1)vnHα,ω12\displaystyle\frac{p-1}{2(p+1)}\|v_{n}\|_{H_{\alpha,\omega}^{1}}^{2} dα(ω),\displaystyle\to d_{\alpha}(\omega), p12(p+1)vnLp+1p+1\displaystyle\frac{p-1}{2(p+1)}\|v_{n}\|_{L^{p+1}}^{p+1} dα(ω).\displaystyle\to d_{\alpha}(\omega).

This implies that (vn)n(v_{n})_{n} is bounded in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}), and so there exists v0Hα1(2)v_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}) and a subsequence of (vn)(v_{n}) such that vnv0v_{n}\rightharpoonup v_{0} weakly in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}). From the definition of Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}), we see that fnf0f_{n}\rightharpoonup f_{0} weakly in H1(2)H^{1}(\mathbb{R}^{2}) and cnc0c_{n}\to c_{0} for some (f0,c0)H1(2)×(f_{0},c_{0})\in H^{1}(\mathbb{R}^{2})\times\mathbb{C}.

Now we show that c00c_{0}\neq 0. Suppose that c0=0c_{0}=0. Then by (3.7) we have

(3.8) p12(p+1)(fnL22+ωfnL22)\displaystyle\frac{p-1}{2(p+1)}(\|\nabla f_{n}\|_{L^{2}}^{2}+\omega\|f_{n}\|_{L^{2}}^{2}) dα(ω),\displaystyle\to d_{\alpha}(\omega), p12(p+1)fnLp+1p+1\displaystyle\frac{p-1}{2(p+1)}\|f_{n}\|_{L^{p+1}}^{p+1} dα(ω).\displaystyle\to d_{\alpha}(\omega).

Let

λn:=(fnL22+ωfnL22fnLp+1p+1)1/(p1).\displaystyle\lambda_{n}\mathrel{\mathop{:}}=\left(\frac{\|\nabla f_{n}\|_{L^{2}}^{2}+\omega\|f_{n}\|_{L^{2}}^{2}}{\|f_{n}\|_{L^{p+1}}^{p+1}}\right)^{1/(p-1)}.

We have K,ω(λnfn)=0K_{\infty,\omega}(\lambda_{n}f_{n})=0. Moreover, (3.8) and Lemma 3.4 imply λn1\lambda_{n}\to 1. From the definition of d(ω)d_{\infty}(\omega) and Lemma 3.5, we obtain

d(ω)p12(p+1)λnfnLp+1p+1\displaystyle d_{\infty}(\omega)\leq\frac{p-1}{2(p+1)}\|\lambda_{n}f_{n}\|_{L^{p+1}}^{p+1} dα(ω)<d(ω).\displaystyle\to d_{\alpha}(\omega)<d_{\infty}(\omega).

This is a contradiction, which implies c00c_{0}\neq 0.

We show the strong convergence. By the Brezis–Lieb Lemma [15], we have

(3.9) vnHα,ω12vnv0Hα,ω12\displaystyle\|v_{n}\|_{H_{\alpha,\omega}^{1}}^{2}-\|v_{n}-v_{0}\|_{H_{\alpha,\omega}^{1}}^{2} v0Hα,ω12,\displaystyle\to\|v_{0}\|_{H_{\alpha,\omega}^{1}}^{2},
(3.10) Kω(vn)Kω(vnv0)\displaystyle K_{\omega}(v_{n})-K_{\omega}(v_{n}-v_{0}) Kω(v0).\displaystyle\to K_{\omega}(v_{0}).

Since v0Hα,ω1>0\|v_{0}\|_{H_{\alpha,\omega}^{1}}>0, by (3.9), we have

limnp12(p+1)vnv0Hα,ω12<limnp12(p+1)vnHα,ω12=dα(ω).\displaystyle\lim_{n\to\infty}\frac{p-1}{2(p+1)}\|v_{n}-v_{0}\|_{H_{\alpha,\omega}^{1}}^{2}<\lim_{n\to\infty}\frac{p-1}{2(p+1)}\|v_{n}\|_{H_{\alpha,\omega}^{1}}^{2}=d_{\alpha}(\omega).

From this inequality and Lemma 3.3, we have Kω(vnv0)>0K_{\omega}(v_{n}-v_{0})>0 for large nn. Therefore, from Kω(vn)0K_{\omega}(v_{n})\to 0 and (3.10), we obtain Kω(v0)0K_{\omega}(v_{0})\leq 0. Thus, from (3.6), the lower semicontinuity of norms, we deduce that

dα(ω)p12(p+1)v0Hα,ω12p12(p+1)lim infnvnHα,ω12=dα(ω).\displaystyle d_{\alpha}(\omega)\leq\frac{p-1}{2(p+1)}\|v_{0}\|_{H_{\alpha,\omega}^{1}}^{2}\leq\frac{p-1}{2(p+1)}\liminf_{n\to\infty}\|v_{n}\|_{H_{\alpha,\omega}^{1}}^{2}=d_{\alpha}(\omega).

From (3.9), we have vnv0Hα,ω120\|v_{n}-v_{0}\|_{H_{\alpha,\omega}^{1}}^{2}\to 0. Therefore, vnv0v_{n}\to v_{0} in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}). This completes the proof. ∎

Lemma 3.8.

𝒢ωω\mathcal{G}_{\omega}\subset\mathcal{M}_{\omega}.

Proof.

Let ϕ𝒢ω\phi\in\mathcal{G}_{\omega}. Then ϕ𝒦ω\phi\in\mathcal{K}_{\omega}. Since ω\mathcal{M}_{\omega}\neq\emptyset by Lemma 3.6, we can take ψω\psi\in\mathcal{M}_{\omega}. Moreover, by Lemma 3.2 we have ψ𝒢ω\psi\in\mathcal{G}_{\omega}. Therefore, for each v𝒦ωv\in\mathcal{K}_{\omega} we obtain

Sω(ϕ)=Sω(ψ)Sω(v).S_{\omega}(\phi)=S_{\omega}(\psi)\leq S_{\omega}(v).

This implies ϕω\phi\in\mathcal{M}_{\omega}. This completes the proof. ∎

Proof of Proposition 3.1.

The assertion follows from Lemmas 3.2, 3.6, and 3.8. ∎

4. Symmetry of ground states

In this section, we prove Theorem 1.3 based on the argument in [20, Proof of Theorem 8.1.4] but need suitable modifications. We note that if ϕ𝒜ω\phi\in\mathcal{A}_{\omega}, then we have

(Δα+ω)ϕL2ϕL2ppϕHα,ω1p,\|(-\Delta_{\alpha}+\omega)\phi\|_{L^{2}}\leq\|\phi\|_{L^{2p}}^{p}\lesssim\|\phi\|_{H_{\alpha,\omega}^{1}}^{p},

which implies ϕD(Δα)\phi\in D(-\Delta_{\alpha}). In particular, we can decompose ϕ=f+f(0)βα(ω)1Gω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}, and by (1.3) and (1.16), we have the relation

(4.1) (Δ+ω)f|ϕ|p1ϕ=0.(-\Delta+\omega)f-|\phi|^{p-1}\phi=0.

Moreover, by the same argument in the proof of Lemma 3.6, we see that if ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega}, then f(0)0f(0)\neq 0.

Lemma 4.1.

If ψHα1(2)\psi\in H_{\alpha}^{1}(\mathbb{R}^{2}) satisfies

(4.2) p12(p+1)ψHα,ω12dα(ω)p12(p+1)ψLp+1p+1,\frac{p-1}{2(p+1)}\|\psi\|_{H_{\alpha,\omega}^{1}}^{2}\leq d_{\alpha}(\omega)\leq\frac{p-1}{2(p+1)}\|\psi\|_{L^{p+1}}^{p+1},

then ψ𝒢ω\psi\in\mathcal{G}_{\omega}.

Proof.

From the assumption (4.2), we have Kω(ψ)0K_{\omega}(\psi)\leq 0 and Sω(ψ)dα(ω)S_{\omega}(\psi)\leq d_{\alpha}(\omega). On the other hand, by the first inequality in (4.2) and Lemma 3.3, we have Kω(ψ)0K_{\omega}(\psi)\geq 0. Thus, Kω(ψ)=0K_{\omega}(\psi)=0. Moreover, by the definition of dα(ω)d_{\alpha}(\omega), we obtain dα(ω)Sω(ψ)d_{\alpha}(\omega)\leq S_{\omega}(\psi). Therefore, ψω=𝒢ω\psi\in\mathcal{M}_{\omega}=\mathcal{G}_{\omega}. ∎

Throughout this section, we denote the Schwartz symmetrization of ff by ff^{*}.

Lemma 4.2.

Let ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega}. Then

ϕLp+1p+1f+|f(0)|βα(ω)GωLp+1p+1.\|\phi\|_{L^{p+1}}^{p+1}\leq\left\|f^{*}+\frac{|f(0)|}{\beta_{\alpha}(\omega)}G_{\omega}\right\|_{L^{p+1}}^{p+1}.
Proof.

We note that

ϕLp+1p+1=f+f(0)βα(ω)GωLp+1p+1|f|+|f(0)|βα(ω)GωLp+1p+1.\displaystyle\|\phi\|_{L^{p+1}}^{p+1}=\left\|f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}\right\|_{L^{p+1}}^{p+1}\leq\left\||f|+\frac{|f(0)|}{\beta_{\alpha}(\omega)}G_{\omega}\right\|_{L^{p+1}}^{p+1}.

After that, we only have to show that

(4.3) |f|+|f(0)|βα(ω)GωLp+1p+1f+|f(0)|βα(ω)GωLp+1p+1.\left\||f|+\frac{|f(0)|}{\beta_{\alpha}(\omega)}G_{\omega}\right\|_{L^{p+1}}^{p+1}\leq\left\|f^{*}+\frac{|f(0)|}{\beta_{\alpha}(\omega)}G_{\omega}\right\|_{L^{p+1}}^{p+1}.

To prove (4.3) we use [11, Theorem 2.2]. We denote

F(g,h):=(g+h)p+1F(g,h)\mathrel{\mathop{:}}=(g+h)^{p+1}

for g,h+g,h\in\mathbb{R}_{+}. Then we have

Fgh(g,h)=p(p+1)(g+h)p10F_{gh}(g,h)=p(p+1)(g+h)^{p-1}\geq 0

for all g,h0g,h\in\mathbb{R}_{\geq 0}. Therefore, from [11, Theorem 2.2], we obtain

(4.4) 2F(g,h)𝑑x2F(g,h)𝑑x\int_{\mathbb{R}^{2}}F(g,h)\,dx\leq\int_{\mathbb{R}^{2}}F(g^{*},h^{*})\,dx

for all g,h𝒮0g,h\in\mathcal{S}_{0}. Here 𝒮0\mathcal{S}_{0} is the set of all measurable functions w:2w\colon\mathbb{R}^{2}\to\mathbb{R} which satisfy

(4.5) w0,2({x2:w(x)>a})<for all a>0,w\geq 0,\quad\mathcal{L}^{2}(\{x\in\mathbb{R}^{2}\colon\,w(x)>a\})<\infty\quad\text{for all $a>0$},

where 2(A)\mathcal{L}^{2}(A) is the Lebesgue measure of the set AA.

Moreover, we see that |f||f| and GωG_{\omega} satisfy (4.5) because fH2(2)f\in H^{2}(\mathbb{R}^{2}) and GωG_{\omega} is decreasing. Thus, we can use the inequality (4.4) for g=|f|g=|f| and h=|f(0)|βα(ω)1Gωh=|f(0)|\beta_{\alpha}(\omega)^{-1}G_{\omega}. Since Gω=GωG_{\omega}^{*}=G_{\omega}, we obtain (4.3). This completes the proof. ∎

Lemma 4.3.

If ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega}, then f+|f(0)|βα(ω)1Gω𝒢ωf^{*}+|f(0)|\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega}.

Proof.

Let ψ:=f+|f(0)|βα(ω)1Gω\psi\mathrel{\mathop{:}}=f^{*}+|f(0)|\beta_{\alpha}(\omega)^{-1}G_{\omega}. We have

(4.6) p12(p+1)ψHα,ω12\displaystyle\frac{p-1}{2(p+1)}\|\psi\|_{H_{\alpha,\omega}^{1}}^{2} =p12(p+1)(fL22+ωfL22+|f(0)|2βα(ω))\displaystyle=\frac{p-1}{2(p+1)}{\left(\|\nabla f^{*}\|_{L^{2}}^{2}+\omega\|f^{*}\|_{L^{2}}^{2}+\frac{|f(0)|^{2}}{\beta_{\alpha}(\omega)}\right)}
p12(p+1)(fL22+ωfL22+|f(0)|2βα(ω))\displaystyle\leq\frac{p-1}{2(p+1)}{\left(\|\nabla f\|_{L^{2}}^{2}+\omega\|f\|_{L^{2}}^{2}+\frac{|f(0)|^{2}}{\beta_{\alpha}(\omega)}\right)}
=p12(p+1)ϕHα,ω12=dα(ω).\displaystyle=\frac{p-1}{2(p+1)}\|\phi\|_{H_{\alpha,\omega}^{1}}^{2}=d_{\alpha}(\omega).

Moreover, from Lemma 4.2, we have

p12(p+1)ψLp+1p+1p12(p+1)ϕLp+1p+1=dα(ω).\frac{p-1}{2(p+1)}\|\psi\|_{L^{p+1}}^{p+1}\geq\frac{p-1}{2(p+1)}\|\phi\|_{L^{p+1}}^{p+1}=d_{\alpha}(\omega).

Therefore, Lemma 4.1 implies ψ𝒢ω\psi\in\mathcal{G}_{\omega}. ∎

Lemma 4.4.

If ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega} and f(0)>0f(0)>0, then ff is a positive function.

Proof.

Let

g:=|Ref|,\displaystyle g\mathrel{\mathop{:}}=\lvert\operatorname{Re}f\rvert, h:=|Imf|,\displaystyle h\mathrel{\mathop{:}}=\lvert\operatorname{Im}f\rvert, ψ:=g+ih+f(0)βα(ω)Gω.\displaystyle\psi\mathrel{\mathop{:}}=g+ih+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}.

Then we see that

ψHα,ω1=ϕHα,ω1,\displaystyle\|\psi\|_{H_{\alpha,\omega}^{1}}=\|\phi\|_{H_{\alpha,\omega}^{1}}, ψLp+1ϕLp+1.\displaystyle\|\psi\|_{L^{p+1}}\geq\|\phi\|_{L^{p+1}}.

Therefore, Lemma 4.1 implies ψ𝒢ω\psi\in\mathcal{G}_{\omega}. Since

ψ=g+ih+(g+ih)(0)βα(ω)GωD(Δα),\psi=g+ih+\frac{(g+ih)(0)}{\beta_{\alpha}(\omega)}G_{\omega}\in D(-\Delta_{\alpha}),

we can use (1.3) and obtain

(Δ+ω)(g+ih)=(Δα+ω)ψ=|ψ|p1ψ=|ψ|p1(g+ih+f(0)βα(ω)Gω).(-\Delta+\omega)(g+ih)=(-\Delta_{\alpha}+\omega)\psi=|\psi|^{p-1}\psi=|\psi|^{p-1}\biggl{(}g+ih+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}\biggr{)}.

We decompose it into the real part and the imaginary part as

(4.7) (Δ+ω)g\displaystyle(-\Delta+\omega)g =|ψ|p1(g+f(0)βα(ω)Gω)\displaystyle=|\psi|^{p-1}\biggl{(}g+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}\biggr{)}
(4.8) (Δ+ω)h\displaystyle(-\Delta+\omega)h =|ψ|p1h.\displaystyle=|\psi|^{p-1}h.

Since each right-hand side of (4.7) and (4.8) is in L2(2)L^{2}(\mathbb{R}^{2}), we see that g,hH2(2)H1(2)C(2)g,h\in H^{2}(\mathbb{R}^{2})\subset H^{1}(\mathbb{R}^{2})\cap C(\mathbb{R}^{2}). Therefore, since g(0)=f(0)>0g(0)=f(0)>0 and g,Gω0g,G_{\omega}\geq 0 in 2\mathbb{R}^{2}, applying the strong maximal principle (e.g., [21, Theorem 3.1.2]) to the solution gg of (4.7), we have g>0g>0 in 2\mathbb{R}^{2}. Similarly, by using h(0)=0h(0)=0 and (4.8), we obtain h0h\equiv 0 in 2\mathbb{R}^{2}.

From the continuity of ff and the positivity of gg, we see that the sign of f(x)f(x) does not depend on xx. Therefore, by f(0)>0f(0)>0 we have f>0f>0 in 2\mathbb{R}^{2}. This completes the proof. ∎

Lemma 4.5.

If ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega} and ff is positive, then ff is a radial and strictly decreasing function.

Proof.

We denote

ψ:=f+f(0)βα(ω)Gω.\psi\mathrel{\mathop{:}}=f^{*}+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}.

From Lemma 4.3 we have ψ𝒢ω\psi\in\mathcal{G}_{\omega}. In particular, ψD(Δα)\psi\in D(-\Delta_{\alpha}) and so f(0)=f(0)f^{*}(0)=f(0). Moreover, since ψHα,ω12=ϕHα,ω12\|\psi\|_{H_{\alpha,\omega}^{1}}^{2}=\|\phi\|_{H_{\alpha,\omega}^{1}}^{2} by ϕ,ψω\phi,\psi\in\mathcal{M}_{\omega}, we see that

(4.9) fL2=fL2.\|\nabla f^{*}\|_{L^{2}}=\|\nabla f\|_{L^{2}}.

Now we show that ff^{*} is strictly decreasing. Since ψ=f+f(0)βα(ω)1Gω\psi=f^{*}+f^{*}(0)\beta_{\alpha}(\omega)^{-1}G_{\omega} is a positive radial solution of (4.1), f(r)f^{*}(r) satisfies

(4.10) f′′1rf+ωf=(f+f(0)βα(ω)Gω)p,r>0.-f^{\prime\prime}-\frac{1}{r}f^{\prime}+\omega f=\left(f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{\omega}\right)^{p},\quad r>0.

Suppose that ff^{*} is not strictly decreasing. Then ff^{*} is constant in some interval I=(r1,r2)I=(r_{1},r_{2}). From the equation (4.10), ff^{*} satisfies

ωf=(f+f(0)βα(ω)Gω)pin (r1,r2).\omega f^{*}=\left(f^{*}+\frac{f^{*}(0)}{\beta_{\alpha}(\omega)}G_{\omega}\right)^{p}\quad\text{in $(r_{1},r_{2})$}.

The left hand side is a constant whereas the right hand side is not a constant on (r1,r2)(r_{1},r_{2}) since GωG_{\omega} is a strictly decreasing function and f(0)=f(0)>0f^{*}(0)=f(0)>0. This is a contradiction. Thus, ff^{*} is strict decreasing.

Since f(r)f^{*}(r) is strictly decreasing, we see that the Lebesgue measure of the set

(4.11) {x2:f(x)=0, 0<f(|x|)<fL}\{x\in\mathbb{R}^{2}\colon\,\nabla f^{*}(x)=0,\ 0<f^{*}(|x|)<\|f^{*}\|_{L^{\infty}}\}

is zero. Therefore, combining (4.9), we can use [16, Theorem 1.1] to see that there exists y2y\in\mathbb{R}^{2} such that f(y)=ff(\cdot-y)=f^{*}. Moreover, from f(0)=f(0)f(0)=f^{*}(0) we obtain f=ff=f^{*}. This completes the proof. ∎

Proof of Theorem 1.3.

Let ϕ=f+f(0)βα(ω)1Gω𝒢ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega}. Since f(0)0f(0)\neq 0, there exists θ\theta\in\mathbb{R} such that eiθf(0)>0e^{i\theta}f(0)>0. By Lemmas 4.4 and 4.5 we see that eiθfe^{i\theta}f is a positive, radial, and decreasing function. This completes the proof. ∎

5. Rescaled limit

In this section, we prove that a family of rescaled positive ground states converges to the positive radial ground state of (1.20). The argument is based on [32].

Let ϕ=f+f(0)βα(ω)1Gω𝒜ω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{A}_{\omega} and define the rescaling

ϕ~(x):=ω1/(p1)ϕ(x/ω),\displaystyle\widetilde{\phi}(x)\mathrel{\mathop{:}}=\omega^{-1/(p-1)}\phi(x/\sqrt{\omega}),
α~=α~(ω):=α+14πlnω.\displaystyle\widetilde{\alpha}=\widetilde{\alpha}(\omega)\mathrel{\mathop{:}}=\alpha+\frac{1}{4\pi}\ln\omega.

Since

βα(ω)\displaystyle\beta_{\alpha}(\omega) =α+γ2π+12πlnω2\displaystyle=\alpha+\frac{\gamma}{2\pi}+\frac{1}{2\pi}\ln\frac{\sqrt{\omega}}{2}
=α+14πlnω+γ2π+12πln12=βα~(1),\displaystyle=\alpha+\frac{1}{4\pi}\ln\omega+\frac{\gamma}{2\pi}+\frac{1}{2\pi}\ln\frac{1}{2}=\beta_{\widetilde{\alpha}}(1),
Gω(λx)\displaystyle G_{\omega}(\lambda x) =Gλ2ω(x)for λ>0,\displaystyle=G_{\lambda^{2}\omega}(x)\quad\text{for $\lambda>0$},

we have the decomposition

(5.1) ϕ~=f~+f~(0)βα~(1)G1D(Δα~).\widetilde{\phi}=\widetilde{f}+\frac{\widetilde{f}(0)}{\beta_{\widetilde{\alpha}}(1)}G_{1}\in D(-\Delta_{\widetilde{\alpha}}).

From this expression, (1.3), and f(x)=ω1/(p1)f~(ωx)f(x)=\omega^{1/(p-1)}\widetilde{f}(\sqrt{\omega}\,x), we have the relations

(Δα+ω)ϕ\displaystyle(-\Delta_{\alpha}+\omega)\phi =(Δ+ω)f\displaystyle=(-\Delta+\omega)f
=ω1/(p1)ω(Δ+1)f~(ω)\displaystyle=\omega^{1/(p-1)}\omega(-\Delta+1)\widetilde{f}(\sqrt{\omega}\mathrel{\cdot})
=ωp/(p1)(Δα~+1)ϕ~(ω),\displaystyle=\omega^{p/(p-1)}(-\Delta_{\widetilde{\alpha}}+1)\widetilde{\phi}(\sqrt{\omega}\mathrel{\cdot}),
ωϕ\displaystyle\omega\phi =ωp/(p1)ϕ~(ω),\displaystyle=\omega^{p/(p-1)}\widetilde{\phi}(\sqrt{\omega}\mathrel{\cdot}),
|ϕ|p1ϕ\displaystyle|\phi|^{p-1}\phi =ωp/(p1)|ϕ~(ω)|p1ϕ~(ω).\displaystyle=\omega^{p/(p-1)}|\widetilde{\phi}(\sqrt{\omega}\mathrel{\cdot})|^{p-1}\widetilde{\phi}(\sqrt{\omega}\mathrel{\cdot}).

Therefore, ϕ~\widetilde{\phi} is a solution of

(5.2) Δα~ϕ~+ϕ~|ϕ~|p1ϕ~=0.\mathopen{}-\Delta_{\widetilde{\alpha}}\widetilde{\phi}+\widetilde{\phi}-|\widetilde{\phi}|^{p-1}\widetilde{\phi}=0.

Noting that βα~(1)=βα(ω)\beta_{\widetilde{\alpha}}(1)=\beta_{\alpha}(\omega), we denote

vH~α,ω1:=vHα~(ω),11=fH12+βα(ω)|c|2for v=f+cG1Hα1(2).\|v\|_{\widetilde{H}_{\alpha,\omega}^{1}}\mathrel{\mathop{:}}=\|v\|_{H_{\widetilde{\alpha}(\omega),1}^{1}}=\sqrt{\|f\|_{H^{1}}^{2}+\beta_{\alpha}(\omega)|c|^{2}}\quad\text{for }v=f+cG_{1}\in H_{\alpha}^{1}(\mathbb{R}^{2}).

The action and the Nehari functional corresponding to (5.2) are given by

S~α,ω(v)=12vH~α,ω121p+1vLp+1p+1,\displaystyle\widetilde{S}_{\alpha,\omega}(v)=\frac{1}{2}\|v\|_{\widetilde{H}_{\alpha,\omega}^{1}}^{2}-\frac{1}{p+1}\|v\|_{L^{p+1}}^{p+1}, K~α,ω(v)=vH~α,ω12vLp+1p+1\displaystyle\widetilde{K}_{\alpha,\omega}(v)=\|v\|_{\widetilde{H}_{\alpha,\omega}^{1}}^{2}-\|v\|_{L^{p+1}}^{p+1}

for vHα1(2)v\in H_{\alpha}^{1}(\mathbb{R}^{2}), respectively. The action and the Nehari functional corresponding to the limit equation (5.2) are given by

S~(v)=12vH121p+1vLp+1p+1,\displaystyle\widetilde{S}_{\infty}(v)=\frac{1}{2}\|v\|_{H^{1}}^{2}-\frac{1}{p+1}\|v\|_{L^{p+1}}^{p+1}, K~(v)=vH12vLp+1p+1\displaystyle\widetilde{K}_{\infty}(v)=\|v\|_{H^{1}}^{2}-\|v\|_{L^{p+1}}^{p+1}

for vH1(2)v\in H^{1}(\mathbb{R}^{2}), respectively. Note that since

eα~=eαω1,e_{\widetilde{\alpha}}=e_{\alpha}\omega^{-1},

we have

ω>eα1>eα~(ω).\omega>-e_{\alpha}\iff 1>-e_{\widetilde{\alpha}(\omega)}.

In what follow, we only consider the ground state with the positive regular part:

(5.3) ϕω=fω+fω(0)βα(ω)Gω𝒢ω,fω>0,fωHrad2(2).\phi_{\omega}=f_{\omega}+\frac{f_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{\omega}\in\mathcal{G}_{\omega},\quad f_{\omega}>0,\quad f_{\omega}\in H_{\mathrm{rad}}^{2}(\mathbb{R}^{2}).

Note that from Theorem 1.3, for any ground state ψω𝒢ω\psi_{\omega}\in\mathcal{G}_{\omega} there exists θ\theta\in\mathbb{R} such that ϕω:=eiθψω\phi_{\omega}\mathrel{\mathop{:}}=e^{i\theta}\psi_{\omega} satisfies (5.3).

Remark 5.1.

If ϕ\phi is decompose as ϕ=f+f(0)βα(ω)1Gω\phi=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}, then it is natural to decompose ϕ~\widetilde{\phi} as (5.1). However, since βα~(1)=βα(ω)\beta_{\widetilde{\alpha}}(1)=\beta_{\alpha}(\omega), for simplicity of notations, we always decompose it as

ϕ~=f~+f~(0)βα(ω)G1.\widetilde{\phi}=\widetilde{f}+\frac{\widetilde{f}(0)}{\beta_{\alpha}(\omega)}G_{1}.

We denote

d~α(ω)\displaystyle\widetilde{d}_{\alpha}(\omega) :=S~α,ω(ϕ~ω),\displaystyle\mathrel{\mathop{:}}=\widetilde{S}_{\alpha,\omega}(\widetilde{\phi}_{\omega}),
d~()\displaystyle\widetilde{d}(\infty) :=S~(ϕ~)=p12(p+1)ϕ~Lp+1p+1,\displaystyle\mathrel{\mathop{:}}=\widetilde{S}_{\infty}(\widetilde{\phi}_{\infty})=\frac{p-1}{2(p+1)}\|\widetilde{\phi}_{\infty}\|_{L^{p+1}}^{p+1},

where ϕ~\widetilde{\phi}_{\infty} is the unique positive radial solution of (1.20).

Proposition 5.2.

Let (ϕω)ω>eα(\phi_{\omega})_{\omega>-e_{\alpha}} be a family of ground states for (1.16) satisfying (5.3). Then

(5.4) f~ωϕ~in H2(2) as ω.\widetilde{f}_{\omega}\to\widetilde{\phi}_{\infty}\quad\text{in $H^{2}(\mathbb{R}^{2})$ as $\omega\to\infty$}.

In particular, ϕ~ωϕ~\widetilde{\phi}_{\omega}\to\widetilde{\phi}_{\infty} in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) as ω\omega\to\infty.

Proof.

We divide the proof into several steps.

  1. Step 1.

    ω<ω<d~α(ω)<d~α(ω)d~()\omega<\omega^{\prime}<\infty\implies\widetilde{d}_{\alpha}(\omega)<\widetilde{d}_{\alpha}(\omega^{\prime})\leq\widetilde{d}(\infty).

    Since

    K~ω(ϕ~ω)<K~ω(ϕ~ω)=0,\widetilde{K}_{\omega}(\widetilde{\phi}_{\omega^{\prime}})<\widetilde{K}_{\omega^{\prime}}(\widetilde{\phi}_{\omega^{\prime}})=0,

    by Lemma 3.3 we have

    d~α(ω)<p12(p+1)ϕ~ωLp+1p+1=d~(ω).\widetilde{d}_{\alpha}(\omega)<\frac{p-1}{2(p+1)}\|\widetilde{\phi}_{\omega^{\prime}}\|_{L^{p+1}}^{p+1}=\widetilde{d}(\omega^{\prime}).

    Similarly, since

    K~ω(ϕ~)=K~(ϕ~)=0,\widetilde{K}_{\omega}(\widetilde{\phi}_{\infty})=\widetilde{K}_{\infty}(\widetilde{\phi}_{\infty})=0,

    we obtain

    d~α(ω)p12(p+1)ϕ~Lp+1p+1=d~().\widetilde{d}_{\alpha}(\omega)\leq\frac{p-1}{2(p+1)}\|\widetilde{\phi}_{\infty}\|_{L^{p+1}}^{p+1}=\widetilde{d}(\infty).
  2. Step 2.

    supω>eαϕ~ωH~α,ω1<\sup_{\omega>-e_{\alpha}}\|\widetilde{\phi}_{\omega}\|_{\widetilde{H}_{\alpha,\omega}^{1}}<\infty and infω>1eαϕ~ωLp+1>0\inf_{\omega>1-e_{\alpha}}\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}>0.

    This follows from the expression

    d~α(ω)=p12(p+1)ϕ~ωH~ω12=p12(p+1)ϕ~ωLp+1p+1\widetilde{d}_{\alpha}(\omega)=\frac{p-1}{2(p+1)}\|\widetilde{\phi}_{\omega}\|_{\widetilde{H}_{\omega}^{1}}^{2}=\frac{p-1}{2(p+1)}\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}^{p+1}

    and Step 1.

  3. Step 3.

    Weak convergence to a positive radial function.

    If we decompose ϕ~ω=f~ω+f~ω(0)βα(ω)1G1\widetilde{\phi}_{\omega}=\widetilde{f}_{\omega}+\widetilde{f}_{\omega}(0)\beta_{\alpha}(\omega)^{-1}G_{1}, then by Step 2, there exist nonnegative radial function fH1(2)f_{\infty}\in H^{1}(\mathbb{R}^{2}), a constant cc_{\infty}\in\mathbb{R}, and a subsequence (ωj)j(\omega_{j})_{j\in\mathbb{N}} such that ωj\omega_{j}\to\infty and

    (5.5) f~ωjf~weakly in H1(2),\displaystyle\widetilde{f}_{\omega_{j}}\rightharpoonup\widetilde{f}_{\infty}\quad\text{weakly in }H^{1}(\mathbb{R}^{2}),
    (5.6) f~ωj(0)2βα(ωj)c\displaystyle\frac{\widetilde{f}_{\omega_{j}}(0)^{2}}{\beta_{\alpha}(\omega_{j})}\to c_{\infty}

    as jj\to\infty. By (5.5) and the radial compactness lemma, we see that

    f~ωjf~in Lq(2) as j for any q>2.\widetilde{f}_{\omega_{j}}\to\widetilde{f}_{\infty}\quad\text{in $L^{q}(\mathbb{R}^{2})$ as $j\to\infty$ for any $q>2$}.

    Since βα(ωj)\beta_{\alpha}(\omega_{j})\to\infty, by (5.6) we have

    f~ωj(0)βα(ωj)0as j.\frac{\widetilde{f}_{\omega_{j}}(0)}{\beta_{\alpha}(\omega_{j})}\to 0\quad\text{as $j\to\infty$}.

    Therefore, we obtain

    ϕ~ωjf~in Lq(2) as j for any q>2.\widetilde{\phi}_{\omega_{j}}\to\widetilde{f}_{\infty}\quad\text{in $L^{q}(\mathbb{R}^{2})$ as $j\to\infty$ for any $q>2$}.

    Moreover, from Step 2 again, we see that f~0\widetilde{f}_{\infty}\neq 0.

  4. Step 4.

    f~=ϕ~\widetilde{f}_{\infty}=\widetilde{\phi}_{\infty}.

    From the equation

    Δf~ω+f~ωϕ~ωp=0-\Delta\widetilde{f}_{\omega}+\widetilde{f}_{\omega}-\widetilde{\phi}_{\omega}^{p}=0

    and Step 3, we see that f~\widetilde{f}_{\infty} is a weak solution of Equation (1.20). Therefore, the uniqueness of the nonnegative, radial, and decreasing solutions of (1.20), we obtain f~=ϕ~\widetilde{f}_{\infty}=\widetilde{\phi}_{\infty}.

  5. Step 5.

    Strong convergence in H2(2)H^{2}(\mathbb{R}^{2}), i.e.,

    (5.7) f~ωjϕ~in H2(2).\widetilde{f}_{\omega_{j}}\to\widetilde{\phi}_{\infty}\quad\text{in }H^{2}(\mathbb{R}^{2}).

    From the equations we have

    |(Δ+1)(f~ωjϕ~)|=|ϕ~ωjpϕ~p|(ϕ~ωjp1+ϕ~p1)|ϕ~ωjϕ~|.\displaystyle|(-\Delta+1)(\widetilde{f}_{\omega_{j}}-\widetilde{\phi}_{\infty})|=|\widetilde{\phi}_{\omega_{j}}^{p}-\widetilde{\phi}_{\infty}^{p}|\lesssim(\widetilde{\phi}_{\omega_{j}}^{p-1}+\widetilde{\phi}_{\infty}^{p-1})|\widetilde{\phi}_{\omega_{j}}-\widetilde{\phi}_{\infty}|.

    Thus, by Steps 3 and 4, we obtain

    f~ωjϕ~H22\displaystyle\|\widetilde{f}_{\omega_{j}}-\widetilde{\phi}_{\infty}\|_{H^{2}}^{2} (ϕ~ωj2(p1)+ϕ~2(p1))|ϕ~ωjϕ~|2𝑑x\displaystyle\lesssim\int(\widetilde{\phi}_{\omega_{j}}^{2(p-1)}+\widetilde{\phi}_{\infty}^{2(p-1)})|\widetilde{\phi}_{\omega_{j}}-\widetilde{\phi}_{\infty}|^{2}\,dx
    (ϕ~ωjL2(p+1)2(p1)+ϕ~L2(p+1)2(p1))ϕ~ωjϕ~Lp+120.\displaystyle\lesssim(\|\widetilde{\phi}_{\omega_{j}}\|_{L^{2(p+1)}}^{2(p-1)}+\|\widetilde{\phi}_{\infty}\|_{L^{2(p+1)}}^{2(p-1)})\|\widetilde{\phi}_{\omega_{j}}-\widetilde{\phi}_{\infty}\|_{L^{p+1}}^{2}\to 0.

    This means that (5.7) holds.

  6. Step 6.

    Conclusion.

    The above argument works if we start with any subsequence of (ϕ~ω)ω>eα(\widetilde{\phi}_{\omega})_{\omega>-e_{\alpha}}. Thus we have

    f~ωϕ~in H2(2) as ω\widetilde{f}_{\omega}\to\widetilde{\phi}_{\infty}\quad\text{in $H^{2}(\mathbb{R}^{2})$ as $\omega\to\infty$. }

This completes the proof. ∎

6. Nondegeneracy and lower boundedness of linearized operator in radial function space

In this section, we prove lower boundedness and nondegeneracy of the linearlized operator around rescaled ground states with the large frequency in the radial function space. We follow the argument in [28, 30].

We denote

Drad(Δα;)\displaystyle D_{\mathrm{rad}}(-\Delta_{\alpha};\mathbb{R}) :={f+f(0)βα(λ)Gλ:fHrad2(2;)},\displaystyle\mathrel{\mathop{:}}=\left\{f+\frac{f(0)}{\beta_{\alpha}(\lambda)}G_{\lambda}\colon\,f\in H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})\right\},
Hα,rad1(2;)\displaystyle H_{\alpha,\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R}) :={f+cGλ:fHrad1(2;),c}.\displaystyle\mathrel{\mathop{:}}=\{f+cG_{\lambda}\colon\,f\in H_{\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R}),\ c\in\mathbb{R}\}.

Let (ϕω)ω>eα(\phi_{\omega})_{\omega>-e_{\alpha}} be a family of positive ground states. We define the linearized operator around ϕ~ω\widetilde{\phi}_{\omega} by

L~ωv=L~α,ωv\displaystyle\widetilde{L}_{\omega}v=\widetilde{L}_{\alpha,\omega}v :=(Δα~(ω)+1)vpϕ~ωp1vfor vHα1(2;).\displaystyle\mathrel{\mathop{:}}=(-\Delta_{\widetilde{\alpha}(\omega)}+1)v-p\widetilde{\phi}_{\omega}^{p-1}v\quad\text{for }v\in H_{\alpha}^{1}(\mathbb{R}^{2};\mathbb{R}).

We denote

L~f:=(Δ+1)fpϕ~p1ffor fH2(2;).\widetilde{L}_{\infty}f\mathrel{\mathop{:}}=(-\Delta+1)f-p\widetilde{\phi}_{\infty}^{p-1}f\quad\text{for $f\in H^{2}(\mathbb{R}^{2};\mathbb{R})$}.

It is known (see, e.g., [28]) that there exists C>0C>0 such that

(6.1) fH2\displaystyle\|f\|_{H^{2}} CL~fL2for all fHrad2(2;),\displaystyle\leq C\|\widetilde{L}_{\infty}f\|_{L^{2}}\quad\text{for all $f\in H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})$},
(6.2) fH1\displaystyle\|f\|_{H^{1}} CL~fHrad1for all fHrad1(2;).\displaystyle\leq C\|\widetilde{L}_{\infty}f\|_{H_{\mathrm{rad}}^{-1}}\quad\text{for all $f\in H_{\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R})$}.
Lemma 6.1.

There exist ω1>eα\omega_{1}>-e_{\alpha} and C>0C>0 such that for all ω>ω1\omega>\omega_{1},

(6.3) vH~α,ω1CL~ωvH~α,ω,rad1for all v=f+cG1Hα,rad1(2;).\|v\|_{\widetilde{H}_{\alpha,\omega}^{1}}\leq C\|\widetilde{L}_{\omega}v\|_{\widetilde{H}_{\alpha,\omega,\mathrm{rad}}^{-1}}\quad\text{for all }v=f+cG_{1}\in H_{\alpha,\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R}).

In particular, if vHα,rad1(2;)v\in H_{\alpha,\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R}) satisfies L~ωv=0\widetilde{L}_{\omega}v=0, then v=0v=0.

Remark 6.2.

Lemma 6.1 means that zero is not an eigenvalue of the operator L~ω:Drad(Δα~;)Lrad2(2;)\widetilde{L}_{\omega}\colon D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R})\to L_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R}). In fact, by essential spectral theorem, we see that zero is its resolvent.

Proof of Lemma 6.1.

For v=f+cG1v=f+cG_{1}, w=g+dG1Hα,rad1(2;)w=g+dG_{1}\in H_{\alpha,\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R}), we have the expression

L~ωv,w=\displaystyle\langle\widetilde{L}_{\omega}v,w\rangle={} L~ω(f+cG1),g+dG1\displaystyle\langle\widetilde{L}_{\omega}(f+cG_{1}),g+dG_{1}\rangle
=\displaystyle={} (f,g)L2+(f,g)L2+cdβα(ω)p(ϕ~ωp1v,w)L2\displaystyle(\nabla f,\nabla g)_{L^{2}}+(f,g)_{L^{2}}+cd\beta_{\alpha}(\omega)-p(\widetilde{\phi}_{\omega}^{p-1}v,w)_{L^{2}}
=\displaystyle={} L~f,g+cdβα(ω)+p(ϕ~p1f,g)L2p(ϕ~ωp1v,w)L2\displaystyle\langle\widetilde{L}_{\infty}f,g\rangle+cd\beta_{\alpha}(\omega)+p(\widetilde{\phi}_{\infty}^{p-1}f,g)_{L^{2}}-p(\widetilde{\phi}_{\omega}^{p-1}v,w)_{L^{2}}
=\displaystyle={} L~f,g+cdβα(ω)+p(ϕ~p1ϕ~ωp1)fg\displaystyle\langle\widetilde{L}_{\infty}f,g\rangle+cd\beta_{\alpha}(\omega)+p\int(\widetilde{\phi}_{\infty}^{p-1}-\widetilde{\phi}_{\omega}^{p-1})fg
pϕ~ωp1(df+cg)G1pcdϕ~ωp1G12.\displaystyle-p\int\widetilde{\phi}_{\omega}^{p-1}(df+cg)G_{1}-pcd\int\widetilde{\phi}_{\omega}^{p-1}G_{1}^{2}.

Therefore, using the estimate (6.2), we have

L~ωvH~α,ω,rad1\displaystyle\|\widetilde{L}_{\omega}v\|_{\widetilde{H}_{\alpha,\omega,\mathrm{rad}}^{-1}} =sup{L~ωv,w:wH~α,ω,rad11}\displaystyle=\sup\{\langle\widetilde{L}_{\omega}v,w\rangle\colon\,\|w\|_{\widetilde{H}_{\alpha,\omega,\mathrm{rad}}^{1}}\leq 1\}
sup{L~ω(f+cG1),g+dG1:gHrad121/2,d2βα(ω)1/2}\displaystyle\geq\sup\biggl{\{}\langle\widetilde{L}_{\omega}(f+cG_{1}),g+dG_{1}\rangle\colon\,\begin{subarray}{c}\|g\|_{H_{\mathrm{rad}}^{1}}^{2}\leq 1/2,\\ d^{2}\beta_{\alpha}(\omega)\leq 1/2\end{subarray}\biggr{\}}
=sup{L~f,g+cdβα(ω)+p(ϕ~p1ϕ~ωp1)fgpϕ~ωp1(df+cg)G1pcdϕ~ωp1G12:gHrad121/2,|d|(2βα(ω))1/2}\displaystyle=\sup\biggl{\{}\begin{aligned} &\langle\widetilde{L}_{\infty}f,g\rangle+cd\beta_{\alpha}(\omega)+p\int(\widetilde{\phi}_{\infty}^{p-1}-\widetilde{\phi}_{\omega}^{p-1})fg\\ &-p\int\widetilde{\phi}_{\omega}^{p-1}(df+cg)G_{1}-pcd\int\widetilde{\phi}_{\omega}^{p-1}G_{1}^{2}\colon\,\begin{subarray}{c}\|g\|_{H_{\mathrm{rad}}^{1}}\leq 2^{-1/2},\\ |d|\leq(2\beta_{\alpha}(\omega))^{-1/2}\end{subarray}\biggr{\}}\end{aligned}
L~fHrad1+βα(ω)1/2|c|\displaystyle\gtrsim\|\widetilde{L}_{\infty}f\|_{H_{\mathrm{rad}}^{-1}}+\beta_{\alpha}(\omega)^{1/2}|c|
ϕ~ϕ~ωLp+1fLp+1fH1βα(ω)1/2|c||c|βα(ω)1/2\displaystyle-\|\widetilde{\phi}_{\infty}-\widetilde{\phi}_{\omega}\|_{L^{p+1}}\|f\|_{L^{p+1}}-\frac{\|f\|_{H^{1}}}{\beta_{\alpha}(\omega)^{1/2}}-|c|-\frac{|c|}{\beta_{\alpha}(\omega)^{1/2}}
fH1+βα(ω)1/2|c|vH~α,ω1\displaystyle\gtrsim\|f\|_{H^{1}}+\beta_{\alpha}(\omega)^{1/2}|c|\simeq\|v\|_{\widetilde{H}_{\alpha,\omega}^{1}}

for large ω\omega, where we used the fact βα(ω)\beta_{\alpha}(\omega)\to\infty as ω\omega\to\infty. This completes the proof. ∎

7. Uniqueness of ground states for large frequencies

In this section, we prove the uniqueness of ground states for large frequencies (Theorem 1.4). The proof is based on [30].

Lemma 7.1.

There exist δ>0\delta>0 and ω1>eα\omega_{1}>-e_{\alpha} such that the following holds. If (ϕω)ω>eα(\phi_{\omega})_{\omega>-e_{\alpha}} is a family of positive ground states with ϕω𝒢ω\phi_{\omega}\in\mathcal{G}_{\omega}, ω0>ω1\omega_{0}>\omega_{1}, and ψ𝒜ω0\psi\in\mathcal{A}_{\omega_{0}} is a real-valued radial solution satisfying ψ~ϕ~Hα1<δ\|\widetilde{\psi}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}<\delta, then ψ=ϕω0\psi=\phi_{\omega_{0}}.

Proof.

We take a small δ>0\delta>0 to be chosen later, and let ψ~ϕ~Hα1<δ\|\widetilde{\psi}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}<\delta. We define the operator

L~ωv:=(Δα~(ω)+1)vVω(x)v,\widetilde{L}_{\omega}^{*}v\mathrel{\mathop{:}}=(-\Delta_{\widetilde{\alpha}(\omega)}+1)v-V_{\omega}(x)v,

where

Vω(x):={ϕ~ω(x)p|ψ~(x)|p1ψ~(x)ϕ~ω(x)ψ~(x)if ϕω(x)ψ(x),pϕ~ω(x)p1if ϕω(x)=ψ(x).V_{\omega}(x)\mathrel{\mathop{:}}=\left\{\begin{array}[]{@{}c l @{}}\dfrac{\widetilde{\phi}_{\omega}(x)^{p}-|\widetilde{\psi}(x)|^{p-1}\widetilde{\psi}(x)}{\widetilde{\phi}_{\omega}(x)-\widetilde{\psi}(x)}&\text{if }\phi_{\omega}(x)\neq\psi(x),\\[12.0pt] p\widetilde{\phi}_{\omega}(x)^{p-1}&\text{if }\phi_{\omega}(x)=\psi(x).\end{array}\right.

First, we show that there exist ω1>eα\omega_{1}>-e_{\alpha} and C>0C>0 such that if ω>ω1\omega>\omega_{1}, then

(7.1) fH2CL~ωvL2for all v=f+f(0)βα(ω)G1Drad(Δα~;).\|f\|_{H^{2}}\leq C\|\widetilde{L}_{\omega}^{*}v\|_{L^{2}}\quad\text{for all }v=f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{1}\in D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R}).

To prove this, we will show that

(7.2) Vωpϕ~p1Lqδfor all q2 and large ω.\|V_{\omega}-p\widetilde{\phi}_{\infty}^{p-1}\|_{L^{q}}\lesssim\delta\quad\text{for all $q\geq 2$ and large $\omega$}.

We can rewrite

Vω(x)=p01|ϕ~ω(x)+srω(x)|p1ds,rω(x):=ψ~(x)ϕ~ω(x).V_{\omega}(x)=p\int_{0}^{1}|\widetilde{\phi}_{\omega}(x)+sr_{\omega}(x)|^{p-1}\,ds,\quad r_{\omega}(x)\mathrel{\mathop{:}}=\widetilde{\psi}(x)-\widetilde{\phi}_{\omega}(x).

Then we have

|Vωpϕ~p1|\displaystyle|V_{\omega}-p\widetilde{\phi}_{\infty}^{p-1}|
p01||ϕ~ω+srω|p1ϕ~p1|𝑑s\displaystyle\leq p\int_{0}^{1}\left||\widetilde{\phi}_{\omega}+sr_{\omega}|^{p-1}-\widetilde{\phi}_{\infty}^{p-1}\right|\,ds
{(ϕ~ωp2+|rω|p2+ϕ~p2)(|ϕ~ωϕ~|+|rω|)if p>2|ϕ~ωϕ~|+|rω|if 1<p2.\displaystyle\lesssim\left\{\begin{aligned} &(\widetilde{\phi}_{\omega}^{p-2}+|r_{\omega}|^{p-2}+\widetilde{\phi}_{\infty}^{p-2})(|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}|+|r_{\omega}|)&&\text{if $p>2$}\\ &|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}|+|r_{\omega}|&&\text{if $1<p\leq 2$}.\end{aligned}\right.

We note that Proposition 5.2 implies ϕ~ωϕ~Hα1<δ\|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}<\delta for large ω\omega. Therefore, if p>2p>2, then

Vωpϕ~p1Lq\displaystyle\|V_{\omega}-p\widetilde{\phi}_{\infty}^{p-1}\|_{L^{q}}\lesssim{} (ϕ~ωL(p1)qp2+rωL(p1)qp2+ϕ~L(p1)qp2)\displaystyle(\|\widetilde{\phi}_{\omega}\|_{L^{(p-1)q}}^{p-2}+\|r_{\omega}\|_{L^{(p-1)q}}^{p-2}+\|\widetilde{\phi}_{\infty}\|_{L^{(p-1)q}}^{p-2})
(ϕ~ωϕ~L(p1)q+rωL(p1)q)\displaystyle\cdot(\|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}\|_{L^{(p-1)q}}+\|r_{\omega}\|_{L^{(p-1)q}})
\displaystyle\lesssim{} ϕ~ωϕ~Hα1+ψ~ϕ~Hα1δ.\displaystyle\|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}+\|\widetilde{\psi}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}\lesssim\delta.

If 1<p21<p\leq 2, we have

Vωpϕ~p1Lq\displaystyle\|V_{\omega}-p\widetilde{\phi}_{\infty}^{p-1}\|_{L^{q}} ϕ~ωϕ~Lq+rωLq\displaystyle\lesssim\|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}\|_{L^{q}}+\|r_{\omega}\|_{L^{q}}
ϕ~ωϕ~Hα1+ψ~ϕ~Hα1δ.\displaystyle\lesssim\|\widetilde{\phi}_{\omega}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}+\|\widetilde{\psi}-\widetilde{\phi}_{\infty}\|_{H_{\alpha}^{1}}\lesssim\delta.

Therefore, we have (7.2).

Next, we show the estimate (7.1). By using the expression

L~ωv\displaystyle\widetilde{L}_{\omega}^{*}v =L~f+pϕ~p1fVω(x)v\displaystyle=\widetilde{L}_{\infty}f+p\widetilde{\phi}_{\infty}^{p-1}f-V_{\omega}(x)v
=L~f+(pϕ~p1Vω(x))ff(0)βα(ω)Vω(x)G1\displaystyle=\widetilde{L}_{\infty}f+(p\widetilde{\phi}_{\infty}^{p-1}-V_{\omega}(x))f-\frac{f(0)}{\beta_{\alpha}(\omega)}V_{\omega}(x)G_{1}

for v=f+f(0)βα(ω)1G1Drad(Δα~;)v=f+f(0)\beta_{\alpha}(\omega)^{-1}G_{1}\in D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R}), (6.1), (7.2), and the embedding |f(0)|fH2|f(0)|\lesssim\|f\|_{H^{2}} we obtain

L~ωvL2\displaystyle\|\widetilde{L}_{\omega}^{*}v\|_{L^{2}} L~fL2(pϕ~p1Vω)fL2|f(0)|βα(ω)VωG1L2\displaystyle\geq\|\widetilde{L}_{\infty}f\|_{L^{2}}-\|(p\widetilde{\phi}_{\infty}^{p-1}-V_{\omega})f\|_{L^{2}}-\frac{|f(0)|}{\beta_{\alpha}(\omega)}\|V_{\omega}G_{1}\|_{L^{2}}
fH2δfH2fH2\displaystyle\gtrsim\|f\|_{H^{2}}-\delta\|f\|_{H^{2}}\simeq\|f\|_{H^{2}}

if δ\delta is small and ω\omega is large sufficiently. This implies that the estimate (7.1) holds.

Finally, from the equation (5.2), we have L~ω(ϕ~ωψ~)=0\widetilde{L}_{\omega}^{*}(\widetilde{\phi}_{\omega}-\widetilde{\psi})=0. By (7.1) we obtain ϕ~ω=ψ~\widetilde{\phi}_{\omega}=\widetilde{\psi}. This completes the proof. ∎

Proof of Theorem 1.4.

The assertion follows from Theorem 1.3, Proposition 5.2, and Lemma 7.1. ∎

8. Regularity of ωϕω\omega\mapsto\phi_{\omega}

In this section, we verify the differentiability of ωϕω\omega\mapsto\phi_{\omega} for large ω\omega following the argument in [55, Section 6] with modifications.

Proposition 8.1.

Let ω1>eα\omega_{1}>-e_{\alpha} be as in Lemma 6.1 and Lemma 7.1 and let ϕω=fω+fω(0)βα(ω)1Gω𝒢ω\phi_{\omega}=f_{\omega}+f_{\omega}(0)\beta_{\alpha}(\omega)^{-1}G_{\omega}\in\mathcal{G}_{\omega} be the unique positive radial ground state given in Theorem 1.4. Then the map ωf~ω\omega\mapsto\widetilde{f}_{\omega} is in C1((ω1,),Hrad2(2;))C^{1}((\omega_{1},\infty),H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})).

Proof.

We define the function FF by

F(ω,f):=f(1Δ)1[|f+f(0)βα(ω)G1|p1(f+f(0)βα(ω)G1)],\displaystyle F(\omega,f)\mathrel{\mathop{:}}=f-(1-\Delta)^{-1}\left[\left|f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{1}\right|^{p-1}\left(f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{1}\right)\right],
(ω,f)(eα,)×Hrad2(2;).\displaystyle(\omega,f)\in(-e_{\alpha},\infty)\times H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R}).

Then we have

Fω(ω,f)=\displaystyle\frac{\partial F}{\partial\omega}(\omega,f)={} pf(0)4πωβα(ω)2(1Δ)1[|f+f(0)βα(ω)G1|p1G1],\displaystyle-\frac{pf(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}(1-\Delta)^{-1}\left[\left|f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{1}\right|^{p-1}G_{1}\right],

and for wHrad2(2;)w\in H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R}),

Ff(ω,f)w=\displaystyle\frac{\partial F}{\partial f}(\omega,f)w={} wp(1Δ)1[|f+f(0)βα(ω)G1|p1(w+w(0)βα(ω)G1)].\displaystyle w-p(1-\Delta)^{-1}\left[\left|f+\frac{f(0)}{\beta_{\alpha}(\omega)}G_{1}\right|^{p-1}\left(w+\frac{w(0)}{\beta_{\alpha}(\omega)}G_{1}\right)\right].

From this expression we have

FC1((eα,)×Hrad2(2;),Hrad2(2;)).F\in C^{1}\bigl{(}(-e_{\alpha},\infty)\times H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R}),H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})\bigr{)}.

Let ω0>ω1\omega_{0}>\omega_{1}. We have

F(ω0,f~ω0)=(1Δ)1S~ω0(ϕ~ω0)=0.F(\omega_{0},\widetilde{f}_{\omega_{0}})=(1-\Delta)^{-1}\widetilde{S}_{\omega_{0}}^{\prime}(\widetilde{\phi}_{\omega_{0}})=0.

Moreover, since the operator L~ω0=(1Δα~)pϕ~ω0p1:Drad(Δα~;)L2(2;)\widetilde{L}_{\omega_{0}}=(1-\Delta_{\widetilde{\alpha}})-p\widetilde{\phi}_{\omega_{0}}^{p-1}\colon D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R})\to L^{2}(\mathbb{R}^{2};\mathbb{R}) is invertible by Remark 6.2 and since the map τω:Hrad2(2;)Drad(Δα~;)\tau_{\omega}\colon H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})\to D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R}); ww+w(0)βα~(1)1G1w\mapsto w+w(0)\beta_{\widetilde{\alpha}}(1)^{-1}G_{1} is also invertible by the definition of Drad(Δα~;)D_{\mathrm{rad}}(-\Delta_{\widetilde{\alpha}};\mathbb{R}), we see that the operator

Ff(ω0,f~ω0)\displaystyle\frac{\partial F}{\partial f}(\omega_{0},\widetilde{f}_{\omega_{0}}) =I(1Δ)1pϕ~ω0p1\displaystyle=I-(1-\Delta)^{-1}p\widetilde{\phi}_{\omega_{0}}^{p-1}
=(1Δ)1L~ω0τω0:Hrad2(2;)Hrad2(2;)\displaystyle=(1-\Delta)^{-1}\widetilde{L}_{\omega_{0}}\tau_{\omega_{0}}\colon H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})\to H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R})

is also invertible. Therefore, by the implicit function theorem, there exists a C1C^{1}-curve ωgω\omega\mapsto g_{\omega} defined on a neighborhood of ω0\omega_{0} into Hrad2(2;)H_{\mathrm{rad}}^{2}(\mathbb{R}^{2};\mathbb{R}) such that F(ω,gω)=0F(\omega,g_{\omega})=0 and gω0=f~ω0g_{\omega_{0}}=\widetilde{f}_{\omega_{0}}. From Lemma 7.1 we have gω=f~ωg_{\omega}=\widetilde{f}_{\omega} for ω\omega around ω0\omega_{0}. This completes the proof. ∎

Note that by a standard elliptic regularity argument, one can obtain the spatially exponential decay of fωf_{\omega}. Thus, by Proposition 8.1 and the definition of the rescaling f~ω\widetilde{f}_{\omega}, we obtain the regularity of ωfω\omega\mapsto f_{\omega}. In particular, we have

Corollary 8.2.

Let ω1>eα\omega_{1}>-e_{\alpha} be as in Lemma 6.1 and Lemma 7.1 and let ϕω\phi_{\omega} be the unique positive radial ground state given in Theorem 1.4. Then the map ωϕω\omega\mapsto\phi_{\omega} is in C1((ω1,),Hα,rad1(2;))C^{1}((\omega_{1},\infty),H_{\alpha,\mathrm{rad}}^{1}(\mathbb{R}^{2};\mathbb{R})).

9. Stability and instability for large frequencies

In this section, we calculate the value ddωϕωL22\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2} for large ω\omega based on [28, 30]. We prove the following.

Proposition 9.1.

Let (ϕω)ω>ω1(\phi_{\omega})_{\omega>\omega_{1}} be the family of the unique positive ground states obtained in Theorem 1.4. Then there exists ω=ω(p)(ω1,)\omega^{*}=\omega^{*}(p)\in(\omega_{1},\infty) such that the following is true.

  • If 1<p31<p\leq 3, then ddωϕωL22>0\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}>0 for all ω>ω\omega>\omega^{*}.

  • If p>3p>3, then ddωϕωL22<0\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}<0 for all ω>ω\omega>\omega^{*}.

We note that the rescaled ground state ϕ~ω=f~ω+f~(0)βα(ω)1G1\widetilde{\phi}_{\omega}=\widetilde{f}_{\omega}+\widetilde{f}(0)\beta_{\alpha}(\omega)^{-1}G_{1} satisfies the equation

(9.1) (Δ+1)f~ωϕ~ωp=0.(-\Delta+1)\widetilde{f}_{\omega}-\widetilde{\phi}_{\omega}^{p}=0.

Moreover, for ω>ω1\omega>\omega_{1}, where ω1\omega_{1} is as in Proposition 8.1, the derivative ωf~ω\partial_{\omega}\widetilde{f}_{\omega} is in H2(2)H^{2}(\mathbb{R}^{2}), and ωf~ω(0)\partial_{\omega}\widetilde{f}_{\omega}(0) makes sense for

We prepare some lemmas.

Lemma 9.2.

For ω>ω1\omega>\omega_{1}, the following Pohozaev identity holds:

(9.2) ϕ~ωL22=f~ω(0)24πβα(ω)2+2p+1ϕ~ωLp+1p+1.\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}=\frac{\widetilde{f}_{\omega}(0)^{2}}{4\pi\beta_{\alpha}(\omega)^{2}}+\frac{2}{p+1}\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}^{p+1}.

In particular,

(9.3) ddωϕ~ωL22=f~ω(0)ωf~(0)2πβα(ω)2f~ω(0)28π2ωβα(ω)3+2ϕ~ωpωϕ~ω.\frac{d}{d\omega}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}=\frac{\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}(0)}{2\pi\beta_{\alpha}(\omega)^{2}}-\frac{\widetilde{f}_{\omega}(0)^{2}}{8\pi^{2}\omega\beta_{\alpha}(\omega)^{3}}+2\int\widetilde{\phi}_{\omega}^{p}\partial_{\omega}\widetilde{\phi}_{\omega}.
Proof.

By multiplying xϕ~ωx\cdot\nabla\widetilde{\phi}_{\omega} with the equation (9.1) and integrating it, we have

(9.4) (Δ+1)f~ω,xf~ω+f~ω(0)βα(ω)(Δ+1)f~ω,xG1=ϕ~ωp,xϕ~ω.\langle(-\Delta+1)\widetilde{f}_{\omega},x\cdot\nabla\widetilde{f}_{\omega}\rangle+\frac{\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}\langle(-\Delta+1)\widetilde{f}_{\omega},x\cdot\nabla G_{1}\rangle=\langle\widetilde{\phi}_{\omega}^{p},x\cdot\nabla\widetilde{\phi}_{\omega}\rangle.

From properties of the scaling, we have

(9.5) (Δ+1)f,xf=12ddλf(λ)H12|λ=1=fL22,\displaystyle\langle(-\Delta+1)f,x\cdot\nabla f\rangle=\left.\frac{1}{2}\frac{d}{d\lambda}\|f(\lambda\cdot)\|_{H^{1}}^{2}\right|_{\lambda=1}=-\|f\|_{L^{2}}^{2},
(9.6) ϕp,xϕ=1p+1ddλϕ(λ)Lp+1p+1|λ=1=2p+1ϕLp+1p+1.\displaystyle\langle\phi^{p},x\cdot\nabla\phi\rangle=\left.\frac{1}{p+1}\frac{d}{d\lambda}\|\phi(\lambda\cdot)\|_{L^{p+1}}^{p+1}\right|_{\lambda=1}=-\frac{2}{p+1}\|\phi\|_{L^{p+1}}^{p+1}.

Moreover, by Lemma 2.3 we have

(9.7) (Δ+1)f,xG1=2(f,G1)L2.\langle(-\Delta+1)f,x\cdot\nabla G_{1}\rangle=-2(f,G_{1})_{L^{2}}.

By using (9.5), (9.6), (9.7), we can rewrite (9.4) as

(9.8) f~ωL22+2f~ω(0)βα(ω)(f~ω,G1)L2=2p+1ϕ~ωLp+1p+1.\|\widetilde{f}_{\omega}\|_{L^{2}}^{2}+2\frac{\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}(\widetilde{f}_{\omega},G_{1})_{L^{2}}=\frac{2}{p+1}\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}^{p+1}.

Moreover, from the expression

ϕ~ωL22=f~ω+f~ω(0)βα(ω)G1L22=f~ωL22+2f~ω(0)βα(ω)(f~ω,G1)L2+f~ω(0)24πβα(ω)2,\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}=\left\|\widetilde{f}_{\omega}+\frac{\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}\right\|_{L^{2}}^{2}=\|\widetilde{f}_{\omega}\|_{L^{2}}^{2}+2\frac{\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}(\widetilde{f}_{\omega},G_{1})_{L^{2}}+\frac{\widetilde{f}_{\omega}(0)^{2}}{4\pi\beta_{\alpha}(\omega)^{2}},

we obtain (9.2). By differentiating (9.2) we have (9.3). ∎

Lemma 9.3.

For ω>ω1\omega>\omega_{1},

(p1)ϕ~ωpωϕ~ω=f~ω(0)24πωβα(ω)2.(p-1)\int\widetilde{\phi}_{\omega}^{p}\partial_{\omega}\widetilde{\phi}_{\omega}=\frac{\widetilde{f}_{\omega}(0)^{2}}{4\pi\omega\beta_{\alpha}(\omega)^{2}}.
Proof.

By the equation (9.1), we have (Δ+1)f~ωϕ~ωp,ωϕ~ω=0\langle(-\Delta+1)\widetilde{f}_{\omega}-\widetilde{\phi}_{\omega}^{p},\partial_{\omega}\widetilde{\phi}_{\omega}\rangle=0. This can be rewritten from the expression

(9.9) ωϕ~ω=ωf~ω+ωf~ω(0)βα(ω)G1f~ω(0)4πωβα(ω)2G1\partial_{\omega}\widetilde{\phi}_{\omega}=\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}-\frac{\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}G_{1}

as

(9.10) ϕ~ωpωϕ~ω\displaystyle\int\widetilde{\phi}_{\omega}^{p}\partial_{\omega}\widetilde{\phi}_{\omega} =(Δ+1)f~ω,ωf~ω+ωf~ω(0)βα(ω)G1f~ω(0)4πωβα(ω)2G1\displaystyle=\langle(-\Delta+1)\widetilde{f}_{\omega},\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}-\frac{\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}G_{1}\rangle
=(Δ+1)f~ω,ωf~ω+f~ω(0)ωf~ω(0)βα(ω)f~ω(0)24πωβα(ω)2,\displaystyle=\langle(-\Delta+1)\widetilde{f}_{\omega},\partial_{\omega}\widetilde{f}_{\omega}\rangle+\frac{\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}-\frac{\widetilde{f}_{\omega}(0)^{2}}{4\pi\omega\beta_{\alpha}(\omega)^{2}},

where we used the fact that G1G_{1} is a solution of (Δ+1)G1=δ0(-\Delta+1)G_{1}=\delta_{0}.

By differentiating the equation (9.1) with respect to ω\omega, we have

(9.11) (Δ+1)ωf~ωpϕ~ωp1ωϕ~ω=0.(-\Delta+1)\partial_{\omega}\widetilde{f}_{\omega}-p\widetilde{\phi}_{\omega}^{p-1}\partial_{\omega}\widetilde{\phi}_{\omega}=0.

By multiplying this equation with ϕ~ω\widetilde{\phi}_{\omega} and integrating it, we have

(9.12) pϕ~ωpωϕ~ω\displaystyle p\int\widetilde{\phi}_{\omega}^{p}\partial_{\omega}\widetilde{\phi}_{\omega} =(Δ+1)ωf~ω,ϕ~ω\displaystyle=\langle(-\Delta+1)\partial_{\omega}\widetilde{f}_{\omega},\widetilde{\phi}_{\omega}\rangle
=(Δ+1)ωf~ω,f~ω+f~ω(0)ωf~ω(0)βα(ω).\displaystyle=\langle(-\Delta+1)\partial_{\omega}\widetilde{f}_{\omega},\widetilde{f}_{\omega}\rangle+\frac{\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}.

Therefore, the assertion follows from (9.10) and (9.12). ∎

Lemma 9.4.

There exists C>0C>0 such that for ω>ω1\omega>\omega_{1},

(9.13) |ωf~ω(0)|\displaystyle|\partial_{\omega}\widetilde{f}_{\omega}(0)| Cωβα(ω)3/2.\displaystyle\leq\frac{C}{\omega\beta_{\alpha}(\omega)^{3/2}}.
Proof.

Noting that βα~(1)=βα(ω)\beta_{\widetilde{\alpha}}(1)=\beta_{\alpha}(\omega), we have the relation

(Δ+1)ωf~ω=(Δα~+1)(ωf~ω+ωf~ω(0)βα(ω)G1).(-\Delta+1)\partial_{\omega}\widetilde{f}_{\omega}=(-\Delta_{\widetilde{\alpha}}+1)\left(\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}\right).

Therefore, by the expression (9.9), we can rewrite (9.11) as

L~ω(ωf~ω+ωf~ω(0)βα(ω)G1)=pf~ω(0)4πωβα(ω)2ϕ~ωp1G1.\displaystyle\widetilde{L}_{\omega}\left(\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}\right)=-\frac{p\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}\widetilde{\phi}_{\omega}^{p-1}G_{1}.

Thus, by the invertibility of L~ω\widetilde{L}_{\omega} in the radial space (Lemma 6.1), we obtain

ωf~ω+ωf~ω(0)βα(ω)G1=pf~ω(0)4πωβα(ω)2(L~ω)1(ϕ~ωp1G1).\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}=-\frac{p\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}(\widetilde{L}_{\omega})^{-1}(\widetilde{\phi}_{\omega}^{p-1}G_{1}).

From this expression and the definition of the bilinear form for Δα-\Delta_{\alpha}, we have the estimate

(9.14) |ωf~ω(0)|\displaystyle|\partial_{\omega}\widetilde{f}_{\omega}(0)| =|(Δα~+1)G1,ωf~ω+ωf~ω(0)βα(ω)G1|\displaystyle=\biggl{|}\biggl{\langle}(-\Delta_{\widetilde{\alpha}}+1)G_{1},\partial_{\omega}\widetilde{f}_{\omega}+\frac{\partial_{\omega}\widetilde{f}_{\omega}(0)}{\beta_{\alpha}(\omega)}G_{1}\biggr{\rangle}\biggr{|}
=pf~ω(0)4πωβα(ω)2|(Δα~+1)G1,(L~ω)1(ϕ~ωp1G1)|\displaystyle=\frac{p\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}\left|\left\langle(-\Delta_{\widetilde{\alpha}}+1)G_{1},(\widetilde{L}_{\omega})^{-1}(\widetilde{\phi}_{\omega}^{p-1}G_{1})\right\rangle\right|
pf~ω(0)4πωβα(ω)2(Δα~+1)G1H~α,ω1(L~ω)1(ϕ~ωp1G1)H~α,ω1.\displaystyle\leq\frac{p\widetilde{f}_{\omega}(0)}{4\pi\omega\beta_{\alpha}(\omega)^{2}}\|(-\Delta_{\widetilde{\alpha}}+1)G_{1}\|_{\widetilde{H}_{\alpha,\omega}^{-1}}\|(\widetilde{L}_{\omega})^{-1}(\widetilde{\phi}_{\omega}^{p-1}G_{1})\|_{\widetilde{H}_{\alpha,\omega}^{1}}.

A Direct calculation gives

(9.15) (Δα~+1)G1H~α,ω1\displaystyle\|(-\Delta_{\widetilde{\alpha}}+1)G_{1}\|_{\widetilde{H}_{\alpha,\omega}^{-1}} =sup{(Δα~+1)G1,w:wH~α,ω11}\displaystyle=\sup\{\langle(-\Delta_{\widetilde{\alpha}}+1)G_{1},w\rangle\colon\,\|w\|_{\widetilde{H}_{\alpha,\omega}^{1}}\leq 1\}
=sup{dβα(ω):|d|βα(ω)1/2}=βα(ω)1/2.\displaystyle=\sup\{d\beta_{\alpha}(\omega)\colon\,|d|\leq\beta_{\alpha}(\omega)^{-1/2}\}=\beta_{\alpha}(\omega)^{1/2}.

From Lemma 6.1 we obtain

(9.16) (L~ω)1(ϕ~ωp1G1)H~α,ω1\displaystyle\|(\widetilde{L}_{\omega})^{-1}(\widetilde{\phi}_{\omega}^{p-1}G_{1})\|_{\widetilde{H}_{\alpha,\omega}^{1}} ϕ~ωp1G1H~α,ω,rad1=sup{ϕ~ωp1G1,w:wH~α,ω,rad11}\displaystyle\lesssim\|\widetilde{\phi}_{\omega}^{p-1}G_{1}\|_{\widetilde{H}_{\alpha,\omega,\mathrm{rad}}^{-1}}=\sup\{\langle\widetilde{\phi}_{\omega}^{p-1}G_{1},w\rangle\colon\,\|w\|_{\widetilde{H}_{\alpha,\omega,\mathrm{rad}}^{1}}\leq 1\}
sup{ϕ~ωp1G1,g+dG1:gHrad11,d2βα(ω)1}\displaystyle\leq\sup\Bigl{\{}\langle\widetilde{\phi}_{\omega}^{p-1}G_{1},g+dG_{1}\rangle\colon\,\begin{subarray}{c}\|g\|_{H_{\mathrm{rad}}^{1}}\leq 1,\\ d^{2}\beta_{\alpha}(\omega)\leq 1\end{subarray}\Bigr{\}}
ϕ~ωLp+1p1G1Lp+1+1βα(ω)1/2ϕ~ωLp+1p1G1Lp+12.\displaystyle\lesssim\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}^{p-1}\|G_{1}\|_{L^{p+1}}+\frac{1}{\beta_{\alpha}(\omega)^{1/2}}\|\widetilde{\phi}_{\omega}\|_{L^{p+1}}^{p-1}\|G_{1}\|_{L^{p+1}}^{2}.

Combining the estimates (9.14), (9.15), and (9.16) and using the bounds from Proposition 5.2, we obtain the conclusion. ∎

Proof of Proposition 9.1.

By the definition of the scaling ϕ~ω\widetilde{\phi}_{\omega}, we have

ϕωL22=ω(3p)/(p1)ϕ~ωL22.\displaystyle\|\phi_{\omega}\|_{L^{2}}^{2}=\omega^{(3-p)/(p-1)}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}.

By (9.3) and Lemma 9.3, we have

ddωϕωL22\displaystyle\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2} =ω2(2p)/(p1)(3pp1ϕ~ωL22+ωddωϕ~ωL22)\displaystyle=\omega^{2(2-p)/(p-1)}\left(\frac{3-p}{p-1}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}+\omega\frac{d}{d\omega}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}\right)
=ω2(2p)/(p1)(3pp1ϕ~ωL22+ωf~ω(0)ωf~ω(0)2πβα(ω)2f~ω(0)28π2βα(ω)3+2ωϕ~ωpωϕ~ω)\displaystyle=\omega^{2(2-p)/(p-1)}\biggl{(}\begin{aligned} &\frac{3-p}{p-1}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}+\frac{\omega\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}_{\omega}(0)}{2\pi\beta_{\alpha}(\omega)^{2}}\\ &\quad-\frac{\widetilde{f}_{\omega}(0)^{2}}{8\pi^{2}\beta_{\alpha}(\omega)^{3}}+2\omega\int\widetilde{\phi}_{\omega}^{p}\partial_{\omega}\widetilde{\phi}_{\omega}\biggr{)}\end{aligned}
=ω2(2p)/(p1)(3pp1ϕ~ωL22+ωf~ω(0)ωf~ω(0)2πβα(ω)2f~ω(0)28π2βα(ω)3+f~ω(0)22(p1)πβα(ω)2).\displaystyle=\omega^{2(2-p)/(p-1)}\biggl{(}\begin{aligned} &\frac{3-p}{p-1}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}+\frac{\omega\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}_{\omega}(0)}{2\pi\beta_{\alpha}(\omega)^{2}}\\ &\quad-\frac{\widetilde{f}_{\omega}(0)^{2}}{8\pi^{2}\beta_{\alpha}(\omega)^{3}}+\frac{\widetilde{f}_{\omega}(0)^{2}}{2(p-1)\pi\beta_{\alpha}(\omega)^{2}}\biggr{)}.\end{aligned}

Note that Lemma 9.4 implies

|ωf~ω(0)ωf~ω(0)2πβα(ω)2|1βα(ω)7/2.\left|\frac{\omega\widetilde{f}_{\omega}(0)\partial_{\omega}\widetilde{f}_{\omega}(0)}{2\pi\beta_{\alpha}(\omega)^{2}}\right|\lesssim\frac{1}{\beta_{\alpha}(\omega)^{7/2}}.

Therefore, we obtain

ddωϕωL22=ω2(2p)/(p1)(3pp1ϕ~ωL22+f~ω(0)22(p1)πβα(ω)2+O(1βα(ω)3))\frac{d}{d\omega}\|\phi_{\omega}\|_{L^{2}}^{2}=\omega^{2(2-p)/(p-1)}\left(\frac{3-p}{p-1}\|\widetilde{\phi}_{\omega}\|_{L^{2}}^{2}+\frac{\widetilde{f}_{\omega}(0)^{2}}{2(p-1)\pi\beta_{\alpha}(\omega)^{2}}+O\left(\frac{1}{\beta_{\alpha}(\omega)^{3}}\right)\right)

as ω\omega\to\infty. By Proposition 5.2, ϕ~ωL2\|\widetilde{\phi}_{\omega}\|_{L^{2}} and f~ω(0)\widetilde{f}_{\omega}(0) converge to positive constants as ω\omega\to\infty. Therefore, we deduce the conclusion. ∎

Proof of Theorem 1.7.

The assertion follows from Propositions 9.1 and 1.10. This completes the proof. ∎

Appendix A Review of the properties of Laplace operator with point interaction

Let us review of the properties of the operator Δα-\Delta_{\alpha}. An important feature of the family Δα-\Delta_{\alpha} with α\alpha\in\mathbb{R} is the following explicit formula for the resolvent, valid for every λ>0\lambda>0.

(A.1) (Δα+λ)1g=(Δ+λ)1g+(g,Gλ)L2βα(λ)Gλ.(-\Delta_{\alpha}+\lambda)^{-1}g=(-\Delta+\lambda)^{-1}g+\frac{(g,G_{\lambda})_{L^{2}}}{\beta_{\alpha}(\lambda)}G_{\lambda}.

Identity (A.1) says that the resolvent of Δα-\Delta_{\alpha} is a rank-one perturbation of the free resolvent. As a consequence, it is possible to deduce the spectral properties (1.10) and (1.11) of Δα-\Delta_{\alpha}.

One can apply Strichartz estimates for the non-negative self-adjoint operator Δαeα-\Delta_{\alpha}-e_{\alpha} since [22, Theorem 1.3] guarantees the existence of wave operators

W±=limt±eitΔαeitΔW_{\pm}=\lim_{t\rightarrow\pm\infty}e^{it\Delta_{\alpha}}e^{-it\Delta}

in Lp(2)L^{p}(\mathbb{R}^{2}) for 1<p<1<p<\infty. We know also that W±W_{\pm} are complete in the sense that ran W±=Lac2(Δα)W_{\pm}=L_{\mathrm{ac}}^{2}(-\Delta_{\alpha}), the absolutely continuous subspace of L2(2)L^{2}(\mathbb{R}^{2}) for Δα-\Delta_{\alpha}. In our case this is the space

Lac2(Δα)={fL2(2):(f,χα)L2=0},L_{\mathrm{ac}}^{2}(-\Delta_{\alpha})=\{f\in L^{2}(\mathbb{R}^{2})\colon\,(f,\chi_{\alpha})_{L^{2}}=0\},

so we have

W±W±=1,\displaystyle W_{\pm}^{*}W_{\pm}=1, W±W±=Pac(Δα),\displaystyle W_{\pm}W_{\pm}^{*}=P_{\mathrm{ac}}(-\Delta_{\alpha}),

where Pac(Δα)P_{\mathrm{ac}}(-\Delta_{\alpha}) is the orthogonal projection onto Lac2(Δα)L_{\mathrm{ac}}^{2}(-\Delta_{\alpha}). The wave operators satisfy the intertwining property

f(Δα)Pac(Δα)=W±f(Δ)W±f(-\Delta_{\alpha})P_{\mathrm{ac}}(-\Delta_{\alpha})=W_{\pm}f(-\Delta)W_{\pm}

for any Borel function ff on \mathbb{R}.

By using the intertwining property one can deduce the following Strichartz estimate [22, Corollary 1.5]:

(A.2) eitΔαPac(Δα)fLtr(,Lxq)fL2,\|e^{it\Delta_{\alpha}}P_{ac}(-\Delta_{\alpha})f\|_{L_{t}^{r}(\mathbb{R},L_{x}^{q})}\lesssim\|f\|_{L^{2}},

where (r,q)(r,q) is an admissible Strichartz pair, i.e.

(A.3) 1<r,1<q<,1r+1q=12.1<r\leq\infty,\quad 1<q<\infty,\quad\frac{1}{r}+\frac{1}{q}=\frac{1}{2}.

Since the orthogonal projection on Lac2(Δα)L_{\mathrm{ac}}^{2}(-\Delta_{\alpha}) is given by

Pac(Δα)f=f(f,χα)L2χα,P_{\mathrm{ac}}(-\Delta_{\alpha})f=f-(f,\chi_{\alpha})_{L^{2}}\chi_{\alpha},

we see that

eitΔαf=eitΔαPac(Δα)f+(f,χα)L2eiteαχαe^{it\Delta_{\alpha}}f=e^{it\Delta_{\alpha}}P_{ac}(-\Delta_{\alpha})f+(f,\chi_{\alpha})_{L^{2}}e^{ite_{\alpha}}\chi_{\alpha}

for all fL2(2)f\in L^{2}(\mathbb{R}^{2}). So the property (2.1) guarantees that we have the following (local in time) Strichartz estimate: there exists a constant C>0C>0 such that for any T(0,1]T\in(0,1] we have

(A.4) eitΔαfLtr([0,T],Lxq)CfL2\|e^{it\Delta_{\alpha}}f\|_{L_{t}^{r}([0,T],L_{x}^{q})}\leq C\|f\|_{L^{2}}

for all fL2(2)f\in L^{2}(\mathbb{R}^{2}). By using TTTT^{*} argument and Christ–Kiselev lemma we arrive at the following Strichartz estimate: there exists a constant C>0C>0 so that for any T(0,1]T\in(0,1] we have

(A.5) 0tei(ts)ΔαF(s)𝑑sLtr1([0,T],Lxq1)CFLr2([0,T],Lq2)\biggl{\|}\int_{0}^{t}e^{i(t-s)\Delta_{\alpha}}F(s)\,ds\biggr{\|}_{L_{t}^{r_{1}}([0,T],L_{x}^{q_{1}})}\leq C\|F\|_{L^{r_{2}^{\prime}}([0,T],L^{q_{2}^{\prime}})}

for any FLr2([0,T],Lp2(2)).F\in L^{r_{2}^{\prime}}([0,T],L^{p_{2}^{\prime}}(\mathbb{R}^{2})). Here and below (r1,q1)(r_{1},q_{1}) and (r2,q2)(r_{2},q_{2}) are admissible Strichartz pairs, i.e. 1rj+1qj=12\frac{1}{r_{j}}+\frac{1}{q_{j}}=\frac{1}{2}, qj[2,)q_{j}\in[2,\infty), for j=1,2j=1,2.

Remark A.1.

The L1L^{1}-LL^{\infty} dispersive estimates cannot hold, in fact even for a smooth initial data ff the evolution eitΔαfe^{it\Delta_{\alpha}}f exhibits, for almost every time t0t\neq 0, a non-trivial singular component proportional to GλL(2)G_{\lambda}\not\in L^{\infty}(\mathbb{R}^{2}).

Appendix B Local Well-posedness in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2})

In this section, we establish the local well-posedness in the energy space in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}). To this aim, we apply the abstract theory of Okazawa, Suzuki, and Yokota [52] to construct a weak solution to (1.1) with initial data u0Hα1(2)u_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}). Then we establish the uniqueness of the solution by using the Strichartz estimate obtained by [22].

First, we construct a weak solution to (1.1) by using [52, Theorem 2.2].

Lemma B.1.

For any M>0M>0 there exists TM>0T_{M}>0 such that the following is true. For u0Hα1(2)u_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}) with u0Hα1M\|u_{0}\|_{H_{\alpha}^{1}}\leq M, there exists a local weak solution

uCw([TM,TM],Hα1(2))W1,(TM,TM;Hα1(2))u\in C_{\mathrm{w}}([-T_{M},T_{M}],H_{\alpha}^{1}(\mathbb{R}^{2}))\cap W^{1,\infty}(-T_{M},T_{M};H_{\alpha}^{-1}(\mathbb{R}^{2}))

of (1.1) satisfying

u(t)L2=u0L2,E(u(t))E(u0)\displaystyle\|u(t)\|_{L^{2}}=\|u_{0}\|_{L^{2}},\quad E(u(t))\leq E(u_{0})

for all t[TM,TM]t\in[-T_{M},T_{M}].

Proof.

We will apply [52, Theorem 2.2] as

S\displaystyle S =Δαeα,\displaystyle=-\Delta_{\alpha}-e_{\alpha},
X\displaystyle X =L2(2),XS=Hα1(2),XS=Hα1(2)\displaystyle=L^{2}(\mathbb{R}^{2}),\quad X_{S}=H_{\alpha}^{1}(\mathbb{R}^{2}),\quad X_{S}^{*}=H_{\alpha}^{-1}(\mathbb{R}^{2})
g(v)\displaystyle g(v) =eαv|v|p1v.\displaystyle=e_{\alpha}v-|v|^{p-1}v.

Under this setting, we see that SS is a nonnegative self-adjoint operator in L2(2)L^{2}(\mathbb{R}^{2}) and that Hα1(2)=D((1+S)1/2)H_{\alpha}^{1}(\mathbb{R}^{2})=D((1+S)^{1/2}). After that, we only have to verify [52, (G1)(G5)] given as follows.

(G1): there exists GC1(XS,)G\in C^{1}(X_{S},\mathbb{R}) such that G=gG^{\prime}=g.

(G2): for all M>0M>0 there exists C(M)>0C(M)>0 such that

g(u)g(v)XSC(M)uvXSu,vXS with uXSvXSM.\|g(u)-g(v)\|_{X_{S}^{*}}\leq C(M)\|u-v\|_{X_{S}}\quad\forall u,v\in X_{S}\text{ with $\|u\|_{X_{S}}$,\,$\|v\|_{X_{S}}\leq M$}.

(G3): for all M,δ>0M,\delta>0 there exists Cδ(M)>0C_{\delta}(M)>0 such that

|G(u)G(v)|δ+C(M)uvXu,vXS with uXSvXSM.|G(u)-G(v)|\leq\delta+C(M)\|u-v\|_{X}\quad\forall u,v\in X_{S}\text{ with $\|u\|_{X_{S}}$,\,$\|v\|_{X_{S}}\leq M$}.

(G4):

g(u),iuXS,XS=0uXS.\langle g(u),iu\rangle_{X_{S}^{*},X_{S}}=0\quad\forall u\in X_{S}.

(G5): given a bounded open interval II\subset\mathbb{R}, let (wn)n(w_{n})_{n\in\mathbb{N}} by any bounded sequence in L(I,XS)L^{\infty}(I,X_{S}) such that

{wn(t)w(t)(n)weakly in XS a.a. tI,g(wn)f(n)weakly in L(I,XS).\left\{\begin{aligned} w_{n}(t)&\to w(t)\ (n\to\infty)&&\text{weakly in $X_{S}$ a.a. $t\in I$},\\ g(w_{n})&\to f\ (n\to\infty)&&\text{weakly${}^{*}$ in $L^{\infty}(I,X_{S}^{*})$}.\end{aligned}\right.

Then

If(t),iw(t)XS,XS𝑑t=limnIg(wn(t)),iwn(t)XS,XS𝑑t.\int_{I}\langle f(t),iw(t)\rangle_{X_{S}^{*},X_{S}}\,dt=\lim_{n\to\infty}\int_{I}\langle g(w_{n}(t)),iw_{n}(t)\rangle_{X_{S}^{*},X_{S}}\,dt.

Now we check (G1)(G5).

The conditions (G1) are easily verified as

G(v)=1p+1vLp+1p+1,vHα1(2)G(v)=\frac{1}{p+1}\|v\|_{L^{p+1}}^{p+1},\quad v\in H_{\alpha}^{1}(\mathbb{R}^{2})

by standard inequalities and the embedding Lq(2)Hα1(2)L^{q}(\mathbb{R}^{2})\subset H_{\alpha}^{1}(\mathbb{R}^{2}) obtained in (2.2). Similarly, the condition (G2) also can be verified.

The condition (G3) follows from the following estimate:

|G(u)G(v)|\displaystyle|G(u)-G(v)| (|u|p+|v|p)|uv|𝑑x\displaystyle\lesssim\int(|u|^{p}+|v|^{p})|u-v|\,dx
(uL2pp+vL2pp)uvL2\displaystyle\lesssim(\|u\|_{L^{2p}}^{p}+\|v\|_{L^{2p}}^{p})\|u-v\|_{L^{2}}
MpuvL2\displaystyle\lesssim M^{p}\|u-v\|_{L^{2}}

for u,vHα1(2)u,v\in H_{\alpha}^{1}(\mathbb{R}^{2}) with uHα1,vHα1M\|u\|_{H_{\alpha}^{1}},\,\|v\|_{H_{\alpha}^{1}}\leq M.

The conditions (G4) is clear from the definition of gg.

Finally, we will check the conditions (G5). From [52, Lemma 5.3], it is enough to show that if (un)n(u_{n})_{n\in\mathbb{N}} is a sequence Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) satisfies

{unu(n)weakly in Hα1(2),g(un)f(n)weakly in Hα1(2),\left\{\begin{aligned} u_{n}&\to u\ (n\to\infty)&&\text{weakly in $H_{\alpha}^{1}(\mathbb{R}^{2})$},\\ g(u_{n})&\to f\ (n\to\infty)&&\text{weakly in $H_{\alpha}^{-1}(\mathbb{R}^{2})$},\end{aligned}\right.

then f=g(u)f=g(u).

We follow the argument in [56, Proof of Theorem 1.1]. Let φCc(2)\varphi\in C_{c}^{\infty}(\mathbb{R}^{2}). Then from the weak convergence of (un)n(u_{n})_{n\in\mathbb{N}} in Hα1(2)H_{\alpha}^{1}(\mathbb{R}^{2}) and the compactness Llocp+1(2)Hloc1(2)L_{\mathrm{loc}}^{p+1}(\mathbb{R}^{2})\hookrightarrow H_{\mathrm{loc}}^{1}(\mathbb{R}^{2}) we see that

unuin Llocp+1(2).u_{n}\to u\quad\text{in }L_{\mathrm{loc}}^{p+1}(\mathbb{R}^{2}).

Thus,

|g(un)g(u),φ|\displaystyle|\langle g(u_{n})-g(u),\varphi\rangle| φLp+1(unLp+1p1+uLp+1p1)unuLp+1(suppφ)0\displaystyle\lesssim\|\varphi\|_{L^{p+1}}(\|u_{n}\|_{L^{p+1}}^{p-1}+\|u\|_{L^{p+1}}^{p-1})\|u_{n}-u\|_{L^{p+1}(\operatorname{supp}\varphi)}\to 0

as nn\to\infty. This means that g(un)g(u)g(u_{n})\to g(u) in 𝒟(2)\mathcal{D}^{\prime}(\mathbb{R}^{2}). On the other hand, g(un)fg(u_{n})\to f in Hα1(2)H_{\alpha}^{-1}(\mathbb{R}^{2}) and hence in 𝒟(2)\mathcal{D}^{\prime}(\mathbb{R}^{2}). Therefore we obtain f=g(u)f=g(u). Thus, (G5) is verified.

We have just finished the verification of (G1)(G5). Therefore, [52, Theorem 2.2] implies the conclusion. ∎

Lemma B.2.

Let u0Hα1(2)u_{0}\in H_{\alpha}^{1}(\mathbb{R}^{2}). If u1,u2L(T,T;Hα1(2))u_{1},u_{2}\in L^{\infty}(-T,T;H_{\alpha}^{1}(\mathbb{R}^{2})) are two weak solutions of (1.1) with u1(0)=u2(0)=u0u_{1}(0)=u_{2}(0)=u_{0}, then u1=u2u_{1}=u_{2}.

Proof.

Without loss of generality we can assume T(0,1]T\in(0,1]. Let

r=r(p):=2(p+1)p+1.r=r(p)\mathrel{\mathop{:}}=\frac{2(p+1)}{p+1}.

Then (r,p+1)(r,p+1) is a admissible pair. By the Strichartz estimate (A.5) for

uj(t)=eitΔαu0i0tei(tτ)Δα|uj(τ)|p1uj(τ)𝑑τ,u_{j}(t)=e^{it\Delta_{\alpha}}u_{0}-i\int_{0}^{t}e^{i(t-\tau)\Delta_{\alpha}}|u_{j}(\tau)|^{p-1}u_{j}(\tau)\,d\tau,

we see that

u1u2Lr([0,T],Lxp+1)|u1|p1u1|u2|p1u2L1([0,T],Lx2).\|u_{1}-u_{2}\|_{L^{r}([0,T],L^{p+1}_{x})}\lesssim\||u_{1}|^{p-1}u_{1}-|u_{2}|^{p-1}u_{2}\|_{L^{1}([0,T],L^{2}_{x})}.

From

|u1|p1u1|u2|p1u2Lx2(u1Lx2(p+1)p1+u2Lx2(p+1)p1)u1u2Lxp+1\||u_{1}|^{p-1}u_{1}-|u_{2}|^{p-1}u_{2}\|_{L^{2}_{x}}\lesssim(\|u_{1}\|_{L^{2(p+1)}_{x}}^{p-1}+\|u_{2}\|_{L^{2(p+1)}_{x}}^{p-1})\|u_{1}-u_{2}\|_{L^{p+1}_{x}}

and the Sobolev inequality (2.2) we deduce

ujLx2(p+1)ujHα1(2)1.\|u_{j}\|_{L^{2(p+1)}_{x}}\lesssim\|u_{j}\|_{H^{1}_{\alpha}(\mathbb{R}^{2})}\lesssim 1.

Hence

u1u2L4([0,T],Lx4)\displaystyle\|u_{1}-u_{2}\|_{L^{4}([0,T],L^{4}_{x})} C0Tu1(τ)u2(τ)Lxp+1𝑑τ\displaystyle\leq C\int_{0}^{T}\|u_{1}(\tau)-u_{2}(\tau)\|_{L^{p+1}_{x}}d\tau
CT(p+1)/2pu1u2Lr([0,T],Lxp+1)\displaystyle\leq CT^{(p+1)/2p}\|u_{1}-u_{2}\|_{L^{r}([0,T],L^{p+1}_{x})}

so with TT sufficiently small so that CT(p+1)/(2p)<1CT^{(p+1)/(2p)}<1 we conclude that u1(t)=u2(t)u_{1}(t)=u_{2}(t) for t[0,T]t\in[0,T]. In a similar way, we obtain u1(t)=u2(t)u_{1}(t)=u_{2}(t) for t[T,0]t\in[-T,0]. ∎

Proof of Proposition 1.1.

The assertion follows from Lemmas B.1, B.2, and [52, Theorem 2.3]. ∎

Acknowledgements

NF was supported by JSPS KAKENHI Grant Number JP20K14349. VG was partially supported by Project 2017 “Problemi stazionari e di evoluzione nelle equazioni di campo nonlineari” of INDAM, GNAMPA - Gruppo Nazionale per l’Analisi Matematica, la Probabilita e le loro Applicazioni, by Institute of Mathematics and Informatics, Bulgarian Academy of Sciences, by Top Global University Project, Waseda University and the Project PRA 2018 49 of University of Pisa. MI is supported by JST CREST Grant Number JPMJCR1913, Japan and Grant-in-Aid for Young Scientists Research (No.19K14581), Japan Society for the Promotion of Science.

References

  • [1] R. Adami, R. Carlone, M. Correggi, and L. Tentarelli, Blow-up for the pointwise NLS in dimension two: absence of critical power, J. Differential Equations 269 (2020), 1–37.
  • [2] R. Adami, R. Carlone, M. Correggi, and L. Tentarelli, Stability of the standing waves of the concentrated NLSE in dimension two, Math. Eng. 3 (2021), Paper No. 011, 15.
  • [3] R. Adami, R. Fukuizumi, and J. Holmer, Scattering for the L2L^{2} supercritical point NLS, Trans. Amer. Math. Soc. 374 (2021), 35–60.
  • [4] R. Adami and D. Noja, Existence of dynamics for a 1D NLS equation perturbed with a generalized point defect, J. Phys. A 42 (2009), 495302, 19.
  • [5] R. Adami, D. Noja, and C. Ortoleva, Orbital and asymptotic stability for standing waves of a nonlinear Schrödinger equation with concentrated nonlinearity in dimension three, J. Math. Phys. 54 (2013), 013501, 33.
  • [6] S. Albeverio, F. Gesztesy, and R. H. e. Krohn, The low energy expansion in nonrelativistic scattering theory, Ann. Inst. H. Poincaré Sect. A (N.S.) 37 (1982), 1–28.
  • [7] S. Albeverio, F. Gesztesy, R. H. e. Krohn, and H. Holden, Point interactions in two dimensions: basic properties, approximations and applications to solid state physics, J. Reine Angew. Math. 380 (1987), 87–107.
  • [8] S. Albeverio, F. Gesztesy, R. H. e. Krohn, and H. Holden, Solvable models in quantum mechanics, second ed., AMS Chelsea Publishing, Providence, RI, 2005, With an appendix by Pavel Exner.
  • [9] S. Albeverio, F. Gesztesy, R. H. e. Krohn, and W. Kirsch, On point interactions in one dimension, J. Operator Theory 12 (1984), 101–126.
  • [10] S. Albeverio and R. Høegh-Krohn, Point interactions as limits of short range interactions, J. Operator Theory 6 (1981), 313–339.
  • [11] F. J. Almgren, Jr. and E. H. Lieb, Symmetric decreasing rearrangement is sometimes continuous, J. Amer. Math. Soc. 2 (1989), 683–773.
  • [12] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Rational Mech. Anal. 82 (1983), 313–345.
  • [13] H. Berestycki and T. Cazenave, Instabilité des états stationnaires dans les équations de Schrödinger et de Klein-Gordon non linéaires, C. R. Acad. Sci. Paris Sér. I Math. 293 (1981), 489–492.
  • [14] F. A. Berezin and L. D. Faddeev, Remark on the Schrödinger equation with singular potential, Dokl. Akad. Nauk SSSR 137 (1961), 1011–1014.
  • [15] H. Brézis and E. Lieb, A relation between pointwise convergence of functions and convergence of functionals, Proc. Amer. Math. Soc. 88 (1983), 486–490.
  • [16] J. E. Brothers and W. P. Ziemer, Minimal rearrangements of Sobolev functions, J. Reine Angew. Math. 384 (1988), 153–179.
  • [17] V. S. Buslaev, A. I. Komech, E. A. Kopylova, and D. Stuart, On asymptotic stability of solitary waves in Schrödinger equation coupled to nonlinear oscillator, Comm. Partial Differential Equations 33 (2008), 669–705.
  • [18] C. Cacciapuoti, D. Finco, and D. Noja, Well posedness of the nonlinear schrödinger equation with isolated singularities, arXiv:2012.12391.
  • [19] T. Cazenave and P.-L. Lions, Orbital stability of standing waves for some nonlinear Schrödinger equations, Comm. Math. Phys. 85 (1982), 549–561.
  • [20] T. Cazenave, Semilinear Schrödinger equations, Courant Lecture Notes in Mathematics, vol. 10, New York University, Courant Institute of Mathematical Sciences, New York; American Mathematical Society, Providence, RI, 2003.
  • [21] T. Cazenave, An introduction to semilinear elliptic equations, second ed., Editora do IM-UFRJ, Rio de Janeiro, 2018.
  • [22] H. D. Cornean, A. Michelangeli, and K. Yajima, Two-dimensional Schrödinger operators with point interactions: threshold expansions, zero modes and LpL^{p}-boundedness of wave operators, Rev. Math. Phys. 31 (2019), 1950012, 32.
  • [23] S. Cuccagna and M. Maeda, On stability of small solitons of the 1-D NLS with a trapping delta potential, SIAM J. Math. Anal. 51 (2019), 4311–4331.
  • [24] P. D’Ancona, V. Pierfelice, and A. Teta, Dispersive estimate for the Schrödinger equation with point interactions, Math. Methods Appl. Sci. 29 (2006), 309–323.
  • [25] G. Dell’Antonio, A. Michelangeli, R. Scandone, and K. Yajima, LpL^{p}-boundedness of wave operators for the three-dimensional multi-centre point interaction, Ann. Henri Poincaré 19 (2018), 283–322.
  • [26] NIST Digital Library of Mathematical Functions, http://dlmf.nist.gov/, Release 1.1.2 of 2021-06-15, F. W. J. Olver, A. B. Olde Daalhuis, D. W. Lozier, B. I. Schneider, R. F. Boisvert, C. W. Clark, B. R. Miller, B. V. Saunders, H. S. Cohl, and M. A. McClain, eds.
  • [27] V. Duchêne, J. L. Marzuola, and M. I. Weinstein, Wave operator bounds for one-dimensional Schrödinger operators with singular potentials and applications, J. Math. Phys. 52 (2011), 013505, 17.
  • [28] N. Fukaya, Stability of standing waves for L2{L}^{2}-critical nonlinear Schrödinger equations with attractive inverse-power potential, to appear in RIMS Kôkyûroku Bessatsu.
  • [29] N. Fukaya and M. Ohta, Strong instability of standing waves for nonlinear Schrödinger equations with attractive inverse power potential, Osaka J. Math. 56 (2019), 713–726.
  • [30] R. Fukuizumi, Stability of standing waves for nonlinear Schrödinger equations with critical power nonlinearity and potentials, Adv. Differential Equations 10 (2005), 259–276.
  • [31] R. Fukuizumi and L. Jeanjean, Stability of standing waves for a nonlinear Schrödinger equation with a repulsive Dirac delta potential, Discrete Contin. Dyn. Syst. 21 (2008), 121–136.
  • [32] R. Fukuizumi and M. Ohta, Instability of standing waves for nonlinear Schrödinger equations with potentials, Differential Integral Equations 16 (2003), 691–706.
  • [33] R. Fukuizumi and M. Ohta, Stability of standing waves for nonlinear Schrödinger equations with potentials, Differential Integral Equations 16 (2003), 111–128.
  • [34] R. Fukuizumi, M. Ohta, and T. Ozawa, Nonlinear Schrödinger equation with a point defect, Ann. Inst. H. Poincaré Anal. Non Linéaire 25 (2008), 837–845.
  • [35] V. Georgiev, A. Michelangeli, and R. Scandone, On fractional powers of singular perturbations of the Laplacian, J. Funct. Anal. 275 (2018), 1551–1602.
  • [36] R. H. Goodman, P. J. Holmes, and M. I. Weinstein, Strong NLS soliton-defect interactions, Phys. D 192 (2004), 215–248.
  • [37] M. Grillakis, J. Shatah, and W. Strauss, Stability theory of solitary waves in the presence of symmetry. I, J. Funct. Anal. 74 (1987), 160–197.
  • [38] S. Gustafson, K. Nakanishi, and T.-P. Tsai, Asymptotic stability and completeness in the energy space for nonlinear Schrödinger equations with small solitary waves, Int. Math. Res. Not. (2004), 3559–3584.
  • [39] J. Holmer, J. Marzuola, and M. Zworski, Fast soliton scattering by delta impurities, Comm. Math. Phys. 274 (2007), 187–216.
  • [40] M. Ikeda and T. Inui, Global dynamics below the standing waves for the focusing semilinear Schrödinger equation with a repulsive Dirac delta potential, Anal. PDE 10 (2017), 481–512.
  • [41] M. Kaminaga and M. Ohta, Stability of standing waves for nonlinear Schrödinger equation with attractive delta potential and repulsive nonlinearity, Saitama Math. J. 26 (2009), 39–48 (2010).
  • [42] A. Komech, E. Kopylova, and D. Stuart, On asymptotic stability of solitons in a nonlinear Schrödinger equation, Commun. Pure Appl. Anal. 11 (2012), 1063–1079.
  • [43] M. K. Kwong, Uniqueness of positive solutions of Δuu+up=0\Delta u-u+u^{p}=0 in 𝐑n{\bf R}^{n}, Arch. Rational Mech. Anal. 105 (1989), 243–266.
  • [44] S. Le Coz, Standing waves in nonlinear Schrödinger equations, Analytical and numerical aspects of partial differential equations, Walter de Gruyter, Berlin, 2009, pp. 151–192.
  • [45] S. Le Coz, R. Fukuizumi, G. Fibich, B. Ksherim, and Y. Sivan, Instability of bound states of a nonlinear Schrödinger equation with a Dirac potential, Phys. D 237 (2008), 1103–1128.
  • [46] S. Masaki, J. Murphy, and J.-i. Segata, Asymptotic stability of solitary waves for the 1d NLS with an attractive delta potential, arXiv:2008.11645, August 2020.
  • [47] S. Masaki, J. Murphy, and J.-i. Segata, Stability of small solitary waves for the one-dimensional NLS with an attractive delta potential, Anal. PDE 13 (2020), 1099–1128.
  • [48] A. Michelangeli, A. Ottolini, and R. Scandone, Fractional powers and singular perturbations of quantum differential Hamiltonians, J. Math. Phys. 59 (2018), 072106, 27.
  • [49] A. Michelangeli and R. Scandone, Point-like perturbed fractional Laplacians through shrinking potentials of finite range, Complex Anal. Oper. Theory 13 (2019), 3717–3752.
  • [50] M. Ohta, Instability of solitary waves for nonlinear Schrödinger equations of derivative type, SUT J. Math. 50 (2014), 399–415.
  • [51] M. Ohta and T. Yamaguchi, Strong instability of standing waves for nonlinear Schrödinger equations with a delta potential, Harmonic analysis and nonlinear partial differential equations, RIMS Kôkyûroku Bessatsu, B56, Res. Inst. Math. Sci. (RIMS), Kyoto, 2016, pp. 79–92.
  • [52] N. Okazawa, T. Suzuki, and T. Yokota, Energy methods for abstract nonlinear Schrödinger equations, Evol. Equ. Control Theory 1 (2012), 337–354.
  • [53] H. A. Rose and M. I. Weinstein, On the bound states of the nonlinear Schrödinger equation with a linear potential, Phys. D 30 (1988), 207–218.
  • [54] J. Shatah, Stable standing waves of nonlinear Klein-Gordon equations, Comm. Math. Phys. 91 (1983), 313–327.
  • [55] J. Shatah and W. Strauss, Instability of nonlinear bound states, Comm. Math. Phys. 100 (1985), 173–190.
  • [56] T. Suzuki, Nonlinear Schrödinger equations with inverse-square potentials in two dimensional space, Discrete Contin. Dyn. Syst. (2015), 1019–1024.
  • [57] M. I. Weinstein, Nonlinear Schrödinger equations and sharp interpolation estimates, Comm. Math. Phys. 87 (1982/83), 567–576.
  • [58] M. I. Weinstein, Lyapunov stability of ground states of nonlinear dispersive evolution equations, Comm. Pure Appl. Math. 39 (1986), 51–67.
  • [59] K. Yajima, LpL^{p}-boundedness of wave operators for 2D Schrödinger operators with point interactions, Ann. Henri Poincaré 22 (2021), 2065–2101.