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

Standing wave solutions of a quasilinear Schrödinger equation in the small frequency limit

François Genoud Cours de Mathématiques Spéciales, Ecole Polytechnique Fédérale de Lausanne,
EPFL Station 4, 1015 Lausanne, Switzerland
[email protected]
 and  Simona Rota Nodari Laboratoire J.A. Dieudonné, Université Côte d’Azur, CNRS UMR 7351,
Parc Valrose, 28, avenue Valrose, 06108 Nice Cedex 2, France
[email protected]
Abstract.

This article is concerned with the quasilinear Schrödinger equation

Δuωu+|u|p1u+δΔ(|u|2)u=0,\Delta u-\omega u+|u|^{p-1}u+\delta\Delta(|u|^{2})u=0,

where δ>0\delta>0, N=2N=2 and p>1p>1 or N3N\geqslant 3 and 1<p<3N+2N21<p<\frac{3N+2}{N-2}. After proving uniqueness and non-degeneracy of the positive solution uωu_{\omega} for all ω>0\omega>0, our main results establish the asymptotic behavior of uωu_{\omega} in the limit ω0+\omega\to 0^{+}. Three different regimes arise, termed ‘subcritical’, ‘critical’ and ‘supercritical’, corresponding respectively (when N3N\geqslant 3) to 1<p<N+2N21<p<\frac{N+2}{N-2}, p=N+2N2p=\frac{N+2}{N-2} and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. In each case a limit equation is exhibited which governs, in a suitable scaling, the behavior of uωu_{\omega} in the limit ω0+\omega\to 0^{+}. The critical case is the most challenging, technically speaking. In this case, the limit equation is the famous Lane-Emden-Fowler equation. A substantial part of our efforts is dedicated to the study of the function ωM(ω)=Nuω2\omega\mapsto M(\omega)=\int_{\mathbb{R}^{N}}u_{\omega}^{2}. We find that, for small ω>0\omega>0, M(ω)M(\omega) is increasing if 1<p1+4N1<p\leqslant 1+\frac{4}{N} and decreasing if 1+4N<pN+2N21+\frac{4}{N}<p\leqslant\frac{N+2}{N-2}. In the supercritical case, the monotonicity of M(ω)M(\omega) depends on the dimension, except in the regime p3+4Np\geqslant 3+\frac{4}{N}, where M(ω)M(\omega) is always decreasing close to ω=0\omega=0. The crucial role played by M(ω)M(\omega) for the orbital stability of the standing wave eiωtuωe^{i\omega t}u_{\omega}, and for the uniqueness of normalized ground states, is discussed in the introduction.

1. Introduction

This work is concerned with the quasilinear Schrödinger equation

itφ+Δφ+|φ|p1φ+δΔ(|φ|2)φ=0,i\partial_{t}\varphi+\Delta\varphi+|\varphi|^{p-1}\varphi+\delta\Delta(|\varphi|^{2})\varphi=0, (1.1)

and the associated stationary equation satisfied by standing waves φ(t,x)=eiωtu(x)\varphi(t,x)=e^{i\omega t}u(x),

Δuωu+|u|p1u+δΔ(|u|2)u=0.\Delta u-\omega u+|u|^{p-1}u+\delta\Delta(|u|^{2})u=0. (Eω\mathrm{E}_{\omega})

We consider the problem in N\mathbb{R}^{N} with N3N\geqslant 3, ω>0\omega>0 and δ>0\delta>0.

This equation belongs to a family of quasilinear Schrödinger equations of the form

itφ+Δφ+f(|φ|2)φ+φ(|φ|2)Δ(|φ|2)=0,i\partial_{t}\varphi+\Delta\varphi+f(|\varphi|^{2})\varphi+\varphi\ell^{\prime}(|\varphi|^{2})\Delta\ell(|\varphi|^{2})=0, (1.2)

which appear in several physical situations (see [9] and references therein). Here ff and \ell are given smooth functions. When f(s)=s(p1)/2f(s)=s^{(p-1)/2} and (s)=δs\ell(s)=\sqrt{\delta}s, we obtain (Eω\mathrm{E}_{\omega}) which is relevant in various problems in plasma physics and nonlinear optics (see [43, 30, 20]). The local and global well-posedness of the Cauchy problem associated to (1.2) have been studied by Poppenberg in [42] for smooth initial data (belonging to the space HH^{\infty}). In [12, 13], the authors improved the local well-posedness result for initial data in Hs+2(N)H^{s+2}(\mathbb{R}^{N}) for s=2E(N/2)+2s=2E(N/2)+2 with E(N/2)E(N/2) the integer part of N/2N/2. More precisely, in [12], equation (1.2) is solved locally for smooth nonlinearities \ell and ff such that there exists a positive constant CC_{\ell} with 14σ2(σ)>C2(σ)1-4\sigma\ell^{\prime 2}(\sigma)>C_{\ell}\ell^{\prime 2}(\sigma) for any σ+\sigma\in\mathbb{R}_{+}, while [13] deals with the case (σ)=σ\ell(\sigma)=\sigma and fCs+2(+)f\in C^{s+2}(\mathbb{R}_{+}).

The parameter δ>0\delta>0 is a coupling constant relevant to describe the strength of the quasilinear interaction in the associated physical models. In several works about (Eω\mathrm{E}_{\omega}), it is simply set equal to one. In the series of papers [1, 5, 4, 2], the asymptotic behavior of solutions of (Eω\mathrm{E}_{\omega}) is studied in the limit δ0+\delta\to 0^{+}. It is worth noting here that a simple scaling transforms (Eω\mathrm{E}_{\omega}) with ω=1\omega=1 into

Δuδp12u+|u|p1u+Δ(|u|2)u=0.\Delta u-\delta^{\frac{p-1}{2}}u+|u|^{p-1}u+\Delta(|u|^{2})u=0.

Some of the results obtained in [1, 5, 4, 2] can thus be recovered from the present approach, which we believe to be much more straightforward. Furthermore, as explained below, our main focus in this work is the asymptotic behavior of the function M(ω)M(\omega) defined in (1.11), which is not addressed in detail in [1, 5, 4, 2].

Concerning the existence of solutions to (Eω\mathrm{E}_{\omega}), the following result has been proved by Colin, Jeanjean and Squassina [13].

Theorem 1.1.

Let N=2N=2 and p>1p>1, or N3N\geqslant 3 and 1<p<3N+2N21<p<\frac{3N+2}{N-2}. For any ω>0\omega>0, there exists a solution uωH1(N)C2(N)u_{\omega}\in H^{1}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}) of (Eω\mathrm{E}_{\omega}). Furthermore, uωu_{\omega} is positive, spherically symmetric, radially decreasing and decays exponentially at infinity, together with its derivatives up to second order.

Remark 1.

As pointed out in [13] (see also [47]), a bootstrap argument allows one to show that uωu_{\omega} belongs to s>0Hs(N)\cap_{s>0}H^{s}(\mathbb{R}^{N}) and, in particular, is of class CC^{\infty}.

The main idea used to prove Theorem 1.1 is to remark that finding a solution uu of the equation (Eω\mathrm{E}_{\omega}) is equivalent to finding a solution vv of the semilinear equation

Δv=11+2δr(v)2(|r(v)|p1r(v)ωr(v)),-\Delta v=\frac{1}{\sqrt{1+2\delta r(v)^{2}}}\left(|r(v)|^{p-1}r(v)-\omega r(v)\right), (1.3)

with rr a suitably chosen function (see Section 2 for more details). Variational arguments are then used to deduce an existence result for (1.3). Finally, a solution to (Eω\mathrm{E}_{\omega}) is obtained by setting u=r(v)u=r(v).

Note that that for N3N\geqslant 3, p<3N+2N2p<\frac{3N+2}{N-2} is a necessary condition for the existence of nontrivial solutions uH1(N)L(N)u\in H^{1}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N}). This is a consequence of the following integral identities, which will play an important role in our analysis. They can be proved as sketched in [13]. Let

X:={uH1(N,):N|u|2||u||2dx<}X:=\Big{\{}u\in H^{1}(\mathbb{R}^{N},\mathbb{C}):\int_{\mathbb{R}^{N}}|u|^{2}|\nabla|u||^{2}\,\mathrm{d}x<\infty\Big{\}} (1.4)

and

X~:={uH˙1(N,):N|u|2||u||2dx<}.\widetilde{X}:=\Big{\{}u\in\dot{H}^{1}(\mathbb{R}^{N},\mathbb{C}):\int_{\mathbb{R}^{N}}|u|^{2}|\nabla|u||^{2}\,\mathrm{d}x<\infty\Big{\}}. (1.5)
Proposition 1.1.

Let N3N\geqslant 3. Any solution uXu\in X of (Eω\mathrm{E}_{\omega}) satisfies

12N|u|2dx+22δN|u|2||u||2dx=1p+1N|u|p+1dxω2N|u|2dx\frac{1}{2^{*}}\int_{\mathbb{R}^{N}}|\nabla u|^{2}\,\mathrm{d}x+\frac{2}{2^{*}}\delta\int_{\mathbb{R}^{N}}|u|^{2}|\nabla|u||^{2}\,\mathrm{d}x=\frac{1}{p+1}\int_{\mathbb{R}^{N}}|u|^{p+1}\,\mathrm{d}x-\frac{\omega}{2}\int_{\mathbb{R}^{N}}|u|^{2}\,\mathrm{d}x (1.6)

and

12N|u|2dx+2δN|u|2||u||2dx=12N|u|p+1dxω2N|u|2dx.\frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla u|^{2}\,\mathrm{d}x+2\delta\int_{\mathbb{R}^{N}}|u|^{2}|\nabla|u||^{2}\,\mathrm{d}x=\frac{1}{2}\int_{\mathbb{R}^{N}}|u|^{p+1}\,\mathrm{d}x-\frac{\omega}{2}\int_{\mathbb{R}^{N}}|u|^{2}\,\mathrm{d}x. (1.7)

The above identities still hold true for a solution uX~u\in\widetilde{X} of (Eω\mathrm{E}_{\omega}), in case ω=0\omega=0.

As already noted above, it follows from Proposition 1.1 that, in dimensions N3N\geqslant 3, p<3N+2N2p<\frac{3N+2}{N-2} is a necessary condition for the existence of nontrivial solutions uXu\in X.

The main goal of this paper is to study the qualitative properties of a branch of solutions of (Eω\mathrm{E}_{\omega}), parametrized by ω\omega. In particular, we want to investigate the uniqueness of positive solutions and their non-degeneracy, i.e. the fact that the kernel of the linearized operators is trivial, modulo phase and space translations.

The study of uniqueness of positive solutions to semilinear Schrödinger equations of the form

Δu+g(u)=0,\Delta u+g(u)=0,

with gg a well-behaved nonlinearity, has a very long history, see e.g. [11, 29, 10, 24, 37, 38, 41, 44, 48, 32]. Quasilinear equations have attracted less attention, and fewer results on the uniqueness or non-degeneracy of positive solutions of equations of this type are available, see e.g. [47, 19, 3, 31].

A first important result that will be proved here is the uniqueness and non-degeneracy of the positive solution uωu_{\omega} to (Eω\mathrm{E}_{\omega}), for any ω>0\omega>0.

Theorem 1.2 (Uniqueness and non-degeneracy).

Let N=2N=2 and p>1p>1, or N3N\geqslant 3 and 1<p<3N+2N21<p<\frac{3N+2}{N-2}. For any ω>0\omega>0, the positive solution uωu_{\omega} to the nonlinear equation (Eω\mathrm{E}_{\omega}) is unique, modulo space translation. Moreover, it is non-degenerate:

{ker(L+)=span{x1uω,,xNuω},ker(L)=span{uω},\left\{\begin{aligned} &\ker(L_{+})=\mathrm{span}\{\partial_{x_{1}}u_{\omega},\ldots,\partial_{x_{N}}u_{\omega}\},\\ &\ker(L_{-})=\mathrm{span}\{u_{\omega}\},\end{aligned}\right. (1.8)

where the linear operators L+L_{+} and LL_{-} are defined by

L+=(1+2δuω2)Δ4δuωuωδ(4uωΔuω+2|uω|2)puωp1+ω,\displaystyle L_{+}=-(1+2\delta u_{\omega}^{2})\Delta-4\delta u_{\omega}\nabla u_{\omega}\cdot\nabla-\delta(4u_{\omega}\Delta u_{\omega}+2|\nabla u_{\omega}|^{2})-pu_{\omega}^{p-1}+\omega, (1.9)
L=Δδ(2uωΔuω+2|uω|2)uωp1+ω.\displaystyle L_{-}=-\Delta-\delta(2u_{\omega}\Delta u_{\omega}+2|\nabla u_{\omega}|^{2})-u_{\omega}^{p-1}+\omega. (1.10)

As a consequence of the non-degeneracy of uωu_{\omega}, the following proposition will also be established.

Proposition 1.2.

Let ω>0\omega>0 and uωu_{\omega} be the unique positive solution of (Eω\mathrm{E}_{\omega}). Then the map ωuω\omega\mapsto u_{\omega} is of class C1((0,),H1(N))C^{1}((0,\infty),H^{1}(\mathbb{R}^{N})).

The non-degeneracy property is also crucial for the study of the mass

M(ω)=Nuω(x)2dxM(\omega)=\int_{\mathbb{R}^{N}}u_{\omega}(x)^{2}\,\mathrm{d}x (1.11)

of the unique positive solution uωu_{\omega}. An important motivation for studying this quantity is the central role played by the function ωM(ω)\omega\mapsto M(\omega) in the Grillakis-Shatah-Strauss theory of stability [52, 21, 22, 14, 15], which can be applied to the standing wave solutions eiωtuω(x)e^{i\omega t}u_{\omega}(x) of the time-dependent equation (1.1). In particular, the standing wave is expected to be orbitally stable when M(ω)>0M^{\prime}(\omega)>0 and unstable when M(ω)<0M^{\prime}(\omega)<0. Therefore, it is important to be able to identify the regions where MM is increasing, which correspond to stable solutions, and those where MM is decreasing, corresponding to unstable solutions. For the classical nonlinear Schrödinger equation (NLS) with a single-power nonlinearity, which corresponds to (Eω\mathrm{E}_{\omega}) with δ=0\delta=0, N=2N=2 and p>1p>1, or N3N\geqslant 3 and p<N+2N2p<\frac{N+2}{N-2}, the mass is an explicit function of ω\omega. Indeed, for δ=0\delta=0, the solution uωu_{\omega} for ω>0\omega>0 can be obtained from the unique positive solution to ΔQQ+Qp=0\Delta Q-Q+Q^{p}=0 by the simple scaling uω(x)=ω1p1Q(ωx)u_{\omega}(x)=\omega^{\frac{1}{p-1}}Q(\sqrt{\omega}x). This leads to

MNLS(ω)=ω4+NNp2(p1)NQ(x)2dx.M_{\mathrm{NLS}}(\omega)=\omega^{\frac{4+N-Np}{2(p-1)}}\int_{\mathbb{R}^{N}}Q(x)^{2}\,\,\mathrm{d}x.

However, when δ>0\delta>0, the presence of the quasilinear term prevents one from using a scaling argument and any hope of obtaining a simple expression for M(ω)M(\omega) vanishes. A similar situation occurs in the case of the double-power nonlinearity considered in [32]. Those two problems seem rather different at first sight, but they share a common feature, in the form of an extra term living at another spatial scale.

Hence, in the same spirit as the works of Lewin, Rota Nodari [32] and Moroz, Muratov [39], our main theorem regarding (Eω\mathrm{E}_{\omega}) establishes the behavior of uωu_{\omega} and M(ω)M(\omega) as ω0+\omega\to 0^{+}.

Theorem 1.3.

(i) Suppose N=2N=2 and p>1p>1, or N3N\geqslant 3 and 1<p<N+2N21<p<\frac{N+2}{N-2}. Then, as ω0+\omega\to 0^{+}, the rescaled function

1ω1p1uω(xω)\frac{1}{\omega^{\frac{1}{p-1}}}u_{\omega}\left(\frac{x}{\sqrt{\omega}}\right) (1.12)

converges in H1(N)C2(N)H^{1}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}) to the ground state QQ of the stationary nonlinear Schrödinger equation

ΔQQ+|Q|p1Q=0.\Delta Q-Q+|Q|^{p-1}Q=0. (1.13)

Furthermore, as ω0+\omega\to 0^{+},

M(ω)\displaystyle M(\omega) =ω4N(p1)2(p1)NQ2dx\displaystyle=\omega^{\frac{4-N(p-1)}{2(p-1)}}\int_{\mathbb{R}^{N}}Q^{2}\,\mathrm{d}x
+ω8N(p1)2(p1)2(p1)+8N(p1)4(p1)δN|Q2|2dx+o(ω8N(p1)2(p1)).\displaystyle+\omega^{\frac{8-N(p-1)}{2(p-1)}}\frac{2(p-1)+8-N(p-1)}{4(p-1)}\delta\int_{\mathbb{R}^{N}}|\nabla Q^{2}|^{2}\,\mathrm{d}x+o\Big{(}\omega^{\frac{8-N(p-1)}{2(p-1)}}\Big{)}. (1.14)

In a neighborhood of ω=0\omega=0, ωM(ω)\omega\mapsto M(\omega) is increasing if p1+4Np\leqslant 1+\frac{4}{N} and decreasing if p>1+4Np>1+\frac{4}{N}.

(ii) Suppose N3N\geqslant 3 and p=N+2N2p=\frac{N+2}{N-2}. Then there exists a function ωλω:(0,)(0,)\omega\mapsto\lambda_{\omega}:(0,\infty)\to(0,\infty) such that, as ω0+\omega\to 0^{+}, λω\lambda_{\omega}\to\infty and the rescaled function

λωN22uω(λωx)\lambda_{\omega}^{\frac{N-2}{2}}u_{\omega}(\lambda_{\omega}x) (1.15)

converges in H˙1(N)C2(N)\dot{H}^{1}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}) to the function

U(x)=(1+|x|2N(N2))N22,U(x)=\left(1+\frac{|x|^{2}}{N(N-2)}\right)^{-\frac{N-2}{2}}, (1.16)

which is the (up to dilations) unique positive radial-decreasing solution of the Lane-Emden-Fowler equation

ΔU+|U|4N2U=0.\Delta U+|U|^{\frac{4}{N-2}}U=0. (1.17)

The scaling function ωλω\omega\mapsto\lambda_{\omega} can be chosen so that, as ω0+\omega\to 0^{+},

{ω14λωω14if N=3,(ωlog1ω)1/4λω(ωlog1ω)1/4if N=4,ω1Nλωω1Nif N5.\begin{cases}\omega^{-\frac{1}{4}}\lesssim\lambda_{\omega}\lesssim\omega^{-\frac{1}{4}}&\text{if }N=3,\\ \left(\omega\log\frac{1}{\omega}\right)^{-1/4}\lesssim\lambda_{\omega}\lesssim\left(\omega\log\frac{1}{\omega}\right)^{-1/4}&\text{if }N=4,\\ \omega^{-\frac{1}{N}}\lesssim\lambda_{\omega}\lesssim\omega^{-\frac{1}{N}}&\text{if }N\geqslant 5.\end{cases} (1.18)

Furthermore,

limω0+M(ω)=limω0+M(ω)=+.\lim_{\omega\to 0^{+}}M(\omega)=-\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=+\infty. (1.19)

In a neighborhood of ω=0\omega=0, ωM(ω)\omega\mapsto M(\omega) is decreasing.

(iii) Suppose N3N\geqslant 3 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Then, as ω0+\omega\to 0^{+}, uωu_{\omega} converges in H˙1(N)C2(N)\dot{H}^{1}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}) to the unique positive radial-decreasing solution u0H˙1(N)Lp+1(N)u_{0}\in\dot{H}^{1}(\mathbb{R}^{N})\cap L^{p+1}(\mathbb{R}^{N}) of the equation

Δu+|u|p1u+δΔ(|u|2)u=0.\Delta u+|u|^{p-1}u+\delta\Delta(|u|^{2})u=0. (1.20)

Furthermore, u0(x)=O(|x|(N2))u_{0}(x)=O(|x|^{-(N-2)}) as |x||x|\to\infty. If N5N\geqslant 5, we have u0L2(N)u_{0}\in L^{2}(\mathbb{R}^{N}) and uωu0u_{\omega}\to u_{0} in L2(N)L^{2}(\mathbb{R}^{N}) as ω0+\omega\to 0^{+}.
Finally,

limω0+M(ω)={+ifN{3,4},u0L22ifN5,\lim_{\omega\to 0^{+}}M(\omega)=\begin{cases}+\infty&\text{if}\ N\in\{3,4\},\\ \|u_{0}\|_{L^{2}}^{2}&\text{if}\ N\geqslant 5,\end{cases} (1.21)

and

limω0+M(ω)={ifN{3,4,5,6},M(0)ifN7.\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=\begin{cases}-\infty&\text{if}\ N\in\{3,4,5,6\},\\ M^{\prime}(0)\in\mathbb{R}&\text{if}\ N\geqslant 7.\end{cases} (1.22)

In a neighborhood of ω=0\omega=0, ωM(ω)\omega\mapsto M(\omega) is decreasing in dimension N{3,4,5,6}N\in\{3,4,5,6\}. In dimensions N7N\geqslant 7, ωM(ω)\omega\mapsto M(\omega) is decreasing for p3+4Np\geqslant 3+\frac{4}{N}.

Remark 2.

In the subcritical case (i), the classical NLS scaling (1.12) kills the quasilinear term in the limit ω0\omega\to 0, yielding the single-power NLS (1.13) as limit equation. As in the case of the NLS with a double-power nonlinearity [32], it seems natural to use an implicit function argument to recover the branch of solutions uωu_{\omega} starting from the ground state QQ of (1.13). In the case δ>0\delta>0, the presence of the quasilinear term in (Eω\mathrm{E}_{\omega}) requires a clever choice of functional framework.

In the supercritical case (iii), the limit equation is formally obtained by letting ω=0\omega=0 in (Eω\mathrm{E}_{\omega}), which yields (1.20). However, (1.20) has no nontrivial H˙1\dot{H}^{1} solutions if pN+2N2p\leqslant\frac{N+2}{N-2}, as can be seen by combining (1.7) and (1.6) (see (4.41)). In the critical case (ii), revealing the asymptotic behavior of uωu_{\omega} as ω0\omega\to 0 is more delicate. Our approach follows that laid down in [39], based on the concentration-compactness principle. It shows that, in the rescaled variable (1.15), only two terms survive as ω0\omega\to 0, thus yielding the limit equation (1.17). As in [39], both in the critical and the supercritical cases, we will take advantage of the variational characterization of solutions of the auxiliary semilinear problem (1.3) to deduce the desired convergence of uωu_{\omega}. The main difficulty is thus to deal with the function rr whose explicit expression is not known.

Remark 3.

In the supercritical case, as for the double-power nonlinearity [32], although we are able to prove that MM^{\prime} admits a finite limit when ω0\omega\to 0 in dimension N7N\geqslant 7, we cannot determine its sign in the full range of parameters. However, if

N+2N2<p<2+4N116N+2N2(N2) or  2+4N+116N+2N2(N2)<p<3N+2N2,\frac{N+2}{N-2}<p<2+\frac{4}{N}-\sqrt{1-16\frac{N+2}{N^{2}(N-2)}}\ \text{ or }\ 2+\frac{4}{N}+\sqrt{1-16\frac{N+2}{N^{2}(N-2)}}<p<\frac{3N+2}{N-2},

then M(ω)<0M^{\prime}(\omega)<0 for ω\omega small enough, see Proposition 4.6 below. This condition is probably not optimal but it allows us to conclude that ωM(ω)\omega\mapsto M(\omega) is a decreasing function in a neighborhood of ω=0\omega=0 whenever p3+4Np\geqslant 3+\frac{4}{N}, for any N7N\geqslant 7.

As a consequence of Theorem 1.3, for any N2N\geqslant 2 and 3+4N<p<3N+2N23+\frac{4}{N}<p<\frac{3N+2}{N-2} (3+4N<p<3+\frac{4}{N}<p<\infty if N=2N=2), the positive solution uωu_{\omega}, for ω\omega small enough, lies on an unstable branch of solutions, which is in agreement with the result of instability by blow-up obtained by Colin, Jeanjean, Squassina [13, Theorem 1.5]. 111For simplicity, we speak of the stability of uωu_{\omega} when actually referring to the orbital stability of the associated standing wave eiωtuωe^{i\omega t}u_{\omega} of (1.1).

In [13], it has been conjectured that whenever 1<p<3+4N1<p<3+\frac{4}{N}, the positive solution to (Eω\mathrm{E}_{\omega}) is orbitally stable. On the one hand, our analysis for ω\omega close to 0 confirms this conjecture for 1<p1+4N1<p\leqslant 1+\frac{4}{N} since ωM(ω)\omega\mapsto M(\omega) is an increasing function in a neighborhood of 0 in this case. On the other hand, when 1+4N<p<3+4N1+\frac{4}{N}<p<3+\frac{4}{N}, the function ωM(ω)\omega\mapsto M(\omega) is decreasing in a neighborhood of 0, at least in dimension N{2,3,4,5,6}N\in\{2,3,4,5,6\}. This implies that uωu_{\omega}, the positive solution of (Eω\mathrm{E}_{\omega}), should be unstable at least for ω\omega small enough in the latter case.

The original conjecture from Colin, Jeanjean and Squassina was supported by the results obtained for the so-called normalized solutions, i.e. solutions of (Eω\mathrm{E}_{\omega}) with a prescribed L2L^{2}-norm. In particular, for a given λ>0\lambda>0, one can look for solutions of

I(λ)=inf{(u):uX,N|u|2dx=λ},I(\lambda)=\inf\left\{\mathcal{E}(u):u\in X,\int_{\mathbb{R}^{N}}|u|^{2}\,\,\mathrm{d}x=\lambda\right\}, (1.23)

where \mathcal{E} is the energy functional defined by

(u)=12N|u|2dx+δN|u|2|u|2dx1p+1N|u|p+1dx,\mathcal{E}(u)=\frac{1}{2}\int_{\mathbb{R}^{N}}|\nabla u|^{2}\,\mathrm{d}x+\delta\int_{\mathbb{R}^{N}}|u|^{2}|\nabla u|^{2}\,\,\mathrm{d}x-\frac{1}{p+1}\int_{\mathbb{R}^{N}}|u|^{p+1}\,\,\mathrm{d}x, (1.24)

with δ>0\delta>0. To each minimizer uu of (1.23) corresponds a Lagrange multiplier ω>0\omega>0 such that u=uωu=u_{\omega} (after an appropriate space translation), where uωu_{\omega} is the unique positive solution to (Eω\mathrm{E}_{\omega}).

The minimization problem (1.23) has been studied extensively over the last decade (see [13, 25, 26, 54]). The known results can be summarized as follows:

  1. (i)

    For all λ>0\lambda>0, I(λ)(,0]I(\lambda)\in(-\infty,0] if 1<p<3+4N1<p<3+\frac{4}{N} and I(λ)=I(\lambda)=-\infty if p>3+4Np>3+\frac{4}{N}.

  2. (ii)

    If 1<p<1+4N1<p<1+\frac{4}{N}, then for all λ>0\lambda>0, I(λ)<0I(\lambda)<0 and I(λ)I(\lambda) admits a minimizer.

  3. (iii)

    If 1+4Np<3+4N1+\frac{4}{N}\leqslant p<3+\frac{4}{N}, there exists λc(0,)\lambda_{c}\in(0,\infty) such that I(λ)=0I(\lambda)=0 for all 0<λλc0<\lambda\leqslant\lambda_{c} and λI(λ)\lambda\mapsto I(\lambda) is negative and strictly decreasing on (λc,)(\lambda_{c},\infty). Moreover, if 1+4N<p<3+4N1+\frac{4}{N}<p<3+\frac{4}{N} then I(λ)I(\lambda) admits a minimizer if and only if λ[λc,)\lambda\in[\lambda_{c},\infty). If p=1+4Np=1+\frac{4}{N}, then I(λ)I(\lambda) admits a minimizer if and only if λ(λc,)\lambda\in(\lambda_{c},\infty).

  4. (iv)

    If p=3+4Np=3+\frac{4}{N}, then either I(λ)=0I(\lambda)=0 or I(λ)=I(\lambda)=-\infty. As a consequence, I(λ)I(\lambda) has no minimizers for all λ(0,)\lambda\in(0,\infty).

  5. (v)

    The set of minimizers of (1.23), when it is not empty, is orbitally stable.

Unfortunately, the orbital stability of the set of minimizers does not allow one to deduce the orbital stability of a single solution uωu_{\omega} of (Eω\mathrm{E}_{\omega}) for a fixed ω>0\omega>0. On the one hand, for fixed ω>0\omega>0, one should first determine whether a solution uωu_{\omega} of (Eω\mathrm{E}_{\omega}) is also a solution of the minimization problem (1.23) with λ=M(ω)\lambda=M(\omega). Thanks to the non-degeneracy of uωu_{\omega} and the spectral properties of L+L_{+}, we know from [52, App. E] and [27, Theorem 5.3.2] (see also [35, 36]) that, uωu_{\omega} is a strict local minimum of \mathcal{E} at fixed mass λ=M(ω)\lambda=M(\omega), if M(ω)>0M^{\prime}(\omega)>0, whereas the solution uωu_{\omega} is not a local minimum when M(ω)<0M^{\prime}(\omega)<0. As a consequence, a solution uωu_{\omega} lying on a decreasing branch of the function MM cannot be a solution of (1.23), so its stability cannot be deduced from the orbital stability of the set of minimizers. On the other hand, the set of minimizers may contain more than one solution. More precisely, the uniqueness of positive solutions to (Eω\mathrm{E}_{\omega}) at fixed ω\omega does not imply the uniqueness of energy minimizers. As already mentioned, any minimizer, when it exists, is positive and solves (Eω\mathrm{E}_{\omega}) for some Lagrange multiplier ω>0\omega>0. The difficulty here is that the Lagrange multiplier is a priori not uniquely determined : for a given mass λ>0\lambda>0, there can be several minimizers that share the same energy I(λ)I(\lambda) but give rise to different Lagrange multipliers ω\omega. In other words, there may not be a one-to-one mapping λω(λ)\lambda\mapsto\omega(\lambda). Nevertheless, any candidate to be a Lagrange multiplier in this situation must be a solution to the equation M(ω)=λM(\omega)=\lambda, hence the importance of studying the behavior of the function MM and its variations. In particular, if one is able to find a region of λ\lambda’s where the function MM is one-to-one, then the uniqueness of energy minimizers follows for such λ\lambda’s.

To conclude, we mention the following result, that is a corollary of Theorem 1.3.

Corollary 1.1.

Let N=2N=2 and p>1p>1, or N3N\geqslant 3 and 1<p<3N+2N21<p<\frac{3N+2}{N-2}. For any ω>0\omega>0, let uωu_{\omega} be the unique positive solution to (Eω\mathrm{E}_{\omega}) and define E(ω):=(uω)E(\omega):=\mathcal{E}(u_{\omega}).

  1. (i)

    E(ω)=ω2M(ω)E^{\prime}(\omega)=-\frac{\omega}{2}M^{\prime}(\omega), for all ω>0\omega>0.

  2. (ii)
    limω0+E(ω)={0 if N=2 or N3 and p<N+2N2,1NN|U|2dx if N3 and p=N+2N2,2(3N+2)p(N2)N|u0|2dx if N3 and p>N+2N2,\lim_{\omega\to 0^{+}}E(\omega)=\begin{cases}0&\text{ if }N=2\text{ or }N\geqslant 3\text{ and }p<\frac{N+2}{N-2},\\ \frac{1}{N}\int_{\mathbb{R}^{N}}|\nabla U|^{2}\,\,\mathrm{d}x&\text{ if }N\geqslant 3\text{ and }p=\frac{N+2}{N-2},\\ \frac{2}{(3N+2)-p(N-2)}\int_{\mathbb{R}^{N}}|\nabla u_{0}|^{2}\,\,\mathrm{d}x&\text{ if }N\geqslant 3\text{ and }p>\frac{N+2}{N-2},\end{cases}

    where UU is the unique positive solution to (1.17) and u0u_{0} is the unique positive solution to (1.20).

  3. (iii)

    In a neighborhood of ω=0\omega=0, E(ω)<0E(\omega)<0 for p1+4Np\geqslant 1+\frac{4}{N} and E(ω)>0E(\omega)>0 for p>1+4Np>1+\frac{4}{N}.

Thanks to this result, we can conclude that if 1+4N<p<3+4N1+\frac{4}{N}<p<3+\frac{4}{N}, then the solution uωu_{\omega}, for ω\omega close to 0, cannot be a global minimizer of I(λ)I(\lambda) since its energy is strictly positive. Thus, we cannot deduce its stability from the orbital stability of the set of minimizers. In fact, our conjecture is that uωu_{\omega}, for ω\omega small, is unstable for any p>1+4Np>1+\frac{4}{N} (at least in dimensions N6N\leqslant 6).

All the previous remarks emphasize the importance of studying the function MM and, in particular, the number of sign changes of MM^{\prime}. Nevertheless, getting a complete picture on the behavior of MM seems out of reach for the moment. As in [32], a first step will be to understand the exact behavior of M(ω)M(\omega) at the two endpoints of its interval of definition : the current paper was devoted to the limit ω0+\omega\to 0^{+}, while the limit ω\omega\to\infty will be discussed in a follow-up paper currently in preparation.

Organization of the paper

Section 2 is devoted to basic properties of the change of variables rr and of the auxiliary semilinear equation (1.3). In Section 3, the uniqueness and non-degeneracy stated in Theorem 1.2 are proved, as well as Proposition 1.2. Finally, the proof of Theorem 1.3 is given in Section 4, where it is split into three subsections, dealing with cases (i), (ii) and (iii), respectively.

2. Reformulation of the problem: from a quasilinear to a semilinear equation

To prove Theorem 1.2, and Theorem 1.3 (ii) and (iii), we will use a change of variables v=h(u)v=h(u) borrowed from [34], given by

h(t):=122δ(ln(2δt+1+2δt2)+2δt1+2δt2),t0.h(t):=\frac{1}{2\sqrt{2}\delta}\left(\ln\left(\sqrt{2}\delta t+\sqrt{1+2\delta t^{2}}\right)+\sqrt{2}\delta t\sqrt{1+2\delta t^{2}}\right),\quad t\geqslant 0.

Since h(t)=1+2δt20h^{\prime}(t)=\sqrt{1+2\delta t^{2}}\geqslant 0 for all t0t\geqslant 0, it follows that hh has an inverse function r:[0,)[0,)r:[0,\infty)\to[0,\infty), which is C1C^{1} on (0,)(0,\infty) and satisfies the first order Cauchy problem

r(s)=11+2δr(s)2,s(0,),r(0)=0.r^{\prime}(s)=\frac{1}{\sqrt{1+2\delta r(s)^{2}}},\quad s\in(0,\infty),\qquad r(0)=0. (2.1)

We now extend rr to the whole real line by letting r(s)=r(s)r(s)=-r(-s) when s<0s<0.

Lemma 2.1.

The odd function r:r:\mathbb{R}\to\mathbb{R} has the following properties:

rC1()r\in C^{1}(\mathbb{R}) and rr is strictly increasing on \mathbb{R}, with r(s)>0r^{\prime}(s)>0 for all ss\in\mathbb{R};

|r(s)|1|r^{\prime}(s)|\leq 1 for all ss\in\mathbb{R} and rr is a Lipschitz function on \mathbb{R};

lims0+r(s)s=1andlims+r(s)s=(2δ)1/4.\lim_{s\to 0^{+}}\frac{r(s)}{s}=1\quad\text{and}\quad\lim_{s\to+\infty}\frac{r(s)}{\sqrt{s}}=\Big{(}\frac{2}{\delta}\Big{)}^{1/4}.

for all s+s\in\mathbb{R}_{+},

12r(s)s1+2δr2(s)r(s).\frac{1}{2}r(s)\leqslant\frac{s}{\sqrt{1+2\delta r^{2}(s)}}\leqslant r(s).
Proof.

(i) and (ii) follow immediately from (2.1). To prove (iii), integrating (2.1) yields

0sr(σ)1+2δr(σ)2dσ=s2δr(s)1+2δr(s)2+arcsinh(2δr(s))=22δs\int_{0}^{s}r^{\prime}(\sigma)\sqrt{1+2\delta r(\sigma)^{2}}\,\mathrm{d}\sigma=s\iff\sqrt{2\delta}r(s)\sqrt{1+2\delta r(s)^{2}}+\mathrm{arcsinh}\big{(}\sqrt{2\delta}r(s)\big{)}=2\sqrt{2\delta}s

(whence the formula for hh if you rather decided to define rr by (2.1)). Since, by construction, r(s)0r(s)\to 0 as s0+s\to 0^{+} and r(s)r(s)\to\infty as s+s\to+\infty, the limits in (iii) then easily follow from the above relation. To prove (iv), we consider the function gλ:+g_{\lambda}:\mathbb{R}_{+}\to\mathbb{R} defined by gλ(s)=λr(s)1+2δr2(s)sg_{\lambda}(s)=\lambda r(s)\sqrt{1+2\delta r^{2}(s)}-s for λ{12,1}\lambda\in\{\tfrac{1}{2},1\}. Clearly, gλ(0)=0g_{\lambda}(0)=0 and

gλ(s)=λr(s)1+2δr2(s)+λ2δr2(s)(r(s))21=(λ1)+λ2δr2(s)1+2δr2(s).g^{\prime}_{\lambda}(s)=\lambda r^{\prime}(s)\sqrt{1+2\delta r^{2}(s)}+\lambda 2\delta r^{2}(s)(r^{\prime}(s))^{2}-1=(\lambda-1)+\lambda\frac{2\delta r^{2}(s)}{1+2\delta r^{2}(s)}.

As a consequence, we get

{g1/2(s)=12+122δr2(s)1+2δr2(s)0,g1(s)=2δr2(s)1+2δr2(s)0,\left\{\begin{aligned} &g_{1/2}^{\prime}(s)=-\frac{1}{2}+\frac{1}{2}\frac{2\delta r^{2}(s)}{1+2\delta r^{2}(s)}\leqslant 0,\\ &g_{1}^{\prime}(s)=\frac{2\delta r^{2}(s)}{1+2\delta r^{2}(s)}\geqslant 0,\end{aligned}\right.

which implies (iv). ∎

Remark 4.

Since r:r:\mathbb{R}\to\mathbb{R} is a smooth function such that r(0)=0r(0)=0 we have (see [23, Theorem 2.87]):

s>0,vHs(N)L(N)r(v)Hs(N)L(N),\forall s>0,\ v\in H^{s}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N})\implies r(v)\in H^{s}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N}),
q1,vLq(N)r(v)Lq(N).\forall q\geqslant 1,\ v\in L^{q}(\mathbb{R}^{N})\implies r(v)\in L^{q}(\mathbb{R}^{N}).

For ω0\omega\geqslant 0, we now define fω:f_{\omega}:\mathbb{R}\to\mathbb{R},

fω(s):=11+2δr(s)2(|r(s)|p1r(s)ωr(s)),f_{\omega}(s):=\frac{1}{\sqrt{1+2\delta r(s)^{2}}}\left(|r(s)|^{p-1}r(s)-\omega r(s)\right), (2.2)

and we consider the nonlinear elliptic equation on N\mathbb{R}^{N}

Δv=fω(v).-\Delta v=f_{\omega}(v). (Sω\mathrm{S}_{\omega})

Note that the nonlinearity fωf_{\omega} is of the form fω(s)=r(s)Pω(r(s))f_{\omega}(s)=r^{\prime}(s)P_{\omega}(r(s)) with

Pω(τ)=|τ|p1τωτ.P_{\omega}(\tau)=|\tau|^{p-1}\tau-\omega\tau. (2.3)

As a consequence, for any s+s\in\mathbb{R}_{+},

fω(s)=r′′(s)Pω(r(s))+(r(s))2Pω(r(s)).f^{\prime}_{\omega}(s)=r^{\prime\prime}(s)P_{\omega}(r(s))+(r^{\prime}(s))^{2}P^{\prime}_{\omega}(r(s)). (2.4)

Furthermore, it follows from Lemma 2.1 that there exists a constant C0>0C_{0}>0 such that

s0,f(s)r(s)p1+2δr(s)2C0sp12.\forall s\geqslant 0,\quad f(s)\leqslant\frac{r(s)^{p}}{\sqrt{1+2\delta r(s)^{2}}}\leqslant C_{0}s^{\frac{p-1}{2}}. (2.5)

The two following lemmas are well-known, see [34] and [8], respectively.

Lemma 2.2.

u=r(v)u=r(v) is a classical solution of (Eω\mathrm{E}_{\omega}) if and only if vv is a classical solution of (Sω\mathrm{S}_{\omega}).

Lemma 2.3.

Any weak solution of (Eω\mathrm{E}_{\omega}) is a classical C2C^{2} solution of (Eω\mathrm{E}_{\omega}). The same holds for (Sω\mathrm{S}_{\omega}).

For all ω>0\omega>0, we attack problem (Sω\mathrm{S}_{\omega}) via a standard variational approach. Let222The normalization factor 22^{*} in (2.6) is introduced for coherence with the classical minimization problem for the critical Sobolev inequality which will be used in Section 4.3.

mω:=inf{N|z|2dx:zH1(N), 2NFω(z)dx=1},m_{\omega}:=\inf\Big{\{}\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x:\,z\in H^{1}(\mathbb{R}^{N}),\ 2^{*}\int_{\mathbb{R}^{N}}F_{\omega}(z)\,\mathrm{d}x=1\Big{\}}, (2.6)

where, using (2.1),

Fω(s):=0sfω(σ)dσ=1p+1|r(s)|p+1ω2r(s)2.F_{\omega}(s):=\int_{0}^{s}f_{\omega}(\sigma)\,\mathrm{d}\sigma=\frac{1}{p+1}|r(s)|^{p+1}-\frac{\omega}{2}r(s)^{2}. (2.7)

It follows by classical results of Berestycki-Lions [8, Theorem 2] and Berestycki-Gallouët-Kavian [7] that, for all ω>0\omega>0, the minimization problem (2.6) has a solution zωH1(N)z_{\omega}\in H^{1}(\mathbb{R}^{N}) which is spherically symmetric and radially nonincreasing. Furthermore, there is a corresponding Lagrange multiplier θω>0\theta_{\omega}>0 such that

Δzω=θωfω(zω).-\Delta z_{\omega}=\theta_{\omega}f_{\omega}(z_{\omega}). (2.8)

Then, by elliptic regularity, zωC2Lz_{\omega}\in C^{2}\cap L^{\infty}, with zω(x)0z_{\omega}(x)\to 0 exponentially as |x||x|\to\infty. Hence, zωLqz_{\omega}\in L^{q} for all q1q\geqslant 1. Furthermore, the following classical integral identities (Nehari and Pohozaev) provide a relation between mωm_{\omega} and θω\theta_{\omega}:

N|z|2dx=θωNfω(z)zdx,N|z|2dx=2θωNFω(z)dx.\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x=\theta_{\omega}\int_{\mathbb{R}^{N}}f_{\omega}(z)z\,\mathrm{d}x,\quad\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x=2^{*}\theta_{\omega}\int_{\mathbb{R}^{N}}F_{\omega}(z)\,\mathrm{d}x. (2.9)

Indeed, if zωH1Lp+1z_{\omega}\in H^{1}\cap L^{p+1} satisfies N|z|2dx=mω\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x=m_{\omega} and 2NFω(zω)dx=12^{*}\int_{\mathbb{R}^{N}}F_{\omega}(z_{\omega})\,\mathrm{d}x=1, it follows that

mω=θω>0.m_{\omega}=\theta_{\omega}>0.

Once a solution zωz_{\omega} of (2.8) is obtained, one gets a solution vωv_{\omega} of (Sω\mathrm{S}_{\omega}) by a simple dilation,

vω(x):=zω(θω1/2x),xN.v_{\omega}(x):=z_{\omega}({\theta_{\omega}}^{-1/2}x),\quad x\in\mathbb{R}^{N}.

Finally, uω=r(vω)u_{\omega}=r(v_{\omega}) is a solution of (Eω\mathrm{E}_{\omega}) for all ω>0\omega>0.

3. Uniqueness and non-degeneracy

Equipped with the change of variables in Lemma 2.2, we can now prove uniqueness and non-degeneracy of uωu_{\omega}. Our proof is a consequence of results from McLeod [37] and Lewin, Rota Nodari [32] (see also Adachi et al. [3]).

Proof of Theorem 1.2.

It is straightforward to check that fωf_{\omega} satisfies the hypotheses of [17, Theorem 4.1]. As a consequence, if uωu_{\omega} is a positive solution to (Eω\mathrm{E}_{\omega}), then it is radial decreasing with respect to some point in N\mathbb{R}^{N}. Next a direct computation shows that fωf_{\omega} satisfies the hypotheses of [37, Theorem 2] when 3<p<3N+2N23<p<\tfrac{3N+2}{N-2} for N3N\geqslant 3 or p>3p>3 for N=2N=2, and of [32, Theorem 1] when 1<p31<p\leqslant 3 (see also [3]). Thus, we deduce that the solution vωv_{\omega} is the unique positive radial solution of (Sω\mathrm{S}_{\omega}), modulo translations. Furthermore, it is non-degenerate:

{ker(Δfω(vω))=span{x1vω,,xNvω},ker(Δfω(vω)vω)=span{vω}.\left\{\begin{aligned} &\ker(-\Delta-f_{\omega}^{\prime}(v_{\omega}))=\mathrm{span}\{\partial_{x_{1}}v_{\omega},\ldots,\partial_{x_{N}}v_{\omega}\},\\ &\ker\left(-\Delta-\frac{f_{\omega}(v_{\omega})}{v_{\omega}}\right)=\mathrm{span}\{v_{\omega}\}.\end{aligned}\right. (3.1)

Since fωf_{\omega} satisfies the hypotheses of [18, Theorem 2], all positive classical solutions of (Sω\mathrm{S}_{\omega}) that tend to zero at infinity are radial decreasing about some point in N\mathbb{R}^{N}. As a consequence, the solution vωv_{\omega} is the unique positive solution of (Sω\mathrm{S}_{\omega}), modulo translations.

Thanks to the monotonicity of r(s)r(s), we can conclude that uω=r(vω)u_{\omega}=r(v_{\omega}) is the unique positive solution of (Eω\mathrm{E}_{\omega}), modulo translations.

To prove the non-degeneracy of uωu_{\omega}, we have to compute the linearized operator of (Eω\mathrm{E}_{\omega}) around uωu_{\omega}. A straightforward computation gives u=L+Re(u)+iLImu\mathcal{L}u=L_{+}\operatorname{Re}(u)+iL_{-}\operatorname{Im}{u} with L+L_{+} and LL_{-} defined by (1.9) and (1.10) respectively. Note the first eigenvalue of the operators L+L_{+} and LL_{-}, when it exists, is necessarily simple with a positive eigenvalue. This can be proved for instance by using that w,L+/w|w|,L+/|w|\langle w,L_{+/-}w\rangle\geqslant\langle|w|,L_{+/-}|w|\rangle and Harnack’s inequality.

Since uωu_{\omega} is a solution of (Eω\mathrm{E}_{\omega}), it is clear that Luω=0L_{-}u_{\omega}=0. Moreover, uω>0u_{\omega}>0. As a consequence, uωu_{\omega} is the first eigenfunction of LL_{-} and ker(L)=span{uω}\ker(L_{-})=\mathrm{span}\{u_{\omega}\}.

Next, recall that uω=r(vω)u_{\omega}=r(v_{\omega}) and let η=wr(vω)\eta=\frac{w}{r^{\prime}(v_{\omega})}. As a consequence,

L+w=\displaystyle L_{+}w= ((1+2δuω2)w)δ(4uωΔuω+2|uω|2)wpuωp1w+ωw\displaystyle-\nabla\cdot\left((1+2\delta u_{\omega}^{2})\nabla w\right)-\delta(4u_{\omega}\Delta u_{\omega}+2|\nabla u_{\omega}|^{2})w-pu_{\omega}^{p-1}w+\omega w
=\displaystyle= (1(r(vω))2w)δ(4r(vω)r(vω)Δvω+4r(vω)r′′(vω)|vω|2+2(r(vω))2|vω|2)w\displaystyle-\nabla\cdot\left(\frac{1}{(r^{\prime}(v_{\omega}))^{2}}\nabla w\right)-\delta(4r(v_{\omega})r^{\prime}(v_{\omega})\Delta v_{\omega}+4r(v_{\omega})r^{\prime\prime}(v_{\omega})|\nabla v_{\omega}|^{2}+2(r^{\prime}(v_{\omega}))^{2}|\nabla v_{\omega}|^{2})w
Pω(r(vω))w\displaystyle-P^{\prime}_{\omega}(r(v_{\omega}))w
=\displaystyle= (1r(vω)η+r′′(vω)(r(vω))2ηvω)δ(4r(vω)r′′(vω)+2(r(vω))2)|vω|2r(vω)η\displaystyle-\nabla\cdot\left(\frac{1}{r^{\prime}(v_{\omega})}\nabla\eta+\frac{r^{\prime\prime}(v_{\omega})}{(r^{\prime}(v_{\omega}))^{2}}\eta\nabla v_{\omega}\right)-\delta(4r(v_{\omega})r^{\prime\prime}(v_{\omega})+2(r^{\prime}(v_{\omega}))^{2})|\nabla v_{\omega}|^{2}r^{\prime}(v_{\omega})\eta
+2r′′(s)(r(s))2(Δvω)ηPω(r(vω))r(vω)η\displaystyle+2\frac{r^{\prime\prime}(s)}{(r^{\prime}(s))^{2}}(\Delta v_{\omega})\eta-P^{\prime}_{\omega}(r(v_{\omega}))r^{\prime}(v_{\omega})\eta
=\displaystyle= Δηr(vω)(r′′(vω)(r(vω))2)η|vω|2δ(4r(vω)r′′(vω)+2(r(vω))2)|vω|2r(vω)η\displaystyle-\frac{\Delta\eta}{r^{\prime}(v_{\omega})}-\left(\frac{r^{\prime\prime}(v_{\omega})}{(r^{\prime}(v_{\omega}))^{2}}\right)^{\prime}\eta|\nabla v_{\omega}|^{2}-\delta(4r(v_{\omega})r^{\prime\prime}(v_{\omega})+2(r^{\prime}(v_{\omega}))^{2})|\nabla v_{\omega}|^{2}r^{\prime}(v_{\omega})\eta
+r′′(s)(r(s))2(Δvω)ηPω(r(vω))r(vω)η\displaystyle+\frac{r^{\prime\prime}(s)}{(r^{\prime}(s))^{2}}(\Delta v_{\omega})\eta-P^{\prime}_{\omega}(r(v_{\omega}))r^{\prime}(v_{\omega})\eta
=\displaystyle= Δηr(vω)r′′(s)(r(s))Pω(r(vω))ηPω(r(vω))r(vω)η=Δηfω(vω)ηr(vω)\displaystyle-\frac{\Delta\eta}{r^{\prime}(v_{\omega})}-\frac{r^{\prime\prime}(s)}{(r^{\prime}(s))}P_{\omega}(r(v_{\omega}))\eta-P^{\prime}_{\omega}(r(v_{\omega}))r^{\prime}(v_{\omega})\eta=\frac{-\Delta\eta-f_{\omega}^{\prime}(v_{\omega})\eta}{r^{\prime}(v_{\omega})}

where we use the fact that r′′(s)=2δr(s)(r(s))4r^{\prime\prime}(s)=-2\delta r(s)(r^{\prime}(s))^{4} and vωv_{\omega} is a solution of (Sω\mathrm{S}_{\omega}).

Since 0<vωvω(0)0<v_{\omega}\leqslant v_{\omega}(0), the multiplier r(vω)r^{\prime}(v_{\omega}) is bounded away from 0 and we deduce that ηL2(N)\eta\in L^{2}(\mathbb{R}^{N}) if and only if wL2(N)w\in L^{2}(\mathbb{R}^{N}). Hence, wker(L+)w\in\ker(L_{+}) if and only if ηker(Δfω)\eta\in\ker(-\Delta-f^{\prime}_{\omega}). As a consequence, the non-degeneracy of vωv_{\omega} implies that wker(L+)w\in\ker(L_{+}) if and only if η=wr(vω)span{x1vω,,xNvω}\eta=\frac{w}{r^{\prime}(v_{\omega})}\in\mathrm{span}\{\partial_{x_{1}}v_{\omega},\ldots,\partial_{x_{N}}v_{\omega}\}. As a conclusion, using that xiuω=r(vω)xivω\partial_{x_{i}}u_{\omega}=r^{\prime}(v_{\omega})\partial_{x_{i}}v_{\omega} for all i=1,,Ni=1,\ldots,N, we deduce that

ker(L+)=span{x1uω,,xNuω}.\ker(L_{+})=\mathrm{span}\{\partial_{x_{1}}u_{\omega},\ldots,\partial_{x_{N}}u_{\omega}\}.

This concludes the proof of Theorem 1.2. ∎

The next proposition will be useful to prove parts (ii) and (iii) of Theorem 1.3.

Proposition 3.1.

The linearized operator L+L_{+} has exactly one negative eigenvalue.

Proof.

We shall use again the relation

L+w=Δηfω(vω)ηr(vω),L_{+}w=\frac{-\Delta\eta-f_{\omega}^{\prime}(v_{\omega})\eta}{r^{\prime}(v_{\omega})},

with η=wr(vω)\eta=\frac{w}{r^{\prime}(v_{\omega})}. In this proof we will also denote by differentiation with respect to r(0,)r\in(0,\infty). We shall use the same notation for vωv_{\omega} and its radial counterpart, i.e. vω(r)v_{\omega}(r), with r=|x|r=|x|, and similarly for other spherically symmetric functions. On the one hand, by taking w=vωr(vω)w=v^{\prime}_{\omega}r^{\prime}(v_{\omega}), we see that

L+w,w=(Δfω(vω))vω,vω=(N1)N(vω(x))2|x|2dx<0.\langle L_{+}w,w\rangle=\langle(-\Delta-f_{\omega}^{\prime}(v_{\omega}))v^{\prime}_{\omega},v^{\prime}_{\omega}\rangle=-(N-1)\int_{\mathbb{R}^{N}}\frac{(v^{\prime}_{\omega}(x))^{2}}{|x|^{2}}\,\mathrm{d}x<0.

As a consequence, L+L_{+} has at least one negative eigenvalue. Let λ1<0\lambda_{1}<0 be the first eigenvalue of L+L_{+}.

Note that, for any eigenvalue λ\lambda of L+L_{+}, there exists wL2(N)w\in L^{2}(\mathbb{R}^{N}) such that L+w=λwL_{+}w=\lambda w while the corresponding η\eta solves

(Δfω(vω)λr(vω)2)η=0,(-\Delta-f_{\omega}^{\prime}(v_{\omega})-\lambda r^{\prime}(v_{\omega})^{2})\eta=0,

that is ηker(Δfω(vω)λr(vω)2)\eta\in\ker(-\Delta-f_{\omega}^{\prime}(v_{\omega})-\lambda r^{\prime}(v_{\omega})^{2}). Since the operator Δfω(vω)λr(vω)2-\Delta-f_{\omega}^{\prime}(v_{\omega})-\lambda r^{\prime}(v_{\omega})^{2} commutes with space rotations, it may be written as a direct sum 0Aλ()𝟙\bigoplus_{\ell\geqslant 0}A^{(\ell)}_{\lambda}\otimes\mathds{1} with

Aλ()η=η′′N1sη+(+N2)s2ηfω(vω)ηλr(vω)2η.A^{(\ell)}_{\lambda}\eta=-\eta^{\prime\prime}-\frac{N-1}{s}\eta^{\prime}+\frac{\ell(\ell+N-2)}{s^{2}}\eta-f_{\omega}^{\prime}(v_{\omega})\eta-\lambda r^{\prime}(v_{\omega})^{2}\eta.

For any λ<0\lambda<0 and 1\ell\geqslant 1, Aλ()>A0()0A^{(\ell)}_{\lambda}>A^{(\ell)}_{0}\geqslant 0 in the sense of quadratic forms. As a consequence, ker(Aλ())={0}\ker(A^{(\ell)}_{\lambda})=\{0\} for all λ<0\lambda<0 and 1\ell\geqslant 1. Hence, to the first eigenvalue λ1<0\lambda_{1}<0 of L+L_{+}, there corresponds η1L2(N)\eta_{1}\in L^{2}(\mathbb{R}^{N}), η1>0\eta_{1}>0, such that

η1′′+N1sη1+fω(vω)η1+λ1r(vω)2η1=0.\eta_{1}^{\prime\prime}+\frac{N-1}{s}\eta_{1}^{\prime}+f_{\omega}^{\prime}(v_{\omega})\eta_{1}+\lambda_{1}r^{\prime}(v_{\omega})^{2}\eta_{1}=0.

Moreover, using that fω(vω)+λ1r(vω)2L(N)f_{\omega}^{\prime}(v_{\omega})+\lambda_{1}r^{\prime}(v_{\omega})^{2}\in L^{\infty}(\mathbb{R}^{N}), we can easily show that η1H1(N)\eta_{1}\in H^{1}(\mathbb{R}^{N}). A classical bootstrap argument shows that η1W2,q(N)\eta_{1}\in W^{2,q}(\mathbb{R}^{N}) for any q[2,+)q\in[2,+\infty). This, thanks to Sobolev inequalities, implies η1C1(N)\eta_{1}\in C^{1}(\mathbb{R}^{N}). Finally, using the ODE satisfied by η1\eta_{1} and the fact that fω(vω)+λ1r(vω)2L(N)f_{\omega}^{\prime}(v_{\omega})+\lambda_{1}r^{\prime}(v_{\omega})^{2}\in L^{\infty}(\mathbb{R}^{N}), we can then show that η1(0)=0\eta_{1}^{\prime}(0)=0.

Next, suppose, by contradiction, that L+L_{+} has a second eigenvalue λ2(λ1,0)\lambda_{2}\in(\lambda_{1},0). As above, there exists η2L2(N)\eta_{2}\in L^{2}(\mathbb{R}^{N}) such that

{η2′′+N1sη2+fω(vω)η2+λ2r(vω)2η2=0,η2(0)=0.\begin{cases}\eta_{2}^{\prime\prime}+\frac{N-1}{s}\eta_{2}^{\prime}+f_{\omega}^{\prime}(v_{\omega})\eta_{2}+\lambda_{2}r^{\prime}(v_{\omega})^{2}\eta_{2}=0,\\ \eta_{2}^{\prime}(0)=0.\end{cases}

Furthermore, from the proof of [32, Lemma 7.3], we know that the unique solution to

{η0′′+N1sη0+fω(vω)η0=0,η0(0)=1,η0(0)=0,\begin{cases}\eta_{0}^{\prime\prime}+\frac{N-1}{s}\eta_{0}^{\prime}+f_{\omega}^{\prime}(v_{\omega})\eta_{0}=0,\\ \eta_{0}(0)=1,\ \eta_{0}^{\prime}(0)=0,\end{cases}

vanishes exactly once and is such that lims+η0(s)==lims+η0(s)\lim_{s\to+\infty}\eta_{0}(s)=-\infty=\lim_{s\to+\infty}\eta^{\prime}_{0}(s).

Since fω(vω)+λ2r(vω)2>fω(vω)+λ1r(vω)2f_{\omega}^{\prime}(v_{\omega})+\lambda_{2}r^{\prime}(v_{\omega})^{2}>f_{\omega}^{\prime}(v_{\omega})+\lambda_{1}r^{\prime}(v_{\omega})^{2}, using Sturm’s comparison theorem, we deduce that η2\eta_{2} vanishes at least once in (0,+)(0,+\infty). Let μ2>0\mu_{2}>0 such that η2(μ2)=0\eta_{2}(\mu_{2})=0. Since fω(vω)>fω(vω)+λ2r(vω)2f_{\omega}^{\prime}(v_{\omega})>f_{\omega}^{\prime}(v_{\omega})+\lambda_{2}r^{\prime}(v_{\omega})^{2}, we can apply twice Sturm’s comparison theorem and deduce that η0\eta_{0} has at least one zero in (0,μ2)(0,\mu_{2}) and at least one zero in (μ2,+)(\mu_{2},+\infty). This contradicts the fact that η0\eta_{0} vanishes exactly once and concludes the proof. ∎

We conclude this section with the proof of Proposition 1.2, which is a consequence of the implicit function theorem.

Proof of Proposition 1.2.

Let ss\in\mathbb{N} such that s>Ns>N and define G:×W2,sHrad2LsLrad2G:\mathbb{R}\times W^{2,s}\cap H^{2}_{\mathrm{rad}}\to L^{s}\cap L^{2}_{\mathrm{rad}} as

G(ω,u)=Δu+ωuδΔ(u2)uup.G(\omega,u)=-\Delta u+\omega u-\delta\Delta(u^{2})u-u^{p}.

Using the fact that W1,s(N)L(N)W^{1,s}(\mathbb{R}^{N})\subset L^{\infty}(\mathbb{R}^{N}), we deduce that u,xjuL(N)u,\partial_{x_{j}}u\in L^{\infty}(\mathbb{R}^{N}) for any uW2,sHrad2u\in W^{2,s}\cap H^{2}_{\mathrm{rad}} and any j{1,,N}j\in\{1,\ldots,N\}. Then we can easily show that GG is well-defined and continuously Fréchet differentiable.

For any ω0>0\omega_{0}>0, let uω0u_{\omega_{0}} be the unique positive solution to (Eω\mathrm{E}_{\omega}). Then G(ω0,uω0)=0G(\omega_{0},u_{\omega_{0}})=0 and, since uω0u_{\omega_{0}} is non-degenerate, DuG(ω0,uω0)=L+D_{u}G(\omega_{0},u_{\omega_{0}})=L_{+} is one-to-one. Hence, it remains to prove that L+L_{+} is an isomorphism from W2,sHrad2W^{2,s}\cap H^{2}_{\mathrm{rad}} onto LsLrad2L^{s}\cap L^{2}_{\mathrm{rad}}. Since uω0u_{\omega_{0}} is in the Schwartz space 𝒮(N)\mathcal{S}(\mathbb{R}^{N}), the operator L+L_{+} can be seen as a compact perturbation of (1+2δuω02)Δ+ω0-(1+2\delta u_{\omega_{0}}^{2})\Delta+\omega_{0} which is itself an isomorphism from W2,sHrad2W^{2,s}\cap H^{2}_{\mathrm{rad}} onto LsLrad2L^{s}\cap L^{2}_{\mathrm{rad}}.

Applying the implicit function theorem, we conclude that there exists a C1C^{1} map ωu(ω)W2,sHrad2\omega\mapsto u(\omega)\in W^{2,s}\cap H^{2}_{\mathrm{rad}} defined in a neighborhood of ω0\omega_{0} of solutions to (Eω\mathrm{E}_{\omega}).

To conclude that u(ω)=uωu(\omega)=u_{\omega}, it suffices to prove that u(ω)u(\omega) is positive, which can be done via a spectral theory argument. For each ω>0\omega>0, consider the operator

Hω=Δ+Vω,H_{\omega}=-\Delta+V_{\omega},

where

Vω(x):=ωδΔ(u(ω)(x)2)u(ω)(x)p1,xN.V_{\omega}(x):=\omega-\delta\Delta(u(\omega)(x)^{2})-u(\omega)(x)^{p-1},\quad x\in\mathbb{R}^{N}.

Note that VωLV_{\omega}\in L^{\infty} and HωH_{\omega} has a zero eigenvalue with eigenfunction u(ω)u(\omega). Moreover, when ω=ω0\omega=\omega_{0}, u(ω0)=uω0>0u(\omega_{0})=u_{\omega_{0}}>0 and 0 is an isolated simple eigenvalue at the bottom of the spectrum. Hence, for ω\omega close to ω0\omega_{0}, the zero eigenvalue of HωH_{\omega} must also be at the bottom of the spectrum. Hence, by applying [46, Theorem XIII.46], we deduce that u(ω)u(\omega) must be positive for any ω\omega close to ω0\omega_{0}. ∎

4. Proof of Theorem 1.3

We shall decompose the proof of Theorem 1.3 into three parts: first (i), then (iii), and finally (ii), which is more involved.

4.1. Subcritical case

In the subcritical case, the behavior of uωu_{\omega} as ω0+\omega\to 0^{+} can be determined by a straightforward application of the implicit function theorem.

Proof of Theorem 1.3 (i).

When N=2N=2 or N3N\geqslant 3 and 1<p<N+2N21<p<\frac{N+2}{N-2}, the rescaled function u~ω\tilde{u}_{\omega} defined by (1.12) solves

Δu~ωu~ω+u~ωp+ω2p1δΔ(u~ω2)u~ω=0\Delta\tilde{u}_{\omega}-\tilde{u}_{\omega}+\tilde{u}_{\omega}^{p}+\omega^{\frac{2}{p-1}}\delta\Delta(\tilde{u}_{\omega}^{2})\tilde{u}_{\omega}=0 (4.1)

and it converges to the unique positive solution QQ to the NLS equation (1.13). More precisely, the implicit function theorem gives

u~ωQ+ω2p1δ(Q)rad1(Δ(Q2)Q)H2(N)L(N)=o(ω2p1),\|\tilde{u}_{\omega}-Q+\omega^{\frac{2}{p-1}}\delta(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)\|_{H^{2}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N})}=o\left(\omega^{\frac{2}{p-1}}\right), (4.2)

where Q:=ΔpQp1+1\mathcal{L}_{Q}:=-\Delta-pQ^{p-1}+1 and H2(N)L(N)=max{H2(N),L(N)}\|\cdot\|_{H^{2}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N})}=\max\{\|\cdot\|_{H^{2}(\mathbb{R}^{N})},\|\cdot\|_{L^{\infty}(\mathbb{R}^{N})}\}. Using (4.2), we obtain

M(ω)=\displaystyle M(\omega)= Nuω2(x)dx=ω2p1N2Nu~ω2(x)dx\displaystyle\,\int_{\mathbb{R}^{N}}u_{\omega}^{2}(x)\,\mathrm{d}x=\omega^{\frac{2}{p-1}-\frac{N}{2}}\int_{\mathbb{R}^{N}}\tilde{u}_{\omega}^{2}(x)\,\mathrm{d}x
=\displaystyle= ω4N(p1)2(p1)NQ2dx2ω8N(p1)2(p1)Q,δ(Q)rad1(Δ(Q2)Q)+o(ω8N(p1)2(p1)).\displaystyle\,\omega^{\frac{4-N(p-1)}{2(p-1)}}\int_{\mathbb{R}^{N}}Q^{2}\,\mathrm{d}x-2\omega^{\frac{8-N(p-1)}{2(p-1)}}\langle Q,\delta(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)\rangle+o\left(\omega^{\frac{8-N(p-1)}{2(p-1)}}\right). (4.3)

To compute the derivative of MM, we note that L+ωuω=uωL_{+}\partial_{\omega}u_{\omega}=-u_{\omega} and, thanks to the non-degeneracy of uωu_{\omega}, we can write

M(ω)=2Nuω(x)ωuω(x)dx=2Nuω(x)(L+)1uω(x)dx.M^{\prime}(\omega)=2\int_{\mathbb{R}^{N}}u_{\omega}(x)\partial_{\omega}u_{\omega}(x)\,\mathrm{d}x=-2\int_{\mathbb{R}^{N}}u_{\omega}(x)(L_{+})^{-1}u_{\omega}(x)\,\mathrm{d}x.

Using the change of variables defined in (1.12), we obtain

M(ω)=2ω4N(p1)2(p1)1Nu~ω(x)(~(ω))1u~ω(x)dx.M^{\prime}(\omega)=-2\omega^{\frac{4-N(p-1)}{2(p-1)}-1}\int_{\mathbb{R}^{N}}\tilde{u}_{\omega}(x)(\tilde{\mathcal{L}}(\omega))^{-1}\tilde{u}_{\omega}(x)\,\mathrm{d}x.

with

~(ω)=(1+2δω2p1u~ω2)Δ4δω2p1u~ωu~ωω2p1δ(4u~ωΔu~ω+2|u~ω|2)pu~ωp1+1.\tilde{\mathcal{L}}(\omega)=-(1+2\delta\omega^{\frac{2}{p-1}}\tilde{u}_{\omega}^{2})\Delta-4\delta\omega^{\frac{2}{p-1}}\tilde{u}_{\omega}\nabla\tilde{u}_{\omega}\cdot\nabla-\omega^{\frac{2}{p-1}}\delta(4\tilde{u}_{\omega}\Delta\tilde{u}_{\omega}+2|\nabla\tilde{u}_{\omega}|^{2})-p\tilde{u}_{\omega}^{p-1}+1.

Since u~ω\tilde{u}_{\omega} converges to QQ in LL^{\infty} as ω0\omega\to 0, we deduce that ~(ω)\tilde{\mathcal{L}}(\omega) converges to Q\mathcal{L}_{Q} in the norm resolvent sense (see [45, Theorem VIII.25]). Since 0ρ((Q)rad)ρ(~(ω)rad)0\in\rho((\mathcal{L}_{Q})_{\mathrm{rad}})\cap\rho(\tilde{\mathcal{L}}(\omega)_{\mathrm{rad}}), we obtain the convergence

(~(ω)rad)1((Q)rad)1(\tilde{\mathcal{L}}(\omega)_{\mathrm{rad}})^{-1}\to((\mathcal{L}_{Q})_{\mathrm{rad}})^{-1} (4.4)

in norm. More precisely, using (4.2), we have

Q\displaystyle\mathcal{L}_{Q} ~(ω)=ω2p1(2δu~ω2Δ+4δu~ωu~ω+δ(4u~ωΔu~ω+2|u~ω|2))+p(uωp1Qp1)\displaystyle\,-\tilde{\mathcal{L}}(\omega)=\omega^{\frac{2}{p-1}}\left(2\delta\tilde{u}_{\omega}^{2}\Delta+4\delta\tilde{u}_{\omega}\nabla\tilde{u}_{\omega}\cdot\nabla+\delta(4\tilde{u}_{\omega}\Delta\tilde{u}_{\omega}+2|\nabla\tilde{u}_{\omega}|^{2})\right)+p(u_{\omega}^{p-1}-Q^{p-1})
=ω2p1δ(2Q2Δ+4QQ+4QΔQ+2|Q|2p(p1)Qp2(Q)rad1(Δ(Q2)Q))+o(ω2p1)\displaystyle\,=\omega^{\frac{2}{p-1}}\delta\left(2Q^{2}\Delta+4Q\nabla Q\cdot\nabla+4Q\Delta Q+2|\nabla Q|^{2}-p(p-1)Q^{p-2}(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)\right)+o(\omega^{\frac{2}{p-1}})
=ω2p1δ(2(Q2)+Θ)+o(ω2p1),\displaystyle\,=-\omega^{\frac{2}{p-1}}\delta(-2\nabla\cdot(Q^{2}\nabla)+\Theta)+o(\omega^{\frac{2}{p-1}}),

where Θ:=(4QΔQ+2|Q|2p(p1)Qp2(Q)rad1(Δ(Q2)Q))\Theta:=-(4Q\Delta Q+2|\nabla Q|^{2}-p(p-1)Q^{p-2}(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)) is understood as a multiplication operator. By iterating the resolvent identity, we obtain

(~(ω)rad)1((Q)rad)1+ω2p1δ(Q)rad1(2(Q2)+Θ)(Q)rad1=o(ω2p1).\|(\tilde{\mathcal{L}}(\omega)_{\mathrm{rad}})^{-1}-((\mathcal{L}_{Q})_{\mathrm{rad}})^{-1}+\omega^{\frac{2}{p-1}}\delta(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-2\nabla\cdot(Q^{2}\nabla)+\Theta)(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}\|=o(\omega^{\frac{2}{p-1}}).

Note that (Q)rad1(Δ(Q2)Q)(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q) decays at the same rate as QQ so that Qp2(Q)rad1(Δ(Q2)Q)Q^{p-2}(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q) tends to 0 at infinity even if p<2p<2. As a consequence,

ω14N(p1)2(p1)M(ω)=\displaystyle\omega^{1-\frac{4-N(p-1)}{2(p-1)}}M^{\prime}(\omega)= 2Qω2p1(Q)rad1(Δ(Q2)Q),(~(ω)rad)1(Qω2p1(Q)rad1(Δ(Q2)Q))\displaystyle\,-2\left\langle Q-\omega^{\frac{2}{p-1}}(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q),(\tilde{\mathcal{L}}(\omega)_{\mathrm{rad}})^{-1}(Q-\omega^{\frac{2}{p-1}}(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q))\right\rangle
+o(ω2p1)\displaystyle\,+o(\omega^{\frac{2}{p-1}})
=\displaystyle= 2Q,(Q)rad1Q+2ω2p1δQ,(Q)rad1(2(Q2)+Θ)(Q)rad1Q\displaystyle\,-2\left\langle Q,(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q\right\rangle+2\omega^{\frac{2}{p-1}}\delta\left\langle Q,(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-2\nabla\cdot(Q^{2}\nabla)+\Theta)(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q\right\rangle
+4ω2p1δ(Q)rad1Q,(Q)rad1(Δ(Q2)Q)+o(ω2p1).\displaystyle\,+4\omega^{\frac{2}{p-1}}\delta\left\langle(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q,(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)\right\rangle+o(\omega^{\frac{2}{p-1}}). (4.5)

With a scaling argument, we can easily compute

(Δ+pQp1+1)(rQ2+Qp1)=Q(-\Delta+pQ^{p-1}+1)\left(\frac{rQ^{\prime}}{2}+\frac{Q}{p-1}\right)=-Q

for the NLS case, so that

(Q)rad1Q=(rQ2+Qp1).(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q=-\left(\frac{rQ^{\prime}}{2}+\frac{Q}{p-1}\right).

This leads to

2Q,(Q)rad1Q=2Q,rQ2+Qp1=4N(p1)2(p1)QL22-2\left\langle Q,(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q\right\rangle=2\left\langle Q,\frac{rQ^{\prime}}{2}+\frac{Q}{p-1}\right\rangle=\frac{4-N(p-1)}{2(p-1)}\|Q\|^{2}_{L^{2}}

and gives the first-order term in MM^{\prime} whenever p1+4Np\neq 1+\frac{4}{N}.

To compute the next-order term in MM and MM^{\prime} we proceed as follows. Let Qω(x)=ω1p1Q(ω1/2x)Q_{\omega}(x)=\omega^{\frac{1}{p-1}}Q(\omega^{1/2}x). We know that

ΔQωQωp+ωQω=0 and ωQω=(Qω)rad1Qω-\Delta Q_{\omega}-Q^{p}_{\omega}+\omega Q_{\omega}=0\text{ and }\partial_{\omega}Q_{\omega}=-(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega}

with

Qω=ΔpQωp1+ω.\mathcal{L}_{Q_{\omega}}=-\Delta-pQ^{p-1}_{\omega}+\omega.

On the one hand,

2Qω,(Qω)rad1(Δ(Qω2)Qω)\displaystyle-2\left\langle{Q_{\omega}},(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2}_{\omega})Q_{\omega})\right\rangle =2ωQω,Δ(Qω2)Qω=12ωNQω2(Δ(Qω2))dx\displaystyle\,=2\left\langle\partial_{\omega}{Q_{\omega}},-\Delta(Q^{2}_{\omega})Q_{\omega}\right\rangle=\frac{1}{2}\partial_{\omega}\int_{\mathbb{R}^{N}}Q_{\omega}^{2}(-\Delta(Q^{2}_{\omega}))\,\mathrm{d}x
=12ωN|Qω2|2dx.\displaystyle\,=\frac{1}{2}\partial_{\omega}\int_{\mathbb{R}^{N}}\left|\nabla Q_{\omega}^{2}\right|^{2}\,\mathrm{d}x.

On the other hand,

N|Qω2|2dx=ω4p1+1N2N|Q2|2dx\displaystyle\int_{\mathbb{R}^{N}}\left|\nabla Q_{\omega}^{2}\right|^{2}\,\mathrm{d}x=\omega^{\frac{4}{p-1}+1-\frac{N}{2}}\int_{\mathbb{R}^{N}}\left|\nabla Q^{2}\right|^{2}\,\mathrm{d}x

so that

2Q,(Q)rad1(Δ(Q2)Q)=12(4p1+1N2)N|Q2|2dx.\displaystyle-2\left\langle{Q},(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2})Q)\right\rangle=\frac{1}{2}\left(\frac{4}{p-1}+1-\frac{N}{2}\right)\int_{\mathbb{R}^{N}}\left|\nabla Q^{2}\right|^{2}\,\mathrm{d}x. (4.6)

Formulas (4.1) and (4.6) prove (1.3). (Note that, alternatively, (4.6) can be obtained by using directly the explicit expression of (Q)rad1Q(\mathcal{L}_{Q})_{\mathrm{rad}}^{-1}Q.)

Similarly, we consider ω2Qω\partial^{2}_{\omega}Q_{\omega} and we compute

Qωω2Qω=p(p1)Qωp2(ωQω)22ωQω=p(p1)Qωp2((Qω)rad1Qω)2+2(Qω)rad1Qω.\displaystyle\mathcal{L}_{Q_{\omega}}\partial^{2}_{\omega}Q_{\omega}=p(p-1)Q_{\omega}^{p-2}(\partial_{\omega}Q_{\omega})^{2}-2\partial_{\omega}Q_{\omega}=p(p-1)Q_{\omega}^{p-2}((\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega})^{2}+2(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega}.

Hence,

14ω2N|Qω2|2dx=\displaystyle\frac{1}{4}\partial^{2}_{\omega}\int_{\mathbb{R}^{N}}\left|\nabla Q_{\omega}^{2}\right|^{2}\,\mathrm{d}x= ω2Qω,Δ(Qω2)Qω+(ωQω)2,Δ(Qω2)+2QωωQω,Δ(QωωQω)\displaystyle\,\left\langle\partial^{2}_{\omega}{Q_{\omega}},-\Delta(Q^{2}_{\omega})Q_{\omega}\right\rangle+\left\langle(\partial_{\omega}{Q_{\omega}})^{2},-\Delta(Q^{2}_{\omega})\right\rangle+2\left\langle Q_{\omega}\partial_{\omega}{Q_{\omega}},-\Delta(Q_{\omega}\partial_{\omega}Q_{\omega})\right\rangle
=\displaystyle= p(p1)Qωp2((Qω)rad1Qω)2+2(Qω)rad1Qω,(Qω)rad1(Δ(Qω2)Qω)\displaystyle\,\left\langle p(p-1)Q_{\omega}^{p-2}((\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega})^{2}+2(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega},\right(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2}_{\omega})Q_{\omega})\rangle
+((Qω)rad1Qω)2,2|Qω|24QωΔQω+2ωQω,(Qω2ωQω)\displaystyle\,+\left\langle((\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega})^{2},-2|\nabla Q_{\omega}|^{2}-4Q_{\omega}\Delta Q_{\omega}\right\rangle+2\left\langle\partial_{\omega}Q_{\omega},-\nabla\cdot(Q_{\omega}^{2}\nabla\partial_{\omega}Q_{\omega})\right\rangle
=\displaystyle= (Qω)rad1Qω,2(Qω2(Qω)rad1Qω)+(Qω)rad1Qω,Θ(ω)(Qω)rad1Qω\displaystyle\,\left\langle(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega},-2\nabla\cdot(Q_{\omega}^{2}\nabla(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega})\right\rangle+\left\langle(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega},\Theta(\omega)(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega}\right\rangle
+2(Qω)rad1Qω,(Qω)rad1(Δ(Qω2)Qω)\displaystyle\,+2\left\langle(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}Q_{\omega},\right(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}(-\Delta(Q^{2}_{\omega})Q_{\omega})\rangle

with Θ(ω):=(4QΔQω+2|Qω|2p(p1)Qωp2(Qω)rad1(Δ(Qω2)Qω))\Theta(\omega):=-(4Q\Delta Q_{\omega}+2|\nabla Q_{\omega}|^{2}-p(p-1)Q_{\omega}^{p-2}(\mathcal{L}_{Q_{\omega}})_{\mathrm{rad}}^{-1}(-\Delta(Q_{\omega}^{2})Q_{\omega})). As a consequence, it is enough to evaluate 12ω2N|Qω2|2dx\frac{1}{2}\partial^{2}_{\omega}\int_{\mathbb{R}^{N}}\left|\nabla Q_{\omega}^{2}\right|^{2}\,\mathrm{d}x at ω=1\omega=1 to conclude. In particular, this gives

ω14N(p1)2(p1)M(ω)\displaystyle\omega^{1-\frac{4-N(p-1)}{2(p-1)}}M^{\prime}(\omega)
=4N(p1)2(p1)QL22+ω2p1δ(8(N2)(p1))(8N(p1))8(p1)2(Q2)L22+o(ω2p1).\displaystyle=\frac{4-N(p-1)}{2(p-1)}\|Q\|^{2}_{L^{2}}+\omega^{\frac{2}{p-1}}\delta\frac{(8-(N-2)(p-1))(8-N(p-1))}{8(p-1)^{2}}\|\nabla(Q^{2})\|^{2}_{L^{2}}+o(\omega^{\frac{2}{p-1}}).

For p=1+4Np=1+\frac{4}{N}, we obtain

M(ω)=ωN21δN(N+2)4(Q2)L22+o(ωN21).\displaystyle M^{\prime}(\omega)=\omega^{\frac{N}{2}-1}\delta\frac{N(N+2)}{4}\|\nabla(Q^{2})\|^{2}_{L^{2}}+o(\omega^{\frac{N}{2}-1}).

The monotonicity properties of the map ωM(ω)\omega\mapsto M(\omega) follow from these formulas. This concludes the proof of part (i) of Theorem 1.3. ∎

4.2. Supercritical case

To prove part (iii) of Theorem 1.3, we shall take advantage of another variational characterization of solutions of (Sω\mathrm{S}_{\omega}). We define the functional Jω:KωJ_{\omega}:K_{\omega}\to\mathbb{R} by

Jω(z):=N|z|2dx(2NFω(z)dx)12/N,J_{\omega}(z):=\frac{\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x}{\left(2^{*}\int_{\mathbb{R}^{N}}F_{\omega}(z)\,\mathrm{d}x\right)^{1-2/N}},

where

Kω:={zH1(N):NFω(z)dx>0}.K_{\omega}:=\{z\in H^{1}(\mathbb{R}^{N}):\,\textstyle\int_{\mathbb{R}^{N}}F_{\omega}(z)\,\mathrm{d}x>0\}.

It follows from [8] that KωøK_{\omega}\neq\mbox{\Large\o}. We next observe that JωJ_{\omega} is invariant under dilations: Jω(zλ)=Jω(z)J_{\omega}(z_{\lambda})=J_{\omega}(z) if zKωz\in K_{\omega} and

zλ(x):=z(λ1/2x),λ>0,xN.z_{\lambda}(x):=z(\lambda^{-1/2}x),\quad\lambda>0,\ x\in\mathbb{R}^{N}.
Proposition 4.1.

Let mωm_{\omega} be defined by (2.6). There holds

infzKωJω(z)=mω.\inf_{z\in K_{\omega}}J_{\omega}(z)=m_{\omega}.

Furthermore, suppose φKω\varphi\in K_{\omega} satisfies Jω(φ)=mωJ_{\omega}(\varphi)=m_{\omega}. Then, letting λ=(2NFω(φ)dx)2/N\lambda=(2^{*}\int_{\mathbb{R}^{N}}F_{\omega}(\varphi)\,\mathrm{d}x)^{-2/N}, the dilation φλ\varphi_{\lambda} is a minimizer for (2.6).

Proof.

Let zH1z\in H^{1} be a minimizer for (2.6):

|z|2=mω,2Fω(z)=1.\int|\nabla z|^{2}=m_{\omega},\quad 2^{*}\int F_{\omega}(z)=1.

It follows that zKωz\in K_{\omega} and Jω(z)=mωJ_{\omega}(z)=m_{\omega}. Hence, infzKωJω(z)mω\inf_{z\in K_{\omega}}J_{\omega}(z)\leqslant m_{\omega}. Suppose by contradiction that infzKωJω(z)<mω\inf_{z\in K_{\omega}}J_{\omega}(z)<m_{\omega}. One can then find zKωz\in K_{\omega} such that Jω(z)<mωJ_{\omega}(z)<m_{\omega}. Letting λ=(2Fω(φ))2/N\lambda=(2^{*}\int F_{\omega}(\varphi))^{-2/N}, the dilation zλz_{\lambda} satisfies 2Fω(zλ)=12^{*}\int F_{\omega}(z_{\lambda})=1, so that

mω|zλ|2=Jω(zλ)=Jω(z)<mω.m_{\omega}\leqslant\int|\nabla z_{\lambda}|^{2}=J_{\omega}(z_{\lambda})=J_{\omega}(z)<m_{\omega}.

This contradiction shows that, indeed, infzKωJω(z)=mω\inf_{z\in K_{\omega}}J_{\omega}(z)=m_{\omega}.

Finally, if φKω\varphi\in K_{\omega} is such that Jω(φ)=mωJ_{\omega}(\varphi)=m_{\omega}, again the dilation factor λ=(2Fω(φ))2/N\lambda=(2^{*}\int F_{\omega}(\varphi))^{-2/N} yields 2Fω(φλ)=12^{*}\int F_{\omega}(\varphi_{\lambda})=1, and so φλ\varphi_{\lambda} is a minimizer for (2.6). ∎

We shall now consider the so-called ‘zero-mass case’, ω=0\omega=0. We let

m0:=inf{N|z|2dx:zH˙1(N)Lp+1(N), 2NF0(z)dx=1},m_{0}:=\inf\Big{\{}\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x:\,z\in\dot{H}^{1}(\mathbb{R}^{N})\cap L^{p+1}(\mathbb{R}^{N}),\ 2^{*}\int_{\mathbb{R}^{N}}F_{0}(z)\,\mathrm{d}x=1\Big{\}}, (4.7)

where f0f_{0} and F0F_{0} are still defined by (2.2) and (2.7), respectively. The existence of a spherically symmetric and radially nonincreasing minimizer z0H˙1Lp+1C2z_{0}\in\dot{H}^{1}\cap L^{p+1}\cap C^{2} of (4.7) follows from [8]. It is also known (see [6, 16]) that z0(x)z_{0}(x) decays like |x|(N2)|x|^{-(N-2)} as |x||x|\to\infty. Furthermore, as in the case ω>0\omega>0, there exists a Lagrange multiplier θ0>0\theta_{0}>0 such that

Δz0=θ0f0(z0).-\Delta z_{0}=\theta_{0}f_{0}(z_{0}). (4.8)

The integral identities (2.9) still hold for ω=0\omega=0, and it follows that

m0=θ0>0.m_{0}=\theta_{0}>0.

Again, the dilation

v0(x):=z0(θ01/2x),xN,v_{0}(x):=z_{0}({\theta_{0}}^{-1/2}x),\quad x\in\mathbb{R}^{N},

produces a solution of

Δv=f0(v),-\Delta v=f_{0}(v), (P0\mathrm{P}_{0})

and the change of variables u0=r(v0)u_{0}=r(v_{0}) yields a solution of (1.20). Note that, as in the case ω>0\omega>0, the above changes of variables give a one-to-one correspondence between u0u_{0} and z0z_{0}. Furthermore, u0u_{0} and z0z_{0} have the same regularity and decay at infinity. The main difference here compared to the case ω>0\omega>0 is that, in dimensions N=3,4N=3,4, z0L2z_{0}\not\in L^{2}, due to its slow decay as |x||x|\to\infty. Only in dimensions N5N\geqslant 5 does z0L2z_{0}\in L^{2}.

We now define J0:K0J_{0}:K_{0}\to\mathbb{R} by

J0(z):=N|z|2dx(2NF0(z)dx)12/N,J_{0}(z):=\frac{\int_{\mathbb{R}^{N}}|\nabla z|^{2}\,\mathrm{d}x}{\left(2^{*}\int_{\mathbb{R}^{N}}F_{0}(z)\,\mathrm{d}x\right)^{1-2/N}},

where

K˙0:={zH˙1(N)Lp+1(N):NF0(z)dx>0}.\dot{K}_{0}:=\{z\in\dot{H}^{1}(\mathbb{R}^{N})\cap L^{p+1}(\mathbb{R}^{N}):\,\textstyle\int_{\mathbb{R}^{N}}F_{0}(z)\,\mathrm{d}x>0\}.

The following result is proved in the exact same way as Proposition 4.1.

Proposition 4.2.

There holds

infzK˙0J0(z)=m0.\inf_{z\in\dot{K}_{0}}J_{0}(z)=m_{0}.

Furthermore, suppose φK˙0\varphi\in\dot{K}_{0} satisfies J0(φ)=m0J_{0}(\varphi)=m_{0}. Then, letting λ=(2NF0(φ)dx)2/N\lambda=(2^{*}\int_{\mathbb{R}^{N}}F_{0}(\varphi)\,\mathrm{d}x)^{-2/N}, the dilation φλ\varphi_{\lambda} is a minimizer for (4.7).

We also have uniqueness, up to translations, and non-degeneracy of the minimizer z0z_{0}, as a consequence of the following theorem.

Theorem 4.1.

Suppose N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Then the following properties hold true.

(4.8) has a unique positive radial solution z0z_{0} in H˙1\dot{H}^{1}.

z0z_{0} non-degenerate:

ker(L0)=span{iz0:i=1,,N},\ker(L_{0})=\mathrm{span}\{\partial_{i}z_{0}:\,i=1,\dots,N\},

where

L0=Δθ0f0(z0).L_{0}=-\Delta-\theta_{0}f_{0}^{\prime}(z_{0}). (4.9)

The proof of the uniqueness is a direct application of [51, Theorem 2] while the non-degeneracy in L2L^{2} can be proved as in [32, Lemma A.1].

We can now state and prove the first main result of this section.

Proposition 4.3.

Suppose N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Consider a sequence (ωn)(0,)(\omega_{n})\subset(0,\infty), such that ωn0\omega_{n}\to 0. Then

vωnv0inH˙1(N)Lq(N)C2(N),q2,v_{\omega_{n}}\to v_{0}\quad\text{in}\quad\dot{H}^{1}(\mathbb{R}^{N})\cap L^{q}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}),\quad\forall q\geqslant 2^{*}, (4.10)

and

ωnr(vωn)L220.\omega_{n}\|r(v_{\omega_{n}})\|_{L^{2}}^{2}\to 0. (4.11)

We will prove Proposition 4.3 using several lemmas. The first one is technical.

Lemma 4.1.

Let N{3,4}N\in\{3,4\}. Consider a function zC(N)z\in C(\mathbb{R}^{N}), a number ρ>0\rho>0 and ηρC0(N)\eta_{\rho}\in C_{0}^{\infty}(\mathbb{R}^{N}) a cut-off function such that ηρ1\eta_{\rho}\equiv 1 on Bρ(0)B_{\rho}(0) and ηρ0\eta_{\rho}\equiv 0 on NB2ρ(0)\mathbb{R}^{N}\setminus B_{2\rho}(0). Suppose there exists a constant C>0C>0 such that

lim|x||z(x)||x|N2=C.\lim_{|x|\to\infty}\frac{|z(x)|}{|x|^{N-2}}=C.

Then we have the following asymptotics as ρ\rho\to\infty:

N|ηρz|2dx={O(ρ)ifN=3,O(log(ρ))ifN=4.\int_{\mathbb{R}^{N}}|\eta_{\rho}z|^{2}\,\mathrm{d}x=\begin{cases}O(\rho)&\text{if}\ N=3,\\ O(\log(\rho))&\text{if}\ N=4.\end{cases} (4.12)
Proof.

Let f:(0,)f:(0,\infty)\to\mathbb{R} be a continuous function such that 0<f(ρ)<ρ0<f(\rho)<\rho for all ρ>0\rho>0. For ρ1\rho\gg 1, we have

N|ηρz|2dxf(ρ)|x|<ρ|ηρz|2dx=f(ρ)|x|<ρ|z(x)|2dxf(ρ)ρr2(N2)rN1dr.\int_{\mathbb{R}^{N}}|\eta_{\rho}z|^{2}\,\mathrm{d}x\geqslant\int_{f(\rho)\leqslant|x|<\rho}|\eta_{\rho}z|^{2}\,\mathrm{d}x=\int_{f(\rho)\leqslant|x|<\rho}|z(x)|^{2}\,\mathrm{d}x\gtrsim\int_{f(\rho)}^{\rho}r^{-2(N-2)}r^{N-1}\,\mathrm{d}r.

If N=3N=3, choosing f(ρ)=ρ/2f(\rho)=\rho/2 yields

N|ηρz|2dxρ/2ρdr=ρ2.\int_{\mathbb{R}^{N}}|\eta_{\rho}z|^{2}\,\mathrm{d}x\gtrsim\int_{\rho/2}^{\rho}\,\mathrm{d}r=\frac{\rho}{2}.

If N=4N=4, we choose f(ρ)=ρf(\rho)=\sqrt{\rho} and we get

N|ηρz|2dxρρr1dr=logρ2.\int_{\mathbb{R}^{N}}|\eta_{\rho}z|^{2}\,\mathrm{d}x\gtrsim\int_{\sqrt{\rho}}^{\rho}r^{-1}\,\mathrm{d}r=\frac{\log\rho}{2}.

This gives the desired lower bounds. The upper bounds are straightforward. ∎

Lemma 4.2.

As ω0\omega\to 0, there holds

1mωm01+o(1).1\leqslant\frac{m_{\omega}}{m_{0}}\leqslant 1+o(1).
Proof.

Firstly, for ω>0\omega>0 and z>0z>0,

F0(z)=1p+1r(z)p+1>1p+1r(z)p+1ω2r(z)2=Fω(z).F_{0}(z)=\frac{1}{p+1}r(z)^{p+1}>\frac{1}{p+1}r(z)^{p+1}-\frac{\omega}{2}r(z)^{2}=F_{\omega}(z).

Hence, using the minimizer zωz_{\omega} as a test function for J0J_{0}, we have

m0J0(zω)=|zω|2(2F0(zω))12/N<Jω(zω)=mω.m_{0}\leqslant J_{0}(z_{\omega})=\frac{\int|\nabla z_{\omega}|^{2}}{\left(2^{*}\int F_{0}(z_{\omega})\right)^{1-2/N}}<J_{\omega}(z_{\omega})=m_{\omega}.

Now, if N5N\geqslant 5, we have that z0L2z_{0}\in L^{2}. Furthermore, since z0K˙0z_{0}\in\dot{K}_{0}, it follows by continuity that, for ω>0\omega>0 small enough, z0Kωz_{0}\in K_{\omega}. Therefore, we can use z0z_{0} as a test function for JωJ_{\omega}, which yields

mωJω(z0)=|z0|2(2Fω(z0))12/N=|z0|2(2F0(z0))12/N(2F0(z0))12/N(2Fω(z0))12/N=m0(F0(z0)Fω(z0))12/N.m_{\omega}\leqslant J_{\omega}(z_{0})=\frac{\int|\nabla z_{0}|^{2}}{\left(2^{*}\int F_{\omega}(z_{0})\right)^{1-2/N}}=\frac{\int|\nabla z_{0}|^{2}}{\left(2^{*}\int F_{0}(z_{0})\right)^{1-2/N}}\cdot\frac{\left(2^{*}\int F_{0}(z_{0})\right)^{1-2/N}}{\left(2^{*}\int F_{\omega}(z_{0})\right)^{1-2/N}}=m_{0}\cdot\left(\frac{\int F_{0}(z_{0})}{\int F_{\omega}(z_{0})}\right)^{1-2/N}.

By Remark 4, r(z0)Lp+1r(z_{0})\in L^{p+1} and we have

F0(z0)Fω(z0)\displaystyle\frac{\int F_{0}(z_{0})}{\int F_{\omega}(z_{0})} =1p+1r(z0)p+11p+1r(z0)p+1ω2r(z0)2\displaystyle=\frac{\frac{1}{p+1}\int r(z_{0})^{p+1}}{\frac{1}{p+1}\int r(z_{0})^{p+1}-\frac{\omega}{2}\int r(z_{0})^{2}}
=11(p+12)(r(z0)2r(z0)p+1)ω=1+O(ω),asω0,\displaystyle=\frac{1}{1-\left(\frac{p+1}{2}\right)\left(\frac{\int r(z_{0})^{2}}{\int r(z_{0})^{p+1}}\right)\omega}=1+O(\omega),\quad\text{as}\ \omega\to 0,

which concludes the proof in case N5N\geqslant 5.

For N{3,4}N\in\{3,4\}, we let R>0R>0 and we introduce a cut-off function ηRC0()\eta_{R}\in C_{0}^{\infty}(\mathbb{R}) such that ηR(s)=1\eta_{R}(s)=1 for |s|<R|s|<R, 0<ηR<10<\eta_{R}<1 for R<|s|<2RR<|s|<2R, ηR(s)=0\eta_{R}(s)=0 for |s|>2R|s|>2R, and |ηR(s)|2/R|\eta_{R}^{\prime}(s)|\leqslant 2/R for all ss\in\mathbb{R}. We shall simply write ηRz0\eta_{R}z_{0} for the function xηR(|x|)z0(x)x\mapsto\eta_{R}(|x|)z_{0}(x). We have ηRz0L2(N)\eta_{R}z_{0}\in L^{2}(\mathbb{R}^{N}) and we now use it as a test function for JωJ_{\omega}. As above, we have

mωJω(ηRz0)=|ηRz0|2(2F0(ηRz0))12/N(F0(ηRz0)Fω(ηRz0))12/N.m_{\omega}\leqslant J_{\omega}(\eta_{R}z_{0})=\frac{\int|\nabla\eta_{R}z_{0}|^{2}}{\left(2^{*}\int F_{0}(\eta_{R}z_{0})\right)^{1-2/N}}\cdot\left(\frac{\int F_{0}(\eta_{R}z_{0})}{\int F_{\omega}(\eta_{R}z_{0})}\right)^{1-2/N}.

First note that N|ηRz0|2dx|z0|2dx=m0\int_{\mathbb{R}^{N}}|\nabla\eta_{R}z_{0}|^{2}\,\mathrm{d}x\to\int|\nabla z_{0}|^{2}\,\mathrm{d}x=m_{0} as RR\to\infty by dominated convergence. Next,

Nr(ηRz0(x))p+1dx=Nr(z0(x))p+1dx+NBR(0)(r(ηRz0(x))p+1r(z0(x))p+1)dx.\int_{\mathbb{R}^{N}}r(\eta_{R}z_{0}(x))^{p+1}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}r(z_{0}(x))^{p+1}\,\mathrm{d}x+\int_{\mathbb{R}^{N}\setminus B_{R}(0)}(r(\eta_{R}z_{0}(x))^{p+1}-r(z_{0}(x))^{p+1})\,\mathrm{d}x.

Now, for any xNBR(0)x\in\mathbb{R}^{N}\setminus B_{R}(0), there exists τ(x)(z0(x)(1ηR(|x|))z0(x),z0(x))\tau(x)\in(z_{0}(x)-(1-\eta_{R}(|x|))z_{0}(x),z_{0}(x)) such that

r(z0(x))p+1r(ηRz0(x))p+1=(p+1)r(τx)r(τx)p(1ηR(|x|))z0(x).r(z_{0}(x))^{p+1}-r(\eta_{R}z_{0}(x))^{p+1}=(p+1)r^{\prime}(\tau_{x})r(\tau_{x})^{p}(1-\eta_{R}(|x|))z_{0}(x).

Since z0(x)z_{0}(x) decays like |x|(N2)|x|^{-(N-2)} as |x||x|\to\infty, so does τ(x)\tau(x). Hence,

r(ηRz0(x))p+1r(z0(x))p+1=O(|x|(p+1)(N2)).r(\eta_{R}z_{0}(x))^{p+1}-r(z_{0}(x))^{p+1}=O(|x|^{-(p+1)(N-2)}).

This leads to

Nr(ηRz0(x))p+1dx=Nr(z0(x))p+1dx+O(RN(p+1)(N2)),\int_{\mathbb{R}^{N}}r(\eta_{R}z_{0}(x))^{p+1}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}r(z_{0}(x))^{p+1}\,\mathrm{d}x+O(R^{N-(p+1)(N-2)}), (4.13)

where we observe that N(p+1)(N2)<0N-(p+1)(N-2)<0 since p>N+2N2p>\frac{N+2}{N-2}. As consequence, we have

|ηRz0|2(2F0(ηRz0))12/Nm0,asR.\frac{\int|\nabla\eta_{R}z_{0}|^{2}}{\left(2^{*}\int F_{0}(\eta_{R}z_{0})\right)^{1-2/N}}\to m_{0},\quad\text{as}\ R\to\infty.

Furthermore, by Lemma 4.1,

fN(R):=Nr(ηRz0)2dx={O(R)ifN=3,O(log(R))ifN=4,asR.f_{N}(R):=\int_{\mathbb{R}^{N}}r(\eta_{R}z_{0})^{2}\,\mathrm{d}x=\begin{cases}O(R)&\text{if}\ N=3,\\ O(\log(R))&\text{if}\ N=4,\end{cases}\quad\text{as}\ R\to\infty. (4.14)

For ω,R>0\omega,R>0, by (4.13) and (4.14) we have

F0(ηRz0)Fω(ηRz0)=(1(p+12)(fN(R)r(z0)p+1+O(RN(p+1)(N2)))ω)1.\frac{\int F_{0}(\eta_{R}z_{0})}{\int F_{\omega}(\eta_{R}z_{0})}=\left(1-\left(\frac{p+1}{2}\right)\left(\frac{f_{N}(R)}{\int r(z_{0})^{p+1}+O(R^{N-(p+1)(N-2)})}\right)\omega\right)^{-1}.

We now conclude the proof in the following way.

If N=3N=3, we let R=ω1/2R=\omega^{-1/2} and we have

F0(ηRz0)Fω(ηRz0)=1+O(ωf3(ω1/2))=1+O(ω1/2),asω0.\frac{\int F_{0}(\eta_{R}z_{0})}{\int F_{\omega}(\eta_{R}z_{0})}=1+O(\omega f_{3}(\omega^{-1/2}))=1+O(\omega^{1/2}),\quad\text{as}\ \omega\to 0.

If N=4N=4, we let R=ω1R=\omega^{-1} and we have

F0(ηRz0)Fω(ηRz0)=1+O(ωf4(ω1))=1+O(ωlog(ω1)),asω0,\frac{\int F_{0}(\eta_{R}z_{0})}{\int F_{\omega}(\eta_{R}z_{0})}=1+O(\omega f_{4}(\omega^{-1}))=1+O(\omega\log(\omega^{-1})),\quad\text{as}\ \omega\to 0,

which completes the proof. ∎

Lemma 4.3.

Consider a sequence (ωn)(0,)(\omega_{n})\subset(0,\infty), such that ωn0\omega_{n}\to 0. For all nn\in\mathbb{N}, let zn:=zωnz_{n}:=z_{\omega_{n}} be a minimizer for (2.6) with ω=ωn\omega=\omega_{n}. Then

znL22m0=z0L22,ωnr(zn)L220,r(zn)Lp+1p+1p+12.\|\nabla z_{n}\|_{L^{2}}^{2}\to m_{0}=\|\nabla z_{0}\|_{L^{2}}^{2},\quad\omega_{n}\|r(z_{n})\|_{L^{2}}^{2}\to 0,\quad\|r(z_{n})\|_{L^{p+1}}^{p+1}\to\frac{p+1}{2^{*}}. (4.15)
Proof.

Since znL22=mωn\|\nabla z_{n}\|_{L^{2}}^{2}=m_{\omega_{n}}, the first limit follows directly from Lemma 4.2. Furthermore,

12=NFωn(zn)dx=1p+1Nr(zn)p+1dxωn2Nr(zn)2dx=NF0(zn)dxωn2Nr(zn)2dx.\frac{1}{2^{*}}=\int_{\mathbb{R}^{N}}F_{\omega_{n}}(z_{n})\,\mathrm{d}x=\frac{1}{p+1}\int_{\mathbb{R}^{N}}r(z_{n})^{p+1}\,\mathrm{d}x-\frac{\omega_{n}}{2}\int_{\mathbb{R}^{N}}r(z_{n})^{2}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}F_{0}(z_{n})\,\mathrm{d}x-\frac{\omega_{n}}{2}\int_{\mathbb{R}^{N}}r(z_{n})^{2}\,\mathrm{d}x. (4.16)

Hence,

J0(zn)=|zn|2(2F0(zn))12/N=|zn|2(1+22ωnr(zn)2)12/N=mωn(1+22ωnr(zn)2)12/N.J_{0}(z_{n})=\frac{\int|\nabla z_{n}|^{2}}{\left(2^{*}\int F_{0}(z_{n})\right)^{1-2/N}}=\frac{\int|\nabla z_{n}|^{2}}{\left(1+\frac{2^{*}}{2}\omega_{n}\int r(z_{n})^{2}\right)^{1-2/N}}=\frac{m_{\omega_{n}}}{\left(1+\frac{2^{*}}{2}\omega_{n}\int r(z_{n})^{2}\right)^{1-2/N}}.

Now, suppose by contradiction that lim supωnr(zn)L22>0\limsup\omega_{n}\|r(z_{n})\|_{L^{2}}^{2}>0. Then, using Lemma 4.2, there is a subsequence along which, for nn large enough,

m0J0(zn)=mωn(1+22ωnr(zn)2)12/Nm0(1+o(1))(1+22ωnr(zn)2)12/N<m0.m_{0}\leqslant J_{0}(z_{n})=\frac{m_{\omega_{n}}}{\left(1+\frac{2^{*}}{2}\omega_{n}\int r(z_{n})^{2}\right)^{1-2/N}}\leqslant\frac{m_{0}(1+o(1))}{\left(1+\frac{2^{*}}{2}\omega_{n}\int r(z_{n})^{2}\right)^{1-2/N}}<m_{0}.

This contradiction proves that, indeed, ωnr(zn)L220\omega_{n}\|r(z_{n})\|_{L^{2}}^{2}\to 0 as nn\to\infty. Finally, the third limit follows directly from (4.16). ∎

The next two lemmas provide classical results that are crucial in our analysis, and which will be proved in the Appendix for completeness.

Lemma 4.4 (Radial Lemma).

Let s1s\geqslant 1 and uLs(N)u\in L^{s}(\mathbb{R}^{N}) be a radial nonincreasing function. Then,

x0,|u(x)|CN,suLs|x|N/s,\forall x\neq 0,\quad|u(x)|\leqslant C_{N,s}\|u\|_{L^{s}}|x|^{-N/s}, (4.17)

where CN,s=(N/|𝕊N1|)1/sC_{N,s}=(N/|\mathbb{S}^{N-1}|)^{1/s}.

Let {un}H˙1(N)\{u_{n}\}\subset\dot{H}^{1}(\mathbb{R}^{N}) be a sequence of radial nonincreasing functions, such that

supnunH˙1<.\sup_{n\in\mathbb{N}}\|u_{n}\|_{\dot{H}^{1}}<\infty.

Then there exists uH˙1(N)u\in\dot{H}^{1}(\mathbb{R}^{N}) such that, up to a subsequence:

R>0,q>2,unuinL(NBR(0))Lq(NBR(0)).\forall R>0,\ \forall q>2^{*},\quad u_{n}\to u\ \text{in}\ L^{\infty}(\mathbb{R}^{N}\setminus B_{R}(0))\cap L^{q}(\mathbb{R}^{N}\setminus B_{R}(0)). (4.18)
Lemma 4.5.

Let N+2N2p<3N+2N2\frac{N+2}{N-2}\leqslant p<\frac{3N+2}{N-2}, N3N\geqslant 3. Let zωH˙1(N)z_{\omega}\in\dot{H}^{1}(\mathbb{R}^{N}) be the minimizer obtained above, which solves the Euler-Lagrange equation

Δzω=mωfω(zω).-\Delta z_{\omega}=m_{\omega}f_{\omega}(z_{\omega}). (4.19)

There exists η0>0\eta_{0}>0 and M0>0M_{0}>0 such that

sup0<ω<η0zωLM0.\sup_{0<\omega<\eta_{0}}\|z_{\omega}\|_{L^{\infty}}\leqslant M_{0}. (4.20)

There exists η>0\eta>0 such that for all s2s\geqslant 2^{*} there exists a constant KN,s>0K_{N,s}>0 such that

ω(0,η),zωLs(N)KN,s.\forall\omega\in(0,\eta),\quad\|z_{\omega}\|_{L^{s}(\mathbb{R}^{N})}\leqslant K_{N,s}. (4.21)

Note that the result in (ii) is trivial for s=2s=2^{*} since H˙1(N)L2(N)\dot{H}^{1}(\mathbb{R}^{N})\hookrightarrow L^{2^{*}}(\mathbb{R}^{N}). For s>2s>2^{*}, the proof follows a Moser-type iteration argument, which will be given in the Appendix.

We are now in a position to prove Proposition 4.3.

Proof of Proposition 4.3.

Suppose ωn0\omega_{n}\to 0 and let znz_{n} be defined as in Lemma 4.3. By (4.15), {zn}\{z_{n}\} is bounded in H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}). Hence, there exists zH˙1(N)z_{*}\in\dot{H}^{1}(\mathbb{R}^{N}) and a subsequence of {zn}\{z_{n}\} (still denoted by {zn}\{z_{n}\}) such that

znzweakly inH1˙andznza.e. inN.z_{n}\rightharpoonup z_{*}\quad\text{weakly in}\ \dot{H^{1}}\quad\text{and}\quad z_{n}\to z_{*}\quad\text{a.e.~{}in}\ \mathbb{R}^{N}.

Furthermore, by Lemma 4.2 and Lemma 4.5 (ii), for all s2s\geqslant 2^{*} there exists a constant Cs>0C_{s}>0 such that

znLs(N)Cs,n.\|z_{n}\|_{L^{s}(\mathbb{R}^{N})}\leqslant C_{s},\quad\forall n\in\mathbb{N}. (4.22)

Let q>2q>2^{*}. We will first show that, up to a further subsequence,

znzinLq(N).z_{n}\to z_{*}\quad\text{in}\ L^{q}(\mathbb{R}^{N}). (4.23)

Let wn:=znzw_{n}:=z_{n}-z_{*}. By the Radial Lemma, we already know that, up to a subsequence,

wn0inLq(NB1(0)).w_{n}\to 0\quad\text{in}\ L^{q}(\mathbb{R}^{N}\setminus B_{1}(0)). (4.24)

We now fix s>qs>q. Since {zn}\{z_{n}\} is bounded in Ls(N)L^{s}(\mathbb{R}^{N}), there exists zLs(N)z_{**}\in L^{s}(\mathbb{R}^{N}) such that, up to a subsequence, znzz_{n}\rightharpoonup z_{**} weakly in LsL^{s} and znzz_{n}\to z_{**} a.e. in N\mathbb{R}^{N}. But then z=zz_{**}=z_{*} a.e. and we conclude that zLs(N)z_{*}\in L^{s}(\mathbb{R}^{N}). As a consequence, by (4.17) and (4.22), there is a constant C>0C>0 such that

n,x0,|wn(x)|Cs,NznzLs|x|N/sC|x|N/s.\forall n\in\mathbb{N},\ \forall x\neq 0,\quad|w_{n}(x)|\leqslant C_{s,N}\|z_{n}-z_{*}\|_{L^{s}}|x|^{-N/s}\leqslant C|x|^{-N/s}.

Hence, |wn(x)|qCq|x|Nqs|w_{n}(x)|^{q}\leqslant C^{q}|x|^{-N\frac{q}{s}} for xB1(0){0}x\in B_{1}(0)\setminus\{0\}. Since |x|NqsL1(B1(0))|x|^{-N\frac{q}{s}}\in L^{1}(B_{1}(0)) and |wn|q0|w_{n}|^{q}\to 0 a.e. in B1(0)B_{1}(0), it follows by dominated convergence that

wn0inLq(B1(0)).w_{n}\to 0\quad\text{in}\ L^{q}(B_{1}(0)). (4.25)

Thus, (4.23) follows from (4.24) and (4.25).

Next, using (4.23) with q=p+1q=p+1, we have

r(z)r(zn)Lp+1zznLp+10,\|r(z_{*})-r(z_{n})\|_{L^{p+1}}\leqslant\|z_{*}-z_{n}\|_{L^{p+1}}\to 0,

so that

r(z)Lp+1=limnr(zn)Lp+1.\|r(z_{*})\|_{L^{p+1}}=\lim_{n\to\infty}\|r(z_{n})\|_{L^{p+1}}.

Hence, using the constraint Fωn(zn)=12\int F_{\omega_{n}}(z_{n})=\frac{1}{2^{*}} for all nn\in\mathbb{N}, it follows by Lemma 4.3 that

NF0(z)dx=1p+1r(z)Lp+1p+1=1p+1limnr(zn)Lp+1p+1=limnNFωn(zn)dx=12.\int_{\mathbb{R}^{N}}F_{0}(z_{*})\,\mathrm{d}x=\frac{1}{p+1}\|r(z_{*})\|_{L^{p+1}}^{p+1}=\frac{1}{p+1}\lim_{n\to\infty}\|r(z_{n})\|_{L^{p+1}}^{p+1}=\lim_{n\to\infty}\int_{\mathbb{R}^{N}}F_{\omega_{n}}(z_{n})\,\mathrm{d}x=\frac{1}{2^{*}}. (4.26)

We deduce that z0z_{*}\neq 0 and zz_{*} satisfies the constraint in (4.7).

Furthermore, by weak lower semi-continuity of zzL2z\mapsto\|\nabla z\|_{L^{2}} on H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}),

m0zL22lim infnznL22=limnmωn=m0.m_{0}\leqslant\|\nabla z_{*}\|_{L^{2}}^{2}\leqslant\liminf_{n\to\infty}\|\nabla z_{n}\|_{L^{2}}^{2}=\lim_{n\to\infty}m_{\omega_{n}}=m_{0}.

Hence, zz_{*} is a minimizer for (4.7). It follows that zz_{*} is a positive solution of (4.8) with θ0=m0\theta_{0}=m_{0}, and so z=z0z_{*}=z_{0} by Theorem 4.1.

We now prove that znz0z_{n}\to z_{0} in H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}). We have

znz0L22=znL22+z0L222Nznz0dx.\|\nabla z_{n}-\nabla z_{0}\|_{L^{2}}^{2}=\|\nabla z_{n}\|_{L^{2}}^{2}+\|\nabla z_{0}\|_{L^{2}}^{2}-2\int_{\mathbb{R}^{N}}\nabla z_{n}\cdot\nabla z_{0}\,\mathrm{d}x.

On the one hand,

znL22=mωnm0andz0L22=m0.\|\nabla z_{n}\|_{L^{2}}^{2}=m_{\omega_{n}}\to m_{0}\quad\text{and}\quad\|\nabla z_{0}\|_{L^{2}}^{2}=m_{0}.

On the other, using (4.8) we find

Nznz0dx\displaystyle\int_{\mathbb{R}^{N}}\nabla z_{n}\cdot\nabla z_{0}\,\mathrm{d}x =NznΔz0dx\displaystyle=-\int_{\mathbb{R}^{N}}z_{n}\Delta z_{0}\,\mathrm{d}x
=Nz0Δz0dxNΔz0(znz0)dx\displaystyle=-\int_{\mathbb{R}^{N}}z_{0}\Delta z_{0}\,\mathrm{d}x-\int_{\mathbb{R}^{N}}\Delta z_{0}(z_{n}-z_{0})\,\mathrm{d}x
=N|z0|2dx+m0Nf0(z0)(znz0)dx\displaystyle=\int_{\mathbb{R}^{N}}|\nabla z_{0}|^{2}\,\mathrm{d}x+m_{0}\int_{\mathbb{R}^{N}}f_{0}(z_{0})(z_{n}-z_{0})\,\mathrm{d}x
=m0+m0Nf0(z0)(znz0)dx.\displaystyle=m_{0}+m_{0}\int_{\mathbb{R}^{N}}f_{0}(z_{0})(z_{n}-z_{0})\,\mathrm{d}x.

By Hölder’s inequality, we have

|Nf0(z0)(znz0)dx|\displaystyle\left|\int_{\mathbb{R}^{N}}f_{0}(z_{0})(z_{n}-z_{0})\,\mathrm{d}x\right| Nr(z0)p|znz0|dx\displaystyle\leqslant\int_{\mathbb{R}^{N}}r(z_{0})^{p}|z_{n}-z_{0}|\,\mathrm{d}x
z0Lp+1pznz0Lp+10.\displaystyle\leqslant\|z_{0}\|_{L^{p+1}}^{p}\|z_{n}-z_{0}\|_{L^{p+1}}\to 0.

Therefore,

znz0L22=znL22+z0L222Nznz0dx0.\|\nabla z_{n}-\nabla z_{0}\|_{L^{2}}^{2}=\|\nabla z_{n}\|_{L^{2}}^{2}+\|\nabla z_{0}\|_{L^{2}}^{2}-2\int_{\mathbb{R}^{N}}\nabla z_{n}\cdot\nabla z_{0}\,\mathrm{d}x\to 0.

We conclude that

znz0inH˙1(N)Lq(N),q2.z_{n}\to z_{0}\quad\text{in}\quad\dot{H}^{1}(\mathbb{R}^{N})\cap L^{q}(\mathbb{R}^{N}),\quad\forall q\geqslant 2^{*}.

Next, by Lemma 4.5 (i), the sequence {zn}\{z_{n}\} is bounded in L(N)L^{\infty}(\mathbb{R}^{N}). It then follows by standard elliptic theory arguments (see e.g. the proof of Proposition 4.14 below) that znz0z_{n}\to z_{0} in C2(N)C^{2}(\mathbb{R}^{N}).

To complete the proof, we now turn to {vωn}\{v_{\omega_{n}}\}. Observing that

vωn(x)=zn(mωn1/2x)andv0(x)=z0(m01/2x),v_{\omega_{n}}(x)=z_{n}(m_{\omega_{n}}^{-1/2}x)\quad\text{and}\quad v_{0}(x)=z_{0}(m_{0}^{-1/2}x),

we deduce from the conclusions obtained for the sequence {zn}\{z_{n}\} that, up to a subsequence,

vωnv0inH˙1(N)Lq(N)C2(N),q2.v_{\omega_{n}}\to v_{0}\quad\text{in}\quad\dot{H}^{1}(\mathbb{R}^{N})\cap L^{q}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}),\quad\forall q\geqslant 2^{*}.

Finally, since the only possible limit point is v0v_{0}, a proof by contradiction shows that, in fact, the whole sequence {vωn}\{v_{\omega_{n}}\} converges to v0v_{0}. ∎

Going back to the variable uω=r(vω)u_{\omega}=r(v_{\omega}), we obtain the following.

Proposition 4.4.

Suppose N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Consider a sequence (ωn)(0,)(\omega_{n})\subset(0,\infty), such that ωn0\omega_{n}\to 0. Then

uωnu0=r(v0)inH˙1(N)Lq(N)C2(N),q2,u_{\omega_{n}}\to u_{0}=r(v_{0})\quad\text{in}\quad\dot{H}^{1}(\mathbb{R}^{N})\cap L^{q}(\mathbb{R}^{N})\cap C^{2}(\mathbb{R}^{N}),\quad\forall q\geqslant 2^{*}, (4.27)

and

ωnuωnL220.\omega_{n}\|u_{\omega_{n}}\|_{L^{2}}^{2}\to 0. (4.28)
Proof.

First, (4.28) is a direct reformulation of (4.11). Next, in (4.27), the LqL^{q}-convergence of uωnu_{\omega_{n}} follows directly from the LqL^{q}-convergence of vωnv_{\omega_{n}} in (4.10) and the fact that rr is Lipschitz. The H˙1\dot{H}^{1}-convergence follows from the H˙1\dot{H}^{1}-convergence of vωnv_{\omega_{n}} by dominated convergence. Finally, C2C^{2}-convergence follows from the C2C^{2}-convergence of vωnv_{\omega_{n}}, using the LL^{\infty}-bound on vωnv_{\omega_{n}} (Lemma 4.5 (i)) and the fact that rr, rr^{\prime} and r′′r^{\prime\prime} are Lipschitz on compact subsets of \mathbb{R}. ∎

4.2.1. Asymptotic behavior of M(ω)=uωL22M(\omega)=\|u_{\omega}\|_{L^{2}}^{2} as ω0+\omega\to 0^{+}

Proposition 4.5.

Let N3N\geqslant 3 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. As ω0+\omega\to 0^{+}, we have

limω0uωL2={+if N{3,4},u0L2if N5.\lim_{\omega\to 0}\|u_{\omega}\|_{L^{2}}=\begin{cases}+\infty&\text{if }N\in\{3,4\},\\ \|u_{0}\|_{L^{2}}&\text{if }N\geqslant 5.\end{cases}

More precisely,

vωv0 and uωu0 in Lq(N)v_{\omega}\to v_{0}\text{ and }u_{\omega}\to u_{0}\text{ in }L^{q}(\mathbb{R}^{N}) (4.29)

for all q>NN2q>\frac{N}{N-2}.
Moreover, letting ηω=r(vω)r(vω)\eta_{\omega}=r(v_{\omega})r^{\prime}(v_{\omega}) for any ω>0\omega>0, there holds

ηωη0\eta_{\omega}\to\eta_{0} (4.30)

in Lq(N)L^{q}(\mathbb{R}^{N}) for all q>NN2q>\frac{N}{N-2}.

Remark 5.

The strong convergence of vωv_{\omega}, uωu_{\omega} and ηω\eta_{\omega} holds in Lq(N)L^{q}(\mathbb{R}^{N}) for all q>NN2q>\frac{N}{N-2}. This includes L2(N)L^{2}(\mathbb{R}^{N}) only in dimensions N5N\geqslant 5.

Proof.

Let us start with the case N{3,4}N\in\{3,4\}. First, we show that vωL2\|v_{\omega}\|_{L^{2}}\to\infty. Indeed, suppose by contradiction there is a sequence ωn0\omega_{n}\to 0 such that vnL2\|v_{n}\|_{L^{2}} is bounded, where vn:=vωnv_{n}:=v_{\omega_{n}}. Then using Proposition 4.3 {vn}\{v_{n}\} is bounded in H1H^{1} and (using Rellich-Kondrachov and Radial Lemma) up to a subsequence, vnv0v_{n}\to v_{0} in L2L^{2}, hence v0L2v_{0}\in L^{2}, a contradiction. As a consequence, vωL2\|v_{\omega}\|_{L^{2}}\to\infty as ω0\omega\to 0.

Next, by Lemma 4.5 (i), there exist η0,M0>0\eta_{0},M_{0}>0 such that vωLM0\|v_{\omega}\|_{L^{\infty}}\leqslant M_{0} for all 0<ω<η00<\omega<\eta_{0}. Then, by Lemma 2.1, for all 0<ω<η00<\omega<\eta_{0},

r(vω)2vω2M02r(v_{\omega})^{2}\leqslant v_{\omega}^{2}\leqslant M_{0}^{2}

and

uωL22=N|r(vω)|2dxNvω21+2δr(vω)2dx11+2δM02Nvω2dx.\|u_{\omega}\|_{L^{2}}^{2}=\int_{\mathbb{R}^{N}}|r(v_{\omega})|^{2}\,\mathrm{d}x\geqslant\int_{\mathbb{R}^{N}}\frac{v_{\omega}^{2}}{1+2\delta r(v_{\omega})^{2}}\,\mathrm{d}x\geqslant\frac{1}{1+2\delta M_{0}^{2}}\int_{\mathbb{R}^{N}}v_{\omega}^{2}\,\mathrm{d}x.

This implies, uωL2\|u_{\omega}\|_{L^{2}}\to\infty as ω0\omega\to 0.

Since vωv0v_{\omega}\to v_{0} a.e. and {vω}\{v_{\omega}\} is bounded in LL^{\infty}, dominated convergence yields vωv0Lq(BR(0))0\|v_{\omega}-v_{0}\|_{L^{q}(B_{R}(0))}\to 0 as ω0\omega\to 0, for any R>0R>0 and q>NN2q>\frac{N}{N-2}. Furthermore, a maximum principle argument as in [32] allows one to improve the classical bound

vω(x)C0|x|N22(|x|1)v_{\omega}(x)\leqslant\frac{C_{0}}{|x|^{\frac{N-2}{2}}}\quad(|x|\geqslant 1)

to

vω(x)Cε|x|N2ε(|x|Rε)v_{\omega}(x)\leqslant\frac{C_{\varepsilon}}{|x|^{N-2-\varepsilon}}\quad(|x|\geqslant R_{\varepsilon})

with Rε,CεR_{\varepsilon},C_{\varepsilon} independent of ω\omega. This provides an upper bound in Lq(|x|Rε)L^{q}(|x|\geqslant R_{\varepsilon}), for q(N2)>Nq(N-2)>N, if ε>0\varepsilon>0 is chosen small enough, and it follows by dominated convergence that vωv0Lq(NBRε(0))0\|v_{\omega}-v_{0}\|_{L^{q}(\mathbb{R}^{N}\setminus B_{R_{\varepsilon}}(0))}\to 0 as ω0\omega\to 0.

Next, by Lemma 2.1,

uωu0Lq=r(vω)r(v0)Lqvωv0Lq0,ω0.\|u_{\omega}-u_{0}\|_{L^{q}}=\|r(v_{\omega})-r(v_{0})\|_{L^{q}}\leqslant\|v_{\omega}-v_{0}\|_{L^{q}}\to 0,\quad\omega\to 0.

Finally, let q>NN2q>\frac{N}{N-2} and ηω=r(vω)r(vω)\eta_{\omega}=r(v_{\omega})r^{\prime}(v_{\omega}),

ηωη0Lqq\displaystyle\|\eta_{\omega}-\eta_{0}\|_{L^{q}}^{q} =N|r(vω)r(vω)r(v0)r(v0)|qdx\displaystyle=\int_{\mathbb{R}^{N}}|r(v_{\omega})r^{\prime}(v_{\omega})-r(v_{0})r^{\prime}(v_{0})|^{q}\,\mathrm{d}x
N|r(vω)|q|r(vω)r(v0)|qdx+N|r(v0)|q|r(vω)r(v0)|qdx\displaystyle\leqslant\int_{\mathbb{R}^{N}}|r^{\prime}(v_{\omega})|^{q}|r(v_{\omega})-r(v_{0})|^{q}\,\mathrm{d}x+\int_{\mathbb{R}^{N}}|r(v_{0})|^{q}|r^{\prime}(v_{\omega})-r^{\prime}(v_{0})|^{q}\,\mathrm{d}x
N|r(vω)r(v0)|qdx+N|r(v0)|q|r(vω)r(v0)|qdx.\displaystyle\leqslant\int_{\mathbb{R}^{N}}|r(v_{\omega})-r(v_{0})|^{q}\,\mathrm{d}x+\int_{\mathbb{R}^{N}}|r(v_{0})|^{q}|r^{\prime}(v_{\omega})-r^{\prime}(v_{0})|^{q}\,\mathrm{d}x.

The first term of the right-hand side of this inequality is simply uωu0Lqq\|u_{\omega}-u_{0}\|_{L^{q}}^{q} and goes to zero as ω0\omega\to 0. Since vωv0v_{\omega}\to v_{0} a.e. and rC1()r\in C^{1}(\mathbb{R}), the second term also goes to zero by dominated convergence. We conclude that ηωη0Lq0\|\eta_{\omega}-\eta_{0}\|_{L^{q}}\to 0 as ω0\omega\to 0. ∎

4.2.2. Asymptotic behavior of M(ω)M^{\prime}(\omega) as ω0\omega\to 0.

Following ideas from [32] (see also [28]), we derive an upper bound on M(ω)M^{\prime}(\omega).

Proposition 4.6.

Let N3N\geqslant 3 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Then, for ω>0\omega>0 small enough, we have

M(ω)2\displaystyle\frac{M^{\prime}(\omega)}{2} N2N+2[(p1p+1)2(p3N+2N2)β2(ω)p1p+1(p3)β(ω)4N]\displaystyle\,\frac{N-2}{N+2}\left[-\left(\frac{p-1}{p+1}\right)^{2}\left(p-\frac{3N+2}{N-2}\right)\beta^{2}(\omega)-\frac{p-1}{p+1}(p-3)\beta(\omega)-\frac{4}{N}\right]
<M2(ω)2T(ω)[p1p+1(N+2N2(p1))β(ω)2]\displaystyle<\frac{M^{2}(\omega)}{2T(\omega)}\left[\frac{p-1}{p+1}\left(N+2-\frac{N}{2}(p-1)\right)\beta(\omega)-2\right] (4.31)

where

T(ω)=N|uω(x)|2dx,β(ω)=T(ω)1RNuω(x)p+1dx.T(\omega)=\int_{\mathbb{R}^{N}}|\nabla u_{\omega}(x)|^{2}\,\mathrm{d}x,\quad\beta(\omega)=T(\omega)^{-1}\int_{R^{N}}u_{\omega}(x)^{p+1}\,\mathrm{d}x.

Moreover,

  1. (i)

    if N{3,4}N\in\{3,4\}, then

    limω0+M(ω)=;\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=-\infty;
  2. (ii)

    if N=5N=5, then

    M(ω)<0M^{\prime}(\omega)<0

    for all ω\omega small enough;

  3. (iii)

    if N6N\geqslant 6 and

    p<2(N+2)N116N+2N2(N2) or 2(N+2)N+116N+2N2(N2)<pp<\frac{2(N+2)}{N}-\sqrt{1-16\frac{N+2}{N^{2}(N-2)}}\ \text{ or }\ \frac{2(N+2)}{N}+\sqrt{1-16\frac{N+2}{N^{2}(N-2)}}<p (4.32)

    then

    M(ω)<0M^{\prime}(\omega)<0

    for all ω\omega small enough.

Remark 6.

For N6N\geqslant 6, the condition (4.32) is probably not optimal. As an illustration, for N=7N=7, (4.32) is satisfied whenever 1.8<p1.921.8<p\leqslant 1.92 or 3.22p<4.63.22\leqslant p<4.6. For N=8N=8, we need 53<p1.73\tfrac{5}{3}<p\leqslant 1.73 or 3.27p<1333.27\leqslant p<\tfrac{13}{3}.

Proof.

For any ω>0\omega>0, let (ω)=L+\mathcal{L}(\omega)=L_{+} be defined by (1.9). We know from Proposition 3.1 that, for any ω>0\omega>0, (ω)\mathcal{L}(\omega) has exactly one negative eigenvalue. Then we define the symmetric matrix L=(Lij)L=(L_{ij}) given by the restriction of (ω)\mathcal{L}(\omega) to the finite dimensional space spanned by {ωuω,uω,xuω+N2uω}\left\{\partial_{\omega}u_{\omega},u_{\omega},x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right\}.

Since

(ω)ωuω=uω,\displaystyle\mathcal{L}(\omega)\partial_{\omega}u_{\omega}=-u_{\omega},
(ω)uω=2δuωΔuω2+(1p)uωp,\displaystyle\mathcal{L}(\omega)u_{\omega}=-2\delta u_{\omega}\Delta u_{\omega}^{2}+(1-p)u_{\omega}^{p},
(ω)(xuω+N2uω)=NδuωΔuω2+(2+N2(1p))uωp2ωuω,\displaystyle\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)=-N\delta u_{\omega}\Delta u_{\omega}^{2}+\left(2+\frac{N}{2}(1-p)\right)u_{\omega}^{p}-2\omega u_{\omega},

straightforward computations give

L11:=\displaystyle L_{11}:= ωuω,(ω)ωuω=M(ω)2,L12:=ωuω,(ω)uω=M(ω),\displaystyle\,\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)\partial_{\omega}u_{\omega}\rangle=-\frac{M^{\prime}(\omega)}{2},\quad L_{12}:=\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)u_{\omega}\rangle=-M(\omega),
L13:=\displaystyle L_{13}:= ωuω,(ω)(xuω+N2uω)=uω,xuω+N2uω=0,\displaystyle\,\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=-\langle u_{\omega},x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\rangle=0,
L22:=\displaystyle L_{22}:= uω,(ω)uω=8δNuω2|uω|2(p1)Nuωp+1,\displaystyle\,\langle u_{\omega},\mathcal{L}(\omega)u_{\omega}\rangle=8\delta\int_{\mathbb{R}^{N}}u^{2}_{\omega}|\nabla u_{\omega}|^{2}-(p-1)\int_{\mathbb{R}^{N}}u_{\omega}^{p+1},
L23:=\displaystyle L_{23}:= uω,(ω)(xuω+N2uω)=4δNNuω2|uω|2+(2+N2(1p))Nuωp+12ωM(ω),\displaystyle\,\langle u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=4\delta N\int_{\mathbb{R}^{N}}u_{\omega}^{2}|\nabla u_{\omega}|^{2}+\left(2+\frac{N}{2}(1-p)\right)\int_{\mathbb{R}^{N}}u_{\omega}^{p+1}-2\omega M(\omega),
L33:=\displaystyle L_{33}:= xuω+N2uω,(ω)(xuω+N2uω)=2δN(1+N2)Nuω2|uω|2\displaystyle\,\langle x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=2\delta N\left(1+\frac{N}{2}\right)\int_{\mathbb{R}^{N}}u_{\omega}^{2}|\nabla u_{\omega}|^{2}
+N2p1p+1(2+N2(1p))Nuωp+1.\displaystyle+\frac{N}{2}\frac{p-1}{p+1}\left(2+\frac{N}{2}(1-p)\right)\int_{\mathbb{R}^{N}}u_{\omega}^{p+1}.

Next, for any ω>0\omega>0, let Q(ω)=nuω(x)2|uω(x)|2dxQ(\omega)=\int_{\mathbb{R}^{n}}u_{\omega}(x)^{2}|\nabla u_{\omega}(x)|^{2}\,\mathrm{d}x. As a consequence of Proposition 1.1, we can write

T(ω)+2δQ(ω)=2(1p+1T(ω)β(ω)ω2M(ω)),\displaystyle T(\omega)+2\delta Q(\omega)=2^{*}\left(\frac{1}{p+1}T(\omega)\beta(\omega)-\frac{\omega}{2}M(\omega)\right),
T(ω)+4δQ(ω)=T(ω)β(ω)ωM(ω),\displaystyle T(\omega)+4\delta Q(\omega)=T(\omega)\beta(\omega)-\omega M(\omega),

with 2=2NN22^{*}=\frac{2N}{N-2}, and deduce that

{2δQ(ω)=(12p+1)T(ω)β(ω)+2N2ωM(ω),ωM(ω)=N2N+2(22p+11)T(ω)β(ω)N2N+2T(ω).\begin{cases}&2\delta Q(\omega)=\left(1-\frac{2^{*}}{p+1}\right)T(\omega)\beta(\omega)+\frac{2}{N-2}\omega M(\omega),\\ &\omega M(\omega)=\frac{N-2}{N+2}\left(2\frac{2^{*}}{p+1}-1\right)T(\omega)\beta(\omega)-\frac{N-2}{N+2}T(\omega).\end{cases} (4.33)

As a consequence,

2δQ(ω)=NN+2p1p+1T(ω)β(ω)2N+2T(ω)2\delta Q(\omega)=\frac{N}{N+2}\frac{p-1}{p+1}T(\omega)\beta(\omega)-\frac{2}{N+2}T(\omega)

and

L22=\displaystyle L_{22}= T(ω)[p1p+1(2N2N+2(p1))β(ω)8N+2],\displaystyle\,T(\omega)\left[\frac{p-1}{p+1}\left(2\frac{N-2}{N+2}-(p-1)\right)\beta(\omega)-\frac{8}{N+2}\right],
L23=\displaystyle L_{23}= T(ω)[N2p1p+1(3p)β(ω)2],\displaystyle\,T(\omega)\left[\frac{N}{2}\frac{p-1}{p+1}(3-p)\beta(\omega)-2\right],
L33=\displaystyle L_{33}= T(ω)[N2p1p+1(N+2N2(p1))β(ω)N].\displaystyle\,T(\omega)\left[\frac{N}{2}\frac{p-1}{p+1}\left(N+2-\frac{N}{2}(p-1)\right)\beta(\omega)-N\right].

Moreover,

(3N+2N2p)β(ω)p+1=1+N+2N2ωM(ω)T(ω).\left(\frac{3N+2}{N-2}-p\right)\frac{\beta(\omega)}{p+1}=1+\frac{N+2}{N-2}\frac{\omega M(\omega)}{T(\omega)}. (4.34)

In particular, using N+2N2<p<3N+2N+2\frac{N+2}{N-2}<p<\frac{3N+2}{N+2}, this leads to

1=limω0+(3N+2N2p)β(ω)p+11=\lim_{\omega\to 0^{+}}\left(\frac{3N+2}{N-2}-p\right)\frac{\beta(\omega)}{p+1} (4.35)

since limω0+ωM(ω)=0\lim_{\omega\to 0^{+}}\omega M(\omega)=0.

Now, a tedious but straightforward computation gives

L22L33\displaystyle L_{22}L_{33} L232=N(N2)N+2T2(ω)[(p1p+1)2(p3N+2N2)β2(ω)+p1p+1(p3)β(ω)+4N]\displaystyle-L^{2}_{23}=\frac{N(N-2)}{N+2}T^{2}(\omega)\left[\left(\frac{p-1}{p+1}\right)^{2}\left(p-\frac{3N+2}{N-2}\right)\beta^{2}(\omega)+\frac{p-1}{p+1}(p-3)\beta(\omega)+\frac{4}{N}\right]

and, using again N+2N2<p<3N+2N+2\frac{N+2}{N-2}<p<\frac{3N+2}{N+2}, we obtain from (4.35) that

L22L33\displaystyle L_{22}L_{33}- L232\displaystyle\,L^{2}_{23}
ω0\displaystyle\underset{\omega\to 0}{\sim} N(N2)N+2T2(0)(3N+2N2p)1[(p1)2+(p1)(p3)+4N(3N+2N2p)]\displaystyle\frac{N(N-2)}{N+2}T^{2}(0)\left(\frac{3N+2}{N-2}-p\right)^{-1}\left[-(p-1)^{2}+(p-1)(p-3)+\frac{4}{N}\left(\frac{3N+2}{N-2}-p\right)\right]
=2N(N2)N+2T2(0)(3N+2N2p)1[(p1)+2N(3N+2N2p)]\displaystyle=\frac{2N(N-2)}{N+2}T^{2}(0)\left(\frac{3N+2}{N-2}-p\right)^{-1}\left[-(p-1)+\frac{2}{N}\left(\frac{3N+2}{N-2}-p\right)\right]
=2(N2)T2(0)(3N+2N2p)1(N+2N2p)<0.\displaystyle=2(N-2)T^{2}(0)\left(\frac{3N+2}{N-2}-p\right)^{-1}\left(\frac{N+2}{N-2}-p\right)<0.

Since (ω)\mathcal{L}(\omega) has a unique negative eigenvalue, we deduce that the determinant of LL is negative:

0>det(L)=M(ω)2(L232L22L33)M(ω)2T(ω)[N2p1p+1(N+2N2(p1))β(ω)N].\displaystyle 0>\det(L)=\frac{M^{\prime}(\omega)}{2}(L^{2}_{23}-L_{22}L_{33})-M(\omega)^{2}T(\omega)\left[\frac{N}{2}\frac{p-1}{p+1}\left(N+2-\frac{N}{2}(p-1)\right)\beta(\omega)-N\right].

This gives the estimate (4.6).

Using again (4.35), we have

[p1p+1(N+2N2(p1))β(ω)2]ω0(3N+2N2p)1C(p)\displaystyle\left[\frac{p-1}{p+1}\left(N+2-\frac{N}{2}(p-1)\right)\beta(\omega)-2\right]\underset{\omega\to 0}{\sim}\left(\frac{3N+2}{N-2}-p\right)^{-1}C(p)

with

C(p)=(p1)(N+2N2(p1))2(3N+2N2p)=N2p2+2(N+2)p3N2+10N2(N2).\displaystyle C(p)=(p-1)\left(N+2-\frac{N}{2}(p-1)\right)-2\left(\frac{3N+2}{N-2}-p\right)=-\frac{N}{2}p^{2}+2(N+2)p-\frac{3N^{2}+10N}{2(N-2)}.

The function C(p)C(p) is a second order polynomial and its maximum on (1,+)\left(1,+\infty\right) is reached at p(N)=2(N+2)Np_{*}(N)=\frac{2(N+2)}{N}. In particular,

C(p(N))=N32N216N322N(N2)<0\displaystyle C(p_{*}(N))=\frac{N^{3}-2N^{2}-16N-32}{2N(N-2)}<0

for N{3,4,5}N\in\{3,4,5\}. This, together with Proposition 4.5, concludes the proof for N{3,4,5}N\in\{3,4,5\}.

For N6N\geqslant 6, C(p(N))>0C(p_{*}(N))>0 and C(p)C(p) vanishes twice in (1,+)\left(1,+\infty\right) at

p(N)=p(N)2NC(p(N)) and p+(N)=p(N)+2NC(p(N)).p_{-}(N)=p_{*}(N)-\sqrt{\frac{2}{N}C(p_{*}(N))}\text{ and }p_{+}(N)=p_{*}(N)+\sqrt{\frac{2}{N}C(p_{*}(N))}.

Note that p(N)>N+2N2p_{-}(N)>\frac{N+2}{N-2} and p+(N)<3N+2N2p_{+}(N)<\frac{3N+2}{N-2}, so that C(p)<0C(p)<0 for p(N+2N2,p(N))p\in\left(\frac{N+2}{N-2},p_{-}(N)\right) or p(p+(N),3N+2N2)p\in\left(p_{+}(N),\frac{3N+2}{N-2}\right). As a consequence, for such pp’s, M(ω)<0M^{\prime}(\omega)<0 whenever ω\omega is small enough. ∎

Proposition 4.6 gives a complete description of the asymptotic behavior of M(ω)M^{\prime}(\omega) for N{3,4}N\in\{3,4\}. We conclude this section with the study of M(ω)M^{\prime}(\omega) for N5N\geqslant 5. We know from above that uωu0u_{\omega}\to u_{0} in H˙1(N)L(N)\dot{H}^{1}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N}), u0u_{0} being the unique positive radial-decreasing solution to (1.20). More precisely, for any ω>0\omega>0, uω=r(vω)u_{\omega}=r(v_{\omega}) and vωv0v_{\omega}\to v_{0} in H˙1(N)L(N)\dot{H}^{1}(\mathbb{R}^{N})\cap L^{\infty}(\mathbb{R}^{N}) with u0=r(v0)u_{0}=r(v_{0}) and v0v_{0} the unique positive radial-decreasing solution to

Δv=f0(v).-\Delta v=f_{0}(v).

To discuss the asymptotic behavior of M(ω)M^{\prime}(\omega) as ω0\omega\to 0, we proceed as follows. On the one hand,

M(ω)=ωNuω2dx=2Nuωωuωdx=2Nr(vω)r(vω)ωvωdx=2Nηωωvωdx,\displaystyle M^{\prime}(\omega)=\partial_{\omega}\int_{\mathbb{R}^{N}}u_{\omega}^{2}\,\mathrm{d}x=2\int_{\mathbb{R}^{N}}u_{\omega}\partial_{\omega}u_{\omega}\,\mathrm{d}x=2\int_{\mathbb{R}^{N}}r(v_{\omega})r^{\prime}(v_{\omega})\partial_{\omega}v_{\omega}\,\mathrm{d}x=2\int_{\mathbb{R}^{N}}\eta_{\omega}\partial_{\omega}v_{\omega}\,\mathrm{d}x,

with ηω=r(vω)r(vω)\eta_{\omega}=r(v_{\omega})r^{\prime}(v_{\omega}). Now, since vωv_{\omega} is a solution to (Sω\mathrm{S}_{\omega}), we have

(Δfω(vω))ωvω=r(vω)r(vω).(-\Delta-f^{\prime}_{\omega}(v_{\omega}))\partial_{\omega}v_{\omega}=-r(v_{\omega})r^{\prime}(v_{\omega}).

This implies

M(ω)=2Nηω(Lω)rad1ηω\displaystyle M^{\prime}(\omega)=-2\int_{\mathbb{R}^{N}}\eta_{\omega}(L_{\omega})^{-1}_{\mathrm{rad}}\eta_{\omega}

with Lω:=Δfω(vω)L_{\omega}:=-\Delta-f^{\prime}_{\omega}(v_{\omega}).

On the other hand, from Theorem 4.1, we know that the limiting linearized operator

L0:=Δf0(v0)L_{0}:=-\Delta-f^{\prime}_{0}(v_{0})

has a trivial kernel ker(L0)=span{iv0:i=1,,N}\ker(L_{0})=\mathrm{span}\{\partial_{i}v_{0}:\,i=1,\dots,N\} so that (L0)rad1(L_{0})^{-1}_{\mathrm{rad}} can be defined using functional calculus. Moreover, since vωv_{\omega} converges to v0v_{0} in LL^{\infty} as ω0\omega\to 0, we deduce that LωL_{\omega} converges to L0L_{0} in the norm resolvent sense (see [45, Theorem VIII.25]). More precisely, we have the following lemma.

Lemma 4.6.

Let N5N\geqslant 5 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. For any ω0\omega\geqslant 0, define Vω:=fω(vω)V_{\omega}:=-f^{\prime}_{\omega}(v_{\omega}). Then, as ω0\omega\to 0,

VωV0 in L(N), and (Vωω)V0 in Lq(N)V_{\omega}\to V_{0}\text{ in }L^{\infty}(\mathbb{R}^{N}),\text{ and }\ (V_{\omega}-\omega)\to V_{0}\text{ in }L^{q}(\mathbb{R}^{N}) (4.36)

for all q>NN2q>\frac{N}{N-2} if p2p\geqslant 2, for all qN4q\geqslant\frac{N}{4} if p<2p<2.

Proof.

First of all, using fω(s)=f0(s)ωr(s)r(s)f_{\omega}(s)=f_{0}(s)-\omega r^{\prime}(s)r(s) and the properties of r(s)r(s), we have

fω(vω)=f0(vω)ω1(1+2δr2(vω))2f^{\prime}_{\omega}(v_{\omega})=f^{\prime}_{0}(v_{\omega})-\omega\frac{1}{(1+2\delta r^{2}(v_{\omega}))^{2}}

so that

V0Vω=fω(vω)f0(v0)=f0(vω)f0(v0)ω1(1+2δr2(vω))2V_{0}-V_{\omega}=f^{\prime}_{\omega}(v_{\omega})-f^{\prime}_{0}(v_{0})=f^{\prime}_{0}(v_{\omega})-f^{\prime}_{0}(v_{0})-\omega\frac{1}{(1+2\delta r^{2}(v_{\omega}))^{2}}

and

V0(Vωω)=fω(vω)+ωf0(v0)=f0(vω)f0(v0)+ω(11(1+2δr2(vω))2).V_{0}-(V_{\omega}-\omega)=f^{\prime}_{\omega}(v_{\omega})+\omega-f^{\prime}_{0}(v_{0})=f^{\prime}_{0}(v_{\omega})-f^{\prime}_{0}(v_{0})+\omega\left(1-\frac{1}{(1+2\delta r^{2}(v_{\omega}))^{2}}\right).

Next, for any s(0,)s\in(0,\infty), f0(s)f^{\prime}_{0}(s) can be written as

f0(s)=r(s)2hp(s)r(s)p1withhp(s)=p2δr2(s)1+2δr2(s).f^{\prime}_{0}(s)=r^{\prime}(s)^{2}h_{p}(s)r(s)^{p-1}\ \text{with}\ h_{p}(s)=p-\frac{2\delta r^{2}(s)}{1+2\delta r^{2}(s)}.

A direct computation gives, for any t,s(0,)t,s\in(0,\infty),

|r(s)2r(t)2|2δ(|r(s)|+|r(t)|)|st|,\displaystyle|r^{\prime}(s)^{2}-r^{\prime}(t)^{2}|\leqslant 2\delta(|r(s)|+|r(t)|)\,|s-t|,
|hp(s)hp(t)|2δ(|r(s)|+|r(t)|)|st|,\displaystyle|h_{p}(s)-h_{p}(t)|\leqslant 2\delta(|r(s)|+|r(t)|)\,|s-t|,

while

|r(s)p1r(t)p1|{|st|p1 if p<2,(p1)(|r(s)|p2+|r(t)|p2)|st| if p2.|r(s)^{p-1}-r(t)^{p-1}|\leqslant\begin{cases}|s-t|^{p-1}&\text{ if }p<2,\\ (p-1)(|r(s)|^{p-2}+|r(t)|^{p-2})\,|s-t|&\text{ if }p\geqslant 2.\end{cases}

As a consequence, for any xNx\in\mathbb{R}^{N},

|f0(vω(x))f0(v0(x))|{|vω(x)v0(x)|p1 if p<2,|vω(x)v0(x)| if p2.|f^{\prime}_{0}(v_{\omega}(x))-f^{\prime}_{0}(v_{0}(x))|\lesssim\begin{cases}|v_{\omega}(x)-v_{0}(x)|^{p-1}&\text{ if }p<2,\\ |v_{\omega}(x)-v_{0}(x)|&\text{ if }p\geqslant 2.\end{cases} (4.37)

where we used that vωL\|v_{\omega}\|_{L^{\infty}} is uniformly bounded for ω\omega small enough. This, together with the LL^{\infty}-convergence of vωv_{\omega}, implies

f0(vω)f0(v0)L0 as ω0.\|f^{\prime}_{0}(v_{\omega})-f^{\prime}_{0}(v_{0})\|_{L^{\infty}}\to 0\text{ as }\omega\to 0.

Finally,

V0VωLf0(vω)f0(v0)L+ω0 as ω0.\|V_{0}-V_{\omega}\|_{L^{\infty}}\leqslant\|f^{\prime}_{0}(v_{\omega})-f^{\prime}_{0}(v_{0})\|_{L^{\infty}}+\omega\to 0\text{ as }\omega\to 0.

Next, using Proposition 4.5 and (4.37), we deduce that

f0(vω)f0(v0)Lq0 as ω0\|f^{\prime}_{0}(v_{\omega})-f^{\prime}_{0}(v_{0})\|_{L^{q}}\to 0\text{ as }\omega\to 0

for any q>NN2q>\frac{N}{N-2} if p2p\geqslant 2 or qN4q\geqslant\frac{N}{4} if p<2p<2. Hence, to prove that (Vωω)V0(V_{\omega}-\omega)\to V_{0} in Lq(N)L^{q}(\mathbb{R}^{N}), it is enough to show that

N(11(1+2δr2(vω(x)))2)qdx\int_{\mathbb{R}^{N}}\left(1-\frac{1}{(1+2\delta r^{2}(v_{\omega}(x)))^{2}}\right)^{q}\,\,\mathrm{d}x

is uniformly bounded. This is the case for any q>N2(N2)q>\frac{N}{2(N-2)}. Indeed, using again that vωL\|v_{\omega}\|_{L^{\infty}} is uniformly bounded for ω\omega small enough and that there exist C,R>0C,R>0 independent of ω\omega such that

vω(x)C|x|N2v_{\omega}(x)\leqslant\frac{C}{|x|^{N-2}}

for all xNBR(0)x\in\mathbb{R}^{N}\setminus B_{R}(0) (see proof of Proposition 4.5), we have

N(11(1+2δr2(vω(x)))2)qdxNr2q(vω(x))dx1+NBR(0)1|x|2q(N2)dx1\int_{\mathbb{R}^{N}}\left(1-\frac{1}{(1+2\delta r^{2}(v_{\omega}(x)))^{2}}\right)^{q}\,\,\mathrm{d}x\lesssim\int_{\mathbb{R}^{N}}r^{2q}(v_{\omega}(x))\,\,\mathrm{d}x\lesssim 1+\int_{\mathbb{R}^{N}\setminus B_{R}(0)}\frac{1}{|x|^{2q(N-2)}}\,\,\mathrm{d}x\lesssim 1

for any q>N2(N2)q>\frac{N}{2(N-2)}. Since, in any case, N2(N2)<min{NN2,N4}\frac{N}{2(N-2)}<\min\{\frac{N}{N-2},\frac{N}{4}\}, this completes the proof. ∎

Since L0L_{0} is the limit, in the norm resolvent sense, of LωL_{\omega} and LωL_{\omega} has exactly one negative eigenvalue (see Proposition 3.1), we deduce that (L0)rad1(L_{0})^{-1}_{\mathrm{rad}} has one negative eigenvalue and is otherwise positive and unbounded from above. As a consequence, we expect that, in the limit ω0\omega\to 0, M(ω)M^{\prime}(\omega) behaves like 2η0,(L0)rad1η0-2\langle\eta_{0},(L_{0})^{-1}_{\mathrm{rad}}\eta_{0}\rangle.

As in [32, Lemma 4.2], the first step to give a proper interpretation of the quadratic form

η0,(L0)rad1η0\langle\eta_{0},(L_{0})^{-1}_{\mathrm{rad}}\eta_{0}\rangle

is to prove that the quadratic form domain of (L0)rad1(L_{0})^{-1}_{\mathrm{rad}} is the same as for the free Laplacian if N5N\geqslant 5. More precisely, we have the following lemma.

Lemma 4.7.

Let N5N\geqslant 5 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. There exists a constant C>0C>0 such that

1C(Δ)rad1C(L0)rad1C(Δ)rad1.\frac{1}{C}(-\Delta)_{\mathrm{rad}}^{-1}-C\leqslant(L_{0})^{-1}_{\mathrm{rad}}\leqslant C(-\Delta)_{\mathrm{rad}}^{-1}.

The proof of this lemma is exactly the same as in [32, Lemma 4.2] and we do not reproduce it here. The key ingredients are the regularity of the potential V0=f0(v0)V_{0}=-f^{\prime}_{0}(v_{0}) and the fact L0L_{0} has exactly one negative eigenvalue. In particular, since r(s)0sr(s)\sim_{0}s and v0(|x|)+|x|2Nv_{0}(|x|)\sim_{+\infty}|x|^{2-N}, we have that V0V_{0} behaves like p|x|(2N)(p1)p|x|^{(2-N)(p-1)} at infinity. As a consequence, V0Lq(N)V_{0}\in L^{q}(\mathbb{R}^{N}) for any qN4q\geqslant\frac{N}{4} which is the regularity used in the proof of the lemma. Using that L0L_{0} has exactly one negative eigenvalue, we then deduce that the operator 1+K0=1+(Δ)12V0(Δ)121+K_{0}=1+(-\Delta)^{-\frac{1}{2}}V_{0}(-\Delta)^{-\frac{1}{2}} is invertible (within the sector of radial functions) and that (L0)rad1(L_{0})^{-1}_{\mathrm{rad}} can be written as

(Δ)12(1+K0)1(Δ)12.(-\Delta)^{-\frac{1}{2}}(1+K_{0})^{-1}(-\Delta)^{-\frac{1}{2}}.

Next, we give the upper bound on M(ω)M^{\prime}(\omega).

Proposition 4.7.

Let N5N\geqslant 5 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Then

lim supω0+M(ω)2η0,(L0)rad1η0[,+)\limsup_{\omega\to 0^{+}}M^{\prime}(\omega)\leqslant-2\langle\eta_{0},(L_{0})^{-1}_{\mathrm{rad}}\eta_{0}\rangle\in[-\infty,+\infty) (4.38)

in the sense of quadratic forms and where η0=r(v0)r(v0)\eta_{0}=r(v_{0})r^{\prime}(v_{0}). In dimensions N{5,6}N\in\{5,6\} the right side equals -\infty whereas it is finite for N7N\geqslant 7.

As above, the proof of this lemma is exactly the same as in [32, Lemma 4.3]. Once again we use the convergence of LωL_{\omega} in the norm resolvent sense and the strong convergence of ηω\eta_{\omega} in L2(N)L^{2}(\mathbb{R}^{N}) for N5N\geqslant 5. The fact that the right side of (4.38) is infinite in dimension N{5,6}N\in\{5,6\} and finite for N7N\geqslant 7 depends on the behavior of (Δ)12η0L2\|(-\Delta)^{-\frac{1}{2}}\eta_{0}\|_{L^{2}}. In particular, it depends on the behavior at infinity of η0=r(v0)r(v0)\eta_{0}=r(v_{0})r^{\prime}(v_{0}). On the one hand, we have

η0(x)=r(v0(x))r(v0(x))r(v0(x))v0(x).\eta_{0}(x)=r(v_{0}(x))r^{\prime}(v_{0}(x))\leqslant r(v_{0}(x))\leqslant v_{0}(x).

On the other hand, using Lemma 2.1, we have

η0(x)=r(v0(x))r(v0(x))v0(x)v0(x)r(v0(x))2v0(x)11+2δv0L2v0(x).\eta_{0}(x)=r^{\prime}(v_{0}(x))\frac{r(v_{0}(x))}{v_{0}(x)}v_{0}(x)\geqslant r^{\prime}(v_{0}(x))^{2}v_{0}(x)\geqslant\frac{1}{1+2\delta\|v_{0}\|^{2}_{L^{\infty}}}v_{0}(x).

As a consequence, using the decay of v0v_{0} at infinity, we deduce that, for any |x|1|x|\geqslant 1,

c1|x|N2η0(x)C1|x|N2c\frac{1}{|x|^{N-2}}\leqslant\eta_{0}(x)\leqslant C\frac{1}{|x|^{N-2}}

as in [32, Lemma 4.3].

Proposition 4.7 gives (1.22) for N6N\leqslant 6. It remains to prove it in dimensions N7N\geqslant 7.

Proposition 4.8.

Let N7N\geqslant 7 and N+2N2<p<3N+2N2\frac{N+2}{N-2}<p<\frac{3N+2}{N-2}. Then

limω0+M(ω)=2η0,(L0)rad1η0.\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=-2\langle\eta_{0},(L_{0})^{-1}_{\mathrm{rad}}\eta_{0}\rangle. (4.39)

The proof of this lemma is similar to that of [32, Lemma 4.4]. The main difference is linked to the particular form of fω(vω)f_{\omega}(v_{\omega}) which contains a term of the form ωr(vω)r(vω)\omega r^{\prime}(v_{\omega})r(v_{\omega}).

Here the quadratic form η0,(L0)rad1η0\langle\eta_{0},(L_{0})^{-1}_{\mathrm{rad}}\eta_{0}\rangle should be interpreted as (Δ)12η0,(1+K0)1(Δ)12η0\langle(-\Delta)^{-\frac{1}{2}}\eta_{0},(1+K_{0})^{-1}(-\Delta)^{-\frac{1}{2}}\eta_{0}\rangle with K0=(Δ)12V0(Δ)12K_{0}=(-\Delta)^{-\frac{1}{2}}V_{0}(-\Delta)^{-\frac{1}{2}}.

Proof.

We start by noticing that LωL_{\omega} can be written as

Lω=Δfω(vω)=Δ+ω(fω(vω)+ω)=Δ+ω+Vω=(Δ+ω)12(1+Kω)(Δ+ω)12L_{\omega}=-\Delta-f^{\prime}_{\omega}(v_{\omega})=-\Delta+\omega-(f^{\prime}_{\omega}(v_{\omega})+\omega)=-\Delta+\omega+V_{\omega}=(-\Delta+\omega)^{\frac{1}{2}}(1+K_{\omega})(-\Delta+\omega)^{\frac{1}{2}} (4.40)

where

Kω:=(Δ+ω)12Vω(Δ+ω)12=(ΔΔ+ω)12(Δ)12Vω(Δ)12(ΔΔ+ω)12.K_{\omega}:=(-\Delta+\omega)^{-\frac{1}{2}}V_{\omega}(-\Delta+\omega)^{-\frac{1}{2}}=\left(\frac{-\Delta}{-\Delta+\omega}\right)^{\frac{1}{2}}(-\Delta)^{-\frac{1}{2}}V_{\omega}(-\Delta)^{-\frac{1}{2}}\left(\frac{-\Delta}{-\Delta+\omega}\right)^{\frac{1}{2}}.

Thanks to Lemma 4.6, fω(vω)+ωf0(v0)=V0f^{\prime}_{\omega}(v_{\omega})+\omega\to f^{\prime}_{0}(v_{0})=-V_{0} in LN/2(N)L^{N/2}(\mathbb{R}^{N}). This, together with the Hardy-Littlewood-Sobolev inequality, implies that

(Δ)12Vω(Δ)12(Δ)12V0(Δ)12=K0(-\Delta)^{-\frac{1}{2}}V_{\omega}(-\Delta)^{-\frac{1}{2}}\to(-\Delta)^{-\frac{1}{2}}V_{0}(-\Delta)^{-\frac{1}{2}}=K_{0}

in operator norm. Since the operator K0K_{0} is compact and (ΔΔ+ω)12\left(\frac{-\Delta}{-\Delta+\omega}\right)^{\frac{1}{2}} converges strongly to the identity, we deduce that KωK0K_{\omega}\to K_{0} in operator norm. This implies that the spectrum of KωK_{\omega} converges to the spectrum of K0K_{0} and, since (1+K0)(1+K_{0}) is invertible, we deduce that (1+Kω)1(1+K_{\omega})^{-1} is bounded and converges to (1+K0)1(1+K_{0})^{-1} in operator norm. As a consequence, we can invert (4.40) and write

(Lω)rad1=(Δ+ω)12(1+Kω)1(Δ+ω)12(L_{\omega})^{-1}_{\mathrm{rad}}=(-\Delta+\omega)^{-\frac{1}{2}}(1+K_{\omega})^{-1}(-\Delta+\omega)^{-\frac{1}{2}}

and

M(ω)=2(Δ+ω)12ηω,(1+Kω)1(Δ+ω)12ηω.M^{\prime}(\omega)=-2\langle(-\Delta+\omega)^{-\frac{1}{2}}\eta_{\omega},(1+K_{\omega})^{-1}(-\Delta+\omega)^{-\frac{1}{2}}\eta_{\omega}\rangle.

To conclude, it remains to prove that (Δ+ω)12ηω(-\Delta+\omega)^{-\frac{1}{2}}\eta_{\omega} converges to (Δ)12η0(-\Delta)^{-\frac{1}{2}}\eta_{0} in L2(N)L^{2}(\mathbb{R}^{N}). From the HLS inequality, we have

(Δ)12η0L2η0L2NN+2\|(-\Delta)^{-\frac{1}{2}}\eta_{0}\|_{L^{2}}\lesssim\|\eta_{0}\|_{L^{\frac{2N}{N+2}}}

which is finite for N7N\geqslant 7. More precisely, if N7N\geqslant 7, Proposition 4.5 gives the strong convergence of ηω\eta_{\omega} towards η0\eta_{0} in L2NN+2(N){L^{\frac{2N}{N+2}}}(\mathbb{R}^{N}) as ω\omega goes to 0. As a consequence,

(Δ+ω)12ηω(Δ)12η0L2\displaystyle\|(-\Delta+\omega)^{-\frac{1}{2}}\eta_{\omega}-(-\Delta)^{-\frac{1}{2}}\eta_{0}\|_{L^{2}} (Δ+ω)12(ηωη0)L2+((Δ+ω)12(Δ)12)η0L2\displaystyle\leqslant\|(-\Delta+\omega)^{-\frac{1}{2}}(\eta_{\omega}-\eta_{0})\|_{L^{2}}+\|((-\Delta+\omega)^{-\frac{1}{2}}-(-\Delta)^{-\frac{1}{2}})\eta_{0}\|_{L^{2}}
ηωη0L2NN+2+((ΔΔ+ω)121)(Δ)12η0L2\displaystyle\lesssim\|\eta_{\omega}-\eta_{0}\|_{L^{\frac{2N}{N+2}}}+\left\|\left(\left(\frac{-\Delta}{-\Delta+\omega}\right)^{\frac{1}{2}}-1\right)(-\Delta)^{-\frac{1}{2}}\eta_{0}\right\|_{L^{2}}

which tends to zeros. Hence, we can pass to the limit and obtain

limω0M(ω)=2(Δ)12η0,(1+K0)1(Δ)12η0\lim_{\omega\to 0}M^{\prime}(\omega)=-2\langle(-\Delta)^{-\frac{1}{2}}\eta_{0},(1+K_{0})^{-1}(-\Delta)^{-\frac{1}{2}}\eta_{0}\rangle

as expected. ∎

To sum up, we finally gather the main results for the supercritical case.

Proof of Theorem 1.3 (iii).

The convergence of uωu_{\omega} to u0u_{0} is a consequence of Proposition 4.4. The limits in (1.21) follow from Proposition 4.5, while those in (1.22) follow from Proposition 4.6, Proposition 4.7 and Proposition 4.8. The proof of the supercritical case is complete. ∎

4.3. Critical case

We now consider the case p=N+2N2p=\frac{N+2}{N-2}, i.e. p+1=2p+1=2^{*}. Combining the two identities from Proposition 1.1 with ω=0\omega=0, we find that

2N|u|2||u||2dx=(12p+1)N|u|p+1dx2\int_{\mathbb{R}^{N}}|u|^{2}|\nabla|u||^{2}\,\mathrm{d}x=\Big{(}1-\frac{2^{*}}{p+1}\Big{)}\int_{\mathbb{R}^{N}}|u|^{p+1}\,\mathrm{d}x (4.41)

for any solution uX~u\in\widetilde{X}. Hence, for p+1=2p+1=2^{*}, the limit equation (1.20) has no nontrivial solution in H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}). In fact, we will see that, in the critical case, a rescaled version of the solution vωv_{\omega} converges, as ω0\omega\to 0, to the unique (up to scaling) radial positive solution of the critical Lane-Emden-Fowler equation

Δw+wN+2N2=0.\Delta w+w^{\frac{N+2}{N-2}}=0.

On the formal level, considering a scaling parameter λω>0\lambda_{\omega}>0, if u>0u>0 is a solution of (Eω\mathrm{E}_{\omega}), it follows that

w(x):=λω2/(p1)u(λωx)w(x):=\lambda_{\omega}^{2/(p-1)}u(\lambda_{\omega}x)

satisfies

Δwωλω2w+wp+λω4/(p1)Δ(w2)w=0.\Delta w-\omega\lambda_{\omega}^{2}w+w^{p}+\lambda_{\omega}^{-4/(p-1)}\Delta(w^{2})w=0.

Thus, choosing λω\lambda_{\omega}\to\infty such that ωλω20\omega\lambda_{\omega}^{2}\to 0 yields the limit equation (1.17).

In order to establish rigorously the convergence of uωu_{\omega} in the limit ω0\omega\to 0 we shall rather work with zω=h(uω)(mω)z_{\omega}=h(u_{\omega})(\sqrt{m_{\omega}}\cdot), as in Subsection 4.2. We will compare, in the limit ω0\omega\to 0, the minimization problem infzKωJω(z)=mω\inf_{z\in K_{\omega}}J_{\omega}(z)=m_{\omega} and its minimizers zωz_{\omega} with the following one:

m:=infH˙1{0}J,m_{*}:=\inf_{\dot{H}^{1}\setminus\{0\}}J_{*},

where

J(w):=|w|2(|w|p+1)2p+1=|w|2(|w|2)22=|w|2(|w|2NN2)12N.J_{*}(w):=\frac{\int|\nabla w|^{2}}{\left(\int|w|^{p+1}\right)^{\frac{2}{p+1}}}=\frac{\int|\nabla w|^{2}}{\left(\int|w|^{2^{*}}\right)^{\frac{2}{2^{*}}}}=\frac{\int|\nabla w|^{2}}{\left(\int|w|^{\frac{2N}{N-2}}\right)^{1-\frac{2}{N}}}. (4.42)

JJ_{*} is invariant under the scaling

w(x)wλ(x):=λN22w(λ1x)(λ>0).w(x)\mapsto w_{\lambda}(x):=\lambda^{-\frac{N-2}{2}}w(\lambda^{-1}x)\quad(\lambda>0). (4.43)

It is well-known [50] that mm_{*} is attained on the family of positive radial functions

{Wλ}λ>0H˙1(N),W1(x)W(x):=U(mx),\{W_{\lambda}\}_{\lambda>0}\subset\dot{H}^{1}(\mathbb{R}^{N}),\quad W_{1}(x)\equiv W(x):=U(\sqrt{m_{*}}x),

where UU is the Aubin-Talenti function defined in (1.16). The following properties of the family of minimizers {Wλ}λ>0\{W_{\lambda}\}_{\lambda>0} are noteworthy:

ΔWλ=mWλN+2N2inN,-\Delta W_{\lambda}=m_{*}W_{\lambda}^{\frac{N+2}{N-2}}\quad\text{in}\ \mathbb{R}^{N}, (4.44)
N|Wλ|2dx=N|W|2dx=mandNWλ2NN2dx=NW2NN2dx=1,λ>0.\int_{\mathbb{R}^{N}}|\nabla W_{\lambda}|^{2}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}|\nabla W|^{2}\,\mathrm{d}x=m_{*}\quad\text{and}\quad\int_{\mathbb{R}^{N}}W_{\lambda}^{\frac{2N}{N-2}}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}W^{\frac{2N}{N-2}}\,\mathrm{d}x=1,\quad\forall\lambda>0. (4.45)
Proposition 4.9.

Suppose p=N+2N2p=\frac{N+2}{N-2}. There exists a scaling function ωλω:(0,)(0,)\omega\mapsto\lambda_{\omega}:(0,\infty)\to(0,\infty) with the following properties. For any sequence (ωn)(0,)(\omega_{n})\subset(0,\infty) such that ωn0\omega_{n}\to 0, letting

λn=λωn,mn=mωn,zn=zωn,\lambda_{n}=\lambda_{\omega_{n}},\quad m_{n}=m_{\omega_{n}},\quad z_{n}=z_{\omega_{n}},

and

wn:=λnN22zn(λn)w_{n}:=\lambda_{n}^{\frac{N-2}{2}}z_{n}(\lambda_{n}\cdot) (4.46)

we have

wnWinH˙1(N)L2(N),asn.w_{n}\to W\quad\text{in}\ \dot{H}^{1}(\mathbb{R}^{N})\cap L^{2^{*}}(\mathbb{R}^{N}),\quad\text{as}\ n\to\infty. (4.47)

The proof of Proposition 4.9 will use several lemmas. The following bounds are direct consequences of Lemma 2.1 and (4.45).

Lemma 4.8.

For any λ>0\lambda>0 and WλW_{\lambda} defined as in (4.44), we have

Nr(Wλ(x))2dx(11+2δλ(N2))22\int_{\mathbb{R}^{N}}r(W_{\lambda}(x))^{2^{*}}\,\mathrm{d}x\geqslant\left(\frac{1}{1+2\delta\lambda^{-(N-2)}}\right)^{\frac{2^{*}}{2}} (4.48)

and

Nr(Wλ(x))2dx4λ2NW(x)2dx.\int_{\mathbb{R}^{N}}r(W_{\lambda}(x))^{2}\,\mathrm{d}x\leqslant 4\lambda^{2}\int_{\mathbb{R}^{N}}W(x)^{2}\,\mathrm{d}x. (4.49)

We next establish the lemma corresponding to Lemma 4.2 in the supercritical case.

Lemma 4.9.

As ω0\omega\to 0, there holds

1mωm1+o(1).1\leqslant\frac{m_{\omega}}{m_{*}}\leqslant 1+o(1). (4.50)
Proof.

First, since zωH˙1z_{\omega}\in\dot{H}^{1} satisfies

N|zω|2dx=mω\int_{\mathbb{R}^{N}}|\nabla z_{\omega}|^{2}\,\mathrm{d}x=m_{\omega}

and

12=NFω(zω)dx=12Nr(zω)2dxω2Nr(zω)2dx<12Nr(zω)2dx12Nzω2dx,\frac{1}{2^{*}}=\int_{\mathbb{R}^{N}}F_{\omega}(z_{\omega})\,\mathrm{d}x=\frac{1}{2^{*}}\int_{\mathbb{R}^{N}}r(z_{\omega})^{2^{*}}\,\mathrm{d}x-\frac{\omega}{2}\int_{\mathbb{R}^{N}}r(z_{\omega})^{2}\,\mathrm{d}x<\frac{1}{2^{*}}\int_{\mathbb{R}^{N}}r(z_{\omega})^{2^{*}}\,\mathrm{d}x\leqslant\frac{1}{2^{*}}\int_{\mathbb{R}^{N}}z_{\omega}^{2^{*}}\,\mathrm{d}x,

we obtain

mJ(zω)=|zω|2(zω2)22=mω(zω2)22<mω.m_{*}\leqslant J_{*}(z_{\omega})=\frac{\int|\nabla z_{\omega}|^{2}}{\left(\int z_{\omega}^{2^{*}}\right)^{\frac{2}{2^{*}}}}=\frac{m_{\omega}}{\left(\int z_{\omega}^{2^{*}}\right)^{\frac{2}{2^{*}}}}<m_{\omega}. (4.51)

Now, if N5N\geqslant 5, we have WL2W\in L^{2} and we can use the family {Wλ:λ>0}H1\{W_{\lambda}:\lambda>0\}\subset H^{1} as test functions for JωJ_{\omega}. By Lemma 4.8, we see that, if λ1\lambda\gg 1 and ωλ21\omega\lambda^{2}\ll 1,

NFω(Wλ)dx=12Nr(Wλ)2dxω2Nr(Wλ)2dx>0,\int_{\mathbb{R}^{N}}F_{\omega}(W_{\lambda})\,\mathrm{d}x=\frac{1}{2^{*}}\int_{\mathbb{R}^{N}}r(W_{\lambda})^{2^{*}}\,\mathrm{d}x-\frac{\omega}{2}\int_{\mathbb{R}^{N}}r(W_{\lambda})^{2}\,\mathrm{d}x>0,

i.e. WλKωW_{\lambda}\in K_{\omega}. Choosing λ=ω1/N\lambda=\omega^{-1/{N}}, it follows from Lemma 4.8 that

mωJω(Wλ)\displaystyle m_{\omega}\leqslant J_{\omega}(W_{\lambda}) =|W|2(W2)22(W2r(Wλ)222ωr(Wλ)2)22\displaystyle=\frac{\int|\nabla W|^{2}}{\left(\int W^{2^{*}}\right)^{\frac{2}{2^{*}}}}\cdot\left(\frac{\int W^{2^{*}}}{\int r(W_{\lambda})^{2^{*}}-\frac{2^{*}}{2}\omega\int r(W_{\lambda})^{2}}\right)^{\frac{2}{2^{*}}}
=m(r(Wλ)222ωr(Wλ)2)22\displaystyle=m_{*}\left(\int r(W_{\lambda})^{2^{*}}-\frac{2^{*}}{2}\omega\int r(W_{\lambda})^{2}\right)^{-\frac{2}{2^{*}}}
m((11+2δλ(N2))2222ωλ2W2)22\displaystyle\leqslant m_{*}\left(\left(\frac{1}{1+2\delta\lambda^{-(N-2)}}\right)^{\frac{2^{*}}{2}}-2\cdot 2^{*}\omega\lambda^{2}\int W^{2}\right)^{-\frac{2}{2^{*}}}
=m(1+O(λ(N2))+O(ω12N))22\displaystyle=m_{*}\left(1+O(\lambda^{-(N-2)})+O(\omega^{1-\frac{2}{N}})\right)^{-\frac{2}{2^{*}}}
=m(1+O(ω12N))22\displaystyle=m_{*}\left(1+O(\omega^{1-\frac{2}{N}})\right)^{-\frac{2}{2^{*}}}
=m(1+O(ω12N)),asω0.\displaystyle=m_{*}(1+O(\omega^{1-\frac{2}{N}})),\quad\text{as}\ \omega\to 0.

We deduce that

1mωm1+O(ω12N),asω0,1\leqslant\frac{m_{\omega}}{m_{*}}\leqslant 1+O(\omega^{1-\frac{2}{N}}),\quad\text{as}\ \omega\to 0,

which concludes the proof in case N5N\geqslant 5.

For N{3,4}N\in\{3,4\}, we take RλR\gg\lambda and use the same cut-off function ηR\eta_{R} as in the proof of Lemma 4.2. Then WλL2W_{\lambda}\not\in L^{2} but ηRWλH1\eta_{R}W_{\lambda}\in H^{1} can be used as a test function for JωJ_{\omega}. We have

mωJω(ηRWλ)=|ηRWλ|2(|ηRWλ|2)22((ηRWλ)2r(ηRWλ)222ωr(ηRWλ)2)22.m_{\omega}\leqslant J_{\omega}(\eta_{R}W_{\lambda})=\frac{\int|\nabla\eta_{R}W_{\lambda}|^{2}}{\left(\int|\eta_{R}W_{\lambda}|^{2^{*}}\right)^{\frac{2}{2^{*}}}}\cdot\left(\frac{\int(\eta_{R}W_{\lambda})^{2^{*}}}{\int r(\eta_{R}W_{\lambda})^{2^{*}}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}}\right)^{\frac{2}{2^{*}}}.

First, by dominated convergence, |ηRWλ|2|Wλ|2=m\int|\nabla\eta_{R}W_{\lambda}|^{2}\to\int|\nabla W_{\lambda}|^{2}=m^{*} as RR\to\infty.

Next, similarly to (4.48), we obtain

Nr(ηRWλ)2dx(11+2δλ(N2))22N(ηRWλ)2dx.\int_{\mathbb{R}^{N}}r(\eta_{R}W_{\lambda})^{2^{*}}\,\mathrm{d}x\geqslant\left(\frac{1}{1+2\delta\lambda^{-(N-2)}}\right)^{\frac{2^{*}}{2}}\int_{\mathbb{R}^{N}}(\eta_{R}W_{\lambda})^{2^{*}}\,\mathrm{d}x.

Furthermore, by (4.45),

N(ηRWλ)2dx=NWλ2dx+O(1(R/λ)N)=1+O(1(R/λ)N),R/λ1.\int_{\mathbb{R}^{N}}(\eta_{R}W_{\lambda})^{2^{*}}\,\mathrm{d}x=\int_{\mathbb{R}^{N}}W_{\lambda}^{2^{*}}\,\mathrm{d}x+O\Big{(}\frac{1}{(R/\lambda)^{N}}\Big{)}=1+O\Big{(}\frac{1}{(R/\lambda)^{N}}\Big{)},\quad R/\lambda\gg 1.

Using Lemma 4.1 and proceeding as in (4.49), we have

gN(R,λ):=Nr(ηRWλ)2dx4N(ηRWλ)2dx={O(λR)ifN=3,O(λ2log(R/λ))ifN=4,R/λ1.g_{N}(R,\lambda):=\int_{\mathbb{R}^{N}}r(\eta_{R}W_{\lambda})^{2}\,\mathrm{d}x\leqslant 4\int_{\mathbb{R}^{N}}(\eta_{R}W_{\lambda})^{2}\,\mathrm{d}x=\begin{cases}O(\lambda R)&\text{if}\ N=3,\\ O(\lambda^{2}\log(R/\lambda))&\text{if}\ N=4,\end{cases}\quad R/\lambda\gg 1.

Now

(ηRWλ)p+1r(ηRWλ)p+122ωr(ηRWλ)2\displaystyle\frac{\int(\eta_{R}W_{\lambda})^{p+1}}{\int r(\eta_{R}W_{\lambda})^{p+1}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}} 1+O(1(R/λ)N)(11+2δλ(N2))22(1+O(1(R/λ)N))22ωgN(R,λ)\displaystyle\leqslant\frac{1+O\left(\frac{1}{(R/\lambda)^{N}}\right)}{\left(\frac{1}{1+2\delta\lambda^{-(N-2)}}\right)^{\frac{2^{*}}{2}}\left(1+O\Big{(}\frac{1}{(R/\lambda)^{N}}\Big{)}\right)-\frac{2^{*}}{2}\omega g_{N}(R,\lambda)}
=1+O(1(R/λ)N)1+O(λ(N2))+O(1(R/λ)N)22ωgN(R,λ).\displaystyle=\frac{1+O\Big{(}\frac{1}{(R/\lambda)^{N}}\Big{)}}{1+O\left(\lambda^{-(N-2)}\right)+O\Big{(}\frac{1}{(R/\lambda)^{N}}\Big{)}-\frac{2^{*}}{2}\omega g_{N}(R,\lambda)}.

We now conclude the proof in the following way. Let R=λαR=\lambda^{\alpha} with α>1\alpha>1 to be chosen later.

If N=3N=3, we let λ=ω1/4\lambda=\omega^{-1/4}, so that

1(R/λ)N=1λ3(α1)=ω34(α1),1λN2=ω14\frac{1}{(R/\lambda)^{N}}=\frac{1}{\lambda^{3(\alpha-1)}}=\omega^{\frac{3}{4}(\alpha-1)},\quad\frac{1}{\lambda^{N-2}}=\omega^{\frac{1}{4}}

and

ωg3(λα,λ)=O(ωλα+1)=O(ω114(α+1)),\omega g_{3}(\lambda^{\alpha},\lambda)=O(\omega\lambda^{\alpha+1})=O(\omega^{1-\frac{1}{4}(\alpha+1)}),

provided that 1<α<31<\alpha<3. This implies, by taking 43α2\frac{4}{3}\leqslant\alpha\leqslant 2,

(ηRWλ)p+1r(ηRWλ)p+122ωr(ηRWλ)21+O(ω14)+O(ω34(α1))+O(ω114(α+1))=1+O(ω14).\frac{\int(\eta_{R}W_{\lambda})^{p+1}}{\int r(\eta_{R}W_{\lambda})^{p+1}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}}\leqslant 1+O(\omega^{\frac{1}{4}})+O(\omega^{\frac{3}{4}(\alpha-1)})+O(\omega^{1-\frac{1}{4}(\alpha+1)})=1+O(\omega^{\frac{1}{4}}).

If N=4N=4, we let λ=ω1/6\lambda=\omega^{-1/6}, so that

1(R/λ)N=1λ4(α1)=ω23(α1),1λN2=ω13\frac{1}{(R/\lambda)^{N}}=\frac{1}{\lambda^{4(\alpha-1)}}=\omega^{\frac{2}{3}(\alpha-1)},\quad\frac{1}{\lambda^{N-2}}=\omega^{\frac{1}{3}}

and

ωg4(λα,λ)=O(ωλ2logλ)=O(ω23log1ω)=o(ω13).\omega g_{4}(\lambda^{\alpha},\lambda)=O(\omega\lambda^{2}\log\lambda)=O\Big{(}\omega^{\frac{2}{3}}\log\frac{1}{\omega}\Big{)}=o(\omega^{\frac{1}{3}}).

This implies, by taking α32\alpha\geqslant\frac{3}{2},

(ηRWλ)p+1r(ηRWλ)p+122ωr(ηRWλ)21+O(ω13)+O(ω23(α1))=1+O(ω13).\frac{\int(\eta_{R}W_{\lambda})^{p+1}}{\int r(\eta_{R}W_{\lambda})^{p+1}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}}\leqslant 1+O(\omega^{\frac{1}{3}})+O(\omega^{\frac{2}{3}(\alpha-1)})=1+O(\omega^{\frac{1}{3}}).

The proof is complete.

Remark 7.

Observe from the proof of Lemma 4.9 that the rate of decay of the remainder in the right-hand side of (4.50) depends on the dimension NN. As will be discussed in detail below (see Remark 8), the decay rate obtained here for N=4N=4 is not optimal.

We can now determine the behavior of zωz_{\omega} as ω0\omega\to 0.

Lemma 4.10.

Consider a sequence (ωn)(0,)(\omega_{n})\subset(0,\infty), such that ωn0\omega_{n}\to 0. For all nn\in\mathbb{N}, let zn:=zωnz_{n}:=z_{\omega_{n}} be a minimizer for (2.6) with ω=ωn\omega=\omega_{n}. Then, as nn\to\infty,

znL22m=W1L22,ωnr(zn)L220,r(zn)L221\|\nabla z_{n}\|_{L^{2}}^{2}\to m_{*}=\|\nabla W_{1}\|_{L^{2}}^{2},\quad\omega_{n}\|r(z_{n})\|_{L^{2}}^{2}\to 0,\quad\|r(z_{n})\|_{L^{2^{*}}}^{2^{*}}\to 1 (4.52)

and

znL221.\|z_{n}\|_{L^{2^{*}}}^{2^{*}}\to 1. (4.53)
Proof.

First, the limits in (4.52) are proved in the same way as Lemma 4.3, using Lemma 4.9 instead of Lemma 4.2. Next, (4.53) follows directly from (4.50) and (4.51). ∎

We can now give the

Proof of Proposition 4.9.

By Lemma 4.10, the radial sequence {zn}H˙1(N)L2(N)\{z_{n}\}\subset\dot{H}^{1}(\mathbb{R}^{N})\cap L^{2^{*}}(\mathbb{R}^{N}) satisfies

znL22m,znL221,n.\|\nabla z_{n}\|_{L^{2}}^{2}\to m_{*},\quad\|z_{n}\|_{L^{2^{*}}}^{2^{*}}\to 1,\quad n\to\infty.

Then, by the concentration-compactness principle (see the original paper [33], or [53] for a more concise exposition), there exists a scaling sequence (λn)(\lambda_{n}) such that

wn:=λnN22zn(λn)WinH˙1(N)L2(N),n.w_{n}:=\lambda_{n}^{\frac{N-2}{2}}z_{n}(\lambda_{n}\cdot)\to W\quad\text{in}\quad\dot{H}^{1}(\mathbb{R}^{N})\cap L^{2^{*}}(\mathbb{R}^{N}),\quad n\to\infty.

4.3.1. Further convergence properties of the solutions

We first prove the upper bounds in (1.18).

Proposition 4.10.

The scaling function ωλω:(0,)(0,)\omega\mapsto\lambda_{\omega}:(0,\infty)\to(0,\infty) can be chosen so that, as ω0\omega\to 0,

{λωω38if N=3,λωω13if N=4,λωω1Nif N5.\begin{cases}\lambda_{\omega}\lesssim\omega^{-\frac{3}{8}}&\text{if }N=3,\\ \lambda_{\omega}\lesssim\omega^{-\frac{1}{3}}&\text{if }N=4,\\ \lambda_{\omega}\lesssim\omega^{-\frac{1}{N}}&\text{if }N\geqslant 5.\end{cases} (4.54)
Proof.

Consider δω:=mωm\delta_{\omega}:=m_{\omega}-m_{*}. By the proof of Lemma 4.9, we have

0δω{ω14if N=3,ω13if N=4,ω12Nif N5.0\leqslant\delta_{\omega}\lesssim\begin{cases}\omega^{\frac{1}{4}}&\text{if }N=3,\\ \omega^{\frac{1}{3}}&\text{if }N=4,\\ \omega^{1-\frac{2}{N}}&\text{if }N\geqslant 5.\end{cases} (4.55)

The upper bounds (4.54) follow directly from (4.55) and the following estimates:

ωλω2wωL22δω,\omega\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}\lesssim\delta_{\omega}, (4.56)
wωL21.\|w_{\omega}\|_{L^{2}}\gtrsim 1. (4.57)

To prove (4.56), we first observe that

λω2wωL22=zωL22.\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}=\|z_{\omega}\|_{L^{2}}^{2}. (4.58)

By Lemma 4.5 (i), there exists η0>0\eta_{0}>0 and M0>0M_{0}>0 such that

zωLM0,0<ω<η0.\|z_{\omega}\|_{L^{\infty}}\leqslant M_{0},\quad 0<\omega<\eta_{0}.

Hence,

r(zω)2zω2M02r(z_{\omega})^{2}\leqslant z_{\omega}^{2}\leqslant M_{0}^{2}

and it follows by Lemma 2.1 that

Nr(zω)2dxNzω21+2δr(zω)2dx11+2δM02Nzω2dx.\int_{\mathbb{R}^{N}}r(z_{\omega})^{2}\,\mathrm{d}x\geqslant\int_{\mathbb{R}^{N}}\frac{z_{\omega}^{2}}{1+2\delta r(z_{\omega})^{2}}\,\mathrm{d}x\geqslant\frac{1}{1+2\delta M_{0}^{2}}\int_{\mathbb{R}^{N}}z_{\omega}^{2}\,\mathrm{d}x.

Thus,

zωL22(1+2δM02)r(zω)L22\|z_{\omega}\|_{L^{2}}^{2}\leqslant(1+2\delta M_{0}^{2})\|r(z_{\omega})\|_{L^{2}}^{2}

and (4.58) yields

ωλω2wωL22ωr(zω)L22.\omega\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}\lesssim\omega\|r(z_{\omega})\|_{L^{2}}^{2}.

The proof of (4.56) will be complete if we show that

ωr(zω)L22δω.\omega\|r(z_{\omega})\|_{L^{2}}^{2}\lesssim\delta_{\omega}. (4.59)

To this end, we use zωz_{\omega} as a test function for JJ_{*}:

mJ(zω)=|zω|2zωL22=mω(zω2)2/2mω(r(zω)2)2/2.m_{*}\leqslant J_{*}(z_{\omega})=\frac{\int|\nabla z_{\omega}|^{2}}{\|z_{\omega}\|_{L^{2^{*}}}^{2}}=\frac{m_{\omega}}{(\int z_{\omega}^{2^{*}})^{2/2^{*}}}\leqslant\frac{m_{\omega}}{(\int r(z_{\omega})^{2^{*}})^{2/2^{*}}}.

Since

1=Nr(zω)2dx22ωNr(zω)2dx,1=\int_{\mathbb{R}^{N}}r(z_{\omega})^{2^{*}}\,\mathrm{d}x-\frac{2^{*}}{2}\omega\int_{\mathbb{R}^{N}}r(z_{\omega})^{2}\,\mathrm{d}x,

we deduce that

mmω(1+22ωr(zω)2)2/2.m_{*}\leqslant\frac{m_{\omega}}{(1+\frac{2^{*}}{2}\omega\int r(z_{\omega})^{2})^{2/2^{*}}}.

It follows by a first order Taylor expansion that

22ωNr(zω)2dxmω22m22=22mω221(mωm)+o(mωm).\frac{2^{*}}{2}\omega\int_{\mathbb{R}^{N}}r(z_{\omega})^{2}\,\mathrm{d}x\leqslant m_{\omega}^{\frac{2^{*}}{2}}-m_{*}^{\frac{2^{*}}{2}}=\frac{2^{*}}{2}m_{\omega}^{\frac{2^{*}}{2}-1}(m_{\omega}-m_{*})+o(m_{\omega}-m_{*}).

This proves (4.59) and completes the proof of (4.56).

To prove (4.57), we observe that wωWw_{\omega}\to W in H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}) as ω0\omega\to 0 implies wωWw_{\omega}\to W in L2(N)L^{2^{*}}(\mathbb{R}^{N}) and in Lloc2(N)L^{2}_{\mathrm{loc}}(\mathbb{R}^{N}). Since

χB1(0)wωL2χB1(0)WL2χB1(0)(Wwω)L2,\|\chi_{B_{1}(0)}w_{\omega}\|_{L^{2}}\geqslant\|\chi_{B_{1}(0)}W\|_{L^{2}}-\|\chi_{B_{1}(0)}(W-w_{\omega})\|_{L^{2}},

it follows that

wωL2χB1(0)WL2+o(1),ω0.\|w_{\omega}\|_{L^{2}}\geqslant\|\chi_{B_{1}(0)}W\|_{L^{2}}+o(1),\quad\omega\to 0.

This proves (4.57) and completes the proof of the proposition. ∎

We next turn to the lower bounds in (1.18), which are more involved.

Proposition 4.11.

The scaling function ωλω:(0,)(0,)\omega\mapsto\lambda_{\omega}:(0,\infty)\to(0,\infty) can be chosen so that, as ω0\omega\to 0,

{λωω14if N=3,λωω16if N=4,λωω1Nif N5.\begin{cases}\lambda_{\omega}\gtrsim\omega^{-\frac{1}{4}}&\text{if }N=3,\\ \lambda_{\omega}\gtrsim\omega^{-\frac{1}{6}}&\text{if }N=4,\\ \lambda_{\omega}\gtrsim\omega^{-\frac{1}{N}}&\text{if }N\geqslant 5.\end{cases} (4.60)
Remark 8.

For N5N\geqslant 5, the upper bounds (4.54) and the lower bounds (4.60) come with the same power of ω1\omega^{-1}. For N=3,4N=3,4, they do not match. This is due to the rough estimate (4.57) used to derive the upper bounds from (4.56) and also, in case N=4N=4, to the non-optimal decay of δω\delta_{\omega} that was already pointed out in Remark 7. For N=3,4N=3,4, estimate (4.57) will be improved in Lemma 4.16 and the optimal lower bounds on λω\lambda_{\omega} will be given in Proposition 4.12 and Proposition 4.13, respectively.

The proof of Proposition 4.11 will use the following lemmas.

Lemma 4.11.
limω0λω=.\lim_{\omega\to 0}\lambda_{\omega}=\infty.
Proof.

In this proof we use the Nehari and Pohozaev integral identities derived from the equation

Δwω=λω1+N2mωfω(λωN22wω),-\Delta w_{\omega}=\lambda_{\omega}^{1+\frac{N}{2}}m_{\omega}f_{\omega}(\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}), (4.61)

which respectively read

Nr(λωN22wω)2(1+2δr(λωN22wω)2)|wω|2dx=λωNmωNr(λωN22wω)2dx22ωmωλωNNr(λωN22wω)2dx\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\Big{(}1+2\delta r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\Big{)}|\nabla w_{\omega}|^{2}\,\mathrm{d}x\\ =\lambda_{\omega}^{N}{m_{\omega}}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2^{*}}\,\mathrm{d}x-\frac{2^{*}}{2}\omega\,{m_{\omega}}\lambda_{\omega}^{N}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\,\mathrm{d}x

and

Nr(λωN22wω)2(1+4δr(λωN22wω)2)|wω|2dx=λωNmωNr(λωN22wω)2dxωmωλωNNr(λωN22wω)2dx.\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\Big{(}1+4\delta r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\Big{)}|\nabla w_{\omega}|^{2}\,\mathrm{d}x\\ =\lambda_{\omega}^{N}{m_{\omega}}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2^{*}}\,\mathrm{d}x-\omega\,{m_{\omega}}\lambda_{\omega}^{N}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\,\mathrm{d}x.

We deduce from these identities that

δNr(λωN22wω)2r(λωN22wω)2|wω|2dx=1N2ωmωλωNNr(λωN22wω)2dx.\delta\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x=\frac{1}{N-2}\omega\,{m_{\omega}}\lambda_{\omega}^{N}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}\big{)}^{2}\,\mathrm{d}x. (4.62)

Suppose by contradiction there exist [0,)\ell\in[0,\infty) and a sequence (ωn)(0,)(\omega_{n})\subset(0,\infty) such that ωn0\omega_{n}\to 0 and λωn\lambda_{\omega_{n}}\to\ell, as nn\to\infty. For the remainder of the proof, we shall again abbreviate the notation as zn:=zωnz_{n}:=z_{\omega_{n}}, wn:=wωnw_{n}:=w_{\omega_{n}}, λn:=λωn\lambda_{n}:=\lambda_{\omega_{n}} and mn:=mωnm_{n}:=m_{\omega_{n}}.

First suppose that =0\ell=0. From Lemma 4.5 (i), we know there exists n0n_{0}\in\mathbb{N} such that znLM0\|z_{n}\|_{L^{\infty}}\leqslant M_{0} for all nn0n\geqslant n_{0}. It follows that

|λnN22zn(λnx)|M0λnN220,xN,\big{|}\lambda_{n}^{\frac{N-2}{2}}z_{n}(\lambda_{n}x)\big{|}\leqslant M_{0}\lambda_{n}^{\frac{N-2}{2}}\to 0,\quad\forall x\in\mathbb{R}^{N},

whence wn0w_{n}\to 0 pointwise as nn\to\infty, which contradicts wnWw_{n}\to W.

Now suppose that (0,)\ell\in(0,\infty). It follows from Fatou’s Lemma that

Nr(N22W)2r(N22W)2|W|2dxlim infnNr(λnN22wn)2r(λnN22wn)2|wn|2dx.\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\ell^{-\frac{N-2}{2}}W\big{)}^{2}r\big{(}\ell^{-\frac{N-2}{2}}W\big{)}^{2}|\nabla W|^{2}\,\mathrm{d}x\leqslant\liminf_{n\to\infty}\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\lambda_{n}^{-\frac{N-2}{2}}w_{n}\big{)}^{2}r\big{(}\lambda_{n}^{-\frac{N-2}{2}}w_{n}\big{)}^{2}|\nabla w_{n}|^{2}\,\mathrm{d}x. (4.63)

On the other hand, by Lemma 4.9 and (4.59),

ωnmnλnNNr(λnN22wn(x))2dx\displaystyle\omega_{n}{m_{n}}\lambda_{n}^{N}\int_{\mathbb{R}^{N}}r\big{(}\lambda_{n}^{-\frac{N-2}{2}}w_{n}(x)\big{)}^{2}\,\mathrm{d}x =mnωnλnNNr(zn(λnx))2dx\displaystyle={m_{n}}\omega_{n}\lambda_{n}^{N}\int_{\mathbb{R}^{N}}r\big{(}z_{n}(\lambda_{n}x)\big{)}^{2}\,\mathrm{d}x
=mnωnNr(zn(y))2dy0,n.\displaystyle={m_{n}}\omega_{n}\int_{\mathbb{R}^{N}}r\big{(}z_{n}(y)\big{)}^{2}\,\mathrm{d}y\to 0,\quad n\to\infty. (4.64)

It follows by (4.62), (4.63) and (4.3.1) that

Nr(N22W)2r(N22W)2|W|2dx=0.\int_{\mathbb{R}^{N}}r^{\prime}\big{(}\ell^{-\frac{N-2}{2}}W\big{)}^{2}r\big{(}\ell^{-\frac{N-2}{2}}W\big{)}^{2}|\nabla W|^{2}\,\mathrm{d}x=0.

This contradiction shows that, indeed, λω\lambda_{\omega}\to\infty as ω0\omega\to 0. ∎

Lemma 4.12.

As ω0\omega\to 0, there holds

λω2wωL22uωL22λω2wωL22.\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}\lesssim\|u_{\omega}\|_{L^{2}}^{2}\lesssim\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}.
Proof.

We have

1λω2uωL22=1λω2r(vω)L22=1λω2mωN2Nr(zω(x))2dx=1λω2mωN2Nzω(x)2gω(x)dx,\displaystyle\frac{1}{\lambda_{\omega}^{2}}\|u_{\omega}\|_{L^{2}}^{2}=\frac{1}{\lambda_{\omega}^{2}}\|r(v_{\omega})\|_{L^{2}}^{2}=\frac{1}{\lambda_{\omega}^{2}}m_{\omega}^{\frac{N}{2}}\int_{\mathbb{R}^{N}}r(z_{\omega}(x))^{2}\,\mathrm{d}x=\frac{1}{\lambda_{\omega}^{2}}m_{\omega}^{\frac{N}{2}}\int_{\mathbb{R}^{N}}z_{\omega}(x)^{2}g_{\omega}(x)\,\mathrm{d}x,

where

gω(x)=r(zω(x))2zω(x)2.g_{\omega}(x)=\frac{r(z_{\omega}(x))^{2}}{z_{\omega}(x)^{2}}.

Using Lemma 2.1 (iv), we deduce that, for all xNx\in\mathbb{R}^{N},

r(zω(x))2gω(x)4r(zω(x))2\displaystyle r^{\prime}(z_{\omega}(x))^{2}\leqslant g_{\omega}(x)\leqslant 4r^{\prime}(z_{\omega}(x))^{2}

with

r(zω(x))2=11+2δr(zω(x))2.r^{\prime}(z_{\omega}(x))^{2}=\frac{1}{1+2\delta r(z_{\omega}(x))^{2}}.

Thanks to Lemma 4.20, for all ω(0,η0)\omega\in(0,\eta_{0}), r(zω)LzωLM0\|r(z_{\omega})\|_{L^{\infty}}\leqslant\|z_{\omega}\|_{L^{\infty}}\leqslant M_{0}, so that

11+2δM02r(zω(x))2gω(x)4,xN.\frac{1}{1+2\delta M_{0}^{2}}\leqslant r^{\prime}(z_{\omega}(x))^{2}\leqslant g_{\omega}(x)\leqslant 4,\quad x\in\mathbb{R}^{N}. (4.65)

As a consequence, since mωmm_{\omega}\geqslant m_{*} by Lemma 4.9,

1λω2uωL22mN21+2δM021λω2Nzω(x)2dx=mN21+2δM02Nwω(x)2dx.\displaystyle\frac{1}{\lambda_{\omega}^{2}}\|u_{\omega}\|_{L^{2}}^{2}\geqslant\frac{m_{*}^{\frac{N}{2}}}{1+2\delta M_{0}^{2}}\frac{1}{\lambda_{\omega}^{2}}\int_{\mathbb{R}^{N}}z_{\omega}(x)^{2}\,\mathrm{d}x=\frac{m_{*}^{\frac{N}{2}}}{1+2\delta M_{0}^{2}}\int_{\mathbb{R}^{N}}w_{\omega}(x)^{2}\,\mathrm{d}x.

On the other hand, using again Lemma 4.9 we obtain, for ω\omega small enough,

1λω2uωL224(m+1)N21λω2Nzω(x)2dx=4(m+1)N2Nwω(x)2dx.\displaystyle\frac{1}{\lambda_{\omega}^{2}}\|u_{\omega}\|_{L^{2}}^{2}\leqslant 4(m_{*}+1)^{\frac{N}{2}}\frac{1}{\lambda_{\omega}^{2}}\int_{\mathbb{R}^{N}}z_{\omega}(x)^{2}\,\mathrm{d}x=4(m_{*}+1)^{\frac{N}{2}}\int_{\mathbb{R}^{N}}w_{\omega}(x)^{2}\,\mathrm{d}x.

Lemma 4.13.

As ω0\omega\to 0, there holds

λω(N2)Nwω2|wω|2dxωλω2wωL22.\lambda_{\omega}^{-(N-2)}\int_{\mathbb{R}^{N}}w_{\omega}^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x\lesssim\omega\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}. (4.66)
Proof.

Firstly, by (1.6) and (1.7), we deduce using p+1=2=2NN2p+1=2^{*}=\frac{2N}{N-2} that

δNuω2|uω|2dx=1N2ωuωL22.\delta\int_{\mathbb{R}^{N}}u_{\omega}^{2}|\nabla u_{\omega}|^{2}\,\mathrm{d}x=\frac{1}{N-2}\omega\|u_{\omega}\|_{L^{2}}^{2}. (4.67)

We will now exploit this identity using the relations

uω(x)=r(vω(x))=r(zω(mω1/2x))=r(λωN22wω(λω1mω1/2x)).u_{\omega}(x)=r\big{(}v_{\omega}(x)\big{)}=r\big{(}z_{\omega}(m_{\omega}^{-1/2}x)\big{)}=r\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}(\lambda_{\omega}^{-1}m_{\omega}^{-1/2}x)\big{)}.

We have

uω(x)=mω1/2λωN/2r(λωN22wω(λω1mω1/2x))wω(λω1mω1/2x).\nabla u_{\omega}(x)=m_{\omega}^{-1/2}\lambda_{\omega}^{-N/2}r^{\prime}\big{(}\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}(\lambda_{\omega}^{-1}m_{\omega}^{-1/2}x)\big{)}\nabla w_{\omega}(\lambda_{\omega}^{-1}m_{\omega}^{-1/2}x).

On the one hand, by (4.65), there holds, for all ω(0,η0)\omega\in(0,\eta_{0}):

Nuω2|uω|2dx\displaystyle\int_{\mathbb{R}^{N}}u_{\omega}^{2}|\nabla u_{\omega}|^{2}\,\mathrm{d}x =mω1λωNNr(zω(mω1/2x))2r(zω(mω1/2x))2|wω(λω1mω1/2x)|2dx\displaystyle=m_{\omega}^{-1}\lambda_{\omega}^{-N}\int_{\mathbb{R}^{N}}r\big{(}z_{\omega}(m_{\omega}^{-1/2}x)\big{)}^{2}r^{\prime}\big{(}z_{\omega}(m_{\omega}^{-1/2}x)\big{)}^{2}|\nabla w_{\omega}(\lambda_{\omega}^{-1}m_{\omega}^{-1/2}x)|^{2}\,\mathrm{d}x
=mωN21λωNNr(zω(y))2r(zω(y))2|wω(λω1y)|2dy\displaystyle=m_{\omega}^{\frac{N}{2}-1}\lambda_{\omega}^{-N}\int_{\mathbb{R}^{N}}r\big{(}z_{\omega}(y)\big{)}^{2}r^{\prime}\big{(}z_{\omega}(y)\big{)}^{2}|\nabla w_{\omega}(\lambda_{\omega}^{-1}y)|^{2}\,\mathrm{d}y
=mωN21λωNNzω(y)2gω(y)r(zω(y))2|wω(λω1y)|2dy\displaystyle=m_{\omega}^{\frac{N}{2}-1}\lambda_{\omega}^{-N}\int_{\mathbb{R}^{N}}z_{\omega}(y)^{2}g_{\omega}(y)r^{\prime}\big{(}z_{\omega}(y)\big{)}^{2}|\nabla w_{\omega}(\lambda_{\omega}^{-1}y)|^{2}\,\mathrm{d}y
1(1+2δM02)2mωN21λωNNzω(y)2|wω(λω1y)|2dy\displaystyle\geqslant\frac{1}{(1+2\delta M_{0}^{2})^{2}}m_{\omega}^{\frac{N}{2}-1}\lambda_{\omega}^{-N}\int_{\mathbb{R}^{N}}z_{\omega}(y)^{2}|\nabla w_{\omega}(\lambda_{\omega}^{-1}y)|^{2}\,\mathrm{d}y
=1(1+2δM02)2mωN21λωNNλω(N2)wω(λω1y)2|wω(λω1y)|2dy\displaystyle=\frac{1}{(1+2\delta M_{0}^{2})^{2}}m_{\omega}^{\frac{N}{2}-1}\lambda_{\omega}^{-N}\int_{\mathbb{R}^{N}}\lambda_{\omega}^{-(N-2)}w_{\omega}(\lambda_{\omega}^{-1}y)^{2}|\nabla w_{\omega}(\lambda_{\omega}^{-1}y)|^{2}\,\mathrm{d}y
=1(1+2δM02)2mωN21Nλω(N2)wω(x)2|wω(x)|2dx.\displaystyle=\frac{1}{(1+2\delta M_{0}^{2})^{2}}m_{\omega}^{\frac{N}{2}-1}\int_{\mathbb{R}^{N}}\lambda_{\omega}^{-(N-2)}w_{\omega}(x)^{2}|\nabla w_{\omega}(x)|^{2}\,\mathrm{d}x.

Since mωmm_{\omega}\geqslant m_{*} by Lemma 4.9, we deduce that

Nuω2|uω|2dxλω(N2)Nwω2|wω|2dx.\int_{\mathbb{R}^{N}}u_{\omega}^{2}|\nabla u_{\omega}|^{2}\,\mathrm{d}x\gtrsim\lambda_{\omega}^{-(N-2)}\int_{\mathbb{R}^{N}}w_{\omega}^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x.

On the other hand, by Lemma 4.12, we have that

1N2ωuωL22ωλω2wωL22,ω0.\frac{1}{N-2}\omega\|u_{\omega}\|_{L^{2}}^{2}\lesssim\omega\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2},\quad\omega\to 0.

Hence, the result follows by (4.67). ∎

Lemma 4.14.

Consider 2s<4NN22\leqslant s<\frac{4N}{N-2}, s<q<4NN2s<q<\frac{4N}{N-2} and let θ=(N2)(qs)2s+N(4s)\theta=\frac{(N-2)(q-s)}{2s+N(4-s)}. Then there exists a constant C=C(N,s,q)>0C=C(N,s,q)>0 such that

uC0(N),N|u|qdxC(N|u|2|u|2dx)θN2N(N|u|sdx)1θ.\forall u\in C_{0}^{\infty}(\mathbb{R}^{N}),\quad\int_{\mathbb{R}^{N}}|u|^{q}\,\mathrm{d}x\leqslant C\Big{(}\int_{\mathbb{R}^{N}}|u|^{2}|\nabla u|^{2}\,\mathrm{d}x\Big{)}^{\theta\frac{N-2}{N}}\Big{(}\int_{\mathbb{R}^{N}}|u|^{s}\,\mathrm{d}x\Big{)}^{1-\theta}. (4.68)
Proof.

Apply the classical Gagliardo-Nirenberg inequality (see [40]) to |u|2|u|^{2}. ∎

We are now in a position to prove Proposition 4.11.

Proof of Proposition 4.11.

Extending (4.68) to XX by density and applying it to wωw_{\omega} with exponents q>2q>2^{*} and s=2s=2^{*}, we obtain

N|wω|qdxC(N|wω|2|wω|2dx)θN2N(N|wω|2dx)1θ.\int_{\mathbb{R}^{N}}|w_{\omega}|^{q}\,\mathrm{d}x\leqslant C\Big{(}\int_{\mathbb{R}^{N}}|w_{\omega}|^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x\Big{)}^{\theta\frac{N-2}{N}}\Big{(}\int_{\mathbb{R}^{N}}|w_{\omega}|^{2^{*}}\,\mathrm{d}x\Big{)}^{1-\theta}. (4.69)

Furthermore, the Hölder inequality yields a constant C>0C>0 such that

χB1(0)wωLqCχB1(0)wωL2C(χB1(0)WL2χB1(0)(Wwω)L2).\|\chi_{B_{1}(0)}w_{\omega}\|_{L^{q}}\geqslant C\|\chi_{B_{1}(0)}w_{\omega}\|_{L^{2^{*}}}\geqslant C\big{(}\|\chi_{B_{1}(0)}W\|_{L^{2^{*}}}-\|\chi_{B_{1}(0)}(W-w_{\omega})\|_{L^{2^{*}}}\big{)}.

Since wωWw_{\omega}\to W in L2(N)L^{2^{*}}(\mathbb{R}^{N}), it follows that

wωLqχB1(0)wωLqCχB1(0)WL2+o(1),ω0.\|w_{\omega}\|_{L^{q}}\geqslant\|\chi_{B_{1}(0)}w_{\omega}\|_{L^{q}}\geqslant C\|\chi_{B_{1}(0)}W\|_{L^{2^{*}}}+o(1),\quad\omega\to 0.

Therefore, (4.69) shows that N|wω|2|wω|2dx\int_{\mathbb{R}^{N}}|w_{\omega}|^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x is bounded away from zero. Hence, by (4.56) and (4.66),

λω(N2)δω.\lambda_{\omega}^{-(N-2)}\lesssim\delta_{\omega}. (4.70)

The conclusion now follows from (4.55). ∎

As was pointed out in Remark 8, for all N5N\geqslant 5 the upper and lower bounds on λω\lambda_{\omega} have the same blow-up rate. We shall now tighten the upper bounds on λω\lambda_{\omega} for N=3,4N=3,4.

Lemma 4.15.

For all R>0R>0, there exist η1=η1(R)(0,η0)\eta_{1}=\eta_{1}(R)\in(0,\eta_{0}) and C=C(R)>0C=C(R)>0 such that

ω(0,η1),xNBR(0),|wω(x)|Cemωωλω|x||x|N2.\forall\omega\in(0,\eta_{1}),\ \forall x\in\mathbb{R}^{N}\setminus B_{R}(0),\quad|w_{\omega}(x)|\geqslant C\frac{e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}}{|x|^{N-2}}. (4.71)
Proof.

The rescaled minimizer wωw_{\omega} defined in (4.46) solves the equation (4.61). As a consequence,

Δwω+mωωλω2wω\displaystyle-\Delta w_{\omega}+m_{\omega}\omega\lambda^{2}_{\omega}w_{\omega} =mωλω1+N2fω(λωN22wω)+mωωλω2wω\displaystyle=m_{\omega}\lambda_{\omega}^{1+\frac{N}{2}}f_{\omega}(\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega})+m_{\omega}\omega\lambda^{2}_{\omega}w_{\omega}
mωωλω2(λωN22r(λωN22wω)r(λωN22wω)+wω)\displaystyle\geqslant m_{\omega}\omega\lambda_{\omega}^{2}\left(-\lambda_{\omega}^{\frac{N-2}{2}}r^{\prime}(\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega})r(\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega})+w_{\omega}\right)
=mωωλω2k(λωN22wω)wω\displaystyle=m_{\omega}\omega\lambda_{\omega}^{2}k(\lambda_{\omega}^{-\frac{N-2}{2}}w_{\omega}){w_{\omega}}

with k:(0,)k:(0,\infty)\to\mathbb{R} defined by k(s)=1r(s)r(s)sk(s)=1-\frac{r^{\prime}(s)r(s)}{s}. Since 0r(s)10\leqslant r^{\prime}(s)\leqslant 1 and 0r(s)/s10\leqslant r(s)/s\leqslant 1 for all s(0,)s\in(0,\infty), we deduce that k(s)0k(s)\geqslant 0 for all s(0,)s\in(0,\infty). Hence,

Δwω(x)+mωωλω2wω(x)0-\Delta w_{\omega}(x)+m_{\omega}\omega\lambda^{2}_{\omega}w_{\omega}(x)\geqslant 0

for all xNx\in\mathbb{R}^{N}. For all xN{0}x\in\mathbb{R}^{N}\setminus\{0\}, let

hω(x)=|x|(N2)emωωλω|x|.h_{\omega}(x)=|x|^{-(N-2)}e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}.

A direct computation shows that, for all R>0R>0,

Δhω(x)+mωωλω2hω(x)=\displaystyle-\Delta h_{\omega}(x)+m_{\omega}\omega\lambda^{2}_{\omega}h_{\omega}(x)= |x|(N2)(ωmωλω2+(N1)|x|1mωωλω)emωωλω|x|\displaystyle\,|x|^{-(N-2)}\left(-\omega m_{\omega}\lambda^{2}_{\omega}+(N-1)|x|^{-1}\sqrt{m_{\omega}\omega}\lambda_{\omega}\right)e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}
2(N2)mωωλω|x|(N1)emωωλω|x|+mωωλω2|x|(N2)emωωλω|x|\displaystyle-2(N-2)\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|^{-(N-1)}e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}+m_{\omega}\omega\lambda_{\omega}^{2}|x|^{-(N-2)}e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}
=\displaystyle= (3N)mωωλω|x|(N1)emωωλω|x|0\displaystyle\,(3-N)\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|^{-(N-1)}e^{-\sqrt{m_{\omega}\omega}\lambda_{\omega}|x|}\leqslant 0

on NBR(0)\mathbb{R}^{N}\setminus B_{R}(0), provided that N3N\geqslant 3.

Next, thanks to Proposition 4.9 and Lemma 4.4, we have, for all R>0R>0,

(wωW)χB2R(0)BR/2(0)L0\|(w_{\omega}-W)\chi_{B_{2R}(0)\setminus B_{R/2}(0)}\|_{L^{\infty}}\to 0

as ω\omega goes to 0. As a consequence,

wω(R)12W(R)w_{\omega}(R)\geqslant\frac{1}{2}W(R)

for all ω\omega sufficiently small. Finally, we choose C=C(R)>0C=C(R)>0 such that

12W(R)CR(N2).\frac{1}{2}W(R)\geqslant CR^{-(N-2)}.

This implies wω(R)Chω(R)w_{\omega}(R)\geqslant Ch_{\omega}(R) and the conclusion follows by applying the maximum principle to the function wωChωw_{\omega}-Ch_{\omega}. ∎

As a consequence of (4.71), using the same arguments as in [39, Lemma 4.9, Lemma 4.11], we obtain the following.

Lemma 4.16.

As ω0\omega\to 0, there holds

wωL22{ω1/2λω1if N=3,log1ωλωif N=4.\|w_{\omega}\|_{L^{2}}^{2}\gtrsim\begin{cases}\omega^{-1/2}\lambda_{\omega}^{-1}&\text{if }N=3,\\ \log\frac{1}{\sqrt{\omega}\lambda_{\omega}}&\text{if }N=4.\end{cases} (4.72)

This is enough to conclude in the case N=3N=3.

Proposition 4.12.

Let N=3N=3. As ω0\omega\to 0, there holds

λωω14.\lambda_{\omega}\lesssim\omega^{-\frac{1}{4}}. (4.73)
Proof.

As in the proof of Proposition 4.10, we use estimate (4.56) but now with the refined estimate (4.72) instead of (4.57). This leads to

ωλω2ω1/2λω1ωλω2wωL22δω\omega\lambda^{2}_{\omega}\omega^{-1/2}\lambda_{\omega}^{-1}\lesssim\omega\lambda^{2}_{\omega}\|w_{\omega}\|^{2}_{L^{2}}\lesssim\delta_{\omega}

which implies

λωω1/2δω.\lambda_{\omega}\lesssim\omega^{-1/2}\delta_{\omega}.

The result then follows from (4.55). ∎

The case N=4N=4 is slightly more involved. Indeed, to obtain a matching lower and upper bound for λω\lambda_{\omega} we need to improve the upper bound for δω\delta_{\omega}.

Proposition 4.13.

Let N=4N=4. As ω0\omega\to 0, there holds

δω(ωlog1ω)12 and (ωlog1ω)14λω(ωlog1ω)14.\delta_{\omega}\lesssim\left(\omega\log\frac{1}{\omega}\right)^{\frac{1}{2}}\text{ and }\left(\omega\log\frac{1}{\omega}\right)^{-\frac{1}{4}}\lesssim\lambda_{\omega}\lesssim\left(\omega\log\frac{1}{\omega}\right)^{-\frac{1}{4}}. (4.74)
Remark 9.

Heuristically, we know from the proof of Proposition 4.11 that, for N=4N=4,

λω2δω1.\lambda_{\omega}^{2}\gtrsim\delta_{\omega}^{-1}.

On the other hand, estimates (4.56) and (4.72) imply that

δωωλω2wωL22ωλω2log1ωλω.\displaystyle\delta_{\omega}\gtrsim\omega\lambda^{2}_{\omega}\|w_{\omega}\|^{2}_{L^{2}}\gtrsim\omega\lambda^{2}_{\omega}\log\frac{1}{\sqrt{\omega}\lambda_{\omega}}.

Now, since ω13ωλωω16\omega^{\frac{1}{3}}\lesssim\sqrt{\omega}\lambda_{\omega}\lesssim\omega^{\frac{1}{6}}, we deduce

log1ωλωlog1ω\log\frac{1}{\sqrt{\omega}\lambda_{\omega}}\gtrsim\log\frac{1}{\omega}

so that

δωωδω1log1ωδω2ωlog1ω.\delta_{\omega}\gtrsim\omega\delta_{\omega}^{-1}\log\frac{1}{\omega}\implies\delta_{\omega}^{2}\gtrsim\omega\log\frac{1}{\omega}.

Hence, the bound (4.55) can be improved in order to obtain a matching upper and lower bound for λω\lambda_{\omega}.

Proof.

From the proof of Lemma 4.9, we know that, for N=4N=4,

(ηRWλ)2r(ηRWλ)222ωr(ηRWλ)21+O((R/λ)4)+O(λ2)+O(ωλ2log(R/λ))\frac{\int(\eta_{R}W_{\lambda})^{2^{*}}}{\int r(\eta_{R}W_{\lambda})^{2^{*}}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}}\leqslant 1+O((R/\lambda)^{-4})+O(\lambda^{-2})+O(\omega\lambda^{2}\log(R/\lambda))

provided that λ1\lambda\gg 1, Rλ1\frac{R}{\lambda}\gg 1 and ωλ2log(R/λ)1\omega\lambda^{2}\log(R/\lambda)\ll 1. Hence, we let λ=(ωlog1ω)1/4\lambda=\left(\omega\log\frac{1}{\omega}\right)^{-1/4} and R=λαR=\lambda^{\alpha} with α>1\alpha>1 to be chosen later. We have

1(R/λ)N=1λ4(α1)=(ωlog1ω)(α1),1λ2=1λ4(α1)=(ωlog1ω)12\frac{1}{(R/\lambda)^{N}}=\frac{1}{\lambda^{4(\alpha-1)}}=\left(\omega\log\frac{1}{\omega}\right)^{(\alpha-1)},\quad\frac{1}{\lambda^{2}}=\frac{1}{\lambda^{4(\alpha-1)}}=\left(\omega\log\frac{1}{\omega}\right)^{\frac{1}{2}}

and

O(ωλ2log(R/λ))=O(ωλ2logλ)=O((ωlog1ω)1/2ωlog1ω)=O((ωlog1ω)1/2).O(\omega\lambda^{2}\log(R/\lambda))=O(\omega\lambda^{2}\log\lambda)=O\left(\left(\omega\log\frac{1}{\omega}\right)^{-1/2}\omega\log\frac{1}{\omega}\right)=O\left(\left(\omega\log\frac{1}{\omega}\right)^{1/2}\right).

This implies, by taking α32\alpha\geqslant\frac{3}{2},

(ηRWλ)2r(ηRWλ)222ωr(ηRWλ)21+O((ωlog1ω)1/2)\frac{\int(\eta_{R}W_{\lambda})^{2^{*}}}{\int r(\eta_{R}W_{\lambda})^{2^{*}}-\frac{2^{*}}{2}\omega\int r(\eta_{R}W_{\lambda})^{2}}\leqslant 1+O\left(\left(\omega\log\frac{1}{\omega}\right)^{1/2}\right)

so that

δω(ωlog1ω)1/2.\delta_{\omega}\lesssim\left(\omega\log\frac{1}{\omega}\right)^{1/2}.

Arguing as in the proof of Proposition 4.11, we get the lower bound

λωδω12(ωlog1ω)1/4.\lambda_{\omega}\gtrsim\delta_{\omega}^{-\frac{1}{2}}\gtrsim\left(\omega\log\frac{1}{\omega}\right)^{-1/4}.

On the other hand, estimates (4.59), (4.56) and (4.72) imply that

(ωlog1ω)1/2δωωλω2wωL22ωλω2log1ωλωωλω2log1ω\displaystyle\left(\omega\log\frac{1}{\omega}\right)^{1/2}\gtrsim\delta_{\omega}\gtrsim\omega\lambda^{2}_{\omega}\|w_{\omega}\|^{2}_{L^{2}}\gtrsim\omega\lambda^{2}_{\omega}\log\frac{1}{\sqrt{\omega}\lambda_{\omega}}\gtrsim\omega\lambda^{2}_{\omega}\log\frac{1}{\omega}

which leads to

λω(ωlog1ω)1/4,\lambda_{\omega}\lesssim\left(\omega\log\frac{1}{\omega}\right)^{-1/4},

as claimed. ∎

Lemma 4.17.

For all q(2,4NN2)q\in(2^{*},\frac{4N}{N-2}),

wωLq1,ω0.\|w_{\omega}\|_{L^{q}}\lesssim 1,\quad\omega\to 0. (4.75)
Proof.

We shall again use (4.69). We already know that wωL2\|w_{\omega}\|_{L^{2^{*}}} is bounded. Furthermore, by (4.55), (4.56), (4.66), (4.73) and (4.74), we have

N|wω|2|wω|2dxλωN2δω{ω14ω14=1,if N=3,(ωlog1ω)12(ωlog1ω)12=1,if N=4,ωN2NωN2N=1,if N5,\int_{\mathbb{R}^{N}}|w_{\omega}|^{2}|\nabla w_{\omega}|^{2}\,\mathrm{d}x\lesssim\lambda_{\omega}^{N-2}\delta_{\omega}\lesssim\begin{cases}\omega^{-\frac{1}{4}}\omega^{\frac{1}{4}}=1,&\text{if }N=3,\\ \left(\omega\log\frac{1}{\omega}\right)^{-\frac{1}{2}}\left(\omega\log\frac{1}{\omega}\right)^{\frac{1}{2}}=1,&\text{if }N=4,\\ \omega^{-\frac{N-2}{N}}\omega^{\frac{N-2}{N}}=1,&\text{if }N\geqslant 5,\end{cases}

which completes the proof. ∎

As a consequence of Lemma 4.17, the following LL^{\infty}-bound can be proved similarly to Lemma 4.5 (i), using (4.61) instead of (5.7).

Lemma 4.18.

There exists η2>0\eta_{2}>0 and M2>0M_{2}>0 such that

sup0<ω<η2wωLM2.\sup_{0<\omega<\eta_{2}}\|w_{\omega}\|_{L^{\infty}}\leqslant M_{2}. (4.76)

Thanks to this estimate, we can now establish C2C^{2}-convergence of wωw_{\omega}.

Proposition 4.14.

Consider (ωn)(0,)(\omega_{n})\subset(0,\infty) such that ωn0\omega_{n}\to 0 as nn\to\infty. Then

wωnWinC2(N),as n.w_{\omega_{n}}\to W\quad\text{in}\quad C^{2}(\mathbb{R}^{N}),\ \text{as }n\to\infty.
Proof.

Fix an arbitrary s>max{N,2NN2}s>\max\{N,\frac{2N}{N-2}\}. We will first prove that

wωnWW2,s(N)0,n.\|w_{\omega_{n}}-W\|_{W^{2,s}(\mathbb{R}^{N})}\to 0,\quad n\to\infty. (4.77)

This convergence follows by interpolation between

wωnWLs(N)0\|w_{\omega_{n}}-W\|_{L^{s}(\mathbb{R}^{N})}\to 0 (4.78)

and

ΔwωnΔWLs(N)0.\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(\mathbb{R}^{N})}\to 0. (4.79)

To prove (4.78), we split wωnWLs(N)s\|w_{\omega_{n}}-W\|_{L^{s}(\mathbb{R}^{N})}^{s} into

wωnWLs(N)s=wωnWLs(B1(0))s+wωnWLs(NB1(0))s.\|w_{\omega_{n}}-W\|_{L^{s}(\mathbb{R}^{N})}^{s}=\|w_{\omega_{n}}-W\|_{L^{s}(B_{1}(0))}^{s}+\|w_{\omega_{n}}-W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{1}(0))}^{s}.

On the one hand, thanks to Lemma 4.18, wωnWLs(B1(0))s0\|w_{\omega_{n}}-W\|_{L^{s}(B_{1}(0))}^{s}\to 0 by dominated convergence. On the other, wωnWLs(NB1(0))s0\|w_{\omega_{n}}-W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{1}(0))}^{s}\to 0 by Lemma 4.4 (ii).

To prove (4.79), let R>0R>0 and write

ΔwωnΔWLs(N)s=ΔwωnΔWLs(BR(0))s+ΔwωnΔWLs(NBR(0))s.\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(\mathbb{R}^{N})}^{s}=\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(B_{R}(0))}^{s}+\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s}.

Using the elliptic equations (4.61) and (4.44) satisfied by wωnw_{\omega_{n}} and WW, respectively, we have

ΔwωnΔW=(λωn1+N2mωnfωn(λωnN22wωn)mWN+2N2).\Delta w_{\omega_{n}}-\Delta W=-\Big{(}\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})-m_{*}W^{\frac{N+2}{N-2}}\Big{)}.

Since ωλω20\omega\lambda_{\omega}^{2}\to 0 as ω0\omega\to 0 by (4.54), it is easy to show that

λωn1+N2mωnfωn(λωnN22wωn)mWN+2N2a.e. on N,n.\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})\to m_{*}W^{\frac{N+2}{N-2}}\quad\text{a.e. on }\mathbb{R}^{N},\quad n\to\infty. (4.80)

It follows by dominated convergence and Lemma 4.18 that ΔwωnΔWLs(BR(0))s0\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(B_{R}(0))}^{s}\to 0, for any fixed R>0R>0.

As for ΔwωnΔWLs(NBR(0))s\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s}, we now show that it can be made as small as desired by choosing R>0R>0 large enough. On the one hand, since W(x)|x|N22W(x)\lesssim|x|^{-\frac{N-2}{2}}, we deduce that

WLs(NBR(0))s=|x|RWN+2N2sdxRNN+22s.\|W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s}=\int_{|x|\geqslant R}W^{\frac{N+2}{N-2}s}\,\mathrm{d}x\lesssim R^{N-\frac{N+2}{2}s}. (4.81)

On the other hand, since

|fω(s)|r(s)N+2N2+ωr(s)sN+2N2+ωs,|f_{\omega}(s)|\leqslant r(s)^{\frac{N+2}{N-2}}+\omega r(s)\leqslant s^{\frac{N+2}{N-2}}+\omega s,

we have that

λωn1+N2mωnfωn(λωnN22wωn)Ls(NBR(0))s\displaystyle\|\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s}
=|x|Rλωn(1+N2)smωns|fωn(λωnN22wωn)|sdx\displaystyle=\int_{|x|\geqslant R}\lambda_{\omega_{n}}^{(1+\frac{N}{2})s}m_{\omega_{n}}^{s}\big{|}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})\big{|}^{s}\,\mathrm{d}x
|x|RλωnN+22s(λωnN22wωn)N+2N2sdx+|x|RλωnN+22s(ωnλωnN22wωn)sdx\displaystyle\lesssim\int_{|x|\geqslant R}\lambda_{\omega_{n}}^{\frac{N+2}{2}s}\big{(}\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}}\big{)}^{\frac{N+2}{N-2}s}\,\mathrm{d}x+\int_{|x|\geqslant R}\lambda_{\omega_{n}}^{\frac{N+2}{2}s}\big{(}\omega_{n}\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}}\big{)}^{s}\,\mathrm{d}x
|x|RwωnN+2N2sdx+|x|R(ωnλωn2)swωnsdx.\displaystyle\lesssim\int_{|x|\geqslant R}w_{\omega_{n}}^{\frac{N+2}{N-2}s}\,\mathrm{d}x+\int_{|x|\geqslant R}(\omega_{n}\lambda_{\omega_{n}}^{2})^{s}w_{\omega_{n}}^{s}\,\mathrm{d}x.

Furthermore, since {wωn}\{w_{\omega_{n}}\} is bounded in L2(N)L^{2^{*}}(\mathbb{R}^{N}), it follows by Lemma 4.4 (i) that there exists a constant CNC_{N} such that

n,|wωn(x)|CN|x|N22.\forall n\in\mathbb{N},\quad|w_{\omega_{n}}(x)|\leqslant C_{N}|x|^{-\frac{N-2}{2}}.

Thus, using again ωλω20\omega\lambda_{\omega}^{2}\to 0 as ω0\omega\to 0, we find that

λωn1+N2mωnfωn(λωnN22wωn)Ls(NBR(0))s\displaystyle\|\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s} |x|R|x|N+22sdx+|x|R|x|N22sdx\displaystyle\lesssim\int_{|x|\geqslant R}|x|^{-\frac{N+2}{2}s}\,\mathrm{d}x+\int_{|x|\geqslant R}|x|^{-\frac{N-2}{2}s}\,\mathrm{d}x
RNN+22s+RNN22s.\displaystyle\lesssim R^{N-\frac{N+2}{2}s}+R^{N-\frac{N-2}{2}s}. (4.82)

Since s>2NN2s>\frac{2N}{N-2}, it follows from (4.81) and (4.3.1) that, given any ε>0\varepsilon>0, we can choose R>0R>0 so large that

ΔwωnΔWLs(NBR(0))s<ε.\|\Delta w_{\omega_{n}}-\Delta W\|_{L^{s}(\mathbb{R}^{N}\setminus B_{R}(0))}^{s}<\varepsilon.

This completes the proof of (4.79) and thus of (4.77).

Since s>Ns>N, we deduce that wωnWw_{\omega_{n}}\to W in C1(N)C^{1}(\mathbb{R}^{N}). We now bootstrap this to wωnWw_{\omega_{n}}\to W in C2(N)C^{2}(\mathbb{R}^{N}) by using the ODE’s satisfied by the radial functions wωnw_{\omega_{n}} and WW:

wωn′′N1rwωn=λωn1+N2mωnfωn(λωnN22wωn),-w_{\omega_{n}}^{\prime\prime}-\frac{N-1}{r}w_{\omega_{n}}^{\prime}=\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}}), (4.83)
W′′N1rW=mWN+2N2,-W^{\prime\prime}-\frac{N-1}{r}W^{\prime}=m_{*}W^{\frac{N+2}{N-2}}, (4.84)

where denotes differentiation with respect to r(0,)r\in(0,\infty). We use the same notation for wωnw_{\omega_{n}}, WW and their radial counterparts, i.e. wωn(r)w_{\omega_{n}}(r), W(r)W(r), with r=|x|r=|x|.

We first note that, using wωnWL0\|w_{\omega_{n}}-W\|_{L^{\infty}}\to 0, the pointwise convergence in (4.80) can be improved to

λωn1+N2mωnfωn(λωnN22wωn)mWN+2N2L0,n.\Big{\|}\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})-m_{*}W^{\frac{N+2}{N-2}}\Big{\|}_{L^{\infty}}\to 0,\quad n\to\infty. (4.85)

Subtracting (4.83) from (4.84), multiplying by rN1r^{N-1} and integrating yields

rN1(wωn(r)W(r))=0rsN1(λωn1+N2mωnfωn(λωnN22wωn(s))mWN+2N2(s))ds.r^{N-1}\left(w_{\omega_{n}}^{\prime}(r)-W^{\prime}(r)\right)=-\int_{0}^{r}s^{N-1}\left(\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}}(s))-m_{*}W^{\frac{N+2}{N-2}}(s)\right)\,\mathrm{d}s.

With the change of variables s=rts=rt, this identity becomes

wωn(r)W(r)r=01tN1(λωn1+N2mωnfωn(λωnN22wωn(rt))mWN+2N2(rt))dt\frac{w_{\omega_{n}}^{\prime}(r)-W^{\prime}(r)}{r}=-\int_{0}^{1}t^{N-1}\left(\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}}(rt))-m_{*}W^{\frac{N+2}{N-2}}(rt)\right)\,\mathrm{d}t

and it follows that

|wωn(r)W(r)r|λωn1+N2mωnfωn(λωnN22wωn)mWN+2N2L01tN1dt.\left|\frac{w_{\omega_{n}}^{\prime}(r)-W^{\prime}(r)}{r}\right|\leqslant\Big{\|}\lambda_{\omega_{n}}^{1+\frac{N}{2}}m_{\omega_{n}}f_{\omega_{n}}(\lambda_{\omega_{n}}^{-\frac{N-2}{2}}w_{\omega_{n}})-m_{*}W^{\frac{N+2}{N-2}}\Big{\|}_{L^{\infty}}\int_{0}^{1}t^{N-1}\,\mathrm{d}t.

Hence, by (4.85),

wωn(r)W(r)rL0,n.\left\|\frac{w_{\omega_{n}}^{\prime}(r)-W^{\prime}(r)}{r}\right\|_{L^{\infty}}\to 0,\quad n\to\infty. (4.86)

It then follows from (4.83), (4.84) and (4.86) that

wωn′′(r)W′′(r)L0,n.\left\|w_{\omega_{n}}^{\prime\prime}(r)-W^{\prime\prime}(r)\right\|_{L^{\infty}}\to 0,\quad n\to\infty.

Since we already know that wωnWw_{\omega_{n}}\to W in C1(N)C^{1}(\mathbb{R}^{N}), this completes the proof. ∎

4.3.2. Asymptotic behavior of M(ω)=uωL22M(\omega)=\|u_{\omega}\|^{2}_{L^{2}} and M(ω)M^{\prime}(\omega) as ω0\omega\to 0.

As a consequence of Lemma 4.12 and the exact asymptotic behavior of λω\lambda_{\omega}, we deduce the following proposition which gives the behavior of M(ω)M(\omega) as ω\omega goes to 0.

Proposition 4.15.

Let N3N\geqslant 3 and p=N+2N2p=\frac{N+2}{N-2}. As ω0\omega\to 0, we have

limω0uωL22=+.\lim_{\omega\to 0}\|u_{\omega}\|^{2}_{L^{2}}=+\infty.

More precisely,

{ω34M(ω)ω34if N=3,(1ωlog1ω)12M(ω)(1ωlog1ω)12if N=4,ω2NM(ω)ω2Nif N5.\begin{cases}\omega^{-\frac{3}{4}}\lesssim M(\omega)\lesssim\omega^{-\frac{3}{4}}&\text{if }N=3,\\ \left(\frac{1}{\omega}\log\frac{1}{\omega}\right)^{\frac{1}{2}}\lesssim M(\omega)\lesssim\left(\frac{1}{\omega}\log\frac{1}{\omega}\right)^{\frac{1}{2}}&\text{if }N=4,\\ \omega^{-\frac{2}{N}}\lesssim M(\omega)\lesssim\omega^{-\frac{2}{N}}&\text{if }N\geqslant 5.\end{cases} (4.87)
Proof.

From Lemma 4.12, we know that, as ω0\omega\to 0,

λω2wωL22M(ω)λω2wωL22.\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}\gtrsim M(\omega)\gtrsim\lambda_{\omega}^{2}\|w_{\omega}\|_{L^{2}}^{2}.

Moreover, using (4.56) and (4.57) for N5N\geqslant 5 or (4.72) for N{3,4}N\in\{3,4\}, together with the exact behavior of λω\lambda_{\omega}, we deduce that

{ω14wωL22ω14if N=3,log1ωwωL22log1ωif N=4,1wωL221if N5.\begin{cases}\omega^{-\frac{1}{4}}\lesssim\|w_{\omega}\|_{L^{2}}^{2}\lesssim\omega^{-\frac{1}{4}}&\text{if }N=3,\\ \log\frac{1}{\omega}\lesssim\|w_{\omega}\|_{L^{2}}^{2}\lesssim\log\frac{1}{\omega}&\text{if }N=4,\\ 1\lesssim\|w_{\omega}\|_{L^{2}}^{2}\lesssim 1&\text{if }N\geqslant 5.\\ \end{cases}

Combining these estimates with (1.18) completes the proof. ∎

To deduce the asymptotic behavior of M(ω)M^{\prime}(\omega), we proceed as in Proposition 4.6.

Proposition 4.16.

Let N3N\geqslant 3 and p=N+2N2p=\frac{N+2}{N-2}. Then, for ω>0\omega>0 small enough, we have

M(ω)2\displaystyle\frac{M^{\prime}(\omega)}{2} [2β(ω)+(N2)]<14M(ω)ω[N(N2)(β(ω)1)4]\displaystyle\,\left[2\beta(\omega)+(N-2)\right]<\frac{1}{4}\frac{M(\omega)}{\omega}\left[N(N-2)(\beta(\omega)-1)-4\right] (4.88)

where

T(ω)=N|uω(x)|2dx,β(ω)=T(ω)1Nuω(x)2dx.T(\omega)=\int_{\mathbb{R}^{N}}|\nabla u_{\omega}(x)|^{2}\,\mathrm{d}x,\quad\beta(\omega)=T(\omega)^{-1}\int_{\mathbb{R}^{N}}u_{\omega}(x)^{2^{*}}\,\mathrm{d}x.

Moreover,

limω0+M(ω)=.\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=-\infty.
Proof.

As in the proof of Proposition 4.6, for any ω>0\omega>0, let (ω)=L+\mathcal{L}(\omega)=L_{+} be defined by (1.9) and L=(Lij)L=(L_{ij}) be the symmetric matrix given by the restriction of (ω)\mathcal{L}(\omega) to the finite dimensional space spanned by {ωuω,uω,xuω+N2uω}\left\{\partial_{\omega}u_{\omega},u_{\omega},x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right\}.

When p=N+2N2p=\frac{N+2}{N-2}, the same arguments detailed above give

L11:=\displaystyle L_{11}:= ωuω,(ω)ωuω=M(ω)2,L12:=ωuω,(ω)uω=M(ω),\displaystyle\,\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)\partial_{\omega}u_{\omega}\rangle=-\frac{M^{\prime}(\omega)}{2},\quad L_{12}:=\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)u_{\omega}\rangle=-M(\omega),
L13:=\displaystyle L_{13}:= ωuω,(ω)(xuω+N2uω)=uω,xuω+N2uω=0,\displaystyle\,\langle\partial_{\omega}u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=-\langle u_{\omega},x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\rangle=0,
L22:=\displaystyle L_{22}:= uω,(ω)uω=8δQ(ω)4N2β(ω)T(ω),\displaystyle\,\langle u_{\omega},\mathcal{L}(\omega)u_{\omega}\rangle=8\delta Q(\omega)-\frac{4}{N-2}\beta(\omega)T(\omega),
L23:=\displaystyle L_{23}:= uω,(ω)(xuω+N2uω)=4δNQ(ω)4N2β(ω)T(ω)2ωM(ω),\displaystyle\,\langle u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=4\delta NQ(\omega)-\frac{4}{N-2}\beta(\omega)T(\omega)-2\omega M(\omega),
L33:=\displaystyle L_{33}:= xuω+N2uω,(ω)(xuω+N2uω)=2δN(N+2)2Q(ω)4N2β(ω)T(ω),\displaystyle\,\langle x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega},\mathcal{L}(\omega)\left(x\cdot\nabla u_{\omega}+\frac{N}{2}u_{\omega}\right)\rangle=2\delta\frac{N(N+2)}{2}Q(\omega)-\frac{4}{N-2}\beta(\omega)T(\omega),

with Q(ω)=Nuω(x)2|uω(x)|2dxQ(\omega)=\int_{\mathbb{R}^{N}}u_{\omega}(x)^{2}|\nabla u_{\omega}(x)|^{2}\,\mathrm{d}x.

In the critical case, the integral identities of Proposition 1.1 give

{2δQ(ω)=2N2ωM(ω),β(ω)=1+N+2N2ωM(ω)T(ω).\begin{cases}&2\delta Q(\omega)=\frac{2}{N-2}\omega M(\omega),\\ &\beta(\omega)=1+\frac{N+2}{N-2}\frac{\omega M(\omega)}{T(\omega)}.\end{cases} (4.89)

As a consequence,

L22=\displaystyle L_{22}= 4N2T(ω)[2ωM(ω)T(ω)β(ω)],\displaystyle\,\frac{4}{N-2}T(\omega)\left[2\frac{\omega M(\omega)}{T(\omega)}-\beta(\omega)\right],
L23=\displaystyle L_{23}= 2N2T(ω)[(N+2)ωM(ω)T(ω)2β(ω)],\displaystyle\,\frac{2}{N-2}T(\omega)\left[(N+2)\frac{\omega M(\omega)}{T(\omega)}-2\beta(\omega)\right],
L33=\displaystyle L_{33}= 1N2T(ω)[N(N+2)ωM(ω)T(ω)4β(ω)],\displaystyle\,\frac{1}{N-2}T(\omega)\left[N(N+2)\frac{\omega M(\omega)}{T(\omega)}-4\beta(\omega)\right],

and

L22L33\displaystyle L_{22}L_{33} L232=4N2ωM(ω)T(ω)[(N+2)ωM(ω)T(ω)Nβ(ω)]\displaystyle-L^{2}_{23}=\frac{4}{N-2}\omega M(\omega)T(\omega)\left[(N+2)\frac{\omega M(\omega)}{T(\omega)}-N\beta(\omega)\right]
=4N2ωM(ω)T(ω)[2β(ω)(N2)]<0\displaystyle=\frac{4}{N-2}\omega M(\omega)T(\omega)\left[-2\beta(\omega)-(N-2)\right]<0

for all ω>0\omega>0.

Since (ω)\mathcal{L}(\omega) has a unique negative eigenvalue, we deduce that the determinant of LL is negative:

0>det(L)=M(ω)2(L232L22L33)1N2M(ω)2T(ω)[N(N2)(β(ω)1)4].\displaystyle 0>\det(L)=\frac{M^{\prime}(\omega)}{2}(L^{2}_{23}-L_{22}L_{33})-\frac{1}{N-2}M(\omega)^{2}T(\omega)\left[N(N-2)(\beta(\omega)-1)-4\right].

This gives the estimate (4.88).

In the limit ω0\omega\to 0, ωM(ω)\omega M(\omega) goes 0 which implies limω0β(ω)=1\lim_{\omega\to 0}\beta(\omega)=1. As a consequence,

[N(N2)(β(ω)1)4]ω04\displaystyle\left[N(N-2)(\beta(\omega)-1)-4\right]\underset{\omega\to 0}{\sim}-4

and, for ω>0\omega>0 small enough,

M(ω)<cM(ω)ωM^{\prime}(\omega)<-c\frac{M(\omega)}{\omega}

for some positive constant cc. This implies,

limω0+M(ω)=.\lim_{\omega\to 0^{+}}M^{\prime}(\omega)=-\infty.

Collecting the main results of Section 4.3, we can now complete the proof of Theorem 1.3.

Proof of Theorem 1.3 (ii).

The proof follows from Propositions 4.9, 4.14, 4.15 and 4.16. ∎

5. Appendix

Proof of Lemma 4.4.

For the whole proof, we denote by u~(r),v~(r)\widetilde{u}(r),\widetilde{v}(r) etc., the functions on +\mathbb{R}_{+} defined by u~(r)=u(x),v~(r)=v(x)\widetilde{u}(r)=u(x),\widetilde{v}(r)=v(x), etc., for r=|x|r=|x|. Any radial uH˙1(N)u\in\dot{H}^{1}(\mathbb{R}^{N}) has a representative such that u~\widetilde{u} is continuous on +\mathbb{R}_{+}, and u~L2(+;rN1dr)\widetilde{u}^{\prime}\in L^{2}(\mathbb{R}_{+};r^{N-1}\,\mathrm{d}r) is related to u\nabla u by

uL22=|𝕊N1|+|u~(r)|2rN1dr.\|\nabla u\|_{L^{2}}^{2}=|\mathbb{S}^{N-1}|\int_{\mathbb{R}_{+}}|\widetilde{u}^{\prime}(r)|^{2}r^{N-1}\,\mathrm{d}r.

Part (i) follows from the estimate

uLss|𝕊N1|0r|u~(t)|stN1dt|𝕊N1||u~(r)|srNN,r>0.\|u\|_{L^{s}}^{s}\geqslant|\mathbb{S}^{N-1}|\int_{0}^{r}|\widetilde{u}(t)|^{s}t^{N-1}\,\mathrm{d}t\geqslant|\mathbb{S}^{N-1}||\widetilde{u}(r)|^{s}\frac{r^{N}}{N},\quad\forall r>0.

To prove (ii), we fix an arbitrary R>0R>0. We first address the convergence in Lq(NBR(0))L^{q}(\mathbb{R}^{N}\setminus B_{R}(0)). Since unH˙1\|u_{n}\|_{\dot{H}^{1}} is bounded, there exists uH˙1(N)u\in\dot{H}^{1}(\mathbb{R}^{N}) such that, up to a subsequence, unu_{n} converges to uu, weakly in H˙1(N)\dot{H}^{1}(\mathbb{R}^{N}) and a.e. on N\mathbb{R}^{N}, as nn\to\infty. Applying (4.17) with s=2NN2s=\frac{2N}{N-2} and the Sobolev embedding theorem, there is a constant CN>0C_{N}>0 (independent of nn) such that

|un(x)u(x)|CNunuL2|x|N22.|u_{n}(x)-u(x)|\leqslant C_{N}\|\nabla u_{n}-\nabla u\|_{L^{2}}|x|^{-\frac{N-2}{2}}.

Hence, for nn large,

|un(x)u(x)|quL2q|x|N22qL1(NBR(0)),q>2.|u_{n}(x)-u(x)|^{q}\lesssim\|\nabla u\|_{L^{2}}^{q}|x|^{-\frac{N-2}{2}q}\in L^{1}(\mathbb{R}^{N}\setminus B_{R}(0)),\quad\forall q>2^{*}.

Dominated convergence then implies unuu_{n}\to u in Lq(NBR(0))L^{q}(\mathbb{R}^{N}\setminus B_{R}(0)).

To prove convergence in L(NBR(0))L^{\infty}(\mathbb{R}^{N}\setminus B_{R}(0)), we let vn=unuv_{n}=u_{n}-u and we show that

limnsuprR|v~n(r)|=0.\lim_{n\to\infty}\sup_{r\geqslant R}|\widetilde{v}_{n}(r)|=0.

We first apply (4.17) again with s=2NN2s=\frac{2N}{N-2}, using the fact that vnL2\|\nabla v_{n}\|_{L^{2}} is bounded. Given ε>0\varepsilon>0, there exist Rε>RR_{\varepsilon}>R and a constant C>0C>0 such that,

n,supr>Rε|v~n(r)|=sup|x|>Rε|vn(x)|CRεN22<ε.\forall n\in\mathbb{N},\quad\sup_{r>R_{\varepsilon}}|\widetilde{v}_{n}(r)|=\sup_{|x|>R_{\varepsilon}}|v_{n}(x)|\leqslant CR_{\varepsilon}^{-\frac{N-2}{2}}<\varepsilon.

Hence, we only need to show that v~n(r)0\widetilde{v}_{n}(r)\to 0 as nn\to\infty, uniformly for r[R,Rε]r\in[R,R_{\varepsilon}]. By an Arzela-Ascoli type argument which will be made explicit below, this is a consequence of the equicontinuity of the sequence {v~n}\{\widetilde{v}_{n}\} on [R,Rε][R,R_{\varepsilon}], which we prove now.

Let s,tR,s<ts,t\geqslant R,\ s<t. Since {vn}\{v_{n}\} is bounded in H˙1\dot{H}^{1}, there exists M>0M>0 such that,

n,|v~n(t)v~n(s)|\displaystyle\forall n\in\mathbb{N},\quad|\widetilde{v}_{n}(t)-\widetilde{v}_{n}(s)| st|v~n(σ)|dσ\displaystyle\leqslant\int_{s}^{t}|\widetilde{v}_{n}^{\prime}(\sigma)|\,\mathrm{d}\sigma (5.1)
(ts)1/2(st|v~n(σ)|2dσ)1/2\displaystyle\leqslant(t-s)^{1/2}\Big{(}\int_{s}^{t}|\widetilde{v}_{n}^{\prime}(\sigma)|^{2}\,\mathrm{d}\sigma\Big{)}^{1/2} (5.2)
M(ts)1/2.\displaystyle\leqslant M(t-s)^{1/2}. (5.3)

Hence, {v~n}\{\widetilde{v}_{n}\} is equicontinuous on [R,)[R,\infty).

We now claim that v~n0\widetilde{v}_{n}\to 0 as nn\to\infty for all r[R,Rε]r\in[R,R_{\varepsilon}]. We will prove below that the convergence is uniform. To complete the proof, the claim can be proved by similar arguments, using convergence almost everywhere and equicontinuity of {v~n}\{\widetilde{v}_{n}\} on [R,Rε][R,R_{\varepsilon}].

Suppose by contradiction that {v~n}\{\widetilde{v}_{n}\} does not converge uniformly to 0 on [R,Rε][R,R_{\varepsilon}]: there exists ε>0\varepsilon>0 such that, for all nn\in\mathbb{N} there exists rn[R,Rε]r_{n}\in[R,R_{\varepsilon}] such that

|v~n(rn)|ε.|\widetilde{v}_{n}(r_{n})|\geqslant\varepsilon.

There exists a subsequence (rnj)[R,Rε](r_{n_{j}})\subset[R,R_{\varepsilon}] and a point r[R,Rε]r^{*}\in[R,R_{\varepsilon}] such that rnjrr_{n_{j}}\to r^{*} as jj\to\infty. Suppose without loss of generality that r(R,Rε)r^{*}\in(R,R_{\varepsilon}). By equicontinuity, there exists δ>0\delta>0 such that

n,r(rδ,r+δ)[R,Rε],|v~n(r)v~n(r)|<ε2.\forall n\in\mathbb{N},\ \forall r\in(r^{*}-\delta,r^{*}+\delta)\subset[R,R_{\varepsilon}],\quad|\widetilde{v}_{n}(r)-\widetilde{v}_{n}(r^{*})|<\frac{\varepsilon}{2}.

Now choose NεN_{\varepsilon}\in\mathbb{N} such that rnj(rδ,r+δ)r_{n_{j}}\in(r^{*}-\delta,r^{*}+\delta) for all jNεj\geqslant N_{\varepsilon}. It follows that,

jNε,|v~nj(r)|||v~nj(rnj)||v~nj(rnj)v~nj(r)||ε2.\forall j\geqslant N_{\varepsilon},\quad|\widetilde{v}_{n_{j}}(r^{*})|\geqslant||\widetilde{v}_{n_{j}}(r_{n_{j}})|-|\widetilde{v}_{n_{j}}(r_{n_{j}})-\widetilde{v}_{n_{j}}(r^{*})||\geqslant\frac{\varepsilon}{2}.

This contradicts the pointwise convergence v~nj(r)0\widetilde{v}_{n_{j}}(r^{*})\to 0 and finishes the proof. ∎

Proof of Lemma 4.5.

We start by proving part (ii), which will be used in the proof of part (i). To prove estimate (4.21), we follow the scheme of proof laid down in [1, Lemma 5.5]. Let

q0=2=2NN2andq=p12.q_{0}=2^{*}=\frac{2N}{N-2}\quad\text{and}\quad q=\frac{p-1}{2}.

Since p<3N+2N2p<\frac{3N+2}{N-2}, there holds

q0q>1.q_{0}-q>1.

Multiplying both sides of (4.22) by zωq0qz_{\omega}^{q_{0}-q} and integrating by parts, one has

Nzω(zωq0q)dx=mωNfω(zω)zωq0qdx.\int_{\mathbb{R}^{N}}\nabla z_{\omega}\cdot\nabla(z_{\omega}^{q_{0}-q})\,\mathrm{d}x=m_{\omega}\int_{\mathbb{R}^{N}}f_{\omega}(z_{\omega})z_{\omega}^{q_{0}-q}\,\mathrm{d}x. (5.4)

Since zωL22=mω\|\nabla z_{\omega}\|_{L^{2}}^{2}=m_{\omega}, it follows by Lemma 4.2/4.9 and the Sobolev embedding theorem that there exists η>0\eta>0 and M>0M>0 such that

ω(0,η),zωLq0M.\forall\omega\in(0,\eta),\quad\|z_{\omega}\|_{L^{q_{0}}}\leqslant M.

Hence, using (2.5) and Lemma 4.2/4.9, there is a constant CC such that, for all ω(0,η)\omega\in(0,\eta),

mωNfω(zω)zωq0qdxmωC0Nzωqzωq0qdxCNzωq0dxCMq0,m_{\omega}\int_{\mathbb{R}^{N}}f_{\omega}(z_{\omega})z_{\omega}^{q_{0}-q}\,\mathrm{d}x\leqslant m_{\omega}C_{0}\int_{\mathbb{R}^{N}}z_{\omega}^{q}z_{\omega}^{q_{0}-q}\,\mathrm{d}x\leqslant C\int_{\mathbb{R}^{N}}z_{\omega}^{q_{0}}\,\mathrm{d}x\leqslant CM^{q_{0}}, (5.5)

On the other hand,

Nzω(zωq0q)dx=(q0q)Nzωq0q1|zω|2dx.\int_{\mathbb{R}^{N}}\nabla z_{\omega}\cdot\nabla(z_{\omega}^{q_{0}-q})\,\mathrm{d}x=(q_{0}-q)\int_{\mathbb{R}^{N}}z_{\omega}^{q_{0}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x.

Hence, by (5.4) and (5.5),

(q0q)Nzωq0q1|zω|2dxCMq0.(q_{0}-q)\int_{\mathbb{R}^{N}}z_{\omega}^{q_{0}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x\leqslant CM^{q_{0}}. (5.6)

Let

q1=q02(q0q+1)>q0.q_{1}=\frac{q_{0}}{2}(q_{0}-q+1)>q_{0}.

A direct calculation shows that

(q0q)Nzωq0q1|zω|2dx=4(q0q)(q0q+1)2N|(zωq0q+12)|2dx.(q_{0}-q)\int_{\mathbb{R}^{N}}z_{\omega}^{q_{0}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x=\frac{4(q_{0}-q)}{(q_{0}-q+1)^{2}}\int_{\mathbb{R}^{N}}\Big{|}\nabla\Big{(}z_{\omega}^{\frac{q_{0}-q+1}{2}}\Big{)}\Big{|}^{2}\,\mathrm{d}x.

By the Sobolev embedding, there is a constant CSC_{S} such that

zωq0q+12Lq0CS(zωq0q+12)L2.\Big{\|}z_{\omega}^{\frac{q_{0}-q+1}{2}}\Big{\|}_{L^{q_{0}}}\leqslant C_{S}\Big{\|}\nabla\Big{(}z_{\omega}^{\frac{q_{0}-q+1}{2}}\Big{)}\Big{\|}_{L^{2}}.

Therefore,

(q0q)Nzωq0q1|zω|2dx4(q0q)(q0q+1)21CS2(Nzωq02(q0q+1)dx)2q0.(q_{0}-q)\int_{\mathbb{R}^{N}}z_{\omega}^{q_{0}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x\geqslant\frac{4(q_{0}-q)}{(q_{0}-q+1)^{2}}\frac{1}{C_{S}^{2}}\Big{(}\int_{\mathbb{R}^{N}}z_{\omega}^{\frac{q_{0}}{2}(q_{0}-q+1)}\,\mathrm{d}x\Big{)}^{\frac{2}{q_{0}}}.

By (5.6), there exists a constant C1>0C_{1}>0 such that

ω(0,η),Nzωq1dxC1.\forall\omega\in(0,\eta),\quad\int_{\mathbb{R}^{N}}z_{\omega}^{q_{1}}\,\mathrm{d}x\leqslant C_{1}.

The proof now follows by iteration. Let

qi=q02(qi1q+1)>qi1,i.q_{i}=\frac{q_{0}}{2}(q_{i-1}-q+1)>q_{i-1},\quad i\in\mathbb{N}^{*}.

For each ii\in\mathbb{N}^{*}, multiplying both sides of (4.22) by zωqi1qz_{\omega}^{q_{i-1}-q} yields

(qi1q)Nzωqi1q1|zω|2dxCMqi1(q_{i-1}-q)\int_{\mathbb{R}^{N}}z_{\omega}^{q_{i-1}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x\leqslant CM^{q_{i-1}}

instead of (5.6)\eqref{eq:moser_estimate_1}. Next, the above argument shows that

(Nzωqidx)2q0Nzωqi1q1|zω|2dx,\Big{(}\int_{\mathbb{R}^{N}}z_{\omega}^{q_{i}}\,\mathrm{d}x\Big{)}^{\frac{2}{q_{0}}}\lesssim\int_{\mathbb{R}^{N}}z_{\omega}^{q_{i-1}-q-1}|\nabla z_{\omega}|^{2}\,\mathrm{d}x,

and we conclude that there exists a constant Ci>0C_{i}>0 such that

ω(0,η),NzωqidxCi.\forall\omega\in(0,\eta),\quad\int_{\mathbb{R}^{N}}z_{\omega}^{q_{i}}\,\mathrm{d}x\leqslant C_{i}.

Finally, it is an easy exercise to show that qiq_{i}\to\infty as ii\to\infty. The conclusion thus follows by Hölder interpolation.

We next prove part (i). First recall that, for any ω>0\omega>0, zωz_{\omega} is positive, radial decreasing, and satisfies the elliptic equation

Δzω=cω(x)zω,-\Delta z_{\omega}=c_{\omega}(x)z_{\omega}, (5.7)

where

cω(x):=mωfω(zω(x))zω(x),xN.c_{\omega}(x):=\frac{m_{\omega}f_{\omega}(z_{\omega}(x))}{z_{\omega}(x)},\quad x\in\mathbb{R}^{N}.

By Lemma 4.9, there exist η0>0\eta_{0}>0 and C>0C>0 such that

ω(0,η0),|cω(x)|C(r(zω(x))p1+ω)C(zω(x)4N2+ω).\displaystyle\forall\omega\in(0,\eta_{0}),\quad|c_{\omega}(x)|\leqslant C\big{(}r(z_{\omega}(x))^{p-1}+\omega\big{)}\leqslant C\big{(}z_{\omega}(x)^{\frac{4}{N-2}}+\omega\big{)}.

Furthermore, by part (ii) proved above, for any s2s\geqslant 2^{*} there exists a constant Cs>0C_{s}>0 such that

ω(0,η0),zωLs(N)Cs.\forall\omega\in(0,\eta_{0}),\quad\|z_{\omega}\|_{L^{s}(\mathbb{R}^{N})}\leqslant C_{s}. (5.8)

Hence, by the Radial Lemma, taking CsC_{s} larger if needed,

ω(0,η0),zω(x)Cs|x|N/s.\forall\omega\in(0,\eta_{0}),\quad z_{\omega}(x)\leqslant C_{s}|x|^{-N/s}.

Thus,

zω(x)4N2=O(|x|4Ns(N2)),z_{\omega}(x)^{\frac{4}{N-2}}=O\big{(}|x|^{-\frac{4N}{s(N-2)}}\big{)},

uniformly for ω(0,η0)\omega\in(0,\eta_{0}). Choosing s>2NN2s>\frac{2N}{N-2}, it is possible to find q(N2,sN24)q\in(\frac{N}{2},s\frac{N-2}{4}), so that Nqs4NN2>0N-\frac{q}{s}\frac{4N}{N-2}>0 and hence cωLq(B1(0))\|c_{\omega}\|_{L^{q}(B_{1}(0))} is bounded, for ω(0,η0)\omega\in(0,\eta_{0}). Then, by Theorem 5.1 and Remark 5.1 in [49], there is a constant K>0K>0, independent of ω\omega, such that

ω(0,η0),zωL(B1/2(0))KzωL2(B1(0))2.\forall\omega\in(0,\eta_{0}),\quad\|z_{\omega}\|_{L^{\infty}(B_{1/2}(0))}\leqslant K\|z_{\omega}\|_{L^{2}(B_{1}(0))}^{2}. (5.9)

Since zωz_{\omega} converges in L2(B1(0))L^{2}(B_{1}(0)) as ω0\omega\to 0, it follows that zωL(B1/2(0))\|z_{\omega}\|_{L^{\infty}(B_{1/2}(0))} is bounded, for ω(0,η0)\omega\in(0,\eta_{0}). Finally, it follows by part (ii) of Lemma 4.4 that zωL(NB1/2(0))\|z_{\omega}\|_{L^{\infty}(\mathbb{R}^{N}\setminus B_{1/2}(0))} is also bounded, which completes the proof. ∎

References

  • [1] S. Adachi, M. Shibata, and T. Watanabe, Blow-up phenomena and asymptotic profiles of ground states of quasilinear elliptic equations with H1H^{1}-supercritical nonlinearities, J. Differential Equations, 256 (2014), pp. 1492–1514.
  • [2]  , Blow-up phenomena and asymptotic profiles of ground states of quasilinear elliptic equations with H1H^{1}-supercritical nonlinearities, J. Differential Equations, 256 (2014), pp. 1492–1514.
  • [3] S. Adachi, M. Shibata, and T. Watanabe, A note on the uniqueness and the non-degeneracy of positive radial solutions for semilinear elliptic problems and its application, Acta Math. Sci. Ser. B (Engl. Ed.), 38 (2018), pp. 1121–1142.
  • [4] S. Adachi, M. Shibata, and T. Watanabe, Asymptotic property of ground states for a class of quasilinear Schrödinger equation with H1H^{1}-critical growth, Calc. Var. Partial Differential Equations, 58 (2019), pp. Paper No. 88, 29.
  • [5] S. Adachi and T. Watanabe, Asymptotic properties of ground states of quasilinear Schrödinger equations with h1h^{1}-subcritical exponent, Adv. Nonlinear Stud., 12 (2012), pp. 255–279.
  • [6]  , Asymptotic uniqueness of ground states for a class of quasilinear Schrödinger equations with h1-supercritical exponent, J. Differential Equations, 260 (2016), pp. 3086–3118.
  • [7] H. Berestycki, T. Gallouet, and O. Kavian, Équations de champs scalaires euclidiens non linéaires dans le plan, C. R. Acad. Sci. Paris Ser I Math., 297 (1983), pp. 307–310.
  • [8] H. Berestycki and P.-L. Lions, Nonlinear scalar field equations. I. Existence of a ground state, Arch. Rational Mech. Anal., 82 (1983), pp. 313–345.
  • [9] L. Brüll, H. Lange, and E. de Jager, Stationary, oscillatory and solitary wave type solution of singular nonlinear Schrödinger equations, Mathematical Methods in the Applied Sciences, 8 (1986), pp. 559–575.
  • [10] C. C. Chen and C. S. Lin, Uniqueness of the ground state solutions of Δu+f(u)=0\Delta u+f(u)=0 in 𝐑n,n3{\bf R}^{n},\;n\geq 3, Comm. Partial Differential Equations, 16 (1991), pp. 1549–1572.
  • [11] C. V. Coffman, Uniqueness of the ground state solution for Δuu+u3=0\Delta u-u+u^{3}=0 and a variational characterization of other solutions, Arch. Rational Mech. Anal., 46 (1972), pp. 81–95.
  • [12] M. Colin, On the local well posedness of quasilinear Schrödinger equations in arbitrary space dimension, Comm. Partial Differential Equations, 27 (2002), pp. 325–354.
  • [13] M. Colin, L. Jeanjean, and M. Squassina, Stability and instability results for standing waves of quasi-linear Schrödinger equations, Nonlinearity, 23 (2010), p. 1353.
  • [14] S. De Bièvre, F. Genoud, and S. Rota Nodari, Orbital stability: analysis meets geometry, in Nonlinear optical and atomic systems. At the interface of physics and mathematics, Cham: Springer; Lille: Centre européen pour les mathématiques, la physique et leurs interactions (CEMPI), 2015, pp. 147–273.
  • [15] S. De Bièvre and S. Rota Nodari, Orbital stability via the energy-momentum method: the case of higher dimensional symmetry groups, Arch. Rational Mech. Anal., 231 (2019), pp. 233–284.
  • [16] M. Flucher and S. Müller, Radial symmetry and decay rate of variational ground states in the zero mass case, SIAM J. Math. Anal., 29 (1998), pp. 712–719.
  • [17] R. L. Frank, Ground states of semi-linear PDE. Lecture notes from the “Summerschool on Current Topics in Mathematical Physics”, CIRM Marseille, Sept. 2013.
  • [18] B. Gidas, W. M. Ni, and L. Nirenberg, Symmetry of positive solutions of nonlinear elliptic equations in 𝐑n{\bf R}^{n}, Adv. in Math. Suppl. Stud., 7 (1981), pp. 369–402.
  • [19] F. Gladiali and M. Squassina, Uniqueness of ground states for a class of quasi-linear elliptic equations, Adv. Nonlinear Anal., 1 (2012), pp. 159–179.
  • [20] M. V. Goldman, Strong turbulence of plasma waves, Rev. Mod. Phys., 56 (1984), pp. 709–735.
  • [21] M. Grillakis, J. Shatah, and W. Strauss, Stability theory of solitary waves in the presence of symmetry. I, J. Funct. Anal., 74 (1987), pp. 160–197.
  • [22]  , Stability theory of solitary waves in the presence of symmetry. II, J. Funct. Anal., 94 (1990), pp. 308–348.
  • [23] R. D. Hajer Bahouri, Jean-Yves Chemin, Fourier Analysis and Nonlinear Partial Differential Equations, Grundlehren der mathematischen Wissenschaften, Springer Berlin, Heidelberg, 2011.
  • [24] J. Jang, Uniqueness of positive radial solutions of Δu+f(u)=0\Delta u+f(u)=0 in N\mathbb{R}^{N}, N2N\geq 2, Nonlinear Anal., 73 (2010), pp. 2189–2198.
  • [25] L. Jeanjean and T. Luo, Sharp nonexistence results of prescribed L2L^{2}-norm solutions for some class of Schrödinger–Poisson and quasi-linear equations, Z. Angew. Math., 64 (2013), pp. 937–954.
  • [26] L. Jeanjean, T. Luo, and Z.-Q. Wang, Multiple normalized solutions for quasi-linear Schrödinger equations, J. Differential Equations, 259 (2015), pp. 3894–3928.
  • [27] T. Kapitula and K. Promislow, Orbital Stability of Waves in Hamiltonian Systems, Springer New York, New York, NY, 2013, pp. 117–157.
  • [28] R. Killip, T. Oh, O. Pocovnicu, and M. Vişan, Solitons and scattering for the cubic–quintic nonlinear Schrödinger equation on 3\mathbb{R}^{3}, Arch. Rational Mech. Anal., 225 (2017), pp. 469–548.
  • [29] 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), pp. 243–266.
  • [30] E. W. Laedke, K. H. Spatschek, and L. Stenflo, Evolution theorem for a class of perturbed envelope soliton solutions, J. Math. Phys., 24 (1983), pp. 2764–2769.
  • [31] M. Lewin and S. Rota Nodari, Uniqueness and non-degeneracy for a nuclear nonlinear Schrödinger equation, NoDEA Nonlinear Differential Equations Appl., 22 (2015), pp. 673–698.
  • [32] M. Lewin and S. Rota Nodari, The double-power nonlinear Schrödinger equation and its generalizations: uniqueness, non-degeneracy and applications, Calc. Var. Partial Differential Equations, 59 (2020), pp. Paper No. 197, 49.
  • [33] P.-L. Lions, The concentration-compactness principle in the calculus of variations. The limit case. I, Rev. Mat. Iberoamericana, 1 (1985), pp. 145–201.
  • [34] J.-Q. Liu, Y.-Q. Wang, and Z.-Q. Wang, Soliton solutions for quasilinear Schrödinger equations, ii, J. Differential Equations, 187 (2003), pp. 473–493.
  • [35] J. H. Maddocks, Restricted quadratic forms and their application to bifurcation and stability in constrained variational principles, SIAM J. Math. Anal., 16 (1985), pp. 47–68.
  • [36]  , Errata: Restricted quadratic forms and their application to bifurcation and stability in constrained variational principles, SIAM J. Math. Anal., 19 (1988), pp. 1256–1257.
  • [37] K. McLeod, Uniqueness of positive radial solutions of Δu+f(u)=0-\Delta u+f(u)=0 in RnR^{n}, II, Trans. Amer. Math. Soc., 339 (1993), pp. 495–505.
  • [38] K. McLeod and J. Serrin, Uniqueness of positive radial solutions of Δu+f(u)=0\Delta u+f(u)=0 in 𝐑n{\bf R}^{n}, Arch. Rational Mech. Anal., 99 (1987), pp. 115–145.
  • [39] V. Moroz and C. B. Muratov, Asymptotic properties of ground states of scalar field equations with a vanishing parameter, J. Eur. Math. Soc. (JEMS), 16 (2014).
  • [40] L. Nirenberg, On elliptic partial differential equations, Ann. Scuola Norm. Sup. Pisa Cl. Sci., Ser. 3, 13 (1959), pp. 115–162.
  • [41] L. A. Peletier and J. Serrin, Uniqueness of positive solutions of semilinear equations in 𝐑n{\bf R}^{n}, Arch. Rational Mech. Anal., 81 (1983), pp. 181–197.
  • [42] M. Poppenberg, On the local well posedness of quasilinear Schrödinger equations in arbitrary space dimension, J. Differential Equations, 172 (2001), pp. 83–115.
  • [43] M. Porkolab and M. V. Goldman, Upper‐hybrid solitons and oscillating‐two‐stream instabilities, The Physics of Fluids, 19 (1976), pp. 872–881.
  • [44] P. Pucci and J. Serrin, Uniqueness of ground states for quasilinear elliptic operators, Indiana Univ. Math. J., 47 (1998), pp. 501–528.
  • [45] M. Reed and B. Simon, Methods of Modern Mathematical Physics. I. Functional analysis, Academic Press, 1972.
  • [46]  , Methods of Modern Mathematical Physics. IV. Functional analysis, Academic Press, 1972.
  • [47] A. Selvitella, Uniqueness and nondegeneracy of the ground state for a quasilinear Schrödinger equation with a small parameter, Nonlinear Anal., 74 (2011), pp. 1731–1737.
  • [48] J. Serrin and M. Tang, Uniqueness of ground states for quasilinear elliptic equations, Indiana Univ. Math. J., 49 (2000), pp. 897–923.
  • [49] G. Stampacchia, Le problème de Dirichlet pour les équations elliptiques du second ordre à coefficients discontinus, Ann. Inst. Fourier (Grenoble), 15 (1965), pp. 189–258.
  • [50] G. Talenti, Best constant in Sobolev inequality, Ann. Mat. Pura Appl. (4), 110 (1976), pp. 353–372.
  • [51] M. Tang, Uniqueness and global structure of positive radial solutions for quasilinear elliptic equations, Comm. Partial Differential Equations, 26 (2001), pp. 909–938.
  • [52] M. I. Weinstein, Modulational stability of ground states of nonlinear Schrödinger equations, SIAM J. Math. Anal., 16 (1985), pp. 472–491.
  • [53] M. Willem, Minimax theorems, vol. 24 of Progress in Nonlinear Differential Equations and their Applications, Birkhäuser Boston, Inc., Boston, MA, 1996.
  • [54] H. Ye and Y. Yu, The existence of normalized solutions for L2L^{2}-critical quasilinear Schrödinger equations, J. Math. Anal. Appl., 497 (2021), p. 124839.