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

Scattering for critical radial Neumann waves outside a ball

Thomas Duyckaerts and David Lafontaine Thomas Duyckaerts, LAGA, UMR 7539, Institut Galilée, Université Sorbonne Paris Nord, 93430 - Villetaneuse, France [email protected] David Lafontaine, Department of Mathematical Sciences, University of Bath, Bath, BA2 7AY, United Kingdom [email protected]
Abstract.

We show that the solutions of the three-dimensional critical defocusing nonlinear wave equation with Neumann boundary conditions outside a ball and radial initial data scatter. This is to our knowledge the first result of scattering for a nonlinear wave equation with Neumann boundary conditions. Our proof uses the scheme of concentration-compactness/rigidity introduced by Kenig and Merle, extending it to our setup, together with the so-called channels of energy method to rule out compact-flow solutions. We also obtain, for the focusing equation, the same exact scattering/blow-up dichotomy below the energy of the ground-state as in 3\mathbb{R}^{3}.

Key words and phrases:
Wave equation. Exterior domain. Scattering
2010 Mathematics Subject Classification:
Primary: 35L70; Secondary: 34B15, 34D05.
T. Duyckaerts is supported by the Institut Universitaire de France and partially supported by the Labex MME-DII. D. Lafontaine acknowledges support from EPSRC Grant EP/1025995/1.

1. Introduction

This work concerns the energy-critical wave equation outside an obstacle of 3\mathbb{R}^{3} with Neumann boundary condition:

(1.1) t2uΔu+ιu5\displaystyle\partial_{t}^{2}u-\Delta u+\iota u^{5} =0,in Ω\displaystyle=0,\quad\text{in }\Omega
(1.2) nu\displaystyle\partial_{n}u =0,in Ω\displaystyle=0,\quad\text{in }\partial\Omega
(1.3) ut=0\displaystyle\vec{u}_{\restriction t=0} =(u0,u1)H˙1(Ω)×L2(Ω),\displaystyle=(u_{0},u_{1})\in\dot{H}^{1}(\Omega)\times L^{2}(\Omega),

where Ω=3K\Omega=\mathbb{R}^{3}\setminus K, KK is a compact subset of 3\mathbb{R}^{3} with smooth boundary, nu\partial_{n}u is the normal derivative of uu on the boundary Ω\partial\Omega of Ω\Omega, u\vec{u} denotes (u,tu)(u,\partial_{t}u), and ι{±1}\iota\in\{\pm 1\}. In our main result we will treat the case where KK is the unit ball of 3\mathbb{R}^{3} and the initial data (u0,u1)(u_{0},u_{1}) is assumed to be radial.

The equation (1.1), (1.2), (1.3) is locally well-posed (see [BP09]). The energy

(u(t))=12Ω|u(t,x)|2𝑑x+12Ω|tu(t,x)|2𝑑x+ι6Ωu6(t,x)𝑑x\mathscr{E}(\vec{u}(t))=\frac{1}{2}\int_{\Omega}|\nabla u(t,x)|^{2}dx+\frac{1}{2}\int_{\Omega}|\partial_{t}u(t,x)|^{2}dx+\frac{\iota}{6}\int_{\Omega}u^{6}(t,x)dx

is conserved. When ι=1\iota=1 (defocusing case), the energy yields a uniform bound of the norm of the solution in H˙1×L2\dot{H}^{1}\times L^{2} and solutions are expected to be global and to scatter to linear solutions (see definition below). When ι=1\iota=-1 (focusing case), one can easily construct, using the differential equation u′′=u5u^{\prime\prime}=u^{5} and finite-speed of propagation, solutions with initial data in H˙1×L2\dot{H}^{1}\times L^{2} that blow up in finite time.

We first consider the defocusing case ι=1\iota=1. When there is no obstacle (Ω=3\Omega=\mathbb{R}^{3}), global existence was obtained for smooth radial data by Struwe [Str88], and extended to smooth non-radial data by Grillakis [Gri90]. Global existence for data in the energy space was then proved by Shatah and Struwe [SS94]. Bahouri and Shatah [BS98] have shown that any solution uu to the defocusing equation scatters to a linear solution, i.e. there exists a solution uLu_{L} of the free wave equation

(1.4) t2uLΔuL=0\partial_{t}^{2}u_{L}-\Delta u_{L}=0

on ×3\mathbb{R}\times\mathbb{R}^{3} such that

limt+u(t)uL(t)H˙×L2=0.\lim_{t\to+\infty}\|\vec{u}(t)-\vec{u}_{L}(t)\|_{\dot{H}\times L^{2}}=0.

The scattering is proved as a consequence of the fact that the L6L^{6} norm of the solution goes to 0, which is obtained by multipliers techniques involving integration by parts on the wave cone {|x||t|}\{|x|\leq|t|\}.

The equation (1.1) with Dirichlet boundary condition:

(1.5) uΩ=0,ut=0=(u0,u1)H˙01(Ω)×L2(Ω),u_{\restriction\partial\Omega}=0,\quad\vec{u}_{\restriction t=0}=(u_{0},u_{1})\in\dot{H}^{1}_{0}(\Omega)\times L^{2}(\Omega),

where H˙01(Ω)={fH1(Ω),fΩ=0}\dot{H}^{1}_{0}(\Omega)=\{f\in H^{1}(\Omega),\;f_{\restriction\partial\Omega}=0\}, was studied in several articles. The global well-posedness is proved in [BLP08]. The local well-posedness follows from a local-in-time Strichartz estimate, which is a direct consequence of a spectral projector estimate of Smith and Sogge [SS07]. The global well-posedness is obtained by the same arguments as in the case without obstacle, observing that the boundary term appearing in the integration by parts can be dealt with a commutator estimate.

The asymptotic behaviour of equations (1.1) and (1.4) with Dirichlet boundary conditions (1.5) is not known in general, and depends on geometrical assumptions on the obstacle. When KK is non-trapping, for the linear equation (1.4), [MRS77] proved the exponential decay of the local energy in odd dimensions, polynomial in even dimensions, for compactly supported initial data. A related estimate is the integrability of the local energy, introduced in [Bur03]

(1.6) (χu,χtu)L2(,H1×L2)χ(u0,u1)H˙1×L2,\|(\chi u,\chi\partial_{t}u)\|_{L^{2}(\mathbb{R},H^{1}\times L^{2})}\lesssim_{\chi}\|(u_{0},u_{1})\|_{\dot{H}^{1}\times L^{2}},

where χ\chi is an arbitrary smooth compactly supported function. In odd space dimensions, the exponential decay of the local energy was first used by [SS00] to show global-in-time Strichartz estimates. This result was then extended independently to all space dimensions by [Bur03] and [Met04]. The general argument of Burq [Bur03] shows that (1.6), together with the local-in-time Strichartz estimates, imply global Strichartz estimates. When the obstacle KK is moreover assumed to be star-shaped, the same computation as in the article of Bahouri and Shatah [BS98] yields that any solution scatter to a linear solution. The only difference with the case without obstacle is that boundary terms appear in the integration by parts. The key point is that when Ω\Omega is star-shaped and uu satisfies Dirichlet boundary conditions, these boundary terms come with a good sign, so that the proof is still valid in this case. This argument can be extended to illuminated obstacles, that are generalisations of star-shaped obstacles, as done in [AS13, AS14] adapting the multiplier so that the boundary term as the right sign, and in [Laf17] showing that it decays to zero. However, the case of a general non-trapping obstacle seems at the moment out of hand due to the rigidity of the Morawetz multiplier arguments used for now.

Much less is known in the case of Neumann boundary conditions. Note that these boundary conditions are more challenging than the Dirichlet boundary conditions, as they do not make sense in the energy space. Also, the strong Huygens principle is lost in this case (see Proposition 2.4 below).

Local-in-time Strichartz inequalities for the linear wave equation with Neumann boundary condition were obtained by Blair, Smith and Sogge [BSS09], and global existence for equation (1.1)-(1.2) with ι=1\iota=1 by Burq and Planchon [BP09]. Exponential decay of the local energy in the three-dimensional case was shown by [Mor75]. Combined with the local-in-time Strichartz estimates [BSS09], this should lead to global in time Strichartz estimates by the arguments of [SS00]. We give a direct proof of (1.6) (see Proposition 2.6) when the obstacle is the unit ball and the solution is radial, which implies global Strichartz estimates by the main result of [Bur03].

The asymptotic behaviour of the solutions of the nonlinear equation (1.1)-(1.3) was to our knowledge not previously investigated. Assuming the global Strichartz estimates for the linear wave equation, the proof of scattering in [BS98] does not work anymore since the boundary terms appearing in the integration by parts do not seem to have any specific signs and cannot be controlled.

The main result of this article is that the scattering to a linear solution holds for the defocusing wave equation with Neumann boundary conditions, when KK is the unit ball of 3\mathbb{R}^{3} and (u0,u1)(u_{0},u_{1}) is radially symmetric. We thus consider the equation

(1.7) t2uΔu+u5\displaystyle\partial_{t}^{2}u-\Delta u+u^{5} =0,in 3B(0,1)\displaystyle=0,\quad\text{in }\mathbb{R}^{3}\setminus B(0,1)
(1.8) ru\displaystyle\partial_{r}u =0,for r=1\displaystyle=0,\quad\text{for }r=1
(1.9) ut=0\displaystyle\vec{u}_{\restriction t=0} =(u0,u1)(Bc),\displaystyle=(u_{0},u_{1})\in\mathcal{H}(B^{c}),

where B(0,1)B(0,1) is the unit ball of 3\mathbb{R}^{3} and (Bc)\mathcal{H}(B^{c}) is the space of radial functions in (H˙1×L2)(3B(0,1))(\dot{H}^{1}\times L^{2})\left(\mathbb{R}^{3}\setminus B(0,1)\right), and the corresponding linear wave equation:

(1.10) t2uLΔuL\displaystyle\partial_{t}^{2}u_{L}-\Delta u_{L} =0,in 3B(0,1).\displaystyle=0,\quad\text{in }\mathbb{R}^{3}\setminus B(0,1).

with the boundary condition (1.8).

Theorem 1.1.

Let uu be a solution of (1.7) with Neumann boundary condition (1.8) and initial data (1.9). Then uu is global and there exists a solution uLu_{L} of (1.10), (1.8), with initial data in (Bc)\mathcal{H}{(B^{c})}, such that

limt+u(t)uL(t)(Bc)=0.\lim_{t\to+\infty}\|\vec{u}(t)-\vec{u}_{L}(t)\|_{\mathcal{H}{(B^{c})}}=0.

Our proof uses and extends the by now standard compactness/rigidity scheme introduced by Kenig and Merle in [KM06], [KM08] to study the focusing energy-critical Schrödinger and wave equations on N\mathbb{R}^{N}. The compactness step consists in constructing, in a contradiction argument, a global nonzero solution ucu_{c} of (1.7), (1.8) such that

{uc(t),t}\{\vec{u}_{c}(t),\;t\in\mathbb{R}\}

has compact closure in \mathcal{H}. The essential tool of this construction is a profile decomposition (in the spirit of the one introduced by Bahouri and Gérard [BG99] on the whole space), describing the defect of compactness of the Strichartz inequality uLL5(,L10)(u0,u1)\|u_{L}\|_{L^{5}(\mathbb{R},L^{10})}\lesssim\|(u_{0},u_{1})\|_{\mathcal{H}} for solutions of (1.10), (1.8). We construct this profile decomposition, which is new for the wave equation with Neumann boundary conditions, in Section 4. In this decomposition, the linear wave equation on the whole space appears as a limiting equation for dilating profiles, as shown in Section 3. The knowledge of the fact that any solution of the defocusing equation on the whole space scatters is essential to rule out these profiles and obtain the critical solution ucu_{c}, constructed in Section 5.

The second step of the proof (the rigidity argument), carried out in Section 6 consists in ruling out the existence of the critical solution. Since no monotonicity formula is available due to the Neumann boundary condition, we use the channels of energy method introduced in [DKM11], [DKM13] to classify solutions of the focusing energy-critical wave equation on 3\mathbb{R}^{3}. Using this method, we prove that ucu_{c} must be independent of time, a contradiction with the well-known fact that there is no stationary solution of (1.7) with boundary conditions (1.8) in H˙1\dot{H}^{1}. This idea was first used in the context of the supercritical wave equation in [DKM12].

Our method also gives scattering for solutions of the focusing wave equation:

(1.11) t2uΔuu5=0,in 3B(0,1),\partial_{t}^{2}u-\Delta u-u^{5}=0,\quad\text{in }\mathbb{R}^{3}\setminus B(0,1),

with Neumann boundary condition (1.8) below a natural energy threshold. Let WW be the ground state of the equation on 3\mathbb{R}^{3}:

W=(1+|x|2/3)1/2W=\left(1+|x|^{2}/3\right)^{-1/2}

and recall that WW is (up to scaling and sign change), the unique radial, stationary solution of ΔW=W5-\Delta W=W^{5} [Poh65, GNN81]. Denote by 3(W,0)\mathscr{E}_{\mathbb{R}^{3}}(W,0) the energy of the solution (W,0)(W,0) on the whole space 3\mathbb{R}^{3}:

3(W,0):=123|W|2163W6.\mathscr{E}_{\mathbb{R}^{3}}(W,0):=\frac{1}{2}\int_{\mathbb{R}^{3}}|\nabla W|^{2}-\frac{1}{6}\int_{\mathbb{R}^{3}}W^{6}.

Then we have the following:

Theorem 1.2.

Let uu be a solution of (1.11), (1.8) with initial data (1.9). Assume

(u0,u1)<3(W,0),3B(0,1)|u0(x)|2𝑑x<3|W(x)|2𝑑x.\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0),\quad\int_{\mathbb{R}^{3}\setminus B(0,1)}|\nabla u_{0}(x)|^{2}dx<\int_{\mathbb{R}^{3}}|\nabla W(x)|^{2}dx.

Then uu is global,

t,3B(0,1)|u(t,x)|2𝑑x<3|W(x)|2𝑑x,\forall t\in\mathbb{R},\quad\int_{\mathbb{R}^{3}\setminus B(0,1)}|\nabla u(t,x)|^{2}dx<\int_{\mathbb{R}^{3}}|\nabla W(x)|^{2}dx,

and uu scatters to a linear solution.

Finally, we have exactly the same dichotomy as in 3\mathbb{R}^{3} for the solutions below the energy threshold 3(W,0)\mathscr{E}_{\mathbb{R}^{3}}(W,0), indeed, with the same proof as in [KM08], one obtains:

Theorem 1.3.

Let uu be a solution of (1.11), (1.8) with initial data (1.9). Assume

(u0,u1)<3(W,0),3B(0,1)|u0|2𝑑x>3|W|2𝑑x.\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0),\quad\int_{\mathbb{R}^{3}\setminus B(0,1)}|\nabla u_{0}|^{2}dx>\int_{\mathbb{R}^{3}}|\nabla W|^{2}dx.

Then uu blows up in finite time.

Noting that by variational arguments (see Proposition 7.1), using that WW is a maximizer to the critical Sobolev inequilality on 3\mathbb{R}^{3}, one cannot have (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0) and {|x|>1}|u0|2=3|W|2\int_{\{|x|>1\}}|\nabla u_{0}|^{2}=\int_{\mathbb{R}^{3}}|\nabla W|^{2}, we see that Theorems 1.2 and 1.3 describe all solutions of (1.11), (1.8) such that (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0). Let us also mention that Theorems 1.2 and 1.3 cannot be generalized to non-symmetric solutions or other domains than the exterior of a ball, see Subsection 7.4.

We finally give a few more references related to this problem. The study of the linear wave equation outside an obstacle was initiated by Morawetz in [Mor61], and considered in the 60’s and 70’s by Lax and Phillips, Morawetz, Ralston and Strauss, and many other contributors: for an extensive discussion, see for example [MRS77] and references therein.

For resolvent estimates in general non-trapping geometries, leading in particular to (1.6), see [Bur02] and references therein. For a general discussion about local energy decay estimates, one can look at the recent paper [BB18].

The focusing nonlinear wave equation with a superquintic nonlinearity outside the unit ball of 3\mathbb{R}^{3} with Dirichlet boundary conditions was considered in [DY19].

The nonlinear Schrödinger equation outside a non-trapping obstacle with Dirichlet boundary conditions was first considered in [BGT04]. The scattering for the three-dimensional defocusing cubic Schrödinger equation outside a star-shaped obstacle was shown by Planchon and Vega [PV09], and by the same authors [PV12] for the analogous equation in two space dimensions. The energy-critical case outside a strictly convex obstacle in dimension three was treated in [KVZ16]. A scattering result for a nonlinear Schrödinger equation in a model case of weakly trapping geometry can be found in [Laf19]. To our knowledge, there is no work on the nonlinear Schrödinger equation outside an obstacle with Neumann boundary conditions.

Notations

We will use the following notations:

  • If uu is a function of time and space, u\vec{u} is understood to be (u,tu)(u,\partial_{t}u).

  • Conversely, if u(Bc)\vec{u}\in\mathcal{H}(B^{c}), uu is understood to be the first component of u\vec{u}.

  • B(0,R)B(0,R) is the ball centered in 0 of radius RR, B=B(0,1)B=B(0,1), Bc:=3\B(0,1)B^{c}:=\mathbb{R}^{3}\backslash B(0,1) is the domain we are interested in.

  • S3S_{\mathbb{R}^{3}} and SNS_{N} are the linear flow of the wave equation respectively in 3\mathbb{R}^{3} and in BcB^{c} with Neumann boundary condition. If (u0,u1)(u_{0},u_{1}) is the initial data we will denote by SN(t)(u0,u1)S_{N}(t)(u_{0},u_{1}) or (SN(u0,u1))(t)\big{(}S_{N}(u_{0},u_{1})\big{)}(t) the solution of (1.10), (1.8), (1.9) at time tt, and (SN(u0,u1))(t,r)\big{(}S_{N}(u_{0},u_{1})\big{)}(t,r) the solution at time tt, location x=|r|x=|r|. We use similar notations for S3S_{\mathbb{R}^{3}}, and the flows 𝒮N\mathscr{S}_{N} and 𝒮3\mathscr{S}_{\mathbb{R}^{3}} defined below. The arrowed versions S3\vec{S}_{\mathbb{R}^{3}} and SN\vec{S}_{N} denote the flows together with their first temporal derivative.

  • 𝒮3\mathscr{S}_{\mathbb{R}^{3}} and 𝒮N\mathscr{S}_{N} are the corresponding nonlinear flows for the defocusing energy critical wave equation (1.1).

  • We will make the following convention: if (u0,u1)(3)(u_{0},u_{1})\in\mathcal{H}(\mathbb{R}^{3}), SN(t)(u0,u1)S_{N}(t)(u_{0},u_{1}) and 𝒮N(t)(u0,u1)\mathscr{S}_{N}(t)(u_{0},u_{1}) will denote the flows applied to the restriction of (u0,u1)(u_{0},u_{1}) to BcB^{c}.

  • Throughout the paper, we often deal with solutions of linear and nonlinear equations both in BcB^{c} with Neumann boundary conditions and in 3\mathbb{R}^{3}. In such situations, the letter uu has been chosen for the Neumann solutions, whereas vv stands for 3\mathbb{R}^{3} solutions.

  • LpLq:=Lp(,Lq(Bc))L^{p}L^{q}:=L^{p}(\mathbb{R},L^{q}(B^{c})).

  • H˙1(Bc)\dot{H}^{1}(B^{c}) is the space of radial functions fL6(Bc)f\in L^{6}(B^{c}) such that |f|L2(Bc)|\nabla f|\in L^{2}(B^{c}).

  • (Bc)\mathcal{H}(B^{c}) is the space of radial functions in H˙1(Bc)×L2(Bc)\dot{H}^{1}(B^{c})\times L^{2}(B^{c}).

  • Finally, (3)\mathcal{H}(\mathbb{R}^{3}) is the space of radial functions in H˙1(3)×L2(3)\dot{H}^{1}(\mathbb{R}^{3})\times L^{2}(\mathbb{R}^{3}).

2. Preliminaries

2.1. The functional setting

Definition 2.1.

We define the extension operator 𝒫\mathcal{P} from H˙1(Bc)\dot{H}^{1}(B^{c}) to H˙rad1(3)\dot{H}_{\text{rad}}^{1}(\mathbb{R}^{3}) by

𝒫u(r):={u(r)r1,u(1)r<1,\mathcal{P}u(r):=\begin{cases}u(r)&r\geq 1,\\ u(1)&r<1,\end{cases}

which is well-defined since by the radial Sobolev embedding, for uH˙1(Bc)u\in\dot{H}^{1}(B^{c}), the function ru(r)r\mapsto u(r) is continuous on [1,)[1,\infty). Similarly, we define the extension operator 𝒫{\vec{\mathcal{P}}} from \mathcal{H} to (3)\mathcal{H}(\mathbb{R}^{3}) by

𝒫(f,g)(r):={(f(r),g(r))r1,(f(1),0)r<1.\mathcal{\vec{\mathcal{P}}}(f,g)(r):=\begin{cases}(f(r),g(r))&r\geq 1,\\ (f(1),0)&r<1.\end{cases}

Let us recall that:

Lemma 2.2.

For uH˙1(Bc)u\in\dot{H}^{1}(B^{c}) we have

(2.1) 1u(r)2𝑑r41u(r)2r2𝑑r,\int_{1}^{\infty}u(r)^{2}dr\leq 4\int_{1}^{\infty}u^{\prime}(r)^{2}r^{2}dr,

in particular, for any compact KBcK\subset B^{c}

(2.2) uL2(K)uH˙1.\|u\|_{L^{2}(K)}\lesssim\|u\|_{\dot{H}^{1}}.

Moreover

(2.3) |u(1)|uH˙1(Bc).|u(1)|\lesssim\|u\|_{\dot{H}^{1}(B^{c})}.
Proof.

Integrating by parts, we get

(2.4) 21ru(r)u(r)𝑑r=1u(r)2𝑑ru(1)2.2\int_{1}^{\infty}ru(r)u^{\prime}(r)dr=-\int_{1}^{\infty}u(r)^{2}dr-u(1)^{2}.

Note that the integration by parts is justified by approximating uu by smooth compactly supported functions. Thus

1u(r)2𝑑r21ru(r)u(r)𝑑r,\int_{1}^{\infty}u(r)^{2}dr\leq-2\int_{1}^{\infty}ru(r)u^{\prime}(r)dr,

and (2.1) follows by the Cauchy-Schwarz inequality. The estimate (2.2) follows immediately. Bounding the left-hand side of (2.4) by the Cauchy-Schwarz inequality, and using (2.1), we obtain (2.3). ∎

Remark 2.3.

With the same proof, one can generalize (2.3) to |u(r)|uH˙1(Bc)|u(r)|\lesssim\|u\|_{\dot{H}^{1}(B^{c})}. This implies readily that a radial solution of the defocusing critical wave equation with Neumann boundary condition (1.7), (1.8) is uniformly bounded, thus global (giving a short proof of the result of [BP09] in the radial case). Similarly, any radial solution of the focusing equation (1.11), (1.8) that is bounded in (Bc)\mathcal{H}(B^{c}), is global.

2.2. Linear estimates

In the present radial case, we can derive an explicit formula for the linear flow:

Proposition 2.4 (The linear group).

For any (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}), we have, for almost every r1r\geq 1 and tt\in\mathbb{R}, and for every r1r\geq 1 and tt\in\mathbb{R} if we have additionally (u0,u1)C1×C0(u_{0},u_{1})\in C^{1}\times C^{0}:

(2.5) (SN(u0,u1))(t,r)=1r(φ+(rt)+φ(r+t))\big{(}S_{N}(u_{0},u_{1})\big{)}(t,r)=\frac{1}{r}\left(\varphi_{+}(r-t)+\varphi_{-}(r+t)\right)

where, denoting (ζ0,ζ1):=(ru0,ru1)(\zeta_{0},\zeta_{1}):=(ru_{0},ru_{1}), for s1s\geq 1,

(2.6) φ+(s)\displaystyle\varphi_{+}(s) =12ζ0(s)121sζ1(σ)𝑑σ,\displaystyle=\frac{1}{2}\zeta_{0}(s)-\frac{1}{2}\int_{1}^{s}\zeta_{1}(\sigma)d\sigma,
(2.7) φ(s)\displaystyle\varphi_{-}(s) =12ζ0(s)+121sζ1(σ)𝑑σ,\displaystyle=\frac{1}{2}\zeta_{0}(s)+\frac{1}{2}\int_{1}^{s}\zeta_{1}(\sigma)d\sigma,

and, for s(,1]s\in(-\infty,1]

(2.8) φ+(s)\displaystyle\varphi_{+}(s) =12ses+σ2(ζ0(σ)+ζ1(σ))𝑑σ12ζ0(2s)1212sζ1(σ)𝑑σ+es1ζ0(1),\displaystyle=\int_{1}^{2-s}e^{s+\sigma-2}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma-\frac{1}{2}\zeta_{0}(2-s)-\frac{1}{2}\int_{1}^{2-s}\zeta_{1}(\sigma)d\sigma+e^{s-1}\zeta_{0}(1),
(2.9) φ(s)\displaystyle\varphi_{-}(s) =12ses+σ2(ζ0(σ)ζ1(σ))𝑑σ12ζ0(2s)+1212sζ1(σ)𝑑σ+es1ζ0(1).\displaystyle=\int_{1}^{2-s}e^{s+\sigma-2}(\zeta^{\prime}_{0}(\sigma)-\zeta_{1}(\sigma))d\sigma-\frac{1}{2}\zeta_{0}(2-s)+\frac{1}{2}\int_{1}^{2-s}\zeta_{1}(\sigma)d\sigma+e^{s-1}\zeta_{0}(1).

Moreover, for fL1(,L2(3,Bc))f\in L^{1}(\mathbb{R},L^{2}(\mathbb{R}^{3},B^{c})) radial, we have, for t0t\geq 0 and rt<1r-t<1

(2.10) 0t(SN(0,f(τ)))(tτ,r)dτ=1r01r+t(12r+tτert+τ+σ2σf(τ,σ)dσ+2r+tτr+tτσf(τ,σ)dσ)dτ+12r1r+ttrt+τr+tτσf(τ,σ)dσdτ.\int_{0}^{t}\Big{(}S_{N}(0,f(\tau))\Big{)}(t-\tau,r)\ d\tau=\frac{1}{r}\int_{0}^{1-r+t}\Big{(}\int_{1}^{2-r+t-\tau}e^{r-t+\tau+\sigma-2}\sigma f(\tau,\sigma)\ d\sigma\\ +\int_{2-r+t-\tau}^{r+t-\tau}\sigma f(\tau,\sigma)\ d\sigma\Big{)}d\tau+\frac{1}{2r}\int_{1-r+t}^{t}\int_{r-t+\tau}^{r+t-\tau}\sigma f(\tau,\sigma)\ d\sigma\ d\tau.
Proof.

Observe that, arguing by density, it suffices to consider smooth (u0,u1)(u_{0},u_{1}), for which r(SN(u0,u1))(1,t)=0\partial_{r}(S_{N}(u_{0},u_{1}))(1,t)=0 for all t0t\neq 0. Let us denote ζ(t,r)=rSN(u0,u1)(t,r).\zeta(t,r)=rS_{N}(u_{0},u_{1})(t,r). Then ζ\zeta is solution of the one dimensional problem

(2.11) t2ζr2ζ=0,\displaystyle\partial_{t}^{2}\zeta-\partial_{r}^{2}\zeta=0,
(2.12) rζζr=1=0t0,\displaystyle{\partial_{r}\zeta-\zeta}_{\restriction r=1}=0\quad\forall t\neq 0,
(2.13) ζt=0=(ru0,ru1).\displaystyle\zeta_{\restriction t=0}=(ru_{0},ru_{1}).

By (2.11), ζ(r)=φ+(rt)+φ(r+t)\zeta(r)=\varphi_{+}(r-t)+\varphi_{-}(r+t). The boundary condition (2.12) gives

(2.14) t,φ+(1t)+φ(1+t)=φ+(1t)+φ(1+t),\forall t,\ \varphi_{+}^{\prime}(1-t)+\varphi_{-}^{\prime}(1+t)=\varphi_{+}(1-t)+\varphi_{-}(1+t),

and the initial condition (2.13) gives (2.6) and (2.7). Then, integrating (2.14) for t0t\geq 0 gives (2.8), and integrating it for t0t\leq 0 gives (2.9). The identity (2.10) is then a straightforward computation. ∎

As a first consequence of Proposition 2.4, we prove that any radial solution of the linear wave equation on BcB^{c} with Neumann boundary conditions is asymptotically close to a solution of the linear wave equation on 3\mathbb{R}^{3}.

Proposition 2.5.

Let (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}). Then

(2.15) limt1|x|SN(t)(u0,u1)L2(Bc)=0\lim_{t\to\infty}\left\|\frac{1}{|x|}S_{N}(t)(u_{0},u_{1})\right\|_{L^{2}(B^{c})}=0

and there exists (v0,v1)(3)({v}_{0},{v}_{1})\in\mathcal{H}(\mathbb{R}^{3}) such that

(2.16) limt+SN(t)(u0,u1)S3(t)(v0,v1)(Bc)=0.\lim_{t\to+\infty}\left\|\vec{S}_{N}(t)(u_{0},u_{1})-\vec{S}_{\mathbb{R}^{3}}(t)({v}_{0},{v}_{1})\right\|_{\mathcal{H}(B^{c})}=0.
Proof.
Step 1.

We first prove (2.15). By a straightforward density argument and the conservation of the energy, we can assume that (u0,u1)(u_{0},u_{1}) is smooth and compactly supported. We let φ+\varphi_{+} and φ\varphi_{-} be as in Proposition 2.4, and

φ¯+(s)=φ+(s)+121+ζ1(σ)𝑑σ,φ¯(s)=φ(s)121+ζ1(σ)𝑑σ.\bar{\varphi}_{+}(s)=\varphi_{+}(s)+\frac{1}{2}\int_{1}^{+\infty}\zeta_{1}(\sigma)\,d\sigma,\quad\bar{\varphi}_{-}(s)=\varphi_{-}(s)-\frac{1}{2}\int_{1}^{+\infty}\zeta_{1}(\sigma)d\sigma.

By (2.5),

(2.17) (SN(u0,u1))(t,r)=1r(φ¯+(rt)+φ¯(r+t))\big{(}S_{N}(u_{0},u_{1})\big{)}(t,r)=\frac{1}{r}\left(\bar{\varphi}_{+}(r-t)+\bar{\varphi}_{-}(r+t)\right)

We claim that there exists C>0C>0 (depending on uu) such that

(2.18) |φ¯+(s)|+|φ¯(s)|Ces11sC.\left|\bar{\varphi}_{+}(s)\right|+\left|\bar{\varphi}_{-}(s)\right|\leq Ce^{s}1\!\!1_{s\leq C}.

Note that (2.17) and (2.18) imply easily (2.15). Using that ζ0\zeta_{0}, ζ0\zeta_{0}^{\prime} and ζ1\zeta_{1} are bounded and compactly supported, the bound of φ¯+\bar{\varphi}_{+} in (2.18) follows from the fact that if s1s\geq 1,

φ¯+(s)=12ζ0(s)+12s+ζ1(σ)𝑑σ\bar{\varphi}_{+}(s)=\frac{1}{2}\zeta_{0}(s)+\frac{1}{2}\int_{s}^{+\infty}\zeta_{1}(\sigma)\,d\sigma

and if s<1s<1,

φ¯+(s)=12ses+σ2(ζ0(σ)+ζ1(σ))𝑑σ12ζ0(2s)+122s+ζ1(σ)𝑑σ+es1ζ0(1).\bar{\varphi}_{+}(s)=\int_{1}^{2-s}e^{s+\sigma-2}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma-\frac{1}{2}\zeta_{0}(2-s)+\frac{1}{2}\int_{2-s}^{+\infty}\zeta_{1}(\sigma)d\sigma+e^{s-1}\zeta_{0}(1).

The proof of the bound of φ¯\bar{\varphi}_{-} in (2.17) is very similar and we omit it.

Step 2.

We next prove that there exists (v0,v1)(3)({v}_{0},{v}_{1})\in\mathcal{H}(\mathbb{R}^{3}) such that (2.16) holds. We recall (see e.g. [DKM19, Theorem 2.1]) that for any GL2()G\in L^{2}(\mathbb{R}), there exists a radial solution v(t)=S3(t)(v0,v1)v(t)=S_{\mathbb{R}^{3}}(t)({v}_{0},{v}_{1}) of the linear wave equation on 3\mathbb{R}^{3} such that

(2.19) limt0+|rtv(t,r)G(rt)|2𝑑r\displaystyle\lim_{t\to\infty}\int_{0}^{+\infty}\left|r\partial_{t}v(t,r)-G(r-t)\right|^{2}dr =0,\displaystyle=0,
(2.20) limt0+|rrv(t,r)+G(rt)|2𝑑r\displaystyle\lim_{t\to\infty}\int_{0}^{+\infty}\left|r\partial_{r}v(t,r)+G(r-t)\right|^{2}dr =0.\displaystyle=0.

Denote by u(t,r)=(SN(u0,u1))(t,r)u(t,r)=\left(S_{N}(u_{0},u_{1})\right)(t,r). Let φ+(s)\varphi_{+}(s) be as in Proposition 2.4. We will prove that φ+L2()\varphi_{+}^{\prime}\in L^{2}(\mathbb{R}) and that

(2.21) limt1+|rtu(t,r)+φ+(rt)|2𝑑r\displaystyle\lim_{t\to\infty}\int_{1}^{+\infty}\left|r\partial_{t}u(t,r)+\varphi_{+}^{\prime}(r-t)\right|^{2}dr =0,\displaystyle=0,
(2.22) limt1+|rru(t,r)φ+(rt)|2𝑑r\displaystyle\lim_{t\to\infty}\int_{1}^{+\infty}\left|r\partial_{r}u(t,r)-\varphi_{+}^{\prime}(r-t)\right|^{2}dr =0.\displaystyle=0.

Letting (v0,v1)(v_{0},v_{1}) be such that (2.19) and (2.20) are satisfied with G=φ+G=-\varphi_{+}^{\prime} we see that (2.21) and (2.22) imply the desired conclusion (2.16).

By the definition of φ+\varphi_{+}, we have

φ+(s)={12ζ0(s)12ζ1(s) if s112ζ0(2s)12ζ1(2s)+es1u0(1) if s1,\varphi_{+}^{\prime}(s)=\begin{cases}\frac{1}{2}\zeta_{0}^{\prime}(s)-\frac{1}{2}\zeta_{1}(s)&\text{ if }s\geq 1\\ -\frac{1}{2}\zeta_{0}^{\prime}(2-s)-\frac{1}{2}\zeta_{1}(2-s)+e^{s-1}u_{0}(1)&\text{ if }s\leq 1,\end{cases}

where (ζ0,ζ1)=(ru0,ru1)(\zeta_{0},\zeta_{1})=(ru_{0},ru_{1}). Since ζ0\zeta_{0}^{\prime} and ζ1\zeta_{1} are in L2([1,+))L^{2}([1,+\infty)), we obtain that φ+L2()\varphi_{+}^{\prime}\in L^{2}(\mathbb{R}). The same proof yields that φL2()\varphi_{-}^{\prime}\in L^{2}(\mathbb{R}). By Proposition 2.4,

tu(t,r)=1r(φ+(rt)+φ(r+t)),\partial_{t}u(t,r)=\frac{1}{r}\left(-\varphi_{+}^{\prime}(r-t)+\varphi_{-}^{\prime}(r+t)\right),

and thus

1+|rtu(t,r)+φ+(rt)|2𝑑r=1+|φ(t+r)|2𝑑rr0.\int_{1}^{+\infty}|r\partial_{t}u(t,r)+\varphi_{+}^{\prime}(r-t)|^{2}dr=\int_{1}^{+\infty}|\varphi_{-}^{\prime}(t+r)|^{2}dr\underset{r\to\infty}{\longrightarrow}0.

Similarly

ru(t,r)=1r(φ+(rt)+φ(r+t))1ru(t,r),\partial_{r}u(t,r)=\frac{1}{r}\left(\varphi_{+}^{\prime}(r-t)+\varphi_{-}^{\prime}(r+t)\right)-\frac{1}{r}u(t,r),

and thus, using (2.15),

1+|rru(t,r)φ+(rt)|2𝑑r21+|φ(t+r)|2𝑑r+21+|u(t,r)|2𝑑rr0.\int_{1}^{+\infty}|r\partial_{r}u(t,r)-\varphi_{+}^{\prime}(r-t)|^{2}dr\leq 2\int_{1}^{+\infty}|\varphi_{-}^{\prime}(t+r)|^{2}dr+2\int_{1}^{+\infty}|u(t,r)|^{2}\,dr\\ \underset{r\to\infty}{\longrightarrow}0.

This concludes the proof.

An other consequence of Proposition 2.4 is the local decay of energy:

Proposition 2.6 (Local energy decay).

Let χCc\chi\in C_{c}^{\infty}. For any (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c})

(χu,χtu)L2(,(Bc))χ(u0,u1)(Bc).\|(\chi u,\chi\partial_{t}u)\|_{L^{2}(\mathbb{R},\mathcal{H}(B^{c}))}\lesssim_{\chi}\|(u_{0},u_{1})\|_{\mathcal{H}(B^{c})}.

where u=SN(u0,u1)u=S_{N}(u_{0},u_{1}).

Proof.

Let ζ(t,r):=ru(t,r)\zeta(t,r):=ru(t,r) and R>0R>0 be arbitrary. Note that

1Rr2((ru)2+u2+(tu)2)𝑑rR1R((rζ)2+ζ2+(tζ)2)𝑑r,\int_{1}^{R}r^{2}\left((\partial_{r}u)^{2}+u^{2}+(\partial_{t}u)^{2}\right)dr\lesssim_{R}\int_{1}^{R}\left((\partial_{r}\zeta)^{2}+\zeta^{2}+(\partial_{t}\zeta)^{2}\right)dr,

thus, to obtain the proposition, it is sufficient to show that

(2.23) 1R((rζ)2+ζ2+(tζ)2)𝑑r𝑑tRu0H˙1(Bc)2+u1L22.\int_{-\infty}^{\infty}\int_{1}^{R}\left((\partial_{r}\zeta)^{2}+\zeta^{2}+(\partial_{t}\zeta)^{2}\right)drdt\lesssim_{R}\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2}+\|u_{1}\|_{L^{2}}^{2}.

To this purpose, observe that, by conservation of energy

(2.24) R+1R11R((rζ)2+ζ2+(tζ)2)𝑑r𝑑tRu0H˙1(Bc)2+u1L22,\int_{-R+1}^{R-1}\int_{1}^{R}\left((\partial_{r}\zeta)^{2}+\zeta^{2}+(\partial_{t}\zeta)^{2}\right)drdt\lesssim_{R}\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2}+\|u_{1}\|_{L^{2}}^{2},

where we used (2.2) to bound the u2u^{2} term. Thus, it suffices to bound the integrals tR1\int_{t\geq R-1} and tR+1\int_{t\leq-R+1}. We will deal with the first one, the proof of the bound for the second one being similar. Thus, let us suppose that tR1t\geq R-1. In particular, tr1t\geq r-1, so ζ\zeta writes, by Proposition 2.4, for such tt’s, for all r[1,R]r\in[1,R]:

ζ(t,r)=12r+tert+σ2(ζ0(σ)+ζ1(σ))𝑑σ+122r+tr+tζ1(σ)𝑑σ+12ζ0(r+t)12ζ0(2r+t)+ert1ζ0(1).\zeta(t,r)=\int_{1}^{2-r+t}e^{r-t+\sigma-2}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma+\frac{1}{2}\int_{2-r+t}^{r+t}\zeta_{1}(\sigma)d\sigma\\ +\frac{1}{2}\zeta_{0}(r+t)-\frac{1}{2}\zeta_{0}(2-r+t)+e^{r-t-1}\zeta_{0}(1).

Thus, we have, for tR1t\geq R-1 and 1rR1\leq r\leq R

(2.25) (rζ(t,r))2+(tζ)2+ζ2R(12r+teσt(ζ0(σ)+ζ1(σ))𝑑σ)2+(2r+tr+t(ζ0(σ)+ζ1(σ))𝑑σ)2+ζ0(2r+t)2+ζ0(r+t)2+ζ1(r+t)2+ζ1(2r+t)2+e2tζ0(1)2.(\partial_{r}\zeta(t,r))^{2}+(\partial_{t}\zeta)^{2}+\zeta^{2}\\ \lesssim_{R}\left(\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma\right)^{2}+\left(\int_{2-r+t}^{r+t}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma\right)^{2}\\ \ +\zeta^{\prime}_{0}(2-r+t)^{2}+\zeta^{\prime}_{0}(r+t)^{2}+\zeta_{1}(r+t)^{2}+\zeta_{1}(2-r+t)^{2}\\ \ +e^{-2t}\zeta_{0}(1)^{2}.

By the Cauchy-Schwarz inequality

(12r+teσt(ζ0(σ)+ζ1(σ))𝑑σ)2(12r+teσt𝑑σ)(12r+teσt(ζ0(σ)+ζ1(σ))2𝑑σ)R12r+teσt(ζ0(σ)2+ζ1(σ)2)𝑑σ,\left(\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma\right)^{2}\\ \leq\left(\int_{1}^{2-r+t}e^{\sigma-t}d\sigma\right)\left(\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))^{2}d\sigma\right)\\ \lesssim_{R}\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})d\sigma,

and therefore,

R11R(12r+teσt(ζ0(σ)+ζ1(σ))𝑑σ)2𝑑σ𝑑r𝑑tRR11R12r+teσt(ζ0(σ)2+ζ1(σ)2)𝑑σ𝑑r𝑑tRR11eσt(ζ0(σ)2+ζ1(σ)2)11σ2+t𝑑σ𝑑t=1(R1eσt11σ2+t𝑑t)(ζ0(σ)2+ζ1(σ)2)𝑑σ1(σ2eσt𝑑t)(ζ0(σ)2+ζ1(σ)2)𝑑σ 1(ζ0(σ)2+ζ1(σ)2)dσu0H˙1(Bc)2+u1L22,\int_{R-1}^{\infty}\int_{1}^{R}\left(\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)+\zeta_{1}(\sigma))d\sigma\right)^{2}d\sigma drdt\\ \lesssim_{R}\int_{R-1}^{\infty}\int_{1}^{R}\int_{1}^{2-r+t}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})\ d\sigma drdt\\ \lesssim_{R}\int_{R-1}^{\infty}\int_{1}^{\infty}e^{\sigma-t}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})1\!\!1_{\sigma\leq 2+t}\ d\sigma dt\\ \ =\int_{1}^{\infty}\Big{(}\int_{R-1}^{\infty}e^{\sigma-t}1\!\!1_{\sigma\leq 2+t}dt\Big{)}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})d\sigma\\ \ \leq\int_{1}^{\infty}\Big{(}\int_{\sigma-2}^{\infty}e^{\sigma-t}dt\Big{)}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})d\sigma\\ \text{\ $\lesssim\int_{1}^{\infty}$}(\zeta^{\prime}_{0}(\sigma)^{2}+\zeta_{1}(\sigma)^{2})d\sigma\lesssim\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2}+\|u_{1}\|_{L^{2}}^{2},

where we used (2.1) and the Cauchy-Schwarz inequality to obtain the last bound. As 2r+tr+t𝑑σR1,\int_{2-r+t}^{r+t}d\sigma\lesssim_{R}1, the term coming from the second term in the first line (2.25) is handled in the same way. Moreover,

R11Rζ0(2r+t)2𝑑r𝑑t=1RR1ζ0(2r+t)2𝑑t𝑑rR1ζ0(s)2𝑑sRu0H˙1(Bc)2,\int_{R-1}^{\infty}\int_{1}^{R}\zeta^{\prime}_{0}(2-r+t)^{2}drdt=\int_{1}^{R}\int_{R-1}^{\infty}\zeta^{\prime}_{0}(2-r+t)^{2}dtdr\leq R\int_{1}^{\infty}\zeta^{\prime}_{0}(s)^{2}ds\lesssim_{R}\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2},

and all the terms of the second line of (2.25) are dealt with similarly. Finally, the remark that, by Lemma (2.3),

ζ0(1)2=u0(1)2u0H˙1(Bc)2,\zeta_{0}(1)^{2}=u_{0}(1)^{2}\lesssim\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2},

permits to handle the term coming from the third line of (2.25). We just showed that

R1+1R(rζ)2+(tζ)2+ζ2drdtRu0H˙1(Bc)2+u1L22.\int_{R-1}^{+\infty}\int_{1}^{R}(\partial_{r}\zeta)^{2}+(\partial_{t}\zeta)^{2}+\zeta^{2}\ drdt\lesssim_{R}\|u_{0}\|_{\dot{H}^{1}(B^{c})}^{2}+\|u_{1}\|_{L^{2}}^{2}.

Dealing with the part R+1\int_{-\infty}^{-R+1} in the same way and using (2.24), the estimate (2.23) on ζ\zeta, and hence the proposition, follow. ∎

The integrability of the local energy allows us to obtain the following crucial global Strichartz estimates for the Neumann flow:

Proposition 2.7 (Strichartz estimates for the Neumann flow).

For any couple (p,q)(p,q) verifying

(2.26) 1p+3q=12,3p+2q1,2<p,q<,\frac{1}{p}+\frac{3}{q}=\frac{1}{2},\hskip 10.00002pt\frac{3}{p}+\frac{2}{q}\leq 1,\hskip 10.00002pt2<p\leq\infty,\hskip 10.00002ptq<\infty,

there exists a constant C>0C>0 such that, for all (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}) and all fL1(,Lrad2(r1)),f\in L^{1}(\mathbb{R},L_{\text{rad}}^{2}(r\geq 1)), if uu verifies

t2uΔNu\displaystyle\partial_{t}^{2}u-\Delta_{N}u =f, in Bc,\displaystyle=f,\text{ in }B^{c},
nu\displaystyle\partial_{n}u =0 on B(0,1),\displaystyle=0\text{ on }\partial B(0,1),
ut=0\displaystyle\vec{u}_{\restriction t=0} =(u0,u1),\displaystyle=(u_{0},u_{1}),

then, for all T>0T>0

uLp([T,T],Lq(r1))C((u0,u1)(Bc)+fL1([T,T],L2(r1))).\|u\|_{L^{p}([-T,T],L^{q}(r\geq 1))}\leq C\Big{(}\|(u_{0},u_{1})\|_{\mathcal{H}(B^{c})}+\|f\|_{L^{1}([-T,T],L_{\text{}}^{2}(r\geq 1))}\Big{)}.
Proof.

The main result of [Bur03] shows that the local energy decay of Proposition 2.6 combined with local in time Strichartz estimates implies global in time ones. Such local estimates where shown by [BSS09] for the above range of couples (p,q)(p,q), hence the proposition follows. ∎

Let us also recall the Strichartz estimates in 3\mathbb{R}^{3}:

Proposition 2.8 (Strichartz estimates in 3\mathbb{R}^{3}, [GV87, GV95], [LS95], [KT98]).

For any couple (p,q)(p,q) verifying

(2.27) 1p+3q=12,1p+1q12,2<p,q<,\frac{1}{p}+\frac{3}{q}=\frac{1}{2},\hskip 10.00002pt\frac{1}{p}+\frac{1}{q}\leq\frac{1}{2},\hskip 10.00002pt2<p\leq\infty,\hskip 10.00002ptq<\infty,

there exists a constant C>0C>0 such that, for all (u0,u1)(3)(u_{0},u_{1})\in\mathcal{H}(\mathbb{R}^{3}) and all fL1(,L2(3)),f\in L^{1}(\mathbb{R},L^{2}(\mathbb{R}^{3})), if vv verifies

t2vΔv\displaystyle\partial_{t}^{2}v-\Delta v =f,\displaystyle=f,
vt=0\displaystyle\vec{v}_{\restriction t=0} =(v0,v1),\displaystyle=(v_{0},v_{1}),

then, for all T>0T>0

vLp([T,T],Lq(3))C((v0,v1)(3)+fL1([T,T],L2(3))).\|v\|_{L^{p}([-T,T],L^{q}(\mathbb{R}^{3}))}\leq C\Big{(}\|(v_{0},v_{1})\|_{\mathcal{H}(\mathbb{R}^{3})}+\|f\|_{L^{1}([-T,T],L_{\text{}}^{2}(\mathbb{R}^{3}))}\Big{)}.
Remark 2.9.

Observe the loss in the range of admissibles couples (2.26) in Proposition 2.7 compared to the free case (2.27). This is because we used the local-in-time Strichartz estimates of [BSS09], valid in a general geometrical setup. It is likely that the above Strichartz estimates, outside a ball, could be extended to the full range of couples (2.27), using for the local-in-time estimates a construction similar to the one done by [SS95] for Dirichlet boundary conditions. However, the range of exponents (2.26) is sufficient for our analysis and we don’t pursue this question here.

As a last consequence of the explicit formula for the linear group given by Proposition 2.4, we have

Lemma 2.10.

Let (u0,u1)(C1×C0)(Bc)(u_{0},u_{1})\in(C^{1}\times C^{0})\cap\mathcal{H}(B^{c}). Then

  1. (1)

    we have

    SN()(u0,u1)C0(×Bc)C1({t±r1}),S_{N}(\cdot)(u_{0},u_{1})\in C^{0}(\mathbb{R}\times B^{c})\cap C^{1}(\{t\pm r\neq 1\}),

    with

    r(SN(u0,u1))(1,t)=0t0,\partial_{r}\big{(}S_{N}(u_{0},u_{1})\big{)}(1,t)=0\hskip 8.5359pt\forall t\neq 0,
  2. (2)

    if in addition fL1(,L2(Bc))f\in L^{1}(\mathbb{R},L^{2}(B^{c})) is radial and continuous and uu is defined by

    u(t):=SN(t)(u0,u1)+0tSN(tτ)(0,f(τ))𝑑τ,u(t):=S_{N}(t)(u_{0},u_{1})+\int_{0}^{t}S_{N}(t-\tau)(0,f(\tau))\ d\tau,

    then uC0(×Bc)C1({t±r1})u\in C^{0}(\mathbb{R}\times B^{c})\cap C^{1}(\{t\pm r\neq 1\}) and

    ru(1,t)=0t0.\partial_{r}u(1,t)=0\hskip 8.5359pt\forall t\neq 0.
Proof.

The explicit formulas (2.5), (2.6), (2.7), (2.8), (2.9) give the first part of the lemma, and (2.10) then gives the second part for t>0t>0. The case t<0t<0 is given by a similar computation. ∎

2.3. Perturbative theory

Definition 2.11.

We say that a solution uu of the nonlinear wave equation (1.7), with Neumann boundary conditions (1.8) scatters in the future when there exists a solution uLu_{L} of the linear wave equation (1.10) with Neumann boundary conditions (1.8) such that

limt+u(t)uL(t)(Bc)=0.\lim_{t\to+\infty}\left\|\vec{u}(t)-\vec{u}_{L}(t)\right\|_{\mathcal{H}(B^{c})}=0.

We define similarly scattering in the past. We say that the solution scatters when it scatters both in the future and in the past.

In a classical way, we have

Proposition 2.12.

Let (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}) and u(t)=𝒮N(t)(u0,u1)u(t)=\mathscr{S}_{N}(t)(u_{0},u_{1}).

(2.28) uL5([0,+),L10)u scatters in the future.u\in L^{5}\left([0,+\infty),L^{10}\right)\implies u\text{ scatters in the future}.

A similar property holds in the past. Moreover, there exists ϵ0>0\epsilon_{0}>0 such that, for any (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}),

(2.29) (u0,u1)(Bc)ϵ0𝒮N()(u0,u1)L5L10,\|(u_{0},u_{1})\|_{\mathcal{H}(B^{c})}\leq\epsilon_{0}\implies\mathscr{S}_{N}(\cdot)(u_{0},u_{1})\in L^{5}L^{10},

and 𝒮N()(u0,u1)\mathscr{S}_{N}(\cdot)(u_{0},u_{1}) scatters. And, for any (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}), there exists a solution U±L5(±,L10)U^{\pm}\in L^{5}(\mathbb{R}_{\pm},L^{10}) of (1.7)-(1.8) such that

(2.30) U±(t)SN(t)(u0,u1)(Bc)0, as t±.\|\vec{U}^{\pm}(t)-\vec{S}_{N}(t)(u_{0},u_{1})\|_{\mathcal{H}(B^{c})}\longrightarrow 0,\text{ as }t\longrightarrow\pm\infty.
Sketch of proof.

Observe that (5,10)(5,10) is Strichartz-admissible in the sense of Proposition 2.7. The properties (2.28) and (2.29) are then classical consequences of the global in time Strichartz estimates. Finally, (2.30) can be proved by a fixed point argument using the Strichartz estimates. ∎

In addition,

Proposition 2.13 (Perturbation).

For any M>0M>0, there exists ϵ(M)>0\epsilon(M)>0 such that, for any 0<ϵϵ(M)0<\epsilon\leq\epsilon(M), and all (u0,u1),(u~0,u~1)(Bc)(u_{0},u_{1}),\,(\tilde{u}_{0},\tilde{u}_{1})\in\mathcal{H}(B^{c}), eL1L2e\in L^{1}L^{2} and uL5L10u\in L^{5}L^{10} verifying

uL5L10M,SN()((u0,u1)(u~0,u~1))L5L10ϵ,eL1L2ϵ,\|u\|_{L^{5}L^{10}}\leq M,\hskip 10.00002pt\|S_{N}(\cdot)\big{(}(u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1})\big{)}\|_{L^{5}L^{10}}\leq\epsilon,\hskip 10.00002pt\|e\|_{L^{1}L^{2}}\leq\epsilon,

if u,u~u,\tilde{u} are solutions of

{t2uΔNu=u5 in B(0,1)c,ut=0=(u0,u1),nu=0 on B(0,1),\left\{\begin{aligned} \partial_{t}^{2}u-\Delta_{N}u&=-u^{5}\text{ in }B(0,1)^{c},\\ \vec{u}_{\restriction t=0}&=(u_{0},u_{1}),\\ \partial_{n}u&=0\text{ on }\partial B(0,1),\end{aligned}\right. {t2u~ΔNu~=u~5+e in B(0,1)c,u~t=0=(u~0,u~1),nu~=0 on B(0,1),\left\{\begin{aligned} \partial_{t}^{2}\tilde{u}-\Delta_{N}\tilde{u}&=-\tilde{u}^{5}+e\text{ in }B(0,1)^{c},\\ \vec{\tilde{u}}_{\restriction t=0}&=(\tilde{u}_{0},\tilde{u}_{1}),\\ \partial_{n}\tilde{u}&=0\text{ on }\partial B(0,1),\end{aligned}\right.

then u~L5L10\tilde{u}\in L^{5}L^{10} and we have

uu~L5L10ϵ.\|u-\tilde{u}\|_{L^{5}L^{10}}\lesssim\epsilon.

In addition, the same statement holds for the corresponding equations in 3\mathbb{R}^{3}.

Proof.

The proof is classical and similar to Proposition 4.7 of [FXC11] , we give it for completeness. Let us denote w=uu~w=u-\tilde{u}. Then ww is solution of

t2wΔNw=u5+u~5e,wt=0=(u0,u1)(u~0,u~1).\displaystyle\partial_{t}^{2}w-\Delta_{N}w=-u^{5}+\tilde{u}^{5}-e,\hskip 10.00002pt\vec{w}_{\restriction t=0}=(u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}).

Let T>0T>0. By the Strichartz inequality for the Neumann flow (Proposition 2.7) applied to ww, we get, with an implicit constant independent of TT

uu~L5(T,T)L10\displaystyle\|u-\tilde{u}\|_{L^{5}(-T,T)L^{10}} u~5u5L1(T,T)L2+eL1L2\displaystyle\lesssim\|\tilde{u}^{5}-u^{5}\|_{L^{1}(-T,T)L^{2}}+\|e\|_{L^{1}L^{2}}
+SN()((u0,u1)(u~0,u~1))L5L10\displaystyle\hskip 10.00002pt+\|S_{N}(\cdot)((u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}))\|_{L^{5}L^{10}}
|uu~|(|u|4+|uu~|4)L1((T,T)L2)+eL1L2\displaystyle\lesssim\||u-\tilde{u}|(|u|^{4}+|u-\tilde{u}|^{4})\|_{L^{1}((-T,T)L^{2})}+\|e\|_{L^{1}L^{2}}
+SN()((u0,u1)(u~0,u~1))L5L10\displaystyle\hskip 10.00002pt+\|S_{N}(\cdot)((u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}))\|_{L^{5}L^{10}}
C(TTuu~L10uL104+uu~L5((T,T),L10)5\displaystyle\leq C\Big{(}\int_{-T}^{T}\|u-\tilde{u}\|_{L^{10}}\|u\|_{L^{10}}^{4}+\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}^{5}
+eL1L2+SN()((u0,u1)(u~0,u~1))L5L10).\displaystyle\hskip 10.00002pt+\|e\|_{L^{1}L^{2}}+\|S_{N}(\cdot)((u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}))\|_{L^{5}L^{10}}\Big{)}.

We apply the Grönwall-type lemma of [FXC11, Lemma 8.1], with

φ=uu~L10,γ=5,f=CuL104,β=1,\displaystyle\varphi=\|u-\tilde{u}\|_{L^{10}},\;\gamma=5,\;f=C\|u\|_{L^{10}}^{4},\;\beta=1,
η=C(uu~L5((T,T),L10)5+eL1L2+SN()((u0,u1)(u~0,u~1))L5L10).\displaystyle\eta=C\big{(}\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}^{5}+\|e\|_{L^{1}L^{2}}+\|S_{N}(\cdot)((u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}))\|_{L^{5}L^{10}}\big{)}.

We obtain, for all T>0T>0

uu~L5((T,T),L10)(eL1L2+SN()((u0,u1)(u~0,u~1))L5L10+uu~L5((T,T),L10)5)×Φ(CM4),\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}\leq\Big{(}\|e\|_{L^{1}L^{2}}+\|S_{N}(\cdot)((u_{0},u_{1})-(\tilde{u}_{0},\tilde{u}_{1}))\|_{L^{5}L^{10}}\\ +\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}^{5}\Big{)}\times\Phi(CM^{4}),

where Φ(s)=2Γ(3+2s)\Phi(s)=2\Gamma(3+2s), Γ\Gamma being the Gamma function. Let CM:=6Φ(CM4)C_{M}:=6\Phi(CM^{4}) and ϵ(M)>0\epsilon(M)>0 be sufficiently small so that, for any ϵϵ(M)\epsilon\leq\epsilon(M)

ϵ5CM5ϵ,i.e. ϵ1/CM5/4.\epsilon^{5}C_{M}^{5}\leq\epsilon,\ \text{i.e. }\epsilon\leq 1/C_{M}^{5/4}.

Then, given T>0T>0 so that uu~L5((T,T),L10)CMϵ\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}\leq C_{M}\epsilon, we have

uu~L5((T,T),L10)Φ(CM4)(2ϵ+CM5ϵ5),\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}\leq\Phi(CM^{4})\left(2\epsilon+C_{M}^{5}\epsilon^{5}\right),

and thus uu~L5((T,T),L10)3Φ(CM4)ϵ12CMϵ\|u-\tilde{u}\|_{L^{5}((-T,T),L^{10})}\leq 3\Phi(CM^{4})\epsilon\leq\frac{1}{2}C_{M}\epsilon. It easily follows that we can make TT goes to infinity, thus uu~L5(,L10)12CMϵ\|u-\tilde{u}\|_{L^{5}(\mathbb{R},L^{10})}\leq\frac{1}{2}C_{M}\epsilon and the lemma follows. The same proof works for the problem in 3\mathbb{R}^{3} using the corresponding Strichartz estimates. ∎

3. Comparison between Neumann and 3\mathbb{R}^{3} evolutions for dilating profiles

Let us introduce the following notation for the scaling associated to the equation

Definition 3.1.

For λ>0\lambda>0, σλ\sigma_{\lambda} denotes the rescaling on H˙1(3)\dot{H}^{1}(\mathbb{R}^{3}), given by

σλ(f)=1λ1/2f(λ)\sigma_{\lambda}(f)=\frac{1}{\lambda^{1/2}}f\left(\frac{\cdot}{\lambda}\right)

and on (3)\mathcal{H}(\mathbb{R}^{3}) given by

σλ(f,g):=(1λ1/2f(λ),1λ3/2g(λ)).\sigma_{\lambda}(f,g):=\left(\frac{1}{\lambda^{1/2}}f\left(\frac{\cdot}{\lambda}\right),\frac{1}{\lambda^{3/2}}g\left(\frac{\cdot}{\lambda}\right)\right).

The aim of this section is to show that a dilating profile (λ\lambda\rightarrow\infty) does not see the obstacle, in the sense that for such profiles, the associated Neumann and 3\mathbb{R}^{3} evolutions are asymptoticaly the same.

Lemma 3.2 (Comparison of linear evolutions for dilating profiles).

Let ψ(3){{\vec{\psi}}}\in\mathcal{H}(\mathbb{R}^{3}), fL1(,L2(3))f\in L^{1}(\mathbb{R},L^{2}(\mathbb{R}^{3})) be radial, (λn)n1(\lambda_{n})_{n\geq 1} a sequence of positive real numbers such that λn+\lambda_{n}\longrightarrow+\infty, (tn)n1(t_{n})_{n\geq 1} a sequence of times, vv be the solution in the sense of Duhamel of

t2vΔv\displaystyle\partial_{t}^{2}v-\Delta v =fin 3,\displaystyle=f\quad\text{in }\mathbb{R}^{3},
vt=0\displaystyle\vec{v}_{\restriction t=0} =ψ,\displaystyle=\vec{\psi},

and vn:=σλnvv_{n}:=\sigma_{\lambda_{n}}v. Finally, let fn:=1λn52f(tnλn,λn)f_{n}:=\frac{1}{\lambda_{n}^{\frac{5}{2}}}f(\frac{\cdot-t_{n}}{\lambda_{n}},\frac{\cdot}{\lambda_{n}}) and unu_{n} be the solution in the sense of Duhamel of

t2unΔun\displaystyle\partial_{t}^{2}u_{n}-\Delta u_{n} =fnin Bc,\displaystyle=f_{n}\quad\text{in }B^{c},
run\displaystyle\partial_{r}u_{n} =0for r=1,\displaystyle=0\quad\text{for }r=1,
unt=tn\displaystyle\vec{u}_{n\restriction t=-t_{n}} =vnt=tn.\displaystyle=\vec{v}_{n\restriction t=-t_{n}}.

Then, as nn goes to infinity

(3.1) suptun(t)vn(t)(Bc)0,\sup_{t\in\mathbb{R}}\|u_{n}{(t)}-v_{n}{(t)}\|_{\mathcal{H}(B^{c})}\longrightarrow 0,

and

(3.2) unvnL5L100.\|u_{n}-v_{n}\|_{L^{5}L^{10}}\longrightarrow 0.
Proof.

Observe that, by interpolation, it suffices to obtain (3.1): indeed, if (3.1) holds, by Sobolev embedding we have unvnLL60\|u_{n}-v_{n}\|_{L^{\infty}L^{6}}\longrightarrow 0, and then (3.2) follows by Hölder inequality, Minkowski inequality, Strichartz estimates for both flows (Propositions 2.7 and and 2.8) and conservation of energy. Moreover, arguing by density, we can assume that ψ\vec{\psi} and ff are smooth and compactly supported. We will argue in three steps:

  1. (1)

    tn=0t_{n}=0 n\forall n and f=0f=0 ,

  2. (2)

    tn=0t_{n}=0 n\forall n and ψ=0\vec{\psi}=\vec{0},

  3. (3)

    general case.

Step 1: tn=0t_{n}=0 and f=0f=0

We have, using the equations satisfied by unu_{n} and vnv_{n}

(3.3) ddt(12Bc|(unvn)|2+12Bc|t(unvn)|2)=Bcr(unvn)t(unvn)=Bcrvnt(unvn)=4πrvn(t,1)t(unvn)(t,1).\frac{d}{dt}\Big{(}\frac{1}{2}\int_{B^{c}}|\nabla(u_{n}-v_{n})|^{2}+\frac{1}{2}\int_{B^{c}}|\partial_{t}(u_{n}-v_{n})|^{2}\Big{)}\\ =-\int_{\partial B^{c}}\partial_{r}(u_{n}-v_{n})\partial_{t}(u_{n}-v_{n})=\int_{\partial B^{c}}\partial_{r}v_{n}\partial_{t}(u_{n}-v_{n})\\ =4\pi\partial_{r}v_{n}(t,1)\partial_{t}(u_{n}-v_{n})(t,1).

We now claim that, for large nn

(3.4) |tvn(t,1)|+|rvn(t,1)|\displaystyle|\partial_{t}v_{n}(t,1)|+|\partial_{r}v_{n}(t,1)| 1λn3211[Cλn,Cλn],\displaystyle\lesssim\frac{1}{\lambda_{n}^{\frac{3}{2}}}1\!\!1_{[-C\lambda_{n},C\lambda_{n}]},
(3.5) |tun(t,1)|\displaystyle|\partial_{t}u_{n}(t,1)| 1λn32+e|t|λn12,\displaystyle\lesssim\frac{1}{\lambda_{n}^{\frac{3}{2}}}+\frac{e^{-|t|}}{\lambda_{n}^{\frac{1}{2}}},

where the constant C>0C>0 and the implicit constants depend on φ\vec{\varphi}. Observe that integrating (3.3), (3.4) and (3.5) give (3.1).

Let us first show (3.4). Observe that

vn(t,x)=1λn12v(tλn,xλn),v_{n}(t,x)=\frac{1}{\lambda_{n}^{\frac{1}{2}}}v(\frac{t}{\lambda_{n}},\frac{x}{\lambda_{n}}),

where v:=S3(t)ψv:=S_{\mathbb{R}^{3}}(t)\vec{\psi}. As ψCc\vec{\psi}\in C^{\infty}_{c}, v\vec{v} is bounded in any Sobolev space Hσ(3)×Hσ1(3)H^{\sigma}(\mathbb{R}^{3})\times H^{\sigma-1}(\mathbb{R}^{3}) for σ1\sigma\geq 1. As a consequence,

(3.6) |tvn(t,1)|+|rvn(t,1)|1λn32.|\partial_{t}v_{n}(t,1)|+|\partial_{r}v_{n}(t,1)|\lesssim\frac{1}{\lambda_{n}^{\frac{3}{2}}}.

Furthermore, by the strong Huygens principle, vv is supported in {|t||x|+C}\big{\{}|t|\leq|x|+C\big{\}}, and thus

(3.7) vn(t,1)=0 for |t|1+Cλn.v_{n}(t,1)=0\text{ for }|t|\geq 1+C\lambda_{n}.

Together with (3.6), (3.7) gives (3.4).

We now show (3.5). By Proposition 2.4, we have for t0t\geq 0

(3.8) tun(t,1)=φ+,n(1t)+φ,n(1+t),\partial_{t}u_{n}(t,1)=-\varphi^{\prime}_{+,n}(1-t)+\varphi^{\prime}_{-,n}(1+t),

where, denoting ψ=(ψ0,ψ1)\vec{\psi}=(\psi_{0},\psi_{1})

(3.9) φ,n(s)=12(1λn32ψ0(sλn)+1λn32ψ1(sλn))\varphi^{\prime}_{-,n}(s)=\frac{1}{2}\Big{(}\frac{1}{\lambda_{n}^{\frac{3}{2}}}\psi_{0}^{\prime}(\frac{s}{\lambda_{n}})+\frac{1}{\lambda_{n}^{\frac{3}{2}}}\psi_{1}^{\prime}(\frac{s}{\lambda_{n}})\Big{)}

and

(3.10) φ+,n(s)=12(1λn32ψ0(2sλn)+1λn32ψ1(2sλn))+1λn3212ses+σ2(ψ0(σλn)+ψ1(σλn))𝑑σ+es11λn12ψ0(1λn).\varphi^{\prime}_{+,n}(s)=-\frac{1}{2}\Big{(}\frac{1}{\lambda_{n}^{\frac{3}{2}}}\psi_{0}^{\prime}(\frac{2-s}{\lambda_{n}})+\frac{1}{\lambda_{n}^{\frac{3}{2}}}\psi_{1}^{\prime}(\frac{2-s}{\lambda_{n}})\Big{)}\\ +\frac{1}{\lambda_{n}^{\frac{3}{2}}}\int_{1}^{2-s}e^{s+\sigma-2}\Big{(}\psi_{0}^{\prime}(\frac{\sigma}{\lambda_{n}})+\psi_{1}^{\prime}(\frac{\sigma}{\lambda_{n}})\Big{)}\,d\sigma+e^{s-1}\frac{1}{\lambda_{n}^{\frac{1}{2}}}\psi_{0}(\frac{1}{\lambda_{n}}).

This last identity (3.10) with (3.8) and (3.9) gives (3.5) for t0t\geq 0. The argument for t0t\leq 0 is similar and Step 1 follows.

Step 2: tn=0t_{n}=0 and ψ=0{\vec{\psi}}={\vec{0}}

As in the first step, we have

(3.11) ddt(12Bc|(unvn)|2+12Bc|t(unvn)|2)=4πrvn(t,1)t(unvn)(t,1).\frac{d}{dt}\Big{(}\frac{1}{2}\int_{B^{c}}|\nabla(u_{n}-v_{n})|^{2}+\frac{1}{2}\int_{B^{c}}|\partial_{t}(u_{n}-v_{n})|^{2}\Big{)}\\ =4\pi\partial_{r}v_{n}(t,1)\partial_{t}(u_{n}-v_{n})(t,1).

Let us show that

(3.12) |tvn(t,1)|+|rvn(t,1)|\displaystyle|\partial_{t}v_{n}(t,1)|+|\partial_{r}v_{n}(t,1)| 1λn3211[Cλn,Cλn],\displaystyle\lesssim\frac{1}{\lambda_{n}^{\frac{3}{2}}}1\!\!1_{[-C\lambda_{n},C\lambda_{n}]},
(3.13) |tun(t,1)|\displaystyle|\partial_{t}u_{n}(t,1)| 1λn72t2,\displaystyle\lesssim\frac{1}{\lambda_{n}^{\frac{7}{2}}}t^{2},

which, together with (3.11), implies (3.1).

We first show (3.12). We have

vn(t,x)=1λn12v(tλn,xλn),v_{n}(t,x)=\frac{1}{\lambda_{n}^{\frac{1}{2}}}v(\frac{t}{\lambda_{n}},\frac{x}{\lambda_{n}}),

where v:=S3(t)ψv:=S_{\mathbb{R}^{3}}(t)\vec{\psi}. As tv\partial_{t}v and rv\partial_{r}v are bounded,

(3.14) |tvn(t,1)|+|rvn(t,1)|1λn32.|\partial_{t}v_{n}(t,1)|+|\partial_{r}v_{n}(t,1)|\lesssim\frac{1}{\lambda_{n}^{\frac{3}{2}}}.

In addition, as we assumed ff to be compactly supported in time and space,

vn(t,1)=0 for |t|1+Cλn,v_{n}(t,1)=0\text{ for }|t|\geq 1+C\lambda_{n},

which, with (3.14), gives (3.12).

In order to prove (3.13), we will need

Claim.

Let fC0(×Bc)f\in C^{0}(\mathbb{R}\times B^{c}) be radial and bounded:

(t,x)×Bc,|f(t,x)|M,\forall(t,x)\in\mathbb{R}\times B^{c},\,|f(t,x)|\leq M,

and ww be the solution of

t2wΔw\displaystyle\partial_{t}^{2}w-\Delta w =fin Bc,\displaystyle=f\quad\text{in }B^{c},
rw\displaystyle\partial_{r}w =0for r=1,\displaystyle=0\quad\text{for }r=1,
wt=0\displaystyle\vec{w}_{\restriction t=0} =0.\displaystyle=\vec{0}.

Then we have

(t,x)×Bc,|w(t,x)|12Mt2.\forall(t,x)\in\mathbb{R}\times B^{c},\,|w(t,x)|\leq\frac{1}{2}Mt^{2}.

To obtain (3.13) from the claim, we apply it to w:=tunw:=\partial_{t}u_{n}, observing that as unu_{n} is a regular solution, tun\partial_{t}u_{n} is in C0(,D(ΔN))C^{0}(\mathbb{R},D(-\Delta_{N})), and thus satisfies Neumann boundary conditions. Let us now prove the claim to achieve the proof of Step 2. Let

z(t,r):=12Mt2w(t,r).z(t,r):=\frac{1}{2}Mt^{2}-w(t,r).

By the formulas of Proposition 2.4, we see that if u1u_{1} is positive for t0t\geq 0, then so is SN(0,u1)(t,r)S_{N}(0,u_{1})(t,r). Thus, by the Duhamel formula, as (t2Δ)z0(\partial_{t}^{2}-\Delta)z\geq 0, we have z0z\geq 0 for t0t\geq 0, from which we obtain w12Mt2w\leq\frac{1}{2}Mt^{2} for t0t\geq 0. Considering z~(,r):=12Mt2+w(t,r)\tilde{z}(,r):=\frac{1}{2}Mt^{2}+w(t,r), we obtain as well w12Mt2-w\leq\frac{1}{2}Mt^{2} for t0t\geq 0. The negative times are obtained in a similar fashion.

Step 3: general case

By the two first steps, we obtain the case tn=0t_{n}=0. Now, let wnw_{n} be solution of the Neumann problem with initial condition at t=0t=0

t2wnΔwn\displaystyle\partial_{t}^{2}w_{n}-\Delta w_{n} =fnin Bc,\displaystyle=f_{n}\quad\text{in }B^{c},
rwn\displaystyle\partial_{r}w_{n} =0for r=1,\displaystyle=0\quad\text{for }r=1,
wnt=0\displaystyle\vec{w}_{n\restriction t={0}} =vnt=0.\displaystyle=\vec{v}_{n\restriction t={0}}.

By the case tn=0t_{n}=0, we have, as nn\longrightarrow\infty

(3.15) suptwnvn(Bc)0,\sup_{t\in\mathbb{R}}\|w_{n}-v_{n}\|_{\mathcal{H}(B^{c})}\longrightarrow 0,

and in particular, as by definition un(tn)=vn(tn)u_{n}(-t_{n})=v_{n}(-t_{n})

(3.16) wn(tn)un(tn)(Bc)0.\|w_{n}(-t_{n})-u_{n}(-t_{n})\|_{\mathcal{H}(B^{c})}\longrightarrow 0.

From (3.16), as wnunw_{n}-u_{n} is solution of the homogeneous linear wave equation with Neumann boundary conditions in BcB^{c}, it follows from conservation of energy that

(3.17) suptwnun(Bc)=wn(tn)un(tn)(Bc)0.\sup_{t\in\mathbb{R}}\|w_{n}-u_{n}\|_{\mathcal{H}(B^{c})}=\|w_{n}(-t_{n})-u_{n}(-t_{n})\|_{\mathcal{H}(B^{c})}\longrightarrow 0.

The result (3.1) follows from (3.15) and (3.17). ∎

The following lemma will play a key role in the comparison between the 3\mathbb{R}^{3} and Neumann dynamics in the nonlinear profile decomposition introduced in section 5 (see in particular (5.9)).

Lemma 3.3 (Comparison of nonlinear evolutions for dilating profiles).

\̇newline Let VL5(,L10(3))V\in L^{5}(\mathbb{R},L^{10}(\mathbb{R}^{3})) be a solution of the critical defocusing nonlinear wave equation in 3\mathbb{R}^{3}, (i.e. (1.1) with Ω=3\Omega=\mathbb{R}^{3} and ι=1\iota=1), (λn)n(\lambda_{n})_{n} a sequence of positive real numbers such that λn+\lambda_{n}\longrightarrow+\infty, and (tn)n(t_{n})_{n}\in\mathbb{R}^{\mathbb{N}}. We denote

Vn(t,x):=1λn1/2V(ttnλn,xλn)=𝒮3(t)σλn(V(tnλn))V_{n}(t,x):=\frac{1}{\lambda_{n}^{1/2}}V\left(\frac{t-t_{n}}{\lambda_{n}},\frac{x}{\lambda_{n}}\right)=\mathscr{S}_{\mathbb{R}^{3}}(t)\sigma_{\lambda_{n}}\left(\vec{V}\left(\frac{-t_{n}}{\lambda_{n}}\right)\right)

and let UnU_{n} be the solution of the nonlinear Neumann problem

{t2UnΔUn+Un5=0in Bc,rUn=0for r=1,Unt=0=Vnt=0.\left\{\begin{aligned} \partial_{t}^{2}U_{n}-\Delta U_{n}+U_{n}^{5}&=0\quad\text{in }B^{c},\\ \partial_{r}U_{n}&=0\quad\text{for }r=1,\\ \vec{U}_{n\restriction t=0}&=\vec{V}_{n\restriction t=0}.\end{aligned}\right.

Then

lim supnUnL5L10<,\limsup_{n\in\mathbb{N}}\|U_{n}\|_{L^{5}L^{10}}<\infty,

and, as nn\longrightarrow\infty

suptUn(t)Vn(t)(Bc)+UnVnL5L100.\sup_{t\in\mathbb{R}}\big{\|}\vec{U}_{n}(t)-\vec{V}_{n}(t)\big{\|}_{\mathcal{H}(B^{c})}+\|U_{n}-V_{n}\|_{L^{5}L^{10}}\longrightarrow 0.
Remark 3.4.

The conclusion of the proposition implies

limn𝒮3(t)σλn(V(tnλn))𝒮N(t)σλn(V(tnλn))(Bc)=0.\lim_{n\to\infty}\left\|\vec{\mathscr{S}}_{\mathbb{R}^{3}}(t)\sigma_{\lambda_{n}}\left(\vec{V}\left(\frac{-t_{n}}{\lambda_{n}}\right)\right)-\vec{\mathscr{S}}_{N}(t)\sigma_{\lambda_{n}}\left(\vec{V}\left(\frac{-t_{n}}{\lambda_{n}}\right)\right)\right\|_{\mathcal{H}(B^{c})}=0.
Proof.

Observe that, by energy estimates, it suffices to show UnVnL5L100\|U_{n}-V_{n}\|_{L^{5}L^{10}}\longrightarrow 0. Let ZnZ_{n} be the solution of the nonlinear Neumann problem

{t2ZnΔZn+Vn5=0in Bc,rZn=0for r=1,Znt=0=Unt=0.\left\{\begin{aligned} \partial_{t}^{2}Z_{n}-\Delta Z_{n}+V_{n}^{5}&=0\quad\text{in }B^{c},\\ \partial_{r}Z_{n}&=0\quad\text{for }r=1,\\ \vec{Z}_{n\restriction t=0}&=\vec{U}_{n\restriction t=0}.\end{aligned}\right.

By Lemma 3.2 applied to Zn(+tn)Z_{n}(\cdot+t_{n}) and Vn(+tn)V_{n}(\cdot+t_{n}), we get

(3.18) ZnVnL5L100.\|Z_{n}-V_{n}\|_{L^{5}L^{10}}\longrightarrow 0.

Let T>0T>0 and observe that

{t2(ZnUn)+Δ(ZnUn)=Un5Vn5in Bc,r(ZnUn)=0for r=1,ZnUnt=0=0,\left\{\begin{aligned} \partial_{t}^{2}(Z_{n}-U_{n})+\Delta(Z_{n}-U_{n})&=U_{n}^{5}-V_{n}^{5}\quad\text{in }B^{c},\\ \partial_{r}(Z_{n}-U_{n})&=0\quad\text{for }r=1,\\ \vec{Z}_{n}-\vec{U}_{n\restriction t=0}&=\vec{0},\end{aligned}\right.

and therefore, we have, by the global Strichartz estimates for the Neumann flow (Proposition 2.7), together with Hölder and Minkowski inequalities, with an implicit constant which is independent of T>0T>0

(3.19) ZnUnL5(T,T)L10Un5Vn5L1(T,T)L2TT[Vn(t)L104Un(t)Vn(t)L10+Un(t)Vn(t)L105]𝑑tTT[Vn(t)L104Zn(t)Un(t)L10+Zn(t)Un(t)L105]𝑑t+ϵn(T),\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}\lesssim\|U_{n}^{5}-V_{n}^{5}\|_{L^{1}(-T,T)L^{2}}\\ \lesssim\int_{-T}^{T}\Big{[}\|V_{n}(t)\|^{4}_{L^{10}}\|U_{n}(t)-V_{n}(t)\|_{L^{10}}+\|U_{n}(t)-V_{n}(t)\|^{5}_{L^{10}}\Big{]}\;dt\\ \lesssim\int_{-T}^{T}\Big{[}\|V_{n}(t)\|_{L^{10}}^{4}\|Z_{n}(t)-U_{n}(t)\|_{L^{10}}+\|Z_{n}(t)-U_{n}(t)\|^{5}_{L^{10}}\Big{]}\,dt\;+\epsilon_{n}(T),

where we decomposed Un(t)Vn(t)=Un(t)Zn(t)+Zn(t)Vn(t)U_{n}(t)-V_{n}(t)=U_{n}(t)-Z_{n}(t)+Z_{n}(t)-V_{n}(t) in the last line, and

ϵn(T)=TTVn(t)Zn(t)L105+Vn(t)L104Vn(t)Zn(t)L10dt.\epsilon_{n}(T)=\int_{-T}^{T}\|V_{n}(t)-Z_{n}(t)\|_{L^{10}}^{5}+\|V_{n}(t)\|_{L^{10}}^{4}\|V_{n}(t)-Z_{n}(t)\|_{L^{10}}\;dt.

By Hölder inequality and (3.18)

(3.20) ϵn:=supT>0ϵn(T)VnZnL5(,L10)5+VL5(,L10(3))4VnZnL5(,L10)0.\epsilon^{\prime}_{n}:=\sup_{T>0}\epsilon_{n}(T)\leq\|{V}_{n}-{Z}_{n}\|_{L^{5}(\mathbb{R},L^{10})}^{5}+\|V\|_{L^{5}(\mathbb{R},L^{10}(\mathbb{R}^{3}))}^{4}\|{V}_{n}-{Z}_{n}\|_{L^{5}(\mathbb{R},L^{10})}\longrightarrow 0.

By (3.19), we have, with an implicit constant independent of TT

(3.21) ZnUnL5(T,T)L10TTVn(t)L104Zn(t)Un(t)L10𝑑t+ϵn+ZnUnL5(T,T)L105.\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}\lesssim\int_{-T}^{T}\|V_{n}(t)\|_{L^{10}}^{4}\|Z_{n}(t)-U_{n}(t)\|_{L^{10}}\;dt\\ +\epsilon^{\prime}_{n}+\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}^{5}.

Now, VnL104L54()\|V_{n}\|_{L^{10}}^{4}\in L^{\frac{5}{4}}(\mathbb{R}) and VnL104L54()=VL5L104.\left\|\|V_{n}\|_{L^{10}}^{4}\right\|_{L^{\frac{5}{4}}(\mathbb{R})}=\|V\|_{L^{5}L^{10}}^{4}. Thus we get, by (3.21), using the Gronwall-type lemma of [FXC11, Lemma 8.1], for all T>0T>0, with C>0C>0 independent of T>0T>0:

(3.22) ZnUnL5(T,T)L10C(ϵn+ZnUnL5(T,T)L105).\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}\leq C\big{(}\epsilon^{\prime}_{n}+\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}^{5}\big{)}.

Let ϵ>0\epsilon>0 be small enough so that 2Cϵ512ϵ,2C\epsilon^{5}\leq\frac{1}{2}\epsilon, and nn large enough so that ϵnϵ5.\epsilon^{\prime}_{n}\leq\epsilon^{5}. From (3.22), it follows that if TT is such that ZnUnL5(T,T)L10ϵ\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}\leq\epsilon, we have

ZnUnL5(T,T)L1012ϵ.\|Z_{n}-U_{n}\|_{L^{5}(-T,T)L^{10}}\leq\frac{1}{2}\epsilon.

We can therefore send TT to infinity to obtain:

ZnUnL5(,L10)0,\|Z_{n}-U_{n}\|_{L^{5}(\mathbb{R},L^{10})}\longrightarrow 0,

and the lemma follows using (3.18)(\ref{eq:dilnl_ZU}). ∎

4. Linear profile decomposition

We recall that by convention, if (u0,u1)(3)(u_{0},u_{1})\in\mathcal{H}(\mathbb{R}^{3}), SN(t)(u0,u1)S_{N}(t)(u_{0},u_{1}) (respectively 𝒮N(t)(u0,u1)\mathscr{S}_{N}(t)(u_{0},u_{1})) denotes the flow of the linear (respectively nonlinear) wave equation with Neumann boundary condition applied to the restriction of (u0,u1)(u_{0},u_{1}) to BcB^{c}. The aim of this section is to show

Proposition 4.1 (Linear profile decomposition).

Let (ϕn)n1(\vec{\phi}_{n})_{n\geq 1} be a bounded sequence in (Bc)\mathcal{H}(B^{c}). Then, up to a subsequence, there exists sequences of real parameters (tj,n)j,n1(t_{j,n})_{j,n\geq 1}, (λj,n)j,n1(\lambda_{j,n})_{j,n\geq 1} and a sequence (ψj)j1(\vec{\psi}^{j})_{j\geq 1} of elements of (3)\mathcal{H}(\mathbb{R}^{3}) such that

(4.1) jklimn+|tj,ntk,n|λj,n+|logλj,nλk,n|=+,j\neq k\implies\lim_{n\rightarrow+\infty}\frac{|t_{j,n}-t_{k,n}|}{\lambda_{j,n}}+\big{|}\log\frac{\lambda_{j,n}}{\lambda_{k,n}}\big{|}=+\infty,

there exists a partition (Jcomp,Jdiff)(J_{\text{comp}},J_{\text{diff}}) of \mathbb{N} such that

(4.2) jJcompn,λj,n=1,j\in J_{\text{comp}}\implies\forall n,\ \lambda_{j,n}=1,
(4.3) jJdiffλj,nn+,j\in J_{\text{diff}}\implies\lambda_{j,n}\underset{n\rightarrow\infty}{\longrightarrow}+\infty,

moreover

(4.4) j,tj,n/λj,n± or n,tj,n=0,\forall j,\ t_{j,n}{/\lambda_{j,n}}\longrightarrow\pm\infty\text{ or }\forall n,\;t_{j,n}=0,

and, for all J1J\geq 1,

(4.5) ϕn=j=1JSN(tj,n)σλj,nψj+wnJ,\vec{\phi}_{n}=\sum_{j=1}^{J}\vec{S}_{N}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}+\vec{w}_{n}^{J},

where the remainder enjoys the decay

(4.6) limJ+lim supn+SN()wnJL5L10=0.\hskip 10.00002pt\lim_{J\rightarrow+\infty}\limsup_{n\rightarrow+\infty}\|S_{N}(\cdot)\vec{w}_{n}^{J}\|_{L^{5}L^{10}}=0.

In addition, this decomposition verifies the Pythagorean expansion,

(4.7) J,ϕn(Bc)2=jJcomp1jJψj(Bc)2+jJdiff1jJψj(3)2+wnJ(Bc)2+on(1),\forall J,\hskip 10.00002pt\|\vec{\phi}_{n}\|_{\mathcal{H}(B^{c})}^{2}=\sum_{\begin{subarray}{c}j\in J_{\text{comp}}\\ 1\leq j\leq J\end{subarray}}\|\vec{\psi}^{j}\|_{\mathcal{H}(B^{c})}^{2}\ +\sum_{\begin{subarray}{c}j\in J_{\text{diff}}\\ 1\leq j\leq J\end{subarray}}\|\vec{\psi}^{j}\|_{\mathcal{H}(\mathbb{R}^{3})}^{2}\ +\|\vec{w}_{n}^{J}\|_{\mathcal{H}(B^{c})}^{2}\ +o_{n}(1),

as well as the L6L^{6} version of it:

(4.8) J,ϕnL66=j=1JSN(tj,n)σλj,nψjL66+wnJL66+on(1).\forall J,\hskip 10.00002pt\|\phi_{n}\|_{L^{6}}^{6}=\sum_{j=1}^{J}\|S_{N}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}\|_{L^{6}}^{6}\ +\|w_{n}^{J}\|_{L^{6}}^{6}\ +o_{n}(1).

Recall from (2.1) the definition of the extension operator 𝒫\mathcal{P}. Proposition 4.1 will be a consequence of:

Lemma 4.2.

Let (fn)n1(f_{n})_{n\geq 1} be a bounded sequence in H˙1(Bc)\dot{H}^{1}(B^{c}) such that for all sequence of real numbers (λn)n1(\lambda_{n})_{n\geq 1} verifying

limnλn=+orn,λn=1,\lim_{n}\lambda_{n}=+\infty\quad\text{or}\quad\forall n,\;\lambda_{n}=1,

we have, as nn goes to infinity

λn12𝒫(fn)(λn)0 in H˙1(3).\lambda_{n}^{\frac{1}{2}}\mathcal{P}(f_{n})(\lambda_{n}\cdot)\rightharpoonup 0\text{ in }\dot{H}^{1}(\mathbb{R}^{3}).

Then, up to a subsequence, as nn goes to infinity

fnL6(Bc)0.\|f_{n}\|_{L^{6}(B^{c})}\longrightarrow 0.
Proof.

As (𝒫(fn))n1(\mathcal{P}(f_{n}))_{n\geq 1} is a bounded sequence in H˙rad1(3)\dot{H}_{\text{rad}}^{1}(\mathbb{R}^{3}), we may apply the elliptic profile decomposition of [Gér98], and up to a subsequence

𝒫(fn)=j=1J1λj,n1/2φj(λj,n)+wnJ,\mathcal{P}(f_{n})=\sum_{j=1}^{J}\frac{1}{\lambda_{j,n}^{1/2}}\varphi_{j}\left(\frac{\cdot}{\lambda_{j,n}}\right)+w_{n}^{J},

with

limJ+lim supn+wnJL6=0.\lim_{J\rightarrow+\infty}\limsup_{n\rightarrow+\infty}\|w_{n}^{J}\|_{L^{6}}=0.

Remark that

φj=weaklimnλj,n1/2𝒫(fn)(λj,n) in H˙1(3).\varphi_{j}=\operatorname*{weak\,lim}_{n\rightarrow\infty}\lambda_{j,n}^{1/2}\mathcal{P}(f_{n})(\lambda_{j,n}\cdot)\text{ in }\dot{H}^{1}(\mathbb{R}^{3}).

Thus, for all jj such that lim infnλj,n>0\liminf_{n}\lambda_{j,n}>0, we have φj=0\varphi_{j}=0 by hypothesis. Indeed in this case, extracting subsequences, we can assume that λj,n\lambda_{j,n} has a limit λ(0,){+}\lambda_{\infty}\in(0,\infty)\cup\{+\infty\}. If this limit is finite, we may furthermore assume, rescaling φj\varphi_{j} if necessary, that λj,n=1\lambda_{j,n}=1 for all nn.

On the other hand, if jj is such that λj,nn0\lambda_{j,n}\underset{n\rightarrow\infty}{\longrightarrow}0, observe that

λj,n1/2𝒫(fn)(λj,n)=λj,n1/2fn(1) on {r1λj,n}.\lambda_{j,n}^{1/2}\mathcal{P}(f_{n})(\lambda_{j,n}\cdot)=\lambda_{j,n}^{1/2}f_{n}(1)\text{ on }\Big{\{}r\leq\frac{1}{\lambda_{j,n}}\Big{\}}.

By Lemma 2.2,

|fn(1)|fnH˙1(Bc)|f_{n}(1)|\lesssim\|f_{n}\|_{\dot{H}^{1}(B^{c})}

which is bounded independently of nn, and we deduce that λj,n1/2𝒫(fn)(λn)\lambda_{j,n}^{1/2}\mathcal{P}(f_{n})(\lambda_{n}\cdot) goes to zero as nn goes to infinity, uniformly on every compact of 3\mathbb{R}^{3}, and thus in the sense of distributions as well. By the uniqueness of the limit, we conclude that φj=0\varphi_{j}=0. Therefore φj=0\varphi_{j}=0 for all jj and the lemma follows. ∎

Before showing Proposition 4.1, let us observe that

Lemma 4.3.

Let (Rn)n\big{(}\vec{R}_{n}\big{)}_{n} be a sequence in (3)\mathcal{H}(\mathbb{R}^{3}). For j=1,2j=1,2, let (λj,n)n(+),(tj,n)n(\lambda_{j,n})_{n}\in(\mathbb{R}_{+}^{*})^{\mathbb{N}},\ (t_{j,n})_{n}\in\mathbb{R}^{\mathbb{N}} be such that

(4.9) limnλj,n=+orn,λj,n=1.\lim_{n\to\infty}\lambda_{j,n}=+\infty\quad\text{or}\quad\forall n,\;\lambda_{j,n}=1.

Then

  1. (1)

    If

    M,n,|t1,nt2,n|+|logλ1,nλ2,n|M,\exists M,\forall n,\ |t_{1,n}-t_{2,n}|+\big{|}\log\frac{\lambda_{1,n}}{\lambda_{2,n}}\big{|}\leq M,

    then, up to a subsequence, weakly in (3)\mathcal{H}(\mathbb{R}^{3})

    Rn0σλ2,n1𝒫(SN(t1,nt2,n)σλ1,nRn)0.\vec{R}_{n}\rightharpoonup 0\implies\sigma_{\lambda_{2,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n}-t_{2,n})\sigma_{\lambda_{1,n}}\vec{R}_{n})\rightharpoonup 0.
  2. (2)

    If

    |t1,nt2,n|λ1,n+|logλ1,nλ2,n|+,\frac{|t_{1,n}-t_{2,n}|}{\lambda_{1,n}}+\big{|}\log\frac{\lambda_{1,n}}{\lambda_{2,n}}\big{|}\longrightarrow+\infty,

    then, for all ψ(3)\vec{\psi}\in\mathcal{H}(\mathbb{R}^{3}), up to a subsequence, weakly in (3)\mathcal{H}(\mathbb{R}^{3}):

    σλ2,n1𝒫(SN(t1,nt2,n)σλ1,nψ)0.\sigma_{\lambda_{2,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n}-t_{2,n})\sigma_{\lambda_{1,n}}\vec{\psi})\rightharpoonup 0.
Proof.

Let us show the first point. Up to the extraction of a subsequence, we have

t1,nt2,nτ,t_{1,n}-t_{2,n}\longrightarrow\tau\in\mathbb{R},

and additionally

either (λ1,n,λ2,n)(+,+), or n,(λ1,n,λ2,n)=(1,1).\text{either }(\lambda_{1,n},\lambda_{2,n})\longrightarrow(+\infty,+\infty),\text{ or }\forall n,\;(\lambda_{1,n},\lambda_{2,n})=(1,1).

In the first situation, Lemma 3.2 allows us to replace SNS_{N} by S3S_{\mathbb{R}^{3}}, for which the result is known. In the second situation, we have, for any test function ξ(Bc)\vec{\xi}\in\mathcal{H}(B^{c}),

σλ1,n1𝒫(SN(t2,nt1,n)σλ2,nξ)𝒫(SN(τ)ξ)\sigma_{\lambda_{1,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{2,n}-t_{1,n})\sigma_{\lambda_{2,n}}\vec{\xi})\longrightarrow\vec{\mathcal{P}}(\vec{S}_{N}(-\tau)\vec{\xi})

strongly in (Bc)\mathcal{H}(B^{c}), and the first point follows.

Let us now deal with the second point. We are in one of the three following situations:

  1. (i)

    λ1,n\lambda_{1,n}\longrightarrow\infty,

  2. (ii)

    λ2,n\lambda_{2,n}\longrightarrow\infty, n,λ1,n=1\forall n,\;\lambda_{1,n}=1 and M>0,|t1,nt2,n|M\exists M>0,\;|t_{1,n}-t_{2,n}|\leq M,

  3. (iii)

    n,λ1,n=1\forall n,\;\lambda_{1,n}=1 and |t1,nt2,n||t_{1,n}-t_{2,n}|\longrightarrow\infty.

In the situation (i), we can use again Lemma 3.2 to replace SNS_{N} by S3S_{\mathbb{R}^{3}}, and the result follows.

In the situation (ii), up to a subsequence, 𝒫(SN(t1,nt2,n)σλ1,nψ)\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n}-t_{2,n})\sigma_{\lambda_{1,n}}\vec{\psi}) is converging strongly in (3)\mathcal{H}(\mathbb{R}^{3})

𝒫(SN(t1,nt2,n)σλ1,nψ)ξ.\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n}-t_{2,n})\sigma_{\lambda_{1,n}}\vec{\psi})\longrightarrow\vec{\xi}.

By a density argument, we can assume that ξ\vec{\xi} is smooth and compactly supported. Then, by definition of the scaling σ\sigma

r0,σλ2,n1ξ(r)0\forall r\neq 0,\hskip 5.69046pt\sigma^{-1}_{\lambda_{2,n}}\vec{\xi}(r)\longrightarrow 0

and the result follows.

In the situation (iii), we use this time Proposition 2.5 to compare the solution to a solution in 3\mathbb{R}^{3}, for which the result is known.∎

We are now in position to prove the main result of this section.

Proof of Proposition 4.1.

We will first construct the profiles and the parameters by induction, so that the expansion (4.5) holds together with the orthogonality of the parameters (4.1), (4.2), (4.3), and the Pythagorean expansion (4.7), (4.8). Then, we will show the decay of the remainder (4.6).

For α=(αn)n\vec{\alpha}=(\vec{\alpha}_{n})_{n} a bounded sequence in (Bc)\mathcal{H}(B^{c}), let us denote by Λ(α)\Lambda(\vec{\alpha}) the set of all ψ(3)\vec{\psi}\in\mathcal{H}(\mathbb{R}^{3}) such that there exist an extraction {nk}k\{n_{k}\}_{k} and sequences (λnk)k(0,)(\lambda_{n_{k}})_{k}\in(0,\infty)^{\mathbb{N}} and (tnk)k(t_{n_{k}})_{k}\in\mathbb{R}^{\mathbb{N}}, with

limkλnk=ork,λnk=1,\displaystyle\lim_{k\to\infty}\lambda_{n_{k}}=\infty\quad\text{or}\quad\forall k,\;\lambda_{n_{k}}=1,
ψ=weaklimk(σλnk1𝒫(SN(tnk)αnk)) in (3).\displaystyle\vec{\psi}=\operatorname*{weak\,lim}_{k\to\infty}\Big{(}\sigma_{\lambda_{n_{k}}}^{-1}\vec{\mathcal{P}}\big{(}\vec{S}_{N}(t_{n_{k}})\vec{\alpha}_{n_{k}}\big{)}\Big{)}\text{ in }\mathcal{H}(\mathbb{R}^{3}).

We denote

(4.10) η(α):=supψΛ(α)ψ(3),\eta(\vec{\alpha}):=\sup_{\vec{\psi}\in\Lambda(\vec{\alpha})}\|\vec{\psi}\|_{\mathcal{H}(\mathbb{R}^{3})},

and observe that, by definition of 𝒫\vec{\mathcal{P}} and ψ\vec{\psi}

(4.11) ψ(3)=ψ(Bc) if ψ is associated with λnk=1.\|\vec{\psi}\|_{\mathcal{H}(\mathbb{R}^{3})}=\|\vec{\psi}\|_{\mathcal{H}(B^{c})}\;\text{ if }\vec{\psi}\text{ is associated with }\lambda_{n_{k}}=1.

Extraction of the first profile. If η((ϕn)n1)=0\eta((\vec{\phi}_{n})_{n\geq 1})=0, then the decomposition holds. Otherwise, there exists ψ1(3)\vec{\psi}^{1}\in\mathcal{H}(\mathbb{R}^{3}) and (λ1,n)n1(+),(t1,n)n1(\lambda_{1,n})_{n\geq 1}\in(\mathbb{R}_{+}^{*})^{\mathbb{N}},\ (t_{1,n})_{n\geq 1}\in\mathbb{R}^{\mathbb{N}} with λ1,n+\lambda_{1,n}\rightarrow+\infty or n,λ1,n=1\forall n,\;\lambda_{1,n}=1, such that, up to an extraction

(4.12) ψ1=weaklimnσλ1,n1𝒫(SN(t1,n)ϕn) in (3),\vec{\psi}^{1}=\operatorname*{weak\,lim}_{{n\to\infty}}\sigma_{\lambda_{1,n}}^{-1}\vec{\mathcal{P}}\big{(}\vec{S}_{N}(t_{1,n})\vec{\phi}_{n}\big{)}\text{ in }\mathcal{H}(\mathbb{R}^{3}),

and

12η((un)n1)ψ1(3).\frac{1}{2}\eta((\vec{u}_{n})_{n\geq 1})\leq\|\vec{\psi}^{1}\|_{\mathcal{H}(\mathbb{R}^{3})}.

Let us denote

(4.13) wn1:=ϕnSN(t1,n)σλ1,nψ1.\vec{w}_{n}^{1}:=\vec{\phi}_{n}-\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}.

Observe that, if t1,n/λ1,nt_{1,n}/\lambda_{1,n}, has a finite limit τ¯1\bar{\tau}_{1} , we can harmlessly assume that t1,n=0t_{1,n}=0 for all nn. Indeed, if λ1,n=1\lambda_{1,n}=1 for all nn, we see by (4.12) that

𝒫(SN(τ¯1)(ψ1))=weaklimn𝒫(ϕn).\vec{\mathcal{P}}\left(\vec{S}_{N}(-\bar{\tau}_{1})(\vec{\psi}^{1})\right)=\operatorname*{weak\,lim}_{n\to\infty}\vec{\mathcal{P}}\big{(}\vec{\phi}_{n}\big{)}.

If λ1,n+\lambda_{1,n}\to+\infty, we have, by (4.12) and Lemma 3.2,

ψ1=weaklimnσλ1,n1(S3(t1,n)ϕn)=weaklimn(S3(t1,n/λ1,n)σλ1,n1ϕn)=weaklimn(S3(τ¯1)σλ1,n1ϕn).\vec{\psi}^{1}=\operatorname*{weak\,lim}_{n\to\infty}\sigma_{\lambda_{1,n}}^{-1}\big{(}\vec{S}_{\mathbb{R}^{3}}(t_{1,n})\vec{\phi}_{n}\big{)}=\operatorname*{weak\,lim}_{n\to\infty}\left(\vec{S}_{\mathbb{R}^{3}}(t_{1,n}/\lambda_{1,n})\sigma_{\lambda_{1,n}}^{-1}\vec{\phi}_{n}\right)\\ =\operatorname*{weak\,lim}_{n\to\infty}\left(\vec{S}_{\mathbb{R}^{3}}(\bar{\tau}_{1})\sigma_{\lambda_{1,n}}^{-1}\vec{\phi}_{n}\right).

In both cases, we see that we can assume t1,n=0t_{1,n}=0 by modifying the limiting profile ψ1\vec{\psi}^{1}.

Now, we have, by the definition of wn1\vec{w}_{n}^{1} (4.13) and the weak convergence (4.12)

(4.14) SN(t1,n)σλ1,nψ1,wn1(Bc)=σλ1,nψ1,SN(t1,n)ϕnσλ1,nψ1(Bc)=σλ1,nψ1,𝒫(SN(t1,n)ϕn)σλ1,nψ1(3)=ψ1,σλ1,n1𝒫(SN(t1,n)ϕn)ψ1(3)0 as n goes to infinity,\Big{\langle}\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1},\vec{w}_{n}^{1}\Big{\rangle}_{\mathcal{H}(B^{c})}=\Big{\langle}\sigma_{\lambda_{1,n}}\vec{\psi}^{1},\vec{S}_{N}(t_{1,n})\vec{\phi}_{n}-\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\Big{\rangle}_{\mathcal{H}(B^{c})}\\ =\Big{\langle}\sigma_{\lambda_{1,n}}\vec{\psi}^{1},\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n})\vec{\phi}_{n})-\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\Big{\rangle}_{\mathcal{H}(\mathbb{R}^{3})}=\Big{\langle}\vec{\psi}^{1},\sigma_{\lambda_{1,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{1,n})\vec{\phi}_{n})-\vec{\psi}^{1}\Big{\rangle}_{\mathcal{H}(\mathbb{R}^{3})}\\ \longrightarrow 0\text{ as }n\text{ goes to infinity,}

and therefore,

(4.15) ϕn(Bc)2=SN(t1,n)σλ1,nψ1(Bc)2+wn1(Bc)2+on(1).\|\vec{\phi}_{n}\|_{\mathcal{H}(B^{c})}^{2}=\|\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}(B^{c})}^{2}+\|w_{n}^{1}\|_{\mathcal{H}(B^{c})}^{2}+o_{n}(1).

But, by conservation of energy

(4.16) SN(t1,n)σλ1,nψ1(Bc)2\displaystyle\|\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}(B^{c})}^{2} =σλ1,nψ1(Bc)2.\displaystyle=\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}(B^{c})}^{2}.

Now, remark that, if λ1,n\lambda_{1,n}\longrightarrow\infty, then, as nn goes to infinity, we have

σλ1,nψ1(B(0,1))20\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}(B(0,1))}^{2}\longrightarrow 0

and thus, as σλ1,n\sigma_{\lambda_{1,n}} is an isometry on (3)\mathcal{H}(\mathbb{R}^{3}),

σλ1,nψ1(Bc)2\displaystyle\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}{(B^{c})}}^{2} =σλ1,nψ1(3)2+on(1)\displaystyle=\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\|_{\mathcal{H}(\mathbb{R}^{3})}^{2}+o_{n}(1)
(4.17) =ψ1(3)2+on(1)if λ1,n,\displaystyle=\|\vec{\psi}^{1}\|_{\mathcal{H}(\mathbb{R}^{3})}^{2}+o_{n}(1)\hskip 10.00002pt\text{\text{if }$\lambda_{1,n}\longrightarrow\infty$},

and thus, combining (4.17) with (4.15) and (4.16), the decomposition (4.5) with Pythagorean expansion (4.7) holds at rank J=1J=1.

Let us now show the L6L^{6} Pythagorean expansion (4.8).

First case: t1,n=0t_{1,n}=0. Let

fn:=||ϕn|6|σλ1,nψ1|6|wn1|6|,f_{n}:=\bigg{|}\int|{\phi_{{}n}}|^{6}-|\sigma_{\lambda_{1,n}}{{\psi}^{1}}|^{6}-|{w_{n}^{1}}|^{6}\bigg{|},

and observe that, as for any z,wz,w\in\mathbb{R}

||z+w|6|z|6|w|6||z||w|(|z|4+|w|4),\big{|}|z+w|^{6}-|z|^{6}-|w|^{6}\big{|}\lesssim|z||w|\big{(}|z|^{4}+|w|^{4}\big{)},

we have, by (4.13)

fn|σλ1,nψ1||wn1|gn,gn:=|σλ1,nψ1|4+|wn1|4.f_{n}\lesssim\int\left|\sigma_{\lambda_{1,n}}{{\psi}^{1}}\right|\,|w_{n}^{1}|\;g_{n},\hskip 5.69046ptg_{n}:=\left|\sigma_{\lambda_{1,n}}{{\psi}^{1}}\right|^{4}+|w_{n}^{1}|^{4}.

On the other hand, by Sobolev embedding, conservation of energy and scale invariance

σλ1,nψ1L6σλ1,nψ1(Bc)σλ1,nψ1(3)=ψ1(3),\left\|\sigma_{\lambda_{1,n}}{\psi}^{1}\right\|_{L^{6}}\lesssim\left\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\right\|_{\mathcal{H}(B^{c})}\\ \leq\left\|\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\right\|_{\mathcal{H}(\mathbb{R}^{3})}=\left\|{\vec{\psi}}^{1}\right\|_{\mathcal{H}(\mathbb{R}^{3})},

Together with (4.13) and Sobolev embedding, it follows that supngnL3/2<\sup_{n}\|g_{n}\|_{L^{3/2}}<\infty, and we get, by Hölder inequality

(4.18) fn(Bc|σλ1,nψ1|3|wn1|3)13(3|σλ1,nψ1|3|w~n1|3)13=(3|ψ1|3|σλ1,n1w~n1|3)13,f_{n}\lesssim\Big{(}\int_{B^{c}}|\sigma_{\lambda_{1,n}}{{\psi}^{1}}|^{3}|w_{n}^{1}|^{3}\Big{)}^{\frac{1}{3}}\\ {\leq\Big{(}\int_{\mathbb{R}^{3}}|\sigma_{\lambda_{1,n}}{{\psi}^{1}}|^{3}|\tilde{w}_{n}^{1}|^{3}\Big{)}^{\frac{1}{3}}=\Big{(}\int_{\mathbb{R}^{3}}|{{\psi}^{1}}|^{3}|\sigma_{\lambda_{1,n}^{-1}}\tilde{w}_{n}^{1}|^{3}\Big{)}^{\frac{1}{3}},}

where w~n1:=𝒫ϕnσλ1,nψ1\vec{\tilde{w}}_{n}^{1}:=\vec{\mathcal{P}}\vec{\phi}_{n}-\sigma_{\lambda_{1,n}}\vec{\psi}^{1} extends the definition of wn1\vec{w}_{n}^{1} to 3\mathbb{R}^{3} in the present case t1,n=0t_{1,n}=0. Now, observe that by (4.12) and (4.13), σλ1,n1w~n10\sigma_{\lambda_{1,n}^{-1}}\vec{\tilde{w}}_{n}^{1}\rightharpoonup 0 weakly in (3)\mathcal{H}(\mathbb{R}^{3}). By Rellich theorem, for any compact K3K\subset\mathbb{R}^{3}, σλ1,n1w~n1\sigma_{\lambda_{1,n}^{-1}}\tilde{w}_{n}^{1} strongly converges to 0 in L4(K)L^{4}(K). It follows that |σλ1,n1w~n1|3|\sigma_{\lambda_{1,n}^{-1}}\tilde{w}_{n}^{1}|^{3} converges strongly to 0 in L4/3(K)L^{4/3}(K). By Sobolev embedding, |σλ1,n1w~n1|3|\sigma_{\lambda_{1,n}^{-1}}\tilde{w}_{n}^{1}|^{3} is bounded in L2(3)L^{2}(\mathbb{R}^{3}), thus has a weakly convergent subsequence in L2(3)L^{2}(\mathbb{R}^{3}). By uniqueness of the limit in the sense of distributions, this weak limit is zero and (4.8) follows from (4.18).

Second case: t1,n/λ1,n±t_{1,n}/\lambda_{1,n}\longrightarrow\pm\infty. In this case, we have

SN(t1,n/λ1,n)ψ1L6n0,\|{S_{N}}(-t_{1,n}/\lambda_{1,n})\vec{\psi}^{1}\|_{L^{6}}\underset{n\to\infty}{\longrightarrow}0,

which can be proved easily from the corresponding property for the free flow S3S_{\mathbb{R}^{3}}, and Proposition 2.5. The L6L^{6} Pythagorean expansion (4.8) follows immediately.

Extraction of the subsequent profiles. Let us show how to extract the second profile, the extraction of the JJ’th from the J1J-1’th being the same for arbitrary J2J\geq 2. If η(wn1)=0\eta(\vec{w}_{n}^{1})=0, then we are done, otherwise, there exists ψ2(3)\vec{\psi}^{2}\in\mathcal{H}(\mathbb{R}^{3}) and (λ2,n)n1(+),(t2,n)n1(\lambda_{2,n})_{n\geq 1}\in(\mathbb{R}_{+}^{*})^{\mathbb{N}},\ (t_{2,n})_{n\geq 1}\in\mathbb{R}^{\mathbb{N}} with λ2,n+\lambda_{2,n}\rightarrow+\infty or λ2,n=1\lambda_{2,n}=1, such that

(4.19) ψ2=weaklimσλ2,n1𝒫(SN(t2,n)wn1) in (3),\vec{\psi}^{2}=\operatorname*{weak\,lim}\sigma_{\lambda_{2,n}}^{-1}\vec{\mathcal{P}}\big{(}\vec{S}_{N}(t_{2,n})\vec{w}_{n}^{1}\big{)}\text{ in }\mathcal{H}(\mathbb{R}^{3}),

and

12η((wn1)n1)ψ2(3).\frac{1}{2}\eta((\vec{w}_{n}^{1})_{n\geq 1})\leq\|\vec{\psi}^{2}\|_{\mathcal{H}(\mathbb{R}^{3})}.

We take

(4.20) wn2\displaystyle\vec{w}_{n}^{2} :=wn1SN(t2,n)σλ2,nψ2\displaystyle:=\vec{w}_{n}^{1}-\vec{S}_{N}(-t_{2,n})\sigma_{\lambda_{2,n}}\vec{\psi}^{2}
=unSN(t2,n)σλ2,nψ2SN(t1,n)σλ1,nψ1.\displaystyle\ =\vec{u}_{n}-\vec{S}_{N}(-t_{2,n})\sigma_{\lambda_{2,n}}\vec{\psi}^{2}-\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}.

Let us first show the orthogonality condition (4.1). Denoting

rn1:=σλ1,n1𝒫(SN(t1,n)wn1)=σλ1,n1𝒫(SN(t1,n)un)σλ1,n1𝒫σλ1,nψ1,\vec{r}_{n}^{1}:=\sigma_{\lambda_{1,n}}^{-1}\vec{\mathcal{P}}\left(\vec{S}_{N}(t_{1,n})\vec{w}_{n}^{1}\right){=\sigma_{\lambda_{1,n}}^{-1}\vec{\mathcal{P}}\left(\vec{S}_{N}(t_{1,n})\vec{u}_{n}\right)-\sigma_{\lambda_{1,n}^{-1}}\vec{\mathcal{P}}\sigma_{\lambda_{1,n}}\vec{\psi}^{1}},

we have, by (4.12) and (4.13)

rn10 weakly in (3),\vec{r}_{n}^{1}\rightharpoonup 0\text{ weakly in }\mathcal{H}(\mathbb{R}^{3}),

and in addition, by (4.19)

σλ2,n1𝒫(SN(t2,nt1,n)σλ1,nrn1)ψ20,\sigma_{\lambda_{2,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{2,n}-t_{1,n})\sigma_{\lambda_{1,n}}\vec{r}_{n}^{1})\rightharpoonup\vec{\psi}^{2}\neq 0,

therefore, by Lemma 4.3, the orthogonality condition (4.1) for (j,k)=(1,2)(j,k)=(1,2) follows.

To show the Pythagorean expansion (4.7), using the arguments of the case J=1J=1, it suffices to show that the newly arising mixed term goes to zero, namely that

SN(t2,n)σλ2,nψ2,SN(t1,n)σλ1,nψ1(Bc)0n.\langle\vec{S}_{N}(-t_{2,n})\sigma_{\lambda_{2,n}}\vec{\psi}^{2},\vec{S}_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\rangle_{\mathcal{H}(B^{c})}\underset{{n\to\infty}}{\longrightarrow 0}.

Noting that the left-hand side of the previous line equals

ψ2,σλ2,n1𝒫(SN(t2,nt1,n)σλ1,nψ1)(Bc),\langle\vec{\psi}^{2},\sigma_{\lambda_{2,n}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{2,n}-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1})\rangle_{\mathcal{H}(B^{c})},

the result follows by the orthogonality condition together with Lemma 4.3.

Finally, (4.19) and (4.20) imply by the exact same arguments as in the extraction of the first profile that

wn1L66=SN(t2,n)σλ2,nψ2L66+wn2L66+on(1),\|w_{n}^{1}\|_{L^{6}}^{6}=\|{S_{N}}(-t_{2,n})\sigma_{\lambda_{2,n}}\vec{\psi}^{2}\|_{L^{6}}^{6}+\|w_{n}^{2}\|_{L^{6}}^{6}+o_{n}(1),

from which the L6L^{6} Pythagorean expansion (4.8) follows using the decomposition proved at the previous rank, which readed

ϕnL66=SN(t1,n)σλ1,nψ1L66+wn1L66+on(1).\|\phi_{n}\|_{L^{6}}^{6}=\Big{\|}{S_{N}}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}\Big{\|}_{L^{6}}^{6}+\left\|w_{n}^{1}\right\|_{L^{6}}^{6}+o_{n}(1).

Labeling. We define JdiffJ_{\text{diff}} and JcompJ_{\text{comp}} as follows: if λj,n=1\lambda_{j,n}=1 for all nn, then jJcompj\in J_{\text{comp}}, otherwise, jJdiffj\in J_{\text{diff}}.

Decay of the remainder. In order to obtain (4.6), it suffices to show that

(4.21) limJ+lim supn+SN()wnJLL6=0.\lim_{J\rightarrow+\infty}\limsup_{n\rightarrow+\infty}\|S_{N}(\cdot)\vec{w}_{n}^{J}\|_{L^{\infty}L^{6}}=0.

Indeed, if (4.21) holds, Strichartz estimates of Proposition 2.7 together with Hölder inequality, conservation of energy, and the fact that, by the Pythagorean expansion (4.7),

J,lim supn+wnJ(Bc)lim supn+ϕn(Bc),\forall J,\ \limsup_{n\rightarrow+\infty}\|\vec{w}_{n}^{J}\|_{\mathcal{H}(B^{c})}\leq\limsup_{n\rightarrow+\infty}\|\vec{\phi}_{n}\|_{\mathcal{H}(B^{c})},

yields (4.6).

Let us show (4.21). To this purpose, observe that, by the Pythagorean expansion (4.7),

J,j=1,jJcompJψj(Bc)2+j=1,jJdiffJψj(3)2lim supn1ϕn(Bc)2,\forall J,\hskip 10.00002pt\sum_{j=1,\ j\in J_{\text{comp}}}^{J}\|\vec{\psi}^{j}\|_{\mathcal{H}(B^{c})}^{2}\ +\sum_{j=1,\ j\in J_{\text{diff}}}^{J}\|\vec{\psi}^{j}\|_{\mathcal{H}(\mathbb{R}^{3})}^{2}\leq\limsup_{n\geq 1}\|\vec{\phi}_{n}\|_{\mathcal{H}(B^{c})}^{2},

and thus both series in jj are convergent. Because, by (4.11), the profiles are constructed in such a way that

η((wnj)n1)2{ψj(3)if jJdiff,ψj(Bc)if jJcomp,j0\eta((\vec{w}_{n}^{j})_{n\geq 1})\leq 2\begin{cases}\|\vec{\psi}^{j}\|_{\mathcal{H}(\mathbb{R}^{3})}&\text{if }j\in J_{\text{diff}},\\ \|\vec{\psi}^{j}\|_{\mathcal{H}(B^{c})}&\text{if }j\in J_{\text{comp}},\ j\neq 0\end{cases}

it follows that

(4.22) η((wnJ)n1)J0.\eta\Big{(}(\vec{w}_{n}^{J})_{n\geq 1}\Big{)}\underset{J\rightarrow\infty}{\longrightarrow}0.

Arguing by contradiction, the LL6L^{\infty}L^{6} decay of SN()wnJS_{N}(\cdot)\vec{w}_{n}^{J} follows by Lemma 4.2: indeed, if the decay of the remainder (4.21) does not hold, by a diagonal argument, there exists ϵ0>0\epsilon_{0}>0 and sequences Jk+J_{k}\longrightarrow+\infty, nk+n_{k}\longrightarrow+\infty, and tkt_{k} such that

η((wnkJk)k)=0andk,SN(tk)wnkJkL6(Bc)ϵ0.{\eta\Big{(}(\vec{w}_{n_{k}}^{J_{k}})_{k}\Big{)}=0\quad\text{and}}\quad\forall k,\ \Big{\|}S_{N}(t_{k})\vec{w}_{n_{k}}^{J_{k}}\Big{\|}_{L^{6}(B^{c})}\geq\epsilon_{0}.

Using Lemma 4.2, it follows that there exists ψ(3)\vec{\psi}\in\mathcal{H}(\mathbb{R}^{3}), ψ0\vec{\psi}\neq 0, and a sequence (λk)k(\lambda_{k})_{k} with

limkλk= or k,λk=1\lim_{k}\lambda_{k}=\infty\text{ or }\forall k,\;\lambda_{k}=1

such that, after extraction

σλk1𝒫(SN(tk)wnkJk)ψ\sigma_{\lambda_{k}}^{-1}\vec{\mathcal{P}}(\vec{S}_{N}(t_{k})\vec{w}_{n_{k}}^{J_{k}})\rightharpoonup\vec{\psi}

weakly in (3)\mathcal{H}(\mathbb{R}^{3}). This contradicts the definition (4.10) of η\eta and ends the proof of the proposition. ∎

5. Construction of a compact flow solution

Let us define the critical energy EcE_{c} by

(5.1) Ec:=sup{E>0,u(Bc),(u)E𝒮N()uL5L10},E_{c}:=\sup\big{\{}E>0,\hskip 10.00002pt\forall\vec{u}\in\mathcal{H}(B^{c}),\ \mathscr{E}(\vec{u})\leq E\implies\mathscr{S}_{N}(\cdot)\vec{u}\in L^{5}L^{10}\big{\}},

where, for u(Bc)\vec{u}\in\mathcal{H}(B^{c}), \mathscr{E} is as before the conserved energy

(u):=12u(Bc)2+16uL66.\mathscr{E}(\vec{u}):=\frac{1}{2}\|\vec{u}\|_{\mathcal{H}(B^{c})}^{2}+\frac{1}{6}\|u\|_{L^{6}}^{6}.

Observe that Ec>0E_{c}>0 by Proposition 2.12. The aim of this section is to show

Theorem 5.1.

If Ec<+E_{c}<+\infty, then there exists uc(Bc)\vec{u}_{c}\in\mathcal{H}(B^{c}), uc0\vec{u}_{c}\neq\vec{0}, such that the nonlinear flow {𝒮N(t)uc,t}\big{\{}\vec{\mathscr{S}}_{N}(t)\vec{u}_{c},\ t\in\mathbb{R}\big{\}} has a compact closure in (Bc)\mathcal{H}(B^{c}).

Proof.

If Ec<+E_{c}<+\infty, let u0n\vec{u}_{0}^{n} be a minimising sequence for EcE_{c}, in the sense that

(5.2) (u0n)Ec,limn(u0n)=Ec,𝒮N()u0nL5L10.\mathscr{E}(\vec{u}_{0}^{n})\geq E_{c},\ \lim_{n\rightarrow\infty}\mathscr{E}(\vec{u}_{0}^{n})=E_{c},\ \mathscr{S}_{N}(\cdot)\vec{u}_{0}^{n}\notin L^{5}L^{10}.

Translating un=𝒮N()u0nu_{n}=\mathscr{S}_{N}(\cdot)\vec{u}_{0}^{n} in time if necessary, we may assume

(5.3) limnunL5((0,+),L10)=limnunL5((,0),L10)=+,\lim_{n\to\infty}\|u_{n}\|_{L^{5}\left((0,+\infty),L^{10}\right)}=\lim_{n\to\infty}\|u_{n}\|_{L^{5}\left((-\infty,0),L^{10}\right)}=+\infty,

where by convention unL5((,0),L10)=+\|u_{n}\|_{L^{5}\left((-\infty,0),L^{10}\right)}=+\infty if unL5((,0),L10)u_{n}\notin L^{5}\left((-\infty,0),L^{10}\right), and similary for L5((0,),L10)L^{5}\left((0,\infty),L^{10}\right). As u0n\vec{u}_{0}^{n} is bounded in (Bc)\mathcal{H}(B^{c}), we can, up to a subsequence, decompose it into profiles according to Proposition 4.1:

(5.4) u0n=j=1JSN(tj,n)σλj,nψj+wnJ.\vec{u}_{0}^{n}=\sum_{j=1}^{J}\vec{S}_{N}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}+\vec{w}_{n}^{J}.

To each profile (ψj,(λj,n)n1,(tj,n)n1)(\vec{\psi}^{j},(\lambda_{j,n})_{n\geq 1},(t_{j,n})_{n\geq 1}) we associate a family of nonlinear Neumann profiles (Unj)n1({U_{n}^{j}})_{n\geq 1}, and additionally, for jJdiffj\in J_{\text{diff}}, a free nonlinear profile VjV^{j} and its rescaled family (Vnj)n1({V_{n}^{j}})_{n\geq 1}, in the following way:

  • If jJcomp j\in J_{\text{comp }} i.e. λj,n=1\lambda_{j,n}=1, let Uj{U^{j}} be the only solution of the critical nonlinear wave equation with Neumann boundary conditions (1.7)–(1.8), given by Proposition 2.12 such that

    (5.5) limnUj(tj,n)SN(tj,n)ψj(Bc)=0,\lim_{n\rightarrow\infty}\left\|{\vec{U}^{j}}({-t_{j,n}})-\vec{S}_{N}(-t_{j,n})\vec{\psi}^{j}\right\|_{\mathcal{H}(B^{c})}=0,

    and we set

    (5.6) Unj(t):=Uj(ttj,n).U_{n}^{j}(t):=U^{j}({t-t_{j,n}}).

    Notice that, if tj,n±-t_{j,n}\to\pm\infty, UjL5(±,L10(Bc))U^{j}\in L^{5}(\mathbb{R}_{\pm},L^{10}(B^{c})) by construction.

  • If jJdiff,j\in J_{\text{diff}}, i.e. λj,n\lambda_{j,n}\rightarrow\infty, by Lemma 3.2,

    limnSN(tj,n)σλj,nψjS3(tj,n)σλj,nψj(Bc)=0.\lim_{n\to\infty}\left\|\vec{S}_{N}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}-\vec{S}_{\mathbb{R}^{3}}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}\right\|_{\mathcal{H}{(B^{c})}}=0.

    Furthermore, denoting by VLj(t):=S3(t)ψj(t)V_{L}^{j}(t):=S_{\mathbb{R}^{3}}(t){\psi}^{j}(t), we have

    S3(ttj,n)σλj,nψj=1λj,n1/2VLj(ttj,nλj,n,xλj,n).S_{\mathbb{R}^{3}}(t-t_{j,n})\sigma_{\lambda_{j,n}}{\psi}^{j}=\frac{1}{\lambda_{j,n}^{1/2}}V^{j}_{L}\left(\frac{t-t_{j,n}}{\lambda_{j,n}},\frac{x}{\lambda_{j,n}}\right).

    We define the free nonlinear profile VjV^{j} as the unique solution of the critical nonlinear wave equation on 3\mathbb{R}^{3} such that if tj,n=0t_{j,n}=0 for all nn, Vj(0)=ψj\vec{V}^{j}(0)={\psi}^{j} and if limntj,n/λj,n=±\lim_{n\to\infty}-t_{j,n}/\lambda_{j,n}=\pm\infty, limt±Vj(t)VLj(t)(Bc)=0\lim_{t\to\pm\infty}\left\|\vec{V}^{j}(t)-\vec{V}_{L}^{j}(t)\right\|_{\mathcal{H}{(B^{c})}}=0. In other terms:

    (5.7) limn±Vj(tj,n/λj,n)VLj(tj,n/λj,n)(Bc)=0.\lim_{n\to\pm\infty}\left\|\vec{V}^{j}(-t_{j,n}/\lambda_{j,n})-\vec{V}_{L}^{j}(-t_{j,n}/\lambda_{j,n})\right\|_{\mathcal{H}{(B^{c})}}=0.

    Furthermore, we set

    Vnj(t):=1λj,n1/2Vj(ttj,nλj,n),{V}_{n}^{j}(t):=\frac{1}{\lambda_{j,n}^{1/2}}V^{j}\Big{(}\frac{t-t_{j,n}}{\lambda_{j,n}}\Big{)},

    and we then define the associated family of nonlinear Neumann profiles as

    (5.8) Unj(t):=𝒮N(t)(Vnj(0))=𝒮N(t)(σλj,n(Vj(tj,nλj,n))).U_{n}^{j}(t):=\mathscr{S}_{N}(t)\left(\vec{V}_{n}^{j}(0)\right)=\mathscr{S}_{N}(t)\left(\sigma_{\lambda_{j,n}}\Big{(}\vec{V}^{j}\Big{(}\frac{-t_{j,n}}{\lambda_{j,n}}\Big{)}\Big{)}\right).

    Observe that, as a solution of a defocusing nonlinear wave equation in 3\mathbb{R}^{3}, for which the scattering is well known, we have VjL5L10(3)V^{j}\in L^{5}L^{10}(\mathbb{R}^{3}). Furthermore, as Unj(0)=Vnj(0)\vec{U}_{n}^{j}(0)=\vec{V}_{n}^{j}(0), Lemma 3.3 (used with tn=tj,nt_{n}=t_{j,n}) yields

    jJdiff,supnUnjL5(,L10(Bc))<,\forall j\in J_{\text{diff}},\hskip 10.00002pt\sup_{n}\|U_{n}^{j}\|_{L^{5}(\mathbb{R},L^{10}(B^{c}))}<\infty,

    and

    (5.9) jJdiff,suptVnj(t)Unj(t)(Bc)+VnjUnjLt5Lx10n0.\forall j\in J_{\text{diff}},\hskip 10.00002pt\sup_{t}\left\|\vec{V}_{n}^{j}(t)-\vec{U}_{n}^{j}(t)\right\|_{\mathcal{H}{(B^{c})}}+\|V_{n}^{j}-U_{n}^{j}\|_{L_{t}^{5}L_{x}^{10}}\underset{n\to\infty}{\longrightarrow}0.

Let us assume from now on, by contradiction, that the decomposition (5.4) has strictly more than one non trivial profile, i.e

(5.10) J>1.J>1.

Then, by the Pythagorean expansion (4.7) together with its L6L^{6} version (4.8)

jJcomp,lim supn(SN(tj,n)ψj)<Ec.\forall j\in J_{\text{comp}},\hskip 10.00002pt\limsup_{n\rightarrow\infty}\;\mathscr{E}\left(S_{N}(-t_{j,n})\vec{\psi}^{j}\right)<E_{c}.

Hence, by (5.5), (Uj)<EC\mathscr{E}(U^{j})<E_{C}, and UjL5L10(Bc){U^{j}}\in L^{5}L^{10}(B^{c}) by the definition of the critical energy. Summing up, we have

(5.11) jJcomp,UjL5(,L10(Bc));jJdiff,VjL5(,L10(Bc)).\forall j\in J_{\text{comp}},\hskip 3.00003ptU^{j}\in{L^{5}(\mathbb{R},L^{10}(B^{c}))};\;\forall j\in J_{\text{diff}},\hskip 3.00003ptV^{j}\in{L^{5}(\mathbb{R},L^{10}(B^{c}))}.

Let un:=𝒮Nu0nu_{n}:=\mathscr{S}_{N}\vec{u}_{0}^{n}. We will show the following nonlinear profile decomposition:

Proposition 5.2.

We have

(5.12) J,un(t)\displaystyle\forall J,\ u_{n}(t) =1jJUnj(t)+RnJ(t)\displaystyle=\sum_{1\leq j\leq J}U_{n}^{j}(t)+R_{n}^{J}(t)
=jJcomp1jJUnj(t)+jJdiff1jJVnj(t)+R~nJ(t),\displaystyle=\sum_{\begin{subarray}{c}j\in J_{\text{comp}}\\ 1\leq j\leq J\end{subarray}}U_{n}^{j}(t)+\sum_{\begin{subarray}{c}j\in J_{\text{diff}}\\ 1\leq j\leq J\end{subarray}}V_{n}^{j}(t)+\tilde{R}_{n}^{J}(t),

where

limJlim supnRnJL5L10=limJlim supnR~nJL5L10=0.\lim_{J\rightarrow\infty}\limsup_{n\rightarrow\infty}\|R_{n}^{J}\|_{L^{5}L^{10}}=\lim_{J\rightarrow\infty}\limsup_{n\rightarrow\infty}\|\tilde{R}_{n}^{J}\|_{L^{5}L^{10}}=0.

To this purpose, let

(5.13) u~nJ:=j=1JUnj+znJ,\tilde{u}_{n}^{J}:=\sum_{j=1}^{J}U_{n}^{j}+z_{n}^{J},

where

(5.14) znJ(t):=SN(t)wnJ,z_{n}^{J}(t):=S_{N}(t)\vec{w}_{n}^{J},

verifies, by the decay of the remainder of the linear profile decomposition

(5.15) limJlim supnznJL5L10=0.\lim_{J\to\infty}\limsup_{n\to\infty}\|z_{n}^{J}\|_{L^{5}L^{10}}=0.

Observe that u~nJ\tilde{u}_{n}^{J} is solution in BcB^{c} of the following nonlinear wave equation with Neumann boundary conditions:

(5.16) (t2ΔN)u~nJ+(u~nJ)5=enJ,with enJ:=(u~nJ)5j=1J(Unj)5.(\partial_{t}^{2}-\Delta_{N}){\tilde{u}}_{n}^{J}+({\tilde{u}}_{n}^{J})^{5}=e_{n}^{J},\hskip 10.00002pt\text{with }e_{n}^{J}:=({\tilde{u}}_{n}^{J})^{5}-\sum_{j=1}^{J}(U_{n}^{j})^{5}.

Let us show

Lemma 5.3.

We have

(5.17) limJlim supnenJL1L2=0,\lim_{J\to\infty}\limsup_{n\rightarrow\infty}\|e_{n}^{J}\|_{L^{1}L^{2}}=0,

and

(5.18) u~nt=0J=un+αnJ;limJlim supnSN()αnJL5L10=0.\vec{\tilde{u}}_{n\restriction t=0}^{J}=\vec{u}_{n}+\vec{\alpha}_{n}^{J};\hskip 10.00002pt\ \lim_{J\to\infty}\limsup_{n\to\infty}\|S_{N}(\cdot)\alpha_{n}^{J}\|_{L^{5}L^{10}}=0.
Proof.

We will first show (5.17). We have

(5.19) |enJ|J1jkJ|Unj|4|Unk|+|znJ|5+|znJ|j=1J|Unj|4.|e_{n}^{J}|\lesssim_{J}\sum_{1\leq j\neq k\leq J}|U_{n}^{j}|^{4}|U_{n}^{k}|+|z_{n}^{J}|^{5}+|z_{n}^{J}|\sum_{j=1}^{J}|U_{n}^{j}|^{4}.

Let us begin with the mixed terms |Unj|4|Unk||U_{n}^{j}|^{4}|U_{n}^{k}|. We start with the case j,kJdiffj,k\in J_{\text{diff}}. Notice that

|Unj|4|Unk||Vnj|4|Vnk|+|Unj|4|VnkUnk|+|Vnj||VnkUnk|4,|U_{n}^{j}|^{4}|U_{n}^{k}|\leq|V_{n}^{j}|^{4}|V_{n}^{k}|+|U_{n}^{j}|^{4}|V_{n}^{k}-U_{n}^{k}|+|V_{n}^{j}||V_{n}^{k}-U_{n}^{k}|^{4},

thus we get, by Hölder inequality

(5.20) |Unj|4|Unk|L1L2|Vnj|4|Vnk|L1L2+UnjL5L104VnkUnkL5L10+VnjL5L10VnkUnkL5L104.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\leq\left\||V_{n}^{j}|^{4}|V_{n}^{k}|\right\|_{L^{1}L^{2}}+\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{4}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}\\ +\left\|V_{n}^{j}\right\|_{L^{5}L^{10}}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}^{4}.

On the one hand, as VnjV_{n}^{j} and VnkV_{n}^{k} are rescaled solutions of the defocusing critical nonlinear wave equation in 3\mathbb{R}^{3} associated with orthogonal parameters, it is well known that, as nn goes to infinity (see for example [BG99])

(5.21) |Vnj|4|Vnk|L1L20.\left\||V_{n}^{j}|^{4}|V_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0.

On the other hand, as

supnUnjL5L10+VnjL5L10<,\sup_{n}\ \left\|U_{n}^{j}\right\|_{L^{5}L^{10}}+\left\|V_{n}^{j}\right\|_{L^{5}L^{10}}<\infty,

it follows from (5.9) that

(5.22) UnjL5L104VnkUnkL5L10+VnjL5L10VnkUnkL5L1040\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{4}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}+\left\|V_{n}^{j}\right\|_{L^{5}L^{10}}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}^{4}\longrightarrow 0

as nn goes to infinity, and thus (5.20) combined with (5.21) and (5.22) gives

(5.23) |Unj|4|Unk|L1L20,for j,kJdiff.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0,\ \text{for }j,k\in J_{\text{diff}}.

Let us now assume that jJcompj\in J_{\text{comp}} and kJdiffk\in J_{\text{diff}}. We have, in a same way as before

(5.24) |Unj|4|Unk|L1L2|Unj|4|Vnk|L1L2+UnjL5L104VnkUnkL5L10.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\leq\left\||U_{n}^{j}|^{4}|V_{n}^{k}|\right\|_{L^{1}L^{2}}+\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{4}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}.

On the one hand, we already saw that for kJdiffk\in J_{\text{diff}}

(5.25) UnjL5L104VnkUnkL5L10n0.\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{4}\left\|V_{n}^{k}-U_{n}^{k}\right\|_{L^{5}L^{10}}\underset{n\to\infty}{\longrightarrow}0.

On the other hand, by Hölder inequality and change of variables

|Unj|4|Vnk|L1L2UnjL5L103VnkUnjL5/2L5=UjL5L1031λk,n((r1Uj(ttj,n,x)5Vk(ttk,nλk,n,xλk,n)5𝑑x)1/2𝑑t)2/5=UjL5L1031λk,n((r1Uj(s,y)5Vk(s+tj,ntk,nλk,n,yλk,n)5𝑑y)1/2𝑑s)2/5.\left\||U_{n}^{j}|^{4}|V_{n}^{k}|\right\|_{L^{1}L^{2}}\leq\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{3}\left\|V_{n}^{k}U_{n}^{j}\right\|_{L^{5/2}L^{5}}\\ =\left\|U^{j}\right\|_{L^{5}L^{10}}^{3}\frac{1}{\sqrt{\lambda_{k,n}}}\Big{(}\int\Big{(}\int_{r\geq 1}U^{j}({t-t_{j,n}},{x})^{5}V^{k}\Big{(}\frac{t-t_{k,n}}{\lambda_{k,n}},\frac{x}{\lambda_{k,n}}\Big{)}^{5}\ dx\Big{)}^{1/2}dt\Big{)}^{2/5}\\ =\left\|U^{j}\right\|_{L^{5}L^{10}}^{3}\frac{1}{\sqrt{\lambda_{k,n}}}\Big{(}\int\Big{(}\int_{r\geq 1}U^{j}(s,y)^{5}V^{k}\Big{(}\frac{s+t_{j,n}-t_{k,n}}{\lambda_{k,n}},\frac{y}{\lambda_{k,n}}\Big{)}^{5}\ dy\Big{)}^{1/2}ds\Big{)}^{2/5}.

As the above expression is uniformly continuous in VkL5L10V^{k}\in L^{5}L^{10}, we can assume that VkV^{k} is contiuous and compactly supported. Then we get

(5.26) |Unj|4|Vnk|L1L21λk,n0\left\||U_{n}^{j}|^{4}|V_{n}^{k}|\right\|_{L^{1}L^{2}}\lesssim\frac{1}{\sqrt{\lambda_{k,n}}}\longrightarrow 0

and thus by (5.24), (5.25) and (5.26)

(5.27) |Unj|4|Unk|L1L20,for jJcomp,kJdiff.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0,\ \text{for }j\in J_{\text{comp}},\ k\in J_{\text{diff}}.

In a similar fashion we obtain

(5.28) |Unj|4|Unk|L1L20,for kJcomp,jJdiff.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0,\ \text{for }k\in J_{\text{comp}},\ j\in J_{\text{diff}}.

To conclude with the mixed term |Unj|4|Unk||U_{n}^{j}|^{4}|U_{n}^{k}|, let us deal with the case j,kJcompj,k\in J_{\text{comp}}. Then

(5.29) |Unj|4|Unk|L1L2=(r1Uj(ttj,n,x)8Uk(ttk,n,x)2𝑑x)1/2𝑑t.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}=\int\Big{(}\int_{r\geq 1}U^{j}\big{(}{t-t_{j,n}},{x}\big{)}^{8}U^{k}\big{(}{t-t_{k,n}},{x}\big{)}^{2}\ dx\Big{)}^{1/2}dt.

By orthogonality of the parameters,

(5.30) |tj,ntk,n|+,|t_{j,n}-t_{k,n}|\longrightarrow+\infty,

but, by change of variable s=ttj,ns=t-t_{j,n} we obtain from (5.29)

|Unj|4|Unk|L1L2=(r1Uj(s,x)8Uk(s+tj,ntk,n,x)2𝑑x)1/2𝑑s.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}=\int\Big{(}\int_{r\geq 1}U^{j}\big{(}{s},{x}\big{)}^{8}U^{k}\big{(}{s+t_{j,n}-t_{k,n}},{x}\big{)}^{2}\ dx\Big{)}^{1/2}ds.

Again, as this expression is uniformly continuous in (Uj,Uk)L5L10(U^{j},U^{k})\in L^{5}L^{10}, we may assume that both are continuous and compactly supported. But for such functions, the above expression vanishes for nn large enough by (5.30). Thus we have

(5.31) |Unj|4|Unk|L1L20,for j,kJcomp.\left\||U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0,\ \text{for }j,k\in J_{\text{comp}}.

We dealt with all the cases (5.23), (5.27), (5.28), (5.31) and shown that

(5.32) J,1jkJ|Unj|4|Unk|L1L20.\forall J,\ \left\|\sum_{1\leq j\neq k\leq J}|U_{n}^{j}|^{4}|U_{n}^{k}|\right\|_{L^{1}L^{2}}\longrightarrow 0.

Finally, by the decay of the remainder (5.15),

(5.33) limJlim supn|znJ|5L1L2=limJlim supnznJL5L105=0,\lim_{J\to\infty}\limsup_{n\to\infty}\left\||z_{n}^{J}|^{5}\right\|_{L^{1}L^{2}}=\lim_{J\to\infty}\limsup_{n\to\infty}\left\|z_{n}^{J}\right\|_{L^{5}L^{10}}^{5}=0,

and moreover, by Minkowski and Hölder inequalities

(5.34) ||znJ|j=1J|Unj|4L1L2znJL5L10j=1JUnjL5L104.\bigg{\|}||z_{n}^{J}|\sum_{j=1}^{J}|U_{n}^{j}|^{4}\bigg{\|}_{L^{1}L^{2}}\leq\left\|z_{n}^{J}\right\|_{L^{5}L^{10}}\sum_{j=1}^{J}\left\|U_{n}^{j}\right\|_{L^{5}L^{10}}^{4}.

By (5.15),

(5.35) limJlim supn||znJ|j=1J|Unj|4L1L2=0.\lim_{J\to\infty}\limsup_{n\to\infty}\bigg{\|}||z_{n}^{J}|\sum_{j=1}^{J}|U_{n}^{j}|^{4}\bigg{\|}_{L^{1}L^{2}}=0.

Combining (5.32), (5.33), (5.34) and (5.35), we thus proved the L1L2L^{1}L^{2} decay of the error term enJe_{n}^{J}, that is (5.17).

Let us now show (5.18). We have, by the definition of u~nJ{\tilde{u}}_{n}^{J} (5.13), of the remainder (5.14) and of the modified profiles (5.8), (5.6)

(5.36) u~nj(0)=jJcompjJUj(tj,n)+jJdiffjJσλj,n(Vj(tj,nλj,n))+wnJ.\vec{\tilde{u}}_{n}^{j}(0)=\sum_{\begin{subarray}{c}j\in J_{\text{comp}}\\ j\leq J\end{subarray}}\vec{U}^{j}({-t_{j,n}})+\sum_{\begin{subarray}{c}j\in J_{\text{diff}}\\ j\leq J\end{subarray}}\sigma_{\lambda_{j,n}}\Big{(}\vec{V}^{j}\Big{(}\frac{-t_{j,n}}{\lambda_{j,n}}\Big{)}\Big{)}+\vec{w}_{n}^{J}.

As a consequence of the definition (5.5) of UjU^{j} for jJcompj\in J_{\text{comp}}, we have, in (Bc)\mathcal{H}(B^{c}), as nn goes to infinity

(5.37) jJcomp,Uj(tj,n)=SN(tj,n)ψj+on(1).\forall j\in J_{\text{comp}},\hskip 10.00002pt\vec{U}^{j}({-t_{j,n}})=\vec{S}_{N}(-t_{j,n})\vec{\psi}^{j}\ +o_{n}(1).

Let us deal now with the first component of the diffusive profiles, the derivative component being handled in the same fashion. For jJdiffj\in J_{\text{diff}}, by the definition (5.7), this first component verifies, in H˙1\dot{H}^{1}

(5.38) σλj,n(Vj(tj,nλj,n))=1λj,n1/2Vj(tj,nλj,n,λj,n)=1λj,n1/2VLj(tj,nλj,n,λj,n)+on(1)=SN(tj,n)σλj,nψj+on(1),\sigma_{\lambda_{j,n}}\left(V^{j}\Big{(}\frac{-t_{j,n}}{\lambda_{j,n}}\Big{)}\right)=\frac{1}{\lambda_{j,n}^{1/2}}V^{j}\left(\frac{-t_{j,n}}{\lambda_{j,n}},\frac{\cdot}{\lambda_{j,n}}\right)\\ =\frac{1}{\lambda_{j,n}^{1/2}}V^{j}_{L}\left(\frac{-t_{j,n}}{\lambda_{j,n}},\frac{\cdot}{\lambda_{j,n}}\right)+o_{n}(1)=S_{N}(-t_{j,n})\sigma_{\lambda_{j,n}}\vec{\psi}^{j}\ +o_{n}(1),

where at the last line we have used Lemma 3.2. This last expansion (5.38), together with the similar one for the derivative component and (5.37), (5.36), the linear profile decomposition (5.4) and the Strichartz estimates for the Neumann flow (Proposition 2.7) gives (5.18), and ends the proof of the lemma. ∎

The proof of the nonlinear profile decomposition follows:

Proof of Proposition 5.2.

By (5.16) together with (5.17) and (5.18), the perturbative result of Proposition 2.13 gives, together with (5.15)

un=u~nJ+R~nJ,u_{n}=\tilde{u}_{n}^{J}+\tilde{R}_{n}^{J},

with

limJlim supnR~nJL5L10=0.\lim_{J\rightarrow\infty}\limsup_{n\rightarrow\infty}\|\tilde{R}_{n}^{J}\|_{L^{5}L^{10}}=0.

But (5.9) enables us to replace all the UnjU_{n}^{j} by VnjV_{n}^{j} for jJdiffj\in J_{\text{diff}} in the definition (5.13) of u~nJ\tilde{u}_{n}^{J} and ends the proof of the nonlinear profile decomposition. ∎

We are now in position to end the proof of the theorem. Indeed, by Proposition 5.2 together with (5.11), unu_{n} is in L5L10L^{5}L^{10} for nn large enough, and (5.2) is contradicted. Therefore the assumption (5.10) cannot hold, that is, J=1J=1: there is only one non-trivial profile in the decomposition (5.2):

(5.39) u0n=SN(t1,n)σλ1,nψ1+wn,SN()wnL5L100\vec{u}_{0}^{n}=S_{N}(-t_{1,n})\sigma_{\lambda_{1,n}}\vec{\psi}^{1}+\vec{w}_{n},\hskip 10.00002pt\|S_{N}(\cdot)\vec{w}_{n}\|_{L^{5}L^{10}}\longrightarrow 0

Let us show that it is the time-compact (t1,n=0t_{1,n}=0), scaling-compact (λ1,n=1\lambda_{1,n}=1) one.

As noticed before, as the scattering in the free space 3\mathbb{R}^{3} is well known, we have VjL5L10V^{j}\in L^{5}L^{10} for any jJdiffj\in J_{\text{diff}}. Therefore, if 1Jdiff1\in J_{\text{diff}}, the same proof as before yields the decomposition:

(5.40) un(t)=1λ1,n1/2V1(tt1,nλ1,n,λ1,n)+Rn(t)u_{n}(t)=\frac{1}{\lambda_{1,n}^{1/2}}V^{1}\Big{(}\frac{t-t_{1,n}}{\lambda_{1,n}},\frac{\cdot}{\lambda_{1,n}}\Big{)}+R_{n}(t)

with

(5.41) lim supnRnL5L10=0,\limsup_{n\rightarrow\infty}\|R_{n}\|_{L^{5}L^{10}}=0,

proving that unL5L10u_{n}\in L^{5}L^{10}, a contradiction. Thus 1Jcomp1\in J_{\text{comp}} i.e. λ1,n=1\lambda_{1,n}=1.

It remains to eliminate the case t1,n±t_{1,n}\longrightarrow\pm\infty. Recall that

(5.42) unL5((,0)L10),unL5((0,+)L10).\|u_{n}\|_{L^{5}\left((-\infty,0)L^{10}\right)}\longrightarrow\infty,\hskip 10.00002pt\|u_{n}\|_{L^{5}\left((0,+\infty)L^{10}\right)}\longrightarrow\infty.

Let us for example assume, by contradiction, that t1,n+t_{1,n}\longrightarrow+\infty. This implies

limnSN(t1,n)ψ1L5((,0)L10)=0,\lim_{n\to\infty}\left\|S_{N}(\cdot-t_{1,n})\vec{\psi}^{1}\right\|_{L^{5}\left((-\infty,0)L^{10}\right)}=0,

and we obtain, by the small data well-posedness theory, that for large nn, unL5((,0),L10)u_{n}\in L^{5}((-\infty,0),L^{10}) with

limnunL5((,0),L10)=0,\lim_{n\to\infty}\|u_{n}\|_{L^{5}((-\infty,0),L^{10})}=0,

contradicting (5.42). The case t1,nt_{1,n}\longrightarrow-\infty is eliminated in the same way.

Therefore, u0n\vec{u}_{0}^{n} writes:

u0n=ψ1+wn,SN()wnL5L100.\vec{u}_{0}^{n}=\vec{\psi}^{1}+\vec{w}_{n},\hskip 10.00002pt\|S_{N}(\cdot)\vec{w}_{n}\|_{L^{5}L^{10}}\longrightarrow 0.

Notice that, by the Pythagorean expansion (4.7) together with its L6L^{6} version (4.8), (ψ1)Ec\mathscr{E}(\vec{\psi}^{1})\leq E_{c}, and therefore

(ψ1)=Ec\mathscr{E}(\vec{\psi}^{1})=E_{c}

otherwise, by (5.40) and the definition of EcE_{c}, unu_{n} scatters. This implies, by the Pythagorean expansion again, together with (5.2)

wn(Bc)0.\|\vec{w}_{n}\|_{\mathcal{H}(B^{c})}\longrightarrow 0.

We take uc\vec{u}_{c} to be this profile:

uc:=ψ1.\vec{u}_{c}:=\vec{\psi}^{1}.

By the conservation of energy, we have (𝒮N(t)uc)=Ec\mathscr{E}(\vec{\mathscr{S}}_{N}(t)\vec{u}_{c})=E_{c} for any tt, and the same argument applied to

𝒮N(tn)uc\vec{\mathscr{S}}_{N}(t_{n})\vec{u}_{c}

for any sequence (tn)n1(t_{n})_{n\geq 1}\in\mathbb{R}^{\mathbb{N}} shows that the flow {t,𝒮N(t)uc}\big{\{}t\in\mathbb{R},\;\vec{\mathscr{S}}_{N}(t)\vec{u}_{c}\big{\}} has a compact closure in (Bc)\mathcal{H}(B^{c}). Indeed this sequence satisfies the same assumptions as un0\vec{u}^{0}_{n} at the beginning of the proof, and will therefore have a convergent subsequence in (Bc)\mathcal{H}(B^{c}) as well. Finally, observe that (uc)=Ec>0\mathscr{E}(\vec{u}_{c})=E_{c}>0 insures in particular that uc0\vec{u}_{c}\neq\vec{0}. ∎

6. Rigidity

In this section we prove:

Theorem 6.1.

Let (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}), radial, and u(t)=𝒮N(t)(u0,u1)u(t)=\mathscr{S}_{N}(t)(u_{0},u_{1}) be a solution of the energy critical defocusing wave equation outside the unit ball with Neumann boundary conditions (1.7)-(1.8)-(1.9). Assume that uu is global and that

K={u(t),t}K=\Big{\{}\vec{u}(t),\;t\in\mathbb{R}\Big{\}}

has compact closure in (Bc)\mathcal{H}(B^{c}). Then u=0u=0.

The proof follows the lines of the proof of [DKM13].

6.1. Preliminaries

We will use the following asymptotic energy property for the wave equation on 3\mathbb{R}^{3}:

Proposition 6.2.

Let R>0R>0. Let (v0,v1)(3)(v_{0},v_{1})\in\mathcal{H}(\mathbb{R}^{3}) and v=S3(v0,v1)v=S_{\mathbb{R}^{3}}(v_{0},v_{1}) be the solution of the linear wave equation on 3\mathbb{R}^{3} with initial data (v0,v1)(v_{0},v_{1}). Then

±limt±R+|t|+|t,r(rv(t,r))|2𝑑x=R+(r(rv0))2+r2v12dr.\sum_{\pm}\lim_{t\to\pm\infty}\int_{R+|t|}^{+\infty}|\partial_{t,r}(rv(t,r))|^{2}\,dx=\int_{R}^{+\infty}\left(\partial_{r}(rv_{0})\right)^{2}+r^{2}v_{1}^{2}\,dr.

We omit the easy proof, which relies on the equation (t2r2)u(t,r)=0(\partial_{t}^{2}-\partial_{r}^{2})u(t,r)=0. We note that by integration by parts,

(6.1) R+(t,r(ru0))2𝑑r+Ru02(R)=R+(t,r(u0))2r2𝑑r.\int_{R}^{+\infty}\left(\partial_{t,r}(ru_{0})\right)^{2}\,dr+Ru_{0}^{2}(R)=\int_{R}^{+\infty}\left(\partial_{t,r}(u_{0})\right)^{2}r^{2}\,dr.
Proposition 6.3.

There exists 𝔷>0\mathfrak{z}>0 and a radial, CC^{\infty} function Z=Z(|x|)Z=Z(|x|) on {x3,|x|>𝔷}\{x\in\mathbb{R}^{3},\;|x|>\mathfrak{z}\} such that

(6.2) ΔZ=Z5 for r>𝔷\displaystyle\Delta Z=Z^{5}\text{ for }r>\mathfrak{z}
(6.3) |rZ(r)+1r|+|Z(r)1r|Cr3\displaystyle\left|rZ^{\prime}(r)+\frac{1}{r}\right|+\left|Z(r)-\frac{1}{r}\right|\leq\frac{C}{r^{3}}
(6.4) limrζ+|Z(r)|=+\displaystyle\lim_{r\to\zeta^{+}}|Z(r)|=+\infty
(6.5) Z(r)0 for r>𝔷.\displaystyle Z^{\prime}(r)\neq 0\text{ for }r>\mathfrak{z}.
Proof.

The existence of 𝔷\mathfrak{z} and ZZ satisfying (6.2), (6.3) and (6.4) is proved in [DY18, Proposition 4.1] and we omit it.

To prove (6.5), we argue by contradiction. Assume that Z(R)=0Z^{\prime}(R)=0 for some R>𝔷R>\mathfrak{z}. Multiplying equation (6.2) by ZZ, integrating by parts for r>Rr>R and using the boundary condition Z(R)=0Z^{\prime}(R)=0, we obtain

{|x>R}|Z|2𝑑x+{|x>R}|Z|6𝑑x=0.\int_{\{|x>R\}}|\nabla Z|^{2}\,dx+\int_{\{|x>R\}}|Z|^{6}\,dx=0.

This proves that Z(r)=0Z(r)=0 for almost every r>Rr>R, contradicting (6.3). ∎

Remark 6.4.

Let {0}\ell\in\mathbb{R}\setminus\{0\} and

Z=1Z(r2).Z_{\ell}=\frac{1}{\ell}Z\left(\frac{r}{\ell^{2}}\right).

Then (6.2), (6.4) and (6.5) hold with ZZ replaced by ZZ_{\ell} and 𝔷\mathfrak{z} by 2𝔷\ell^{2}\mathfrak{z}, and there exists a constant CC_{\ell} such that

(6.6) |rZ(r)+r|+|Z(r)r|Cr3.\left|rZ^{\prime}_{\ell}(r)+\frac{\ell}{r}\right|+\left|Z_{\ell}(r)-\frac{\ell}{r}\right|\leq\frac{C_{\ell}}{r^{3}}.

6.2. Proof of Theorem 6.1

Step 1.

Let (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}) be as in Theorem 6.1. Let ε>0\varepsilon>0 be a small parameter to be specified. In all the proof we fix Rε>1R_{\varepsilon}>1 such that

(6.7) Rε+((ru0)2+u12)r2𝑑rε.\int_{R_{\varepsilon}}^{+\infty}\left((\partial_{r}u_{0})^{2}+u_{1}^{2}\right)r^{2}\,dr\leq\varepsilon.

In this step, we prove

(6.8) RRε,R+(r(ru0))2+r2u12drCR5u010(R).\forall R\geq R_{\varepsilon},\quad\int_{R}^{+\infty}\left(\partial_{r}(ru_{0})\right)^{2}+r^{2}u_{1}^{2}\,dr\leq CR^{5}u_{0}^{10}(R).

Let RRεR\geq R_{\varepsilon}. We define the radial functions v0H˙1(3)v_{0}\in\dot{H}^{1}(\mathbb{R}^{3}), v1L2(3)v_{1}\in L^{2}(\mathbb{R}^{3}) as follows:

(6.9) {(v0,v1)(r)=(u0,u1)(r) if r>R(v0,v1)(r)=(u0(R),0) if r(0,R).\begin{cases}(v_{0},v_{1})(r)=(u_{0},u_{1})(r)&\text{ if }r>R\\ (v_{0},v_{1})(r)=(u_{0}(R),0)&\text{ if }r\in(0,R).\end{cases}

We let v(t)=𝒮3(t)(v0,v1)v(t)=\mathscr{S}_{\mathbb{R}^{3}}(t)(v_{0},v_{1}) be the solution to the quintic wave equation on 3\mathbb{R}^{3} with initial data (v0,v1)(v_{0},v_{1}), and vL(r)=S3(v0,v1)v_{L}(r)=S_{\mathbb{R}^{3}}(v_{0},v_{1}) be the corresponding solution to the free wave equation. We note that by final speed of propagation

v(t,r)=u(t,r),r>R+|t|.v(t,r)=u(t,r),\quad r>R+|t|.

By the small data theory, since ε\varepsilon is small,

(6.10) suptv(t)vL(t)H˙1×L2C(v0,v1)H˙1×L25.\sup_{t\in\mathbb{R}}\left\|\vec{v}(t)-\vec{v}_{L}(t)\right\|_{\dot{H}^{1}\times L^{2}}\leq C\left\|(v_{0},v_{1})\right\|^{5}_{\dot{H}^{1}\times L^{2}}.

By Proposition 6.2,

(6.11) ±limt±R+|t|+|t,r(rvL(t,r))|2𝑑r=R+(r(ru0))2+u12dr.\sum_{\pm}\lim_{t\to\pm\infty}\int_{R+|t|}^{+\infty}|\partial_{t,r}(rv_{L}(t,r))|^{2}\,dr=\int_{R}^{+\infty}\left(\partial_{r}(ru_{0})\right)^{2}+u_{1}^{2}\,dr.

By (6.10), and finite speed of propagation

|R+|t|+|t,r(rvL(t,r))t,r(ru(t,r))|2𝑑r|C(R+((ru0)2+u12)r2𝑑r)5.\left|\int_{R+|t|}^{+\infty}\big{|}\partial_{t,r}(rv_{L}(t,r))-\partial_{t,r}(ru(t,r))\big{|}^{2}\,dr\right|\leq C\left(\int_{R}^{+\infty}\big{(}(\partial_{r}u_{0})^{2}+u_{1}^{2}\big{)}r^{2}dr\right)^{5}.

Combining with (6.11) and using that by the compactness of the closure of KK in (Bc)\mathcal{H}(B^{c})

limt±R+|t|+|t,r(ru(t,r))|2𝑑r=0,\lim_{t\to\pm\infty}\int_{R+|t|}^{+\infty}|\partial_{t,r}(ru(t,r))|^{2}\,dr=0,

we deduce

R+(r(ru0))2+u12drC(R+(ru0)2+u12)r2dr)5.\int_{R}^{+\infty}\left(\partial_{r}(ru_{0})\right)^{2}+u_{1}^{2}\,dr\leq C\left(\int_{R}^{+\infty}\left(\partial_{r}u_{0})^{2}+u_{1}^{2}\right)r^{2}dr\right)^{5}.

Combining with the integration by parts formula (6.1) and the smallness of ε\varepsilon, we deduce (6.8).

Step 2.

In this step we prove that there exists \ell\in\mathbb{R} and C>0C>0 such that for large RR,

(6.12) |u0(r)r|Cr3,r+ρ2u1(ρ)𝑑ρCr5.\left|u_{0}(r)-\frac{\ell}{r}\right|\leq\frac{C}{r^{3}},\quad\int_{r}^{+\infty}\rho^{2}u_{1}(\rho)\,d\rho\leq\frac{C}{r^{5}}.

First fix RR and RR^{\prime} such that RεRR2RR_{\varepsilon}\leq R\leq R^{\prime}\leq 2R. Letting ζ0(r)=ru0(r)\zeta_{0}(r)=ru_{0}(r), we have, using Cauchy-Schwarz, then Step 1

(6.13) |ζ0(R)ζ0(R)|RR|rζ0(r)|𝑑rRRR(rζ0)2𝑑r1R2ζ05(R).\left|\zeta_{0}(R)-\zeta_{0}(R^{\prime})\right|\leq\int_{R}^{R^{\prime}}|\partial_{r}\zeta_{0}(r)|\,dr\leq\sqrt{R}\sqrt{\int_{R}^{R^{\prime}}(\partial_{r}\zeta_{0})^{2}dr}\leq\frac{1}{R^{2}}\zeta_{0}^{5}(R).

Since by the definition (6.7) of RεR_{\varepsilon} and the integration by parts formula (6.1) one has

(6.14) 1Rζ02(R)ε,\frac{1}{R}\zeta_{0}^{2}(R)\leq\varepsilon,

we deduce from (6.13):

(6.15) |ζ0(R)ζ0(R)|ε2ζ0(R).\left|\zeta_{0}(R)-\zeta_{0}(R^{\prime})\right|\leq\varepsilon^{2}\zeta_{0}(R).

We apply this inequality between 2kR2^{k}R and 2k+1R2^{k+1}R for kk\in\mathbb{N} and a fixed RRεR\geq R_{\varepsilon}. This yields

|ζ0(2k+1R)ζ0(2kR)|ε2|ζ0(2kR)|\left|\zeta_{0}\left(2^{k+1}R\right)-\zeta_{0}\left(2^{k}R\right)\right|\lesssim\varepsilon^{2}\left|\zeta_{0}\left(2^{k}R\right)\right|

and thus

|ζ0(2k+1R)|(1+Cε2)|ζ0(2kR)|.\left|\zeta_{0}\left(2^{k+1}R\right)\right|\leq(1+C\varepsilon^{2})\left|\zeta_{0}\left(2^{k}R\right)\right|.

We deduce, by an easy induction:

|ζ0(2kR)|(1+Cε2)k|ζ0(R)|\left|\zeta_{0}\left(2^{k}R\right)\right|\leq(1+C\varepsilon^{2})^{k}|\zeta_{0}(R)|

Combining with (6.13) we obtain

(6.16) |ζ0(2kR)ζ0(2k+1R)|(1+Cε2)5k|ζ0(R)|5122kR2.\left|\zeta_{0}(2^{k}R)-\zeta_{0}(2^{k+1}R)\right|\lesssim\left(1+C\varepsilon^{2}\right)^{5k}|\zeta_{0}(R)|^{5}\frac{1}{2^{2k}R^{2}}.

Chosing ε\varepsilon small, so that (1+Cε2)5<4(1+C\varepsilon^{2})^{5}<4, we see that |ζ0(2kR)ζ0(2k+1R)|\sum\left|\zeta_{0}(2^{k}R)-\zeta_{0}(2^{k+1}R)\right| converges, and thus that ζ0(2kR)\zeta_{0}(2^{k}R) has a limit (R)\ell(R) as kk\to\infty. Summing (6.16) over all kk0k\geq k_{0}, we obtain

(6.17) |ζ0(2k0R)(R)|1R21(1+cε)k0[ζ0(R)|5,\left|\zeta_{0}(2^{k_{0}}R)-\ell(R)\right|\lesssim\frac{1}{R^{2}}\frac{1}{(1+c_{\varepsilon})^{k_{0}}}[\zeta_{0}(R)|^{5},

for some constant cε>0c_{\varepsilon}>0. Combining with (6.13), we see that

limrζ0(r)=(R),\lim_{r\to\infty}\zeta_{0}(r)=\ell(R),

and in particular the limit (R)\ell(R) does not depend on RR. We will simply denote it by \ell. By (6.17) at k0=1k_{0}=1, since ζ0\zeta_{0} is bounded

(6.18) |ζ0(R)|ζ01R2,\left|\zeta_{0}(R)-\ell\right|\lesssim_{\zeta_{0}}\frac{1}{R^{2}},

which yields the first inequality in (6.12).Combining with step 1, we obtain the second inequality in (6.12).

Step 3.

In this step, we assume =0\ell=0 and prove that (u0,u1)(0,0)(u_{0},u_{1})\equiv(0,0). Indeed by (6.15), if RRεR\geq R_{\varepsilon} and kk\in\mathbb{N},

|ζ0(2k+1R)|(1Cε2)|ζ0(2kR)|.\left|\zeta_{0}\left(2^{k+1}R\right)\right|\geq(1-C\varepsilon^{2})\left|\zeta_{0}\left(2^{k}R\right)\right|.

Hence by induction on kk,

|ζ0(2kR)|(1Cε2)k|ζ0(R)|.\left|\zeta_{0}\left(2^{k}R\right)\right|\geq(1-C\varepsilon^{2})^{k}|\zeta_{0}(R)|.

Since by the preceding step and the assumption R=0R=0, |ζ0(2kR)|1/2kR2|\zeta_{0}(2^{k}R)|\lesssim 1/{2^{k}R}^{2}, we deduce, chosing ε\varepsilon small enough and letting kk\to\infty that ζ0(R)=0\zeta_{0}(R)=0. Combining with (6.8) we deduce

RRεR+(rζ0)2+u12(r)dr=0,R\geq R_{\varepsilon}\Longrightarrow\int_{R}^{+\infty}(\partial_{r}\zeta_{0})^{2}+u_{1}^{2}(r)\,dr=0,

that is u0(r)u_{0}(r) and u1(r)u_{1}(r) are 0 for almost every rRεr\geq R_{\varepsilon}. Going back to the definition of RεR_{\varepsilon} we see that we can choose any Rε>1R_{\varepsilon}>1, which concludes this step.

Step 4.

We next assume 0\ell\neq 0. Let ZZ_{\ell} be as in Remark 6.4. In this step we prove that (u0Z,u1)(u_{0}-Z_{\ell},u_{1}) has a bounded support. Let f=uZf=u-Z_{\ell}. Then

(6.19) {t2fΔf=D(f):=k=15(5k)Z5kfk.ft=0=(f0,f1):=(u0Z,u1),\left\{\begin{aligned} \partial_{t}^{2}f-\Delta f&=D_{\ell}(f):=\sum_{k=1}^{5}\binom{5}{k}Z_{\ell}^{5-k}f^{k}.\\ \vec{f}_{\restriction t=0}&=(f_{0},f_{1}):=\left(u_{0}-Z_{\ell},u_{1}\right),\end{aligned}\right.

For ε>0\varepsilon>0 small, we fix Rε1R_{\varepsilon}^{\prime}\gg 1 such that

(6.20) Rε+(|rf0(r)|2+|f1(r)|2)r2𝑑rε2\displaystyle\int_{R_{\varepsilon}^{\prime}}^{+\infty}\left(|\partial_{r}f_{0}(r)|^{2}+|f_{1}(r)|^{2}\right)r^{2}dr\leq\varepsilon^{2}
(6.21) (Rε+|t|+Z10(r)r2𝑑r)12𝑑tε5.\displaystyle\int_{\mathbb{R}}\left(\int_{R_{\varepsilon}^{\prime}+|t|}^{+\infty}Z_{\ell}^{10}(r)r^{2}\,dr\right)^{\frac{1}{2}}dt\leq\varepsilon^{5}.

Let fLf_{L} be the solution of t2fL=ΔfL\partial_{t}^{2}f_{L}=\Delta f_{L} with

fLt=0=(f~0,f~1),\vec{f}_{L\restriction t=0}=(\tilde{f}_{0},\tilde{f}_{1}),

where (f~0,f~1)(\tilde{f}_{0},\tilde{f}_{1}) coincides with (f0,f1)(f_{0},f_{1}) for r>Rεr>R_{\varepsilon}^{\prime} and is defined as in (6.9). Using (6.19) and the assumptions (6.20) and (6.21) on RεR_{\varepsilon}^{\prime}, we obtain

(6.22) supt11{|x|>|t|+Rε}|t,x(f~(t)f~L(t))|L2ε4(f~0,f~1)H˙1×L2.\sup_{t\in\mathbb{R}}\left\|1\!\!1_{\{|x|>|t|+R_{\varepsilon}^{\prime}\}}\big{|}\nabla_{t,x}(\tilde{f}(t)-\tilde{f}_{L}(t))\big{|}\,\right\|_{L^{2}}\lesssim\varepsilon^{4}\Big{\|}(\tilde{f}_{0},\tilde{f}_{1})\Big{\|}_{\dot{H}^{1}\times L^{2}}.

Let RRεR\geq R_{\varepsilon}^{\prime}. Using that by Proposition 6.2,

±limt±R+(t,r(rf~L(t,r)))2𝑑rR+((r(rf~))2+r2f~12)𝑑r,\sum_{\pm}\lim_{t\to\pm\infty}\int_{R}^{+\infty}\left(\partial_{t,r}\big{(}r\tilde{f}_{L}(t,r)\big{)}\right)^{2}dr\gtrsim\int_{R}^{+\infty}\left(\left(\partial_{r}(r\tilde{f})\right)^{2}+r^{2}\tilde{f}_{1}^{2}\right)dr,

and since

±limt±R+(t,r(rf~(t,r)))2𝑑r=0,\sum_{\pm}\lim_{t\to\pm\infty}\int_{R}^{+\infty}\left(\partial_{t,r}\big{(}r\tilde{f}(t,r)\big{)}\right)^{2}dr=0,

we deduce from (6.22)

ε8R+((rf0)2+f12)r2𝑑rR+((r(rf0))2+r2f12)𝑑r,\varepsilon^{8}\int_{R}^{+\infty}\left((\partial_{r}f_{0})^{2}+f_{1}^{2}\right)r^{2}dr\gtrsim\int_{R}^{+\infty}\left(\big{(}\partial_{r}(rf_{0})\big{)}^{2}+r^{2}f_{1}^{2}\right)dr,

and thus

(6.23) ε8Rf02(R)R((r(rf0))2+r2f12)𝑑r.\varepsilon^{8}Rf_{0}^{2}(R)\gtrsim\int_{R}^{\infty}\left(\big{(}\partial_{r}(rf_{0})\big{)}^{2}+r^{2}f_{1}^{2}\right)dr.

Letting g0=rf0g_{0}=rf_{0}, we deduce by Cauchy-Schwarz that for RRεR\geq R_{\varepsilon}^{\prime}, kk\in\mathbb{N},

|g0(2k+1R)g0(2kR)|2kR2k+1R|rg0|𝑑rε4|g0(2kR)|.\left|g_{0}\left(2^{k+1}R\right)-g_{0}\left(2^{k}R\right)\right|\lesssim\int_{2^{k}R}^{2^{k+1}R}|\partial_{r}g_{0}|dr\lesssim\varepsilon^{4}\left|g_{0}(2^{k}R)\right|.

This yields by an easy induction |g0(2kR)|(1Cε4)k|g0(R)||g_{0}(2^{k}R)|\geq\left(1-C\varepsilon^{4}\right)^{k}|g_{0}(R)|, where C>0C>0 is a constant which is independent of ε\varepsilon. Since by Step 2,

C(2kR)2|g0(2kR)|,\frac{C}{\left(2^{k}R\right)^{2}}\geq\left|g_{0}\left(2^{k}R\right)\right|,

we obtain choosing ε\varepsilon small enough that g0(R)=0g_{0}(R)=0 for large RR. Combining with (6.23), we deduce that (f0(r),f1(r))=0(f_{0}(r),f_{1}(r))=0 a.e. for large RR, concluding this step.

Step 5.

In this step we still assume 0\ell\neq 0 and deduce a contradiction. We let

ρ=inf{R>c:R+((rf0)2+f12)r2𝑑r=0}\rho=\inf\left\{R>c\;:\;\int_{R}^{+\infty}\left((\partial_{r}f_{0})^{2}+f_{1}^{2}\right)r^{2}dr=0\right\}

and prove that ρ=max(1,𝔷2)\rho=\max(1,\mathfrak{z}\ell^{2}), i.e. that u0(r)=Z(r)u_{0}(r)=Z_{\ell}(r) almost everywhere for r>max(1,𝔷2)r>\max(1,\mathfrak{z}\ell^{2}). If 𝔷21\mathfrak{z}\ell^{2}\geq 1, we deduce

limr𝔷2|u0(r)|=+,\lim_{r\to\mathfrak{z}\ell^{2}}|u_{0}(r)|=+\infty,

a contradiction with the radial Sobolev embedding theorem. If 𝔷21\mathfrak{z}\ell^{2}\leq 1, we obtain

u0(r)=Z(r)u_{0}(r)=Z_{\ell}(r)

for all r>1r>1. Translating the solution in time, the same proof yields that for all tt in the domain of definition of uu,

(6.24) u(t,r)=Z(r),u(t,r)=Z_{\ell}(r),

a contradiction with the Neumann boundary condition, as given by Lemma 2.10. Note that by finite speed of propagation, the limit \ell in (6.24) is independent of tt.

To prove that ρ=max(1,𝔷2)\rho=\max(1,\mathfrak{z}\ell^{2}), we argue by contradiction, assuming ρ>max(1,𝔷2)\rho>\max(1,\mathfrak{z}\ell^{2}). By the preceding step and finite speed of propagation, the essential support of ff is included in {rρ+|t|}\{r\leq\rho+|t|\}. Thus ff is solution of

{t2fΔf=11{|x|ρ+|t|}D(f).ft=0=(f0,f1):=(u0Z,u1),\left\{\begin{aligned} \partial_{t}^{2}f-\Delta f&=1\!\!1_{\{|x|\leq\rho+|t|\}}D_{\ell}(f).\\ \vec{f}_{\restriction t=0}&=(f_{0},f_{1}):=\left(u_{0}-Z_{\ell},u_{1}\right),\end{aligned}\right.

Fix Rε′′(1,ρ)R_{\varepsilon}^{\prime\prime}\in(1,\rho) such that,

Rε′′+(|rf0(r)|2+|f1(r)|2)r2𝑑rε2\displaystyle\int_{R_{\varepsilon}^{\prime\prime}}^{+\infty}\left(|\partial_{r}f_{0}(r)|^{2}+|f_{1}(r)|^{2}\right)r^{2}dr\leq\varepsilon^{2}
(Rε′′+|t|ρ+|t|Z10(r)r2𝑑r)12𝑑tε5.\displaystyle\int_{\mathbb{R}}\left(\int_{R_{\varepsilon}^{\prime\prime}+|t|}^{\rho+|t|}Z_{\ell}^{10}(r)r^{2}\,dr\right)^{\frac{1}{2}}dt\leq\varepsilon^{5}.

The same argument as in the preceding step, replacing RεR_{\varepsilon}^{\prime} by Rε′′R_{\varepsilon}^{\prime\prime}, yields that (f0,f1)=0(f_{0},f_{1})=0 for almost every r>Rε′′r>R_{\varepsilon}^{\prime\prime}, which contradicts the definition of ρ\rho. The proof is complete.

We are now in position to conclude

Proof of Theorem 1.1.

By contradiction, assume that EcE_{c}, as defined by (5.1), is finite. Then Theorem 5.1 shows that there exists a solution uc\vec{u}_{c} to (1.7)-(1.8)-(1.9) such that {uc(t),t}\big{\{}\vec{u}_{c}(t),\ t\in\mathbb{R}\big{\}} has a compact closure in (Bc)\mathcal{H}{(B^{c})}, but by Theorem 6.1, such a solution cannot exist. Thus Ec=+E_{c}=+\infty, and by Proposition 2.12, all the solutions of (1.7)-(1.8)-(1.9) scatter. ∎

7. Focusing case

In this section we sketch the proofs of Theorems 1.2 and 1.3. Subsection 7.1 is dedicated to the proof of a trapping property for solutions below the energy of the 3\mathbb{R}^{3} ground state WW that is important in the proof of both of these results. Subsection 7.2 concerns Theorem 1.2 and Subsection 7.3 Theorem 1.3. Finally, in Subsection 7.4, we comment on the assumptions of these two theorems, and prove that the exact analog of Theorem 1.2 is not true when 3B(0,1)\mathbb{R}^{3}\setminus B(0,1) is replaced by a more general domain.

7.1. Trapping by the energy

Recall that

W(x)=1(1+|x|23)12W(x)=\frac{1}{\left(1+\frac{|x|^{2}}{3}\right)^{\frac{1}{2}}}

is the ground state of the focusing critical wave equation on 3\mathbb{R}^{3}. If (f,g)(3)(f,g)\in\mathcal{H}(\mathbb{R}^{3}), we denote by

3(f,g)=123|f|2+123|g|2163|f|6.\mathscr{E}_{\mathbb{R}^{3}}(f,g)=\frac{1}{2}\int_{\mathbb{R}^{3}}|\nabla f|^{2}+\frac{1}{2}\int_{\mathbb{R}^{3}}|g|^{2}-\frac{1}{6}\int_{\mathbb{R}^{3}}|f|^{6}.
Proposition 7.1.

Let uu be a solution of (1.11) with Neumann boundary condition (1.8) and initial data (1.9). Let II be its maximal interval of existence. Assume (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0). Then the sign of Bc|u(t)|23|W|2\int_{B^{c}}|\nabla u(t)|^{2}-\int_{\mathbb{R}^{3}}|\nabla W|^{2} is independent of tIt\in I, and there exists δ>0\delta>0 depending only on (u0,u1)\mathscr{E}(u_{0},u_{1}) such that

(7.1) tI,|Bc|u(t,x)|23|W(x)|2𝑑x|δ.\forall t\in I,\quad\left|\int_{B^{c}}|\nabla u(t,x)|^{2}-\int_{\mathbb{R}^{3}}|\nabla W(x)|^{2}dx\right|\geq\delta.
Proof.

For (f,g)(f,g)\in\mathcal{H}, we denote by (f~,g¯):=𝒫(f,g)(\tilde{f},\bar{g}):=\vec{\mathcal{P}}(f,g), the extension of (f,g)(f,g) to (3)\mathcal{H}(\mathbb{R}^{3}) by (f(1),0)(f(1),0), as defined in definition 2.1. Observe that (f~,g¯)(\tilde{f},\bar{g}) verifies

3|f~|2=Bc|f|2 and 3f~6Bcf6,\int_{\mathbb{R}^{3}}\big{|}\nabla\tilde{f}\big{|}^{2}=\int_{B^{c}}|\nabla f|^{2}\text{ and }\int_{\mathbb{R}^{3}}\tilde{f}^{6}\geq\int_{B^{c}}f^{6},
3|g¯|2=Bc|g|2,\int_{\mathbb{R}^{3}}|\bar{g}|^{2}=\int_{B^{c}}|g|^{2},

and

(7.2) 3(f~,g¯)(f,g).\mathscr{E}_{\mathbb{R}^{3}}\left(\tilde{f},\bar{g}\right)\leq\mathscr{E}\left(f,g\right).

Let uu satisfy the assumptions of Proposition 7.1. Then by conservation of the energy and (7.2),

tI,3(u~(t),tu¯(t))(u0,u1)<3(W,0).\forall t\in I,\quad\mathscr{E}_{\mathbb{R}^{3}}\left(\tilde{u}(t),\overline{\partial_{t}u}(t)\right)\leq\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0).

The conclusion of the proposition then follows from the variational properties of the ground-state WW on 3\mathbb{R}^{3}, see e.g. [KM06, Lemma 3.4]. ∎

7.2. Scattering

Note that by Proposition 7.1 and the radial Sobolev inequality (see Remark 2.3), any solution of (1.11)-(1.8)-(1.9) that satisfies (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0), Bc|u0|2<3|W|2\int_{B^{c}}|\nabla u_{0}|^{2}<\int_{\mathbb{R}^{3}}|\nabla W|^{2} is global.

Using Proposition 7.1, the proof of Theorem 1.2 follows exactly the same lines as the proof of Theorem 1.1.

Recall that according to [KM08], any solution of the quintic focusing wave equation on 3\mathbb{R}^{3} with initial data (v0,v1)(H˙1×L2)(3)(v_{0},v_{1})\in(\dot{H}^{1}\times L^{2})\left(\mathbb{R}^{3}\right) such that

3|v0|2<3|W|2 and 3(v0,v1)<3(W,0)\int_{\mathbb{R}^{3}}|\nabla v_{0}|^{2}<\int_{\mathbb{R}^{3}}|\nabla W|^{2}\text{ and }\mathscr{E}_{\mathbb{R}^{3}}(v_{0},v_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0)

scatters to a linear solution.

Arguing by contradiction and using the arguments of Sections 3, 4, and 5, we see that it is sufficient to prove:

Theorem 7.2.

Let (u0,u1)(Bc)(u_{0},u_{1})\in\mathcal{H}(B^{c}), radial, and u(t)u(t) be a solution of the energy critical focusing wave equation outside the unit ball with Neumann boundary conditions (1.11)-(1.8)-(1.9). Assume that uu is global and that

K={u(t),t}K=\Big{\{}\vec{u}(t),\;t\in\mathbb{R}\Big{\}}

has compact closure in (Bc)\mathcal{H}(B^{c}). Then u0u\equiv 0.

Note that it would be sufficient to prove Theorem 7.2 with the additional assumptions (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0), Bc|u0|2<3|W|2\int_{B^{c}}|\nabla u_{0}|^{2}<\int_{\mathbb{R}^{3}}|\nabla W|^{2}, but these assumptions are not needed to obtain the conclusion of the theorem.

The proof of Theorem 7.2 is the same as the proof of the Theorem 6.1 in Section 6, except that in Steps 4 and 5 the solution ZZ_{\ell} of the elliptic equation ΔZ=Z5\Delta Z_{\ell}=Z_{\ell}^{5} must be replaced by the solution WW_{\ell} of the elliptic equation ΔW=W5-\Delta W_{\ell}=W_{\ell}^{5}, where

(7.3) W(x)=3W(3x2)=3(1+3|x|24)1/2,W_{\ell}(x)=\frac{\sqrt{3}}{\ell}W\left(\frac{3x}{\ell^{2}}\right)=\frac{\sqrt{3}}{\ell\left(1+\frac{3|x|^{2}}{\ell^{4}}\right)^{1/2}},

so that

|W(x)|x||1|x|3,|x|1.\left|W_{\ell}(x)-\frac{\ell}{|x|}\right|\lesssim\frac{1}{|x|^{3}},\quad|x|\gg 1.

Also, since W(x)W_{\ell}(x) is defined for all x3x\in\mathbb{R}^{3}, whereas Z(x)Z_{\ell}(x) is only defined for r>𝔷2r>\mathfrak{z}\ell^{2}, we must replace max(1,𝔷2)\max(1,\mathfrak{z}\ell^{2}) everywhere in these two steps of the proof by 11. The key point to obtain the contradiction is that rW(1)0\partial_{r}W_{\ell}(1)\neq 0 for any 0\ell\neq 0, i.e. that WW_{\ell} is not a stationary solution of the focusing wave equation on BcB^{c} with Neumann boundary condition, which can be easily checked on the explicit formula (7.3) .

7.3. Blow-up

Using Proposition 7.1, the proof of Theorem 1.3 is very close to the proof of its analog on the whole space 3\mathbb{R}^{3}, see Theorem 3.7 and the proof of Theorem 1.1, (ii) in section 7 of [KM08]. Let us mention that this argument is inspired by the work of H.A. Levine [Lev74].

Let us first assume that u0H1(Bc)=H˙1(Bc)L2(Bc)u_{0}\in H^{1}(B^{c})=\dot{H}^{1}(B^{c})\cap L^{2}(B^{c}). Using the equation satisfied by uu, one sees that u(t)L2(Bc)u(t)\in L^{2}(B^{c}) for all tt and, denoting y(t)=Bcu2(t,x)𝑑xy(t)=\int_{B^{c}}u^{2}(t,x)dx, that

y(t)=2Bcutu,y′′(t)=2Bcu62Bc|u|2+Bc(tu)2.y^{\prime}(t)=2\int_{B^{c}}u\partial_{t}u,\quad y^{\prime\prime}(t)=2\int_{B^{c}}u^{6}-2\int_{B^{c}}|\nabla u|^{2}+\int_{B^{c}}(\partial_{t}u)^{2}.

Note that we have used the boundary condition nuBc=0\partial_{n}u_{\restriction\partial B^{c}}=0 which implies BcuΔu=Bc|u|2\int_{B^{c}}u\Delta u=-\int_{B^{c}}|\nabla u|^{2}.

Recall that 3(W,0)=133|W|2\mathscr{E}_{\mathbb{R}^{3}}(W,0)=\frac{1}{3}\int_{\mathbb{R}^{3}}|\nabla W|^{2}. As in the proof of Theorem 3.7 of [KM08], one can write, for tt in the domain of existence of uu,

y′′(t)\displaystyle y^{\prime\prime}(t) =12(u0,u1)+4Bc|u|2+8Bc(tu)2\displaystyle=-12\mathscr{E}(u_{0},u_{1})+4\int_{B^{c}}|\nabla u|^{2}+8\int_{B^{c}}(\partial_{t}u)^{2}
=8Bc(tu)2+4Bc|u|243|W|2+123(W,0)12(u0,u1)\displaystyle=8\int_{B^{c}}(\partial_{t}u)^{2}+4\int_{B^{c}}|\nabla u|^{2}-4\int_{\mathbb{R}^{3}}|\nabla W|^{2}+12\mathscr{E}_{\mathbb{R}^{3}}(W,0)-12\mathscr{E}(u_{0},u_{1})
8Bc(tu)2+δ0,\displaystyle\geq 8\int_{B^{c}}(\partial_{t}u)^{2}+\delta_{0},

where δ0=123(W,0)12(u0,u1)>0\delta_{0}=12\mathscr{E}_{\mathbb{R}^{3}}(W,0)-12\mathscr{E}(u_{0},u_{1})>0 and we have used that by Proposition 7.1, Bc|u(t)|2>3|W|2\int_{B^{c}}|\nabla u(t)|^{2}>\int_{\mathbb{R}^{3}}|\nabla W|^{2} for all tt.

The end of the proof that uu blows up in finite time is exactly as the end of the proof of Theorem 3.7, p.165 of [KM08] and we omit it.

To treat the general case u0H˙1(Bc)u_{0}\in\dot{H}^{1}(B^{c}) one should use a localized version of Bcu2(t)\int_{B^{c}}u^{2}(t). These bring out new terms in the preceding computation, that can be controled using finite speed of propagation. We refer to [KM08, p.205-206] for the details.

7.4. Comments on the assumptions

Consider the nonlinear focusing wave equation (1.1) with ι=1\iota=-1, and Neumann boundary condition (1.2) in a general open domain Ω\Omega of 3\mathbb{R}^{3}. We claim that the analog of Theorem 1.2 does not hold in general. Indeed, first consider the case of a half-plane:

Ω={(x1,x2,x3)2,x1>0}.\Omega=\left\{(x_{1},x_{2},x_{3})\in\mathbb{R}^{2},\;x_{1}>0\right\}.

Let ww be the restriction of WW to Ω\Omega. Then ww is a solution of Δw=w5-\Delta w=w^{5}. Since WW is radial, ww satisfies in addition the Neumann boundary condition (1.2). This yields a non-scattering solution ww of (1.1), (1.2) such that

Ω|w|2=123|W|2,(w(0))=123(W,0),\int_{\Omega}|\nabla w|^{2}=\frac{1}{2}\int_{\mathbb{R}^{3}}|\nabla W|^{2},\quad\mathscr{E}(\vec{w}(0))=\frac{1}{2}\mathscr{E}_{\mathbb{R}^{3}}(W,0),

which proves that one cannot generalise Theorem 1.2 in this setting. Similarly, for ε>0\varepsilon>0, the solution wεw_{\varepsilon} of (1.1), (1.2) with initial data ((1+ε)w,0)((1+\varepsilon)w,0) blows up in finite time by [KM08]. This solution satisfies

Ω|w|2=(1+ε)223|W|2,(w(0))<123(W,0),\int_{\Omega}|\nabla w|^{2}=\frac{(1+\varepsilon)^{2}}{2}\int_{\mathbb{R}^{3}}|\nabla W|^{2},\quad\mathscr{E}(\vec{w}(0))<\frac{1}{2}\mathscr{E}_{\mathbb{R}^{3}}(W,0),

which shows that the assumptions (u0,u1)<3(W,0)\mathscr{E}(u_{0},u_{1})<\mathscr{E}_{\mathbb{R}^{3}}(W,0), Ω|u0|2<3|W|2\int_{\Omega}|\nabla u_{0}|^{2}<\int_{\mathbb{R}^{3}}|\nabla W|^{2} is not sufficient to ensure global existence on the half-plane.

We now give a similar example when Ω\Omega is an exterior domain. Assume that Ω=3K\Omega=\mathbb{R}^{3}\setminus K, where KK is bounded subset of 3\mathbb{R}^{3} with a smooth boundary K=Ω\partial K=\partial\Omega containing a portion of a plane. Without loss of generality, we can assume (translating and rescaling Ω\Omega):

{0}×[1,+1]2Ω,B(0,1){x1>0}Ω.\{0\}\times[-1,+1]^{2}\subset\partial\Omega,\quad B(0,1)\cap\{x_{1}>0\}\subset\Omega.

According to [KST09], for all ε>0\varepsilon>0, there exists a radial solution zz of the focusing critical wave equation on 3\mathbb{R}^{3}, blowing-up in finite time T>0T>0 and such that

lim suptT3|(z(t,x)1tW(xt2))|2+(tz(t,x))2dxε,(z(0))(W,0)+ε.\limsup_{t\to T}\int_{\mathbb{R}^{3}}\left|\nabla\left(z(t,x)-\frac{1}{t}W\left(\frac{x}{t^{2}}\right)\right)\right|^{2}+(\partial_{t}z(t,x))^{2}\,dx\leq\varepsilon,\quad\mathscr{E}(\vec{z}(0))\leq\mathscr{E}(W,0)+\varepsilon.

using finite speed of propagation, time translating and rescaling the solution, we can assume that the support of z(t)\vec{z}(t) is included in B(0,1)B(0,1) for all t[0,T)t\in[0,T). The restriction uu of zz to x1>0x_{1}>0 is then a solution of (1.1), (1.2), (1.3) that satisfies

(u0,u1)123(W,0)+ε,lim suptTΩ|u(t)|2+Ω(tu(t))2123|W|2+ε,\mathscr{E}(u_{0},u_{1})\leq\frac{1}{2}\mathscr{E}_{\mathbb{R}^{3}}(W,0)+\varepsilon,\quad\limsup_{t\to T}\int_{\Omega}|\nabla u(t)|^{2}+\int_{\Omega}(\partial_{t}u(t))^{2}\leq\frac{1}{2}\int_{\mathbb{R}^{3}}|\nabla W|^{2}+\varepsilon,

proving that a generalization of Theorem 1.2 is hopeless in this setting also.

In view of this example, we conjecture that Theorem 1.3 cannot be either generalised to other geometries, and that the radiality assumptions in Theorems 1.2 and 1.3 is also necessary. More precisely, a natural conjecture is that the energy threshold to ensure energy trapping and a blow-up scattering/dichotomy in the case of Neumann boundary condition is exactly 123(W,0)\frac{1}{2}\mathscr{E}_{\mathbb{R}^{3}}(W,0). This is of course the case when Ω\Omega is a half-plane, since one can then use the result on 3\mathbb{R}^{3} after extending the solution by symmetry to the whole space.

Acknowledgements

The authors thank Fabrice Planchon for interesting discussions about the problem.

References

  • [AS13] Farah Abou Shakra, Asymptotics of the critical nonlinear wave equation for a class of non-star-shaped obstacles, J. Hyperbolic Differ. Equ. 10 (2013), no. 3, 495–522. MR 3104079
  • [AS14] by same author, Asymptotics of wave models for non star-shaped geometries, Discrete Contin. Dyn. Syst. Ser. S 7 (2014), no. 2, 347–362. MR 3109476
  • [BB18] Jean-Marc Bouclet and Nicolas Burq, Sharp resolvent and time decay estimates for dispersive equations on asymptotically Euclidean backgrounds, Preprint arXiv:1810.01711, 2018.
  • [BG99] Hajer Bahouri and Patrick Gérard, High frequency approximation of solutions to critical nonlinear wave equations, Amer. J. Math. 121 (1999), no. 1, 131–175. MR MR1705001 (2000i:35123)
  • [BGT04] N. Burq, P. Gérard, and N. Tzvetkov, On nonlinear Schrödinger equations in exterior domains, Ann. Inst. H. Poincaré Anal. Non Linéaire 21 (2004), no. 3, 295–318. MR 2068304
  • [BLP08] Nicolas Burq, Gilles Lebeau, and Fabrice Planchon, Global existence for energy critical waves in 3-D domains, J. Amer. Math. Soc. 21 (2008), no. 3, 831–845. MR 2393429
  • [BP09] Nicolas Burq and Fabrice Planchon, Global existence for energy critical waves in 3-D domains: Neumann boundary conditions, Amer. J. Math. 131 (2009), no. 6, 1715–1742. MR 2567505
  • [BS98] Hajer Bahouri and Jalal Shatah, Decay estimates for the critical semilinear wave equation, Ann. Inst. H. Poincaré Anal. Non Linéaire 15 (1998), no. 6, 783–789. MR MR1650958 (99h:35136)
  • [BSS09] Matthew D. Blair, Hart F. Smith, and Christopher D. Sogge, Strichartz estimates for the wave equation on manifolds with boundary, Ann. Inst. H. Poincaré Anal. Non Linéaire 26 (2009), no. 5, 1817–1829. MR 2566711
  • [Bur02] N. Burq, Semi-classical estimates for the resolvent in nontrapping geometries, International Mathematics Research Notices 2002 (2002), no. 5, 221–241.
  • [Bur03] Nicolas Burq, Global Strichartz estimates for nontrapping geometries: about an article by H. F. Smith and C. D. Sogge: “Global Strichartz estimates for nontrapping perturbations of the Laplacian” [Comm. Partial Differential Equation 25 (2000), no. 11-12 2171–2183; MR1789924 (2001j:35180)], Comm. Partial Differential Equations 28 (2003), no. 9-10, 1675–1683. MR 2001179
  • [DKM11] Thomas Duyckaerts, Carlos Kenig, and Frank Merle, Universality of blow-up profile for small radial type II blow-up solutions of the energy-critical wave equation, J. Eur. Math. Soc. (JEMS) 13 (2011), no. 3, 533–599. MR 2781926
  • [DKM12] by same author, Scattering for radial, bounded solutions of focusing supercritical wave equations, IMRN (2012).
  • [DKM13] by same author, Classification of radial solutions of the focusing, energy-critical wave equation, Cambridge Journal of Mathematics 1 (2013), no. 1, 75–144.
  • [DKM19] by same author, Scattering profile for global solutions of the energy-critical wave equation, J. Eur. Math. Soc. (JEMS) 21 (2019), no. 7, 2117–2162.
  • [DY18] Thomas Duyckaerts and Jianwei Yang, Blow-up of a critical Sobolev norm for energy-subcritical and energy-supercritical wave equations, Anal. PDE 11 (2018), no. 4, 983–1028.
  • [DY19] by same author, Scattering to a stationary solution for the superquintic radial wave equation outside an obstacle, Preprint arXiv:1910.00811, 2019.
  • [FXC11] DaoYuan Fang, Jian Xie, and Thierry Cazenave, Scattering for the focusing energy-subcritical nonlinear Schrödinger equation, Sci. China Math. 54 (2011), no. 10, 2037–2062. MR 2838120
  • [Gér98] Patrick Gérard, Description du défaut de compacité de l’injection de Sobolev, ESAIM Control Optim. Calc. Var. 3 (1998), 213–233. MR 1632171
  • [GNN81] B. Gidas, Wei Ming Ni, and L. Nirenberg, Symmetry of positive solutions of nonlinear elliptic equations in 𝐑n{\bf R}^{n}, Mathematical analysis and applications, Part A, Adv. in Math. Suppl. Stud., vol. 7, Academic Press, New York, 1981, pp. 369–402. MR MR634248 (84a:35083)
  • [Gri90] Manoussos G. Grillakis, Regularity and asymptotic behaviour of the wave equation with a critical nonlinearity, Ann. of Math. (2) 132 (1990), no. 3, 485–509. MR 1078267 (92c:35080)
  • [GV87] J. Ginibre and G. Velo, Conformal invariance and time decay for nonlinear wave equations. I, II, Ann. Inst. H. Poincaré Phys. Théor. 47 (1987), no. 3, 221–261, 263–276. MR 921307
  • [GV95] by same author, Generalized Strichartz inequalities for the wave equation, Partial differential operators and mathematical physics (Holzhau, 1994), Oper. Theory Adv. Appl., vol. 78, Birkhäuser, Basel, 1995, pp. 153–160. MR 1365328
  • [KM06] Carlos E. Kenig and Frank Merle, Global well-posedness, scattering and blow-up for the energy-critical, focusing, non-linear Schrödinger equation in the radial case, Invent. Math. 166 (2006), no. 3, 645–675. MR MR2257393 (2007g:35232)
  • [KM08] by same author, Global well-posedness, scattering and blow-up for the energy-critical focusing non-linear wave equation, Acta Math. 201 (2008), no. 2, 147–212. MR MR2461508
  • [KST09] Joachim Krieger, Wilhelm Schlag, and Daniel Tataru, Slow blow-up solutions for the H1(3)H^{1}(\mathbb{R}^{3}) critical focusing semilinear wave equation, Duke Math. J. 147 (2009), no. 1, 1–53. MR MR2494455
  • [KT98] Markus Keel and Terence Tao, Endpoint Strichartz estimates, Amer. J. Math. 120 (1998), no. 5, 955–980. MR 1646048
  • [KVZ16] Rowan Killip, Monica Visan, and Xiaoyi Zhang, Quintic NLS in the exterior of a strictly convex obstacle, Amer. J. Math. 138 (2016), no. 5, 1193–1346.
  • [Laf17] D. Lafontaine, About the wave equation outside two strictly convex obstacles, accepted for publication in Communications in PDE, https://arxiv.org/abs/1711.09734 (2017).
  • [Laf19] by same author, Scattering for nls with a sum of two repulsive potential, To appear in Annales de l’Institut Fourier (2019).
  • [Lev74] Howard A. Levine, Instability and nonexistence of global solutions to nonlinear wave equations of the form Putt=Au+(u)Pu_{tt}=-Au+\mathcal{F}(u), Trans. Amer. Math. Soc. 192 (1974), 1–21. MR MR0344697 (49 #9436)
  • [LS95] Hans Lindblad and Christopher D. Sogge, On existence and scattering with minimal regularity for semilinear wave equations, J. Funct. Anal. 130 (1995), no. 2, 357–426. MR 1335386
  • [Met04] Jason L. Metcalfe, Global Strichartz estimates for solutions to the wave equation exterior to a convex obstacle, Trans. Amer. Math. Soc. 356 (2004), no. 12, 4839–4855. MR 2084401
  • [Mor61] Cathleen S. Morawetz, The decay of solutions of the exterior initial-boundary value problem for the wave equation, Comm. Pure Appl. Math. 14 (1961), 561–568. MR MR0132908 (24 #A2744)
  • [Mor75] by same author, Decay for solutions of the exterior problem for the wave equation, Comm. Pure Appl. Math. 28 (1975), 229–264. MR MR0372432 (51 #8641)
  • [MRS77] Cathleen S. Morawetz, James V. Ralston, and Walter A. Strauss, Decay of solutions of the wave equation outside nontrapping obstacles, Comm. Pure Appl. Math. 30 (1977), no. 4, 447–508.
  • [Poh65] SI Pohozaev, On the eigenfunctions of the equation δu+λf(u)=0\delta u+\lambda f(u)=0, Sov. Math. Doklady, vol. 6, 1965, pp. 1408–1411.
  • [PV09] Fabrice Planchon and Luis Vega, Bilinear virial identities and applications, Ann. Sci. Éc. Norm. Supér. (4) 42 (2009), no. 2, 261–290. MR 2518079 (2010b:35441)
  • [PV12] by same author, Scattering for solutions of NLS in the exterior of a 2D star-shaped obstacle, Math. Res. Lett. 19 (2012), no. 4, 887–897. MR 3008422
  • [SS94] Jalal Shatah and Michael Struwe, Well-posedness in the energy space for semilinear wave equations with critical growth, Internat. Math. Res. Notices (1994), no. 7, 303ff., approx. 7 pp. (electronic). MR MR1283026 (95e:35132)
  • [SS95] Hart F. Smith and Christopher D. Sogge, On the critical semilinear wave equation outside convex obstacles, J. Amer. Math. Soc. 8 (1995), no. 4, 879–916. MR 1308407
  • [SS00] by same author, Global Strichartz estimates for nontrapping perturbations of the Laplacian, Comm. Partial Differential Equations 25 (2000), no. 11-12, 2171–2183. MR 1789924
  • [SS07] by same author, On the LpL^{p} norm of spectral clusters for compact manifolds with boundary, Acta Math. 198 (2007), no. 1, 107–153. MR 2316270
  • [Str88] Michael Struwe, Globally regular solutions to the u5u^{5} Klein-Gordon equation, Ann. Scuola Norm. Sup. Pisa Cl. Sci. (4) 15 (1988), no. 3, 495–513 (1989). MR 1015805 (90j:35142)