06.07.2013 Views

Besov spaces and self-similar solutions for the wave-map equation

Besov spaces and self-similar solutions for the wave-map equation

Besov spaces and self-similar solutions for the wave-map equation

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

<strong>Besov</strong> <strong>spaces</strong> <strong>and</strong> <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong> <strong>for</strong> <strong>the</strong> <strong>wave</strong>-<strong>map</strong><br />

<strong>equation</strong><br />

P. Germain<br />

June 27, 2007<br />

Abstract<br />

Self-<strong>similar</strong> <strong>solutions</strong> play a crucial role in <strong>the</strong> blow-up <strong>the</strong>ory <strong>for</strong> <strong>the</strong> <strong>wave</strong>-<strong>map</strong><br />

<strong>equation</strong>; in space-dimension d, <strong>the</strong>y correspond to data in <strong>the</strong> infinite energy space<br />

˙B d/2<br />

2,∞ × ˙ B d/2−1<br />

2,∞ at <strong>the</strong> time of <strong>the</strong> blow-up. However, <strong>solutions</strong> to this <strong>equation</strong> are<br />

generally considered <strong>for</strong> data in <strong>the</strong> st<strong>and</strong>ard finite energy <strong>spaces</strong> ˙ Hd/2 × ˙ Hd/2−1 .<br />

We show in this article that it is possible to build up <strong>solutions</strong> of <strong>the</strong> covariant<br />

<strong>wave</strong>-<strong>map</strong> <strong>equation</strong> <strong>for</strong> data which are small <strong>and</strong> of infinite energy, or large <strong>and</strong> <strong>self</strong><strong>similar</strong>.<br />

This provides us with a general framework which includes in particular <strong>the</strong><br />

blowing up <strong>solutions</strong> of Shatah [15] <strong>and</strong> Bizoń [3]. As an application, we describe more<br />

precisely <strong>the</strong> regularity breakdown phenomenon.<br />

1 Introduction<br />

1.1 Presentation of <strong>the</strong> <strong>equation</strong><br />

We start with <strong>the</strong> <strong>wave</strong> <strong>map</strong> <strong>equation</strong> <strong>for</strong> a <strong>map</strong> u from <strong>the</strong> Minkowski space R 1+d to a<br />

general manifold N. Greek indices correspond to <strong>the</strong> Minkowski space, <strong>and</strong> <strong>the</strong>y are raised<br />

according to its metric. Latin indices correspond to <strong>the</strong> target manifold, whose Christoffel<br />

symbols are denoted by Γ. The Cauchy problem takes <strong>the</strong>n <strong>the</strong> <strong>for</strong>m<br />

(1)<br />

⎧<br />

⎨<br />

⎩<br />

∂ α ∂αu k = −Γ k i,j ∂α u i ∂αu j<br />

u(0) = u0<br />

ut(0) = u1 .<br />

The problem <strong>for</strong> small data has been studied in a series of articles, see in particular<br />

[19] [18] [8] [17] [20] ; without going into <strong>the</strong> details, it is globally well-posed <strong>for</strong> (u0, u1)<br />

small in ˙ H d/2 × ˙ H d/2−1 ; notice that <strong>the</strong>se <strong>spaces</strong> are at <strong>the</strong> scaling of <strong>the</strong> <strong>equation</strong>.<br />

1.2 Reduction of <strong>the</strong> problem<br />

We will focus on <strong>the</strong> case where <strong>the</strong> target manifold N is spherically symmetric,<br />

more precisely it is a rotationnally symmetric warped product manifold (we follow Shatah<br />

<strong>and</strong> Tahvilar-Zadeh [16]) defined as<br />

N = [0, φ ∗ ) ×G S d ,<br />

where φ ∗ ∈ R + ∪ {∞} <strong>and</strong> G : R → R is smooth <strong>and</strong> odd G(0) = 0, G ′ (0) = 1. On N we<br />

have <strong>the</strong> polar coordinates<br />

(φ, χ) ∈ [0, φ ⋆ ) × S d−1 ;<br />

1


in <strong>the</strong>se coordinates <strong>the</strong> metric reads<br />

dφ 2 + G 2 (φ)dχ 2 ,<br />

where dχ 2 corresponds to <strong>the</strong> st<strong>and</strong>ard metric of S d−1 .<br />

We fur<strong>the</strong>r assume that <strong>the</strong> data (<strong>and</strong> hence <strong>the</strong> solution) is covariant; in o<strong>the</strong>r<br />

words, if we adopt polar coordinates on R d <strong>and</strong> on N, u is of <strong>the</strong> <strong>for</strong>m<br />

u :R 1+d −→ N<br />

(t, r, ω) −→ (φ(t, r) , ω) ,<br />

<strong>for</strong> some function φ. The <strong>equation</strong> 1) becomes<br />

d − 1<br />

(2) φtt − φrr −<br />

r φr + 1<br />

f(φ) = 0<br />

r2 where f(φ) = (d − 1)G(φ)G ′ (φ).<br />

We per<strong>for</strong>m a final trans<strong>for</strong>mation by setting<br />

v = φ<br />

r<br />

so that <strong>the</strong> Cauchy problem we will consider reads<br />

(3)<br />

⎧<br />

⎨<br />

⎩<br />

vtt − vrr − n−1<br />

r vr = v 3 Z(rv)<br />

v(0) = v0<br />

vt(0) = v1 .<br />

where n = d + 2 <strong>and</strong> Z(x) = 1<br />

((d − 1)x − f(x)) is a smooth, even function. It will be<br />

x3 very helpful in <strong>the</strong> following to consider <strong>the</strong> above <strong>equation</strong> as a <strong>wave</strong> <strong>equation</strong> <strong>for</strong> a radial<br />

function, set in R1+n .<br />

Notice that <strong>the</strong> above problem is invariant by <strong>the</strong> trans<strong>for</strong>mation<br />

(4)<br />

(v0(x) , v1(x)) ↦→ (λv0(λx) , λ 2 v1(λx))<br />

v(t, x) ↦→ λv(λt, λx) ,<br />

<strong>for</strong> any λ > 0, so that scale invariant <strong>spaces</strong> <strong>for</strong> <strong>the</strong> data <strong>and</strong> solution are typically<br />

(v0, v1) ∈ ˙ H n<br />

2 −1 × ˙ H n<br />

2 −2<br />

1.3 Infinite energy <strong>and</strong> <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong><br />

Self-<strong>similar</strong> <strong>solutions</strong><br />

, v ∈ L ∞ (R, ˙ H n<br />

2 −1 ) .<br />

In order to explain ideas better, we discuss in this section <strong>the</strong> case of <strong>the</strong> sphere N = S d .<br />

Let us first consider <strong>the</strong> case of finite energy data<br />

φ0 ∈ ˙ H d<br />

2 , φ1 ∈ ˙ H d<br />

2 −1<br />

(notice that <strong>the</strong>se <strong>spaces</strong> are scale invariant). For small data (see <strong>the</strong> references in Section<br />

1.1), <strong>the</strong> <strong>equation</strong> is well-posed globally, <strong>and</strong> scattering occurs. For large data, we<br />

have to distinguish between d = 2 (<strong>the</strong> critical case) <strong>and</strong> d ≥ 3 (<strong>the</strong> super-critical case).<br />

2


• In dimension 2, it has been proved by Shatah <strong>and</strong> Tahvildar-Zadeh [16] that <strong>self</strong><strong>similar</strong>blow<br />

up cannot occur, ie blow up cannot occur along a profile of <strong>the</strong> type<br />

f , it has to be faster. Indeed, instances of blow-up of <strong>the</strong> <strong>for</strong>m f<br />

x<br />

T −t<br />

x<br />

(T −t)| log(T −t)| −1/2<br />

+ radiation were built up recently by Rodnianski <strong>and</strong> Sterbenz [13]; <strong>and</strong> Krieger,<br />

Schlag <strong>and</strong> Tataru [9] proved that ano<strong>the</strong>r possible blow-up regime is of <strong>the</strong> <strong>for</strong>m<br />

f<br />

x<br />

(T −t) s<br />

+ radiation, where s > 3<br />

2 .<br />

• In dimension 3 (or higher) <strong>the</strong> picture is very different, since examples of exact <strong>self</strong>-<br />

<strong>similar</strong> blow up (i.e. <strong>solutions</strong> of <strong>the</strong> <strong>for</strong>m f<br />

x<br />

T −t<br />

<br />

) are known, see Shatah [15],<br />

Shatah <strong>and</strong> Tahvildar-Zadeh [16] <strong>and</strong> Bizon [3]. Fur<strong>the</strong>rmore, numerical experiments<br />

[4] seem to indicate that <strong>self</strong>-<strong>similar</strong> blow-up is a generic situation.<br />

<br />

So consider in <strong>the</strong> case of dimension 3 a blowing-up <strong>self</strong>-<strong>similar</strong> solution φ , with<br />

φ ∈ ˙ H d/2 . It leaves <strong>the</strong> space<br />

L ∞ t<br />

˙H d<br />

2<br />

x<br />

x<br />

T −t<br />

as it approaches <strong>the</strong> blow-up time T . Fur<strong>the</strong>rmore, at time T , <strong>the</strong> trace of <strong>the</strong> solution<br />

looks like<br />

φ(T ) = c0 <strong>and</strong> ∂tφ(T ) = c1<br />

|x| ,<br />

so that<br />

(φ(T ) , ∂tφ(T )) does not belong to ˙ H d<br />

2 × ˙ H d<br />

2 −1 .<br />

The idea of <strong>the</strong> present article is <strong>the</strong> following observation: if one replaces ˙ H d<br />

2 <strong>and</strong> L∞ d<br />

H˙ 2<br />

by <strong>spaces</strong> that can accomodate <strong>self</strong>-<strong>similar</strong> data, respectively <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong>, <strong>the</strong>re<br />

is in some sense no blow-up any more. Such <strong>spaces</strong> will be provided by<br />

˙B d<br />

2<br />

2,∞ × ˙ B d<br />

2 −1<br />

2,∞ <strong>and</strong> L∞ d<br />

B˙ 2<br />

(see <strong>the</strong> next section <strong>for</strong> a definition of <strong>the</strong>se <strong>Besov</strong> <strong>spaces</strong>). If one works with this new<br />

functional framework, one has<br />

<br />

x<br />

φ ∈ L<br />

T − t<br />

∞<br />

<br />

R, ˙ B d<br />

2 −1<br />

<br />

2,∞<br />

<strong>and</strong><br />

2,∞<br />

<strong>for</strong> any time t, (φ(t) , ∂tφ(t)) ∈ ˙ B d<br />

2<br />

2,∞ .<br />

The above discussion shows that this new functional framework could be very interesting<br />

<strong>for</strong> <strong>the</strong> study of <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong>.<br />

Infinite energy <strong>solutions</strong><br />

Also notice that it is interesting <strong>and</strong> natural to study infinite energy <strong>solutions</strong> <strong>for</strong> <strong>the</strong><br />

following reasons<br />

• First, harmonic <strong>map</strong>s (that is, stationary <strong>wave</strong>-<strong>map</strong>s) of S 3 are of infinite energy.<br />

• Second, data or <strong>solutions</strong> of globally finite energy necessarily correspond to functions<br />

which at spatial infinity focus to a point; by allowing <strong>the</strong> energy to be infinite, we<br />

can study <strong>the</strong> global behaviour of much more general data.<br />

3


• Finally (<strong>and</strong> this is somehow <strong>the</strong> dual of <strong>the</strong> previous point), infinite energy data<br />

enable us to consider (covariant) <strong>wave</strong>-<strong>map</strong>s such that φ(0) is not necessarily <strong>the</strong><br />

north or <strong>the</strong> south pole of <strong>the</strong> sphere S d , so that a singularity occurs at 0.<br />

Plan of <strong>the</strong> article<br />

In Section 2, we will prove existence, uniqueness <strong>and</strong> scattering of <strong>solutions</strong> <strong>for</strong> data that<br />

are of infinite energy but small. We shall use a classical fixed-point argument.<br />

In Section 3, we will study <strong>the</strong> case of large data. However, since we cannot resort to<br />

perturbation techniques anymore, we will have to restrict to <strong>the</strong> case of <strong>self</strong>-<strong>similar</strong> data.<br />

Matters reduce <strong>the</strong>n to studying a (singular) ODE.<br />

In Section 4, we shall see that we can apply <strong>the</strong> results of <strong>the</strong> last two sections in order<br />

to describe a little more precisely <strong>the</strong> blow-up phenomenon <strong>for</strong> <strong>the</strong> <strong>wave</strong>-<strong>map</strong> <strong>equation</strong> on<br />

<strong>the</strong> sphere S 3 .<br />

1.4 Littlewood-Paley decomposition <strong>and</strong> associated <strong>spaces</strong><br />

We shall present in this section <strong>the</strong> main harmonic analysis tools that will be used in <strong>the</strong><br />

sequel.<br />

A Littlewood-Paley decomposition is defined in <strong>the</strong> following way: consider Ψ a function<br />

supported in <strong>the</strong> annulus centered in 0, of small radius 3/4, <strong>and</strong> big radius 8/3, such that<br />

<br />

<br />

ξ<br />

Ψ<br />

2j <br />

= 1 <strong>for</strong> ξ = 0 .<br />

j∈Z<br />

Then <strong>the</strong> Fourier multipliers ∆j <strong>and</strong> Sj are defined by<br />

<br />

D<br />

∆j = Ψ<br />

2j <br />

SN = 1 − <br />

<br />

D<br />

Ψ<br />

2j <br />

j≥N+1<br />

Thus any distribution can be decomposed (modulo polynomials) into a sum of elementary<br />

elements that are localized in frequency:<br />

f = <br />

∆jf or f = SNf + <br />

∆jf .<br />

j∈Z<br />

We can now define <strong>the</strong> (homogeneous) <strong>Besov</strong> <strong>spaces</strong> ˙ B s q,r<br />

f ˙ B s q,r<br />

def<br />

=<br />

<br />

<br />

2 js ∆jf q<br />

L <br />

x ℓr j<br />

.<br />

j≥N+1<br />

<br />

<br />

which are given by <strong>the</strong> norm<br />

Notice that <strong>the</strong> homogeneous L 2 -based Sobolev <strong>spaces</strong> identify with <strong>Besov</strong> <strong>spaces</strong><br />

˙H s = ˙ B s 2,2 .<br />

A variant of <strong>Besov</strong> <strong>spaces</strong> which are particularly well suited <strong>for</strong> <strong>the</strong> study of PDEs is given<br />

by <strong>the</strong> <strong>spaces</strong> L p (R, ˙ B s q,r ) - we will later omit <strong>the</strong> R in order to make <strong>the</strong> notations lighter.<br />

4<br />

.


These <strong>spaces</strong> were introduced by Chemin <strong>and</strong> Lerner [5] in <strong>the</strong> context of PDEs (notice<br />

that very <strong>similar</strong> <strong>spaces</strong> were considered earlier by Triebel [14] from a harmonic analysis<br />

point of view). The norm of <strong>the</strong>se <strong>spaces</strong> is defined as follows<br />

fe Lp (R, B˙ s )<br />

q,r<br />

def<br />

<br />

<br />

= 2 js ∆jf p<br />

L<br />

<br />

<br />

t Lq <br />

x ℓr j<br />

In Section 2, we will build up a solution to (3) in Chemin-Lerner <strong>spaces</strong>.<br />

Let us record <strong>the</strong> two following results, which extend classical <strong>the</strong>orems on <strong>Besov</strong> <strong>spaces</strong>.<br />

Theorem 1.1 (Sobolev embedding) Suppose that<br />

we have <strong>the</strong>n<br />

q ≥ Q <strong>and</strong> s − n<br />

q<br />

L p ˙ B S Q,r ↩→ L p ˙ B s q,r .<br />

The proof simply relies on Bernstein inequality.<br />

= S − n<br />

Q ,<br />

Proposition 1.1 Suppose s > 0 <strong>and</strong><br />

f = <br />

fk with Supp fk ⊂ B(0, C2 k ) .<br />

One has <strong>the</strong>n<br />

k<br />

f L p (R, ˙ B s q,r) ∼<br />

2 Small data in <strong>Besov</strong> <strong>spaces</strong><br />

<br />

<br />

2 js fjf L p<br />

<br />

<br />

t Lq <br />

x ℓr j<br />

2.1 Main result: existence, uniqueness, <strong>and</strong> scattering<br />

Well-posedness <strong>for</strong> <strong>the</strong> <strong>wave</strong>-<strong>map</strong> system is only known <strong>for</strong> data in Sobolev <strong>spaces</strong>. We<br />

want to extend this to <strong>Besov</strong> <strong>spaces</strong>, in <strong>the</strong> same spirit as Planchon [12] did <strong>for</strong> <strong>the</strong> <strong>wave</strong><br />

<strong>equation</strong> with a power non-linearity.<br />

Theorem 2.1 Let n ≥ 5 (i.e. d ≥ 3), <strong>and</strong> consider data (v0, v1) ∈ ˙ B n/2−1<br />

2,∞ × ˙ B n/2−2<br />

2,∞ radial<br />

<strong>and</strong> small enough. Then <strong>the</strong>re exists a global solution v of (3), which is unique in a ball<br />

of sufficiently small radius of<br />

(5) X def<br />

<br />

∞<br />

= L R, ˙B n<br />

2 −1<br />

<br />

∩ L 2<br />

<br />

<br />

∩ L 3<br />

<br />

<br />

.<br />

2,∞<br />

R, ˙B n<br />

2 −2<br />

( 1<br />

2<br />

− 3<br />

2n) −1 ,∞<br />

.<br />

.<br />

R, ˙B n<br />

2 −2<br />

( 1<br />

2<br />

− 4<br />

3n) −1 ,∞<br />

Fur<strong>the</strong>rmore, a weak scattering takes place, in <strong>the</strong> sense that <strong>the</strong>re exists (v + 0 , v+ 1 ) <strong>and</strong><br />

(v − 0 , v− 1 ) in ˙ B n/2−1<br />

2,∞ × ˙ B n/2−2<br />

2,∞ such that<br />

∀j ,<br />

<br />

+<br />

∆j W (t)(v 0 , v + 1 ) − v + 2 <br />

+<br />

∆j∂t W (t)(v 0 , v + 1 ) − v <br />

−→ 0 2<br />

∆j<br />

−<br />

W (t)(v0 , v − 1 ) − v + 2 <br />

∆j∂t<br />

−<br />

W (t)(v0 , v − 1 ) − v −→ 0 2<br />

as t → +∞ ,<br />

where we denote by W (t)(f, g) <strong>the</strong> free solution of <strong>the</strong> <strong>wave</strong> <strong>equation</strong> associated to <strong>the</strong><br />

initial data (f, g).<br />

In general, no classical scattering result <strong>for</strong> <strong>the</strong> norm of ˙ B n/2−1<br />

2,∞<br />

5<br />

× ˙ B n/2−2<br />

2,∞<br />

holds.


The proof of <strong>the</strong> above <strong>the</strong>orem will proceed in three steps: first, dealing with existence<br />

<strong>and</strong> uniqueness in <strong>the</strong> case Z = Id (Section 2.2), <strong>the</strong>n in <strong>the</strong> general case (Section 2.3),<br />

<strong>and</strong> finally proving <strong>the</strong> scattering result (Section 2.4).<br />

Let us make a few comments concerning this <strong>the</strong>orem.<br />

• Actually, we will prove <strong>the</strong> above <strong>the</strong>orem <strong>for</strong> any smooth function Z in (3) ; when<br />

<strong>the</strong> <strong>equation</strong> (3) is derived from <strong>the</strong> <strong>wave</strong>-<strong>map</strong> system however, we get a function Z<br />

which is even<br />

• The reason why we have to take such unpleasant exponents in <strong>the</strong> definition of X is<br />

that we need positive regularity indices so that Proposition 2.2 holds true.<br />

• How does <strong>the</strong> regularity of v0 <strong>and</strong> v1 translate <strong>for</strong> φ0 <strong>and</strong> φ1? It is easy to see that<br />

(v0, v1) ∈ ˙ B n/2−1<br />

2,∞ × ˙ B n/2−2<br />

2,∞ does not imply (φ0, φ1) ∈ ˙ B d/2<br />

2,∞ × ˙ B n/2−1<br />

2,∞ . On <strong>the</strong> o<strong>the</strong>r<br />

h<strong>and</strong>, it is proved in [16] that (v0, v1) ∈ ˙ Hn/2−1 × ˙ Hn/2−2 is equivalent to (φ0, φ1) ∈<br />

˙H d/2−1 × ˙H d/2−2 . As a conclusion, <strong>the</strong> data of <strong>the</strong> above <strong>the</strong>orem correspond to<br />

(φ0, φ1) in a space lying (strictly) between ˙ Hd/2−1 × ˙ Hd/2−2 <strong>and</strong> ˙ B d/2<br />

2,∞ × ˙ B n/2−1<br />

2,∞ .<br />

2.2 Proof of existence <strong>and</strong> uniqueness in <strong>the</strong> case Z = Id<br />

The <strong>equation</strong> under consideration is <strong>for</strong> now<br />

(6)<br />

⎧<br />

⎨<br />

⎩<br />

v = rv 4<br />

v(0) = v0<br />

vt(0) = v1 .<br />

where we denote <strong>for</strong> <strong>the</strong> <strong>wave</strong> operator = ∂ 2 t − ∆. The initial data (v0, v1) are radial<br />

<strong>and</strong> belong to ˙ B n/2−1<br />

2,∞<br />

× ˙ B n/2−2<br />

2,∞ .<br />

We will use a fixed point argument in <strong>the</strong> space<br />

X = <br />

∞<br />

L R, ˙ B n<br />

2 −1<br />

2,∞<br />

<br />

∩ L 2<br />

<br />

R, ˙ B n<br />

2 −2<br />

Strichartz estimates will be a crucial tool.<br />

( 1<br />

2<br />

− 3<br />

2n) −1 ,∞<br />

Lemma 2.1 (Strichartz estimate) The solution u of<br />

(7)<br />

satisfies <strong>the</strong> estimate<br />

uX ≤ C<br />

⎧<br />

⎨<br />

⎩<br />

<br />

u0 ˙B −1+ n 2<br />

2,∞<br />

u = F<br />

u(0) = u0<br />

ut(0) = u1 .<br />

<br />

+ u1 ˙B −2+ n 2<br />

2,∞<br />

∩ L 3<br />

<br />

R, ˙ B n<br />

2 −2<br />

( 1<br />

2<br />

+ F eL 1 ˙ B −2+ n 2<br />

2,∞<br />

− 4<br />

3n) −1 ,∞<br />

Proof of Lemma 2.1: We notice first that we can localize in frequency <strong>the</strong> <strong>equation</strong> (7).<br />

The classical energy estimate <strong>for</strong> <strong>the</strong> <strong>wave</strong> <strong>equation</strong> gives <strong>the</strong>n<br />

∆juL∞H˙ 1 ≤ C <br />

∆ju0H˙ 1 + ∆ju12 + ∆jF L1L2 .<br />

6<br />

<br />

.<br />

<br />

.


Using <strong>the</strong> frequency localization, we can rewrite this as<br />

2 j ∆juL∞L2 ≤ C 2 j <br />

∆ju02 + ∆ju12 + ∆jF L1L2 ,<br />

n<br />

j( <strong>and</strong> it suffices to multiply <strong>the</strong> above by 2 2 −2) <strong>and</strong> take <strong>the</strong> supremum over j to obtain<br />

u −1+<br />

eL ∞ ( B˙ n <br />

≤ C u0 2<br />

−1+<br />

2,∞ ) ˙B n 2<br />

2,∞<br />

+ u1 ˙B −2+ n 2<br />

2,∞<br />

+ F eL 1 ˙ B −2+ n 2<br />

2,∞<br />

Next, we would like to obtain <strong>the</strong> same estimate <strong>for</strong> <strong>the</strong> L2 n<br />

B˙ 2 −2<br />

( 1 3<br />

− 2 2n) −1 ,∞<br />

apply Strichartz’ inequality [7] to each one of <strong>the</strong> dyadic components<br />

∆ju<br />

L2L ( 1 2<br />

−<br />

3<br />

2n) −1 ≤ C 2 j <br />

∆ju02 + ∆ju12 + ∆jF L1L2 ,<br />

<br />

.<br />

norm. We just<br />

n<br />

j( <strong>and</strong> multiplying both sides by 2 2 −2 ), <strong>the</strong>n taking <strong>the</strong> supremum over j, one gets<br />

u<br />

eL 2 ˙B n 2 −2<br />

( 1 2<br />

−<br />

3<br />

2n) −1 ,∞<br />

≤ C<br />

<br />

u0 ˙B −1+ n 2<br />

2,∞<br />

+ u1 ˙B −2+ n 2<br />

2,∞<br />

It remains to show this inequality <strong>for</strong> <strong>the</strong> L3 n<br />

B˙ 2 −2<br />

( 1 4<br />

− 2 3n) −1 ,∞<br />

<strong>similar</strong> to <strong>the</strong> above one, <strong>and</strong> we <strong>the</strong>re<strong>for</strong>e omit it. <br />

+ F eL 1 ˙B −2+ n 2<br />

2,∞<br />

<br />

norm; but <strong>the</strong> proof is very<br />

The above lemma shows that in order to run a fixed point argument <strong>and</strong> solve (6) in <strong>the</strong><br />

space X <strong>for</strong> small data, it suffices to control rv 4 in L 1 ˙ B −2+ n<br />

2<br />

2,∞ .<br />

Proposition 2.1 There holds<br />

rv 4 eL 1 ˙ B −2+ n 2<br />

2,∞<br />

≤ Cv 4 X<br />

Proof of Proposition 2.1: First of all we observe that by Sobolev embedding (Theorem<br />

1.1),<br />

X ↩→ L ∞ ˙ B −1+ n<br />

2<br />

2,∞<br />

∩ L 2 ˙ B − 1<br />

2<br />

∞,∞ ∩ L 3 −<br />

B˙ 2<br />

3<br />

∞,∞ ,<br />

<strong>and</strong> we will be using in <strong>the</strong> proof of Proposition 2.1 only <strong>the</strong> norms appearing in <strong>the</strong> right<br />

h<strong>and</strong> side of <strong>the</strong> above embedding.<br />

We now split v 4 using <strong>the</strong> paraproduct algorithm introduced by Bony [1], that is we write<br />

v 4 = <br />

j∈Z<br />

(Sj+1v) 4 − (Sjv) 4 .<br />

Next we observe that <strong>the</strong> summ<strong>and</strong>s in <strong>the</strong> above <strong>equation</strong> are of <strong>the</strong> <strong>for</strong>m (∆jv)ABC,<br />

where A, B <strong>and</strong> C can be ei<strong>the</strong>r Sjv or ∆jv. The most difficult case is <strong>the</strong> one where<br />

A = B = C = Sjv, so we will try to estimate in <strong>the</strong> following<br />

r <br />

∆jv(Sjv) 3 ,<br />

j<br />

7<br />

.


that we rewrite as<br />

r <br />

∆jv(Sjv) 3 = <br />

j<br />

Estimate <strong>for</strong> I eL 1 ˙B −2+ n 2<br />

2,∞<br />

j<br />

Sj+100<br />

= I + II .<br />

We will use <strong>the</strong> following estimates:<br />

• Simply by definition of X,<br />

• By Lemma 5.1, we have<br />

<br />

r∆jv(Sjv) 3 + <br />

(1 − Sj+100) r∆jv(Sjv) 3<br />

n<br />

j(1−<br />

∆jvL∞ L2 ≤ CvX2 2 ) .<br />

rSjvL ∞ L ∞ ≤ Cv L ∞ ˙ B n 2 −1<br />

2,∞<br />

j<br />

≤ CvX .<br />

• Due to <strong>the</strong> negative regularity of L2 −<br />

B˙ 1<br />

2<br />

∞,∞, we can bound Sjv as follows<br />

SjvL2 L∞ ≤ <br />

<br />

∆kvL2 L∞ ≤ CvX<br />

k


• By definition of X,<br />

• With <strong>the</strong> same justification as above<br />

Combining <strong>the</strong>se two estimates, we get<br />

n<br />

1−<br />

∆jvL∞ L2 ≤ C2 2 vX .<br />

2<br />

j<br />

SjvL3 L∞ ≤ C2 3 vX .<br />

n<br />

j(3−<br />

(8) fjL1 L2 ≤ 2 2 ) v 4 X .<br />

Now to estimate ∆k(rfj) we use Plancherel <strong>the</strong>orem (in space) which yields<br />

<br />

<br />

∆k(rfj)2 = <br />

ψ <br />

ξ<br />

2k <br />

1<br />

|ξ − η| n+1 <br />

<br />

fj(η) dη<br />

.<br />

Notice that, due to <strong>the</strong> Fourier localization of <strong>the</strong> Littlewood-Paley operators, in <strong>the</strong> above<br />

expression, η ≤ C2j <strong>and</strong> ξ ∼ 2k . As a result, ξ − η is of order 2k <strong>and</strong> we don’t have to<br />

worry about <strong>the</strong> Fourier trans<strong>for</strong>m at 0 of |x|. We can bound <strong>the</strong> above by<br />

<br />

<br />

∆k(rfj)2 ≤ <br />

ψ <br />

ξ<br />

2k <br />

2<br />

<br />

<br />

<br />

<br />

<br />

|η|≤C2 j<br />

2<br />

1<br />

|ξ − η| n+1 <br />

<br />

<br />

fj(η) dη<br />

.<br />

<br />

∞<br />

We now use Hölder’s inequality <strong>and</strong> <strong>the</strong> fact that ξ − η ∼ 2 k to estimate <strong>the</strong> integral in<br />

<strong>the</strong> above expression <strong>and</strong> get<br />

n<br />

k<br />

∆k(rfj)2 ≤ C2 2 2 −k(n+1) n<br />

n<br />

j k(−1−<br />

2 2 fj2 = C2 2 ) n<br />

j<br />

2 2 fj2 .<br />

Integrating in time <strong>and</strong> using (8),<br />

It suffices now to sum over k to get<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

∆k(rfj) <br />

<br />

<strong>and</strong> Proposition 1.1 yields<br />

n<br />

k(−1−<br />

∆k(rfj)L1L2 ≤ C2 2 ) 2 3j v 4 X .<br />

k>j+100<br />

L 1 L 2<br />

II eL 1 ˙B −2+ n 2<br />

2,∞<br />

n<br />

j(2−<br />

≤ C2 2 ) v 4 X ,<br />

≤ Cv 4 X .<br />

This concludes <strong>the</strong> proof of Theorem 2.1 in <strong>the</strong> case where Z = Id.<br />

2.3 Proof of existence <strong>and</strong> uniqueness in <strong>the</strong> general case<br />

We want now to consider <strong>the</strong> non-linearity Z(rv)v 3 instead of rv 4 .<br />

We observe first that we can assume that Z vanishes at 0. If it is not <strong>the</strong> case, we split<br />

<strong>the</strong> non-linearity into two parts as follows<br />

Z(0)v 3 + (Z(rv) − Z(0)) v 3 .<br />

9


The cubic term is easy to h<strong>and</strong>le, so we will focus on <strong>the</strong> o<strong>the</strong>r one, (Z(rv) − Z(0)) v 3 ,<br />

which is <strong>the</strong> same as assuming that Z vanishes at 0. We fur<strong>the</strong>r notice that, in order to<br />

apply <strong>the</strong> results of Section 2.2, it would suffice to show that Z(rv)<br />

has <strong>the</strong> same regularity<br />

as v. This is <strong>the</strong> content of <strong>the</strong> next Proposition, which will <strong>the</strong>re<strong>for</strong>e complete <strong>the</strong> proof<br />

of Theorem 2.1.<br />

Proposition 2.2 Suppose that v is a radial function. We have <strong>the</strong>n<br />

<br />

<br />

<br />

Z(rv) <br />

<br />

≤ CvX .<br />

r<br />

X<br />

Proof of Proposition 2.2: The proof is classical, it uses <strong>the</strong> first linearization method<br />

of Meyer [11]. We write, using Taylor’s <strong>for</strong>mula<br />

Z(rv)<br />

r<br />

Z (rSj+1v) − Z (rSjv)<br />

=<br />

r<br />

j∈Z<br />

= <br />

∆jv<br />

j<br />

1<br />

0<br />

Z ′ (rSjv + sr∆jv) ds .<br />

We now observe that rv is bounded in L ∞ by Lemma 5.1, <strong>and</strong> that <strong>the</strong> regularity index<br />

of each of <strong>the</strong> <strong>spaces</strong> whose intersection is X has a positive regularity index. It <strong>the</strong>re<strong>for</strong>e<br />

suffices to follow <strong>the</strong> lines of [2] in order to prove <strong>the</strong> proposition. <br />

2.4 Proof of <strong>the</strong> scattering result<br />

We consider a solution v as built up in <strong>the</strong> preceding sections, <strong>and</strong> we only prove <strong>the</strong><br />

scattering as t → ∞, <strong>the</strong> o<strong>the</strong>r case being identical. Denoting F = Z(rv)v 3 , we have seen<br />

that F ∈ L1B˙ d/2−1<br />

2,∞ . Applying <strong>the</strong> frequency localization operator to <strong>the</strong> <strong>equation</strong> satisfied<br />

by v, we get<br />

∆jv = ∆jF ,<br />

2<br />

with ∆jF 2 ≤ C2<br />

j( n −2) . So by <strong>the</strong> classical energy scattering result, <strong>the</strong>re exists<br />

(v + 0,j , v+ 1,j ) such that<br />

<strong>and</strong><br />

It suffices to set<br />

<br />

<br />

∆jv − W (t)(v + 0,j , v+ 1,j )<br />

<br />

<br />

+ ∂t∆jv − ∂tW (t)(v<br />

2 + 0,j , v+ 1,j )<br />

<br />

<br />

−→ 0 ,<br />

2<br />

v + 0,j2 n<br />

j(1−<br />

≤ C2 2 ) , v + 1,j2 n<br />

j(2−<br />

≤ C2 2 ) .<br />

v + 0<br />

to get <strong>the</strong> weak scattering result.<br />

= <br />

j∈Z<br />

v + 0,j , v + 1<br />

= <br />

j∈Z<br />

v + 1,j<br />

Let us now prove that scattering <strong>for</strong> <strong>the</strong> norm of B˙ n/2−1<br />

2,∞<br />

general. To do so, we consider <strong>self</strong>-<strong>similar</strong> data<br />

v0(x) = c<br />

|x|<br />

, v1(x) = c<br />

,<br />

|x| 2<br />

10<br />

r<br />

× ˙ B n/2−2<br />

2,∞<br />

does not occur in


where c is small enough so that <strong>the</strong> existence <strong>and</strong> uniqueness result applies. By uniqueness,<br />

<strong>the</strong> solution v it<strong>self</strong> is <strong>self</strong>-<strong>similar</strong>, ie is left unchanged by <strong>the</strong> trans<strong>for</strong>mation (4). Let us<br />

argue by contradiction <strong>and</strong> suppose that <strong>the</strong>re exists (f0 , f1) ∈ ˙ B n<br />

2 −1<br />

2,∞ × ˙ B n<br />

2 −2<br />

2,∞ such that<br />

W (t)(f0 , f1) − v ˙B n 2 −1<br />

2,∞<br />

n<br />

= sup 2<br />

j( 2<br />

j<br />

−1) (∆jW (t)(f0 , f1) − v)2 −→ 0 as t → ∞ .<br />

It is easy to see that (f0 , f1) has to be scaling invariant as well, <strong>and</strong> a small computation<br />

yields <strong>the</strong>n<br />

∆j (v − W (f0 , f1)) (t, x) = 2 j ∆0 (v − W (f0 , f1)) (2 j t , 2 j x) .<br />

This implies that<br />

d<br />

sup 2<br />

j( 2<br />

j<br />

−1) <br />

∆j (W (t)(f0 , f1) − v) (t)2 = sup ∆0 (W (f0 , f1) − v) (2<br />

j<br />

j t) , 2<br />

<strong>and</strong> we see that this last quantity goes to zero as t goes to infinity only if u = W (f0 , f1),<br />

which implies u = 0.<br />

3 Large <strong>self</strong>-<strong>similar</strong> data<br />

3.1 Main result<br />

For <strong>the</strong> sake of simplicity, we will restrict our investigations in this section to a classical<br />

case: in <strong>the</strong> following, we take d = 3 (hence n = 5) <strong>and</strong> N = S 3 ; already a great<br />

wealth of behaviours can be observed. In this case, G(φ) = sin(φ), <strong>and</strong> <strong>equation</strong> (2)<br />

becomes<br />

(9) φtt − φrr − 2<br />

r φr + 1<br />

sin(2φ) = 0 .<br />

r2 We want to consider <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong>; since we always deal with corotational <strong>solutions</strong>,<br />

<strong>the</strong> data must have <strong>the</strong> following <strong>for</strong>m<br />

φ0 = c0 , φ1 = c1<br />

|x| .<br />

If <strong>for</strong> instance uniqueness holds, it is easy to see that <strong>the</strong> solution φ is it<strong>self</strong> <strong>self</strong>-<strong>similar</strong>,<br />

i.e. given by a profile f,<br />

<br />

x<br />

<br />

φ(x, t) = f .<br />

t<br />

We will look <strong>for</strong> a solution under that <strong>for</strong>m, so that <strong>the</strong> problem reduces to finding f.<br />

The interest of <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong> is that <strong>the</strong>y provide, if <strong>the</strong> profile f is regular, an<br />

example of blow up of (9) at t = 0. The first example of such a blowup was given by<br />

Shatah [15]: <strong>for</strong> c0 = 0, c1 = 2, <strong>the</strong> solution corresponding to <strong>the</strong> profile<br />

(10) f0(ρ) = 2 tan −1 (ρ)<br />

solves (9). Numerical simulations of Bizoń, Chmaj <strong>and</strong> Tabor [4] suggest that f0 is a<br />

generic blowup profile.<br />

Bizoń [3] showed that f0 was actually <strong>the</strong> “lowest-energy” member of a family of regular<br />

profiles (fn)n∈N. The profiles fn <strong>for</strong> n > 1 should be less stable than f0.<br />

11


Finally, it was proved in [16] that <strong>for</strong> given data <strong>the</strong> profile may not be unique. Their<br />

example was <strong>the</strong> following<br />

¯f0(ρ) =<br />

π<br />

2<br />

<strong>for</strong> ρ < 1<br />

f0(ρ) <strong>for</strong> rho > 1<br />

It is easily seen that <strong>the</strong> <strong>solutions</strong> corresponding to <strong>the</strong>se data share <strong>the</strong> same initial data.<br />

The initial data considered by Bizoń <strong>and</strong> Shatah <strong>for</strong>m a countable set, namely only those<br />

data that yield regular profiles; in <strong>the</strong> following <strong>the</strong>orem, we consider any initial data.<br />

We now switch to <strong>the</strong> v unknown function <strong>and</strong> observe that <strong>the</strong> Cauchy problem becomes<br />

(11)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

vtt − vrr − n−1<br />

r vr = v 3 Z(rv)<br />

v(0) = c0<br />

|x|<br />

vt(0) = c1<br />

|x| 2 .<br />

What regularity should be expected <strong>for</strong> <strong>the</strong> <strong>self</strong>-<strong>similar</strong> <strong>solutions</strong> of (11)? For c0 <strong>and</strong> c1<br />

small, we know that <strong>the</strong> solution v = φ<br />

r<br />

in (5) <strong>and</strong> which reads <strong>for</strong> n = 5<br />

built up in Theorem 2.1 belongs to X defined<br />

X = <br />

∞<br />

L R, ˙ B 3/2<br />

<br />

2,∞ ∩ <br />

2<br />

L R, ˙ B 1/2<br />

<br />

5,∞ ∩ <br />

3<br />

L R, ˙ B 1/2<br />

<br />

30/7,∞ .<br />

How does it translate <strong>for</strong> <strong>the</strong> profile f? By an argument given <strong>for</strong> instance in [12], v<br />

belongs to X if <strong>and</strong> only if, when considered as a function over R 5 .<br />

with<br />

f(ρ)<br />

ρ<br />

∈ Y<br />

Y def<br />

= ˙ B 3/2<br />

2,∞ (R5 ) ∩ ˙ B 1/2<br />

5,2 (R5 ) ∩ ˙ B 1/2<br />

30/7,3 .<br />

For c0 <strong>and</strong> c1 large, <strong>the</strong> <strong>the</strong>orem which follows tells us that we can always build up <strong>self</strong>-<br />

<strong>similar</strong> <strong>solutions</strong> such that v ∈ X or equivalently f<br />

ρ<br />

∈ Y .<br />

Theorem 3.1 (i) For any real numbers c0 <strong>and</strong> c1, <strong>the</strong>re exists a solution v of (11) given<br />

by a profile f.<br />

(ii) If one denotes v = φ<br />

r , one has<br />

v ∈ X <strong>and</strong><br />

(as explained above, both functions are here considered as functions on R 1+5 ).<br />

(iii) v or f are unique outside of <strong>the</strong> ball of radius one, but uniqueness <strong>for</strong> v ∈ X or f ∈ Y<br />

does not hold in general.<br />

(iv) However, if v ∈ X or f ∈ Y , one has f(0) = k π<br />

2 , with k an integer.<br />

So in contrast with <strong>the</strong> result of Theorem 2.1, which dealt with small data, uniqueness<br />

does not hold anymore <strong>for</strong> large data.<br />

The proof of <strong>the</strong> above <strong>the</strong>orem will be given in <strong>the</strong> following subsections 3.2, 3.3, 3.4,<br />

<strong>and</strong> 3.5.<br />

12<br />

f<br />

ρ<br />

∈ Y<br />

.


3.2 The <strong>equation</strong> satisfied by <strong>the</strong> profile f<br />

Due to <strong>the</strong> <strong>self</strong>-<strong>similar</strong>ity assumption, <strong>the</strong> <strong>equation</strong> satisfied by f reduces to an ordinary<br />

differential <strong>equation</strong>. It is given by <strong>the</strong> following proposition.<br />

Proposition 3.1 Suppose that φ is given by φ(x, t) = f( |x|<br />

t ) (we consider here <strong>the</strong> radial<br />

function f as a function of one real variable). Then<br />

(i) Suppose f ′ is locally integrable, <strong>and</strong> f, f ′ <strong>and</strong> ρf ′′ (ρ) are continuous at 0. Then v<br />

solves (11) if <strong>and</strong> only if f solves<br />

⎧<br />

⎨ ρ<br />

(12)<br />

⎩<br />

2 (1 − ρ2 )f ′′ (ρ) + 2ρ(1 − ρ2 )f ′ (ρ) − sin(2f(ρ)) = 0 in S ′ (0, ∞)<br />

limρ→∞ f(ρ) = c0<br />

limρ→∞ ρ 2 f ′ (ρ) = −c1 .<br />

(ii) The first <strong>equation</strong> of (12) is satisfied if <strong>and</strong> only if it holds true in <strong>the</strong> classical sense<br />

on (0, 1)∪(1, ∞), <strong>and</strong> f(1−) = f(1+) (<strong>the</strong>se limit values make sense because if <strong>the</strong> <strong>equation</strong><br />

is satisfied on (0, 1) ∪ (1, ∞), f ′ is locally integrable on (0, 1) ∪ (1, ∞)).<br />

Proof of Proposition 3.1: We begin by proving (i). It is first very easy to check that<br />

<strong>the</strong> initial conditions of (11) <strong>and</strong> <strong>the</strong> boundary conditions of (12) are equivalent.<br />

Next, we will work with <strong>the</strong> φ <strong>equation</strong>, which is easier to h<strong>and</strong>le in this case. It it <strong>the</strong>n<br />

easy to see that v solves (11). By definition, φ is a solution of (9) if <strong>and</strong> only if <strong>for</strong> any<br />

test function ψ ∈ D([0, ∞) × [0, ∞)),<br />

(13)<br />

∞ ∞ <br />

−ψtφt + ψrφr + 1<br />

<br />

sin(2φ)<br />

r2 0<br />

0<br />

ɛ<br />

0<br />

r 2 ∞<br />

dr dt =<br />

0<br />

ψ(r, 0) c1<br />

r r2 dr<br />

<strong>and</strong> <strong>the</strong> initial conditions are verified. It is very easy to check that <strong>the</strong> initial conditions<br />

of (9) <strong>and</strong> <strong>the</strong> boundary conditions of (12) are equivalent. So we will focus on <strong>the</strong> <strong>equation</strong><br />

above, <strong>and</strong> prove that it is equivalent to <strong>the</strong> first <strong>equation</strong> of (12). We begin by looking at<br />

∞ ∞<br />

− ψtφtr<br />

0 0<br />

2 ∞ ∞ r<br />

dr dt =<br />

0 0 t2 f ′ r<br />

<br />

ψtr<br />

t<br />

2 dr dt<br />

∞ ∞ r<br />

=<br />

t2 f ′ r<br />

<br />

ψtr<br />

t<br />

2 dr dt + O(ɛ) ,<br />

where <strong>the</strong> 0(ɛ) comes from <strong>the</strong> finite limit of ρ2f ′ (ρ) at ∞. Integrating by parts in t, we<br />

find<br />

∞ ∞<br />

− ψtφtr<br />

0 0<br />

2 dr dt<br />

∞ r<br />

= O(ɛ) −<br />

ɛ2 f ′ r<br />

<br />

ψr<br />

ɛ<br />

2 ∞ ∞ r<br />

dr dt + 2<br />

t3 f ′ r<br />

<br />

ψr<br />

t<br />

2 ∞ ∞ r<br />

dr dt +<br />

2<br />

t4 f ′′ r<br />

<br />

ψr<br />

t<br />

2 dr dt .<br />

0<br />

0<br />

ɛ<br />

0<br />

Integrating by parts in r <strong>the</strong> last term of <strong>the</strong> right-h<strong>and</strong> side, we get, using <strong>the</strong> fact that<br />

f ′ has a finite limit at 0,<br />

∞ ∞<br />

− ψtφtr<br />

0 0<br />

2 dr dt<br />

∞ r<br />

= O(ɛ) −<br />

ɛ2 f ′ r<br />

<br />

ψr<br />

ɛ<br />

2 ∞ ∞ r<br />

dr dt − 2<br />

t3 f ′ r<br />

<br />

ψr<br />

t<br />

2 ∞ ∞ r<br />

dr dt −<br />

2<br />

t3 f ′ r<br />

<br />

ψrr<br />

t<br />

2 dr dt .<br />

ɛ<br />

0<br />

13<br />

ɛ<br />

ɛ<br />

0<br />

0


We now per<strong>for</strong>m <strong>the</strong> change of variable ρ = r<br />

above equality, <strong>the</strong> left-h<strong>and</strong> side of (13) becomes<br />

∞ ∞ <br />

−ψtφt + ψrφr +<br />

0 0<br />

1<br />

<br />

sin(2φ) r<br />

r2 2 dr dt<br />

∞ r<br />

= O(ɛ) −<br />

ɛ2 f ′ r<br />

<br />

ψr<br />

ɛ<br />

2 ∞ ∞<br />

dr dt +<br />

0<br />

ɛ<br />

0<br />

t , so that in particular ψr = 1<br />

t ψρ. Using <strong>the</strong><br />

−2ρ 3 f ′ ψ + ρ 2 (1 − ρ 2 )f ′ ψρ + ψ sin(2f) dρdt .<br />

Obviously, <strong>the</strong> limit as ɛ goes to 0 of <strong>the</strong> two first terms of <strong>the</strong> right h<strong>and</strong> side is<br />

∞<br />

0<br />

ψ(r, 0) c1<br />

r r2 dr. So (13) holds true if <strong>and</strong> only if <strong>for</strong> any t<br />

∞<br />

This is equivalent to<br />

0<br />

−2ρ 3 f ′ ψ + ρ 2 (1 − ρ 2 )f ′ ψρ + ψ sin(2f) dρ = 0 .<br />

ρ 2 (1 − ρ 2 )f ′′ (ρ) + 2ρ(1 − ρ 2 )f ′ (ρ) − sin(2f(ρ)) = 0 in S ′ (0, ∞)<br />

Let us now prove (ii). It is convenient to switch to <strong>the</strong> function<br />

<br />

1<br />

g(z) = f ,<br />

z<br />

<strong>and</strong> one can check that <strong>the</strong> first <strong>equation</strong> of (12) is equivalent to<br />

(z 2 − 1)g ′′ (z) = sin(2g(z)) .<br />

Let us assume that this <strong>equation</strong> is satisfied classically on (0, 1)∪(1, ∞), <strong>and</strong> that g(1−) =<br />

g(1+). We will show that it is <strong>the</strong>n satisfied in S ′ . Of course, <strong>the</strong> only problematic point<br />

is 1, so we will try to describe more precisely <strong>the</strong> behaviour of g near 1. First of all, we<br />

observe that g ′ is integrable on (0, 1)∪(1, ∞), so that g is continuous. We have fur<strong>the</strong>rmore<br />

<strong>the</strong> following bound<br />

(14) |g ′ (z)| ≤ C| log |z − 1||<br />

<strong>for</strong> z = 1; g ′ could a priori have a δ singularity at 1, but this is excluded by <strong>the</strong> continuity<br />

of g.<br />

Now we would like to show that<br />

∞ 2 ′′<br />

(z − 1)g (z) − sin(2g(z)) ψ(z) dz = 0 ,<br />

0<br />

if ψ ∈ D((0, ∞)). We can cut <strong>the</strong> integral into three pieces,<br />

∞<br />

=<br />

0<br />

1−ɛ<br />

0<br />

+<br />

1+ɛ<br />

1−ɛ<br />

∞<br />

+ .<br />

The first <strong>and</strong> third pieces are 0 because <strong>the</strong> <strong>equation</strong> is satisfied in <strong>the</strong> classical sense, <strong>and</strong><br />

<strong>for</strong> <strong>the</strong> second we have, integrating by parts,<br />

1+ɛ<br />

1−ɛ<br />

(z 2 − 1)g ′′ (z) − sin(2g(z)) ψ(z) dz<br />

= O(ɛ) + g ′ (z)(z 2 − 1)ψ(z) 1+ɛ<br />

1−ɛ −<br />

1+ɛ<br />

g<br />

1−ɛ<br />

′ (z) (z 2 − 1)ψ(z) ′<br />

dz .<br />

Due to (14), we see that <strong>the</strong> right-h<strong>and</strong> side goes to zero as ɛ goes to 0. <br />

14<br />

1+ɛ


3.3 Solving <strong>the</strong> <strong>equation</strong> <strong>for</strong> f<br />

We will here show <strong>the</strong> existence of <strong>solutions</strong> <strong>for</strong> <strong>the</strong> system (12).<br />

Proposition 3.2 For any real numbers c0 <strong>and</strong> c1, <strong>the</strong> system (12) has a solution f ∈ L ∞ .<br />

It is unique on [1, ∞), but not in general on [0, 1].<br />

Remark 3.1 As stated in <strong>the</strong> above proposition, <strong>solutions</strong> to <strong>the</strong> first <strong>equation</strong> of (12)<br />

ρ 2 (1 − ρ 2 )f ′′ (ρ) + 2ρ(1 − ρ 2 )f ′ (ρ) − sin(2f(ρ)) = 0 ,<br />

(with prescribed data at infinity) are not unique in S ′ . Never<strong>the</strong>less, one can prove that<br />

<strong>solutions</strong> to <strong>the</strong> <strong>equation</strong><br />

f ′′ (ρ) + 2<br />

ρ f ′ (ρ) − sin(2f(ρ))<br />

ρ2 (1 − ρ2 = 0 ,<br />

)<br />

considered in ([16]) <strong>and</strong> ([3]) are unique in S ′ . Notice that <strong>for</strong>mally, <strong>the</strong> second <strong>equation</strong><br />

is nothing but <strong>the</strong> first divided by ρ 2 (1 − ρ 2 ).<br />

Proof of proposition 3.2: The problem (12) becomes, <strong>for</strong> <strong>the</strong> new unknown function<br />

g(z) = f <br />

1<br />

z<br />

⎧<br />

⎨<br />

⎩<br />

(z 2 − 1)g ′′ (z) = sin(2g(z))<br />

g(0) = c0<br />

g ′ (0) = −c1<br />

It is clear that this problem admits a unique solution on [0, 1], which translates into a<br />

unique solution <strong>for</strong> (12) on [1, ∞). Notice that g ′ is locally integrable, so that g(1−) =<br />

f(1+) is well defined.<br />

By proposition 3.1, it suffices, in order to get a global solution of<br />

(15) ρ 2 (1 − ρ 2 )f ′′ (ρ) + 2ρ(1 − ρ 2 )f ′ (ρ) − sin(2f(ρ)) = 0 ,<br />

(with <strong>the</strong> given prescribed data at infinity) to solve this <strong>equation</strong> on [0, 1], in such a way<br />

that f(1−) = f(1+), <strong>and</strong> fur<strong>the</strong>rmore f, f ′ , ρf ′′ have finite limits at 0.<br />

We begin with <strong>the</strong> following lemma<br />

Lemma 3.1 For any α ∈ R, <strong>the</strong> Cauchy problem<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

f ′′ + 2<br />

ρf ′ − sin(2f)<br />

ρ2 (1−ρ2 = 0 )<br />

f(0) = 0<br />

f ′ (0) = α<br />

admits a solution on [0, 1). Fur<strong>the</strong>rmore, it depends continuously on α.<br />

Proof: We only prove <strong>the</strong> existence part of <strong>the</strong> lemma; <strong>the</strong> continuous dependence on<br />

<strong>the</strong> data can be h<strong>and</strong>led in a <strong>similar</strong> fashion. We set f = H + αx, so that <strong>the</strong> problem<br />

becomes ⎧ ⎪⎨<br />

⎪⎩<br />

H ′′ + 2<br />

ρH′ + 2α sin(2(H+αρ))<br />

ρ − ρ2 (1−ρ2 = 0 )<br />

H(0) = 0<br />

H ′ (0) = 0<br />

15


Setting h = H ′ , <strong>the</strong> problem becomes<br />

<br />

h = 1<br />

ρ2 ρ<br />

0<br />

H = ρ<br />

0 h<br />

We use <strong>the</strong> following iteration scheme<br />

<br />

hn+1 = 1<br />

ρ2 ρ<br />

0<br />

Hn+1 = ρ<br />

0 hn ,<br />

sin(2αρ+2H)<br />

1−ρ 2<br />

sin(2αρ+2H n )<br />

1−ρ 2<br />

<br />

− 2αρ dρ<br />

<br />

− 2αρ dρ<br />

<strong>and</strong> we set h 0 = H 0 = 0. It is easy to see that, on a small enough interval,<br />

|h 1 (ρ)| ≤ Cρ 2<br />

(<strong>and</strong> H 1 = 0) .<br />

The difference of two consecutive iterates is given by<br />

<br />

hn+1 − hn = 1<br />

ρ2 ρ<br />

0<br />

Hn+1 − Hn = ρ <br />

0 hn − hn−1 .<br />

sin(2αρ+2H n )−sin(2αρ+2H n−1 )<br />

1−ρ 2<br />

On a small enough (but fixed) interval, this implies that<br />

<br />

|Hn − Hn−1 | ≤ Cρ3 |hn − hn−1 | ≤ Cρ2 <br />

|Hn+1 − Hn | ≤ δCρ3 =⇒<br />

|hn+1 − hn | ≤ δCρ2 where δ < 1. This implies that <strong>the</strong> iteration scheme converges on a small interval [0, ɛ],<br />

with ɛ > 0. It is <strong>the</strong>n straight<strong>for</strong>ward to extend <strong>the</strong> solution to [0, 1]. <br />

With <strong>the</strong> help of this lemma, it is easy to complete <strong>the</strong> proof of <strong>the</strong> proposition. It suffices<br />

to show that, by choosing α (as in <strong>the</strong> statement of <strong>the</strong> lemma) in an appropriate way,<br />

one can reach any given value f(1−). But we know two particular <strong>solutions</strong>, <strong>the</strong> solution<br />

corresponding to α = 0 <strong>for</strong> which f(1) = 0, <strong>and</strong> f0 (defined in (10)) <strong>for</strong> which α = 2,<br />

f(1) = π . The continuous dependence on α of <strong>the</strong> <strong>solutions</strong> built up in <strong>the</strong> above lemma<br />

2<br />

tells us that any real number between 0 <strong>and</strong> π<br />

2<br />

<br />

dρ<br />

can be reached by f(1−).<br />

Then <strong>the</strong> symmetries of <strong>the</strong> <strong>equation</strong> (<strong>for</strong> instance, if f is a solution, so is π−f <strong>and</strong> 2π+f)<br />

enable to reach any desired f(1) with f(0) = kπ <strong>and</strong> f ′ (0) judiciously chosen. <br />

3.4 Regularity questions<br />

In this section, we complete <strong>the</strong> proof of Theorem 3.1 by examining <strong>the</strong> regularity of <strong>the</strong><br />

<strong>solutions</strong>.<br />

Proposition 3.3 Any of <strong>the</strong> <strong>solutions</strong> v built up in Proposition 3.2 satisfies<br />

v ∈ X which is equivalent to<br />

f<br />

ρ<br />

∈ Y .<br />

(here f is considered as a function on R 5 ; <strong>for</strong> <strong>the</strong> definitions of X <strong>and</strong> Y , see section 3.1).<br />

Proof: We will prove that f<br />

ρ belongs to Y . The space Y is <strong>the</strong> intersection of three <strong>Besov</strong><br />

<strong>spaces</strong>; <strong>the</strong> hardest to deal with is <strong>the</strong> first one, B˙ 3/2<br />

, so we will focus on it, <strong>the</strong> proof <strong>for</strong><br />

<strong>the</strong> two o<strong>the</strong>r ones being <strong>similar</strong>.<br />

The function f<br />

ρ (seen as a function on R5 ) may be singular at three points: 0, 1 <strong>and</strong> ∞,<br />

it is C ∞ elsewhere, so we will examine separately <strong>the</strong>se three points.<br />

16<br />

2,∞<br />

,


• Let us begin with <strong>the</strong> behaviour at infinity, that is we look at χf with χ a smooth<br />

function equal to 0 on B(0, 2) <strong>and</strong> 1 outside of B(0, 3).<br />

It is first of all clear that χ c0<br />

ρ ∈ ˙ B 3/2<br />

f−c0<br />

2,∞ ; so we are left with χ ρ . Using <strong>the</strong> <strong>equation</strong><br />

satisfied by f <strong>and</strong> <strong>the</strong> data at ∞, we see that<br />

<br />

<br />

<br />

′′ <br />

<br />

f − c0 <br />

χ ≤<br />

ρ<br />

C<br />

.<br />

〈ρ〉 4<br />

In order to show that χ f−c0<br />

, it suffices to see that <strong>the</strong> second<br />

derivative belongs to ˙ B −1/2<br />

2,∞ . Using <strong>the</strong> above bound along with Young’s inequality,<br />

we get<br />

<br />

<br />

<br />

∆j ′′ <br />

f − c0 2 <br />

χ = <br />

ρ 25j ′′ <br />

ψ(2 j f − c0 2<br />

·) ⋆ χ<br />

ρ<br />

<br />

<br />

≤ 2 5j <br />

ψ(2 j <br />

·) <br />

χ<br />

ρ belongs to ˙ B 3/2<br />

2,∞<br />

10/9<br />

′′ <br />

f − c0<br />

5/3<br />

≤ C2<br />

ρ<br />

j/2 .<br />

• We will now examine <strong>the</strong> function in a neighborhood of <strong>the</strong> point 0.<br />

Consider first <strong>the</strong> profile identically equal to kπ<br />

2 . Then it is straight<strong>for</strong>ward to see<br />

that f<br />

ρ<br />

= kπ<br />

2ρ belongs to ˙ B 3/2<br />

2,∞ (R5 ).<br />

We suppose now f(0) = 0 <strong>and</strong> look at one of <strong>the</strong> <strong>solutions</strong> built up in Lemma 3.1.<br />

An examination of <strong>the</strong> <strong>equation</strong> satisfied by f implies that, close to 0,<br />

f(ρ) ∼ αρ , f ′ (ρ) ∼ αρ <strong>and</strong> f ′′ (ρ) ∼ − α<br />

ρ .<br />

We localize <strong>the</strong> function around 0 by considering φ f<br />

ρ , with φ a smooth function equal<br />

to 1 on B(0, 1/2) <strong>and</strong> 0 outside of B(0, 3/4). As in <strong>the</strong> last point, it suffices to see<br />

that<br />

′′<br />

f<br />

φ ∈ L<br />

ρ<br />

5/3 ,<br />

<strong>and</strong> it is easy to see that this holds, since, by <strong>the</strong> estimates above on <strong>the</strong> behaviour<br />

of f close to 0, we have f<br />

ρ ∈ L5/3 .<br />

• We finally investigate <strong>the</strong> situation at 1. Since f is radial, it belongs around ρ = 1<br />

(as a function on R5 ) to ˙ B 3/2<br />

2,∞ if <strong>and</strong> only if it does so as a function on R. So we have<br />

reduced <strong>the</strong> question to a one-dimensional problem. To simplify matters a little bit<br />

more, we will work with <strong>the</strong> function g(z) = f <br />

1<br />

z ; its regularity at 1 is obviously<br />

<strong>the</strong> same as <strong>the</strong> regularity of f, <strong>and</strong> it satisfies <strong>the</strong> simple <strong>equation</strong><br />

(16) (z 2 − 1)g ′′ (z) = sin(2g(z)) .<br />

This implies at once that g ′ behaves like sin(2g(1)| log |z − 1|| around 1, <strong>and</strong> integrating<br />

one more time we see that<br />

Coming back to (16), we see that<br />

g(z) = g(1) + O(|z − 1| log |z − 1|) .<br />

g ′′ (z) = sin(2g(1)<br />

z 2 − 1<br />

17<br />

+ O(log |z − 1|) .


So that, integrating one last time we get <strong>the</strong> desired description of <strong>the</strong> singularity<br />

at 1<br />

g ′ (1 + δ) = sin(2g(1)) log |δ| + Csign δ + φ ,<br />

where C± are constants, <strong>and</strong> φ is a continuous function around 1, that we will ignore<br />

in <strong>the</strong> following. So (shifting <strong>the</strong> singularity to 0, we are now in one dimension) <strong>the</strong><br />

problem reduces to showing that log |z| <strong>and</strong> <strong>the</strong> Heaviside function H(z) belong to<br />

˙B 1/2<br />

1<br />

2,∞ (R). Differentiating, this is equivalent to <strong>the</strong> fact that z <strong>and</strong> <strong>the</strong> Dirac function<br />

δ belong to ˙ B −1/2<br />

2,∞ (R).<br />

But it is well known that <strong>the</strong> Fourier trans<strong>for</strong>m of <strong>the</strong>se two functions are, respec-<br />

tively, sign(ξ) <strong>and</strong> 1. It follows that<br />

1<br />

∆j<br />

z 2<br />

<br />

<br />

= <br />

ψ <br />

ξ<br />

2j <br />

<br />

sign(ξ) ∼ 2 j/2 <strong>and</strong><br />

<br />

<br />

∆jδ2 = <br />

ψ 2<br />

<strong>and</strong> this completes <strong>the</strong> study of <strong>the</strong> singularity of f at |ρ| = 1. <br />

3.5 Uniqueness issues<br />

<br />

ξ<br />

2j <br />

2<br />

∼ 2 j/2 ,<br />

That <strong>the</strong> solution is unique outside <strong>the</strong> ball of radius one is obvious. The first known<br />

example of non-uniqueness consisted of <strong>the</strong> function ¯ f0 of Shatah, defined in Section 3.1.<br />

As proved in <strong>the</strong> previous section, it belongs to X.<br />

Due to <strong>the</strong> second point of Proposition 3.1, it very easy to build up new examples of<br />

non-unique <strong>solutions</strong>. The profile is imposed <strong>for</strong> ρ ∈ (1, ∞), this also imposes f(1+).<br />

Then we can complete it by any of <strong>the</strong> <strong>solutions</strong> on (0, 1) of Lemma 3.1, provided that<br />

f(1−) = f(1+). That <strong>the</strong> resulting solution belongs to X has been proved in <strong>the</strong> previous<br />

section.<br />

As a last result, we would like to prove that, though no uniqueness holds on (0, 1), not<br />

any value can be taken at 0 by <strong>the</strong> profile.<br />

If v = f <br />

r<br />

t belongs to X, <strong>the</strong>n, by Lemma 5.1, f ∈ L<br />

r<br />

∞ . We have <strong>the</strong>n <strong>the</strong> following<br />

lemma.<br />

Lemma 3.2 Suppose f solves (15) on (0, 1)) <strong>and</strong> is bounded in L∞ . Then f is continuous<br />

at 0, <strong>and</strong> f(0) = k π<br />

2 , with k an integer.<br />

If k is odd, <strong>the</strong>n f is constant on (0, 1).<br />

Proof: Take f as in <strong>the</strong> lemma, <strong>and</strong> let<br />

A small computation shows that<br />

M(ρ) = 1<br />

2 (ρ2 − ρ 4 )f ′2 + cos 2 f .<br />

M ′ = −f ′2 ρ ,<br />

so that M is a decreasing quantity. If M is larger than 1 <strong>for</strong> some ɛ > 0, <strong>the</strong>n it must be<br />

larger than 1 on (0, ɛ), <strong>and</strong> this implies that f blows up at 0. So M is less than one on<br />

(0, 1/2), let us denote M0 <strong>for</strong> its limit at 0. We will show that f also has a limit at 0.<br />

We observe that <strong>the</strong> integral of M ′ is finite, this implies that<br />

(17) ρf ′2 ∈ L 1<br />

<br />

0, 1<br />

<br />

;<br />

2<br />

18


also <strong>the</strong> boundedness of M implies that<br />

(18) |f ′ (ρ)| ≤ C<br />

ρ .<br />

Fix ɛ > 0, we can suppose that ρ is taken close enough to 0 so that<br />

|M(ρ) − M0| ≤ ɛ .<br />

Also, we must have cos 2 f(ρ n ) − M(ρ n ) ≤ ɛ<br />

<strong>for</strong> a sequence ρ n going to 0, due to (17). We argue by contradiction <strong>and</strong> assume (possibly<br />

considering a subsequence of (ρ n )) that <strong>for</strong> a sequence τ n going to zero, with ρ n+1 ≤ τ n ≤<br />

ρ n , we have<br />

<br />

cos 2 f(τ n ) − M(τ n ) ≥ 2ɛ <strong>and</strong> hence<br />

n<br />

In<br />

n<br />

1<br />

2 (ρ2 − ρ 4 )f ′2 ≥ ɛ .<br />

Due to (18), f ′ remains larger than ɛ on an interval In centered at τn <strong>and</strong> of size comparable<br />

to τ n . But <strong>the</strong>n<br />

1/2<br />

ρf<br />

0<br />

′2 dρ ≥ <br />

<br />

ρf ′2 dρ ≥ <br />

<br />

1 ɛ<br />

dρ = ∞ ,<br />

2 In ρ<br />

contradicting (17). So f has a finite limit at 0, that we denote f(0). Equation (15) can<br />

be rewritten as<br />

d <br />

2 ′<br />

ρ f<br />

dρ<br />

= sin(2f)<br />

,<br />

1 − ρ2 <strong>and</strong> so f cannot remain bounded unless sin(2f(0)) = 0 i.e. f(0) = k π<br />

2 , with k an integer.<br />

If k is odd, we have M0 = 0, hence M remains equal to 0 on (0, 1), that is, f is constant.<br />

<br />

4 An application to <strong>the</strong> development of singularities<br />

4.1 A refined version of Theorem 2.1<br />

Let us begin by noticing that <strong>the</strong> proof of Theorem 2.1 can be adapted (as <strong>for</strong> instance<br />

in [12] or [6]) to yield <strong>the</strong> following <strong>the</strong>orem.<br />

Theorem 4.1 (i) Take (v0, v1) ∈ ˙ B 3/2<br />

2,r × ˙ B 1/2<br />

2,r with 1 < r < ∞. Then <strong>the</strong>re exists T > 0<br />

<strong>and</strong> a solution of (3) in<br />

(19) Xr([0, T ]) def<br />

= <br />

∞<br />

L [0, T ], ˙ B 3/2<br />

<br />

∩ <br />

2<br />

L [0, T ], ˙ B 1/2<br />

<br />

which is unique in a ball of sufficiently small radius of <strong>the</strong> above space.<br />

(ii) Consider now data of <strong>the</strong> <strong>for</strong>m<br />

with<br />

2,r<br />

5,r<br />

v0 = w0 + w0 , v1 = w1 + w1 ,<br />

∩ <br />

3<br />

L [0, T ], ˙ B 1/2<br />

<br />

30/7,r<br />

( w0, w1) ∈ ˙ B 3/2<br />

2,∞ × ˙ B 1/2<br />

2,∞ <strong>and</strong> (w0, w1) ∈ ˙ H 3/2 × ˙ H 1/2 ,<br />

19<br />

,


Assume that all <strong>the</strong> norms of <strong>the</strong> above functions are small enough, <strong>the</strong>n <strong>the</strong>re exists by<br />

Theorem 2.1 a unique solution v ∈ X = X∞ of (3).<br />

Denote by w ∈ X <strong>the</strong> unique solution of (3) with data ( w0, w1); <strong>the</strong>n v reads<br />

v = w + w with v ∈ X2 ↩→ L ∞ (R, ˙ H 3/2 ) .<br />

If r = 2, <strong>the</strong> assertion (i) corresponds simply to <strong>the</strong> well posedness result <strong>for</strong> data in<br />

˙H 3/2 × ˙ H 1/2 proved in [16].<br />

4.2 Regularity breakdown<br />

As we discussed at <strong>the</strong> beginning of Section 3, we expect <strong>for</strong> <strong>the</strong> covariant <strong>wave</strong>-<strong>map</strong><br />

<strong>equation</strong> with N = S 3 <strong>the</strong> development of singularities along <strong>self</strong>-<strong>similar</strong> blow-up profiles.<br />

Consider a solution v which blows up at 0, in <strong>the</strong> sense that v ∈ L ∞ loc ((−ɛ, 0) ∪ (0, ɛ), ˙ H 3/2 .<br />

We expect <strong>the</strong> value at 0 of <strong>the</strong> function to be<br />

with<br />

v(0) = c0<br />

|x| + w0(x) , ∂tv(0) = c1<br />

+ w1(x)<br />

|x| 2<br />

(w0 , w1) ∈ ˙ H 3/2 × ˙ H 1/2 .<br />

We would like to underst<strong>and</strong> what are <strong>the</strong> possible values of c0 <strong>and</strong> c1. The idea is to<br />

look at <strong>the</strong> problem not as a regularity breakdown problem, but as an “instantaneous<br />

smoothing” problem. More precisely, we consider <strong>the</strong> Cauchy problem<br />

(20)<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

vtt − vrr − n−1<br />

v(0) = c0<br />

|x| + w0<br />

vt(0) = c1<br />

|x| 2 + w1(x)<br />

r vr = v 3 Z(rv)<br />

<strong>and</strong> we ask <strong>for</strong> which (c0, c1) ∈ R 2 , (w0, w1) ∈ ˙ H 3/2 × ˙ H 1/2 <strong>the</strong> solution v belongs to<br />

L ∞ loc ((−ɛ, 0) ∪ (0, ɛ), ˙ H 3/2 ) <strong>for</strong> some ɛ > 0.<br />

Two obvious possibilities are (c0, c1) = 0, <strong>and</strong> <strong>the</strong> Shatah-Bizoń finite energy <strong>self</strong> <strong>similar</strong><br />

<strong>solutions</strong> fn (see section 3.1), in which case (w0, w1) = 0.<br />

The following <strong>the</strong>orem states that, <strong>for</strong> small data, <strong>the</strong>se are <strong>the</strong> only possibilities.<br />

Theorem 4.2 If (c0, c1) is small enough, <strong>and</strong> (w0 , w1) ∈ ˙ H 3/2 × ˙ H 1/2 , Theorem 2.1 gives<br />

a local in time solution v of (20).<br />

• If (c0, c1) = (0, 0), <strong>the</strong>re exists ɛ > 0 such that v ∈ L ∞ ((−ɛ, ɛ), ˙ H 3/2 ).<br />

• If (c0, c1) = (0, 0), v does not belong to L ∞ loc ((−ɛ, 0) ∪ (0, ɛ), ˙ H 3/2 ) <strong>for</strong> any ɛ > 0.<br />

Remark 4.1 If we do not assume (c0, c1) to be small, we do not know if a solution exists,<br />

<strong>and</strong> even if it did, <strong>the</strong>re are good chances that it would not be unique, so it is not clear<br />

how to generalize <strong>the</strong> above <strong>the</strong>orem to large data.<br />

Proof: The first point of <strong>the</strong> <strong>the</strong>orem is contained in Theorem 4.1; so we focus on <strong>the</strong><br />

second one, <strong>and</strong> assume that (c0, c1) = (0, 0).<br />

By finite speed of propagation, we can localize matters around 0, <strong>and</strong> hence assume that<br />

<strong>the</strong> norm of (w0 , w1) is small in ˙ H 3/2 × ˙ H 1/2 . By point (ii) of Theorem 4.1, <strong>the</strong> solution<br />

v reads<br />

w + w ,<br />

20


where w ∈ L ∞ (R, ˙ H 3/2 ), <strong>and</strong> w is <strong>the</strong> unique solution of (3) with data<br />

c0<br />

|x| ,<br />

c1<br />

|x| 2<br />

<br />

. We<br />

will prove that w is given by a profile, <strong>and</strong> that this profile does not belong to ˙ H 3/2 .<br />

We follow <strong>the</strong> construction of Section 3 in order to build up a profile f <strong>for</strong> <strong>the</strong> data we<br />

are interested in. First solving from infinity to 1, we have that f(1) has to be close to 0<br />

since <strong>the</strong> data are small. Then in order to solve on (0, 1) we apply Lemma 3.1, where α is<br />

taken small, since g(1) is small. The solution corresponding to this profile has a small X<br />

norm <strong>for</strong> small enough data; by <strong>the</strong> uniqueness in Theorem 2.1, it is equal to w.<br />

All we have to do now is proving that f (<strong>and</strong> hence w) does not belong to ˙ H 3/2 . If f(1)<br />

is not zero, this is obvious, since f ′ behaves like | log |ρ − 1|| close to 1, see Section 3.4.<br />

If f(1) = 0, we observe first that f is 0 on [0, 1]. In order to examine <strong>the</strong> situation on<br />

[1, ∞), we will resort to <strong>the</strong> following lemma. Its proof is very <strong>similar</strong> to that of Lemma 3.1,<br />

we <strong>the</strong>re<strong>for</strong>e omit it.<br />

Lemma 4.1 For any β ∈ R, <strong>the</strong> Cauchy problem<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

f ′′ + 2<br />

ρf ′ − sin(2f)<br />

ρ2 (1−ρ2 = 0 )<br />

f(1) = 0<br />

f ′ (1+) = β<br />

admits a unique solution on (1, ∞).<br />

Now we switch to g(z) = f <br />

1<br />

z , which verifies<br />

g ′′ = (z 2 − 1) sin(2g(z)) .<br />

Integrating two times, we see that g is bounded by<br />

|g(z)| ≤ |z − 1|| log |z − 1|| .<br />

But this implies that g ′′ is integrable, hence g ′ has a limit at 1−, hence f is given on (1, ∞)<br />

by a solution of <strong>the</strong> above lemma, <strong>for</strong> some β. This β cannot be 0, o<strong>the</strong>rwise f would be<br />

identically 0. Hence<br />

f ′ (1−) = f ′ (1+) ;<br />

this is <strong>the</strong> desired singularity, f does not belong to ˙ H 3/2 , see Section 3.4. <br />

5 Appendix : generalized Hardy inequality <strong>and</strong> <strong>Besov</strong> <strong>spaces</strong><br />

It is a consequence of Lemma 1.2. of [16] that if n ≥ 4 <strong>and</strong> f is radial, one has<br />

|f(x)| ≤ C f ˙ H n 2 −1<br />

|x|<br />

The next lemma extends this result to <strong>the</strong> <strong>Besov</strong> space ˙B n<br />

2 −1<br />

2,∞ .<br />

Lemma 5.1 Suppose n ≥ 5. If f is a radial function,<br />

We even have<br />

f n<br />

˙B 2<br />

−1<br />

2,∞<br />

|f(x)| ≤ C<br />

|x|<br />

f n<br />

˙B 2<br />

−1<br />

2,∞<br />

|Sjf(x)| ≤ C<br />

|x|<br />

21<br />

.<br />

.<br />

<strong>for</strong> any j .


Proof : The first inequality can be proved by real interpolation. Indeed, we know (see<br />

Lemma 1.2 of [16]) that, <strong>for</strong> a radial function v <strong>and</strong> α small enough,<br />

rvL∞ (r1+αdr) = <br />

1+α<br />

r v∞ ≤ Cv<br />

˙H d 2<br />

−1−α .<br />

The conclusion follows from <strong>the</strong> two following real interpolation facts<br />

d<br />

˙H 2 −1−α d<br />

, H˙ 2 −1+α<br />

1<br />

2 ,∞ = ˙ B d<br />

2 −1<br />

2,∞<br />

∞ 1−α ∞ 1+α<br />

L (r dr) , L (r dr) 1<br />

2 ,∞ = L∞ (r dr) .<br />

To prove <strong>the</strong> second inequality of Lemma 5.1, we use <strong>the</strong> fact that, denoting <strong>the</strong> st<strong>and</strong>ard<br />

maximal function M,<br />

|Sjf(x)| ≤ CMf(x) ,<br />

<strong>and</strong> that<br />

M<br />

<br />

1<br />

≤<br />

|x|<br />

C<br />

|x| .<br />

Acknowledgements: The author is indebted to Jalal Shatah <strong>for</strong> very interesting <strong>and</strong><br />

useful discussions during <strong>the</strong> writing of this paper.<br />

References<br />

[1] J.-M. Bony, Calcul symbolique et propagation des singularits pour les quations aux<br />

drives partielles non linaires. Ann. Sci. Ecole Norm. Sup. (4) 14 (1981), no. 2,<br />

209–246<br />

[2] H. Bahouri, J.-Y. Chemin, Equations d’ondes quasilinéaires et estimations de<br />

Strichartz. Amer. J. Math. 121 (1999), no. 6, 1337–1377<br />

[3] P. Bizoń, Equivariant <strong>self</strong>-<strong>similar</strong> <strong>wave</strong> <strong>map</strong>s from Minkowski spacetime into 3sphere.<br />

Comm. Math. Phys. 215 (2000), no. 1, 45–56<br />

[4] P. Bizoń, T. Chmaj, Z. Tabor, Dispersion <strong>and</strong> collapse of <strong>wave</strong> <strong>map</strong>s. Nonlinearity<br />

13 (2000), no. 4, 1411–1423<br />

[5] J.-Y. Chemin, N. Lerner, Flot de champs de vecteurs non lipschitziens et quations<br />

de Navier-Stokes. J. Differential Equations 121 (1995), no. 2, 314–328<br />

[6] P. Germain, Global infinite energy <strong>solutions</strong> of <strong>the</strong> critical semilinear <strong>wave</strong> <strong>equation</strong>,<br />

to appear in Rev. Mat. Ibreoamericana<br />

[7] M. Keel, T. Tao, Endpoint Strichartz estimates., Amer. J. Math. 120 (1998), no.<br />

5, 955–980.<br />

[8] S. Klainerman, I. Rodnianski, On <strong>the</strong> global regularity of <strong>wave</strong> <strong>map</strong>s in <strong>the</strong> critical<br />

Sobolev norm. Internat. Math. Res. Notices 2001, no. 13, 655–677<br />

22


[9] J. Krieger, W. Schlag, D. Tataru, Slow blow-up <strong>solutions</strong> <strong>for</strong> <strong>the</strong> H 1 (R 3 ) critical<br />

focusing semi-linear <strong>wave</strong> <strong>equation</strong> in R 3 , preprint available on arxiv<br />

[10] P.G. Lemarié-Rieusset, Recent developments in <strong>the</strong> Navier-Stokes problem., Chapman<br />

& Hall/CRC (2002).<br />

[11] Y. Meyer, R.R. Coifman, Ondelettes et oprateurs III : Oprateurs multilinaires.<br />

Actualits Mathmatiques. Hermann, Paris, 1991<br />

[12] F. Planchon, Self-<strong>similar</strong> <strong>solutions</strong> <strong>and</strong> semi-linear <strong>wave</strong> <strong>equation</strong>s in <strong>Besov</strong> <strong>spaces</strong>.<br />

J. Math. Pures Appl. (9) 79 (2000), no. 8, 809–820<br />

[13] I. Rodnianski, J. Sterbenz, On <strong>the</strong> Formation of Singularities in <strong>the</strong> Critical O(3)<br />

Sigma-Model, preprint available on arxiv<br />

[14] H.-J. Schmeisser, H. Triebel, Topics in Fourier analysis <strong>and</strong> function <strong>spaces</strong>. A<br />

Wiley-Interscience Publication. John Wiley & Sons, Ltd., Chichester, 1987<br />

[15] J. Shatah, Weak <strong>solutions</strong> <strong>and</strong> development of singularities of <strong>the</strong> SU(2) σ-model.<br />

Comm. Pure Appl. Math. 41 (1988), no. 4, 459–469<br />

[16] J. Shatah, S. Tahvildar-Zadeh, On <strong>the</strong> Cauchy problem <strong>for</strong> equivariant <strong>wave</strong> <strong>map</strong>s.<br />

Comm. Pure Appl. Math. 47 (1994), no. 5, 719–754<br />

[17] J. Shatah, M. Struwe, The Cauchy problem <strong>for</strong> <strong>wave</strong> <strong>map</strong>s. Int. Math. Res. Not.<br />

2002, no. 11, 555–571<br />

[18] T. Tao, Global regularity of <strong>wave</strong> <strong>map</strong>s. I. Small critical Sobolev norm in high<br />

dimension. Internat. Math. Res. Notices 2001, no. 6, 299–328.<br />

[19] T. Tao, Global regularity of <strong>wave</strong> <strong>map</strong>s. II. Small energy in two dimensions. Comm.<br />

Math. Phys. 224 (2001), no. 2, 443–544<br />

[20] D. Tataru, Rough <strong>solutions</strong> <strong>for</strong> <strong>the</strong> <strong>wave</strong> <strong>map</strong>s <strong>equation</strong>. Amer. J. Math. 127 (2005),<br />

no. 2, 293–377.<br />

23


Pierre GERMAIN<br />

Courant Institute of Ma<strong>the</strong>matical Sciences<br />

New York University<br />

New York, NY 10012-1185<br />

USA<br />

pgermain@math.nyu.edu<br />

24

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!