• Aucun résultat trouvé

On the L p norm of spectral clusters for compact manifolds with boundary

N/A
N/A
Protected

Academic year: 2022

Partager "On the L p norm of spectral clusters for compact manifolds with boundary"

Copied!
47
0
0

Texte intégral

(1)

c

2007 by Institut Mittag-Leffler. All rights reserved

On the L p norm of spectral clusters for compact manifolds with boundary

by

Hart F. Smith

University of Washington Seattle, WA, U.S.A.

Christopher D. Sogge

Johns Hopkins University Baltimore, MD, U.S.A.

1. Introduction

LetM be a compact 2-dimensional manifold with boundary, and let P be an elliptic, second-order differential operator onM, self-adjoint with respect to a density dµ, and with vanishing zero-order term, so that in local coordinates

(P f)(x) =%(x)−1

n

X

j,k=1

j(%(x)gjk(x)∂kf(x)), dµ=%(x)dx. (1.1)

We take the gjk’s to be positive, so that the Dirichlet eigenvalues ofP can be written as {−λ2j}j=0.

Let χλ be the projection of L2(dµ) onto the subspace spanned by the Dirichlet eigenfunctions for whichλj∈[λ, λ+1]. In the case whereM is compact without boundary of dimensionn>2, and the coefficients ofP areCfunctions, Sogge [14] established the following bounds:

λfkLq(M)6Cλ((n−1)/2)(1/2−1/q)kfkL2(M), 26q6qn, (1.2) kχλfkLq(M)6Cλn(1/2−1/q)−1/2kfkL2(M), qn6q6∞. (1.3) Furthermore, the exponent ofλis sharp on every such manifold (see, e.g., [15]). In the case of a sphere, the examples which prove sharpness are in fact eigenfunctions. For (1.2), the appropriate example is an eigenfunction which concentrates in aλ−1/2-diameter tube about a geodesic. For (1.3), the example is a zonal eigenfunction ofL2 norm λ(n−1)/2 which takes on values comparable toλon aλ−1-diameter ball about each of the north and

The authors were supported by the National Science Foundation, Grants DMS-0140499, DMS- 0099642, and DMS-0354668.

(2)

south poles. Approximate spectral clusters with similar properties can be constructed in the interior of any smooth manifold, showing that for spectral clusters (though not necessarily eigenfunctions) the exponents in (1.2) and (1.3) are also lower bounds on manifolds with boundary.

In [13], the authors showed that, on a manifold of dimension n>2 for which the boundary is everywhere strictly geodesically concave (such as the complement inRn of a strictly convex set), the estimates (1.2) and (1.3) both hold.

On the other hand, Grieser [5] observed that in the unit disk{x:|x|61} there are eigenfunctions of the Laplacian, for Dirichlet as well as for Neumann boundary conditions, of eigenvalue−λ2that concentrate within aλ−2/3-neighborhood of the boundary. These are the classical Rayleigh whispering gallery modes (see [10] and [9]). The Fourier–

Airy calculus of Melrose and Taylor allows one to construct an approximate spectral cluster with similar localization properties near any boundary point ofM at which the boundary is strictly convex (the gliding case). Consequently, ifM is of dimension 2 and the boundary has a point of strict convexity with respect to the metric g (for instance, any smoothly bounded planar domain endowed with the standard Laplacian and either Dirichlet or Neumann conditions), then the following bounds cannot be improved upon:

λfkLq(M)6Cλ(2/3)(1/2−1/q)kfkL2(M), 26q68, (1.4) kχλfkLq(M)6Cλ2(1/2−1/q)−1/2kfkL2(M), 86q6∞. (1.5) In this paper we show that the estimates (1.4) and (1.5) hold on any 2-dimensional compact manifold with boundary, forP as above and either Dirichlet or Neumann con- ditions assumed. Estimate (1.4) follows by interpolation of the trivial caseq=2 with the caseq=6, so we restrict attention toq>6 for (1.4). Forq>6, the above estimates are an immediate consequence of the following theorem (see for example [8] or [11]).

Theorem 1.1. Suppose that usolves the Cauchy problem on R×M,

t2u(t, x) =P u(t, x), u(0, x) =f(x), ∂tu(0, x) = 0, (1.6) and satisfies either Dirichlet conditions:

u(t, x) = 0 if x∈∂M ,

or Neumann conditions, where Nx is a unit normal field with respect to g:

Nx·∇xu(t, x) = 0 if x∈∂M. Then,the following bounds hold for 66q68:

kukLqxL2

t(M×[−1,1])6CkfkHγ(q)(M), γ(q) =2 3

1 2−1

q

,

(3)

and the following bounds hold for 86q6∞:

kukLqxL2

t(M×[−1,1])6CkfkHδ(q)(M), δ(q) = 2 1

2−1 q

−1 2.

In the statement of the theorem, the space Hs(M) refers to the Sobolev space of ordersonM determined, respectively, by Dirichlet or Neumann eigenfunctions.

Our approach to proving Theorem 1.1 is to work in geodesic normal coordinates near∂M, and to extend both the operatorP and the solutionuacross the boundary, to obtainuas a solution to a wave equation on an open set, but for an operator with coef- ficients of Lipschitz regularity. We then adapt a frequency dependent scaling argument, originally developed to handle Lipschitz metrics, to metrics with the particular type of codimension-1 singularities that the extendedP will have.

We remark that, for operators of the type (1.1) with %and the gjk’s of Lipschitz regularity, the estimate (1.4) is known on the range 26q66, as established by the first author in [12], along with a weaker version of (1.5) having larger exponent if q <∞.

It is not currently known what the sharp exponents are for general Lipschitz P, since the known counterexamples satisfy the estimates (1.5). The estimates for q=∞ were established for eigenfunctions recently by Grieser [6], while the sup-norm estimates for spectral clusters were obtained by the second author in [16].

For q=∞, the squarefunction estimate of Theorem 2.1 below was shown in [12] to hold for operators P with Lipschitz coefficients, which in particular implies the q=∞

case of Theorem 1.1 for P on a manifold with boundary. Our proof here of the case q <∞, however, depends crucially on the fact that if u is appropriately microlocalized away from directions tangent to∂M, then better squarefunction estimates hold than do for directions near to tangent. In other words, we exploit the fact that the more highly localized eigenfunctions considered in [5] are associated only to gliding directions along

∂M, not directions transverse to∂M.

A historical curiosity is that the critical L2!L8 bounds for χλ have an analog in Euclidean space which seems to be the first restriction theorem for the Fourier transform.

To explain this, we first notice that, by duality, ourL2!L8bounds are equivalent to the statement thatχλ:L8/7!L2 with norm O(λ1/4). The Euclidean analog would say that ifχλ:L8/7(R2)!L2(R2) denotes the projection onto Fourier frequencies |ξ|∈[λ, λ+1], then this operator also has norm O(λ1/4). An easy scaling argument shows then that the latter result is equivalent to the following Fourier restriction theorem for the circle

Z

0

|f(cosˆ θ,sinθ)|21/2

6CkfkL8/7(R2), f∈C0(R2).

(4)

Motivated by a question of DeLeeuw [4], Stein [17] proved this by a now standard T T argument, using the fact that convolution with cdθ maps L8/7(R2)!L8(R2) by the Hardy–Littlewood–Sobolev theorem, as |cdθ|6C|x|−1/2. Since this argument does not use the oscillations ofcdθ, one can strengthen the above restriction theorem to show that, forj>1, one has the uniform bounds

Z 2−j

0

|fˆ(cosθ,sinθ)|21/2

6C2−j/8kfkL8/7(R2), f∈C0(R2). (1.7) By the Knapp example, there is no small angle improvement for the critical

L6/5(R2)−!L2(S1)

restriction theorem of Stein–Tomas. A key step for us is that in the setting of compact manifolds with boundary, we also get the sameO(2−j/8) improvement in ourL8-estimates when microlocalized to regions of phase space that correspond to bicharacteristics that are of angle comparable to 2−j from tangency to the boundary.

In higher dimensions the natural analog of (1.4)–(1.5) would say that kχλfkLq(M)6Cλ(2/3+(n−2)/2)(1/2−1/q)kfkL2(M), 26q66n+4

3n−4, (1.8)

λfkLq(M)6Cλn(1/2−1/q)−1/2kfkL2(M), 6n+4

3n−46q6∞. (1.9) By higher dimensional versions of the Rayleigh whispering gallery modes, this would be sharp if true. At present we are unable to prove this estimate but, as we shall indicate in the final section, we can prove the bounds in (1.9) for the smaller range of exponents q>4 ifn>4, andq>5 ifn=3. We hope to return to the problem of proving sharp results in higher dimensions in a future work.

Notation. We use the following notation. The symbol a.b means that a6Cb, where C is a constant that depends only on globally fixed parameters (or on N, α andβ, in case of inequalities involving general integers).

For convenience, we will letx3 serve as substitute for the time variablet. We use d=(d1, d2, d3) to denote the gradient operator, andD=−id.

2. Dyadic localization arguments

The estimates of Theorem 1.1 hold if u is supported away from ∂M, by the results of [8]. Consequently, by finite propagation velocity and the use of a smooth partition of unity, we may assume that, forT small, the solutionu(t, x) in Theorem 1.1 is for|t|6T

(5)

supported in a suitably small coordinate patch centered on the boundary. Note that if we establish Theorem 1.1 on the set|t|6T for some smallT, it then holds for T=1 by energy conservation.

We work in boundary normal coordinates for the Riemannian metric gjkthat is dual to gjk of (1.1). Thus,x2>0 will define the manifold M, andx1is a coordinate function on∂M which we choose so that∂x1 is of unit length along∂M. In these coordinates,

g22(x1, x2) = 1, g11(x1,0) = 1 and g12(x1, x2) = g21(x1, x2) = 0. (2.1) The metric gjk forP is the inverse to gjk, and the same equalities hold for it.

We now extend the coefficient g11and%in an even manner across the boundary, so that

g11(x1,−x2) = g11(x1, x2) and %(x1,−x2) =%(x1, x2). (2.2) The extended functions are then piecewise smooth, and of Lipschitz regularity across x2=0. Because g is diagonal, the operatorP is preserved under the reflectionx27!−x2. After multiplying %(x) by a constant, and rescaling variables if necessary, we may assume that on the ball|x|<1 the function%(x) isCN(R2+)-close to the function 1, and gjk(x) isCN(R2+)-close to the Euclidean metric, whereN is suitably large, andc0 will be taken suitably small:

k%−1kCN(R2+)6c0 and kgjk−δjkkCN(R2+)6c0. (2.3) We may then extend %and gjk globally, preserving conditions (2.1)–(2.3), so thatP is defined globally onR2and such that

%(x) = 1 and gjk(x) =δjk for|x|>34. (2.4) We then extend the initial data f and the solutionuto be odd inx2 (respectively, even in x2 in case of Neumann conditions). This extension map is seen to map the Dirichlet (respectively, Neumann) Sobolev spaceH2(Rn+) toH2(Rn), and henceHδ(Rn+) to Hδ(Rn) for 06δ62. The extended solution u thus solves the extended equation

2tu=P u on R×R2, with the extended initial data f. The result of Theorem 1.1 is therefore a direct consequence of the following one.

Theorem 2.1. Suppose that the operator P takes the form (1.1),and that %and g satisfy conditions (2.1)–(2.4)above. Let usolve the Cauchy problem on R×R2:

t2u(t, x) =P u(t, x), u(0, x) =f(x), ∂tu(0, x) =g(x). (2.5)

(6)

Then the following bounds hold for 66q68:

kukLqxL2

t(R2×[−1,1]).(kfkHγ(q)+kgkHγ(q)−1), γ(q) =2 3

1 2−1

q

,

and the following bounds hold for 86q6∞:

kukLqxL2

t(R2×[−1,1]).(kfkHδ(q)+kgkHδ(q)−1), δ(q) = 2 1

2−1 q

−1 2.

We begin by reducing matters to compactly supported u satisfying an inhomoge- neous equation. Henceforth, we will use the notationx3=t. Letφ(x) be a smooth even function onR3, equal to 1 for|x|632, and vanishing for|x|>2. We may then write

3

X

j=1

Dj(ajj(x)Dj(φu)(x)) =

3

X

j=1

DjFj(x), where

a33(x) =%(x) and ajj(x) =−%(x)gjj(x) forj= 1,2.

We express this equation concisely asDAD(φu)=DF, and observe that, for 06δ62, kφukHδ(R3)+kFkHδ(R3).kfkHδ+kgkHδ−1.

This is a consequence of energy estimates, which hold separately onR3+andR3, together with the fact thatDAD(φu) is compactly supported and has integral 0, so may be written asDF.

We may thus assume that u(x) is supported in the ball |x|62, and need to show that

kukLqL2.kukHγ(q)+kFkHγ(q), 66q68, (2.6) kukLqL2.kukHδ(q)+kFkHδ(q), 86q6∞, (2.7) whereDADu=DF.

Next let Γ(ξ) be a multiplier of order 0, supported in the set ξ:143|6|(ξ1, ξ2)|64|ξ3|}, which equals 1 on the set

ξ:123|6|(ξ1, ξ2)|62|ξ3|}.

The operatorDAD is elliptic on the support of 1−Γ, and we may write DAD(1−Γ(D))u= (1−Γ(D))DF−D[A,Γ(D)]Du.

(7)

As a consequence of the Coifman–Meyer commutator theorem [2] (see also [20, Proposi- tion 3.6.B]) the operator [A,Γ(D)] mapsHδ−1!Hδ for 06δ61. Hence, the right-hand side of the above belongs to Hδ−1, and, by Sobolev embedding and elliptic regularity (see, for example, [20, Theorem 2.2.B], which applies in the Sobolev setting), we have

k(1−Γ(D))ukLqL2.k(1−Γ(D))ukHδ(q)+1.kukHδ(q)+kFkHδ(q).

Indeed, there is an extra 12 derivative inδ(q)+1 beyond the Sobolev indexn(1/2−1/q), so this holds for all 26q6∞. Sinceγ(q)>δ(q) forq68, this implies that (2.6) and (2.7) hold forureplaced by (1−Γ(D))uon the left-hand side.

It thus remains to establish (2.6) and (2.7) with u replaced on the left by Γ(D)u.

We take a Littlewood–Paley decomposition inξto write Γ(D)u=

X

k=1

Γk(D)u=

X

k=1

uk,

with ˆuk supported in a region where|(ξ1, ξ2)|≈|ξ3|and|ξ|≈2k. Since these regions have finite overlap in theξ3 axis, we have

kΓ(D)ukLqL2.kukkLq`2kL2.kukk`2 kLqL2, where we useq>2 at the last step.

Now let Ak denote the matrix of coefficients obtained by truncating the frequencies ofaii(x) to|ξ|6c2k for a fixed smallc. We then haveDAkDuk=DFk, where

Fk= Γk(D)F+[A,Γk(D)]Du+(Ak−A)Duk. (2.8) Note that the inhomogeneity Fk is now localized in frequency to |ξ3|≈|ξ|≈2k, by the frequency localizations ofAk anduk.

We claim that, for 06δ61,

X

k=1

22kδkFkk2L2.kuk2Hδ+kFk2Hδ.

This follows by orthogonality for the first term on the right of (2.8), and the last term is handled by the boundkA−AkkL.2−k. The middle term is handled by the Coifman–

Meyer commutator theorem, which yields thatP

k=1εk[A,Γk(D)] maps Hδ−1!Hδ for all sequencesεk=±1.

We thus are reduced to establishing uniform estimates for each dyadically localized pieceuk. We thus fix a frequency scale λ=2k for the rest of this paper. We then need to prove the following estimates, where we now setDAλDuλ=Fλ,

kuλkLqL2(R3)γ(q)(kuλkL2(R3)−1kFλkL2(R3)), 66q68, kuλkLqL2(R3)δ(q)(kuλkL2(R3)−1kFλkL2(R3)), 86q6∞.

(8)

Since we are usingx3 orthogonality to make this reduction, we must control the norms of theuλ’s globally. However, sinceuis supported in the ball of radius 2, it is easy to see that the norm ofuλover|x|>3 is bounded byλ−1kukL2, so in fact it suffices to establish the above estimate with the left-hand side norm taken over the cube of sidelength 3.

If we let vλ denote the localization ofuλ to frequencies where|ξ2|>183|, then the squarefunction estimates hold forvλas on an open manifold:

kvλkLqL2(R3)δ(q)(kuλkL2(R3)+kFλkL2(R3)), 66q6∞.

This will follow as a consequence of the techniques that we use to handle the part ofuλ with frequencies localized to angles≈1 from theξ3 axis.

Consequently, we will assume that supp(ˆuλ)⊆

ξ:|ξ1| ∈1

2λ,2λ

,|ξ2|6101λand|ξ3| ∈1

2λ,2λ .

On this region, the operator DAλD is hyperbolic with respect to the x1 direction. We can thus takep(x, ξ0), a positive elliptic symbol in ξ0=(ξ2, ξ3), so that

a11λ (x)(ξ12−p(x, ξ0)2) =

3

X

j=1

ajjλ(x)ξj2 if |ξ2|619λ and |ξ3| ∈1

3λ,3λ ,

and such that

p(x, ξ0) =|ξ0| if ξ0∈/

18λ,18λ

×1 4λ,4λ

.

We also smoothly setp(x, ξ0)=1 nearξ0=0. Thus, p(x, ξ0), dxp(x, ξ0)∈S1,11 ,

andp(x, ξ0) differs from|ξ0|by a symbol supported in the dyadic shell|ξ0|≈λ.

Next, let pλ(x0, ξ) be obtained by truncating the symbol p(x, ξ0) to x0-frequencies less thancλ, where cis a small constant. Then, uniformly overλ,

pλ(x, ξ0)−p(x, ξ0)∈S1,10 and supp(pλ−p)⊂ {ξ0:|ξ0| ≈λ}.

Furthermore, the symbol-composition rule holds forpλ to first order. Consequently, we can write

(D1+pλ(x, D0))(D1−pλ(x, D0))uλ=Fλ0, where

kFλ0kL2(R3).λkuλkL2(R3)+kFλkL2(R3).

(9)

The functionuλcan be written as the sum of four pieces with disjoint Fourier transforms, according to the possible signs of ξ1 and ξ3. We restrict attention to the piece u+λ, supported whereξ1>0 and ξ3>0. Estimates for the other pieces will follow similarly.

Since pλ is x0-frequency localized, Fλ0 also splits into four disjoint pieces. The symbol ξ1+p(x, ξ0) is elliptic on the regionξ1>0, hence we may write

D1u+λ−pλ(x, D0)u+λ=Fλ00, where

kFλ00kL2(R3).kuλkL2(R3)−1kFλkL2(R3). Finally, we have that

pλ(x, D0)−pλ(x, D0)∈Op(S01,1),

and is dyadically supported inξ0. We have thus reduced the proof of Theorem 2.1 to the proof of the following result.

Theorem 2.2. Suppose that the x0-Fourier transform of uλ satisfies the support condition

supp(ˆuλ)⊆

ξ0:|ξ2|6101λand ξ31

2λ,2λ , and that

D1uλ−Pλuλ=Fλ,

where Pλ=12(pλ(x, D0)+pλ(x, D0)). Then, for S=[0,1]×R2,

kuλkLqL2(S)γ(q)(kuλkLL2(S)+kFλkL2(S)), 66q68, kuλkLqL2(S)δ(q)(kuλkLL2(S)+kFλkL2(S)), 86q6∞.

(2.9)

The use of the LL2 norm ofuλ is allowed by Duhamel and energy bounds. Here, as in what follows, we are using the shorthand mixed-norm notation thatLpLq=Lpx

1Lqx0. 3. The angular localization

In this section we take a further decomposition ofuλ, by decomposing its Fourier trans- form dyadically in theξ2 variable. The reductions of the previous section required only the fact that the coefficientsajj(x) were Lipschitz functions. The reduction to estimates for angular pieces depends on the fact that the singularities of ajj(x), and hence the points where thex2-derivatives ofpλ(x, ξ0) are large, occur only atx2=0. Consequently, various error terms that arise in this further reduction will be highly concentrated at x2=0, which we express through weightedL2 estimates.

(10)

We will take a dyadic decomposition of theξ2variable, from scaleξ2≈λ2/3toξ2≈λ.

Thus, for 16j <Nλ=13log2λ, letβj0)=βj2, ξ3) denote a smooth cutoff satisfying supp(βj)⊂[2−j−2λ,2−j+1λ]×1

4λ,4λ ,

andβNλ supported in [−λ2/3, λ2/31

4λ,4λ

, such that, withβ−j2, ξ3)=βj(−ξ2, ξ3),

Nλ

X

j=1

βj0)+

−1

X

j=1−Nλ

βj0) = 1 if |ξ2|618λ and ξ31

2λ,2λ .

Let

uj(x) =βj(D0)uλ(x).

If we define

θj= 2−|j|,

thenuj has frequencies localized toξ2≈±θjξ3, or|ξ2|.λ−1/3ξ3in casej=Nλ.

On the microlocal support ofuj, the bicharacteristic equation for the principal sym- bolξ1−pλ(x, ξ0) satisfiesdx2/dx1≈±θj, respectively asj >0 orj <0. A bicharacteristic curve passing through the microlocal support ofuj will satisfy this condition on an in- terval ofx1-length less thanεθj, ifεis a small constant. It is thus natural that we will have good estimates foruj on slabs of widthεθj in thex1variable, and it turns out that this is sufficient to prove Theorem 2.2.

In proving estimates foruj, it is convenient to work with the symbolpj obtained by truncatingp(x, ξ0) tox0-frequencies less thancθ−1/2j λ1/2. This finer truncation than that ofpλis chosen so that, after rescaling the space byθj, the rescaled symbolpjjx,·) will bex0-frequency truncated atµ1/2, whereµ=θjλis the frequency scale of the rescaled so- lutionujjx). This square root truncation is consistent with the wave packet techniques we use, and is standard in the construction of parametrices for rough metrics.

The energy of the induced error term (Pλ−Pj)uwill be large at x2=0, but decays away from x2=0 at a rate that is integrable along bicharacteristic curves that traverse the boundary at angleθj. This error term can thus be considered as a bounded driving force, and we call this termGj below.

In the next two sections we will establish the following result.

Theorem 3.1. Let Sj,k denote the slab x1∈[kεθj,(k+1)εθj] for 06k6ε−12|j|. Then, if

D1uj−Pjuj=Fj+Gj,

(11)

it holds uniformly over j, k, and 66q6∞, that

kujkLqL2(Sj,k)δ(q)θ1/2−3/qj kujkLL2(Sj,k)+kFjkL1L2(Sj,k)

1/4θ1/4j khλ1/2θ−1/2j x2i−1ujkL2(Sj,k)

−1/4θj−1/4khλ1/2θ−1/2j x2i2GjkL2(Sj,k)

.

For j=Nλ, it holds that

kujkLqL2(Sj,k)δ(q)θj1/2−3/q(kujkLL2(Sj,k)+kFj+GjkL1L2(Sj,k)).

The gain of the factorθ1/2−3/qj reflects the fact that, forq >6, there is an improvement in the squarefunction estimates if the solution is localized to a small conic set in frequency.

The terms Gj arise naturally in both the linearization step of Lemma 4.4 and the paradifferential smoothing (6.2). They reflect the fact that the singularities ofd2ajj(x) are localized tox2=0. The weightedL2bound onuj is a characteristic energy estimate.

If θj≈1, then the weightedL2 bound on Gj dominates theL1x

2L2x

1,x3 norm ofGj, and exchangingx1 and x2 we could treat Gj and Fj the same. In this case the bound onuj would be dominated by the Lx

2L2x

1,x3 norm. For smallθj’s, however, we cannot usex2 as our “time” variable, and we are forced to work with the weighted L2 norms.

These weighted norms can be thought of as an energy norm along the bicharacteristic flow at angleθj. Precisely, if one replaced x2 by θj(x1−c) in the weight, then the weighted L2norms ofuj andGj would behave like theLL2andL1L2 norms, respectively. The crossing pointc differs, however, for different bicharacteristics.

The proof of Theorem 3.1 is contained in §4 and§5. In§6 we establish the appro- priate bounds on the norms occuring on the right side if, as above,ujj(D0)uλ, while Fj and Gj are defined in (6.1)–(6.2) below.

To state the bounds required, letcj,kdenote the term occuring inside the parentheses on the right-hand side of Theorem 3.1. In§6, we show that, ifD1uλ−Pλuλ=Fλ, then we have a uniform summability condition

X

j

c2j,k(j).kuλk2LL2(S)+kFλk2L2(S), (3.1)

wherek(j) denotes any sequence of values forksuch that the slabsSj,k(j)are nested, in that forj >0 we haveSj+1,k(j+1)⊂Sj,k(j)(with the analogous condition forj <0.)

In the remainder of this section we show how Theorem 2.2 follows from Theorem 3.1 together with the bound (3.1).

We first remark that, if q is a fixed index with q6=8, the bounds of Theorem 2.2 hold (with constant depending onq) under the weaker assumption that the cj,k’s are

(12)

uniformly bounded by the right-hand side of (3.1). To see this, we sum over the 2jε−1 slabs and write

kujkLqL2(S)6 2jε−1

X

k=1

kujkqLqL2(Sj,k)

1/q

δ(q)θ1/2−4/qj kcj,kk`j `k

δ(q)θ1/2−4/qj (kuλkLL2(S)+kFλkL2(S)).

The values ofθj=2−|j| vary dyadically fromλ−1/3 to 1. Forq >8 we can sum overj to obtain

kuλkLqL2(S)δ(q)(kuλkLL2(S)+kFλkL2(S)), and for 66q <8 the sum yields

kuλkLqL2(S)δ(q)−(1/3)(1/2−4/q)(kuλkLL2(S)+kFλkL2(S)).

The above exponent of λequals γ(q), yielding the desired bound. The geometric sum, however, increases asq!8, and yields a logarithmic loss inλatq=8.

To obtain the bound atq=8, and hence uniform bounds overqin Theorem 2.2, we use the following worst-case branching argument. We consider terms withj >0 here, the negative terms being controlled by the same argument.

LetS1,k(1) denote the slab at scale ε2−1 that maximizes kuλkL8L2(S1,k). Since the decomposition ofuλintouj’s is a Littlewood–Paley decomposition in theξ2variable, we have

kuλk2L8L2(S1,k(1)).

Nλ

X

j=1

|uj|2 1/2

2

L8L2(S1,k(1))

.

By the Minkowski inequality,

Nλ

X

j=1

|uj|2 1/2

2

L8L2(S1,k(1))

6ku1k2L8L2(S1,k(1))+

X

S2,k⊂S1,k(1)

Nλ

X

j=2

|uj|2 1/2

8

L8L2(S2,k)

2/8

6ku1k2L8L2(S1,k(1))+22/8

Nλ

X

j=2

|uj|2 1/2

2

L8L2(S2,k(2))

,

where k(2) is chosen to maximize k(P

j=2|uj|2)1/2kL8(S2,k) among the two slabs S2,k contained inS1,k(1). Repeating this procedure yields a nested sequence such that

ε1/4kuλk2L8L2(S)6ku1k2L8L2(S1,k(1))+22/8ku2k2L8L2(S2,k(2))+24/8ku3k2L8L2(S3,k(3))+...

2δ(8)(c21,k(1)+c22,k(2)+c23,k(3)+...),

where the last inequality holds by Theorem 3.1 sinceθj1/2−3/8=2−j/8. The case q=8 of Theorem 2.2 follows by (3.1).

(13)

4. The wave packet transform

The purpose of this section and the next one is to establish Theorem 3.1. We assume for these two sections that we have fixedλandθj, and considerj >0 so thatξ2>0 (except for the termj=Nλ, where|ξ2|6λ2/3).

We will rescale the space byθj. Thus, we work with the function u(x) =ujjx),

which, forj6=Nλ, is supported in the set ξ:ξ21

4θjµ,2θjµ

andξ31

4µ,4µ , where

µ=θjλ

is now the frequency scale foru(x). Forj=Nλ, we have|ξ2|6µ1/2and θNλ−1/2. Letq(x, ξ0) denote the rescaled symbol

q(x, ξ0) =θjpjjx, θ−1j ξ0),

which is truncated to x0-frequencies less than cµ1/2. For|ξ0|≈µ, the symbol q satisfies the estimates

|∂xβξα0q(x, ξ0)|.

µ1−|α|, if|β|= 0,

c0(1+µ(|β|−1)/2θj1/2x2i−N1−|α|, if|β|>1. (4.1) This follows from (6.32).

In the remainder of this section and the next one, we will drop the index j. The quantities θ and µ are the two relevant parameters for our purposes. After rescaling the estimates of Theorem 3.1, and translating Sj,k in x1 to x1=0, we are reduced to establishing the following result. Here,S denotes the (x1, x0) slab [0, ε]×R2.

Theorem 4.1. Suppose that u(ξ)ˆ is supported in the set ξ:ξ21

4θµ,2θµ

and ξ31

4µ,4µ , respectively

ξ:|ξ2|6µ1/2 and ξ31

4µ,4µ in case θ=µ−1/2. Suppose that usatisfies D1u−q(x, D0)u=F+G

on the slabS, where q satisfies (4.1),and is truncated to x0-frequencies less than cµ1/2. Then the following bounds hold, uniformly over θ, µ,and 66q6∞:

kukLqL2(S)δ(q)θ1/2−3/q kukLL2(S)+kFkL1L2(S)1/4θ1/2khµ1/2x2i−1ukL2(S)

−1/4θj−1/2khµ1/2x2i2GkL2(S)

, and forθ=µ−1/2

kukLqL2(S)δ(q)θ1/2−3/q(kukLL2(S)+kF+GkL1L2(S)).

(14)

We introduce at this point the phase-space transform which will be used to establish Theorem 4.1. This transform is essentially the C´ordoba–Fefferman wave packet trans- form [3]. The precise form here is a simple modification of the Fourier–Bros–Iagolnitzer transform used by Tataru in [18] and [19] to establish Strichartz estimates for low reg- ularity metrics; the difference is that in our applications we use a Schwartz function with compactly supported Fourier transform, instead of the Gaussian, as the fundamen- tal wave packet. This is useful in that it strictly localizes the frequency support of the transformed functions. Our transform will act on thex0=(x2, x3) variables.

We use the notion of the previous sections: x=(x1, x2, x3)=(x1, x0), where x3 de- notes the variablet.

Fix a real, radial Schwartz functiong(z0)∈S(R2), with kgkL2(R2)=(2π)−1, and as- sume that its Fourier transform ˆg(ζ0) is supported in the ball{ζ0:|ζ0|6c}. Forµ>1, we defineTµ:S0(R2)!C(R4) by the rule

(Tµf)(x0, ξ0) =µ1/2 Z

e−ihξ0,y0−x0ig(µ1/2(y0−x0))f(y0)dy0. A simple calculation shows that

f(y0) =µ1/2 Z

eihξ0,y0−x0ig(µ1/2(y0−x0))(Tµf)(x0, ξ0)dx00, so thatTµTµ=I. In particular,

kTµfkL2(R4)=kfkL2(R2). (4.2) It will be useful to note that this holds in a more general setting.

Lemma4.2. Suppose that gx00(y0)is a family of Schwartz functions onR2, depend- ing on the parameters x0 and ξ0, with uniform bounds over x0 and ξ0 on each Schwartz norm of g. Then the operator

(Tµf)(x0, ξ0) =µ1/2 Z

e−ihξ0,y0−x0igx001/2(y0−x0))f(y0)dy0 satisfies the bound

kTµfkL2(R4).kfkL2(R2). Proof. Tµ is bounded if and only if Tµ is bounded. Since

kTµFk2L26kTµTµFkL2kFkL2,

it suffices to see thatTµTµ is bounded onL2(dy00).

(15)

The operatorTµTµ is an integral operator with kernel K(y0, η0;x0, ξ0) =µeihη0,y0i−ihξ0,x0i

Z

eihξ0−η0,z0igy001/2(z0−y0))gx001/2(z0−x0))dz0. A simple integration by parts argument shows that

|K(y0, η0;x0, ξ0)|.(1+µ−1/20−ξ0|+µ1/2|y0−x0|)−N,

with constants depending only on uniform bounds for a finite collection of seminorms of gx00 depending on N. The L2(R4) boundedness of TµTµ then follows by Schur’s lemma.

A corollary of Lemma 4.2 is that, forN positive or negative,

khµ1/2x2iNTµfkL2(R4).khµ1/2x2iNfkL2(R2), (4.3) by consideringgx0(y)=hµ1/2x2iN1/2x2−y2i−Ng(y0).

The next two lemmas relate the conjugation of q(x, D0), by the wave packet trans- form, to the Hamiltonian flow underq. These results are analogous to [18, Theorem 1], in that the error term is one order better where the metric has two bounded derivatives.

In our case the second derivatives are large along the boundaryx2=0, which leads to larger errors there. A key fact for our paper is that the errors are suitably integrable along the Hamiltonian flow ofq.

Lemma 4.3. Let q(x, ξ0)satisfy the estimates (4.1). Suppose that |ξ0|≈µ. Then,if q(y, Dy0) acts on the y0 variable, and y1=x1,we can write

(q(y, Dy0)−idξ0q(x, ξ0)·dx0+idx0q(x, ξ0)·dξ0)[eihξ0,y0−x0ig(µ1/2(y0−x0))]

=eihξ0,y0−x0igx,ξ01/2(y0−x0)), wheregx,ξ0(·)denotes a family of Schwartz functions onR2depending on the parameters xand ξ0,each of which has Fourier transform supported in the ball of radius 2c. If k · k denotes any of the Schwartz seminorms, we have

kgx,ξ0k.1+c0µ1/2θhµ1/2x2i−3, where c0 is the small constant of (2.3).

Proof. LettingFdenote the Fourier transform with respect toy0, we write F(q(y, Dy0)−idξ0q(x, ξ0)·dx0+idx0q(x, ξ0)·dξ0)[eihξ0,y0−x0ig(µ1/2(y0−x0))](η0)

=e−ihη0,x0iµ−1gdx,ξ0−1/20−ξ0)),

(16)

wheregdx,ξ00) is equal to Z

e−ihη0,y0i[q(x+µ−1/2y0, ξ01/2η0)−q(x, ξ0)−dx00q(x, ξ0)·(µ−1/2y0, µ1/2η0)]g(y0)dy0

= Z 1

0

(1−σ) Z

e−ihη0,y0iσ2(q(x+σµ−1/2y0, ξ0+σµ1/2η0))g(y0)dy0

dσ.

The spectral restriction onqandgimplies that this vanishes for|η0|>2c. Consequently, it suffices to establishCbounds inη0for the term in brackets, uniformly overσ∈[0,1] and

0|62c. Since the effect of differentiating the integrand with respect toη0 is innocuous, as the rapid decrease ing(y0) counters any polynomial in y0, we content ourselves with establishing uniform pointwise bounds on the term in brackets. Note that

0+σµ1/2η0| ≈µ.

The effect of∂σ2 is to bring out factors ofµ±1/2, and to differentiateqtwice. Ifqis differentiated at most once inx0, then the bounds

|∂x0ξ0q(x, ξ0)|.1 and |∂ξ20q(x, ξ0)|.µ−1 for|ξ0| ≈µ,

yield bounds of size 1 on the term. If q is differentiated twice in x0, then by (4.1) we have the bounds, for|ξ0|≈µ,

µ−1|∂x20q(x+σµ−1/2y0, ξ0)|.c0+c0µ1/2θhµ1/2x2+σy2i−3 .1+c0µ1/2θhµ1/2x2i−3hy2i3.

The rapid decrease ofg(y0) absorbs the term hy2i3, leading to the desired bounds.

We now take the wave packet transform of the solutionu(x) with respect to thex0 variables, and introduce the notation ˜u(x, ξ0)=(Tµu)(x, ξ0). The functionsFe(x, ξ0) and G(x, ξe 0) in the next lemma, though, include terms in addition to the transforms of F andGof Theorem 4.1. LetSedenote the (x1, x0, ξ0) slab [0, ε]×R4=S×R2ξ0.

Lemma4.4. Under the above conditions, we may write

(d1−dξ0q(x, ξ0)·dx0+dx0q(x, ξ0)·dξ0)˜u(x, ξ0) =Fe(x, ξ0)+G(x, ξe 0), where

kFekL1L2(S)e−1/4θ−1/2khµ1/2x2i2Gke L2(S)e .kukLL2(S)+kFkL1L2(S)

1/4θ1/2khµ1/2x2i−1ukL2(S)

−1/4θ−1/2khµ1/2x2i2GkL2(S). (4.4)

Références

Documents relatifs

Abstract Let M be a compact n-dimensional Riemannian manifold with nonnega- tive Ricci curvature and mean convex boundary ∂M.. Applying this result to the universal cover M, ˜ we

En déduire que tout nombre entier impair est la différence de deux

Characteristic boundary conditions for the Navier-Stokes equations (NSCBC) are implemented for the first time with discontinuous spectral methods, namely the spectral di ff erence

The goal of this section is to prove Theorem 3.5 and to give some examples of the existence of solutions to the Brenier-Schr¨odinger problem.. Theorem 3.5 describes a relation

RUPP Masters in Mathematics Program: Number Theory.. Problem Set

Surprisingly, the case of measures concentrated on the boundary of a strictly convex domain turns out to be easier. We can indeed prove in some cases that any optimal γ in this case

First, the estimates of Smith and Sogge [24] for spectral clusters turned out to be strong enough to prove certain Strichartz estimates for the linear wave equations with

We then prove Strichartz estimates for the frac- tional Schr¨ odinger and wave equations on compact Riemannian manifolds without boundary (M, g). This result extends the