• Aucun résultat trouvé

RIESZ TRANSFORM AND L

N/A
N/A
Protected

Academic year: 2022

Partager "RIESZ TRANSFORM AND L"

Copied!
27
0
0

Texte intégral

(1)

RIESZ TRANSFORM AND L COHOMOLOGY FOR MANIFOLDS WITH EUCLIDEAN ENDS

GILLES CARRON, THIERRY COULHON, AND ANDREW HASSELL

Abstract. LetM be a smooth Riemannian manifold which is the union of a compact part and a finite number of Euclidean ends,Rn\B(0, R) for some R >0, each of which carries the standard metric. Our main result is that the Riesz transform onMis bounded fromLp(M)Lp(M;TM) for 1< p < n and unbounded forpnif there is more than one end. It follows from known results that in such a case the Riesz transform onMis bounded for 1< p2 and unbounded forp > n; the result is new for 2< pn. We also give some heat kernel estimates on such manifolds.

We then consider the implications of boundedness of the Riesz transform inLpfor somep >2 for a more general class of manifolds. Assume thatM is an-dimensional complete manifold satisfying the Nash inequality and with anO(rn) upper bound on the volume growth of geodesic balls. We show that boundedness of the Riesz transform onLpfor somep >2 implies a Hodge-de Rham interpretation of theLpcohomology in degree 1, and that the map from L2 toLpcohomology in this degree is injective.

1. Introduction

Let M be a complete Riemannian manifold with infinite measure. The Riesz transformT onM is the operator

(1.1) f →d∆−1/2f,

where ∆ is the positive Laplace operator onM. The Riesz transform is always a bounded map fromL2(M) to L2(M;TM). It is of interest to figure out the range ofpfor whichT extends to a bounded mapLp(M)→Lp(M;TM). Equivalently, we can ask whether

kdfkp≤ k∆1/2fkpfor allf ∈Cc(M).

It has been shown in [14] that the Riesz transform is bounded onLp for 1< p <2 as soon as the manifold satisfies the doubling property as well as a natural heat kernel upper bound. The situation for p >2 is more complicated: there is some understanding of what happens in the more restricted class of manifolds satisfying upper and lower Gaussian estimates for the heat kernel (see [5], [4]), and it is also known that the above more general assumptions do not imply the boundedness of the Riesz transform for allp >2. The counterexample is simply the connected sum of two copies of Rn, where one easily sees that the Riesz transform is unbounded

Key words and phrases. Riesz transform,Lpcohomology.

T.C. is supported in part by the European Commission (IHP Network “Harmonic Analysis and Related Problems” 2002-2006, Contract HPRN-CT-2001-00273-HARP). He also acknowledges for this work the support of Macquarie University and Australian National University.

A.H. is supported in part by an Australian Research Council Fellowship and acknowledges the support of Universit´e de Nantes.

1

(2)

for p > n (see [14]), precisely because the boundedness for suchpwould imply a lower Gaussian heat kernel estimate on this manifold, which is known to be false.

The first aim of the present article is to find out what happens in the remaining range 2< p≤n, and to treat more generally by the same token manifolds with a finite number of Euclidean ends. Our second aim is to give, in a somewhat more general class of manifolds, a cohomological consequence of the boundedness of the Riesz transform for some p > 2, which explains the unboundedness in the range p≥nin the examples just mentioned.

On Rn, it is a classical result in harmonic analysis that the Riesz transform is bounded onLpfor all 1< p <∞. In this case, the kernel of the Riesz transform is given by a multiple of

d 1

|z−z|n−1 =c(z−z)idzi

|z−z|n+1 .

(We shall usez for a Euclidean coordinate throughout this paper, while the prime denotes the ‘right variable’ of a kernelK(z, z); also, ˆzdenotes |z|z and we writer, r for|z|,|z|.) It is useful to compactifyRn to a compact manifold with boundaryZ by adding a sphere at infinity, and using|z|−1as a boundary defining function1for it. If we consider this kernel of ∆−1/2 near the ‘right boundary’ ofZ×Z, namely Z×∂Z, where|z| → ∞, we see that it has an expansion given by

1

|z−z|n−1 = 1

|z|n−1 1−2z·zˆ

|z| + |z|2

|z|2 n−12

∼ 1

|z|n−1 + (n−1)z·zˆ

|z|n +. . . The leading power in this expansion is|z|−n+1and the coefficient multiplying it is 1. This kernel, by itself, is not bounded onLp forp≥nbecause the decay of the kernel puts it only inLn/(n−1)+ǫ, ǫ >0, as a function ofz withz fixed, so it can only be boundedly paired with elements inLpforp < n. However, when we apply a z-derivative, the leading term is killed (sinced1 = 0) and the kernel ofT decays at one order better, namely|z|−n. This allows pairing with elements ofLpfor any p <∞.

LetM be a manifold with Euclidean ends, and assume the number of ends is at least two. It is now relatively easy to explain why the Riesz transform onM is not bounded for p≥n. For simplicity we shall assume here that M has exactly two Euclidean ends. Let us compactify M to a compact manifold M in the analogous way to Z above. The boundary of M is then the disjoint union of two (n−1)- spheres, which we shall denote ∂M+ and ∂M. It turns out that the kernel of

−1/2 has a similar expansion atM ×∂M, of the form (1.2)

X

j=n−1

|z|−jaj(z), aj∈ C(M ×∂M),

but it is no longer true that the leading terman−1= 1. Rather,an−1is the harmonic function on M that equals 1 on ∂M+, and zero on ∂M. As a consequence, applying the derivative operator in the left variable z does not make this leading term disappear, sincedan−16= 0. Hence the kernel ofT only decays to ordern−1 at the right boundary ofM×M, and therefore can only be paired boundedly with elements ofLp forp < n.

1This means that smooth functions near the boundary of the compactification are given pre- cisely by smooth functions of ˆzand 1/|z|.

(3)

In this paper we shall prove

Theorem 1.1. LetM be a completeCRiemannian manifold of dimension n≥3 which is the union of a compact part and a finite number of Euclidean ends. Then the Riesz transform is bounded fromLp(M)toLp(M;TM)for 1< p < n, and is unbounded on Lp for all other values ofpif the number of ends is at least two.

Our method is to analyze the kernel ∆−1/2 based on the formula

(1.3) ∆−1/2= 2

π Z

0

(∆ +k2)−1dk.

Since M is a manifold which is conic at infinity, the Laplacian ∆ on M lies in the class of scattering differential operators [27] and we can use methods from the scattering calculus to analyze the kernel of ∆−1/2. We shall analyze the kernel of

−1/2rather precisely and work out the leading term in the expansion at the right boundary. From this, it will be straightforward to analyze the kernel ofT and to prove the theorem.

The plan of the paper is as follows. We briefly describe the scattering calculus in section 2. In sections 3 – 5 we prove the theorem using the analysis of the resolvent of the Laplacian on asymptotically conic spaces in [23] as a model. We give some large time asymptotics on derivatives of the heat kernel on manifolds with Euclidean ends in section 6.

In section 7 we change point of view and consider a much more general class of manifolds, namely complete manifolds M of dimension n satisfying the Nash inequality and with a uniform upper boundO(rn) on the volume of geodesic balls of radiusr. Weassumethat the Riesz transform onM is bounded onLp for some p >2 and give several geometric and topological consequences: a Hodge-de Rham interpretation of theLp cohomology ofM (Proposition 7.3), injectivity of the map from L2 to Lp cohomology (Lemma 7.4) and derive a contradiction if p > n and M has at least two ends (Corollary 7.5), thus generalizing the unboundedness part of Theorem 1.1 for p > n to this larger class of manifolds. In the final section we discuss our results in the context of previously known examples and pose some open problems.

Acknowledgements. The final version of this paper is an improvement on the original version thanks to the efforts of two anonymous referees who found numerous typographical errors, as well as pointing out that the original proof of Lemma 3.2 was wrong. We thank them for their careful reading of the manuscript.

2. Scattering Calculus

As noted above, we shall use the scattering calculus [27] to analyze the kernel of ∆−1/2 on manifolds with several Euclidean ends. The scattering calculus is expressed in terms of compactifications on M and, especially, of the double space M2 which carries the kernel of the resolvent of ∆ and of the operator ∆−1/2. The spaceM is compactified by adding a sphere at infinitySn−1for each Euclidean end, and declaringr−1 = 1/|z|and ˆz =z/|z| to be local coordinates near a boundary point; in particular,r−1 is taken to be a defining function for the boundary (which we shall sometimes refer to as ‘infinity’). We sometimes usex=r−1to denote this boundary defining function andy= ˆzto denote boundary coordinates, extending to a collar neighbourhood of the boundary, as is customary when using the scattering

(4)

calculus. The metric takes the form

dr2+r2h(y, dy) = dx2

x4 +h(y, dy) x2

at each end, and is therefore ascattering metricas defined in [27] (of a particularly simple form, being an exact conic metric near infinity).

We are mostly interested in the case when the number of ends is at least two. In fact, for the sake of clear exposition we shall assume from now on that the number of ends is exactly two, although all proofs in this paper generalize in an obvious way to any finite number of ends. We shall label these ends + and −, thus for example we shall usez+ as the Euclidean variable on the positive end, andz for the variable on the negative end; when it is not necessary to stipulate which end is being considered, we shall just usez.

It is not so obvious which compactification ofM2is most appropriate for dealing with the Schwartz kernels of operators such as the Laplacian, or functions of the Laplacian, on manifolds with Euclidean ends. There are several different asymptotic regimes of interest when dealing with such kernels. One regime, the ‘near-diagonal’

regime, is when the two variablesz, zremain a finite distance apart as they both go to infinity. Another is when they both go to infinity with the ratior/r approaching a limit and with ˆz,zˆ both approaching a limit. Finally there is the case that one variable approaches a limit, while the other remains fixed. The kernel has different behaviour in each of these asymptotic regimes, so they need to be represented by distinct parts of the boundary of the compactification. The space (M)2 is thus too

‘small’ a compactification ofM2for our purposes, since it only has the third regime distinguished; the first two are squashed into the corner.

It turns out that there is a space denoted M2sc, the ‘scattering double space’, which satisfies these criteria. It is obtained by performing two blowups on M2. The first is blowing up the corner (∂M)2, creating the so-called b-double space, and the second is blowing up the boundary of the diagonal (which lifts to the b- double space to be transverse to the boundary, hence this blowup is well-defined).

Each asymptotic regime is represented by a boundary hypersurface of M2sc. The first, ‘near-diagonal’ regime is represented by the boundary hypersuface created by the second blowup, denoted sf for ‘scattering face’; the second is represented by the boundary hypersurface created by the first blowup, denoted bf for ‘b-face’

(since it is present in the b-calculus) and the third regime is represented by the two boundary hypersurfaces∂M ×M, M ×∂M of (M)2, denoted lb and rb for ‘left boundary’ and ‘right boundary’. Note that when M has k ends, then bf has k2 components and sf, lb and rb each havekcomponents.

The structure of sf and bf is as follows. Each component of sf is naturally diffeomorphic toRn×Sn−1, andz−zand ˆzare coordinates on the interior of each component of sf. Each component of bf is naturally isomorphic to a blowup of the spaceSn−1×Sn−1×[0,1], with coordinates (ˆz,ˆz,|z|(|z|+|z|)−1); the blowup is of the submanifold{zˆ= ˆz,|z|(|z|+|z|)−1= 1/2}which corresponds to the boundary of the diagonal.

The set ofscattering vector fieldsonM consists of all smooth vector fields onM of bounded length with respect to the scattering metric. Equivalently, they can be described as vector fieldsV of the formV =xW, wherexis a boundary defining function for M and W is a smooth vector field which is tangent to the boundary ofM. For a manifold with Euclidean ends, every scattering vector fieldV has the

(5)

lb

rb bf

sf

diagsc

Figure 1. The double scattering space

form Paizi on the ends (with respect to Euclidean coordinates zi on the end), where ai∈ C(M). In particular, the partial derivatives∂zi are scattering vector fields. The description V = xW shows that scattering vector fields V map any conormal space on M to itself, and indeed improve the order of vanishing at the boundary by 1; for example, V(xαC(M)) ⊂xα+1C(M). Similarly, acting ‘on the left factor’ on functions onM2sc, we have

(2.1) V β· C(M2sc)

⊂ρlbρbfβ· C(M2sc)

whereβ is any product of powers of boundary defining functions onM2sc.

Thescattering calculusis an algebra of pseudodifferential operators onM which is defined by the properties of their Schwartz kernels. Namely, A is a scattering psuedodifferential operator of ordermonM iff the kernel ofA, when lifted toM2sc, is conormal2of ordermat the diagonal ofM2sc smoothly up to the boundary sf, is smooth elsewhere at sf, and is rapidly decreasing at bf, lb and rb. All scattering vector fields are in the scattering calculus, together with all the differential operators they generate over C(M), such as the Laplace operator. The resolvent of the Laplacian (∆−λ2)−1 is a scattering pseudodifferential operator of order−2 onM for Reλ6= 0. In fact, the structure of the resolvent on the spectrum, i.e. the kernel of (∆−(λ±i0)2)−1 for real λ, can also be described on M2sc, although here the kernel is no longer rapidly decreasing at bf, lb and rb, rather it is a ‘Legendrian distribution’ [23]. Our approach is partly modelled on the analysis in this paper.

However, we take advantage of the assumption here that M has exact Euclidean ends, which leads to great simplifications over the analysis of [23] since we can exploit the well-known explicit formulae for the resolvent of the Laplacian on Rn and use these as ingredients for a parametrix of the resolvent kernel onM, thereby avoiding the need to use Legendrian distributions in this paper.

2In other words, the kernel ofAhas a singularity at the diagonal characteristic of pseudodif- ferential operators of orderm, and this holds smoothly up to the boundary, in the sense that it could be extended across the boundary as a conormal distribution. See [24], section 18.2, for the precise definition.

(6)

Notation. We writezfor a Euclidean variablez= (z1, z2, . . . , zn)∈Rnand write hzi=p

1 +|z|2, while xis used for|z|−1, or, sometimes, where more convenient, forhzi−1. For a manifold with cornersX, we write ˙C(X) for the space of smooth functions which vanish to infinite order at the boundary of X. We use notation [X;S1, S2, . . . Sn] to denote the blowup ofX at the submanifoldsS1,S2, . . . (in that order). For any boundary hypersurface H on a manifold with cornersX we use the notation ρH to denote a boundary defining function forH, that is, a smooth nonnegative function onX that vanishes exactly to first order atH.

3. Parametrix construction

To analyze the operator ∆−1/2 we return to the formula (1.3). We first observe that the off-diagonal terms in the kernel of ∆−1/2 come from a neighbourhood of zero in the integral (1.3). Indeed, let s0(k) be a cutoff function equal to 1 in a neighbhourhood of k = 0 and equal to zero outside a compact set, and let s1(k) = 1−s0(k). Then we may insert the factor 1 =s0(k) +s1(k) into the integral (1.3). With the factorsi inserted, the integral gives a functiongi(∆) of ∆, i= 0 or 1, where

g1(t) = Z

0

s1(k) 1 k2+t2dk

is easily checked to be a classical symbol of order−1. By the symbolic functional calculus [22], this term is a scattering pseudodifferential operator of order−1, hence dg1(∆) is a scattering pseudodifferential operator of order zero. It is therefore bounded onLpfor all 1< p <∞[30]. So we are reduced to studyingdg0(∆), given by the integral (1.3) with factors0(k) inserted.

We shall write down a fairly explicit parametrix for (∆ +k2)−1 for smallk. In doing so, we need to consider the different asymptotics that this kernel takes when k = 0 and k 6= 0. Indeed, onRn the kernel decays as |z−z|−(n−1)/2 for k 6= 0 and |z−z|−n+2 for k = 0, as |z−z| → ∞, which (except when n = 3) is a different rate. This can be encoded geometrically by blowing up at the boundary whenk= 0. Consider the space

(3.1) M2sc,k= [M2sc×[0, k0]; bf×{0}; lb×{0}; rb×{0}].

We shall denote the boundary hypersurfaces which are the lifts of bf×[0, k0], lb×[0, k0] and rb×[0, k0] to M2sc,k by bf, lb and rb, and sf×[0, k0] by sf; this is of course an abuse of notation, but in the context it will always be clear whether it is a boundary hypersurface ofM2sc or M2sc,k that is referred to. We shall denote the new boundary hypersurfaces corresponding to the three blowups by bf0, lb0 and rb0, according as they arise from the first, second or third blowups in (3.1) respectively, and we shall denoteM2sc× {0} by zf, for ‘zero face’. We also define

sc,k to be ∆sc×[0, k0] ⊂M2sc,k. Letχ be a smooth function on M2sc,k which is equal to one in a neighbourhood of ∆sc,k, and whose support meets the boundary ofM2sc,k only at sf and zf.

We recall the well-known expression for the resolvent kernel (∆ +k2)−1 onRn forn≥3:

(3.2) (∆ +k2)−1= e−k|z−z|

|z−z|n−2fn(k|z−z|),

(7)

rb bf

rb bf

bf0 sf

sf

rb

rb0 bf0

bf zf

Figure 2. Blowing upM2sc×[0, k0] to produceM2sc,k

where fn(t) is symbolic of order (n−3)/2 ast → ∞, while it is O(1) and has a classical expansion in powers and logarithms as t→0. In fact,fn is a polynomial of order (n−3)/2 whenn≥3 is odd. It is straightforward to check

Lemma 3.1. Let Z be the compactification of Rn. Then the resolvent kernel(∆ + k2)−1 is such that χ(∆ +k2)−1 is conormal at∆sc,k, and

(3.3) (1−χ)(∆ +k2)−1∈ρ0sfbfρlbρrb)bf0ρlb0ρrb0)n−2C(M2sc,k).

(8)

zf ff

bb k

{ k=0 } x {x=0}

Figure 3. Blowing upM×[0, k0] to produceMk

Here, ‘conormal to∆sc,k’ means that the kernel is conormal inz−z which defines

sc,k and smooth in the remaining variables x = |z|−1,z, k, uniformly up to theˆ boundary.

For example, let us check the statement of the lemma near the triple intersection bf0∩rb0∩rb. Coordinates near this codimension three corner are ˆz,zˆand bound- ary defining functions ρrb0 =k/x=k|z|for rb0,ρrb=x/k= 1/(k|z|) for rb and ρbf0=xfor bf0. Near this corner,|z|is much larger than |z| so we may expand

k|z−z|=k|z| 1−2z·zˆ

|z| + |z|2

|z|2 1/2

= 1 ρrb

1−2ˆz·zˆρrbρrb0+ (ρrbρrb0)21/2

, from which it is easy to check that (3.3) holds.

We also need a single space version of this space. Let (3.4) Mk = [M×[0, k0];∂M× {0}].

Denote the boundary hypersurfaces bb, zf and ff which arise from ∂M ×[0, k0], M × {0} and from the blowup, respectively, and denote corresponding boundary defining functions by ρbb, ρzf and ρff. Again, it will always be clear in context whether zf refers to the zero-face ofMk or M2sc,k. We have

Lemma 3.2. Let v∈C˙(M). Then there is a function

u∈e−1/ρbbρ(n−1)/2bb ρn−2ff C(Mk) +e−1/ρbbρ(n−1)/2bb ρn−2ff ρn−2zf logρzf C(Mk) such that (∆ +k2)u|zf is equal to v, and (∆ +k2)u vanishes to infinite order at both ff andbb.

Proof. We first use results from [29] to show that we can solve ∆f =v onM. The Laplacian on an asymptotically Euclidean manifoldM may be written in the form

∆ =xn/2+1P x−n/2+1,

where P is an elliptic b-differential operator on M. A short computation shows that near infinity,P takes the form

−(x∂x)2+ n−2 2

2

+ ∆Sn1, x=|z|−1,

where ∆Sn−1 is the standard Laplacian on the (n−1)-sphere. This is a strictly positive operator, soP is ‘totally elliptic’, and hence is Fredholm acting between

(9)

the b-Sobolev spaces3Hb2(M)→L2b(M). Thus ∆ itself is Fredholm acting between xn/2−1Hb2(M)→xn/2+1L2b(M). AlsoP is self-adjoint with respect to the measure induced bygb, so its index acting betweenHb2(M)→L2b(M) is equal to zero, hence it is invertible if and only if its null space is trivial. This is therefore also true for

∆ : xn/2−1Hb2(M)→ xn/2+1L2b(M). It is also shown in [29], section 5.25, that if P g∈C˙(M), with g in L2b(M), then g is conormal at x= 0 with an asymptotic expansion of the form

g∼X

j

xn/2−1+jajφj(ˆz),

where φj is a spherical harmonic with eigenvaluej(j+n−2) : ∆Sn−1φj =j(j+ n−2)φj. Therefore, iff =xn/2−1gand ∆f = 0, thenf is conormal atx= 0 with an expansion

(3.5) f ∼X

j

xn−2+jajφj(ˆz).

In particular, such a function tends to zero at infinity. By the maximum prin- ciple, a harmonic function on M tending to zero at infinity vanishes identically.

Therefore there is no nontrivial solution to ∆f = 0, with f ∈ xn/2−1Hb2(M).

Hence ∆ :xn/2−1Hb2(M)→xn/2+1L2b(M) is invertible, and we can solve ∆f =v, v∈C˙(M)⊂xn/2+1L2b(M), wheref ∈xn/2−1Hb2(M) has an expansion (3.5).

Let f be as in the previous paragraph. We first find a formal expansion for u near the corner zf∩ff ofMk. Coordinates near this corner arex, y andκ=k/x.

Let us look for an expansion foruof the form

(3.6) u=X

j

xn−2+jφjaj(κ), κ=k x,

whereaj(0) is given by the expansion (3.5) forf, so that u|zf =f. The operator

∆ +k2 may be written

(3.7) −(x2x)2+ (n−1)x3x+x2Sn−1+k2.

Here ∂x indicates the derivative keepingk fixed. When we switch to using coordi- nates (x, y, κ), then we must replacex∂x byx∂x−κ∂κ, getting

(3.8) x2

−(x∂x−κ∂κ)2+ (n−2)(x∂x−κ∂κ) + ∆Sn−12 .

Acting on thejth term, we may replace ∆Sn−1 byj(j+n−2) andx∂xby (j+n−2), getting the operator

(3.9) Rj ≡x2

−(κ∂κ)2+ (n−2 + 2j)(κ∂κ) +κ2 .

We want to solveRj(aj(κ)) = 0. The equation is a regular singular ODE atκ= 0, and the indicial roots are 0 and n−2 + 2j; since n ≥ 3 and j ≥0, the second indicial root is positive and therefore there is a solution which is continuous up to κ= 0 for every j. Thus we can match the solution to the initial condition aj(0) determined by the coefficient in (3.5). The solution will then be smooth nearκ= 0

3Here L2b(M) is the L2-space with respect to the b-metric gb = x2g; thus L2b(M) = hzin/2L2(M). Also Hb2(M) is the b-Sobolev space of order 2, defined as the set of functions gL2b(M) such thatQgL2b(M) for all b-differential operators of order 2 onM (near infinity, such operators take the formhzi2P

i,jaijzizj, withaij∈ C(M)).

(10)

forn odd, while for evennthe solution will have an expansion in integral powers ofκtogether with log termsκn−2+2j+2klogκfork∈N.

Because the second indicial root is positive, the coefficient of κn−2+2j is not determined by the above considerations. We need to choose the solution so that it has exponential decay at the bb face. To do this, we let aj(κ) = κβj˜aj(κ) where βj= (n−2 + 2j)/2. The equation for ˜aj(κ) is

(3.10) h

κ2κ2+κ∂κ−(κ2j2)i

˜

aj(κ) = 0.

This is a modified Bessel equation and a particular solution which decays asκ→ ∞ is the Macdonald functionKβj(κ) (see [1], section 9.6)4. If we take ˜aj(κ) to be a suitable constant times Kβj(κ), then aj(κ) = κβj˜aj(κ) has the correct behaviour at κ= 0 and as κ→ ∞ has the form e−κκn/2−3/2+j times a smooth function of κ−1.

We now wish to Borel sum the seriesP

xn−2+jaj(κ)φj(ˆz) at the face ff. This series is Borel summable forκ≤2, say, since the functionsaj(κ)φj(ˆz) lie in a fixed Fr´echet space asj→ ∞, namely, the spaceC([0,2]×Sn−1)+κn−2logκC([0,2]× Xn−1). For κlarge, say κ > 1, we write the series in terms of coordinatesk and ρ= 1/κwhich are valid near the corner ff∩bb. We get a series of the form

X

j=0

kn−2+je−1/ρρ(n−1)/2bj(ρ)φj(ˆz),

where bj is smooth in ρ. This is also Borel summable, and moreover the sum will (for ρ <1) be of the form e−1/ρρ(n−1)/2 times a smooth function on Mk, in particular vanishing to infinite order at bb.

Letu1be a Borel sum of the series. Thenu1is in the space given in the statement of the lemma, (∆ +k2)u1 vanishes to infinite order at ff, andu1|zf−f vanishes to infinite order at zf∩ff. Let u2 be a function on Mk equal tof −u1|zf at zf and vanishing to infinite order at ff and bb. Thenu1+u2satisfies the conditions of the

lemma.

Remark 3.3. The conclusion of the lemma holds if we letkvary along a rayk=sγ, where|argγ|< π/2, ands≥0, except that we must replacee−1/ρbb bye−γ/ρbb.

We now use this lemma to define a harmonic function onM which will be key to the parametrix construction. We begin by choosing a smooth functionφ:Rn→R which is equal to 1 for|z|large and is supported in{|z|> R}. Using this we define functions φ± on M, with φ± supported on the± end of M, in the obvious way.

Then let u± be the function given by Lemma 3.2 from the function v = −∆φ±. Then

(3.11)

Φ±±+u±|zf is a harmonic function equal to 1 at∂M± and 0 at∂M. We now define our parametrix. It is based on the resolvent kernel for Rn, but there is a crucial additional term (G3 below) which corrects the leading order co- efficient of the kernel at the face rb0 (see the discussion of this coefficient in the Introduction). We now write φ± to denote this function of the left variable on M2 and φ± denote this function of the right variable on M2. Let Gint(k) be a parametrix, modulo smoothing operators, for (∆ +k2)−1 in the interior ofM. We

4Kβ0(κ) is related tofn(κ) in (3.2) byKβ0(κ) =eκκ(n2)/2fn(κ).

(11)

may assume that it is localized sufficiently close to the diagonal. We recall that the resolvent of the Laplacian onRn has the form (3.2). Using this notation we define

(3.12)

G(k) =˜ G1(k) +G2(k) +G3(k), where G1(k) = e−k|z−z|

|z−z|n−2fn(k|z−z|)

φ+φ+φ G2(k) =Gint(k)

1−φ+φ+−φφ G3(k) = e−k|z|

|z|n−2fn(k|z|)

u+φ++uφ

.

Remark 3.4. We have abused notation somewhat in the formula for G1(k). The factor φ+φ+φ means that it is localized to the region where either both variables are in the + end, or both variables are in the−end of M. In either case we may use the Euclidean coordinatez∈Rn in both the left and right variable; the formula is written in terms of this variable. (Thus, a given pair (z, z) corresponds in the formula forG1(k) totwopoints, one on the ++ end and one on the−−end ofM2.)

Remark 3.5. Scattering vector fields (such as∂zi near the ends onM — see Sec- tion 2) act nicely on conormal spaces onM2sc,k. For example, if V is a scattering vector field acting on the left factor and β is any product of powers of defining functions onM2sc,k, we have analogously to (2.1)

(3.13) V β· C(M2sc,k)

⊂ρlbρlb0ρbfρbf0β· C(M2sc,k).

This, and the fact that our parametrix is conormal on M2sc,k, makes it very easy to estimate the size of the error terms from applying the operator ∆ +k2 to our parametrix.

Remark 3.6. Notice that G2(k) is supported away from every boundary hypersur- face ofM2sc,kexcept for zf. Also, thanks to Lemma 3.2 applied tou±, (∆+k2)G3(k) is rapidly vanishing at all boundaries except rb0 and zf.

4. Error term and resolvent

In this section, we correct the parametrix to the exact resolvent. The main point is that we obtain information about the regularity of the kernel of the error term on M2sc,k, and therefore of the resolvent itself on this space. This allows us to determine the regularity of ∆−1/2 and d∆−1/2 and compute its behaviour to leading order at the boundary hypersurfaces ofM2sc.

Applying ∆ +k2to our parametrix (on the left), we get

(4.1) (∆ +k2) ˜G(k)≡(∆ +k2)(G1(k) +G2(k) +G3(k)) = Id + ˜E(k),

where this equation defines ˜E(k). We may think of ˜E(k) either as a kernel onM2sc,k or, by restricting to M2×[0, k0] which is the interior of M2sc,k together with the interior of zf, a family of kernels parametrized bykacting on functions onM.

By construction, the complete symbol of the diagonal singularity of ˜G(k) is the inverse, modulo symbols of order −∞, of the complete symbol of ∆ +k2. Thus E(k) is smooth at the diagonal. Also, we see due to the properties of˜ u±that ˜E(k)

(12)

vanishes to infinite order at bf,bf0,lb, rb and lb0. The crucial property of ˜E(k) is the order of vanishing at rb0. To calculate this, we need to determine the leading coefficient of the expansion of ˜G(k) at rb0. These terms come fromG1(k) andG3(k).

Since|z−z|=|z| −z·zˆ+O(ρrb0ρ−1bf0) at rb0, we havek|z−z|=k|z|+O(ρrb0);

note thatk|z|vanishes on rb0, whilek|z|is finite in the interior of rb0. Hence (4.2) G1(k) = e−k|z|

|z|n−2fn(k|z|)φ±(z) +O(ρn−1rb0 ) at rb0.

If we combine this withG3(k), then using (3.11) we see that the leading coefficient becomes

e−k|z|

|z|n−2fn(k|z|)Φ±(z) +O(ρn−1rb0 ) =ρn−2rb0 e−k|z|fn(k|z|)Φ±(z) +O(ρn−1rb0 ) at rb0. The leading term annihilated by the operator ∆ +k2 (since Φ± is harmonic and k = 0 at rb0), so the error term ˜E(k) is O(ρn−1rb0 ) at rb0 — an improvement of one order over what might be expected, and the main point of introducing the correction termG3(k). Thus, we have

(4.3) E(k)˜ ∈ρsfρbfρbf0ρlbρlb0ρrbρn−1rb0 L(M2sc,k).

Note that both x = hzi−1 and x = hzi−1 are smooth on M2sc,k, the former function vanishing simply at lb,lb0,bf,bf0,sf and the latter vanishing simply at rb,rb0,bf,bf0,sf. Thus (4.3) implies that the kernel of ˜E(k) is (hzihzi)1−ntimes a bounded function onM2sc,k. This implies that ˜E(k) is Hilbert-Schmidt, uniformly for k ∈ [0, k0], hence compact for each k ∈ [0, k0]. Therefore Id + ˜E(0) has finite dimensional null space and cokernel of the same dimension on L2(M). We next show that we can modify our parametrix by the addition of a finite rank term so that the new error term is invertible for smallk. The correction term will be

(4.4) G4(k) =

N

X

i=1

φii,·i N = dim (Id + ˜E(0))

where the φi, ψi are in ˙C(M) and independent of k. Since ˜E(0) maps into C˙(M), the null space is contained in ˙C(M). Thus we chooseψi to span the null space of Id + ˜E(0), and we would like to choose φi so that ∆φi span a space supplementary to the range of Id + ˜E(0). This is possible since ∆ has trivialL2- null space (as proved in Lemma 3.2), and ∆ is self-adjoint, hence the range of ∆ on C˙(M) is dense in L2(M). Choosing suchφi, we define G4(k) (which is actually independent ofk) by (4.4). We now define

G(k) =G1(k) +G2(k) +G3(k) +G4(k) = ˜G(k) +G4(k) and defineE(k) by setting

(4.5) (∆ +k2)G(k) = Id +E(k) =⇒ E(k) = ˜E(k) + (∆ +k2)G4(k);

E(k) enjoys all the properties of ˜E(k) listed above. In addition, sinceE(0) is such that Id +E(0) is invertible, it follows that actually Id +E(k) is invertible for all sufficiently smallk; we assume thatk0>0 is chosen so that Id +E(k) is invertible for allk∈[0, k0].

We now analyze the inverse of Id +E(k). Let us write (Id +E(k))−1= Id +S(k),

(13)

where this equation defines S(k). Since E(k) is Hilbert-Schmidt, uniformly on L2(M), S(k) is also uniformly Hilbert-Schmidt. The regularity (4.3) of ˜E(k) on M2sc,k, and the fact thatx/(x+k)∈ C(M2sc,k) vanishes simply at rb,bf,sf, imply that

(4.6) E(k)∈xN(x)n−1 x x+k

N

L(M2×[0, k0]) for allN.

Using this and the formula

S(k) =−E(k) +E(k)2+E(k)S(k)E(k) shows that

(4.7) S(k)∈xN(x)n−1 x x+k

N

L(M2×[0, k0]) for allN.

We are particularly interested in the kernelG(k)S(k), which we shall callG5(k), since the addition ofG5(k) will correct the parametrixG(k) to the exact resolvent kernel.

Lemma 4.1. Let l= 0,1,2. . .. Then the kernel (4.8)

Z k0

0

s0(k)∇(l)G(k)S(k)dk is in

(4.9) hzi−(n−1+l)hzi−(n−1)L(M2)∩ hzi−(n−2+l)hzi−nL(M2).

Remark 4.2.Much more precise statements can be made about the kernels in (4.3), (4.7) and (4.8), for example by using Melrose’s Pushforward Theorem [28], which shows that these kernels are actually conormal, with respect to the boundary and the diagonal, on M2sc. However, the L statements will suffice for our purposes and are more straightforward to prove.

Proof. Let us break upG(k) into two partsG(k) =χG(k) + (1−χ)G(k), whereχ is as in Lemma 3.1. ThusχG(k) is a smooth family of scattering pseudodifferential operators, while (1−χ)G(k) has no singularity at the diagonal.

We first consider (1−χ)G(k) which is localized away from the diagonal. Letmx

denote the multiplication operator byx=hzi−1 onM. Then we have

(l)(1−χ)G(k)S(k) = ∇(l)(1−χ)G(k)mn+lx

m−(n+l)x S(k) .

By (3.13), the kernel∇(l)G(k) decays to ordern−2 +lat lb0 and to order∞at lb and bf. If we multiply this kernel byhzi−(n+l), which corresponds to composing withmn+lx on the right, then it also decays to ordern+lat bf0 and sf. This means that we can write

(4.10) ∇(l)(1−χ)G(k)mn+lx ∈xn−2+l x x+k

2

L(M2×[0, k0]), x=hzi−1 sincexis a product of boundary defining functions for lb,lb0,bf,bf0,sf, andx/x+k vanishes to first order at lb and bf. In a similar way, using (4.7), we find that (4.11) m−(n+l)x S(k)∈(x)n−1 x

x+k 2

L1(M;L(M×[0, k0]));

note that composing with m−(n+l)x on the left is harmless here because the kernel S(k) vanishes to infinite order on every boundary hypersurface wherex−(n+l)blows

(14)

up. For this same reason the kernel isL1in the left variable, uniformly in the right variable and ink. It follows that the composition

(4.12) ∇(l)(1−χ)G(k)S(k)∈xn−2+l(x)n−1 x x+k

2 x x+k

2

L(M2×[0, k0]).

Now we integrate ink. If we ignore the (x/x+k)2 factor (which is bounded) then we find that

Z k0

0

s0(k)∇(l)(1−χ)G(k)S(k)dk∈xn−2+l(x)nL(M2), because

Z k0

0

x x+k

2

dk≤ Z

0

x x+k

2

dk=x since Z

0

1 1 +k

2

dk= 1.

In exactly the same way we show that Z k0

0

s0(k)∇(l)(1−χ)G(k)S(k)dk∈xn−1+l(x)n−1L(M2).

Finally we consider the integral (4.8) with G(k) replaced by χG(k). We may regard ∇(l)χG(k) as a smooth family of scattering pseudodifferential operators, andS(k) as an element of

(4.13) C˙ M;ρrbρn−1rb0 L(M×[0, k0]) .

To see this, note that anyz-derivative∇(l)z E(k) ofE(k) also lies in the space (4.6), using the explicit formula forGi(k) and Lemma 3.2. Using this and (4.7) we obtain S(k)∈(4.13). Since scattering pseudodifferential operators map ˙C(M) to itself continuously, it follows that∇(l)χG(k)S(k) is also an element of the space (4.13).

Performing the k integral we get an extra vanishing factor at rb⊂M2sc, yielding x(x)nL(M2), which proves the Lemma for this piece. This completes the proof.

5. Riesz Transform

Recall that in Section 3 we split ∆−1/2 = g0(∆) +g1(∆), where dg1(∆) was bounded fromLpto Lp for all 1< p <∞, and

(5.1) g0(∆) = 2

π Z

0

s0(k)(∆ +k2)−1dk.

Hence, it remains to analyzedg0(∆). Let us decompose

(5.2) (∆ +k2)−1=G1(k) +G2(k) +G3(k) +G4(k) +G5(k)

as in the previous section and writedg0(∆) =T1+T2+T3+T4+T5correspondingly.

The easiest kernel to deal with isT4; this kernel is in ˙C(M2), hence is bounded from Lp to Lp for 1 ≤ p≤ ∞. The kernel T2 is bounded onLp for 1 < p < ∞ because it is a classical zero order pseudodifferential operator with proper support;

see Chapter VI, section 5 of [30]. Next considerT5. Lemma 4.1, withl= 1, shows thatT5is inLp(M;Lp(M)), wherep′−1= 1−p−1, for allp∈(1,∞), which implies thatT5 is bounded onLp for 1< p <∞.

(15)

The kernelT1we decompose further as asT1=T1,1+T1,2+T1,3, where

(5.3)

T1,1=

φ+φ+φZ 0

dz

e−k|z−z|fn(k|z−z|)

|z−z|n−2 s0(k)dk, T1,2

(dφ+++ (dφZ 0

e−k|z−z|fn(k|z−z|)

|z−z|n−2 s0(k)dk, T1,3= (1−χ)

(dφ+++ (dφZ 0

e−k|z−z|fn(k|z−z|)

|z−z|n−2 s0(k)dk and χis as defined above Lemma 3.1. Then T1,2, likeT2, is a classical zero order pseudodifferential operator with proper support, hence bounded on allLp. As for T1,1, if we ignore the s0(k) factor (which we may since the difference, replacing s0(k) by 1 is a zeroth order pseudodifferential operator) then the integral gives precisely the Riesz kernel onRn, and the factor (φ+φ+φ) may be regarded as composition with bounded operators. HenceT1,1 is bounded onLpfor 1< p <∞, because the Riesz kernel onRn has this property. Thus we are left withT1,3+T3. Lemma 5.1. The kernelT1,3+T3 is bounded pointwise by

(5.4) Chzi−n+1hzi−n+1

and by

(5.5) Chzi−n+2hzi−n

for someC >0. Moreover, for some cn6= 0we have a bound

(5.6)

T1,3+T3−cn|z|−n+1±

≤Chzi−nhzi−n+2.

Proof. We first note that, as with T1,2, we may ignore the factor s0(k) in the expression above for T1,3 (the difference, with s0(k) replaced by 1, is compactly supported inzand rapidly decreasing inhzi). Thus we consider the integral

(1−χ)

+φ++dφφZ 0

e−k|z−z|fn(k|z−z|)

|z−z|n−2 dk.

The value of this integral is (5.7) cn(1−χ)

+φ++dφφ 1

|z−z|n−1, cn= Z

0

e−κfn(κ)dκ, where cn 6= 0 since fn(κ) > 0 for κ > 0 (see [1], section 9.6). This satisfies the estimates (5.4), (5.5) (note that|dφ±|= 0 forhzilarge) and ashzi → ∞we have

(5.8)

T1,3−cn|z|−n+1±

≤Chzi−nhzi−∞. Now considerT3. The kernel ofT3 is given by

(5.9)

Z 0

e−k|z|fn(k|z|)

|z|n−2

du+φ++duφ

s0(k)dk.

By Lemma 3.2 we may estimate |du±| ≤Chzi−n+1, since dimproves the order of vanishing of smooth functions on Mk at ff by 1. Substituting this into (5.9) and estimating gives (5.4). On the other hand, we can also estimate by Lemma 3.2

|du±| ≤Ce−k|z| k|z|+ 1

|z| n−1

,

(16)

since (k|z|+ 1)/|z|=k+|z|−1 is a boundary defining function for ff⊂Mk. If we use this estimate fordu± and use a uniform bound |e−k|z|fn(k|z|)| ≤C in (5.9) then we end up with (5.5).

To obtain (5.6), definef± =u±|zf, the restriction ofu± to the face zf; we may regardf±as a function onM, constant ink. Then we can estimate by Lemma 3.2

du±−df±

≤Ckhzi−(n−2).

If we substitutedf± fordu± in (5.9), then the integral is equal to cn|z|−n+1df±

with the same cn as in (5.8). This combines with the main term in (5.8) to give the main term in (5.6). The error term, substitutingCkhzi−(n−2)fordu± in (5.9), gives usChzi−nhzi−n+2. This completes the proof.

This lemma implies that, for 1< p < n,T1,3+T3is an element ofLp(M;Lp(M)).

Moreover, forp≥n, this is not true since the function (x)n−1=|z|−n+1is not in Lp then, and forz in a fixed compact set inM we have

T1,3+T3=cn|z|−n+1±(z) +O(|z|−n), |z| → ∞

by Lemma 5.1. The coefficient of (x)n−1 iscn±(z) which does not vanish iden- tically. Therefore T1,3+T3 cannot be applied to any bounded function equal to x(logx)−1 near infinity, which lies in Lp for p ≥n. This completes the proof of Theorem 1.1.

Remark 5.2. If M has one Euclidean end then the same argument shows that the Riesz transform is bounded onLp for all 1< p <∞. In this case, the parametrix G(k) can be taken to be (compare with (3.12))˜

e−k|z−z|

|z−z|n−2fn(k|z−z|)φφ+Gint(k) 1−φφ

+ e−k|z|

|z|n−2fn(k|z|)(1−φ)φ. In this case the role of Φ± in the computation above is played by the constant function 1. The argument is the same as above, except that the gradient of 1 vanishes so that instead of (5.6) we get ahzi−nhzi−n+2bound on the kernelT1,3+T3

itself. This and (5.5) imply the boundedness ofT1,3+T3 on Lp for allp strictly between 1 and∞.

6. Heat kernel

As part of the analysis of the heat kernel we analyzed the structure of the re- solvent (∆ +k2)−1for realk, including an analysis of the asymptotics of its kernel whenk→0. This analysis remains valid for any cone{k=iλ|Imλ≥ǫReλ}for anyǫ >0. We can use this to obtain information about the heat kernelH(t, z, z) ofe−t∆onM via the contour integral

(6.1) e−t∆= 1

2πi Z

Γ

e−tλ2 ∆−λ2−1

2λ dλ where Γ = Γ+∪Γ, Γ±={λ=se±iπ/12|s∈R+}.

Let us focus on the heat kernel in the following asymptotic regime: We fix a point z ∈ M, which we think of as being in the ‘compact part’ ofM (where the metric is not flat), and fix an end ofM and a pointω ∈Sn−1 which we think of as a point at infinity for this end. Consider the behaviour of the heat kernelH(t, z, z)

Références

Documents relatifs

Starting from the current knowledge based on intestinal metagenomics, we point out on the role of food in shaping gut microbiota composition and functions, which is therefore

One can therefore argue that the dynamism of the l- clitics in Romani third person affirmative clauses could account for the higher rates of selection of estar in

The proposi- tion gives in this case the right range of p’s where the Riesz transform is bounded on L p.. Second example: Let M be a connected sum of several (say l ≥ 2) copies of

Based on the obtained results we can conclude that considering complete document does not always increase the classification accuracy5. Instead, the accuracy depends on the nature

The proof of Theorem 0.7 in Section 1 uses methods of the first author in [1] adapted to the present situation and in particular a Calder´on-Zygmund lemma for Sobolev functions,

[ 45 ] 41 10–20 Chiari osteotomy 4.0 (7.2) Improved Harris hip score, improvement of acetabular angle of Sharp, increase of LCE angle, improved femoral head coverage,

In case 3, (D) is again obvious since such a manifold has polynomial volume growth, (DUE) is well-known (it can be extracted from the work of Varopoulos, see for instance [99],

We study the validity of the L p inequality for the Riesz transform when p &gt; 2 and of its reverse inequality when p &lt; 2 on complete Riemannian manifolds under the