lectures on resonances - centre national de la …sjostrand.perso.math.cnrs.fr/coursgbg.pdf · this...

169
Lectures on resonances JohannesSj¨ostrand * 1 Introduction This text is the written version of lectures on resonances given at the Mathe- matics department of the Gothenburg University and Chalmers in the Spring semesters of 2000 to 2002. Some of the first chapters were also given in a graduate course at Ecole Polytechnique. Despite the somewhat preliminary form of these notes we hope they can be of some use as an introduction to the subject, and some guidance for a deeper study (necessarily involving the reading of research papers). We hope to improve and complete the text gradually. Acknowledgements. I am grateful to the people at the Mathematics de- partment of Chalmers and the Gothenburg University for giving me very nice working conditions, and the impulse to prepare this course. Thanks also to many other collegues and students at that department for stimulating dis- cussions, and for making me feel well-come in Gothenburg. Thanks also to J.F. Bony for helping me to find a presentation of the material in Chapter 12, even though I finally settled for something less ambitious than what was originally planned. Last but not least, I would like to thank those who have courageously followed the lectures and showed their interest. Palaiseau, 31.5.2002 the author * Centre de Math´ ematiques, Ecole Polytechnique, F-91128 Palaiseau cedex, UMR 7640, CNRS 1

Upload: dangnhu

Post on 16-Sep-2018

213 views

Category:

Documents


0 download

TRANSCRIPT

Lectures on resonances

Johannes Sjostrand∗

1 Introduction

This text is the written version of lectures on resonances given at the Mathe-matics department of the Gothenburg University and Chalmers in the Springsemesters of 2000 to 2002. Some of the first chapters were also given in agraduate course at Ecole Polytechnique. Despite the somewhat preliminaryform of these notes we hope they can be of some use as an introductionto the subject, and some guidance for a deeper study (necessarily involvingthe reading of research papers). We hope to improve and complete the textgradually.

Acknowledgements. I am grateful to the people at the Mathematics de-partment of Chalmers and the Gothenburg University for giving me very niceworking conditions, and the impulse to prepare this course. Thanks also tomany other collegues and students at that department for stimulating dis-cussions, and for making me feel well-come in Gothenburg. Thanks also toJ.F. Bony for helping me to find a presentation of the material in Chapter12, even though I finally settled for something less ambitious than what wasoriginally planned. Last but not least, I would like to thank those who havecourageously followed the lectures and showed their interest.

Palaiseau, 31.5.2002the author

∗Centre de Mathematiques, Ecole Polytechnique, F-91128 Palaiseau cedex, UMR 7640,CNRS

1

2 Compactly supported perturbations of −∆Rn,

definition of resonances and their multi-

plicity.

2.1 The free resolvent.

We start by recalling some facts about the “free Laplacian” P0 = −∆ =∑n1 − ∂2

∂x2j

and its resolvent. First of all, −∆ is an unbounded selfadjoint

operator L2(Rn) → L2(Rn) with domain H2(Rn) (the standard Sobolevspace). By Fourier transform it is unitarily equivalent to the (unbounded)multiplication operator L2(R2) 3 u(ξ) 7→ ξ2u(ξ) ∈ L2(Rn). The spectrum of−∆ is equal to [0,+∞[ (and it is purely absolutely continuous).

For Imλ > 0, we know that λ2 6∈ [0,+∞[, so P0−λ2 : H2(R2)→ L2(Rn)is bijective with inverse R0(λ), which is a convolution operator with locallyintegrable kernel

R0(λ;x− y) =i

4(

1

2π|x− y|)n−2

2 (un−2

2 H(1)n−2

2

(u))|u=λ|x−y|

, (2.1)

where H(1)n−2

2

is a Hankel function. We can deduce (see Vainberg [91]) that

R0(λ) =

E(λ) + λn−2E1(λ), n oddF (λ) + F1(λ)λn−2 log λ, n even,

(2.2)

where E(λ), E1(λ), F (λ), F1(λ) are entire functions with values in the spaceof convolution operators and in the space of classical pseudodifferential op-erators of order −2: Hcomp(Rn)→ H2

loc(Rn).

If χ1, χ2 ∈ C∞0 (Rn) with suppχ1∩ χ2 = ∅, then χ1(any of E,E1, F, F1)χ2

is an entire function in λ with values in L(H−s, Hs), for every s > 0.Further for n even:

R0(eimπz) = R0(z)− (im

4π)(−1)

n−22

(m+1)T (z) (2.3)

for 0 ≤ arg ≤ π (z 6= 0,) m ∈ Z, and eimπz is viewed as a point on thelogarithmic covering space of C \ 0. Here T (z) is the convolution operatorwith kernel

T (z, x− y) = (z

2π)n−2

∫Sn−1

eiz(x−y)wL(dw), (2.4)

2

where L(dw) denotes the standard Lebesgue measure on Sn−1.For n odd we have

R0(z)−R0(−z) = T (z), (2.5)

where

T (z, x− y) =i

2

zn−2

(2π)n−1

∫Sn−1

eiz(x−y)wL(dw). (2.6)

The basic facts to remember are that R0(λ), initially defined for Imλ >0 and viewed as a convolution operator H0

comp(Rn) → H2loc(R

n), can beextended holomorphically to:– λ ∈ C, when n is odd ≥ 3,– λ ∈ C \ 0, when n = 1, and the extension has a simple pole at λ = 0,– to the logarithmic (universal) covering space of C \ 0, when n is even.

These properties have an interesting relation to the strong Huygens prin-ciple for the wave equation. It is well known that the Cauchy problem inR×Rn:

(∂2t −∆)U0(t) = 0, U0(0) = 0, U ′0(0) = δ, (2.7)

has a unique solution U0 ∈ C∞(Rt;D′(Rnx)) and that

suppU0 ⊂ (t, x); |x| ≤ |t|.

This is the Huygens principle which is valid more generally for wave equationswith variable coefficients and more general hyperbolic equations, and whichsays that the support of a solution cannot travel faster than the natural wavepropagation speed (which is one in our special case). When n is odd and ≥ 3,we have the strong Huygens principle:

suppU0 = (t, x); |x| = |t|

This principle has the important practical consequence that we can commu-nicate and send intelligible signals using sound waves and electromagneticwaves, since we are lucky enough to live in a three dimensional world.

By Duhamel’s principle, we have the forward fundamental solution of thewave operator E0 ∈ S ′(R×Rn) with

(∂2t −∆)E0 = δRn+1 , E0 ⊂ (t, x); t ≥ |x|, (2.8)

given byE0 = H(t)U0(t), (2.9)

3

where H(t) = 1[0,+∞[(t) is the Heaviside function. Viewing U0 as a convolu-tion operator, solving (∂2 −∆)U0(t) = 0, U0(0) = 0, U ′0(0) = 1 (1 indicatingthe identity operator) we can write

U0(t) =sin t√−∆√−∆

. (2.10)

Without any general spectral theory, (2.10) simply says that for the Fouriertransform with respect to x, we have

U0(t)(ξ) =sin t|ξ||ξ|

,

a fact that can be easily verified by taking the Fourier transform with respectto x of (2.8).

On the other hand, for Imλ > 0, it is easy to see that

R0(λ) =∫ ∞

0eitλU0(t)dt. (2.11)

This will be verified in a more general situation in Chapter 3. The sameformula holds on the level of convolution kernels. If n is odd ≥ 3, we seethat U0(t, x) = 0 for t > |x| by the strong Huygens principle, so in a region|x| < R we have

R0(λ, x) =∫ R+1

0eitλU0(t, x)dt, (2.12)

which can be extended holomorphically in λ to all of C.Using (2.11), we can also derive explicit formulas for R0, from correspond-

ing explicit formulas for the fundamental solution. In odd dimensions thecomputations are easier. It is known (see volume 1 of [42]) that we have fort > 0:

〈U0(t), χ〉 =1

4πn−1

2

(1

2t

∂t)n−3

2 tn−2∫|ω|=1

χ(tω)L(dω), (2.13)

where L(dω) denotes the Lebesgue measure, and n is odd and ≥ 3. Taken = 3 for simplicity, so that

〈U0(t), χ〉 =1

4πt∫|ω|=1

χ(tω)L(dω). (2.14)

Then

〈R0(λ), χ〉 =1

∫|ω|=1

∫ ∞0

eitλχ(tω)tdtL(dω). (2.15)

4

With x = tω, t = |x|, |ω| = 1, we have t2dtdω = dx, so

〈R0(λ), χ〉 =1

∫Rn

1

|x|ei|x|λχ(x)dx, (2.16)

or in other terms,

R0(λ, x) =ei|x|λ

4π|x|. (2.17)

For higher odd values of n one gets a similar formula with the same exponen-tial factor and a more complicated polynomial expression in front. For evenn the expression is even more complicated and contains special functions.For n = 1 the computation is of course easy and is left as an exercise. In allcases, R0(λ, x) is smooth and even analytic for x 6= 0, locally integrable inx, and exponentially decaying (for λ in the upper halfplane) when x → ∞.These properties can also be deduced by means of Fourier inversion.

vHere is a more general way to understand the link between the dimension

and the extendability properties. Starting from the physical halfplane Imλ >0, we use Fourier inversion to write

R0(λ, x) =1

(2π)n

∫eix·ξ

1

ξ2 − λ2dξ. (2.18)

To avoid uninteresting convergence problems, we take φ ∈ C∞0 (Rn), so thatφ = F−1φ is of class S on Rn and extend to an entire function on Cn. Inthe sense of distributions, we get

〈R0(λ), φ〉 =∫ 1

(ξ2 − λ2)φ(ξ)dξ. (2.19)

Now pass to polar coordinates, but let r vary in R rather than in [0,∞[, sothat we get the double covering

R× Sn−1 3 (r, ω) 7→ rω ∈ Rn. (2.20)

Then the pull back of dξ becomes

|r|n−1drdL(ω) =

rn−1drL(dω) for n odd.(sgn r)rn−1drL(dω) for n even.

5

We can then write

〈R0(λ), φ〉 =1

2

∫Sn−1

∫γ

φ(rω)

r2 − λ2rn−1drL(dω), (2.21)

where γ is the positively oriented real axis when n is odd and the union of[0,−∞[ and [0,+∞[, both oriented from 0 to∞, when n is even. Let λ startin the upper half plane and turn once around 0 in the clockwise direction.[[[[ Figur!]]]] From this simple argument we see that R0(λ) extends to aholomorphic function on C \ 0 when n is odd, while we get a log typesingularity when n is even. From this one can verify or at least understandqualitatively the earlier statements involving T (λ).

2.2 Definition of resonances in the case of P = −∆+V ,when V is bounded and compactly supported.

Let V ∈ L∞comp(Rn; R). Then P := P0 + V is an unbounded selfadjointoperator in L2(Rn) with domain H2. [0,+∞[ is the essential spectrum, so in]−∞, 0[ there can only be isolated eigenvalues of finite multiplicity. We thenknow that the resolvent (P − λ2)−1 exists as a bounded operator H0 → H2,when Imλ > 0, λ 6= iµj, where −µ2

j are the negative eigenvalues. We wantto find a meromorphic extension of the resolvent as a continuous operatorL2

comp(Rn)→ H2comp(Rn) across R \ 0. Write

P − λ2 = (P0 − λ2) + V = (1 + V R0(λ))(P0 − λ2). (2.22)

For |λ| 1, arg λ = π4, it is clear that (1 + V R0(λ))−1 exists. Moreover

V R0(λ) : L2 → L2 is compact and depends holomorphically on λ for Imλ >0. We conclude by analytic Fredholm theory (see Chapter 5) that (1 +V R0(λ))−1 exists except for λ in some discrete set, where the inverse haspoles of finite multiplicity, with finite rank coefficients in the singular terms ofthe Laurent expansions. (Actually these poles are precisely the iµj, because(2.22) shows that P − λ2 is bijective H2 → H0 precisely when 1 + V R0(λ) isbijective.)

After crossing R \ 0, V R0 is in general no more a bounded operator inL2, but we can still try to solve

(1 + V R0(λ))u = v (2.23)

6

with u, v of compact support. Notice that the support of such a solution u willbe contained in the union of the supports of v and V . Let 1suppV ≺ χ ∈ C∞0where we write χ1 ≺ χ2 for two functions χ1, χ2 if the support of χ1 iscontained in the interior of the support of 1 − χ2. Then 1 + V R0(λ)χ is aholomorphic family of Fredholm operators (see Chapter 5) and we shall seethat in order to solve (2.23) it suffices to invert 1 + V R0(λ)χ.

Indeed, we have

1 + V R0(λ) = (1 + V R0(λ)(1− χ))(1 + V R0(λ)χ), (2.24)

1 = (1 + V R0(λ)(1− χ))(1− V R0(1− χ)), (2.25)

so(1 + V R0(λ)(1− χ))−1 = 1− V R0(λ)(1− χ) on L2

comp. (2.26)

Hence (1+V R0(λ)) : L2comp → L2

comp is invertible precisely when 1+V R0(λ)χis, and

(1 + V R0(λ))−1 = (1 + V R0(λ)χ)−1(1− V R0(λ)(1− χ)) : L2comp → L2

comp.(2.27)

Inverting (2.22), we get

(P − λ2)−1 = R0(λ)(1 + V R0(λ))−1, (2.28)

and by analytic Fredholm theory (see chapter 5), we know that (1+V R0(λ)(1−χ))−1 has a meromorphic extension in λ to the complex plane when n ≥ 3is odd and to other sets already described in the other cases, as a functionwith values in the space L(L2(Rn), L2(Rn)) of bounded operators from L2

to L2. By a meromorphic operator valued function of z ∈ Ω, with values inL(H1,H2), where Hj are complex Hilbert spaces and Ω is an open set in thecomplex plane or in a covering surface over some open set of C, we meana function A(z) which is holomorphic in z on Ω \ S, where S is a discretesubset of Ω and such that if z0 ∈ S, then for z in a neighborhood of z0, wehave the Laurent series representation

A(z) =N∑1

Aj(z − z0)j

+B(z), (2.29)

where N is finite (so that we have poles of finite order), B(z) a holomorphicoperator valued function in a neighborhood of z0, and Aj are bounded oper-ators, independent of z and of finite rank. We leave as an exercise to verify

7

that compositions and sums of meromorphic operator valaued functions arestill meromorphic. Meromorphic operator valued functiosn are also conservedunder changes of variables z = κ(λ) if κ : Λ → Ω is a biholomorphic mapbeteween two open sets.

Proposition 2.1 (P − λ2)−1 : H0comp → H2

loc has a meromorphic extensionfrom the open upper half plane to:C \ 0, when n = 1,C, when n ≥ 3 is odd,the logarithmic covering space (C \ 0)∗ of C \ 0, when n is even.

By definition, the resonances or scattering poles of P are the poles of thisextension, with the exception of the iµj. Since H0

comp, H2loc are not Hilbert

or even Banach spaces, we have to extend the definition of meromorphic andholomorphic functions with values in operators between such spaces. LetΩ be as above. A holomorphic function with values in L(H0

comp, H2loc) is a

function A(z) with values in the space of linear operators H0comp → H2

loc,such that χ1A(z)χ2 is holomorphic for all χj ∈ C∞0 (Rn). Correspondingly ameromorphic function is one which is holomorphic on Ω \ S, where S ⊂ Ωis discrete, and such that if z0 ∈ S, then near z0, we have (2.29), now withAj : H0

comp → H2loc (continuous in the sense that χ1Ajχ2 is bounded for all χj

as above) of finite rank, and B(z) holomorphic with values in L(H0comp, H

2loc)

for z in a neighborhood of z0.

2.3 Black box frame work and definition of resonancesfor general compactly supported perturbations of−∆.

The black box frame work (introduced by Sjostrand and Zworski in [80])permits to give a uniform treatment of all sorts of compactly supported (andlater also long range) perturbations of −∆, like exterior obstacle problems,metric pertubations, Schrodinger operators, or a combination of such prob-lems.

Let H be a complex Hilbert space with the orthogonal decomposition

H = HR0 ⊕ L2(Rn \B(0, R0)), (2.30)

where R0 > 0 and we use the standard notation B(x, r) = y ∈ Rn; |y−x| <r. We denote by

u 7→ u|B(0,R0)or 1B(0,R0u

8

u 7→ u|Rn\B(0,R0)or 1Rn\B(0,R0)u,

the corresponding orthogonal projections. If χ is a bounded continuous func-tion on B(0, R0); χ ∈ Cb(Rn), and χ = C0 = Const. on B(0, R0), then foru ∈ H, we can define χu ∈ H by

χu = C0u+ 1Rn\B(0,R0)(χ− C0)u. (2.31)

Let

Hcomp = u ∈ H; u|Rn\B(0,R0)has bounded support,

Hloc = HR0 ⊕ L2loc(R

n \B(0, R0)).

Notice that if χ ∈ C(Rn) is constant near B(0, R0), then u 7→ χu is welldefined: Hcomp/loc → Hcomp/loc.

We now consider a selfadjoint unbounded operator P : H → H withdomain D ⊂ H. We assume

u ∈ D ⇒ u|Rn\B(0,R0)∈ H2(Rn \B(0, R0)), (2.32)

and that the restriction operator is continuous,

u ∈ H2(Rn \B(0, R0)), u = 0 near ∂B(0, R0)⇒ u ∈ D, (2.33)

where as often we identify H0(Rn \ B(0, R0)) with a subspace of H in thenatural way,

(Pu)|Rn\B(0,R0)= −∆(u|Rn\B(0,R0)

), ∀u ∈ D. (2.34)

Notice that if u is as in (2.33) (with vanishing HR0 component) then Pu =−∆u (i.e. the HR0 component of Pu vanishes) since for every v ∈ D:

(Pu|v) = (u|Pv) = (u| −∆(v|Rn\B(0,R0))) = (−∆u|v)L2(Rn\B).

We also assume that

1B(0,R0)(i− P )−1 is compact:H → H. (2.35)

DefineDcomp = D∩Hcomp, Dloc = u ∈ Hloc; χu ∈ D, ∀χ ∈ C∞0 (Rn) with χ =Const. near B(0, R0).

We should keep i mind the following two examples:

9

1) Let P = −∆+V (x), where V ∈ L∞comp(Rn). Then we can takeH = L2(Rn)and the domain of our self-adjoint operator P is D = H2(Rn).

2) Let O ⊂ Rn be a bounded open set with smooth boundary. Let P bethe unbounded self-adjoint realization of −∆ on L2(Rn \ O) with domainD = u ∈ H2(Rn \ O); u|∂O = 0. P is often called the exterior DirichletLaplacian. Similarly we can define the exterior Neumann Laplacian by takingD = u ∈ H2(Rn\O); ∂νu|∂O = 0, where ∂νu denotes the exterior derivativein the unit normal direction.

Theorem 2.2 a) P has only discrete spectrum in ]−∞, 0[,

b) (P − λ2)−1 : Hcomp → Dloc has a meromorphic extension from λ ∈C; Imλ > 0, λ2 6∈ σ(P )∩]−∞, 0[ to:C \ 0 when n = 1,C when n ≥ 3 is odd,(C \ 0)∗ when n is even.

Here meromorphic functions with values in the space of continuous op-erators Hcomp → Dloc are defined in the same way as in the special case ofsuch functions with values in the continuous operators: H0

comp → H2loc.

Proof. We first “enlarge” the compactness property (2.35). If z, w ∈ ρ(P ),the resolvent set of P , then we have the resolvent identity

(z − P )−1 = (w − P )−1 − (w − P )−1(z − w)(z − P )−1, (2.36)

implying that

1B(0,R0)(z−P )−1 = 1B(0,R0)(i−P )−1−1B(0,R0)(i−P )−1(i−w)(z−P )−1 (2.37)

is compact. Passing to the adjoints, we see that (z−P )−11B(0,R0) is compacttoo.

Let B(x, r0, r1) denote the shell y ∈ Rn; r0 < |y − x| < r1. Then1B(0,R0,R) : H2(Rn \B(0, R0))→ L2(Rn \B(0, R0)) is compact for R0 < R <∞. It follows that

1B(0,R)(z − P )−1, (z − P )−11B(0,R) : H → H (2.38)

are compact for z ∈ ρ(P ), R0 < R <∞.

Let χ0, χ1, χ2 ∈ C∞0 with 1B(0,R0) ≺ χ0 ≺ χ1 ≺ χ2. We first let λ ∈ Cwith Imλ > 0 and take µ with the same properties and such that µ2 6∈ σ(P ).We will fix µ later. Put

Q(λ, µ) = (1−χ0)R0(λ)(1−χ1)+χ2R(µ)χ1 = Q0(λ, µ)+Q1(λ, µ) : H → D,

10

where R0(λ) = (P0 − λ2)−1, R(µ) = (P − µ2)−1, and the last equality abovegives the obvious definition of Q0, Q1. Then:

(P − λ2)Q0 = (1− χ1) + [∆, χ0]R0(λ)(1− χ1) = (1− χ1) +K0(λ), (2.39)

(P − λ2)Q1 = χ2(P − λ2)R(µ)χ1 − [∆, χ0]R0(µ)χ1 (2.40)

= χ1 + χ2(µ2 − λ2)R(µ)χ1 − [∆, χ2]R(µ)χ1 = χ1 +K1(λ, µ),

where the last equations define the operators K0, K1. Hence,

(P − λ2)Q = 1 +K0(λ) +K1(λ, µ) = 1 +K(λ, µ), (2.41)

where the last equality defines K.Using (2.38) and the fact that [∆, χj] is a first order operator with smooth

compactly supported coefficients and therefore compact: H2(Rn\B(0, R0))→L2(Rn \B(0, R0)), we see that

K(λ, µ) is compact. (2.42)

Now choose µ with arg µ = π4, so that Im (µ2) = |µ|2. Then

R(µ) =

O( 1|µ|2 ) : H → H,

O(1) : H → D,

and,

(1− χ0)R(µ) =

O( 1|µ|2 ) : H → L2

O(1) : H → H2,

so by interpolation, (1− χ0)R(µ) = O( 1|µ|) : H → H1. It follows that

K(µ, µ) = O(1

|µ|) : H → H, (2.43)

and consequently 1 + K(µ, µ) : H → H is invertible with a bounded inverse(1 + K(µ, µ))−1 of norm ≤ 2 when |µ| is large enough (and arg µ = π/4).We fix such a value of µ and apply analytic Fredholm theory with respectto λ. It follows that (1 + K(λ, µ))−1 exists except for λ in a discrete set inthe open upper half plane, and that the poles are of finite order and withcoefficients of finite rank for the singular terms in the Laurent expansion ateach such pole. (See chapter 5.)

11

Writing(P − λ2)−1 = Q(λ, µ)(1 +K(λ, µ))−1, (2.44)

we see that (P − λ2)−1 : H → D is meromorphic in the upper half planewith only finite rank singular terms in the Laurent expansion. Of course, wealready know that the only possible singularities are situated on the negativeimaginary axis and correspond to the negative spectrum of P , so we see thatthat this spectrum is purely discrete and we have obtained part (a) of thetheorem. (The spectral projection onto any finite part of the spectrum isgiven by 1

2πi

∫γ(λ

2−P )−1d(λ2), where γ is simple closed contour in the upperhalf plane which avoids the poles and encircles a certain number of squareroots of negative eigenvalues.)

We next consider the meromophic extension and let n ≥ 3 be odd inorder to fix the ideas. We keep µ fixed as before. Then C 3 λ 7→ Q(λ, µ) :Hcomp → Dloc is holomorphic. Let χ2 ≺ χ ∈ C∞0 (Rn), so that χK = K. Toinvert 1 +K : Hcomp → Hcomp it suffices to invert 1 +Kχ : H → H. Indeed,we observe that (1− χ)K = 0 and hence

(1 +K(1− χ))−1 = 1−K(1− χ), (1 +K) = (1 +K(1− χ))(1 +Kχ),

implying(1 +K)−1 = (1 +Kχ)−1(1−K(1− χ)). (2.45)

Using (2.45) in (2.44) and applying analytic Fredholm theory to (1 +Kχ)−1,we get the desired meromorphic extension

R(λ) = Q(λ, µ)(1 +K)−1 = Q(1 +Kχ)−1(1−K(1− χ)). (2.46)

#

In the following we shall sometimes use the notation R(λ) = (P − λ)−1

for the meromorphic extension of the resolvent, even though it is no morean inverse, strictly speaking. The poles of the meromorphic extension notalready in the physical halfplane C+ := λ ∈ C; Imλ > 0 that was thestarting point of the extension, will be called resonances (and sometimesscattering poles). If λ0 is a non-vanishing resonance, we define its multiplicitym(λ0) by

m(λ0) = rank1

2πi

∫γ(λ2 − P )−1d(λ2), (2.47)

12

where γ is a small positively oriented circle centered at λ0. With z0 = λ20,

z = λ2, we have for z close to z0:

(z − P )−1 = (z − z0)−NΠ−N + ..+ (z − z0)−1Π−1 + Hol (z) : Hcomp → Dloc ,(2.48)

where the last term is holomorphic in a neighborhood of z = z0. From thisexpression we see that m(λ0) is the rank of Π−1 in view of the standardformula:

Π−j =1

2πi

∫γ(z − z0)j−1(z − P )−1dz, j ≥ 1. (2.49)

Here γ denotes the contour in the z plane which corresponds to γ in the λplane.

Use that

(P − z0)(z − P )−1 = (z − z0)(z − P )−1 − 1

by holomorphic extension from the physical sheet, to get for j ≥ 1:

(P − z0)Π−j =1

2πi

∫γ(z − z0)j−1(P − z0)(z − P )−1dz (2.50)

=1

2πi

∫γ(z − z0)j(z − P )−1dz − 1

2πi

∫γ(z − z0)j−1dz = Π−(j+1)

Consequently (P − z0) maps Π−j(Hcomp) into Π−(j+1)(Hcomp) (also includedin Dloc). Since Π−j are of finite rank and Dcomp is dense in Hcomp [[detaillercet argument]], we have Π−j(Hcomp) = Π−j(Dcomp). Moreover, on Dcomp wehave Π−jP = PΠ−j, so (2.50) becomes:

Π−(j+1) = Π−j(P − z0) on Dcomp. (2.51)

Consequently,

Π−(j+1)(Hcomp) = Π−(j+1)(Dcomp) ⊂ Π−j(Hcomp),

and (P − z0)|Π−1(Hcomp)is a nilpotent operator. The elements of the kernel

of this operator will be called resonant states. The elements of Π−1(Hcomp)will be called generalized resonant states.

If Π−1v is a resonant state, then 0 = Π−2v = Π−3v = .. and (z − P )−1vhas a simple pole at z0.

13

Proposition 2.3 Let 1B(0,R0) ≺ χ ∈ C∞0 (Rn). Then with γ as in (2.47),

m(λ0) = rank∫γ(λ2 − P )−1χd(λ2) (2.52)

Proof. Clearly m(λ0) is at least as large as the RHS of (2.52). Let χ ∈C∞0 (Rn), with 1B(0,R0) ≺ χ ≺ χ. From (2.44), we get

(P − λ2)−1 = Q− (P − λ2)−1K(λ, µ) = Q− (P − λ2)−1χK(λ, µ),

provided we let the cutoffs χ1, χ2, χ3 have their supports sufficiently close toB(0, R0). Hence

1

2πi

∫γ(λ2 − P )−1d(λ2) = − 1

2πi

∫γ(λ2 − P )−1χK(λ, µ)d(λ2). (2.53)

Recall that µ is fixed and write the Taylor expansion of K at λ20 in the

variable λ2:

K(λ, µ) =N−1∑j=0

Kj(µ)(λ2 − λ20)j + (λ2 − λ2

0)N)Hol (λ),

where Hol (λ) is function which is holomorphic for λ close to λ0. Use this in(2.53) together with (2.49):

1

2πi

∫γ(λ2 − P )−1d(λ2) = (2.54)

−N−1∑

0

1

2πi

∫γ(λ2 − λ2

0)j(λ2 − P )−1d(λ2)χKj = −N−1∑

0

Π−(1+j)χKj.

Here we use (2.50):

Π−(j+1)χ = Π−j(P − z0)χ = Π−jχ(P − z0)χ,

if χ ≺ χ, soR(Π−(j+1)χ) = R(Π−jχ).

Iterating, we getR(Π−(j+1)χ) ⊂ R(Π−1χ),

and using this in (2.54), we get

R(Π−1) ⊂ R(Π−1χ),

which implies (2.52). #

14

2.4 Absence of real resonances.

We end this chapter by discussing the relation with outgoing solutions to thefree Helmholtz equation, and then applying it to show that real non-vanishingresonances will not appear in standard situations. (Rellich’s theorem.) Weshall first see that resonant states associeted to such resonances are outgoingstates in a sense that we define below.

Let 1B(0,R0) ≺ χ ∈ C∞0 (Rn). For z = λ2, Imλ > 0, we have for v ∈ Hcomp

with 1supp v ≺ χ:

(z − P )−1v = χ(z − P )−1v + (z − P0)−1[P, χ](z − P )−1v, (2.55)

where we notice that the second term of the RHS vanishes outside supp (1−χ). In fact, let u = (z − P )−1v and compute

(z − P0)((1− χ)u) = (z − P )((1− χ)u) = [P, χ]u

which implies that

(1− χ)(z − P )−1v = (z − P0)−1[P, χ](z − P )−1v.

The relation (2.55) extends by analytic continuation to the non-physicalsheet(s). Let z0 = λ2

0 where λ0 is a resonance. Integrating (2.55) along asmall contour γ around z0, we get for z ∈ Rn \B(0, R0):

Π−1v − χΠ−1v =N∑j=1

1

2πi

∫(z − P0)−1[P, χ](z − z0)−jΠ−jvdz (2.56)

=N∑j=1

1

(j − 1)!(∂j−1z (z − P0)−1)z=z0 [P, χ]Π−jv.

Choose v ∈ Hcomp so that Π−1v is a resonant state. Then (2.56) simplifies to

Π−1v = χΠ−1v + (z0 − P0)−1[P, χ]Π−1v, (2.57)

where we recall that (z0−P0)−1 denotes the holomorphic extension from thephysical sheet to the point z0.

For simplicity, we now restrict the attention to a real positive z0 = zor equivalently to a real non-vanishing λ. Let R0(λ) = (P0 − λ2)−1 be thebranch of the resolvent obtained by extension from the upper half plane:R0(λ) = (P0 − (λ+ i0)2)−1 : E ′ → S ′.

15

Definition. We say that u ∈ S ′(Rn) is outgoing if there exists w ∈ E ′(Rn)such that

u(x) = R0(λ)w for |x| 1. (2.58)

We have just seen that a resonant state (or rather it’s component inL2

loc(Rn \B(0, R0))) is outgoing.

Exercise: Let u ∈ S(Rn) be outgoing, so that u(x) = R0(λ)w(x), |x| 1,for some w ∈ E ′. Then (P0 − λ2)u = v ∈ E ′. Show that u = R0(λ)v.

Solution: As in (2.55) we have for χ ∈ C∞0 with χ = 1 on a sufficiently largeball:

R0(λ)w = χR0(λ)w −R0(λ)[P0, χ]R0(λ)w,

and hence u = −R0(λ)[P0, χ]u near ∞.On the other hand, let 1supp v ≺ χ. Then

(P0 − λ2)χu = v + [P0, χ]u ∈ E ′, (2.59)

and on E ′, we have R0(λ)(P0−λ2) = 1 by analytic extension from the physicalhalf plane. So applying R0(λ) to (2.59), we get

χu = R0(λ)v +R0(λ)[P0, χ]u.

HenceR0(λ)v = −R0(λ)[P0, χ]u = u near ∞.

u−R0(λ)v is therefore in E ′ and since it satisfies (P0− λ2)(u−R0(λ)v) = 0,it has to be 0. (Notice or recall that every non-vanishing partial differentialoperator with constant coefficients is injective E ′(Rn)→ E ′(Rn).) #

Theorem 2.4 If λ ∈ R \ 0 is a resonance and u ∈ Dloc a correspondingresonant state, then u has compact support.

We notice that in most standard situations, there cannot be any non-trivial solution of the equation (P − λ2)u which has compact support andhence the theorem implies that that there are no real non-vanishing reso-nances. For the proof of the theorem we need

16

Proposition 2.5 Let v ∈ C∞0 (Rn), 1supp v ≺ χ ∈ C∞0 (Rn; R), 0 6= λ ∈ R.Then

1

2i([−∆, χ]R0(λ)v|R0(λ)v) = C(λ)

∫∂B(0,|λ|)

|v(ξ)|2L(dξ) = Im (R0(λ)v|v),

(2.60)where C(λ) 6= 0 can be computed, and L(dξ) denotes the standard Lebesguemeasure on the sphere of radius |λ|.

Proof. This is a simple computation, where we assume that λ > 0 forsimplicity:

1

2i([−∆, χ]R0(λ)v|R0(λ)v) =

1

2i([−∆− λ2, χ]R0(λ)v|R0(λ)v)

=1

2i

((χR0(λ)v|(−∆− λ2)R0(λ)v)− (χ(−∆− λ2)R0(λ)v|R0(λ)v)

)=

1

2i

((R0(λ)v|χv)− (χv|R0(λ)v)

)= Im (R0(λ)v|χv) = Im (R0(λ)v|v) =

Im1

(2π)n

∫Rn

v(ξ)v(ξ)

(ξ2 − (λ+ i0)2)dξ =

1

(2π)n

∫Rn|v(ξ)|2 Im

1

(|ξ|2 − (λ+ i0)2)dξ

(2π)n2λ

∫∂B(0,λ)

|v(ξ)|2L(dξ).

#

Notice that by density the first equality in (2.60) holds also for v ∈ E ′.Proof of Theorem 2.4. Let u be an outgoing solution of (P − λ2)u = 0with λ ∈ R+. Let 1B(0,R0) ≺ χ ∈ C∞0 be real-valued and consider

1

2i([P, χ]u|u) =

1

2i((χu|Pu)− (χPu|u))

=1

2i(λ2(χu|u)− λ2(χu|u)) = 0.

On the other hand, if we write u = R0(λ)v, v ∈ C∞0 , for |x| 1, (we cantake v = −[P0, χ]u for a suitable χ), then if we let χ ∈ C∞0 be a real-valuedcut-off which is equal to 1 on a sufficiently large ball, we have

0 = (1

2i[P, χ]u|u) =

1

2i([−∆, χ]R0,+(λ)v|R0,+(λ)v)

= C(λ)∫∂B(0,λ)

|v(ξ)|2L(dξ),

17

so v(ξ) = 0 on B(0, λ) and hence also on the complexification of this ball.By the Paley-Wiener theorem plus easy division estimates (see Lemma 2.6below), we see that R0,+(λ)v ∈ E ′. Hence u has compact support. #

Remark. In many situations, we know that there are no non-trivial solutionsof (P − λ2)u = 0 with compact support, and it then follows from Theorem2.4 that there are no real non-vanishing resonances.

Lemma 2.6 Let v ∈ E ′(Rn) with v(ξ) = 0 on the real sphere ξ2 = λ2, forsome λ > 0. Then R0(λ)v ∈ E ′(Rn) (which implies that suppR0(λ)v iscontained in the union of the support of v and the bounded components of thecomplement of the support of v).

Proof. We recall a weak form of the Paley-Wiener Theorem: If u ∈ S ′(Rn),then u belongs to E ′ if and only if

∃C0, C1, N0 > 0, such that u(ξ) is entire and (2.61)

|u(ξ)| ≤ C1〈ξ〉N0eC0|Im ξ|, ξ ∈ Cn.

If v satisfies the assumptions of the lemma, then u := v(ξ)(ξ2−λ2)

is an entire

function and it is easy to see that u is the Fourier transform of R0(λ)v. Itsuffices, then to show that u(ξ) satisfies (2.61) for some suitable constants.

Let δ > 0 be small (to be fixed later) and let ξ0 ∈ Cn with |ξ20 − λ2| ≤ δ.

Writeξ2 − λ2 = 2ξ0(ξ − ξ0) + (ξ2

0 − λ2) + (ξ − ξ0)2.

Take

ξ − ξ0 = zξ0

|ξ0|, with z ∈ C, |z| = ε > 0,

where ε > 0 will be fixed independently of ξ0.Then

|ξ2 − λ2| ≥ 2|ξ0|ε− δ − ε2.λ being fixed, we first choose δ ≤ δ1(λ) small enough, so that |ξ0| ≥ |λ|

2and

get|ξ2 − λ2| ≥ |λ|ε− δ − ε2.

Choose ε < |λ|/2 so that |ξ2 − λ2| ≥ |λ|ε/2 − δ and finally δ ≤ δ2(λ, ε) sothat |ξ2 − λ2| ≥ |λ|ε/4. Consider

u(ξ0 + zξ0

|ξ0|) =

v(ξ0 + z ξ0

|ξ0|))

(ξ0 + z ξ0

|ξ0|)2 − λ2

18

which is holomorphic for |z| < 2ε. We have

|u(ξ0 + zξ0

|ξ0|)| ≤ 4

|v(ξ0 + z ξ0

|ξ0|)||λ|ε

for |z| = ε, so by the maximum principle we get for |z| ≤ ε:

|u(ξ0 + zξ0

|ξ0|)| ≤ 4

|λ|εsup|z|=ε|v(ξ0 + z

ξ0

|ξ0|)|.

Especially we get the last estimate for u(ξ0), and we conclude that if v satisfies(2.61), then u also satisfies (2.61) in the region |ξ2 − λ2| ≤ δ, and henceeverywhere, since |u(ξ)| ≤ δ−1|v(ξ)| in the region where |ξ2 − λ2| ≥ δ. #

Remark. It is a little easier to show that if

e0 ∈ D, (P − λ2)e0 = 0, λ ∈ R \ 0,

so that λ2 is a positive eigenvalue, then e0 has compact support. Take χ asusual, equal to one near B(0, R0). Then

(P0 − λ2)(1− χ)e0 = [∆, χ]e0 ∈ H1comp,

so if F denotes the Fourier transformation, then

(ξ2 − λ2)F((1− χ)e0)(ξ) = (F [∆, χ]e0)(ξ).

Here the right hand side is entire and in order to have F(1− χ)e0 ∈ L2, wemust have F [∆, χ]e0)(ξ) = 0 when ξ2 = λ2. As before, we conclude by meansof Paley-Wiener, that

(F((1− χ)e0)(ξ) =F([∆, χ]e0)(ξ))

ξ2 − λ2

is an entire function and satisfies (2.61). Hence (1−χ)e0 has compact supportand so does e0.

[[[Missing piece in this presentation: Show that if (P − λ2)u = 0 and u isoutgoing, then u is a resonant state and λ is a resonance.]]]

19

3 Expansion of solutions to the wave equa-

tion in exponentially decaying resonant modes,

when t→ +∞.

In this chapter we show (in the case of odd dimension n ≥ 3) that if a certainnon-trapping assumption for the associated wave equation is fulfilled, thenfor every N > 0 there are at most finitely many resonances above the curve

Im z = −N log+ |Re z|

(This fact is true also in even dimensions, but a little more difficult to provein that case.) Moreover if the space dimension is odd and ≥ 3 we shall showthat the solution of the associated wave equation with compactly supportedinitial data has an asymptotic expansion in exponentially decaying modes:e−iλjtuj(x, t) as t → +∞, where λj is a resonance or a square root of anegative eigen-value and uj is a polynomial of degree ≤ ω(λj)−1 in t, whereω(λj) is the order of λj as a pole of the meromorohic extension of (P −λ2)−1.This result is due to Lax-Phillips (see [50]) in a more restricted setting, andthe idea of the proof we give is due to Vainberg [91]. Vainberg’s proof wasrecently adapted to the black box setting by Tang and Zworski [90], but ourproof (and result) differs slightly. At one point we shall use a basic result onthe propagation of singularities for the wave operator. See [42].

Let P be a bb operator as in the preceding chapter, satisfying the addi-tional assumption:

P ≥ −C, (3.1)

for some constant C. Let Dj denote the domain of 〈P 〉j when j ≥ 0 and letDj = (D−j)∗ be the dual space of D−j, when j ≤ 0.

Recall that by the spectral theorem, P is unitarily equivalent to the multi-plication operator P : u(m) 7→ p(m)u(m) in H = L2(M,µ) where M is someset, µ a measure on M and p(m) a real-valued measurable function on M .Then Dj becomes 〈p(m)〉−ju(m); u ∈ L2(M ;µ). From this representationand the Holder inequality, we deduce the interpolation inequality

‖u‖Dθj+(1−θ)k ≤ ‖u‖θDj‖u‖1−θDk , j, k ∈ R, 0 ≤ θ ≤ 1, u ∈ Dmax(j,k). (3.2)

For g ∈ Dj, we can solve the abstract Cauchy problem

(P + ∂2t )u = 0, u(0) = 0, ∂tu(0) = g, (3.3)

20

by the functional formula:

u(t) =sin t√P√

Pg =: U(t)g. (3.4)

(If we apply the unitary equivalence above, then U(t) can be identified with

the multiplication operator u(m) 7→ sin(t√p(m))√

p(m)u(m).) Note that f(t, E) :=

sin t√E√

Eis a well-defined smooth function on R×R, independent of the choice

of branch of√E, which satisfies the following estimates for E ≥ −Const.:

∂kt f(t, E) = Ok(1)e|t|C〈E〉(k−1)/2. (3.5)

It follows thatU(t) ∈ Cm(Rt;L(Dj,Dj+

1−m2 )), (3.6)

where the mth derivative is merely strongly continuous in t with values inL(Dj,Dj−m−1

2 ), while the derivatives of lower order are continuous in the

norm of L(Dj,Dj+ 1−m2 ). Moreover,

‖∂kt U‖L(Dj ,Dj+1−k

2 )≤ Ck,je

C|t|. (3.7)

We take a break in the general discussion in order to recall some basicexistence and uniqueness results for the abstract equation (3.3).

Proposition 3.1 If v ∈ C0(R;Dj− 12 ), f ∈ Dj, g ∈ Dj− 1

2 , then

u(t) := U ′(t)f + U(t)g +∫ t

0U(t− s)v(s)ds

belongs to C0(R;Dj) ∩ C1(R;Dj− 12 ) ∩ C2(R;Dj−1) and solves the Cauchy

problem,(∂2t + P )u = v, u(0) = f, u′(0) = g. (3.8)

Proof. Notice that U(0) = 0, U ′(0) = 1, so u1 := U ′(t)f +U(t)g solves (3.8)with v = 0. Let

u2(t) =∫ t

0U(t− s)v(s).

Then u2(0) = 0,

u′2(t) = U(0)v(t) +∫ t

0U ′(t− s)v(s)ds =

∫ t

0U ′(t− s)v(s)ds,

21

so u′2(0) = 0. Since U ′(0) = 1, we get

∂2t u2(t) = v(t) +

∫ t

0U ′′(t− s)v(s)ds,

and hence

∂2t u2(t) + Pu2(t) = v(t) +

∫ t

0(∂2t + P )(U(t− s)v(s))ds = v(t),

since (∂2t +P )(U(t− s)v(s)) = 0. Further it is easy to see that u2(t) belongs

to the required space. #

Next look at the uniqueness of solutions of (3.8):

Proposition 3.2 Let I be an interval containing 0, and let u ∈ C2(I;Dk)be a solution of (∂2

t + P )u = 0. Then

u(t) = U ′(t)u(0) + U(t)u′(0). (3.9)

Proof. Let t ∈ I, and put v(s) = U(s − t)φ, φ ∈ DN for some sufficientlylarge N . Then for s between 0 and t:

d

ds[(u′(s)|v(s))− (u(s)|v′(s))] = (u′′(s)|v(s))− (u(s)|v′′(s))

= −(Pu(s)|v(s)) + (u(s)|Pv(s)) = 0.

Hence(u′(t)|v(t))− (u(t)|v′(t)) = (u′(0)|v(0))− (u(0)|v′(0)),

i.e.−(u(t)|φ) = (u′(0)|U(−t)φ)− (u(0)|U ′(−t)φ).

U(t) is odd in t, so U(−t) = −U(t), U ′(−t) = U ′(t). Hence

(u(t)|φ) = (u′(t)|U(t)φ) + (u(0)|U ′(t)φ).

Use also that U(t) and U ′(t) are symmetric:

(u(t)|φ) = (U(t)u′(0) + U ′(t)u(0)|φ).

Since DN is dense in D−k, the space dual to Dk, we get (3.9). #

We now return to the main discussion. Outside B(0, R0), (3.3) is just thestandard wave equation, for which Huygens’ principle applies and says that

22

supports of solutions cannot propagate with speed > 1: If u ∈ C2(R;D′(Rn))and

(−∆ + ∂2t )u = 0, u(0) = u0, ∂tu(0) = u1, (3.10)

thensuppu ⊂ (t, x); distRn(x, suppu0 ∪ suppu1) ≤ |t|. (3.11)

In the bb setting, we define the support of u ∈ Hloc to be suppu|Rn\B(0,R0if

u vanishes near B(0, R0) and B(0, R0) ∪ suppu|Rn\B(0,R0)otherwise. Notice

that suppu is closed.Let dist bb(x, y) = min(|x−y|, dist (x,B(0, R0))+dist (y,B(0, R0))). Then

for (3.3) we have

suppu ⊂ (t, x); distbb(x, supp g) ≤ |t|. (3.12)

We now introduce an abstract non-trapping condition, which says thatsingularities in compactly supported initial data are washed away from anybounded region after a suitable time:

For all a > R0, ∃Ta > 0, such that t 7→ χU(t)χ (3.13)

belongs to C∞(]Ta, Ta + 1[;L(D0;DN))

for every N ≥ 0 and every χ ∈ C∞0 (B(0, a)) with 1B(0,R0) ≺ χ.

In practice, the verification of this non-trapping condition is importantand depends on some result on the propagation of singularities for solutionsof the wave equation. The condition (3.13) is known to hold in the followingtwo cases: 1) P = −∆ +V , V ∈ C∞0 (Rn) and 2) −P is the exterior Dirichletor Neumann Laplacian on Rn \ O where O ⊂⊂ Rn is open with smoothboundary and non-trapping in the geometric sense that no maximal opticalray can be contained in a bounded set. For the second example we refer to[56], while the first example follows from the general result of Hormanderon progation of singularities for solutions of equations of principal type. See[42].

Let a > R0 + 1 and let 1B(0,R0≺ χa−1 ∈ C∞0 (B(0, a− 1)). We will work

for t ≥ 0 in the following. Let ψa(t) ∈ C∞0 (]−∞, Ta + 1[; [0, 1]) be equal to1 near ]−∞, Ta]. Then

(P + ∂2t )(ψa(t)U(t)χa−1) = (3.14)

(∂2t (ψa(t)) + 2∂t(ψa(t))∂t)U(t)χa−1 =: Va(t).

23

In view of (3.13), we see that if 1B(0,R0) ≺ χ ∈ C∞0 (B(0, a)), then

χVa(t) ∈ C∞0 (]Ta, Ta + 1[;L(H,DN)), ∀N ∈ N. (3.15)

We will choose χ ≺ χa−1.The ”exterior part” (1− χ)Va(t)g will not be smooth in general, and we

will add a correction for it, by solving the free wave equation. Let U0(t) bethe free wave ”group” defined as in (2.7) (with P replaced by −∆ and withHreplaced by L2(Rn)). As explained in chapter 2, we have the correspondingforward fundamental solution E0 given by

E0 = H(t)U0(t), (3.16)

so that (∂2t −∆)E0 = δRn+1 , and

suppE0 = (t, x) ∈ R×Rn; |x| = t, for n odd ≥ 3,(t, x) ∈ R×Rn; |x| ≤ t, otherwise.

(3.17)

Notice that Vag vanishes outside some bounded set, by Huygens’ principle.Let χ ∈ C∞0 (B(0, a−1)) be equal to 1 near B(0, R0). (We will take χ ≺ χa−1.)When n ≥ 3 is odd, the corrected truncated solution of the wave equation(associated to P ) is then by definition:

Ua(t) = ψa(t)U(t)χa−1 − (1− χ)(E0∗)(1− χ)Va(t), (3.18)

where E0∗ is the operator of convolution with E0, and χ ∈ C∞0 with 1B(0,R0) ≺χ ≺ χ. We get

(P + ∂2t )Ua(t)g = χVa(t)g + [−∆, χ](E0∗)(1− χ)Va(t)g. (3.19)

The second term has uniformly compact support by the strong Huygensprinciple. Let us consider also

Ua(t)g := ψa(t)U(t)χa−1g − (E0∗)(1− χ)Va(t)g (3.20)

as a distribution in (R× (Rn \B(0, R0))) ∪ (]Ta + 1,∞[×Rn), so that

(∂2t −∆)Ua(t)g ∈ C∞(R× (Rn \B(0, R0))) ∪ (]Ta + 1,∞[×Rn)) (3.21)

Now we shall use a fact about propagation of singularities for solutions ofthe wave equation. Let Ω ⊂ R×Rn be open, u ∈ D′(Ω), (∂2

t −∆)u ∈ C∞(Ω):

24

If (t0, x0) ∈ sing supp (u) then there exists a maximally extended light rayI 3 t 7→ (t0 + t, x0 + tω) in Ω, where I is an open interval containing 0,ω ∈ Sn−1, which passes through (t0, x0) and is contained in sing suppu.

Let (t0, x0) ∈ sing supp (Ua(t)g), with R0 < |x0| < a − 1, t0 > Ta. Fort > Ta, r0 < |x| < a, the term ψa(t)U(t)χa−1g is in C∞, so (t0, x0) mustbelong to sing suppE0∗(1−χ)Va(t)g ⊂ (t, x); (a−|x|)+ ≤ t−Ta. We deducethat t0 > Ta + 1. Also from the fact above, we see that there is a maximallyextended ray in (R × (Rn \ B(0, R0))) ∪ (]Ta + 1,∞[×Rn) passing through(t0, x0), contained in sing supp (Ua(t)g), which is well defined at t = Ta + 1and with a corresponding point (Ta + 1, x(Ta + 1)) with |x(Ta + 1)| ≥ a− 1.But then it is clear that the ray can be further extended backwards for alltimes t < Ta + 1 and the corresponding x(t) stays outside B(0, a− 1). Thisray must be in sing supp Ua(t)g and we reach a contradiction for |t| close to0, since we are then outside the support of Ua(t)g.

In conclusion: Ua(t)g and (E0∗)(1 − χ)Va(t)g are C∞ for t > Ta, R0 <|x| < a− 1. This implies that the last term in (3.19) is C∞ for |x| < a− 1.Moreover, by the strong Huygens principle, it has compact support. So

(P + ∂2t )Ua(t)g = Wa(t)g, (3.22)

where Wa(t)g ∈ C∞0 (]0,∞[;DNcomp) has its support in some fixed compact set

contained in ]Ta, Ta[×suppχ. (When n is even, we loose the compact supportproperty in t.)

By the closed graph theorem,

Wa ∈ C∞0 (]0,∞[;L(H,DNcomp)). (3.23)

We also have the Cauchy data,

Ua(t)g = 0, (∂tUa)(0)g = χa−1g. (3.24)

Finally Ua(t) has the same general regularity properties as Ua:

Ua(t) ∈ Cm(Rt;L(Dj,Dj+1−m

2 )), m ≥ 0, j ≥ 0. (3.25)

(From (3.22), (3.24), we deduce that Ua−Uχa−1 ∈ C∞(R;L(H,Dj)), for allj ≥ 0.)

When Imλ ≥ C 0, we express R(λ) = (P − λ2)−1 by the integralformula (cf. (3.7)):

R(λ) =∫ ∞

0eitλU(t)dt. (3.26)

25

In fact,

−λ2R(λ) =∫ ∞

0∂2t (e

itλ)U(t)dt (3.27)

= −∫ ∞

0∂t(e

itλ)∂tU(t)dt = ∂tU(0) +∫ ∞

0eitλ∂2

tU(t)dt

= 1 +∫ ∞

0eitλ∂2

tU(t)dt,

and it follows that(P − λ2)

∫ ∞0

eitλU(t)dt = 1.

PutRa(λ) =

∫ ∞0

eitλUa(t)dt : H → Dloc, (3.28)

and notice that this operator is well defined and holomorphic for λ ∈ C,since by the strong Huygens principle,

(Ua(t)g)(x) = 0, for t ≥ Ta + |x|. (3.29)

The same calculation gives

(P − λ2)Ra(λ) = χa−1 +∫ ∞

0eitλWa(t)dt. (3.30)

Since χ ≺ χ ≺ χa−1, the x-space projection of suppWa(t) is contained in theregion where χa−1 = 1. Writing

Sa(λ) =∫ ∞

0eitλWa(t)gdt, (3.31)

we get(P − λ2)Ra(λ) = χa−1(1 + Sa(λ)). (3.32)

Lemma 3.3 For every N > 0, 1 + Sa(λ) : H → H, Hcomp → Hcomp isinvertible with inverse O(1) in L(H,H), for

Imλ ≥ −N log |λ|, |λ| ≥ C(N).

Proof. We use the compact support property and (3.23) for Wa, and inte-grate by parts in (3.31):

Sa(λ) =1

(iλ)M

∫ ∞0

eitλ(−∂t)MWa(t)dt.

26

The H → H norm of this operator is ≤ CM |λ|−MeTa(Imλ)− , with the standardnotation b = b+ − b−, b± = max(±b, 0). If Imλ ≥ −N log |λ|, we get

‖Sa(λ)‖L(H,H) ≤ CM |λ|TaN−M ,

which is ≤ 12

if we first take M > TaN , and then |λ| large enough. Noticethat (1 + Sa(λ))−1 = 1− Sa(λ)(1 + Sa(λ))−1 maps Hcomp → Hcomp. #

When Imλ ≥ C 0, we get from (3.32) the relation

R(λ)χa−1 = Ra(λ)(1 + Sa(λ))−1, (3.33)

for R(λ) = (P − λ2)−1, and we can use this relation to get a meromorphicextension of R(λ)χa−1 to C, (since Sa(λ) is compact and holomorphic inλ). If (as before), we let R(λ) denote also the meromorphic extension ofR(λ) : Hcomp → Hloc, then (3.33) extends to λ ∈ C, and we conclude thatR(λ)χa−1 has at most finitely many poles in Imλ ≥ −N log+ |λ|, for everyN ≥ 0. Applying Proposition 2.3, we then get the following result (whichremains true for n even).

Proposition 3.4 In addition to the general bb assumptions of chapter 2, weassume that P ≥ −C and that the non-trapping assumption (3.13) holds.We also assume that n ≥ 3 is odd. Let R(λ) be the meromorphic extensionof (P − λ2)−1 : Hcomp → Dloc. Then for every N ≥ 0, R(λ) has at mostfinitely many poles in Imλ ≥ −N log+ |λ|.

We next estimate the truncated extension of R(λ) and start with Ra.If χ ∈ C∞0 (Rn) is equal to 1 near B(0, R0), then (by the strong Huygensprinciple) χUa(t) vanishes for t ≥ Tχ for some sufficiently large Tχ > 0.Hence

χRa(λ) =∫ Tχ

0eitλχUa(t)dt, (3.34)

and using (3.25), we get after an integration by parts:

‖χRa(λ)‖L(Dj ,Dj) ≤Cj|λ|eTχ(Imλ)− , j ≥ 0. (3.35)

Combining this with (3.33) and Lemma 3.3, we get

‖χR(λ)χa−1‖L(H,H) ≤C

|λ|eTχ(Imλ)− , (3.36)

27

for |λ| ≥ C(N), Imλ ≥ −N log+ |λ|. Here we will need to improve theexponent −1 in |λ|−1 by replacing ‖ · ‖L(H,H) by ‖ · ‖L(Dθ,H)

As a preparation for the promised asymptotic expansion of the wave groupfor t→ +∞, we recall that

R(λ) =∫ ∞

0eitλU(t)dt, Imλ 0,

and we shall establish a corresponding ”Fourier inversion” formula, whichcould be obtained by selfadjoint functional calculus, but which we prefer toderive by direct arguments. For this, we work in a region Imλ ≥ C 0 andstart by getting some further estimates on R(λ).

First, by (3.7), (3.26), we get

‖R(λ)‖L(Dj ,Dj+

12 )≤ O(1), (3.37)

and after integration by parts in (3.26),

‖R(λ)‖L(Dj ,Dj) ≤O(1)

〈λ〉. (3.38)

From the first of the two equivalent identities

PR(λ) = 1 + λ2R(λ), R(λ) = − 1

λ2+

1

λ2PR(λ), (3.39)

and (3.38), we get‖R(λ)‖L(Dj ,Dj+1) ≤ O(1)〈λ〉. (3.40)

Use (3.39), (3.37) to get

‖R(λ)‖L(Dj ,Dj−1/2) ≤O(1)

〈λ〉2. (3.41)

The last two estimates are the most extreme ones, and (3.37), (3.38) can beobtained from these two by interpolation (cf (3.2)) and we get more generally

‖R(λ)‖L(Dj ,Dj−α) ≤O(1)

〈λ〉1+2α, −1 ≤ α ≤ 1

2. (3.42)

We claim that U(t) = U(t), where

U(t) =1

∫Im =C

e−itλR(λ)dλ, t ≥ 0. (3.43)

28

The integral converges in L(Dj,Dj−ε) for every ε > 0 and becomes a contin-uous function of t with values in that space. Using (3.39), we get

U(t)g =1

∫Imλ=C

e−itλ(− 1

λ2+R(λ)

λ2P )gdλ

=1

∫γe−itλ

1

λ2dλ g +

1

∫Imλ=C

e−itλR(λ)

λ2Pgdλ

= tg +1

∫Imλ=C

e−itλR(λ)

λ2Pgdλ,

where γ is a small positively oriented circle around 0.In the last integral, we gained a sufficiently large negative power of λ, to

be able to apply ∂2 and get

∂2t U(t)g = − 1

∫Imλ=C

e−itλR(λ)Pgdλ = −PU(t)g,

i.e.(P + ∂2

t )U(t)g = 0. (3.44)

Further,

U(0)g =1

∫Imλ=C

R(λ)

λ2Pgdλ = 0,

by contour deformation upwards, and

∂tU(0)g = g − i

∫Imλ=C

R(λ)

λPgdλ = g,

by the same contour deformation.We have then verified that U(t) solves the same Cauchy problem as U ,

so we have U = U :

U(t) =1

∫Imλ=C

e−itλR(λ)dλ, t ≥ 0, (3.45)

where the integral converges in L(Dj,Dj−ε) for every ε > 0.We want to multiply this to the left by χ and to the right by χa−1, and

then make contour deformation, and use an estimate like (3.36), but with alittle faster decay than |λ|−1, so we start by obtaining such improved decay.For g ∈ D we write

χR(λ)χa−1g = χ(− 1

λ2+R(λ)

λ2χa−1P )χa−1g,

29

where χa−1 ≺ χa−1 ∈ C∞0 (B(0, a−1)). Using (3.36) (with χa−1 there replacedby χa−1) we get for the same λ as there,

‖χR(λ)χa−1‖L(D,H) ≤C

|λ|2eTχ(Imλ)− . (3.46)

(A more optimal estimate would probably be to have L(D1/2,H) instead ofL(D,H).) Interpolating between (3.36) and (3.46), we get

‖χR(λ)χa−1‖L(Dθ,H) ≤C

|λ|1+θeTχ(Imλ)− , (3.47)

for λ as in (3.36) and 0 ≤ θ ≤ 1.With this in mind, we write

χU(t)χa−1g =1

∫Imλ=C

e−itλχR(λ)χa−1dλ (3.48)

=1

∫Imλ=−A

e−itλχR(λ)χa−1dλ+∑

Imλ>−Aresidues,

where we choose A > 0 such that no resonances has the imaginary part −A.The L(Dθ,H)-norm of the integral in the last expression in (3.48) can beestimated by

OA(1)e−tA+TχA = OA(1)e−At.

Let λ0 ∈ C, be a pole of the meromorphic extension of (P − λ2)−1, sothat in the case λ0 6= 0 (cf (2.48)) we have near λ0:

−R(λ) =ω(λ0)∑

1

Π−j(λ2 − λ2

0)j+ Hol(λ) (3.49)

In the case Imλ0 ≤ 0 we discussed some of the structure of the ”spectralprojections” Π−j. When Imλ0 > 0 we know that λ0 is purely imaginaryand that λ2

0 is a negative eigenvalue of P . In this case Π−1 is the ordinaryorthogonal spectral projection, while Π−j = 0 for j ≥ 2. In the case λ0 = 0things are a little more complicated and we make no further comments inthat case. For λ0 6= 0 the corresponding contribution to the sum of residuesin (3.48) is

ω(λ0)∑1

1

∫γ0

e−itλ1

(λ2 − λ20)jdλχΠ−jχa−1, (3.50)

30

where γ0 is a small positively oriented circle around λ0. The integral is ofthe form e−itλ0 times a polynomial of degree at most j − 1 in t. For j = 1,we get

1

2π2πi

e−itλ0

2λ0

χΠ−1χa−1 =ie−itλ0

2λ0

χΠ−1χa−1. (3.51)

We have proved the following result essentially due to Lax and Phillips witha more straight forward proof due to Vainberg [91], extended to the black boxframework by Tang-Zworski [90], that we have modified a little to decreasethe number of cut-offs:

Theorem 3.5 Let n ≥ 3 be odd. Let P be a black box operator as in chapter2 and assume (3.1), (3.13). Then for every A > 0 such that −A is notthe imaginary part of any resonance, and for every χ ∈ C∞0 (Rn) equal to aconstant near B(0, R0), we have

χU(t)χ =∑

λ∈Res (P )∪iµ;µ>0,−µ2∈σ(P ))e−itλχpλ,P (t)χ+ rχ,A(t),

where pλ,P (t) is a polynomial in t with values in the operators Hcomp → Dloc.When λ0 is a simple pole of the resolvent and λ0 6= 0, we have pλ,P (t) =i

2λΠ−1,λ, where Π−1,λ is the corresponding spectral projection. The remainder

term satisfies ‖rχ,A(t)‖L(Dθ,H) ≤ Cχ,A,θe−tA, for every θ > 0.

According to Tang-Zworski, Burq, there is a trick of C. Morawetz, whichallows us to replace Dθ by H in the above theorem.

4 Absence of resonances exponentially close

to R.

In this chapter we shall descibe a recent result of N. Burq [15] which says thatunder quite general assumptions there are no resonances exponentially closeto the real axis. We start by formulating the precise theorem. Let O ⊂⊂ Rn

be an open set with a smooth (C∞) boundary ∂O and assume that Rn \ Ois connected. Let

P =∑|α|≤2

aα(x)Dα (4.1)

be an elliptic formally selfadjoint operator with smooth coefficients on Rn\O.We assume that

P = −∆ for |x| > R0 − 1, (4.2)

31

for some sufficiently large R0 > 0, that we choose large enough so thatO ⊂ B(0, R0 − 1).

We explain the main ideas of the proof of the following result of N. Burq[15].

Theorem 4.1 Let P also denote the Dirichlet realization of P in L2(Rn\O).Then there exists a constant C > 0, such that P has no resonances in the setof λ ∈ C such that

|λ| ≥ C, Imλ ≥ −e−C|Reλ|. (4.3)

Actually, Burq also establishes this result for the Neumann realizationof P and even in the more general case when P is realized with Dirichletconditions on one connected part of the boundary and with Neumann condi-tions on the other part. As a corollary, Burq shows that the local energy ofsolutions to the wave equation with compactly supported initial data decaysat least as fast as (log(1 + t))−k, t → +∞, where k ≥ 0 depends on thesmoothness of the Cauchy data at t = 0. Here smoothness of the data ismeasured in terms of being in the domain of a certain power of the generatorof the wave group, when the latter is considered as a matrix solution to a firstorder system. Earlier results in this direction were obtained by E. Harrell[36] and C. Fernandez and R. Lavine [24]. After the work [15] there havebeen extensions by G. Vodev [97, 98], and by Burq [16]. Vodev observedsome simplifications in Burq’s original proof, and we shall use one of themin our presentation, as well as a new estimate (that will not be explainedin detail here) that is planned to appear in a joint work with Burq. Thisestimate permits to give a treatment near infinity which is more natural andcloser to the intuition.

The first step in the proof is to notice that it suffices to show that for realλ with |λ| ≥ 1, we have

‖R(λ)v‖H1(B(0,R0)\O) ≤ C0eC0|λ|‖v‖L2 , (4.4)

for v ∈ L2comp(B(0, R0) \ O) for some C0 > 0. In fact, assuming (4.4),

we shall extend R(µ) holomorphically to µ in an exponentially small disccentered at λ as an operator L2

comp(Rn \ O) → (H2 ∩ H10 )loc(R

n \ O). Let1B(0,R0−1) ≺ χ0 ≺ χ1 ≺ χ2 ≺ 1B(0,R0), with χj ∈ C∞0 . As an approximationfor R(µ)χ1, we take

A(µ) := χ2R(λ)χ1 −R0(µ)[P0, χ2]R(λ)χ1.

32

Then

(P − µ2)A(µ) = (4.5)

χ1 + [P0, χ2]R(λ)χ1 + χ2(λ2 − µ2)R(λ)χ1 − [P0, χ2]R(λ)χ1

= χ1 + χ2(λ2 − µ2)R(λ)χ1.

For the exterior part of the approximate resolvent, we first compute

(P − µ2)(1− χ0)R0(µ)(1− χ1) = 1− χ1 − [P0, χ0]R0(µ)(1− χ1).

Correcting for the last term, we are led to the approximation of R(µ)(1−χ1) :

B(µ) := (1− χ0)R0(µ)(1− χ1) + A(µ)[P0, χ0]R0(µ)(1− χ1).

Then

(P − µ2)B(µ) = 1− χ1 − [P0, χ0]R0(µ)(1− χ1)

+(χ1 + χ2(λ2 − µ2)R(λ)χ1)[P0, χ0]R0(µ)(1− χ1)

= (1− χ1) + χ2(λ2 − µ2)R(λ)[P0, χ0]R0(µ)(1− χ1).

Put R(µ) = A(µ) +B(µ). Then

(P − µ2)R(µ) = 1 +K, (4.6)

K = χ2(λ2 − µ2)R(λ)(χ1 + [P0, χ0]R0(µ)(1− χ1)). (4.7)

Using the Fourier transform and complex deformation of integration contoursas in section 2.1, we see that

[P0, χ0]R0(µ) : L2comp(B(0, R0) \ O)→ L2

comp(B(0, R0) \ O)

is of norm ≤ CeC|µ|. It follows that

K : L2comp(B(0, R0) \ O)→ L2

comp(B(0, R0) \ O)

and has operator normO(1)|λ|eC0|λ||λ−µ|, so if |λ−µ| ≤ e−2C0|λ| and |λ| 1,1 +K has a bounded inverse and we get the holomorphic extension

R(µ) = R(µ)(1 +K)−1 : L2comp(B(0, R0) \ O)→ (H2 ∩H1

0 )loc(B(0, R0) \ O).(4.8)

33

Proposition 2.3 then shows that R(µ) extends holomorphically also as anoperator on L2

comp(Rn \ O).One of the main steps of the proof is to establish weighted L2 estimates

in a bounded domain for (P − λ2) = λ2(h2P − 1), with h = 1/|λ|. Theseestimates are so called Carleman estimes, developed with enormous successby Hormander and others in different branches of partial differential equa-tions, including uniqueness problems and Cauchy-Riemann type equations.In the context of spectral problems, they have been used by G. Lebeau andL. Robbiano [52, 53]. Much works in a more general frame work, so to startwith, we take a more general semi-classical operator.

4.1 Local Carleman estimates away from the bound-ary.

Let Ω ⊂ Rn be some open set. Let P be a semiclassical differential operatorof order 2, formally self-adjoint and of the form:

P =∑|α|≤2

aα(x;h)(hDx)α, aα(x;h) ∈ C∞(Ω). (4.9)

Assume thataα(x;h) = a0

α(x) +O(h) in C∞(Ω), (4.10)

and that aα is independent of h for |α| = 2. The semiclassical principalsymbol of P is then by definition:

p(x, ξ) =∑|α|≤2

a0α(x)ξα = a(x, ξ) +

∑|α|≤1

a0α(x)ξα. (4.11)

Here a(x, ξ) =∑|α|=2(...) is equal to the ordinary principal symbol of the

operator h−2P . We assume that P is elliptic in the classical sense:

a(x, ξ) ≥ 1

C|ξ|2. (4.12)

Let φ ∈ C∞(Ω; R) and consider

Pφ = eφ(x)/hPe−φ(x)/h. (4.13)

This operator is also of the form (4.9) with the same ordinary principalsymbol h2a(x, ξ) and with the new semiclassical principal symbol

pφ(x, ξ) = p(x, ξ + iφ′(x)). (4.14)

34

Assume that we have

pφ(x, ξ) = 0⇒ 1

ipφ, pφ < 0, (4.15)

where

f, g =n∑j=1

(∂f

∂ξj(x, ξ)

∂g

∂xj(x, ξ)− ∂f

∂xj(x, ξ)

∂g

∂ξj(x, ξ)),

denotes the Poisson bracket of two differentiable functions f, g. Notice that1if, f = −2Re f, Im f is a real-valued function.

The basic Carleman estimate is then given by:

Lemma 4.2 Let W ⊂⊂ Ω be open. Then there exist constants C0, h0 > 0such that for all 0 < h ≤ h0 , u ∈ C∞0 (W ):

h1/2‖u‖H2 ≤ C0‖Pφu‖. (4.16)

Here we let Hs denote the semi-classical Sobolev space equipped with the h-dependent norm ‖u‖Hs = ‖〈hD〉su‖.

We will only give an outline of the proof, mainly for readers who arefamiliar with pseudodifferential operator machinery especially in the semi-classical setting. However, we point out that some basic ideas are older thanthe theory of pseudodifferential operators. Start by looking at the trivialidentity:

1

h‖Pφu‖2 ≥ 1

h‖Pφu‖2 − 1

h‖P ∗φu‖2 (4.17)

=1

h((P ∗φPφ − PφP ∗φ)u|u) = (

1

h[P ∗φ , Pφ]u|u).

Here the commutator has the form

1

h[P ∗φ , Pφ] =

∑|α|≤3

bα(x;h)(hDx)α,

withbα(x;h) = b0

α(x) +O(h) in C∞(Ω),

and the associated semi-classical principal symbol is

1

ipφ, pφ

35

which is > 0 on p−1φ (0). Since |pφ(x, ξ)| ∼ 〈ξ〉2 for large ξ, we can find

constants b, C > 0, such that

1

ipφ, pφ+ b|pφ(x, ξ)|2 ≥ 1

C〈ξ〉4 on W ×Rn,

where we choose W ⊂⊂ Ω open with W ⊂⊂ W .Using h-pseudodifferential calculus and/or a suitable version of Garding’s

inequility (see for instance [21]), we deduce that

(1

h[P ∗φ , Pφ]u|u) + b‖Pφu‖2 ≥ 1

2C‖u‖2

H2 , u ∈ C∞0 (W ),

We get

‖u‖2H2 ≤ 2C((

1

h[P ∗φ , Pφ]u|u) + b‖Pφu‖2). (4.18)

Combining this with (4.17), we get (4.16). #

Actually, we shall not use the full strength of (4.16) but only the weakerestimate

h1/2‖u‖H1 ≤ OW (1)‖Pφu‖, u ∈ C∞0 (W ). (4.19)

Taking squares and recalling the definition of Pφ, we get

h∫

(e2φ/h|u|2 + |h∇x(eφ/hu)|2)dx ≤ OW (1)

∫e2φ/h|Pu|2dx. (4.20)

This can also be written:

h∫e2φ/h(|u|2 + |h∇xu|2)dx ≤ OW (1)

∫e2φ/h|Pu|2dx. (4.21)

4.2 Local Carleman estimates at the boundary

Now let Ω ⊂ Rn be closed, bounded and have a smooth boundary ∂Ω, Letx0 ∈ ∂Ω and assume that P, φ are as above in a neighborhood W (in Rn)of x0. We want to have (4.21) for u ∈ C∞0 (Ω ∩W ) with u|∂Ω

= 0. Usingthe ellipticity, it is quite classical that we can find local coordinates x1, .., xncentered at x0, such that Ω is defined in W by xn ≥ 0, and the (new) principalsymbol becomes

a(x, ξ) = ξ2n + r(x, ξ′), ξ′ = (ξ1, .., ξn−1). (4.22)

36

SoP = (hDxn)2 +R(x, hDx′ ;h) + b(x;h)hDxn . (4.23)

The Lebesgue measure becomes λ(x)dx for some smooth and strictly posi-tive function λ(x). Define Pφ as above. It is now practical to organize thecomputations a little differently, so we write

Pφ = Q2 + iQ1, Q2 =1

2(Pφ + P ∗φ), Q1 =

1

2i(Pφ − P ∗φ).

Here the * indicates that the we take the formal adjoint in L2(W ;λ(x)dx1..dxn)and norms and scalar products will be in this space for a while. The subscriptj in Qj indicates that we have a differential operator of order j.

Then for all u ∈ C∞0 (Ω ∩W ) with u|∂Ω= 0, we get

‖Pφu‖2 = ‖Q2u‖2 + ‖Q1u‖2 + i[(Q1u|Q2u)− (Q2u|Q1u)

]. (4.24)

If we could move the Qj from the right to the left in each of the two scalarproducts, the bracket in (4.24) would become

(i[Q2, Q1]u|u) =1

2([P ∗φ , Pφ]u|u),

and we can hope to estimate this from below as before. (The presence of thefactor 1/2 is explained by the fact that ‖Q2u‖2 + ‖Q1u‖2 is not in generalequal to ‖P ∗φu‖2, but rather to 1

2‖Pφu‖2 + 1

2‖P ∗φu‖2.) However, we have to

take into account the boundary terms, when carrying out the integrations byparts that pull over the Qj to the left. Because of the Dirichlet condition,we get no boundary integral, when moving over Q1, while Q2 will give riseto a boundary term from its (hDxn)2 term. Notice that Q2 = (hDxn)2 plusterms which contain hDxn at most to the power 1. Hence,

i(Q1u|Q2u) = i(Q2Q1u|u) + h∫xn=0

(Q1u)hDxnuλ(x′, 0)dx′. (4.25)

Only the hDxn term in Q1 can give a non-vanishing contribution to theboundary integral (because of the Dirichlet condition) so we compute thisterm, using (4.23):

Pφ = (hDxn + i∂xnφ)2 +Rφ(x, hDx′ ;h) + b(x;h)(hDxn + i∂xnφ)

= (hDxn)2 − (∂xnφ)2 + i((∂xnφ) hDxn + (hDxn) (∂xnφ)) +

Rφ(x, hDx′ ;h) + bhDxn + ib∂xnφ

= P − (∂xnφ)2 + (Rφ −R) +

i((∂xnφ) hDxn + (hDxn) (∂xnφ)) + ib∂xnφ.

37

Now use that (hDxn)∗ = hDxn + hµ(x) for some smooth function µ (whichdepends on λ) and

P ∗φ = P − (∂xnφ)2 + (R∗φ −R∗)− i((hDxn)∗ (∂xnφ) + (∂xnφ)(hDxn)∗)

−ib∂xnφ,

to conclude that the hDxn term in Pφ − P ∗φ is 4i(∂xnφ)hDxn , so

Q1 =1

2i(Pφ − P ∗φ) = 2(∂xnφ)hDxn + a tangential operator, (4.26)

where by a tangential operator we mean one which contains differentiationswith respect to x1, .., xn−1 but not with respect to xn.

Now (4.25) becomes

i(Q1u|Q2u) = i(Q2Q1u|u) + 2h∫xn=0

(∂xnφ)|hDxnu|2λ(x′, 0)dx′. (4.27)

Using this in (4.24), we get

‖Pφu‖2 = ‖Q2u‖2 + ‖Q1u‖2 + (4.28)

2h∫xn=0

(∂xnφ)|hDxnu|2λ(x′, 0)dx′ + (i[Q2, Q1]u|u),

so we can hope to have (4.21) also in the boundary case, provided that

∂xnφ|xn=0> 0. (4.29)

A coordinate free formulation of the last condition is that

∂φ

∂ν |∂Ω

> 0, (4.30)

where ν denotes the interior unit normal of ∂Ω, for the metric given by thequadratic form on the tangent space which is dual to a(x, ξ).

The treatment of the last term in (4.28) becomes more technical becauseof the boundary (and requires also the use of tangential pseudodifferentialoperators and a division trick going back to a work of V. Ivrii). This wascarried out by Lebeau–Robbiano [52, 53] also in the case of Neumann bound-ary conditions. We will not go into the technical details and simply state theresult, as it was formulated by Burq [15]: Let P be as above, but of a morerestricted form:

P =∑|α|=2

aα(x)(hDx)α + h

∑|α|≤1

bα(x;h)(hDx)α − 1, (4.31)

with bα(x;h) = O(1) in C∞(Ω). Then we have

38

Proposition 4.3 Let P, φ,Ω be as above (including (4.30), (4.31)) and letW be a sufficiently small neighborhood of x0 ∈ ∂Ω. Then for h > 0 smallenough, and all u ∈ C∞0 (Ω ∩W ) with u|∂Ω

= 0, we have (4.20), (4.21), withΩ ∩W as the integration domain.

4.3 Gluing local Carleman estimates for a given φ.

Let O ⊂ Rn be closed with smooth boundary, and let W ⊂⊂ Rn be open.Let P be of the form (4.9) near W and let φ be smooth and real near W .Assume that W is covered by finitely many open sets W1, ..,WN such that(4.20) (or equivalently (4.21)) holds for u ∈ C∞0 (Ω∩Wj) with u|∂Ω

= 0. Thenwe shall see that it also holds for all u ∈ C∞0 (Ω ∩W ) with u|∂Ω

= 0.

Indeed, let 0 ≤ χj ∈ C∞0 (Wj) with∑N

1 χ2j = 1 near W and apply (4.21)

to χju:

h∫e2φ/h(|χju|2 + |χjh∇xu|2)dx ≤ (4.32)

O(1)∫e2φ/h|χjPu|2dx +O(1)

∫e2φ/h|[P, χj]u|2dx

+O(1)∫e2φ/h|[h∇x, χj]u|2dx.

Summing and using that [P, χj] = h∑|α|≤1 bα,j(x;h)hDx, we get

h∫e2φ/h(|u|2 + |h∇u|2)2dx ≤ O(1)

∫e2φ/h|Pu|2dx+ (4.33)

O(1)h2∫e2φ/h(|u|2 + |h∇xu|2)dx.

For h small enough, the last term can be absorbed by the first member ofthe inequality. #

4.4 Construction of φ satisfying the Poisson bracketcondition.

Our function φ will satisfy

|dφ| ≥ Const. > 0. (4.34)

39

Write

p(x, ξ) = a(x, ξ) + p1(x, ξ) + p0(x), a(x, ξ) = 〈A(x)ξ, ξ〉. (4.35)

We will replace φ by τφ for τ > 0 large enough. Then

pτφ(x, ξ) = p(x, ξ + iτφ′x) = (4.36)

a(x, ξ)− τ 2a(x, φ′) + 2iτ〈A(x)ξ, φ′(x)〉+ p1(x, ξ) + iτp1(x, φ′(x)) + p0(x)

= τ 2(q2(x, ξ) + 2iq1(x, ξ)),

with

q2(x, ξ) = a(x,ξ

τ)− a(x, φ′x) +

1

τp1(x,

ξ

τ) +

1

τ 2p0(x),

q1(x, ξ) = 〈A(x)ξ

τ, φ′(x)〉+

1

2τp1(x, φ′(x)).

Let us look for the commons zeros of the real and imaginary parts: Firstnotice that at such a common zero, we must have |ξ| ∼ τ |φ′(x)|, uniformlywith respect to φ, for τ sufficiently large. Then we get

a(x, ξτ)− a(x, φ′(x) = Oφ(1/τ)

〈A(x) ξτ, φ′(x)〉 = Oφ(1/τ).

(4.37)

At such common zeros, we compute

∂q2

∂ξ=

1

τ2A(x)

ξ

τ+Oφ(

1

τ 2),

∂q2

∂x=

∂a

∂x(x,

ξ

τ)− ∂a

∂x(x, φ′(x))− 2φ′′(x)A(x)φ′(x) +Oφ(

1

τ),

∂q1

∂ξ=

1

τA(x)φ′(x),

∂q1

∂x= φ′′(x)A(x)

ξ

τ+ 〈∂A

∂x

ξ

τ, φ′(x)〉+Oφ(

1

τ),

leading to

q2, q1 = (4.38)

= 〈1τ

2A(x)ξ

τ, φ′′(x)A(x)

ξ

τ〉+ 〈1

τ2A(x)

ξ

τ, 〈∂A(x)

∂x

ξ

τ, φ′(x)〉〉

−〈∂a∂x

(x,ξ

τ)− ∂a

∂x(x, φ′(x))− 2φ′′(x)A(x)φ′(x),

1

τA(x)φ′(x)〉+Oφ(

1

τ 2)

=2

τ

[〈A(x)

ξ

τ, φ′′(x)A(x)

ξ

τ〉+ 〈A(x)φ′(x), φ′′(x)A(x)φ′(x)〉+O(|φ′(x)|3)

]+Oφ(

1

τ 2).

40

Now look for φ of the form

φ(x) = eβψ(x), (4.39)

where dψ(x) 6= 0. Then

φ′(x) = eβψ(x)βψ′(x), (4.40)

φ′′(x) = eβψ(x)β2ψ′(x)⊗ tψ′(x) + eβψ(x)βψ′′(x), (4.41)

and (4.38) gives with ξ = τ |φ′(x)|η, |η| ∼ 1:

q2, q1 =2

τe3βψ

[β4|ψ′(x)|2〈A(x)η, (ψ′(x)⊗ tψ′(x))A(x)η〉 (4.42)

+β4〈A(x)ψ′(x), (ψ′(x)⊗ tψ′(x))A(x)ψ′(x)〉+O(β3)]

+Oβ,ψ(1

τ 2).

Here ψ′(x)⊗ tψ′(x) ≥ 0 and

〈A(x)ψ′(x), (ψ′(x)⊗ tψ′(x))A(x)ψ′(x)〉 = 〈A(x)ψ′(x), ψ′(x)〉2 > 0,

so with β > 0 large enough, and for τ large enough depending on β, we get(4.15) with φ replaced by τφ.

4.5 Estimates in a ball.

Recall the situation in Burq’s theorem: O ⊂⊂ Rn is open with smoothboundary, Ω = Rn \ O is connected, and P =

∑|α|≤2 aα(x;h)Dα is a sec-

ond order, elliptic, formally self-adjoint operator with smooth coefficients inC∞(Ω). Moreover P is equal to −∆ outside some ball B(0, R0), with R0 > 0.We may assume that O ⊂ B(0, R0). We also recall that we take the Dirich-let realization of P , i.e. we realize P as a self-adjoint operator L2 → L2

with domain H2 ∩H10 )(Ω). We are interested in estimates for P − λ2, when

λ→ +∞. Put h = 1/λ, and write

(P − λ2) = λ2(h2P − 1). (4.43)

h2P − 1 is then then of the type that we considered above.

Lemma 4.4 Let R2 > R0. Then there exists ψ ∈ C∞(B(0, R2)\O) such thatψ = Const and ∂ψ

∂ν> 0 on ∂O, and dψ(x) is 6= 0 for all x ∈ B(0, R2) \ O).

41

Proof. First, we can find a smooth real function ψ0 on B(0, R2) \ O) suchthat ψ0 = 0 and ∂ψ0

∂ν> 0 on ∂O. Recall that if M is a smooth manifold and f :

M → R is a smooth function, then x0 ∈M is called a critical point, if df(x0)vanishes. Such a critical point is non-degenerate if the Hessian f ′′(x0) whichcan be defined in local coordinates or more generally as a linear map from thetangent space Tx0M into itself, is invertible. A Morse function is a smoothfunction f as above for which all critical points are non-degenerate. It is awell-known fact in differential topology that the Morse functions on M forma set which is dense in C∞(M ; R). Let ψ1 be a smooth real-valued extensionof ψ0 to a neighborhood of B(0, R2)\O, and let ψ2 be a Morse function on thesame neighborhood which is very close to ψ0 in the norm of C1(B(0, R2)\O).Let ψ3 be the restriction of ψ2 to B(0, R2)\O. Then the critical points of ψ3

are of finite number and away from ∂O. We can modify ψ3 into a new smoothfunction ψ4 on B(0, R2) \ O), which has the same critical points as ψ3 and

such that ψ4(x) = 0, ∂ψ4(x)∂ν

> 0 on ∂O. Let x1, .., xN ∈ B(0, R2) \ O) be thecritical points of ψ4. Let [0, 1] 3 t 7→ xj(t) ∈ B(0, R2+1)\O be smooth curveswith xj(0) = xj, xj(1) ∈ ∂B(0, R2 + 1). Then we can construct a smoothfamily of diffeomorphisms κt : B(0, R2) \ O → Ωt \ O, where Ωt ⊃ B(0, R2),such that κt(x) = x for x near ∂O and such that κt(xj) = xj(t). Then thelemma follows, if we let ψ be the restriction to B(0, R2) \ O) of ψ4 κ−1

1 . #

From ψ, we form φ = τeβψ as in section 4.4, and get the estimate:

h∫e2φ/h(|u|2 + |h∇u|2)dx ≤ O(1)

∫e2φ/h|(h2P − 1)u|2dx, (4.44)

for all u ∈ C∞0 (B(0, R2) \ O) with u|∂O = 0. Recalling that h = 1/λ, we get

λ3∫e2λφ(|u|2 + |1

λ∇u|2)dx ≤ O(1)

∫e2λφ|(P − λ2)u|2dx. (4.45)

Here we want to remove the assumption that u has compact support.Take R1 with R0 < R1 < R2, and take χ with 1B(0,R1) ≺ χ ∈ C∞0 (B(0, R2)),and apply (4.45) to χu:

λ3∫B(0,R1)\O

e2λφ(|u|2 + |1λ∇u|2)dx ≤ (4.46)

O(1)∫B(0,R2)\O

e2λφ|(P − λ2)u|2dx+O(1)λ2∫B(0,R1,R2)

e2λφ(|u|2 + |1λ∇u|2)dx,

where B(0, R1, R2) = x ∈ Rn; R1 < |x| < R2 denotes the open shell ofcenter 0 and radii R1, R2. To the left we estimate the exponential factor from

42

below by e−λC and to the left we estime the exponenential factors from aboveby eλC , where C > 0 is a sufficiently large constant. It follows that there isa constant C > 0 such that∫B(0,R1)\O

(|u|2 + |1λ∇u|2)dx ≤ (4.47)

O(1)e2λC∫B(0,R2)\O

|(P − λ2)u|2dx+O(1)e2λC∫B(0,R1,R2)

(|u|2 + |1λ∇u|2)dx.

4.6 Estimate in a shell.

We work in the region |x| > R0, where h2P − 1 = −h2∆ − 1 has the semi-classical symbol p = ξ2− 1. In polar coordinates, x = rω, ω ∈ Sn−1, r > R0,we get

p = ρ2 +(ω∗)2

r2− 1,

where (ω∗)2 ≥ 0 denotes the principal symbol of −∆Sn−1 . We want to makeCarleman estimates with radial weight φ = φ(r). We get

pφ = (ρ+ iφ′)2 +1

r2(ω∗)2 − 1,

so Re pφ = ρ2 − (φ′)2 + (ω∗)2

r2 − 1Im pφ = 2ρφ′,

(4.48)

1

4Re pφ, Im pφ = ρ2φ′′ + (φ′′φ′ +

(ω∗)2

r3)φ′.

We need the last Poisson bracket to be > 0 on the joint characteristics whereboth expressions in (4.48) are equal to 0. We shall have φ′ > 0, so pφ = 0 isequivalent to

ρ = 0,(ω∗)2

r2= 1 + (φ′)2. (4.49)

On this set, we need

φ′ > 0, φ′′φ′ +1 + (φ′)2

r> 0. (4.50)

With ψ = (φ′)2 we get the linear system of inequalities,

1

2ψ′ +

1 + ψ

r> 0, ψ > 0.

43

This will follow from

ψ′ ≥ −2

r, ψ > 0.

Start by taking ψ = 2(A − log r), A 0, for 1 ≤ r ≤ B < eA. Choosing Bvery close to eA, we may have ψ(B) > 0 as small as we like. We can extendthe definition of ψ to all r in such a way that ψ > 0 and

−2

r≤ ψ′ ≤ 0,

with equality to the right for r ≥ eA. Then ψ = ε2 for r ≥ eA with ε > 0 assmall as we like.

In conclusion, we can find a smooth real function φ on [1,+∞[ satisfying(4.50), and such that

φ′(r) =√

2(A− log r), 1 ≤ r ≤ eA − 1, (4.51)

φ′ is decreasing on [eA − 1,+∞[ and equal to ε for r ≥ eA. (4.52)

Let R3 > R2. Then for u ∈ C∞0 (B(0, R0, R3)) (the open shell), we havethe Carleman estimate,

λ3∫e2λφ(|u|2 + |1

λ∇u|2)dx ≤ OA,ε,R3,R0(1)

∫e2λφ|(P − λ2)u|2dx. (4.53)

4.7 Estimate in a ball ∪ shell.

We shall combine (4.47), valid for functions on B(0, R2) \ O, with (4.53).We choose R1, R2 so that R0 < R1 − 2 and R2 = R1 + 1. Choose φ asin section 4.6 and such that φ(R1) − φ(R1 − 1) is larger than the constantC in the exponents in (4.47). Modifying φ by a constant, we get φ ≤ −ε0,|x| ≤ R1 − 1, φ ≥ C + ε0, |x| ≥ R1, for some ε0 > 0.

Let u ∈ C∞(B(0, R3) \ O), u|∂O = 0, with

(P − λ2)u = v ∈ C∞0 (B(0, R0) \ O). (4.54)

Then from (4.47), (4.53), we get:∫B(0,R1)\O

(|u|2 +1

λ∇u|2)dx ≤ (4.55)

O(1)e2λC∫|v|2dx+O(1)e2λC

∫B(0,R1,R2)

(|u|2 + |1λ∇u|2)dx,

44

∫B(0,R1,R3−1)

e2λφ(|u|2 + |1λ∇u|2)dx ≤ O(1)× (4.56)

( ∫B(0,R1−2,R1−1)

e2λφ(|u|2 + |1λ∇u|2)dx+

∫B(0,R3−1,R3)

e2λφ(|u|2 + |1λ∇u|2)dx

).

There is a δ > 0 such that the last term of the right hand side of (4.55)can be estimated by e−λδ times the left hand side of (4.56) and the first termof the right hand side of (4.56) can be estimated by e−λδ times the left handside of (4.55), when λ > 0 is large enough. So, if we sum (4.55), (4.56), weget rid of these two contributions to the the right hand sides and with

φ+(x) =

0, |x| < R1,φ(x), |x| ≥ R1,

we get ∫B(0,R3−1)\O

e2λφ+(|u|2 + |1λ∇u|2)dx ≤ (4.57)

O(1)e2λC∫|v|2dx+O(1)

∫B(0,R3−1,R3)

e2λφ+(|u|2 + |1λ∇u|2)dx,

for u, v as in (4.54).

4.8 Estimates for outgoing solutions.

In the previous estimate, we may assume that R3 is as large as we like, andthat φ(x) = Const. + ε|x|, in B(0, R3

4,∞) with ε as small as we like. (The

choice of ε, R3, will only affect the ”O(1)” factors and the required largenessof λ, but not the constant C in the exponent.) Let u be as before, but assumein addition that u is outgoing: u = R(λ)v, where v ∈ L2

comp(B(0, R0) \ O).For 1supp v∪O ≺ χ ∈ C∞0 (Rn; [0, 1]), we have

(1

i[P, χ]u|u) =

1

i((χu|Pu))− (Pu|χu)) (4.58)

=λ2

i((χu|u)− (u|χu)) +

1

i((χu|v)− (v|χu))

=1

i((u|v)− (u|v)) = 2Im (u|v).

By the Cauchy–Schwartz inequality, we have

(1

i[P, χ]u|u) ≤ 2‖u‖B(0,R0)‖v‖B(0,R0). (4.59)

45

(Recall that u is outgoing, so that u = R0(λ)w for |x| ≥ R0, for some

w ∈ C∞0 (B(0, R0) \ O).

Then by Proposition 2.5

(1

i[P, χ]u|u) = C(λ)

∫∂B(0,λ)

|w(ξ)|2L∂B(0,λ)(dξ),

where C(λ) > 0 and we assume that λ > 0 in order to fix the ideas. So thescalar product in (4.59) is ≥ 0.)

The semi-classical principal symbol of 1i[hP, χ] is −Hp(χ) = −2ξ ·∂xχ(x),

which is ≥ 0 on the outgoing trajectories (or ”rays”). Here by an outgo-ing ray, we mean a maximal(ly extended) integral curve (necessarily a linesegment) of Hp = 2ξ · ∂x in

(x, ξ) ∈ R2n; ξ2 = 1,R3

2< |x| < 2R3, (4.60)

such that the backward extension (for negative times, as an integral curve ofHξ2), passes over B(0, R3

4). By an incoming ray, we mean a maximal ray in

the set (4.60) which is not an outgoing one.If we work in the shell R3

2≤ |x| ≤ 2R3, and consider relative sizes com-

pared to ‖u‖H1(B(0,

R34,R33

)), then by using phase space analysis, one can show

that ”u is O(e−λR3/C)” near the incoming trajectories (this is a consequenceof the fact that u is outgoing, see [38]) and ”Oδ(eδλR3)” for every δ > 0, nearthe outgoing trajectories. Also u is exponentially localized to the energy sur-face ξ2 = 1, when we restrict x to a shell as above. In view of these facts, oneexpects (and can indeed show) that (1

i[P, χ]u|u) behaves like the square of a

norm of u. More precisely we have ([17]) for every δ > 0 and for λ ≥ λ(δ):∫B(0,R3−1,R3)

(|u|2 + |1λ∇u|2)dx ≤ (4.61)

e−2λR3/C0

∫B(0,

R34,R33

)(|u|2 + |1

λ∇u|2)dx+ e2R3δλ(

1

i[P, χ]u|u),

where C0 > 0. Combining this with (4.57), (4.59) leads to∫B(0,R3−1)\O

e2λφ+(|u|2 + |1λ∇u|2)dx ≤ (4.62)

O(1)e2λC∫|v|2dx+O(1)e

2λ(φ+(R3)−R3C0

)∫B(0,

R44,R33

)(|u|2 + |1

λ∇u|2)dx

+O(1)e2λ(φ+(R3)+δ)‖u‖B(0,R0)‖v‖B(0,R0).

46

Since |∇φ| ≤ ε for R3

4≤ |x|, the second term to the right can be absorbed

by the left hand side. So can the u-contribution to the last term, if we usethe standard inequality ”ab ≤ ε

2a2 + 1

2εb2”:

O(1)e2λ(φ+(R3)+δ)‖u‖ ≤ 1

2‖u‖2 +O(1)e4λ(φ+(R3)+δ)‖v‖2.

With a new constant C depending on R3, we obtain:∫B(0,R3−1)\O

(|u|2 + |1λ∇u|2)dx ≤ e2λC

∫|v|2dx. (4.63)

This implies (4.4) and concludes our outline of the proof of Burq’s theorem.

5 Review of non-selfadjoint spectral theory

[Missing piece: Fredholm theory via Grushin problems as in an appendix of[38].]

In this chapter we review some of the standard theory for non-selfadjointoperators, and we will essentially follow the book [31]. Let H be a separablecomplex Hilbert space. If A ∈ L(H,H) is compact, we let s1(A) ≥ s2(A) ≥... 0 be the eigenvalues of (A∗A)1/2. They are called the singular values ofA. We notice that sj(A

∗) = sj(A). In fact, this follows from the intertwiningrelations:

A(A∗A) = (AA∗)A, (A∗A)A∗ = A∗(AA∗).

The singular values appear naturally in the polar decomposition: If A ∈L(H,H), then

‖Au‖2 = (Au|Au) = (A∗Au|u) = ((A∗A)1/2u|(A∗A)1/2u) = ‖(A∗A)1/2u‖2.

The operator

U : R((A∗A)1/2) 3 (A∗A)1/2u 7→ Au ∈ R(A)

is isometric and bijective. It extends to a unitary operator that we alsodenote by U from R((A∗A)1/2 to R(A) and to a partial isometry if we putU = 0 on the orthogonal space (R(A∗A)1/2)⊥ = N ((A∗A)1/2) = N (A), weget the polar decomposition :

A = U(A∗A)1/2. (5.1)

47

This leads to the Schmidt decomposition of A: Let e1, e2, .. be an orthonomralfamily of eigenvectors of (A∗A)1/2 associated to the eigenvalues s1(A), s2(A), ..that are > 0. Then

Au =∑

sj(A)(u|ej)fj, (5.2)

where fj = Uej is also an orthonormal family.Recall the mini-max characterization of the the sj:

sj(A) = infE⊂H;E is a closed

subspace ofHof codimension≤j−1

supu∈E\0

((A∗A)1/2u|u)

‖u‖2. (5.3)

From that we get the following characterization of the singular values whichis due to Allaverdiev:

Theorem 5.1 Let A ∈ L(H,H) be compact. Then

sn+1(A) = minK∈L(H,H)K of rank≤n

‖A−K‖, n = 0, 1, .. .

The minimum is realized by an operator K for which s1(K) = s1(A),..,sn(K) =sn(A), sn+1(K) = 0, s1(A−K) = sn+1(A), s2(A−K) = sn+2(A),.. .

Proof. If K is of rank ≤ n, then N (K) is of codimension ≤ n and

sn+1(A) ≤ sup0 6=u∈N (K)

‖Au‖‖u‖

= sup06=u∈K

‖(A−K)u‖‖u‖

≤ ‖A−K‖.

To get the minimizing operator write the polar decompositionA = U(A∗A)1/2

and take K = U(A∗A)1/2Pn, where Pn is the orthogonal projection onto thespace spanned by e1, .., en. Then

A−K = U(A∗A)1/2(1− Pn),

(A−K)∗(A−K) = (1−Pn)(A∗A)1/2U∗U(A∗A)1/2(1−Pn) = (A∗A)(1−Pn),

and we get the statement about the singular values of A − K. Especiallysn+1(A) = ‖A −K‖. The statement about the singular values of K can beobtained similarly. #

The following corollary is due to Ky Fan:

48

Corollary 5.2 Let A,B ∈ L(H) be compact. Then for n,m ≥ 0:

sm+n−1(A+B) ≤ sm(A) + sn(B) (5.4)

sm+n−1(AB) ≤ sm(A)sn(B). (5.5)

Proof. Let KA, KB be operators of rank ≤ m− 1 and ≤ n− 1 respectively,such that

sm(A) = ‖A−KA‖, sn(B) = ‖B −KB‖.

Then

sm+n−1(A+B) ≤ ‖A+B−(KA+KB)‖ ≤ ‖A−KA‖+‖B−KB‖ = sm(A)+sn(B).

The proof for AB is essentially the same. #

Corollary 5.3 We have |sn(A)− sn(B)| ≤ ‖A−B‖.

Proof. Let K be an operator of rank n− 1. Then

sn(A) ≤ ‖A−K‖ = ‖B −K + A−B‖ ≤ ‖B −K‖+ ‖A−B‖.

Varying K, we get sn(A) ≤ sn(B)+‖A−B‖, and we have the same inequalitywith A and B exchanged. #

We now discuss Weyl inequalities, and start with the following result toH. Weyl:

Theorem 5.4 Let A ∈ L(H,H) be compact and let λ1(A), λ2(A), .. be thenon-vanishing eigenvalues of A arranged in such a way that |λ1| ≥ |λ2| ≥ ...and repeated according to their multiplicity (which by definition is the rankof the spectral projection). The for every n ≥ 1 for which λn(A) is defined,we have

|λ1(A) · .. · λn(A)| ≤ s1(A) · .. · sn(A). (5.6)

(For notational reasons, we assume that there are infinitely many non-vanishingeigenvalues of A.)

Proof. For n = 1, (5.6) just says that |λ1(A)| ≤ ‖A‖. Approaching A bya sequence of finite rank operators, we can assume that A is of finite rankand replace H by the finite dimensional space R(A) + (N (A))⊥, that we

49

denote by H from now on. [[[Haer maaste vi saega naagot om kontinuetenav egenvaerden som funktioner av A.]]] Introduce the space

∧nH = H ∧ ... ∧H (5.7)

generated by n-fold exterior products of vectors in H. ∧nH is a Hilbert spacewith a scalar product that satisfies

(u1 ∧ .. ∧ un|v1 ∧ .. ∧ vn) = det((uj|vk)), uj, vj ∈ H. (5.8)

Further, there is a linear operator ∧nA : ∧nH → ∧nH which is uniquelydetermined by the condition,

(∧nA)(u1 ∧ .. ∧ un) = Au1 ∧ .. ∧ Aun, uj ∈ H. (5.9)

Using a basis of generalized eigenvectors, we see that the eigenvalues of∧nA are the values λj1 · .. · λjn , with jν 6= jµ, for ν 6= µ. The eigenvalue ofgreatest modulus is then λ1 · .. · λn. On the other hand the adjoint of ∧nA is∧nA∗. We also have (∧nA)(∧nB) = ∧n(AB). Then (∧nA)∗(∧nA) = ∧n(A∗A)and this operator has the eigenvalues sj1(A)·..·sjn(A)2 out of which the largestone is

(s1(A) · .. · sn(A))2 = ‖ ∧n A‖2 ≥ |λ1| · .. · |λn|2.The proof is complete. #

In the same spirit we have the inequality of A. Horn:

n∏1

sj(AB) ≤ (n∏1

(sj(A)sj(B)) (5.10)

Proof. As before it suffices to treat the case when H is of finite dimension.The largest eigenvalue of

(∧nAB)∗(∧nAB) = ∧n((AB)∗AB)

is equal to (s1(AB) · .. · sn(AB))2. On the other hand,

((∧nAB)∗(∧nAB)u|u) = ‖(∧nAB)u‖2 = ‖(∧nA) (∧nB)u‖2

≤ ‖ ∧n A‖2‖ ∧n B‖2‖u‖2 ≤ (s1(A) · .. · sn(A))n(s1(B) · .. · sn(B))2‖u‖2,

and taking the supremum over all normalized u, we obtain the requiredinequality. #

We next need a convexity inequality, due to Weyl and Littlewood–Polya.

50

Lemma 5.5 Let Φ(x) be a convex function on R, which tends to 0, whenx→ −∞, with Φ(−∞) := limx→−∞Φ(x) = 0. Let a1 ≥ .. ≥ aN , b1 ≥ .. ≥ bNbe real numbers with

k∑1

aj ≤k∑1

bj, 1 ≤ k ≤ N.

Then,k∑1

Φ(aj) ≤k∑1

Φ(bj), 1 ≤ k ≤ N.

Proof. Approaching Φ by a sequence of smooth functions, we can reducethe proof to the case when Φ ∈ C∞. Then Φ′ ≥ 0,

Φ′(x)→ 0, x→ −∞. (5.11)

Letting y → −∞ in the identity

Φ(x) = Φ(y) +∫ x

yΦ′(t)dt, (5.12)

we get

Φ(x) =∫ x

−∞Φ′(t)dt.

From the convergence of the last integral, we conclude that∫ 0y Φ′(t)dt ≤ C,

implying that |y|Φ′(y), is a bounded function for y ≤ 0, which tends to 0when y → −∞.

Integration by parts in (5.12) gives

Φ(x) = Φ(y) + [(t− x)Φ′(t)]xt=y −∫ x

y(t− x)Φ′′(t)dt

= Φ(y) + (x− y)Φ′(y) +∫ x

y(x− t)Φ′′(t)dt.

Letting y tend to −∞, we get

Φ(x) =∫ x

−∞(x− t)Φ′′(t)dt =

∫(x− t)+Φ′′(t)dt.

Hencek∑1

Φ(aj) =∫

(k∑

j =1

(aj − t)+)Φ′′(t)dt,

51

for every t. Let k(t) ≤ k be the largest k ≤ k with ak≥ t. Then

k∑j=1

(aj − t)+ =k(t)∑j =1

(aj − t) ≤k(t)∑j=1

(bj − t) ≤k∑j=1

(bj − t)+.

Hence∑k

1 Φ(aj) ≤∑k

1 Φ(bk). #

As a consequence we get the following result of Weyl :

Theorem 5.6 Let A : H → H be a compact operator, and f(x) ≥ 0 afunction on [0,∞[ with f(0) = 0 such that f(et) is convex. Let λj and sjbe the eigenvalues and singular values of A, arranged with |λ1| ≥ |λ2| ≥ ...,s1 ≥ s2 ≥ .... Then for every k ≥ 1 :

k∑1

f(|λj|) ≤k∑1

f(sj). (5.13)

Proof. We know that

k∑1

log |λj| ≤k∑1

log sj,

and it suffices to apply the preceding convexity lemma. #

Corollary 5.7 For every p > 0, we have

n∑1

|λj(A)|p ≤n∑1

sj(A)p.

For every r > 0, we have

n∏1

(1 + r|λj(A)|) ≤n∏1

(1 + rsj(A)).

Let A,B be compact operators. With Φ(t) = et, aj = log sj(AB), bj =log(sj(A)sj(B)), we get from Horn’s inequality (5.10) and Lemma 5.5 :

Corollary 5.8∑n

1 sj(AB) ≤ ∑n1 sj(A)sj(B).

Let C∞ ⊂ L(H) be the subspace of compact operators. The followingLemma is due to Ky Fan.

52

Lemma 5.9 Let A ∈ C∞. Then for every 1 ≤ n ∈ N, we have

n∑1

sj(A) = maxn∑j=1

(UAφj|φj),

where the maximum is taken over the set of all unitary operators U and allorthonormal systems φ1, .., φn.

We leave the proof as an exercise, or else see [31].

Corollary 5.10 If A,B ∈ S∞, then

n∑1

sj(A+B) ≤n∑1

sj(A) +n∑1

sj(B).

We are no ready to discuss the Schatten–von Neumann classes.

Definition. For 1 ≤ p ≤ ∞, we put

Cp = A ∈ C∞;∞∑1

sj(A)p <∞.

Theorem 5.11 Cp is a closed two-sided ideal in L(H,H) equipped with thenorm

‖A‖Cp = ‖(sj(A))∞1 ‖`p .If p1 ≤ p2, then Cp1 ⊂ Cp2. The space of finite rank operators is dense in Cpfor every p.

We will only recall the proof of the fact that ‖ · ‖Cp satisfies the triangleinequality. Let A,B ∈ Cp, and put ξj = sj(A + B), ηj = sj(A) + sj(B).According to Corollary 5.10, we have

∑n1 ξj ≤

∑n1 ηj, ∀n and hence,

‖A+B‖C1 ≤ ‖A‖C1 + ‖B‖C1 ,

by letting n tend to ∞.It remains to treat the case p > 1. ξj and ηj are both decreasing sequen-

cies. It suffices to show that

‖ξ‖`p ≤ ‖η‖`p . (5.14)

We have‖ξ‖`p = sup

ζ∈`q,‖ζ‖`q=1

〈ξ, ζ〉,

53

by Holder, where q ∈ [1,+∞[ is the conjugate index, given by p−1 + q−1 = 1and 〈·, ·〉 denotes the real scalar product on `2(1, 2, ..). We also know that

the supremum is attained by a ζ = ζ0 of the form ζ0j = (Const. > 0)ξ

p/qj and

in particular ζ0j is a decreasing sequence. We use partial summation with

Ξj =∑j

1 ξk, Ξ0 = 0:

〈ξ, ζ0〉(n) : =n∑1

ξjζ0j =

n∑j=1

(Ξj − Ξj−1)ζ0j

=n∑j=1

Ξjζ0j −

n−1∑j=1

Ξζ0j+1 = Ξnζ

0n +

n−1∑j=1

Ξj(ζ0j − ζ0

j+1)

= (n∑1

ξk)ζ0n +

n−1∑j=1

j∑k=1

ξk) (ζ0j − ζ0

j+1)︸ ︷︷ ︸≥0

The last expression is≤ the same expression with ξ replaced by η and runningthe same calculation backwards the latter expression is equal to 〈η, ζ0〉(n)

Hence 〈ξ, ζ0〉(n) ≤ 〈η, ζ0〉 ≤ ‖η‖`p . Letting n tend to infinity, we get (5.14),and this completes the proof of the triangle-inequality for the Cp-norms. #

We notice that ‖A‖ = s1(A) ≤ ‖A‖Cp . The space C1 is the space ofnuclear or trace-class operators, and C2 is the space of Hilbert-Schmidt op-erators. We have the following Holder type result :

Theorem 5.12 Let p, q ∈ [1,∞] be conjugate indices; p−1 + q−1 = 1. IfA ∈ Cp, B ∈ Cq, then AB ∈ C1 and ‖AB‖C1 ≤ ‖A‖Cp‖B‖Cq .

Proof. We know that

n∑1

sj(AB) ≤n∑1

sj(A)sj(B)

and letting n tend to ∞, we get from the usual Holder inequuality:

∞∑1

sj(AB) ≤∞∑1

sj(A)sj(B) ≤ ‖s·(A)‖`p‖s·(B)‖`q = ‖A‖Cq‖B‖Cq .

#

We next discuss the trace of a nuclear operator. If A ∈ L(H,H) is of finiterank, we choose a finite dimensional subspace H′ ⊂ H such that R(A) ⊂ H′,

54

(H′)⊥ ⊂ N (A) We then define the trace of A, trA as the trace trA|H′ of therestriction of A to H′. We check that this does not depend on the choice ofH′ and that trA is the sum of the finitely many non-vanishing eigenvaluesof A (each counted with its algebraic multiplicity). We see that

A 7→ trA (5.15)

is a linear functional on the space of finite rank operators. Moreover,

|trA| ≤∑|λj(A)| ≤ ‖A‖C1 . (5.16)

We can then extend (5.15) to a continuous linear functional on C1 and westill have

|trA| ≤ ‖A‖C1 . (5.17)

In the case of finite rank operators, we also have

trAB = trBA. (5.18)

Let now A ∈ Cp, B ∈ Cq, where p, q ∈ [1,∞] are conjugate indices and chooseAν , Bν , ν = 1, 2, .. of finite rank, so that ‖A−Aν‖Cp → 0, ‖B −Bν‖Cq → 0.Then

‖AB − AνBν‖C1 = ‖(A− Aν)B + Aν(B −Bν)‖C1

≤ ‖A− Aν‖Cp‖B‖Cq + ‖Aν‖Cp‖B −Bν‖Cq → 0, ν →∞.

Using this also for BA and the cyclictiy of the trace (5.18) for finite rankoperators, we obtain it also in the case A ∈ Cp, B ∈ Cq, where p, q ∈ [1,∞]are conjugate indices.

Remark. There is a simple way of extending most of the theory to the caseof operators A : H1 → H2, where H1,H2 are two different Hilbert spaces.Consider namely the corresponding operator(

0 0A 0

): H1 ⊕H2 → H1 ⊕H2, (5.19)

and say that A belongs to Cp if the operator in (5.19) does. We leave theseextensions as an exercise for the reader, and notice simply that the cyclicity ofthe trace still holds in this setting, namely if A : H1 → H2, and B : H2 → H1

belong to Cp and Cq respectively, where p and q are conjugate indices.

55

We next discuss determinants of trace-class perturbations of the iden-tity operator. Let first A ∈ L(H,H) be of finite rank and chose a finitedimensional Hilbert space as above. Then we define

det(1− A) = det((1− A)|H′) =∏j

(1− λj(A)), (5.20)

where λj(A) denote the non-vanishing eigenvalues, repeated according totheir multiplicity. We remark that

| det(1− A)| ≤∏j

(1 + |λj(A)|) ≤∏j

(1 + sj(A)) ≤ e∑

jsj(A),

where the the second inequlity follows from Corollary 5.7 . We want to extendthe definition to the case when A ∈ C1. Let first I be a compact interval andlet I 3 t 7→ At be a C1family of finite rank rank operators, withR(At) ⊂ H′,N (At) ⊂ (H′)⊥ for some finite dimensional subspaceH′ which is independentof t. We first assume that 1−At is invertible for all t ∈ I, or in other wordsthat 1− λj(At) 6= 0 for all t and j. Then det(1− At) 6= 0 and by a classicalformula,

∂tlog det(1− At) = −tr ((1− At)−1 ∂

∂t(At)) = −tr ((

∂tAt)(1− At)−1).

Hence,

∣∣∣ ∂∂t det(1− At)det(1− At)

∣∣∣ =∣∣∣ ∂∂t

log det(1− At)∣∣∣ ≤ ‖(1− At)−1‖‖ ∂

∂tAt‖C1 .

In particular, if I = [0, 1], At = tA1 + (1− t)A0, we get

| log det(1−A1)−log det(1−A0)| ≤ sup0≤t≤1

‖(1−(tA1+(1−t)A0))−1‖‖A1−A0‖C1 .

(5.21)Now let A ∈ C1. If 1−A1 is not invertible, we put det(1−A) = 0. Assume

then that 1 − A is invertible. Let Aν be a sequence of finite rank operatorswhich converges to A in C1. For ν large enough, we have ‖(1−Aν)−1‖ ≤ C0

for some fixed constant C0 and more generally ‖(1− (tAν + (1− t)A0))−1‖ ≤C0. Then

| log det(1− Aν)− log det(1− Aµ)| ≤ C0‖Aν − Aµ‖C1

56

and consequently limν→∞(log det(1−Aν)− log det(1−Aµ)) exists. We thenput

det(1−A) = det(1−Aµ) exp limν→∞

(log det(1−Aν)− log det(1−Aµ)). (5.22)

Notice that

| det(1− A)| ≤∏

(1 + sj(A)) ≤ e‖A‖C1 . (5.23)

Using approximation by finite rank operators, we also see that

det((1− A)(1−B)) = det(1− A) det(1−B). (5.24)

By the same argument, we can extend (5.21) to general trace clacc operatorsfor which 1− (tA1 + (1− t)A0) is invertible.

We now add a complex variable z ∈ C and consider the function det(1−zA). If A is of finite rank, then this is an entire function of z. If A ∈ C1,let Aν → A be a sequence of finite rank operators. Then | det(1 − zAν)| ≤e|z|‖Aν‖C1 . If zλj(A) 6= 1, ∀j, then det(1 − zAν) → det(1 − zA) with locallyuniform convergence in C \ ∪j1/λj, and it follows that det(1 − zA) is aholomorphic function on this set, which verifies

| det(1− zA)| ≤ e‖A‖C1|z|. (5.25)

It follows that det(1 − zAν) converges to an entire function f(z) locallyuniformly on C. If z = 1/λj(A) where λj(A) is of multiplicity m, then exaclym eigenvlues of Aν will converge to λj(A) while the others will stay awayfrom a neighborhood of this point (when ν is large enough). Considering theargument variation (Rouche), we conclude that f(z) vanishes to the order mat 1/λj(A) and in particular we have f(z) = det(1− zA) also at that point.In conclusion, we have

Proposition 5.13 Let A ∈ C1. Then DA(z) := det(1 − zA) is an en-tire function whose zeros counted with multiplicity coincide with the values1/λ1(A), 1/λ2(A), .. counted with the multiplicities of λ1(1), λ2(A), ...

Observe thatDA(0) = 1 (5.26)

Also observe that DA(z) is of subexponental growth in the sense that forevery ε > 0 there exists a constant Cε > 0 such that

|DA(z)| ≤ Cεeε|z|. (5.27)

57

In fact, by a limiting argument, we have

|DA(z)| ≤∞∏1

(1 + |z|sj(A)) ≤N∏1

(1 + |z|sj(A))e∑∞

N+1sj(A))|z|.

Here the prefactor is of polynomial growth for every fixed N and for a givenε > 0, we can always choose N > 0 so large that the exponent is ≤ ε|z|.

The next observation is that W (z) =∏∞

1 (1 − zλj(A)) is also an entirefunction of subexponential growth. This follows by the same argument, if werecall that

∑∞1 |λj(A)| is convergent. We now use a special case of a theorem

of Hadamard (see [2]): Since DA(z) and W (z) have the same zeros (countedwith multiplicity), we have

DA(z) = W (z)eg(z),

where g(z) is an entire function. It is also clear that we can choose g withg(0) = 0. The function

Re g(z) := log |DA(z)| −∞∑1

log |1− λz| (5.28)

is harmonic. Let R ≥ 2. For |z| ≤ R/2, we have

Re g(z) =∫|w|=R

PR(z, w)Re g(w)`(dw), (5.29)

where PR is the Poisson kernel for the disc of radius R and ` denotes thelength element on the boundary of this disc. By a scaling argument, it iseasy to see that

1

CR≤ PR(z, w) ≤ C

R, |z| ≤ R

2, |w| = R, (5.30)

where C > 0 is independent of R. Using the subexponential growth of DA(z),we get ∫

|w|=RPR(z, w) log |DA(w)|`(dw) ≤ C

RεRR ≤ CεR, (5.31)

for R ≥ Rε large enough. On the other hand,

|∫|w|=R

PR(z, w)∞∑1

log |1− λjw|`(dw)| ≤∞∑1

1

R

∫|w|=R

| log |1− λjw||`(dw)

(5.32)

58

If |λj|R ≤ 12, we write | log |1 − λjw|| ≤ C|λj|R and the sum over the

corresponding js in (5.32) can be bounded by

∑|λj |≤ 1

2R

C

R2π|λj|RR = 2πCR

∑|λj |≤ 1

2R

|λj| = o(R), R→∞. (5.33)

If 12≤ |λj|R ≤ T , where T 1 is independent of R, then with a =

|λj|R ∈ [12, T ]:

1

R

∫|w|=R

| log |1− λjw||`(dw) =∫|ζ|=1| log |1− aζ||`(dζ).

With ζ = eiθ, we get

|1− aζ|2 = (1− a cos θ)2 + a2(sin θ)2

= 1 + a2 − 2a cos θ = a(1

a+ a− 2 cos θ) ≥ 2a(1− cos θ).

Consequently,1

R

∫|w|=R

| log |1− λjw||`(dw) ≤ CT .

Let us estimate the number of λj in this case:∑12≤|λj |R≤T

1 ≤ 2R∑

12R≤|λj |TR

|λj| = oT (R).

Hence

∑12≤|λj |R≤T

1

R

∫|w|=R

| log |1− λjw||`(dw) = oT (R), R→∞. (5.34)

It remains to consider the case |λj|R ≥ T . Here log |1−λjR| ∼ log(|λj|R).Hence, with constants C and Cδ that are independent of T :

∑|λj |R≥T

1

R

∫|w|=R

| log |1− λjw||`(dw) ≤ C∑

|λj |R≥Tlog(|λj|R)

≤ Cδ∑|λj |≥TR

|λj|δRδ = CδRδ∑|λj |≥TR

|λj|δ.

59

Put δ = 1/p, 1 < p < ∞, and let q be the conjugate index. Then, if Ndenotes the number of λj with |λj| ≥ T/R, we get from Holder’s inequlity:

∑|λj |≥TR

|λj|1p ≤ N

1q (

∑|λj |≥TR

|λj|)1p . (5.35)

Here NT/R ≤ ∑ |λj| ≤ C, so N ≤ CR/T and the expression (5.35) isbounded by CR1/q/T 1/q. Hence

∑|λj |R≥T

1

R

∫|w|=R

| log |1− λjw||`(dw) ≤ CR1p

+ 1q /T

1q =

CR

T1q

. (5.36)

Combining the three cases, we find

|∫|w|=R

PR(z, w)(∞∑1

log |1− λjw|)`(dw)| ≤ oT (R) +CR

T1q

= o(R), R→∞

(5.37)Combining this with (5.28), (5.29) and (5.31), we get

Re g(z) ≤ o(R), on |z| ≤ R

2.

Now we can apply Harnack’s inequality to the function Re g − o(R), whichis ≤ 0 on the disc |z| ≤ R/2 and ≥ −o(R) at 0 and conclude that

Re g ≥ −o(R) on the disc |z| ≤ R

4.

Since g is harmonic, it follows from the last two estimates that Re g = 0.Hence g is constant and since we have chosen g with g(0) = 0, we get g(z) = 0∀z ∈ C. We have then showed

Theorem 5.14 Let A : H → H be a trace class operator with the non-vanishing eigenvalues λ1(A), λ2(A), .., 0 ≤ N ≤ ∞ repeated according tomultiplicity. Then DA(z) = det(1− zA) satisfies

DA(z) =N∏j=1

(1− λjz), z ∈ C. (5.38)

From this we get the important Lidskii’s theorem as a corollary :

60

Corollary 5.15 If A ∈ C1 we have

trA =N∑1

λj(A), (5.39)

where λj(A) are the non-vanishing eigenvalues as in Theorem 5.14.

Proof. We know the result when A is of finite rank. In this case, we alsoknow that

D′ADA

=∂

∂zlog det(1− zA) = −tr ((1− zA)−1A),

away from the zeros of z 7→ det(1− zA). In particular,

D′A(0)

DA(0)= −trA, (5.40)

when A is of finite rank.When A ∈ C1, let Aν be a sequence of finite rank operators converging

to A in the C1 norm. Then DAν (z) → DA(z), D′Aν (z) → D′A(z), whenν → ∞, uniformly for z in a neighborhood of 0. Since DA(z) 6= 0 in such aneighborhood, we have also

D′Aν (0)

DAν (0)→ D′A(0)

DA(0).

By (5.40) we know that the right hand side of the last relation is equal to−trAν and we also know that this quantity converges to−trA. Consequently(5.40) remains valid for general trace class operators. In view of Theorem5.14, we know on the other hand that

D′A(0)

DA(0)= −

N∑1

λj(A),

and the Corollary follows. #

The last proof also shows that

D′A(z)

DA(z)=

∂zlogDA(z) = −tr (1− zA)−1A = −

N∑1

λj(A)

1− zλj(A), (5.41)

for all z with 1− zλj(A) 6= 0, ∀j.We now turn to some other questions, that will be of use.

61

Proposition 5.16 Let H0,H1,H2 be complex separable Hilbert spaces. LetΩ ⊂ Cn be an open set and let Ω 3 z 7→ K(z) ∈ C1(H1) be a holomorphicfunction. Then det(1−K(z)) is a holomorphic function on Ω and if z0 ∈ Ωis a zero of order m ≥ 1 of this function, and A(z) ∈ L(H1,H2), B(z) ∈L(H0,H1) depend holomorphically on z ∈ Ω, then

rank (∫γA(z)(1−K(z))−1B(z)dz) ≤ m, (5.42)

if γ is the positively oriented boundary of a sufficiently small circle centeredat z0.

Proof. We know that 1 is an eigenvalue of K(z0) of a certain multiplicityN0, defined to be the rank of the spectral projection

Π(z0) =1

2πi

∫α(λ−K(z0))−1dλ,

where α is the positively oriented boundary of a small disc centered at λ = 1.For z close to z0, we put

Π(z) =1

2πi

∫α(λ−K(z))−1dλ,

and we notice that this is the sum of the spectral projections correspondingto the N0 eigenvalues λj0(z), λj0+1(z), .., λj0+N0−1(z) (repeated according tomultiplicity) that are close to λj0(z0). m is then also the order of vanishingof (1−λj0(z)) · .. · (1−λj0+N0−1(z)) (if we note that det(1−K(z)) = det(1−K(z)Π(z)) det(1 − K(z)(1 − Π(z))) ). The range of Π(z) is of constantdimension N0 and we can find a basis e1(z), .., eN0(z) of this space whichdepends holomorphically on z (possibly after restricting z to a new evensmaller neighborhood of z0.

DefineR+ : H1 → CN0 byR+(z)u(j) = aj(u, z), where Π(z)u =∑N0

1 aj(u, z)ej.Define R−(z) : CN0 → H1 by R−(z)u− =

∑N01 u−(j)ej(z). Then(

1−K(z) R−(z)R+(z) 0

): H1 ×CN0 → H1 ×CN0

is bijective with inverse (E(z) E+(z)E−(z) E−+(z)

)

62

where E−+ is the matrix of the restriction of K(z)−1 toR(Π(z)) with respectto the basis e1(z), .., eN0(z). Hence detE−+(z) =

∏N0−1ν=0 (1 − λj0(z))...(1 −

λj0+N0−1(z)) has the same order of vanishing at z = z0 as det(1−K(z)). Itthen suffices to apply Lemma [[... som skall ingaa i en foersta del av dettakapitel daer vi behandlar Fredholm teori med hjaelp av Grushin problem]].

We end this chapter by recalling Jensen’s formula and the standard ap-plication to getting bounds on the number of zeros of holomorphic functions.Let f(z) be a holomorphic function on the open disc D(0, R) with a contin-uous extension to the corresponding closed disc. Assume that f(0) 6= 0 andf(z) 6= 0 for |z| = R.

Assume first that f(z) has no zeros at all. Then log |f(z)| = Re log f(z) isa harmonic function in the open disc which is continuous up to the boundary,and the mean value property of harmonic functions tells us that

log |f(0)| = 1

∫ 2π

0log |f(Reiθ)|dθ.

We now allow f to vanish and let z1, .., zN be the zeros repeated accordingto their multiplicity. Since f is not allowed to vanish at 0 or at the boundaryof the disc of radius R, we have 0 < |zj| < R. Then

F (z) := f(z)N∏j=1

R2 − zjzR(z − zj)

is holomorphic in the open disc, continuous up to the boundary and has nozeros in the closed disc of radius R. Moreover |F (z)| = |f(z)| when |z| = R,so according to the preceding paragraph, we have

log |F (0)| = 1

∫ 2π

0log |f(Reiθ)|dθ.

Expanding the left hand side, we get Jensen’s formula :

log |f(0)|+N∑1

logR

|zj|=

1

∫ 2π

0log |f(Reiθ)|dθ. (5.43)

A standard application of this formula is to notice that if N(R/2) is thenumber of zeros zj of f with |zj| ≤ R/2, then we get

N(R

2) log 2 ≤ 1

∫ 2π

0log |f(Reiθ)|dθ − log |f(0)|. (5.44)

63

6 Global upper bounds on the number of res-

onances

In this chapter we shall show that in standard situations and for odd dimen-sions ≥ 3, the number of resonances in a disc D(0, r) (of center 0 and radiusr) is O(rn), r →∞. It will be convenient to formulate the result in the blackbox setting of section 2.3 of chapter 2, and after that we will mention someof the history of this result. The methods come to a large extent from thetheory of non-selfadjoint operators, and can therfore be viewed as classical.

Let P : H → H be as in chapter 2, section 2.3, so that H is given by theorthogonal decomposition (2.30). Introduce the torus M = (R/2RZ)n withR > R0 and identify B(0, R0) with its image under the natural projectionRn →M . Put H] → H]. The domain D] = D(P ]) is by definition

D] = χu1 + (1− χ)u2; u1 ∈ D, u2 ∈ H2(M). (6.1)

Here χ ∈ C∞0 has its support in a small neighborhood of B(0, R0) and isequal to one near that set. We view χ both as an element of C∞0 (Rn) andas an element of C∞0 (M). It is easy to see that the definition (6.1) does notdepend on the choice of χ. For D(P ]) 3 u = χu1 + (1− χ)u2, we put

P ]u = P (χu1) + (−∆)((1− χ)u2), (6.2)

and again, this does not depend on the representation of u.

Lemma 6.1 P ] is self-adjoint and has purely discrete spectrum.

Proof. It is easy to see that P ] is symmetric and we leave the details to thereader. Let u ∈ D((P ])∗), (P ])∗u = v, v ∈ H]. We claim that u ∈ D(P ]):First we use that

(u| −∆φ) = (v|φ), ∀φ ∈ C∞0 (M \B(0, R0)),

so that −∆u = v in M \B(0, R0) and hence

u|M\B(0,R0)∈ H2

loc(M \B(0, R0)).

After subtracting an exterior part from u, we may assume that χu = u, withχ as above. Now we can view u, v as elements of H, and from the relation(P ])∗u = v, it follows that

(u|Pφ) = (v|φ),

64

first for all φ ∈ D with support in a neighborhodd of suppχ, then for allφ ∈ D. Since P is self-adjoint, we get u ∈ D and the claim follows.

Let vj ∈ H], j = 1, 2, ... be a bounded sequence and consider uj = (P ] +i)−1vj ∈ D]. With χ as before, we first see that (1−χ)(P ]+i)−1vj = (1−χ)ujis bounded in H2(M) and hence has a convergent subsequence in H]. As forχuj, we compute (P + i)χuj = [−∆, χ]uj + χvj =: wj, so wj is bounded inH. Now write χuj = (P + i)−1wj = (P + i)−1χwj, where χ ≺ χ ∈ C∞0 (Rn).Since (P + i)−1χ is compact, χuj has a convergent subsequence in H and inH]. It follows that (P ] + i)−1 : H] → H] is compact. Hence P ] has purelydiscrete spectrum. #

Let N(P ], I) be the number of eigenvalues of P ] in the interval I. Weassume

N(P ], [−λ, λ]) = O(1)Φ(λ), λ ≥ 1, (6.3)

where Φ : [1,∞[→ [1,∞[ is continuous strictly increasing, with Φ(1) = 1.We also assume:

Φ(t) ≥ tn2 , (6.4)

∀C ≥ 1, ∃C(C) ≥ 1, such that (6.5)

Φ(Ct) ≤ C(C)Φ(t), φ(Cλ) ≤ C(C)φ(λ), λ, t ≥ 1,

where φ(λ) = Φ−1(λ).Let λ1, λ2, λ3, ... be the eigenvalues of P ] repeated according to multiplic-

ity and arranged so that |λ1| ≤ |λ2| ≤ ... . The eigenvalues of the compactnormal operator (i−P ])−1 are then (i−λj)−1, and the corresponding singularvalues are given by

sj((i− P ])−1) =1

|i− λj|=

1

〈λj〉. (6.6)

It is easy to see that the property (6.3) is equivalent to

sj((i− P ])−1) ≤ O(1)

φ(j). (6.7)

For the same operator P , let M be a second torus and P the corre-sponding operator analogous to P ]. We identify M and M by means of adiffeomorphism which acts like the identity near B(0, R0). In this way, wecan view both P and P ] as operators on M (i.e. as operators H] → H].

65

(Notice however, that P will not be self-adjoint with respect to the scalarproduct of H] but for another one which gives an equivalent norm.) We havethe resolvent identity:

(i− P )−1 = (i− P ])−1 + (i− P )−1(P − P ])(i− P ])−1. (6.8)

Now recall some facts from Chapter 5 about singular values. If K : H →H is a compact operator, then the jth singular value (i.e. the jth eigenvalueof√K∗K counted with multiplicities in decreasing order) obeys Ky Fan’s

identitysj(K) = inf

R∈L(H,H),rank (R)≤j−1

‖K −R‖. (6.9)

From this we deduce the wellknown inequalities:

sj+k−1(K + L) ≤ sj(K) + sk(L), (6.10)

sj+k−1(KL) ≤ sj(K)sk(L), (6.11)

for j, k ≥ 1. Applying this to (6.8), we see that P also satisfies (6.7). Thisshows that the assumption (6.3) does not depend on the choice of M .

Later on we will also need the following information which follows from(6.9): If K is a compact operator, then for every j ≥ 1, there exists a boundedoperator R of rank ≤ j − 1, such that

sν(K) = sν(R), 1 ≤ ν ≤ j − 1, (6.12)

sj(K) = ‖K −R‖ = s1(K −R),

sν(K) = sν−j+1(K −R), ν ≥ j.

The same argument shows that

sj((P − z)−1χ), sj(χ(P − z)−1) ≤ C

φ(j), (6.13)

for χ ∈ C∞0 (Rn), 1B(0,R0) ≺ χ and for z in any fixed comapct subset of theresolvent set. This estimate is the starting point of the proof of the followingresult:

Theorem 6.2 Let n ≥ 3 be odd. Under the assumptions above, the numberN(r) of resonances in the open disc D(0, r) satisfies

N(r) ≤ CΦ(r2), r ≥ 1. (6.14)

66

Before starting the proof we give some examples and make some historicalremarks. Let g be a smooth Riemannian metric on Rn which coincides withthe standard Euclidean metric on Rn outside a bounded set, let O ⊂⊂ Rn

be open with smooth boundary, let V ∈ L∞comp(Rn \ O), and let P be theDirichlet or Neumann realization in L2(Rn \ O) of the Schrodinger operator−∆g +V (x), where ∆g denotes the Laplace-Beltrami operator for the metricg. Then we can take Φ(r) = rn/2 and we get N(r) ≤ Crn. In the case,V = 0 and ∆g = ∆ everywhere, this was proved by Melrose [54]. Zworski[101] obtained it for −∆ + V without any obstacle, and Vodev [92] obtainedit for −∆g on L2(Rn). Theorem 6.2 as stated above is essentially due toSjostrand–Zworski [80], but in the proof below we shall mainly follow Vodev[93], who extended the result to certain non-selfadjoint operators and whoalso obtained sharp results in the even-dimensional case [95, 96]. In thecontext of hyperbolic manifolds, analogous results have been obtained byGuillope–Zworski [34], Froese–Hislop [25].

Proof. Let1B(0,R0) ≺ χ0 ≺ χ1 ≺ χ2, χj ∈ C∞0 (Rn)

as in Chapter 2. Let µ ∈ C, arg µ = π4, |µ| = r 1. Then we know that the

resonances λ are contained in the set of λ, such that (1+K(λ, µ)χ) : H → His not invertible, where χ2 ≺ χ ∈ C∞0 (Rn). Here K is the same as in Chapter2, (2.39), (2.40), (2.41):

K(λ, µ) = [∆, χ0]R0(λ)(1−χ1)− [∆, χ2]R(µ)χ1 +(µ2−λ2)χ2R(µ)χ1. (6.15)

We will restrict the attention to λ ∈ C; |λ − µ| < 3r. Since K(λ, µ)χ isnot necessarily of trace class, we shall first decompose it into one term withsmall norm, and one term which is of trace class. This decomposition onlyconcerns the last term in (6.15).

WriteR(µ)χ1 = (P − i|µ|2)−1(P − i)(P − i)−1χ,

and conclude that

sj(χ2R(µ)χ1) ≤ min(1

r2,O(1)

φ(j)), j = 1, 2, ... (6.16)

For j(r) = O(1)Φ(r2), we can decompose :

χ2R(µ)χ1 = Aµ +Bµ, (6.17)

67

where

‖A‖ ≤ 1

100r2, (6.18)

sj(B) =sj(χ2R(µ)χ1), 1 ≤ j ≤ j(r),0, j ≥ j(r) + 1.

(6.19)

We have already seen in Chapter 2, that if |µ| = r is large enough, then

‖[∆, χ2]R(µ)χ1‖ ≤1

4. (6.20)

Write

1 +K(λ, µ) = (6.21)

1 +D + (µ2 − λ2)B + [∆, χ0]R0(λ)(1− χ1)χ,

where D = (µ2 − λ2)A− [∆, χ2]R(µ)χ1 and ‖D‖ ≤ 1/2,

1 +K(λ, µ)χ = (1 +D)(1 + K), (6.22)

K = (µ2 − λ2)(1 +D)−1B + (1 +D)−1[∆, χ0]R0(λ)(1− χ1)χ.

We shall take the determinant of 1 + K, and in order to estimate thisdeterminant, we shall first estimate the singular values of the last term inthe expression for K.

For Imλ ≥ 0, α > 0, we have

sj([∆, χ0]R0(λ)(1− χ1)χ) ≤ Cα〈λ〉αj−α/n, (6.23)

and for Imλ < 0 :

sj([∆, χ0]R0(λ)(1− χ1)χ) ≤eC〈λ〉, ∀jj−

nn−1 + Cα〈λ〉αj

−αn , j ≥ C〈λ〉n−1.

(6.24)

We postpone the proof of these two estimates and finish the proof of thetheorem. Let

h(λ, µ) = det(1 + K)

For λ = µ, we get

h(µ, µ) = det(1 + (1 +D)−1[∆, χ0]R0(µ)(1− χ1)χ).

68

In this case we can use the off diagonal exponential decay of the distributionkernel of R0(µ), to see that the distribution kernel k(x, y) of [∆, χ0]R0(µ)(1−χ1)χ satsifies: ∑

|α|≤2n+1

‖∂αx,yk(x, y)‖L1(R2n) ≤ O(1)e−r/C0 , (6.25)

for some C0 > 0. On the other hand it is well known that the trace classnorm of [∆, χ0]R0(µ)(1− χ1)χ is bounded by a constant times the left handside of (6.25), so

‖(1 +D)−1[∆, χ0]R0(µ)(1− χ1)χ‖tr ≤ O(1)e−r/C0 . (6.26)

Use also the Weyl inequality

| det(1 + K)| ≤∞∏1

(1 + sj(K)) ≤ e‖K‖tr , (6.27)

to conclude that1

C≤ |h(µ, µ)| ≤ C, (6.28)

for r = |µ| sufficiently large.We shall next find upper bounds for |h(λ, µ)|. It follows from (6.23),

(6.24), that (6.24) holds also in the upper half plane. We may assume that‖(1 +D)−1‖ ≤ 2 and get

|h(λ, µ)| ≤∞∏j=1

(1 + 2sj((µ2 − λ2)B + [∆, χ0]R0(λ)(1− χ1)χ)).

Use that for j = 2k − 1: sj(A1 + A2) ≤ sk(A1) + sk(A2) by Corollary 5.2:

|h(λ, µ)| ≤∞∏k=1

(1 + 2s2k−1((µ2 − λ2)B) + [∆, χ0]R0(λ)(1− χ1)χ))2

≤∞∏k=1

(1 + 2sk((µ2 − λ2)B))2

∞∏k=1

(1 + 2sk([∆, χ0]R0(λ)(1− χ1)χ))2

= F1F2,

where the last equation defines the two factors Fj in the natural way. Clearly,F1 ≤ exp(O(1)r2∑∞

1 sk(B)). Hence by (6.16), (6.19):

∞∑1

sk(B) ≤j(r)∑

1

min(1

r2,O(1)

φ(k)) ≤

O(1)Φ(r2)∑1

1

r2≤ O(1)

Φ(r2)

r2.

69

HenceF1 ≤ exp(C1Φ(r2)). (6.29)

Using (6.24):

F2 ≤ (∏

1≤k≤C〈λ〉n−1

e2C〈λ〉)×

exp(4∑

k>C〈λ〉n−1

j−nn−1 + Cα1

∑C〈λ〉n−1≤j≤〈λ〉n

〈λ〉α1j−α1n + Cα2

∑j>〈λ〉n

〈λ〉α2j−α2n ),

where we are free to choose α1, α2 > 0. Take α1 < n < α2. Then evaluatingthe three sums in the last exponent, we get

F2 ≤ expC〈λ〉n. (6.30)

Consequently, for |λ− µ| ≤ 3r, arg µ = π4, |µ| = r:

|h(λ, µ)| ≤ eC(rn+Φ(r2)) ≤ eCΦ(r2). (6.31)

by Jensen’s formula for the disc D(µ, 3r), we see that the number of zerosof λ→ h(λ, µ) in D(µ, 2r) is ≤ O(1)Φ(r2). Now write (cf (2.46))

R(λ) = Q(λ, µ)(1 +K(λ, µ)χ)−1(1−K(λ, µ)(1− χ))

= Q(λ, µ)(1 + K(λ, µ))−1(1 +D)−1(1−K(λ, µ)(1− χ))

= A(λ, µ)(1 + K(λ, µ))−1B(λ, µ),

where A,B are holomorphic in λ. After multiplying to the left and to theright by functions in C∞0 , we can apply the propositions 5.16, 2.3, and con-clude that the number of resonances in D(µ, 2r) is bounded by the numberof zeros of h(·, µ) in that set and hence by O(1)Φ(r2). The theorem follows.#

It remains to prove (6.23), (6.24). We start with the first of these esti-mates. For |λ| ≤ 1, we use that [∆, χ0]R0(λ)(1 − χ1)χ has a distributionkernel in C∞0 . If Ω is a large ball, we can consider our operator on Ω andwrite

[∆, χ0]R0(λ)(1− χ1)χ = (1−∆Ω)−N(1−∆Ω)N [∆, χ0]R0(λ)(1− χ1)χ,

where ∆Ω denotes the Dirichlet Laplacian in Ω. Here

(1−∆Ω)N [∆, χ0]R0(λ)(1− χ1)χ

70

is uniformly bounded and the jth singular value of (1 − ∆Ω)−N is ∼ j−2nN

by the known (Weyl) asymptotics for the eigenvalues of −∆Ω.For |λ| ≥ 1, we follow the same idea, but we have to be more precise. Let

us first show that if χ1, χ2 ∈ C∞0 , then

χ1R0(λ)χ2 = Os(1

|λ|) : Hs → Hs, s ∈ R, Imλ ≥ 0. (6.32)

Indeed, on the Fourier transform side, the kernel of χ1R0(λ)χ2 becomes

K(ξ, η) =∫χ1(ξ − ζ)

1

ζ2 − λ2χ2(ζ − η)

(2π)n, ξ, η ∈ Rn.

If |ζ2 − λ2| ≥ |λ|C

, ζ ∈ Rn, then |K(ξ, η)| ≤ CN〈ξ−η〉−N|λ| and χ1R(λ)χ2 =

Os( 1|λ|) : Hs → Hs, ∀s.If |ζ2−λ2| is smaller than |λ|

C, then |ζ| ∼ |λ|, and hence |∇ζ(ζ

2−λ2)| ∼ |λ|.We then replace Rn

ζ by the deformed contour:

Γ : Rn 3 ζ 7→ ζ ± iεχ(|ζ|2 − |λ|2

|λ|2)ζ

|ζ|= ζ + iO(ε) =: ζ ,

for a suitable χ, and along Γ, we have

ζ2 = ζ2 ± 2iεχ(...)ζ2

|ζ|− ε2χ2(...),

and we get |ζ2 − λ2| ≥ |λ|/C. Using Paley–Wiener, we obtain (6.32).Next we show that if χ1, χ2 ∈ C∞0 (Rn), suppχ1 ∩ suppχ2 = ∅, then

χ1R0(λ)χ2 = Os,m(〈λ〉m−1) : Hs → Hs+m, for m ≥ 0. (6.33)

In fact, if m = 2k, we notice that

(−∆ + 1)R0(λ) = (1 + λ2)R0(λ) + 1,

and hence outside suppχ2:

(−∆ + 1)R0(λ)χ2 = (1 + λ2)R0(λ)χ2.

By iteration, we get toutside suppχ2:

(−∆ + 1)kR0(λ)χ2 = (1 + λ2)kR0(λ)χ2,

71

soχ1(−∆ + 1)kR0(λ)χ2 = (1 + λ2)kχ1R0(λ)χ2.

This gives (6.33) for m = 2k, and by interpolation, we get it for generalm ≥ 0.

Proof of (6.23). Let Ω be as before, so that the jth eigenvalue of (−∆Ω +1)1/2 is ∼ j1/n. Then as before,

[∆, χ0]R0(λ)(1− χ1)χ = (1−∆Ω)−α2 (1−∆Ω)

α2 [∆, χ0]R0(λ)(1− χ1)χ︸ ︷︷ ︸

=O(〈λ〉α) in operator norm

.

It suffices to use sj((1−∆Ω)−α2 ) ∼ j−

αn #

Proof of (6.24). Thanks to (2.5), (6.23) it suffices to show that

sj(q1T (λ)q0) ≤eC〈λ〉, ∀j,j−

nn−1 , j ≥ C〈λ〉n−1,

(6.34)

where q1, q0 are differential operators with C∞0 coefficients of order ≤ 1 and0 respectively and where T (λ) is given by (2.6). Clearly, ‖T (λ)‖ ≤ eC〈λ〉 sothe first estimate in (6.34) is obvious.

We observe that T (λ) = λn−2E(λ)∗E(λ), where

E(λ)u(ω) =∫e−iλω·yχ(y)u(y)dy. (6.35)

WriteT (λ) = λn−2E(λ)∗(1−∆Sn−1)−k(1−∆Sn−1)kE(λ). (6.36)

Here

sj((1−∆Sn−1)−k) ≤ (j

C

)− 2kn−1 , (6.37)

‖(1−∆Sn−1)kE(λ)q0‖ ≤ e2|λ|(Ck)2k, (6.38)

where the last estimate follows from the Cauchy inequalities. We deducethat

sj(q1T (λ)q0) ≤ Ce3|λ|(Cj

) 2kn−1 (Ck)2k (6.39)

≤ Ce3|λ|( Ck

j1/(n−1)

)2k.

72

Choose k so thatCk

j1/(n−1)=

1

e,

i.e. so that

k =j1/(n−1)

Ce.

Actually we have to modify this choice a little bit since k should be aninteger.) Then (6.39) gives

sj(T (λ)) ≤ C3|λ|e−2 j

1n−1

Ce . (6.40)

For j1

n−1 |λ|, we get

sj(T (λ)) ≤ Ce− j

1n−1

Ce ,

which is even better than (6.34). This completes the proof of (6.24). #

7 Resonances for long range perturbations of

−h2∆, analytic distorsions

7.1 Operators of black box type

Our operators will depend on a Planck’s constant h ∈]0, h0], h0 > 0. Let Hbe a complex separable Hilbert space with an orthogonal decomposition

H = HR0 ⊕ L2(Rn \B(0, R0)), B(0, R0) = x ∈ Rn; |x| < R0, R0 > 0.(7.1)

The orthogonal projections will be denoted by u 7→ u|B(0,R0), u 7→ u|Rn\B(0,R0)

.

Sometimes we let the characteristic functions 1B(0,R0), 1Rn\B(0,R0) indicate thesame projections.

Consider an unbounded self-adjoint operator depending on h:

P : H → H (7.2)

with domain D = D(P ). Recall that D is a Hilbert space with norm ‖u‖D =‖(P + i)u‖H. Let Hs(Rn) be the standard Sobolev space, but equipped

73

with the h-dependent norm ‖u‖Hs = ‖〈hD〉su‖L2 . Assume that for every1B(0,R0) ≺ χ ∈ C∞0 (Rn) independent of h,

the multiplication operators 1− χ : L2(Rn)→ H, H → L2(Rn) (7.3)

restrict to bounded operators H2(Rn)→ D, D → H2(Rn),

with norms bounded uniformly in h.

Assumeχ(P + i)−1 is compact, (7.4)

for every χ as in (7.3). This assumption will be strengthened later. Furtherassume that

(Pu)|Rn\B(0,R0)= Q(u|Rn\B(0,R0)

), u ∈ D, (7.5)

where

Qu =∑|α|≤2

aα(x;h)(hDx)αu, (7.6)

aα(x;h) = aα(x) is independent of h for |α| = 2,

and aα is bounded in C∞b (Rn) when h varies .

We further assume that Q is formally self-adjoint on Rn with∑|α|=2

aα(x)ξα ≥ 1

C|ξ|2, (7.7)

∑|α|≤2

aα(x;h)ξα → ξ2, |x| → ∞, uniformly w.r.t. h. (7.8)

Example 1. Let O ⊂⊂ Rn be an open set with smooth boundary and letP be the Dirichlet or Neumann realization of −∆ in L2(Rn \ O). Then (cf.Chapter 2), h2P satisfies the assumptions above, with H = L2(Rn \ O).(Take R0 large enough so that the ball of radius R0 contains O.)

Example 2. P = −h2∆ +V (x), where V ∈ C∞b (Rn) with V (x)→ 0, x→∞.H = L2(Rn).

Let R > R0, T = (R/RZ)n, R > 2R. We can view B(0, R) as a subsetof T . Put

H] := HR0 ⊕ L2(T \B(0, R0)), (7.9)

with orthogonal decomposition. We define a self-adjoint operator

P ] : H] → H]

74

as follows: The domain D] = D(P ]) is by definition

u ∈ H];χu ∈ D, (1− χ)u ∈ H2(T )

where 1B(0,R0) ≺ χ ∈ C∞0 (B(0, R); [0, 1]). (This does not depend on thechoice of χ.) Let

Q]u =∑|α|≤2

a]α(x;h)(hD)α

be a formally selfadjoint differential operator on T satisfying (7.6), (7.7) andwith a]α(x;h) = aα(x;h) for |x| < R. With χ as in the definition of D], weput

P ]u = Pχu+Q](1− χ)u, u ∈ D]. (7.10)

We have

Proposition 7.1 P ] is self-adjoint with discrete spectrum.

Proof. It is easy to see that P ] is symmetric. Let u ∈ D((P ])∗), and write(P ])∗u = v, with v ∈ H]. Then

(u|Q]φ) = (v|φ),

for all u ∈ C∞0 (T \ B(0, R0)), so Q]u = v in T \ B(0, R0), in the sense ofdistributions. From the ellipticity of the operator in this region, it followsthat

u|T\B(0,R0)∈ H2

loc(T \B(0, R0)).

After subtracting an exterior part from u, of class H2 and hence belongingto D(P ]), we may assume that χu = u, with χ as in the definition of D(P ]).

We now have(u|Pφ) = (v|φ), (7.11)

for all φ ∈ D with support in B(0, 2R). (Here we may view u, v as elementsof H since their supports are contained in B(0, R).) Because of the supportproperties of u, v, we see that (7.11) extends to general φ ∈ D and theselfadjointness of P implies that u ∈ D, and hence u ∈ D] by the supportproperty of u. This finishes the proof of the self-adjointness of P ].

To prove that P ] has discrete spectrum, we shall use the assumption (7.4).Let vj ∈ H], j = 1, 2, .. be a bounded sequence and put uj = (P ] + i)−1vj ∈D]. With χ as above, we notice that (1 − χ)(P ] + i)−1vj = (1 − χ)uj is a

75

bounded sequence in H2(T ) and hence has a convergent subsequence in H].As for χuj, we compute

(P + i)χuj = [Q,χ]uj︸ ︷︷ ︸bounded inH

+ χvj︸︷︷︸bounded inH

=: wj,

so χuj = (P + i)−1wj = (P + i)−1χwj, if χ ≺ χ ∈ C∞0 . Now we noticethat by the resolvent identity, (7.4) implies the compactness of χ(P − z)−1

for every z ∈ C \ σ(P ). Here we can replace χ by χ and the compactnessof χ(P − i)−1 implies that of (P + i)−1χ by passing to adjoints. Hence χujalso has a convergent subsequence. It follows that (P ] + i)−1 is compact orequivalently, that P ] has a purely discrete spectrum. #

Later on we shall need assumptions on the density of eigenvalues of P ].Let N(P ], I) denote the number of eigenvalues of P ] in the interval I. Weassume

N(P ], [−λ, λ]) = O(1)Φ(λ

h2), λ ≥ 1. (7.12)

Here Φ : [1,∞[→ [1,∞[ is a continuous strictly increasing function withΦ(t) ≥ tn/2 such that:

For every C ≥ 1, there exists C(C) > 0, such that Φ(Ct) ≤ C(C)Φ(t).(7.13)

Let us next verify the invariance of the last assumption. Consider two self-adjoint reference operators P ], P [ for the same operator P , on two differenttorii which both contain B(0, R0). Then P ] and P [ coincide near B(0, R0).We make all the general assumptions above for both operators, except (7.12),and we shall see in the end that if P ] satisfies (7.12), then so does P [. Afterapplying a diffeomorphism which is equal to the identity near B(0, R0), wemay assume that T ] = T [ =: T so that P ] and P [ live in the same Hilbertspace H] = H[ and have the same domain D] = D[. Since we cannot takethe diffeomorphism with Jacobian 1 in general, the two Hilbert spaces willcarry different though equivalent norms uniformly with respect to h.

Let z vary in a bounded subset Ω of C \ R. Let Q], Q[ be ellipticdifferential operators on T , self-adjoint for the respctive inner products, whichcoincide with P ] and P [ respectively in T \ B(0, R0). Let ψ0 ≺ ψ1 ≺ ψ2

belong to C∞(T ; [0, 1]) with ψ0 equal to 1 near the closure of B(0, R0) andwith P ] = P [ near suppψ2. As an approximation for (z − P [)−1, we try

E[ = ψ2(z − P ])−1ψ1 + (1− ψ0)(z −Q[)−1(1− ψ1). (7.14)

76

Then

(z−P [)E[(z) = 1− [Q], ψ2](z−P ])−1ψ1 +[Q], ψ0](z−Q[)−1(1−ψ1). (7.15)

For the last two terms we use telescopic formulas: Let θ0 ≺ θ1 ≺ .. ≺ θN bein C∞(T ) with supp θN ∩B(0, R0) = ∅. Iterating the identity

(z − P ])−1θj = θj+1(z −Q])−1θj + (z − P ])−1[Q], θj+1](z −Q])−1θj,

and similarly with P ] replaced by Q], and where θj may be replaced by[Q], θj], we get

(z − P ])−1θ0 =

∑Nj=1 θj(z −Q])−1[Q], θj−1](z −Q])−1[Q], θj−2]..[Q], θ1](z −Q])−1θ0

+(z − P ])−1[Q], θN ](z −Q])−1[Q], θN−1]..[Q], θ1](z −Q])−1θ0.(7.16)

The adjoint formula is:

θ0(z − P ])−1 =

∑Nj=1 θ0(z −Q])−1[θ1, Q

]]..[θj−2, Q]](z −Q])−1[θj−1, Q

]](z −Q])−1θj

+θ0(z −Q])−1[θ1, Q]]..[θN−1, Q

]](z −Q])−1[θN , Q]](z − P ])−1.

(7.17)For the second term to the right in (7.15), take θ0, .., θN as above with

θ0 = 1 near supp[Q], ψ2] and with supp θN ∩ suppψ1 = ∅. Then (7.17) gives

[Q], ψ2](z − P ])−1ψ1 =

±[Q], ψ2]θ0(z −Q])−1[Q], θ1]..[Q], θN−1](z −Q])−1[Q], θN ](z − P ])−1ψ1.(7.18)

Here ‖[Q], θN ](z−P ])−1ψ1‖L(H,L2) ≤ C/|Im z|, where we used the assumption(7.3) which implies the corresponding fact for P ]. On the other hand,

[Q], ψ2]θ0(z −Q])−1[Q], θ1]...[Q], θN−1](z −Q])−1

= ON(1) hN

|Im z|N : L2 → HN ,

77

where we equip HN with the h-dependent norm ‖〈hD〉Nu‖L2 . ChoosingN > n it follows that the last operator is of trace class as an operator inL2 with trace class norm bounded by ON(1)hN−n/|Im z|N . We have thenshowed that [Q], ψ2](z−P ])−1ψ1 is trace negligible (cf.[80]) in the sense thatfor every N ∈ N there exists M(N) ≥ 0, such that

‖[Q], ψ2](z − P ])−1ψ1‖tr = ON(1)hN

|Im z|M(N).

Here ‖ · ‖tr = ‖ · ‖C1 denotes the trace class norm. Similarly we get the samefact for[Q], ψ0](z −Q[)−1(1− ψ1).

From (7.14), (7.15), we get

(z − P [)−1 = ψ2(z − P ])−1ψ1 + (1− ψ0)(z −Q[)−1(1− ψ1) +R[(z), (7.19)

where R[ is trace negligible. Actually we can replace (z−P ])−1 by (z−P [)−1,changing only the trace negligible part. Similarly,

(z − P ])−1 = ψ2(z − P ])−1ψ1 + (1− ψ0)(z −Q])−1(1− ψ1) +R](z), (7.20)

where R](z) is trace negligible. If f ∈ C∞0 (R), we have the simple Cauchyformula ([22]), much exploited since the work of [39]:

f(P ]) = − 1

π

∫ ∂f

∂z(z − P ])−1L(dz), (7.21)

valid more generally for P ] replaced by any self-adjoint operator. Here f ∈C∞0 (C) is an almost analytic extension of f , i.e. with ∂f/∂z vanishing toinfinite order on R, and L(dz) is the Lebesgue measure on C. It follows fromthis and (7.19), (7.20), that

f(P ])− f(P [) = (1− ψ0)(f(Q])− f(Q[))(1− ψ1) +K,

where ‖K‖tr = O(h∞). On the other hand, we know that ‖f(Q])‖tr ,‖f(Q[)‖tr are O(h−n), so we conclude that

‖f(P ])− f(P [)‖tr = O(h−n). (7.22)

Let I1 ⊂⊂ I2 ⊂⊂ I3 ⊂⊂ R be fixed open intervals. Choosing f ∈ C∞0 (R)first with 1I1 ≺ f ≺ 1I2 and then with 1I2 ≺ f ≺ 1I3 , we get

N(I1, P[)−O(h−n) ≤ N(I2, P

]) ≤ N(I3, P[) +O(h−n). (7.23)

78

It is now clear that if P ] satisfies (7.12) with λ = 2, then P [ does so withλ = 1.

More generally, assume that P ] satisfies (7.12) for all λ ≥ 1. Then we canview λ−1P ] as a reference operator associated to λ−1P , and all assumptionsabove, including (7.12) are satisfied uniformly with h replaced by h = h/

√λ.

The previous argument gives (7.12) for λ−1P [ with h replaced by h and withλ replaced by 1/2. Making the inverse substitutions, we recover (7.12) forP [. #

7.2 Analytic distorsions.

This classical approach to resonances was initiated by Aguilar-Combes [1],Balslev-Combes [5], and then followed by many others, [70, 43]. Here werecall the approach of [80] with some minor modifications of [76], and werefer to these papers for more details and references.

A smooth (C∞) manifold Γ ⊂ Cn is called totally real if

TxΓ ∩ i(TxΓ) = 0, ∀x ∈ Γ. (7.24)

If Γ is totally real, then dim Γ ≤ n, and a natural example of a maximallytotally real (m.t.r.) manifold (i.e. of maximal dimension n) is Γ = Rn.(The opposite extreme of totally real manifolds is given by complex mani-folds, for which TxΓ = iTxΓ.) Totally real manifolds are mapped to totallyreal manifolds under holomorphic diffeomorphisms, so the notion extends tosubmanifolds of complex manifolds.

Let Γ ⊂ Cn be smooth of real dimension n. Locally we have Γ = f(Rn),where f : Rn → Cn is smooth with injective differential. Let f : Cn → Cn

be an almost analytic extension of f so that ∂f vanishes to infinite order onRn. Let y ∈ Rn. Then since df is complex linear, iTf(y)Γ = (df(y))(iTyR

n).Hence Γ is totally real near f(y) iff

df(y)(TyRn) ∩ df(y)(iTyR

n) = 0,

i.e. iff df is injective, or more explicitly iff

det (∂fj(y)

∂yk) 6= 0 (7.25)

If Γ is m.t.r. and u ∈ C∞(Γ), then locally we can find an almost analyticextension u ∈ C∞(Cn) with u|Γ = u, ∂u = 0 to infinite order on Γ. If u is a

79

second a.a. extension of the same function u, then u− u vanishes to∞ orderon Γ. In fact, let f be as above and let f be an a.a. extension, so that f isa local diffeomorphism according to (7.25). Then the inverse of f is a.a. onΓ and we can relate a.a. extensions from Rn and from Γ (locally) by meansof f and its inverse.

Let Ω ⊂ Cn be open, Γ ⊂ Ω a m.t.r. manifold. Let

P (x,Dx) =∑|α|≤m

aα(x)Dαx (7.26)

be an differential operator with holomorphic coefficient on Ω. We can definePΓ : C∞(Γ)→ C∞(Γ) by

PΓu = (P (x,Dx)u)|Γ, (7.27)

where u is a local a.a. extension of u. If f is as above then

(PΓu) f = Q(y,Dy)(u f), (7.28)

Q(y,Dy) =∑|α|≤m

(aα f)(y)((t∂yf)−1Dy)α. (7.29)

In particular PΓ is a differential operator of order m with smooth coefficients.The principal symbol q of Q is related to the principal symbol p of P by theusual relation

q(y, η) = p(f(y), (t∂f(y))−1η).

More invariantly, if we identify T ∗Γ with a submanifold of Cn ×Cn via themap

T ∗Γ 3 (x, dφ(x)) 7→ (x, ∂xφ(x)) ∈ Γn ×Cn,

for φ ∈ C∞(C; R), then the principal symbol pΓ of PΓ is given by

pΓ = p|T ∗Γ. (7.30)

The quickest way to see this is to consider

pΓ(x, dφ(x)) = limh→0

e−iφ(x)/hPΓ(eiφ(x)/h).

This remark also carries over to principal symbols in the sense of h-differentialoperators: Let

80

P (x, hDx) =∑|α|≤m

aα(x;h)(hDx)α

with aα holomorphic in Ω, aα = a0α(x) +O(h), and define the corresponding

principal symbol byp(x, ξ) =

∑|α|≤m

a0α(x)ξα.

Correspondingly we can consider h-differential operators on manifolds anddefine their principal symbols in the sense of h-differential operators. (7.30)still holds.

We need a deformation result.

Lemma 7.2 Let ω ⊂ Rn be open and let f : [0, 1]×ω 3 (t, y) 7→ f(t, y) ∈ Cn

be a smooth proper map such that det(∂f∂y

) 6= 0, ∀ (t, y) and such that f(t, ·)is injective. Assume that f(t, y) = f(0, y) for all y ∈ ω \K, where K is somecompact subset of ω. Let P (x,Dx) be a differential operator with holomorphiccoefficients defined in a neighborhood of f([0, 1]× ω) such that PΓt is ellipticfor 0 ≤ t ≤ 1, where Γt = f(t× ω). If u0 ∈ D′(Γ0) and PΓ0u0 extends to aholomorphic function in a neighborhood of f([0, 1] × ω), then u0 extends toa possibly multivalued holomorphic function in a neighborhood of the sameset. More precisely, for every t ∈ [0, 1], there is a holomorphic function utdefined in a neighborhood of Γt such that ut = us near Γs when |t−s| is smallenough, and u0 is an extension of u.

This result is standard and was proved in [80]. We follow that work:

Proof. We shall use an FBI type argument. Let Γ be a m.t.r. manifold. Weuse Lebeau’s resolution of the identity [51]: When Γ = Rn, we can representthe Dirac measure δ as

δ(x) =∫ei(x·ξ+iλ|ξ|x

2/2)bλ(x, ξ)dξ, (7.31)

where λ > 0 and bλ(x, ξ) = 1(2π)n

(1 + iλx·ξ2|ξ| ). This integral as well as the

similar ones below should be interpreted as oscillatory integrals, i.e. as thelimit in D′, when ε→ 0 of

δε(x) =∫ei(x·ξ+iλ|ξ|x

2/2)bλ(x, ξ)e−εξ2/2dξ,

81

and where the precise expression of the convergence factor exp(−εξ2/2) isnot of great importance, since it suffices to have a family χε(ξ) of functionsin S which is bounded in C∞b and converges to 1 uniformly on every compactset when ε→ 0. This means that if u ∈ C∞0 (Rn):

u(x) =∫∫

ei((x−y)·ξ+iλ|ξ|(x−y)2/2)bλ(x− y, ξ)u(y)dydξ. (7.32)

Put φα(x, y, α) = (x − y) · αξ + iλ|αξ|((x − αx)2 + (y − αx)2), aλ(x, y, α) =c(λ)|αξ|

n2 b(x−y, αξ) for a suitable c(λ) > 0, α = (αx, αξ). Then by evaluating

a Gaussian integral in αx, we get from (7.32) :

u(x) =∫∫

y∈Rnα∈R2n

eiφλ(x,y,α)aλ(x, y, α)u(y)dydα. (7.33)

We now pass to the case of general Γ, replacing |αξ| by its holomorphic

extension√α2ξ , and work near some fixed point x0 ∈ Γ. We may assume

x0 = 0, T ∗x0Γ = Rn. With λ > 0 sufficiently large, φ = φλ, a = aλ, we put

Au(x) =∫∫

Γ×T ∗Γeiφ(x,y,α)a(x, y, α)χ1(αx)χ(y)u(y)dydα, x ∈ Γ, (7.34)

where χ ∈ C∞0 (Γ) is equal to 1 near 0 and χ ≺ χ1 ∈ C∞0 (Γ). Notice that

Imφ ∼ |αξ|(|x− αx|2 + |y − αx|2).

By Stokes’ formula, we can change the integration contour to y ∈ Γ, αx ∈Γ, αξ ∈ T ∗y (Γ). (This is a change in αξ only, and we can use the intermediatecontours given by y, αx ∈ Γ, αξ ∈ T ∗[ty+(1−t)αx]Γ, where [z] denotes the pointin Γ which is closest to z, for z in a small complex neighborhood of x0.) Thenwe can integrate out αx and get

Au(x) ≡∫y∈Γ

∫ξ∈T ∗y Γ

ei((x−y)·ξ+iλ|ξ|(x−y)2/2)bλ(x− y, ξ)χ(y)u(y)dydξ =: Au(x),

(7.35)modulo a term Ku(x), which extends holomorphically to a u-independentneighborhood of x0. By analytic continuation from (7.32) we see that thedistribution kernel of A vanishes outside the diagonal and since A is a pseu-dodifferential operator of order 0, we have

Au = a(x)u(x). (7.36)

82

Testing A on oscillatory functions and using Stokes’ formula, we see thata = χ(x), so we end up with

χu(x) ≡ Au(x). (7.37)

Using the ellipticity of P , we can construct an analytic symbol c(x, y, α),with αξ as the large parameter (see for instance [73]) of order n

2−m, defined

in a conic neighborhood of x0, x0× T ∗x0Γ, holomorphic in x, y, αx and with

∂αξc exponentially small (O(e−|αξ|/C) for some C > 0), such that

tP (y,Dy)(eiφ(x,y,α)c(x, y, α)) = eiφ(x,y,α)(a(x, y, α)− r(x, y, α)), (7.38)

where r is holomorphic in x, y and of exponential decrease:

|r(x, y, α)| ≤ Ce−|αξ|/C . (7.39)

Using this, we get for x in a neighborhood of x0 and assuming suppχ small:

Au(x) =∫∫

eiφ(x,y,α)r(x, y, α)χ1(αx)(χu)(y)dydα (7.40)

+∫∫

eiφ(x,y,α)c(x, y, α)χ(y)χ1(αx)Pu(y)dydα

+∫∫

eiφ(x,y,α)c(x, y, α)χ1(αx)[P, χ]u(y)dydα.

Notice that (tP )Γ = t(PΓ).The first and the last integrals extend holomorphically to a u-independent

neighborhood of x0. For the middle integral, we get the same conclusion, ifwe assume that Pu extends to some fixed neighborhood Ω of x0. To see thiswe notice that the exponential factor eiφ(x,y,α) become exponentially small, ifwe make a small α-dependent deformation of the integration in y near y = x0

and restrict x to sa small complex neighborhood of that point.Summing up, we have proved: Let P be holomorphic near Γ with PΓ el-

liptic. If Γ is m.t.r., x0 ∈ Γ, u ∈ D′(Γ) and Pu extends to some neighborhoodΩ of x0, then u extends to another neighborhood Ω of x0, which does notdepend on u. Also Ω will not collapse, if we vary Γ reasonably. From thisthe lemma follows #

We now return to the operators in section 7.1 and add an assumption:

83

There exist θ0 ∈]0, π[, ε > 0, R1 ≥ R0, such that the (7.41)

coefficients aα(x;h) of Q extend holomorphically in x to

rω; ω ∈ Cn, dist (ω, Sn−1) < ε, r ∈ ei[0,θ0]]R1,+∞[,and the relevant parts of (7.6), (7.8) remain valid

in the sense that aα and all its x-derivatives are bounded for x in this largerset, uniformly in h, and the convergence in (7.8) remains valid when x tendsto infinity in the larger set above, again uniformly in h.

For given ε0 > 0, R1 > R0, we can construct a smooth function[0, π]× [0,∞[3 (θ, t) 7→ fθ(t) ∈ C, injective for every θ, with the properties

(i) fθ(t) = t for 0 ≤ t ≤ R1,

(ii) 0 ≤ arg fθ(t) ≤ θ, ∂tfθ 6= 0,

(iii) arg fθ(t) ≤ arg ∂tfθ(t) ≤ arg fθ(t) + ε0,

(iv) fθ(t) = eiθt, for t ≥ T0, where T0 only depends on ε0 and R1.

Consider the map

κθ : Rn 3 x = tω 7→ fθ(t)ω ∈ Cn, t = |x|.

The image Γθ is a m.t.r. manifold which coincides with Rn along B(0, R1).We choose R1 at least as large as R1 in the assumption (7.41).

Let Hθ = HR0 ⊕ L2(Γθ \ B(0, R0)). By means of κθ we can identify Hθ

with H and similarly we can define Dθ = D. Let χ ∈ C∞0 (B(0, R1)) be equalto 1 near B(0, R0) and define the unbounded operator Pθ : H → H withdomain D by

Pθu = P (χu) +QΓθ(1− χ)u.

Here we assume that 0 ≤ θ ≤ θ0.

Parametrizing Γθ by means of κθ we get:

−∆Γθ = (1

f ′(t)Dt)

2 − n− 1

f(t)f ′(t)iDt + (f(t))−2D2

ω, f = fθ,

84

where −D2ω denotes the Laplacian on Sn−1. If ω∗2 denotes the principal

symbol of D2ω and τ is the dual variable of t, then the principal symbol of

−∆Γθ is

p0,θ =τ 2

f ′(t)2+

ω∗2

f(t)2, (7.42)

so pointwise −∆Γθ is elliptic (if ε0 > 0 is small enough) and the principalsymbol p0,θ takes its values in a sector of angle ≤ 2ε0 and globally it takesits values in the sector

−2(θ + ε0) ≤ arg p0,θ ≤ 0. (7.43)

Since θ ≤ θ0, this sector is of angle < 2π (when ε0 is small enough).

Choosing R1 large enough, we get

In Rn \B(0, R0), h−2Pθ is an elliptic differential operator (7.44)

whose principal symbol in the classical sense over each fixed

point in Γθ takes its values in a sector of angle ≤ 3ε0,

and globally in a sector − 2θ − 3ε0 ≤ arg z ≤ ε0

The coefficients of Pθ − e−2iθ(−h2∆Rn) (7.45)

(as a semiclassical differential operator) tend to 0 uniformly w.r.t.

h, when Γθ 3 x→∞, and we identify Γθ and Rn by means of κθ.

Lemma 7.3 If z ∈ C\0, arg z 6= −2θ, then Pθ−z : D → H is a Fredholmoperator of index 0.

The proof [75] is a minor modification of the corresponding one in [80]:

Proof. This has nothing to do with the smallness of h, so we assume h = 1for simplicity. We first show that Pθ − z is a Fredholm operator, and forthat it suffices to invert Pθ − z modulo compact operators. On Γθ ' Rn, weintroduce a smooth partition of unity :

1 = θ1 + θ2 + θ3, (7.46)

where 1B(0,R0) ≺ θ1 ∈ C∞0 (B(0, R1)), θ2 ∈ C∞0 , supp θ3 ⊂ x ∈ Rn; |x| ≥ R,R 1. Further, 0 ≤ θj ≤ 1. Let χ1, χ2, χ3 have the same properties as θjwith the exception of (7.46) and with θj ≺ χj. Let z0 ∈ C \R, and put

E(z) = χ1(P − z0)−1θ1 +QQθ2 + χ3(−e−2iθ∆− z)−1θ3,

85

where Q is a properly supported parametrix of Pθ − z (i.e. a pseudodifferen-tial operator which inverts the latter operator modulo smoothing operatorsoutside a small neighborhood of the closure of B(0, R0) with the propertythat the distribution kernel is supported in a thin neighborhood of the diag-onal). Here θ3, χ3 have their support in the region where Γθ = eiθRn. Thenwe get

(Pθ − z)E(z) = 1 +K(z),

with

K(z) = (z0 − z)χj(P − z0)−1θ1 + [Pθ, χ1](P − z0)−1θ1 + ((Pθ − z)Q− 1)θ2

+[Pθ, χ3](−e−2iθ∆− z)−1θ3 + χ3(Pθ − (−e−2iθ∆))(−e−2iθ∆− z)−1θ3

= K ′(z) +K ′′(z),

where K ′′(z) is the last term in the middle member. With R large enough,we can arrange so that ‖K ′′(z)‖L(L2,L2) < 1/2. On the other hand, K ′(z) iscompact. Now write

(Pθ − z)E(z)(1 +K ′′(z))−1 = 1 +K ′(1 +K ′′)−1,

to get a right inverse modulo a compact operator.As an approximate left inverse, we put

F (z) = θ1(P − z0)−1χ1 + θ2Q+ θ3(−e−2iθ∆− z)−1χ3.

ThenF (Pθ − z) = 1 + L(z), L(z) = L′(z) + L′′(z), (7.47)

where

L′(z) = I + II + III + IV ,

L′′ = θ3(−e−2iθ∆− z)−1(Pθ − (−e−2iθ∆))χ3,

I = (z0 − z)θ1(P − z0)−1χ1,

II = −θ1(P − z0)−1[P, χ1],

III = θ2(Q(Pθ − z)− 1),

IV = −θ3(−e−2iθ∆− z)−1[Pθ, χ3].

We can arrange so that ‖L′′‖L(Dθ,Dθ) is a small as we like, and we turn tothe proof that L′ is compact: Dθ → Dθ.

86

Clearly, III, IV have this compactness property. For I and II, we first seethat they are bounded: Dθ → Dθ, so it is enough to show that (P − z0)I,(P − z0)II are compact: Dθ → Hθ. We have

(P − z0)I = (z − z0)θ1 + (z0 − z)[P, θ1](P − z0)−1χ1,

where θ1 is compact: D → H, and [P, θ1](P − z0)−1χ1 is even compact as anoperator : H → H. Similarly,

(P − z0)II = θ1[P, χ1]︸ ︷︷ ︸=0

− [P, θ1](P − z0)−1[P, χ1]︸ ︷︷ ︸compact:D→H

is compact. Hence L′ is compact Dθ → Dθ and we invert (Pθ − z) to the leftmodulo a compact operator Dθ → Dθ, by means of (1 + L′′)−1F :

(1 + L′′)−1F (Pθ − z) = 1 + (1 + L′′)−1L′.

We have then showed that Pθ − z : Dθ → Hθ is Fredholm. This operatordepends contiuously on (θ, z), so the index is constant under deformation in(θ, z) with z 6∈ e−2iθ[0,∞[. Deforming to θ = 0, z = i, we see that the indexis 0. #

Corollary 7.4 A point z in C\e−2iθ[0,∞[ belongs to σ(Pθ) iff N (Pθ−z) 6= 0.

Proposition 7.5 Assume that 0 ≤ θ1 < θ2 ≤ θ0, and let z0 ∈ C\e−2i[θ1,θ2][0,∞[.Then

dimN (Pθ2 − z0) = dimN (Pθ1 − z0).

Proof. We follow [80] ([75]). Again the statement does not depend on thesmallness of h, so we take h = 1 for simplicity. We shall prove that the twokernels are ”equal” via holomorphic extension. We shall apply Lemma 7.2to suitable deformations of the contours. Let χ ∈ C∞0 (]1

2, 3

2[; [0, 1]) be equal

to 1 near 1 and put for T ≥ 1:

Fθ1,θ1,T (t) = fθ1(t) + χ(t

T)(fθ2(t)− fθ1(t)).

Let Γθ1,θ2,T be the corresponding submanifold of Cn. If θ2 − θ1 is smallenough, Lemma 7.2 is applicable to the family of m.t.r. manifolds,

[0, 1] 3 t 7→ Γθ1,θ1+t(θ2−θ1),T .

87

We conclude that if (Pθ1 − z0)uθ1 = 0, u ∈ Dθ1 , then uθ1 extends to aholomorphic function in a neighborhood of ∪θ1≤θ≤θ2(Γθ \ B(0, R1)), and byrestriction, we get an element uθ ∈ Dloc

θ with (Pθ − z0)uθ = 0.We also need to control uθ near infinity, and for that we shall use the

assumption on z0 which assures ellipticity near ∞.Let 0 < a1 < a2 < b2 < b1, Ωj = z ∈ Rn; aj < |z| < bj, j = 1, 2. The

change of variables y = Tz transforms −∆y into − 1T 2 ∆z. We take T ≥ 1

and view the latter operator as semi-classical one with h = 1/T , and use thestandard a priori estimate∑|α|≤2

‖(T−1Dz)αu‖L2(Ω2) ≤ C(‖(−T−2e−2iθ∆z − z0)u‖L2(ω1) + ‖u‖L2(Ω1\Ω2)),

with C independent of T . In the y-variables this becomes∑|α|≤2

‖Dαy u‖L2(TΩ2) ≤ C(‖(−e−2iθ∆y− z0)u‖L2(TΩ1) + ‖u‖L2(T (Ω1\Ω2))). (7.48)

For θ2−θ1 small enough and T large enough, and for t/T near suppχ, wemay view PΓθ1,θ2,T

as a small perturbation of −e−2iθ1∆y. We conclude that

‖uθ2‖H2(TΩ) ≤ C‖uθ1‖L2(TV ).

Now cover a neighborhood of ∞ in Γθ2 by a sequence TkΩ, where Tk =2kT0. If uθ1 ∈ Dθ1 , it follows that uθ2 ∈ Dθ2 .

To sum up, if θ2 − θ1 is small enough, we have constructed an injec-tive linear map (by holomorphic extension and restriction): N (Pθ1 − z0) →N (Pθ2 − z0) The inverse map is constructed the same way, so dimN (Pθ1 −z0) = dimN (Pθ2 − z0).

When θ2− θ1 is larger, we introduce intermediate θ-values and apply theabove. #

Two extensions of the lemma are possible:

1) We may consider inhomogeneous equations for (Pθ− z) assuming suit-able holomorphic extension properties of the right hand side.

2) We may consider N ((Pθ − z)k) for k ∈ N.

The preceding result shows that for 0 ≤ θ1 ≤ θ ≤ θ2 ≤ θ0, the spectrumof Pθ in C \ e−2i[θ1,θ2][0,+∞[ is independent of θ. In particular Pθ will haveno spectrum in a small sector ei]0,ε[]0,+∞[. Analytic Fredholm theory showsthat the spectrum of Pθ in C \ e−2iθ[0,+∞[ is discrete. In particular for

88

θ = 0, we see that the spectrum in ] − ∞, 0[ is discrete. If 0 < θ < π2,

the spectrum of Pθ outside e−2iθ[0,+∞[ consists of the negative eigenvaluesof P0 and the eigenvalues in e−i[0,2θ[]0,+∞[. When θ ≥ π

2, we only get the

latter eigenvalues. By definition, the resonances of P are the eigenvalues ofPθ in the sector e−2i[0,θ[]0,+∞[, and as we have seen, they are independentof θ in the sense that replacing θ by a larger value will not change the set ofresonances in the sector e−2i[0,θ[]0,+∞[.

By analytic Fredholm theory, if z0 ∈ e−i[0,2θ[]0,+∞[, 0 ≤ θ ≤ θ0 is aresonance, then the spectral projection

πθ,z0 =1

2πi

∫γ(z0)

(z − Pθ)−1dz (7.49)

is of finite rank, where γ(z0) denotes the positively oriented boundary of asmall disc centered at z0, which contains no other resonances than z0. Thisrank is independent of the choice of θ for which z0 belongs to the sectorwhere the resonances are defined, and by definition it is the multiplicityof the resonance z0. The image Fθ,z0 is contained in the domain of anypower of P and is invariant under Pθ. (Pθ − z0)|Fθ,z0

is nilpotent and Fθ,z0 =

N ((Pθ−z0)k0) for some k0 ∈ N. The proof of Proposition 2.4 gives a bijectionFθ1,z0 → Fθ2,z0 under the assumptions there.

To end this chapter we mention that resonances defined by means ofcomplex distorsions, coincide with other definitions. (See [37] for generalresults in this direction.) In particular, one can show that the resolvent(z − P )−1 : Hcomp → Dloc, defined first for Im z > 0, can be extended mero-morphically across ]0,+∞[ to the sector e−i[0,2θ0[]0,+∞[ and the poles arejust the resonances. [[[Add the proof of this statement]]]. The multiplicity ofa resonance is the rank of the formal spectral projection, obtained by replac-ing the resolvent of Pθ in the integral (7.49) by the meromorphic extensionof the resolvent of P . See [80].

The set of resonances will be denoted by ResP and the elements of thisset will be counted (repeated) according to their multiplicity.

89

8 The local trace formula and local upper

bounds

8.1 Some trace estimates

Let P· : H → H, · = 0, 1, be two selfadjoint operators satisfying the assump-tions in the preceding section. Here

H· = H·,R0 ⊕ L2(Rn \B(0, R0)) (8.1)

is an orthogonal sum. In particular for the reference operators P ]· , we have

corresponding estimates #(σ(P·) ∩ [−λ, λ]) ≤ O(1)Φ·(λh2 ). In view of (7.13),

we notice that Φ· are of at most polynomial growth:

Φ·(t) = O(tn·/2) (8.2)

for some exponents n· ∈ [n,+∞[.

To the earlier assumptions, we add:

|[a·,α(x;h)]10| ≤ O(1)〈x〉−n, for some n > n (8.3)

not only in the real domain but also in the complex domain appearing in(7.41). Here θ0 > 0 is assumed to be the same for both operators P1, P0, andwe write [a·]

10 = a1 − a0.

Proposition 8.1 Let f ∈ C∞0 (R) be independent of h or vary in a boundedsubset of C∞0 (R). Let χ ∈ C∞0 (Rn) be equal to 1 near B(0, R0). Thenχf(P·), f(P·)χ, [(1− χ)f(P·)(1− χ)]10 are of trace class and

′′tr [f(P·)]10′′ := (8.4)

[tr (χf(P·)χ+ χf(P·)(1− χ) + (1− χ)f(P·)χ)]10 + tr [(1− χ)f(P·)(1− χ)]10

is independent of the choice of χ and is O(1)Φmax (h−2). Here Φmax :=max(Φ0,Φ1).

Let ε0 > 0 be fixed and small enough, so that 2θ0 +ε0 < 2π. Let W ⊂⊂ Ωbe open relatively compact subsets of ei]−2θ0,ε0]]0,+∞[. We assume that thesesets are independent of h and that Ω is simply connected. Also assume thatthe intersections J+, I+ between Ω, W and ]0,+∞[ are intervals. Let Ω−,W− denote the intersections of ei]−2θ0,0]]0,+∞[ with Ω and W respectively.The local trace formula ([75, 76]) is given in the following result:

90

Theorem 8.2 Fix Ω, W as above. Then for h > 0 sufficiently small thefollowing holds: Let f = f(z;h) be holomorphic for z ∈ Ω and satisfy|f(z;h)| ≤ 1, for z ∈ Ω \ W . Let χ ∈ C∞0 (J+) be independent of h, withχ = 1 near I+. Then uniformly with respect to f :

′′tr [(χf)(P·)]10′′ =

[∑

λ∈ResP·∩W−f(λ)]10 − [

∑µ∈σ(P·)∩]−∞,0[∩W−

f(µ)]10 +O(1)Φmax( 1h2 ).

(8.5)

The proof also shows that the number of resonances of P· in Ω− isO(1)Φ·(h−2).

Later in these notes we will also discuss the more classical trace formulaof Bardos-Guillot-Ralston [6] with an important improvement of R. Melrose[55] and further generalized in [85]. That formula relates all the resoances tothe trace of the wave-group for operators as in Chapter 2, when the dimensionis odd. In the cited works, it is proved by using the Lax–Phillips theory andin particular translation represenations, but following the work [35], Zworskialso gave a proof in [102] which uses more general scattering theory and whichis closer to the proof that we shall give for the local trace formula (following[76]).

It should also be pointed out that the left hand side in the local traceformula is the same as in the Birman–Krein formula for the scattering phase.See [...].

In this section we shall prove Proposition 8.1, by adapting arguments from[80] and the local trace formula will be proved in the next section. Whenworking with the P· separately, we generally drop the subscript “·”. If χ is asin the Proposition, choose a torus T for the definition of a reference operatorP ] as in Chapter 7, and choose χ ∈ C∞0 (Rn) with χ ≺ χ. We may arrangeso that supp χ is naturally included in T and P ] = P near suppχ. Recallthat trace negligible operators were introduced in Chapter 7.

Lemma 8.3 The following operators are trace negligible for z in any fixedbounded set in C \R:

χ(z − P )−1χ− χ(z − P ])−1χ,(1− χ)(z − P )−1(1− χ)− (1− χ)(z −Q)−1(1− χ),χ(z − P )−1(1− χ),χ(z −Q)−1(1− χ).

(8.6)

91

Proof. We start with χ(z − P )−1(1 − χ). Let χ ≺ θN ≺ ... ≺ θ1 ≺ χ,θ ∈ C∞0 . Then we have the telescopic formula:

(z − P )−1(1− χ) =

∑Nj=1(1− θj)(z −Q)−1[Q, 1− θj−1]...(z −Q)−1[Q, 1− θ1](z −Q)−1(1− χ)

+(z − P )−1[Q, 1− θN ](z −Q)−1[Q, 1− θN−1]...[Q, 1− θ1](z −Q)−1(1− χ),

so

χ(z − P )−1(1− χ) =

χ(z − P )−1[Q, 1− θN ](z −Q)−1[Q, 1− θN−1]...[Q, 1− θ1](z −Q)−1(1− χ).

We conclude as in the proof of the invariance of (7.12), and see that χ(z −P )−1(1−χ) is trace negligible. The operator χ(z−Q)−1(1−χ) can be treatedin the same way.

Let χ0 ∈ C∞0 (Rn) be equal to one near B(0, R0) and satisfy χ0 ≺ χ.Write

(z−P )−1(1−χ) = (1−χ0)(z−Q)−1(1−χ)−(z−P )−1[Q,χ0](z−Q)−1(1−χ),

so that

(1− χ)(z − P )−1(1− χ)− (1− χ)(z −Q)−1(1− χ) =

−((1− χ)(z − P )−1χ1)([Q,χ0](z −Q)−1(1− χ)),

where we inserted a cutoff χ1 ∈ C∞0 with χ0 ≺ χ1 ≺ χ, and indicated afactorization of the last term into 2 factors. The first of the two factors istrace negligible, while the second one is bounded of norm O(|Im z|−1), so theoperator above is trace negligible.

The argument for χ(z − P )−1χ − χ(z − P ])−1χ is the same except thatwe use χ instead of 1− χ0 and χ instead of 1− χ. #

Let χ1 ∈ C∞0 (Rn) with χ ≺ χ1, and choose the reference torus largeenough, to include suppχ1. Consider the short telescopic formula:

(z − P )−1χ = χ1(z − P ])−1χ+ (z − P )−1[P, χ1](z − P ])−1χ.

92

The proof of Lemma 8.3 shows that [P, χ1](z − P ])−1χ is trace negligible, sothe same holds for (z − P )−1[P, χ1](z − P ])−1χ.

Let f ∈ C∞0 (R). Then,

f(P )χ = χ1f(P ])χ+O(h∞)

in trace norm. The trace norm of χ1f(P ])χ is O(1)Φ(h−2), so the same holdsfor f(P )χ and by duality for χf(P ).

From the lemma it follows that the trace class norm of (1−χ)f(P )(1−χ)−(1− χ)f(Q)(1− χ) is O(h∞), so in order to prove that the trace class normof [(1− χ)f(P·)(1− χ)]10 is O(h−n), as we shall, it suffices to prove the samefact about [f(Q·)]

10.

Let 1 =∑j∈Zn χj(x) be a partition of unity with χj(x) = χ0(x−j) ∈ C∞0 .

For N ∈ N, consider

(i+Q1)−N(z −Q1)−1 − (i+Q0)−N(z −Q0)−1, (8.7)

which is a linear combination of terms

(i+Q1)−k(Q1 −Q0)(i+Q0)−(N+1−k)(z −Q0)−1 (8.8)

for 1 ≤ k ≤ N , and of the term

(i+Q1)−N(z −Q1)−1(Q1 −Q0)(z −Q0)−1. (8.9)

For (8.8), we consider

χν(i+Q1)−k(Q1 −Q0)χρ(i+Q0)−(N+1−k)(z −Q0)−1χµ. (8.10)

Using a Combes-Thomas argument, (see [66], [39], [75], section 7) [explain..]we see that this operator is bounded: L2 → H2(N+1) of norm

O(1) 1|Im z|〈ρ〉

−ne−1Ch

((|ν−ρ|−C)++|Im z|(|ρ−µ|−C)+),

with t+ = max(t, 0), t ∈ R.

(8.11)

If N ≥ N(n) is large enough, this is also a bound for hn times the trace normfor operators L2 → L2, and summing over ν, µ, we get

‖(i+Q1)−k(Q1 −Q0)χρ(i+Q0)−(N+1−k)(z −Q0)−1‖tr

= O(h−n)〈ρ〉−n 1|Im z|(1 + h

|Im z|)n.

(8.12)

93

For (8.9) we get similarly

‖(i+Q1)−N(z −Q1)−1(Q1 −Q0)χρ(z −Q0)−1‖tr

= O(h−n)〈ρ〉−n 1|Im z|2 (1 + h

|Im z|)2n.

(8.13)

Summing over ρ, we get

‖(i+Q1)−N(z −Q1)−1 − (i+Q0)−N(z −Q0)−1‖tr

= O(h−n) 1|Im z|2 (1 + h

|Im z|)2n.

(8.14)

For f ∈ C∞0 (R), we write f(t) = (t+ i)−Ng(t) with g ∈ C∞0 (R). Then

[f(Q·)]10 = [(i+Q·)

−Ng(Q·)]10

= − 1π

∫ ∂g∂z

(z)[(i+Q·)−N(z −Q·)−1]10L(dz),

(8.15)

where g ∈ C∞0 (C) is an almost analytic extension of g. Using (8.14), we get

‖[f(Q·)]10‖tr = O(h−n), (8.16)

and we conclude (as already mentioned) that

‖[(1− χ)f(P·)(1− χ)]10‖tr = O(h−n). (8.17)

We have then proved everything in Proposition 8.1, except that the righthand side of (8.4) is independent of the choice of χ, which we leave to thereader. #

Remark. Let f ∈ C∞0 (]−∞, 0[). Possibly after increasing R0, we may assumethat (z−Q)−1 exists and is equal to O(1) for z in some fixed complex neigh-borhood of supp f so that f(Q) = 0. Lemma 8.3 together with the operatorCauchy formula then shows that (1− χ)f(P )(1− χ) is trace negligible, nowin the sense that the trace class norm is O(h∞). Choose χ0 ∈ C∞0 (Rn) with1B(0,R0) < χ0 < χ. Then (1− χ0)f(P )(1− χ0) is also trace negligible and

tr f(P ) = trχf(P )χ+ tr (1− χ)f(P )χ+ trχf(P )(1− χ) +O(h∞) .

Here trχf(P )(1−χ) = trχf(P )(1−χ)(1−χ0)2 = trχ(1−χ0)f(P )(1−χ0)(1−χ) = O(h∞). Similarly tr (1 − χ)f(P )χ = O(h∞) so tr f(P ) = trχf(P )χ +

94

O(h∞) Similarly tr f(P ]) = trχf(P ])χ+O(h∞) and from Lemma ltrace1.3and the Cauchy formula we have trχf(P )χ = trχf(P ])χ + O(h∞). Weconclude that

tr f(P ) = tr f(P ]) +O(h∞) .

Remark. If P0, P1 satisfy the assumptions of Proposition ltrace1.1, thenuniformly with respect to λ ≥ 1, the same holds for the operators λ−1P0,λ−1P1, provided that we replace h by the new semi-classical parameter h/

√λ.

The same holds for Theorem 8.2. In particular, if P0, P1 are independent ofh and satisfy all assumptions of Proposition ltrace1.1 (or Theorem ltrace1.2)which make sense for h = 1, then λ−1P0, λ−1P1 do the same with the semi-classical parameter h = 1/

√λ.

8.2 Proof of the local trace formula.

We choose m ∈ N large enough to get nice trace class properties for someoperators below. We take θ = θ0. Let Ω+ be the intersection of Ω withthe sector 0 ≤ arg z < ε0 (for some ε0 > 0), define W+ similarly andrecall that W−,Ω− were defined prior to Theorem 8.2. Fix a point z0 ∈ei[3ε0,min (π,2π−2θ−3ε0)]]0,+∞[ away from σ(P·) and σ(P·,θ) (where we use theobservation that we can always replace ε0 by some smaller number withoutchanging anything in the conclusion of Theorem 8.2) and write

f(z) = (z − z0)−mg(z). (8.18)

Then

(χf)(P·) = (P· − z0)−m(χg)(P·) = (8.19)

− 1

π

∫g(z)

∂χ

∂z(z)(P· − z0)−m(z − P·)−1L(dz),

where χ ∈ C∞0 (Ω) is an almost analytic extension of χ with support in asmall neighborhood of J+ and equal to 1 near I+. We first look at

I−· := − 1

π

∫Ω−g(z)

∂χ

∂z(z)(P· − z0)−m(z − P·)−1L(dz). (8.20)

Let χ ∈ C∞0 (Ω) be equal to 1 near W−, equal to χ near J+ and be almostanalytic near R−, in case Ω intersects that set. If Ω∩R− = ∅, we can replace

95

χ by χ in the last formula. More generally, we have

I−· = −(χ1R−f)(P·)−1

π

∫Ω−g(z)

∂χ(z)

∂z(P· − z0)−m(z − P·)−1L(dz). (8.21)

The support of ∂χ(z)/∂z in Ω− does not intersect W−, so |f | ≤ 1 thereand consequently g = O(1) uniformly w.r.t. h on that set. Choosing mlarge enough, it follows from the first remark after (8.17) and the proof ofProposition 8.1, that

′′tr [I−· ]10′′ = −[

∑µ∈σ(P·)∩R−∩W

f(µ)]10 +O(1)Φmax(1

h2). (8.22)

Next look at

I+· := − 1

π

∫Ω+

g(z)∂χ(z)

∂z(P· − z0)−m(z − P·)−1L(dz).

Green’s formula gives for every δ > 0:

I+· = − 1

π

∫Ω+∩Im z≤δ

g(z)∂χ(z)

∂z(P· − z0)−m(z − P·)−1L(dz) (8.23)

+1

2πi

∫Ω+∩Im z=δ

g(z)χ(z)(P· − z0)−m(z − P·)−1dz,

where the integration contour in the last integral is oriented in the directionof decreasing Re z. If δ > 0 is small enough (independent of h), the integrandin the first integral in (8.23) has its support in the region where g = O(1)and as in the proof of Proposition 8.1, we get

′′tr [I+· ]10′′ = ′′tr [J·]

10′′ +O(1)Φmax(

1

h2), (8.24)

where J· denotes the last integral in (8.23).

Committing another error O(1)Φmax( 1h2 ) in the evaluation of ′′tr [J·]

10′′, we

may replace J· by

J· :=1

2πi

∫γg(z)(P· − z0)−m(z − P·)−1dz, (8.25)

where γ is the segment from b to a, where Im b = Im a = δ, Re a,Re b ∈ J+,Re a < inf I+, Re b > sup I+.

96

Proposition 8.4 We have

′′tr [J·]10′′ = ′′tr [J·]

10′′, (8.26)

where

J· =1

2πi

∫γg(z)(P·,θ − z0)−m(z − P·,θ)−1L(dz), (8.27)

and where P·,θ = P·|Γθ is defined as in section 3, with the parameter ε0 therechosen small enough depending on δ.

Proof. Take χ ∈ C∞(R; [0, 1]) with χ(t) = 0 for t ≤ 0, χ(t) = 1 for t ≥ 1.As in the proof of Proposition 7.5 (but now with a different χ), we put

fθ1,θ2,T (t) = fθ1(t) + χ(t

T)(fθ2(t)− fθ1(t)),

and let Γθ1,θ2,T be the corresponding contour. As in Chapter 7 we assumethat0 ≤ θ1 < θ2 ≤ θ with θ2 − θ1 small enough, and we let P·,θ1,θ2,T denote therestriction of P· to Γθ1,θ2,T . The proof of Proposition 8.1 can be applied toshow that

′′tr [(P·,θ1,θ2,T − z0)−m(z − P·,θ1,θ2,T )−1]10′′ (8.28)

is O(1)Φmax( 1h2 ), depends continuously on T > 0, is equal to

′′tr [(P·,θ2 − z0)−m(z − P·,θ2)−1]10′′

for small T , and tends to

′′tr [(P·,θ1 − z0)−m(z − P·,θ1)−1]10′′,

when T → +∞.

To get the proposition, it suffices to show that the expression (8.28) isindependent of T . Let Γt = f(t, y); y ∈ ω, 0 ≤ t ≤ 1, be a smooth familyof m.t.r. manifolds in Cn, where ω is open and f in C∞, with det(∂f

∂y) 6= 0

everywhere.

If u is holomorphic near Γt for some t, let ut(y) = u(f(t, y)), so that ut isthe restriction u|Γt , expressed in the parametrization f(t, ·). Then

∂ut(y)

∂t= (

∂f

∂t(t, y) · ∂u

∂x)(f(t, y)). (8.29)

97

Since Γt is m.t.r., there exists a complex vector field νt(y,∂∂y

) on Γt (tangent

to Γt), such that∂ut∂t

= νtut. (8.30)

Moreover νt depends smoothly on (t, y) and vanishes outside the support of∂f∂t

(t, ·).Let P be a differential operator with holomorphic coefficients, defined

near Γt. Define Pt = PΓt as in section 3, so that

Ptut = (Pu)t, (8.31)

when u is holomorphic near Γt. Then

νt(Ptut) = ∂t(Ptut) = (∂tPt)(ut) + Pt∂tut = (∂tPt)(ut) + Ptνtut,

and since the functions of the type ut are “sufficiently dense”, we concludethat in the sense of differential operators,

∂tPt = [νt, Pt]. (8.32)

This applies to Γθ1,θ2,T and passes to the resolvents and their compositions:

∂T ((P·,θ1,θ2,T − z0)−m(z − P·,θ1,θ2,T )−1) =

[ν, (P·,θ1,θ2,T − z0)−m(z − P·,θ1,θ2,T )−1],(8.33)

where ν = νθ1,θ2,T is a smooth complex vector field whose support is compactand disjoint from B(0, R0).

When considering (8.28), we may choose the cutoff χ in (8.4) with supportdisjoint from that of ν, so with ψ equal to χ or to 1− χ,

trχ[ν, (P·,θ1,θ2,T − z0)−m(z − P·,θ1,θ2,T )−1]ψ =

trχν (...)−m(..)−1ψ − trχ(...)−m(..)−1 νψ

= 0− tr (...)−m(..)−1ν ψχ = 0,

where we used the cyclicity of the trace. The T -derivative of the expression(8.28) then reduces to the following expression, where we drop θ1, θ2, T for

98

simplicity:

tr ([(1− χ)[ν, (P· − z0)−m(z − P·)−1](1− χ)]10)

= tr [ν, [(1− χ)(P· − z0)−m(z − P·)−1(1− χ)]10].(8.34)

Here [(1−χ)(P·−z0)−m(z−P·)−1(1−χ)]10 is of trace class also as an operatorL2 → H1, so the expression (8.34) vanishes by the cyclicity of the trace. Itfollows that the T -derivative of (8.28) is zero and the proof is complete. #

It remains to study “tr [J·]10”. We shall use a finite rank perturbation of

P·,θ. Let F be a smooth mapping from a neighborhood of ei[−2(θ+ε0),ε0][0,∞[into itself such that

F (z) = z, for |z| large and for z near e−2iθ[0,∞[, (8.35)

Ω is disjoint from the image of F. (8.36)

If θ < π2, we further extend F to be a smooth map from a neighborhood of

ei[−π−ε0,ε0][0,∞[ to itself, by putting F (z) = z for z near ei[−π−ε0,−2θ][0,∞[.

In the appendix (with a suitable choice of Γθ) we shall construct an op-erator P·,θ : D· → H· with the following properties:

K· := P·,θ − P·,θ is of rank O(1)Φmax( 1h2 ) and

O(1) : D(PN· )→ D(PM

· ) for all N,M ∈ N.(8.37)

Notice here that D(P 0· ) = H·.

K· is compactly supported in the sense that K· = χK·χ (8.38)

if χ ∈ C∞0 is equal to 1 on B(0, R) for some sufficiently large R.

For every N ∈ N, (P·,θ − z)−1 = O(1) : D(PN· )→ D(PN+1

· ), (8.39)

uniformly for z ∈ Ω.

Since we shall always work with operators on Γθ in the remainder of theproof, we drop the subscript θ most of the time and write P· instead of P·,θ.

Write

99

z − P· = (1 + K·)(z − P·), where K· = K·(z − P·)−1, (8.40)

so that

(z − P·)−1 = (z − P·)−1(1 + K·)−1. (8.41)

Using the resolvent identity,

(z − P·)−1 − (z − P·)−1 = −(z − P·)−1K·(z − P·)−1,

we can decompose the right hand side of (8.26) as I + II, where

I = ′′tr [1

2πi

∫γg(z)(P· − z0)−m(z − P·)−1dz]10

′′, (8.42)

II = −′′tr [1

2πi

∫γg(z)(P· − z0)−m(z − P·)−1K·(z − P·)−1dz]10

′′. (8.43)

Let γ be a smooth contour from b to a (as was the case with γ) andcontained in Ω \ W . The integrand in I is holomorphic in Ω, so we mayreplace γ by γ in (8.42). Along γ, we have g = O(1) (uniformly w.r.t. h) andif we choose a cutoff χ in Lemma 8.3 equal to 1 on a sufficiently large bounded(sub)set (of Γθ), we can apply the proof of Proposition 8.1, to obtain:

′′tr [(P· − z0)−m(z − P·)−1]10′′ = O(1)Φmax(

1

h2). (8.44)

We therefore get

I = O(1)Φmax(1

h2). (8.45)

Since K· are of finite rank, we can write

II = −[tr1

2πi

∫γg(z)(P· − z0)−m(z − P·)−1K·(z − P )−1dz]10. (8.46)

Here we can replace (P·− z0)−m by (z− z0)−m because the difference has anintegrand which is holomorphic in Ω, and the contour γ can then be replacedby γ. It follows that the difference gives a contribution O(1)Φmax( 1

h2 ) to II.Recalling (8.18), we get:

II = −[1

2πi

∫γf(z)tr ((z − P·)−1K·(z − P·)−1)dz]10 +O(1)Φmax(

1

h2). (8.47)

100

Now use (8.41) and the cyclicity of the trace:

−tr ((z − P·)−1K·(z − P·)−1) = −tr ((z − P·)−1(1 + K·)−1K·(z − P·)−1)

= −tr ((1 + K·)−1K·(z − P·)−2) = tr ((1 + K·(z))−1 ∂

∂zK·(z)).

(8.48)

Remark. In the last equality in (8.48) we used that K· is independent of z. Itmay be of interest to see what would happen if K· = K·(z) (and P· = P·(z))depends holomorphically on z. Differentiating the first identity in (8.40), weget

−K·(z − P·)−1 =∂K·∂z− (1 + K·)

∂P·∂z

(z − P·)−1. (8.49)

Using this in the third expression in (8.48), we get

−tr ((z − P·)−1K·(z − P·)−1) = (8.50)

tr ((1 + K·)−1∂K·

∂z)− tr ((1 + K·)

−1(1 + K·)∂P·∂z

(z − P·)−1)

= tr ((1 + K·)−1∂K·

∂z)− tr(

∂P·∂z

(z − P·)−1).

Here the last term is holomorphic in Ω and we get (8.51) below also in thismore general case.

Substitution of (8.48) into (8.47) gives

II = [1

2πi

∫γf(z)tr ((1 + K·(z))−1∂K·

∂z)dz]10 +O(1)Φmax(

1

h2). (8.51)

With (8.40) in mind we consider an identity of the form

A(z) = B(z)C(z), (8.52)

where C(z) : H1 → H2, B(z) : H2 → H3 depend holomorphically on zin some domain, and ∂zB(z) is of trace class. Restricting the attentionto some subdomain, where B and C have bounded inverses, we get afterdifferentiating the equation (8.52) and using the cyclicity of the trace:

tr (B−1∂zB) = tr (A−1∂zA− C−1∂zC). (8.53)

101

Let Γ be a relatively compact subdomain with smooth boundary and assumethat C−1 exists everywhere on the closure of Γ. Also assume that B(z)−1

exists everywhere on the boundary of Γ. Then for f holomorphic in Γ andcontinuous up the boundary, we get

tr (1

2πi

∫∂Γf(z)B(z)−1∂B(z)dz) = (8.54)

tr1

2πi

∫∂Γf(z)(A−1∂zA− C−1∂zC)dz = tr (

1

2πi

∫∂Γf(z)A−1∂zAdz).

We apply this to (8.40) with Γ ⊂⊂ Ω and ∂Γ avoiding the resonances, andget:

tr1

2πi

∫∂Γf(z)(1 + K·(z))−1∂zK·dz =

∑λ∈Res (P·)∩Γ

f(λ). (8.55)

Notice that we could have taken the trace inside the integral, that tr ((1 +K·))

−1∂zK·) = ∂z log det(1+ K·(z)) and that the resonances are precisely thezeros of det(1 + K·(z)) and that the multiplicities agree.

The remainder of the proof will have common features with a proof ofLidskii’s theorem, close to the one in [31], of which we gave a variant inChapter 5. Put

D·(z;h) := det(1 + K·) = O(1)eO(1)Φ·(h−2). (8.56)

Since we will work separately with the two operators P·, we drop the subscript· most of the time in the following. In Ω+,δ := z ∈ Ω+; Im z > δ with δ > 0(suitably chosen and independent of h), not only (z−P ) : D → H is uniformlyinvertible, but also (z − P ), and hence

(1 + K)−1 = (z − P )(z − P )−1 = O(1) : H → H.

It follows that (1 + K)−1 = 1 + (P − P )(z − P )−1 with

‖(P − P )(z − P )−1‖tr = O(1)Φ(1

h2),

giving an upper bound for the determinant of (1 + K)−1 as in (8.56):

|D(z;h)| ≥ e−CΦ(h−2), z ∈ Ω+,δ. (8.57)

102

Let N = N(P,Ω;h) be the number of resonances in Ω counted with theirmultiplicity. (8.56) remains valid in a slightly larger domain and combiningthis with (8.57) and Jensen’s inequality, we get

N(P,Ω;h) ≤ CΦ(1

h2). (8.58)

At this point, we could apply an estimate of H. Cartan about lower boundsof holomorphic functions outside small unions of circles around the zeros asin [81], but we prefer to keep the argument of [75, 76], which is simple, moreor less standard and gives precisely what is needed.

Let zj, j = 1, .., N be the resonances in Ω repeated according to their multi-plicity and put

Dw(z;h) =N∏j=1

(z − zj). (8.59)

Thanks to (8.58), we have

|Dw(z;h)| ≤ eCΦ(h−2) in Ω, |Dw(z;h)| ≥ e−CΦ(h−2) in Ω+,δ. (8.60)

In Ω \ Ω+,δ we can get the same lower bound, if we avoid to go too closeto the zj. For that, we first establish the simple

Lemma 8.5 Let x1, .., xN ∈ R and let I ⊂ R be an interval of length |I| ∈]0,+∞[. Then there exists x ∈ I, such that

N∏j=1

|x− xj| ≥ e−N(1+log 2|I| ).

Proof. Consider F (x) =N∑j=1

log 1|x−xj | . We have

∫I

log1

|x− xj|dx ≤ 2

∫ |I|/20

log1

tdt = |I|(1 + log

2

|I|),

since the first integral takes its largest possible value when xj is the midpointof I. It follows that ∫

IF (x)dx ≤ N |I|(1 + log

2

|I|).

103

We can therefore find x ∈ I, such that F (x) ≤ N log(1 + log 2|I|), i.e.

N∏j=1

|x− xj| = e−F (x) ≥ e−N(1+log 2|I| ).

#

Let γ = γα, α ∈ J be a smooth family of smooth curves joining b to a,where J is a bounded interval of length 6= 0. We let γα move transversallyin Ω+,δ/2 when α varies in such a way that for every z ∈ Ω− we have:

1) If z belongs to γα for some α, then this α is unique, α = α(z) anddist (z, γβ) ≥ |β − α(h)| for all β ∈ J .

2) If z belongs to no γβ, then dist (z, γβ) ≥ dist (β, ∂J) for every β ∈ J .

It follows from Lemma 8.5, that if I ⊂ J is an interval of length > 0, thenwe can find α ∈ I such that

|Dw(z;h)| ≥ e−CΦ(h−2), z ∈ γα, (8.61)

where C = C|I|. We factorize D:

D(z;h) = G(z;h)Dw(z;h) z ∈ Ω, (8.62)

where G and 1/G are holomorphic in Ω. Combining (8.60), (8.61), (8.62),we get

|G(z;h)| ≤ eCΦ(h−2), z ∈ Ω+,δ ∪ γα. (8.63)

The maximum principle gives

|G(z;h)| ≤ eCΦ(h−2), z ∈ Ω, (8.64)

where Ω ⊂⊂ Ω is any simply connected relatively open h-independent setwith W in its (relative) interior. In fact, it suffices to choose the family γαwith γα ∩ Ω ∩ Ω− = ∅.

The relations (8.60, (8.57) imply that

|G(z;h)| ≥ e−CΦ(h−2), z ∈ Ω+,δ. (8.65)

Choose C > 0 large enough, so that

`(z;h) := CΦ(1

h2)− log |G(z;h)| ≥ 0. (8.66)

104

Notice that ` is harmonic. Harnack’s inequality tells us that for every K ⊂⊂Ω, there is a constant CK ≥ 1 such that for every non-negative harmonicfunction ˜ on Ω :

supK

˜≤ CK infK

˜, (8.67)

This applies to ` and after an arbitrarily small decrease of Ω, we have (8.67)with K = Ω, ˜= ` and if we use (8.65), we get `(z;h) ≤ CΦ( 1

h2 ) on Ω+,δ and

hence by (8.67) (with K = Ω):

`(z;h) ≤ CΦ(1

h2) on Ω. (8.68)

We conclude that log |G| ≥ −CΦ( 1h2 ) on Ω and with (8.64) we get

| log |G(z;h)|| ≤ CΦ(1

h2), z ∈ Ω. (8.69)

Since log |G(z;h)| = Re logG is harmonic, we get after an arbitrarily smalldecrease of Ω:

∇(Re logG) = O(1)Φ(1

h2), z ∈ Ω. (8.70)

The Cauchy Riemann equations for logG imply

d

dzlogG = O(1)Φ(

1

h2), z ∈ Ω. (8.71)

Let γ be a curve from b to a in Ω \W and which avoids the resonances.From (8.58), (8.59) we get∫

γf(z;h)

d

dzlogDw(z;h)dz = O(1)Φ(

1

h2). (8.72)

Sinced

dzlogD =

d

dzlogG+

d

dzlogDw,

we get from (8.71), (8.72):∫γf(z;h)

d

dzlogDdz = O(1)Φ(

1

h2). (8.73)

105

We finally return to (8.51), where we recall (8.55) and the subsequent identity.Together with (8.73), this gives:

II = (8.74)[ ∑Res(P·)3z

between γ and γ

f(z;h) +1

2πi

∫γf(z;h)

d

dz(logD·(z))dz

]10

+O(1)Φmax(1

h2)

=[ ∑

Res(P·)3zbetween γ and γ

f(z;h)]1

0+O(1)Φmax(

1

h2).

Here we reintroduce the subscripts “.” and recall that Φmax = max(Φ0,Φ1).

Since the total number of zj’s is O(1)Φ( 1h2 ) and |f | ≤ 1 for z ∈ Ω \W , we

may replace the last sum in (8.74) by∑z∈W−∩Res(P·)

f(z;h)

and the proof of Theorem 8.2 is complete. #

Remark. Applying Lemma 8.3 or rather the equivalent estimate on∫I F (x)dx

in its proof, we see that if we take a family of contours γα, α ∈ J , |J | > 0,as earlier in the proof, then (for every h) we can find α ∈ J , such that (8.73)strengthens to ∫

γα| ddz

logD||dz| = O(1)Φ(1

h2). (8.75)

In fact, in view of (8.71) it suffices to establish this with D replaced by Dw.

Appendix: construction of P·,θ.

We shall construct an operator P·,θ, which satisfies (8.37)–(8.39) and forthat we shall use the function F which was introduced prior to those equa-tions. Since we shall work with the two operators P· separately we drop thesubscript ·. Let pθ be the semi-classical principal symbol of Pθ, definedas

∑|α|≤2 bα(x;h)ξα modO(h〈ξ〉/〈x〉), where Pθ =

∑|α|≤2 bα(x;h)(hDx)

α.Choosing Γθ suitably (more precisely with R1 sufficiently large and ε0 suffi-ciently small in the construction of fθ in section 3), we see that F pθ is awell-defined symbol with values away from Ω, such that F pθ = pθ, when|x| ≥ R2, R2 > R1. Here we recall that Γθ coincides with Rn for |x| ≤ R1.

106

Let P ] be a reference operator as in section 2 and assume that the corre-sponding torus T is large enough so that B(0, R1) is naturally contained inT . If f = F|R, then

f(P ]) = P ] + g(P ]) = P ] +K], (A.1)

where f(E) = E + g(E), g ∈ C∞0 (R), and

K] = O(1) : H] → H], rankK] = O(1)Φ(1

h2). (A.2)

More generally K] is uniformly O(1) as a bounded operator from the domainof any power of P ] to the domain of any other power of P ]. Moreover

(z − f(P ]))−1 = O(1), (A.3)

uniformly for z ∈ Ω and in h. Let Q] be a self-adjoint differential operatoras in the construction of P ], so that Q] = P ] in T \B(0, R0). As in the proofof Lemma 8.3, we see that if χ ∈ C∞0 (T \B(0, R0)), then

χ(z − P ])−1χ− χ(z −Q])−1χ is O(hN

|Im z|M(N)) : H−N → HN , ∀N ∈ N.

(A.4)Using the Cauchy formula (7.21) for g(P ]), g(Q]), we see that

χf(P ])χ− χf(Q])χ = O(hN) : H−N → HN , ∀N ∈ N (A.5)

for χ as in (A.4).

Using a result of Helffer-Robert, which can also be proved with the Cauchyformula (7.21) (see [21] and further references there), we know that χf(Q])χis a h-pseudo-differential operator on T of the natural class, with leadingsymbol χ(x)2f(p](x, ξ)), where p] is the semiclassical leading symbol of Q].

Let 1 = χ0 + χ1 + χ2, where χ0 ∈ C∞0 (B(0, R1); [0, 1]) is equal to 1near B(0, R0), χ1 ∈ C∞0 (Γθ; [0, 1]), χ0 + χ1 = 1 near B(0, R2). Then χ2 ∈C∞b (Γθ; [0, 1]) has its support disjoint from B(0, R2). Let χj ≺ χj, whereχj ∈ C∞(Γθ; [0, 1]) has its support close to that of χj, and define

Pθ := χ0f(P ])χ0 + χ1RFχ1 + χ2Pθχ2, (A.6)

where RF is an h-pseudodifferential operator with leading symbol F (pθ) suchthat the total (Weyl) symbol (with Γθ identified with Rn) of RF − Pθ (for xnear suppχ1) has compact support in ξ.

107

Lemma 8.6 If h > 0 is small enough, then z − Pθ is invertible for everyz ∈ Ω and, for every N ∈ N, the inverse satisfies

(z − Pθ)−1 = ON(1) : DN → DN+1, (A.7)

uniformly for z ∈ Ω, where DN is the domain of PN , so that D0 = H.

Proof. The operator z − Pθ : D → H is Fredholm of index 0, for z ∈ Ω,so to get invertibility and the estimate (A.7) in the case N = 0, it is enoughto show the a priori estimate:

‖u‖2 ≤ C‖(z − Pθ)u‖2, (A.8)

for u ∈ D. Let ψ0, ψ1, ψ2 ∈ C∞b (Cθ; [0, 1]) have the same support propertiesas χ0, χ1, χ2 and with

1 = ψ20 + ψ2

1 + ψ22. (A.9)

Assume that ψ0 ≺ χ0. Then for some r1 > 0 and for all z ∈ Ω

‖(z − Pθ)ψ0u‖2 = ‖(z − f(P ]))ψ0u‖2 +O(h∞)‖u‖2,

and for j = 1, 2:‖(z − Pθ)ψju‖2 ≥ r2

1‖ψju‖2.

Then

‖(z − Pθ)u‖2 =2∑0

‖ψj(z − Pθ)u‖2 ≥2∑0

(‖(z − Pθ)ψju‖ − ‖[ψj, Pθ]u‖)2

≥ r21‖u‖2 −O(h)‖u‖2

D.

But‖u‖Dθ ≤ O(1)(‖(z − Pθ)u‖+ ‖u‖),

so we get (A.8) and the stronger a priori estimate

‖u‖D ≤ C‖(z − Pθ)u‖, (A.10)

which is precisely (A.7) for N = 0.

If (z − Pθ)u = v, with u, v ∈ D, then

(z0 − Pθ)u = v + ((z0 − z) + (Pθ − Pθ))u ∈ D, (A.11)

108

and applying (z− Pθ) and using (A.10), we see that ‖(z0−Pθ)u‖D ≤ C‖v‖D.This gives the case N = 1 of (A.7) and we can continue this argument byiteration. #

WritePθ = Pθ + χ0g(P ])χ0 + χ1(RF − Pθ)χ1.

We can find TF of finite rank O(h−n) and = O(1) : Hs1 → Hs2 for alls1, s2 ∈ R such that

χ1((RF − Pθ)− TF )χ1 = O(h∞) : Hs1 → Hs2

for all s1, s2 ∈ R. Put

Pθ = Pθ + χ0g(P ])χ0 + χ1TFχ1. (A.12)

Then it easy to see that (8.37)-(8.39) hold.

9 Clouds of resonances for −h2∆ + V (x)

The applications that we develop in this chapter exploit very much the semi-classical nature of the operator. In chapter 10 we shall get other applications,also for operators that are not semi-classical to start with, and which exploitthe existence of closed classical trajectories. The proof of the results in thischapter will also use a semi-classical trace formula of D. Robert, and we startby discussing certain distributions on the real axis that appear in connectionwith that formula.

Let V0, V1 be continuous real-valued functions on Rn which tend to 0,when |x| → ∞, and satisfy

|V1(x)− V0(x)| ≤ C〈x〉−n, (9.1)

where n > n.For E > 0, let ν+,·(E) =

∫V·(x)≥E dx. This is a decreasing function of

E, so µ+,·(E) := − ddEν+,·(E) is a positive measure on ]0,+∞[ with support

equal to ]0, supV·]. Similarly, for E < 0, we put ν−,·(E) =∫V·(x)≤E dx (an

increasing function of E,) µ−,·(E) = ddEν−,·(E) which is a positive measure

on ]−∞, 0[ with support in [inf V·, 0[.

109

For φ ∈ C∞0 (R), we put

〈µ, φ〉 =∫

[φ V·]10dx =∫

(φ(V1(x))− φ(V0(x)))dx

Clearly µ is a distribution of order ≤ 1 with support contained in [min inf V·,max supV·],

∫µ(E)dE = 0, and

µ|R± = [µ±,·]10. (9.2)

For f ∈ C∞0 (R), we put

〈ω, f〉 =∫∫

[f(ξ2 + V·(x))]10dxdξ,

so that ω is a distribution on R of order ≤ 1 with

suppω ⊂ [min inf V·,+∞[.

We have with E+ := max(E, 0) and ∗ denoting convolution:∫f(ξ2 + V·(x))dξ =

∫ ∞0

f(E + V·(x))d(Vol ξ ∈ Rn; |ξ| ≤√E)

= Vol (BRn(0, 1))∫f(V·(x) + E)d(E

n/2+ )

=n

2Vol (BRn(0, 1))

∫f(V·(x) + E)E

n2−1

+ dE

= Cn(f ∗ (−·)n2−1

+ )(V·(x)),

where Cn > 0. Then

〈ω, f〉 = Cn〈µ, f ∗ (−·)n2−1

+ 〉 = Cn〈µ ∗ (·)n2−1

+ , f〉,

soω = Cnµ ∗ (·)

n2−1

+ . (9.3)

In general, if α > −1 and if Hα(x) = xα+, then Hα(ξ) = Cα(ξ − i0)−1−α,where Cα 6= 0. Let Eα ∈ S ′(R) be the inverse Fourier transform of C−1

α (ξ −i0)1+α. Then suppEα ⊂ [0,+∞[, Hα ∗ Eα = δ. When 1 + α ∈ N, then Eαis a constant times some derivative of the delta function, and in general, wehave

Eα|]0,+∞[(x) = Cαx

−2−α.

110

We can now invert (9.3) and get

µ =1

CnEn

2−1 ∗ ω. (9.4)

If near some point t0 ∈ R, we know that µ extends to a holomorphic function,then by contour deformation in (9.3), we see that ω will have the sameproperty. (It is easy to see how to get such a bounded extension by contourdeformation, and an easy way to see that this extension is also holomorphic,is to regularize ω and En

2−1 by convolving with Gaussians, and then let the

Gaussians converge to the delta-function.) The converse implication holdsby (9.4). It follows that µ and ω have the same analytic singular support.Since µ is a real distribution we know that its analytic wavefront set WFa(µ)is of the form (x, ξ) ∈ R×(R\0);x ∈ sing supp a(µ). The same holds forω and since these two distributions have the same analytic singular support,we conclude that

WFa(µ) = WFa(ω). (9.5)

We refer to [73] for basic properties of the analytic wavefront set WFa.We now discuss resonances close to analytic singularities of µ on ]0,+∞[.

Let P· = −h2∆ + V·(x), where V· ∈ C∞(Rn; R), · = 0, 1, and assume thatthe general assumptions in Chapter 8 are verified, with Φ·(

1h2 ) = h−n. Then

we can define µ as above.

Theorem 9.1 Let 0 < E0 ∈ sing supp a(µ). Then for every complex neigh-borhood W of E0, there exist h0 = h0(W ) > 0, and C = C(W ) > 0, suchthat for h < h ≤ h0, we have

∑10 ](Res (P·) ∩W ) ≥ 1

C(W )h−n.

The following corollary is stronger than the theorem:

Corollary 9.2 Let P1 = −h2∆ + V1(x), V1 ∈ C∞(Rn; R) satisfy all theassumptions of the local trace formula (Th. 8.2) that make sense for onesingle operator. Let 0 < E0 ∈ sing supp a(ν+,1). Then for every complexneighborhood W of E0, there exist h0 = h0(W ) > 0, and C = C(W ) > 0,such that for h < h ≤ h0, we have ](Res (P1) ∩W ) ≥ 1

C(W )h−n.

Proof of the Corollary. It suffices to construct P0 = −h2∆ + V0(x), withV0 ∈ C∞b (Rn; R) such that (P1, P0) satisfies the assumptions of the localtrace formula and such that:

sup supp ν+,0 < E0, (9.6)

111

∃ a complex neighborhood W0 of E0 such that (9.7)

ResP0 ∩W0 = ∅, when h > 0 is small enough.

We shall produce V0 from V1 by cut-off and regularization. Put K(x) =Cne

−x2/2 with Cn > 0 chosen so that∫Rn K(x)dx = 1. Put Kλ(x) =

λ−nK(λ−1x), for λ > 0. We make an x-dependent choice of λ:

λ(R, x) = R〈R−1x〉−N0 ,

where N0 > 0 is sufficiently large, depending on the dimension and R ≥ 1 is alarge parameter. Put KR(x, y) = Kλ(R,x)(x− y). Let χ ∈ C∞0 (B(0, 2); [0, 1])be equal to 1 on B(0, 1). For R large enough, put

V0(x) =∫KR(x, y)(1− χ(R−1y))V1(y)dy.

Let θ > 0 be small but independent of R. Then in the domain |Imx| <θ〈Rex〉:

|V0(x)| ≤ ε(R), ε(R)→ 0, R→∞,|V1(x)− V0(x)| ≤ C(R)〈x〉−n−1.

Then we have (9.6) and using that the resonances near E0 can be viewed aseigenvalues of P0|eiθ/2Rn , we see that (9.7) also holds. #

Proof of Theorem 9.1. We have seen that WFa(µ) = WFa(ω) and sinceµ, ω are real, it is clear that (E0, 1), (E0,−1) ∈ WFa(ω). Considering thedefinition of WFa(ω) by means of the FBI-transform ([73]), we see that thereexist sequences (αj, βj)→ (E0, 1) in R2, λj → +∞, εj 0, such that

|∫eiλj(βj(αj−E)+ i

2(αj−E)2)χ(E)ω(E)dE| ≥ e−εjλj , (9.8)

where χ ∈ C∞0 (]0,∞[) is equal to 1 near E0 and has its support in a smallneighborhood of E0. Let a, b, a/b be small and positive and let

Ω =]E0 − b, E0 + b[+i]− a, a], W =]E0 −b

2, E0 +

b

2[+i]− a

2, a].

Let I, J be the intersections of W , Ω with the real line and choose χ ∈ C∞0 (J)equal to 1 near the closure of I. Let

fj(E) = eiλj(βj(αj−E)+ i2

(αj−E)2).

112

Then|fj |Ω\W | ≤ e

− 1C0λj ,

and (9.8) reads

|∫

(χfj)(E)ω(E)dE| ≥ e−εjλj . (9.9)

The trace formula gives

′′tr [(χfj)(P·)]10′′ = [

∑Res (P·)∩W−

fj(λ)]10 +O(h−n)e− 1C0λj . (9.10)

On the other hand we have a trace formula of D. Robert [67], which in oursituation implies:

tr [(χfj)(P·)]10 =

1

(2πh)n

∫(χfj)(E)ω(E)dE +Oj(h1−n). (9.11)

(See the appendix of this chapter for an outline of a proof.)Combining (9.10), (9.11), we get:

[∑

Res (P·)∩W−

fj(λ)]10 =1

(2πh)n

∫(χfj)(E)ω(E)dE +O(h−n)e

− 1C0λj +Oj(h1−n),

so by (9.9) :

|[∑

Res (P·)∩W−

fj(λ)]10| ≥e−εjλj −O(1)e

− 1C0λj

(2πh)n+Oj(h1−n).

Choose first j large enough to see that the previous expression is

≥12e−εjλj

(2πh)n+Oj(h1−n).

Then fix j and choose h small enough to see that it can be bounded frombelow by (2πh)−ne−εjλj/3. The theorem follows. #

We next consider resonances generated by analytic singularities on thenegative axis.

113

Theorem 9.3 We make the same general assumptions as in Theorem 9.1and assume that the angle of scaling θ0 is > π/2. Let 0 > E1 ∈ sing supp a(µ).Let γ : [0, 1] → C be a C1 curve with γ(0) ∈] sup supp (µ),+∞[, γ(1) = E1,Im γ(t) < 0 for 0 < t < 1. Also assume that γ is injective and γ′(t) 6= 0, ∀t.Then for every neighborhood W of γ([0, 1]), there exist constants C, h0 > 0,such that

1∑0

#(Res (P·) ∩W ) ≥ 1

Ch−n, 0 < h ≤ h0(W ).

Corollary 9.4 Let P1 = −h2∆ + V1(x) satisfy all the assumptions of thelocal trace formula that make sense for one of the operators. Let the angle ofscaling satisfy θ0 > π/2, and assume that V1(x) = O(〈x〉−n), for some n > nin the region appearing in the corresponding complex scaling assumption. Let0 > E1 ∈ sing supp a(ν−,1), and let γ : [0, 1] → C be a C1 curve with γ(0) ∈]sup supp (ν+,1),+∞[, γ(1) = E1, Im γ(t) < 0 for 0 < t < 1, γ injective andγ′(t) 6= 0 everywhere. Then for every neighborhood W of γ([0, 1]), there existC, h0 > 0 such that

#(Res (P1) ∩W ) ≥ 1

Ch−n, 0 < h ≤ h0(W ).

To get the corollary, it suffices to apply Theorem 9.3 with P0 = −h2∆. Ifwe combine the two corollaries, we get the following general statement:

Consider P1 = −h2∆ + V1(x) with angle of scaling θ0 > π2

and with

V1 = O(〈x〉−n), n > n in the corresponding scaling region. If V1 6≡ 0, then#(P1 ∩W ) ≥ 1

C(W )h−n, 0 < h ≤ h0(W ), where W is either any fixed complex

neighborhood of some point on ]0,+∞[ or any neighborhood of a curve as inCorollary 9.4.

We refer to [77] for the proof of Theorem 9.3, and we here only explainsome of the ideas. Let γ, W be as in the theorem and assume that W is openin z ∈ C \ 0; −2θ0 < arg z ≤ ε0 for some small ε0 > 0 (as in the localtrace formula). We may assume that the intersection of W with the real lineis the union of two intervals I± ⊂ R±, where R± denote the open half-axes.Let Ω ⊃⊃ W have the same properties as W and let J± be the correspondingintervals. Let χ ∈ C∞0 (Ω) be equal to 1 on W , almost holomorphic at J±.For a suitable sequence f = fj (that will not be constructed in detail here),with |f | ≤ 1 on Ω \W , we write the trace formula:

[∑

λ∈Res∩W−fj(λ)]10 = ′′tr [χ+fj(P·)]

10′′ + ′′tr [χ−fj(P·)]

10′′ +O(h−n).

114

Using the semiclassical trace formula, we get:

′′tr [χ±fj(P·)]10′′ =

1

(2πh)n

∫(χ±fj)(E)ω(E)dE +Oj(h1−n).

ω(E) extends analytically from J+ to a function ω+ on Ω ∩ Im z < 0 andby Stokes’ formula, we get∫

χ+fj(E)dE = −∫

(χ−fj)(E)ω+(E − i0)dE +O(1).

It follows that

[∑

λ∈ResP·∩W−fj(λ)]10 = (9.12)

1

(2πh)n

∫(χ−fj)(E)(ω(E)− ω+(E − i0))dE +O(h−n) +Oj(h1−n).

Here ω(E)− ω+(E − i0) can be obtained from µ by convolution and wecan show that (E1,±1) both belong to the analytic wavefront set of ω(E)−ω+(E − i0).

Approximating Gaussian functions by suitable functions fj as above, wecan arrange so that the first term in the right hand side of (9.12) is O(h−n)which permits us to conclude. #

We shall next consider clouds of resonances for systems, following the thearticle of L. Nedelec [59]. Schrodinger operators with matrix-valued poten-tials and similar systems appear frequently in quantum mechanics, for in-stance when one reduces the dimension by means of the Born-Oppenheimerreduction and similar methods that lead to a so called effective Hamiltonian.Although these effective Hamiltonians are not in general exactly equal toSchrodinger operators with matrix-valued potential, the latter form impor-tant models for more general systems that could also be studied with moreor less the same methods.

LetP = −h2∆ + V (x), (9.13)

with V (x) ∈ C∞b (Rn; Hermitian r × rmatrices ), satifying

(H1) ∃θ0 ∈ [0, π2[, ε > 0, R > 0, such that V extends to a holomorphic

function on rω; ω ∈ Cn, dist (ω, Sn−1) < ε, r ∈ ei[0,θ0]]R,+∞[.

115

(H2) There exists a Hermitian r × r-matrix V0 and a number n > n such

that ‖V (x)− V0‖ ≤ C〈x〉−n in the set in (H1).

Put P0 = −h2∆ + V0, and let λ1,0 ≤, .. ≤ λr,0 be the eigenvalues of V0,x ∈ Rn. As in the scalar case, we can then define the resonances of P asthe eigenvalues of the scaled operator Pθ = P|Γθ , 0 < θ ≤ θ0 in the set

R \∪rj=1(λj,0 + e−2iθ]0,+∞[), and the resonances are independent of θ in thesence that for 0 < θ1 < θ2 ≤ θ0, θ1 ≤ θ ≤ θ2, and for every 1 ≤ j ≤ r − 1,the eigenvalues of Pθ in the triangle bounded by the interval ]λj,0, λj+1,0[ andthe half-rays λj,0 + e−2iθ1 ]0,+∞[, λj+1,0 + e−2iθ2 ]0,+∞[ are independent of θ.A similar statement holds in the sector λr,0 + e−2i[0,θ0[]0,+∞[.

Assume λj,0 < λj+1,0 for some 1 ≤ j ≤ r and let W ⊂⊂ Ω be open andrelatively compact in

(λj,0 + ei]−2θ0,ε0]]0,∞[) ∩ (]λj,0, λj+1,0[+e−2iθ0R).

Here ε0 > 0 is small and we use the convention that λr+1,0 = +∞. Assumethat I := W ∩R, J := Ω ∩R are intervals.

The local trace formula still holds:

Theorem 9.5 Let f = f(z;h) be holomorphic for z ∈ Ω with |f(z;h)| ≤ 1,z ∈ Ω \W . Let 1I ≺ χ ∈ C∞0 (J) be independent of h. Let P1 := P . Then[χf(P·)]

10 is of trace class and

tr [χf(P·)]∞0 =

∑z∈Res (P )∩W−

f(z;h) +O(h−n),

where W− = W ∩ e−2i[0,θ0[]0,+∞[

The proof (see [59]) is essentially the same as in the scalar case.On the other hand, we have Robert’s trace formula:

Theorem 9.6 Let g ∈ C∞0 (R) be independent of h. Then [g(P·)]10 is of trace

class and

tr [g(P·)]10 =

1

(2πh)n

∫ ∫tr (g(p(x, ξ))− g(p0(x, ξ)))dxdξ +O(h1−n),

where (x, ξ) = ξ2 + V (x), p0(x, ξ) = ξ2 + V0(x).

116

Here the integral can also be written∑j

∫ ∫(g(ξ2 + λj(x))− g(ξ2 + λj,0(x)))dxdξ =: 〈ω, g〉,

where λ1(x) ≤ ... ≤ λr(x) denote the eigenvalues of V (x). As in the scalarcase,

ω = Cn(E)n2−1

+ ∗ µ, where 〈µ, φ〉 =∑j

∫(φ(λj(x))− φ(λj,0))dx.

For E ∈]λj,0, λj+1,0[ we put

νj(E) = −j∑

k=1

∫λk(x)≥E

dx+r∑

k=j+1

∫λk(x)≤E

dx

Then on this interval, we have

µ =d

dEνj, sing suppa(ω) = sing suppa(µ) = sing suppa(νj).

The arguments in the scalar case extend (see [59]) and give

Theorem 9.7 Let ]λj,0, λj+1,0[3 E0 ∈ sing suppa(νj). Then for any neigh-borhood W of E0, there exist h0, C > 0, such that

#(Res (P ) ∩W ) ≥ 1

Ch−n, 0 < h ≤ h0.

If λk(x) is a simple eigenvalue near some point x0, then λk(x) is a smoothfunction there, and if E0 = λk(x0) is a critical value, we expect in generalthat E0 will belong to the analytic singular support of µ. This is essentiallyas in the scalar case. Multiple eigenvalues can also give rise to analyticsingularities of µ and we now describe such a situation:

Let E0 ∈]λj,0, λj+1,0[, 1 ≤ j ≤ r, with the convention λr+1,0 = +∞, andassume:

0) n = 2, V is real and symmetric.

1) For x = 0, there are exactly two eigenvalues λk(0), λk+1(0) that are equalto E0.

2) If E0 = λν(x) for some ν and some x 6= 0, then λν(x) is a simple eigenvalue,and dλν(x) 6= 0.

117

3) Let a(x)I +(c(x) b(x)b(x) −c(x)

)be a smooth representative of the restriction

of V (x) to the 2-dimensional eigenspace corresponding to the two eigenvaluesλk(x), λk+1(x), for x in a neighborhood of 0. Then dc(0), db(0) are linearlyindependent and da(0) = βdb(0) + γdc(0) with β2 + γ2 < 1. Under theseassumtions it is proved in [59] that E0 ∈ Sing suppaµ.

Appendix.

We first recall the operator version of the Cauchy-Riemann-Green-Stokesformula, which was used (in a slightly different operator version) by Dynkin[22] and has been used very much since the work [39]. Let P : H → H besome unbounded selfadjoint operator, and let f ∈ C∞0 (R). Let f ∈ C∞0 (C)be an almost holomorphic extension of f with the property that ∂f vanishesto infinite order on the real axis). Then since 1/(πz) is a fundamental solutionof the Cauchy-Riemann operator ∂, we have

f(E) = − 1

π

∫∂f(z)

1

z − EL(dz), E ∈ R. (A.1))

The operator version follows by replacing “E” by “P”:

f(P ) = − 1

π

∫∂f(z)(z − P )−1L(dz), E ∈ R. (A.2))

The proof is a straight forward application of the spectral theorem. ([39],[21], [20]).

Let P· be as in the main text of this chapter. Let χ ∈ C∞0 (R) be conve-niently chosen, so that the infimum of the spectrum of P. := −h2∆ +V.(x) +χ(hD) is ≥ 1 + sup supp f . The resolvent identity gives

(z − P.)−1 = (z − P.)−1 + (z − P.)−1χ(hD)(z − P·)−1. (A.3)

Using this in (A.2) we get, since f(P·) = 0:

f(P·) = − 1

π

∫∂f(z)(z − P.)−1χ(hD)(z − P·)−1L(dz). (A.4)

As for instance in the proof of the invariance of the condition (7.12), we seethat [(z − P·)−1χ(hD)(z − P.)−1]10 is of trace class with corresponding norm≤ O(1)h−n|Im z|−N0 , for z ∈ neigh supp (f). A convenient way to see this, isto put Pt = (1− t)P0 + tP1, and to write

[(z− P·)−1χ(hD)(z−P·)−1]10 =∫ 1

0

∂t((z− Pt)−1χ(hD)(z−Pt)−1)dt, (A.5)

118

then develop the right hand side and use the same arguments as in Chapter7.

It follows that for any fixed δ > 0:

tr [f(P.)]10 = − 1

πtr∫|Im z|>hδ

∂f(z)[(z−P.)−1χ(hD)(z−P·)−1]10L(dz)+O(h∞).

(A.6)Let 0 < δ < 1

2. Then (z − Pt)

−1 is an h-pseudodifferential operator withsymbol a of class Sδ(h

−δ〈ξ〉−2) for |Im z| > hδ, in the sense that

∂αx∂βξ a = O(h−δ(|α|+|β|)h−δ〈ξ〉−2).

(The detailed proof of this uses a semi-classical version of a lemma of R.Beals, see [39] and also [21].) In this class, the symbol has an asymptoticexpansion:

∼ 1

(z − pt(x, ξ))+ha1(x, ξ, z)

(z − pt)3+h2a2(x, ξ, z)

(z − pt)5+ ..., .

[[Nasta version, skriv lite mer om egenskaperna hos aj.]] (See [21].) Usingthis in (A.5), we see that [(z − P·)−1χ(hD)(z − P·)−1]10 has a symbol of class

Sδ(h−2δ〈ξ〉−N〈x〉−n) for every n > 0. The symbol is equal to

[(z − p·)−1χ(ξ)(z − p·)−1]10 +∫O(h)〈x〉−n〈ξ〉−N |z − pt|−N0dt. (A.7)

We plug this into (A.6) and use that an h-pseudodifferential operator withsymbol a in Sδ(m) is of trace class if m is integrable, and has the trace

1(2πh)n

∫ ∫a(x, ξ)dxdξ. It follows that

tr [f(P·)]10 = − 1

π

∫ ∫ ∫∂f(z)[(z−p·)−1χ(ξ)(z−p·)−1]10

dxdξ

(2πh)nL(dz)+O(h1−n).

(A.8)Here

− 1

π

∫∂f(z)[(z − p·(x, ξ))−1χ(ξ)(z − p·)]10L(dx) =

− 1

π

∫∂f(z)[(z − p·)−1 − (z − p·)−1]10L(dx) =

[f(p.(x, ξ))− f(p.(x, ξ))]10 = [f(p·(x, ξ))]

10,

since f(p·(x, ξ)) = 0 by construction.

119

It follows that

tr [f(P·)]10 =

1

(2πh)n

∫ ∫[f(p·(x, ξ))]

10dxdξ +O(h1−n),

as claimed.

10 Resonances generated by closed trajecto-

ries.

In this chapter we show how closed classical trajectories can give rise tosingularities in the local trace-formula, leading to the existence of many res-onances. Here enters also some version of the so called Gutzwiller traceformula. This strategy was used by Bardos-Lebeu-Rauch [7], in order toshow that there are infinitely many resonances associated to certain types ofstrictly convex obstacles, and later by Ikawa [45] in the case of an obstaclecomposed by several strictly convex disjoint parts. Quantitative results withlower bounds on some counting function for the resonances were obtainedby Sjostrand and Zworski [84]. The lower bounds in [84] were originally ob-tained by using the exact Bardos-Guillot-Ralston trace-formula, and hencelimited to the case of odd space-dimensions, however the local trace-formulacan be used here as well ([75]), and as a consequence we avoid the restrictionon the dimension. Here we prefer to discuss an analogous recent result byJ.F. Bony [11], which treats explicitly the semi-classical situation. It is alsoa bit sharper than the corresponding result in [84].

LetP =

∑|α|≤2

aα(x;h)(hDx)α : L2(Rn)→ L2(Rn) (10.1)

satisfy the general assumptions of Chapter 7. Assume for simplicity that

aα(x;h) ∼∑j≥0

ajα(x)hj (10.2)

in C∞b and that this also holds in the complex domain (as in (7.41)). Let

p = p0(x, ξ) =∑|α|≤2

a0α(x)ξα (10.3)

120

be the semi-classical principal symbol of P . Let

Hp =n∑j=1

(∂p

∂ξj

∂xj− ∂p

∂xj

∂ξj) (10.4)

be the corresponding Hamilton field.Let E0 > 0, T0 > 0 and assume that γ : R 3 t 7→ γ(t) ∈ p−1(E0) is

a T0-periodic Hp-trajectory along which dp 6= 0. Let us recall some prop-erties of the associated Poincare map: At a point ρ0 = γ(t0), the spaceTρ0(p−1(E0))/RHp(ρ0) is a symplectic vectorspace: If σ =

∑dξj ∧ dxj is the

standard symplectic form that we also view as a bilinear form: 〈σ, t ∧ s〉 =tξ · sx − tx · sξ, t = (tx, tξ), s = (sx, sξ), then the symplectic orthogonal spaceof Tρ0(p−1(E0)) is RHp ⊂ Tρ0(p−1(E0)) and we get a natural non-degeneratealternate bilinear form on Tρ0(p−1(E0))/RHp(ρ0). A more or less equivalentversion of this, is that if Σ ⊂ p−1(E0) is a hypersurface (i.e. of dimension(2n − 1) − 1 = 2(n − 1)) passing through ρ0 transversally to Hp(ρ0), thennear ρ0, (Σ, σ|Σ) is a symplectic manifold.

The (non-linear) Poincare map is then obtained in the following way:Start at ρ ∈ Σ∩neigh (ρ0) and follow the Hp-trajectory until you hit Σ againfor a unique time t(ρ) = T0 +O(|ρ−ρ0|), at the point κ(ρ) = exp(t(ρ)Hp)(ρ).κ is then the non-linear Poincare map. (If ρ0 = γ(s0) is a different basepoint, and Σ ⊂ p−1(E0) a hypersurface transversal to Hp(ρ0), let κ be thecorresponding Poincare map. Then

κ = α−1 κ α,

where α : Σ ∩ neigh (ρ0)→ Σ ∩ neigh (ρ0) is given by α(ρ) = exp(r(ρ)Hp(ρ),with t(ρ) = s0 − t0 + O(|ρ − ρ0|), . Then it is easy to check that thenon-linear Poincare map κ : Σ ∩ neigh (ρ0) → Σ ∩ neigh (ρ0) is a canoni-cal transformation (or symplectomorphism, i.e. a map which conserves thesymplectic form) and the corresponding linearized Poincare map dκ(ρ0) :Tρ0Σ → Tρ0Σ will of course be a symplectomorphism also. Notice thatTρ0Σ ' Tρ0(p−1(E0))/(RHp(ρ0)). Since dκ(ρ0) is a real symplectomorphism,we know that its eigenvalues can be grouped into the following families:

1) The value 1 with even multiplicity,

2) The value –1 with even multiplicity.

3) The values λ and λ−1 for some λ ∈]−∞,−1[∪]1,+∞[

4) The values λ, λ for some λ with |λ| = 1, Imλ > 0.

121

5) The values λ, λ−1, λ, λ−1

, for some λ ∈ C with =λ > 0, |λ| > 1.

For each of the groups in 3)–5), the eigenvalues have the same multiplictity.Assume that the (linearized) Poincare map is non-degenerate in the sense

that 1 is not an eigenvalue of dκ(ρ0). Then by the implicit function theorem,applied to the map ρ 7→ κ(ρ)− ρ, we see that ρ0 is an isolated fixed point ofκ and this situation is stable under small perturbations of p − E0. We cantherefore vary E ∈ neigh (E0,R) and see that for each such energy, there isa unique closed trajectory γE in p−1(E) of period T (E) = T0 +O(E − E0).Moreover this trajectory depends smoothly on E in the obvious sense.

The set S = ∪E∈neigh (E0,R)R(γE) (where “R”) stands for “range of”, isa smooth symplectic 2-dimensional manifold. Write P = dκE(ρE), whereρE ∈ R(γE) and κE denote a fixed point and Poincare map at energy E.

Assume that γ0 = γE0 is (up to trivial reparametrizations) the only T0

periodic Hp-trajectory in p−1(E0). (This assumption can be weakened toallow for finitely many trajectories, obtained from γ0 by the action of somefinite symmetry group.) Notice that this assumption implies that E0 is nota critical value of p.

Under the above assumptions we have the following theorem, due to J.F.Bony [11].

Theorem 10.1 Let ε, µ > 0. Then ∃C > 0 such that for every interval[a(h), b(h)] ⊂ [E0 − 1

C, E0 + 1

C], 0 < h ≤ 1

C, we have∑

λ∈ResP∩WeT (Reλ)(Imλ)(1−µ)/h ≥ (10.5)

1

2πh

∫ b(h)

a(h)T ∗(E)| det(1− PγE)|−1/2dE −O(h1/C)

b(h)− a(h)

h,

where

W = λ ∈ C; a(h)− Ch log1

h≤ Reλ ≤ b(h) + Ch log

1

h,

−h log 1

h

T (Reλ)(n− 1 + ε

log(b(h)− a(h))

log h) ≤ Imλ ≤ 0

and T ∗(E) > 0, is the primitive period, i.e. the smallest positive period ofγE.

Corollary 10.2 Under the same assumptions, we have the same lower boundon #(Res (P ) ∩W ).

122

Outline of the proof. Let Pj be two operators as above, with [aα,j]10 =

O(1)〈x〉−n, for some n > n. We have then an extension of the local traceformula: Let

Ω =]− b, b[+i]− a, a], W =]− b

2,b

2[+i]− a

2, a], a, b > 0.

Let I, J be the intersections of W , Ω with the real line and choose χ ∈ C∞0 (J)equal to 1 near the closure of I. Put Ωδ = E0 + δΩ, Wδ = E0 + δW ,χδ(E) = χ(E−E0

δ).

Theorem 10.3 Let Ch < δ < 1/C, with C 1 sufficiently large. Letf(z;h) be holomorphic in Ωδ with |f(z;h)| ≤ 1 for z ∈ Ωδ \Wδ. Then

tr [(χδf)(P·;h)]10 = [∑

λ∈ResP

f(λ;h)]10 +O(1)δ−1h−n. (10.6)

The proof is an adaptation of the usual one. See [11]. Another basicingredient is the Gutzwiller trace formula, which in this version was obtainedby D. Robert [67] and to which J.F. Bony [11] has given a detailed proof,involving the approximation of e−itP·/h by Fourier integral operators.

We now describe that formula in the special case that we will need it. LetA > 0 be small and let g(t) ∈ C∞0 (neigh (T ([E0 − A,E0 + A]); R)) be equalto 1 near T ([E0 − A,E0 + A]). Put

fg(F ) = fg,C,E(F ) =∫e−it(F−E)/hg(t)e−(t−T (E))2C(log h−1)/2dt.

Let

f(F ) = f1(F ) =

√2π√

C log h−1e−iT (E)(F−E)/h− (F−E)2

2Ch2 log(h−1) .

Let 1[E0−2A,E0+2A] ≺ χ ∈ C∞0 (]E0 − 3A,E0 + 3A[), and put P1 = P . Itis possible to construct P0 as in the local trace-formula, so that P0 has noresonances in a fixed h-independent neighborhood of [E0−A,E0 +A], and sothat the principal symbol p0 has no closed classical trajectories with energyin [E0 − 3A,E0 + 3A] with period in some fixed neighborhood of T ([E0 −3A,E0 + 3A])

Theorem 10.4 For E ∈ [E0 − A,E0 + A] and h small enough, we have

tr [(χ2fg)(Pj)]10 =

eiS(γE)/h+i(σ(γE)π/4−S1(γE))T ∗(E)| det(1− PγE)|−12 +O(h log h−1).

123

Here S(γE) =∫γEξ · dx is the action of γE and σ(γE) and S1(γE) are real-

valued.

We can now end the outline of the proof of Theorem 10.1. Combiningthe last two theorems we get

∑λ∈Res (P )∩WE,h

|fg,E(λ)| ≥ T ∗(E)| det(1− PγE)|−1/2 −O(h log1

h).

[[Har blir mitt handskrivna manus alltfor otydligt. Skall aterkomma till dettasenare! Beviset avslutas i alla fall med att man integrerar olikheten m.a.p.E.]]

11 From quasi-modes to resonances.

We shall describe a result of P. Stefanov which culminates a series of works,started by Stefanov–Vodev [88] and continued by Tang–Zworski [89]. Actu-ally many of the basic ideas can be traced back much further to works ofAgmon, Agranovich and others in connection with the completeness of theset of generalized eigenfunctions for non-selfadjoint operators.

Let P = P (h) : H → H be a a semiclassical black box operator as inchapter 7, for which we can define the resonances in ei]−2θ0,0]]0,+∞[. Asalready mentioned in that chapter, we can also define the resonances as thepoles of the meromorphic extension from the upper half-plane across ]0,+∞[of (z − P )−1 : Hcomp → Dloc.

We need some remarks about truncated spectral projections. Let

1B(0,R0) ≺ χ ∈ C∞0 (Rn)

be such that “the exterior” operator Q has analytic coefficients in a neigh-borhood of supp (1 − χ). We choose Γθ so that Γθ = Rn near suppχ. Letz0 ∈ ei]−2θ,0]]0.+∞[ be a resonance, 0 < θ ≤ θ0. Let

πθ,z0 =1

2πi

∫∂D(z0,ε)

(z − Pθ)−1dz, Pθ = P|Γθ , 0 < ε 1, (11.1)

be the corresponding spectral projection. We know that rank πθ,z0 is inde-pendent of the choice of θ with −1

2arg z0 < θ ≤ θ0.

Lemma 11.1 rank (χπθ,z0) = rank (πθ,z0).

124

Proof. We know that R(πθ,z0) = N ((Pθ − z0)N0) for some 1 ≤ N0 ≤ m(z0).Let u ∈ R(πθ,z0) with χu = 0. From the extension lemma 7.2, we concludethat u = 0. It follows that rank (πθ,z0) ≤ rank (χπθ,z0), and the lemma followssince the opposite inequality is obvious. #

Recall that Hθ = HR0 ⊕ L2(Γθ \ B(0, R0)). Let Γθ = Γθ. If u ∈ L2(Γθ),then u∗(x) := u(x) ∈ L2(Γ−θ), and we see that L2(Γ−θ) and L2(Γθ) becomemutually dual spaces for the scalar product (u|v) =

∫Γθu(x)v∗(x)dx. This

extends in the natural way to a duality between Hθ and H−θ, and P−θ :Hθ → H−θ becomes the adjoint of Pθ. If z0 is an eigenvalue of Pθ, then z0 isan eigenvalue of P−θ, and we get π−θ,z0 = π∗θ,z0 for the corresponding spectralprojections. It follows that

rank (πθ,z0χ) = rank (χπ−θ,z0) = rank (π−θ,z0) = rank (πθ,z0).

Hence R(πθ,z0) = R(πθ,z0χ), and we get:

rank (πθ,z0) = rank (χπθ,z0) = rank (πθ,z0χ) = rank (χπθ,z0χ). (11.2)

Let R(z) : Hcomp → Dloc denote the meromorphic extension of (z − P )−1

from ei]0,ε0[]0,+∞[ to ei]−2θ0,ε0[]0,+∞[. From Chapter 7 it follows that if z0

is a resonance, then

rankπz0 ≤ rankπθ,z0 , −1

2arg z0 < θ ≤ θ0, (11.3)

where

πz0 =1

2πi

∫∂D(z0,ε)

R(z)dz.

On the other hand, since χ(z − Pθ)−1χ = χR(z)χ, it is clear that

χπz0χ = χπθ,z0χ, (11.4)

so from (11.2), (11.3), we get

rank (πz0) = rank (χπz0χ) = rank (πθ,z0) = rank (χπθ,z0χ). (11.5)

Recall that among our assumptions we have (7.12):

N(P ], [−λ, λ]) = O(1)Φ(λ

h2), λ ≥ 1.

125

We also made assumptions on Φ which imply that Φ(t) = O(tn]/2) for some

n] ≥ n. The following result is a simple long-range extension of a result ofP. Stefanov [87]. (The main novelty of Stefonov’s result compared to that ofTang-Zworski [89] was that Stefanov is able to treat multiplicities in the casewhen quasi-modes are very close to each other.)

Theorem 11.2 Let H ⊂]0, h0] have 0 as an accumulation point. Let 0 <a0 ≤ a(h) ≤ b(h) ≤ b0 < ∞ be two functions on H. Assume that for everyh ∈ H, ∃m(h) ∈ 1, 2, ... and Ej(h) ∈ [a(h), b(h)], uj(h) ∈ D, 1 ≤ j ≤m(h), such that suppuj ∈ K ⊂⊂ Rn, where K is independent of h,

‖(P (h)− Ej(h)))uj(h)‖H ≤ R(h), 1 ≤ j ≤ m(h),

|(ui(h)|uj(h))− δi,j| ≤ R(h),

where R(h) = O(h∞). Let S(h) be a function with

max(h−n]−1R(h), e−D/h) ≤ S(h) = O(h∞),

for some fixed constant D > 0.Then for every k ∈ N, ∃h(S, k) > 0, such that for H 3 h < h(S, k), P (h)

has at least m(h) resonances in

[a(h)− 6hk, b(h) + 6hk] + i[0, 2S(h)h−1−n] ].

The difficulty is that the problem is non-selfadjoint. Under the same assump-tions, it is much easier to prove that P ](h) has at least m(h) eigenvalues in[a(h) − hk, b(h) + hk]. In particular, we know that m(h) = O(h−n

]). The

following lemma is essentially from [89].

Lemma 11.3 For z in some fixed compact subset of ei]−2θ0,0]]0,∞[, thereexists A > 0 such that we have

‖χ(z − P (h))−1χ‖H→H ≤ AeAh−n] log g(h)−1

,

if dist (z,ResP (h)) ≥ g(h) > 0. (Here (z−P (h))−1 denotes the meromorphicextension of the resolvent, when Im z ≤ 0.)

Proof. Recall from Chapter 8 that

z − Pθ = (1 + K(z))(z − Pθ), θ = θ0,

126

where ‖K‖ = O(1). Also recall that χ(z − P (h))−1χ = χ(z − Pθ)−1χ. It istherefore enough to show that

‖(1 + K(z))−1‖ ≤ AeAh−n] log 1

g , when dist (z,ResP ) ≥ g(h). (11.6)

Recall thatdet(1 + K(z)) = GDw,

where |G| ≥ A−1e−Ah−n]

It is easy to see that |Dw| ≥ A−1e−Ah−n] log 1

g , andwe get (11.6) with a new constant A. #

Notice that ‖χ(z − P )−1χ‖ ≤ 1/|Im z|, for Im z > 0. We state thefollowing consequence of the maximum principle essentially from [89].

Lemma 11.4 For h ∈ H, let F (z;h) be holomorphic near Ω5(h) = [E(h)−5hk, E(h) + 5hk] + i[−S(h)h−1−n] , S(h)], with E(h) ∈ R, S(h) as in thetheorem. Assume that

|F (z;h)| ≤ A exp(Ah−n]

log1

hS(h)) on Ω5(h),

|F (z;h)| ≤ 1

|Im z|on Ω5(h) ∩ z; Im z > 0.

Then there exist h1 = h1(S,A, k) > 0, B = B(S,A, k) > 0 (independent ofh) such that

|F (z;h)| ≤ B

S(h), z ∈ [E(h)− hk, E(h) + hk], h ∈ H, h ≤ h1.

Proof. This is a routine argument for subharmonic functions. Consider thesubharmonic function

f(z) = log |F (z;h)| − log (1

S(h))

Im z + S(h)h−1−n]

S(h) + S(h)h−1−n]

−((logA) + Ah−n]

log (1

hS(h)))−Im z + S(h))

(S(h) + S(h)h−1−n] ,

which is ≤ 0 on the horizontal segments of ∂Ω5(h) and satisfies

|f(z)| ≤ (logA) + Ah−n]

log (1

hS(h))

127

on the vertical segments.After the change of coordinates, z = (S + Sh−1−n])w, Ω5 becomes a

rectangle of vertical “width” 1 and of horizontal “length” 10hk

S(1+h−1−n] ). In such

a long rectangle, the Poisson kernel decays exponentially (see for instance[76]), and it follows that in the rectangle of the same center and same width

and of length 2hk

S(1+h−1−n] ), we have

|f(z)| ≤ ((logA) + Ah−n]

log (1

hS))e−h

k/(S(1+h−1−n] ))

≤ O(h∞) log (1

hS) ≤ log

1

S.

Restricting this further to real z, we get

log |F (z;h)| ≤ O(1) log1

S,

and the lemma follows. #

Let z0 be a resonance. Then we know that for z ∈ neigh (z0):

(z − Pθ)−1 =N∑j=1

π(j)z0,θ

(z − z0)j+ hol (z),

whereπ

(1)z0,θ

= πz0,θ

is the spectral projection and R(π(j)z0,θ

) ⊂ R(πz0,θ) = R(πz0,θχ), j ≥ 1, if we

choose χ as in the beginning of this chapter. It follows that R(χπ(j)z0,θχ) ⊂

R(χπz0,θχ), and recalling that χ(z − P )−1χ = χ(z − Pθ)−1χ, we get

χ(z − P )−1χ =N∑j=1

(z − z0)−jAj + hol (z), z ∈ neigh (z0), (11.7)

with Aj = χπ(j)z0,θχ, R(Aj) ⊂ R(A1), rank (A1) = m(z0).

End of the proof of the theorem. Let z1(h), ..., zM(h)(h), be the distinctresonances of P (h) in

Ω6(h) := [a(h)− 6hk, b(h) + 6hk] + i[−2S(h)h−1−n] , S(h)],

128

and let

A(j)(h) = A1,zj(h) =1

2πi

∫∂D(zj ,ε)

χ(z − P )−1χdz.

Here we choose χ as above, with the additional property that 1K ≺ χ. Letπ(h) be the orthogonal projection onto

∑M(h)1 A(j)(h)H, π′(h) = 1 − π(h).

Then π′(h)χ(z − P (h))−1χ is holomorphic for z in a neighborhood of Ω6(h).On the other hand, we have

‖π′(h)χ(z − P )−1χ‖H→H ≤ AeAh−n] log 1

hS(h) (11.8)

in Ω6(h) \ ∪M(h)j=1 D(zj, hS(h)). Put

Ω5(h) = [a(h)− 5hk, b(h) + 5hk] + i[−S(h)h−1−n] , S(h)]. (11.9)

Then (11.8), the maximum principle and the upper bound on M(h) implythat (11.8) holds in all of Ω5(h), without any discs removed. (Notice thatwe cannot have any connected chain of discs D(zj, hS(h)) from Ω5(h) to∂Ω6(h) \R.)

Now for Im z > 0, we have

‖π′(h)χ(z − P (h))−1χ‖H→H ≤1

|Im z|, (11.10)

and together with (11.8) and Lemma 11.4, this implies:

‖π′(h)χ(z − P (h))−1χ‖H→H ≤B

S(h), z ∈ [a(h)− hk, b(h) + hk]. (11.11)

Now, use the quasimodes and conclude that for Im z > 0:

π′(h)uj(h) = π′(h)χ(z − P )−1χ(z − P )uj,

where we also used that χ = 1 near suppuj. Let z tend to Ej and obtain:

‖π′uj‖ ≤B

S(h)R(h) ≤ Bhn

]+1.

In other words,πuj = uj + wj,

with‖wj‖ ≤ Bhn

]+1.

It follows thataj,k := (πuj|πuk)− δj,k = O(hn

]+1),

so the matrix 1 + (aj,k) is invertible and hence (πuj)m(h)j=1 are linearly inde-

pendent. Hence dimR(π) ≥ m(h), so M(h) ≥ m(h). #

129

12 Microlocal analysis and dynamical bounds

In this chapter, we develop some amount of analytic microlocal analysis inorder to study upper bounds for the number of resonances for second orderoperators in the semi-classical limit. In the case the coefficients are analyticeverywhere, the first and most basic result (Theorem 12.13) is that if thereare no trapped classical trajectories at some positive energy, say 1, thenthere is a complex neighborhood of 1, independent of h, which containsno resonances when h is small enough. The second main result (Theorem12.15) is un upper bound on the number of resonances in certain possiblyh-dependent neighborhoods of 1 in terms of a certain escape function, whichmeasures the amount of trapped trajectories. When the classical dynamicsis hyperbolic this leads to a more explicit estimate in terms of the dimensionof the set of trapped trajectories. (See Theorem 12.22.). The first resultis implicit in Helffer-Sjostrand [38] and is really a direct consequence of thetheory developed there which is a direct microlocal treatment of resonancesby means of certain FBI-transforms and suitable weighted spaces. The othertwo results were obtained in [72]. The theory of [38] permits to treat operatorswith a quite general behaviour near infinity and that was of importance insome of the main applications there. In [72] we used the same theory as thegeneral frame-work.

Lahmar-Benbernou and Martinez [10] noticed that if we retrict the atten-tion to operators that converge to the Laplacian near infinity, then one canuse a Bargman transform (which is a special case of FBI-transforms) after asimple complex scaling, and get a simpler version of the theory. This simpli-fies the treatment near infinity. We follow that approach here, which permitsto get faster to the main ideas of [38], [72], and especially the approach of[72] is followed without any essential changes.

12.1 Trajectories and escape functions

We will consider a simplified version of scattering for classical particles (Si-mon, Hunziker) and the relation with the existence of certain escape func-tions. We follow the appendix of [29]. Consider the classical Hamiltonian

p(x, ξ) =∑|α|≤2

aα(x)ξα, (12.1)

130

where aα ∈ C∞(Rn; R), ∂βxaα(x) = oβ(1)〈x〉−|β|, x→∞, β 6= 0,

p(x, ξ)− ξ2 → 0, x→∞, (12.2)

∑|α|=2

aα(x)ξα ≥ 1

C|ξ|2. (12.3)

We shall mainly work near the energy surface p−1(1).Let

G(x, ξ) = x · ξ. (12.4)

If p0(x, ξ) = ξ2, then Hp0G = 2ξ · ∂x(x · ξ) = 2ξ2, which is ≥ 1/C onp−1

0 ([1−ε, 1+ε]) if ε ≤ 1/2. If we restrict the attention to p−1([1−ε, 1+ε]) =:Σεp, then

|(p′ξ − p′0,ξ)(x, ξ)| = o(1), |(p′x − p′0,x)(x, ξ)| = o(1)〈x〉−1, |x| → ∞,

while G′x, 〈x〉−1G′ξ remain bounded, so

|(HpG−Hp0G)(x, ξ)| → 0, Σεp 3 (x, ξ)→∞. (12.5)

Also note that the Hp-flow is complete in the sense that a trajectory t 7→exp(tHp)(x, ξ) cannot reach infinity in finite time.

Fix ε ∈]0, 12]. Clearly there is a compact set K ⊂ Σε

p, such that

HpG ≥1

C, on Σε

p \ K, (12.6)

for some C > 0. Choose T > 0 large enough so that

K ⊂ ρ ∈ Σεp; 1− T < G(ρ) < T − 1. (12.7)

Define the outgoing (+) and incoming (–) tails:

Γ+ = ρ ∈ Σεp; exp(tHp) 6→ ∞, t→ −∞,

Γ− = ρ ∈ Σεp; exp(tHp) 6→ ∞, t→ +∞.

Notice that if ρ ∈ Σεp \ Γ+, then G(exp tHp(ρ)) ≤ 1 − T for some t(ρ) < 0,

and that G(exp tHp(ρ)) −∞, t(ρ) > t→ −∞. It follows that Γ+ is closed(and the same holds for Γ−).

131

Proposition 12.1 For every s ∈ R, Γ+∩ρ ∈ Σεp;G(ρ) ≤ s and Γ−∩ρ ∈

Σεp;G(ρ) ≥ s are compact.

Proof. We only consider the first of the two sets. Notice that Γ+ ∩ ρ ∈Σεp; G(ρ) ≤ s is empty when s ≤ 1 − T . Clearly there is a compact set

K ⊂ Σεp, depending on s, such that if ρ ∈ K and |t| ≤ (s + T − 1)C, then

exp tHp(ρ) ∈ K. Then if µ ∈ Σεp \ K, G(µ) ≤ s, we know that exp(tHp)(µ)

reaches the region G ≤ 1−T for some t ∈ [−C(s+T − 1), 0]. Hence µ 6∈ Γ+.#

Define the trapped set K := Γ+∩Γ−. It is easy to see that K is compact.

Proposition 12.2 If Γ− 6= ∅ (or if Γ+ 6= ∅), then K 6= ∅.

Proof. Let ρ ∈ Γ−. Then exp tHp(ρ); t ≥ 0 is contained in a compactset L, so we have exp tjHp(ρ) → ρ0 ∈ Σε

p, for some sequence tj → +∞.Put ρj = exp tjHp(ρ). For every S > 0, we have exp tHp(ρj) → exp tHp(ρ0)uniformly for |t| < S. But exp tHp(ρj) = exp(tj + t)Hp(ρj), for |t| < S andj large enough so that tj ≥ S. Hence exp tHp(ρ0) ∈ L, |t| < S. Since S canbe chosen arbitrarily large, we conclude that ρ0 ∈ K. #

Though we will not need it in the following, we also consider the truetails:

T± = Γ± \K.

On Σεp we have the symplectic volume which is Hp-invariant. For any s <

T − 1, the set T− ∩ ρ ∈ Σεp; G ≥ s is bounded and hence of finite volume.

On the other hand, for s < 1− T , we have

exp(tHp)(T− ∩ G ≥ s) ∅, t→ +∞.

Hence

Vol (T− ∩ G ≥ s) = Vol (exp(tHp)(T− ∩ G ≥ s))→ 0, t→ +∞.

We get:

Proposition 12.3Vol (T±) = 0.

In [29] we also proved:

132

Proposition 12.4 The following three statements are equivalent:

(a) T+ 6= ∅(b) T− 6= ∅(c) If Kα = ρ ∈ Σε

p; dist (ρ,K) ≤ α, then K and Kα \ K are non-emptyfor every α > 0.

Example. A potential well in an island. [[[Elaborate.]]]

We next construct a modification G of G with G = G outside a boundedset and with HpG ≥ 0 everywhere on Σε

p. Increasing K if necessary, we

may assume that K ⊂ K. Let HT = ρ ∈ Σεp; G(ρ) = T. We have a

diffeomorphism κ+ : R × HT → Σεp \ Γ−, given by κ+(t, ρ) = exp(tHp)(ρ).

Similarly, we have a diffeomorphism κ− : R × H−T → Σεp \ Γ+ (with the

obvious definition of H−T ).Choose 0 < f+ ∈ C∞0 (Σε

p \ Γ−) such that f+ ≥ Hp(G) with equalityin G ≥ T and outside a compact set in −T ≤ G ≤ T. Let G+ ∈C∞(Σε

p \ Γ−) be the solution of

HpG+ = f+, G+|HT= T.

Then G+ = G in G ≥ T and outside a compact set in −T ≤ G ≤ T.We also have G+ ≤ G and choosing f+ large enough, we may assume that

lim supρ→Γ−∪H−T

G+(ρ) ≤ −T. (12.8)

Let f−, G− have the analogous properties.Choose θ± ∈ C∞(R; [0, 1]) with 1 = θ+ + θ−, supp θ+ ⊂]1 − T,+∞[,

supp θ− ⊂]−∞, T − 1[, and put

χ+(t) =∫ t

−∞θ+(s)ds, χ−(t) =

∫ t

+∞θ−(s)ds.

We may further arrange so that θ−(−t) = θ+(t). Then χ′± = θ± and we have

suppχ+ ⊂]1− T,+∞[, suppχ− ⊂]−∞, T − 1[,

χ+(t) + χ−(t) = t.

Put G = χ+ G+ + χ− G− and notice that χ+ G+ extends to a smoothfunction (= 0) near G ≤ −T ∪ Γ−. Similarly χ− G− is smooth and well-defined on all of Σε

p. By construction we have G = G in G ≤ −T ∪ G ≥T ∪ (−T < G < T \ a compact set).

133

We haveHpG = (χ′+ G+)f+ + (χ′− G−)f− ≥ 0, (12.9)

∇G = (χ′+ G+)∇G+ + (χ′− G−)∇G−, (12.10)

and get

Proposition 12.5 Given G = x · ξ, we can find a new “escape” functionG ∈ C∞(R2n; R), equal to G outside a compact set, such that on Σε

p we have

G = 0 in a neighbohood of K = Γ+ ∩ Γ− and such that locally uniformly onΣεp:

Hp(G) ≥ 1

C0

|∇G|. (12.11)

We may arrange so that Hp(G) > 0 in Σεp outside an arbitrarily small neigh-

borhood of K, and so that Hp(G) > 0 everywhere on Σεp in the case when

K = ∅.

12.2 Bargmann transforms and pseudodifferential op-erators

We first review some well-known facts about Bargmann transforms (whichcan be viewed as linearized models for FBI-tranforms) following [78]. Letφ(x, y) be a holomorphic quadratic form (i.e. a homogenous polynomial ofdegree 2) on Cn

x ×Cny such that

detφ′′xy 6= 0, Imφ′′yy > 0. (12.12)

In the case of a classical Bargmann transform, we have φ(x, y) = i2(x− y)2.

PutΦ(x) = sup

y∈Rn−Imφ(x, y) (12.13)

Notice that the supremum is attained at a unique point y(x) ∈ Rn. In thespecial case above, we get Φ(x) = 1

2(Imx)2.

For u ∈ S ′(Rn), 0 < h ≤ 1, the function

Tu(x;h) = Ch−3n/4∫eiφ(x,y)/hu(y)dy, x ∈ Cn (12.14)

is entire and satisfies |Tu(x;h)| ≤ Oh(1)〈x〉NueΦ(x)/h for some Nu ∈ R. Ifu ∈ S(Rn), then |Tu(x;h)| ≤ ON,h(1)〈x〉−NeΦ(x)/h for every N ∈ R.

134

We view T as a semiclassical Fourier integral operator with associatedlinear canonical transformation

κT : C2n 3 (y,−φ′y(x, y)) 7→ (x, φ′x(x, y)) ∈ C2n. (12.15)

For every x ∈ Cn there is a unique point (y(x), η(x)) ∈ R2n with πx κT (y(x), η(x)) = x, where πx is the natural projection C2n

x,ξ → Cnx. Indeed,

y(x) must be the same as above and η(x) = −φ′y(x, y(x)). Let ξ(x) ∈ Cn bethe point with (x, ξ(x)) = κT (y(x), η(x)), so that ξ(x) = φ′x(x, y(x)).

Compare with

2

i

∂Φ

∂x(x) =

2

i

∂x(−Imφ(x, y(x))) =

2

i

∂x(−Imφ)y=y(x) =

2

i

∂x(i

2(φ(x, y)− φ(x, y)))y=y(x) = (

∂xφ)(x, y(x)) = ξ(x).

We conclude that

κT (R2n) = ΛΦ := (x, 2

i

∂Φ

∂x(x));x ∈ Cn

x. (12.16)

Since κT is canonical for the complex symplectic form σ =∑n

1 dηj ∧ dyj, andR2n is I-Lagrangian, i.e. a Lagrangian manifold for the real symplectic formImσ, the same holds for σ|ΛΦ

. Further, the real 2-form σ|R2n is nondegenerate,

so the same holds for σ|ΛΦ. We say that ΛΦ, like R2n is an IR-manifold:

Lagrangian for Imσ and symplectic for the restriction of Reσ.What does this means in terms of Φ? In general, let F (x) be a smooth

real-valued function on some open set in Cn and define ΛF as above. Considerthe restriction to ΛF of the fundamental one form ξ · dx (whose exteriordifferential equals σ):

ξ · dx|ΛF '2

i

∂F

∂x· dx ' 2

i∂F, (12.17)

−Im ξ · dx|ΛF 'i

2(2

i

∂F

∂x· dx− 2

−i∂F

∂x· dx) = dF,

so−Imσ|ΛF = d(−Im ξ · dx)|ΛF ' d2F = 0.

Hence ΛF is I-Lagrangian for all sufficiently smooth F . Conversely, everysmooth I-Lagrangian manifold of the form ξ = ξ(x) is locally of the form ΛF

for a suitable smooth function F (exercise). From (12.17) we get

σ|ΛF ' d(2

i

∂F

∂x· dx) =

2

i

∑j,k

∂2F

∂xj∂xkdxj ∧ dxk (=

2

i∂∂F ). (12.18)

135

We check that ΛF is R-symplectic iff det( ∂2F

∂x∂x) 6= 0. Returning to ΛΦ we then

know that det ∂2Φ∂x∂x

6= 0. On the other hand, (12.13) tells us that Φ(x) is thesupremum of a family of pluriharmonic functions, so it is plurisubharmonicand then also strictly plurisubharmonic:

∂2Φ

∂x∂x> 0. (12.19)

Remark. IR-manifolds in C2nx,ξ are totally real of maximal real dimension 2n.

(More details should be given here in later versions, and in the meanwhilewe refer to [73, 78].)

We shall show that T is bounded: L2(R2) → HΦ(Cn) := Hol (Cn) ∩L2(Cn; e−2Φ(x)/hL(dx)), where L(dx) denotes the Lebesgue measure on Cn

and Hol (Ω) denotes the space of holomorphic functions on the open setΩ ⊂ Cn. The formal adjoint of T is given by

T ∗v(y) = Ch−3n4

∫e−iφ

∗(x,y)/hv(x)e−2Φ(x)/hL(dx), (12.20)

where φ∗(x, y) := φ(x, y) is holomorphic. Then we get for v ∈ Hol (Cn), with|v| ≤ ON,h(1)〈x〉−NeΦ(x)/h for all N :

TT ∗v(x) = |C|2Cφh−3n2

+n2

∫e

2h

Ψ(x,z)u(z)e−2h

Φ(z)L(dz), Cφ > 0, (12.21)

where

Ψ(x,w) = vc.yi

2(φ(x, y)− φ∗(w, y)).

Here “vc.y” means “critical value with respect to y of” .Ψ(x,w) is holomorphic in both variables and

Ψ(x, x) = vc.yi

2(φ(x, y)− φ∗(x, y)) = Φ(x). (12.22)

Consider the strictly plurisubharmonic quadratic form

F (x, y) =1

2(Φ(x)+Φ(y))−Re Ψ(x, y) =

1

2Φ(x)+

1

2Φ(y)−1

2Ψ(x, y)−1

2Ψ(x, y),

which vanishes on the anti-diagonal D = y = x. Computing ∂x and ∂x of(12.22), we get

∂xΨ = ∂xΦ, ∂yΨ = ∂xΦ on D.

136

It follows that ∂xF = 0, ∂yF = 0 on D and since F is real-valued, that ∇F =0 on D. The strict plurisubharmonicity of F now implies that F (x, y) ∼|x− y|2, or in other words,

−Φ(x) + 2Re Ψ(x, z)− Φ(z) ∼ −|x− z|2. (12.23)

Let Πu denote the right hand side of (12.21). Using (12.23) we see that

Π = O(1) : L2(e−2Φ/hL(dx))→ Hol ∩ L2(e−2Φ/hL(dx)) =: HΦ.

Taking ∂x∂x of (12.22) and using (12.19), we get

det ∂x∂yΨ(x, y) 6= 0.

For λ > 0, u ∈ Hol (Cn), |u| ≤ Oh(1)〈x〉N0eΦ(x)/h, consider

Iλu(x) =1

(2πh)n

∫∫Γλ(x)

eih

(x−y)·θu(y)dydθ, (12.24)

Γλ(x) : θ =2

i

∂Φ

∂x(x+ y

2) + i

λ

2(x− y).

It is easy to see that the integral converges and is independent of λ (byStokes’ formula). Letting first λ become very large we can further replace Γλby

Γt,λ : θ = t2

i

∂Φ

∂x(x+ y

2) + i

λ

2(x− y), 0 ≤ t ≤ 1.

Then let λ→ +∞ and consider

I0,λu(x) =1

(2πh)n

∫∫Γ0,λ(x)

eih

(x−y)·θu(y)dydθ

=λn

(2πh)n

∫∫e−

λh|x−y|2u(y)L(dy)→ u(x), λ→∞.

We obtainIλu(x) = u(x), u ∈ HΦ. (12.25)

On the other hand, we can use the Kuranishi trick to write Iλu(x) as

Iλu(x) =C

hn

∫∫Γλ(x)

e2h

(Ψ(x,z)−Ψ(y,z))u(y)dydz. (12.26)

137

This is done by a change of variables, using that 2(Ψ(x, z) − Ψ(y, z)) =i(x − y) · θ(x, y, z) with θ(x, y, z) = 2

iΨ′x(

x+y2, z). Here Γλ(x) is the image

contour in the (y, z)-variables.If we replace Γλ(x) by the contour z = y (which can be justified by means

of Stokes’ formula), we get

C

hn

∫∫e

2h

Ψ(x,y)u(y)e−2h

Φ(y)dydy, (12.27)

which is a non-vanishing constant times Πu(x) in (12.21). Both Π and (12.27)are positive semi-definite on HΦ, so C in (12.27) is C > 0. Adjusting thecontant C in the definition of T , we get Πu(x) precisely. Using Stokes’formula, we check quite easily that Πu(x) = Iλu(x) for u ∈ Hol (Cn) with|u(x)| ≤ C(h)〈x〉N0eΦ(x)/h for some N0 > 0. In view of (12.25) we have shownthat

Πu = u, (12.28)

for such functions u, and in particular for u ∈ HΦ. It is also clear that Π isself-adjoint in L2

Φ := L2(Cn; e−2Φ/hL(dx)). Finally Π maps the latter spaceinto HΦ, so

Π is the orthogonal projection : L2Φ → HΦ. (12.29)

We next claim that

T is unitary:L2(Rn)→ HΦ. (12.30)

We have already seen that T is bounded: L2(Rn)→ HΦ and that TT ∗ = 1 onHΦ, so it only remains to see that T is injective: If Tu = 0, then (∂αxTu)(0) =0 for all α ∈ Nn, and hence∫

eihφ(0,y)p(y)u(y)dy = 0,

for every polynomial p. If we let F denote the Fourier transform, then thismeans that

p(Dη)(F(u(y)eihφ(0,y)))(0) = 0,

for all polynomials p, and hence that F(u(y)eihφ(0,y))(η) = 0, since this func-

tion is entire.We next look at the action of pseudodifferential operators. Let S(ΛΦ,m)

be the space of all C∞-functions a on ΛΦ such that ∂αa = Oα(m) for every

138

multiindex α, when identifying ΛΦ linearly with R2n. Here m is an orderfunction with

m(x) ≤ C〈x− y〉N0m(y), x, y ∈ ΛΦ,

for some constants C,N0 > 0. When our symbols depend on h, it is assumedthat the above estimates are uniform.

If u ∈ ON,h(1)〈x〉−NeΦ(x)/h for every N , we put

Oph(a)u(x) =1

(2πh)n

∫∫Γ(x)

eih

(x−y)·θa(x+ y

2, θ)u(y)dydθ, (12.31)

where Γ(x) is the contour: θ = 2i∂Φ∂x

(x+y2

). Notice that on this contour,

−Φ(x) + Re (i(x− y) · θ) + Φ(y) = 0,

so the integral converges. Using this parametrization, we get

Oph(a)u(x) = h−n∫k(x, y;h)u(y)dy,

where ∂xk = ∂yk, so by integration by parts, we see that Oph(a)u(x) isholomorphic if u is.

Let m be a second order function on ΛΦ. Both m and m will be viewedas functions on Cn

x, by the natural identification πx : ΛΦ → Cnx, where

πx : (x, ξ) 7→ x is the projection onto the x-space. Let

L2Φ,m = L2(Cn; m2e−2Φ/hL(dx)), HΦ,m = Hol(Cn) ∩ L2

Φ,m.

Proposition 12.6 Oph(a) extends to a bounded operator: HΦ,m → HΦ, mm

,

whose norm is uniformly bounded with respect to h.

Proof. For 0 ≤ t ≤ 1, let Γt(x) be the contour

θ =2

i

∂Φ

∂x(x+ y

2) + it

x− y〈x− y〉

,

parametrized by y ∈ Cn and let G[0,1](x) be the (n+ 1)-contour obtained byletting also t ∈ [0, 1] be a parametrizing variable.

Now we have an almost holomorphic extension a of a to C∞(C2n) suchthat

∂a(x, ξ) = ON(1)m(x)|ξ − 2

i

∂Φ

∂x|N , |ξ − 2

i

∂Φ

∂x| ≤ O(1),

139

for every N ≥ 0 and such that a(x, ξ) = O(1)m(x). For simplicity we shallwrite a instead of a. Then Stokes’ formula gives

Oph(a)u(x) =1

(2πh)n

∫∫Γ1(x)

eih

(x−y)·θa(x+ y

2, θ)u(y)dydθ (12.32)

+1

(2πh)n

∫G[0,1](x)

eih

(x−y)·θu(y)∂y,θ(a(x+ y

2, θ)) ∧ dydθ.

For the first term to the right, we notice that along Γ1(x) we have dydθ =O(1)dydy and

−Φ(x) + Re (i

h(x− y) · θ) + Φ(y) = −|x− y|

2

〈x− y〉.

Hence the first term to the right in (12.32) can be written

m(x)

m(x)e

Φ(x)h h−n

∫e−

1h|x−y|2〈x−y〉 f(x, y;h)u(y)e−

Φ(y)h m(y)L(dy),

with

f(x, y;h) = O(1)m(x+ y

2)

m(x)

m(x)m(y)≤ O(1)〈x− y〉N0 ,

so the reduced kernel

h−ne−1h|x−y|2〈x−y〉 f(x, y;h)

is of modulus

≤ Ch−ne−1h|x−y|2〈x−y〉 〈x− y〉N0

which is the kernel of a uniformly bounded convolution operator: L2 → L2.Thus the first term of the right hand side of (12.32) is

O(1) : L2Φ,m → L2

Φ, mm

.

For the last term in (12.32), we notice that along G[0,1](x):

∂y,θ(a(x+ y

2, θ)) ∧ dydθ = ON(1)m(

x+ y

2)tN|x− y|N

〈x− y〉NL(dy)dt,

for every N ≥ 0.

140

For every t we then get the reduced operator

v 7→ h−n∫e−

th|x−y|2〈x−y〉 kt(x, y;h)v(y)L(dy),

with

|kt(x, y;h)| = ON(1)m(x+ y

2)m(x)

m(x)

1

m(y)tN|x− y|N

〈x− y〉N≤ ON(1)tN

|x− y|N

〈x− y〉N|x−y|N0 ,

for some fixed N0 > 0. We estimate the L2-norm of this operator by theL1-norm of the corresponding convolution kernel:

ON(1)h−n∫e−

th|x|2〈x〉 tN

|x|N+N0

〈x〉NL(dx).

Cut the integral into two parts: |x| ≤ 1, |x| ≥ 1: For∫|x|≤1 ... we get the

bound:

O(1)h−n∫ ∞

0e−

t2hr2

tNrN+N0+2n−1dr =

O(1)h−n∫ ∞

0e−

tr2

2h (tr2

h)N+N0+2n

2dr

r· h

N+N0+2n

2 tN−N+N0+2n

2 = O(1)hN+N0

2 ,

if N is large enough.For

∫|x|≥1, we get the bound:

O(1)h−n∫ ∞

1e−

t2hrtNrN0+2ndr

r,

and the change of variables r = htρ gives

O(1)h−n∫ ∞th

e−ρ2 (h

tρ)N0+2ndρ

ρtN ≤

O(1)h−n∫ ∞th

e−ρ2ρN0+2ndρ

ρhN0+2ntN−N0−2n.

The last integral is O((1 + th)−M) for every M ≥ 0. For t ≤ h1/2, we gain

as many powers of h as we like from tN−N0−2n. For t > h1/2 the integral isO(( t

h)−M) ≤ O(hM/2) for all M and we also gain as many powers of h as we

like.

141

We conclude that the last term in (12.32) defines an operator which is

O(h∞) : L2Φ,m → L2

Φ, mm

.

This completes the proof of the proposition. #

Exercise: Show that TS(Rn) = ∩N≥0HΦ,〈x〉N . Hint: We already know thatTS ⊂ ∩N≥0.... It then suffices to show that T ∗(∩N≥0...) ⊂ S, since TT ∗ = 1on HΦ.

Exercise: Show that ∩N≥0HΦ,〈x〉N is dense in HΦ. Hint: Let χ ∈ C∞0 (Cn) beequal to 1 near 0 and consider Π(χ( x

R)u), R→∞.

We end this section by discussing the link with h-pseudodifferential op-erators on Rn. If m is an order function on R2n and a ∈ S(R2n,m), thenwe can define Oph(a) : S(Rn)→ S(Rn) and we have boundedness results inSobolev spaces analogous to the HΦ,m above. A classical way of developingthe theory for the Weyl quantization on Rn (see [21] for more details) is toobserve that if `(x, ξ) is a real linear form on R2n, then Oph(`) = `(x, hDx)and that this operator is essentially self-adjoint from C∞0 (Rn). Moreover

e−i`(x,hDx) = Oph(e−i`(x,ξ)). (12.33)

Imitating a proof of an important invariance property for the Weyl-quantization under conjugation by certain metaplectic operators, one thenobserves that if k is the linear form on ΛΦ with

k κT = `, (12.34)

thenk(x, hDx)Tu = T`(x, hDx)u, ∀u ∈ S. (12.35)

Now on the transform side, we have from (12.35) and the unitarity of T , thatk(x, hDx) is essentially self-adjoint on HΦ from TS(Rn) and we also checkthat

k(x,Dx) = Oph(k), e−ik(x,hDx) = Oph(e−ik(x,ξ)).

Again from (12.35), it follows that

e−ik(x,hDx)T = Te−i`(x,hDx). (12.36)

If a ∈ S(R2n) and b ∈ S(ΛΦ) are related by b κT = a, then we canrepresent a as a superposition of linear exponentals of the type e−i`(x,ξ) (by

142

Fourier’s inversion formula) and we get a corresponding superposition of b inexponentials e−ik(x,ξ). Passing the quantizations, we get

Oph(b) T = T Oph(a). (12.37)

By a density argument this relation extends to the case b ∈ S(ΛΦ,m), a ∈S(R2n,m κT ).

For the later applications, we will need very little calculus of pseudo-differential operators. For short presentations, see [32] for the classical way(with further references given there) with a maximal use of the method ofstationary phase, and [21] for the approach sketched here with essentially noexplicit use of stationary phase.

12.3 Pseudodifferential operators with holomorphic sym-bols

Assume that a(x, ξ) is holomorphic in a tubular neighborhood ΛΦ +W of ΛΦ,where W is an open bounded neighborhood of 0 ∈ C2n, and of class S(m)there, where m is an order function defined first on ΛΦ and then extendedto a full neighborhood by putting m(x, ξ) = m(x, 2

i∂Φ∂x

(x)).In this case, we have no remainder term in a formula like (12.32) and we

get

Oph(a)u(x) =1

(2πh)n

∫∫Γc(x)

eih

(x−y)·θa(x+ y

2, θ)u(y)dydθ, u ∈ HΦ,m,

(12.38)if c > 0 is small enough, so that Γc(x) ⊂ ΛΦ +W .

LetΦ(x) = Φ(x) + f(x), (12.39)

where f ∈ C1,1(Cn) has support in some fixed compact set and sufficientlysmall norm: ‖f‖C1,1 = ‖∇f‖Lip +supCn |f(x)|, where ‖g‖Lip = supx 6=y |g(x)−g(y)|/|x − y|. Recall that the gradient of a Lipschitz function g belongs toL∞ and that ‖g‖Lip = ‖∇g‖L∞ (with the right choice of pointwise norm), sothe same holds for the derivatives of functions of class C1,1.

We can define the Hilbert space HΦ,m

, which coincides with HΦ,m as aspace. The corresponding norms are equivalent for every fixed h but notuniformly when h → 0. We have an associated Lipschitz manifold Λ

Φ: ξ =

143

2iδΦ∂x

(x) which coincides with ΛΦ outside a compact set and which is close toΛΦ in the sense of Lipschitz graphs (since f is small in C1,1).

If c > 0 is small enough, we can use Stokes’ formula to replace Γc(x) in(12.40) by

Γc(x) : θ =2

i

∂Φ

∂x(x+ y

2) + ic

x− y〈x− y〉

. (12.40)

Lemma 12.7 Oph(a) is bounded: HΦ,m→ H

Φ,m/m, uniformly with respect

to h.

Proof. On the level of the amplitudes, the estimates are the same as before,so we only have to check that along Γc(x), we have

|e1h

(i(x−y)·θ−Φ(x)+Φ(y))| ≤ e−c2|x−y|2〈x−y〉 (12.41)

Along Γc(x), we have using notation from R2n ' Cn:

Re (i(x− y) · θ) = 〈∇Φ(x+ y

2), x− y〉 − c |x− y|

2

〈x− y〉, (12.42)

while Taylor’s formula with integral remainder gives:

Φ(x)− Φ(y)− 〈∇Φ(x+ y

2), x− y〉 = (12.43)

〈∫ 1

0(1− t)(Φ′′(x+ y

2+ t

x− y2

)− Φ′′(x+ y

2− tx− y

2))dt

x− y2

,x− y

2〉

= 〈∫ 1

0(1− t)(f ′′(x+ y

2+ t

x− y2

)− f ′′(x+ y

2− tx− y

2))dt

x− y2

,x− y

2〉.

Since the support of f belongs to some fixed compact set, we see that∫(1− t)(f ′′(x+ y

2+ t

x− y2

)−f ′′(x+ y

2− tx− y

2))dt = O(1)‖f ′′‖L∞/〈x−y〉.

Combining this with (12.42), (12.43), we get

Re (i(x− y) · θ)− Φ(x) + Φ(y) = −(c+O(1)‖f ′′‖L∞)|x− y|2

〈x− y〉, (12.44)

which gives (12.41), when ‖f ′′‖L∞ is small enough. #

144

Although not necessary in the following, we end this section with a briefdiscussion of other quantizations. (See [78] for more details on the holomor-phic side and [21] for the real case.) For t ∈ [0, 1], put (formally):

Oph,t(a)u(x) =1

(2πh)n

∫∫Γt(x)

ei(x−y)·θ/ha(tx+ (1− t)y, θ)u(y)dydθ, (12.45)

where Γt(x): θ = 2i∂Φ∂x

(tx+ (1− t)y). Along Γt we get

−Φ(x) + Re (i(x− y) · θ) + Φ(y) = (2t− 1)Φ(x− y).

Hence,

Proposition 12.8 Let Φ be strictly convex and let a ∈ S(ΛΦ,m). ThenOph,t(a) is a well-defined and uniformly bounded operator: HΦ,m → HΦ,m/m

for 0 ≤ t ≤ 12. When 0 ≤ t < 1

2it is enough to assume that a = O(m(x)) to

have the same conclusion, and when t = 0 we even get a uniformly boundedoperator: L2

Φ,m→ HΦ,m/m.

Next we are interested in describing the same operator with differentquantizations. This is very classical in the real case, when Oph,t is well-defined for 0 ≤ t ≤ 1. If at ∈ S(R2n,m) depends smoothly on t, we get foru ∈ S:

∂tOph,t(at)u(x) (12.46)

=1

(2πh)n

∫∫eih

(x−y)·θ∂t(at(tx+ (1− t)y, θ))u(y)dydθ

=1

(2πh)n

∫∫eih

(x−y)·θ(∂tat)(tx+ (1− t)y, θ)u(y)dydθ +

1

(2πh)n

∫∫eih

(x−y)·θ(x− y) · (∂xat)(tx+ (1− t)y, θ)u(y)dydθ,

where the last term equals

1

(2πh)n

∫∫(h

i∂θ)(e

ih

(x−y)·θ) · ∂xat(tx+ (1− t)y, θ)u(y)dydθ = (12.47)

− 1

(2πh)n

∫∫eih

(x−y)·θ(h

i∂θ · ∂xat)(tx+ (1− t)y, θ)u(y)dydθ.

We conclude that Oph,t(at) is independent of t if at fulfills the Schrodingertype equation

∂tat =h

i(∂θ · ∂x)at. (12.48)

145

The Cauchy problem for this equation is well-posed in the sense that if as =at=s is given in S(R2n,m) for some s ∈ [0, 1], then we get a unique solutionat ∈ C∞([0, 1];S(R2n,m)).

The same discussion can be carried out in the complex framework, exceptthat (12.48) is now of heat equation type (on ΛΦ) and can be solved only inthe forward time direction. Thus if as ∈ S(ΛΦ,m) is given, we can find at ∈C∞([s, 1];S(ΛΦ,m)) which is entire for t > s such that Oph,s(as) = Oph,t(at)for t ≥ s. Notice that Oph,0 maps weighted L2-spaces into weighted L2-spacesof holomorphic functions.

We point out that the strict convexity condition on Φ in the propositiondissappears if we choose to represent our operators as in (12.26). The caset = 0 gives Toplitz operators.

The above calculations show that if a ∈ S(ΛΦ,m) then for s, t ∈ [0, 1] wehave

Oph,t(a)−Oph,s(a) = O(h) : HΦ,m → HΦ,m/m. (12.49)

If we also know that a has a holomorphic extension of class S(m) to a tubularneighborhood of ΛΦ, then we can replace Φ by Φ = Φ+f in the above result,provided that the C1,1-norm of f is suffiently small.

12.4 Approximation by multiplication operators

The main result of this section is

Proposition 12.9 Let a be holomorphic and of class S(m) in a tubularneighborhood of ΛΦ. Let Φ ∈ C1,1 be a small perturbation of Φ as before.

Then with ξ(x) = 2i∂Φ∂x

(x):

Oph(a)u(x) = a(x, ξ(x))u(x) +n∑1

(∂ξja(x, ξ(x)))(hDxj − ξj(x))u(x) +Ru(x),

(12.50)where R is of norm O(h): H

Φ,m→ L2

Φ,m/m.

Proof. We write Oph(a)u(x) in (12.38) with the contour (12.40). We alsoassume that m = m = 1 for simplicty. By Taylor expansion, we get

a(x+ y

2, θ) = a(x, ξ(x)) + (∂ξa)(x, ξ(x)) · (θ − ξ(x)) (12.51)

−1

2(∂xa)(x, ξ(x)) · (x− y) +O(|x− y|2) + |θ − ξ(x)|2).

146

Here the third term to the right gives no contribution to (12.38):∫∫eih

(x−y)·θ(x− y) · (∂xa)(x, ξ(x))u(y)dydθ =

(∂xa(x, ξ(x)) ·∫∫ h

i∂θ(e

ih

(x−y)·θu(y))dydθ = 0.

The first two terms to the right in (12.51) lead to the corresponding termsin (12.50). Along the contour (12.40), we have θ− ξ(x) = O(|x− y|), so theremainder term in (12.51) gives contribution O(|x− y|2) in (12.38) and thatgives an operator R with norm O(h). #

We use the above result to study certain scalar products.

Proposition 12.10 Let a be as before and let ψ(x) be locally Lipschitz on Cn

with ψ(x), |∇ψ(x)| = O(mψ(x)), where mψ together with m1,m2 are orderfunctions with mψm ≤ m1m2. Then

(ψOph(a)u|v)L2

Φ

=∫ψ(x)a(x, ξ(x))u(x)v(x)e−2Φ(x)/hL(dx) (12.52)

+O(h)‖u‖HΦ,m1

‖v‖HΦ,m2

for u ∈ HΦ,m1

, v ∈ HΦ,m2

.

That the action of a pseudodifferential operator can be approximatedby a multiplication operator with an error O(h1/2) after a suitable integraltransform was used in [79] and in [19]. That the error becomes only O(h) forscalar products was observed in [19] and used there to give a short proof of thesharp Garding inequality. Here we are in a more general situation, and thelimited regularity in the weights is quite essential for the later applications.(We follow [74].)

Proof. Assume mψ = m = m1 = m2 = 1 for simplicity. Substituting (12.50)into (12.52), we see that we only have to estimate∫

ψ(x)(∂ξja)(x, ξ(x))(hDxj − ξj(x))u(x)v(x)e−2Φ(x)/hL(dx). (12.53)

Here (∂ξja)(x, ξ(x)) is Lipshitz, so we can make an integration by parts in(12.53) and get O(h)‖u‖H

Φ‖v‖H

Φplus∫

ψ(x)(∂ξja)(x, ξ(x))u(x)v(x)(−hDxj − ξj(x))(e−2Φ(x)/h)L(dx), (12.54)

147

which vanishes, since (−hDxj − ξj(x))(e−2Φ(x)/h) = 0. We also used that

hDxjv(x) = 0 since v is holomorphic. #This can be iterated. Let b be second symbol like a with associated order

functions mb and write m = ma for the order function associated to a.

Proposition 12.11 Let a, b, ψ,ma,mb,mψ,m1,m2 be as above with mψmamb ≤m1m2. Then,

(ψOph(a)u|Oph(b)v)L2

Φ

= (12.55)∫ψ(x)a(x, ξ(x))u(x)b(x, ξ(x))v(x)e−2Φ(x)/hL(dx) +O(h)‖u‖H

Φ,m1

‖v‖HΦ,m2

,

for u ∈ HΦ,m1

, v ∈ HΦ,m2

.

Corollary 12.12 Under the appropriate part of the same assumptions, wehave

‖Oph(a)u‖2H

Φ= ‖a(x, ξ(x))u(x)‖2

L2

Φ

+O(h)‖u‖2H

Φ,ma

,

for u ∈ HΦ,ma

.

12.5 Absence of resonances in the analytic non-trappingcase

LetP =

∑|α|≤2

aα(x;h)(hDx)α : H2(Rn)→ L2(R2) (12.56)

satisfy the general assumptions for “Q” in Chapter 7, that we here recall:

aα(x;h) is independent of h for |α| = 2,

aα(·;h) is bounded in C∞b , uniformly with repect to h, and aα(·;h) = aα,0(·)+O(h) in this space, where aα,0 is independent of h .∑|α|=2 aα(x)ξα ≥ 1

C|ξ|2.∑

|α|≤2 aα(x;h)ξα → ξ2, |x| → ∞, uniformly with respect to h.

In addition to the assumption that the coefficients extend holomorphicallyto some angle, we assume them to be analytic everywhere. Thus we assume

(A) There exists a constant C > 0 such that the coefficients aα extendholomorphically in x to x ∈ Cn; |Imx| < C−1〈Rex〉, and the relevantparts of the preceding assumptions extend to this complex domain.

148

Recall that if θ > 0 is small enough, then the resonances of P in e−i[0,2θ[]0,+∞[are defined to be the eigenvalues in this sector of P|eiθRn .

Let p(x, ξ) =∑|α|≤2 aα,0(x)ξα and make the non-trapping assumption

(NT) There are no trapped Hp-trajectories in p−1(1).

By a trapped trajectory we mean a (maximal) integral curve R 3 t 7→exp(tHp)(ρ0) ∈ p−1(1) such that | exp(tHp)(ρ0)| 6→ ∞ both when t → −∞and when t→ +∞. According to section 12.1, this means that “K” there isempty, if “ε” there is small enough. According to Proposition 12.5,, we canfind an “escape” function

G(x, ξ) + f(x, ξ), (12.57)

with f ∈ C∞0 (R2n; R), such that

HpG > 0 on p−1(1). (12.58)

Consider the IR-manifold:

ΛεG : Im (x, ξ) = εHG(Re (x, ξ)), 0 < ε 1. (12.59)

(We leave as an exercise to verify that this is really an IR-manifold.)Outside a bounded set, we have HG = x ·δx−ξ ·∂ξ and here ΛεG becomes:

Imx = εRex, Im ξ = −εRe ξ. (12.60)

Let Γε ⊂ Cn be given byImx = εRex, (12.61)

so that Γε = eiθ(ε)Rn, with θ(ε) = arctg ε. Then the projection of ΛεG \(compact set) is equal to Γε and we claim that ΛεG coincides with T ∗Γεoutside a bounded set. (We view T ∗Γε as a subset of C2n

x,ξ as in Chapter 7.)Indeed for (x, ξ) ∈ ΛεG \ (compact set), we have Im ξ = −εRe ξ in additionto (12.61) and hence

ξ · dx|Γε = (1− iε)Re ξ · (1 + iε)dRex = (1 + ε2)Re ξ · dRex

is real. Also notice that the parametrization κ : Rn 3 y 7→ (1+ iε)y = x ∈ Γεinduces an identification of T ∗Rn = R2n

x,ξ and T ∗Γε via the map

(y, η) 7→ (κ(y), tdκ(y)−1η) = ((1 + iε)y, (1 + iε)−1η).

149

Next parametrize ΛεG by R2n 3 (x, ξ) 7→ (x, ξ) + iεHG(x, ξ) and computepε = p|ΛεG by Taylor expansion:

pε ' p((x, ξ) + iεHG(x, ξ)) = p(x, ξ)− iεHpG(x, ξ) +O(ε2〈ξ〉2). (12.62)

This estimate is uniform since

HG = O(〈x〉) ∂∂x

+O(〈ξ〉) ∂∂ξ, ∂αx∂

βξG = O(〈x〉1−|α|〈ξ〉1−|β|)

∂αx∂βξ p = O(〈ξ〉2−|β|〈x〉−|α|),

and we can work in boxes |x−x0| < 12〈x0〉, |ξ−ξ0| < 1

2〈ξ0〉, using the rescaled

variables (y, η), given by

x = x0 + 〈x0〉y, ξ = ξ0 + 〈ξ0〉η,

for which

∂αy ∂βηHG = O(1)

∂y+O(1)

∂η, ∂αy ∂

βη p = O(〈ξ0〉2).

Using this together with (12.58), we see that for ε > 0 small enough, we havewith the parametrization used in (12.62):

Re pε(x, ξ) = p(x, ξ)+O(ε〈ξ〉2), Im pε(x, ξ) ≤ −ε

C+Cε|p(x, ξ)−1|. (12.63)

In particular, we have

|pε(x, ξ)− 1| ≥ ε〈ξ〉2/C. (12.64)

Let T : L2(Rn) → HΦ00(Cn) be a Bargman transform as in (12.14). Re-

placing the integration contour Rn there by Γε, we get a new Bargman tran-form

T : L2(Γε)→ HΦ0ε(Cn), (12.65)

by identifying Γε with Rn in the natural way. Viewing Γε ⊂ Cn and T ∗Γε ⊂C2n as subspaces in the natural way, we justify the writing “T” rather than“Tε” and we have an associated canonical tranformation

κT : C2n → C2n with κT (T ∗Γε) = ΛΦ0ε, (12.66)

150

where Φ0ε = supy∈Γε −Imφ(x, y). Clearly Φ0

ε(x) = Φ00(x) + O(ε)|x|2 is a

quadratic form which is strictly plurisubharmonic.Now recall that Λε = ΛεG coincides with T ∗Γε outside a bounded set and

that both manifolds are ε-perturbations of R2n in the natural sense. ThenκT (Λε) is an IR-manifold, equal to ΛΦ0

εoutside a bounded set and ε-close to

that manifold everywhere, so

κT (Λε) = ΛΦε , (12.67)

where Φε(x) = Φ0ε(x) + εg(x, ε), where g is C∞ in both arguments and has

its x-support in a fixed compact set.

Theorem 12.13 Under the assumptions above, and in particular (A) and(NT), there exist constants h0, C0 > 0 such that P has no resonances in thedisc D(1, C−1

0 ), for 0 < h < h0.

Corollary 12.14 Let V ∈ C∞(Rn; R) extend to a holomorphic function in|Imx| < 〈Re 〉/C which tends to 0 when |x| → ∞ in that domain. Thenthere exist C0, C1 > 0, such that P = −∆ + V (x) has no resonances inz ∈ C; Re z > C1, Im z > −Re z/C0.

Proof of the Corollary. It will follow from the proof of the theorem,that we we have uniformity with respect to additional parameters if theassumptions hold uniformly.

Consider P − z with Re z 1, |Im z| < Re z/C0 and write this operatoras

(Re z)(−h2∆ +1

Re zV − 1− iIm z

Re z), h2 =

1

Re z.

Theorem 12.13 implies that we have no resonances when Re z is large andIm z/Re z is small. #

Proof of the theorem. Let ε > 0 be sufficiently small but fixed. Let

Q = T P|Γε T−1. (12.68)

Then (using the same letters for symbols and operators)

Q = Oph(Q) = Qw(x, hDx;h), (12.69)

where on the symbol level,

(Q κT )(x, ξ;h) = P (x, ξ;h). (12.70)

151

Here P denote the Weyl-symbol of our operator P which is holomorphic ina tubular neighborhood of T ∗Γε and satisfies P (x, ξ;h) = p(x, ξ) +O(h〈ξ〉2)there. (See for instance [21].) We see that Q(·;h) is holomorphic in a tubularneighborhood of ΛΦ0

εof the form ΛΦ0

ε+W , with W independent of ε, and in

this neighborhood we have

Q(x, ξ;h) = O(m(x, ξ)), Q = q +O(hm), (12.71)

where m is the order function with mκT = 〈ξ〉2, and q is given by qκT = p.For ε > 0 small enough, ΛΦε is well inside ΛΦ0

ε+W , so we can consider

Qw : HΦε,m → HΦε . (12.72)

Let z ∈ D(0, C−10 ) with C−1

0 ε (ε being fixed). Then

|q(x, ξ)− z| ∼ m(x, ξ) on ΛΦε , (12.73)

and Corollary 12.12 together with (12.71) imply that

‖(Qw − z)u‖2HΦε≥ (1−Oε(h))

Cε‖u‖2

HΦε ,m. (12.74)

If z is a resonance, let v ∈ H2(Γε) be a corresponding eigenfunction of P|Γε , sothat (P−z)v = 0 on Γε. Applying T we get u := Tv ∈ HΦε,m, (Qw−z)Tv = 0and (12.74) implies that if h > 0 is small enough, then u = Tv = 0 and sov = 0 in contradiction with the assumption that z is a resonance. #

12.6 Dynamical upper bounds in terms of escape func-tions

We drop the non-trapping assumtion (NT), and keep all the other assump-tions of the preceding section (including (A)). Let K be the union of trappedtrajectories in Σε0

p = p−1([1− ε0, 1 + ε0]), where ε0 > 0 is small and fixed. LetG be a real-valued function on R2n, such that

G and HpG are of class C1,1, G = x · ξ for |(x, ξ)| large, (12.75)

HpG is ≥ O(1)−1|∇G|2 locally uniformly in Σε0p ,

HpG > 0 outside a small neighborhood of K in Σε0p .

152

In section 12.1 we saw that such smooth functions always exist, but ourestimates will depend on a specific such function, and good choices of suchfunctions seem to require limited regularity as we shall see in section 12.7.

For 0 ≤ δ ≤ r ≤ 1, δ ≤ C−10 , let M(r, δ) be the number of resonances of

P in D(1 + ir, r + δ). Let W be a small neighborhood of K in Σε0p , outside

which HpG is bounded from below by some positive constant. Put

V (r, δ) = Vol (ρ ∈ W ; (p(ρ)− 1)2 + 2HpG+ (HpG)2 ≤ 2rδ + δ2) (12.76)

= Vol (ρ ∈ W ; |(p(ρ)− iHpG)− (1 + ir)|2 ≤ (r + δ)2).

Notice that this quantity increases when r or δ increases. Also, if

V (r, δ) = Vol (ρ ∈ W ; (p(ρ)− 1)2 +HpG ≤ rδ),

then V (r, δ) ≤ V (r, 3δ), V (r, δ) ≤ V (r, 3δ).

Theorem 12.15 For C1, C0 > 0 sufficiently large and for C0h ≤ δ ≤ r ≤ 1,δ ≤ 1/C0, rδ ≥ C0h, we have

M(r, δ) ≤ CV (r, Cδ)h−n. (12.77)

Define Γε as in (12.61) and let T : L2(Γε) → HΦ0ε(Cn) be a Bargman

transform as in section 12.5. As there, we define Φε, so that

κT (ΛεG) = ΛΦε , κT (T ∗Γε) = ΛΦ0ε, Φε = Φ0

ε + εg(·, ε), (12.78)

where g(·, ε) is bounded in C1,1 when ε varies, and has uniformly compactsupport. (We recall the definition of ΛεG below.)

Having already fixed a small ε0 > 0, we will let ε > 0 be small and (later)fixed. Let m be the order function with m κT = 〈ξ〉2.

By abuse of notation, we let P also denote the conjugated operatorTP|ΓεT

−1. Then (on the transform side)

P = O(1) : HΦ0ε ,m→ HΦ0

ε, HΦε,m → HΦε ,

uniformly in ε, h.Consider the Hermitian form:

q(u, u) = ((P − 1− ir)u|(P − 1− ir)u)HΦε, 0 < r ≤ 1, (12.79)

formally equal to (Qu|u)HΦε, with Q = (P − 1 − ir)∗(P − 1 − ir), where ∗

indicates that we take the adjoint in the Hilbert space HΦε .

153

Let ΛεG be the the image of

R2n 3 (x, ξ) 7→ (x, ξ) + iεHG(x, ξ), (12.80)

and identify R2n and ΛεG by means of this map. We also identify ΛεG andΛΦε by means of κT , and ΛΦε and Cn

x by means of the natural projectionπx : (x, ξ) 7→ x. Consequently we shall view pε := p|Λε also as a function on

R2n, ΛΦε or Cn whenever convenient.According to Corollary 12.12, we have

q(u, u) =∫Cn|pε − 1− ir|2|u|2e−2Φε/hL(dx) +O(h)‖u‖2

HΦε ,m(12.81)

(where pε is viewed as a function on Cn).If we now consider pε as a function on R2n, we know that the values of

pε on the complement of Σε0p and on the complement of W in Σε0

p avoid a set

of the form z; |Re z − 1| < C−10 , Im z > −ε/C0 and consequently we have

here|pε − 1− ir|2 ≥ (r +

ε

C0

)2, (12.82)

when ε > 0 is small enough. In the complement of Σε0p , we even have

|pε − 1− ir|2 ≥ r2 +〈ξ〉2

O(1). (12.83)

In a neighborhood of K in Σε0p , we take a closer look:

pε ' p((x, ξ)+iHG(x, ξ)) = p(x, ξ)−iεHpG(x, ξ)+O(ε2)|∇G|2+iO(ε3)|∇G|3,(12.84)

where the first remainder term is real and the second one is purely imaginary.Hence, using (12.75):

−Im pε = ε(1 +O(ε2))HpG, (12.85)

Re pε = p+O(ε2)HpG = p+O(ε)Im pε. (12.86)

It follows that near K in Σε0p , we have

|pε − 1− ir|2 = (Re pε − 1)2 + (r − Im pε)2

= (p− 1 +O(ε)Im pε)2 + (r − Im pε)

2

≥ 1

2(p− 1)2 −O(ε2)(−Im pε)

2 + r2 + 2r(−Im pε) + (−Im pε)2

≥ r2 +1

2(p− 1)2 + 2r(−Im pε) +

1

2(−Im pε)

2 :

154

|pε − 1− ir|2 ≥ r2 +1

2(p− 1)2 + 2r(−Im pε) +

1

2(−Im pε)

2, (12.87)

near K in Σε0p .

Using these estimates and a variant of (12.81), we shall first prove:

Proposition 12.16 For 0 ≤ r ≤ 1, we have

q(u, u) ≥ (r2 − Crh)‖u‖2HΦε

. (12.88)

Proof. We drop the subscript HΦε for the corresponding norms and scalarproducts. Consider,

q(u, u)− r2‖u‖2 = ((P − 1− ir)u|(P − 1− ir)u)− r2‖u‖2 (12.89)

= ‖(P − 1)u‖2 + ir[((P − 1)u|u)− (u|(P − 1)u)]

≥ r(‖(P − 1)u‖2 − 2Im (Pu|u))

= r(∫

(|pε − 1|2 − 2Im pε)|u|2e−2Φε/hL(dx) +O(h)‖u‖2HΦε,m

).

From (12.85), (12.86) it is clear that |pε − 1|2 − 2Im pε is ≥ 0 in Σε0p and

≥ O(1)−1〈ξ〉2 in the complement of Σε0p . Using this in (12.89), we get q(u, u)−

r2‖u‖2 ≥ −Crh‖u‖2. #

We want to make finite rank perturbations of q for which the lower bound(12.88) improves. Fix a sufficiently small ε > 0 for which the previous esti-mates hold. We start by combining (12.82), (12.83), (12.85), (12.87) to seethat for δ > 0 small enough, we have

|pε − 1− ir|2 ≥ r2 + rδ in R2n \WCδ,r, (12.90)

where

WCδ,r = ρ ∈ R2n; (p(ρ)−1)2 + 2rHpG+ (HpG)2 < 2Crδ+ (Cδ)2, (12.91)

and C > 0 is sufficiently large and independent of δ. (Recall that ε > 0 hasnow been fixed.)

Using this in (12.81), we get

q(u, u) ≥ (r2 + rδ −O(h))‖u‖2 − (rδ +O(h))‖1WCδ,ru‖2

L2Φ, (12.92)

where Φ = Φε.

155

We shall next show that the last term in (12.92) is very small on certainsubspaces of HΦ of finite and explicitly controled codimension. Let χ0 ∈C∞0 (B(0, 1)) (where B(0, 1) is the open unit ball in Cn) be a radial function

with∫χ0(x)L(dx) = 1. Let ξ(x) = 2

i∂Φ(x)∂x

. Using the mean-value propertyfor the holomorphic function y 7→ ei(x−y)·ξ(x)/hu(y), we get the representationof the identity operator on HΦ:

Iu(x)(= u(x)) = h−n∫eih

(x−y)·ξ(x)χ0(x− yh1/2

)u(y)L(dy). (12.93)

Let

K(x, y;h) = eih

(x−y)·ξ(x)χ0(x− yh1/2

)h−n.

Then the reduced kernel e−(Φ(x)−Φ(y))/hK(x, y;h) is O(1)h−nχ0( x−yh1/2 ) and we

deduce that the last member of (12.93) defines a bounded operator: L2Φ → L2

Φ

of norm O(1) uniformly with respect to h.Let x be close to x. We shall compare the kernel K(x, y;h) with the rank

1 kernel

Kx(x, y;h) = ei(x−y)·ξ(x)/hχ0(x− yh1/2

)h−n.

If |x − x| ≤ εh1/2 (with a new ε > 0, the earlier one being now fixed andforgotten), we check that the difference betwen the corresponding reducedkernels is O(εh−n) and has its support in a domain |x− y| ≤ Const. h1/2.

Now let W ⊂⊂ Cn be a compact set which can be covered by M(W, εh1/2)

balls, Bj = B(xj, εh1/2). Let W = ∪Mj=1Wj be a partition of W with Wj ⊂ Bj,Wj ∩Wk = ∅ for j 6= k. For x ∈ Wj, we put x(x) = xj. Then W 3 x 7→ x(x)

takes at most M(W, εh1/2) values and |x− x(x)| ≤ εh1/2. For x 6∈ W , we putχu(x) = 0 and for x ∈ W , we put

χu(x) = h−n∫eih

(x−y)·ξ(x(x))χ(x(x)− yh1/2

)u(y)L(dy).

Then the discussion above gives:

Lemma 12.17 χ : L2Φ → L2

Φ is of finite rank ≤ M(W, εh1/2) and

‖1Wu− χu‖L2Φ≤ Cε‖u‖Φ, u ∈ HΦ(Cn). (12.94)

We next estimate M(W, r):

156

Lemma 12.18 Let W ⊂ Cn be compact and let M(W, r) be the minimalnumber of closed balls of radius r > 0, centered at points of W , that coverW . Then

Vol (W )r−2n ≤ Vol (B(0, 1))M(W, r) ≤ 22nVol (W +B(0,r

2))r−2n. (12.95)

Proof. Clearly

Vol (W ) ≤ M(W, r)Vol (B(0, r)) = M(W, r)Vol (B(0, 1))r2n,

which gives the first inequality in (12.95). To prove the second inequality,let A(r) be the maximum number of closed disjoint balls of radius r/2 thatcan be centered at points of W , and let B(xj, r/2), j = 1, ..., A(r) be such afamily of balls. Then,

A(r)Vol (B(0,r

2)) ≤ Vol (W +B(0,

r

2)),

so thatA(r)Vol (B(0, 1)) ≤ 22nr−2nVol (W +B(0,

r

2)). (12.96)

The second part of (12.95) will then follow, if we prove that M(W, r) ≤ A(r),and to do so, it is enough to prove that the B(xj, r) cover W . If that werenot the case, there would be a point x ∈ W which is not in any of theB(xj, r) and hence B(x, r/2) ∩B(xj, r/2) = ∅ for every j, which contradictsthe maximality of the disjoint family B(xj, r/2). #

Lemma 12.19 Let q ≥ 0 be a C1,1-function on Cn with lim inf |x|→∞ q > 0.If λ0 > 0 is sufficiently small, there exists a constant C1 > 0 such that forh ≤ λ ≤ λ0, we have for 0 ≤ ε ≤ 1:

Vol (Wλ +B(0, εh1/2)) ≤ Vol (Wλ+C1εh1/2λ1/2), (12.97)

where Wλ = x; q(x) < λ.

Proof. We prove the corresponding inclusions for the sets. If x ∈ Kλ, wehave q(x) ≤ λ and |q′(x)| ≤ Cλ1/2, so if |y| ≤ εh1/2, then

q(x+ y) ≤ λ+ C(λ1/2εh1/2 + ε2h) ≤ λ+ C1εh1/2λ1/2.

#

157

We now return to (12.92) and apply the last three lemmas with q =(p(ρ)− 1)2 + 2rHpG + (HpG)2 in (12.91). We conclude that there exists anoperator χr,δ = O(1) : L2

Φ → L2Φ of rank

≤ Vol (WCδ,r +B(0, εh1/2))ε−2nh−n

≤ Vol (WCδ+C2δ1/2r−1/2εh1/2,r)ε−2nh−n,

when rδ ≥ C0h, such that

‖(1WCδ,r− χ)u‖ ≤ O(ε)‖u‖, u ∈ HΦ.

Choose ε > 0, so that the last “O(ε)” is ≤ 1/3 and use it in (12.92):

q(u, u) ≥ (r2 +2rδ

3−O(h))‖u‖2 − (rδ +O(h))‖χu‖2, u ∈ HΦ,m. (12.98)

Notice thatrank (χ) ≤ CV (r, Cδ)h−n, (12.99)

for a new sufficiently large constant C > 0.

Proof of Theorem 12.15. Consider still the Hermitian form (12.79):

q(u, u) = ((P − 1− ir)u|(P − 1− ir)u)HΦ, u ∈ HΦ,m.

From (12.81) it follows that q is a closed quadratic form with domain HΦ,m.Consequently, by the Lax-Milgram theorem, there exists a unique self-adjointoperator Q ≥ 0 in HΦ such that

D(Q) ⊂ HΦ,m, (12.100)

q(u, u) = (Qu|u), u ∈ D(Q). (12.101)

Formally, Q = (P −1− ir)∗(P −1− ir). Let µ1 ≤ µ2 ≤ ... be an enumerationof first all the discrete eigenvalues of Q1/2 below the essential spectrum,repeated according to their multiplicity, and then if there are only finitelymany such eigenvalues, the infimum of the essential spectrum repeated “adeternam”. Then we have the max-min formula

µ2N = sup

E is a subspace of HΦof codimension≤N−1

infu∈E‖u‖=1

q(u, u), (12.102)

with the convention that q(u, u) = +∞ if u ∈ HΦ \HΦ,m.

158

Proposition 12.16 tells us that

µ1 ≥√r2 − Crh ≥ r −O(h), (12.103)

where we now restrict all parameters as in Theorem 12.15. On the otherhand, (12.98), (12.99) show that if E = N (χ) (the null-space of χ), then

infu∈E‖u‖=1

q(u, u) ≥ r2 +2rδ

3−O(h),

and since codimN (χ) ≤ CV (r, Cδ)h−n, we conclude from (12.102), that

µ2N ≥ r2 +

2rδ

3−O(h), for N ≥ CV (r, Cδ)h−n. (12.104)

Multiplying δ by a constant (affecting the “C” above), we may assume that,

µN ≥ r + 2δ, for N ≥ CV (r, Cδ)h−n. (12.105)

Let z1, z2, ... be the resonances of P in D(1 + ir, r + 2/C0) repeated ac-cording to their multiplicity and arranged so that j 7→ |zj − 1− ir|. Then wehave the Weyl inequalities:

µ1µ2...µN ≤N∏1

|zj − 1− ir|, (12.106)

viewing zj − 1 − ir as the eigenvalues of P − 1 − ir. Notice that theseinequalities are “opposite” to the corresponding ones for a compact operator,where the singular values are enumerated in decreasing order and similarlyfor the moduli of the eigenvalues. See the appendix (dans la prochaine versionc’est promis!) for a proof.

Let M(r, δ) be as in the theorem. If M(r, δ) ≤ CV (r, Cδ)h−n, we aredone. If not, we apply (12.106) with N = M(r, δ) and get from (12.103),(12.105), with N = [CV (r, Cδ)h−n]:

(r − Ch)N(r + 2δ)M−N ≤ (r + δ)M ,

(r + 2δ

r + δ)M ≤ (

r + 2δ

r − Ch)N ,

M log (r + 2δ

r + δ) ≤ N log (

r + 2δ

r − Ch).

159

Here the logarithms are of the order of magnitude δ/r, so we get M ≤ O(1)N .#

By replacing the energy “1” be a real energy E which varies in a neigh-borhood of E, one obtains a variant of the main theorem. If J ⊂ neigh (1,R)is an interval and 0 < ε 1, let

R(J, ε) = ρ ∈ W ; p(ρ) ∈ J, HpG ≤ ε. (12.107)

Then we have

Theorem 12.20 We make the same assumptions as in Theorem 12.15. Thenthere are constants C0, C > 0 such that for every subinterval I of [1−ε0/2, 1+ε0/2], we have when C0h ≤ δ ≤ C−1

0 ,

M3(I, δ) ≤ CVol (R(I + [−Cδ, Cδ], Cδ))h−n, (12.108)

where δ = max(δ, h1/2). Here M3(I, δ) denotes the number of resonances inthe rectangle I − i[0, δ[.

In a later version of these notes we will put in the detailed proof, butin this version we only give the outline: Cover I − i[0, δ[ by a “minimal”number of discs D(Ej + ir, r + 2δ) and remember the constraints 1 ≥ r ≥ δ,rδ ≥ C0h. If δ ≥ h1/2, we take r ∼ δ and if C0h ≤ δ ≤ h1/2, we taker ∼ h/δ. Then estimate the number of resonances in D(Ej + ir, r + 2δ) byCVol (ρ ∈ W ; p − iHpG ∈ D(Ej + ir, r + Cδ). When summing over j wenotice that every point in the lower half plane can belong to at most a fixedfinite number of the discs D(Ej + it, r + Cδ). #

12.7 Quick review of the case when the classical flowis hyperbolic

Recall that K,Γ± were defined in Σε0p for some small fixed ε0 > 0. Let

K, Γ± be the corresponding sets in Σ2ε0p . We make the following hyperbolicity

assumptions about the classical dynamics:

(Hyp 1) In a neighborhood Ωρ0 of every point ρ0 ∈ K, we can represent Γ±as a union of closed disjoint C1-manifolds of dimension n + 1, such that ifρ ∈ Ωρ0 ∩ Γ+ and if E+

ρ = Tρ(Γ+,ρ), where Γ+,ρ is the corresponding leaf,

then E+ρ depends continuously on ρ ∈ Ωρ0 ∩ Γ+, and contains Hp(ρ). Same

assumption about E−ρ = Tρ(Γ−,ρ).

160

(Hyp 2) E+ρ and E−ρ are transversal for every ρ ∈ K, and are indpendent of

the choice of Ωρ0 , containing ρ.

(Hyp 3) There exists a constant C > 0, such that with Φt = exp tHp:

‖dΦt(ν)‖ ≤ Ce−t/C‖ν‖, ν ∈ Tρ(R2n)/E+ρ , ρ ∈ K, t ≥ 0, (12.109)

‖dΦ−t(ν)‖ ≤ Ce−t/C‖ν‖, ν ∈ Tρ(R2n)/E−ρ , ρ ∈ K, t ≥ 0, (12.110)

where dΦt is considered as a map Tρ(R2n)/E±ρ → TΦt(ρ)(R

2n)/E±Φt(ρ).

Theorem 12.21 Under the above assumptions and after an arbitrarily smalldecrease of ε0, we can find G as in Theorem 12.15 such that

HpG ≥1

Cdist (·, K)2, near K in Σε0

p and > 0 in Σε0p \K. (12.111)

Let L ⊂ R2n be bounded and put Lε = ρ ∈ R2n; dist (ρ, L) < ε, forε > 0. We define the Minkowski codimension of L; codim (L), to be thesupremum of all d ≥ 0, such that Vol (Lε) ≤ O(εd) when ε → 0. Clearly0 ≤ codim (L) ≤ 2n when L 6= ∅. We say that L is of pure dimension ifthe supremum is attained, so that Vol (Lε) ≤ O(εcodim (L)). Put dim (L) =2n− codim (L). The following result is a consequence of Theorem 12.20 andTheorem 12.21:

Theorem 12.22 Under the above assumptions, let d = codim (K) if K isof pure codimension, and otherwise let 0 ≤ d < codim K. Then there is aconstant C0 > 0 such that for 0 < h ≤ C−1

0 , C0h ≤ δ ≤ C−10 , the number of

resonances of P in the rectangle ]− ε02, ε0

2[−i[0, δ[ is ≤ C0δ

d/2h−n.

Remains to add to this section: The proof of Theorem 12.21 and a discus-sion of the example of several strictly convex obstacles and the correspondingexample of approximating potentials.

References

[1] J. Aguilar, J.M. Combes, A class of analytic perturbations for one-bodySchrodinger Hamiltonians. Comm. Math. Phys., 22(1971), 269–279.

[2] L. Ahlfors, Complex analysis. An introduction to the theory of analyticfunctions of one complex variable. Third edition. International Seriesin Pure and Applied Mathematics. McGraw-Hill Book Co., New York,1978.

161

[3] A.A. Arsenev, On the resonance properties of the scattering amplitudefor the Schrodinger equation with a trapping potential. Teoret. Mat. Fiz.104(2)(1995), 214–232, and Theoret. and Math. Phys. 104(2)(1995),935–949 (1996).

[4] A.A. Arsenev, The singularities of the analytic continuation and theresonance properties of the solution of a scattering problem for theHelmholtz equation, (Russian) Dokl. Akad. Nauk SSSR 197(1971), 511–512.

[5] E. Balslev, J.M. Combes, Spectral properties of many-body Schrodingeroperators with dilatation-analytic interactions, Comm. Math. Phys.,22(1971), 280–294.

[6] C. Bardos, J.C. Guillot, J. Ralston, La relation de Poisson pourl’equation des ondes dans un ouvert non-borne, Comm. PDE 7(1982),905–958.

[7] C. Bardos, G. Lebeau, J. Rauch, Scattering frequencies and Gevrey 3singularities, Invent. Math. 90 (1987), no. 1, 77–114.

[8] J.T. Beale, Scattering frequencies of resonators, Comm. Pure Appl.Math. 26(1973), 549–563.

[9] A. Lahmar-Benbernou, A. Martinez, Semiclassical asymptotics of theresidues of the scattering matrix for shape resonances, Asymptot. Anal.20(1)(1999), 13–38.

[10] A. Lahmar-Benbernou, A. Martinez, On Helffer–Sjostrand’s theory ofresonances, International Math. Research Notices, 2002, no 13, 697–717.

[11] J.F. Bony, Minoration du nombre de resonances engendrees par unetrajectoire fermee, Comm. PDE.,

[12] P. Briet, J.M. Combes, P. Duclos, On the location of resonances forSchrodinger operators in the semiclassical limit. I. Resonance free do-mains, J. Math. Anal. Appl. 126(1)(1987), 90–99.

162

[13] P. Briet,J.M. Combes, P. Duclos, On the location of resonances forSchrodinger operators in the semiclassical limit. II. Barrier top reso-nances, Comm. Partial Differential Equations 12 (1987), no. 2, 201–222.Erratum in Comm. Partial Differential Equations 13 (1988), no. 3, 377–381.

[14] Russell Brown, P.D. Hislop, A. Martinez, Eigenvalues and resonancesfor domains with tubes: Neumann boundary conditions, J. DifferentialEquations 115(2)(1995), 458–476.

[15] N. Burq, Acta Math. 1998.

[16] N. Burq, Lower bounds for shape resonance widths of long rangeSchrodinger operators,Preprint Universite de Paris Sud, 2000.

[17] N. Burq, J. Sjostrand, In preparation.

[18] J.M. Combes, P. Duclos,M. Klein, R. Seiler, The shape resonance,Comm. Math. Phys. 110(2) (1987), 215–236.

[19] A. Cordoba, C. Fefferman,

[20] E.B. Davies, bok om spektral teori med HS-formeln

[21] M. Dimassi, J. Sjostrand, Spectral asymptotics in the semi-classicallimit, London Math. Soc. Lecture Note Series, 268. Cambridge Uni-versity Press, Cambridge, 1999.

[22] Dynkin,

[23] C.A. Fernandez, Resonances in scattering by a resonator, Indiana Univ.Math. J., 34(1)(1985), 115–125.

[24] C. Fernandez, R. Lavine, Lower bounds for resonance width in potentialand obstacle scattering, Comm. Math. Phys., 128(1990), 263–284.

[25] R. Froese, P. Hislop, ”upper bounds”

[26] C. Gerard, Asymptotique des poles de la matrice de scattering pour deuxobstacles strictement convexes. Mem. Soc. Math. France (N.S.) No. 31,(1988),

163

[27] C. Gerard, A. Martinez, Prolongement meromorphe de la matrice descattering pour des problemes a deux corps a longue portee, Ann. Inst.H. Poincare Phys. Theor. 51(1)(1989), 81–110.

[28] C. Gerard, A. Martinez, D. Robert, Breit-Wigner formulas for the scat-tering phase and the total scattering cross-section in the semi-classicallimit, Comm. Math. Phys. 121(2)(1989), 323–336.

[29] C. Gerard, J. Sjostrand, Semiclassical resonances generated by a closedtrajectory of hyperbolic type, Comm. Math.Phys., 108(1987), 391-421.

[30] C. Gerard, J. Sjostrand, Resonances en limite semiclassique et exposantsde Lyapunov, Comm. Math. Phys. 116(1988), 193-213.

[31] I.C. Gohberg, M.G. Kreın Introduction to the theory of linear non-selfadjoint operators. Translations of Mathematical Monographs, Vol.18 American Mathematical Society, Providence, R.I. 1969.

[32] A. Grigis, J. Sjostrand,

[33] V. Guillemin, R. Melrose, The Poisson summation formula for manifoldswith boundary, Adv. Math. 32(1979), 128–148.

[34] L. Guillope, M. Zworski, Polynomial bounds on the number of resonancesfor some complete spaces of constant negative curvature near infinity,Asympt. Anal. 11(1)(1995), 1–22.

[35] L. Guillope, M. Zworski, “traceformula for hyperbolic surfaces”.

[36] E. Harrel, General lower bounds for resonances in one dimension,Comm. Math. Phys., 86(1982), 221–225.

[37] B. Helffer, A. Martinez, Comparaison entre les diverses notions deresonances, Helv. Phys. Acta 60(8)(1987), 992–1003.

[38] B. Helffer, J. Sjostrand, Resonances en limite semi-classique, Mem. Soc.Math. France (N.S.) No. 24-25, (1986),

[39] B. Helffer, J. Sjostrand, Equation de Schrodinger avec champmagnetique et equation de Harper, Springer Lecture Notes in Physics,345(1989), 118–197.

164

[40] P.D. Hislop, I.M. Sigal, Semiclassical theory of shape resonances in quan-tum mechanics. Mem. Amer. Math. Soc. 78 (1989), no. 399.

[41] P.D. Hislop, A. Martinez, Scattering resonances of a Helmholtz res-onator, Indiana Univ. Math. J. 40(2)(1991), 767–788.

[42] L. Hormander, The analysis of linear partial differential operators, I–IV,Springer Verlag, 1983–1985.

[43] W. Hunziker, Distortion analyticity and molecular resonance curves.Ann. Inst. H. Poincare Phys. Theor. 45(4)(1986), 339–358.

[44] M. Ikawa, On the scattering matrix for two convex obstacles, Hyperbolicequations and related topics (Katata/Kyoto, 1984), 63–84, AcademicPress, Boston, MA, 1986.

[45] M. Ikawa, On scattering poles for several convex bodies, Quantum andchaos: how incompatible? (Kyoto, 1993). Progr. Theoret. Phys. Suppl.No. 116, (1994), 107–119.

[46] A. Intissar, A polynomial bound on the number of the scattering poles fora potential in even-dimensional spaces Rn, Comm. Partial DifferentialEquations 11(4)(1986), 367–396.

[47] G. Jona-Lasinio, C. Presilla, J. Sjostrand, On Schrodinger equationswith concentrated nonlinearities, Ann. Physics 240(1)(1995), 1-21.

[48] N. Kaidi, P. Kerdelhue, Forme normale de Birkhoff et resonances,preprint 1998, Universite de Paris Sud.

[49] A. Knauf, Qualitative aspects of classical potential scattering. Regul.Chaotic Dyn., 4(1)(1999), 3–22.

[50] P. Lax, R. Phillips, Scattering theory, Second edition, Pure and AppliedMathematics, 26. Academic Press, Inc., Boston, MA, 1989.

[51] G. Lebeau, Fonctions harmoniques et spectre singulier, Ann. Sci. EcoleNorm. Sup. (4)13(2)(1980), 269–291.

[52] G. Lebeau, L. Robbiano, Controle exact de l’equation de la chaleur,Comm. Partial Differential Equations 20(1-2)(1995), 335–356.

165

[53] G. Lebeau, L. Robbiano, Stabilization of the wave equation by the bound-ary, Partial differential equations and mathematical physics (Copen-hagen, 1995; Lund, 1995), 207–210, Progr. Nonlinear Differential Equa-tions Appl., 21, Birkhauser Boston, Boston, MA, 1996.

[54] R. Melrose, Polynomial bound on the distribution of poles in scatteringby an obstacle, Proc. Journees e.d.p. St Jean de Monts, 1984, Soc. Math.de France.

[55] R. Melrose, Scattering theory and the trace of the wave group, J. Funct.Anal., 45(1982), 29–40.

[56] R. Melrose, J. Sjostrand, Singularities in boundary value problems, Iand II, Comm. Pure Appl. Math., 31(1978), 593–617, and 35(1982),129–168.

[57] L. Nedelec, Resonances semi-classiques pour l’operateur de Schrodingermatriciel en dimension deux, Ann. Inst. H. Poincare Phys. Theor.65(2)(1996), 129–162.

[58] L. Nedelec, Asymptotique du nombre de resonances de l’operateurde Schrodinger avec potentiel lineaire et matriciel, Math. Res. Lett.4(2,3)(1997), 309–320.

[59] L. Nedelec, Resonances for matrix Schrodinger operators, Duke Math.J., to appear

[60] V. Petkov, Sur la conjecture de Lax et Phillips pour un nombrefini d’obstacles strictement convexes, Seminaire sur les Equations auxDerivees Partielles, 1995–1996, Exp. No. XI, Ecole Polytech., Palaiseau.

[61] V. Petkov, G. Vodev, Upper bounds on the number of scattering polesand the Lax-Phillips conjecture, Asymptotic Anal. 7(2)(1993), 97–104.

[62] V. Petkov, M. Zworski, Breit-Wigner approximation and the distributionof resonances, Comm. Math. Phys. 204(2)(1999), 329–351. Correction,to appear.

[63] G. Popov, On the contribution of degenerate periodic trajectories to thewave-trace, Comm. Math. Phys. 196(2)(1998), 363–383.

166

[64] C. Presilla, J. Sjostrand, Nonlinear resonant tunneling in systems cou-pled to quantum reservoirs, Physical Review B, 55(15)(15 April 1997-I),9310-9313.

[65] C. Presilla, J. Sjostrand, Transport properties in resonant tunneling het-erostructures, J. Math. Physics, 37(10)(1996), 4816-4844.

[66] M. Reed, B. Simon, Methods of modern mathematical physics, I–IV,Academic press, New-York, 1978.

[67] D. Robert, ”semiklassisk spaar-formel foer h-oberoende funktioner aven h-pseudo”

[68] M Rouleux, Resonances for a semi-classical Schrodinger operator neara non-trapping energy level, Publ. Res. Inst. Math. Sci. 34(6)(1998),487–523.

[69] M. Rouleux, Absence of resonances for semiclassical Schrodinger oper-ators with Gevrey coefficients, in preparation

[70] B. Simon, Phys. Lett. A 71 (1979), no. 2–3, 211–214.

[71] J. Sjostrand, Semiclassical resonances generated by a non-degeneratecritical point, Springer Lect. Notes in Math., 1256, 402-429.

[72] J. Sjostrand, Geometric bounds on the density of resonances for semi-classical problems, Duke Math. J., 60(1)(1990), 1-57.

[73] J. Sjostrand, Singularites analytiques microlocales. Asterisque, 95(1982).

[74] J. Sjostrand, Density of resonances for strictly convex analytic obstacles,Canad. J. Math. 48(2)(1996), 397–447.

[75] J. Sjostrand, A traceformula and review of some estimates for res-onances, p. 377-437 dans Microlocal Analysis and Spectral Theory,NATO ASI Series C, vol. 490, Kluwer 1997.

[76] J. Sjostrand, Resonances for bottles and trace formulae. Math. Nachr. aparaitre

[77] J. Sjostrand, A trace formula for resonances and application to semi-classical Schrodinger operators. Seminaire e.d.p., Ecole Polytechnique,1996-97, expose no 2.

167

[78] J. Sjostrand, Function spaces associated to global I-Lagrangian mani-folds., pages 369-423 in Structure of solutions of differential equations,Katata/Kyoto, 1995, World Scientific 1996

[79] J. Sjostrand, Ann Inst. Fourier ≈ 1976.

[80] J. Sjostrand, M. Zworski, Complex scaling and the distribution of scat-tering poles, J. Amer. Math. Soc. 4 (1991), no. 4, 729–769.

[81] J. Sjostrand, M. Zworski, Distribution of scattering poles near the realaxis, Comm. Partial Differential Equations 17(5,6)(1992), 1021–1035.

[82] J. Sjostrand, M. Zworski, Estimates on the number of scattering polesnear the real axis for strictly convex obstacles, Ann. Inst. Fourier,43(3)(1993), 769-790.

[83] J. Sjostrand, M. Zworski, The complex scaling method for scattering bystrictly convex obstacles, Ark. f. Matematik, 33(1)(1995), 135-172.

[84] J. Sjostrand, M. Zworski, Lower bounds on the number of scatteringpoles. Comm. Partial Differential Equations 18(5,6)(1993), 847–857.

[85] J. Sjostrand, M. Zworski, Lower bounds on the number of scatteringpoles. II. J. Funct. Anal. 123(2)(1994), 336–367.

[86] J. Sjostrand, M. Zworski, Asymptotic distribution of resonances for con-vex obstacles, Acta Math. 183(2)(1999), 191–253.

[87] P. Stefanov, Quasimodes and resonances: sharp lower bounds. DukeMath. J. 99(1)(1999), 75–92.

[88] P. Stefanov, G. Vodev, Neumann resonances in linear elasticity for anarbitrary body. Comm. Math. Phys. 176(3)(1996), 645–659.

[89] S.H. Tang, M. Zworski, From quasimodes to reasonances, Math. Res.Lett. 5(3)(1998), 261–272.

[90] S.H. Tang, M. Zworski, ”resonance expansions

[91] B.R. Vaınberg, Asymptotic methods in equations of mathematicalphysics, Gordon & Breach Science Publishers, New York, 1989.

168

[92] G. Vodev, Sharp polynomial bounds on the number of scattering poles formetric perturbations of the Laplacian in Rn, Math. Ann. 291(1)(1991),39–49.

[93] G. Vodev, Sharp bounds on the number of scattering poles for perturba-tions of the Laplacian, Comm. Math. Phys. 146(1)(1992), 205–216.

[94] G. Vodev, On the distribution of scattering poles for perturbations of theLaplacian. Ann. Inst. Fourier (Grenoble) 42(3)(1992), 625–635.

[95] G. Vodev, Sharp bounds on the number of scattering poles in even-dimensional spaces, Duke Math. J. 74(1)(1994), 1–17.

[96] G. Vodev, Sharp bounds on the number of scattering poles in the two-dimensional case, Math. Nachr. 170(1994), 287–297.

[97] G. Vodev, extension du th. de Burq

[98] G. Vodev, extension du th. de Burq

[99] M. Zerzeri, Majoration du nombre de resonances pres de l’axe reel pourune perturbation, a support compact, abstraite, du laplacien, preprint,Universite de Paris Nord, 1999, Comm. PDE, to appear

[100] M. Zworski, Distribution of poles for scattering on the real line J. Funct.Anal. 73(2)(1987), 277–296.

[101] M. Zworski, Sharp polynomial bounds on the number of scattering poles,Duke Math. J. 59(2)(1989), 311–323.

[102] M. Zworski, Poisson formulae for resonances, Sem. Equationsaux Derivees Partielles, 1996–1997, Exp. No. XIII, Ecole Polytech.,Palaiseau, 1997.

[103] M. Zworski, Dimension of the limit set and the density of resonancesfor convex co-compact hyperbolic surfaces, Inv. Math. 136(2)(1999), 353–409.

169