• Aucun résultat trouvé

A note on regularity for free convolutions

N/A
N/A
Protected

Academic year: 2022

Partager "A note on regularity for free convolutions"

Copied!
14
0
0

Texte intégral

(1)

www.elsevier.com/locate/anihpb

A note on regularity for free convolutions

Serban Teodor Belinschi

a,b,1

aMathematics Department, Indiana University, Bloomington, IN 47405, USA

bInstitute of Mathematics of the Romanian Academy, P.O. Box 1-764, RO-014700, Bucharest, Romania

Received 5 January 2004; received in revised form 8 April 2005; accepted 18 May 2005 Available online 7 December 2005

Abstract

Letμν andμν denote the free additive convolution and the free multiplicative convolution, respectively, of the Borel probability measuresμandν. We analyze the boundary behavior of the functionsGμν(z)= 1

ztd(μν)(t)andψμν(z)=

zt

1ztd(μν)(t). We prove that, under certain conditions, these functions extend continuously to the boundary of their natural domains as functions with values in the extended complex planeC∪ {∞}. As a consequence, we obtain thatμν(respectively μν) can never be purely singular, unlessμorνis concentrated in one point.

Crown Copyright©2005 Published by Elsevier SAS. All rights reserved.

Résumé

Soitμνetμνla convolution additive libre et, respectivement, la convolution multiplicative libre des mesures boréliennes de probabilitéμetν. Nous étudions le comportement à la frontière des fonctionsGμν(z)= 1

ztd(μν)(t)etψμν(z)=

zt

1ztd(μν)(t). Nous démontrons que, dans certaines conditions, ces fonctions peuvent être prolongées à la frontière de leurs domaines naturels de définition, comme fonctions continues prenant des valeurs dans la compactificationC∪ {∞}du plan complexeC. Une conséquence de ce fait est queμν(et, respectivement,μν) peut être purement singulière seulement siμou νest concentrée en un point.

Crown Copyright©2005 Published by Elsevier SAS. All rights reserved.

MSC:primary 46L54; secondary 30D40

Keywords:Free convolutions; Cluster sets of analytic functions

1. Introduction

Free convolutions appear as natural analogues of the classical convolutions in the context of free probability theory.

Denote byMthe set of Borel probability measures supported on the real lineR, byM+the ones supported on the interval [0,+∞), and by M the ones supported on the unit circle in the complex planeC having nonzero first moment.

E-mail addresses:sbelinsc@math.uwaterloo.ca, teodor.belinschi@imar.ro (S.T. Belinschi).

1 Present address: Department of Pure Mathematics, University of Waterloo, Waterloo, Ontario N2L 3G1, Canada.

0246-0203/$ – see front matter Crown Copyright©2005 Published by Elsevier SAS. All rights reserved.

doi:10.1016/j.anihpb.2005.05.004

(2)

Forμ, νMone defines the free additive convolutionμν of μ andν in the following way:μν is the probability distribution ofX+Y, whereXandY are free selfadjoint random variables with distributionμandν.

Similarly, given μ, νM, the free multiplicative convolutionμν ofμ andν is defined as the probability distribution ofXY, whereXandY are free unitary random variables with distributionμandν.

Forμ, νM+, one still can defineμν, but in a slightly different way:μνis the distribution ofX1/2Y X1/2, whereXandY are free positive random variables with distributionsμandν. (We refer to [16] for an introduction to the area of free probability theory.)

LetC+= {z∈C: z >0}denote the upper half of the complex plane,D= {z∈C: |z|<1}the open unit disk, and letT= {z∈C: |z| =1}denote the unit circle in the complex plane.

For anyμM, let Gμ(z)=

−∞

1

ztdμ(t ), z=0,

denote the Cauchy transform ofμ, and letFμ(z)=Gμ1(z). By results of [5] (see Corollary 5.8), there exists a number M0, depending onμ, such thatFμ has a right inverseFμ1defined on {x+iy: y > M,|x|< y}; the function φμ(z)=Fμ1(z)zhas the remarkable property that

φμ(z)+φν(z)=φμν(z) (1)

forz∈ {x+iy: y > M,|x|< y}, withMlarge enough.

Similar results have been proved for free multiplicative convolutions. ForμM+(orμM), let ψμ(z)=

V

zt

1−ztdμ(t ),

wherezbelongs either to the open unit diskD(ifμM, in which caseV =T), or toC\ [0,∞)(ifμM+, in which caseV = [0,+∞)). The functionψμdetermines uniquely the measureμ, and it is univalent in a neighborhood of zero (ifμM) or in the left half-plane iC+= {z∈C: z <0}(ifμM+).

Letημ(z)=ψμ(z)/(1+ψμ(z)). The functionΣμ(z)=1zημ1(z)satisfies the equation

Σμν(z)=Σμ(z)Σν(z) (2)

for z in a neighborhood of zero (if μ, νM), or in a neighborhood of the interval (ε,0)for some ε >0 (if μ, νM+). For proofs we refer to [5] and [14].

Previous results related to regularity questions for free convolutions of probability measures seem to indicate that these operations typically do not favor the existence of a large singular part (with respect to the Lebesgue measure). It has been proved in [6] and, respectively, in [1], that after free additive, respectively multiplicative, convolution, atoms usually disappear. More precisely,cis an atom forμνif and only if there exist atomsa ofμandbofνwith the property thata+b=candμ({a})+ν({b}) >1. Moreover,(μν)({c})=μ({a})+ν({b})−1. A similar result holds for free multiplicative convolution. The singular continuous part has been shown to disappear altogether for the free additive convolution of a measure with itself. Also, in this case, the density of the absolutely continuous part is continuous, and analytic outside a closed set of measure zero inR(see [2], Theorem 3.4). Similar results for free multiplicative convolutions can be found in [3]. Biane has shown in [7] that ifμ=2π t1

4t−x2dxis the semicircular distribution, thenμν is absolutely continuous with respect to the Lebesgue measure, its density is a continuous function, analytic wherever strictly positive. These facts contrast sharply with the behavior of classical convolution.

In this paper we show that, roughly speaking, the singular continuous part of the free convolution of two measures, none of them a point mass, if existing, has its (topological) support included in the support of the absolutely continuous part. Moreover, the Lebesgue measure of the singular continuous support is always zero. It is easy to see that these results have no classical counterpart. Indeed, consider a probability measureνwhich is purely singular continuous and has its topological support equal to the interval[0,1]. Its classical convolution with4+δ4)/2 will give a purely singular measure.

An important tool in proving results related to regularity for free convolutions has been subordination of analytic functions. The following result has been first proved in [15] under some more restrictive hypotheses, and later in [8]

in full generality (see Theorems 3.1, 3.5, and 3.6):

(3)

Theorem 1.1.

(a) Letμ, νMbe two Borel probability measures on the real line. Then there is an analytic functionωμ,νonC\R such that

(a1) Fμν(z)=Fμμ,ν(z)),z∈C\R;

(a2) ωμ,ν(z)¯ =ωμ,ν(z),ωμ,ν(z)zforz∈C+, andlimy→+∞μ,ν(iy)/iy)=1.

The mapωμ,ν is uniquely determined by properties(a1)and(a2).

(b) Letμ, νM. There exists an analytic functionωμ,νdefined onDsuch that (b1) |ωμ,ν(z)||z|,z∈D;

(b2) ψμμ,ν(z))=ψμν(z),z∈D.

The mapωμ,ν is uniquely determined by properties(b1)and(b2).

(c) Letμ, νM+be different fromδ0. There exists an analytic functionωμ,ν defined onC\ [0,+∞), such that (c1) Ifz∈C+, thenωμ,ν(z)∈C+, ωμ,ν(z)¯ =ωμ,ν(z)andarg(ωμ,ν(z))arg(z);

(c2) ψμμ,ν(z))=ψμν(z),z∈C\ [0,+∞).

The mapωμ,ν is uniquely determined by properties(c1)and(c2).

The results stated in the following lemma have been proved in [6] and [1], and are direct consequences of Theo- rem 1.1, equalities (1) and (2), and of analytic continuation.

Lemma 1.2.

(a) Letμ, νM. With the notations from Theorem1.1, the following equality holds:

Fμν(z)+z=ωμ,ν(z)+ων,μ(z), z∈C\R.

(b) Letμ, νM+(orμ, νM). With the notations from Theorem1.1, the following equality holds:

ψμν(z)= ωμ,ν(z)ων,μ(z) zωμ,ν(z)ων,μ(z), or, equivalently,

ημν(z)=1

μ,ν(z)ων,μ(z),

for allz∈C\ [0,+∞)(orz∈D, respectively).

For a function f:D→C∪ {∞}, and a pointx∈T, we say that the nontangential limit of f at x exists if the limit limzx, zΓα(x)f (z)exists for allα(0, π ), whereΓα(x)denotes the angular domain having vertexx, angular openingπα, and being bisected by the radius ofDthat ends atx. A similar definition holds for maps defined on the upper half-plane. Iff:C+→C∪ {∞}andx∈R, then the nontangential limit off atxexists if limzx, zΓα(x)f (z) exists for allα(0, π ), whereΓα(x)denotes the angular domain having vertexx, angular openingπα, and being bisected by the perpendicular on Ratx. We say that nontangential limit at infinity exists if limz→∞, zΓα(0)f (z) exists, for allα(0, π ). We shall denote nontangential limits bylimzαf (z), or

z−→αlim

f (z).

The following three theorems describe properties of analytic functions in the unit disc related to their nontangential boundary behavior.

Theorem 1.3.Letf:D→Cbe a bounded analytic function. Then the set of pointse∈Tat which the radial limit limr1f (re)off fails to exist is of linear measure zero.

Theorem 1.4. Letf:D→Cbe a bounded analytic function, and let e ∈T. If there exists a pathγ:[0,1)→D such thatlimt1γ (t )=e and=limt1f (γ (t ))exists inC, then the nontangential limit off ate exists, and equals.

(4)

Theorem 1.5.Letf:D→Cbe an analytic function. Assume that there exists a setAof nonzero linear measure inT such that the nontangential limit off exists at each point ofA, and equals zero. Thenf (z)=0for allz∈D.

Theorem 1.3 is due to Fatou, Theorem 1.5 to Privalov, and Theorem 1.4 to Lindelöf. For proofs, we refer to [10]

(see Theorems 2.1, 2.3, and 8.1). Observe that, by Theorems 1.3 and 1.4, the existence of radial and nontangential limits are equivalent for bounded analytic maps defined in the unit disc.

Let us notice that, in particular, all the above theorems apply to self-maps of the unit disc. The upper half-plane is known to be conformally equivalent to the unit disc via a rational transformation of the extended complex plane:

the mapg(z)=(z−i)/(z+i)is a conformal bijection ofC∪ {∞}onto itself that carriesC+ontoDandR∪ {∞}

ontoT; a subset ofRhas Lebesgue measure zero if and only if its image viagis a set of linear measure zero inT.

Moreover, such rational transformations preserve angles. Thus, even if the mapgdoes not necessarily carry an angular domain onto an angular domain, we have however thatf has nontangential limit atg(x)∈Tif and only iffghas nontangential limit atx∈R∪ {∞}. In particular, all the above theorems apply to analytic self-maps of the upper half-plane, or to analytic maps fromC+into−C+.

Main ingredients of our proofs will be two results from the theory of cluster sets of analytic functions. Given a domain (i.e. an open connected set)D⊆C∪ {∞}, the cluster set of a functionf:D→C∪ {∞}at the pointPD is

C(f, P )=

z∈C∪ {∞} | ∃{zn}n∈ND\P such that lim

n→∞zn=P , lim

n→∞f (zn)=z

.

A useful, but obvious property ofC(f, P )is the following:

Lemma 1.6.LetD⊂Cbe a domain andf:D→C∪ {∞}be continuous onD. IfDis locally connected atPD, thenC(f, P )is either one point, or a continuum.

(This result appears in [10], as Theorem 1.1.)

The first one is the following theorem of Seidel, which describes the behavior of certain analytic functions near the boundary of their domain of definition. For proof, we refer to [10], Theorem 5.4.

Theorem 1.7.Letf:D→Dbe an analytic function such that the radial limitf (e)=limr1f (re)has modulus1 for almost everyθ1, θ2). Ifθ1, θ2)is such thatf does not extend analytically throughe, thenC(f,e)= D.

The second result is the following theorem of Carathéodory (Theorem 5.5 in [10]):

Theorem 1.8.Letf (z)be analytic and bounded in|z|<1. Assume that for almost everyθ1, θ2)the radial limit f (e)belongs to a setW in the plane. Then, for everyθ1, θ2)the cluster setC(f,e)is contained in the closed convex hull ofW.

Theorem 1.7 can be applied to self-maps of the upper half-planeC+, via a conformal mapping, but in that case one must consider meromorphic, instead of analytic, extensions.

Theorems 1.7 and 1.8 allow us to prove the following Proposition 1.9.

(a) Letf be an analytic self-map ofDsuch that|limr1f (re)| =1for almost everyθ1, θ2). Suppose that there is a pointθ01, θ2)such that the functionf cannot be continued analytically throughe. Then for any t1< t2there is a setE1, θ2)of nonzero Lebesgue measure such thatlimr1f (re)exists for allθE, and the set{limr1f (re): θE}is dense in the arcA= {eit: t1< t < t2}.

(b) Let f be an analytic self-map of C+ such that limy0f (x+iy) exists and belongs to R for almost every x(a, b). Suppose thatx0(a, b)is such thatf cannot be continued meromorphically throughx0. Then for anyc < dthere is a setE(a, b)of nonzero Lebesgue measure such thatlimy0f (x+iy)exists for allxE, and the set{limy0f (x+iy): xE}is dense in the interval(c, d).

(5)

Proof. Let f and θ0 be as in the hypothesis of (a). Using Theorems 1.7 and 1.8, we conclude that, on the one hand, C(f,e0)= D, and on the other, that C(f,e0) equals the closure of the convex hull of the set {limr1f (re): θG} for any set G1, θ2)with the property that limr1f (re)exists for all θG, and 1, θ2)\Ghas zero linear measure. This implies that the set

rlim1f re

: θ1< θ < θ2, lim

r1f re

exists and belongs toA

is dense inA. It remains to prove that the setEof thoseθ1, θ2)such that limr1f (re)exists and belongs toA has nonzero linear measure. If this were not true, then, according to Theorem 1.8, we could takeG\Einstead ofG in the previous argument, and obtain a contradiction. This proves (a).

Part (b) follows directly from (a), by using the conformal mappingz(z−i)/(z+i)and its inverse. 2 Consider now a Borel probability measureμonR. It is a well-known fact that one can draw informations about a measure from the behavior of its Poisson integral near the support of that measure. For example, the value atx∈Rof the density function of the absolutely continuous part with respect to the Lebesgue measure of a finite measure can be obtained as nontangential limit atx of the Poisson integral of the measure for almost allx∈R(for more details we refer to [13]). A simple computation shows that the imaginary part of the Cauchy transform ofμis (up to a multiple of−π) equal to the Poisson integral ofμ:

Gμ(x+iy)= −

R

y

(xt )2+y2dμ(t ), x∈R, y >0.

On the other hand, the reciprocal of the Cauchy transform of μ,Fμ, mapsC+ into itself. Thus, it will be natural to investigate the boundary behavior of the analytic map Fμ (and, implicitly, ofGμ) in order to draw conclusions aboutμ.

The following result is well known. Since we do not know any reference, we give here a full proof. We do not claim any paternity of this proof.

Lemma 1.10.Letμbe a Borel probability measure onR, and denote byμscits singular continuous part. Then, for μsc-almost allx∈R, the nontangential limit of the imaginary part of the Cauchy transformGμofμatxis infinite.

Proof. According to Theorem 1.4, it is enough to show that, forμsc-almost allx∈R, the imaginary part ofGμ(x+iy) tends to infinity asy approaches zero. As we have seen before, for anyy >0 andx∈R,

Gμ(x+iy)=

R

y

(xt )2+y2dμ(t ).

Let us observe that y2/((xt )2+y2)1/2 for all t∈R such that |xt|y. Thus,for any given y >0, the following holds:

Gμ(x+iy)=

R

y

(xt )2+y2dμ(t )=1 y

R

y2

(xt )2+y2dμ(t )

1

y

x+y

xy

y2

(xt )2+y2dμ(t )μsc((xy, x+y])

2y .

Now we can apply de La Vallée Poussin’s theorem (Theorem 9.6 in [12]) to conclude that limy0μsc((xy, x+y])/

(2y)= ∞forμsc-almost allx∈R(see also Theorem 31.6 in [9]). 2

To study properties of the free multiplicative convolutions of probability measures, one finds more convenient to use the functionsψμandημdefined at the beginning of the introduction.

Remark 1.11.The formulaGμ(1/z)=z(ψμ(z)+1)together with Lemma 1.10 above, allows us to conclude that for μsc-almost allx(0,+∞), the nontangential limit ofψμat 1/xis infinite.

(6)

Remark 1.12.For measuresμM, we have ψμ(z)=

T

z

¯

tzdμ(t )= −1 2+1

2

T

t+z

tzdμ(¯t ), z∈D,

and hence the functionψμsatisfies the conditionψμ(z)−1/2 for allz∈D. The real part of 1

π

ψμ(z)+1

2 = 1

T

t+z

tzdμ(¯t ), z∈D,

is the Poisson integral of the measure dμ(¯t ), and thus, by Lemma 1.10 and a conformal transformation, forμsc-almost allx∈T, the nontangential limit ofψμatx¯is infinite (a detailed presentation of Poisson integrals of measures on the unit circle can be found in [11]).

We conclude the introduction with a proposition which essentially characterizes analytic functions that are Cauchy transforms of probability measures.

Proposition 1.13.

(a) For everyμM, we haveFμ(z)z,z∈C+. The inequality is strict if and only ifμis not a point mass.

(b) For everyμM+=δ0, we haveπ >argημ(z)argz,z∈C+. The second inequality is strict if and only if μis not a point mass.

(c) For everyμM, we have|ημ(z)||z|,z∈D. The inequality is strict if and only ifμis not a point mass.

Part (a) of the above proposition can be found also in [5], while parts (b) and (c) appear in [3] as Propositions 2.2 and 3.2.

2. Boundary behavior ofGμν

In the following we shall denote by suppτ the topological support of the measure τ, and by τac (respectively τs,τsc,τa) the absolutely continuous (respectively singular, singular continuous, atomic) part of the measureτ with respect to the Lebesgue measure. We shall fix two Borel probability measuresμ, ν on the real line, none of them concentrated in one point. For simplicity, denote the subordination functionsωμ,νandων,μprovided by Theorem 1.1 byω1andω2, respectively.

Lemma 2.1([4]). Assume that there existsa∈Rsuch thatFμ(z)=Fν(a+Fμ(z)z)for all z∈C+. Then the functionshμ(z)=Fμ(z)z+a,hν(z)=Fν(z)z+a,z∈C+, are conformal self-maps of the upper half-plane, inverse to each other.

Proof. Lethμ,hν be as in the statement of the lemma. We have:

z=Fν

Fμ(z)z+a

Fμ(z)z+a

+a=hν hμ(z)

, z∈C+.

Since, by Proposition 1.13, bothhμandhν are nonconstant analytic self-maps of the upper half-plane, applyinghμ in both sides of the previous equality yieldshμ(hν(w))=w, for allw in the open nonempty sethμ(C+), and, by analytic continuation, for allw∈C+. This shows that bothhμ, hν are conformal self-maps ofC+ inverse to each other and completes the proof. 2

We effectively use in the proof the fact that neither one ofμ, νis concentrated at one point.

Remark 2.2.Any functionFμsatisfying the conditions of Lemma 2.1 must be of the form Fμ(z)=z2+mz+p

z+r , m, p, r∈R,

(7)

and thus, μis purely atomic with at most two atoms. Using relation (1), it is easy to see that, if bothμ andν are convex combinations of two atoms, then for anyx∈Rlimzx,z>0Fμν(z)exists inC+∪R∪ {∞}.

The next theorem describes the boundary behavior of the Cauchy transform ofμν.

Theorem 2.3.Letμ, νM, none of them concentrated in one point. Then the set C+=

a∈R: C(Fμν, a)∩C+is infinite is empty. Ifμandνhave compact support, then the set

C=

a∈R: C(Fμν, a)is infinite is empty.

Let us note thatCis empty if and only if limzx,z>0Gμν(z)exists for allx∈R, or, equivalently, if the restriction ofGμνto the upper half-plane extends continuously toC+∪Ras function with values inC∪ {∞}.

Proof. We shall consider the restriction of the functions F and ω to the upper half-plane. By Lemma 2.1 and Remark 2.2, it is enough to show that if aC+ (or, in the second case, ifaC), then Fμ andFν satisfy either Fμ(z)=Fν(a+Fμ(z)z), orFν(z)=Fμ(a+Fν(z)z).

SupposeaC+. The relation ω1(z)+ω2(z)=z+Fμν(z)

from Lemma 1.2(a), together with Gμω1=Gνω2=Gμν, assures us that at least one of C(ω1, a)∩C+, C(ω2, a)∩C+will be infinite. Without loss of generality, assume thatC(ω1, a)∩C+is infinite, hence, according to Lemma 1.6, a continuum.

For anyzC(ω1, a)∩C+there exists a sequence{zn}n∈N⊂C+converging toasuch that limn→∞ω1(zn)=z.

Using Lemma 1.2(a), we have:

nlim→∞ω2(zn)= lim

n→∞zn+Fμν(zn)ω1(zn)=a+ lim

n→∞Fμ

ω1(zn)

z=a+Fμ(z)z.

Now the subordination formula from Theorem 1.1(a), will give Fμ(z)= lim

n→∞Fμ ω1(zn)

= lim

n→∞Fν ω2(zn)

=Fν

a+Fμ(z)z ,

for allzC(ω1, a)∩C+. SinceC(ω1, a)∩C+, is a continuum inC+, the identity theorem for analytic functions will imply that

Fμ(z)=Fν

a+Fμ(z)z

for allz∈C+. By Remark 2.2, we haveC+= ∅.

Let us assume now thatμandν have compact support, and suppose thatC(Fμν, a)⊆R∪ {∞}has more than one point. Then, by Lemma 1.6,C(Fμν, a)must be either a closed interval, or the complement of an open (possibly empty) interval inR∪ {∞}. It is easy to see thatC(ω1, a),C(ω2, a)do not contain points inC+. As before, at least one ofC(ω1, a),C(ω2, a)⊆R∪ {∞}will be nontrivial. Without loss of generality, assumeC(ω1, a)is nontrivial.

We claim that for any cinC(ω1, a)\ {∞}, with the possible exception of three points, there exists a sequence {z(c)n }n∈Nconverging toasuch that limn→∞ω1(z(c)n )=c, andω1(z(c)n )=cfor alln. Let{cn}n∈Nbe a dense sequence in C(ω1, a), and considerzn∈C+, such that|zna|<1/n, and |ω1(zn)cn|<1/n,n∈N. We define a path γ:[0,1] →C+∪ {a}such thatγ (1−1/n)=zn,γ (1)=a, andγ is linear on the intervals[1−1/n,1−1/(n+1)], n∈N. It will suffice to show that there exists at most one pointcin the interior ofC(ω1, a)such thatω1(γ ([0,1)))∩ {c+it: t∈ [0, ε)} = ∅for someε >0. Indeed, assume to the contrary thatc < care two such points. The set

K=

c+it: t(0,1)

c+it: t(0,1)

s+i, csc

separatesC+into two components, and the pathω1(γ (t ))contains infinitely many points in either of the components, hence it crossesKinfinitely many times. By our assumption, crossings cannot be close tocorc, and this implies the existence of a point inC(ω1, a)K⊂C+, contradicting the fact thatC(ω1, a)∩C+= ∅. This proves our claim.

(8)

We shall next show that, in fact, bothC(ω1, a), C(ω2, a)⊆R∪ {∞}contain more than one point. Indeed, suppose, for example, that limzaω2(z)=ω2(a)∈R∪ {∞}. Ifω2(a)= ∞, then

−∞

tdν(t )= lim

z→∞Fν(z)z=lim

zaFν ω2(z)

ω2(z)=lim

zaω1(z)z, so thatC(ω1, a)= {a

−∞tdν(t )}, which is a contradiction. Assumeω2(a)∈R. By Theorems 1.3, 1.4, and 1.5, the nontangential limit ofFμatxexists and is finite for allx∈R, with the possible exception of a subset of Lebesgue measure zero. Thus, for almost allcC(ω1, a), we have:

Fμ(c)=lim

z−→c

Fμ(z)= lim

n→∞Fμ ω1

z(c)n

= lim

n→∞Fμν z(c)n

= lim

n→∞ω1 z(c)n

+ω2 z(c)n

z(c)n

=c+ω2(a)a.

Using Privalov’s theorem (Theorem 1.5), we conclude thatFμ(z)=z(aω2(a))for allz∈C+. This contradicts the fact thatμis not concentrated at the pointaω2(a). So indeed bothC(ω1, a),C(ω2, a)are infinite.

Assume that the nontangential limit ofFμexists atcC(ω1, a), and there exists a sequence{z(c)n }n∈Nconverging toasuch that limn→∞ω1(z(c)n )=candω1(z(c)n )=cfor alln. Then

Fμ(c)=lim

z−→c

Fμ(z)= lim

n→∞Fμ ω1

z(c)n

= lim

n→∞Fμν z(c)n

= lim

n→∞Fν

z(c)n +Fμ ω1

z(c)n

ω1 z(c)n

.

If there exists a subsetEofC(ω1, a)of nonzero Lebesgue measure such thatFμ(c)∈C+for allcE, then we have Fμ(c)=lim

z−→c

Fμ(z)=lim

z−→c

Fν

a+Fμ(z)z

=Fν

a+Fμ(c)c

, cE.

By Theorem 1.5, this implies that the equalityFμ(z)=Fν(a+Fμ(z)z)holds for allz∈C+, and thus, by Re- mark 2.2, we obtain a contradiction.

IfFμ(c)∈Rfor almost allcC(ω1, a), then we use Theorem 1.7 to conclude that eitherC(Fμ, c)=C+∪R∪{∞}

for somecC(ω1, a), orFμextends meromorphically through the interior ofC(ω1, a).

In the former case, since suppν is compact, we use Proposition 1.9(b) to conclude that there exists a set EC(ω1, a)of nonzero Lebesgue measure such that

z−→clim

Fμ(z)z+a∈R\suppν, cE.

But this means that Fμ(c)=lim

z−→c

Fμ(z)= lim

n→∞Fμ

ω1

z(c)n

= lim

n→∞Fν

z(c)n +Fμ

ω1

z(c)n

ω1

z(c)n

=Fν

a+Fμ(c)c ,

for allcE. SinceEis of positive Lebesgue measure, we apply Privalov’s theorem to conclude thatFμ(z)=Fν(a+ Fμ(z)z)for allz∈C+and obtain a contradiction.

In the latter case, we look atC(ω2, a). If either there is a subset of nonzero Lebesgue measure ofC(ω2, a)on which the nontangential limit ofFν has imaginary part strictly greater than zero, or there is acC(ω2, a)at which C(Fν, c)=C+∪R∪ {∞}, then by the same argument as before we have the equality

Fν(z)=Fμ

a+Fν(z)z

, z∈C+,

and thenC(Fμν, a)must be trivial by Lemma 2.1 and Remark 2.2.

The only case that remains to be analyzed is whenFμextends analytically through the interior ofC(ω1, a)andFν

extends analytically through the interior ofC(ω2, a). By dropping if necessary to a subsequence, we may assume that bothc=limn→∞ω1(z(c)n )and limn→∞ω2(z(c)n )exist, wherez(c)n is such thatω1(z(c)n )=c.

Suppose there were a point dC(ω2, a) and a set VdC(ω1, a) of nonzero Lebesgue measure such that limn→∞ω2(z(c)n )=d for allcVd. Taking limit asn→ ∞in the equality

Fμ ω1

z(c)n

+zn(c)=Fμν z(c)n

+z(c)n =ω1 z(c)n

+ω2 z(c)n

(9)

gives

Fμ(c)+a=c+d, for allcVd.

Applying again Privalov’s theorem, we obtain thatFμ(z)=z(ad)for allz∈C+. This contradicts the fact that μis not concentrated at the pointad. Thus, there exists a setEC(ω1, a)of positive Lebesgue measure such that {c=limn→∞ω2(z(c)n ): cE} ⊆IntC(ω2, a). Then

Fμ(c)=lim

zcFμ(z)= lim

n→∞Fμ ω1

z(c)n

= lim

n→∞Fμν z(c)n

= lim

n→∞Fν

z(c)n +Fμ ω1

z(c)n

ω1 z(c)n

=Fν

a+Fμ(c)c

for allcE. Privalov’s theorem implies thatFμ(z)=Fν(a+Fμ(z)z)for allz∈C+. This concludes the proof. 2 Remark 2.4.The proof of Theorem 2.3 implies that, whenever μ, νhave compact support and neither of them is concentrated in one point, the restrictions to the upper half-plane ofω1andω2can be continuously extended toR.

An immediate consequence of Theorem 2.3 is the following

Corollary 2.5.Ifμandνhave compact support and neither of them is concentrated in one point, thensupp(μν)s is closed and of zero Lebesgue measure. Moreover,supp(μν)sc⊂supp(μν)ac. In particular,μνcan never be singular(the last statement holds also for Borel probability measures with noncompact support onR).

Proof. Suppose first thatμ, νhave compact support. According to Theorem 2.3,Fμν extends continuously toR.

Thus, the preimage of zero under the extension ofFμν toC+∪Rmust be a closed set. This fact, together with Lemma 1.10, imply that the set supp(μν)sc is included in the closed set Fμ1ν({0}). According to Privalov’s theorem (Theorem 1.5),Fμ1ν({0})must be of zero Lebesgue measure. Since by Theorem 7.4 of [6],μνcan have only finitely many atoms, we conclude that supp(μν)sis a closed set of zero Lebesgue measure.

Assume now thatx∈supp(μν)sc. Asμν can have only finitely many atoms, there is no loss of generality to assume thatx is not an atom ofμν. Recall thatπ1Gμν is the Poisson integral ofμν. Assume to the contrary that x does not belong to supp(μν)ac. Thus, there exists an open neighborhoodV ⊆Rof x such that the nontangential limit of the Poisson integral of μν ata equals zero for almost all aV with respect to the Lebesgue measure. We conclude thatlimzaGμν(z)∈Rfor Lebesgue-almost allaV. On the other hand, the setV ∩supp(μν)sc is infinite, without isolated points, and the nontangential limit ofGμν at(μν)sc-almost all points inV ∩supp(μν)sc is infinite, by Lemma 1.10. Thus, the identity theorem for analytic functions implies thatGν)sc does not extend meromorphically through anyaV ∩supp(μν)sc. We conclude by Theorem 1.7, applied to the upper half-plane, thatC(Gμν, x)=C+∪R∪ {∞}, and thusC(Fμν, x)=C+∪R∪ {∞}. This contradicts Theorem 2.3.

Finally, suppose that μ and ν have arbitrary support, and assume that μν is purely singular. According to Theorem 7.4 of [6], μν cannot be purely atomic, and can have only finitely many atoms. Thus, it must have a singular continuous component. As before, we conclude that for anyx∈supp(μν)which is not an atom ofμν we haveC(Fμν, x)=C+∪R∪ {∞}, contradicting Theorem 2.3. 2

3. Boundary behavior ofψμν

3.1. Measures supported on[0,+∞)

As in Section 2, letμ, νM+, neither of them a point mass, and denote the subordination functions provided by Theorem 1.1ωμ,ν=ω1, andων,μ=ω2. We have the following analogue of Lemma 2.1:

Lemma 3.1.If there existsa(0,+∞)such thatημ(z)=ην(azημ(z)), then there existα1, α2, β1, β2>0,t, s(0,1) such thatα1=α21=β2, andμ=t δα1+(1t )δα2=β1+(1s)δβ2.

(10)

Proof. Consider the analytic functionshμ(z)=μ(z)/z,hν(z)=ν(z)/z,z∈C\ [0,+∞). The equationημ(z)= ην(azημ(z))implies

hν hμ(z)

= a hμ(z)ην

hμ(z)

= a

hμ(z)ημ(z)=z, z∈C\R.

Since, by Proposition 1.13,hμis not constant, applyinghμto the previous equality, we obtainhμ(hν(w))=wfor all win the nonempty open sethμ(C\R). Sohμandhνare conformal maps inverse to each other.

Sincez(ψμ(z)+1)=Gμ(1/z), we have hμ(z)=a

1

z− 1

Gμ(1/z) =a 1

zFμ 1

z

, z∈C\ [0,+∞), sohμ(C+)⊆C+, by Proposition 1.13.

As we also haveημ((−∞,0))⊂(−∞,0), so thathμmaps(−∞,0)into(0,), andμ, νare not point masses, it follows that

hμ(z)=cz+d

z+b , c >0, b <0, d < cb, z= −b.

Then

hν(z)=−bz+d

zc , z=c.

Thus,ημ(z)=(cz2+dz)/(az+ab)andην(z)=(bz2+dz)/(azac)extend analytically through[0,+∞)\{−b}, and[0,+∞)\ {c}, respectively. This implies that bothμandν must be convex combination of two distinct Dirac measures,μ=t δα1 +(1t )δα2, andν=β1+(1s)δβ2 for someα1, α2, β1, β2>0,t, s(0,1), and concludes the proof. 2

It is not obvious that, forμandνas in Lemma 3.1, C(ημν, x)is finite for allx∈ [0,+∞). To prove this, we shall compute explicitlyημν. Let us note that

ημ(z)= ψμ(z)

1+ψμ(z)= t zα1(12)+(1t )zα2(11)

(11)(12)+t zα1(12)+(1t )zα2(11)

=z(t α1+(1t )α2)z2α1α2 1−z((1t )α1+t α2) ,

so thatα1α2= −c/(ab),α1+α2=(da)/(ab), andt=(1+1)/(bα12). This implies

b= 1

(t−1)α1t α2, c= 1α2

(1t )α1+t α2, and d=a+ a(α1+α2) (t−1)α1t α2.

Thus,α1, α2, t, anda determine uniquelyb, c, andd. For simplicity, we shall computeημν in terms ofb, c, d.

Using successively Theorem 1.1(c), Lemmas 3.1, and 1.2, we have ημ

ω1(z)

=ην ω2(z)

=ημ

ν2(z)) ω2(z) =ημ

1(z)

z , z∈C\ [0,+∞).

Sinceημ(z)=(cz2+dz)/(az+ab), the previous equality will imply that the subordination functionω1satisfies acω1(z)2+bc(a+z)ω1(z)+bdz=0, z∈C\ [0,+∞).

This, together with the conditionω1((−∞,0))⊆(−∞,0)(required by Theorem 1.1(c1)), shows that ω1(z)=−bc(a+z)

b2c2(a+z)2−4abcdz

2ac , z∈C\ [0,+∞).

A similar computation gives ω2(z)=bc(a+z)+

b2c2(a+z)2−4abcdz

2ab , z∈C\ [0,+∞).

Since, by Lemma 1.2(b)ημν(z)=ω1(z)ω2(z)/zfor allz∈C\ [0,+∞), we conclude thatC(ημν, x)is finite for allx∈ [0,+∞). Now we are ready to prove the following

Références

Documents relatifs

Finally we give a bound on the Hausdorff measure of the free boundary, generalizing to non-smooth fixed boundaries ~03A9, a result of Brezis, Kinder-.. lehrer [6]

We provide a rather complete description of the sharp regularity theory to a family of heterogeneous, two-phase free boundary problems, J γ → min, ruled by nonlinear,

For a one-phase free boundary problem involving a fractional Laplacian, we prove that “flat free boundaries” are C 1,α. We recover the regularity results of Caffarelli for

For properly defined weak solutions to such systems, we prove various regularity properties: higher integrability, higher differentiability, partial regularity of the spatial

In the sequel we show how the bisection and concentration properties can be improved as stated in the following theorem which represents the main result of this paper.. T

The difference of these values converges to zero as the convolutions approach u, and we prove that the rate of convergence to zero is connected to regularity: u ∈ W 1 , 2 if and only

On the contrary, convergence in W 1 (or in duality with some other class F ) ensures good estimates simultaneously for

After that, it is easy to prove global in time estimates for the solutions to (1.6) just by combining the estimates for the paralinearized system, and standard product laws in