• Aucun résultat trouvé

INTEGRAL KERNELS ON COMPLEX SYMMETRIC SPACES AND FOR THE DYSON BROWNIAN MOTION

N/A
N/A
Protected

Academic year: 2021

Partager "INTEGRAL KERNELS ON COMPLEX SYMMETRIC SPACES AND FOR THE DYSON BROWNIAN MOTION"

Copied!
28
0
0

Texte intégral

(1)

HAL Id: hal-01789322

https://hal.archives-ouvertes.fr/hal-01789322v2

Preprint submitted on 18 Dec 2020

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

INTEGRAL KERNELS ON COMPLEX SYMMETRIC SPACES AND FOR THE DYSON BROWNIAN

MOTION

Piotr Graczyk, P Sawyer

To cite this version:

Piotr Graczyk, P Sawyer. INTEGRAL KERNELS ON COMPLEX SYMMETRIC SPACES AND

FOR THE DYSON BROWNIAN MOTION. 2020. �hal-01789322v2�

(2)

INTEGRAL KERNELS ON COMPLEX SYMMETRIC SPACES AND FOR THE DYSON BROWNIAN MO- TION

P. Graczyk 1 and P. Sawyer

1

LAREMA, UFR Sciences, Universit´ e d’Angers, 2 bd Lavoisier, 49045 Angers cedex 01, France

2

Department of Mathematics and Computer Science, Laurentian University, Sudbury, Canada P3E 2C6

Key words Poisson kernel, Newton kernel, heat kernel, spherical functions, Dunkl processes MSC (2010) 31B05, 31B25, 60J50, 53C35

In this article, we consider flat and curved Riemannian symmetric spaces in the complex case and we study their basic integral kernels, in potential and spherical analysis: heat, Newton, Poisson kernels and spherical functions, i.e. the kernel of the spherical Fourier transform.

We introduce and exploit a simple new method of construction of these W -invariant kernels by alternating sum formulas. We then use the alternating sum representation of these kernels to obtain their asymptotic behavior. We apply our results to the Dyson Brownian Motion on R

d

.

Copyright line will be provided by the publisher

1 Introduction and notations

Analysis on Riemannian symmetric spaces of Euclidean type, also called flat symmetric spaces, continues to develop in recent years ([21, 23, 45, 46]). Its importance is due to its relationship with Dunkl analysis ([9, 12, 42]) together with the correspondence of the complex case with the parameter k = 1, in which symmetric spaces of Euclidean type constitute the “geometric case”, frequently used as a model case in most challenging open problems of Dunkl theory. The analysis on flat complex symmetric spaces coincides with Weyl group invariant Dunkl analysis associated with multiplicity k = 1, see [9]. In particular, the heat kernel p W t (X, Y ) is a special case of the heat kernel in the Weyl group invariant Dunkl setting. We employ this intimate connection to Dunkl theory in our paper in Section 3 as one of main tools of the proof of main theorems. This connection appears also in Proposition 2.8.

Another important aspect of this paper is to apply analysis on symmetric spaces of Euclidean type to potential theory and to stochastic analysis of Dyson Brownian Motion, one of the most important models of non-colliding particles, see the recent survey [29]. We expect further applications of our results and techniques to other non-intersecting stochastic path problems related to root systems and to multivariate stochastic processes related to Laplace-Beltrami operators on symmetric spaces, to Dunkl Laplacians and to Schr¨ odinger operators, see the discussion in Section 5. We thank an anonymous referee for pointing out to us such further stochastic applications.

The objective of this paper is to study basic integral kernels, in potential theory and spherical analysis:

heat, Newton, Poisson kernels, Green function and spherical functions (i.e. the kernel of the spherical Fourier transform), in the set-up of flat and curved symmetric spaces of complex type.

Our main results on the exact form and asymptotics of the heat, Poisson and Newton kernels (Theo- rems 2.2, 3.11, 3.13 and Corollaries 5.6 and 5.7) are crucial for the future development of the potential theory on flat and curved symmetric spaces of complex type, and for the potential theory of the Dyson

The first author thanks Laurentian University of Sudbury for its hospitality and financial support during his visits to Sudbury. The second author thanks LAREMA for its hospitality and the R´ egion Pays de la Loire for its financial support on several occasions via the projects Matpyl, G´ eanpyl and D´ efimaths.

Corresponding author: e-mail: psawyer@laurentian.ca, Phone: +1-613-408-5820

Copyright line will be provided by the publisher

(3)

Brownian Motion. These results are a starting point of research and a source of conjectures for the corresponding kernels in the Weyl-invariant Dunkl setting (for the rank one case, refer to [19]).

The main result on asymptotics of the spherical functions contained in Theorem 4.5 is important from the point of view of spherical analysis on symmetric spaces, because it generalizes significantly the results of Helgason in [23], of Narayanan, Pasquale and Pusti in [36] and of Schapira in [44], for the flat and curved symmetric spaces in the complex case, cf. Remark 4.9.

We recall now some basic terminology and facts about symmetric spaces associated to Cartan motion groups.

Let G be a semisimple Lie group and let g = k ⊕ p be the Cartan decomposition of G. We recall the definition of the Cartan motion group and the flat symmetric space associated with the semisimple Lie group G with maximal compact subgroup K. The Cartan motion group is the semi-direct product 5

G 0 = K o p where the multiplication is defined by (k 1 , X 1 ) · (k 2 , X 2 ) = (k 1 k 2 , Ad(k 1 )(X 2 ) + X 1 ).

The associated flat symmetric space is then M = p ' G 0 /K (the action of G 0 on p is given by (k, X) · Y = Ad(k)(Y ) + X ).

We tacitly identify K-invariant measures, functions, differential operators on M with W -invariant measures etc. on a.

The spherical functions for the symmetric space M are then given by ψ λ (X ) =

Z

K

e λ( Ad (k)(X)) dk (1)

where λ is a complex linear functional on a ⊂ p, a Cartan subalgebra of the Lie algebra of G. To extend λ to X ∈ Ad(K)a = p, one uses λ(X ) = λ(π

a

(X)) where π

a

is the orthogonal projection with respect to the Killing form (denoted throughout this paper by h·, ·i). Note also that the spherical function for the symmetric space G/K is given by

φ λ (g) = Z

K

e (λ−ρ)(H(g k)) dk (2)

where λ is a complex linear functional on a and the map H is defined via the Iwasawa decomposition of G, namely g = k e H(g) n ∈ K A N and ρ = (1/2) P

α>0 m α α. Note that in [23, 24, 25], λ is replaced by i λ.

Throughout this paper, we suppose that G is a semisimple complex Lie group. The complex root systems are respectively A n−1 for n ≥ 2 (where p consists of the n × n hermitian matrices with trace 0), B n for n ≥ 2 (where p = i so(2 n + 1)), C n for n ≥ 3 (where p = i sp(n)) and D n for n ≥ 4 (where p = i so(2 n)) for the classical cases and the exceptional root systems E 6 , E 7 , E 8 , F 4 and G 2 .

Let ∆ be the Laplace-Beltrami operator on M and ∆ W its restriction to W -invariant functions on a where W is the corresponding Weyl group. Recall the formula

W f = π

−1

Rd

(π f), (3)

where π(X ) = Q

α>0 α(X) (see [25, Chap. II, Theorem 5.37]) in the Euclidean case.

In Section 2, we introduce and exploit a simple new method of construction of important K-invariant kernels on the space M .

We show in Theorem 2.2 that if K(X, Y ) is an Euclidean kernel (heat kernel, potential kernel, Poisson kernel, . . . ) for the Laplacian ∆

Rd

, then the corresponding kernel acting on W -invariant functions on M is given by the alternating sum

K W (X, Y ) = 1

|W | π(X) π(Y ) X

w∈W

(w)K(X, w · Y ). (4)

Here and in Theorem 2.2 below, K(X, Y ) is an Euclidean kernel on the Cartan subalgebra a which

is isomorphic to R d where d is the dimension of a and with the underlying scalar product being the

Killing form on a.

(4)

The proof of Theorem 2.2 is short and easy and uses the simple form of the operator ∆ W given in (3).

It is well-known that the spherical functions of the space can be written explicitly as such alternating sums ([25, Chap. IV, Proposition 4.10]).

The alternating sum formulas (4) also include determinantal formulas for transition probabilities p W t (x, y) (equivalently, for heat kernels) of Karlin-McGregor type, proven for Dyson Brownian Motions in Weyl chambers [18] and exploited in stochastic analysis (refer to [31, 32]).

The fact that alternating sums formulas (4) are true for many further analytic and stochastic kernels beyond spherical functions and heat kernels, was surprisingly not published or exploited (we asked experts of the field for an existing reference).

The approach with formulas (4) will allow us to provide asymptotics for kernels K W , using our knowledge of the kernels K(X, Y ) on R d as given in Table 1.

In Section 3, we discuss the asymptotic behaviour of the Poisson kernel especially when one or both arguments are singular. These results translate well to the Newton kernel.

In Section 4, we compute asymptotics for the spherical functions ψ λ (Y ) which can prove challenging when either λ or Y are singular (i.e. such that at least one of the nonzero root vanish on X or Y ).

Our results depend on a property we call “Killing-max” namely the property that for X , Y ∈ a + , hX, w · Y i = hX, Y i if and only if w ∈ W X W Y where W X = {w ∈ W : w · X = X}. It is known that this property is verified when either X or Y is non singular [23]. We prove in Appendix A, using the classification of Lie algebras, that the Killing-max holds in almost all cases (only in the cases related to the root systems E 6 , E 7 and E 8 is the question left unanswered).

We conclude with Section 5 where we apply the previous results to the heat kernel and Poisson and Newton kernels for the Dyson Brownian Motion.

Acknowledgements. We thank M. Denkowski for advice with Lemma 2.3 and J.-J. Loeb for useful discussions. We are grateful to both anonymous referees for their insightful comments and remarks that greatly helped to improve the paper.

2 Kernels on flat symmetric spaces in the complex case

2.1 Definitions

We first recall the classical integral kernels on R d in Table 1.

The integral kernels on the flat symmetric space M are considered with respect to the invariant measure µ(dY ) = π 2 (Y ) dY on M . Their definition is analogous to the classical R d and Riemannian manifold case, with the W -invariance imposed on the operator, boundary problem and solutions. The Dunkl-Poisson, Newton and Green kernels and their W -invariant versions were introduced and studied in [16] and [19].

Definition 2.1 We define a kernel K W (X, Y ) for the operator ∆ W and a boundary problem P as the fundamental solution of this PDE problem, which is W -invariant in X -variable, for each Y . Equivalently, K W (X, Y ) is an integral reproducing kernel for the W -invariant solutions of the problem P and this kernel is W -invariant in X .

The uniqueness of K W (X, Y ) may be deduced, as in the classical case, from the uniqueness of the spherical Fourier transform. Another approach for the existence of Poisson, Newton and Green kernels is available from the point of view of stochastic diffusion processes [7]. Note that W -invariant Dunkl processes are diffusions.

2.2 The method of alternating sums for constructing kernels on M This method will be introduced and used in the proof of Theorem 2.2 below.

Theorem 2.2 Let M be a symmetric space of Euclidean type with G a complex simple Lie group of rank d. Then the following formulas hold for X , Y ∈ a, a Cartan subalgebra associated with M .

Copyright line will be provided by the publisher

(5)

PDE Kernel Solution ∆

Rd

u(X, t) = ∂t u(X, t)

lim t→0

+

u(X, t) = f (X) p t (X, Y ) = e

|X−Y|

2 4t

(4 π t) d/2 u(X, t) = Z

Rd

f (Y ) p t (X, Y ) dY X ∈ R d , t > 0

Rd

u(X ) = f(X ) on R d ,

|u(X )| → 0 as X → ∞

f ∈ C c (R d ) N(X, Y ) = Φ(X − Y ), u(X ) = Z

Rd

f (Y ) N (X, Y ) dY ∆

Rd

u(X ) = 0 on B(X 0 , r)

u(X ) = f(X ) on ∂B(X 0 , r) P(X, Y ) = r 2 − |X − X 0 | 2

w d r |X − Y | d u(X ) = Z

∂B(X

0

,r)

f (Y ) P(X, Y ) dY ∆

Rd

u(X ) = f(X ) on B = B(0, 1)

u(X ) = 0 on ∂B G B (X, Y ) = Φ(X − Y ) u(X ) =

Z

B

f (Y ) G B (X, Y ) dY

−Φ(|X | (Y − X/|X | 2 ))

where w d = 2 π d/2 /Γ(d/2) (the surface area of a sphere of radius 1 in R d ) and Φ(X) = 1

2 π ln |X| if d = 2

1

(2−d) w

d

|X| 2−d if d ≥ 3 .

Table 1 The heat kernel p

t

, the Newton kernel N , the Poisson kernel P and the Green kernel G

B

for R

d

1. The heat kernel on M is given by p W t (X, Y ) = 1

|W | (4 π t) d/2 π(X) π(Y ) X

w∈W

(w)e

|X−w·Y|

2

4t

(5)

2. The Newton kernel on M is given by N W (X, Y ) = 1

2 π |W | π(X ) π(Y ) X

w∈W

(w) ln |X − w · Y | when d = 2, (6)

N W (X, Y ) = 1

|W | (2 − d) w d π(X) π(Y ) X

w∈W

(w)

|X − w · Y | d−2 when d ≥ 3.

3. The Poisson kernel of the open unit ball B is given for X ∈ B and Y ∈ ∂B by P W (X, Y ) = 1 − |X| 2

|W | w d π(X) π(Y ) X

w∈W

(w)

|X − w · Y | d (7)

4. The Green function of the unit ball is given by G W B (X, Y ) = 1

|W | π(X) π(Y ) X

w∈W

(w)G B (X, Y ), (8)

where G B (X, Y ) is the classical Green function of the unit ball B in R d (refer to Table 1).

P r o o f. It is based on the following steps:

1. Write a kernel on R d where d is the rank of M ; 2. Exploit formula (3);

3. Apply the W -invariance (the kernels on M must be W -invariant).

(6)

We give the proof in the Poisson kernel case; the other proofs are similar. The Poisson kernel of B(0, 1) in the Euclidean case is

P (X, Y ) = 1 − |X| 2 w d |X − Y | d .

If u is harmonic with respect to ∆ W then π u is harmonic with respect to ∆

Rd

. Hence π(X) u(X ) =

Z

∂B

1 − |X | 2

w d |X − Y | d π(Y ) f (Y ) dY.

This is equivalent to u(X ) =

Z

∂B

1 − |X| 2 w d π(X ) π(Y )

1

|X − Y | d f (Y ) π(Y ) 2 dY.

The reproducing kernel w 1−|X|

2

d

π(X) π(Y ) 1

|X−Y|d

is not W -invariant. We write the last equation |W | times, replacing X by w · X

u(X ) = u(w · X) = Z

∂B

1 − |X | 2 w d π(w · X ) π(Y )

1

|w · X − Y | d f (Y ) π(Y ) 2 dY

= Z

∂B

1 − |X| 2 w d π(X ) π(Y )

(w)

|X − w · Y | d f (Y ) π(Y ) 2 dY and we sum up the |W | equations. We obtain

u(X ) = 1

|W | w d

Z

∂B

1 − |X| 2 π(X ) π(Y )

X

w∈W

(w)

|X − w · Y | d f (Y ) π(Y ) 2 dY.

The formula for the Newton kernel requires more care. Let ˜ u be the solution of the inhomogeneous Laplace equation on R d , then

u(X ) = P

w∈W (w) ˜ u(wX ) π(X)

solves the corresponding problem for ∆ W . We need however to show that lim X→∞ |u(X )| = 0. It is useful to note that the function ˆ u(X) = P

w∈W (w) ˜ u(wX) is skew-symmetric.

For J ⊆ {1, 2, . . . , n}, les A J = {x ∈ R n : |x i | > 1/2 for i ∈ J , |x i | < 1 for i ∈ J c }. Note that R d is the union of the open sets A J . Now, on A J with |J| ≥ 1 (so that X → ∞),

X→∞ lim

ˆ u(X ) π(x 1 , . . . , x d )

= lim

(x

i

)

i∈J→∞

ˆ u(X ) π(x 1 , . . . , x d )

= lim

∃i∈J,xi→0

ˆ

u((1/x i ) i∈J , (x i ) i∈J

c

) π((1/x i ) i∈J , (x i ) i∈J

c

)

= lim

|xi|<2,j∈J,|xi|>1,i∈Jc

,∃i∈J,x

i→0

Y

i∈J

|x i | d−1

ˆ

u((1/x i ) i∈J , (x i ) i∈J

c

) π((x i ) i∈J ) π((x i ) i∈J

c

) Q

i∈J,j∈J

c

(1 − x i x j ) . Observe that ˆ u((1/x i ) i∈J , (x i ) i∈J

c

) is continuous since lim X→∞ u(X ˜ ) = 0 and skew-symmetric in (x i ) i∈J and in (x i ) i∈J

c

. We remark also that it is zero when 1 − x i x j = 0, i ∈ J , j ∈ J c . Using Lemma 2.3 below, we can conclude that the term

ˆ

u((1/x i ) i∈J , (x i ) i∈J

c

) π((x i ) i∈J ) π((x i ) i∈J

c

) Q

i∈J,j∈J

c

(1 − x i x j )

is an analytic function so that it remains bounded and that the limit is 0. We are grateful to the anonymous referee for pointing out the need for additional justification in the Newton kernel case.

Copyright line will be provided by the publisher

(7)

Lemma 2.3 Let f : R 2 → R, h : R 2 → R two analytical functions such that h

−1

({0}) ⊂ f

−1

({0}).

Suppose that for each z 0 = (x 0 , y 0 ) ∈ h

−1

({0}), the order of y 0 as a zero of h(x 0 , ·) is one (i.e.

h(x 0 , y) = (y − y 0 )˜ h(y), ˜ h analytical, h(y ˜ 0 ) 6= 0).

Then there exists an analytical function g : R 2 → R such that f = hg.

P r o o f. We apply the Weierstrass division theorem ([8, Th. 0.43(2)], [33, Th.6.1.3(1)]). For each z 0 = (x 0 , y 0 ) ∈ h

−1

({0}) there exists a neighbourhood V z

0

and analytical functions v z

0

(x, y) and b 1 (x) such that

f (x, y) = h(x, y)v z

0

(x, y) + b 1 (x).

For all (x, y) ∈ h

−1

(0) ∩ V z

0

the last equality gives 0 = 0 + b 1 (x), so that f (x, y) = h(x, y)v z

0

(x, y), (x, y) ∈ V z

0

. An application of the principle of identity ends the proof.

Remark 2.4 The properties of factorization of analytical functions of several real variables are not as straightforward as one might hope. For example, consider f (x, y) = y 3 which is zero whenever x 2 + y 2 = 0. However, it is not true that f divided by x 2 + y 2 is analytic or even defined.

For the root systems of type A, we obtain the following determinantal formula for the heat kernel on M . This formula may be also deduced from the formula for the transition function of the Dyson Brownian Motion, based on the Doob transform and Karlin-MacGregor formula, see Section 5.

Corollary 2.5 Consider the flat complex symmetric space M with the root system Σ = A d−1 . Let g t (u, v) =

1

4 π t e

−|u−v|2

/4t be the 1-dimensional classical heat kernel. The heat kernel on M is given by p W t (X, Y ) = 1

|W | π(X) π(Y ) det (g t (x i , y j )) (9)

where x 1 , . . . , x d are the coordinates of X and y 1 , . . . , y d are the coordinates of Y . P r o o f. Formula (9) follows from Theorem 2.2 (1) and the definition of determinant.

Remark 2.6 In [18], Grabiner computes determinant formulas for the transition probabilities of the Dyson Brownian motion in the Weyl chambers of A n−1 , B n , C n and D n .

Note that the alternating sum formula (4) reduces to a determinant if and only if the kernel K(X, Y ) has a multiplicative form

K(X, Y ) =

d

Y

i=1

k(x i , y i ).

This holds true for the transition probabilities of the Brownian Motion on R d or, more generally, of any multidimensional stochastic process X (t) with independent identically distributed components X i (t).

Let us resume the method of alternating sums, applied in the proof of Theorem 2.2. An Euclidean kernel K(X, Y ) (heat kernel, potential kernel, Poisson kernel, . . . ) for the Laplacian ∆

Rd

is transformed in the following way into the kernel K W acting on W -invariant functions on M :

K W (X, Y ) = 1

|W | π(X) π(Y ) X

w∈W

(w)K(X, w · Y ). (10)

Formula (5) is immediate from the explicit form of the heat kernel in Dunkl theory (refer to [40]) together with Proposition 2.7 below. The formulas (6)-(9) are new.

However, in the harmonic analysis of flat symmetric spaces of complex type, the alternating sum

formula (11) for a spherical function on M given below is well known (see [25, Chap. IV, Proposition

4.8 and Chap. II, Theorem 5.35]). Dunkl had provided a proof for the root system A n−1 in [10] using

a similar approach as ours.

(8)

Proposition 2.7 Given λ ∈ a

C

(the dual of the complexification of a), the spherical function ψ λ (X ) on M is given by the formula

ψ λ (X ) = π(ρ) 2 γ π(λ) π(X )

X

w∈W

(w)e

hλ,w·Xi

, (11)

where ρ = 1 2 P

α∈Σ

+

m α α = P

α∈Σ

+

α and γ = |Σ + | is the number of positive roots.

We finish this section with a relation between the heat kernel and spherical functions which will be useful in stochastic applications of our results, see Proposition 4.10 and Corollary 5.8. Proposition 2.8 is an immediate consequence of well-known results by R¨ osler in Dunkl theory (see for instance [40, Lemma 4.5]) and of [9], ensuring that heat kernel and spherical functions on flat complex symmetric spaces coincide with their Weyl group invariant analogues in Dunkl analysis when the multiplicity k = 1.

Proposition 2.8 Let M be a flat symmetric space of complex type. The following formula holds p W t (X, Y ) = 1

|W | 2 d π d/2 π(ρ) t

d2−γ

e

−|X|

2−|Y|2

4t

ψ X

Y 2t

. (12)

Remark 2.9 We provide here a simple explanation for the constant occurring in (12). From (5) and (11),

p W t (X, Y ) = 1

|W | (4 π t) d/2 π(X) π(Y ) X

w∈W

(w)e

|X−w·Y|

2 4t

= 1

|W | 2 d π d/2 π(ρ) t

d2−γ

π(ρ) 2 γ π(X) π Y 2t

X

w∈W

(w) e

hX,w·Y2t i

= 1

|W | 2 d π d/2 π(ρ) t

d2−γ

e

−|X|

2−|Y|2

4t

ψ X

Y 2t

.

Note that the constants in [40] lead to the same constant as in (12) even though the functional ρ is not used in the context of Dunkl theory. The same phenomenon will appear for the constant for the Poisson kernel, see Remark 3.5.

3 Asymptotic behavior of the kernels

To simplify the notation, we will write f Y

0

g if lim X→Y

0

f(X) g(X) = 1.

The main results of this Section are Theorems 3.11 and 3.13 which give asymptotics of the Poisson and Newton kernels of the flat complex symmetric space M . In their proofs, we need some knowledge of Dunkl analysis on R d .

Consider R d with a root system Σ. The basic information on the Dunkl analysis in this context can be found in [42]. Denote the Dunkl Laplacian by ∆ k and the intertwining operator by V k .

Recall now the formula of Dunkl ([11, 12]) for the Dunkl-Poisson kernel of the unit open ball B = B(0, 1).

P k (X, Y ) = 2 2 γ (d/2) γ

π(ρ) |W | w d V k

1 − |X| 2

(1 − 2hX, ·i + |X| 2 ) γ+d/2

(Y ), X ∈ B, Y ∈ ∂B, γ = X

α∈Σ

k α . (13) The constant in (13) is different from the one given in [11, 12]. Our constant is explained below in Remark 3.5.

The flat complex symmetric spaces M correspond to the formula (13) in the W -invariant case and with k α = 1. Then γ = |Σ + | expresses the number of positive roots.

A formula for the Dunkl-Newton kernel N k (X, Y ), analogous to (13), was proven in [16].

Copyright line will be provided by the publisher

(9)

3.1 Poisson kernel of the flat complex symmetric space The following technical results will prove useful further on.

Lemma 3.1

∂(π) |X |

−d

= 2 γ

γ−1

Y

k=0

(−d/2 − k) π(X) |X|

−d−2

γ

∂(π) log |X | = (−2) γ−1 (γ − 1)! π(X) |X|

−2

γ .

P r o o f. We see easily that |X| d+2 γ ∂(π) |X |

−d

is a skew polynomial of degree at most γ. It must therefore be a constant multiple of π(X ). Note from the definition of ∂(π) that

∂(π) f (X) = Y

α>0

∂t α

t

α

=0

f (X + X

α>0

t α H α )

where H α is defined by the relation hX, H α i = α(X ) for X ∈ a. Hence,

∂(π) |X |

−d

= Y

α>0

∂t α

t

α

=0

hX + X

α>0

t α H α , X + X

α>0

t α H α i

−d/2

.

After applying the operators ∂t

α

t

α

=0

, we will be left with the term (−d/2) (−d/2 − 1) · · · (d/2 − (γ − 1)) hX, Xi

−d/2−γ

Y

α>0

(2 α(X ))

and other terms which do not have the right form. This tells us that desired constant is 2 γ Q γ−1

k=0 (−d/2−

k).

A similar reasoning applies for the computation of ∂(π) log |X |.

Proposition 3.2 Let T (X, Y ) = π(X)π(Y 1 ) P

w∈W (w)

|X−w·Y|d

. Then T (0, Y ) = 2

2γ

π(ρ) (d/2)

γ

|Y |

−d−2

γ . P r o o f. Note first that ∂(π) X |X − Y |

−d

= 2 γ Q γ−1

k=0 (−d/2 − k) π(X − Y ) |X − Y |

−d−2

γ . Consider B(X, Y ) = π(X ) T (X, Y ) = π(Y 1 ) P

w∈W (w) |X −w·Y |

−d

. We apply the differential operator ∂(π)| X=0 to B . We find

∂(π)(π) T(0, Y ) = 2 γ

γ−1

Y

k=0

(−d/2 − k) 1 π(Y )

X

w∈W

(w) π(X − w · Y ) |X − w · Y |

−d−2

γ X=0

= (−1) γ 2 γ

γ−1

Y

k=0

(−d/2 − k) |W | |Y |

−d−2

γ .

Finally,

T (0, Y ) = (−1) γ 2 γ Q γ−1

k=0 (−d/2 − k) |W |

∂(π)(π) |Y |

−d−2

γ = (−1) γ 2 γ Q γ−1

k=0 (−d/2 − k) |W |

π(ρ) |W |/2 γ |Y |

−d−2

γ

= 2 2 γ (d/2) γ

π(ρ) |Y |

−d−2

γ .

(10)

Corollary 3.3 We have

P W (0, Y ) = 2 2 γ (d/2) γ π(ρ) |W | w d

N W (0, Y ) = −2 2 γ−1 (γ − 1)!

2 π |W | π(ρ) |Y |

−2

γ if d = 2 N W (0, Y ) = 2 2 γ ((d − 2)/2) γ

|W | (2 − d) w d π(ρ) |Y | 2−d−2 γ if d ≥ 3

Proposition 3.4 The Poisson kernel of the unit ball on the flat complex symmetric space M is given by

P W (X, Y ) = 2 2 γ (d/2) γ

π(ρ) |W | w d A

1 − |X| 2

(1 − 2 hX, ·i + |X| 2 ) γ+d/2

(Y ), (14)

where A

denotes the dual Abel transform on M .

Proposition 3.4 will be essential to establish (17) in Theorem 3.11. Recall that the dual of the Abel transform can be defined by the equation

A

(f)(X ) = Z

K

f (π

a

(Ad(k) X )) dk

where, as before, π

a

is the orthogonal projection from p to a with respect to the Killing form. Note in particular that A

(e λ(·) )(X ) = ψ λ (X ). Note also (see [25, Ch. IV, Theorem 10.11]) that unless C(X ) reduces to {X }, there exists a density K(H, X) such that

A

(f)(X ) = Z

C(X)

f (H ) K(H, X) dH.

Proof of Proposition 3.4. It should be noted that for Weyl-invariant f , A

(f ) = V k (f ) (refer to [9]).

Since the argument of A

in (14) is not Weyl-invariant, some proof is needed. Let K(Z, Y ) be the kernel of the dual Abel transform. Using (13), we have

P W (X, Y ) = P

w,w

0∈W

P k (w · X, w 0 · Y )

|W | 2 (with k = 1)

= 2 2 γ (d/2) γ π(ρ) |W | 3 w d

X

w,w

0∈W

Z

C(w

0·Y

)

1 − |w · X | 2

(1 − 2 hw · X, Zi + |w · X| 2 ) γ+d/2w

0·Y

(Z )

= 2 2 γ (d/2) γ

π(ρ) |W | 3 w d (1 − |X | 2 ) Z

C(Y )

X

w,w

0∈W

1

(1 − 2 hw · X, Zi + |X | 2 ) γ+d/2 dµ Y (w

−1

0 · Z)

= 2 2 γ (d/2) γ π(ρ) |W | 3 w d

(1 − |X | 2 ) Z

C(Y )

Weyl-invariant

z }| {

X

w,w

0∈W

1

(1 − 2 hw · X, w 0 · Zi + |X| 2 ) γ+d/2Y (Z)

= 2 2 γ (d/2) γ

π(ρ) |W | 3 w d (1 − |X | 2 ) Z

C(Y )

X

w,w

0∈W

1

(1 − 2 hw · X, w 0 · Zi + |X| 2 ) γ+d/2 K(Z, Y ) dZ

= 2 2 γ (d/2) γ

π(ρ) |W | 3 w d

(1 − |X | 2 ) Z

C(Y )

X

w,w

0∈W

1

(1 − 2 hX, w

−1

w 0 · Z i + |X| 2 ) γ+d/2 K(Z, Y ) dZ

= 2 2 γ (d/2) γ π(ρ) |W | 3 w d

(1 − |X | 2 ) Z

C(Y )

X

w,w

0∈W

1

(1 − 2 hX, Zi + |X | 2 ) γ+d/2 K(w 0

−1

w Z, Y ) dZ

= 2 2 γ (d/2) γ

π(ρ) |W | w d (1 − |X | 2 ) Z

C(Y )

1

(1 − 2 hX, Zi + |X| 2 ) γ+d/2 K(Z, Y ) dZ.

Copyright line will be provided by the publisher

(11)

Remark 3.5 Note that our normalizing constant is different from what is found in [11, 12]. We ex- plain here how they correspond in the complex case. In [11], the Poisson kernel P W (X, Y ) is normalized in the following manner:

u(X ) = c

0

d Z

∂B(0,1)

f (y) P W (X, Y ) π(Y ) 2 dY w d where c

0

d is such that

1 = c

0

d Z

∂B(0,1)

π(Y ) 2 dY w d

where, reading through [11, Page 1215], c

0

d = 2 γ (d/2) γ

Y

α>0

|α| 2

2 (hα, ρi/|α| 2 + 1)

−1

. Our different normalizations come down to the equality

2 2 γ (d/2) γ

π(ρ) |W | w d

= c

0

d w d

which gives the interesting equality π(ρ) |W |

2 γ = Y

α>0

|α| 2

2 (hα, ρi/|α| 2 + 1)

.

This equality is easily verified directly for the classical Lie algebras and for g 2 (the other exceptional Lie algebras require more work). It should be noted that in [11], Dunkl used the notation ν h instead of ρ but refers to the fact that Opdam uses ρ in [38].

Corollary 3.6 The Newton kernel of the flat complex symmetric space M is given by N W (X, Y ) = 2 2 γ ((d − 2)/2) γ

|W | (2 − d) w d π(ρ) A

(|Y | 2 − 2 hX, ·i + |X | 2 ) (2−d−2 γ)/2 (Y ).

P r o o f. We apply the same computations as for the Poisson kernel to formula [16, (6.1)] (the constant has been adjusted to follow our conventions as per Remark 3.5).

We now start to study the asymptotic behavior of the Poisson kernel P W (X, Y ). Let us introduce some notations. We define

R(X, Y ) = X

w∈W

(w)

|X − w · Y | d and T (X, Y ) = R(X, Y ) π(X )π(Y ) and therefore,

P W (X, Y ) = 1 − |X| 2

|W | w d

T(X, Y ).

The function R(X, Y ) is defined for X, Y ∈ a such that X 6∈ W · Y = {w · Y | w ∈ W }. We will denote this domain by

D := {(X, Y ) ∈ a 2 | X 6∈ W · Y }

The function T (X, Y ) is, for now, defined for non-singular X , Y ∈ a (i.e. such no nonzero root vanish on X or on Y ) such that X 6∈ W Y . We will see in Proposition 3.10 that the function T (X, Y ) extends by continuity to an analytic function on the domain D.

Studying the properties of P W (X, Y ) is equivalent to studying the properties of T (X, Y ) and R(X, Y ).

We will give some of them in Proposition 3.10. We start by introducing two auxiliary results.

(12)

Lemma 3.7 Assume a 1 , . . . , a d are not all 0 and let U be an open set. Let q be an analytic function on U which is 0 whenever P d

k=1 a k x k = 0. Then q(X ) = P d

k=1 a k x k

r(X ) where r is a analytic function on U .

P r o o f. The Lemma follows from Lemma 2.3. We give here an elementary proof.

Using a change of variable, we can assume that a 1 = 1 and a i = 0 for i > 1. It is also enough to show that for every X 0 = (b 1 , . . . , b d ) ∈ U , there exists > 0 such that the result holds in the ball B(X 0 , ).

If X 0 6= 0, then pick > 0 small enough so that (x 1 , . . . , x d ) ∈ B(X 0 , ) implies x 1 6= 0. Then we can pick r(X) = q(X )/x 1 .

Suppose now that b 1 = 0. We then have q(x 1 , . . . , x d ) = x 1

r(X)

z }| { Z 1

0

∂x 1

q(t x 1 , x 2 , . . . , x d ) dt for (x 1 , . . . , x d ) ∈ B(X 0 , ) ⊂ U.

Proposition 3.8 Let p(X ) = Q d

i=1 hα i , X i where no α i ’s is a multiple of another α j and let U be an open set. If q is an analytic function on U which is 0 whenever α i (X) = 0 for some i then q(X ) = p(X) r(X) where r is an analytic function on U .

P r o o f. We use induction on n. Lemma 3.7 shows that the result is true for n = 1. Assume it is true for n − 1, n ≥ 2 and write q(X) = Q n−1

i=1 hα i , Xi r(X). Since q(X) = 0 when hα n , Xi = 0, we conclude that r(X ) = 0 on the set {X|hα n , Xi = 0 and hα i , Xi 6= 0, i < n}. By continuity, we deduce that r(X) = 0 when hα n , Xi = 0 and, using Lemma 3.7 once more, we can conclude.

Remark 3.9 We thank the referee for pointing out that Lemma 3.7 and Proposition 3.8 can be also be proven with the help of the Weierstrass division theorem, via Lemma 2.3.

Proposition 3.10

1. (Symmetry in X and Y ) R(X, Y ) = R(Y, X) and T (X, Y ) = T (Y, X).

2. (Skew-symmetry) R(w 0 X, Y ) = (w 0 ) R(X, Y ) and R(X, w 0 Y ) = (w 0 ) R(X, Y ).

3. (Nullity of R on singular arguments) R(X, Y ) is zero whenever at least one of X or Y is singular.

4. (analytic factorization of R, analytic extension of T to D.) There exists a function f analytic on D such that R(X, Y ) = π(X )π(Y )f (X, Y ) on D. Equivalently, the function T extends to an analytic function on D.

5. (Non-nullity of T and P W ) When X ∈ B and Y ∈ ∂B then T (X, Y ) > 0 and P W (X, Y ) > 0.

P r o o f. The proof of (1) and (2) is straightforward.

(3) Suppose α(Y ) = 0. We use Property 2 and (σ α ) = −1 where σ α is the reflection with respect to the hyperplane {α = 0}. Since R(X, Y )/π(Y ) is analytic, the statement follows.

(4) This follows from Proposition 3.8.

(5) This follows from Proposition 3.4. The dual Abel integral transform of a strictly positive function is strictly positive.

Theorem 3.11 Let Y 0 ∈ ∂B, Σ

0

= {α ∈ Σ| α(Y 0 ) = 0} and Σ

0

+ = Σ

0

∩ Σ + . Then P W (X, Y 0 ) Y

0

2 2 γ

0

(d/2) γ

0

|W | w d π

0

0

) (π

00

(Y 0 )) 2

1 − |X | 2

|X − Y 0 |

0

+d (15)

where γ

0

= |Σ

0

+ | is the number of positive roots annihilating Y 0 , π

0

(Y ) = Q

α∈Σ

0+

hα, Y i and π

00

(Y ) = Q

α∈Σ

+\Σ0+

hα, Y i.

Copyright line will be provided by the publisher

(13)

P r o o f. Let W

0

= {w ∈ W |w · Y = Y }. In this proof, we consider X ∈ V = B(Y 0 , ) with > 0 fixed and chosen in such a way that

α( ¯ V ) ⊂ (0, ∞) for α ∈ Σ + \ Σ

0

+ and w V ∩ V = ∅ for every w ∈ W \ W

0

. Using Theorem 2.2, we have

P W (X, Y ) = 1

|W | w d

1 − |X | 2 π(X) π(Y )

X

w∈W

(w)

|X − w · Y | d . We consider X ∈ V \ {Y 0 } and we deal with

T (X, Y 0 )= |W | w d

1 − |X | 2 P W (X, Y 0 ) = 1 π(X ) π(Y 0 )

X

w∈W

(w)

|X − w · Y 0 | d = R(X, Y 0 )

π(X ) π(Y 0 ) . (16) By Proposition 3.10 applied to the root systems Σ and Σ

0

, all the expressions in (16) are well defined for X ∈ V \ {Y 0 }, if needed in the limit sense.

We decompose the sum P

w∈W into two terms, the first being the sum over the subgroup W

0

= {w ∈ W |w · Y 0 = Y 0 } which is the Weyl group of the root subsystem Σ

0

. We obtain

T (X, Y 0 ) = P

w∈W (w)|X − w · Y 0 |

−d

π(X ) π(Y 0 ) =

P

w∈W

0

(w)|X − w · Y 0 |

−d

π(X) π(Y 0 ) +

P

w∈W

\W0

(w)|X − w · Y 0 |

−d

π(X) π(Y 0 ) . By Proposition 3.10, all the expressions in the last formula are well defined for X ∈ V \ {Y 0 }, if needed

in the limit sense. Denote T 1 (X, Y 0 ) =

P

w∈W

0

(w)|X − w · Y 0 |

−d

π(X) π(Y 0 ) and T 2 (X, Y 0 ) = P

w∈W

\W0

(w)|X − w · Y 0 |

−d

π(X) π(Y 0 ) . Let π

0

(X ) = Q

α∈Σ

0+

α(X) and π

00

(X) = Q

α∈Σ

+\Σ0+

α(X). Observe that by Theorem 2.2, π

00

(X ) π

00

(Y 0 ) T 1 (X, Y 0 ) =

P

w∈W

0

(w)|X − w · Y 0 |

−d

π

0

(X ) π

0

(Y 0 ) = |W

0

| w d

1 − |X | 2 P W

0

(X, Y 0 )

where P W

0

(X, Y ) is the Poisson kernel for the flat symmetric space (R d , Σ

0

) corresponding to the complex root system Σ

0

. The convex hull C

0

(Y 0 ) = conv(W

0

Y 0 ) = {Y 0 }, so by Proposition 3.4 and the properties of A

,

1

1 − |X| 2 P W

0

(X, Y 0 ) = 2 2 γ

0

(d/2) γ

0

π(ρ

0

) |W

0

| w d

Z

C(Y

0

)

1

(1 − 2 hX, Zi + |X| 2 ) γ

0

+d/2 δ

{Y0}

(dZ )

= 2 2 γ

0

(d/2) γ

0

π

0

0

) |W

0

| w d

1

|X − Y 0 |

0

+d (17)

where X ∈ B ∩ V .

We now prove that the function X 7→ T 2 (X, Y 0 ) is bounded on V , which, together with (17), will conclude the proof. We denote by

N (X, Y ) = X

w∈W

\W0

(w)|X − w · Y |

−d

the numerator of T 2 . Observe that N(X, Y ) is an analytic function on V × V . The function T 2 (X, Y ) =

P

w∈W

\W0

(w)|X − w · Y |

−d

π(X) π(Y )

(14)

is well defined and analytic for (X, Y ) ∈ V × V \ D with D = {(X, Y ) ∈ a × a : X = Y }, since T (X, Y ) and T 1 (X, Y ) have these properties by Proposition 3.10 and T 2 = T − T 1 .

This implies that if X

0

∈ V or Y

0

∈ V are singular (i.e. α(X

0

) = 0 or α(Y

0

) = 0 for some α ∈ Σ

0

+ ) and X

0

6= Y

0

then the numerator N (X

0

, Y

0

) = 0 since otherwise the limit N (X, Y )/π(X)π(Y ) could not exist when (X, Y ) → (X

0

, Y

0

).

We deduce that if X

0

∈ V or Y

0

∈ V and α(X

0

) = 0 or α(Y

0

) = 0 for some α ∈ Σ

0

+ ) then N (X

0

, Y

0

) = 0. This is also true for X

0

= Y

0

since such points are limits when t tends to 1 of (t X

0

, Y

0

) with singular t X

0

6= Y

0

and N(t X

0

, Y

0

) converges to N(X

0

, Y

0

).

By Proposition 3.8, there exists a function F(X, Y ) analytic on V × V such that N (X, Y ) = π

0

(X)π

0

(Y )F(X, Y ), X, Y ∈ V

and, finally,

T 2 (X, Y ) = F (X, Y )

π

00

(X)π

00

(Y ) , X, Y ∈ V

(we have min X∈ V ¯ π

00

(X) > 0 since π

00

( ¯ V ) ⊂ (0, ∞)). In particular, the function X 7→ T 2 (X, Y 0 ) is bounded on V .

Remark 3.12 For the asymptotic properties of P W , besides the alternating sum formula, the ap- proach via the Dunkl formula (13) and dual Abel transform, i.e. Proposition 3.4 is needed. We use it to compute the leading term T 1 (X, Y ) in T (X, Y ).

3.2 Asymptotic behavior of the Newton kernel on flat complex symmetric spaces

Using the same approach as in the proof of Theorem 3.11 together with Corollary 3.6, we conclude that Theorem 3.13 Let Y 0 ∈ a + . If d = 2 and α, β are the simple roots then

N W (X, 0) = −2 2 γ−1 (γ − 1)!

2 π |W | π(ρ) |X |

−2

γ (case Y 0 = 0), N W (X, Y 0 ) Y

0

−2 2 γ

0−1

0

− 1)!

2 π |W | π

00

(Y 0 ) 2 hα, αi |X − Y 0 |

−2

where Y 0 6= 0, α(Y 0 ) 6= 0 and β(Y 0 ) = 0.

If d ≥ 3

N W (X, Y 0 ) Y

0

2 2 γ

0

((d − 2)/2) γ

0

|W | (2 − d) w d π

0

0

) (π

00

(Y 0 )) 2

1

|X − Y 0 |

0

+d−2 . (18) Here γ

0

= |Σ

0

+ | is the number of positive roots annihilating Y 0 and π

00

(Y ) = Q

α∈Σ

+\Σ0+

hα, Y i.

Remark 3.14 In the paper [20] exact estimates of the Poisson and Newton kernels P W and N W were proven complementing the results of Theorem 3.11 and Theorem 3.13. For the Poisson kernel it is proven that

P W (X, Y ) P

Rd

(X, Y ) Q

α∈Σ

+

|X − σ α Y | 2

where σ α is the symmetry with respect to the hyperplane perpendicular to α.

Copyright line will be provided by the publisher

(15)

4 Asymptotic behavior of spherical functions on flat complex sym- metric spaces

In this section we consider spherical functions on M , satisfying the formula ψ λ (Y ) = π(ρ)

2 γ π(λ) π(Y ) X

w∈W

(w)e

hλ,w·Yi

, λ ∈ a

C

, Y ∈ a

R

. (19)

Note that our notation is different from that of Helgason (in his notation the function given by (19) is denoted ψ

−i λ

).

The following technical lemma will prove useful later in this section.

Lemma 4.1 Suppose G 1 and G 2 are subgroups of the finite group G. Then |G 1 G 2 | |G 1 ∩ G 2 | =

|G 1 | |G 2 |.

P r o o f. The group G 1 × G 2 acts on the set G 1 G 2 ⊂ G via (g 1 , g 2 )(g) = g 1 g g 2

−1

. Clearly the action is transitive. The stabilizer of e ∈ G 1 G 2 (e being the identity) is easily seen to be isomorphic to G 1 ∩G 2 . The orbit-stabilizer theorem ([43, Theorem 5.8]) implies then that |G 1 G 2 | |G 1 ∩ G 2 | = |G 1 | |G 2 |.

We introduce here some notation. If X ∈ a, we denote by Σ + X the positive root system Σ + X = {α ∈ Σ + : α(X ) = 0} and by W X the Weyl group generated by the symmetries s α with α ∈ Σ + X (consequently, W X = {w ∈ W : w · X = X}). We also write π X (Y ) = Q

α∈Σ

+X

α(Y ) and c X = ∂(π X )(π X ) (this derivative is constant on a).

For X ∈ a we define the polynomial π

0

X (Y ) by π(Y ) = π X (Y )π X

0

(Y ). Denote W (λ 0 , Y 0 ) = {w ∈ W : hλ 0 , w · Y 0 i = hλ 0 , Y 0 i}.

Remark 4.2 We conjecture that the property W (λ 0 , Y 0 ) = W λ

0

W Y

0

is valid for all root systems. In Appendix A, we provide a series of proofs that cover all cases except for the exceptional root systems of type E. We also point out that if one of λ 0 or Y 0 is regular then this property is also verified, see [23].

Denote the Weyl subgroup W λ

0

,Y

0

= W λ

0

∩ W Y

0

= {w ∈ W : w · λ 0 = λ 0 and w · Y = Y }. The group W λ

0

,Y

0

corresponds to the root system Σ + λ

0

,Y

0

= Σ + λ

0

∩ Σ + Y

0

. We write π 0 (Y ) = π λ

0

,Y

0

(Y ) = Q

α∈Σ

+λ

0,Y0

α(Y ) and c λ

0

,Y

0

= ∂(π λ

0

,Y

0

)(π λ

0

,Y

0

). Denote by M the set of positive roots that are neither in Σ + λ

0

nor in Σ + Y

0

, i.e. M = Σ + \ (Σ + λ

0

∪ Σ + Y

0

). We also write π

M

(X) = Q

α∈M α(X ).

Proposition 4.3

(i) If w ∈ W Y then π Y (w · X) = (w) π Y (X ).

(ii) If w ∈ W Y then π Y (∂)[f (w · Y )] = (w)(π Y (∂)f )(w · Y ).

P r o o f. The property (i) is well known [24]. The property (ii) is straightforward for f (X ) = e

hZ,Xi

and extends by linear density.

Proposition 4.4 Let λ 0 , Y 0 be singular. The asymptotics of ψ λ

0

(t Y 0 ) when t → ∞ is given by the following formula:

ψ λ

0

(t Y 0 ) ∼ C(λ 0 , Y 0 )t

+Y0|−|Σ+|

X

w∈W (λ

0

,Y

0

)

(w)π Y

0

(∂ Y )

π λ

0

(w · Y )e

0

,w·Y

i

Y =t Y

0

(20) where C(λ 0 , Y 0 ) = (c λ

0

c Y

0

π

0

λ

0

(λ 0 ) π Y

00

(Y 0 ))

−1

.

When W (λ 0 , Y 0 ) = W λ

0

W Y

0

, the last formula simplifies to ψ λ

0

(t Y 0 ) ∼ C 10 , Y 0 )t

+Y0|−|Σ+|

π Y

0

(∂ Y )

π λ

0

(Y )e

0

,Y

i

Y =t Y

0

(21)

where C 1 (λ 0 , Y 0 ) = C(λ 0 , Y 0 )|W λ

0

| |W Y

0

|/|W λ

0

,Y

0

|.

(16)

P r o o f. We start with the alternating sum formula for the spherical function ψ λ , written in the following way

π(λ)π(Y )ψ λ (Y ) = X

w∈W

(w)e

hλ,w·Yi

(22)

We write π(λ) = π λ

0

(λ)π λ

00

(λ) and π(Y ) = π Y

0

(Y )π

0

Y

0

(Y ). We apply the operator L = π Y

0

(∂ Y )π λ

0

(∂ λ ) to both sides of (22). Using the fact that π λ

0

(∂ λ ) e

hλ,w·Yi

= π λ

0

(w · Y )e

hλ,w·Yi

, we obtain

c λ

0

c Y

0

π

0

λ

0

(λ 0 ) π Y

00

(t Y 0 ) ψ λ

0

(t Y 0 ) = X

w∈W

(w) π Y

0

(∂ Y )

π λ

0

(w · Y ) e

0

,w·Y

i

Y =t Y

0

. In order to get the exact asymptotics of ψ λ

0

(t Y 0 ), we only need to deal with w ∈ W such that hλ 0 , w · Y 0 i = hλ 0 , Y 0 i. This gives the asymptotics (20).

We now assume that W (λ 0 , Y 0 ) = W λ

0

W Y

0

. The asymptotics (20) simplify, since by Proposition 4.3, we obtain for w = w 1 w 2 with w 1 ∈ W λ

0

and w 2 ∈ W Y

0

π Y

0

(∂ Y )

π λ

0

(w · Y ) e

0

,w·Y

i

= (w 1 )π Y

0

(∂ Y )

π λ

0

(w 2 Y ) e

0

,w

2

Y

i

= (w 1 ) (w 2Y

0

(∂ Y )

π λ

0

(Y ) e

0

,Y

i

= (w) π Y

0

(∂ Y )

π λ

0

(Y ) e

0

,Y

i

.

Using Lemma 4.1, we have |W λ

0

W Y

0

| = |W λ

0

||W Y

0

|/|W λ

0

,Y

0

|. We obtain the formula (21).

Theorem 4.5 Let λ 0 , Y 0 be singular. Assume that W (λ 0 , Y 0 ) = W λ

0

W Y

0

. Then the asymptotics of ψ λ

0

(t Y 0 ) when t → ∞ are given by the following formula:

ψ λ

0

(t Y 0 ) ∼ D(λ 0 , Y 0 ) t

−m

e t

0

,Y

0i

(23) where m is the number of positive roots that are neither in Σ + λ

0

nor in Σ + Y

0

i.e.

m = cardM = |Σ| + − (|Σ + λ

0

| + |Σ + Y

0

| − |Σ + λ

0

∩ Σ + Y

0

|) and

D(λ 0 , Y 0 ) = c λ

0

,Y

0

c λ

0

c Y

0

|W λ

0

| |W Y

0

|

|W Y

0

∩ W λ

0

|

1

π

M

0 ) π

M

(Y 0 ) .

Remark 4.6 When Y 0 is regular, the method of proof used in Theorem 3.11 for the asymptotics of the Poisson kernel could have been used here. When both λ 0 and Y 0 are singular, that approach fails to apply.

P r o o f. Using Leibniz formula, we have π Y

0

(∂ Y )

π λ

0

(Y ) e

0

,Y

i

Y =t Y

0

= π 0 (∂ Y ) Y

α∈Σ

+Y

0\Σ+λ

0

Y (A α )

π λ

0

(Y )e

0

,Y

i

Y =t Y

0

= π 0 (∂ Y )

 Y

α∈Σ

+Y

0\Σ+λ

0

0 , αi π λ

0

(Y ) e

0

,Y

i

+ P(Y ) e

0

,Y

i

Y =t Y

0

Copyright line will be provided by the publisher

(17)

The number of factors in each term P(Y ) of the form hη, Y i where η is a root, is strictly less than the number of factors in π λ

0

i.e. less than |Σ + λ

0

|.

In the expression in the last line, all derivatives involving the term e

0

,Y

i

give 0 since β (λ 0 ) = 0 for β ∈ Σ + Y

0

∩ Σ + λ

0

.

In the derivatives of π λ

0

(Y ), any term that contains hβ, Y i with β ∈ Σ + Y

0

∩ Σ + λ

0

will be zero when Y is replaced by t Y 0 . Thus, for a nonzero result, the operator π 0 (∂ Y ) must be applied to π 0 (Y ), what gives c λ

0

,Y

0

> 0. We obtain

π Y

0

(∂ Y )

π λ

0

(Y ) e

0

,Y

i

Y =t Y

0

= Y

α∈Σ

+Y

0\Σ+λ

0

hλ 0 , αi Y

γ∈Σ

+λ

0\Σ+Y

0

hγ, t Y 0 i c λ

0

,Y

0

e t

0

,Y

0i

+ π 0 (∂ Y )P(t Y 0 )e thλ

0

,Y

0i

= c λ

0

,Y

0

t

+λ0|−|Σ+λ0∩Σ+Y0|

Y

α∈Σ

+Y

0\Σ+λ

0

hλ 0 , αi Y

γ∈Σ

+λ

0\Σ+Y

0

hγ, Y 0 i e t

0

,Y

0i

+ negligible terms.

We labeled as “negligible terms” the terms with the derivatives involving P (Y ). They have the number of factors of the form hη, t Y 0 i strictly less then |Σ + λ

0

| − |Σ + λ

0

,Y

0

|, so strictly less than the term Q

γ∈Σ

+λ

0\Σ+Y

0

hγ, t Y 0 i. The rest follows from the definition of C(λ 0 , Y 0 ).

Remark 4.7 We can give a more explicit expression for the constant D, using the formula

∂(π) π = |W | π(ρ) 2 γ , where ρ = 1 2 P

α∈Σ

+

m α α = P

α∈Σ

+

α and γ = |Σ + | is the number of positive roots.

For X ∈ a, denote p X = π XX ). Analogously, we define p X

1

,X

2

for the root system annihilating both elements X 1 , X 2 ∈ a. We have

D(λ 0 , Y 0 ) = 2 γ

λ0,Y0−γλ0−γY0

π

M

(λ 0 ) π

M

(Y 0 )

p λ

0

,Y

0

p λ

0

p Y

0

, and therefore

t→∞ lim

ψ λ

0

(t Y 0 )

t

−m

e thλ

0

,Y

0i

= 2 γ

λ0,Y0−γλ0−γY0

π

M

(λ 0 ) π

M

(Y 0 )

p λ

0

,Y

0

p λ

0

p Y

0

Remark 4.8 As a quick application of Theorem 4.5, we find, in the flat complex case, a simple proof of a general result of Vogel and Voit: for symmetric spaces with subexponential (here polynomial) growth, the set of bounded spherical functions coincides with the support of the Plancherel measure of the associated Gelfand pair (G 0 , K), G 0 the Cartan motion group. See for instance Sections 3.2 and 3.3. of [41] for details. A proof in the flat complex case was also proposed by Helgason in [23].

Remark 4.9 Taking into account the relationship between the spherical functions in the flat case and those in the curved case for the complex Lie groups, the estimates of spherical functions in [36, 44]

extend to the flat case.

In this case, Theorem 4.5 completes the estimates of [36, 44] providing the exact asymptotics. We conjecture that asymptotics with appropriate constants and not only estimates hold in the results of Narayanan, Pasquale and Pusti [36] and Schapira [44].

The asymptotic expansion given in [2, Proposition 3.8] for regular λ 0 and Y 0 implies asymptotics of spherical functions in this case. Theorem 4.5 strengthens this result to singular λ 0 and Y 0 .

Proposition 4.10 Let X and Y be singular and m

0

= |Σ + X ∪ Σ + Y |. With the same notation as in Theorem 4.5, we have

p W t (X, Y ) ∼ D(X, Y ) 2 m−d

|W | π d/2 π(ρ) t

d2−m0

e

−|X−Y|

2 4t

as t → 0 + .

(18)

P r o o f. From Theorem 4.5, we have for t > 0 close to 0, ψ X (Y /(2 t)) ∼ D(X, Y ) (2 t) m e

hX,Y /(2

t)i . Combined with (12), this leads us to

p W t (X, Y ) ∼ 2 m

|W | 2 d π d/2 π(ρ) t

d2−γ

e

−|X|

2−|Y|2

4t

D(X, Y ) t m e

hX,Y /(2

t)i

= D(X, Y ) 2 m

|W | 2 d π d/2 π(ρ) t

d2−(γ−m)

e

−|X−Y|

2 4t

.

5 Applications to the Dyson Brownian motion and stochastic anal- ysis

5.1 Definition and transition density of the Dyson Brownian motion

When a probabilist looks at formula (3), he or she sees in it the generator of the Doob h-transform (refer to [39]) of the Brownian Motion on R d with the excessive function h(X ) = π(X). For the root system A d on R d , the operator ∆ W restricted to functions on a + , is the generator of the Dyson Brownian Motion on a + ⊂ R d ([14]), i.e. the d Brownian independent particles B t (1) , . . . , B t (d) conditioned not to collide. More generally, for any root system Σ on R d , the construction of a Dyson Brownian Motion as a Brownian Motion conditioned not to touch the walls of the positive Weyl chamber, can be done ith a starting point X ∈ a + ([18]).

Let us recall basic facts about the Doob h-transform and the Dyson Brownian Motion. Let Σ be a root system on R d and π(X ) = Q

α>0 hα, Xi. It is known that π is ∆

Rd

-harmonic on R d ([18]), so in particular π is excessive.

Definition 5.1 Let Σ be a root system on R d and π(X) = Q

α>0 hα, Xi. The Dyson Brownian Motion D t Σ on the positive Weyl chamber a + is defined as the h-Doob transform of the Brownian Motion on R d , with h = π, i.e. its transition density is equal to

p D t (X, Y ) = π(Y )

π(X) p killed t (X, Y ), X ∈ a + , Y ∈ a + , (24)

where p killed t (X, Y ) is the transition density of the Brownian Motion killed at the first strictly positive time of touching ∂a + .

The infinitesimal generator of D Σ t is given by the formula ([39])

D f = π

−1

Rd

(πf ), supp f ⊂ a + , (25)

which coincides on a + with formula (3) for ∆ W . The only differences with the symmetric flat complex case are that the domain of kernels K D (X, Y ) is restrained to X ∈ a + , Y ∈ a + , and that no invariant measure π 2 (Y ) dY appears for the integral kernels in the Dyson Brownian Motion case. Consequently, we obtain

Corollary 5.2 The transition density and the heat kernel of the Dyson Brownian Motion D Σ t on a + ⊂ R d is given by the formula

p D t (X, Y ) = π(Y ) π(X )

X

w∈W

(w) h t (X − w · Y ), X ∈ a + , Y ∈ a + ,

where h t (X − Y ) = (4 π t) 1

d/2

e

|X−Y|

2

4t

is the Euclidean heat kernel on R d .

Copyright line will be provided by the publisher

Références

Documents relatifs

the normal rule of quantization Ref.. j~ is called the contravariant Weyl symbol of the operator A. [1 ]) the isometry group acts on the covariant symbols of..

This is to say, we get an expression for the Itô excursion measure of the reflected Langevin process as a limit of known measures.. This result resembles the classical approximation

L’accès aux archives de la revue « Annali della Scuola Normale Superiore di Pisa, Classe di Scienze » ( http://www.sns.it/it/edizioni/riviste/annaliscienze/ ) implique l’accord

— We find a characterization of the spherical Fourier transform of K- invariant probability measures on Riemannian symmetric spaces of noncompact type in the complex case.. To do

For complete Riemannian manifolds (M^g) of dimension n {n &gt; 3) that are projectively symmetric and projective homogeneous (it is shown with an example that projective homogeneity

In his paper [T2] Terras obtained a central limit theorem for rotation- invariant random variables on the space of n x n real positive definite (symmetric) matrices in the case n =

Denote by Ao the Abel transform with respect to the Riemannian sym- metric space GQ/KQ. Since it is well known how to invert the transform Ao^ at least for &#34;good&#34; functions,

In this section we first state and comment a global estimate for the heat kernel on X = G/K (same upper and lower bound, up to constants) that was conjectured by the first author