03.11.2014 Views

arbres aléatoires, conditionnement et cartes planaires - DMA - Ens

arbres aléatoires, conditionnement et cartes planaires - DMA - Ens

arbres aléatoires, conditionnement et cartes planaires - DMA - Ens

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

THÈSE DE DOCTORAT DE L’UNIVERSITÉ PARIS 6<br />

Spécialité<br />

Mathématiques<br />

Présentée par<br />

Mathilde Weill<br />

Pour obtenir le grade de<br />

DOCTEUR de l’UNIVERSITÉ PARIS 6<br />

Suj<strong>et</strong> de la thèse :<br />

ARBRES ALÉATOIRES, CONDITIONNEMENT<br />

ET<br />

CARTES PLANAIRES<br />

soutenue le 8 Décembre 2006 devant le jury composé de<br />

M. Jean Bertoin Examinateur<br />

M. Philippe Chassaing Rapporteur<br />

M. Thomas Duquesne Examinateur<br />

M. Jean-François Le Gall Directeur de thèse<br />

M. Wendelin Werner Examinateur


Remerciements<br />

Mes remerciements vont tout d’abord à Jean-François Le Gall qui a dirigé ma thèse pendant<br />

ces trois dernières années. Il m’a guidée attentivement tout au long de mon travail, prodigué de<br />

nombreux conseils, patiemment expliqué <strong>et</strong> appris beaucoup de mathématiques. Son exigence<br />

de rigueur <strong>et</strong> de clarté ont été <strong>et</strong> resteront pour moi un modèle. Si j’ajoute que son cours sur<br />

le mouvement brownien est l’un de ceux qui m’ont donné le goût des probabilités lors de ma<br />

première année d’études à l’ENS, cela ne donnera encore qu’une faible idée de ce que le présent<br />

travail lui doit.<br />

Philippe Chassaing <strong>et</strong> Jim Pitman ont accepté de rapporter ma thèse <strong>et</strong> je leur en suis très<br />

reconnaissante. Philippe Chassaing a de plus accepté de participer au jury, ce dont je le remercie<br />

tout particulièrement.<br />

Jean Bertoin est le second responsable de mon goût pour les probabilités. Son cours “Probabilités<br />

2” est l’un des meilleurs souvenirs de ma première année d’études à l’ENS. Sa présence<br />

dans le jury est un grand plaisir pour moi <strong>et</strong> je l’en remercie chaleureusement.<br />

Thomas Duquesne <strong>et</strong> Wendelin Werner ont accepté de faire partie du jury <strong>et</strong> je les en remercie<br />

vivement. L’intérêt amical qu’ils ont manifesté pour ma thèse tout au long de son élaboration<br />

m’a été précieux.<br />

Je tiens ici à remercier tous les membres du département de mathématiques de l’ENS, qui m’a<br />

offert un cadre idéal pour la rédaction de c<strong>et</strong>te thèse : Patricia qui partage mon bureau depuis<br />

trois ans ; Florent, Nathanaël <strong>et</strong> Mathieu en souvenir du mois de Juill<strong>et</strong> 2003 ; mes collègues<br />

caïmans <strong>et</strong> en particulier Marie, Arnaud <strong>et</strong> Raphaël ; les voisins du “passage vert” Gilles <strong>et</strong> les<br />

Philippe; Benoît, David, Frédéric <strong>et</strong> Guy du “passage bleu”, François <strong>et</strong> Marc... Je souhaite<br />

aussi remercier Bénédicte, Lara, Laurence <strong>et</strong> Zaïna pour leur aide, <strong>et</strong> le spi pour sa disponibilité.<br />

J’adresse enfin toute ma gratitude aux élèves pour avoir rendu mon travail de caïman si agréable.<br />

Je n’oublie pas les thésards <strong>et</strong> post-doctorants du laboratoire de Paris 6 que j’ai croisés<br />

pendant ces années : Christina, Eulalia, Emmanuel <strong>et</strong> les autres.<br />

Je termine par une pensée pour mes amis, les matheux : Béné <strong>et</strong> Greg, <strong>et</strong> les non-matheux :<br />

Hélhél, Olivia <strong>et</strong> Bab<strong>et</strong>h ; <strong>et</strong> bien sûr pour toute ma famille, en particulier mes parents, mon<br />

frère Romain, <strong>et</strong> Thierry, pour leur soutien constant.<br />

3


Table des matières<br />

Chapitre 1. Introduction 7<br />

1.1. Arbres généalogiques aléatoires 7<br />

1.2. Arbres continus régénératifs 11<br />

1.3. Arbres spatiaux aléatoires 12<br />

1.4. Arbre brownien conditionné 14<br />

1.5. Arbres spatiaux <strong>et</strong> <strong>cartes</strong> <strong>planaires</strong> 18<br />

1.6. Résultats asymptotiques pour de grandes <strong>cartes</strong> biparties enracinées aléatoires 23<br />

Chapitre 2. Regenerative real trees 27<br />

2.1. Introduction 27<br />

2.2. Preliminaries 29<br />

2.3. Proof of Theorem 2.1.1 35<br />

2.4. Proof of Theorem 2.1.2 44<br />

Chapitre 3. Conditioned Brownian trees 51<br />

3.1. Introduction 51<br />

3.2. Preliminaries 55<br />

3.3. Conditioning and re-rooting of trees 64<br />

3.4. Other conditionings 75<br />

3.5. Finite-dimensional marginal distributions under N 0 83<br />

Chapitre 4. Asymptotics for rooted planar maps and scaling limits of two-type spatial<br />

trees 89<br />

4.1. Introduction 89<br />

4.2. Preliminaries 90<br />

4.3. A conditional limit theorem for two-type spatial trees 98<br />

4.4. Separating vertices in a 2κ-angulation 117<br />

Bibliographie 123<br />

5


CHAPITRE 1<br />

Introduction<br />

Ce travail de thèse est consacré à l’étude d’<strong>arbres</strong> aléatoires. Dans l’introduction à ce travail,<br />

nous présenterons dans la partie 1.1 les <strong>arbres</strong> généalogiques aléatoires, c’est-à-dire les <strong>arbres</strong><br />

décrivant la généalogie d’une population aléatoire. Puis dans la partie 1.3, nous enrichirons la<br />

structure d’<strong>arbres</strong> généalogiques pour en faire des <strong>arbres</strong> spatiaux. Un arbre spatial sera alors<br />

un arbre généalogique pour lequel on a attribué à chacun de ses somm<strong>et</strong>s une position spatiale.<br />

Enfin dans la partie 1.5, nous présenterons les liens entre certains <strong>arbres</strong> spatiaux <strong>et</strong> les <strong>cartes</strong><br />

<strong>planaires</strong> biparties.<br />

Les parties 1.2, 1.4 <strong>et</strong> 1.6 présentent les contributions originales de ce travail de thèse.<br />

1.1. Arbres généalogiques aléatoires<br />

1.1.1. Arbres de Galton-Watson. Un arbre de Galton-Watson est un arbre discr<strong>et</strong><br />

aléatoire qui décrit la généalogie d’une population gouvernée par un processus de Galton-Watson<br />

(ou processus de branchement discr<strong>et</strong>). Ce modèle d’<strong>arbres</strong> généalogiques a été introduit par Neveu<br />

[48].<br />

Commençons par présenter le formalisme des <strong>arbres</strong> discr<strong>et</strong>s. On note U l’ensemble des mots<br />

d’entiers défini par<br />

U = ⋃ n≥0<br />

N n ,<br />

où par convention N = {1,2,...} <strong>et</strong> N 0 = {∅}. Un élément de U est une suite u = u 1 ...u n<br />

<strong>et</strong> l’on pose |u| = n de sorte que |u| représente la génération de u. En particulier, |∅| = 0. De<br />

plus, si u = u 1 ...u n ∈ U \ {∅} = U ∗ alors on note ǔ le père de u c’est-à-dire ǔ = u 1 ... u n−1 .<br />

Enfin si u = u 1 ...u n ∈ U <strong>et</strong> v = v 1 ...v m ∈ U, on définit la concaténation de u <strong>et</strong> v par<br />

uv = u 1 ...u n v 1 ...v m . En particulier u∅ = ∅u = u.<br />

Un arbre planaire est un sous-ensemble fini A de U vérifiant les propriétés suivantes :<br />

• ∅ ∈ A,<br />

• si u ∈ A \ {∅} alors ǔ ∈ A,<br />

• pour tout u ∈ A, il existe un nombre k u (A) ≥ 0 tel que uj ∈ A si <strong>et</strong> seulement si<br />

1 ≤ j ≤ k u (A).<br />

On note A l’ensemble des <strong>arbres</strong> <strong>planaires</strong>. Si A ∈ A, on note H(A) la hauteur de A c’est-à-dire<br />

H(A) = max{|u| : u ∈ A}.<br />

Tout arbre planaire est codé par un processus appelé processus de contour. Pour définir le<br />

processus de contour d’un arbre planaire A, imaginons une particule se déplaçant tout autour<br />

de A dans le sens des aiguilles d’une montre à vitesse 1. Chaque arête est visitée deux fois par la<br />

particule si bien qu’il lui faut un temps 2(#A − 1) pour parcourir entièrement A. Pour chaque<br />

7


t ∈ [0,2(#A − 1)] entier, on définit C(t) comme la hauteur du somm<strong>et</strong> visité par la particule au<br />

temps t, puis on interpole linéairement C sur l’intervalle [0,2(#A − 1)].<br />

Il existe une deuxième façon de coder un arbre planaire. Si A est un arbre planaire, énumérons<br />

les somm<strong>et</strong>s de A dans l’ordre lexicographique<br />

u(0) = ∅ ≺ u(1) ≺ ... ≺ u(#A − 1).<br />

Pour chaque n ∈ {0,1,... ,#A − 1}, on définit H n comme la hauteur du somm<strong>et</strong> u(n). Le<br />

processus H = (H n , 0 ≤ n ≤ #A − 1) est appelé processus des hauteurs de A.<br />

Fig. 1. Un arbre, son processus de contour <strong>et</strong> son processus des hauteurs.<br />

Nous sommes maintenant en mesure de définir les <strong>arbres</strong> de Galton-Watson. Si A est un arbre<br />

planaire <strong>et</strong> u ∈ A, on note τ u A = {v ∈ U : uv ∈ A} le sous-arbre de A issu de u. Soit µ une loi<br />

de reproduction critique ou sous-critique, ce qui signifie que ∑ k≥1 kµ(k) ≤ 1. Alors il existe une<br />

unique mesure de probabilité Π µ sur A vérifiant les deux propriétés suivantes :<br />

• Π µ (k ∅ (A) = k) = µ(k) pour tout k ≥ 0,<br />

• si µ(k) > 0, sous la mesure Π µ (· | k ∅ (A) = k), les sous <strong>arbres</strong> τ 1 A,...,τ k A sont<br />

indépendants <strong>et</strong> de loi Π µ .<br />

La mesure Π µ est par définition la loi d’un arbre de Galton-Watson de loi de reproduction µ. La<br />

deuxième propriété est une propriété de régénération. Les <strong>arbres</strong> de Galton-Watson sont ainsi<br />

les seuls <strong>arbres</strong> discr<strong>et</strong>s régénératifs. Pour p ≥ 1 <strong>et</strong> A ∈ A, on pose<br />

A p = A ∩ {u ∈ U : |u| ≤ p}.<br />

Soit p ≥ 1, soit T ∈ A tel que H(T) = p <strong>et</strong> tel que v 1 ≺ ... ≺ v m est la liste des somm<strong>et</strong>s de T à<br />

la génération p énumérés dans l’ordre lexicographique. On a alors plus généralement la propriété<br />

de régénération suivante :<br />

• si Π µ (A p = T) > 0, sous la mesure Π µ (· | A p = T), les sous <strong>arbres</strong> τ v1 A,... ,τ vm A sont<br />

indépendants <strong>et</strong> de loi Π µ .<br />

Terminons c<strong>et</strong>te sous-partie en présentant un lien entre les <strong>arbres</strong> de Galton-Watson <strong>et</strong> les<br />

marches aléatoires (voir par exemple [41]). Si µ est une loi de reproduction, on définit une mesure<br />

ν sur {−1,0,1,...} par la formule<br />

ν(k) = µ(k + 1), k ≥ −1.<br />

8


On peut alors montrer qu’il existe une marche aléatoire (X n ) n≥0 de loi de sauts ν telle que sous<br />

Π µ , presque sûrement, pour tout n ∈ {1,... ,#A − 1},<br />

{<br />

}<br />

H n = # j ∈ {1,... ,n} : X j = inf X k .<br />

j≤k≤n<br />

1.1.2. Arbres continus aléatoires. En vue d’étudier la structure limite de certains <strong>arbres</strong><br />

discr<strong>et</strong>s aléatoires, Aldous a introduit la notion d’arbre continu aléatoire.<br />

L’ensemble des <strong>arbres</strong> continus que nous considérerons est l’ensemble des <strong>arbres</strong> réels compacts<br />

enracinés. Un arbre réel est un espace métrique (T ,d) tel que pour chaque couple de points<br />

(σ 1 ,σ 2 ) de T , il existe un unique arc noté [[σ 1 ,σ 2 ]] reliant σ 1 à σ 2 , c<strong>et</strong> arc étant isométrique à un<br />

segment. Dans la suite, nous ne considérerons que des <strong>arbres</strong> réels compacts <strong>et</strong> enracinés, c’està-dire<br />

ayant un point ρ distingué, que nous appellerons la racine. Deux <strong>arbres</strong> réels compacts <strong>et</strong><br />

enracinés T <strong>et</strong> T ′ de racines respectives ρ <strong>et</strong> ρ ′ sont dits isométriques s’il existe une isométrie<br />

φ de T sur T ′ telle que φ(ρ) = ρ ′ . Nous noterons T l’ensemble des classes d’isométrie d’<strong>arbres</strong><br />

réels compacts enracinés.<br />

L’espace T est muni de la distance de Gromov-HausdorffGH définie de la façon suivante :<br />

si T <strong>et</strong> T ′ sont deux <strong>arbres</strong> réels compacts de racines respectives ρ <strong>et</strong> ρ ′ , alors<br />

GH(T , T ′ ) = inf { δ Haus (ϕ(T ),ϕ ′ (T ′ )) ∨ δ(ϕ(ρ),ϕ ′ (ρ ′ )) } ,<br />

où l’infimum est pris sur tous les plongements isométriques ϕ : T → E <strong>et</strong> ϕ ′ : T ′ → E dans un<br />

même espace métrique (E,δ). On rappelle que si (E,δ) est un espace métrique, on note δ Haus<br />

la distance de Hausdorff définie sur l’ensemble des sous-ensembles compacts de E.<br />

On peut donner une définition équivalente de la distance de Gromov-Hausdorff. Pour ce<br />

faire, il nous faut introduire la notion de correspondance entre deux espaces métriques. Une<br />

correspondance entre deux espaces métriques (E,δ) <strong>et</strong> (E ′ ,δ ′ ) est un sous-ensemble R de E ×E ′<br />

tel que pour tout x ∈ E (respectivement tout x ′ ∈ E ′ ), il existe x ′ ∈ E ′ (respectivement x ∈ E)<br />

tel que (x,x ′ ) ∈ R. La distorsion de la correspondance R est alors définie par<br />

dis(R) = sup { |δ(x,y) − δ ′ (x ′ ,y ′ )| : (x,x ′ ),(y,y ′ ) ∈ R } .<br />

Si T <strong>et</strong> T ′ sont deux <strong>arbres</strong> réels compacts de racines respectives ρ <strong>et</strong> ρ ′ , on note C(T , T ′ )<br />

l’ensemble des correspondances entre T <strong>et</strong> T ′ . On a alors<br />

GH(T , T ′ ) = 1 2 inf { dis(R) : R ∈ C(T , T ′ ),(ρ,ρ ′ ) ∈ R } .<br />

Evans, Pitman & Winter [24] ont montré que l’espace (T,GH) est un espace polonais.<br />

On peut construire un arbre réel par un codage similaire au codage d’un arbre planaire par<br />

son processus de contour. Plus précisément, si f : [0,+∞) → [0,+∞) est une fonction continue<br />

à support compact telle que f(0) = 0, on peut construire l’arbre réel compact codé par f de la<br />

façon suivante. Soit d f la pseudo-distance sur R + donnée par la relation<br />

d f (s,t) = f(s) + f(t) − 2 inf<br />

u∈[s,t] f(u),<br />

avec 0 ≤ s ≤ t. On peut alors définir une relation d’équivalence sur R + en posant s ∼ t si<br />

d f (s,t) = 0. L’arbre codé par f est alors l’espace quotient<br />

T f = [0,+∞)/ ∼<br />

muni de la distance d f <strong>et</strong> enraciné en la classe d’équivalence de 0. Pour tout s ∈ [0,+∞), on<br />

notera ṡ la classe d’équivalence de s. Ainsi ṡ est un somm<strong>et</strong> de T f à distance f(s) de la racine.<br />

9


Remarquons qu’un arbre planaire peut-être interprété comme une union de segments de<br />

longueur 1 de la façon suivante. Soit A ∈ A. On note (ǫ u ,u ∈ A \ {∅}) la base canonique de<br />

R A\{∅} . On définit une famille (l u ,u ∈ A \ {∅}) d’éléments de R A\{∅} par l u = 0 si |u| = 1, <strong>et</strong><br />

par récurrence l u = lǔ + ǫǔ si |u| ≥ 2. On pose alors<br />

T A =<br />

⋃<br />

[l u ,l u + ǫ u ].<br />

u∈A\{∅}<br />

On munit T A de la distance d A de sorte que pour tout couple de points (x,y) de T A , la distance<br />

d A (x,y) est la longueur du plus court chemin entre x <strong>et</strong> y, <strong>et</strong> l’on enracine T A en 0. Ainsi<br />

(T A ,d A ) est un arbre réel compact enraciné. Par ailleurs, notons C le processus de contour de<br />

A. On vérifie que l’arbre réel T A coïncide avec l’arbre réel T C à isométrie près. Remarquons<br />

que l’arbre planaire A ne peut être reconstruit à partir de l’arbre réel T A . En eff<strong>et</strong>, il n’y a pas<br />

d’ordre entre les enfants d’un même individu dans T A .<br />

Un premier exemple d’arbre continu aléatoire est le CRT d’Aldous [3], [4], [5]. Le CRT<br />

est défini comme l’arbre codé par √ 2e, où e = (e(t),0 ≤ t ≤ 1) est l’excursion brownienne<br />

normalisée. Si T est un arbre réel muni de la distance d, <strong>et</strong> si r > 0, alors on note rT l’arbre<br />

T muni de la distance rd. On peut alors énoncer un premier résultat de convergence dû à<br />

Aldous. Soit (A n ) n≥1 une suite d’<strong>arbres</strong> <strong>planaires</strong> aléatoires telle que chaque A n est uniformément<br />

distribué sur l’ensemble des <strong>arbres</strong> <strong>planaires</strong> à n somm<strong>et</strong>s. Alors<br />

n −1/2 T An<br />

(loi)<br />

−→<br />

n→∞ T√ 2e .<br />

Plus généralement, si µ est une loi de reproduction critique <strong>et</strong> de variance σ 2 finie, alors la loi<br />

de l’arbre réel n −1/2 T A sous Π µ (· | #A = n) converge au sens de la convergence étroite des<br />

mesures sur T vers la loi de l’arbre réel T 2e/σ quand n → ∞.<br />

1.1.3. Arbres de Lévy. Les <strong>arbres</strong> de Lévy, introduits par Duquesne & Le Gall [22],<br />

sont des analogues continus des <strong>arbres</strong> de Galton-Watson, au sens où un arbre de Lévy est un<br />

arbre continu aléatoire qui décrit la généalogie d’un processus de branchement à espace d’états<br />

continu.<br />

Soit Y = (Y t ,t ≥ 0) un processus de branchement à espace d’états continu s’éteignant<br />

presque sûrement, de mécanisme de branchement ψ, <strong>et</strong> soit X un processus de Lévy d’exposant<br />

de Laplace ψ. Duquesne & Le Gall ont montré que l’on peut définir un processus H = (H t ,t ≥ 0)<br />

par la convergence en probabilité suivante, pour tout t ≥ 0,<br />

1<br />

H t = lim<br />

ε→0 ε<br />

∫ t<br />

0½{X s≤inf s≤u≤t X u+ε} ds.<br />

Informellement, H t mesure la taille de l’ensemble {s ∈ [0,t] : X s = inf s≤u≤t X u }. Le processus H<br />

est appelé processus des hauteurs du processus Y par analogie avec le modèle discr<strong>et</strong>. Par ailleurs,<br />

notons I t = inf s∈[0,t] X s pour t ≥ 0, <strong>et</strong> N la mesure des excursions du processus (X t − I t ,t ≥ 0).<br />

Le processus H peut être défini sous N, <strong>et</strong> adm<strong>et</strong> sous N une modification continue. L’arbre<br />

de Lévy associé à H est alors l’arbre codé par H sous la mesure N <strong>et</strong> l’on note Θ ψ la “loi” de<br />

l’arbre T H sous N.<br />

Les <strong>arbres</strong> stables forment une classe importante d’<strong>arbres</strong> de Lévy. Ils sont associés aux<br />

mécanismes de branchement ψ(u) = u α pour 1 < α ≤ 2. Duquesne & Le Gall ont calculé la<br />

dimension de Hausdorff d’un arbre stable [22] <strong>et</strong> ont obtenu des résultats précis concernant la<br />

mesure de Hausdorff d’un arbre stable [23]. Par ailleurs, Miermont [46], [47] <strong>et</strong> Haas & Miermont<br />

[27] ont utilisé la structure des <strong>arbres</strong> stables pour étudier les processus de fragmentation stables.<br />

10


Enfin, les <strong>arbres</strong> de Lévy apparaissent comme les seules limites possibles pour des suites<br />

d’<strong>arbres</strong> de Galton-Watson convenablement changés d’échelle. Nous renvoyons le lecteur aux<br />

theorèmes 2.2.1 <strong>et</strong> 2.2.3 pour des énoncés précis.<br />

1.2. Arbres continus régénératifs<br />

L’obj<strong>et</strong> de c<strong>et</strong>te partie est de présenter le chapitre 2 de ce travail de thèse, qui est la version<br />

d’un article [53] soumis pour publication. Il s’agit de caractériser les <strong>arbres</strong> de Lévy par une<br />

propriété de régénération.<br />

Introduisons pour cela quelques notations. Si (T ,d) est un arbre réel compact enraciné en ρ,<br />

on note H(T ) sa hauteur c’est-à-dire<br />

H(T ) = max {d(ρ,σ) : σ ∈ T } .<br />

De plus, pour t,h > 0, on définit Z T (t,t + h) comme le nombre de sous-<strong>arbres</strong> de T issus du<br />

niveau t <strong>et</strong> de hauteur strictement supérieure à h. Duquesne & Le Gall [22] ont montré que la<br />

mesure Θ ψ vérifie la propriété de régénération suivante :<br />

(R) pour t,h > 0 <strong>et</strong> p ∈ N, sous la mesure Θ ψ (· | H(T ) > t) <strong>et</strong> conditionnellement à<br />

l’événement {Z T (t,t + h) = p}, les p sous-<strong>arbres</strong> de T issus du niveau t <strong>et</strong> de hauteur<br />

strictement supérieure à h sont indépendants <strong>et</strong> de loi Θ ψ (· | H(T ) > h).<br />

La propriété (R) est l’analogue pour des <strong>arbres</strong> continus de la propriété de régénération des<br />

<strong>arbres</strong> de Galton-Watson. Nous avons montré que c<strong>et</strong>te propriété caractérise les “lois” des <strong>arbres</strong><br />

de Lévy parmi les mesures infinies sur T.<br />

Théorème 1.2.1. Soit Θ une mesure infinie sur (T,GH) telle que Θ(H(T ) = 0) = 0 <strong>et</strong><br />

0 < Θ(H(T ) > t) < +∞ pour tout t > 0, <strong>et</strong> satisfaisant la propriété (R). Alors il existe un<br />

processus de branchement à espace d’états continu s’éteignant presque sûrement, de mécanisme<br />

de branchement ψ, tel que Θ = Θ ψ .<br />

L’idée principale de la preuve du théorème 1.2.1 est d’approcher les <strong>arbres</strong> régénératifs par<br />

des <strong>arbres</strong> de Galton-Watson. C<strong>et</strong>te idée repose sur un procédé de discrétisation des <strong>arbres</strong> réels<br />

que nous allons décrire brièvement.<br />

On se donne ε > 0, n ≥ 1 <strong>et</strong> un arbre réel T ∈ T de hauteur H(T ) vérifiant (n + 1)ε <<br />

H(T ) ≤ (n + 2)ε. Pour chaque k ∈ {0,1,... ,n}, on note σ1 k,... ,σk m k<br />

les somm<strong>et</strong>s de T à la<br />

génération kε (c’est-à-dire à distance kε de la racine) dont sont issus les sous-<strong>arbres</strong> de T au<br />

dessus du niveau kε <strong>et</strong> de hauteur strictement supérieure à ε. Chaque somm<strong>et</strong> σi<br />

k+1 appartient à<br />

un sous-arbre issu d’un élément σj k i<br />

de l’ensemble {σl k : 1 ≤ l ≤ m k }. On dit alors que l’individu<br />

σi k+1 est issu de σj k i<br />

. Une difficulté provient du fait que l’ordre entre les individus σi k+1<br />

1<br />

,... ,σi k+1<br />

l j<br />

issus d’un même parent σj k n’est pas défini. En ordonnant uniformément ces individus, on peut<br />

construire un arbre planaire A ε (T ) rendant compte de la généalogie dans l’arbre réel T de la<br />

collection de points {σi k : 1 ≤ i ≤ m k, 1 ≤ k ≤ n}.<br />

La propriété de régénération (R) nous assure alors que si T est distribué selon la mesure<br />

de probabilité Θ(· | (n + 1)ε < H(T ) ≤ (n + 2)ε), l’arbre planaire A ε (T ) est un arbre de<br />

Galton-Watson conditionné à avoir une hauteur égale à n.<br />

Dans ce travail, nous nous sommes également intéressés aux mesures de probabilité sur T<br />

vérifiant la propriété de régénération (R).<br />

11


Théorème 1.2.2. Soit Θ une mesure de probabilité sur (T,GH) telle que Θ(H(T ) = 0) = 0<br />

<strong>et</strong> 0 < Θ(H(T ) > t) pour tout t > 0, <strong>et</strong> satisfaisant la propriété (R). Alors il existe a > 0 <strong>et</strong> µ,<br />

une mesure de probabilité critique ou sous-critique sur Z + \ {1} tels que Θ est la loi de l’arbre<br />

généalogique d’un processus de branchement en temps continu, à espace d’états discr<strong>et</strong>, de loi de<br />

reproduction µ <strong>et</strong> de taux de branchement a.<br />

1.3. Arbres spatiaux aléatoires<br />

1.3.1. Arbres de Galton-Watson spatiaux. Un arbre spatial discr<strong>et</strong> est un couple<br />

(A,U) où A ∈ A <strong>et</strong> U = (U v ,v ∈ A) est une application de A dans R. On note Ω l’ensemble des<br />

<strong>arbres</strong> spatiaux discr<strong>et</strong>s.<br />

Rappelons que l’on peut coder un arbre planaire par son processus de contour C. De même,<br />

on peut définir le processus de contour spatial d’un arbre spatial discr<strong>et</strong> (A,U) en posant pour<br />

tout t ∈ [0,2(#A − 1)] entier, V (t) = U v où v est le somm<strong>et</strong> visité par C au temps t, puis on<br />

interpole linéairement V sur l’intervalle [0,2(#A − 1)]. Le couple de processus (C,V ) code alors<br />

l’arbre spatial (A,U).<br />

Soit µ une loi de reproduction <strong>et</strong> soit γ une mesure de probabilité sur R. Construisons à<br />

présent sur l’ensemble Ω la loi d’un arbre de Galton-Watson spatial de loi de reproduction µ <strong>et</strong><br />

dont les déplacements spatiaux sont régis par γ. Soit A ∈ A <strong>et</strong> soit x ∈ R. On définit la mesure<br />

R x (A,dU) sur R A de la façon suivante. Considérons une suite (Y u ,u ∈ U) de variables aléatoires<br />

indépendantes <strong>et</strong> de loi γ. On pose U ∅ = x <strong>et</strong> pour tout v ∈ A \ {∅},<br />

U v = ∑<br />

Y v ′,<br />

v ′ ∈]∅,v]<br />

où ]∅,v] est l’ensemble des ancêtres de v privé de la racine ∅ (noter que v ∈]∅,v]). La mesure<br />

R x (A,dU) est alors la loi de (U v ,v ∈ A). Posons<br />

P x (dAdU) = Π µ (dA)R x (A,dU).<br />

On dit que la mesure P x est la loi d’un arbre de Galton-Watson spatial issu de x, de loi de<br />

reproduction µ <strong>et</strong> de loi de déplacement spatial γ.<br />

1.3.2. Arbre brownien <strong>et</strong> serpent brownien. L’arbre brownien est un arbre réel spatial<br />

aléatoire. Un arbre réel spatial est un couple (T ,Y ) où T est un arbre réel compact <strong>et</strong> enraciné<br />

<strong>et</strong> Y = (Y σ ,σ ∈ T ) est une application continue de T dans R. Deux <strong>arbres</strong> réels spatiaux (T ,Y )<br />

<strong>et</strong> (T ′ ,Y ′ ) sont dits isométriques si T <strong>et</strong> T ′ sont isométriques <strong>et</strong> si pour tout σ ∈ T ,<br />

Y ′ φ(σ) = Y σ,<br />

où φ est une isométrie de T sur T ′ telle que φ(ρ) = ρ ′ . On notera T sp l’ensemble des classes<br />

d’isométrie d’<strong>arbres</strong> réels spatiaux.<br />

Rappelons que C(T , T ′ ) est l’ensemble des correspondances entre T <strong>et</strong> T ′ <strong>et</strong> que si R ∈<br />

C(T , T ′ ), alors on note dis(R) la distorsion de R. On définit sur T sp une distancesp par la<br />

formule suivante :<br />

(<br />

sp (T ,Y ),(T ′ ,Y ′ ) ) {<br />

}<br />

= 1 2 inf dis(R) + sup |Y σ − Y σ ′ ′| : R ∈ C(T , T ′ ),(ρ,ρ ′ ) ∈ R .<br />

(σ,σ ′ )∈R<br />

On vérifie alors que l’espace T sp muni de la distancesp est un espace polonais.<br />

12


Nous pouvons maintenant définir l’arbre brownien. Si T ∈ T <strong>et</strong> x ∈ R, on note Q x (T ,dY ) la<br />

loi du processus gaussien (Y σ ,σ ∈ T ) caractérisé par<br />

• E(Y σ ) = x,<br />

• Cov(Y σ ,Y σ ′) = d(ρ,σ ∧ σ ′ ),<br />

où σ ∧ σ ′ est le plus proche ancêtre commun à σ <strong>et</strong> σ ′ , c’est-à-dire le somm<strong>et</strong> de T vérifiant<br />

la relation [[ρ,σ ∧ σ ′ ]] = [[ρ,σ]] ∩ [[ρ,σ ′ ]] (sous une hypothèse faible sur T toujours réalisée dans<br />

la suite, (Y σ ,σ ∈ T ) a une modification continue). Notons n la loi de l’excursion brownienne<br />

normalisée. On définit une mesure de probabilité N x (dT dY ) sur T sp en posant<br />

∫<br />

∫ ∫<br />

N x (dT dY )F(T ,Y ) = n(de) Q x (T e ,dY )F(T e ,Y ).<br />

La mesure N x est la loi de l’arbre brownien issu de x 1 .<br />

Le serpent brownien, introduit par Le Gall, est un obj<strong>et</strong> intimement lié à l’arbre brownien.<br />

Nous renvoyons le lecteur à [37] pour une présentation détaillée du serpent brownien <strong>et</strong> de<br />

certaines de ses applications. Le serpent brownien est un processus de Markov à valeurs dans<br />

l’espace des trajectoires arrêtées<br />

W = ⋃<br />

C([0,t], R).<br />

t∈R +<br />

Pour tout w ∈ W, on pose ζ w = t si w ∈ C([0,t], R). Autrement dit, ζ w représente le temps<br />

de vie de la trajectoire w. De plus, on note ŵ = w(ζ w ) le point terminal de w. On définit une<br />

distance d W sur W de la façon suivante :<br />

d W (w,w ′ ) = sup<br />

t≥0<br />

∣<br />

∣w(t ∧ ζ w ) − w ′ (t ∧ ζ w ′) ∣ + |ζ w − ζ w ′|.<br />

On peut alors montrer que l’espace W muni de la distance d W est un espace polonais.<br />

Construisons à présent le serpent brownien “conditionnellement à son temps de vie”. Pour<br />

cela, on se donne une fonction continue f : [0,1] −→ [0,+∞) telle que f(0) = 0, <strong>et</strong> l’on pose<br />

pour tous 0 ≤ s ≤ s ′ ≤ 1,<br />

m(s,s ′ ) = inf<br />

u∈[s,s ′ ] f(u).<br />

Soit x ∈ R. Il existe alors un processus de Markov inhomogène (W s ,0 ≤ s ≤ 1) à valeurs dans W<br />

tel que W 0 = x presque sûrement, <strong>et</strong> dont le noyau de transition est caractérisé par la description<br />

suivante : pour tous 0 ≤ s ≤ s ′ ≤ 1,<br />

• W s ′(t) = W s (t) pour tout t ≤ m(s,s ′ ), presque sûrement,<br />

• (W s ′(m(s,s ′ ) + t) − W s (m(s,s ′ )),0 ≤ t ≤ f(s ′ ) − m(s,s ′ )) est indépendant de W s <strong>et</strong><br />

suit la loi d’un mouvement brownien partant de 0.<br />

On note θ f x la loi du processus (W s ,s ≥ 0) <strong>et</strong> l’on pose,<br />

N x (dζ dW) = n(dζ)θ ζ x(dW).<br />

Le serpent brownien est alors le processus canonique ((ζ s ,0 ≤ s ≤ 1),(W s ,0 ≤ s ≤ 1)) de<br />

l’espace C([0,1], R + ) × C([0,1], W) muni de la mesure de probabilité N x .<br />

Remarquons que l’on peut définir sous N x un processus Y = (Y σ ,σ ∈ T ζ ) en posant pour<br />

tout s ∈ [0,1] tel que σ = ṡ,<br />

Y σ = Ŵs.<br />

1 Les notations n <strong>et</strong> Nx introduites ici n’ont pas les mêmes significations que dans le chapitre 3.<br />

13


Rappelons que si s ∈ [0,1], alors ṡ est la classe d’équivalence de s pour la relation d’équivalence<br />

définie par ζ. En outre, on voit que la loi de (T ζ ,Y ) est la mesure N x .<br />

Une mesure aléatoire est naturellement associée à l’arbre brownien. Il s’agit de la mesure Z<br />

sur R définie par<br />

∫ 1<br />

∫ 1 )<br />

〈Z,ϕ〉 = ϕ(Yṡ)ds = ϕ<br />

(Ŵs ds.<br />

0<br />

C<strong>et</strong>te mesure aléatoire est appelée Integrated Super-Brownian Excursion (ISE). La mesure ISE<br />

en grande dimension intervient dans divers résultats asymptotiques de modèles de mécanique<br />

statistique (voir par exemple [18], [28], [29]).<br />

Enfin, l’arbre brownien apparaît comme limite de suites d’<strong>arbres</strong> spatiaux discr<strong>et</strong>s convenablement<br />

renormalisés. Plus précisément, Janson & Marckert [31] ont montré le résultat suivant.<br />

Soit µ une loi de reproduction critique telle qu’il existe η > 0 satisfaisant<br />

∑<br />

e ηk µ(k) < +∞.<br />

k≥0<br />

Soit γ une mesure de probabilité sur R, de moyenne nulle <strong>et</strong> vérifiant la condition de moments<br />

suivante : quand y → +∞,<br />

γ ({|u| > y}) = o ( y −4) .<br />

Les lois µ <strong>et</strong> γ ont donc des variances finies <strong>et</strong> l’on pose Var(µ) = ϑ 2 µ <strong>et</strong> Var(γ) = ϑ 2 γ. Rappelons<br />

que si A est un arbre planaire, on note C le processus de contour de A <strong>et</strong> V le processus de<br />

contour spatial de A. Alors, pour tout x ∈ R, la loi sous la mesure P x (· | #A = n) de<br />

⎛<br />

⎝(<br />

ϑµ<br />

2<br />

)<br />

C(2nt)<br />

√ n<br />

0≤t≤1<br />

,<br />

(<br />

(<br />

1 ϑµ<br />

ϑ γ 2<br />

0<br />

) )<br />

1/2<br />

V (2nt)<br />

n 1/4<br />

0≤t≤1<br />

converge quand n → ∞ vers la loi sous la mesure de probabilité N 0 du couple<br />

(<br />

)<br />

(ζ s ,0 ≤ s ≤ 1),(Ŵs,0 ≤ s ≤ 1) .<br />

1.4. Arbre brownien conditionné<br />

Nous présentons dans c<strong>et</strong>te partie le chapitre 3 de ce travail de thèse, qui est une version d’un<br />

article [42] écrit en collaboration avec Jean-François Le Gall <strong>et</strong> paru aux annales de l’institut<br />

Henri Poincaré. Il s’agit de définir l’arbre brownien issu de 0 conditionné à rester positif.<br />

Pour (T ,Y ) ∈ T sp , on note X T ,Y = {Y σ : σ ∈ T }. On cherche donc à conditionner l’arbre<br />

brownien issu de 0 par l’événement {X T ,Y ⊂ [0,+∞)}. Ce <strong>conditionnement</strong> est un <strong>conditionnement</strong><br />

dégénéré car<br />

N 0 (X T ,Y ⊂ [0,+∞)) = 0.<br />

En revanche, pour tout ε > 0, on a<br />

N 0 (X T ,Y ⊂ (−ε,+∞)) > 0.<br />

L’idée est donc de construire la loi de l’arbre brownien conditionné à rester positif comme la<br />

limite en un certain sens des mesures N 0 (dT dY | X T ,Y ⊂ (−ε,+∞)).<br />

Théorème 1.4.1. On a<br />

N 0 (X T ,Y ⊂ (−ε, ∞))<br />

lim<br />

ε↓0 ε 4 = 2<br />

21 .<br />

⎞<br />

⎠<br />

14


De plus, il existe une mesure N 0 sur l’espace T sp telle que<br />

lim<br />

ε↓0<br />

N 0 (dT dY | X T ,Y ⊂ (−ε, ∞)) = N 0 (dT dY ),<br />

au sens de la convergence étroite des mesures sur T sp .<br />

Le deuxième théorème principal de ce travail donne une représentation explicite de l’arbre réel<br />

spatial de loi N 0 au moyen d’une transformation de l’arbre brownien analogue à la transformation<br />

de Vervaat du pont brownien.<br />

Rappelons brièvement en quoi consiste la transformation de Vervaat du pont brownien. Soit<br />

(B t ,t ∈ [0,1]) un pont brownien sur [0,1]. Il est connu que presque sûrement, il existe un unique<br />

instant t ∗ ∈ [0,1] tel que<br />

B t∗ = min<br />

t∈[0,1] B t.<br />

On définit un processus B ∗ = (B ∗ t ,t ∈ [0,1]) par la transformation<br />

B ∗ t = B {t∗+t} − B t∗ ,<br />

où {t ∗ + t} est la partie fractionnaire de t ∗ + t. Vervaat [52] a montré que la loi du processus B ∗<br />

est alors la loi n de l’excursion brownienne normalisée.<br />

La transformation que nous effectuons sur les <strong>arbres</strong> réels spatiaux est une opération de<br />

“réenracinement au minimum”. Présentons tout d’abord en quoi consiste le réenracinement d’un<br />

arbre réel spatial en l’un de ses somm<strong>et</strong>s. Si (T ,Y ) ∈ T sp <strong>et</strong> σ ∈ T , on définit l’arbre réenraciné<br />

(T [σ] ,Y [σ] ) de la façon suivante. L’arbre T [σ] est l’arbre T , mais sa racine est le somm<strong>et</strong> σ <strong>et</strong><br />

non plus ρ. Puis on translate les positions spatiales en posant pour tout σ ′ ∈ T ,<br />

Y [σ]<br />

σ ′ = Y σ ′ − Y σ .<br />

Par ailleurs, on montre (voir la proposition 3.2.5) que sous la mesure N 0 , il existe presque<br />

sûrement un unique somm<strong>et</strong> σ ∗ ∈ T tel que<br />

Y σ∗ = min {Y σ : σ ∈ T }.<br />

On est alors en mesure d’énoncer le deuxième théorème principal de ce travail.<br />

Théorème 1.4.2. La mesure N 0 est la loi sous N 0 de l’arbre réenraciné (T [σ∗] ,Y [σ∗] ).<br />

Le point de départ de la preuve de ce théorème est une propriété d’invariance de l’opération<br />

de réenracinement sous N 0 . Marckert & Mokkadem [45] (voir aussi le théorème 3.2.3) ont montré<br />

que pour toute fonction mesurable positive sur T sp <strong>et</strong> pour tout s ∈ [0,1], on a<br />

(<br />

N 0<br />

(F T [ṡ] ,Y [ṡ])) = N 0 (F(T ,Y )) .<br />

Les théorèmes 1.4.1 <strong>et</strong> 1.4.2 peuvent être exprimés en termes du serpent brownien. Tout<br />

d’abord, notons que la proposition 3.2.5 nous assure que N 0 presque sûrement, il existe un<br />

unique s ∗ ∈ [0,1] tel que<br />

}<br />

Ŵ s∗ = min{Ŵs : s ∈ [0,1] = min {W s (t) : t ∈ [0,ζ s ], s ∈ [0,1]} .<br />

De plus, pour tout s ∈ [0,1], on pose<br />

ζ [s]<br />

r = ζ s + ζ {s+r} − 2 inf<br />

[s,{s+r}] ζ,<br />

Ŵ [s]<br />

r<br />

= Ŵ{s+r} − Ŵs.<br />

15


On définit ensuite W [s] à partir de Ŵ [s] de la façon suivante : pour tout r ∈ [0,1] <strong>et</strong> tout<br />

t ∈ [0,ζ r [s] ],<br />

[s]<br />

(t) = Ŵ n<br />

o.<br />

W [s]<br />

r<br />

sup<br />

u≤r:ζ u [s] =t<br />

Notons X = {Ŵs : s ∈ [0,1]}. On a alors l’existence d’une mesure N 0 sur C(R + , R + )×C(R + , W)<br />

telle que<br />

N 0 (dζ dW | X ⊂ (−ε,+∞)) −→<br />

ε↓0<br />

N 0 (dζ dW),<br />

au sens de la convergence étroite des mesures sur C(R + , R + ) × C(R + , W). De plus N 0 est la loi<br />

sous N 0 du serpent réenraciné (ζ [s∗] ,W [s∗] ).<br />

De même que précédemment, on définit sous N 0 un processus Y = (Y σ ,σ ∈ T ζ ) en posant<br />

pour tout s ∈ [0,1] tel que σ = ṡ,<br />

Y σ = Ŵs.<br />

Alors la loi de (T ζ ,Y ) sous N 0 est la mesure N 0 .<br />

Une des motivations de ce travail était d’exhiber un arbre aléatoire qui serait une limite<br />

possible de suites d’<strong>arbres</strong> discr<strong>et</strong>s conditionnés à rester positifs, <strong>et</strong> convenablement renormalisés.<br />

Le Gall [39] a montré le résultat suivant. Notons, pour (A,U) ∈ Ω,<br />

U = min{U v : v ∈ A \ {∅}}.<br />

Supposons que les lois µ <strong>et</strong> γ vérifient les mêmes hypothèses que celles du résultat de Janson<br />

& Marckert <strong>et</strong> que la loi γ est symétrique. Alors pour tout x > 0, la loi sous la mesure de<br />

probabilité P x (· | #A = n + 1,U > 0) de<br />

⎛<br />

⎝(<br />

ϑµ<br />

2<br />

)<br />

C(2nt)<br />

√ n<br />

0≤t≤1<br />

,<br />

(<br />

(<br />

1 ϑµ<br />

ϑ γ 2<br />

) )<br />

1/2<br />

V (2nt)<br />

n 1/4<br />

0≤t≤1<br />

converge quand n → ∞ vers la loi sous la mesure de probabilité N 0 du couple<br />

(<br />

)<br />

(ζ s ,0 ≤ s ≤ 1),(Ŵs,0 ≤ s ≤ 1) .<br />

Dans ce travail nous nous sommes également intéressés au <strong>conditionnement</strong> de l’arbre brownien<br />

ayant une hauteur fixée. Pour h > 0, on note n h la loi de l’excursion brownienne de hauteur<br />

h. On définit une mesure de probabilité N h x (dT dY ) en posant pour x ∈ R,<br />

∫<br />

∫ ∫<br />

N h x(dT dY )F(T ,Y ) = n h (de) Q x (T e ,dY )F(T e ,Y ).<br />

Rappelons que si f : [0,+∞) −→ [0,+∞) est une fonction continue à support compact telle que<br />

f(0) = 0, on dit que f est de hauteur h si<br />

max f(s) = h.<br />

s∈[0,+∞)<br />

Remarquons que si f est de hauteur h, alors H(T f ) = h. On dit que N h x est la loi de l’arbre<br />

brownien de hauteur h issu de x.<br />

Théorème 1.4.3. Pour tout h > 0, il existe une mesure de probabilité N h 0 sur T sp telle que<br />

lim<br />

ε→0 Nh 0(dT dY | X T ,Y ⊂ (−ε,+∞)) = N h 0(dT dY ),<br />

au sens de la convergence étroite des mesures sur T sp .<br />

⎞<br />

⎠<br />

16


Décrivons l’arbre réel spatial de loi N h 0 . Tout d’abord, on peut montrer que Nh 0 presque<br />

sûrement, il existe un unique somm<strong>et</strong> σ ∈ T tel que<br />

d(ρ,σ) = h.<br />

Pour tout t ∈ [0,h], soit σ t l’unique point de l’arc [[ρ,σ]] tel que d(ρ,σ t ) = t. On définit alors un<br />

processus (R h t ,0 ≤ t ≤ h) par la relation,<br />

R h t = Y σ t<br />

, t ∈ [0,h].<br />

On montre que la loi du processus (Rt h ,0 ≤ t ≤ h) est absolument continue par rapport à la loi<br />

de (β t ,0 ≤ t ≤ h) où β est un processus de Bessel de dimension 9. Nous renvoyons à la partie<br />

3.4 pour l’expression explicite de la densité de la loi de (Rt h ,0 ≤ t ≤ h) par rapport à la loi du<br />

processus de Bessel (β t ,0 ≤ t ≤ h). De plus, on montre que le processus (Rt∧h h ,t ≥ 0) converge<br />

en loi vers (β t ,t ≥ 0) quand h → ∞.<br />

Par ailleurs, notons (T i ,i ∈ I) la famille des sous-<strong>arbres</strong> de T issus de l’arc [[ρ,σ]]. Soit σ i le<br />

point de [[ρ,σ]] dont est issu le sous-arbre T i . On pose<br />

On définit alors une mesure ponctuelle M par<br />

d i = d(ρ,σ i ).<br />

M = ∑ i∈I<br />

δ (di ,T i ).<br />

Puis, si l > 0, on note n


1.5. Arbres spatiaux <strong>et</strong> <strong>cartes</strong> <strong>planaires</strong><br />

1.5.1. Lois de Boltzmann sur les <strong>cartes</strong> <strong>planaires</strong>. Une carte planaire M est un plongement<br />

d’un graphe connexe G dans la sphère bidimensionnelle S 2 , c’est à dire un “dessin” de<br />

G dans S 2 tel que les arêtes ne se rencontrent qu’au niveau des somm<strong>et</strong>s. Une face de M est<br />

une composante connexe de S 2 \ M, <strong>et</strong> son degré est le nombre d’arêtes de M incluses dans sa<br />

ferm<strong>et</strong>ure. Une carte planaire est dite bipartie si chacune de ses faces est de degré pair. Une<br />

carte planaire est une 2κ-angulation si chacune de ses faces est de degré 2κ.<br />

Si M est une carte planaire, on notera respectivement V M , E M <strong>et</strong> F M l’ensemble des somm<strong>et</strong>s,<br />

des arêtes <strong>et</strong> des faces de M. L’ensemble V M est muni de la distance de graphe, c’est-à-dire que la<br />

distance entre deux somm<strong>et</strong>s de M est la longueur du plus court chemin entre ces deux somm<strong>et</strong>s.<br />

Une carte planaire pointée est un couple (M,τ) où M est une carte planaire <strong>et</strong> τ est un<br />

somm<strong>et</strong> distingué de M. Une carte planaire enracinée est un couple (M,⃗e ) où M est une carte<br />

planaire <strong>et</strong> ⃗e est une arête orientée de M. On appelle somm<strong>et</strong> racine le point dont est issue l’arête<br />

orientée ⃗e. Une carte planaire enracinée <strong>et</strong> pointée est un tripl<strong>et</strong> (M,τ,e) où (M,τ) est une carte<br />

planaire pointée <strong>et</strong> e est une arête non orientée de M. Remarquons que l’on peut interpréter une<br />

carte enracinée (M,⃗e ) comme une carte enracinée <strong>et</strong> pointée en conservant l’arête non-orientée<br />

e <strong>et</strong> en choisissant le somm<strong>et</strong> racine comme point distingué.<br />

On identifie deux <strong>cartes</strong> <strong>planaires</strong> pointées (respectivement enracinées, enracincées <strong>et</strong> pointées)<br />

M <strong>et</strong> M ′ s’il existe un homéomorphisme de S 2 préservant l’orientation, qui envoie M sur M ′ <strong>et</strong><br />

qui préserve le point distingué (respectivement l’arête orientée, le point distingué <strong>et</strong> l’arête nonorientée).<br />

On note respectivement M p , M r <strong>et</strong> M r,p , l’ensemble des <strong>cartes</strong> <strong>planaires</strong> pointées,<br />

enracinées <strong>et</strong> enracinées <strong>et</strong> pointées après l’identification précédente.<br />

On définit à partir d’une suite q = (q i ) i≥1 de poids positifs telle que q i > 0 pour au moins un<br />

indice i ≥ 2, une mesure de Boltzmann sur l’ensemble M r,p de la façon suivante. Pour M ∈ M r,p ,<br />

on pose<br />

W q (M) = ∏<br />

q deg(f)/2 ,<br />

f∈F M<br />

où deg(f) est le degré de la face f. Si la suite q vérifie<br />

Z q =<br />

∑<br />

W q (M) < ∞,<br />

M∈M r,p<br />

on dit que q est admissible <strong>et</strong> l’on définit la distribution de Boltzmann B r,p<br />

q sur l’ensemble M r,p<br />

par<br />

B r,p<br />

q (M) = W q(M)<br />

.<br />

Z q<br />

De plus, on définit pour M ∈ M r<br />

Z q (r) = ∑ ∏<br />

q deg(f)/2 .<br />

f∈F M<br />

La condition Z q < ∞ implique que Z q<br />

(r)<br />

sur l’ensemble M r par<br />

M∈M r<br />

< ∞. Alors on définit la distribution de Boltzmann B r q<br />

B r q(M) = W q(M)<br />

Z q<br />

(r) .<br />

18


On pose N(k) = ( 2k−1<br />

k−1<br />

)<br />

pour k ≥ 1. Pour tout suite q de poids comme précédemment on<br />

définit<br />

f q (x) = ∑ k≥0<br />

N(k + 1)q k+1 x k , x ≥ 0,<br />

<strong>et</strong> l’on note R q le rayon de convergence de la série entière f q . D’après la proposition 1 dans [43],<br />

la suite q est admissible si <strong>et</strong> seulement si l’équation<br />

(1.5.1) f q (x) = 1 − 1 x , x > 0,<br />

adm<strong>et</strong> au moins une solution. Dans ce cas Z q est la solution de (1.5.1) vérifiant<br />

On dit que la suite q est régulière si<br />

Z 2 q f ′ q (Z q) ≤ 1.<br />

Z 2 q f ′ q (Z q) = 1.<br />

Si de plus Z q < R q on dit que q est critique régulière.<br />

Présentons finalement un cas particulier de mesure de Boltzmann ne chargeant que l’ensemble<br />

des 2κ-angulations. Soit κ ≥ 2. On pose<br />

α κ =<br />

(κ − 1)κ−1<br />

κ κ N(κ) .<br />

Soit q κ la suite de poids définie par q κ = α κ <strong>et</strong> q i = 0 pour tout i ∈ N \ {κ}. D’après la partie<br />

1.5 dans [43], la suite q κ est critique régulière <strong>et</strong><br />

Z qκ =<br />

κ<br />

κ − 1 .<br />

Pour tout n ≥ 1, notons respectivement U n κ <strong>et</strong> U n κ la mesure uniforme sur l’ensemble des 2κangulations<br />

enracinées <strong>et</strong> pointées <strong>et</strong> la mesure uniforme sur l’ensemble des 2κ-angulations enracinées.<br />

On voit alors que<br />

B r,p<br />

q κ<br />

(· | #F M = n) = U n κ,<br />

B r q κ<br />

(· | #F M = n) = U n κ .<br />

1.5.2. Loi de Boltzmann sur l’ensemble des mobiles. Dans c<strong>et</strong>te partie, on interprétera<br />

les <strong>arbres</strong> spatiaux comme des <strong>arbres</strong> spatiaux à deux types en déclarant que les somm<strong>et</strong>s<br />

des générations paires sont de type 0 <strong>et</strong> les somm<strong>et</strong>s des générations impaires sont de type 1.<br />

Pour (A,U) ∈ A, on pose<br />

A 0<br />

A 1<br />

= {u ∈ A : |u| est pair},<br />

= {u ∈ A : |u| est impair}.<br />

Un mobile (enraciné) est un arbre spatial à deux types (A,U) tel que les propriétés suivantes<br />

sont vérifiées :<br />

(a) U v = Uˇv pour tout v ∈ A 1 .<br />

(b) Soit v ∈ A 1 tel que k v (A) = k ≥ 1. On note v (0) = ˇv le père de v <strong>et</strong> v (j) = vj pour tout<br />

j ∈ {1,... ,k}. Alors pour tout j ∈ {0,... ,k},<br />

où par convention v (k+1) = v (0) .<br />

U v(j+1) ≥ U v(j) − 1,<br />

19


Si de plus U v ≥ 1 pour tout v ∈ A, alors (A,U) est un mobile bien étiqu<strong>et</strong>é.<br />

Construisons à présent une loi de Galton-Watson sur l’ensemble des mobiles de la façon<br />

expliquée dans [43]. Soit q une suite de poids critique régulière. Notons µ q 0 la loi géométrique<br />

de paramètre f q (Z q ) c’est-à-dire la mesure sur {0,1,2,...} définie pour k ≥ 0 par<br />

<strong>et</strong> µ q 1<br />

µ q 0 (k) = f q(Z q ) k<br />

Z q<br />

,<br />

la mesure sur {0,1,2,...} définie pour k ≥ 0 par<br />

µ q 1 (k) = Zk qN(k + 1)q k+1<br />

.<br />

f q (Z q )<br />

On pose µ q = (µ q 0 ,µq 1 ). On définit une mesure de probabilité P µ q sur l’ensemble des <strong>arbres</strong><br />

<strong>planaires</strong> par la formule suivante : pour t ∈ A,<br />

P µ q(t) = ∏ u∈t 0 µ q 0 (k u(t)) ∏ u∈t 1 µ q 1 (k u(t)).<br />

Pour tout k ≥ 1 on note ν0 k la mesure de Dirac en 0 ∈ Rk <strong>et</strong> ν1 k la mesure uniforme sur<br />

l’ensemble A k défini par<br />

{<br />

}<br />

A k = (x 1 ,...,x k ) ∈ Z k : x 1 ≥ −1,x 2 − x 1 ≥ −1,... ,x k − x k−1 ≥ −1, −x k ≥ −1 .<br />

La mesure ν1 k est alors la loi du vecteur (X 1,X 1 + X 2 ,...,X 1 + ... + X k ) où (X 1 ,...,X k+1 ) est<br />

uniformément distribué sur l’ensemble B k defini par<br />

{<br />

}<br />

B k = (x 1 ,...,x k+1 ) ∈ {−1,0,1,2,...} k+1 : x 1 + ... + x k+1 = 0 .<br />

Soit A ∈ A. On définit une mesure de probabilité R ν,1 (A,dU) sur R A de la façon suivante.<br />

Soit (Y u ,u ∈ A) une suite de vecteurs aléatoires indépendants telle que pour u ∈ A vérifiant<br />

k u (A) = k, le vecteur Y u = (Y u1 ,... ,Y uk ) est de loi ν0 k si u ∈ A0 <strong>et</strong> de loi ν1 k si u ∈ A1 . On<br />

définit U ∅ = 1 <strong>et</strong> pour tout v ∈ A \ {∅},<br />

U v = 1 + ∑<br />

Y u ,<br />

u∈]∅,v]<br />

où ]∅,v] représente l’ensemble des ancêtres de v privé de la racine ∅. La mesure R ν,1 (A,dU)<br />

est alors la loi de (U v ,v ∈ A).<br />

Soit P µ,ν,1 la mesure de probabilité définie par<br />

P µ,ν,1 (dAdU) = P µ (dA)R ν,1 (A,dU).<br />

On vérifie facilement que la mesure P µ,ν,1 ne charge que l’ensemble des mobiles (A,U) tels que<br />

U ∅ = 1.<br />

1.5.3. La bijection de Bouttier, di Francesco & Guitter. Bouttier, di Francesco &<br />

Guitter [11] ont établi une bijection entre l’ensemble des mobiles bien étiqu<strong>et</strong>és <strong>et</strong> l’ensemble<br />

des <strong>cartes</strong> biparties enracinées. C<strong>et</strong>te bijection est une généralisation de la bijection construite<br />

par Schaeffer [51] entre les <strong>arbres</strong> bien étiqu<strong>et</strong>és <strong>et</strong> les quadrangulations enracinées (voir aussi<br />

[15] pour une étude antérieure de c<strong>et</strong>te bijection).<br />

Présentons dans un premier temps une version de la bijection de Bouttier, di Francesco &<br />

Guitter entre l’ensemble des mobiles <strong>et</strong> l’ensemble des <strong>cartes</strong> biparties enracinées <strong>et</strong> pointées.<br />

Soit (A,U) un mobile tel que U ∅ = 1. On pose ξ = #A−1 <strong>et</strong> l’on définit une suite w 0 ,w 1 ,... ,w ξ<br />

20


de somm<strong>et</strong>s de A 0 de telle sorte que pour tout k ∈ {0,1,... ,ξ}, le somm<strong>et</strong> w k est le somm<strong>et</strong><br />

visité par le contour de A au temps 2k. De plus on définit un mobile bien étiqu<strong>et</strong>é (A,U + ) par<br />

la relation pour v ∈ A<br />

U + v = U v − min{U w : w ∈ A} + 1.<br />

Notons que<br />

min { U + w : w ∈ A} = 1.<br />

Supposons maintenant que l’arbre planaire A est dessiné dans le plan <strong>et</strong> ajoutons un somm<strong>et</strong><br />

∂. On associe au mobile (A,U) une carte planaire dont l’ensemble des somm<strong>et</strong>s est<br />

A 0 ∪ {∂},<br />

<strong>et</strong> dont les arêtes sont obtenues par le procédé suivant :<br />

• si U + w k<br />

= 1, on trace une arête entre w k <strong>et</strong> ∂,<br />

• si U + w k<br />

≥ 2, on trace une arête entre w k <strong>et</strong> le premier somm<strong>et</strong> parmi w k+1 ,...,w ξ−1 ,<br />

w 0 ,...,w k−1 dont l’étiqu<strong>et</strong>te est U + w k<br />

− 1.<br />

La condition (b) de la définition d’un mobile garantit que si U + w k<br />

≥ 2, alors il existe un somm<strong>et</strong><br />

parmi w k+1 ,...,w ξ−1 ,w 0 ,...,w k−1 dont l’étiqu<strong>et</strong>te est U + w k<br />

− 1. La carte planaire ainsi obtenue<br />

est une carte bipartie que l’on pointe en ∂ <strong>et</strong> que l’on enracine en l’arête obtenue pour k = 0<br />

dans la construction précédente.<br />

Bouttier, di Francesco & Guitter [11] ont montré que la construction précédente fournit une<br />

bijection Ψ r,p entre l’ensemble des mobiles (A,U) tels que U ∅ = 1 <strong>et</strong> l’ensemble des <strong>cartes</strong><br />

biparties enracinées <strong>et</strong> pointées.<br />

On remarque que si (A,U) est un mobile bien étiqu<strong>et</strong>é tel que U ∅ = 1 alors U + v = U v pour<br />

tout v ∈ A. En particulier U + ∅ = 1. On en déduit que l’arête distinguée de la carte Ψ r,p((A,U))<br />

contient le point ∂. On peut ainsi indentifier Ψ r,p ((A,U)) à une carte enracinée. Cela implique<br />

que la bijection Ψ r,p induit une bijection Ψ r entre l’ensemble des mobiles bien étiqu<strong>et</strong>és (A,U)<br />

tels que U ∅ = 1 <strong>et</strong> l’ensemble M r (voir la figure 2).<br />

La bijection Ψ r vérifie les deux propriétés suivantes : soit (A,U) un mobile tel que U ∅ = 1<br />

<strong>et</strong> soit M = Ψ((A,U)),<br />

• pour tout k ≥ 1, l’ensemble {f ∈ F M : deg(f) = 2k} est en bijection avec l’ensemble<br />

{v ∈ A 1 : k v (A) = k},<br />

• pour tout l ≥ 1, l’ensemble {a ∈ V M : d(∂,a) = l} est en bijection avec l’ensemble<br />

{v ∈ A 0 : U v = l}.<br />

Enfin signalons la propriété suivante (voir la proposition 10 dans [43]). La distribution de<br />

Boltzmann conditionnée B r,p<br />

q (· | #F M = n) est l’image de la mesure P µ q ,ν,1(· | #A 1 = n) par<br />

l’application Ψ r,p . De plus, pour tout mobile (A,U) on définit<br />

U = min { U v : v ∈ A 0 \ {∅} } .<br />

La distribution de Boltzmann conditionnée B r q (· | #F M = n) est alors l’image de la mesure<br />

P µ q ,ν,1(· | #A 1 = n, U ≥ 1) par l’application Ψ r .<br />

21


Fig. 2. Construction d’une carte bipartie enracinée à partir d’un mobile bien étiqu<strong>et</strong>é<br />

1.5.4. Un théorème limite pour les mobiles. En vue d’étudier certaines propriétés<br />

asymptotiques des grandes <strong>cartes</strong> biparties enracinées <strong>et</strong> pointées, Marckert & Miermont [43]<br />

ont établi un théorème limite pour les mobiles.<br />

Avant d’énoncer ce résultat, rappelons que si (A,U) est un arbre spatial, on note C son<br />

processus de contour <strong>et</strong> V son processus de contour spatial. De plus, pour toute suite de poids<br />

critique régulière q, on pose<br />

ρ q = 2 + Zqf 3 q(Z ′′ q ).<br />

Marckert & Miermont [43] ont montré le théorème suivant : si q est une suite critique régulière,<br />

alors la loi sous la mesure P µ q ,ν,1(· | #A 1 = n) de<br />

( (αq ( )<br />

C(2(#A − 1)t) βq V (2(#A − 1)t)<br />

n<br />

)0≤t≤1<br />

1/2 ,<br />

n<br />

)0≤t≤1<br />

1/4<br />

converge (au sens de la convergence étroite des mesures sur C([0,1], R 2 ) quand n → ∞ vers la<br />

loi de (ζ,Ŵ) sous la mesure N 0, où les constantes α q <strong>et</strong> β q sont définies par les formules<br />

√<br />

ρq (Z q − 1)<br />

α q =<br />

,<br />

4<br />

( ) 9(Zq − 1) 1/4<br />

β q =<br />

.<br />

4ρ q<br />

Ce résultat est un cas particulier du théorème 11 dans [43] qui traite d’<strong>arbres</strong> de Galton-Watson<br />

spatiaux à deux types plus généraux. Signalons que dans le cas particulier q = q κ les constantes<br />

apparaissant dans le résultat précédent sont données par<br />

α qκ = 1 ( ) κ 1/2<br />

,<br />

4 κ − 1<br />

β qκ =<br />

(<br />

9<br />

4κ(κ − 1)<br />

) 1/4<br />

.<br />

22


1.6. Résultats asymptotiques pour de grandes <strong>cartes</strong> biparties enracinées<br />

aléatoires<br />

C<strong>et</strong>te partie récapitule les résultats du chapitre 4 de ce travail de thèse, dans lequel nous nous<br />

intéressons à certaines propriétés asymptotiques de grandes <strong>cartes</strong> <strong>planaires</strong> biparties enracinées<br />

aléatoires.<br />

Introduisons au préalable quelques notations. Soit M ∈ M r . On note o le somm<strong>et</strong> racine de<br />

M c’est-à-dire le somm<strong>et</strong> dont est issue l’arête racine. Soit<br />

R M = max {d(o,a) : a ∈ V M } ,<br />

le rayon de la carte M. Le profil de M est la mesure de probabilité sur {0,1,2,...} définie par<br />

λ M (k) = #{a ∈ V M : d(o,a) = k}<br />

, k ≥ 0.<br />

#V M<br />

Si M a n faces, on définit enfin le profil de M changé d’échelle par la relation suivante,<br />

où A est un borélien de R + .<br />

λ (n)<br />

M (A) = λ M<br />

(<br />

n 1/4 A<br />

)<br />

,<br />

Théorème 1.6.1. Soit q une suite de poids critique régulière.<br />

(i) La loi de n −1/4 R M sous la mesure de probabilité B r q(· | #F M = n) converge quand<br />

n → ∞ vers la loi sous la mesure N 0 de la variable aléatoire<br />

(<br />

)<br />

1<br />

sup Ŵ(s) − inf Ŵ(s) .<br />

β q 0≤s≤1<br />

0≤s≤1<br />

(ii) La loi de la mesure aléatoire λ (n)<br />

M sous la mesure de probabilité Br q (· | #F M = n)<br />

converge quand n → ∞ vers la loi sous la mesure N 0 de la mesure aléatoire I définie<br />

par<br />

〈I,g〉 =<br />

∫ 1<br />

0<br />

( ( 1<br />

g<br />

β q<br />

Ŵ(t) −<br />

inf<br />

0≤s≤1<br />

))<br />

Ŵ(s) dt.<br />

(iii) La loi de n −1/4 d(o,a) où a est uniformément distribué sur V M , sous la mesure de<br />

probabilité B r q(· | #F M = n), converge quand n → ∞ vers la loi sous la mesure N 0 de<br />

la variable aléatoire ( )<br />

1<br />

sup Ŵ(s) .<br />

β q<br />

0≤s≤1<br />

Le théorème 1.6.1 est une généralisation des résultats obtenus par Chassaing & Schaeffer [14]<br />

qui traitent le cas particulier q = q 2 des quadrangulations (voir aussi le théorème 8.2 dans [39]).<br />

Le théorème 1.6.1 est aussi très proche du théorème 3 de [43] qui a largement motivé notre<br />

étude. La différence vient de ce que [43] considère des <strong>cartes</strong> enracinées <strong>et</strong> pointées, <strong>et</strong> étudie<br />

les distances à partir du somm<strong>et</strong> distingué <strong>et</strong> non du somm<strong>et</strong> racine comme nous le faisons ici.<br />

Signalons que dans le cas q = q 2 , la constante qui apparaît dans le théorème 1.6.1 vaut<br />

( )<br />

1 8 1/4<br />

= .<br />

β q2 9<br />

La preuve du théorème 1.6.1 repose sur un théorème limite pour les mobiles bien étiqu<strong>et</strong>és.<br />

23


Théorème 1.6.2. Soit q une suite de poids critique régulière. La loi sous la mesure de<br />

probabilité P µ q ,ν,1(· | #A 1 = n, U ≥ 1) de<br />

( (αq ( )<br />

C(2(#A − 1)t) βq V (2(#A − 1)t)<br />

n<br />

)0≤t≤1<br />

1/2 ,<br />

n<br />

)0≤t≤1<br />

1/4<br />

converge (au sens de la convergence étroite des mesures sur C([0,1], R) 2 ) quand n → ∞ vers la<br />

loi de (ζ,Ŵ) sous la mesure N 0.<br />

Le théorème 1.6.1 se déduit du théorème 1.6.2 en utilisant les propriétés vérifiées par la<br />

bijection de Bouttier, di Francesco & Guitter. En particulier on observe que la loi de R M sous<br />

la mesure B r q coïncide avec la loi de la variable aléatoire<br />

sup {U v : v ∈ A}<br />

sous la mesure P µ q ,ν,1(· | #A 1 = n,U ≥ 1). Or le théorème 1.6.2 implique que la loi sous la<br />

mesure P µ q ,ν,1(· | #A 1 = n, U ≥ 1) de<br />

1<br />

n<br />

supU 1/4 v<br />

v∈A<br />

converge quand n → ∞ vers la loi sous N 0 de la variable aléatoire<br />

1<br />

sup Ŵ s ,<br />

β q 0≤s≤1<br />

c’est-à-dire d’après le théorème 1.4.2, la loi sous N 0 de la variable aléatoire<br />

(<br />

)<br />

1<br />

sup Ŵ s − inf Ŵ s .<br />

β q 0≤s≤1<br />

0≤s≤1<br />

La preuve du théorème 1.6.2 s’articule de manière analogue à la preuve du théorème 2.2 dans<br />

[39]. En particulier, une étape cruciale est le calcul d’estimations de la probabilité<br />

P µ q ,ν,1<br />

(<br />

U ≥ 1 | #A 1 = n ) .<br />

Proposition 1.6.3. Il existe des constantes γ > 0 <strong>et</strong> ˜γ > 0 telles que pour tout n suffisamment<br />

grand,<br />

γ<br />

n ≤ P (<br />

µ q ,ν,1 U ≥ 1 | #A 1 = n ) ≤ ˜γ n .<br />

C<strong>et</strong>te proposition nous perm<strong>et</strong> d’obtenir des résultats asymptotiques concernant les points<br />

de disconnection dans les grandes 2κ-angulations enracinées uniformes.<br />

Soit M une carte planaire <strong>et</strong> soit σ 0 un somm<strong>et</strong> de M. Soit σ ∈ V M \ {σ 0 }. On note S σ 0,σ<br />

M<br />

l’ensemble des somm<strong>et</strong>s a de M tel que tout chemin allant de σ à a passe par σ 0 . On dit que σ 0<br />

est un point de disconnection de M s’il existe σ ∈ V M \ {σ 0 } tel que S σ 0,σ<br />

M ≠ {σ 0} <strong>et</strong> l’on note<br />

D M l’ensemble des points de disconnection de la carte M.<br />

Soit (A,U) un mobile <strong>et</strong> soit v ∈ A 0 . Rappelons que τ v A = {w ∈ U,vw ∈ A}. Supposons que<br />

(τ v A) 0 ≠ {v} <strong>et</strong> que la condition suivante est satisfaite :<br />

inf { U vw : w ∈ (τ v A) 0 \ {∅} } > U v .<br />

On remarque alors que dans la construction de la carte M = Ψ r,p ((A,U)) l’ensemble S v,τ<br />

M<br />

est en<br />

bijection avec (τ v A) 0 (où l’on a noté τ le point distingué de la carte enracinée <strong>et</strong> pointée M).<br />

Le somm<strong>et</strong> v est donc un point de disconnection de M.<br />

24


Rappelons que U n κ <strong>et</strong> U n κ désignent respectivement la mesure uniforme sur l’ensemble des<br />

2κ-angulations enracinées <strong>et</strong> pointées à n faces <strong>et</strong> la mesure uniforme sur l’ensemble des 2κangulations<br />

enracinées à n faces. On montre alors le résultat suivant en utilisant la remarque<br />

précédente <strong>et</strong> la proposition 1.6.3.<br />

Théorème 1.6.4. Pour tout ε > 0,<br />

lim<br />

n→∞ Un κ<br />

(<br />

∃ σ 0 ∈ D M : σ 0 ≠ τ, n 1/2−ε ≤ #S σ 0,τ<br />

M ≤ 2n1/2−ε) = 1.<br />

Remarquons que si M est une 2κ-angulation à n faces alors d’après la formule d’Euler<br />

#V M = n(κ − 1) + 2.<br />

On déduit alors du théorème 1.6.4 le résultat suivant.<br />

Théorème 1.6.5. Pour tout ε > 0,<br />

lim<br />

n→∞ U n κ<br />

(<br />

∃ σ 0 ∈ D M : ∃ σ ∈ V M \ {σ 0 }, n 1/2−ε ≤ #S σ 0,σ<br />

M ≤ 2n1/2−ε) = 1.<br />

25


CHAPITRE 2<br />

Regenerative real trees<br />

2.1. Introduction<br />

Galton-Watson trees are well known to be characterized among all random discr<strong>et</strong>e trees by<br />

a regenerative property. More precisely, if µ is a probability measure on Z + , the law Π µ of the<br />

Galton-Watson tree with offspring distribution µ is uniquely d<strong>et</strong>ermined by the following two<br />

conditions : under the probability measure Π µ ,<br />

(i) the ancestor has p children with probability µ(p),<br />

(ii) if µ(p) > 0, then conditionally on the event that the ancestor has p children, the p subtrees<br />

which describe the genealogy of the descendants of these children are independent<br />

and distributed according to Π µ .<br />

The aim of this work is to study σ-finite measures satisfying an analogue of this property on<br />

the space of equivalence classes of rooted compact R-trees.<br />

An R-tree is a m<strong>et</strong>ric space (T ,d) such that for any two points σ 1 and σ 2 in T there is a<br />

unique arc with endpoints σ 1 and σ 2 , and furthermore this arc is isom<strong>et</strong>ric to a compact interval<br />

of the real line. We denote this arc by [[σ 1 ,σ 2 ]]. In this work, all R-trees are supposed to be<br />

compact. A rooted R-tree is an R-tree with a distinguished vertex called the root. Say that two<br />

rooted R-trees are equivalent if there is a root-preserving isom<strong>et</strong>ry that maps one onto the other.<br />

It was noted in [24] that the s<strong>et</strong> T of all equivalence classes of rooted compact R-trees equipped<br />

with the pointed Gromov-Hausdorff distanceGH (see e.g. Chapter 7 in [12]) is a Polish space.<br />

Hence it is legitimate to consider random variables with values in T, that is, random R-trees.<br />

A particularly important example is the CRT, which was introduced by Aldous [3], [5] with a<br />

different formalism. Striking applications of the concept of random R-trees can be found in the<br />

recent papers [24] and [25].<br />

L<strong>et</strong> T be an R-tree. We write H(T ) for the height of the R-tree T , that is, the maximal<br />

distance from the root to a vertex of T . For every t ≥ 0, we denote by T ≤t the s<strong>et</strong> of all vertices<br />

of T which are at distance at most t from the root, and by T >t the s<strong>et</strong> of all vertices which are<br />

at distance greater than t from the root. To each connected component of T >t there corresponds<br />

a “subtree” of T above level t (see section 2.2.2.3 for a more precise definition). For every h > 0,<br />

we define Z(t,t+h)(T ) as the number of subtrees of T above level t with height greater than h.<br />

L<strong>et</strong> Θ be a σ-finite measure on T such that 0 < Θ(H(T ) > t) < ∞ for every t > 0 and<br />

Θ(H(T ) = 0) = 0. We say that Θ satisfies the regenerative property (R) if the following holds :<br />

(R) for every t,h > 0 and p ∈ N, under the probability measure Θ(· | H(T ) > t) and<br />

conditionally on the event {Z(t,t + h) = p}, the p subtrees of T above level t with<br />

height greater than h are independent and distributed according to the probability<br />

measure Θ(· | H(T ) > h).<br />

This is a natural analogue of the regenerative property stated above for Galton-Watson trees.<br />

Beware that, unlike the discr<strong>et</strong>e case, there is no natural order on the subtrees above a given<br />

27


level. So, the preceding property should be understood in the sense that the unordered collection<br />

of the p subtrees in consideration is distributed as the unordered collection of p independent<br />

copies of Θ(· | H(T ) > h).<br />

Property (R) is known to be satisfied by a wide class of infinite measures on T, namely the<br />

“laws” of Lévy trees. Lévy trees have been introduced by T. Duquesne and J.F. Le Gall in<br />

[22]. Their precise definition is recalled in section 2.2.3, but l<strong>et</strong> us immediately give an informal<br />

presentation.<br />

L<strong>et</strong> Y be a critical or subcritical continuous-state branching process. The distribution of Y<br />

is characterized by its branching mechanism function ψ. Assume that Y becomes extinct a.s.,<br />

which is equivalent to the condition ∫ ∞<br />

1<br />

ψ(u) −1 du < ∞. The ψ-Lévy tree is a random variable<br />

taking values in (T,GH), which describes the genealogy of a population evolving according to<br />

Y and starting with infinitesimally small mass. More precisely, the “law” of the Lévy tree is<br />

defined in [22] as a σ-finite measure on the space (T,GH), such that 0 < Θ ψ (H(T ) > t) < ∞<br />

for every t > 0. As a consequence of Theorem 4.2 of [22], the measure Θ ψ satisfies Property<br />

(R). In the special case ψ(u) = u α , 1 < α ≤ 2 corresponding to the so-called stable trees, this<br />

was used by Miermont [46], [47] to introduce and to study certain fragmentation processes.<br />

In the present work we describe all σ-finite measures on T that satisfy Property (R). We<br />

show that the only infinite measures satisfying Property (R) are the measures Θ ψ associated<br />

with Lévy trees. On the other hand, if Θ is a finite measure satisfying Property (R), we can<br />

obviously restrict our attention to the case Θ(T) = 1 and we obtain that Θ must be the law of<br />

the genealogical tree associated with a continuous-time discr<strong>et</strong>e-state branching process.<br />

Theorem 2.1.1. L<strong>et</strong> Θ be an infinite measure on (T,GH) such that Θ(H(T ) = 0) = 0 and<br />

0 < Θ(H(T ) > t) < +∞ for every t > 0. Assume that Θ satisfies Property (R). Then, there<br />

exists a continuous-state branching process, whose branching mechanism is denoted by ψ, which<br />

becomes extinct almost surely, such that Θ = Θ ψ .<br />

Theorem 2.1.2. L<strong>et</strong> Θ be a probability measure on (T,GH) such that Θ(H(T ) = 0) = 0<br />

and 0 < Θ(H(T ) > t) for every t > 0. Assume that Θ satisfies Property (R). Then there<br />

exists a > 0 and a critical or subcritical probability measure γ on Z + \{1} such that Θ is the<br />

law of the genealogical tree for a discr<strong>et</strong>e-space continuous-time branching process with offspring<br />

distribution γ, where branchings occur at rate a.<br />

In other words, Θ in Theorem 2.1.2 can be described in the following way : there exists a real<br />

random variable J such that under Θ :<br />

(i) J is distributed according to the exponential distribution with param<strong>et</strong>er a and there<br />

exists σ J ∈ T such that T ≤J = [[ρ,σ J ]],<br />

(ii) the number of subtrees above level J is distributed according to γ and is independent<br />

of J,<br />

(iii) for every p ≥ 2, conditionally on J and given the event that the number of subtrees<br />

above level J is equal to p, these p subtrees are independent and distributed according<br />

to Θ.<br />

Theorem 2.1.1 is proved in section 2.3, after some preliminary results have been established in<br />

section 2.2. A key idea of the proof is to use the regenerative property to embed discr<strong>et</strong>e Galton-<br />

Watson trees in our random real trees (Lemma 2.3.3). A technical difficulty comes from the<br />

fact that real trees are not ordered whereas Galton-Watson trees are usually defined as random<br />

ordered discr<strong>et</strong>e trees (cf subsection 2.2.2.4 below). To overcome this difficulty, we assign a<br />

28


andom ordering to the discr<strong>et</strong>e trees embedded in real trees. Another major ingredient of the<br />

proof of Theorem 2.1.1 is the construction of a “local time” L t at every level t of a random real<br />

tree governed by Θ. The local time process is then shown to be a continuous-state branching<br />

process with branching mechanism ψ, which makes it possible to identify Θ with Θ ψ . Theorem<br />

2.1.2 is proved in section 2.4. Several arguments are similar to the proof of Theorem 2.1.1, so<br />

that we have skipped some d<strong>et</strong>ails.<br />

2.2. Preliminaries<br />

In this section, we recall some basic facts about branching processes, R-trees and Lévy trees.<br />

2.2.1. Branching processes.<br />

2.2.1.1. Continuous-state branching processes. A (continuous-time) continuous-state branching<br />

process (in short a CSBP) is a Markov process Y = (Y t ,t ≥ 0) with values in the positive<br />

half-line [0,+∞), with a Feller semigroup (Q t ,t ≥ 0) satisfying the following branching property :<br />

for every t ≥ 0 and x,x ′ ≥ 0,<br />

Q t (x, ·) ∗ Q t (x ′ , ·) = Q t (x + x ′ , ·).<br />

Informally, this means that the union of two independent populations started respectively at x<br />

and x ′ will evolve like a single population started at x + x ′ .<br />

We will consider only the critical or subcritical case, meaning that for every t ≥ 0 and x ≥ 0,<br />

∫<br />

yQ t (x,dy) ≤ x.<br />

[0,+∞)<br />

Then, if we exclude the trivial case where Q t (x, ·) = δ 0 for every t > 0 and x ≥ 0, the Laplace<br />

functional of the semigroup can be written in the following form : for every λ ≥ 0,<br />

∫<br />

e −λy Q t (x,dy) = exp(−xu(t,λ)),<br />

[0,+∞)<br />

where the function (u(t,λ),t ≥ 0,λ ≥ 0) is d<strong>et</strong>ermined by the differential equation<br />

du(t,λ)<br />

= −ψ(u(t,λ)), u(0,λ) = λ,<br />

dt<br />

and ψ : R + −→ R + is of the form<br />

∫<br />

(2.2.1) ψ(u) = αu + βu 2 + (e −ur − 1 + ur)π(dr),<br />

(0,+∞)<br />

where α,β ≥ 0 and π is a σ-finite measure on (0,+∞) such that ∫ (0,+∞) (r ∧ r2 )π(dr) < ∞. The<br />

process Y is called the ψ-continuous-state branching process (in short the ψ-CSBP).<br />

Continuous-state branching processes may also be obtained as weak limits of rescaled Galton-<br />

Watson processes. We recall that an offspring distribution is a probability measure on Z + . An<br />

offspring distribution µ is said to be critical if ∑ i≥0 iµ(i) = 1 and subcritical if ∑ i≥0<br />

iµ(i) < 1.<br />

L<strong>et</strong> us state a result that can be derived from [26] and [30].<br />

Theorem 2.2.1. L<strong>et</strong> (µ n ) n≥1 be a sequence of offspring distributions. For every n ≥ 1, denote<br />

by X n a Galton-Watson process with offspring distribution µ n , started at X n 0 = n. L<strong>et</strong> (m n) n≥1<br />

be a nondecreasing sequence of positive integers converging to infinity. We define a sequence of<br />

processes (Y n ) n≥1 by s<strong>et</strong>ting, for every t ≥ 0 and n ≥ 1,<br />

Y n<br />

t = n −1 X n [m nt] .<br />

29


Assume that, for every t ≥ 0, the sequence (Y n<br />

t ) n≥1 converges in distribution to Y t where Y =<br />

(Y t ,t ≥ 0) is an almost surely finite process such that P(Y δ > 0) > 0 for some δ > 0. Then, Y<br />

is a continuous-state branching process and the sequence of processes (Y n ) n≥1 converges to Y in<br />

distribution in the Skorokhod space D(R + ).<br />

Proof : It follows from the proof of Theorem 1 of [30] that Y is a CSBP. Then, thanks to<br />

Theorem 2 of [30], there exists a sequence of offspring distributions (ν n ) n≥1 and a nondecreasing<br />

sequence of positive integers (c n ) n≥1 such that we can construct for every n ≥ 1 a Galton-Watson<br />

process Z n started at c n and with offspring distribution ν n satisfying<br />

where the symbol<br />

(<br />

c −1<br />

n Zn [nt] ,t ≥ 0 )<br />

d<br />

−→<br />

n→∞ (Y t,t ≥ 0),<br />

d<br />

−→ indicates convergence in distribution in D(R + ).<br />

L<strong>et</strong> (m nk ) k≥1 be a strictly increasing subsequence of (m n ) n≥1 . For n ≥ 1, we s<strong>et</strong> B n = X n k<br />

and b n = n k if n = m nk for some k ≥ 1, and we s<strong>et</strong> B n = Z n and b n = c n if there is no<br />

k ≥ 1 such that n = m nk . Then, for every t ≥ 0, (b −1<br />

n Bn [nt] ) n≥1 converges in distribution to Y t .<br />

Applying Theorem 4.1 of [26], we obtain that<br />

( )<br />

b −1<br />

n Bn [nt] ,t ≥ 0 d<br />

−→ (Y t,t ≥ 0).<br />

n→∞<br />

In particular, we have,<br />

( )<br />

(2.2.2)<br />

Y n k d<br />

t ,t ≥ 0 −→ (Y t,t ≥ 0).<br />

k→∞<br />

As (2.2.2) holds for every strictly increasing subsequence of (m n ) n≥1 , we obtain the desired<br />

result.<br />

□<br />

2.2.1.2. Discr<strong>et</strong>e-state branching processes. A (continuous-time) discr<strong>et</strong>e-state branching process<br />

(in short DSBP) is a continuous-time Markov chain Y = (Y t ,t ≥ 0) with values in Z + whose<br />

transition probabilities (P t (i,j),t ≥ 0) i≥0,j≥0 satisfy the following branching property : for every<br />

i ∈ Z + , t ≥ 0 and |s| ≤ 1,<br />

⎛ ⎞i<br />

∞∑<br />

∞∑<br />

P t (i,j)s j = ⎝ P t (1,j)s j ⎠ .<br />

j=0<br />

We exclude the trivial case where P t (i,i) = 1 for every t ≥ 0 and i ∈ Z + . Then, there exist<br />

a > 0 and an offspring distribution γ with γ(1) = 0 such that the generator of Y can be written<br />

of the form ⎛<br />

⎞<br />

0 0 0 0 0 ...<br />

aγ(0) −a aγ(2) aγ(3) aγ(4) ...<br />

Q =<br />

0 2aγ(0) −2a 2aγ(2) aγ(3) ...<br />

.<br />

⎜<br />

⎝ 0 0 3aγ(0) −3a 3aγ(2) ... ⎟<br />

⎠<br />

.<br />

. . .. . .. . .. . ..<br />

Furthermore, it is well known that Y becomes extinct almost surely if and only if γ is critical<br />

or subcritical. We refer the reader to [7] and [49] for more d<strong>et</strong>ails.<br />

2.2.2. D<strong>et</strong>erministic trees.<br />

j=0<br />

30


2.2.2.1. The space (T,GH) of rooted compact R-trees. We start with a basic definition.<br />

Definition 2.2.1. A m<strong>et</strong>ric space (T ,d) is an R-tree if the following two properties hold for<br />

every σ 1 ,σ 2 ∈ T .<br />

(i) There is a unique isom<strong>et</strong>ric map f σ1 ,σ 2<br />

from [0,d(σ 1 ,σ 2 )] into T such that f σ1 ,σ 2<br />

(0) = σ 1<br />

and f σ1 ,σ 2<br />

(d(σ 1 ,σ 2 )) = σ 2 .<br />

(ii) If q is a continuous injective map from [0,1] into T such that q(0) = σ 1 and q(1) = σ 2 ,<br />

we have<br />

q([0,1]) = f σ1 ,σ 2<br />

([0,d(σ 1 ,σ 2 )]).<br />

A rooted R-tree is an R-tree with a distinguished vertex ρ = ρ(T ) called the root.<br />

In what follows, R-trees will always be rooted.<br />

L<strong>et</strong> (T ,d) be an R-tree with root ρ, and σ,σ 1 ,σ 2 ∈ T . We write [[σ 1 ,σ 2 ]] for the range of the<br />

map f σ1 ,σ 2<br />

. In particular, [[ρ,σ]] is the path going from the root to σ and can be interpr<strong>et</strong>ed as<br />

the ancestral line of σ.<br />

The height H(T ) of the R-tree T is defined by H(T ) = sup{d(ρ,σ) : σ ∈ T }. In particular,<br />

if T is compact, its height H(T ) is finite.<br />

Two rooted R-trees T and T ′ are called equivalent if there is a root-preserving isom<strong>et</strong>ry that<br />

maps T onto T ′ . We denote by T the s<strong>et</strong> of all equivalence classes of rooted compact R-trees.<br />

We often abuse notation and identify a rooted compact R-tree with its equivalence class.<br />

The s<strong>et</strong> T can be equipped with the pointed Gromov-Hausdorff distance, which is defined<br />

as follows. If (E,δ) is a m<strong>et</strong>ric space, we use the notation δ Haus for the usual Hausdorff m<strong>et</strong>ric<br />

b<strong>et</strong>ween compact subs<strong>et</strong>s of E. Then, if T and T ′ are two rooted compact R-trees with respective<br />

roots ρ and ρ ′ , we define the distanceGH(T , T ′ ) as<br />

{<br />

GH(T , T ′ (<br />

) = inf δ Haus φ(T ),φ ′ (T ′ ) ) }<br />

∨ δ(φ(ρ),φ ′ (ρ ′ )) ,<br />

where the infimum is over all isom<strong>et</strong>ric embeddings φ : T −→ E and φ ′ : T ′ −→ E into a<br />

common m<strong>et</strong>ric space (E,δ). We see thatGH(T , T ′ ) only depends on the equivalence classes<br />

of T and T ′ . Furthermore, according to Theorem 2 in [24],GH defines a m<strong>et</strong>ric on T that<br />

makes it compl<strong>et</strong>e and separable. Notice thatGH(T , T ′ ) makes sense more generally if T and<br />

T ′ are pointed compact m<strong>et</strong>ric spaces (see e.g. Chapter 7 in [12]). We will use this in the proof<br />

of Lemma 2.2.2 below.<br />

We equip T with its Borel σ-field.<br />

If T ∈ T, we s<strong>et</strong> T ≤t = {σ ∈ T : d(ρ,σ) ≤ t} for every t ≥ 0. Plainly, T ≤t is an R-tree whose<br />

root is the same as the root of T . Note that the mapping T ↦−→ T ≤t from T into T is Lipschitz<br />

for the Gromov-Hausdorff m<strong>et</strong>ric.<br />

2.2.2.2. The R-tree coded by a function. We now recall a construction of rooted compact<br />

R-trees which is described in [22]. L<strong>et</strong> g : [0,+∞) −→ [0,+∞) be a continuous function with<br />

compact support satisfying g(0) = 0. We exclude the trivial case where g is identically zero. For<br />

every s,t ≥ 0, we s<strong>et</strong><br />

m g (s,t) = inf g(r)<br />

r∈[s∧t,s∨t]<br />

and<br />

d g (s,t) = g(s) + g(t) − 2m g (s,t).<br />

31


We define an equivalence relation ∼ on [0,+∞) by declaring that s ∼ t if and only if d g (s,t) = 0<br />

(or equivalently if and only if g(s) = g(t) = m g (s,t)). L<strong>et</strong> T g be the quotient space<br />

T g = [0,+∞)/ ∼ .<br />

Then, d g induces a m<strong>et</strong>ric on T g and we keep the notation d g for this m<strong>et</strong>ric. According to<br />

Theorem 2.1 of [22], the m<strong>et</strong>ric space (T g ,d g ) is a compact R-tree. By convention, its root is the<br />

equivalence class of 0 for ∼ and is denoted by ρ g .<br />

2.2.2.3. Subtrees of a tree above a fixed level. L<strong>et</strong> (T ,d) ∈ T and t > 0. Denote by T i,◦ ,i ∈ I<br />

the connected components of the open s<strong>et</strong> T >t = {σ ∈ T : d(ρ(T ),σ) > t}. L<strong>et</strong> i ∈ I. Then<br />

the ancestor of σ at level t, that is, the unique vertex on the line segment [[ρ,σ]] at distance<br />

t from ρ, must be the same for all σ ∈ T i,◦ . We denote by σ i this common ancestor and s<strong>et</strong><br />

T i = T i,◦ ∪ {σ i }. Thus T i is a compact rooted R-tree with root σ i . The trees T i ,i ∈ I are called<br />

the subtrees of T above level t. We now consider, for every h > 0,<br />

Z(t,t + h)(T ) = #{i ∈ I : H(T i ) > h}.<br />

By a compactness argument, we can easily verify that Z(t,t + h)(T ) < ∞.<br />

2.2.2.4. Discr<strong>et</strong>e trees and real trees. We start with some formalism for discr<strong>et</strong>e trees. We<br />

first introduce the s<strong>et</strong> of labels<br />

U = ⋃ n≥0<br />

N n ,<br />

where by convention N 0 = {∅}. An element of U is a sequence u = u 1 ...u n , and we s<strong>et</strong> |u| = n<br />

so that |u| represents the generation of u. In particular, |∅| = 0. If u = u 1 ...u n and v = v 1 ...v m<br />

belong to U, we write uv = u 1 ...u n v 1 ...v m for the concatenation of u and v. In particular,<br />

∅u = u∅ = u. The mapping π : U\{∅} −→ U is defined by π(u 1 ...u n ) = u 1 ... u n−1 (π(u) is<br />

the father of u). Note that π k (u) = ∅ if k = |u|.<br />

A rooted ordered tree θ is a finite subs<strong>et</strong> of U such that<br />

(i) ∅ ∈ θ,<br />

(ii) u ∈ θ\{∅} ⇒ π(u) ∈ θ,<br />

(iii) for every u ∈ θ, there exists a number k u (θ) ≥ 0 such that uj ∈ θ if and only if<br />

1 ≤ j ≤ k u (θ).<br />

We denote by A the s<strong>et</strong> of all rooted ordered trees. If θ ∈ A, we write H(θ) for the height of θ, that<br />

is H(θ) = max{|u| : u ∈ θ}. And for every u ∈ θ, we define τ u θ ∈ A by τ u θ = {v ∈ U : uv ∈ θ}.<br />

This is the tree θ shifted at u.<br />

L<strong>et</strong> us define an equivalence relation on A by s<strong>et</strong>ting θ ≡ θ ′ if and only if we can find a<br />

permutation ϕ u of the s<strong>et</strong> {1,... ,k u (θ)} for every u ∈ θ such that k u (θ) ≥ 1, in such a way that<br />

{<br />

θ ′ (<br />

= {∅} ∪ ϕ ∅ u<br />

1 ) (<br />

ϕ u 1 u<br />

2 ) ...ϕ u 1 ...u n−1 (un ) : u 1 ...u n ∈ θ, n ≥ 1}<br />

.<br />

In other words θ ≡ θ ′ if they correspond to the same unordered tree. L<strong>et</strong> A = A/ ≡ be the<br />

associated quotient space and l<strong>et</strong>Ô:A−→ A be the canonical projection. It is immediate that<br />

if θ ≡ θ ′ , then k ∅ (θ) = k ∅ (θ ′ ). So, for every ξ ∈ A, we may define k ∅ (ξ) = k ∅ (θ) where θ is any<br />

representative of ξ.<br />

L<strong>et</strong> us fix ξ ∈ A such that k ∅ (ξ) = k > 0 and choose a representative θ of ξ. We may<br />

define {ξ 1 ,... ,ξ k } = {Ô(τ 1 θ),...,Ô(τ k θ)} as the unordered family of subtrees of ξ above the<br />

32


first generation. Then, if F : A k −→ R + is any symm<strong>et</strong>ric measurable function, we have<br />

( #Ô−1 (ξ) ) −1<br />

∑<br />

F(τ 1 θ,...,τ k θ)<br />

(2.2.3)<br />

=<br />

θ∈Ô−1 (ξ)<br />

( ) −1 ( ) −1 ∑<br />

#Ô−1 (ξ 1 ) ... #Ô−1 (ξ k )<br />

θ 1 ∈Ô−1 (ξ 1 )<br />

...<br />

∑<br />

θ k ∈Ô−1 (ξ k )<br />

F(θ 1 ,...,θ k ).<br />

Note that the right-hand side of (2.2.3) is well defined since it is symm<strong>et</strong>ric in {ξ 1 ,...,ξ k }. The<br />

identity (2.2.3) is a simple combinatorial fact, whose proof is left to the reader.<br />

A marked tree is a pair T = (θ, {h u } u∈θ ) where θ ∈ A and h u ≥ 0 for every u ∈ θ. We denote<br />

by M the s<strong>et</strong> of all marked trees. We can associate with every marked tree T = (θ, {h u } u∈θ ) ∈ M,<br />

an R-tree T T in the following way. L<strong>et</strong> R θ be the vector space of all mappings from θ into R.<br />

Write (e u ,u ∈ θ) for the canonical basis of R θ . We define l ∅ = 0 and l u = ∑ |u|<br />

k=1 h π k (u)e π k (u) for<br />

u ∈ θ \ {∅}. L<strong>et</strong> us s<strong>et</strong><br />

T T =<br />

u∈θ[l ⋃ u ,l u + h u e u ].<br />

T T is a connected union of line segments in R θ . It is equipped with the distance d T such that<br />

d T (a,b) is the length of the shortest path in T T b<strong>et</strong>ween a and b, and can be rooted in ρ(T T ) = 0<br />

so that it becomes a rooted compact R-tree.<br />

If θ ∈ A, we write T θ for the R-tree T T where T = (θ, {h u } u∈θ ) with h ∅ = 0 and h u = 1<br />

for every u ∈ θ \ {∅}, and we write d θ for the associated distance. We then s<strong>et</strong> m ∅ = 0 and<br />

m u = ∑ |u|−1<br />

k=0 e π k (u) = l u + e u for every u ∈ θ \ {∅}.<br />

It is easily checked that T θ = T θ′ if θ ≡ θ ′ . Thus for every ξ ∈ A, we may write T ξ for the<br />

tree T θ where θ is any representative of ξ.<br />

We will now explain how to approximate a general tree T in T by a discr<strong>et</strong>e type tree. L<strong>et</strong><br />

ε > 0 and s<strong>et</strong> T (ε) = {T ∈ T : H(T ) > ε}. For every T in T (ε) , we can construct by induction<br />

an element ξ ε (T ) of A in the following way.<br />

• If T ∈ T (ε) satisfies H(T ) ≤ 2ε, we s<strong>et</strong> ξ ε (T ) =Ô({∅}).<br />

• L<strong>et</strong> n be a positive integer. Assume that we have defined ξ ε (T ) for every T ∈ T (ε) such<br />

that H(T ) ≤ (n + 1)ε. L<strong>et</strong> T be an R-tree such that (n + 1)ε < H(T ) ≤ (n + 2)ε. We<br />

s<strong>et</strong> k = Z(ε,2ε)(T ) and we denote by T 1 ,... , T k the k subtrees of T above level ε with<br />

height greater than ε. Then ε < H(T i ) ≤ (n + 1)ε for every i ∈ {1,... ,k}, so we can<br />

define ξ ε (T i ). L<strong>et</strong> us choose a representative θ i of ξ ε (T i ) for every i ∈ {1,... ,k}. We<br />

s<strong>et</strong><br />

ξ ε (T ) =Ô({∅} ∪ 1θ 1 ∪ ... ∪ kθ k) ,<br />

where iθ i = {iu : u ∈ θ i }. Clearly this does not depend on the choice of the representatives<br />

θ i .<br />

If r > 0 and T is a compact rooted R-tree with m<strong>et</strong>ric d, we write rT for the same tree<br />

equipped with the m<strong>et</strong>ric rd.<br />

Lemma 2.2.2. For every ε > 0 and every T ∈ T (ε) , we have<br />

( )<br />

(2.2.4) GH εT ξε (T ) , T ≤ 4ε.<br />

Proof : L<strong>et</strong> ε > 0 and T ∈ T. L<strong>et</strong> θ be any representative of ξ ε (T ). Recall the notation<br />

(m u ,u ∈ θ). We can construct a mapping φ : θ −→ T such that :<br />

33


(i) for every σ ∈ T , there exists u ∈ θ such that d(σ,φ(u)) ≤ 2ε,<br />

(ii) for every u ∈ θ, d(φ(u),ρ) = ε|u| where ρ denotes the root of T ,<br />

(iii) for every u,u ′ ∈ θ, 0 ≤ εd θ (m u ,m u ′) − d(φ(u),φ(u ′ )) ≤ 2ε.<br />

To be specific, we always take φ(∅) = ρ, which suffices for the construction if H(T ) ≤ 2ε. If<br />

(n + 1)ε < H(T ) ≤ (n + 2)ε for some n ≥ 1, we have as above<br />

θ = {∅} ∪ 1θ 1 ∪ ... ∪ kθ k ,<br />

where θ 1 ,... ,θ k are representatives of respectively ξ ε (T 1 ),... ,ξ ε (T k ), if T 1 ,... , T k are the<br />

subtrees of T above level ε with height greater than ε. With an obvious notation we define<br />

φ(ju) = φ j (u) for every j ∈ {1,... ,k} and u ∈ θ j . Properties (i)-(iii) are then easily checked by<br />

induction.<br />

L<strong>et</strong> us now s<strong>et</strong> Φ = {φ(u),u ∈ θ} and M = {m u ,u ∈ θ}. We equip Φ with the m<strong>et</strong>ric<br />

induced by d and M with the m<strong>et</strong>ric induced by εd θ . Then, Φ and M can be viewed as pointed<br />

compact m<strong>et</strong>ric spaces with respective roots ρ and 0. It is immediate thatGH(εT θ ,M) ≤ ε.<br />

Furthermore, Property (i) above implies thatGH(T ,Φ) ≤ 2ε. At last, according to Lemma 2.3<br />

in [24] and Property (iii) above, we haveGH(Φ,M) ≤ ε. Lemma 2.2.2 then follows from the<br />

triangle inequality forGH.<br />

□<br />

2.2.3. Lévy trees. Roughly speaking, a Lévy tree is a T-valued random variable which is<br />

associated with a CSBP in such a way that it describes the genealogy of a population evolving<br />

according to this CSBP.<br />

2.2.3.1. The measure Θ ψ . We consider on a probability space (Ω,P) a ψ-CSBP Y = (Y t ,t ≥<br />

0), where the function ψ is of the form (2.2.1), and we suppose that Y becomes extinct almost<br />

surely. This condition is equivalent to<br />

∫ ∞<br />

du<br />

(2.2.5)<br />

ψ(u) < ∞.<br />

This implies that at least one of the following two conditions holds :<br />

∫<br />

(2.2.6) β > 0 or rπ(dr) = ∞.<br />

1<br />

(0,1)<br />

The Lévy tree associated to Y will be defined as the tree coded by the so-called height<br />

process, which is a functional of the Lévy process with Laplace exponent ψ. L<strong>et</strong> us denote by<br />

X = (X t ,t ≥ 0) a Lévy process on (Ω,P) with Laplace exponent ψ. This means that X is a<br />

Lévy process with no negative jumps, and that for every λ,t ≥ 0,<br />

E(exp(−λX t )) = exp(tψ(λ)).<br />

Then, X does not drift to +∞ and has paths of infinite variation (by (2.2.6)).<br />

We can define the height process H = (H t ,t ≥ 0) by the following approximation :<br />

1<br />

H t = lim<br />

ε→0 ε<br />

∫ t<br />

0½{X s≤I s t +ε} ds,<br />

where It s = inf{X r : s ≤ r ≤ t} and the convergence holds in probability (see Chapter 1 in [21]).<br />

Informally, we can say that H measures the size of the s<strong>et</strong> {s ∈ [0,t] : X s− ≤ It s }. Thanks to<br />

condition (2.2.5), we know that the process H has a continuous modification (see Theorem 4.7<br />

in [21]). From now on, we consider only this modification.<br />

34


L<strong>et</strong> us now s<strong>et</strong> I t = inf {X s : 0 ≤ s ≤ t} for every t ≥ 0, and consider the process X − I =<br />

(X t −I t ,t ≥ 0). We recall that X −I is a strong Markov process for which the point 0 is regular.<br />

The process −I is a local time for X − I at level 0. We write N for the associated excursion<br />

measure. We l<strong>et</strong> ∆(de) be the “law” of (H s ,s ≥ 0) under N. This makes sense because the<br />

values of the height process in an excursion of X −I away from 0 only depend on that excursion<br />

(see section 1.2 in [21]). Then, ∆(de) is a σ-finite measure on C([0, ∞)), and is supported on<br />

functions with compact support such that e(0) = 0.<br />

The Lévy tree is the tree (T e ,d e ) coded by the function e, in the sense of section 2.2.2.2,<br />

under the measure ∆(de). We denote by Θ ψ the σ-finite measure on T which is the “law”of the<br />

Lévy tree, that is the image of ∆(de) under the mapping e ↦−→ T e .<br />

2.2.3.2. A discr<strong>et</strong>e approximation of the Lévy tree. L<strong>et</strong> us now recall that the Lévy tree is<br />

the limit in the Gromov-Hausdorff distance of suitably rescaled Galton-Watson trees.<br />

We start by recalling the definition of Galton-Watson trees which was given informally in<br />

the introduction above. L<strong>et</strong> µ be a critical or subcritical offspring distribution. We exclude the<br />

trivial case where µ(1) = 1. Then, there exists a unique probability measure Π µ on A such that :<br />

(i) for every p ≥ 0, Π µ (k ∅ = p) = µ(p),<br />

(ii) for every p ≥ 1 with µ(p) > 0, under the probability measure Π µ (· | k ∅ = p), the shifted<br />

trees τ 1 θ,...,τ p θ are independent and distributed according to Π µ .<br />

Recall that if r > 0 and T is a compact rooted R-tree with m<strong>et</strong>ric d, we write rT for the<br />

same tree equipped with the m<strong>et</strong>ric rd. The following result is Theorem 4.1 in [22].<br />

Theorem 2.2.3. L<strong>et</strong> (µ n ) n≥1 be a sequence of critical or subcritical offspring distributions.<br />

For every n ≥ 1, l<strong>et</strong> us denote by X n a Galton-Watson process with offspring distribution µ n ,<br />

started at X n 0 = n. L<strong>et</strong> (m n) n≥1 be a nondecreasing sequence of positive integers converging to<br />

infinity. We define a sequence of processes (Y n ) n≥1 by s<strong>et</strong>ting, for every t ≥ 0 and n ≥ 1,<br />

Y n<br />

t = n −1 X n [m nt] .<br />

Assume that, for every t ≥ 0, the sequence (Yt n ) n≥1 converges in distribution to Y t where Y =<br />

(Y t ,t ≥ 0) is a ψ-CSBP which becomes extinct almost surely. Assume furthermore that for every<br />

δ > 0,<br />

lim inf P(Y n<br />

n→∞<br />

δ = 0) > 0.<br />

Then, for every a > 0, the law of the R-tree m −1<br />

n T θ under Π µn (· | H(θ) ≥ [am n ]) converges<br />

as n → ∞ to the probability measure Θ ψ (· | H(T ) > a) in the sense of weak convergence of<br />

measures in the space T.<br />

2.3. Proof of Theorem 2.1.1<br />

L<strong>et</strong> Θ be an infinite measure on (T,GH) satisfying the assumptions of Theorem 1.1. Clearly<br />

Θ is σ-finite.<br />

We start with two important lemmas that will be used throughout this section. L<strong>et</strong> us first<br />

define v : (0, ∞) −→ (0, ∞) by v(t) = Θ(H(T ) > t) for every t > 0. For every t > 0, we denote<br />

by Θ t the probability measure Θ(· | H(T ) > t).<br />

Lemma 2.3.1. The function v is nonincreasing, continuous and verifies<br />

v(t)<br />

v(t)<br />

−→ ∞,<br />

t→0<br />

−→ 0.<br />

t→∞<br />

35


Proof : We only have to prove the continuity of v. To this end, we argue by contradiction<br />

and assume that there exists t > 0 such that Θ(H(T ) = t) > 0. L<strong>et</strong> s > 0 and u ∈ (0,t) such<br />

that v(u) > v(t). From the regenerative property (R), we have<br />

Θ s (H(T ) = s + t) = Θ s( )<br />

Θ s (H(T ) = s + t | Z(s,s + u))<br />

= Θ s ( Z(s,s + u)Θ u (H(T ) = t)(Θ u (H(T ) ≤ t)) Z(s,s+u)−1)<br />

=<br />

> 0.<br />

Θ(H(T ) = t)<br />

v(u)<br />

(<br />

Θ<br />

(Z(s,s s + u) 1 − v(t) ) )<br />

Z(s,s+u)−1<br />

v(u)<br />

We have shown that Θ(H(T ) = t + s) > 0 for every s > 0. This is absurd since Θ is σ-finite. □<br />

Lemma 2.3.2. For every t > 0 and 0 < a < b, the conditional law of the random variable<br />

Z(t,t+b), under the probability measure Θ t and given Z(t,t+a), is a binomial distribution with<br />

param<strong>et</strong>ers Z(t,t+a) and v(b)/v(a) (where we define the binomial distribution with param<strong>et</strong>ers<br />

0 and p ∈ [0,1] as the Dirac measure δ 0 ).<br />

Proof : This is a straightforward consequence of the regenerative property.<br />

□<br />

2.3.1. The CSBP derived from Θ. In this section, we consider a random forest of trees<br />

derived from a Poisson point measure with intensity Θ. We associate with this forest a family<br />

of Galton-Watson processes. We then construct local times at every level a > 0 as limits of the<br />

rescaled Galton-Watson processes. Finally we show that the local time process is a CSBP.<br />

L<strong>et</strong> us now fix the framework. We consider a probability space (Ω, P) and on this space a<br />

Poisson point measure N = ∑ i∈I δ T i<br />

on T, whose intensity is the measure Θ.<br />

2.3.1.1. A family of Galton-Watson trees. We start with some notation that we need in<br />

the first lemma. We consider on another probability space (Ω ′ , P ′ ), a collection (θ ξ ,ξ ∈ A) of<br />

independent A-valued random variables such that for every ξ ∈ A, θ ξ is distributed uniformly<br />

overÔ−1 (ξ). In what follows, to simplify notation, we identify an element ξ of the s<strong>et</strong> A with<br />

the subs<strong>et</strong>Ô−1 (ξ) of A.<br />

Recall the definition of ξ ε (T ) before Lemma 2.2.2.<br />

Lemma 2.3.3. L<strong>et</strong> us define for every ε > 0, a mapping θ (ε) from T (ε) × Ω ′ into A by<br />

θ (ε) (T ,ω) = θ ξ ε (T )(ω).<br />

Then for every positive integer p, the law of the random variable θ (ε) under the probability<br />

measure Θ pε ⊗ P ′ is Π µε (· | H(θ) ≥ p − 1) where µ ε denotes the law of Z(ε,2ε) under Θ ε .<br />

Proof : Since {H(T ) > pε} × Ω ′ = {H(θ (ε) ) ≥ p − 1} for every p ≥ 1, it suffices to show<br />

the result for p = 1. L<strong>et</strong> k be a nonnegative integer. According to the construction of ξ ε (T ), we<br />

have<br />

( Θ ε ⊗ P ′ k ∅<br />

(θ (ε)) )<br />

= k = Θ ε (Z(ε,2ε) = k) = µ ε (k).<br />

36


L<strong>et</strong> us fix k ≥ 1 with µ ε (k) > 0. L<strong>et</strong> F : A k −→ R + be a symm<strong>et</strong>ric measurable function. Then<br />

we have<br />

( Θ ε ⊗ P ′ F<br />

(τ 1 θ (ε) ,...,τ k θ (ε)) ∣ (<br />

∣ k∅ θ (ε)) )<br />

= k<br />

⎛<br />

= Θ ε ⊗ P ′ ⎝ ∑<br />

⎞<br />

F (τ 1 θ,...,τ k θ)½{θ ξ ε (T ) =θ}<br />

∣ Z(ε,2ε) = k ⎠<br />

(2.3.1)<br />

θ∈ξ ε (T )<br />

= Θ ε ⎛<br />

⎝(#ξ ε (T )) −1<br />

∑<br />

θ∈ξ ε (T )<br />

F (τ 1 θ,...,τ k θ)<br />

⎞<br />

∣ Z(ε,2ε) = k ⎠ .<br />

On the event {Z(ε,2ε) = k}, we write T 1 ,...,T k for the k subtrees of T above level ε with<br />

height greater than ε. Then, Formula (2.2.3) and the regenerative property yield<br />

⎛<br />

⎞<br />

∑<br />

Θ ε ⎝(#ξ ε (T )) −1<br />

F (τ 1 θ,...,τ k θ)<br />

∣ Z(ε,2ε) = k ⎠<br />

θ∈ξ ε (T )<br />

= Θ ε ⎛<br />

⎝(#ξ ε (T 1 )) −1 ...(#ξ ε (T k )) −1<br />

=<br />

=<br />

∫<br />

∫<br />

∑<br />

θ 1 ∈ξ ε (T 1 )<br />

...<br />

Θ ε (dT 1 )... Θ ε (dT k )(#ξ ε (T 1 )) −1 ...(#ξ ε (T k )) −1<br />

∑<br />

θ k ∈ξ ε (T k )<br />

∑<br />

θ 1 ∈ξ ε (T 1 )<br />

⎞<br />

F (θ 1 ,... ,θ k )<br />

∣ Z(ε,2ε) = k ⎠<br />

...<br />

∑<br />

θ k ∈ξ ε (T k )<br />

Θ ε ⊗ P ′ (dT 1 ,dω ′ 1 )... Θε ⊗ P ′ (dT k ,dω ′ k )F (<br />

θ (ε) (T 1 ,ω ′ 1 ),... ,θ(ε) (T k ,ω ′ k ) )<br />

,<br />

F (θ 1 ,... ,θ k )<br />

as in (2.3.1). We ( have thus proved that<br />

Θ ε ⊗ P ′ F<br />

(τ 1 θ (ε) ,... ,τ k θ (ε))∣ (<br />

∣ k∅ θ (ε)) )<br />

= k<br />

∫<br />

(<br />

)<br />

(2.3.2) = Θ ε ⊗ P ′ (dT 1 ,dω 1 ′ )... Θε ⊗ P ′ (dT k ,dω k ′ )F θ (ε) (T 1 ,ω 1 ′ ),...,θ(ε) (T k ,ω k ′ ) .<br />

Note that for every permutation ϕ of the s<strong>et</strong> {1,... ,k}, the k-tuples (τ ϕ(1) θ (ε) ,...,τ ϕ(k) θ (ε) ) and<br />

(τ 1 θ (ε) ,...,τ k θ (ε) ) have the same distribution under Θ ε ⊗P ′ . Then, (2.3.2) means that the law of<br />

θ (ε) under Θ ε ⊗ P ′ satisfies the branching property of the Galton-Watson trees. This compl<strong>et</strong>es<br />

the proof of the desired result.<br />

□<br />

Recall that ∑ i∈I δ T i<br />

is a Poisson point measure on T with intensity Θ. L<strong>et</strong> us now s<strong>et</strong>, for<br />

every t,h > 0,<br />

Z(t,t + h) = ∑ Z(t,t + h)(T i ).<br />

i∈I<br />

For every ε > 0, we define a process X ε = (Xk ε ,k ≥ 0) on (Ω, P) by the formula<br />

= Z(kε,(k + 1)ε), k ≥ 0.<br />

X ε k<br />

Proposition 2.3.4. For every ε > 0, the process X ε is a Galton-Watson process whose initial<br />

distribution is the Poisson distribution with param<strong>et</strong>er v(ε) and whose offspring distribution is<br />

µ ε .<br />

Proof : First note that X0 ε = N(H(T ) > ε) is Poisson with param<strong>et</strong>er Θ(H(T ) > ε) = v(ε).<br />

Then l<strong>et</strong> p be a positive integer. We know from a classical property of Poisson measures that,<br />

37


under the probability measure P and conditionally on the event {X0 ε = p}, the atoms of N that<br />

belong to the s<strong>et</strong> T ε are distributed as p i.i.d. variables with distribution Θ ε . Furthermore, it<br />

follows from Lemma 2.3.3 that under Θ ε , the process (Z(kε,(k + 1)ε)) k≥0 is a Galton-Watson<br />

process started at one with offspring distribution µ ε . This compl<strong>et</strong>es the proof. □<br />

As a consequence, we g<strong>et</strong> the next proposition, which we will use throughout this work.<br />

Proposition 2.3.5. For every t > 0 and h > 0, we have Θ(Z(t,t + h)) ≤ v(h).<br />

Proof : Since compact R-trees have finite height, the Galton-Watson process X ε dies out P<br />

a.s. This implies that µ ε is critical or subcritical so that (Xk ε ,k ≥ 0) is a supermartingale. L<strong>et</strong><br />

t,h > 0. We can find ε > 0 and k ∈ N such that t = kε and ε ≤ h. Thus we have,<br />

(2.3.3) Θ(Z(t,t + ε)) = Θ(Z(kε,(k + 1)ε)) = E(X ε k ) ≤ E(X ε 0) = v(ε).<br />

Using Lemma 2.3.2 and (2.3.3), we g<strong>et</strong><br />

2.3.1.2. A local time process.<br />

Θ(Z(t,t + h)) = Θ<br />

(<br />

Z(t,t + ε) v(h) )<br />

≤ v(h).<br />

v(ε)<br />

Proposition 2.3.6. For every t ≥ 0, there exists a random variable L t on the space T such<br />

that Θ a.e.,<br />

Z(t,t + h)<br />

v(h)<br />

−→ L t.<br />

h→0<br />

Proof : L<strong>et</strong> us start with the case t = 0. As Z(0,h) =½{H(T )>h} for every h > 0, Lemma<br />

2.3.1 gives v(h) −1 Z(0,h) → 0 Θ a.e. as h → 0, so we s<strong>et</strong> L 0 = 0.<br />

L<strong>et</strong> us now fix t > 0. Thanks to Lemma 2.3.1, we can define a decreasing sequence (ε n ) n≥1<br />

by the condition v(ε n ) = n 4 for every n ≥ 1. We claim that there exists a random variable L t<br />

on the space T such that, Θ a.e.,<br />

Z(t,t + ε n )<br />

(2.3.4)<br />

n 4 −→ L t .<br />

n→∞<br />

Indeed, using Lemma 2.3.2, we have, for every n ≥ 1,<br />

( ∣∣∣∣<br />

Θ t Z(t,t + ε n )<br />

n 4 − Z(t,t + ε ∣ )<br />

n+1) ∣∣∣<br />

2<br />

(n + 1) 4<br />

( ( ∣∣∣∣<br />

= Θ t 1<br />

n 4<br />

n 8Θt Z(t,t + ε n ) −<br />

(n + 1) 4Z(t,t + ε ∣ ))<br />

n+1) ∣ ∣∣∣<br />

∣2<br />

Z(t,t + ε n+1 )<br />

( )<br />

Z(t,t +<br />

≤ Θ t εn+1 )<br />

4n 8<br />

(2.3.5)<br />

≤<br />

(n + 1)4<br />

4v(t)n 8 ,<br />

where the last bound follows from Proposition 2.3.5 and the definition of ε n+1 . Thanks to the<br />

Cauchy-Schwarz inequality, we g<strong>et</strong><br />

(2.3.6) Θ t (∣ ∣∣∣ Z(t,t + ε n )<br />

n 4<br />

− Z(t,t + ε ∣)<br />

n+1) ∣∣∣<br />

(n + 1) 4 ≤<br />

(n + 1)2<br />

2n 4√ v(t) ≤ 2<br />

n 2√ v(t) .<br />

□<br />

38


The bound (2.3.6) implies<br />

(<br />

∑ ∞ Θ<br />

Z(t,t + ε n )<br />

∣ n 4<br />

In particular, Θ a.e.,<br />

Our claim (2.3.4) follows.<br />

n=1<br />

∞∑<br />

Z(t,t + ε n )<br />

∣ n 4<br />

n=1<br />

− Z(t,t + ε ∣)<br />

n+1) ∣∣∣<br />

(n + 1) 4 < ∞.<br />

− Z(t,t + ε n+1)<br />

(n + 1) 4 ∣ ∣∣∣<br />

< ∞.<br />

For every h ∈ (0,ε 1 ], we can find n ≥ 1 such that ε n+1 ≤ h ≤ ε n . Then, we have Z(t,t+ε n ) ≤<br />

Z(t,t + h) ≤ Z(t,t + ε n+1 ) Θ a.e., and n 4 ≤ v(h) ≤ (n + 1) 4 so that<br />

We then deduce from (2.3.4) that Θ a.e.,<br />

which compl<strong>et</strong>es the proof.<br />

Z(t,t + ε n ) Z(t,t + h)<br />

(n + 1) 4 ≤ ≤ Z(t,t + ε n+1)<br />

v(h) n 4 .<br />

Z(t,t + h)<br />

v(h)<br />

−→<br />

h→0 L t<br />

Definition 2.3.1. We define a process L = (L t ,t ≥ 0) on (Ω, P) by s<strong>et</strong>ting L 0 = 1 and for<br />

every t > 0,<br />

L t = ∑ L t (T i ).<br />

i∈I<br />

□<br />

Notice that L t (T ) = 0 if H(T ) ≤ t so that the above sum is finite a.s.<br />

Corollary 2.3.7. For every t ≥ 0, we have P a.s.<br />

Z(t,t + h)<br />

v(h)<br />

−→ L t.<br />

h→0<br />

Moreover, this convergence holds in L 1 (P) uniformly in t ∈ [0, ∞).<br />

Proof : The first assertion is an immediate consequence of Proposition 2.3.6. L<strong>et</strong> us focus<br />

on the second assertion. From Lemma 2.3.2 we have<br />

(<br />

)<br />

Θ n −4 Z(t,t + ε n ) − (n + 1) −4 Z(t,t + ε n+1 ) = 0<br />

for every t ≥ 0 and n ≥ 1. Thus, from the second moment formula for Poisson measures, we g<strong>et</strong>,<br />

for every t ≥ 0 and n ≥ 1,<br />

( (Z(t,t + εn )<br />

E<br />

n 4 − Z(t,t + ε ) ) ( 2 (Z(t,t<br />

n+1)<br />

+ εn )<br />

(n + 1) 4 = Θ<br />

n 4 − Z(t,t + ε ) ) 2<br />

n+1)<br />

(n + 1) 4 .<br />

Now, we have<br />

( (Z(0,εn<br />

)<br />

Θ<br />

n 4<br />

− Z(0,ε ) ) ( 2 ) ) 2<br />

n+1)<br />

(½{H(T )>ε<br />

(n + 1) 4 = Θ<br />

n} )>ε n+1 }<br />

−½{H(T<br />

n 4 (n + 1) 4 = 1 n 4 − 1<br />

(n + 1) 4<br />

and for every t > 0, thanks to the bound (2.3.5),<br />

( (Z(t,t + εn )<br />

Θ<br />

n 4 − Z(t,t + ε ) ) 2<br />

n+1) (n + 1)4<br />

(n + 1) 4 ≤<br />

4n 8 .<br />

39


So for every t ≥ 0 and n ≥ 1, we have from the Cauchy-Schwarz inequality<br />

(∣ ∣∣∣ Z(t,t + ε n )<br />

(2.3.7) E<br />

n 4 − Z(t,t + ε ∣)<br />

n+1) ∣∣∣ (n + 1)2<br />

(n + 1) 4 ≤<br />

n 4 .<br />

Then n −4 Z(t,t + ε n ) → L t in L 1 as n → ∞ and, for every n ≥ 2,<br />

(∣ ∣)<br />

∣∣∣ Z(t,t + ε n ) ∣∣∣<br />

∞∑ (k + 1) 2 ∞∑ 4<br />

E<br />

n 4 − L t ≤<br />

k 4 ≤<br />

k 2 ≤ 8 n .<br />

k=n<br />

In the same way as in the proof of (2.3.7), we have the following inequality : if h ∈ (0,ε 1 ], t ≥ 0<br />

and n is a positive integer such that ε n+1 ≤ h ≤ ε n ,<br />

(∣ ) √ ∣∣∣ Z(t,t + ε n ) Z(t,t + h)<br />

E<br />

n 4 − v(h)<br />

v(h) ∣ ≤<br />

n 4 ≤ √ 16 .<br />

v(h)<br />

Then, for every h ∈ (0,ε 2 ] and t ≥ 0, we g<strong>et</strong><br />

(∣ ∣) ∣∣∣ Z(t,t + h) ∣∣∣ (<br />

E<br />

− L t ≤ 16 v(h) −1/2 + v(h) −1/4) ,<br />

v(h)<br />

which compl<strong>et</strong>es the proof.<br />

We will now establish a regularity property of the process (L t ,t ≥ 0).<br />

Proposition 2.3.8. The process (L t ,t ≥ 0) admits a modification, denoted by ( ˜L t ,t ≥ 0),<br />

which is right-continuous with left-limits, and which has no fixed discontinuities.<br />

Proof : We start with two lemmas.<br />

Lemma 2.3.9. There exists λ ≥ 0 such that E(L t ) = e −λt for every t ≥ 0.<br />

Proof : We claim that the function t ∈ [0,+∞) ↦−→ E(L t ) is multiplicative, meaning that<br />

for every t,s ≥ 0, E(L t+s ) = E(L t )E(L s ). As L 0 = 1 by definition, E(L 0 ) = 1. L<strong>et</strong> t,s > 0 and<br />

0 < h < s. L<strong>et</strong> us denote by T 1 ,... , T Z(t,t+h) the subtrees of T above level t with height greater<br />

than h. Then, using the regenerative property, we can write<br />

⎛<br />

Θ(Z(t + s,t + s + h)) = Θ ⎝<br />

Z(t,t+h)<br />

∑<br />

i=1<br />

⎞<br />

k=n<br />

(<br />

)<br />

Z(s,s + h)(T i ) ⎠ = Θ Z(t,t + h)Θ h (Z(s,s + h)) ,<br />

which implies<br />

( )<br />

Z(s,s + h)<br />

(2.3.8) E(Z(t + s,t + s + h)) = E(Z(t,t + h))E<br />

.<br />

v(h)<br />

Thus, dividing by v(h) and l<strong>et</strong>ting h → 0 in (2.3.8), we g<strong>et</strong> our claim from Corollary 2.3.7.<br />

Moreover, thanks to Proposition 2.3.5 and Corollary 2.3.7, we know that E(L t ) ≤ 1 for every<br />

t ≥ 0. Then, we obtain in particular that the function t ∈ [0, ∞) ↦−→ E(L t ) is nonincreasing.<br />

To compl<strong>et</strong>e the proof, we have to check that E(L t ) > 0 for every t > 0. If we assume that<br />

E(L t ) = 0 for some t > 0 then L t = 0, Θ a.e. L<strong>et</strong> s,h > 0 such that 0 < h < s. With the same<br />

notation as in the beginning of the proof, we can write<br />

Θ(H(T ) > t + s) = Θ ( ∃i ∈ {1,... ,Z(t,t + h)} : H(T i ) > s )<br />

( (<br />

= Θ 1 − 1 − v(s) ) ) Z(t,t+h)<br />

(2.3.9)<br />

.<br />

v(h)<br />

□<br />

40


Now, thanks to Proposition 2.3.6, Θ a.e.,<br />

(<br />

1 − v(s) ) Z(t,t+h)<br />

−→ exp(−L t v(s)) = 1.<br />

v(h) h→0<br />

Moreover, Θ a.e.,<br />

(<br />

1 − 1 − v(s) ) Z(t,t+h)<br />

≤½{H(T<br />

v(h)<br />

)>t}.<br />

Then, using dominated convergence in (2.3.9) as h → 0, we obtain Θ(H(T ) > t + s) = 0 which<br />

contradicts the assumptions of Theorem 2.1.1.<br />

□<br />

Lemma 2.3.10. L<strong>et</strong> us denote by D = {k2 −n ,k ≥ 1,n ≥ 0} the s<strong>et</strong> of positive dyadic numbers<br />

and define G t = σ(L s ,s ∈ D,s ≤ t) for every t ∈ D. Then (L t ,t ∈ D) is a nonnegative<br />

supermartingale with respect to the filtration (G t ,t ∈ D).<br />

Proof : L<strong>et</strong> p be a positive integer, l<strong>et</strong> s 1 ,... ,s p ,s,t ∈ D such that s 1 < ... < s p ≤ s < t<br />

and l<strong>et</strong> f : R p −→ R + be a bounded continuous function. We can find a positive integer n<br />

such that 2 n t, 2 n s, and 2 n s i for i ∈ {1,... ,p} are nonnegative integers. The process X 2−n is a<br />

subcritical Galton-Watson process, so<br />

E<br />

(<br />

X 2−n<br />

2 n t f<br />

(<br />

X 2−n<br />

2 n s 1<br />

,... , X 2−n<br />

2 n s p<br />

))<br />

≤ E<br />

( (<br />

))<br />

X2 2−n<br />

n s f X2 2−n<br />

n s 1<br />

,... , X2 2−n<br />

n s p<br />

.<br />

Therefore we have also,<br />

( Z(t,t + 2 −n (<br />

) Z(s1 ,s 1 + 2 −n )<br />

E<br />

v(2 −n f<br />

) v(2 −n ,..., Z(s p,s p + 2 −n ))<br />

)<br />

) v(2 −n )<br />

( Z(s,s + 2 −n (<br />

) Z(s1 ,s 1 + 2 −n )<br />

≤ E<br />

v(2 −n f<br />

) v(2 −n ,..., Z(s p,s p + 2 −n ))<br />

)<br />

(2.3.10)<br />

) v(2 −n .<br />

)<br />

We can then use Corollary 2.3.7 to obtain<br />

E ( L t f ( L s1 ,...,L sp<br />

))<br />

≤ E<br />

(<br />

Ls f ( L s1 ,... , L sp<br />

))<br />

.<br />

We now compl<strong>et</strong>e the proof of Proposition 2.3.8. L<strong>et</strong> us s<strong>et</strong>, for every t ≥ 0,<br />

˜G t =<br />

⋂<br />

G s .<br />

s>t,s∈D<br />

From Lemma 2.3.10 and classical results on supermartingales, we can define a rightcontinuous<br />

supermartingale ( ˜L t ,t ≥ 0) with respect to the filtration (˜G t ,t ≥ 0) by s<strong>et</strong>ting,<br />

for every t ≥ 0,<br />

(2.3.11) ˜Lt = lim<br />

s↓t,s∈D L s,<br />

where the limit holds P a.s. and in L 1 (see e.g. Chapter VI in [16] for more d<strong>et</strong>ails). We claim<br />

that ( ˜L t ,t ≥ 0) is a càdlàg modification of (L t ,t ≥ 0) with no fixed discontinuities.<br />

We first prove that ( ˜L t ,t ≥ 0) is a modification of (L t ,t ≥ 0). For every t ≥ 0 and every<br />

sequence (s n ) n≥0 in D such that s n ↓ t as n ↑ ∞, we have thanks to (2.3.11) and Lemma 2.3.9,<br />

( )<br />

E ˜Lt = lim E(L s<br />

n→∞ n<br />

) = E(L t ).<br />

□<br />

41


L<strong>et</strong> us now show that for every t ≥ 0, L t ≤ ˜L t P a.s. L<strong>et</strong> α,ε > 0 and δ ∈ (0,1). Thanks to<br />

Corollary 2.3.7, we can find h 0 > 0 such that for every h ∈ (0,h 0 ) and n ≥ 0,<br />

(∣ ∣) ∣∣∣ Z(t,t + h) ∣∣∣<br />

E<br />

− L t ≤ εα,<br />

v(h)<br />

(∣ ∣) ∣∣∣ Z(s n ,s n + h) ∣∣∣<br />

E<br />

− L sn ≤ εα.<br />

v(h)<br />

We choose h ∈ (0,h 0 ) and n 0 ≥ 0 such that s n − t + h ≤ h 0 and v(h) ≤ (1 + δ)v(s n − t + h) for<br />

every n ≥ n 0 . We notice that Z(t,s n + h) ≤ Z(s n ,s n + h) so that, for every n ≥ n 0 ,<br />

P(L t > (1 + δ)L sn + ε)<br />

(<br />

≤ P L t − Z(t,s n + h)<br />

v(s n − t + h) > (1 + δ)L s n<br />

− (1 + δ) Z(s )<br />

n,s n + h)<br />

+ ε<br />

v(h)<br />

(∣ )<br />

(∣ ∣)<br />

∣∣∣<br />

≤ 2ε −1 Z(t,s n + h)<br />

E<br />

v(s n − t + h) − L ∣∣∣ t∣<br />

+ 2ε −1 Z(s n ,s n + h) ∣∣∣<br />

(1 + δ)E<br />

− L sn<br />

v(h)<br />

≤ 6α.<br />

We have thus shown that<br />

(2.3.12) P(L t > (1 + δ)L sn + ε) −→<br />

n→∞<br />

0.<br />

So, P(L t − (1 + δ) ˜L t > ε) = 0 for every ε > 0, implying that L t ≤ (1 + δ) ˜L t , P a.s. This leads<br />

us to the claim L t ≤ ˜L t a.s. Since we saw that E(L t ) = E( ˜L t ), we have L t = ˜L t P a.s. for every<br />

t ≥ 0.<br />

Now, ( ˜L t ,t ≥ 0) is a right-continuous supermartingale. Thus, ( ˜L t ,t ≥ 0) is also left-limited<br />

and we have E( ˜L t ) ≤ E( ˜L t− ) for every t > 0. Moreover, we can prove in the same way as we did<br />

for (2.3.12) that, for every t > 0 and every sequence (s n ,n ≥ 0) in D such that s n ↑ t as n ↑ ∞,<br />

(<br />

P ˜Lsn > (1 + δ) ˜L<br />

)<br />

t + ε<br />

−→ 0,<br />

n→∞<br />

implying that ˜L t− ≤ ˜L t , P a.s. So, L t = L t− P a.s. for every t > 0 meaning that ( ˜L t ,t ≥ 0) has<br />

no fixed discontinuities.<br />

□<br />

From now on, to simplify notation, we replace (L t ,t ≥ 0) by its càdlàg modification ( ˜L t ,t ≥ 0).<br />

2.3.1.3. The CSBP. We will prove that the suitably rescaled family of Galton-Watson processes<br />

(X ε ) ε>0 converges to the local time L.<br />

Thanks to Lemma 2.3.1, we can define a sequence (η n ) n≥1 by the condition v(η n ) = n for every<br />

n ≥ 1. We s<strong>et</strong> m n = [ηn −1 ] where [x] denotes the integer part of x. We recall from Proposition<br />

2.3.4 that X ηn is a Galton-Watson process on (Ω, P) whose initial distribution is the Poisson<br />

distribution with param<strong>et</strong>er n. For every n ≥ 1, we define a process Y n = (Yt n ,t ≥ 0) on (Ω, P)<br />

by the following formula,<br />

Yt n = n −1 X ηn<br />

[m , t ≥ 0.<br />

nt]<br />

Proposition 2.3.11. For every t ≥ 0, Y n t → L t in probability as n → ∞.<br />

Proof : L<strong>et</strong> t ≥ 0 and l<strong>et</strong> δ > 0. We can write<br />

P(|Yt n − L t| > 2δ) ≤ P (∣ ) (∣ ∣ )<br />

∣ Y<br />

n<br />

t − L ηn[m<br />

∣ nt] > δ + P ∣Lηn[m nt] − L t > δ<br />

≤ δ −1 E (∣ ∣ Y<br />

n<br />

t − L ηn[m<br />

∣ ) nt] + P (∣ ∣ )<br />

∣ Lηn[mnt] − L t > δ .<br />

42


Now, Corollary 2.3.7 and Proposition 2.3.8 imply respectively that<br />

E (∣ ∣Yt n − L ∣ ) η n[m nt] −→<br />

which compl<strong>et</strong>es the proof.<br />

n→∞<br />

0,<br />

P (∣ ∣L ηn[m nt] − L t<br />

∣ ∣ > δ ) −→<br />

n→∞<br />

0,<br />

Corollary 2.3.12. For every t ≥ 0, the law of Y n t under P(· | X ηn<br />

0 = n) converges weakly<br />

to the law of L t under P as n → ∞.<br />

Proof : For positive integers n and k, we denote by µ n the offspring distribution of the<br />

Galton-Watson process X ηn , by f n the generating function of µ n and by fk n the k-th iterate<br />

fk n = fn ◦ ... ◦ f n of f n . L<strong>et</strong> λ > 0 and t ≥ 0. We have,<br />

∞<br />

E(exp(−λYt n )) = ∑ (<br />

e −nnp f[m n<br />

p! nt]<br />

(e −λ/n)) p ( ( (<br />

= exp −n 1 − f[m n<br />

nt]<br />

e −λ/n))) .<br />

p=0<br />

From Proposition 2.3.11, it holds that<br />

( ( (<br />

exp −n 1 − f[m n<br />

nt]<br />

e −λ/n))) −→ E(exp(−λL t )).<br />

n→∞<br />

L<strong>et</strong> us s<strong>et</strong> u(t,λ) = − log(E[exp(−λL t )]). It follows that,<br />

( (<br />

n 1 − f[m n<br />

nt]<br />

e −λ/n)) −→ u(t,λ).<br />

n→∞<br />

Thus, we obtain,<br />

(<br />

)<br />

E exp(−λYt n ) ∣ X ηn<br />

0 = n =<br />

( (<br />

f[m n<br />

nt] e −λ/n)) n<br />

−→ exp(−u(t,λ)) = E[exp(−λL t )].<br />

n→∞<br />

At this point, we can use Theorem 2.2.1 to assert that (L t ,t ≥ 0) is a CSBP and that the law<br />

of (Yt n ,t ≥ 0) under the probability measure P(· | X ηn<br />

0 = n) converges to the law of (L t ,t ≥ 0)<br />

as n → ∞ in the space of probability measures on the Skorokhod space D(R + ). To verify the<br />

assumptions of Theorem 2.2.1, we need to check that there exists δ > 0 such that P(L δ > 0) > 0.<br />

This is obvious from Lemma 2.3.9.<br />

2.3.2. Identification of the measure Θ. In the previous section, we have constructed<br />

from Θ a CSBP L, which becomes extinct almost surely. We denote by ψ the associated branching<br />

mechanism. We can consider the σ-finite measure Θ ψ , which is the law of the Lévy tree associated<br />

with L. Our goal is to show that the measures Θ and Θ ψ coincide.<br />

Recall that µ n denotes the offspring distribution of the Galton-Watson process X ηn .<br />

Lemma 2.3.13. For every a > 0, the law of the R-tree η n T θ under Π µ n(· | H(θ) ≥ [am n ])<br />

converges as n → ∞ to the probability measure Θ ψ (· | H(T ) > a) in the sense of weak convergence<br />

of measures in the space T.<br />

Proof : We first check that, for every δ > 0,<br />

(2.3.13) lim inf<br />

n→∞ P(Yn δ = 0) > 0.<br />

Indeed, we have<br />

) (<br />

)<br />

P(Yδ (N(H(T n = 0) = P ) > ([m n δ] + 1)η n ) = 0 = exp − v(([m n δ] + 1)η n ) .<br />

□<br />

□<br />

43


As v is continuous, it follows that P(Yδ n = 0) → exp(−v(δ)) as n → ∞ implying (2.3.13).<br />

We recall that the law of Y n under the probability measure P(· | X ηn<br />

0 = n) converges to the<br />

law of (L t ,t ≥ 0). Then, thanks to (2.3.13), we can apply Theorem 2.2.3 to g<strong>et</strong> that, for every<br />

a > 0, the law of the R-tree m −1<br />

n T θ under Π µ n(· | H(θ) ≥ [am n ]) converges to the probability<br />

measure Θ ψ (· | H(T ) > a) in the sense of weak convergence of measures in the space T. As<br />

m −1<br />

n η n → 1 as n → ∞, we g<strong>et</strong> the desired result.<br />

□<br />

We can now compl<strong>et</strong>e the proof of Theorem 2.1.1. Indeed, thanks to Lemmas 2.2.2 and 2.3.3,<br />

we can construct on the same probability space (Ω,P), a sequence of T-valued random variables<br />

(T n ) n≥1 distributed according to Θ(· | H(T ) > ([am n ] + 1)η n ) and a sequence of A-valued<br />

random variables (θ n ) n≥1 distributed according to Π µ n(· | H(θ) ≥ [am n ]) such that for every<br />

n ≥ 1, P a.s., )<br />

GH<br />

(T n ,η n T θn ≤ 4η n .<br />

Then, using Lemma 2.3.13, we have Θ(· | H(T ) > ([am n ]+1)η n ) → Θ ψ (· | H(T ) > a) as n → ∞<br />

in the sense of weak convergence of measures on the space T. So we g<strong>et</strong><br />

for every a > 0, and thus Θ = Θ ψ .<br />

Θ(· |H(T ) > a) = Θ ψ (· |H(T ) > a)<br />

2.4. Proof of Theorem 2.1.2<br />

L<strong>et</strong> Θ be a probability measure on (T,GH) satisfying the assumptions of Theorem 2.1.2.<br />

In this case, we define v : [0, ∞) −→ (0, ∞) by v(t) = Θ(H(T ) > t) for every t ≥ 0. Note<br />

that v(0) = 1 is well defined here. For every t > 0, we denote by Θ t the probability measure<br />

Θ(· | H(T ) > t). The following two results are proved in a similar way to Lemma 2.3.1 and<br />

Lemma 2.3.2.<br />

Lemma 2.4.1. The function v is nonincreasing, continuous and goes to 0 as t → ∞.<br />

Lemma 2.4.2. For every t > 0 and 0 < a < b, the conditional law of the random variable<br />

Z(t,t+b), under the probability measure Θ t and given Z(t,t+a), is a binomial distribution with<br />

param<strong>et</strong>ers Z(t,t + a) and v(b)/v(a).<br />

2.4.1. The DSBP derived from Θ. We will follow the same strategy as in section 2.3<br />

but instead of a CSBP we will now construct an integer-valued branching process.<br />

2.4.1.1. A family of Galton-Watson trees. We recall that µ ε denotes the law of Z(ε,2ε) under<br />

the probability measure Θ ε , and that (θ ξ ,ξ ∈ A) is a sequence of independent A-valued random<br />

variables defined on a probability space (Ω ′ , P ′ ) such that for every ξ ∈ A, θ ξ is distributed<br />

uniformly overÔ−1 (ξ). The following lemma is proved in the same way as Lemma 2.3.3.<br />

Lemma 2.4.3. L<strong>et</strong> us define for every ε > 0, a mapping θ (ε) from T (ε) × Ω ′ into A by<br />

θ (ε) (T ,ω) = θ ξ ε (T )(ω).<br />

Then for every positive integer p, the law of the random variable θ (ε) under the probability<br />

measure Θ pε ⊗ P ′ is Π µε (· | H(θ) ≥ p − 1).<br />

For every ε > 0, we define a process X ε = (Xk ε ,k ≥ 0) on T by the formula<br />

= Z(kε,(k + 1)ε), k ≥ 0.<br />

X ε k<br />

We show in the same way as Proposition 2.3.4 and Proposition 2.3.5 the following two results.<br />

44


Proposition 2.4.4. For every ε > 0, the process X ε is under Θ a Galton-Watson process<br />

whose initial distribution is the Bernoulli distribution with param<strong>et</strong>er v(ε) and whose offspring<br />

distribution is µ ε .<br />

Proposition 2.4.5. For every t > 0 and h > 0, we have Θ(Z(t,t + h)) ≤ v(h) ≤ 1.<br />

The next proposition however is particular to the finite case and will be useful in the rest of<br />

this section.<br />

Proposition 2.4.6. The family of probability measures (µ ε ) ε>0 converges to the Dirac measure<br />

δ 1 as ε → 0. In other words,<br />

Proof : We first note that<br />

Θ ε (Z(ε,2ε) = 1) −→<br />

ε→0<br />

1.<br />

2Θ ε (Z(ε,2ε) ≥ 1) − Θ ε (Z(ε,2ε)) ≤ Θ ε (Z(ε,2ε) = 1) ≤ Θ ε (Z(ε,2ε) ≥ 1).<br />

Moreover,<br />

Θ ε (Z(ε,2ε) ≥ 1) = Θ ε (H(T ) > 2ε) = v(2ε)/v(ε)<br />

and Θ ε (Z(ε,2ε)) ≤ 1. So,<br />

(2.4.1)<br />

2v(2ε)<br />

v(ε)<br />

− 1 ≤ Θ ε (Z(ε,2ε) = 1) ≤ v(2ε)<br />

v(ε) .<br />

We l<strong>et</strong> ε → 0 in (2.4.1) and we use Lemma 2.4.1 to obtain the desired result.<br />

2.4.1.2. Construction of the DSBP.<br />

Proposition 2.4.7. For every t ≥ 0, there exists an integer-valued random variable L t on<br />

the space T such that Θ(L t ) ≤ 1 and Θ a.s.,<br />

Z(t,t + h) ↑ L t .<br />

h↓0<br />

Proof : L<strong>et</strong> t ≥ 0. The function h ∈ (0, ∞) ↦−→ Z(t,t + h) ∈ Z + is nonincreasing so that<br />

there exists a random variable L t with values in Z + ∪ {∞} such that, Θ a.s.,<br />

Z(t,t + h) ↑ L t .<br />

h↓0<br />

Thanks to the monotone convergence theorem, we have<br />

Θ(Z(t,t + h)) −→<br />

h↓0<br />

Θ(L t ).<br />

Now, by Proposition 2.4.5, Θ(Z(t,t + h)) ≤ 1 for every h > 0. Then, Θ(L t ) ≤ 1 which implies<br />

in particular that L t < ∞ Θ a.s.<br />

□<br />

Proposition 2.4.8. For every t > 0, the following two convergences hold Θ a.s.,<br />

(2.4.2)<br />

(2.4.3)<br />

Z(t − h,t)<br />

Z(t − h,t + h)<br />

↑<br />

h↓0<br />

↑<br />

h↓0<br />

L t ,<br />

L t .<br />

□<br />

45


Proof : L<strong>et</strong> t > 0 be fixed throughout this proof. By the same arguments as in the proof<br />

of Proposition 2.4.7, we can find a Z + -valued random variable L t such that Θ(L t ) ≤ 1 and<br />

Z(t − h,t) ↑ L t as h ↓ 0, Θ a.s. If h ∈ (0,t), we write T 1 ,...,T Z(t−h,t) for the subtrees of T<br />

above level t − h with height greater than h. Then, from the regenerative property,<br />

Θ (|Z(t,t + h) − Z(t − h,t)| ≥ 1)<br />

( (∣ ∣ ∣∣∣∣ Z(t−h,t)<br />

))<br />

∑<br />

∣∣∣∣<br />

= Θ Θ (Z(h,2h)(T i ) − 1)<br />

∣ ≥ 1 Z(t − h,t)<br />

i=1<br />

( (<br />

))<br />

≤ Θ Θ |Z(h,2h)(T i ) − 1| ≥ 1 for some i ∈ {1,... ,Z(t − h,t)} | Z(t − h,t)<br />

(<br />

)<br />

(2.4.4) ≤ Θ Z(t − h,t)Θ h (|Z(h,2h) − 1| ≥ 1) .<br />

Since Z(t−h,t)Θ h (|Z(h,2h) − 1| ≥ 1) ≤ L t Θ a.s., Proposition 2.4.6 and the dominated convergence<br />

theorem imply that the right-hand side of (2.4.4) goes to 0 as h → 0. Thus L t = L t Θ<br />

a.s.<br />

Likewise, there exists a random variable ̂L t with values in Z + such that, Θ a.s., Z(t−h,t+h) ↑<br />

̂L t as h ↓ 0. L<strong>et</strong> us now notice that, for every h > 0, Θ a.s.,<br />

Z(t − h,t + h) ≤ Z(t − h,t),<br />

implying that ̂L t ≤ L t , Θ a.s. Moreover, thanks to Lemma 2.4.2, we have<br />

( (v(2h) ) ) ( Z(t−h,t) (v(2h) ) )<br />

Lt<br />

(2.4.5) Θ(Z(t−h,t) ≥ Z(t−h,t+h)+1) = 1−Θ<br />

≤ 1−Θ<br />

.<br />

v(h)<br />

v(h)<br />

The right-hand side of (2.4.5) tends to 0 as h → 0. So L t = ̂L t Θ a.s.<br />

We will now establish a regularity property of the process (L t ,t ≥ 0).<br />

Proposition 2.4.9. The process (L t ,t ≥ 0) admits a modification which is right-continuous<br />

with left limits, and which has no fixed discontinuities.<br />

Proof : We start the proof with three lemmas. The first one in proved in a similar but easier<br />

way as Lemma 2.3.9.<br />

Lemma 2.4.10. There exists λ ≥ 0 such that Θ(L t ) = e −λt for every t ≥ 0.<br />

For every n ≥ 1 and every t ≥ 0 we s<strong>et</strong> Y n<br />

t = X 1/n<br />

[nt] .<br />

Lemma 2.4.11. For every t ≥ 0, Y n<br />

t<br />

→ L t as n → ∞, Θ a.s.<br />

This lemma is an immediate consequence of Proposition 2.4.8.<br />

Lemma 2.4.12. L<strong>et</strong> us define G t = σ(L s ,s ≤ t) for every t ≥ 0. Then (L t ,t ≥ 0) is a<br />

nonnegative supermartingale with respect to the filtration (G t ,t ≥ 0).<br />

Proof : L<strong>et</strong> s,t,s 1 ,...,s p ≥ 0 such that 0 ≤ s 1 ≤ ... ≤ s p ≤ s < t and l<strong>et</strong> f : R p → R +<br />

be a bounded measurable function. For every n ≥ 1, the offspring distribution µ 1/n is critical or<br />

subcritical so that (X 1/n ,k ≥ 0) is a supermartingale. Thus we have<br />

Θ<br />

k<br />

(<br />

X 1/n<br />

[nt] f (<br />

X 1/n<br />

[ns 1 ] ,...,X1/n [ns p]<br />

))<br />

≤ Θ<br />

( (<br />

X 1/n<br />

[ns] f X 1/n<br />

[ns 1 ],... ,X1/n<br />

[ns p]<br />

))<br />

.<br />

□<br />

46


Since f is bounded and Xu<br />

1/n ≤ L u Θ a.s. for every u ≥ 0, Lemma 2.4.11 yields<br />

L<strong>et</strong> us s<strong>et</strong>, for every t ≥ 0,<br />

Θ ( L t f ( L s1 ,...,L sp<br />

))<br />

≤ Θ<br />

(<br />

Ls f ( L s1 ,...,L sp<br />

))<br />

.<br />

˜G t = ⋂ s>tG s .<br />

□<br />

Recall that D denotes the s<strong>et</strong> of positive dyadic numbers. From Lemma 2.4.12 and classical<br />

results on supermartingales, we can define a right-continuous supermartingale (˜L t ,t ≥ 0) with<br />

respect to the filtration ( ˜G t ,t ≥ 0) by s<strong>et</strong>ting, for every t ≥ 0,<br />

(2.4.6) ˜Lt = lim<br />

s↓t,s∈D L s<br />

where the limit holds Θ a.s. and in L 1 . In a way similar to Section 2.3 we can prove that<br />

(˜L t ,t ≥ 0) is a càdlàg modification of (L t ,t ≥ 0) with no fixed discontinuities.<br />

From now on, to simplify notation we replace (L t ,t ≥ 0) by its càdlàg modification (˜L t ,t ≥ 0).<br />

Proposition 2.4.13. (L t ,t ≥ 0) is a DSBP which becomes extinct Θ a.s.<br />

Proof : By the same arguments as in the proof of (2.4.2), we can prove that, for every<br />

0 < s < t, the following convergence holds in probability under Θ,<br />

( )<br />

[nt] − [ns] [nt] − [ns] + 1<br />

(2.4.7) Z , −→<br />

n n<br />

L t−s.<br />

n→∞<br />

L<strong>et</strong> s,t,s 1 ,... ,s p ≥ 0 such that 0 ≤ s 1 ≤ ... ≤ s p ≤ s < t, λ > 0 and l<strong>et</strong> f : R p −→ R be a<br />

bounded measurable function. For every n ≥ 1, under Θ 1/n , (X 1/n<br />

k<br />

,k ≥ 0) is a Galton-Watson<br />

process started at one so that<br />

( ) ( ( ( ( ))) )<br />

Θ 1/n f(Ys n<br />

1<br />

,...,Ys n<br />

p<br />

)exp (−λYt n ) = Θ 1/n f(Ys n<br />

1<br />

,... ,Ys n<br />

p<br />

) Θ 1/n exp −λX 1/n Y n<br />

s<br />

[nt]−[ns]<br />

.<br />

From Lemma 2.4.11, (2.4.7) and dominated convergence, we g<strong>et</strong><br />

Θ ( f(L s1 ,...,L sp )exp(−λL t ) ) = Θ ( f(L s1 ,... ,L sp )(Θ(exp(−λL t−s ))) Ls) .<br />

Then, (L t ,t ≥ 0) is a continuous-time Markov chain with values in Z + satisfying the branching<br />

property. Furthermore, since H(T ) < ∞ Θ a.s., it is immediate that (L t ,t ≥ 0) becomes extinct<br />

Θ a.s.<br />

□<br />

2.4.2. Identification of the probability measure Θ. L<strong>et</strong> us now define, for every T ∈ T<br />

and t ≥ 0,<br />

N t (T ) = #{σ ∈ T : d(ρ,σ) = t},<br />

where we recall that ρ denotes the root of T .<br />

Proposition 2.4.14. For every t ≥ 0, N t = L t Θ a.s.<br />

Note that for every t ≥ 0, L t is the number of subtrees of T above level t.<br />

Proof : Since Θ(H(T ) = 0) = 0, we have L 0 = 1 = N 0 Θ a.s. Thanks to Proposition 2.4.7,<br />

Proposition 2.4.8 and Proposition 2.4.9, for every t > 0, Θ a.s., there exists h 0 > 0 such that for<br />

every h ∈ (0,h 0 ],<br />

L t = L t−h = L t+h = Z(t − h,t + h).<br />

47


The remaining part of the argument is d<strong>et</strong>erministic. We fix t,h 0 > 0 and a (d<strong>et</strong>erministic)<br />

tree T ∈ T. We assume that there is a positive integer p such that for every h ∈ (0,h 0 ],<br />

L t = L t−h = L t+h = Z(t − h,t + h) = p,<br />

and we will verify that N t (T ) = p. We denote by T 1 ,...,T p the p subtrees of T above level t−h 0<br />

and we write ρ i for the root of the subtree T i . For every i ∈ {1,... ,p}, we have H(T i ) > 2h 0<br />

so that there exists x i ∈ T i such that d(ρ i ,x i ) = 2h 0 . L<strong>et</strong> us prove that for every i ∈ {1,... ,p},<br />

(2.4.8) T i ≤2h 0<br />

= { σ ∈ T i : d(ρ i ,σ) ≤ 2h 0<br />

}<br />

= [[ρi ,x i ]].<br />

To this end, we argue by contradiction and assume that we can find i ∈ {1,... ,p} and t i ∈ T i<br />

such that d(ρ i ,t i ) ≤ 2h 0 and t i /∈ [[ρ i ,x i ]]. L<strong>et</strong> z i be the unique vertex of T i satisfying<br />

[[ρ i ,z i ]] = [[ρ i ,x i ]] ∩ [[ρ i ,t i ]].<br />

We choose c > 0 such that d(ρ i ,z i ) < c < d(ρ i ,t i ). Then it is not difficult to see that T has<br />

at least p + 1 subtrees above level t − h 0 + c. This is a contradiction since L t−h0 +c = p. So<br />

N t (T ) = p, which compl<strong>et</strong>es the proof.<br />

□<br />

Proposition 2.4.14 means that (L t ,t ≥ 0) is a modification of the process (N t ,t ≥ 0) which<br />

describes the evolution of the number of individuals in the tree. L<strong>et</strong> us denote by Q the generator<br />

of (L t ,t ≥ 0) which is of the form<br />

⎛<br />

Q =<br />

⎜<br />

⎝<br />

⎞<br />

0 0 0 0 0 ...<br />

aγ(0) −a aγ(2) aγ(3) aγ(4) ...<br />

0 2aγ(0) −2a 2aγ(2) aγ(3) ...<br />

0 0 3aγ(0) −3a 3aγ(2) ... ⎟<br />

⎠<br />

. .. . .. . .. . ..<br />

.<br />

.<br />

where a > 0 and γ is a critical or subcritical offspring distribution with γ(1) = 0.<br />

For every t ≥ 0 we l<strong>et</strong> F t be the σ-field on T generated by the mapping T ↦−→ T ≤t and<br />

compl<strong>et</strong>ed with respect to Θ. Thus (F t ,t ≥ 0) is a filtration on T.<br />

Lemma 2.4.15. L<strong>et</strong> t > 0 and p ∈ N. Under Θ, conditionally on F t and given {L t = p}, the<br />

p subtrees of T above level t are independent and distributed according to Θ.<br />

Proof : Thanks to Lemma 2.2.2 and Lemma 2.4.3, we can construct on the same probability<br />

space (Ω,P), a sequence of T-valued random variables (T n ) n≥1 distributed according to Θ 1/n<br />

and a sequence of A-valued random variables (θ n ) n≥1 distributed according to Π µ1/n such that,<br />

for every n ≥ 1,<br />

)<br />

(2.4.9) GH<br />

(T n ,n −1 T θn ≤ 4n −1 .<br />

For every n ≥ 1 and k ≥ 0, we define X n k = #{u ∈ θ n : |u| = k}. L<strong>et</strong> t ≥ 0 and p ≥ 1, l<strong>et</strong><br />

g : T −→ R be a bounded continuous function and l<strong>et</strong> G : T p −→ R be a bounded continuous<br />

symm<strong>et</strong>ric function. For n ≥ 1, on the event {X n [nt]<br />

= p}, we s<strong>et</strong> {u n 1 ,... ,un p } = {u ∈ θ n :<br />

|u| = [nt]} and θ i n = τ u n<br />

i<br />

θ n for every i ∈ {1,... ,p}. Then we can write, thanks to the branching<br />

property of Galton-Watson trees,<br />

( ) (<br />

)<br />

E g n<br />

(½nX )<br />

−1 T θn<br />

n<br />

[nt] =po ≤[nt]<br />

G n −1 T θ1 n,... ,n −1 T θp n<br />

( ) ) ) ⊗p (<br />

))<br />

(2.4.10) = g n E(½nX −1 T θn<br />

n<br />

[nt] =po ≤[nt]<br />

(Π µ1/n G<br />

(n −1 T θ 1<br />

,... ,n −1 T θp ,<br />

,<br />

48


where θ 1 ,... ,θ p denote the coordinate variables under the product measure (Π µ1/n ) ⊗p . As a<br />

consequence of (2.4.9), we see that the law of n −1 T θ under Π µ1/n converges to Θ in the sense of<br />

weak convergence of measures on the space T. Then, thanks to Lemma 2.4.11, the right-hand<br />

side of (2.4.10) converges as n → ∞ to<br />

Θ (½{L t=p}g(T ≤t ) ) Θ ⊗p (G(T 1 ,... , T p )).<br />

Similarly, the left-hand side of (2.4.10) converges as n → ∞ to<br />

Θ (½{L t=p}g(T ≤t )G(T 1 ,... , T p ) ) ,<br />

where T 1 ,...,T p are the p subtrees of T above level t on the event {L t = p}. This compl<strong>et</strong>es<br />

the proof.<br />

□<br />

L<strong>et</strong> us define J = inf{t ≥ 0 : L t ≠ 1}. Then J is an (F t ) t≥0 -stopping time.<br />

Lemma 2.4.16. L<strong>et</strong> p ∈ N. Under Θ, given {L J = p}, the p subtrees of T above level J are<br />

independent and distributed according to Θ, and are independent of J.<br />

Proof : L<strong>et</strong> p ∈ N, l<strong>et</strong> f : R + −→ R be a bounded continuous function and l<strong>et</strong> G : T p −→ R<br />

be a bounded continuous symm<strong>et</strong>ric function. On the event {L J = p}, we denote by T 1 ,...,T p<br />

the p subtrees of T above level J. L<strong>et</strong> n ≥ 1 and k ≥ 0. On the event {L (k+1)/n = p}, we<br />

denote by T 1,(n,k) ,... , T p,(n,k) the p subtrees of T above level (k + 1)/n. On the one hand, the<br />

right-continuity of the mapping t ↦−→ L t gives<br />

Θ<br />

( ∞<br />

∑<br />

k=1½{L (k+1)/n =p} G (<br />

T 1,(n,k) ,... , T p,(n,k)) f ((k + 1)/n)½{k/n


CHAPITRE 3<br />

Conditioned Brownian trees<br />

3.1. Introduction<br />

In this work, we define and study a continuous tree of one-dimensional Brownian paths<br />

started from the origin, which is conditioned to remain in the positive half-line. An important<br />

motivation for introducing this object comes from its relation with analogous discr<strong>et</strong>e models<br />

which are discussed in several recent papers.<br />

In order to present our main results, l<strong>et</strong> us briefly describe a construction of unconditioned<br />

Brownian trees. We start from a positive Brownian excursion conditioned to have duration 1 (a<br />

normalized Brownian excursion in short), which is denoted by (e(s),0 ≤ s ≤ 1). This random<br />

function can be viewed as coding a continuous tree via the following simple prescriptions. For<br />

every s,s ′ ∈ [0,1], we s<strong>et</strong><br />

m e (s,s ′ ) := inf e(r).<br />

s∧s ′ ≤r≤s∨s ′<br />

We then define an equivalence relation on [0,1] by s<strong>et</strong>ting s ∼ s ′ if and only if e(s) = e(s ′ ) =<br />

m e (s,s ′ ). Finally we put<br />

d e (s,s ′ ) = e(s) + e(s ′ ) − 2m e (s,s ′ )<br />

and note that d e (s,s ′ ) only depends on the equivalence classes of s and s ′ . Then the quotient<br />

space T e := [0,1]/ ∼ equipped with the m<strong>et</strong>ric d e is a compact R-tree (see e.g. Section 2 of<br />

[22]). In other words, it is a compact m<strong>et</strong>ric space such that for any two points σ and σ ′ there is<br />

a unique arc with endpoints σ and σ ′ and furthermore this arc is isom<strong>et</strong>ric to a compact interval<br />

of the real line. We view T e as a rooted R-tree, whose root ρ is the equivalence class of 0. For<br />

every σ ∈ T e , the ancestral line of σ is the line segment joining ρ to σ. This line segment is<br />

denoted by [[ρ,σ]]. We write ṡ for the equivalence class of s, which is a vertex in T e at generation<br />

e(s) = d e (0,s).<br />

Up to unimportant scaling constants, T e is the Continuum Random Tree (CRT) introduced<br />

by Aldous [3]. The preceding presentation is indeed a reformulation of Corollary 22 in [5], which<br />

was proved via a discr<strong>et</strong>e approximation (a more direct approach was given in [35]). As Aldous<br />

[5] has shown, the CRT is the scaling limit of critical Galton-Watson trees conditioned to have<br />

a large fixed progeny (see [20] and [22] for recent generalizations of Aldous’ result). The fact<br />

that Brownian excursions can be used to model continuous genealogies had been used before, in<br />

particular in the Brownian snake approach to superprocesses (see [34]).<br />

We can now combine the branching structure of the CRT with independent spatial motions.<br />

We restrict ourselves to spatial displacements given by linear Brownian motions, which is the<br />

case of interest in this work. Conditionally given e, we introduce a centered Gaussian process<br />

(V σ ,σ ∈ T e ) with covariance<br />

cov(Vṡ,Vṡ′) = m e (s,s ′ ) , s,s ′ ∈ [0,1].<br />

51


This definition should become clear if we observe that m e (s,s ′ ) is the generation of the most<br />

recent common ancestor to ṡ and ṡ ′ in the tree T e . It is easy to verify that the process (V σ ,σ ∈<br />

T e ) has a continuous modification. The random measure Z on R defined by<br />

〈Z,ϕ〉 =<br />

∫ 1<br />

0<br />

ϕ(Vṡ)ds<br />

is then the one-dimensional Integrated Super-Brownian Excursion (ISE, see Aldous [6]). Note<br />

that ISE in higher dimensions, and related Brownian trees, have appeared recently in various<br />

asymptotic results for statistical mechanics models (see e.g. [18],[28],[29]). The support, or<br />

range, of ISE is<br />

R := {V σ : σ ∈ T e }.<br />

For our purposes, it is also convenient to reinterpr<strong>et</strong> the preceding notions in terms of the<br />

Brownian snake. The Brownian snake (W s ,0 ≤ s ≤ 1) driven by the normalized excursion e is<br />

obtained as follows (see subsection 3.2.1 for a more d<strong>et</strong>ailed presentation). For every s ∈ [0,1],<br />

W s = (W s (t),0 ≤ t ≤ e(s)) is the finite path which gives the spatial positions along the ancestral<br />

line of ṡ : W s (t) = V σ if σ is the vertex at distance t from the root on the segment [[ρ,ṡ]]. Note<br />

that W s only depends on the equivalent class ṡ. We view W s as a random element of the space<br />

W of finite paths.<br />

Our first goal is to give a precise definition of the Brownian tree (V σ ,σ ∈ T e ) conditioned to<br />

remain positive. Equivalently this amounts to conditioning ISE to put no mass on the negative<br />

half-line. Our first theorem gives a precise meaning to this conditioning in terms of the Brownian<br />

snake. We denote by N (1)<br />

0 the distribution of (W s ,0 ≤ s ≤ 1) on the canonical space C([0,1], W)<br />

of continuous functions from [0,1] into W, and we abuse notation by still writing (W s ,0 ≤ s ≤ 1)<br />

for the canonical process on this space. The range R is then defined under N (1)<br />

0 by<br />

}<br />

R =<br />

{Ŵs : 0 ≤ s ≤ 1<br />

where Ŵs denotes the endpoint of the path W s .<br />

Theorem 3.1.1. We have<br />

lim ε −4 N (1)<br />

2<br />

0<br />

(R ⊂] − ε, ∞[) =<br />

ε↓0 21 .<br />

There exists a probability measure on C([0,1], W), which is denoted by N (1)<br />

0 , such that<br />

lim N (1)<br />

0 ε↓0<br />

(· | R ⊂] − ε, ∞[) = N(1) 0 ,<br />

in the sense of weak convergence in the space of probability measures on C([0,1], W).<br />

Our second theorem gives an explicit representation of the conditioned measures N (1)<br />

0 , which<br />

is analogous to a famous theorem of Vervaat [52] relating the normalized Brownian excursion<br />

to the Brownian bridge. To state this result, we need the notion of re-rooting. For s ∈ [0,1], we<br />

write T [s]<br />

e for the “same” tree T e but with root ṡ instead of ρ = ˙0. We then shift the spatial<br />

positions by s<strong>et</strong>ting V σ<br />

[s] = V σ − Vṡ for every σ ∈ T e , in such a way that the spatial position of<br />

the new root is still the origin. (Notice that both T e<br />

[s] and V [s] only depend on ṡ, and we could<br />

as well define T [σ]<br />

e and V [σ] for σ ∈ T e .) Finally, the re-rooted snake W [s] = (W r [s] ,0 ≤ r ≤ 1) is<br />

defined analogously as before : For every r ∈ [0,1], W r<br />

[s] is the path giving the spatial positions<br />

V σ [s] along the ancestral line (in the re-rooted tree) of the vertex s + r mod. 1.<br />

52


Theorem 3.1.2. L<strong>et</strong> s ∗ be the unique time of the minimum of Ŵ on [0,1]. The probability<br />

measure N (1)<br />

0 is the law under N (1)<br />

0 of the re-rooted snake W [s∗] .<br />

If we want to define one-dimensional ISE conditioned to put no mass on the negative half-line,<br />

the most natural way is to condition to put no mass on ] − ∞, −ε[ and then to l<strong>et</strong> ε go to 0.<br />

As a consequence of the previous two theorems, this is equivalent to shifting the unconditioned<br />

ISE to the right, so that the left-most point of its support becomes the origin. Another m<strong>et</strong>hod<br />

would be to condition the mass in ]−∞,0] to be less than ε and then to l<strong>et</strong> ε go to 0. Proposition<br />

3.3.7 below shows that this leads to the same measure N (1)<br />

0 .<br />

Both Theorem 3.1.1 and Theorem 3.1.2 could be presented in a different and perhaps more<br />

elegant manner by using the formalism of spatial trees as in Section 5 of [22]. In this formalism,<br />

a spatial tree is a pair (T,U) where T is a compact rooted R-tree (in fact an equivalent class<br />

of such objects modulo root-preserving isom<strong>et</strong>ries) and U is a continuous mapping from T into<br />

R d . Then the second assertion of Theorem 3.1.1 can be rephrased by saying that the conditional<br />

distribution of the spatial tree (T e ,V ) knowing that R ⊂] − ε, ∞[ has a limit when ε goes to 0,<br />

and Theorem 3.1.2 says that this limit is the distribution of (T [σ∗]<br />

e ,V [σ∗] ) where σ ∗ is the unique<br />

vertex minimizing V . We have chosen the above presentation because the Brownian snake plays<br />

a fundamental role in our proofs and also because the resulting statements are stronger than<br />

the ones in terms of spatial trees.<br />

L<strong>et</strong> us discuss the relationship of the above theorems with previous results. The first assertion<br />

of Theorem 3.1.1 is closely related to some estimates of Abraham and Werner [1]. In particular,<br />

Abraham and Werner proved that the probability for a Brownian snake driven by a Brownian<br />

excursion of height 1 not to hit the s<strong>et</strong> ]−∞, −ε[ behaves like a constant times ε 4 (see Section 3.4<br />

below). The d-dimensional Brownian snake conditioned not to exit a domain D was studied by<br />

Abraham and Serl<strong>et</strong> [2], who observed that this conditioning gives rise to a particular instance<br />

of the Brownian snake with drift. The s<strong>et</strong>ting in [2] is different from the present work, in that<br />

the initial point of the snake lies inside the domain, and not at its boundary as here. We also<br />

mention the paper [32] by Jansons and Rogers, who establish a decomposition at the minimum<br />

for a Brownian tree where branchings occur only at discr<strong>et</strong>e times.<br />

An important motivation for the present work came from several recent papers that discuss<br />

asymptotics for planar maps. Cori and Vauquelin [15] proved that there exists a bijection b<strong>et</strong>ween<br />

rooted planar quadrangulations and certain discr<strong>et</strong>e trees called well-labelled trees (see<br />

also Chassaing & Schaeffer [14] for a more tractable description of this bijection). Roughly, a<br />

well-labelled tree consists of a (discr<strong>et</strong>e) plane tree whose vertices are given labels which are<br />

positive integers, with the constraints that the label of the root is 1 and the labels of two neighboring<br />

vertices can differ by at most 1. Our conditioned Brownian snake should then be viewed<br />

as a continuous model for well-labelled trees. This idea was exploited in [14] and especially in<br />

Marckert and Mokkadem [45], where the re-rooted snake W [s∗] appears in the description of the<br />

Brownian map, which is the continuous object describing scaling limits of planar quadrangulations.<br />

In contrast with the present work, the re-rooted snake W [s∗] is not interpr<strong>et</strong>ed in [45] as<br />

a conditioned object, but rather as a scaling limit of re-rooted discr<strong>et</strong>e snakes. Closely related<br />

models of discr<strong>et</strong>e labelled trees are also of interest in theor<strong>et</strong>ical physics : See in particular [9]<br />

and [10]. The article [39], which was motivated by [14] and [45], proves that our conditioned<br />

Brownian tree is the scaling limit of discr<strong>et</strong>e spatial trees conditioned to remain positive. To be<br />

specific, consider a Galton-Watson tree whose offspring distribution is critical and has (small)<br />

exponential moments, and condition this tree to have exactly n vertices (in the special case of<br />

53


the geom<strong>et</strong>ric distribution, this gives rise to a tree that is uniformly distributed over the s<strong>et</strong> of<br />

plane trees with n vertices). This branching structure is combined with a spatial displacement<br />

which is a symm<strong>et</strong>ric random walk with bounded jump size on Z. Assuming that the root is at<br />

the origin of Z, the spatial tree is then conditioned to remain on the positive side. According to<br />

the main theorem of [39], the scaling limit of this conditioned discr<strong>et</strong>e tree when n → ∞ leads<br />

to the measure N (1)<br />

0 discussed above. The convergence here, and the precise form of the scaling<br />

transformation, are as in Theorem 2 of [31], which discusses scaling limits for unconditioned<br />

discr<strong>et</strong>e snakes.<br />

L<strong>et</strong> us now describe the other contributions of this paper. Although the preceding theorems<br />

have been stated for the measure N (1)<br />

0 , a more fundamental object is the excursion measure N 0 of<br />

the Brownian snake (see e.g. [37]). Roughly speaking, N 0 is obtained by the same construction as<br />

above, but instead of considering a normalized Brownian excursion, we now l<strong>et</strong> e be distributed<br />

according to the (infinite) Itô measure of Brownian excursions. If σ(e) denotes the duration<br />

of excursion e, we have N (1)<br />

0 = N 0 (· | σ = 1). It turns out that many calculations are more<br />

tractable under the infinite measure N 0 than under N (1)<br />

0 . For this reason, both Theorems 3.1.1<br />

and Theorem 3.1.2 are proved in Section 3.3 as consequences of Theorem 3.3.1, which deals<br />

with N 0 . Motivated by Theorem 3.3.1 we introduce another infinite measure denoted by N 0 ,<br />

which should be interpr<strong>et</strong>ed as N 0 conditioned on the event {R ⊂ [0, ∞[}, even though the<br />

conditioning requires some care as we are dealing with infinite measures. In the same way as<br />

for unconditioned measures, we have N (1)<br />

0 = N 0 (· | σ = 1). Another motivation for considering<br />

the measure N 0 comes from connections with superprocesses : Analogously to Chapter IV of<br />

[37] in the unconditioned case, N 0 could be used to define and to analyse a one-dimensional<br />

super-Brownian motion started from the Dirac measure δ 0 and conditioned never to charge the<br />

negative half-line.<br />

In Section 3.4, we present a different approach that leads to the same limiting measures. If<br />

H(e) stands for the height of excursion e, we consider for every h > 0 the measure N h 0 := N 0(· |<br />

H = h). In the above construction this amounts to replacing the normalized excursion e by a<br />

Brownian excursion with height h. By using a famous decomposition theorem of Williams, we<br />

can then analyse the behavior of the measure N h 0 conditioned on the event that the range does<br />

not intersect ] − ∞, −ε[ and show that it has a limit denoted by N h 0 when ε → 0. The m<strong>et</strong>hod<br />

also provides information about the Brownian tree under N h 0 : This Brownian tree consists of a<br />

spine whose distribution is absolutely continuous with respect to that of the nine-dimensional<br />

Bessel process, and as usual a Poisson collection of subtrees originating from the spine, which<br />

are Brownian snake excursions conditioned not to hit the negative half-line. The connection with<br />

the measures N (1)<br />

0 and N 0 is made by proving that N h 0 = N 0(· | H = h). Several arguments in<br />

this section have been inspired by Abraham and Werner’s paper [1]. It should also be noted that<br />

a discr<strong>et</strong>e version of the nine-dimensional Bessel process already appears in the paper [13] by<br />

Chassaing and Durhuus.<br />

At the end of Section 3.4, we also discuss the limiting behavior of the measures N h 0 as h → ∞.<br />

This leads to a probability measure N ∞ 0 that should be viewed as the law of an infinite Brownian<br />

snake excursion conditioned to stay positive. We again g<strong>et</strong> a description of the Brownian tree<br />

coded by N ∞ 0 in terms of a spine and conditioned Brownian snake excursions originating from<br />

this spine. Moreover, the description is simpler in the sense that the spine is exactly distributed<br />

as a nine-dimensional Bessel process started at the origin.<br />

54


Section 3.5 gives an explicit formula for the finite-dimensional marginal distributions of the<br />

Brownian tree under N 0 , that is for<br />

)<br />

N 0<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F(W s1 ,...,W sp )<br />

where p ≥ 1 is an integer and F is a symm<strong>et</strong>ric nonnegative measurable function on W p . In<br />

a way similar to the corresponding result for the unconditioned Brownian snake (see (3.2.1)<br />

below), this formula involves combining the branching structure of certain discr<strong>et</strong>e trees with<br />

spatial displacements. Here however because of the conditioning, the spatial displacements turn<br />

out to be given by nine-dimensional Bessel processes rather than linear Brownian motions. In<br />

the same way as the finite-dimensional marginal distributions of the CRT can be derived from<br />

the analogous formula under the Itô measure (see Chapter III of [37]), one might hope to derive<br />

the expression of the finite-dimensional marginals under N (1)<br />

0 from the case of N 0 . This idea<br />

apparently leads to untractable calculations, but we still expect Theorem 3.5.1 to have useful<br />

applications in future work about conditioned trees.<br />

Basic facts about the Brownian snake are recalled in Section 3.2, which also establishes a few<br />

important preliminary results, some of which are of independent interest. In particular, we state<br />

and prove a general version of the invariance property of N 0 under re-rooting (Theorem 3.2.3).<br />

This result is clearly related to the invariance of the CRT under uniform re-rooting, which was<br />

observed by Aldous [4] (and generalized to Lévy trees in Proposition 4.8 of [22]). An equivalent<br />

form of Theorem 3.2.3 already appears as Proposition 4.9 of [45] : See the discussion after the<br />

statement of this theorem in subsection 3.2.3.<br />

3.2. Preliminaries<br />

In this section, we recall the basic facts about the Brownian snake that we will use later,<br />

and we also establish a few important preliminary results. We refer to [37] for a more d<strong>et</strong>ailed<br />

presentation of the Brownian snake and its connections with partial differential equations. In<br />

the first four subsections below, we deal with the d-dimensional Brownian snake since the proofs<br />

are not more difficult in that case, and the results may have other applications.<br />

3.2.1. The Brownian snake. The (d-dimensional) Brownian snake is a Markov process<br />

taking values in the space W of finite paths in R d . Here a finite path is simply a continuous<br />

mapping w : [0,ζ] −→ R d , where ζ = ζ (w) is a nonnegative real number called the lif<strong>et</strong>ime of w.<br />

The s<strong>et</strong> W is a Polish space when equipped with the distance<br />

d(w,w ′ ) = |ζ (w) − ζ (w ′ )| + sup |w(t ∧ ζ (w) ) − w ′ (t ∧ ζ (w ′ ))|.<br />

t≥0<br />

The endpoint (or tip) of the path w is denoted by ŵ. The range of w is denoted by w[0,ζ (w) ].<br />

In this work, it will be convenient to use the canonical space Ω := C(R + , W) of continuous<br />

functions from R + into W, which is equipped with the topology of uniform convergence on every<br />

compact subs<strong>et</strong> of R + . The canonical process on Ω is then denoted by<br />

and we write ζ s = ζ (Ws) for the lif<strong>et</strong>ime of W s .<br />

W s (ω) = ω(s) , ω ∈ Ω ,<br />

L<strong>et</strong> w ∈ W. The law of the Brownian snake started from w is the probability measure P w<br />

on Ω which can be characterized as follows. First, the process (ζ s ) s≥0 is under P w a reflected<br />

55


Brownian motion in [0, ∞[ started from ζ (w) . Secondly, the conditional distribution of (W s ) s≥0<br />

knowing (ζ s ) s≥0 , which is denoted by Θ ζ w, is characterized by the following properties :<br />

(i) W 0 = w, Θ ζ w a.s.<br />

(ii) The process (W s ) s≥0 is time-inhomogeneous Markov under Θ ζ w. Moreover, if 0 ≤ s ≤ s ′ ,<br />

• W s ′(t) = W s (t) for every t ≤ m(s,s ′ ) := inf [s,s ′ ] ζ r , Θ ζ w a.s.<br />

• (W s ′(m(s,s ′ ) + t) − W s ′(m(s,s ′ ))) 0≤t≤ζs ′ −m(s,s ′ ) is independent of W s and distributed<br />

as a d-dimensional Brownian motion started at 0 under Θ ζ w.<br />

Informally, the value W s of the Brownian snake at time s is a random path with a random<br />

lif<strong>et</strong>ime ζ s evolving like reflecting Brownian motion in [0, ∞[. When ζ s decreases, the path is<br />

erased from its tip, and when ζ s increases, the path is extended by adding “little pieces” of<br />

Brownian paths at its tip.<br />

Excursion measures play a fundamental role throughout this work. We denote by n(de) the<br />

Itô measure of positive Brownian excursions. This is a σ-finite measure on the space C(R + , R + )<br />

of continuous functions from R + into R + . We write<br />

σ(e) = inf{s > 0 : e(s) = 0}<br />

for the duration of excursion e. For s > 0, n (s) will denote the conditioned measure n(· | σ = s).<br />

Our normalization of the excursion measure is fixed by the relation<br />

∫ ∞<br />

ds<br />

n =<br />

2 √ 2πs n (s).<br />

3<br />

0<br />

If x ∈ R d , the excursion measure N x of the Brownian snake from x is then defined by<br />

∫<br />

N x = n(de) Θx<br />

e<br />

C(R + ,R + )<br />

where x denotes the trivial element of W with lif<strong>et</strong>ime 0 and initial point x. Alternatively, we<br />

can view N x as the excursion measure of the Brownian snake from the regular point x. With a<br />

slight abuse of notation we will also write σ(ω) = inf{s > 0 : ζ s (ω) = 0} for ω ∈ Ω. We can then<br />

consider the conditioned measures<br />

∫<br />

N x (s) = N x(· | σ = s) = n (s) (de) Θx e .<br />

C(R + ,R + )<br />

Note that in contrast to the introduction we now view N (s)<br />

x as a measure on Ω rather than on<br />

C([0,s], W). The range R = R(ω) is defined by R = {Ŵs : s ≥ 0}.<br />

Lemma 3.2.1. Suppose that d = 1 and l<strong>et</strong> x > 0.<br />

(i) We have<br />

N x (R∩] − ∞,0] ≠ ∅) = 3<br />

2x 2.<br />

(ii) For every λ > 0,<br />

√<br />

N x<br />

(1 −½{R∩]−∞,0]=∅} e −λσ) λ<br />

(<br />

)<br />

= 3(coth(2 1/4 xλ 1/4 )) 2 − 2<br />

2<br />

where coth(y) = cosh(y)/sinh(y).<br />

Proof : (i) According to Section VI.1 of [37], the function u(x) = N x (R∩] − ∞,0] ≠ ∅)<br />

solves u ′′ = 4u 2 in ]0, ∞[, with boundary condition u(0+) = +∞. The desired result follows.<br />

(ii) See Lemma 7 in [17].<br />

56<br />


3.2.2. Finite-dimensional marginal distributions. In this subsection we state a result<br />

giving information about the joint distribution of the values of the Brownian snake at a finite<br />

number of times and its range. In order to state this result, we need some formalism for trees.<br />

We first introduce the s<strong>et</strong> of labels<br />

∞⋃<br />

U = {1,2} n<br />

n=0<br />

where by convention {1,2} 0 = {∅}. An element of U is thus a sequence u = u 1 ...u n of elements<br />

of {1,2}, and we s<strong>et</strong> |u| = n, so that |u| represents the “generation” of u. In particular, |∅| = 0.<br />

The mapping π : U\{∅} −→ U is defined by π(u 1 ... u n ) = u 1 ...u n−1 (π(u) is the “father” of<br />

u). In particular, if k = |u|, we have π k (u) = ∅.<br />

A binary (plane) tree T is a finite subs<strong>et</strong> of U such that :<br />

(i) ∅ ∈ T .<br />

(ii) u ∈ T \{∅} ⇒ π(u) ∈ T .<br />

(iii) For every u ∈ T , either u1 ∈ T and u2 ∈ T , or u1 /∈ T and u2 /∈ T (u is called a leaf<br />

in the second case).<br />

We denote by A the s<strong>et</strong> of all binary trees. A marked tree is then a pair (T ,(h u ) u∈T ) where<br />

T ∈ A and h u ≥ 0 for every u ∈ T . We denote by T the space of all marked trees. In this<br />

work it will be convenient to view marked trees as R-trees in the sense of [22] or [24] (see<br />

also Section 3.1 above). This can be achieved through the following explicit construction. L<strong>et</strong><br />

θ = (T ,(h u ) u∈T ) be a marked tree and l<strong>et</strong> R T be the vector space of all mappings from T into<br />

R. Write (ε u ,u ∈ T ) for the canonical basis of R T . Then consider the mapping<br />

p θ : ⋃<br />

{u} × [0,h u ] −→ R T<br />

defined by<br />

u∈T<br />

|u|<br />

∑<br />

p θ (u,l) = h π k (u) ε π k (u) + lε u .<br />

k=1<br />

As a s<strong>et</strong>, the R-tree associated with θ is the range ˜θ of p θ . Note that this is a connected union<br />

of line segments in R T . It is equipped with the distance d θ such that d θ (a,b) is the length of the<br />

shortest path in ˜θ going from a to b. By definition, the range of this path is the segment b<strong>et</strong>ween<br />

a and b and is denoted by [[a,b]]. Finally, we will write L θ for the (one-dimensional) Lebesgue<br />

measure on ˜θ.<br />

By definition, leaves of ˜θ are points of the form p θ (u,h u ) where u is leaf of θ. Points of the<br />

form p θ (u,h u ) when u is not a leaf are called nodes of ˜θ. We write L(θ) for the s<strong>et</strong> of leaves of<br />

˜θ, and I(θ) for the s<strong>et</strong> of its nodes. The root of ˜θ is just the point 0 = p θ (∅,0).<br />

We will consider Brownian motion indexed by ˜θ, with initial point x ∈ R d . Formally, we may<br />

consider, under the probability measure Q θ x , a collection (ξu ) u∈T of independent d-dimensional<br />

Brownian motions all started at 0 except ξ ∅ which starts at x, and define a continuous process<br />

(V a ,a ∈ ˜θ) by s<strong>et</strong>ting<br />

|u|<br />

∑ ( )<br />

V pθ (u,l) = ξ πk (u)<br />

h π k (u) + ξ u (l),<br />

k=1<br />

57


for every u ∈ T and l ∈ [0,h u ]. Finally, with every leaf a of ˜θ we associate a stopped path w (a)<br />

with lif<strong>et</strong>ime d θ (0,a) : For every t ∈ [0,d θ (0,a)], w (a) (t) = V r(a,t) where r(a,t) is the unique<br />

element of [[0,a]] such that d θ (0,r(a,t)) = t.<br />

For every integer p ≥ 1, denote by A p the s<strong>et</strong> of all binary trees with p leaves, and by T p the<br />

corresponding s<strong>et</strong> of marked trees. The uniform measure Λ p on T p is defined by<br />

∫<br />

Λ p (dθ)F(θ) = ∑ ∫ ∏<br />

dh v F(T ,(h v ) v∈T ).<br />

T p T ∈A p v∈T<br />

With this notation, Proposition IV.2 of [37] states that, for every integer p ≥ 1 and every<br />

symm<strong>et</strong>ric nonnegative measurable function F on W p ,<br />

( ∫ ) ∫ [ ( (<br />

(3.2.1) N x ds 1 ...ds p F(W s1 ,... ,W sp ) = 2 p−1 p! Λ p (dθ)Q θ x F w (a)) .<br />

]0,σ[ a∈L(θ))]<br />

p<br />

We will need a stronger result concerning the case where the function F also depends on the<br />

range R of the Brownian snake. To state this result, denote by K the space of all compact subs<strong>et</strong>s<br />

of R d , which is equipped with the Hausdorff m<strong>et</strong>ric and the associated Borel σ-field. Suppose<br />

that under the probability measure Q θ x (for each choice of θ in T), in addition to the process<br />

(V a ,a ∈ ˜θ), we are also given an independent Poisson point measure on ˜θ × Ω, denoted by<br />

∑<br />

δ (ai ,ω i ),<br />

with intensity 4 L θ (da) ⊗ N 0 (dω).<br />

i∈I<br />

Theorem 3.2.2. For every nonnegative measurable function F on W p × K × R + , which is<br />

symm<strong>et</strong>ric in the first p variables, we have<br />

)<br />

N x<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F(W s1 ,...,W sp , R,σ)<br />

∫<br />

= 2 p−1 p!<br />

Λ p (dθ)Q θ x<br />

where cl(A) denotes the closure of the s<strong>et</strong> A.<br />

[<br />

F<br />

( (<br />

w (a)) a∈L(θ) ,cl ( ⋃<br />

i∈I<br />

(V ai + R(ω i ))<br />

)<br />

, ∑ i∈I<br />

σ(ω i )<br />

Remark. It is immediate to see that<br />

( ) ⎛<br />

⋃<br />

cl (V ai + R(ω i )) = ⎝ ⋃ [ w (a) 0,ζ (w )] ⎞ ( )<br />

⋃<br />

⎠ (a) ∪ (V ai + R(ω i )) , Q θ x a.e.<br />

i∈I<br />

i∈I<br />

a∈L(θ)<br />

Proof : Consider first the case p = 1. L<strong>et</strong> F 1 be a nonnegative measurable function on W,<br />

and l<strong>et</strong> F 2 and F 3 be two nonnegative measurable functions on Ω. By applying the Markov<br />

property under N x at time s, then using the time-reversal invariance of N x (which is easy from<br />

the analogous property for the Itô measure n(de)), and finally using the Markov property at<br />

)]<br />

,<br />

58


time s once again, we g<strong>et</strong><br />

(∫ σ ( ) ( ) )<br />

N x ds F 1 (W s )F 2 (W (s−r) +) r≥0 F 3 (W s+r ) r≥0<br />

0<br />

= N x<br />

(∫ σ<br />

0<br />

= N x<br />

(∫ σ<br />

0<br />

= N x<br />

(∫ σ<br />

0<br />

ds F 1 (W s )F 2<br />

(<br />

(W (s−r) +) r≥0<br />

)<br />

E Ws<br />

[<br />

F 3<br />

(<br />

(W r∧σ ) r≥0<br />

)] )<br />

ds F 1 (W s )F 2<br />

(<br />

(W s+r ) r≥0<br />

)<br />

E Ws<br />

[<br />

F 3<br />

(<br />

(W r∧σ ) r≥0<br />

)] )<br />

ds F 1 (W s ) E Ws<br />

[<br />

F 2<br />

(<br />

(W r∧σ ) r≥0<br />

)]<br />

E Ws<br />

[<br />

F 3<br />

((W r∧σ ) r≥0<br />

)] ) .<br />

We then use the case p = 1 of (3.2.1) to see that the last quantity is equal to<br />

∫ ∞<br />

∫<br />

[ ( )] [ (<br />

dt Px t (dw)F 1(w) E w F 2 (W r∧σ ) r≥0 E w F 3 (W r∧σ ) r≥0<br />

)],<br />

0<br />

where Px t denotes the law of Brownian motion started at x and stopped at time t (this law is<br />

viewed as a probability measure on W). Now if we specialize to the case where F 2 is a function<br />

of the form F 2 (ω) = G 2 ({Ŵs(ω) : s ≥ 0},σ), an immediate application of Lemma V.2 in [37]<br />

shows that<br />

⎡ ⎛ ⎛<br />

⎞ ⎞⎤<br />

[ ( )]<br />

E w F 2 (W r∧σ ) r≥0 = E ⎣G 2<br />

⎝cl (w(t j ) + R(ω j )) ⎠ , ∑ σ(ω j ) ⎠⎦ ,<br />

j∈J<br />

⎝ ⋃ j∈J<br />

where ∑ j∈J δ (t j ,ω j ) is a Poisson point measure on [0,ζ (w) ]×Ω with intensity 2 dt N 0 (dω). Applying<br />

the same observation to F 3 , we easily g<strong>et</strong> the case p = 1 of the theorem.<br />

The general case can be derived along similar lines by using Theorem 3 in [35]. Roughly<br />

speaking, the case p = 1 amounts to combining Bismut’s decomposition of the Brownian excursion<br />

(Lemma 1 in [35]) with the spatial displacements of the Brownian snake. For general p, the<br />

second assertion of Theorem 3 in [35] provides the analogue of Bismut’s decomposition, which<br />

when combined with spatial displacements leads to the statement of Theorem 3.2.2. D<strong>et</strong>ails are<br />

left to the reader.<br />

□<br />

3.2.3. The re-rooting theorem. In this subsection, we state and prove an important<br />

invariance property of the Brownian snake under N 0 , which plays a major role in Section 3.3<br />

below. We first need to introduce some notation. For every s,r ∈ [0,σ], we s<strong>et</strong><br />

{<br />

s + r if s + r ≤ σ ,<br />

s ⊕ r =<br />

s + r − σ if s + r > σ .<br />

We also use the following convenient notation for closed intervals : if u,v ∈ R, [u,v] = [v,u] =<br />

[u ∧ v,u ∨ v].<br />

L<strong>et</strong> s ∈ [0,σ[. In order to define the re-rooted snake W [s] , we first s<strong>et</strong><br />

if r ∈ [0,σ], and ζ r<br />

[s]<br />

that<br />

ζ [s]<br />

r = ζ s + ζ s⊕r − 2 inf<br />

u∈[s,s⊕r] ζ u ,<br />

= 0 if r > σ. We also want to define the stopped paths W [s]<br />

r , in such a way<br />

Ŵ [s]<br />

r = Ŵs⊕r − Ŵs ,<br />

59


if r ∈ [0,σ], and Ŵ r<br />

[s] = 0 if r > σ. To this end, we may notice that Ŵ [s] satisfies the property<br />

Ŵ [s]<br />

r<br />

= Ŵ [s]<br />

r ′<br />

if ζ [s]<br />

r<br />

= ζ [s]<br />

r ′<br />

= inf<br />

u∈[r,r ′ ] ζ[s] u<br />

and so in the terminology of [44], (W r [s] ) 0≤r≤σ is uniquely d<strong>et</strong>ermined as the snake whose tour<br />

is (ζ r [s] ,Ŵ r [s] ) 0≤r≤σ (see the homeomorphism theorem of [44]). We have the explicit formula, for<br />

r ≥ 0, and 0 ≤ t ≤ ζ r [s] ,<br />

(3.2.2) W [s]<br />

r (t) = Ŵ [s]<br />

sup{u≤r : ζ [s]<br />

u =t} .<br />

As explained in the introduction, (ζ r [s] ) r≥0 codes the same R-tree as the one coded by (ζ r ) r≥0 ,<br />

but with a new root which is the vertex originally labelled by s, and Ŵ r<br />

[s] gives the spatial<br />

displacements along the line segment from the (new) root to the vertex coded by r (in the<br />

coding given by ζ [s] ).<br />

Theorem 3.2.3. For every nonnegative measurable function F on R + × Ω,<br />

(∫ σ<br />

N 0 ds F<br />

(s,W [s])) (∫ σ<br />

)<br />

= N 0 ds F(s,W) .<br />

0<br />

Remark. For every s ∈ [0,σ[, the duration of the re-rooted snake excursion W [s] is the same<br />

as that of the original one. Using this simple observation, and replacing F by½{1−ε 0). Via<br />

a continuity argument, it follows that, for every s ∈ [0,1[, and every nonnegative measurable<br />

function G on Ω,<br />

(3.2.3) N (1)<br />

0<br />

( (<br />

G W [s])) = N (1)<br />

0 (G(W)).<br />

The identity (3.2.3) appears as Proposition 4.9 of [45], which is proved via discr<strong>et</strong>e approximations.<br />

Note that conversely, it would be easy to derive Theorem 3.2.3 from (3.2.3). We have<br />

chosen to give an independent proof of Theorem 3.2.3 because this result plays a major role<br />

in the present work, and also because the proof below fits in b<strong>et</strong>ter with our general strategy,<br />

which is to deal first with unnormalized excursion measures before conditioning with respect to<br />

the duration.<br />

Proof : By (3.2.2), W [s] can be written N 0 a.e. as Φ(ζ [s] ,Ŵ [s] ), where the d<strong>et</strong>erministic<br />

function Φ does not depend on s. Also note that when s = 0, W = W [0] = Φ(ζ,Ŵ), N 0 a.e. In<br />

view of these considerations, it will be sufficient to treat the case when<br />

)<br />

F(s,W) = F 1 (s,ζ)F 2<br />

(s,Ŵ<br />

where F 1 and F 2 are nonnegative measurable functions defined respectively on R + ×C(R + , R + )<br />

and on R + × C(R + , R d ). We first deal with the special case F 2 = 1.<br />

For s ∈ [0,σ[ and r ≥ 0, s<strong>et</strong><br />

ζ 1,s<br />

r = ζ (s−r) + − ζ s ,<br />

ζ 2,s<br />

r = ζ s+r − ζ s .<br />

0<br />

60


L<strong>et</strong> G be a nonnegative measurable function on R + × C(R + , R) × R + × C(R + , R). From the<br />

Bismut decomposition of the Brownian excursion (see e.g. Lemma 1 in [35]), we have<br />

(∫ σ (<br />

) ) ∫ ∞ [ (<br />

)]<br />

N 0 ds G s,(ζr 1,s ) r≥0 ,σ − s,(ζr 2,s ) r≥0 = daE G T a ,(B r∧Ta ) r≥0 ,T a ′ ,(B′ r∧T a ′) r≥0 ,<br />

0<br />

where B and B ′ are two independent linear Brownian motions started at 0, and<br />

Now observe that<br />

T a = inf{r ≥ 0 : B r = −a},<br />

T ′ a = inf{r ≥ 0 : B ′ r = −a}.<br />

0<br />

ζ r [s] = ζr<br />

2,s − 2 inf<br />

0≤u≤r ζ2,s u , if 0 ≤ r ≤ σ − s ,<br />

ζ [s]<br />

σ−r = ζ1,s r − 2 inf<br />

0≤u≤r ζ1,s u , if 0 ≤ r ≤ s ,<br />

and note that R t := B t − 2inf r≤t B r and R r ′ := B′ t − 2inf r≤t B r ′ are two independent threedimensional<br />

Bessel processes, for which<br />

L a := sup{t ≥ 0 : R t ≤ a} = T a ,<br />

L ′ a := sup{t ≥ 0 : R ′ t ≤ a} = T ′ a .<br />

(This is Pitman’s theorem, see e.g. [50], Theorem VI.3.5.) It follows that<br />

(∫ σ (<br />

) )<br />

N 0 ds G σ − s,(ζ [s]<br />

r∧(σ−s) ) r≥0,s,(ζ [s]<br />

σ−(r∧s) ) r≥0<br />

0<br />

∫ ∞ [<br />

))<br />

= daE G<br />

(L ′ a,(R r∧L ′ ′ ) r≥0,L<br />

a a ,(R r∧La ) r≥0<br />

0<br />

(∫ σ<br />

)<br />

= N 0 ds G<br />

(s,(ζ )<br />

r∧s ) r≥0 ,σ − s,(ζ (σ−r)∨s ) r≥0<br />

0<br />

where the last equality is again a consequence of the Bismut decomposition, tog<strong>et</strong>her with the<br />

Williams reversal theorem ([50], Corollary XII.4.4). Changing s into σ − s in the last integral<br />

gives the desired result when F 2 = 1.<br />

L<strong>et</strong> us consider the general case. For simplicity we take d = 1, but the argument can obviously<br />

be extended. From the definition of the Brownian snake, we have<br />

(∫ σ<br />

)<br />

N 0 ds F 1 (s,ζ)F 2<br />

(s,Ŵ<br />

) ( ∫ σ<br />

( )])<br />

= N 0 ds F 1 (s,ζ)Θ ζ 0<br />

[F 2 s,Ŵ<br />

0<br />

and Ŵ is under Θζ 0 a centered Gaussian process with covariance<br />

)<br />

cov Θ<br />

ζ<br />

(Ŵs ,Ŵs ′ = inf ζ r.<br />

0<br />

r∈[s,s ′ ]<br />

We have in particular<br />

(∫ σ (<br />

N 0 ds F 1 s,ζ [s]) (<br />

F 2 s,Ŵ [s]))<br />

0<br />

(∫ σ<br />

= N 0 ds F 1<br />

(s,ζ [s]) ( )<br />

Θ ζ 0<br />

[F 2 s,(Ŵs⊕r − Ŵs<br />

0<br />

0<br />

r≥0<br />

)])<br />

.<br />

Now note that (Ŵs⊕r − Ŵs) r≥0 is under Θ ζ 0 a Gaussian process with covariance<br />

)<br />

cov<br />

(Ŵs⊕r − Ŵs,Ŵs⊕r ′ − Ŵs = inf ζ<br />

[s⊕r,s⊕r ′ u − inf ζ u − inf ζ<br />

] [s⊕r,s] [s⊕r ′ u + ζ s = inf<br />

,s]<br />

61<br />

[r,r ′ ] ζ[s] u ,


where the last equality follows from an elementary verification. Hence,<br />

( ) )] ( )]<br />

Θ ζ 0<br />

[F 2 s,(Ŵs⊕r − Ŵs = Θ ζ[s]<br />

0<br />

[F 2 s,Ŵ ,<br />

and, using the first part of the proof,<br />

(∫ σ<br />

N 0<br />

0<br />

This compl<strong>et</strong>es the proof.<br />

r≥0<br />

ds F 1<br />

(<br />

s,ζ [s]) F 2<br />

(<br />

s,Ŵ [s])) = N 0<br />

(∫ σ<br />

0<br />

(<br />

ds F 1 s,ζ [s]) ( )]<br />

Θ ζ[s]<br />

0<br />

[F )<br />

2 s,Ŵ<br />

(∫ σ<br />

= N 0 ds F 1 (s,ζ)Θ ζ 0<br />

0<br />

( )] [F )<br />

2 s,Ŵ<br />

(∫ σ (s,Ŵ) )<br />

= N 0 ds F 1 (s,ζ)F 2 .<br />

0<br />

□<br />

3.2.4. The special Markov property. L<strong>et</strong> D be a domain in R d , and fix a point x ∈ D.<br />

For every w ∈ W, we s<strong>et</strong><br />

τ(w) := inf{t ≥ 0 : w(t) /∈ D}<br />

where inf ∅ = +∞ as usual. The random s<strong>et</strong><br />

{s ≥ 0 : τ(W s ) < ζ s }<br />

is open N x a.e., and can thus be written as a disjoint union of open intervals ]a i ,b i [, i ∈ I. It is<br />

easy to verify that N x a.e. for every i ∈ I and every s ∈]a i ,b i [,<br />

τ (W s ) = τ (W ai ) = τ (W bi ) = ζ ai = ζ bi<br />

and moreover the paths W s , s ∈ [a i ,b i ] coincide up to their exit time from D.<br />

For every i ∈ I, we define a random element W (i) of Ω by s<strong>et</strong>ting for every s ≥ 0<br />

W (i)<br />

s (t) = W (ai +s)∧b i<br />

(ζ ai + t),<br />

for 0 ≤ t ≤ ζ (W<br />

(i)<br />

s ) := ζ (a i +s)∧b i<br />

− ζ ai .<br />

Informally, the W (i) ’s represent the excursions of the Brownian snake outside D (the word<br />

“outside” is a bit misleading since these excursions may come back into D even though they<br />

start from the boundary of D).<br />

Finally, we also need a process that contains the information given by the Brownian snake<br />

paths before they exit D. We s<strong>et</strong> ˜W s D = W η D s<br />

, where for every s ≥ 0,<br />

∫ r<br />

}<br />

ηs {r D := inf ≥ 0 : du½{τ(W u)≥ζ u} > s .<br />

0<br />

The σ-field E D is by definition generated by the process ˜W D and by the class of N x -negligible<br />

subs<strong>et</strong>s of Ω (the point x is fixed throughout this subsection). The following statement is proved<br />

in [36] (Proposition 2.3 and Theorem 2.4).<br />

Theorem 3.2.4. There exists a random finite measure denoted by Z D , which is E D -measurable<br />

and N x a.e. supported on ∂D, such that the following holds. Under N x , conditionally on E D , the<br />

point measure<br />

N := ∑ i∈I<br />

δ (W (i) )<br />

is Poisson with intensity ∫ ∂D ZD (dy) N y (·).<br />

62


We will apply this theorem to the case d = 1, x = 0 and D =]c, ∞[ for some c < 0. In that<br />

case, the measure Z D is a random multiple of the Dirac measure at c : Z D = L c δ c for some<br />

nonnegative random variable L c . From Lemma 3.2.1(i) and Theorem 3.2.4, it is easy to verify<br />

that {L c > 0} = {R∩] − ∞,c] ≠ ∅} = {R∩] − ∞,c[≠ ∅} , N 0 a.e. Moreover, as a simple<br />

consequence of the special Markov property, the process (L −r ) r>0 is a nonnegative martingale<br />

under N 0 (it is indeed a critical continuous-state branching process). In particular the variable<br />

is finite N 0 a.e., for every r < 0.<br />

L ∗,r :=<br />

sup L c<br />

c∈Q∩]−∞,r]<br />

3.2.5. Uniqueness of the minimum. From now on we assume that d = 1. In this subsection,<br />

we consider the Brownian snake under its excursion measure N 0 . We use the notation<br />

W = inf Ŵ s .<br />

s≥0<br />

Note that the law of W under N 0 is given by Lemma 3.2.1(i) and an obvious translation argument.<br />

Proposition 3.2.5. There exists N 0 a.e. a unique instant s ∗ ∈]0,σ[ such that Ŵs ∗<br />

= W.<br />

This result already appears as Lemma 16 in [45], where its proof is attributed to T. Duquesne.<br />

We provide a short proof for the sake of compl<strong>et</strong>eness and also because this result plays a major<br />

role throughout this work.<br />

Proof : S<strong>et</strong><br />

{<br />

}<br />

λ := inf s ≥ 0 : Ŵs = W ,<br />

{<br />

}<br />

ρ := sup s ≥ 0 : Ŵs = W<br />

so that 0 < λ ≤ ρ < σ. We have to prove that λ = ρ. To this end we fix δ > 0 and we verify<br />

that N 0 (ρ − λ > δ) = 0.<br />

Fix two rational numbers q < 0 and ε > 0. We first g<strong>et</strong> an upper bound on the quantity<br />

N 0 (q − ε ≤ W < q , ρ − λ > δ).<br />

Denote by (W (i) ) i∈I the excursions of the Brownian snake outside ]q, ∞[, and by N the corresponding<br />

point measure, as in the previous subsection. Since the law of W under N 0 has no<br />

atoms, the numbers W (i) , i ∈ I are distinct N 0 a.e. Therefore, on the event {W < q}, the whole<br />

interval [ρ,λ] must be contained in a single excursion interval below level q. Hence,<br />

}<br />

N 0 (q − ε ≤ W < q , ρ − λ > δ) ≤ N 0<br />

({∀i ∈ I : W (i) ≥ q − ε ∩<br />

{<br />

∃i ∈ I : σ<br />

Introduce the events A ε := {W < q − ε} and B ε,δ = {W ≥ q − ε , σ > δ}. We g<strong>et</strong><br />

N 0 (q − ε ≤ W < q , ρ − λ > δ) ≤ N 0 (N(A ε ) = 0, N(B ε,δ ) ≥ 1).<br />

(<br />

W (i)) > δ<br />

From the special Markov property (and the remarks of the end of subsection 3.2.3), we know<br />

that conditionally on L q , N is a Poisson point measure with intensity L q N q . Since the s<strong>et</strong>s A ε<br />

and B ε,δ are disjoint, independence properties of Poisson point measures give<br />

)<br />

N 0 (q − ε ≤ W < q , ρ − λ > δ) ≤ N 0<br />

(½{N(A ε)=0} (1 − exp(−L q N q (B ε,δ )))<br />

)<br />

= N 0<br />

(½{q−ε≤W


where c(ε,δ) = N q (B ε,δ ) does not depend on q by an obvious translation argument.<br />

We can apply the preceding bound with q replaced by q − ε,q − 2ε, <strong>et</strong>c. By summing the<br />

resulting estimates we g<strong>et</strong><br />

)<br />

N 0 (W < q , ρ − λ > δ) ≤ N 0<br />

(½{W δ) = 0.<br />

3.2.6. Bessel processes. Throughout this work, (ξ t ) t≥0 will stand for a linear Brownian<br />

motion started at x under the probability measure P x . The notation ξ[0,t] will stand for the<br />

range of ξ over the time interval [0,t]. For every δ > 0, (R t ) t≥0 will denote a Bessel process<br />

of dimension δ started at x under the probability measure P x (δ) . We will use repeatedly the<br />

following simple facts. First, if λ > 0, the process R (λ)<br />

t := λ −1 R λ 2 t is under P x<br />

(δ) a Bessel process<br />

of dimension δ started at x/λ. Secondly, if 0 ≤ x ≤ x ′ and t ≥ 0, the law of R t under P x<br />

(δ)<br />

is stochastically bounded by the law of R t under P (δ)<br />

x<br />

. The latter fact follows from standard<br />

′<br />

comparison theorems applied to squared Bessel processes.<br />

Absolute continuity relations b<strong>et</strong>ween Bessel processes, which are consequences of the Girsanov<br />

theorem, were first observed by Yor [55]. We state a special case of these relations, which<br />

will play an important role in this work. This special case appears in Exercise XI.1.22 of [50].<br />

Proposition 3.2.6. L<strong>et</strong> t > 0 and l<strong>et</strong> F be a nonnegative measurable function on C([0,t], R).<br />

Then, for every x > 0 and λ > 0,<br />

∫ t<br />

) ]<br />

E x<br />

[½{ξ[0,t]⊂]0,∞[} exp<br />

(− λ2 dr<br />

[<br />

]<br />

2 0 ξr<br />

2 F((ξ r ) 0≤r≤t ) = x ν+1 2 E<br />

(2+2ν)<br />

x (R t ) −ν−1 2 F((Rr ) 0≤r≤t ) ,<br />

√<br />

where ν = λ 2 + 1 4 .<br />

We shall be concerned by the case when λ 2 /2 = 6, and then 2 + 2ν = 9 and ν + 1/2 = 4.<br />

Taking F = 1 in that case, we see that<br />

(3.2.4) x 4 E x<br />

(9) [ ]<br />

R<br />

−4<br />

t ≤ 1.<br />

3.3. Conditioning and re-rooting of trees<br />

This section contains the proof of Theorem 3.1.1 and Theorem 3.1.2 which were stated in the<br />

introduction. Both will be derived as consequences of Theorem 3.3.1 below. Recall the notation<br />

s ∗ for the unique time of the minimum of Ŵ under N 0, and W [s] for the snake re-rooted at s.<br />

Theorem 3.3.1. L<strong>et</strong> ϕ : R + −→ R + be a continuous function such that ϕ(s) ≤ C(1 ∧ s) for<br />

some finite constant C. L<strong>et</strong> F : Ω −→ R + be a bounded continuous function. Then,<br />

(<br />

)<br />

lim<br />

ε→0 ε−4 N 0 σ ϕ(σ)F(W)½{W >−ε} = 2 ( (<br />

21 N 0 ϕ(σ)F W [s∗])) .<br />

□<br />

64


The proof of Theorem 3.3.1 occupies most of the remainder of this section. This proof will<br />

depend on a series of lemmas. To motivate these lemmas, we first observe that, from the rerooting<br />

identity in Theorem 3.2.3, we have<br />

)<br />

∫ σ (<br />

)<br />

N 0<br />

(σ ϕ(σ)F(W)½{W >−ε} = N 0<br />

(ϕ(σ) ds F W [s])½{W[s] >−ε}<br />

0<br />

∫ σ (<br />

)<br />

(3.3.1)<br />

= N 0<br />

(ϕ(σ) ds F W [s])½{ W c s


Similarly,<br />

N 0<br />

(½{W>−ε}e −σ ∫ σ<br />

=<br />

=<br />

=<br />

∫ δ<br />

0<br />

∫ δ<br />

0<br />

∫ δ<br />

0<br />

0<br />

ds½{ζ s−ε} exp<br />

)<br />

(<br />

−4<br />

(<br />

daE ε<br />

[½{ξ a<br />

>0} exp −4<br />

∫ a<br />

0<br />

∫ a<br />

∫ a<br />

daE ε<br />

[½{ξ a<br />

>0} exp<br />

(−2 3/2 dt<br />

0<br />

)]<br />

dt N ξt (1 −½{W>−ε)}e −σ )<br />

)]<br />

dt N ξt (1 −½{W>0)}e −σ )<br />

0<br />

( ) ))] 2<br />

3coth<br />

(2 1/4 ξ t − 2<br />

using Lemma 3.2.1(ii) in the last equality. For every x > 0, s<strong>et</strong><br />

h(x) = − 3 ( ( ) ) 2<br />

2x 2 + 2−1/2 3coth 2 1/4 x − 2 > 0.<br />

A Taylor expansion shows that h(x) ≤ C x 2 . (Here and later, C,C ′ ,C ′′ denote constants whose<br />

exact value is unimportant.) Then,<br />

∫ σ<br />

) ∫ δ<br />

( ∫ a ∫<br />

N 0<br />

(½{W>−ε}e −σ dt a<br />

)]<br />

= daE ε<br />

[½{ξ a<br />

>0} exp −6 − 4 dt h(ξ t )<br />

0<br />

ds½{ζ s−ε}(1 − e −σ )<br />

=<br />

≤<br />

∫ δ<br />

0<br />

4C<br />

∫ σ<br />

0<br />

ds½{ζ s−ε}(1 − e −σ ) ds½{ζ s


Lemma 3.3.3. For every δ > 0,<br />

( ( ∫ σ<br />

) ) 2<br />

sup N 0 ε −4 ds½{ W<br />

0


To simplify notation, we have written E instead of E a,b,c , and η[0,a] obviously denotes the s<strong>et</strong><br />

{η t : 0 ≤ t ≤ a}, with a similar notation for η ′ [0,b] and η ′′ [0,c].<br />

In the preceding formula for Jε<br />

a,b,c<br />

quantity depending on y = (η<br />

b ′ ∧ η′′<br />

(3.3.9)<br />

J a,b,c<br />

ε<br />

E<br />

[½{η[0,a]⊂]−y,∞[} exp<br />

(<br />

−6<br />

c<br />

∫ a<br />

0<br />

, conditioning with respect to the pair (η ′ ,η ′′ ) leads to a<br />

) + ε, of the form<br />

)]<br />

( ∫<br />

dt<br />

a<br />

)]<br />

dt<br />

(η t + y) 2 = E y<br />

[½{ξ[0,a]⊂]0,∞[} exp −6<br />

0<br />

= y 4 E (9)<br />

y<br />

[ ]<br />

R<br />

−4<br />

a<br />

using Proposition 3.2.6 as in the proof of Lemma 3.3.2 above. Hence,<br />

= E<br />

[½{|η b ′ − η′′ c | < ε,η′ [0,b] ⊂] − ε, ∞[,η ′′ [0,c] ⊂] − ε, ∞[}<br />

( (∫<br />

× ((η b ′ ∧ η′′ c ) + ε)4 E (9) [ ] b<br />

(η R<br />

−4<br />

b ′ ∧η′′ c )+ε a exp −6<br />

0<br />

∫<br />

dt c<br />

(η t ′ + + ε)2<br />

0<br />

ξ 2 t<br />

dt<br />

(η ′′<br />

t + ε)2 )) ] .<br />

Recall that our goal is to bound ∫ dadbdc½{a+b≥δ,a+c≥δ} Jε<br />

a,b,c . First consider the integral over<br />

the s<strong>et</strong> {a < δ/2}. Then plainly we have b > δ/2 and c > δ/2, and we can use (3.2.4) to bound<br />

∫<br />

{a 0,<br />

∫ ∞<br />

∫<br />

(3.3.10)<br />

dbE y<br />

(9) [ ] ∞<br />

R<br />

−4<br />

b ≤ dbE (9) [ ]<br />

0 R<br />

−4<br />

b = C<br />

′<br />

δ < ∞.<br />

δ/2<br />

0<br />

))]<br />

dt<br />

(η t ′′ + ε)2<br />

We still have to g<strong>et</strong> a similar bound for the integral over the s<strong>et</strong> {a ≥ δ/2}. Applying the<br />

bound (3.3.10) with y = (η<br />

b ′ ∧ η′′ c ) + ε, we see that it is enough to prove that<br />

∫ [½{|η b ′ − η′′ c | < ε, η′ [0,b] ⊂] − ε, ∞[, η ′′ [0,c] ⊂] − ε, ∞[}<br />

(3.3.11)<br />

[0,∞[ 2 dbdcE<br />

( (∫ b<br />

×((η b ′ ∧ η′′ c ) + ε)4 exp −6<br />

0<br />

∫<br />

dt c<br />

(η t ′ + + ε)2<br />

From Proposition 3.2.6 again, the left-hand side of (3.3.11) is equal to<br />

∫<br />

(<br />

(3.3.12) ε 8 dbdcE ε (9) ⊗ E ε<br />

[½{|R (9)<br />

[0,∞[ 2 b −R e c|


where R and ˜R are two independent nine-dimensional Bessel processes started at ε under the<br />

probability measure P ε (9) ⊗ P ε<br />

(9) . The quantity (3.3.12) is bounded above by ε 8 (I1 ε + Iε 2 ), where<br />

∫<br />

(<br />

I1 ε = dbdcE ε (9) ⊗ E ε<br />

[½{R (9)<br />

b


N = ∑ i∈I δ ω i with intensity xN 0, under the probability measure P (x) . To simplify notation, we<br />

write Ws i = W s (ω i ). We then s<strong>et</strong><br />

( )<br />

W = inf inf Ŵs<br />

i = inf W i .<br />

i∈I s≥0 i∈I<br />

Lemma 3.3.5. For every x > 0 and ε > 0,<br />

E (x)<br />

[ ∑<br />

i∈I<br />

∫ σ(ω i )<br />

ds½{ W<br />

0<br />

c s


which gives the formula of the lemma, with<br />

[∫ ∞<br />

(<br />

g(u) = E u<br />

(9) dbR −4<br />

b<br />

exp − 3 )]<br />

0<br />

2Rb<br />

2 .<br />

The fact that g is nondecreasing follows from the strong Markov property of R. The continuity<br />

of g is easy from a similar argument. Finally, the value of g(0) is obtained from the explicit<br />

formula for the Green function of nine-dimensional Brownian motion.<br />

□<br />

Recall our notation E ]a,∞[ for the σ-field generated by the Brownian snake paths before their<br />

first exit from ]a, ∞[, and L a for the total mass of the exit measure Z ]a,∞[ .<br />

Lemma 3.3.6. L<strong>et</strong> a < 0. For every bounded E ]a,∞[ -measurable function Φ on Ω,<br />

)<br />

lim N 0<br />

= 2<br />

ε→0 21 N 0<br />

(½{W


It remains to study<br />

ε −4 N 0<br />

(½{W


and with a suitable choice of Φ), we g<strong>et</strong>, for every integer n ≥ n 0 and every i ∈ Z such that<br />

−2 n ≤ i2 −n < a,<br />

( (<br />

lim ε −4 i2<br />

N 0<br />

(½{W


Proof of Theorems 3.1.1 and 3.1.2 : Both Theorems 3.1.1 and 3.1.2 follow from the<br />

convergence<br />

( )<br />

(3.3.20) lim ε −4 N (1)<br />

ε→0<br />

0 F(W)½{W>−ε} = 2 ( (<br />

21 N(1) 0 F W [s∗])) ,<br />

which holds for every bounded continuous function F on Ω = C(R + , W) (take F = 1 to recover<br />

the first assertion of Theorem 3.1.1). We will now derive (3.3.20) from Theorem 3.3.1.<br />

For every λ > 0, l<strong>et</strong> us introduce the scaling operator θ λ defined on Ω by<br />

ζ s ◦ θ λ = λ 1/2 ζ s/λ ,<br />

W s ◦ θ λ (t) = λ 1/4 W s/λ (λ −1/2 t) .<br />

Note that, for every r > 0, the image of N (r)<br />

0 under θ 1/r is N (1)<br />

0 .<br />

L<strong>et</strong> δ ∈]0,1[. It follows from Theorem 3.3.1 that the law of the pair (σ,W) under<br />

µ ε,δ := ε −4 N 0 (· ∩ {W > −ε, 1 − δ < σ < 1})<br />

converges weakly as ε → 0 towards the law of (σ,W [s∗] ) under the measure µ δ having density<br />

2/(21σ) with respect to N 0 (· ∩ {1 − δ < σ < 1}).<br />

Since the mapping (r,ω) → θ r ω is continuous, it follows that the law of W ◦ θ 1/σ under µ ε,δ<br />

converges as ε → 0 towards the law of W [s∗] ◦ θ 1/σ under µ δ . Thus,<br />

lim µ ε,δ(F(W ◦ θ 1/σ )) = µ δ (F(W [s∗] ◦ θ 1/σ )),<br />

ε→0<br />

or equivalently<br />

) ( 2<br />

lim<br />

ε→0 ε−4 N 0<br />

(½{W>−ε, 1−δ


Proposition 3.3.7. For any bounded continuous function F on Ω,<br />

lim N (1)<br />

0 ε↓0<br />

(F(W) | Z(] − ∞,0]) ≤ ε) = N(1) 0 (F).<br />

Proof : From the re-rooting theorem 3.2.3 and the remark after this statement we have,<br />

(∫<br />

(3.3.21) N (1) ( 1 (<br />

)<br />

(1)<br />

0 F(W)½{Z(]−∞,0])≤ε})<br />

= N 0 ds F W [s])½{Z(]−∞, W c s])≤ε}<br />

.<br />

Now, since the measure Z has no atoms,<br />

∫ 1<br />

∫<br />

ds½{Z(]−∞, W c = s])≤ε}<br />

0<br />

0<br />

Z(dy)½{Z(]−∞,y])≤ε} = ε.<br />

In particular by taking F = 1 in (3.3.21) we see that the law of Z(]−∞,0]) under N (1)<br />

0 is uniform<br />

over [0,1] (this fact was already observed in Section 3.2 of [6]). Hence,<br />

N (1)<br />

0 (F(W) | Z(] − ∞,0]) ≤ ε) = ε−1 N (1)<br />

0<br />

(∫ 1<br />

ds F<br />

0<br />

(<br />

W [s])½{Z(]−∞, c W s])≤ε}<br />

On the other hand the closed s<strong>et</strong>s {s ∈ [0,1] : Z(] − ∞,Ŵs]) ≤ ε} decrease to the singl<strong>et</strong>on<br />

{s ∗ } as ε ↓ 0, N (1)<br />

0<br />

a.e. It follows that<br />

∫ 1<br />

lim ε −1 ds F<br />

(W [s])½{Z(]−∞, c ε↓0 W = F s])≤ε}<br />

(W [s∗])<br />

0<br />

N (1)<br />

0 a.e. Using dominated convergence and Theorem 3.1.2, we g<strong>et</strong> Proposition 3.3.7. □<br />

Remark. Up to some point, the approximation given in Proposition 3.3.7 may seem as<br />

“reasonable” as the one we used in Theorem 3.1.1 to define N (1)<br />

0 . In applications such that the<br />

ones developed in [39], the approximation given by Theorem 3.1.1 turns out to be more useful.<br />

3.4. Other conditionings<br />

Motivated by Theorem 3.3.1, we define a σ-finite measure N 0 on Ω by s<strong>et</strong>ting<br />

0<br />

N 0 (F) = N 0<br />

( 1<br />

σ F (W [s∗])) .<br />

Theorem 3.3.1 shows that, up to the multiplicative constant 2/21, N 0 is the limit in an appropriate<br />

sense of the measures ε −4 N 0 (· ∩ {W > −ε}) as ε → 0. We have also<br />

∫ ∞ ( (<br />

N 0 (F) =<br />

F W [s∗])) ∫ ∞<br />

=<br />

dr<br />

2 √ 2πr 5 N(r) 0<br />

0<br />

dr<br />

2 √ 2πr 5 N(r) 0 (F),<br />

where N (r)<br />

0 can be defined equivalently as the law of W [s∗] under N (r)<br />

0 , or as the image of N(1) 0<br />

under the scaling operator θ r .<br />

We will now describe a different approach to N 0 , which involves conditioning the Brownian<br />

snake excursion on its height H = sup s≥0 ζ s , rather than on its length as in Theorem 3.1.1. This<br />

will give more insight in the behavior of the Brownian snake under N 0 . Eventually, this will lead<br />

to a construction of a Brownian snake excursion with infinite length conditioned to stay on the<br />

positive side. We rely on some ideas from [1].<br />

)<br />

.<br />

75


For every h > 0, we s<strong>et</strong> N h 0 = N 0(· | H = h). Then,<br />

N 0 =<br />

∫ ∞<br />

0<br />

dh<br />

2h 2 Nh 0 .<br />

From Theorem 1 in [1] we know that there exists a constant c 0 > 0 such that<br />

(3.4.1) lim<br />

ε→0<br />

ε −4 N 1 0(W > −ε) = c 0 .<br />

A simple scaling argument then implies that, for every h > 0,<br />

(3.4.2) lim<br />

ε→0<br />

ε −4 N h 0(W > −ε) = c 0<br />

h 2.<br />

Theorem 3.4.1. For every h > 0, there exists a probability measure N h 0 on Ω such that<br />

lim<br />

ε→0 Nh 0(· | W > −ε) = N h 0<br />

in the sense of weak convergence on the space of probability measures on Ω. Moreover,<br />

N 0 = 21c 0<br />

4<br />

∫ ∞<br />

0<br />

dh<br />

h 4 Nh 0.<br />

Remark. Our proof of the first part of Theorem 3.4.1 does not use Section 3.3. This proof<br />

thus gives another approach to the conditioned measure N 0 , which does not depend on the<br />

re-rooting m<strong>et</strong>hod that played a crucial role in Section 3.3.<br />

Before proving Theorem 3.4.1, we will establish an important preliminary result. We first<br />

introduce some notation. Following [1], we s<strong>et</strong> for every ε > 0,<br />

and, for every x > 0,<br />

f(ε) = N 1 0 (W > −ε)<br />

G(x) = 4<br />

∫ x<br />

0<br />

u(1 − f(u))du.<br />

The function G is obviously nondecreasing. It is also bounded since<br />

G(∞) = 4<br />

∫ ∞<br />

0<br />

u N 1 0(W ≤ −u)du = 2<br />

by a scaling argument and Lemma 3.2.1(i).<br />

∫ ∞<br />

0<br />

r −2 N r 0(W ≤ −1)dr = 4 N 0 (W ≤ −1) = 6<br />

By well-known properties of Brownian excursions, there exists N h 0 a.s. a unique time α ∈]0,σ[<br />

such that ζ α = h. The next proposition discusses the law of W α under N h 0 (· | W > −ε).<br />

Proposition 3.4.2. L<strong>et</strong> Φ be a bounded continuous function on W. Then,<br />

[<br />

(∫<br />

lim ε −4 N h ( h<br />

( ))]<br />

(9)<br />

0 Φ(Wα )½{W>−ε})<br />

= E 0 Φ(R t ,0 ≤ t ≤ h)R −4 dt Rt<br />

ε↓0<br />

h<br />

exp G √ . h − t<br />

Remarks. (i) From the bound G(x) ≤ 6 ∧ (2x 2 ), it is immediate to verify that<br />

∫ h<br />

( )<br />

dt Rt<br />

G √ < ∞ , P (9)<br />

h − t<br />

0 a.s.<br />

0<br />

R 2 t<br />

(ii) By taking Φ = 1, we see that the constant c 0 in (3.4.1) is given by<br />

[ (∫ 1<br />

( ))]<br />

c 0 = E (9)<br />

0 R1 −4 dt Rt<br />

exp G √ , 1 − t<br />

0<br />

R 2 t<br />

0<br />

R 2 t<br />

76


as it was already observed in [1]. The fact that the quantity in the right-hand side is finite<br />

follows from the proof below.<br />

Proof : Our main tool is Williams’ decomposition of the Brownian excursion at its maximum<br />

(see e.g. Theorem XII.4.5 in [50]). For every s ≥ 0, we s<strong>et</strong><br />

ρ s = ζ s∧α ,<br />

ρ ′ s = ζ (σ−s)∨α.<br />

Under the probability measure N h 0 , the processes (ρ s) s≥0 and (ρ ′ s) s≥0 are two independent threedimensional<br />

Bessel processes started at 0 and stopped at their first hitting time of h.<br />

We also need to introduce the excursions of ρ and ρ ′ above their future infimum. S<strong>et</strong><br />

ρ s<br />

= inf<br />

r≥s ρ r<br />

and l<strong>et</strong> (a j ,b j ), j ∈ J be the connected components of the open s<strong>et</strong> {s ≥ 0 : ρ s > ρ s<br />

}. For every<br />

j ∈ J, define<br />

Then, by excursion theory,<br />

ζ j s = ρ (aj +s)∧b j<br />

− ρ aj , s ≥ 0<br />

h j = ρ aj .<br />

∑<br />

δ (hj ,ζ j )(dr de)<br />

is a Poisson point measure on R + × C(R + , R + ) with intensity<br />

j∈J<br />

2½[0,h](r)½[0,h−r](H(e))dr n(de)<br />

where H(e) = sup s≥0 e(s) as previously. The same result obviously holds for the analogous point<br />

measure<br />

∑<br />

j∈J ′ δ (h ′<br />

j ,ζ ′j )(dr de)<br />

obtained by replacing ρ with ρ ′ .<br />

We can combine the preceding assertions with the spatial displacements of the Brownian<br />

snake, in a way very similar to the proof of Lemma V.5 in [37]. For every j ∈ J, we s<strong>et</strong><br />

W j s (t) = W (aj +s)∧b j<br />

(h j + t) − Ŵa j<br />

, 0 ≤ t ≤ ζ j s, s ≥ 0.<br />

Note that by the properties of the Brownian snake Ŵa j<br />

= Ŵb j<br />

= W α (h j ). Then,<br />

N := ∑ j∈J<br />

δ (hj ,W j )<br />

is under N h 0 a Poisson point measure on R + × Ω, with intensity<br />

(3.4.3) 2½[0,h](r)½[0,h−r](H(ω))dr N 0 (dω).<br />

The same holds for the analogous point measure<br />

N ′ := ∑ j∈J ′ δ (h ′<br />

j ,W ′j ).<br />

Moreover N and N ′ are independent and the pair (N,N ′ ) is independent of W α . All these<br />

assertions easily follow from properties of the Brownian snake.<br />

77


Now note that the range of the Brownian snake under N h 0 can be decomposed as<br />

( ⋃<br />

) ( ⋃<br />

)<br />

{W α (t) : 0 ≤ t ≤ h} ∪ (W α (h j ) + R(W j )) ∪ (W α (h ′ j) + R(W ′j )) .<br />

j∈J<br />

j∈J ′<br />

Using this observation and conditioning with respect to W α , we g<strong>et</strong><br />

(<br />

Φ(Wα )½{W>−ε})<br />

N h 0<br />

= N h 0<br />

(<br />

(<br />

Φ(W α )½{W α(t)>−ε, 0≤t≤h} exp −4<br />

= E 0<br />

[Φ(ξ t , 0 ≤ t ≤ h)½{ξ[0,h]⊂]−ε,∞[} exp<br />

Then, for every x > −ε,<br />

N x (H < h − t, W ≤ −ε) =<br />

and we obtain<br />

(3.4.4)<br />

0<br />

N h 0 (Φ(W α)½{W>−ε})<br />

∫ h−t<br />

= E 0<br />

[Φ(ξ t , 0 ≤ t ≤ h)½{ξ[0,h]⊂]−ε,∞[} exp<br />

0<br />

∫ h<br />

0<br />

(<br />

−4<br />

))<br />

dt N Wα(t)(H < h − t, W ≤ −ε)<br />

∫ h<br />

0<br />

)]<br />

dt N ξt (H < h − t, W ≤ −ε) .<br />

∫<br />

du<br />

h−t<br />

( ( ))<br />

2u 2 Nu x (W ≤ −ε) = du x + ε<br />

2u 2 1 − f √ u<br />

(<br />

−2<br />

∫ h<br />

(<br />

= E ε<br />

[Φ(ξ t − ε, 0 ≤ t ≤ h)½{ξ[0,h]⊂]0,∞[} exp −2<br />

0<br />

∫ h<br />

0<br />

∫ h−t<br />

dt<br />

0<br />

0<br />

∫ h−t<br />

dt<br />

0<br />

(<br />

du<br />

u 2 1 − f<br />

(<br />

ξt + ε<br />

√ u<br />

)))]<br />

( ( )))]<br />

du ξt<br />

u 2 1 − f √ . u<br />

For every x > 0, the change of variable v = x/ √ u gives<br />

∫ h−t<br />

( ( )) ∫<br />

du x<br />

∞<br />

(<br />

u 2 1 − f √u = 2x −2 dv v(1 − f(v)) = x −2 3 − 1 ( )) x<br />

2 G √ . h − t<br />

x/ √ h−t<br />

By substituting this into (3.4.4) and using Proposition 3.2.6 once more, we g<strong>et</strong><br />

(<br />

Φ(Wα )½{W>−ε})<br />

N h 0<br />

= E ε<br />

[Φ(ξ t − ε, 0 ≤ t ≤ h)½{ξ[0,h]⊂]0,∞[} exp<br />

( ∫ h<br />

−6<br />

[<br />

(∫ h<br />

( ))]<br />

= ε 4 E ε<br />

(9) Φ(R t − ε,0 ≤ t ≤ h)R −4 dt Rt<br />

(3.4.5)<br />

h<br />

exp<br />

0 Rt<br />

2 G √ . h − t<br />

In view of (3.4.5), the proof of Proposition 3.4.2 reduces to checking that<br />

(3.4.6)<br />

lim E ε<br />

(9)<br />

ε↓0<br />

= E (9)<br />

0<br />

[<br />

Φ(R t − ε,0 ≤ t ≤ h)R −4<br />

h<br />

exp<br />

[<br />

Φ(R t ,0 ≤ t ≤ h)R −4<br />

h<br />

exp<br />

(∫ h<br />

0<br />

(∫ h<br />

0<br />

0<br />

dt<br />

R 2 t<br />

dt<br />

ξ 2 t<br />

dt<br />

R 2 t<br />

G<br />

∫ h<br />

( ))]<br />

dt ξt<br />

+<br />

0 ξt<br />

2 G √<br />

h − t<br />

G<br />

(<br />

Rt<br />

√<br />

h − t<br />

))]<br />

(<br />

Rt<br />

√<br />

h − t<br />

))]<br />

This follows from a dominated convergence argument, which at the same time will prove that the<br />

quantity in the right-hand side of (3.4.6) is well-defined. Note that we may define on a common<br />

probability space, a nine-dimensional Bessel process X ε = (Xt ε ,t ≥ 0) started at ε, for every<br />

ε ≥ 0, in such a way that the inequality X ε ≥ X 0 holds a.s. for every ε > 0. Since<br />

( ) X<br />

ε<br />

G √ t<br />

≤ 4(Xε t ) 2<br />

, ∀t ∈ [0,h/2]<br />

h − t h<br />

78


we first g<strong>et</strong><br />

(∫ h<br />

( ))<br />

(Xh ε dt X<br />

ε )−4 exp<br />

(Xt ε G √ t<br />

)2 h − t<br />

(3.4.7)<br />

0<br />

≤<br />

≤<br />

e 2 (X ε h )−4 exp<br />

( ∫ h<br />

e 2 ( (<br />

Xh<br />

0 ) −4<br />

exp 6<br />

h/2<br />

∫ h<br />

( ) )<br />

dt X<br />

ε<br />

(Xt ε G √ t<br />

)2 h − t<br />

h/2<br />

)<br />

dt<br />

(Xt 0 ,<br />

)2<br />

using the bounds G ≤ 6 and X ε ≥ X 0 . Then, an application of Itô’s formula shows that<br />

( ∫ )<br />

t<br />

(Xt 0 ) −4 dr<br />

exp 6<br />

(Xr 0)2 is a local martingale on the time interval [h/2, ∞[, and so<br />

[ ( ∫ )]<br />

(X ) h<br />

0 −4 dt<br />

[<br />

E h exp 6<br />

(Xt 0 ≤ E (X 0<br />

)2 h/2 )−4] < ∞.<br />

h/2<br />

Tog<strong>et</strong>her with (3.4.7), this shows that the random variables appearing in the left-hand side of<br />

(3.4.6) are uniformly integrable. The convergence (3.4.6) easily follows. □<br />

Proof of Theorem 3.4.1 : We first explain how the first part of Theorem 3.4.1 can be<br />

deduced from Proposition 3.4.2. Recall the notation (N,N ′ ) from the proof of this proposition.<br />

We first observe that we can find a measurable functional Γ such that<br />

h/2<br />

W = Γ(W α , N,N ′ ) ,<br />

L<strong>et</strong> us make this functional more explicit. We have first<br />

α = ∑ j∈J<br />

σ(W j ).<br />

N h 0 a.s.<br />

For every l ∈ [0,h], we s<strong>et</strong><br />

τ l = ∑ j∈J½{h j ≤l} σ(W j ).<br />

Then, if s ∈ [0,α], there is a unique l such that τ l− ≤ s ≤ τ l , and :<br />

• Either there is a (unique) j ∈ J such that l = h j , and<br />

• Or there is no such j, and<br />

ζ s = l + ζ j s−τ l−<br />

,<br />

{<br />

Wα (t) if t ≤ l ,<br />

W s (t) =<br />

W α (l) + W j s−τ l−<br />

(t − l) if l < t ≤ ζ s ;<br />

ζ s = l ,<br />

W s (t) = W α (t) , t ≤ l .<br />

The previous formulas identify (W s ,0 ≤ s ≤ α) as a measurable function of the pair (W α , N),<br />

and in a similar way we can recover (W σ−s ,0 ≤ s ≤ σ − α) as the same measurable function of<br />

(W α , N ′ ).<br />

To simplify notation, write N h,(ε) for the conditional probability N h 0 (· | W > −ε). From<br />

elementary properties of Poisson measures, we g<strong>et</strong> that under the probability measure N h,(ε)<br />

0<br />

79


and conditionally given W α , the point measures N and N ′ are independent and Poisson with<br />

intensity<br />

µ h ε(W α ;dr dω) := 2½[0,h](r)½[0,h−r](H(ω))½{R(ω)⊂]−ε−W α(r),∞[} dr N 0 (dω).<br />

As a consequence of Proposition 3.4.2, the law of W α under N h,(ε)<br />

0 converges as ε → 0 to the law<br />

of the process Y h = (Yt h ,0 ≤ t ≤ h) such that<br />

[<br />

(3.4.8) E Φ<br />

(Y h)] [<br />

(∫ h<br />

( ))]<br />

= h2<br />

E (9)<br />

0 Φ(R t ,0 ≤ t ≤ h)R −4 dt Rt<br />

c<br />

h<br />

exp G √ .<br />

0 h − t<br />

Suppose that on the same probability space where Y h is defined, we are also given two random<br />

point measures M and M ′ on R + × Ω, which conditionally given Y h are independent Poisson<br />

point measures with intensity<br />

(3.4.9) µ h 0 (Y h ;dr dω) := 2½[0,h](r)½[0,h−r](H(ω))½{R(ω)⊂]−Yr h ,∞[} dr N 0(dω).<br />

From the continuity properties of the “reconstruction mapping” Γ, it should now be clear that<br />

the probability measures N h,(ε) converge as ε → 0 to the measure N h 0 defined as the law of<br />

Γ(Y h , M, M ′ ). Here we leave some easy technical d<strong>et</strong>ails to the reader.<br />

L<strong>et</strong> us prove the second assertion of Theorem 3.4.1. L<strong>et</strong> us fix s 1 > 0, and l<strong>et</strong> ψ be a<br />

continuous function on R + with compact support contained in ]0, ∞[. L<strong>et</strong> F be a bounded<br />

continuous function on Ω. It follows from Theorem 3.3.1 that<br />

(<br />

lim<br />

ε→0 ε−4 N 0 ψ(ζ s1 )F(W)½{W>−ε}<br />

)<br />

= 2<br />

21 N 0<br />

= 2<br />

(<br />

σ −1 ψ<br />

0<br />

R 2 t<br />

( )<br />

ζ s [s∗]<br />

1<br />

F<br />

(W [s∗]))<br />

(3.4.10)<br />

21 N 0(ψ(ζ s1 )F(W)).<br />

To see this, apply Theorem 3.3.1 with a function ϕ such that sϕ(s) vanishes on a neighborhood<br />

of 0 and is identically equal to 1 on [s 1 , ∞[.<br />

On the other hand, we have also<br />

) ∫ ∞<br />

ε −4 N 0<br />

(ψ(ζ s1 )F(W)½{W>−ε} = ε −4 dh<br />

(<br />

)<br />

2h 2 Nh 0 ψ(ζ s1 )F(W)½{W >−ε}<br />

(3.4.11)<br />

=<br />

∫ ∞<br />

0<br />

0<br />

dh<br />

2h 2 ε−4 N h 0 (W > −ε) × Nh,(ε) 0 (ψ(ζ s1 )F(W)).<br />

We pass to the limit ε → 0 in the right-hand side of (3.4.11), using (3.4.2) and the first assertion<br />

of Theorem 3.4.1, which gives<br />

lim<br />

ε→0 Nh,(ε) 0 (ψ(ζ s1 )F(W)) = N h 0 (ψ(ζ s 1<br />

)F(W)).<br />

To justify dominated convergence, first note that<br />

(3.4.12) ε −4 N h 0 (W > −ε) = ε−4 N 1 0<br />

(<br />

W > − ε √<br />

h<br />

)<br />

≤ C h 2.<br />

Furthermore, by comparing the intensity measures in (3.4.3) and (3.4.9), we g<strong>et</strong> that the distribution<br />

of σ under N h,(ε)<br />

0 is stochastically bounded by the distribution of σ under N h 0<br />

(<br />

. )<br />

Hence,<br />

N h,(ε)<br />

0 (ψ(ζ s1 )F(W)) ≤ C ′ N h,(ε)<br />

0 (σ > s 1 ) ≤ C ′ N h 0(σ > s 1 ) ≤ C (s1 ) exp − C′ (s 1 )<br />

h 2 ,<br />

where C (s1 ) and C ′ (s 1 ) are positive constants depending on s 1.<br />

80


The previous observations allow us to apply the dominated convergence theorem to the righthand<br />

side of (3.4.11), and to g<strong>et</strong><br />

(<br />

)<br />

lim ε −4 N 0 ψ(ζ s1 )F(W)½{W>−ε} = c ∫ ∞<br />

0 dh<br />

ε↓0 2 h 4 Nh 0 (ψ(ζ s 1<br />

)F(W)).<br />

Comparing with (3.4.10) now compl<strong>et</strong>es the proof.<br />

At this point we have obtained two distinct descriptions of N 0 :<br />

• The law of σ under N 0 has density (8π) −1/2 s −5/2 , and the conditional distribution<br />

N 0 (· | σ = s) is the law under N (s)<br />

0 of the re-rooted snake W [s∗] .<br />

• The law of H under N 0 has density 21c 0<br />

4<br />

h −4 , and the conditional distribution N 0 (· |<br />

H = h) can be reconstructed from the “spine” Y h and the Poisson point measures M<br />

and M ′ as explained in the proof of Theorem 3.4.1.<br />

If we think of analogous results for the Itô measure of Brownian excursions, it is tempting<br />

to look for a more Markovian description of N 0 . It is relatively easy to see that the process<br />

((ζ s ,W s ),s > 0) is Markovian under N 0 , and to describe its transition kernels (informally, this<br />

is the Brownian snake conditioned not to exit ]0, ∞[ – compare with [2]). One would then like<br />

to have an explicit formula for entrance laws, that is for the law of (ζ s ,W s ) under N 0 , for each<br />

fixed s > 0. Such explicit expressions seem difficult to obtain. See however the calculations in<br />

Section 3.5.<br />

In the final part of this section, we investigate the limiting behavior of the measures N 0 (· |<br />

H = h) as h → ∞. This leads to a (one-dimensional) Brownian snake conditioned to stay<br />

positive and to live forever. The motivation for introducing such a process comes from the fact<br />

that it is expected to appear in scaling limits of discr<strong>et</strong>e trees coding random quadrangulations :<br />

See the recent work of Chassaing and Durhuus [13].<br />

Before stating our result, we give a description of the limiting process. L<strong>et</strong> Z = (Z t ,t ≥ 0)<br />

be a nine-dimensional Bessel process started at 0. Conditionally given Z, l<strong>et</strong><br />

P = ∑ i∈I<br />

be a Poisson point measure on R + × Ω with intensity<br />

δ (hi ,ω i )<br />

2½{R(ω)⊂]−Z r,∞[} dr N 0 (dω).<br />

We may and will assume that P is constructed in the following way. Start from a Poisson point<br />

measure<br />

Q = ∑ δ (hj ,ω j )<br />

j∈J<br />

with intensity 2 dr N 0 (dω), and assume that Q is independent of Z. Then s<strong>et</strong><br />

P = ∑ j∈J½{R(ω j )⊂]−Z hj ,∞[} δ (hj ,ω j ) .<br />

0<br />

□<br />

We then construct our conditioned snake W ∞ from the pair (Z, P). This is very similar to<br />

the reconstruction mapping that was already used in the proof of Theorem 3.4.1. To simplify<br />

notation, we put<br />

σ i = σ(ω i ) , ζ i s = ζ s(ω i ) , W i s = W s(ω i )<br />

81


for every i ∈ I and s ≥ 0. For every l ∈ [0,h], we s<strong>et</strong><br />

τ l = ∑ i∈I½{h i ≤l} σ i .<br />

Then, if s ≥ 0, there is a unique l such that τ l− ≤ s ≤ τ l , and :<br />

• Either there is a (unique) i ∈ I such that l = h i , and we s<strong>et</strong><br />

ζs ∞ = l + ζs−τ i l−<br />

,<br />

{<br />

Ws ∞ Zt if t ≤ l ,<br />

(t) =<br />

Z l + Ws−τ i l−<br />

(t − l) if l < t ≤ ζs ∞ ;<br />

• Or there is no such i, and we s<strong>et</strong><br />

ζ ∞ s = l ,<br />

W ∞ s (t) = Z t , t ≤ l .<br />

It is easy to verify that these prescriptions define a continuous process W ∞ with values in W.<br />

We denote by N ∞ 0 the law of W ∞ .<br />

Theorem 3.4.3. The probability measures N h 0 converge to N ∞ 0 when h → ∞.<br />

Proof : We rely on the explicit description of N h 0 obtained in the proof of Theorem 3.4.1.<br />

L<strong>et</strong> Y h = (Y h<br />

t ,0 ≤ t ≤ h) be as in (3.4.8).<br />

Lemma 3.4.4. The processes (Yt∧h h ,t ≥ 0) converge in distribution to Z as h → ∞.<br />

Proof : L<strong>et</strong> A > 0 and l<strong>et</strong> Φ be a bounded continuous function on C([0,A], R + ). By (3.4.8),<br />

if h ≥ A,<br />

[<br />

] [<br />

(∫ h<br />

( ))]<br />

E Φ(Yt h ,0 ≤ t ≤ A) = h2<br />

E (9)<br />

0 Φ(R t ,0 ≤ t ≤ A)R −4 dt Rt<br />

c<br />

h<br />

exp<br />

0 0 Rt<br />

2 G √ . h − t<br />

We apply the Markov property at time A in the right-hand side, and write h = A+a to simplify<br />

notation :<br />

[<br />

[<br />

] (∫ A<br />

E Φ(Yt h ,0 ≤ t ≤ A) = h2<br />

E (9)<br />

dt<br />

0 Φ(R t ,0 ≤ t ≤ A) exp<br />

c 0<br />

(3.4.13)<br />

From the bound 0 ≤ G(x) ≤ 2x 2 , it is immediate that<br />

(∫ A<br />

dt<br />

(3.4.14) 1 ≤ exp G<br />

0<br />

R 2 t<br />

On the other hand, a scaling argument gives<br />

h 2 [ (∫ a<br />

E (9)<br />

c<br />

R A<br />

Ra<br />

−4 dt<br />

exp<br />

0 0 Rt<br />

2 G<br />

( ) h 2<br />

= c −1<br />

0 E (9)<br />

a R A / √ a<br />

0<br />

R 2 t<br />

[ (∫ a<br />

× E (9)<br />

R A<br />

Ra<br />

−4 exp<br />

(<br />

Rt<br />

√<br />

h − t<br />

))<br />

≤<br />

(<br />

Rt<br />

√ a − t<br />

))]<br />

[ (∫ 1<br />

R1 −4 exp<br />

0<br />

dt<br />

R 2 t<br />

0<br />

(<br />

1 + A a<br />

) 2<br />

.<br />

( ))<br />

Rt<br />

G √<br />

h − t<br />

dt<br />

R 2 t<br />

( ))]<br />

Rt<br />

G √ . 1 − t<br />

( ))] ]<br />

Rt<br />

G √ . a − t<br />

82


From (3.4.6), we know that<br />

(3.4.15)<br />

[ (∫ 1<br />

lim E x<br />

(9) R1 −4 dt<br />

exp<br />

x↓0 0<br />

R 2 t<br />

( ))] [ (∫ 1<br />

Rt<br />

G √ = E (9)<br />

1 − t<br />

0 R1 −4 exp<br />

0<br />

dt<br />

R 2 t<br />

( ))]<br />

Rt<br />

G √ = c 0 .<br />

1 − t<br />

We can use (3.4.14) and (3.4.15) to pass to the limit h → ∞ in the right-hand side of (3.4.13).<br />

The justification of dominated convergence is easy thanks to the bounds we obtained when<br />

proving (3.4.6). It follows that<br />

lim E[Φ(Y t h ,0 ≤ t ≤ A)] = E (9)<br />

0 [Φ(R t,0 ≤ t ≤ A)]<br />

h→∞<br />

which was the desired result.<br />

We can now compl<strong>et</strong>e the proof of Theorem 3.4.3. By Lemma 3.4.4 and the Skorokhod<br />

representation theorem, we may assume that (Yt∧h h ) t≥0 converges to (Z t ) t≥0 uniformly on every<br />

compact subs<strong>et</strong> of R + , a.s.<br />

Recall the description of N h 0 as the law of Γ(Y h , M, M ′ ) in the proof of Theorem 3.4.1 :<br />

According to this description, we can construct a process (W h s ) s≤αh having the distribution of<br />

(W s ) s≤α under N h 0, by the same formulas we used to define W ∞ from the pair (Z, P), provided<br />

that Z is replaced by Y h , the point measure P is replaced by<br />

N h := ∑ j∈J½{R(ω j )⊂]−Y h h j<br />

,∞[}½{h j


and then, for every u ∈ T \{∅}, constructing ξ u as the solution of<br />

⎧<br />

⎪⎨ dξ u t = dξt u + 4<br />

ξ u dt , 0 ≤ t ≤ h u ,<br />

t<br />

⎪⎩ ξ u 0 = ξ π(u)<br />

h π(u)<br />

.<br />

We then define (V a ,a ∈ ˜θ) by the formula V pθ (u,l) = ξ u l for every u ∈ T and l ∈ [0,h u ]. Finally,<br />

for every leaf a of ˜θ, we define the stopped path w (a) from (V a ,a ∈ ˜θ) in the same way as w (a)<br />

was defined from (V a ,a ∈ ˜θ). Recall the notation L(θ) for the s<strong>et</strong> of leaves of ˜θ, and I(θ) for the<br />

s<strong>et</strong> of its nodes.<br />

Theorem 3.5.1. L<strong>et</strong> p ≥ 1 be an integer. L<strong>et</strong> F be a symm<strong>et</strong>ric nonnegative measurable<br />

function on W p . Then,<br />

N 0<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F ( W s1 ,...,W sp<br />

) )<br />

∫<br />

= 2 p−1 p!<br />

⎡<br />

( ( ) ∏<br />

Λ p (dθ)Q θ 0<br />

⎣F w (a)) a∈L(θ)<br />

a∈I(θ)<br />

(<br />

V a<br />

) 4<br />

∏<br />

a∈L(θ)<br />

(<br />

V a<br />

) −4<br />

⎤<br />

Proof : We may assume that F is continuous and bounded above by 1, and that there exist<br />

positive constants δ and M such that F(w 1 ,... ,w p ) = 0 as soon as ζ (wi ) /∈ [δ,M] for some i.<br />

The proof will be divided in several steps.<br />

Step 1. To simplify notation, we write R(V ) = {V a ,a ∈ ˜θ}, for the range of V , or equivalently<br />

for the union of the ranges of w (a) for a ∈ L(θ). We first apply Theorem 3.2.2 to compute<br />

)<br />

N 0<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F ( W s1 ,...,W sp<br />

)½{R⊂]−ε,∞[}<br />

= p!2 p−1 ∫<br />

= p!2 p−1 ∫<br />

= p!2 p−1 ∫<br />

[<br />

(<br />

Λ p (dθ)Q θ 0 F((w (a) ) a∈L(θ) )½{R(V )⊂]−ε,∞[} exp<br />

[<br />

Λ p (dθ)Q θ 0 F((w (a) ) a∈L(θ) )½{R(V )⊂]−ε,∞[} exp<br />

( ( (<br />

Λ p (dθ)Q θ ε F −ε + w (a)) a∈L(θ)<br />

We then use Proposition 3.2.6 inductively to see that<br />

[ ( (<br />

ε −4 Q θ ε F −ε + w (a)) a∈L(θ)<br />

⎡<br />

( ( ) ∏<br />

= Q θ ε<br />

⎣F −ε + w (a)) a∈L(θ)<br />

∫<br />

− 4<br />

∫<br />

(<br />

−6<br />

)½{R(V )⊂]0,∞[} exp<br />

(<br />

−6<br />

)½{R(V )⊂]0,∞[} exp<br />

(<br />

−6<br />

a∈I(θ)<br />

(<br />

V a<br />

) 4<br />

∏<br />

⎦.<br />

)]<br />

L θ (da) N 0 (r ⊂] − ε − V a , ∞[)<br />

)]<br />

L θ (da)<br />

(V a + ε) 2<br />

a∈L(θ)<br />

∫ )]<br />

Lθ (da)<br />

(V a ) 2 .<br />

∫ )]<br />

Lθ (da)<br />

(V a ) 2<br />

(<br />

V a<br />

) −4<br />

⎤<br />

⎦ .<br />

84


We have thus proved that<br />

)<br />

(3.5.1)<br />

ε −4 N 0<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F ( W s1 ,... ,W sp<br />

)½{W>−ε}<br />

⎡<br />

∫ ( (<br />

= p!2 p−1 Λ p (dθ)Q θ ε<br />

⎣F −ε + w (a)) ) ∏<br />

a∈L(θ)<br />

a∈I(θ)<br />

⎤<br />

( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

⎦.<br />

a∈L(θ)<br />

Step 2. We focus on the right-hand side of (3.5.1). Our goal is to prove that<br />

⎡<br />

⎤<br />

∫ ( ( ) ∏<br />

lim Λ p (dθ)Q θ ε<br />

⎣F −ε + w (a)) ( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

⎦<br />

ε→0<br />

a∈L(θ)<br />

a∈I(θ) a∈L(θ)<br />

⎡<br />

⎤<br />

∫ ( ( ) ∏<br />

= Λ p (dθ)Q θ 0<br />

⎣F −ε + w (a)) ( ) 4<br />

∏ ( ) −4<br />

(3.5.2)<br />

V a V a<br />

⎦ .<br />

a∈L(θ)<br />

We first state a lemma.<br />

Lemma 3.5.2. We have<br />

⎡<br />

Q θ ε<br />

⎣ ∏<br />

a∈I(θ)<br />

where D(θ) = max{d θ (0,a) : a ∈ L(θ)}.<br />

(<br />

V a<br />

) 4<br />

∏<br />

a∈L(θ)<br />

(<br />

V a<br />

) −4<br />

⎤<br />

a∈I(θ)<br />

⎦ ≤ E (9)<br />

ε<br />

[ ]<br />

R −4<br />

D(θ)<br />

a∈L(θ)<br />

Proof : We argue by induction on p. If p = 1, the result is immediate, with an equality. L<strong>et</strong><br />

p ≥ 2 and l<strong>et</strong> us assume that the result holds at order 1,2,... ,p − 1. L<strong>et</strong> θ = (T ,(h u ,u ∈ T ))<br />

be a marked tree with p leaves. Write h = h ∅ . By decomposing θ at its first branching point,<br />

we g<strong>et</strong> two marked trees θ ′ ∈ T j , and θ ′′ ∈ T p−j , for some j ∈ {1,... ,p − 1}, in such a way that<br />

⎡<br />

⎤<br />

⎣ ∏ ( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

⎦<br />

Q θ ε<br />

a∈I(θ)<br />

= E (9)<br />

ε<br />

⎡<br />

a∈L(θ)<br />

⎡<br />

⎣R h 4 ⎣ ∏<br />

Qθ′ R h<br />

a∈I(θ ′ )<br />

[ [ ]<br />

≤ E ε<br />

(9) Rh 4 E(9) R h<br />

R −4<br />

D(θ ′ )<br />

(<br />

V a<br />

) 4<br />

∏<br />

a∈L(θ ′ )<br />

[ ]]<br />

E (9)<br />

R h<br />

R −4<br />

D(θ ′′ )<br />

(<br />

V a<br />

) −4<br />

⎤<br />

.<br />

⎡<br />

⎦Q θ′′ ⎣ ∏<br />

R h<br />

a∈I(θ ′′ )<br />

(<br />

V a<br />

) 4<br />

∏<br />

a∈L(θ ′′ )<br />

⎤⎤<br />

( ) −4<br />

V a<br />

⎦⎦<br />

We have used the induction hypothesis in the last inequality. We now observe that D(θ) =<br />

h+max{D(θ ′ ),D(θ ′′ )}. Assume for definiteness that D(θ) = h+D(θ ′ ). Using the bound (3.2.4)<br />

and the Markov property we g<strong>et</strong><br />

E (9)<br />

ε<br />

[ [<br />

Rh 4 E(9) R h<br />

R −4<br />

D(θ ′ )<br />

]<br />

[<br />

E (9)<br />

R h<br />

R −4<br />

D(θ ′′ )<br />

]]<br />

This compl<strong>et</strong>es the proof of the lemma.<br />

[ ]]<br />

≤ E ε<br />

[E (9)<br />

R 9 h<br />

R −4<br />

D(θ ′ )<br />

[ ]<br />

= E ε<br />

(9) R −4<br />

h+D(θ ′ )<br />

[ ]<br />

= E ε<br />

(9) R −4<br />

D(θ)<br />

.<br />

□<br />

85


Q θ ε<br />

As a consequence of Lemma 3.5.2, we g<strong>et</strong> the bound<br />

⎡<br />

⎤<br />

( (<br />

⎣F −ε + w (a)) ) ∏ ( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

a∈L(θ)<br />

a∈I(θ)<br />

a∈L(θ)<br />

⎦ ≤ E (9)<br />

ε<br />

[ ] ∏<br />

R −4<br />

D(θ)<br />

[ ]<br />

≤ E (9) ∏<br />

0 R −4<br />

D(θ)<br />

= E(9) 0<br />

[ ]<br />

R<br />

−4<br />

1<br />

D(θ) 2<br />

a∈L(θ)½{δ≤d θ (0,a)≤M}<br />

a∈L(θ)½{δ≤d θ (0,a)≤M}<br />

∏<br />

a∈L(θ)½{δ≤d θ (0,a)≤M}.<br />

The last quantity is clearly integrable with respect to the measure Λ p (dθ). In addition, using<br />

the continuity of F, it is easy to verify that<br />

⎡<br />

⎤<br />

( ( ) ∏<br />

lim<br />

ε→0 Qθ ε<br />

⎣F −ε + w (a)) ( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

⎦<br />

a∈L(θ)<br />

a∈I(θ) a∈L(θ)<br />

⎡<br />

⎤<br />

( (<br />

= Q θ 0<br />

⎣F w (a)) ) ∏ ( ) 4<br />

∏ ( ) −4<br />

V a V a<br />

⎦<br />

a∈L(θ)<br />

a∈I(θ)<br />

a∈L(θ)<br />

An application of the dominated convergence theorem now leads to (3.5.2).<br />

Step 3. We now consider the left-hand side of formula (3.5.1). For every 0 < b < 1, we consider<br />

the continuous function φ b : R + −→ [0,1] such that φ b (s) = 1 for every s ∈ [b,1/b], φ b (s) = 0<br />

for every s ∈ R + \]b/2,2/b[, and φ b is linear on [b/2,b] and on [1/b,2/b]. From Theorem 3.3.1<br />

and the definition of N 0 , we g<strong>et</strong><br />

( ∫<br />

lim<br />

ε→0 ε−4 N 0 φ b (σ) ds 1 ... ds p F ( )<br />

W s1 ,... ,W sp<br />

)½{W>−ε}<br />

]0,σ[ p ∫<br />

= N 0<br />

(φ b (σ) ds 1 ...ds p F ( ) )<br />

(3.5.3)<br />

W s1 ,...,W sp .<br />

]0,σ[ p<br />

Lemma 3.5.3. The following convergence holds :<br />

∫<br />

lim sup ε −4 N 0<br />

((1 − φ b (σ)) ds 1 ...ds p F ( )<br />

W s1 ,... ,W sp<br />

)½{W>−ε} = 0.<br />

b→0 ε∈(0,1)<br />

]0,σ[ p<br />

Proof : We first observe that<br />

∫<br />

ε −4 N 0<br />

(½{σ−ε}<br />

]0,σ[<br />

(<br />

p ) ∫ ∞<br />

≤ b p ε −4 N 0 sup ζ s > δ , W > −ε = b p ε −4 dh<br />

(3.5.4)<br />

s∈[0,σ]<br />

2h 2 Nh 0(W > −ε) ≤ b p C<br />

6δ 3<br />

δ<br />

86


where the constant C is such that ε −4 N h 0 (W > −ε) ≤ Ch−2 , for every h > 0 and 0 < ε < 1 (cf<br />

(3.4.12)). On the other hand, the Cauchy-Schwarz inequality gives<br />

∫<br />

ε −4 N 0<br />

(½{σ>1/b} ds 1 ...ds p F ( )<br />

W s1 ,... ,W sp<br />

)½{W>−ε}<br />

]0,σ[ p<br />

⎛ ⎛( ∫<br />

≤ ⎝ε −4 N 0<br />

⎝ ds 1 ...ds p F ( ) ) ⎞<br />

2½{W>−ε}⎞ 1/2<br />

W s1 ,... ,W sp<br />

⎠⎠<br />

]0,σ[ p<br />

×<br />

(ε −4 N 0<br />

(<br />

σ > 1 b ,W > −ε )) 1/2<br />

.<br />

Note that we may write<br />

⎛( ∫<br />

N 0<br />

⎝ ds 1 ...ds p F ( ) ) 2½{W>−ε}⎞<br />

W s1 ,... ,W sp<br />

⎠<br />

]0,σ[ p<br />

= N 0<br />

( ∫<br />

]0,σ[ 2p ds 1 ...ds 2p G ( W s1 ,...,W s2p<br />

)½{W>−ε}<br />

where G is a nonnegative symm<strong>et</strong>ric function on W 2p , which is also bounded by 1. As a consequence<br />

of (3.5.1) and Lemma 3.5.2, we then g<strong>et</strong><br />

⎛( ∫<br />

ε −4 N 0<br />

⎝ ds 1 ...ds p F ( ) ) 2½{W>−ε}⎞<br />

W s1 ,... ,W sp<br />

⎠<br />

]0,σ[ p<br />

≤ (2p)!2 2p−1 E (9)<br />

0<br />

[ ] ∫ ∏<br />

R<br />

−4<br />

1 Λ 2p (dθ)D(θ) −2 θ (∅,a)≤M} = C(p,δ,M) < ∞.<br />

a∈L(θ)½{δ≤d<br />

From Theorem 3.1.1 and a simple scaling argument, we have<br />

(<br />

ε −4 N 0 σ > 1 ) ∫ ∞<br />

b , W > −ε = ε −4 ds<br />

√<br />

2πs<br />

3 N(s) 0 (W > −ε) ≤ C′ b 1/2 .<br />

By combining these estimates, we g<strong>et</strong><br />

(3.5.5)<br />

∫<br />

ε −4 N 0<br />

(½{σ>1/b} ds 1 ...ds p F ( )<br />

W s1 ,... ,W sp<br />

)½{W>−ε} ≤ (C ′ C(p,δ,M)) 1/2 b 1/4 .<br />

]0,σ[ p<br />

Lemma 3.5.3 follows from (3.5.4) and (3.5.5).<br />

b −1<br />

We can now compl<strong>et</strong>e the proof of Theorem 3.5.1. First, by monotone convergence,<br />

( ∫<br />

lim N 0 φ b (σ) ds 1 ...ds p F ( ) ) ( ∫<br />

W s1 ,... ,W sp = N 0 ds 1 ...ds p F ( ) )<br />

W s1 ,...,W sp .<br />

b→0 ]0,σ[ p ]0,σ[ p<br />

)<br />

,<br />

□<br />

87


From (3.5.3) and Lemma 3.5.3, it then follows that<br />

lim<br />

ε→0 ε−4 N 0<br />

( ∫<br />

]0,σ[ p ds 1 ...ds p F ( W s1 ,... ,W sp<br />

)½{W>−ε}<br />

= N 0<br />

( ∫<br />

]0,σ[ p ds 1 ... ds p F ( W s1 ,... ,W sp<br />

) ) .<br />

)<br />

Combining this with (3.5.1) and (3.5.2) gives Theorem 3.5.1.<br />

□<br />

Acknowledgement. The first author wishes to thank Philippe Chassaing for a stimulating<br />

conversation which motivated the present work. We also thank the referee for several useful<br />

remarks.<br />

88


CHAPITRE 4<br />

Asymptotics for rooted planar maps and scaling limits of<br />

two-type spatial trees<br />

4.1. Introduction<br />

The main goal of the present work is to investigate asymptotic properties of large rooted<br />

bipartite planar maps under the so-called Boltzmann distributions. This s<strong>et</strong>ting includes as a<br />

special case the asymptotics as n → ∞ of the uniform distribution over rooted 2κ-angulations<br />

with n faces, for any fixed integer κ ≥ 2. Boltzmann distributions over planar maps that are both<br />

rooted and pointed have been considered recently by Marckert & Miermont [43] who discuss<br />

in particular the profile of distances from the distinguished point in the map. Here we deal<br />

with rooted maps and we investigate distances from the root vertex, so that our results do not<br />

follow from the ones in [43], although many statements look similar. The specific results that<br />

are obtained in the present work have found applications in the paper [40], which investigates<br />

scaling limits of large planar maps.<br />

L<strong>et</strong> us briefly discuss Boltzmann distributions over rooted bipartite planar maps. We consider<br />

a sequence q = (q i ) i≥1 of weights (nonnegative real numbers) satisfying certain regularity<br />

properties. Then, for each integer n ≥ 2, we choose a random rooted bipartite map M n with<br />

n faces whose distribution is specified as follows : the probability that M n is equal to a given<br />

bipartite planar map m is proportional to<br />

n∏<br />

i=1<br />

q deg(fi )/2<br />

where f 1 ,...,f n are the faces of m and deg(f i ) is the degree (that is the number of adjacent<br />

edges) of the face f i . In particular we may take q κ = 1 and q i = 0 for i ≠ κ, and we g<strong>et</strong> the<br />

uniform distribution over rooted 2κ-angulations with n faces.<br />

Theorem 4.2.5 below provides asymptotics for the radius and the profile of distances from<br />

the root vertex in the random map M n when n → ∞. The limiting distributions are described<br />

in terms of the one-dimensional Brownian snake driven by a normalized excursion. In particular<br />

if R n denotes the radius of M n (that is the maximal distance from the root), then n −1/4 R n<br />

converges to a multiple of the range of the Brownian snake. In the special case of quadrangulations<br />

(q 2 = 1 and q i = 0 for i ≠ 2), these results were obtained earlier by Chassaing & Schaeffer<br />

[14]. As was mentioned above, very similar results have been obtained by Marckert & Miermont<br />

[43] in the s<strong>et</strong>ting of Boltzmann distributions over rooted pointed bipartite planar maps, but<br />

considering distances from the distinguished point rather than from the root.<br />

Similarly as in [14] or [43], bijection b<strong>et</strong>ween trees and maps serve as a major tool in our<br />

approach. In the case of quadrangulations, these bijections were studied by Cori & Vauquelin<br />

[15] and then by Schaeffer [51]. They have been recently extended to bipartite planar maps by<br />

Bouttier, di Francesco & Guitter [11]. More precisely, Bouttier, di Francesco & Guitter show that<br />

89


ipartite planar maps are in one-to-one correspondence with well-labelled mobiles, where a welllabelled<br />

mobile is a two-type spatial tree whose vertices are assigned positive labels satisfying<br />

certain compatibility conditions (see section 4.2.4 for a precise definition). This bijection has the<br />

nice feature that labels in the mobile correspond to distances from the root in the map. Then<br />

the above mentioned asymptotics for random maps reduce to a limit theorem for well-labelled<br />

mobiles, which is stated as Theorem 4.3.3 below. This statement can be viewed as a conditional<br />

version of Theorem 11 in [43]. The fact that [43] deals with distances from the distinguished<br />

point in the map (rather than from the root) makes it possible there to drop the positivity<br />

constraint on labels. In the present work this constraint makes the proof significantly more<br />

difficult. We rely on some ideas from Le Gall [39] who established a similar conditional theorem<br />

for well-labelled trees. Although many arguments in Section 4.3 below are analogous to the ones<br />

in [39], there are significant additional difficulties because we deal with two-type trees and we<br />

condition on the number of vertices of type 1 rather than on the total number of vertices.<br />

A key step in the proof of Theorem 4.3.3 consists in the derivation of estimates for the probability<br />

that a two-type spatial tree remains on the positive half-line. As another application of<br />

these estimates, we derive some information about separating vertices of uniform 2κ-angulations.<br />

We show that with a probability close to one when n → ∞ a random rooted 2κ-angulation with<br />

n faces will have a vertex whose removal disconnects the map into two components both having<br />

size greater that n 1/2−ε . Related combinatorial results are obtained in [8]. More precisely, in<br />

a vari<strong>et</strong>y of different models, Proposition 5 in [8] asserts that the second largest nonseparable<br />

component of a random map of size n has size at most O(n 2/3 ). This suggests that n 1/2−ε in<br />

our result could be replaced by n 2/3−ε .<br />

The paper is organized as follows. In section 4.2, we recall some preliminary results and we<br />

state our asymptotics for large random rooted planar maps. Section 4.3 is devoted to the proof of<br />

Theorem 4.3.3 and to the derivation of Theorem 4.2.5 from asymptotics for well-labelled mobiles.<br />

Finally Section 4.4 discusses the application to separating vertices of uniform 2κ-angulations.<br />

4.2. Preliminaries<br />

4.2.1. Boltzmann laws on planar maps. A planar map is a proper embedding, without<br />

edge crossings, of a connected graph in the 2-dimensional sphere S 2 . Loops and multiple edges<br />

are allowed. A planar map is said to be bipartite if all its faces have even degree. In this paper,<br />

we will only be concerned with bipartite maps. The s<strong>et</strong> of vertices will always be equipped with<br />

the graph distance : if a and a ′ are two vertices, d(a,a ′ ) is the minimal number of edges on a<br />

path from a to a ′ . If M is a planar map, we write F M for the s<strong>et</strong> of its faces, and V M for the<br />

s<strong>et</strong> of its vertices.<br />

A pointed planar map is a pair (M,τ) where M is a planar map and τ is a distinguished<br />

vertex. Note that, since M is bipartite, if a and a ′ are two neighbouring vertices, then we have<br />

|d(τ,a) − d(τ,a ′ )| = 1. A rooted planar map is a pair (M,⃗e ) where M is a planar map and ⃗e is<br />

a distinguished oriented edge. The origin of ⃗e is called the root vertex. At last, a rooted pointed<br />

planar map is a triple (M,e,τ) where (M,τ) is a pointed planar map and e is a distinguished<br />

non-oriented edge. We can always orient e in such a way that its origin a and its end point<br />

a ′ satisfy d(τ,a ′ ) = d(τ,a) + 1. Note that a rooted planar map can be interpr<strong>et</strong>ed as a rooted<br />

pointed planar map by choosing the root vertex as the distinguished point.<br />

Two pointed maps (resp. two rooted maps, two rooted pointed maps) are identified if there<br />

exists an orientation-preserving homeomorphism of the sphere that sends the first map to the<br />

second one and preserves the distinguished point (resp. the root edge, the distinguished point<br />

90


and the root edge). L<strong>et</strong> us denote by M p (resp. M r , M r,p ) the s<strong>et</strong> of all pointed bipartite maps<br />

(resp. the s<strong>et</strong> of all rooted bipartite maps, the s<strong>et</strong> of all rooted pointed bipartite maps) up to<br />

the preceding indentification.<br />

L<strong>et</strong> us recall some definitions and propositions that can be found in [43]. L<strong>et</strong> q = (q i ,i ≥ 1)<br />

be a sequence of nonnegative weights such that q i > 0 for at least one i > 1. For any planar<br />

map M, we define W q (M) by<br />

W q (M) = ∏<br />

q deg(f)/2 ,<br />

f∈F M<br />

where we have written deg(f) for the degree of the face f. We require q to be admissible that is<br />

Z q =<br />

∑<br />

W q (M) < ∞.<br />

M∈M r,p<br />

Note that the sum is over the s<strong>et</strong> M r,p of all rooted pointed bipartite planar maps which is<br />

countable thanks to the identification that was explained above. For k ≥ 1, we s<strong>et</strong> N(k) = ( )<br />

2k−1<br />

k−1 .<br />

For every weight sequence q, we define<br />

f q (x) = ∑ k≥0<br />

N(k + 1)q k+1 x k , x ≥ 0.<br />

L<strong>et</strong> R q be the radius of convergence of the power series f q . Consider the equation<br />

(4.2.1) f q (x) = 1 − x −1 , x > 0.<br />

From Proposition 1 in [43], a sequence q is admissible if and only if equation (4.2.1) has at least<br />

one solution, and then Z q is the solution of (4.2.1) that satisfies Z 2 q f ′ q (Z q) ≤ 1. An admissible<br />

weight sequence q is said to be critical if it satisfies<br />

(Z q ) 2 f ′ q (Z q ) = 1,<br />

which means that the graphs of the functions x ↦−→ f q (x) and x ↦−→ 1 − 1/x are tangent at the<br />

left of x = Z q . Furthermore, if Z q < R q , then q is said to be regular critical. This means that<br />

the graphs are tangent both at the left and at the right of Z q . In what follows, we will only be<br />

concerned with regular critical weight sequences.<br />

L<strong>et</strong> q be a regular critical weight sequence. We define the Boltzmann distribution B r,p<br />

q on the<br />

s<strong>et</strong> M r,p by<br />

B r,p<br />

q (M) = W q(M)<br />

.<br />

Z q<br />

L<strong>et</strong> us now define Z (r)<br />

q<br />

by<br />

Z (r)<br />

q = ∑<br />

q deg(f)/2 .<br />

M∈M r f∈F M<br />

Note that the sum is over the s<strong>et</strong> M r of all rooted bipartite planar maps. From the fact that<br />

Z q < ∞ it easily follows that Z q<br />

(r) < ∞. We then define the Boltzmann distribution B r q on the<br />

s<strong>et</strong> M r by<br />

B r q (M) = W q(M)<br />

Z q<br />

(r) .<br />

∏<br />

91


L<strong>et</strong> us turn to the special case of 2κ-angulations. A 2κ-angulation is a bipartite planar map<br />

such that all faces have a degree equal to 2κ. If κ = 2, we recognize the well-known quadrangulations.<br />

L<strong>et</strong> us s<strong>et</strong><br />

(κ − 1)κ−1<br />

α κ =<br />

κ κ N(κ) .<br />

We denote by q κ the weight sequence defined by q κ = α κ and q i = 0 for every i ∈ N \ {κ}. It is<br />

proved in Section 1.5 of [43] that q κ is a regular critical weight sequence, and<br />

Z qκ =<br />

κ<br />

κ − 1 .<br />

For every n ≥ 1, we denote by U n κ (resp. U n κ ) the uniform distribution on the s<strong>et</strong> of all rooted<br />

pointed 2κ-angulations with n faces (resp. on the s<strong>et</strong> of all rooted 2κ-angulations with n faces).<br />

We have<br />

B r,p<br />

q κ<br />

(· | #F M = n) = U n κ,<br />

B r q κ<br />

(· | #F M = n) = U n κ .<br />

4.2.2. Two-type spatial Galton-Watson trees. We start with some formalism for discr<strong>et</strong>e<br />

trees. S<strong>et</strong><br />

U = ⋃ n≥0<br />

N n ,<br />

where N = {1,2,3,...} and by convention N 0 = {∅}. An element of U is a sequence u = u 1 ...u n ,<br />

and we s<strong>et</strong> |u| = n so that |u| represents the generation of u. In particular, |∅| = 0. If u = u 1 ...u n<br />

and v = v 1 ...v m belong to U, we write uv = u 1 ...u n v 1 ...v m for the concatenation of u and<br />

v. In particular, ∅u = u∅ = u. If v is of the form v = uj for u ∈ U and j ∈ N, we say that v is<br />

a child of u, or that u is the father of v, and we write u = ˇv. More generally if v is of the form<br />

v = uw for u,w ∈ U, we say that v is a descendant of u, or that u is an ancestor of v. The s<strong>et</strong> U<br />

comes with the natural lexicographical order such that u v if either u is an ancestor of v, or if<br />

u = wa and v = wb with a ∈ U ∗ and b ∈ U ∗ satisfying a 1 < b 1 , where we have s<strong>et</strong> U ∗ = U \ {∅}.<br />

And we write u ≺ v if u v and u ≠ v.<br />

A plane tree T is a finite subs<strong>et</strong> of U such that<br />

(i) ∅ ∈ T ,<br />

(ii) u ∈ T \ {∅} ⇒ ǔ ∈ T ,<br />

(iii) for every u ∈ T , there exists a number k u (T ) ≥ 0 such that uj ∈ T if and only if<br />

1 ≤ j ≤ k u (T ).<br />

We denote by T the s<strong>et</strong> of all plane trees.<br />

L<strong>et</strong> T be a plane tree and l<strong>et</strong> ζ = #T − 1. The search-depth sequence of T is the sequence<br />

u 0 ,u 1 ,...,u 2ζ of vertices of T wich is obtained by induction as follows. First u 0 = ∅, and then<br />

for every i ∈ {0,1,... ,2ζ − 1}, u i+1 is either the first child of u i that has not y<strong>et</strong> appeared in<br />

the sequence u 0 ,u 1 ,...,u i , or the father of u i if all children of u i already appear in the sequence<br />

u 0 ,u 1 ,...,u i . It is easy to verify that u 2ζ = ∅ and that all vertices of T appear in the sequence<br />

u 0 ,u 1 ,...,u 2ζ (of course some of them appear more that once). We can now define the contour<br />

function of T . For every k ∈ {0,1,... ,2ζ}, we l<strong>et</strong> C(k) denote the distance from the root of<br />

the vertex u k . We extend the definition of C to the line interval [0,2ζ] by interpolating linearly<br />

b<strong>et</strong>ween successive integers. Clearly T is uniquely d<strong>et</strong>ermined by its contour function C.<br />

A discr<strong>et</strong>e spatial tree is a pair (T ,U) where T ∈ T and U = (U v ,v ∈ T ) is a mapping from<br />

the s<strong>et</strong> T into R. If v is a vertex of T , we say that U v is the label of v. We denote by Ω the s<strong>et</strong> of<br />

92


all discr<strong>et</strong>e spatial trees. If (T ,U) ∈ Ω we define the spatial contour function of (T ,U) as follows.<br />

First if k is an integer, we put V (k) = U uk with the preceding notation. We then compl<strong>et</strong>e the<br />

definition of V by interpolating linearly b<strong>et</strong>ween successive integers. Clearly (T ,U) is uniquely<br />

d<strong>et</strong>ermined by the pair (C,V ).<br />

L<strong>et</strong> (T ,U) ∈ Ω. We interpr<strong>et</strong> (T ,U) as a two-type (spatial) tree by declaring that vertices of<br />

even generations are of type 0 and vertices of odd generations are of type 1. We then s<strong>et</strong><br />

T 0<br />

T 1<br />

= {u ∈ T : |u| is even},<br />

= {u ∈ T : |u| is odd}.<br />

L<strong>et</strong> us turn to random trees. We want to consider a particular family of two-type Galton-<br />

Watson trees, in which vertices of type 0 only give birth to vertices of type 1 and vice-versa.<br />

L<strong>et</strong> µ = (µ 0 ,µ 1 ) be a pair of offspring distributions, that is a pair of probability distributions<br />

on Z + . If m 0 and m 1 are the respective means of µ 0 and µ 1 we assume that m 0 m 1 ≤ 1 and we<br />

exclude the trivial case µ 0 = µ 1 = δ 1 where δ 1 stands for the Dirac mass at 1. We denote by P µ<br />

the law of a two-type Galton-Watson tree with offspring distribution µ, meaning that for every<br />

t ∈ T,<br />

P µ (t) = ∏ u∈t 0 µ 0 (k u (t)) ∏ u∈t 1 µ 1 (k u (t)) ,<br />

where t 0 (resp. t 1 ) is as above the s<strong>et</strong> of all vertices of t with even (resp. odd) generation. The<br />

fact that this formula defines a probability measure on T is justified in [43].<br />

L<strong>et</strong> us now recall from [43] how one can couple plane trees with a spatial displacement in order<br />

to turn them into random elements of Ω. To this end, l<strong>et</strong> ν0 k,νk 1 be probability distributions on Rk<br />

for every k ≥ 1. We s<strong>et</strong> ν = (ν0 k,νk 1 ) k≥1. For every T ∈ T and x ∈ R, we denote by R ν,x (T ,dU)<br />

the probability measure on R T which is characterized as follows. L<strong>et</strong> (Y u ,u ∈ T ) be a family<br />

of independent random variables such that for u ∈ T with k u (T ) = k, Y u = (Y u1 ,... ,Y uk ) is<br />

distributed according to ν0 k if u ∈ T 0 and according to ν1 k if u ∈ T 1 . We s<strong>et</strong> X ∅ = x and for<br />

every v ∈ T \ {∅},<br />

X v = x + ∑<br />

Y u ,<br />

u∈]∅,v]<br />

where ]∅,v] is the s<strong>et</strong> of all ancestors of v distinct from the root ∅. Then R ν,x (T ,dU) is the law<br />

of (X v ,v ∈ T ). We finally define for every x ∈ R a probability measure P µ,ν,x on Ω by s<strong>et</strong>ting<br />

P µ,ν,x (dT dU) = P µ (dT )R ν,x (T ,dU).<br />

4.2.3. The Brownian snake and the conditioned Brownian snake. L<strong>et</strong> x ∈ R. The<br />

Brownian snake with initial point x is a pair (e,r x ), where e = (e(s),0 ≤ s ≤ 1) is a normalized<br />

Brownian excursion and r x = (r x (s),0 ≤ s ≤ 1) is a real-valued process such that, conditionally<br />

given e, r x is Gaussian with mean and covariance given by<br />

• E[r x (s)] = x for every s ∈ [0,1],<br />

• Cov(r x (s),r x (s ′ )) = inf<br />

s≤t≤s ′ e(t) for every 0 ≤ s ≤ s′ ≤ 1.<br />

We know from [37] that r x admits a continuous modification. From now on we consider only this<br />

modification. In the terminology of [37] r x is the terminal point process of the one-dimensional<br />

Brownian snake driven by the normalized Brownian excursion e and with initial point x.<br />

93


Write P for the probability measure under which the collection (e,r x ) x∈R is defined. Note<br />

that for every x > 0, we have ( )<br />

P inf<br />

s∈[0,1] rx (s) ≥ 0 > 0.<br />

We may then define for every x > 0 a pair (e x ,r x ) which is distributed as the pair (e,r x ) under<br />

the conditioning that inf s∈[0,1] r x (s) ≥ 0.<br />

We equip C([0,1], R) 2 with the norm ‖(f,g)‖ = ‖f‖ u ∨ ‖g‖ u where ‖f‖ u stands for the<br />

supremum norm of f. The following theorem is a consequence of Theorem 3.1.1.<br />

Theorem 4.2.1. There exists a pair (e 0 ,r 0 ) such that (e x ,r x ) converges in distribution as<br />

x ↓ 0 towards (e 0 ,r 0 ).<br />

The pair (e 0 ,r 0 ) is the so-called conditioned Brownian snake with initial point 0.<br />

Theorem 3.1.2 provides a useful construction of the conditioned object (e 0 ,r 0 ) from the<br />

unconditioned one (e,r 0 ). In order to present this construction, first recall that there is a.s. a<br />

unique s ∗ in (0,1) such that<br />

r 0 (s ∗ ) = inf<br />

s∈[0,1] r0 (s)<br />

(see Lemma 16 in [45] or Proposition 3.2.5). For every s ∈ [0, ∞), write {s} for the fractional part<br />

of s. According to Theorem 3.1.2, the conditioned snake (e 0 ,r 0 ) may be constructed explicitly<br />

as follows : for every s ∈ [0,1],<br />

e 0 (s) = e(s ∗ ) + e({s ∗ + s}) − 2<br />

r 0 (s) = r 0 ({s ∗ + s}) − r 0 (s ∗ ).<br />

inf e(t),<br />

s∧{s ∗+s}≤t≤s∨{s ∗+s}<br />

4.2.4. The Bouttier-di Francesco-Guitter bijection. We start with a definition. A<br />

(rooted) mobile is a two-type spatial tree (T ,U) whose labels U v only take integer values and<br />

such that the following properties hold :<br />

(a) U v = Uˇv for every v ∈ T 1 .<br />

(b) L<strong>et</strong> v ∈ T 1 such that k = k v (T ) ≥ 1. L<strong>et</strong> v (0) = ˇv be the father of v and l<strong>et</strong> v (j) = vj<br />

for every j ∈ {1,... ,k}. Then for every j ∈ {0,... ,k},<br />

where by convention v (k+1) = v (0) .<br />

U v(j+1) ≥ U v(j) − 1,<br />

Furthermore, if U v ≥ 1 for every v ∈ T , then we say that (T ,U) is a well-labelled mobile.<br />

L<strong>et</strong> T mob<br />

1 denotes the s<strong>et</strong> of all mobiles such that U ∅ = 1. We will now describe the Bouttierdi<br />

Francesco-Guitter bijection from T mob<br />

1 onto M r,p . This bijection can be found in section 2<br />

in [11]. Note that [11] deals with pointed planar maps rather than with rooted pointed planar<br />

maps. It is however easy to verify that the results described below are simple consequences of<br />

[11].<br />

L<strong>et</strong> (T ,U) ∈ T mob<br />

1 . Recall that ζ = #T −1. L<strong>et</strong> u 0 ,u 1 ,... ,u 2ζ be the search-depth sequence of<br />

T . It is immediate to see that u k ∈ T 0 if k is even and that u k ∈ T 1 if k is odd. The search-depth<br />

sequence of T 0 is the sequence w 0 ,w 1 ,... ,w ζ defined by w k = u 2k for every k ∈ {0,1,... ,ζ}.<br />

Notice that w 0 = w ζ = ∅. Although (T ,U) is not necesseraly well labelled, we may s<strong>et</strong> for every<br />

v ∈ T ,<br />

U + v = U v − min{U w : w ∈ T } + 1,<br />

94


and then (T ,U + ) is a well-labelled mobile. Notice that min{U + v : v ∈ T } = 1.<br />

Suppose that the tree T is drawn in the plane and add an extra vertex ∂. We associate with<br />

(T ,U + ) a bipartite planar map whose s<strong>et</strong> of vertices is<br />

T 0 ∪ {∂},<br />

and whose edges are obtained by the following device : for every k ∈ {0,1,... ,ζ},<br />

• if U + w k<br />

= 1, draw an edge b<strong>et</strong>ween w k and ∂ ;<br />

• if U + w k<br />

≥ 2, draw an edge b<strong>et</strong>ween w k and the first vertex in the sequence w k+1 ,...,w ζ−1 ,<br />

w 0 ,w 1 ,... ,w k−1 whose label is U + w k<br />

− 1.<br />

Notice that condition (b) in the definition of a mobile entails that U + w k+1<br />

≥ U + w k<br />

− 1 for every<br />

k ∈ {0,1,... ,ζ − 1} and recall that min{U + w 0<br />

,U + w 1<br />

,... ,U + w ζ−1<br />

} = 1. The preceding properties<br />

ensure that whenever U + w k<br />

≥ 2 there is at least one vertex among w k+1 ,... ,w ζ−1 ,w 0 ,... ,w k−1<br />

with label U + w k<br />

− 1. The construction can be made in such a way that edges do not intersect<br />

(see section 2 in [11] for an example). The resulting planar graph is a bipartite planar map. We<br />

view this map as a rooted pointed planar map by declaring that the distinguished vertex is ∂<br />

and that the root edge is the one corresponding to k = 0 in the preceding construction.<br />

It follows from [11] that the preceding construction yields a bijection Ψ r,p b<strong>et</strong>ween T mob<br />

1 and<br />

M r,p . Furthermore it is not difficult to see that Ψ r,p satisfies the following two properties : l<strong>et</strong><br />

(T ,U) ∈ T mob<br />

1 and l<strong>et</strong> M = Ψ r,p ((T ,U)),<br />

(i) for every k ≥ 1, the s<strong>et</strong> {f ∈ F M : deg(f) = 2k} is in one-to-one correspondence with<br />

the s<strong>et</strong> {v ∈ T 1 : k v (T ) = k − 1},<br />

(ii) for every l ≥ 1, the s<strong>et</strong> {a ∈ V M : d(∂,a) = l} is in one-to-one correspondence with the<br />

s<strong>et</strong> {v ∈ T 0 : U v − min{U w : w ∈ T } + 1 = l}.<br />

We observe that if (T ,U) is a well-labelled mobile then U v<br />

+ = U v for every v ∈ T . In<br />

particular U ∅ + = 1. This implies that the root edge of the planar map Ψ r,p((T ,U)) contains<br />

the distinguished point ∂. Then Ψ r,p ((T ,U)) can be identified to a rooted planar map, whose<br />

root is an oriented edge b<strong>et</strong>ween the root vertex ∂ and w 0 . Write T mob<br />

1 for the s<strong>et</strong> of all welllabelled<br />

mobiles such that U ∅ = 1. Thus Ψ r,p induces a bijection Ψ r from the s<strong>et</strong> T mob<br />

1 onto<br />

the s<strong>et</strong> M r . Furthermore Ψ r satisfies the following two properties : l<strong>et</strong> (T ,U) ∈ T mob<br />

1 and l<strong>et</strong><br />

M = Ψ r ((T ,U)),<br />

(i) for every k ≥ 1, the s<strong>et</strong> {f ∈ F M : deg(f) = 2k} is in one-to-one correspondence with<br />

the s<strong>et</strong> {v ∈ T 1 : k v (T ) = k − 1},<br />

(ii) for every l ≥ 1, the s<strong>et</strong> {a ∈ V M : d(∂,a) = l} is in one-to-one correspondence with the<br />

s<strong>et</strong> {v ∈ T 0 : U v = l}.<br />

4.2.5. Boltzmann distribution on two-type spatial trees. L<strong>et</strong> q be a regular critical<br />

weight sequence. We recall the following definitions from [43]. L<strong>et</strong> µ q 0 be the geom<strong>et</strong>ric<br />

distribution with param<strong>et</strong>er f q (Z q ) that is<br />

and l<strong>et</strong> µ q 1<br />

µ q 0 (k) = Z−1 q f q (Z q ) k , k ≥ 0,<br />

be the probability measure defined by<br />

µ q 1 (k) = Zk q N(k + 1)q k+1<br />

, k ≥ 0.<br />

f q (Z q )<br />

95


From [43], we know that µ 1 has small exponential moments, and that the two-type Galton-<br />

Watson tree associated with µ q = (µ q 0 ,µq 1 ) is critical.<br />

Also, for every k ≥ 0, l<strong>et</strong> ν0 k be the Dirac mass at 0 ∈ Rk and l<strong>et</strong> ν1 k be the uniform distribution<br />

on the s<strong>et</strong> A k defined by<br />

{<br />

}<br />

A k = (x 1 ,...,x k ) ∈ Z k : x 1 ≥ −1,x 2 − x 1 ≥ −1,... ,x k − x k−1 ≥ −1, −x k ≥ −1 .<br />

We can say equivalently that ν1 k is the law of (X 1,...,X 1 + ... + X k ) where (X 1 ,... ,X k+1 ) is<br />

uniformly distributed on the s<strong>et</strong> B k defined by<br />

{<br />

}<br />

B k = (x 1 ,...,x k+1 ) ∈ {−1,0,1,2,...} k+1 : x 1 + ... + x k+1 = 0 .<br />

Notice that #A k = #B k = N(k+1). We s<strong>et</strong> ν = ((ν k 0 ,νk 1 )) k≥1. The following result is Proposition<br />

10 in [43]. However, we provide a short proof for the sake of compl<strong>et</strong>eness.<br />

Proposition 4.2.2. The Boltzmann distribution B r,p<br />

q is the image of the probability measure<br />

P µ q ,ν,1 under the mapping Ψ r,p .<br />

Proof : By construction, the probability measure P µ q ,ν,1 is supported on the s<strong>et</strong> T mob<br />

1 . L<strong>et</strong><br />

(Θ,u) ∈ T mob<br />

1 . We have by the choice of ν,<br />

P µ q ,ν,1 ((t,u)) = P µ q(t)R ν,1 (t, {u})<br />

( ∏ ) −1Pµ<br />

= N(k v (t) + 1) q(t).<br />

v∈t 1<br />

Now,<br />

P µ q(t) =<br />

∏ µ q 0 (k v(t)) ∏<br />

µ q 1 (k v(t))<br />

v∈t 0 v∈t 1<br />

so that we arrive at<br />

= ∏ (<br />

Zq −1 f q(Z q ) kv(t)) ∏ Zq<br />

kv(t) N(k v (t) + 1)q kv(t)+1<br />

f q (Z q )<br />

v∈t 0 v∈t 1<br />

= Z −#t0<br />

q<br />

= Z −1<br />

q<br />

f q (Z q ) #t1 Z #t0 −1<br />

q f q (Z q ) ∏ −#t1 N(k v (t) + 1) ∏<br />

v∈t 1<br />

∏<br />

v∈t 1 q kv(t)+1<br />

∏<br />

v∈t 1 N(k v (t) + 1),<br />

P µ q ,ν,1 ((t,u)) = Z −1<br />

q<br />

∏<br />

v∈t 1 q kv(t)+1.<br />

We s<strong>et</strong> m = Ψ r,p ((t,u)). We have from the property (ii) satisfied by Ψ r,p ,<br />

∏<br />

Zq<br />

−1 q kv(t)+1 = B r,p<br />

q (m),<br />

v∈t 1<br />

which leads us to the desired result.<br />

v∈t 1 q kv(t)+1<br />

L<strong>et</strong> us introduce some notation. As µ q 0 (1) > 0, we have P µ(T 1 = n) > 0 for every n ≥ 1.<br />

Then we may define, for every n ≥ 1 and x ∈ R,<br />

P n µ q = P µ q (· | #T 1 = n ) ,<br />

P n µ q ,ν,x = P µ q ,ν,x<br />

(· | #T 1 = n ) .<br />

□<br />

96


Furthermore, we s<strong>et</strong> for every (T ,U) ∈ Ω,<br />

U = min { U v : v ∈ T 0 \ {∅} } ,<br />

with the convention min ∅ = ∞. Finally we define for every n ≥ 1 and x ≥ 0,<br />

P µ q ,ν,x = P µ q ,ν,x(· | U > 0),<br />

P n µ q ,ν,x = P µ q ,ν,x<br />

(· | #T 1 = n ) .<br />

Corollary 4.2.3. The probability measure B r,p<br />

q (· | #F M = n) is the image of P n µ q ,ν,1 under<br />

the mapping Ψ r,p . The probability measure B r q is the image of P µ q ,ν,1 under the mapping Ψ r .<br />

The probability measure B r q(· | #F M = n) is the image of P n µ q ,ν,1 under the mapping Ψ r .<br />

Proof : The first assertion is a simple consequence of Proposition 4.2.2 tog<strong>et</strong>her with the<br />

property (i) satisfied by Ψ r,p . Recall from section 4.2.1 that we can identify the s<strong>et</strong> M r to<br />

a subs<strong>et</strong> of M r,p in the following way. L<strong>et</strong> Υ : M r −→ M r,p be the mapping defined by<br />

Υ((M,⃗e )) = (M,e,o) for every (M,⃗e ) ∈ M r , where o denotes the root vertex of the map<br />

(M,⃗e ). We easily check that B r,p<br />

q (· | M ∈ Υ(M r )) is the image of B r q under the mapping Υ.<br />

This tog<strong>et</strong>her with Proposition 4.2.2 yields the second assertion. The third assertion follows. □<br />

At last, if q = q κ , we s<strong>et</strong> µ κ 0 = µqκ 0 , µκ 1 = µqκ 1 and µ κ = (µ κ 0 ,µκ 1 ). We then verify that µκ 0 is<br />

the geom<strong>et</strong>ric distribution with param<strong>et</strong>er 1/κ and that µ κ 1 is the Dirac mass at κ − 1. Recall<br />

the notation U n κ and U n κ .<br />

Corollary 4.2.4. The probability measure U n κ is the image of P n µ κ ,ν,1 under the mapping<br />

Ψ r,p . The probability measure U n κ is the image of P n µ κ ,ν,1 under the mapping Ψ r.<br />

4.2.6. Statement of the main result. We first need to introduce some notation. L<strong>et</strong><br />

M ∈ M r . We denote by o its root vertex. The radius R M is the maximal distance b<strong>et</strong>ween o<br />

and another vertex of M that is<br />

R M = max{d(o,a) : a ∈ V M }.<br />

The normalized profile of M is the probability measure λ M on {0,1,2,...} defined by<br />

λ M (k) = #{a ∈ V M : d(o,a) = k}<br />

, k ≥ 0.<br />

#V M<br />

Note that R M is the supremum of the support of λ M . It is also convenient to introduce the<br />

rescaled profile. If M has n faces, this is the probability measure on R + defined by<br />

λ (n)<br />

M (A) = λ M<br />

(<br />

n 1/4 A<br />

for any Borel subs<strong>et</strong> A of R + . At last, if q is a regular critical weight sequence, we s<strong>et</strong><br />

ρ q = 2 + Z 3 q f ′′<br />

q (Z q).<br />

Recall from section 4.2.3 that (e,r 0 ) denotes the Brownian snake with initial point 0.<br />

Theorem 4.2.5. L<strong>et</strong> q be a regular critical weight sequence.<br />

(i) The law of n −1/4 R M under the probability measure B r q(· | #F M = n) converges as<br />

n → ∞ to the law of the random variable<br />

( ) 1/4 (<br />

)<br />

4ρq<br />

r 0 (s) − inf<br />

9(Z q − 1)<br />

0≤s≤1 r0 (s) .<br />

sup<br />

0≤s≤1<br />

)<br />

97


(ii) The law of the random measure λ (n)<br />

M under the probability measure Br q (· | #F M = n)<br />

converges as n → ∞ to the law of the random probability measure I defined by<br />

∫ (<br />

1<br />

( ) 1/4 (<br />

) ) 4ρq<br />

〈I,g〉 = g<br />

r 0 (t) − inf<br />

9(Z q − 1)<br />

0≤s≤1 r0 (s) dt.<br />

0<br />

(iii) The law of the rescaled distance n −1/4 d(o,a) where a is a vertex chosen uniformly at<br />

random among all vertices of M, under the probability measure B r q (· | #F M = n)<br />

converges as n → ∞ to the law of the random variable<br />

( 4ρq<br />

9(Z q − 1)<br />

) 1/4 (<br />

sup r 0 (s)<br />

0≤s≤1<br />

In the case q = q κ , the constant appearing in Theorem 4.2.5 is (4κ(κ − 1)/9) 1/4 . It is equal<br />

to (8/9) 1/4 when κ = 2. The results stated in Theorem 4.2.5 in the special case q = q 2 were<br />

obtained by Chassaing & Schaeffer [14] (see also Theorem 8.2 in [39]).<br />

Obviously Theorem 4.2.5 is related to Theorem 3 proved by Marckert & Miermont [43]. Note<br />

however that [43] deals with rooted pointed maps instead of rooted maps as we do and studies<br />

distances from the distinguished point of the map rather than from the root vertex.<br />

)<br />

.<br />

4.3. A conditional limit theorem for two-type spatial trees<br />

Recall first some notation. L<strong>et</strong> q be a regular critical weight sequence, l<strong>et</strong> µ q = (µ q 0 ,µq 1 ) be<br />

the pair of offspring distributions associated with q and l<strong>et</strong> ν = (ν k 0 ,νk 1 ) k≥1 be the family of<br />

probability measures defined before Proposition 4.2.2.<br />

If (T ,U) ∈ Ω, we denote by C its contour function and by V its spatial contour function.<br />

Recall that C([0,1], R) 2 is equipped with the norm ‖(f,g)‖ = ‖f‖ u ∨ ‖g‖ u . The following<br />

result is a special case of Theorem 11 in [43].<br />

Theorem 4.3.1. L<strong>et</strong> q be a regular critical weight sequence. The law under P n µ q ,ν,0<br />

⎛<br />

of<br />

(√ ) ( (9(Zq ) ) ⎞<br />

ρq (Z<br />

⎝ q − 1) C(2(#T − 1)t)<br />

− 1) 1/4<br />

V (2(#T − 1)t)<br />

4 n 1/2 ,<br />

⎠<br />

4ρ q n 1/4<br />

0≤t≤1<br />

0≤t≤1<br />

converges as n → ∞ to the law of (e,r 0 ). The convergence holds in the sense of weak convergence<br />

of probability measures on C([0,1], R) 2 .<br />

Note that Theorem 11 in [43] deals with the so-called height-process instead of the contour<br />

process. However, we can deduce Theorem 4.3.1 from [43] by classical arguments (see e.g. [38]).<br />

In this section, we will prove a conditional version of Theorem 4.3.1. Before stating this result,<br />

we establish a corollary of Theorem 4.3.1. To this end we s<strong>et</strong><br />

Q µ q = P µ q(· | k ∅ (T ) = 1),<br />

Q µ q ,ν = P µ q ,ν,0(· | k ∅ (T ) = 1).<br />

Notice that this conditioning makes sense since µ q 0 (1) > 0. We may also define for every n ≥ 1,<br />

Q n µ q = Q µ<br />

(· q | #T 1 = n ) ,<br />

Q n µ q ,ν = Q µ q ,ν<br />

(· | #T 1 = n ) .<br />

98


Corollary 4.3.2. L<strong>et</strong> q be a regular critical weight sequence. The law under Q n µ q ,ν of<br />

⎛(√ ) ( (9(Zq ) ) ⎞<br />

ρq (Z<br />

⎝ q − 1) C(2(#T − 1)t)<br />

− 1) 1/4<br />

V (2(#T − 1)t)<br />

4 n 1/2 ,<br />

⎠<br />

4ρ q n 1/4<br />

0≤t≤1<br />

0≤t≤1<br />

converges as n → ∞ to the law of (e,r 0 ). The convergence holds in the sense of weak convergence<br />

of probability measures on C([0,1], R) 2 .<br />

Proof : We first introduce some notation. If (T ,U) ∈ Ω and w 0 ∈ T , we define a spatial<br />

tree (T [w 0] ,U [w 0] ) by s<strong>et</strong>ting<br />

T [w 0] = {v : w 0 v ∈ T },<br />

and for every v ∈ T [w 0]<br />

U [w 0]<br />

v = U w0 v − U w0 .<br />

Denote by C [w 0] the contour function and by V [w 0] the spatial contour function of (T [w 0] ,U [w 0] ).<br />

As a consequence of Theorem 11 in [43], the law under Q n µ q ,ν<br />

⎛<br />

of<br />

(√<br />

ρq (Z<br />

⎝ q − 1) C [1] ( 2 ( #T [1] − 1 ) t ) ) ( (9(Zq )<br />

− 1) 1/4<br />

V [1] ( 2 ( #T [1] − 1 ) t ) )<br />

4<br />

n 1/2 ,<br />

4ρ q n 1/4<br />

0≤t≤1<br />

converges as n → ∞ to the law of (e,r 0 ). We then easily g<strong>et</strong> the desired result.<br />

0≤t≤1<br />

Recall from section 4.2.3 that (e 0 ,r 0 ) denotes the conditioned Brownian snake with initial<br />

point 0.<br />

Theorem 4.3.3. L<strong>et</strong> q be a regular critical weight sequence. For every x ≥ 0, the law under<br />

P n µ q ,ν,x<br />

⎛<br />

of<br />

(√ ) ( (9(Zq ) ) ⎞<br />

ρq (Z<br />

⎝ q − 1) C(2(#T − 1)t)<br />

− 1) 1/4<br />

V (2(#T − 1)t)<br />

4 n 1/2 ,<br />

⎠<br />

4ρ q n 1/4<br />

0≤t≤1<br />

0≤t≤1<br />

converges as n → ∞ to the law of (e 0 ,r 0 ). The convergence holds in the sense of weak convergence<br />

of probability measures on C([0,1], R) 2 .<br />

To prove Theorem 4.3.3, we will follow the lines of the proof of Theorem 2.2 in [39]. From<br />

now on, we s<strong>et</strong> µ = µ q to simplify notation.<br />

4.3.1. Rerooting two-type spatial trees. If T ∈ T, we say that a vertex v ∈ T is a leaf<br />

of T if k v (T ) = 0 meaning that v has no child. We denote by ∂T the s<strong>et</strong> of all leaves of T and<br />

we write ∂ 0 T = ∂T ∩ T 0 for the s<strong>et</strong> of leaves of T which are of type 0.<br />

L<strong>et</strong> us recall some notation that can be found in section 3 in [39]. Recall that U ∗ = U \ {∅}.<br />

If v 0 ∈ U ∗ and T ∈ T are such that v 0 ∈ T , we define k = k(v 0 , T ) and l = l(v 0 , T ) in the<br />

following way. Write ζ = #T − 1 and u 0 ,u 1 ,...,u 2ζ for the search-depth sequence of T . Then<br />

we s<strong>et</strong><br />

k = min{i ∈ {0,1,... ,2ζ} : u i = v 0 },<br />

l = max{i ∈ {0,1,... ,2ζ} : u i = v 0 },<br />

⎞<br />

⎠<br />

□<br />

99


which means that k is the time of the first visit of v 0 in the evolution of the contour of T and<br />

that l is the time of the last visit of v 0 . Note that l ≥ k and that l = k if and only if v 0 ∈ ∂T .<br />

For every t ∈ [0,2ζ − (l − k)], we s<strong>et</strong><br />

Ĉ (v0) (t) = C(k) + C([[k − t]]) − 2 inf C(s),<br />

s∈[k∧[k−t],k∨[k−t]]<br />

where C is the contour function of T and [[k−t]] stands for the unique element of [0,2ζ) such that<br />

[[k−t]]−(k−t) = 0 or 2ζ. Then there exists a unique plane tree ̂T (v 0) ∈ T whose contour function<br />

is Ĉ(v 0) . Informally, ̂T (v 0) is obtained from T by removing all vertices that are descendants of<br />

v 0 and by re-rooting the resulting tree at v 0 . Furthermore, if v 0 = u 1 ...u n , then we see that<br />

̂v 0 = 1u n ...u 2 belongs to ̂T (v 0) . In fact, ̂v 0 is the vertex of ̂T (v 0) corresponding to the root of<br />

the initial tree. At last notice that k ∅ (̂T (v 0) ) = 1.<br />

If T ∈ T and w 0 ∈ T , we s<strong>et</strong><br />

T (w 0) = T \ {w 0 u ∈ T : u ∈ U ∗ }.<br />

The following lemma is an analogue of Lemma 3.1 in [39] for two-type Galton-Watson trees.<br />

Note that in what follows, two-type trees will always be re-rooted at a vertex of type 0.<br />

Recall the definition of the probability measure Q µ .<br />

Lemma 4.3.4. L<strong>et</strong> v 0 ∈ U ∗ be of the form v 0 = 1u 2 ... u 2p for some p ∈ N. Assume that<br />

Q µ (v 0 ∈ T ) > 0. Then the law of the re-rooted tree ̂T (v 0) under Q µ (· | v 0 ∈ T ) coincides with<br />

the law of the tree T (bv 0) under Q µ (· | ̂v 0 ∈ T ).<br />

Proof : We first notice that<br />

Q µ (v 0 ∈ T ) = µ 1<br />

({<br />

u 2 ,u 2 + 1,... }) µ 0<br />

({<br />

u 3 ,u 3 + 1,... }) ... µ 1<br />

({<br />

u 2p ,u 2p + 1,... }) ,<br />

so that<br />

Q µ (v 0 ∈ T ) = Q µ (̂v 0 ∈ T ) > 0.<br />

In particular, both conditionings of Lemma 4.3.4 make sense. L<strong>et</strong> t be a two-type tree such that<br />

̂v 0 ∈ ∂ 0 t and k ∅ (t) = 1. Since the trees t and ̂t (bv 0) represent the same graph, we have<br />

) ∏<br />

Q µ<br />

(T (bv 0) = t = µ 0 (k u (t)) ∏ µ 1 (k u (t))<br />

u∈t 1<br />

=<br />

=<br />

which implies the desired result.<br />

u∈t 0 \{∅,bv 0 }<br />

∏<br />

u∈t 0 \{∅,bv 0 }<br />

∏<br />

u∈bt (bv 0 ),0 \{∅,v 0 }<br />

(<br />

= Q µ T (v0) = ̂t )<br />

(bv 0)<br />

)<br />

(v<br />

= Q µ<br />

(̂T 0 ) = t ,<br />

µ 0 (deg(u) − 1) ∏ u∈t 1 µ 1 (deg(u) − 1)<br />

µ 0 (deg(u) − 1)<br />

∏<br />

u∈bt (bv 0 ),1 µ 1 (deg(u) − 1)<br />

Before stating a spatial version of Lemma 4.3.4, we establish a symm<strong>et</strong>ry property of the<br />

collection of measures ν. To this end, we l<strong>et</strong> ˜ν 1 k be the image measure of νk 1 under the mapping<br />

(x 1 ,... ,x k ) ∈ R k ↦−→ (x k ,...,x 1 ) and we s<strong>et</strong> ˜ν = ((ν0 k, ˜νk 1 )) k≥1.<br />

□<br />

100


Lemma 4.3.5. For every k ≥ 1 and every j ∈ {1,... ,k}, the measures ν1 k and ˜νk 1 are invariant<br />

under the mapping φ j : R k −→ R k defined by<br />

φ j (x 1 ,...,x k ) = (x j+1 − x j ,... ,x k − x j , −x j ,x 1 − x j ,...,x j−1 − x j ).<br />

Proof : Recall the definition of the s<strong>et</strong>s A k and B k . L<strong>et</strong> ρ k be the uniform distribution on<br />

B k . Then ν k 1 is the image measure of ρk under the mapping ϕ k : B k −→ A k defined by<br />

For every (x 1 ,...,x k+1 ) ∈ R k+1 we s<strong>et</strong><br />

ϕ k (x 1 ,... ,x k+1 ) = (x 1 ,x 1 + x 2 ,... ,x 1 + ... + x k ).<br />

p j (x 1 ,... ,x k+1 ) = (x j+1 ,...,x k+1 ,x 1 ,... ,x j ).<br />

It is immediate that ρ k is invariant under the mapping p j . Furthermore φ j ◦ ϕ k (x) = ϕ k ◦ p j (x)<br />

for every x ∈ B k , which implies that ν k 1 is invariant under φ j.<br />

At last for every (x 1 ,... ,x k ) ∈ R k we s<strong>et</strong><br />

S(x 1 ,...,x k ) = (x k ,...,x 1 ).<br />

Then φ j ◦ S = S ◦ φ k−j+1 , which implies that ˜ν k 1 is invariant under φ j.<br />

If (T ,U) ∈ Ω and v 0 ∈ T 0 , the re-rooted spatial tree (̂T (v0),Û(v0) ) is defined as follows. For<br />

every vertex v ∈ ̂T (v0),0 , we s<strong>et</strong><br />

Û (v 0)<br />

v = U v − U v0 ,<br />

where v is the vertex of the initial tree T corresponding to v, and for every vertex v ∈ ̂T (v0),1 ,<br />

we s<strong>et</strong><br />

Û (v 0)<br />

v = Û(v 0)<br />

ˇv .<br />

Note that, since v 0 ∈ T 0 , v ∈ ̂T (v0) is of type 0 if and only if v ∈ T is of type 0.<br />

If (T ,U) ∈ Ω and w 0 ∈ T , we also consider the spatial tree (T (w 0) ,U (w 0) ) where U (w 0) is the<br />

restriction of U to the tree T (w 0) .<br />

Recall the definition of the probability measure Q µ,ν .<br />

Lemma 4.3.6. L<strong>et</strong> v 0 ∈ U ∗ be of the form v 0 = 1u 2 ... u 2p for some p ∈ N. Assume that<br />

Q µ (v 0 ∈ T ) > 0. Then the law of the re-rooted spatial tree (̂T (v 0) ,Û(v 0) ) under Q µ,eν (· | v 0 ∈ T )<br />

coincides with the law of the spatial tree (T (bv 0) ,U (bv 0) ) under Q µ,ν (· | ̂v 0 ∈ T ).<br />

Lemma 4.3.6 is a consequence of Lemma 4.3.4 and Lemma 4.3.5. We leave d<strong>et</strong>ails to the<br />

reader.<br />

If (T ,U) ∈ Ω, we denote by ∆ 0 = ∆ 0 (T ,U) the s<strong>et</strong> of all vertices of type 0 with minimal<br />

spatial position :<br />

∆ 0 = { v ∈ T 0 : U v = min{U w : w ∈ T } } .<br />

We also denote by v m the first element of ∆ 0 in the lexicographical order. The following two<br />

lemmas can be proved from Lemma 4.3.6 in the same way as Lemma 3.3 and Lemma 3.4 in [39].<br />

Lemma 4.3.7. For any nonnegative measurable functional F on Ω,<br />

(<br />

(v<br />

Q µ,eν<br />

(F<br />

(̂T m) ,Û(vm))½{#∆ 0 =1,v m∈∂ 0 T })<br />

= Q µ,ν F(T ,U)(#∂0 T )½{U>0})<br />

.<br />

Lemma 4.3.8. For any nonnegative measurable functional F on Ω,<br />

⎛<br />

Q µ,eν<br />

⎝<br />

∑ (v<br />

F<br />

(̂T )⎞ 0 ) ( ) ,Û(v 0) ⎠ = Q µ,ν F(T ,U)(#∂0 T )½{U≥0} .<br />

v 0 ∈∆ 0 ∩∂ 0 T<br />

101<br />


4.3.2. Estimates for the probability of staying on the positive side. In this section<br />

we will derive upper and lower bounds for the probability P n µ,ν,x(U > 0) as n → ∞. We first<br />

state a lemma which is a direct consequence of Lemma 17 in [43].<br />

Lemma 4.3.9. There exist constants c 0 > 0 and c 1 > 0 such that<br />

n 3/2 (<br />

P µ #T 1 = n )<br />

−→ c 0 ,<br />

n→∞<br />

n 3/2 (<br />

Q µ #T 1 = n )<br />

−→ c 1 .<br />

n→∞<br />

We now establish a preliminary estimate concerning the number of leaves of type 0 in a tree<br />

with n vertices of type 1.<br />

Lemma 4.3.10. There exists a constant β 0 > 0 such that for every n sufficiently large,<br />

)<br />

P µ<br />

(|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 ,#T 1 = n ≤ e −nβ0 .<br />

Proof : L<strong>et</strong> T be a two-type tree. Recall that ζ = #T − 1. L<strong>et</strong><br />

v(0) = ∅ ≺ v(1) ≺ ... ≺ v(ζ)<br />

be the vertices of T listed in lexicographical order. For every n ∈ {0,1,... ,ζ} we define R n =<br />

(R n (k)) k≥1 as follows. For every k ∈ {1,... , |v(n)|}, R n (k) is the number of younger brothers of<br />

the ancestor of v(n) at generation k. Here younger brothers are those brothers which have not<br />

y<strong>et</strong> been visited at time n in search-depth sequence. For every k > |v(n)|, we s<strong>et</strong> R n (k) = 0.<br />

Standard arguments (see e.g. [41] for similar results) show that (R n , |v(n)|) 0≤n≤ζ has the same<br />

distribution as (R n ′ ,h′ n ) 0≤n≤T ′ −1, where (R n ′ ,h′ n ) n≥0 is a Markov chain whose transition kernel<br />

is given by :<br />

(<br />

)<br />

• S ((r 1 ,...,r h ,0,...),h),((r 1 ,...,r h ,k − 1,0,...),h + 1) = µ i (k) for k ≥ 1, h ≥ 0 and<br />

r 1 ,...,r h ≥ 0,<br />

(<br />

)<br />

• S ((r 1 ,...,r h ,0,...),h),((r 1 ,...,r l − 1,0,...),l) = µ i (0), where<br />

l = inf{m ≥ 1 : r m > 0},<br />

for h ≥ 1 and r 1 ,...,r h ≥ 0 such that {m ≥ 1 : r m > 0} ≠ ∅,<br />

• S((0,h),(0,0)) = µ i (0) for every h ≥ 0,<br />

where i = 0 if h is even, and i = 1 if h is odd, and finally<br />

T ′ = inf { n ≥ 1 : (R ′ n,h ′ n) = (0,0) } .<br />

Write P ′ for the probability measure under which (R n ′ ,h′ n ) n≥0 is defined. We define a sequence<br />

of stopping times (τ j ′) j≥0 by τ 0 ′ = inf{n ≥ 0 : h′ n is odd} and τ j+1 ′ = inf{n > τ j ′ : h′ n is odd} for<br />

every j ≥ 0. At last we s<strong>et</strong> for every j ≥ 0,<br />

} ( ( ))<br />

X j =½{<br />

′ h ′ τ j ′ +1 = h′ τ j ′ + 1 1 + R<br />

τ ′ j ′ +1 h ′ τ j ′ +1 .<br />

Since #T 0 = 1 + ∑ u∈T 1 k u(T ), we have<br />

⎛<br />

)<br />

P µ<br />

(|#T 0 − m 1 n| > n 3/4 ,#T 1 = n = P ′ ⎝ ∣ ⎞<br />

n−1<br />

∑<br />

∣ X j ′ − m 1n + 1∣ > n 3/4 ,τ n−1 ′ < T ′ < τ n<br />

′ ⎠<br />

≤<br />

j=0<br />

⎛<br />

P ′ ⎝ ∣ ⎞<br />

n−1<br />

∑<br />

∣ X j ′ − m 1 n∣ > n 3/4 − 1⎠ .<br />

102<br />

j=0


Thanks to the strong Markov property, the random variables X j ′ are independent and distributed<br />

according to µ 1 . A standard moderate deviations inequality ensures the existence of a positive<br />

constant β 1 > 0 such that for every n sufficiently large,<br />

( ∣∣#T<br />

(4.3.1) P 0 µ − m 1 n ∣ )<br />

> n 3/4 ,#T 1 = n ≤ e −nβ1 .<br />

In the same way as previously, we define another sequence of stopping times (θ j ′) j≥0 by θ 0 ′ = 0<br />

and θ j+1 ′ = inf{n > θ′ j : h′ n is even} for every j ≥ 0 and we s<strong>et</strong> for every j ≥ 0,<br />

}<br />

Y j ′ =½{h ′ θ j ′ +1 ≤ h′ θ j<br />

′ .<br />

Using the sequences (θ j ′) j≥0 and (Y j ′)<br />

j≥0, an argument similar to the proof of (4.3.1) shows that<br />

there exists a positive constant β 2 > 0 such that for every n sufficiently large,<br />

)<br />

(4.3.2) P µ<br />

(|#(∂ 0 T ) − µ 0 (0)n| > n 5/8 ,#T 0 = n ≤ e −nβ2 .<br />

From (4.3.1), we g<strong>et</strong> for n sufficiently large,<br />

)<br />

P µ<br />

(|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 ,#T 1 = n<br />

≤ e −nβ1 + P µ<br />

(|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 , |#T 0 − m 1 n| ≤ n 3/4) .<br />

However, for n sufficiently large,<br />

P µ<br />

(<br />

|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 , |#T 0 − m 1 n| ≤ n 3/4)<br />

=<br />

≤<br />

≤<br />

⌊n 3/4 +m 1 n⌋<br />

∑<br />

k=⌈−n 3/4 +m 1 n⌉<br />

⌊n 3/4 +m 1 n⌋<br />

∑<br />

k=⌈−n 3/4 +m 1 n⌉<br />

⌊n 3/4 +m 1 n⌋<br />

∑<br />

k=⌈−n 3/4 +m 1 n⌉<br />

)<br />

P µ<br />

(|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 ,#T 0 = k<br />

)<br />

P µ<br />

(|(#∂ 0 T ) − µ 0 (0)k| > (1 − µ 0 (0))n 3/4 ,#T 0 = k<br />

)<br />

P µ<br />

(|(#∂ 0 T ) − µ 0 (0)k| > k 5/8 ,#T 0 = k .<br />

At last, we use (4.3.2) to obtain for n sufficiently large,<br />

P µ<br />

(|(#∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 , |#T 0 − m 1 n| ≤ n 3/4) ≤ (2n 3/4 + 1)e −Cnβ2 ,<br />

where C is a positive constant. The desired result follows by combining this last estimate with<br />

(4.3.1). □<br />

We will now state a lemma which plays a crucial role in the proof of the main result of this<br />

section. To this end, recall the definition of v m and s<strong>et</strong> for every n ≥ 1,<br />

Q n µ = Q µ<br />

(· | #T 1 = n ) ,<br />

Q n µ,ν = Q µ,ν<br />

(· | #T 1 = n ) .<br />

Lemma 4.3.11. There exists a constant c > 0 such that for every n sufficiently large,<br />

Q n µ,ν (v m ∈ ∂ 0 T ) ≥ c.<br />

103


Proof : The proof of this lemma is similar to the proof of Lemma 4.3 in [39]. Nevertheless,<br />

we give a few d<strong>et</strong>ails to explain how this proof can be adapted to our context.<br />

Choose p ≥ 1 such that µ 1 (p) > 0. Under Q µ,ν (· | k 1 (T ) = p, k 11 (T ) = ... = k 1p (T ) = 2),<br />

we can define 2p spatial trees {(T ij ,U ij ),i = 1,... ,p,j = 1,2} as follows. For i ∈ {1,... ,p} and<br />

j = 1,2, we s<strong>et</strong><br />

T ij = {∅} ∪ {1v : 1ijv ∈ T },<br />

U ij<br />

∅<br />

= 0 and Uij 1v = U 1ijv − U 1i if 1ijv ∈ T . Then under the probability measure Q µ,ν (· |<br />

k 1 (T ) = p, k 11 (T ) = ... = k 1p (T ) = 2), the trees {(T ij ,U ij ),i = 1,... ,p,j = 1,2} are<br />

independent and distributed according to Q µ,ν . Furthermore, we notice that under the measure<br />

Q µ,ν (· | k 1 (T ) = p, k 11 (T ) = ... = k 1p (T ) = 2), we have with an obvious notation<br />

( {#T 11,1 + #T 12,1 = n − 2p + 1 } ∩ {U 11 < 0} ∩ { v 11<br />

m ∈ ∂ 0 T 11} ∩<br />

∩ { U 12 ≥ 0 } ∩<br />

⋂<br />

2≤i≤p,j=1,2<br />

So we have for n ≥ 1 + 2p,<br />

Q µ,ν<br />

(<br />

#T 1 = n,v m ∈ ∂ 0 T )<br />

(4.3.3)<br />

⋂<br />

2≤i≤p<br />

{U 1i ≥ 0}<br />

{<br />

T ij = {∅,1,11,... ,1p}, U ij ≥ 0 } ) ⊂ { #T 1 = n,v m ∈ ∂ 0 T } .<br />

n−2p<br />

∑ (<br />

≥ C(µ,ν,p) Q µ,ν #T 1 = j,v m ∈ ∂ 0 T ) (<br />

Q µ,ν #T 1 = n − j − 2p + 1, U ≥ 0 ) ,<br />

j=1<br />

where<br />

C(µ,ν,p) = µ 1 (p) 2p−1 µ 0 (2) p µ 0 (0) 2p(p−1) ν p (<br />

1 (−∞,0) × [0,+∞)<br />

p−1 ) ν p 1 ([0,+∞)p ) 2(p−1) .<br />

From (4.3.3), we are now able to g<strong>et</strong> the result by following the lines of the proof of Lemma 4.3<br />

in [39].<br />

□<br />

We can now state the main result of this section.<br />

Proposition 4.3.12. L<strong>et</strong> K > 0. There exist constants γ 1 > 0, γ 2 > 0, ˜γ 1 > 0 and ˜γ 2 > 0<br />

such that for every n sufficiently large and for every x ∈ [0,K],<br />

˜γ 1<br />

n ≤ Qn µ,ν(U > 0) ≤ ˜γ 2<br />

n ,<br />

γ 1<br />

n ≤ Pn µ,ν,x (U > 0) ≤ γ 2<br />

n .<br />

Proof : The proof of Proposition 4.3.12 is similar to the proof of Proposition 4.2 in [39].<br />

The major difference comes from the fact that we cannot easily g<strong>et</strong> an upper bound for #(∂ 0 T )<br />

on the event {#T 1 = n}. In what follows, we will explain how to circumvent this difficulty.<br />

We first use Lemma 4.3.7 with F(T ,U) =½{T 1 =n}. Since #̂T (v 0),1 = #T 1 if v 0 ∈ ∂ 0 T , we<br />

g<strong>et</strong><br />

(4.3.4) Q µ,ν<br />

(<br />

#(∂0 T )½{#T 1 =n, U>0})<br />

≤ Qµ<br />

(<br />

#T 1 = n ) .<br />

On the other hand, we have<br />

( )<br />

Q µ,ν #(∂0 T )½{#T 1 =n, U>0}<br />

≥ m 1 µ 0 (0)nQ µ,ν<br />

(<br />

#T 1 = n, U > 0 ) − Q µ,ν<br />

(<br />

|#(∂0 T ) − m 1 µ 0 (0)n|½{#T 1 =n, U>0})<br />

.<br />

(4.3.5)<br />

104


Now thanks to Lemma 4.3.10, we have for n sufficiently large,<br />

( )<br />

Q µ,ν |#(∂0 T ) − m 1 µ 0 (0)n|½{#T 1 =n, U>0}<br />

≤ n 3/4 (<br />

Q µ,ν #T 1 = n, U > 0 ) ( )<br />

+ Q µ,ν #(∂0 T )½{#T 1 =n, U>0}<br />

(4.3.6)<br />

+ m 1 µ 0 (0)nQ µ,ν (|#(∂ 0 T ) − m 1 µ 0 (0)n| > n 3/4 ,#T 1 = n)<br />

≤ n 3/4 Q µ,ν<br />

(<br />

#T 1 = n, U > 0 ) + Q µ,ν<br />

(<br />

#(∂0 T )½{#T 1 =n, U>0})<br />

+ m1 µ 0 (0)ne −nβ0 .<br />

From (4.3.4), (4.3.5) and (4.3.6) we g<strong>et</strong> for n sufficiently large<br />

(m 1 µ 0 (0)n − n 3/4 )Q µ,ν<br />

(<br />

#T 1 = n, U > 0 ) ≤ 2Q µ<br />

(<br />

#T 1 = n ) + m 1 µ 0 (0)ne −nβ0 .<br />

Using Lemma 4.3.9 it follows that<br />

lim supnQ n µ,ν (U > 0) ≤ 2<br />

n→∞<br />

m 1 µ 0 (0) ,<br />

which ensures the existence of ˜γ 2 .<br />

L<strong>et</strong> us now use Lemma 4.3.8 with<br />

F(T ,U) =½{T 1 =n,#(∂ 0 T )≤m 1 µ 0 (0)n+n 3/4 } .<br />

Since #(∂ 0 T ) = #(∂ 0 ̂T<br />

(v 0 ) ) if v 0 ∈ ∂ 0 T , we have for n sufficiently large,<br />

)<br />

Q µ,ν<br />

(#(∂ 0 T )½{#T 1 =n,#(∂ 0 T )≤m 1 µ 0 (0)+n 3/4 , U≥0}<br />

)<br />

= Q µ,eν<br />

(#(∆ 0 ∩ ∂ 0 T )½{#T 1 =n,#(∂ 0 T )≤m 1 µ 0 (0)+n 3/4 }<br />

)<br />

= Q µ,ν<br />

(#(∆ 0 ∩ ∂ 0 T )½{#T 1 =n,#(∂ 0 T )≤m 1 µ 0 (0)+n 3/4 }<br />

(<br />

≥ Q µ,ν #(∆ 0 ∩ ∂ 0 T ) ≥ 1, #T 1 = n, #(∂ 0 T ) ≤ m 1 µ 0 (0) + n 3/4)<br />

(<br />

≥ Q µ,ν v m ∈ ∂ 0 T , #T 1 = n, #(∂ 0 T ) ≤ m 1 µ 0 (0) + n 3/4)<br />

(<br />

≥ Q µ,ν vm ∈ ∂ 0 T , #T 1 = n ) )<br />

− Q µ,ν<br />

(#(∂ 0 T ) > m 1 µ 0 (0) + n 3/4 ,#T 1 = n<br />

(4.3.7)<br />

≥ c Q µ,ν<br />

(<br />

#T 1 = n ) − e −nβ0 ,<br />

where the last inequality comes from Lemma 4.3.10 and Lemma 4.3.11. On the other hand,<br />

)<br />

Q µ,ν<br />

(#(∂ 0 T )½{#T 1 =n,#(∂ 0 T )≤m 1 µ 0 (0)+n 3/4 , U≥0}<br />

(<br />

≤ m 1 µ 0 (0)n + n 3/4) (<br />

(4.3.8)<br />

Q µ,ν #T 1 = n, U ≥ 0 ) .<br />

Then (4.3.7), (4.3.8) and Lemma 4.3.9 imply that for n sufficiently large,<br />

c<br />

(4.3.9) lim inf<br />

n→∞ nQn µ,ν(U ≥ 0) ≥<br />

m 1 µ 0 (0) .<br />

105


Recall that p ≥ 1 is such that µ 1 (p) > 0. Also recall the definition of the spatial tree (T [w 0] ,U [w 0] ).<br />

From the proof of Corallary 4.3.2, we have<br />

(<br />

P µ,ν,0 U > 0,#T 1 = n ) ≥ P µ,ν,0<br />

(k ∅ (T ) = 1,k 1 (T ) = p,U 11 > 0,... ,U 1p > 0,<br />

)<br />

#T [11],1 = n − 1, U [11] ≥ 0, T [12] = ... = T [1p] = {∅}<br />

(<br />

≥ C 2 (µ,ν,p)P µ,ν,0 U ≥ 0,#T 1 = n − 1 )<br />

(4.3.10)<br />

≥ µ 0 (1)C 2 (µ,ν,p)Q µ,ν<br />

(<br />

U ≥ 0,#T 1 = n − 1 ) ,<br />

where we have s<strong>et</strong><br />

C 2 (µ,ν,p) = µ 0 (1)µ 1 (p)µ 0 (0) p−1 ν p 1 ((0,+∞)p ) .<br />

We then deduce the existence of γ 1 from (4.3.9), (4.3.10) and Lemma 4.3.9. A similar argument<br />

gives the existence of ˜γ 1 .<br />

At last, we define m ∈ N by the condition (m − 1)p < K ≤ mp. For every l ∈ N, we define<br />

1 l ∈ U by 1 l = 11... 1, |1 l | = l. Notice that ν p 1 ({(p,p − 1,... ,1)}) = N(p + 1)−1 . By arguing on<br />

the event<br />

{k ∅ (T ) = k 11 (T ) = ... = k 1 2m−2 = 1,k 1 (T ) = ... = k 1 2m−1(T ) = p},<br />

we see that for every n ≥ m,<br />

(4.3.11) Q µ,ν<br />

(<br />

#T 1 = n, U > 0 ) ≥ C 3 (µ,ν,p,m)P µ,ν,K<br />

(<br />

#T 1 = n − m, U > 0 ) ,<br />

with<br />

C 3 (µ,ν,p,m) = µ 0 (1) m−1 µ 1 (p) m µ 0 (0) m(p−1) N(p + 1) −m .<br />

Thanks to Lemma 4.3.9, (4.3.11) yields for every n sufficiently large,<br />

which gives the existence of γ 2 .<br />

P n µ,ν,K (U > 0) ≤ 2c 1<br />

c 0<br />

˜γ 2<br />

C 3 (µ,ν,p,m)<br />

4.3.3. Asymptotic properties of conditioned trees. We first introduce a specific notation<br />

for rescaled contour and spatial contour processes. For every n ≥ 1 and every t ∈ [0,1],<br />

we s<strong>et</strong><br />

√<br />

C (n) ρq (Z q − 1) C(2(#T − 1)t)<br />

(t) =<br />

4 n 1/2 ,<br />

( )<br />

V (n) 9(Zq − 1) 1/4<br />

V (2(#T − 1)t)<br />

(t) =<br />

4ρ q n 1/4 .<br />

In this section, we will g<strong>et</strong> some information about asymptotic properties of the pair (C (n) ,V (n) )<br />

under P n µ,ν,x. We will consider the conditioned measure<br />

Q n µ,ν = Qn µ,ν (· | U > 0).<br />

Berfore stating the main result of this section, we will establish three lemmas. The first one is<br />

the analogue of Lemma 6.2 in [39] for two-type spatial trees and can be proved in a very similar<br />

way.<br />

1<br />

n ,<br />

□<br />

106


Lemma 4.3.13. There exists a constant c > 0 such that, for every measurable function F on<br />

Ω with 0 ≤ F ≤ 1,<br />

(<br />

( )<br />

Qµ,ν(F(T n ,U)) ≤ c Q n (v<br />

µ,eν F<br />

(̂T m) ,Û(vm))) + O n 5/2 e −nβ 0<br />

,<br />

( )<br />

where the constant β 0 is defined in Lemma 4.3.10 and the estimate O n 5/2 e −nβ 0<br />

for the remainder<br />

holds uniformly in F.<br />

Recall the notation ˇv for the “father” of the vertex v ∈ T \ {∅}.<br />

Lemma 4.3.14. For every ε > 0,<br />

(<br />

(4.3.12) Q n µ,ν<br />

Likewise, for every ε > 0 and x ≥ 0,<br />

(<br />

(4.3.13) P n µ,ν,x<br />

sup<br />

v∈T \{∅}<br />

sup<br />

v∈T \{∅}<br />

|U v − Uˇv |<br />

n 1/4<br />

|U v − Uˇv |<br />

n 1/4<br />

> ε<br />

> ε<br />

)<br />

)<br />

−→ 0.<br />

n→∞<br />

−→ 0.<br />

n→∞<br />

Proof : L<strong>et</strong> ε > 0. First notice that the probability measure ν1 k is supported on the s<strong>et</strong><br />

{−k, −k + 1,... ,k} k . Then we have Q n µ,ν a.s. or P n µ,ν,x a.s.,<br />

sup |U v − Uˇv | = sup |U v − Uˇv | ≤ sup k v (T ).<br />

v∈T \{∅}<br />

v∈T 0 \{∅}<br />

v∈T 1<br />

Now, from Lemma 16 in [43], there exists a constant α 0 > 0 such that for all n sufficiently large,<br />

(<br />

)<br />

Q µ sup k v (T ) > εn 1/4 ≤ e −nα0 ,<br />

v∈T<br />

(<br />

1 )<br />

P µ sup k v (T ) > εn 1/4 ≤ e −nα0 .<br />

v∈T 1<br />

Our assertions (4.3.12) and (4.3.13) easily follow using also Lemma 4.3.9 and Proposition<br />

4.3.12. □<br />

Recall the definition of the re-rooted tree (̂T (v 0) ,Û(v 0) ). Its contour and spatial contour functions<br />

(Ĉ(v 0) , ̂V (v 0) ) are defined on the line interval [0,2(#̂T (v 0) − 1)]. We extend these functions<br />

to the line interval [0,2(#T − 1)] by s<strong>et</strong>ting Ĉ(v 0) (t) = 0 and ̂V (v 0) (t) = 0 for every<br />

t ∈ [2(#̂T (v 0) − 1),2(#T − 1)]. Also recall the definition of v m .<br />

At last, recall that (e 0 ,r 0 ) denotes the conditioned Brownian snake.<br />

Lemma 4.3.15. The law under Q n µ,ν<br />

⎛<br />

of<br />

(√ )<br />

ρq (Z<br />

⎝ q − 1) Ĉ (vm) (2(#T − 1)t)<br />

4 n 1/2<br />

0≤t≤1<br />

,<br />

( (9(Zq<br />

− 1)<br />

4ρ q<br />

) )<br />

1/4 ̂V<br />

(v m) (2(#T − 1)t)<br />

n 1/4<br />

0≤t≤1<br />

converges to the law of (e 0 ,r 0 ). The convergence holds in the sense of weak convergence of<br />

probability measures on the space C([0,1], R) 2 .<br />

Proof : From Corollary 4.3.2 and the Skorokhod representation theorem, we can construct<br />

on a suitable probability space a sequence a sequence (T n ,U n ) and a Brownian snake (,Ö0 ),<br />

⎞<br />

⎠<br />

107


such that each pair (T n ,U n ) is distributed according to Q n µ,ν, and such that if we write (C n ,V n )<br />

for the contour and spatial contour functions of (T n ,U n ) and ζ n = #T n − 1, we have<br />

⎛(√ ) ( (9(Zq ) ) ⎞<br />

ρq (Z<br />

(4.3.14) ⎝ q − 1) C n (2ζ n t)<br />

− 1) 1/4<br />

V n (2ζ n t)<br />

4 n 1/2 ,<br />

⎠<br />

4ρ q n 1/4 −→ (,Ö0 ),<br />

n→∞<br />

uniformly on [0,1], a.s.<br />

0≤t≤1<br />

0≤t≤1<br />

Then if (T ,U) ∈ Ω and v 0 ∈ T , we introduce a new spatial tree (̂T (v 0) ,Ũ(v 0) ) by s<strong>et</strong>ting for<br />

every w ∈ ̂T (v 0)<br />

Ũ (v 0)<br />

w = U w − U v0 ,<br />

where w is the vertex corresponding to w in the initial tree (in contrast with the definition of<br />

Û (v 0)<br />

w , Ũ(v 0)<br />

v does not necesseraly coincide with Ũ(v 0)<br />

ˇv when v is of type 1). We denote by Ṽ (v0) the<br />

spatial contour function of (̂T (v0),Ũ(v0) ), and we s<strong>et</strong> Ṽ (v0) (t) = 0 for t ∈ [2(#̂T (v0) − 1),2(#T −<br />

1)]. Note that, if w ∈ ̂T (v0) is either a vertex of type 0 or a vertex of type 1 which does not<br />

belong to the ancestral line of ̂v 0 , then<br />

Û (v 0)<br />

w = Ũ(v 0)<br />

w ,<br />

whereas if w ∈ ̂T (v 0),1 belongs to the ancestral line of ̂v 0 , then<br />

Û (v 0)<br />

w = Ũ(v 0)<br />

ˇw .<br />

Then we have<br />

∣ ∣∣ (4.3.15) sup Û w − Ũw∣ ≤<br />

w∈ bT (v 0 )<br />

sup<br />

w∈T \{∅}<br />

|U w − U ˇw |.<br />

Write vm n for the first vertex realizing the minimal spatial position in T n. In the same way as<br />

in the derivation of (18) in the proof of Proposition 6.1 in [39], it follows from (4.3.14) that<br />

⎛(√ ) (<br />

ρq (Z<br />

⎝ q − 1) Ĉ (vn m<br />

n )<br />

(9(Zq ) ) ⎞<br />

(2ζ n t)<br />

− 1) 1/4 (v Ṽ<br />

m n n ) (2ζ n t)<br />

4 n 1/2 ,<br />

⎠<br />

4ρ q n 1/4 −→ (0 ,Ö0 ),<br />

n→∞<br />

0≤t≤1<br />

0≤t≤1<br />

uniformly on [0,1], a.s., where the conditioned pair (0,Ö0 ) is constructed from the unconditioned<br />

one (,Ö0 ) as explained in section 4.2.3. L<strong>et</strong> ε > 0. We deduce from (4.3.15) that<br />

(<br />

∣ ) (<br />

)<br />

P ′ ̂V (vn m)<br />

n (2ζ n t)<br />

sup<br />

∣ n 1/4 − Ṽ (vn m)<br />

n (2ζ n t) ∣∣∣∣<br />

n 1/4 > ε ≤ Q n |U w − U ˇw |<br />

µ,ν<br />

n 1/4 > ε ,<br />

t∈[0,1]<br />

sup<br />

w∈T \{∅}<br />

where we have written P ′ for the probability measure under which the sequence (T n ,U n ) n≥1 and<br />

the Brownian snake (,Ö0 ) are defined. From (4.3.12) we g<strong>et</strong><br />

(<br />

∣ )<br />

P ′ ̂V (vn m<br />

n ) (2ζ n t)<br />

sup<br />

∣ n 1/4 − Ṽ (vn m<br />

n ) (2ζ n t) ∣∣∣∣<br />

n 1/4 > ε −→ 0,<br />

n→∞<br />

t∈[0,1]<br />

and the desired result follows.<br />

The following proposition can be proved using Lemma 4.3.13 and Lemma 4.3.15 in the same<br />

way as Proposition 6.1 in [39].<br />

□<br />

108


Proposition 4.3.16. For every b > 0 and ε ∈ (0,1/10), we can find α,δ ∈ (0,ε) such that<br />

for all n sufficiently large,<br />

(<br />

)<br />

(<br />

)<br />

Q n µ,ν inf V (n) (t) ≥ 2α, sup C (n) (t) + V (n) (t) ≤ ε ≥ 1 − b.<br />

t∈[δ/2,1−δ/2] 2<br />

t∈[0,4δ]∪[1−4δ,1]<br />

Consequently, if K > 0, we have also for all n sufficiently large, for every x ∈ [0,K],<br />

(<br />

)<br />

(<br />

)<br />

P n µ,ν,x inf V (n) (t) ≥ α, sup C (n) (t) + V (n) (t) ≤ ε ≥ 1 − γ 3 b,<br />

t∈[δ,1−δ]<br />

t∈[0,3δ]∪[1−3δ,1]<br />

where the constant γ 3 only depends on µ,ν,K.<br />

4.3.4. Proof of Theorem 4.3.3. The proof below is similar to Section 7 in [39]. We<br />

provide d<strong>et</strong>ails because the fact that we deal with two-type trees creates nontrivial additional<br />

difficulties.<br />

On a suitable probability space (Ω,P) we can define a collection of processes (e,r z ) z≥0 such<br />

that (e,r z ) is a Brownian snake with initial point z for every z ≥ 0. Recall from section 4.2.3<br />

the definition of (e z ,r z ) and the construction of the conditioned Brownian snake (e 0 ,r 0 ).<br />

Recall that C([0,1], R) 2 is equipped with the norm ‖(f,g)‖ = ‖f‖ u ∨ ‖g‖ u . For every f ∈<br />

C([0,1], R) and r > 0, we s<strong>et</strong><br />

ω f (r) =<br />

sup |f(s) − f(t)|.<br />

s,t∈[0,1],|t−s|≤r<br />

L<strong>et</strong> x ≥ 0 be fixed throughout this section and l<strong>et</strong> F be a bounded Lipschitz function. We<br />

have to prove that ( (<br />

E n µ,ν,x F C (n) ,V (n))) −→ E ( F(e 0 ,r 0 ) ) .<br />

n→∞<br />

We may and will assume that 0 ≤ F ≤ 1 and that the Lipschitz constant of F is less than 1.<br />

The first lemma we have to prove gives a spatial Markov property for our spatial trees. We<br />

use the notation of section 5 in [39]. L<strong>et</strong> recall briefly this notation. We fix a > 0. If (T ,U)<br />

is a mobile and v ∈ T , we say that v is an exit vertex from (−∞,a) if U v ≥ a and U v ′ < a<br />

for every ancestor v ′ of v distinct from v. Notice that, since U v = Uˇv for every v ∈ T 1 , an exit<br />

vertex is necessarily of type 0. We denote by v 1 ,... ,v M the exit vertices from (−∞,a) listed in<br />

lexicographical order. For v ∈ T , recall that T [v] = {w ∈ U : vw ∈ T }. For every w ∈ T [v] we<br />

s<strong>et</strong><br />

U [v]<br />

w = U vw = U w [v] + U v .<br />

At last, we denote by T a the subtree of T consisting of those vertices which are not strict<br />

descendants of v 1 ,...,v M . Note in particular that v 1 ,...,v M ∈ T a . We also write U a for the<br />

restriction of U to T a . The tree (T a ,U a ) corresponds to the tree (T ,U) which has been truncated<br />

at the first exit time from (−∞,a). The following lemma is an easy application of classical<br />

properties of Galton-Watson trees. We leave d<strong>et</strong>ails of the proof to the reader.<br />

Lemma 4.3.17. L<strong>et</strong> x ∈ [0,a) and p ∈ {1,... ,n}. L<strong>et</strong> n 1 ,...,n p be positive integers such that<br />

n 1 + ... + n p ≤ n. Assume that<br />

)<br />

(M = p, #T [v1],1 = n 1 ,..., #T [vp],1 = n p > 0.<br />

P n µ,ν,x<br />

109


Then, under the probability measure Pµ,ν,x(· n | M = p, #T [v1],1 = n 1 ,... , #T [vp],1 = n p ), and<br />

conditionally on (T a ,U a ), the spatial trees<br />

( ) T [v1] ,U [v 1]<br />

,... ,<br />

(T [vp] ,U [vp])<br />

are independent and distributed respectively according to P n µ,ν,U v1<br />

,...,P n µ,ν,U vp<br />

.<br />

The next lemma is analogous to Lemma 7.1 in [39] and can be proved in the same way using<br />

Theorem 4.3.1.<br />

Lemma 4.3.18. L<strong>et</strong> 0 < c ′ < c ′′ . Then<br />

(<br />

F<br />

∣ ∣∣E n<br />

sup µ,ν,y<br />

c ′ n 1/4 ≤y≤c ′′ n 1/4<br />

where B = (9(Z q − 1)/(4ρ q )) 1/4 .<br />

(<br />

C (n) ,V (n))) − E<br />

( (<br />

F e By/n1/4 ,r By/n1/4))∣ ∣ −→ 0,<br />

n→∞<br />

We can now follow the lines of section 7 in [39]. L<strong>et</strong> b > 0. We will prove that for n sufficiently<br />

large,<br />

∣ ( ( ∣∣E n<br />

µ,ν,x F C (n) ,V (n))) − E ( F ( e 0 ,r 0))∣ ∣ ≤ 17b.<br />

We can choose ε ∈ (0,b ∧ 1/10) in such a way that<br />

(4.3.16)<br />

∣ E (F (e z ,r z )) − E ( F ( e 0 ,r 0))∣ ∣ ≤ b,<br />

for every z ∈ (0,2ε). By taking ε smaller if necessary, we may also assume that,<br />

(( ) )<br />

E 3ε sup e 0 (t) ∧ 1 ≤ b, E (ω e 0(6ε) ∧ 1) ≤ b,<br />

(4.3.17)<br />

0≤t≤1<br />

((<br />

E 3ε sup r 0 (t)<br />

0≤t≤1<br />

For α,δ > 0, we denote by Γ n = Γ n (α,δ) the event<br />

{<br />

Γ n =<br />

)<br />

inf V (n) (t) ≥ α,<br />

t∈[δ,1−δ]<br />

)<br />

∧ 1 ≤ b, E (ω r 0(6ε) ∧ 1) ≤ b.<br />

sup<br />

t∈[0,3δ]∪[1−3δ,1]<br />

}<br />

(<br />

)<br />

C (n) (t) + V (n) (t) ≤ ε .<br />

From Proposition 4.3.16, we may fix α,δ ∈ (0,ε) such that, for all n sufficiently large,<br />

(4.3.18) P n µ,ν,x(Γ n ) > 1 − b.<br />

We also require that δ satisfies the following bound<br />

(4.3.19) 4δ(m 1 + 1) < 3ε.<br />

Recall the notation ζ = #T − 1 and B = (9(Z q − 1)/(4ρ q )) 1/4 . On the event Γ n , we have for<br />

every t ∈ [2ζδ,2ζ(1 − δ)],<br />

(4.3.20) V (t) ≥ αB −1 n 1/4 .<br />

This incites us to apply Lemma 4.3.17 with a n = αn 1/4 , where α = αB −1 . Once again, we use<br />

the notation of [39]. We write v1 n,... ,vn M n<br />

for the exit vertices from (−∞,αn 1/4 ) of the spatial<br />

tree (T ,U), listed in lexicographical order. Consider the spatial trees<br />

(<br />

T [vn 1 ] ,U [vn 1 ]) ,... ,<br />

(<br />

T [vn Mn ] ,U [vn Mn ]) .<br />

110


The contour functions of these spatial trees can be obtained in the following way. S<strong>et</strong><br />

{<br />

k1 n = inf k ≥ 0 : V (k) ≥ αn 1/4} , l1 n = inf {k ≥ k1 n : C(k + 1) < C(k1)} n ,<br />

and by induction on i,<br />

{<br />

ki+1 n = inf k > li n : V (k) ≥ αn 1/4} , li+1 n = inf { k ≥ ki+1 n : C(k + 1) < C(ki n ) } .<br />

Then ki n ≤ li n < ∞ if and only if i ≤ M n . Furthermore, (C(ki n + t) − C(kn i ),0 ≤ t ≤ ln i − kn i ) is<br />

the contour function of T [vn i ] and (V (ki n + t),0 ≤ t ≤ ln i − kn i ) is the spatial contour function of<br />

(T [vn i ] ,U [vn i ] ). Using (4.3.20), we see that on the event Γ n , all integer points of [2ζδ,2ζ(1 − δ)]<br />

must be contained in a single interval [ki n,ln i ], so that for this particular interval we have<br />

li n − ki n ≥ 2ζ(1 − δ) − 2ζδ − 2 ≥ 2ζ(1 − 3δ),<br />

if n is sufficiently large, P n µ,ν,x a.s. Hence if<br />

E n = {∃i ∈ {1,... ,M n } : l n i − kn i > 2ζ(1 − 3δ)} ,<br />

then, for all n sufficiently large, Γ n ⊂ E n so that<br />

(4.3.21) P n µ,ν,x(E n ) > 1 − b.<br />

As in [39], on the event E n , we denote by i n the unique integer i ∈ {1,... ,M n } such that<br />

l n i −kn i > 2ζ(1 −3δ). We also define ζ n = #T [vn in ] −1 and Y n = U v n<br />

in<br />

. Note that ζ n = (l n i −kn i )/2<br />

so that ζ n > ζ(1 − 3δ). Furthermore, we s<strong>et</strong><br />

p n = #T [vn in ],1 .<br />

We need to prove a lemma providing an estimate of the probability for p n to be close to n. Note<br />

that p n ≤ n = #T 1 , P n µ,ν,x a.s. Recall that m 1 denotes the mean of µ 1 .<br />

Lemma 4.3.19. For every n sufficiently large,<br />

P n µ,ν,x (Γ n ∩ {p n ≥ (1 − 4δ(m 1 + 1))n}) ≥ 1 − 2b.<br />

Proof : In the same way as in the proof of the bound (4.3.1) we can verify that there exists<br />

a constant β 3 > 0 such that for all n sufficiently large,<br />

)<br />

P µ<br />

(|ζ − (m 1 + 1)n| > n 3/4 ,#T 1 = n ≤ e −nβ3 .<br />

So Lemma 4.3.9 and Proposition 4.3.12 imply that for all n sufficiently large,<br />

(<br />

(4.3.22) P n µ,ν,x |ζ − (m 1 + 1)n| > n 3/4) ≤ b.<br />

Now, on the event Γ n , we have<br />

((<br />

n − p n = # T \ T [vn ]) in ∩ T 1) (<br />

≤ # T \ T [vn ]) in = ζ − ζ n ≤ 3δζ,<br />

since we saw that Γ n ⊂ E n and that ζ n > (1 − 3δ)ζ on E n . If n is sufficiently large, we have<br />

3δ(m 1 + 1)n + 3δn 3/4 ≤ 4δ(m 1 + 1)n so we obtain that<br />

( {<br />

Γ n ∩ ζ ≤ (m 1 + 1)n + n 3/4}) ⊂ (Γ n ∩ {p n ≥ (1 − 4δ(m 1 + 1))n}) ,<br />

for all n sufficiently large. The desired result then follows from (4.3.18) and (4.3.22).<br />

□<br />

111


L<strong>et</strong> us now define on the event E n , for every t ∈ [0,1],<br />

√<br />

˜C (n) ρq (Z q − 1) C ( ki n (t) =<br />

n<br />

+ 2ζ n t ) − C ( k n )<br />

i n<br />

4<br />

p 1/2 ,<br />

n<br />

( )<br />

Ṽ (n) 9(Zq − 1) 1/4<br />

V ( ki n (t) =<br />

n<br />

+ 2ζ n t )<br />

.<br />

4ρ q<br />

Note that ˜C (n) and Ṽ (n) are rescaled versions of the contour and the spatial contour functions<br />

of (T [vn in ] ,U [vn in ] ). On the event En, c we take ˜C (n) (t) = Ṽ (n) (t) = 0 for every t ∈ [0,1]. Straightforward<br />

calculations show that on the event Γ n , for every t ∈ [0,1],<br />

S<strong>et</strong><br />

∣<br />

∣C (n) (t) − ˜C (n) (t) ∣ ≤ ε +<br />

∣<br />

∣V (n) (t) − Ṽ (n) (t) ∣ ≤ ε +<br />

(<br />

(<br />

1 − p1/2 n<br />

n 1/2 )<br />

)<br />

1 − p1/4 n<br />

n 1/4<br />

p 1/4<br />

n<br />

˜Γ n = Γ n ∩ {p n ≥ (1 − 4δ(m 1 + 1))n}.<br />

We then g<strong>et</strong> that on the event ˜Γ n , for every t ∈ [0,1],<br />

∣<br />

∣C (n) (t) − ˜C (n) (4.3.23)<br />

(t) ∣ ≤<br />

(4.3.24)<br />

Likewise, we s<strong>et</strong><br />

∣<br />

∣V (n) (t) − Ṽ (n) (t) ∣ ≤<br />

sup ˜C (n) (s) + ω eC (n)(6δ),<br />

s∈[0,1]<br />

sup Ṽ (n) (s) + ω eV (n)(6δ).<br />

s∈[0,1]<br />

ε + 4δ(m 1 + 1) sup ˜C (n) (s) + ω eC (n)(6δ),<br />

s∈[0,1]<br />

ε + 4δ(m 1 + 1) sup Ṽ (n) (s) + ω eV (n)(6δ).<br />

s∈[0,1]<br />

Ẽ n = E n ∩ {p n ≥ (1 − 4δ(m 1 + 1))n}.<br />

Lemma 4.3.17 implies that, under the probability measure Pµ,ν,x(· n | Ẽ n ) and conditionally on<br />

the σ-field G n defined by<br />

((<br />

G n = σ T αn1/4 ,U αn1/4) ))<br />

, M n ,<br />

(#T [vn i ],1 , 1 ≤ i ≤ M n ,<br />

the spatial tree (T [vn in ] ,U [vn in ] ) is distributed according to P pn<br />

µ,ν,Y n<br />

(recall that Y n = U v n<br />

in<br />

). Note<br />

that Ẽn ∈ G n , and that Y n and p n are G n -measurable. Thus we have<br />

( (½e En<br />

Eµ,ν,x(<br />

n ˜C(n) ,Ṽ (n)) ∣ ))<br />

∣ Gn<br />

(4.3.25)<br />

(<br />

E n µ,ν,x<br />

(½e En<br />

F ˜C(n) ,Ṽ (n))) = E n µ,ν,x<br />

= E n µ,ν,x<br />

F<br />

(<br />

(½e En<br />

E p µ,ν,Y n<br />

F<br />

(C (p) ,V (p))) p=p n<br />

)<br />

.<br />

From Lemma 4.3.18, we g<strong>et</strong> for every p sufficiently large,<br />

∣ ( ( ∣∣E p<br />

sup µ,ν,y F C (p) ,V (p))) ( (<br />

− E F e By/p1/4 ,r By/p1/4))∣ ∣ ≤ b,<br />

α<br />

2 p1/4 ≤y≤ 3α 2 p1/4<br />

which implies using (4.3.16), since 3αB/2 ≤ 2αB = 2α < 2ε, that for every p sufficiently large,<br />

∣ ∣∣E p<br />

(4.3.26) sup µ,ν,y<br />

α<br />

2 p1/4 ≤y≤ 3α 2 p1/4<br />

(<br />

F<br />

(<br />

C (p) ,V (p))) − E ( F ( e 0 ,r 0))∣ ∣ ∣ ≤ 2b.<br />

112


Furthermore Lemma 4.3.14 implies that<br />

)<br />

P n µ,ν,x<br />

({|n −1/4 Y n − α| > η} ∩ E n −→ 0.<br />

→∞<br />

So we g<strong>et</strong> for every n sufficiently large,<br />

(<br />

)<br />

∣ E n µ,ν,x<br />

(½e En<br />

E p µ,ν,Y n<br />

F<br />

(C (p) ,V (p))) − P n µ,ν,x<br />

p=p n<br />

{<br />

})<br />

≤ 2P n µ,ν,x<br />

(Ẽn ∩ |n −1/4 Y n − α| > α/4<br />

⎛<br />

+ E n µ,ν,x<br />

(4.3.27) ≤ 2b + E n µ,ν,x<br />

∣<br />

⎝ ∣∣E p<br />

sup<br />

α<br />

2 p1/4 ≤y≤ 3α 2 p1/4<br />

⎛<br />

⎝½e En<br />

⎛<br />

⎝½eE n<br />

⎛<br />

µ,ν,y<br />

∣<br />

⎝ ∣∣E p<br />

sup<br />

α<br />

2 p1/4 ≤y≤ 3α 2 p1/4<br />

(Ẽn<br />

)<br />

E ( F ( e 0 ,r 0))∣ ∣ ∣∣<br />

⎞<br />

( (<br />

F C (p) ,V (p))) − E ( F ( e 0 ,r 0))∣ ∣ ⎠<br />

µ,ν,y<br />

⎞<br />

⎠<br />

p=p n<br />

⎞<br />

( (<br />

F C (p) ,V (p))) − E ( F ( e 0 ,r 0))∣ ∣ ⎠<br />

⎠ .<br />

p=p n<br />

⎞<br />

Thus we use (4.3.25), (4.3.26), (4.3.27) and the fact that p n ≥ 1 − 4δ(m 1 + 1)n on Ẽn, to obtain<br />

that for every n sufficiently large,<br />

(<br />

∣<br />

(4.3.28) ∣E n µ,ν,x<br />

(½eE n<br />

F ˜C(n) ,Ṽ (n))) − Pµ,ν,x(Ẽn)E n ( F ( e 0 ,r 0))∣ ∣ ≤ 4b.<br />

From Lemma 4.3.19, we have Pµ,ν,x(Ẽn) n ≥ 1 −2b. Furthermore, 0 ≤ F ≤ 1 so that (4.3.28) gives<br />

(<br />

∣<br />

(4.3.29)<br />

∣E n µ,ν,x F( ˜C<br />

)<br />

(n) ,Ṽ (n) ) − E(F(e 0 ,r 0 )) ∣ ≤ 8b.<br />

On the other hand, since ˜Γ n ⊂ Ẽ n and F is a Lipschitz function whose Lipschitz constant is less<br />

than 1, we have using (4.3.23) and (4.3.24), for n sufficiently large,<br />

∣ (<br />

E n ∣∣F<br />

µ,ν,x<br />

(½e ˜C(n) Γn<br />

,Ṽ (n)) − F<br />

(C (n) ,V (n))∣ )<br />

∣ ((<br />

)<br />

)<br />

(4.3.30)<br />

≤<br />

2ε + E n µ,ν,x<br />

+ E n µ,ν,x<br />

((<br />

4δ(m 1 + 1) sup<br />

s∈[0,1]<br />

˜C (n) (s)<br />

4δ(m 1 + 1) sup Ṽ (n) (s)<br />

s∈[0,1]<br />

)<br />

∧ 1 + ω eC (n)(6δ) ∧ 1<br />

∧ 1 + ω eV (n)(6δ) ∧ 1<br />

By the same arguments we used to derive (4.3.29), we can bound the right-hand side of (4.3.30),<br />

for n sufficiently large, by<br />

((<br />

)<br />

)<br />

b + 2ε + E<br />

+ E<br />

4δ(m 1 + 1) sup<br />

((<br />

s∈[0,1]<br />

e 0 (s)<br />

4δ(m 1 + 1) sup r 0 (s)<br />

s∈[0,1]<br />

∧ 1 + ω e 0(6δ) ∧ 1<br />

)<br />

∧ 1 + ω r 0(6δ) ∧ 1<br />

From (4.3.17) tog<strong>et</strong>her with (4.3.19), the latter quantity is bounded above by 7b. Since<br />

) (˜Γn ≥ 1 − 2b,<br />

P n µ,ν,x<br />

we g<strong>et</strong> for all n sufficiently large,<br />

(∣ (<br />

E n ∣∣F ˜C(n) µ,ν,x ,Ṽ (n)) − F<br />

(C (n) ,V (n))∣ )<br />

∣ ≤ 9b,<br />

)<br />

.<br />

)<br />

.<br />

113


which implies tog<strong>et</strong>her with (4.3.29) that for all n sufficiently large,<br />

( (<br />

∣<br />

F C (n) ,V (n))) − E(F ( e 0 ,r 0 ) )∣ ∣ ≤ 17b.<br />

∣E n µ,ν,x<br />

This compl<strong>et</strong>es the proof of Theorem 4.3.3<br />

4.3.5. Proof of Theorem 4.2.5. In this section we derive Theorem 4.2.5 from Theorem<br />

4.3.3. We first need to prove a lemma. Recall that if T ∈ T, we s<strong>et</strong> ζ = #T − 1 and we<br />

denote by v(0) = ∅ ≺ v(1) ≺ ... ≺ v(ζ) the list of vertices of T in lexicographical order. For<br />

n ∈ {0,1,... ,ζ}, we s<strong>et</strong> as in [43],<br />

J T (n) = # ( T 0 ∩ {v(0),v(1),... ,v(n)} ) .<br />

We extend J T to the real interval [0,ζ] by s<strong>et</strong>ting J T (t) = J T (⌊t⌋) for every t ∈ [0,ζ], and we<br />

s<strong>et</strong> for every t ∈ [0,1]<br />

J T (t) = J T (ζt)<br />

#T 0 .<br />

We also define for k ∈ {0,1,... ,2ζ},<br />

{<br />

}<br />

K T (k) = 1 + # l ∈ {1,... ,k} : C(l) = max C and C(l) is even .<br />

[l−1,l]<br />

Note that K T (k) is the number of vertices of type 0 in the search-depth sequence up to time k.<br />

As previously, we extend K T to the real interval [0,2ζ] by s<strong>et</strong>ting K T (t) = K T (⌊t⌋) for every<br />

t ∈ [0,2ζ], and we s<strong>et</strong> for every t ∈ [0,1]<br />

(4.3.31) P n µ,ν,1<br />

K T (t) = K T (2ζt)<br />

#T 0 .<br />

Lemma 4.3.20. The law under P n µ,ν,1 of ( J T (t),0 ≤ t ≤ 1 ) converges as n → ∞ to the Dirac<br />

mass at the identity mapping of [0,1]. In other words, for every η > 0,<br />

(<br />

∣ J T (t) − t ∣ )<br />

> η<br />

(4.3.32) P n µ,ν,1<br />

sup<br />

t∈[0,1]<br />

sup<br />

t∈[0,1]<br />

−→ 0.<br />

n→∞<br />

Consequently, the law under P n µ,ν,1 of ( K T (t),0 ≤ t ≤ 1 ) converges as n → ∞ to the Dirac mass<br />

at the identity mapping of [0,1]. In other words, for every η > 0,<br />

(<br />

∣ KT (t) − t ∣ )<br />

> η<br />

−→ 0.<br />

n→∞<br />

Proof : For T ∈ T, we l<strong>et</strong> v 0 (0) = ∅ ≺ v 0 (1) ≺ ... ≺ v 0 (#T 0 − 1) be the list of vertices of<br />

T of type 0 in lexicographical order. We define as in [43]<br />

G T (k) = # { u ∈ T : u ≺ v 0 (k) } , 0 ≤ k ≤ #T 0 − 1,<br />

and we s<strong>et</strong> G T (#T 0 ) = ζ. Note that v 0 (k) does not belong to the s<strong>et</strong> {u ∈ T : u ≺ v 0 (k)}.<br />

Recall that m 0 denotes the mean of the offspring distribution µ 0 . From the second assertion of<br />

Lemma 18 in [43] there exists a constant ε > 0 such that for n sufficiently large,<br />

(<br />

)<br />

Pµ<br />

n sup |G T (k) − (1 + m 0 )k| ≥ n 3/4 ≤ e −nε .<br />

0≤k≤#T 0<br />

114


Then Lemma 4.3.9 and Proposition 4.3.12 imply that there exists a constant ε ′ > 0 such that<br />

for n sufficiently large,<br />

(<br />

)<br />

(4.3.33) P n µ,ν,1 sup |G T (k) − (1 + m 0 )k| ≥ n 3/4 ≤ e −nε′ .<br />

0≤k≤#T 0<br />

From our definitions, we have for every 0 ≤ k ≤ #T 0 − 1 and 0 ≤ n ≤ ζ,<br />

{G T (k) > n} = {J T (n) ≤ k} .<br />

It then follows from (4.3.33) that, for every η > 0<br />

(<br />

P n µ,ν,1<br />

n −1<br />

sup |J T (((1 + m 0 )k) ∧ ζ) − k| > η<br />

0≤k≤#T 0<br />

Also from the bound (4.3.1) of Lemma 4.3.10 we g<strong>et</strong> for every η > 0,<br />

(∣ ∣∣∣<br />

P n µ,ν,1 n −1 #T 0 − 1 ∣ ) ∣∣∣<br />

> η −→<br />

m 0.<br />

0 n→∞<br />

)<br />

−→ 0.<br />

n→∞<br />

The first assertion of Lemma 4.3.20 follows from the last two convergences.<br />

L<strong>et</strong> us s<strong>et</strong> j n = 2n − |v(n)| for n ∈ {0,... ,ζ}. It is well known and easy to check by induction<br />

that j n is the first time at which v(n) appears in the search-depth sequence. It is also convenient<br />

to s<strong>et</strong> j ζ+1 = 2ζ. Then we have K T (k) = J T (n) for every k ∈ {j n ,... ,j n+1 − 1} and every<br />

n ∈ {0,... ,ζ}. L<strong>et</strong> us define a random function ϕ : [0,2ζ] −→ Z + by s<strong>et</strong>ting ϕ(t) = n if<br />

t ∈ [j n ,j n+1 ) and 0 ≤ n ≤ ζ, and ϕ(2ζ) = ζ. From our definitions, we have for every t ∈ [0,2ζ],<br />

(4.3.34) K T (t) = J T (ϕ(t)).<br />

Furthermore, we easily check from the equality j n = 2n − |v(n)| that<br />

(4.3.35) sup<br />

∣ ϕ(t) − t 2∣ ≤ max C.<br />

[0,2ζ]<br />

t∈[0,2ζ]<br />

We s<strong>et</strong> ϕ ζ (t) = ζ −1 ϕ(2ζt) for t ∈ [0,1]. So (4.3.35) gives<br />

which implies that for every η > 0,<br />

(<br />

(4.3.36) P n µ,ν,1<br />

sup<br />

t∈[0,1]<br />

sup |ϕ ζ (t) − t| ≤ 1<br />

t∈[0,1] ζ max C,<br />

[0,2ζ]<br />

sup |ϕ ζ (t) − t| > η<br />

t∈[0,1]<br />

)<br />

−→ 0.<br />

n→∞<br />

On the other hand we g<strong>et</strong> from (4.3.34) that K T (t) = J T (ϕ ζ (t)) for every t ∈ [0,1] and thus<br />

∣ KT (t) − J T (t) ∣ ≤ 2 sup ∣ J T (t) − t ∣ + sup |ϕ ζ (t) − t|.<br />

t∈[0,1]<br />

The desired result then follows from (4.3.31) and (4.3.36).<br />

t∈[0,1]<br />

We are now able to compl<strong>et</strong>e the proof of Theorem 4.2.5. The proof of (i) is similar to the<br />

proof of the first part of Theorem 8.2 in [39], and is therefore omitted.<br />

L<strong>et</strong> us turn to (ii). By Corollary 4.2.3 and properties of the Bouttier-di Francesco-Guitter<br />

bijection, the law of λ (n)<br />

M<br />

under Br q (· | #F M = n) is the law under P n µ,ν,1 of the probability<br />

□<br />

115


measure I n defined by<br />

〈I n ,g〉 =<br />

⎛<br />

1<br />

⎝g(0)<br />

#T 0 + ∑ ( ) ⎞<br />

g n −1/4 U v<br />

⎠.<br />

+ 1<br />

v∈T 0<br />

It is more convenient for our purposes to replace I n by a new probability measure I n ′ defined by<br />

〈I n ′ ,g〉 = 1 ∑ )<br />

#T 0 g<br />

(n −1/4 U v .<br />

v∈T 0<br />

L<strong>et</strong> g be a bounded continuous function. Clearly, we have for every η > 0,<br />

(4.3.37) P n (∣ ∣〈In<br />

µ,ν,1 ,g〉 − 〈I n,g〉 ′ ∣ ) > η −→ 0.<br />

n→∞<br />

Furthermore, we have from our definitions<br />

(4.3.38) 〈I ′ n,g〉 = 1<br />

#T 0 g (<br />

n −1/4) +<br />

∫ 1<br />

0<br />

( )<br />

g n −1/4 V (2ζt) dK T (t),<br />

where the first term in the right-hand side corresponds to v = ∅ in the definition of I n ′ . Then<br />

from Theorem 4.3.3, (4.3.32) and the Skorokhod representation theorem, we can construct on<br />

a suitable probability space, a sequence (T n ,U n ) n≥1 and a conditioned Brownian snake (0,Ö0 ),<br />

such that each pair (T n ,U n ) is distributed according to P n µ,ν,1 , and such that if we write (C n,V n )<br />

for the contour functions of (T n ,U n ), ζ n = #T n − 1 and K n = K Tn , we have,<br />

( )<br />

Vn (2ζ n t)<br />

,K n (t)<br />

n 1/4<br />

−→<br />

n→∞<br />

( ( 4ρq<br />

9(Z q − 1)<br />

) 1/4Ö0 (t), t)<br />

uniformly in t ∈ [0,1], a.s. Now g is Lipschitz, which implies that a.s.,<br />

∫ 1 ( ) ∫ (<br />

1<br />

( ) 1/4Ö0 (4.3.39)<br />

g n −1/4 4ρq<br />

V<br />

∣<br />

n (2ζ n t) dK n (t) − g<br />

(t))<br />

dK n (t)<br />

9(Z q − 1)<br />

∣ −→<br />

0<br />

0<br />

,<br />

n→∞ 0.<br />

Furthermore, the sequence of measures dK n converges weakly to the uniform measure dt on<br />

[0,1] a.s., so that a.s.<br />

∫ (<br />

1<br />

( ) 1/4Ö0 ∫ (<br />

1<br />

( ) 4ρq<br />

4ρq<br />

(4.3.40) g<br />

(t))<br />

dK n (t) −→ g<br />

(t)) 1/4Ö0 dt.<br />

9(Z q − 1)<br />

n→∞ 9(Z q − 1)<br />

0<br />

Then (4.3.39) and (4.3.40) imply that a.s.,<br />

∫ 1 ( )<br />

g n −1/4 V n (2ζ n t) dK n (t) −→<br />

0<br />

n→∞<br />

∫ 1<br />

which tog<strong>et</strong>her with (4.3.37) and (4.3.38) yields the desired result.<br />

0<br />

g<br />

0<br />

( ( ) 4ρq<br />

(t)) 1/4Ö0 dt,<br />

9(Z q − 1)<br />

Finally, the proof of (iii) from (ii) is similar to the proof of the third part of Theorem 8.2 in<br />

[39]. This compl<strong>et</strong>es the proof of Theorem 4.2.5.<br />

116


4.4. Separating vertices in a 2κ-angulation<br />

In this section, we use the estimates of Proposition 4.3.12 to derive a result concerning<br />

separating vertices in rooted 2κ-angulations. Recall that in a 2κ-angulation, all faces have a<br />

degree equal to 2κ.<br />

L<strong>et</strong> M be a planar map and l<strong>et</strong> σ 0 ∈ V M . L<strong>et</strong> σ be a vertex of M different from σ 0 . We<br />

denote by S σ 0,σ<br />

M the s<strong>et</strong> of all vertices a of M such that any path from σ to a goes through σ 0.<br />

The vertex σ 0 is called a separating vertex of M if there exists a vertex σ of M different from<br />

σ 0 such that S σ 0,σ<br />

M ≠ {σ 0}. We denote by D M the s<strong>et</strong> of all separating vertices of M.<br />

Recall that U n κ stands for the uniform probability measure on the s<strong>et</strong> of all rooted 2κangulations<br />

with n faces. Our goal is to prove the following theorem.<br />

Theorem 4.4.1. For every ε > 0,<br />

lim<br />

n→∞ U n κ<br />

(<br />

∃ σ 0 ∈ D M : ∃ σ ∈ V M \ {σ 0 }, n 1/2−ε ≤ #S σ 0,σ<br />

M ≤ 2n1/2−ε) = 1.<br />

Theorem 4.4.1 is a consequence of the following theorem. Recall that U n κ denotes the uniform<br />

probability measure on the s<strong>et</strong> of all rooted pointed 2κ-angulations with n faces. If M is a rooted<br />

pointed bipartite planar map, we denote by τ its distinguished point.<br />

Theorem 4.4.2. For every ε > 0,<br />

lim<br />

n→∞ Un κ<br />

(<br />

∃ σ 0 ∈ D M : σ 0 ≠ τ, n 1/2−ε ≤ #S σ 0,τ<br />

M ≤ 2n1/2−ε) = 1.<br />

Theorem 4.4.1 can be deduced from Theorem 4.4.2 but not as directly as one could think.<br />

Indeed the canonical surjection from the s<strong>et</strong> of rooted pointed 2κ-angulations with n faces onto<br />

the s<strong>et</strong> of rooted 2κ-angulations with n faces does not map the uniform measure U n κ to the<br />

uniform measure U n κ . Nevertheless a simple argument allows us to circumvent this difficulty.<br />

L<strong>et</strong> ˜M r,p be the s<strong>et</strong> of all triples (M,⃗e,τ) where (M,⃗e ) ∈ M r and τ is a distinguished vertex<br />

of the map M. We denote by s the canonical surjection from the s<strong>et</strong> ˜M r,p onto the s<strong>et</strong> M r,p<br />

which is obtained by “forg<strong>et</strong>ting” the orientation of ⃗e. We observe that for every (M,e,τ) ∈ M r,p<br />

# ( s −1 ((M,e,τ)) ) = 2.<br />

Denote by Ũ κ n the uniform measure on the s<strong>et</strong> of all triples (M,⃗e,τ) ∈ ˜M r,p such that M is a<br />

2κ-angulation with n faces. Then the image measure of the measure Ũ κ n under the mapping s is<br />

the measure U n κ. Thus we obtain from Theorem 4.4.2 that<br />

(4.4.1) lim<br />

n→∞ Ũ n κ<br />

(<br />

∃ σ 0 ∈ D M : ∃ σ ∈ V M \ {σ 0 }, n 1/2−ε ≤ #S σ 0,σ<br />

M ≤ 2n1/2−ε) = 1.<br />

On the other hand l<strong>et</strong> p be the canonical projection from the s<strong>et</strong> ˜M r,p onto the s<strong>et</strong> M r . If M<br />

is a 2κ-angulation with n faces, we have thanks to Euler formula<br />

#V M = (κ − 1)n + 2.<br />

Thus the image measure of the measure Ũ n κ under the mapping p is the measure U n κ . This<br />

remark tog<strong>et</strong>her with (4.4.1) implies Theorem 4.4.1.<br />

The remainder of this section is devoted to the proof of Theorem 4.4.2. We first need to state<br />

a lemma. Recall the definition of the spatial tree (T [v] ,U [v] ) for (T ,U) ∈ Ω and v ∈ T .<br />

117


Lemma 4.4.3. L<strong>et</strong> (T ,U) ∈ T mob<br />

1 and l<strong>et</strong> M = Ψ r,p ((T ,U)). Suppose that we can find v ∈ T 0<br />

such that T [v],0 ≠ {∅} and U [v] > 0. Then there exists σ 0 ∈ D M such that σ 0 ≠ τ and<br />

#S σ 0,τ<br />

M = #T [v],0 .<br />

Proof : L<strong>et</strong> (T ,U) ∈ T mob<br />

1 . Write w 0 ,w 1 ,...,w ζ for the search-depth sequence of T 0 (see<br />

Section 4.2.4). Recall from Section 4.2.4 the definition of (U + v ,v ∈ T ) and the construction of<br />

the planar map Ψ r,p ((T ,U)). For every i ∈ {0,1,... ,ζ}, we s<strong>et</strong> s i = ∂ if U + w i<br />

= 1, whereas if<br />

U + w i<br />

≥ 2, we denote by s i the first vertex in the sequence w i+1 ,... ,w ζ−1 ,w 0 ,w 1 ,... ,w i−1 whose<br />

label is U + w i<br />

− 1.<br />

Suppose that there exists v ∈ T 0 such that T [v],0 ≠ {∅} and U [v] > 0. We s<strong>et</strong><br />

k = min{i ∈ {0,1,... ,ζ} : w i = v},<br />

l = max{i ∈ {0,1,... ,ζ} : w i = v}.<br />

The vertices w k ,w k+1 ,... ,w l are exactly the descendants of v in T 0 . The condition U [v] > 0<br />

ensures that for every i ∈ {k + 1,... ,l − 1}, we have<br />

U + w i<br />

> U + w l<br />

= U + w k<br />

.<br />

This implies that s i is a descendant of v for every i ∈ {k + 1,... ,l − 1}. Furthermore s k = s l<br />

and s i is not a strict descendant of v if i ∈ {0,1,... ,ζ} \ {k,k + 1... ,l}. From the construction<br />

of edges in the map Ψ r,p ((T ,U)) we see that any path from ∂ to a vertex that is a descendant<br />

of v must go through v. It follows that v is a separating vertex of the map M = Ψ r,p ((T ,U))<br />

and that the s<strong>et</strong> T [v],0 is in one-to-one correspondence with the s<strong>et</strong> S v,τ<br />

□<br />

Thanks to Corollary 4.2.4 and Lemma 4.4.3, it suffices to prove the following proposition in<br />

order to g<strong>et</strong> Theorem 4.4.2. Recall the definition of µ κ = (µ κ 0 ,µκ 1 ).<br />

Proposition 4.4.4. For every ε > 0,<br />

lim<br />

n→∞ Pn µ κ ,ν,1<br />

M .<br />

(<br />

∃ v 0 ∈ T 0 : n 1/2−ε ≤ #T [v 0],0 ≤ 2n 1/2−ε , U [v 0] > 0<br />

Proof : For n ≥ 1 and ε > 0, we denote by Λ n,ε the event<br />

{<br />

}<br />

Λ n,ε = ∃ v 0 ∈ T 0 : n 1/2−ε ≤ #T [v0],1 ≤ 2n 1/2−ε , U [v0] > 0 .<br />

L<strong>et</strong> ε > 0 and α > 0. We will prove that for all n sufficiently large,<br />

(4.4.2) P n µ κ ,ν,1 (Λ n,ε) ≥ 1 − 3α.<br />

We first state a lemma. For T ∈ T and k ≥ 0, we s<strong>et</strong><br />

Z 0 (k) = #{v ∈ T : |v| = 2k}.<br />

)<br />

= 1.<br />

Lemma 4.4.5. There exist constants β > 0, γ > 0 and M > 0 such that for all n sufficiently<br />

large, (<br />

Pµ n κ inf<br />

γ √ n≤k≤2γ √ Z 0 (k) > β √ n, sup Z 0 (k) < M √ )<br />

n ≥ 1 − α.<br />

n<br />

k≥0<br />

We postpone the proof of Lemma 4.4.5 and compl<strong>et</strong>e that of Proposition 4.4.4. To this end,<br />

we introduce some notation. For n ≥ 1, we define an integer K n by the condition<br />

⌈γ √ n ⌉ + K n ⌈n 1/4 ⌉ ≤ ⌊2γ √ n⌋ < ⌈γ √ n ⌉ + (K n + 1)⌈n 1/4 ⌉,<br />

118


and we s<strong>et</strong> for j ∈ {0,... ,K n },<br />

k (n)<br />

j<br />

= ⌈γ √ n ⌉ + j⌈n 1/4 ⌉.<br />

If T ∈ T, we write H(T ) for the height of T , that is the maximal generation of an individual in<br />

T . For k ≥ 0, and N,P ≥ 1 we s<strong>et</strong><br />

{<br />

}<br />

Z 0 (k,N,P) = # v ∈ T : |v| = 2k,N ≤ #T [v],0 ≤ 2N, H(T [v] ) ≤ 2P .<br />

We denote by Γ n , C n,ε and E n,ε the events<br />

⋂ {<br />

Γ n = β √ ( )<br />

n < Z 0 k (n)<br />

j<br />

0≤j≤K n<br />

C n,ε =<br />

E n,ε =<br />

< M √ }<br />

n ,<br />

⋂ { ( Z 0 k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) > n 1/4} ,<br />

0≤j≤K n<br />

⋂ {<br />

0≤j≤K n<br />

n 1/4 < Z 0 ( k (n)<br />

The first step is to prove that for all n sufficiently large,<br />

j<br />

,n 1/2−ε ,n 1/4) < M √ }<br />

n .<br />

(4.4.3) P n<br />

µ κ (E n,ε) ≥ 1 − 2α.<br />

Since Γ n ∩ C n,ε ⊂ E n,ε , it suffices to prove that for all n sufficiently large,<br />

(4.4.4) Pµ n κ (Γ n ∩ C n,ε ) ≥ 1 − 2α.<br />

We first observe that<br />

∑K n<br />

(4.4.5) P µ κ(Γ n ∩ Cn,ε) c ≤<br />

j=0<br />

P µ κ<br />

(<br />

β √ n < Z 0 ( k (n)<br />

j<br />

L<strong>et</strong> j ∈ {0,... ,K n }. We have<br />

(<br />

P µ κ β √ ( )<br />

n < Z 0 k (n)<br />

j<br />

< M √ ( n, Z 0 k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) ≤ n 1/4)<br />

(4.4.6)<br />

=<br />

⌊M √ n ⌋<br />

∑<br />

q=⌈β √ n ⌉<br />

P µ κ<br />

( (<br />

Z<br />

k (n)<br />

j<br />

)<br />

< M √ n, Z 0 ( k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) ≤ n 1/4) .<br />

,n 1/2−ε ,n 1/4) ∣ ( ) ) (<br />

≤ n 1/4 ∣∣ Z<br />

0<br />

k (n)<br />

j<br />

= q P µ κ<br />

Z 0 ( k (n)<br />

j<br />

)<br />

)<br />

= q .<br />

Now, under the probability measure P µ κ(· | Z 0 (k (n)<br />

j<br />

) = q), the q subtrees of T above level 2k (n)<br />

j<br />

are independent and distributed according to P µ κ. Consider on a probability space (Ω ′ ,P ′ ) a<br />

sequence of Bernoulli variables (B i ,i ≥ 1) with param<strong>et</strong>er p n defined by<br />

(<br />

p n = P µ κ n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε , H(T ) ≤ 2n 1/4) .<br />

Then (4.4.6) gives<br />

(<br />

P µ κ β √ ( )<br />

n < Z 0 k (n)<br />

j<br />

< M √ (<br />

n, Z k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) ≤ n 1/4)<br />

⌊M √ n ⌋<br />

)<br />

∑<br />

≤<br />

B i ≤ n 1/4<br />

(4.4.7)<br />

≤<br />

( q∑<br />

q=⌈β √ n ⌉P ′ i=1<br />

(<br />

⌊M √ n ⌋<br />

∑<br />

q=⌈β √ n ⌉<br />

eexp<br />

− qp nn −1/4<br />

2<br />

)<br />

,<br />

119


where the last bound follows from a simple exponential inequality. However, from Lemma 14 in<br />

[43], there exists η > 0 such that for all n sufficiently large,<br />

p n = P µ κ<br />

(n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε) − P µ<br />

(n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε , H(T ) > 2n 1/4)<br />

(<br />

(4.4.8) ≥ P µ κ n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε) − e −nη .<br />

Under the probability measure P µ κ, we have #T 0 = (κ −1)#T 1 +1 a.s., so we g<strong>et</strong> from Lemma<br />

4.3.9 that there exists a constant c κ such that,<br />

(<br />

(4.4.9) n 1/4−ε/2 P µ κ n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε) −→ c κ .<br />

n→∞<br />

It then follows from (4.4.8) that<br />

n 1/4−ε/2 p n −→ c κ .<br />

n→∞<br />

From (4.4.7), we obtain that there exists a constant c ′ > 0 such that<br />

(<br />

P µ κ β √ n < Z 0 < M √ (<br />

)<br />

n, Z k (n)<br />

k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) ≤ n 1/4 ≤ M √ ne −c′ n ε ,<br />

j<br />

which tog<strong>et</strong>her with (4.4.5) implies that<br />

(<br />

P µ κ Γn ∩ Cn,ε<br />

c ) √ ≤ MKn ne<br />

−n ε .<br />

Since K n ∼ γn 1/4 as n → ∞, we g<strong>et</strong> from Lemma 4.3.9 that, for all n sufficiently large,<br />

(4.4.10) Pµ n ( κ Γn ∩ Cn,ε<br />

c ) P µ κ(Γ n ∩ Cn,ε)<br />

c<br />

≤<br />

P µ κ(#T 1 = n) ≤ α.<br />

On the other hand, Lemma 4.4.5 implies that<br />

The bound (4.4.4) now follows.<br />

S<strong>et</strong><br />

I n =<br />

For (p 0 ,... ,p Kn ) ∈ I n , we s<strong>et</strong><br />

E n,ε (p 0 ,... ,p Kn ) =<br />

P n<br />

µ κ(Γ n) ≥ 1 − α.<br />

{<br />

⌈n 1/4 ⌉, ⌈n 1/4 ⌉ + 1,... , ⌊M √ n⌋} Kn+1<br />

.<br />

⋂ { ( Z 0 k (n)<br />

j<br />

,n 1/2−ε ,n 1/4) }<br />

= p j .<br />

0≤j≤K n<br />

On the event E n,ε (p 0 ,... ,p Kn ) for every j ∈ {1,... ,K n }, l<strong>et</strong> v j 1 ≺ ... ≺ vj p j<br />

be the list in<br />

lexicographical order of those vertices v ∈ T at generation 2k (n)<br />

j<br />

such that n 1/2−ε ≤ #T [v],0 ≤<br />

2n 1/2−ε and H(T [v] ) ≤ 2n 1/4 . Note that for every j,j ′ ∈ {1,... ,K n } such that j < j ′ , we have<br />

for every i ∈ {1,... ,p j },<br />

{<br />

max |v j i v| : v ∈ T [vj]} i ≤ 2k (n)<br />

j<br />

+ 2n 1/4 < 2k (n)<br />

j<br />

. ′<br />

Then, it is not difficult to check that under P µ κ ,ν,1(· | E n,ε (p 0 ,... ,p Kn )), the spatial trees<br />

{(T [vj i ] ,U [vj i ] ), i = 1,... ,p j , j = 1,... ,K n } are independent and distributed according to the<br />

probability measure P µ κ ,ν,0(· | n 1/2−ε ≤ #T 0 ≤ n 1/2−ε , H(T ) ≤ 2n 1/4 ). S<strong>et</strong><br />

π n = P µ,ν,0<br />

(U > 0 | n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε , H(T ) ≤ 2n 1/4) .<br />

120


Since the events E n,ε (p 0 ,... ,p Kn ), (p 0 ,... ,p Kn ) ∈ I n are disjoints we have,<br />

∑<br />

P µ κ ,ν,1(E n,ε ∩ Λ c n,ε) = P µ κ(E n,ε (p 0 ,... ,p Kn ))<br />

(p 0 ,...,p Kn )∈I n<br />

}<br />

)<br />

× P µ κ ,ν,1<br />

({U [vj i ] ≤ 0 : 0 ≤ j ≤ K n ;1 ≤ i ≤ p j | E n,ε (p 0 ,...,p Kn )<br />

∑<br />

=<br />

(4.4.11)<br />

≤<br />

(p 0 ,...,p Kn )∈I n<br />

P µ κ(E n,ε (p 0 ,... ,p Kn ))(1 − π n ) p 0+...+p Kn<br />

P µ κ(E n,ε )(1 − π n ) ⌈n1/4 ⌉(K n+1)<br />

≤ (1 − π n ) ⌈n1/4 ⌉(K n+1) .<br />

Now, from Lemma 14 in [43] and (4.4.9) there exists c ′′ > 0 such that for all n sufficiently large,<br />

π n ≥ P µ,ν,0<br />

(U > 0 | n 1/2−ε ≤ #T 0 ≤ 2n 1/2−ε) − c ′′ n 1/2 e −nη ,<br />

where η was introduced before (4.4.8). Since under the probability measure P µ κ, we have #T 0 =<br />

1 + (κ − 1)#T 1 a.s., we g<strong>et</strong> from Proposition 4.3.12 that there exists a constant c > 0 such that<br />

for all n sufficiently large,<br />

π n ≥<br />

c<br />

n 1/2−ε.<br />

So, there exists a constant c ′ > 0 such that (4.4.11) becomes for all n sufficiently large,<br />

(4.4.12) P µ κ ,ν,1<br />

(<br />

En,ε ∩ Λ c n,ε)<br />

≤ e<br />

−c ′ n ε .<br />

Finally (4.4.12) tog<strong>et</strong>her with Lemma 4.3.9 implies that if n is sufficiently large,<br />

Using also (4.4.3), we obtain our claim (4.4.2).<br />

P n µ κ ,ν,1<br />

(<br />

En,ε ∩ Λ c n,ε)<br />

≤ α.<br />

Proof of Lemma 4.4.5 : Under P µ κ, Z 0 = (Z 0 (k),k ≥ 0) is a critical Galton-Watson<br />

process with offspring law µ κ 0,κ supported on (κ − 1)Z + and defined by<br />

µ κ 0,κ ((κ − 1)k) = µκ 0 (k), k ≥ 0.<br />

Note that the total progeny of Z 0 is #T 0 and recall that under the probability measure P µ κ,<br />

we have #T 0 = 1 + (κ − 1)#T 1 a.s.<br />

Define a function L = (L(t),t ≥ 0) by interpolating linearly Z 0 b<strong>et</strong>ween successive integers.<br />

For n ≥ 1 and t ≥ 0, we s<strong>et</strong><br />

l n (t) = √ 1 L(t √ n). n<br />

Denote by l(t) the total local time of e at level t that is,<br />

1<br />

l(t) = lim<br />

ε→0 ε<br />

∫ 1<br />

0½{t≤e(s)≤t+ε} ds,<br />

where the convergence holds a.s. From Theorem 1.1 in [19], the law of (l n (t),t ≥ 0) under Pµ n κ<br />

converges to the law of (l(t),t ≥ 0). So we have<br />

lim inf P n<br />

n→∞<br />

µ κ (<br />

)<br />

inf l n(t) > β, sup l n (t) < M<br />

γ≤t≤2γ<br />

t≥0<br />

(<br />

≥ P<br />

inf l(t) > β, sup l(t) < M<br />

γ≤t≤2γ t≥0<br />

)<br />

.<br />

□<br />

121


However we can find β > 0, γ > 0 and M > 0 such that<br />

(<br />

)<br />

P inf l(t) > β, sup l(t) < M ≥ 1 − α<br />

γ≤t≤2γ t≥0<br />

2 ,<br />

and the desired result follows.<br />

□<br />

122


Bibliographie<br />

[1] Abraham, R., Werner, W. (1997) Avoiding probabilities for Brownian snakes and super-Brownian motion.<br />

Electron. J. Probab. 2 no. 3, 27 pp.<br />

[2] Abraham, R., Serl<strong>et</strong>, L. (2002) Representations of the Brownian snake with drift. Stochastics and Stochastics<br />

Reports 73, 287-308.<br />

[3] Aldous, D. (1991) The continuum random tree I. Ann. Probab. 19, 1-28.<br />

[4] Aldous, D. (1991) The continuum random tree. II. An overview. Stochastic analysis (Durham, 1990), 23-70,<br />

London Math. Soc. Lecture Note Ser. 167. Cambridge Univ. Press, Cambridge, 1991.<br />

[5] Aldous, D. (1993) The continuum random tree III. Ann. Probab. 21, 248-289.<br />

[6] Aldous, D. (1993) Tree-based models for random distribution of mass. J. Stat. Phys. 73, 625-641.<br />

[7] Athreya, K.B., Ney, P.E. (1972) Branching Processes. Springer, Berlin.<br />

[8] Banderier, C., Flajol<strong>et</strong>, P., Schaeffer G., Soria M. (2001) Random maps, coalescing saddles, singularity<br />

analysis, and Airy phenomena. Random Struct. Alg. 19, 194-246.<br />

[9] Bouttier, J., Di Francesco, P., Guitter, E. (2003) Random trees b<strong>et</strong>ween two walls : exact partition<br />

function. J. Phys. A 36, 12349-12366.<br />

[10] Bouttier, J., Di Francesco, P., Guitter, E. (2003) Statistics of planar graphs viewed from a vertex : a<br />

study via labeled trees. Nuclear Phys. B 675, 631-660.<br />

[11] Bouttier J., Di Francesco P., Guitter E. (2004) Planar maps as labeled mobiles. Electron. J. Combin.<br />

11, R69<br />

[12] Burago, D., Burago, Y., Ivanov, S. (2001) A Course in M<strong>et</strong>ric Geom<strong>et</strong>ry. Graduate studies in mathematics,<br />

vol.33, AMS, Boston.<br />

[13] Chassaing, P., Durhuus, B. (2006) Local limit of labelled trees and and expected volume growth in a<br />

random quadrangulation. Ann. Probab., 34.<br />

[14] Chassaing, P., Schaeffer, G. (2004) Random planar lattices and integrated superBrownian excursion.<br />

Probab. Th. Rel. Fields 128, 161-212.<br />

[15] Cori, R., Vauquelin, B. (1981) Planar trees are well labeled trees. Canad. J. Math. 33, 1023-1042.<br />

[16] Dellacherie, C., Meyer, P.A. (1980) Probabilités <strong>et</strong> Potentiels, Chapitres V à VIII : Théorie des Martingales.<br />

Hermann, Paris.<br />

[17] Delmas, J.F. (2003) Computation of moments for the length of the one dimensional ISE support. Electron.<br />

J. Probab. 8 no. 17, 15 pp.<br />

[18] Derbez, E., Slade, G. (1998) The scaling limit of lattice trees in high dimensions. Comm. Math. Phys.<br />

198, 69-104.<br />

[19] Drmota, M., Gittenberger, B. (1997) On the profile of random trees. Random Struct. Alg. 10, 421-451.<br />

[20] Duquesne, T. (2003) A limit theorem for the contour process of conditioned Galton-Watson trees. Ann.<br />

Probab. 31, 996-1027.<br />

[21] Duquesne, T., Le Gall, J.F. (2002) Random Trees, Lévy Processes and Spatial Branching Processes,<br />

Astérisque 281.<br />

[22] Duquesne, T., Le Gall, J.F. (2005) Probabilistic and fractal aspects of Lévy trees. Probab. Th. Rel. Fields<br />

131, 553-603.<br />

[23] Duquesne, T., Le Gall, J.F. (2005) The Hausdorff measure of stable trees.<br />

[24] Evans, S.N., Pitman, J.W., Winter, A. (2003) Rayleigh processes, real trees and root growth with regrafting.<br />

Probab. Th. Rel. Fields, 134, 81-126.<br />

123


[25] Evans, S.N., Winter, A. Subtree prune and re-graft : A reversible tree valued Markov process. Ann.<br />

Probab., to appear.<br />

[26] Grimvall, A. (1974) On the convergence of sequences of branching processes. Ann. Probab. 2 1027-1045.<br />

[27] Haas B., Miermont G. (2004) The genealogy of self-similar fragmentations with negative index as a continuum<br />

random tree. Electr. J. Probab. 9, 57-97.<br />

[28] Hara, T., Slade, G. (2000) The scaling limit of the incipient infinite cluster in high-dimensional percolation.<br />

II. Integrated super-Brownian excursion. Probabilistic techniques in equilibrium and nonequilibrium<br />

statistical physics. J. Math. Phys. 41 (2000), 1244-1293.<br />

[29] van der Hofstad, R., Slade, G. (2003) Convergence of critical oriented percolation to super-Brownian<br />

motion above 4 + 1 dimensions. Ann. Inst. H. Poincaré Probab. Statist. 20, 413-485.<br />

[30] Lamperti, J. (1966) The limit of a sequence of branching processes. Z. Wahrsch. verw. Gebi<strong>et</strong>e 7 271-288.<br />

[31] Janson, S., Marckert, J.F. (2005) Convergence of discr<strong>et</strong>e snakes. J. Theor<strong>et</strong>. Probab., 18, 615-647.<br />

[32] Jansons, K.M., Rogers, L.C.G. (1992) Decomposing the branching Brownian path. Ann. Probab. 2, 973-<br />

986.<br />

[33] Kallenberg, O. (1975) Random Measures. Academic Press, London.<br />

[34] Le Gall, J.F. (1991) Brownian excursions, trees and measure-valued branching processes. Ann. Probab.<br />

19, 1399-1439.<br />

[35] Le Gall, J.F. (1993) The uniform random tree in a Brownian excursion. Probab. Th. Rel. Fields 96, 369-383.<br />

[36] Le Gall, J.F. (1995) The Brownian snake and solutions of ∆u = u 2 in a domain. Probab. Th. Rel. Fields<br />

102, 393-432.<br />

[37] Le Gall, J.F. (1999) Spatial Branching Processes, Random Snakes and Partial Differential Equations. Lectures<br />

in Mathematics ETH Zürich. Birkhäuser, Boston.<br />

[38] Le Gall, J.F. (2005) Random trees and applications. Probab. Surveys 2, 245-311.<br />

[39] Le Gall, J.F. (2006) A conditional limit theorem for tree-indexed random walk. Stoch. Process. Appl. 116,<br />

539-567.<br />

[40] Le Gall, J.F. (2006) The topological structure of scaling limits of large planar maps. arXiv :<br />

math.PR/0607567.<br />

[41] Le Gall, J.F, Le Jan, Y. (1998) Branching processes in Lévy processes : the exploration process. Ann. Probab.<br />

26, 213-252.<br />

[42] Le Gall, J.F., Weill, M. (2006) Conditioned Brownian trees. Ann. Inst. H. Poincaré Probab. Statist. 42,<br />

455-489 .<br />

[43] Marckert, J.F., Miermont, G. (2006) Invariance principles for random bipartite planar maps. Ann. Probab.,<br />

to appear.<br />

[44] Marckert, J.F., Mokkadem, A. (2004) State spaces of the snake and its tour - Convergence of the discr<strong>et</strong>e<br />

snake. J. Theor<strong>et</strong>. Probability 16, 1015-1046.<br />

[45] Marckert, J.F., Mokkadem, A. (2004) Limits of normalized quadrangulations. The Brownian map.<br />

Ann. Probab, to appear.<br />

[46] Miermont, G. (2003) Self-similar fragmentations derived from the stable tree I : splitting at heights. Probab.<br />

Th. Rel. Fields, 127, 423-454.<br />

[47] Miermont, G. (2005) Self-similar fragmentations derived from the stable tree II : splitting at nodes. Probab.<br />

Th. Rel. Fields, 131, 341-375.<br />

[48] Neveu, J. (1986) Arbres <strong>et</strong> processus de Galton-Watson. Ann. Inst. H. Poincaré Probab. Statist. 22, 199-207<br />

[49] Norris, J. (1997) Markov Chains. Cambridge University Press, Cambridge.<br />

[50] Revuz, D., Yor, M. (1991) Continuous Martingales and Brownian Motion. Springer, Berlin-Heidelberg-New<br />

York.<br />

[51] Schaeffer G. (1998) Conjugaison d’<strong>arbres</strong> <strong>et</strong> <strong>cartes</strong> aléatoires, Thèse, Université de Bordeaux I.<br />

[52] Vervaat, W. (1979) A relation b<strong>et</strong>ween Brownian bridge and Brownian excursion. Ann. Probab. 7, 143-149.<br />

[53] Weill, M. (2006) Regenerative real trees. Ann. Probab., to appear.<br />

[54] Weill, M. (2006) Asymptotics for rooted planar maps and scaling limits of two-type spatial trees. Preprint.<br />

[55] Yor, M. (1980) Loi de l’indice du lac<strong>et</strong> brownien, <strong>et</strong> distribution de Hartman-Watson. Z. Wahrsch. verw.<br />

Gebi<strong>et</strong>e 53, 71-95.<br />

124

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!