• Aucun résultat trouvé

Hyperbolic Radial Spanning Tree

N/A
N/A
Protected

Academic year: 2021

Partager "Hyperbolic Radial Spanning Tree"

Copied!
32
0
0

Texte intégral

(1)

HAL Id: hal-03024280

https://hal.archives-ouvertes.fr/hal-03024280

Submitted on 25 Nov 2020

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

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

Hyperbolic Radial Spanning Tree

Lucas Flammant

To cite this version:

Lucas Flammant. Hyperbolic Radial Spanning Tree. Electronic Journal of Probability, Institute of Mathematical Statistics (IMS), 2020. �hal-03024280�

(2)

Hyperbolic Radial Spanning Tree

Lucas Flammant

November 25, 2020

We define and analyze an extention to thed-dimensional hyperbolic space of the Ra- dial Spanning Tree (RST) introduced by Baccelli and Bordenave in the two-dimensional Euclidean space (2007). In particular, we will focus on the description of the infinite branches of the tree. The properties shown for the two-dimensional Euclidean RST are extended to the hyperbolic case in every dimension: almost surely, every infinite branch admits an asymptotic direction and each asymptotic direction is reached by at least one infinite branch. Moreover, the branch converging to any deterministic asymptotic direction is unique almost surely. Our strategy mainly relies on the two following ingre- dients. First, the hyperbolic metric allows us to control fluctuations of the branches in the hyperbolic DSF without using planarity arguments. Then, we couple the hyperbolic RST with the hyperbolic DSF introduced and studied in Flammant 2019.

Key words: continuum percolation, hyperbolic space, stochastic geometry, random geometric tree, Radial Spanning Tree, Poisson point processes.

AMS 2010 Subject Classification: Primary 60D05, 60K35, 82B21.

Acknowledgments. This work has been supervised by David Coupier and Viet Chi Tran, who helped a lot to write this paper. It has been supported by the LAMAV (Université Polytechnique des Hauts de France) and the Laboratoire P. Painlevé (Université de Lille). It has also benefitted from the GdR GeoSto 3477, the Labex CEMPI (ANR-11-LABX-0007-01) and the ANR PPPP (ANR-16-CE40-0016).

Univ. de Valenciennes, CNRS, EA 4015 - LAMAV, F-59313 Valenciennes Cedex 9, France

(3)

Contents

1 Introduction 3

2 Definitions, notations and basic properties 4

2.1 The hyperbolic space . . . . 4

2.2 The hyperbolic RST . . . . 7

2.3 Sketch of proofs . . . . 8

3 Proof of Theorem 1.1 10 3.1 The existence part: proof of (i),(ii),(iv) and (v) . . . . 10

3.2 The uniqueness part: proof of (iii) . . . . 11

4 Proof of Proposition 2.5 15 4.1 Step 1: a deterministic upper-bound of MBDrr0 0 (·) . . . . 17

4.2 Step 2: a control of the tail decay ofM(z). . . . 18

4.3 Step 3: conclusion . . . . 19

4.4 Proof of Lemmas 4.1, 4.2, 4.3, 4.4 . . . . 20

5 Proof of Propositions 2.6 and 2.7 23

6 Appendix A: proof of Proposition 2.2 28

7 Appendix B: computing distances and angles in the half-space model 29

(4)

1 Introduction

Geometric random trees are well studied in the literature since it interacts with many other fields, such as communication networks, particles systems or population dynamics. Several works have established scaling limits for two-dimensional radial trees [10, 9] and translation invariant forests [11, 22, 14]. In addition, random spanning trees appear in the context or first passage percolation [15]. A complete introduction to geometric random graphs is given in Penrose 2003 [20].

The Radial Spanning Tree (RST) is a random tree whose introduction in the two-dimensional Euclidean space has been motivated by applications for communication networks [1]. The set of vertices is given by a homogeneous Poisson Point Process (PPP)N of intensityλinR2. The RST rooted at the origin 0 is the graph obtained by connecting each point z∈ N to its parent A(z), defined as the closest point tozamong all pointsz0∈ N ∪ {0}that are closer to the origin thanz.

This defines a random tree rooted at the origin with a radial structure. An infinite backward path is defined as a sequence of Poisson points(zn)n≥0 (N ∪ {0})N with z0 = 0 andzn =A(zn+1) for anyn0. Given an infinite path, we will say that the forward direction is towards0and the backward direction is towards infinity.

The topological properties of the bi-dimensional Euclidean RST are well-understood. Baccelli and Bordenave showed that almost surely, any infinite backward path admits an asympotic direc- tion; moreover, a.s., every asymptotic direction is reached by at least one infinite backward path and there exists a.s. a unique infinite path in any given deterministic asymptotic direction [1].

These results on the infinite paths are completed by Baccelli, Coupier & Tran [2].

For any integer d2, the hyperbolic space Hd is a d-dimensional Riemannian manifold with constant negative curvature, that can be chosen equal to−1without loss of generality. It admits a set of ideal boundary points Hd, and Hd :=HdHd denotes the hyperbolic space endowed with its boundary. It is a non-amenable space, i.e. the measure of the boundary of a large subset is not negligible with respect to its volume. The hyperbolic space is defined in more details in [7]

and [19].

There is a growing interest for the study of random models in a hyperbolic setting. Benjamini and Schramm establish percolation results on regular tilings and Voronoï tesselation in the hyper- bolic plane [3]. Mean characteristics of the Poisson-Voronoï tessellation have also been considered in a general Riemannian manifold by Calka et al. [6]. This interest is explained by at least two reasons. First, hyperbolic random graphs are well-fitted to modelize social networks [5]. In ad- dition, strong differences have been noticed for properties of random models depending whether they are considered in an Euclidean or hyperbolic setting. Indeed, some hyperbolic random graphs admits a non-degenerate regime with infinitely many unbounded components in the hyperbolic space [23, 16], which is generally not the case in the Euclidean space. In addition, behaviours of non-amenable spaces are well studied in a discrete context [4, 18, 21].

Thus it is natural to consider and study the hyperbolic RST, which we define in the same way as the Euclidean RST. A simulation of the two-dimensional hyperbolic RST is given in Figure 1.

In this paper, we extend the results of Baccelli and his coauthors to hyperbolic geometryin every dimension. Here is our main result:

Theorem 1.1. For any dimensiond1 and any intensityλ, the following happens:

(i) almost surely, any infinite backward path(zn)n∈Nadmits an asymptotic direction, i.e. there existszHd+1 such thatlimn→∞zn =z (in the sense of the topology ofHd+1);

(ii) almost surely, for anyIHd+1, there exists an infinite backward path(zn)with asymptotic direction I (i.e. such thatlimn→∞zn=I);

(iii) for any deterministic boundary point I Hd+1, the path with asymptotic direction I is almost surely unique;

(iv) the set of boundary points with two infinite backward paths is dense inHd+1; (v) this set is moreover countable in the bi-dimensional case (i.e. d= 1).

(5)

Establishing the results announced in Theorem 1.1 in every dimension constitutes the main originality of this paper. For the two reasons explained further, the proofs of Baccelli and Bordenave in the 2D-Euclidean setting [1] cannot be generalised to higher dimensions.

In both contextsR2andHd+1, for anyd1, the proofs of (i), (ii), (iv) and (v) of Theorem 1.1 follow the strategy of Howard and Newman [15], which is to show that the tree isstraight, that is, the descendents subtree of a vertex far from the origin is included in a thin cone. To prove that the 2D-Euclidean RST is straight, Baccelli and Bordenave used a translation invariant model derived from the RST: the Directed Spanning Forest (DSF), which constitutes a local approximation of the RST far from the origin [1]. They exploit the theory of Markov chains to upper-bound fluctuations of trajectories in the DSF and then, they deduce the straightness of the RST via planarity. This strategy cannot be generalised to higher dimensions. However, in Hd, we manage to control the angular deviations of branches in the RST without resorting to an auxiliary model, which required planarity in the Euclidean setting. The hyperbolic metric guarantees that angular deviations decay exponentially fast with the distance to the origin, which is strong enough to show straightness.

In addition, in the Euclidean context, the uniqueness part (point (iii) in Theorem 1.1) is only proved in dimension2since it strongly uses planarity [15, 1], and the strategy of proof cannot be generalised to higher dimensions. To prove (iii) inHd, our strategy consists in exploiting the link existing between the hyperbolic RST and another random graph, the hyperbolic DSF, defined and studied in Flammant 2019 [12], which is the hyperbolic counterpart of the Euclidean DSF used by Baccelli and Bordenave. Roughly speaking, the hyperbolic DSF can be defined as the limit of the hyperbolic RST when the origin point tends to an ideal boundary point. Similarly to the Euclidean setting, it constitutes a local approximation of the RST far from the origin. The proof of (iii) exploits the coalescence of the hyperbolic DSF (i.e. it is almost surely a tree) [12, Theorem 1.1], which is a non-trivial fact obtained by exploiting the mass-transport principle, and a local coupling between the two models.

After defining the hyperbolic RST and giving its basic properties, we define two quantities that encode angular fluctuations along trajectories, the Cumulative angular Forward Deviations (CFD) and the Maximal Backward Deviations (MBD). We then establish upper-bounds of these quantities: first, we upper-bound the Maximal Backward Deviations in a thin annulus of width δ >0(Proposition 2.5) and then we deduce a global control of MBD in the whole space (Proposition 2.6), that roughly says that angular deviations decay exponentially fast with the distance to the origin. From this upper-bound, we deduce that the RST is straight in the sense of Howard &

Newman (Proposition 2.7). The points (i), (ii), (iv) and (v) in Theorem 1.1 can be deduced from straightness and the upper-bound of MBD given by Proposition 2.6. The point (iii) (the uniqueness part) is done by exploiting a local coupling existing between the RST and the DSF far from the origin.

The rest of paper is organized as follows. In Section 2, we set some reminders of hyperbolic geometry and we define the hyperbolic RST. Then, we give its basic properties and a road-map of the proofs. We also announce the upper-bounds of angular deviations (Propositions 2.5 and 2.6) and the straightness property (Proposition 2.7). The proof of Theorem 1.1 is done in Section 3. Proposition 2.5 is proved in Section 4 and the proofs of Propositions 2.6 and 2.7 are done in Section 5.

2 Definitions, notations and basic properties

We denote byNthe set of non-negative integers and byNthe set of positive integers. In the rest of the paper,c(resp. C) will be some small (resp. large) constant whose value can change from a line to another.

2.1 The hyperbolic space

We refer to [7] or [19] for a complete introduction to hyperbolic geometry. FordN, the(d+ 1)- dimensional hyperbolic space, denoted byHd+1, is a(d+ 1)-dimensional Riemannian manifold of constant negative curvature−1 that can be defined by several isometric models. One of them is

(6)

Figure 1: Simulation of the two-dimensional hyperbolic RST, withλ = 30, in the Poincaré disc model. The edges are represented by geodesics. The different connected components of the RST (appart from the root) are represented with different colors.

the open-ball model consisting in the unit open ball

I={(x1, ..., xd+1)Rd+1, x21+...+x2d+1<1} (2.1) endow with the following metric:

ds2I := 4 dx21+...+dx2n+1

1x21...x2d+1. (2.2)

This model is rotation invariant. The metric becomes smaller as we get closer to the boundary unit sphere∂I, and this boundary is at infinite distance from the center0.

The volume measure on(I, ds2I), denoted by VolI, is given by dVolI = 2d+1 dx1...dxd+1

1x21...x2d+1. (2.3)

In this model (I, ds2I), the geodesics are of two types: the diameters of I and the arcs that are perpendicular to the boundary unit sphere ∂I. We refer to discussion [7, P.80] for a proof.

Moreover, this model is conformal, which means that the hyperbolic angle between two geodesics corresponds to their Euclidean angle in the open-ball representation.

An important fact about hyperbolic geometry is that all points and all directions play the same role. More precisely, Hd+1 is homogeneous and isotropic. It means that the group of isometries

(7)

ofHd+1 acts transitively on the unit tangent bundle ofHd+1: given two pointsx, y Hd+1 and two unit tangent vectorsuTxHd+1, v TyHd+1, there exists an isometry g of Hd+1 such that g(x) =y and that pushes forward uon v. The notations Tx, Ty and the vocabulary relating to Riemannian geometry are defined in [17]. We refer to [19], Proposition 1.2.1 p.5, for a proof.

Figure 2: Geodesics in the open ball model

The hyperbolic spaceHd+1 is naturally equipped with a set of points at infinity, and the most natural way to identify these points is to use the open-ball model. In (I, ds2I), the set of points at infinity is identified by the boundary unit sphere∂I. We denote by Hd+1 the boundary set (represented by the boundary unit sphere in(I, ds2I)) and byHd+1:=Hd+1Hd+1the hyperbolic spaceHd+1 plus the set of points at infinity, with the topology given by the closed ball. A point zHd+1 is calledideal point or point at infinity.

We denote byd(·,·)the hyperbolic distance inHd+1, and byk · kthe Euclidean norm inRd, with the conventionk∞k=∞. Let us denote by Vol the volume measure onHd+1. Forz1, z2H

d+1, let us denote by[z1, z2]the geodesic betweenz1 andz2. Moreover, we set the notations:

[z1, z2[:= [z1, z2]\{z2}, ]z1, z2] := [z1, z2]\{z1}, ]z1, z2[\({z2} ∪ {z2}).

Let us denote by[z1, z2)(resp. (z1, z2]) the semi-geodesic passing threwz2 (resp. z1) and ending atz1 (resp. z2). Forz1, z2, z3H

d+1,z\1z2z3 is the measure of the corresponding (non-oriented) hyperbolic angle. For any subsetB Hd+1, B denotes the closure of B in Hd+1. For any point z Hd+1 and θ >0, Cone(z, θ) :={z0 Hd+1, z0zd0 θ} is defined as the cone of apex0 and apertureθ(ifθπthen Cone(z, θ)is the hole spaceHd+1). In addition, forr >0 andzHd+1, we define

BS(r)(z, θ) :=Cone(z, θ)S(r). (2.4)

Let 0 Hd+1 be some arbitrary origin point (it can be thought as the center of the ball in the open-ball representation), which will plays the role of the root of the RST. ForzHd+1 and r > 0, we denote by B(z, r) := {z0 H, d(z, z0)< r} (resp. S(z, r) := {z0 H, d(z, z0) = r}) the hyperbolic ball (resp. sphere) centered at z of radius r, and we set B(r) := B(0, r) (resp.

S(r) :=s(0, r)). For xRd andr >0, let us also denote byBRd(x, r) :={x0Rd, kx0xk< r}

the Euclidean ball centered atxof radiusr.

Let us denote by Sd the unit Euclidean sphere in Rd+1 and by ν its d-dimensional volume measure. Since the RST is a rooted graph, a convenient way to represent points in Hd+1 is to use polar coordinates. Recall that0 is the origin point. For any point z Hd+1, we denote by

(8)

z= (r;u)its polar coordinates w.r.t. 0: ris its distance to0anduU T0Hd+1'Sdis its direction (U T0Hd+1 is the unitary tangent space of0 in Hd+1). In polar coordinates, the volume measure Vol is given by

dVol(r;u) = sinh(r)d dr dν(u). (2.5)

A direct consequence of this is that the volume of a ball of radiusris given by:

Vol(B(r)) = Z r

0

sinh(r)d dr dν(Sd) =Sdsinh(r)dedr whenr→ ∞, (2.6) whereSd is thed-dimensional volume ofSd.

The hyperbolic law of cosines [24, p.13] is a well adapted tool to compute distances using polar coordinates. Givenz1= (r1;u1),z2= (r2, u2)Hd+1, the hyperbolic law of cosines gives,

coshd(z1, z2) = cosh(r1) cosh(r2)− hu1, u2isinh(r1) sinh(r2). (2.7)

2.2 The hyperbolic RST

In the rest of the paper, the dimensiondand the intensityλ >0are fixed. LetN be a homogeneous PPP of intensityλinHd+1. The definition of the hyperbolic RST is similar to the Euclidean case.

The set of vertices isN ∪ {0}. Each vertexz∈ N is connected to the closest Poisson point among those that are closer to the origin thanz:

Definition 2.1(Radial Spanning Tree inHd+1). For anyz= (r;u)∈ N, theparent ofzis defined as

A(z) := argmin

z0∈N ∩B(r)

d(z0, z).

We callRadial Spanning Tree (RST) inHd+1 rooted at0the oriented graph(V, ~E)where V :=N ∪ {0}, E~ :={(z, A(z)), z∈ N }.

It is possible to assume thatN ∪ {0} does not contain isosceles triangles, since this event has probability1. Thus the ancestor A(z)is well-defined.

Forz ∈ N ∪ {0} and k N, let us define A(k)(z) = A..A k times and A(−k)(z) = {z0 N, A(k)(z0) =z}(in particularA(−1)(z)is the set of daughters ofz). For z∈ N andr0, let us define

B+(z, r) :=B(z, r)B(0, d(0, z))andB+(z) :=B+(z, d(z, A(z))).

By definition of the parent,B+(z)∩ N = for allz∈ N.

Definition 2.1 does not specify the shape of edges, but the results announced in Theorem 1.1 only concern the graph structure of the hyperbolic RST, so their veracity does not depend on the geometry of edges. It is more natural to represent edges with hyperbolic geodesics, but we do another choice which will appear more convenient for the proofs. Given z1 = (r1;u1), z2 = (r2;u2)Hd+1 such that0/ [z1, z2], we define a path[z1, z2], in an isotropic way, verifying the two following conditions:

i) the distance to the origin0 is monotonous along the path[z1, z2], ii) the distance toz1 is also monotonous along this path.

It will be necessary for the proofs that the shape of edges satisfy conditions (i) and (ii), and the geodesic [z1, z2] does not verify condition (i) in general. Since 0 / [z1, z2], u1 and u2 are not antipodal, thus one can consider the unique geodesic pathγu1,u2 : [0,1]U T0Hd+1 on the sphere with constant speed connectingu1 tou2. Hence we define the path[z1, z2] as

[0,1] Hd+1

t 7→ (1t)r1+tr2;γu1,u2r1,r2,u\1,u2(t))

, (2.8)

(9)

whereφr1,r2,u\1,u2 : [0,1][0,1]is defined as:

φr1,r2,u\1,u2(t) := 1

u\1, u2arccos

(1t) sinh(r1) +tcos(u[1u2) sinh(r2) sinh((1t)r1+tr2)

.

This functionφr1,r2,u\1,u2 is built to ensure that the distance to the originz1is monotonous along the path[z1, z2]. Indeed, by the hyperbolic law of cosines (2.7),

coshd z1,((1t)r1+tr2;γu1,u2(φ(t))

= cosh(r1) cosh((1t)r1+tr2)cos(φ(t)(u\1, u2)) sinh(r1) sinh((1t)r1+tr2)

=t

cosh(r1) cosh(r2)cos(u\1, u2) sinh(r1) sinh(r2) is non-decreasing int.

We define [z1, z2[:= [z1, z2]\{z2} and ]z1, z2] := [z1, z2]\{z1}. It is possible to assume that N does not contain two points z1, z2 such that 0[z1, z2] since this event has probability1. Let us now define the random set RST by connecting each pointz∈ N toA(z)by the path[z, A(z)]:

RST:= G

z∈N

[z, A(z)[.

It may exists some pointsz belonging to several paths[z1, A(z1)[, ...,[zk, A(zk)[; in that case,z is counted with multiplicitykin RST. Formally, RST=S

z∈N[z, A(z)[×{z} ⊂Hd+1×Hd+1., i.e.

an elementz= (z0, z00)RST is the data of a pointzHd+1 and an edge[z00, A(z00)] containing z0. Forz= (z0, z00)RST, we define

z=z00, z=A(z00).

In the following, we will commit an abuse of notations by considering that RST Hd+1 and identifying an elementz= (z0, z00)RST to the corresponding pointz0 Hd+1. GivenzRST, letn:= min{k0, A(k)(z) = 0}be the number of steps required to reach the origin fromz; we define thetrajectory fromz as

π(z) := [z, z]

n−1

[

k=0

h

A(k)(z), A(k+1)(z)i

.

Forr >0, we define thelevel ras

Lr:=RSTS(r).

For0< rr0 and forz0∈ Lr0, theancestor at level rofz0, denoted byArr0(z0)is the intersection point ofπ(z0)andS(r). For0< rr0 and forz∈ Lr, theset of descendents at level r0 is defined asDrr0(z) :={z0 ∈ Lr0, zπ(z0)}(we extend the notation forz /∈ Lr by settingDrr0(z) :=∅). For z= (r;u)RST, thedescendents subtree ofzis defined asD(z) :=S

r0≥rDrr0(z). In addition, we callinfinite backward path a sequence(zi)i∈N Hd+1N

such thatz0= 0andzi=A(zi+1)for all i0.

Let us end this section with basic properties about RST proved in Appendix 6.

Proposition 2.2. The RST is a tree and it has finite degree a.s. Moreover, in the bi-dimensional case (d= 1), the representation of the RST obtained by connecting each vertex z∈ N to its parent A(z)by the geodesic[z, A(z)](instead of[z, A(z)]) is planar, i.e. their is no two pointsz1, z2∈ N such that[z1, A(z1)][z2, A(z2)]6=∅.

2.3 Sketch of proofs

In order to prove our main result (Theorem 1.1), the key point is to upper-bound angular deviations of trajectories. We first introduce two quantities, the Cumulative Forward angular Deviations (CFD)andMaximal Backward Deviations (MBD) to quantify those fluctuations.

(10)

0 r

r0

z2∈ Lr

z1∈ Lr0

Arr0(z1)

Drr0(z2)

Figure 3: Representation of levelsrandr0, the ancestorArr0(·)and the set of descendentsDrr0(·)

Definition 2.3 (Cumulative Forward angular Deviations). Let0< rr0 andz0 S(r0). Ifz0/ RST, we set CFDrr0(z0) = 0 by convention and we now suppose that z0 RST. Letz :=Arr0(z0).

We define theCumulative Forward angular Deviations ofz0 between levelsr andr0 as

CFDrr0(z0) :=

zd00zifz=z0, z[00z0+

n−1

X

k=0

A(k)(z0\)0A(k+1)(z0) +zd0z else,

wherenis the unique non negative integer such thatA(n)(z0) =z.

Definition 2.4(Maximal Backward angular Deviations). Let0< rr0 andzS(r). We define theMaximal Backward angular Deviations between levelsrandr0 as

MBDrr0(z) :=

0 ifz /RST, sup

r00∈[r,r0]

max

z00∈Drr00(z)

CFDrr00(z00)ifzRST.

We extend the definition tor0 =by setting:

MBDr (z) := lim

r0→∞MBDrr0(z), the limit exists sincer07→MBDrr0(z)is non-decreasing.

These quantities will be upper-bounded in two steps. First, a percolation argument is used to control angular deviations in any annulus of widthδ >0 for some smallδ > 0 (Proposition 2.5) and then we deduce a global control of angular deviations (Proposition 2.6). Recall thatBS(r)(·,·) is defined in (2.4).

(11)

Proposition 2.5. There existsδ >0 such that, for anyp1, there exists C=C(d, p)>0 such that for anyr >0,θ0 and any direction uSd,

E

X

z∈BS(r)(u,θ)∩RST

MBDr+δr (z)p

der(d−p), (2.9)

Proposition 2.6. For anyp large enough, there exists some constantCfl >0 such that, for any 0< r0<∞,A >0 and any directionuS1,

E

X

z∈BS(r0 )(u,Ae−r0)∩RST

MBDr0(z)p

CflAde−r0p.

These controls of angular deviations will be first used to show that the RST isstraight (Propo- sition 2.7). The straightness property is the key to show (i), (ii) and (iv) in Theorem 1.1.

Proposition 2.7 (straightness property). Almost surely, the following happens. For any ε >0, there exists someR0>0, such that, for any radius r0 R0, for anyz RST with d(0, z)r0, the descendents subtreeD(z)is contained in a cone of apex 0 and aperture e−(1−ε)r0, i.e. for any z0, z00∈ D(z),z[00z00e−(1−ε)r0.

The proof of (iii) in Theorem 1.1 exploits the controls of angular deviations (Proposition 2.6) and the link existing between the RST and the hyperbolic Directed Spanning Forest introduced in [12]: the DSF approximates locally the RST far from the origin. The unicity of the infinite backward path with some given deterministic asymptotic direction has been shown for the DSF [13], and the local coupling existing between the two models permits to show that this property remains true for the RST.

3 Proof of Theorem 1.1

Here we assume that Propositions 2.6 and 2.7 are proved and we show that it implies Theorem 1.1.

3.1 The existence part: proof of (i),(ii),(iv) and (v)

It will be shown in this section that any infinite backward path admits an asymptotic direction and that any ideal boundary point is the asymptotic direction of an infinite backward path (point (i) and (ii) in Theorem 1.1. The strategy consists, in exploiting the straightness property (Proposition 2.7):

Let(zn)be an infinite backward path, we prove that (zn)admits an asymptotic direction. For n 0, let us decompose zn in polar coordinates: zn = (rn;un). Proposition 2.7 immediately implies that the sequence(un)n≥0 is a Cauchy sequence in U T0Hd+1 'Sd, thus it converges, so (zn)converges to some boundary pointzHd+1.

Let Ψ = {limn→∞zn, (zn)is an infinite backward path} ⊂ Hd+1 be the set of asymptotic directions reached by at least one infinite backward path. In order to prove thatΨ =Hd+1, we proceed in two steps: we first show thatΨis dense in Hd+1, then we show that Ψis closed in

Hd+1.

Since the RST is an infinite tree with finite degree a.s. (Proposition 2.2), there exists an infinite backward path from0and the corresponding infinite backward path converges to an ideal boundary point by the previous paragraph, thusΨ6=almost surely.

We denote by Stab(0)the set of isometries that fix0, in particular it contains rotations centered at0. LetB be an open subset ofHd+1. SinceHd+1'Sd is compact, there exists finitely many isometriesγ1, ..., γk Stab(0)such thatS

i=1,...,kγiB =Hd+1. The random set RST is invariant in distribution by Stab(0), so the eventsγiB6=∅}all have the same probability. SinceΨ6= almost surely,P

hS

i=1,...,kγiB6=∅}i

= 1 therefore PB 6=∅)>0. In addition, for any

Références

Documents relatifs

Such strategies are: In general if two edges e 1 and e 2 generate the same increment on the number of models of A Σ e 1 is preferred over e 2 if e 1 could bring about a change of

— If G is a strongly connected finite directed graph, the set T G of rooted directed spanning trees of G is naturally equipped with a structure of directed graph: there is a

Then, we prove that in each (deterministic) direction, there exists with probability one a unique semi-infinite path, framed by an infinite number of other semi-infinite paths of

Consider the graph in Figure 2 where the weights on the edges represent the fiability of the corresponding communication link (where 0 is the minimum and 1 is the maximum fiability

Le protocole spanning tree utilise les BPDU pour élire le commutateur root et le port root pour le réseau commuté, aussi bien que le port root et le port désigné pour chaque

Si des boucles ou des switchs en parallèle sont présents, l’algorithme désactive les liens vers les switchs non nécessaires et crée alors une structure en arbre. Cette structure

Macro RecoverEdge(v) returns the local internal edge of v which has the common ancestor nearest from the root among the internal edges that were not taken into account by its

un nouveau commutateur avec une priorité plus basse (donc plus intéressante dans le cadre de l’élec- tion du root bridge). Ce nouveau commutateur a toutes les