01.07.2013 Views

Random Processes in Hyperbolic Spaces Hyperbolic Brownian ...

Random Processes in Hyperbolic Spaces Hyperbolic Brownian ...

Random Processes in Hyperbolic Spaces Hyperbolic Brownian ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Dottorato di Ricerca <strong>in</strong> Statistica Metodologica<br />

Tesi di Dottorato XXII Ciclo – 2006/2009<br />

Dipartimento di Statistica, Probabilità e Statistiche Applicate<br />

<strong>Random</strong> <strong>Processes</strong> <strong>in</strong> <strong>Hyperbolic</strong> <strong>Spaces</strong><br />

<strong>Hyperbolic</strong> <strong>Brownian</strong> Motion and <strong>Processes</strong> with F<strong>in</strong>ite<br />

Velocity <strong>in</strong> the <strong>Hyperbolic</strong> Plane<br />

Valent<strong>in</strong>a Cammarota


Contents<br />

Introduction v<br />

1 Tools of Riemannian Geometry 1<br />

1.1 Sett<strong>in</strong>gs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1<br />

1.2 Laplace operator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6<br />

1.3 Sectional curvature . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8<br />

1.4 <strong>Hyperbolic</strong> space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10<br />

2 <strong>Hyperbolic</strong> Geometry 13<br />

2.1 Möbius transformations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13<br />

2.2 Po<strong>in</strong>caré extension of Möbius transformations . . . . . . . . . . . . . . . . . . . . . 16<br />

2.3 <strong>Hyperbolic</strong> metric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

2.4 Geodesics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

2.5 <strong>Hyperbolic</strong> trigonometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

3 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion 29<br />

3.1 Sett<strong>in</strong>gs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.2.1 Transition function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31<br />

3.2.2 Hitt<strong>in</strong>g distribution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

3.2.3 <strong>Hyperbolic</strong> <strong>Brownian</strong> bridge . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

3.2.4 Fractional hyperbolic <strong>Brownian</strong> motion . . . . . . . . . . . . . . . . . . . . 43<br />

3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.1 Transition function <strong>in</strong> H 3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44<br />

3.3.2 The Millson recursive formula . . . . . . . . . . . . . . . . . . . . . . . . . . 45<br />

3.3.3 Bounds for the hyperbolic heat kernel . . . . . . . . . . . . . . . . . . . . . 47<br />

iii


CONTENTS CONTENTS<br />

3.3.4 Gruet’s formula for the heat kernel . . . . . . . . . . . . . . . . . . . . . . . 49<br />

3.3.5 <strong>Hyperbolic</strong> branch<strong>in</strong>g <strong>Brownian</strong> motion . . . . . . . . . . . . . . . . . . . . 52<br />

3.A Appendix . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

4 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass 59<br />

4.1 Description of the planar random motion on H 2 . . . . . . . . . . . . . . . . . . . . 59<br />

4.2 Equations related to the mean hyperbolic distance . . . . . . . . . . . . . . . . . . 62<br />

4.3 About the higher moments of the hyperbolic distance . . . . . . . . . . . . . . . . 69<br />

4.4 Motions with jumps backwards to the start<strong>in</strong>g po<strong>in</strong>t . . . . . . . . . . . . . . . . . 74<br />

4.5 Motion at f<strong>in</strong>ite velocity on the surface of a three-dimensional sphere . . . . . . . . 80<br />

5 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2 83<br />

5.1 Description of the randomly mov<strong>in</strong>g and branch<strong>in</strong>g model . . . . . . . . . . . . . . 83<br />

5.2 Mean hyperbolic distance: the Laplace transform approach . . . . . . . . . . . . . 86<br />

5.3 Equation govern<strong>in</strong>g the hyperbolic distance . . . . . . . . . . . . . . . . . . . . . . 92<br />

5.4 Mean hyperbolic distance of the k-th spl<strong>in</strong>ter . . . . . . . . . . . . . . . . . . . . . 97<br />

iv<br />

101


Introduction<br />

Euclid wrote his famous Elements around 300 B.C. In the first book of the Elements Euclid develops<br />

plane geometry start<strong>in</strong>g with the follow<strong>in</strong>g five postulates<br />

• A straight l<strong>in</strong>e may be drawn from any po<strong>in</strong>t to any other po<strong>in</strong>t.<br />

• A f<strong>in</strong>ite straight l<strong>in</strong>e may be extended cont<strong>in</strong>uously <strong>in</strong> a straight l<strong>in</strong>e.<br />

• A circle may be drawn with any center and any radius.<br />

• All right angles are equal.<br />

• If a straight l<strong>in</strong>e fall<strong>in</strong>g on two straight l<strong>in</strong>es makes the <strong>in</strong>terior angles on the same side less<br />

than two right angles, the two straight l<strong>in</strong>es, if extended <strong>in</strong>def<strong>in</strong>itely, meet on the side on<br />

which the angles are less than two right angles.<br />

Euclid’s fifth postulate is equivalent to the modern postulate of Euclidean geometry<br />

• Through a po<strong>in</strong>t outside a given <strong>in</strong>f<strong>in</strong>ite l<strong>in</strong>e there is one and only one <strong>in</strong>f<strong>in</strong>ite straight l<strong>in</strong>e<br />

parallel to the given l<strong>in</strong>e.<br />

For two thousand years it was believed that the fifth postulate could be derived from the other<br />

four postulates and for centuries the mathematicians attempted to prove the fifth postulate. It<br />

was not until the n<strong>in</strong>eteenth century that the fifth postulate was f<strong>in</strong>ally shown to be <strong>in</strong>dependent.<br />

The proof of this <strong>in</strong>dependence was the result of a completely unexpected discovery: the denial of<br />

the fifth postulate leads to a new consistent geometry.<br />

It was Gauss who first made this discovery. It was known from the eighteenth century that the<br />

fifth postulate is equivalent to the theorem that the sum of the angles of a triangle is 180 o , Gauss,<br />

assum<strong>in</strong>g that the sum of the angles of a triangle is less than 180 o , discovered the non-Euclidean<br />

geometry.<br />

Only few years later Lobachevsky published the first account of non-Euclidean geometry <strong>in</strong> 1829<br />

<strong>in</strong> a paper entitled On the pr<strong>in</strong>ciple of geometry. In 1832, Bolay published an <strong>in</strong>dependent account<br />

of non-Euclidean geometry <strong>in</strong> a paper entitled The absolute science of space.<br />

Today the non-Euclidean geometry of Gauss, Lobachevsky and Bolay is called hyperbolic geometry,<br />

and the term non-Euclidean geometry refers to any geometry that is not Euclidean.<br />

After def<strong>in</strong><strong>in</strong>g spaces of variable curvature and not<strong>in</strong>g that the Euclidean space is a space of<br />

zero curvature, Riemann considered spaces of constant non zero curvature. An example of a ndimensional<br />

Riemannian space of constant positive curvature is the n-dimensional sphere <strong>in</strong> the


(n + 1)-dimensional Euclidean space. In the sett<strong>in</strong>g of Riemannian geometry an example of a<br />

n-dimensional space of constant negative curvature is the n-dimensional Lobachevskian space. For<br />

a comprehensive history of non-Euclidean geometry see Rosenfeld (38).<br />

In Chapter 1 are <strong>in</strong>troduced some fundamental results of Riemannian geometry, the aim of<br />

this chapter is to prove that the hyperbolic space is a space with constant negative curvature equal<br />

to −1 and to obta<strong>in</strong> an explicit expression, <strong>in</strong> local coord<strong>in</strong>ate, of the Laplace operator <strong>in</strong> the<br />

n-dimensional hyperbolic space. On Riemannian geometry see, e.g., Chavel (8), (9) and Spivak<br />

(40).<br />

Chapter 2 is devoted to hyperbolic geometry. In this chapter are <strong>in</strong>troduced the half-space<br />

model and the disk model of the hyperbolic geometry. We give some explicit formulas for the<br />

hyperbolic metric and a characterization of the geodesics (that is curves of shortest hyperbolic<br />

length) <strong>in</strong> these models. In the last section of the chapter it is proved the famous Lobachevsky’s<br />

formula and the <strong>Hyperbolic</strong> Pythagorean Theorem. For a complete account on hyperbolic geometry<br />

see, for example, Beardon (2), Kulczychi (23), Ratcliffe (35) and Terras (41).<br />

<strong>Brownian</strong> motion on hyperbolic spaces has been studied s<strong>in</strong>ce the end of the Fifties. In 1959<br />

Gertsenshte<strong>in</strong> and Vasiliev (15) gave the explicit form of the transition function for the hyperbolic<br />

<strong>Brownian</strong> motion <strong>in</strong> two and three dimensional hyperbolic spaces. The studies on hyperbolic <strong>Brownian</strong><br />

motion have been revitalized by mathematical f<strong>in</strong>ance s<strong>in</strong>ce some f<strong>in</strong>ancial products (Asian<br />

options) have a strict connection with the stochastic representation of the hyperbolic <strong>Brownian</strong><br />

motion (see, for example, Yor (43)). In Chapter 3 we give a description of the pr<strong>in</strong>cipal results<br />

obta<strong>in</strong>ed <strong>in</strong> the literature about two dimensional and multi-dimensional hyperbolic <strong>Brownian</strong><br />

motion: stochastic representation, transition function, hyperbolic <strong>Brownian</strong> bridge, hitt<strong>in</strong>g distribution,<br />

fractional hyperbolic <strong>Brownian</strong> motion, the Millson recursive formula, uniform estimations<br />

for the transition function and branch<strong>in</strong>g hyperbolic <strong>Brownian</strong> motion.<br />

More recently also works concern<strong>in</strong>g two-dimensional random motions at f<strong>in</strong>ite velocity on planar<br />

hyperbolic spaces have been <strong>in</strong>troduced and analyzed, see Ors<strong>in</strong>gher and De Gregorio (34). While<br />

<strong>in</strong> (34) the components of the motion are supposed to be <strong>in</strong>dependent, we present <strong>in</strong> Chapter 4 a<br />

planar random motion with <strong>in</strong>teract<strong>in</strong>g components. We study the motion of particle that runs on<br />

the geodesic l<strong>in</strong>es of the hyperbolic plane chang<strong>in</strong>g direction at Poisson-paced times. The motion<br />

considered here is the non-Euclidean counterpart of the planar motion with orthogonal deviations<br />

studied <strong>in</strong> Ors<strong>in</strong>gher (33). The ma<strong>in</strong> object of the <strong>in</strong>vestigation is the hyperbolic distance η(t)<br />

at time t of the mov<strong>in</strong>g po<strong>in</strong>t from the orig<strong>in</strong>. We are able to give explicit expressions for its<br />

mean value, also under the condition that the number of changes of direction is known. In the<br />

case of motion <strong>in</strong> the hyperbolic plane with <strong>in</strong>dependent components an explicit expression for the<br />

distribution of the hyperbolic distance η(t) has been obta<strong>in</strong>ed (Ors<strong>in</strong>gher and De Gregorio (34)).<br />

Here, however, the components of motion are dependent and this excludes any possibility of f<strong>in</strong>d<strong>in</strong>g<br />

the distribution of the hyperbolic distance.<br />

We obta<strong>in</strong> the follow<strong>in</strong>g explicit formula for the mean value of the hyperbolic distance which<br />

reads<br />

<br />

λt −<br />

E{cosh η(t)} = e 2 cosh t λ t <br />

<br />

λ2 + 4c2 + √ s<strong>in</strong>h λ2 + 4c2 2<br />

λ2 + 4c2 2<br />

Section 4.4 is devoted to motions on the hyperbolic plane where the return to the start<strong>in</strong>g po<strong>in</strong>t<br />

is admitted and occurs at the <strong>in</strong>stants of changes of direction. The mean distance from the orig<strong>in</strong> of<br />

vi


vii CHAPTER 0. INTRODUCTION<br />

these jump<strong>in</strong>g-back motions is obta<strong>in</strong>ed explicitly by exploit<strong>in</strong>g their relationship with the motion<br />

without jumps. In the case where the return to the start<strong>in</strong>g po<strong>in</strong>t occurs at the first Poisson event<br />

T1, the mean value of the hyperbolic distance η1(t) reads<br />

E{cosh η1(t)|N(t) ≥ 1} =<br />

λ<br />

√ λ 2 + 4c 2<br />

s<strong>in</strong>h t<br />

√<br />

2 λ2 + 4c2 s<strong>in</strong>h λt<br />

2<br />

The last section of the chapter considers the motion at f<strong>in</strong>ite velocity, with orthogonal deviations<br />

at Poisson times, on the unit-radius sphere. The ma<strong>in</strong> results concern the mean value<br />

E{cos d(P0Pt)}, where d(P0Pt) is the distance of the current po<strong>in</strong>t Pt from the start<strong>in</strong>g position<br />

P0. We take profit of the analogy of the spherical motion with its counterpart on the hyperbolic<br />

plane to discuss the different situations due to the f<strong>in</strong>iteness of the space where the random motion<br />

develops. See also Cammarota and Ors<strong>in</strong>gher (5) and (7).<br />

In Chapter 5 we study a random motion of a cloud of particles mov<strong>in</strong>g at f<strong>in</strong>ite velocity on<br />

geodesic l<strong>in</strong>es of the hyperbolic plane. Such particles are generated by successive splitt<strong>in</strong>g out of<br />

a unit-mass particle <strong>in</strong>itially placed at the orig<strong>in</strong> O. The dis<strong>in</strong>tegration process of the unit-mass<br />

particle is governed by an underly<strong>in</strong>g Poisson process of rate λ as follows. The orig<strong>in</strong>al particle<br />

keeps mov<strong>in</strong>g on the ma<strong>in</strong> geodesic l<strong>in</strong>e with constant hyperbolic velocity c; at the first Poisson<br />

event it breaks <strong>in</strong>to two parts of equal mass 1/2. One particle cont<strong>in</strong>ues its motion on the ma<strong>in</strong><br />

geodesic l<strong>in</strong>e, while the other one deviates orthogonally. At the second Poisson event the deviated<br />

particle is separated <strong>in</strong>to two pieces each of mass 1/2 2 ; one piece cont<strong>in</strong>ues its motion on the same<br />

geodesic l<strong>in</strong>e whereas the other one starts mov<strong>in</strong>g on the geodesic l<strong>in</strong>e orthogonal to that jo<strong>in</strong><strong>in</strong>g<br />

its position with the orig<strong>in</strong> O. In general, at the k-th Poisson event, only the deviat<strong>in</strong>g particle of<br />

mass 1/2 k breaks <strong>in</strong>to two fragments of equal mass 1/2 k+1 ; the first cont<strong>in</strong>ues its motion on the<br />

same geodesic l<strong>in</strong>e while the other one deviates orthogonally.<br />

If up to time t, N(t) Poisson events have occurred, we have N(t) + 1 particles runn<strong>in</strong>g, at a<br />

constant hyperbolic velocity c, along different geodesic l<strong>in</strong>es with a mass depend<strong>in</strong>g on the <strong>in</strong>stant<br />

of separation from the generat<strong>in</strong>g particle.<br />

In the above branch<strong>in</strong>g process each particle can reproduce only once and the particle splitt<strong>in</strong>g<br />

at time Tk of the k-th Poisson event and which never more dis<strong>in</strong>tegrates will preserve its mass,<br />

equal to 1/2 k , for the successive time <strong>in</strong>terval (Tk, t).<br />

Our ma<strong>in</strong> result concerns the dynamics of the center of mass cm of the cloud of particles perform<strong>in</strong>g<br />

the branch<strong>in</strong>g and diffusion process. In particular, we are able to give an exact expression<br />

for the mean hyperbolic distance from the orig<strong>in</strong> O, ηcm(t), of the center of mass cm at any time<br />

t > 0<br />

E{cosh ηcm(t)} = 23c2 <br />

3 −<br />

e 22 λt<br />

√<br />

λ2 + 24c2 t −<br />

e 22 √<br />

λ2 +24c2 3 √ λ2 + 24c2 + 5λ +<br />

+ λ + 2c<br />

2(λ + 3c) ect λ − 2c<br />

+<br />

2(λ − 3c) e−ct .<br />

.<br />

e t<br />

22 √<br />

λ2 +24c2 3 √ λ 2 + 2 4 c 2 − 5λ<br />

We give two different and <strong>in</strong>dependent proofs of the above result: our first technique is based on<br />

Laplace transforms, while the other one br<strong>in</strong>gs about the follow<strong>in</strong>g non-homogeneous second-order


differential equation<br />

3 −<br />

22 λt<br />

d2 dt2 u − c2u = λc2e √<br />

λ2 + 24c2 t −<br />

e 22 √<br />

λ2 +24c2 − e t<br />

22 √<br />

λ2 +24c2 <br />

.<br />

We also exam<strong>in</strong>e the mean hyperbolic distance of each <strong>in</strong>dividual particle which stops chang<strong>in</strong>g<br />

direction after the k-th Poisson event. Our ma<strong>in</strong> result shows that<br />

where g(s; k, λ) = e−λs λ k s k−1<br />

Γ(k)<br />

E{cosh ηk(t)I {N(t)≥k}} = 1<br />

2 k<br />

t<br />

0<br />

cosh c(t − s)h(k, c, s)g(s; k, λ) ds<br />

is a Gamma distribution and h(k, c, s) = k<br />

r=0<br />

viii<br />

k cs(2Yr,k−1)<br />

r E{e }<br />

with Yr,k ∼ Beta(r, k − r) so that (2Yr,k − 1) ∈ (−1, 1) and with the assumption that Y0,k = 1 and<br />

Yk,k = −1. Thus the last formula shows that the mean hyperbolic distance, at time t, of a particle<br />

generated at the k-th Poisson event can be seen as the mean hyperbolic distance of a particle which<br />

never deviates from the ma<strong>in</strong> geodesic l<strong>in</strong>e and which stops mov<strong>in</strong>g at a random time with law<br />

h(k, c, s)g(s; k, λ). See also Cammarota and Ors<strong>in</strong>gher (6).


Chapter 1<br />

Tools of Riemannian Geometry<br />

1.1 Sett<strong>in</strong>gs<br />

Def<strong>in</strong>ition 1.1.1. A topological space is a set X together with a collection O of subsets of X<br />

satisfy<strong>in</strong>g the follow<strong>in</strong>g properties<br />

• ∅, X ∈ O<br />

• the <strong>in</strong>tersection of any f<strong>in</strong>ite collection of sets <strong>in</strong> O is also <strong>in</strong> O,<br />

• the union of any collection of sets <strong>in</strong> O is also <strong>in</strong> O.<br />

The sets <strong>in</strong> O are called open.<br />

Def<strong>in</strong>ition 1.1.2. A topological space is Hausdorff if for any two dist<strong>in</strong>ct po<strong>in</strong>ts x1, x2 ∈ X there<br />

exists open sets O1, O2 ∈ O such that x1 ∈ O1, x2 ∈ O2 and O1 ∩ O2 = ∅.<br />

The Hausdorff axiom is essential for the uniqueness of the limit of convergent sequences.<br />

Def<strong>in</strong>ition 1.1.3. A map between two topological spaces is called cont<strong>in</strong>uos if the <strong>in</strong>verse image<br />

of any open set is open. A bijective map which is cont<strong>in</strong>uous <strong>in</strong> both directions is called<br />

homeomorphism.<br />

Def<strong>in</strong>ition 1.1.4. A n-dimensional manifold is a Hausdorff topological space M such that for each<br />

x ∈ M there is a neighborhood U of x that is homeomorphic to an open subset O of R n . Such a<br />

homeomorphism ϕ : U → O is called coord<strong>in</strong>ate chart.<br />

The important po<strong>in</strong>t here is that locally the topology of a manifold is the same as that of R n .<br />

Def<strong>in</strong>ition 1.1.5. Consider a family (Ui)i∈I of open sets such that M = <br />

i∈I Ui. An atlas is the<br />

family {Ui, ϕi}i∈I of charts.<br />

Remark 1.1.1. Each ϕ −1<br />

i (·) is referred to as parametrization and the set ϕi(Ui) is called parameter<br />

doma<strong>in</strong>.<br />

Def<strong>in</strong>ition 1.1.6. A n-dimensional manifold M is C r differentiable if, for Ui ∩ Uj = ∅, the<br />

composition<br />

ϕj ◦ ϕ −1<br />

i<br />

: ϕi(Ui ∩ Uj) → ϕj(Ui ∩ Uj)


1.1 Sett<strong>in</strong>gs 2<br />

is C r differentiable for arbitrary i, j.<br />

In what follows we will assume that M and N are two C ∞ differentiable manifolds.<br />

Def<strong>in</strong>ition 1.1.7. Let M be a n-dimensional manifold and N be a m-dimensional manifold. A<br />

function f : M → N is said to be differentiable if for all charts ϕi : Ui → Rn , ψi : Vi → Rm with<br />

f(Ui) ⊂ Vi the composition ψi ◦ f ◦ ϕ −1<br />

i is also differentiable.<br />

The tangent space of a n-dimensional manifold M at a po<strong>in</strong>t p ∈ M is go<strong>in</strong>g to be thought of<br />

as the n-dimensional set of “directions” which, start<strong>in</strong>g at p, po<strong>in</strong>t <strong>in</strong> all directions of M. Without<br />

mak<strong>in</strong>g reference to the ambient space this notion has to be <strong>in</strong>tr<strong>in</strong>sically def<strong>in</strong>ed as follows.<br />

Def<strong>in</strong>ition 1.1.8. A tangent vector X(p) at p ∈ M is a derivative operator def<strong>in</strong>ed on the set of<br />

germs of functions Fp(M) := {f : M → R|f differentiable}/ ∼, where the equivalence relation ∼<br />

is def<strong>in</strong>ed by declar<strong>in</strong>g f ∼ f ∗ if and only if f and f ∗ co<strong>in</strong>cide <strong>in</strong> a neighborhood of p.<br />

Def<strong>in</strong>ition 1.1.8 means more precisely that the tangent vector X(p) is a map X(p) : Fp(M) → R<br />

with the two follow<strong>in</strong>g properties<br />

X(p)(αf + βg) = αX(p)f + βX(p)g, (1.1)<br />

X(p)(fg) = g(p)X(p)f + f(p)X(p)g, (1.2)<br />

where α, β ∈ R and f, g ∈ Fp(M). The value X(p)f is called directional derivative of f at p <strong>in</strong> the<br />

direction X(p).<br />

Def<strong>in</strong>ition 1.1.9. The tangent space TpM of M at p is def<strong>in</strong>ed as the set of all tangent vectors<br />

at the po<strong>in</strong>t p.<br />

Let M be a n-dimensional manifold, p ∈ U ⊂ M and let ϕ : U → R n be a chart. We denote by<br />

(u 1 , . . . , u n ) the standard coord<strong>in</strong>ates of R n and by (x 1 , . . . , x n ) the correspond<strong>in</strong>g coord<strong>in</strong>ates <strong>in</strong><br />

M. Thus x i (p) is the function given by the i-th coord<strong>in</strong>ate of ϕ(p) <strong>in</strong> R n , that is<br />

x i (p) = u i (ϕ(p)).<br />

On the other hand it is possible to see the coord<strong>in</strong>ates (x 1 , . . . , x n ) as variables with respect to<br />

which we can form derivatives. For a function f : M → R, f ∈ C 1 , we set<br />

∂f<br />

∂xi <br />

<br />

<br />

p<br />

:= ∂(f ◦ ϕ−1 )<br />

∂u i<br />

In view of Def<strong>in</strong>ition 1.1.8, special tangent vectors at p are the partial derivatives ∂i(p), with<br />

i = 1, . . . , n, def<strong>in</strong>ed by<br />

∂i(p)f := ∂f<br />

∂xi <br />

<br />

.<br />

It is possible to prove that the tangent space TpM at p of the n-dimensional manifold M is<br />

a n-dimensional vector space on R spanned, <strong>in</strong> appropriately chosen coord<strong>in</strong>ate (x 1 , . . . , x n ), <strong>in</strong> a<br />

given chart by the tangent vectors ∂i(p), i = 1, . . . , n,<br />

p<br />

<br />

<br />

<br />

ϕ(p)<br />

TpM = Span{∂1(p), . . . , ∂n(p)}.<br />

.


3 Tools of Riemannian Geometry<br />

For a proof see, for example, Kühnel (24) Theorem 5.6. Therefore for every tangent vector X(p) ∈<br />

TpM one has<br />

n<br />

X(p) =<br />

i=1<br />

ξ i p ∂i(p), (1.3)<br />

where ξ i p is the directional derivative of the coord<strong>in</strong>ate function x i <strong>in</strong> the direction X(p). For a<br />

function f : M → R, f ∈ C 1 we have<br />

X(p)f =<br />

n<br />

i=1<br />

ξ i p ∂i(p)f.<br />

Def<strong>in</strong>ition 1.1.10. A differentiable vector field X on a n-dimensional manifold M is an association<br />

p ∈ U ⊂ M → X(p) ∈ TpM such that <strong>in</strong> every chart ϕ : U → R n , with coord<strong>in</strong>ates (x 1 , . . . , x n ), the<br />

coefficients ξ i p : U → R <strong>in</strong> the representation (1.3) valid at the each po<strong>in</strong>t p ∈ U are differentiable<br />

functions.<br />

Common notations for the vector field and the directional derivative of a function f : M → R,<br />

f ∈ C 1 <strong>in</strong> the direction of the vector field are<br />

X =<br />

n<br />

ξ i ∂i, Xf =<br />

i=1<br />

n<br />

ξ i ∂if.<br />

The <strong>in</strong>terpretation of X as a derivative operator permits us to consider the iterates of X. If<br />

X and Y are two differentiable vector fields on M and f : M → R, f ∈ C 2 , we can consider the<br />

functions X(Y f) and Y (Xf).<br />

Proposition 1.1.1. Let X and Y be two differentiable vector fields on a C ∞ manifold M. Then<br />

there exists a unique vector filed [X, Y ], which is referred to as the Lie bracket of X and Y , such<br />

that for all f : M → R, f ∈ C 2 ,<br />

i=1<br />

[X, Y ]f = X(Y f) − Y (Xf).<br />

For a proof see, for example, Do Carmo (14) Lemma 5.2.<br />

We <strong>in</strong>troduce now a metric structure on M.<br />

Def<strong>in</strong>ition 1.1.11. An <strong>in</strong>ner product on a vector space V over a field F is a bil<strong>in</strong>ear form<br />

such that is<br />

• symmetric: 〈v, w〉 = 〈w, v〉<br />

• positive def<strong>in</strong>ite: 〈v, v〉 > 0 for all v = 0.<br />

(v, w) ∈ V × V → 〈v, w〉 ∈ F,<br />

Def<strong>in</strong>ition 1.1.12. Given a C ∞ manifold M, def<strong>in</strong>e a Riemannian metric g on M to be a mapp<strong>in</strong>g<br />

that associates with each p ∈ M an <strong>in</strong>ner product<br />

gp : MpT × MpT → R


1.1 Sett<strong>in</strong>gs 4<br />

satisfy<strong>in</strong>g the follow<strong>in</strong>g differentiability property. If U is any open set <strong>in</strong> M and X, Y are two<br />

differentiable vector fields on U, then the function g(X, Y ) : U → R given by<br />

is differentiable on U.<br />

g(X, Y )(p) := gp(X(p), Y (p))<br />

Each tangent space is then equipped with an <strong>in</strong>ner product g that varies smoothly from po<strong>in</strong>t to<br />

po<strong>in</strong>t.<br />

Def<strong>in</strong>ition 1.1.13. A Riemannian manifold is a C ∞ manifold M with a given Riemannian metric<br />

g.<br />

For a given Riemannian metric g we def<strong>in</strong>e<br />

〈·, ·〉 := g(·, ·), gjk = gkj := 〈∂j, ∂k〉, G := (gjk), G −1 := (g jk ), g := det G,<br />

where j, k = 1 . . . , n.<br />

Def<strong>in</strong>ition 1.1.14. Let X and Y be two differentiable vector fields on a manifold M, a connection<br />

is a mapp<strong>in</strong>g that associates to X and Y the differentiable vector field ▽XY that satisfies the<br />

follow<strong>in</strong>g properties<br />

▽fX+gY Z = f▽XZ + g▽Y Z, (1.4)<br />

▽X(Y + Z) = ▽XY + ▽XZ, (1.5)<br />

▽X(fY ) = f▽XY + (Xf)Y. (1.6)<br />

where X, Y and Z are differentiable vector fields, f, g are two differentiable functions on M and<br />

fX denotes the vector field (fX)(p) := f(p)X(p).<br />

The follow<strong>in</strong>g theorem holds (for a proof see, e.g., Do Carmo (14) Theorem 3.6).<br />

Theorem 1.1.1 (Levi-Civita). Given a Riemannian manifold M, there exists a unique connection<br />

▽ on M satisfy<strong>in</strong>g the conditions<br />

for all X, Y and Z differentiable vector fields on M.<br />

▽XY − ▽Y X = [X, Y ], (1.7)<br />

X〈Y, Z〉 = 〈▽XY, Z〉 + 〈Y, ▽XZ〉, (1.8)<br />

The connection def<strong>in</strong>ed by Theorem 1.1.1 is referred to as the Levi-Civita (or Riemannian) connection.<br />

Let X and Y be two differentiable vector fields on M such that<br />

X =<br />

n<br />

ξ i ∂i, Y =<br />

i=1<br />

In view of (1.4), (1.5) and (1.6) we have<br />

i<br />

n<br />

η j ∂j.<br />

j=1<br />

▽XY = ▽P i ξi <br />

∂iY = ξ i ▽∂iY = <br />

ξ i ▽∂i(η j ∂j) = <br />

= <br />

ξ i η j ▽∂i∂j + <br />

ξ i (∂iη j )∂j = <br />

ξ i η j Γ k ji∂k + <br />

i,j<br />

i,j<br />

i,j<br />

i,j,k<br />

i,j<br />

ξ i [η j ▽∂i∂j + (∂iη j )∂j]<br />

i,j<br />

ξ i (∂iη j )∂j, (1.9)


5 Tools of Riemannian Geometry<br />

where the functions Γ k ij<br />

def<strong>in</strong>ed by<br />

From (1.9) if follows that<br />

: U ⊂ M → R, i, j, k = 1, . . . , n, referred to as Christoffel symbols, are<br />

▽∂j ∂i =<br />

n<br />

k=1<br />

Γ k ij∂k.<br />

▽XY − ▽Y X = <br />

ξ i η j Γ k ji∂k + <br />

ξ i (∂iη j )∂j − <br />

η j ξ i Γ k ij∂k − <br />

i,j,k<br />

i,j<br />

= <br />

ξ i (∂iη j )∂j − <br />

i,j<br />

= <br />

i,j<br />

i,j<br />

[ξ i (∂iη j ) − η i (∂iξ j )]∂j<br />

η j (∂jξ i )∂i<br />

i,j,k<br />

i,j<br />

η j (∂jξ i )∂i<br />

(1.10)<br />

In order to obta<strong>in</strong> an explicit expression of the Γ k ij , by sett<strong>in</strong>g X = ∂i and Y = ∂j, we have<br />

from (1.10) that<br />

▽∂i ∂j − ▽∂j ∂i = [∂i − ∂i]∂j = 0. (1.11)<br />

But, s<strong>in</strong>ce<br />

▽∂i∂j − ▽∂j ∂i = <br />

(Γ k ji − Γ k ij)∂k, (1.12)<br />

we obta<strong>in</strong> from (1.11) and (1.12) that, for all i, j, k = 1, . . . , n,<br />

If we set X = ∂i, Y = ∂j and Z = ∂k <strong>in</strong> (2.20) then it becomes<br />

that is<br />

k<br />

Γ k ij = Γ k ji. (1.13)<br />

∂i〈∂j, ∂k〉 = 〈▽∂i ∂j, ∂k〉 + 〈∂j, ▽∂i ∂k〉<br />

∂igjk = 〈 <br />

Γ l ji∂l, ∂k〉 + 〈∂j, <br />

Γ l ki∂l〉 = <br />

Γ l ji glk + <br />

l<br />

l<br />

l<br />

l<br />

Γ l ki gjl<br />

= <br />

[Γ l ji glk + gjl Γ l ki]. (1.14)<br />

l<br />

From (1.13) and (1.14) we deduce the follow<strong>in</strong>g explicit expression for the Γ k ij<br />

Γ k ij = 1<br />

2<br />

<br />

g kl {∂i glj + ∂j gil − ∂l gij} (1.15)<br />

l


1.2 Laplace operator 6<br />

<strong>in</strong> fact<br />

<br />

g kl {∂i glj + ∂j gil − ∂l gij} = <br />

g kl<br />

<br />

<br />

[Γ m li gmj + glm Γ m ji] + <br />

l<br />

1.2 Laplace operator<br />

l<br />

− <br />

m<br />

m<br />

[Γ m il gmj + gim Γ m jl ]<br />

<br />

m<br />

[Γ m ij gml + gim Γ m lj ]<br />

= <br />

g kl Γ m li gmj + glmΓ m ji + Γ m ij gml + gimΓ m lj − Γ m il gmj − gimΓ m jl<br />

l,m<br />

= <br />

g kl glmΓ m ji + Γ m <br />

ij gml = 2<br />

l,m<br />

= 2 Γ k ij.<br />

l,m<br />

g kl Γ m ij glm<br />

Def<strong>in</strong>ition 1.2.1. Given a function f : M → R, f ∈ C 1 , the gradient of f, grad f, is the vector<br />

field on M such that<br />

〈grad f, X〉 = Xf (1.16)<br />

for all differentiable vector fields X on M.<br />

Given two functions f, h : M → R, f, h ∈ C 1 , it follows that<br />

In fact, <strong>in</strong> view of (1.1), we have<br />

grad (f + h) = grad f + grad h, (1.17)<br />

grad (fh) = h(grad f) + f(grad h). (1.18)<br />

〈grad (f + h), X〉 = X(f + h) = Xf + Xh = 〈grad f, X〉 + 〈grad h, X〉<br />

and <strong>in</strong> view (1.2), we have<br />

= 〈grad f + grad h, X〉,<br />

〈grad (fh), X〉 = X(fh) = h(Xf) + f(Xh) = h〈grad f, X〉 + f〈grad h, X〉<br />

= 〈h grad f + f grad h, X〉.<br />

We now calculate the expression of gradf <strong>in</strong> local coord<strong>in</strong>ates. Observ<strong>in</strong>g that, for j, k, l = 1, . . . , n,<br />

we have<br />

<br />

gjkg kl ∂l = <br />

δjl∂l = ∂j,<br />

k,l<br />

where δjl is the Kronecker delta, and<br />

<br />

gjkg kl ∂lf = <br />

〈∂j, ∂k〉g kl ∂lf = 〈∂j, <br />

g kl (∂lf)∂k〉,<br />

k,l<br />

k,l<br />

l<br />

k,l


7 Tools of Riemannian Geometry<br />

we obta<strong>in</strong><br />

Xf =<br />

n<br />

ξ j ∂jf = <br />

j=1<br />

j,k,l<br />

ξ j gjkg kl ∂lf =<br />

Now, <strong>in</strong> view of Def<strong>in</strong>ition 1.2.1 and formula (1.19), we f<strong>in</strong>ally have<br />

n<br />

ξ j 〈∂j, <br />

(g kl ∂lf)∂k〉 = 〈X, <br />

g kl (∂lf)∂k〉. (1.19)<br />

j=1<br />

k,l<br />

grad f = <br />

g kl (∂lf)∂k = <br />

k,l<br />

k,l<br />

g kl ∂(f ◦ ϕ−1 )<br />

∂u l<br />

∂<br />

k,l<br />

. (1.20)<br />

∂xk From formula (1.20) and <strong>in</strong> view of the representation (1.3) it follows that, if f ∈ C k with k ≥ 1,<br />

then grad f is a C k−1 vector field.<br />

Def<strong>in</strong>ition 1.2.2. Given a differentiable vector field Y on M, def<strong>in</strong>e the real-valued function<br />

divergence of Y , div Y , by<br />

where X(p) ranges over MpT .<br />

(divY )(p) = trace(X(p) → ▽ X(p)Y )<br />

To obta<strong>in</strong> the expression of div Y <strong>in</strong> local coord<strong>in</strong>ates we observe that, <strong>in</strong> view of (1.9), we have<br />

therefore<br />

▽XY = <br />

ξ i [ <br />

div Y =<br />

i,k<br />

j<br />

⎧<br />

n ⎨<br />

m=1<br />

Insert<strong>in</strong>g formula (1.15) <strong>in</strong> (1.21), we obta<strong>in</strong><br />

div Y = <br />

<br />

<br />

j<br />

= 1<br />

2<br />

l<br />

η l Γ j j<br />

lj + ∂jη<br />

<br />

⎩<br />

j<br />

= <br />

j,l<br />

η j Γ k ji + ∂iη k ]∂k,<br />

η j Γ m jm + ∂mη m<br />

η l Γ j<br />

lj<br />

<br />

η j<br />

<br />

<br />

g lk <br />

{∂jgkl + ∂lgjk − ∂kgjl} + <br />

∂jη j<br />

j,l<br />

k<br />

<br />

j<br />

= 1 <br />

η g<br />

2<br />

j,l k<br />

lk ∂jgkl + <br />

∂jη<br />

j<br />

j<br />

= <br />

⎧<br />

⎨<br />

1 <br />

ηj g<br />

⎩2<br />

j<br />

k,l<br />

kl ∂jglk + ∂jη j<br />

⎫<br />

⎬<br />

⎭<br />

= <br />

<br />

1<br />

2 ηj tr(G −1 ∂jG) + ∂jη j<br />

<br />

,<br />

j<br />

⎫<br />

⎬<br />

. (1.21)<br />

⎭<br />

<br />

+ ∂jη j = <br />

η j Γ l jl + <br />

∂jη j<br />

s<strong>in</strong>ce g kl = g lk because G is symmetric and where ∂jG is the matrix obta<strong>in</strong>ed from G by differentiat<strong>in</strong>g<br />

each of the entries with respect to the j-th coord<strong>in</strong>ate. With the standard formula for<br />

j<br />

j<br />

j,l<br />

j


1.3 Sectional curvature 8<br />

differentiat<strong>in</strong>g determ<strong>in</strong>ants we have that tr(G −1 ∂jG) = ∂j(ln detG) and then<br />

div Y = div ( <br />

η<br />

j<br />

j ∂j) = <br />

<br />

1<br />

2<br />

j<br />

ηj tr(G −1 ∂jG) + ∂jη j<br />

<br />

= <br />

<br />

1<br />

2<br />

j<br />

ηj ∂j(ln g) + ∂jη j<br />

<br />

= j<br />

η ∂j(ln √ g) + ∂jη j = <br />

<br />

j 1<br />

η √ ∂j(<br />

g √ g) + ∂jη j<br />

<br />

=<br />

=<br />

j<br />

1<br />

√ g<br />

1<br />

√ g<br />

j<br />

η ∂j( √ g) + (∂jη j ) √ g <br />

j<br />

j<br />

<br />

∂j(η j√ g). (1.22)<br />

j<br />

Def<strong>in</strong>ition 1.2.3. For every function f : M → R, f ∈ C 2 , we def<strong>in</strong>e the Laplacian of f, ∆f, by<br />

∆f = div(grad f).<br />

By comb<strong>in</strong><strong>in</strong>g (1.20) and (1.22), we obta<strong>in</strong> the explicit expression of ∆f <strong>in</strong> local coord<strong>in</strong>ates:<br />

<br />

<br />

∆f = div(grad f) = div ( <br />

g kl <br />

∂lf)∂k<br />

=<br />

=<br />

1<br />

√ g<br />

1<br />

√ g<br />

<br />

∂k( <br />

g kl ∂lf √ g)<br />

k<br />

l<br />

k<br />

l<br />

<br />

∂k(g kl√ g ∂lf). (1.23)<br />

Next we consider the follow<strong>in</strong>g <strong>in</strong>variance property of the Laplace operator.<br />

k,l<br />

Theorem 1.2.1. Let Φ be a diffeomorphism of the Riemannian manifold M onto itself. Then Φ<br />

leaves the Laplace operator ∆ <strong>in</strong>variant, that is<br />

for all f ∈ C 2 , if and only if Φ is an isometry.<br />

∆f = (∆(f ◦ Φ)) ◦ Φ −1<br />

The proof of Theorem 1.2.1 follows by tak<strong>in</strong>g <strong>in</strong>to account the representation <strong>in</strong> (1.23) and by<br />

observ<strong>in</strong>g that, for p ∈ M, we have gij(p) = gij(Φ(p)) if and only if Φ is an isometry. For a proof<br />

see Helgason (19) Proposition 2.4, Chapter 2.<br />

1.3 Sectional curvature<br />

Def<strong>in</strong>ition 1.3.1. Let X, Y and Z be three differentiable vector fields on a Riemannian manifold<br />

M. The curvature R is a correspondence that associates to X, Y and Z the differentiable vector<br />

field R(X, Y )Z given by<br />

where ▽ is the Levi-Civita connection.<br />

R(X, Y )Z = ▽Y ▽XZ − ▽X▽Y Z + ▽ [X,Y ]Z,


9 Tools of Riemannian Geometry<br />

S<strong>in</strong>ce, if M = R n then R(X, Y )Z = 0 for all X, Y , Z, we are able to th<strong>in</strong>k of R as a way of<br />

measur<strong>in</strong>g how mach M deviates from be<strong>in</strong>g Euclidean. In view of (1.4), (1.5) and (1.6), we obta<strong>in</strong><br />

that<br />

R(fX1 + gX2, Y )Z = fR(X1, Y )Z + gR(X2, Y )Z, (1.24)<br />

R(X, fY1 + gY2)Z = fR(X, Y1)Z + gR(X, Y2)Z, (1.25)<br />

R(X, Y )(fZ1 + gZ2) = fR(X, Y )Z1 + gR(X, Y )Z2. (1.26)<br />

Given a coord<strong>in</strong>ate system (U, ϕ) around a po<strong>in</strong>t p ∈ M and sett<strong>in</strong>g<br />

X =<br />

n<br />

ξ i ∂i, Y =<br />

i=1<br />

from (1.24), (1.25) and (1.26) we have<br />

n<br />

η j ∂j, Z =<br />

j=1<br />

R(X, Y )Z = <br />

ξ i R(∂i, Y )Z = <br />

i<br />

= <br />

i,j,k<br />

i,j<br />

ξ i η j ζ k R(∂i, ∂j)∂k.<br />

n<br />

ζ k ∂k,<br />

k=1<br />

ξ i η j R(∂i, ∂j)Z<br />

S<strong>in</strong>ce, <strong>in</strong> view of (1.7) and (1.11) we have [∂i, ∂j] = 0, apply<strong>in</strong>g Def<strong>in</strong>ition 1.3.1 we obta<strong>in</strong> that<br />

R(∂i, ∂j)∂k = ▽∂j ▽∂i∂k − ▽∂i▽∂j ∂k = ▽∂j ( <br />

Γ l ik∂l) − ▽∂i( <br />

and, if we put<br />

we have that<br />

l<br />

l<br />

Γ l jk∂l)<br />

= <br />

(Γ l ik▽∂j ∂l + (∂jΓ l ik)∂l) − <br />

(Γ l jk▽∂i∂l + (∂iΓ l jk)∂l)<br />

l<br />

= <br />

l<br />

Γ l ik<br />

<br />

Γ s jl∂s − <br />

s<br />

l<br />

Γ l jk<br />

R(∂i, ∂j)∂k = <br />

R l ijk = <br />

Γ s ikΓ l js − <br />

s<br />

s<br />

l<br />

<br />

Γ s il∂s + <br />

(∂jΓ l ik)∂l − <br />

(∂iΓ l jk)∂l<br />

l<br />

s<br />

R l ijk∂l,<br />

l<br />

Γ s jkΓ l is + ∂jΓ l ik − ∂iΓ l jk. (1.27)<br />

Theorem 1.3.1. Let σ ∈ TpM be a two-dimensional subspace of the tangent space TpM and let<br />

x, y ∈ σ be two l<strong>in</strong>early <strong>in</strong>dependent vectors, then<br />

K(x, y) =<br />

does not depend on the choice of the vectors x, y ∈ σ.<br />

〈R(x, y)x, y〉<br />

|x| 2 |y| 2 − 〈x, y〉 2<br />

For a proof of Theorem 1.3.1 see Do Carmo (14) Proposition 3.1. We can give now the follow<strong>in</strong>g<br />

def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 1.3.2. Given a po<strong>in</strong>t p ∈ M and a two-dimensional subspace σ ∈ TpM where {x, y} is<br />

any basis of σ, the real number K(σ) = K(x, y) is called sectional curvature of the two dimensional<br />

section σ at p.<br />

The importance of the sectional curvature comes from the fact that the knowledge of K(σ),<br />

l


1.4 <strong>Hyperbolic</strong> space 10<br />

for all σ ∈ TpM, determ<strong>in</strong>es the curvature R completely. Among the Riemannian manifolds those<br />

with constant sectional curvature are the most simple and play a fundamental role.<br />

1.4 <strong>Hyperbolic</strong> space<br />

For a Riemmanian manifold M and a chart x : U → Rn , we write the Riemannian metric <strong>in</strong> the<br />

chart as<br />

ds 2 =<br />

n<br />

gjk(x)dx j dx k . (1.28)<br />

j,k=1<br />

If M is the Euclidean space with the usual <strong>in</strong>ner product and x is the identity map, formula (1.28)<br />

gives the usual Euclidean distance element<br />

Consider the half-space of R n given by<br />

ds 2 =<br />

n<br />

(dx j ) 2 = |dx| 2 .<br />

j=1<br />

H n = {(x 1 , . . . , x n ) ∈ R n : x n > 0},<br />

with n ≥ 2, endowed with the Riemannian metric given by<br />

We have that<br />

gij = δij<br />

(x n ) 2 , gij = (x n ) 2 δij, g =<br />

ds 2 =<br />

n<br />

j=1<br />

(dx j ) 2<br />

(x n )<br />

1<br />

(xn .<br />

) 2n<br />

|dx|2<br />

= 2 (xn . (1.29)<br />

) 2<br />

H n is a model for the n-dimensional hyperbolic space. We start by prov<strong>in</strong>g that the hyperbolic<br />

space has constant sectional curvature equal to −1.<br />

We put fi = ∂i(log x n ) and<br />

Rijkm = 〈R(∂i, ∂j)∂k, ∂m〉 = 〈 <br />

R l ijk∂l, ∂m〉 = <br />

We have to prove that for all i, j = 1, . . . , n,<br />

K(∂i, ∂j) =<br />

l<br />

Rijij<br />

|∂2 i ||∂j| 2 = −1.<br />

− 〈∂i, ∂j〉 2<br />

l<br />

R l ijkglm.


11 Tools of Riemannian Geometry<br />

To calculate Rijij, <strong>in</strong> view of (1.27), we have that<br />

Rijij = <br />

R l iji glj = <br />

=<br />

=<br />

l<br />

1<br />

(x n ) 2<br />

1<br />

(x n ) 2<br />

<br />

⎡<br />

s<br />

⎣ <br />

l<br />

Γ s iiΓ j<br />

js<br />

R l iji<br />

− <br />

δlj<br />

(xn Rj iji<br />

=<br />

) 2 (xn ) 2<br />

s<br />

Γ s jiΓ j j j<br />

is + ∂jΓii − ∂iΓji (Γ<br />

s=i,j<br />

s iiΓ j<br />

js − ΓsjiΓ j<br />

is ) + ΓjiiΓjjj<br />

− Γjji<br />

Γjij<br />

+ ΓiiiΓ j<br />

ji − ΓijiΓ j j j<br />

ii + ∂jΓii − ∂iΓji where, <strong>in</strong> view of (1.15), the Christoffel symbols for the hyperbolic metric are given by<br />

Γ k ij = 1<br />

2<br />

<br />

g kl {∂iglj + ∂jgil − ∂lgij}<br />

l<br />

= 1 <br />

(x<br />

2<br />

l<br />

n ) 2 <br />

δkl ∂i<br />

(xn )<br />

= 1<br />

2 (xn ) 2<br />

<br />

δkj<br />

∂i<br />

(xn δik<br />

+ ∂j<br />

) 2 (xn )<br />

= −δkjfi − δikfj + δijfk.<br />

δlj<br />

+ ∂j 2<br />

<br />

δil δij<br />

− ∂l 2 (xn ) 2<br />

(x n )<br />

δij<br />

− ∂k 2 (xn ) 2<br />

The last formula implies that Γk ij = 0 if i = j = k, Γiij = −fj, Γ j<br />

ij = −fi, Γk ii = fk and Γi ii = −fi<br />

and thus<br />

Rijij =<br />

1<br />

(xn ) 2<br />

<br />

− <br />

s<br />

<br />

(fs) 2 + (fi) 2 + (fj) 2 + fii + fjj<br />

F<strong>in</strong>ally, observ<strong>in</strong>g that |∂i| 2 |∂j| 2 − 〈∂i, ∂j〉 2 = gii gjj, we have<br />

K(∂i, ∂j) = Rijij<br />

gii gjj<br />

= (x n ) 4 Rijij = (x n ) 2<br />

<br />

− <br />

s<br />

<br />

(fs) 2 + (fi) 2 + (fj) 2 + fii + fjj<br />

and s<strong>in</strong>ce fn = 1<br />

x n and fi = 0 for i = n if follows that K(∂i, ∂j) = −1 for all i, j = 1, . . . , n.<br />

In view of (1.23), the Laplace operator ∆ <strong>in</strong> the hyperbolic space is given by<br />

∆ = (x n ) n<br />

= (x n ) n<br />

= (x n ) n<br />

= (x n ) 2<br />

n ∂<br />

∂x<br />

j=1<br />

j<br />

<br />

(x n ) 2 (x n −n ∂<br />

)<br />

∂xj <br />

⎡<br />

n−1 <br />

⎣ (x<br />

j=1<br />

n ) 2−n ∂2<br />

∂(xj ) 2 + (2 − n)(xn )<br />

⎡<br />

n<br />

⎣ (x n ) 2−n ∂2<br />

∂(xj ) 2 + (2 − n)(xn )<br />

j=1<br />

n<br />

j=1<br />

= (x n ) n<br />

n<br />

j=1<br />

2−n−1 ∂<br />

1−n ∂<br />

<br />

∂<br />

∂xj <br />

(x n 2−n ∂<br />

)<br />

∂xj <br />

∂xn + (xn ) 2−n<br />

∂(xn ) 2<br />

⎦<br />

⎤<br />

∂xn ⎦<br />

∂2 ∂(xj ∂<br />

+ (2 − n)xn . (1.30)<br />

) 2 ∂xn There are several models for the hyperbolic space, the ball model, for example, consists of the<br />

.<br />

∂ 2<br />

⎤<br />

<br />

⎤<br />

⎦ ,


1.4 <strong>Hyperbolic</strong> space 12<br />

open unit ball B n = {x ∈ R n : |x| < 1} endowed with the Riemannian metric<br />

ds 2 = 4|dx|2<br />

(1 − |x| 2 .<br />

) 2<br />

We will see <strong>in</strong> the next chapter that the fractional transformation z → z−i<br />

z+i<br />

maps isometrically H2 onto B2 .<br />

of the complex plane


Chapter 2<br />

<strong>Hyperbolic</strong> Geometry<br />

2.1 Möbius transformations<br />

The reflection (or <strong>in</strong>version) <strong>in</strong> the sphere S(a, r) = {x ∈ R n : |x − a| = r} where a ∈ R n and<br />

r > 0, is a function φ(·) such that, for a, x ∈ R n , x = a<br />

φ(x) = a +<br />

r2<br />

(x − a), (2.1)<br />

|x − a| 2<br />

it is natural to def<strong>in</strong>e φ(a) = ∞ and φ(∞) = a. We observe that φ(x) = x if and only if x ∈ S(a, r).<br />

We have that φ ◦ φ(x) = x and so φ(·) is a bijection of R n onto itself.<br />

A plane P (a, t) <strong>in</strong> R n is a set of the form<br />

P (a, t) = {x ∈ R n : 〈x, a〉 = t} ∪ {∞}<br />

where a ∈ R n , a = 0, t ∈ R and 〈x, a〉 = aixi. The reflection φ(·) <strong>in</strong> the plane P (a, t) is def<strong>in</strong>ed,<br />

for every x ∈ R n , by<br />

φ(x) = x + λa<br />

and φ(∞) = ∞. The parameter λ is chosen so that 1<br />

2 (x + φ(x)) ∈ P (a, t), this gives the explicit<br />

formula<br />

φ(x) = x − 2a<br />

(〈x, a〉 − t) .<br />

|a| 2 (2.2)<br />

In fact 1<br />

λa<br />

λ<br />

2 (x + φ(x)) ∈ P (a, t) if and only if 〈x + 2 , a〉 = t that is 〈x, a〉 + 2 |a|2 =t. Aga<strong>in</strong> φ(x) = x<br />

if and only if x ∈ P (a, t), we have φ ◦ φ(x) = x and so φ(·) is a bijection of R n onto itself.<br />

It is possible to prove that the reflection <strong>in</strong> a sphere and the reflection <strong>in</strong> a plane are two cont<strong>in</strong>uous<br />

functions <strong>in</strong> R n onto itself with respect to the chordal metric d(x, y) def<strong>in</strong>ed by<br />

⎧<br />

⎨<br />

d(x, y) =<br />

⎩<br />

2|x−y|<br />

(1+|x| 2 ) 1/2 (1+|y| 2 ) 1/2 , if x, y = ∞,<br />

2<br />

(1+|x| 2 ) 1/2 , if y = ∞.


2.1 Möbius transformations 14<br />

The chordal metric <strong>in</strong> R n <strong>in</strong>duces the same topology as does the Euclidean metric, this implies<br />

that a function is cont<strong>in</strong>uous with respect to both or to neither of these metrics.<br />

We have observed that for any reflection φ(·) we have that φ(x) = φ −1 (x), this implies that φ −1 (x)<br />

is cont<strong>in</strong>uous on R n with respect to the chordal metric too and then each reflection φ(·) is a<br />

homeomorphism of R n onto itself with respect to the chordal metric.<br />

Remark 2.1.1. If φ(·) is the reflection <strong>in</strong> the sphere S(a, r), we have<br />

|φ(y) − φ(x)| = r 2<br />

<br />

<br />

<br />

y − a x − a<br />

−<br />

|y − a| 2 |x − a| 2<br />

<br />

<br />

<br />

1 2〈x − a, y − a〉<br />

= r2 −<br />

|x − a| 2 |x − a| 2 1<br />

+<br />

|y − a| 2 |y − a| 2<br />

1/2 =<br />

r2 2 2<br />

|y − a| − 2〈x − a, y − a〉 + |x − a|<br />

|x − a||y − a|<br />

1/2 This shows that<br />

=<br />

r2 |x − y|<br />

. (2.3)<br />

|x − a||y − a|<br />

|φ(x + h) − φ(x)| r<br />

lim<br />

=<br />

h→0 |h|<br />

2<br />

. (2.4)<br />

|x − a| 2<br />

Remark 2.1.2. Any reflection <strong>in</strong> a plane is a Euclidean isometry, <strong>in</strong> fact<br />

|φ(x) − φ(y)| 2 <br />

2<br />

<br />

= <br />

2a<br />

2a<br />

<br />

x − (〈x, a〉 − t) − y + (〈y, a〉 − t) <br />

|a| 2 |a| 2 <br />

<br />

2<br />

<br />

= <br />

2a <br />

x − y − 〈x − y, a〉 <br />

|a| 2 <br />

= |x − y| 2 <br />

+ 2 − 2<br />

<br />

〈x − y, a〉 〈x − y, a〉 +<br />

|a| 2 4<br />

|a| 4 〈x − y, a〉2 |a| 2<br />

= |x − y| 2 . (2.5)<br />

Def<strong>in</strong>ition 2.1.1. A Möbius transformation act<strong>in</strong>g <strong>in</strong> R n is a f<strong>in</strong>ite composition of reflections <strong>in</strong><br />

spheres and planes.<br />

It follows that each Möbius transformation is a homeomorphism of R n onto itself. The composition<br />

of two Möbius transformations is aga<strong>in</strong> a Möbius transformation and so is the <strong>in</strong>verse of a<br />

Möbius transformation. In fact if φ = φ1 ◦ · · · ◦ φm, where the φj, j =, 1 . . . , m, are reflections,<br />

then φ −1 = φm ◦ · · · ◦ φ1. F<strong>in</strong>ally, s<strong>in</strong>ce for any Möbius transformation we have φ ◦ φ −1 (x) = x,<br />

the identity map is a Möbius transformation. We can give the follow<strong>in</strong>g def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 2.1.2. The group of Möbius transformations act<strong>in</strong>g <strong>in</strong> R n is called General Möbius<br />

Group and is denoted by GM(R n ).<br />

Example 2.1.1. The translation φ(x) = x + a, a, x ∈ Rn , is a Möbius transformation s<strong>in</strong>ce it<br />

is the reflection <strong>in</strong> the plane 〈x, a〉 = 0 followed by the reflection <strong>in</strong> the plane 〈x, a〉 = |a|2<br />

2 . The<br />

magnification φ(x) = kx, k > 0 is a Möbius transformation s<strong>in</strong>ce it is the reflection <strong>in</strong> the sphere


15 <strong>Hyperbolic</strong> Geometry<br />

S(0, 1) followed by the reflection <strong>in</strong> the sphere S(0, √ k). The transformations of the form<br />

φ(x) =<br />

ax + b<br />

cx + d<br />

where a, b, c, d are real and such that ad − bc = 0 are Möbius transformations. In fact φ(x) =<br />

φ4 ◦ φ3 ◦ φ2 ◦ φ1(x) where φ1(x) = x + d<br />

c is a translation, φ2(x) = 1<br />

x is an <strong>in</strong>version and a reflection<br />

with respect to the real axis, φ3(x) = − ad−bc<br />

c 2<br />

is a translation.<br />

(2.6)<br />

x is a magnification and a rotation and φ4(x) = x+ a<br />

c<br />

In what follows we will denote with Σ either a sphere S(a, r) or a plane P (a, t) <strong>in</strong> R n .<br />

Theorem 2.1.1. For every Möbius transformation φ(·) and every sphere or plane Σ we have that<br />

φ(Σ) is also a sphere or a plane.<br />

The <strong>in</strong>versive product (Σ, Σ ′ ), given <strong>in</strong> Def<strong>in</strong>ition 2.1.3, is a real expression which depends only<br />

on Σ and Σ ′ and which is <strong>in</strong>variant under all Möbius transformations. When Σ and Σ ′ <strong>in</strong>tersect<br />

it is a function of their angle of <strong>in</strong>tersection and when Σ and Σ ′ are disjo<strong>in</strong>t it is a function of the<br />

hyperbolic distance between them.<br />

S<strong>in</strong>ce the equation def<strong>in</strong><strong>in</strong>g a sphere S(a, r) is<br />

and the equation def<strong>in</strong><strong>in</strong>g a plane P (a, t) is<br />

|x| 2 − 2〈x, a〉 + |a| 2 − r 2 = 0<br />

−2〈x, a〉 + 2t = 0,<br />

we have that every sphere or plane Σ can be written <strong>in</strong> a common form as<br />

a0|x| 2 − 2〈x, a〉 + an+1 = 0,<br />

where a = (a1, . . . , an) and the coefficient vector (a0, . . . , an+1) is determ<strong>in</strong>ed with<strong>in</strong> a real nonzero<br />

multiple. S<strong>in</strong>ce for Σ = S(a, r) and Σ = P (a, t) we have that |a| 2 > a0an+1, we can give the<br />

follow<strong>in</strong>g def<strong>in</strong>ition.<br />

Def<strong>in</strong>ition 2.1.3. Let Σ and Σ ′ have coefficient vectors (a0, . . . , an+1) and (b0, . . . , bn+1). The<br />

<strong>in</strong>versive product (Σ, Σ ′ ) of Σ and Σ ′ is def<strong>in</strong>ed by<br />

(Σ, Σ ′ ) =<br />

|2〈a, b〉 − a0bn+1 − an+1b0|<br />

2(|a| 2 − a0an+1) 1/2 (|b| 2 .<br />

− b0bn+1)<br />

1/2<br />

Remark 2.1.3. It is helpful to obta<strong>in</strong> the explicit expression of (Σ, Σ ′ ) <strong>in</strong> the follow<strong>in</strong>g three<br />

cases.<br />

• If Σ = S(a, r) and Σ ′ = S(b, t) then (Σ, Σ ′ ) = r2 +t 2 −|a−b| 2<br />

2rt<br />

• If Σ = S(a, r) and Σ ′ = P (b, t) then (Σ, Σ ′ ) = |〈a,b〉−t|<br />

r|b| .<br />

• If Σ = P (a, r) and Σ ′ = P (b, t) then (Σ, Σ ′ ) = |〈a,b〉|<br />

|a||b| .<br />

<br />

<br />

<br />

<br />

.


2.2 Po<strong>in</strong>caré extension of Möbius transformations 16<br />

Theorem 2.1.2. For any Σ and Σ ′ and any Möbius transformation φ(·) we have that<br />

(Σ, Σ ′ ) = (φ(Σ), φ(Σ ′ )).<br />

Remark 2.1.4. It is simple to see that, if Σ and Σ ′ <strong>in</strong>tersect, then (Σ, Σ ′ ) = cos θ where θ is one<br />

of the angles of <strong>in</strong>tersection between Σ and Σ ′ .<br />

For example let Σ = S(a, r) and Σ ′ = S(b, t) be two spheres that meat at some po<strong>in</strong>t c ∈ R n ,<br />

then there exists a translation φ(·) such that φ(c) = 0. It follows that |φ(a)| = r and |φ(b)| = t<br />

and <strong>in</strong> view of Theorem 2.1.2 we have<br />

(Σ, Σ ′ ) = (φ(Σ), φ(Σ ′ <br />

<br />

)) = <br />

r<br />

<br />

2 + t2 − |φ(a) − φ(a)| 2<br />

<br />

<br />

<br />

2rt <br />

= |〈φ(a), φ(b)〉|<br />

|φ(a)||φ(b)|<br />

= cos θ.<br />

Theorem 2.1.3. Let Σ be any sphere or plane, σ the reflection <strong>in</strong> Σ and I the identity map. If<br />

φ(·) is a Möbius transformation that fixes each x ∈ Σ then either φ = I or φ = σ.<br />

2.2 Po<strong>in</strong>caré extension of Möbius transformations<br />

Po<strong>in</strong>caré observed that each Möbius transformation φ(·) act<strong>in</strong>g <strong>in</strong> R n has a natural extension to a<br />

Möbius transformation ˜ φ(·) act<strong>in</strong>g <strong>in</strong> R n+1 . This extension depends on the embedd<strong>in</strong>g<br />

x = (x1, . . . , xn) ↦→ ˜x = (x1, . . . , xn, 0),<br />

of R n <strong>in</strong>to R n+1 . In particular if φ(·) is the reflection <strong>in</strong> S(a, r), a ∈ R n , then ˜ φ(·) is def<strong>in</strong>ed as the<br />

reflection <strong>in</strong> S(ã, r) and <strong>in</strong> view of (2.1) we have that for every x ∈ R n+1<br />

˜φ(x) = ã +<br />

and <strong>in</strong> particular for x ∈ R n we have that<br />

˜φ(˜x) = ˜ φ(x1, . . . , xn, 0) = ã +<br />

r2<br />

<br />

(x − ã) = φ(x),<br />

|x − ã| 2 r2xn+1 |x − ã| 2<br />

<br />

(2.7)<br />

r2<br />

|˜x − ã| 2 (˜x − ã) = (φ(x), 0) = φ(x). (2.8)<br />

If φ(·) is the reflection <strong>in</strong> P (a, t), a ∈ R n , then ˜ φ(·) is def<strong>in</strong>ed as the reflection <strong>in</strong> P (ã, t) and for<br />

every x ∈ R n+1 we have<br />

For x ∈ R n it follows that<br />

˜φ(x) = x − 2ã<br />

|ã| 2 (〈x, ã〉 − t) = (φ(x), xn+1). (2.9)<br />

˜φ(˜x) = ˜ φ(x1, . . . , xn, 0) = ˜x − 2ã<br />

|ã| 2 (〈˜x, ã〉 − t) = (φ(x), 0) = φ(x). (2.10)<br />

S<strong>in</strong>ce for every x ∈ R n and every reflection φ(·) we have that ˜ φ(˜x) = φ(x), we can see ˜ φ(·) as an<br />

extension of φ(·), such extension is called Po<strong>in</strong>caré extension of φ(·).<br />

Remark 2.2.1. In view of (2.7), (2.8), (2.9) and (2.10) we observe that each Po<strong>in</strong>caré extension<br />

˜φ(·) leaves <strong>in</strong>variant the plane xn+1 = 0 and the half-spaces xn+1 > 0 and xn+1 < 0.


17 <strong>Hyperbolic</strong> Geometry<br />

S<strong>in</strong>ce each Möbius transformation φ = φ1 ◦ · · · ◦ φm <strong>in</strong> R n is def<strong>in</strong>ed as the composition of a<br />

f<strong>in</strong>ite number of reflections φj, j = 1, . . . , m, there is at last one Möbius transformation ˜ φ =<br />

˜φ1 ◦ · · · ◦ ˜ φm which extends the action of φ(·) to R n+1 <strong>in</strong> the sense that ˜ φ(x, 0) = φ(x). This<br />

Möbius transformation preserves the plane xn+1 = 0 and the half-spaces xn+1 > 0 and xn+1 < 0.<br />

In fact there is at most one extension ˜ φ(·) of a Möbius transformation φ(·). If ˜ φ1 and ˜ φ2 are two<br />

such extensions, then ˜ φ −1<br />

2 ◦ ˜ φ1 fixes each po<strong>in</strong>t of the plane xn+1 = 0, then by Theorem 2.1.3 we<br />

have that either ˜ φ −1<br />

2 ◦ ˜ φ1 = I or ˜ φ −1<br />

2 ◦ ˜ φ1 = σ where σ is the reflection <strong>in</strong> the plane xn+1 = 0. But<br />

s<strong>in</strong>ce ˜ φ −1<br />

2 ◦ ˜ φ1 preserves the half-spaces xn+1 > 0 and xn+1 < 0 we must have that ˜ φ −1<br />

2 ◦ ˜ φ1 = I<br />

and then ˜ φ1 = ˜ φ2.<br />

2.3 <strong>Hyperbolic</strong> metric<br />

We now focus on the action of the Po<strong>in</strong>caré extension ˜ φ(·) of a Möbius transformation <strong>in</strong> the upper<br />

half-plane<br />

H n+1 = {x ∈ R n+1 : xn+1 > 0}.<br />

If ˜ φ(·) is the reflection <strong>in</strong> the sphere S(ã, r), a ∈ R n , then by (2.3) for every y, x ∈ H n+1 we have<br />

and from (2.7) it follows that<br />

| ˜ φ(y) − ˜ φ(x)|<br />

|y − x|<br />

=<br />

r2 , (2.11)<br />

|x − ã||y − ã|<br />

[ ˜ φ(x)]n+1 = r2xn+1 , (2.12)<br />

|x − ã| 2<br />

where [ ˜ φ(x)]n+1 denotes the (n + 1)-th component of ˜ φ(x). Formula (2.11) and formula (2.12)<br />

imply that<br />

and this shows that<br />

is <strong>in</strong>variant under ˜ φ(·).<br />

| ˜ φ(y) − ˜ φ(x)| 2<br />

[ ˜ φ(y)]n+1[ ˜ =<br />

φ(x)]n+1<br />

r 4 |y−x| 2<br />

|x−ã| 2 |y−ã| 2<br />

r 2 xn+1<br />

|x−ã| 2<br />

|y − x| 2<br />

xn+1yn+1<br />

r 2 yn+1<br />

|y−ã| 2<br />

= |y − x|2<br />

xn+1yn+1<br />

(2.13)<br />

Let ˜ φ(·) be the reflection <strong>in</strong> the plane P (ã, t), a ∈ R n , s<strong>in</strong>ce <strong>in</strong> view of (2.5) every reflection <strong>in</strong> a<br />

plane is a Euclidean isometry, we have that | ˜ φ(y) − ˜ φ(x)| = |y − x| and from (2.9) we have that<br />

[ ˜ φ(x)]n+1 = xn+1, this implies that<br />

| ˜ φ(y) − ˜ φ(x)| 2<br />

[ ˜ φ(y)]n+1[ ˜ =<br />

φ(x)]n+1<br />

|y − x|2<br />

xn+1yn+1<br />

Thus (2.13) is <strong>in</strong>variant under all Po<strong>in</strong>caré extensions of any φ(·) <strong>in</strong> GM(R n ) and this implies that<br />

such Po<strong>in</strong>caré extensions are isometry of H n+1 endowed with the Riemannian metric derived from<br />

the differential<br />

.<br />

ds = |dx|<br />

. (2.14)<br />

xn+1


2.3 <strong>Hyperbolic</strong> metric 18<br />

Given a positive cont<strong>in</strong>uous function λ(x) on D ∈ Rn , let ρ(x, y) be the function def<strong>in</strong>ed for<br />

x, y ∈ D by<br />

<br />

ρ(x, y) = <strong>in</strong>f λ(γ(t)) |γ<br />

γ<br />

′ (t)|dt,<br />

where the <strong>in</strong>fimum is taken over all curves γ(t) <strong>in</strong> D with derivative γ ′ (t), jo<strong>in</strong><strong>in</strong>g x to y <strong>in</strong> D.<br />

The function ρ(x, y) is a metric <strong>in</strong> D s<strong>in</strong>ce it is symmetric, non negative and satisfies the triangle<br />

<strong>in</strong>equality. In view of (2.14) the function η(x, y), def<strong>in</strong>ed by<br />

b<br />

η(x, y) = <strong>in</strong>f<br />

a<br />

|γ ′ (t)|<br />

dt,<br />

[γ(t)]n+1<br />

where γ : [a, b] → H n+1 is a differentiable curve that jo<strong>in</strong>s x to y <strong>in</strong> H n+1 , is the hyperbolic metric<br />

on H n+1 . The couple (H n+1 , η) is a model of the hyperbolic space and<br />

is the hyperbolic length of the curve γ.<br />

||γ|| =<br />

b<br />

a<br />

|γ ′ (t)|<br />

dt<br />

[γ(t)]n+1<br />

We shall now map isometrically H n+1 onto B n+1 = {x ∈ R n+1 : |x| < 1} to obta<strong>in</strong> a second model<br />

of the hyperbolic space.<br />

Let φ0 denote the reflection <strong>in</strong> the sphere S(en+1, √ 2), for every x ∈ R n+1<br />

then<br />

|φ0(x)| 2 =<br />

<br />

en+1 +<br />

= |en+1| 2 +<br />

= 1 +<br />

φ0(x) = en+1 +<br />

2(x − en+1)<br />

|x − en+1| 2 , en+1 +<br />

2(x − en+1)<br />

,<br />

|x − en+1|<br />

2<br />

2(x − en+1)<br />

|x − en+1| 2<br />

<br />

4<br />

|x − en+1| 2 〈en+1,<br />

4<br />

2<br />

x − en+1〉 +<br />

|x − en+1|<br />

|x − en+1|<br />

4<br />

4<br />

|x − en+1| 2 (〈en+1,<br />

4<br />

x〉 − 〈en+1, en+1〉) +<br />

|x − en+1| 2<br />

= 1 + 4〈en+1, x〉 4xn+1<br />

= 1 + . (2.15)<br />

|x − en+1|<br />

2 |x − en+1|<br />

2<br />

This shows that φ0 maps the lower half-space xn+1 < 0 onto B n+1 . Now let σ be the reflection <strong>in</strong><br />

the plane xn+1 = 0 and φ = φ0 ◦ σ. S<strong>in</strong>ce for x ∈ R n+1 we have<br />

σ(x) = x − 2xn+1en+1 = (x1, . . . , xn, −xn+1),<br />

we obta<strong>in</strong> that φ(·) maps the half-space H n+1 onto B n+1 . Observ<strong>in</strong>g that the reflection <strong>in</strong> a plane<br />

is a Euclidean isometry and <strong>in</strong> view of (2.4), we f<strong>in</strong>d that<br />

|φ(y) − φ(x)|<br />

lim<br />

y→x |y − x|<br />

|φ0(σ(y)) − φ0(σ(x))| |φ0(σ(y)) − φ0(σ(x))|<br />

= lim<br />

= lim<br />

y→x |y − x|<br />

y→x |σ(y) − σ(x)|<br />

2<br />

=<br />

.<br />

|σ(x) − en+1|<br />

2


19 <strong>Hyperbolic</strong> Geometry<br />

S<strong>in</strong>ce, <strong>in</strong> view of (2.15), we have that<br />

it follows that<br />

1 − |φ(x)| 2 = 1 − |φ0(σ(x))| 2 = − 4[σ(x)]n+1 4xn+1<br />

=<br />

,<br />

|σ(x) − en+1|<br />

2 |σ(x) − en+1|<br />

2<br />

|φ(y) − φ(x)|<br />

lim<br />

=<br />

y→x |y − x|<br />

1 − |φ(x)|2<br />

2xn+1<br />

. (2.16)<br />

In general, given a positive cont<strong>in</strong>uous function λ(x) on D ∈ Rn , a metric ρ(x, y) = <strong>in</strong>f <br />

γ λ(γ(t)) |γ′ (t)|dt<br />

and a bijection f of D onto D1 ∈ Rn such that<br />

|f(y) − f(x)|<br />

lim<br />

= µ(x),<br />

y→x |y − x|<br />

it is possible to def<strong>in</strong>e a positive a cont<strong>in</strong>uous function λ1(x) = λ(x)/µ(x) on D1 and a metric<br />

ρ1(x, y) = <strong>in</strong>f <br />

γ λ1(γ(t)) |γ ′ (t)|dt on D1. It follows that f is an isometry of (D, ρ) onto (D1, ρ1)<br />

(see e.g. Beardon (2) page 7).<br />

S<strong>in</strong>ce φ0 ◦ σ is a bijection of H n+1 onto B n+1 that satisfies (2.16), tak<strong>in</strong>g <strong>in</strong>to account (2.14), we<br />

def<strong>in</strong>e λ(x) = 1<br />

1−|φ(x)|2<br />

2<br />

, µ(x) = and λ1(x) =<br />

xn+1 2xn+1<br />

1−|φ(x)| 2 . It follows that φ = φ0◦σ is an isometry<br />

between (H n+1 , η) and (B n+1 , η1) where η1 is the metric obta<strong>in</strong>ed from the differential<br />

ds = 2|dx|<br />

. (2.17)<br />

1 − |x| 2<br />

The upper-half plane H 2 = {z = x + iy ∈ C : Im[z] > 0} with the metric η derived from the<br />

differential ds = |dz|<br />

Im[z] and the disk B2 = {z = x+iy ∈ C : |z| < 1} with the metric η1 derived from<br />

the differential ds = 2|du|<br />

1−|u| 2 are two models for the hyperbolic plane. We give now a description of<br />

the hyperbolic metric <strong>in</strong> the upper-half plane H 2 and <strong>in</strong> the disk B 2 .<br />

Now let<br />

g(z) =<br />

az + b<br />

cz + d<br />

(2.18)<br />

be a Möbius transformation of the form (2.6), such that a, b, c, d are real and ad − bc > 0. S<strong>in</strong>ce<br />

for every z = x + iy ∈ H 2 we have<br />

g(z) =<br />

ax + b + iay<br />

cx + d + icy<br />

= (ax + b)(cx + d) + acy2<br />

(cx + d) 2 + c 2 y 2<br />

it follows that g(z) maps H 2 onto itself. Observ<strong>in</strong>g that<br />

y(ad − bc)<br />

+ i<br />

(cx + d) 2 + c2y2 |g ′ <br />

<br />

(z)| = <br />

a(cz + d) − c(az + b)<br />

(cz + d) 2<br />

<br />

<br />

<br />

=<br />

<br />

<br />

<br />

ad − bc<br />

(cz<br />

+ d) 2<br />

<br />

<br />

<br />

=<br />

ad − bc<br />

(cx + d) 2 + c2y2


2.3 <strong>Hyperbolic</strong> metric 20<br />

we have that<br />

|g ′ (z)|<br />

Im[g(z)] =<br />

ad − bc<br />

(cx + d) 2 + c2y2 (cx + d) 2 + c2y2 =<br />

y(ad − bc)<br />

1 1<br />

= . (2.19)<br />

y Im[z]<br />

From (2.19) it follows that, for z, w ∈ H 2 ,<br />

η(g(z), g(w)) = <strong>in</strong>f<br />

′ ′ <br />

|g (γ(t))| |γ (t)|<br />

dt = <strong>in</strong>f<br />

Im[g(γ(t))]<br />

|γ ′ (t)|<br />

dt (2.20)<br />

Im[γ(t)]<br />

= η(z, w) (2.21)<br />

and then each g(z) of the form (2.18) is an isometry of (H 2 , η). The follow<strong>in</strong>g theorem holds.<br />

Theorem 2.3.1. The group of isometries of (H 2 , η) is precisely the group of the maps of the form<br />

g(z) =<br />

where a, b, c, d ∈ R and ad − bc > 0.<br />

az + b<br />

a(−¯z) + b<br />

, q(z) =<br />

cz + d c(−¯z) + d ,<br />

This result will be used <strong>in</strong> the next theorem to obta<strong>in</strong> an explicit expression for the hyperbolic<br />

metric η.<br />

Theorem 2.3.2. For every z, w ∈ H 2 we have that<br />

Proof<br />

We have seen <strong>in</strong> (2.20) that<br />

and from the <strong>in</strong>variance of (2.13) we have<br />

it follows that is sufficient to prove that<br />

cosh η(z, w) = 1 +<br />

cosh η(z, w) = cosh η(g(z), g(w))<br />

|z − w| 2 |g(z) − g(w)|2<br />

=<br />

Im[z]Im[w] Im[g(z)]Im[g(w)]<br />

cosh η(g(z), g(w)) = 1 +<br />

|z − w|2<br />

. (2.22)<br />

2Im[z]Im[w]<br />

|g(z) − g(w)|2<br />

2Im[g(z)]Im[g(w)]<br />

for some convenient transformation g(z). Let l be the unique Euclidean circle or l<strong>in</strong>e which conta<strong>in</strong>s<br />

z and w and is orthogonal to the real axis, if l is a Euclidean circle its center (x0, 0) is on the real<br />

axis. l meets the real axis at some po<strong>in</strong>t α ∈ R. We choose<br />

g1(z) = − 1<br />

(z − α) + β and g2(z) = z − x0.<br />

These are two η-isometries of the form (2.18). If l is a Euclidean l<strong>in</strong>e we choose g(z) = g1(z) with<br />

β = 0 s<strong>in</strong>ce it maps l onto the imag<strong>in</strong>ary axis. If l is a Euclidean circle we choose β = − 1<br />

2α and<br />

g(z) = g2 ◦ g1(z) s<strong>in</strong>ce it maps l onto the imag<strong>in</strong>ary axis. We can now assume that g(z) = ip and<br />

g(w) = iq for some p and q such that 0 < p < q and we have to prove that<br />

η(ip, iq) = log(q/p)


21 <strong>Hyperbolic</strong> Geometry<br />

s<strong>in</strong>ce<br />

1 +<br />

|ip − iq|2 (p − q)2<br />

= 1 +<br />

2Im[ip]Im[iq] 2pq = p2 + q2 = cosh log(q/p).<br />

2pq<br />

But, if γ(t) = x(t) + iy(t) with t ∈ [0, 1] is any smooth curve jo<strong>in</strong><strong>in</strong>g z to w such that γ(1) = q and<br />

γ(0) = p, we have that<br />

<br />

||γ|| =<br />

1<br />

0<br />

|x ′ (t) + iy ′ 1<br />

(t)|<br />

dt ≥<br />

y(t)<br />

0<br />

In general, if we do not assume that p < q, we have that<br />

y ′ (t)<br />

dt = log(q/p). (2.23)<br />

y(t)<br />

η(ip, iq) = | log(p/q)|. (2.24)<br />

Remark 2.3.1. Formula (2.22) can be easily generalized to the (n + 1)-dimensional case. For<br />

z, w ∈ Hn+1 , we have<br />

|z − w|2<br />

cosh η(z, w) = 1 + . (2.25)<br />

2zn+1wn+1<br />

Remark 2.3.2. From Theorem 2.3.2 we obta<strong>in</strong> that the hyperbolic sphere with center y ∈ H n+1<br />

and hyperbolic radius r is the Euclidean sphere with center (y1, . . . , yn, yn+1 cosh r) and radius<br />

yn+1 s<strong>in</strong>h r. In fact {x ∈ Hn+1 : η(x, y) = r} = {x ∈ Hn+1 : cosh r = 1 + |x−y|2<br />

} this implies<br />

2xn+1yn+1<br />

that a po<strong>in</strong>t x on the hyperbolic sphere satisfies the equations<br />

2xn+1yn+1 cosh r = 2xn+1yn+1 + (x1 − y1) 2 + · · · + (xn+1 − yn+1) 2 ,<br />

2xn+1yn+1 cosh r = (x1 − y1) 2 + · · · + (xn − yn) 2 + x 2 n+1 + y 2 n+1,<br />

(yn+1 s<strong>in</strong>h r) 2 = (x1 − y1) 2 + · · · + (xn − yn) 2 + (xn+1 − yn+1 cosh r) 2 .<br />

Remark 2.3.3. It is possible to prove that the follow<strong>in</strong>g equations are equivalent to formula<br />

(2.22).<br />

s<strong>in</strong>h<br />

cosh<br />

tanh<br />

η(z, w)<br />

2<br />

η(z, w)<br />

2<br />

η(z, w)<br />

2<br />

2 x<br />

Start<strong>in</strong>g from (2.22) and apply<strong>in</strong>g s<strong>in</strong>h 2<br />

=<br />

=<br />

=<br />

|z − w|<br />

,<br />

2(Im[z]Im[w]) 1/2<br />

(2.26)<br />

|z − ¯w|<br />

,<br />

2(Im[z]Im[w]) 1/2<br />

<br />

<br />

<br />

z − w <br />

<br />

z<br />

− ¯w .<br />

(2.27)<br />

(2.28)<br />

1<br />

x<br />

= 2 (cosh x − 1) and cosh2 2<br />

1 = 2 (cosh x + 1) we obta<strong>in</strong><br />

easily (2.26) and (2.27). Formula (2.28) follows from (2.26) and (2.27). It is possible to obta<strong>in</strong> the<br />

follow<strong>in</strong>g explicit expression for the hyperbolic metric η<br />

observ<strong>in</strong>g that<br />

<br />

cosh log<br />

η(z, w) = log<br />

<br />

|z − ¯w| + |z − w|<br />

= 1 +<br />

|z − ¯w| − |z − w|<br />

|z − ¯w| + |z − w|<br />

, (2.29)<br />

|z − ¯w| − |z − w|<br />

2|z − w| 2<br />

|z − ¯w| 2 |z − w|2<br />

= 1 +<br />

− |z − w| 2 2Im[z]Im[w] .


2.3 <strong>Hyperbolic</strong> metric 22<br />

We now consider the disk model (B 2 , η1). The map<br />

is a bijection of H 2 onto B 2 , <strong>in</strong> fact<br />

f(z) =<br />

z − i<br />

z + i<br />

|f(z)| 2 = x2 + (y − 1) 2<br />

x2 4y<br />

= 1 −<br />

+ (y + 1) 2 x2 + y2 + 1 + 2y .<br />

(2.30)<br />

Thus the metric ρ ∗ (u, v) = η(f −1 (u), f −1 (v)), u, v ∈ B 2 , is a metric on B 2 . However, s<strong>in</strong>ce we have<br />

that<br />

2|f ′ (z)|<br />

=<br />

1 − |f(z)| 2<br />

2<br />

1 − 1 +<br />

2<br />

x2 +(y+1) 2<br />

4y<br />

x2 +y2 +1+2y<br />

= 1 1<br />

=<br />

y Im[z] ,<br />

we can identify ρ ∗ with the metric η1 derived from the differential (2.17), then f is an isometry<br />

between (H 2 , η) and (B 2 , η1). We can derive an explicit expression for the hyperbolic metric η1<br />

<strong>in</strong> B 2 by simply rewrit<strong>in</strong>g Theorem 2.3.2 by means of f. In particular for u = f(z) ∈ B 2 and<br />

v = f(w) ∈ B 2 we have that<br />

s<strong>in</strong>h η1(u, v)<br />

2<br />

=<br />

|u − v|<br />

(1 − |u| 2 ) 1/2 (1 − |v| 2 ) 1/2<br />

<strong>in</strong> fact, <strong>in</strong> view of (2.26), we have, for u = a + ib and v = α + iβ, that<br />

s<strong>in</strong>h 2 η1(u, v)<br />

2<br />

= s<strong>in</strong>h 2 η(f −1 (z), f −1 (w))<br />

=<br />

=<br />

=<br />

<br />

<br />

<br />

4Im<br />

−i u+1<br />

u−1<br />

−i u+1<br />

u−1<br />

2<br />

<br />

v+1 <br />

+ i v−1<br />

<br />

Im<br />

2<br />

−i v+1<br />

v−1<br />

|u−v| 2<br />

[(a−1) 2 +b2 ][(α−1) 2 +β2 ]<br />

1−a2−b2 (a−1) 2 +b2 1−α2−β2 (α−1) 2 +β2 |u − v| 2<br />

(1 − |u| 2 )(1 − |v| 2 ) .<br />

Start<strong>in</strong>g from formula (2.31) we obta<strong>in</strong> that<br />

<strong>in</strong> fact<br />

cosh 2 η1(u, v)<br />

2<br />

cosh η1(u, v)<br />

2<br />

= 1 + s<strong>in</strong>h 2 η1(u, v)<br />

2<br />

=<br />

|1 − u¯v| 2<br />

(1 − |u| 2 )(1 − |v| 2 )<br />

=<br />

= 1 +<br />

=<br />

<br />

η<br />

2<br />

= s<strong>in</strong>h<br />

=<br />

|1 − u¯v|<br />

(1 − |u| 2 ) 1/2 (1 − |v| 2 ) 1/2<br />

−i u+1<br />

u−1<br />

<br />

v+1 , −i v−1<br />

2<br />

4|u−v| 2<br />

|(u−1)(v−1)| 2<br />

4 1−a2−b2 (a−1) 2 +b2 1−α2−β2 (α−1) 2 +β2 |u − v| 2<br />

(1 − a 2 − b 2 )(1 − α 2 − β 2 )<br />

(2.31)<br />

(2.32)<br />

|u − v| 2<br />

(1 − |u| 2 )(1 − |v| 2 ) = (1 − |u|2 )(1 − |v| 2 ) + |u − v| 2<br />

(1 − |u| 2 )(1 − |v| 2 )


23 <strong>Hyperbolic</strong> Geometry<br />

s<strong>in</strong>ce<br />

|1 − u¯v| 2 = |1 − ūv| 2 = |ū| 2<br />

2<br />

<br />

<br />

1 <br />

− v<br />

ū = |u| 2<br />

2<br />

<br />

<br />

1 <br />

− v<br />

ū = |u| 2<br />

<br />

1 <br />

<br />

ū<br />

= 1 + |u| 2 |v| 2 − 2|u| 2<br />

<br />

1<br />

, v = 1 + |u|<br />

ū 2 |v| 2 − 2 〈u, v〉<br />

= 1 + |u| 2 |v| 2 − 2 〈u, v〉 ± |u| 2 ± |v| 2<br />

= |u − v| 2 + (|u| 2 − 1)(|v| 2 − 1).<br />

From (2.31) and (2.32) it follows that<br />

and from (2.33) we obta<strong>in</strong> that<br />

s<strong>in</strong>ce<br />

In the particular case u = 0 we have<br />

tanh η1(u, v)<br />

2<br />

η1(u, v) = log<br />

= |u − v|<br />

|1 − u¯v|<br />

|1 − u¯v| + |u − v|<br />

|1 − u¯v| − |u − v|<br />

<br />

1 |1 − u¯v| + |u − v|<br />

tanh log =<br />

2 |1 − u¯v| − |u − v|<br />

|u − v|<br />

|1 − u¯v| .<br />

η1(0, v) =<br />

1 + |v|<br />

1 − |v| .<br />

2<br />

+ |v| 2 <br />

1<br />

− 2 , v<br />

ū <br />

(2.33)<br />

It follows that an hyperbolic circle <strong>in</strong> B 2 with center 0 and radius r is a Euclidean circle with center<br />

0 and radius R = r−1<br />

r+1 , <strong>in</strong> fact<br />

v ∈ B 2 : η1(0, v) = r =<br />

<br />

v ∈ B 2 :<br />

<br />

1 + |v|<br />

= r = v ∈ B<br />

1 − |v| 2 <br />

r − 1<br />

: |v| = . (2.34)<br />

r + 1<br />

In the next theorem is obta<strong>in</strong>ed the length of an hyperbolic circle and the area of an hyperbolic<br />

disk.<br />

Theorem 2.3.3. The area of a hyperbolic disk of hyperbolic radius r is 4π s<strong>in</strong>h 2 (r/2). The length<br />

of a hyperbolic circle of hyperbolic radius r is 2π s<strong>in</strong>h r.<br />

Proof<br />

We use the model B2 . Let C = {v ∈ B2 : η1(0, v) = r} and D = {v ∈ B2 : η1(0, v) ≤ r} be the<br />

circle and disk with center <strong>in</strong> the orig<strong>in</strong> and hyperbolic radius r. In view of (2.31) and (2.34) we<br />

have that<br />

s<strong>in</strong>h 2 r<br />

2 = s<strong>in</strong>h2 η1(0, v)<br />

=<br />

2<br />

|v|2 R2<br />

= .<br />

1 − |v| 2 1 − R2 Then<br />

<br />

h − area(D) = 4<br />

D<br />

= 4π s<strong>in</strong>h 2 r<br />

2 .<br />

1<br />

(1 − |v| 2 ) 2 dv1dv2 = 4<br />

2π<br />

0<br />

dθ<br />

R<br />

0<br />

s<br />

ds<br />

(1 − s2 = 4(2π)<br />

) 2 2(1 − R2 )<br />

R 2


2.4 Geodesics 24<br />

In view of (2.34) we have<br />

then<br />

<br />

<br />

h − length(C) = 2<br />

C<br />

2.4 Geodesics<br />

<br />

1 + |v|<br />

s<strong>in</strong>h r = s<strong>in</strong>h η1(0, v) = s<strong>in</strong>h log =<br />

1 − |v|<br />

2R<br />

1 − R2 2π<br />

|dv|<br />

= 2<br />

1 − |v| 2<br />

0<br />

R<br />

R<br />

dθ = 4π = 2π s<strong>in</strong>h r.<br />

1 − R2 1 − R2 The circle of po<strong>in</strong>ts at <strong>in</strong>f<strong>in</strong>ity is represented by the x-axis and denoted by ∂H 2 <strong>in</strong> the half-plane<br />

model and by ∂B 2 = {u : |u| = 1} <strong>in</strong> the disk model.<br />

Def<strong>in</strong>ition 2.4.1. An hyperbolic l<strong>in</strong>e is the <strong>in</strong>tersection of the half-plane H 2 or the disk B 2 with<br />

a Euclidean circle or a straight l<strong>in</strong>e which is orthogonal to the circle of the po<strong>in</strong>ts at <strong>in</strong>f<strong>in</strong>ity.<br />

The follow<strong>in</strong>g facts follow easily from Def<strong>in</strong>ition 2.4.1.<br />

1. There is a unique hyperbolic l<strong>in</strong>e through two dist<strong>in</strong>ct po<strong>in</strong>ts of the hyperbolic plane.<br />

2. Two dist<strong>in</strong>ct hyperbolic l<strong>in</strong>es <strong>in</strong>tersect <strong>in</strong> at most one po<strong>in</strong>t of the hyperbolic plan.<br />

3. The reflection <strong>in</strong> a hyperbolic l<strong>in</strong>e is a hyperbolic isometry. In fact we have seen that (2.13) is<br />

<strong>in</strong>variant under the action of Po<strong>in</strong>caré extensions of Möbius transformations and <strong>in</strong> particular<br />

under reflections <strong>in</strong> spheres and planes of the form S(ã, r) = {z = x+iy ∈ H 2 : |x−a| 2 +y 2 =<br />

r 2 } and P (ã, t) = {z = x + iy ∈ H 2 : ax = t} for a, t ∈ R and r > 0, but these are the<br />

hyperbolic l<strong>in</strong>es of H 2 .<br />

4. Given two hyperbolic l<strong>in</strong>es l1 and l2 there is a hyperbolic isometry g(·) such that g(l1) = l2.<br />

In fact, as <strong>in</strong> the proof of Theorem 2.3.2, for every hyperbolic l<strong>in</strong>e li it is possible to f<strong>in</strong>d an<br />

isometry gi(·) such that gi(li) maps li onto the imag<strong>in</strong>ary axis of H2 . It follows that it is<br />

sufficient to chose g = g −1<br />

2 ◦ g1.<br />

5. Given any hyperbolic l<strong>in</strong>e l and any po<strong>in</strong>t w ∈ H 2 , there is a unique hyperbolic l<strong>in</strong>e through<br />

w orthogonal to l. Let g(z) be the η-isometry such that g(l) maps l onto the imag<strong>in</strong>ary axis.<br />

We observe that, s<strong>in</strong>ce g(z) is Möbius, it preserves the angles, then {z ∈ H 2 : |z| = |g(w)|} is<br />

the unique hyperbolic l<strong>in</strong>e which conta<strong>in</strong>s g(w) and is orthogonal to the positive imag<strong>in</strong>ary<br />

axis.<br />

Given two dist<strong>in</strong>ct po<strong>in</strong>ts z and w on an hyperbolic l<strong>in</strong>e l, we denote by [z, w] the closed hyperbolic<br />

segment on l.<br />

In the follow<strong>in</strong>g theorem we prove that the hyperbolic l<strong>in</strong>es are geodesics, that is curves of shortest<br />

hyperbolic length.


25 <strong>Hyperbolic</strong> Geometry<br />

Theorem 2.4.1. Let z and w be two po<strong>in</strong>ts <strong>in</strong> H 2 . A curve γ jo<strong>in</strong><strong>in</strong>g z to w satisfies<br />

||γ|| = η(z, w)<br />

if and only if γ is a parametrization of [z, w] as a simple curve.<br />

Proof<br />

Let g(·) be a η-isometry such that g(ip) = z and g(iq) = w. Let γ1(t) = x1(t) + x1(t), t ∈ [0, 1],<br />

be any curve jo<strong>in</strong><strong>in</strong>g ip to iq. In view of formula (2.23) we have that ||γ1|| = η(ip, iq) if and only<br />

if x1(t) = 0 and y ′ 1(t) > 0, that is if and only if γ1 is a parametrization of [ip, iq] as a simple<br />

curve. In fact, if ||γ1|| = η(ip, iq) this implies that x ′ 1(t) = 0 and y ′ 1(t) > 0 that is x1(t) = 0 and<br />

y ′ 1(t) > 0 s<strong>in</strong>ce x1(0) = 0. On the other side, if x1(t) = 0 and y ′ 1(t) > 0, than <strong>in</strong> view of (2.23)<br />

we immediately have ||γ1|| = η(ip, iq). The theorem follows by tak<strong>in</strong>g <strong>in</strong>to account that g(·) is a<br />

η-isometry and then maps hyperbolic l<strong>in</strong>es <strong>in</strong> hyperbolic l<strong>in</strong>es. <br />

Theorem 2.4.2. Let z and w be two dist<strong>in</strong>ct po<strong>in</strong>ts <strong>in</strong> the hyperbolic plane. Then<br />

if and only if ζ ∈ [z, w].<br />

η(z, w) = η(z, ζ) + η(ζ, w)<br />

Proof<br />

If ζ ∈ [z, w] and g(·) is such that g(z) = ip, g(ζ) = iq and g(w) = iv than <strong>in</strong> view of (2.24) we<br />

have that η(z, w) = η(z, ζ) + η(ζ, w). On the other side if ζ /∈ [z, w] and γ is a curve made of the<br />

segments [z, ζ] and [ζ, w], then <strong>in</strong> view of Theorem 2.4.1 we have that ||γ|| > η(z, w). <br />

Def<strong>in</strong>ition 2.4.2. Let l1 and l2 be two dist<strong>in</strong>ct geodesics. If l1 ∩ l1 = ∅, then l1 and l2 are <strong>in</strong>tersect<strong>in</strong>g.<br />

If l1 ∩ l1 = ∅, then l1 and l2 are parallel.<br />

In the hyperbolic geometry for any given hyperbolic l<strong>in</strong>e l and a po<strong>in</strong>t w not on l, there are <strong>in</strong>f<strong>in</strong>itely<br />

many hyperbolic l<strong>in</strong>es through w which are parallel to l.<br />

For each po<strong>in</strong>t ζ and any hyperbolic l<strong>in</strong>e l the distance of ζ from l is def<strong>in</strong>ed by<br />

η(ζ, l) = <strong>in</strong>f η(ζ, w).<br />

w∈l<br />

Theorem 2.4.3. Let l1be the unique geodesic through ζ orthogonal to l and let ζ1 be the po<strong>in</strong>t of<br />

<strong>in</strong>tersection between l1 and l. We have that η(ζ, l) = η(ζ, ζ1).<br />

Proof<br />

We may assume that l is the positive imag<strong>in</strong>ary axis, then l1 = {z ∈ H2 : |ζ| = |z|}. We need to<br />

show that η(ζ, l) = η(ζ, i|ζ|). S<strong>in</strong>ce each po<strong>in</strong>t on l is of the form z = ip, p > 0, from Theorem<br />

2.3.2 we have that<br />

cosh η(ζ, ip) = x2 + y2 + p2 =<br />

2yp<br />

|ζ|<br />

<br />

|ζ| p<br />

+ ≥<br />

2y p |ζ|<br />

|ζ|<br />

y ,


2.5 <strong>Hyperbolic</strong> trigonometry 26<br />

for ζ = x + iy. The equality holds if and only if p = |ζ|. <br />

Def<strong>in</strong>ition 2.4.3. A horocycle is a curve <strong>in</strong> the hyperbolic plane that cuts at right angles all the<br />

parallels pass<strong>in</strong>g through the same po<strong>in</strong>t at <strong>in</strong>f<strong>in</strong>ity. In H 2 and B 2 the horocycles are represented<br />

by Euclidean circles tangent to the circle of po<strong>in</strong>ts at <strong>in</strong>f<strong>in</strong>ity.<br />

2.5 <strong>Hyperbolic</strong> trigonometry<br />

In this section we focus on some fundamental results of the hyperbolic trigonometry. With za, zb<br />

and zc we denote the vertices of an hyperbolic triangle T , the sides opposite these vertices will<br />

have lengths a, b and c respectively and the <strong>in</strong>terior angles at the vertices will be α, β and γ. If,<br />

for example, the vertex za is on the circle of po<strong>in</strong>ts at <strong>in</strong>f<strong>in</strong>ity, then we have α = 0, b = c = ∞. As<br />

isometries preserve lengths and angles, we note that the trigonometric formulae rema<strong>in</strong> <strong>in</strong>variant<br />

under isometries. In what follows we will use (H 2 , η) as model of the hyperbolic plane.<br />

Theorem 2.5.1. Let T be a hyperbolic triangle with vertices za, zb, zc and let l be the geodesic<br />

conta<strong>in</strong><strong>in</strong>g the longest side, say [zb, zc], of T . Then the geodesic l1 through za and orthogonal to l<br />

meets l at a po<strong>in</strong>t w ∈ [zb, zc].<br />

Proof<br />

We may assume that l is the positive imag<strong>in</strong>ary axis, it follows that w = i|za|. In view of Theorem<br />

2.4.3 we have that η(za, zb) ≥ η(w, zb) and that η(za, zc) ≥ η(w, zc). As [zb, zc] is the longest side<br />

of T we deduce that<br />

max{η(za, w), η(zc, w)} ≤ max{η(za, zc), η(za, zb)} ≤ η(zb, zc).<br />

S<strong>in</strong>ce the po<strong>in</strong>ts zb, zc and w are coll<strong>in</strong>ear, from Theorem 2.4.2, we have that w must lie between<br />

zb and zc or be equal to one of them. <br />

Theorem 2.5.2. Let T be a right triangle with angles α = 0, β = 0 and γ = π/2. Then<br />

cosh b s<strong>in</strong> α = 1, (2.35)<br />

s<strong>in</strong>h b tan α = 1, (2.36)<br />

tanh b sec α = 1. (2.37)<br />

Proof<br />

In H 2 we may assume that za = x + iy with x 2 + y 2 = 1 and zc = i. S<strong>in</strong>ce y = s<strong>in</strong> α, from Theorem<br />

2.3.2 we obta<strong>in</strong> (2.35), <strong>in</strong> fact<br />

cosh b = cosh η(i, x + iy) = 1 +<br />

|i − x − iy|2<br />

2y<br />

= x2 + y 2 + 1<br />

2y<br />

= 1 1<br />

=<br />

y s<strong>in</strong> α .<br />

Formulae (2.36) and (2.37) follow easily by apply<strong>in</strong>g s<strong>in</strong>h 2 x = cosh 2 x − 1 and tanh x =<br />

s<strong>in</strong>h x<br />

cosh x .


27 <strong>Hyperbolic</strong> Geometry<br />

Remark 2.5.1. The famous Lobachevsky’s formula<br />

cot α<br />

2<br />

follows easily from Theorem 2.5.2, <strong>in</strong> fact, <strong>in</strong> view of (2.37), we have tanh b = cos α, then<br />

= eb<br />

cot α<br />

2 =<br />

<br />

1 + cos α<br />

1 − cos α =<br />

<br />

eb e−b = eb .<br />

Theorem 2.5.3 (<strong>Hyperbolic</strong> Pythagorean Theorem). For any triangle with angles α, β and γ =<br />

π/2, we have<br />

cosh c = cosh a cosh b.<br />

Proof<br />

For any right triangle with angles α, β and π/2, we may assume, by apply<strong>in</strong>g a suitable isometry,<br />

that za = x + iy with x 2 + y 2 = 1, zb = it for t > 0 and zc = i. Us<strong>in</strong>g Theorem 2.3.2 we have<br />

<br />

cosh c = cosh η(it, x + iy) = 1 +<br />

cosh b = cosh η(i, x + iy) = 1 +<br />

cosh a = cosh η(i, it) =<br />

1 + t2<br />

.<br />

2t<br />

|it − x − iy|2<br />

2yt<br />

|i − x − iy|2<br />

2y<br />

= 1 + t2<br />

= 1<br />

y ,<br />

2yt ,


2.5 <strong>Hyperbolic</strong> trigonometry 28


Chapter 3<br />

<strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

3.1 Sett<strong>in</strong>gs<br />

The n-dimensional hyperbolic <strong>Brownian</strong> motion, n ≥ 2, is a diffusion process on the upper halfspace<br />

H n = {z = (x, y) = (x1, · · · , xn−1, y) : x ∈ R n−1 , y > 0}<br />

associated with the Laplace operator ∆n given by (see formula (1.30))<br />

∆n = y 2<br />

n−1<br />

<br />

∂ 2<br />

∂x<br />

i=1<br />

2 i<br />

+ ∂2<br />

∂y2 <br />

− y(n − 2) ∂<br />

.<br />

∂y<br />

(3.1)<br />

For a fixed z ′ = (x ′ , y ′ ) ∈ H n , for all t > 0 and z = (x, y) ∈ H n , we denote with Gn(z, z ′ , t) the<br />

transition function of the hyperbolic <strong>Brownian</strong> motion <strong>in</strong> H n with start<strong>in</strong>g po<strong>in</strong>t z ′ . The transition<br />

function Gn(z, z ′ , t) is solution to the follow<strong>in</strong>g heat-type equation<br />

with <strong>in</strong>itial condition<br />

∂<br />

∂t un(z, z ′ , t) = ∆n un(z, z ′ , t), (3.2)<br />

un(z, z ′ , 0) =<br />

n−1 <br />

i=1<br />

δ(xi − x ′ i)δ(y − y ′ ).<br />

We have seen <strong>in</strong> Theorem 1.2.1 that if Φ : H n → H n is a diffeomorphism, then ∆f = (∆(f ◦<br />

Φ)) ◦ Φ −1 if and only if Φ is an isometry, this <strong>in</strong>variance property implies that<br />

u(z, z ′ , t) = u(Φ(z), Φ(z ′ ), t).<br />

Let η(z, z ′ ) be the hyperbolic distance between z and the start<strong>in</strong>g po<strong>in</strong>t z ′ , given by formula<br />

(2.25), s<strong>in</strong>ce the Laplace operator commutes with rotations and translations, we have that the<br />

heat kernel Gn(z, z ′ , t) is a function of η(z, z ′ ) and, without restrictions, we can assume that the<br />

start<strong>in</strong>g po<strong>in</strong>t z ′ is the orig<strong>in</strong> O = (0, 1) of H n (see for example Davies (10) page 178 or Borthwick<br />

(3) Section 4.1).<br />

S<strong>in</strong>ce, for a fixed time t, the transition function Gn(z, z ′ , t) is <strong>in</strong> fact a function of the hyperbolic


3.1 Sett<strong>in</strong>gs 30<br />

distance η = η(z) = η(z, O), we denote it by Gn(η, t) and we take <strong>in</strong>to account only the radial part<br />

of the Laplace operator (3.2) with <strong>in</strong>itial condition un(η, 0) = δ(η).<br />

It is possible to prove that the follow<strong>in</strong>g relation holds<br />

∆n η(z) = (n − 1) coth η(z). (3.3)<br />

The proof of relation (3.3) is postponed to Proposition 3.A.1 <strong>in</strong> the Appendix 3.A. In view of (3.3),<br />

for a smooth function f on R, we deduce that<br />

∆nf(η) = f ′′ (η) + (n − 1) coth η f ′′ (η). (3.4)<br />

For a proof of (3.4) see Proposition 3.A.2 <strong>in</strong> the Appendix 3.A. S<strong>in</strong>ce we can rewrite formula (3.4)<br />

<strong>in</strong> the follow<strong>in</strong>g form<br />

1<br />

∆nf(η) =<br />

s<strong>in</strong>h n−1 <br />

d<br />

s<strong>in</strong>h<br />

η dη<br />

n−1 η d<br />

dη f(η)<br />

<br />

,<br />

we f<strong>in</strong>ally note that the heat kernel Gn(η, t) is obta<strong>in</strong>ed by resolv<strong>in</strong>g the follow<strong>in</strong>g Cauchy problem<br />

⎧<br />

⎨ ∂<br />

∂t<br />

⎩<br />

un(η, t) =<br />

un(η, 0) = δ(η).<br />

1<br />

s<strong>in</strong>h n−1 η<br />

<br />

∂<br />

∂η s<strong>in</strong>h n−1 η ∂<br />

<br />

∂η un(η, t),<br />

S<strong>in</strong>ce an hyperbolic sphere <strong>in</strong> H n with hyperbolic center O and hyperbolic radius η is a Euclidean<br />

sphere with Euclidean center (0, cosh r) ∈ H n and Euclidean radius s<strong>in</strong>h r, we have that the volume<br />

element <strong>in</strong> H n is given by<br />

dv = s<strong>in</strong>h n−1 η dη dΩn,<br />

where dΩn is the surface element of the n-dimensional unit sphere<br />

n−1 <br />

dΩn = (s<strong>in</strong> θi) n−1−i dθi.<br />

i=1<br />

To obta<strong>in</strong> the normalized heat kernel we impose that<br />

<br />

Gn(η, t) dv = 2πn/2<br />

Γ( n<br />

2 )<br />

∞<br />

0<br />

(3.5)<br />

Gn(η, t) s<strong>in</strong>h n−1 η dη = 1. (3.6)<br />

The probability density pn(η, t) to be at time t at a distance η from the start<strong>in</strong>g po<strong>in</strong>t, normalized<br />

with respect to s<strong>in</strong>h n−1 η dη, is given by<br />

pn(η, t) = 2πn/2<br />

Γ( n<br />

2 ) Gn(η, t).<br />

We denote by (B (n) , B (n) , P (n) ) the n-dimensional standard Wiener space where B (n) is the<br />

space of all R n valued cont<strong>in</strong>uous paths on [0, ∞), with (B1(s), . . . , Bn(s)) ∈ B (n) satisfy<strong>in</strong>g<br />

(B1(0), . . . , Bn(0)) = 0 and (B (n) (s), s ≥ 0) is the canonical filtration (see Revuz and Yor (36)<br />

page 35).<br />

Tak<strong>in</strong>g <strong>in</strong>to account the generator ∆n/2, where ∆n is given by (3.1), it follows that the <strong>Brownian</strong>


31 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

motion on H n may be obta<strong>in</strong>ed by solv<strong>in</strong>g the stochastic differential equations<br />

⎧<br />

⎨dXi(s)<br />

= Y (s)dBi(s), i = 1, ..., n − 1,<br />

⎩dY<br />

(s) = Y (s)dBn(s) − n−2Y<br />

(s)ds.<br />

By Itô’s formula we have that the solution {Z(t) = (X(t), Y (t)), t ≥ 0} satisfy<strong>in</strong>g Z(0) = z = (x, y)<br />

may be represented as follows:<br />

2<br />

⎧<br />

⎨Xi(t)<br />

= xi + y<br />

⎩<br />

t<br />

0 exp{Bµ (s)} dBi(s), i = 1, ..., n − 1,<br />

Y (t) = y exp{B µ (t)},<br />

where µ = (n − 1)/2 and B µ (t) = Bn(t) − µt has constant drift −µ. S<strong>in</strong>ce µ is assumed to be<br />

positive, Y (t) is the usual geometric <strong>Brownian</strong> motion with negative drift and then {Z(t), t ≥ 0}<br />

converges almost surely to the boundary of H n as t → ∞.<br />

3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2<br />

3.2.1 Transition function<br />

The 2-dimensional hyperbolic <strong>Brownian</strong> motion is a diffusion process on the upper half-plane<br />

H 2 = {z = (x, y) : x ∈ R, y > 0}<br />

associated with the Laplace operator ∆2 given by<br />

(3.7)<br />

∆2 = y 2<br />

2 ∂ ∂2<br />

+<br />

∂x2 ∂y2 <br />

. (3.8)<br />

The position of po<strong>in</strong>ts <strong>in</strong> H 2 can be given either <strong>in</strong> terms of Cartesian coord<strong>in</strong>ates (x, y) or by<br />

means of the hyperbolic coord<strong>in</strong>ates (η, α). The coord<strong>in</strong>ate η represents the hyperbolic distance<br />

of (x, y) from the orig<strong>in</strong> O = (0, 1) of H 2 . The coord<strong>in</strong>ate α is the slope of the tangent <strong>in</strong> O to<br />

the geodesic through O and (x, y) with Euclidean center (x0, 0) and Euclidean radius r (see Figure<br />

5.3).<br />

The formulas which relates the hyperbolic coord<strong>in</strong>ate to the cartesian coord<strong>in</strong>ates are<br />

x =<br />

y =<br />

s<strong>in</strong>h η cos α<br />

cosh η−s<strong>in</strong>h η s<strong>in</strong> α ,<br />

1<br />

cosh η−s<strong>in</strong>h η s<strong>in</strong> α ,<br />

with η > 0 and α ∈ [− π π<br />

2 , 2 ]. See, for example, Rogers and Williams (37), page 213 or Terras (41)<br />

formula (3.7). For a proof of the relations <strong>in</strong> (3.9) see Proposition 3.A.3.<br />

In Proposition 3.A.4 it is showed that the Laplace operator (3.8) <strong>in</strong> hyperbolic coord<strong>in</strong>ates takes<br />

the form<br />

<br />

1 ∂<br />

s<strong>in</strong>h η<br />

s<strong>in</strong>h η ∂η<br />

∂<br />

<br />

1<br />

+<br />

∂η s<strong>in</strong>h 2 ∂<br />

η<br />

2<br />

.<br />

∂α2 (3.10)<br />

S<strong>in</strong>ce, for a fixed t, the transition function p2(z, z ′ , t) is a function of the hyperbolic distance η,<br />

we can take <strong>in</strong>to account only the radial part of the Laplace operator (3.10).<br />

(3.9)


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 32<br />

Figure 3.1: <strong>Hyperbolic</strong> coord<strong>in</strong>ates <strong>in</strong> H 2 .<br />

In the next theorem, follow<strong>in</strong>g the proof given <strong>in</strong> Lao and Ors<strong>in</strong>gher (26), we give the explicit<br />

form of the normalized heat kernel G2(η, t). In the literature there are several proofs of formula<br />

(3.11), for example Buser obta<strong>in</strong> the heat kernel G2(η, t) by us<strong>in</strong>g the Abel transform, see Buser<br />

(4) Theorem 7.4.1. An approach based on the Mehler transform is given <strong>in</strong> Chavel (9) Section X.2.<br />

Other transform techniques are described <strong>in</strong> Terras (41) Section 3.2 and Helgason (19) page 29.<br />

Theorem 3.2.1. The normalized heat kernel G2(η, t), t > 0, has the follow<strong>in</strong>g explicit expression<br />

G2(η, t) =<br />

t − e 4<br />

25/2 (πt) 3/2<br />

∞<br />

η<br />

ϕ2<br />

− ϕe 4t<br />

√ dϕ, η > 0. (3.11)<br />

cosh ϕ − cosh η<br />

Proof<br />

In order to obta<strong>in</strong> formula (3.11) we must solve the <strong>in</strong>itial-value problem<br />

⎧<br />

<br />

⎨ ∂u 1 ∂<br />

∂t = s<strong>in</strong>h η ∂η s<strong>in</strong>h η<br />

⎩<br />

∂<br />

<br />

∂η u,<br />

u(η, 0) = δ(η).<br />

It is now convenient to perform the change of variable cosh η = y which yields<br />

We now put<br />

that gets the follow<strong>in</strong>g equations<br />

∂u<br />

∂t =<br />

<br />

(y 2 − 1) ∂2<br />

<br />

∂<br />

+ 2y u.<br />

∂y2 ∂y<br />

u(t, y) = T (t)F (y)<br />

⎧<br />

⎨T<br />

⎩<br />

′ (t) = −T (t) ω,<br />

(y 2 − 1)F ′′ + 2yF ′ + ωF = 0,<br />

(3.12)<br />

(3.13)<br />

where ω is an arbitrary positive constant. The method of separation of variables produces two<br />

ord<strong>in</strong>ary equations, one of which is strictly related to Legendre equation. The bounded solution of


33 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

equation (3.12), because of its homogeneity and l<strong>in</strong>earity, has thus the form<br />

u(η, t) =<br />

∞<br />

0<br />

T (t, ω)F (η, ω)Q(ω)dω, (3.14)<br />

where Q(ω), ω > 0, is an arbitrary function which is determ<strong>in</strong>ed by apply<strong>in</strong>g the <strong>in</strong>itial condition.<br />

The equation satisfied by F , for ω = −ν(ν + 1), becomes the classical Legendre equation<br />

which is a special case of the hypergeometric equation<br />

(1 − z 2 ) d2u du<br />

− 2z + ν(ν + 1)u = 0 (3.15)<br />

dz2 dz<br />

t(1 − t) d2u du<br />

+ [γ − (α + β + 1)t] − αβu = 0, (3.16)<br />

dt2 dt<br />

for z = 1 − 2t, α = −ν, β = ν + 1 and γ = 1. A solution to (3.16) is the hypergeometric function<br />

2F1(α, β, γ; t) =<br />

=<br />

=<br />

∞ (α)k(β)k<br />

t<br />

k!(γ)k<br />

k<br />

∞ α(α + 1)...(α + k − 1)β(β + 1)...(β + k − 1)<br />

t<br />

k!γ(γ + 1)...(γ + k − 1)<br />

k<br />

∞ Γ(α + k)Γ(β + k) Γ(γ)<br />

k!Γ(γ + k) Γ(α)Γ(β) tk , |t| < 1.<br />

k=0<br />

k=0<br />

k=0<br />

This means that a solution to the Legendre equation (3.15) is the Legendre polynomial<br />

Pν(z) = 2F1<br />

=<br />

∞<br />

k=0<br />

<br />

∞<br />

k 1 − z (−ν)k(ν + 1)k 1 − z<br />

−ν, ν + 1, 1; =<br />

2<br />

k!(1)k 2<br />

k=0<br />

(−ν)k(ν + 1)k<br />

(k!) 2<br />

k 1 − z<br />

, |1 − z| < 2. (3.17)<br />

2<br />

By follow<strong>in</strong>g Lebedev (27), page 165, it is possible to represent the Legendre polynomials <strong>in</strong> a more<br />

suitable form by consider<strong>in</strong>g that<br />

By substitut<strong>in</strong>g (3.18) <strong>in</strong> (3.17) we get that<br />

Pν(z) =<br />

π<br />

2 2<br />

s<strong>in</strong><br />

π 0<br />

2k <br />

1<br />

2 k<br />

ϕ dϕ = . (3.18)<br />

Γ(k + 1)<br />

∞ (−ν)k(ν + 1)k<br />

<br />

1<br />

k=0 2 k k!<br />

k 1 − z 2<br />

2 π<br />

= 2<br />

π ∞ 2 (−ν)k(ν + 1)k<br />

<br />

π<br />

1<br />

0<br />

2 k k!<br />

<br />

s<strong>in</strong> 2 ϕ<br />

k=0<br />

= 2<br />

π<br />

2<br />

2F1<br />

π 0<br />

<br />

−ν, ν + 1, 1<br />

2 ; s<strong>in</strong>2 ϕ<br />

π<br />

2<br />

0<br />

1 − z<br />

2<br />

s<strong>in</strong> 2k ϕ dϕ<br />

k 1 − z<br />

dϕ<br />

2<br />

<br />

dϕ. (3.19)


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 34<br />

The function <strong>in</strong> the <strong>in</strong>tegrand of (3.19) can be rewritten as<br />

<br />

fν(ω) = 2F1 −ν, ν + 1, 1<br />

<br />

; −ω<br />

2<br />

√ √ 2ν+1 √ √ −2ν−1 1 + ω + ω + 1 + ω + ω<br />

=<br />

2 √ ,<br />

1 + ω<br />

(3.20)<br />

<strong>in</strong> fact fν(ω) solves the hypergeometric equation (3.16) for α = −ν, β = ν + 1, γ = 1<br />

2 and t = −ω,<br />

that is<br />

ω(1 + ω) d2 <br />

fν 1 dfν<br />

+ + 2ω<br />

dω2 2 dω − ν(ν + 1)fν = 0,<br />

which can also be rewritten <strong>in</strong> the more convenient form<br />

√ √ω√ √ ′<br />

ω 1 + ω 1 + ωfν<br />

<br />

′<br />

− ν + 1<br />

2 fν = 0. (3.21)<br />

2<br />

It is now a relatively simple matter to show that (3.20) is a solution to (3.21) and also to check<br />

that<br />

fν(ω) = f−ν−1(ω).<br />

where<br />

By some calculations it can be seen that<br />

fν<br />

s<strong>in</strong> 2 ϕ<br />

Therefore the Legendre polynomial (3.19)<br />

Pν(cosh η) = 2<br />

π<br />

2<br />

π<br />

= 2<br />

π<br />

2<br />

0<br />

π<br />

2<br />

<strong>in</strong> view of (3.22), takes the follow<strong>in</strong>g form<br />

0<br />

<br />

(cosh η − 1) = cosh θ(ν + 1<br />

2 )<br />

cosh θ<br />

2<br />

s<strong>in</strong>h θ η<br />

= s<strong>in</strong>h s<strong>in</strong> ϕ. (3.22)<br />

2 2<br />

2F1<br />

fν<br />

Pν(cosh η) = 1<br />

η<br />

π −η<br />

<br />

−ν, ν + 1, 1<br />

2 ; s<strong>in</strong>2 <br />

ϕ<br />

(1 − cosh η) dϕ<br />

2<br />

2 <br />

s<strong>in</strong> ϕ<br />

(cosh η − 1) dϕ,<br />

2<br />

1<br />

(ν+ e 2)θ<br />

dθ. (3.23)<br />

2(cosh η − cosh θ)<br />

We skip the next step which consists <strong>in</strong> apply<strong>in</strong>g a contour <strong>in</strong>tegration to the complex-valued<br />

function. The detailes of this transformation can be found <strong>in</strong> Lebedev (27), page 172.<br />

1<br />

(ν+ e 2)t<br />

f(t) =<br />

π , t ∈ C,<br />

2(cosh η − cosh t)<br />

and permits us to express the Legendre function (3.23) by means of an <strong>in</strong>tegral on [η, ∞) as follows<br />

Pν(cosh η) = 2<br />

<br />

cot π ν +<br />

π 1<br />

∞ s<strong>in</strong>h[<br />

2 η<br />

ν + 1<br />

<br />

2 ϕ]<br />

√ dϕ. (3.24)<br />

2 cosh ϕ − 2 cosh η


35 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

If we now choose ω = 1<br />

4 + x2 <strong>in</strong> the second equation of (3.13) we note that it co<strong>in</strong>cides with<br />

(3.15) when ν = − 1<br />

2 ± ix. From (3.14), by consider<strong>in</strong>g (3.24), we get that<br />

u(η, t) =<br />

= 2<br />

π<br />

∞<br />

t −<br />

e 4<br />

0<br />

−x2t P− 1<br />

2 +ix(cosh η) x Q(x) dx (3.25)<br />

∞<br />

t −<br />

x Q(x)e 4<br />

0<br />

−x2 ∞<br />

t s<strong>in</strong>(xϕ)<br />

coth(πx) dx √ dϕ.<br />

η 2 cosh ϕ − 2 cosh η<br />

We show now that if we choose Q(x) = k tanh(πx), for some real constant k, then the <strong>in</strong>itial<br />

condition u(η, 0) = δ(η) is satisfied. We have<br />

u(η, t) =<br />

=<br />

=<br />

√<br />

2k<br />

π<br />

√<br />

2k<br />

π<br />

∞<br />

0<br />

e− t<br />

4<br />

t − ke 4<br />

√ <br />

π (2t) 3<br />

t −<br />

x tanh(πx) e 4 −x2 ∞<br />

t s<strong>in</strong>(xϕ)<br />

coth(πx) dx √ dϕ<br />

cosh ϕ − cosh η<br />

η<br />

∞<br />

∞<br />

dϕ<br />

√ xe<br />

η cosh ϕ − cosh η 0<br />

−x2t s<strong>in</strong>(xϕ)dϕ<br />

∞<br />

ϕ2<br />

− ϕe 4t<br />

√ dϕ (3.26)<br />

cosh ϕ − cosh η<br />

η<br />

In order to check that (3.26) satisfies the <strong>in</strong>itial condition it suffices to write that<br />

lim<br />

t→0<br />

t→0<br />

+ u(η, t) = lim<br />

∞<br />

k<br />

+<br />

= k<br />

∞<br />

η/ √ 2<br />

=<br />

η/ √ 2<br />

ϕ2<br />

− ϕe 2t<br />

√<br />

2πt3 δ(ϕ)<br />

<br />

cosh( √ 2ϕ) − cosh η<br />

dϕ<br />

<br />

cosh( √ 2ϕ) − cosh η<br />

dϕ<br />

⎧<br />

⎨0,<br />

for η > 0,<br />

⎩k<br />

∞<br />

0<br />

δ(ϕ)<br />

√ cosh( √ 2ϕ)−1 dϕ = ∞, for η = 0.<br />

In the last step we took <strong>in</strong>to account the behaviour of the distribution of the first-passage time of<br />

a standard <strong>Brownian</strong> motion. We f<strong>in</strong>ally observe that if G2(η, t) = u(η, t), where u(η, t) is given<br />

by (3.26) with k = 1<br />

2π , then it satisfies the condition (3.6). In fact<br />

<br />

G2(η, t) dv =<br />

t − 1 e 4<br />

√ <br />

π (2t) 3<br />

= 1<br />

t − e 4<br />

√ <br />

π (2t) 3<br />

=<br />

t − e 4<br />

√ <br />

π (2t) 3<br />

t<br />

2e− 4<br />

= √ <br />

π (2t) 3<br />

√<br />

2<br />

= √ <br />

π (2t) 3<br />

= 1<br />

t √ <br />

√2t<br />

4πt<br />

= 1<br />

√ 2π<br />

∞ ∞<br />

ϕ2<br />

− ϕe 4t<br />

s<strong>in</strong>h η dη √ dϕ<br />

0<br />

η cosh ϕ − cosh η<br />

∞ ϕ<br />

ϕ2<br />

− s<strong>in</strong>h η<br />

ϕe 4t dϕ √ dη<br />

0<br />

0 cosh ϕ − cosh η<br />

∞<br />

η=ϕ 0<br />

∞<br />

<br />

ϕ2<br />

−<br />

ϕe 4t dϕ −2 cosh ϕ − cosh η<br />

ϕe<br />

ϕ2<br />

− 4t<br />

cosh ϕ − 1dϕ =<br />

η=0<br />

∞<br />

t − 2e 4<br />

√ <br />

π (2t) 3<br />

<br />

(ϕ+t)2<br />

−<br />

ϕe 4t dϕ<br />

ϕ2<br />

− e<br />

ϕe 4t<br />

ϕ<br />

ϕ<br />

2 − − e 2<br />

√ dϕ<br />

2<br />

0<br />

∞<br />

∞<br />

(ϕ−t)2<br />

−<br />

ϕe 4t dϕ −<br />

0<br />

0<br />

0<br />

∞<br />

− √ (<br />

t<br />

2<br />

√ y2<br />

−<br />

2ty + t)e 2 dy − √ ∞<br />

2t √ (<br />

t<br />

2<br />

√ y2<br />

−<br />

2ty − t)e 2 dy<br />

∞<br />

y2<br />

−<br />

e 2 dy = 1.<br />

−∞


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 36<br />

<br />

Remark 3.2.1. In view of Theorem 3.2.1 we obta<strong>in</strong> that the probability density p2(η, t) to be at<br />

time t at a distance η, normalized with respect to the area element s<strong>in</strong>h η dη, is given by<br />

p2(η, t) = 2π G2(η, t) =<br />

t − ∞<br />

e 4<br />

√ <br />

π (2t) 3<br />

η<br />

ϕ2<br />

− ϕe 4t<br />

√ dϕ. (3.27)<br />

cosh ϕ − cosh η<br />

Remark 3.2.2. Sometime, <strong>in</strong> the literature, <strong>in</strong>stead of equation (3.2), it is considered the equation<br />

∂ 1<br />

∂t u = 2∆2u, <strong>in</strong> this case the probability density (3.27) must be replaced by<br />

p2(η, t) =<br />

t − ∞<br />

e 8<br />

√ √ 3<br />

π t η<br />

ϕ2<br />

− ϕe 2t<br />

√ dϕ.<br />

cosh ϕ − cosh η<br />

Remark 3.2.3. From (3.27) also the distribution function of η(t) can be obta<strong>in</strong>ed:<br />

where ft(ϕ) =<br />

Pr{η(t) > η} =<br />

=<br />

=<br />

t − e 4<br />

√ <br />

π (2t) 3<br />

t − e 4<br />

√ <br />

π (2t) 3<br />

t − 2e 4<br />

√ <br />

π (2t) 3<br />

∞<br />

η<br />

∞<br />

η<br />

∞<br />

η<br />

= 2 √ ∞<br />

t −<br />

2e 4<br />

η<br />

∞<br />

dη<br />

η<br />

ϕ2<br />

−<br />

ϕe 4t dϕ<br />

ϕe<br />

ϕ2<br />

− 4t<br />

ϕ2<br />

− ϕe 4t s<strong>in</strong>h η<br />

√ dϕ<br />

cosh ϕ − cosh η<br />

ϕ<br />

η<br />

s<strong>in</strong>h η<br />

√ cosh ϕ − cosh η dη<br />

cosh ϕ − cosh ηdϕ<br />

ft(ϕ) cosh ϕ − cosh ηdϕ,<br />

ϕ2<br />

ϕe− 4t<br />

√ , t > 0, ϕ > 0, as a function of t, is the distribution of the first-passage<br />

2π(2t) 3<br />

time for the standard <strong>Brownian</strong> motion, with <strong>in</strong>f<strong>in</strong>itesimal variance equal to 2, through level ϕ.<br />

3.2.2 Hitt<strong>in</strong>g distribution<br />

In several papers (see Baldi et al. (1), Matsumoto and Yor (30)) it is considered the diffusion on<br />

H 2 associated to the operator<br />

L = y2<br />

2<br />

2 ∂ ∂2<br />

+<br />

∂x2 ∂y2 <br />

− νy ∂<br />

∂x −<br />

<br />

µ − 1<br />

<br />

y<br />

2<br />

∂<br />

∂y<br />

where ν ∈ R and µ > 0. The coefficients µ and ν measure respectively the vertical and the<br />

horizontal component of the drift. In view of (3.1) we obta<strong>in</strong> that, for µ = 1/2 and ν = 0, the<br />

operator L co<strong>in</strong>cides with ∆2/2 and <strong>in</strong> this case we have the standard two dimensional hyperbolic<br />

<strong>Brownian</strong> motion.<br />

Unlike ∆2, that is <strong>in</strong>variant under hyperbolic isometries, the differential operator L is <strong>in</strong>variant<br />

under transformations of the form z → az + b with a > 0 and b ∈ R.<br />

The diffusion process {Z(t) = (X(t), Y (t)), t > 0} associated to L and such that Z(0) = z =


37 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

(x, y) ∈ H 2 may be realized as the solution of the stochastic differential equations<br />

⎧<br />

⎨dX(t)<br />

= Y (t)dB1(t) − νY (t)dt,<br />

⎩dY<br />

(t) = Y (t)dB2(t) − µ − 1<br />

<br />

Y (t)dt,<br />

where B1(t) and B2(t) are two <strong>in</strong>dependent one-dimensional <strong>Brownian</strong> motions. Z(t) is represented<br />

as follows ⎧<br />

⎨X(t)<br />

= x + y<br />

⎩<br />

t<br />

0 exp{Bµ 2 (s)} dBν 1 (s),<br />

Y (t) = y exp{B µ<br />

2 (t)},<br />

where B ν 1 (t) = B1(t)−νt and B µ<br />

2 (t) = B2(t)−µt. S<strong>in</strong>ce µ is assumed to be positive, Y (t) converges<br />

to 0 as t tends to ∞. It is possible to show that the distribution of the horizontal component X(t)<br />

converges weakly as t tends to <strong>in</strong>f<strong>in</strong>ity to a given limit<strong>in</strong>g distribution.<br />

Let ha = {z ∈ H2 : Im{z} = a}, with 0 ≤ a < y, be the horocycle through ∞. If a = 0 we<br />

have that h0 is the boundary of the hyperbolic half plane H2. The hitt<strong>in</strong>g distribution on ha, with<br />

a > 0, of the diffusion process associated to L is given by the law of the random variable<br />

τa<br />

Xτa = x + y exp{B<br />

0<br />

µ<br />

2 (s)} dBν 1 (s)<br />

where τa = <strong>in</strong>f{t ≥ 0 : Y (t) = a} is the first hitt<strong>in</strong>g time on ha. With a slight abuse of term<strong>in</strong>ology<br />

the law of the random variable<br />

is the hitt<strong>in</strong>g distribution on h0.<br />

Hitt<strong>in</strong>g distribution on h0<br />

X∞ = x + y<br />

∞<br />

0<br />

exp{B µ<br />

2 (s)} dBν 1 (s)<br />

Let pz(ξ) be the density of the hitt<strong>in</strong>g distribution on h0 for the process associated to the operator<br />

L and start<strong>in</strong>g at z = (x, y). We po<strong>in</strong>t out that pz(ξ) is the Poisson kernel of L <strong>in</strong> the doma<strong>in</strong> H 2<br />

and satisfies the follow<strong>in</strong>g conditions:<br />

⎧<br />

Lpz(ξ) = 0, ∀ξ ∈ h0,<br />

⎪⎨ pz(ξ) > 0, ∀ξ ∈ h0,<br />

∞<br />

−∞<br />

⎪⎩<br />

pz(ξ)dξ = 1,<br />

limy→0 + pz(ξ) = 0, if ξ = x.<br />

S<strong>in</strong>ce L is <strong>in</strong>variant under the action of the map z → az + b then so is pz(ξ)dξ on R. That is<br />

pz(ξ) dξ = paz+b(aξ + b) d(aξ + b) = a paz+b(aξ + b) dξ.<br />

Sett<strong>in</strong>g a = 1 x<br />

y , b = − y and f(x) = pi(x) we have<br />

pz(ξ) = px+iy(ξ) = 1<br />

y pi<br />

<br />

ξ − x<br />

y<br />

2<br />

= 1<br />

y f<br />

<br />

ξ − x<br />

y<br />

(3.28)<br />

(3.29)


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 38<br />

then the hitt<strong>in</strong>g distribution with arbitrary start<strong>in</strong>g po<strong>in</strong>t z is obta<strong>in</strong>ed by rescal<strong>in</strong>g the one with<br />

start<strong>in</strong>g po<strong>in</strong>t z = i.<br />

Theorem 3.2.2. The horizontal component X(t), with start<strong>in</strong>g po<strong>in</strong>t z = i, converges <strong>in</strong> distribution,<br />

as t → ∞, to the random variable X∞ with density<br />

f(x) = 22µ−1 Γ( 1<br />

2 + µ − iν) 2<br />

πΓ(2µ)<br />

−2ν arctan x e<br />

(1 + x2 .<br />

) µ+1/2<br />

Proof<br />

The first condition <strong>in</strong> (3.28), Lpz(ξ) = 0, ∀ξ ∈ h0, can be rewritten for f(x) us<strong>in</strong>g formula (3.29).<br />

We have<br />

<br />

∂ 1<br />

∂x y f<br />

<br />

ξ − x<br />

y<br />

∂2 ∂x2 <br />

1<br />

y f<br />

<br />

ξ − x<br />

y<br />

<br />

∂ 1<br />

∂y y f<br />

<br />

ξ − x<br />

y<br />

∂2 ∂y2 <br />

1<br />

y f<br />

<br />

ξ − x<br />

y<br />

= − 1<br />

y2 f ′ (u)| ξ−x ,<br />

y<br />

= 1<br />

y3 f ′′ (u)| ξ−x ,<br />

y<br />

= − 1<br />

<br />

ξ − x ξ − x<br />

f −<br />

y2 y y3 f ′ (u)| ξ−x ,<br />

y<br />

= (ξ − x)2<br />

y 5<br />

f ′′ (u)| ξ−x<br />

y<br />

ξ − x<br />

+ 4<br />

y4 f ′ (u)| ξ−x<br />

y<br />

+ 2<br />

f<br />

y3 ξ − x<br />

then we obta<strong>in</strong><br />

Lpz(ξ) = y2<br />

2 ∂<br />

2 ∂x2 <br />

1<br />

y f<br />

<br />

ξ − x<br />

+<br />

y<br />

∂2<br />

∂y2 <br />

1<br />

y f<br />

<br />

ξ − x<br />

− νy<br />

y<br />

∂<br />

<br />

1<br />

∂x y f<br />

<br />

ξ − x<br />

y<br />

<br />

− µ − 1<br />

<br />

y<br />

2<br />

∂<br />

<br />

1<br />

∂y y f<br />

=<br />

<br />

ξ − x<br />

y<br />

1<br />

2y f ′′ (u)| ξ−x +<br />

y<br />

(ξ − x)2<br />

2y3 f ′′ ξ − x<br />

(u)| ξ−x + 2<br />

y y2 f ′ (u)| ξ−x +<br />

y<br />

1<br />

y f<br />

<br />

ξ − x<br />

+<br />

y<br />

ν<br />

y f ′ (u)| ξ−x<br />

y<br />

<br />

− µ − 1<br />

<br />

−<br />

2<br />

1<br />

y f<br />

<br />

ξ − x ξ − x<br />

−<br />

y y2 f ′ <br />

(u)| ξ−x .<br />

y<br />

Sett<strong>in</strong>g y = 1 and x = 0 we have that f(ξ) solves the second-order l<strong>in</strong>ear differential equation<br />

Mf(ξ) = 0 where<br />

Mf(ξ)<br />

= 1<br />

2 f ′′ (ξ) + ξ2<br />

2 f ′′ (ξ) + 2ξf ′ (ξ) + f(ξ) + νf ′ <br />

(ξ) − µ − 1<br />

<br />

[−f(ξ) − ξf<br />

2<br />

′ (ξ)]<br />

= 1 + ξ2<br />

f<br />

2<br />

′′ <br />

(ξ) + ν + µ + 3<br />

<br />

ξ f<br />

2<br />

′ <br />

(ξ) + µ + 1<br />

<br />

f(ξ) (3.30)<br />

2<br />

= d<br />

<br />

1 + ξ2<br />

ξf(ξ) + f<br />

dξ<br />

2<br />

′ <br />

(ξ) + νf(ξ) + µ − 1<br />

<br />

ξf(ξ)<br />

2<br />

= d<br />

<br />

2 d 1 + ξ<br />

f(ξ) + νf(ξ) + µ −<br />

dξ dξ 2<br />

1<br />

<br />

ξf(ξ) .<br />

2<br />

With the change of variable ξ = tan φ<br />

2<br />

and by impos<strong>in</strong>g that f(ξ) = 2<br />

1+ξ 2<br />

y<br />

<br />

.<br />

g(2 arctan ξ) =


39 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

2 φ<br />

2 cos 2 g(φ), we obta<strong>in</strong><br />

<br />

2<br />

d d 1 + ξ<br />

f(ξ) + νf(ξ) + µ −<br />

dξ dξ 2<br />

1<br />

<br />

ξf(ξ)<br />

2<br />

<br />

<br />

2 φ d 2 φ d<br />

= 2 cos 2 cos g(φ) + ν + µ −<br />

2 dφ 2 dφ 1<br />

<br />

tan<br />

2<br />

φ<br />

<br />

2 φ<br />

2 cos<br />

2 2 g(φ)<br />

<br />

<br />

2 φ d d<br />

= 4 cos g(φ) + ν + µ −<br />

2 dφ dφ 1<br />

<br />

tan<br />

2<br />

φ<br />

<br />

2 φ<br />

g(φ) cos .<br />

2<br />

2<br />

s<strong>in</strong>ce d<br />

dξ<br />

= 2 cos2 φ<br />

2<br />

d<br />

dφ<br />

2 φ d<br />

4 cos<br />

2 dφ<br />

and is proportional to g1(φ) that satisfies<br />

. The function g(φ) is solution to the differential equation<br />

<br />

d<br />

g(φ) + ν + µ −<br />

dφ 1<br />

<br />

tan<br />

2<br />

φ<br />

<br />

g(φ) cos<br />

2<br />

<br />

= 0<br />

2<br />

2 φ<br />

d<br />

dφ g1(φ)<br />

<br />

+ ν + µ − 1<br />

<br />

tan<br />

2<br />

φ<br />

<br />

−2 φ<br />

g1(φ) = k cos<br />

2<br />

2 .<br />

The function g(φ) has the follow<strong>in</strong>g form<br />

−F (φ)<br />

g(φ) = c1g1(φ) = c1e<br />

where c, c1 are two real constants and<br />

Then<br />

F (φ) =<br />

<br />

c<br />

2 +<br />

φ<br />

0<br />

F (ψ) k<br />

e<br />

cos2 ψ<br />

2<br />

<br />

ν + µ − 1<br />

<br />

tan<br />

2<br />

φ<br />

<br />

<br />

dφ = νφ − 2 µ −<br />

2<br />

1<br />

<br />

log cos<br />

2<br />

φ<br />

<br />

.<br />

2<br />

g(φ) = c1e −νφ <br />

2µ−1 φ c<br />

cos + k<br />

2 2<br />

φ<br />

0<br />

dψ<br />

<br />

e νψ −2µ−1 ψ<br />

cos<br />

2 dψ<br />

<br />

c, c1, k ∈ R. S<strong>in</strong>ce µ > 0 and φ ∈ (−π, π) the <strong>in</strong>tegral takes values of either sign. Tak<strong>in</strong>g <strong>in</strong>to<br />

account the second condition <strong>in</strong> (3.28) we have that k = 0 and<br />

g(φ) = c2<br />

2 e−νφ 2µ−1 φ<br />

cos<br />

2<br />

for some real constant c2. Recall<strong>in</strong>g that φ = 2 arctan ξ and f(ξ) = 2<br />

1+ξ 2 g(2 arctan ξ) we have<br />

and f<strong>in</strong>ally<br />

g(φ) =<br />

1 + ξ2<br />

2<br />

= c2 arctan ξ<br />

e−2ν<br />

2<br />

f(ξ) = c2<br />

2 e−νφ cos<br />

<br />

1<br />

<br />

1 + ξ2 f(ξ) = c2<br />

2µ−1 φ<br />

2<br />

2µ−1<br />

−2ν arctan ξ e<br />

(1 + ξ2 1<br />

) µ+ 2<br />

= c2<br />

2 e−2ν arctan ξ cos 2µ−1 (arctan ξ)<br />

By impos<strong>in</strong>g the third condition <strong>in</strong> (3.28) that is π<br />

g(φ)dφ = 1 we obta<strong>in</strong> the value of the constant<br />

−π<br />

.


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 40<br />

c2. With the change of variable u = φ<br />

2<br />

obta<strong>in</strong><br />

1 =<br />

π<br />

−π<br />

g(φ)dφ = c2<br />

2<br />

=<br />

π<br />

−π<br />

and by Gradshteyn and Ryzhik (16) formula 3.892.2. we<br />

e −νφ 2µ−1 φ<br />

cos dφ = c2<br />

2<br />

c2 π<br />

2 2µ−1 2µB 2µ+i2ν+1<br />

2<br />

, 2µ−i2ν+1.<br />

2<br />

π/2<br />

−π/2<br />

e −2νu cos 2µ−1 u du<br />

It follows that the value of the constant c2 is (see Gradshteyn and Ryzhik (16) formula 8.384.1 and<br />

8.326.1)<br />

<br />

c2 = 22µ−1 <br />

2µ<br />

B µ +<br />

π<br />

1<br />

<br />

1<br />

+ iν, µ + − iν =<br />

2 2 22µ−1 Γ(µ +<br />

π<br />

1<br />

2<br />

= 22µ−1<br />

π<br />

|Γ(µ + 1<br />

2 − iν)|2<br />

Γ(2µ)<br />

For a different proof of Theorem 3.2.2 see Baldi et al. (1) page 595.<br />

1<br />

+ iν)Γ(µ + 2 − iν)<br />

Γ(2µ)<br />

. (3.31)<br />

Remark 3.2.1. We note that if ν = 0 and µ = 1<br />

2 , that is if we consider the hyperbolic <strong>Brownian</strong><br />

motion on H2 without drift, the limit<strong>in</strong>g distribution is the Cauchy distribution as <strong>in</strong> the case<br />

of the hitt<strong>in</strong>g distribution on the x-axis for the Euclidean <strong>Brownian</strong> motion on R2 . In fact f(x)<br />

becomes<br />

f(x) =<br />

1<br />

π(1 + x 2 ) .<br />

In general f(x) belongs to the type IV family of Pearson distributions (see Johnson et al. (20)).<br />

Characteristic function of the hitt<strong>in</strong>g distribution on h0<br />

In view of (3.29) the characteristic function uz(λ) of X z ∞ for arbitrary start<strong>in</strong>g po<strong>in</strong>t z = (x, y) ∈ H 2<br />

can be derived easily from the special case u(λ) = ui(λ) of start<strong>in</strong>g po<strong>in</strong>t i. In fact<br />

uz(λ) =<br />

∞<br />

−∞<br />

= e iλx u(λy).<br />

e iλξ pz(ξ)dξ =<br />

∞<br />

−∞<br />

iλξ 1<br />

e<br />

y pi<br />

ξ − x<br />

y<br />

<br />

dξ = e iλx<br />

∞<br />

−∞<br />

e iλyw pi(w)dw<br />

Theorem 3.2.3. The characteristic function of the random variable X z ∞, with start<strong>in</strong>g po<strong>in</strong>t z = i,<br />

is given by<br />

u(λ) =<br />

1 Γ( 2 + µ + iν)<br />

e<br />

Γ(2µ)<br />

−λ <br />

1<br />

Φ − µ + iν, 1 − 2µ; 2λ<br />

2<br />

where Φ is the confluent hypergeometric function of the second k<strong>in</strong>d.<br />

Proof<br />

S<strong>in</strong>ce f(ξ) is a real-valued function we have u(−λ) = u(λ). Start<strong>in</strong>g from the differential equation<br />

(3.30)<br />

1 + ξ 2<br />

2<br />

f ′′ <br />

(ξ) + ν + µ + 3<br />

<br />

ξ f<br />

2<br />

′ <br />

(ξ) + µ + 1<br />

<br />

f(ξ) = 0<br />

2


41 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

and apply<strong>in</strong>g the <strong>in</strong>verse Fourier transform to both members, we get<br />

∞<br />

−∞<br />

e iλx<br />

2 1 + ξ<br />

f<br />

2<br />

′′ <br />

(ξ) + ν + µ + 3<br />

<br />

ξ f<br />

2<br />

′ <br />

(ξ) + µ + 1<br />

<br />

f(ξ) dξ = 0.<br />

2<br />

S<strong>in</strong>ce, for k = 0, 1, 2, the function xkf (k) (x) is <strong>in</strong>tegrable, we have that the function λku (k) (λ) exists,<br />

is cont<strong>in</strong>uous <strong>in</strong> R and vanishes at <strong>in</strong>f<strong>in</strong>ity. In particular u(λ) is twice cont<strong>in</strong>uously differentiable<br />

outside 0. We have<br />

∞<br />

e<br />

−∞<br />

iλx f (k) (x)dx = (−iλ) k u(λ) = (−1) k u (k) (λ),<br />

∞<br />

−∞<br />

∞<br />

−∞<br />

e iλx xf ′ (x)dx = −i ∂<br />

∞<br />

e<br />

∂λ −∞<br />

iλx f ′ (x)dx = −i ∂<br />

∂λ [−iλu(λ)] = −u(λ) − λu′ (λ),<br />

e iλx x 2 f ′′ (x)dx = − ∂2<br />

∂λ2 ∞<br />

e<br />

−∞<br />

iλx f ′′ (x)dx = − ∂2<br />

∂λ2 [(iλ)2u(λ)] = 2u(λ) + 4λu ′ (λ) + λ 2 u ′′ (λ).<br />

It follows that u(λ) is <strong>in</strong> the kernel of the operator N given by<br />

Nu(λ) = λ2<br />

2 u′′ <br />

(λ) − µ − 1<br />

<br />

λu<br />

2<br />

′ <br />

2 λ<br />

(λ) − + iνλ u(λ).<br />

2<br />

Tak<strong>in</strong>g <strong>in</strong>to account the third condition <strong>in</strong> (3.28) we have u(0) = ∞<br />

−∞ pz(ξ)dξ = 1. With the<br />

change of variable u(λ) = e −λ v(2λ) and ω = 2λ we get<br />

ωv ′′ (ω) + (1 − 2µ − ω)v ′ <br />

1<br />

(ω) − − µ + iν v(ω) = 0.<br />

2<br />

S<strong>in</strong>ce the confluent hypergeometric differential equation ωv ′′ (ω) + (b − ω)v ′ (ω) − av(ω) = 0 has<br />

solution<br />

Φ(a, b; ω) = 21−b Γ(1 − a)e ω<br />

2<br />

π<br />

π<br />

2<br />

0<br />

<br />

ω<br />

<br />

cos tan θ + (2a − b)θ cos<br />

2 −b θdθ<br />

where Φ(a, b; ω) is the confluent hypergeometric function of the second k<strong>in</strong>d with Re{b} < 1 and a<br />

not a positive <strong>in</strong>teger (see Gradshteyn and Ryzhik (16) formula 9.216.1), the characteristic function<br />

u(λ) we are look<strong>in</strong>g for is given by<br />

u(λ) = e −λ v(2λ) = Ke −λ <br />

1<br />

Φ − µ + iν, 1 − 2µ; 2λ .<br />

2<br />

for some real constant K. In particular we note that Φ(a, b; ω) has a f<strong>in</strong>ite non-zero limit for<br />

ω → 0 + ω − and that limω→∞ e 2 Φ(a, b; ω) = 0. To obta<strong>in</strong> the value of the constant K we observe<br />

that<br />

<br />

1<br />

1 = u(0) = KΦ − µ + iν, 1 − 2µ; 0 = K<br />

2 22µ<br />

π Γ<br />

π<br />

1<br />

2<br />

+ µ − iν cos(2iνθ) cos<br />

2 0<br />

2µ−1 θ dθ.


3.2 <strong>Hyperbolic</strong> <strong>Brownian</strong> motion <strong>in</strong> H 2 42<br />

Recall<strong>in</strong>g that e −2νθ = 2 cos(2iνθ) − e 2νθ , we have<br />

π/2<br />

−π/2<br />

e −2νθ cos 2µ−1 θdθ =<br />

= 2<br />

π/2<br />

−π/2<br />

π/2<br />

then <strong>in</strong> view of formula (3.31) we obta<strong>in</strong><br />

<br />

0<br />

[2 cos(2iνθ) − e 2νθ ] cos 2µ−1 π/2<br />

θdθ = 2<br />

cos(2iνθ) cos 2µ−1 θdθ,<br />

−π/2<br />

1 = K 22µ<br />

π Γ<br />

π<br />

1<br />

2<br />

+ µ − iν cos(2iνθ) cos<br />

2 0<br />

2µ−1 θ dθ<br />

= K 22µ<br />

π Γ<br />

π/2<br />

1<br />

1<br />

+ µ − iν e<br />

2 2 −π/2<br />

−2νθ cos 2µ−1 θdθ<br />

= K 22µ<br />

π Γ<br />

<br />

1<br />

1 π<br />

+ µ − iν<br />

2 2 22µ−1 Γ(2µ)<br />

.<br />

|Γ(1/2 + µ − iν)| 2<br />

cos(2iνθ) cos 2µ−1 θdθ<br />

It is possible to obta<strong>in</strong> the characteristic function u(λ) by means of the Feynman-Kac formula too,<br />

see Baldi et al. (1) Section 3.<br />

Hitt<strong>in</strong>g distribution on ha with a > 0<br />

It is possible to show that the characteristic function of the hitt<strong>in</strong>g distribution on ha, with a > 0,<br />

can be derived from the characteristic function of the hitt<strong>in</strong>g distribution on h0. We know that<br />

u(λy) =<br />

∞<br />

−∞<br />

= Eiy<br />

<br />

e iλyω <br />

f(ω)dω = E e iλyXi<br />

<br />

∞ = E e iλy R ∞<br />

0 exp{Bµ 2 (s)}dBν 1 (s)<br />

e iλ R ∞<br />

0<br />

Y (s)dBν<br />

1 (s)<br />

where Y (s) = y exp{B µ<br />

2 (s)} is the y-component of the hyperbolic <strong>Brownian</strong> motion start<strong>in</strong>g at<br />

z = (0, y). By condition<strong>in</strong>g with respect to the σ-algebra Fτa and us<strong>in</strong>g the strong Markov<br />

property one has, for a < y, that<br />

<br />

<br />

u(λy) = Eiy Eiy<br />

<br />

= Eiy e iλ R τa<br />

0<br />

It follows that<br />

<br />

<br />

e iλ R ∞<br />

Y (s)dBν<br />

0 1 (s)<br />

Fτa<br />

Y (s)dBν 1 (s)<br />

Eia<br />

<br />

<br />

= Eiy e iλ R τa<br />

0<br />

e iλ R ∞<br />

0<br />

Y (s)dBν<br />

1 (s)<br />

= Eiy<br />

Eiy[e iλXτa ] = u(λy)<br />

u(λa) .<br />

Y (s)dBν 1 (s) Eiy<br />

<br />

iλXτa e u(λa).<br />

e iλ R ∞<br />

Y (s)dB(ν)<br />

τa 1 (s) <br />

<br />

Fτa<br />

<br />

iλXτa<br />

Then the characteristic function Eiy e of the hitt<strong>in</strong>g distribution on ha of the hyperbolic<br />

<strong>Brownian</strong> motion start<strong>in</strong>g at iy is given by the ratio of the characteristic functions u(λy) and u(λa)<br />

of the hitt<strong>in</strong>g distribution on h0 for the hyperbolic <strong>Brownian</strong> motion start<strong>in</strong>g at i.<br />

The densities of the random variables X∞ and Xτa belong to the doma<strong>in</strong> of attraction of a<br />

stable law of which it is possible to determ<strong>in</strong>e the parameters, see Baldi et al. (1) Section 6.


43 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

3.2.3 <strong>Hyperbolic</strong> <strong>Brownian</strong> bridge<br />

Let {Bz1 (t), t ≥ 0} be a Euclidean <strong>Brownian</strong> motion <strong>in</strong> Rn start<strong>in</strong>g from z1 ∈ R n , n ≥ 2, then the<br />

Euclidean <strong>Brownian</strong> bridge {W (t), t ∈ [0, 1]} between z1 and z2 ∈ R n is given by<br />

W (t) = Bz1 (t) + t[z2 − Bz1 (t)]<br />

for every t ∈ [0, 1]. For the <strong>Brownian</strong> bridge <strong>in</strong> the Euclidean space we have that the probability<br />

that it stays uniformly close to the l<strong>in</strong>e jo<strong>in</strong><strong>in</strong>g z1 and z2 does non depend on the distance between<br />

z1 and z2.<br />

Let U(t) be the hyperbolic <strong>Brownian</strong> bridge <strong>in</strong> H n , n ≥ 2, between the orig<strong>in</strong> O and z ′ ∈ H n ,<br />

Eberle (see (13)) proved that on the hyperbolic space H n the sample paths of the hyperbolic<br />

<strong>Brownian</strong> bridge actually concentrate, accord<strong>in</strong>g to some exponential rate, around the geodesic<br />

through O and z ′ when z ′ tends to <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> f<strong>in</strong>ite time. Simon <strong>in</strong> (39) has found the exact<br />

exponential rate of concentration <strong>in</strong> the case of the hyperbolic <strong>Brownian</strong> bridge on H 2 .<br />

Let lz ′ be the unique geodesic segment jo<strong>in</strong><strong>in</strong>g O to z′ and def<strong>in</strong>e the follow<strong>in</strong>g function on H 2<br />

d O z ′(·) = <strong>in</strong>f{η(·, z), z ∈ lz ′}.<br />

For every a > 0, it is possible to prove the follow<strong>in</strong>g<br />

lim<br />

η(O,z ′ )→∞ −<br />

1<br />

η(O, z ′ ) log<br />

<br />

P sup d<br />

0≤t≤1<br />

O <br />

z ′(U(t)) > a = 2 log cosh a.<br />

The proof of the above result is based on the comparison between the law of the hyperbolic bridge<br />

and that of the hyperbolic <strong>Brownian</strong> motion conditioned to tend toward <strong>in</strong>f<strong>in</strong>ity <strong>in</strong> the direction<br />

of z ′ .<br />

3.2.4 Fractional hyperbolic <strong>Brownian</strong> motion<br />

A generalization of the classical hyperbolic <strong>Brownian</strong> motion <strong>in</strong> H 2 is given <strong>in</strong> Lao and Ors<strong>in</strong>gher<br />

(26) where it is considered the time-fractional equation<br />

with α ∈ (0, 1], subject to the <strong>in</strong>itial condition<br />

∂α ∂tα uα = 1<br />

<br />

s<strong>in</strong>h η<br />

s<strong>in</strong>h η<br />

∂<br />

<br />

uα, (3.32)<br />

∂η<br />

uα(η, 0) = δ(η).<br />

We assume that the fractional derivative appear<strong>in</strong>g <strong>in</strong> (3.32) is understand <strong>in</strong> the sense of Dzerbashyan-<br />

Caputo that is, for f ∈ C m ,<br />

dαf =<br />

dtα t<br />

1<br />

Γ(m − α) 0<br />

f (m) (s)<br />

ds, m − 1 < α ≤ m.<br />

(t − s) α+1−m


3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion 44<br />

Proceed<strong>in</strong>g as <strong>in</strong> the proof of Theorem 3.2.1, we get<br />

⎧<br />

⎨<br />

⎩<br />

d α T<br />

dtα = −ωT,<br />

(y2 − 1)F ′′ + 2yF ′ + ωF = 0.<br />

For the first equation, by apply<strong>in</strong>g the Laplace transform to both members, we obta<strong>in</strong> that<br />

where Eα,1 = ∞<br />

k=0<br />

x k<br />

Γ(αk+1)<br />

T (t, ω) = Eα,1(ωt α )<br />

is the Mittag-Leffler function. By mak<strong>in</strong>g use of the representation<br />

(3.25) and choos<strong>in</strong>g aga<strong>in</strong> Q(x) = tanh(πx), we have<br />

uα(η, t) = 2<br />

π<br />

∞<br />

0<br />

xEα,1<br />

<br />

− tα<br />

4 − x2t α<br />

∞<br />

s<strong>in</strong>(xϕ)<br />

dx √ dϕ, (3.33)<br />

η 2 cosh ϕ − 2 cosh η<br />

for 0 < α ≤ 1. We observe that the classical result (3.11) can be derived from (3.33) as a special<br />

case for α = 1 and that for α = 1/2 we get that<br />

∞<br />

u1/2(η, t) = 2<br />

0<br />

τ2 − e 4t<br />

√ p2(η, τ)dτ. (3.34)<br />

2π2t<br />

Formula (3.34) implies that the fractional hyperbolic <strong>Brownian</strong> motion with α = 1/2 is equal <strong>in</strong><br />

distribution to the classical hyperbolic <strong>Brownian</strong> motion at a reflected <strong>Brownian</strong> time.<br />

3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion<br />

3.3.1 Transition function <strong>in</strong> H 3<br />

We consider now the behavior of the transition function of the hyperbolic brownian motion <strong>in</strong> H 3 .<br />

We follow the proof given <strong>in</strong> Ors<strong>in</strong>gher and De Gregorio (32).<br />

Theorem 3.3.1. The normalized heat kernel G3(η, t), t > 0, has the follow<strong>in</strong>g explicit expression<br />

G3(η, t) =<br />

e −t<br />

2 3 (πt) 3/2<br />

Proof<br />

We show that (3.35) satisfies the Cauchy problem (3.5). Let us write<br />

We start by observ<strong>in</strong>g that<br />

h(η, t) = e−t<br />

t 3/2<br />

∂<br />

∂t h(η, t) = e−tt −3/2 η2<br />

−<br />

e 4t<br />

η2<br />

− ηe 4t<br />

, η > 0. (3.35)<br />

s<strong>in</strong>h η<br />

η2<br />

− ηe 4t<br />

s<strong>in</strong>h η .<br />

<br />

η<br />

−1 −<br />

s<strong>in</strong>h η<br />

3 η2<br />

+<br />

2t 4t2 <br />

.


45 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

On the other hand,<br />

and we have<br />

<br />

s<strong>in</strong>h 2 η ∂<br />

∂<br />

∂η h(η, t) = e−tt −3/2 η2<br />

−<br />

e 4t<br />

∂η<br />

<br />

∂<br />

s<strong>in</strong>h<br />

∂η<br />

2 η ∂<br />

∂η<br />

<br />

1<br />

s<strong>in</strong>h η<br />

cosh η<br />

1 − η<br />

s<strong>in</strong>h η<br />

<br />

η2<br />

− ,<br />

2t<br />

<br />

h(η, t) = e −t t −3/2 <br />

η2<br />

−<br />

e 4t s<strong>in</strong>h η − η cosh η − η2<br />

<br />

h(η, t) = e −t t −3/2 <br />

η2<br />

−<br />

e 4t η s<strong>in</strong>h η −1 − 3 η2<br />

+<br />

2t 4t2 2t<br />

<br />

s<strong>in</strong>h η<br />

S<strong>in</strong>ce lim t→0 + h(η, t) = δ(η), we obta<strong>in</strong> that the function h(η, t) a is solution to the Cauchy problem<br />

(3.5). By impos<strong>in</strong>g the condition (3.6) we obta<strong>in</strong> that the normalized heat kernel G3(η, t) is given<br />

by<br />

<strong>in</strong> fact<br />

<br />

<br />

G3(η, t)dv = 4π<br />

=<br />

∞<br />

1<br />

2 √ π<br />

= t−3/2<br />

4 √ π<br />

0<br />

e−tt−3/2 ∞<br />

G3(η, t) =<br />

G3(η, t) s<strong>in</strong>h 2 η dη<br />

2<br />

∞<br />

0<br />

= t−3/2 √ 2t<br />

4 √ π<br />

= t−3/2 √ 2t<br />

4 √ π<br />

0<br />

h(η, t)<br />

23 ,<br />

π3/2 η2<br />

−<br />

ηe 4t (e η − e −η )dη<br />

(η−2t)2<br />

(η+2t)2<br />

−<br />

η(e 4t −<br />

− e 4t )dη<br />

∞<br />

− √ (2t + w<br />

2t<br />

√ w2 −<br />

2t)e<br />

∞<br />

w2 −<br />

2t e 2 dw = 1.<br />

−∞<br />

2 dw −<br />

∞<br />

√ 2t<br />

<br />

.<br />

(w √ 2t − 2t)e<br />

− w2<br />

<br />

2 dw<br />

The probability density p3(η, t) to be at time t at a distance η from the start<strong>in</strong>g po<strong>in</strong>t is then<br />

given by<br />

p3(η, t) = 2π3/2 h(η, t)<br />

√ G3(η, t) =<br />

π<br />

2<br />

2<br />

√ e−t<br />

=<br />

π 2 √ πt3/2 η2<br />

− ηe 4t<br />

s<strong>in</strong>h η .<br />

In what follows we will see that <strong>in</strong> the odd-dimensional hyperbolic spaces it is always possible<br />

to express the heat kernel Gn(η, t) <strong>in</strong> a closed form.<br />

3.3.2 The Millson recursive formula<br />

The follow<strong>in</strong>g recursive formula holds for all n = 1, 2, . . .<br />

Gn+2(η, t) = − e−nt ∂<br />

2π s<strong>in</strong>h η ∂η Gn(η, t). (3.36)<br />

Formula (3.36) is attributed to Millson (unpublished) by Debiard, Gaveau and Mazet <strong>in</strong> (12)<br />

page 369, it is also reported <strong>in</strong> Chavel, see (9) page 151, <strong>in</strong> the general sett<strong>in</strong>g of Riemannian


3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion 46<br />

manifolds and <strong>in</strong> Grigor’yan and Noguchi, see (17) page 664, and Davies, see (10) page 178, for the<br />

hyperbolic spaces. For a proof of Theorem (3.3.2) see Davies (11) Theorem 2.1 and Matsumoto<br />

(28) Corollary 3.2.<br />

Theorem 3.3.2. For n = 1, 2, . . . , the follow<strong>in</strong>g relation holds<br />

Gn+2(η, t) = − e−nt ∂<br />

2π s<strong>in</strong>h η ∂η Gn(η, t).<br />

Remark 3.3.1. We note that, if Gn(η, t) satisfies ∂<br />

∂t Gn =<br />

function<br />

satisfies the follow<strong>in</strong>g differential equation<br />

We first observe that<br />

1<br />

s<strong>in</strong>h n−1 η<br />

Gn+2(η, t) = − e−nt ∂<br />

2π s<strong>in</strong>h η ∂η Gn(η, t),<br />

∂<br />

∂t Gn+2 =<br />

1<br />

s<strong>in</strong>h n+1 η<br />

<br />

∂<br />

s<strong>in</strong>h<br />

∂η<br />

n+1 η ∂<br />

<br />

Gn+2.<br />

∂η<br />

<br />

∂<br />

∂η s<strong>in</strong>h n−1 η ∂<br />

<br />

∂η Gn, then the<br />

∂<br />

∂t Gn+2<br />

= 1<br />

<br />

∂<br />

−<br />

2π ∂t<br />

e−nt ∂<br />

s<strong>in</strong>h η ∂η Gn<br />

<br />

= 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn − 1 e<br />

2π<br />

−nt <br />

∂ ∂<br />

s<strong>in</strong>h η ∂η ∂t Gn<br />

<br />

= 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn − 1 e<br />

2π<br />

−nt <br />

∂ 1<br />

s<strong>in</strong>h η ∂η s<strong>in</strong>h n−1 <br />

∂<br />

s<strong>in</strong>h<br />

η ∂η<br />

n−1 η ∂<br />

∂η Gn<br />

<br />

= 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn − 1 e<br />

2π<br />

−nt <br />

∂ cosh η ∂<br />

(n − 1)<br />

s<strong>in</strong>h η ∂η s<strong>in</strong>h η ∂η Gn + ∂2<br />

<br />

Gn<br />

∂η2 = 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn − 1 e<br />

2π<br />

−nt <br />

1<br />

(1 − n)<br />

s<strong>in</strong>h η s<strong>in</strong>h 2 ∂<br />

η ∂η Gn<br />

cosh η ∂<br />

+ (n − 1)<br />

s<strong>in</strong>h η<br />

2<br />

∂η2 Gn + ∂3<br />

<br />

Gn<br />

∂η3 = 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn<br />

n − 1 e<br />

+<br />

2π<br />

−nt<br />

s<strong>in</strong>h 3 ∂<br />

η ∂η Gn<br />

n − 1 cosh η<br />

− e−nt<br />

2π s<strong>in</strong>h 2 ∂<br />

η<br />

2<br />

∂η2 Gn − 1 e<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η<br />

3<br />

Gn.<br />

∂η3 On the other side we have<br />

1<br />

s<strong>in</strong>h n+1 <br />

∂<br />

s<strong>in</strong>h<br />

η ∂η<br />

n+1 η ∂<br />

<br />

Gn+2<br />

∂η<br />

= 1 1<br />

2π s<strong>in</strong>h n+1 <br />

∂<br />

s<strong>in</strong>h<br />

η ∂η<br />

n+1 η ∂<br />

<br />

−<br />

∂η<br />

e−nt ∂<br />

s<strong>in</strong>h η ∂η Gn<br />

<br />

= 1 1<br />

2π s<strong>in</strong>h n+1 <br />

∂<br />

e<br />

η ∂η<br />

−nt s<strong>in</strong>h n−1 η cosh η ∂<br />

= n − 1<br />

2π e−nt cosh2 η<br />

s<strong>in</strong>h 3 η<br />

− e−nt<br />

2π<br />

= n e−nt<br />

2π<br />

1<br />

s<strong>in</strong>h η<br />

cosh 2 η<br />

s<strong>in</strong>h 3 η<br />

∂<br />

∂η Gn + e−nt<br />

2π<br />

∂3 ∂η3 Gn − n e−nt<br />

2π<br />

= 1 ne<br />

2π<br />

−nt ∂<br />

s<strong>in</strong>h η ∂η Gn<br />

n − 1<br />

+<br />

2π<br />

cosh η<br />

s<strong>in</strong>h 2 η<br />

1<br />

s<strong>in</strong>h η<br />

∂η Gn − e −nt s<strong>in</strong>h n η ∂2<br />

Gn<br />

∂η2 ∂<br />

∂η Gn + e−nt<br />

2π<br />

∂2 Gn<br />

∂η2 ∂<br />

∂η Gn − e−nt<br />

2π s<strong>in</strong>h 3 ∂<br />

η ∂η Gn − (n − 1) e−nt<br />

2π<br />

e −nt<br />

s<strong>in</strong>h 3 η<br />

∂<br />

∂η Gn −<br />

cosh η<br />

s<strong>in</strong>h 2 η<br />

cosh η<br />

s<strong>in</strong>h 2 η<br />

n − 1 cosh η<br />

e−nt<br />

2π s<strong>in</strong>h 2 η<br />

<br />

∂2 Gn<br />

∂η2 ∂2 ∂η2 Gn − e−nt ∂<br />

2π s<strong>in</strong>h η<br />

3<br />

Gn<br />

∂η3 ∂ 2<br />

∂η 2 Gn − 1<br />

2π<br />

e −nt<br />

s<strong>in</strong>h η<br />

∂3 Gn.<br />

∂η3


47 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

We f<strong>in</strong>ally note that, if lim t→0 + Gn(η, 0) = δ(η), then<br />

lim<br />

t→0 + Gn+2(η, t) = lim<br />

t→0<br />

2π s<strong>in</strong>h η<br />

+ − e−nt<br />

∂<br />

∂η Gn(η, t) = δ(η).<br />

In view of Theorem 3.2.1 and Theorem 3.3.1 we have the explicit form for G2(η, t) and G3(η, t),<br />

now, apply<strong>in</strong>g recursively the Millson formula <strong>in</strong> (3.36), we can obta<strong>in</strong> an explicit form for Gn(η, t)<br />

for all n = 1, 2, . . . We start by consider<strong>in</strong>g the odd dimensional case.<br />

• n = 2k + 1<br />

G2k+1(η, t) = − e−[2(k−1)+1]t ∂<br />

2π s<strong>in</strong>h η ∂η G2(k−1)+1 = − e−(2k−1)t<br />

<br />

∂<br />

−<br />

2π s<strong>in</strong>h η ∂η<br />

e−[2(k−2)+1]t ∂<br />

2π s<strong>in</strong>h η ∂η G <br />

2(k−2)+1<br />

= e−2(2k−2)t<br />

(2π) 2<br />

<br />

− 1<br />

2 ∂<br />

G2(k−2)+1 s<strong>in</strong>h η ∂η<br />

= e−2(2k−2)t<br />

(2π) 2<br />

<br />

− 1<br />

2 <br />

∂<br />

−<br />

s<strong>in</strong>h η ∂η<br />

e−[2(k−3)+1]t ∂<br />

2π s<strong>in</strong>h η ∂η G <br />

2(k−3)+1<br />

= e−3(2k−3)t<br />

(2π) 3<br />

<br />

− 1<br />

3 ∂<br />

G2(k−3)+1 s<strong>in</strong>h η ∂η<br />

= e−(k−1)[2k−(k−1)]t<br />

(2π) k−1<br />

<br />

− 1<br />

k−1 ∂<br />

G2[k−(k−1)]+1 s<strong>in</strong>h η ∂η<br />

= e−(k2 −1)t<br />

(2π) k−1<br />

<br />

− 1<br />

k−1 ∂<br />

G3. (3.37)<br />

s<strong>in</strong>h η ∂η<br />

In the even dimensional case we obta<strong>in</strong><br />

• n = 2k + 2<br />

<br />

− 1<br />

s<strong>in</strong>h η<br />

2 ∂<br />

G2(k−2)+2 ∂η<br />

G2k+2(η, t) = − e−2kt ∂<br />

2π s<strong>in</strong>h η ∂η G2k = e−2(2k−1)t<br />

(2π) 2<br />

= e−3(2k−2)t<br />

(2π) 3<br />

<br />

− 1<br />

3 ∂<br />

G2(k−3)+2 s<strong>in</strong>h η ∂η<br />

= e−k(2k−(k−1))t<br />

(2π) k<br />

<br />

− 1<br />

k ∂<br />

G2<br />

s<strong>in</strong>h η ∂η<br />

= e−(k2 +k)t<br />

(2π) k<br />

<br />

− 1<br />

k ∂<br />

G2. (3.38)<br />

s<strong>in</strong>h η ∂η<br />

Formulas (3.37) and (3.38) are obta<strong>in</strong>ed with a different approach <strong>in</strong> Grigor’yan (17), Theorem<br />

1.1, where the proof is obta<strong>in</strong>ed by <strong>in</strong>vestigat<strong>in</strong>g the relation between the heat equation and the<br />

wave equation <strong>in</strong> H n .<br />

3.3.3 Bounds for the hyperbolic heat kernel<br />

Davies and Mandouvalos <strong>in</strong> (11) obta<strong>in</strong>ed sharp upper and lower bounds for the heat kernel Gn(η, t)<br />

on the hyperbolic space H n for all n ≥ 2.<br />

Theorem 3.3.3. For all n ≥ 2 there exists a positive constant cn such that<br />

c −1<br />

n hn(η, t) ≤ Gn(η, t) ≤ cn hn(η, t)


3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion 48<br />

for all t > 0 and η > 0, where<br />

<br />

n −<br />

hn(η, t) = t 2 exp − (n − 1)2t (n − 1)η<br />

− −<br />

4 2<br />

η2<br />

<br />

(1 + η + t)<br />

4t<br />

n−3<br />

2 (1 + η).<br />

The proof of Theorem 3.3.3 is based on the Millson formula (3.36) that relates the heat kernel on<br />

hyperbolic spaces of different dimension. In particular, for n = 3, we obta<strong>in</strong> the uniform estimate<br />

<br />

3 −<br />

G3(η, t) ∼ t 2 exp −t − η − η2<br />

<br />

(1 + η)<br />

4t<br />

that is consistent with (3.35) because of the uniform equivalence<br />

η<br />

s<strong>in</strong>h η ∼ (1 + η)e−η . (3.39)<br />

Recall<strong>in</strong>g that dv = s<strong>in</strong>h n−1 η dη dΩn and s<strong>in</strong>ce, <strong>in</strong> view of (3.39), we have<br />

we obta<strong>in</strong> that<br />

s<strong>in</strong>h n−1 η Gn(η, t) ∼<br />

=<br />

s<strong>in</strong>h n−1 η ∼<br />

If we fix η > 0, Theorem 3.3.3 implies that<br />

n−1 η<br />

e<br />

1 + η<br />

(n−1)η ,<br />

n−1 <br />

η<br />

n −<br />

t 2 exp −<br />

1 + η<br />

(n − 1)2t +<br />

4<br />

(n − 1)η<br />

−<br />

2<br />

η2<br />

<br />

(1 + η + t)<br />

4t<br />

n−3<br />

2 (1 + η)<br />

n−1 <br />

<br />

η<br />

n − [η − (n − 1)t]2<br />

t 2 exp − (1 + η + t)<br />

1 + η<br />

4t<br />

n−3<br />

2 (1 + η). (3.40)<br />

n<br />

3<br />

−<br />

Gn(η, t) ∼ t 2 −<br />

if t → 0 and Gn(η, t) ∼ t 2 e − (n−1)2t 4 if t → ∞.<br />

n − For large t, s<strong>in</strong>ce t 2 (1 + η + t) n−3<br />

2 goes to zero, we have that the dom<strong>in</strong>ant term <strong>in</strong> formula (3.40)<br />

is exp{−[η − (n − 1)t] 2 /4t} and it has its maximum around η = (n − 1)t. Intuitively formula (3.40)<br />

states that the diffusion on the hyperbolic space behaves like a motion with asymptotic velocity<br />

η/t = n − 1 directed away from the start<strong>in</strong>g po<strong>in</strong>t. This behavior is totally different from that of<br />

the <strong>Brownian</strong> motion <strong>in</strong> the Euclidean space.<br />

This result is <strong>in</strong> agreement with Theorem 2.1 <strong>in</strong> Matsumoto (29) where a central limit theorem<br />

for the radial component of the hyperbolic <strong>Brownian</strong> motion is shown. Start<strong>in</strong>g from the stochastic<br />

representation of the hyperbolic distance {η(Z(t), z), t ≥ 0}, where {Z(t), t ≥ 0} is the hyperbolic<br />

<strong>Brownian</strong> motion start<strong>in</strong>g at z ∈ H n , Matsumoto proved the follow<strong>in</strong>g theorem.<br />

Theorem 3.3.4. The probability distribution of the random variable<br />

η(Z(t), z) − (n − 1)t<br />

√ 2t<br />

converges weakly, as t → ∞, to the standard Normal distribution.


49 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

For the particular case n = 2, <strong>in</strong> Lao and Ors<strong>in</strong>gher (26), it is proved that<br />

E{cosh η(Z(t), z)} = e t .<br />

3.3.4 Gruet’s formula for the heat kernel<br />

While the explicit expressions for Gn(η, t) given <strong>in</strong> (3.37) and (3.38) have different forms for odd<br />

and even dimensions, Gruet <strong>in</strong> (18) gives a unique <strong>in</strong>tegral representation for the normalized heat<br />

kernel Gn(η, t) that holds for every dimension n (see Matsumoto (28) and Matsumoto and Yor<br />

(31)). This representation is based on the Yor’s formula on the distribution of the <strong>in</strong>tegral of<br />

geometric <strong>Brownian</strong> motion (see Yor (43)).<br />

Let {B(t), t ≥ 0} be a standard one-dimensional <strong>Brownian</strong> motion start<strong>in</strong>g from the orig<strong>in</strong> and<br />

def<strong>in</strong>ed on the probability space (Ω, F, P ), set B µ (t) = B(t) + µt the correspond<strong>in</strong>g <strong>Brownian</strong><br />

motion with constant drift µ ∈ R. We consider the exponential functional A µ (t) def<strong>in</strong>ed by<br />

A µ (t) =<br />

t<br />

0<br />

exp{2B µ (s)}ds =<br />

t<br />

0<br />

exp{2B(s) + 2µs}ds.<br />

Yor’s formula gives an explicit expression for the density of the jo<strong>in</strong>t distribution of (A µ (t), B µ (t)).<br />

For r > 0 and t > 0 we have<br />

where<br />

P{A µ (t) ∈ du, B µ (t) ∈ dx} = e µx− µ2 x<br />

t 1 + e2x e dudx<br />

2 exp − θ , t , (3.41)<br />

2u u u<br />

θ(r, t) =<br />

re π2<br />

2t<br />

√ 2π 3 t<br />

∞<br />

0<br />

ξ2<br />

−<br />

e 2t −r cosh ξ s<strong>in</strong>h(ξ) s<strong>in</strong><br />

<br />

πξ<br />

dξ. (3.42)<br />

t<br />

Yor (see Yor (44) or Matsumoto and Yor (30)) has shown that the follow<strong>in</strong>g relation holds<br />

Iα(r) =<br />

∞<br />

where Iα(r) is the modified Bessel function.<br />

0<br />

e − α2 t<br />

2 θ(r, t)dt, α > 0, (3.43)<br />

The proof of the Gruet’s formula (3.44) is based on the explicit expression of the jo<strong>in</strong>t density<br />

of (A µ (t), B µ (t)). In this section we will assume that the heat kernel Gn(η, t) is solution to<br />

∂ 1<br />

∂t u = 2∆2u. Theorem 3.3.5. For every n ≥ 2 and t > 0 it holds that<br />

Gn(η, t) =<br />

e−(n−1)2 t<br />

8<br />

π(2π) n<br />

2<br />

∞<br />

n + 1<br />

√ Γ<br />

t 2<br />

0<br />

e π2−ξ 2<br />

2t s<strong>in</strong>h ξ s<strong>in</strong> πξ<br />

t<br />

(cosh η + cosh ξ) n+1 dξ. (3.44)<br />

2<br />

Proof<br />

We denote by ˜ δz ′ and δz ′ the Dirac delta function concentrated at z′ with respect to the volume<br />

element dv = dxdy<br />

yn and the Lebesgue measure dz = dxdy respectively. Apply<strong>in</strong>g the generalized<br />

expectation <strong>in</strong> the sense of Watanabe (see (42) page 16), it is possible to represent the heat kernel


3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion 50<br />

as<br />

Gn(z, z ′ <br />

, t) =<br />

B (n)<br />

˜δz ′(Z(t))dP (n) = (y ′ ) n<br />

<br />

B (n)<br />

δz ′(Z(t))dP (n) , (3.45)<br />

where Z(t) = (X(t), Y (t)) is given by (3.7). Let F µ (t) be the R n−1 -valued random variable<br />

F µ (t) =<br />

t<br />

from (3.7) and (3.45) it follows that<br />

Gn(z, z ′ , t) = (y ′ ) n<br />

<br />

0<br />

=<br />

exp{B µ (s)}dB1(s), ...,<br />

B (n)<br />

′ n <br />

y<br />

y<br />

B (n)<br />

t<br />

0<br />

exp{B µ <br />

(s)}dBn−1(s) ,<br />

δz ′(x + yF µ (t), y exp{B µ (t)})dP (n)<br />

δ “ x ′ −x<br />

y , y′<br />

” µ µ (n)<br />

(F (t), exp{B (t)})dP . (3.46)<br />

y<br />

We note that the conditional distribution of F µ (t) given {B(s), 0 ≤ s ≤ t} is the (n−1)-dimensional<br />

centered Gaussian distribution with covariance matrix A µ (t)In−1, where In−1 is the (n − 1)dimensional<br />

unit matrix. Therefore, tak<strong>in</strong>g the conditional expectation on the last member of<br />

(3.46), we obta<strong>in</strong>, us<strong>in</strong>g the same notation for the Dirac delta function on R n and R, that<br />

Gn(z, z ′ , t) =<br />

=<br />

′ n <br />

y<br />

y<br />

B (n)<br />

′ n−1 <br />

y<br />

y<br />

B (n)<br />

e − ||x′ −x|| 2<br />

2y 2 A µ (t)<br />

(2πA µ (t)) n−1<br />

2<br />

e − ||x′ −x|| 2<br />

2y 2 A µ (t)<br />

(2πA µ (t)) n−1<br />

2<br />

δ y ′ (exp{B<br />

y<br />

µ (t)})dP (n)<br />

δ y<br />

log ′ (B<br />

y<br />

µ (t))dP (n)<br />

where it is used the relation δc(ex ) = c−1δlog c(x), for c > 0. The last generalized expectation is<br />

the conditional expectation under the condition B µ (t) = log y′<br />

y and can be computed by apply<strong>in</strong>g<br />

formula (3.41) and recall<strong>in</strong>g that µ = (n − 1)/2, we have<br />

<br />

P A µ (t) ∈ du| B µ ′ y<br />

(t) = log =<br />

y<br />

It follows that<br />

Gn(z, z ′ ′ n−1 ∞<br />

y<br />

, t) =<br />

y<br />

0<br />

e − ||x′ −x|| 2<br />

2y2u (2πu) n−1<br />

2<br />

by means of the change of variable v = y′<br />

yu<br />

Gn(z, z ′ , t) =<br />

= e−(n−1)2 t<br />

∞<br />

0<br />

<br />

v<br />

n−1 <br />

2<br />

exp −<br />

2π<br />

v<br />

2y ′<br />

8<br />

(2π) n−1<br />

2<br />

∞<br />

0<br />

<br />

y<br />

y ′<br />

n−1<br />

2<br />

e − (n−1)2 <br />

t<br />

8 exp − 1 + (y′ /y) 2<br />

′ y du<br />

θ , t<br />

2u yu u .<br />

<br />

y<br />

y ′<br />

n−1<br />

2<br />

e − (n−1)2 <br />

t<br />

8 exp − 1 + (y′ /y) 2<br />

′ y du<br />

θ , t<br />

2u yu u ,<br />

and recall<strong>in</strong>g formula (2.25), we obta<strong>in</strong><br />

||x ′ − x|| 2 + y 2 + y ′2<br />

y<br />

<br />

− (n − 1)2 <br />

t θ(v, t)<br />

dv<br />

8 v<br />

exp {−v cosh η(z, z ′ )} v n−3<br />

2 θ(v, t)dv. (3.47)<br />

F<strong>in</strong>ally, tak<strong>in</strong>g <strong>in</strong>to account formula (3.42) and chang<strong>in</strong>g the order of <strong>in</strong>tegration it is possible to


51 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

obta<strong>in</strong> the representation (3.44), <strong>in</strong> fact<br />

Gn(η, t) = e− (n−1)2t π2<br />

8 + 2t<br />

(2π) n−1<br />

2<br />

<br />

√ 2tπ 3<br />

= e− (n−1)2t π2<br />

8 + 2t<br />

= e−(n−1)2 t<br />

∞<br />

0<br />

∞<br />

(2π) n−1 √<br />

2 2tπ3 0<br />

∞<br />

8 n + 1<br />

√ Γ<br />

t 2<br />

π(2π) n<br />

2<br />

ξ2<br />

−<br />

e 2t s<strong>in</strong>h(ξ) s<strong>in</strong><br />

ξ2<br />

−<br />

e 2t s<strong>in</strong>h(ξ) s<strong>in</strong><br />

0<br />

πξ<br />

t<br />

πξ<br />

t<br />

<br />

dξ<br />

∞<br />

0<br />

<br />

dξ Γ<br />

e π2−ξ 2<br />

2t s<strong>in</strong>h ξ s<strong>in</strong> πξ<br />

t<br />

(cosh η + cosh ξ) n+1 dξ.<br />

2<br />

v n−1<br />

2 exp{−v[cosh η + cosh ξ]}dv<br />

n + 1<br />

2<br />

<br />

n+1<br />

−<br />

(cosh η + cosh ξ) 2<br />

It is no simple to see directly the co<strong>in</strong>cidence of Gruet’s formula with the classical formulas <strong>in</strong><br />

(3.11) and (3.35) obta<strong>in</strong>ed for the particular cases n = 2 and n = 3, nevertheless it is simple to see<br />

the co<strong>in</strong>cidence of their Laplace transforms <strong>in</strong> t.<br />

We first compute the Laplace transform of Gruet’s formula (3.44). Start<strong>in</strong>g from formula (3.47)<br />

we have<br />

∞<br />

0<br />

α2 −<br />

e 2 t Gn(η, t)dt =<br />

=<br />

recall<strong>in</strong>g formula (3.43) we get<br />

∞<br />

0<br />

∞<br />

0<br />

1<br />

2 t e− (n−1)2t ∞<br />

8<br />

α2 −<br />

e<br />

(2π) n−1<br />

2<br />

α2 −<br />

e 2 t Gn(η, t)dt =<br />

=<br />

=<br />

∞<br />

0<br />

1<br />

(2π) n−1<br />

2<br />

1<br />

(2π) n−1<br />

2<br />

2<br />

(2π) n<br />

2<br />

(2π) n−1<br />

2<br />

0<br />

exp{−v cosh η}v n−3<br />

2 θ(v, t)dv dt<br />

v n−3<br />

2 exp{−v cosh η}dv<br />

∞<br />

0<br />

2<br />

π<br />

n−2<br />

−iπ e 2<br />

s<strong>in</strong>h n−2<br />

2 r<br />

∞<br />

v n−3<br />

2 exp{−v cosh η}dv I q<br />

n−2<br />

−iπ e 2<br />

s<strong>in</strong>h n−2<br />

2 r<br />

n−2<br />

2 Qq n−2<br />

2 Qq α 2 + (n−1)2<br />

4<br />

0<br />

α 2 + (n−1)2<br />

4<br />

1 − 2<br />

e −<br />

„<br />

α 2<br />

«<br />

(n−1)2<br />

2 + 8 t<br />

θ(v, t)dt,<br />

1 − 2<br />

(cosh η)<br />

α 2 + (n−1)2<br />

4<br />

(cosh η)<br />

where Iα(v) is the modified Bessel function and Q m n (x) is the associated Legendre function of the<br />

second k<strong>in</strong>d. In the second equality we have used formula 6.622.3 <strong>in</strong> Gradshteyn and Ryzhik (16).<br />

For n = 2 we obta<strong>in</strong><br />

∞<br />

For n = 3 we have the simple expression<br />

0<br />

α2 −<br />

e 2 t G2(η, t)dt = 1<br />

∞<br />

0<br />

π Q√ α 2 + 1<br />

4<br />

α2 −<br />

e 2 t G3(η, t)dt = e−η√ α2 +1<br />

2π s<strong>in</strong>h η<br />

1 (cosh η).<br />

− 2<br />

s<strong>in</strong>ce, <strong>in</strong> view of Gradshteyn and Ryzhik (16) formula 8.754.4, we have Q 1<br />

2<br />

ν− 1<br />

2<br />

(v)<br />

(cosh r) = π<br />

2 s<strong>in</strong>h ν ie−νr .<br />

On the other hand, start<strong>in</strong>g from the representation of G2(η, t) given <strong>in</strong> (3.11) and replac<strong>in</strong>g t


3.3 Multidimensional hyperbolic <strong>Brownian</strong> motion 52<br />

with t/2, we obta<strong>in</strong><br />

∞<br />

0<br />

α2 −<br />

e 2 t G2(η, t)dt =<br />

where we have used that ∞<br />

0<br />

=<br />

√<br />

2<br />

(2π) 3/2<br />

√<br />

2<br />

(2π) 3/2<br />

1<br />

=<br />

π √ 2<br />

= 1<br />

e −ta− b t<br />

t 3/2<br />

∞<br />

η<br />

∞<br />

η<br />

∞<br />

η<br />

∞<br />

ϕ<br />

√ dϕ<br />

cosh ϕ − cosh η<br />

√<br />

ϕ<br />

2π<br />

√ dϕ<br />

cosh ϕ − cosh η ϕ e−2<br />

e −ϕ√ α 2 + 1<br />

4<br />

√ cosh ϕ − cosh η dϕ<br />

π Q√ α2 + 1 1 (cosh η)<br />

4 − 2<br />

0<br />

t −<br />

e 2(α 2 + 1<br />

4) − e ϕ2<br />

2t<br />

t 3/2<br />

r “ α 2<br />

2<br />

”<br />

1 ϕ2 + 8 2<br />

dt = π<br />

b e−2√ ab and the last step follows from formula 3.544 of<br />

Gradshteyn and Ryzhik (16) and Qn(x) is the Legendre function of the second k<strong>in</strong>d. Concern<strong>in</strong>g<br />

the Laplace transform of G3(η, t) given <strong>in</strong> formula (3.35), when t is replaced by t/2, we have<br />

∞<br />

0<br />

α2 −<br />

e 2 t G3(η, t)dt =<br />

1<br />

(2π) 3/2<br />

η<br />

∞ t − e<br />

s<strong>in</strong>h η 0<br />

= e−η√ α 2 +1<br />

2π s<strong>in</strong>h η .<br />

2 (α2 +1)− η2<br />

2t<br />

Remark 3.3.2. We note that the Millson formula (3.36) follows from the representation (3.44) of<br />

the heat kernel. If we differentiate both sides of (3.44) with respect to η we have<br />

∂<br />

∂η Gn(η, t)<br />

= e−(n−1)2 t<br />

8<br />

π √ t(2π) n<br />

2<br />

∞<br />

n + 1<br />

Γ<br />

2<br />

= − s<strong>in</strong>h η 2πe nt<br />

2 Gn+2(η, t).<br />

0<br />

e π2 −ξ 2<br />

2t s<strong>in</strong>h ξ s<strong>in</strong><br />

<br />

πξ<br />

−<br />

t<br />

3.3.5 <strong>Hyperbolic</strong> branch<strong>in</strong>g <strong>Brownian</strong> motion<br />

t 3/2<br />

<br />

n + 1<br />

2<br />

dt<br />

s<strong>in</strong>h η<br />

dt<br />

(cosh η + cosh ξ) n+3<br />

2<br />

The hyperbolic branch<strong>in</strong>g <strong>Brownian</strong> motion is a branch<strong>in</strong>g diffusion <strong>in</strong> which <strong>in</strong>dividual particles<br />

follow <strong>in</strong>dependent <strong>Brownian</strong> paths <strong>in</strong> the hyperbolic space H n and undergo b<strong>in</strong>ary fission at<br />

exponentially distributed random times <strong>in</strong>dependent of the motion.<br />

Lalley and Sellke (25) considered the hyperbolic branch<strong>in</strong>g <strong>Brownian</strong> motion <strong>in</strong> the hyperbolic<br />

plane when the rate λ of fission is assumed to be constant. Unlike the branch<strong>in</strong>g <strong>Brownian</strong> motion<br />

<strong>in</strong> the Euclidean plane, the hyperbolic branch<strong>in</strong>g <strong>Brownian</strong> motion <strong>in</strong> H 2 exhibits a phase transition<br />

<strong>in</strong> λ. For λ ≤ 1/8 the number of particles <strong>in</strong> any compact region of H 2 is eventually zero with<br />

probability one, but for λ > 1/8 the number of particles <strong>in</strong> any open set grows to <strong>in</strong>f<strong>in</strong>ity with<br />

probability one. This phase transition manifests itself <strong>in</strong> the behavior of the process at <strong>in</strong>f<strong>in</strong>ity.<br />

Let Λ be the set of all limit po<strong>in</strong>ts <strong>in</strong> the boundary of H 2 (it is irrelevant whether ∂H 2 is viewed<br />

as the real axis of the half-plane model or the unit circle of the disk model) of particle trails. In<br />

(25), Theorem 1 and Proposition 10, it is shown that for λ ∈ (0, 1/8] the Hausdorff dimension of<br />

Λ is, with probability one, equal to<br />

1<br />

2 (1 − √ 1 − 8λ)<br />


53 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

and that for λ > 1/8 the complement of Λ <strong>in</strong> ∂H 2 has Lebesgue measure zero. Note that as λ<br />

goes to 1/8 the Hausdorff dimension <strong>in</strong>creases cont<strong>in</strong>uously to 1/2 and not to 1, so the Hausdorff<br />

dimension of Λ is discont<strong>in</strong>uous at the critical value λ = 1/8.<br />

Karpelevich, Pechersky and Suhov (21) generalized the Lalley and Sellke result for an hyperbolic<br />

branch<strong>in</strong>g <strong>Brownian</strong> motion <strong>in</strong> H n , with n ≥ 2. Let v = λ(χ − 1) be the fission potential where<br />

χ > 1 is the mean number of offspr<strong>in</strong>g <strong>in</strong> a s<strong>in</strong>gle fission. The Hausdorff dimension h(Λ) of the<br />

limit<strong>in</strong>g set Λ on ∂H n is a function of v:<br />

⎧<br />

⎨ 1<br />

2<br />

h(Λ) =<br />

⎩<br />

(n − 1 − (n − 1) − 8v), if 0 ≤ v ≤ (n−1)2<br />

8 ,<br />

n − 1, if v > (n−1)2<br />

.<br />

It follows that h(Λ) monotonically <strong>in</strong>creases with v from 0 to (n − 1)/2 and then jumps to n − 1,<br />

we have that Λ = ∂H n for v > (n − 1) 2 /8. Kelbert and Suhov <strong>in</strong> (22) considered the case of a<br />

variable fission mechanism.<br />

3.A Appendix<br />

Proposition 3.A.1. If z = (x, y) and z ′ = (x ′ , y ′ ) are <strong>in</strong> H n , ∆n is given by formula (3.1) and<br />

η(z, z ′ ) is given by (2.25), we have that<br />

∆n η(z, z ′ ) = (n − 1) coth η(z, z ′ ).<br />

Proof<br />

S<strong>in</strong>ce coth(arcosh(x)) = x<br />

√ x 2 −1 and s<strong>in</strong>ce, <strong>in</strong> view of formula (2.25), we have<br />

we have to prove that<br />

∆n η(z, z ′ ) = (n − 1)<br />

η(z, z ′ ) = arcosh |x − x′ | 2 + y2 + y ′2<br />

2yy ′ ,<br />

|x − x ′ | 2 + y 2 + y ′2<br />

<br />

|x − x ′ | 2 + (y + y ′ ) 2 |x − x ′ | 2 + (y − y ′ .<br />

) 2<br />

Us<strong>in</strong>g the follow<strong>in</strong>g notation δ+ = |x − x ′ | 2 + (y + y ′ ) 2 and δ− = |x − x ′ | 2 + (y − y ′ ) 2 , we obta<strong>in</strong><br />

that<br />

∂<br />

∂y η(z, z′ ) = y2 − y ′2 − |x − x ′ | 2<br />

y ,<br />

δ− δ+<br />

∂2 ∂y2 η(z, z′ ) = 2y2δ− δ+ − (y2 − y ′2 − |x − x ′ | 2 ) ∂<br />

y2δ−δ+ =<br />

2<br />

−<br />

δ− δ+<br />

y2 − y ′2 − |x − x ′ | 2<br />

y2δ− δ+<br />

8<br />

∂y (y δ− δ+)<br />

− y2 − y ′2 − |x − x ′ | 2<br />

δ 3/2<br />

− δ 1/2<br />

+<br />

+ y2 − y ′2 − |x − x ′ | 2<br />

y δ 3/2<br />

− δ 1/2<br />

y<br />

+<br />

′ − y2 − y ′2 − |x − x ′ | 2<br />

δ 1/2<br />

− δ 3/2<br />

−<br />

+<br />

y2 − y ′2 − |x − x ′ | 2<br />

y δ 1/2<br />

− δ 3/2<br />

+<br />

y ′ ,


3.A Appendix 54<br />

s<strong>in</strong>ce<br />

∂<br />

∂y (y δ− δ+) = δ− δ+ + y(y − y′ ) δ+<br />

δ−<br />

On the other side, for i = 1, . . . , n − 1, we have<br />

s<strong>in</strong>ce<br />

∂<br />

∂xi<br />

∂ 2<br />

∂x 2 i<br />

η(z, z ′ ) = 2(xi − x ′ i )<br />

<br />

δ− δ+<br />

η(z, z ′ <br />

δ− δ+ − (xi − x<br />

) = 2<br />

′ ∂<br />

i ) ∂xi (δ− δ+)<br />

∂<br />

∂xi<br />

So, f<strong>in</strong>ally, we obta<strong>in</strong> that<br />

∆nη(z, z ′ ) =<br />

δ−δ+<br />

( δ− δ+) = δ1/2 +<br />

2(n − 1)y2<br />

δ− δ+<br />

2δ 1/2<br />

−<br />

=<br />

2(xi − x ′ i) + δ1/2 −<br />

2δ 1/2<br />

+<br />

+ y(y + y′ ) δ−<br />

.<br />

δ+<br />

2<br />

δ− δ+<br />

2(xi − x ′ i).<br />

− 2y 2 δ+<br />

n−1<br />

+ δ−<br />

<br />

(xi − x ′ i) 2 + 2y2<br />

δ 3/2<br />

+ δ 3/2<br />

−<br />

i=1<br />

δ− δ+<br />

− 2(xi − x ′ i) 2 δ+ + δ−<br />

,<br />

δ 3/2<br />

− δ 3/2<br />

+<br />

− y2 − y ′2 − |x − x ′ | 2<br />

δ− δ+<br />

− y2 − y ′2 − |x − x ′ | 2<br />

δ 3/2<br />

− δ 1/2<br />

y<br />

+<br />

2 + y2 − y ′2 − |x − x ′ | 2<br />

δ 3/2<br />

− δ 1/2<br />

yy<br />

+<br />

′ − y2 − y ′2 − |x − x ′ | 2<br />

δ 1/2<br />

− δ 3/2<br />

y<br />

+<br />

2<br />

− y2 − y ′2 − |x − x ′ | 2<br />

δ 1/2<br />

− δ 3/2<br />

yy<br />

+<br />

′ − (n − 2) y2 − y ′2 − |x − x ′ | 2<br />

<br />

δ− δ+<br />

= 2(n − 1)y2 + (n − 1)(−y 2 + y ′2 + |x − x ′ | 2 ) + 2y 2<br />

δ− δ+<br />

+ −y2 (y2 − y ′2 − |x − x ′ | 2 ) + yy ′ (y2 − y ′2 − |x − x ′ | 2 ) − 2y2 |x − x ′ | 2<br />

δ 3/2<br />

− δ 1/2<br />

+<br />

− −y2 (y2 − y ′2 − |x − x ′ | 2 ) + yy ′ (y2 − y ′2 − |x − x ′ | 2 ) − 2y2 |x − x ′ | 2<br />

δ 3/2<br />

+ δ 1/2<br />

−<br />

= (n − 1) |x − x′ | 2 + y 2 + y ′2<br />

δ− δ+<br />

and the proposition is proved. <br />

Proposition 3.A.2. If f is a smooth function on R, we have<br />

Proof<br />

We have<br />

∆nf(η(z, z ′ )) = y 2<br />

n−1<br />

<br />

∆nf(η) = f ′′ (η) + (n − 1) coth η f ′′ (η).<br />

∂ 2<br />

∂x<br />

i=1<br />

2 i<br />

f(η(z, z ′ )) + ∂2<br />

∂y 2 f(η(z, z′ ))<br />

<br />

− (n − 2)y ∂<br />

∂y f(η(z, z′ ))


55 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

where<br />

∂ ∂f ∂η<br />

f(η) =<br />

∂y ∂η ∂y ,<br />

We have than<br />

∆nf(η(z, z ′ )) = y 2<br />

In fact<br />

n−1 <br />

<br />

∂η<br />

i=1<br />

∂xi<br />

∂ 2<br />

∂y 2 f(η) = ∂2 f<br />

∂η 2<br />

n−1<br />

<br />

= ∂2f y2<br />

∂η2 = ∂2f y2<br />

∂η2 <br />

∂ 2 f<br />

∂η<br />

i=1<br />

2<br />

<br />

n−1<br />

∂η<br />

∂xi<br />

<br />

<br />

∂η<br />

i=1<br />

<br />

n−1<br />

∂xi<br />

<br />

<br />

∂η<br />

i=1<br />

= ∂2f ∂f<br />

+<br />

∂η2 ∂η ∆nη.<br />

2<br />

+<br />

2 ∂η<br />

∂y<br />

=<br />

∂xi<br />

2<br />

2 ∂η<br />

+<br />

∂y<br />

∂f<br />

∂η<br />

2<br />

2<br />

∂2η ,<br />

∂y2 ∂ 2<br />

∂x 2 i<br />

f(η) = ∂2f ∂η2 2 ∂η<br />

+<br />

∂xi<br />

∂f<br />

∂η<br />

∂2η ∂x2 .<br />

i<br />

+ ∂f ∂<br />

∂η<br />

2η ∂x2 <br />

+<br />

i<br />

∂2f ∂η2 2 ∂η<br />

+<br />

∂y<br />

∂f ∂<br />

∂η<br />

2η ∂y2 <br />

− (n − 2)y ∂f ∂η<br />

∂η ∂y<br />

<br />

2<br />

∂η<br />

+ +<br />

∂y<br />

∂f<br />

<br />

y<br />

∂η<br />

2<br />

<br />

n−1 ∂<br />

i=1<br />

2η ∂x2 +<br />

i<br />

∂2η ∂y2 <br />

− (n − 2)y ∂η<br />

<br />

∂y<br />

<br />

2<br />

∂η<br />

+ +<br />

∂y<br />

∂f<br />

∂η ∆nη<br />

n−1 <br />

4<br />

(xi − x<br />

δ+δ−<br />

i=1<br />

′ i) 2 + (y2 − y ′2 − |x − x ′ | 2 ) 2<br />

y2δ+δ− = y4 + y ′4 + |x − x ′ | 4 − 2y 2 y ′2 + 2y 2 |x − x ′ | 2 + 2y ′ |x − x ′ | 2<br />

y 2 δ+δ−<br />

= 1<br />

.<br />

y2 In view of Proposition 3.A.1 we obta<strong>in</strong><br />

and the proposition is proved. <br />

∆nf(η(z, z ′ )) = ∂2f ∂f<br />

+<br />

∂η2 ∂η ∆nη = ∂2f ∂f<br />

+ (n − 1) coth η<br />

∂η2 ∂η<br />

Proposition 3.A.3. The hyperbolic coord<strong>in</strong>ates (η, α) and the cartesian coord<strong>in</strong>ates (x, y) <strong>in</strong> H 2<br />

satisfy the follow<strong>in</strong>g relationships<br />

Proof<br />

We first observe that<br />

x =<br />

y =<br />

and from these relations we obta<strong>in</strong><br />

s<strong>in</strong>h η cos α<br />

cosh η−s<strong>in</strong>h η s<strong>in</strong> α , η > 0,<br />

1<br />

cosh η−s<strong>in</strong>h η s<strong>in</strong> α , π<br />

π<br />

− 2 < α ≤ 2 .<br />

(x − x0) 2 + y 2 = r 2 , x 2 0 + 1 = r 2 , x0 = r s<strong>in</strong> α,<br />

tan α = x2 + y 2 − 1<br />

2x<br />

S<strong>in</strong>ce, <strong>in</strong> view of (2.25), we have<br />

(3.48)<br />

and x = tan α ±<br />

<br />

tan2 α + 1 − y2 . (3.49)<br />

cosh η = x2 + y2 − 1<br />

, (3.50)<br />

2y


3.A Appendix 56<br />

substitut<strong>in</strong>g (3.49) <strong>in</strong> (3.50) we have<br />

that is<br />

which implies<br />

2y cosh η = y 2 <br />

+ 1 + tan α ± tan2 α − y2 2 + 1<br />

= 2 + 2 tan 2 <br />

α ± 2 tan α tan2 α1 − y2 ,<br />

y cosh η − 1 − s<strong>in</strong>2 α<br />

cos2 α<br />

=<br />

<br />

s<strong>in</strong> α<br />

± tan<br />

cos α<br />

2 α + 1 − y2 ,<br />

(y cosh η cos 2 α − 1) 2 = s<strong>in</strong> 2 α cos 2 α(tan 2 α + 1 − y 2 ),<br />

y 2 [cosh 2 η cos 4 α + s<strong>in</strong> 2 α cos 2 α] + cos 2 α − 2y cosh η cos 2 α = 0,<br />

y 2 [cosh 2 + s<strong>in</strong> 2 (1 − cos 2 η)] + 1 − 2y cosh η = 0,<br />

[y cosh η − 1] 2 − y 2 s<strong>in</strong> 2 α s<strong>in</strong>h 2 η = 0,<br />

(y cosh η − 1 − y s<strong>in</strong> α s<strong>in</strong>h η)(y cosh η − 1 + y s<strong>in</strong> α s<strong>in</strong>h η) = 0.<br />

The last equation and formula (3.49) imply both relations <strong>in</strong> (3.9). <br />

Proposition 3.A.4. The hyperbolic Laplacian <strong>in</strong> cartesian coord<strong>in</strong>ates<br />

∆2 = y 2<br />

2 ∂ ∂2<br />

+<br />

∂x2 ∂y2 <br />

is expressed <strong>in</strong> hyperbolic coord<strong>in</strong>ates (η, α) by means of the differential operator<br />

Proof<br />

By deriv<strong>in</strong>g (3.9) we get<br />

and<br />

∂2 ∂2<br />

=<br />

∂η2 ∂x2 ∂ ∂<br />

=<br />

∂η ∂x<br />

∆2 = 1<br />

<br />

∂<br />

s<strong>in</strong>h η<br />

s<strong>in</strong>h η ∂η<br />

∂<br />

<br />

+<br />

∂η<br />

1<br />

s<strong>in</strong>h 2 η<br />

∂2 .<br />

∂α2 cos α<br />

∂ − s<strong>in</strong>h η + cosh η s<strong>in</strong> α<br />

2 +<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) ∂y<br />

2 , (3.51)<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α)<br />

∂ ∂ s<strong>in</strong>h<br />

=<br />

∂α ∂x<br />

2 η − s<strong>in</strong>h η cosh η s<strong>in</strong> α<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 2<br />

∂ cos α s<strong>in</strong>h η<br />

+<br />

∂y (cosh η − s<strong>in</strong>h η s<strong>in</strong> α)<br />

cos 2 α<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α)<br />

∂2<br />

4 +<br />

∂y2 +2 ∂2 cos α (− s<strong>in</strong>h η + cosh η s<strong>in</strong> α)<br />

∂x∂y (cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

+ ∂ −2 cos α (s<strong>in</strong>h η − cosh η s<strong>in</strong> α)<br />

∂x (cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 3 + ∂<br />

∂y<br />

(− s<strong>in</strong>h η + cosh η s<strong>in</strong> α) 2<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

− cos2 α + (s<strong>in</strong>h η − cosh η s<strong>in</strong> α) 2<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 3 ,(3.52)<br />

2 ,


57 <strong>Hyperbolic</strong> <strong>Brownian</strong> Motion<br />

∂2 ∂2<br />

=<br />

∂α2 ∂x2 s<strong>in</strong>h 2 η (s<strong>in</strong>h η − cosh η s<strong>in</strong> α) 2<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4 + ∂2<br />

∂y2 s<strong>in</strong>h 2 η cos2 α<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

+2 ∂2 s<strong>in</strong>h<br />

∂x∂y<br />

2 η cos α (s<strong>in</strong>h η − cosh η s<strong>in</strong> α)<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

+ ∂ s<strong>in</strong>h η cos α<br />

∂x<br />

s<strong>in</strong>h 2 η − 1 − s<strong>in</strong>h η cosh η s<strong>in</strong> α <br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 3<br />

+ ∂ s<strong>in</strong>h η<br />

∂y<br />

− s<strong>in</strong> α cosh η + s<strong>in</strong>h η + s<strong>in</strong>h η cos2 α <br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 3 . (3.53)<br />

By multiply<strong>in</strong>g (3.52) by s<strong>in</strong>h 2 η and then summ<strong>in</strong>g (3.53) we obta<strong>in</strong> that<br />

∂ 2<br />

∂η2 s<strong>in</strong>h2 η + ∂2<br />

∂α<br />

2 = ∂2<br />

∂x2 s<strong>in</strong>h2 η cos2 α + (s<strong>in</strong>h η − cosh η s<strong>in</strong> α) 2<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

+ ∂2<br />

∂y2 s<strong>in</strong>h2 η cos2 α + (s<strong>in</strong>h η − cosh η s<strong>in</strong> α) 2<br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 4<br />

+ ∂ s<strong>in</strong>h η cos α<br />

∂x<br />

− cosh 2 η + s<strong>in</strong>h η cosh η s<strong>in</strong> α <br />

(cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 3<br />

+ ∂ s<strong>in</strong>h η cosh η (s<strong>in</strong>h η − cosh η s<strong>in</strong> α)<br />

∂y (cosh η − s<strong>in</strong>h η s<strong>in</strong> α) 2<br />

= s<strong>in</strong>h 2 η y 2<br />

2 ∂ ∂2<br />

+<br />

∂x2 ∂y2 <br />

2 ∂<br />

− s<strong>in</strong>h η cosh η cos α y<br />

∂x<br />

2 ∂<br />

+ s<strong>in</strong>h η cosh η(s<strong>in</strong>h η − cosh η s<strong>in</strong> α) y<br />

∂y<br />

= s<strong>in</strong>h 2 η y 2<br />

2 ∂ ∂2<br />

+<br />

∂x2 ∂y2 <br />

− s<strong>in</strong>h η cosh η ∂<br />

∂η .<br />

Where <strong>in</strong> the last equality we have used formula (3.51).


3.A Appendix 58


Chapter 4<br />

Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a<br />

Pythagorean Compass<br />

4.1 Description of the planar random motion on H 2<br />

We start our analysis by consider<strong>in</strong>g a particle located at the orig<strong>in</strong> O of H 2 . The particle <strong>in</strong>itially<br />

moves on the half-circumference with center at (0, 0) and radius 1. The motion of the particle<br />

develops on the geodesic l<strong>in</strong>es represented by half-circles with the center located on the x axis.<br />

Changes of direction are governed by a homogeneous Poisson process of rate λ.<br />

At the occurrence of the first Poisson event, the particle starts mov<strong>in</strong>g on the circumference<br />

orthogonal to the previous one.<br />

After hav<strong>in</strong>g reached the po<strong>in</strong>t P2, where the second Poisson event happened, the particle<br />

cont<strong>in</strong>ues its motion on the circumference orthogonal to that jo<strong>in</strong><strong>in</strong>g O with P2 (see Figure 4.1).<br />

In general, at the n-th Poisson event, the particle is located at the po<strong>in</strong>t Pn and starts mov<strong>in</strong>g<br />

on the circumference orthogonal to the geodesic curve pass<strong>in</strong>g through Pn and the orig<strong>in</strong> O (consult<br />

aga<strong>in</strong> Figure 4.1).<br />

At each Poisson event the particle moves from the reached position P clockwise or counterclockwise<br />

(with probability 1<br />

2 ) on the circumference orthogonal to the geodesic l<strong>in</strong>e pass<strong>in</strong>g through<br />

P and O.<br />

The hyperbolic length of the arc run by the particle dur<strong>in</strong>g the <strong>in</strong>ter-time between two successive<br />

changes of direction, occurr<strong>in</strong>g at tk−1 and tk respectively, is given by c(tk − tk−1), with k ≥ 1 and<br />

t0 = 0. The velocity c is assumed to be the constant hyperbolic velocity<br />

c = ds<br />

dt<br />

<br />

1 dx2 + dy2 =<br />

y dt2 .<br />

The Cartesian coord<strong>in</strong>ates of the po<strong>in</strong>ts Pk, where the changes of direction occur, can be explicitly<br />

evaluated, but they are not important <strong>in</strong> our analysis because we study only the evolution<br />

of the hyperbolic distance from the orig<strong>in</strong> of the mov<strong>in</strong>g particle.<br />

The construction outl<strong>in</strong>ed above shows that the arcs OPk−1, Pk−1Pk, and OPk form right


4.1 Description of the planar random motion on H 2 60<br />

Figure 4.1: In the first three figures a sample path where the particle chooses the outward direction<br />

is depicted. In the last one a trajectory with one step mov<strong>in</strong>g towards the orig<strong>in</strong> is depicted.<br />

triangles with the vertex of the right angle at Pk−1.<br />

In force of the <strong>Hyperbolic</strong> Pythagorean Theorem we have that<br />

cosh d(OPk) = cosh d(OPk−1) cosh d(Pk−1Pk).<br />

The hyperbolic distance η(t) of the mov<strong>in</strong>g po<strong>in</strong>t Pt after n changes of direction is thus given by<br />

cosh η(t) = cosh d(OPt)<br />

= cosh d(PnPt) cosh d(OPn)<br />

n<br />

= cosh c(t − tn) cosh c(tk − tk−1)<br />

=<br />

n+1 <br />

k=1<br />

k=1<br />

cosh c(tk − tk−1), (4.1)<br />

where t0 = 0 and tn+1 = t. The <strong>in</strong>stants tk, k = 0, 1, . . . , n are uniformly distributed <strong>in</strong> the set<br />

T = {0 = t0 < t1 < · · · < tk < · · · < tn < tn+1 = t}.<br />

This means that cosh η(t), def<strong>in</strong>ed <strong>in</strong> (4.1), can be viewed as the hyperbolic distance from O of<br />

the mov<strong>in</strong>g particle for fixed time po<strong>in</strong>ts of the underly<strong>in</strong>g Poisson process and for a fixed number<br />

N(t) = n of changes of direction.<br />

We remark that the geodesic distance (4.1) depends on how much time the particle spends on<br />

each geodesic curve (but not on the chosen direction). Of course, (4.1) depends on the number n<br />

of changes of direction and on the speed c of the mov<strong>in</strong>g particle, as well.<br />

The set of possible positions at different times t is depicted <strong>in</strong> Figure 4.2. The vertices A and<br />

B are reached when the particle never changes direction, whereas C and D are reached if the<br />

deviation occurs immediately after the start.<br />

The ensemble of po<strong>in</strong>ts hav<strong>in</strong>g the same hyperbolic distance from O at time t, forms the circle


61 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

0.8 1.0 1.2 1.4<br />

0 2 4 6 8<br />

A<br />

C<br />

D<br />

−0.4 −0.2 0.0 0.2 0.4<br />

(a) N=2, t=6<br />

C<br />

A D<br />

−4 −2 0 2 4<br />

B<br />

(c) N=2, t=40<br />

B<br />

0 1 2 3 4<br />

0 2 4 6 8 10 12 14<br />

A<br />

C<br />

D<br />

−2 −1 0 1 2<br />

(b) N=2, t=25<br />

C<br />

A D<br />

−6 −4 −2 0 2 4 6<br />

B<br />

(d) N=2, t=50<br />

Figure 4.2: The set of all possible po<strong>in</strong>ts reachable by the process for different values of t is drawn.<br />

In each doma<strong>in</strong> a trajectory of the process, with c = 0.05 and N(t) = 2, is depicted.<br />

with center C = (0, cosh η(t)) and radius s<strong>in</strong>h η(t). S<strong>in</strong>ce<br />

cosh η(t) =<br />

n+1 <br />

k=1<br />

cosh c(tk − tk−1),<br />

the ord<strong>in</strong>ate of the center C is obta<strong>in</strong>ed by successively multiply<strong>in</strong>g the ord<strong>in</strong>ates of the centers<br />

of equally distant po<strong>in</strong>ts at each step. For the radius, however, such a f<strong>in</strong>e <strong>in</strong>terpretation is not<br />

possible (the radii do not exhibit the same multiplicative behavior) but nevertheless we will study<br />

their product<br />

because<br />

n+1 <br />

k=1<br />

s<strong>in</strong>h η(t) ≥<br />

s<strong>in</strong>h c(tk − tk−1) (4.2)<br />

n+1 <br />

k=1<br />

s<strong>in</strong>h c(tk − tk−1)<br />

and (4.2) represents a lower bound for the circle of equally distant po<strong>in</strong>ts at time t.<br />

B


4.2 Equations related to the mean hyperbolic distance 62<br />

4.2 Equations related to the mean hyperbolic distance<br />

In this section we study the conditional and unconditional mean values of the hyperbolic distance<br />

η(t). Our first result concerns the derivation of the equations satisfied by the mean values<br />

where<br />

and by<br />

En(t) = E{cosh η(t)|N(t) = n} (4.3)<br />

= n!<br />

t n<br />

t<br />

0<br />

dt1<br />

= n!<br />

In(t),<br />

tn In(t) =<br />

t<br />

0<br />

t<br />

t1<br />

t<br />

dt1 · · ·<br />

dt2 · · ·<br />

tn−1<br />

t<br />

dtn<br />

tn−1<br />

n+1 <br />

k=1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh c(tk − tk−1)<br />

cosh c(tk − tk−1),<br />

E(t) = E{cosh η(t)} (4.4)<br />

∞<br />

= E{cosh η(t)|N(t) = n}P{N(t) = n}<br />

n=0<br />

= e −λt<br />

∞<br />

λ n In(t).<br />

n=0<br />

At first, we state the follow<strong>in</strong>g result concern<strong>in</strong>g the evaluation of the <strong>in</strong>tegrals In(t), n ≥ 1.<br />

Lemma 4.2.1. The functions<br />

In(t) =<br />

t<br />

0<br />

dt1<br />

t<br />

t1<br />

t<br />

dt2 · · ·<br />

tn−1<br />

dtn<br />

n+1 <br />

with t0 = 0 and tn+1 = t, satisfy the difference-differential equations<br />

where t > 0, n ≥ 1, and I0(t) = cosh ct.<br />

Proof<br />

We first note that<br />

d<br />

dt In =<br />

t<br />

t<br />

dt1 · · ·<br />

0<br />

t<br />

+c<br />

0<br />

= In−1 + c<br />

tn−2<br />

t<br />

dt1 · · ·<br />

t<br />

0<br />

k=1<br />

d 2<br />

dt 2 In = d<br />

dt In−1 + c 2 In<br />

dtn−1<br />

dtn<br />

n<br />

cosh c(tk − tk−1)<br />

k=1<br />

tn−1 k=1<br />

t<br />

dt1 · · ·<br />

tn−1<br />

cosh c(tk − tk−1), (4.5)<br />

n<br />

cosh c(tk − tk−1) s<strong>in</strong>h c(t − tn)<br />

dtn<br />

n<br />

cosh c(tk − tk−1) s<strong>in</strong>h c(t − tn)<br />

k=1<br />

(4.6)


63 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

and therefore<br />

<br />

d 2<br />

dt 2 In = d<br />

dt In−1 + c 2<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh c(tk − tk−1)<br />

= d<br />

dt In−1 + c 2 In. (4.7)<br />

In view of Lemma 4.2.1 we can prove also the follow<strong>in</strong>g:<br />

Theorem 4.2.2. The mean value E(t) = E{cosh η(t)} satisfies the second-order l<strong>in</strong>ear homogeneous<br />

differential equation<br />

d2 d<br />

E(t) = −λ<br />

dt2 dt E(t) + c2E(t) with <strong>in</strong>itial conditions <br />

E(0) = 1,<br />

d<br />

dt E(t) t=0 = 0.<br />

The explicit value of the mean hyperbolic distance is therefore<br />

λt −<br />

E(t) = e 2<br />

Proof<br />

From (4.4), it follows that<br />

<br />

cosh t√ λ 2 + 4c 2<br />

2<br />

+<br />

d<br />

E(t) = −λE(t) + e−λt<br />

dt<br />

(4.8)<br />

λ<br />

√<br />

λ2 + 4c2 s<strong>in</strong>h t√λ2 + 4c2 <br />

. (4.9)<br />

2<br />

∞<br />

n=0<br />

and thus, <strong>in</strong> view of (4.7) and by lett<strong>in</strong>g I−1 = 0, we have that<br />

d2 <br />

d<br />

d<br />

E(t) = −λ E(t) − λ E(t) + λE(t)<br />

dt2 dt dt<br />

n d<br />

λ<br />

dt In<br />

+ e −λt<br />

= −2λ d<br />

dt E(t) − λ2 E(t) + c 2 E(t) + e −λt<br />

= −2λ d<br />

dt E(t) − λ2 E(t) + c 2 E(t) + λ<br />

∞<br />

n=0<br />

∞<br />

n=0<br />

λ n<br />

<br />

d<br />

dt In−1 + c 2 <br />

In<br />

n d<br />

λ<br />

dt In−1<br />

<br />

d<br />

E(t) + λE(t)<br />

dt<br />

<br />

(4.10)<br />

= −λ d<br />

dt E(t) + c2 E(t). (4.11)<br />

While it is straightforward to see that the first condition <strong>in</strong> (4.8) is verified, the second one<br />

needs some explanations: if we write<br />

d<br />

dt E(t)<br />

<br />

<br />

<br />

t=0<br />

E(∆t) − 1<br />

= lim<br />

∆t↓0 ∆t<br />

(4.12)


4.2 Equations related to the mean hyperbolic distance 64<br />

and observe that<br />

E(∆t) = (1 − λ∆t) cosh c∆t + λ<br />

∆t<br />

= (1 − λ∆t) cosh c∆t + λ∆t<br />

2<br />

0<br />

cosh ct1 cosh c(∆t − t1)dt1 + o(∆t)<br />

λ<br />

cosh c∆t + s<strong>in</strong>h c∆t + o(∆t),<br />

2c<br />

by substitut<strong>in</strong>g (4.13) <strong>in</strong> (4.12), the second condition emerges. The <strong>in</strong>tegral <strong>in</strong> (4.13) represents<br />

the mean value E{cosh η(∆t)|N(∆t) = 1} and is <strong>in</strong> fact evaluated by apply<strong>in</strong>g the Pythagorean<br />

hyperbolic theorem, as <strong>in</strong> (4.3), for k = 1 and t = ∆t.<br />

The general solution to equation (4.11) has the form<br />

<br />

λt −<br />

E(t) = e 2<br />

Ae t<br />

2<br />

√ λ 2 +4c 2<br />

+ Be<br />

t − 2<br />

√ λ 2 +4c 2 <br />

. (4.13)<br />

By impos<strong>in</strong>g the <strong>in</strong>itial conditions, the constants A and B can be evaluated and co<strong>in</strong>cide with:<br />

From (4.13) and (4.14) we obta<strong>in</strong><br />

E(t) =<br />

e− λt<br />

2<br />

so that (4.9) emerges. <br />

2<br />

A = λ + √ λ2 + 4c2 2 √ λ2 + 4c2 √<br />

λ2 + 4c2 − λ<br />

, B =<br />

2 √ λ2 + 4c2 . (4.14)<br />

<br />

λ + √ λ 2 + 4c 2<br />

√ λ 2 + 4c 2<br />

e t<br />

√<br />

√<br />

2 λ2 +4c2 λ2 + 4c2 − λ<br />

+ √ e<br />

λ2 + 4c2 √<br />

t − 2 λ2 +4c2 Remark 4.2.1. The mean value E(t) tends to <strong>in</strong>f<strong>in</strong>ity as t → ∞ so that the mov<strong>in</strong>g particle, <strong>in</strong><br />

the long run, either reaches the x axis or moves away towards the <strong>in</strong>f<strong>in</strong>ity.<br />

Of course, if c = 0 we have that E(t) = 1, and for λ → ∞ we have aga<strong>in</strong> that E(t) = 1 because<br />

<strong>in</strong> both cases the particle cannot leave the start<strong>in</strong>g po<strong>in</strong>t.<br />

If λ → 0 we get E(t) = cosh ct because the particle will simply move on the basic geodesic l<strong>in</strong>e<br />

and its hyperbolic distance grows l<strong>in</strong>early with t.<br />

We note that the hyperbolic distance itself tends to <strong>in</strong>f<strong>in</strong>ity as t → ∞ because<br />

lim cosh η(t) =<br />

t→∞<br />

∞<br />

cosh d(PkPk−1) = ∞ (4.15)<br />

k=1<br />

and (4.15) is the <strong>in</strong>f<strong>in</strong>ite product of terms bigger than one.<br />

Remark 4.2.2. By tak<strong>in</strong>g <strong>in</strong>to account the difference-differential equation (4.6), or directly from<br />

(4.3), it follows that the conditional mean values En(t) satisfy the follow<strong>in</strong>g equation with nonconstant<br />

coefficients<br />

d 2<br />

dt2 En + 2n<br />

t<br />

d<br />

dt En − n d<br />

t dt En−1 + n2 − n<br />

t2 (En − En−1) − c 2 En = 0. (4.16)<br />

In order to obta<strong>in</strong> the explicit value of the conditional mean value En(t) it is convenient to<br />

perform a series expansion of E(t), <strong>in</strong>stead of solv<strong>in</strong>g the difference-differential equation (4.16). In<br />

this way we can prove the follow<strong>in</strong>g result.


65 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

Theorem 4.2.3. The conditional mean values En(t), n ≥ 1, can be expressed as<br />

En(t) =<br />

[ n<br />

2 ]<br />

<br />

r=0<br />

+<br />

1<br />

2 n<br />

[ n−1<br />

2 ]<br />

<br />

r=0<br />

n!<br />

(n − 2r)!<br />

1<br />

2 n<br />

∞<br />

2j<br />

r + j (ct)<br />

j (2r + 2j)!<br />

j=0<br />

n!<br />

(n − 2r − 1)!<br />

∞<br />

<br />

r + j (ct)<br />

j<br />

2j<br />

(2r + 2j + 1)!<br />

j=0<br />

Proof<br />

By expand<strong>in</strong>g the hyperbolic functions <strong>in</strong> (4.9) we have that<br />

λt −<br />

E(t) = e 2<br />

+<br />

∞<br />

k=0<br />

1<br />

(2k)!<br />

λ<br />

√ λ 2 + 4c 2<br />

∞<br />

k=0<br />

<br />

t <br />

2k λ2 + 4c2 2<br />

1<br />

(2k + 1)!<br />

<br />

t <br />

<br />

2k+1<br />

λ2 + 4c2 .<br />

2<br />

(4.17)<br />

(4.18)<br />

By apply<strong>in</strong>g the Newton b<strong>in</strong>omial formula to the terms <strong>in</strong> the round brackets and by expand<strong>in</strong>g<br />

e λt<br />

2 it follows that<br />

E(t) = e −λt<br />

+ λ<br />

∞<br />

m=0<br />

∞<br />

k=0<br />

1<br />

m!<br />

1<br />

(2k + 1)!<br />

<br />

m ∞<br />

λt<br />

2<br />

k=0<br />

t<br />

2<br />

1<br />

(2k)!<br />

2k+1 k<br />

r=0<br />

t<br />

2<br />

F<strong>in</strong>ally, <strong>in</strong>terchang<strong>in</strong>g the summation order, it results that<br />

S<strong>in</strong>ce<br />

E(t) = e −λt<br />

⎡<br />

∞ ∞<br />

2r+m ⎣<br />

1 λt (2r + m)!<br />

m!r! 2 (2r + m)!<br />

+<br />

∞<br />

m=0 r=0<br />

∞<br />

m=0 r=0<br />

1<br />

m!r!<br />

2k k<br />

r=0<br />

<br />

k<br />

λ<br />

r<br />

2r (2c) 2k−2r<br />

∞<br />

j=0<br />

2r+m+1 λt (2r + m + 1)!<br />

2 (2r + m + 1)!<br />

E(t) = e −λt<br />

∞ (λt) n<br />

n=0<br />

<br />

k<br />

λ<br />

r<br />

2r (2c) 2k−2r<br />

(r + j)!<br />

j!<br />

∞<br />

j=0<br />

<br />

(r + j)!<br />

j!<br />

.<br />

(ct) 2j<br />

(2r + 2j)!<br />

(ct) 2j<br />

⎤<br />

⎦ .<br />

(2r + 2j + 1)!<br />

n! En(t), (4.19)


4.2 Equations related to the mean hyperbolic distance 66<br />

from (4.19) and (4.19), we have that<br />

En(t) = <br />

=<br />

m, r: 2r+m=n<br />

+<br />

[ n<br />

2 ]<br />

<br />

r=0<br />

+<br />

<br />

m, r: 2r+m+1=n<br />

1<br />

2 n<br />

[ n−1<br />

2 ]<br />

<br />

r=0<br />

n!<br />

(n − 2r)!<br />

1<br />

2 n<br />

1<br />

22r+m ∞<br />

<br />

(2r + m)! r + j 2j (ct)<br />

m! j (2r + 2j)!<br />

j=0<br />

1<br />

22r+m+1 ∞<br />

<br />

(2r + m + 1)! r + j (ct)<br />

m!<br />

j<br />

2j<br />

(2r + 2j + 1)!<br />

∞<br />

2j<br />

r + j (ct)<br />

j (2r + 2j)!<br />

j=0<br />

n!<br />

(n − 2r − 1)!<br />

j=0<br />

∞<br />

<br />

r + j (ct)<br />

j<br />

2j<br />

(2r + 2j + 1)! ,<br />

and this represents the explicit form of the conditional mean values. <br />

j=0<br />

Remark 4.2.3. We check formula (4.17) by evaluat<strong>in</strong>g the mean value En(t) for n = 0, 1, 2, 3.<br />

It can be noted that for n = 0 only the term r = 0 of the first sum <strong>in</strong> (4.17) must be considered,<br />

so that<br />

∞ (ct)<br />

E{cosh η(t)|N(t) = 0} =<br />

2j<br />

= cosh ct.<br />

(2j)!<br />

For n = 1 both sums of (4.17) contribute to the mean value with the r = 0 term<br />

E{cosh η(t)|N(t) = 1} = 1<br />

2<br />

j=0<br />

∞<br />

j=0<br />

(ct) 2j<br />

(2j)!<br />

+ 1<br />

2<br />

∞<br />

j=0<br />

= 1 1<br />

cosh ct + s<strong>in</strong>h ct.<br />

2 2ct<br />

(ct) 2j<br />

(2j + 1)!<br />

For n = 2 we have two terms <strong>in</strong> the first sum (correspond<strong>in</strong>g to r = 0, 1) and the term r = 0 <strong>in</strong><br />

the second sum, so that<br />

E{cosh η(t)|N(t) = 2} = 1<br />

2 2<br />

+ 1<br />

2<br />

∞<br />

j=0<br />

(ct) 2j<br />

(2j)!<br />

= 1<br />

cosh ct +<br />

22 For n = 3 we need to consider two terms <strong>in</strong> both sums<br />

+ 1<br />

2<br />

∞<br />

2j<br />

j + 1 (ct)<br />

j (2j + 2)!<br />

j=0<br />

∞ (ct)<br />

j=0<br />

2j<br />

(2j + 1)!<br />

<br />

1<br />

22 <br />

1<br />

+ s<strong>in</strong>h ct.<br />

ct 2ct<br />

E{cosh η(t)|N(t) = 3} = 1<br />

23 ∞ (ct)<br />

j=0<br />

2j 3!<br />

+<br />

(2j)! 23 ∞<br />

2j<br />

j + 1 (ct)<br />

j (2j + 2)!<br />

j=0<br />

+ 3<br />

23 ∞ (ct)<br />

j=0<br />

2j 3<br />

+<br />

(2j + 1)! 22 ∞<br />

2j<br />

j + 1 (ct)<br />

j (2j + 3)!<br />

j=0<br />

<br />

1 3<br />

= +<br />

23 23 (ct) 2<br />

<br />

6<br />

cosh ct +<br />

23 3<br />

−<br />

ct (2ct) 3<br />

<br />

s<strong>in</strong>h ct.


67 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

The same results can be obta<strong>in</strong>ed directly from (4.3) by successive <strong>in</strong>tegrations.<br />

For each step the ensemble of po<strong>in</strong>ts with hyperbolic distance equal to c(tk − tk−1) forms a<br />

Euclidean circumference Ck with radius s<strong>in</strong>h c(tk −tk−1) and center located at (0, cosh c(tk −tk−1)).<br />

At time t, if n steps have occurred, the set of po<strong>in</strong>ts Ct with hyperbolic distance equal to η(t) is a<br />

circumference with center at (0, cosh η(t)) and radius s<strong>in</strong>h η(t). Clearly<br />

cosh η(t) =<br />

n+1 <br />

k=1<br />

cosh c(tk − tk−1)<br />

so that the ord<strong>in</strong>ate of the center of Ct is equal to the product of the ord<strong>in</strong>ates of Ck. However<br />

s<strong>in</strong>h η(t) =<br />

=<br />

≥<br />

<br />

1 + cosh 2 η(t)<br />

<br />

<br />

n+1 <br />

1 + cosh 2 c(tk − tk−1)<br />

n+1 <br />

k=1<br />

k=1<br />

s<strong>in</strong>h c(tk − tk−1)<br />

and this shows that the quantity n+1<br />

k=1 s<strong>in</strong>h c(tk − tk−1) represents a lower bound of the radius of<br />

the circle Ct.<br />

Theorem 4.2.4. The functions<br />

Jn(t) =<br />

t<br />

0<br />

dt1<br />

t<br />

t1<br />

t<br />

dt2 · · ·<br />

tn−1<br />

where t0 = 0, tn+1 = t > 0, and n ≥ 1, take the form<br />

where J0(t) = s<strong>in</strong>h ct.<br />

Jn(t) = t2n+1 c n+1<br />

n!<br />

∞<br />

r=0<br />

dtn<br />

(n + r)!<br />

r!<br />

n+1 <br />

k=1<br />

s<strong>in</strong>h c(tk − tk−1),<br />

(ct) 2r<br />

, (4.20)<br />

(2r + 2n + 1)!<br />

Proof<br />

We first note that the functions Jn(t), n ≥ 1, t > 0 satisfy the difference-differential equations<br />

S<strong>in</strong>ce<br />

we have that<br />

d<br />

dt Jn = c<br />

d 2<br />

dt 2 Jn = cJn−1 + c 2 Jn, n ≥ 1, t > 0. (4.21)<br />

t<br />

0<br />

d 2<br />

dt 2 Jn = c<br />

t<br />

dt1 · · ·<br />

t<br />

+c 2<br />

tn−1<br />

0<br />

t<br />

dtn<br />

n<br />

s<strong>in</strong>h c(tk − tk−1) cosh c(t − tn),<br />

k=1<br />

t<br />

dt1 · · ·<br />

0<br />

tn−2<br />

t<br />

dt1 · · ·<br />

= c Jn−1 + c 2 Jn.<br />

dtn−1<br />

tn−1<br />

dtn<br />

n<br />

s<strong>in</strong>h c(tk − tk−1)<br />

k=1<br />

n+1 <br />

k=1<br />

s<strong>in</strong>h c(tk − tk−1)


4.2 Equations related to the mean hyperbolic distance 68<br />

From (4.21), we have that the generat<strong>in</strong>g function<br />

satisfies the differential equation<br />

In fact, by (4.21), we have<br />

∞<br />

n=0<br />

G(s, t) =<br />

n d2<br />

s<br />

dt2 Jn = cs<br />

∞<br />

n=0<br />

s n Jn<br />

(4.22)<br />

d2 G = c(s + c)G. (4.23)<br />

dt2 ∞<br />

s n−1 Jn−1 + c 2<br />

∞<br />

n=0<br />

n=0<br />

s n Jn<br />

and this easily yields (4.23). Consider<strong>in</strong>g that the general solution to (4.23) is<br />

and that G(s, t) satisfies the <strong>in</strong>itial conditions<br />

it follows that<br />

G(s, t) = Ae t√ c(s+c) + Be −t √ c(s+c)<br />

<br />

G(s, t) =<br />

G(s, 0) = 0,<br />

d<br />

dt G(s, t) t=0 = c,<br />

By expand<strong>in</strong>g the s<strong>in</strong>h function <strong>in</strong> (4.26) we obta<strong>in</strong> that<br />

G(s, t) =<br />

=<br />

=<br />

c<br />

s + c<br />

k=0 j=0<br />

(4.24)<br />

(4.25)<br />

√ c<br />

√ s + c s<strong>in</strong>h t c(s + c). (4.26)<br />

∞ (t<br />

k=0<br />

c(s + c)) 2k+1 ∞<br />

=<br />

(2k + 1)!<br />

(2k + 1)!<br />

k=0<br />

∞ k<br />

<br />

k<br />

s<br />

j<br />

j c k−j t2k+1ck+1 (2k + 1)! =<br />

∞<br />

s<br />

j=0<br />

j<br />

⎧<br />

⎨ ∞<br />

<br />

k<br />

⎩ j<br />

k=j<br />

∞<br />

s j<br />

<br />

t2j+1cj+1 ∞ (j + r)! (ct)<br />

j! r!<br />

2r<br />

<br />

(2r + 2j + 1)!<br />

j=0<br />

r=0<br />

and, <strong>in</strong> view of (4.22), formula (4.20) appears. <br />

Remark 4.2.4. We consider the quantity<br />

∞<br />

n=0<br />

n!<br />

t n Jn(t)P{N(t) = n} = e −λt<br />

∞<br />

n=0<br />

t 2k+1 c k+1 (s + c) k<br />

c k−j t2k+1 c k+1<br />

(2k + 1)!<br />

⎫<br />

⎬<br />

⎭<br />

λ n Jn(t) = e −λt G(λ, t) (4.27)<br />

= e −λt<br />

√<br />

c<br />

√ s<strong>in</strong>h t<br />

λ + c c(λ + c)<br />

which represents a lower bound for mean values of the radius of the circle C of po<strong>in</strong>ts with equal<br />

hyperbolic distance from the orig<strong>in</strong> at time t. We note that the bound (4.27) <strong>in</strong>creases if<br />

c 2 + cλ − λ 2 > 0.


69 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

For large values of λ the radius of the circle C tends to decrease because the particle often changes<br />

direction and hardly leaves the start<strong>in</strong>g po<strong>in</strong>t O.<br />

4.3 About the higher moments of the hyperbolic distance<br />

In this section we study the conditional and unconditional higher moments of the hyperbolic<br />

distance η(t). Our first results concern the derivation of the equations satisfied by the secondorder<br />

moments<br />

where<br />

and by<br />

Mn(t) = E{cosh 2 η(t)|N(t) = n}<br />

= n!<br />

t n<br />

t<br />

0<br />

dt1<br />

= n!<br />

Un(t),<br />

tn Un(t) =<br />

t<br />

0<br />

t<br />

t1<br />

t<br />

dt1 · · ·<br />

dt2 · · ·<br />

tn−1<br />

t<br />

dtn<br />

tn−1<br />

n+1 <br />

k=1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh 2 c(tk − tk−1)<br />

cosh 2 c(tk − tk−1),<br />

M(t) = E{cosh 2 η(t)} (4.28)<br />

=<br />

∞<br />

E{cosh 2 η(t)|N(t) = n}P{N(t) = n}<br />

n=0<br />

= e −λt<br />

∞<br />

λ n Un(t).<br />

n=0<br />

At first, we state the follow<strong>in</strong>g results concern<strong>in</strong>g the evaluation of the <strong>in</strong>tegrals Un(t), n ≥ 1.<br />

Lemma 4.3.1. The functions<br />

Un(t) =<br />

t<br />

0<br />

dt1<br />

t<br />

t1<br />

t<br />

dt2 · · ·<br />

tn−1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh 2 c(tk − tk−1), (4.29)<br />

with t0 = 0 and tn+1 = t, satisfy the follow<strong>in</strong>g third-order difference-differential equations<br />

where t > 0, n ≥ 1, and U0(t) = cosh 2 ct.<br />

Proof<br />

d3 dt3 Un = d2<br />

dt2 Un−1<br />

2 d<br />

+ 4c<br />

dt Un − 2c 2 Un−1, (4.30)


4.3 About the higher moments of the hyperbolic distance 70<br />

We first note that<br />

d<br />

dt Un =<br />

t<br />

0<br />

+2c<br />

= Un−1<br />

+2c<br />

t<br />

dt1 · · ·<br />

t<br />

0<br />

t<br />

A further derivation yields<br />

0<br />

tn−2<br />

t<br />

dt1 · · ·<br />

dtn−1<br />

tn−1<br />

t<br />

dt1 · · ·<br />

tn−1<br />

d 2<br />

dt 2 Un = d<br />

dt Un−1 + 2c 2<br />

= d<br />

+2c 2<br />

t<br />

0<br />

dtn<br />

dtn<br />

n<br />

cosh 2 c(tk − tk−1)<br />

k=1<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

n<br />

cosh 2 c(tk − tk−1) cosh c(t − tn) s<strong>in</strong>h c(t − tn)<br />

k=1<br />

n<br />

cosh 2 c(tk − tk−1) cosh c(t − tn) s<strong>in</strong>h c(t − tn).<br />

k=1<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dt Un−1 + 2c 2 Un<br />

t t<br />

+2c 2<br />

0<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

dtn<br />

tn−1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh 2 c(tk − tk−1)<br />

n<br />

cosh 2 c(tk − tk−1) s<strong>in</strong>h 2 c(t − tn)<br />

k=1<br />

n<br />

cosh 2 c(tk − tk−1) s<strong>in</strong>h 2 c(t − tn).<br />

S<strong>in</strong>ce it is not possible to express the <strong>in</strong>tegral <strong>in</strong> (4.31) <strong>in</strong> terms of Un and its first two derivatives, a<br />

further derivation is necessary, that, <strong>in</strong> view of (4.31), leads to the follow<strong>in</strong>g third-order differencedifferential<br />

equation<br />

<br />

d3 dt3 Un = d2<br />

dt2 Un−1<br />

2 d<br />

+ 2c<br />

dt Un + 2 2 c 3<br />

× s<strong>in</strong>h c(t − tn) cosh c(t − tn)<br />

k=1<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

= d2<br />

dt2 Un−1<br />

2 d<br />

+ 2c<br />

dt Un<br />

2 d<br />

+ 2c<br />

dt Un − 2c 2 Un−1.<br />

In view of Lemma 4.3.1 we can prove also the follow<strong>in</strong>g:<br />

dtn<br />

n<br />

cosh 2 c(tk − tk−1)<br />

Theorem 4.3.2. The function M(t) = E{cosh 2 η(t)} satisfies the third-order l<strong>in</strong>ear differential<br />

equation<br />

d3 d2<br />

M(t) = −2λ<br />

dt3 dt2 M(t) + (4c2 − λ 2 ) d<br />

dt M(t) + 2c2λM(t), (4.31)<br />

with <strong>in</strong>itial conditions ⎧⎪ ⎨<br />

Proof<br />

⎪⎩<br />

M(0) = 1,<br />

d<br />

dtM(t) = 0, t=0<br />

d 2<br />

<br />

<br />

= 2c2 .<br />

dt2 M(t) <br />

t=0<br />

k=1<br />

(4.32)


71 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

By multiply<strong>in</strong>g both members of (4.30) by λ n and summ<strong>in</strong>g up we have that<br />

and also<br />

d 3<br />

dt 3<br />

∞<br />

n=0<br />

λ n Un = λ d2<br />

dt 2<br />

∞<br />

n=1<br />

λ n−1 2 d<br />

Un−1 + 4c<br />

dt<br />

∞<br />

λ n Un − 2c 2 ∞<br />

λ<br />

n=0<br />

n=1<br />

λ n−1 Un−1,<br />

d3 dt3 λt d<br />

e M(t) = λ 2<br />

dt2 λt 2<br />

e M(t) + 4c d λt 2 λt<br />

e M(t) − 2c λe M(t),<br />

dt<br />

so that, after some manipulations, equation (4.31) appears. While the first condition <strong>in</strong> (4.32) is<br />

obvious, the second one can be <strong>in</strong>ferred from (4.31) as follows<br />

and also<br />

i.e.,<br />

d λt<br />

e M(t)<br />

dt<br />

=<br />

∞ d<br />

λ<br />

dt<br />

n=0<br />

n Un<br />

=<br />

∞<br />

λ n ∞<br />

Un−1 + 2c λ n<br />

n=0<br />

n=0<br />

t<br />

× cosh c(t − tn) s<strong>in</strong>h c(t − tn)<br />

λe λt λt d<br />

M(t) + e<br />

dt M(t) = λeλtM(t) + 2c<br />

0<br />

∞<br />

λ n<br />

n=0<br />

t<br />

dt1 · · ·<br />

t<br />

× cosh c(t − tn) s<strong>in</strong>h c(t − tn),<br />

d<br />

dt M(t)<br />

<br />

<br />

<br />

t=0<br />

0<br />

tn−1<br />

t<br />

dt1 · · ·<br />

dtn<br />

tn−1<br />

n<br />

cosh 2 c(tk − tk−1)<br />

k=1<br />

dtn<br />

n<br />

cosh 2 c(tk − tk−1)<br />

k=1<br />

= 0 (4.33)<br />

s<strong>in</strong>ce 2c cosh ct s<strong>in</strong>h ct| t=0 =0. By differentiat<strong>in</strong>g twice (4.28) and by tak<strong>in</strong>g <strong>in</strong>to account (4.31), we<br />

have that<br />

λ 2 e λt λt d<br />

M + 2λe<br />

= e λt<br />

dt<br />

<br />

λ 2 M + λ d<br />

dt M<br />

+2c 2 e λt M + 2c 2<br />

d2<br />

M + eλt M 2<br />

<br />

∞<br />

λ n<br />

n=0<br />

t<br />

0<br />

dt<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n<br />

cosh 2 c(tk − tk−1) s<strong>in</strong>h 2 c(t − tn),<br />

and therefore, by consider<strong>in</strong>g (4.33), we obta<strong>in</strong> the second condition of (4.32). <br />

In order to solve the differential equation (4.31) we need to first solve the related third-order<br />

algebraic equation<br />

r 3 + 2λr 2 − (4c 2 − λ 2 )r − 2c 2 λ = 0<br />

which can be reduced to the standard form by means of the change of variable<br />

This leads to<br />

k=1<br />

s = r + 2λ<br />

. (4.34)<br />

3<br />

s 3 <br />

2 λ<br />

− s + 4c2 +<br />

3 2λ<br />

<br />

c<br />

3<br />

2 − λ2<br />

32 <br />

= 0 (4.35)


4.3 About the higher moments of the hyperbolic distance 72<br />

to which the well-known Cardano formula can be applied. In fact, for the third-order equation<br />

the solution can be expressed as<br />

<br />

s = 3<br />

− q<br />

2 +<br />

s 3 + ps + q = 0, (4.36)<br />

p 3<br />

By compar<strong>in</strong>g (4.35) and (4.36) it results<br />

p 3<br />

3<br />

3 + q2<br />

3<br />

3 + q2<br />

<br />

3<br />

+<br />

22 − q<br />

2 −<br />

p 3<br />

3<br />

3 + q2<br />

c2<br />

= −<br />

22 33 3 2 2 2 2 2 2<br />

(2 c + λ ) + λ (λ − 3c ) ,<br />

− q<br />

<br />

= −λ c<br />

2 3<br />

2 − λ2<br />

32 <br />

.<br />

. (4.37)<br />

22 The simplest case is that of c = λ<br />

3 for which the solutions of (4.35) are s1 = 0, s2 = √ 7c and<br />

s3 = − √ 7c. After some calculations we get that<br />

E{cosh 2 η(t)} = e−2ct<br />

7<br />

<br />

1 + 6 cosh √ 7ct + 2 √ 7 s<strong>in</strong>h √ <br />

7ct .<br />

Follow<strong>in</strong>g Lemma (4.3.1) we can prove a more general result:<br />

Theorem 4.3.3. The functions<br />

K m t<br />

n (t) =<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n+1 <br />

k=1<br />

cosh m c(tk − tk−1),<br />

with t0 = 0 and tn+1 = t, are solutions of difference-differential equations of order m + 1.<br />

Proof<br />

For m = 1 and m = 2 this statement has already been shown above s<strong>in</strong>ce, <strong>in</strong> Theorem 4.2.2 and<br />

Theorem 4.3.2, we have obta<strong>in</strong>ed that<br />

and<br />

We easily see that<br />

d<br />

dt Km n = K m n−1 + c m<br />

and<br />

d 2<br />

dt 2 K1 n − d<br />

dt K1 n−1 − c 2 K 1 n = 0,<br />

d3 dt3 K2 n − d2<br />

dt2 K2 2 d<br />

n−1 − 4c<br />

dt K2 n + 2c 2 K 2 n−1 = 0.<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n<br />

k=1<br />

d 2<br />

dt 2 Km n = d<br />

dt Km n−1 + c 2 mK m n + c 2 m(m − 1)<br />

× cosh m−2 c(t − tn) s<strong>in</strong>h 2 c(t − tn).<br />

cosh m c(tk − tk−1) cosh m−1 c(t − tn) s<strong>in</strong>h c(t − tn),<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n<br />

cosh m c(tk − tk−1)<br />

k=1


73 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

In view of (4.38) it also results<br />

d 3<br />

dt 3 Km n = d2<br />

dt 2 Km n−1 + c 2 m d<br />

+c 3 m(m − 1)(m − 2)<br />

dt Km n + 2c 2 <br />

d<br />

(m − 1)<br />

t<br />

0<br />

dt1 · · ·<br />

t<br />

tn−1<br />

dt Km n − K m n−1<br />

dtn<br />

<br />

n<br />

cosh m c(tk − tk−1)<br />

× cosh m−3 c(t − tn) s<strong>in</strong>h 3 c(t − tn). (4.38)<br />

After (m − 1) derivatives the follow<strong>in</strong>g equation is obta<strong>in</strong>ed<br />

dm−1 dtm−1 Km n = dm−2<br />

dtm−2 Km n−1 + c 2 m dm−3<br />

dtm−3 Km n + · · · + c m−1 m(m − 1) · · · (m − (m − 1) + 1)<br />

t t<br />

×<br />

0<br />

dt1 · · ·<br />

and the next derivative gives<br />

tn−1<br />

dtn<br />

n<br />

cosh m c(tk − tk−1)<br />

k=1<br />

× cosh c(t − tn) s<strong>in</strong>h m−1 c(t − tn), (4.39)<br />

dm dtm Km n = dm−1<br />

dtm−1 Km n−1 + c 2 m dm−2<br />

dtm−2 Km n + · · · + c m m(m − 1) · · · 2<br />

t t n<br />

× dt1 · · · dtn cosh<br />

0<br />

tn−1 k=1<br />

m c(tk − tk−1) s<strong>in</strong>h m c(t − tn)<br />

k=1<br />

+c m m(m − 1) · · · 2 · (m − 1)<br />

t t n<br />

× dt1 · · · dtn cosh m c(tk − tk−1)<br />

0<br />

tn−1<br />

k=1<br />

× cosh 2 c(t − tn) s<strong>in</strong>h m−2 c(t − tn). (4.40)<br />

The second <strong>in</strong>tegral of (4.40) can be expressed <strong>in</strong> terms of the derivatives of order (m − 2) and<br />

lower.<br />

By further differentiat<strong>in</strong>g equation (4.40) it turns out that, because of (4.39), the derivative of the<br />

first <strong>in</strong>tegral <strong>in</strong> (4.40) can be expressed <strong>in</strong> terms of the derivatives of order (m − 1) and lower. The<br />

theorem is thus proved. <br />

Likewise Theorem 4.2.4, the follow<strong>in</strong>g theorem holds:<br />

Theorem 4.3.4. The function<br />

Vn(t) =<br />

t<br />

0<br />

dt1<br />

t<br />

t1<br />

t<br />

dt2 · · ·<br />

tn−1<br />

dtn<br />

n+1 <br />

k=1<br />

s<strong>in</strong>h 2 c(tk − tk−1)<br />

with t0 = 0 and tn+1 = t, satisfies the third-order difference-differential equation<br />

where t > 0, n ≥ 1, and V0(t) = s<strong>in</strong>h 2 ct.<br />

Proof<br />

d3 dt3 Vn<br />

2 d<br />

= 4c<br />

dt Vn + 2c 2 Vn−1


4.4 Motions with jumps backwards to the start<strong>in</strong>g po<strong>in</strong>t 74<br />

We first note that<br />

d<br />

dt Vn = 2c<br />

and therefore<br />

and<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tn−1<br />

d 2<br />

dt 2 Vn = 2c 2 Vn + 2c 2<br />

t<br />

dtn<br />

0<br />

n<br />

s<strong>in</strong>h 2 c(tk − tk−1) s<strong>in</strong>h c(t − tn) cosh c(t − tn) (4.41)<br />

k=1<br />

t<br />

dt1 · · ·<br />

tn−1<br />

d3 dt3 Vn<br />

2 d<br />

= 2c<br />

dt Vn + 2c 2 Vn−1 + 4c 3<br />

dtn<br />

t<br />

0<br />

n<br />

s<strong>in</strong>h 2 c(tk − tk−1) cosh 2 c(t − tn),<br />

k=1<br />

t<br />

dt1 · · ·<br />

tn−1<br />

dtn<br />

n<br />

s<strong>in</strong>h 2 c(tk − tk−1)<br />

× s<strong>in</strong>h c(t − tn) cosh c(t − tn). (4.42)<br />

F<strong>in</strong>ally, by substitut<strong>in</strong>g (4.41) <strong>in</strong> (4.42), we obta<strong>in</strong><br />

<br />

d3 dt3 Vn<br />

2 d<br />

= 4c<br />

dt Vn + 2c 2 Vn−1.<br />

4.4 Motions with jumps backwards to the start<strong>in</strong>g po<strong>in</strong>t<br />

We here exam<strong>in</strong>e the planar motion dealt with so far assum<strong>in</strong>g now that, at the <strong>in</strong>stants of changes<br />

of direction, the particle can return to the start<strong>in</strong>g po<strong>in</strong>t and commence its motion from scratch.<br />

The new motion and the orig<strong>in</strong>al one are governed by the same Poisson process so that changes<br />

of direction occur simultaneously <strong>in</strong> the orig<strong>in</strong>al as well as <strong>in</strong> the new motion start<strong>in</strong>g afresh from<br />

the orig<strong>in</strong>. This implies that the arcs of the orig<strong>in</strong>al sample path and those of the new trajectories<br />

have the same hyperbolic length. However, the angles formed by successive segments differ <strong>in</strong> order<br />

to make the <strong>Hyperbolic</strong> Pythagorean Theorem applicable to the trajectories of the new motion.<br />

In order to make our description clearer, we consider the case where, <strong>in</strong> the <strong>in</strong>terval (0, t),<br />

N(t) = n Poisson events (n ≥ 1) occur and we assume that the jump to the orig<strong>in</strong> happens at the<br />

first change of direction, i.e., at the <strong>in</strong>stant t1. The <strong>in</strong>stants of changes of direction for the new<br />

motion are<br />

t ′ k = tk+1 − t1<br />

where k = 0, · · · , n with t ′ 0 = 0 and t ′ n = t − t1 and the hyperbolic lengths of the correspond<strong>in</strong>g<br />

arcs are<br />

c(t ′ k − t ′ k−1) = c(tk+1 − tk).<br />

Therefore, at the <strong>in</strong>stant t, the hyperbolic distance from the orig<strong>in</strong> of the particle perform<strong>in</strong>g the<br />

k=1


75 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

motion which has jumped back to O at time t1 is<br />

n<br />

k=1<br />

cosh c(t ′ k − t ′ k−1) =<br />

=<br />

n<br />

cosh c(tk+1 − tk)<br />

k=1<br />

n+1 <br />

k=2<br />

cosh c(tk − tk−1) (4.43)<br />

where 0 = t ′ 0 < t ′ 1 < · · · < t ′ n = t − t1 and tk+1 = t ′ k + t1. Formula (4.43) shows that the new<br />

motion has an hyperbolic distance equal to that of the orig<strong>in</strong>al motion where the first step has been<br />

deleted. However, the distance between the position Pt and the orig<strong>in</strong> O of the mov<strong>in</strong>g particle<br />

which jumped back to O after hav<strong>in</strong>g reached the position P1, is different from the distance of<br />

Pt from P1 s<strong>in</strong>ce the angle between successive steps must be readjusted <strong>in</strong> order to apply the<br />

<strong>Hyperbolic</strong> Pythagorean Theorem.<br />

If we denote by T1 the random <strong>in</strong>stant of the return to the start<strong>in</strong>g po<strong>in</strong>t (occurr<strong>in</strong>g at the first<br />

Poisson event), we have that<br />

E{cosh η1(t)I {N(t)≥1}|N(t) = n} = E{cosh η(t − T1)I {T1≤t}|N(t) = n}<br />

By observ<strong>in</strong>g that<br />

and that<br />

=<br />

=<br />

t<br />

0<br />

t<br />

0<br />

E{cosh η(t − T1)I {T1∈dt1}|N(t) = n}dt1<br />

E{cosh η(t − T1)|T1 = t1, N(t) = n}P{T1 ∈ dt1|N(t) = n}dt1.<br />

E{cosh η(t − T1)|T1 = t1, N(t) = n} = E{cosh η(t − t1)|N(t) = n − 1}<br />

=<br />

(n − 1)!<br />

(t − t1) n−1 In−1(t − t1),<br />

with 0 < t1 < t, formula (4.44) becomes<br />

P{T1 ∈ dt1|N(t) = n} = n!<br />

tn (t − t1) n−1<br />

(n − 1)! dt1<br />

E{cosh η1(t)I {N(t)≥1}|N(t) = n} = n!<br />

t n<br />

t<br />

0<br />

In−1(t − t1)dt1. (4.44)<br />

From (4.44) we have that the mean hyperbolic distance for the particle which returns to O at time<br />

T1 has the form:<br />

E{cosh η1(t)|N(t) ≥ 1} =<br />

=<br />

e −λt<br />

P{N(t) ≥ 1}<br />

λe −λt<br />

P{N(t) ≥ 1}<br />

∞<br />

λ n<br />

n=1<br />

t<br />

0<br />

t<br />

0<br />

In−1(t − t1)dt1<br />

e λ(t−t1) E(t − t1)dt1<br />

We give here a general expression for the mean value of the hyperbolic distance of a particle<br />

which returns to the orig<strong>in</strong> for the last time at the k-th Poisson event Tk. We shall denote the<br />

distance by the follow<strong>in</strong>g equivalent notation η(t−Tk) = ηk(t) where the first expression underl<strong>in</strong>es<br />

that the particle starts from scratch at time Tk and then moves away for the rema<strong>in</strong><strong>in</strong>g <strong>in</strong>terval of


4.4 Motions with jumps backwards to the start<strong>in</strong>g po<strong>in</strong>t 76<br />

length t − Tk. In the general case we have the result stated <strong>in</strong> the next theorem:<br />

Theorem 4.4.1. If N(t) ≥ k, then the mean value of the hyperbolic distance ηk is equal to<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

where E(t) is given by (4.9).<br />

Proof<br />

We start by observ<strong>in</strong>g that<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

=<br />

=<br />

=<br />

λ k e −λt<br />

P{N(t) ≥ k}<br />

t<br />

0<br />

t<br />

dt1 · · ·<br />

tk−1<br />

λke−λt t<br />

e<br />

P{N(t) ≥ k}(k − 1)! 0<br />

λ(t−tk) k−1<br />

tk ∞<br />

E{cosh ηk(t)I {N(t)=n}|N(t) ≥ k}<br />

n=k<br />

n=k<br />

e λ(t−tk) E(t − tk)dtk<br />

E(t − tk)dtk,<br />

∞<br />

P{N(t) = n}<br />

E{cosh ηk(t)I {N(t)≥k}|N(t) = n}<br />

P{N(t) ≥ k}<br />

∞<br />

E{cosh ηk(t)I {N(t)≥k}|N(t) = n}P{N(t) = n|N(t) ≥ k}.<br />

n=k<br />

S<strong>in</strong>ce Tk = <strong>in</strong>f{t : N(t) = k}, the conditional mean value <strong>in</strong>side the sum can be developed as<br />

follows<br />

E{cosh ηk(t)I {N(t)≥k}|N(t) = n} = E{cosh η(t − Tk)I {Tk≤t}|N(t) = n}<br />

In view of (4.3) we have that<br />

=<br />

=<br />

t<br />

0<br />

t<br />

0<br />

E{cosh η(t − tk)I {Tk∈dtk}|N(t) = n}dtk<br />

E{cosh η(t − tk)|Tk = tk, N(t) = n}P{Tk ∈ dtk|N(t) = n}dtk.<br />

E{cosh η(t − Tk)|Tk = tk, N(t) = n} = E{cosh η(t − tk)|N(t − tk) = n − k}<br />

=<br />

(n − k)!<br />

(t − tk) n−k In−k(t − tk),<br />

and on the base of well-known properties of the Poisson process we have that<br />

where 0 < tk < t. In conclusion we have that<br />

P{Tk ∈ dtk|N(t) = n} = n!<br />

tn (t − tk) n−k t<br />

(n − k)!<br />

k−1<br />

k<br />

(k − 1)! dtk<br />

E{cosh ηk(t)I {N(t)≥k}|N(t) = n} = n!<br />

t n<br />

1<br />

(k − 1)!<br />

t<br />

0<br />

t k−1<br />

k<br />

In−k(t − tk)dtk


77 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

and, from this and (4.45), it follows that<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

=<br />

∞<br />

n=k<br />

n!<br />

t n (k − 1)!<br />

t<br />

0<br />

t k−1<br />

k<br />

λ k e −λt<br />

P{N(t) ≥ k}(k − 1)!<br />

e −λt (λt) n<br />

In−k(t − tk)dtk<br />

n!P{N(t) ≥ k}<br />

t<br />

e<br />

0<br />

λ(t−tk) k−1<br />

tk F<strong>in</strong>ally, <strong>in</strong> view of Cauchy formula of multiple <strong>in</strong>tegrals, we obta<strong>in</strong> that<br />

<br />

P{N(t) ≥ k}(k − 1)!<br />

=<br />

λke−λt t<br />

e<br />

0<br />

λ(t−tk) k−1<br />

tk λke−λt t t<br />

dt1 · · ·<br />

P{N(t) ≥ k} 0<br />

tk−1<br />

E(t − tk)dtk<br />

e λ(t−tk) E(t − tk)dtk.<br />

E(t − tk)dtk.<br />

Theorem 4.4.2. The mean of the hyperbolic distance of the mov<strong>in</strong>g particle return<strong>in</strong>g to the orig<strong>in</strong><br />

at the k-th change of direction is<br />

where<br />

E{cosh ηk(t)|N(t) ≥ k}<br />

=<br />

λ k e −λt<br />

√ λ 2 + 4c 2 P{N(t) ≥ k}<br />

<br />

e At<br />

A<br />

k−1 − eBt<br />

k−1 <br />

<br />

1 1<br />

+ −<br />

Bk−1 Bi Ai <br />

k−i−1 t<br />

(k − i − 1)!<br />

i=1<br />

A = 1<br />

2 (λ + λ 2 + 4c 2 ), B = 1<br />

2 (λ − λ 2 + 4c 2 ). (4.45)<br />

For k = 1, the sum <strong>in</strong> (4.45) is <strong>in</strong>tended to be zero.<br />

Proof<br />

We can prove (4.45) by apply<strong>in</strong>g both formulas <strong>in</strong> (4.45). We start our proof by employ<strong>in</strong>g the<br />

first one:<br />

E{cosh ηk(t)|N(t) ≥ k} (4.46)<br />

=<br />

λ k e −λt<br />

P{N(t) ≥ k}<br />

t<br />

0<br />

dt1 · · ·<br />

Therefore, <strong>in</strong> view of (4.15), formula (4.46) becomes<br />

E{cosh ηk(t)|N(t) ≥ k}<br />

=<br />

λ k e −λt<br />

P{N(t) ≥ k}<br />

t<br />

× e (t−tk ) √<br />

2 λ2 +4c2 +<br />

0<br />

dt1 · · ·<br />

<br />

t<br />

tk−1<br />

t<br />

e λ(t−tk)<br />

−λ + √ λ 2 + 4c 2<br />

√ λ 2 + 4c 2<br />

tk−1<br />

e λ(t−tk) E(t − tk)dtk.<br />

<br />

λ − e 2 (t−tk)<br />

<br />

2<br />

<br />

e − (t−tk ) √<br />

2 λ2 +4c2 λ + √ λ2 + 4c2 <br />

√<br />

λ2 + 4c2 <br />

dtk.


4.4 Motions with jumps backwards to the start<strong>in</strong>g po<strong>in</strong>t 78<br />

By <strong>in</strong>troduc<strong>in</strong>g A and B as <strong>in</strong> (4.45), we can easily determ<strong>in</strong>e the k-fold <strong>in</strong>tegral<br />

E{cosh ηk(t)|N(t) ≥ k}<br />

=<br />

=<br />

=<br />

λ k e −λt<br />

√ λ 2 + 4c 2 P{N(t) ≥ k}<br />

λ k e −λt<br />

√ λ 2 + 4c 2 P{N(t) ≥ k}<br />

t<br />

0<br />

λke−λt t<br />

√<br />

λ2 + 4c2P{N(t) ≥ k} 0<br />

t<br />

At the j-th stage the <strong>in</strong>tegral becomes<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

0<br />

t<br />

dt1 · · ·<br />

dt1 · · ·<br />

dt1 · · ·<br />

tk−1<br />

t<br />

tk−2<br />

t<br />

tk−3<br />

λ k e −λt<br />

√ λ 2 + 4c 2 P{N(t) ≥ k}<br />

i=1<br />

At the k − 1-th stage the <strong>in</strong>tegral becomes<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

<br />

Ae A(t−tk) − Be B(t−tk) <br />

dtk<br />

<br />

e A(t−tk−1) − e B(t−tk−1) <br />

dtk−1<br />

e A(t−tk−2)<br />

t<br />

0<br />

A<br />

t<br />

dt1 · · ·<br />

j−1<br />

<br />

1 1<br />

+ −<br />

Bi Ai <br />

(t − tk−j) j−i−1<br />

<br />

.<br />

(j − i − 1)!<br />

λ k e −λt<br />

√ λ 2 + 4c 2 P{N(t) ≥ k}<br />

i=1<br />

t<br />

− eB(t−tk−2)<br />

B<br />

tk−j−1<br />

dt1<br />

0<br />

k−2 <br />

<br />

1 1<br />

+ −<br />

Bi Ai <br />

(t − t1) k−i−2<br />

<br />

.<br />

(k − i − 2)!<br />

+ 1<br />

<br />

1<br />

− dtk−2.<br />

B A<br />

e A(t−tk−j)<br />

e A(t−t1)<br />

A j−1<br />

eB(t−t1)<br />

−<br />

Ak−2 Bk−2 eB(t−tk−j)<br />

−<br />

Bj−1 At the k-th <strong>in</strong>tegration we obta<strong>in</strong> formula (4.45). By means of the second formula <strong>in</strong> (4.45) and<br />

by repeated <strong>in</strong>tegrations by parts we can obta<strong>in</strong> aga<strong>in</strong> result (4.45). <br />

Remark 4.4.1. For k = 1 we have that<br />

E{cosh η1(t)|N(t) ≥ 1} =<br />

λ<br />

√ λ 2 + 4c 2<br />

s<strong>in</strong>h t<br />

√<br />

2 λ2 + 4c2 s<strong>in</strong>h λt<br />

2<br />

. (4.47)<br />

It is clear that the mean value (4.47) tends to <strong>in</strong>f<strong>in</strong>ity as t → ∞. Furthermore, if λ, c → ∞<br />

(so that c2<br />

λ → 1) then E{cosh η1(t)I {N(t)≥1}} → e t . It can also be checked that if c = 0 then<br />

E{cosh η1(t)I {N(t)≥1}} = 1, s<strong>in</strong>ce the particle never leaves the start<strong>in</strong>g po<strong>in</strong>t.<br />

For k = 2 formula (4.45) yields<br />

E{cosh η2(t)|N(t) ≥ 2} =<br />

=<br />

λ2 λt − e 2<br />

c2 <br />

P{N(t) ≥ 2}<br />

−<br />

λ<br />

√ λ 2 + 4c 2<br />

cosh t <br />

λ2 + 4c2 2<br />

t <br />

s<strong>in</strong>h λ2 + 4c2 −<br />

− e<br />

2<br />

λt<br />

<br />

2<br />

λ 2<br />

c 2 P{N(t) ≥ 2} [E{cosh η(t)} (4.48)<br />

− P{N(t) ≥ 1}E{cosh η1(t)|N(t) ≥ 1} − e −λt


79 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

S<strong>in</strong>ce<br />

lim<br />

c→0<br />

1<br />

c2 <br />

cosh t<br />

2<br />

we have, as expected, that<br />

λ 2 + 4c 2 −<br />

λ<br />

√ λ 2 + 4c 2<br />

s<strong>in</strong>h t<br />

2<br />

<br />

λ2 + 4c2 −<br />

− e λt<br />

<br />

2 = e λt<br />

2<br />

P{N(t) ≥ 2},<br />

λ2 lim<br />

c→0 E{cosh η2(t)|N(t) ≥ 2} = 1. (4.49)<br />

Also, when λ → ∞, we obta<strong>in</strong> the same limit as <strong>in</strong> (4.49). The expression (4.48) suggests the<br />

follow<strong>in</strong>g decomposition<br />

E{cosh η(t)} = c2<br />

λ 2 P{N(t) ≥ 2}E{cosh η2(t)|N(t) ≥ 2}<br />

+P{N(t) ≥ 1}E{cosh η1(t)|N(t) ≥ 1} + e −λt<br />

= c2<br />

λ 2 E{cosh η2(t)I {N(t)≥2}} + E{cosh η1(t)I {N(t)≥1}} + e −λt<br />

Remark 4.4.2. The result <strong>in</strong> (4.45) appears as the mean hyperbolic distance of a motion start<strong>in</strong>g<br />

from the orig<strong>in</strong> and runn<strong>in</strong>g, without returns, until time t − Tk, where Tk has a truncated Gamma<br />

distribution (Erlang distribution) with density<br />

P{Tk ∈ dtk} =<br />

In other words we can write (4.45) as<br />

λ k e −λtk t k−1<br />

k<br />

(k − 1)!P{Tk ≤ t} dtk<br />

0 < tk < t.<br />

E{cosh ηk(t)|N(t) ≥ k} = E{E{cosh η(t − Tk)}} (4.50)<br />

=<br />

t<br />

0<br />

E{cosh η(t − Tk)}P{Tk ∈ dtk}.<br />

Furthermore the expression (4.45) conta<strong>in</strong>s a fractional <strong>in</strong>tegral of order k for the function g(s) =<br />

e λs E(s)<br />

E{cosh ηk(t)|N(t) ≥ k} =<br />

λ k<br />

∞ (λt)<br />

j=k<br />

j<br />

j!<br />

If the mean value (4.50) is taken with respect to<br />

then we have that<br />

P{Tν ∈ ds} = λν e −λs s ν−1<br />

E{E{cosh η(t − Tν)}} = λν e −λt<br />

1<br />

Γ(k)<br />

t<br />

(t − s)<br />

0<br />

k−1 e λs E(s)ds<br />

ds 0 < s < t,<br />

Γ(ν)P{Tν ≤ t}<br />

P{Tν ≤ t}<br />

1<br />

Γ(ν)<br />

t<br />

(t − s)<br />

0<br />

ν−1 e λs E(s)ds<br />

that also conta<strong>in</strong>s a fractional <strong>in</strong>tegral of order ν <strong>in</strong> the sense of Riemann-Liouville. The expression<br />

(4.51) can be <strong>in</strong>terpreted as the mean hyperbolic distance at time t where the particle can jump<br />

back to the orig<strong>in</strong> at an arbitrary <strong>in</strong>stant (different from the <strong>in</strong>stants of change of direction).<br />

<br />

.


4.5 Motion at f<strong>in</strong>ite velocity on the surface of a three-dimensional sphere 80<br />

4.5 Motion at f<strong>in</strong>ite velocity on the surface of a three-dimensional<br />

sphere<br />

Let P0 be a po<strong>in</strong>t on the equator of a three-dimensional sphere. Let us assume that the particle<br />

starts mov<strong>in</strong>g from P0 along the equator <strong>in</strong> one of the two possible directions (clockwise or counterclockwise)<br />

with velocity c.<br />

At the first Poisson event (occurr<strong>in</strong>g at time T1) it starts mov<strong>in</strong>g on the meridian jo<strong>in</strong><strong>in</strong>g the<br />

north pole PN with the position reached at time T1 (denoted by P1) along one of the two possible<br />

directions (see Figure 4.3).<br />

At the second Poisson event the particle is located at P2 and its distance from the start<strong>in</strong>g po<strong>in</strong>t<br />

P0 is the length of the hypothenuse of a right spherical triangle with cathetus P0P1 and P1P2; the<br />

hypothenuse belongs to the equatorial circumference through P0 and P2.<br />

Now the particle cont<strong>in</strong>ues its motion (<strong>in</strong> one of the two possible directions) along the equatorial<br />

circumference orthogonal to the hypothenuse through P0 and P2 until the third Poisson event<br />

occurs.<br />

In general, the distance d(P0Pt) of the po<strong>in</strong>t Pt from the orig<strong>in</strong> P0 is the length of the shortest<br />

arc of the equatorial circumference through P0 and Pt and therefore it takes values <strong>in</strong> the <strong>in</strong>terval<br />

[0, π]. Counter-clockwise motions cover the arcs <strong>in</strong> [−π, 0] so that the distance is also def<strong>in</strong>ed <strong>in</strong><br />

[0, π] or <strong>in</strong> [−π/2, π/2] with a shift that avoids negative values for the cos<strong>in</strong>e.<br />

P0<br />

Pt<br />

PN<br />

Figure 4.3: Motion on the surface of a three-dimensional sphere.<br />

By means of the spherical Pythagorean relationship we have that the Euclidean distance d(P0P2)<br />

satisfies<br />

cos d(P0P2) = cos d(P0P1) cos d(P1P2)<br />

and, after three displacements,<br />

cos d(P0P3) = cos d(P0P2) cos d(P2P3)<br />

P2<br />

P1<br />

= cos d(P0P1) cos d(P1P2) cos d(P2P3).<br />

After n displacements the position Pt on the sphere at time t is given by<br />

cos d(P0Pt) =<br />

n<br />

cos d(PkPk−1) cos d(PnPt).<br />

k=1


81 Travell<strong>in</strong>g <strong>Random</strong>ly on H 2 with a Pythagorean Compass<br />

S<strong>in</strong>ce d(PkPk−1) is represented by the amplitude of the arc run <strong>in</strong> the <strong>in</strong>terval (tk, tk−1), it results<br />

d(PkPk−1) = c(tk − tk−1).<br />

The mean value E{cos d(P0Pt)|N(t) = n} is given by<br />

where t0 = 0, tn+1 = t, and<br />

En(t) = E{cos d(P0Pt)|N(t) = n}<br />

= n!<br />

t n<br />

t<br />

0<br />

dt1<br />

= n!<br />

Hn(t),<br />

tn Hn(t) =<br />

t<br />

0<br />

t<br />

t1<br />

dt2 · · ·<br />

t<br />

dt1 · · ·<br />

The mean value E{cos d(P0Pt)} is given by<br />

tn−1<br />

t<br />

dtn<br />

tn−1<br />

n+1 <br />

k=1<br />

dtn<br />

n+1 <br />

k=1<br />

cos c(tk − tk−1)<br />

cos c(tk − tk−1).<br />

E(t) = E{cos d(P0Pt)}<br />

∞<br />

= E{cos d(P0Pt)|N(t) = n}P{N(t) = n}<br />

n=0<br />

= e −λt<br />

∞<br />

λ n Hn(t).<br />

n=0<br />

By steps similar to those of the hyperbolic case we have that Hn(t), t ≥ 0, satisfies the differencedifferential<br />

equation<br />

d2 dt2 Hn = d<br />

dt Hn−1 − c 2 Hn,<br />

where H0(t) = cos ct, and therefore we can prove the follow<strong>in</strong>g:<br />

Theorem 4.5.1. The mean value E(t) = E{cos d(P0Pt)} satisfies<br />

with <strong>in</strong>itial conditions <br />

and has the form<br />

Proof<br />

⎧<br />

⎪⎨<br />

E(t) =<br />

⎪⎩<br />

<br />

λt − e 2 cosh t<br />

λt − e 2<br />

λt − e 2<br />

√<br />

2 λ2 − 4c2 +<br />

<br />

λt<br />

<br />

1 + 2<br />

√<br />

4c2 − λ2 +<br />

cos t<br />

2<br />

d2 d<br />

E = −λ<br />

dt2 dt E − c2E (4.51)<br />

E(0) = 1,<br />

d<br />

dt E(t) t=0 = 0,<br />

λ<br />

√ λ 2 −4c 2<br />

λ<br />

√ 4c 2 −λ 2<br />

√ <br />

t s<strong>in</strong>h 2 λ2 − 4c2 √ <br />

t s<strong>in</strong> 2 4c2 − λ2 0 < 2c < λ,<br />

λ = 2c > 0,<br />

2c > λ > 0.<br />

(4.52)<br />

(4.53)


4.5 Motion at f<strong>in</strong>ite velocity on the surface of a three-dimensional sphere 82<br />

The solution to the problem (4.51)-(4.52) is given by<br />

E(t) =<br />

so that (4.53) emerges. <br />

e− λt<br />

2<br />

+<br />

2<br />

<br />

e t<br />

√<br />

2 λ2−4c2 λ<br />

√ λ 2 − 4c 2<br />

<br />

+ e<br />

e t<br />

√<br />

2 λ2−4c2 √<br />

t − 2 λ2−4c2 <br />

− e<br />

t − 2<br />

√ λ 2 −4c 2 <br />

,<br />

For large values of λ the first expression furnishes E(t) ∼ 1 and therefore the particle hardly<br />

leaves the start<strong>in</strong>g po<strong>in</strong>t.<br />

If λ<br />

2 < c, the mean value exhibits an oscillat<strong>in</strong>g behavior; <strong>in</strong> particular, the oscillations decrease as<br />

time goes on, and this means that the particle moves further and further reach<strong>in</strong>g <strong>in</strong> the limit the<br />

poles of the sphere.<br />

Remark 4.5.1. By assum<strong>in</strong>g that c is replaced by ic <strong>in</strong> (4.53) we formally extract from the first<br />

and the third expression <strong>in</strong> (4.53) the hyperbolic mean distance (4.9). This is because the space<br />

H 2 can be regarded as a sphere with imag<strong>in</strong>ary radius. Clearly the <strong>in</strong>termediate case λ = 2c has<br />

no correspondence for the motion on H 2 because the Poisson rate must be a real positive number.


Chapter 5<br />

Cascades of Particles Mov<strong>in</strong>g at<br />

F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

5.1 Description of the randomly mov<strong>in</strong>g and branch<strong>in</strong>g model<br />

We assume that a unit-mass particle is placed at time t = 0 at the orig<strong>in</strong> O of H 2 and starts<br />

mov<strong>in</strong>g on the ma<strong>in</strong> geodesic l<strong>in</strong>e represented <strong>in</strong> H 2 by the half-circle of radius 1 pass<strong>in</strong>g through<br />

O. This particle chooses with probability 1/2 one of the two possible directions and moves with<br />

constant hyperbolic velocity equal to c (see Figure 5.1 (a)). The hyperbolic velocity<br />

c = ds<br />

dt<br />

<br />

dx<br />

1<br />

=<br />

y dt<br />

2<br />

+<br />

2 dy<br />

dt<br />

is assumed to be constant. For an Euclidean observer, the closer to the x-axis is the mov<strong>in</strong>g particle<br />

the slower it moves.<br />

A Poisson process of rate λ governs the changes of direction. At the first Poisson event the<br />

particle splits <strong>in</strong>to two pieces of equal mass: one cont<strong>in</strong>ues its motion on the same geodesic l<strong>in</strong>e<br />

while the other one starts mov<strong>in</strong>g (<strong>in</strong> one of the two possible directions) on the geodesic l<strong>in</strong>e<br />

orthogonal to the previous one (see Figure 5.1 (b)). In general, at the k-th Poisson event, the<br />

deviat<strong>in</strong>g particle of mass 1/2 k undergoes a further decomposition: one spl<strong>in</strong>ter of mass 1/2 k+1<br />

cont<strong>in</strong>ues undisturbed its motion, while the other one, also of mass 1/2 k+1 , is forced to move onto<br />

the geodesic l<strong>in</strong>e orthogonal to that jo<strong>in</strong><strong>in</strong>g O with the position it occupied at the time where the<br />

splitt<strong>in</strong>g took place.<br />

Therefore, if no Poisson event occurs (i.e., {N(t) = 0}, where N(t) is the number of Poisson<br />

events <strong>in</strong> [0, t]), the unit-mass particle is located, at time t, on the first geodesic l<strong>in</strong>e at an hyperbolic<br />

distance from O equal to η0(t) = ct (see Figure 5.1 (a)).<br />

If one Poisson event happens, at time S1 < t (i.e., {N(t) = 1}), then, at time t, one fragment<br />

of mass 1/2 will be at distance η0(t) = ct on the first geodesic l<strong>in</strong>e, while the other spl<strong>in</strong>ter, also<br />

of mass 1/2, will be located at hyperbolic distance η1(t) given by<br />

cosh η1(t) = cosh c S1 cosh c(t − S1), (5.1)


5.1 Description of the randomly mov<strong>in</strong>g and branch<strong>in</strong>g model 84<br />

1<br />

1<br />

1 2 3<br />

●<br />

O<br />

●<br />

0<br />

(a) N(t)=0<br />

1 2 3<br />

●<br />

O<br />

●<br />

0<br />

(d) N(t)=3<br />

● 1<br />

●<br />

1 2<br />

●<br />

1 2 2<br />

1<br />

1<br />

● 1 2 4<br />

1 2 4<br />

●<br />

O<br />

●<br />

0<br />

(b) N(t)=1<br />

1 2 3<br />

●<br />

O<br />

●<br />

0<br />

(e) N(t)=4<br />

●<br />

●<br />

1 2<br />

1 2<br />

●<br />

1 2<br />

●<br />

1 2 2<br />

1<br />

1<br />

4<br />

∏ coshc(Sj − Sj−1)<br />

j=1<br />

5<br />

∏<br />

j=1<br />

coshc(S j − Sj−1)<br />

●<br />

1 2 2<br />

●<br />

O<br />

●<br />

0<br />

(c) N(t)=2<br />

3<br />

∏ coshc(Sj − Sj−1)<br />

j=1<br />

●<br />

●<br />

0<br />

(f) N(t)=4<br />

●<br />

●<br />

1 2<br />

coshct<br />

●<br />

1 2 2<br />

coshcS 1coshc(t − S1)<br />

Figure 5.1: In (a), the trajectory of the unit-mass particle <strong>in</strong>itially placed at the orig<strong>in</strong> O of H 2<br />

and mov<strong>in</strong>g on the ma<strong>in</strong> geodesic l<strong>in</strong>e is shown. When N(t) = 0, no dis<strong>in</strong>tegration occurs. In (b),<br />

(c), (d), and (e), the trajectories of the particles generated by N(t) = 1, 2, 3, 4 Poisson events are<br />

plotted. The relevant mass associated with each particle is also <strong>in</strong>dicated by a suitable label. In<br />

(f), for each particle the relevant hyperbolic cos<strong>in</strong>e of the hyperbolic distance from the orig<strong>in</strong> is<br />

also reported.<br />

where <strong>in</strong> formula (5.1) the <strong>Hyperbolic</strong> Pythagorean Theorem 2.5.3 has been applied (see Figure<br />

5.1 (b)).<br />

If N(t) = n, the process described above produces n + 1 spl<strong>in</strong>ters. The particle generated at<br />

the k-th Poisson event, with k = 0, · · · n − 1, (and which will not break up after the k-th event)<br />

has mass 1/2 k+1 while the last spl<strong>in</strong>ter, which has changed direction at all fission events, has mass<br />

1/2 n . For every k = 0, · · · n, the particle generated at the k-th Poisson event is located, at time t,<br />

at the hyperbolic distance ηk(t) from O. Such a distance, <strong>in</strong> force of the <strong>Hyperbolic</strong> Pythagorean<br />

Theorem, reads<br />

cosh ηk(t) =<br />

k+1 <br />

j=1<br />

cosh c(Sj − Sj−1),<br />

where S0 = 0, Sk+1 = t and the random times Sj with j = 1, · · · k represent the <strong>in</strong>stants where<br />

the Poisson events happen and the deviations of motion occur. (see Figure 5.1 (f)).<br />

In general, if the number of splits recorded is N(t), the number of particles is N(t) + 1 and each<br />

runs on a different geodesic l<strong>in</strong>e of H 2 . The hyperbolic distance of the center of mass cm of the<br />

O<br />

●<br />


85 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

cloud of mov<strong>in</strong>g particles, at time t > 0, is denoted by ηcm(t) and is represented by<br />

cosh ηcm(t) =<br />

N(t)−1 <br />

k=0<br />

<br />

1<br />

2k+1 k+1<br />

j=1<br />

cosh c(Sj − Sj−1)I {N(t)>0} + 1<br />

2N(t) N(t)+1 <br />

j=1<br />

cosh c(Sj − Sj−1), (5.2)<br />

where S1, S2 . . . S N(t) are the random times at which the dis<strong>in</strong>tegrations occur, and S0 = 0,<br />

S N(t)+1 = t. The second term <strong>in</strong> (5.2) refers to the spl<strong>in</strong>ter which underwent all dis<strong>in</strong>tegrations<br />

occurred until time t, while the first one is related to those particles produced dur<strong>in</strong>g the<br />

branch<strong>in</strong>g process.<br />

The assumption that the particles deviate on geodesic l<strong>in</strong>es orthogonal to those jo<strong>in</strong><strong>in</strong>g the<br />

orig<strong>in</strong> O with their current position is crucial s<strong>in</strong>ce it makes the <strong>Hyperbolic</strong> Pythagorean Theorem<br />

applicable. Otherwise it should be necessary to apply the Carnot hyperbolic formula and this<br />

would make the analytic treatment of the problem extremely difficult.<br />

Under the condition that N(t) = n, the hyperbolic distance of the center of mass of the cloud<br />

of particles at time t is the random variable<br />

n−1 <br />

k=0<br />

<br />

1<br />

2k+1 k+1<br />

j=1<br />

cosh c(Sj − Sj−1)I {n>0} + 1<br />

2n n+1<br />

<br />

cosh c(Sj − Sj−1). (5.3)<br />

We observe that the n <strong>in</strong>stants S1, · · · , Sn where the fissions take place are uniformly distributed<br />

under the condition that N(t) = n, and possess density<br />

j=1<br />

P{S1 ∈ ds1, · · · , Sn ∈ dsn} = n!<br />

t n ds1 · · · dsn<br />

for 0 = s0 < s1 < · · · < sn+1 = t. Therefore, under the condition that the number of splitt<strong>in</strong>g<br />

events is N(t) = n, the hyperbolic distance ηk(t) of the k-th spl<strong>in</strong>ter is for k = 0, · · · n − 1<br />

E{cosh ηk(t)|N(t) = n} = n!<br />

t n<br />

and for k = n<br />

= n!<br />

t n<br />

t<br />

0<br />

t<br />

E{cosh ηn(t)|N(t) = n} = n!<br />

t n<br />

0<br />

t<br />

ds1 · · ·<br />

sk−1<br />

t<br />

ds1 · · ·<br />

sk−1<br />

t<br />

dsk · · ·<br />

sn−1<br />

dsn<br />

k+1 <br />

j=1<br />

(t − sk)<br />

dsk<br />

n−k k+1<br />

(n − k)!<br />

j=1<br />

cosh c(sj − sj−1)<br />

<br />

cosh c(sj − sj−1)<br />

= n!<br />

t n Gn,k(t), (5.4)<br />

t<br />

0<br />

t<br />

ds1 · · ·<br />

sn−1<br />

dsn<br />

n+1 <br />

j=1<br />

cosh c(sj − sj−1)<br />

= n!<br />

t n Gn,n(t). (5.5)<br />

The branch<strong>in</strong>g process described above can be adapted to the Po<strong>in</strong>caré disk B: s<strong>in</strong>ce the mapp<strong>in</strong>g<br />

(2.30) preserves the hyperbolic distance, the trajectories of splitt<strong>in</strong>g and mov<strong>in</strong>g particles can be<br />

conveniently depicted <strong>in</strong> B as well as H 2 (see Figure 5.1 and 5.2).<br />

We restrict ourselves to the mean hyperbolic distance because this leads to f<strong>in</strong>e explicit results.<br />

The analysis of the distribution of the hyperbolic distance, even <strong>in</strong> the case of a random motion of


5.2 Mean hyperbolic distance: the Laplace transform approach 86<br />

an <strong>in</strong>dividual particle that changes direction at all Poisson events, implies a much more complicated<br />

analysis and is almost <strong>in</strong>tractable s<strong>in</strong>ce multiple <strong>in</strong>tegrals of the form<br />

must be evaluated.<br />

0<br />

0<br />

E{e iα cosh η(t) |N(t) = n} = n!<br />

t n<br />

●<br />

O<br />

0<br />

(a) N(t)=0<br />

● 1 2 3<br />

●<br />

1 2 3<br />

●<br />

O<br />

0<br />

(d) N(t)=3<br />

●<br />

1 2 2<br />

1<br />

●<br />

1 2<br />

●<br />

0<br />

0<br />

t<br />

0<br />

t<br />

ds1 · · ·<br />

●<br />

O<br />

0<br />

(b) N(t)=1<br />

1 2<br />

●<br />

3<br />

1 2 3<br />

●<br />

●<br />

1 2 4<br />

●<br />

O<br />

0<br />

(e) N(t)=4<br />

sn−1<br />

●<br />

●<br />

1 2<br />

1 2 2<br />

e iα Q n+1<br />

j=1 cosh c(sj−sj−1) dsn<br />

1 2<br />

●<br />

1 2<br />

●<br />

0<br />

0<br />

1 2 2<br />

●<br />

●<br />

O<br />

0<br />

(c) N(t)=2<br />

3<br />

∏ coshc(Sj − Sj−1) 4<br />

j=1<br />

∏ coshc(Sj − Sj−1) ●<br />

j=1 5<br />

●<br />

∏ coshc(Sj − Sj−1) j=1<br />

0<br />

(f) N(t)=4<br />

●<br />

1 2 2<br />

1 2<br />

●<br />

●<br />

coshcS1coshc(t − S1) Figure 5.2: Same trajectories as <strong>in</strong> Figure 5.1 represented through the Po<strong>in</strong>caré disk model.<br />

5.2 Mean hyperbolic distance: the Laplace transform ap-<br />

proach<br />

We are able to obta<strong>in</strong> the explicit form of the mean-value of (5.2) by means of two different and<br />

<strong>in</strong>dependent approaches. In this section we present the Laplace-transform derivation which leads<br />

to our first theorem.<br />

Theorem 5.2.1. The mean-value of the hyperbolic distance (5.2) of the center of mass is<br />

E{cosh ηcm(t)} = 23c2 <br />

3 −<br />

e 22 λt<br />

√<br />

λ2 + 24c2 t −<br />

e 22 √<br />

λ2 +24c2 3 √ λ2 + 24c2 + 5λ +<br />

e t<br />

22 √<br />

λ2 +24c2 3 √ λ 2 + 2 4 c 2 − 5λ<br />

+ λ + 2c<br />

2(λ + 3c) ect λ − 2c<br />

+<br />

2(λ − 3c) e−ct , t > 0.<br />

●<br />

O<br />

<br />

coshct<br />

●<br />

●<br />

(5.6)


87 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

Proof<br />

In view of (5.3), (5.4) and (5.5), we first note that<br />

E{cosh ηcm(t)|N(t) = n} = n!<br />

tn n−1 <br />

k=0<br />

=<br />

+ n!<br />

tn 1<br />

2n ⎧<br />

⎪⎨<br />

⎪⎩<br />

1<br />

2 k+1<br />

t<br />

0<br />

n!<br />

tn n−1 k=0<br />

t<br />

0<br />

t<br />

ds1 · · ·<br />

t<br />

ds1 · · ·<br />

Our task is therefore to study the Laplace transform,<br />

∞<br />

0<br />

e −µt E{cosh ηcm(t)}dt =<br />

=<br />

sn−1<br />

sn−1<br />

dsn<br />

1<br />

2k+1 Gn,k(t) + n!<br />

tn dsn<br />

n+1 <br />

j=1<br />

k+1 <br />

j=1<br />

cosh c(sj − sj−1)<br />

cosh c(sj − sj−1)<br />

1<br />

2 n Gn,n(t), n ≥ 1,<br />

G0,0(t), n = 0.<br />

∞<br />

0<br />

∞<br />

e −µt<br />

λ n<br />

n−1 <br />

∞<br />

E{cosh ηcm(t)|N(t) = n}P{N(t) = n}dt<br />

n=0<br />

1<br />

2k+1 n=1 k=0<br />

∞ λ<br />

+<br />

n=0<br />

n<br />

2n ∞<br />

0<br />

∞<br />

0<br />

e −(λ+µ)t Gn,k(t)dt<br />

e −(λ+µ)t Gn,n(t)dt, (5.7)<br />

where µ > 0. We evaluate the Laplace transform appear<strong>in</strong>g <strong>in</strong> (5.7) <strong>in</strong> the follow<strong>in</strong>g way. If<br />

γ = λ + µ and c < γ, then we have, by successively <strong>in</strong>vert<strong>in</strong>g the <strong>in</strong>ner <strong>in</strong>tegrals, that<br />

∞<br />

e −γt Gn,k(t)dt<br />

∞<br />

=<br />

0<br />

e −γt<br />

⎧<br />

⎨ t<br />

⎩ 0<br />

∞ ∞<br />

= ds1 e −γt ⎧<br />

⎨<br />

dt<br />

⎩<br />

0<br />

=<br />

=<br />

=<br />

0<br />

∞<br />

0<br />

∞<br />

ds1<br />

0<br />

∞<br />

×<br />

0<br />

∞<br />

0<br />

s1<br />

∞<br />

s1<br />

∞<br />

ds1 · · ·<br />

t<br />

ds1 · · ·<br />

t<br />

sk−1<br />

s1<br />

∞<br />

ds2 · · ·<br />

sk−1<br />

k−1 <br />

dsk−1<br />

sk−2 j=1<br />

−γw wn−k<br />

e<br />

⎫<br />

n−k k+1<br />

(t − sk) <br />

⎬<br />

dsk<br />

cosh c(sj − sj−1)<br />

(n − k)!<br />

⎭<br />

j=1<br />

dt<br />

⎫<br />

t<br />

n−k k+1<br />

(t − sk) <br />

⎬<br />

ds2 · · · dsk<br />

cosh c(sj − sj−1)<br />

sk−1 (n − k)!<br />

⎭<br />

j=1<br />

k<br />

<br />

∞<br />

n−k<br />

−γt (t − sk)<br />

dsk cosh c(sj − sj−1) e cosh c(t − sk)dt<br />

(n − k)!<br />

j=1<br />

cosh c(sj − sj−1)<br />

∞<br />

sk−1<br />

cosh cw dw<br />

(n − k)!<br />

e −γw k ∞<br />

−γw wn−k<br />

cosh cw e cosh cw dw.<br />

(n − k)!<br />

0<br />

sk<br />

e −γsk cosh c(sk − sk−1)dsk<br />

In the last step above the change of variable sj − sj−1 = wj applied k times leads to the f<strong>in</strong>al<br />

expression. For k = n, from the previous calculations, we obta<strong>in</strong><br />

∞<br />

0<br />

e −γt Gn,n(t)dt =<br />

∞<br />

0<br />

e −γw n+1 cosh cw dw .


5.2 Mean hyperbolic distance: the Laplace transform approach 88<br />

S<strong>in</strong>ce<br />

we have that<br />

∞<br />

0<br />

∞<br />

0<br />

∞<br />

e<br />

0<br />

−γw γ<br />

cosh cw dw =<br />

γ2 ,<br />

− c2 <br />

−γw wn−k<br />

1 1<br />

1<br />

e cosh cw dw = +<br />

(n − k)! 2 (γ − c) n−k+1 (γ + c) n−k+1<br />

<br />

,<br />

e −µt E{cosh ηcm(t)}dt<br />

∞<br />

= λ<br />

n=1<br />

n<br />

n−1 1<br />

2<br />

k=0<br />

k+1<br />

<br />

γ<br />

γ2 − c2 k <br />

1 1<br />

1<br />

+<br />

2 (γ − c) n−k+1 (γ + c) n−k+1<br />

<br />

∞<br />

n <br />

λ γ<br />

+<br />

2 γ<br />

n=0<br />

2 − c2 n+1 = 1<br />

22 ∞<br />

λ n<br />

<br />

1<br />

(γ − c) n+1<br />

n−1 <br />

k γ<br />

1<br />

+<br />

2(γ + c) (γ + c) n+1<br />

n−1 <br />

<br />

k<br />

γ<br />

2(γ − c)<br />

n=1<br />

k=0<br />

2γ<br />

+<br />

2γ2 − 2c2 , (5.8)<br />

− γλ<br />

where the last sum converges if µ satisfies the <strong>in</strong>equality 2c 2 < λ 2 + 2µ 2 + 3λµ. The double sum<br />

<strong>in</strong> (5.8) can be calculated by <strong>in</strong>vert<strong>in</strong>g the order of summation <strong>in</strong> the follow<strong>in</strong>g way:<br />

∞<br />

λ<br />

n=1<br />

n<br />

<br />

1<br />

(γ − c) n+1<br />

n−1 <br />

k γ<br />

1<br />

+<br />

2(γ + c) (γ + c)<br />

k=0<br />

n+1<br />

n−1 <br />

<br />

k<br />

γ<br />

2(γ − c)<br />

k=0<br />

∞<br />

k ∞ n 1 γ<br />

λ<br />

=<br />

+<br />

γ − c 2(γ + c) γ − c<br />

k=0<br />

n=k+1<br />

1<br />

∞<br />

k ∞ n γ<br />

λ<br />

γ + c 2(γ − c) γ + c<br />

k=0<br />

n=k+1<br />

∞<br />

k k+1 ∞ r 1 γ λ<br />

λ<br />

=<br />

γ − c 2(γ + c) γ − c γ − c<br />

k=0<br />

r=0<br />

+ 1<br />

∞<br />

k k+1 ∞ r γ λ<br />

λ<br />

γ + c 2(γ − c) γ + c γ + c<br />

k=0<br />

r=0<br />

λ<br />

=<br />

(γ − c) 2<br />

∞<br />

<br />

γλ<br />

2(γ<br />

k=0<br />

2 − c2 k γ − c<br />

) γ − c − λ +<br />

λ<br />

(γ + c) 2<br />

∞<br />

<br />

γλ<br />

2(γ<br />

k=0<br />

2 − c2 k γ + c<br />

) γ + c − λ<br />

<br />

1 1 1 1 2(γ<br />

= λ<br />

+<br />

γ − c γ − c − λ γ + c γ + c − λ<br />

2 − c2 )<br />

2γ2 − 2c2 . (5.9)<br />

− γλ<br />

The <strong>in</strong>version of the Laplace transform is made possible by suitably rearrang<strong>in</strong>g the expression<br />

(5.9) as follows:<br />

∞<br />

e<br />

0<br />

−µt E{cosh ηcm(t)}dt<br />

= λ<br />

<br />

1 1 1 1 γ<br />

+<br />

2 γ − c γ − c − λ γ + c γ + c − λ<br />

2 − c2 2γ2 − 2c2 − γλ +<br />

2γ<br />

2γ2 − 2c2 − γλ<br />

= λ<br />

<br />

1 1<br />

2 λ + µ − c µ − c +<br />

<br />

1 1 (λ + µ)<br />

λ + µ + c µ + c<br />

2 − c2 λ2 + 2µ 2 2λ + 2µ<br />

+<br />

+ 3λµ − 2c2 λ2 + 2µ 2 + 3λµ − 2c2 = λ<br />

<br />

λ + µ + c λ + µ − c<br />

1<br />

+<br />

2 µ − c µ + c λ2 + 2µ 2 2λ + 2µ<br />

+<br />

+ 3λµ − 2c2 λ2 + 2µ 2 . (5.10)<br />

+ 3λµ − 2c2 k=0


89 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

By means of the decomposition<br />

2µ 2 + 3λµ + λ 2 − 2c 2 = 2<br />

<br />

µ + 3λ − √ λ 2 + 2 4 c 2<br />

2 2<br />

the expression (5.10) can be further worked out by writ<strong>in</strong>g<br />

∞<br />

0<br />

e −µt E{cosh ηcm(t)}dt<br />

= λ<br />

2 2<br />

=<br />

+<br />

<br />

+<br />

+<br />

<br />

<br />

2<br />

µ + 3λ−√ λ 2 +2 4 c 2<br />

2 2<br />

3<br />

2<br />

µ + 3λ−√ λ 2 +2 4 c 2<br />

2 2<br />

1<br />

µ + 3λ−√λ2 +24c2 22 <br />

<br />

1<br />

1<br />

<br />

λ λ + 2 + 2µ<br />

<br />

−<br />

µ + 3λ−√ λ 2 +2 4 c 2<br />

2 2<br />

1<br />

µ + 3λ−√ λ 2 +2 4 c 2<br />

2 2<br />

µ + 3λ+√ λ 2 +2 4 c 2<br />

2 2<br />

µ + 3λ+√λ2 +24c2 22 1<br />

µ + 3λ+√ λ 2 +2 4 c 2<br />

2 2<br />

−<br />

+<br />

1<br />

µ + 3λ+√ λ 2 +2 4 c 2<br />

2 2<br />

1<br />

µ + 3λ+√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

<br />

<br />

<br />

<br />

λ + 2c λ − 2c<br />

2 + +<br />

µ − c µ + c<br />

<br />

<br />

λ 1<br />

√ +<br />

λ2 + 24c2 2<br />

<br />

λ<br />

2 √ λ2 + 24c2 <br />

λ + 2c<br />

µ − c<br />

<br />

λ<br />

22 (2µ + 3<br />

2λ). µ + 3λ + √ λ 2 + 2 4 c 2<br />

2 2<br />

1<br />

µ + 3λ−√λ2 +24c2 22 + λ − 2c<br />

µ + c<br />

<br />

+<br />

<br />

,<br />

1<br />

µ + 3λ+√ λ 2 +2 4 c 2<br />

2 2<br />

It is now a simple matter to <strong>in</strong>vert the Laplace transform (5.11) and we arrive at the mean<br />

hyperbolic distance of the center of mass <strong>in</strong> an <strong>in</strong>tegral form<br />

E{cosh ηcm(t)}<br />

=<br />

1<br />

2 +<br />

λ<br />

√ λ 2 + 2 4 c 2<br />

+ λ(λ + 2c)<br />

2 √ λ2 + 24c2 + λ(λ − 2c)<br />

2 √ λ2 + 24c2 + λ<br />

2 3<br />

t<br />

0<br />

t<br />

√<br />

3λ− λ2 +24c2 −t<br />

e 22 <br />

1<br />

+<br />

2 −<br />

e cs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 ds −<br />

√<br />

λ<br />

3λ+ λ2 +24c2 −t<br />

√ e 2<br />

λ2 + 24c2 2<br />

t<br />

0<br />

0<br />

t<br />

e<br />

0<br />

−cs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 t<br />

ds −<br />

0<br />

3 −<br />

e 22 λs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 t<br />

ds +<br />

0<br />

e cs √<br />

3λ+ λ2 +24c2 −(t−s)<br />

e 22 <br />

ds<br />

e −cs √<br />

3λ+ λ2 +24c2 −(t−s)<br />

e 22 <br />

ds<br />

3 −<br />

e 22 λs 3λ+<br />

−(t−s)<br />

e<br />

√<br />

λ2 +24c2 22 ds<br />

<br />

. (5.12)<br />

The expression (5.12) can be further developed and simplified by observ<strong>in</strong>g that, after some simple<br />

calculations, we have that<br />

λ(λ + 2c)<br />

2 √ λ 2 + 2 4 c 2<br />

=<br />

t<br />

0<br />

e cs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 ds −<br />

t<br />

λ(λ + 2c)<br />

2 √ λ2 + 24c2 22 22c + 3λ − √ λ2 + 24c2 <br />

e ct 3λ−<br />

−t<br />

− e<br />

λ(λ + 2c)<br />

−<br />

2 √ λ2 + 24c2 22 22c + 3λ + √ λ2 + 24c2 <br />

e ct 3λ+<br />

−t<br />

− e<br />

0<br />

e cs √<br />

3λ+ λ2 +24c2 −(t−s)<br />

e 22 <br />

ds<br />

√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

. (5.13)<br />

<br />

(5.11)


5.2 Mean hyperbolic distance: the Laplace transform approach 90<br />

Similar manipulations yield<br />

and also<br />

λ(λ − 2c)<br />

2 √ λ 2 + 2 4 c 2<br />

=<br />

t<br />

0<br />

e −cs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 ds −<br />

t<br />

λ(λ − 2c)<br />

2 √ λ2 + 24c2 22 −22c + 3λ − √ λ2 + 24c2 <br />

e −ct 3λ−<br />

−t<br />

− e<br />

λ(λ − 2c)<br />

−<br />

2 √ λ2 + 24c2 22 −22c + 3λ + √ λ2 + 24c2 <br />

e −ct 3λ+<br />

−t<br />

− e<br />

λ<br />

23 t<br />

3 −<br />

e 2<br />

0<br />

2 λs √<br />

3λ− λ2 +24c2 −(t−s)<br />

e 22 ds +<br />

=<br />

λ<br />

2 √ λ 2 + 2 4 c<br />

<br />

3<br />

e− 22 λt<br />

2<br />

e t<br />

0<br />

t<br />

0<br />

√ λ 2 +2 4 c 2<br />

2 2 − e −t<br />

e −cs √<br />

3λ+ λ2 +24c2 −(t−s)<br />

e 22 <br />

ds<br />

√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

, (5.14)<br />

3 −<br />

e 22 λs √<br />

3λ+ λ2 +24c2 −(t−s)<br />

e 22 <br />

ds<br />

√ λ 2 +2 4 c 2<br />

2 2<br />

<br />

. (5.15)<br />

By <strong>in</strong>sert<strong>in</strong>g results (5.15), (5.13), and (5.14) <strong>in</strong>to (5.12) we now obta<strong>in</strong> the f<strong>in</strong>al formula<br />

E{cosh ηcm(t)}<br />

= e− 3<br />

22 λt <br />

<br />

3λ<br />

1 + √ e<br />

2<br />

λ2 + 24c2 t<br />

22 <br />

<br />

√<br />

λ2 +24c2 3λ<br />

t −<br />

+ 1 − √ e 2<br />

λ2 + 24c2 2<br />

<br />

√<br />

λ2 +24c2 + λ + 2c<br />

2(λ + 3c) ect √<br />

2λ<br />

3λ− λ2 +24c2 + √ e−t 2<br />

λ2 + 24c2 2<br />

<br />

λ − 2c<br />

22c − 3λ + √ λ2 + 24 λ + 2c<br />

−<br />

c2 22c + 3λ − √ λ2 + 24c2 <br />

+ λ − 2c<br />

2(λ − 3c) e−ct √<br />

2λ<br />

3λ+ λ2 +24c2 + √ e−t 2<br />

λ2 + 24c2 2<br />

<br />

λ + 2c<br />

22c + 3λ + √ λ2 + 24 λ − 2c<br />

−<br />

c2 22c − 3λ − √ λ2 + 24c2 <br />

3 −<br />

= e 22 λt e t<br />

22 <br />

√<br />

λ2 +24c2 1<br />

2 +<br />

3λ<br />

2 √ λ2 + 24 <br />

2 3λ<br />

− √<br />

c2 λ2 + 24c2 2 − λ √ λ2 + 24c2 − 23c2 5λ − 3 √ λ2 + 24c2 <br />

3 −<br />

+e 22 λt t −<br />

e 22 <br />

√<br />

λ2 +24c2 1<br />

2 −<br />

3λ<br />

2 √ λ2 + 24 <br />

2 3λ<br />

+ √<br />

c2 λ2 + 24c2 2 + λ √ λ2 + 24c2 − 23c2 5λ + 3 √ λ2 + 24c2 <br />

<br />

+ λ + 2c<br />

2(λ + 3c) ect λ − 2c<br />

+<br />

2(λ − 3c) e−ct<br />

= 23c2 3<br />

t<br />

−<br />

e 22 λt −<br />

e 2<br />

√<br />

λ2 + 24c2 2<br />

√<br />

λ2 +24c2 5λ + 3 √ λ2 + 24 e<br />

−<br />

c2 t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ2 + 24c2 <br />

+ λ + 2c<br />

2(λ + 3c) ect λ − 2c<br />

+<br />

2(λ − 3c) e−ct .<br />

Remark 5.2.1. Apparently a critical po<strong>in</strong>t <strong>in</strong> formula (5.6) is λ = 3c. We show that the mean<br />

hyperbolic distance (5.6) is f<strong>in</strong>ite for λ = 3c. We first write ε = λ−3c and r(ε) = √ ε 2 + 6cε + 5 2 c 2<br />

and evaluate the limit<br />

lim<br />

λ→3c<br />

3 −<br />

22 λt<br />

e t<br />

22 √<br />

λ2 +24c2 λ − 2c<br />

2(λ − 3c) e−ct − 23c2e √<br />

λ2 + 24c2 5λ − 3 √ λ2 + 24c2 ε + c<br />

= lim<br />

ε→0 2ε e−ct −<br />

23c2 t<br />

e− 2<br />

r(15c + 5ε − 3r) 2 (9c+3ε−r) , (5.16)<br />

which refers to the components of (5.6) with diverg<strong>in</strong>g coefficients. We expand the second expo-


91 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

1.04<br />

1.03<br />

1.02<br />

1.01<br />

* * * * * * * * * * * * * * * * * * * * * * * * *<br />

1.00<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

λ = 0.1c<br />

λ = 4c<br />

λ = 1c<br />

λ = 5c<br />

λ = 2c<br />

λ = 6c<br />

* λ = 3c<br />

* * * * * * * * * * * * * * * * * * * * * * * * * * * * * *<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

*<br />

0.10 0.15 0.20 0.25 0.30<br />

(a)<br />

*<br />

*<br />

*<br />

*<br />

1.036<br />

1.034<br />

1.032<br />

1.030<br />

1.028<br />

*<br />

*<br />

1.026<br />

*<br />

*<br />

*<br />

*<br />

0.250 0.260 0.270<br />

Figure 5.3: In (a) and (b) the mean-value E{cosh ηcm(t)} is plotted for c = 1 and different values<br />

of λ, the curve correspond<strong>in</strong>g to λ = 3c is obta<strong>in</strong>ed by plott<strong>in</strong>g the limit value E{cosh ηcm(t)} =<br />

e−ct 53<br />

2252 + 3<br />

10ct ct 5 + e 22 7ct<br />

3 + e− 2<br />

22<br />

523 .<br />

nential <strong>in</strong> (5.16) as<br />

t −<br />

e 22 (9c+3ε−r) = e −ct t −<br />

e 22 (5c+3ε−r) = e −ct<br />

<br />

1 − t<br />

t2<br />

(5c + 3ε − r) +<br />

22 25 (5c + 3ε − r)2 + o(ε 3 <br />

) ,<br />

where we have taken <strong>in</strong>to account that r(ε) → 5c for ε → 0. It is now convenient to write (5.16)<br />

<strong>in</strong> the form<br />

e −ct <br />

ε + c<br />

lim<br />

ε→0 2ε −<br />

23c2 r(15c + 5ε − 3r) + 2tc2 (5c + 3ε − r)<br />

r(15c + 5ε − 3r) − t2c2 (5c + 3ε − r) 2<br />

22r(15c + 5ε − 3r) + o(ε2 <br />

) .<br />

By consider<strong>in</strong>g that r(ε) ∼ 5c + 3<br />

5ε for ε → 0, we can easily realize that<br />

<br />

ε + c<br />

lim<br />

ε→0 2ε −<br />

23c2 <br />

=<br />

r(15c + 5ε − 3r)<br />

2 · 7<br />

2tc<br />

, lim<br />

52 ε→0<br />

2 (5c + 3ε − r) 3 t<br />

= ct, lim<br />

r(15c + 5ε − 3r) 2 · 5 ε→0<br />

2c2 (5c + 3ε − r) 2<br />

22 = 0.<br />

r(15c + 5ε − 3r)<br />

This suffices to show that near λ = 3c the mean hyperbolic distance is f<strong>in</strong>ite. Consider<strong>in</strong>g also the<br />

two additional terms of (5.6) leads to the follow<strong>in</strong>g asymptotic estimate of the mean hyperbolic<br />

distance of the center of mass for large values of t and near λ ∼ 3c. In particular, we have that<br />

E{cosh ηcm(t)} ∼ 5<br />

12 ect .<br />

In Figure 5.3, the function E{cosh ηcm(t)} is plotted for c = 1 and different values of λ <strong>in</strong> the<br />

neighborhood of λ = 3c (<strong>in</strong>clud<strong>in</strong>g the limit<strong>in</strong>g case λ = 3c).<br />

Remark 5.2.2. We now show that the center of mass (as well as each <strong>in</strong>dividual particle) goes<br />

further and further away from the orig<strong>in</strong> O of H 2 (or equivalently it migrates towards the frontier<br />

of the Po<strong>in</strong>caré disc B) as time t <strong>in</strong>creases. In order to show this result we work on (5.12) and,<br />

(b)<br />

*<br />

*<br />

*<br />

*


5.3 Equation govern<strong>in</strong>g the hyperbolic distance 92<br />

after some manipulations, we obta<strong>in</strong> that<br />

d<br />

dt E{cosh ηcm(t)} = 22c2 3 −<br />

e 22 λt<br />

t<br />

√ s<strong>in</strong>h<br />

λ2 + 24c2 22 <br />

λ2 + 24c2 2cλ<br />

+<br />

2<br />

√<br />

λ2 + 24c2 + 22 c 2 λ<br />

√ λ 2 + 2 4 c 2<br />

t<br />

0<br />

t<br />

0<br />

3 −<br />

e 22 λs s<strong>in</strong>h c(t − s) s<strong>in</strong>h s<br />

22 3 −<br />

e 22 λs cosh c(t − s) s<strong>in</strong>h s<br />

22 λ 2 + 2 4 c 2 ds<br />

λ 2 + 2 4 c 2 ds. (5.17)<br />

The first term <strong>in</strong> the right hand side of (5.17) is produced by the derivatives of all terms of (5.12)<br />

while the <strong>in</strong>tegrals stem from the third and fourth term only. From (5.17) it is easy to check that<br />

d 2<br />

dt 2 E{cosh ηcm(t)} > 0 for all t so that the center of mass gets off from the start<strong>in</strong>g po<strong>in</strong>t with<br />

positive acceleration.<br />

0<br />

0<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

● ●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

● ● ●<br />

O<br />

●<br />

0<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

c=0.05, t=40, N(t)=20<br />

(a)<br />

●<br />

●<br />

●<br />

● ● ● ●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●● ●<br />

● ●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ● ●● ● ● ●<br />

0<br />

c=0.05, t=50, N(t)=300<br />

(d)<br />

●<br />

●<br />

●<br />

● ●● ● ●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●●●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

O<br />

●<br />

● ●<br />

●<br />

●<br />

● ● ●●● ● ● ● ●<br />

●<br />

● ● ●●<br />

●<br />

●<br />

●<br />

● ●<br />

● ●●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

0<br />

0<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

O<br />

●<br />

0<br />

c=0.05, t=40, N(t)=20<br />

(b)<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

● ●<br />

●<br />

●<br />

●●<br />

● ●<br />

●<br />

● ●<br />

●<br />

●<br />

● ●<br />

0<br />

c=0.05, t=50, N(t)=300<br />

(e)<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●● ●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●●●<br />

●●● ● ●● ●<br />

●<br />

●● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●● O<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

0<br />

0<br />

●<br />

●<br />

●<br />

O<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

0<br />

c=0.05, t=40, N(t)=20<br />

(c)<br />

0<br />

●<br />

●<br />

●<br />

●<br />

●<br />

c=0.05, t=50, N(t)=300<br />

(f)<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

● ●<br />

● ●<br />

● ●<br />

●<br />

●<br />

● ●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

● ●<br />

●<br />

●●<br />

● ●●<br />

●<br />

●<br />

●●<br />

●<br />

●●<br />

● ● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

O ●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

● ●●<br />

●<br />

● ● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●●<br />

●<br />

●<br />

●<br />

● ● ●●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

● ●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●●<br />

●<br />

●●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

● ● ●<br />

●<br />

●●<br />

●●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

Figure 5.4: In (a), (b), and (c) the trajectories and the positions of the spl<strong>in</strong>ters at time t = 40,<br />

when c = 0.05 and N(t) = 20, are shown. In (d), (e) and (f) only the positions of the spl<strong>in</strong>ters at<br />

time t = 50, when c = 0.05 and N(t) = 300, are drawn. In (a) and (d) each spl<strong>in</strong>ter chooses the<br />

clockwise direction, <strong>in</strong> (b) and (e) each spl<strong>in</strong>ter chooses the counterclockwise direction, <strong>in</strong> (c) and<br />

(f) the clockwise and counterclockwise directions are alternatively chosen.<br />

5.3 Equation govern<strong>in</strong>g the hyperbolic distance<br />

We are able to confirm result (5.6) by a completely different method based on a system of differential<br />

equations govern<strong>in</strong>g all the quantities appear<strong>in</strong>g <strong>in</strong> the mean hyperbolic distance of the center of


93 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

mass. In view of formulas (5.4) and (5.5) we can write the mean hyperbolic distance of the center<br />

of mass <strong>in</strong> the follow<strong>in</strong>g form<br />

E{cosh ηcm(t)} = e −λt<br />

∞<br />

n=1<br />

λ n<br />

n−1 <br />

k=0<br />

1<br />

2 k+1 Gn,k(t) + e −λt<br />

∞<br />

n=0<br />

λ n<br />

2 n Gn,n(t). (5.18)<br />

We start by deriv<strong>in</strong>g the difference-differential equations govern<strong>in</strong>g the functions Gn,k(t), with<br />

t > 0 and 0 ≤ k ≤ n. As it has been po<strong>in</strong>ted out <strong>in</strong> the previous chapter, the j-th moment of<br />

the conditional hyperbolic distance of a s<strong>in</strong>gle particle leads to a difference-differential equation of<br />

order equal to j +1. The same type of phenomenon occurs <strong>in</strong> the present case where the hyperbolic<br />

distance of a cloud of po<strong>in</strong>ts is envisaged.<br />

Lemma 5.3.1. The functions represented by the follow<strong>in</strong>g multiple <strong>in</strong>tegrals<br />

Gn,k(t) =<br />

t<br />

0<br />

t<br />

ds1 · · ·<br />

sk−1<br />

dsk<br />

are solutions to the difference-differential equations<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

d 2<br />

n−k<br />

(t − sk) k<br />

cosh c(sj − sj−1) cosh c(t − sk), 0 ≤ k ≤ n,<br />

(n − k)!<br />

j=1<br />

dt2 Gn,k = 2 d<br />

dtGn−1,k − Gn−2,k + c2Gn,k, k ≤ n − 2,<br />

d 2<br />

dt2 Gn,n−1 = 2 d<br />

dtGn−1,n−1 − Gn−2,n−2 + c2Gn,n−1, k = n − 1,<br />

d 2<br />

dt 2 Gn,n = d<br />

dt Gn−1,n−1 + c 2 Gn,n, k = n.<br />

Proof<br />

We start by consider<strong>in</strong>g the case k ≤ n − 2 where the first derivative reads<br />

d<br />

dt Gn,k = Gn−1,k + c<br />

t<br />

0<br />

t<br />

ds1 · · ·<br />

sk−1<br />

and, from (5.20), the second derivative becomes<br />

d 2<br />

dt 2 Gn,k = d<br />

dt Gn−1,k + c 2 Gn,k<br />

+c<br />

t<br />

0<br />

ds1 · · ·<br />

t<br />

sk−1<br />

(t − sk)<br />

dsk<br />

n−k<br />

(n − k)!<br />

(t − sk)<br />

dsk<br />

n−k−1<br />

(n − k − 1)!<br />

k<br />

j=1<br />

(5.19)<br />

cosh c(sj − sj−1) s<strong>in</strong>h c(t − sk), (5.20)<br />

k<br />

cosh c(sj − sj−1) s<strong>in</strong>h c(t − sk)<br />

= 2 d<br />

dt Gn−1,k − Gn−2,k + c 2 Gn,k. (5.21)<br />

The expression (5.20), for k = n − 1, is qualitatively different and must be handled carefully. In<br />

this case we have that<br />

d<br />

dt Gn,n−1 = Gn−1,n−1<br />

t<br />

+c<br />

0<br />

ds1 · · ·<br />

t<br />

sn−2<br />

<br />

j=1<br />

n−1<br />

dsn−1(t − sn−1)<br />

j=1<br />

cosh c(sj − sj−1) s<strong>in</strong>h c(t − sn−1), (5.22)


5.3 Equation govern<strong>in</strong>g the hyperbolic distance 94<br />

and thus, from (5.22),<br />

d 2<br />

dt 2 Gn,n−1 = d<br />

dt Gn−1,n−1 + c 2 Gn,n−1<br />

+c<br />

t<br />

0<br />

ds1 · · ·<br />

= d<br />

dt Gn−1,n−1 +<br />

t<br />

n−1 <br />

dsn−1<br />

sn−2 j=1<br />

d<br />

dt Gn−1,n−1 − Gn−2,n−2<br />

= 2 d<br />

dt Gn−1,n−1 − Gn−2,n−2 + c 2 Gn,n−1.<br />

cosh c(sj − sj−1) s<strong>in</strong>h c(t − sn−1)<br />

<br />

+ c 2 Gn,n−1<br />

We omit the derivation of the last formula <strong>in</strong> (5.19) because it co<strong>in</strong>cides with the result of Lemma<br />

4.2.1. <br />

The results of Lemma 5.3.1 permit us to obta<strong>in</strong> the equation govern<strong>in</strong>g the mean-value of<br />

the hyperbolic distance of the center of mass of the randomly mov<strong>in</strong>g spl<strong>in</strong>ters produced by the<br />

dis<strong>in</strong>tegration process.<br />

Theorem 5.3.2. The mean-value E{cosh ηcm(t)} <strong>in</strong> (5.2) is solution to the non-homogeneus<br />

second-order l<strong>in</strong>ear equation<br />

3 −<br />

22 λt<br />

d2 dt2 u − c2u = λc2e √<br />

λ2 + 24c2 Proof<br />

We beg<strong>in</strong> by successively deriv<strong>in</strong>g the expression (5.18) as follows<br />

and<br />

d<br />

u = −λu + e−λt<br />

dt<br />

∞<br />

n=1<br />

d2 d<br />

u = −2λ<br />

dt2 dt u − λ2u + e −λt<br />

λ n<br />

n−1 <br />

k=0<br />

∞<br />

n=1<br />

t −<br />

e 22 √<br />

λ2 +24c2 − e t<br />

22 √<br />

λ2 +24c2 <br />

. (5.23)<br />

1<br />

2 k+1<br />

λ n<br />

n−1 <br />

k=0<br />

d<br />

dt Gn,k + e −λt<br />

1<br />

2 k+1<br />

∞<br />

n=0<br />

λ n<br />

2 n<br />

d 2<br />

dt 2 Gn,k + e −λt<br />

d<br />

dt Gn,n,<br />

∞<br />

n=0<br />

λn 2n d2 dt2 Gn,n. (5.24)<br />

We must now <strong>in</strong>sert all the expressions of Lemma 5.3.1 <strong>in</strong>to (5.24) by accurately tak<strong>in</strong>g <strong>in</strong>to account


95 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

the constra<strong>in</strong>s on n and k. It is convenient to write the last two terms of (5.24) as follows<br />

∞<br />

n=1<br />

=<br />

=<br />

λ n<br />

n−1 <br />

k=0<br />

∞<br />

n=2<br />

+<br />

∞<br />

n=2<br />

+<br />

+<br />

1<br />

2 k+1<br />

λ n<br />

n−2 <br />

∞<br />

n=1<br />

k=0<br />

λ n<br />

2 n<br />

λ n<br />

n−2 <br />

∞<br />

n=2<br />

∞<br />

n=1<br />

k=0<br />

λ n<br />

2 n<br />

λ n<br />

2 n<br />

d 2<br />

dt 2 Gn,k +<br />

1<br />

2 k+1<br />

d 2<br />

∞<br />

n=0<br />

d 2<br />

dt 2 Gn,k +<br />

λn 2n d2 Gn,n<br />

dt2 ∞<br />

n=2<br />

dt2 Gn,n + d2<br />

G0,0<br />

dt2 λ n<br />

2 n<br />

d2 dt2 Gn,n−1 + λ d<br />

2<br />

2<br />

<br />

2 d<br />

dt Gn−1,k − Gn−2,k + c 2 <br />

Gn,k<br />

G1,0<br />

dt2 1<br />

2k+1 <br />

2 d<br />

dt Gn−1,n−1 − Gn−2,n−2 + c 2 <br />

Gn,n−1 + λ d<br />

dt G0,0 + λ<br />

2 c2G1,0 <br />

d<br />

dt Gn−1,n−1 + c 2 <br />

Gn,n + c 2 G0,0. (5.25)<br />

By regroup<strong>in</strong>g the above terms we notice that (5.25) can take the form<br />

c 2<br />

∞<br />

n=1<br />

λ n<br />

n−1 <br />

−λ 2<br />

1<br />

∞<br />

2<br />

k=0<br />

k+1 Gn,k +<br />

n=0<br />

∞<br />

n=1<br />

λ n<br />

n−1 <br />

k=0<br />

1<br />

2 k+1 Gn,k +<br />

λn <br />

Gn,n + 2λ<br />

2n d<br />

dt<br />

∞<br />

n=0<br />

= c 2 e λt u + 2λ d<br />

dt [eλtu] − λ 2 e λt u − λ<br />

2<br />

λn Gn,n<br />

2n ∞<br />

n=0<br />

<br />

∞<br />

n=1<br />

− λ<br />

2<br />

λ n<br />

n−1 <br />

∞<br />

n=0<br />

k=0<br />

λn 2n d<br />

dt Gn,n + λ2<br />

22 Multiply<strong>in</strong>g by e −λt the right hand side of (5.26), we have that<br />

e −λt<br />

∞<br />

n=1<br />

λ n<br />

n−1 <br />

k=0<br />

1<br />

2 k+1<br />

d 2<br />

dt 2 Gn,k + e −λt<br />

∞<br />

n=0<br />

= c 2 u + 2λ 2 u + 2λ d<br />

dt u − λ2 u − λ<br />

2 e−λt<br />

By <strong>in</strong>sert<strong>in</strong>g result (5.27) <strong>in</strong>to (5.24) we f<strong>in</strong>ally obta<strong>in</strong><br />

d 2<br />

dt 2 u − c2 u = − λ<br />

2 e−λt<br />

∞<br />

n=0<br />

λn 2n d2 Gn,n<br />

dt2 ∞<br />

n=0<br />

1<br />

2 k+1 Gn,k +<br />

λn 2n d<br />

dt Gn,n + λ2<br />

22 ∞<br />

n=0<br />

∞<br />

n=0<br />

∞<br />

n=0<br />

λn <br />

Gn,n<br />

2n λn Gn,n<br />

2n λ n<br />

2 n Gn,n. (5.26)<br />

∞<br />

λn 2n d<br />

dt Gn,n + λ2 λ<br />

e−λt<br />

22 n=0<br />

n<br />

2n Gn,n. (5.27)<br />

∞<br />

λn 2n d<br />

dt Gn,n + λ2 λ<br />

e−λt<br />

22 n=0<br />

n<br />

2n Gn,n. (5.28)<br />

λt − ∞<br />

The series e 2<br />

n=0 λn<br />

2n Gn,n has been studied <strong>in</strong> the previous chapter and represents the mean<br />

hyperbolic distance of the particle mov<strong>in</strong>g <strong>in</strong> H2 and chang<strong>in</strong>g direction at each Poisson event<br />

when the rate of the Poisson process is λ/2. We have that<br />

λt −<br />

e 2<br />

S<strong>in</strong>ce<br />

∞<br />

n=0<br />

λ n<br />

2 n Gn,n =<br />

λt −<br />

e 2<br />

e− λt<br />

4<br />

∞<br />

n=0<br />

2<br />

λ n<br />

2 n<br />

<br />

1 +<br />

<br />

λ<br />

√<br />

λ2 + 24c2 d<br />

dt Gn,n = d<br />

<br />

λt −<br />

e 2<br />

dt<br />

e t<br />

<br />

√<br />

4 λ2 +24c2 + 1 −<br />

∞<br />

n=0<br />

λn <br />

Gn,n<br />

2n + λ λt<br />

e− 2<br />

2<br />

<br />

λ<br />

√ e<br />

λ2 + 24c2 ∞<br />

n=0<br />

λn Gn,n,<br />

2n <br />

√<br />

t − 4 λ2 +24c2 .


5.3 Equation govern<strong>in</strong>g the hyperbolic distance 96<br />

the right-hand side of (5.28) can be rewritten as<br />

− λ<br />

2 e−λt<br />

∞ λ<br />

n=0<br />

n<br />

2n d<br />

dt Gn,n + λ2<br />

∞ λ<br />

e−λt<br />

22 n=0<br />

n<br />

2n Gn,n = − λ<br />

<br />

∞<br />

λt d λt<br />

e− 2 − λ<br />

e 2<br />

2 dt<br />

n=0<br />

n<br />

<br />

Gn,n<br />

2n = − λ<br />

<br />

λt<br />

e− 2 −<br />

2 λ λt<br />

e− 2<br />

23 2<br />

<br />

<br />

λ<br />

1 + √ e<br />

λ2 + 24c2 t<br />

22 <br />

<br />

√<br />

λ2 +24c2 λ<br />

t −<br />

+ 1 − √ e 2<br />

λ2 + 24c2 2<br />

<br />

√<br />

λ2 +24c2 √<br />

λ2 + 24c2 +<br />

23 λt −<br />

e 22 <br />

<br />

λ<br />

1 + √ e<br />

λ2 + 24c2 t<br />

22 <br />

<br />

√<br />

λ2 +24c2 λ<br />

t −<br />

− 1 − √ e 2<br />

λ2 + 24c2 2<br />

√<br />

λ2 +24c2 <br />

= λ2 3λt<br />

e− 2<br />

24 2<br />

<br />

<br />

λ<br />

1 + √ e<br />

λ2 + 24c2 t<br />

22 <br />

<br />

√<br />

λ2 +24c2 λ<br />

t −<br />

+ 1 − √ e 2<br />

λ2 + 24c2 2<br />

<br />

√<br />

λ2 +24c2 <br />

<br />

λ<br />

1 + √ e<br />

λ2 + 24c2 t<br />

22 <br />

<br />

√<br />

λ2 +24c2 λ<br />

t −<br />

− 1 − √ e 2<br />

λ2 + 24c2 2<br />

<br />

√<br />

λ2 +24c2 − λ√λ2 + 24c2 24 3λt −<br />

e 22 3λt −<br />

= −e 22 e t<br />

22 √<br />

λ2 +24c2 λc2 3λt<br />

√ + e− 2<br />

λ2 + 24c2 2 t −<br />

e 22 √<br />

λ2 +24c2 λc2 √<br />

λ2 + 24c2 .<br />

This lets equation (5.23) appear. <br />

The expression (5.6) of the mean-value E{cosh ηcm(t)} derived <strong>in</strong> Theorem 5.2.1 satisfies the<br />

non-homogeneous second-order l<strong>in</strong>ear equation (5.23) as shown <strong>in</strong> the next theorem. This proves<br />

that the mean hyperbolic distance (5.6) is obta<strong>in</strong>ed by two different and <strong>in</strong>dependent methods.<br />

Theorem 5.3.3. The mean hyperbolic distance E{cosh ηcm(t)} <strong>in</strong> (5.6) is the solution to the<br />

follow<strong>in</strong>g Cauchy problem<br />

⎧<br />

⎪⎨<br />

⎪⎩<br />

d 2<br />

dt2 u − c2u = λc2e u(0) = 1,<br />

d<br />

dtu(t) = 0.<br />

t=0<br />

− 3<br />

22 λt<br />

√<br />

λ2 +24c2 t −<br />

e 22 √<br />

λ2 +24c2 − e t<br />

22 √<br />

λ2 +24c2 <br />

,<br />

Proof<br />

In order to perform the calculations it is convenient to write (5.6) as<br />

where<br />

K =<br />

E{cosh ηcm(t)} = KA(t)B(t) +<br />

23c2 3<br />

√ , A(t) = e− 2<br />

λ2 + 24c2 2 λt , B(t) =<br />

and B(t) has derivatives<br />

B ′ (t) = −<br />

√ <br />

λ2 + 24c2 2 2<br />

t −<br />

e 22 √<br />

λ2 +24c2 5λ + 3 √ λ2 + 24 +<br />

c2 λ + 2c<br />

2(λ + 3c) ect λ − 2c<br />

+<br />

2(λ − 3c) e−ct<br />

t<br />

e− 22 √<br />

λ2 +24c2 5λ + 3 √ λ2 + 24 −<br />

c2 e t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ 2 + 2 4 c 2<br />

<br />

e t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ2 + 24 ,<br />

c2 , B ′′ (t) = λ2 + 24c2 24 B(t).


97 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

Therefore<br />

d2 dt2 E{cosh ηcm(t)} − c 2 E{cosh ηcm(t)} = λKA<br />

<br />

5λ<br />

B − 3B′<br />

2 22 = 22λc2 <br />

3 −<br />

e 22 λt<br />

5λ<br />

√<br />

λ2 + 24c2 22 <br />

t −<br />

e 22 √<br />

λ2 +24c2 5λ + 3 √ λ2 + 24 e<br />

−<br />

c2 t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ2 + 24c2 <br />

√<br />

λ2 + 24c2 + 3<br />

22 <br />

t −<br />

e 22 √<br />

λ2 +24c2 5λ + 3 √ λ2 + 24 e<br />

+<br />

c2 t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ2 + 24c2 <br />

= λc2 <br />

3<br />

√<br />

−<br />

e 22 λt<br />

t − e 4 λ2 +24c2 √<br />

λ2 + 24c2 5λ + 3 √ λ2 + 24c2 (5λ + 3λ2 + 24c2 ) +<br />

3 −<br />

22 λt<br />

= λc2e √<br />

λ2 + 24c2 <br />

e<br />

√<br />

t − 4 λ2 +24c2 − e t<br />

√<br />

4 λ2 +24c2 <br />

.<br />

e t<br />

√<br />

4 λ2 +24c2 5λ − 3 √ λ2 + 24c2 (−5λ + 3λ2 + 24c2 )<br />

We can easily check that E{cosh ηcm(0)} = 1 and we have that d<br />

dt E{cosh ηcm(t)} t=0 = 0 s<strong>in</strong>ce<br />

d<br />

dt E{cosh ηcm(t)}<br />

<br />

= KA − 3λ<br />

<br />

λ + 2c<br />

B + B′ + c<br />

22 2(λ + 3c) ect λ − 2c<br />

− c<br />

2(λ − 3c) e−ct<br />

= 2c2 <br />

3 −<br />

e 22 λt<br />

e<br />

√<br />

λ2 + 24c2 t<br />

22 √<br />

λ2 +24c2 5λ − 3 √ λ2 + 24c2 (3λ − λ2 + 24c2 t<br />

e− 2<br />

) − 2<br />

√<br />

λ2 +24c2 5λ + 3 √ λ2 + 24c2 (3λ + λ2 + 24c2 <br />

)<br />

<br />

λ + 2c<br />

+c<br />

2(λ + 3c) ect λ − 2c<br />

− c<br />

2(λ − 3c) e−ct .<br />

Remark 5.3.1. From Theorem 5.3.3 an <strong>in</strong>terest<strong>in</strong>g relationship between the hyperbolic mean<br />

distance of the center of mass and its acceleration can be extracted<br />

3 −<br />

22 λt<br />

c 2 E{cosh ηcm(t)} − d2<br />

dt2 E{cosh ηcm(t)} = 2λc2e √<br />

λ2 + 24c2 t<br />

s<strong>in</strong>h<br />

22 <br />

λ2 + 24c2 . (5.29)<br />

For large values of t, if √ λ 2 + 2 4 c 2 − 3λ > 0, the mean distance and its second derivative tend to<br />

co<strong>in</strong>cide while <strong>in</strong> the opposite case they tend to diverge. In view of Remark 5.2.2 this means that<br />

there is a different rate of growth for the mean distance and its acceleration only if √ λ 2 + 2 4 c 2 −<br />

3λ < 0.<br />

5.4 Mean hyperbolic distance of the k-th spl<strong>in</strong>ter<br />

If N(t) = n ≥ k, the k-th spl<strong>in</strong>ter, at time t, is located at the hyperbolic distance ηk(t) from O<br />

equal to<br />

cosh ηk(t) =<br />

k+1 <br />

j=1<br />

cosh c(Sj − Sj−1), 0 ≤ k ≤ n, (5.30)<br />

The expression (5.30) refers to sample paths of particles which change direction until the k-th<br />

Poisson event and then rema<strong>in</strong> on the same geodesic l<strong>in</strong>e until time t. Clearly η0(t) represents the<br />

distance of the particle which never changed direction and ηn(t) is the distance of the particle that<br />

changed direction at all Poisson events.


5.4 Mean hyperbolic distance of the k-th spl<strong>in</strong>ter 98<br />

1<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

0<br />

●<br />

c=0.05, t=50, N(t)=300<br />

(a)<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●●●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●●●●●●<br />

●●●●●●● ● ●● ● ●● ●●●● ●● ●<br />

● ● ● ● ●● ●<br />

●●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●●● ●●●●<br />

●<br />

● ●●●●●<br />

● ● ●● ●●●●● ●●<br />

● ● ●●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● O<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

1<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●● ●<br />

●●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

O<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●●<br />

●●<br />

●●<br />

●<br />

●●<br />

●<br />

●●<br />

●●<br />

● ●●● ●●<br />

●<br />

●<br />

●<br />

●<br />

●●<br />

● ●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

● ●<br />

●●<br />

● ●<br />

● ●●<br />

●●<br />

●<br />

●●<br />

● ●<br />

● ●●● ●<br />

●<br />

● ●<br />

●<br />

●● ●<br />

●●<br />

●●●<br />

●● ●●<br />

● ●●●●<br />

0<br />

c=0.05, t=50, N(t)=300<br />

(b)<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

1<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ● ● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●●<br />

●<br />

●●<br />

●<br />

●<br />

●●<br />

●<br />

● ●●<br />

● ●●<br />

●<br />

● ● ●<br />

●<br />

●●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● O ● ●<br />

●<br />

●<br />

●<br />

● ●<br />

● ●<br />

●<br />

●<br />

●<br />

●●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ●<br />

●<br />

●<br />

● ●●<br />

● ●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

● ● ●<br />

0<br />

c=0.05, t=50, N(t)=300<br />

(c)<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

●<br />

Figure 5.5: The position of the spl<strong>in</strong>ters <strong>in</strong> the hyperbolic half-plane H 2 at time t = 50, when<br />

c = 0.05 and N(t) = 300, are drawn. In (a) each spl<strong>in</strong>ter chooses the clockwise direction, <strong>in</strong> (b)<br />

each spl<strong>in</strong>ter chooses the counterclockwise direction and <strong>in</strong> (c) the clockwise and counterclockwise<br />

directions are alternatively chosen.<br />

The mean-value of (5.30) becomes<br />

E{cosh ηk(t)I {N(t)≥k}} =<br />

=<br />

∞<br />

E{cosh ηk(t)I {N(t)=n}}<br />

n=k<br />

●<br />

●<br />

●<br />

● ●<br />

∞<br />

E{cosh ηk(t)|N(t) = n}P{N(t) = n}. (5.31)<br />

n=k<br />

We now evaluate the Laplace transform of (5.31). By lett<strong>in</strong>g γ = λ + µ and by consider<strong>in</strong>g (5.4)<br />

we have that<br />

∞<br />

e<br />

0<br />

−µt E{cosh ηk(t)I {N(t)≥k}}dt =<br />

∞<br />

λ n<br />

n=k<br />

∞<br />

0<br />

e −γt Gn,k(t)dt<br />

∞<br />

= λ<br />

n=k<br />

n<br />

<br />

γ<br />

γ2 − c2 k <br />

1 1<br />

1<br />

+<br />

2 (γ − c) n−k+1 (γ + c) n−k+1<br />

<br />

= 1<br />

<br />

γλ<br />

2 γ2 − c2 <br />

k ∞<br />

n−k 1 λ<br />

+<br />

γ − c γ − c<br />

n=k<br />

1<br />

∞<br />

<br />

n−k<br />

λ<br />

γ + c γ + c<br />

n=k<br />

= 1<br />

<br />

γλ<br />

2 γ2 − c2 k <br />

1<br />

γ − c − λ +<br />

<br />

1<br />

γ + c − λ<br />

= 1<br />

<br />

λ(λ + µ)<br />

2 (λ + µ) 2 − c2 k <br />

1 1<br />

+<br />

µ + c µ − c<br />

= 1<br />

∞<br />

e<br />

2<br />

−µt t<br />

dt e −c(t−s) t<br />

fk(s)ds + e c(t−s) <br />

fk(s)ds<br />

=<br />

0<br />

∞<br />

0<br />

e −µt dt<br />

t<br />

0<br />

0<br />

0<br />

cosh c(t − s)fk(s)ds (5.32)<br />

●<br />

●<br />


99 Cascades of Particles Mov<strong>in</strong>g at F<strong>in</strong>ite Velocity <strong>in</strong> H 2<br />

where fk(s) is the <strong>in</strong>verse Laplace transform of the function<br />

<br />

λ(λ + µ)<br />

(λ + µ) 2 − c2 k = λk<br />

2k <br />

1<br />

λ + µ − c +<br />

k 1<br />

=<br />

λ + µ + c<br />

λk<br />

2k <br />

λ(λ+µ)<br />

(λ+µ) 2−c2 k . Let us write<br />

k<br />

r=0<br />

<br />

k<br />

r<br />

1<br />

(λ + µ − c) r<br />

<br />

j<br />

u If k = 0 we have f0(s) = δ(s). When k = 1, 2, . . . , recall<strong>in</strong>g that L j! eαu <br />

1+(u) =<br />

s > α and j = 0, 1, 2 . . . , we obta<strong>in</strong> the <strong>in</strong>verse Laplace transform<br />

fk(s) = λk<br />

2 k<br />

sk−1 (k − 1)! es(c−λ)<br />

k−1 <br />

s<br />

k w<br />

r<br />

r−1 (s − w)k−r−1<br />

ew(c−λ)<br />

(r − 1)! (k − r − 1)! e−(s−w)(λ+c) dw<br />

+ λk<br />

2 k<br />

+ λk<br />

2 k<br />

r=1<br />

0<br />

s k−1<br />

(k − 1)! e−s(c+λ) .<br />

1<br />

.<br />

(λ + µ + c) k−r<br />

1<br />

(s−α) j+1 with<br />

By <strong>in</strong>vert<strong>in</strong>g the Laplace transform (5.32) we obta<strong>in</strong> for k = 0 that E{cosh η0(t)I {N(t)≥0}} = cosh ct,<br />

while for k = 1, 2, . . . we have<br />

E{cosh ηk(t)I {N(t)≥k}}<br />

t<br />

cosh c(t − s)e<br />

0<br />

cs e−λsλksk−1 k−1<br />

= 1<br />

2k (k − 1)! ds<br />

+ 1<br />

2k <br />

t<br />

k<br />

cosh c(t − s)e<br />

r 0<br />

−λs λ k s k−1<br />

+ 1<br />

2k = 1<br />

2 k<br />

1<br />

r=1<br />

t<br />

cosh c(t − s)e<br />

0<br />

−cs e−λsλksk−1 (k − 1)! ds<br />

t<br />

cosh c(t − s)e<br />

0<br />

cs e−λsλksk−1 (k − 1)! ds<br />

k−1 <br />

cosh c(t − s)E{e cs(2Yr,k−1) e<br />

} −λsλksk−1 + 1<br />

2 k<br />

+ 1<br />

2k = 1<br />

2 k<br />

t<br />

k<br />

r 0<br />

r=1<br />

t<br />

cosh c(t − s)e<br />

0<br />

−cs e−λsλksk−1 t<br />

<br />

k−1<br />

0<br />

r=1<br />

0<br />

e cs(2y−1) yr−1 (1 − y) k−r−1<br />

dy ds<br />

(r − 1)!(k − r − 1)!<br />

Γ(k)<br />

(k − 1)! ds<br />

cosh c(t − s) e cs <br />

<br />

k<br />

+ E{e<br />

r<br />

cs(2Yr,k−1)<br />

<br />

−cs<br />

} + e g(s; k, λ) ds (5.33)<br />

where g(s; k, λ) = e−λsλ k s k−1<br />

Γ(k) , Yr,k ∼ Beta(r, k − r) and (2Yr,k − 1) ∈ (−1, 1). If we adopt the<br />

convention that Y0,k = 1 and Yk,k = −1 it is possible to simplify the last expression rewrit<strong>in</strong>g it as<br />

E{cosh ηk(t)I {N(t)≥k}} = 1<br />

2 k<br />

where h(k, c, s) = k<br />

r=0<br />

= 1<br />

2 k<br />

t<br />

0<br />

t<br />

0<br />

cosh c(t − s)<br />

k<br />

r=0<br />

ds<br />

<br />

k<br />

E{e<br />

r<br />

cs(2Yr,k−1)<br />

}g(s; k, λ) ds<br />

cosh c(t − s) h(k, c, s) g(s; k, λ) ds<br />

k cs(2Yr,k−1)<br />

r E{e }. The last result suggests regard<strong>in</strong>g the mean hyperbolic<br />

distance of a particle generated at the k-th Poisson event as the mean hyperbolic distance of a<br />

particle mov<strong>in</strong>g on the first geodesic l<strong>in</strong>e and stopp<strong>in</strong>g at a randomly distributed time; the law of<br />

the stopp<strong>in</strong>g time is a suitable comb<strong>in</strong>ation of a Beta and Gamma distributions.


5.4 Mean hyperbolic distance of the k-th spl<strong>in</strong>ter 100<br />

Remark 5.4.1. In some particular cases it is possible to check easily that formula (5.33) gives<br />

exactly the same expression of the mean hyperbolic distance of the k-th particle as that obta<strong>in</strong>ed<br />

start<strong>in</strong>g from formula (5.31). In particular for k = 1 we have<br />

E{cosh η1(t)I {N(t)≥1}}<br />

= 1<br />

2<br />

t<br />

For k = 2 we obta<strong>in</strong><br />

E{cosh η2(t)I {N(t)≥2}}<br />

= 1<br />

22 = 1<br />

2<br />

t<br />

0<br />

t<br />

0<br />

cosh c(t − s) e cs + e −cs g(s; 1, λ)ds =<br />

cosh c(t − s)<br />

where Y1,2 ∼ Beta(1, 1).<br />

0<br />

t<br />

<br />

e cs + 2E{e cs(2Y1,2−1) } + e −cs<br />

g(s; 2, λ) ds<br />

cosh c(t − s) cosh cs λ2 se −λs<br />

Γ(2)<br />

ds + 1<br />

2<br />

0<br />

cosh c(t − s) cosh cs λe −λs ds.<br />

t<br />

cosh c(t − s) E{e<br />

0<br />

cs(2Y1,2−1) } λ2se−λs Γ(2)<br />

ds


102


Bibliography<br />

[1] Baldi, P., Casadio Tarabusi, E., Figà-Talamanca, A., Yor, M.: Non-symmetric hitt<strong>in</strong>g distributions<br />

on the hyperbolic half-plane and subord<strong>in</strong>ated perpetuities. Rev. Mat. Iberoam., 17,<br />

587–605 (2001)<br />

[2] Beardon, A., F.: The Geometry of Discrete Groups. Spr<strong>in</strong>ger-Verlag, New York (1983)<br />

[3] Borthwick, D.: Spectral Theory of Inf<strong>in</strong>ite-Area <strong>Hyperbolic</strong> Surfaces. Birkhäuser, Boston (2007)<br />

[4] Buser, P.: Geometry and Spectra of Compact Riemannian Surfaces. Birkhäuser, Boston (1992)<br />

[5] Cammarota, V., Ors<strong>in</strong>gher, E.: Travell<strong>in</strong>g randomly on the Po<strong>in</strong>caré half-plane with a<br />

Pythagorean compass. J. Stat. Phys., 130, 455–482 (2008)<br />

[6] Cammarota, V., Ors<strong>in</strong>gher, E.: Cascades of particles mov<strong>in</strong>g at f<strong>in</strong>ite velocity <strong>in</strong> hyperbolic<br />

spaces. J. Stat. Phys., 133, 1137–1159 (2008)<br />

[7] Cammarota, V., Ors<strong>in</strong>gher, E.: Angular processes related to Cauchy random walks. Under<br />

revision Theory of Probability and its Applications.<br />

[8] Chavel, I.: Riemannian Geometry: a Modern Introduction. Cambridge University Press, Cambridge<br />

(1996)<br />

[9] Chavel, I.: Eigenvalues <strong>in</strong> Riemannian Geometry. Academic Press, New York, London (1984)<br />

[10] Davies, E., B.: Heat Kernels and Spectral Theory. Cambridge University Press, Cambridge<br />

(1989)<br />

[11] Davies, E., B., Mandouvalos, N.: Heat kernel bounds on hyperbolic space and Kle<strong>in</strong>ian groups.<br />

Proc. London Math. Soc., 57, 182–208 (1988)<br />

[12] Debiard, A., Gaveau, B., Mazet, E.: Théorèmes de comparison en géométrie rimannienne.<br />

Publ. Kyoto Univ., 12, 391–425 (1976)<br />

[13] Eberle, A.: Absence of spectral gaps on a class of loop spaces. J. Math. Pures Appl., 81,<br />

915–955 (2002)<br />

[14] do Carmo, M., P.: Riemannian Geometry. Birkhäuser, Boston (1992)<br />

[15] Gertsenshte<strong>in</strong>, M., E., Vasiliev, V., B.: Limit theorems for the compositions of distributions<br />

<strong>in</strong> the Lobachevsky plane and space. Theory Probab. Appl., 3, 391–398 (1959).<br />

[16] Gradshteyn, I., S., Ryzhik, I., M.: Table of Integrals, Series, and Products. Alan Jeffrey and<br />

Daniel Zwill<strong>in</strong>ger eds., Seventh edition (2007)<br />

103


BIBLIOGRAPHY 104<br />

[17] Grigor’yan, A., Noguchi, M.: The heat kernel on hyperbolic space. Bull. London Math. Soc.,<br />

30, no. 6, 643–650 (1998)<br />

[18] Gruet, J.-C.: Semi-groupe du mouvement brownien hyperbolique. Stochastics Stochastics<br />

Rep., 56, 53–61 (1996)<br />

[19] Helgason, S.: Groups and Geometric Analysis. Academic Press, New York, London (1984)<br />

[20] Johnson, N., L., Kotz, S., Balakrishnan, N.: Cont<strong>in</strong>uous Univariate Distributions. John Wiley<br />

& Sons, New York (1995)<br />

[21] Karpelevich, F., I., Pechersky, E., A., Suhov, Yu., M.: A phase transition for hyperbolic<br />

branch<strong>in</strong>g processes. Commun. Math. Phys., 195, 627–642 (1998)<br />

[22] Kelbert, M., Suhov, Yu., M.: Branch<strong>in</strong>g diffusions on H d with variable fission: the Hausdorff<br />

dimension of the limit<strong>in</strong>g set. Theory Probab. Appl., 51, no 1, 155–167 (2007)<br />

[23] Kulczychi, S.: Non-Euclidean Geometry. Pergamon Press, Oxford (1961)<br />

[24] Kühnel, W.: Differential Geometry. American Mathematical Society (2002)<br />

[25] Lalley, S., P., Sellke, T.: <strong>Hyperbolic</strong> branch<strong>in</strong>g <strong>Brownian</strong> motion. Probab. Theory Relat. Fields,<br />

108, 171–192 (1997)<br />

[26] Lao, L., Ors<strong>in</strong>gher, E.: <strong>Hyperbolic</strong> and fractional hyperbolic <strong>Brownian</strong> motion. Stochastics,<br />

79, no. 6, 505–522 (2007)<br />

[27] Lebedev, N., N.: Special Functions and Their Applications. Dover Publications, New York<br />

(1972)<br />

[28] Matsumoto, H.: Closed form formulae for the heat kernels and the green functions for the<br />

Laplacians on the symmetric spaces of rank one. Bull. Sci. Math., 125, no 6-7, 553–581 (2001)<br />

[29] Matsumoto, H.: Limit<strong>in</strong>g behaviors of the <strong>Brownian</strong> motions on hyperbolic spaces.<br />

arXiv:0901.4028v2 [math.PR]<br />

[30] Matsumoto, H., Yor, M.: Exponential functionals of <strong>Brownian</strong> motion, I: Probability laws at<br />

fixed time. Probab. Surveys, 2, 312–347 (2005)<br />

[31] Matsumoto, H., Yor, M.: Exponential functionals of <strong>Brownian</strong> motion, II: Some related diffusion<br />

processes. Probab. Surveys, 2, 348–384 (2005)<br />

[32] Ors<strong>in</strong>gher, E., De Gregorio, A.: <strong>Random</strong> Motions at f<strong>in</strong>ite velocity <strong>in</strong> a non-euclidean space.<br />

Adv. Appl. Prob., 39, 588–611 (2007)<br />

[33] Ors<strong>in</strong>gher, E.: Exact jo<strong>in</strong>t distribution <strong>in</strong> a model of planar random motion with constant<br />

velocity. Stoch. Stoch. Rep., 69, 1–10 (2000)<br />

[34] Ors<strong>in</strong>gher, E., De Gregorio, A.: <strong>Random</strong> motions at f<strong>in</strong>ite velocity <strong>in</strong> a non-Euclidean space.<br />

Adv. Appl. Prob., 32, no 2, 588–611 (2007).<br />

[35] Ratcliffe, J., G.: Foundations of <strong>Hyperbolic</strong> Manifolds. Spr<strong>in</strong>ger (2006)<br />

[36] Revuz, D., Yor, M.: Cont<strong>in</strong>uous Mart<strong>in</strong>gales and <strong>Brownian</strong> Motion. Spr<strong>in</strong>ger-Verlag, Berl<strong>in</strong><br />

(1999)


105 BIBLIOGRAPHY<br />

[37] Rogers, L., C., G., Williams, D.: Diffusions, Markov <strong>Processes</strong>, and Mart<strong>in</strong>gales. Wiley,<br />

Chirchester (1987)<br />

[38] Rosenfeld, B., A.: A History of Non-Euclidean Geometry. Spr<strong>in</strong>ger-Velgar, New York (1988)<br />

[39] Simon, T.: Concentration of the <strong>Brownian</strong> bridge on the hyperbolic plane. Ann. Probab., 30,<br />

4, 1977-1989 (2002)<br />

[40] Spivak, M.: A Comprehensive Introduction to Differential Geometry. Publish or Perish, Houston<br />

(1979)<br />

[41] Terras, A.: Harmonic Analysis on Symmetric <strong>Spaces</strong> and Applications. I. Spr<strong>in</strong>ger-Verlag,<br />

New York (1985)<br />

[42] Watanabe, S.: Analysis of Wiener functionals (Malliav<strong>in</strong> calculus) and its applications to heat<br />

kernels. Ann. Probab., 16, 1–39 (1987)<br />

[43] Yor, M.: On some exponential functionals of <strong>Brownian</strong> motion. Adv. <strong>in</strong> Appl. Probab., 24,<br />

509–531 (1992)<br />

[44] Yor, M.: Loi de l’<strong>in</strong>dice du lacet brownien, et distribution de Hartman-Watson. Z. Wahrsch.<br />

Verw. Gebiete., 53, 71–95 (1980)

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!