• 検索結果がありません。

From Sine kernel to Poisson statistics

N/A
N/A
Protected

Academic year: 2022

シェア "From Sine kernel to Poisson statistics"

Copied!
26
0
0

読み込み中.... (全文を見る)

全文

(1)

El e c t ro nic J

o f

Pr

ob a bi l i t y

Electron. J. Probab.19(2014), no. 114, 1–25.

ISSN:1083-6489 DOI:10.1214/EJP.v19-3742

From Sine kernel to Poisson statistics

Romain Allez

*

Laure Dumaz

Abstract

We study the Sineβprocess introduced in [B. Valkó and B. Virág. Invent. math.177 463-508 (2009)] when the inverse temperatureβtends to0. This point process has been shown to be the scaling limit of the eigenvalues point process in the bulk of β-ensembles and its law is characterized in terms of the winding numbers of the Brownian carrousel at different angular speeds. After a careful analysis of this family of coupled diffusion processes, we prove that the Sineβ point process converges weakly to a Poisson point process onR. Thus, the Sineβ point processes establish a smooth crossover between the rigid clock (or picket fence) process (corresponding to β=∞) and the Poisson process.

Keywords:Random matrices ; Diffusions ; Poisson point process ; Exit time problem.

AMS MSC 2010:60G55 ; 60G17 ; 60B20 ; 60J60.

Submitted to EJP on August 14, 2014, final version accepted on December 11, 2014.

1 Introduction and main result

Although random matrices were originally introduced by John Wishart [24] in 1928 as a tool to study population dynamics in biology through principal component analysis, they became very popular much later in 1951 when Wigner [23] postulated that the fluctuations in positions of the energy levels of heavy nuclei are well described (in terms of statistical properties) by the eigenvalues of a very large Hermitian random matrix. Random matrix theory (RMT) is now an active research area in mathematics and theoretical physics with applications in statistics, biology, financial mathematics, engineering and telecommunications, number theory etc. (see [6, 7, 18, 13, 1] for a state of the art).

The classical models of Hermitian random matrices are the Gaussian orthogonal, unitary and symplectic ensembles. It is well known that the joint law of the eigenvalues of the matrices in those Gaussian ensembles is the Boltzmann-Gibbs equilibrium measure of a one-dimensional repulsive Coulomb gas confined in a harmonic well. More precisely, this joint law has a probability densityPβ onRN (N is the dimension of the square matrices) given by

Pβ1,· · · , λN) = 1 ZN

Y

i<j

i−λj|βexp(−N β 4

N

X

i=1

λ2i) (1.1)

*Weierstrass Institute, Berlin, Germany. E-mail:romain.allez@gmail.com

Statistical Laboratory, Centre for Math. Sciences, Cambridge, UK. E-mail:L.dumaz@statslab.cam.ac.uk

(2)

where the inverse temperatureβ = 1for the Gaussian orthogonal ensemble, respectively β= 2,4for the unitary and symplectic ensembles. The linear statistics of the point pro- cesses with joint probability density functions (jpdf)Pβ, β= 1,2,4have been extensively studied in the literature with different methods [6, 18].

In 2002, Dumitriu and Edelman [11] came up with a new explicit ensemble of random tri-diagonal matrices whose eigenvalues are distributed according to the jpdfPβ for anyβ > 0 (see also [5, 2] where invariant ensembles associated to general β were constructed).

Those tri-diagonal matrices have been very useful in the last decade, leading to important progress on the understanding of the local eigenvalues statistics in the limit of large dimensionN for generalβ >0. At the edge of the spectrum, it was first proved [20] that the largest eigenvalues converge jointly (when zooming in the edge-scaling region of widthN−2/3around2) to the low lying eigenvalues of a random Schrödinger operator called the stochastic Airy operator (see also [12]). Similar results were proved for the bulk in [22] by Valkó and Virág. Forλbelonging to the Wigner sea(−2,2), the authors of [22] consider the point process

ΛN := (2πN ρ(λ) (λi−λ))i=1,···,N (1.2) where(λ1,· · · , λN)is distributed according toPβandρ(λ) = 1

4−λ2 is the Wigner semi-circle density. Indeed, the mean level spacing around levelλfor the points with lawPβis approximately1/(N ρ(λ))whenN1. The mean point spacing ofΛN defined in (1.2) is therefore of order2πand in this scaling, one can now investigate the limiting statistics of this point process whenN → ∞. The authors of [22] precisely answer this question proving that the point processΛN converges in law1to a point process Sineβ

onRfirst introduced in [22] and characterized in terms of a family(αλ)λ∈Rof coupled one-dimensional diffusion processes, thestochastic sine equations. As expected for the eigenvalues statistics in the bulk, the point process Sineβ is translation-invariant in law. The family of diffusions(αλ)λ∈Rcan be interpreted as the hyperbolic angle of the Brownian carousel with parameterλand its law is characterized as follows: Given a (driving) complex Brownian motion(Zt)t≥0, the diffusionsαλ, λ∈Rsatisfy

λ=λβ

4eβ4tdt+ Re((e−iαλ−1)dZt), αλ(0) = 0. (1.3) Note that all the diffusionsαλ, λ∈Rare adapted to the filtration of the Brownian motion (Zt). This coupling induces a strong interaction between the diffusions which makes the joint law difficult to analyse, as we shall see. A key feature shared by the processes αλ, λ∈Ris that they all converge almost surely ast→ ∞to a limitαλ(∞)which is an integer multiple of2π. The characterization of the law of the Sineβpoint process2can now be enunciated as follows

(Sineβ([λ, λ0]))λ<λ0

(d)=

αλ0(∞)−αλ(∞) 2π

λ<λ0

. (1.4)

In this paper, we are interested in the limiting law of the Sineβ process when the inverse temperatureβgoes to0. Theorem 1.1 is the main result of this paper and gives the convergence asβ →0of the Sineβprocess towards a Poisson point process onR. This convergence at the continuous (N=∞) level seems rather natural since takingβ→0 amounts to decreasing the electrostatic repulsion (and hence the correlation) at the

1The convergence is with respect to vague topology for the counting measure of the point process.

2IfAis a Borel set ofR, Sineβ(A)insideAdenotes the number of points insideA. In other words, Sineβis the counting measure of the point process.

(3)

0 20 40 60 80 100

01234

Figure 1: (Color online). Sample paths of two diffusionsαλ/(2π)(blue curve) andαλ0/(2π) (red curve) forλ < λ0 and for a small value ofβ. We see that wheneverbαλ/(2π)cjumps,λ0/(2π)cjumps at the same time in agreement with Lemma 3.4.

discrete level, i.e. between theN points distributed according toPβ. If one takesβ= 0 abruptly for fixedN, the probability densityP0corresponds (up to a rescaling depending onβ) to the joint law of independent Gaussian variables and it is then straightforward to check the convergence (asN → ∞) of the point process with lawP0towards a Poisson process. Theorem 1.1 exchanges the order of the limitsβ→0andN → ∞, describing the statistics whenN → ∞first and thenβ →0. It also gives the precise rate of the convergence.

Theorem 1.1.As β → 0, the Sineβ point process converges weakly in the space of Radon measure (equipped with the topology of vague convergence [15]) to a Poisson point process onRwith intensity . In particular, we have, for anyk∈Nandλ < λ0,

P[Sineβ[λ, λ0] =k]→β→0exp

λ−λ0

0−λ)k (2π)kk! ,

and the numbers of points of Sineβ inside two disjoint intervals are asymptotically independent.

Let us briefly discuss some implications of Theorem 1.1 and mention a few related questions on the spectral statistics of random matrices and random Schrödinger opera- tors.

In [16], the authors have shown that the circularβ-ensemble, which was later shown to be Sineβin [19], interpolates between Poisson and clock distributions on the circle (point process with rigid spacings like the numerals on a clock) by considering random CMV matrices. Theorem 1.1 provides a more precise description of this interpolating process on the Poisson process side.

In our study, we are led to examine a time homogeneous family of diffusions(θλ)λ∈R

defined as

λ=λβ

4dt+ Re((e−iθλ−1)dZt), θλ(0) = 0. (1.5) This family of coupled diffusions also appears in [17] to describe the law of the limiting point process of a certain critical random discrete Schrödinger operator. Our result can

(4)

be extended in this context to prove that this critical Schrödinger operator continuously interpolates between the extended (clock/picket fence) and localized (Poisson) regimes.

More precisely, one could prove using our ideas that the random spectrum has Poissonian statistics in the limit of large temperature.

Let us also compare the results stated in Theorem 1.1 with those of a previous work [4] where we consider the stochastic Airy ensemble, Airyβ, obtained in the scaling limit ofβ-ensembles at the edge of the spectrum. In this context, we proved that the number of points Airyβ(]− ∞, λ])inside the interval]− ∞, λ]displays Poisson statistics in the smallβlimit. This permitted us to obtain the limiting distributions asβ→0of each of the lowest eigenvalues (individually) of the Airyβ ensemble. In particular, we obtained the weak convergence of theT W(β)distribution towards the Gumbel distribution. Although the Sineβ and Airyβ characterizations in law (in terms of a family of coupled diffusions) look very similar, the analysis of the limiting marginal statistics of the number of points inside a finite closed interval Airyβ[λ, λ0]forλ < λ0 and the asymptotic independence whenβ → 0 of the respective numbers of points of the Airyβ point process into two disjoint intervals remain open even after our study [4]. In this aspect, Theorem 1.1 gives a much more powerful and complete description of the Sineβ process in the smallβlimit.

In this case, we are able to prove the asymptotic independence between the number of points of the Sineβprocess in two disjoint intervals. This part of the proof requires new ideas in order to obtain estimates on the relative positions between two coupled diffusionsαλandαλ0. The nice feature of the Sineβprocess is its translation invariance in law. This property makes the analysis of Sineβeasier than the one of the Airyβprocess.

The non-homogeneous intensity of the Airyβ process is governed by the edge-scaling crossover spectral density ofβ-ensembles computed explicitly in [8, 14] forβ= 2(see also [4]).

Other spectral statistics of random matrices at high temperature, i.e. whenβ→0, have been investigated in [2, 3]. In [2], the authors study the empirical eigenvalue density in the limit of large dimensionN forβ-ensembles whenβtends to0withN as β= 2c/N wherec >0is a constant. The authors compute the limiting spectral density ρc(λ)explicitly in terms of parabolic cylinder function and establish a Gauss-Wigner crossover, in the sense that the familyρc interpolate between the Gaussian probability distribution (c= 0) and the Wigner semi-circle (c→+∞). The case of Gaussian Wishart matrices has also been studied in [3].

It would be interesting to have a description of the crossover statistics of the β- ensembles obtained in the double scaling limits whenβ tends to0withN → ∞(this question is briefly discussed in [4] for the statistics at the edge of the spectrum).

Organization of the paper. We start in section 2 by looking at the limiting marginal distributions of the random variables Sineβ[λ, λ0]forλ < λ0. We first study a classical problem on the exit time of a diffusion trapped in the well of a stationary potential (obtained by neglecting the slow evolution with time). Then, we prove that the jump process ofbαλ/(2π)cconverges weakly to an (inhomogeneous) Poisson point process by first approximatingαλwith diffusions processes with piecewise constant drifts on a subdivision of small intervals and then by using the convergence of the exit time of the stationary well established previously. This requires estimates on the sample paths of a single diffusion, in a spirit similar to [4, 10]. In Section 3, we investigate the asymptotic independence of the numbers of points of Sineβ in two disjoint intervals. We prove a crucial estimate regarding the typical relative positions of two diffusionsαλ andα0λfor λ < λ0. Loosely speaking, the main point is to use this estimate to prove that, in the limitβ→0, the jumps of the processbαλ/(2π)cimmediately follow those ofbαλ0/(2π)c (see Fig. 1) while the processesbαλ/(2π)candb(αλ0−αλ)/(2π)cnever jump at the same time (see Fig. 4). The asymptotic independence follows essentially from the fact that

(5)

two Poisson point processes adapted to the same filtration are independent if and only if they never jump simultaneously.

We gather in the next paragraph important properties of the family (αλ) already established in [Section 2.2, [22]] that we will use throughout the paper.

First properties of the coupling of the diffusionsαλ:

(i) For allλ < λ0λ0−αλhas the same distribution asαλ0−λ; (ii) “Increasing property”:αλ(t)is increasing inλ;

(iii) bαλ(t)/(2π)cis non-decreasing int; (iv) E[αλ(t)] =λβ4Rt

0e−βs/4ds;

(v) limt→∞αλ(t)/(2π)exists and is an integer a.s.

We will also use the following notation:

{x}=x−2πj x 2π

k .

Acknowledgments. Special thanks are addressed to Chris Janjigian and Benedek Valkó. We have benefited from insightful and precise comments from them. Their detailed feedback has helped us to improve the second version of this manuscript, especially the proofs of Lemmas 3.2 and 3.4. We are also grateful to them for pointing out references [16, 19, 17] and the connections with our work.

We thank Stéphane Benoist and Antoine Dahlqvist for useful comments and discus- sions.

R. A. received funding from the European Research Council under the European Union’s Seventh Framework Programme (FP7/2007-2013) / ERC grant agreement nr.

258237 and thanks the Statslab in DPMMS, Cambridge for its hospitality. The work of L. D. was supported by the Engineering and Physical Sciences Research Council under grant EP/103372X/1 and L.D. thanks the hospitality of the maths department of TU and the Weierstrass institute in Berlin.

2 Limiting marginal distributions

We are first interested in the limiting law of the random variable Sineβ[0, λ]when β→0for a single fixedλ. In this case, we re-write the diffusionαλin a more convenient way:

λ=λβ

4eβ4tdt+ 2 sin(αλ

2 )dBt, (2.1)

whereB is a real standard Brownian motion (which depends onλ). Let us introduce the change of variableRλ:= log(tan(αλ/4)). A straightforward computation (see [22]) shows that:

dRλ= 1 2

λβ

4eβ4tcosh(Rλ) + tanh(Rλ)

dt+dBt, Rλ(0) =−∞. (2.2) 2.1 Trapping in the stationary potential

In this subsection, we study an exit time problem for a Langevin diffusionSλevolving in a stationary potentialVβ defined forr∈Ras

Vβ(r) :=−1 2

λβ

4 sinh(r) + log cosh(r)

.

(6)

This problem is relevant to our study thanks to the slow variation with time of the non-stationary potential asβ →0in which the diffusionRλevolves. The diffusionSλ satisfies the following stochastic differential equation

dSλ=−Vβ0(Sλ)dt+dBt, Sλ(0) =−∞. (2.3) In the whole paper, the diffusionsRλ andSλdefined respectively in (2.2) and (2.3) are coupled, driven by the same Brownian motionB. Under this coupling, we have almost surely

Rλ(t)≤Sλ(t) for allt≤ζwhereζis the first explosion (stopping) time

ζ:= inf{t>0 :Sλ(t) = +∞}.

' log(1/β)/2 ' log(1/β )

(0, 0)

aβ

bβ

Figure 2: The potentialVβ(r)as a function ofr.

In this paragraph we investigate the limiting law of the stopping timeζthrough its Laplace transform defined forξ >0as

gξ(r) :=Er[e−ξζ], whereris the initial position of the diffusionSλ.

We know from classical diffusion theory (see also [4]) thatgξ satisfies 1

2gξ00(r)−Vβ0(r)gξ0(r) =ξgξ(r), withgξ(r)→1asr→+∞. (2.4)

(7)

Let us first examine the expectation of the first explosion time of Sλ. Due to the strong barrier separating the local minimum and the local maximum (see Fig. 2), it is natural to expect the asymptotic of this mean exit time not to depend on the starting point of the diffusion, whenβ → 0, as long as it is located in the well (“memory-loss property”). This is the purpose of Proposition 2.1. We then show that this first exit time properly rescaled by its mean value converges in law to an exponential distribution (Proposition 2.2) when the starting point is in the well.

Proposition 2.1.Suppose that the diffusionSλstarts fromr:=rβ such thatrβ→ −∞. Then its expected exit time denotedtβ(r) :=Er[ζ]has the following equivalent when β→0:

tβ(rβ)∼ 8π βλ.

Proof of Proposition 2.1. From (2.4), it is easy to see that the expected exit timetβ(r) satisfies the boundary value problem

1

2t00β(r)−Vβ0(r)t0β(r) =−1, withtβ(r)→0asr→+∞. (2.5) Solving (2.5) explicitly we obtain the integral form

tβ(r) = 2 Z +∞

r

dx Z x

−∞

exp (2 [Vβ(x)−Vβ(y)])dy . (2.6) By extracting carefully the asymptotic behavior of this integral in the limitβ →0(see Appendix A.1), we obtained the desired result.

The Laplace transform of the exit time satisfies the fixed point equation gξ(r) = 1−2ξ

Z +∞

r

dx Z x

−∞

exp(2[Vβ(x)−Vβ(y)])gξ(y)dy.

With classical arguments similar to those of [4], we prove the following proposition in Appendix A.1.

Proposition 2.2.Conditionally onSλ(0) =r:=rβsuch thatrβ → −∞whenβ→0, the first explosion timeζof the diffusion(Sλ(t))t>0converges weakly when rescaled by the expected exit time8π/(βλ)towards an exponential distribution with mean1.

The processθλsuch thatSλ:= log tan(θλ/4)satisfies dθλ=λβ

4dt+ 2 sin(θλ 2 )dBt.

Proposition (2.2) translates into a convergence of the stopping timeζ := inf{t >0 : θλ(t)>2π}.

Proposition 2.3.Conditionally onθλ(0) =θ0:=θβ0 such thatθ0β→0+ whenβ →0, the stopping time βλζ converges weakly whenβ →0towards an exponential distribution with mean1.

2.2 Convergence of the jump process

We consider the diffusionαλdefined in (1.5) or equivalently for a singleλ, in (2.1).

Fork∈N, let

ζk:= inf{t>0 :αλ(t)>2kπ}. (2.7) Note that those stopping times correspond to the jumps of the processbαλ/(2π)c. We denote byFtthe filtration associated to the diffusion processαλ. In this paragraph, we

(8)

will sometimes omit the subscript λinαλ to simplify the notations. We consider the rescaled empirical measureµβλ of the β4ζkdefined onR+

µβλ[0, t] =

+∞

X

k=1

δζk

0,8πt β

. (2.8)

We divide the time intervalR+into random small intervalsIk := [Snkβ,Sk+1n β], k∈N, independent of the diffusionαwhereS0= 0and:

Sk :=

k

X

i=1

τi,

where theτiare i.i.d. random variables with mean1uniformly distributed on[1/2,3/2]. For eachk∈N, conditionally onFSk

n

β

, we define the two diffusion processes α+n andαn such that, fort∈Ik,

+n =λβ

4eSkndt+ 2 sin(α+n

2 )dBt, α+n(Sk n

β ) =α(Sk n

8π β ), dαn =λβ

4eSk+1n dt+ 2 sin(αn

2 )dBt, αn(Sk

n 8π

β ) =α(Sk

n 8π

β ). By the increasing property, it follows that for anyk∈Nandt∈Ik, we have

αn(t)≤α(t)≤α+n(t).

Theorem 2.4.Asβ →0, the empirical measureµβλ converges weakly in the space of Radon measures (equipped with the topology of vague convergence [15]) to a Poisson point process onR+ with inhomogeneous intensityλ e−2πtdt. In particular, we have, for anyk∈N,

P[µβλ[0, t] =k]→β→0exp(−λ

2π(1−e−2πt))(λ)k(1−e−2πt)k

k! . (2.9)

The proof of Theorem 2.4 is done in the next paragraph 2.3. It uses a careful analysis of the behaviour of a single diffusionαλin the smallβ limit.

Remark 2.5.Theorem 2.4 establishes the convergence of the rescaled empirical mea- sure of the jumps of the processbαλ/2πcunder the initial conditionαλ(0) = 0. It is easy to generalize this result to other starting timess0and positionsαλ(s0)6= 0such thats0

scales with1/βandαλ(s0)is close enough to0modulo2π.

More precisely, fix a time s > 0 and denote by µβλ,s,x the rescaled counting mea- sure of the jumps ofbαλ/2πcon the time interval[8πs/β,+∞)with the initial condition αλ(8πs/β) =x. Then, for any family of starting points(xβ)such that{xβ}64 arctan(β1/4), the measureµβλ,s,x

β converges weakly in the space of Radon measures towards a Poisson point process on[s,+∞)with inhomogeneous intensityλ e−2πtdt.

Thanks to the equality in law

µβλ(R+)(d)= αλ(∞) 2π

(d)= Sineβ[0, λ],

we easily deduce from Theorem 2.4 the convergence of the marginals of the Sineβpoint process.

Corollary 2.6.Letλ < λ0. The random variableSineβ[λ, λ0]converges weakly asβ→0 to a Poisson law with parameter λ0−λ.

(9)

2.3 Estimates for a single diffusion and proof of Theorem 2.4

We analyse in this paragraph the diffusion αλ and the jumps ofbαλ/(2π)c in the limitβ→0. The results we obtain are derived using the diffusionRλ:= log(tan(αλ/4)) satisfying the SDE (2.2). We defer the proof of those technical lemmas in Appendix A.2.

In the following, for allt >0andx∈R+, the law of the diffusion processαλstarting fromxat timetis denotedPx,t. Whent= 0, we use the short cutsPxforPx,0and and PforP0,0.

In the following Lemma, we first prove that if the diffusionαλstarts just below2π modulo2π, thenbαλ/(2π)cwill jump in a time of order logβ1 much shorter than the typical time between two jumps (of order1/β) with probability going to1.

Lemma 2.7.Let0< ε <1ands >0. Denotesˆ:= 8πs/β. We define the first reaching time

ζλ := inf{t≥sˆ:αλ(t) = 2π+ 2πbαλ(ˆs)/(2π)c}. (2.10) Then, there exists a constantc >0(depending only ons,εandλ) such that for allβ >0 small enough,

P

ζλ <9 log 1 β

λ(ˆs)}= 2π−4 arctan(βε)

>1−βc.

Lemma 2.7 permits us to show that the time spent byαλaway from0modulo2πis negligible compared to the time scale1/β. This is the content of Lemma 2.8.

Lemma 2.8.Lett > s≥0,x∈R+. Let us define Ξβ(s, t, x) :=Ex,8πs/β

"

Z βt

βs

1{{αλ(u)}≥4 arctan(β1/4)}du

#

. (2.11)

Then, for allt≥s >0, there existsC >0(depending only ontandλ) such that for all β >0andx∈R+,

Ξβ(s, t, x)6 C

√β log1 β . We can now prove Theorem 2.4 using the previous estimates:

Proof of Theorem 2.4.

The proof follows the same lines as the proof of [Theorem 4.1 in [4]]. The idea is to approximate the number of jumps of the diffusionαby those of stationary diffusions and to use the increasing property. To this end, we will use the subdivision ofR+introduced above and the diffusionsαn+andαn.

From Kallenberg’s theorem [15], we just need to see that, for any finite unionIof disjoint and bounded intervals, we have whenβ→0,

E[µβλ(I)]−→λ Z

I

e−2πtdt , (2.12)

P[µβλ(I) = 0]−→exp(−λ Z

I

e−2πtdt). (2.13)

Denote by[t1;t2] the right most interval ofI, byJ the union of disjoint and bounded intervals such thatI=J∪[t1;t2]and byt0the supremum ofJ.

Let us use the random subdivision introduced above. It is crucial to control the position of the diffusion at the starting point of all the sub-intervals. As a consequence of

(10)

Lemma 2.8, with large probability, the diffusionαis close to0modulo2πin the beginning of each of the sub-intervals overlapping[0, t2]3.

More precisely, denote byCk :={{α((8π/β)Sk/n)} 64 arctan(β1/4)}and consider:

C:=

b2nt2c+1

\

k=0

Ck.

Every sub-interval intersecting [0, t2] is taken care of in this event as by definition Sk∈[k/2,3k/2]. Its probability is bounded from above by

P[Cc]6

b2nt2c+1

X

k=0

P[{α((8π/β)Sk/n)}>4 arctan(β1/4)]

62E[

Z 3nt2+3 0

1{{α((8π/β)u/n)}>4 arctan(β1/4)}du]

6 n

4πβΞβ(0,3t2+ 1,0)→β→00

where we used the notation and result of Lemma 2.8 for the last line.

We now turn to the proof of 2.12. Note that thanks to the linearity of the expectation, we simply need to prove (2.12) for intervalsIof the formI= [0, t2]. The upper bound simply follows from the SDE form:

E[α(t)] =λβ 4

Z t 0

e−βs/4ds . We immediately derive:

E[µβ[0, t2]] =E[bα((8π/β)t2)/(2π)c]6λ Z t2

0

e−2πsds .

For the lower bound, denote byNk+,NkandNkthe number of jumps ofα+nn and αin the sub-interval[(8π/β)Sk/n,(8π/β)Sk+1/n].

E[µβ[0, t2]]>

b2nt2c+1

X

k=0

E[Nk1{(Sk/n)<t2}|Ck]−

b2nt2c+1

X

k=0

E[Nk1{(Sk/n)<t2}|Ck]P[Ckc]. The second sum of the RHS can be simply bounded from above by:

E[number of jumps ofθλin[0,3t2(8π)/β]]× sup

k∈{0,···,b2nt2c+1}

P[Ckc] P[Ck].

The first expectation is bounded from above by a number independent of β and the second term (bounded byP[Cc]/(1−P[Cc])) tends to0asβ →0. Moreover, thanks to Proposition 2.3, we have

E Nk+

Ck

β→0E[λ(τk+1/n) exp(−2πSk/n)].

Using the convergence of the Riemann sum whenn → ∞, it gives the desired lower bound.

Let us now examine the convergence 2.13 for a single interval[t1, t2]first. By the Markov property:

P[µβ[t1, t2] = 0]6EhY

k=0

P

Nk+= 0

Ck,(τi)i

1{Sk+1/n>t1, Sk/n6t2}i

+P[Cc].

3For the generalization of Remark 2.5, the initial position moduloat times0belongs to the interval [0,4 arctan(β1/4))so that the eventC0occurs almost surely by definition.

(11)

Using the convergence in each of the sub-intervals (given by Proposition 2.3) P

Nk+= 0 Ck

β→0E[exp(−λ(τk+1/n) exp(−2πSk/n)))], we obtain:

lim sup

β→0 P[µβ[t1, t2] = 0]6E[

Y

k=0

exp(−λ(τk+1/n) exp(−2πSk/n)))1{Sk+1/n>t1, Sk/n6t2}]. Thanks to the convergence of the Riemann sum when n → ∞, we deduce the upper bound. The lower bound can be done using the same techniques as above. To generalize the result to any finite union of intervalsI, we use the simple Markov property which gives

P[µβλ(I) = 0] =E[1β(J)=0}P[µβ([t1, t2]) = 0|α(8πt0/β)]].

To conclude, let us introduce an independent random time U sampled uniformly in [t0, t1]. With probability going to1,α(8πU/β)belongs to an interval of the type[2π,2π+ 4 arctan(β1/4)). We can then iterate the previous arguments to obtain the result.

3 Asymptotic spatial independence

3.1 Ordering of two diffusions

Forλ < λ0, we now control the expected time spent by the process{αλ0}below the process{αλ}whereαλandαλ0 are coupled according to (1.5). The following Lemma is a crucial step towards the proof of the asymptotic independence of the limiting point process on disjoint intervals and will be used for the proof of Lemmas 3.4 and 3.5.

Lemma 3.1.Lett >0, λ < λ0and Θβ(t) :=E

"

Z βt 0

1{{α

λ0(u)}≤{αλ(u)}}du

# .

Then, there exists two constantsc, C >0(depending only ontandλ0) such that for all β >0,

Θβ(t)6 C β1−c. Proof. We set

Eu:={{αλ0(u)}<{αλ(u)}} .

Before evaluating the probability of the eventEu, we need to introduce foru >0the last jump time of the processbαλ0cbefore timeu, i.e.

ζ(u):= sup{ζkλ0 : k∈N, ζkλ0 ≤u}.

The main idea is to prove that any timeusuch thatEu occurs is associated to a jump time ofαλ0 right beforeu.

We setu0:=u−9 logβ1 and bound from above the probability of the eventEuby P[Eu]≤Ph

Eu∩n

ζ(u)∈[u0, u]oi +Ph

Eu∩n

ζ(u)< u0oi

≤Ph Eu∩n

ζ(u)∈[u0, u]oi +P

\

s∈[u0,u]

Es

 (3.1)

(12)

u 2π

0

ζ (u)

{ α λ } 2π

{ α λ

0

} 2π

u

0

Figure 3: Representation of the eventEu∩ {ζ(u)< u0}. where we have noticed the inclusionEu∩{ζ(u)< u0} ⊆T

s∈[u0,u]Esfor the second line (see Fig. 3). Indeed the definition ofζ(u)implies that the processbαλ0chas not jumped during the time interval[ζ(u), u]so that the relative ordering (modulo2π){αλ0(t)}<{αλ(t)} at timet=uhas to be preserved for allt∈[ζ(u), u).

We now tackle the second probability of (3.1) using the fact that αλ is close to0 modulo2πwith large probability:

P

\

s∈[u0,u]

Es

≤P

\

s∈[u0,u]

Es∩n

λ(u0)}≤4 arctan(β1/4)o

+Ph

λ(u0)}≥4 arctan(β1/4)i . We introduce the stopping time

ζ>(u0) := inf{t≥u0:{αλ0(t)}>{αλ(t)}} and notice that, conditionally on the eventAu0 :=Eu0

λ(u0)}≤4 arctan(β1/4) , we have almost surely for allt∈[u0, ζ>(u0)],

λ0(t)−αλ(t)}= 2π−({αλ(t)}− {αλ0(t)}).

We define a diffusion processαeλ0−λand its associated first reaching time of2πsuch that dαeλ0−λ(t) = (λ0−λ)β

4eβ4tdt+ Re((e−iαeλ0 −λ(t)−1)dZt), t≥u0

withαeλ0−λ(u0) = 2π−({αλ(u0)}− {αλ0(u0)}), andζe := inf{t≥u0:αeλ0−λ(t) = 2π}.

Endowed with those definitions, we have the following upper-bound

P

\

s∈[u0,u]

Es∩ Au0

=P[{ζ>(u0)≥u} ∩ Au0]

≤P

≥9 log1 β

≤βc

where we have used the equality in law(αeλ0−λ(t), t≥u0) = (αλ0(t)−αλ(t), t≥u0)for the second line and Lemma 2.7 (as well as the increasing property) to obtain the last

(13)

upper-bound. Gathering the above inequalities, we obtain P[Eu]≤Ph

ζ(u)∈[u0, u]i

c+Ph

λ(u0)}≥4 arctan(β1/4)i .

Now, to conclude, we just have to integrate this latter inequality with respect tou. First notice that we have almost surely

Z βt 9 log(1/β)

1{ζ(u)[u−9 logβ1,u]}du≤9 log1

β µβλ0[0,8π β t].

Integrating the other terms as well with respect tou∈[0,βt]and taking the expectation, we get

Θβ(t)≤9 log1 β

1 +E

µβλ0[0,8π β t]

+ Ξβ(0, t,0)

≤9 log1 β

1 + λ

+ 8π t βc−1+ Ξβ(0, t,0)

whereΞβ(0, t,0)is defined in (2.11). The conclusion now follows from Lemma (2.8).

3.2 Limiting coupled Poisson processes

Lemma 3.1 shall be an important tool to prove the asymptotic independence between αλ(∞)andαλ0(∞)−αλ(∞)forλ < λ0.

Theorem 2.4 gives the weak convergence of the random measuresµβλandµβλ0 in the space of measures onR+ equipped with the topology of vague convergence denoted M+(R+). Due to the equality in law

αλ0−αλ

(d)= αλ0−λ, (3.2)

Theorem 2.4 also implies the weak convergence of the (positive) random measureµβλ0−λ

such that for allt≥0,

µβλ0−λ[0, t] :=

αλ0(t)−αλ(t) 2π

towards a Poisson measurePλ0−λwith intensity(λ0−λ)e−2πtdt.

We now fixλ < λ0< λ00and work with the three diffusionsαλλ0 andαλ00coupled according to (1.5). We are interested in the limiting joint distribution of the triplet of random measures(µβλ, µβλ0, µβλ00−λ0)according to this coupling.

From the above convergences, it is straightforward to deduce the relative-compactness of the family of the triplets of (random) measures

{ µβλ, µβλ0, µβλ00−λ0

, β >0} (3.3)

for the weak topology over(M+(R+))3equipped with the product topology of vague convergence.

Let us take a sequenceβk→0whenk→ ∞such that the processes µβλk, µβλk0, µβλk00−λ0

(3.4)

converge jointly weakly in the product space whenk→ ∞to a triplet (Pλk),Pλ0k),Pλ00k−λ) 0)

(14)

of point measures onR+whose marginal distributions are given respectively by the law of the Poisson measuresPλ andPλ0 andPλ00−λ0 and whose joint law dependsa priori on the chosen sub-sequence(βk). In the following, we study this triplet and we drop the superscript(βk)to ease the notations. We shall in fact see later that all the possible limit points have the same law. Therefore the law of the triplet does not depend on the subsequence(βk)and the weak convergence of (3.3) holds (see Remark 3.6).

In the next Lemma, we regard the point measures Pλ,Pλ0 andPλ00−λ0 as Poisson point processes and prove that they are indeed jointly Poisson processes on a common filtration. This is an important step for our needs.

Lemma 3.2.Letλ < λ0< λ00andF:= (Ft)t≥0be the natural filtration associated to the process(Pλ,Pλ0,Pλ00−λ0)i.e. such that

Ft:=σ(Pλ(s),Pλ0(s),Pλ00−λ0(s),0≤s≤t).

Then, the marginal processesPλ andPλ0 andPλ00−λ0 are(Ft)-Poisson point processes with respective intensitiesλe−2πtdtandλ0e−2πtdtand(λ00−λ0)e−2πtdt.

The main point used to prove this Lemma is that all the diffusions (αλ, λ ≥0 are measurable with respect to the same driving (complex) Brownian motion(Zt)as they are strong solutions of the SDEs (1.5). A proof can be found in Appendix A.3.

Lemma 3.2 can easily be generalized to finitely manyλ’s:

Lemma 3.3.Let us fix an integerk≥3andλ0< λ1<· · ·< λk. For alli, j ∈ {1,· · · , k} such thatj > i, define the point measuresPλi andPλj−λi as above. LetF:= (Ft)t≥0be the filtration:

Ft:=σ(Pλi(s),Pλj−λi(s),0≤s≤t, i, j ∈ {1,· · · , k}such thatj > i).

Then, for all i, j ∈ {1,· · ·, k} with j > i, the marginal processes Pλi and Pλj−λi are (Ft)-Poisson point processes with respective intensitiesλie−2πtdtand(λj−λi)e−2πtdt. To fix notations, we will denote by(ξλi)i∈N resp.(ξλi0)i∈N,(ξiλ0−λ)i∈N the points associ- ated to the Poisson processesPλandPλ0 andPλ0−λsuch that

Pλ[0, t] =X

i

δξλ

i[0, t], Pλ0[0, t] =X

i

δξλ0

i [0, t], Pλ0−λ[0, t] =X

i

δξλ0 −λ i

[0, t].

Forβ >0, we also recall the notations µβλ[0, t] =X

i

δζλ i[0,8πt

β ], µβλ0[0, t] =X

i

δζλ0 i [0,8πt

β ], µβλ0−λ[0, t] =X

i

δζλ0 −λ i

[0,8πt β ]. Lemma 3.4.Letλ < λ0. Then, we have almost surely

Pλ⊆ Pλ0 i.e. for alli∈N, there existsj≥isuch thatξλiλj0.

Proof. We have to prove that for anyt >0, P

∃i:ξiλ< t, ξiλ6∈ Pλ0

= 0. (3.5)

The probability (3.5) is the increasing limit of the sequence(pn)n∈Ndefined as pn:=P

∃i:ξλi < t,∀j ≥i,|ξiλ−ξjλ0|> 1 2n

. (3.6)

(15)

It suffices to prove thatpn= 0for anyn. Let us introduce the probability pβn:=P

∃i: β

8πζiλ< t,∀j≥i, β

8π|ζiλ−ζjλ0|> 1 2n

. (3.7)

Denote byMp(R+)the space of point measures and recall that it is closed in the space M+(R+)for the vague convergence topology. Notice that the set

{(µ, ν)∈(Mp(R+))2 : µ=

X

i=1

δxi, ν =

X

i=1

δyi,∃isuch thatyi< tand∀j, |yi−xj|> 1 2n} is open in(Mp(R+))2equipped with the product topology of the vague convergence on the space of point measures. It comes from the straightforward fact that ifµk ∈ Mp(R+) converges towardsµ∈ Mp(R+)for the vague topology, the points ofµk belonging to [0, t]for all but finitely manykconverge to points ofµin[0, t].

If n is fixed, we therefore havepn ≤ lim infβ→0pβn using the joint convergence of (µβλ, µβλ0) in the space(Mp(R+))2 (along the subsequence (βk)) and the Portmanteau

theorem. It suffices to check that

lim sup

β→0

pβn = 0.

We need to work with a random subdivision of the interval[0,8πt/β]. As before, we consider a sequence(τk)k∈Nof i.i.d. positive random variables distributed uniformly in [12,32]and form the sumSk =Pk

i=1τi.

Noting that for all x, y such that |x−y| ≤ 1/(2n), there exists k ∈ N such that x, y∈[Sk/n, Sk/n+ 2/n], we obtain

pβn ≤P

∃k≤ b2ntc+ 1 :bαλ

2πcjumps on the interval 8π β

Sk n,(Sk

n + 2 n)

but notbαλ0 2πc

. Due to the increasing property, the event inside the probability can not happen if the process{αλ}starts below{αλ0}at the beginning of the interval. Therefore,

pβn

b2ntc+1

X

k=1

P

λ0(8π β

Sk

n)}≤ {αλ(8π β

Sk n )}

. which can in turn be upper-bounded as follows

b2ntc+1

X

k=1

P

λ0(8π β

Sk

n )}≤ {αλ(8π β

Sk

n )}

≤ βn 4πE

"

Z β(3t+1) 0

1{{αλ0(u)}≤{αλ(u)}}du

#

≤ βn

4πΘβ(3t+ 1) ≤βc (3.8) where we have used Lemma 3.1 to obtain the last inequality which holds forβ small enough (cis a constant which does not depend onβ).

Lemma 3.5.Let λ < λ0. Then the (Ft)-Poisson point processes Pλ and Pλ0−λ are independent.

Proof of Lemma 3.5. From a classical result (see Proposition (1.7) Chapter XII, §1, p.473 in [21]) on Poisson processes, we know that it suffices to prove that the two (Ft)-Poisson processesPλandPλ0−λ do not jump simultaneously, i.e. that fort >0,

Ph

∃i, j∈N:ξiλ< t, ξjλ0−λ< t, ξiλjλ0−λi

= 0. (3.9)

(16)

0 20 40 60 80 100

0.00.51.01.52.02.53.03.5

Figure 4: (Color online). Sample paths of the diffusionsαλ/(2π)(blue curve) andαλ0/(2π) (red curve) together with(αλ0−αλ)/(2π)(green curve) forλ < λ0and for a small value of β. Again we see thatbαλ/(2π)cand b(αλ0 −αλ)/(2π)cnever jump simultaneously whilebαλ0/(2π)candb(αλ0−αλ)/(2π)calways jump at the same times in agreement with Lemma 3.5, Lemma 3.4 and Remark 3.6.

Forn∈N, we consider the probability

pβn :=P

∃i, j∈N: β

8πζiλ< t, β

8πζjλ0−λ< t, β

8π|ζiλ−ζjλ0−λ|< 1 2n

. (3.10)

Ifnis fixed, then we have the convergence (the studied set is open as in the proof Lemma 3.4)

lim inf

β→0 pβn≥pn :=P

∃i, j∈N:ξλi < t, ξjλ0−λ< t, |ξiλ−ξλj0−λ|< 1 2n

.

To prove (3.9), it therefore suffices to prove that

lim sup

n→∞

lim sup

β→0

pβn = 0. (3.11)

For this, we need to work with a random subdivision of the interval [0,8πt/β]. As before, we consider a sequence(τk, k∈N)of i.i.d. positive random variables uniformly distributed in[1/2,3/2]and independent of the processes and form the sumSk=Pk

i=1τi

(S0:= 0).

(17)

8π β

Sk n

α λ

α λ + 2π α λ 0

2π`

2π(` + 1) 2π(` + 2)

Figure 5: Illustration for inclusion (3.15).

The probability (3.10) can be upper-bounded as follows P

∃i, j∈N, k≤ b2ntc+ 1 : 8π β

Sk

n ≤ζiλ, ζjλ0−λ≤8π β (Sk

n +2 n)]

=P

∃k≤ b2ntc+ 1 :bαλ0−αλ

2π candλ

2πcboth jump on the interval 8π β

Sk

n ,(Sk

n +2 n)

b2ntc+1

X

k=1

P

λ0(8π β

Sk

n)}≤ {αλ(8π β

Sk n )}

(3.12)

+

b2ntc+1

X

k=1

P

λ0

2πcjumps two times during the interval 8π β

Sk

n,(Sk

n + 2 n)

. (3.13)

For this bound, we have used the fact that, conditionally on {αλ(8π

β Sk

n )}≤ {αλ0(8π β

Sk

n )}, (3.14)

the increasing property and the equality in law (3.2) impose

λ0−αλ

2π candλ

2πcboth jump on the interval 8π β

Sk

n,(Sk

n + 2 n)

(3.15)

λ0

2πcjumps two times during the interval8π β

Sk

n,(Sk

n + 2 n)

.

Indeed, the equality in law (3.2) implies thatbαλ0(t)−α λ(t)cis increasing with respect tot (once the differenceαλ0 −αλ has reached the value2kπwherek ∈N, it remains forever above this value). This fact and the increasing property imply that under the event (3.15), the processbαλ0chas to jump two times on the interval β[Snk,(Snk +n2)]

(see Fig. 5). We now estimate the two sums in (3.12). The first one was already tackled in (3.8).

参照

関連したドキュメント

Keywords: continuous time random walk, Brownian motion, collision time, skew Young tableaux, tandem queue.. AMS 2000 Subject Classification: Primary:

BOUNDARY INVARIANTS AND THE BERGMAN KERNEL 153 defining function r = r F , which was constructed in [F2] as a smooth approx- imate solution to the (complex) Monge-Amp` ere

Abstract. Recently, the Riemann problem in the interior domain of a smooth Jordan curve was solved by transforming its boundary condition to a Fredholm integral equation of the

modular proof of soundness using U-simulations.. &amp; RIMS, Kyoto U.). Equivalence

Then it follows immediately from a suitable version of “Hensel’s Lemma” [cf., e.g., the argument of [4], Lemma 2.1] that S may be obtained, as the notation suggests, as the m A

Definition An embeddable tiled surface is a tiled surface which is actually achieved as the graph of singular leaves of some embedded orientable surface with closed braid

In order to be able to apply the Cartan–K¨ ahler theorem to prove existence of solutions in the real-analytic category, one needs a stronger result than Proposition 2.3; one needs

Our method of proof can also be used to recover the rational homotopy of L K(2) S 0 as well as the chromatic splitting conjecture at primes p &gt; 3 [16]; we only need to use the