• Aucun résultat trouvé

Addendum to " Sum rules via large deviations "

N/A
N/A
Protected

Academic year: 2021

Partager "Addendum to " Sum rules via large deviations ""

Copied!
8
0
0

Texte intégral

(1)

HAL Id: hal-01369130

https://hal.archives-ouvertes.fr/hal-01369130v2

Preprint submitted on 24 Feb 2017

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or

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

Addendum to ” Sum rules via large deviations ”

Fabrice Gamboa, Jan Nagel, Alain Rouault

To cite this version:

Fabrice Gamboa, Jan Nagel, Alain Rouault. Addendum to ” Sum rules via large deviations ”. 2017. �hal-01369130v2�

(2)

Addendum to ”Sum rules via large deviations”

Fabrice Gamboa∗ Jan Nagel† Alain Rouault‡

February 24, 2017

Abstract

In these notes we fill a gap in a proof in Section 4 of Gamboa, Nagel, Rouault [Sum rules via large deviations, J. Funct. Anal. 270 (2016), 509-559]. We prove a general theorem which combines a LDP with a convex rate function and a LDP with a non-convex one. This result will be used to prove LDPs for spectral matrix measures and for spectral measures on the unit circle.

Keywords : Large deviations, random measures MSC 2010: 60F10

1

Introduction

In Section 4 of [4], we studied large deviations for a pair of random variables with values in topological vector spaces by means of the joint normalized generating function. However, in some cases, the rate function of one of the marginals is not convex, which invalidates this way of proof. Actually, it is possible to state a general theorem which combines a LDP with a convex rate function and a LDP with a non-convex one. It is used to prove LDPs in [3] for spectral matrix measures and in [5] for spectral measures on the unit circle.

Since this theorem may have its own interest, we give it in a general setting in Section 3 ∗Universit´e Paul Sabatier, Institut de Math´ematiques de Toulouse, 31062 Toulouse Cedex 9, France, e-mail:

gamboa@math.univ-toulouse.fr

Technische Universitat M¨unchen, Fakult¨at f¨ur Mathematik, Boltzmannstr. 3, 85748 Garching, Germany,

e-mail: jan.nagel@tum.de

Laboratoire de Math´ematiques de Versailles, UVSQ, CNRS, Universit Paris-Saclay, 78035-Versailles Cedex

(3)

after recalling classical results in Section 2. We come back in Section 4 to the framework of [4].

In the sequel, we assume that X and Y are Hausdorff topological vector spaces. X∗ is the topological dual of X and X is endowed with the weak topology. We denote by Cb(Y) the set of all bounded continuous functions ϕ : Y → R. A point x ∈ X is called an exposed point of a function F on X , if there exists x∗ ∈ X∗ (called an exposing hyperplane for x) such that

F(x) − hx∗, xi < F (z) − hx∗, zi (1.1)

for all z 6= x.

2

Some classical results in large deviations

Let us recall two well known results in the theory of large deviations which will be combined in order to solve our problem. The first result is the inverse of Varadhan’s lemma (Theorem 4.4.2 in [1]), the second one is a version of the so-called Baldi’s theorem (Theorem 4.5.20 in [1]). The latter differs from the version in [1] in a straightforward condition to identify the rate function, which was applied for instance in [6] (see also [2]). The proof of our Theorem 3.1 will be quite similar to the proof of these two classical theorems.

Theorem 2.1 (Bryc’s Inverse Varadhan Lemma) Suppose that the sequence (Yn) of

random variables in Y is exponentially tight and that the limit

Λ(ϕ) := lim n→∞

1 nlog Ee

nϕ(Yn)

exists for every ϕ∈ Cb(Y). Then (Yn) satisfies the LDP with the good rate function

I(y) = sup ϕ∈Cb(Y)

{ϕ(y) − Λ(ϕ)} .

Furthermore, for every ϕ∈ Cb(Y),

Λ(ϕ) = sup y∈Y

{ϕ(y) − I(y)} .

Theorem 2.2 (A version of Baldi’s Theorem) Suppose that the sequence (Xn) of

ran-dom variables in X is exponentially tight and that :

1. There is a set D ⊂ X∗ and a function GX : D → R such that for all x∗ ∈ D

(2.1) lim

n→∞ 1

nlog E exp (nhx

(4)

2. If F denotes the set of exposed points x of

G∗X(x) = sup x∗∈D{hx

, xi − GX(x)},

with an exposing hyperplane xsatisfying x∈ D and γx∈ D for some γ > 1,

then for every x ∈ {G∗

X < ∞} there exists a sequence (xk)k with xk ∈ F such that limk→∞xk= x and

lim k→∞G

X(xk) = G∗X(x).

Then (Xn) satisfies the LDP with good rate function G∗X.

3

A general theorem

Our extension is the following combination of the two above theorems. The main point is that the rate function does not need to be convex, but we still only need to control linear functionals of Xn.

Theorem 3.1 Assume that Xn ∈ X and Yn ∈ Y are defined on the same probabilistic

space. Moreover, we assume that the two sequences (Xn) and (Yn) are exponentially tight.

Assume further that:

1. There is a set D ⊂ X∗ and functions G

X : D → R, J : Cb(Y) → R such that for all x∗ ∈ D and ϕ ∈ Cb(Y) (3.1) lim n→∞ 1 nlog E exp (nhx ∗, Xni + nϕ(Yn)) = GX(x) + J(ϕ) .

2. If F denotes the set of exposed points x of

G∗X(x) = sup x∗∈D{hx

, xi − GX(x)}

with an exposing hyperplane xsatisfying x∈ D and γx∈ D for some γ > 1,

then for every x ∈ {G∗

X < ∞} there exists a sequence (xk)k with xk ∈ F such that limk→∞xk= x and

lim k→∞G

X(xk) = G∗X(x).

Then, the pair (Xn, Yn) satisfies the LDP with speed n and good rate function I(x, y) = G∗X(x) + IY(y) ,

where

(5)

Let us note that in view of Varadhan’s Lemma we have

J(ϕ) = sup y∈Y

{ϕ(y) − IY(y)}.

Proof:

Upperbound: The proof follows the lines of the proof of part (b) of Theorem 4.5.3 in [1]. Note that since the sequence (Xn, Yn) is exponentially tight it suffices to show the upper bound for compact sets. Furthermore, the rate is necessarily good, since, if in (3.1) we set x∗ = 0 (resp. ϕ = 0) it reduces to Theorem 2.1 (resp. Theorem 2.2).

Lowerbound: As usual, it is enough to consider a neighbourhood ∆1× ∆2 of (x, y) where I(x, y) < ∞, take lim infn→∞n1 log P((Xn, Yn) ∈ ∆1× ∆2) and get a lower bound tending to I(x, y) when the infimum over all neighborhoods is taken. Actually, due to the density assumption 2. it is enough to study the lower bound of P(Xn∈ ∆1, Yn∈ ∆2) when x∈ F and IY(y) < ∞.

As in [1] (Proof of Lemma 4.4.6), let ϕ : Y → [0, 1] be a continuous function, such that ϕ(y) = 1 and ϕ vanishes on the complement ∆c

2 of ∆2. For m > 0, define ϕm := m(ϕ − 1). Note that J(ϕm) ≥ −IY(y) . We have P(Xn∈ ∆1, Yn ∈ ∆2) = E h 1 {Xn∈∆1}1 {Yn∈∆2}e nhx∗,X ni+nϕm(Yn) e−nhx∗,Xni−nϕm(Yn) i .

Now −ϕm≥ 0 and on ∆1, −hx∗, Xni ≥ −hx∗, xi − δ for a δ > 0, so that

(3.2) P(Xn∈ ∆1, Yn∈ ∆2) ≥ E h 1 {Xn∈∆1}1 {Yn∈∆2}e nhx∗,X ni+nϕm(Yn) i e−nhx∗,xi−nδ. Denoting ℓn= 1 nlog Ee nhx∗,X ni , Ln:= 1 nlog Ee nhx∗,X ni+nϕm(Yn)

and eP the new probability on X × Y such that deP dP = e nhx∗ ,Xni+nϕm(Yn)−nLn, we get (3.3) P(Xn∈ ∆1, Yn∈ ∆2) ≥ eP(Xn∈ ∆1, Yn∈ ∆2)e−nhx ∗ ,xi−nδ+nLn .

For the exponential term we have

(3.4) lim inf n→∞ 1 nlog e −nhx∗ ,xi−nδ+nLn ≥ hx∗, xi − δ + GX(x∗) + J(ϕm) ≥ −G∗X(x) − IY(y) − δ.

(6)

We may choose δ arbitrarily small by choosing ∆1sufficiently small, so that it will be enough to prove that (3.5) P(Xne ∈ ∆1, Yn∈ ∆2) −−−→ n→∞ 1 or equivalently, that (3.6) eP(Xn∈ ∆c1) + eP(Yn∈ ∆c2) −−−→ n→∞ 0 .

For the first term, note that under eP the moment generating function of Xn satisfies

lim n→∞ 1 nlog eE[e nhz∗,X ni ] = lim n→∞ 1 nlog E[e nhz∗ +x∗,X ni+ϕm(Yn)−nLn ] = GX(z∗+ x∗) + J(ϕm) − GX(x∗) − J(ϕm) = GX(z∗+ x∗) − GX(x∗) =: eGX(z∗),

for z∗ ∈ eD := {z∗ : x∗+ z∗ ∈ D}. We may then follow the argument on p.159-160 in [1] (as an auxiliary result in their proof of the lower bound). Using that x∗ ∈ D is an exposing hyperplane, we get lim sup n→∞ 1 nlog eP(Xn∈ ∆ c 1) < 0. Considering the second term in (3.6), we have, on ∆c

2 deP dP = e −nm+nhx∗,X ni−nLn so that e P(Yn∈ ∆c2) ≤ e−nm+nℓn−nLn . Taking the logarithm, this implies

lim sup n→∞ 1 nlog eP(Yn∈ ∆ c 2) ≤ −m + GX(x∗) − GX(x∗) − J(ϕm) = −m − sup z∈Y {ϕm(z) − IY(z)} ≤ −m + IY(y)

which tends to −∞ when m → ∞.

To summarize, we have proved (3.6), i.e. (3.5), which with (3.3) and (3.4) gives

lim ∆1↓x,∆2↓y lim inf n→∞ 1 nlog P(Xn∈ ∆1, Yn∈ ∆2) ≥ −G ∗ X(x) − IY(y) ,

(7)

4

Joint LDP for measure and truncated

eigenval-ues

In Section 4 of [4], we studied the joint moment generating function of a non-negative measure ˜µ(n)I(j) on a compact set [α−, α+] and a collection of j extremal support points λ±M(j) ∈ R2j, restricted to the compact set [−M, M ]. For the sake of a clearer notation, we drop here the dependency on j. It is shown in Theorem 4.1 in [4], that λ±M satisfies the LDP with speed n and good rate IM,λ±. Furthermore, the sequence of ˜µ(n)

I is exponentially tight and if Gn(f, s) = E  exp  n Z f d˜µ(n)I + nhs, λ±Mi  ,

then for all f such that log(1 − f ) is continuous and bounded and all s ∈ R2j,

lim n→∞

1

nlog Gn(f, s) = G(f ) + H(s), (4.1)

with G∗ strictly convex on a set of points dense in {G∗ <∞}.

However, the rate IM,λ± might be non-convex and hence the dual H∗ is not strictly

convex on a dense set. The convergence in (4.1) is therefore not enough to conclude the joint LDP for (˜µ(n)I , λ±M) directly from the classical Theorem 2.2.

To show the joint LDP, we will apply Theorem 3.1. Indeed, let D be the set of bounded continuous functions f from [α−, α+] to R such that supxf(x) < 1. If we define for ϕ : R2j → R continuous and bounded and f ∈ D

ˆ Gn(f, ϕ) = E  exp  n Z f dµ˜(n)I + nϕ(λ±M)  ,

then the same arguments as in Section 4 of [4] show

lim n→∞ 1 nlog ˆGn(f, ϕ) = G(f ) + J(ϕ), (4.2) where G(f ) = − Z log(1 − f ) dµV

for a probability measure µV. Moreover, in Section 4 of [4] it is shown that every measure on [α−, α+] with a strictly positive continuous density h with respect to µ

V is an exposed point and the exposing hyperplane is the function f = 1 − h−1. Since f is continuous and strictly less than 1, γf ∈ D for γ > 1 small enough. By the same arguments as in [6], any measure µ with G∗(µ) < ∞ may be approximated weakly by measures µnwith such a strictly positive continuous density such that G∗(µn) converges to G∗(µ). This approximation is also made

(8)

more precise for matrix valued measures in [3]. The assumptions of Theorem 3.1 are then satisfied, which yields the LDP for (˜µ(n)I , λ±M) with good rate

I(µ, λ) = G∗(µ) + IM,λ±.

After taking the limit M → ∞, this proves the statement of Theorem 4.2 of [4].

References

[1] A. Dembo and O. Zeitouni. Large Deviations Techniques and Applications. Springer, 1998.

[2] H. Dette and F. Gamboa. Asymptotic properties of the algebraic moment range process.

Acta Mathematica Hungarica, 116:247–264, 2007.

[3] F. Gamboa, J. Nagel, and A. Rouault. Sum rules and large deviations for spectral matrix measures. arXiv preprint arXiv:1601.08135, 2016.

[4] F. Gamboa, J. Nagel, and A. Rouault. Sum rules via large deviations. J. Funct. Anal., 270(2):509 – 559, 2016.

[5] F. Gamboa, J. Nagel, and A. Rouault. Sum rules and large deviations for spectral measures on the unit circle. Random Matrices Theory Appl., 6(1), 2017.

[6] F. Gamboa, A. Rouault, and M. Zani. A functional large deviation principle for quadratic forms of Gaussian stationary processes. Stat. and Probab. Letters, 43:299–308, 1999.

Références

Documents relatifs

These results apply to the fluctuations of Brownian motion, Birkhoff averages on hyperbolic dynamics, as well as branching random walks.. Also, they lead to new insights into

[AGZ10] G. An introduction to random matrices. Cambridge University Press, Cambridge, 2010... Large deviations and sum rules for spectral theory - A pedagogical approach.

When we further impose that these boxes are visited by another independent walk, one can sum over all possible centers of the boxes, and this yields some bound on the number of

We shall apply Shao’s method in [14] to a more general class of self-normalized large deviations, where the power function x 7→ x p used as a generator function for the normalizing

In a previous work ([10]), we proved that the sequence (µ (N) w ) satisfies a Large Deviation Principle (denoted hereafter LDP), with speed N and good rate function given by

Since Σ T is invariant, the Verblunsky coefficients are Hermitian (see Lemma 4.1 in [5]).. An introduction to random matrices. Large deviations and sum rules for spectral theory -

In many random combinatorial problems, the distribution of the interesting statistic is the law of an empirical mean built on an independent and identically distributed (i.i.d.)

Large deviations for these random spectral measures arising from the classical ensembles of ran- dom matrix theory have been considered before in Gamboa and Rouault (2011)..