• Aucun résultat trouvé

No temporal distributional limit theorem for a.e. irrational translation

N/A
N/A
Protected

Academic year: 2022

Partager "No temporal distributional limit theorem for a.e. irrational translation"

Copied!
22
0
0

Texte intégral

(1)

D M I T R Y D O L G O P Y AT O M R I S A R IG

N O T E M P O R A L DIS T R I B U T IO N A L L I M I T T H E O R E M F O R A. E.

IR R AT IO N A L T R A N S L AT IO N

U N E T R A N S L AT IO N IR R AT IO N N E L L E T Y PIQ U E N E S AT IS F A I T PA S D E T H É O R È M E L I M I T E T E M P O R E L

Abstract. — Bromberg and Ulcigrai constructed piecewise smooth functions on the circle such that the set ofαfor which the sumPn−1

k=0f(x+ mod 1) satisfies a temporal distributional limit theorem along the orbit of a.e.xhas Hausdorff dimension one. We show that the Lebesgue measure of this set is equal to zero.

Résumé. — Bromberg et Ulcigrai ont construit des fonctions lisses par morceaux sur le cercle pour lesquelles l’ensemble desαtels que la sommePn−1

k=0f(x+ mod 1) satisfait un théorème limite temporel le long de l’orbite de presque toutxest un ensemble de dimension de Hausdorff 1. Nous montrons que cet ensemble est de mesure nulle.

1. Introduction and statement of main result

1.1. Background

SupposeT :XXis a map,f :X →Ris a function, andx0Xis afixed initial condition. We say that theT–ergodic sums Sn =f(x0) +f(T x0) +· · ·+f(Tn−1x0)

2010Mathematics Subject Classification:37D25, 37D35.

DOI:https://doi.org/10.5802/ahl.4

(*) This work was partially supported by the ISF grants 199/14, 1149/18 and by the BSF grant 2016105. The authors thank Adam Kanigowski for simplifying the statement and the proof of Lemma 3.2.

(2)

satisfy a temporal distributional limit theorem (TDLT) on the orbit of x0, if there exists a non-constant real valued random variable Y, centering constants AN ∈ R and scaling constantsBN → ∞s.t.

(1.1) SnAN

BN −−−→

N→∞ Y in distribution,

whennis sampled uniformly from {1, . . . , N}andx0 is fixed. Equivalently, for every Borel set E ⊂Rs.t. P(Y ∈∂E) = 0,

1 N Card

16n6N : SnAN

BNE

−−−→N→∞ P(Y ∈E).

We allow and expect AN, BN, Y to depend on T, f, x0.

Such limit theorems have been discovered for several zero entropy uniquely ergodic transformations, including systems where the more traditional spatial limit theorems, withx0 is sampled from a measure onX, fail [Bec10, Bec11, ADDS15, DS17b, PS17, DS18]. Of particular interest are TDLT for

Rα : [0,1]→[0,1], Rα(x) =x+α mod 1, fβ(x) := 1[0,β)(x)−β,

because the Rα–ergodic sums of fβ along the orbit of x represent the discrepancy of the sequence x+ mod 1 with respect to [0, β) [Sch78, CK76, Bec10]. Another source of interest is the connection to the “deterministic random walk” [AK82, ADDS15].

The validity of the TDLT forRα andfβ depends on the diophantine properties ofα andβ. Recall thatα∈(0,1) is badly approximable if for some c >0,|qα−p|>c/|q|

for all irreducible fractions p/q. Equivalently, the digits in the continued fraction expansion ofα are bounded [Khi63]. Say that β ∈(0,1) is badly approximable with respect to α if for some C > 0, |qα−βp| > C/|q| for all p, q ∈ Z, q 6= 0. If α is badly approximable then everyβ ∈Q∩(0,1) is badly approximable with respect to α. The recent paper [BU18] shows:

Theorem 1.1 (Bromberg–Ulcigrai [BU18]). — Suppose α is badly approximable and β is badly approximable with respect to α, e.g. β ∈ Q∩(0,1). Then the Rα- ergodic sums of fβ satisfy a temporal distributional limit theorem with Gaussian limit on the orbit of every initial condition.

The set of badly approximableαhas Hausdorff dimension one [Jar29], but Lebesgue measure zero [Khi24]. This leads to the following question:Is there a β s.t. the Rα ergodic sums of fβ satisfy a temporal distributional limit theorem for a.e. α and a.e.

initial condition?

In this paper we answer this question negatively.

1.2. Main result

To state our result in its most general form, we need the following terminology.

LetT :=R/Z. We say that f :T→R is piecewise smooth if there exists a finite set S ⊂ T s.t. f is continuously differentiable on T\S and ∃ ψ : T → R with bounded variation s.t. f0 = ψ on T\S. For example: fβ(x) = 1[0,β)(x)−β (take S={0, β},ψ ≡0). We show:

(3)

Theorem 1.2. — Letf be a piecewise smooth function of zero mean. Then there is a set of full measureE ⊂ T×T s.t. if (α, x)∈ E then the Rα–ergodic sums of f do not satisfy a TDLT on the orbit of x.

The conditionRTf = 0 is necessary: By Weyl’s equidistribution theorem, for every α 6∈ Q, f Riemann integrable s.t. R

Tf = 1, and x0 ∈ T, Sn/N −−−→dist

N→∞ U[0,1] as n∼U(1, . . . , N). See Section 1.4 for the notation.

This paper has a companion [DS17a] which gives a different proof of Theorem 1.2, in the special case f(x) ={x} − 12. Unlike the proof given below, [DS17a] does not identify the set ofαwhere the TDLT fails, but it does give more information on the different scaling limits for the distributions of Sn, n ∼U(1, . . . , Nk) along different subsequences Nk → ∞. [DS17a] also shows that if we randomize both n and α by sampling (n, α) uniformly from {1, . . . , N} ×T, then SnN1 PNk=1Sk/

lnN converges in distribution to the Cauchy distribution.

The methods of [DS17a] are specific for f(x) ={x} −12, and we do not know how to apply them to other functions such asfβ(x) = 1[0,β)(x)−β.

1.3. The structure of the proof Supposef is piecewise smooth and has mean zero.

We shall see below that iff is continuous, then for a.e. α, f is an Rα-coboundary, therefore Sn are bounded, hence (1.1) cannot hold with BN → ∞, Y non-constant.

We remark that (1.1) does hold withBN ≡1,AN =f(x0),Y = distribution of minus the transfer function, but this is not a TDLT since no actual scaling is involved.

The heart of the proof is to show that if f is discontinuous, then for a.e. α, the temporal distributions of the ergodic sums have different asymptotic scaling behavior on different subsequences. The proof of this has three independent parts:

(1) A reduction to the case f(x) = Pdm=1bmh(x+βm),h(x) := {x} −12.

(2) A proof that if N ⊂ N has positive lower density, then there exists M > 1 s.t. the following set has full Lebesgue measure in (0,1):

A(N, M) :=

(

α ∈(0,1) : ∃nk ↑ ∞, rk6M s.t. for all k:

rkqnk ∈ N, ank+1/(a1+· · ·+ank)→ ∞

)

.

Here an andqn are the partial quotients and principal denominators of α, see Section 3.1.

(3) Construction of N =N(b1, . . . , bd;β1, . . . , βd)⊆N with positive density, s.t.

for everyα∈ A(N, M) and a.e.x, one can analyze the temporal distributions of the Birkhoff sums of Pdm=1bmh(x+βm).

1.4. Notation

n ∼U(1, . . . , N) means that n is a random variable taking values in {1, . . . , N}, each with probability N1. U[a, b] is the uniform distribution on [a, b]. Lebesgue’s measure is denoted by mes. N = {1,2,3, . . .} and N0 = N∪ {0}. If x ∈ R, then kxk:= dist(x,Z) and {x} is the unique number in [0,1) s.t.x ∈ {x}+Z. Card(·) is the cardinality. Ifε >0, then a=b±ε means that |a−b|6ε.

(4)

2. Reduction to the case f (x) =

Pdm=1

b

m

h(x + β

m

)

Let h(x) = {x} − 12, and let G denote the collection of all non-identically zero functions of the formf(x) = Pdm=1bmh(x+βm), whered∈N, bi, βi ∈R.We explain how to reduce the proof of Theorem 1.2 from the case of a general piecewise smooth f(x) to the casef ∈ G.

The following proposition was proved in [DS17a]. Let C(T) denote the space of continuous real-valued functions on T with the sup norm.

Proposition 2.1. — Iff(t) is differentiable on T\ {β1, . . . , βd} and f0 extends to a function with bounded variation on T, then there ared ∈N0,b1, . . . , bd∈R s.t.

for a.e. α∈T there isϕαC(T) s.t.

f(x) =

d

X

i=1

bih(x+βi) +

Z

T

f(t) dt+ϕα(x)−ϕα(x+α) (x6=β1, . . . , βd).

The following proposition was proved in [DS17b]. Let (Ω,B, µ) be a probability space, and let T : Ω→Ω be a probability preserving map.

Proposition 2.2. — Suppose f = g +ϕϕT µ-a.e. with f, g, ϕ : Ω → R measurable. If the ergodic sums of g satisfy a TDLT along the orbit of a.e.x, then so do the ergodic sums of f.

These results show that if Theorem 1.2 holds for every f ∈ G, then Theorem 1.2 holds for anydiscontinuous piecewise smooth function with zero mean. As for con- tinuous piecewise smooth functions with zero mean, these are Rα-cohomologous to g ≡ 0 for a.e. α because the bi in Proposition 2.1 must all vanish. Since the zero function does not satisfy the TDLT, continuous piecewise smooth functions do not satisfy a TDLT.

3. The set A has full measure

3.1. Statement and plan of proof

Let α be an irrational number, with continued fraction expansion denoted by [a0;a1, a2, a3, . . .] :=a0 + 1

a1 +. . ., a0 ∈ Z, ai ∈N (i> 1). We call an the quotients of α. Let pn/qn denote the principal convergents of α, determined recursively by

qn+1 =an+1qn+qn−1, pn+1 =an+1pn+pn−1

andp0 =a0, q0 = 1;p1 = 1+a1a0,q1 =a1. We callqntheprincipal denominatorsand ai the partial quotients ofα. Sometimes – but not always! – we will writeqk=qk(α), pk =pk(α), ak =ak(α).

Given N ⊂ Nand M >1, let A =A(N, M)⊂(0,1) denote the set of irrational α∈(0,1) s.t. for some subsequence nk↑ ∞,

(3.1) ∃rk 6M s.t. rkqnk ∈ N, ank+1

(a0+· · ·+ank) −−−→

k→∞ ∞.

Thelower density of N is d(N) := lim inf N1 Card(N ∩[1, N]). The purpose of this section is to prove:

(5)

Theorem 3.1. — If a setN has positive lower density, then there exists M such that A(N, M) has full Lebesgue measure in (0,1).

The proof consists of the following three lemmas:

Lemma 3.2. — For almost all α there is n0 = n0(α) s.t. if k > n0 and ak+1 >

1

4k(lnk)(ln lnk), then ak+1/(a1+· · ·+ak)> 18ln lnk.

Lemma 3.3. — Supposeα ∈(0,1)\Q and(p, q) ∈N0×N satisfy gcd(p, q) = 1 and|qα−p|6 qL1 whereL>4. Then there existsk s.t.q=qk(α)andak+1(α)> 12L.

Lemma 3.4. — Suppose ψ :R+ →Ris a non-decreasing function s.t.

(3.2) X

n

1

nψ(n) =∞.

Suppose N ⊂ N has positive lower density. For all M sufficiently large, for a.e.

α∈(0,1)there are infinitely many pairs(m, n)∈N0×Ns.t.n∈ N,gcd(m, n)6M, and |nα−m|6 nψ(n)1 .

Remark 3.5. — By the monotonicity of ψ, if ek−1 < n < ek then ψek−1 6 ψ(n)6ψek. Hence (3.2) holds iff Pkψ(e1k) =∞.

Remark 3.6. — If N = N, then Lemma 3.4 holds with M = 1 by the classical Khinchine Theorem. We do not know if Lemma 3.4 holds withM = 1 for every set N with positive lower density.

Proof of Theorem 3.1 given Lemmas 3.2–3.4. — We apply these lemmas with ψ(t) =c(lnt) (ln lnt) (ln ln lnt) and c >1/ln(1+

5 2 ).

Fix M > 1 as in Lemma 3.4. Then ∃ Ω ⊂ (0,1) of full measure s.t. for every α∈Ω there are infinitely many (m, n)∈N0×Nas follows. Let m :=m/gcd(m, n), n :=n/gcd(m, n),p:= gcd(m, n), then

(1) pn ∈ N, p6M, |nαm|= |nα−m|p 6 nψ(n1 ) (∵n 6n);

(2) ∃ k s.t. n = qk(α) and ak+1(α) > 12ψ(qk) (∵ Lemma 3.3). By its recursive definition, qk> k-th Fibonacci number> 13(1+

5

3 )k. So for all k large enough, ak+1(α)> 12ψ(qk)> 14k(lnk)(ln lnk);

(3) ak+1/(a1+· · ·+ak)> 18ln lnk → ∞ (∵Lemma 3.2).

So everyα ∈Ω belongs to A =A(N, M), andA has full measure.

Next we prove Lemmas 3.2–3.4.

3.2. Proof of Lemma 3.2 By [DV86], for almost everyα

(a1+· · ·+ak+1)−maxj6k+1aj

klnk → 1

ln 2 <2.

So ifk is large enough, and ak+1 > 14k(lnk)(ln lnk) then

j6k+1maxaj =ak+1, a1+· · ·+ak

klnk 62, and ak+1

a1+· · ·+ak > 1

8ln lnk.

(6)

3.3. Proof of Lemma 3.3

For every (p, q) as in the lemma, |qα−p|< 2q1. A classical result in the theory of continued fractions [Khi63, Theorem 19] says that in this case∃k s.t.q=qk(α), p= pk(α).

To estimate ak+1 = ak+1(α) we recall the following facts, valid for the principal denominators of any irrationalα∈(0,1) [Khi63]:

(1) |qkαpk|> q 1

k+qk+1;

(2) qk+1+qk <(ak+1+ 2)qk, whence by (a) ak+1 > q 1

k|qkα−pk|−2.

In our case, |qkαpk|=|qα−p|6 qk1L, so ak+1 > L−2> L2. 3.4. Preparations for the proof of Lemma 3.4

Let (Ω,F,P) be a probability space, and Ak ∈ F be measurable events. Given D >1, we say that Ak are D-quasi-independent, if

(3.3) P(Ak1Ak2)6DP(Ak1)P(Ak2) for all k1 6=k2.

The following proposition is a slight variation on Sullivan’s Borel–Cantelli Lemma from ([Sul82]):

Proposition 3.7. — For every D > 1 there exists a constant δ(D) > 0 such that the following holds in any probability space:

(a) If Ak are D-quasi-independent measurable events s.t. limk→∞P(Ak) = 0 but

P

kP(Ak) = ∞, thenP(Ak occurs infinitely often)>δ(D).

(b) The quasi-independence assumption in(a)can be weakened to the assumption that for somer ∈N,P(Ak1Ak2)6DP(Ak1)P(Ak2) for all |k2k1|>r.

(c) One can take δ(D) = 2D1 .

Proof. — Since P(Ak)→0 but PP(Ak) =∞, there is an increasing sequence Nj such that limj→∞PNk=Nj+1

j+1P(Ak) = D1.

Let Bj be the event that at least one of events {Ak}Nk=Nj+1j+1 occurs. Since Bj =

UNj+1 k=Nj+1

Ak\Sk−1j=N

j+1Aj, P(Bj)>

Nj+1

X

k=Nj+1

P(Ak)− X

Nj+16k1<k26Nj+1

P(Ak1Ak2)

>

Nj+1

X

k=Nj+1

P(Ak)−D X

Nj+16k1<k26Nj+1

P(Ak1)P(Ak2)

>

Nj+1

X

k=Nj+1

P(Ak)− D 2

Nj+1

X

k=Nj+1

P(Ak)

2

.

Since limj→∞PNk=Nj+1

j+1P(Ak) = D1 and D>1,lim infP(Bj)> 2D1 .

LetE denote the event that Aj happens infinitely often. E is also the event that Bj happens infinitely often, therefore E = Tn=1Sj=n+1Bj. In a probability space,

(7)

the measure of a decreasing intersection of sets is the limit of the measure of these sets. SoP(E)>lim infP(Bj)> 2D1 , proving (a) and (c).

Part (b) follows from part (a) by applying it to the sets{Akr+`}where 06`6r−1

is chosen to getPkP(Akr+`) = ∞.

Themultiplicity of a collection of measurable sets{Ek}is defined to be the largest K s.t. there areK different ki with P(TKi=1Eki)>0.

Proposition 3.8. — Let Ek be measurable sets in a finite measure space. If the multiplicity of {Ek} is less than K, then

mes [

k

Ek

!

> 1 K

X

k

mes(Ek).

Proof. — 1S

iEi > K1 Pi1Ei almost everywhere.

Proposition 3.9. — For every non-empty open intervalI ⊂[0,1], Card

(m, n)∈ {0, . . . , N}2 : m

nI, gcd(m, n) = 1

∼3 mes(I)N22, as N→ ∞.

Proof. — This classical fact due to Dirichlet follows from the inclusion-exclusion principle and the identityζ(2) =π2/6, see [HW08, Theorem 459].

Proposition 3.10. — Suppose α = [0;a1, a2, . . .] and α = [0;a`+1, a`+2, . . .].

Then the principal convergents p`/q` of α and the principal convergents p`/q` of α are related by

pl+l pl+l+1 ql+l ql+l+1

!

= pl−1 pl ql−1 ql

! pl pl+1 ql ql+1

!

. Proof. — Since a0 = 0, the recurrence relations forpn/qn imply

pn pn+1 qn qn+1

!

= pn−1 pn qn−1 qn

! 0 1 1 an+1

!

, p0 p1 q0 q1

!

= 0 1

1 a1

!

. So

pn pn+1 qn qn+1

!

= 0 1

1 a1

!

· · · 0 1 1 an+1

!

. It follows that

pl+l pl+l+1 ql+l ql+l+1

!

= pl−1 pl ql−1 ql

! pl pl+1 ql ql+1

!

,

where pi/qi are the principal convergents ofα:= [0;al+1, al+2, . . .].

3.5. Proof of Lemma 3.4

Without loss of generality, limt→∞ψ(t) = ∞, otherwise replace ψ(t) by the bigger monotone functionψ(t) + lnt.

(8)

FixM >1, to be determined later. Let

k:={(m, n)∈N2 :n∈ N, n∈[ek−1, ek],0< m < n,gcd(m, n)6M}, Am,n,k:=

(

α∈T:|nα−m|6 1 ekψ(ek)

)

, Ak:= [

(m,n)∈Ωk

Am,n,k,

A:={α∈T:α belongs to infinitely many Ak}.

The lemma is equivalent to saying that A has full Lebesgue measure for a suitable choice of M.

We will prove a slightly different statement. Fixε >0 small. Given an non-empty intervalI ⊂[ε,1−ε], let

k(I) :=

(m, n)∈Ωk : m nI

, Ak(I) := [

(m,n)∈Ωk(I)

Am,n,k,

A(I) := {α∈T:α belongs to infinitely many Ak(I)}.

We will prove that there exists a positive constantδ =δ(ε, M) s.t. for all intervals I ⊂[ε,1−ε], mes(A(I)I)>δmes(I). It then follows by a standard density point argument (see below) that A ∩[ε,1−ε] has full measure. Since ε is arbitrary, the lemma is proved.

Claim 1. — There existsK =K(ε) such that for every k > K, the multiplicity of {Am,n,k}(m,n)∈Ωk(I) is uniformly bounded byM.

Proof. — Suppose (mi, ni) ∈ Ωk(I) (i = 1,2) and Am1,n1,kAm2,n2,k 6= ∅. Then there is α s.t. |niαmi| 6δk := ekψ(e1 k) (i = 1,2). Choose K =K(ε) so large that k > Kδk< 4eεk.

Ifk > K, thenα> mniiδk >minIε2 > ε2. Let ri := gcd(mi, ni) and (ni, mi) :=

1

ri(ni, mi). Then |niαmi| 6 δk and mi 6 ni 6 ni 6 ek, so |n2m1n1m2| =

1

α|m1(n2αm2)−m2(n1αm1)|6 2eε/2kδk <1. Son2m1 =n1m2. Since gcd(ni, mi) = 1, (n1, m1) = (n2, m2). It follows that (n2, m2)∈ {(rn1, rm1) :r= 1, . . . , M}. So the multiplicity of {Am,n,k}(m,n)∈Ωk(I) is uniformly bounded by M. Claim 2. — Let d(N) := lim inf N1 Card(N ∩ [1, N]) > 0, then there exists M =M(N) and Kf=K(ε,f N,|I|) s.t. for allk > K,f

(3.4) d(N) mes(I)

4M ψ(ek) 6mes(Ak(I))6 6 mes(I) ψ(ek) . In particular,mes(Ak(I))−−−→

k→∞ 0 and Pmes(Ak(I)) =∞.

Proof. — mes(Am,n,k) = meshmnrm,n,mn +rm,ni= 2rm,nwhererm,n = nekψ(e1 k). Sincen ∈[ek−1, ek],

(3.5) Card(Ωk(I))

M e2kψ(ek) 6mesAk(I)6 e Card(Ωk(I)) e2kψ(ek) ,

(9)

where the lower bound uses Claim 1 and Proposition 3.8.

Card(Ωk(I)) satisfies the boundsAB 6Card(Ωk(I))6A where A:= Card

(m, n) :n∈ N, n ∈[ek−1, ek], m nI

B := Card

(m, n) :n∈ N, n ∈[ek−1, ek], m

nI,gcd(m, n)>M

. Choose Kf=Kf(ε,N,|I|)> K(ε) s.t. for all k >Kf

(1) Card{n ∈ N : 06n 6ek}> 12d(N)

(2) Card{n ∈[ek−1, ek]∩N:p|n}62(ekek−1)/p for all p>1;

(3) For all n > eK−1e , p>1, n

p

2mes(I)6Card

m∈N: m

nI, p|m

6 2n

p mes(I).

Ifk >Kf, then 12d(N)e2kmes(I)6A62e2kmes(I) and B 6

X

p=M

Card

(m, n) :n ∈[ek−1, ek], m

nI, p|m, p|n

6

X

p=M

2(ekek−1)

p · 2ekmes(I)

p <4e2kmes(I)

X

p=M

1 p2 6 1

4d(N)e2kmes(I), provided we choose M s.t.

X

p=M

p−2 < 1

16d(N).

Together we get 14d(N)e2kmes(I) 6 Card(Ωk(I)) 6 2e2kmes(I). The claim now

follows from (3.5).

Claim 3. — There exists D=D(N, M), r=r(M), andKˆ = ˆK(ε,N, I) s.t. for all k1, k2 >Kˆ s.t.|k1k2|> r(M),

(3.6) mes(Ak1(I)∩ Ak2(I)|I)6Dmes(Ak1(I)|I) mes(Ak2(I)|I) Proof. — By Claim 2, if k1, k2 are large enough, then

(3.7) mes(Ak1(I)|I) mes(Ak2(I)|I)> d(N) 5M

!2

1 ψ(ek1)ψ(ek2),

where we put 5 instead of 4 in the denominator to deal with edge effects arising from mes(Ak(I)\I) =Oekψ(e1 k)

.

To prove the claim, it remains to bound mes (Ak1(I)∩ Ak2(I)|I) from above by

const

R1R2, whereRi :=ψ(eki).

A cylinder is a set of the form

[[a1, . . . , an]] ={α∈(0,1)\Q:ai(α) = ai (16i6n)}.

Equivalently, α ∈[[a1, . . . , an]] iff α has an infinite continued fraction expansion of the formα = [0;a1, . . . , an,∗,∗, . . .].

(10)

Our plan is to cover Aki(I) by unions of cylinders of total measure O(1/Ri), and then use the following well-known fact: There is a constant G > 1 s.t. for any (a1, . . . , an, b1, . . . , bm)∈Nn+m,

(3.8) G−1 6 mes[[a1, . . . , an;b1, . . . , bm]]

mes[[a1, . . . , an]] mes[[b1, . . . , bm]] 6G.

This is because the invariant measure ln 21 1+xdx of T : (0,1)→(0,1), T(x) = {1x} (the Gauss map) is a Gibbs–Markov measure, thanks to the bounded distortion of T, see [ADU93, Section 2].

To cover Aki(I) by cylinders, it is enough to cover Am,n,ki by cylinders for every (m, n) ∈ Ωki(I). Suppose αAm,n,ki. Then r := gcd(m, n) 6 M and (m, n) :=

1

r(m, n) satisfies

gcd(m, n) = 1, n[

|ki−ki|6lnM

[eki−1, eki], |nαm|< 1 nRi.

Assume ki is so large that Ri = ψ(eki) > 4. Then Lemma 3.3 gives ali+1 > R2i. ThusAki(I)⊂ Cki(I, Ri) where

Ck(I, R) := [

k∈[k−lnM,k]

α∈(0,1)\Q:∃` s.t.

q`(α)∈[ek−1, ek], a`+1(α)>R/2 p`(α)/q`(α)∈I

. This is a union of cylinders, becauseq`(α), p`(α), a`+1(α) are constant on cylinders of length`+ 1.

We claim that for somec(M) which only depends on M, for all ki large enough, (3.9) mes(Cki(I, Ri))6 c(M) mes(I)

Ri .

Every rational mn ∈(0,1) has two finite continued fraction expansions: [0;a1, . . . , a`] and [0;a1, . . . , a`−1,1] with a` >1. We write `=`(mn) andai =ai(mn). With this notation

Cki(I, Ri) = [

ki∈[ki−lnM,ki] n∈[eki−1,eki]

[

gcd(m,n)=1 m/n∈I

[

b>Ri/2

a1

m n

, . . . , a`(m

n)

m n

, b

a1

m n

, . . . , a`(m

n)

m n

−1,1, b

. We have [[a1, . . . , a`]] = (pq`+p`−1

`+q`−1,pq`

`) or (pq`

`,pq`+p`−1

`+q`−1), depending on the parity of`, see for instance [Khi63]. Since |p`q`−1p`−1q`|= 1 and q`+1 =a`+1q`+q`−1, we have mes([[a1, . . . , a`, b]]) = q 1

`+1(q`+1+q`) = (bq 1

`+q`−1)((b+1)q`+q`−1) 6 b(b+1)q1 2

`

, leading to mes(Cki(I, Ri))6 X

ki∈[ki−lnM,ki]

X

n∈[eki−1,ek

i]

X

gcd(m,n)=1 m/n∈I

X

b>Ri/2

2 n2b(b+ 1) 6 8 lnM

e2(ki−1−lnM)Ri

ekiM

X

n=1

#

m ∈N: m

nI,gcd(m, n) = 1

6 c(M)

Ri mes(I) where c(M) only depends onM. The last step uses Proposition 3.9.

(11)

Next we cover Ak1(I)∩ Ak2(I) by cylinders. Suppose without loss of generality that k2 > k1. Arguing as before one sees that if

(3.10) k2 > k1+ lnM + 1,

then Ak1(I)∩ Ak2(I) can be covered by sets [[a1, . . . , a`, b, a1, . . . , al, b]] as follows:

The convergents pi/qi of (every) α in [[a1, . . . , a`, b, a1, . . . , al, b]], (16i6l+l+ 2), satisfy

(1) ql ∈[ek1−1, ek1], k1 ∈[k1−lnM, k1], pl/qlI, b>R1/2;

(2) ql+l+1 ∈[ek2−1, ek2], k2 ∈[k2−lnM, k2], pl/qlI, b>R2/2 (3) k2 > k1 (this is where (3.10) is used).

We claim that

(3.11) [[a1, . . . , a`, b]]⊂ Ck1(I),

(3.12) b 6ek2−k1+1,

(3.13) [[a1, . . . , a`, b]][

|r|63

Ck2−k1+r−lnb([0,1], R2).

(3.11) follows from (1). Next,ek2>ql+1>bql>bek1−1 proving (3.12). To prove (3.13), letpi/qi, 16i6`+ 2, be the principal convergents of (every) α∈[[b, a1, . . . , a`, b]].

By Proposition 3.10, ql+1+l = ql−1pl+1 +qlql+1, whence qlql+1 6 ql+1+l 6 2qlql+1. Sinceql ∈[ek1−1, ek1] and ql+l+1 ∈[ek2−1, ek2],

(3.14) ek2−k1−2 6 ql+1+l

2ql 6ql+1 6 ql+1+l

ql 6ek2−k1+1.

Next, letpei/qei(16i6l) denote the principal convergents of (every)αe∈[[a1, . . . , a`, b]].

Then pql+1

l+1

= 1/(b + epl

eql), so ql+1 = bqel +pel, whence bqel 6 ql+1 6 (b + 1)qel. Thus qel ∈ [(b+ 1)−1ql+1, b−1ql+1]. It follows that the l-th principal convergent of every αe ∈[[a1, . . . , a`, b]] satisfies

(3.15) qel ∈[ek2−k1−3−lnb, ek2−k1+1−lnb].

It is now easy to see (3.13).

By (3.13),Ak1(I)∩ Ak2(I)∩IS|r|63U[a,b]⊂Ck

1(I)

U

[a0,b0]⊂Ck

2−k1+r−lnb([0,1])[[a, b, a0, b]].

Now arguing as in the proof of (3.9) and using (3.8) we obtain (3.16) mes(Ak1(I)∩ Ak2(I)∩I)

6 X

ki∈[ki−lnM,ki] i=1,2;|r|63

X

n∈[ek1−1

,ek1]

X

gcd(m,n)=1 m/n∈I

[exp(k2−k1+1)]

X

b=[R1/2]

2Gmes(Ck2−k1+r−lnb([0,1],R2)) n2b(b+1)

6 const mes(I) R1R2 .

(3.16) uses the estimate mes(Ck2−k1+r−lnb([0,1], R2)) =O(1/R2) which is also valid when k2k1 +r−lnb is small, provided we choose M large enough so that the asymptotic in Proposition 3.9 holds for all N > M with I = [0,1]. See the proof of (3.9).

(12)

Combining (3.16) with (3.7), we find that under (3.10) Aki(I) are D-quasi-indep-

endent for sufficiently largeD, proving Claim 3.

Claims 2 and 3 allow us to apply Sullivan’s Borel–Cantelli Lemma (Proposition 3.7).

We obtain δ = δ(M) s.t. for every interval I ⊂ [ε,1−ε], mes(A ∩I) > δmes(I).

This means that [ε,1−ε]\ Ahas no Lebesgue density points, and therefore must have measure zero. So A has full measure in [ε,1−ε]. Since ε is arbitrary, A has

full measure.

4. Proof of Theorem 1.2

As explained in Section 2, it is enough to prove Theorem 1.2 forf(x) of the form

Pd

m=1bmh(x+βm) 6≡ 0 with h(x) = {x} − 12. Without loss of generality, βi are different andbi 6= 0. Notice that

f(x) =

d

X

m=1

bm

X

j=1

sin(2πj(x+βm))

πj .

Thereforekfk22 = 12

P

n 1

n2D(β1n, . . . , βdn), where D:Td→R is D(γ1, . . . , γd) :=

Z 1 0

" d X

m=1

bmsin(2π(y+γm))

#2

dy.

(4.1)

Since f 6≡ 0, D(β1n, . . . , βdn) > 0 for some n. Let T denote the closure in Td of O:={(β1n, . . . , βdn) modZ:n ∈Z}.This is a minimal set for the translation by (β1, . . . , βd) onTd, so a standard compactness argument shows that for everyε0 >0,

the set

(4.2) N :={n∈N: D(β1n, . . . , βdn)> ε0}

is syndetic: its gaps are bounded. Thus N has positive lower density.

By Theorem 3.1, if M is sufficiently large then the set A := A(N, M) has full measure inT. Let

Sn(α, x) :=

n−1

X

k=0

f(x+kα).

The proof of Theorem 1.2 for f(x) above consists of two parts:

Theorem 4.1. — Supposeα∈ A, then for a.e.x∈[0,1), there exist Ak(x)∈R and Bk(x), Nk(x)→ ∞ such that

Sn(α, x)−Ak(x) Bk(x)

−−−→dist

k→∞ U[0,1], as n∼U(0, . . . , Nk(x)).

Theorem 4.2. — Supposeα∈ A, then for a.e.x∈[0,1), there are noAN(x)∈R and BN(x)→ ∞ such that

Sn(α, x)−AN(x) BN(x)

−−−→dist

N→∞ U[0,1], as n∼U(0, . . . , N).

Références

Documents relatifs

For many chaotic systems it has been proved that the central limit theorem holds for H¨older continuous functions.. If the action of K is H¨older continuous, then the central

In his paper [T2] Terras obtained a central limit theorem for rotation- invariant random variables on the space of n x n real positive definite (symmetric) matrices in the case n =

In this paper we prove a central limit theorem of Lindeberg-Feller type on the space Pn- It generalizes a theorem obtained by Faraut ([!]) for n = 2. To state and prove this theorem

In study of the central limit theorem for dependent random variables, the case of martingale difference sequences has played an important role, cf.. Hall and

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

The main result by Lasserre [11] (restated below) we borrow is Theorem 3.1 that state that the Handelman representation cannot give exact solution to the minimization problem when

The proof relies on the correct aymptotics of the variance of the eigenvalue counting function of GUE matrices due to Gustavsson, and its extension to large families of Wigner

Debbasch, Malik and Rivet introduced in [DMR] a relativistic diffusion in Minkowski space, they called Relativistic Ornstein-Uhlenbeck Process (ROUP), to describe the motion of a