• Aucun résultat trouvé

Polynomial decay of correlations for a class of smooth flows on the two torus

N/A
N/A
Protected

Academic year: 2022

Partager "Polynomial decay of correlations for a class of smooth flows on the two torus"

Copied!
17
0
0

Texte intégral

(1)

POLYNOMIAL DECAY OF CORRELATIONS FOR A CLASS OF SMOOTH FLOWS ON THE TWO TORUS

by Bassam Fayad

Abstract. — Koˇcergin introduced in 1975 a class ofsmooth flows on the two torus that are mixing. When these flows have one fixed point, they can be viewed as special flows over an irrational rotation ofthe circle, with a ceiling function having a power- like singularity. Under a Diophantine condition on the rotation’s angle, we prove that the special flows actually have at−η-speed ofmixing, for someη >0.

esum´e (D´ecroissance polynomiale des corr´elations pour une classe de flots lisses surT2)

Koˇcergin a introduit en 1975 une classe de flotsCsur le tore `a deux dimensions qui sont m´elangeants. Quand ces flots ont un seul point fixe, ils correspondent `a des flots sp´eciaux au-dessus d’une rotation irrationnelle du cercle, dont la fonction de suspension pr´esente une singularit´e en puissance fractionnaire. Sous une condition diophantienne sur l’angle de la rotation, on prouve que ces flots sp´eciaux ont une vitesse de m´elange ent−η, pour un certainη >0.

1. Introduction

1.1. Koˇcergin gave in [5], examples of C measure-preserving flows on the two torus that are mixing. He starts by proving that special flows over irra- tional rotations of the circle (or over interval exchange transformations) with

Texte re¸cu le 16 mai 2000, r´evis´e le 2 novembre 2000, accept´e le 5 janvier 2001

Bassam Fayad, Centre de Math´ematiques, UMR 7640 du CNRS, ´Ecole polytechnique, 91128 Palaiseau Cedex (France) E-mail :fayadb@math.polytechnique.fr

2000 Mathematics Subject Classification. — 37E35.

Key words and phrases. — Flows on the torus, special flows, speed ofmixing, correlations.

Part ofthis work was carried on during my stay at CAMS (Beirut, Lebanon) in the summer of1999. I wish to thank the members ofthis institution for their warm hospitality.

BULLETIN DE LA SOCI ´ET ´E MATH ´EMATIQUE DE FRANCE 0037-9484/2001/487/$ 5.00 c Soci´et´e Math´ematique de France

(2)

a ceiling function having a power-like singularity are mixing. It is possible to identify these special flows with smooth flows on the two torus, having one fixed point or more. In the same article, Koˇcergin describes in some examples how to realize this identification by smoothly gluing to an irrational flow Rt(1,α) a small neighborhood of the fixed point of an adequatly chosen Hamiltonian flow in the plane. We will prove that under a Diophantine condition on the rota- tion number, the special flow has a polynomial decay of correlations between rectangles. We will see later why the arithmetical condition is required.

1.2. First we give the definition of a special flow over an irrational rotation.

Given a strictly positive function ϕ L1(T1), the special flow constructed overRα and under the functionϕis the quotient flow of the action

T1×R−→T1×R, (x, s)−→(x, s+t),

by the relation

x, s+ϕ(x)

Rα(x), s . This flow acts on the space

MRα =T1×R/∼,

is uniquely ergodic and preserves the normalized Lebesgue measure onMRα, i.e. the product of the Haar measure on the basisT1with the Lebesgue measure on the fibers divided by the constant

T1ϕ(x) dx. In the sequel, we will simply denote byM the spaceMRα, and byµthe invariant measure described above.

We callrectangles inM the sets B=

t0+ t=t0

Tt(I),

when the union is disjoint and Iis an interval ofT1,t0RandR+. It is immediate that the collection of rectangles generates the σ-algebra of Borel sets on M. There is of course a slight abuse in calling rectangles the latter sets, because under the action of the flow, when t0 or are large, they get distorted and do not have rectangular shapes anymore. Nevertheless, whent0= 0 andinfT1ϕ, we have real rectangles that we callrectangles on the basis.

1.3. Description of the flow under consideration. — In what follows we will consider the special flow{Tt}constructed over an irrational rotation Rα, and under a ceiling functionϕ.

We assume the following hypothesis on ϕ:

o

(3)

ϕ∈C3(T/{0}), and infT1ϕ=c >0;

there exists 0< γ <1 such that lim

x0+

ϕ(x)

xγ = 1, lim

x0

ϕ(x) (−x)γ = 1, lim

x0+

ϕ(x)

−γxγ1 = 1, lim

x0

ϕ(x)

γ(−x)γ1 = 1, lim

x0+

ϕ(x)

γ(γ+ 1)xγ2 = 1, lim

x0

ϕ(x)

γ(γ+ 1)(−x)γ2 = 1.

For commodity, we will also suppose that

T1ϕ(x) dx= 1.

It follows from Koˇcergin’s result that these special flows, for anyαand any γ∈]0,1[, are mixing. But to force estimates on the decay of correlations, our techniques require that the exponent should be at least less than 12, and we will assumeγ≤25.

Remark. — The exponent obtained in one of the smooth examples given by Koˇcergin isγ= 13 <25.

As for the rotation number α, we require that the sequence {qn}n∈N of denominators of its convergents satisfies for some positive constantCα

(CD-) qn+1≤Cαqn1+,

where is small compared toγ,=1001 γ being enough for our purpose.

Finally, we introduce the number η:= 1

50γ.

In the sequel, we will often use the fact that 2≤η γ.

1.4. Statement. — Under the above assumptions onαandϕ, we will show the following:

Theorem 1.1. — For any two rectangles A and B, for any η0 < η, we have for t large enough

µ(A∩TtB)−µ(A)µ(B)≤ 1 tη0· (1)

In the statement, the rectangles and the measure µare as defined in (1.2).

(4)

1.5. Remarks. — For a fixed >0, the set of rotation numbers satisfying a Diophantine condition (CD-) is of total Lebesgue measure.

To simplify the presentation we considered only one power-like and symmet- rical singularity

ϕ(x)∼

0 ϕ(−x)∼

0|x|γ.

From the proof it will appear clearly that the same result holds when there are finitely many singularities, and not all of them being necessarily power-like (some of them could be logarithmic for example), under the condition that the strongest one should be power-like (with exponent γ 25). Furthermore, our assumption of symmetry is not necessary, to the contrary, symmetry plays in general against mixing. When the singularity is logarithmic for example, the symmetry impedes mixing as proved by Koˇcergin in [4]; while Khanin and Sinai proved mixing in the case of asymmetrical logarithmic singularities [3].

Our estimates are far from optimal and η = 50γ is certainly not the best polynomial rate of decay one can obtain. A faster speed of mixing thant12 would be very interesting because it would imply a Lebesgue spectrum for the flow. But since it appeared very hard to obtain faster decay than t12 by the techniques involved in this paper we wrote the proof with the aboveη=50γ.

Correlations between functions. — Through the proof of the theorem, it will appear that (1) is valid whent∈[qn, qn+1],nsufficiently large, for any pair of squares A and B with side of length equal to qnη. Hence we could establish for any couple of complex functions of class C1 on T2, the same decay of correlations obtained for rectangles.

1.6. Plan of the proof. — The property underlying mixing for a special flow over a rigid transformation is the uniform stretchof the Birkhoff sums of the ceiling function,

ϕm(x) :=ϕ(x) +ϕ Rα(x)

+· · ·+ϕ

Rmα1(x) .

When the Birkhoff sums have large derivatives, the image of a small intervalJ T1 by the flow is stretched with time in the vertical direction along the fibers, and asttends to infinity the interval actually breaks down into a lot of almost vertical curves whose projections on the circle follow the trajectory of Rα. By unique ergodicity of the rotation these projections become more uniformly distributed on the circle as their number increases, and so will be Tt(J) in the whole space (see [6], [5], [3], [2] and [1]).

For each t, we want to cover the circle excepted a small set with intervals being stretched as described above (here, we want the exceptional set to have measure less than tη). The first intervals to be over ruled are those that come too close to the singularity before timetand eventually get trapped in its neighborhood (Lemma 2.3). Other intervals must be automatically discarded,

o

(5)

those where there is no stretch at all, i.e. where the first derivative of ϕm is small (the singularity being symmetric this likely occurs): a lower bound on the second derivatives due to the convex average behavior ofϕwill allow us to estimate the size of such bad intervals (Lemma 2.5 and Step 2 in Section 3.1).

For the remaining part of the circle, we seek a good control on the stretch ofϕm, whenmis comparable witht, and here the Diophantine condition onα is required to insure the uniformity of stretch (Lemmas 2.4 and 2.5). Still, uniform stretch (Properties (P1)–(P2) in (3.1)) of an intervalJ is not enough by itself to estimate the asymptotic repartition of Tt(J) in the space as t goes to infinity. We need in addition to make sure that the pieces of Tt(J) (those almost vertical curves) do not enter in a too small neighborhood of the singularity, otherwise a lot of measure can be lost there (See (2.4)).

A “good” partition is finally constructed for each time t (Proposition 3.1) and Lemmas 3.3 and 3.2 give a precise description ofTt(J), for an intervalJ in this partition.

To conclude, we need a good estimate on the asymptotic distribution of the trajectories of the rotation on the circle that we obtain using again the Diophantine condition onα.

2. Preliminary estimates and lemmas

2.1. A Fubini Lemma. — We begin by a Fubini lemma that reduces our problem to studying the image under the special flow of intervals on its basisT1. Given ν > 0 and a finite partial partition P = {I0, ..., Im} of T1, we say that P is ν-fineif, for any interval I on the circle, there exists a collection of atoms fromP such that the symmetrical difference between their union andI has Lebesgue measure less thanν.

We recall that a rectangle on the basis is a subsetB =

0tTt(I),where I is an interval ofT1 and R+ is the height ofB. We also recall that η is the fixed number 501γ (2≤η γ≤25).

Lemma 2.1. — If there exist partial partitions ofT1,Pt, that aretη-fine and such that for any rectangleB on the basis with height less than c= infT1ϕ, we have when t is large enough

λ(Ji(t)∩TtB)−λ(Ji(t))µ(B)≤tηλ(Ji(t)), (2)

for every Ji(t)∈ Pt, then Theorem1.1is true.

In the statement of the Lemma,λis the Haar measure on the circle and µ is the normalized measure onM invariant by the special flow.

(6)

Proof. — Take two rectangles inM,

A=

t1st1+1

Ts(I) and B=

t2st2+2

Ts(I)

(1 and2 are the heights ofA andB). Since the rectangleB can always be decomposed into a finite disjoint union of rectangles with height less thanc, it is enough to prove Theorem 1.1 when2=δ < c. On the other hand one has

µ

A∩TtB

=µ

A∩T(tt2)Tt2B ,

hence we can also assume forBthatt2= 0. From the hypothesis of the Lemma, whent is large enough, there exist atomsJ1(t), ..., Jl(t) fromPt such that

λ

I

i=0

Ji(t) ≤tη and for each one of which (2) holds. Hence,

λ(I∩TtB)−λ(I)µ(B)≤3tη. Since A =

t1st1+1Ts(I), we apply the Fubini lemma and obtain the es- timation of the Theorem 1 when t is large enough (we took η0 < η to avoid having constants in the statement of the theorem). Of course the same conclu- sion holds if we establish (2) with some constant in front oftη.

2.2. Avoiding the singularity at time t. — We recall that ϕm(x) =

m1

k=0

ϕ Rαk(x)

.

For anyx∈T1(a point on the basis), and any positive timet, there is a unique integermsuch that

0≤t−ϕm(x)< ϕ Rmα(x)

.

We denote this integer byN(x, t). Since infT1ϕ=c >0, we have:

Lemma 2.2. — For everyx∈T1 and any time t, we have N(x, t) t

(3)

Proof. — One hast≥ϕN(x,t)(x)≥N(x, t)cby definition ofN(x, t).

Lett∈[qn, qn+1]. From the lemma above, we know we will avoid going too close to the singularity at timet if we consider only the set

Fn =T1

2[qn+1/c]

k=0

Rαk 1

q1+ηn+1 , 1

q1+ηn+1

.

(The notation [r] stands for the integer part of r R.) For this set we can state the following, when nis large enough

o

(7)

Lemma 2.3. — For any x, y∈Fn, for anyt∈[qn, qn+1], we have N(x, t)−N(y, t)≤qn+1γ+2η,

(4)

t

2 ≤N(x, t) t (5)

Proof. — By definition of the numberN(x, t), we have t−ϕN(x,t)(x)≤ϕ

RN(x,t)α (x) .

Sincex∈Fn, (3) implies thatRαN(x,t)(x)T1[−qn+11η, qn+11η], hence t−ϕN(x,t)(x)2qn+1(1+η)γ

(6)

because ϕ(|x|)0|x|γ andϕ∈C3(T1/0).

Now, on one hand t qn qn+11/(1+), and the other γ 25, 1 andη 1, hence (6) leads to

ϕN(x,t)(x) 9 10t.

(7)

We want now a bound on m(x)−m|, when x∈Fn andm≤2qn+1/c. First, we writems expansion:

m=

m

s=0

msqs, 0≤ms≤as+1, (8)

where as+1 denote the partial quotients of the continued fraction expansion ofα. Sinceαis-Diophantine,as+1≤qs, by definition. Furthermore, because m 2qn+1/c, we have m ≤n+ 1 and supms qn in the expression of m.

Let ˜ϕ be the function equal to ϕ outside [−qn+11η, qn+11η], constant on the latter interval and of integral one on the torus. Applying the Denjoy-Koksma inequality to ˜ϕwe have, for anys∈N,

ϕ˜qs(x)−qsVar ˜ϕ,

but the variations of ˜ϕon the circle are of the order ofq(1+η)γn+1 ,hence ϕ˜qs(x)−qs≤Cq(1+η)γn+1

whereC is a constant depending only onϕ. Usingm’s expansion, we have for any xon the circle

ϕ˜m(x)−m≤C(n+ 2)qnqn+1(1+η)γ, so when nis large enough, we have

ϕ˜m(x)−m≤q(1+η)γ+2n+1 (9)

whennis large enough.

(8)

On the other hand, form≤2qn+1/cwe haveRmα(x)T1[−qn+11η, qn+11η] for anyx∈Fn, so ˜ϕm(x) =ϕm(x). The above equation then becomes for any x∈Fn andm≤2qn+1/c

ϕm(x)−m≤qn+1(1+η)γ+2=o(qn+1γ+2η).

(10)

Finally, (4) follows from (6) and (10), while (5) follows from (7) and (10).

2.3. Stretch estimates. — To obtain sharp estimates for the stretch re- quires a good control on the first and second derivatives of the Birkhoff sums of ϕ. From Lemma 2.3 we know that when t [qn, qn+1], we have to con- sider those sums form∈[qn/2, qn+1/c]. A lower bound on the first derivatives will guarantee good stretch and an upper bound on the second will insure its uniformity. Here again, we will exploit the Diophantine condition onα. Never- theless, because the singularity is symmetrical, we will not be able to minorate directly the first derivatives of ϕ. It is a fact in this case that at some points of the circle the Birkhoff sums are not stretching at all. The lower bound on the second derivatives that comes from the convex average behavior of ϕwill allow us to overcome this difficulty.

We will denote the first and second derivatives ofϕmbyϕmandϕm. Whenn is sufficiently large, we have

Lemma 2.4. — For any m≤2qn+1/c andx∈Fn, ϕm(x)≤qn+11+γ+2η. (11)

We also have, for sufficiently largen,

Lemma 2.5. — For any m∈[qn/2,2qn+1/c] andx∈Fn, qn+12+γ ≤ϕm(x)≤qn+12+γ+3η. (12)

Proof of Lemma 2.4. — Writeϕ=ϕ+ϕ,where ϕ=χ[1/qn+1,1/qn+1]ϕ

L being the characteristic function of a set L). Reasoning exactly as in the proof of (9), the Diophantine condition onαand the Denjoy-Koksma inequality imply form≤2qn+1/cand for any x∈T1

ϕm(x)−m

T1

ϕ(y) dy ≤q1+γ+2n+1 . But

T1ϕ(y) dy =O(qn+1γ ), hencem

T1ϕ(y) dy =O(q1+γn+1) and we have for sufficiently largen

ϕm(x)2q1+γ+2n+1 . (13)

On the other hand under m iterations, m≤2qn+1/c, any pointx enters less than 8[1/c] times in the interval In := [1/qn+1,1/qn+1] ( where [r] denotes

o

(9)

the integer part of r R. We used the fact that successive [14qn+1] iterates ofIn byRα are disjoint). Meanwhile,x∈Fn andRαm(x) still remains outside [−qn+11η, qn+11η] form≤2qn+1/c, hence

ϕm(x) 9

cγ q(1+η)(1+γ)n+1 , (14)

because (x)| ∼γ|x|γ1 in the neighborhood of 0.

Lemma 2.4 then follows from (13) and (14) whennis sufficiently large since 2η > η( 1 +γ)>2.

Proof of Lemma 2.5. — The right hand side in (12) is obtained in exactly the same way as (11). For the left hand side consider the decomposition ϕ = ϕ+ϕ, with this time ϕ = χKϕ, where K is a fixed interval containing the singularity and on which the second derivative is always positive. It is possible to do so since we assumedϕ(|x|)0γ(γ+ 1)|x|2γ. Clearly, for anym∈N andx∈T1,

ϕm(x)≤Cm, (15)

where Chere is the supremum of the second derivative of ϕoutsideK.

On the other hand inmiterations,m≥ 12qn≥qn1, the orbit of any pointx enters at least once in the interval [1/qn1,1/qn1], since the union of the firstqn1inverse iterates of this interval cover the circle. From the hypothesis on the singularity of ϕand the Diophantine condition on αwe conclude that forx∈T1,m∈[12qn, qn+1]

ϕm(x)≥qn2+γ1≥qn+1(2+γ)/(1+)2.

Since (1 +)2 is of the order of 1−η, and ( 2 +γ)(1−η)>2 +γ−3η, the last equation and (15) yield the left hand side in (12).

2.4. Central lemma for the asymptotic distribution of the pieces ofTt(J). — In the introduction we explained how an intervalIbreaks down under stretch into many almost vertical curves whose projection on the basis follows the trajectory of the rotation.

From Lemma 2.3, we know more or less where do these parts of Tt(J) end up at timet. To have sharp estimates on their distribution in the whole space, we want to make sure that the parts stay relatively far from the singularity. We will define to this end, for each timet∈[qn, qn+1], a subset ofFn(see (18)) and a partition of this subset into intervals (see (3.1)) that after they get stretched will have a good asymptotic distribution in the space.

The following lemma, is essential in establishing the description of Tt(J) that will given in Lemma 3.2 and Lemma 3.3.

(10)

Letβ be a positive number such that β > γ+ 4η, (16)

βγ < γ−18η, (17)

We can resume the conditions (16) and (17) by asking that γ+ 4η < β <118η

γ,

which is easy to realize for someβ sinceη γ, andγ≤ 25 is also far from 1.

2.4.1. Denote byγ the exponentγ+ 2η that appeared in (4).

2.4.2. Define the set Tn,β=T1

2[qn+1γ ]

k=2[qγn+1]

Rαk

1 qn+1β, 1

qn+1β

. (18)

Remark 2.6. — With our choice ofβ, we clearly haveλ(Tn,β)18qn+1. Lemma 2.7. — For any x∈Tn,β and forD∈Nsuch that

qγn+115η ≤D≤2qγn+1 , we have for sufficiently large n

ϕD(x)−D≤qn+1η D, (19)

ϕD(x)≤q(γ+1)β+ηn+1 . (20)

Proof. — Reasoning the same way as in the previous lemmas we obtain, for D≤qγn+1 andx∈Tn,β, equation (20) and

ϕD(x)−D≤qn+1γβ+2. (21)

With our choice of β (cf. (17)), we have γβ+ 2≤γβ+η < γ−16η. Hence, the lower bound onD and (21) imply (19).

3. Proof of the theorem on correlations.

3.1. Construction of the good partition at time t. — We fix nowt in [qn, qn+1] and pick an arbitraryx0∈Fn.Define

Ft,β=Fn

N(x0,t)+2[qn+1γ ]

k=N(x0,t)2[qn+1γ ]

Rαk

1 qn+1β, 1

qn+1β

.

Proposition 3.1. — It is possible to construct a partial partition Pt of Ft,β consisting of intervals, with the following properties:

o

(11)

(P0) the partial partitionPt istη-fine on the circle;

(P1) the sizes of the intervals in Pt vary betweenqn+11 and2qn+11; (P2) for any J ∈ Pt and anym∈[N(x0, t)−qn+1γ , N(x0, t) +qγn+1 ]

inf

xJ

ϕm(x)≥qn+11+γ.

Claim. — As a direct consequence of(P1),(P2) and the upper bound in(12) we will have for large n:

(P2)For any J∈ Pt and anym∈[N(x0, t)−qn+1γ , N(x0, t) +qγn+1 ], sup

xJ

ϕm(x)· |J| ≤2qn+1η inf

xJ

ϕm(x).

Proof of the claim. — Since N(x0, t)≤ qn+1/c, m satisfies the hypothesis of Lemma 2.5 and we have supxJm(x)||J| ≤q2+γ+3ηn+1 2qn+11.

Proof of Proposition 3.1. — We will do the construction in three steps.

First step. — We cover the circle with disjoint intervals of sizes varying between qn+11η and 2qn+11η. (We do not care about the ending points of the intervals because their total measure is always zero and we will indifferently talk about segments or intervals.) In this first step we just take away from the above intervals these that are not completely included in Ft,β, or equivalently, these that intersect either

2[qn+1/c]

k=0

Rαk

1/qn+11+η,1/q1+ηn+1 or

N(x0,t)+2[qn+1γ ]

k=N(x0,t)2[qn+1γ ]

Rαk[1/qn+1β,1/qn+1β].

The total measure of the discarded segments is bounded by 2

qn+1

c 2

qn+11+η + 4 qn+1

c 2

qn+11+η + 4[qγn+1 ] 2

qn+1β + 8[qn+1γ ] 2

qn+11+η =O(qn+1η ) due to our choice of β. The partial partition we obtained is hence qn+1η -fine (up to a constant).

Second step. — Consider any intervalLof the ones we kept in the first step.

It follows from Lemma 2.2 and Lemma 2.5 that for m0 = N(x0, t), ϕm0 is greater than qn+12+γ on L. So, after taking off eventually an interval of size 4qn+11there will remain fromLone or two intervals,L andLr, wherem0| is always greater than 2qn+11+γ. If one of those intervals,LorLrhas size less thanqn+11 we discard it (see Fig. 1).

(12)

L L Lr Lr

Figure 1. Different configurations forϕmoverL

The total measure we might loose by this operation onL is 5qn+11 which is at most a proportion 5qn+1η of the intervalL. Hence we remain with a partial partition that isqn+1η fine (up to a constant).

Moreover, we claim that the bound on ϕm0 leads to (P2). Indeed, for any m∈[m0−qγn+1 , m0+qn+1γ ],

ϕm(x) =ϕmm0(Rαm0x) +ϕm0(x);

but Lr and L are in Ft,β, so Rmα0(x) Tn,β, therefore (20), of Lemma 2.7, implies that

ϕmm

0(Rαm0x)≤q(γ+1)β+ηn+1

whenx∈Lr,x∈L. From the choice ofβ(cf. (17) it follows that (γ+1)β+η <

1 +γ−5η. Hence, ϕmm0(Rmα0x) ϕm0(x) and the claim follows.

Third step. — In this step we just divide each intervalLandLrwe get from the first two steps into intervals of size varying between qn+11 and 2qn+11. It is possible to do so since the sizes of L and Lr are larger than qn+11. This givesPt.

3.2. Propertiesof a good interval. — From now on J will designate an interval of Pt. Fix a rectangle B = δ

s=0TsV. In the light of Lemma 2.1, we will finish if we prove that, under the conditions (P1)–(P2)–(P2) on an intervalJ, the measure of the intersectionTt(J)

B satisfies (2). Lemmas 3.2 and 3.4, that we will state below, will be the ingredients of our proof. Let J ∈ Ptbe fixed

J = [x1, x2], m1=N(x1, t)m2=N(x2, t).

(22)

Sincex1, x2∈J ⊂Fn, Lemma 2.3 implies that m1, m2

N(x0, t)−qn+1γ , N(x0, t) +qn+1γ .

From (P2), it follows thatϕmis monotone onJ and we will supposeϕm1(x1) ϕm1(x2), the other case being similar.

o

(13)

Define form∈N, Jm=

x∈J;N(x, t) =m , (23)

=

x∈J; 0≤t−ϕm(x)< ϕ(Rmα(x)) Jm,δ =

x∈J; 0≤t−ϕm(x)≤δ . (24)

The set Jmis the part of J such that the projection ofTs(Jm) onT1, when s runs through [0, t], is translatedm times by Rα. AndJm,δ is the part of Jm that lies at time tin the band T1×[0, δ].

Finally, denote by ∆ϕm1|J the quantityϕm1(x1)−ϕm1(x2).

Form∈]m1, m2[, we have the following asymptotic estimate ofλ(Jm,δ) (λ being the Haar measure onT1)

Lemma 3.2. — For any msuch that m1< m < m2 it is true that λ(Jm,δ) δ|J|

∆ϕm1|J

≤qn+1η δ|J|

∆ϕm1|J

·

Formi=m1, m2 we have

λ(Jmi) δ|J|

∆ϕm1|J ≤qn+1η δ|J|

∆ϕm1|J· Form not in[m1, m2],Jm,δ is empty.

Using Lemma 2.7 we will prove the following estimate onm2−m1: Lemma 3.3. — We have

m2−m1

∆ϕm1|J 1≤qn+1η . (25)

Proof of Lemma 3.2. — First we will prove the following uniform estimate, for any m∈[m1, m2] and for any x∈J

1

ϕm(x) |J|

∆ϕm1|J

3qn+1η |J|

∆ϕm1|J

· (26)

By the mean value theorem, there existsx1∈J such that

∆ϕm1|Jm

1(x1)· |J|,

(14)

hence E:= 1

ϕm(x) |J|

∆ϕm1|J

= m(x)−ϕm

1(x1)| · |J|

m(x)|∆ϕm1|J

m(x)−ϕm(x1)| · |J|

m(x)|∆ϕm1|J +mm1(Rmα1x1)| · |J|

m(x)|∆ϕm1|J

supxJm(x)| · |J|2

m(x)|∆ϕm1|J +mm1(Rmα1x1)| · |J|

m(x)|∆ϕm1|J · Using (P2) we obtain

E 2qn+1η |J|

∆ϕm1|J +mm1(Rmα1x1)| · |J|

m(x)|∆ϕm1|J .

Exactly as in our proof of (P2) in Proposition 3.1, it follows from (20) that ϕmm1(Rαm1x1)≤qn+1η inf

xJm(x)|. This ends the proof of (26).

Now, form∈]m1, m2[ we have immediately from (26) thatϕmis decreasing like ϕm1. Also for suchm, we have

t−ϕm(x1)<0, (27)

t−ϕm(x2)> δ , (28)

the first inequality following directly from m1 := N(x1, t) (in the definition given in Section 2.2,N(x, t) is the larger integer msuch thatt−ϕm(x)0).

For the second one, we use our assumption δ < c (indeed, t−ϕm2(x2) 0 implies in this case (28) for any m≤m21).

In conclusion, for eachm∈]m1, m2[, by monotonicity ofϕm and the mean value theorem, there exists xm∈J such that

λ(Jm,δ) = −δ ϕm(xm)·

Using (26) we obtain the first point of Lemma 3.2. The second point follows from (26) in a similar fashion. As for the third one we have to do the following argumentation: by definition of m2 = N(x2, t) one has t−ϕm2+1(x2) < 0, but ϕm2+1 is decreasing on J, therefore, for any x J, t−ϕm2+1(x) < 0.

The same will clearly hold for any m m2+ 1, andJm,δ will be empty for all m > m2. When m < m1, we use in addition that ϕ is greater thanδ to obtaint−ϕm(x)> δ ,for anyx∈J, which impliesJm,δis empty for m < m1. The proof of Lemma 3.2 is accomplished.

Proof of Lemma 3.3. — SinceJ ⊂Ft,β⊂Fn, Lemma 2.3 implies m1−N(x0, t)≤qn+1γ

o

(15)

hence, from Lemma 2.4 and conditions (P1) and (P2) we obtain qγn+114η ∆ϕm1|J2qn+1γ.

(29)

NowJ ⊂Ft,β implies

Rmα1(J)⊂Tn,β. (30)

In particularϕ(Rαm1(x1))2qn+1γβ , and by the definition ofm1

0≤t−ϕm1(x1)≤ϕ

Rmα1(x1)

2qn+1γβ . Together with the similar equation ont−ϕm2(x2) this implies

ϕm2(x2)−ϕm1(x1)4qn+1γβ . (31)

On the other hand, we have

ϕm2(x2)−ϕm1(x1) =ϕm1(x2) +ϕm2m1(Rmα1x2)−ϕm1(x1),

=ϕm2m1(Rmα1x2)∆ϕm1|J. It follows from our choice ofβ (17), that

qn+1γβ qn+1qn+1γ14η.

Hence, the left hand side of (29) and (31) imply whennis large ϕm2m1(Rmα1x2)∆ϕm1|J≤qn+1∆ϕm1|J. (32)

From (30), we know that Rmα1(x2)∈Tn,β; to apply Lemma 2.7 and finish, we still need to prove thatqn+1γ15η≤m2−m12qn+1γ . From (29) and (32) we have

1

2qγn+114η ≤ϕm2m1(Rαm1x2)4qn+1γ,

the right hand side of which implies thatm2−m1≤c14qn+1γ qn+1γ . Next, becauseRαm1(x2)∈Tn,β, (21) is valid and implies the lower bound onm2−m1

from the lower bound onϕm2m1(Rmα1x2) and ∆ϕm1|J.

3.3. Asymptotic distribution on the circle of the rotation’s orbit Sinceαis-Diophantine, we have:

Lemma 3.4. — Let χV be the characteristic function of an interval V T1, then for N large enough, we have, for anyx∈T1

N1

i=0

χV

Riα(x)

−N|V|≤N2|V|, where|V|=λ(V)is the Lebesgue measure of V.

Références

Documents relatifs

Throughout this section, S indicates a minimal surface of general type over the field of complex numbers, and with a ample canonical sheaf.. We use K- semistability in the sense

Moreover, the correlations of H¨older functions (as defined in [14]) decay as indicated in (28): in the exponential case, this is proved in [14]. Young treats the case of 1/n γ ,

Furthermore such a maximal lVln must necessarily be irreducible. Otherwise it must contain a ring of three regions.. Thus in this remaining case we are also led

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

Tile Green functions G • are continuous plurisub- harmonic and they give the rate of escape to infinity.. Of course, this result holds for polynomial automorphisms

Ambiguous ideals in real quadratic fields and Diophantine equations Given the square-free integer D &gt; 1, the determination of the ambiguous ideal classes in

This conjecture, which is similar in spirit to the Hodge conjecture, is one of the central conjectures about algebraic independence and transcendental numbers, and is related to many

The second one follows formally from the fact that f ∗ is right adjoint to the left-exact functor f −1.. The first equality follows easily from the definition, the second by