• Aucun résultat trouvé

Using Apéry’s proof thatζ (3)is irrational, we obtain an upper bound for theϕ-exponent of irrationality ofζ (3), for a givenϕ

N/A
N/A
Protected

Academic year: 2022

Partager "Using Apéry’s proof thatζ (3)is irrational, we obtain an upper bound for theϕ-exponent of irrationality ofζ (3), for a givenϕ"

Copied!
15
0
0

Texte intégral

(1)

Indag. Mathem., N.S., 20 (2), 201–215 June, 2009

Restricted rational approximation and Apéry-type constructions

by Stéphane Fischler

Univ. Paris-Sud, Laboratoire de Mathématiques d’Orsay, Orsay Cedex, F-91405, France and CNRS, Orsay Cedex, F-91405, France

Communicated by Prof. M.S. Keane

ABSTRACT

Letξ be a real irrational number, andϕbe a function (satisfying some assumptions). In this text we study theϕ-exponent of irrationalityofξ, defined as the supremum of the set ofμfor which there are infinitely manyq1such thatqis a multiple ofϕ(q)and|ξpq|q−μfor somepZ. We obtain general results on this exponent (a lower bound, the Haussdorff dimension of the set where it is large, . . .) and connect it to sequences of small linear forms in 1 andξwith integer coefficients, with geometric behaviour and a divisibility property of the coefficients. Using Apéry’s proof thatζ (3)is irrational, we obtain an upper bound for theϕ-exponent of irrationality ofζ (3), for a givenϕ.

1. INTRODUCTION

Apéry has proved [2] (see also [9] for a survey) that for ξ =ζ (3), α=e3(1+

√2)4<1, andβ=e3(1+√

2)4>1, the following holds:

There exist two integer sequences(un)n1and(vn)n1such that un0,|unξvn|1/nαandu1/nnβ

(1.1)

and also, withδn=dn3wheredn=lcm(1,2, . . . , n): δndividesunfor anyn1.

(1.2)

E-mail: stephane.fischler@math.u-psud.fr (S. Fischler).

(2)

Sinceα <1, (1.1) implies the irrationality ofζ (3). It is well known that (1.1) impliesμ(ξ )1−logβ

logα, (1.3)

whereμ(ξ )is the exponent of irrationality ofξ, so thatμ(ζ (3))13.4178202. . . . It is proved in [11] (together with more precise results connected to [10]) that, conversely, forξ ∈R\Q,

ifμ(ξ ) <1−logβ

logα with0< α <1< β, then (1.1) holds.

(1.4)

In this paper, we generalize both implications (1.3) and (1.4), to take into account the divisibility property (1.2), under the assumption thatδn divides δn+1 for any n1 (which is the case in Apéry’s construction since δn =dn3). For instance, Apéry’s construction implies the following result (the analogue of which, where ζ (3)is replaced withlog(2), has been proved by Dubitskas [6] in a stronger version, see further):

Theorem 1. For anyε >0, there are only finitely many integersq1satisfying both

ζ (3)p q

< 1

q2+ε for somep∈Z and

dn3dividesq, withn=

logq log((1+√

2)4)

.

To state our results more precisely, let us denote byE the set of all functions ϕ:N→N(withN= {1,2,3, . . .}) such that:

• For anyq1,ϕ(q+1)is a multiple ofϕ(q).

• The limitγϕ:=limq→∞logϕ(q)

logq exists and satisfies0γϕ<1.

The following definition generalizes that of the usual exponent of irrationality μ(ξ )(which is obtained as a special case whenϕis the function1defined by1(q)= 1for anyq).

Definition 1. ForϕE andξ ∈R\Q, theϕ-exponent of irrationalityofξ is the supremum, denoted by μϕ(ξ ), of the set of real numbers μ for which there are infinitely manyq1such that

q is a multiple ofϕ(q) and ξp q

1

qμ for somep∈Z.

(3)

Of course, when this set isR, we haveμϕ(ξ )= +∞.

If we letϕ(q)=dn3where n= [log((1log+q2)4)], then ϕE and Theorem 1 means thatμϕ(ζ (3))2.

In the case oflog(2), Dubitskas’ result impliesμϕ(log(2))2whereϕ(q)=dn

withn= [log(3logq+22)]. On the other hand, Rivoal has proved [21] that log(2)− p

2ndn

1

(2ndn)1.948967 for anyp∈Zandnsufficiently large, so that only finitely many convergents in the continued fraction expansion oflog(2) have a denominator of the form2ndn. Maybe Rivoal’s methods (which apply also tolog(r)for other positive rational numbersr) can lead to upper bounds less than 2 forμϕ(log(r)), for suitableϕEandr∈Q,r >0.

The main result of this paper is the following generalization of (1.3) and (1.4), which allows one to deduce Theorem 1 from Apéry’s construction:

Theorem 2. Letξ∈R\Q,0< α <1,β >1, and(δn)n1be a sequence of positive integers such thatδndividesδn+1for anyn1, andδn1/ntends toδasn→ ∞.

Define a functionϕE by ϕ(q)=δn withn=

logq log(δ/α)

.

Then we have the following implications:

(i) If (1.1)and(1.2)hold thenμϕ(ξ )loglogβδloglogαα. (ii) Ifμϕ(ξ ) <loglogβδloglogαα then(1.1)and(1.2)hold.

We also prove various results (of independent interest) on the ϕ-exponent of irrationalityμϕ(ξ ), namely:

• For anyξ ∈R\Q, we haveμϕ(ξ )2−γϕ, and equality holds for almost any ξ with respect to Lebesgue measure.

• For μ >2−γϕ, the set of real numbers ξ ∈R\Q such thatμϕ(ξ )μ has Hausdorff dimension 2μγϕ.

• For anyξ ∈R\Q, we haveμϕ(ξ )= +∞if, and only if,μ(ξ )= +∞(that is, if and only ifξ is a Liouville number).

In the case ofζ (3), we obtain the following result as a consequence of Theorem 1 and this Hausdorff dimension computation.

Corollary 1. For anyq1, letϕ(q)=dn3wheren= [log((1log+q2)4

)]. LetSdenote the set of allξ ∈R\Qsuch thatμϕ(ξ ) >2. Thenζ (3) /S andS has Hausdorff dimension0.5745. . . .

(4)

As far as we know, this is the largest known Hausdorff dimension for a subset of R, defined by Diophantine conditions, which does not contain ζ (3). It is worthwile noticing that variants of Apéry’s construction (due to Hata, and Rhin and Viola, . . .) give better bounds for the (usual) irrationality exponentμ(ζ (3)), but do not seem to allow any improvement on Corollary 1 (see Section 4).

In this text, we consider asymptotic estimates like (1.1) since these can be easily used to work with exponents likeμϕ(ξ ). However, in the case ofζ (3)(and alsoζ (2) andlog 2), more precise estimates are known. They enable us to prove the following result, which refines Theorem 1 and is analogous to Dubitskas’ result [6] forlog 2. Theorem 3. There existsc >0such that for anyq1and anyp∈Zwe have

ζ (3)p q

c(logq)3 q2

provided that

dn3dividesq, withn=

logq log((1+√

2)4)

.

Corollary 2. Only finitely many convergents p/q in the continued fraction expansion ofζ (3)are such thatdn3dividesq, withn= [log((1log+q2)4)].

The structure of this text is as follows. In Section 2, we prove the general results stated above (and some others) about the ϕ-exponent of irrationality μϕ(ξ ). In Section 3, we give a proof of Theorem 2. Finally, in Section 4 we apply our results to particular numbersξ, especiallyζ (3),ζ (2), andlog 2, and in Section 5 we prove Theorem 3 and Corollary 2.

2. GENERAL PROPERTIES

The definition ofμϕ(ξ )makes sense because for anyϕEthere are infinitely many integersq such thatq is a multiple ofϕ(q). More precise statements are given in the proofs of Lemmas 1 and 2 and also in the statement of Proposition 1.

Let us start with examples of functions inE. Letb1, . . . , br 2be pairwise dis- tinct integers, andε1, . . . , εr>0be such thatri=1εilogbi <1. Then the function defined byϕ(q)=r

i=1bi[εilogq]belongs toE, and satisfiesγϕ=r

i=1εilogbi. 2.1. A lower bound

The following lemma generalizes the lower boundμ(ξ )2 which holds for any ξ∈R\Q. The proof uses the same tool as Dirichlet’s proof, namely the pigeonhole principle.

Lemma 1. For anyϕEand anyξ∈R\Q, we haveμϕ(ξ )2−γϕ.

(5)

Proof. Letε >0, andQbe sufficiently large in terms ofε. Consider, for0n [Q/ϕ(Q)], the fractional part of nϕ(Q)ξ. Since ξ /∈Q, this gives [Q/ϕ(Q)] +1 pairwise distinct points in[0,1]. Thanks to the pigeonhole principle, two of them lie within a distance less than or equal to [Q/ϕ(Q)]1. The difference of the corresponding integers n yields an integer m, with 1mQ/ϕ(Q), such that

|mϕ(Q)ξp|[Q/ϕ(Q)]1 for somep∈Z. Now letq=mϕ(Q). ThenqQ so thatϕ(q)dividesϕ(Q), and alsoϕ(Q)dividesq by definition ofq. Finallyϕ(q) dividesq, and sinceQis sufficiently large in terms ofεwe have:

ξp q

1 q

Q ϕ(Q)

1

1

qQ1γϕε 1 q2γϕε. This concludes the proof of Lemma 1. 2

2.2. Comparisons betweenμϕ(ξ )for variousϕ

In this section, we show howμϕ(ξ )and μϕ(ξ ) are connected forϕ, ϕE. It is specially interesting whenϕ=1since in this caseμϕ(ξ )is the classical exponent of irrationalityμ(ξ ).

Let us start with the following remark.

Remark 1. Letξ ∈R\QandϕE. Then μ1ϕ(ξ )γ

ϕ is the supremum of the set ofμ for which there are infinitely manyqsuch that

q is a multiple ofϕ(q) and ξp q

1

(q/ϕ(q))μ for somep∈Z.

The proof of this fact is easy, since for anyε >0 and any q sufficiently large in terms ofεwe haveq1γϕεq/ϕ(q)q1γϕ+ε.

This remark is crucial in the proof (given further) of the following lemma.

Lemma 2. Letξ ∈R\Q, andϕ, ϕE be such that ϕ(q)dividesϕ(q)for any q1. Thenμϕ(ξ )is finite if, and only if,μϕ(ξ )is finite, and in this case we have:

1−γϕ

1−γϕμϕ(ξ )μϕ(ξ )μϕ(ξ ).

Lettingϕ=1, we obtain the following corollary.

Corollary 3. Letξ∈R\QandϕE. Thenμϕ(ξ )is infinite if, and only if,μ(ξ ) is infinite (that is, if and only ifξ is a Liouville number). Otherwise we have

(1γϕ)μ(ξ )μϕ(ξ )μ(ξ ).

Remark 2. LetϕEbe such thatγϕ=0. Then for anyξwe haveμϕ(ξ )=μ(ξ )so thatμϕis nothing but the usual exponent of irrationality. More generally, ifϕ, ϕE

(6)

are such thatϕ(q)dividesϕ(q)for anyq1, andγϕ=γϕ, then Lemma 2 shows thatμϕ(ξ )=μϕ(ξ )for anyξ.

Proof of Lemma 2. Ifq is a multiple ofϕ(q)thenq is a multiple ofϕ(q), so the second inequality is trivial. Let us prove the first one, that is μ1ϕγ(ξ )

ϕ μ1ϕ(ξ )γϕ. This follows immediately from Remark 1 and the following fact.If ϕ, ϕE are such thatϕ(q)dividesϕ(q)for anyq1, and ifq1is such thatϕ(q)dividesq, then there exists an integer multipleq ofqsuch thatϕ(q)dividesq andq/ϕ(q)= q(q). To prove this fact, we let q be the least integer such that q q and

q

ϕ(q)ϕ(q)q. Such an integer exists sinceϕE. Ifq=qthenϕ(q(q)so thatϕ(q)=ϕ(q), and the conclusion holds withq=q. Otherwise, we haveq > q and, by minimality,qϕq(q)ϕ(q)ϕq(q)ϕ(q−1) > q−1so thatq=ϕq(q)ϕ(q). Nowq< qimplies thatϕ(q)dividesϕ(q); sinceϕ(q)dividesϕ(q), we obtain thatq=qϕϕ(q)(q) is a multiple ofq. This concludes the proof of the fact, and that of Lemma 2. 2

2.3. A special set of functions

In this subsection, we focus on specific functionsϕE, of major importance to us since they are the ones involved in Theorem 2. Actually, since our interest lies only on the exponents of irrationalityμϕ(ξ ), Lemma 4 below shows that we do not lose anything by considering only these functions (and even only a part of them). Let us start by the following lemma, in which these functionsϕ are defined. We omit the proof, since it is very easy.

Lemma 3. Let n)n1 be a sequence of positive integers such thatδn divides δn+1for anyn1, andδn1/ntends toδasn→ ∞. Letα∈Rbe such that0< α < δ. Define a functionϕby

ϕ(q)=δn withn=

logq log(δ/α)

.

Then we haveϕEandγϕ=log(δ/α)logδ .

Whenα=δ/2, the definition ofϕin this lemma means ϕ(q)=δn=ϕ

2n when2nq <2n+1.

The following lemma shows that we would not lose too much by considering only functionsϕobtained in this way. The number2in2n is not important, it could be replaced with any other number greater than one.

Lemma 4. LetϕE. Define a functionϕ by letting, for any integersq1and n0:

ϕ(q)=ϕ

2n when2nq <2n+1.

(7)

Then we haveϕE,γϕ=γϕ, andμϕ(ξ )=μϕ(ξ )for anyξ.

Proof. The propertiesϕEandγϕ=γϕ are obvious, andμϕ(ξ )=μϕ(ξ )follows from Remark 2 sinceϕ(q)dividesϕ(q)for anyq1. 2

The functionϕ of Lemma 4 is useful to prove the following statement, which shows “how many” integersq are multiples ofϕ(q). This statement will be helpful in the proof of metric results in Section 2.4.

Proposition 1. LetϕE, and denote by Qϕ the set of allq1such thatϕ(q) dividesq. Then for any ε >0, the series qQϕq1+γϕε is convergent and the seriesqQϕq1+γϕ+ε is divergent.

Proof. Letε >0; we may assumeε <1−γϕ. For anyn0andq such that2n q <2n+1, we let ϕ1(q)=ϕ(2n) and ϕ2(q)=ϕ(2n+1). As in Lemma 4, we have ϕ1, ϕ2Ewithγϕ1=γϕ2=γϕ, andQϕ2QϕQϕ1. This implies

qQϕ

q1+γϕε

qQϕ1

q1+γϕε

=

n0

qQϕ1 2nq<2n+1

q1+γϕε

n0

2n 1+γϕε 2n

ϕ(2n)<+∞

since, fornsufficiently large,(2n)γϕε< 1

n2ϕ(2n). In the same way,

qQϕ

q1+γϕ+ε

qQϕ2

q1+γϕ+ε

=

n0

qQϕ2 2nq<2n+1

q1+γϕ+ε

n0

2n+1 1+γϕ+ε 2n

ϕ(2n+1)= +∞

since(2n+1)γϕ+εϕ(2n+1)for n sufficiently large. This concludes the proof of Proposition 1. 2

2.4. Metric results

Proposition 2. LetϕE. For almost anyξ∈Rin the sense of Lebesgue measure, we haveμϕ(ξ )=2−γϕ.

(8)

Proof. Thanks to Lemma 1, it is enough to prove that for any ε >0 the set of all ξ ∈ [0,1] with μϕ(ξ ) >2−γϕ +ε has Lebesgue measure 0. Now this set is contained, for anyq01, in the union of[pq 1

q2−γϕ+ε,pq+ 1

q2−γϕ+ε]with0pq andqQϕ,qq0(whereQϕis defined in the statement of Proposition 1), so that is has measure less than or equal to

qQϕ qq0

2(q+1)

q2γϕ+ε 4

qQϕ qq0

1 q1γϕ+ε.

Now Proposition 1 proves that this upper bound is finite, and tends to 0 asq0tends to infinity. This concludes the proof of Proposition 2. 2

Proposition 3. LetϕE andμ >2−γϕ. Then the set of all real numbersξ ∈R such thatμϕ(ξ )μhas Lebesgue measure zero and Hausdorff dimension2μγϕ.

In the special caseϕ =1, we obtain the classical theorem of Jarník [15] and Besicovitch [3] stating that the set of ξ ∈R such that μ(ξ )μ has Hausdorff dimension2/μ(see for instance [5], p. 104, Theorem 5.2, or Chapter 10 of [8]).

Proposition 3 follows immediately from Proposition 1 and the following theorem due to Borosh and Fraenkel [4].

Theorem 4 (Borosh–Fraenkel). Letν∈ [0,1], andQbe a subset ofNsuch that, for anyε >0, the seriesqQqνε is convergent and the seriesqQqν+ε is divergent. Letμ > ν+1. Then the set of allξ ∈Rfor which there are infinitely manyqQsuch that

ξp q

1

qμ for somep∈Z has Hausdorff dimensionν+μ1.

Actually in [4] it is assumed that the series qQqν is divergent, but this assumption is not necessary (see the remark before Lemma 2.1 of [22], p. 72).

3. A TRANSFERENCE THEOREM

In this section, we prove Theorem 2 stated in the Introduction. Let us recall the following from Lemma 3 stated in Section 2.3. Letn)n1be a sequence of positive integers such thatδndividesδn+1for anyn1, andδn1/ntends toδasn→ ∞. Let α∈Rbe such that0< α < δ. Define a functionϕby

ϕ(q)=δn withn=

logq log(δ/α)

(3.1) .

Then we haveϕE andγϕ=log(δ/α)logδ . We can now re-state Theorem 2 as follows.

(9)

Theorem 5. Letξ∈R\Q,0< α <1,β >1, and(δn)n1be a sequence of positive integers such thatδndividesδn+1for anyn1, andδ1/nn tends toδasn→ ∞. Let ϕE be the function defined by(3.1). Then the following implications hold.

(i) If there exist two integer sequences (un) and (vn) such that un0, |unξvn|1/nα,u1/nnβandδndividesunfor anyn, then we have

μϕ(ξ )logβ−logα logδ−logα. (3.2)

(ii) If we have

μϕ(ξ ) <logβ−logα logδ−logα

then there exist two integer sequences(un)and(vn)such thatun0,|unξvn|1/nα,u1/nnβandδndividesunfor anyn.

In the special case whereβ=δβ0andα=δ/β0, we have loglogβδloglogαα =2.This is the situation with Apéry’s construction forξ =ζ (3)and ξ =ζ (2), and also with Alladi and Robinson’s [1] forξ =log 2 (see Section 4 for more details). In this case, thanks to Lemma 3 and Proposition 3, the upper bound (3.2) means thatξ lies outside a set of Hausdorff dimension1−2 loglogδβ0 whereas the usual bound μ(ξ ) 1−loglogβα means thatξ lies outside a set of Hausdorff dimension 1−loglogβδ0. This means this Hausdorff dimension has come twice closer to 1.

Putting part (i) of Theorem 5 with the lower bound of Lemma 1, we obtain the following corollary which is a special case of the linear independence criteria of [12].

Corollary 4. Letξ ∈R,0< α <1,β >1, and(δn)n1be a sequence of positive integers such thatδndividesδn+1for anyn1, andδn1/ntends toδasn→ ∞.

If there exist two integer sequences (un) and (vn) such that un0, |unξvn|1/nα,u1/nnβ andδndividesunfor anyn, then we have

δαβ.

Proof of (i) of Theorem 5. If δ =1, we have γϕ =0 thanks to Lemma 3, so thatμϕ(ξ )=μ(ξ )using Corollary 3, and assertion (i) follows from the classical implication (1.3). So we may assumeδ >1.

Letμ > logβlogδloglogαα,ε >0 be sufficiently small, andq be sufficiently large such thatϕ(q)dividesqand|ξp/q|< qμfor somep∈Z. Let

n=

(logq)log(δ−ε)−logδ+logα log(α+ε)log(δ/α)

+1

andm= [log(δ/α)logq ]. We havelog(δ−ε)−logδ+logα <log(α+ε) <0so thatmn, andϕ(q)=δm dividesδn. Thereforeδmis a common divisor ofq andun, and the

(10)

determinantuq

n p vn

is an integer multiple ofδm. Now this determinant is equal (up to a sign) touq

n p unξvn

; we shall prove it has absolute value less thanδm, so that it is zero.

Sincelog(α+ε) <0, the lower boundn(logq)log(δlog(αε)+ε)log(δ/α)logδ+logα yields n

logq log(α+ε) <log(δ−ε) log(δ/α) −1,

so thatlogq+nlog(α+ε) < (m+1)log(δ−ε)andq(α+ε)n< (δε)m+1. On the other hand, with this choice ofnwe haven(logq)log(δlog(αε)+ε)log(δ/α)logδ+logα+1so that

n

logq log(β+ε)−log(δ−ε) log(δ/α)

1

log(δ/α)

log(β+ε)(log(δε)−log(δ/α))

log(α+ε) −log(δ−ε)

.

Ifεis sufficiently small, the right-hand side is close enough to loglog(δ/α)βlogδ to ensure that it is less thanμ−1. Therefore we have

μ−1> n

logq log(β+ε)−log(δ−ε) log(δ/α) ,

so that−1)logq > nlog(β+ε)(m+1)log(δ−ε)and+ε)n< qμ1ε)m+1.

Using these two estimates, we obtain the following upper bound for the absolute value ofuq

n −p unξvn

:

q(α+ε)n++ε)n

qμ1 <2(δ−ε)m+1< δm. Therefore this determinant is zero, and

1

qμ1>|p| = q

un|unξvn|> q αε

β+ε n

henceqμ< ((β+ε)/(αε))n, therefore μlogq < n

log(β+ε)−log(α−ε)

< (logq)

log(β+ε)−log(α−ε) log(δ−ε)−logδ+logα log(α+2ε)log(δ/α) which contradicts the assumption onμforεsufficiently small. This concludes the proof of (i) of Theorem 5. 2

The following lemma is essentially a special case of the one proved in [10] (in the proof of Lemma 7.3, on p. 39). We give the proof since (as announced in [10]) it is really easier than the one of [10].

(11)

Lemma 5. Letε and Qbe real numbers such that 0< ε <1 and Q >1. Let ξ∈R\Qbe such that0< ξ <1. Then at least one of the following two assertions holds:

(i) There exist integerspandq such that1q <2ε and ξp

q < 3

qQ. (3.3)

(ii) There exist integerspandq such thatQq2Qand p+ε

q ξp+3ε

q .

This lemma is useful whenεis really bigger than1/Q. In this case,p/qis a very precise approximation toξ in case (i), whereas in case (ii) it is not precise at all but we have a very good control upon the exact size ofqandp(namely, not only upper bounds as usual, but also lower bounds).

Proof of Lemma 5. Let F be the set of all fractions p/q with 0pq and 1q 2Q. ForfF, we writef = ˜p/q˜ as a fraction in its lowest terms, and alsof =p/q whereq is the least denominator off such thatQq2Q. With this convention, forfFwe denote byIf the interval[p+qε,p+q]. Let us assume that (ii) does not hold, i.e.ξ does not belong to any of these intervalsIf. Assertion (i) holds withq=1andp=0ifξε/Q=minI0, so we can assumeξ >minI0

and thereforeξ >maxI0. Letf be the greatest fraction inF such thatmaxIf <

ξ. Sinceξ <1, there is a least element fF such thatf> f. Thanks to our assumption onξ, we haveξ /If so that ξ <minIf. Lettingf =p/q andf= p/qwith the same convention as above, we have

p+3ε

q < ξ <p+ε q . (3.4)

SinceQq, q2Q, this gives p

qp q >

qε q ε

2Q. (3.5)

Now we write pq =pq˜˜ and pq=pq as fractions in their lowest terms. Since they are consecutive Farey fractions, it is well known (see for instance [13]) thatq˜+ q>2Qand pqpq =q˜1q. Letm=min(q,˜ q)andM=max(q,˜ q). ThenM > Q so that pq pq =q˜1q < mQ1 . Thanks to (3.5), this impliesm <2/ε. Now we use Equation (3.4) to bound from above the distance ofξ to the fraction (either pq˜˜ orpq) with denominatorm. Ifm= ˜q we obtain

ξp˜

˜ q

<p qp

q + ε q < 1

mQ+ 2

mQ= 3

mQ

(12)

whereas ifm=qwe obtain ξp

q

<max ε

q, p qp

q −3ε q

< 2 mQ.

So in both cases assertion (i) holds. This concludes the proof of Lemma 5. 2 Proof of (ii) of Theorem 5. First of all, let us notice that the assumptions of (ii) implyδ < β, sinceμϕ(ξ )1thanks to Lemma 1. Letμbe such thatμϕ(ξ ) < μ <

logβlogα

logδlogα. Letnbe sufficiently large. We denote byξnbe the fractional part ofδnξ, Qn=3βn

δn and εn=3αnδn δn

so thatεn<1andQn>1. Let us apply Lemma 5 toξn,εnandQn. In the first case, we obtain integersunandvnsuch thatun<23αn δδn

n and

|unξnvn| 3 Qn = δn

βn < 1

nδn)μ1 < 1 (unδn)μ1 sinceμ−1<log(β/δ)log(δ/α).

By definition ofξn, there is an integerv˜n such thatunξnvn=unδnξ− ˜vn. So we have

ξv˜n unδn

= 1

unδn|unξvn|< 1 (unδn)μ.

Moreover we haveϕ(unδn)=δkwith k=

log(unδn) log(δ/α)

log(23αnδn) log(δ/α)

n

so thatδk dividesδn, and finallyδk=ϕ(unδn)dividesunδn. Sinceμϕ(ξ ) < μ, this is possible only for a finite number of values of n. Therefore, as soon as n is sufficiently large, statement (ii) in Lemma 5 holds; letpn and qn be the integers provided by this lemma. We have

Qnqn2Qn and εnqnξnpnn,

so thatlimqn1/n=β/δandlim|qnξnpn|1/n=α. As above, there is an integerp˜n such thatqnξnpn=qnδnξ − ˜pn. Lettingun=δnqn and vn= ˜pn concludes the proof of (ii) of Theorem 5. 2

4. APPLICATION TO PARTICULAR NUMBERSξ

Let us start by summarizing Theorem 2 and Proposition 3 in the following corollary (which contains Corollary 1 as a special case).

(13)

Corollary 5. Letξ∈R\Q,0< α <1,β >1andδ1. Assume there exist integer sequences(un)n1,(vn)n1and(δn)n1such thatδndividesunandδn+1for anyn, and

un0, |unξvn|1/nα, u1/nnβ and δn1/nδ.

LetϕEbe defined byϕ(q)=δnwheren= [log(δ/α)logq ]. Then we have

μϕ(ξ )logβ−logα logδ−logα.

Moreover, the setS of allξ∈R\Qsuch thatμϕ) >loglogβ−δloglogαα, which does not containξ, has Hausdorff dimension loglogδβ2 loglogαα.

Whenβ=δβ0andα=δ/β0, the Hausdorff dimension ofSis1−2 loglogδβ0 (see the remark after the statement of Theorem 5). The linear forms constructed by Apéry forξ =ζ (3)satisfy the assumptions of Corollary 5 withδ=e3andβ0=(1+√

2)4 so that S has Hausdorff dimension1−2 log(13+2)4 =0.5745. . . .This Hausdorff dimension is larger than what we have been able to deduce from other constructions of linear forms in 1 andζ (3)(due to Dvornicich and Viola [7], Hata [14], Rhin and Viola [19]), eventhough these constructions yield better upper bounds forμ(ζ (3)). It would be pleasant to have a precise statement showing that Apéry’s linear forms are the ones, among a given set, that give the largest Hausdorff dimension forS. Trying to find a point of view from which Apéry’s linear forms would be “better”

than its further refinements was the starting point of [10] (see also [11]).

Forξ=ζ (2), the situation is similar. Apéry’s linear forms correspond toδ=e2 andβ0=((

5−1)/2)5, so thatS has Hausdorff dimension1−2 log(((25+1)/2)5

) = 0.5843. . . .

Forξ=log(2), Alladi–Robinson’s linear forms [1] giveδ=eandβ0=3+2√ 2, so thatShas Hausdorff dimension0.7163. . . .

Letξ be an algebraic irrational number, andϕE. Assume there exists a finite setSof primes such that, for anyq1, all prime factors ofϕ(q)belong toS(this is the case for instance whenϕ is constructed as in the beginning of Section 2).

Then Ridout’s theorem [20] implies μϕ(ξ )=2−γϕ. It would be interesting to generalize this result to other functions ϕ (for instance the one of Corollary 1).

Indeed, whenξ is an algebraic irrational number orξ ∈ {log 2, ζ (2), ζ (3)}, it seems natural to imagine thatμϕ(ξ )=2−γϕ for anyϕE.

5. REFINED RESULTS FORζ (3)

In this section, we prove Theorem 3 stated in the introduction (of which Corollary 2 is an immediate consequence). We follow Dubitskas’ proof [6] in the case oflog 2.

It is known that Apéry’s linear forms are such that, for somec1, c2>0, unc1dn3(

2+1)4n

n3/2 and unζ (3)vnc2dn3(√ 2−1)4n

n3/2 (5.1)

(14)

(forunthis is due to Cohen [18], see also Example 3.2 of [23] or [16]; forunζ (3)vnsee [17]).

Letq be a sufficiently large positive integer. Letnbe such that 3c2

(√ 2−1)4n

n3/2 1 q <3c2

(

2−1)4(n1) (n−1)3/2 . Then we have q > (3c2)1(n−1)3/2(

2+1)4(n1) > (

2+1)4(n+1) since n is sufficiently large, so that[ logq

log((1+

2)4)]n+1anddn3+1dividesq. Now the determinant uvn

n un+1 vn+1

is non-zero (this classical fact [18] can be deduced from the estimates (5.1)), so that at least one amonguvn

n q p

anduvn+1

n+1 q p

is non-zero. Let us assume thatuvn

n q p

=0(otherwise the proof is similar). Sincedn3 divides the coefficients in the first row, this determinant has absolute value greater than or equal todn3. We obtain in this way (sincenis large enough)

dn3un|qζ (3)p| +q|unζ (3)vn| 2c1dn3(

2+1)4n

n3/2 |qζ (3)p| +2 3dn3 so that

|qζ (3)p| 1 6c1

n3/2√ 2−1 4n n3×3c2

(

2−1)4(n1) (n−1)3/2 n3

q (logq)3

q ,

thereby concluding the proof.

ACKNOWLEDGEMENTS

I am grateful to Tanguy Rivoal and Wadim Zudilin for providing me with very useful references connected to this work. I would like also to thank the referee for his careful reading of the paper.

REFERENCES

[1] Alladi K., Robinson M. – Legendre polynomials and irrationality, J. Reine Angew. Math.318 (1980) 137–155.

[2] Apéry R. – Irrationalité deζ (2)etζ (3), in: Journées Arithmétiques (Luminy, 1978), Astérisque, vol. 61, 1979, pp. 11–13.

[3] Besicovitch A.S. – Sets of fractional dimension (IV): on rational approximation to real numbers, J. London Math. Soc.9(1934) 126–131.

[4] Borosh I., Fraenkel A.S. – A generalization of Jarník’s theorem on diophantine approximations, Indag. Mathem.34(1972) 193–201.

[5] Bugeaud Y. – Approximation by Algebraic Numbers, Cambridge Tracts in Math., vol. 160, Cambridge University Press, 2004.

Références

Documents relatifs

PROPOSITION 3. Let A be a Gorenstein noetherian local ring con- taining and let I be an unmixed ideal of finite projective dimen-.. Consequently the ring A/I is

This was first proved by Comets and Yoshida [3, Theorem 2.4.4], for a model of directed polymers in which the random walk is replaced by a Brownian motion and the environment is

Because the box dimension of a set is always greater or equal to its Hausdorff dimension, the inequality of our theorem, also gives an upper bound for the Hausdorff dimension.. If the

The main result of Section 2 is Theorem 2.3 in which it is shown that the symbolic power algebra of monomial ideal I of codimension 2 which is generically a complete intersection

We would like to solve this problem by using the One-dimensional Convex Integration Theory.. But there is no obvious choice

In this paper, our first goal is to show that indeed there is a sharp estimate for λ 1 (M) under Ricci curvature lower bound. To prove this, we will develop a new argument, that

An expansion is developed for the Weil-Petersson Riemann cur- vature tensor in the thin region of the Teichm¨ uller and moduli spaces.. The tensor is evaluated on the gradients

In 1979, Ejiri [Eji79] obtained the following rigidity theorem for n( ≥ 4)- dimensional oriented compact simply connected minimal submanifolds with pinched Ricci curvatures in