• Aucun résultat trouvé

Sign changes of error terms related to arithmetical functions

N/A
N/A
Protected

Academic year: 2021

Partager "Sign changes of error terms related to arithmetical functions"

Copied!
26
0
0

Texte intégral

(1)

Paulo J. ALMEIDA

Sign changes of error terms related to arithmetical functions Tome 19, no1 (2007), p. 1-25.

<http://jtnb.cedram.org/item?id=JTNB_2007__19_1_1_0>

© Université Bordeaux 1, 2007, tous droits réservés.

L’accès aux articles de la revue « Journal de Théorie des Nom-bres de Bordeaux » (http://jtnb.cedram.org/), implique l’accord avec les conditions générales d’utilisation (http://jtnb.cedram. org/legal/). Toute reproduction en tout ou partie cet article sous quelque forme que ce soit pour tout usage autre que l’utilisation à fin strictement personnelle du copiste est constitutive d’une infrac-tion pénale. Toute copie ou impression de ce fichier doit contenir la présente mention de copyright.

cedram

Article mis en ligne dans le cadre du

Centre de diffusion des revues académiques de mathématiques http://www.cedram.org/

(2)

de Bordeaux 19 (2007), 1-25

Sign changes of error terms related to

arithmetical functions

par Paulo J. ALMEIDA

Résumé. Soit H(x) =Pn≤x φ(n)

n −

6

π2x. Motivé par une

conjec-ture de Erdös, Lau a développé une nouvelle méthode et il a dé-montré que #{n ≤ T : H(n)H(n+1) < 0}  T. Nous considérons des fonctions arithmétiques f (n) =P

d|n bd

d dont l’addition peut être exprimée comme P

n≤xf (n) = αx + P (log(x)) + E(x). Ici P (x) est un polynôme, E(x) = −P

n≤y(x) bn

nψ x

n + o(1) avec ψ(x) = x − bxc − 1/2. Nous généralisons la méthode de Lau et démontrons des résultats sur le nombre de changements de signe pour ces termes d’erreur.

Abstract. Let H(x) =Pn≤x φ(n)

n −

6

π2x. Motivated by a

con-jecture of Erdös, Lau developed a new method and proved that #{n ≤ T : H(n)H(n + 1) < 0}  T. We consider arithmetical functions f (n) =P

d|n bd

d whose summation can be expressed as P

n≤xf (n) = αx + P (log(x)) + E(x), where P (x) is a polynomial, E(x) = −P n≤y(x) bn nψ x n + o(1) and ψ(x) = x − bxc − 1/2. We generalize Lau’s method and prove results about the number of sign changes for these error terms.

1. Introduction

We say that an arithmetical function f (x) has a sign change on integers at x = n, if f (n)f (n + 1) < 0. The number of sign changes on integers of f (x) on the interval [1, T ] is defined as

Nf(T ) = #{n ≤ T, n integer : f (n)f (n + 1) < 0}.

We also define zf(T ) = #{n ≤ T, n integer : f (n) = 0}. Throughout this

work, ψ(x) = x − bxc − 1/2 and f (n) will be an arithmetical function such

Manuscrit reçu le 8 janvier 2006.

This work was funded by Fundação para a Ciência e a Tecnologia grant number SFRH/BD/4691/2001, support from my advisor and from the department of mathematics of University of Georgia.

(3)

that

f (n) =X

d|n

bd

d for some sequence of real numbers bn.

The motivation for our work was a paper by Y.-K. Lau [5], where he proves that the error term, H(x), given by

X n≤x φ(n) n = 6 π2x + H(x)

has a positive proportion of sign changes on integers solving a conjecture stated by P. Erdös in 1967.

An important tool that Lau used to prove his theorem, was that the error term H(x) can be expressed as

(1) H(x) = − X n≤ x log5 x µ(n) n ψ x n  + O  1 log20x  , (S. Chowla [3])

We generalize Lau’s result in the following way

Theorem 1.1. Suppose H(x) is a function that can be expressed as

(2) H(x) = − X n≤y(x) bn nψ x n  + O  1 k(x)  ,

where each bn is a real number and

(i) y(x) increasing, x14  y(x)  x

(log x)5+ D2

, for some D > 0, and

(3) X

n≤x

b4n x logDx;

(ii) k(x) is an increasing function, satisfying limx→∞k(x) = ∞.

(iii) H(x) = H(bxc) − α{x} + θ(x), where α 6= 0 and θ(x) = o(1).

Let ≺ ∈ {<, =, ≤}. If #{1 ≤ n ≤ T : αH(n) ≺ 0}  T then there exists a positive constant c0 and c0T disjoint subintervals of [1, T ], with each of

them having at least two integers, m and n, such that αH(m) > 0 and

αH(n) ≺ 0. In particular,

(1) #{n ≤ T : αH(n) > 0}  T ;

(2) if #{n ≤ T : αH(n) < 0}  T , then NH(T )  T or zH(T )  T .

We consider arithmetical functions f (n) for which, the error term of the summation function satisfies the conditions of Theorem 1.1. A first class is described in the following result

(4)

Theorem 1.2. Let f (n) be an arithmetical function and suppose the

se-quence bn satisfies condition (3) and

(4) X n≤x bn= Bx + O x logAx !

for some B real, D > 0 and A > 6 + D2, respectively. Let α =P∞n=1 bn

n2, γb = limx→∞  X n≤x bn n − B log x  and H(x) = X n≤x f (n) − αx +B log 2πx 2 + γb 2 .

If α 6= 0, then Theorem 1.1 is valid for the error term H(x). Moreover, if

f (n) is a rational function, then, except when α = 0, or B = 0 and α is

rational, we have

NH(T )  T if and only if #{n ≤ T : αH(n) < 0}  T.

Notice that this class of arithmetical functions is closed for addition, i.e., if f (n) and g(n) are members of the class then also is (f + g)(n). In the case considered by Lau, it was known that H(x) has a positive proportion of negative values (Y.-F. S. Pétermann [6]), so the second part of Theorem 1.2 generalizes Lau’s result. Another example is f (n) = φ(n)n .

Using a result of U. Balakrishnan and Y.-F. S. Pétermann [2] we are able to apply Theorem 1.1 to more general arithmetical functions:

Theorem 1.3. Let f (n) be an arithmetical function and suppose the

se-quence bn satisfies condition (3) and

(5) ∞ X n=1 bn ns = ζ β(s)g(s)

for some β real, D > 0, and a function g(s) with a Dirichlet series expan-sion absolutely convergent for σ > 1 − λ, for some λ > 0. Let α = ζβ(2)g(2)

and H(x) = ( P n≤xf (n) − αx, if β < 0, P n≤xf (n) − αx − Pbβc j=0Bj(log x)β−j if β > 0,

where the constants Bj are well defined. If α 6= 0, then Theorem 1.1 is valid

for the error term H(x).

Theorem 1.3 is valid for the following examples, where r 6= 0 is real:

φ(n) n r , σ(n) n r , φ(n) σ(n) r .

(5)

2. Main Lemma

The main tool used by Y.-K. Lau was his Main Lemma, where he proved that if H(x) =P n≤x φ(n) n − 6 π2x then Z 2T T Z t+h t H(u) du !2 dt  T h,

for sufficiently large T and any 1 ≤ h  log4T . Lau’s argument depends essentially on the formula (1). In this section, we obtain a generalization of Lau’s Main Lemma.

Main Lemma. Suppose H(x) is a function that can be expressed as (2)

and satisfies conditions (i) and (ii) of Theorem 1.1. Then, for all large T and h ≤ min log T, k2(T )

, we have (6) Z 2T T Z t+h t H(u) du !2 dt  T h32. For any positive integer N , define

(7) HN(x) = − X d≤N bd dψ x d  . The Main Lemma will follow from the next result.

Lemma 2.1. Assume the conditions of the Main Lemma and take D > 0

satisfying condition (i). Let E = 4 +D2, then

(a) For any δ > 0, large T , any Y  T and N ≤ y(T ), we have

Z T +Y T (H(u) − HN(u))2 du  Y N1−δ + Y k2(T )+ y(T + Y ) (log T ) E;

(b) For all large T , N ≤ y(T ) and 1 ≤ h ≤ min log T, k2(T )

, we have Z 2T T Z t+h t HN(u) du !2 dt  T h32 + N3(log N )E. Now we prove the Main Lemma:

Proof. Take N = T14 and δ > 0 small. Cauchy’s inequality gives us

Z t+h t H(u) du !2 ≤ 2 Z t+h t HN(u) du !2 + 2 Z t+h t (H(u) − HN(u)) du !2 .

(6)

Since N = T14 then, for sufficiently large T , N3logEN  T . So, using part (b) of Lemma 2.1 we have Z 2T T Z t+h t HN(u) du !2 dt  T h32 + N3logEN  T h 3 2. Using Cauchy’s inequality and interchanging the integrals,

Z 2T T Z t+h t (H(u) − HN(u)) du !2 dt ≤ h Z 2T T Z t+h t (H(u) − HN(u))2 du ! dt ≤ h Z 2T +h T Z min(u,2T ) max(u−h,T ) (H(u) − HN(u))2dt ! du ≤ h2 Z 2T +h T (H(u) − HN(u))2 du  h2 T + h N1−δ + T + h k2(T )+ y(2T + h) (log T ) E  T + T h  T h32

since y(2T + h)  (log T )TE+1 and h ≤ min(log T, k2(T )). Hence

Z 2T T Z t+h t H(u) du !2 dt  T h32.  3. Step I

In this section, we will prove part (a) of Lemma 2.1. Using expression (2) and Cauchy’s inequality, we obtain

Z T +Y T (H(u) − HN(u))2du ≤ 2 Z T +Y T   y(u) X m=N +1 bm mψ u m    2 du + O  Y k2(T )  . Let η(T, m, n) = max T, y−1(m), y−1(n) , then 2 Z T +Y T   y(u) X m=N +1 bm mψ u m    2 du = 2 y(T +Y ) X m,n=N +1 bmbn mn Z T +Y η(T,m,n) ψ u m  ψ u n  du.

(7)

The Fourier series of ψ(u) = u − buc −12, when u is not an integer, is given by (8) ψ(u) = −1 π ∞ X k=1 sin(2πku) k , so we obtain 2 π2 y(T +Y ) X m,n=N +1 bmbn mn ∞ X k,l=1 1 kl Z T +Y η(T,m,n) sin  2πku m  sin  2πlu n  du. Now, the integral above is equal to

1 2 Z T +Y η(T,m,n) cos  2πu k m + l n  − cos  2πu k m− l n  du. For the first term we get

Z T +Y η(T,m,n) cos  2πu k m + l n  du   1 k m + l n . If mk = nl, then Z T +Y η(T,m,n) cos  2πu k m − l n  du ≤ Y, otherwise Z T +Y η(T,m,n) cos  2πu k m− l n  du  1 k m − l n .

Part (a) of Lemma 2.1 will now follow from the next three lemmas.

Lemma 3.1. Let E = 4 +D2 as in Lemma 2.1. Then

X m,n≤X |bmbn| ∞ X k,l=1 kn6=lm 1 kl | kn − lm|  X (log X) E .

Lemma 3.2. If D > 0 satisfies condition (3), then X m,n≤X |bmbn| ∞ X k,l=1 1 kl ( kn + lm)  X (log X) 1+D2 .

Lemma 3.3. For any δ > 0, X N <m,n≤X |bmbn| mn ∞ X k,l=1 kn=lm 1 kl  1 N1−δ.

(8)

In order to finish the proof of part (a) of Lemma 2.1 we just need to take X = y(T + Y ) in the previous lemmas. Hence

Z T +Y T (H(u) − HN(u))2 du  Y N1−δ + y(T + Y ) (log T ) E + Y k2(T ).  Before we prove the three lemmas above, we need the following technical result

Lemma 3.4. Let bn be a sequence satisfying condition (3). Then

X n≤N b2n N logD2 N, X n≤N |bn|  N log D 4 N, X n≤N b2n n  (log N ) 1+D 2 , X n≤N |bn| n  (log N ) 1+D4 , X n>N b4n n2τ (n)  1 N1−δ, for any δ > 0.

Proof. Follows from Cauchy’s inequality, partial summations and the fact

that, for any  > 0, τ (n) = O(n). 

Remark. If H(x) can be expressed in the form (2), then

(9) |H(x)| ≤ X n≤y(x) |bn| n + O  1 k(x)   (log x)1+D4 .

Proof of Lemma 3.2 : Since the arithmetical mean is greater or equal to

the geometrical mean, we have

X m,n≤X |bmbn| ∞ X k,l=1 1 kl(kn + lm) ≤ 2 X m,n≤X |bmbn| ∞ X k,l=1 1 kl√knlm  X m≤X |bm| √ m 2  X m≤X 1 X M ≤X b2M M   X(log X)1+D2. 

Proof of Lemma 3.3 : For the second sum, take d = (m, n), m = dα and

n = dβ. Since kn = lm, then α|k and β|l. Taking k = αγ, we also have l = βγ. As (10) ∞ X k,l=1 kn=lm 1 kl = 1 αβ ∞ X γ=1 1 γ2 = π2 6 (m, n)2 mn .

(9)

Then, X N <m,n≤X |bmbn| mn ∞ X k,l=1 kn=lm 1 kl = π2 6 X N <m,n≤X |bmbn|(m, n)2 m2n2 ≤ π 2 6 X d≤X d X N <m≤X d|m |bm| m2 !2 .

The next step is to estimate the inner sum using Hölder inequality

X N <m≤X d|m |bm| m2 !2 ≤ X N <m≤X d|m b4m m2 ! 1 2 X N <M ≤X d|M 1 M2 ! 3 2 . Set M = βd, then (11) X N <M ≤X d|M 1 M2 ! 3 2 = 1 d3 X N d<β≤ X d 1 β2 ! 3 2  1 d3 min ( 1, d 3 N3 ) ! 1 2 .

To complete the proof of Lemma 3.3 we use Cauchy’s inequality and Lemma 3.4. For any δ > 0,

X N <m,n≤X |bmbn| mn ∞ X k,l=1 kn=lm 1 kl  X d≤X 1 d  min ( 1, d 3 N3 ) 12 X N <m≤X d|m b4m m2 12 ! , therefore X N <m,n≤X |bmbn| mn ∞ X k,l=1 kn=lm 1 kl !2  X d≤X 1 d2min ( 1, d 3 N3 ) X D≤X X N <m≤X D|m b4m m2 !  1 N3 X d≤N d +X d>N 1 d2  X N <m≤X b4m m2 X D|m 1  1 N X N <m≤X b4m m2τ (m)  1 N2−δ. 

Proof of Lemma 3.1 : This lemma is a generalization of Hilfssatz 6 in [9] of

A. Walfisz. Notice first that

X m,n≤X  |bmbn| ∞ X k,l=1 kn6=lm 1 kl | kn − lm|  ≤ 2 X m≤n≤X  |bmbn| ∞ X k,l=1 kn6=lm 1 kl | kn − lm|  .

(10)

Like in [9] we begin by separating the interior sum into four terms: ∞ X k,l=1 kn6=lm 1 kl | kn − lm| = ∞ X k,l=1 lm≤kn2  1 kl | kn − lm|  + ∞ X k,l=1 kn 2 <lm<kn  1 kl | kn − lm|  + ∞ X k,l=1 kn<lm<2 kn  1 kl | kn − lm|  + ∞ X k,l=1 lm≥2 kn  1 kl | kn − lm|  .

For the first term, we use Lemma 3.4,

X m≤n≤X |bmbn| ∞ X k,l=1 lm≤kn2 1 kl | kn − lm| ≤ 2 X m≤n≤X |bmbn| ∞ X k,l=1 lm≤kn2 1 k2ln = 2 X m≤n≤X |bm||bn| n ∞ X k=1 1 k2 X l≤2mkn 1 l  X m≤n≤X |bm| |bn| n ∞ X k=1 log k k2 + log X k2   log X X n≤X |bn| n X m≤n |bm|  X (log X)1+ D 2.

From the forth inequality of Lemma 3.4, we get

X m≤n≤X |bmbn| ∞ X k,l=1 kn≤lm 2 1 kl | kn − lm|  X (log X) 1+D2 .

The estimation of the third term is more complicated and we have to use a different approach. In this case, 1l < 2mkn, so that

X m≤n≤X |bmbn| ∞ X k,l=1 kn 2 <lm<kn 1 kl | kn − lm| < 2 X m≤n≤X |bm||bn| n ∞ X k=1 1 k2 X kn 2m<l< kn m m kn − lm

(11)

< 2 X m≤n≤X |bm||bn| n ∞ X k=1 1 k2 X l≤knm−1 1 kn m − l ! + 2 X m≤n≤X |bm| |bn| n ∞ X k=1 1 k2 X kn m−1<l< kn m m kn − lm. Now, taking L =hknm − li, X m≤n≤X |bm| |bn| n ∞ X k=1 1 k2 X l≤knm−1 1 kn m − l ! ≤ X m≤n≤X |bm| |bn| n ∞ X k=1 1 k2 X L≤knm−1 1 L  X (log X)1+D2 ,

as in the first term. If there exists an integer l with knm − 1 < l < kn m, then

m - kn. In this case, kn − lm = mnknmoand m < n. So, we have to estimate

(12) X m<n≤X |bm||bn| n ∞ X k=1 m- kn 1 k2nkn m o.

Notice that the fractional part of knm is at least m1. So, when k ≥ m,

X m<n≤X |bm| |bn| n ∞ X k=m m k2  X m<n≤X |bm| |bn| n  X (log X) D 2 . We are left with the estimation of

X m<n≤X |bm| |bn| n X k<m m- kn 1 k2nkn m o.

Since m - kn, given k and n, we can take ak,n, such that 1 ≤ ak,n < m and

ak,n ≡ kn mod m. Then, X m<n≤X |bm| |bn| n X k<m m- kn 1 k2nkn m o ≤ X m<n≤X |bm| |bn| n X k<m m k2a k,n ≤ X a,k≤X 1 ak2 X max(a,k)<m≤X m|bm| X m<n≤X kn≡a mod m |bn| n .

We need to estimate the inner sums. In order to do that, we partition the interval [1, X] in intervals of the form [M, 2M ) and apply Cauchy’s

(12)

inequality. Take 1 ≤ P ≤ Q ≤ X, then, X P ≤m<2P m|bm| X Q≤n<2Q kn≡a mod m |bn| n  P Q X P ≤m<2P |bm| X Q≤n<2Q kn≡a mod m |bn|.

Next, we apply Cauchy’s inequality twice, first to the first sum on the right and afterwards to the second sum:

X P ≤m<2P |bm| X Q≤n<2Q kn≡a mod m |bn| !!2 ≤ X P ≤M <2P b2M X P ≤m<2P X Q≤n<2Q kn≡a mod m |bn| !2  P logD2 P X P ≤m<2P X Q≤n<2Q kn≡a mod m b2n X Q≤N <2Q kN ≡a mod m 1 !  P logD2 P X P ≤m<2P  1 + Qm (k,m)  X Q≤n<2Q kn≡a mod m b2n ! . Since m ≤ 2P ≤ 2Q, we have mQ ≥ 1

2. Using also (k, m) ≤ k, we obtain

1 + Qm (k,m) ≤ 1 +Qk m ≤ Q m(k + 2) ≤ 3 Qk m . Therefore, X P ≤m<2P  |bm| X Q≤n<2Q kn≡a mod m |bn|  !2  P logD2 P X P ≤m<2P 3Qk m X Q≤n<2Q kn≡a mod m b2n !  P3kQ P log D 2 P X Q≤n<2Q b2n X P ≤m<2P m|kn−a 1 !  kQ (log P )D2 X Q≤n<2Q b2n τ (kn − a).

(13)

By a theorem of S. Ramanujan [7],P

n≤Xτ2(n) ∼ X log3X and by another

application of Cauchy inequality and condition (3), we get

 X Q≤n<2Q b2n τ (kn − a)2 ≤ X Q≤n<2Q b4n X Q≤n<2Q τ2(kn − a)  Q logDQ X kQ−a≤N <2kQ−a τ2(N )  kQ2logD+3X. Therefore, X P ≤m<2P  m|bm| X Q≤n<2Q kn≡a mod m |bn| n   P Q  kQ(log P )D2 X Q≤n<2Q b2nτ (kn − a) 1 2  P Q  kQ(log X)D2(kQ2logD+3X) 1 2 12  P k34(log X)1+ D 2.

The number of pairs of intervals of the form ([P, 2P ), [Q, 2Q)) to be con-sidered is at most  log2X, hence

X a,k≤X 1 ak2 X a<m≤X m|bm| X m<n≤X kn≡amod m |bn| n  X a,k≤X k34 ak2 X P,Q P (log X)1+D2  X (log X)4+D2 . The fourth term is treated as the third, hence

X m≤n≤X |bmbn| ∞ X k,l=1 kn<lm<2 kn 1 kl | kn − lm|  X (log X) 4+D2 .

This completes the proof of Lemma 3.1. 

4. Step II

In this section we prove part (b) of Lemma 2.1. From equation (8), we get Z b a ψ(u) du = 1 2π2 ∞ X k=1 cos(2πkb) k2 − cos(2πka) k2  .

(14)

Using the definition of HN stated in (7), we obtain Z t+h t HN(u) du = − X m≤N bm m Z t+h t ψ(u m) du = − 1 2π2 X m≤N bm ∞ X k=1 cos2πk(t+h)m − cos2πktm k2 .

As usual, let us write e(t) for e2πit, then

Z t+h t HN(u)du = 1 4π2 X m≤N bm ∞ X k=1  ekhm− 1emkt e−k(2t+h)m − 1 k2 . Therefore, 16π4 Z 2T T Z t+h t HN(u)du 2 dt = Z 2T T X m≤N bm ∞ X k=1  ekhm− 1emkt e−k(2t+h)m − 1 k2 2 dt = X m,n≤N bmbn ∞ X k,l=1  ekhm− 1 e−lhn− 1 (kl)2 × Z 2T T e kt m  e  −lt n  e  −k (2t + h) m  − 1  e  l (2t + h) n  − 1  dt.

After multiplying the terms inside the integral above, we obtain the follow-ing four terms that we will estimate below:

Z 2T T e kt m − lt n  dt + e lh n − kh m  Z 2T T e lt n − kt m  dt − e  −kh m  Z 2T T e  −lt n − kt m  dt − e lh n  Z 2T T e lt n+ kt m  dt.

(15)

Notice that, R2T T e2πirtdt ≤ 1

π|r|, for any r 6= 0. We begin with the last

term and use |e(t) − 1)| ≤ 2 and Lemma 3.2. Then

X m,n≤N bmbn ∞ X k,l=1  ekhm− 1 e−lh n  − 1elhn (kl)2 Z 2T T e lt n + kt m  dt ≤ 4 π X m,n≤N |bmbn| ∞ X k,l=1 1 (kl)2l n+ k m   X m,n≤N |bmbn| ∞ X k,l=1 m k  n l  1 kl(lm + kn)  N 3(log N )1+D2 .

The third term is treated similarly to obtain

X m,n≤N bmbn ∞ X k,l=1  ekhm− 1 e−lhn− 1e−khm (kl)2 Z 2T T e  −lt n − kt m  dt  N3(log N )1+D2 . Now, if kn = lm, then Z 2T T e kt m − lt n  dt + e lh n − kh m  Z 2T T e lt n − kt m  dt = 2T. If kn 6= lm then, Z 2T T e kt m − lt n  dt + e lh n − kh m  Z 2T T e lt n − kt m  dt  1 k m − l n .

Let us study first the case when kn 6= lm,

X m,n≤N bmbn ∞ X k,l=1 kn6=lm  ekhm− 1 e−lh n  − 1 (kl)2 Z 2T T e kt m − lt n  dt  X m,n≤N |bmbn| ∞ X k,l=1 kn6=lm 1 (kl)2 k m− l n  X m,n≤N |bmbn| ∞ X k,l=1 kn6=lm m k  n l  1 kl |kn − lm|  N 3logEN,

(16)

by Lemma 3.1. Similarly, X m,n≤N bmbn ∞ X k,l=1 kn6=lm  ekhm− 1 e−lhn− 1elhn −khm (kl)2 Z 2T T e lt n − kt m  dt  N3logEN.

If kn = ml, we will use |e(t) − 1|  min(1, |t|) instead. The expression obtained has some similarities with Lemma 3.3. We are going to use the same argument to prove:

(13) X m,n≤N |bmbn| ∞ X k,l=1 kn=lm 1 (kl)2 min  1,kh m  min  1,lh n   h32.

As in Lemma 3.3, take d = (m, n), α = md, β = nd and γ = αk. So l = βγ and then ∞ X k,l=1 kn=lm 1 (kl)2 min  1,kh m  min  1,lh n  = 1 α2β2 ∞ X γ=1 1 γ4  min  1,hγ d 2 . If d ≤ h, we obtainP∞γ=1γ14  min1,hγd2= π904, and if h < d ≤ N , ∞ X γ=1 1 γ4  min  1,hγ d 2 = h d 2 X γ≤d h 1 γ2 + X γ>d h 1 γ4  h d 2 + h d 3  h d 2 . Therefore, X m,n≤N |bmbn| ∞ X k,l=1 kn=lm 1 (kl)2min  1,kh m  min  1,lh n  = X m,n≤N |bmbn| (m, n)4 m2n2 ∞ X γ=1 1 γ4  min  1, hγ (m, n) 2 ≤ X d≤N d4 X m,n≤N d=(m,n) | bmbn| m2n2 ∞ X γ=1 1 γ4  min  1,hγ d 2  X d≤h  d2 X m≤N d|m |bm| m2 2 + h2 X h<d≤N  d X m≤N d|m |bm| m2 2 .

(17)

Since d|m we have m > h and so, h2P h<d≤N d P m≤N d|m |bm| m2 !2  h1+δ. To

estimate the first term we begin with Hölder inequality:

X d≤h d4 X m≤N d|m |bm| m2 2 ≤ X d≤h d4 X m≤N d|m |bm|4 m2 12 X M ≤N d|M 1 M2 32 .

The third sum is Od13

 (similar to (11)). Then X d≤h d4 X m≤N d|m |bm|4 m2 12 X M ≤N d|M 1 M2 32  X d≤h d X d≤m≤N d|m |bm|4 m2 12  X d≤h d2 1 2h X D≤h  X D≤m≤N D|m |bm|4 m2 i12  h32  X m≤N |bm|4 m2 X D≤h D|m 1 1 2  h32  X m≤N |bm|4 m2 τ (m) 12 .

The last part of Lemma 3.4 implies P m≤N

|bm|4

m2 τ (m) = O(1), where the underlying constant doesn’t depend on N . Therefore, we obtain inequality

(13) and part (b) of Lemma 2.1, follows. 

5. A general Theorem

In this section, we prove Theorem 1.1, from which the main Theorems 1.2 and 1.3, will be deduced.

Proof of Theorem 1.1 : From the Main Lemma, we have, for all large T

and h ≤ min log T, k2(T )

, Z 2T T Z t+h t H(u) du !2 dt  T h32.

Assume #{n ≤ T : αH(n) ≺ 0}  T . Let c > 0 be a constant and T be sufficiently large, such that #{n ≤ 2T : αH(n) ≺ 0} > cT. Divide the interval [1, 2T ] into subintervals of length h, where h is a sufficiently large integer satisfying h ≤ log T . Then more than cT /h of those subintervals must have at least one integer n with αH(n) ≺ 0. Let C be the set of the subintervals which satisfy this property. Write C = {Jr| 1 ≤ r ≤ R}, where

(18)

where R > cT /h. Define Ks = J3s−2, for 1 ≤ s ≤ R/3, and let D be the

set of these subintervals. We have #(D) > cT /3h. Notice that any two members of D are separated by a distance of at least 2h.

Let M be the number of subintervals K in D for which there exists an integer n in K such that αH(n) ≺ 0 and αH(m) ≤ 0 for every integer m ∈ (n, n + 2h), and let S be the set of the corresponding values of n.

Lemma 5.1. For some absolute constant c1, we have M ≤ c1 T h32

.

Proof. Since H(x) = H(bxc) − α{x} + θ(x), then

αH(x) − αH(bxc) = −α2{x} + αθ(x). So, if x is sufficiently large and not an integer then

(14) −5

2{x} < αH(x) − αH(bxc) < −3

2{x}.

Let n1 be the smallest integer such that any non integer x > n1 satisfies

condition (14). If #{n ∈ S : n ≥ n1} = 0 then M ≤ n1, so, the lemma is

clearly true for sufficiently large T . Otherwise,

#{n ∈ S : n ≥ n1} ≥ M − n1  M.

Take n ∈ S with n ≥ n1 and t ∈ [n, n + h]. For any integer m ∈ [t, t + h],

αH(m) ≤ 0. Now, for any 1 ≤ j < h,

Z [t]+j+1 [t]+j H(u) du = Z [t]+j+1 [t]+j (H(u) − H([t] + j)) du + Z [t]+j+1 [t]+j H([t] + j) du. Therefore, by (14), Z [t]+j+1 [t]+j αH(u) du < Z 1 0  −3 4α 2xdx + αH([t] + j) < −3 8α 2,

because αH([t] + j) ≤ 0. Since [t] ≥ n, we also have

Z [t]+1 t αH(u) du < Z 1 {t}  −3 4α 2xdx + αH([t]) (1 − {t}) < 0 and Z t+h [t]+h αH(u) du < Z {t} 0  −3 4α 2xdx + αH([t] + h){t} ≤ 0. Hence, Z t+h t H(u) du ≥ 3 8|α|(h − 1).

(19)

Take an integer r = r(T ) such that 2r > (log T )3+D2. Using (6) and (9), we obtain Z 2T 0 Z t+h t H(u) du !2 dt = Z T 2r 0 Z t+h t H(u) du !2 dt + r X j=0 Z T 2j−1 T 2j Z t+h t H(u) du !2 dt  T 2rh 2(log T )2+D2 + h3 2 r X j=0 T 2j  T h32, due to h ≤ log T . On the other hand,

Z 2T 0 Z t+h t H(u) du !2 dt ≥ X n∈S Z n+h n Z t+h t H(u) du !2 dt ≥ X n∈S n≥n1 Z n+h n 3 8|α|(h − 1) 2 dt  M h3. Hence M ≤ c1 T h32

for some absolute constant c1. 

Take c0 = c/6h. If h is a suitably large integer such that c1T < c0T h

3 2, then there are at least c0T intervals K in D such that αH(n) ≺ 0 for some

integer n ∈ K and αH(m) > 0 for some integer m lying in (n, n + 2h). Now, suppose #{n ≤ T : αH(n) ≤ 0}  T . Take T sufficiently large and take the order relation ‘ ≺’ to be ‘ ≤’. Therefore, we have c0T integers m in

the interval [1, 2T ], for which αH(m) is positive. In this case, #{n ≤ T : αH(n) > 0}  T.

If we don’t have #{n ≤ T : αH(n) ≤ 0}  T , then #{n ≤ T : αH(n) > 0} = T (1 + o(1)).

Hence, part 1 of Theorem 1.1 is proved. Next, we prove part 2. Take ‘ ≺’ to be ‘ <’. Then, there exists a positive constant c0 and c0T disjoint

subin-tervals of [1, T ], with each of them having at least two integers, m and n, such that H(m) > 0 and H(n) < 0. Therefore, in each of those intervals we have at least one l with either H(l) = 0 or H(l)H(l + 1) < 0. Whence, zH(T ) > c20T or NH(T ) > c20T . 

(20)

6. A class of arithmetical functions

In this section, we consider arithmetical functions f (n), such that the sequence bnsatisfies conditions (3) and (4). We begin with some elementary

results about this class of arithmetical functions. Using condition (4), we immediately obtain the following lemma:

Lemma 6.1. Let bn be a sequence of real numbers satisfying (4), for some constants B and A > 1, then there exist constants γb and α such that

X n≤x bn n = B log x + γb+ O 1 logA−1x ! , (15) ∞ X n=1 bn n2 = α, (16) X n>x bn n2 = B x + O 1 x logA−1x ! . (17)

Next, we calculate the sum of f (n) and describe the error term H(x).

Lemma 6.2. Let bn be a sequence of real numbers as in Lemma 6.1, then

(18) X n≤x f (n) =X n≤x X d|n bd d = αx − B log 2πx 2 − γb 2 + H(x), where, (19) H(x) = − X n≤ x logC x bn nψ x n  + O 1 logCx ! + O 1 logA−C−1x ! , for any 0 < C < A − 1. Proof. We have X n≤x f (n) = X n≤x X d|n bd d = X d≤x bd d X n≤x d|n 1 = X m≤x X d≤mx bd d.

We separate the double sum above in two parts. Let 0 < C < A − 1 and y = logCx. Then (20) X m≤x X d≤mx bd d = X n≤y X d≤xn bd d + X d≤xy bd d X y<n≤xd 1.

(21)

In order to evaluate the first term on the right, we start with an application of formula (15): X n≤y X d≤xn bd d = X n≤byc B log x − B log n + γb+ O 1 logA−1(nx) !! = Bbyc log x − B X n≤byc log n + γbbyc + O y logA−1(xy) ! .

Recall that by Stirling formula,P

n≤byclog n = byc logbyc − byc + logbyc 2 + log 2π 2 + O  1 y  . Hence, X n≤y X d≤nx bd

d = Bbyc (log x − logbyc + 1) + γbbyc −

B(log 2πy) 2 + O y logA−1x ! + O 1 y  . For the second term, we get

X d≤x y bd d X y<n≤x d 1 = X d≤x y bd d x d  − byc  = x ∞ X d=1 bd d2 − x X d>x y bd d2 − X d≤x y bd d ψ x d  − 1 2 + byc  X d≤x y bd d = αx − By − X d≤xy bd d ψ x d  −γb+ B log x 2 − Bbyc(log x − log y) +B log y

2 − γbbyc + O y logA−1x

!

.

Notice also that Bbyc(log y − logbyc) = Bbyc− log1 −{y}y = B{y} + O1y. Joining everything together, we obtain

H(x) = X n≤x f (n) −  αx − B log 2πx 2 − γb 2  = − X d≤ x logC x bd dψ x d  + O 1 logA−C−1x ! + O 1 logCx ! . 

(22)

Proof of Theorem 1.2 : We just have to show that H(x) satisfies the

con-ditions of Theorem 1.1. From Lemma 6.2, for any x H(x) − H(bxc) = −α{x} −B 2 (log 2πbxc − log 2πx) = −α{x} +B 2 {x} x + O  1 x2  . In Lemma 6.2, we also obtained

H(x) = − X n≤ x logC x bn nψ x n  + O 1 logCx ! + O 1 logA−C−1x ! ,

for any 0 < C < A − 1. Take C = 5 +D2, y(x) = x

logCx and k(x) = minlogCx, logA−C−1x. Since A > 6 + D/2, then C < A − 1 and A − C − 1 > 0. The first part of Theorem 1.2 now follows from Theorem 1.1.

Suppose that f (n) takes only rational values. In order to prove the second part of Theorem 1.2, we use the following result of A. Baker [1].

Proposition. Let α1, . . . , αnand β0, . . . , βndenote nonzero algebraic

num-bers. Then β0+ β1log α1+ · · · + βnlog αn6= 0.

Using the result above, we obtain the next lemma.

Lemma 6.3. Let f (n) be a rational valued arithmetical function and

sup-pose the sequence bn satisfies condition (4) for some real B and A > 1. Let

r be a real number and suppose H(x) is given by (18). Then

(1) If B = 0 and α is irrational then #{n integer : H(n) = r} ≤ 1; (2) If B is a nonzero algebraic number then #{n integer : H(n) = r} ≤

2;

(3) If B is transcendental then there exists a constant C that depends

on r and on the function f (n), such that

#{n ≤ T, n integer : H(n) = r} < (log T )C.

Proof. Suppose that B = 0 and α is irrational. Suppose also that there are

two integers, say M 6= N , such that H(M ) = H(N ). Then

X n≤M f (n) − αM +γb 2 = X n≤N f (n) − αN + γb 2. But this implies that α is rational, a contradiction.

Next, suppose B 6= 0 is algebraic number and that there are M > N > Q integers, satisfying H(M ) = H(N ) = H(Q). We have

X n≤M f (n) − αM + B log 2πM 2 + γb 2 = X n≤N f (n) − αN +B log 2πN 2 + γb 2

(23)

which implies α = B M − N log M N  + 1 M − N X N <n≤M f (n). Consequently B log      M N M −N1  M Q  1 M −Q     = 1 M − Q X Q<n≤M f (n) − 1 M − N X N <n≤M f (n).

We are going to prove that (21) M N M −N1 6= M Q M −Q1 .

Since B is a nonzero algebraic number and the values of f (n) are rational, for any integer n, the proposition implies

B log      M N M −N1  M Q M −Q1     6= 1 M − Q X Q<n≤M f (n) − 1 M − N X N <n≤M f (n),

and so we get a contradiction, which implies #{n integer : H(n) = r} ≤ 2, for any real number r. In fact, instead of proving (21), we are going to prove that

(22) MN −QQM −N < NM −Q,

for any positive integers M > N > Q. Clearly, this implies (21). The inequality (22) is just a particular case of the geometric mean-analytic mean inequality (23) n Y i=1 ui !1 n ≤ 1 n n X i=1 ui,

where equality happens only if u1 = u2 = · · · = un. In fact, if we take

n = M − Q, ui = M for 1 ≤ i ≤ N − Q and ui = Q for N − Q < i ≤ M − Q,

we derive  MN −QQM −N 1 M −Q < 1 M − Q((N − Q)M + (M − N )Q) = N. Hence, we obtain (22) and part 2 of the Lemma.

Finally we prove part 3. Suppose r is a real number such that #{n ≤ T : H(n) = r} ≥ 4.

(24)

Let Q < N < M be the three smallest positive integers in the above set, then 0 6= B log     M N M −N1 M Q M −Q1     = 1 M − Q X Q<n≤M f (n) − 1 M − N X N <n≤M f (n).

Suppose L is such that H(L) = r. Then L > N > Q, and as in part 2:

0 6= B log     L N L−N1 L Q L−Q1     = 1 L − Q X Q<n≤L f (n) − 1 L − N X N <n≤L f (n).

After we cross multiply the two expressions above, we obtain

log      M N M −N1  M Q M −Q1     = r1log      L N L−N1  L Q L−Q1     ,

for some rational r1. Therefore, there are four rational numbers r2, r3, r4

and r5, such that

Lr2 = Mr3Nr4Qr5.

Now, any prime dividing L must divide M N Q. Notice that, if p is a prime, k is an integer and pk≤ x then k ≤ log x

log p. Therefore, the number of integers

smaller than x, which have all prime divisors smaller than M is smaller than (log x)π(M ). This finishes the proof.  Except when α = 0, or B = 0 and α is rational, we cannot have zH(T ) 

T . Hence, we obtain the second part of Theorem 1.2. 

Example. We finish this section by proving that Theorem 1.2 is valid for

the arithmetical function φ(n)n . Notice that n φ(n) = Y p|n  1 + 1 p − 1  =X d|n µ2(d) φ(d). Let bn= µ 2(n)n φ(n) , then f (n) = P d|nbdd. In [8], R. Sitaramachandrarao proved that X n≤x µ2(n)n φ(n) = x + O  x12  ,

(25)

so condition (4) is satisfied for any A and with B = 1. By Merten’s Theorem Q p|n  1 −1p−1 ≤Q p≤n  1 −1p−1 ∼ eγlog n, hence X n≤x b4n= X n≤x µ2(n) n 4 φ4(n) = X n≤x Olog4n= Ox log4x, and condition (3) is satisfied for D ≥ 4. In this case,

α = ζ(2)ζ(3) ζ(6) , γb = γ + X p log p p(p − 1) and H(x) = X n≤x n φ(n)− ζ(2)ζ(3) ζ(6) x + log x 2 + log 2π + γ +P p log p p(p−1) 2 .

Since B = 1 we can apply Theorem 6.3, and so z(T ) ≤ 2. Therefore, if #{n ≤ T : αH(n) < 0}  T, then NH(T )  T .

7. Second class of arithmetical functions

Given a sequence of real numbers bn, and a complex number s, we define

the Dirichlet series B(s) =P∞n=1bn

ns. In this section, we consider arithmeti-cal functions f (n), such that the sequence bnsatisfies conditions (3) and (5)

for some D > 0, β real and a function g(s) with a Dirichlet series expansion absolutely convergent for σ > 1 − λ, for some λ > 0.

U. Balakrishnan and Y.-F. S. Pétermann [2] proved that:

Proposition. Let f (n) be a complex valued arithmetical function satisfying ∞ X n=1 f (n) ns = ζ(s)ζ β(s + 1)g(s + 1),

for a complex number β, and g(s) having a Dirichlet series expansion

g(s) = ∞ X n=1 cn ns,

which is absolutely convergent in the half plane σ > 1 − λ for some λ > 0. Let β0 be the real part of β. If

ζβ(s)g(s) = ∞ X n=1 bn ns

(26)

then there is a real number b, 0 < b < 1/2, and constants Bj, such that,

taking y(x) = x exp−(log x)b and α = ζβ(2)g(2), X n≤x f (n) = ( αx −P n≤y(x)bnnψ x n  + o(1) if β0< 0, αx +P[β0] j=0Bj(log x)β−j−Pn≤y(x) bnnψ x n  + o(1) if β0≥ 0,

The real version of the previous proposition allows us to prove Theo-rem 1.3:

Proof. Notice that, for any c > 0, logcbxc = logcx − c{x}x logc−1x + Ox1. So, H(x) = H(bxc) − α{x} + o(1). From the previous proposition, there is an increasing function k(x), with limx→∞k(x) = ∞, such that

H(x) = − X n≤y(x) bn nψ x n  + O  1 k(x)  ,

where y(x) = x exp−(log x)b, for some 0 < b < 1/2. Hence, the result

follows from Theorem 1.1. 

Acknowledgements: I would like to thank my advisor Andrew Granville

for his guidance and encouragement in this research and to the Referee for his useful comments.

References

[1] A. Baker, Linear forms in the logarithms of algebraic numbers (III). Mathematika 14 (1967), 220–228.

[2] U. Balakrishnan, Y.-F. S. Pétermann, The Dirichlet series of ζ(s)ζα(s + 1)f (s + 1):

On an error term associated with its coefficients. Acta Arithmetica 75 (1996), 39–69.

[3] S. Chowla, Contributions to the analytic theory of numbers. Math. Zeit. 35 (1932), 279– 299.

[4] G. H. Hardy, E. M. Wright, An Introduction to the Theory of Numbers. 5th ed. Oxford 1979.

[5] Y.-K. Lau, Sign changes of error terms related to the Euler function. Mathematika 46 (1999), 391–395.

[6] Y.-F. S. Pétermann, On the distribution of values of an error term related to the Euler

function. Théorie des Nombres, (Quebec, PQ, 1987), 785–797.

[7] S. Ramanujan, Collected papers. Cambridge, 1927, 133–135.

[8] R. Sitaramachandrarao, On an error term of Landau - II. Rocky Mount. J. of Math.

15 2 (1985), 579–588.

[9] A. Walfisz, Teilerprobleme II. Math. Zeit. 34 (1931), 448–472. Paulo J. Almeida

Departamento de Matemática Universidade de Aveiro

Campus Universitário de Santiago 3810-193 Aveiro, Portugal

Références

Documents relatifs

Using further properties of the induction functor one shows that the cohomology of a Deligne–Lusztig variety X(wF ) depends only on the conjugacy class of wF in the braid group..

USER GENERATED CONTEXTS: THINKING ABOUT CHANGES IN MASS COMMUNICATION IN TERMS OF AGENCY, INNOVATION, TRUST AND RISK..

Kaytoue, M., Kuznetsov, S.O., Napoli, A., Duplessis, S.: Mining Gene Expression Data with Pattern Structures in Formal Concept Analysis.. Madeira, S.C., Oliveira, A.L.:

The initial purpose of the contextual inquiry was to obtain field information about service management, with special attention to service architecture, in a way that can be used

To control the terms around the central point d ∈ [x 1/2−ε , x 1/2+ε ] we will intro- duce a short sieve weight to maintain the feature that our sums are essentially supported

For continuous Galerkin finite element methods, there now exists a large amount of literature on a posteriori error estimations for (positive definite) problems in mechanics

Gene expression data can be represented as a matrix, where rows and columns represent genes and experiments respectively.. Each cell contains the numeric expression level of a

We extend the applicability of the Generalized Linear Sampling Method (GLSM) [2] and the Factorization Method (FM)[14] to the case of inhomogeneities where the contrast change