• Nie Znaleziono Wyników

Normality of numbers generated by the values of polynomials at primes

N/A
N/A
Protected

Academic year: 2021

Share "Normality of numbers generated by the values of polynomials at primes"

Copied!
12
0
0

Pełen tekst

(1)

LXXXI.4 (1997)

Normality of numbers generated by the values of polynomials at primes

by

Yoshinobu Nakai (Kofu) and Iekata Shiokawa (Yokohama) To the memory of Norikata Nakagoshi

1. Introduction. Let r ≥ 2 be a fixed integer and let θ = 0.a

1

a

2

. . . be the r-adic expansion of a real number θ with 0 < θ < 1. Then θ is said to be normal to base r if, for any block b

1

. . . b

l

∈ {0, 1, . . . , r − 1}

l

,

n

−1

N (θ; b

1

. . . b

l

; n) = r

−l

+ o(1)

as n → ∞, where N (θ, b

1

. . . b

l

; n) is the number of indices i ≤ n − l + 1 such that a

i

= b

1

, a

i+1

= b

2

, . . . , a

i+l−1

= b

l

. Let (m)

r

denote the r- adic expansion of an integer m ≥ 1. For any infinite sequence {m

1

, m

2

, . . .}

of positive integers, we consider the number 0.(m

1

)

r

(m

2

)

r

. . . whose r-adic expansion is obtained by the concatenation of the strings (m

1

)

r

, (m

2

)

r

, . . . of r-adic digits, which will be written simply as 0.m

1

m

2

. . . (r).

Copeland and Erd˝os [1] proved that the number 0.m

1

m

2

. . . (r) is normal to base r for any increasing sequence {m

1

, m

2

, . . .} of positive integers such that, for every positive % < 1, the number of m

i

’s up to x exceeds x

%

provided x is sufficiently large. In particular, the normality of the number

0.23571113 . . . (r)

defined by the primes was established. Davenport and Erd˝os [2] proved that the number

0.f (1)f (2) . . . f (n) . . . (r)

is normal to base r, where f (x) is any nonconstant polynomial taking posi- tive integral values at all positive integers.

In this paper, we prove the following

Theorem. Let f (x) be as above. Then the number α(f ) = 0.f (2)f (3)f (5)f (7)f (11)f (13) . . . (r)

1991 Mathematics Subject Classification: 11K, 11L.

[345]

(2)

defined by the values of f (x) at primes is normal to base r. More precisely, for any block b

1

. . . b

l

∈ {0, 1, . . . , r − 1}

l

, we have

(1) n

−1

N (α(f ); b

1

. . . b

l

; n) = r

−l

+ O

 1 log n



as n → ∞, where the implied constant depends possibly on r, f , and l.

2. Preliminary of the proof of the Theorem. Let α(f ) = 0.a

1

a

2

. . . a

n

. . . be the r-adic expansion of the number α(f ) given in the Theorem. Then each a

n

belongs to the corresponding string (f (p

ν

))

r

, where p

ν

is the νth prime and ν = ν(n) is defined by

ν−1

X

i=1

([log

r

f (p

i

)] + 1) < n ≤ X

ν

i=1

([log

r

f (p

i

)] + 1).

Here [t] denotes the greatest integer not exceeding the real number t. We put x = x(n) = p

ν(n)

, so that

n = X

p≤x

log

r

f (p) + O(π(x)) + O(log

r

f (x)) (2)

= dx log r + O

 x log x

 ,

where d ≥ 1 is the degree of the polynomial f (t), p runs through prime numbers, and π(x) is the number of primes not exceeding x. We used here the prime number theorem:

π(x) = Li x + O

 x

(log x)

G

 , where G is a positive constant given arbitrarily and

Li x =

x

\

2

dt log t . Then we have

N (α(f ); b

1

. . . b

l

; n) = X

p≤x

N (f (p); b

1

. . . b

l

) + O(π(x)) + O(log

r

f (x))

= X

p≤x

N (f (p); b

1

. . . b

l

) + O

 n log n



with x = x(n) = p

ν(n)

.

Let j

0

be a large constant. Then for each integer j ≥ j

0

, there is an integer n

j

such that

r

j−2

≤ f (n

j

) < r

j−1

≤ f (n

j

+ 1) < r

j

.

(3)

We note that

n

j

 r

j/d

and that n

j

< n ≤ n

j+1

if and only if the r-adic expansion of f (n) is of length j; namely,

(3) (f (n))

r

= c

j−1

. . . c

1

c

0

∈ {0, 1, . . . , r − 1}

j

, c

j−1

6= 0.

For any x > r

j0

, we define an integer J = J(x) by n

J

< x ≤ n

J+1

, so that

(4) J = log

r

f (x) + O(1)  log x.

Let n be an integer with n

j

< n ≤ n

j+1

and j

0

< j ≤ J, so that (f (n))

r

can be written as in (3). We denote by N

(f (n); b

1

. . . b

l

) the number of occurrences of the block b

1

. . . b

l

appearing in the string 0 . . . 0 | {z }

J−j

c

j−1

. . . c

1

c

0

of length J. Then we have 0 ≤ X

p≤x

N

(f (p); b

1

. . . b

l

) − X

p≤x

N (f (p); b

1

. . . b

l

)

J−1

X

j=j0+1

(J − j)(π(n

j+1

) − π(n

j

)) + O(1)

J−1

X

j=j0+1

π(n

j+1

) + O(1) 

J−1

X

j=1

r

j/d

J  x

log x and so

(5) N (α(f ); b

1

. . . b

l

; n) = X

p≤x

N

(f (p); b

1

. . . b

l

) + O

 n log n



with x = x(n) = p

ν(n)

.

We shall prove in Sections 4 and 5 that

(6) X

p≤x

N

(f (p); b

1

. . . b

l

) = r

−l

π(x) log

r

f (x) + O

 x log x



which, combined with (5) and (2), yields (1).

(4)

3. Lemmas

Lemma 1 ([9; 4.19]). Let F (x) be a real function, k times differentiable, and satisfying |F

(k)

(x)| ≥ λ > 0 throughout the interval [a, b]. Then

b

\

a

e(F (x)) dx

≤ c(k)λ

−1/k

. Lemma 2 ([3; p. 66, Theorem 10]). Let

F (t) = h

q t

d

+ α

1

t

d−1

+ . . . + α

k

,

where h, q are coprime integers and α

i

’s are real. Suppose that (log x)

σ

≤ q ≤ x

d

(log x)

−σ

,

where σ > 2

6d

0

+ 1) with σ

0

> 0. Then

X

p≤x

e(F (p))

≤ c(d)x(log x)

−σ0

as x → ∞, where p runs through the primes.

Lemma 3 ([3; p. 2, Lemma 1.3 and p. 5, Lemma 1.6]). Let F (x) = b

0

x

d

+ b

1

x

d−1

+ . . . + b

d−1

x + b

d

be a polynomial with integral coefficients and let q be a positive integer. Let D be the greatest common divisor of q, b

0

, b

1

, . . . , and b

d−1

. Then

X

q n=1

e

 F (n) q



≤ d

3ω(q/D)

D

1/d

q

1−1/d

as q → ∞, where ω(n) is the number of distinct prime divisors of n.

Lemma 4 ([6; Corollary of Lemma]). Let F (x) be a polynomial with real coefficients with leading term Ax

d

, where A 6= 0 and d ≥ 2. Let a/q be a rational number with (a, q) = 1 such that |A − a/q| < q

−2

. Assume that

(log Q)

H

≤ q ≤ Q

d

/(log Q)

H

, where H > d

2

+ 2

d

G with G ≥ 0. Then

X

1≤n≤Q

e(F (n))

 Q(log Q)

−G

.

Lemma 5 ([7; Theorem], cf. [8; Theorem 1]). Let f (t) and b

1

. . . b

l

be as in Theorem. Then

X

n≤y

N (f (n); b

1

. . . b

l

) = r

−l

y log

r

f (y) + O(y)

as y → ∞, where the implied constant depends possibly on r, f , and l.

(5)

4. Proof of the Theorem. We have to prove the inequality (6). We write

X

p≤x

N

(f (p); b

1

. . . b

l

) = X

p≤x

X

J m=l

I

 f (p) r

m

 , where

I(t) =

 

 

 

 1 if

X

l k=1

b

k

r

−k

≤ t − [t] <

X

l k=1

b

k

r

−k

+ r

−l

, 0 otherwise.

There are functions I

(t) and I

+

(t) such that I

(t) ≤ I(t) ≤ I

+

(t), having Fourier expansion of the form

I

±

(t) = r

−l

± J

−1

+ X

ν=−∞

ν6=0

A

±

(ν)e(νt)

with

|A

±

(ν)|  min(|ν|

−1

, Jν

−2

),

where e(x) = e

2πix

([10; Chap. 2, Lemma 2]). We choose a large constant c

0

and put

(7) M = [c

0

log

r

J].

Then it follows that

(8) X

p≤x

N

(f (p); b

1

. . . b

l

)

Q

 X

l≤m≤dM

+ X

dM <m≤J−M

+ X

J−M <m≤J

 X

p≤x

I

±

 f (p) r

m



= P

1

+ π(x)

r

l

(J − dM ) + P

2

+ P

3

+ O(π(x)), where d is the degree of the polynomial f (x),

P

1

= P

1(±)

= X

l≤m≤dM

X

p≤x

I

±

 f (p) r

m

 , P

2

= P

2(±)

= X

dM <m≤J−M

X

1≤|ν|≤J2

A

±

(ν) X

p≤x

e

 ν r

m

f (p)

 ,

P

3

= P

3(±)

= X

J−M <m≤J

X

1≤|ν|≤J2

A

±

(ν) X

p≤x

e

 ν r

m

f (p)

 .

We first estimate P

2

. Suppose that dM ≤ m ≤ J − M . Then, writing

the leading coefficient of the polynomial νr

−m

f (t) as a/q with (a, q) = 1,

(6)

we have

(log x)

σ

≤ q ≤ x

d

(log x)

−σ

with a large constant σ, so that by Lemma 2,

X

p≤x

e

 ν r

m

f (p)



 x(log x)

−σ0

,

where σ

0

> 3 is a constant. Therefore we obtain

(9) P

2

 x(log x)

2−σ0

 x log x . Next we estimate P

3

. We appeal to the prime number theorem of the form referred to in Section 2. Then it follows that

X

p≤x

e

 ν r

m

f (p)



=

x

\

2

e

 ν r

m

f (t)



dπ(t) + O(1)

=

x

\

2

e

 ν r

m

f (t)

 dt log t + O

 x

(log x)

G



=

x

\

x(log x)−G

e

 ν r

m

f (t)

 dt log t + O

 x

(log x)

G



 1

log x sup

ξ

ξ

\

x(log x)−G

e

 ν r

m

f (t)

 dt

+ O

 x

(log x)

G



 1

log x

 |ν|

r

m



−1/d

+ O

 x

(log x)

G

 ,

using the second mean-value theorem and Lemma 1 with |νr

−m

f

(d)

(t)| 

|ν|r

−m

. Therefore we have P

3

 X

1≤|ν|≤J2

|ν|

−1

X

J−M ≤m≤J

 1 log x

 |ν|

r

m



−1/d

+ O

 x

(log x)

G



(10)

 1

log x X

1≤|ν|≤J2

1

|ν|

1+1/d

X

m≤J

r

−m/d

+ O

 x

(log x)

G−2



 x

log x .

To prove the Theorem, it remains to show that

(11) P

1

= π(x)

r

l

dM + O

 x log x



,

(7)

since this together with (4), (8), (9), and (10) implies X

p≤x

N

(f (p); b

1

. . . b

l

) = π(x)

r

l

J + O(π(x))

= π(x)

r

l

log

r

f (x) + O

 x log x

 , which is the inequality (6).

5. Proof of Theorem (continued). We shall prove the inequality (11) in three steps.

F i r s t s t e p. Suppose that l ≤ m ≤ dM , where M is given by (7) with (4). We appeal to the prime number theorem for arithmetic progres- sions of the following form ([4; Sect. 17]): Let π(x; q, a) be the number of primes p ≤ x in an arithmetic progression p ≡ a (mod q) with (a, q) = 1 and let ϕ(n) be the Euler function. Then

π(x; q, a) = 1

ϕ(q) Li x + O(xe

−c

log x

)

uniformly in 1 ≤ q ≤ (log x)

H

, where c > 0 is a constant which depends on a constant H > 0 given arbitrary. (A weaker result O(x(log x)

−G

) is enough for our purpose.) Let B denote the least common multiple of all denominators of the coefficients, other than the constant term, of f (t). Then

X

p≤x

I

±

 f (p) r

m



= X

p≤x (p,Br)=1

I

±

 f (p) r

m



+ O(1)

= X

a mod Brm (a,Br)=1

I

±

 f (a) r

m



π(x; Br

m

, a) + O(1)

= X

a mod Brm (a,Br)=1

I

±

 f (a) r

m

 1

ϕ(Br

m

) Li x + O

 x

(log x)

G



+ O(1)

= π(x) ϕ(Br

m

)

X

a mod Brm (a,Br)=1

I

±

 f (a) r

m

 + O



r

m

x (log x)

G

 .

Hence we have (12) P

1

Q X

l≤m≤dM

π(x) ϕ(Br

m

)

X

a mod Brm (a,Br)=1

I

±

 f (a) r

m

 + O



M r

dM

x (log x)

G



(8)

= X

l≤m≤dM

π(x) ϕ(Br

m

)

X

a mod Brm

I

±

 f (a) r

m

 X

b|(a,Br)

µ(b) + O

 x log x



= X

b|Br

µ(b) X

l≤m≤dM

π(x) ϕ(Br

m

)

X

a mod Brm b|a

I

±

 f (a) r

m

 + O

 x log x



= π(x) Br ϕ(Br)

X

b|Br

µ(b) X

l≤m≤dM

1 Br

m

X

1≤n≤Brm/b

I

±

 f (bn) r

m



+ O

 x log x

 ,

where µ(n) is the M¨obius function. Note that Br = O(1).

S e c o n d s t e p. We shall prove that, for each b | Br,

(13) X

l≤m≤dM

1 Br

m

X

1≤n≤Brm/b

I

±

 f (bn) r

m



= X

l≤m≤dM

1 Br

M

X

1≤n≤BrM/b

I

±

 f (bn) r

m



+ O(1).

If l ≤ m ≤ M , then we have 1

Br

m

X

1≤n≤Brm/b

I

±

 f (bn) r

m



= 1

Br

M

X

1≤n≤BrM/b

I

±

 f (bn) r

m

 ,

so that

(14) X

l≤m<M

1 Br

m

X

1≤n≤Brm/b

I

±

 f (bn) r

m



= X

l≤m≤M

1 Br

M

X

1≤n<BrM/b

I

±

 f (bn) r

m

 .

If d = 1, (14) implies (13). So in what follows we assume d ≥ 2 and M ≤ m ≤ dM . We have

X

1≤n≤Brm/b

I

±

 f (bn) r

m



Q Br

m

b · 1

r

l

+ O

 r

m

J

 + O

 X

1≤|ν|≤J2

1

|ν|

X

1≤n≤Brm/b

e

 ν r

m

f (bn)





= Br

m

b · 1

r

l

+ O

 r

m

J



+ O(r

m(1−1/d)

J

2/d

log J),

(9)

since, by Lemma 3,

X

1≤n≤Brm/b

e

 ν r

m

f (bn)



 (r

m

, ν)

1/d

r

m(1−1/d)

. Hence we get

(15) X

M ≤m≤dM

1 Br

m

X

1≤n≤Brm/b

I

±

 f (bn) r

m



= (d − 1)M

br

l

+ O(1).

In the rest of this step, we shall prove the inequality

(16) X

M ≤m≤dM

1 Br

M

X

1≤n≤BrM/b

I

±

 f (bn) r

m



= (d − 1)M

br

l

+ O(1), which together with (15) and (14) yields (13).

P r o o f o f (16). It is easily seen that

(17) X

M ≤m≤dM

1 Br

M

X

1≤n≤BrM/b

I

±

 f (bn) r

m



Q 1

Br

M

X

M ≤m≤dM

X

1≤n≤BrM/b

 1 r

l

+ O

 1 J



+ X

1≤|ν|≤J2

A

±

(ν)e

 ν r

m

f (bn)



= (d − 1)M

br

l

+ O(1) + O

 X

1≤|ν|≤J2

1

|ν| · 1 Br

M

X

M ≤m≤dM

X

1≤n≤BrM/b

e

 ν r

m

f (bn)



 .

We estimate the last sum. Let H be a large constant. For any ν, m, b, we can choose, by Dirichlet’s theorem, coprime integers a and q = q(ν, m, b) such that

1 ≤ q ≤ Q

d

/(log Q)

H

, Q = Br

M

/b

and

ν

r

m

b

d

a q

< (log Q)

H

qQ

d

(≤ 1/q

2

).

If

(log Q)

H

≤ q ≤ Q

d

/(log Q)

H

,

(10)

then by Lemma 4,

X

1≤n≤BrM/b

e

 ν r

m

f (bn)



 Q

(log Q)

G

 r

M

(log J)

2

. Hence the contribution of these sums in the last term in (17) is

 1

Br

M

(d − 1)M log J · r

M

(log J)

2

= O(1).

Otherwise, we have

1 ≤ q ≤ (log Q)

H

( M

H

).

In particular, (ν/r

m

)b

d

6= a/q, since m ≥ M . Hence 1

qr

m

ν

r

m

b

d

a q

 M

H

qr

dM

, so that

(dM ≥) m ≥ dM − H

1

log M, with a large constant H

1

. From this it follows that

d dt · ν

r

m

f (bt)  ν

r

m

t

d−1

 J

2

r

−M +H1log M

= o(1)

throughout the interval [1, Br

M

/b]. Thus by a van der Corput’s lemma ([9;

Lemma 4.8]) we have X

1≤n≤BrM/b

e

 ν r

m

f (bn)



=

BrM

\

/b 1

e

 ν r

m

f (bt)



dt + O(1)

 ν

r

m

f

(d)

(t)

−1/d

+ O(1) 

 |ν|

r

m



−1/d

,

using again Lemma 1. Hence the contribution of these sums to the last term in (17) is

 1

Br

M

X

M ≤m≤dM

X

1≤|ν|≤J2

1

|ν|

 |ν|

r

m



−1/d

= O(1).

Combining these results, we obtain (16).

(11)

T h i r d s t e p. It follows from (12) with (13) that P

1

Q π(x) Br ϕ(Br)

X

b|Br

µ(b) 1 Br

M

X

l≤m≤dM

X

1≤n≤BrM/b

I

±

 f (bn) r

m



+ O

 x log x



Q π(x) Br ϕ(Br)

X

b|Br

µ(b) 1 Br

M

X

l≤m≤dM

X

1≤n≤BrM/b

I

 f (bn) r

m



+ O

 x log x

 .

We put, in Lemma 5, y = Br

M

/b, so that log

r

f (by) = dM + O(1). Then we have

X

l≤m≤dM

X

1≤n≤BrM/b

I

 f (bn) r

m



= X

n≤y

N (f (bn); b

1

. . . b

l

) + O(r

M

)

= r

−l

y log

r

f (by) + O(r

M

)

= r

−l

Br

M

b dM + O(r

M

).

Therefore we obtain P

1

R Br ϕ(Br)

X

b|Br

µ(b) b · dM

r

l

π(x) + O

 x log x



= r

−l

dM π(x) + O

 x log x

 ,

which is (11). The proof of the Theorem is now complete.

References

[1] A. H. C o p e l a n d and P. E r d ˝o s, Notes on normal numbers, Bull. Amer. Math. Soc.

52 (1946), 857–860.

[2] H. D a v e n p o r t and P. E r d ˝o s, Note on normal decimals, Canad. J. Math. 4 (1952), 58–63.

[3] L.-K. H u a, Additive Theory of Prime Numbers, Transl. Math. Monograph 13, Amer.

Math. Soc., Providence, RI, 1965.

[4] M. N. H u x l e y, The Distribution of Prime Numbers, Oxford Math. Monograph, Oxford Univ. Press, 1972.

[5] Y.-N. N a k a i and I. S h i o k a w a, A class of normal numbers, Japan. J. Math. 16 (1990), 17–29.

[6] —, —, A class of normal numbers II , in: Number Theory and Cryptography, J. H. Loxton (ed.), London Math. Soc. Lecture Note Ser. 154, Cambridge Univ.

Press, 1990, 204–210.

(12)

[7] Y.-N. N a k a i and I. S h i o k a w a, Discrepancy estimates for a class of normal num- bers, Acta Arith. 62 (1992), 271–284.

[8] J. S c h i f f e r, Discrepancy of normal numbers, ibid. 47 (1986), 175–186.

[9] E. C. T i t c h m a r s h, The Theory of the Riemann Zeta-Function, 2nd ed., revised by D. R. Heath-Brown, Oxford Univ. Press, 1986.

[10] I. M. V i n o g r a d o v, The Method of Trigonometrical Sums in Number Theory, Nauka, 1971 (in Russian).

Department of Mathematics Department of Mathematics

Faculty of Education Keio University

Yamanashi University Hiyoshi, Yokohama, 223 Japan

Kofu, 400 Japan E-mail: shiokawa@math.keio.ac.jp

E-mail: nakai@grape.kkb.yamanashi.ac.jp

Received on 28.6.1996

and in revised form on 16.12.1996 (3014)

Cytaty

Powiązane dokumenty

Before leaving this section, we note that it is possible to improve on the result of Heath-Brown for gaps between squarefull numbers by combining the above estimate for S 1 (x 3/13

Gelfond method together with the knowledge of the number of zeros of certain meromorphic functions involving ℘(z) and ζ(z).. The Ramanujan functions are closely connected to

The following lemma is an important tool for various constructions in Banach spaces.. It enables one to generalize constructions in Hilbert spaces

Applying process II to the ν-summation, we have to consider the prime term and the almost-prime term... Then we have an

We did not use Watt’s mean-value bound (Theorem 2 of [12]) in prov- ing Lemma 6, because the hypothesis T ≥ K 4 (in our notation) limits the former’s usefulness in this problem to

In the present paper we show that if instead of primes one asks for almost primes of some fixed order r (that is, numbers with at most r prime factors, counted with multiplicity, or P

For any set X let |X| denote its cardinality and for any integer n, larger than one, let ω(n) denote the number of distinct prime factors of n and let P (n) denote the greatest

This abstract result provides an elementary proof of the existence of bifurcation intervals for some eigenvalue problems with nondifferentiable nonlinearities1. All the results