• Nie Znaleziono Wyników

for cyclic fields of prime degree l

N/A
N/A
Protected

Academic year: 2021

Share "for cyclic fields of prime degree l"

Copied!
18
0
0

Pełen tekst

(1)

LXXIV.4 (1996)

Congruence of Ankeny–Artin–Chowla type modulo p

2

for cyclic fields of prime degree l

by

Stanislav Jakubec (Bratislava)

Introduction. Let p ≡ 1 (mod 4), let T + U

p > 1 be the fundamental unit, and let h be the class number of Q(

p). The following congruence (Ankeny–Artin–Chowla congruence) holds:

h U

T ≡ B

(p−1)/2

(mod p).

For a cubic field K ⊂ Q(ζ

p

+ ζ

p−1

), p ≡ 1 (mod 3) the analogous congruence was proved by Feng Ke Qin in [1]. In general, for fields of the lth degree, the analogous congruence is proved in [5].

The aim of this paper is to prove a congruence of Ankeny–Artin–Chowla type modulo p

2

for real Abelian fields of a prime degree l and a prime conductor p. The importance of such a congruence can be demonstrated by the following example. Let K be a cubic field. In [5], the following congruence is proved:

(1) h

K

S

1

S

2

≡ −

34

B

(p−1)/3

B

2(p−1)/3

(mod p).

As is well known, h

K

< p for a cubic field. If B

(p−1)/3

B

2(p−1)/3

≡ 0 (mod p), then S

1

S

2

≡ 0 (mod p), hence the congruence (1) does not provide any information about h

K

. Note that such a prime exists, e.g. p = 5479. We have

B

(p−1)/3

= B

1826

≡ 0 (mod 5479).

In this paper two applications of the main theorem (Theorem 1), for a quadratic and for a cubic field, will be given.

Let l and p be primes such that p ≡ 1 (mod l) and let K ⊂ Q(ζ

p

+ ζ

p−1

) with [K : Q] = l. Let a be a primitive root modulo p. As is well known, the conjugates of the unit

1991 Mathematics Subject Classification: Primary 11R29.

[293]

(2)

η

a

= N

Q(ζ

p−1p )/K



ζ

p(1−a)/2

1 − ζ

pa

1 − ζ

p

 ,

generate the group of cyclotomic units C(K) of K. Consider the unit η

2

= N

Q(ζ

pp−1)/K

p

+ ζ

p−1

).

It is easy to prove that if 2 is not an lth power modulo p, then the conjugates of the unit η

2

generate the group C(K). Let hεi be the group generated by all conjugates of the unit ε.

According to ([3], Lemma 1, p. 69) for a cyclic field K with [K : Q] = l, there is a unit δ such that [U

K

: hδi] = f , where (p, f ) = 1.

The following is taken from [5]. According to [8] and [9] (see also [10], p. 284), we have h

K

= [U

K

: C(K)], where C(K) = hη

2

i is the group of cyclotomic units of K. From [U

K

: hδi] = f we have η

2f

∈ hδi. Let [hδi : hη

2f

i] = e. Clearly [hη

2

i : hη

2f

i] = f

l−1

.

Consider two towers of groups

f2

i ⊂ hδi ⊂ U

K

,

f2

i ⊂ hη

2

i ⊂ U

K

. This implies ef = h

K

f

l−1

and hence e = h

K

f

l−2

. Let

η

2f

= δ

c0

σ(δ)

c1

. . . σ

l−2

(δ)

cl−2

. It is easy to prove that

e = N

Q(ζl)/Q

(c

0

+ c

1

ζ

l

+ . . . + c

l−2

ζ

ll−2

).

Note that a unit δ for which f = 1 is called a strong Minkowski unit. As is well known, a strong Minkowski unit exists for a cyclic fields K with l < 23.

The problem of existence of a strong Minkowski unit for cyclic fields of small non-prime degree is solved in [7].

Let a be a fixed primitive root modulo p, let χ be the Dirichlet character of order n, n | p − 1, χ(x) = ζ

nindax

. Let g be such that g ≡ a

(p−1)/n

(mod p) and g

n

≡ 1 (mod p

p

). Denote by p a prime divisor of Q(ζ

n

) such that p | p and 1/g ≡ ζ

n

(mod p

p

).

Define the rational numbers A

0

(n), A

1

(n), . . . , A

n−1

(n) in the following way:

(2)

A

0

(n) = −1/n,

τ (χ

i

)

n

≡ n

n

A

i

(n)

n

(−p)

i

(mod p

2+i

), A

i

(n) ≡ (p − 1)/n

(i(p − 1)/n)! (mod p), where τ (χ) is the Gauss sum.

Put m = (p − 1)/2, and

G

j

(X) = A

0

(m)X

j

+ A

1

(m)X

j−1

+ . . . + A

j

(m), F

j

(X) = 1

(p − 1)! X

j

+ 1

(p + 1)! X

j−1

+ 1

(p + 3)! X

j−2

+ . . . + 1

(p + 2j − 1)! .

(3)

Define

E

n

= E

2n

(2n)! for n = 1, 2, 3, . . . ,

where E

2n

are the Euler numbers, i.e. E

0

= 1, E

2

= −1, E

4

= 5, E

6

= −61, E

8

= 1385, E

10

= −50521, E

12

= 2702765, E

14

= −199360981, . . .

Consider the formal expressions G

j

(E

) and F

j

(E

), where (E

)

k

= E

k

.

Let β

0

, β

1

, . . . , β

l−1

be the integral basis of the field K formed by the Gauss periods. Let

δ = x

0

β

0

+ x

1

β

1

+ . . . + x

l−1

β

l−1

. Associate with the unit δ the polynomial f (X) as follows:

f (X) = X

l−1

+ d

1

X

l−2

+ d

2

X

l−3

+ . . . + d

l−1

, where

d

i

= −lA

i

(l) x

0

+ x

1

g

i

+ x

2

g

2i

+ . . . + x

l−1

g

i(l−1)

x

0

+ x

1

+ . . . + x

l−1

for i = 1, . . . , l − 1. Put

S

j

= S

j

(d

1

, . . . , d

l−1

) = sum of the jth powers of the roots of f (X) for j = 1, . . . , 2l − 1. Hence

S

1

= −d

1

, S

2

= d

21

− 2d

2

, S

3

= −d

31

+ 3d

1

d

2

− 3d

3

, . . . Define the numbers T

1

, . . . , T

2l−1

as follows:

T

i

= − 1

(i(p − 1)/l)! 2

i(p−1)/l−1

(2

i(p−1)/l

−1)B

i(p−1)/l

−i p − 1

4l G

i(p−1)/(2l)

(E

) for i = 1, . . . , l − 1, and

T

l

= 1 − q

2

2 , where q

2

= 2

p−1

− 1

p ,

T

l+i

= − 1

(p − 1 + i(p − 1)/l)! 2

p−1+i(p−1)/l−1

(2

p−1+i(p−1)/l

− 1)

× B

(p−1+i(p−1)/l)

+

 p − 1

2 + i p − 1 2l



F

i(p−1)/(2l)

(E

) for i = 1, . . . , l − 1.

Define

α

i

= c

0

+ c

1

g

i

+ c

2

g

2i

+ . . . + c

l−2

g

(l−2)i

for i = 1, . . . , 2l − 1.

(4)

Let X

1

, . . . , X

2l−1

∈ Q and let

g(X) = X

2l−1

+ Y

1

X

2l−2

+ . . . + Y

2l−1

be a polynomial such that

X

j

= sum of the jth powers of the roots of g(X).

Define the mapping Φ : Q

2l−1

→ Q

l

as follows:

Φ(X

1

, . . . , X

2l−1

) = (1 − pY

l

, Y

1

− pY

l+1

, . . . , Y

l−1

− pY

2l−1

).

Now the main theorem of this paper be formulated.

Theorem 1. Let l and p be primes with p ≡ 1 (mod l) and let K ⊂ Q(ζ

p

+ ζ

p−1

) with [K : Q] = l. Suppose that 2 is not an l-th power mod- ulo p. Let δ be a unit of K such that [U

K

: hδi] = f , (f, p) = 1. Let η

2f

= δ

c0

σ(δ)

c1

. . . σ

l−2

(δ)

cl−2

and α

i

= c

0

+ c

1

g

i

+ . . . + c

l−2

g

(l−2)i

for i = 1, . . . , 2l − 1. The following congruence holds:

(3) ε

 x

0

+ x

1

+ . . . + x

l−1

−l



αl

Φ(α

1

S

1

, . . . , α

2l−1

S

2l−1

)

≡ (2 + 2p)

f (p−1)/(2l)

Φ(f T

1

, . . . , f T

2l−1

) (mod p

2

), where ε = ±1.

R e m a r k. The class number h

K

appears in the preceding congruence implicitly, via the congruence

h

K

f

l−2

≡ α

1

. . . α

l−1

(mod p

2

).

This congruence can be proved in the following way:

We have the congruence 1/g ≡ ζ

l

(mod p

p

) and hence σ

−1

(c

0

+ c

1

ζ

li

+ . . . + c

l−2

ζ

l(l−2)i

) ≡ α

i

(mod p

p

);

this yields

h

K

f

l−2

= e = N

Q(ζl)/Q

(c

0

+ c

1

ζ

l

+ . . . + c

l−2

ζ

ll−2

) ≡ α

1

. . . α

l−1

(mod p

2

).

The congruence (3) gives l congruences (one in each component). If B

(p−1)/l

B

2(p−1)/l

. . . B

(l−1)(p−1)/l

6≡ 0 (mod p),

then from the congruence (3), the numbers α

1

, . . . , α

l−1

modulo p

2

can be calculated. Using the congruence

h

K

f

l−2

≡ α

1

. . . α

l−1

(mod p

2

), also h

K

can be calculated modulo p

2

. If

B

(p−1)/l

B

2(p−1)/l

. . . B

(l−1)(p−1)/l

≡ 0 (mod p),

then the numbers α

1

, . . . , α

l−1

and hence also h

K

can be calculated at most

modulo p.

(5)

Before proving Theorem 1, we show its applications to quadratic and cubic fields.

The quadratic case K = Q(

p), p ≡ 5 (mod 8). Let δ = x

0

β

0

+ x

1

β

1

= x

0

−1 + p 2 + x

1

−1 − p 2 > 1 be a fundamental unit of Q(

p). Then

d

1

= −2A

1

(2) x

0

− x

1

x

0

+ x

1

. Hence

S

1

= 2A

1

(2) x

0

− x

1

x

0

+ x

1

, S

2

= 4A

1

(2)

2

(x

0

− x

1

)

2

(x

0

+ x

1

)

2

, S

3

= 8A

1

(2)

3

(x

0

− x

1

)

3

(x

0

+ x

1

)

3

. For A

1

(2) we have

τ (χ)

2

≡ 4A

1

(2)

2

(−p) (mod p

3

), A

1

(2) ≡ (p − 1)/2

((p − 1)/2)! (mod p).

Hence

A

1

(2)

2

≡ − 1

4 (mod p

2

), A

1

(2) ≡ (p − 1)/2

((p − 1)/2)! (mod p).

Therefore

S

1

= 2A

1

(2) x

0

− x

1

x

0

+ x

1

, S

2

= − (x

0

− x

1

)

2

(x

0

+ x

1

)

2

, S

3

= −2A

1

(2) (x

0

− x

1

)

3

(x

0

+ x

1

)

3

. Let

x

0

+ x

1

−2 + x

0

− x

1

2

p = T + U p > 1.

It can be proved that |η

2

| = |N

Q(ζ

pp−1)/Q(√

p)

p

+ ζ

p−1

)| < 1, hence it is necessary to start from the unit T − U

p. We have S

1

= 2A

1

(2) U

T , S

2

= − U

2

T

2

, S

3

= −2A

1

(2) U

3

T

3

. For the numbers T

1

, T

2

, T

3

we have

T

1

= − 1

((p − 1)/2)! 2

(p−1)/2−1

(2

(p−1)/2

− 1)B

(p−1)/2

p − 1

8 G

(p−1)/4

(E

), T

2

=

12

(1 − q

2

),

T

3

= − 1

(3(p − 1)/2)! 2

3(p−1)/2−1

(2

3(p−1)/2

− 1)B

3(p−1)/2

+ (3(p − 1)/4)F

(p−1)/4

(E

).

(6)

It is easy to see that Φ(X

1

, X

2

, X

3

) =



1 − p X

12

− X

2

2 , −X

1

− p



1

6 X

13

+ 1

2 X

1

X

2

1 3 X

3



. Hence

ε

 x

0

+ x

1

−2



α2

Φ(α

1

S

1

, α

2

S

2

, α

3

S

3

) ≡ (2+2p)

(p−1)/4

Φ(T

1

, T

2

, T

3

) (mod p

2

).

Since (x

0

+ x

1

)/(−2) = T and α

1

= α

2

= α

3

= h, we get

εT

h

Φ(hS

1

, hS

2

, hS

3

) ≡ (2 + 2p)

(p−1)/4

Φ(T

1

, T

2

, T

3

) (mod p

2

), where ε = ±1. It can be proved that ε = (−1)

1+r

, where r is the number of quadratic residues modulo p in the interval (p/4, p/2).

The cubic case. Let p be a prime such that p ≡ 1 (mod 3), p 6= a

2

+ 27b

2

and let

δ = x

0

β

0

+ x

1

β

1

+ x

2

β

2

. Then

d

1

= −3A

1

(3) x

0

+ x

1

g + x

2

g

2

x

0

+ x

1

+ x

2

, d

2

= −3A

2

(3) x

0

+ x

1

g

2

+ x

2

g x

0

+ x

1

+ x

2

. For the numbers A

1

(3), A

2

(3) we have

τ (χ)

3

≡ 27A

1

(3)

3

(−p) (mod p

3

), A

1

(3) ≡ (p − 1)/3

((p − 1)/3)! (mod p), τ (χ

2

)

3

≡ 27A

2

(3)

3

(−p)

2

(mod p

4

), A

2

(3) ≡ (p − 1)/3

(2(p − 1)/3)! (mod p).

As is well known, τ (χ)

3

= pJ(χ, χ), where J(χ, χ) is the Jacobi sum.

Let J(χ, χ) = a + bζ

3

, a ≡ −1, b ≡ 0 (mod 3) and p = a

2

− ab + b

2

. Hence

 a + b 1

g



≡ 27A

1

(3)

3

(mod p

2

), A

1

(3) ≡ (p − 1)/3

((p − 1)/3)! (mod p).

The number A

2

(3) is determined by the congruence

−1 ≡ 27

2

A

1

(3)

3

A

2

(3)

3

(mod p

2

), A

2

(3) ≡ (p − 1)/3

(2(p − 1)/3)! (mod p).

For the numbers T

1

, . . . , T

5

we have T

i

= − 1

(i(p − 1)/3)! 2

i(p−1)/3−1

(2

i(p−1)/3

−1)B

i(p−1)/3

−i p − 1

12 G

i(p−1)/6

(E

) for i = 1, 2, and

T

3

=

12

(1 − q

2

),

(7)

T

3+i

= − 1

(p − 1 + i(p − 1)/3)! 2

p−1+i(p−1)/3−1

(2

p−1+i(p−1)/3

− 1)

× B

(p−1+i(p−1)/3)

+

 p − 1

2 + i p − 1 6



F

i(p−1)/6

(E

) for i = 1, 2.

It is easy to prove that for Φ(X

1

, . . . , X

5

) we have Φ(X

1

, . . . , X

5

)

=

 1 − p



1

6 X

13

+ 1

2 X

1

X

2

1 3 X

3

 ,

− X

1

− p

 1

24 X

14

+ 1

3 X

1

X

3

+ 1

8 X

22

1

4 X

12

X

2

1 4 X

4

 , 1

2 (X

12

− X

2

) − p



1

120 X

15

+ 1

4 X

1

X

4

+ 1 6 X

2

X

3

+ 1

12 X

13

X

2

1

6 X

12

X

3

1

8 X

1

X

22

1 5 X

5



. Hence

±

 −1 3



α3

(x

0

+ x

1

+ x

2

)

α3

Φ(α

1

S

1

, . . . , α

5

S

5

)

≡ (2 + 2p)

(p−1)/6

Φ(T

1

, . . . , T

5

) (mod p

2

).

P r o o f o f T h e o r e m 1. Let g

1

(X), . . . , g

r

(X) be polynomials such that g

i

(X) 6≡ 0 (mod X). Let

g(X) ≡ g

1

(X)

b1

. . . g

r

(X)

br

(mod X

M

).

Let s

j

be the homomorphism defined in [4]. We have

s

j

(g(X)) = b

1

s

j

(g

1

(X)) + b

2

s

j

(g

2

(X)) + . . . + b

r

s

j

(g

r

(X)), for j = 1, . . . , M − 1. Define

X

j

= b

1

s

j

(g

1

(X)) + b

2

s

j

(g

2

(X)) + . . . + b

r

s

j

(g

r

(X)) for j = 1, . . . , M − 1. Let

g(X) ≡ C

0

+ C

1

X + C

2

X

2

+ . . . + C

M −1

X

M −1

(mod X

M

).

Consider the reciprocal polynomial F (X) = X

M −1

+ C

1

C

0

X

M −2

+ . . . + C

M −1

C

0

.

(8)

By the definition of the homomorphism s

j

(see [4]) we have X

j

= sum of the jth powers of the roots of F (X).

The numbers C

1

/C

0

, C

2

/C

0

, . . . , C

M −1

/C

0

can be calculated by the Newton recurrence formula.

According to [2] and [4] we have:

Proposition 1. There is a number π ∈ Q(ζ

p

+ ζ

p−1

), π | p such that (i) N

Q(ζp−1

p )/Q

(π) = (−1)

m

p,

(ii) σ(π) ≡ gπ (mod π

2m+1

), g ≡ a

2p

(mod p

2

), (iii) ζ

p

+ ζ

p−1

P

2m

i=0

a

i

π

i

(mod π

2m+1

), where 0 ≤ a

i

< p and a

i

(2/(2i)!) (mod p) for i = 1, . . . , m.

The numbers a

m+i

for i = 1, . . . , m, are defined by a

m+1

= 2 p − 1 − p(p + 1)B

p−1

p .

If 2 is a primitive root modulo p then the coefficients a

m+2

, a

m+3

, . . . , a

2m

are given by the recurrence formula a

m+1+s

1

4

s+1

− 4

 4

p(s+1)

a

s+1

− b

s+1

p + b

m+s+1



(mod p), where b

s+1

and b

m+1+s

are the coefficients of X

s+1

and X

m+1+s

, respec- tively, in the polynomial

 2 + 2

2! X + . . . + 2

(p − 1)! X

m

+ a

m+1

X

m+1

+ . . . + a

m+s

X

m+s



2

. Proposition 2. Let K ⊂ Q(ζ

p

+ ζ

p−1

) with [K : Q] = n. There is a number π ∈ K with π | p such that

(i) N

K/Q

(π) = (−1)

n

p,

(ii) σ(π) ≡ gπ (mod π

n+1

), g ≡ a

p(p−1)/n

(mod p

2

), (iii) β

0

P

2n

i=0

a

i

π

i

(mod π

2n+1

), where 0 ≤ a

i

< p and where a

i

≡ ((p − 1)/n)/(i(p − 1)/n)! (mod p) for i = 1, . . . , n.

From now on we will use the following transformation. Let γ ≡ c

0

+ c

1

π + c

2

π

2

+ . . . + c

2l−1

π

2l−1

(mod π

2l

).

From π

l

≡ −p (mod π

2l

) we have

γ ≡ (c

0

− pc

l

) + (c

1

− pc

l+1

)π + . . . + (c

l−1

− pc

2l−1

l−1

(mod π

2l

).

Lemma 1. Let α, β ∈ K with αβ 6≡ 0 (mod π). Let

α ≡ e

0

+ e

1

π + . . . + e

l−1

π

l−1

(mod π

2l

),

β ≡ b

0

+ b

1

π + . . . + b

l−1

π

l−1

(mod π

2l

).

(9)

Relate to the conjugation σ

i

(α) the polynomial

f

i

(X) = e

0

+ e

1

g

i

X + e

2

g

2i

X

2

+ . . . + e

l−1

g

i(l−1)

X

l−1

. Let

F (X) ≡ f

0

(X)

c0

f

1

(X)

c1

. . . f

l−2

(X)

cl−2

≡ d

0

+ d

1

X + . . . + d

2l−1

X

2l−1

(mod X

2l

).

Define X

i

= s

i

(F (X)). Then

α

c0

σ(α)

c1

. . . σ

l−2

(α)

cl−2

≡ β (mod p

2

) if and only if

d

0

Φ(X

1

, . . . , X

2l−2

) ≡ (b

0

, b

1

, . . . , b

l−1

) (mod p

2

).

P r o o f. Clearly

α

c0

σ(α)

c1

. . . σ

l−2

(α)

cl−2

≡ d

0

+ d

1

π + . . . + d

2l−1

π

2l−1

(mod π

2l

).

The coefficients d

1

/d

0

, d

2

/d

0

, . . . , d

2l−1

/d

0

can be expressed using the num- bers X

1

, . . . , X

2l−1

by the Newton recurrence formula. Clearly

d

0

+ d

1

π + . . . + d

2l−1

π

2l−1

≡ (d

0

− pd

l

) + (d

1

− pd

l+1

)π + . . . + (d

l−1

− pd

2l−1

l−1

(mod π

2l

).

From the definition of the mapping Φ it follows that

(d

0

−pd

l

, d

1

−pd

l+1

, . . . , d

l−1

−pd

2l−1

) ≡ d

0

Φ(X

1

, X

2

, . . . , X

2l−1

) (mod p

2

).

Lemma 1 is proved.

Proposition 2 gives

β

0

≡ −1/l + (a

1

− pa

l+1

)π + . . . + (a

l−1

− pa

2l−1

l−1

(mod π

2l

), where 0 ≤ a

i

< p and a

i

≡ ((p − 1)/n)/(i(p − 1)/n)! (mod p) for i = 1, . . . , l − 1. According to [4] we have

a

i

− pa

l+i

≡ A

i

(l) (mod p

2

).

Let δ = x

0

β

0

+ x

1

β

1

+ . . . + x

l−1

β

l−1

. Then δ ≡ − 1

l (x

0

+ x

1

+ . . . + x

l−1

) +

X

l−1 i=1

A

i

(l)(x

0

+ x

1

g

i

+ x

2

g

2i

+ . . . + x

l−1

g

i(l−1)

i

(mod π

2l

).

Let

δ

c0

σ(δ)

c1

. . . σ

l−2

(δ)

cl−2

= η

2f

.

Then X

j

corresponding to the product on the left-hand side is equal to

c

0

s

j

(δ) + c

1

g

j

s

j

(δ) + . . . + c

l−2

g

j(l−2)

s

j

(δ) = α

j

S

j

(10)

for j = 1, . . . , 2l − 1. Hence the number corresponding to the left-hand side is 

x

0

+ x

1

+ . . . + x

l−1

−l



αl

Φ(α

1

S

1

, . . . , α

2l−1

S

2l−1

).

Now we prove that the right-hand side is equal to (2 + 2p)

f (p−1)/l

Φ(f T

1

, . . . , f T

2l−1

).

By Proposition 1 we have ζ

p

+ ζ

p−1

X

2m i=0

a

i

π

i

(mod π

2m+1

),

where 0 ≤ a

i

< p and a

i

≡ (2/(2i)!) (mod p) for i = 1, . . . , m, and hence ζ

p

+ ζ

p−1

≡ A

0

(m) + A

1

(m)π + . . . + A

m−1

(m)π

m−1

(mod π

2m

).

Consider the polynomial

g(X) = A

0

(m)X

m−1

+ A

1

(m)X

m−2

+ . . . + A

m−1

(m).

Now we shall calculate the numbers

s

i

= sum of the ith powers of the roots of g(X)

for i = 1, . . . , 2m − 1 modulo p

2

. It is easy to see that for i > m − 1 it is enough to determine s

i

modulo p. Let W

1

, W

2

, . . . be a linearly recurrent sequence modulo p of order m − 1 defined by

W

i

= −1

(2i)! 2

2i−1

(2

2i

− 1)B

2i

for i = 1, . . . , m − 1.

For i > m − 1 we have W

i

= −

 1

2! W

i−1

+ 1

4! W

i−2

+ . . . + 1

(2m − 2)! W

i−m+1

 . Lemma 2. The following congruence holds:

s

n+1

−1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

n + 1

2 G

n+1

(E

) (mod p

2

) for n + 1 < m.

P r o o f. It is easy to see that A

0

(m) ≡ 2 + 2p (mod p

2

). Clearly (2 + 2p) 1 − p

2 ≡ 1 (mod p

2

).

Consider the polynomial

g

(X) = X

m−1

+ C

1

X

m−2

+ . . . + C

m−1

, where

C

i

1 − p

2 A

i

(m) (mod p

2

).

(11)

Obviously s

n

is equal to the sum of the nth powers of the roots of g

(X).

Write

C

1

= c

1

+ b

1

p, C

2

= c

2

+ b

2

p, . . . , C

m−1

= c

m−1

+ b

m−1

p.

It is known that there exists a polynomial f

n

(x

1

, . . . , x

n

) such that s

n

= f

n

(c

1

+ b

1

p, . . . , c

n

+ b

n

p).

Write s

n

= r

n

+ t

n

p. Clearly r

1

+ t

1

p = −(c

1

+ b

1

p). Put t

1

= −b

1

. According to the Newton recurrent formula we have

r

2

+ t

2

p + (c

1

+ b

1

p)(r

1

+ t

1

p) + 2(c

2

+ pb

2

) = 0.

The last equation can be rewritten in the form

r

2

+ c

1

r

1

+ 2c

2

+ p(t

2

+ c

1

t

1

+ r

1

b

1

+ 2b

2

) = 0.

Define F

2

= r

1

b

1

+ 2b

2

. Hence r

2

+ c

1

r

1

+ 2c

2

+ p(t

2

+ c

1

t

1

+ F

2

) = 0. Put t

2

= −(c

1

t

1

+ F

2

).

Further we have

r

3

+ t

3

p + (c

1

+ b

1

p)(r

2

+ t

2

p) + (c

2

+ b

2

p)(r

1

+ t

1

p) + 3(c

3

+ b

3

p) = 0, hence

r

3

+ c

1

r

2

+ c

2

r

1

+ 3c

3

+ p(t

3

+ c

1

t

2

+ c

2

t

1

+ b

1

r

2

+ b

2

r

1

+ 3b

3

) = 0.

Define F

3

= b

1

r

2

+ b

2

r

1

+ 3b

3

. Then we put t

3

= t

1

(c

21

− c

2

) + c

1

F

2

− F

3

,

t

4

= t

1

(−c

31

+ 2c

1

c

2

− c

3

) − (c

21

− c

2

)F

2

+ c

1

F

3

− F

4

,

t

5

= t

1

(c

41

− 3c

21

c

2

+ 2c

1

c

3

+ c

22

− c

4

) − (−c

31

+ 2c

1

c

2

− c

3

)F

2

− (c

21

− c

2

)F

3

+ c

1

F

4

− F

5

. Consider the numbers

K

2

= −c

1

, K

4

= c

21

− c

2

, K

6

= −c

31

+ 2c

1

c

2

− c

3

, . . .

The numbers c

1

, c

2

, c

3

, . . . are equal to 1/2!, 1/4!, 1/6!, . . . modulo p. Define K

2n

= K

2n

/(2n)!.

Then

K

2n

(2n)! + 1

(2)! · K

2n−2

(2n − 2)! + . . . + 1

(2n − 2)! · K

2

2! + 1

(2n)! = 0, hence

1 (2n)!

 K

2n

+

 2n 2



K

2n−2

+ . . . + 1



= 0,

and therefore K

2n

= E

2n

, the Euler number. Using induction by n we put (4) t

n+1

= t

1

E

2n

(2n)! − F

2

E

2n−2

(2n − 2)! − . . . − F

n+1

,

(12)

where

F

2

= b

1

r

1

+ 2b

2

, F

3

= b

1

r

2

+ b

2

r

1

+ 3b

3

, F

4

= b

1

r

3

+ b

2

r

2

+ b

3

r

1

+ 4b

4

, . . . Since t

1

= −b

1

, we have

t

n+1

= − b

1

E

2n

(2n)! E

2n−2

(2n − 2)! (b

1

r

1

+ 2b

2

)

− . . . − (b

1

r

n

+ b

2

r

n−1

+ . . . + (n + 1)b

n+1

).

In the last equation we first cancel the brackets by multiplication and then make new brackets by factoring out b

1

, . . . , b

n+1

. The summand obtained by factoring out b

1

is

(5) b

1

 E

2n

(2n)! + r

1

E

2n−2

(2n − 2)! + . . . + r

n

 . According to [4, Lemma 3],

r

i

≡ − 1

(2i)! 2

2i−1

(2

2i

− 1)B

2i

(mod p).

By substitution into (5) we get the sum E

2n

(2n)! E

2n−2

(2n − 2)! · 1

2! 2(2

2

− 1)B

2

E

2n−4

(2n − 4)! · 1

4! 2

3

(2

4

− 1)B

4

− . . . − 1

(2n)! 2

2n−1

(2

2n

− 1)B

2n

. The following identity holds:

(6) E

2n

(2n)! E

2n−2

(2n − 2)! · 1

2! 2(2

2

− 1)B

2

E

2n−4

(2n − 4)! · 1

4! 2

3

(2

4

− 1)B

4

− . . . − 1

(2n)! 2

2n−1

(2

2n

− 1)B

2n

= (n + 1) E

2n

(2n)! . This identity can be proved in the following way. For the functions sec x and tan x, we have

sec x = 1 − E

2

2! x

2

+ E

4

4! x

4

E

6

6! x

6

+ . . . , tan x = 2

2

(2

2

− 1)B

2

x

2! − 2

4

(2

4

− 1)B

4

x

3

4! + . . . Then

tan x · sec x = 2

2

(2

2

− 1) 2! B

2

x −

 E

2

B

2

2

2

(2

2

− 1)

2!2! + 2

4

(2

4

− 1)B

4

4!



x

3

+ . . . The identity (6) follows from the equation

d(sec x)

dx = tan x · sec x.

(13)

Using the identity (6), by induction (n + 1 → 1) we have (7) t

n+1

= −(n + 1)

 b

1

E

2n

(2n)! + b

2

E

2n−2

(2n − 2)! + . . . + b

n+1

 . The numbers a

1

, . . . , a

m−1

from Proposition 1 are

a

i

2

(2i)! (mod p), 0 < a

i

< p.

Define the numbers D

i

as follows:

1 − p

2 a

i

+ pD

i

1

(2i)! (mod p

2

), 0 ≤ D

i

< p, hence

D

i

1 p

 1

(2i)! 1 − p 2 a

i



(mod p).

Let v

n+1

be the sum of the (n+1)th powers of the roots of the polynomial X

m−1

+ 1 − p

2 a

1

X

m−2

+ . . . + 1 − p 2 a

m−1

.

According to [4, Lemma 3], the sum of the (n + 1)th powers of the roots of the polynomial

X

m−1

+ 1

2! X

m−2

+ 1

4! X

m−3

+ . . . + 1 (2m − 2)! , is equal to

−1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

. Hence

v

n+1

= −1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

+ p(n + 1)



D

1

E

2n

(2n)! + D

2

E

2n−2

(2n − 2)! + . . . + D

n+1

 . Therefore

s

n+1

= −1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

(8)

+ p(n + 1)



D

1

E

2n

(2n)! + D

2

E

2n−2

(2n − 2)! + . . . + D

n+1



− p(n + 1)

 b

1

E

2n

(2n)! + b

2

E

2n−2

(2n − 2)! + . . . + b

n+1

 . Since

D

i

1 p

 1

(2i)! 1 − p 2 a

i



,

(14)

from (8) we get s

n+1

= −1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

(n + 1)E

2n+2

(2n + 2)!

(9)

+ (n + 1) p − 1 2

 a

1

E

2n

(2n)! + a

2

E

2n−2

(2n − 2)! + . . . + a

n+1



+ p n + 1 2



a

m+1

E

2n

(2n)! + a

m+2

E

2n−2

(2n − 2)! + . . . + a

m+n+1

 , where a

1

, . . . , a

m

, a

m+1

, a

m+2

, . . . are the numbers from Proposition 1. Since A

i

(m) = a

i

− pa

m+i

, from the equality (9) we get

s

n+1

= −1

(2n + 2)! 2

2n+1

(2

2n+2

− 1)B

2n+2

n + 1

2 G

n+1

(E

).

Lemma 3. The following congruence holds:

W

m+j

≡ − 1

(p − 1 + 2j)! 2

p−1+2j−1

(2

p−1+2j

− 1)B

p−1+2j

+ (m + j)F

j

(E

) (mod p) for j = 1, . . . , m − 1.

P r o o f. Let n > 2p − 3. Consider the polynomial h(X) = X

n

+ 1

2! X

n−1

+ 1

4! X

n−2

+ . . . + 1 (2n)! .

Let W

i

be the sum of the ith powers of the roots of h(X). By [4, Lemma 3], W

i

= − 1

(2i)! 2

2i−1

(2

2i

− 1)B

2i

. The characteristic polynomial of the sequence W

i

is

X

m−1

+ 1

2! X

m−2

+ 1

4! X

m−3

+ . . . + 1 (2m − 2)! .

Hence W

i

is the sum of the ith powers of the roots of this polynomial.

Let

c

1

= 1

2! , c

2

= 1 4! , . . . Obviously

W

m

= W

m

+ mc

m

,

W

m+1

= W

m+1

+ c

m

W

1

+ (m + 1)c

m+1

− c

1

mc

m

. We can write

G

m

= mc

m

, G

m+1

= (m + 1)c

m+1

− c

1

G

m

.

(15)

By induction we can prove that

W

m+j

= W

m+j

+ c

m

W

j

+ (c

m+1

+ c

m

E

1

)W

j−1

(10)

+ (c

m+2

+ c

m+1

E

1

+ c

m

E

2

)W

j−2

+ . . .

+ (c

m+j−1

+ c

m+j−2

E

1

+ . . . + c

m

E

j−1

)W

1

+ G

m+j

. Now we rewrite the right side of (10). We cancel the brackets by multi- plication and make new brackets by factoring out c

m

, c

m+1

, . . .

E.g., the summand obtained by factoring out c

m

is c

m

(W

j

+ E

1

W

j−1

+ E

2

W

j−2

+ . . . + E

j−1

W

1

).

From (6) we get

c

m

(W

j

+ E

1

W

j−1

+ E

2

W

j−2

+ . . . + E

j−1

W

1

) = c

m

jE

j

. By repeating this procedure with each summand we have

(11) W

m+j

= W

m+j

+ jc

m

E

j

+ (j − 1)c

m+1

E

j−1

+ . . . + c

m+j−1

E

1

+ G

m+j

. By induction it is easy to prove

(12) G

m+j

= (m + 1)c

m+j

+ (m + j − 1)c

m+j−1

E

1

+ . . . + mc

m

E

j

. Substituting (12) into (11) we get

W

m+j

= W

m+j

+ (m + j)(c

m+j

+ c

m+j−1

E

1

+ . . . + c

m

E

j

).

Since η

2

= N

Q(ζ

p−1p )/K

p

+ ζ

p−1

) we have η

f2

≡ r

0

+

X

l−1 i=1

r

i(p−1)/(2l)

π

i(p−1)/(2l)

(mod π

2m

),

where r

0

= (2 + 2p)

f (p−1)/(2l)

. From Lemmas 1–3 we deduce that the right- hand side is equal to

(2 + 2p)

f (p−1)/(2l)

Φ(f T

1

, . . . , f T

2l−1

).

It remains to prove that T

l

= (1 − q

2

)/2. From the proof of Lemma 3 we have

T

l

−1

(p − 1)! 2

p−2

(2

p−1

− 1)B

p−1

+ p − 1

2 · 1

(p − 1)! , hence

T

l

−1

(p − 1)! 2

p−2

2

p−1

− 1

p pB

p−1

+ p − 1

2 · 1

(p − 1)! .

From pB

p−1

≡ −1 (mod p) the required congruence follows. Theorem 1 is

proved.

(16)

Example 1 (p = 13, l = 3, K ⊂ Q(ζ

13

+ ζ

13−1

), [K : Q] = 3). By Proposition 1 we have

ζ

13

+ ζ

13−1

≡ 2 + π + 12π

2

+ 3π

3

+ 4π

4

+ 9π

5

+ 11π

6

+ 2π

7

+ 12π

8

+ 8π

9

+ 10π

10

+ 4π

11

(mod π

12

), hence

ζ

13

+ ζ

13−1

≡ A

0

(m) + A

1

(m)π + . . . + A

5

(m)π

5

≡ 28 + 144π + 25π

2

+ 68π

3

+ 43π

4

+ 126π

5

(mod π

12

).

It is easy to see that η

2

= N

Q(ζ

pp−1)/K

p

+ ζ

p−1

)

≡ (28 + 144π + 25π

2

+ 68π

3

+ 43π

4

+ 126π

5

)

× (28 − 144π + 25π

2

− 68π

3

+ 43π

4

− 126π

5

)

≡ 56 + 86π

2

+ 50π

4

(mod π

12

).

The number β

0

= ζ

13

+ ζ

135

+ ζ

138

+ ζ

1312

is the Gauss period. By Proposition 2 we have

β

0

≡ −

13

+ A

1

(3)π + A

2

(3)π

2

(mod π

6

)

(π is a prime divisor of the field K, [K : Q] = 3, π | p). For A

1

(3), A

2

(3) we have

τ (χ)

3

≡ 27A

1

(3)

3

(−13) (mod p

3

), A

1

(3) ≡ 4

4! (mod 13),

−1 ≡ 27

2

A

1

(3)

3

A

2

(3)

3

(mod 169), A

2

(3) ≡ 4

8! (mod 13).

The number g satisfies g ≡ 2

4·13

≡ 146 (mod 169), and τ (χ)

3

= pJ(χ, χ), J(χ, χ) = −4 − 3ζ

3

. Hence A

1

(3) ≡ 4/4! ≡ 11 (mod 13). Thus



− 4 − 3 1 146



≡ 27(11 + 13k)

3

(mod 169).

It follows that A

1

(3) = 50, A

2

(3) = 86.

The fundamental unit of the field K is δ = β

2

. Hence

d

1

= −3 · 50 · 146

2

≡ 80 (mod 169), d

2

= −3 · 86 · 146 ≡ 19 (mod 169).

Therefore

S

1

= −80, S

2

= 109, S

3

= 2, S

4

= 5, S

5

= 4.

Cytaty

Powiązane dokumenty

Magnetoelastic properties, magnetic anisotropy and magnetic damping properties of several series of quaternary Co 2 YZ epitaxially grown thin films of Heusler alloys, including Co 2

A prime number is a natural number greater than 1, which cannot be written as a product of two smaller natural numbers.. Equivalent definition:

The only quantity of interest here is the magnitude of the total energy functional and its components, e.g. the term describing van der Waals interactions. Static calculations

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

The new tool here is an improved version of a result about enumerating certain lattice points due to E.. A result about enumerating certain

the numerical value of B, Theorem 0 qualitatively settles Baker’s problem on the bound for small prime solutions of the equation (1.1).. Therefore, it remains to estimate the infimum

We see that the λ-root is an extension of the primitive root to all moduli, and we extend the notation g ∗ (q) to mean the least prime λ-root (mod q)..

[r]