• Nie Znaleziono Wyników

of characteristic zero

N/A
N/A
Protected

Academic year: 2021

Share "of characteristic zero"

Copied!
10
0
0

Pełen tekst

(1)

Polynomial imaginary decompositions for finite extensions of fields

of characteristic zero

Andrzej Nowicki and Stanis law Spodzieja

December 27, 2001

Abstract

Let k be a field of characteristic zero, L = k[ξ] a finite field ex- tension of degree m > 1, and f (z) a polynomial in one variable z over L. Then there exist unique polynomials u0, . . . , um−1 belong- ing to k[x0, . . . , xm−1] such that f (x0 + ξx1 + · · · + ξm−1xm−1) = u0+ ξu1+ · · · + ξm−1um−1. We prove that if f (z) 6∈ L, then the poly- nomials u0, . . . , um1 are algebraically independent over k and they have no common divisors in k[x0, . . . , xm−1] of positive degrees. Some other properties of polynomials u0, . . . , um−1 are also given.

1 Introduction

If z, x, y are variables and f (z) is a polynomial from C[z], then there exist unique polynomials u(x, y), v(x, y) belonging to R[x, y] such that f (x + iy) = u(x, y) + iv(x, y). We will show that if f is nonzero, then the polynomials u(x, y) and v(x, y) are coprime. We will also show that the same is true if instead of the extension R ⊂ C we consider a finite field extension of characteristic zero.

More exactly, assume that k is a field and L = k[ξ] is a finite field exten- sion of degree m > 1. Let z and x0, . . . , xn−1 be variables and let f (z) be a polynomial from L[z]. Then there exist unique polynomials u0, . . . , um−1, belonging to k[x] := k[x0, . . . , xm−1], such that

f (x0+ ξx1+ · · · + ξm−1xm−1) = u0+ ξu1 + · · · + ξm−1um−1.

Supported by KBN Grant 2 PO3A 007 18

(2)

This representation we call the imaginary decomposition of f , and the poly- nomials u0, . . . , um−1 we call the imaginary parts of f . We will show that if f is nonzero, then the imaginary parts of f have no common divisors in k[x] of positive degrees. Moreover, we prove that a sequence (u0, . . . , um−1) of poly- nomials from k[x] forms imaginary parts of a polynomial f (z) ∈ L[z] if and only if the polynomials u0, . . . , um−1 satisfy some generalizations of Cauchy - Riemann equations. Some other properties concerning the divisibility of imaginary parts are also given.

2 Notations and preliminaries

Throughout the paper k is a field of characteristic zero and L = k[ξ] is a finite field extension of degree m > 1. Assume that

ϕ(t) = tm− am−1tm−1 − · · · − a1t − a0

(with a0, . . . , am−1 ∈ k) is the minimal polynomial of ξ over k. Let x = (x0, x1, . . . , xm−1), where x0, . . . , xm−1are variables, and let k[x] := k[x0, . . . , xm−1], L[x] := L[x0, . . . , xm−1] be the polynomial rings. We denote by M the set k[X]m, that is,

M := {(u0, u1, . . . , um−1); u0, . . . , um−1 ∈ k[x]} .

Let u = (u0, . . . , um−1) ∈ M . We use the following notations. We denote by u = (u0, . . . , um−1) the element from M defined by:













u0 = a0um−1, u1 = u0+ a1um−1, u2 = u1+ a2um−1,

...

um−1 = um−2+ am−1um−1.

Moreover, we denote by [u] the polynomial from L[x] defined as [u] := u0+ ξu1+ · · · + ξm−1um−1.

In particular, [x] := x0+ξx1+· · ·+ξm−1xm−1. If the polynomials u0, . . . , um−1 are imaginary parts of a polynomial f (z) ∈ L[z], then f ([x]) = [u]. Observe that the equality ξm = a0+ a1ξ + · · · + am−1ξm−1 implies that [u]ξ = [u] .

If i ∈ {0, . . . , m − 1}, then we denote by ∂x∂u

i the element from M defined as

∂u

∂xi :=

∂u0

∂xi, . . . , ∂u∂xm−1

i

 .

(3)

Lemma 2.1. If u0, . . . , um−1 are imaginary parts of a polynomial f (z) ∈ L[z], then

h∂u

∂xi

i

= f0([x])ξi

for i ∈ {0, . . . , m − 1}, where f0(z) means the ordinary derivative of f with respect to z.

Proof. h

∂u

∂xi

i

= ∂x

i[u] = ∂x

if ([x]) = f0([x])∂x

i[x] = f0([x])ξi.  As a consequence of this lemma we obtain the following proposition.

Proposition 2.2. If u0, . . . , um−1 ∈ k[x] are imaginary parts of a polynomial f (z) ∈ L[z] r L, then u0, . . . , um−1 are algebraically independent over k.

Proof. Assume that f ([x]) = [u], where u := (u0, . . . , um−1) and f (z) ∈ L[z] r L. Suppose that u0, . . . , um−1 are algebraically dependent over k.

Then the vectors ∂x∂u

0, . . . ,∂x∂u

m−1 are linearly dependent over the field k(x) :=

k(x0, . . . , xm−1). Hence, there exist a nonzero sequence α = (α0, . . . , αm−1) of polynomials from k[x] such thatPm−1

i=0 αi∂x∂u

i = 0, that is,Pm−1 i=0 αih

∂u

∂xi

i

= 0.

Now, by Lemma 2.1 we get:

0 =

m−1

P

i=0

αih

∂u

∂xi

i

=

m−1

P

i=0

αif0([x])ξi = f0([x])

m−1 P

i=0

αiξi



= f0([x])[α].

Hence, in the polynomial ring L[x] we have the equality f0([x])[α] = 0. But f0([x]) 6= 0 and [α] 6= 0. So, we have a contradiction. 

3 Generalization of the Cauchy-Riemann equa- tion

We introduce the following generalization of the Cauchy - Riemann equa- tion.

Definition 3.1. Let u ∈ M . We say that u is a ξ-sequence if

∂u

∂xi = ∂u

∂xi−1, for all i = 1, 2, . . . , m − 1.

The following proposition show that the imaginary parts of a polynomial from L[z] form a ξ-sequence.

(4)

Proposition 3.2. Let f (z) ∈ L[z] and let u be the element from M such that

f ([x]) = [u].

Then u is a ξ-sequence.

Proof. If i ∈ {1, . . . , m − 1} then, by Lemma 2.1, we have:

h∂u

∂xi

i

= f0([x])ξi = f0([x])ξi−1· ξ = h

∂u

∂xi−1

i· ξ = h

∂u

∂xi−1

i

=h

∂u

∂xi−1

i , and hence ∂x∂u

i = ∂x∂u

i−1. 

We will show that the converse of the above proposition is also true. For a proof of this fact we need several lemmas.

Lemma 3.3. If u ∈ M is a ξ-sequence, then each partial derivative ∂x∂u

j, for j = 0, . . . , m − 1, is also a ξ-sequence.

Proof. Let j ∈ {0, . . . , m − 1} and put w := ∂x∂u

j =

∂u0

∂xj, . . . , ∂u∂xm−1

j

 . Then for every i ∈ {1, . . . , m − 1} we have:

∂w

∂xi = ∂x2u

i∂xj = ∂x

j

∂u

∂xi = ∂x

j

∂u

∂xi−1 = ∂x

i−1

∂u

∂xj = ∂x∂w

i−1, and hence, w is a ξ-sequence. 

Lemma 3.4. If u ∈ M is a ξ-sequence, then h∂u

∂xi

i ξ =h

∂u

∂xi+1

i , for i = 0, 1, . . . , m − 2. In particular,

h∂u

∂x0

i ξ =h

∂u

∂x1

i , h

∂u

∂x0

i

ξ2 =h

∂u

∂x2

i

, . . . , h

∂u

∂x0

i

ξm−1 =h

∂u

∂xm−1

i . Proof. h

∂u

∂xi

i ξ =h

∂u

∂xi

i

=h

∂u

∂xi

i

=h

∂u

∂xi+1

i .  Consider the usual gradation k[x] = L

s>0k[x]s, where each k[x]s is the subgroup of k[x] containing zero and all homogeneous polynomials from k[x]

of degree s. We say that an element u = (u0, . . . , um−1) from M is homoge- neous of degree s, if all the polynomials u0, . . . , um−1 belong to k[x]s.

Every polynomial h ∈ k[x] has a presentation of the form h = h(0) + h(1) + · · · + h(r), where each h(j), for j = 0, . . . , r, is a unique homogeneous polynomial belonging to k[x]j.

(5)

Let u = (u0, . . . , um−1) ∈ M . Then there exists a common integer r > 0 such that

ui = u(0)i + u(1)i + · · · + u(r)i , for all i = 0, . . . , m − 1, and we have

u = u(0)+ u(1)+ · · · + u(r), where u(j) = 

u(j)0 , u(j)1 , . . . , u(j)m−1

for j = 0, 1, . . . , r. Then each u(j), for j = 0, . . . , r, is a homogeneous element of M of degree j. We call it the homogeneous component of u of degree j. Since

∂xi (k[x]s) ⊆ k[x]s−1,

for every s > 0 and i = 0, 1, . . . , m − 1 (where k[x]−1 = 0), we obtain the following lemma.

Lemma 3.5. Let u ∈ M . If u is a ξ-sequence, then each homogeneous component of u is also a ξ-sequence. 

Note also:

Lemma 3.6. Let u be a homogeneous element of M of degree s > 0. If u is a ξ-sequence, then

[x]h

∂u

∂x0

i

= s[u].

Proof. As a consequence of Lemma 3.4 we have:

[x]h

∂u

∂x0

i

= (x0+ ξx1+ ξ2x2+ · · · + ξm−1xm−1)h

∂u

∂x0

i

= x0h

∂u

∂x0

i

+ x1h

∂u

∂x0

i

ξ + x2h

∂u

∂x0

i

ξ2+ · · · + xm−1h

∂u

∂x0

i ξm−1

= x0h

∂u

∂x0

i

+ x1h

∂u

∂x1

i

+ x2h

∂u

∂x2

i

+ · · · + xm−1h

∂u

∂xm−1

i

= h

x0 ∂u

∂x0 + x1 ∂u

∂x1 + x2 ∂u

∂x2 + · · · + xm−1 ∂u

∂xm−1

i

= [su] = s[u].

We used also the well-known Euler equality. 

Lemma 3.7. Let u be a homogeneous element of M of degree s > 0. If u is a ξ-sequence, then there exists a unique b ∈ L such that [u] = b[x]s.

(6)

Proof. We use an induction with respect to s. It is obvious for s = 0.

Let s > 0 and assume that it is true for all homogeneous ξ-sequences of degree s − 1.

Let u be a homogeneous ξ-sequence of degree s. Then, by Lemma 3.3, the partial derivative ∂x∂u

0 is a homogeneous ξ-sequence of degree s − 1 and hence, by induction, there exists an element c ∈ L such that

h∂u

∂x0

i

= c[x]s−1. Put b := 1sc. Then, by Lemma 3.6, we have:

b[x]s = 1sc[x]s−1[x] = 1sh

∂u

∂x0

i

[x] = 1ss[u] = [u].

The uniqueness is obvious. 

Now we are ready to prove the following theorem.

Theorem 3.8. Let k be a field of characteristic zero and let L = k[ξ] be a finite field extension of degree m > 1. Let z, x0, x1, . . . , xm−1 be vari- ables and let u = (u0, . . . , um−) be a sequence of polynomials belonging to k[x] := k[x0, . . . , xm−1]. Then the following two conditions are equivalent.

(1) u is a ξ-sequence.

(2) There exists a polynomial f (z) ∈ L[z] such that the polynomials u0, . . . , um−1 are the imaginary parts of f (z), that is,

f (x0+ ξx1+ · · · + ξm−1xm−1) = u0+ ξu1 + · · · + ξm−1um−1.

Proof. The implication (2) ⇒ (1) we already proved (see Proposition 3.2). Assume now that u is a ξ-sequence. Let u = u(0)+ u(1)+ · · · + u(r) be the homogeneous decomposition of u. Then each u(j), for j = 0, . . . , r, is (by Lemma 3.5) a homogeneous ξ-sequence of degree j and so, by Lemma 3.7, there exists bj ∈ L such thatu(j) = bj[x]j. Put f (z) := b0+ b1z + · · · + brzr. Then

f ([x]) = b0[x]0+ b1[x]1+ · · · + br[x]r

= u(0) + u(1) + · · · + u(r)

= u(0)+ u(1)+ · · · + u(r)

= [u].

This completes the proof. 

Corollary 3.9. If u ∈ M is a ξ-sequence, then u is also a ξ-sequence.

Proof. By Theorem 3.8 there exists a polynomial f (z) ∈ L[z] such that f ([x]) = [u]. Consider the polynomial g(z) := ξf (z) ∈ L[z]. Since g([x]) = ξf (z) = ξ[u] = [u], the sequence u is, again by Theorem 3.8, a ξ-sequence. 

(7)

4 Divisibility

In this section we use the same notations as in the previous sections.

Proposition 4.1. Let u = (u0, . . . , um−1) be a ξ-sequence. Assume that the polynomials u1, u2, . . . , um−1 belong to k. Then u0 ∈ k.

Proof. Since ∂x∂u

i = ∂x∂u

i−1 for all i = 1, . . . , m − 1 (see Definition 3.1), we have

∂u0

∂xi = ∂x∂u0

i−1 = ∂a∂x0um−1

i−1 = 0, (for i = 1, . . . , m − 1) and moreover,

0 = ∂u∂x1

1 = ∂u∂x1

0 = ∂(u0+a∂x1um−1)

0 = ∂u∂x0

0 + a1∂u∂xm−1

0 = ∂u∂x0

0. Therefore, ∂u∂x0

0 = ∂u∂x0

1 = · · · = ∂x∂u0

m−1 = 0 and hence, since char(k) = 0, the polynomial u0 belongs to k. 

Proposition 4.2. Let u = (u0, . . . , um−1) be a nonzero homogeneous ξ- sequence. Then gcd(u0, . . . , um−1) = 1.

Proof. We know, by Lemma 3.7, that [u] = b[x]s for some b ∈ L and s > 0. Let 0 6= h ∈ k[x] := k[x0, . . . , xm−1] be a common divisor of all the polynomials u0, . . . , um−1. We will show that h ∈ k.

Put u0 = v0h, . . . , um−1 = vm−1h with v0, . . . , vm−1 ∈ k[x]. Then in the polynomial ring L[x] := L[x0, . . . , xm−1] we have the equality

b[x]s = [v]h.

But L[x] is a UFD and [x] = x0 + ξx1 + · · · + ξm−1xm−1 is an irreducible polynomial in L[x], so h = c[x]r for some nonzero c ∈ L and 0 6 r 6 s. This means (by Theorem 3.8) that (h, 0, 0, . . . , 0) is a ξ-sequence. Now Proposition 4.1 implies that h ∈ k. 

Theorem 4.3. Let k be a field of characteristic zero and let L = k[ξ] be a finite field extension of degree m > 1. Let z, x0, x1, . . . , xm−1 be variables and let u = (u0, . . . , um−1) be a sequence of polynomials belonging to k[x] :=

k[x0, . . . , xm−1]. If the polynomials u0, . . . , um−1 are the imaginary parts of a nonzero polynomial f (z) ∈ L[z], then gcd(u0, . . . , um−1) = 1.

(8)

Proof. Let 0 6= h ∈ k[x] := k[x0, . . . , xm−1] be a common divisor of all the polynomials u0, . . . , um−1. Denote by h the homogeneous component of highest degree of h, and let u0, . . . , um−1 be the homogeneous components of highest degree of u0, . . . , um−1, respectively. Then 0 6= h is a common divisor of all the polynomials u0, . . . , um−1 and moreover, by Lemma 3.5, (u0, . . . , um−1) is a homogeneous ξ-sequence. This implies, by Proposition 4.2, that h ∈ k. Therefore, h ∈ k and so, gcd(u0, . . . , um−1) = 1. 

As a consequence of theorems 3.8 and 4.3 we have:

Theorem 4.4. Let k be a field of characteristic zero and let k ( L be a finite field extension. If (u0, . . . , um−1) is a nonzero ξ-sequence, then the polynomials u0, . . . , um−1 have no common divisors of positive degrees. 

5 Quadratic extensions

Throughout this section L = k[ξ] is a quadratic field extension of k. We assume that ξ2 = r, where r is a nonzero element from k such that the polynomial t2 − r is irreducible in k[t]. Every element of L has a unique presentation of the form a + bξ with a, b ∈ k.

Let x, y, z be variables. If w ∈ k[x, y], then we denote by wx and wy

the partial derivatives ∂w∂x and ∂w∂y, respectively. In this case a pair (u, v) of polynomials from k[x, y] is a ξ-sequence iff uy = rvx and vy = ux. Such a pair we will call a ξ-pair. By Theorem 3.8 we have:

Proposition 5.1. Let (u, v) be a pair of polynomials from k[x, y]. The fol- lowing two conditions are equivalent.

(1) There exists a polynomial f (z) ∈ L[z] such that f (x + ξy) = u + ξv.

(2) uy = rvx and vy = ux. 

Let ∆ : k[x, y] → k[x, y] be a generalization of the Laplace operator defined by

∆ := r∂x22∂y22,

that is, ∆(w) = rwxx− wyy for w ∈ k[x, y]. It is easy to check that if (u, v) is a ξ-pair, then ∆(u) = ∆(v) = 0. Note the following observation.

Proposition 5.2. Let u ∈ k[x, y]. The following two conditions are equiva- lent.

(1) There exists a polynomial v ∈ k[x, y] such that (u, v) is a ξ-pair.

(2) ∆(u) = 0.

(9)

Proof. The implication (1) ⇒ (2) is obvious. We prove the implication (2) ⇒ (1). Let f := 1ruy and g := ux. Since fy = gx and char(k) = 0, there exists a polynomial v ∈ k[x, y] such that vx = f and vy = g. Then uy = rvx and vy = ux and hence, by Proposition 5.1, (u, v) is a ξ-pair. 

Consider two sequences (pn) and (qn) of polynomials from k[x, y] defined as follows:

p0 = 1, q0 = 0, pn+1 = xpn+ ryqn,

qn+1 = ypn+ xqn.

In particular, p1 = x, p2 = x2+ry2, p3 = x(x2+3ry2), p4 = x4+6rx2y2+r2y4, q1 = y, q2 = 2xy, q3 = y(3x2+ ry2), q4 = 4xy(x2+ ry2). Note the following matrix presentation of these sequences.

Proposition 5.3. If A = x ry y x



, then An= pn rqn qn pn



, for all n > 0.



It is easy to check that, for any nonnegative n, the pair (pn, qn) is a ξ-pair such that

(∗) pn+ ξqn = (x + ξy)n.

We present some observations concerning the sequence (pn) and (qn).

As a consequence of (∗) and Proposition 2.2 we obtain

Proposition 5.4. If n > 1, then the polynomials pn and qn are algebraically independent over k. 

As a consequence of (∗) and Theorem 4.3 we obtain Proposition 5.5. gcd(pn, qn) = 1. 

The equality (∗) implies the following proposition.

Proposition 5.6. If n and m are nonnegative integers, then pn+m = pnpm+ rqnqm and qn+m= pnqm+ pmqn. 

In particular, for n = m, we get

Proposition 5.7. p2n = p2n+ rq2n and q2n = 2pnqn. 

Using a simple induction and the above propositions it is easy to prove the next proposition.

(10)

Proposition 5.8.

(1) y | qn, for n ∈ N.

(2) If n | m, then qn| qm. (3) pn | q2kn, for n, k ∈ N.

(4) pn | p(2k+1)n, for n, k ∈ N.

(5) gcd(pkn, qn) = 1, for n, k ∈ N.

(6) gcd(qkn+r, qn) = gcd(qr, qn), for n, k, r ∈ N. 

Now, by Proposition 5.8 and the Euclid algorithm, we have:

Proposition 5.9. If n, m ∈ N, then

gcd (qn, qm) = qgcd(n,m). 

Note that a similar proposition is well known for some classical sequences of integers. Such a property have the sequences of Fibonacci numbers, Mersenne numbers and others (see for example, [1], [2]).

References

[1] W. Sierpi´nski, Elementary Theory of Numbers, Hafner Publishing Com- pany, New York, 1964.

[2] N. N. Vorobiev, The Fibonacci numbers, (Russian), Nauka, Moskow, 1978.

Faculty of Mathematics and Computer Science, Nicolaus Copernicus University,

87-100 Toru´n, Poland

e-mail: anow@mat.uni.torun.pl Faculty of Mathematics

University of L´od´z S. Banacha 22 90-238 L´od´z, Poland

e-mail: spodziej@imul.uni.lodz.pl

Cytaty

Powiązane dokumenty

By application of the operators T and q (they are derivations in the Hurwitz algebra A(C)), we can obtain various classical relations on the Bell exponential polynomials and

In papers devoted to the necessary and sufficient conditions for SLLN some authors give such conditions for certain classes of random variables, expressing them in terms of

The concepts of fuzzy sets and fuzzy set operations were first introduced by Zadeh [11] and subsequently several au- thors have discussed various aspects of the theory and

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

In 1994 Alford, Granville and Pomerance [1] proved that there exist infinitely many Carmichael numbers and that there are more than x 2/7 Carmichael numbers up to x, for

Let Z, N, Q be the sets of integers, positive integers and rational numbers respectively.. In this note we prove the following

Littlewood was the first to observe that this fact is “atypical” among sequences with growth similar to the square numbers.. In

They investigated the arithmetic properties of Q A and, in particular, they proved the following theorem:..