• Nie Znaleziono Wyników

2 The derivation ∆

N/A
N/A
Protected

Academic year: 2021

Share "2 The derivation ∆"

Copied!
14
0
0

Pełen tekst

(1)

Algebras of constants

for some extensions of derivations

Andrzej J. Maciejewski Andrzej Nowicki March 6, 2002

Abstract

Let A = B[t] be the polynomial ring in one variable over a unique factorization domain B, where k is a field of characteristic zero. Let δ be a derivation of B, ϕ ∈ B with δ(ϕ) 6= 0, and let s be a positive integer. Denote by d the derivation of A such that d(b) = tsδ(b) for b ∈ B and d(t) = s+11 δ(ϕ). We describe the rings of constants of d.

Several applications of our description are also given.

1 Introduction

Throughout this paper k is a field of characteristic zero. Let R be a commutative algebra over k with identity. A k-linear mapping d : R → R is called a k-derivation of R if it satisfies the Leibniz rule: d(ab) = ad(b) + bd(a) for all a, b ∈ R.

Let d be a k-derivation of R. We denote by Rd the kernel of d, that is, Rd = Ker d = {a ∈ R; d(a) = 0}.

This set is a k-subalgebra of R which we call the ring of constants of d. If R is a domain and k is a field, then we denote by R0 the field of quotients of R and we denote also by d the unique extension of d to R0. In this case Rd0 is a subfield of R0 containing k.

Assume that R = k[X] := k[x1, . . . , xn] is the polynomial ring over k in n ≥ 1 variables. We know a description of all k-derivations of R. If d is a k- derivation of k[X], then we have the polynomials f1 := d(x1), . . . , fn := d(xn), belonging to k[X], and then

d = f1

∂x1 + · · · + fn

∂xn.

(2)

Every k-derivation d of k[X] is uniquely determined by a sequence (f1, . . . , fn) of polynomials from k[X].

Let d be a k-derivation of k[X] and consider the ring of constants k[X]d. In this case the ring k[X]dcoincides with the k-algebra of all polynomial first integrals of the following system of ordinary differential equations:

dxi

dt = fi(x1(t), . . . , xn(t)), i = 1, . . . , n,

where f1 = d(x1), . . . , fn = d(xn). It is known ([17], [15]) that if n ≤ 3, then k[X]dis finitely generated over k. If n ≥ 5, then there exists a k-derivation of k[X] for which the ring of constants is not finitely generated over k ([2], [1]).

For n = 4 the problem is open. Every example of k-derivation of k[X] with nonfinitely generated ring of constants is a counterexample to the famous fourteenth problem of Hilbert (see [12], [14], [3], [6]). We present two new such examples for n = 6 (see Examples 5.9 and 5.10).

The problem of a description of the ring k[X]d is known to be difficult even for n = 2. The importance of the study of these rings comes from the fact that many important mathematical problems can be transformed into problems concerning kernels of certain derivations of k[X] (see for example [3], [14]). In this paper we describe the rings of constants of some classes of derivations of the polynomial ring k[X].

Let A = B[t] be the polynomial algebra in one variable t over a unique factorization k-domain B. We will consider the following four derivations δ,

∆, D and d.

The derivation δ : B → B is an arbitrary nonzero derivation of B.

The next derivation ∆ : A → A is the unique derivation such that ∆(t) = 0 and ∆(b) = δ(b) for b ∈ B..

Let ϕ ∈ B and δ(ϕ) 6= 0. The derivation D : A → A is defined as:

D(t) = δ(ϕ) and D(b) = δ(b) for b ∈ B.

If s is a positive integer, then d : A → A is the unique derivation such that

d(t) = 1

s + 1δ(ϕ) and d(b) = tsδ(b) for all b ∈ B.

Throughout this article all algebras are commutative k-algebras with identity and all derivations are k-derivations.

The aim of this note is to describe the algebras of constants of the above derivations. The main result is Theorem 4.5 which states that Ado = Boδ(h) and Ad= Bδ[h], where h = ts+1− ϕ.

In Section 5 we present several applications of this theorem. Further ap- plications we present in Section 6, where we discuss the algebras of constants

(3)

of the derivation db = ysz∂x − xsz∂y + (bysz − xys)∂z, where s is a positive integer and b ∈ k. If s = 2 then the derivations of the form db are asso- ciated with systems of differential equations which appear in descriptions of 3-dimensional homogeneous sub-Riemannian spaces which are Lie groups (see [4] for details). Some investigations concerning db one may find also in [7] and [9].

2 The derivation ∆

It is obvious that A is equal to Bδ[t]. In this case we do not need to assume that B is a UFD. However, this assumption is important in our proof of the following proposition.

Proposition 2.1. Ao = Boδ(t).

Proof. Let F ∈ A0 . Put F = P/Q, where P and Q are nonzero relatively prime polynomials from A = B[t]. Let

P = pntn+ · · · + p1t + p0, Q = qmtm+ · · · + q1t + q0, with p0, . . . , pn, q0, . . . , qm ∈ B and pn6= 0, qm 6= 0. Then

∆(P ) = δ(pn)tn+ · · · + δ(p1)t + δ(p0), ∆(Q) = δ(qm)tm+ · · · + δ(q1)t + δ(q0).

Since F ∈ A0, we have

(i) P ∆(Q) = ∆(P )Q.

Suppose that ∆(P ) = 0. Then ∆(Q) = 0. Hence δ(pn) = · · · = δ(p1) = δ(p0) = 0 and δ(qm) = · · · = δ(q1) = δ(q0) = 0, and hence P, Q ∈ Bδ[t], so F ∈ Boδ(t). If ∆(Q) = 0, then ∆(P ) = 0 and again F ∈ Boδ(t).

Now we may assume that ∆(P ) 6= 0 and ∆(Q) 6= 0.

Since A is a unique factorization domain and the polynomials P and Q are relatively prime, it follows from the equality (i) that

∆(P ) = M P and ∆(Q) = M Q

for some nonzero M ∈ A. Comparing the degrees with respect to t we deduce that M ∈ B. Thus,

δ(pi) = M pi, δ(qj) = M qj,

for all i = 0, 1, . . . , n and j = 0, 1, . . . , m. In particular, all the fractions

pn

qm, pqn−1

m , . . . ,qp0

m and qpm

n, qm−1p

n , . . . ,pq0

n belong to Boδ. Hence pn = anqm,

(4)

pn−1 = an−1qm, . . . , p0 = a0qm and qm = bmpn, qm−1 = bm−1pn, . . . , q0 = b0pn, for some a0, . . . , an, b0, . . . , bm ∈ Bδo. This means that P = qmP1 and Q = pnQ1, where P1, Q1 ∈ Boδ[t] r {0}. Thus we have

F = QP = qpm

n

P1

Q1 = bmP1

Q1 ∈ Boδ(t).

Therefore, Ao ⊆ Boδ(t). The opposite inclusion is obvious. 

3 The derivation D

Proposition 3.1. AD = Bδ[ω], ADo = Bδo(ω), where ω = t − ϕ.

Proof. Consider the automorphism σ : A → A such that σ(b) = b for all b ∈ B, and σ(t) = ω = t − ϕ. Then D = σ∆σ−1 and we have

AD = σ A = σ Bδ[t] = Bδ[σ(t)] = Bδ[ω].

Similarly, by Proposition 2.1, ADo = Boδ(ω). 

In the next section we will use the following lemma.

Lemma 3.2. If f ∈ Ao satisfies the equality tD(f ) = λδ(ϕ)f for some λ ∈ Q r Z, then f = 0.

Proof. Suppose that f 6= 0 and put f = uv, where u and v are nonzero relatively prime polynomials from A = B[t]. Since tD(uv) = λδ(ϕ)uv, we have tv12(D(u)v − uD(v)) = λδ(ϕ)uv and so

(1) t(D(u)v − uD(v)) = λδ(ϕ)uv.

This implies that t divides u or v.

Assume that t | u. Let u = tmu1, where m ≥ 1, u1 ∈ A, t - u1. Then, by (1), we get t(tmD(u1)v + mtm−1u1δ(ϕ)v − tmu1D(v)) = λδ(ϕ)tmu1v, and hence

t(D(u1)v − u1D(v)) = (λ − m)δ(ϕ)u1v.

This implies, that t | v (since t - u1 and λ − m 6= 0), but it is impossible because t | u and the polynomials u and v are relatively prime.

A similar argument we use when t | v. In this case we put v = tmv1, where m ≥ 1, v1 ∈ A, t - v1 and, by (1), we obtain

t(D(u)v1− uD(v1)) = (λ + m)δ(ϕ)uv1.

This implies that t | u and again we have a contradiction with the assumption that u and v are relatively prime. 

(5)

4 The derivation d

In this section we denote by p the polynomial ts+1.

Since t is an algebraic element over the field Bo(p), Ao is an algebraic extension of Bo(p) of the degree s + 1. This implies that every element g ∈ Ao has a unique presentation of the form

(∗) g = gs(p)ts+ gs−1(p)ts−1+ · · · + g1(p)t + g0(p),

where g0, g1, . . . , gs∈ Ao. It is clear that if g ∈ A, then g0, g1, . . . , gs ∈ A.

Lemma 4.1. Let g ∈ Ao and g0, g1, . . . , gs ∈ Ao as in (∗). If g ∈ Ado, then gi(p)ti ∈ Ado

for every i = 0, 1, . . . , s.

Proof. Replacing A and d by A ⊗kk and d ⊗k1, respectively, where k is an algebraic closure of k, we may assume that the field k is algebraically closed.

Denote by ε a primitive root of unity of degree s + 1 and consider the automorphism τ : Ao → Ao such that

τ (t) = εt and ∀b∈Bo τ (b) = b.

Then τ dτ−1 = εsd. In fact, if b ∈ Bo, then τ dτ−1(b) = τ d(b) = τ (tsδ(ϕ)) = (εt)sδ(ϕ) = εstsδ(ϕ) = εsd(b). Moreover, τ dτ−1(t) = τ d(εst) = εsτ (d(t)) = εsd(t).

Let g = g(t) ∈ Ado. The equality τ dτ−1 = εsd implies that g(εt) = τ (g(t)) ∈ Ado, and consequently g(εit) = τi(g(t)) ∈ Ado for any integer i. This means, in particular, that every element of the form

s)igs(p)ts+ (εs−1)igs−1(p)ts−1+ · · · + (ε1)ig1(p)t + g0(p), i ∈ {0, 1, . . . , s}, belongs of Ado. Now, using a determinant of Vandermonde we deduce that gi(p)ti ∈ Ado for i = 0, 1, . . . , s. 

Lemma 4.2. If f ∈ Ao, then d(f (p)) = D(f )(p)ts.

(6)

Proof. It is sufficient to show that the equality d(f (p)) = D(f )(p)ts holds when f = btn with b ∈ Bo and n ≥ 0. In this case we have:

d(f (p)) = d (b(ts+1)n)

= d(b)(ts+1)n+ (s + 1)nbd(t)t(s+1)n−1

= tsδ(b)(ts+1)n+ (s + 1)nbs+11 δ(ϕ)t(s+1)n−1

= (δ(b)tn) (p)ts+ (nbtn−1δ(ϕ)) (p)ts

= (D(b)tn+ bD(tn)) (p)ts

= D(btn)(p)ts

= D(f )(p)ts. This completes the proof. 

Lemma 4.3. Let f ∈ Ao and 1 ≤ r ≤ s. If f (p)tr ∈ Ado, then f = 0.

Proof. Assume that d(f (p)tr) = 0. Then, by Lemma 4.2, we have 0 = d(f (p))tr+ rtr−1f (p)d(t)

= D(f )(p)tstr+ rtr−1f (p)s+11 δ(ϕ)

= tr−1· D(f )t + s+1r δ(ϕ)f (p).

This implies that D(f )t + s+1r δ(ϕ)f (p) = 0, and hence tD(f ) = λδ(ϕ)f

for λ = −s+1r . Since 1 ≤ r ≤ s, the coefficient λ belongs to Q r Z. Therefore, by Lemma 3.2, f = 0. 

Using the above lemmas we may prove the following proposition.

Proposition 4.4. If g ∈ Ado, then g = f (p) for some f ∈ ADo. If g ∈ Ad, then g = f (p) for some f ∈ AD.

Proof. Assume that g ∈ Ado and let g0, . . . , gs ∈ Ao be such as in (∗).

Put f = g0. Then, by Lemmas 4.1 and 4.3, g = f (p). Moreover, Lemma 4.2 implies that f ∈ ADo. It is clear that if g ∈ A, then f ∈ A. 

Now we are ready to prove the following theorem which is the main result of this article.

(7)

Theorem 4.5. Let A = B[t] be a polynomial algebra in a single variable t over a unique factorization k-domain B, where k is a field of characteristic zero. Let δ : B → B be a nonzero derivation and let ϕ be an element of B such that δ(ϕ) 6= 0. Moreover, let s be a positive integer. Consider the derivation d : A → A defined by

d(t) = 1

s + 1δ(ϕ) and ∀b∈B d(b) = tsδ(b).

Then Ado = Boδ(h) and Ad = Bδ[h], where h = ts+1− ϕ.

Proof. First observe that h = ω(p) where, as in Proposition 3.1, ω = t − ϕ.

Now let g ∈ Ad. Then, by Proposition 4.4, g = f (p) for some f ∈ AD. We know, by Proposition 3.1, that AD = Bδ[ω]. Hence f = γ(ω) for some γ ∈ Bδ[t], and hence g = f (p) = γ(ω(p)) = γ(h) ∈ Bδ[h]. Thus, Ad⊆ Bδ[h].

Since the opposite inclusion is obvious, we have Ad= Bδ[h].

Repeating the same arguments, thanks to Propositions 3.1 and 4.4, we obtain the equality Ado = Boδ(h). 

5 Examples

Example 5.1. Let d be the derivation of k[x, y, z] defined by

d(x) = yz d(y) = zx d(z) = xy.

Then k[x, y, z]d = k[x2− y2, z2− x2] and k(x, y, z)d= k(x2− y2, z2− x2).

Proof. Consider the derivation δ = y∂x + x∂y of the polynomial algebra B = k[x, y] and let ϕ = x2. It is easy to check that k[x, y]δ = k[x2− y2] and k(x, y)δ = k(x2− y2). Moreover, 12δ(ϕ) = xy = d(z). Using Theorem 4.5 for B, δ, ϕ as above and t = z, A = k[x, y, z] and s = 1, we obtain:

k[x, y, z]d= k[x, y]δ[z2− ϕ] = k[x2− y2][z2− x2] = k[x2− y2, z2− x2].

The same we do for k(x, y, z)d. 

The next example is a generalization of Example 5.1.

(8)

Example 5.2. Let dn be the derivation of k[X] = k[x1, . . . , xn], (n ≥ 2) defined by









dn(x1) = x2x3· · · xn dn(x2) = x1x3· · · xn

...

dn(xn) = x1x2· · · xn−1.

Then k[X]dn = k[h2, h3, . . . , hn] and k(X)dn = k(h2, h3, . . . , hn), where hi = x21− x2i, i = 2, 3, . . . , n. 

Proof. We use an induction with respect to n. This is clear for n = 2. Now assume that it is true for n − 1. Let B = k[x1, . . . , xn−1], δ = dn−1, t = xn, A = k[x1, . . . , xn], s = 1 and ϕ = x21. Then A = B[t], t2− ϕ = x2n− x21 = −hn and

1

1+sδ(ϕ) = 122x1dn−1(x1) = x1x2. . . xn−1= dn(xn) = dn(t).

Hence, by induction and Theorem 4.5, we get

k[X]dn = Adn = Bδ[t2− ϕ] = k[h2, . . . , hn−1][hn] = k[h2, h3, . . . , hn].

The same we do for k(X)dn.  By a similar way we obtain:

Example 5.3. Let dn be the derivation of k[X] = k[x1, . . . , xn], (n ≥ 2) defined by













dn(x1) = +x2x3· · · xn

dn(x2) = −x1x3· · · xn dn(x3) = −x1x2· · · xn

...

dn(xn) = −x1x2· · · xn−1.

Then k[X]dn = k[h2, h3, . . . , hn] and k(X)dn = k(h2, h3, . . . , hn), where hi = x21+ x2i, i = 2, 3, . . . , n. 

Using Theorem 4.5 it is easy to prove the following examples which de- scribe the algebras of constants of a derivation d of a polynomial algebra over k. In Examples 5.4 and 5.5 the mapping d is a derivation of k[x, y, z].

(9)

Example 5.4. If d = xy∂x + zy∂y + (x2 + y2)∂z , then k[x, y, z]d = k[f ], k(x, y, z)d= k(xy, f ), where f = z2− x2− y2. 

Example 5.5. Let s be a positive integer. If d = zsy∂x + zsx∂y + xsy∂z , then k[x, y, z]d= k[f, g], k(x, y, z)d = k(f, g), where f = zs+1− xs+1, g = x2− y2.



In the next examples d is a derivation of the polynomial algebra k[x, y, z, t].

Example 5.6 is connected with a diagonal derivation described in [16].

Example 5.6. Let d = tx∂x + ty∂y + −ntz∂x + (x2+ y2− nz2)∂t, where n ≥ 1. Then k[x, y, z, t]d = k[f, xnz, xn−1yz, xn−2y2z, . . . , ynz], k(x, y, z, t)d = k(x/y, x/z, f ), with f = t2− x2− y2− z2. 

The next two examples concern a Jouanolou derivation (see [8], [11], [10]

or [14]).

Example 5.7. Let d = ty2 ∂∂x+ tz2 ∂∂y+ +tx2 ∂∂z + (xy2+ yz2+ zx2)∂t. Then k[x, y, z, t]d= k[f ], k(x, y, z, t)d= k(f ), with f = t2− x2− y2− z2.  Example 5.8. If d = t2y2 ∂∂x + t2z2 ∂∂y + +t2x2 ∂∂z + (x2y2 + y2z2 + z2x2)∂t, then k[x, y, z, t]d = k[f ], k(x, y, z, t)d= k(f ), where f = t3− x3− y3− z3.  Note also two examples of derivations of k[x1, . . . , x6] with nonfinitely generated ring of constants.

Example 5.9. Let d be the derivation of k[X] := k[x1, x2, . . . , x6] defined by d = x21x6

∂x3 + (x1x3 + x4)x6

∂x4 + x4x6

∂x5 + x21

∂x6. Then k[X]d is not finitely generated over k.

Proof. Consider the derivation δ = x21

∂x3 + (x1x3+ x4) ∂

∂x4 + x4

∂x5

of the polynomial ring B = k[x1, . . . , x5] and let ϕ = 2x3, t = x6 and s = 1. Then k[X] = B[t] and, by Theorem 4.5, k[X]d = Bδ[x6 − 2x3].

Daigle and Freudenburg ([1], [5]) proved that Bδ is not finitely generated over k. This implies that k[X]d is also not finitely generated over k, because k[X]d/(x6− 2x3) ≈ Bδ. 

(10)

Example 5.10. Let d be the derivation of k[X] := k[x1, x2, . . . , x6] defined by

d = x21x26

∂x3 + (x1x3+ x4)x26

∂x4 + x4x26

∂x5 + x4x25

∂x6. Then k[X]d is not finitely generated over k.

Proof. We repeat the proof of Example 5.9 with ϕ = x35 and s = 2. 

6 Constants of the derivation d

b

Let s ≥ 1 be a fixed integer. If b ∈ k, then we denote by db the derivation of k[x, y, z] defined by

db(x) = ysz db(y) = −xsz

db(z) = −xys+ bysz.

For any b ∈ k, the derivation db has a nontrivial constant polynomial, namely db(h) = 0 where

h = xs+1+ ys+1.

Thus, k[h] ⊆ k[x, y, z]d and k(h) ⊆ k(x, y, z)d. We will characterize all the cases in which k[x, y, z]d= k[h] and all the cases in which k(x, y, z)d= k(h).

For this aim, we consider the following derivation δb : k[x, z] → k[x, z] defined by

 δb(x) = z

δb(z) = −x + bz.

The derivation δb is linear with matrix M =

 0 1

−1 b

 .

Let α, β ∈ k be the eigenvalues of M . Then α and β are roots of the polynomial x2 − xb + 1. It is known (see for example [13] or [14]) that k[x, z]d6= k if and only if the eigenvalues α and β are linearly dependent over N.

We know also (see [13] or [14]) that k(x, z)d 6= k if and only if the eigen- values α and β are linearly dependent over Z and the Jordan matrix of M is not of the form

 c 1 0 c



for some nonzero c ∈ k (in our case this means that b 6= ±2).

(11)

Lemma 6.1. The following conditions are equivalent.

(1) The eigenvalues α and β are linearly dependent over N.

(2) There exist positive integers n, m such that −b2 = (m−n)mn 2.

Proof. It is clear if b = 0. Let us assume that b 6= 0. Observe that α + β = b and αβ = 1. In particular, α 6= 0 and β 6= 0.

(1) ⇒ (2). Assume that nα + mβ = 0 for some n, m ∈ N with (n, m) 6=

(0, 0). Then n 6= 0, m 6= 0 and n 6= m (because b 6= 0). Moreover, 0 = nα + m(b − α) and 0 = n(b − β) + mβ. Hence α = m−nm b, β = n−mn b and we have 1 = αβ = m−nm b · n−mn b = −(m−n)nm 2, that is, −b2 = (m−n)mn 2.

(2) ⇒ (1). Assume that −b2 = (m−n)mn 2, where m, n ∈ N, m > 0, n > 0.

Since b 6= 0, we have n 6= m. Put α0 = m−nm b, β0 = n−mn b. Then α0+ β0 = b, α0β0 = 1, hence α0 and β0 are the roots of x2−bx+1, that is, {α0, β0} = {α, β}.

Now the linear dependence over N of α and β follows from the equality nα0+ mβ0 = 0. 

Repeating the above proof we get

Lemma 6.2. The following conditions are equivalent.

(1) The eigenvalues α and β are linearly dependent over Z.

(2) There exist nonzero integers n, m such that −b2 = (m−n)mn 2. 

Now, using the above facts and Theorem 4.5 (for B = k[x, z], t = y, A = k[x, y, z], δ = δb, d = db and ϕ = −xs+1), we get the following proposition.

Proposition 6.3. (1) k[x, y, z]db 6= k[h] if and only if there exist positive integers n, m such that −b2 = (m−n)mn 2.

(2) k(x, y, z)db 6= k(h) if and only if b 6= ±2 and there exist nonzero integers n, m such that −b2 = (m−n)mn 2. 

Using the remarks given before Lemma 6.1 it is not difficult to present a description of the algebras of constants of the derivation δb. Therefore, as a consequence of Theorem 4.5 we get

Proposition 6.4. (0) If b = 0, then k[x, y, z]d = k[h, f ] and k(x, y, z)d = k(h, f ) where f = xs+1+ zs+1.

(1) If b 6= 0 and k[x, y, z]db 6= k[h], then there exist relatively prime positive integers m, n such that b2 = −(m−n)mn 2, and then k[x, y, z]db = k[h, f ] and k(x, y, z)db = k(h, f ), where

f = (ax + z)n(x + az)m, with a = m−nn b.

(12)

(2) If b 6= 0 and k(x, y, z)db 6= k(h), then there exist nonzero relatively prime integers m, n such that b2 = −(m−n)mn 2, and then k(x, y, z)db = k(h, F ), where

F = (ax + z)n(x + az)m, with a = m−nn b. 

Corollary 6.5. If b ∈ R, then C[x, y, z]db = C[h], where R and C are fields of real and complex numbers, respectively.

Proof. Suppose that C[x, y, z]db 6= C[h]. Then, by Proposition 6.3, there exist positive integers m, n such that b2 = −(n−m)nm 2 and then b2 < 0.  Corollary 6.6. If 0 6= b ∈ Q, then k[x, y, z]d = k[h] and the following conditions are equivalent

(1) k(x, y, z)db 6= k(h).

(2) There exist relatively prime positive integers p and q such that (p, q) 6=

(1, 1) and b = ±p2pq+q2.

If p, q are such as in (2), then k(x, y, z)db = k(h, F ) where F is the rational function defined as follows

F =

(px − qz)p2(qx − pz)−q2, if b = p2pq+q2, (px + qz)p2(qx + pz)−q2, if b = −p2pq+q2.

Proof. For a proof that k[x, y, z]db = k[h] we use the same argument as in the proof of Corollary 6.5.

(2) ⇒ (1). If b = ±p2pq+q2, p, q ∈ N with (p, q) 6= (1, 1), then b 6= ±2 and for n = −p2, m = q2 we have: −b2 = −(p2p+q2q22)2 = (m−n)mn 2. Hence, by Proposition 6.3, k(x, y, z)db 6= k(h).

(1) ⇒ (2) The condition (1) implies, by Proposition 6.3, that b 6= ±2 and −b2 = (m−n)mn 2 for some nonzero integers n, m. We may assume that n i m are relatively prime and that n = −n1 < 0 and m > 0. Let b = ±uv, where u and v are relatively prime positive integers. Then uv22 = (m+nmn1)2

1 , and hence u = m + n1 and v2 = mn1. Since gcd(m, n1) = 1, there exists relatively prime positive integers p and q such that m = p2 and n1 = q2. Thus, b = ±uv = ±m+npq 1 = ±p2pq+q2.

The remaining part of this corollary is a consequence of Proposition 6.4.



(13)

Example 6.7. If b = 52, then k[x, y, z]db = k[h] and k(x, y, z)db = k(h, F ), where

F = (2x−z)x−2z4. 

Corollary 6.8. If 0 6= b ∈ Z, then k(x, y, z)db = k(h).

Proof. This is a consequence of Corollary 6.6, because if p and q are relatively prime positive numbers and (p, q) 6= (1, 1), then the number

p2+q2

pq = pq + qp is not integer.  Example 6.9. Let k = C and b =

2

2 iyz. Then k[x, y, z]db = k[h, f ] and k(x, y, z)db = k(h, f ), where

f = 2x3+ 3z2x +√

2iz3. 

Acknowledgement.

The authors wish to thank the referee for pointing out several errors in the first version of this paper and for writ- ing a very constructive report.

References

[1] D. Daigle, G. Freudenburg, A counterexample to Hilbert’s fourteenth problem in dimension five, J. of Algebra 221(1999), 528-535.

[2] H.G.J. Derksen, The kernel of a derivation, J. Pure Appl. Algebra 84(1993), 13 – 16.

[3] A. van den Essen, Polynomial automorphisms and the Jacobian Conjec- ture, Progress in Mathematics 190, 2000.

[4] E. Falbel and C. Gorodski. Sub-Riemannian homogeneous spaces in di- mensions 3 and 4. Geom. Dedicata, 62(3)(1996), 227 - 252.

[5] G. Freudenburg, Recent progress on Hilbert’s fourteenth problem via tri- angular derivations, Annales Polonici Math., 76(2001).

[6] G. Freudenburg, A survey of counterexamples to Hilbert’s fourteenth problem, Seredica Math. J., 27(2001), 171 - 192.

(14)

[7] P. Hoyer. ¨Uber die Integration eines Differentialgleichungs von der Form

˙x1 = a1x2x3 + a2x3x1 + a3x1x2, ˙x2 = b1x2x3+ b2x3x1 + b3x1x2, ˙x3 = c1x2x3+c2x3x1+c3x1x2 durch eliptische Funktionen. PhD thesis, K¨onigl.

Friedrich-Wilhelms Univ., Berlin, 1879.

[8] J.-P. Jouanolou, Equations de Pfaff alg´ebriques, Lect. Notes in Math., 708, Springer-Verlag, Berlin (1979).

[9] A. J. Maciejewski and M. Przybylska, Studies of Hoyer system, Rep.

Math. Phys., 2000.

[10] J. Moulin Ollagnier, A. Maciejewski, A. Nowicki, J. -M. Strelcyn, Around Jouanolou non-integrability theorem, Indag. Mathem., N. S., 11(2000), 239 - 254.

[11] J. Moulin Ollagnier, A. Nowicki, J. -M. Strelcyn, On the non-existence of constants of derivations: The proof of a theorem of Jouanolou and its development, Bull. Sci. Math., 119(1995), 195 - 233.

[12] M. Nagata, Lectures on the Fourteenth Problem of Hilbert, Lect. Notes 31, Tata Institute, Bombay, 1965.

[13] A. Nowicki, On the non-existence of rational first integrals for systems of linear differential equations, Linear Algebra and Its Applications, 235(1996), 107 - 120.

[14] A. Nowicki, Polynomial derivations and their rings of constants, N.

Copernicus University Press, Toru´n, 1994.

[15] A. Nowicki, M. Nagata, Rings of constants for k-derivations on k[x1, . . . , xn], J. Math. Kyoto Univ., 28(1988), 111 - 118.

[16] A. Nowicki, J. -M. Strelcyn, Generators of rings of constants for some diagonal derivations in polynomial rings, J. Pure Appl. Algebra, 101(1995), 207 - 212.

[17] O. Zariski, Interpretations algebraico–geometriques du quatorzieme prob- leme de Hilbert, Bull. Sci. Math., 78(1954), 155 – 168.

Author’s addresses

Andrzej J. Maciejewski, Institute of Astronomy, University of Zielona G´ora, Lubuska 2, 65-265 Zielona G´ora, Poland; e-mail: maciejka@astri.uni.torun.pl Andrzej Nowicki, Faculty of Mathematics and Computer Science, N. Coper- nicus University, 87–100 Toru´n, Poland, e-mail: anow@mat.uni.torun.pl

Cytaty

Powiązane dokumenty

and [9]. Generally, if X is an algebraic set of pure dimension n ≥ 1, X is said to be uniruled if every component of X is uniruled. Points at which a polynomial map is not proper.

Hasse considered only abelian number fields L/Q, hence he was able to describe these fields in terms of their character groups X(L); as we are interested in results on

Keeping the type of option constant, in-the-money options experience the largest absolute change in value and out-of-the-money options the smallest absolute change in

(1 point) The sides of a rectangle has been measured to be 40cm and 50cm correct to the nearest 10cmA. The lower bound for the area of the rectangle is (select all

[r]

If the matrix ½aij is nilpotent, then the algebra of constants k½X d is finitely generated.. u It turns out that the assumption of nilpotency of the matrix ½aij is

NowicKi, Quasi-prime and d-prime ideals in commutative differential rings.. Now~c~:I, Derivations satisfying polynomial

If CH holds, then there is a compact 0-dimensional, hereditarily separable and hereditarily Lindel¨ of space X and a Radon proba- bility measure µ on X such that the measure algebra