• Nie Znaleziono Wyników

The Abhyankar - Moh theorem

N/A
N/A
Protected

Academic year: 2021

Share "The Abhyankar - Moh theorem"

Copied!
20
0
0

Pełen tekst

(1)

The Abhyankar - Moh theorem

Andrzej Nowicki

N. Copernicus University, Faculty of Mathematics and Informatics, 87–100 Toru´n, Poland

(e-mail: anow@mat.uni.torun.pl) November 2, 1997

In this note we present a proof of the following theorem.

Theorem 0.1. Let k[t] be the ring of polynomials in one variable t over a field k of charac- teristic zero, and let f, g ∈ k[t] r k. If k[f, g] = k[t], then deg f | deg g or deg g | deg f .

The above theorem appeared in the fifties with an incorrect proof in Segre’s paper [12]

which discusses a subject of the Jacobian conjecture. The first time it was proved in 1975 by Abhyankar and Moh [3] (see also [1]). The history of this theorem and its applications are presented in the papers [3] and [4]. There appeared also other proofs in the years 1977 – 1982. They were publicated by Miyanishi [7] (see also [8]), Ganong [5] and Rudolph [11].

Abhyankar and Moh proved this theorem basing on the theory of approximate roots of polynomials which was developed by them. In 1986 Richman [10] presented a new interesting proof which does not use approximate roots. There is however a certain gap in his proof, noticed and corrected, in 1991, by Kang [6] (see also [9]).

The purpose of this article is a presentation of the full proof of Theorem 0.1 according to Richman.

We assumed in Theorem 0.1 that k is a field of characteristic zero. With a certain additional assumption which is ”char(k) - gcd(deg f, deg g)”, this theorem is also true for fields of positive characteristics (see for example [3]). In such version Richman proves it in [10]. But we do not take up this case here.

We stress that it is easy to prove Theorem 0.1 when the degree of f or g is a prime number (see Corollary 1.7).

The author would like to thank Ilona Nowosad and Maciej Smoczy´nski for their help in the preparation of this note.

1 Preliminary facts

Throughout this article k is a field of characteristic zero, k[t] is the polynomial ring in one variable t over k and k(t) is the field of rational functions in the variable t over k. If K ⊆ L is a finite field extension, then we denote by (L : K) the degree of this extension.

Proposition 1.1. If R is a ring such that k ( R ⊆ k[t], then the ring k[t] is integral over R.

(2)

Proof. Let f = antn+ · · · + a1t + a0 where an. . . , a0 ∈ k, an 6= 0, be a polynomial belonging to R r k. Then we have the equality tn+ a−1n an−1tn−1+ · · · + a−1n (a0− f ) = 0, which implies that t is integral over R (even over k[f ]). 

In a similar way we prove:

Proposition 1.2. If g ∈ k[t] r k, then k(g) ⊆ k(t) is a finite field extension and (k(t) : k(g)) = deg g. 

Assume now that f and g are polynomials belonging to k[t] r k, and consider the rings k ( k[g] ⊆ k[f, g] ⊆ k[t] and the fields k ( k(g) ⊆ k(f, g) ⊆ k(t). Proposition 1.1 implies that k[f, g] is integral over k[g]. In particular, the polynomial f is integral over k[g]. Therefore there exists a polynomial W ∈ k[g][X] such that W (f ) = 0.

Proposition 1.3. Let f, g ∈ k[t] r k and let W ∈ k[g][X] be the monic polynomial of the minimal degree such that W (f ) = 0. Then deg W = (k(f, g) : k(g)).

Proof. Let N = (k(f, g) : k(g)) = (k(g)(f ) : k(g)) and let B ∈ k(g)[X] be a minimal polynomial for f over k(g). Since W (f ) = 0 and W ∈ k[g][X] ⊂ k(g)[X], we obtain W = AB for some A ∈ k(g)[X]. Let a, b be elements from k[g] such that aA, bB ∈ k[g][X]. Then in the ring k[g][X] we have the equality: abW = aA · bB.

The ring k[g] is a unique factorization domain (because k[g] is a ring isomorphic to k[t]).

Therefore there exist elements a1, b1 ∈ k[g] and primitive polynomials A1, B1 ∈ k[g][X] such that aA = a1A1, bB = b1B1. Then we have the equality:

abW = a1b1(A1B1).

The polynomial W is primitive (because it is monic) and the polynomial A1B1is also primitive (by Gauss Lemma). Hence W = c(A1B1) where c is an invertible element of k[g]. Thus there exists an invertible element d ∈ k[g] such that dB1is a monic polynomial belonging to k[g][X].

Let H = dB1. Then H ∈ k[g][X] is a monic polynomial, H(f ) = 0, deg H = deg B = N and so, the minimality of W implies that deg W = N = (k(f, g) : k(g)). 

Corollary 1.4. Let f, g ∈ k[t] r k. Then

k[f, g] = k[g]fN −1+ k[g]fN −2+ · · · + k[g]f + k[g], where N = (k(f, g) : k(g)).

Proof. Let E = k[g]fN −1+· · ·+k[g]f +k[g]. We know, by Proposition 1.3, that fN ∈ E.

It is easy to show (using a simple induction) that fn∈ E for each n > N.  Corollary 1.5. Let f, g ∈ k[t]. If k(f, g) = k(g), then k[f, g] = k[g].

Proof. It follows from Corollary 1.4 for N = 1.  Corollary 1.6. If g ∈ k[t], then k(g) ∩ k[t] = k[g].

Proof. Let f ∈ k(g) ∩ k[t]. Then k(f, g) = k(g) and so, by Corollary 1.5, k[f, g] = k[g].

Thus f ∈ k[g]. 

The next corollary is a special case of Theorem 0.1.

(3)

Corollary 1.7. Let f, g ∈ k[t] rk. Assume that one of the numbers deg f and deg g is prime.

If k[f, g] = k[t], then deg f | deg g or deg g | deg f .

Proof. Let deg g = p, where p is prime. Then (by the assumption and Proposition 1.2) (k(f, g) : k(g)) = 1 · (k(f, g) : k(g))

= (k(t) : k(f, g))(k(f, g) : k(g))

= (k(t) : k(g))

= deg g = p,

and hence k[f, g] = k[g]fp−1+ · · · + k[g]f + k[g] (Corollary 1.4). Since t ∈ k[f, g], there exist nonzero elements a1, . . . , as∈ k[g] such that

t = a1fi1 + · · · + asfis, where s > 1 and p > i1 > · · · is> 0.

Assume that deg g - deg f . Then the degrees of the polynomials a1fi1, . . . , asfis are pairwise incongruent modulo p (because the degrees of the polynomials a1, . . . , asare divisible by p and p - deg f ). Therefore there exists r ∈ {1, . . . , s} such that ir > 0 and 1 = deg t = deg(arfir). This implies that deg f = 1 and we have: deg f | deg g. 

2 Notations and definitions

For the rest of the article we assume that f and g are polynomials belonging to k[t] r k.

We denote by N the number (k(f, g) : k(g)). If N = 1, then (see Corollary 1.5) f ∈ k[g] and then deg g | deg f . Therefore we assume that N > 1. The equalities

deg g = (k(t) : k(g)) = (k(t) : k(f, g))(k(f, g) : k(g)) imply that N 6 deg g.

If n is a positive integer then we denote:

An = fn+ k[g]fn−1+ · · · + k[g]f + k[g], An = fn+ k(g)fn−1+ · · · + k(g)f + k(g), Ln = k[g]fn+ k[g]fn−1+ · · · + k[g]f + k[g], Ln = k(g)fn+ k(g)fn−1+ · · · + k(g)f + k(g).

Moreover, we denote: L0= k[g] and L0= k(g). Observe that An= fn+ Ln−1, An= fn+ Ln−1.

If n > N then An = k[f, g] and An = k(f, g) (see Corollary 1.4). If n > N − 1 then Ln= k[f, g] and Ln= k(f, g).

Proposition 2.1. If n is a positive integer, then An∩ k[f, g] = An and Ln∩ k[f, g] = Ln.

(4)

Proof. We will show that An∩ k[f, g] = An. If n > N then the equality is obvious because in this case An = k(f, g). Assume now that n < N and let u ∈ An ∩ k[f, g].

Then u = fn+ an−1fn−1+ · · · + a1f + a0, for some an−1, . . . , a0 ∈ k(g). On the other hand u = bN −1fN −1 + · · · + b1f + b0, where bN −1, . . . , b0 ∈ k[g], because u ∈ k[f, g] = k[g]fN −1+ · · · + k[g]f + k[g]. The polynomials fN −1, . . . , f1, f0 form a basis of the space k(g)(f ) over k(g). The representation of the element u is then unique. This implies that the elements an−1, . . . , a0 belong to k[g], thus u ∈ An. Therefore we have the inclusion An∩ k[f, g] ⊆ An. The opposite inclusion is obvious. In a similar way we prove that Ln∩ k[f, g] = Ln. 

Definition 2.2. A sequence (h1, . . . , hN −1) of elements of the field k(f, g) is called an α- system, if:

(a) hn∈ An for each n = 1, . . . , N − 1,

(b) the numbers 0, deg h1, . . . , deg hN −1 are pairwise incongruent modulo deg g.

Definition 2.3. A sequence (h1, . . . , hN −1) of elements of the ring k[f, g] is called a β-system, if:

(a) hn∈ An for each n = 1, . . . , N − 1,

(b) the numbers 0, deg h1, . . . , deg hN −1 are pairwise incongruent modulo deg g.

3 Properties of α and β-systems

Every α-system (h1, . . . , hN −1), such that h1, . . . , hN −1∈ k[f, g], is (by Proposition 2.1) a β-system. It is easy to prove the following two propositions.

Proposition 3.1. Let (h1, . . . , hN −1) be an α-system and let n ∈ {1, . . . , N − 1}. Then:

(1) An= hn+ Ln−1,

(2) Ln= k(g)hn+ · · · + k(g)h1+ k(g). 

Proposition 3.2. Let (h1, . . . , hN −1) be a β-system. Then Ln= k[g]hn+ · · · + k[g]h1+ k[g]

for all n ∈ {1, . . . , N − 1}.  Now we prove:

Proposition 3.3. Let (h1, . . . , hN −1) be a β-system. Let h0 = 1, n < N and let w ∈ Ln. Then there exist uniquely determined integers s and r such that deg w = deg gshr, s > 0 and r ∈ {0, 1, . . . , n}.

Proof. We know from the previous proposition that Ln= k[g]hn+ · · · + k[g]h1+ k[g]h0. Thus there exist polynomials un, . . . , u0 ∈ k[g] such that w = unhn+ · · · + u0h0. From the definition of β-system it follows that the degrees of the polynomials unhn, . . . , u0h0 are pairwise distinct. Therefore deg w = deg urhr, for a certain r ∈ {0, 1, . . . , n}. The number deg ur is divisible by deg g (because ur ∈ k[g]). Thus there exists a non-negative integer s satisfying the equality deg ur = s deg g. Then we have: deg w = deg urhr= deg gshr.

Suppose now that deg gs1hr1 = deg gs2hr2 for some integers r1, s1, r2, s2such that s1, s2 >

0 and r1, r2∈ {0, . . . , n}. Then

deg hr1 − deg hr2 = (s1− s2) deg g,

and hence deg hr1 ≡ deg hr2 (mod deg g). Therefore r1 = r2 and s1 = s2.  The following theorem play an essential role in the proof of Theorem 0.1.

(5)

Theorem 3.4. There exists a β-system.

A proof of this theorem we will present in Section 10.

4 Proof of Theorem 0.1

Proof. Denote: df = deg f , dg = deg g, d = gcd(df, dg). Let df = ad, dg = bd where a and b are relatively prime natural numbers. There exist integers m and n such that 1 = ma + nb and 0 6 m < b. Then d = mdf + ndg, and therefore

d ≡ deg fm (mod dg). (4.1)

Let us recall that we denote by N the number (k(f, g) : k(g)). In our case: N = dg (because dg = (k(t) : k(g)) = (k(t) : k(f, g))N = 1 · N = N ).

Let (h0 = 1, h1, . . . , hN −1) be a β-system. Since the polynomial td belongs to k[f, g] = LN −1 (see Corollary 1.4), then, by Proposition 3.3, there exist integers r and s such that

d = deg td= deg gshr, s > 0 i r ∈ {0, 1, . . . , N − 1}. (4.2) The polynomial fm belongs to Lm. Applying Proposition 3.3 once more (this time for the polynomial fm) we observe that deg fm = deg gs1hr1, where 0 6 r16 m and s1 > 0.

From (4.1) and (4.2) it follows that deg hr1 ≡ deg hr (mod dg) and consequently r1= r.

Therefore we have the inequality r < b which implies that hr∈ Lb−1= k[g]fb−1+ · · · + k[g]f0. Observe now that the degrees of the polynomials f0, f1, . . . , fb−1are pairwise incongruent modulo dg. In fact, suppose that deg fi ≡ deg fj (mod dg) for some i, j ∈ {0, 1, . . . , b − 1}.

Then (i − j)df = udg, where u ∈ Z, and we have the equality (i − j)a = ub which implies that b | i − j (because the numbers a and b are relatively prime), and so i = j.

Repeating the proof of Proposition 3.3 we see that deg hr= deg gpfi where p, i are some non-negative integers. Thus we have

d = deg gshr= sdg+ deg hr= sdg+ deg gpfi= (s + p)dg+ idf, where s + p > 0, i > 0 and s + p + i > 0. Consequently:

d = (s + p)dg+ idf > min(df, dg) > gcd(df, dg) = d, i.e. gcd(df, dg) = min(df, dg). Therefore df | dg lub dg | df. 

In the above proof the assumption ”k[t] = k[f, g]” we use only for a statement, that there exists a nonzero polynomial w belonging to k[f, g] such that deg w = d, where d = gcd(deg f, deg g). In our case w = td. Therefore we have proved:

Theorem 4.1 ([10]). Let f, g ∈ k[t] r k and let d = gcd(deg f, deg g). If k[f, g] contains a nonzero polynomial of the degree d then deg f | deg g or deg g | deg f . 

In the proof we have applied Theorem 3.4 saying that there exists at least one β-system.

This fact we have though not proved yet. The remaining part of the article leads to the proof of this fact.

(6)

5 The existence of α-systems

In this section we will prove that there exists at least one α-system.

Let S be a subset of k[t] r {0} such that each two different elements of S have different degrees. Denote by deg S the set {deg s; s ∈ S}. Recall that deg 0 = −∞.

If h ∈ k[t], then we denote by R(h, S) a polynomial belonging to k[t], which we define in the following inductive way.

Definition 5.1.

(1) R(0, S) = 0.

(2) Let h 6= 0, deg h = m > 0 and assume that R(h0, S) is already determined for all h0∈ k[t] such that deg h0 < deg h. Then:

(a) If deg h 6∈ deg S, then we put: R(h, S) = h.

(b) If deg h ∈ deg S, then there exists a unique element s ∈ S such that deg h = deg s and moreover, there exists a unique nonzero element a ∈ k such that deg(h − as) < deg h. In this case we put: R(h, S) = R(h − as, S).

Example 5.2. Let S = {x2+ 1, 2x + 1}. Then deg S = {2, 1} and we have:

(1) R(x3, S) = x3, R(x5+ 3x, S) = x5 + 3x (because the degrees of the polynomials x3 and x5+ 3x do not belong to deg S),

(2) R(x2+ x, S) = R((x2+ x) − (x2+ 1), S) = R(x − 1, S) = R((x − 1) − 1/2(2x + 1), S) = R(−3/2, S) = −3/2. 

Therefore we have the function R( , S) : k[t] −→ k[t]. It is easy to prove:

Proposition 5.3. Let h ∈ k[t]. Then:

(1) deg R(h, S) 6∈ deg S;

(2) if deg h 6∈ deg S, then R(h, S) = h;

(3) R(h, S) = 0 ⇐⇒ h = a1s1+ · · · + apsp where a1, . . . , ap ∈ k i s1, . . . , sp∈ S.  Now we will prove:

Proposition 5.4 ([10]). There exist polynomials h1, . . . , hN −1∈ k[t] such that:

(1) hn∈ Lnr Ln−1, for n = 1, . . . , N − 1;

(2) the numbers 0, deg h1, . . . , deg hN −1 are pairwise incongruent modulo deg g.

Proof. We construct the polynomials h1, . . . , hN −1 using an induction.

Let S = {gn; n = 0, 1, . . . }. Let h1 = R(f, S). Then h1 ∈ L1r L0 and the numbers 0 and deg h1 are incongruent modulo deg g. Assume that the polynomials h1, . . . , hn are already constructed. If n + 1 = N , then the construction is finished. Assume therefore that n + 1 < N and let h0 = 1. Consider the set

T = {hrgi; r = 0, . . . , n, i > 0}.

It is clear that T ⊂ k[t] r {0} and that the elements of T have pairwise different degrees. Let y0= R(fn+1, T ).

(7)

Then y0 ∈ fn+1 + Ln (in particular, y0 6= 0) and the elements of the set T ∪ {y0} (see Proposition 5.3) have pairwise different degrees. Thus we can consider the polynomial

y1 = R(gfn+1, T ∪ {y0}).

This polynomial belongs to the set (g + a0)fn+1+ Ln for some a0 ∈ k. In particular y1 6= 0.

Moreover, the elements of T ∪ {y0, y1} have pairwise different degrees.

Continuing this proceeding we can construct an infinite sequence y0, y1, . . . of nonzero polynomials belonging to k[t] such that

yi = R(gifn+1, T ∪ {y0, . . . , yi−1}) for i = 0, 1, . . . .

Now it is easy to show that the degrees of the polynomials from T ∪ {y0, . . . , yi} are pairwise different. In particular, the degree of each of the polynomials y0, y1, . . . is not divisible by deg g. It is also easy to verify that yi ∈ (gi+ ai−1gi−1+ · · · + a0)fn+1+ Ln, for some a0, . . . , ai−1∈ k. This implies that the polynomials y0, y1, . . . belong to Ln+1r Ln.

Now we will prove that among the polynomials y0, y1, . . . there exists at least one whose degree is incongruent modulo deg g to any of the numbers deg h0, deg h1, . . . , deg hn .

Assume that it is not true. Then for every non-negative integer j there exists a number mj ∈ {0, . . . , n} such that

deg yj ≡ deg hmj (mod deg g).

Since the set {0, . . . , n} is finite, there exists an infinite subset U of non-negative integers and there exists s ∈ {0, . . . , n} such that deg yu≡ deg hs (mod deg g) for all u ∈ U . Then

deg yu+ pudeg g = deg hs, where pu ∈ Z. (5.1) Observe that pu > 0. In fact, suppose that pu 6 0. Then deg yu = deg(g−puhs), g−puhs∈ T and we have a contradiction because the degrees of the elements of T ∪ {y0, . . . , yu} are pairwise different. Thus every integer of the form pu is positive. Now the equality (5.1) implies that deg yu < deg hs for all u ∈ U . But it is a contradiction, because the set U is infinite and the degrees of the polynomials y0, y1, . . . are pairwise different.

The obtained contradiction proves that there exists a non-negative integer j such that the numbers deg yj, 0, deg h1, . . . , deg hn are pairwise incongruent modulo deg g. We define now the polynomial hn+1 putting hn+1= yj. 

Let h1, . . . , hN −1 be such polynomials as in the above proposition. The fact that hn ∈ Lnr Ln−1 (for n = 1, . . . , N − 1) implies that hn has a nonzero coefficient at fn, belonging to k[g]. Dividing hn by this nonzero coefficient we obtain a rational function h0n belonging to An. The numbers 0, deg h01, . . . , deg h0N −1 are obviously pairwise incongruent modulo deg g.

Thus we have:

Proposition 5.5. There exists at least one α-system. 

6 Definitions of the sequence c(1), ..., c(p)

Let us recall (see Section 2) that if n = 0, 1, . . . , N − 1, then we denote by Ln the set k(g)fn+ · · · + k(g)f1+ k(g)f0. Let us introduce also the following two notations.

Un = {deg ϕ; ϕ ∈ Ln},

Gn = the ideal in Z generated by Un.

(8)

We have then the sequence of ideals G0 ⊆ G1 ⊆ · · · ⊆ GN −1. Observe that G0 is a principal ideal generated by deg g. Moreover we have

Proposition 6.1. G06= G1.

Proof. We know by Proposition 5.4 that there exists h1 ∈ L1rL0such that deg g - deg h1. This means that deg h1 ∈ G1r G0. 

Now we define a sequence c(1), c(2), . . . , c(p) of certain natural numbers.

Definition 6.2.

(1) c(1) = 1.

(2) Assume that the numbers c(1), . . . , c(m) are already determined and let S be the set of all natural numbers n ∈ {c(m) + 1, c(m) + 2, . . . , N − 1} such that Gc(m)( Gn. If S = ∅, then we end the definition and the number m we denote by p. If S 6= ∅, then c(m + 1) is the smallest element of S.

Moreover, we fix: c(p + 1) = N .

The sequence c(1), . . . , c(p) has the following properties:

(i) c(1) = 1,

(ii) c(1) < c(2) < · · · < c(p) 6 N − 1, (iii) G0 ( Gc(1)( · · · ( Gc(p),

(iv) Gc(n)= Gc(n)+1 = · · · = Gc(n+1)−1, for n = 1, . . . , p.

We will prove now that the sequence c(1), . . . , c(p) can be defined also in a different way;

using an α-system. For that purpose first prove the following proposition.

Proposition 6.3. Let (h1, . . . , hN −1) be an α-system and let n ∈ {1, . . . , N − 1}. Denote h0= g. Then:

(1) Un=Sn

j=0{deg hj+ Z deg g}, (2) Gn= (deg h0, deg h1, . . . , deg hn).

Proof. (1) Let 0 6 j 6 n, s ∈ Z. Then deg hj + s deg g = deg(gshj) ∈ Un, and thus

nj=0{deg hj+ Z deg g} ⊆ Un. The reverse inclusion follows from the fact that all the numbers deg g, deg h1, . . . , deg hN −1 are pairwise incongruent modulo deg g.

(2) It follows from (1) and Proposition 3.1. 

Let (h1, . . . , hN −1) be an α-system. Put h0 = g and denote:

d0 = deg h0= deg g d1 = deg h1

...

dN −1 = deg hN −1.

The numbers d0, d1, . . . , dN −1 are integers (they can be negative) and none of them are congruent modulo d0 = deg g. Now the sequence c(1), . . . , c(p) can be defined as follows.

(9)

Definition 6.4.

(1) c(1) = 1.

(2) Assume that the numbers c(1), . . . , c(m) are already defined. If the numbers d0, d1, . . . , dN −1 belong to the ideal (d0, d1, . . . , dc(m)), then c(m) is the last element of the sequence and in this case the number m we denote by p. In the opposite case c(m + 1) = i, where i is the smallest of the numbers c(m) + 1, c(m) + 2, . . . , N − 1 such that di 6∈ (d0, d1, . . . , dc(m)).

We see (by Proposition 6.3) that the above two definitions describe the same sequence c(1), . . . , c(p). By Definition 6.2 it follows that this sequence does not depend on a selection of an α-system.

Note also an obvious proposition.

Proposition 6.5. Let n ∈ {1, . . . , N − 1}. The following conditions are equivalent:

(1) n appears in the sequence c(1), . . . , c(p);

(2) there exists ϕ ∈ An such that deg ϕ 6∈ Gn; (3) Gn−16= Gn. 

7 The numbers e

1

, ..., e

p

Assume that (h1, . . . , hN −1) is a fixed α-system. Put h0 = g and let di = deg hi for i = 0, 1, . . . , N − 1. In the previous section we defined the number p and the sequence c(1), . . . , c(p). Additionally we assume that c(p + 1) = N .

Consider now the ideals D0, D1, . . . , Dp, of the ring Z, defined as follows:

Definition 7.1.

Dj =

( (d0), for j = 0,

(d0, d1, . . . , dc(j)), for j ∈ {1, . . . , p}.

Let s ∈ {1, . . . , p}. We know (see Definition 6.4) that dc(s) 6∈ Ds−1. Since the ideal Ds−1

is principal (because every ideal of Z is principal), there exists a natural number n such that ndc(s) ∈ Ds−1.

Definition 7.2. We denote by es the smallest natural number n such that ndc(s)∈ Ds−1. Thus, we have natural numbers e1, . . . , ep. They are greater than 1. We will prove:

Proposition 7.3 ([10]). If s ∈ {1, . . . , p}, then esc(s) = c(s + 1).

Before the proof of this proposition we will introduce a new set Bs and we will prove several lemmas.

Let s be an element of the set {1, . . . , p}. Denote:

Bs= {hnhjc(s) ; 0 6 n < c(s), 0 6 j < es}.

In particular we have:

B1 = {g = gh01, gh11, . . . , ghe11−1}.

(10)

Lemma 7.4.

(1) The degrees of the elements of Bs are pairwise incongruent modulo d0 = deg g.

(2) |Bs| = esc(s) (where |Bs| denotes the cardinality of the set Bs).

(3) The set Bs is linearly independent over k(g).

(4) esc(s) 6 N .

(5) The set Bs is a basis of the linear space Lesc(s)−1 over k(g).

Proof. (1). Suppose that

deg(hnhac(s)) ≡ deg(hmhbc(s)) (mod d0),

where n, m ∈ {0, 1, . . . , c(s) − 1} and 0 6 a, b < es. If a = b, then deg hn ≡ deg hm modd0

and hence, by the definition of an α-system, n = m. Therefore a 6= b.

Let a > b. Then (a − b)dc(s) ∈ Ds−1 and we have a contradiction with the property of the minimality of the number es. The same argument we use in the case b > a.

(2). It follows from (1) that the elements of the form hnhjc(s) (where 0 6 n < c(s) and 0 6 j < es) are pairwise different. Of course there are c(s)es such elements.

(3). Let α1a1 + · · · + αnan = 0, where α1, . . . , αn ∈ k(g) and a1, . . . , an are pairwise different elements of Bs.

Suppose that α1 6= 0. Then a1 = −α−11 α2a2− · · · − α−11 αnan. Since the degrees of the elements of the form α−11 αi are divisible by d0, the number deg a1 is, by (1), equal to one of the numbers deg(α−11 α2a2), . . . , deg(α−11 αnan). But it is a contradiction with (1).

(4). It follows from (2), (3) and from the fact that Bs ⊆ k(f, g) and N = (k(f, g) : k(g)).

(5). If 0 6 n < N , then the polynomials f0, f1, . . . , fnare linearly independent over k(g).

Therefore the dimension over k(g), of every space of the form Ln, is equal to n + 1. Since esc(s) − 1 < N (see (4)), we have:

dimk(g)Lesc(s)−1 = esc(s).

We know from (2) and (3) that the dimension of the subspace generated by Bs is also equal to esc(s). Therefore, it suffices to prove that Bs⊆ Lesc(s)−1.

Let u = hnhjc(s), where 0 6 n < c(s) and 0 6 j < es. Then we have:

u ∈ An(Ac(s))j ⊆ An+jc(s)⊆ Ln+jc(s)⊆ Lesc(s)−1, and so Bs⊆ Lesc(s)−1. 

Lemma 7.5. esc(s) 6 c(s + 1).

Proof. If s = p, then the inequality follows from Lemma 7.4(4) (because c(p + 1) = N ). Assume now that s < p and let m be a natural number smaller than esc(s). Then hm ∈ Lesc(s)−1 and so, by Lemma 7.4(5), hm = α1a1+ · · · + αnan, where α1, . . . , αn ∈ k(g) and a1, . . . , an are pairwise different elements of Bs. We know, by Lemma 7.4(1), that the elements a1, . . . , as have pairwise incongruent degrees modulo d0. This implies that dm= deg hm = deg(αjaj), for some j ∈ {1, . . . , n}. It is obvious that deg(αjaj) ∈ Ds. Hence dm ∈ Ds. So, if m < esc(s), then dm ∈ Ds. By the definition of the sequence c(1), . . . , c(p) we know that dc(s+1) 6∈ Ds. Therefore c(s + 1) > esc(s). 

(11)

Lemma 7.6. For every w ∈ Ds there exists b ∈ Bs such that w ≡ deg b (mod d0)).

Proof. (Induction with respect to s). Let s = 1. Recall that c(1) = 1, B1 = {gh01 = g, gh11, . . . , ghe11−1} and D1 = (d0, d1). Let w ∈ D1. Then there exist integers a and m such that w = ad0 + md1. Since e1d1 ∈ (d0), we may assume that 0 6 m < e1. Then w = a deg g + deg hm1 = (a − 1) deg g + deg(ghm1 ) which implies that w ≡ b (mod d0), where b = ghm1 ∈ B1.

Assume now that the lemma is true for some s ∈ {1, . . . , p − 1} and let w ∈ Ds+1. Since Ds+1 = Ds+ (dc(s+1)), we have w = w0+ mdc(s+1), where w0 ∈ Ds and m ∈ Z. We know that es+1dc(s+1) ∈ Ds. Therefore we may assume that 0 6 m < es+1. Moreover, by the induction, w0 ≡ deg b0 (mod d0) for some b0 ∈ Bs. Using Lemmas 7.4, 7.5 and Proposition 3.1 we obtain:

b0 ∈ Bs⊆ Lesc(s)−1 ⊆ Lc(s+1)−1= k(g)hc(s+1)−1+ · · · + k(g)h0.

Therefore, deg b0 ≡ deg hr (mod d0), for some r < c(s + 1) (because h0, . . . , hc(s+1)−1 have pairwise incongruent degrees). Now we have:

w = w0+ mdc(s+1) ≡ deg b0+ deg hmc(s+1) ≡ deg hr+ deg hmc(s+1) = deg b, where b = hrhmc(s+1) ∈ Bs+1. 

Now we are ready to prove the announced proposition.

Proof of Proposition 7.3. Let n = esc(s). Then (Lemma 7.4) n 6 N . Assume that n = N and suppose that s < p. Then (by Lemma 7.5) we obtain the following contradiction:

N = esc(s) 6 c(s + 1) 6 c(p) 6 N − 1. Thus, if n = N , then s = p and epc(p) = N = c(p + 1).

Assume now that n < N . We will show that dn 6∈ Ds. Suppose that it is not true.

Let dn ∈ Ds. Then there exists an element b ∈ Bs (see Lemma 7.6) such that dn ≡ deg b (mod d0). But, by Lemma 7.4 and Proposition 3.1,

Bs⊂ Lesc(s)−1 = Ln−1= k(g)hn−1+ · · · + k(g)h0,

thus deg b ≡ deg hj for some j < n, and hence deg hn = dn ≡ deg hj. Now we have a contradiction with the fact that the polynomials h0, . . . , hnhave pairwise incongruent degrees.

Therefore dn6∈ Ds. Thus n > c(s + 1) (see Definition 6.4), esc(s) > c(s + 1) and so, by Lemma 7.5, esc(s) = c(s + 1). 

Corollary 7.7. If s ∈ {1, . . . , p}, then e1e2. . . es = c(s + 1).

Proof. e1 = e1c(1) = c(2), e1e2 = c(2)e2= c(3), etc. 

8 The rational functions of the form w(H, s)

From Section 5 we aim at the proof that there exists a β-system. This fact is essential in the proof of Abhyankar–Moh theorem which is presented in Section 4. We already know (see Section 5) that there exists an α-system.

In the present section with every α-system H we associate certain, uniquely determined, rational functions w(H, 1), . . . , w(H, p), belonging to the spaces Lc(1)−1, . . . , Lc(p)−1, re- spectively. We will prove that the degrees of these functions are lower than the numbers dc(1), . . . , dc(s), respectively. This will allow us to prove (in the next section) that there exists

(12)

such an α-system for which all the above functions are equal zero. This fact will be very useful in the proof of the theorem on the existence of a β-system.

Let H = (h1, . . . , hN −1) be an α-system and let s be a fixed natural number belonging to the set {1, . . . , p}. We assume additionally that h0= g and hN = 0. Before we introduce the rational function mentioned above, we will prove the following two propositions.

Proposition 8.1. For each w ∈ Lc(s+1)−1 there exist uniquely determined elements w1, w2, . . . , we, belonging to Lc(s)−1, such that

w = w1he−1+ w2he−2+ · · · + we−1h1+ weh0, (8.1) where h = hc(s) and e = es.

Proof. As in Section 7, let B = Bs= {hnhj; 0 6 n < c(s), 0 6 j < e}. We know that B is a basis of the space Lec(s)−1 = Lc(s+1)−1over k(g) (see Lemma 7.4 and Proposition 7.3).

Thus

w = a1b1+ · · · + arbr,

where a1, . . . , ar ∈ k(g) and b1, . . . , br are pairwise different elements of B. Excluding, in the above equality, all the elements of the form h0, . . . , he−1 and assembling the rest of them respectively, we obtain an equality of the form (8.1) in which the elements w1, . . . , webelong to the space k(g)h0+ · · · + k(g)hc(s)−1= Lc(s)−1.

The uniqueness follows from the fact that the sets of the form Bsare linearly independent over k(g). 

Proposition 8.2. hec(s)s − hc(s+1)∈ Lc(s+1)−1.

Proof. Let H = hec(s)s − hc(s+1). If s = p, then H = hec(p)p (because hc(p+1)= hN = 0) and then H ∈ k(f, g) = LN −1= Lc(p+1)−1. For s < p we have:

hec(s)s ∈ (Ac(s))es ⊆ Aesc(s) = Ac(s+1)

(see Proposition 7.3) and hc(s+1) ∈ Ac(s+1). But Ac(s+1) = fc(s+1) + Lc(s+1)−1, so H ∈ Lc(s+1)−1. 

As a consequence of the above two propositions we get

Corollary 8.3. There exist uniquely determined elements w1, w2, . . . , we, belonging to Lc(s)−1, such that

he− hc(s+1)= w1he−1+ w2he−2+ · · · + we−1h1+ weh0, (8.2) where h = hc(s) and e = es. 

Definition 8.4. The element w1∈ Lc(s)−1 from Corollary 8.3 we will denote by w(H, s).

Proposition 8.5. deg w(H, s) < deg hc(s).

(13)

Proof. Let h = hc(s), e = es, B = Bs = {hnhj; 0 6 n < c(s), 0 6 j < e} and let

he− hc(s+1) = a1b1+ · · · + arbr, (8.3) where a1, . . . , ar∈ k(g) and b1, . . . , br are pairwise different elements of B. Since the degrees of b1, . . . , brare pairwise incongruent modulo d0 (Lemma 7.4), there exist q ∈ {1, . . . , r} such that

deg(he− hc(s+1)) = deg(aqbq) (8.4) and deg(he− hc(s+1)) > deg(aibi) for all i 6= q. This implies that the number deg(he− hc(s+1)) belongs to the ideal Ds. Moreover we have:

deg he= edc(s)∈ Ds and deg hc(s+1) = dc(s+1) 6∈ Ds.

Hence deg hc(s+1)< deg he and so, deg(he− hc(s+1)) = deg he= e deg h. Thus we have:

deg(aqbq) = e deg h and deg(aibi) < e deg h for i 6= q. (8.5) Observe that if deg bj 6∈ Ds−1(where j ∈ {1, . . . , r}), then j 6= q. In fact, if j = q then, by (8.5), deg(ajbj) = e deg h = esdc(s) ∈ Ds−1 (see the definition of the number es). Moreover, deg aj ∈ D0 ⊆ Ds−1. Then we have a contradiction: deg bj = deg(ajbj) − deg aj ∈ Ds−1. Therefore j 6= q and, by (8.5), we know that

if deg bj 6∈ Ds−1, then deg(ajbj) < e deg h. (8.6) Let us return to the equality (8.3) and look at the elements b1, . . . , br belonging to B.

Choose of these elements all those which contain the factor he−1. If such elements do not exist then from the uniqueness of the presentations (8.3) and (8.2) it follows that w(H, s) = 0 and then our proposition is proved, because then deg w(H, s) = −∞ < deg h (since h 6= 0).

Assume now that {b1, . . . , bm}, where m 6 r, is the set of all such elements among b1, . . . , br, which contain the factor he−1. Then we have:

a1b1+ · · · + ambm = w(H, s)he−1. (8.7) Let j ∈ {1, . . . , m}. Then bj = hnhe−1, for some n such that 0 6 n < c(s). This implies that deg bj 6∈ Ds−1. In fact, suppose that the number deg bj = deg hn+ (e − 1)dc(s) belongs to the ideal Ds−1. Then the number (e − 1)dc(s) also belongs to this ideal (because deg hn = dn∈ Ds−1). It is a contradiction with the property of the minimality of e.

Now we deduce, by (8.6), that deg(ajbj) < e deg h, which means (see (8.7)) that deg(w(H, s) + (e − 1) deg h = deg(w(H, s)he−1< e deg h,

and hence deg w(H, s) < deg hc(s). 

9 New α-systems

Let H = (h1, . . . , hN −1) be an α-system. Let h0 = g, hN = 0 and assume that s is a fixed number from the set {1, . . . , p}. Denote by H the sequence (h1, . . . , hN −1) of rational functions belonging to the field k(f, g), which are defined as follows:

(14)

Definition 9.1.

hn=

( hn, for n 6= c(s),

hc(s)− (1/es)w(H, s), for n = c(s).

Proposition 9.2. H is an α-system.

Proof. The facts that hc(s) ∈ Ac(s) = fc(s)+ Lc(s)−1 and w(H, s) ∈ Lc(s)−1 imply that hc(s) ∈ Ac(s). Thus hn ∈ An, for n ∈ {1, . . . , N − 1}. We know (see Proposition 8.5) that deg w(H, s) < deg hc(s). It implies that deg hc(s) = deg hc(s), and therefore the numbers 0, deg h1, . . . , hN −1are pairwise incongruent modulo deg g. 

In the proof of the next proposition we will use the following lemma.

Lemma 9.3. If a2, a3, . . . , ae ∈ Lc(s)−1, then there exist uniquely determined elements b2, b3, . . . , be∈ Lc(s)−1 such that

a2he−2+ a3he−3+ · · · + ae = b2he−2+ b3he−3+ · · · + be, where h = hc(s), h = hc(s) and e = es.

Proof. Let u = a2he−2+ · · · + ae. Observe that u ∈ L(e−1)c(s)−1. In fact,

u ∈ Lc(s)−1(Ac(s))e−2 ⊆ Lc(s)−1(Lc(s))e−2 ⊆ Lc(s)−1+(e−2)c(s)= L(e−1)c(s)−1. (9.1) The space L(e−1)c(s)−1 is obviously included in Lec(s)−1 = Lc(s+1)−1, therefore u ∈ Lc(s+1)−1 and, by Proposition 8.1, there exist uniquely determined elements b1, . . . , be ∈ Lc(s)−1 such that

u = b1he−1+ b2he−2+ · · · + be.

We must show that b1 = 0. Suppose the contrary. Let b1 ∈ Lnr Ln−1, where 0 6 n < c(s) (we assume that L−1= 0). Then b1he−1 ∈ Ln+(e−1)c(s)r V , where V = Ln−1+(e−1)c(s) and

b1he−1= u − u, (9.2)

where u = b2he−2+ · · · + be. In the same way as in (9.1) we show that u ∈ L(e−1)c(s)−1. But L(e−1)c(s)−1 ⊆ V . Thus the right side of the equality (9.2) belongs to V and the left side does not belong to V . The obtained contradiction implies that b1 = 0 and it ends the proof of our lemma. 

Proposition 9.4 (compare [6] page 403). Let w = w(H, s) and w = w(H, s). Assume that w 6= 0.

(1) If w ∈ L0 = k(g), then w = 0.

(2) If w ∈ Lj r Lj−1 where 0 < j < c(s), then w = 0 or w ∈ Lir Li−1 for some i such that 0 6 i < j.

Proof. Let e = es, h = hc(s), h = hc(s)= h − (1/e)w and let

R =

e−2

X

r=0

e r

 hr(−1

ew)e−r

(15)

(recall that e > 2; see Section 7). Then

he− hc(s+1)= whe−1+ w2he−2+ · · · + we−1h1+ weh0, where w2, . . . , we∈ Lc(s)−1 and

he− hc(s+1) = (h − 1ew)e− hc(s+1)

= hee1he−1(1e)w1+ R − hc(s+1)

= (he− hc(s+1)) − whe−1+ R

= (w2he−2+ · · · + we−1h1+ we) + R.

It follows from Lemma 9.3 that there exist elements b2, . . . , be∈ Lc(s)−1 such that w2he−2+

· · · + we−1h1+ we= b2he−2+ · · · + be−1h1+ be, and therefore

he− hc(s+1)= (b2he−2+ · · · + be−1h1+ be) + R. (9.3) Now we will deal with the component R. Assume that w ∈ Lj, where 0 6 j < c(s) and let m ∈ {0, 1, . . . , e − 2}. Then

hmwe−m∈ Lmc(s)Le−mj ⊆ Lmc(s)+j(e−m)

and it is easy to show that mc(s) + j(e − m) 6 (e − 1)c(s) + j − 1. This implies that

R ∈ L(e−1)c(s)+j−1. (9.4)

Observe that (e − 1)c(s) + j − 1 6 ec(s) − 1 = c(s + 1) − 1. Thus R ∈ Lc(s+1)−1 and so, by Proposition 8.1, there exist uniquely determined elements z1, . . . , ze∈ Lc(s)−1 such that

R = z1he−1+ · · · + ze−1h1+ ze. (9.5) Now from (9.3) and (9.5) we obtain the equality

w = w(H, s) = z1. Observe also that

z2he−2+ · · · + ze∈ Lc(s)−1(Lc(s))e−2⊆ L(e−1)c(s)−1. Thus, by (9.4), we have:

whe−1= z1he−1 ∈ L(e−1)c(s)+j−1. (9.6) Now it is not difficult to prove the properties (1) and (2). Assume that w 6= 0. Then w ∈ Lir Li−1for some i such that 0 6 i < c(s), assuming that L−1= 0. Since he−1 ∈ A(e−1)c(s), then

whe−1 ∈ L(e−1)c(s)+ir L(e−1)c(s)+i−1. (9.7) If j = 0, then the property (9.6) contradicts with the property (9.7). Thus, if j = 0 (that is, if w ∈ k(g)), then w = 0. Therefore we have (1).

Assume now that j > 0. Then, by (9.6) and (9.7), we get (e − 1)c(s) + i − 1 < (e − 1)c(s) + j − 1, and so i < j. Therefore we have (2). 

Starting from an arbitrary α-system we may construct a new α-system H. By the same way, an α-system H emerges of an α-system H, etc. We can make such a conversion suffi- ciently many times (for each s = 1, . . . , p). Then, by Proposition 9.4, we obtain:

(16)

Proposition 9.5. There exists an α-system H such that w(H, s) = 0, for each s ∈ {1, . . . , p}.



Note the following lemma.

Lemma 9.6 ([2]). Let e, n be natural numbers such that en 6 N . Let u, v ∈ k(f, g).

Assume that u ∈ An, v ∈ Aenand ue− v ∈ Len−n−1 (assuming that L−1 = 0). Then u ∈ An. A proof of this lemma we will present in the last section in which we will prove a more general fact (Proposition 11.1). Now, using the above lemma, we will prove the following proposition.

Proposition 9.7. There exists an α-system H = (h1, . . . , hN −1) such that each element of the form hc(s) (dla s = 1, . . . , p) belongs to the ring k[f, g].

Proof. Let H be an α-system such as in Proposition 9.5. We will show that if s ∈ {1, . . . , p}, then hc(s) ∈ k[f, g].

Assume first that s = p. Let u = hc(p), v = hc(p+1) = hN = 0, e = ep, n = c(p). Then u, v ∈ k(f, g), en = epc(p) = c(p + 1) = N , u ∈ An and the equality N = (k(g)(f ) : k(g)) implies that v = 0 ∈ AN = Aen. Moreover (by Corollary 8.3 and Definition 8.4),

ue− v = hec(p)− hc(p+1)= w2he−2c(p)+ · · · + we−1hc(p)+ we, where w2, . . . , we are some elements belonging to Lc(p)−1. This implies that

ue− v ∈ Lc(p)−1(Lc(p))e−2 ⊆ Lec(p)−c(p)−1= Len−n−1.

Therefore all the assumptions of Lemma 9.6 are satisfied. Thus hc(p)= u ∈ An⊆ k[f, g].

Now let s + 1 6 p and assume that hc(s+1) ∈ k[f, g]. Denote: u = hc(s), v = hc(s+1), e = es

and n = c(s). Then en = esc(s) = c(s + 1) 6 c(p) < N , u ∈ An and v ∈ k[f, g] ∩ Ac(s+1) = Ac(s+1)= Aen(see Proposition 2.1). As above (thanks to the fact that w(H, s) = 0) we show that ue− v ∈ Lc(s)es−c(s)−1 = Len−n−1. Thus all the assumptions of Lemma 9.6 are again satisfied. Therefore hc(s) = u ∈ An⊆ k[f, g]. 

10 The existence of β-systems

Now we are ready to prove the theorem on the existence of a β-system announced before.

The proof of Theorem 3.4. Let H = (h1, . . . , hN −1) be an α-system such as in Propo- sition 9.7. Let

H0 = {hjc(1)1 · · · hjc(p)p ; 0 6 js < es for s = 1, . . . , p}, J = {j1c(1) + · · · + jpc(p); 0 6 js< es for s = 1, . . . , p}.

The fact that hc(1), . . . , hc(p)∈ k[f, g], implies that H0 ⊂ k[f, g]. Each element of the form hjc(1)1 · · · hjc(p)p belongs to the set Aj1c(1)+···+jpc(p). We will show that the elements of the set H0 one can arrange in a sequence (h00 = 1, h01, . . . , h0N −1) which will be a β-system. For this purpose it is enough to prove the following four properties:

(1) Each element of J is smaller than N . (2) The elements of J are pairwise different.

(17)

(3) The cardinality of J equals N .

(4) The degrees of polynomials belonging to H0 are pairwise incongruent modulo deg g.

Proof (1). We know (see Proposition 7.3) that esc(s) = c(s + 1), for s = 1, . . . , p. We also know that c(p + 1) = N . Thus we have:

j1c(1) + · · · + jpc(p) < e1c(1) + j2c(2) + · · · + jpc(p)

= c(2) + j2c(2) + · · · + jpc(p) 6 e2c(2) + j3c(3) + · · · + jpc(p)

= c(3) + j3c(3) + · · · + jpc(p) ...

6 epc(p) = c(p + 1) = N.

Proof (2). Let

j1c(1) + · · · + jpc(p) = i1c(1) + · · · + ipc(p), (10.1) where i1, j1< e1, . . . , ip, jp < ep. Since c(1) = 1, c(2) = e1, c(3) = e1e2, . . . , c(p) = e1· · · ep−1 (see Corollary 7.7) then, by the equality (10.1), the number |i1 − j1| is divisible by e1, and thus i1= j1 (because i1, j1< e1). Now we have the equality

j2c(2) + · · · + jpc(p) = i2c(2) + · · · + ipc(p)

from which it follows (after the division by e1) that the number |i2 − j2| is divisible by e2, that means i2 = j2. Repeating this procedure we obtain that is= js for all s = 1, . . . , p.

Proof (3). From (2) it follows that the set J contains exactly e1· · · ep= N elements.

Proof (4). Let

a = hic(1)1 · · · hic(p)p , b = hjc(1)1 · · · hjc(p)p ,

where 0 6 is, js< ee, for s = 1, . . . , p, and assume that deg a ≡ deg b (mod deg g). Then (i1dc(1)+ · · · + ipdc(p)) − (j1dc(1)+ · · · + jpdc(p)) ∈ (d0),

and thus |ip− jp|dc(p)∈ Dp−1and, from the property of the minimality of ep, we obtain the equality ip = jp. In a similar way we deduce that ip−1= jp−1. Repeating this procedure we can see that a = b. It ends the proof of the property (4) and the proof of Theorem 3.4. 

11 The proof of Lemma 9.6

It has remained to prove Lemma 9.6. We will do it in the present section. For this purpose first we will prove the following proposition.

Proposition 11.1. Let R ⊂ P be commutative rings containing the field Q of rational num- bers. Let F ∈ P [x] be a monic polynomial of the degree n > 1 and let e > 1 be a natural number. Consider the polynomial

Fe = xne+ bne−1xne−1+ bne−2xne−2+ · · · + b0. If bne−1, bne−2, . . . , bne−n ∈ R, then F ∈ R[x].

(18)

For the proof of this proposition we need certain lemmas. Let n, e be positive integers and let

L(n, e) = {(in, . . . , i0) ∈ Zn+1; in+ · · · + i0= e, is> 0 for s = 0, . . . , n}.

Consider the following partial order 6 on the set L(n, e):

(in, . . . , i0) 6 (jn, . . . , j0) ⇐⇒









in 6 jn

in−1+ in 6 jn+ jn−1

...

i0+ · · · + in−1+ in 6 jn+ jn−1+ · · · + j0. Consider also the function ϕ : L(n, e) −→ N ∪ {0} defined by the formula:

ϕ(in, . . . , i0) = nin+ (n − 1)in−1+ · · · + 1i1+ 0i0. Note the following obvious lemma.

Lemma 11.2. Let i, j ∈ L(n, e). Then:

(1) i 6 j ⇒ ϕ(i) 6 ϕ(j);

(2) i < j ⇒ ϕ(i) < ϕ(j);

(3) If i 6= j and ϕ(i) = ϕ(j), then i, j are not comparable with respect to 6.  Let πn−1, . . . , π0 be elements of L(n, e) defined as follows:

πn−1 = (e − 1, 1, 0, 0, . . . , 0, 0), πn−2 = (e − 1, 0, 1, 0, . . . , 0, 0), ...

π1 = (e − 1, 0, 0, 0, . . . , 1, 0), π0 = (e − 1, 0, 0, 0, . . . , 0, 1).

The next lemma is also obvious.

Lemma 11.3. Let i = (in, . . . , i0) ∈ L(n, e) and let s ∈ {0, 1, . . . , n}. Then:

(1) if is> 1, then ϕ(πs) > ϕ(i);

(2) if ir> 0 for some r < s, then ϕ(πs) > ϕ(i). 

Proof of Proposition 11.1. Let F = xn+ an−1xn−1+ · · · + a1x + a0, where an−1, . . . , a0 ∈ P . Then:

Fe= X

in+···+i0=e

hin, . . . , i0iain−1n−1· · · a0i0xϕ(in,...,i0), (11.1) where hi,. . . , i0i = e!(in! · · · i0!)−1. We will show, using an induction, that the elements an−1, . . . , a0 belong to R.

First we prove that an−1 ∈ R. For this purpose study the coefficient standing by xne−1 in (11.1). Since ne − 1 = ϕ(πn−1), we have in (11.1) the component

he − 1, 1, 0, . . . , 0ia1n−1xne−1.

Let i = (in, . . . , i0) be such an element of L(n, e) which satisfies the equality ϕ(i) = ne − 1 = ϕ(πn−1).

Cytaty

Powiązane dokumenty

Let G be Green’s function of C\J (that trivially coincides with Siciak’s function defined before (3)).. The implication 3)⇒2) is proved.. This was the most

We prove the ` p -spectral radius formula for n-tuples of commuting Banach algebra elements.. This generalizes results of some

This model has been constructed by Wieczorek in [3], where the author was interested mainly in the existence and properties of competitive equilib- ria. In the present paper we

The space X of all countable ordinal numbers, endowed with the order topology, is sequentially compact and therefore countably compact4. This shows that Theorem 2 is false if R is

only Σ 1 2 graphs and require that the universe is a generic extension of L, the constructible universe.) The proofs in [7] involve a technique quite close to arguments in [9] but

In the case of k = 1 and real variables, applying the Banach contrac- tion principle, the Neumann series and the Fourier series methods resulted in getting certain existence

In this section we shall present some considerations concerning convergence of recurrence sequences, and their applications to solving equations in Banach

D i b l´ık, On existence and asymptotic behaviour of solutions of singular Cauchy problem for certain system of ordinary differential equations, Fasc. H a l e, Theory of