User:D.Lazard/Bézout theorem

Introduction edit

Bézout's theorem states: if   polynomials in   variables have a finite number of common zeros, including the zeros at infinity, then the number of these zeros is the product of the degrees of the polynomials, if the zeros, including those at infinity are counted with their multiplicities. Despite its apparently simple statement, this theorem needed almost a century for being completely proved. The main difficulty was to give an accurate definition of multiplicities, and this requires some machiney of commutative algebra. Most proofs of this theorem proceed by recurrence on the number of polynomials, and use the concept of degree of a polynomial ideal. These proofs obtain the theorem as a corollary of if a homogeneous polynomial   of degree   is not a zero divisor modulo a homogeneous ideal   of degree   then the degree of the ideal   is  

If some hypotheses are relaxed, such as counting multiplicities or working with homogeneous polynomials, one gets only inequalities, commonly called Bézout inequalities. For example, if   polynomials in   variables have a finite number of common zeros, then the number of these zeros, counted with their multiplicities is at most the product of the   largest degrees of the polynomials. Surprisingly, this Bézout inequality is not a corollary of Bézout's theorem, and seems to have not been proved before 1983 (cite MW).

All these results require the definition of the degree of a polynomial ideal. Many definitions have been given, either in terms of algebraic geometry or in terms of commutative algebra. Most, but not all, deal with homogeneous ideals, and it is rather difficult to compare them. One of the objectives of this article is to give a general definition in terms of elementary commutative algebra and to prove that all other definitions are special cases. In fact, we give two different definitions that are equal for equi-dimensional ideals, but not in general.

Basically, the degree of an algebraic variety is the number of points of its intersection with a generic linear variety of a convenient dimension. This definition is not algebraic but can easily be translated into an algebraic one. However the resulting definition is not intrinsic, involving auxiliary generic polynomials, which makes proofs unnecessarily complicated. Therefore, we use the definition through Hilbert series, which provides a simpler presentation of the theory. In the case of homogeneous ideals, this approach is not new, although is seems unknown by many specialists of algebraic geometry, and we do not know any published presentation of it. Personally, we have learnt it from an early version by Carlo Traverso alone of [Robbiano-Traverso]. In the first part of our article, we extend this approach to a unified presentation for the homogeneous and the non-homogeneous cases.

In the case of non-equidimensional ideals, this definition of the degree does not depend on the components of lower dimension and the embedded components. For taking the isolated components into account, Masser and Wüstholz have introduced another notion of degree, which is the sum of the degrees of the isolated components of any dimemsion. They have proved that, for an ideal generated by polynomials of degrees   the degree of the intersection of the isolated components of height   is at most   So, if the height of the ideal is   the Masser–Wüstholz degree is at most  

The main new result of this article is that   bounds not only the degree of the intersection of the isolated components of height   but also the sum of the degrees of all isolated primary components of height at most   Moreover, although Masser–Wüstholz proof involves analytic geometry, our proof is purely algebraic.

Gradation and Hilbert series edit

In this article, we work with the polynomial ring   in   indeterminates over a fiels $K$. This ring is a graded by the degree, and this gradation extends to homogeneous ideals and quotients by such ideals. In all these graded modules, the homogeneous part of degree $d$ is a finite-dimensional $K$-vector space for every integer $d$.

Definition —  If

 

is a graded object, where $A_d$ is a finite-dimensional $K$-vector space for evry $d$, then the Hilbert series of $A$ is the formal power series

 

The main property of Hilbert series is to be additive under exact sequence,

Proposition —  If

 

is an exact sequence of homomorphisms of graded modules, then

 

Proof. Results immediately from the similar formula for dimension of vector spaces.

Proposition —  If $A$ is a graded algebra, and $x$ a homogeneous element of degree $d$ that is not a zero divisor in $A$, then  

Proof. Results immediately from the exact sequence

 

where $A^{[d]}$ is $A$ with its gradation shifted by $d$, which multiply its Hilbert series by $t^d$.

Corollary —  The Hilbert series of   is

 

Proof. Proof by recurrence on $n$, using the preceding proposition with   The recurrence starts with the trivial result  

Corollary —  If   is a regular sequence of homogeneous elements in   of respective degrees   (this means that each   is not a zero divisor modulo the preceding  s) then

 

Theorem —  The Hilbert series of any graded   module $M$ (without elements of negative degree) can be summed as

 

where   is a polynomial not divisible by  , and   is a nonnegative integer not larger than  .

Proof. If   is a free graded module whose basis elements have degrees  , then the form of  , and the addivity of Hilbert series under direct sums show that  . In the general case, this results from Hilbert's syzygy theorem, which asserts that every graded   module has a free resolution of length at most  . The additivity of Hilbert series under exact sequences inplies thus that every Hilbert series is an alternating sum of Hilbert series of free modules.  

Definition or theorem —  If   is a homogeneous ideal in  , and   where   is a polynomial not divisible by  , then   is the dimension of  , and   is the degree of  

In this article, we take the preceding statement as a definition. However, we will prove, below, that this definition is equivalent with the usual ones. In particular, the dimension of   is the Krull dimension of  .

General settings edit

  • n: number of variables
  • k, ground field, supposed to be infinite for some results
  •  
  •   polynomials of degrees   such that   (this can always be achieved by permuting the polynomials)
  • Degree of an ideal, denoted   as defined in Hilbert series and Hilbert polynomial
  • Strong degree of an ideal, denoted   the sum of the degrees of the isolated primary components

Lemmas edit

Lemma 1 (Regularity lemma) — By replacing   by   for sufficiently generic scalars   one can suppose that, for every i, and every associated prime p of  , if   then   for every   (must be restated for including the homogeneous case)

Proof

Standard, see, for example, [Kaplanski, Commutative rings]. Idea: if some  

Degree of an affine ideal edit

When considering non-homogeneous polynomials, that is the filtration by the degree that is considered, rather than the gradation.

Deg vs deg edit

By definition of Deg, one has Deg I = deg I for every primary ideal I.

Lemma — Let I be a strictly equi-dimensional ideal of dimension d, that is an ideal whose all associated primes have dimension d. Let J be an ideal of dimension at most d that has no associated prime in common with I. Then, if dim J = d, then

 

and if dim J < d, then

 

Proof: The exact sequence

 

implies that

 

As all quotient rings have dimension at most d, the Hilbert series can be written   and one has P(1) = 0 if the dimension is smaller than d. Otherwise P(1) is the degree of the ideal.

One has dim (I + J) < d. In fact, every minimal prime p of I + J must contain a minimal prime p1 of I and a minimal prime p2 of J. If the dimension of p would be d, this would imply that both p1 and p2 have dimension d, and therefore that they are equal to p, which is excluded by the hypotheses.

The result follows immediately by removing the denominators   in the above equation between Hulbert series, and substituting t for 1 in the resulting equality of polynomials.

Proposition — If I is an ideal of dimension d, then deg I is the sum of the degrees of the (isolated) primary components of I of dimension d (by definition, deg q = Deg q for a primary ideal q)

Proof: Let   be the primary components of dimension d, and   be the intersection of all other primary components. The dimension of   is thus smaller than d. The result is obtained by applying recursively the above lemma to   and   for          

Corollary — For every ideal I, one has deg I ≤ Deg I, and the equality is true if and only if all minimal primes have the same dimension.

Corollary — Let   be the intersection of the isolated primary components of an ideakl   Then   and  

Strong Bézout inequality edit

The strong Bézout inequality bounds the MW-degree of an ideal in terms of the degrees of its generators. It is

Theorem (Strong Bézout inequality) — Let   be nonzero polynomials of respective degrees   which are sorted in order that   If the height of the ideal   is   then

 

Moreover for every   the sum of the degrees of the isolated primary components of   of height at most   is at most  

Technical lemmas edit

Lemma 5.1 — If   and   are two ideals, and   is a polynomial, then   and   have the same minimal primes.

Moreover, if there is no inclusion between the associated primes of   and   (that is, if no associated prime of one ideal contains the other ideal), then   and a minimal primary decomposition of   is the union of minimal primary resolutions of   and  

Proof: One has

 

As the inteersection and the product of two ideals have the same radical,   and   have the same radical and thus the same minimal primes.

The hypothesis of the last assertion implies that all above inclusions are equalities, and the result follows thus immediatly.

Thus, it remains to prove that if two ideals   and   satisfy the hypothesis of the last assertion, then   In fact, no associated prime of either ideal can contain   So, there is an element   that does not belong to either associated prime. If   then   If   it follows that   The hypothesis implies thus that the inverse image in   of a primary decomposition of this localized ideal is a primary decomposition of both   and   which are thus equal.

Lemma 5.2 —  Let   be a minimal prime of an ideal   There is an element   that belongs to all other associated primes of   If   is the multiplicative set of the powers of   then the  -primary component of   is  

Proof: As   is a minimal prime, each other associated prime contains an element that is not in   The product of these elements is the desired element   The last assertion results of the classical property of stability of primary decompositions under localization.

Lemma 5.3 —  Let   be two ideals that have a common minimal prime   Let   and   be their respective  -primary components. Then   and  

Proof: The first assertion results from Lemma 5.2, since localization and intersection preserve inclusion.

By definition of Hilbert series, the inclusion   implies that each coefficient of the Hilbert series   is not smaller than the corresponding coefficient of   Thus   for   Writing these series as a rational fractions with denominator   one see that the same inequality applies to the numerators, and thus to their limits when   which are the degrees of the ideals.

Lemma 5.4 —  Let   be an ideal, and   be a polynomial. If   is a minimal prime of   such that   then   is a minimal prime of   Moreover, if   and   are the  -primary components of   and   respectively, then  

Proof: By hypothesis,   Any minimal prime   of   contains   and thus some minimal prime of   Therefore,   cannot be strictly included in   that is,   is a minimal prime of   Then, the inequality on the degrees follows directly from Lemma 5.3.

Recursion edit

Input:   of degrees   satisfying the regularity condition, that is, for   and every associated prime p of   if  , then   for all  

If   the hypothesis can be achieved by adding to   a sufficiently generic linear combination of   In the homogeneous case, the coefficients of   in the linear combination must be a homogeneous polynomial of degrees  

Let   for   We have first to study the minimal primes of   By Krull's height theorem, such the height of such a minimal prime is at most  

Lemma 6.1 — If   if a minimal prime of   of height   then   is a minimal prime of   and   for  

Proof: Let us choose recursively, for   a minimal prime   of   that is contained in   As the height of   is less than   these minimal primes cannot be all distinct. Thus, let   be the lowest index such   This implies that   and, by regularity hypothesis,   for   Therefore   Finally,   since the height of   is at most   and it is at least  , as   is a strictly increasing sequence of primes.  

Lemma 6.2 — Let   be a minimal prime of   whose height is less than   It is also a minimal prime of   Let   and   be the corresponding primary components of   and   Then  

Proof: The first assertion results immediately from lemma 6.1, and the second one is a special case of lemma 5.3. 

Corollary 6.3 — If   then  

Now, we have to study, for   the isolated primary components of   whose height is   (the case   being trivial). So, let   such that   has a minimal prime of height   Let   be a polynomial that does not belong to any minimal prime of height   of   but belongs to all other associated primes of   Let   the multiplicative set generated by  

The ideal   is the intersection of the isolated primary components of height   of   We will prove the following lemma by recursion on  

Lemma 6.4 —  For every   the sequence   is a regular sequence in   and the ideal   is unmixed of height   (that is, all its primary components have height  

Moreover,   is the intersection of some isolated primary components of height   of   whose associated prime does not contain  

Proof: The cases   and   are trivial. So we suppose that the assertions are true for some   and we prove them for  

The definition of   as the inverse image of a localization implies that   is the intersection of some primary components of   and that the minimal primes of   are also minimal primes of   None of these minimal primes can contain   In fact, if such a prime would contains   it would contain   by the regularity condition, and thus it would be a minimal prime of   of height   this is impossible since   has been chosen for having a non-empty intersection with such a prime. As   is unmixed, we can deduce that   is not a zero divisor modulo   and, using the recurrence hypothesis, that   is a regular sequence in  

This shows that the primary components of   are primary components of   that they have height   and that their associated primes do not contain    

Lemma 6.5 —  Let   the intersection of all primary components of height   of   and   be the intersection of the isolated primary components of height   of   whose associated prime do not contain   Then  

Proof: Using previous notations, and the last assertion of lemma 6.4, we have   and the two ideals have the same height. Thus

 

In the preceding lemma, we have proved that   is not a zero divisor modulo   So,

 

It results from the proof ofth epreceding lemma that   and that these ideals have the same height. So

 

The result follows immediately, by combining these inequalities.  

End of the proof edit

Given an ideal   let us denote by   the sum of the degrees of its isolated primary components of height at most   The strong Bézout inequality results immediately, by recurrence on   from the following result.

Proposition —  Using notations of preceding section, one has for every   and every  

If   and   then

 

Otherwise, that is, if   or  

 

(By Krull's height theorem, one has always  )

Proof: The second inequality is obtained by summing the inequalities of lemma 6.2 over the minimal primes of   whose height is   The first inequality is obtained by adding to the inequality of lemma 6.5 the inequalities of lemma 6.2, relaxed to   This gives the result since no minimal prime involved in lemma 6.2 is a minimal primes of the ideal   of lemma 6.5.