• No results found

The two theorems sound different, but are equivalent nevertheless.

N/A
N/A
Protected

Academic year: 2021

Share "The two theorems sound different, but are equivalent nevertheless."

Copied!
9
0
0

Loading.... (view fulltext now)

Full text

(1)

5. Hilbert’s Nullstellensatz

Gauss proved several times that the field of complex numbers C is algebraically closed, i.e., if C ⊂ K is any algebraic extension of fields then necessarily C = K, or equivalently every non-constant polynomial f (X) ∈ C[X] has a root in C. This is in contrast with the field of real numbers R, where for example the polynomial X2+ 1 has no real root.

Now consider a polynomial ring over a field k with several variables k[X1, . . . , Xn] and I an

ideal. We say that v = (v1, v2, . . . , vn) ∈ kn is a common zero of the ideal if every polynomial

f (X1, . . . , Xn) in I vanishes on v, i.e. when we substitute vi for the variable Xi for all i, we obtain

0:

f (v) = f (v1, . . . , vn) = 0.

The evaluation-at-v map

evv : k[X1, . . . , Xn] → k : f 7→ f (v)

is a surjective ring homomorphism; its kernel Mv is therefore a maximal ideal. Hence v is a common

zero for I if and only if I ⊆ Mv.

Write V(I) ⊂ kn for the collection of all the common zeros of I, called the zero set of I. This zero set V(I) can be empty, for example when I contains the constant function 1, i.e. when I = k[X1, . . . , Xn]. For the ideal I = (X12+ 1)C R[X1] the zero set V(I) in R1 is empty. We can

reformulate : A field k is algebraically closed if and only if the zero set V(I) ⊂ k1 is non-empty for any non-trivial ideal I ⊂ k[X1], i.e. 1 6∈ I.

One version of a theorem of Hilbert, called Hilbert’s Nullstellensatz, generalizes this for all n. Theorem 5.1. A field k is algebraically closed if and only if the zero set V(I) ⊂ kn is non-empty for any n ≥ 1 and non-trivial ideal I ⊂ k[X1, . . . , Xn], i.e. 1 6∈ I.

In particular, if I ⊂ C[X1, . . . , Xn] is an ideal and I 6= C[X1, . . . , Xn] then there does exist a

common zero for I in Cn.

An algebraic version of this theorem is as follows.

Theorem 5.2. A field k is algebraically closed if and only if for any n ≥ 1 any maximal ideal M of k[X1, . . . , Xn] is of the form M = Mv for some v ∈ V .

The two theorems sound different, but are equivalent nevertheless.

Proof of equivalence of the two theorems. Let k be algebraically closed, MC k[X1, . . . , Xn] a

max-imal ideal and suppose Theorem 5.1 holds. Then there exists a v ∈ kn that is a common zero of M, or M ⊆ Mv. Since M is maximal it follows that M = Mv.

On the other hand let I be a non-trivial ideal of k[X1, . . . , Xn] and suppose Theorem 5.2 holds.

By Krull’s theorem there exist a maximal ideal M containing I. By assumption there is a v ∈ kn such that M = Mv. So I ⊂ Mv, and so v is a common zero for I. 

To prove the theorems we will use two useful lemmas.

Lemma 5.1. Let R ⊂ S be an integral extension of integral domains. Then R is a field if and only S is a field.

(2)

Proof. Suppose R is a field and s ∈ S, s 6= 0. Let sd+ r1sd−1+ . . . + rd = 0 be an integrality

relation of smallest degree d, where ri ∈ R. Put s0 := sd−1+ r1sd−2+ . . . + rd−1 then ss0+ rd= 0

and by minimality of the degree s0 6= 0. Suppose rd= 0 then s0s = 0. But by assumption S has no

zero divisors. So rd6= 0 and rdhas an inverse in R. We have ss0= −rdor s−1= −r−1d s

0 exists. So

S is a field.

Suppose S is a field and r ∈ R, r 6= 0. Now r−1 ∈ S exists and satisfies an integrality relation over R, say (r−1)d+ r1(r−1)d−1+ . . . + rd= 0. Multiplying by rd−1 we obtain r−1 = −(r1+ r2r +

. . . + rdrd−1) ∈ R. So R is a field. 

The second lemma uses Noether normalization.

Lemma 5.2. Let MC k[X1, . . . , Xn] be a maximal ideal, where k is a field. Then the field extension

k ⊂ K := k[X1, . . . , Xn]/M is finite, i.e. dimkK < ∞ and hence in particular the extension is

algebraic.

Proof. Since K is an affine k-algebra, Noether normalization says that there is a d ≥ 0 and y1, . . . , yd ∈ K that are algebraically independent over k and such that k[y1, . . . , yd] ⊂ K is

in-tegral. Since K is a field, by the previous lemma k[y1, . . . , yd] is a field. This is only possible when

d = 0, hence k ⊂ K is an integral extension. 

Now we are ready to prove the Nullstellensatz.

Proof of Theorem 5.2. Let k be algebraically closed and M a maximal ideal of k[X1, . . . , Xn]. Put

K := k[X1, . . . , Xn]/M and xithe class of Xi in K. By Lemma 5.2 the extension k ⊆ K is algebraic.

Let s ∈ K have minimum polynomial f ∈ k[X]. Since k is algebraically closed f (X) has a zero c ∈ k, so by polynomial division there is a monic polynomial g(X) ∈ k[X] such that f (X) = (X − c)g(X). If s 6= c then g(s) = 0 and so g is divisible by f , contradiction. So c = s. We proved that k = K. In particular there are vi ∈ k such that xi = vi, and for any polynomial f (X1, . . . , Xn) we get

f (x1, . . . , xn) = f (v) or stated differently f − f (v) = f (X1, . . . , Xn) − f (v) ∈ M. In particular

when f ∈ Mv we get f = f − f (v) ∈ M, or Mv⊆ M. Since Mv is a maximal ideal, it follows that

Mv = M. 

5.1. If I ⊆ J C k[X1, . . . , Xn] then clearly V(I) ⊇ V(J ). Different ideals can have the same zero

set, for example (f1, f2)) and (f110, f22) have the same zero-set. Let IC R be an ideal, the radical of

I, written √I, is defined as:

I = {f ∈ R; ∃n ≥ 1 : fn∈ I}. It contains I and is an ideal itself and V(I) = V(√I).

Let Y ⊂ kn be any subset of the n-dimensional affine space over the field k. Then we define its

vanishing ideal

I(Y ) := {f ∈ k[X1, . . . , Xn]; ∀v ∈ Y : f (v) = 0} = ∩v∈YMv.

If Y ⊂ X then I(Y ) ⊇ I(X). Associated to this ideal we have a zero set called the (Zariski)-closure of Y :

Y := V(I(Y )). And we could repeat, but Y = Y , by the next lemma.

(3)

Lemma 5.3. Let k be any field, Y ⊂ kn a subset and J C k[X1, . . . , Xn] an ideal.

(i) I(Y ) = I(V(I(Y ))); (ii) V(J ) = V(I(V(J ))).

Proof. (i) It is clear from the definition (I hope) that Y ⊆ V(I(Y )), hence I(Y ) ⊇ I(V(I(Y ))). Let f ∈ I(Y ) then f (v) = 0 for all v ∈ V(I(Y )), hence f ∈ I(V(I(Y ))).

(ii) It is clear from the definition (I hope) that J ⊆ I(V(J )), hence V (J ) ⊇ V(I(V(J ))). Let v ∈ V (J ) then f (v) = 0 for all f ∈ I(V(J )), hence v ∈ V(I(V(J ))). 

Given a field k and an ideal J C k[X1, . . . , Xn] then at least

I ⊆ I(V(J ). Another version of the Nullstellensatz is the following.

Theorem 5.3. A field k is algebraically closed if and only if for any n ≥ 1 and any ideal J C k[X1, . . . , Xn] we have that

I(V(J )) = √

J .

Proof. Suppose k is not algebraically closed, then there is a non-constant f ∈ k[X1] without zeros

in k. So I(V(f )) = I(∅) = k[X1] 6=p(f).

The other direction uses a trick due to Rabinovich, cf. [2, §15.3]. Let g ∈ I(V(J )). By Hilbert’s basis theorem J is finitely generated, say by f1, . . . , fm. Consider the ideal ˜J of k[X1, . . . , Xn, Xn+1]

generated by f1, . . . , fm and Xn+1g − 1. Suppose (v1, . . . , vn, vn+1) ∈ kn+1 is a common zero for

the ideal ˜J . Then fi(v1, . . . , vn) = 0 for all i and vn+1 · g(v1, . . . , vn) − 1 = 0. In particular

(v1, . . . , vn) ∈ knis a common zero for J , so g(v1, . . . , vn) = 0 as well. But then −1 = vn+1· 0 − 1 =

vn+1· g(v1, . . . , vn) − 1 = 0 gives a contradiction. We conclude that the ideal ˜J does not have a

common zero in kn+1. By the first version of the Nullstellensatz, Theorem 5.1, we conclude that 1 ∈ ˜J . So there are hi ∈ k[X1, . . . , Xn+1] such that

1 =

n

X

i=1

hifi+ hn+1(Xn+1g − 1).

Let N > 1 be larger then the xn+1-degree of any of the hi’s, then gi := xNhi n+1

becomes a polynomial in x1

n+1 with coefficients in k[X1, . . . , Xn], i.e., gi is an element of k[X1, . . . , Xn, 1 Xn+1]. Say  1 Xn+1 N = n X i=1 gi(X1, . . . , Xn, 1 Xn+1 ) · fi+ gn+1(X1, . . . , Xn, 1 Xn+1 )(Xn+1g − 1).

Now substitute Xn+1 := 1g to get

gN = n X i=1 gi(X1, . . . , Xn, g) · fi+ gn+1(X1, . . . , Xn, g)( 1 gg − 1) = n X i=1 gi(X1, . . . , Xn, g) · fi,

where each gi ∈ k[X1, . . . , Xn]. So a power of g is contained in the ideal (f1, . . . , fm) = J , or

g ∈√J . 

Remark. So when k is not algebraically closed then generally I(V(J )) is bigger than √J . Can we nevertheless give a description of I(V(J )) for a given field? For the field of real numbers such a description is known. But for Q?!

(4)

5.2. Generators of maximal ideals of polynomial rings. Any maximal ideal of a polynomial ring over a field with n variables can be generated by n polynomials, and can be constructed iteratively from minimal polynomials, as shown in the proof of the next proposition.

Proposition 5.1. Let m ⊂ k[X1, . . . , Xn] be a maximal ideal, where k is a field. Then m can be

generated by n polynomials.

Proof. We shall use induction on n; the case n = 0 is trivial and if n = 1 every ideal is generated by one element. Suppose the result is true for less than n variables. Put K = k[X1, . . . , Xn]/m

and xi := Xi, the class of Xi in K. Put m0 := k[X1, . . . , Xn−1] ∩ m, then k[X1, . . . , Xn1]/m 0 '

k[x1, . . . , xn−1]. Since the xi’s are algebraic over k, each k[x1, . . . , xn−1] = k(x1, . . . , xn−1) is a field,

by Lemma 5.1, and so m0 is even a maximal ideal of k[X1, . . . , Xn−1].

By induction we know generators F1, . . . , Fn−1 for m0. Since xn is algebraic over the field

k(x1, . . . , xn−1), it has a minimum polynomial

f (T ) = Td+ r1Td−1+ . . . + rd∈ k(x1, . . . , xn−1)[T ].

Choose polynomials h1, . . . , hd∈ k[X1, . . . , Xn−1] such that hi(x1, . . . , xn−1) = ri, and put

Fn:= Xnd+ h1Xnd−1+ . . . + hd.

Then Fn(x1, . . . , xn) = 0, i.e., Fn ∈ m. We have m ⊇ (F1, . . . , Fn) and shall prove that m =

(F1, . . . , Fn).

Let F ∈ m, hence F (x1, . . . , xn) = 0. Expand F as a polynomial in Xn, say

F = f0Xnm+ f1Xnm−1+ . . . + fm,

with fi ∈ k[X1, . . . , Xn−1]. Since F (x1, . . . , xn−1, Xn) is divisible by the minimal polynomial f (Xn),

there exists a polynomial h(Xn) ∈ k(x1, . . . , xn−1)[Xn] such that

F (x1, . . . , xn−1, Xn) = f (Xn)h(Xn).

We can choose a polynomial H ∈ k[X1, . . . , Xn] such that H(x1, . . . , xn−1, Xn) = h(Xn).

Further-more, since Fm is a monic polynomial we can assume that H has the same main coefficient as F ,

i.e., H is of the form

H = f0Xnm−d+ plus lower order terms in Xn.

So

F − FnH ∈ m

and the Xn-degree of F − FnH is smaller than the Xn-degree of F . By repeating this procedure,

we can find a polynomial G such that

F − FnG ∈ m

has Xn-degree is 0, i.e., F − FnG ∈ m0 = (F1, . . . , Fn−1) or

m⊆ (F1, . . . , Fn)C k[X1, . . . , Xn].

(5)

6. Cohen-Seidenberg theorems and Noether normalisation

Integral extension of rings have more nice properties than we have already mentioned. We will need them to get more corollaries from the Noether normalization, Theorem 4.1, and Hilbert’s Nullstellensatz.

We collect the needed results in one omnibus theorem, due to Cohen and Seidenberg (see also [2, §15.3, Theorem 26]).

Let R be a ring and

p0 ⊂ p1 ⊂ . . . ⊂ pm ⊂ . . . ⊂ pn

a sequence of prime ideals. We shall call it a chain of prime ideals of length n (there are n chains) if for all 0 ≤ i < n we have pi 6= pi+1. The chain is saturated if it has the property that if q is any

prime ideal such that pi ⊆ q ⊆ pi+1, for some 0 ≤ i < n, then q = pi or q = pi+1. We shall say

that R has finite (Krull) dimension d if R has a chain of prime ideals of length d, but no chain of length > d.

Theorem 6.1 (Cohen-Seidenberg). Let R ⊂ S be an integral extension of rings.

(i) Let PC S be a prime ideal, put p = P ∩ R. Then P is a maximal ideal in S if and only if p is a maximal ideal of R.

(ii) (Incompatibility theorem) Suppose Q ⊇ P are prime ideals of S such that Q ∩ R = P ∩ R. Then P = Q.

(iii) (Lying-over theorem) Suppose p is a prime ideal of R. Then there is a prime ideal P of S such that P ∩ R = p.

(iv) (Going-up theorem) Let m < n . Suppose there is a chain of prime ideals in S P0⊂ P1 ⊂ . . . ⊂ Pm,

and a chain of prime ideals in R

p0 ⊂ p1 ⊂ . . . ⊂ pm ⊂ . . . ⊂ pn

such that Pi∩ R = pi for 1 ≤ i ≤ m. Then the chain of prime ideals in S can be extended to a

chain

P0⊂ P1 ⊂ . . . ⊂ Pm ⊂ . . . ⊂ Pn such that Pi∩ R = pi for all i.

(v)(Going-down theorem) Suppose S is an integral domain and that R is integrally closed in its fraction field and still R ⊂ S an integral extension. Let 0 < m ≤ n. Suppose there is a chain of prime ideals in S

Pm ⊂ Pm+1⊂ . . . ⊂ Pn,

and a chain of prime ideals in R

p0 ⊂ p1 ⊂ . . . ⊂ pm ⊂ . . . ⊂ pn

such that Pi∩ R = pi for m ≤ i ≤ n. Then the chain of prime ideals in S can be extended to a

chain

P0⊂ P1 ⊂ . . . ⊂ Pm ⊂ . . . ⊂ Pn such that Pi∩ R = pi for all i.

(6)

The following lemma provides examples of domains that are integrally closed in their field of fractions.

Lemma 6.1. Let R be a factorial domain. Then R is integrally closed in its field of fractions. In particular, this is the case for a polynomial ring over a field.

Proof. Let the fraction ab be integral over R, we can suppose that a and b have no common irre-ducible factors. There is an integrality relation over R:

(a b) d+ r 1( a b) d−1+ . . . + r d= 0, where ri∈ R. So ad= −b(r1ad−1+ r2ad−2b + . . . + rdbd−1).

So any irreducible factor of b is also an irreducible factor of a, but a and b have no irreducible factors in common. We conclude that b is a unit in R, so ab = b−1a ∈ R. We proved that R is

integrally closed in its field of fractions. 

Corollary 6.1. Suppose R ⊂ S is an integral extension. Then they have the same Krull-dimension. Proof. Suppose

p0 ⊂ p1 ⊂ . . . ⊂ pd

is a chain of prime ideals in R. By Lying-over there is a prime ideal P0 in S, such that P0∩ R = p0,

then by Going-up there is a chain of prime ideals in S P0 ⊂ P1⊂ . . . ⊂ Pd

lying over our chain. So the Krull-dimension of S is at least as big as the Krull-dimension of R. On the other hand, suppose

P0 ⊂ P1⊂ . . . ⊂ Pd

is a chain of prime ideals in S. Put pi := Pi∩ R. Then pi is a prime ideal. If pi = pi+1 we get a

contradiction with the Incompatibility theorem, so we get a chain on prime ideals in R. p0 ⊂ p1 ⊂ . . . ⊂ pd

So the Krull-dimension of R is at least as big as the Krull-dimension of S.  Corollary 6.2. (i) If S is an affine ring over a field k, then the Krull-dimension of S equals the dimension of R as affine k-algebra.

(ii) In particular, a polynomial ring over a field k[X1, . . . , Xn] has Krull-dimension n. In fact

every maximal saturated chain of prime ideals in k[X1, . . . , Xn] has length n.

Proof. By Noether normalization there are elements x1, . . . , xn in S that are algebraically

inde-pendent over k and such that R := k[x1, . . . , xn] ⊂ S is an integral extension of rings and we can

apply the corollary. So it suffices to prove the special case of a polynomial ring with n-variables S = k[X1, . . . , Xn].

Put P0 = (0) and Pi = (X1, . . . , Xi), for 1 ≤ i ≤ n. Then S/Pi is itself isomorphic to a

polynomial ring, so Pi is a prime ideal. So we get a chain of prime ideals of length n

(7)

So the Krull-dimension of k[X1, . . . , Xn] is at least n.

Let

P0⊂ P1 ⊂ . . . ⊂ Pe

be any maximal saturated chain of prime ideals of S. By Noether normalisation, there are elements x1, . . . , xn algebraically independent over k, such that R := k[x1, . . . , xn] ⊂ S is integral and there

are integers n ≥ d0 ≥ d1 ≥ . . . ≥ de≥ 0 such that

Pi∩ R = (xdi+1, xdi+2, . . . , xn)

Since there are only n + 1 different integers between 0 and n, if e > n we have e + 1 > n + 1 ideals Pi, there must be therefore an i such that Pi∩ R = Pi+1∩ R. But that is a contradiction with

the Incompatibility theorem.

Suppose now that e < n. If d0 < n then (0) 6= p0. By Going-down there exists a prime ideal

P−1⊂ P0 such that P−1∩ R = (0), and we have extended our chain. But this is impossible, since

the original chain was already a maximal saturated chain. So d0 = n. Let i ≥ 1 be minimal such

that di > n − i; in particular di−1 = n − i + 1, and there is a linear prime ideal, say q strictly

between pi−1 and pi. Now R/pi−1 ⊂ S/Pi−1 is also integral, and since R/pi−1 also a polynomial

ring we can use Going-down again starting with 0 6= q/pi−1 ⊂ pi/pi−1. By Going-down and the

correspondence theorem we get a prime ideal Q in S containing Pi−1 and contained in Pi, such

that Q ∩ R = q. We conclude again that our chain was not saturated, which is a contradiction.

Hence e = n. 

We say that a ring is catenary if for every two prime ideals p ⊆ p0 every saturated chain of prime ideals starting at p and ending in p0 has the same length.

Corollary 6.3. Any affine ring over a field is catenary.

Proof. Since any affine k-algebra is the quotient of a polynomial ring k[X1, . . . , Xn], using the

correspondence theorem it suffices to prove that polynomial rings are catenary rings. Fix two prime ideals p ⊆ p0 of the polynomial ring. Suppose there are two saturated chains between p and p0 with different lengths. We can complete those two chains to two maximal saturated chains of different lengths. But we just proved, that two maximal saturated chains of prime ideals in a

polynomial ring have the same length. Contradiction. 

6.1. Let IC k[X1, . . . , Xn] be an ideal. We proved that if k is algebraically closed,p(I) = ∩m⊆Im,

where the intersection is over all maximal ideal containing m. This remains true if k is not alge-braically closed (we shall use some facts from field theory to prove this).

Proposition 6.1. Let I be an ideal of an affine k-algebra S. Then the radical of I is the intersection of all maximal ideals containing I:

I = ∩m⊇Im.

Proof. By the correspondence theorem we can assume that S = k[X1, . . . , Xn] is a polynomial ring

over the field k. Let I be generated by f1, . . . , fm.

There exists an algebraic extension k ⊂ L, where L is an algebraically closed field. We shall not prove this here.

(8)

Then k[X1, . . . , Xn] ⊂ L[X1, . . . , Xn] is an integral extension of rings. Let h ∈ ∩m⊇Im. Let M be

a maximal ideal of L[X1, . . . , Xn] that contains every fi, then its intersection m = M∩k[X1, . . . , Xn]

is a maximal ideal (by the Cohen-Seidenberg theorem) containing I. So if we write Ie for the ideal in L[X1, . . . , Xn] generated by I, we get

f ∈ ∩m⊇Im⊂ ∩M⊇IeM=

√ Ie

by the Nullstellensatz. So there is an N and h1, . . . , hmıL[X1, . . . , Xn] such that

fN =

m

X

i=1

hifi.

Let K be the finite dimensional algebraic extension of k generated by the finitely many coefficients of the hi’s, so now hi∈ K[X1, . . . , Xn]. Let b1, . . . , bs be a k-basis for K, where b1 = 1. Any z ∈ K

can be written uniquely as z =Ps

i=1cibi, with ci ∈ k. Extend the projection

π : K → k : π(z) = c1b1= c1

to a k[X1, . . . , Xn]-linear projection π : K[X1, . . . , Xn.

Now apply π to the equation of fN, to obtain: fN =

m

X

i=1

π(hi)fi.

(9)

D´epartement de math´ematiques et de statistique, Universit´e de Montr´eal, C.P. 6128, succursale Centre-ville, Montr´eal (Qu´ebec), Canada H3C 3J7

References

Related documents

Consequently, the local dynamic properties of the general equilibrium model with heterogeneous agents and those of the planner’s problem with the welfare weights fixed at their

discharged from the trust, or refuses or becomes, in the opinion of a principal civil court of original jurisdiction, unfit or personally incapable to act in the trust, or accepts

The FSC logo, the initials ‘FSC’ and the name ‘Forest Stewardship Council’ are registered trademarks, and therefore a trademark symbol must accompany the.. trademarks in

In this section, I analyse my hypotheses; local ethno-political exclusion and diversity reduces the likelihood of nonviolent action (H1a, H1b) and higher food price spikes increase

· The study tour gave me insights about how the European Commission works and how CSR can be of positive influence on the growth of companies.

Material and Methods: The dysfunction and quality of life scores from 25 children with ECC were evaluated before treatment (T0), one month (T1) and three months after treatment

Even for those who choose not to incorporate social or economic justice ideals into their clinical program, practice ready can, and should, mean preparing law students to