Commutative Algebra/Irreducibility, algebraic sets and varieties

From testwiki
Revision as of 21:54, 17 February 2020 by imported>Eugcomax
(diff) ← Older revision | Latest revision (diff) | Newer revision → (diff)
Jump to navigation Jump to search

Irreducibility

Template:TextBox

Some people (topologists) call irreducible spaces hyperconnected.

Template:TextBox

Proof 1: We prove 1. 2. 3. 4. 1.

1. 2.: Assume that X=AB, where A, B are proper and closed. Define O:=XA and U:=XB. Then O,U are open and

OU=(XA)(XB)=X(AB)=

by one of deMorgan's rules, contradicting 1.

2. 3.: Assume that UX is open but not dense. Then A:=U is closed and proper in X, and so is B:=XU. Furthermore, X=AB, contradicting 2.

3. 4.: Let AX be closed such that A. By definition of the closure, AA, which is why A is a non-dense open set, contradicting 3.

4. 1.: Let O,UX be open and non-empty such that OU=. Define A:=XO. Then A is a proper, closed subset of X, since OXA. Furthermore, OA, which is why A has non-empty interior.

Proof 2: We prove 1. 4. 3. 2. 1.

1. 4.: Assume we have a proper closed subset A of X with nonempty interior. Then XA and A are two disjoint nonempty open subsets of X.

4. 3.: Let OX be open. If O was not dense in X, then O would be a proper closed subset of X with nonempty interior.

3. 2.: Assume X=AB, A,BX proper and closed. Set O:=XB. Then AO, and hence O is not dense within X.

2. 1.: Let O,UX be open. If they are disjoint, then X=(XO)(XU).

Remaining arrows:

1. 3.: Assume UX open, not dense. Then XU is nonempty and disjoint from U.

3. 1.: Let O,UX be open. If they are disjoint, then OXU and thus O is not dense.

2. 4.: Let AX be proper and closed with nonempty interior. Then X=A(XA).

4. 2.: Let X=AB, A,BX proper and closed. Then XAA.


We shall go on to prove a couple of properties of irreducible spaces.

Theorem 21.3:

Every irreducible space X is connected and locally connected.

Proof:

1. Connectedness: Assume X=U˙O, U,O open, non-empty. This certainly contradicts irreducibility.

2. Local connectedness: Let xVX, where V is open. But any open subset of X is connected as in 1., which is why we have local connectedness.

Theorem 21.4:

Let X be an irreducible space. Then X is Hausdorff if and only if |X|1.

Proof:

If |X|1, then X is trivially Hausdorff. Assume that X is Hausdorff and contains two distinct points xy. Then we find Ux,UyX open such that xUx, yUy and UxUy=, contradicting irreducibility.

Theorem 21.5:

Let X,Y be topological spaces, where X is irreducible, and let f:XY be a continuous function (i.e. a morphism in the category of topological spaces). Then f(X) is irreducible with the subspace topology induced by Y.

Proof: Let O,U be two disjoint non-empty open subsets of f(X). Since we are working with the subspace topology, we may write O=f(X)V, U=f(X)W, where V,WY are open. We have

f1(O)=f1(f(X)V)=f1(V) and similarly f1(U)=f1(W).

Hence, f1(O) and f1(U) are open in X by continuity, and since they further are disjoint (since if xf1(O), then f(x)O and thus f(x)U) and non-empty (since e.g. if yO, since Of(X), y=f(x) for an xX and hence xf1(O)), we have a contradiction.

Corollary 21.6:

If X is irreducible, Y is Hausdorff and f:XY is continuous, then f is constant.

Proof: Follows from theorems 21.4 and 21.5.

We may now connect irreducible spaces with Noetherian spaces.

Template:TextBox

Proof:

First we prove existence. Let AX be closed. Then either A is irreducible, and we are done, or A can be written as the union of two proper closed subsets A=B1B2. Now again either B1 and B2 are irreducible, or they can be written as the union of two proper closed subsets again. The process of thus splitting up the sets must eventually terminate with all involved subsets being irreducible, since X is Noetherian and otherwise we would have an infinite properly descending chain of closed subsets, contradiction. To get the last condition satisfied, we unite any subset contained within another with the greater subset (this can be done successively since there are only finitely many of them). Hence, we have a decomposition of the desired form.

We proceed to proving uniqueness up to order. Let A=B1Bn=C1Cm be two such decompositions. For k{1,,n}, we may thus write Bk=(BkC1)(BkCm). Assume that there does not exist j{1,,m} such that BkCjBk=(BkCj). Then we may define S1:=(BkC1) and then successively

Sl+1:=Sl(BkCl+1)

for 1l<m. Then we set l=1 and increase l until Sl(BkCl+1) is a decomposition of Bk into two proper closed subsets (such an l exists since it equals the first l such that Sl(BkCl+1)=Bk). Thus, our assumption was false; there does exist j{1,,m} such that BkCj. Thus, each Bk is contained within a Cj, and by symmetry Cj is contained within some Bk. Since by transitivity of this implies BkBk, k=k and Cj=Bk. For a fixed k, we set σ(k)=j, where j is thus defined (j is unique since otherwise there exist two equals among the C-sets). In a symmetric fashion, we may define τ(j)=k, where Bk=Cj. Then τ and σ are inverse to each other, and hence follows n=m (sets with a bijection between them have equal cardinality) and the definition of σ, for example, implies that both decompositions are equal except for order.

Exercises

  • Exercise 21.1.1: Let X be an irreducible topological space, and let OX be open. Prove that O is irreducible.

Algebraic sets and varieties

Template:TextBox

The following picture depicts three algebraic sets (apart from the cube lines):

The orange surface is the set V(f1), the blue surface is the set V(f2), and the green line is the intersection of the two, equal to the set V({f1,f2}), where

f1(x,y,z)=z+y3 and
f2(x,y,z)=x+z2.

Three immediate lemmata are apparent.

Lemma 21.9:

STV(T)V(S).

Proof: Being in V(T) is the stronger condition.

Lemma 21.10 (formulas for algebraic sets):

Let 𝔽 be a field and set R:=𝔽[x1,,xn]. Then the following rules hold for algebraic sets of 𝔽n:

  1. V(S)=V(S) (SR a set)
  2. V(R)= and V()=𝔽n
  3. V(I1)V(Ik)=V(I1Ik) (I1,,IkR ideals)
  4. jJV(Sj)=V(jJSj) (SjR sets)

Proof:

1. Let i:=j=1krjsjS. If x=(x1,,xn)V(S) follows i(x)=0. This proves . The other direction follows from lemma 21.9.

2. V(R)= follows from the constant functions being contained within R, and V() gives no condition on the points of 𝔽n to be contained within it.

3. follows by

xV(I1)V(Ik)j{1,,k}:xV(Ij)fIj:f(x)=0fI1Ik:f(x)=0,

since clearly I1IkIj.

We will first prove for the case k=2. Indeed, let xV(I1)V(I2), that is, neither xV(I1) nor xV(I2). Hence, we find a polynomial fI1 such that f(x)0 and a polynomial gS2 such that g(x)0. The polynomial fg is contained within I1I2 and (fg)(x)=f(x)g(x)0, since every field is an integral domain. Thus, xV(I1I2).

Assume holds for k1 many sets. Then we have

V(I1Ik)=V((I1Ik1)Ik)V(I1Ik1)V(Ik)V(I1)V(Ik1)V(Ik).

4.

xjJV(Sj)jJ:xV(Sj)jJ:fSj:f(x)=0fjJSj:f(x)=0xV(jJSj).

From this lemma we see that the algebraic sets form the closed sets of a topology, much like the Zariski-closed sets we got to know in chapter 14. We shall soon find a name for that topology, but we shall first define it in a different way to justify the name we will give.

Lemma 21.11:

Let 𝔽 be a field and I𝔽[x1,,xn]. Then

V(I)=V(r(I));

we recall that r(I) is the radical of I.

Proof: "" follows from lemma 21.9. Let on the other hand xV(I) and gr(I). Then gkI for a suitable k. Thus, gk(x)=g(x)k=0. Assume g(x)0. Then g(x)k0, contradiction. Hence, xV(r(I)).

From calculus, we all know that there is a natural topology on n, namely the one induced by the Euclidean norm. However, there exists also a different topology on n, and in fact, on 𝔽n for any field 𝔽. This topology is called the Zariski topology on 𝔽n. Now the Zariski topology actually is a topology on SpecR, for R a ring, isn't it? Yes, and if R=𝔽[x1,,xn], then 𝔽n is in bijective correspondence with a subset of SpecR. Through this correspondence we will define the Zariski topology. So let's establish this correspondence by beginning with the following lemma.

Lemma 21.12:

Let 𝔽 be a field and set R:=𝔽[x1,,xn]. If (α1,,αn)𝔽n, then the ideal

x1α1,,xnαn

is a maximal ideal of R.

Proof:

Set

φ:𝔽[x1,,xn]𝔽,φ(f):=f(α1,,αn).

This is a surjective ring homomorphism. We claim that its kernel is given by x1α1,,xnαn. This is actually not trivial and requires explanation. The relation x1α1,,xnαnkerφ is trivial. We shall now prove the other direction, which isn't. For a given f𝔽[x1,,xn], we define f~(x1,x2,,xn):=f(x1+α1,,xn+αn); hence,

f(x1,x2,,xn)=f(x1+α1α1,,xn+αnαn)=f~(x1α1,x2α2,,xnαn).

Furthermore, f(α1,,αn)=0 if and only if f~(0,,0)=0. The latter condition is satisfied if and only if f~ has no constant, and this happens if and only if f~ is contained within the ideal x1,,xn. This means we can write f~ as an 𝔽[x1,,xn]-linear combination of x1,,xn, and inserting xjαj for xj gives the desired statement.

Hence, by the first isomorphism theorem for rings,

𝔽[x1,,xn]/x1α1,,xnαn𝔽.

Thus, 𝔽[x1,,xn]/x1α1,,xnαn is a field and hence x1α1,,xnαn is maximal.

Lemma 21.13:

Let 𝔽 be a field. Define

𝔽:={x1α1,,xnαn|(α1,,αn)𝔽n}

(according to the previous lemma this is a subset of Spec𝔽[x1,,xn], as maximal ideals are prime). Then the function

Φ:𝔽n𝔽,f((α1,,αn)):=x1α1,,xnαn

is a bijection.

Proof:

The function is certainly surjective. Let x1α1,,xnαn=x1β1,,xnβn, and assume βjαj for a certain j{1,,n}. Then xjβjx1α1,,xnαn, and thus

0αjβj=xjβj(xjαj)x1α1,,xnαn.

Thus, x1α1,,xnαn contains a unit and therefore equals 𝔽[x1,,xn], contradicting its maximality that was established in the last lemma.

Template:TextBox

It is easy to check that the sets Φ1(O), O𝔽 really do form a topology.

There is a very simple different way to characterise the Zariski topology:

Template:TextBox

Proof:

Unfortunately, for a set T𝔽[x1,,xn], the notation V(T) is now ambiguous; it could refer to the algebraic set associated to T, or to the set of prime ideals p of 𝔽[x1,,xn] satisfying Tp. Hence, we shall write the latter as V~(T) for the remainder of this wikibook.

Let A𝔽n be closed w.r.t. the Zariski topology; that is, A=Φ1(V~(T)𝔽), where Φ is the function from lemma 21.13 and TR:=𝔽[x1,,xn]. We claim that A=V(T). Indeed, for α𝔽n,

αV(T)fT:f(α)=0fT:((x1α1)|f(x1α1)|f)fT:fx1α1,,xnαnTΦ(α)Φ(α)V~(T)αΦ1(V~(T)𝔽).

Let now V(S) be an algebraic set. We claim V(S)=Φ1(V~(S)𝔽). Indeed, the above equivalences prove also this identity (with S replacing T).

In fact, we could have defined the Zariski topology in this way (that is, just defining the closed sets to be the algebraic sets), but then we would have hidden the connection to the Zariski topology we already knew.

We shall now go on to give the next important definition, which also shows why we dealt with irreducible spaces.

Template:TextBox

Often, we shall just write variety for algebraic variety.

We have an easy characterisation of algebraic varieties. But in order to prove it, we need a definition with theorem first.

Template:TextBox

Proof:

Let first T be any set such that V(T)=V(S). Then for all fT and xV(S)=V(T), f(x)=0 and hence fI(V(S)). Thus TI(V(S)).

Therefore, SI(V(S)), and hence V(I(V(S)))V(S) by lemma 21.9. On the other hand, if xV(S), then f(x)=0 for all fI(V(S)) by definition. Hence xV(I(V(S))). This proves V(S)V(I(V(S))).

Template:TextBox

Proof:

Let first p𝔽[x1,,xn] be a prime ideal. Assume that V(p)=V(S)V(T), where V(S),V(T) are two proper closed subsets of V(p) (according to lemma 21.10, all subsets of V(p) closed w.r.t. the subspace topology of V(p) have this form). Then there exist xV(p)V(T) and yV(p)V(S). Hence, there is gT such that g(x)0 and fS such that f(y)0. Furthermore, fgp since for all zV(S)V(T) either f(z)=0 or g(z)=0, but neither fp nor gp.

Let now V(S) be an algebraic set, and assume that I:=I(V(S)) is not prime. Let fgI such that neither fI nor gI. Set Jf:=I+f and Jg:=I+g. Then Jf and Jg are strictly larger than I. According to 21.17, V(Jf)V(S) and V(Jg)V(S), since otherwise JfI or JgI respectively. Hence, both V(Jf) and V(Jg) are proper subsets of V(S). But if xV(S)=V(I), then fg(x)=f(x)g(x)=0. Hence, either f(x)=0 or g(x)=0, and thus either xJf or xJg. Thus, V(S) is the union of two proper closed subsets,

V(S)=V(Jg)V(Jf),

and is not irreducible. Hence, if irreducibility is present, then I(V(S)) is prime and from 21.17 V(I(V(S)))=V(S).

Template:TextBox

Proof:

Let O1O2Ok be an ascending chain of open sets. Let Φ and 𝔽 be given as in lemma 21.13 and definition 21.14. Set Uj=Φ(Oj) for all j. Then, since Φ, being a function, preserves inclusion,

U1U2Uk.

Since 𝔽 is a Noetherian ring, so is 𝔽[x1,,xn] (by repeated application of Hilbert's basis theorem). Hence, the above ascending chain of the Uj eventually stabilizes at some N. Since Φ is a bijection, Oj=Φ1(Φ(Oj))=Φ1(Uj). Hence, the Oj stabilize at N as well.

Template:TextBox

That is, we can decompose algebraic sets into algebraic varieties.

Proof:

Combine theorems 21.19, 21.7 and 21.18.

Exercises

  • Exercise 21.2.1: Let f,g𝔽[x1,,xn]. Prove that V({fg})=V({f})V({g}).

Template:BookCat