Math 8203 (Algebraic Geometry)
Fall 1997

The Zariski Topology and Regular Functions

Copyright (©) 1997 by Joel Roberts.

Contents

The Zariski topology.

Let X be a variety. Thus, we can be considering X \subset An or X \subset Pn. In either case, we claim that the subsets of X which are algebraic (as subsets of An or Pn respectively) are the closed sets of a topology. In order to verify this fact, it is clearly sufficient to consider the case where X is An or Pn itself.

PROOF IN THE AFFINE CASE. We have V(0) = An while V(1) is empty. If Xa = V(Sa), where a runs through some index set and the Sa are subsets of K[x1, ...,xn], then:

 \cap V(S_a) = V(\cup S_a).

Thus, an arbitrary intersection of algebraic sets is algebraic. To show that a finite union of algebraic sets is algebraic, the key step is to understand the union of two algebraic sets. So, if X1 = V(S1) and X2 = V(S2), then we can check that X1 \cupX2 = V(T) where:

T = {fg | f in S1 and g in S2 }.
The key step is to observe that if p lies in V(T) but p is not in X1 , then f(p) is nonzero for some f0 in S1. Since f0(p)g(p) = 0 for every g in S2, it follows that p is a point of X2.

THE NOETHERIAN PROPERTY. By definition, a topological space is Noetherian if every descending sequence of closed subsets eventually stabilizes. Thus, if:

X1\supseteqX2\supseteq. . . \supseteqXk\supseteqXk+1 \supseteq. . . ,
then there exists M such that
XM = XM+1 = . . .

PROPOSITION. The Zariski topology is Noetherian.

Before proving this in the affine case, we state and prove a lemma.

LEMMA. Every algebraic set in An is of the form X = V(I), where I is an ideal in K[x1, ...,xn].

TO PROVE THE LEMMA, we show that if S is a subset of K[x1, ...,xn], then V(S) = V(I), where I is the ideal generated by S. This follows from the observation that elements of I are sums of the form:

g1f1 + . . . + gmfm ,
with m varying, fj in S and gj in K[x1, ...,xn].

TO PROVE THE PROPOSITION, we consider a descending sequence:

X1\supseteqX2\supseteq. . .\supseteqXk\supseteqXk+1\supseteq. . . ,
with Xk = V(Ik). We may not have started with an ascendending sequence of ideals, but can arrange that by noting that
Xk = V(Jk), where Jk = I1 + ... + Ik .
Note that Xk = X1 \cap ... \cap Xk .) We now have an ascending sequence of ideals:
J1 \subseteq J2 \subseteq . . . \subseteq Jk \subseteq Jk+1 \subseteq . . .
It is known that K[x1,...,xn] is Noetherian, so that every ascending sequence of ideals eventually stabilizes. Therefore our descending sequence of closed subsets also does.

DISTINGUISHED OPEN SETS (AFFINE CASE). If f is an element of the polynomial ring K[x1,...,x n], we define:

D(f) = {p | p lies in An and f(p) is nonzero}
This is clearly an open subset, since D(f) is the complement in An of the closed set V(f). These are of interest because of the following result.

PROPOSITION. The family of distinguished open sets is a basis of the Zariski topology of An.

PROOF. Every ideal is a sum of principal ideals. Therefore, the closed set V(I) is the intersection of closed sets V(f), where f runs through some set of generators. Therefore the complement of V(I) is the union of the distinguished open sets V(f).

Regular functions: affine case.

Let X be a variety, and let U be an open subset. If f is a K-valued function defined on U, we will say that f is regular at a point p of U if there is an open neighborhood V on which f is expressible as a quotient of polynomials f = g/h, where h(p) is nonzero. We say that f is regular on U if it is regular at every point of U.

FOR INSTANCE:

  1. Let X = A2 . Then f(x,y) = (x + y)/ (x2 + 3xy) is regular on the open set U = D(x2 + 3xy ).
  2. Let X = V(xy - zw) in A 3. We define two functions:

Now, let U = D(z) D(w). Since xy - zw = 0 at every point of X, we have x/z = y/w at points of D(z)\capD(w). Therefore, f1 and f2 can be "glued together" to give a function f which is regular at every point of U.

SOME RINGS ASSOCIATED TO AN AFFINE VARIETY: If X and U are as above, we define \Oh(U) to be the set of regular functions on U. It is a commutative ring, since we can add and multiply the values of functions. We will usually call \Oh(U) by its obvious name, specifically the ring of regular functions on U. As a particular case we have \Oh(X), the ring of globally regular functions on X.

On the other hand, for X \subset An, we define the coordinate ring of X to be the quotient A(X) = K[x 1, ...,xn]/I(X), where I(X) is the defining ideal of X.

We have a map K[x1 ,...,xn] - > \Oh(X), by evaluating polynomials at points of X. One checks easily that this is a homomorphism. The kernel is exactly I(X), which follows easily from the definition of I(X). Therefore, we at least have an injective homomorphism:

A(X) - > \Oh(X).
More generally, let U = D(f). Then we have a homomorphism:
K[x1, ...,xn, 1/f] - > \Oh(U),
by evaluating expressions of the form g/fd at points of U. This provides an injective homomorphism
A(X) [1/f] - > \Oh(U).
To give a naïve but rigorous definition of A(X) [1/f] we can observe that:
K[x1, ...,xn, 1/f] \cong K[x 1,...,xn + 1]/(1 - xn + 1f(x1, ...,xn))
so that:
A(X) [1/f] = K[x1, ...,xn + 1]/(I(X) + (1 - xn + 1f)).

As observed in the text (with appropriate sketches), we can understand this isomorphism geometrically by observing that if X \subset An, then we have a bijective correspondence between D(f) and the subvariety Y \subset An, whose defining ideal is:

(I(X) + (1 - xn + 1f)).
The map is given by:
(x1, ...,xn) - > (x1, ...,xn , 1/f(x1, ...,xn)).

THE LOCAL RING AT A POINT. Let p be a point of an affine variety X. We define the local ring of X at p to be the set \OhX,p of germs of regular functions at p. By definition, a germ is represented by a pair, (U,g), where U is an open neighborhood of p, and g is regular on U. Two pairs (U,g), and (V,h) represent the same germ if g and h agree on some open neighborhood of p contained in U«V. If we think about it briefly, we see that germs can be added and multiplied in a natural way, so that \OhX,p is indeed a commutative ring.

A basic property of \OhX,p is that the set of nonunit (or equivalently, noninvertible) elements in \OhX,p is an ideal. Specifically, the germ represented by (U,g) is invertible if and only if g(p) is nonzero: the inverse is represented by 1/g (on a possibly smaller neighborhood). Hence the nonunit elements are represented by germs (U,g) such that g(p) = 0. The germs of functions which vanish at p obviously form an ideal. This property of \OhX,p is one way to characterize a ocal ring:

PROPOSITION. Let \Oh be a commutative ring with 1. Then the following statements are equivalent:

  1. The set of nonunit elements in \Oh is an ideal.
  2. \Oh has exacly one maximal ideal.

THE PROOF that (1) implies (2) involves the observation that an ideal in a commutative ring with 1 which is distinct from the unit ideal consists entirely of nonunit elements. To prove that (2) implies (1), let m be the unique maximal ideal, and let a be a nonunit element. Then (a) is a nonunit ideal, which is contained in some maximal ideal, by Zorn's lemma (etc. ...). So, (a) \subset m, which forces m to coincide with the set of nonunits.

A SPECIFIC CASE. Let X = An. For any p in An, elements of \OhAn, p can be identified as quotients of polynomials:

f(x1, ...,xn)/ g(x1, ...,xn),
where g(p) is nonzero. The set of quotients of this type obviously is a subring of the fraction field of K[x1, ...,xn]. This fraction field is called the field of rational functions in n variables and denoted as follows:
K(x1, ...,xn) = frac(K[x1, ...,xn])

FOR A GENERAL AFFINE VARIETY X \subset An, we also can identify elements of \OhX,p as quotients of the form f/g where f and g are elements of the coordinate ring A(X) and g(p) is nonzero, provided that we do the following:

  1. identify elements of the coordinate ring as regular functions on X by means of the inclusion of A(X) in \Oh(X) discussed above;
  2. consider two quotients g1/ h1 and g2/ h2 as being equal if and only if there exists f in A(X) such that f(p) is nonzero and:
    f(g1 h2 - g2 h1) = 0.

While we can consider condition (2) as a technical algebraic condition that adapts the construction of field of fractions to a situation where zerodivisors may exist, it also has the intuitive content of saying that the difference of the two quotients g1/ h1 - g2/ h2 vanishes identically in some neighborhood of p, specifically D(f).

TO VERIFY THAT this set of quotients is a ring, we work formally with the set of ordered pairs and the equivalence relation between two pairs (g1, h1) and (g2, h2) given as in condition (2). This involves some calculations, but they are fairly routine except for transitivity. Thus, suppose that g1/ h1 = g2/ h2 and g2/ h2 = g3/ h3 as formal quotients. Then f( g1 h2 - g2 h1) = 0 and f* (g1 h2 - g2 h1) = 0 in A(X), where f and f* are elements such that f(p) and f*(p) are nonzero. After some calculation, we find that:

ff*h2 (g1 h3 - g3 h1) = 0.
¡¡Please check this for yourself!! Since h2 is a denominator, h2(p) is nonzero, so that g1/ h1 = g3/ h3 as formal quotients.

IT IS A PLEASANT SURPRISE that the identification of \OhX,p with this ring of quotients follows fairly immediately from the definitions (or at least without major obstacles), even though we will need the Nullstellensatz to verify the "global" identification of A(X) and \Oh(X). Explicitly, we map our ring of quotients to \OhX,p by sending the quotient g/h to the germ represented by (U,f), where U = D(h) and f is given by the quotient g/h at points of U. Surjectivity follows from the definitions (more or less). In verifying injectivity, we use the fact that distinguished open subsets form a basis of the topology.

Regular functions: the projective case

Let X \subset Pn be a projective variety. There are two ways to define the concept of a regular function on an open subset U \subset X. These two approaches are easily seen to be equivalent.

METHOD 1. Consider the standard affine covering {U0,...,Un } of Pn. If f is defined on U \subset X, we say that f is regular on U if f is regular on U\capUi for i = 0,..., n.

METHOD 2. Again, let f be defined on U \subset X. We say that f is regular at a point p in U if there is an open neighborhood V of p in U and homogeneous polynomials G (X0,..., Xn) and H (X0,..., Xn) with degree(G) = degree(H) such that f(q) = G(q)/H(q) for every point q in V.

Explicitly, if q = [a0,..., an], we are requiring that

\f(q) = {G(a_0,...,a_n  \over  H(a_),...,a_n)}

ONE CAN CHECK that the value of this expression does not depend on the choice of homogeneous coordinate vector used to represent q.

TO SEE WHY the two methods are equivalent, note that the numerator and denominator in a quotient of (usual) polynomials in x1,...,xn can be "homogenized", and that we can adjust degrees by multiplying by a suitable power of X0. For instance, suppose that f is represented in a neighborhood of some point of U0 \subset P2 by the following quotient:

<x^3 + 5xy + y^3 \over x^2 - 3y}.

Setting x = X/Z and y = Y/Z, we transform this into the following quotient of homogeneous polynomials:

homogenization of the above expression ...

SOME RINGS ASSOCIATED TO A PROJECTIVE VARIETY. We can define the local ring of germs of regular functions \OhX,p at a point p in X by methods very similar to what we did in the affine case. So, we won't discuss the details right now.

A concept related to the coordinate ring of an affine variety is the homogeneous coordinate ring of a projective variety. To get started, consider X \subset Pn and define I(X) to be the ideal in K[X0,..., Xn] generated by all homogeneous polynomials that vanish identically along X. We define the homogeneous coordinate ring to be

S(X) = K[X0,..., Xn]/I(X) .
Since homogeneous polynomials do not define regular functions on Pn, we do not attempt to describe elements of S(X)) as regu lar functions on X.

A SPECIFIC INSTANCE. Let X be the twisted cubic curve in P3. One can show that I(X) is generated by the following three homogeneous polynomials of degree 2 in the polynomial ring K[W, X,Y,Z]:

  • X2 - WY ,
  • Y2 - XZ ,
  • WZ - XY .
We won't try to verify this right now. Let's just note that it was shown in the text (and also in class) that these polynomials vanish at all points of the twisted cubic (and in fact that their common zeroset is exactly the twisted cubic). Showing that they generate I(X) requires somewhat more work.

ONE MORE DEFINITION: We define a quasiprojective variety to be a nonempty open subset of a projective variety. This includes affine varieties as a special case. The concept of a regular function is the same as in the projective case, since the definition of regular function is essentially local.

Irreducible varieties

We can work either with affine varieties or with projective varieties. Thus, if we consider an affine variety X \subset An, we say that X is irreducible if it is nonempty and is not the union of two properly smaller nonempty affine varieties. Thus:

if X = Y\cupZ, where Y and Z are nonempty affine varieties,
then either Y = X or Z = X.
The definition in the case of a projective variety X \subset Pn is completely parallel.

Some examples:

  1. A set consisting of one point is obviously irreducible.
  2. A1 is irreducible Indeed, the proper closed subsets of A1 are finite sets, while A1 is infi nite. (We are assuming that our field is algebraically closed, and thus infinite.)
  3. By extension, any curve with a rational parametrization is irreducible .
  4. An and Pn are irreducible.

PROOF OF 4). We will consider the case of An, since the case of Pn is similar.Suppose that An = X\cupY, where X and Y are affine varieties, both nonempty and distinct from An. Let f and g be nonzero elements of I(X) and I(Y) respectively, i.e. f and g are nonzero polynomials which vanish identically along X and Y respectively. It follows that

An = V(f)\cupV(g),
so that
An = V(fg).
Therefore, we can deduce the conclusion from the following:

LEMMA. Let K be an algebraically closed field, and let f(x1,..., xn) be a nonzero element of K[x1,..., xn]. Then V(f) is nonempty, and V(f) \neq An.

PROOF. We proceed by induction on n, the case n = 1 being immediate. So, in the inductive step, we can assume (by renumbering the variables) that xn actually occurs nontrivially in f. Therefore, we can write:
f(x1,..., xn) = g0( x1,..., xn-1) + g1( x1,..., xn-1) xn + ... + gd( x1,..., xn-1 ) xnd,
where gi is homogeneous of degree = i, while d > 0 and gd is nonzero. By the inductive hypothesis, there exists (a1,..., an-1 ) in An-1 such that gd( a1,..., an-1 ) \neq 0. Hence, \varphi(x) := f (a1,..., an-1,x) is a nonzero polynomial of degree = d > 0. It follows that there exist finitely many values of an in K such that f(a1,..., an) = \varphi (an) = 0, and infinitely many values of the last coordinate such that the corresponding value of f is nonzero. This proves our claim.

We note that irreducibility is really a property of closed sets in the Zariski topology, and does not really involve any other aspects of varieties. Here are a couple of basic properties of irreducible varieties.

PROPOSITION 1. If X is an affine or projective variety, then X is irreducible if and only if every nonempty open subset U \subset X is dense in X.

PROOF. Every open subset is dense if and only if every two nonempty open subsets have a nonempty intersection. By taking complements, this is equivalent to the statement that the union of two proper closed subsets is not equal to X.

PROPOSITION 2. Let X \subset An be an irreducible affine variety. If X \subseteq Y\cupZ, where Y and Z are affine varieties, then either X \subseteq Y or X \subseteq Z.

PROOF. If X \subseteq Y\cupZ, then X = (X\capY) \cup ( X\capZ). By irreducibility,

either X = X\capY or X = X\capZ.
This proves our claim.

Now we can state and prove our main result.

THEOREM. Let X be an affine variety (respectively a projective variety). Then X is the union of finitely many irreducible affine varieties (respectively finitely many irreducible projective varieties). Thus:

X = Y1 \cup ... \cup Ym,
where Y1 ,..., Ym are irreducible. If we require that Yi does not contain Yj when i \neq j, then Y1 ,..., Ym are uniquely determined. The corresponding statements are also true in the projective case.

DEFINITION: The irreducible closed subsets which occur in this uniquely determined decomposition are called the irreducible components of X.

PROOF OF THE THEOREM. In proving the existence of the decomposition for Zariski closed subsets of An, we use the fact that the Zariski topology of An is Noetherian. By definition, this means that every descending closed subsets stabilizes after fin itely many steps. An equivalent characterization is that every nonempty collection of closed subsets has a minimal member, i.e., there is a member of the collection which is not contained in any other set in the collection.

So, assume that there exists a Zariski-closed subset of An which is not the union of finitely many irreducible closed subsets, and let S be the collection of closed subsets which are not finite unions of irreducibles. Then S contains a minimal member, say X0 . If X0 were irreducible, this would immediately contradict the fact that no member of S is a union of finitely many irreducible closed subsets. So, we have X0 = Y\cupZ, where Y and Z are proper closed subsets of X0 . Since X0 is a minimal member of S, neither Y nor Z is a member of S; therefore, both Y and Z are unions of finitely many irreducible closed subsets. This implies that S is a union of finitely many irreducible closed subsets, thus contradicting the hypothesis that S is nonempty and thereby proving the first part of the theorem.

To prove uniqueness, suppose that some X \subset An has two different decompositions as a union of irreducible closed subsets, both satisfying the condition that there be no inclusion relation among different subsets in the collection. Specifically, suppose that:

X = Y1 \cup ... \cup Ym
and also:
X = Z1 \cup ... \cup Zp ,
where there is no inclusion relation among the various Y's and also no such relation among the Z's

.

Consider a particular subset Yi from the first decomposition. Since Yi \subseteq Z1 \cup ... \cup Zp, we can use Proposition 2 to show that Yi \subseteq Zj for some j. Similarly, we show that Zj \subseteq Yk for some k. Thus, Yi \subseteq Zj \subseteq Yk . Because there are no non trivial inclusions among the Y's, it follows that i = k, and Yi = Zj . Hence, all of the Yi occur among Z1, ... ,Zp . Now, we can show by similar methods that all of the Zj occur among Z1, ... , Zp . Therefore, the same irreducible closed subsets occur in the two decompositions.

Another example.

We would now like to claim that if f is an irreducible polynomial in n variables (respectively, if F is an irreducible homogeneous polynomial in n + 1 variables), then the zeroset V(f) is an irreducible closed subset of An (respectively that the zeroset V(F) is an irreducible closed subset of Pn ). A little work is required for this, so that it will be saved for Exercises 7 and 8, which are linked below. It will follow from this claim that if f is an arbitrary polynomial in n variables, then the irreducible components of the hypersurface V(f) correspond bijectively to the distinct irreducible factors of f [and similarly in the projective case.]

Related Exercises

Last updated November 18, 1997.

Please send comments and/or corrections to:

Joel Roberts
351 Vincent Hall
625-1076
e-mail: roberts@math.umn.edu
http://www.math.umn.edu/~roberts