Algebraic Set
Algebraic Set
Algebraic Sets
From the algebraic point of view, the most natural functions to consider
on An are those defined by evaluating a polynomial in k[x1 , . . . , xn ] at a point.
Analogously, the simplest geometric figures in An are the zero sets of a single
polynomial.
V(f ) = {p ∈ An | f (p) = 0}
1.1.3 Examples.
(i) In R1 , V(x2 + 1) = ∅, but in C1 , V(x2 + 1) = {±i}. Generally, if n = 1
then V(F ) is the set of roots of F in k. If k is algebraically closed and F
is non-constant then V(F ) is non-empty.
(ii) In Z1p , by Fermat’s Little Theorem, V(xp − x) = Z1p .
(iii) By Fermat’s Last Theorem, if n > 2 then V(xn + y n − 1) is finite in Q2 .
1
(iv) In R2 , V(x2 +y 2 −1) = the unit circle in R2 , and in Q2 it gives the rational
points on the unit circle. Notice the circle admits a parameterization by
rational functions as follows:
1 − t2
2t
(x, y) = , , t ∈ R.
1 + t2 1 + t2
1.2.2 Examples.
(i) For any a, b ∈ k, {(a, b)} is an algebraic set in k2 since {(a, b)} = V(x −
a, y − b).
(ii) In R2 , V(y−x2 , x−y 2 ) is only 2 points, but in C2 it is 4 points. Generally,
Bézout’s Theorem tells us that the number of intersection points of a
curve of degree m with a curve of degree n is mn in projective space over
an algebraically closed field.
(iii) The twisted cubic is the rational curve {(t, t2 , t3 ) | t ∈ R} ⊆ R3 . It is an
algebraic curve; indeed, it is easy to verify that it is V(y − x2 , z − x3 ).
(iv) Not all curves in R2 are algebraic. For example, let
X = {(x, y) | y − sin x = 0}
and suppose that X is algebraic, so that X = V(S) for some S ⊆ R[x, y].
Then there is F ∈ S such that F 6= 0 and so
\
X = V(S) = V(f ) ⊆ V(F ).
f ∈S
2
Remark. The method used in the last example works in more generality. Sup-
pose that C is an algebraic affine plane curve and L is a line not contained C.
Then L ∩ C is either ∅ or a finite set of points.
Proof: Clearly these sets are all algebraic. Conversely, the zero set of any
non-zero polynomial is finite, so if S contains a non-zero polynomial F then
V(S) ⊆ V(F ) is finite. If S = ∅ or S = {0} then V(S) = A1 .
Proof:
(i) Since S ⊆ T ,
\ \
V(T ) = V(f ) ⊆ V(f ) = V(S).
f ∈T f ∈S
(ii) From (i), V(hSi) ⊆ V(S). If x ∈ V(S) and f ∈ I then we can write f as
f = gq f1 + · · · + gm fm ,
since x ∈ V(S).
Since every algebraic set is the zero set of an ideal of polynomials, we now
turn our attention to ideals in polynomial rings. If a ring R is such that all of
its ideals are finitely generated it is said to be Noetherian 2 . For example, all
fields are Noetherian. The Hilbert Basis Theorem states that all polynomial
rings with coefficients in a Noetherian ring are Noetherian.
1 The ideal generated by S is the intersection of all ideals containing S. More concretely,
( n )
X
hSi = ak sk : a1 , . . . , an ∈ R and s1 , . . . , sn ∈ S .
k=1
2 Some readers may be more familiar with a different definition of Noetherian in terms of
ascending chains of ideals. This definition is equivalent to ours by Proposition A.0.9.
3
1.2.5 Theorem (Hilbert Basis Theorem). If R is Noetherian, then R[x1 , . . . , xn ]
is Noetherian.
1.2.6 Corollary. Every algebraic set X in An is the zero set of a finite set of
polynomials.
Tn
Remark. Since V(f1 , . . . , fn ) = k=1 V(fk ), the preceding corollary shows that
every algebraic set is the intersection of finitely many hypersurfaces.
1.2.7 Proposition.
S T
(i) If {Iα } is a collection of ideals then V( α Iα ) = α V(Iα ), so the inter-
section of any collection of algebraic sets is an algebraic set.
(ii) If I and J are ideals then V(IJ) = V (I) ∪ V(J), so the finite union of
algebraic sets is an algebraic set.3
(iii) V(0) = An , V(1) = ∅, and V(x1 − a1 , . . . , xn − an ) = {(a1 , . . . , an )}, so
any finite set of points is algebraic.
Proof:
(i) We have
!
[ \ \ \ \
V Iα = V(f ) = V(f ) = V(Iα ).
α f ∈∪α Iα α f ∈Iα α
n
( )
X
IJ = ak bk : a1 , . . . , an ∈ I and b1 , . . . , bn ∈ J .
k=1
4
(ii) Since (gh)(x) = 0 if and only if g(x) = 0 or h(x) = 0,
\
V(IJ) = V(f )
f ∈IJ
\
= V(gh)
g∈I,h∈J
\
= V(g) ∪ V(h)
g∈I,h∈J
\ \
= V(g) ∪ V(h)
g∈I h∈J
= V(I) ∪ V(J).
(iii) This is clear.
Remark. Note that finiteness of the union in property (ii) is required; for ex-
ample, consider Z in R. It is not an algebraic set, because a polynomial over a
field can only have finitely many roots, but it is the union of (infinitely many)
algebraic sets, namely V(x − n) for n ∈ Z.
1.2.9 Example. The non-empty open sets in the Zariski topology on the affine
line A1 are precisely the complements of finite sets of points. However, this is
not true for An when k is infinite and n > 1. For example, V(x2 + y 2 − 1), the
unit circle in R2 , is closed but is not finite. Moreover, note that the Zariski
topology on An is Hausdorff5 if and only if k is finite, in which case it is identical
to the discrete topology.
4 A topology on a set X is a collection τ of subsets of X that satisfies the following
properties:
(i) ∅, X ∈ τ , S
(ii) if Gi ∈ τ for every i ∈ I then i∈I Gi ∈ τ ,
(iii) if G1 , G2 ∈ τ then G1 ∩ G2 ∈ τ .
The sets in τ are said to be open, and their complements are said to be closed.
5 Recall that a topology is said to be Hausdorff if distinct points always have disjoint
open neighbourhoods.
5
We have associated an algebraic subset of An to any ideal in k[x1 , . . . , xn ]
by taking the common zeros of its members. We would now like to do the
converse and associate an ideal in k[x1 , . . . , xn ] to any subset of An .
1.2.11 Examples.
(i) The following ideals of k[x] correspond to the algebraic sets of A1 : I(∅) =
h1i, I({a1 , . . . , an }) = h(x − a1 ) · · · (x − an )i, and
(
1 0 if k is infinite,
I(A ) = pn
hx − xi if k has pn elements.
hx − a, y − bi ⊆ I({(a, b)}),
so we need only prove the reverse inequality. Assume that f ∈ I({(a, b)}).
By the division algorithm, there is g(x, y) ∈ k[x, y] and r(y) ∈ k[y] such
that
f (x, y) = (x − a)g(x, y) + r(y).
But 0 = f (a, b) = r(b), so y − b divides r(y) and we can write we can
write r(y) = (y − b)h(y), and hence
1.2.12 Proposition.
(i) If X ⊆ Y ⊆ An then I(Y ) ⊆ I(X).
(ii) I(∅) = k[x1 , . . . , xn ].
I({(a1 , . . . , an )}) = hx1 −a1 , . . . , xn −an i for any point (a1 , . . . , an ) ∈ An .
I(An ) = 0 if k is infinite.
(iii) S ⊆ I(V(S)) for any set of polynomials S ⊆ k[x1 , . . . , xn ].
X ⊆ V(I(X)) for any set of points X ⊆ An .
(iv) V(I(V(S))) = V(S) for any set of polynomials S ⊆ k[x1 , . . . , xn ].
I(V(I(X))) = I(X) for any set of points X ⊆ An .
Proof:
(i) If f is zero on every point of Y then it is certainly zero on every point of
X.
6
(ii) That I(∅) = k[x1 , . . . , xn ] and I(An ) = 0 if k is infinite are clear. Fix
(a1 , . . . , an ) ∈ An , and define ϕ : k[x1 , . . . , xn ] → k by ϕ(f ) = f (a1 , . . . , an ).
Clearly, ϕ is a surjective homomorphism, and
ker(ϕ) = hx1 − a1 , . . . , xn − an i.
We have
k[x1 , . . . , xn ]/hx1 − a1 , . . . , xn − an i ∼
= k,
so hx1 − a1 , . . . , xn − an i is a maximal ideal. The ideal I({(a1 , . . . , an )})
is proper and contains hx1 − a1 , . . . , xn − an i, a maximal ideal, so it must
be equal to that maximal ideal.
(iii) These follow from the definitions of I and V.
(iv) From (iii), V(S) ⊆ V(I(V(S))), and by Proposition 1.2.4 (i), V(I(V(S))) ⊆
V(S) since S ⊆ V(I(S)). Therefore V(S) = V(I(V(S))). The proof of
the second part is similar.
Remarks.
(i) As is shown in the proof of part (ii) of the last proposition, the ideal
hx1 − a1 , . . . , xn − an i of any point (a1 , ..., an ) ∈ An is maximal.
(ii) Equality does not always hold in part (iii) of the last proposition, as
shown by the following examples:
(a) Consider I = hx2 + 1i ⊆ R[x]. Then 1 ∈ / I, so I 6= R[x]. But
V(I) = ∅, so I(V(I)) = R[x] % I.
(b) Consider X = [0, 1] ⊆ R. Then I(X) = 0 and V(I(X)) = R % X.
These examples also show that not every ideal of k[x1 , . . . , xn ] is the ideal
of a set of points and that not every subset of An is algebraic.
7
1.2.14 Proposition.
(i) If X ⊆ An , then X = V(I(X)).
(ii) If I ⊆ k[x1 , . . . , xn ] is an ideal, then I = I(V(I)).
Proof: We will only prove (i), as the proof of (ii) is very similar. By part
(iii) of Proposition 1.2.12, we have X ⊆ V(I(X)) Since V(I(X)) is an algebraic
set, X ⊆ V(I(X)). Conversely, since X ⊆ X, V(I(X)) ⊆ V(I(X)). By part
(ii) of Proposition 1.2.7, we have V(I(X)) = X, because X is an algebraic set.
Therefore, V(I(X)) ⊆ X, showing that X = V(I(X)).
1.2.15 Examples.
(i) If X = (0, 1) ⊆ R, then the closure of X in the metric topology is [0, 1],
whereas the closure of X in the Zariski topology is R.
(ii) If k is infinite and X ⊆ A1 is any infinite set of points then X = A1 .
In particular, the Zariski closure of any non-empty open set is the whole
line, or every non-empty open set is Zariski dense in the affine line.
(iii) Let I = hx2 i. Then I = I(V(I)) = hxi, so that I 6= I and I is not an
ideal of a set of points.
(ii) Let I = hx2 + y 2 − 1, x − 1i ⊆ R[x, y]. The set V(x2 + y 2 − 1) is the unit
circle, and V(x − 1) is the vertical line that is tangent to the circle at
(1, 0). Because it only intersects the circle at one point, the intersection is
with “multiplicity two”. Therefore, our intuition would lead us to think
that I is not a closed ideal. This is indeed the case, as
8
The zero sets of the generators of I are a vertical line through (1, 0) and a
horizontal line through the origin, which intersect once at the point (1, 0)
with “multiplicity one”, again confirming our intuition.
1.3.2 Examples.
(i) If X ⊆ An then I(X) is radical, because f (x) = 0 whenever f n (x) = 0.
(ii) Every prime ideal is radical. For a proof, see Proposition A.0.15. How-
ever, not every proper radical ideal is prime. For example, the ideal
but y ∈
/ I, simply because of the degrees of the y terms in the generators.
Hence I is not radical. We already examined this example geometrically
above.
(iv) Let I = hy − x2 , y − x3 i. If u = x(x − 1), then
u2 = [(y − x2 ) − (y − x3 )](x − 1) ∈ I,
We saw in the first of the above examples that if I is the ideal of a set of
points then I is radical. Is the converse true? That is, if I is radical is it true
that I = I?
√
1.3.3 Proposition. If I is an ideal of k[x1 , . . . , xn ], then I ⊆ I ⊆ I. In
particular, a closed ideal is radical.
9
√ √
Proof: Clearly, I ⊆ I. Suppose f ∈ I. Then f n ∈ I for some n ≥ 1. Since
f n (x) = 0 if and only if f (x) = 0, √
we have f ∈ I(V(I)). By Proposition 1.2.14,
I = I(V(I)), so f ∈ I. Therefore, I ⊆ I.
√
It follows
√ from the previous proposition that if I = I then I = I if and
only if
√ I = I(V(I)). However, if k is not algebraically closed, it often happens
that I 6= I(V(I)):
1.3.4 Example. The polynomial x2 + 1 ∈ R[x] is irreducible, so the ideal
hx2 + 1i is maximal. Hence it is radical, and it is obviously proper. However,
10
Proof: If X is irreducible then suppose that f, g ∈ k[x1 , . . . , xn ] are such that
f g ∈ I(X). Then hf gi ⊆ I(X), so X = V(I(X)) ⊆ V(f g) = V(f )∪V(g). Hence
X = (X ∩ V(f )) ∪ (X ∩ V(g)), so without loss of generality, X = X ∩ V(f ) ⊆
V(f ). Therefore f ∈ I(X) and I(X) is prime.
Suppose that I(X) is prime but is reducible, with X = X1 ∪ X2 . Then
I(X) = I(X1 ) ∩ I(X2 ). If I(X) = I(X1 ) then X = X1 , which is not allowed.
Hence there is f ∈ I(X1 ) \ I(X). But for any g ∈ I(X2 ), f g ∈ I(X1 ) ∩ I(X2 ) =
I(X), so since f ∈ / I(X) and I(X) is prime, g ∈ I(X). This implies that
I(X) = I(X2 ) (and hence X = X2 ), a contradiction.
1.4.3 Examples.
(i) An is irreducible for all n ≥ 1, because I(An ) = {0}, which is a prime
ideal.
(ii) If (a1 , . . . , an ) ∈ An , then {x} is irreducible, because
The correspondence between algebraic sets and ideals of sets of points takes
irreducible algebraic sets to prime ideals, and prime ideals that are ideals of sets
of points to irreducible algebraic sets. If k is algebraically closed, by combining
the results of this chapter we have the following correspondence:
Geometry Algebra
affine space An polynomial ring k[x1 , . . . , xn ]
algebraic set radical ideal
irreducible algebraic set prime ideal
point maximal ideal
Remark. If k is not algebraically closed then there are more prime ideals than
irreducible algebraic sets.
11
(i) distinct prime ideals may give the same algebraic set, e.g. V(hx2 + y 2 i) =
{(0, 0)} = V(hx, yi) in R2 ;
(ii) a prime ideal may have a reducible zero set, e.g. V(hx2 + y 2 (y − 1)2 i) =
{(0, 0), (0, 1)} in R2 .
X = X1 ∪ · · · ∪ Xm ,
12
Proof: By Proposition 1.4.4, X is the finite union of irreducible algebraic sets.
By possibly removing some constituents of this union, we have an irredundant
decomposition X = X1 ∪ · · · ∪ Xm . Suppose that X also has an irredundant
decompostion X = Y1 ∪ · · · ∪ Yn . Then for any i, Xi is contained in some
Yj0 by Lemma 1.4.5. Similarly, Yj0 ⊆ Xi0 for some i0 , but this implies that
Xi ⊆ Yj0 ⊆ Xi0 , and since the decomposition is irredundant, Xi = Xi0 = Yj0 .
Therefore every Xi corresponds to a Yj , and vice-versa.
1.4.7 Examples.
(i) Suppose that f ∈ k[x1 , . . . , xn ] and f = f1r1 . . . fm
rm
then
V(f ) = V(f1 ) ∪ · · · ∪ V(fm ).
If k is algebraically closed then this is a decomposition and I(V(f )) =
hf1 · · · fm i.
(ii) Consider X = V(y 4 − x2 , y 4 − x2 y 2 + xy 2 − x3 ) ⊆ C2 . Notice that
y 4 − x2 = (y 2 − x)(y 2 + x),
and
y 4 − x2 y 2 + xy 2 − x3 = (y 2 + x)(y − x)(y + x),
so V(y 2 + x) is an irreducible component of X. The other 3 points in X
are (0, 0), (1, 1) and (1, −1). But (0, 0) ∈ V(y 2 + x), so the decomposition
of X is V(y 2 + x) ∪ {(1, 1)} ∪ {(1, −1)}.
(iii) Consider X = V(x2 + y 2 (y − 1)2 ) ⊆ R2 . X = {(0, 0), (0, 1)}, so X is
reducible. But f (x, y) = x2 + y 2 (y − 1)2 is irreducible in R[x, y]. Indeed,
f (x, y) = (x + iy(y − 1))(x − iy(y − 1)).
Since R[x, y] ⊆ C[x, y] are UFDs, if f factors in R[x, y] the decompostion
must agree with the decomposition we have, up to constant multiple, but
this is impossible.
13
1.5.1 Proposition. If f, g ∈ k[x, y] have no common factors then V(f, g) =
V(f ) ∩ V(g) is at most a finite set of points.
Proof: Since f and g have no common factor in k[x, y] = k[x][y], they have
no common factors in k(x)[y]. Therefore gcd(f, g) exists and is 1 in k(x)[y], so
there are s, t ∈ k(x)[y] such that sf + tg = 1. Hence there is d ∈ k[x] such that
ds = a, dt = b, where a, b ∈ k[x][y] = k[x, y]. Then af + bg = d ∈ k[x]. Now if
(x0 , y0 ) ∈ V(f, g) then d(x0 ) = 0, so there are at most finitely many possible
values for x0 . Similarily, there are at most finitely many possible values for y0 ,
so V(f, g) is finite.
Proof: We have already seen that all algebraic subsets of A2 of these forms
are irreducible. Let X ⊆ A2 be an irreducible algebraic set. Assume that X
is not A2 or a single point. Then I(X) 6= 0, so there is at least on non-zero
polynomial f ∈ I(X). Moreover, any irreducible factor of f is in the prime
ideal I(X), since X is assumed to be irreducible. We may therefore assume
that f is irreducible. Then Corollary 1.5.2 implies that X = V (f ) since X is
infinite.
1.5.4 Examples.
(i) In R2 , V(y − x2 ) is irreducible because f = y − x2 is an irreducible
polynomial and V(y − x2 ) is infinite.
(ii) In R2 , V(y 2 − x2 (x − 1)) is also irreducible for the same reasons. Hence
it is connected in the Zariski topology. However, it is not connected in
the metric topology.
14
Appendix A
A.0.5 Definition. A principal ring is a ring for which every ideal is generated
by a single element. A principal integral domain is called a principal ideal
domain, or PID for short.
Proof: Since k[x] is clearly an integral domain, we only need to show that
it is principal. Let I be an ideal of k[x], and let f be a monic polynomial of
minimum degree in I. First, we show that f is unique, i.e. if g is another monic
polynomial in I such that deg(g) = deg(f ), then f = g. Let h = f − g. Then
h ∈ I, and since deg(h) < deg(f ) we must have h = 0, so g = f .
We now show that I = hf i. Since f ∈ I, we have hf i ⊆ I. To establish the
reverse inclusion, fix g ∈ I. By the division algorithm, there exist q, r ∈ k[x]
such that r is monic, g = qf + r, and either r = 0 or deg(r) < deg(f ). Since I
is an ideal, r = g − qf ∈ I. By the minimality of the degree of f , we can not
have deg(r) < deg(f ), so r = 0. Therefore, g = qf and g ∈ hf i. Since g ∈ I
was arbitrary, this shows that I ⊆ hf i, and thus I = hf i.
15
(i) R is Noetherian,
(ii) R satisfies the ascending chain condition on ideals, i.e. if
I0 ⊆ I1 ⊆ · · · ⊆ In ⊆ · · ·
Ik = Ik+1 = · · · = Ik+n = · · · .
I0 ⊆ I1 ⊆ · · · ⊆ In ⊆ · · ·
In general, the union of ideals is not an ideal, but the union of an increasing
chain of ideals can easily be seen to be an ideal. Thus I is an ideal. Since R
is Noetherian, I is finitely generated, i.e. there exist a1 , . . . , am ∈ I such that
I = ha1 , . . . , am i. Let k ∈ N be such that a1 , . . . , am ∈ Ik . Then
I = Ik = Ik+1 = · · · = Ik+n = · · · .
I0 ( I1 ( · · · ( In ( · · · ,
Proof: Suppose R[x] is not Noetherian, and let I is an ideal of R[x] that is not
finitely generated. Let f0 be a polynomial of minimum degree in I. Continuing
by induction, let fk+1 be a polynomial of minimum degree in I \ hf0 , . . . , fk i.
For every k ∈ N, let dk = deg(fk ), and let ak be the leading coefficient of fk ,
and let J = h{ak : k ∈ N}i. Since R is Noetherian and
16
is an increasing chain of ideals whose union is J, there exists an n ∈ N such
that J = ha0 , . . . , an i.
Let I0 = hf0 , . . . , fn i. By construction, fn+1 ∈
/ I0 . Since J = ha0 , . . . , an i
and an+1 ∈ J, there exist b0 , . . . , bn ∈ R such that an+1 = b0 a0 + · · · + bn an .
Then, as fn+1 ∈ I \ I0 , we have
g = mn+1 − xdn+1 −d0 b0 f0 − · · · − xdn+1 −dn bn fn ∈ I,
so deg(g) < deg(fn+1 ). However, g ∈/ I0 , as fn+1 ∈
/ I0 , contradicting the mini-
mality of deg(fn+1 ). Therefore, our assumption that R[x] is not Noetherian is
false.
17
A.0.15 Proposition. Let R be a ring, and I a prime ideal of R. Then I is
radical.
18
Appendix B
Transcendence Bases
B.0.17 Examples.
(i) The empty set is algebraically independent. If K = F , it is also a tran-
scendence basis of K over F .
(ii) Let F a fixed field and let K = F (x1 , . . . , xn ) be the fraction field of the
ring F [x1 , . . . , xn ]. We claim that {x1 , . . . , xn } is a transcendence basis
of K over F . It is clearly algebraically independent, as if f ∈ F [t1 , . . . , tn ]
is such that f (x1 , . . . , xn ) = 0, we have that f = f (t1 , . . . , tn ) = 0. To
show that {x1 , . . . , xn } is a maximal algebraically independent set, we
will show that {x1 , . . . , xn , p/q} is algebraically dependent over F for any
p, q ∈ F [x1 , . . . , xn ], q 6= 0. Define f ∈ F [t1 , . . . , tn+1 ] by
19
Proof:
(i) Let P be the partial order of algebraically independent subsets S of K that
contain U , ordered by inclusion. If C is a chain in P , then C is clearly
algebraically independent,Sas any possible algebraic dependence involves
finitely many elements of C, which could all be chosen to be in the same
member of C. Therefore, by Zorn’s Lemma, P has a maximal element.
However, by definition, such a maximal element is a transcendence basis
of K containing U .
(ii) For the sake of sanity, we will assume that B1 is finite. In the infinite
case, it is argued using either multiple applications of Zorn’s Lemma or
transfinite induction. Suppose B1 = {x1 , . . . , xm }, where m ≥ 1 is the
minimal cardinality of any transcendence basis. It suffices to show that
if w1 , . . . , wn are algebraically independent elements of K then n ≤ m,
as we could then swap B1 and B2 to get the opposite inequality. If
every wi is an xj , there is nothing to prove, so by possibly reordering the
wi ’s, we can assume that w1 6= xi for i = 1, . . . , m. Since {x1 , . . . , xm }
is a transcendence basis, {w1 , x1 , . . . , xm } is algebraically dependent, so
there is a non-zero polynomial f1 ∈ F [t1 , . . . , tm+1 ], which can clearly be
chosen to be irreducible, such that f1 (w1 , x1 , . . . , xm ) = 0. After possibly
renumbering the xj ’s we may write
k
X
f1 = gj (w1 , x2 , . . . , xm )xj1
j=1
20
Appendix C
One fundamental fact about integral extensions is that they are transitive,
i.e. the composition of integral extensions is also an integral extension. In the
proof we use a special case of a notion from the theory of modules. If R and
S are rings and R ⊆ S, we say that S is a finitely generated R-module if there
exists a finite set A ⊆ S such that S = RA.
Proof: Suppose that t is integral over R. There then exist r0 , ..., rn−1 ∈ R
such that
tn + rn−1 tn−1 + · · · + r1 t + r0 = 0,
or
tn = −(rn−1 tn−1 + · · · + r1 t + r0 ),
21
so tn and all higher powers of t can be expressed as R-linear combinations of
tn−1 , . . . , t, 1. Then R[t] = R{tn−1 , . . . , t, 1} is a finitely generated R-module.
Suppose R[t] is a finitely generated R-module. Since t ∈ R[t] and R[t] ⊆ T ,
this means that S = R[t] satisfies the conditions of (iii).
Suppose there exists a subring S of T such that t ∈ S and S is a finitely
generated R-module. Let A ⊆ S be a finite set such that S = RA. Enumerate
the elements of A by A = {a1 , . . . , an }. For i = 1, . . . , n, the element tai is an
element of S and can thus be written as R-linear combinations of a1 , . . . , an ,
i.e. for some coefficients cij ∈ R,
n
X
tai = cij aj .
j=1
where δij is the Kronecker delta. Let B be the n × n matrix whose i, j entry
is δij t − cij , and let v be the n × 1 column vector whose entries are a1 , . . . , an .
These equations then simply amount to saying that Bv = 0; it follows from
Cramer’s Rule that (det B)ai = 0 for i = 1, . . . , n. Since 1 ∈ S is an R-linear
combination of a1 , . . . , an , it follows that det B = 0. But B = tI − C, where C
is the matrix whose i, j entry is cij . Thus, t is a root of the monic polynomial
det(xI − C) ∈ R[x], i.e. t is integral over R.
Proof: Fix t ∈ T . Since T is integral over S, there exist s0 , ..., sn−1 ∈ S with
tn + sn−1 tn−1 + · · · + s1 t + s0 = 0.
22
Proof: Suppose R is a field and fix a non-zero s ∈ S. Then, s is integral over
R, so there exist a0 , a1 , . . . , an−1 ∈ R such that
sn + an−1 sn−1 + · · · + a1 s + a0 = 0.
Then
r−1 = −(am−1 + · · · + a1 rm−2 + a0 rm−1 ) ∈ R.
Therefore, R is a field.
Most rings we deal with in these notes are also vector spaces over some
field k, where the vector space addition is the same as addition in the ring,
and the scalar multiplication coincides with multiplication in the ring, after
identifying the scalar α ∈ k with the element α · 1 of the ring. Such rings are
called k-algebras. We deal with k-algebras elsewhere in these notes, but repeat
some basic facts about them here.
C.0.24 Examples.
(i) k[x1 , . . . , xn ] is a k-algebra.
(ii) Let A be a k-algebra and I an ideal of A. Then I is also a vector subspace
of A, and the ring quotient of A by I agrees with the vector space quotient
of A by I. Hence A/I is also a k-algebra.
ϕ(an xn + · · · + a1 x + a0 ) = an xn + · · · + a1 x + a0
23
Let A be a k-algebra. We say that A is finitely generated if A = k[a1 , . . . , an ]
for some a1 , . . . , an ∈ A. Moreover, y1 , . . . , yn ∈ A are said to be algebraically
independent if there is no non-zero f ∈ k[x1 , . . . , xn ] such that f (y1 , . . . , yn ) =
0, or equivalently, if the k-algebra homomorphism
ϕ : k[x1 , . . . , xn ] → k[y1 , . . . , yn ]
y2 = r2 − r1m2 , . . . , yn = rn − r1mn .
24
In the particular case of fields, the Noether Normalization Lemma can be
restated more naturally in the language of field extensions.
I ⊆ hx1 − a1 , . . . , xn − an i.
Thus
{(a1 , . . . , an )} = V(hx1 − a1 , . . . , xn − an i) ⊆ V(I),
so (a1 , . . . , an ) ∈ V(I), and V(I) 6= ∅.
25
is not proper, i.e. 1 ∈ I 0 . Therefore, there exist h1 , . . . , hn+1 ∈ k[x1 , . . . , xn+1 ]
such that
1 = h1 f1 + · · · + hn fn + hn+1 (1 − xn+1 g).
Working in the field of fractions of k[x1 , . . . , xn+1 ], substitute g −1 for xn+1 and
multiply both sides by an appropriate power g k of g to clear denominators on
the right-hand side and give
26