lecture-notesAG
lecture-notesAG
Contents
1. Notation 3
2. What is algebraic geometry ? 3
3. Affine algebraic varieties 6
3.1. From algebra to geometry 6
3.2. From geometry to algebra 8
3.3. Hilbert’s Nullstellensatz: statement and consequences 9
3.4. Proof of the Nullstellensatz 10
3.5. Irreducibility 12
3.6. Morphisms (part I) 14
3.7. Exercises 16
4. Projective algebraic varieties 18
4.1. Projective spaces and subspaces 18
4.2. Covering by affine spaces and Zariski topology 18
4.3. Projective algebraic varieties 19
4.4. Ideals of projective varieties 20
4.5. Cones and projective Nullstellensatz 21
4.6. Projective coordinate ring 21
4.7. A word about morphisms of projective varieties 22
4.8. Exercises 23
5. Sheaves, ringed spaces, and affine algebraic varieties 24
5.1. Sheaves 24
5.2. Ringed spaces 25
5.3. Affine algebraic varieties as ringed spaces 26
5.4. Affine algebraic varieties, revisited 30
5.5. The equivalence between morphisms and coordinate rings 31
6. Algebraic varieties 33
6.1. Definition and basic properties 33
6.2. Local rings 36
7. Projective algebraic varieties, revisited 39
7.1. Graded rings and modules 39
7.2. The sheaf of regular functions on a projective variety 39
7.3. From affine to projective and back: homogenization and dehomogenization 40
1
2 LEONARDO CONSTANTIN MIHALCEA
8. Morphisms (II) 44
8.1. Morphisms to affine varieties 44
8.2. Morphisms to projective varieties 44
8.3. Images of morphisms 47
9. Products 49
9.1. Categorical product 49
9.2. Products of varieties 50
9.3. Separated varieties and the Haussdorff axiom 52
10. Dimension 55
10.1. Topology 55
10.2. Krull dimension and dimension of algebraic varieties 55
10.3. Expected dimension of systems of equations 57
10.4. Projective versions 59
11. The fibres of a morphism 60
12. Sheaves of modules 62
12.1. Definitions 62
12.2. Quasi-coherent sheaves on affine varieties 63
12.3. Quasi-coherent sheaves on projective varieties 66
13. Hilbert polynomials and Bézout’s theorem 69
13.1. The Hilbert function and the Hilbert polynomial 69
13.2. Examples of Hilbert polynomials; arithmetic genus 70
13.3. Bézout’s theorem 72
13.4. Degree of the union of two projective varieties 73
14. Start of semester 2: goals 75
15. Schemes 76
15.1. Spectrum of a ring 76
15.2. The structure sheaf of Spec(R) 79
15.3. Preschemes and morphisms of preschemes 82
16. Products of preschemes 87
16.1. Examples of fibre products 87
16.2. Preschemes and algebraic varieties 90
17. Proj(R) and projective schemes 91
17.1. Proj(R) and its structure sheaf 91
17.2. Projective subschemes 92
18. More properties of schemes 94
18.1. Integral schemes; open and closed embeddings 94
18.2. Separated morphisms; schemes 95
18.3. Proper morphisms 97
19. Relative differentials 99
19.1. (Quasi-)coherent sheaves 99
19.2. The module of relative differentials 100
19.3. The sheaf of relative differentials 102
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 3
1. Notation
Throughout the notes: N = {0, 1, 2, . . .} (the set of natural numbers); Z = N ∪ −N
(the set of integers); Q = {a/b : a, b ∈ Z, b ̸= 0} (the set of rational numbers); R is
the set of real numbers, and C := R + iR is the set of complex numbers; here i2 = −1.
All fields are commutative. For k a field, k[x1 , . . . , xn ] denotes the polynomial ring with
coefficients in k.
C = {(x, y, z) = (t3 , t4 , t5 ) : t ∈ C} ⊂ C3 .
Note that there are three equations, so we would expect only two equations.
However, if we leave out the last equation, the z-axis satisfies the first two;
similarly we cannot remove any of the other two equations.
Further, one may actually show that the common locus of
3
x = yz
y 2 = xz
z 2 = x2 y + ε
a∈/ V ∪ W , then a ∈/ V (f g : f ∈ S1 , g ∈ S2 ). If a ∈
/ V ∪ W then there exist f ∈ S1
and g ∈ S2 such that f (a) ̸= 0 and g(a) ̸= 0. Then f (a)g(a) = (f g)(a) ̸= 0, and we are
done. □
Example 3.3. (An example of a space which is not an algebraic variety.) Let
C := {(x, sin(x)) : x ∈ R} ⊂ R2 .
If this were an affine variety, then C is the common zero locus of some polynomials
Pi (x, y) ∈ k[x, y]. But then C ∩ {(x, 0) : x ∈ R} is also an algebraic variety, with
equations Pi (x, y) = 0 and y = 0 (by Proposition 3.2). Observe that this intersection
consists of the infinite discrete set {(kπ, 0); k ∈ Z}. On the other side, the zero locus
of the system Pi (x, 0) = 0 is either ∅, the whole x-axis, or finitely many points on the
x-axis. This is a contradiction.
(A shorter way to say this is that C ∩ R is an affine variety in R, identified to the
x-axis. But we know the affine varieties in R from Lemma 3.1, and C ∩ R is not of that
form.)
Consider the function
V : { subsets of k[x1 , . . . , xn ]} → { subsets of An }; S 7→ V (S).
We study some properties of V .
Proposition 3.4. (1) The function V is decreasing, i.e. for S1 ⊂ S2 , V (S1 ) ⊃ V (S2 ).
(2) Let ⟨S⟩ be the ideal generated by S. Then V (S) = V (⟨S⟩).
(3) Let S be a nonempty subset in k[x1 , . . . , xn ]. There exist finitely many elements
P1 , . . . , Ps ∈ S such that V (S) = V (P1 , . . T
. , Ps ). S
(4) For sets Si ∈ k[x1 , . . . , xn ] (i ∈ I), V (Si ) = V ( Si ).
(5) For finitely many sets S1 , . . . , Sp ⊂ k[x1 , . . . , xn ],
V (S1 ) ∪ V (S2 ) ∪ . . . ∪ V (Sp )
is an affine variety.
Proof. Part (1) follows by definition: if we impose more equations, the resulting variety
will be smaller.
For part (2), S ⊂ ⟨S⟩, therefore V (⟨S⟩) ⊂ P V (S). Conversely, take x ∈ V (S) and
f ∈ ⟨S⟩. Then f is a finite combination f = ai Pi for some Pi ∈ S (by definition),
therefore f (x) = 0, proving the reverse inclusion.
Part (3) utilizes the Hilbert Basis Theorem Theorem 26.1: the ideal ⟨S⟩ = ⟨P1 , . . . , Ps ⟩
for some Pi ∈ S, therefore
V (S) = V (⟨S⟩) = V (⟨P1 , . . . , Ps ⟩) = V (P1 , . . . , Ps ).
S T
S part (4), observe that since V (Si ) ⊃ V ( Si ) (by
For Spart(1)), it follows that V (Si ) ⊃
V ( Si ). We prove the reverse inclusion. If x ∈ V ( Si ), then for any i ∈ I, and any
P ∈ Si , P (x) = 0. Then x ∈ V (Si ), and since i is arbitrary, it is also in the intersection
T
V (Si ).
8 LEONARDO CONSTANTIN MIHALCEA
Remark 3.4. Once we know about morphisms of affine varieties, we will see that the
equivalence above may be upgraded to an equivalence of contravariant functors between
the category of
{ finitely generated k-algebras } = {k[x1 , . . . , xn ]/I}
to
{ (reduced) affine varieties with morphisms} = {Spec(k[x1 , . . . , xn ]/I)}.
Corollary 3.10. There is a one-to-one correspondence between maximal ideals in k[x1 , . . . , xn ]
and points in An . Any maximal ideal is of the form ⟨x1 − a1 , . . . , xn − an ⟩ for some
(a1 , . . . , an ) ∈ An , and the correspondence is given by
⟨x1 − a1 , . . . , xn − an ⟩ ↔ (a1 , . . . , an ).
Proof. By Corollary 3.9, the maximal ideals corresponds to minimal, non-empty, affine
varieties, i.e. single points. These ideals are precisely of the form stated. □
Example 3.11. Let n = 1, and let I ⊂ k[x] be an ideal. Since any ideal in k[x] is
principal, I = ⟨P (x)⟩. Because k is algebraically closed
P (x) = (x − a1 )n1 · . . . · (x − as )ns ,
for some ni ∈ N. Then the Nullstellensatz says that
p
I(V (P )) = I({a1 , . . . , as }) = ⟨(x − a1 ) · . . . · (x − as )⟩ = ⟨P ⟩.
Example 3.12. Take n = 2. If P (x, y) ∈ k[x, y] is a non-constant irreducible polynomial
then I(V (P )) = ⟨P ⟩ and V (P ) is a curve. (We will see this curve is irreducible.)
3.4. Proof of the Nullstellensatz. We start by proving a weaker version of Nullstel-
lensatz.
Theorem 3.2 (Weak Nullstellensatz). Let k be algebraically closed and let I be an ideal
such that I ̸= k[x1 , . . . , xn ]. Then V (I) ̸= ∅. Further, if I is maximal, then V (I) is a
single point in An .
The theorem has the following interpretation. Consider a system of polynomial equa-
tions
P1 (x1 , . . . , xn ) = 0
P2 (x1 , . . . , xn ) = 0
..
.
P (x , . . . , x ) = 0.
s 1 n
P
If there exist a1 , . . . , as ∈ k[x1 , . . . , xn ] such that ai Pi = 1, then obviously the system
has no solution. The theorem says that the converse is also true.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 11
Proof of Theorem 3.2. Since I ̸= k[x1 , . . . , xn ], we may find a maximal ideal J which
contains I. Since V (I) ⊃ V (J), we may assume I is maximal to start with. Define:
K := k[x1 , . . . , xn ]/I.
This is a k-algebra and since I is maximal K ⊃ k is also a field.
Lemma 3.13. K = k.
Proof. (In the case k is uncountable.) We will prove this in the case when k is uncount-
able. The main steps for the general proof are outlined in Remark 3.5 below.
If the extension K ⊃ k is algebraic, we are done, since k is algebraically closed.
Therefore we assume that there exist t ∈ K which is transcendent over k. Then we have
field extensions
P (t)
k ⊊ k(t) = { : P, Q ∈ k[x]} ⊂ K.
Q(t)
We claim that the elements
1
S := { : a ∈ k} ⊂ k(t)
t−a
are linearly independent over k. Indeed, assume that there exists a non-trivial finite
combination X ci
= 0; ci ∈ k.
t − ai
We may assume that ai ’s are distinct (otherwise collect the like terms). Consider a
coefficient ci0 , and multiply both sides by (t−a1t−a)·...·(t−as )
i
∈ k[t] to obtain
0
Now we use another result in commutative algebra, stating that for an integral exten-
sion of integral domains A ⊂ B, if B is a field, then A is also a field; see Lemma 26.1.
We apply this to the integral extension k[y1 , . . . , yN ] ⊂ K; then k[y1 , . . . , yN ] is a field.
This forces N = 0, thus K/k is an integral extension, and since k is algebraically closed,
k = K.
√
Proof of the strong
√ Nullstellensatz Theorem 3.1. We need to show that J = I(V (J)).
The inclusions J ⊂ I(V (J)) follows from definition.
To prove the converse, let J = ⟨f1 , . . . , fs ⟩ for some fi ∈ k[x1 , . . . , xn ]. We need to
prove the following. Let g ∈ k[x1 , . . . , xn ]; if g(a) = 0 for any a such that f1 (a) = . . . =
fs (a), then g N ∈ J for some number N . Define the ideal:
B := ⟨f1 , . . . , fs , 1 − gxn+1 ⟩ ⊂ k[x1 , . . . , xn , xn+1 ].
We first show that we must have that B = k[x1 , . . . , xn , xn+1 ]. Indeed, if B ̸= ⟨1⟩,
then there exists a maximal ideal m ⊃ B. By the Weak Nullstellensatz Theorem 3.2, it
follows that
m = ⟨x1 − a1 , . . . , xn − an , xn+1 − an+1 ⟩.
Then
f1 (a1 , . . . , an ) = . . . = fs (a1 , . . . , an ) = 1 − an+1 g(a1 , . . . , an ) = 0.
But the first s equalities imply that g(a1 , . . . , an ) = 0, which is impossible.
We now consider the situation when B = k[x1 , . . . , xn , xn+1 ] = ⟨1⟩. Then there exist
hi ∈ k[x1 , . . . , xn+1 ] such that
X
hs+1 (x, xn+1 )(1 − xn+1 g(x)) + hi (x, xn+1 )fi (x) = 1.
1
This is an identity in k[x1 , . . . , xn , xn+1 ]. We specialize to xn+1 = g(x)
to obtain
X 1
hi (x, )fi (x) = 1.
g(x)
After clearing denominators (which are simply powers of g), we obtain an identity of the
form X
h′i (x)fi (x)g(x)ni = g(x)N .
This shows that g N ∈ ⟨f1 , . . . , fs ⟩ = J, and we are done. □
3.5. Irreducibility. Let k be an arbitrary field. In this section we explore the first
properties of affine varieties, in relation to the algebra/geometry dictionary. We start
with the following definition.
Definition 3.3 (Coordinate ring). Let X ⊂ An be an affine variety and I := I(X). The
ring
Γ(X) := k[x1 , . . . , xn ]/I
is called the coordinate ring of X.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 13
In a sense which will become precise later, Γ(X) is the ring of (algebraic) regular
functions on X. Some examples (to avoid arithmetic issues, here we assume k is of
characteristic 0):
• Γ(An ) = k[x1 , . . . , xn ];
• If C1 = V (y − x2 ) ⊂ A2 is a parabola, Γ(C1 ) = k[x, y]/⟨y − x2 ⟩;
• If C2 is the union of x and y axes in A2 , then Γ(C2 ) = k[x, y]/⟨xy⟩.
The parabola and the union of lines are somehow different: the parabola is made of
‘one part’, the other is a union of two parts. We want to make this precise.
Definition 3.4. A topological space X is called reducible if it can be written as
X = X1 ∪ X2 where X1 , X2 ⊊ X are closed subsets. If X is not reducible, it is called
irreducible.
• The real affine line A1 = R with Euclidean topology is reducible: R = (−∞, 0] ∪
[0, ∞);
• The affine line A1k over any infinite field with Zariski topology is irreducible.
(Because the non-trivial closed sets are just finite sets of points.)
• The union of lines C2 from above is obviously reducible, but one can show that
the parabola is irreducible.
Proposition 3.14. Let X be an irreducible space. Then the following hold:
(1) If ∅ =
̸ U ⊂ X is an open subset, then U is dense in X.
(2) Any two nonempty subsets in X intersect.
(3) Any non-empty subset U ⊂ X (with induced topology) is irreducible.
Proof. Write X = U ∪ (X \ U ). Since X is irreducible one of the two closed subsets must
be X, forcing X = U .
We leave (2) and (3) as exercises. □
Proposition 3.15. Let X ⊂ An be an affine variety (with Zariski topology) and let
I(X). Then the following are equivalent:
(1) X is irreducible;
(2) The ideal I(X) is a prime ideal;
(3) The coordinate ring of X, Γ(X) is an integral domain.
Proof. Obviously (2) ⇔ (3), so we will prove (1) ⇔ (2).
”⇒” Assume that X is irreducible, and take f, g ∈ k[x1 , . . . , xn ] such that f g ∈ I(X).
Then ⟨f g⟩ ⊂ I(X), therefore X = V (I(X)) ⊂ V (f g) = V (f ) ∪ V (g). It follows that
X = (X ∩ V (f )) ∪ (X ∩ V (g)).
By irreducibility it follows that either X ⊂ V (f ) or X ⊂ V (g), i.e. f ∈ I(X) or
g ∈ I(X).
”⇐” Assume that I(X) is prime, and write X = X1 ∪ X2 where X1 , X2 are closed.
First observe that
(3.2) I(X1 ∪ X2 ) = I(X1 ) ∩ I(X2 )
14 LEONARDO CONSTANTIN MIHALCEA
Observe that this morphism is well defined. For that, it suffices to show that F ∗ (Q) ∈
I(X) for any Q ∈ I(Y ). Utilizing that F ∗ is an algebra homomorphism, and Equa-
tion (3.4), we obtain
F ∗ (Q) = Q(f1 (x), . . . , fs (x)) ∈ I(X),
as needed. We denote by Hom(X, Y ) the set of morphisms from X to Y .
Given a k-algebra homomorphisms φ : Γ(Y ) → Γ(X) we may define a morphism
Fφ : X → Y by
Fφ (x) = (f1 (x), . . . , fs (x))
where fi (x) are any lifts to k[x1 , . . . , xp ] of φ(xi ). Clearly, Fφ∗ = φ.
Lemma 3.16. The assignments F 7→ F ∗ and φ → Fφ are compatible with composition,
i.e.
(F ◦ G)∗ = G∗ ◦ F ∗ ; Fφ◦ψ = Fψ ◦ Fφ .
Proposition 3.17. Let X ⊂ Ap , Y ⊂ As be two affine varieties. There is a one-to-
one correspondence between morphisms F = (f1 , . . . , fs ) ∈ Hom(X, Y ) and k-algebra
homomorphisms Homk−alg (Γ(Y ), Γ(X)) given by F 7→ F ∗ .
Proof. One needs to check that for any F ∈ Hom(X, Y ), FF ∗ = F , and that for any
φ ∈ Homk−alg (Γ(Y ), Γ(X), Fφ∗ = φ. We already checked the second equality. For the
first, take x ∈ X, and assume that F = (f1 , . . . , fs ). Then F ∗ (xi ) = fi and
FF ∗ (x) = (F ∗ (x1 ), . . . , F ∗ (xs )) = (f1 (x), . . . , fs (x)) = F (x).
□
Remark 3.6. For k algebraically closed, Proposition 3.17 says that the contravariant
functor Γ from the category of affine varieties and morphisms to finitely gener-
ated, reduced, k-algebras and k-algebra homomorphisms, sending X 7→ Γ(X)
is an equivalence of categories.
We now record some algebraic properties of morphisms which correspond to geometric
information. To start, Proposition 3.17 implies that:
Corollary 3.18. A morphism F : X → Y is an isomorphism if and only if F ∗ : Γ(Y ) →
Γ(X) is a k-algebra isomorphism.
In examples below we take char k = 0.
Example 3.19 (Add picture). Consider the parabola C := V (y − x2 ) ⊂ A2 . There
is a morphism F : A1 → C defined by F (t) = (t, t2 ). This is an isomorphism, with
inverse G(x, y) = x. In fact, F induces the isomorphism F ∗ : k[x, y]/⟨y − x2 ⟩ → k[t],
F (x) = t, F (y) = t2 .
Example 3.20 (Add picture). Consider the cusp C = V (y 2 − x3 ) and F : A1 → C
defined by F (t) = (t2 , t3 ). This is a bijective morphism which is not an isomorphism.
16 LEONARDO CONSTANTIN MIHALCEA
Proof. (a) Assume that F = (f1 , . . . , fs ). For any P (x) ∈ Γ(Y ), F ∗ (P (x)) = P (f1 , . . . , fs ).
Then P (x) ∈ ker(F ∗ ) if and only if P (f1 , . . . , fs ) ∈ I(X), i.e. for all x ∈ X, F (x) ∈
V (P ) ∩ Y .
If F ∗ is not injective then P may be chosen outside I(Y ), therefore the open set
DP = Y \ V (P ) is non-empty, and disjoint from the image. Conversely, if the image is
not dense then we can find an open set U ⊂ Y such that Im(F ) ⊂ Y \ U . The latter is
a nonempty closed subset not equal to Y , therefore there exists P ∈ I(Y \ U ) \ I(Y ).
Then 0 ̸= P ∈ ker(F ∗ ).
To prove (b), observe that by (a) Γ(Y ) is a subalgebra of Γ(X). Since X is irreducible,
Γ(X) is a domain, therefore so must be Γ(Y ). □
3.7. Exercises. 1. Prove that the topology on A2 = A1 ×A1 is not the product topology.
2. Prove that the distinguished open sets Df form a basis of open sets for the Zariski
topology.
3. (a) Prove that if k is infinite, then I(An ) = (0). Find a counterexample for this in
a finite field different from k = Fp from Example 3.6 above.
(b) Fix a ∈ An . Find I(a).
(c) Let k = Fp (the field with p-elements. Find I(A1k ). Is it prime ?
5. Prove Theorem 3.3. (One proof relies on the fact that affine varieties are Noether-
ian.)
18 LEONARDO CONSTANTIN MIHALCEA
The Lemma suggests that if one wants to look at zeros of polynomials which are inde-
pendent of choices of projective coordinates, then it suffices to restrict to homogeneous
polynomials.
Definition 4.1. Let S ⊂ k[x0 , . . . , xn ] be a subset consisting of homogeneous polynomi-
als. Define
Vp (S) := {[x0 , . . . , xn ] : F (x0 : . . . : xn ) = 0, ∀F ∈ S}.
A projective algebraic variety is any subset of Pn of the form Vp (S).
The function Vp sending S 7→ Vp (S) has properties very similar to those from the
affine case:
• VP is decreasing;
• Vp (S) = Vp (⟨S⟩); in particular we may assume that S is finite;
• Vp (0) = Pn ;
• intersections of projective varieties are projective; finite unions of projective va-
rieties are projective
20 LEONARDO CONSTANTIN MIHALCEA
The last property allows us to (re)define Zariski topology on Pn , by requiring that Vp (S)
are closed subsets.
One difference is the behavior with respect to (homogeneous) maximal ideals. Let
m = ⟨x0 , . . . , xn ⟩ ⊂ k[x0 , . . . , xn ]. Then
Vp (m) = ∅.
Because of this, we will call ⟨x0 , . . . , xn ⟩ the irrelevant ideal. In the affine case, the
irrelevant ideal is maximal, and it corresponds to the point (0, . . . , 0) ∈ An+1 . One may
ask what are the ideals corresponding to points. Let [a0 , . . . , an ] ∈ Pn , and pick ai ̸= 0;
w.l.o.g. a0 ̸= 0. Then
a1 an a1 an
[a0 , . . . , an ] = [1 : : . . . : ] = Vp (x1 − x0 , . . . , xn − x0 ).
a0 a0 a0 a0
One may go from an affine variety to a projective variety, by the process called ho-
mogenization. If f (x1 , . . . , xn ) ∈ k[x1 , . . . , xn ], then we may define
x1 xn
F (x0 , . . . , xn ) = xdeg
0
f
f ( , . . . , ).
x0 x0
Note that this is a homogeneous polynomial. Then one may homogenize each of the
generators of the ideal of the affine variety.
To illustrate, consider the parabola P = V (x22 −x1 ) and the hyperbola C = V (x1 x2 −1).
Recall that P and C are not isomorphic when regarded as affine varieties (because their
coordinate rings are not isomorphic - see homework). The homogenizations of the two
equations are:
F1 := x22 − x1 x0 ; F2 = x1 x2 − x20 .
Since F1 , F2 are the same up to rearranging labels, Vp (F1 ) will be isomorphic to Vp (F2 ).
Therefore, the projective completion of the parabola is the same as the projectiviza-
tion of the hyperbola. But let’s see what are the points at infinity are added in each
case. We have
V (F1 ) ∩ (x0 = 0) = {[0 : 1 : 0]} ; V (F2 ) ∩ (x0 = 0) = {[0 : 1 : 0], [0 : 0 : 1]}.
is given by
1 Z0 Z1 . . . Zn−1
νd (P ) = {[Z0 : . . . : Zn ] : rank ≤ 1}.
Z1 Z2 . . . Zn
A similar determinantal description may be given for the image of the Segre embedding.
TODO.
Further, both maps are injective, and in fact are isomorphisms onto their images (to
be proved later). As a consequence, the product of projective spaces is a projective
variety.
However, asking for all morphisms to be of this form is too strong. For example,
consider a potential map
φ : P2 → P1 ; [x0 : x1 : x2 ] 7→ [F0 (x) : F1 (x)],
where F0 , F1 ∈ k[x0 , x1 , x2 ] are homogeneous polynomials of the same degree. We will
see later that codim Vp (F0 ) ∩ Vp (F1 ) ≤ 2; in other words the common zero locus of F0 , F1
will be nonempty in P2 . A similar argument implies that there are no non-constant
morphisms P2 → P1 . In practice, we will write morphisms form any projective variety
φ : Z → Pm as in Equation (4.1) above; we will allow common zero loci, but then we
need to analyze whether one can use other relations (e.g. the equations of Z) to resolve
the indeterminacies.
In general, a morphism will be defined by local data, i.e. φ : Z → Pm is a morphism
iff for any affine covering Vi of Pn and any affine covering Ui of Z, the restrictions of φ
to φ−1 (Vi ) ∩ Uj → Ui give a morphism of affine varieties. More on this is discussed in
§8 below.
This leads us to the next section.
4.8. Exercises. 1. Prove Lemma 4.4.
24 LEONARDO CONSTANTIN MIHALCEA
The restriction map rV U isSgiven by the usual restriction of functions. Both are in fact
sheaves, as a map f : U = Ui → Y is continuous if and only if all f|Ui are continuous.
A morphism of presheaves φ : F → G is a collections of maps φU : F(U ) → G(U )
which commutes with restrictions:
U φ
F(U ) −−−→ G(U )
(5.1) rV U
y
rV U
y
V φ
F(V ) −−−→ G(V )
Example 5.2. (A presheaf which is not a sheaf.) The sheaf of constant K-valued func-
tions on a topological space X is a presheaf, but in general not a sheaf. Indeed, if X
is disconnected, then constant sections over the components may not glue to a global
section, unless all section values are the same.
Definition 5.2. (Stalks) Let F be a sheaf on X and x ∈ X. The collection F(U ), with
x ∈ U forms an inverse system. Define
Fx := lim F(U ).
←−
x∈U
Definition 5.5. (Morphisms of ringed spaces.) Let (X, OX ) and (Y, OY ) be two ringed
spaces. A morphism of ringed spaces φ : (X, OX ) → (Y, OY ) is given by a continuous
map φ : X → Y such that
∀ V ⊂ Y open, and ∀g ∈ OY (V ), g ◦ φ ∈ OX (φ−1 (V )).
For V ⊂ Y open, denote by
φ∗V : OY (V ) → OX (φ−1 (V )); g 7→ g ◦ φ.
This is a homomorphisms of k-algebras.
Remark 5.1. The assumption that OX is a sheaf of functions means that the algebra
homomorphisms φU commute with restriction homomorphisms, i.e. for any V ⊂ U , the
following diagram commutes:
φ∗U
OY (U ) −−−→ OX (φ−1 (U ))
rφ−1 (V )φ−1 (U ) r
V Uy
y
φ∗
V
OY (V ) −−−→ OX (φ−1 (V ))
Without this assumption, the commutation of the diagram will be part of the data of
what it means to be a morphism of ringed spaces.
Example 5.4. If we consider (X, OX ) to be a differential manifold endowed with the
sheaf of C ∞ functions on X, then the condition that φ : (X, OX ) → (Y, OY ) is a
morphism is equivalent to requiring that φ is a differentiable map.
5.3. Affine algebraic varieties as ringed spaces. Let X = V (I) ⊂ An be an affine
variety, where I is a radical ideal. We would like to define a sheaf of regular functions OX
so that (X, OX ) is a ringed space. We already know what the global regular functions
should be:
OX (X) = Γ(X) = k[x1 , . . . , xn ]/I.
In order to define the regular functions in OX (U ) for U ⊂ X arbitrary open set, the
strategy is to cover it by distinguished open sets,
U = Df1 ∪ . . . ∪ Dfp ,
define regular functions OX (Dfp ), then ‘glue’ these functions to give regular functions
on U . To make this precise we need a lemma.
Lemma 5.5. Let X be a topological space, U a basis of open sets in X, and K a set.
Assume we are given an assignment:
U 7→ F(U ) = {f : U → K : functions }, ∀U ∈ U
such that:
• (Restriction) If V ⊂ U are in U, and s ∈ F(U ), then s|V ∈ F(V ).
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 27
where the second equality follows from Equation (5.4). This implies that in Dfi0 ,
P (x) Pi0 (x)
m
=
f (x) fi0 (x)N
□
Remark 5.2. It is possible that Df1 = Df2 for different f1 , f2 ∈ Γ(X). Then
p p
⟨f1 ⟩ = ⟨f2 ⟩
and as in the beginning part of the proof above one checks that Γ(X)f1 = Γ(X)f2 .
30 LEONARDO CONSTANTIN MIHALCEA
Remark 5.3. Finding sections of OX (U ) for arbitrary U is in general much more subtle.
For instance, one can show that
OA2 (A2 \ 0) = OA2 (A2 ).
This corresponds to the classical fact from complex analysis that a multivariable analytic
function may be extended over sets of codimension ≥ 2.
5.4. Affine algebraic varieties, revisited. We are now in position to re-define, and
slightly extend, the notion of affine algebraic variety from the beginning of the notes.
Definition 5.6. An affine algebraic variety is a ringed space (X, OX ) isomorphic
(as ringed spaces) with (V, OV ), where V = V (I) ⊂ An is an ‘old’ affine variety together
with its structure sheaf.
A morphism of affine algebraic variety is a morphism of ringed spaces.
One advantage to viewing affine algebraic varieties as ringed spaces is that this is
an intrinsic notion, independent of coordinates. The next proposition shows that when
viewed this way, the set of ‘new’ affine varieties is strictly larger than the ‘old’ set of
affine varieties.
Proposition 5.6. Let V = V (I) ⊂ An be an affine algebraic variety, and let 0 ̸= f ∈
Γ(V ). Then (Df , (OV )|Df ) is an affine algebraic variety.
Proof. Consider a representative f ∈ k[x1 , . . . , xn ] for the element f ∈ Γ(X). Let
I = ⟨f1 , . . . , fp ⟩, and define an ideal J ⊂ k[x1 , . . . , xn , xn+1 ] by
J = ⟨f1 , . . . , fp , 1 − xn+1 f ⟩.
Define φ : Df → V (J) by
1
(x1 , . . . , xn ) 7→ (x1 , . . . , xn , ).
f (x1 , . . . , xn )
Observe that φ is a continuous bijection. (One way to see continuity is to restrict to
the basis of distinguished open sets.) The inverse is the projection π : V (J) → Df ,
π(x1 , . . . , xn , xn+1 ) = (x1 , . . . , xn ), which is obviously continuous; thus φ is a homeomor-
phism. In particular, as sets, φ(Df ) = V (J). We now need to show that J is a radical
ideal. To do that we define a ring homomorphism
1
φ∗ : k[x1 , . . . , xn , xn+1 ]/J → Γ(V )f ; xi 7→ xi (1 ≤ i ≤ n); xn+1 7→ .
f
This is induced from a similar ring homomorphism k[x1 , . . . , xn , xn+1 ]/J → k[x1 , . . . , xn ]f ,
and it is easy to see it is well defined and surjective. We prove it is injective. Let
P (x1 , . . . , xn , xn+1 ) ∈ k[x1 , . . . , xn+1 ] and write
X j
P (x1 , . . . , xn , xn+1 ) = xn+1 Pj (x1 , . . . , xn ).
j
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 31
Let x = (x1 , . . . , xn ) and let N be the largest power of the powers of xjn+1 . Assume that
φ∗ (P (x1 , . . . , xn , xn+1 ) = 0, i.e.
X f (x)N −j Pj (x)
N
= 0 ∈ Γ(X)f .
j f (x)
This means that X X
f (x)s f (x)N −j Pj (x) = ai (x)fi (x) ∈ I.
j
6. Algebraic varieties
6.1. Definition and basic properties.
Definition 6.1. A topological space X is quasi-compact if any open covering has a
finite subcover.
Note that we do not require that the space is Hausdorff: a quasi-compact space which
is Hausdorff is called compact.
Lemma 6.1. (a) Closed subsets of a quasi-compact space X are quasi-compact.
(b) Finite unions of quasi-compact sets are quasi-compact.
Proof. Same as for the ‘compact’ situation, thus omitted. □
Example 6.2. Affine algebraic varieties are quasi-compact but in general they are not
compact.
Lemma 6.3. Any affine algebraic variety X is quasi-compact. (But in general it is not
Hausdorff, thus not compact.)
Proof. Exercise. □
Lemma 6.4. Let X be an affine algebraic variety and U ⊂ X open. Then U is quasi-
compact.
Proof. We may assume X ⊂ An . Take X \ U = V (f1 , . . . , fp ) for fi ∈ Γ(X). Then
U = Df1 ∪ . . . ∪ Dfp . Each of Dfi is affine by Proposition 5.6, thus quasi-compact. Then
the claim follows from Lemma 6.1. □
Definition 6.2. An algebraic variety is a quasi-compact ringed space (X, OX ) which
is locally isomorphic to an affine variety, in the following sense:
For any x ∈ X there exists an open set U containing x such that (U, (OX )|U ) is an
affine variety.
If U ⊂ X is an open set which is isomorphic to an affine algebraic variety then U is
called an affine open set.
To simplify notation, when we speak about varieties, affine varieties, etc, instead of
writing (X, OX ) we will simply write X, and we omit the associated sheaf of regular
functions.
Lemma 6.5. Let X be an algebraic variety. Then the affine open sets form a basis of
open sets for the topology of X.
Proof. Let U ⊂ X be open. We need S to show that U may be written as S a union of
affine open sets. By definition X = Ui where each Ui is affine, thus U = (U ∩ Ui ).
It suffices to show that U ∩ Ui is a union of affine algebraic varieties. W.l.o.g. we may
34 LEONARDO CONSTANTIN MIHALCEA
To start, take f ∈ Γ(X) and f its image in Γ(Z) = RZ (Z). If f|Z = 0, there is nothing
to do, as Df = ∅. Otherwise,
a
RZ (Df ) = { r : a ∈ Γ(Z)} ⊂ OZ (Df ).
f
Conversely, take V ⊂ Z open and s ∈ OZ (V ). Since the distinguished open sets form a
basis for the topology of Z, it follows that locally s is given by restrictions of sections
si ∈ OX (Dgi ) = Γ(X)gi ,
where V = ∪(Z ∩ Dgi ) and gi ∈ Γ(X). The restrictions of si give sections si of OZ over
Dgi = Dgi ∩ Z. By definition, the restrictions si = (si )|Dgi ∩Z are in RZ (Dgi ), and since
both OZ and RZ are sheaves, it follows that
OZ (V ) ⊂ RZ (V ).
This finishes the proof. □
Remark 6.2. We shall see later that the proof above shows that there is a surjective
homomorphism of sheaves OX → OZ . The proof checks surjectivity on stalks of these
sheaves.
Example 6.9. Let Z ⊂ P2 given by equations x0 x1 = 0, x0 (x0 − x2 ) = 0. Then Z =
Z1 ∪ Z2 has two components, where Z1 ≃ P1 is given by x0 = 0 and Z2 = {[1 : 0 : 1]}.
Take any open set U ⊂ P2 such that Z ⊂ U . This is equivalent to requiring that
P2 \ U ⊂ P2 \ Z, in particular, (P2 \ U ) ∩ V (x0 ) = ∅. By Bézout this implies that P2 \ U
is a finite set, implying that
OP2 (P2 \ U ) = OP2 (P2 ) = k,
i.e. constant sections. This implies that the restriction
OP2 (P2 \ U ) → OZ (Z)
is not surjective. (Indeed, OZ contains more than constant sections, e.g. sections con-
stant on each of the components.)
Proposition 6.10 (Irreducible decomposition). Let X be an algebraic variety. Then
X = X1 ∪. . .∪Xp where Xi is a closed irreducible algebraic variety and no Xi is included
in another Xj .
Proof. Since X is quasi-compact, it can be covered by finitely many affine varieties
Ui . Each Ui has a decomposition by irreducible components (Theorem 3.3). After
taking closures, and possibly removing redundant components one obtains the claimed
decomposition. □
Definition 6.3. A closed subset of a variety with its induced sheaf of functions is called
a subvariety.
36 LEONARDO CONSTANTIN MIHALCEA
Remark 6.3. The notion of subvariety differs among authors. In English circles, a
subvariety is usually closed and irreducible, while in textbooks such as that of Perrin, it
is simply an intersection of an open and closed subset of a variety (with induced ringed
space structure).
6.2. Local rings. The sheaf of regular functions of an algebraic variety may be defined
as an intrinsic object. The main tool to do this is the notion of local ring, which will
turn to be closely related to the notion of stalks, and to rings of fractions.
Definition 6.4. Let (X, OX ) be an algebraic variety, endowed with its sheaf of regular
functions. Fix x ∈ X and consider the set of pairs (U, f ), where x ∈ U ⊂ X is open and
f ∈ OX (U ). Define the equivalence relation
(U, f ) ≃ (V, g) ⇐⇒ ∃x ∈ W ⊂ U ∩ V open s.t.f|W = g|W .
The equivalence class of (U, f ) is called the germ of f at x, denoted by fx .
The set of all germs fx is denoted by OX,x , and it is called the local ring of X at x.
From a topological point of view, the set OX,x is clearly ‘local’, as it depends only on
the neighborhood of x. Algebraically, the ‘local’ terminology is justified by the following
proposition.
Proposition 6.11. The set OX,x is a local k-algebra. More precisely, it is a k-algebra
with the unique maximal ideal
mx := {fx ∈ OX,x : f (x) = 0}.
Furthermore,
OX,x /mx ≃ k.
Proof. Clearly OX,x has a ring structure given by adding and multiplying germs:
(U, f ) + (V, g) = (U ∩ V, f + g); (U, f ) · (V, g) = (U ∩ V, f · g).
(One should check these operations are well defined; this is left as an exercise.) Note
that the 0 of this ring is represented by a germ of the function identically 0 on some
neighborhood of x. There is a surjective ring homomorphism:
ex : OX,x → k; (U, f ) 7→ f (x).
The kernel of ex is ker(ex ) = mx , and by the first isomorphism theorem
OX,x / ker(ex ) = OX,x /mx ≃ k.
This implies that mx is a maximal ideal.
Furthermore, any element in fx ∈ OX,x \ mx is invertible, with inverse (1/f )x . Indeed,
we may choose an open affine set x ∈ U such that f is defined on U (cf. Lemma 6.5).
The hypothesis that f (x) ̸= 0 means that x ∈ Df . Then the inverse 1/f ∈ OU (Df ) is a
regular function on Df , and it determines the germ (1/f )x . This implies that mx is the
unique maximal ideal, and finishes the proof. □
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 37
Recall from Theorem 3.2 that if k is algebraically closed, there is a one-to-one corre-
spondence between points x ∈ An and maximal ideals mx ⊂ k[X1 , . . . , Xn ] given by
x = (x1 , . . . , xn ) ←→ ⟨X1 − x1 , . . . , Xn − xn ⟩.
If J is any ideal in k[X1 , . . . , Xn ], the ideals in the quotient ring k[X1 , . . . , Xn ]/J are in
one-to-one correspondence to the ideals in k[X1 , . . . , Xn ] containing J. Combining the
two gives a one-to-one correspondence between points x ∈ V (J) and maximal ideals in
k[X1 , . . . , Xn ]/J given by
x ∈ V (J) ←→ mx /J ⊂ k[X1 , . . . , Xn ]/J.
Proposition 6.12. Let X be an algebraic variety, x ∈ X, and let x ∈ U ⊂ X be any
affine open set, with coordinate ring R := Γ(U ). Let mx ⊂ R be the maximal ideal
corresponding to x. Then the following hold:
(a) There is an isomorphism Rmx → OX,x , which is local, i.e. it sends the (unique)
maximal ideal in Rmx to mx ⊂ OX,x .
(b) There are one-to-one correspondences between the following three sets:
• irreducible (closed) subvarieties X containing x;
• prime ideals of Rmx ;
• prime ideals p ⊂ mx ⊂ R.
Proof. W.l.o.g. we may assume that U ⊂ An and that x = (x1 , . . . , xn ). Elements of Rmx
P
are (classes of) rational functions of the form Q where P, Q ∈ k[X1 , . . . , Xn ] and Q ∈
/ mx .
This last condition means that Q(x) ̸= 0; in particular, the germ Qx is invertible in OX,x .
Then we may define a ring homomorphism
P Px
φx : Rmx → OX,x ; 7→ .
Q Qx
I leave it as an exercise to check that φx is an isomorphism. Furthermore,
P
φ−1
x (mx ) = { ∈ Rmx : P (x) = 0} ⊂ mx Rmx .
Q
This shows that the isomorphism preserves the maximal ideals, therefore it is local.
To prove (b), note that the set of irreducible subvarieties of X containing x is in bijec-
tion to the set of irreducible subvarieties of U containing x. (This bijection is obtained
by intersecting with U , and the inverse by taking closure in X.) By Nullstellensatz, the
latter is in bijection to the set of prime ideals p in R = Γ(U ) such that p ⊂ mx . The
last equivalence follows from the description of ideals in rings of fractions; see [AM69,
Prop. 3.11]. □
Proposition 6.13. Let f : (X, OX ) → (Y, OY ) be a morphism of varieties. Let x ∈ X
and y = f (x). Then f induces local homomorphism of local rings f # : OY,y → OX,x , i.e.
a homomorphism such that (f # )−1 (mx ) = my .
38 LEONARDO CONSTANTIN MIHALCEA
Definition 7.1. Let X := Vp (J) ⊂ Pn and F ∈ k[x0 , . . . , xn ] of degree > 0. There exists
a unique sheaf of k−valued functions on X, denoted by OX , such that
P
OX (DF+ ) := Γp (X)(F ) = { r : P homogeneous and deg P = r deg(F )}.
F
Proof. The proof is based again on Lemma 5.5, and it follows the outline of Proposi-
tion 5.6. The only observation to make is that elements in Γp (X)(F ) are indeed functions
on X, independent of choices of representatives of points in Pn . □
Remark 7.1. The condition that deg(F ) > 0 is needed. Indeed, if F is a constant,
let’s say F = 1, then Γ(X)(F ) contains only constants, so OX (DF+ ) consists of constant
sections. But this does not lead to a sheaf (e.g. X may be disconnected and there are
functions constant on each of the connected components).
Proposition 7.2. Let X := Vp (J) ⊂ Pn . Then the ringed space (X, OX ) is an algebraic
variety.
Any ringed space which is isomorphic with (Vp (J), OVp (J) ) is called a projective
variety.
Proof. We first reduce the statement to the situation when X = Pn , therefore we assume
for now that (Pn , OPn ) is an algebraic variety. Then any closed subset X of Pn is quasi-
compact.
Next we need to check that X is covered by affine open sets. Clearly
[n
Pn = Dx+i .
i=0
We will prove in Proposition 7.5 below that each (Dx+i , (OPn )|Dx+ ) is isomorphic to the
i
affine variety (An , OAn ). Then X ∩ Dx+i with restricted sheaf of functions is (isomorphic
to) a closed variety of An .
One subtlety is that one needs to check that the restriction of the sheaf OX to Dx+i
coincides with the sheaf of regular functions on X ∩ Dx+i when regarded as a subvariety
in Dx+i . This follows because the sheaf OX is the sheafification of the restriction sheaf of
regular functions OPn (as in Proposition 6.8). □
7.3. From affine to projective and back: homogenization and dehomogeniza-
tion. Consider the polynomial ring k[x0 , . . . , xn ] and fix an indeterminate, x0 for sim-
plicity. We will define two operations (depending on x0 ):
• (dehomogenization) For a homogeneous polynomial F ∈ k[x0 , . . . , xn ] the de-
homogenization is
F♭ := F (1, x1 , . . . , xn ) ∈ k[x1 , . . . , xn ].
This is a ring homomoprhism k[x0 , . . . , xn ] → k[x1 , . . . , xn ] (since the evaluation
map is a ring homomorphism). One may also apply this to ideals: if J = ⟨Pi :
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 41
Using that DF+ ∩ Dx+0 = Dx+0 F , one may check (exercise!) that:
P (x , . . . , x ) P (1, x1 , . . . , xn )
∗ 0 n
φ = .
(x0 F (x0 , . . . , xn ))r F (1, x1 , . . . , xn )r
P (x1 ,...,xn )
To define the map (φ−1 )∗ observe that any fraction of the form (f (x1 ,...,xn ))r
can be
P (x0 ,...,xn )
completed to a unique fraction of the form of degree 0, where P |P and r′ ≥ r.
(x0 f ♯ )r′
♯
(Multiply the denominator with appropriate power of x0 f ♯ , then the numerator with
appropriate power of x0 . Then
P (x , . . . , x ) P (x0 , . . . , xn )
−1 ∗ 1 n
(φ ) = .
(f (x1 , . . . , xn ))r (x0 f ( xx10 , . . . , xxn0 ))r′
and that
deg P
P (x , . . . , x )
−1 ∗ 1 n r deg(f )−deg(P ) x0 P ( xx10 , . . . , xxn0 )
(φ ) = x0 .
(f (x1 , . . . , xn ))r (xdeg
0
f
f ( xx1 , . . . , xxn ))r
0 0
To prove the claim, take F ∈ J and as in Example 7.4 write F = xd0 f ♯ for some
f ∈ k[x1 , . . . , xn ]. Since xd0 f ♯ = 0 on X ⊂ Y it follows that f = 0 on X = X ∩ U0 , i.e.
√
f ∈ I(X) = I(V (I)) = I
by the Nullstellensatz. Then f p ∈ I for some p ∈ Z≥0 and
(f p )♯ = (f ♯ )p ∈ I ♯
√ √ √
thus f ♯ ∈ I ♯ . It follows that F = xd0 f ♯ ∈ I ♯ , giving that J ⊂ I ♯ , as claimed.
To prove (b), we observe first that
⟨f ⟩♯ = ⟨{(f g)♯ = f ♯ g ♯ : g ∈ k[x1 , . . . , xn ]}⟩ = ⟨f ♯ ⟩.
Then by (a), V (f ) = V (f ♯ ). □
Example 7.8. (Homogenization does not commute with generation.) Let I = ⟨x1 , x2 −
x21 ⟩ ⊂ k[x1 , x2 ]. The associated variety is V (I) = {(0, 0)} and by Proposition 7.7,
V (I ♯ ) = {[1 : 0 : 0]} ⊂ P2 .
However, the ideal generated by homogenizations J = ⟨x1 , x2 x0 − x21 ⟩ has zero locus
̸ Vp (I ♯ ).
Vp (J) = {[1 : 0 : 0], [0 : 0 : 1]} =
Geometrically, V (I) is the transversal intersection of a conic Q with an affine line L;
this intersection contains one point, but Q ∩ L ⊊ Q ∩ L, since the latter contains a
second point at infinity.
44 LEONARDO CONSTANTIN MIHALCEA
8. Morphisms (II)
The goal of this section is to give some examples of morphisms to affine varieties, and
morphisms between projective varieties.
8.1. Morphisms to affine varieties. Let φ : (X, OX ) → (Y, OY ) be a map.
Proposition 8.1. The condition that φ is a morphism is local:
S
• φ is a morphism if and only ifSthere exists an open cover Y = Vi , and for each
i, an open cover φ−1 (Vi ) = Uij , such that the restriction φi : Uij → Vi is a
morphism.
Proof. Immediate from definition, using that sections over Uij glue to a section of
OX (φ−1 (Vi )); cf. Definition 5.5. □
Informally, this says that if we have a globally defined map φ : X → Y , the fact that it
is a morphism may be checked in local charts. (This is similar to what happens in other
situations, such as differentiable maps, or holomorphic functions.) One can simplify this
statement in case Y is affine.
Proposition 8.2. If (Y, OY ) is an affine variety, then it suffices to check the morphism
condition for global sections in Γ(Y ) = OY (Y ):
• φ is a morphism if and only if for all f ∈ Γ(Y ), φ∗ (f ) = f ◦ φ ∈ OX (X).
Proof. If X is affine, this follows from Proposition 3.17. In general, cover X by affine
varieties, and apply the previous proposition together with the affine situation. □
Example 8.3. The projection p : An+1 \ 0 → Pn is a morphism. (All components are
global sections on An+1 \ 0, with no common zeros.) Note that we cannot extend this
morphism over 0.
In the simplest case, when n = 1, this map is sending a point (x, y) ∈ A2 to the line
[x : y] ∈ P1 . One may take a modification A f2 → A2 which over (0, 0) has fibre all
possible slopes of lines through (0, 0); in other words, replace (0, 0) ∈ A2 by a P1 . This
will lead to the notion of the blow-up of a point, and to a morphism A f2 → P1 extending
A2 → P1 over (0, 0).
Remark 8.1. If φ : (X, OX ) → (Y, OY ) is a morphism and V ⊂ X, W ⊂ Y are
subvarieties such that φ(V ) ⊂ W , then the restriction φ : V → W is a morphism.
8.2. Morphisms to projective varieties. Many morphisms between projective vari-
eties arise from the following construction.
Consider b + 1 homogeneous polynomials of the same degree F0 , . . . , Fb ∈ k[x0 , . . . , xa ].
Then there is a well defined map:
φ : Pa \ Vp (F0 , . . . , Fb ) → Pb ; [x0 : . . . : xa ] 7→ [F0 (x) : . . . : Fb (x)].
Proposition 8.4. The above map φ is a morphism.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 45
Proof. Cover the image Pb with the standard affine open sets: Pb = Dx+i . The preimage
S
9. Products
In this section we define the product of two varieties. We certainly want that Ar ×As =
A , but recall from a previous homework problem that the topology on Ar+s is not
r+s
the product topology. It turns out that the correct way to define products is to use the
categorical notion of product.
W
pX
p
$
X ×Z Y /& X
pY πX
πY f
g
Y / Z
The product X ×Z Y , if it exists, is unique up to an isomorphism commuting with the
projections.
Example 9.1. Take C = Sets the category of sets and maps between sets. Take A, B, C
sets and f : A → C, g : B → C. Then the product is
There is also the (dual) notion of categorical coproduct, which satisfies a similar
universality property, except that all arrows are reversed.
Example 9.2. Let C =k−Alg be the category of k-algebras (k a field), and k-algebra
homomorphisms. Let A, B, C, D be k-algebras and maps A → B, A → C, B → D
and C → D such that the resulting diagram commutes. (In other words, B, C, D are
A-algebras.) The morphisms B → D and C → D induce an A-bilinear k-algebra homo-
morphism B × C → D. The coproduct B ×A C is the algebra tensor product B ⊗A C.
50 LEONARDO CONSTANTIN MIHALCEA
B ⊗O A C o CO
iB
Bo A
9.2. Products of varieties. The main theorem of this section is the following:
Theorem 9.1. Any two algebraic varieties X, Y have a product (as an algebraic variety).
Note that there is no choice for what should be the underlying set of X × Y : it’s just
the set theoretic product of X and Y . We need to realize this as a ringed space, locally
isomorphic to an affine variety. To do this, the idea is to construct first products of
affine varieties, then ‘glue’ these to get a variety structure.
Remark 9.1. One may also ask where is Z from X ×Z Y : in this case, Z is a point,
regarded as an (affine) algebraic variety. Any variety maps to pt, and X ×pt Y is denoted
simply by X × Y .
Proposition 9.3. Let X, Y be affine algebraic varieties with coordinate rings Γ(X) and
Γ(Y ). Then the following hold:
(1) There is a product X × Y which is an affine algebraic variety with coordinate
ring Γ(X) ⊗k Γ(Y );
(2) A basis for the topology of open sets is given by the distinguished open sets Df ,
where X
f= fi (x)gj (y); fi ∈ Γ(X), gj ∈ Γ(Y ).
(3) The local ring OX×Y,(x,y) is the localization of Γ(X) ⊗k Γ(Y ) at the maximal ideal
mx ⊗ Γ(Y ) + Γ(X) ⊗ my .
Proof. Let X = V (f1 , . . . , fn ) ⊂ Ar and Y = V (g1 , . . . , gm ) ⊂ As . Then the set X ×
Y ⊂ Ar+s is the zero locus of the polynomials fi , gj regarded in k[x1 , . . . , xr ; y1 , . . . , ys ].
Furthermore,
k[x1 , . . . , xr ; y1 , . . . , ys ]/⟨fi ; gj ⟩ ≃ k[x1 , . . . , xr ]/⟨fi ⟩ ⊗k k[y1 , . . . , ys ]/⟨gj ⟩ = Γ(X) ⊗k Γ(Y ).
(There is a natural bilinear map
k[x1 , . . . , xr ]/⟨fi ⟩ × k[y1 , . . . , ys ]/⟨gj ⟩ → k[x1 , . . . , xr ; y1 , . . . , ys ]/⟨fi ; gj ⟩,
which by the universal property induces a map from the tensor product. The latter
is an isomorphism when regarded as k-vector spaces, so it must be an isomorphism in
general.)
We now prove that X × Y is a categorical product. Obviously there are projections
πX : X × Y → X and πY : X × Y → Y , corresponding to maps Γ(X) → Γ(X × Y ),
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 51
We now need to check that the resulting ringed space X ×Y is the categorical product
X × Y . Take any morphisms pX : Z → X and pY : Z → Y which satisfy the required
commutativity requirements. We need to show that the induced set-theoretic map p :
Z → X × Y is a morphism. By Proposition 8.1, it suffices to check that for each U ⊂ X,
V ⊂ Y affine, the restriction p : ZU,V := p−1 (U × V ) → U × V is a morphism. Since
U × V is affine, by Proposition 8.2 it suffices to check that for any s ∈ Γ(U × V ),
p∗U,V (s) ∈ OZ (ZU,V ). This follows as in the proof of Proposition 9.3 because Γ(U × V )
is generated by f ⊗ 1 and 1 ⊗ g for f ∈ Γ(U ), g ∈ Γ(V ), and since p∗U,V (f × 1) = p∗X (f ),
p∗U,V (1 × g) = p∗Y (g) are elements in OZ (ZU,V ). □
Remark 9.2. In addition, one may show the following:
(1) If U × X is open, then U × Y will be an open subset of X × Y .
(2) If Z ⊂ X is a closed subvariety, then Z × Y will be a closed subvariety of X × Y .
(For this, it suffices to show that for any U ⊂ X and V ⊂ Y open affine,
(Z ∩ U ) × Y is closed in U × V .)
For a, b ∈ N define νa,b : Pa × Pb → P(a+1)(b+1)−1 by
(9.1) νa,b ([x0 : . . . : xa ], [y0 : . . . : yb ]) = [. . . : xi yj : . . .].
Note that this is a well defoned map, and an analogous proof as the one from Proposi-
tion 8.4 this is a morphism, called the Segre embedding. (Note that here Pa × Pb is
already given a product variety structure.)
Lemma 9.5. (a) The image of νa,b is given by the determinants of 2 × 2 minors in the
matrix:
Z0,0 Z0,1 . . . Z0,b
Z1,0 Z1,1 . . . Z1,b
. .. .
.. . . . . ..
Za,0 Za,1 . . . Za,b
where Zi,j is the coordinate corresponding to xi yj . (This is another example of a deter-
minantal variety.)
(b) Let Sa,b := νa,b (Pa × Pb ) ⊂ P(a+1)(b+1)−1 be the image of the Segre embedding. Then
νa,b : Pa × Pb → Sa,b is an isomorphism.
Proof. Exercise. □
Theorem 9.3. The product of two projective varieties is projective.
Proof. If X ⊂ Pa and Y ⊂ Pb are closed then so is X × Y ⊂ Pa × Pb (cf. Remark 9.2).
By Lemma 9.5, Pa × Pb is isomorphic to a projective variety, thus it is projective. □
9.3. Separated varieties and the Haussdorff axiom. The following is the analogue
of the Haussdorff condition in algebraic geometry.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 53
Definition 9.1. An algebraic variety X is called separated if for all varieties Z and
f
*
all morphisms Z 4 X the set {z ∈ Z : f (z) = g(z)} is a closed subset of Z.
g
Consider the diagonal map ∆ : X → X ×X given by x 7→ (x, x). This is the morphism
induced by the universality property, where p1 , p2 : X × X → X are the projections:
X
id
∆
# &
X ×X / X
p2
id
p1
X / pt
If X is separated, then by definition
∆(X) = {(a, a) ∈ X × X} = {(a, b) ∈ X × X : p1 (a, b) = p2 (a, b)}
is closed in X × X. The converse is also true:
Proposition 9.6. Let X be a variety. Then X is separated if and only if ∆(X) ⊂ X ×X
is closed.
Proof. We only need to show that if ∆(X) is closed, then X is separated. Let f, g : Z →
X be two morphisms. Then there is an induced morphism (f, g) : Z → X × X, and
{z ∈ Z : f (z) = g(z)} = (f × g)−1 (∆(X)).
The latter is a closed set by hypothesis. □
Example 9.7 (A non-separated variety: the line with 2 origins). The projective line P1
may be seen as the as gluing two copies U0 , U1 ≃ A1 along A1 \ {0} by the map x 7→ 1/x.
We have checked that P1 is an algebraic variety, and, for example, OP1 (P1 ) = k.
We may also glue along A1 \ {0} by x → x; this gives a different variety structure.
We call it X, and we may visualize it as ‘the line with 2 origins’. For example, Γ(X)
will contain non-constant global sections.
7 U0 o
≃
A1 >X
≃
' /
U1
The variety X is not separated; indeed, take i1 , i2 : A1 → X be the compositions. Then
{x ∈ A1 : i1 (x) = i2 (x)} = A1 \ {0},
54 LEONARDO CONSTANTIN MIHALCEA
2As
an aside, observe that Γf = Image((id, f ) : X → X × Y ). Further, the restriction of the first
projection from X × Y to X gives an isomorphism Γf → X.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 55
10. Dimension
The purpose of this section is to introduce the notion of dimension for algebraic vari-
eties. We will then utilize it to study the expected dimensions of systems of polynomial
equations, and dimension of generic fibers of morphisms.
Theorem 10.1. Let R be a finitely generated k-algebra which is an integral domain. Let
K be the fraction field of R. Then the Krull dimension of R is the transcendence degree
K/k.
Proof. See ADDREF. □
Example 10.4. The Krull dimension of k[x1 , . . . , xn ] equals to n. A chain of prime
ideals is given by 0 ⊊ ⟨x1 ⟩ ⊊ ⟨x1 , x2 ⟩ ⊊ . . . ⊊ ⟨x1 , . . . , xn ⟩.
Definition 10.2. Let X be an irreducible affine algebraic variety. Then the coordinate
ring Γ(X) is an integral domain, and the function field K(X) is defined to the fraction
field of Γ(X).
By Theorem 5.1 the function field of X coincides with the function field of any dis-
tinguished open set in X.
Proposition 10.5. Let X be an irreducible algebraic variety, and U ⊂ X a nonempty
open set. Then the following hold:
(1) If X is affine then
dim X = dim Γ(X) = trdegk K(X) < ∞.
(2) For an arbitrary (irreducible) X, dim X = dim U < ∞.
Proof. The equality dim X = dim Γ(X) follows from the Hilbert’s Nullstellensatz, which
gives an order reversing correspondence between closed irreducible subvarieties of X and
prime ideals in Γ(X). The second equality, and the finiteness conclusion, follow from
Theorem 10.1.
We prove the second part in several steps. First, assume that X is affine. Then we
can pick a distinguished open set Df ⊂ U ⊂ X. Since K(Df ) = K(X), it follows from
part (1) and Proposition 10.3 that
dim X = dim Df ≤ dim U ≤ dim X.
For arbitrary X, this also proves that any two open affine open sets U1 , U2 ⊂ X must
have the same dimension. Indeed, their intersection is non-empty, and one can find an
open set V ⊂ U1 ∩ U2 ; this satisfies:
dim U1 = dim V = dim U2 .
Let r := dim V for any ∅ ≠ V ⊂ X open, affine. Find an open affine V such that V ⊂ U .
Then
r = dim V ≤ dim U ≤ dim X,
and it suffices to show that dim X = r.
Take a chain of closed irreducible subsets in X: F0 ⊂ F1 ⊂ . . . ⊂ Fn = X, and take
V open affine such that V ∩ F0 ̸= ∅. Intersecting with V gives a chain F0 ∩ V ⊂ . . . ⊂
Fn ∩ V = V . Each Fi ∩ V is closed and irreducible in V (since Fi are irreducible), and
the sets Fi ∩ V are distinct and non-empty, because Fi = Fi ∩ V . This implies that
dim X ≤ dim V and we are done. □
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 57
Example 10.8. The hypothesis that X is irreducible is essential. For instance, take
X = V (xy) ⊂ A2 ; this is the union of the coordinate axes. Now take
f = x(x + y + 1) ∈ Γ(X) = k[x, y]/⟨xy⟩.
Then V (f ) = V (x) ∪ {(−1, 0)}; this is not pure-dimensional.
Corollary 10.9. Let X be an irreducible variety and let Z be an irreducible component
of V (f1 , . . . , fr ), with each fi ∈ OX (X). Then codim Z ≤ r.
Proof. Induction on r. The case r = 1 follows from Theorem 10.2. Let Z ′ be an
irreducible component of V (f1 , . . . , fr−1 ). By induction, codimZ ′ ≤ r −1. By hypothesis
Z is an irreducible component of Z ′ ∩ V (fr ) ⊂ V (f1 , . . . , fr ). If fr does not vanish
identically on Z ′ , then dim Z = dim Z ′ − 1, as claimed. □
Proposition 10.10. Let X be an affine irreducible variety and let Z ⊂ X be a closed
subvariety of codimension r. Then there exist f1 , . . . , fr ∈ Γ(X) such that Z is a com-
ponent of V (f1 , . . . , fr ).
Proof. See [Mum99, Cor. I.7.4]. □
A subvariety Z ⊂ X of codimension r which is given by r equations is called a
complete intersection. Being a complete intersection is a rather strong constraint.
For example, if X = Pn and Y = Vp (F1 , . . . , Fr has codimension r then by Bèzout
theorem the degree of Y must be
r = deg(Y ) = deg(F1 ) · . . . · deg(Fr ).
This degree may often be calculated by other cohomological methods.
Example 10.11. The degree of the twisted cubic C := νe (P1 ) ⊂ P3 equals to 3. Since C
is not included in a hyperplane, it follows that I(C) cannot be generated by 2 equations.
Example 10.12. Consider the variety of matrices:
a b c
M1 := {A = : rk(A) ≤ 1} ⊂ A5 .
d e f
One may projectivize M1 , and P(M1 ) is the image of P2 × P1 ,→ P5 under the Segre
embedding. In particular, codim M1 = 2. One may show that deg(P(M1 )) = 3, which
implies that Ip (P(M1 )) cannot be generated by 2 equations. Then I(M1 ) (the cone over
P(M1 )) cannot be generated bby 2 equations.
These results show that a codimension 1 irreducible variety may be realized as a
component of the zero locus of a single equation. But does it equal to V (f ) ? This is
not always possible; the next result gives a sufficient condition for this to hold.
Proposition 10.13. Let X be an affine irreducible variety such that Γ(X) is an UFD.
let Z ⊂ X be closed of pure codimension 1. Then there exists f ∈ Γ(X) such that
Z = V (f ).
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 59
Proof. We start by observing that since R := Γ(X) is an UFD, any minimal nonzero
prime must be principal. (Proof: Let p ⊂ R be a minimal prime, and let f ∈ p. Since
R is an UFD, we may decompose f as a product of prime factors. The fact that p is
prime implies that it must contain one of the prime factors of f , say f ′ . But the ideal
⟨f ′ ⟩ is prime and ⟨f ′ ⟩ ⊂ p, thus ⟨f ′ ⟩ = p.)
Now the components of Z correspond to minimal prime ideals in R, generated by
some prime elements f1 , . . . , fr in R. Then Z = V (f1 · . . . · fr ) and we are done. □
10.4. Projective versions. All these results have projective versions, involving projec-
tive coordinate rings and homogeneous polynomials. For instance, the following are the
analogues of Theorem 10.2 and Proposition 10.13.
Theorem 10.4. Let X be an irreducible projective variety with projective coordinate ring
Γp (X) and F ∈ k[x0 , . . . , xn ] be a homogeneous, non-constant polynomial such that F
does not vanish identically on X. Then X ∩ Vp (F ) is nonempty and of pure codimension
1, unless dim X = 0.
Sketch of the proof. Except for the non-emptiness, the claim follows from Theorem 10.2.
To prove non-emptiness, assume that dim X ≥ 1 and let X ∗ ⊂ An+1 be the cone over
X. By definition, the cone is the affine variety given by the homogeneous ideal Ip (X).
One may prove that dim X ∗ = dim X + 1. (Exercise!) Then 0 ∈ X ∗ ∩ V (F ), and by
Theorem 10.2 dim X ∗ ∩ V (F ) ≥ 1. In particular X ∗ ∩ V (F ) contains other points except
for the origin, and since it is k ∗ -stable it must contain a full line through the origin. We
deduce that X ∩ Vp (F ) is non-empty. □
Proposition 10.14. Every closed subset of Pn of pure codimension 1 is a hypersurface,
i.e. it is given by the vanishing of a single homogeneous polynomial.
60 LEONARDO CONSTANTIN MIHALCEA
12.1. Definitions.
Definition 12.1. Let (X, OX ) be a ringed space. An OX -module is a sheaf F such that
for any U ⊂ X open, F(U ) is a OX (U )-module, and the restriction homomorphisms are
maps of modules, i.e. for any V ⊂ U ,
F(U )
res / F(V )
O O
OX (U )
res / OX (V )
Here OX (V ) and F(V ) become OX (U )-modules via the restriction maps.
A homomorphism of OX -modules φ : F → G is given by the data of OX (U )-
module homomorphisms φU : F(U ) → G(U ) for any open set U ⊂ X, compatible with
the restriction maps.
Definition 12.2. Let (X, OX ) be a ringed space and φ : F → G a homomorphism of
OX -modules.
• The kernel of φ, denoted by ker(φ), is the sheaf given by the assignment
U 7→ ker(φU ).
• The image of φ, denoted by Im(φ), is the sheafification of the presheaf
U 7→ Im(φU ).
• A sequence of sheaves and sheaf homomorphisms
0 / F
u / G
v / H / 0
is (short) exact if ker(u) = 0 (the zero sheaf ), Im(v) = H and Im(u) = ker(v).
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 63
Example 12.1 (The exponential sequence). The following sequence of sheaves of abelian
groups is short exact:
/ / e2πiz / /
0 Z OC OC∗ 0
Here Z be the sheaf given by locally constant Z-valued functions, OC is the (additive) sheaf
of holomorphic functions on C, and OC∗ is the (multiplicative) sheaf of non-vanishing
holomorphic functions. All these are sheaves of abelian groups.
Consider the non-vanishing holomorphic function f (z) = z on C∗ = C − 0. Using that
the complex logarithm is defined on any simply connected open subset of C∗ , we see that
z is in the image of the exponential map over any such domain. However, z is not in the
image of any single holomorphic function defined on the whole C∗ . (The fundamental
group of C∗ is Z.) This shows that sheafification is needed in the definition of the image.
Definition 12.3. Let f : X → Y be a morphism of algebraic varieties, and F a sheaf
on X. The direct image of F is the sheaf on Y , denoted by f∗ F, and defined by
V 7→ F(f −1 (V )).
Example 12.2. Let i : Y ,→ X be a closed subvariety and let FY be the sheaf of k-
valued functions on Y . There is a morphism of OX -modules r : OX → FY defined by
sending s ∈ OX (U ) to its restriction on U ∩ Y . By definition, the image of this sheaf
Im(r) = OY , regarded as a OX -module. (Again recall that the definition of OY required
the use of sheafification.)
The kernel of r is an ideal sheaf i.e. a sheaf IY on X where IY (U ) is the ideal of
sections on U ⊂ X vanishing of Y ∩ U . (In particular this is an OX -module.) We have
another short exact sequence
(12.1) 0 / IY / OX / i∗ OY ≃ OY / 0
Remark 12.1. As we shall see later, the sequence from Equation (12.1) may be also
be realized by gluing “local data”. if one assumes that Y, X are affine and Y ⊂ X is
closed, this is the sequence associated with the short exact sequence of Γ(X)-modules
0 → IY → Γ(X) → Γ(Y ) → 0 where IY is an ideal of Y in X.
Definition 12.4. Let F, G be two OX -modules. The tensor product F ⊗OX G is the
sheafification of the presheaf
U 7→ F(U ) ⊗OX (U ) G(U ).
12.2. Quasi-coherent sheaves on affine varieties. Let X ⊂ An be an affine algebraic
variety and let R := Γ(X) = OX (X) be its coordinate ring. For any sheaf F of OX -
modules there is an associated R-module M := F(X); we will also use the notation
Γ(F) = F(X). In other words, taking global sections gives a map:
(12.2) Γ : ( sheaves of OX -modules ) → (Γ(X) − modules)
64 LEONARDO CONSTANTIN MIHALCEA
The purpose of this section is to perform the inverse construction: associate a OX -module
for any R-module. Recall that for an R-module M and f ∈ R, the localization
m
Mf = M ⊗R Rf = { s : m ∈ M, s ∈ Z}/ ≃
f
m1 m2
The equivalence is given by f s1
= f s2
if and if only if there exists f s such that f s (f s2 m1 −
f s1 m2 ) = 0 in M .
As usual, the fact M f defines uniquely a sheaf a OX follows because the distinguished
open sets Df form a basis of open sets for X.
√
Example 12.3. If M = R, then M f = OX . More generally, if I = I ⊂ R is an ideal,
then Ie = IY , the ideal sheaf of Y = V (I).
is exact, where u′ , v ′ are the induced maps. To check exactness, it suffices to show
exactness when restricted to distinguished open sets. For this, it suffices to show that
the localized sequence
uf vf
0 / Mf / Nf / Pf / 0
is exact.3 Clearly vf is surjective, and vf ◦ uf = 0, therefore Im(uf ) ⊂ ker(vf ). We need
to prove the converse inclusion, and also that uf is injective.
Let fns ∈ ker(vf ). Then v(n)
fs
= 0 in Pf . Then there exists a nonnegative integer k such
that f v(n) = v(f n) = 0. It follows that f k m ∈ ker(v) = Im(u), so there exists m ∈ M
k k
e = OX .
Example 12.7. Of course, R
Remark 12.5. The sheaf M f is quasi-coherent; furthermore, if M is finitely generated,
it is coherent. Indeed, we may cover X by distinguished open sets Df+ . Then observe
that Df+ is affine and that M f| + = M
(Df )
g +
(f ) . (Note that Γ(Df ) = (Γp (X))(f ) , and that
Proof. Choose a distinguished open set Df+ ⊂ X, where f is homogeneous of degree > 0.
It suffices to show that φf : M(f ) → N(f ) is surjective. To do this, observe that any
n
element fyr ∈ N(f ) may be written as fyf n
n+r and deg(f y) may be made arbitrarily large.
The hypothesis implies that such elements are in the image of φ; the claim follows from
this. □
Example 12.10. As in the affine case, if Y ⊂ X is closed, the sequence 0 → I →
Γp (X) → Γp (Y ) = Γp (X)/I → 0 gives the natural sequence from Equation (12.1) asso-
ciated to Y .
We now come to the main difference from the affine case: the ability to re-define the
grading on a graded module.
4
L
Definition 12.9. Let R be a graded ring and let M = n∈Z be a graded R-module.
Fix d ∈ Z. The module M (d) is the graded module M , but with the shifted grading
M (d)n := Md+n .
Definition 12.10. Let X ⊂ Pn be a projective algebraic variety and let R := Γp (X) be
] If F is any sheaf of
the projective coordinate ring. Define the sheaf OX (d) := R(d).
OX -modules, define
F(d) := F ⊗OX OX (d).
The sheaves OX (d) are coherent (in fact, locally free of rank 1), and they depend
fundamentally on the grading on Γp (X). Sometimes these are called the Serre twisting
sheaves. Observe also that if M is a graded R-module then by commutativity of tensor
products
]
M f ⊗O OX (d) = M
(d) = M f(d).
X
4We are grading over Z, but this does not make a difference.
68 LEONARDO CONSTANTIN MIHALCEA
This may be given a graded R-module structure by: s ∈ Rd = Γ(OPn (d)) and t ∈
Γ∗ (F(d′ )) then one may multiply: s.t ∈ Γ(F(d + d′ )). A geometric fact which currently
is outside our reach is the following (cf. [Hartshorne, Ch.5 and Ex. II.5.9]):
Theorem 12.2. Let M f ≃ F be a quasi-coherent module. Then the following hold:
(a) There exists an isomorphism of OX -modules Γ^
∗ (F) ≃ F
(b) There exists a graded R-module homomorphism M → Γ∗ (F) such that for d ≫ 0,
Md ≃ Γ(F(d)) as k-vector spaces.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 69
(13.1) 0 / R(−d)
·F / R / R/⟨F ⟩ / 0
If in addition F has no multiple factors, let X = Vp (F ). Then we have a short exact
sequence of coherent OPn -modules:
Therefore PS (x) = 2. We may also consider the ideal I = ⟨x20 ⟩ (a double point). The
Hilbert polynomial is the same. In particular the degree of 2 distinct points equals the
degree of one double point.
Example 13.5. While proving Corollary 13.1 we will also show that if dim X = 0 (i.e.
if X is consists of finitely many points) then hX (z) ≡ dimk Γ(X). Here we utilize that
any such X is affine, and one may show that if dim X = 0 then its coordinate ring Γ(X)
is a finite dimensional vector space (i.e. dimk Γ(X) < ∞).
Next we calculate the Hilbert polynomial of a hypersurface.
Proposition 13.6. Let F ∈ k[x0 , . . . , xn ] be a homogeneous polynomial and let M :=
k[x0 , . . . , xn ]/⟨F ⟩.
z+n z − deg F + n
PM (z) = − .
n n
In particular, if F is reduced (i.e., multiplicity free), then deg Vp (F ) = deg F .
Proof. Taking degree d components in Equation (13.1) we obtain the short exact se-
quence of vector spaces
0 /
(R)d−deg F
·F / Rd / (R/⟨F ⟩)d /0
d+n
− d−deg(F )+n
Then hM (d) = n n
. The result follows after performing the algebraic
calculations. □
Example 13.7 (A conic). Consider the conic C := Vp (y 2 − xz) ⊂ P2 . By Proposi-
tion 13.6,
(z + 2)(z + 1) z(z − 1)
pC (z) = − = 2z + 1.
2 2
Alternatively, one may calculate directly from the definition that
1 d = 0
hC (d) = 3 d = 1
5 d = 3
We also need the following generalization of Lemma 13.12, which is currently outside
our technology.
Theorem 13.4 (Bertini’s theorem). Let X ⊂ Pn be a projective variety. Then there
exists an open set U ⊂ (Pn )∗ such that for all H ∈ U , H ∩ X has codimension 1 and
Ip (X) + H is radical.
Proof of Theorem 13.3. Choose H as in Bertini’s theorem. Then the claim follows by
induction from Bézout’s theorem, since deg(X ∩ H) = deg(X). □
13.4. Degree of the union of two projective varieties. We now turn to the degree
of a union of projective subvarieties.
Lemma 13.13. Let I1 , I2 be two homogeneous ideals in R := k[x0 , . . . , xn ]. Then there
exists a short exact sequence of graded R-modules
(f,f ) (f,−g)
0 / R/(I1 ∩ I2 ) / R/I1 ⊕ R/I2 / R/(I1 + I2 ) / 0
where the first morphism is the diagonal, and the second sends (f, g) ∈ R/I1 ⊕ R/I2 to
f − g ∈ R/(I1 + I2 ).
Proof. Clear from definitions. □
74 LEONARDO CONSTANTIN MIHALCEA
END OF SEMESTER 1
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 75
15. Schemes
Even if one wishes to work with algebraic varieties, the appearance of schemes is
unavoidable. For example:
• the intersection of a conic and a line tangent to it is a double point.
• Consider a degree 2 morphism f : P1 → P1 , e.g. [x : y] 7→ [x2 : y 2 ]. Then a
general fibre has 2 points, but there will be fibres which are double points.
To understand better what we are generalizing, recall that an algebraic variety is (by
definition) covered by affine algebraic varieties, each of which being inside of some affine
space An . Recall the correspondence:
{Affine algebraic varieties X ⊂ An } ↔ { radical ideals I ⊂ k[x1 , . . . , xn ]}
We may generalize this correspondence in several directions:
(1) We may allow the ideals I to be non-radical, leading to non-reduced schemes.
(2) We do not have to consider ideals in a polynomial ring, leading to schemes which
are non-necessarily embedded in an affine space.
(3) We may remove the requirement that I (or the polynomial ring) is finitely gen-
erated. This leads to the theory of schemes of infinite type. We will not work in
this generality.
15.1. Spectrum of a ring. In this section we largely follow Mumford’s treatment from
[Mum99].
Definition 15.1. Let R be a ring. The spectrum of R is
Spec(R) = {p ⊂ R : p is a prime ideal }
Example 15.1. If R = K is a field then Spec(K) = (0). More generally, (0) ∈ Spec(R)
whenever R is an integral domain.
We make Spec(R) into a topological space by defining its closed sets to be of the form
V (I) := {p ∈ Spec(R) : p ⊃ I}
where I ⊂ R is an ideal. Note that V (I) may also be defined by requiring that φf (p) = 0
for all f ∈ I. This topology is again called the Zariski topology.
Proposition 15.2. The closed sets form a topology on Spec(R) such that the corre-
sponding topological space is paracompact. Furthermore, the following hold for any ideals
I, J, Is ⊂ R:
S T
• V ( s Is ) = s V (Is );
• V (I ∩ J) = V (I) ∪ V (J)
Proof. Homework! □
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 77
□
Proposition 15.6. Let p ∈ Spec(R). Then the following hold:
(a) V (p) is irreducible and p is its unique generic point.
(b) Every irreducible closed subset of Spec(R) is of the form V (p) for some p ∈
Spec(R).
Proof. Write V (p) = Z1 ∪ Z2 some closed sets Z1 , Z2 . If p ∈ Zi then {p} ⊂ Zi = Zi , thus
equality must occur. If there is another generic point q then since {p} = {q} it follows
that p ⊂ q and q ⊂ p, thus p = q. This proves (a).
Now take Z = V (I) ⊂ Spec(R)
√ closed and irreducible, where I ⊂ R is an ideal.
By Lemma 15.5 V (I) = V ( I), therefore we may assume that I is radical. We will
78 LEONARDO CONSTANTIN MIHALCEA
as partial ordered sets, and we identify the two orderings. The limits over subsets of
elements in R will be taken with respect to these limits.
15.2. The structure sheaf of Spec(R). In this section we will define the structure
sheaf OSpec(R) . We will be associating Df 7→ Rf for each f ∈ R, and we need to check a
number of compatibilities.
Lemma 15.8. Let f, g, h ∈ R.
(a) If Dg ⊂ Df then there is a natural (restriction) map Rf → Rg . Furthermore, if
Dh ⊂ Dg ⊂ Df , then there is a commutative diagram given by restriction maps
Rf / Rg
Rh
(b) The assignment Df 7→ Rf is well defined, i.e., Rf ≃ Rg if Df = Dg .
(c) Let p ∈ Spec(R). Then the localization Rp is the direct limit
Rp = lim Rf = lim Rf .
−→ −→
f ∈R\p p∈Df
p
Proof. If Dg ⊂ Df then g ∈ ⟨f ⟩, i.e. g n = hf for some positive integer n and h ∈ R.
Then the map Rf → Rg is given by
a ah
7→ n .
f g
If Df = Dg then we get maps Rf → Rg and Rg → Rf which are inverse to each other.
(Check!) We leave part (c) as an exercise. □
We will see that the localization Rp is the stalk at p of the structure sheaf OSpecR and
part (c) is the analogue of the fact that germs of functions a point p have representatives
in a neighborhood of p.
We are ready to define the structure of Spec(R). We Q utilize the same idea as in
Lemma 5.5. The field K from this lemma is replaced by p∈Spec(R) Rp .
Definition 15.3. The structure sheaf OSpecR is Q defined as follows. Let U ⊂ Spec(R) be
open. Then OSpecR (U ) is the set of tuples (sp ) ∈ p∈U Rp such that for each q ∈ U there
exists a distinguished open set Dfi ⊂ U with the property that for all p ∈ Dfi , and all
(sp )p∈Dfi , there exists si ∈ Rfi such that sp is obtained by the restriction Rfi → Rq .
If V ⊂ U are open sets, then the restriction map ρU V : OSpecR (U ) → OSpecR (V ) is the
natural projection.
It is easy to check that each OSpecR (U ) is a ring.
Theorem 15.1. (a) OSpecR is a sheaf.
(b) OSpecR (Df ) = Rf .
(c) The stalk (OSpecR )p = Rp .
80 LEONARDO CONSTANTIN MIHALCEA
Rp
The image of g in Rp factors through Rfi , thus it is equal to 0 ∈ Rp . Then there exists
c ∈ R \ p such that cb = 0, i.e. c ∈ A. But then c ∈ A ⊂ p by assumption, which is a
contradiction. □
S
Lemma 15.10. Assume Df = i Dfi for some fi , and assume we have elements gi ∈ Rfi
such that the restriction to Rfi fj coincide. Then there exists g ∈ Rf such that the
restriction to each Rfi is gi .
Proof. Similar to proof of the gluing property in Equation (5.3) above; see [Mum99, p.
70, Lemma 2]. □
We are now ready to prove Theorem 15.1.
Proof of Theorem 15.1. To prove (a), by Lemma 5.5, it suffices to show the restriction
and gluing property for the basis of open sets. It is clear that the restriction maps ρU V
satisfy the required compatibilites. The gluing property follows from Lemma 15.10.
To prove (b), there is a natural map
Y Y
Rf → OSpecR (Df ) ⊂ Rp = Rp
p∈Df f ∈p
/
□
Remark 15.2. Note that the stalks of OSpecR are Rp , and these are local rings. Then
(SpecR, OSpecR ) is called a locally ringed space.
Remark 15.3. The fact that we have stalks at non-closed points gives an additional
structure. If p1 ⊂ p2 are two prime ideals in R, then p2 ∈ {p1 }, therefore any neighbor-
hood U of p2 contains p1 . Then
Rp2 = (OSpecR )p2 = lim OSpecR (U ) → lim OSpecR (V ) = (OSpecR )p1 = Rp1 .
−→ −→
p2 ∈U p1 ∈V
%
Spec(OX (X))
Remark 15.4. In some situations OX (X) is very small. E.g., if X is a projective
k-variety, then OX (X) = k (This will be shown later.) Therefore it follows that any
morphism from a projective scheme to an affine scheme must be constant!
Corollary 15.20. Spec(Z) is a final object in the category of preschemes, i.e. for any
prescheme X there is a unique morphism X → Spec(Z).
Corollary 15.21. Let X be a prescheme and let x ∈ X be an arbitrary point with
residue field k(x). Then there exists a canonical morphism ix : Spec(k(x)) → X sending
0 7→ x.
Proof. Take U = Spec(R) to be an open affine neighborhood of x. There is a natural
inclusion U ,→ X, and using this we may assume X = Spec(R) is affine. Let px ⊂ R be
86 LEONARDO CONSTANTIN MIHALCEA
the prime ideal corresponding to x. Then the morphism ix is given by the natural map
R → Rpx → k(x). □
To relate preschemes to the algebraic varieties from the previous semester we give the
more general definition:
Definition 15.6. Let R be a ring. Then a prescheme X over R is a morphism πX :
X → Spec(R). A morphism f : X → Y is a morphism of preschemes over R if there is
a commutative diagram
f
X / Y
πX πY
$ {
Spec(R)
There is a category of preschemes and morphisms over R. There is an analogue of
Proposition 15.16 which establishes an equivalence between the category of affine R-
preschemes and morphisms to ring homomorphisms of R-algebras. There is a further
generalization of Theorem 15.2 to the context of R-preschemes. The most important
situation is when R = k and the sheaves of sections satisfy a finite generation condition.
In this case, the schemes over k correspond to the algebraic varieties we studied in the
prior semester.
We also record the following useful lemma:
Lemma 15.22. Let X be a prescheme and Z ⊂ X a non-empty closed irreducible subset.
Then there exists a unique point z ∈ Z such that {z} = Z. (This is called the generic
point of Z.)
Proof. Let U be any affine open scheme which intersects Z. Then Z ∩ U must be dense
in Z, and any dense point in Z must be in Z ∩ U . Observe that the latter is closed
and irreducible in the affine scheme U . By Proposition 15.6, there exists a unique point
dense in Z ∩ U (the generic point), and this finishes the proof. □
Definition 15.7. Let R be a ring. The n-dimensional affine space over R is the affine
scheme
AnR = Spec(R[x1 , . . . , xn ]).
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 87
The fibre f −1 (y) is the prescheme k(y) ×Y X defined by the fibre diagram below:
f −1 (y) ≃ k(y) ×Y X / X
f
iy
k(y) / Y
This is naturally a prescheme over the residue field of y.
Example 16.1. Let f : A1C → A1C defined by f (x) = x2 . Let a ∈ A1C be a closed point,
corresponding to the maximal ideal ma = ⟨t − a⟩ ⊂ C[t]. The residue field is k(a) =
C[x]/ma ≃ C. Note that f is induced by the C-algebra homomorphism f ∗ : C[t] → C[x]
defined by t 7→ x2 .
Then
f −1 (a) ≃ Spec(C[x] ⊗C[t] C[t]/ma ) = Spec(C[x]/⟨x2 − a⟩)
The prescheme f −1 (a) consists of two reduced points if a ̸= 0; if a = 0 then C[x]/⟨x2 ⟩ is
non-reduced, and the fibre f −1 (0) is a single non-reduced point.
Consider now the generic point 0 ∈ Spec(C[t]). The residue field is C[t](0) = C(t) (the
fraction field of C[t]). Then
f −1 (⟨0⟩) = Spec(C[x] ⊗C[t] C(t)) = Spec(C[x]C[x2 ]\0 ).
(This uses that for a multiplicative set S in a ring R and an R-module M , we have
M ⊗R (S −1 R) ≃ S −1 M as S −1 R-modules; see [AM69, Prop. 3.5].) Note that C[x]C[x2 ]\0
is a C(t)-vector space of dimension 2, with basis 1, x.
16.1.2. Base change. More generally, consider the fibre diagram
Y ′ ×Y X / X
f′ f
Y′ / Y
This is called a base change, or base extension (from Y to Y ′ ). It is the natural
algebraic analogue of the pull-back of a vector bundle from topology. It also leads to
the notion of extending the field of a definition. For example, for any ring R,
AnR = AnZ ×Spec(Z) Spec(R).
This follows because R ⊗Z Z[x1 , . . . , xn ] = R[x1 , . . . , xn ]. In other words we’re base
changing from coefficients in Z to coefficients in an arbitrary ring.
One important property of the base extension is that f, f ′ have the same fibres, and
have the same geometric properties (preserves flatness, properness, etc.)
Remark 16.1. The fibre product X ×Z Y depends on the base prescheme Z, and in
general there is no injective map to the set theoretic product. For instance, Spec(C ×C
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 89
is a disjoint union of two points. (The two points are given by the prime ideals generated
by (1, 0) and (0, 1); note also the idempotent 1⊗1+i⊗i
2
in the tensor product, showing
that Spec(C ⊗R C) must be disconnected.) Thus the base extension may not preserve
connectedness.
However, the set theoretic and fibre product will coincide for the reduced, irreducible
preschemes of finite type over algebraically closed fields. (See §16.2 below or [Mum99,
Ch. II, §3, Prop. 5].)
16.1.3. The diagonal map and intersections. Assume we are given a morphism f : X →
Y . The diagonal ∆ : X → X ×Y X is the unique morphism given by the universality
property of fibre products:
(16.1) X
id
∆
$ &/
X ×Y X X
id
f
f
X / Y
X1 ∩ X2 / X1 ×Spec(k) X2
i1 ×i2
X
∆ / X ×X
Equivalently, one may also define it from the fibre diagram
X1 ∩ X2 / X1
i1
i2
X2 / X
Of course, the intersection X1 ∩ X2 may not have good properties (e.g. expected
dimension, singularities etc.)
90 LEONARDO CONSTANTIN MIHALCEA
d
For (d), it suffices to show that I ⊂ I d for d ≫ 0. Since I is finitely generated, take
P1 , . . . , Ps be a set of homogeneous generators. Take D1 large enough such that xdi Pj ∈ I
for d ≥ D1 and all 0 ≤ i ≤ n, 1 ≤ j ≤ s. Take D2 the maximum degrees of Pj . Now let
d
d ≥ D2 + (n + 1)D1 and let P be any element in I , and write
X
P = aj P j
for some homogeneous aj . Then deg aj ≥ d − D2 = (n + 1)D1 . This shows that each
term of aj contains a factor which is a multiple of xD
i
1
for some i. Then aj Pj ∈ I and
we are done. □
Obviously Proj(k[x0 , . . . , xn ]/I) = Vp (I) is closed subset of Pnk ; this is called a pro-
jective subscheme. (We will see in the next section the precise meaning of the word
‘subscheme’; roughly, this means that we have a locally ringed space with an induced
topology, and surjections at the level of stalks, induced from stalks of the structure sheaf
of the ambient space.)
Corollary 17.5. There is a one-to-one correspondence between projective subschemes
of Pn and saturated ideals in k[x0 , . . . , xn ]. The correspondence associates the saturated
homogeneous ideal I to Proj(k[x, . . . , xn ]/I).
94 LEONARDO CONSTANTIN MIHALCEA
f′ ~ f
X
′
where i : Y → Y is an isomorphism.
Example 18.4. (Closed subschemes of affine schemes.) Let X = Spec(R) and I ⊂ R be
an ideal. The projection i∗ : R → R/I induces an injective morphism i : Spec(R/I) →
Spec(R). This is a closed sub(pre)scheme, because the map i∗ induces surjective maps
after localization.
In particular, observe that for any closed subset V (J) ⊂ Spec(R) there are many
closed subscheme structures corresponding to ideals I such that V (I) = V (J). √ The
smallest such √ subscheme corresponds to the largest such ideal, which is necessarily J.
This gives V ( J) the reduced (induced) scheme structure.
Definition 18.5 (Dimension). If X is a prescheme, dim X is the dimension when re-
garded as a (locally) ringed space. In other words, if X is irreducible, then dim X = n
if there is a maximal chain of irreducible closed sets
∅ ⊊ X0 ⊊ X1 ⊊ . . . ⊊ Xn = X.
If X is not irreducible, then dim X is the supremum of dimensions of the irreducible
components of X.
Example 18.5. If Spec(R) is irreducible, then dim Spec(R) = dim R (the Krull dimen-
sion of R).
18.2. Separated morphisms; schemes.
Definition 18.6. Let f : X → Y be a morphism of preschemes. The diagonal mor-
phism is the unique morphism ∆f : X → X ×Y X induced by universality property of
the fibre product X ×Y X; see Eq. (16.1).
The morphism f is separated if ∆f is a closed embedding. In this case X is said to
be separated over Y . A prescheme is called a scheme if it is separated over Spec(Z).
Remark 18.2. Most often we will work with schemes separated over a field k. It follows
from Theorem 18.2 below that if X is separated, then X is separated over k.
Example 18.6. A prescheme which is not separated is the affine line with two origins;
see §9.3 above.
96 LEONARDO CONSTANTIN MIHALCEA
pr2
(i,f ) &
Pnk ×Spec(Z) Y
Definition 18.10. A scheme X over a field k is called complete if the morphism
X → Spec(k) is proper. The scheme is called projective if X → Spec(k) is a projective
morphism; equivalently, there is a closed embedding X ,→ Pnk .
A key property of projective morphisms is that they are proper:
Theorem 18.4. A projective morphism if Noetherian schemes is proper. A quasi-
projective morphism of Noetherian schemes is of finite type and separated.
Remark 18.3. A consequence of the theorem is that any projective morphism is closed.
This is an instance of the elimination theory: assume we have a system of equations
f1 (x0 , . . . , xn ; y1 , . . . , yp ) = 0
f2 (x0 , . . . , xn ; y1 , . . . , yp ) = 0
..
.
f (x , . . . , x ; y , . . . , y ) = 0
s 0 n 1 p
where fi ’s are polynomials in x’s and y ′ over some field k, and such that they are ho-
mogeneous in variables x. Then there exist polynomials g1 (y1 , . . . , yp ), . . . gℓ (y1 , . . . , yp )
such that
g1 (y1 , . . . , yp ) = . . . = gℓ (y1 , . . . , yp ) = 0.
In other words, one may ‘eliminate’ the variables x from the system above. This follows
from the fact that pr2 : Pn ×Spec(k) Ap → Ap is projective, thus it is proper, and thus the
image of any closed subset under pr2 is closed.
If k is algebraically closed, it turns out that (quasi-)projective algebraic varieties from
the first semester correspond to (quasi-)projective schemes over k, in the sense above.
This is proved in [Har77, Prop. II.4.10]. From now on, a variety will mean a
separated, integral scheme of finite type over an algebraically closed field k.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 99
Theorem 19.2. Let K be a finitely generated field over some field k. Then dimK ΩK/k ≥
trdegk K and equality holds if and only if K is separably generated over k.
Theorem 19.3. Let (B, m) be a local ring which contains a field k isomorphic to the
residue field B/m. Consider the map
δ ′ : m/m2 → ΩB/k ⊗B k; m 7→ dm ⊗ 1.
Then δ ′ is a vector space isomorphism.
Let k be algebraically closed and perfect, i.e. either char(k) = 0 or if char(k) = p
then x 7→ xp is a field automorphism. A Noetherian local ring (B, m) is a regular local
ring if the minimum number of generators of m equals the Krull dimension of B.
Theorem 19.4. Let (B, m) be a local ring which contains a perfect field k isomorphic to
the residue field B/m. Assume further that B is the localization of a finitely generated
k-algebra. Then ΩB/k is a free B-module of rank dim B if and only if B is a regular local
ring.
102 LEONARDO CONSTANTIN MIHALCEA
Recall that varieties are separated integral schemes of finite type over an integrally
closed field k.
Definition 19.4. A variety X is called non-singular at x if the local ring OX,x is a
regular local ring. It is non-singular if X is non-singular at every point.
One may show that being non-singular is an open condition, i.e. if X is non-singular
at x then there exists an open set U containing x such that U is non-singular.
The main theorem of this section is the following:
Theorem 19.5. Let X be a variety. Then X is non-singular if and only if the sheaf of
regular differential ΩX/k is a locally free sheaf of rank n = dim X.
Proof. See [Har77, Thm. II.8.15]. □
Definition 19.5. Let X be a non-singular variety of dimension n. The tangent sheaf
is defined as the dual TX := Ω∗X/k . The canonical sheaf ωX is defined as ωX := ∧n ΩX .
Each is a locally free sheaf, of rank n, respectively, rank 1.
The geometric genus of X is defined as pg := dimk ωX (X).
Key tools to understand the sheaves of differentials are the following two theorems.
(We prove the second one.)
Theorem 19.6. Let X be a non-singular variety, and let Y ⊂ X a closed subvariety
defined by the sheaf of ideals I. (I.e., Y → X is a closed embedding.) Then Y is
non-singular if and only if:
(a) ΩX/Y is locally free, and,
(b) There is a short exact sequence of OY -modules:
(19.3) 0 → I/I 2 → (ΩX/k )|Y → ΩY /k → 0.
Here the first map is induced by the natural map I/I 2 ≃ ΩX/Y → ΩX/k , and the second
is the restriction. (Both maps may be defined locally.)
If this is the case then I is locally generated by c := codim(Y, X) elements, and I/I 2
is locally free sheaf of rank c.
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 103
We need to show that M f ≃ ΩPn /k . For that, we define local isomorphisms, and we need to
check that these glue to a global isomorphism. Cover Pn by the standard affine open sets
Ui = Spec[ xx0i , . . . , xxni ] ≃ An for 0 ≤ i ≤ n. Observe that the induced homomorphism of
free Sxi -modules Exi → Sxi is surjective, with kernel the free module of rank n generated
x
by ej − xji ei . Then the restriction of M f|U is a free module of rank n generated by sections
i
ej xj
− 2 ei
xi x i
(we need the extra factor of xi in the denominator to get elements of degree 0; see §12.3
above). Recall from Example 19.5 that ΩAn /k is locally free of rank n with basis given
by sections dxj for 1 ≤ j ≤ n. Then define
xj ej xj
ψi : ΩUi /k → M
f|U
i
d( )= − 2 ei .
xi xi xi
Since both are free modules of the same rank, this is an isomorphism. We need to prove
these isomorphisms glue over the intersections Ui ∩ Uj , i.e.,
xk xk
ψj (d( )) = ψi (d( )).
xi xi
xk xk xj
= · .
xi xj xi
Then
xk x k xj
ψj (d( )) = ψj (d( · ))
xi xj xi
xk xj xj x k
= ψj ( · d( ) + d( ))
xj xi xi xj
xk xj xj xk
= · ψj (d( )) + ψj (d( )).
xj xi xi xj
xi xj
Observe also that xj
· xi
= 1, therefore on Ui ∩ Uj ,
xi xj xj xi xj x2j xi
d( ) + d( ) = 0 =⇒ d( ) = − 2 d( ).
xj xi xi xj xi xi xj
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 105
Then
xk xk xj xj xk
ψj (d( )) = · ψj (d( )) + ψj (d( ))
xi xj xi xi xj
2
xj xk xi xj xk
= − 2 · ψj (d( )) + ψj (d( ))
xi xj xj xi xj
xj xk ei xi xj ek xk
= − 2 ( − 2 ej ) + ( − 2 ej )
xi xj xj xi xj xj
xk ek
= − 2 ei +
xi xi
xk
= ψi (d( )).
xi
One deduces that the morphisms ψi , ψj coincide on a set of module generators of ΩPn /k
over Ui ∩ Uj . This finishes the proof. □
Lemma 19.8. Let I ⊂ R be an ideal in a ring R. Then there is a canonical isomorphism
of R/I-modules
(19.4) I/I 2 → I ⊗R R/I; a + I 2 7→ a ⊗ 1.
Proof. Consider the short exact sequence of R-modules 0 → I → R → R/I → 0. By
[AM69, Ex. 2.2], tensoring with the R-module I preserves exactness, giving
0 → I 2 ≃ I ⊗R I → I ≃ I ⊗R R → I ⊗R R/I → 0.
The statement follows from the first isomorphism theorem applied to the second map.
□
This isomorphism is key to study the important special case when X = Pn and
Y = Vp (F ) ⊂ Pn is a hypersurface given by a homogeneous polynomial F of degree d.
Denote by i : Y ,→ Pn the inclusion. The ideal sheaf Ie = ⟨F
g⟩ induces the standard exact
sequence
(19.5) 0 → I → OPn → i∗ OY → 0.
On the other side, let S := k[x0 , . . . xn ] and consider the short exact sequence of graded
S-modules
0 / S(−d)
·F / S / S/⟨F ⟩ / 0
giving the sequence of OPn -modules
(19.6) 0 → OPn (−d) → OPn → i∗ OY → 0.
Proposition 19.9. (a) There is an isomorphism of locally free OPn -modules
I ≃ OPn (−d).
106 LEONARDO CONSTANTIN MIHALCEA
Therefore a line and a conic have Euler characteristic 2, a cubic has Euler characteristic
0 etc. This is (of course!) consistent with the ‘real picture’ of these curves: complex lines
and conics are homeomorphic to a (real!) sphere S 2 ; an elliptic curve is homeomorphic
to a torus S 1 × S 1 etc.
108 LEONARDO CONSTANTIN MIHALCEA
π pr1
# {
Ui
such that for all i, j, the restriction
of morphisms over Ui ∩ Uj
ψi ψj
(Ui ∩ Uj ) × Ar o π −1 (Ui ) ∩ π −1 (Uj ) / (Ui ∩ Uj ) × Ar
π
pr1 pr1
* t
Ui ∩ Uj
is of the form
and ψi,j is determined by a linear isomorphism (with coefficients in OUi ∩Uj (Ui ∩ Uj )):
∗
ψi,j : OUi ∩Uj (Ui ∩ Uj ) ⊗k k[x1 , . . . , xn ] → OUi ∩Uj (Ui ∩ Uj ) ⊗k k[x1 , . . . , xn ];
Xr
∗ (i,j)
ψi,j (xk ) = ak,ℓ (u)xℓ .
k=1
∗
In other words, ψi,j is an element of GLr (OUi ∩Uj (Ui ∩ Uj )).
20.1.1. From vector bundles to locally free sheaves. Let π : E → X be a vector bundle
of rank r. We define a sheaf E on X by taking local sections of E: for any open set
U ⊂ X,
E(U ) = {s : U → E : π ◦ s = idU }
We need to make E into an OX -module. The sections over the trivializations E|Ui ≃
Ui × Ar are given by morphisms s : Ui → Ar , i.e. by k-algebra homomorphisms
s∗ : k[x1 , . . . , xr ] → OUi (Ui )); xk 7→ fk ∈ OUi (Ui ).
(See also Theorem 15.2.) The fibres Ex = π −1 (x) over points x ∈ X are k-vector
spaces of dimension r, and because of linearity of the transition maps, the vector space
structures arising from different trivializations are the same. Therefore the space of
sections over U has a OX (U )-algebra structure. This may be enlarged to an OX -module
structure by taking trivializations over Ui ’s (i.e., the trivializations ensure that sections
may be multiplied by regular elements in OX (Ui )). Furthermore, there are OX (Ui )-
module isomorphisms
E(Ui ) ≃ OX (Ui )⊕r ; s 7→ (s∗ (x1 ), . . . , s∗ (xr )),
which are again compatible under restriction to intersections Ui ∩ Uj . This makes E a
locally free sheaf of rank r.
20.1.2. From locally free sheaves to vector bundles. Conversely, if one is given a locally
free sheaf E, then it must have trivializations
E(Ui ) ≃ OX (Ui )⊕r ; f 7→ (f1 , . . . , fr ),
where fi ∈ OX (Ui ). These are maps of OX (Ui )-modules, and determine sections si :
Ui → Ar obtained from s∗i (xj ) 7→ fj . We obtain ‘trivial vector bundles’ Ui × Ar → Ui ,
and one needs to prove that these glue over the intersections Ui ∩ Uj . The key fact is
that over Ui ∩ Uj the two structures of OX -modules are given by OX (Ui ∩ Uj )-module
isomorphisms, i.e. precisely by matrices
(gi,j ) ∈ GLr (OX (Ui ∩ Uj )).
This is the same data which determines trivializations of vector bundles.
Finally, we note that if L is a locally free sheaf of rank 1, then the evaluation map
gives an isomorphism of locally free modules L∗ ⊗OX L ≃ OX . (At the level of transition
functions, (gi,j (L∗ ) = gi,j (L)−1 .) Because of this, line bundles are sometimes called
invertible sheaves.
Generally speaking, if φi : E(Ui ) → Ui × Ar and
gi,j = φj ◦ φ−1
i
In what follows we will need the following notion. Let X be a scheme and U =
Spec(R) ⊂ X open affine. Define K(U ) to be the localization RS , where S is the
multiplicative set of non-zero divisors. These form a presheaf of rings on X, and we
denote by KX its sheafification. It is called the sheaf of total quotient rings.
To describe the transition maps, we distinguish between the cases d ≥ 0 and d < 0:
• If d ≥ 0 then
xdi xdi
(ϕ∗j ◦ (ϕ∗i )−1 )(1) = ϕ∗j (xdi ) = ϕ∗i (xdj ) = .
xdj xdj
• If d < 0 then
−d
1 1 xj x−d
j
(ϕ∗j ◦ (ϕ∗i )−1 )(1) = ϕ∗j ( ) = ϕ∗
j ( ) = .
x−d
i x −d −d
j x i x −d
i
Comparing to the transition functions from Equation (20.1) this shows that we have an
isomorphism of invertible sheaves
OPn (d) ≃ LD
As a ‘reality check’, note that the ‘local sections’ Fi ∈ OPn (Ui ) (corresponding to 1 ∈
LD (Ui )) satisfy
Fj = gi,j Fi
therefore they glue to a global section s = F .
112 LEONARDO CONSTANTIN MIHALCEA
(I.e., this is the rank 1 module over OX generated by 1/fi .) As before the transition
function is given by
∗ −1 1 fj 1 fj
gi,j = ϕj ◦ (ϕi ) (1) = ϕj = ϕj · = .
fi fi fj fi
We need the following lemma.
Lemma 21.3. Let R be a ring and T ⊂ R the subset of non-zero divisors. Let M be
a free rank 1 R-module generated by an element m0 . Assume that there is an injective
R-module homomorphism φ : M → RT and let φ(m0 ) = at . Then φ(m0 ) is invertible.
Proof. It suffices to show that a ∈ T . If not, take r ∈ R \ 0 such that ra = 0. Then
φ(rm0 ) = 0, contradicting injectivity. □
Proposition 21.4. Let X be a scheme.
(a) There is a one-to-one correspondence D ↔ LD between Cartier divisors and in-
vertible OX -modules which are subsheaves of KX .
(b) LD1 −D2 ≃ LD1 ⊗ L−1 D2 .
(c) D1 −D2 is a principal Cartier divisor if and only if LD1 ≃ LD2 as abstract invertible
sheaves, i.e., not as subsheaves of KX .
Proof. To prove (a), we need to show that we can recover the Cartier divisor D out of
the invertible sheaf LD together with its embedding in KX . Take fi to be the inverse
of a local generator of LD (Ui ) ⊂ KX (Ui ) as a OX (Ui )-module. Two such choices must
generate the same module, therefore they must differ by an element in OX (Ui )∗ . As such,
they give the same element of KX (Ui )∗ /OX (Ui )∗ , and thus the same Cartier divisor.
Part (b) follows from the formula for the transition functions. Finally, to prove (c),
we utilize part (b). Then it suffices to show that D is principal if and only if LD ≃ OX .
If D is principal, then it is given by a global section f ∈ KX (X)∗ . This global section
generates LD , in the sense that its restriction to each Ui generated LD (Ui ). Then we
obtain a global isomorphism OX → LD by sending 1 7→ f −1 . Conversely, assume that
we are given an isomorphism OX → LD and consider the image of 1 ∈ OX (X). This
must be a global section f in LD , Since LD is a subsheaf of KX , f must be an invertible
element in KX (X)∗ , and this element determines a principal divisor on X. □
There are many common situations when every invertible sheaf on a scheme X is a
subsheaf of KX . Some examples are:
• X is integral (see [Har77, Prop. II.6.15]);
• X is projective over a field.
In all these situations, the (additive) group of Cartier divisors on X modulo equivalence
is isomorphic to the (multiplicative) group of invertible sheaves. This is called the
Picard group of X, denoted by Pic(X).
One can prove that Pic(Pn ) = Z (generated by OPn (1)), and that Pic(An ) = 0.
114 LEONARDO CONSTANTIN MIHALCEA
Definition 21.2. Let D = (Ui , fi ) be a Cartier divisor. We say that D is effective if the
‘local equations’ are regular functions, i.e., fi ∈ OX (Ui ). For an effective Cartier divisor
D one may associate a codimension 1 subscheme Y ⊂ X given locally by the equations
fi = 0. Sometimes we denote by OX (D), respectively by OX (−D), the invertible sheaf
corresponding to an effective divisor D, respectively −D (the inverse of D).
If D is an effective divisor, then OX (D) has a canonical global section sD , which under
the isomorphism OX (D)|Ui = f1i OX (Ui ) is defined locally by the equations (sD )Ui = fi ∈
OX (Ui ). Indeed,
fj
fj = fi = gi,j fi
fi
showing that this satisfies the required glueing properties.
Example 21.5. If X = Pn and F is a homogeneous polynomial of degree d, it determines
a Cartier divisor D = (Ui , xFd ). The corresponding line bundle LD has a section sD = F ,
i
with vanishing locus Vp (F ) ⊂ Pn .
The following follows from definitions:
Corollary 21.6. Let D be an effective Cartier divisor on a scheme X which determines
the codimension 1 subscheme Y . Then OX (−D) ≃ IY .
Definition 21.3. Let D = (Ui , fi ) be a Cartier divisor on a scheme X. The support of
D, denoted by supp(D), consists of those points x ∈ X with the property that the image
of fi in the local ring (OX,x , mx ) is in mx .
One can show that the support of a closed subset of X. If U = X \ supp(D), then
OX (D) has a section sD ∈ OX (D)(U ) which does not vanish anywhere. This section
extends to a section of OX (D) which is ‘meromorphic’, with poles on the locus where D
is not effective.
Often one can write fi ∈ KX (Ui )∗ as fi = abii where ai , bi ∈ OX (Ui )∗ and such that
A = (Ui , ai ) and B = (Ui , bi ) are Cartier divisors. Then the Cartier divisor D = (Ui , fi )
may be written as a difference of two effective divisors: D = A − B. This will allow us
to define the first Chern class
c1 (LD ) = c1 (LA ) − c1 (LB ) = [V (sA )] − [V (sB )]
where sA , sB are the canonical sections.
Definition 21.4. Let D be a Cartier divisor on a scheme X. A meromorphic section
∗
of OX (D) is any global section of OX (D)⊗OX KX . In other words, a meromorphic section
is not required to have local equations as regular functions in OX (Ui ), but in a localization
of it.
From definition, it follows that sD = (fi ) is a meromorphic section of OX (D). The
divisor D is effective precisely when sD is a (‘holomorphic’) section of OX (D).
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 115
f
Z / Y
Theorem 23.3. Let Z ⊂ Y be any closed subscheme. Then there is a well defined
pull-back homomorphism f ∗ : Ak (Y ) → Ak+n (X) defined by:
f ∗ [Z] = [f −1 (Z)] ∈ A∗ (X).
Proof. See [Ful84, Theorem 1.7]. □
Proposition 23.6. Consider a fibre diagram
g′
X′ / X
f′ f
g
Y′ / Y
where g is flat and f is proper. Then g ′ is flat, f ′ is proper and for any α ∈ Z∗ (X),
f∗′ (g ′ )∗ (α) = g ∗ f∗ (α) ∈ Z∗ (Y ′ ).
120 LEONARDO CONSTANTIN MIHALCEA
24.1. Cartier divisors, Weil divisors, pseudodivisors. Assume for now that X is a
variety of dimension n. (Later it will be a scheme.) In particular, there is a well defined
function field K(X). A Weil divisor is an element of Zn−1 (X).
If D = {(Ui .fi )} is a Cartier divisor, and V ⊂ X is a subvariety then one defines
ordV (D) := ordV (fi ).
This is well defined since fi /fj is invertible on the overlaps. Using this, the Cartier
divisor defines a Weyl divisor
X
[D] = ordV (D)[V ],
V
where the sum is over all codimension 1 subvarieties V ⊂ X. Observe that if two
Cartier divisors D ∼ D′ are linearly equivalent then D − D′ is a principal divisor, i.e.,
D′ − D = div(f ) for some f ∈ K(X). But then the associated Weyl divisors [D′ ]
and [D] are rationally equivalent (by definition). This implies that we have a group
homomorphism
CaCl(X) → An−1 (X); D 7→ [D].
Recall that if D = {(Ui , fi )} is a Cartier divisor, its support is defined to be
|D| = {x ∈ X : some fi is not a unit}
(Observe that fi is in general a rational function, i.e. an element of K(X)∗ . The invertible
∗
elements OX,x ⊂ K(X)∗ are a subset of this. Therefore it makes sense to ask whether fi
is in the smaller subset, although fi may not have a well defined image in OX,x .) The
support is a closed subset of X.
Also recall that D is effective if each fi ∈ OX (Ui ) and it is not a zero divisor.
Definition 24.1 (Pseudovisors). Let X be a scheme. A pseudodivisor on X is a triple
(L, Z, s) where L is a line bundle, Z ⊂ X is closed and s is a section of L over X \ Z
122 LEONARDO CONSTANTIN MIHALCEA
24.2. Intersections by divisors and the first Chern class. We are now ready to
define intersections by divisors and the first Chern class of a line bundle. Assume X is
a scheme, D̃ = (L, |D|, s) a pseudivisor on X, and V ⊂ X a k-dimensional subvariety
with inclusion j :,→ X. Then define an intersection product
D̃.[V ] := [j ∗ D̃] ∈ Ak−1 (V )
In other words:
• If V ̸⊂ |D| then D restrict to a unique Cartier divisor DV on V , and D.[V ] =
[DV ];
• If V ⊂ |D|, then restrict LD to V , and take any Cartier divisor DV corresponding
to this restriction. Then D.[V ] = [DV ].
One extends this product by linearity to define an intersection product
D · α ∈ Ak−1 (X)
for any α ∈ Ak (X).
Definition 24.4 (First Chern class). Let L → X be a line bundle over a scheme X.
Define the group homomorphism c1 (L) ∩ : Ak (X) → Ak−1 (X) by
c1 (L) ∩ [V ] = [C],
where C is the Weil divisor associated to the line bundle L|C .
If L = LD is the line bundle associated to a pseudo-divisor D, then c( L)∩[V ] = D·[V ].
We will list formal properties of the Chern classes later ADDREF, once we define Chern
classes of any order; those would be essentially the same as those defined in topology.
Example 24.1. A homogeneous polynomial F ∈ k[x0 , . . . , xn ] of degree d determines a
Cartier divisor with corresponding line bundle OPn (d). Since the Cartier divisor associ-
ated to F is linearly equivalent to that for any xdi , it follows that
c1 (OPn (d)) = dc1 (OPn (1)).
n
Furthermore, if V ⊂ P is a subvariety such that F |V ̸≡ 0,
(24.1) c1 (OPn (d)) ∩ [V ] = [V ∩ Vp (F )].
Example 24.2 (Chow group of Pn ). One may use Proposition 23.7 to show that Ak (Pn )
is generated by [Lk ], the fundamental class of a k-dimensional plane. We prove that
[Lk ] ̸= 0 in Ak (Pn ). From (24.1) it follows that
c1 (OPn (1))k ∩ [Lk ] = 1.[pt] ∈ A0 (Pn ).
One can use again the sequence from Proposition 23.7 to get that [pt] ̸= 0. (Remove
hyperplanes to reduce to A0 (pt) = Z.) This implies that [Lk ] ̸= 0.
Example 24.3 (Degrees). Let α ∈ Ak (Pn ) ≃ Z. The degree of α is
Z
deg(α) = c1 (OPn (1))k ∩ α.
Pn
124 LEONARDO CONSTANTIN MIHALCEA
The splitting principle allows one to formally define the Chern roots of a vector bundle.
If rank(E) = e, these are formal indeterminates x1 , . . . , xe such that
X
c(E) = (1 + x1 )(1 + x2 ) · . . . · (1 + xe ) = ei (x1 , . . . , xe ).
i≥0
Then the Chern class ci (E) = ei (x1 , . . . , xe ). The idea is that xi ’s are only formal, but
their symmetrizations are actual classes. By the splitting principle, one may actually
identify xi = c1 (Ei /Ei−1 ).
The Chern roots are very useful tools relating Chern classes of vector bundles. I will
list several properties below - all follow from the judicious use of the splitting principle.
Lemma 25.1. In all statements x1 , . . . , xe are the Chern roots of E. Then the following
hold:
• (Dual bundles) The Chern roots of the dual bundle E ∨ are −x1 , . . . , −xe . In
particular,
ci (E ∨ ) = (−1)i ci (E).
• (Tensor products) Let F → X be a vector bundle with Chern roots y1 , . . . , yf .
Then the Chern roots of E ⊗ F are xi + yj , for 1 ≤ i ≤ e, 1 ≤ j ≤ f .
• (Symmetric powers) The Chern roots of Symp E are xi1 + xi2 + . . . + xip where
1 ≤ i1 ≤ i2 ≤ . . . ≤ ip ≤ e.
• (Exterior powers) The Chern roots of p E are xi1 + xi2 + . . . + xip where 1 ≤
V
i1 < i2 < . . . < ip ≤ e.
We illustrate this by some examples.
Example 25.2 (The Chern classes of the tangent bundle of Pn ). We use the Euler
sequence
0 / OPn / OPn (1)⊕(n+1) / TPn / 0.
to calculate the Chern classes of the tangent bundle on Pn . Let H = c1 (OPn (1). By
the normalization property this is the class of a hyperplane in Pn . From the Whitney
formula we obtain that the total Chern class
c(O)c(TPn ) = c(OPn (1)⊕(n+1) = c(OPn (1))n+1 = (1 + H)n+1 .
Since c(O) = 1, it follows that
n
n+1
X n+1
c(TPn ) = (1 + H) = Hk.
k=0
k
For instance, cn (TPn ) = (n + 1)[pt], reflecting the fact that the topological Euler charac-
teristic of Pn is n + 1. (This is an instance of the Gauss-Bonnet theorem.)
128 LEONARDO CONSTANTIN MIHALCEA
Example 25.3 (The degree of Gr(2, 4).). Consider the Grassmannian Gr(2, 4) parametriz-
ing linear subspaces of dimension 2 in C4 . We have seen in a previous problem (Fall
final exam!) that this is a quadric hypersurface in P5 . More precisely, if
a1 b1 c1 d1
A=
a2 b2 c2 d2
and if we denote by pi,j the determinant of the 2 × 2 minor in columns i, j, then
p12 p34 − p13 p24 + p14 p23 = 0.
Let D := {V ∈ Gr(2, 4) : dim V ∩ F2 ≥ 1} where F2 = ⟨e1 , e2 ⟩. You also proved that this
variety is given by the equation p34 = 0 (even scheme theoretically). This says that D is
in fact the intersection of the hyperplane divisor p34 = 0 in P5 by the quadric Gr(2, 4).
Then
Z Z
4
deg(Gr(2, 4) = c1 (OP5 (1)) ∩ [Gr(2, 4)] = c1 (OP5 (1))4 ∩ 2c1 (OP5 (1)) = 2.
P5 P5
This has the following enumerative interpretation. The product c1 (OP5 (1))4 represents
the intersection [D]4 . Since D is the variety of projective lines in P3 meeting the line
P(F2 ). To calculate [D]4 , we are allowed to take divisors which are linearly equivalent
to D, then intersect. In particular, we may move the ‘reference line’ P(F2 ) and define
4 linear equivalent Cartier divisors Di by taking 4 lines ‘sufficiently general’. Then we
are saying that there are 2 lines in P3 meeting meeting 4 given lines in general position.
Example 25.4 (Lines on cubic surfaces). Let Σ ⊂ P3 be a general surface, i.e. Σ = Z(F )
where F is a general homogeneous polynomial of degree 3 in variables x0 , x1 , x2 , x3 . We
are interested in how many lines are included in P3 . The space of lines in P3 is the same
as the Grassmannian Gr(2, 4). Let S be the tautological subbundle. The polynomial F
is a section of Sym3 (C4 )∗ , and the condition that a line P(V ) ∈ Gr(2, 4) is included in
Z(F ) means that F |V ≡ 0. In other words, F gives a (general) section
s ∈ H 0 (Gr(2, 4), Sym3 (S ∗ )), V 7→ F |V .
We are interested in the zero locus of this section. By the normalization property (25.1)
[Σ] = [Z(s)] = cr (Sym3 (S ∗ )) ∩ [Gr(2, 4)],
where r = rank(Sym3 S ∗ ) = 4. Let x1 , x2 be the Chern roots of S ∗ . Then by Lemma
Lemma 25.1 the Chern roots of Sym3 S ∗ are 3x1 , 2x1 + x2 , x1 + 2x2 , 3x2 . It follows that
c4 (Sym3 S ∗ ) = 9x1 x2 (x1 + 2x2 )(2x1 + x2 )
= 9x1 x2 (2x21 + 5x1 x2 + 2x22 )
= 9x1 x2 (2(x1 + x2 )2 + x1 x2 )
= 18x1 x2 (x1 + x2 )2 + 9(x1 x2 )2
= 18c2 (S ∗ )c1 (S ∗ )2 + 9c2 (S ∗ )2 .
LECTURE NOTES FOR ALGEBRAIC GEOMETRY 129
Using the intersection theory of Gr(2, 4), one may prove that
c2 (S ∗ )c1 (S ∗ )2 ∩ [Gr(2, 4)] = c2 (S ∗ )2 ∩ [Gr(2, 4)] = [pt].
Therefore
c4 (Sym3 S ∗ ) ∩ [Gr(2, 4)] = 18[pt] + 9[pt] = 27[pt].
The enumerative geometry interpretation of this is that a general cubic surface will con-
tain 27 lines.
25.2. Some answers to enumerative geometry questions. You might enjoy think-
ing about these numbers:
• 2: number of lines passing through 4 given lines in P3 ;
• 92: number of conics in P3 meeting 8 lines;
• 1: non-singular conics tangent to 5 given lines;
• 3264: number of conics tangent to 5 given conics;
• 4407296: number of conics tagent to 8 general quadrics surfaces;
• 2875: number of lines on a quintic threefold in P4 ;
• 5819539783680: number of twisted cubics tangent to 12 quadrics in P3 . (This
was found by H. Schubert in 1870’s.)
THAT’S ALL FOLKS!
130 LEONARDO CONSTANTIN MIHALCEA
26.2. Primary and irreducible ideals. The main reference for this section is [AM69,
Ch.4 and Ch. 7].
Definition 26.1. Let R be an arbitrary ring and I ̸= R an ideal. The ideal I is primary
if
xy ∈ I =⇒ x ∈ I or y n ∈ I.
The ideal I is called irreducible if
I = I1 ∩ I2 =⇒ I = I1 or I = I2 .
√
Proposition 26.5. Let I be a primary ideal. Then the radical I is prime and it is the
smallest prime containing I.
Proof. See [AM69, Prop. 4.1]. □
Proposition 26.6. Let R be a Noetherian ring.
(a) Any ideal I is a finite intersection of irreducible ideals.
(b) Any irreducible ideal is primary.
Proof. See Lemmas 7.11 and 7.12 in [AM69]. □
References
[AM69] M. F. Atiyah and I. G. Macdonald. Introduction to commutative algebra. Addison-Wesley
Publishing Co., Reading, Mass.-London-Don Mills, Ont., 1969.
[CLO15] David A. Cox, John Little, and Donal O’Shea. Ideals, varieties, and algorithms. Undergradu-
ate Texts in Mathematics. Springer, Cham, fourth edition, 2015. An introduction to compu-
tational algebraic geometry and commutative algebra.
[Ful84] William Fulton. Intersection theory. Springer-Verlag, Berlin, 1984.
[Har77] Robin Hartshorne. Algebraic geometry. Graduate Texts in Mathematics, No. 52. Springer-
Verlag, New York-Heidelberg, 1977.
[Mum99] David Mumford. The red book of varieties and schemes, volume 1358 of Lecture Notes in
Mathematics. Springer-Verlag, Berlin, expanded edition, 1999. Includes the Michigan lectures
(1974) on curves and their Jacobians, With contributions by Enrico Arbarello.