Lecture Notes
Lecture Notes
Christian Böhning
Mathematics Institute
University of Warwick
ii
Contents
8 Dimension theory 59
iii
iv CONTENTS
Chapter 1
Definition 1.1. For a k-vector space, P(V ) denotes the set of 1-dimensional
subspaces of V , i.e. the projective space associated to V . For the vector
space k n+1 with componentwise addition and scalar multiplication, we also
write Pn = P(k n+1 ).
Equivalently, Pn is the quotient of k n+1 − {0} by the equivalence relation:
Ui := {(X0 : · · · : Xn ) ∈ Pn | Xi 6= 0} .
1
2 CHAPTER 1. AFFINE AND PROJECTIVE ALGEBRAIC SETS
F (p) = 0 : ⇐⇒ F (P0 , . . . , Pn ) = 0
for all (P0 , . . . , Pn ) ∈ k n+1 − {0} with [(P0 , . . . , Pn )] = p.
This leads to the conclusion that we can assume, without loss of gen-
erality, that the Fα in (2) of Definition 1.2 are homogeneous, which means
the following: S = k[X0 , . . . , Xn ] is a graded ring, which means there is a
decomposition into k-vector subspaces
M
S= Sm
m≥0
U0 only since the other cases are only notationally different. If X0 is defined
by fα (x1 , . . . , xn ) of degree dα (of course not necessarily homogeneous now),
we can define a suitable X by
dα X1 Xn
Fα (X0 , . . . , Xn ) := X0 fα ,..., .
X0 X0
ν : P1 → P3
(X0 : X1 ) 7→ (X03 : X02 X1 : X0 X12 : X13 ) := (Z0 : Z1 : Z2 : Z3 ).
F0 (Z) = Z0 Z2 − Z12
F1 (Z) = Z0 Z3 − Z1 Z2
F2 (Z) = Z1 Z3 − Z22
and those define C, i.e. the zero set of them is exactly C (to see this,
note that if p ∈ P3 , p = (P0 : P1 : P2 : P3 ) lies on Q0 ∩ Q1 ∩ Q2 , then
P0 6= 0 or P3 6= 0; in the former case, P = ν((P0 : P1 )), in the latter
case, P = ν((P2 : P3 )).
4 CHAPTER 1. AFFINE AND PROJECTIVE ALGEBRAIC SETS
νd : P1 → Pd ,
(X0 : X1 ) 7→ (X0d : X0d−1 X1 : · · · : X1d )
G(X0 , X1 )
Hi := ,
(νi X0 − µi X1 )
we get that the image of νd given by
passes through the coordinate points, namely maps (µi : νi ) ∈ P1 to the i-th
coordinate point. We can also assume (µi : νi ) are different from (1 : 0) and
7
(0 : 1); then, given two general additional points pd+2 , pd+3 in Pd , we can
always adjust the µi , νi so that (1 : 0) maps to pd+2 and (0 : 1) maps to pd+3 .
This proves existence.
For uniqueness, note that every rational normal curve passing through the
coordinate points p1 , . . . , pd+1 is the image of a map given by polynomials Hi
as above for certain (µi , νi ). Applying a projectivity in the source, we can also
assume that (1 : 0) and (0 : 1) map to pd+2 , pd+3 . Then such a rational normal
curve is given by polynomials Hi as above, and moreover, the µi are fixed up
to simultaneous rescaling by a constant nonzero factor α, and so are the νj
up to a factor β. Applying the projectivity (X0 : X1 ) 7→ (α−1 X0 : β −1 X1 ) in
the source, we see that the maps corresponding to different α, β all have the
same image. Hence this is the unique rational normal curve meeting all the
requirements.
Another example of projective algebraic sets are finite point sets Γ =
{p1 , . . . , pN } ⊂ Pn : indeed, if q ∈
/ Γ, there is a polynomial vanishing in Γ,
but not in q (take a product of N linear forms). Hence Γ is defined by
polynomials of degree ≤ N .
It is known from courses in linear algebra that
1. Two ordered point sets (p1 , . . . , pn+2 ), (q1 , . . . , qn+2 ) in general posi-
tion can be transformed into each other by a unique projectivity g ∈
PGLn+1 (k).
2. In P1 , one can transform 4 = 1 + 3 ordered points in general position
into another ordered four points in general position if and only if their
cross ratios
(z1 − z2 )(z3 − z4 )
(z1 − z3 )(z2 − z4 )
are the same.
of the Riemann sphere are fractional linear), one gets that (p1 , . . . , pn+3 ) and
(q1 , . . . , qn+3 ) are projectively equivalent on Pn if and only if (p01 , . . . , p0n+3 )
0
and (q10 , . . . , qn+3 ) are projectively equivalent on P1 . The latter means that
9
10 CHAPTER 2. ALGEBRAIC VARIETIES AND REGULAR MAPS
with
f (ϕ(t), ψ(t)) = 0
11
OX (U ) ⊂ Maps(U, k)
S
such that the following is true: if U ⊂ X is open and U = i∈I Vi a cover
of U by open subsets Vi ⊂ U ⊂ X, i ∈ I, I some index set, then a function
f ∈ Maps(U, k) belongs to OX (U ) if and only if all restrictions f |Vi belong
to OX (Vi ) (for all i ∈ I).
This may seem intimidating at first sight, but is totally simple really:
it just means that on any open set U in X we mark certain functions as
distinguished, by painting them red, say; we call these OX (U ). The condition
then means that being “red” is a local property: if we restrict one of our
functions to the open sets of a cover, and if it is red on every one of them,
then it is red globally.
Example 2.5. 1. Let X = R1 with the Euclidean topology, and put for
U ⊂ R open, OX (U ) = {f : U → R | f continuous}. Then OX is a
sheaf of R-valued functions. The property of being continuous is local.
3. Let X = ∆1 t∆2 be the disjoint union of two open discs in R2 , with the
induced topology from R2 . For U ⊂ X open, let OX (U ) = {f : U → R |
f constant}. Then OX is not a sheaf: the property of being constant
is not local. For example, the function which is 1 on ∆1 and 0 on ∆2
is constant locally, but not constant globally.
12 CHAPTER 2. ALGEBRAIC VARIETIES AND REGULAR MAPS
Definition 2.7. Let (X, OX ) and (Y, OY ) be spaces with sheaves of func-
tions. A morphism
ϕ : (X, OX ) → (Y, OY )
is a continuous map ϕ : X → Y with the property that for all U ⊂ Y open,
and for all f ∈ OY (U ), one has ϕ∗ (f ) := f ◦ ϕ ∈ OX (ϕ−1 (U )).
This also looks complicated, but means only that the local functions are
those that can be written, locally, as quotients of two polynomials (homo-
geneous of the same degree in the case of Pn ), with the denominator non-
vanishing on the open under consideration.
Remark 2.10. With these definitions, the natural projection An+1 −{0} → Pn
becomes a morphism, and the natural bijections of the subsets Ui = {Xi 6=
0} ⊂ Pn with An become isomorphisms, as it should be.
We call OX,x the stalk of OX in x; elements of OX,x are called germs (of
functions around x).
These definitions are very slick and smooth, and leave us with a beautiful
category of algebraic varieties, but the drawback is that it takes quite a
while to develop an intuition for what they mean and to work with them.
For example, what is OAn (An ) or OPn (Pn )? Here is a useful example of a
morphism.
14 CHAPTER 2. ALGEBRAIC VARIETIES AND REGULAR MAPS
C : ZY 2 = X(X 2 − Z 2 ).
y 2 = x(x2 − 1).
b2 t2 = at(a2 t2 − 1)
0 = at(t − 1)(a2 t + 1)
However, this is not a well-defined map of UZ ' A2 into itself because the
formula makes no sense for (a, b) = (0, 0). However, (draw a picture!) the
point (0, 0) should map to P1 = (0 : 1 : 0), the unique point at ∞ on C with
respect to the coordinates x, y! To prove that α is a morphism, we must
rather cover C with various open sets as follows: put
P1 = (0 : 1 : 0),
P2 = (0 : 0 : 1)(= P0 ),
Q1 = (1 : 0 : 1),
Q2 = (−1 : 0 : 1)
15
U1 := C − {P1 , P2 }
U2 := C − {P1 , Q1 , Q2 }
U3 = C − {P2 , Q1 , Q2 }
V1 := C − {P1 } = C ∩ UZ = CZ
V2 = C − {P2 , Q1 , Q2 } = C ∩ {Y 6= 0} = C ∩ UY = CY .
Then
1. U1 , U2 , U3 is an open cover of C.
2. V1 , V2 is an affine open cover of C.
3. We have
α(U1 ) ⊂ V1 , α(U2 ) ⊂ V2 , α(U3 ) ⊂ V1 .
Note that x, y, 1/(x2 − 1) ∈ OC (U2 ) and the right hand side of the last
displayed equation is a polynomial in a, b, 1/(a2 − 1), so α |U2 : U2 → V2
is a morphism.
Hilbert’s Nullstellensatz,
primary decomposition and
geometric applications
for the ideal of all polynomials vanishing on X. We call it the ideal of X for
short.
17
18CHAPTER 3. HILBERT’S NULLSTELLENSATZ, PRIMARY DECOMPOSITION
√
Proof. It is clear that I ⊂ I(V (I)), so it suffices to prove the opposite
inclusion.
Step 1. We prove that if V (I) = ∅, then I = A. Suppose by contradiction
that I ( A. Then I ⊂ m for some maximal ideal in A. Now Step 1 will be
complete once we can show that every maximal ideal of A is of the form
m = (x1 − a1 , . . . , xn − an )
for some ai ∈ k. Indeed, then V (I) will contain the point (a1 , . . . , an ), hence
not be empty. Now the assertion that every maximal ideal is of the form
above is equivalent to the claim that the quotient field
L = k[x1 , . . . , xn ]/m
is isomorphic to k via the inclusion k ⊂ L, and since k is algebraically closed,
this is in turn equivalent to showing that L is algebraic over k. We can
assume that x1 , . . . , xl ∈ L are algebraically independent over k and that
xl+1 , . . . , xn are algebraic over k(x1 , . . . , xl ) ⊂ L.
Lemma 3.3. Let R be a Noetherian ring, S ⊃ R a finitely generated R-
algebra. If T ⊂ S is an R-algebra such that S is a finitely T -module, then T
is a finitely generated R-algebra.
Let us first indicate how the Lemma allows us to finish the proof of Step
1. Apply it with R = k, S = L, T = k(x1 , . . . , xl ). The hypotheses are
then satisfies, hence we would get that k(x1 , . . . , xl ) is a finitely generated
k-algebra, which is absurd unless l = 0: if rational functions z1 , . . . , zN
Pi (x1 , . . . , xn )
zi =
Q(x1 , . . . , xn )
were generators, and if f ∈ k[x1 , . . . , xn ] is irreducible, then 1/f must be a
polynomial function of the zi . But then f would have to divide one of the
Qi since A = k[x1 , . . . , xn ] is a UFD, leading to the absurd conclusion that
there are only finitely many irreducible polynomials in A.
Proof. (of Lemma 3.3) Let ξ1 , . . . , ξp be R-algebra generators of S and let
s1 , . . . , sq be T -module generators of S. Thus ∃ tικ , t0ικλ ∈ T such that
X
ξι = tικ sκ (3.1)
κ
X
sι · sκ = t0ικλ sλ . (3.2)
λ
19
(the brackets indicate to take the ideal generated). Then V (J) = ∅, thus by
Step 1, J = k[x1 , . . . , xn ]. Thus if
1 = h0 + h1 xn+1 + · · · + hm xm
n+1 , hi ∈ I.
f m = f m h0 + · · · + hm
and this is in I.
We can use the above to describe rings of regular functions in several
cases more simply:
on Uβ , since
hα hβ
= in Quot(k[x1 , . . . , xn ]/I(X)),
kα kβ
where Quot(k[x1 , . . . , xn ]/I(X)) is the quotient field of k[x1 , . . . , xn ]/I(X).
Namely, hα kβ − hβ kα vanishes on the open Uα ∩ Uβ in X, hence is in I(X),
since I(X) is prime.
Remark 3.5. The statement of Corollary 3.4 also holds without the assump-
tion that I(X) is prime, but the proof is a bit messier then and we omit
it.
Corollary 3.6. We have OPn (Pn ) = k.
Proof. Because of Corollary 3.4, we have in An ' U0 = {X0 6= 0} ⊂ Pn that
an f ∈ OPn (Pn ) can be written as
X1 Xn
f =p ,..., , p ∈ k[x1 , . . . , xn ].
X0 X0
21
The same method of proof even shows that OPn (Pn − H1 ∩ H2 ) = k where
H1 , H2 are two different hyperplanes in Pn .
with fi ∈ k[x1 , . . . , xn ].
Proof. This follows from Corollary 3.4 (and Remark 3.5) since Tj ◦ f , Tj the
j-th coordinate function on Am , is regular on X.
S/I(X)
I = p1 ∩ · · · ∩ ps
I = I λ := {f (λx0 , . . . , λxn ) | f ∈ I}
for infinitely many λ ∈ k ∗ . This implies that for every i there is a j with
pi = pλj for infinitely many λ = λ1 , λ2 , . . . since applying (−)λ gives another
representation of I as an irredundant intersection of primes, and it is unique
λ λ−1
up to reordering. But then pi j 1 = pi for all j, hence pi is homogeneous.
Remark 3.12. As an exercise, one should formulate and prove an extension of
the Nullstellensatz for projective varieties now. Moreover, one should verify
that
Clearing denominators in (3.3), we get one matrix row for A. Doing this for
all i, j, we get A.
Example 3.14. Let us illustrate the preceding Theorem 3.13 in one geomet-
ric situation. Suppose C ⊂ P2 is given by
X 2 + Y 2 − Z 2 = 0.
Then: the line Y − Z = 0 intersects C in one point, namely P = (0 : 1 : 1),
the line Y + Z = 0 intersects C in one point (0 : −1 : 1), and we want
to consider the stereographic projection of that conic from P onto the line
Y = 0:
f : C − {(0 : 1 : 1)} → P1
(X : Y : Z) 7→ (X : Z − Y ).
Thus, geometrically, we construct the line through P and a point R ∈ C−{P }
and let f (R) be the intersection of that line with the line Y = 0. Now
define f˜ by f on C − {P } and f˜(P ) = (1 : 0) (the point at infinity on
P1 ' {Y = 0}. Then f˜ is a morphism, which can be seen as follows: let
(S : T ) be homogeneous coordinates on P1 , with US = {S 6= 0} ' A1 ,
UT = {T 6= 0} ' A1 . Then
f˜−1 (US ) = C − {(0 : 1 : −1)},
f˜−1 (UT ) = C − {(0 : 1 : 1)}.
25
(X : Y : Z) 7→ (Y + Z : X).
The matrix
X Z −Y
A=
Y +Z X
f˜: C → P1
X F0
and
Z −Y F1
define the same rational function in k(t) on A1 via τ since XF1 −F0 (Z −Y ) =
0 on C. Now F1 vanishes in (0 : 1 : 1), but F0 does not, since (0 : 1 : 1) maps
to (1 : 0). Now
∗ X 1
τ =
Z −Y t
has a simple pole in 0 ∈ A1 . Hence the same must hold for
F0 (2t, −t2 + 1, t2 + 1)
∗ F0
τ = .
F1 F1 (2t, −t2 + 1, t2 + 1)
(so all 2 × 2-minors of the matrix (Zij )). Then the map
27
28CHAPTER 4. SEGRE EMBEDDINGS, VERONESE MAPS AND PRODUCTS
and thus Xi = µXi0 for all i. Thus X = X 0 in projective space, and analo-
gously we get Y = Y 0 .
Step 2. The map sn,m is surjective. Suppose Zij satisfy Zij Zkl = Zil Zkj
and not all Z’s are zero, Zi0 j0 6= 0, say. Put
Zij0 Zi0 j
Xi := , Yj = .
Zi0 j0 Zi0 j0
Then
(Zi0 j0 )2 Xi Yj = Zij0 Zi0 j = Zij Zi0 j0
which means s(X, Y ) = [(Zij )].
Step 3. The map sn,m is a homeomorphism. Indeed, the topology we
put on Pn × Pm has a basis consisting of the sets
Thus Σn,m = V (a) is irreducible and this concludes the proof of Theorem
4.2.
We now give Pn ×Pm the structure of a projective variety, and every prod-
uct X × Y of (quasi-)projective varieties the structure of (quasi-)projective
variety that comes from identifying it with the image under sn,m .
30CHAPTER 4. SEGRE EMBEDDINGS, VERONESE MAPS AND PRODUCTS
Example 4.4. The variety Σ1,1 = s1,1 (P1 × P1 ) ⊂ P3 is the image of the map
These are the loci where either the rows or columns of the above matrix
satisfy a linear dependency relation.
In fact, any line L ⊂ P3 on Σ1,1 belongs to one of these two families:
suppose we write L as the image of a map P1 → P3
where the li (λ, µ) are linear forms. The condition that L lie on Σ1,1 means
that
l1 (λ, µ) l2 (λ, µ)
det ≡ 0.
l3 (λ, µ) l4 (λ, µ)
Without loss of generality, after possibly interchanging rows or columns and
transposing, we may assume l1 6= 0 and l1 , l2 linearly independent. Then
l1 l4 = l2 l3 implies that l1 | l3 and l2 | l4 , hence l3 = cl1 and l4 = cl2 , some
c ∈ k.
Remark 4.5. In coordinate free form, the Segre map can be given as
Z
α×β
β
α
X ×Y
prY
prX
{ #
X Y
commutes. To see this note that purely set-theoretically it is clear how we
have to define α × β if it is to exist at all. Then we just have to check that
α × β is a morphism. To see this suppose r0 ∈ Z is a point with α(r0 ) = p,
β(r0 ) = q. Suppose p is in X0 6= 0 in Pn . Then, in a neighborhood of r0 the
map α can be given by
r 7→ (1 : f1 (r) : · · · : fn (r))
which is regular.
Remark 4.7. If f : X → Y is a morphism between projective varieties X ⊂
Pn , Y ⊂ Pm , then
Γf := {(x, f (x)) | x ∈ X} ⊂ X × Y ⊂ Pn × Pm
α0 + · · · + αn = m
Now put
m+n
N := −1
m
and for each decomposition into nonnegative integers i0 + · · · + in = m,
introduce a symbol vi0 ...in , and consider these as homogeneous coordinates in
PN . Then the m-th Veronese map is the map
v m : Pn → PN
vi0 ...in := ui00 . . . uinn
3.
Ann(ω1 ) ∩ Ann(ω2 ) = {0} =⇒ Ann(ω1 ) + Ann(ω2 ) = Ann(ω1 ∧ ω2 ).
35
36CHAPTER 5. GRASSMANNIANS, FLAG MANIFOLDS, SCHUBERT VARIETIES
ω̃ = v1 ∧ · · · ∧ vp .
Thus we can choose a basis of Ann(ω1 ) of the form (v10 , . . . , vq0 , vq+1
0
, . . . , vp0 )
whence
ω1 = det(A)v10 ∧ · · · ∧ vq0 ∧ vq+1
0
∧ . . . vp0
where A is the base change matrix from (v10 , . . . , vp0 ) to (v1 , . . . vp ). Thus there
0
exists ω = ± det(A)vq+1 ∧ . . . vp0 with ω1 = ω ∧ ω2 .
For (2) and (3), first note that
This shows (2), and (3) follows from what has been proven so far (in partic-
ular (5.1)).
is bijective.
Proof. From the proof of Theorem 5.2 we know that if ω 6= 0 and ω is com-
pletely reducible, then dim Ann(ω) = p. Suppose conversely that dim Ann(ω) =
38CHAPTER 5. GRASSMANNIANS, FLAG MANIFOLDS, SCHUBERT VARIETIES
r (we will need the case r = p). Suppose that Ann(ω) = hv1 , . . . , vr i and
that (v1 , . . . , vr , vr+1 , . . . , vn ) is a basis of V . Write
X
ω= ωi1 ...ip vi1 ∧ · · · ∧ vip .
1≤i1 <···<ip ≤n
This is equivalent to
X i i i i
ωi1 i2 ωj1 j2 sgn 1 2 3 4 = 0.
k1 k2 k3 k4
Thus we get one equation for each quadruple of indices 1 ≤ k1 < k2 < k3 <
k4 ≤ n and the sum above then runs over those indices 1 ≤ i1 < i2 ≤ n and
1 ≤ j1 < j2 ≤ n with {i1 , i2 , i3 , i4 } = {k1 , k2 , k3 , k4 }. With sgn we mean the
sign of the permutation
i1 i2 i3 i4
.
k1 k2 k3 k4
40CHAPTER 5. GRASSMANNIANS, FLAG MANIFOLDS, SCHUBERT VARIETIES
Σl (M ) ⊂ Grass(p, V )
where M ⊂ V is an m-dimensional subspace, and Σl (M ) consists of those
p-dimensional W ⊂ V with dim(W ∩ M ) ≥ l.
Proof. Indeed,
Flag(a1 , . . . , at ; V ) = {(W1 , . . . , Wt ) | W1 ⊂ · · · ⊂ Wt }
⊂ Grass(a1 , V ) × · · · × Grass(at , V )
where the latter product is a projective variety via the Segre embedding.
Elements in Flag(a1 , . . . , at ; V ) are called flags of type (a1 , . . . , at ) in V .
Proof. If
then the construction runs analogously: the subspaces W with W ∩Γi1 ...in−p =
{0} form an open subset Uj1 ,...,jp ' Ap(n−p) of Grass(p, V ), where (j1 , . . . , jp )
are the indices complementary to (i1 , . . . , in−p ) in {1, . . . , n}, and Uj1 ,...,jp is
then given by
w1j1 . . . w1jp
det ... .. .. 6= 0.
. .
wp,j1 . . . wpjp
The Uj1 ,...,jp form a cover of Grass(p, V ) by affine spaces.
Step 2. Consider now
is closed in Aa1 (n−a1 ) × Aa2 (n−a2 ) . Without loss of generality we can assume
{l1 , . . . , la2 } = {1, . . . , a2 }. An M ∈ Ul1 ...la2 is given by a matrix
1 . . . 0 m1,1 . . . m1,n−a2
.. . . .. .. .. ..
.
. . . . . .
0 . . . 1 ma2 ,1 . . . ma2 ,n−a2
and the columns corresponding to j1 , . . . , ja1 form the unit matrix. Then
S ⊂ M means that the matrix
s1,1 . . . . . . ... ... s1,n
.. ... ... ... ..
. ... .
sa1 ,1 . . . . . . ... ... sa1 ,n
1 ... 0 m1,1 ... m1,n−a2
. .. . .. .. ..
.. . ..
. . .
0 ... 1 ma2 ,1 . . . ma2 ,n−a2
has rank ≤ a2 , i.e., all (a2 + 1) × (a2 + 1)-minors vanish.This finishes the
proof.
2
Remark 5.10. The subset G = GLn (k) ⊂ An is open, hence a variety. What
is more, the group composition and inverse map are morphisms. The varieties
are homogeneous spaces for G, which means that G acts transitively on them
and the obvious group actions
G × Grass(p, V ) → Grass(p, V ),
G × Flag(a1 , . . . , at , V ) → Flag(a1 , . . . , at , V )
are morphisms.
44CHAPTER 5. GRASSMANNIANS, FLAG MANIFOLDS, SCHUBERT VARIETIES
Chapter 6
f : A2 → A2
(x, y) 7→ (x, xy).
Then Im(f ) = {(a, b) ∈ A2 | a 6= 0} ∪ {(0, 0}. This is not open in its closure
A2 .
Proof. The graph Γf ⊂ X×Y is closed by by Remark 4.7 and f (X) = pr2 (Γf )
where pr2 : X × Y → Y is the projection onto the second factor. So it suffices
to show:
If Z ⊂ X × Y is closed and X is projective, then pr2 (Z) ⊂ Y is closed.
Now if X ⊂ Pn then Z is also closed in Pn × Y and pr2 (Z) is equal to the
image of the projection Pn × Y → Y . Thus we can also assume X = Pn .
45
46CHAPTER 6. IMAGES OF PROJECTIVE VARIETIES UNDER MORPHISMS
S
Now let Y = Ui be a finite cover of Y by affine open subsets. Then
i∈I
[ [
X ×Y = X × Ui , Z = Z ∩ (X × Ui ),
i∈I
[
pr2 (Z) = pr2 (Z ∩ (X × Ui )) ,
i∈I
Fi (T0 , . . . , Tn , t1 , . . . , tm ) = 0, i = 1, . . . , N,
Za := {(T0 : · · · : Tn ) ∈ Pn | Fi (T0 , . . . , Tn , a1 , . . . , am ) = 0 ∀ i = . . . , N }
pr2 (Z) = {a ∈ Am | Za 6= ∅}
\
= {a ∈ Am | (T0 , . . . , Tn )s 6⊂ Ia }.
s≥0
Ys := {a ∈ Am | (T0 , . . . , Tn )s 6⊂ Ia }
47
Proof. In the proof of Theorem 6.2 we saw that, in the situation and notation
of Definition 6.4, the projection X × Y → Y maps closed sets to closed sets.
But X ×Z Y is closed in X × Y , and W is closed in X ×Z Y , hence also in
X ×Y.
An − {point}, Pn − {point}
51
52 CHAPTER 7. FINITE MORPHISMS, NOETHER NORMALIZATION
det(cId − M )ωi = 0, i = 1, . . . , r.
f ∗ : OY (Y ) ,→ OX (X)
tki + a1 tik−1 + · · · + ak = 0 ai ∈ OY
and if x ∈ f −1 (y)
In other words, the fiber is empty if and only if my · OX (X) = OX (X). Hence
the assertion follows from Lemma 7.5 below.
Lemma 7.5. Suppose B ⊃ A is finite. Then, if a ⊂ A is an ideal, we have
the implication
a ( A =⇒ aB ( B.
53
But (gαnα ) = P
1 since the gαnα have no common zeroes on Y , hence there are
hα ∈ A with α gαnα hα = 1. Thus
X XX
b=b gαnα hα = ai,α hα ωi,α .
α i α
Theorem 7.8 tells us that “finite” in this new sense coincides with the
notion of “finite” introduced in Definition 7.2.
Remark that we immediately obtain that for finite f in the sense of Def-
inition 7.9, it continues to be true that all fibers f −1 (y) are finite, and every
such f is surjective.
h : X → Y × Ar , g = pr1 : Y × Ar → Y.
πL : X → Pn−d−1
be the projection with center L; this means that if L is given by linear equa-
tions L0 = · · · = Ln−d−1 = 0, then
Proof. This follows from Theorem 7.11 since f is the composition of the m-th
Veronese embedding of X and a linear projection.
π : Ui → An−d−1
yi 6=0 ∩ π(X)
Gi (x0 , . . . , xn )
g=
Lm i
π̃ : X → Pn−d
x 7→ (z0 : · · · : zn−d ), zj := (Lj (x))m , j ≤ n − d − 1, zn−d := Gi (x).
z0 = · · · = zn−d−1 = F1 = · · · = Fs = 0
(k−1)
where Hj is the homogeneous component of degree k − 1 of Hj . This can
be rephrased by saying that
Φ(Lm m
0 , . . . , Ln−d−1 , Gi ) ≡ 0 on X.
Dimension theory
3. If X is projective, then
f
k(X) = | f, g ∈ k[X0 , . . . , Xn ]/I(X), f, g homogeneous of the same degree, g 6≡ 0 .
g
59
60 CHAPTER 8. DIMENSION THEORY
F0 := F , dim X (i+1) < dim X (i) . If dim X = n, then X (n+1) is empty. This
means that F0 = F, . . . , Fn have no common zeroes on X. Now, if X is
irreducible, we get by Corollary 7.14 that
f : X → Pn
f (X) = (F0 (X) : · · · : Fn (X))
is finite onto its image f (X) ⊂ Pn . Hence dim X = dim f (X) = n, and
by Theorem 8.4, f (X) = Pn . Now, if, arguing by contradiction, we had
dim X (1) < n−1, the already X (n) would be empty. I.e., already F0 , . . . , Fn−1
would have no common zeroes on X. But then (0 : · · · : 0 : 1) wouldn’t be
in the image of f , a contradiction.
We can also regard Theorem 8.7 as giving a strong existence result for
solution sets of polynomial equations:
This implies for example that in P2 , any two curves intersect since by
Theorem 8.6, these are defined by one equation; this is false on P1 × P1 .
Lines of the same ruling do not intersect. This proves also that P2 is not
isomorphic to P1 × P1 although the two are birational.
We can also use Corollary 8.10 to see that any curve of degree ≥ 3 in P2
has an inflection point, and for many similar geometric existence results.
We now prove a strengthening of Theorem 8.7.
64 CHAPTER 8. DIMENSION THEORY
is finite onto its image f (X). Let An(xi 6=0) ⊂ Pn be the standard affine chart,
and Ui ⊂ X its preimage under f . This is open and affine (the latter since
we can realize f as the composition of a Veronese embedding with a linear
projection).
Now it suffices to prove that every component of (X ∩ {F = 0}) ∩ Ui has
dimension n−1 (for all i). The latter set is the zero set, in Ui , of the function
F
ϕ := ∈ OX (Ui ).
Fi
Put U := Ui . Then f restricted to U gives a map
∀ Q ∈ OU (U ) : RQ ≡ 0 on (ϕ = 0) =⇒ Q ≡ 0 on (ϕ = 0). (8.1)
For, if
{u ∈ U | ϕ(u) = 0} = U (1) ∪ · · · ∪ U (t)
65
F (T ) = T k + b1 T k−1 + · · · + bk
namely,
xb1 k−1 x k bk
G(T ) = T k + T + ··· + k . (8.3)
z z
Since v is integral over B,
x i bi
∈ B ∀ i.
zi
But then, since x, z are relatively prime, z i | bi , and substituting v for T in
(8.3) yields x | v k .
Corollary 8.13. Let X ⊂ PN be an irreducible quasi-projective variety, and
F 6= 0 a homogeneous polynomial of positive degree such that F 6≡ 0 on X.
Then every irreducible component of X ∩ {F = 0} has codimension 1.
Proof. It suffices to apply Theorem 8.11 to the closure X ⊂ PN , and remark
that X ∩ {F = 0} = (X ∩ {F = 0}) ∩ X.
The next result also follows directly from Theorem 8.11 now.
Corollary 8.14. If X ⊂ PN is irreducible and quasi-projective, dim X = n,
and Y ⊂ X is the zero set of homogeneous polynomials of positive degrees
F1 , . . . , Fm , then every component of Y has dimension ≥ n − m.
Theorem 8.15. Let X, Y ⊂ PN be irreducible, quasi-projective varieties,
dim X = n, dim Y = m. Then every component Z of X ∩ Y has dimension
≥ n + m − N . If X and Y are projective and n + m ≥ N , then X ∩ Y 6= ∅.
Proof. To prove the first assertion, it suffices to consider the case when
X, Y ⊂ AN are affine. Then
X ∩ Y ' (X × Y ) ∩ ∆AN
and the diagonal ∆AN ⊂ A2N is defined by N equations. Then the first asser-
tion follows from Corollary 8.14. The second assertion follows by considering
the affine cones over X, Y and applying the first assertion.
We can also say that
r
\ r
X
codim Pn Xi ≤ codimPn (Xi )
I=1 i=1
for any finite number r of irreducible quasi-projective subvarieties Xi ⊂ Pn .
The next theorem describes how the dimensions of the fibers of a mor-
phism can vary and has very many applications.
67
f ∗ : OW (W ) ,→ OV (V )
Fi (vi ; v1 , . . . , vn−m ; w1 , . . . , wM ) = 0, i = n − m + 1, . . . , N
69
70 CHAPTER 9. LINES ON SURFACES, DEGREE
where N = d+3
3
and Grass(2, 4) is the Plücker quadric in P5 . We have
q(Φ) = Grass(2, 4) since every line l ∈ P3 lies on some (possibly reducible)
surface of degree d, for example we could take a union of d planes through l.
The next question is: what is dim q −1 (l)? After applying a projectivity, we
may assume l is given by X0 = X1 = 0 in P3 . Then any surface of degree d
containing l is given by an equation of the form
F = X0 G + X1 H, G, H ∈ k[X0 , X1 , X2 , X3 ]d−1
and any such equation defines a surface containing l; thus the fibers of q are
all projective linear subspaces of PN of dimension (the same for all l)
d(d + 1)(d + 5)
dim q −1 (l) = − 1.
6
To see that this is the dimension of q −1 (l), use the exact sequence (where
S = k[X0 , X1 , X2 , X3 ])
X1
β=
−X0 α= X0 X1
0 / Sd−2 / Sd−1 ⊕ Sd−1 / Sd
which gives
dim q −1 (l) + 1 = dim im(α) = 2 dim Sd−1 − dim Sd−2
2(d+2)(d+1)d−(d+1)d(d−1)
= 2 d−1+3 − d−2+3
3 3
= 6
(d+5)d(d+1)
= 6
.
71
Thus there is a point in P19 over which the fiber of p has dimension 0. By
Theorem 8.16 we get dim p(Φ) = 19, whence p(Φ) = P19 . Thus we have
proven:
72 CHAPTER 9. LINES ON SURFACES, DEGREE
Theorem 9.4. Every cubic surface in P3 contains at least one line. There is
a nonempty Zariski open subset in P19 such that the corresponding surfaces
contain finitely many lines.
However, there are cubic surfaces with infinitely many lines, e.g. cones
over cubics in P2 . Hence here we observe again jumps in the fiber dimension.
Moreover, one can show that for “most” cubic surfaces with finitely many
lines, the number and configuration of the lines (i.e., their incidence graph)
are independent of the surface. The number is 27. More concerning this in
Chapter 11.
We start discussing the associated form of Chow and van der Waerden.
Let X ⊂ Pn = P(V ), V an (n + 1)-dimensional k-vector space, be a
projective variety all of whose irreducible components have dimension k (X
is purely k-dimensional). Intuitively, we want to describe such X by “coor-
dinates”, i.e. parametrize them by the points of another variety (we have
already accomplished this in cases of hypersurfaces or if X is a linear sub-
space). Write (Pn )∗ = P(V ∗ ). Look at the incidence correspondence
Γ := {(p, H1 , . . . , Hk+1 ) | p ∈ Hi ∀ i} ⊂ X × (Pn )∗ × · · · × (Pn )∗
where the Hi ⊂ Pn are hyperplanes; here we identify a hyperplane with its
defining equation in (Pn )∗ . Clearly, Γ is a closed subvariety of X × (Pn )∗ ×
· · · × (Pn )∗ equipped with two projections
Γ
ψ ϕ
$
X ((Pn )∗ )k+1 .
Clearly, ψ(Γ) = X, and for p ∈ X the set of hyperplanes containing p
forms an (n − 1)-dimensional projective linear subspace Hp ' Pn−1 ⊂ (Pn )∗ .
Thus ψ −1 (p) is irreducible of dimension (k + 1)(n − 1), and isomorphic to
Pn−1 × · · · × Pn−1 (k + 1 factors). By Theorems Theorem 8.16 and 8.18 Γ is
purely k + (k + 1)(n − 1) = (k + 1)n − 1 dimensional, with one irreducible
component lying over each irreducible component of X. Moreover, there
exist points
y = (H1 , . . . , Hk+1 ) ∈ ((Pn )∗ )k+1
such that ϕ(Γ) 3 y and ϕ−1 (y) is a single point: by the same construction
as the one employed in the proof of Corollary 8.8, we can find a chain
X (0) = X ) X (1) = X ∩ H1 ) · · · ) X (k+1) = X ∩ Hk+1 = {p}
73
\ (l)
X ∩ Hi = {x(1) , . . . , x(c) }, x(l) = (u0 : · · · : u(l)
n ), l = 1, . . . , c
i
|X ∩ H1 ∩ · · · ∩ Hk | = d.
(0) (0)
Proof. The sets of (H1 , . . . , Hk ) for which FX (v0 , . . . , vn ; H1 , . . . , Hk )
1. does not vanish identically
(0) (0)
FX (v0 , . . . , vn ; H1 , . . . , Hk ) must vanish and this defines a closed subvariety
of (Pn )k .
Now in the course of the proof of Theorem 9.10 we saw that the set in (2)
is nonempty. It is Zariski open since the subset of homogeneous polynomials
F ∈ k[t0 , . . . , tn ]d with a multiple factor is closed since the multiplication
map
is a morphism and we can apply Theorem 6.2 to conclude that the image is
closed.
Now being in set (1) means that H1 , . . . , Hk intersect X in finitely many
points, and being in (2) means that their number is d.
0) (0) (k) (k)
The set of all forms F (v0 , . . . , vn ; . . . ; v0 , . . . , vn ) of degree d in each
set of variables forms a projective space PN and by Remark 9.7 we get an
injection
Zk = m1 X1 + · · · + mp Xp
In this way we obtain exactly the closure C n,k,d of Cn,k,d in PN . Then C n,k,d
is called the Chow variety of algebraic cycles of dimension k and degree d
77
Let
g = gd + gd−1 + · · · + g0
be the decomposition into homogeneous components, gd 6= 0. Then the
condition g(x + λy) = 0 can be written as
n
!
d
X ∂gd
gd (y)λ + (y)xi + gd−1 (y) λd−1 + · · · = 0. (9.1)
i=1
∂Xi
2. and such that G̃(λ) = g(x + λy) has degree d (O.K. by (1)) and is
coprime to
n
0
X ∂g
G̃ (λ) = (x + λy)yi .
i=1
∂X i
If we accomplish (2) the proof is complete since then G̃(λ) has d simple roots.
Let Y1 , . . . , Yn , U be further indeterminates. Expand
ϕ := g(X1 + U Y1 , . . . , Xn + U Yn )
in powers of U :
n
!
X ∂g
ϕ = g(X1 , . . . , Xn )+ (X1 , . . . , Xn ) · Yi ·U +· · ·+gd (Y1 , . . . , Yn )U d .
i=1
∂Xi
Then
ψ(Y1 , . . . , Yn , U ) := g(x1 + U Y1 , . . . , xn + U Yn )
is irreducible in k[Y1 , . . . , Yn , U ]. To see this, assume by contradiction ψ =
h1 h2 is a nontrivial decomposition. There are two possibilities now:
79
(I) At least one of the variables Y1 , . . . , Yn occurs in h1 and h2 (it need not
be the same variable though). Then, for suitable u ∈ k ∗ , there would
be a nontrivial decomposition of
Then
G̃(U ) = ψ(y1 , . . . , yn , U ) = g(x1 + y1 U, . . . , xn + yn U )
is of degree d since gd (y) 6= 0 and coprime to
∂ψ
G̃0 (U ) = (y1 , . . . , yn , U )
∂U
since δ(y) 6= 0. Hence l = {x + λy | λ ∈ k} is a line with the desired
properties.
80 CHAPTER 9. LINES ON SURFACES, DEGREE
One vexing point in the theory of Cayley forms as we have been developing
it so far is that we do not know how to compute them explicitly! We can
now say what happens in the case of a hypersurface.
Corollary 9.15. Under the hypotheses of Theorem 9.14 the Cayley form of
X = {G(u0 , . . . , un ) = 0}
is
(0) (n−1)
FX (v0 , . . . , vn(0) ; . . . ; v0 , . . . , vn(n−1) )
= G(∆0 , . . . , ∆n )
(∆0 : · · · : ∆n ).
Put Y
f (λ) := gcd(F1 (λa), . . . , Fm (λa)) = (λ − αi )ki .
i
81
82CHAPTER 10. REGULAR AND SINGULAR POINTS, TANGENT SPACE
I(X) = (Φ1 , . . . , ΦM )
(0) (0) (0)
and p = [µ0 p0 + µ1 p1 ], then multx (X ∩ L) is the highest power of (µ1 µ0 −
(0)
µ0 µ1 ) which divides all Φ’s.
multx (X ∩ L) ≥ 2.
Thus λ2 divides Fi (λa) if and only if Li (a) = 0 for all i; thus L0 is tangent
to X0 in 0 if and only
L1 (a) = · · · = Lm (a) = 0.
Definition 10.2. Let us make An ' U0 into a vector space with origin x = 0
by componentwise addition and scalar multiplication. The sub-vector space
of all points a ∈ An which lie on tangents to X in x is called the (embedded,
affine) tangent space of X in x, denoted by Tx X. Its closure in Pn (i.e. the
locus of points in Pn which lie on tangents L to X in x) is called (embedded,
projective) tangent space, denoted by Tx X.
At the moment it is not a priori clear that X is equal to the union of its
regular and singular points, or that the singular points form a proper closed
subvariety. We will see this later.
83
and
∂ F̃ ∂F (y1 + x1 , . . . , yn + xn ) ∂F (t1 , . . . , tn )
(0) = (0) = (x)
∂yi ∂yi ∂tn
whence n
X ∂F
L= (x)(ti − xi ),
i=1
∂ti
so the formula is valid in the hypersurface case.
In the general case, according to Definition 10.1, we have
\
Tx X = Tx H
H⊃X hypersurface
f (t1 , . . . , tn ) = F (1, t1 , . . . , tn )
This settles the hypersurface case, and the general case follows again by
remarking that \
Tx X = Tx H.
H⊃X hypersurface
85
Remark 10.6. One must be careful that in Theorem 10.5 one has to take the
Fα (or Gβ ) as ideal generators, i.e. defining equations that do not generate
T the
respective ideals may lead to a different answer. In other words, ifTX = i Hi
where the Hi ⊂ An are hypersurfaces, then in general Tx X 6= Tx Hi . To
see this, take for example
H1 = {y − x2 = 0}, H2 = {y = 0} in A2 .
dx (F + G) = dx F + dx G, (10.1)
dx (F G) = F (x)dx G + G(x)dx F. (10.2)
Tx X = {dx F = 0 | F ∈ I(X)}.
dx g := dx G |Tx X ∈ Tx X ∗
In just the same way as above one can show that this is well-defined, sat-
isfies the computational rules (10.1) and (10.2), and gives a k-linear map
∗
dx : mx /m2x → TX,x . The same proof that we used for Theorem 10.8 then also
shows that it is an isomorphism.
Definition 10.11. The vector space (mX,x /m2X,x )∗ is called the Zariski tan-
gent space to X in x. Moreover, for ϕ : X → Y a morphism as in the proof
of Corollary 10.10, the linear map dx ϕ is called the differential of ϕ in x.
dx (ψ ◦ ϕ) = dx ψ ◦ dx ϕ, dx (idX ) = idTx X .
X = {x = y = 0} ∪ {x = z = 0} ∪ {y = z = 0}
with I(X) = (xy, yz, xz) is not isomorphic to any subvariety of A2 (in par-
ticular not isomorphic to the union of three lines through a point in A2 ).
Proof. Indeed,
dim T0 X = 3.
We now want to study the loci of regular and singular points in a variety
X.
T := {(x, a) ∈ X × An | a ∈ Tx X} ⊂ X × An .
π
X
If we put U := ϕ−1 (ψ −1 (W1 )), V := ψ −1 (ϕ−1 (W2 )), then U ' V via ϕ and
ψ.
As a corollary of Theorem 10.5 and Theorem 10.13 we get
1. There are at most three lines passing through a point p ∈ S. If they are
three, they lie in a plane.
91
92 CHAPTER 11. CUBIC SURFACES AND THEIR LINES
f = z 2 L(x, y, z, w) + wQ(x, y, z, w)
f = Ax2 + Bxy + Cy 2 + Dx + Ey + F
where A, B, C ∈ k[z, w]1 , D, E ∈ k[z, w]2 , F ∈ k[z, w]3 . This defines a conic
with coefficients in k[z, w]. It is singular if and only if the discriminant
f |E = w(A(λ, 1)x2 +B(λ, 1)xy+C(λ, 1)y 2 +D(λ, 1)wx+E(λ, 1)wy+F (λ, 1)w2 ).
93
ϕ : S 99K l × m ' P1 × P1 ,
ψ : l × m ' P1 × P1 99K S
2. or the li do not all lie on a quadric and there are exactly one or two
lines which intersect l1 , . . . , l4 .
Proof. It is easy to see as an exercise that l1 , l2 , l3 always lie on a nonsingular
quadric Q ⊂ P3 . In suitable coordinates, Q = {xw − yz =} ' P1 × P1 . Every
line which intersects all of l1 , l2 , l3 lies on Q. If l4 does not lie on Q, then it
has 1 or 2 intersection points with Q. The lines intersecting l1 , . . . , l4 in this
case are those of the ruling on Q to which l1 , l2 , l3 do not belong and which
pass through l4 ∩ Q.
If l4 ⊂ Q, then all of l1 , . . . , l4 belong to one ruling since they are disjoint,
and then the lines of the opposite ruling give infinitely many lines intersecting
all four.
Now let l be a line on S, and let (li , li0 ), i = 1, . . . , 5, be as in Theorem
11.2. Every other line n ⊂ S intersects exactly one of the lines li , li0 for all
i = 1, . . . , 5: namely, n intersects Ei = hli , li0 i, and Ei ∩ S = l ∪ li ∪ li0 . Because
of Proposition 11.1 (1), n cannot intersect both li and li0 .
95
in total
5
1+1+5+5+5+ = 27.
3
We can summarize what we have seen so far in
Theorem 11.6. Every nonsingular cubic surface S ⊂ P3 contains exactly
27 lines {l, m, li , li0 , li00 , lijk }, i ∈ {1, . . . , 5}, i, j, k ∈ {1, . . . , 5} with i < j < k,
as above. The incidence graph Γ of those, whose vertices are lines and edges
denote intersections, can be described as follows:
1. l intersects l1 , . . . , l5 , l10 , . . . , l50 ;
2. l1 intersects l, m, l10 , l100 and l1jk for 6 choices of {j, k} ⊂ {2, 3, 4, 5}.
whenever l0 , l00 , l000 and m0 , m00 , m000 are coplanar lines, i.e. they form triangles
on S.
Proposition 11.8. The module A(S) is free of rank 7, i.e., a lattice A(S) '
Z7 . As a basis we can choose l1 , . . . , l4 , l50 , l500 , l5 .
Proof. The triangles l+li +li0 , i = 1, . . . , 5, contain l, thus l+li +li0 = l+l5 +l50
for i = 1, . . . , 5, and we get
and thus
Then (11.1), (11.2) and (11.3) show that the elements in the statement of
Proposition 11.8 generate. That they are independent will be shown in
Proposition 11.10 below.
Before we can finish the proof of Proposition 11.8 we need a further
definition.
Definition 11.9. Define a bilinear pairing
A(S) × A(S) → Z
3. we put l(m+m0 +m00 ) = 1 for every line l and every triangle m+m0 +m00 .
This is well-defined by (1) and (2) alone: A(S) is generated by lines,
and we know that if l is distinct from m, m0 , m00 , it intersects exactly one of
m, m0 , m00 , hence (3) follows; if on the other hand l = m, say, then lm0 =
lm00 = 1 and l2 = lm = −1 by (2), hence (3) follows as well. So the bilinear
pairing descends to the quotient and is well-defined.
98 CHAPTER 11. CUBIC SURFACES AND THEIR LINES
Proposition 11.10. If
e0 = l5 + l50 + l500 , e1 = l1 , e2 = l2 , e3 = l3 , e4 = l4 ,
e5 = l50 , e6 = l500
then
e20 = 1, e2i = −1, i = 1, . . . , 6, ei ej = 0 i 6= j.
In particular, e0 , . . . , e6 are a basis of A(S).
Proof. Remark that e1 = l1 , . . . , e4 = l4 , e5 = l50 , e6 = l500 are six disjoint lines,
and the first four are also disjoint from l5 ; this implies everything except
e20 = 1. To see this remark that
to do directly by hand; it is easy to see once one knows adjunction and the
genus formula. This gives a symmetric description of the lines.
We also mention without proof that there is a striking connection to the
root system of type E6 : namely, h⊥ ⊂ A(S) is the (negative of the) root
lattice, (−E6 ).
100 CHAPTER 11. CUBIC SURFACES AND THEIR LINES
Chapter 12
Tx Hi ⊂ {v ∈ Tx U | dx Pi (v) = 0}.
dim Tx Hi ≥ dimx Hi ≥ n − 1,
101
102CHAPTER 12. LOCAL PARAMETERS, POWER SERIES METHODS, DIVISORS
Proof. The module M = mx over OX,x is finite. Now the residue classes of
the pi in mx /m2x generate that vector space, hence the submodule M 0 ⊂ M
generated by the pi ’s satisfies M 0 + mx M = M , in other words
m · (M/M 0 ) = (M/M 0 ).
Fi , i = 0, 1, 2, . . . , Fi ∈ k[t1 , . . . , tn ]i
103
with
k
X
f− Fi (p1 , . . . , pn ) ∈ mk+1
x , ∀ k ≥ 0.
i=0
F = F0 + F1 + F2 + . . .
The leading term of a powers series F , lt(F ) is Fi0 where i0 = min{i | Fi 6= 0}.
k
X
f− Fi (p1 , . . . , pn ) ∈ mk+1
x , ∀ k ≥ 0.
i=0
We have seen above that every f ∈ OX,x has at least one Taylor series.
Let IY,x ⊂ OX,x be the ideal of germs of functions in OX,x which vanish
on Y in an open neighborhood of x; if X is affine we have
n u o
IY,x = f = | u, v ∈ OX (X), u ∈ I(Y ), v(x) 6= 0 .
v
Theorem 12.12. The elements f1 , . . . , fm ∈ OX,x are called local equations
for a subvariety Y ⊂ X ⇐⇒ IY,x = (f1 , . . . , fm ).
f = (f0 : · · · : fn )
where fi ∈ k(X), fi ∈ OX,x for all i (we can clear denominators because the
target is a projective space) and the fi without common factor in OX,x (we
can divide by any common factor since the target is projective). Then no
irreducible codimension 1 subvariety Y can be contained in f0 = · · · = fn = 0
since by Theorem 12.13 we would have IY,x = (g) and if fi vanishes on (g = 0),
then g | fi for all i, a contradiction.
The following are immediate consequences of Theorem 12.14.
vY (f ) := vY (h1 ) − vY (h2 )
Definition 12.17. For smooth irreducible projective X let Div(X) the free
abelian group on all codimension 1 irreducible subvarieties. We call Div(X)
the group of (Weil) divisors on X. Any element in it is called a (Weil)
divisor. For f ∈ k(X)∗ we put
X
div(f ) := vY (f )Y ∈ Div(X)
Y ⊂X
where the sum runs over all irreducible codimension 1 subvarieties. A divisor
of the form div(f ) (or the zero divisor 0) is called a principal divisor. The
principal divisors form a subgroup PrincDiv(X) ⊂ Div(X). The quotient
Pic(X) := Div(X)/PrincDiv(X)
is called the Picard group (or more accurately, Weil divisor class group) of
X.
Example 12.19. Using the Segre embedding, one can show that
Pic(Pn × Pm ) = Z ⊕ Z.
By these simple examples one should not be tempted to think that Picard
groups are always finitely generated or torsion-free: this fails in very simple
examples already, like that of plane cubic curves.
Instead of with codimension 1 subvarieties, leading to Pic(X), one can
work with higher-codimensional subvarieties as well, leading to the so-called
Chow groups of X. But these are generally much harder to compute.
We also point out that the power series methods we have started to de-
velop in this Chapter have much more far reaching consequences and develop-
ments, in particular in Zariski’s theory of formal functions and deformation
theory.
Bibliography
[ZS60] Zariski, O. & Samuel, P., Commutative algebra, Vol. II, Van Nos-
trand, New York (1960)
111