Snok PDF
Snok PDF
1
Contents
1 Introduction: Plane Curves 3
2 Algebraic Sets in k n 9
4 Affine Varieties 29
5 Prevarieties 42
6 Algebraic Varieties 50
6.1 Digression: Complex Analytic Spaces . . . . . . . . . . 54
7 Projective Varieties 56
8 Dimension 64
10 Normal Varieties 82
11 Regular Points 87
2
1 Introduction: Plane Curves
In order to give an idea what algebraic geometry is about we discuss in the
introductory section plane curves in a not too rigorous manner:
C = V (f ) := {(x, y) ∈ k 2 ; f (x, y) = 0}
Only the unit circle seems to be an ”honest plane curve”. In order to avoid
points to be plane curves, we shall only consider an algebraically closed field
k. Nevertheless it turns out to be very useful to look at pictures in R2 ,
namely given f ∈ R[X, Y ] ⊂ C[X, Y ] one has V (f ) ⊂ C2 and draws its ”real
trace” V (f ) ∩ R2 . Only one has to understand, which features are due to
the fact that R is not algebraically closed and which ones really reflect the
complex situation.
1. V (f ) ∼
= k,
2. V (f ) ∼
= k ∗ := k \ {0},
3
3. V (f ) ∼
= (k × {0}) ∪ (k × {1}) ⊂ k 2 is the union of two parallel lines,
4. V (f ) ∼
= (k × {0}) ∪ ({0} × k) ⊂ k 2 is the coordinate cross.
f = f2 + f1
f = X 2 + aX + bY
f = XY + aX + bY
V (f ) ∩ ({−b} × k) = ∅
and
ab
t 7→ t − b, −a +
t
provides an isomorphism k ∗ −→ V (f ).
4
1. Every plane curve C ⊂ k 2 is the finite union
C = C1 ∪ ... ∪ Cr
k ⊃ U −→ C, t 7→ (g(t), h(t)),
0 = f (λ(1, t)) = λ2 t2 − λ2 (λ + a)
we find
ϕ(t) = (t2 − a, t(t2 − a)).
Note that for a 6= 0 and char(k) 6= 2 this is unfortunately not an injective
parametrization: The self intersection point (0, 0) has the inverse images ±b
where b2 = a. On the other hand, for a = 0 the above parametrization
becomes ϕ(t) = (t2 , t3 ), it is bijective, but not an isomorphism, since its
inverse can not be defined by the restriction to V (f ) ⊂ k 2 of a polynomial
function on k 2 . In that case
V (f ) = {(x, y); y 2 = x3 }
5
Definition 1.5. Let C = V (f ) ⊂ k 2 with a square free polynomial f . Then
a point (x, y) ∈ V (f ) is called a singular point if ∇f (x, y) = (0, 0). A curve
is called smooth if it has no singular points.
Example 1.6. 1. If C = C1 ∪ ... ∪ Cr with irreducible curves Ci then the
intersection points, i.e. the points in Ci ∩ Cj , j 6= i, are singular points.
2. Any curve has at most finitely many singular points.
3. The noose has the origin as its only singular point. If a 6= 0, it is a ”self
intersection point”: This concept may be defined algebraically, but here
we only give the definition for k = C. A singular point (b, c) ∈ C =
V (f ) (f being irreducible) is called a self intersection point, if there is
a neighbourhood U ⊂ C2 of (b, c) together with holomorphic functions
g, h : U −→ C with g(b, c) = 0 = h(b, c), such that f |U = gh and
V (g) 6= ∅ =
6 V (h). √ take U = {(x,
√ In the case of the noose y); |x| < |a|},
√
g(x, y) = y + x x + a, h(x, y) = y − x x + a, where .. : C \ R<0 is a
suitable branch of the square root.
4. A singular point of an irreducble curve which is not a self intersection
point, is called a cusp singularity, e.g. the origin is a cusp singularity
of Neil’s parabola.
5. Any nonsingular point of a complex plane curve has a neighbourhood
homeomorphic to a disc. This is an easy consequence of the holomor-
phic implicit function theorem.
The ”generic” version of our noose is the curve V (f ) with
f = Y 2 − p(X),
where p ∈ k[X] is a polynomial of degree 3 (resp. 2` + 1) with only simple
roots. In that case V (f ) is a smooth curve, it is called a plane elliptic
curve (resp. hyperelliptic curve for ` ≥ 2). For such curves parametrizations
U −→ V (f ) do not exist at all:
Theorem 1.7. Let C = V (f ) be an elliptic resp. hyperelliptic plane curve
over a field k of characteristic char(k) 6= 2. Then there is no nonconstant
map
(g, h) : U −→ V (f ),
where U ⊂ k is the complement of a finite set and g, h ∈ k(T ) := Q(k[T ])
are rational functions with poles outside U .
6
Proof. We may assume that g 0 6= 0 or h0 6= 0. (For char(k) = p > 0 we have
g 0 = 0 iff g = ĝ(T p ) with some ĝ ∈ k(T ). Then every g ∈ k(T ) can be written
r
g = g0 (T p ) with g00 6= 0. Now replace g with g0 . In order to see all that,
use the partial fraction decomposition of g.) Furthermore that U ⊂ k is the
maximal domain of definition for (g, h), i.e. that at a point a ∈ k \ U one of
the functions g, h has a pole. The equality
h2 = p(g)
shows that then the other function has a pole as well and that, if m, n are
the pole orders of g, h at a, we have 2n = (2` + 1)m. Formal differentiation
of the above equality gives
2hh0 = p0 (g)g 0 .
Let us now consider the rational function
2h0 g0
ϕ= 0 = ∈ k(T ).
p (g) h
Since p has only simple zeros, p0 (g) and h have no common zeros in U , in
particular the function ϕ is defined on U . We show that it can be extended to
k, hence is a polynomial. Consider a point a 6∈ U with m, n as above. Then
the function ϕ = g 0 /h has a removable singularity at a, since the pole order
m + 1 of g 0 is ≤ the pole order n of h: Indeed m + 1 ≤ n ⇐⇒ 1 ≤ (` − 21 )m,
the latter inequality holds, since m > 0 is even.
Finally we want to show that ϕ = 0 - this implies h0 = 0 = g 0 . We com-
pute the degree of ϕ. First of all we can define the degree of a rational func-
tion - the difference of the degrees of numerator and denominator polynomial.
Now denote m, n the degree of g, h. If m > 0, we have deg p(g) = (2` + 1)m
and once again 2n = (2` + 1)m. Since deg g 0 = m − 1 < m + 1 ≤ n, we
see that ϕ has negative degree, hence f = 0 and g 0 = 0. If m < 0, we have
deg p(g) = 0 and then deg h = 0 and deg g 0 ≤ m − 1 < 0, so deg(ϕ) < 0.
If m = 0 = n, we have deg(g 0 ) < 0 and obtain again deg(ϕ) < 0. Finally
in the case m = 0 > n, we have deg p(g) < 0 and thus deg p0 (g) = 0 - the
polynomials p and p0 don’t have common zeros - and are done even then since
deg h0 ≤ n − 1 < 0.
7
where Λ := C \ U is an infinite discrete set, indeed a lattice: A lattice Λ ⊂ C
is a subset of the form
Λ = Zω1 + Zω2 ,
where ω1 , ω2 are linearly independent over R, or, equivalently, ω1 /ω2 6∈ R. A
meromorphic function f on C is called Λ-periodic if f (z +ω) = f (z) holds for
all ω ∈ Λ. There are only constant holomorphic Λ-periodic functions: Such
a function would be globally bounded by its supremum on the parallelogram
{rω1 + sω2 ; 0 ≤ r, s ≤ 1}, and by Liouville’s theorem it follows that it is a
constant. But there are interesting meromorphic functions: The Weierstraß-
℘-function
1 X 1 1
℘(z) := 2 + − ,
z ω∈Λ∗
(z − ω)2 ω 2
where Λ∗ := Λ\{0}, is an example of a Λ-periodic function. It is holomorphic
outside Λ and has poles of order 2 at the lattice points. For
X 1 X 1
a=3 , b = 5
ω∈Λ
ω4 ω∈Λ
ω6
8
of the torus with the one point compactification of V (f ). Indeed, tradition-
ally one means the compactification V (f ) when speaking about an elliptic
curve. And it turns out that for every elliptic curve there is a corresponding
lattice such that we have a homeomorphism of the above type.
There is a further interesting remark: The homeomorphism V (f ) ∼ = C/Λ
endows on V (f ) the structure of an (additive) abelian group with the point
∞ as unit element. The group multiplication, it turns out, can also be
understood in a completely algebraic way: Take any line L ⊂ C2 and a
point a ∈ V (f ) ∩ L. We can define an intersection multiplicity: It is the
multiplicity of the polynomial function f (ϕ(t)) at 0 if ϕ : C −→ L is a
linear parametrization with ϕ(0) = a. Now counting the intersection points
in L ∩ V (f ) with multiplicities, there are at most 3 intersection points - two
for ”vertical” lines L = {a} × C, and three otherwise - and their sum in
V (f ) ∼
= C/Λ equals zero.
Hyperelliptic Curves: Now let us consider the case of a hyperelliptic curve:
The fact that the polynomial p(X) has odd degree guarantees, that the point
∞ ∈ V (f ) has a neighbourhood homeomorphic to a disc – if the degree is
even, it is a self intersection point, has a neighbourhood homeomorphic to
two copies of the unit disc E ⊂ C meeting at their centers, i.e. (E × {0}) ∪
({0} × E) ⊂ C2 with ∞ corresponding to (0, 0).
In that situation we have to replace C with the open unit disc E: There
is a surjective and locally homeomorphic parametrization
(g, h) : E −→ V (f )
given by two meromorphic functions, where the fibre of the point at infinity
consists of the points, where one of the functions g, h or both have a pole.
Indeed, there is a subgroup G ⊂ Aut(E), acting freely on E, such that the
parametrization is G-invariant and induces a homeomorphism
E/G ∼ = V (f ).
But for ` > 1 the extended curve V (f ) is not a group. Indeed V (f ) is
homeomorphic to a sphere with ` handles.
2 Algebraic Sets in k n
Algebraic geometry deals with objects, called ”algebraic varieties”, which
locally look like ”algebraic sets”, i.e. sets which can be described by a system
9
of polynomial equations. For this one has first of all to fix a base field:
In this section we define algebraic sets and prove that they correspond
to certain ideals in polynomial rings, the central result being ”Hilberts Null-
stellensatz”, Th.2.17.
To begin with let us mention the following fact: Given any infinite field
K the polynomial ring
K[T ] := K[T1 , ..., Tn ]
can be regarded as a subring
n
K[T ] ⊂ K K
n
of the ring K K of all functions K n −→ K.
Proof. For n = 1 this follows from the fact, that a polynomial f ∈ K[T ] has
at most deg f zeros. For n > 1 write
r
X
f= fi (T1 , ..., Tn−1 )Tni ∈ (K[T1 , ..., Tn−1 ]) [Tn ]
i=0
and assume f (x) = 0 for all x = (x0 , xn ) ∈ K n = K n−1 × K. Hence for any
x0 ∈ K n−1 we have 0 = f (x0 , Tn ) ∈ K[Tn ] resp. fi (x0 ) = 0 for all x0 ∈ K n−1 .
Using the induction hypothesis we conclude fi = 0 ∈ K[T1 , ..., Tn−1 ] for all
i = 0, ..., r and thus f = 0 ∈ K[T1 , ..., Tn ] as well.
X = N (F ) := N (k n ; F ) := {x ∈ k n ; f (x) = 0, ∀f ∈ F } .
10
Example 2.2. 1. Both k n = N (k n ; 0) and the empty set ∅ = N (k n ; 1)
are algebraic sets.
{a} = N (k n ; T1 − a1 , ..., Tn − an )
3. Arbitrary intersections
\ [
N (Fλ ) = N ( Fλ )
λ∈Λ λ∈Λ
N (F G) = N (F ) ∪ N (G)
4. All finite sets are algebraic; for n = 1 this are all the proper algebraic
sets.
11
Remark 2.3. 1. According to Ex.2.2.1,3 and 4, the algebraic sets satisfy
the axioms of the closed sets of a topology on k n . The corresponding
topology is called the Zariski topology. If we want to express that a
subset X ⊂ k n is Zariski-closed we also write simply X ,→ k n .
a contradiction!
12
Remark 2.5. The ideal
( r )
X X
aF := k[T ]f := gi fi ; gi ∈ k[T ], fi ∈ F, r ∈ N
f ∈F i=1
13
Example 2.8. A PID (:=Principal Ideal Domain) is obviously noetherian.
Theorem 2.9. The polynomial ring R[T ] in one variable over a noetherian
ring R is again noetherian.
Corollary 2.10 (Hilberts Basissatz). The polynomial ring K[T1 , ..., Tn ] over
a field K is noetherian.
Proof. K[T1 , ..., Tn ] = (K[T1 , ..., Tn−1 ])[Tn ].
Proof of 2.9. For an ideal a ⊂ R[T ] we define an ascending chain of ide-
als b0 ⊂ · · · ⊂ bj ⊂ . . . in R by
( )
X
bj := r ∈ R; ∃ rT j + ri T i ∈ a .
i<j
Prop. 2.4 now follows from Corollary 2.9 and the remarks 2.7 and 2.5.
14
Lemma 2.11. Let a, b ⊂ k[T ] and aλ ⊂ k[T ], λ ∈ Λ, be ideals. Then
2. a ⊂ b =⇒ N (a) ⊃ N (b).
P T
3. N ( λ∈Λ aλ ) = λ N (aλ ).
Proof. Exercise!
On the other hand any algebraic set X ,→ k n has an associated ideal, its
vanishing ideal I(X) ,→ k[T ], as well:
4. N (I(X)) = X;
5. I (N (a)) ⊃ a .
Namely:
15
2. For n = 2 and X1 := k×{0} = N (T2 ), X2 := N (T2 −T12 ) we obtain
X1 ∩X2 = {0} and I(X1 ∩X2 ) = (T1 , T2 ), while
The intuitive reason for this is, that the line X1 and the parabola X2
do not meet ”transversally” at their common point 0: Both f = T2
∂f
and f = T2 −T12 satisfy ∂T 1
(0) = 0 and hence this holds as well for all
2
functions f ∈ (T2 , T2 −T1 ).
16
For the proof of Hilberts Nullstellensatz Th.2.17 we need the following
description of maximal ideals m ⊂ k[T ].
Theorem 2.19 (Weak Nullstellensatz). For every maximal ideal m ⊂ k[T ]
there is a unique a ∈ k n with
17
For the inclusion ”⊃” we consider a function g ∈ I (N (a)) \ {0}. We
introduce an additional indeterminate S and consider in the ring k[T, S] :=
k[T1 , ..., Tn , S] the ideal
b := a[S] + (1−gS) ⊂ k[T, S] = (k[T ])[S]
P∞
with a[S] := i=0 a · S i . Its set of zeros is
N (k n ×k; b) = N (k n × k; a[S]) ∩ N (k n ×k; 1−gS)
with
N (k n × k; a[S]) = N (k n ; a) × k
and
N (k n ×k; 1−gS) = Γ1/g := (x, g(x)−1 ); x ∈ kgn ⊂ kgn ×k ⊂ k n × k,
18
We end the first section with some easy topological considerations which
in particular apply to algebraic sets, endowed with the Zariski topology:
Definition 2.23. The Zariski topology on an algebraic set X ,→ k n is the
topology induced by the Zariski topology of k n .
Definition 2.24. 1. A topological space X is called noetherian, if every
decreasing chain X1 ⊃ X2 ⊃ . . . of closed subsets terminates.
19
1. If X is irreducible, any non-empty open subset is dense.
20
3 Integral Ring Extensions
Though the immediate aim of this section is to prove Th.2.20, the notions
we use are very important also later on. So we give a more general presen-
tation, starting with the following class of rings, which is basic for algebraic
geometry:
A∼
= K[T1 , ..., Tn ]/a
The essential point now is to find for a given affine K-algebra A a subal-
gebra B ⊂ A (Th.3.10), such that
21
1. There is a monic polynomial p ∈ R [S] with p(s) = 0.
Proof of 3.4. ”1) =⇒ 2)”: Take an element f (s) ∈ R [s], where f ∈ R [S].
The euclidian algorithm for monic polynomials in R [S] provides aP
decompo-
sition f = qp+r with deg(r) < m := deg(p). Hence f (s) = r(s) ∈ m−1 i
i=0 Rs .
Thus
m−1
X
R [s] = Rsi .
i=0
22
Lemma 3.7. Let M be an R [s]-module with a generator system m1 , . . . , md
over R. For every i = 1, . . . , d we choose elements rij ∈ R, j = 1, . . . , d,
with
Xd
smi = rij mj .
j=1
23
2. If R ,→ R 0 and R 0 ,→ R 00 are integral ring extensions, so is R ,→ R 00 .
3. For every ring extension R ,→ R 0 the integral closure
b := {s ∈ R 0 ; s is integral over R}
R
of R in R 0 is a subring of R 0 .
4. Let R ,→ R 0 be an integral extension of integral domains. Then R is a
field iff R 0 is.
Proof. 1) The second statement follows from condition 3) in 3.4 with M =
R 0 . For the first there is a chain of integral ring extensions
R ,→ R [s1 ] ,→ R [s1 , s2 ] ,→ . . . ,→ R [s1 , . . . , st ] = R 0 ,
which are finite modules over its predecessors. Since that property is ”tran-
sitive” the claim follows. P
2) For c ∈ R 00 let cn + n−1 j
j=0 sj c = 0 be an integral equation over R .
0
24
Remark 3.9. An integral domain R is called integrally closed in its field of
fractions (or normal) if the ring extension Q(R) ⊃ R satisfies
Q(R) ⊃ R
b = R.
ϕn : K[T1 , . . . , Tn ] −→ A, Ti 7→ ti
25
Proof of Th.3.10. We do induction on n, the number of generators of the K-
algebra A. For n = 0 we have A = K, so there is nothing to prove. Now let
A∼ = K[T ]/a with T = (T1 , ..., Tn ). If a = 0, we are done. Otherwise we take
some polynomial f ∈ a \ {0}. Indeed, we may assume that f is Tn -monic:
Lemma 3.12. Let K be an infinite field. For any f ∈ K[T ] \ {0} there is
a linear automorphism F : K n −→ K n such that g := F ∗ (f ) is an ”almost
Tn -monic” polynomial
X
g = cTnm + gi (T 0 )Tni ,
i<m
26
Remark 3.13. We briefly indicate how the above proof has to be modified
if K is a finite field: In that case the pull back automorphism F ∗ : K[T ] −→
K[T ] in the proof of the auxiliary lemma 3.12 has to be replaced with σ :
K[T ] −→ K[T ], where
with exponents mi := bn−i , where b is bigger than all the exponents of in-
determinates Ti in the monomials of the polynomial f . Then the maximal
Tn -power in X
σ(T α ) = Tnm + gi (T 0 )Tni
i<m
Pn
is m := m(α) = i=1 αi bn−i . Since different multi-exponents α give different
maximal Tn -exponents m(α), it follows that the maximal Tn -exponent in
σ(f ) is the maximal m(α) for a monomial T α appearing in f and thus σ(f )
is Tn -monic.
Finally, for use later on, we prove the following important result in the
theory of integral extensions:
Theorem 3.14. Let R ,→ R 0 be an integral ring extension. Then
1. given a prime ideal p in R, there is a prime ideal p0 ,→ R 0 lying above
p, i.e. p0 ∩R = p.
R ,→ R0
↓ 0 0
↓
0
.
R (R ∩ m ) ,→ R /m
27
For general R and p ⊂ R we consider the multiplicative system S := R \ p.
Then m := S −1 p is the only maximal ideal in the local ring S −1 R, and the
ring extension S −1 R ,→ S −1 R 0 is integral. The first part provides a maximal
ideal m0 ,→ S −1 R 0 over m. Thus we obtain a commutative diagram
ι
p ,→ R ,→ R0 ←- p0
↓ ↓ ↓ .
S −1 ι
m ,→ S −1 R ,→ S −1 R 0 ←- m0
28
4 Affine Varieties
In this section we treat all the algebraic subsets X ,→ k n (with the ”embed-
ding dimension” n = n(X) depending on X) simultaneously as objects of a
category T A (no standard terminology!), thereby getting rid of any specific
inclusion into some affine space. From a given embedding X ,→ k n we only
keep the information about the functions obtained as restrictions of a poly-
nomial on k n . These regular functions provide the morphisms X −→ k, and
they are associated to the topological space X as an additional datum.
A first observation is
Proposition 4.2. The regular functions on an algebraic set X ,→ k n form
a k-algebra O(X). Indeed there is an exact sequence
X %
0 −→ I(X) −→ k[T ] −→ O(X) −→ 0
Proof. Exercise!
Example 4.3. Here we shall represent the ring O(X) for some examples
either as subalgebra or as extension of some polynomial ring.
1. The Neil parabola X := N (k 2 ; T22 −T13 ) admits a polynomial parametriza-
tion ϕ : k −→ X, s 7→ (s2 , s3 ), hence there is an induced pull back
map ϕ∗ : O(X) −→ k[S], an injection indeed, the map ϕ being onto
(or rather bijective!). Composing it with the restriction homomor-
phism %X : k[T1 , T2 ] −→ O(X) leads to the homomorphism k[T ] −→
k[S], T1 7→ S 2 , T2 7→ S 3 , whence
∞
O(X) ∼
= k[S 2 , S 3 ] ∼
M
=k·1⊕ k · S i.
i=2
29
2. For char(k) 6= 2 the parametrization
It induces an isomorphism
O(X) ∼
= ϕ∗ (O(X)) = {(f1 , f2 ) ∈ k[S1 ] ⊕ k[S2 ]; f1 (0) = f2 (0)} .
Here k[S1 ] ⊕ k[S2 ] denotes the direct sum of the two polynomial rings
k[Si ] (in one variable Si !), the ring operations are componentwise.
With Hilberts Nullstellensatz Th. 2.17 we can characterize all the alge-
bras O(X) up to isomorphy:
A∼
= O(X),
30
Definition 4.5. A ring R is called reduced, if it does not admit non-zero
nilpotent elements, i.e. if √
0=0
holds in R.
√
Proof of 4.4. An affine algebra A ∼
= k[T ]/a is reduced iff a = a, the latter
being equivalent to a = I(X) for X = N (a).
1. The objects are the pairs (X, A), where X is a noetherian topological
space and A ⊂ C(X) a subalgebra of the algebra C(X) of all k-valued
continuous functions on X (where k is endowed with the Zariski topol-
ogy, so a function f : X −→ k is continuous iff all the level sets
f −1 (c), c ∈ k, are closed.) satisfying in addition
A∗ = {f ∈ A; N (f ) = ∅},
ϕ : X −→ Y ,
ϕ∗ : C(Y ) −→ C(X), f 7→ f ◦ϕ ,
ϕ∗ (B) ⊂ A .
Then the pair (X, O(X)) belongs to this category for any algebraic set
X ,→ k n as a consequence of
31
Remark 4.7. 1. Any maximal ideal m ⊂ O(X) is of the form
m = ma := {f ∈ O(X); f (a) = 0}
32
with the ideal
O(V ) ∼
= S(V ∗ )
33
So it seems, all information about an affine variety X is contained in its
algebra O(X) of regular functions. Indeed:
Theorem 4.12. Denote FRA the category of affine reduced k-algebras. The
functor
O : AV −→ FRA
given by
X 7→ O(X), ϕ 7→ ϕ∗
defines an anti-equivalence of categories.
Sp : FRA −→ AV
X −→ Sp(A), x 7→ mx
is a bijection and hence can be used to define a topology and regular functions
on Sp(A). Indeed these data depend only on A: The closed sets are the sets
N (b) := {m ∈ Sp(A); b ⊂ m}
34
f ∈ A provides a continuous function fb : Sp(A) → k, m 7→ ı−1 (f +m). If we
then declare these functions to be regular:
an isomorphism
∼
=
ΦA : A −→ O(Sp(A)), f 7→ fb.
A comment on notation: Though the elements in the set Sp(A) are ideals,
one usually prefers a geometric notation and denotes x ∈ Sp(A) its points,
such that, strictly speaking, x = mx .
Finally, given an algebra homomorphism σ : B → A between reduced
affine algebras B and A, the inverse image σ −1 (m) of a maximal ideal m ,→ A
is again maximal, since k ,→ B/σ −1 (m) ,→ A/m ∼ = k implies B/σ −1 (m) ∼= k.
Even better, the map
satisfies obviously
gb ◦ Sp(σ) = σ(g)
d
for any g ∈ B and thus
as well as ∼
=
Ψ : idAV −→ Sp ◦ O.
The first one is defined for A ∈ FRA by the above
ΦA : A −→ O(Sp(A)), f 7→ fb,
35
while
ΨX : X −→ Sp(O(X)), x 7→ mx
is the natural equivalence over an affine variety X ∈ AV: We need to check
that given a morphism ϕ : X −→ Y , we have
ΨY ◦ ϕ = Sp(O(ϕ)) ◦ ΨX .
?
mϕ(x) = (ϕ∗ )−1 (mx ),
1. dominant, if Y = ϕ(X);
∼
=
2. a closed embedding, if it admits a factorization X −→ Z ,→ Y , where
Z ,→ Y is a closed subvariety of Y ;
36
Proof. The first statement as well as the implication ”=⇒” of the second
statement is obvious. For the reversed one take Z := ϕ(X). Then X −→ Z is
an isomorphism, since the pull back homomorphism O(Z) −→ O(X) is. For
the third part take a point y ∈ Y . Since ϕ∗ is injective and O(X) ⊃ O(Y )
an integral extension, we may apply Th.3.14 and obtain a maximal ideal
m0 ,→ O(X) withP m0 ∩ O(Y ) = my . But m0 = mx for some x ∈ X and thus
y = ϕ(x). If A = ri=1 Bai , then A/my · A is generated by the residue classes
ai + my · A over B/my ∼ = k, the same holds for O(ϕ−1 (y)) ∼ = A/I(ϕ−1 (y)),
a factor ring of A/my · A. So O(ϕ−1 (y)) is a finite dimensional k-vector
space of a dimension ≤ r, hence the affine variety ϕ−1 (y) is finite with at
most r points. Finally let ϕ : X −→ Y be finite and Z ,→ X be a closed
subset. Then Z −→ ϕ(Z) is a finite dominant map, hence surjective and
thus ϕ(Z) = ϕ(Z).
Remark 4.16. Let us explain for k = C the geometry behind the notion of
a finite map:
37
3. The most suggestive characterization of properness is as follows: A
morphism ϕ : X −→ Y between complex affine varieties is proper if
and only if a point sequence (xn ) ⊂ X has points of accumulation
in X iff the sequence (ϕ(xn )) ⊂ Y has in Y . The implication ”=⇒”
makes use of the continuity of ϕ, while for ”⇐=” consider the compact
set K consisting of the points yn and the accumulation points of that
sequence. One could thus say, that ϕ : X −→ Y is proper iff the space
X has ”no holes over Y ”.
O(Z) := O(X)|Z ∼
= O(X)/I(Z).
But what can we do in order to give open subsets U ⊂ X also some structure,
say, to make them objects in the category T A? We have to look for a good
notion of regular functions on U . We could once again try with the following
algebra
O(U ) := O(X)|U .
But there is some embarrassing fact: If U is dense in X - this is always the
case for irreducible X and non-empty U - we get
O(X)|U ∼
= O(X).
In particular for the subalgebra O(X)|U ⊂ C(U ) it need not be true that
functions without zeros are invertible. So we have to enlarge that algebra.
For certain open subsets that is easily done:
Definition 4.17. Let X be an affine variety.
1. A principal open subset U ⊂ X is an open subset U of the form
U = Xf := X \ N (f )
38
Remark 4.18. 1. Any Zariski open set U ⊂ X is a finite union S of princi-
pal open sets: We can write X \ U = N (f1 , ..., fr ), hence U = ri=1 Xfi .
O(U ) := O(X)f .
39
So a principal open set U ⊂ X defines an object (U, O(U )) ∈ T A. Indeed,
it is again an affine variety:
Proposition 4.21. For a principal open subset Xf ⊂ X of an affine variety
X we have
Xf ∼= Sp(O(X)f ).
X × Y ,→ k m × k n = k m+n .
Γ1/f := N (X × k; 1 − f S) ,→ X × k ∼
= Sp(A[S]),
O(Xf ) ∼
= A[S]/(1 − f S).
40
Finally, as a preparation of the next section, we prove that a locally
regular function on an affine variety is regular:
Proof. Write h|Ui = hi /fimi with functions hi ∈ O(X). Then we have fimi h =
hi on Ui and fimi +1 h = fi hi even on all of X. Since the functions
Pr fimi +1 have
mi +1
no common zero, there are functions gi ∈ O(X) with i=1 gi fi = 1.
Then r r
X X
mi +1
h=( gi f i )h = gi fi hi ∈ O(X).
i=1 i=1
Proof. Exercise!
41
5 Prevarieties
Affine varieties are special objects in the category T A of topological spaces
with a distinguished algebra of regular functions. In order to define algebraic
varieties we have to replace T A with the category of ringed spaces where one
has not only a distinguished subalgebra O(X) on the entire space X, but for
every open set U ⊂ X. That leads to the notion of a sheaf of functions:
Definition 5.1. Let X be a topological space.
1. A sheaf F of (continuous) k-valued functions on X consists of a choice
of a subalgebra F(U ) ⊂ C(U ) for every open subset U ⊂ X such that
(calling functions ∈ F(U ) regular functions)
(a) regular functions restrict to regular functions: F(U )|V ⊂ F(V ),
whenever V ⊂ U are open sets, and
S
(b) locally regular functions are regular: If f ∈ C(U ) for U = i∈I Ui
and f |Ui ∈ F(Ui ) for all i ∈ I, then already f ∈ F(U ).
2. A (k-)structure sheaf on X is a sheaf F of (continuous) k-valued func-
tions on X, such that
F(U )∗ = C(U )∗ ∩ F(U )
for all U ⊂ X, i.e. regular functions without zeros are invertible.
Example 5.2. 1. F(U ) := C(U ) is a sheaf of functions, denoted simply
C.
2. L(U ) := {f : U −→ k locally constant} is called the sheaf of locally
constant functions.
3. Let X be an affine variety. For principal open subsets U ⊂ X we have
already defined O(U ). For arbitrary open U ⊂ X we set then
O(U ) := {f : U −→ k; f |V ∈ O(V ) for all principal open subsets V ⊂ U }
S
and obtain a structure sheaf on X: Let V ⊂ U = i∈I Ui be a principal
open subset and f : U −→ S k with f |Ui ∈ O(Ui ). Since V is quasi-
compact, we may refine V = i∈I V ∩ Ui by a finite covering V = V1 ∪
... ∪ Vr with principal open subsets Vj ⊂ V . Then we have f |Vj ∈ O(Vj )
for j = 1, ..., r and Prop.4.23 tells us that f |V ∈ O(V ). Since that
holds for any principal open subset V ⊂ U , we get f ∈ O(U ).
42
4. Take k = C and endow it with its strong topology. Then we define on
X = Cn (with the strong topology as well) two sheaves of functions H
and E, namely
H(U ) := {f : U −→ C holomorphic}
and
E(U ) := {f : U −→ C differentiable},
where a differentiable function means a C ∞ -function in the ”real sense”,
i.e. if we use C ∼
= R2 and Cn ∼ = R2n .
Note that for Zariski open U ⊂ Cn we have inclusions
43
2. Let us consider the Segre cone X := N (k 4 ; T1 T4 − T2 T3 ) ,→ k 4 . Take
Xi := Xgi with gi := Ti |X . Consider now the open set
U := X \ 0 × 0 × k 2 = X1 ∪ X2 .
ϕ∗ (OY (V )) ⊂ OX (ϕ−1 (V ))
44
Proof. Given an affine variety X ∈ AV define its structure sheaf OX as in
the above example 5.2.3. Now assume that ϕ : X −→ Y is a morphism of
affine varieties. We have to show that ϕ is also a morphism in RS k . For a
principal open subset V = Yg we obviously have U := ϕ−1 (V ) = Xϕ∗ (g) and
OY (V ) := {f : V −→ k locally extendible to X} .
45
Example 5.9. 1. Open subsets U ⊂ X of an affine variety X are preva-
rieties, U being a finite union of principal open subsets, but not nec-
essarily affine (they are called instead ”quasi-affine”): The prevariety
U := k n \ {0} ⊂ X = k n is not affine for n > 1.
OX (X) −→ B := OX (X)/I(Y ) −→ OY (Y )
respectively j : Y −→ X as
Y −→ Z := Sp(B) −→ X.
46
3. There is an open affine cover Y = si=1 Vi , such that Ui := ϕ−1 (Vi ) ⊂ X
S
is affine and ϕ∗ : O(Vi ) −→ O(Ui ) surjective for i = 1, ..., s.
Let us now come back to algebraic geometry and present some important
examples of prevarieties:
U ⊂ X open : ⇐⇒ π −1 (U ) ⊂ Z open.
OX (U ) := {f ∈ C(U ); π ∗ (f ) ∈ OZ (π −1 (U ))}.
Pn = Pn (k) = k n ∪ {∞L ; L ⊂ k n },
The numbers z0 , ..., zn are called the homogeneous coordinates of the point
[z0 , ..., zn ] ∈ Pn , they are unique up to a common nonzero multiple λ ∈ k ∗ . In
47
order to see that the k-ringed space Pn is a prevariety we consider the open
cover n
[
Pn = Ui
i=0
with the open subspaces
Ui := {[z]; zi 6= 0} ∼
= kn,
the latter isomorphism being
n z0 zi−1 zi+1 zn
Ui −→ k , [z] 7→ , ...., , , ..., .
zi zi zi zi
In particular we can identify
∼
=
k n −→ U0 , z = (z1 , ..., zn ) 7→ [1, z],
the point ∞L for L = kz being the point [0.z]. For n = 1 we thus get the
”projective line”
P1 = k ∪ {∞}
with ∞ := [0, 1]. For k = C the projective line P1 is (with respect to the
strong topology) homeomorphic to the 2-sphere S2 , the ”Riemann sphere” of
complex analysis. But note that Pn is not homeomorphic to S2n for n > 1.
We finish this section with the remark that the quotient map π : (k n+1 )∗ −→
Pn belongs to a particularly interesting class of morphisms:
Definition 5.12. A morphism π : E −→ X between prevarieties E S and X
is called a k -principal bundle if there is an open affine cover X = ri=1 Ui
∗
48
Example 5.13. The quotient map π : (k n+1S)∗ −→ Pn is a k ∗ -principal bun-
dle: Consider the standard open cover Pn = ni=0 Ui with the open subspaces
Ui := {[z]; zi 6= 0} ∼
= k n and define
49
6 Algebraic Varieties
In ”real” (i.e. non Zariski) topology reasonable spaces are Hausdorff, and
also in algebraic geometry one needs a corresponding condition: Reasonable
prevarieties, to be called later on simply (algebraic) varieties, should be ”sep-
arated”. But since even affine varieties, the local models for prevarieties, are
very far from being Hausdorff, we can not use literally the same definition
as in topology. First of all here is an example of the anomalies we want to
avoid:
X := Z/ ∼ ,
where
(x, i) ∼ (y, j) : ⇐⇒ x = y ∈ k ∗ or i = j.
Then
X = V0 ∪ V1 = V01 ∪ {o0 , o1 }
with the open subsets Vi := π(k × {i}) ∼ = k, so X is a prevariety, and
∼ ∗
V01 = V0 ∩ V1 = k . So X is an affine line with the two different origins
oi := π(0, i), i = 0, 1.
∆ := {(x.x); x ∈ X} ⊂ X × X
50
Proof. For affine X, Y we have already defined X × Y . In the general case we
have to endow the cartesian product X ×Y with theS structure of a prevariety.
In order to do so, we take open affine coverings X = ri=1 Ui and Y = sµ=1 Vµ
S
and ”patch together” the affine varieties Ui × Vµ : Consider the prevariety Z
defined as the disjoint union
[
Z := Ziµ ,
i,µ
51
2. Let X, Y be affine varieties, f ∈ O(X), g ∈ O(Y ). Then
(X × Y )f ⊗g ∼
= Xf × Yg .
Proof. Do the first part yourself! For the second part, note that both the
RHS and the LHS are affine varieties with the same underlying set, hence it
suffices to check that the regular function algebras agree.
To finish the proof of Prop 6.2 we have to check that the prevariety X ×Y
is a product of X and Y in the category RS k . That follows from the fact,
that given morphisms ϕ : Z −→ X and ψ : Z −→ Y , the map (ϕ, ψ) : Z −→
X × Y is a morphism as well, since the restrictions (ϕ, ψ)|(ϕ,ψ)−1 (Ui ×Vµ ) :
(ϕ, ψ)−1 (Ui × Vµ ) −→ Ui × Vµ are.
5. S
In the general case of a prevariety X with an affine open cover X =
r
i=1 Ui we thus get that
r
[
∆ ,→ Ui × Ui
i=1
52
Sr
i=1 Ui × Ui ⊂ X × X and that the diagonal map
δ : X −→ ∆, x 7→ (x, x),
δ ∗ : O(Ui × Uj ) −→ O(Uij )
Uij ∼
= k i × k ∗ × k j−i−1 × k ∗ × k n−j
53
as inclusions and thus
O(U ) = O(π −1 (U ))0 := {f ∈ O(π −1 (U )); f (λt) = f (t), ∀λ ∈ k ∗ }.
In particular we have
T0 Tn
O(Ui ) = k , ..., ,
Ti Ti
while
Tµ Tν
O(Uij ) = (k[T ]Ti Tj )0 = k ; 0 ≤ µ, ν ≤ n .
Ti Tj
Hence the homomorphism
T0 Tn T0 Tn Tµ Tν
k , ..., ⊗k , ..., −→ k ; 0 ≤ µ, ν ≤ n ,
Ti Ti Tj Tj Ti Tj
Tµ Tν Tµ Tν
⊗ 7→
Ti Tj Ti Tj
is surjective.
54
Definition 6.8. 1. A complex analytic space (or complex analytic variety)
S
X is a C-ringed Hausdorff space admitting an open cover X = i∈I Ui
with open subspaces Ui ∼
= Zi ,→ Wi , where Zi is an analytic subset of
ni
the open subset Wi ⊂ C .
S
2. A complex n-manifold is a complex analytic space X = i∈I Ui with
open subspaces Ui ∼
= Wi , the Wi being open subspaces Wi ⊂ Cn .
3. A Riemann surface is a connected complex 1-manifold.
55
A complex manifold is a complex analytic space. Here are some remarks
how far the latter are from the former:
Example 6.12. For the Neil parabola X := N (C2 ; T22 − T13 ) and the noose
Y := N (C2 ; T22 − T12 (T1 + 1)) the only singular point of Xh resp. Yh is the
origin. The affine curves Z := N (C2 ; T22 − p(T1 )) induces a Riemann surface
Zh .
7 Projective Varieties
Projective space Pn plays a rôle similar to that of affine space k n as an
ambient space for interesting algebraic varieties:
56
Remark 7.2. There are two ways to relate affine varieties to projective
varieties:
1. To every embedded affine variety
Y ,→ k n ∼
= U 0 ⊂ Pn
we can associate its projective closure
Y ,→ Pn = U0 ∪ P(0 × k n ).
C(Y ) = k ∗ · ({1} × Y ),
X 7→ C(X)
57
The counterparts to affine n-cones on the level of ideals are the homoge-
neous ideals:
Proof. The first part as well as the implication ”⇐=” of the second part
follow from the fact that a can be generated by homogeneous
P (q) polynomials.
Finally for an affine cone Z the assumption f = qf ∈ I(Z) implies
P q (q)
fλ = qλ f ∈ I(Z) for all λ ∈ k, where fλ (x) := f (λx). Evaluating
that equality at any point x ∈ Z gives f (q) (x) = 0 for all q ∈ N resp.
f (q) ∈ I(Z).
58
Corollary 7.7. Any projective variety X ,→ Pn can be described as the set
of zeros
Note that homogeneous polynomials are functions only on k n+1 and not
on Pn , but since such a polynomial either vanishes identically on a punctured
line k ∗ · t (t 6= 0) or has no zeros there, the above description of a projective
variety makes sense nevertheless.
satisfies
a) = C(Y ).
N (b
k ∗ · ({1} × Y )
59
as their common set of zeros in k ∗ × k n , and the same holds for the fˆ, f ∈ a.
Since fˆ ∈ k[T0 , ..., Tn ], it follows
k ∗ · ({1} × Y ) ⊂ N (b
a).
Take any g ∈ k[T0 , ..., Tn ] vanishing on k ∗ · ({1} × Y ). Since both ideals are
homogeneous, we may assume, that g is homogeneous (of degree q) and not
divisible by T0 - a factor T0 does not contribute to the zeros on k ∗ × k n . Then
g1 := g(1, T1 , ..., Tn ) vanishes on Y , so (g1 )` ∈ a for some ` > 0. But then
g = gb1 resp. g ∈ I(N (b a)).
Y ,→ k n ∼
= U0 ⊂ Pn .
X := Y ,→ Pn = U0 ∪ P(0 × k n ).
Indeed the projective closure yields a maximal extension, since, as we shall see
in a moment, projective varieties do not admit any further proper extensions:
They are ”complete”.
60
prZ : X × Z −→ Z is a closed map, i.e. maps closed sets to closed sets.
Remark 7.11. 1. In order to check completeness of an algebraic variety
X, it obviously suffices to considers affine ”test varieties” Z.
2. A closed subvariety Y ,→ X of a complete variety is complete, the
induced map Y × Z −→ X × Z being a closed embedding as well.
3. The image ϕ(X) ⊂ Y of a morphism ϕ : X −→ Y with a complete
source X is closed in Y and complete: The graph Γϕ ⊂ X × Y of any
morphism ϕ : X −→ Y is the inverse image
Y ,→ Pn × Z
↓ ↓
?
B ,→ Z
61
and want to see, why we are allowed to remove the question mark. For z ∈ Z
we consider the sectional variety
Yz := {y ∈ Pn ; (y, z) ∈ Y } ,
such that
B = {z ∈ Z; Yz 6= ∅}.
Denote
C(Y ) ,→ k n+1 × Z
the closure of
(π × idZ )−1 (Y ) ,→ (k n+1 \ {0}) × Z,
where π × idZ : (k n+1 \ {0}) × Z −→ Pn × Z. Indeed
[
(π × idZ )−1 (Y ) = (C(Yz )∗ × {z})
z∈B
and !
[
C(Y ) = C(Yz ) × {z} ∪ ({0} × B).
z∈B
We want to show B = B, i.e. the second term is not really needed. The
below lemma assures that V = Z \ B is open:
Lemma 7.13. Let Z = Sp(C) be an affine variety and X ,→ k n+1 × Z be a
relative cone, i.e.
V := {z ∈ Z; Xz ⊂ {0}}
is open.
Proof. We consider the vanishing ideal
a graded ideal
a = (f1 , ..., fr )
62
generated by homogeneous polynomials fi ∈ C[T ] of degree di as well as its
”sectional ideals”
az := {f (z, T ); f ∈ a} = (f1 (z, T ), ..., fr (z, T )) ,→ k[T ], z ∈ Z.
Then
z ∈ V ⇐⇒ N (az ) ⊂ {0}
and furthermore
N (az ) ⊂ {0} ⇐⇒ az ⊃ m` for some ` ∈ N
with the ideal
m := (T0 , ..., Tn ) ,→ k[T0 , ..., Tn ].
The implication ”⇐=” is clear, while for the opposite direction we first note
that Hilberts Nullenstellensatz provides an exponent s ∈ N with Tis ∈ az
for i = 0, ..., n. Now take ` := (n + 1)s: The ideal m` is generated by the
monomials T α with |α| = `, and any such monomial divisible by some Tis .
As a consequence we obtain that
[∞
V = V`
`=1
Thus we are done if we can show that V` ⊂ Z is open for any ` ∈ N. In order
to see that we consider the linear maps
M r
Fz : E := k[T ]`−di −→ L := k[T ]`
i=1
63
8 Dimension
A first basic invariant of an algebraic variety X is its dimension:
while
dim X := max dima X.
a∈X
64
8. An irreducible variety is a curve iff the proper closed subsets are the
finite sets.
9. dim k = 1 and dima k n ≥ n for all a ∈ k n : Take Xi := a + (k i × 0) with
the origin 0 ∈ k n−i .
dimb Y = max
−1
dima X,
a∈ϕ (b)
and in particular
dim X = dim Y.
Proof. First note that for any dominant morphism ϕ : X −→ Y and closed
Z ,→ X we have
I(ϕ(Z))) = I(Z) ∩ O(Y ),
where we treat ϕ∗ : O(Y ) −→ O(X) as inclusion.
”dimb Y ≥ maxa∈ϕ−1 (b) dima X”: Consider a strictly increasing sequence of
irreducible subvarieties X0 = {a} $ X1 $ ... $ Xn on X. Then the (closed)
images Yi := ϕ(Xi ) are as well irreducible closed subvarieties of Y and Yi $
Yi+1 : Otherwise we would have
65
dim X < ∞ for any affine variety X.
Proof. The second part of the statement follows from Prop. 8.3 together with
the Noether normalization theorem. The first part is obtained by induction,
the case n = 1 being obvious. Assume now X0 $ X1 $ ... $ Xm = k n is a
strictly increasing sequence of irreducible subvarieties in k n , n > 1. If we can
show dim Xm−1 < n, we have necessarily m − 1 ≤ n − 1 resp. m ≤ n: Since
Xm−1 6= k n , there is a polynomial f ∈ I(Xm−1 ) \ {0}. As a consequence of
the lemmata 3.12 and 8.3, we see that
dim Xm−1 ≤ dim N (k n ; f ) = n − 1.
66
Theorem 8.6 (Krulls Hauptidealsatz). Let X be an n-dimensional irre-
ducible algebraic variety. Then the zero locus N (f ) ,→ X of a regular func-
tion f ∈ O(X) \ {0} is empty or purely (n−1)-dimensional, i.e. all its
irreducible components Z ,→ N (f ) satisfy dim Z = n − 1.
ϕ|N (f ) : N (f ) −→ N (g).
of the field extension Q(A) ⊃ k(T ), see the below remark with B = K[T ] for
a short presentation. Since f |ϕ∗ (g) we have
N (f ) ⊂ ϕ−1 (N (g)) ,→ X
X ⊃ N (f ) ∪ ϕ−1 (V )
↓ ↓ ↓
k n ⊃ N (g) ∪ V
67
The norm function associated to a field extension: We consider a
finite field extension E ⊃ K fitting in a diagram
Q(A) = E ⊃ K = Q(B)
∪ ∪ ∪
A ⊃ B = Bb
where µa ∈ EndK (E) is the multiplication with a, i.e. µa (x) := ax. Then we
have
1. N(1) = 1,
3. N(A) ⊂ B,
4. N(A∗ ) ⊂ B ∗ , and
5. a|N(a) in A.
The first two statements are obvious, we comment on 3) and 5), while 4)
follows from the multiplicativity 2) and statement 3). We use:
χa = (pa )r ∈ B[S]
with r := [E : K[a]].
Since (−1)n N(a) = χa (0) (with n := [E : K]) is the constant term of the
characteristic polynomial χa and χa (a) = 0, that implies both 3) and 5).
68
Proof. The minimal polynomial pa ∈ K[T ] of an element a ∈ A ⊂ E lies in
B[S]: We may assume - by enlarging E if necessary - that pa splits into linear
factors over E. Then pa |f , if f ∈ B[S] is an integral equation for a ∈ A (i.e.,
a monic polynomial in B[S] with a as zero). As a consequence, all zeros of pa
are integral over R, hence its coefficients - being obtained from these zeros
by additions and multiplications - are both in K = Q(B) and integral over
B, but B is integrally closed in Q(B), so they are already in B.
Now let us consider the characteristic polynomial χa ∈ K[S] of µa ∈
EndK (E). The minimal polynomial pa of a is also the minimal polynomial
of µa , a divisor of the characteristic polynomial of µa . If E = K[a], both
polynomials are monic of the same degree, so they coincide. In the gen-
eral case, choosing a basis x1 , ...,
Lxr of E as K[a]-vector space provides a
µa -invariant decomposition E = di=1 K[a]xi , and the characteristic polyno-
mial of µa |K[a]xi is pa . So, altogether we obtain (pa )r as the characteristic
polynomial of µa .
Finally, in order to see that we may apply the above reasoning with
B = K[T ] we add
69
Proof. We may assume that X is affine, since for a ∈ U ⊂ X with an open
subspace U ⊂ X we have both dim U = dim X and dima U = dima X, and
use induction on dim X. The case dim X = 0 is trivial, since then X is
finite. Now assume dim X > 0. Take a regular function f ∈ ma \ {0} and let
Z ,→ N (f ) be an irreducible component of N (f ) ,→ X containing a. Then,
by induction hypothesis, dima Z = dim Z, and thus
dim X ≥ dima X > dima Z = dim Z = dim X − 1,
so necessarily dima X = dim X.
Corollary 8.10. 1. Let X := N (k n ; f1 , ..., fr ) ,→ k n be a subvariety de-
termined by r equations. Then either X is empty or every irreducible
component X0 ,→ X of X has dimension dim X0 ≥ n − r.
2. Let X := N (Pn ; f1 , ..., fr ) ,→ Pn be a projective variety determined by
r nonconstant homogeneous polynomials. If r ≤ n, then X is non-
empty and every irreducible component X0 ,→ X of X has dimension
dim X0 ≥ n − r.
Proof. 1.) The first part is proved by induction on the number r of equa-
tions, the case r = 0 being trivial. Now for r > 0 every irreducible com-
ponent X0 ,→ X is contained in an irreducible component Z0 of Z :=
N (k n ; f1 , ..., fr−1 ). By induction hypothesis we know that dim Z0 ≥ n−r +1,
while X0 is an irreducible component of N (Z0 ; fr ), so, according to 8.6, has
dimension at least n − r.
2.) With π : (k n+1 )∗ −→ Pn denoting the quotient morphism we have
X = π(N (k n+1 ; f1 , ..., fr ) \ {0}) 6= ∅,
since fi (0) = 0 for i = 1, ..., r and, according to the first part, all the irre-
ducible components of N (k n+1 ; f1 , ..., fr ) have dimension ≥ n + 1 − r ≥ 1.
Finally use that dim C(X) = dim X + 1, since
C(X)Ti ∼
= Xi × k ∗
holds for Xi := X ∩ Ui (with Ui := Pn \ N (Pn ; Ti )).
70
Simultaneously we investigate the sets ϕ(X) obtained as images of mor-
phisms. Here is an easy example showing some possible phenomena:
{(t1 , t−1
∗
1 t2 )} , if (t1 , t2 ) ∈ k × k
−1
ϕ (t1 , t2 ) = {0} × k , if (t1 , t2 ) = (0, 0)
∅ , otherwise
and thus
0 , if (t1 , t2 ) ∈ k ∗ × k
dim ϕ−1 (t1 , t2 ) = 1 , if (t1 , t2 ) = (0, 0) .
−∞ , otherwise
Note furthermore that ϕ(k 2 ) = V ∪ {(0, 0)} is neither open nor closed in
k 2 , not even locally closed and thus does not inherit in a natural way the
structure of an algebraic variety.
1. dim X ≥ dim Y ,
3. there is a nonempty open subset W ⊂ Y , such that all fibers ϕ−1 (y) of
points y ∈ W are pure-d-dimensional.
Proof. 2.) We may assume that Y is affine and take a surjective finite map
π : Y −→ k r (so r = dim Y ). Look at ψ := π ◦ ϕ : X −→ k r . Since
w.l.o.g. ψ(y) = 0, we see with the first part of Cor.8.10, that all irreducible
71
components of ψ −1 (0) ,→ X have dimension at least dim X − r. On the other
hand, if π −1 (0) = {y1 = y, y2 , ..., ys }, then
s
[
ψ −1 (0) = ϕ−1 (yj ),
j=1
72
where we even may assume Tj ∈ A for j = 1, ..., d. If the polynomials
qi ∈ (K[T ])[S], i = 1, ..., r, are integral equations over K[T ] for the (B-
algebra) generators ai ∈ A, we have qi ∈ (Bg [T ])[S] for a suitable
g ∈ B \ {0}: Take g as a common denominator P of the coefficients of the
polynomials fij ∈ K[T ], such that qi = S mi + j<mi fij S j . Then obviously,
Ag is a finite Bg [T ]-module, and we may take W = Yg , with the above
factorization corresponding to the algebra homomorphisms
Bg ,→ Bg [T ] ,→ Ag .
73
are closed:
Sm (ϕ) ,→ X.
Equivalently: The function X −→ N, x 7→ dimx ϕ−1 (ϕ(x)) is upper semicon-
tinuous.
74
r
\
= (X \ Ai ) ∪ (X \ Ui )
i=1
!
[ \ \
= (X \ Ai ) ∩ (X \ Ui ) ∈ C(X).
I∈P({1,...,r}) i∈I i6∈I
2. On any set X the sets which either are finite or cofinite (i.e. have
a finite complement) form an algebra of sets. From this and the
previous point it follows that for an irreducible curve constructible
sets are either finite or cofinite resp. either closed or open. If X
is a reducible
Sr curve, any constructible set is still locally closed: If
X = i=1 Xi is the decomposition into irreducible components, then
A ∈ C(X) ⇐⇒ Ai := A ∩ Xi ∈ C(Xi ) ∀i = 1, ..., r, so A is constructible
iff every Ai ⊂ Xi either is closed or open in Xi . Then it follows that
A is open in its closure A, which is the union of certain irreducible
components of X and finitely many isolated points.
ϕ(X) ∈ C(Y ).
Proof. We use induction on dim X, the case dim X = 0, i.e. |X| < ∞,
being obvious. Furthermore we may assume X to be irreducible and ϕ to be
dominant and hence Y is irreducible as well. Then, according to Prop. 8.12
there is a non-empty open subset W ⊂ ϕ(X), so ϕ(X) = W ∪ ϕ(Z) with
Z := Y \ ϕ−1 (W ). Since then dim Z < dim X the induction hypothesis gives
that ϕ(Z) is constructible, hence so is W ∪ ϕ(Z).
75
9 Rational functions and local rings
On a complete algebraic variety X there are only locally constant regular
functions; so they do not provide any information about the geometry of X.
Instead we have to allow ”poles” and ”points of indeterminacy”: We consider
pairs (U, f ), where U ⊂ X is a dense open subset and f ∈ O(U ) a regular
function on U . Two such pairs are identified using the equivalence relation
In order to see that this really provides an equivalence relation, note that a
finite intersection of dense open subsets is again a dense open subset.
Definition 9.1. A rational function on an algebraic variety X is an equiv-
alence class of pairs (U, f ) with a dense open subset U ⊂ X and f ∈ O(U ).
We denote R(X) the set of all rational functions on X.
Remark 9.2. 1. Since the restriction O(U ) −→ O(W ) of functions is
injective for dense open subsets W ⊂ U , a rational function represented
by (U, f ) has a unique representative (Umax , fmax ) with a maximal open
subset Umax ⊂ X: Set
[
Umax := V , fmax |V := g.
(V,g)∼(U,f )
2. The rational functions form a k-algebra, the sum f +g resp. the product
f g having a domain of definition containing the (dense) intersection
Df ∩ Dg .
R(X) ∼
= S −1 A
76
with the localization of A with respect to the multiplicative set
R(X) ∼
= Q(A),
if X is irreducible.
R(X) ∼
= Q(O(U ))
6. The morphism
f : Df −→ k
representing a nonzero rational function on an irreducible variety X
can be extended to a morphism
fb : U := Df ∪ D1/f −→ P1 = k ∪ {∞},
by defining
fb(x) := [(1/f )(x), 1] ∈ U1 ⊂ P1
for x ∈ D1/f . The points outside Df ∪ D1/f are called ”points of inde-
terminacy” of the rational function f . If X = Sp(A) is affine, a point
a ∈ X is a point of indeterminacy of f ∈ Q(A), iff g(a) = 0 = h(a) for
all representations of f as fraction f = g/h with g, h ∈ A.
For example take X := k 2 and f := T1 /T2 . Then Df = k × k ∗ and
D1/f = k ∗ × k, while the origin is the only point of indeterminacy.
Indeed the morphism fb : (k 2 )∗ −→ P1 turns out to be the quotient
morphism. In particular there is no extension to a morphism k 2 −→ P1 ,
since the restriction fb|k∗ t of fb to a punctured line k ∗ t can be extended
to the entire line kt only with 0 7→ [t], that value depending on the
chosen line.
In the sequel we shall use also the letter f in order to denote fb.
77
7. If X = ri=1 Xi is the decomposition of X into irreducible components,
S
then the component covering
r
a
π:X
e := Xi −→ X
i=1
induces an isomorphism
r
R(X) ∼
M
= R(Xi ),
i=1
S e := π −1 (U ) ∼
since, with U := X \ i6=j Xi ∩ Xj and U = U , we have
∼ ∼ ∼
R(X) = R(U ) = R(U ) = R(X).
e e
In particular R(X) is a finite direct sum of fields.
78
Proof. If K = k(a1 , ..., ar ), we have K = R(X) with X := Sp(k[a1 , ..., ar ]).
Assume now that σ : R(Y ) −→ R(X) is a k-algebra homomorphism. We
may replace X and Y with (nonempty) open affine subsets, i.e., we may
assume X = Sp(A), Y = Sp(B). If B = k[b1 , ..., bs ], choose a common
denominator g ∈ A for the elements σ(b1 ), ..., σ(bs ) ∈ Q(A). So σ(B) ⊂ Ag
and the morphism Xg −→ Y corresponding to the k-algebra homomorphism
σ|B : B −→ Ag defines the rational morphism X −→ Y we are looking for –
the easy proof of uniqueness is left to the reader.
1. R(X) ∼
= R(Y )
In that case we say also, that the varieties X and Y are birationally equiva-
lent.
79
Note that OX,Y is a subring, since any two nonempty open subsets of the
irreducible Y have nonempty intersection and thus
Df +g , Df g ⊃ Df ∩ Dg ←- (Df ∩ Y ) ∩ (Dg ∩ Y ) 6= ∅.
Furthermore:
Remark 9.8. 1. To justify its name, we remark that OX,Y is a local ring
with maximal ideal
OX,Y ∼
= S −1 A
5. The ring OX,Y is never an affine algebra: The only local reduced affine
algebra is the base field k.
80
Remark 9.10. 1. We can also describe OX,a as the direct limit
OX,a := lim OX (U ),
U 3a
taken with respect to the (by inclusion) partially ordered set of open
neighbourhoods of a. With other words the elements in OX.a are equiv-
alence classes of pairs (U, f ), where a ∈ U and f ∈ OX (U ) with
(U, f ) ∼ (V, g) iff f |W = g|W for some open neighbourhood W ⊂ U ∩ V
of the point a.
2. If k = C we can as well consider the stalk
Ha := lim H(U ),
U 3a
81
The proof is analogous to that of Th. 9.4 and thus left as an exercise to
the reader. In particular we have as well the following corollary:
Corollary 9.12. For points a ∈ X, b ∈ Y in the algebraic varieties X, Y the
following statements are equivalent:
1. OX,a ∼= OY,b as k-algebras
2. There are open neighborhoods U of a ∈ X and V of b ∈ Y , such that
U∼= V with an isomorphism of algebraic varieties sending a to b.
Remark 9.13. 1. The local ring OX,a is an integral domain iff a is con-
tained in exactly one irreducible component of X. So all the local rings
of an algebraic variety X are integral domains iff X is the disjoint union
of its irreducible components.
2. For an irreducible variety we have OX,a ⊂ R(X) for all a ∈ X and
\
O(X) = OX,a .
a∈X
10 Normal Varieties
Definition 10.1. A point a ∈ X in an algebraic variety X is called a normal
point, iff the local ring OX,a of X at a is an integral domain, which is integrally
closed in its field of fractions Q(OX,a ) = R(X). An algebraic variety X is
called normal iff all its points are normal.
So a normal variety is the disjoint union of its irreducible components.
Proposition 10.2. 1. An irreducible affine variety is normal iff its alge-
bra of regular functions A is integrally closed in its field of fractions
Q(A).
2. The normal points of an algebraic variety X form a dense open subset.
In the proof we use the following basic facts:
Remark 10.3. 1. If A is normal, i.e. integrally closed in its field of frac-
tions, so is S −1 A for any multiplicative subset S ⊂ A: If f ∈ S −1 A
has integral equation p(f ) = 0 with a monic polynomial p ∈ S −1 A[T ],
then hf has integral equation hn p(h−1 T ), n = deg p. But for a suitable
choice of h ∈ S that polynomial has coefficients in A, and A being
normal, we get hf ∈ A resp. f = hf /h ∈ S −1 A.
82
2. The integral closure A e ⊂ Q(A) of an affine k-algebra without zero
divisors in its field of fractions Q(A) is a finitely generated A-module,
thus in particular an affine algebra.
Proof. The first part follows from the fact, that a localization of a ring inte-
grally closed in its field of fractions enjoys the same property, as well as any
intersection – see Rem.9.13 – of such rings does.
For the second part consider a normal point a ∈ X. Since the irreducible
points of X, i.e. the points lying in only one of the irreducible components
of X, form a dense open subset, we may replace X with an irreducible affine
open neighborhood of a resp. assume that X = Sp(A) is irreducible and
affine. Then we have A e ⊂ OX,a for the integral closure A
e of A in its field of
fractions Q(A). But on the other hand it is a finitely generated A-module,
say Ae = Af1 + ... + Afr with f1 , ..., fr ∈ Q(A). Hence there is a common
denominator g ∈ A for f1 , ..., fr satisfying g(a) 6= 0 and thus Ag = A eg
is integrally closed in Q(Ag ) = Q(A), so the open neighborhood Xg of a
contains only normal points.
Xe −→ X
∪ ∪ ,
∼
=
π −1 (U ) −→ U
m = (π)
83
Remark 10.6. Every nonzero element x ∈ R in a discrete valuation ring R
can be written uniquely
(1) x = eπ n
Indeed there are more rings R, such that the localization Rp with respect
to a prime ideal p ⊂ R is a discrete valuation ring. Of course p has to be a
minimal nonzero prime ideal. Here we shall prove:
Theorem 10.8. For an irreducible normal variety X its local ring OX,Y at
a one codimensional irreducible subvariety Y ,→ X is a discrete valuation
ring.
84
Now, if I(Y ) = (g1 , ..., gr ), then gi ∈ I(N (f )), so Hilberts Nullstellensatz
provides a number m ∈ N with gim ∈ (f ) for i = 1, ..., r. Then we have
I(Y )n ⊂ (f ) ,→ A
1. We have
g·m mn
h·m= ⊂ ⊂ A,
f f
because of g ∈ mn−1 and mn ⊂ (f ),
2. but
h · m 6⊂ m.
Namely: Otherwise M := m is a faithful R[h]-module, which is finitely
generated as an R-module. But that means that h ∈ Q(R) is integral
over R, hence, R being integrally closed in its field of fractions Q(R),
we conclude that h ∈ R, a contradiction.
f = f · h · h−1 = g · h−1 ∈ m · m = m2 ,
a contradiction, since f ∈ m \ m2 .
85
associating to a rational function h ∈ R(X) its ”order” (or ”multiplicity”)
along Y as follows: If OX,Y ⊃ m = (f ), then
∗
n , if h = ef n , with e ∈ OX,Y
ordY (h) := .
∞ , if h = 0
The order function satisfies
ordY (gh) = ordY (g) + ordY (h)
as well as
ordY (g + h) ≥ min {ordY (g), ordY (h)} .
Note that in particular for irreducible Y ,→ X of codimension 1 and a nonzero
f ∈ R(X) we have either Y ∩ Df 6= ∅ or Y ∩ D1/f 6= ∅ depending on whether
ordY (f ) ≥ 0 or ordY (f ) < 0. So the points of indeterminacy of a rational
function f form a closed set of codimension at least 2. For normal curves we
thus find:
Remark 10.9. 1. A rational function f ∈ R(X) on an irreducible normal
curve has no points of indeterminacy, hence defines a morphism f :
X −→ P1 = k ∪ {∞}.
2. Indeed, we may replace P1 with any complete variety Y and obtain
that any morphism ϕ : U −→ Y defined on a non-empty open subset
U ⊂ X of an irreducible normal curve into a complete variety Y can
be extended to a morphism ϕ̂ : X −→ Y .
3. As a consequence we obtain that the category of irreducible normal
complete curves is anti-equivalent to the category of finitely generated
field extensions of our base field k of transcendence degree 1.
4. Finally let us mention, that for k = C the functor X 7→ Xh is an
equivalence between the category of irreducible normal complete curves
and the category of compact Riemann surfaces.
Definition 10.10. For an irreducible affine variety X we denote D(X) the
free abelian group generated by the codimension one irreducible subvarieties
of X, i.e.
( r )
X
D(X) := ni Yi ; ni ∈ Z, Yi ,→ X irreducible, dim X = dim Yi + 1 .
i=1
Pr
The elements D = i=1 ni Yi in D(X) are called Weil divisors on X.
86
Now given a rational function f ∈ R(X) \ {0} on an irreducible normal
variety X we can associate to it a divisor
X
Df := ordY (f ) · Y.
irred. 1-codim Y ,→X
Note that the above sum is finite, indeed ordY (f ) 6= 0, iff Y is an irreducible
component of N (f ) or of P (f ), where N (f ) = f −1 (0) is the set of the zeros
of f and P (f ) = f −1 (∞) the set of poles of f . Then
R(X)∗ −→ D(X), f 7→ Df ,
11 Regular Points
In order to study local properties of an affine algebraic variety X ,→ k n one
could - as in real analysis - think of “approximating“ X near a point a ∈ X
by an affine linear subspace a + Ta X, where Ta X ,→ k n is a vector subspace
of k n : For a polynomial f ∈ k[T ] and a ∈ k n denote
n
X ∂f
da f := (a) · Ti ∈ k]T ]1
i=1
∂T i
Ta X := N (k n ; da f, f ∈ I(X)).
Ta X = N (k n ; da f1 , ..., da fr ),
87
Example 11.1. For a hypersurface X = N (k n ; f ) with squarefree f ∈ k[T ]
(such that I(X) = (f )) we have Ta (X) = N (k n ; da f ). So with the ”singular
set ” S(X) := {a ∈ X; da f = 0} ,→ k n we have
dim X , if a ∈ X \ S(X)
dim Ta X = .
dim X + 1 , if a ∈ S(X)
holds for all f, g ∈ A. Denote Dera (A) the vector space of all derivations of
A at a ∈ X.
Indeed, both, the left and the right hand side consist of the derivations of
k[T ] at a vanishing on I(X).
88
Lemma 11.3. There is a natural isomorphism
Ta X := (mX,a /m2X,a )∗ .
T : (X, a) 7→ Ta X, ϕ 7→ Ta ϕ
89
3. The tangent map Ta ι : Ta Z −→ Ta X induced by a closed embedding
ι : Z ,→ X of a closed subvariety Z containing a is injective: The ho-
momorphisms OX,a −→ OZ,a as well mX,a −→ mZ,a and mX,a /m2X,a −→
mZ,a /m2Z,a are surjective. Hence Ta ι, being the dual of the last map, is
injective.
90
For the next result we need a statement which we shall prove later on:
Proof. Let us first show that the singular locus is closed: S(X) ,→ X, resp.
that the regular locus X\S(X) is open. Since according to 11.10 the reducible
points of X, i.e. those which are contained in several irreducible components
of X, are singular we may assume that X is irreducible, indeed even affine.
But in that case the description of S(X) of Ex. 11.9.1 by minors gives the
result. It remains to show that any irreducible variety X admits smooth
points. We start with irreducible hypersurfaces:
Proposition 11.12. For an irreducible polynomial f ∈ k[T1 , ..., Tn ] there
are points a ∈ N (k n ; f ) with da f 6= 0. In particular the regular locus of
X = N (k n ; f ). is nonempty.
∂f
Proof. If da f = 0 for all a ∈ X, we have ∂T i
∈ I(X) for i = 1, ..., n resp.
∂f
f | ∂Ti . Hence all partial derivatives, having a total degree less than that of
f , have to vanish. And that means that our base field k has characteristic
p > 0 and f ∈ k[T p ] = k[T1p , ..., Tnp ], but k being algebraically closed we have
k[T p ] = (k[T ])p , in particular, f is not irreducible, a contradiction.
In order to apply that in the general case we need
Proposition 11.13. If L = k(a1 , ...., ar ) ⊂ k is a finitely generated field
extension, then there is a subset I ⊂ {1, ..., r}, such that the ai , i ∈ I, are
algebraically independent and L ⊃ E := k(ai ; i ∈ I) is a finite separable field
extension.
Proof. We use induction on r. We may assume that a1 , ..., a` is a maximal
set of over k algebraic independent elements. If ` = r there is nothing to
be shown. Otherwise denote f ∈ k[T1 , ...T` , Tr ] \ {0} a polynomial with
f (a1 , ..., a` , ar ) = 0. We may assume that f is irreducible. Hence, according
∂f
to Prop.11.12 some partial derivative ∂T i
, i = 1, ..., `, r, does not vanish,
indeed we may assume i = r. So ar is separable over k(a1 , ..., a` ) as well as
the field extension L ⊃ k(a1 , ..., ar−1 ). Now apply the induction hypothesis to
91
F := k(a1 , ..., ar−1 ) and use the fact that two successive separable extensions
give again a separable extension.
X⊃U ∼
= V ⊂ N (k n+1 ; f ).
F (X ∩ U ) = N (V ; z1 , ..., z` ) ,→ V,
92
i.e., the point a ∈ X admits with X ∩ U a strong neighborhood, which is, as
a complex analytic variety, isomorphic to an open ball in Cn−` .
In contrast to that it is quite far from being true – even for k = C – that a
smooth point a ∈ X admits a Zariski-neighborhood isomorphic (as algebraic
variety) to an Zariski-open subset of k n (where n = dima X). Indeed, this
would imply that (w.l.o.g. X being irreducible) R(X) ∼ = k(T1 , ..., Tn ) is a
purely transcendent extension of k.
f min := fq , if fq 6= 0.
Then
T C0 (X) := N (I0 (X)) ⊂ T0 X
with the homogeneous ideal
X
I0 (X) := k[T ] · f min .
f ∈I(X)∗
Example 11.16. Since (f g)min = f min · g min , the tangent cone of a hyper-
surface X = N (k n ; f ) with squarefree f is a hypersurface as well:
For example the noose X = N (k 2 ; T22 − T12 (T1 + 1) has tangent cone
a union of two lines (if char(k) 6= 2). In particular, the tangent cone can
detect the two branches of X at 0 in a purely algebraic way!
93
On the other hand for the Neil parabola X = N (T22 − T13 ) we have
I0 (X) = (T22 )
and hence
T C0 (X) = k(1, 0).
So we see, that I0 (X) need not be a radical ideal!
2. The ideal m gives rise to a filtration (mq )q∈N of the ring R, but only if
we can find for every q a complementary additive subgroup Rq ⊂ mq ,
i.e. satisfying mq = Rq ⊕ mq+1 , suchL that Ri · Rj ⊂ Ri+j , then gr(R)
can be identified with the subring ∞ q=0 Rq of R. For example this is
true for R = k[T ] and m = m0 : Take simply Rq = k[T ]q , the vector
subspace of polynomials homogeneous of degree q.
94
3. For the local ring S −1 R with S := R \ m we have
gr(S −1 R) ∼
= grm (R)
gr(OX,0 ) ∼
= grm0 (A).
% : k[T ] −→ A, f 7→ f |X
gr(%) : k[T ] ∼
= gr(k[T ]) −→ gr(A),
gr(%)(f ) = 0 ⇐⇒ f |X ∈ mq+1
0 ⇐⇒ f ∈ mq+1 + I(X),
95
2. A graded reduced affine algebra A = ∞
L
q=0 Aq is (as graded algebra)
isomorphic to the algebra of regular functions on an affine cone iff A,
as a k-algebra, is generated by the elements of degree 1.
Definition 11.22. The tangent cone of an algebraic variety X at a point
a ∈ X is defined as the affine cone
T Ca (X) ,→ Ta X,
π|Z ∗ : Z ∗ −→ (k n )∗
96
b of X ,→ k n with respect to π is obtained from
The ”strict transform ” X
∗
X := X \ {0} as the closure
b := π −1 (X ∗ ) ,→ Z,
X
b ,→ π −1 (X) being proper in general. We shall prove that the
the inclusion X
isomorphism
π −1 (0) ∼
= Pn−1
induces an isomorphism
b ∩ π −1 (0) ∼
X = N (Pn−1 ; I0 (X)).
Z := Γϕ ,→ k n × Pn−1
of the graph
Γϕ ,→ (k n )∗ × Pn−1
of the quotient morphism
ϕ : (k n )∗ −→ Pn−1 , t 7→ [t]
in k n × Pn−1 . We have
Z = N (k n × Pn ; Si Tj − Sj Ti , 1 ≤ i < j ≤ n},
where the points in k n × Pn−1 are denoted (t, [s]) with t, s ∈ k n and Si Tj −
Sj Ti ∈ k[T1 , ..., Tn , S1 , ..., Sn ]. In order to understand the geometry of the
variety Z we consider the restricted projection
pr
p : Z ,→ k n × Pn−1 −→
2
Pn−1 .
97
Above the standard open subset Vi := Pn−1 \ N (Pn−1 ; Si ) there is a trivial-
ization
∼
=
τi : Vi × k −→ Wi := p−1 (Vi ),
([s], λ) 7→ (λs−1
i s, [s]).
π ◦ τi : Vi × k ∗ −→ (k n )Ti
given by
([s], λ) 7→ λs−1
i s.
(π ◦ τi )−1 (XTi ) ,→ Vi × k ∗ ⊂ Vi × k
−`
in Vi × k. A regular function in k[T ]Ti vanishing on
P XTi is of the formn Ti f
with a polynomial f ∈ I(X) ,→ k[T ]. If f = q fq ∈ I(X) ⊂ k with
homogeneous polynomials fq of degree q, we find
X
−` S S
f ◦ π ◦ τi = Λ f Λ = fq Λq−` .
Si q
S i
Now τi−1 (X
b ∩ Wi ) ,→ Vi × k is the zero locus of all regular functions on Vi × k,
whose restriction to Vi ×k ∗ vanishes on (π ◦τi )−1 (XTi ). These functions being
of the above form with only q ≥ `, either vanish on Vi ×0 or yield f min (Si−1 S).
Since that holds for i = 1, ..., n, we see that
(0 × Pn−1 ) ∩ X
b = 0 × N (Pn−1 ; I0 (X))
as desired.
98
Let us now draw some conclusions:
Remark 11.25. If a ∈ Y ,→ X, then obviously T Ca (Y ) ,→ T Ca (X). Fur-
thermore, the closure of a finite union being the finite union of the separate
closures, we see that, if X1 , ..., Xr are the irreducible components of X passing
through a, then
[r
ϕ(T Ca (X)∗ ) = ϕ(T Ca (Xi )∗ )
i=1
n ∗
with the quotient morphism ϕ : (k ) −→ Pn−1 . It follows
r
[
T Ca (X) = T Ca (Xi ).
i=1
Note that the corresponding statement does not hold for the tangent spaces:
The union of the subspaces Ta (Xi ) ,→ Ta X will in general be no subspace
2
anymore - e.g.
Pr consider X = N (k ; T1 T2 ) at a = 0,2 but it is not even true
2
that Ta X = i=1 Ta (Xi ), as the example X = N (k ; T2 (T2 − T1 )) shows.
so X is smooth at a.
”=⇒”: If a ∈ X is a smooth point and dima X = n, the n-dimensional
subvariety T Ca (X) ,→ k n coincides with k n . In particular ker(σ) = {0} and
σ : k[S] −→ gr(OX,a ) is an isomorphism.
99
Remark 11.27. Note that for a smooth point a ∈ X with dima X = n
we have T Ca (X) ∼
= k n , but T Ca (X) ∼
= k n does not imply that a ∈ X is a
smooth point. For example for the Neil parabola X = N (k 2 ; T22 − T13 ) we
have T C0 (X) = k × 0 ∼
= k, while gr(OX,0 ) ∼
= k[T1 , T2 ]/(T22 ).
In order to prove Th, 11.10, we need a further result of commutative
algebra:
D = D ∩ mq M ⊂ D ∩ F = 0.
Mk := {u ∈ M ; ak u ∈ F }
100
form an increasing sequence of submodules Mk ⊂ M ; hence M being noethe-
rian, we have Ms = Ms+1 for some s ∈ N. For b = as m + f ∈ (as M + F ) ∩ D
we then obtain
ab = as+1 m + af ∈ aD ⊂ mD = 0,
i.e. ab = 0 and as+1 m = −af ∈ F resp. m ∈ Ms+1 = Ms and already
as m ∈ F . Hence b ∈ F and thus b ∈ F ∩ D = 0, i.e. b = 0 as desired.
OX,a /(mX,a )n ∼
= S −1 (A/(ma )n ) ∼
= A/(ma )n .
2. For a ∈ k n we have
bkn ,a ∼
O = k[[T1 , ..., Tn ]],
where k[[T1 , ..., Tn ]] denotes the ring of formal power series in n inde-
terminates with coefficients in k. Assuming a = 0 apply the previous
point to X = k n .
Proposition 11.31. A point a ∈ X is a simple point iff the completion O
bX,a
is isomorphic to a formal power series ring.
101
Proof. Let a ∈ X be a smooth point. We may assume that X = Sp(A).
Take functions f1 , ..., fn ∈ ma providing a basis of ma /(ma )2 and consider the
map
k[T1 , ..., Tn ] −→ A, Ti 7→ fi .
It induces surjections k[T ] −→ A/(ma )n for every n ∈ N. On the other hand,
gr(OX,a ) being a polynomial ring in the indeterminates gr(fi ), we conclude
that k[T ]/(m0 )n and A/(ma )n have the same dimension as a k-vector space,
so the induced map
k[T1 , ..., Tn ]/(m0 )n −→ A/(ma )n
is an isomorphism for all n ∈ N. Now apply Ex.11.30.2. On the other hand
gr(OX,a ) ∼
= gr(O
bX,a )
102
Proof of Th.11.33. First let us reduce the problem to the case where
1. ψ is finite,
103
Now for L = (q −1 (a) ∩ U0 ) ∪ {0} with a 6∈ E the conditions 2), 4) and
the first one of 3) are obviously satisfied. It remains to show the finiteness.
After a linear change of coordinates we may assume that L = ken resp. that
πn−1 is the projection onto the first n − 1 coordinates. Then we have to
find an integral equation for Tn |X over k[T1 , ..., Tn−1 ]. Since [0, ..., 0, 1] 6∈ X,b
there is a polynomial f ∈ I(X) ⊂ k[T1 , ..., Tn ], such that its homogeneization
fb = T0r f (T0−1 T ) satisfies fb(0, .., 0, 1) = 1 and thus fb ∈ (k[T0 , ..., Tn−1 ])[Tn ]
as well as f = fb(1, T1 , ..., Tn ) are Tn -monic - this gives the desired integral
equation. Finally let us show the second part of 3): Set Y := ψ(X) ,→ k n−1 .
Since ψ is finite and ψ −1 (0) = {0}, the sets ψ −1 (V ) form a neighbourhood
basis of 0 ∈ X. As a consequence, if O(X) = O(Y )f1 + ... + O(Y )fr , we have
as well OX,0 = OY,0 f1 + ... + OY,0 fr . Since T0 ψ : T0 X −→ T0 Y is injective,
ψ ∗ : mY,0 /m2Y,0 −→ mX,0 /m2X,0 is surjective. Now we apply Nakayama’s
lemma to the finite OX,0 -module mX,0 and obtain mX,0 = OX,0 · ψ ∗ (mY,0 ).
Hence
OX,0 /OX,0 ψ ∗ (mY,0 ) = OX,0 /mX,0 ∼= k.
So 1 is a generator of the finite OY,0 -module OX,0 since it is mod mY,0 , i.e.
ψ ∗ : OY,0 −→ OX,0 is surjective, while injectivity follows from the fact that
ψ : X −→ Y is dominant and ψ −1 (0) = {0}.
The special situation as described in the beginning is treated with the
tools of the proof of Krull’s principal ideal theorem. Let X = Sp(A) and
treat the pull back homomorphism ψ ∗ : k[T ] ,→ A as an inclusion. We
write I(ψ(Z)) = k[T ] · h with some polynomial h ∈ k[T ] and show that
I(Z)0 = OX,0 · h. If ψ −1 (0) = {x1 = 0, x2 , ..., xr }, we are done, if we succeed
in showing
1. Any function g ∈ I(Z) with g(xi ) 6= 0 for i = 2, ..., r lies in (h) ,→ OX,0 .
2. The ideal I(Z)0 is generated by functions g ∈ I(Z) as above.
1.) For the characteristic polynomial χ(T, S) ∈ (k[T1 , ..., Td ])[S] of the mul-
tiplication µg ∈ Endk(T ) (Q(A)) we have
with the extension degree ` := [Q(A) : k(T )]. Due to the Hamilton-Cayley
theorem we have χ(µg ) = 0 resp. χ(g) = 0, hence
104
yields N(g) ∈ I(ψ(Z)) = k[T ]·h; so it suffices to show that gb := p(T, g) ∈ A is
invertible in OX,0 ⊃ A or, equivalently, that gb(0) = p(0, g(0)) = p(0, 0) 6= 0.
To see that we consider the induced multiplication map
over the fiber ψ −1 (0) = N (X; T1 , ..., Td ) with g := g+(T1 , ..., Td ) ∈ A/(T1 , ..., Td ).
Note that the algebra A/(T1 , ..., Td ) need not be reduced, but in any case we
have
Red(A/(T1 , ..., Td )) ∼= O(ψ −1 (0)) ∼
= kr .
If now f (T, S) ∈ (k[T ])[S] denotes the minimal polynomial of Td+1 |X over
k(T1 , ..., Td ) (since k[T1 , ..., Td ] is integrally closed in its field of fractions
k(T1 , ..., Td ), it has coefficients in k[T1 , ..., Td ]), we have
A∼
= k[T1 , ..., Td+1 ]/(f (T, Td+1 )).
Hence
A/(T1 , ..., Td ) ∼
= k[Td+1 ]/(f (0, Td+1 )),
and r
Y
f (0, Td+1 ) = Td+1 (Td+1 − ai )mi ,
i=2
The polynomial f (0, Td+1 ) has a simple zero at the origin, since otherwise
∂f
∂Td+1
(0) = 0 gives 0 × k ⊂ T0 X, which contradicts the injectivity of the
tangent map
T0 ψ : T0 X = ker d0 f −→ T0 k d = k d .
With the Chinese remainder theorem we decompose
r
A/(T1 , ..., Td ) ∼
M
=k⊕ k[Td+1 ]/((Td+1 − ai )mi )
i=2
105
2.) Assume I(Z) = (h1 , ..., hs ). Choose c ∈ k, such that c2 6= hj (xi ) for
i = 1, ..., r, j = 1, ..., s and f ∈ I(Z) with f (xi ) = c for i = 2, ..., r - remember
that the restriction A = O(X) −→ O(Z ∪ {x2 , ..., xr }) is surjective. Take
now gj := hj + f 2 : According to Nakayama’s lemma they generate I(Z)0 ,
since their residue classes g j mod mX,0 I(Z)0 coincide with hj , hence generate
I(Z)0 /mX,0 I(Z)0 .
Trevlig Sommar!
106