Manin
Manin
Summary. This book provides one with a concise but extremely lucid exposition
of basics of algebraic geometry seasoned by illuminating examples.
The preprints of this book proved useful to students majoring in mathematics
and modern mathematical physics, as well as professionals in these fields.
Editor’s preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
Author’s preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
1 Affine schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.1 Equations and rings . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Geometric language: Points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.3 Geometric language, continued . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16
1.4 The Zariski topology on Spec A . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
1.5 The affine schemes (a preliminary definition) . . . . . . . . . . . . . . . . 29
1.6 Topological properties of morphisms and Spm . . . . . . . . . . . . . . . 33
1.7 The closed subschemes and the primary decomposition . . . . . . . 42
1.8 Hilbert’s Nullstellensatz (Theorem on zeroes) . . . . . . . . . . . . . . . 49
1.9 The fiber products . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
1.10 The vector bundles and projective modules . . . . . . . . . . . . . . . . . 55
1.11 The normal bundle and regular embeddings . . . . . . . . . . . . . . . . . 63
1.12 The differentials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
1.13 Digression: Serre’s problem and Seshadri’s theorem . . . . . . . . . . 71
1.14 Digression: ζ-function of a ring . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74
1.15 The affine group schemes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
1.16 Appendix: The language of categories. Representing functors . . 87
1.17 Solutions of selected problems of Chapter 1 . . . . . . . . . . . . . . . . . 96
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 177
Prefaces 5
Editor’s preface
It is for more than 40 years now that I wanted to make these lectures — my first
love — available to the reader: The preprints of these lectures [Ma1, Ma2] of cir-
culation of mere 500+200 copies became bibliographic rarities almost immediately.
Meanwhile the elements of algebraic geometry became everyday language of work-
ing theoretical physicists and the need in a concise manual only increased. Various
(nice) text-books are usually too thick for anybody who does not want to become a
professional algebraic geometer, which makes Manin’s lectures even more appealing.
The methods described in these lectures became working tools of theoretical
physicists whose subject ranges from high energy physics to solid body physics (see
[Del] and [Ef], respectively) — in all questions where supersymmetry naturally arises.
In mathematics, supersymmetry in in-build in everything related to homotopy and
exterior products of differential forms, hence to (co)homology. In certain places,
supersymmetric point of view is inevitable, as in the study of integrability of certain
equations of mathematical physics ([MaG]). Therefore, preparing proceedings of
my Seminar on Supermanifolds (see [SoS]) I urgently needed a concise and clear
introduction into basics of algebraic geometry.
In 1986 Manin wrote me a letter allowing me to include a draft of this trans-
lation as a Chapter in [SoS], and it was preprinted in Reports of Department of
Mathematics of Stockholm University. My 1972 definition of superschemes and su-
permanifolds [L0] was based on these lectures; they are the briefest and clearest
source of the background needed for studying those aspects of supersymmetries
that can not be reduced to linear algebra.
Written at the same time as Macdonald’s lectures [M] and Mumford’s lectures
[M1, M2, M3] Manin’s lectures are more lucid and easier to come to grasps. Later
on there appeared several books illustrating the topic from different positions, e.g.,
[AM], [Sh0], [E1, E2, E3], [H, Kz] and I particularly recommend [Reid] for the first
reading complementary to this book.
For preliminaries on algebra, see [vdW, Lang]; on sheaves, see [God, KaS]; on
topology, see [K, FFG, RF, Bb3]; on number theory, see [Sh1, Sh2].
In this book, N := {1, 2, . . . } and Z+ := {0, 1, 2, . . . }; Fm := Z/m; A× is the
group of invertible elements of the ring A. The term “identity” is only applied to
relations and maps, the element 1 (sometimes denoted by e) such that 1a = a = a1
is called unit (sometimes unity). Other notation is standard.
I advise the reader to digress from the main text and skim the section on cate-
gories as soon as the word “functor” or “category” appears for the first time.
Author’s preface
In 1966–68, at the Department of Mechanics and Mathematics of Moscow State
University, I read a two-year-long course in algebraic geometry. The transcripts of
the lectures of the first year were preprinted [Ma1, Ma2] (they constitute this book),
that of the second year was published in the Russian Mathematical Surveys [Ma3].
These publications bear the remnants of the lecture style with its pros and cons.
Our goal is to teach the reader to practice the geometric language of commutative
algebra. The necessity to separately present algebraic material and later “apply” it
to algebraic geometry constantly discouraged geometers, see a moving account by
O. Zariski and P. Samuel in their preface to [ZS].
The appearance of Grothendieck’s scheme theory opened a lucky possibility not
to draw any line whatever between “geometry” and “algebra”: They appear now as
complimentary aspects of a whole, like varieties and the spaces of functions on these
spaces in other geometric theories. From this point of view,
the commutative algebra coincides with (more pre-
c i s e l y, i s f u n c t o r i a l l y d u a l t o ) t h e t h e o r y o f l o c a l
g e o m e t r i c o b j e c t s — a f f i n e s c h e m e s.
This book is devoted to deciphering the above claim. I tried to consecutively
explain what type of geometric images should be related with, say, primary de-
composition, modules, and nilpotents. In A. Weil’s words, the spacial intuition is
“i n v a l u a b l e i f i t s r e s t r i c t e d n e s s i s t a k e n i n t o a c c o u n t”. I
strived to take into account both terms of this neat formulation.
Certainly, geometric accent considerably influenced the choice of material. In
particular, this chapter should prepare the ground for introducing global objects.
Therefore, the section on vector bundles gives, on “naive” level, constructions be-
longing, essentially, to the sheaf theory.
Finally, I wanted to introduce the categoric notions as soon as possible; they are
not so important in local questions but play ever increasing role in what follows. I
advise the reader to skim through the section “Language of categories” and return
to it as needed. 1)
I am absolutely incapable to edit my old texts; if I start doing it, an irresistible
desire to throw everything away and rewrite completely grips me. But to do some-
thing new is more interesting. Therefore I wish to heartily thank D. Leites who saved
my time.
The following list of sources is by no means exhaustive. It may help the reader to
come to grasps with the working aspects of the theory: [Sh0], [Bb1] (general courses);
[M1], [M2], [Ma3], [MaG], [S1], [S2] (more special questions).
The approach of this book can be extended, to an extent, to non-commutative
geometry [Kas]. My approach was gradually developing in the direction along which
the same guiding principle — construct a matrix with commuting elements satisfying
only the absolutely necessary commutation relations — turned out applicable in ever
wider context of n o n - c o m m u t a t i v e geometries, see [MaT], see also [GK],
[BM] (extension to operads and further).
1
See also [McL, GM].
Chapter 1
Affine schemes
Fi (T ) = 0, where i ∈ I. (1.1)
Why the ground ring or the ring of constants K enters the definition is clear:
The coefficients of the Fi belong to a fixed ring K. We say that the system X
is defined over K.
What should we take for a solution of (1.1)?
To say, “a solution of (1.1) is a set t = (tj )j∈J of elements of K such that
Fi (t) = 0 for all i” is too restrictive: We wish to consider, say, complex roots
of equations with real coefficients. The radical resolution of this predicament
is to consider solutions in any ring, and, “for simplicity”, in all of the rings
simultaneously.
To consider solutions of X belonging to a ring L, we should be able to
substitute the elements of L into Fi , the polynomials with coefficients from K,
i.e., we should be able to multiply the elements from L by the elements from
K and add the results, that is why L must be a K-algebra.
Recall that the set L is said to be a K-algebra if L is endowed with the
structures of a ring and a K-module, interrelated by the following properties:
1. The multiplication K × L −→ L is right and left distributive with respect
to addition;
2. k(l1 l2 ) = (kl1 )l2 for any k ∈ K and l1 , l2 ∈ L.
8 Ch. 1. Affine schemes
1.1.2a. Proposition. 1) The system whose left hand sides are elements of
the ideal (F ) is the largest system equivalent to (F ).
2) There is a one-to-one correspondence X(L) ←→ HomK (K(X), L),
where HomK denotes the set of K-algebra homomorphisms.
Proof. 1) Let P be the ideal in K[T ], where T = (T )j∈J , generated by the
left sides of the system of equations (1.1). It is easy to see that the system
obtained by equating all elements of P to zero is equivalent to our system, call
it (X). At the same time, the larger system is the maximal one since if we add
to it any equation F = 0 not contained in it, we get a system not equivalent
to (X). Indeed, take L = K[T ]/P . In L, the element t, where tj ∼= Tj0 mod P ,
is a solution of the initial system (X) whereas F (t) 6= 0 since F 6∈ P .
2) Let t = (tj )j∈J ∈ X(L). There exists a K-algebra homomorphism
K[T ] −→ L which coincides on K with the structure homomorphism K −→ L
and sends Tj to tj . By definition of X(L), the ideal P lies in the ker-
nel of this homomorphism, so we can consider a through homomorphism
A = K[T ]/P −→ L.
Conversely, any K-algebra homomorphism A −→ L uniquely determines
a through homomorphism K[T ] −→ A −→ L. Let tj be the image of Tj under
1.1 Equations and rings 9
this through homomorphism; then t = (tj )j∈J ∈ X(L) since all elements of P
vanish under this homomorphism.
It is easy to check that the constructed maps X(L) ←→ HomK (K(X), L)
are inverse to each other. u
t
For a non-zero K-algebra L, a system X over a ring K is said to be
consistent over L if X(L) 6= ∅ and inconsistent otherwise. Proposition 1.1.2a
shows that X is inconsistent only if its coordinate ring, or rather, algebra,
K(X), is zero, in other words, if 1 ∈ (X). 1)
1.1.3. Examples from arithmetics.
1.1.3a. The language of congruences. Let n be an integer of the form
4m + 3. The classical proof of the fact that n is not representable as a sum
of two perfect squares is as follows: If it were, the congruence
T12 + T22 ≡ 3 (mod 4) (1.3)
would have been solvable, whereas a simple case-by-case checking (set
T1 = 4a + r1 , T2 = 4b + r2 and consider the eight distinct values of (r1 , r2 ))
establishes that this is not the case.
From our point of view this argument reads as follows: Consider the system
T12 + T22 = n, where K = Z. (1.4)
The reduction modulo 4, i.e., the map Z −→ Z/4, determines a map
X(Z) −→ X(Z/4) and, if X(Z) 6= ∅, then X(Z/4) 6= ∅, which is not the
case.
More generally, in order to study X(Z) for any system X with integer
coefficients, we can consider sets X(Z/m) for any m and try to deduce from
this consideration some information on X(Z).
Usually,
if X(L) = ∅ for a nontrivial (i.e., 1 6= 0) field L, then X(Z) = ∅. (1.5)
In practice one usually tests L = R and the fields L = Fm (:= Z/m) for all
prime m’s.
A number of deepest results of the theory of Diophantine equations are
related with the problem: When is the converse statement true? A prototype
of these results is
Legendre’s theorem ([BSh]). Let
a1 T12 + a2 T22 + a3 T32 = 0, K = Z. (1.6)
If X(Z) = {(0, 0, 0)}, then X(L) = {(0, 0, 0)} for at least one of the rings
L = R or L = Z/mZ, where m > 1.
1
In school, and even in university, one often omits “over L” in the definition
of consistency thus declaring systems inconsistent only partly (over some classes
of rings) as totally inconsistent.
10 Ch. 1. Affine schemes
0 · T + 2 = 0, K = Z. (1.7)
Fig. 1 Fig. 2
being cut along the real axis). Now, routine topological considerations give
Harnak’s estimate, see, e.g., [Ch].
1.1.3d. The algebra of mathematical logic in geometric terms. 2) A
Boolean ring is any ring R (with 1) such that P 2 = P for any P ∈ R. Clearly,
P + Q = (P + Q)2 = P 2 + P Q + QP + Q2 = P + P Q + QP + Q (1.8)
where the bar stands for the negation, ∧ for the conjunction and ∨ for the
disjunction. With respect to the above operations the empty statement ∅ is
the zero, and ¯∅ is the unit. Clearly, P 2 = P and 2P = P + P = 0 for any P .
1.1.4. Summary. We have established the equivalence of the two languages:
That of the systems of equations (which is used in the concrete calculations)
and that of the rings and their morphisms. More exactly, we have established
the following equivalences:
A system of equations X A K-algebra K(X) with a
over a ring K ⇐⇒ system of generators
for unknowns {Tj | j ∈ J}. {tj | j ∈ J}.
A solution of the system ½ ¾
A K-algebra homomorphism
of equations X ⇐⇒
K(X) −→ L.
in a K-algebra L.
Finally, notice that using the language of rings we have no need to consider
a fixed system of generators t = (tj )j∈J of K(X) = K[T ]/(X); we should
rather identify the systems of equations obtained from each other by any
invertible change of unknowns. Every generator of K(X) plays the role of one
of “unknowns”, and the value this unknown takes at a given solution of the
system coincides with its image in L under the corresponding homomorphism.
1.1.5. Exercises. 1) The equation 2T − 4 = 0 is equivalent to the equation
T − 2 = 0, if and only if 2 is invertible in K.
2
For mathematical logic from an algebraist’s point of view, see [Ma4].
12 Ch. 1. Affine schemes
X1 : si (T ) = 0, i = 1, . . . , k ≤ n,
P
X2 : Tji = 0, i = 1, . . . , k ≤ n.
1≤j≤n
.
S (V ∗ ) = K[x1 , . . . , xn ] −→ K, (1.11)
a ∈ A, b ∈ A, a b ∈ p =⇒ either a ∈ p or b ∈ p. (1.12)
The set of all the prime ideals of A is said to be the (prime) spectrum of A
and is denoted by Spec A. The elements of Spec A are called its points.
In what follows, we enrich the set Spec A with additional structures mak-
ing it into a topological space rigged with a sheaf of rings: This will lead to the
definition of an affine scheme. Schemes, i.e., topological spaces with sheaves,
locally isomorphic to affine schemes, are the main characters of algebraic ge-
ometry.
Starting the study of spectra we have to verify, first of all, that there exists
indeed what to study.
1.2.3. Theorem. If A 6= 0, then Spec A 6= ∅.
In the proof of this theorem we need the following:
1.2.4. Lemma (Zorn’s lemma). In a partially ordered set M , let every linearly
ordered subset N ⊂ M contain a maximal element. Then M contains a maximal
element.
For proofs of Zorn’s lemma, see, e.g., [K] or [Hs], where it is proved together with
its equivalence to the choice axiom, the complete order principle, and several other
statements. For an interesting new additions to the list of equivalent statements,
see, e.g., [Bla]. u
t
An ordered set satisfying the condition of Zorn’s lemma is called an inductive
set.
Proof of Theorem 1.2.3. Denote by M the set of all the ideals of A different
from A. Since M contains (0), it follows that M 6= ∅. The set M is partially ordered
with respect
S
to inclusion. In M , take an arbitrary linearly ordered set {Pα }α∈Λ .
Then Pα is also an ideal of A (mind the linear order) and this ideal is different
α S
from A (since the unit element does not belong to Pα ). Therefore M is an inductive
α
14 Ch. 1. Affine schemes
set. Denote by p its maximal element; it is a maximal ideal, and therefore a prime
one: In A/p, every non-zero ideal, in particular, all the principal ones, coincide with
A/p. Therefore every non-zero element of A/p is invertible and A/p is a field. u
t
Fig. 3
(j)
Generally, ak are fixed elements of a ring of constants K whereas “pas-
sage to the absolute case” means that we now w r i t e a g e n e r a t -
(j)
i n g s y s t e m o f a l l r e l a t i o n s b e t w e e n t h e ak O V E R Z ,
and add it to our initial system of equations, bind-
(j)
i n g t h e xi a n d t h e ak t o g e t h e r . A n d a f t e r t h a t w e
may specialize coefficients as well, retaining only
the relations between them.
1.2.8. Exercise. A weak form of Hilbert’s Nullstellensatz (theorem
on zeros). Consider a system of equations {Fi (T ) = 0}, where T = (Tj )j∈J ,
over the ring K. Then either this system has a solution with values in a field,
or there exist polynomials Gi ∈ K[T ] (finitely many of them are 6= 0) such
that X
Gi Fi = 1.
i
Q
Z/5Z
4
3
Z/3Z
2
Z/2Z
1
0
(0) (2) (3) (5) Spec Z
Fig. 4
lie beyond the scope of these lectures, the “line” over the field Z/p — the
“vertical axis” over the point (p) — should be drawn “coiled into a ring”, i.e.,
the points of this “line” should form vertices of a regular p-gon. This does not
simplify the task of an artist.
To distinct elements of A the same functions on the spectrum might T corre-
spond; their difference represents the zero function, i.e., belongs to px .
x∈Spec A
Clearly, all the nilpotents are contained in this intersection. Let us prove the
opposite inclusion. For this, we need a new notion.
The set N of all nilpotent elements of a ring R is an ideal (as is easy to
see), it is called the nilradical of R.
1.3.2. Theorem. A function that vanishes at all the points of Spec A is
represented by a nilpotent element of A. In other words, the nilradical is the
intersection \
p. (1.16)
p∈Spec A
The geometric R-points of At are the following ones: There are two of them
for t > 0, there is one of them for t = 0, there are none for t < 0.
The geometric C-points: There are always two of them except for t = 0
(the case of tangency).
In order to be able to state that over C there are always two intersection
points if proper multiplicities are ascribed to them, we have to assume that
at t = 0 the multiplicity of the intersection point is equal to 2.
Observe that dimR At = 2 regardless of the value of t. The equality:
is not accidental, and we will prove the corresponding theorem when we intro-
duce the projective space that will enable us to take into account the points
that escaped to infinity.
A singularity like the coincidence of the intersection points corresponding
to tangency creates nilpotents in A0 .
2) One-point spectra. Let Spec A consist of one point corresponding to
an ideal p ⊂ A. Then A/p is a field and p consists of nilpotents. The ring A
is Artinian 8) , hence Noetherian, and the standard arguments show that p is
a nilpotent ideal.
Indeed, let f1 , . . . , fn be its generators and fim = 0 for 1 ≤ i ≤ n. Then
mn X
à n !
Y
aij fi = 0 (1.20)
j=1 i=1
M = M1 ⊃ M2 ⊃ · · · ⊃ Mr = 0. (1.21)
V (a1 . . . an ) = V (a1 ∩ · · · ∩ an ).
Using Theorem 1.3.1, we can describe the set of functions that vanish on
V (E). Obviously, it contains all the elements of the ideal (E) generated by E
and all the elements f ∈ A such that f n ∈ (E) for some n. It turns out that
this is all.
The radical r(I) of the ideal I ⊂ A is the set (actually, an ideal in the ring
A) defined to be
In other words,
{x} ∼
= Spec(A/px ), and
only the points corresponding to the maximal ideals are closed.
Fig. 5
On Fig. 5 there are plotted two spectra: The spectrum of the ring of integer
p-adic numbers, Zp , and Spec C[T1 , T2 ]. The arrows indicate the specialization
relation. The picture of Spec Zp does not require comments; note only that
Spec A may be a finite but not discrete topological space. The other picture
is justified by the following statement.
Proposition. Let K be algebraically closed. The following list exhausts the
prime ideals of the ring K[T1 , T2 ]:
a) the maximal ideals (T1 − t1 , T2 − t2 ), where t1 , t2 ∈ K are arbitrary;
22 Ch. 1. Affine schemes
b) the principal ideals (F (T1 , T2 )), where F runs over all irreducible poly-
nomials;
c) (0).
For proof, see [Reid].
The images influenced by this picture can serve as a base for a working
dimension theory of algebraic geometry. We will show this later; at the moment
we will confine ourselves to preliminary definitions and two simple examples.
We say that a sequence of points x0 , x1 , . . . , xn of a topological space X
is a chain of length n with the beginning at x0 and the end at xn if xi 6= xi+1
and xi+1 is a specialization of xi for all 0 ≤ i ≤ n − 1.
The height ht x of x ∈ X is the upper bound of the lengths of the chains
with the beginning at x. The dimension dim X of X is the upper bound of the
heights of its points.
Example. In the space X = Spec K[T1 , . . . , Tn ], where K is a field, there is
a chain of length n corresponding to the chain of prime ideals
The sets D(f ) are called big open sets; they constitute a basis of the Zariski
topology of Spec A, since
\
Spec A \ V (E) = D(f ) for any E ⊂ A. (1.27)
f ∈E
Consider, for example, Spec C[T ]. Its closed points correspond to the ideals
(T − t), where t ∈ C, and constitute, therefore, the “complex line”; the non-
empty open sets are {0} and the sets of all the points of the complex line
9
For discussion of non-integer values of dimension, with examples, see [Mdim]. As
Shander and Palamodov showed, in the super setting, the relation of “border”ing
is partly reversed: The odd codimension of the boundary of a supermanifold might
be negative, see [SoS].
1.4 The Zariski topology on Spec A 23
X = U ∪ (X \ U ). u
t
Therefore f and g vanish on the closed subsets of the whole spectrum that
together cover the whole space (this is a natural way for zero divisors to appear
in the rings of functions).
We only have to verify that V (f ), V (g) 6= Spec A. But this is obvious since
f and g are not nilpotents. ut
Corollary. Let I ⊂ A be an ideal. The closed set V (I) is irreducible if and
only if the radical r(I) is prime.
24 Ch. 1. Affine schemes
To every point x ∈ Spec A the closed set {x} corresponds; x is called a generic
point of this closed set. Every irreducible closed subset has only one generic
point, obviously.
1.4.7. Decomposition into irreducible components.
1.4.7a. Theorem. Let A be a Noetherian
S
ring. Then Spec A can be uniquely repre-
sented in the form of a finite union Xi , where the Xi are maximal closed irreducible
i
subsets.
Proof. The theorem is a geometric reformulation of the ascending chain condition
on the ideals of A: Every descending chain of closed subsets of Spec A stabilizes. u
t
1.4.7b. Noetherian topological space. Since we will encounter spaces with
the property described in Theorem 1.4.7 not homeomorphic to the spectrum of any
ring, let us introduce a special definition: A given topological space X will be called
Noetherian if any descending chain of its closed subsets stabilizes (we say: DCC
holds for X).
Theorem. Let X be a Noetherian topological space. Then X is finite union of its
maximal closed irreducible subsets.
The maximal closed irreducible subsets of a Noetherian topological space X are
called irreducible components of X.
Proof. Consider the set of irreducible closed subsets of X ordered with respect to
inclusion. Let us prove that the set is an inductive one, i.e., if {Xα | α ∈ J} is
a linearly ordered family of irreducible closed subsets of X, then for its maximal
T T
element we may take Xα . The irreducibility of the set Xα follows, for
α α
example, from the fact that if U1 , U2 are non-empty open subsets, then U1 ∩ Xα
and U2 ∩ Xα are non-empty for some α, and therefore U1 ∩ U2 is non-empty since
the Xα are irreducible. S
It follows that X is the union of its irreducible components: X = Xi .
i∈I
So far, we have not used the Noetherian property.
Now, let X be Noetherian and X = X1 ∪ X2 , where X1 , X2 are closed. If one
of the Xi is reducible, we can represent it in the form of the union of two closed
sets, and so on. This process terminates, otherwise we would have got an infinite
descending chain of closed sets (the “Noetherian induction”).
S
In the obtained finite
union, let us leave only the maximal elements: X = Xi . This decomposition
1≤i≤n
coincides with
S
the above one: If YS is an (absolutely) maximal closed subset of X,
then Y ⊂ Xi implies Y = (Xi ∩ Y ), and therefore Xi ∩ Y = Y for some i;
1≤i≤n 1≤i≤n
hence Y = Xi . S
If I 0 is a proper subset of I, then Xi does not coincide with X: Let Xj be
i∈I 0 S S
a discarded component, i.e., j 6∈ I 0 . If Xj ⊂ Xi then Xj = (Xi ∩ Xj ) and,
i∈I 0 i∈I 0
1.4 The Zariski topology on Spec A 25
Then
[
Xi = V (b1 . . . bn ) = V (0) = Spec A,
1≤i≤n
Xi ∩ Xj = V (bi ∪ bj ) = V (A) = ∅ if i 6= j.
Therefore
Q
Spec Ai splits into the disjoint union of its closed subsets
1≤i≤n (1.33)
V (bi ) ∼
= Spec A/bi = Spec Ai .
Note that, for infinite products, the statement (1.33) is false, see Exercise
1.4.14 7e).
1.4.11. The statement converse to (1.30).
`
Proposition. Let X = Spec A = Xi , where the Xi are closed non-
1≤i≤n
intersecting subsets. Then there exists an isomorphism A = ∼ Q Ai such
1≤i≤n
that in the notation of the above subsection Xi = V (bi ).
Proof. We will consider in detail the case n = 2. Let Xi = V (bi ). By Corol-
lary 2.4.12 we have
X1 ∪ X2 = X ⇐⇒ V (b1 b2 ) = X ⇐⇒ b1 b2 ⊂ N,
X1 ∩ X2 = ∅ ⇐⇒ V (b1 + b2 ) = ∅ ⇐⇒ b1 + b2 = A.
T1 (T1 − T22 ) = 0,
T2 (T1 − T22 ) = 0,
in Spec C[T1 , T2 ] consists of the line and a parabola, and hence are pair-
wise isomorphic. The intersection point in each case is the “apex” of the
parabola. Prove that, nevertheless, the rings of functions of these curves are
non-isomorphic.
5) Let A be a Noetherian ring. Construct a graph whose vertices are in one-
to-one correspondence with the irreducible components of the space Spec A,
and any two vertices are connected by an edge if and only if the corresponding
components have a non-empty intersection. Prove that the connected compo-
nents of Spec A are in one-to-one correspondence with the linearly connected
components of the graph.
10
I.e., 1 ∈ S and f, g ∈ S =⇒ f g ∈ S.
1.5 The affine schemes (a preliminary definition) 29
Q
n
6) Finish the proof of Proposition 1.4.11. Is the decomposition A = Ai ,
i=1
whose existence is claimed, uniquely defined? Q
7) Let (Ki )i∈I be a family of fields. Set A = Ki and let πi : A −→ Ki
be the projection homomorphisms. i∈I
Show that the subsets K are non-empty and the family Φa possesses the
following two properties:
α) K1 ∈ Φa & K2 ∈ Φa =⇒ K1 ∩ K2 ∈ Φa ,
β) K1 ∈ Φa & K2 ⊃ K1 =⇒ K2 ∈ Φa .
f ∈ aΦ ⇐⇒ {i | πi (f ) = 0} ∈ Φ.
If X, Y are topological spaces and F (X), F (Y ) are rings not of arbitrary, but
of continuous functions, then the homomorphism f ∗ is uniquely recovered
from a continuous f . Without certain conditions this correspondence does
not have to take place, e.g., the homomorphism f ∗ of rings of, say, continuous
functions is not uniquely recovered from an arbitrary f .
The prime objects of our study are the rings (of “functions”); therefore
important maps of the spaces are — for us — only those obtained from ring
homomorphisms.
Let ϕ : A −→ B be a ring homomorphism. To every prime ideal p ⊂ B,
we assign its pre-image ϕ−1 (p). The ideal ϕ−1 (p) is prime because ϕ induces
the embedding A/ϕ−1 (p) ,→ B/p and, since B/p has no zero divisors, neither
has A/ϕ−1 (p). We have determined a map aϕ : Spec B −→ Spec A, where the
superscript a is for “affine”.
1.5.1. Theorem. 1) aϕ is continuous as a map of topological spaces (with
respect to the Zariski topologies of these spaces).
a
2) (ϕψ) = aψ ◦ aϕ.
Proof. 2) is obvious. To prove 1), it suffices to verify that the pre-image of
a closed set is closed. Indeed,
f a
ϕ (1.36)
² α
²
X / Spec A
1.5.3. Examples. To appreciate the difference between the set Hom(A, B),
which only is of importance for us, and the set of all the continuous maps
Spec B −→ Spec A consider several simple examples.
1.5.3a. A = B = Z. As we have already observed, Spec Z consists of the
closed points (p), where p runs over the primes, and (0). The closure of (0) is
the whole space; the remaining closed sets consist of a finite number of closed
points. There are lots of automorphisms of the space Spec Z: We may permute
the closed points at random; contrariwise Hom(Z, Z) contains only the identity
map.
1.5.3b. A = K[T ], where K is a finite field, B = Z. Obviously, Spec A and
Spec B are isomorphic as topological spaces, whereas Hom(A, B) = ∅.
Examples 1.5.3a and 1.5.3b might make one think that there are much
less homomorphisms of rings than there are continuous maps of their spectra.
The opposite effect is, however, also possible.
1.5.3c. Let K be a field. Then Spec K consists of one point, and therefore
the set of its automorphism consists of one point, whereas the group of auto-
morpisms of K may be even infinite (a Lie group). Therefore
Fig. 6
T1 7→ T1 + F1 ,
T2 7→ T2 + F2 (T1 ),
(1.45)
..........................
Ti 7→ Ti + Fi (T1 , . . . , Ti−1 ),
a
ϕ : Spec B −→ Spec A, px 7→ ϕ−1 (px ). (1.46)
This study gives a partial answer to the question: What is the structure of the
topological space aϕ(Spec B)?
As is known (see, e.g., [Lang]), any homomorphism ϕ : A −→ B factorizes
into the product of the surjective ring homomorphism A −→ A/ Ker ϕ and an
embedding A/ Ker ϕ −→ B. Let us find out the properties of aϕ in these two
cases.
a
1.6.1. The properties of ϕ. The first of these cases is very simple.
Proposition. Let ϕ : A −→ B be a ring epimorphism. Then aϕ is a homeo-
morphism of Spec B onto the closed subset V (Ker ϕ) ⊂ Spec A.
1.6.1a. Exercise. Verify that this is a direct consequence of the definitions.
Hint. Prove the continuity of the inverse map
In particular, let A be a ring of finite type over a field K or over Z, i.e., let
A be a quotient of either K[T1 , . . . , Tn ] or Z[T1 , . . . , Tn ] for n < ∞.
The spectrum of the polynomial ring plays the role of an affine space
(over K or Z, respectively, cf. Example 1.2.1). Therefore s p e c t r a
of the rings of finite typ e corresp ond to affine vari-
e t i e s (“arithmetic affine varieties” if considered over Z): They are embedded
into finite dimensional affine spaces.
Thus, ring epimorphisms correspond to embeddings of spaces. Ring
monomorphisms do not necessarily induce surjective maps of spectra: Only
the closure of aϕ(Spec B) coincides with Spec A. This follows from a trifle
more general result.
1.6.2. Proposition. For any ring homomorphism ϕ : A −→ B and an ideal
b ⊂ B, we have
aϕ(V (b)) = V (ϕ−1 (b)).
In particular, if Ker ϕ = 0, then aϕ(V (0)) = V (0), i.e., the image of Spec B
is dense in Spec A.
Proof. We may assume that b is a radical ideal, since V (r(b)) = V (b) and
ϕ−1 (r(b)) = r(ϕ−1 (b)). The set aϕ(V (b)) is the intersection of all closed sub-
sets containing aϕ(V (b)), i.e., the set of common zeroes of all the functions
f ∈ A which vanish on aϕ(V (b)). But f vanishes on aϕ(V (b)) if and only if
ϕ(f ) vanishes on V (b), i.e., if and only if ϕ(f ) ∈ b (since b is a radical ideal)
or, finally, if and only if f ∈ ϕ−1 (b). Therefore the closure we are interested
in is equal to V (ϕ−1 (b)). u
t
Now, let us give examples of ring monomorphisms for which aϕ(Spec B)
does not actually coincide with Spec A.
1.6 Topological properties of morphisms and Spm 35
Fig. 7
Indeed, aϕ maps a generic point into a generic one. The prime ideal
(f (T1 )) ⊂ A, where f 6= cT1 is an irreducible polynomial, is the pre-image
of the prime ideal (f (T1 )) (mod T1 T2 − 1) ⊂ B. Finally, T1 and T1 T2 − 1
/ aϕ(Spec B).
generate the ideal (1) = K[T1 , T2 ], and therefore (T1 ) ∈ u
t
a
In this example ϕ(Spec B) is open; but it may be neither open nor closed:
2) The projection of a hyperbolic paraboloid onto the plane. Consider the
homomorphism
Fig. 8
closed sets 14) . Such unions are called constructible sets; the image of a con-
structible set under aϕ is always constructible under the described conditions
(Chevalley’s theorem [AM]).
In terms of undetermined coefficients of a (finite) system of equations this
means that the condition for the system’s compatibility is the following:
For example, for the equation M T − N = 0, there are the two statements:
1. M 6= 0;
2. M = 0, N = 0.
1.6.3. Analogs of “finite-sheeted coverings” of Riemannian surfaces.
In the above examples something “escaped to infinity”. Let us describe an im-
portant class of morphisms aϕ for which this does not happen.
Let B be an A-algebra. An element x ∈ B is called integer over A if it satisfies
an equation of the form
and B is called integer over A if all its element are integer over A.
There are two important cases when it is easy to establish whether B is integer
over A.
Case 1. If B has a finite number of generators as an A-module, then B is integer
over A.
Indeed, if A is Noetherian ring, then, for any g ∈ B, the ascending sequence
of A-modules X
Bk = Ag i ⊂ B (1.51)
0≤i≤k
P
stabilizes. Therefore, for some k, we have g k ∈ Ag i , which provides us with
0≤i≤k−1
an equation of integer dependence.
14
A locally closed set is the intersections of a closed and an open set.
1.6 Topological properties of morphisms and Spm 37
The general
P
case reduces to the above one with the help of the following trick.
Let B = Afi . Let
1≤i≤n
X X
fi fj = akij fk , where akij ∈ A, and g = gi fi , where gi ∈ A; (1.52)
1≤k≤n 1≤i≤n
by A0 ⊂ A the minimal subring containing all the akij and gi and set
denote P
B0 = A0 fi . Obviously, A0 is a Noetherian ring, B is an A0 -algebra and g ∈ B0 .
1≤i≤n
Therefore g satisfies the equation of integer dependence with coefficients from A.
Case 2. Let G be a finite subgroup of the group of automorphisms of B and
A = B G the subring of G-invariant elements. Then B is integer over A.
Indeed, all the elementary symmetric
Q
polynomials in s(g), where s ∈ G, belong
to A for any g ∈ B, and g satisfies (g − s(g)) = 0. u
t
s∈G
2) if 0 ∈
/ S, then all the elements from j(S) are invertible in AS , otherwise
AS = 0;
3) every element from AS can be represented in the form j(f )/j(s) with
s ∈ S, f ∈ A.
Notice how drastically AS shrinks when we introduce zero divisors in S.
The following theorem is a main fact on rings of fractions, it describes a
universal character of localization.
1.6.4b. Theorem. Let S be a multiplicative system in a ring A and
j : a 7→ a/1 the canonical homomorphism A −→ AS . For any ring homomor-
phism f : A −→ B such that every element from f (S) is invertible, there exists
a unique homomorphism f 0 : AS −→ B such that f = f 0 ◦ j.
1.6.4c. Exercise. Prove the theorem.
Corollary. Let A be a ring, T and S its multiplicative systems such that
T ⊃ S. Then
1) the following diagram commutes:
jS
A; / AS 3 jS (a)/jS (t)
;; ¡
;; ¡¡
jT ;;
; ¡¡¡ (1.55)
À ¡¡¡ jT (S )
−1
²
AT 3 jT (a)/jT (t)
2) A[T −1 ] ∼
= A[S −1 ][jS (T )−1 ].
1.6.4d. Theorem. Let j : A −→ AS be the canonical homomorphism. Then
the induced map aj : Spec AS −→ Spec A homeomorphically maps Spec AS
onto the subset {x ∈ Spec A | px ∩ S = ∅}.
Proof. We may confine ourselves to the case 0 ∈
/ S. First, let us prove that
there is a one-to-one correspondence
px ∩ {f n | n ∈ Z} = ∅ ⇐⇒ f ∈
/ px . (1.57)
Therefore Spec A splits into the union of an open and a closed set as follows:
Here a certain “duality” between the localization and the passage to the
quotient ring is reflected. Considering D(f ) as Spec Af we “send V (f ) to
infinity”.
If S is generated by a finite number of elements f1 , . . . , fn , then
\
px ∩ S = ∅ ⇐⇒ x ∈ D(fi ) (1.59)
1≤i≤n
and in this case the image of Spec AS is open in Spec A. However, this is not
always the case; cf. Examples 1.6.2.
If S = A\px , then the image of Spec AS in Spec A consists of all the points
y ∈ Spec A whose specialization is x. It is not difficult to see that in general
this set is neither open nor closed in Spec A: It is the intersection of all the
open sets containing x.
The ring Apx = Ox contains the unique maximal ideal px whose spectrum
describes geometrically the “neighborhood of x” in the following sense: We
may trace the behavior of all the irreducible subsets of Spec A through x in
a “vicinity” of x. This is the algebraic version of the germ of neighborhoods
of x in the following sense: We can trace the behavior of the irreducible subsets
of Spec A passing through x in a “vicinity” of x. u
t
1.6.5. Theorem. Let ϕ : A ,→ B be a monomorphism and let B be integer
over A. Then aϕ(Spec B) = Spec A.
Proof. We first prove two particular cases; next we reduce the general state-
ment to these cases.
Case 1. B is a field. Then aϕ(Spec B) = {(0)} ⊂ Spec A and aϕ is
an epimorphism if and only if A has no other prime ideals, i.e., A is a field.
Let us verify this.
40 Ch. 1. Affine schemes
Fig. 9
as desired.
Case 2. A is a local ring. Then, under the conditions of the theorem, the
unique closed point of Spec A belongs to aϕ(Spec B) and, moreover, it is the
image under aϕ of any other closed point of Spec B.
Indeed, let p be a maximal ideal of A, q any maximal ideal of B. Then
B/q is a field integer over the subring A/(A ∩ q) which by the proved above
case 1, should also be a field. This means that A ∩ q is a maximal ideal in A,
and therefore A ∩ q = ϕ−1 (q) = p.
The general case. Let p ⊂ A be a prime ideal; we wish to show that
there exists an ideal q ⊂ B such that q ∩ A = p. Set S = A \ p.
Considering S as a subset of both A and B, we may construct the rings
of quotients AS ⊂ BS . Set pS = {f /s | f ∈ p, s ∈ S}. It is easy to see that
pS ⊂ AS is a prime ideal. It is maximal since AS \ pS consists of invertible
elements s/1. P
The ring BS is integer over AS since if f ∈ B satisfies f n + ai f i = 0,
0≤i≤n−1
then f /s ∈ BS satisfies
X
(f /s)n + ai /sn−i · (f /s)i = 0. (1.62)
0≤i≤n−1
1.6 Topological properties of morphisms and Spm 41
The pre-image of the point T1 = 0 of the line contains, clearly, the generic
point of the T2 -axis which is not closed in the plane.
unlike the closed subsets of the space Spec A, which only correspond to the
radical ideals.
The support of Y = Spec A/a ⊂ X is the space V (a); it is denoted by
supp Y .
To the projection A −→ A/a a scheme embedding Y −→ X corresponds;
it is called a closed embedding of a subscheme.
For any ring L, we had set
and called it the set of L-points of X (cf. definition 1.2.1). Clearly, the
L-points of a subscheme Y constitute a subset Y (L) ⊂ X(L) and the functor
L −→ Y (L) is a subfunctor of the functor L −→ X(L).
1.7 The closed subschemes and the primary decomposition 43
T
The
notation is justified by the fact that, for any T ring L, the set of L-points
Yi (L) of the intersection is naturally identified with Yi (L). Indeed, an L-point
i T i
ϕ : A −→ L belongs to Yi (L) if and only if ai ⊂ Ker ϕ for all i, which is equivalent
P i
to the inclusion ai ⊂ Ker ϕ. This argument shows that
i
\ \
supp Yi = supp Yi . (1.68)
i i
The sum a of all such ideals also satisfies (1.70) and is the unique maximal element
of the set of ideal satisfying (1.70); on
T
the other hand, all the elements of this set
are contained in ai , and therefore in ai .
i
Having failed to define the union
W
of subschemes, let us introduce a notion corre-
sponding to ∩ai : The quasiunion Yi of a family of closed subschemes Yi of a scheme
i
X is the subscheme corresponding to the intersection of all the ideals defining the
subschemes Yi .
It is important to notice that the quasiunion of the subschemes Yi does not
depend on the choice of the closed scheme containing all the Yi , inside of which we
construct this quasiunion.
44 Ch. 1. Affine schemes
The main aim of this section is to construct for the Noetherian affine schemes
the decomposition theory into “irreducible” in some sense components similar to
the one constructed above for Noetherian topological spaces. To this end, we will
use the quasi-union; on supports it coincides with the union (for finite families
of subschemes).
W S
1.7.3a. Lemma. supp Yi = supp Yi .
1≤i≤n 1≤i≤n
S
Proof. The inclusion ⊃ is already proved. Conversely, if x ∈
/ supp Yi , then,
1≤i≤n
for every i, there exists an element fi ∈ ai such that fi (x) 6= 0, therefore
Q
fi (x) 6= 0. Hence, x does not belong to the set of zeroes of all the functions
1≤i≤n
T W
from ai , which is supp Yi . Now, apply Exercise 1.3.4 1).
1≤i≤n 1≤i≤n
u
t
1.7.4. Irreducible schemes. Now, we have to transport to the subschemes the
notion of irreducibility. The first that comes to one’s mind, is to try to imitate the
definition of irreducibility for spaces.
An affine scheme X is said to be reducible if there exists a representation of the
form X = X1 ∨ X2 , where X1 , X2 are proper closed subschemes of X; it is said to
be irreducible otherwise.
An affine scheme X is said to be a Noetherian one if its ring of global functions
is Noetherian or, equivalently, if X satisfies the descending chain condition on closed
subschemes.
1.7.4a. Theorem. Any Noetherian affine scheme X decomposes into the quasiu-
nion of a finite number of closed irreducible subschemes.
Proof. The same arguments as at the end of sec. 1.4.7b lead to the result desired.
If X is reducible, we write X = X1 ∨ X2 and then decompose, if necessary, X1 and
X2 , and so on. Thanks to the Noetherian property, the process terminates. u
t
1.7.5. Primary schemes and primary ideals. The above notion of irre-
ducibility turns out to be too subtle. The following notion of primary affine
schemes is more useful: An ideal q ⊂ A is called primary if any zero divisor in
A/q is nilpotent. A closed subscheme is called primary if it is determined by
a primary ideal.
Proposition. Any irreducible Noetherian scheme is primary.
Remark. The converse statement is false. Indeed, let K be an infinite field.
Consider the ring A = K × V , where V is an ideal with the zero product. The
ideal (0) is primary and, for any subspace V 0 ⊂ V , the ideal (0, V 0 ) is primary.
At the same time, if dimK V > 1, there exists infinitely many representations
of the form (0) = V1 ∩ V2 , where Vi ⊂ V are proper subspaces, i.e., represen-
tations of the form X = Y1 ∨ Y2 , where X = Spec A. This deprives us of any
hope for the uniqueness of the decomposition into quasiunion of irreducible
subschemes. Primary subschemes behave a sight more nicely than irreducible
subschemes as we will see shortly.
1.7 The closed subschemes and the primary decomposition 45
W
Proof. Let Y = Yi , supp Yi = supp Yj for all i, j. Let Yi correspond
1≤i≤n T
to an ideal ai in the ring of global functions A on Y . Then ai = (0).
1≤i≤n
Consider a zero divisor f ∈ A. Let f g = 0, where g 6= 0 and g ∈ ai for some
i. Since ai is primary, f n ∈ ai for some n. But since V (ai ) = V (0), it follows
that ai consists of nilpotents and f is nilpotent; moreover, supp Yi = supp Y .
u
t
W
1.7.8. Theorem (A uniqueness theorem). Let X = Xi be an incom-
1≤i≤n
pressible primary decomposition of a Noetherian affine scheme X. The system
of generic points of irreducible closed sets supp Xi does not depend on such
a decomposition.
This system of generic points is denoted by Prime X (or Prime A if
X = Spec A) and is called the set of prime ideals associated with X (or A).
We will establish a more precise result giving an invariant characterization
of Prime X. Let X = Spec A and Xi = Spec(A/ai ) for an ideal ai ⊂ A.
Proposition. The following two statements are equivalent for any reduced
scheme X (for the general scheme, T the statement is only true when Ann f
in b) is replaced by {g ∈ A | gf ∈ ai }:
a) Any prime ideal p ⊂ A corresponds to a generic point of one of the sets
supp Xi .
b) There exists an element f ∈ A such that Ann f := {g ∈ A | f g = 0} is
primary and p is its radical.
Proof. a) =⇒ b). Let pj be the ideal of W a generic point of supp Xj . Clearly,
T
pj = r(aj ). Since the representation X = Xj is incompressible, ai 6⊃ aj ,
1≤j≤n j6=i
T
where 1 ≤ i ≤ n. Let us select an element f ∈ aj \ ai and show that Ann f is
j6=i
primary and pi its radical.
First of all, Ann(f mod ai ) consists only of nilpotents in A/ai ; therefore
Ann f ⊂ pi (since pi is the pre-image of the nilradical of A/aj with respect to the
naturalThomomorphism A −→ A/ai ). Besides, ai ⊂ Ann f because, by construction,
f ai ⊂ aj = (0); therefore r(Ann f ) = pi .
j
Now let us verify that all the zero divisors in A/ Ann f are nilpotents. Assume
the contrary; then there exist elements g, h ∈ A such that gh ∈ Ann f , h 6∈ Ann f
and g is not nilpotent modulo Ann f ; therefore g is not nilpotent modulo ai as well.
On the other hand, f gh = 0; and,
T since
g mod ai is not nilpotent, f h mod ai = 0
because ai is primary, i.e., f h ∈ aj ∩ ai = 0, implying h ∈ Ann f contrary to
j6=i
the choice of h. This contradiction shows that Ann f is primary.
b) =⇒ a). Let f ∈ A be an element such that T
Ann f is primary, p its radical.
Set si = (ai : f ) = {g ∈ A | gf ∈ ai }. Since ai = (0), it is easy to see that
T T i
Ann f = si and p = r(Ann f ) = r(si ). If f ∈ ai , then si = r(si ) = A. If, on
i i
1.7 The closed subschemes and the primary decomposition 47
X2
X1
Fig. 10
Therefore, innocent at the first glance, the space of an affine scheme may
hide in its depth a complicated structure of embedded primary subschemes
like the one illustrated on Fig. 11. The reader should become used to the
geometric reality of such a structure.
Fig. 11
Clearly, depicting nilpotents by arrows does not make possible to reflect the
details however precisely. It is only obvious that on the embedded components
the nilpotents grow thicker, thus giving away their presence.
1.7.10. Incompressible primary decompositions, cont. 2. The finite set
of prime ideals Spec A which is invariantly connected with every Noetherian ring
A has a number of important properties. In particular, it enables us to refine The-
orem 1.4.8
Theorem. An element f ∈ A is a zero divisor if and only if it vanishes
(as a function) on one of the components of the incompressible primary decom-
position of Spec A. T
In other words, the set of zero divisors of A coincides with p.
p∈Prime A
T
Proof. First, let us show that if f ∈
/ p , then Ann f = (0).
p∈Prime A
S
Let (0) = ai be the incompressible primary decomposition, where pi = r(ai ).
i
Let f g = 0. Since f ∈ / pi , it follows that ai being primary implies that g ∈ ai . This
is true for all i; therefore g = 0.
Conversely, let Ann(f ) = (0). Suppose that f ∈ pi ; since A is Noetherian, it
follows that f k ∈ pki ⊂ ai for some k ≥ 1; i.e., (ai : (f k )) = A. 15) On the other hand,
Ann(f k ) = (0) implying
\ \ \
(0) = Ann(f k ) = (aj : (f k )) = (aj : (f k )) ⊃ aj . (1.72)
j j6=i j6=i
15
The quotient (a : b) of two ideals in A is defined to be the set {x ∈ A | xb ⊂ a}. It
is easy to see that (a : b) is an ideal in A. In particular, the ideal (0 : b) is called
the annihilator of b and is denoted Ann(b).
1.8 Hilbert’s Nullstellensatz (Theorem on zeroes) 49
Proof of Theorem. We will successively widen the class of rings for which this
theorem is true.
a) Let K be algebraically closed. Then the set of closed points of Spec K[T1 , . . . , Tn ]
is dense.
The closure of the set of closed points coincides with the space of zeroes of all the
functions which vanish at all the closed points, and therefore it suffices to prove that
a polynomial F , which belongs to all the maximal ideals of K[T1 , . . . , Tn ], is iden-
tically zero. But such a polynomial satisfies F (t1 , . . . , tn ) = 0 for all t1 , . . . , tn ∈ K,
and an easy induction on n shows that F = 0 (here we actually use not closedness
even, but only the fact that K is infinite).
b) The same as in a) but K is not supposed to be algebraically closed.
Denote by K the algebraic closure of K. We have a natural morphism
The ring B is integer over K, and therefore thanks to the results of 7.4–7.5, we have
c) Theorem 1.8.1 holds for the rings A of finite type without zero divisors over K.
Indeed, by Noether’s normalization theorem there exists a polynomial subalgebra
B ⊂ A such that A is integer over B. By the already proved Spm B = Spec B, and
the literally same argument as in b) shows that Spm A = Spec A.
d) Theorem 1.8.1 holds for any rings A of finite type over a field.
Indeed, any irreducible component of Spec A is homomorphic to Spec A/p, where
p is a prime ideal. The ring A/p satisfies the conditions of c), hence, the closed points
are dense in all the irreducible components of Spec A, and therefore in the whole
space.
e) The same as in d) for the rings A of finite type over Z.
1.8.2. Lemma. No field of characteristic 0 can be a finite type algebra over Z.
Proof. If a field of characteristic 0 is a finite type algebra over Z, then thanks
to Noether’s normalization theorem there exist algebraically independent over Q
element t1 , . . . , tr ∈ K such that K is integral over R = Q[t1 , . . . , tr ]. By Proposi-
tion 1.6.6 the natural map Spm K −→ Spm Q[t1 , . . . , tr ] is surjective; since Spec K
contains only one closed point, so does Spec R, which is only possible if R = Q.
Therefore, K is integral over Q, and hence is a finite extension of the field Q.
Let x1 , . . . , xn generate K as a Z-algebra. Each of the xj is a root of a polynomial
with rational coefficients. If N is the LCM of the denominators of these coefficients,
then, as is easy to see, all the N xj are integral over Z; if y is the product of m of
the generators x1 , . . . , xn (perhaps, with multiplicities), then N m y is integral over
Z. Since all elements of K are linear combinations of such products with integer
coefficients, it follows that for any y ∈ K, there exists a natural m such that N m y
is integral over Z.
Now, let p be a prime not divisor of N . Since 1/p ∈ Q ⊂ K, we see that a
non-integer rational number N m /p is integral over Z; Exercise 1.6.7 shows that this
is impossible. u
t
To complete the proof in case e), denote by ϕ : Z → A a natural homomor-
phism and show that a ϕ(Spm A) ⊂ Spm Z. Indeed, otherwise there exists a maximal
idealp ⊂ A such that ϕ−1 (p) = (0); so Z can be embedded into the field A/p (and
1.8 Hilbert’s Nullstellensatz (Theorem on zeroes) 51
therefore we
N have acquired nilpotents which were previously lacking. The space
of Spec L L consists, however, of one point. ut
K
1.9.5. Examples. Let us give examples which are absolutely parallel to the
set-theoretical construction.
i1 i2
1) Let X = Spec A, let Y1 −→ X and Y2 −→ X be two closed subschemes
of X determined by ideals a1 , a2 ⊂ A. Thanks to results of sec. 1.7.2 their
1.10 The vector bundles and projective modules 55
id /X
X
id (1.81)
² ²
X /S
Proof. Writing down all the necessary diagrams we see that δ = aµ. Since
µ is surjective, its kernel determines a closed subscheme isomorphic to the
image of δ. u
t
The scheme ∆X is said to be the (relative over S) diagonal.
χ0 = id ⊗χ : M 0 = A0 ⊗A M −→ A0 ⊗A B (1.84)
ψ
X0 B /X
BB }}>
BB }} (1.86)
B
χ BB }} ∗
! }} i
Xf
commutes.
This implies that the family of vector spaces Y −→ X is trivial at
x ∈ Spec A = X if and only if, for any element f ∈ A such that f (x) 6= 0, the
family induced over Xf is trivial. Translating this into the language of modules
we find a simple condition which will be used in what follows.
Corollary. An A-module M determines a family of vector spaces trivial at
x ∈ Spec A if and only if there exists f ∈ A such that f (x) 6= 0 and such that
the Af -module Mf := Af ⊗A M is free.
A module M satisfying conditions of Corollary for all the points x ∈ Spec A
is said to be locally free.
Proof of Proposition. a) This is a particular case of Theorem 1.6.4d
b) This statement expresses the known universal property of the rings
of quotients. Indeed, let ψ : A −→ A0 be a ring homomorphism such that
a
ψ(Spec A0 ) ⊂ D(f ). This means that f does not belong to any of the ideals
ψ −1 (p), where p ∈ Spec A0 , i.e., ψ(f ) does not vanish on Spec A0 . Therefore
ψ(f ) is invertible in A0 .
In the category of such A-algebras, the morphism A −→ Af is universal
object (see 1.6.4b) which proves the statement desired.
In particular, if D(f ) = D(g), then the ring Af is canonically isomorphic
to Ag . u
t
58 Ch. 1. Affine schemes
M −→ MS , m 7→ m/1, (1.87)
where F0 , F1 are free modules. Tensoring the sequence by AS we get the exact
sequence
ϕS ψS
(F1 )S −→(F0 )S −→MS −→ 0 (1.89)
(see [Lang]). Here we set ϕS = idAS ⊗A ϕ, and so on.
Let m/s = 0 for m = ψ(n), where n ∈ F0 . Then ψS (n/s) = 0; this implies
1.10 The vector bundles and projective modules 59
as desired. u
t
Observe that we never used the Noetherian property.
1.10.5b. Corollary. Let M be a Noetherian A-module, f ∈ A. There exists
an integer q > 0 such that f q m = 0 for all m ∈ Ker(M −→ Mf ).
Proof. Select the needed value qi for every of a finite number of generators
of the kernel, and set q = max qi . u
t
i
1.10.6. Tensoring exact sequences. In the proof of Lemma 1.10.5 we
have used the following general property of the tensor product:
Tensoring sends short exact sequences into the sequences exact
(1.92)
everywhere except the leftmost term.
Tensoring by AS , however, possesses a stronger property: It completely pre-
serves exactness; this means AS is what is called a flat A-algebra.
ϕS ψS
1.10.6a. Proposition. The sequence MS −→NS −→PS of AS -modules is ex-
ϕ ψ
act for any exact sequence M −→N −→P of A-modules.
Proof. ψ ◦ ϕ = 0 =⇒ ψS ◦ ϕS = 0 =⇒ Ker ϕS ⊃ Im ϕS .
Conversely, let n/s ∈ Ker ψS ; then ψ(n/s) = 0 implying, thanks to the
above, tψ(n) = 0 for some t ∈ S. Therefore tn = ϕ(m) implying
as desired. u
t
1.10.7. Lifts of Af -module homomorphisms. Let ϕ : M −→ N be
an A-module homomorphism. For any f ∈ A, we have an induced Af -module
homomorphism ϕf : Mf −→ Nf . We will say that ψ : Mf −→ Nf can be lifted
to ϕ : M −→ N if ϕf = ψ .
Lemma. Let F be a free Noetherian A-module, M a Noetherian A-module,
f ∈ A, Mf a free A-module. Then, for any homomorphism ϕ : Mf −→ Ff ,
there exists an integer q such that the homomorphism f q ϕ : Mf −→ Ff can
be lifted to a homomorphism M −→ F .
Proof. First of all, Ff is free and has a finite number of generators, so that
ϕ is given by a finite number of coordinate Af -morphisms Mf −→ Af . If
being multiplied by an appropriate power of f each of them can be lifted to
60 Ch. 1. Affine schemes
0 −→ K −→ Ang −→ Mg −→ C −→ 0 (1.98)
is uniquely solvable in A. u
t
Fig. 12
2
The B-module CovectB/A = IB/A /IB/A is called the module of (relative)
17)
differentials of the A-algebra B.
By Proposition 1.9.6 the ideal I determines the diagonal subscheme
∆X ⊂ X × X, where X = Spec B, S = Spec A. Due to the interpretation
S
from sec. 1.11.1 the module CovectB/A represents the co-normal to the diag-
onal.
In Differential Geometry, the normal bundle to the diagonal ∆ is isomor-
phic to the tangent bundle to the manifold X itself. Indeed, transport a vector
field along one of the fibers of the product X ×X “parallel” to the diagonal; we
get a vector field everywhere transversal to the diagonal, see Fig. 12. There-
fore CovectB/A is a candidate for the role of the module “cotangent ” to X
along the fibers of the morphisms X −→ S.
On the other hand, in the interpretation of nilpotents given in §§ 1.5 as
an analog of the elements of the “tangent module” to X (over S), we have
already considered the B-module DB/A of derivations of the A-algebra B
(a vector field on X, i.e., a section of the tangent bundle can be naturally
interpreted as a derivation of the ring of functions on X).
Remark. The higher order “differential neighborhoods of the diagonal” are
n
represented by the schemes Spec(B ⊗A B)/IB/A . They replace the spaces
of jets of the germs of the diagonal considered in differential geometry.
Define a map d = dB/A : B −→ CovectB/A by setting
2
d(b) = (b ⊗ 1 − 1 ⊗ b) (mod IB/A ). (1.106)
In the differential geometry, the tangent and the co-tangent bundles are
dual to each other. In the algebraic setting (over finite fields), this is not the
case, generally: Only “a half” of the duality is preserved:
Therefore, DB/A is recovered from CovectB/A but not the other way round.
This explains the advantage of differentials as compared with derivatives. 18)
Lemma. 1) d is an even A-derivation, and d(ϕ(a)) = 0 for any a ∈ A, where
ϕ : A −→ B is the morphism that defines the A-algebra structure.
2) Let {bi | i ∈ I} be a system of generators of the A-algebra B. Then
{dbi | i ∈ I} is a system of generators of the B-module CovectB/A .
Proof. 1) is subject to a straightforward verification.
2) Notice that
X X
bi ⊗ b0i ∈ IB/A ⇐⇒ bi b0i = 0 ⇐⇒
i i
X X X X (1.112)
bi ⊗ b0i = bi ⊗ b0i − bi b0i ⊗ 1 = bi ⊗ 1(1 ⊗ b0i − b0i ⊗ 1).
i i i i
where
Smblk (M, N ) = Diff k (M, N )/ Diff k−1 (M, N ). (1.111)
If M = N , then, clearly, Diff (M ) = Diff (M, M ) is an associative algebra with
respect to the product and a Lie algebra — denoted by diff(M ) — with respect
to the bracket of operators. Smbl(M ) is a commutative R-algebra with respect to
the product. The bracket in diff(M ) induces a Lie algebra structure in Smbl(M )
(the Poisson bracket), as is not difficult to see, this Lie algebra is isomorphic to
the Poisson Lie algebra in dim M variables over R.
1.12 The differentials 69
such that u and v are mutually inverse. For this, first define a derivation
d0 : (B → B ⊗B CovectB/A )/ Im δ (1.120)
by setting
it follows that v and u are mutually inverse in some cases of our modules,
proving the desired. u
t
1.12.3a. Remark. The difference of the above constructions from similar
ones in the differential geometry is crucial: It well may happen that Ker δ 6= 0
even if the subscheme Y ,→ X is regularly embedded. For example, let
X = Spec Z and Y = Spec Z/p, where p is a prime, S = Spec Z. Then
CovectX/S = 0 and CovectY /S = 0; whereas (p)/(p)2 is a one-dimensional
linear space over Z/p.
Informally speaking, 19)
0 → F → P → P/F → 0 (1.124)
i /P / P/F / 0.
F
1.13.4. Lemma. F 1 6= 0.
Proof. Indeed, j(F 1 ) = pP ∩ F/pF . Let f = pg. Since g ∈
/ Ann P/F ∩ A,
we have gP 6⊂ F =⇒ pgP 6⊂ pF (because p is torsion-free). But
pgP = f P ⊂ pP ∩ F , so, moreover, pP ∩ F 6⊂ pF . u
t
The last step requires some additional arguments.
1.13.5. Lemma. There exists a free A[T ]-submodule F1 ⊂ F with a free
direct complement and such that k[T ] ⊗ F1 = j(F 1 ).
Speaking informally the sequence (1.125) can be lifted to a split exact
sequence of free A[T ]-modules.
1.13.6. Deduction of Theorem 1.13.2. Let F1 ⊂ F be a submodule
whose existence is claimed in Lemma 1.13.5, F2 ⊂ F its free direct comple-
ment. Since F1 /pF1 = Ker i, all elements F1 ⊂ F ⊂ P are divisible by p inside
P . Set
F10 = {m ∈ P | pm ∈ F1 }.
Clearly, F10 is free (the multiplication by p determines an isomorphism
F10 ' F1 ) and is strictly larger than F1 (by Lemma 1.13.4). Therefore the
1.13 Digression: Serre’s problem and Seshadri’s theorem 73
(This equality is obviously well-defined even if it is not known that ν(pa ) and
n(pb ) are finite).
In particular, n(pa ) ≤ ν(pa ) and it suffices to prove that ν(pa ) is finite.
We identify Spec A with a closed subset in Spec Z[T1 , . . . , Tn ]; then N (x)
does not depend on whether we consider x as belonging to Spec A or to
Spec Z[T1 , . . . , Tn ]. Therefore, in obvious notation,
1.14 Digression: ζ-function of a ring 75
where, clearly, pna is just the number of geometric points of the n-dimensional
affine space over a field of pa elements. u
t
1.14.2. ζ-functions of arithmetic rings. Define the ζ-function of any
arithmetic ring A first formally, by setting
Y 1
ζA (s) = −s
. (1.129)
1 − N (x)
x∈Spm A
The relation of the ζ-function with n(pa ) and ν(pa ) is given by the following
obvious identity
Y Y 1 Y
ζA (s) = −as n(pa )
= ζA/pA (s) (1.131)
(1 − p )
p 1≤a<∞ p
or 0 X
ZA (t)
= ν(pa )ta−1 (1.135)
ZA (t)
a
for sufficiently large a with some fixed constants n, ri . Since the νa are the
numbers of solutions of a system of equations with values in finite fields
of growing degree, the statement on rationality bears a direct arithmetic mean-
ing.
1.14.3. Frobenius morphism. In any study of the ζ-function of a ring A
over a field k of characteristic p the following circumstance is of fundamental
importance: ν(pk ) can be viewed as the number of fixed points of a power
of a certain map F acting on the set of geometric points of A.
A Frobenius morphism F : A −→ A is the map g 7→ g p for any g ∈ A,
where p = Char k.
The same term — Frobenius morphism — is applied to the correspond-
ing morphism of spectra, aF , to its powers, F n , to ( aF )n , and to the maps
of some other objects induced by these maps. In particular, let F̄p be the al-
gebraic closure of the Galois field of characteristic p. Then F induces a map
a
F : A(F̄p ) −→ A(F̄p ) of F̄p -points of A into itself.
Proposition. A(Fpb ) coincides with the set of fixed points of F b .
Proof. Let ϕ ∈ A(F̄p ) and let ϕ be represented by ϕ : A −→ F̄p ; let F b (ϕ)
b
be represented by f 7→ ϕ(f )p for any f ∈ A.
The condition ϕ ∈ A(Fpb ) means that Im ϕ ⊂ Fpb ⊂ F̄pb , i.e.,
b
ϕ(f )p = ϕ(f ) for all f . Therefore aF ϕ = ϕ.
The converse statement follows from the Galois theory: Fpb is the field
of invariants for F b . u
t
1.14.4. Lefschetz formula. If an endomorphism F acts on a compact to-
pological space V , then the number ν(F ) of its fixed points (appropriately
defined) satisfies the following famous Lefschetz formula:
X
ν(F ) = (−1)i tr F |H i (V ) , (1.137)
0≤i≤dim V
1.14 Digression: ζ-function of a ring 77
where the summands are the traces of linear operators induced by F on the
spaces of cohomology of V with complex coefficients.
The role of compact topological spaces V is played in our setting by smooth
projective schemes. For them, the essential part of Weil’s conjectures states
that the numbers ν(pa ) are always expressed by Lefschetz type formulas. So
far, we have dealt with Euler’s products and Dirichlet series purely formally.
Now, let us study a little their convergence.
Let A be an arithmetic ring, {xi }i∈I the set of generic points of its irre-
ducible components. Define the dimension of A setting
(
maxi (tr. deg. k(xi )) + 1 if Z ⊂ A
dim A = (1.138)
maxi (tr. deg. k(xi )) if Char A > 0
so the Euler product for A converges if Re s > max dim Ai and in this domain
i
satisfies
X
| ln ζA (s)| ≤ α ln(1 − pni −σ )−1 , where ni = dim Ai . (1.140)
i
is the usual ζ-function with shifted argument whose Euler product, as is well
known, converges absolutely for Re (s − n) > 1, i.e., Re s > n + 1 = dim A.
e) Let A be a ring without zero divisors containing Z. If we can find
a subring Z[T1 , . . . , Tn ] of A, over which A is integer, the arguments from
b) bring about the result. Regrettably, this is not always possible; we can,
however, remedy the situation for the price of localization modulo a finite
number of primes.
More exactly, let us apply Noether’s normalization theorem to A0 = Q⊗Z A
and find a subring Q[T1 , . . . , Tn ] in A0 over which A is integer. Multiplying,
if necessary, Ti by integers, we can assume that Ti ∈ A. Any element of A
over Z[T1 , . . . , Tn ] satisfies an equation whose highest coefficient is an integer.
Consider the set of prime divisors of all such highest coefficients for a finite
system of generators of A over Z[T1 , . . . , Tn ] and denote by S the multiplicative
system generated by this set. Then AS is integer over ZS [T1 , . . . , Tn ], and
Y Y
ζA (s) = ζA/(p) (s) ζA/(p) (s). (1.142)
p∈S p6∈S
The set {p | p ∈ S} is finite and from the above ζA/(p) (s) converges uniformly
for Re s > dim A/(p) ≥ dim A − 1.
For p 6∈ S, we have ζA/(p) (s) = ζAS /(p) (s), and the constant α for the pair
of rings ZS [T1 , . . . , Tn ]/(p) ⊂ AS /(p) determined in b) can be chosen to be
independent of p. Indeed, the class modulo p of the fixed system of integer
generators AS over ZS [T1 , . . . , Tn ] gives a system of generators of AS /(p) for all
p 6∈ S. Therefore the second (infinite) product in (5) for σ = Re s > dim A/(p)
is majored by Y
(1 − pn−σ )−α ,
p6∈S
Prove that either S or P \ S is finite. For the integer domain not of finite type
over Z, give an example when both S and P \ S are infinite.
We will prove simultaneously that (1) the map B 7→ Aut(B 0 /B) is a functor
and (2) this functor is representable.
Select a free basis e1 , . . . , en of K 0 over K. In this basis the multiplication
law in K 0 is given by the formula
X
ei ej = ckij ek . (1.146)
1≤k≤n
N L
Denote e0i := 1 ei ; then B 0 = B e0i , and any endomorphism t of the
K 1≤i≤n
B-module B 0 is given by a matrix (tij ), where tij ∈ B and 1 ≤ i, j ≤ n. The
1.15 The affine group schemes 81
If B has no zero divisors, then the B-points of this K-algebra have a simple
structure: Since T2 must not vanish, T1 vanishes implying that the possible
values of T2 in the quotient ring are ±1. As the conventional Galois group
this√group is isomorphic to Z/2; the automorphisms simply change the sign
of a.
The following case illustrates that when B does have zero divisors the
group of B-points of Aut K × /K can be much larger.
Case 2: Char K = 2. The functor of automorphisms is represented by the
K-algebra
82 Ch. 1. Affine schemes
Obviously, K[T ]/((T − 1)p ) is a local Artinian algebra of length p, and its
spectrum should be considered as a “point of multiplicity p”. This is a nice
23
And also for the group superschemes over any fields.
84 Ch. 1. Affine schemes
agreement with our intuition: All the roots of unity of degree p are glued
together and turn into one root of multiplicity p.
More generally, set
n
µpn /K = Spec K[T ]/((T − 1)p ). (1.159)
and the conventional axioms of the associativity, the left inverse and the left
unit have, respectively, the form
m : X × X −→ X (multiplication, x, y 7→ xy)
i: X −→ X (inversion, x 7→ x−1 ) (1.162)
u: E −→ X (unit, the embedding of E)
that satisfy the axioms of associativity, left inversion and left unit, respectively,
expressed as commutativity of the following diagrams:
(m, idX )
X ×X ×X / X ×X
δ m (1.164)
• u ²
X /E /X
(•, idX ) (u, idX )
X ×O X / E×X / X ×X
δ m (1.165)
idX idX ²
X /X /X
24
An object E is said to be final if card(Hom(X, E)) = 1 for any X ∈ Ob C.
25
Here we denote the group unit by 1; it is the image of E.
1.15 The affine group schemes 85
u∗ : A −→ K co-unit
86 Ch. 1. Affine schemes
which satisfy the axioms of co-associativity, left coinversion and left counit,
respectively, expressed in commutativity of the following diagrams:
∗
m ⊗ idA
A ⊗ AO ⊗ A o A ⊗O A
idA ⊗m ∗
m∗ (1.168)
∗
m
A⊗A o A
∗
i ⊗ idA
A⊗A o A ⊗O A
µ m∗ (1.169)
² u ∗
Ao Ko A
m∗ (T ) = T ⊗ 1 + 1 ⊗ T, i∗ (T ) = −T, u∗ (T ) = 0. (1.171)
• Grf , the subcategory of Gr, whose objects are finite (as the subscript
indicates) groups;
• Ab, the subcategory of Gr of Abelian groups;
• Aff Sch, the category of affine schemes and their morphisms;
• Rings, the category of (commutative) rings (with unit) and their unit
preserving homomorphisms;
• A-Algs, the category of algebras over an algebra A;
• A-Mods, the category of (left) A-modules over a given algebra A;
• Man, the category of manifolds and their morphisms.
The second group of examples. The objects in this series of examples
are still structured sets but the morphisms are no longer structure-preserving
maps of these sets. (No fixed name for some of these categories).
• The main category of homotopic topology: Its object are topological
spaces, the morphisms are the homotopy classes of continuous maps (see, e.g.,
[FFG]).
• Additive relations (See, e.g., [GM]): Its objects are Abelian groups.
A morphism f : X → Y is any subgroup of X × Y and the composition
of ϕ : X → Y and ψ : Y → Z is given by the relation
• /•d • // • k
•? +•
?? ÄÄ
/• ?? Ä
• ?? ÄÄÄ
? ÄÄÄ
•
Fig. 13
Q Q
Ob Ci = Ob Ci ;
³Q Q ´ Q (1.179)
HomQ Ci Xi , Yi = HomCi (Xi , Yi )
such that FX,Z (ϕψ) = FY,Z (ϕ)FX,Y (ψ) for any ϕ, ψ ∈ Hom C provided ϕψ is
defined.
A functor from C◦ into D is called a contravariant functor from C into D.
A functor F : C1 × C2 → D is called a bifunctor, and so on.
Given categories C, D, E and functors F : C → D and G : D → E, define
GF : C → E composing the constituents of F and G in the usual set-theoretical
sense.
The most important examples of functors are just “natural constructions”:
(Co)homology and homotopy are functors Top → Ab; characters of finite
groups constitute a functor Grf → Rings. These examples are too meaningful
to discuss them in passing.
1.16.6. Examples of presheaves. A presheaf of sets (groups, rings, alge-
bras, superalgebras, R-modules, and so on) is a contravariant functor from
the category TopX of open sets of a topological space X with values in Sets
(or Gr, Rings, Algs, Salgs, R-Mods, and so on, respectively).
Let (V, A, e) be a scheme of a diagram, D and DC the associated categories.
A functor from D into a category E is called a diagram of objects from E (of
type (V, A, e)). A functor from DC into E is a commutative diagram of objects.
For an ordered set I considered as a category, a functor from I into a cat-
egory C is a family of objects from C indexed by the elements from I and
connected with morphisms so that these objects constitute either a projective
or an inductive system in C.
Given two functors F, G : C → D, a functor morphism f : F → G is a set
of morphisms f (X) : F (X) → G(X) (one for each X ∈ C) such that for any
ϕ ∈ HomC (X, Y ) the following diagram commutes:
f (X)
F (X) / G(X)
to any morphism ϕ◦ : Y2◦ −→ Y1◦ the functor PX assigns the map of sets
PX (Y2◦ ) → PX (Y1◦ ) which sends ψ : Y2 −→ X into the composition
ϕψ : Y1 −→ Y2 → X.
To any ϕ ∈ HomC (X1 , X2 ), there corresponds a functor morphism
Pϕ : PX1 −→ PX2 which to any Y ∈ C assigns
ϕψ : Y ◦ −→ X1 −→ X2 . (1.184)
Clearly, Pϕψ = Pϕ Pψ .
2) Similarly, define P X ∈ C∗ by setting
P ϕ (Y ) : P X2 (Y ) −→ P X1 (Y ) (1.186)
ψϕ : X1 −→ X2 −→ Y. (1.187)
Clearly, P ϕψ = P ψ P ϕ .
A functor F : C◦ −→ Sets (resp. a functor F : C −→ Sets) is said to be
representable(resp. corepresentable if it is isomorphic to a functor of the form
PX (resp. P X ) for some X ∈ C; then X is called an object that represents F .
1.16.8a. Theorem. 1) The map ϕ 7→ Pϕ defines an isomorphism of sets
HomC (X, Y ) ∼
= HomC∗ (PX , PY ). (1.188)
g(x)
PX (X) / PY (X)
Fig. 1
But, on the other hand, only local geometric objects are associated with
rings; so in order to construct, say, a projective space we have to glue it from
affine ones. Let us learn how to do so.
The gluing procedure we are interested in may be described, topologically,
as follows: Let X1 , X2 be two topological spaces, U1 ⊂ X1 , U2 ⊂ X2 their
open subsets and f : U1 −→ U2 an isomorphism. Then we construct a set
X = (X1 ∪ X2 )/Rf ,
98 Ch. 2. Sheaves, schemes, and projective spaces
where Rf is the equivalence relation which identifies the points that corre-
spond to each other with respect to f . On X, a natural topology is induced,
and we say that X is the result of gluing X1 with X2 by means of f .
On Fig. 1, the two ways to glue two affine lines, X1 = X2 = R with
U1 = U2 = R \ {0}, are illustrated. The results are:
a) The line with a double point (here f = id);
b) P1 , the projective line (here f (x) = x−1 ). Clearly, P1 is homeomorphic to
the circle S 1 .
When we try to apply this construction to the spectra of rings we immedi-
ately encounter the above-mentioned circumstance, namely that the topologi-
cal structure of the open sets reflects but slightly the algebraic data which we
would like to preserve and which is carried by the ring A itself. The theories
of differentiable manifolds and analytic varieties suggest a solution.
In order to glue a differentiable manifold from two balls U1 and U2 , we re-
quire that the isomorphism f : U1 −→ U2 which determines the gluing should
not be just an isomorphism of topological spaces but should also preserve
the differentiable structure. This means that the map f ∗ , sending continuous
functions on U2 into continuous functions on U1 , must induce an isomor-
phism of subrings of differentiable functions, otherwise the gluing would not
be “smooth”.
The analytic case is similar.
Therefore we have to consider the functions of a certain class which are
defined on various open subsets of the topological space X.
The relations between the functions on different open sets are axiomized
by the following definition.
2.1.1. Presheaves. Fix a topological space X. Let P be a law that to every
open set U ⊂ X assigns a set P(U ) and, for any pair of open subsets U ⊂ V ,
V
there is given a restriction map rU : P(V ) −→ P(U ) such that
1) P(∅) consists of one element,
W V
2) rU = rU ◦ rVW for any open subsets (briefly: opens) U ⊂ V ⊂ W .
V
Then the system {P(U ), rU | U, V are opens} is called a presheaf (of sets)
on X.
The elements of P(U ), also often denoted by Γ (U, P), are called the sections
of the presheaf P over U ; a section may be considered as a “function” defined
over U .
Remark. Axiom 1) is convenient in some highbrow considerations of category
theory. Axiom 2) expresses the natural transitivity of restriction.
2.1.2. The category TopX . The objects of TopX are open subsets
of X and morphisms are inclusions. A presheaf of sets on X is a functor
P : Top◦X −→ Sets.
2.1 Basics on sheaves 99
From genuine functions we can construct their products, sums, and mul-
tiply them by scalars; similarly, we may consider presheaves of groups, rings,
and so on. A formal definition is as follows:
Let P be a presheaf of sets on X; if, on every set P (U ), there is given
an algebraic structure (of a group, ring, A-algebra, and so on) and the re-
V
striction maps rU are homomorphisms of this structure, i.e., P is a functor
◦
TopX −→ Gr (Ab, Rings, A-Algs, and so on), then P is called the presheaf
of groups, rings, A-algebras, and so on, respectively.
Finally, we may consider exterior composition laws, e.g., a presheaf of mod-
ules over a presheaf of rings (given on the same topological space).
2.1.2a. Exercise. Give a formal definition of such exterior composition
laws.
2.1.3. Sheaves. The presheaves of continuous (infinitely differentiable, ana-
lytic, and so on) functions on a space X possess additional properties (of “an-
alytic continuation” type) which are axiomized in the following definition.
A presheaf P on a topological space X is called a sheaf if it satisfies
S the
following condition: For any open subset U ⊂ X, its open cover U = Ui ,
i∈I
and a system of sections si ∈ P(Ui ), where i ∈ I, such that
Ui Uj
rU i ∩Uj
(si ) = rUi ∩Uj
(sj ) for any i, j ∈ I, (2.1)
U
there exists a section s ∈ P(U ) such that si = rUi
(s) for any i ∈ I, and such
a section is unique.
In other words, from a set of compatible sections over the Ui a section
over U may be glued and any section over U is uniquely determined by the
set of its restrictions onto the Ui .
Remark. If P is a presheaf of Abelian groups, the following reformulation
of the above condition is useful: S
A presheaf P is a sheaf if, for any U = Ui , the following sequence
i∈I
of Abelian groups is exact
ϕ Y ψ Y
0 −→ P(U ) −→ P(Ui ) −→ P(Ui ∩ Uj ), (2.2)
i∈I i,j∈I
U
ϕ(s) = (. . . , rU , (s), . . .)
i
U
(2.3)
Ui
ψ(. . . , si , . . . , sj , . . .) = (. . . , rUi ∩Uj
(si ) − rUij∩Uj (sj ), . . .).
2.1.4. The relation between sheaves and presheaves. The natural ob-
jects are sheaves but various constructions with them often lead to presheaves
which are not sheaves.
Example. (This example is of a fundamental importance for the cohomology
theory). Let F1 and F2 be two sheaves of Abelian groups and F1 (U ) ⊂ F2 (U )
with compatible restrictions, i.e., F1 ⊂ F2 . As is easy to see, the set of groups
P(U ) = F1 (U )/F2 (U ) and natural restrictions is a presheaf but, in general,
not a sheaf. Here is a particular case:
Let X be a circle, O the sheaf over X for which O(U ) is the group of R-
valued continuous functions on U , and Z e ⊂ O the “constant” presheaf for
e ) = Z for each non-empty U .
which Z(U
The presheaf for which
e ) for every open set U ∈ X
P(U ) = O(U )/Z(U (2.4)
Fig. 2
2.1 Basics on sheaves 101
1
Recall the definition (see. [StE]). Let the objects of a category be sets with alge-
braic structures (such as groups, rings, modules (over a fixed ring), and so on)
and morphisms be the morphisms of these structures. Let (I, ≤) be an inductive
set. Let {Ai | i ∈ I} be a family of objects enumerated by elements of I and
for all i ≤ j, let a family of homomorphisms fij : Ai −→ Aj be given having the
following properties:
a) fii = id |{Ai } ,
b) fik = fjk ◦ fij for all i ≤ j ≤ k.
Then the pair (Ai , fij ) is said to be directed (inductive) system over I. The set
of inductive (or direct) limit A of the inductive system (Ai , fij ) is defined as the
quotient of the disjoint union of the sets A0i modulo an equivalence relation ∼:
a .
lim Ai = Ai ∼,
−→
i
The sheaf P+ is called the sheaf associated with P. Clearly, the algebraic
structures are transplanted from P to P+ .
Proof is a straightforward verification of definitions; so we leave it to the
reader.
2.1.7. Another definition of sheaves. Being equivalent to the previous
one, given in sec. 2.1.3, it is sometimes more convenient to grasp intuitively.
A sheaf F over a topological space X is a pair (YF , r), where YF is a topo-
logical space and r : YF −→ X is an open continuous map onto X such that,
for any y ∈ YF , there exists its open neighborhood (in YF ) and r is a local
homeomorphism in this neighborhood.
The relation between this definition and the previous one is as follows.
Giving r : YF −→ X, we define a sheaf as in sec. 2.1.3: By setting F(U ) to be
the set of local sections of YF over U , i.e., the maps s : U −→ YF such that
r ◦ s = id |U .
Conversely, if F is given by its sections over U for all U ⊂ X, set
[
YF = Fx , (2.7)
x∈X
One should bear in mind that, even for Hausdorff spaces X, the spaces YF
are not Hausdorff, generally.
Fig. 3 depicts a part of the space YF corresponding to the sheaf of continu-
ous functions on the segment [0, 1]. To the graph of every continuous function
an open subset of YF corresponds and the sections corresponding to the graphs
of functions do not intersect in YF unless their graphs intersect.
If the functions f1 , f2 coincide over a (necessarily closed !) set Y , then the
corresponding sections of YF coincide over the interior of Y . In the space YF ,
the points over a (or b) belonging to sections 3 and 4 are distinct, but any
two neighborhoods of these points intersect. u
t
2.2 The structure sheaf on Spec A 103
Fig. 3
Ox = {f /g | f, g ∈ A, g ∈
/ px } (2.10)
2.2.3. The structure sheaf on Spec A: The general case. If A has zero
divisors, it has no quotient field. Therefore an algebraic formalism necessary
for a correct definition of quotient rings and relations between them becomes
more involved. Nevertheless, a sheaf is actually introduced in the same way
as for the ring without zero divisors and with the same results.
Constructing a sheaf on Spec A we have to study the dependence of AS on
S, and the ring homomorphisms AS −→ AS 0 for different S and S 0 . Theorem
1.6.4b stating a universal character of localization, is foundational in what
follows.
2.2.4. The structure sheaf OX over X = Spec A. For every x ∈ X, set
(see sec. 1.6.2)
Ox := AA\px = Apx .
For any open subset U ⊂ X, define the ring of sections of the presheaf OX
over U to be the subring Y
OX (U ) ⊂ Ox (2.14)
x∈U
P
implying that the image of ai gi /f n in Ahj is precisely gj /hj .
1≤i≤r
Therefore the compatible local sections over D(hj ) are the restrictions
of one element from Af , as required. u
t
The sheaf over X = Spec A described above will be sometimes denoted
e The pair (Spec A, A)
by A. e consisting of a topological space and a sheaf over
e
it determines the ring A thanks to Theorem 2.2.4: Namely, A = Γ (Spec A, A).
This pair is the main local object of the algebraic geometry.
i.e., the domain of fU∗ (g) is f −1 (U ), and fU∗ (g) is constant on the pre-image
of every y ∈ U .
In algebraic geometry, the spaces are not Hausdorff ones and their struc-
ture sheaves are not readily recognized as sheaves of functions. Therefore
108 Ch. 2. Sheaves, schemes, and projective spaces
(ϕj |Xi ∩Xj )−1 ◦ θij ◦ ϕi |Xi ∩Xj = id for all i, j. (2.28)
2.3.3a. Projective spaces. Let K be a ring. Define the scheme PnK , the
n-dimensional projective space over K.
Let T0 , . . . , Tn be independent variables. Set
h i h i
T0 Tn T0 Tn
Ui = Spec K ,..., , Uij = Spec K ,..., ⊂ Ui , (2.29)
Ti Ti Ti Ti Tj /Ti
As above, the rings of functions over Uij and Uji can be identified with the
ring whose elements are of the form f (T0 , . . . , Tn )/Tia Tjb , where f is now
an inhomogeneous polynomial, the power of its lowest term being a + b.
Denote by X the scheme obtained after gluing up the Ui and identify
the Ui and Uij with the corresponding open sets in X. We have
[
X= Ui , Uij = Ui ∩ Uj . (2.31)
0≤i≤n
Furthermore,
h i
T0 Tn
V (Ti ) = Spec K T0 , . . . , Tn , , . . . , /(Ti ) (2.35)
Ti Ti
h i
T0 Tn
The ring K T0 , . . . , Tn , , . . . , consists of the elements of the form
Ti Ti
f (T0 , . . . , Tn )
, where f is a polynomial
Tia (2.36)
the degree of its lowest terms being ≥ a.
Hence the ideal (Ti ) of this ring consists of the same elements, but with the
degree of the lowest terms of the numerator being ≥ a + 1. Therefore it is
easy to see that
h i h i
T0 Tn T0 Tn
K T0 , . . . , Tn , , . . . , /(Ti ) ' K ,...,
Ti Ti Ti Ti
and we have: h i
T0 Tn
V (Ti ) = Spec K ,..., . (2.37)
Ti Ti
The affine schemes V (Ti ) are glued together as in the³ preceding
´ example;
S
therefore, from the set-theoretical point of view, X = Di ∪ PnK .
0≤i≤n
Thus, X is obtained from the (n + 1)-dimensional affine space
An+1
K = Spec K[T0 , . . . , Tn ]
and therefore in W ). The stalks of the structure sheaves at these points are
the quotient fields of A and B respectively.
To prove the converse statementh inotice first of all that if A has no zero
f
divisors, then Spec A and Spec A , where f, g ∈ A, have isomorphic big
g
open sets: h i h i ³ h i´
1 1 1 f
A =A , = A . (2.38)
fg f g g f /1
T2 − 1 2T
t1 = , t2 =
T2 + 1 T2 + 1
t2
and its inversion T = establish an isomorphism of rings of quotients
1 + t1
deg r = i ⇐⇒ r ∈ Ri . (2.39)
We associate with X the graded ring R = K[T0 , . . . , Tn ]/(Fi )i∈I . Let us list
several schemes that can be constructed from K[T0 , . . . , Tn ] are related with
the following geometric objects:
Fig. 4
This is the usual definition of the projective space. Though the origin in
An+1
K is a homogeneous prime ideal, R+ = (T0 , . . . , Tn ), it only contains the
vertex of the cone, and therefore is excluded from the definition of points
of PnK .
It is convenient to assign to every straight line through the origin its infinite
point. Then, PnK can be interpreted as a hyperplane in An+1 K moved to infinity.
• Proj R. The above implies that Proj R corresponds to the base of C,
which belongs to the hyperplane through infinity, PnK . u
t
2.4.3. The scheme structure on Proj R. Let us define the structure
sheaf on Proj R and show that the obtained ringed space is locally P affine.
For every f ∈ R, set D+ (f ) = D(f ) ∩ Proj R. Clearly, if f = fi , where
S i∈Z+
fi are homogeneous elements of degree i, then D+ (f ) = D+ (fi ), and
i∈Z+
therefore the sets D+ (f ), where f runs over homogeneous elements of R, form
a basis of a topology of Proj R. For any homogeneous f ∈ R, the localization
Rf , is, clearly, graded:
Ψf
D+ (f ) / Spec(Rf )0
O O
(2.43)
Ψf g
D+ (f g) / Spec(Rf g )0
commute.
(Here the left vertical arrow is the natural embedding, and the right one
is induced by the natural ring homomorphism Rf −→ Rf g .)
ff )0 ) transported onto D+ (f ) via ψf are glued
Corollary. The sheaves ψf∗ (R
together and determine a scheme structure on Proj R.
Proof of Proposition. Since D(f ) ∩ D(g) = D(f g), we obtain a).
To prove b), define ψf as the through map
where the first arrow is the natural embedding, the second one is the isomor-
phism and the last one is induced by the ring monomorphism.
Clearly, ψf is continuous. Let us show that it is one-to-one: construct the
inverse map ϕ : Spec(Rf )0 −→ D+ (f ). Let p ∈ Spec(Rf )0 . Set
D+ (f ) ∩ D+ (g) = D+ (f g) −→ Spec(Rf g )0 =
(2.46)
Spec((Rf )0 )(gd /f e ) −→ Spec(Rf )0
which also shows the possibility of gluing Spec(Rf )0 and Spec(Rg )0 along
D(f g) since
R1 ∼ R2 ⇐⇒ Proj R1 ∼
= Proj R2 (2.50)
The space Proj R for such rings R is most close to the classical notion
of a projective algebraic variety over K. In particular, if dimK R1 = r+1, then
the epimorphism of Z-graded rings K[T0 , . . . , Tr ] −→ R sending T0 , . . . , Tr
into a K-basis of R1 determines an embedding
implying ai ∈ Z. u
t
2.5.4. Hilbert polynomial of the projective space. Let us apply Theo-
rems 2.7.4 and 2.7.5 to calculation of the Hilbert polynomial for the projective
space over a field and prove that it does not depend on the representation
PrK = Proj R.
Let PrK = Proj R; and let temporarily Or (1) denote the invertible sheaf
on PrK constructed with the help of K[T0 , . . . , Tr ], i.e., from the standard
representation PrK as Proj K[T0 , . . . , Tr ]. Then by Theorem 2.7.5 we have
Or (1) ' O(d) for some d ∈ Z.
For r ≥ 1, we have d > 0, since the rank of the space of sections of Or (n)
grows as n −→ ∞.
On the other hand, due to a (yet not proved!) part of Theorem 2.7.4 for
sufficiently large n the map
implying
P d l
1/1 = (fi /g )ai , where ai ∈ (Rg )0 ;
P mi (2.60)
gn = fi bi , where bi ∈ R.
Since all these arguments are reversible, 1) and 2) are equivalent. u
t
120 Ch. 2. Sheaves, schemes, and projective spaces
Q
vanish anywhere on Y except at y. Then bx = ay vanishes everywhere
y∈Y \{x}
P
on Y except at x, whereas bx does not vanish anywhere on Y . This is
x∈X
a contradiction.
Now, let D+ (f ) = X; then Γ (X, OX ) = R(f ) and it suffices to establish
that
dimK R(f ) = dimK Rn for all n ≥ n0 . (2.64)
S
We have (Rf )0 = Rn /f n , and Rn /f n ⊂ Rn+1 /f n+1 . Since dimK (Rf )0 < ∞,
we see that (Rf )0 = Rn /f n for all n ≥ n0 , and therefore
implying ³ ´
X d 1
hr (n)tn = hr t . (2.68)
dt 1−t
n∈Z+
where dots stand for the terms with the poles of orders ≤ k at t = 1. This
and definitions of deg and χ imply 1) and 3).
To prove 2), observe that
³ ´
hR (0)
Res − dt = hR (0) = χ(Proj R),
t=1 t(1 − t)
implying µ³ ¶
´k−1
FR (t) d 1
Res dt = Res d t = 0.
t=1 t t=1 dt 1−t
u
t
2.5.7. Example. Let f ∈ Rd be not a zero divisor. Given Fr (t), it is easy
to calculate FR/f R (t):
implying
∞
X
FR/f R (t) = (hR (n) − hR (n − d))tn = (1 − td )FR (t) + P (t),
n=0
1 1
FPr (t) = F r−1 (t) = . . . = (2.71)
1−t P (1 − t)r+1
implying ³ ´
1 dn 1 n+r
hPr (n) = = ;
n! dtn (1 − t)r+1 r
dim Pr = r,
(2.72)
deg Pr = 1,
χ(Pr ) = 1.
2.5.7a. Theorem (Bezout’s theorem). Let f1 , . . . , fs ∈ R = K[T ], where
T = (T0 , . . . , Tr ), be homogeneous polynomials of degrees d1 , . . . , ds , respec-
tively. Let Y = Proj R/(f1 , . . . , fs ). Then
plane. If we wish to single out one of them, we have to take for f3 , say, the
equation of the plane through the line, and it is not difficult to see that f3 is
a zero divisor in R/(f1 , f2 ). This straight line is not a complete intersection
inside of the quadric because a complete intersection should be of degree ≥ 2
and the degree of the straight line is 1.
2.5.7c. A geometric complete intersection. There is an interest-
ing version of the notion of complete intersection. Let, for definiteness,
R = K[T0 , . . . , Tr ] and p ⊂ R a homogeneous prime ideal. The scheme
X = Proj R/p is called a geometric complete intersection if there exists
an ideal p0 ⊂ p such that X 0 = Proj R/p0 is a complete intersection and
r(p0 ) = p. The latter condition means that the space of X 0 is the same as that
of X and the only difference is in the presence of nilpotents in the structure
sheaf of X 0 .
Is it true that any scheme of the form X = Proj R/p is a geometric com-
plete intersection?
2.5.7d. Problem. The answer is unknown even for the curves in the three-
dimensional space: Is it possible to define any irreducible curve by two equa-
tions? 2)
If dim X = r − 1, where X ⊂ Pr , the answer is positive:
Proposition. If X ⊂ Pr and the dimension of every irreducible component
of X is r − 1, then X is a geometric complete intersection, i.e., X is given by
one equation.
Proof. It suffices to consider the case where X is irreducible.
Let X = Proj K[T0 , . . . , Tr ]/p, where p is a prime ideal, and f ∈ P an ir-
reducible homogeneous element (obviously, it always exists). Then
which to each element i(F) −→ P from the left-hand side group assigns the
through map i(F) −→ P −→ P+ from the right-hand side group (the presheaf
morphism P −→ P+ is described in sec. 2.1.5).
2.6 The presheaves and sheaves of modules 127
Γ (X, OX ) Zp
(2.85)
² ²
Γ ({(0)}, OX ) Qp
Since the only open neighborhood of the point (p) is X, the sheaf
F = (F1 , F2 ) is quasi-coherent if and only if there is an exact sequence of the
(I) (J)
form OX −→ OX −→ F −→ 0. In terms of (F1 , F2 ) this can be expressed as
two exact sequences forming a commutative diagram:
Zp
(I) / Z(J) / F1 /0
p
(2.86)
² ² ²
Qp
(I) / Q(J) / F2 /0
p
N
This immediately implies that F2 ' F1 Qp . It is not difficult to see that
Zp
this condition is also sufficient for quasi-coherentness of F = (F1 , F2 ).
Thus, a quasi-coherent sheaf in this case is uniquely determined by the
module of global sections F1 = Γ (X, OX ); while F2 and the homomorphism
F1 −→ F2 are recovered from F1 . Without this quasi-coherentness condi-
tion we have a greater freedom in defining both F2 and the homomorphism
F1 ⊗Zp Qp −→ F2 : Now the sheaf may suffer a jump at a generic point as
compared with the quasi-coherent case.
In sec. 2.6.4 the result of this example will be generalized to general affine
schemes.
2.6.3. Coherent sheaves. A restriction of “finite type” distinguishes co-
herent sheaves from quasi-coherent ones.
A sheaf F of OX -modules is said to be a sheaf of finite type if it is lo-
(n)
cally isomorphic to the image of OX for some n (in other words, for every
x ∈ X, there exists an open neighborhood U 3 x and a sheaf epimorphism
(n)
OX |U −→ F|U −→ 0). A sheaf of OX -modules is said to be coherent if it is
of finite type and, for every open U and every sheaf morphism
(n)
ϕ : OX |U −→ F|U −→ 0, (2.87)
i : AS ⊗A M ∼
= MS
³ ´ (2.91)
a am
i ⊗m = . u
t
s s
M∼
= Γ (X, M e) ∼
f) −→ Γ (X, N = N. (2.94)
f|D(f ) = M
and by definition of M we have M fi and M fij |D(f f ) = M
fij . It only
i i j
remains to apply Theorem 2.6.5 and the definition of M . u
t
2.6.7. Example. Let (X, OX ) be a scheme, JX ⊂ OX a quasi-coherent sheaf
of ideals. The quotient sheaf OX /JX is obviously quasi-coherent. Define the
support of OX /JX by setting
{sij ∈ (Γ (Ui ∩ Uj , O×
X ) for i, j ∈ I}. (2.104)
sij sji = 1 if i 6= j
(2.105)
sij sjk ski = 1 if i 6= j 6= k 6= i.
All such sets (2.104) constitute a group with respect to multiplication called
the group of 1-dimensional Čech cocycles of the cover (Ui )i∈I with coefficients
in the sheaf O× 1 ×
X and denoted by Z ((Ui )i∈I , OX ).
0 1
Two cocycles (sij ), (sij ) ∈ Z are said to be equivalent if there exist
t i ∈ O× 0 −1 −1
X |Ui such that sij = ti sij tj . The elements ti tj obviously constitute
1 × 1
a subgroup B ((Ui )i∈I , OX ) ⊂ Z of coboundaries.
The corresponding quotient group is called the first Čech cohomology group
of X with coefficients in O× and denoted by H 1 ((Ui )i∈I , O× X ).
The above constructed Čech cocycles for an invertible sheaf is multiplied by
e
a coboundary if we change isomorphisms {ϕi }. Indeed, let ϕ0i : L|Ui −→ OX |Ui
−1
be another set of isomorphisms. Since ϕi ϕ0 i ∈ Aut(L|Ui ), we get ϕ0i = ti ϕi
implying s0ij = ti sij t−1
j .
Proposition. The above map from the set of invertible sheaves L on X trivial
×
on a given cover (Ui )i∈I into the set of first Čech cohomology H 1 ((Ui )i∈I , OX )
determined from the same cover is one-to-one.
2.7 The invertible sheaves and the Picard group 133
H 1 ((Ui )i∈I , O×
X ) −→ Pic X (2.106)
whose image is the set of classes of sheaves trivial over all the Ui .
Let (Uj0 )j∈J be a finer cover. Then a natural monomorphism
H 1 ((Ui )i∈I , O× 1 0 ×
X ) −→ H ((Uj )j∈J , OX ) (2.107)
arises (we leave the task to precisely formulate its definition to the reader).
Since every invertible sheaf is trivial on elements of a sufficiently fine cover,
we get
Pic X ' lim H 1 ((Ui )i∈I , O× 1 ×
X ) = H (X, OX ), (2.108)
−→
where the inductive limit is taken with respect to an ordered system of cov-
erings.
2.7.3. Example. SOn X = Proj R, where R is generated by R1 over R0 , con-
sider a cover X = Uf , where Uf = D+ (f ), and a cocycle sf g ∈ Z 1 (Uf , O×
X)
f ∈R
given by
sf g = (f /g)n ∈ Γ (Uf ∩ Ug , O×
X ). (2.109)
The invertible sheaf determined with the help of this cocycle is denoted by
OX (n); obviously, we have
(
OX (1)⊗n if n ≥ 0,
OX (n) ' (2.110)
OX (−1)⊗n if n ≤ 0,
Proof. Let us construct α and show that its kernel is only supported in small
degrees. The statement on isomorphism will be proved in what follows.
Let h ∈ Rn . For any f ∈ R1 , set
h
α(h)|D+ (f ) = ∈ Γ (D+ (f ), OX ). (2.112)
fn
h
Obviously, the sections are glued together with the help of the cocycle
fn
(f /g)n into a section of Ln over the whole X; denote this section by α(h).
Clearly, α is a homomorphism of graded rings.
h
Let h ∈ Rn ∩ Ker α. This means that = 0 for all f ∈ R1 . Since R1
fn
is Noetherian, there exists an integer m0 such that RmL
h = 0 for m ≥ m0 .
Consider an arbitrary h whose annihilator contains Rm . Obviously,
m≥m0 (h)
all such elements constitute an ideal J ⊂ R.
There are finitely many generators of J since R is Noetherian, and therefore
we can choose one m0 for all generators h; so that Jm = 0 for m ≥ m0 . u
t
2.7.5. Picard groups: Examples.
Proposition. If A is a unique factorization ring, then Pic(Spec A) = {0}.
Proof.
S In the system of all open coverings the finite coverings of the form
D(fi ) constitute a cofinal subsystem 3) , and therefore it suffices to verify
i∈J
that
H 1 (D(fi ), O×
X ) = {0}. (2.113)
Let sij ∈ Z 1 (D(fi ), O× 0 0
X ). Let us represent all sij for i 6= j in the form ti /tj ,
0
where the ti are elements from the quotient field K of A. It is easy to see
that this is indeed possible: Since sij sjk ski = 1 for any k, it follows that
sij = sik /sjk because sik sik = 1.
Now let p be a prime element of A and vp (a) the exponent with which
p enters the decomposition of a ∈ K. Up to multiplication by invertible el-
ements, the set P of primes p ∈ A such that vp (t0i ) 6= 0 for some i is only
finite.
Fix p ∈ P , and divide all the fi into two groups: The one with p dividing
.
fi for i ∈ J1 and the other with fi 6 .. p if i ∈ J2 . Since the fi are coprime,
J1 6= ∅.
.
Since sij is invertible an Afi fj , we see that vp (sij ) = 0 if fij 6 .. p and vp (t0i )
takes the same value — ap — for all i ∈ J1 . Set
3
Recall that a subset B of a partially ordered set A is said to be cofinal if, for
every a ∈ A, there exists b ∈ B such that a = b.
Also, a sequence or net of elements of A is said to be cofinal if its image is
cofinal in A.)
2.7 The invertible sheaves and the Picard group 135
à !
Y
ti = p−ap t0i (2.114)
p∈P
The first step of the above proof establishes that p∗1 = 0 (actually,
the same argument shows that H 1 (X, K e × ) = 0). The second step shows
∗
that p0 is an epimorphism; and it is only here that we have used the fact
that A is a L unique factorization ring, which, in particular, implies that
Ke × /A
e× ' e
Z.
p∈Spec A
Here is an important application of the above statement:
2.7.5a. Theorem. Let A be a unique factorization ring. Then Pic PrA for
r ≥ 1 is an infinite cyclic group with the class of O(1) as its generator.
Proof. Recall that PrA = Proj A[T0 , . . . , Tr ]. By the above
h proposition,i any
r T0 Tr
invertible sheaf L over PA is trivial on D+ (Ti ) = Spec A ,..., , since
Ti Ti
by a theorem of Gauss (cf. [Pr]) the polynomial ring over A preserves the
unique factorization property of A.
Now let (sij | 0 ≤ i, j ≤ r) be a cocycle defining L for the cover
(D+ (Ti ) | 0 ≤ i ≤ r). Since sij is homogeneous of degree 0 and only fac-
torizes in the product of the divisors of Ti Tj (use the unique factorization
property of A[T0 , . . . , Tr ]), we have
µ ¶nij
Ti
sij = εij , where εij ∈ A× . (2.117)
Tj
Since sij sji = 1 and sij sjk ski = 1, it follows that nij = n (does not depend
on i, j), and therefore εij is a cocycle.
In actual fact, εij is automatically a coboundary, since εij = εik /εjk for
any k and εik ∈ Γ (PrA , O× X ). Therefore (sij ) is cohomologic to the cocycle
(Ti /Tj )n defining O(n).
We will now get theorem’s statement if we prove that all the sheaves
O(n) are non-isomorphic. This is true for any A as shown by the following
statement:
136 Ch. 2. Sheaves, schemes, and projective spaces
Lemma.
0 if n < 0
Γ (PrA , O(n)) = L (2.118)
AT0a0 . . . Trar if n ≥ 0.
a0 +...+ar =n
Proof. Let R = A[T0 , . . . , Tr ], deg Ti = 1 for all i. Let us show that the
homomorphism αn : Rn −→ Γ (PrA , O(n)) is an isomorphism for n ≥ 0.
Recall that αn (f )|D+ (Ti ) = f /Tin for any f ∈ Rn . Now the fact that the
Ti are not zero divisors immediately implies that αn is a monomorphism.
Let us prove that αn is an epimorphism. A section of the sheaf O× over
r
PA is represented by the set
½ h i µ ¶n ¾
T0 Tr Ti
fi ∈ A ,..., | 0 ≤ i ≤ n, and fi = fj . (2.119)
Ti Ti Tj
Since Ti are not zero divisors, the compatibility conditions imply that fi Tin
do not depend on i. Obviously, fi Tin is a polynomial since its denominator
can only be a power of Ti .
Now let n < 0; then, in the same notation, we get
and similar divisibility considerations show that this is only possible for fi = 0.
u
t
Theorem is also proved. u
t
2.7.5b. Corollary. O(n) ∼
6 O(m) if n 6= m.
=
Proof. This is an immediate corollary of the above Lemma since the ranks
of the A-modules of sections of a certain power of these sheaves are distinct.
u
t
2.7.6. Hilbert polynomial of the projective space. As an application
of Theorems 2.7.5a and 2.7.4 we can now compute Hilbert polynomials of the
projective spaces over a given field and look which of the numerical charac-
teristics of PrR introduced in § 2.5 do not depend on the representation of PrR
in the form Proj R.
Indeed, let Prk = Proj R; temporarily, denote by OR (1) the invertible sheaf
on Prk constructed with the help of R, and let O(1) be the invertible sheaf
constructed by means of the standard representation Prk = Proj k[T0 , . . . , Tr ].
By Theorem 2.7.5a we have
for some d ∈ Z; for r ≥ 1, we have d > 0 since the rank of the space of sections
of the sheaf OR (n) grows as n −→ ∞.
2.8 The Čech cohomology 137
is an isomorphism. Hence
³ ´
nd + r
hR (n) = .
r
In particular, the degree and the constant term of the Hilbert polynomial do
not depend on R, as claimed.
2.7.7. Exercises. 1) Prove that the curve in Prk can be isomorphic to P1k
only if its degree is equal to 1 or 2.
Hint. By definition, any curve in Prk is of the form Proj k[T0 , T1 , T2 ]/(f ),
where f — is a form. Its Hilbert polynomial is computed in § 2.5.
Try to prove that the curve determined by a quadratic form f , is isomor-
phic to P1k if and only if the following two conditions are fulfilled: 1) rank f = 3;
2) The equation f = 0 has a non-zero solution in k.
f
2) Let r ≥ 1; prove that any automorphism f : Prk −→ Prk over k is linear
(what does it mean?).
Hint. Look how f acts on invertible sheaves and on the Picard group.
and
2.8.3b. Corollary. For any scheme X, its finite cover U by affine schemes,
and a quasi-coherent sheaf F on X, we have
Proof. Apply Cartan’s theorem to X and the family of affine open subsets
of X using Corollary 2.8.3a. u
t
Observe without proof that Serre proved the inversion of Corollary 2.8.3a:
if for a scheme X and any quasi-coherent sheaf of ideals J on X, we have
H 1 (X, J) = 0, then X is an affine scheme.
2.8.4. Properties of Čech complexes. Retain notation of Proposi-
tion 2.8.3. Let Ui0 ...ip = Ui0 ∩ . . . ∩ Uip = D(fi0 . . . fip ). We have
f) ' Mf ...f .
Γ (Ui0 ...ip , M i0 ip
p f
³ Each
´ cochain of C (U, M ) for p < r can be represented by a collection of
r
elements of different localizations of the module M :
p+1
dn+1
An+1 / An dn / An−1
z z
zz Dn−1 zzz f −g
fn+1 −gn+1
zzzD zz n−1 n−1
² |zz n ² |zz ²
A0n+1 / A0n / A0n−1
d0n+1 d0n
If there exists a chain homotopy between f and g, then f and g are said to be
chain homotopic. The complex chain homotopic to the one with zero cohomology
is said to be acyclic.
140 Ch. 2. Sheaves, schemes, and projective spaces
Since the passage to the inductive limit commutes with computing coho-
mology, we see that
f) = H(C p (U, M
Ȟ p (U, M f)) = 0.
S
∞
(n) (n)
we see that Mg = Mg . Each space Mg is an A-module and we can
n=0
replace the union by the inductive limit by considering the system
They are compatible with homomorphisms of the system (2.122) and therefore
define the homomorphism of its limit:
lim M (n) −→ Mg .
−→
Its cokernel is, clearly, zero. Any element of the kernel is represented by a chain
of elements gmn , g 2 mn , . . ., where mn ∈ M (n) = M , such that mn /g n = 0
in Mg ; this means that g n+k mn = 0 for k sufficiently large, and therefore the
whole chain represents the zero class. u
t
2.8 The Čech cohomology 141
2.8.4b. Now, in the whole complex (C p (U, M f)), replace the localizations
(n)
of M by their “approximations” M and appropriately define the cochain
operators.
To make the correct expression of the cochains graphic, we first assume
that the elements fi0 , . . . , fip are not zero divisors in M . Let Cnp (M ) denote
the subgroup of C p (U, M f) consisting of cochains such that
n o
(n) m
s(i0 , . . . , ip ) ∈ Mfi ...fi = n
|m∈M .
0 p (fi0 . . . fip )
Let
mi0 ...ip
s(i0 , . . . , ip ) = .
(fi0 . . . fip )n
Then
p+1
X
mi0 ...ip+1 mi0 ...bik ...ip+1
ds(i0 , . . . , ip+1 ) = = (−1)k ,
(fi0 . . . fip+1 )n
k=0
(fi0 . . . fbik . . . fip+1 )n
implying that
p+1
X
mi0 ...ip+1 = (−1)k fink mi0 ...bik ...ip+1 . (2.123)
k=0
p
The embedding homomorphism Cnp (M ) −→ Cn+1 (M ) is described by the
formula
mi0 ...ip −→ fi0 . . . fip mi0 ...ip . (2.124)
We use formulas (2.123) and (2.124) to define both the differential in the com-
plex Cnp (M ) when the condition on zero divisors is not satisfied, and complex
p
homomorphism Cnp (M ) −→ Cn+1 (M ).
In the general case, denote by Cnp (M ) the group of skew-symmetric func-
tions on [1, r]p+1 with values in M and define the coboundary operator
Cnp (M ) −→ Cnp+1 (M ) by setting:
p+1
X
(dm)(i0 , . . . , ip+1 ) = (−1)k fink m(i0 , . . . , bik , . . . , ip+1 ). (2.125)
k=0
p
Define the group homomorphism ϕn = Cnp (M ) −→ Cn+1 (M ) by the for-
mula:
(ϕn m)(i0 , . . . , ip ) = fi0 . . . , fip m(i0 , . . . , ip ). (2.126)
2.8.4c. Lemma. 1) The collection of sets (Cnp (M )) for n fixed is not a
complex for any n.
2) The collection of homomorphisms ϕn is a complex homomorphism.
f); the inductive limit of differentials is the dif-
3) lim Cnp (M ) = C p (U, M
−→
ferential in the inductive limit.
The first two statements are verified by trivial calculations; the third one
follows from definitions and Lemma 2.8.4c.
This is the end of stage a) of computing cohomology of the Čech complex,
i.e., its approximating by complexes Cnp (M ) which are easier to deal with.
142 Ch. 2. Sheaves, schemes, and projective spaces
2.8.4d. Stage b) of the program. Let us now pass to the proof of acyclic
property of the complex Cnp (M ). In the construction of this complex there
are involved the ring A, the A-module M , the elements f1n , . . . , frn ∈ A that
determine the cover {D(fin )}ri=1 = {D(fi )}ri=1 , and the differential given by
formula (2.125).
Since the complex Cnp (M ) is important in various problems of algebraic
geometry, we will study it in more detail than is strictly necessary for our
purposes.
2.8.5. Koszul complex. Let A be a ring, f = (f1 , . . . , fr ) a collection of
its elements; set f n := (f1n , . . . , frn ).
Consider a free A-module Ae1 ⊕ . . . ⊕ Aer = Ar of rank r and its exterior
p r
powers³ K
´p = ΛA A ; by definition, K0 = A.Clearly, Kp is a free A-module of
r
rank ; the elements ei1 ∧ . . . ∧ eip , where i1 < . . . < ip , constitute its basis.
p
Define the differential d : Kp+1 −→ Kp by setting:
p+1
X
d(ei1 ∧ . . . ∧ eip+1 ) = (−1)k+1 fik ei1 ∧ . . . ∧ ebik ∧ . . . ∧ eip+1 .
k=1
(NB: −1 is raised to power k + 1 in order for the first term enter with its
initial sign.) It is trivial job to verify that d2 = 0; let Kp (f, M ) or briefly
Kp (f ) be the complex obtained. (Observe that it is a chain complex, whereas
the Čech complex is a cochain one.)
The relation between complexes Kp (f ) and Cnp (M ) is as follows.
2.8.5a. Lemma. For p ≥ 0, we have
Kp+1 (f n ) −→ M,
gm : ei0 ∧ . . . ∧ eip 7−→ m(i0 , . . . , ip ).
(i)
Let now K0 = A, and K1i = Aei for all i. Construct a complex by setting
0 ←− A ←− Aei , d(ei ) = fi .
Then (K (1) ⊗ . . . ⊗ K (r) )p is a free A-module with a basis ei1 ⊗ . . . ⊗ eip , where
1 ≤ i1 < . . . < ir ≤ p, and the differential
p
X
d(ei1 ⊗ . . . ⊗ eip ) = (−1)k+1 fik ei1 ⊗ . . . ⊗ ebik ⊗ . . . ⊗ eip .
k=1
dh(ei1 ∧ . . . ∧ eip+1 ) =
µµX r ¶ ¶ µr=p−1
X ¶
d gk ek ∧ ei1 ∧ . . . ∧ eip+1 = d gjk ejk ∧ ei1 ∧ . . . ∧ eip+1 =
k=1 k=1
r=p−1
X p+1
X
(gjk fjk ei1 ∧ . . . ∧ eip+1 ) + gjk (−1)l fil ejk ∧ . . . ∧ ebil ∧ . . . ∧ eip+1 .
k=1 l=1
and
H 0 (U, F) = Z 1 ((f ), M ),
K 0 ((f ), M ) = K 0 ((f n ), M ) = Hom(A, M ) = M,
2.9 Cohomology of the projective space 145
H p (Pr−1 p
A , O(n)) = Ȟ (U, O(n)).
Therefore we can compute the cohomology of the Čech complex of the cover U .
Since Ui0 ,...,ip = D+ (Ti0 . . . Tip ) = Spec R(Ti0 ...Tip ) , we have, for the usual
description of the sheaf O(n):
n o
m(i0 , . . . , ip )
Γ (Ui0 ...ip , O(n)) = | k ∈ Z, m ∈ Rk(p+1)+n .
(Ti0 . . . Tip )k
This formula
Lindicates that it is convenient to compute the direct sum of Čech
complexes C 0 (U, O(n)), and its cohomology, tracing the natural grading,
n∈Z
and at the end separate the homogeneous components in the answer.
2.9.2. Ckp (U, O(n)) Fix k ∈ Z. Denote by Ckp (U, O(n)) the subgroup of
m(i0 , . . . , ip )
chains whose components can be represented as k
, where m is a
L i0 . . . Tip )
(T
form. As p varies, these groups form a complex Ckp (U, O(n)); computing
n∈Z
the action of its differential on the numerators of the cochain’s component we
easily obtain:
L p
Lemma. 1) The complex Ck (U, O(n)) with its grading is isomorphic to the
n∈Z
Koszul complex K p+1 (T1k , . . . , Trk ; R) with the grading in which the elements
g ∈ Hom(Kp+1 (T k ), R) such that
and
⊕C p (U, O(n)) = lim ⊕Ckp (U, O(n))
−→
k
by setting:
where
(−1)σ = ε(i1 . . . bik . . . ir−p ; j1 . . . jp+1 ).
5
Exercise. Determine in which one.
2.9 Cohomology of the projective space 147
where
(−1)τ = ε(i1 , . . . , ir−p ; j1 , . . . , b
jl , . . . , jp+1 ).
Comparing these answers we see how ϕ commutes with the differentials. u
t
Next, we use Lemma 2.8.5a which implies that
and
d(li e0 + li−1 e1 ) = (dli + (−1)i−1 f li−1 )e0 + dli−1 e1 .
This implies a commutative diagram with exact rows
d d d
² ² ²
0 / Li−1 / (L ⊗ K(f ))i−1 / Li−2 /0
All these diagrams can be united into a sequence of complexes. This sequence
is exact in dimensions ≥ 1:
148 Ch. 2. Sheaves, schemes, and projective spaces
where L[−1]i := Li−1 (the complex L, shifted by 1 to the right). In its turn,
this sequence leads to an exact sequence of homology groups
r−1
p
−r
Fig. 5
The following properties of the continuation of the sheaf operation are rather
obvious.
First,
H q (X, F) = H q (PrA , j∗ (F)) for any q,
which is easy to see on the level of Čech complexes: in this case they are just
isomorphic modules with differentials. (For opens on PrA we take, as always,
the sets D+ (Ti ); for opens in Proj R we take, in order to establish an isomor-
phism desired of complexes, the sets D+ (ti ), where ti is the generator of R1
corresponding to Ti .) We may also use of the fact that F −→ j∗ (F) is a fully
faithful functor.
Second, we have j∗ (F(n)) = j∗ (F)(n), where F(n) = F ⊗ OX (n). Now,
heading ) of Serre’s theorem immediately follows from skew-symmetry of
Čech cochains and the fact that Čech cohomology for the cover (D+ (Ti ))
coincide with the usual cohomology.
To prove headings b) and c) of Serre’s theorem we use the following tech-
nical result.
2.10.1a. Lemma. For any coherent sheaf F on PrA , there exists an integer
m such that for a natural p we have the following exact sequence
OpPr −→ F(m) −→ 0.
A
From this Lemma the needed facts are established by a simple descending
induction on q.
Let us define the coherent sheaf E from the exact sequence
0 −→ E −→ OpPr −→ F(m) −→ 0.
A
In the exact cohomology sequence (2.128), the A-module H q+1 (PrA , E(n)) is
Noetherian and vanishes for n ≥ n0 by induction hypothesis; for H q (PrA , O(n))
the same is true thanks to Theorem 2.9.9. This implies the desired for
H q (PrA , F(m + n)). u
t
It remains to prove Lemma 2.10.1a.
Proof of Lemma 2.10.1a. We may assume that F is a coherent sheaf on
PrA . For the standard cover Ui := D+ (Ti ), having identified F|Ui with F(m)|Ui ,
we see that F(m) is glued of the F|D+ (Ti ) by means of the cocycle:
³ ´m ∼
Ti
: (F|Ui )|Ui ∩Uj −→ (F|Uj )|Uj ∩Ui .
Tj
Therefore, for q sufficiently large (and also common for all i), we have
³ ³ ´p ´³ ´q
Ti T0
s0i − s0j = 0. (2.129)
Tj Tj
Set ³ ´q
T0
s00i = s0i .
Ti
2.11 Sheaves on Proj R and graded modules 153
Clearly,
s000 = s0 .
Besides, s000 enters as a component into the section (. . . s00i . . .) of the sheaf
F(p + q). Verification (see (2.129)):
³ ´p+q ³ ´q ³ ´p+q ³ ´q
Ti T0 Ti T0
s00i = s00j ⇐⇒ s0i = s0j . u
t
Tj Ti Tj Ti
αn : Rn −→ Γ (X, OX (n)).
(see Theorem 2.7.4). The definition makes clear that the homomorphisms
are compatible with multiplication and hence provide with a homogeneous
homomorphism of graded rings
α : R −→ Γ∗ (X, OX ).
f = Γ∗^
β: M (X, F) −→ F.
Set ³ ´
m
β = α(m)|D+ (f ) · α(f )−n |D+ (f ) ∈ Γ (D+ (f )).
fn
2.11.4a. Exercise. Verify that the notion is well defined.
It only remains to establish that β is an isomorphism. In the same way as
for the structure sheaf, one can prove that Ker αn = 0 for n sufficiently large.
This implies that β is a monomorphism; indeed,
2.11 Sheaves on Proj R and graded modules 155
where L and L0 are free graded R-modules, i.e., direct sums of R-mod-
ules R(n).
This gives an exact sequence of sheaves
f˜(n)
e
L(n) e 0 (n) −→ M
−→ L f(n) −→ 0,
which immediately implies that the sheaf F is coherent and with the help of
which one constructs the exact cohomology sequence in the bottom line of the
following commutative diagram:
Ln / L0n / Mn /0J
JJ
JJ
JJ
α α α JJ
JJ
² ² ² $
e
Γ (X, L(n)) / Γ (X, L
e 0 (n)) / Γ (X, F(n)) / H 1 (X, f (L)(n)).
e
e
By Serre’s theorem H 1 (X, f (L)(n)) = 0 for n ≥ n0 , and since X = PrA ,
Lemma 2.7.5a shows that the first two vertical arrows are isomorphisms.
Hence, α : Mn −→ Γ (X, F(n)) is also an isomorphism. u
t
2.11.6. Main theorem on correspondence F −→ Γ∗ (X, F). We for-
mulate it without proof: A good deal of it is already verified; to prove the rest
of it does not require any new ideas.
Let R be a graded ring satisfying the conditions of Serre’s theorem. Let
GMR denote the category defined as follows:
156 Ch. 2. Sheaves, schemes, and projective spaces
0 −→ F1 −→ F −→ F2 −→ 0
0 −→ F0 −→ . . . −→ Fk −→ 0
The sheaves Ke and C e correspond to k[T1 , . . . , Tr−1 ]-modules, i.e., are “con-
r−1
centrated” on Pk . This allows one to make the induction step: For r = 0.
the statement is trivial. u
t
158 Ch. 2. Sheaves, schemes, and projective spaces
or
h1 (n) ≤ h2 (nN ) for any n ≥ n0 .
By symmetry, h2 (n) ≤ h1 (nN 0 ), so deg h1 = deg h2 . Theorem is proved. u
t
Let us give now another description of the dimension, often useful. Recall
that f ∈ R is said to be an essential zero divisor in the graded R-module M ,
f
if the kernel of multiplication by f : M −→ M has infinitely many non-zero
homogeneous components, i.e., is non-trivial in the category GMR .
2.12 Applications to the theory of Hilbert polynomials 159
f0
0 −→ N −→ M −→ M (k) −→ M/f0 M −→ 0,
f0
where k = deg f0 and M −→ M (k) is the grading-preserving homomorphism
of multiplication by f0 . Since f0 ∈ R is an inessential zero divisor, we have
Nn = 0 for n ≥ n0 and dim M (k)n − dim Mn = dim(M/f0 M )n . Therefore
f = (d − 1) + 1 = d, proving the desired.
dim M u
t
2.12.9. Proof of Lemma 2.12.7. In R, we have to find an inessential zero
divisor in M .
First of all let us show that there exists a sequence of graded modules Mi
such that
0 = F0 ⊂ F1 ⊂ F2 ⊂ . . . ⊂ F,
such that Fi+1 /Fi ' OX /Ji (ni ), where Ji are coherent sheaves of prime ideals
on X.
2.12.11. Theorem. H q (X, F) = 0 for q > dim F.
Proof. Let the theorem be established for some sheaves F0 and F00 entering
the exact sequence
0 −→ F0 −→ F −→ F00 −→ 0.
Then
χ(F(n)) = χ(F0 (n)) + χ(F00 (n)),
and hence dim F = max(dim F0 , dim F00 ). The exact cohomology sequence
easily implies the validity of the statement of the theorem for F: It suffices to
consider the terms
L
∞
considered as the projective spectrum of the graded ring Γ (X, OX /J(n)).
n=0
Therefore, and thanks to heading 2) of Corollary 2.12.10 it suffices to verify
the statement for OY (n) = F. Set J = Im Γ∗ (X, J) ⊂ R.
Let d = dim Y . By Theorem 2.12.6 there exists a maximal R/J-sequence
f0 , . . . , fd , where fi ∈ R such that dimk R/(f0 , . . . , fd , J) < ∞. So setting
f i = fi (mod J) we have (R/J)n ⊂ (f 0 , . . . , f d ) for n ≥ n0 in R/J. Ge-
S
d
ometrically, this means that Y = D+ (f i ). Computing Čech cohomology
i=0
immediately yields the statement of Theorem. u
t
2.12.12. Remark. Since dim F ≤ dim X for any F (as follows from the
proof of Theorem), then, in particular, H q (X, F) = 0 for any q > dim X.
e 1 −→ F −→ F2 −→ 0
0 −→ F
e = ψ(F
ψ(F) e 1 ) + ψ(F
e 2 ).
e 1 −→ . . . −→ FK −→ 0,
0 −→ F
we have
K
X
e i ) = 0.
(−1)i ψ(F
i=1
e0 ⊂ F
b) Let 0 = F e1 ⊂ . . . ⊂ F
e K = F.
e Then
X
ψ(F)e = ψ(Fe i /F
e i−1 ).
i≥1
2.13.4. K(X), the Grothendieck group. The group K(X) = Z[CohX ]/J
is called the Grothendieck group (of the category CohX or the scheme X).
2.13.4a. Proposition. The function
for which k(F) = [F] (mod J) is additive; the image k(CohX ) generates the
group K(X), and for any additive function ψ : CohX −→ G, there exists a
uniquely determined homomorphism ϕ : K(X) −→ G such that ψ = ϕ ◦ k.
Proof is trivial.
From the point of view of the group K(X), the Riemann-Roch problem is
2.13 The Grothendieck group: First notions 163
is a group isomorphism.
2) Let A be a principal ideal ring, X = Spec A. Then any Noetherian
A-module M has a free projective resolution of length 1:
0 −→ Ar −→ As −→ M −→ 0.
This immediately implies that the group K[X] is cyclic and generated by the
class of ring A. The order of
Nthis class is equal to infinity which is easy to see
passing to linear spaces M K over the ring of quotients K of A; and hence
A
K(X) ' Z.
More generally, the same is true for any affine scheme Spec A provided
any Noetherian A-module has a free resolution of finite length.
This condition is still too strong to lead to interesting notions; however,
even a slight slackening of it defines a very important class of schemes.
2.13.6. Smooth schemes. Let X be a Noetherian scheme, F a coherent
sheaf on it. Let, for any point x ∈ X, there exists an open neighborhood U 3 x
such that the sheaf F|U has in this neighborhood a free resolution consisting
of “free” sheaves OrX |U . Then the scheme X is said to be smooth.
164 Ch. 2. Sheaves, schemes, and projective spaces
In the next section we will prove that the two types of schemes are smooth:
the projective spaces over fields and spectra of local rings whose maximal
ideals are generated by regular sequences.
Smoothness of projective spaces immediately follows from the following
classical Hilbert’s theorem on syzygies. 6)
2.13.7. Theorem. Let R = K[T0 , . . . , Tr ]. Any graded R-module graded free
projective resolution of length ≤ r + 1.
Proof will be given at the end of the chapter; now we use this theorem to
compute K(PrK ).
2.13.8. Theorem. The map xi 7−→ k(O(i)), where k is the function from
eq. (2.132), determines an isomorphism of Abelian groups
. . . −→ R( ) −→ . . . −→ R( r+1
2 ) −→ Rr+1 −→ R −→ K −→ 0,
r+1
i
r+1 ³
X ´
r+1
k(O(n − i)) = 0.
i
i=0
which is only possible if ai = 0 for all³i since, as ´an easy verification shows (say,
n+i+r
by induction on r), the polynomials in n are linearly independent
r
for i = 0, . . . , r. u
t
2.13.9. The group K(PrK ) in formulation of Theorem 2.13.8 turned out en-
dowed with a ring structure. The multiplication in this ring possesses an
invariant meaning: Indeed, as is easy to see
k(F)k(OPr (i)) = k(F(i)) = k(F ⊗ OPr (i)),
so this multiplication corresponds, at least sometimes, to tensor products of
sheaves. There are, however, examples for which k(F)k(G) 6= k(F ⊗ G), So the
general description of multiplication can not be that simple. This question is
studied in detail in the second part of the course [Ma3].
Meanwhile, using our description of K(PrK ), we give a (somewhat naive)
form of the Riemann-Roch theorem for the projective space.
The idea is to select some simple additive function on K(PrK ), and then
“propagate” it using ring multiplication.
Any element of K(PrK ) can be uniquely represented, thanks to Theo-
Pr
rem 2.13.8, as a polynomial ai (l − 1)i , where l = k(O(1)). Introduce an
i=0
additive function κr : K(PrK ) −→ Z by setting
µXr ¶
i
κr ai (l − 1) = ar .
i=0
166 Ch. 2. Sheaves, schemes, and projective spaces
ψ : K(PrK ) −→ Z,
Proof. It suffices to verify the coincidence of the left side with the right side
for the elements li , where i = 0, 1, . . . , r, constituting a Z-basis of the group
K(PrK ). We have:
³ ´
r+i
χ(li ) = χ(O(i)) = dim H 0 (PrK , O(i)) = ,
i
³ ´
r+i
κr (lr+i ) = κr ((1 + (l − 1))r+i ) = . u
t
i
2.13.12. Remark. The usage of κr as a simplest additive function was not
yet motivated. Besides, it is clear that to apply Theorem 2.13.11 is not that
easy: To compute χ(F) with its help, we have to first know what is the class
of the sheaf F in K(PrK ). The only means known to us at the moment is to
consider the resolution of the sheaf F, but this is not “geometric”, besides,
this makes our formula useless: If we know the resolution, we can calculate
χ(F) just by additivity.
Nevertheless, formula (2.133) is very neat; I consider this as a serious
argument in its favor.
f
Orx −→ Osx −→ Fx −→ 0,
f : Ar −→ As
0 −→ Orxn |U −→ . . . −→ Orx0 |U −→ F −→ 0.
e x on
Passing to fibers at point x we obtain a finite resolution of the sheaf F
Spec Ox ; this proves that Spec Ox is smooth.
Conversely, let Spec Ox be a smooth scheme, F a coherent sheaf on X.
Since the only neighborhood of a closed point x in Spec Ox is the whole spec-
trum, there exists a resolution of the fiber
0 −→ Orxn −→ . . . −→ Orx0 −→ Fx −→ 0.
The argument analogous to the above one allows us to continue this se-
quence onto an open neighborhood U of the point x:
It is easy to see that Ker d1 ' M ; so in the next part of the resolution this
segment will be periodically repeated again and again.
170 Ch. 2. Sheaves, schemes, and projective spaces
² ψ
²
F/xF / M/xM /0
θ
² ² ²
ψ
1
SA/xA (M/xM ) / F/xF / M/xM /0
is commutative.
Let us show that θ is an epimorphism. Let f + xF ∈ Ker ψ. Then
ϕ(f ) ∈ xM , and hence
d+1 d
SA (M ) = SA (M 0 ), where M 0 = S 1 (M ).
Algebra Element
flat, 59 integer over A, 36
integer over A, 36 Embedding
Annihilator, 45, 48 closed, of a subscheme, 42
Arrow, 87 Epimorphism
Artinian ring, 18 minimal, 169
Equivalence
Bezout’s theorem, 123, 125 birational, 111
Bundle
conormal, 69 Family
normal, 69 normal (of vector spaces), 63
vector, 56 of vector spaces, 56
stably free, 71 Filter
on a set, 29
Cartesian square, 52 Formula
Cartier’s Theorem, 83 Lefschetz, 76
Category, 87 Function
big, 88 flat, 61
small, 88 Functor, 91
Center representable, 93
of a geometric point, 14 corepresentable, 93
Chain of points, 22 point, 93
Characteristic sheafification, 164
of projective scheme, 118
Chevalley’s theorem, 36 Galois group, 81
Co-normal Genus
to the diagonal, 67 arithmetic, of projective scheme, 118
Cocycle Germ
Čech, 132 of a section, 101
Codimension, 64 of neighborhoods, 39
Index 179